Sunteți pe pagina 1din 278

Geotechnical Finite Element Analysis

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite
Element Analysis
A practical guide

Andrew Lees
BEng PhD CEng MICE

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Published by ICE Publishing, One Great George Street,
Westminster, London SW1P 3AA
Full details of ICE Publishing sales representatives and distributors
can be found at: www.icebookshop.com/bookshop_contact.asp

Other titles by ICE Publishing:


Finite Element Analysis in Geotechnical Engineering: Volume two
Application
D. Potts and L. Zdravkovic. ISBN 978-0-7277-2783-1
Structural Analysis with Finite Elements
P. Rugarli. ISBN 978-0-7277-4093-9
Finite Element Design of Concrete Structures
G. Rombach. ISBN 978-0-7277-3274-3
www.icebookshop.com
A catalogue record for this book is available from the British Library.

ISBN 978-0-7277-6087-6
# Thomas Telford Limited 2016
ICE Publishing is a division of Thomas Telford Ltd, a wholly-owned
subsidiary of the Institution of Civil Engineers (ICE).
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 publisher,
ICE Publishing, One Great George Street, Westminster,
London SW1P 3AA.
This book is published on the understanding that the author is solely
responsible for the statements made and opinions expressed in it
and that its publication does not necessarily imply that such
statements and/or 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 author or publishers.
While every reasonable effort has been undertaken by the author
and the publishers to acknowledge copyright on material
reproduced, if there has been an oversight please contact the
publishers and we will endeavour to correct this upon a reprint.
Commissioning Editor: Laura Balchin
Development Editor: Maria Ines Pinheiro
Production Editor: Rebecca Norris
Market Development Executive: Elizabeth Hobson

Typeset by Academic + Technical, Bristol


Index created by Simon Yapp
Printed and bound in Great Britain by TJ International Ltd, Padstow

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Contents Preface
About the author
vii
ix

01 . . . . . . . . . . . . . . . . . . . . . . . . . . . . How is a geotechnical finite element analysis


set up? 1
1.1. Analysis planning 1
1.2. Geometry 7
1.3. Meshing 17
1.4. Analysis stages 18
1.5. Constitutive models 26
1.6. Groundwater and drainage 27
References 27

02 . . . . . . . . . . . . . . . . . . . . . . . . . . . . How are constitutive models selected? 29


2.1. Introduction 29
2.2. Aspects of ground behaviour 31
2.3. Common constitutive model types 36
2.4. Typical applications 48
References 52

03 . . . . . . . . . . . . . . . . . . . . . . . . . . . . How are soil and rock parameters obtained? 55


3.1. Introduction 55
3.2. Soil and rock sampling and groundwater
measurement 59
3.3. Parameter testing 64
3.4. Parameter derivation and validation 85
Appendix 3.1 Useful equations in the validation of
model or initial state parameters 97
References 99

04 . . . . . . . . . . . . . . . . . . . . . . . . . . . . How are groundwater effects taken into


account? 105
4.1. Introduction 105
4.2. Drained and undrained analyses 109
4.3. Groundwater flow analyses 118
4.4. Consolidation analysis 120
References 123

05 . . . . . . . . . . . . . . . . . . . . . . . . . . . . How are geotechnical structures modelled? 125


5.1. Structural geometry 125
5.2. Structural materials 152
5.3. Soilstructure interaction 156
References 160

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
06 . . . . . . . . . . . . . . . . . . . . . . . . . . . . Can FE analysis be used with design codes? 163
6.1. Introduction 163
6.2. Serviceability limit state (SLS) 167
6.3. Geotechnical ultimate limit state (ULS) 168
6.4. Structural limit states 180
References 181

07 . . . . . . . . . . . . . . . . . . . . . . . . . . . . How is the accuracy of outputs assessed? 183


7.1. Introduction 183
7.2. Assessing accuracy 188
7.3. Managing errors 192
References 197

08 . . . . . . . . . . . . . . . . . . . . . . . . . . . . Examples 199
8.1. Introduction 199
8.2. Raft foundation with settlement-reducing
piles example 199
8.3. Shaft excavation example 225
8.4. Embankment construction example 243
References 261

Index 263

vi

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Preface It soon became clear to me while coordinating the
European Commission Lifelong Learning COGAN
Project on improving competency in geotechnical
numerical analysis that nite element (FE) analysis is
now widely used in geotechnical engineering but, in
contrast to other elds of engineering, there are few full-
time users of such software. Geotechnical FE analysis
places heavy demands on the competency of engineers
but it is difcult to gain sufcient competency when
applying such software part-time between other
engineering tasks. There was an obvious need for a ready
reference for users of geotechnical FE analysis software
to learn about and refresh their knowledge on applying
the technique in practice. This book is intended primarily
to address that need.
Before using this book, it may also be useful to know the
following:

g The book is strictly software neutral. I did not want


to appear to be favouring any particular software.
g I have not endeavoured to cover the essential
background soil mechanics, rock mechanics and
geotechnical engineering knowledge needed to
perform FE analysis since this can be found readily
from other sources.
g Worked examples in FE analysis are complicated to
present and explain. So that readers can access
information quickly, I have avoided putting examples
within the topics in Chapters 1 to 7. Rather, three
examples illustrating application of many of the
topics are presented and described separately in
Chapter 8.
g Some parts of the NAFEMS guidebook Obtaining
Parameters for Geotechnical Analysis which I
authored have been reproduced in this book,
particularly in Chapter 3, with the kind permission of
NAFEMS.
g This book provides the background information
covering about 160 competence statements from the
COGAN Competency Tracker maintained by
NAFEMS. This Competency Tracker is available
online to individuals free of charge for monitoring
and recording competency in geotechnical numerical
analysis.

Andrew Lees
Nicosia
May 2016

vii

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
About the Andrew Lees graduated with a BEng in Civil Engineering
at the University of Southampton in 1996, where he also
author obtained a PhD in the elds of centrifuge modelling and
FE analysis of soilstructure interaction in 2000. He was
then a geotechnical engineer at a major UK consultancy
until 2004 when he took up a lectureship at Frederick
University, Cyprus where he taught geotechnical
engineering until 2015. In 2007, he also set up and
continues to run the successful consultancy Geofem,
specialising in geotechnical FE analysis. In 2016, he was
also appointed Senior Application Technology Manager
at Tensar International, where one of his tasks is to
improve techniques of modelling geogrid-stabilised soils
by FE analysis. He is a member of the NAFEMS
Geotechnical Working Group and authored their rst
guidebook on obtaining parameters for numerical
analysis and is a founding member of the Professional
Simulation Engineer scheme administered by NAFEMS.
He coordinated the European Commission Lifelong
Learning project COGAN on improving competency in
geotechnical numerical analysis. He was convener of the
evolution group advising the Eurocode 7 committee on
the use of numerical methods in accordance with the
design code and has since been involved in the redrafting
of Eurocode 7. He is a member of the British
Geotechnical Association and the Institution of Civil
Engineers.

ix

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.001

Chapter 1
How is a geotechnical finite element
analysis set up?

The following sections in this chapter describe the steps taken and decisions to be made
when setting up a geotechnical nite element (FE) analysis model. In many cases,
readers are referred to sections in subsequent chapters where more detail is provided.
The implementation of these steps is demonstrated in the examples in Chapter 8.

1.1. Analysis planning


1.1.1 Does FE analysis need to be used?
This is an important question because FE analysis usually involves a lot more time and
expense than conventional design methods, so choosing this method needs to be justi-
ed. The mere use of FE methods does not guarantee accurate predictions. Arguably
there is greater scope for error due to the power and complexity of such software.

Non-numerical, or conventional, methods of design are usually quicker and cheaper, but
they have major assumptions (e.g. linear elasticity, uniform ground properties) and
provide limited information (e.g. average settlement of a foundation, limit states). Never-
theless, in spite of the assumptions and probable conservatism, they are often sufcient
to demonstrate a satisfactory design without signicant loss of economy. In such cases
FE analysis cannot normally be justied. However, in other instances there may simply
be no conventional method to calculate the required output, or the greater precision and
detail offered by FE analysis at the design stage could bring signicant economies during
construction.

For example, FE analysis rather than conventional analysis methods might be required
when any of the following need to be considered:

g complex ground behaviour (e.g. non-linear stiffness, hardening soil, anisotropy,


creep), more realistic ground behaviour or changing ground behaviour
(e.g. ground improvement or treatment, consolidation)
g complex hydraulic conditions
g unusual geometry
g soilstructure interaction and internal structural forces in complex structures, and
interactions with adjacent structures
g complex loadings
g the effects of the construction sequence and construction method

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

g applying observational approaches to design


g time effects (e.g. creep, consolidation)
g back-analysis of eld trials or monitored structures.

To help decide whether the use of FE analysis can be justied, a preliminary analysis can
be performed with rudimentary project information and the outputs compared with
appropriate conventional methods to assess the potential economic benet of investing
more time and money at the design stage in FE analysis.

1.1.2 What are the aims of the FE analysis?


Before thinking about building an FE model, the aims of the FE analysis need to be
dened. For example, it may need to be demonstrated that a geotechnical structure has
adequate safety against failure, or that the movement of an adjacent building is small
enough not to cause damage, or to predict the ow of water into a cofferdam. Each
requires a different approach, so the aims need to be dened at the start so that the
decision-making throughout the preparation of the model helps to ensure that the model
provides sufciently accurate predictions. If one of the aims were the prediction of
ground deformations, for example, then software and constitutive models that were
known to produce accurate predictions of ground deformation for the site conditions
would be chosen and parameter testing would focus on obtaining accurate stiffness
parameters for the ground.

From the start, the analysis aims should be discussed with other stakeholders in the
project to help ensure that the FE analysis meets their needs. FE models can take a long
time to prepare and it is frustrating to learn of a new issue near the end of the process
that could have been addressed by the FE model if it had been included in the aims
of the analysis at the start. Some stakeholders will be third parties, particularly if ground
movements might affect adjacent structures, services and infrastructure. So, as part of
the site investigation, check with neighbouring property owners, utility companies and
infrastructure agencies (e.g. highways, railways, metro lines) that their requirements are
covered by the aims of the FE analysis.

Document the aims of the analysis clearly and have them checked by the project stake-
holders so that everyone knows what to expect from the analysis model and to avoid any
misunderstandings. Once agreed, the written aims should be kept close at hand and
referred to whenever decisions are made regarding the FE model and obtaining
parameters.

The outputs from the FE analysis that will be used to meet the specied aims are the
key outputs. Clearly, it is vitally important for these outputs to have sufcient accuracy
because they will inuence the design of the project. Every decision during the design of
the FE model should be made considering its effect on the key outputs.

1.1.3 What information needs to be gathered?


To produce an accurate geotechnical FE model, comprehensive information on the
historical, present day and proposed conditions at the site is needed. This requires an

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

extensive search of information sources, largely as part of the site investigation, as well as
regular communication with members of the project team and third parties. Every
project is different but the types of information gathering often include the following
broad categories:

Ground information
Careful planning of the ground investigation is needed to obtain the information necess-
ary to form a sufciently representative simulation of ground behaviour in the FE
model, and this stage is covered in detail in Chapter 3. Essentially, sample descriptions
and characterisation tests are used to form a ground model representative of site condi-
tions. Then, by referring to the aims of the analysis, the required geotechnical param-
eters can be obtained by dedicated parameter testing.

When interpreting the ndings of the ground investigation and parameter testing results,
it is important to understand the geological history of the site and the mechanisms of
strata formation. The uncertainty in the interpretation of the ground conditions and
parameters needs to be judged in order to select appropriate characteristic values, and
sensitivity analyses are necessary to assess the potential effects of the uncertainties on the
FE model outputs. Regular communication with those undertaking the ground investi-
gation will help in judging the uncertainties.

Historical information
During the desk study stage of a site investigation, information on historical land uses on
and around the site is gathered, but how is this relevant to an FE analysis of todays
situation? Stress history and stress path have signicant effects on the behaviour of the
ground and therefore inuence the input parameters to a model. Also, in order to
recreate the stress path and current stress state in the model accurately, it may be necess-
ary to simulate historical construction stages in the FE model leading up to the present
day situation. Therefore, the gathered historical information should be used to build up a
timeline of signicant loadings (e.g. foundations), unloadings (e.g. excavations), tunnel-
ling and other structures that may exist in the ground (e.g. unused piles or foundations).
When preparing the FE model some of these historical activities may be important
enough to be simulated in the construction stages or may inuence the input parameters
and in situ stresses.

Existing structures and infrastructure information


If the site has existing structures or infrastructure, details of the existing geotechnical
structures (e.g. foundations, retaining walls, slope supports, tunnels, buried services) and
loads from the existing structures and infrastructure will need to be obtained. This may
include structures and infrastructure adjacent to the site where they inuence ground
behaviour or feature in the aims of the FE model.

Ideally, as-built drawings will be available together with designs and load schedules, and
these can be sought from owners of the existing structures and infrastructure. Often such
comprehensive information is not available, particularly for older structures, and some
intrusive investigation of existing geotechnical structures will need to be included in the

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

site investigation. Even with intrusive investigation, assumptions will probably have to
be made regarding existing geotechnical structures, so their type and geometry will need
to be estimated based on experience of similar structures of the same age and by using
design methods appropriate for the period of construction, and different options studied
where there is uncertainty.

Regarding existing loadings, rarely will these be available from the original design of
older structures, so they will have to be estimated based on typical loadings for the type
of structure and its use. Remember that existing loadings can often be favourable: for
instance, an existing structure on a site to be demolished will have pre-loaded the ground
such that settlement of the subsequent structures foundations will be reduced. In such a
case it would be appropriate in an FE model taking account of pre-loading effects to
apply the estimated actual loading rather than an upper bound value typically adopted
for the design of new structures.

Where the aims of the FE analysis include verifying that the settlement or distortion
of adjacent structures and infrastructure are within acceptable limits, the gathered
information could be used to set these limits. Sometimes, particularly for infrastructure,
the owner will provide acceptable deformation limits. On other occasions, the limits
may need to be judged to help ensure that existing structures do not suffer an unaccep-
table level of damage resulting from construction-induced ground movements on the
site.

Proposed structures and infrastructure information


Naturally, information on what is proposed to be constructed on the site will need to
obtained in order to simulate its construction. Consequently, at least the following will
normally be required:

g drawings and building information modelling (BIM) les for the proposed
substructure in order to dene the geometry of the FE model
g sufcient information (drawings and BIM) regarding the superstructure in order
to assess whether this will affect the behaviour of the ground
g proposed loadings on the substructure and foundations and the different load
cases that need to be considered
g limits on acceptable movement and distortion of the substructure and foundations
g proposed construction sequence in order to prepare construction stages in the FE
model
g proposed construction programme in order to estimate time intervals between
construction stages, which will be important for deciding whether to simulate low-
permeability strata as drained, undrained or with consolidation in the various
construction stages, or other temporal effects such as creep.

1.1.4 Which FE analysis software should be used?


Some software will perform certain tasks better than others, so try to choose the soft-
ware most suited to the task. In every case the user needs to know the software very well,
including its strengths and limitations.

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

Look for case studies including FE analysis of similar problems to see which software
was used by others and how well the software performed. Most programs have user
groups providing forums for the exchange of ideas and experiences, as well as ready
access to case studies.

Verify that the software works properly on the computer being used. The range of com-
puters, devices and operating system versions continues to grow and there is always the
potential that one combination of these may not be fully compatible with all features of a
particular FE analysis program (see Section 7.2.1).

1.1.5 How will the FE analysis fit into the design process?
It is likely that the FE analysis will form one part of larger design process, so be aware
of the bigger picture to ensure that the FE analysis work ts within the design process
as seamlessly as possible. A lot of the information gathered for the FE analysis inputs
will be sourced from the main design process while the outputs and recommendations
will be fed back into the subsequent design stages. The FE analysis outputs must meet
the needs of the follow-on designers. For instance, consider the design of a raft foun-
dation where an FE model of the ground volume and foundation has been used to
predict settlement and deformation of the raft. The structural design of the raft may
need to consider multiple load combinations subject to frequent revision so the struc-
tural engineer would prefer to calculate structural forces (bending moment and shear
force) in the raft using his/her own, simpler soilstructure interaction analysis software.
Some of the FE analysis outputs would need to be presented in a form that could
provide input to the structural engineers simpler model (e.g. coefcients of subgrade
reaction for a beam-spring model see Section 5.3.2). Also, outputs of structural forces
from the FE model could be provided to help the structural engineer to validate the
simpler model. In more straightforward cases the structural engineer may use the FE
analysis outputs of structural forces directly in his/her design. Consequently, knowledge
of the wider design process is needed in order to meet the needs of other designers using
the outputs.

Any outputs provided to other designers must be clearly explained to avoid misunder-
standings and delays or errors in the ongoing design process. For example, be clear about
units, axis directions (global and local), sign convention, construction stage, datum
values for outputs, load case and any factors applied to inputs or outputs. Also show
a clear legend for contour and vector plots. Regular communication among the design
team is key to help avoid misunderstandings.

Keep up to date on the wider design process through document management systems
and regular communication to ensure that the FE model stays up to date and that the
outputs and recommendations are relevant to the latest design. It is common for FE
models to be revisited long after they were completed due to delays, changes in design
or issues encountered during construction. This is one of the reasons why a good
write-up of the analysis work is essential (see later in this section) so that engineers can
get up to speed when revisiting an analysis model with minimum delay and without
misunderstandings.

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

The results of the FE analysis may help to make important decisions in the design
process, particularly when considering different design options in the analysis. In order
to present a clear case on the advantages and disadvantages of each option, which of the
many possible outputs should be presented? Being aware of the wider design and nan-
cial issues through discussions with other members of the project team will help in
understanding the key outputs that need to be presented. For example, in an FE analysis
of a basement construction adjacent to other properties, party wall negotiations may be
a critical element to the whole project, so the presentation of FE analysis results could
focus on the predicted foundation movements at the party walls with potential options
to overcome any unacceptable movement, while not forgetting to present any other
outputs considered to be important or ag up potential issues or cost savings to the rest
of the project team.

When adopting an observational approach to design to help manage errors (see


Section 7.3), the FE analysis outputs are compared with site monitoring data during
construction. In such cases, the expected monitoring results based on the FE
analysis outputs should be clearly documented together with ranges of values outside
of which action should be taken on site to modify the construction process. The
project team should be made aware of the importance of the monitoring data both to
the FE analysis model and the project as a whole. Ensure that clear responsibilities
have been assigned for regular viewing and interpretation of the data and that the data
will be fed back into the FE analysis work for validation of the output, as described in
Chapter 7.

As with all engineering design, it is very important to write up calculations in a


clear way so that users of the results can understand the assumptions adopted, to
facilitate checking of work to help avoid errors, to satisfy any approval or licensing
processes and ensure those who revisit the FE model at a later date can get up to speed
quickly. However, write-ups of designs by FE analysis are not straightforward
because the calculations are too complicated to present and are performed by a
computer. Consequently, the requirements for documenting design by FE analysis may
differ from an organisations practice for conventional design. NAFEMS provide
useful guidance on quality assurance procedures for engineering analysis, e.g.
Chillery (2014).

A write-up should include at least the following information. As much of the infor-
mation as possible should be obtained from direct reporting features in the software
to minimise the chance of errors in transferring analysis data to the report:

g background information to the project and how the FE analysis is related to this
summary of information gathered for the FE analysis
g any previous FE analyses superseded by this one
g aims of the FE analysis
g software version and any add-ons plus verication reports
g geometrical assumption (3D, 2D plane strain, 2D axisymmetric) and any axes of
symmetry

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

g plots showing geometry, scale, axis directions, strata, structures, boundary


conditions
g tables showing constitutive models and input parameters for all materials
g derivation of material parameters and their validation
g description of each constitutive model and justication for its selection
g plots of initial stresses, pore water pressure and state variables (elastic, yielding,
etc.)
g tables and geometry plots showing each construction stage, material models
assigned to each element, assumptions (drained/undrained, displacements reset,
etc.), loadings, time (for consolidation analyses), calculation method and
convergence criteria
g as well as presenting the outputs required to meet the aims of the analysis, as a
minimum plots of the deformed mesh, stresses, strains and state variables at key
construction stages should be presented to show satisfactory completion of
calculations
g validation of analysis model
g interpretations, discussions and recommendations based on the analysis results
g any recommended site monitoring, expected values, trigger values and remedial
measures.

1.2. Geometry
1.2.1 2D or 3D?
Whether to build the FE analysis model in three dimensions (3D) or in two dimensions
(2D) using a geometrical assumption (plane strain or axisymmetric) is an important
decision because there can be an enormous difference in the workload between the two
options. Setting up the geometry for a typical 2D analysis model of one section may
take about a day, for instance, while to set up a 3D model of the same structure may
take a whole week due to all the additional geometrical information that must be speci-
ed. So, perform 2D analysis when possible to save time and resources, but only when
the assumptions required to perform 2D analysis will not have a detrimental effect on
the accuracy of the model. The following paragraphs describe some of the effects of the
2D assumptions to assist readers in making the right decision about whether to build a
2D or 3D FE model.

2D plane strain assumption


A 2D plane strain model involves the analysis of a plane, vertical section through the site.
The strain and displacement in the third dimension (i.e. perpendicular to the plane) is
assumed to be zero, hence strains can only occur in directions within the plane and they
are independent of the out-of-plane direction. Consequently, shear stress and shear
strain can be non-zero only in the plane of the analysis, although normal stress perpen-
dicular to the plane is calculated and can be non-zero. This assumption is suited to sites
with a uniform cross-section (including ground conditions) and stress state/loading for a
sufciently long straight dimension for virtually zero strain to be expected in the long
dimension (e.g. straight tunnels, embankments, long excavations, strip foundations),
as shown in Figure 1.1. It is not suited to sites with foundation piling, ground anchors
or similar structural geometries.

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 1.1 Suitable geometries for the plane strain assumption

Plane strain section

Raft foundation

Tunnel

Embankment
Plane strain section

Plane strain section Basement excavation

Raft foundation

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

It is important to visualise what a 2D plane strain model is actually simulating in 3D in order


to understand the limitations of the model. To do this, simply extrude a plane strain model
in the direction perpendicular to the plane. This is illustrated in Figure 1.2 for a basement
example that is probably unsuitable for a plane strain assumption. A section was taken
through the true 3D geometry as shown, which formed the geometry of the 2D plane strain
model. Extruding the section in the out-of-plane direction as shown demonstrates clearly
the geometrical assumptions of the plane strain model and that they are very different from
the true 3D geometry. The excavation is modelled as a long trench instead of the true box-
shape and the strata are assumed horizontal in the out-of-plane direction. Structural
elements are also heavily inuenced by the plane strain assumption, with the struts in the
original geometry now being modelled as continuous slabs. For this reason, linear struc-
tures, such as struts, ground anchors, piles, nails, etc. are not suited to the plane strain
assumption, as described further in Section 5.1.5. Furthermore, the apparent point load
in the plane strain model actually acts as an innite line load in the out-of-plane direction.

2D axisymmetric assumption
A 2D axisymmetric model also involves the analysis of a plane, vertical section through
the site except that one vertical side of the plane (the left hand side usually) is the axis
about which the site has rotational symmetry. The horizontal axis is the radius from the
axis of symmetry, and the strain perpendicular to the plane and in the circumferential or
hoop direction is assumed to be zero; hence displacement, strain and shear stress can
only occur in the analysis plane. All stresses and strains perpendicular to the plane are
zero except for the normal stress in the hoop direction. This assumption is suited to sites
with a vertical structure in the ground with a uniform, radial cross-section (e.g. vertical
shaft, circular cofferdam, single vertical pile, circular spread foundation) and vertical
loading which is uniform around the central axis, as shown in Figure 1.3. If there are any
other features nearby that are not symmetrical about the axis, these cannot be simulated.
Note that torsional loadings (e.g. to simulate pile boring) also cannot be simulated in an
axisymmetric analysis.

To visualise the geometric assumption of a 2D axisymmetric model, extrude the model


through 3608 about the axis of symmetry. The strata and the ground surface can be hori-
zontal or slope only toward or away from the axis of symmetry. Any structure becomes
circular in plan, centred about the axis of symmetry. Point loads applied in axisymmetric
models are treated as circular line loads centred about the axis of symmetry, while line
loads are treated as distributed loads over areas of circles centred about the axis of
symmetry. Care should be taken when specifying the input parameters for and interpret-
ing the outputs from structures in axisymmetric models (see Section 5.1.5).

1.2.2 How detailed does the geometry need to be?


To save time in setting up and running an analysis, the geometry of the FE model needs
to be as simple as possible but without compromising too much on accuracy. As with
many of the decisions to be taken when setting up a geotechnical FE model, it comes
down to a compromise between detail and efciency. Enough detail is required in order
to obtain reasonably accurate key outputs, but not excessive detail such that the task
becomes unnecessarily time-consuming and expensive.

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 1.2 Visualising the plane strain assumption: (a) true 3D geometry; (b) 2D plane strain
model; (c) 3D geometry assumed by plane strain model

(a)

(b)

Extrusion

(c)

10

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

Figure 1.3 Suitable geometries for the axisymmetric assumption

Circular spread foundation


Axisymmetric section

Axisymmetric section

Axisymmetric section

These decisions are taken using judgement and experience, as well as with the help of test
runs when there is any doubt. For example, regarding geometrical detail, an FE analysis
can be run with and without a particular geometrical detail and then the key outputs
compared to see whether that detail had a signicant effect and needed to be included.

11

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

For example, imagine a stratum boundary at a depth of 9.8 m and the base of an excavation
at 10.0 m depth. To avoid the detail in the mesh required to include the 20 cm difference in
elevation between the stratum boundary and excavation oor, it may be acceptable to move
the stratum boundary to the same depth of 10.0 m in the model, particularly given the
uncertainty in ground conditions, without a signicant effect on the outputs.

Further away from the area of interest, less detail is required. In a typical city centre site,
there will be many features in the surrounding ground such as basements, piled foun-
dations, metro tunnels, etc., as illustrated in Figure 1.4, while even in greeneld sites
there may be topographical or geological features nearby that could potentially inuence
ground behaviour in the area of interest. Again, it comes down to judgement which
features around the area of interest need to be included, and if there is any doubt, try
running the analysis with and without certain features to see if they inuence the key
outputs signicantly and need to be included.

Taking advantage of axes of symmetry can also simplify the geometry signicantly by
allowing half, or even more, of the geometry to be omitted. Axisymmetry allows 3D geo-
metry to be simplied to a 2D plane, as described above, while planes of symmetry may
permit only half or a quarter of the geometry to be modelled in, for example, a rectangu-
lar piled raft, as shown by example in Section 8.2. Similarly, a 2D plane strain model may
be simplied further by omitting half the geometry on one side of a vertical axis of sym-
metry. Bear in mind, however, that not only should the geometry be symmetrical about
the axis or plane of symmetry, the construction methods, timing and ground conditions
must be symmetrical too. If construction on one side of a geometrical plane of symmetry
follows a different sequence or timing to the other, then that should not be considered a
plane of symmetry in the FE model and the different sequences should be fully simulated
in a model of the whole geometry. This is due to non-linearities in ground modelling and
soilstructure interaction that do not follow the principle of superposition.

1.2.3 Where should model boundaries be located?


The FE mesh needs to be xed in space in order to establish equilibrium and solve the
global stiffness equation to determine displacement. The xities are applied at the
boundaries to the model, but in eld problems there is often no obvious boundary for
the FE model because the ground extends indenitely. Therefore, some judgement is
required when deciding where to place the model boundaries. The boundaries should not
be placed too close to the area of interest because that would be unrealistic and introduce
a signicant boundary effect, i.e. the xities imposed at the boundary would start to
inuence the key outputs.

The only common situation in the eld where a model boundary would correctly impose
a signicant boundary effect on the area of interest is where a relatively soft soil overlies
a strong or hard layer (e.g. rock or very dense soil) at shallow depth. The top of the hard
layer could form the bottom boundary to the FE model, as shown in Figure 1.5, pro-
vided that the layer is of substantial extent and deformations in the real layer due to the
imposed loads would be insignicant compared with the deformations in the upper
layers.

12

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

Area of interest
Figure 1.4 Simplifying the geometry around the area of interest

Area of interest

13

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 1.5 Example of a clearly defined boundary

Softer soil

Hard soil or rock undergoing


insignificant strain Hard soil or rock represented by
fixed bottom boundary

In other cases, the boundaries to the FE mesh need to be placed sufciently far away
from the area of interest for the xities to be imposed without causing signicant bound-
ary effects on the area of interest. But how far away should this be? This clearly depends
on the particular characteristics of each model, so there are no concrete rules on model
boundary locations. It is best to experiment with different locations, unless the most
appropriate locations are already known from previous experience of similar problems.
By placing the model boundaries progressively nearer or further from the area of interest
in preliminary analysis runs and plotting the key outputs, it should be possible to identify
a threshold boundary position at which boundary effects are no longer signicant, as
illustrated in Figure 1.6.

Model boundaries within the threshold will result in progressively larger boundary
effects on the area of interest as they are located nearer to the area of interest, but any-
where beyond the threshold should result in insignicant boundary effects on the key
outputs. Therefore, the nal model boundary locations should be set anywhere outside
the threshold position.

Another method to test whether the model boundaries are located sufciently distant
from the area of interest is to change the xities (e.g. add and remove vertical xity at
the vertical boundaries) to see if this affects the key outputs. If no signicant effect is
observed, then the boundaries are located sufciently far away.

Figure 1.6 Threshold boundary location

Threshold boundary
location
Increasingly significant
Area of boundary effects
interest

No significant
boundary effects

Increasingly significant
boundary effects

No significant Generic output in


boundary effects area of interest

14

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

Also, when outputs are checked, the stress state should not be on the failure envelope to
a signicant extent at any model boundary, except perhaps on axes of symmetry. As a
general rule, stress changes should be less than 5% at model boundaries, and ideally less
than 1%.

The bottom boundary can usually be placed closer to the area of interest because the
grounds stiffness and strength increase with depth, and even closer when a strain-
dependent stiffness is adopted in the constitutive model. The vertical boundaries usually
need to be located further from the area of interest, particularly for ground models with
linear stiffness because these tend to exaggerate the deformation further away from the
area of interest.

Figure 1.7 shows some rules of thumb that can be used as a rst-guess for model bound-
ary locations when starting to investigate the most appropriate locations. The distances
shown often need to be increased for sloping ground, undrained behaviour and for
groundwater ow analyses. The appropriate distances for groundwater ow analyses can
be estimated from Sichardts empirical formula (Equation 1.1) providing the approxi-
mate radius of inuence R in metres of a well, as described in, for example, Cashman
and Preene (2012).

R = Cs k (1.1)

where s = drawdown in borehole (m), k = permeability (m/s) and C = 3000 for axisym-
metric conditions and 1500 to 2000 for plane strain conditions.

Analysis models of deep tunnels need not include the ground surface if it is sufciently
remote from the area of interest not to inuence the key outputs. A pressure should be
applied to the top surface of the model to represent the total stress from the overlying
ground. There are similar situations where a small detail of a larger analysis model may
need to be studied, in which case the model boundaries can be located closer to the area
of interest than usual, with the total stresses obtained from the larger model applied at
the boundaries of the smaller model.

1.2.4 What fixities are applied at the model boundaries?


Note that the term boundary conditions refers to all conditions imposed on a model in
order to dene a particular problem (e.g. loads, pore pressures, prescribed displacements,
accelerations, etc.) and not just the conditions at the outer boundaries to the FE mesh.

As mentioned in Section 1.2.3, the model needs to be xed in space. As shown in


Figure 1.7, the standard xities applied at the model boundaries are zero displacement
in all directions at the bottom boundary and zero displacement on the vertical sides in
the horizontal direction perpendicular to those boundaries, including on axes of sym-
metry. The top surface has no xities imposed. Structural elements with rotational
degrees of freedom, e.g. beams and shells (see Section 5.1.1) that extend to vertical
boundaries must also be xed rotationally to simulate the restraint from the structure
beyond the boundary. This is particularly important at axes of symmetry.

15

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 1.7 Appropriate first-guess FE mesh boundary locations

B ~3B

~3B

B Largest of 3B or 2D

~2B

B Largest of 3B or 4H

~2B

~5D

~3D

The standard xities should be used at axes of symmetry, but at remote model bound-
aries located sufciently far from the area of interest, the nature of the xities is less
important. The vertical boundaries could be xed in the vertical direction, for instance,
and the bottom boundary allowed to move freely in the horizontal direction. Indeed,
varying these xities provides a means of checking the sensitivity of the key outputs
to boundary effects.

16

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

1.3. Meshing
1.3.1 Which element types should be used?
Having dened the geometry of the model, this is then replaced by an equivalent FE
mesh, with continuum elements used for the ground. The mesh is formed of elements
according to the degree of precision required in the model (a greater number of smaller
elements gives more precision). The elements are connected together at their nodes. The
nodes are the discrete points where the primary unknowns (displacement or excess pore
pressure) are calculated. Nodal displacements are then interpolated by shape functions
or interpolation functions for all locations in each element to obtain the secondary or
derived quantities of strains or strain rates and, via constitutive relationships, stresses
or stress rates. The stresses and strains are calculated at Gauss, stress or integration
points located across the element.

The hierarchy of element types is shown in Table 1.1. The higher order elements have
more nodes and Gauss points so they produce more accurate calculations of stress, par-
ticularly for stiff behaviour. Linear and cubic strain element types are commonly used in
geotechnical FE analysis. The linear strain elements have fast computation times and are
adequate for typical deformation analyses provided a sufcient number are used, but
they may not be suitable for 2D axisymmetric models and they may over-predict failure
loads in all models (although this tendency is reduced when adopting reduced inte-
gration). To predict failure states and for any axisymmetric models, the cubic strain
elements (e.g. 15-noded triangle) are preferred, in spite of their slower computation
times. In groundwater ow analyses, lower order elements are adequate, or even prefer-
able in some programs.

The advantages of triangular (2D) and tetrahedral (3D) elements over quadrilateral (2D)
and hexahedral/brick (3D) elements are that they t into awkward shapes more easily so
are more suited to automatic mesh generators and they are less susceptible to distortion
errors (see Section 1.3.2).

Rock discontinuities, if modelled explicitly, require interface elements with appropriate


material laws to allow slippage and separation along the discontinuity surface. Interface

Table 1.1 Hierarchy of element types

Shape function Variation across element Example elements for continua

Displacement Strain

1st order Linear Constant TRI3, QUAD4, TET4, HEX8


2nd order Quadratic Linear TRI6, QUAD9, TET10, HEX20
3rd order Cubic Quadratic TRI10, QUAD16
4th order Quartic Cubic TRI15

TRI = triangle, QUAD = quadrilateral, TET = tetrahedron, HEX = hexahedron. Number refers to number of nodes
per element

17

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

elements are also used between structures and the ground for the same reasons. Interface
elements and structural elements (beams, shells, etc.) are covered in Section 5.1.

1.3.2 What makes a good FE mesh?


The size and arrangement of elements in a mesh can have a critical effect on the accuracy
of an FE analysis. A poorly formed mesh is a common source of error, so a lot of atten-
tion needs to be paid to mesh quality and automatic mesh generators cannot be relied on
to produce good meshes on their own.

Essentially, large stress concentrations and zones of rapid stress (including pore pressure)
or strain change need smaller elements. These locations typically occur, for example, at
large stiffness changes, discontinuities, foundation corners and pile bases. A very ne
mesh with small elements everywhere would be the most accurate but this would need
long computation times. A good FE mesh is graded with small elements where they are
needed and larger elements remote from the area of interest and where stresses and
strains are more uniform. Thus faster computation times can be achieved without a
signicant loss of accuracy. Examples of graded meshes are shown in Chapter 8.

To check whether the mesh is adversely affecting outputs, try running the analysis with a
ner mesh and compare the key outputs. If the outputs are essentially the same, then the
mesh is not affecting the outputs. If the outputs are different, then theoretically the ner
mesh is closer to the true solution. Experiment with different meshes to determine the
coarsest and hence most computationally efcient mesh that does not inuence signi-
cantly the key outputs. Meshes formed of higher order elements can be coarser because
of the higher number of nodes per element.

Note that the prediction of collapse loads is heavily inuenced by mesh geometry and
element type, particularly when coarse meshes and lower order elements are used.
Higher order elements should be used and meshes made progressively ner until collapse
loads appear uninuenced by mesh geometry.

Some programs have adaptive mesh renement where, based on the outputs from an
initial mesh, more elements are added automatically where the greatest changes in stress
and strain occur. Subsequent analyses and renement are continued until no further
renement is necessary (refer to Sloan, 2013, for example).

The distributions of displacement and stress calculated by the interpolation functions are
only reliable if the element shapes are not excessively distorted. Where the calculated
variables change rapidly, e.g. at stress concentrations, the distribution is even more
sensitive to element shape. Automatic mesh generators cannot control distortion, so this
needs to be checked manually. Distortion is less of a problem for triangular and tetra-
hedral elements provided that the sides of each element are about the same length.

1.4. Analysis stages


1.4.1 How are the initial stresses set up?
Soil and, to a certain extent, rock are frictional materials so their strength and stiffness
are heavily dependent on internal stresses. In terms of FE modelling, the stressstrain

18

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

behaviour predicted by all non-linear constitutive models depends on the current stress
state.

Unless the entire geological history is simulated, which is rare, analyses of eld problems
require the direct establishment of the initial stresses in the ground. These are usually by
far the largest stresses in the model so they are important and this is one aspect that sets
geotechnical FE analysis apart from other sectors of engineering analysis.

The rst stage of any geotechnical FE analysis involves setting up the initial stresses. For
eld problems, these are the stresses in greeneld conditions, i.e. before any signicant
man-made stress changes occurred, which should be relatively homogeneous across the
model. Then signicant stress changes caused by historical constructions or groundwater
changes are simulated in subsequent stages in order to establish the present day stress
state and recent stress history along the correct stress path.

When simulating laboratory tests, on many occasions the self-weight stresses of the
specimen are insignicant compared with the applied stress throughout the specimen.
In such cases, the specimen can be assumed to have zero density and the initial stresses
set to zero. The stresses applied to the specimen in the real test would then be applied in
the simulated test. For the remainder of this section, eld-type situations where the
initial stresses need to be established will be considered.

Except for cases where undrained conditions are simulated in terms of total stress, the
pore water pressure is clearly an important variable in the setting up of initial effective
stresses. In hydrostatic cases and in relatively simple steady-state ow conditions, the
pore pressures can be specied directly in the input data to the FE analysis. For more
complex groundwater ow conditions, a separate groundwater ow analysis (see
Section 4.3) may be required whose output of pore pressure can form some of the input
for the initial stress in the stressstrain FE analysis. The groundwater level should
coincide with element boundaries in cases where material properties (e.g. saturated and
unsaturated weight density) depend on the materials position above or below the
groundwater level.

The vertical effective stress is relatively straightforward to calculate once the pore
pressure prole and ground densities are known. But the horizontal effective stress, as
calculated from the vertical stress using the stress ratio K0 , is heavily dependent on stress
history, stress path, topography and other geological processes experienced by the
ground. Do not underestimate the importance of this stress. There are two horizontal
directions and only one vertical, so it has the strongest inuence on the overall stress
state. It also has a major inuence on the outputs of some FE analyses, e.g. for retaining
walls, cut slopes and piled foundations. Unfortunately, in situ horizontal stress is dif-
cult to measure accurately (see Section 3.3) and careful judgement is needed before using
measured values.

To estimate K0 , or otherwise to help validate measured values, a number of approximate


equations are available which are given in Appendix 3.1. Note that these equations are

19

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

intended for homogeneous ground with horizontal ground surface and layers and in
greeneld conditions where it is reasonable to assume the same in situ stress in both
horizontal directions. With an inclined ground surface or strata and in the vicinity of
man-made structures, such an assumption should not be made. The equations are also
approximate and empirical, so the inuence of K0 values on the key outputs would need
to be considered carefully in a parametric study. Very high (approaching the passive
limit) and very low (approaching the active limit) K0 should be avoided otherwise the
initial stress in the FE model may be in a state of failure.

There are two methods of establishing initial stress in an FE analysis:

Direct specification (K0 method)


This is intended for homogeneous stress proles with horizontal ground surface, strata
and groundwater levels, otherwise equilibrium may not be obtained since the FE analysis
achieves vertical equilibrium while the horizontal stress is based only on the specied K0
or horizontal stress values. Small equilibrium errors may be acceptable, perhaps due to a
small inclination in the layers or ground surface, in which case a plastic nil-step should be
performed following the establishment of initial stress (a plastic nil-step is an additional
analysis stage, with no change in load, intended to restore equilibrium and allow stresses
to return within failure limits). Note that initial stresses for soil layers simulated as
undrained in terms of total stress should be specied in terms of total stress also and the
K0 value would be the stress ratio for total stresses.

Initial stresses for cases with a sloping ground surface but horizontal strata can still be
established with direct specication. This is performed by having a horizontal ground
surface in the initial stage and then creating the slope by activating or deactivating
elements to create the slope in a subsequent analysis stage.

Gravity switch-on
In cases of non-homogeneous stress proles, such as with sloping strata, the initial stress
is generated by activating the self-weight of the ground and by specifying the initial pore
pressures in the model. This is the same method used for activating new volumes of
ground during subsequent analysis stages (even if direct specication was adopted in the
initial stage). A basic constitutive model, such as the linear elastic perfectly plastic
(LEPP) MohrCoulomb model, can be used in the initial stage with the appropriate
parameters to establish the required stress state before changing to an advanced model
with appropriate material parameters for subsequent stages if necessary. Advanced
models may establish horizontal stresses in a complex way, whereas with LEPP models,
K0 can be manipulated more straightforwardly from the equation (for elastic
conditions):

n
K0 = (1.2)
1n

However, K0 values in excess of 1.0 require a Poissons ratio above 0.5 which is not
possible numerically. In such cases, more of the loading history would need to be

20

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

simulated, experimenting with different Poissons ratios for loading and unloading to
achieve the required stress state. Alternatively, increased self-weight can be imposed in
the rst stage before reducing the ground weight in a subsequent stage in order to create
an over-consolidated state. All gravity switch-on stages must be performed with drained
conditions, even when undrained conditions will be simulated in subsequent stages.

It is important to check the outputs from the initial stress analysis stage to ascertain
whether the initial stresses have been established correctly. Also, the model must be in
equilibrium with an error of less than 1% and no signicant plastic yielding. Where there
is a small equilibrium error or a few Gauss points yielding, performing a plastic nil-step
should restore equilibrium and return all stresses within failure limits.

As well as establishing the initial stress state, advanced constitutive models require
certain state parameters that dene, for instance, the initial location and size of the yield
surface. Examples include the initial void ratio and pre-consolidation stress. The former
should be relatively straightforward to measure while the latter often requires a degree of
interpretation of test data (see Section 3.4.1).

Application of the gravity switch-on method to establish the initial stress state is shown
in the example in Section 8.4.

1.4.2 How are the construction stages set up?


Any geotechnical FE analysis of new or existing structures must consider explicitly how
the structures were constructed because this affects stress paths and ground behaviour.
The time periods for construction are also important for temporal effects such as
groundwater ow and excess pore pressure dissipation in low-permeability soils or creep
effects. Furthermore, if outputs only for permanent works were required, the temporary
works stages taken to get there cannot be ignored because of non-linear effects. The
principle of superposition cannot be applied in geotechnical FE analyses.

Construction activities can be complex, with many processes occurring simultaneously


and in different phases across the site. Rather like the creation of the analysis geometry,
it is not possible to simulate every detail of the construction activities. For example, the
placement of each 0.3 m-thick layer of ll in the embankment construction example in
Section 8.4 was not simulated. It was found that 2 m-thick layers could be installed with-
out a signicant loss of accuracy. Judgement is needed to identify the essential elements
of the construction activities that need to be included in the analysis model and which are
likely to have a signicant effect on the key outputs. The most reliable way to test
whether a feature of the construction activities needs to be included in the model is to
run the analysis with and without the feature included, and to check whether the key
outputs change signicantly. If the key outputs are not affected, then the unnecessary
detail could be omitted so that the analysis could run more efciently. If the key outputs
are affected, then that feature would need to be included in the analysis model.

One of the most common activities in construction stages is the deactivation (to simulate
excavation of ground or removal of structural components) and activation (to simulate

21

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

placement of ll or installation of structural components) of elements. All elements must


be present in the FE mesh but may be deactivated at the start where necessary.

On deactivating elements in a construction stage, immediately their material properties


are ignored, stresses and nodal displacements are set to zero, any model boundaries
formed by the deactivated elements become free and permeable and any external loads
applied to those elements are ignored. However, removal of the weight of the deactivated
elements can cause a large out-of-balance force so the equivalent weight of the deacti-
vated elements is applied automatically by most software at the newly exposed ground
surface and reduced in a stepwise fashion in non-linear analyses. Some simpler software,
particularly non-geotechnical-focused programs, may just assume a very low stiffness for
inactive elements, but this could lead to inaccurate predictions.

On activating elements in a construction stage, the material properties of these elements


are taken into account from the start of the construction stage, while the weight of the
activated elements is introduced in a stepwise fashion due to the large out-of-balance
force and the activated element stresses begin to grow. The new nodes also immediately
become active but their initial displacement is set to match the already deformed mesh to
which they are being added. The disadvantage of this is that a false impression of the
deection of newly placed layers can be formed. The deection of nodes occurring prior
to activation of the element needs to be subtracted from the output in order to obtain the
deection of the elements since activation, as demonstrated in the embankment con-
struction example in Section 8.4.4.

The activation of ground elements is the same as the gravity switch-on procedure
described in Section 1.4.1 except that a specic area or volume is activated during the
analysis rather than the whole ground mass at the rst analysis stage. So, in the same
way, it is often necessary to use a different constitutive model or parameters for the
ground during element activation to simulate behaviour during construction and obtain
an appropriate stress state. For example, Poissons ratio may be manipulated with an
LEPP model to obtain an elevated K0 value, perhaps resulting from compaction, or
different drainage conditions may be used for low-permeability lls during construction.
In a subsequent construction stage, the constitutive model may be changed to simulate
the post-placement behaviour more accurately.

Ground improvement
The actual processes of ground improvement, such as compaction and treatment, are too
complex to simulate in routine FE analyses. Therefore, the strength, stiffness and density
properties are usually changed in a construction stage to reect the improvement of the
ground, with the new input parameters obtained from site trials or previous experience
on similar sites.

Installation effects
The term installation effects refers to the effect on the ground around new structures
such as piles and diaphragm walls during their installation. If installation effects are

22

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

likely to inuence key outputs from an FE analysis, then they must be considered in
some way. Unfortunately, simulating installation effects can be complex and, at best,
probably only provides qualitative assessments of installation effects rather than accu-
rate predictions.

In some cases the installation effects may be less signicant, for example during CFA
(continuous ight auger) or cased bored piling, where the support provided should
prevent signicant stress and strain changes in the soil. In other cases, installation effects
may be more signicant. Most commonly, structures are introduced in the ground in FE
analyses as wished in place. This means that line or surface elements are simply acti-
vated in the FE mesh, while area (in 2D meshes) or volume (in 3D meshes) elements have
their material models changed from a soil or rock material to the new structural material.
At the same time, it is possible to alter the properties of the surrounding ground to
attempt to take account of installation effects. Alternatively, additional construction
stages can be added to try to simulate the installation process. Some examples of these
techniques are described in the following:

g Fluid support to pile bores and diaphragm wall excavations: uid provides less
support than casing and results in more ground deformation and stress relief.
Consequently, the wished in place option is likely to be conservative in terms of
earth pressures applied to the structure but not so in terms of ground
deformation. For an approximate assessment of installation effects, uid
support can be simulated with its hydrostatic pressure applied to soil surfaces in
contact with the uid. Fluid concrete can be simulated in a similar way, but
note that pressures increase hydrostatically initially but at a critical depth a
maximum is reached below which the concrete uid pressure stays
approximately constant. This is due to increases in effective stress as its
behaviour changes from a uid (aggregates in suspension) to a granular medium
(aggregates in contact) with pore pressure. The critical depth depends on many
characteristics of the concrete, as described by Clear and Harrison (1985).
Pouring concrete under water or a support uid further complicates the pressure
distribution because it is dependent on both the effective concrete weight and
the initial uid pressure (refer to Lings et al., 1994). Three dimensional analysis
is required for diaphragm wall installation simulations due to the complex stress
redistributions during construction. The approximate nature of these
assessments does not usually warrant their inclusion in the main FE model of
the entire construction sequence. It is usually more appropriate to conduct a
separate, detailed study of installation effects in order to assess the approximate
error in the outputs resulting from adopting the wished in place option in the
main analysis model.
g Driven/displacement piles: the installation effects of displacement piles depend on
the in situ ground density, geological history, installation method and any
installation aids, and include settlement or heave and changes to ground density.
The installation process is too complex to be simulated in routine FE analyses, so,
on pile or wall activation, the adjacent soil parameters need to be modied
appropriately (e.g. Engin et al., 2015).

23

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

g Grouting: compaction and compensation grouting cause displacement of soil, but


in a less dynamic way than for driven piles, so the grouting can be simulated by
widening elements with careful validation using eld tests and monitoring
(see Section 5.1.4).
g Casting thick concrete slabs: some raft foundations and basement slabs have a large
thickness and therefore a signicant self-weight. On casting, the uid concrete
self-weight interacts with the ground in a different way to hardened concrete since it
applies its self-weight without any stiffness so is more inclined to sag. Therefore, in
some situations it may be necessary to consider the installation effects of a thick
concrete slab by applying its self-weight only without structure before substituting
the load for the structural elements of the slab (with self-weight included) in a
subsequent stage to simulate the hardened concrete slabs weight and stiffness.

Further examples of the simulation of installation effects can be found in Hicks et al.
(2013).

1.4.3 Which calculation options should be chosen?


Linear elastic FE analysis is more straightforward and computationally simple than non-
linear analysis. Unfortunately, such analyses are inadequate for geotechnical problems,
except perhaps for intact rock. Geotechnical modelling generally requires the introduc-
tion of plasticity, non-linear elasticity, frictional contact, large displacements or creep, or
a combination of them. Each of these introduces non-linearities to an FE analysis which
require more complex solution methods. In particular, applied loads or displacements
must be divided into increments or steps and equations solved iteratively, ensuring that
equilibrium is satised before moving on to the next iteration or load step to prevent the
solution drifting from the correct equilibrium value. This lengthens computation times
and increases the probability of failing to obtain a nal convergent solution or, worse
still, obtaining an inaccurate solution. Therefore, it is important to exercise caution and
engineering judgement when interpreting outputs from non-linear FE analyses.

This section covers a few issues that are common to most programs, but always consult
the software manuals carefully to learn about the calculation options available and seek
guidance on appropriate selections.

Step size
Non-linear equations need to be solved in calculation steps, but what step size should be
used? Too small and many steps will be required leading to a slow solution. Too big and
a high number of iterations will be required, also leading to a slow solution or no
solution at all. Therefore, the right balance needs to be found for step size and many
programs determine the step size automatically. If the step size is set manually,
experimentation will be required to nd the most efcient step size.

Solution scheme
A widely used method of solving non-linear equations is the NewtonRaphson iterative
method. It establishes the loaddisplacement curve for every degree of freedom using an
initial guess or trial solution for each load increment based on the slope of the curve. It

24

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

calculates the out-of-balance or residual force vector which is the difference between the
external load increment and the corresponding resisting internal force computed using
the straindisplacement and stressstrain relationships. If the residual force exceeds a
particular tolerance, then the displacement is corrected back to the equilibrium solution
and the process repeated successively until the residual force is within the tolerance. The
trial solution must be reasonably close to the true solution in order to achieve conver-
gence, and the slope of the loaddisplacement curve should not change sign. At or near
maxima or minima where the slope changes sign, the arc length method (or Riks method)
can be used to obtain more reliable solutions. There are various stress point algorithms
used to integrate the constitutive equations to obtain stress change (and hence internal
force) and each must make additional assumptions since the constitutive behaviour is
changing in each increment. Users should verify that their software uses an appropriate
stress point algorithm in their FE analysis. The NewtonRaphson method uses the
current slope of the loaddisplacement curve, which is the (tangent) stiffness matrix, at
every iteration. In large FE meshes, calculation of the stiffness matrix is computationally
demanding and can slow down the calculation. An alternative is the Modied Newton
Raphson method where the same slope is used in successive iterations although conver-
gence will be slower the overall calculation may be faster, provided the behaviour is not
overly non-linear, since the stiffness matrix is not re-calculated for every iteration.

Equilibrium error
At the end of each load increment, an equilibrium check is performed by converting the
externally applied loads and internal stresses into equivalent nodal loads and calculating
the difference or out-of-balance load between the external and internal values at each
node. The maximum difference expressed as a ratio or percentage of the out-of-balance
load to external load is termed the maximum equilibrium error. It should be less than
1% preferably, particularly at the initial stage, but denitely less than 5%. While achiev-
ing a low value is a requirement, it eliminates only one of the many potential sources of
error in a non-linear FE analysis so should not be viewed as a guarantee of accuracy.
Increasing the allowable value in the calculation options above this level to achieve con-
vergence merely obtains a false equilibrium and certainly an inaccurate result. Do not be
tempted to do this.

Large deformations
In conventional small deformation analysis, the external loads and internal stresses are
assumed in equilibrium in the original mesh geometry (which is called Total Lagrangian
formulation). So, while nodal deections are calculated, the actual coordinates of the
nodes in the calculations do not change. This is a good approximation for small defor-
mations in most cases, but in some cases the changing geometry of the mesh needs to be
taken into account in the calculation.

Such cases include the analysis of soil reinforcement (e.g. geotextile) where membrane
action can help support loads perpendicular to the reinforcement plane at large strains.
If only the original mesh geometry is considered, then the tension mobilised in the
reinforcement would only act in the original horizontal orientation of the reinforcement
whereas, in reality, the reinforcement bends and so its tension develops a vertical

25

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

component to help support the load. There are also more general cases where large
deformations need to be predicted accurately, often in soft soils, as shown in the
embankment construction example in Section 8.4. Note that for the prediction of
collapse loads (and not for the prediction of collapse deformations), usually the conven-
tional method agrees better with analytical approaches. Using large deformation
methods can lead to stiffer behaviour near failure and a higher predicted failure load.

There are two common options to account for large deformations in the mesh geometry:

g Updated coordinates (only): here the nodal coordinates are updated to match the
calculated deection and equilibrium is satised in the nal deformed geometry.
But this is not a rigorous treatment of large deformation behaviour because no
account is taken of stress changes due to rotation and straining.
g Updated Lagrangian formulation: here, as well as updating coordinates, the stress
changes due to rotation and straining are taken into account.

These methods are slower and less robust, so only use them when necessary. Perform a
conventional analysis rst and check whether deformations are large enough possibly to
require large deformation methods. If so, try using these methods and check whether the
key outputs have changed enough to justify their use.

Groundwater pressures (for pore pressure at Gauss points and external water pressures)
may also be updated due to the changing geometry of the ground under a constant
groundwater level. Check how the software handles distributed loads usually the equiv-
alent nodal loads remain unchanged in spite of the changing nodal coordinates, and the
loads will either follow the initial direction or rotate with the deformed shape.

Note that even these large deformation methods have their limits. Other methods,
beyond the scope of this book, are under development to predict larger deformations
and material ow, including a number of Eulerian methods, such as the material point
method, usefully summarised by Soga et al. (2016).

1.5. Constitutive models


1.5.1 Which constitutive model should be used?
Constitutive models should be selected that simulate each soil or rock stratums behaviour
with sufcient accuracy under all the loading conditions to be imposed. To avoid unnecess-
ary complexity, the simplest constitutive model that satises this requirement should be
selected. Consequently, some compromise is needed and a complex model that recreates all
aspects of ground behaviour may not be necessary. Identify which regions of the FE model
are of greatest interest (where more precise constitutive modelling may be needed) and which
aspects of ground behaviour are the most critical. Then select the constitutive model accord-
ingly. Note that the same soil or rock with different structures (e.g. an embankment or an
excavation) may behave very differently because behaviour is stress-path dependent. So, one
model cannot be said to be suitable for a particular soil or rock in all situations.

More guidance on constitutive models is provided in Chapter 2.

26

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
How is a geotechnical finite element analysis set up?

1.6. Groundwater and drainage


1.6.1 How are the effects of groundwater included in the analysis?
Groundwater is very important in any FE model because it has a direct inuence on
effective stress, time-dependent soil response and the forces acting on structures (e.g.
retaining walls). Sometimes groundwater ow can be neglected and hydrostatic con-
ditions assumed, or the steady-state pore pressure distribution is simple enough to be
specied directly. On other occasions, a groundwater ow calculation is required to
generate the steady-state or transient pore pressure distributions in the ground, either
in a separate analysis or fully coupled with the stressstrain calculations. Including
groundwater effects in an FE model is covered in Sections 4.1 and 4.3.

Submerged surfaces require the application of perpendicular, external water pressures


corresponding with the water level. Some programs create these pressures automatically
once the external water level has been specied. If excavating or lling under water, the
external pressures need to be changed in the same construction stage.

1.6.2 Should a drained, undrained or consolidation analysis be


performed?
The dissipation of excess pore pressures is a time-dependent phenomenon requiring
equations of consolidation (usually Biots equations) for its simulation in a consolidation
analysis. Alternatively, it may be acceptable to simplify the analysis of a soil layer to
wholly drained (when the rate of loading is slower than the rate of drainage) or wholly
undrained (for short-term periods during which no signicant dissipation of excess pore
pressure has occurred). A consolidation analysis is required to dissipate excess pore
pressure either to change from undrained to drained conditions or to obtain temporal
outputs of deformations and structural forces, etc. during consolidation.

Drained, undrained and consolidation analyses are described in more detail in Chapter 4
(Sections 4.2 and 4.4), in particular the issues associated with modelling undrained
behaviour.

REFERENCES
Cashman PM and Preene M (2012) Groundwater Lowering in Construction, A Practical
Guide to Dewatering, 2nd edn. CRC Press, Boca Raton, FL.
Chillery M (2014) NAFEMS Simulation Handbook Quality Management. NAFEMS,
Hamilton.
Clear CA and Harrison RA (1985) Concrete Pressure on Formwork. CIRIA Report R108.
CIRIA, London, UK.
Engin HK, Brinkgreve RBJ and Van Tol AF (2015) Approximation of pile installation
effects: a practical tool. Proceedings of the Institution of Civil Engineers Geotechnical
Engineering 168(4): 319334.
Hicks MA, Dijkstra J, Lloret-Cabot M and Karstunen M (2013) Installation Effects in
Geotechnical Engineering. CRC Press, Leiden, Netherlands.
Lings ML, Ng CWW and Nash DFT (1994) The lateral pressure of wet concrete in
diaphragm wall panels cast under bentonite. Proceedings of the Institution of Civil
Engineers Geotechnical Engineering 107(3): 163172.

27

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Sloan SW (2013) Geotechnical stability analysis. Geotechnique 63(7): 531571.


Soga K, Alonso E, Yerro A, Kumar K and Bandara S (2016) Trends in large-deformation
analysis of landslide mass movements with particular emphasis on the material point
method. Geotechnique 66(3): 248273.

28

Downloaded by [ University College London] on [14/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.029

Chapter 2
How are constitutive models selected?

2.1. Introduction
2.1.1 What is a constitutive model?
A constitutive model is a series of mathematical expressions relating stresses and strains
(or stress rates and strain rates) that are used to model material behaviour in an element.
When implemented into an FE analysis that also ensures equilibrium and compatibility
between each element, the constitutive model allows complex problems to be analysed
and displacement and stress to be calculated everywhere in the model at every construc-
tion stage.

All constitutive models are an approximation of material behaviour. Advanced models


may recreate several aspects of material behaviour but none can recreate all aspects
and knowing what they do not model is just as important as knowing what they do
model.

Many constitutive models have been proposed and published in the literature to recreate
particular aspects of soil and rock behaviour. Most are rarely used. This is often due to
their complexity and the high number of input parameters, some of which may be hard
to obtain. Furthermore, to enter practical application in FE analyses, a constitutive
model needs to work in general stress space whereas during development the model may
have been tested on only limited stress space, stress range and stress paths. Implemen-
tation of constitutive models in FE analysis is not straightforward and requires rigorous
testing to help ensure that the implementation is robust.

This leaves a relatively small number of constitutive models that have been implemented
successfully and applied widely in simulations of actual construction projects. Yet, even
the selection of appropriate models from this short list is not easy. This chapter provides
the background knowledge required to make informed selections of appropriate consti-
tutive models for a particular analysis task.

2.1.2 Why is it important to use appropriate constitutive models?


While FE analysis has allowed the simulation of complex ground behaviour and soil
structure interaction problems, its accuracy depends heavily on the constitutive models
adopted for each material. The constitutive model needs to account for all the important
aspects of material behaviour for a particular problem otherwise the outputs from the
FE analysis will be erroneous.

29

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Soils are usually the softest and weakest material in any soilstructure interaction analy-
sis, so their behaviour governs the deformations and probability of failure. Therefore, it
is important to simulate their behaviour accurately over the range of stresses and strains
they will experience with an appropriate constitutive model. Construction materials,
such as concrete and steel, are stiff in comparison with soils so it is often sufcient to
model these with simple linear elastic constitutive models (see Section 5.2).

2.1.3 How is the appropriateness of a constitutive model judged?


An appropriate constitutive model is one that recreates the important aspects of stress
strain behaviour for the range of stress and strain conditions in the problem to be
analysed. At the same time, it should be economical, i.e. not include other aspects of
material behaviour that are not important, to avoid unnecessary complexity. Further-
more, the model parameters need to be obtainable from the tests that can be (or were)
performed in the site investigation.

The selection of appropriate constitutive models depends primarily on three aspects of


the project:

1 Aims of the FE analysis and required outputs: above all, a constitutive model
needs to be selected that will provide the required outputs accurately, but there is
no need for the model to provide accurate outputs that are not required. For
instance, an FE analysis of a deep excavation may be required to obtain the
structural forces in a retaining wall for its design. In this case, a linear stiffness
varying with depth or a stress-dependent stiffness may be sufcient to model the
elastic behaviour of the soil. There would be no need, in this case, to use a strain-
dependent stiffness because the additional complexity would probably result in
only a marginal increase in the accuracy of the required outputs of retaining wall
forces. On the other hand, if outputs of excavation-induced settlement were also
required from the FE analysis, then a constitutive model with strain-dependent
stiffness would be required in order to obtain that output accurately.
2 Structure type and expected stress path: the type of structure being simulated
inuences the stress path in the ground around the structure. Broadly speaking,
foundations and embankments cause increased loading in the ground while
excavations and tunnelling cause unloading. Of course, many projects will be
more complicated than this with combinations of structure types and a
construction sequence that may result in load reversals (e.g. demolition and
reconstruction of a building) rather than just monotonic loading. The expected
stress path is important because some constitutive models are more suited to
particular stress paths than others. For instance, the Modied Cam Clay (MCC)
model provides realistic predictions of deformation for the compression of soft
clays, but less so for unloading. More guidance on the inuence of structure type
on constitutive model selection is provided in Section 2.4.
3 Soil and rock types: it would be wrong to say that a constitutive model is suited
to a particular soil type in all applications because there are required outputs and
expected stress paths to consider. If a constitutive model was used successfully to
simulate a particular soil on one project, do not expect the same model to

30

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

simulate the same soil successfully on another project with a different stress path
and/or different required outputs. Yet, clearly, the soil or rock type is an
important consideration. Certain soil and rock types will exhibit behaviour less
prevalent in other types, and if this behaviour inuences the required outputs then
a constitutive model that recreates this behaviour needs to be selected. For
instance, soft clays and silts often experience time-dependent creep or stress
relaxation not seen in other soils, while other soils may exhibit particularly
anisotropic properties that could inuence the required outputs. The different
aspects of ground behaviour that need to be considered are covered in Section 2.2.

Further project-specic guidance on the selection of appropriate constitutive models can


be found in published case studies of similar projects in similar ground conditions using
FE analysis where justication of the selection of constitutive models may be provided.

2.2. Aspects of ground behaviour


2.2.1 Which aspects of soil behaviour may need to be considered?
Soil is a complex material consisting of a skeleton of soil grains in frictional contact and
voids lled with air and/or water (or other uids). Forces are transmitted through the soil
skeleton via normal and shear forces at grain contact points. This is a behaviour actually
more suited to discrete element modelling where each contact is modelled. However, due
to limitations on the size of discrete element models and the time taken to set up and run
such analyses, most practical problems are modelled using the principles of continuum
mechanics with, for example, FE analysis. This leaves the engineer with the challenge of
characterising a mixture of solid grains, water and/or air as a continuum. For this reason,
compared with other engineering materials, soil is among the most difcult to model.

The following are some of the aspects of soil behaviour that need to be considered when
selecting a constitutive model:

Soil type
Soft clay and dense gravel, for example, each display very different behaviours with wide
ranges of strength and stiffness. Many constitutive models are more suited to either ne-
grained or granular soils. Each soil type will respond in different ways to most of the
further aspects of soil behaviour that follow.

No tensile strength
Soils have little or no tensile strength, so constitutive models need to include this impor-
tant aspect of behaviour.

Strength changes during shear


Loose soils compress while dense soils dilate during shear, both toward the critical state,
and these changes in density cause changes in the current shear strength of the soil, i.e.
shear hardening in the loose soil and shear softening in the dense soil. Continued shear
can lead to a further reduction in strength toward a residual shear strength. Softening is a
particular issue with stiff plastic clays which are rather brittle and progressive failure is a
common phenomenon in slopes in such clays.

31

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Stress-dependency of stiffness and strength


As the conning stress on a soil increases, the volume of voids decreases, the soil densies
and so it becomes stronger and stiffer. The stress-dependency is non-linear, although a
linear relationship is often assumed particularly for soil shear strength. It is possible to
specify the changing stiffness and strength with depth in the model parameters to take
account of increasing in situ stress, while more advanced models take account of stress
changes during the analysis, which is important when the stress changes are large, e.g. in
embankment construction or excavations.

Stress-path dependency of stiffness


On primary loading (i.e. when loaded to a particular level for the rst time), soil shows a
highly non-linear stiffness (Figure 2.1). When unloaded or reloaded it shows a higher,
more linear stiffness. When reloading changes to primary loading as the previous
maximum stress is exceeded, then there will be a sudden reduction in stiffness. Therefore,
stress history as well as stress path is important. Furthermore, primary loading behav-
iour differs signicantly between deviatoric and compressive loading, as illustrated by
comparing the typical graphs from triaxial and oedometer tests in Figure 2.1. Whereas
stiffness decreases with deviatoric load, it increases under compressive load due to the
increasing density of the soil.

Permanent deformations
Many materials have a signicant elastic stress range within which reversible, elastic
deformations occur. Soils, however, have a narrow elastic stress range such that, even
at quite low stress levels well below the failure stress, permanent deformations occur,
particularly in soft normally consolidated and lightly over-consolidated clays.

Bonding and structure


Natural soils develop a fabric and inter-particle bonding called structure that gives soil
additional strength and stiffness that cannot be explained by void ratio and stress history
alone. Signicant straining causes loss of structure (destructuration) and a change in
behaviour. Yet, many constitutive models are based on the results of laboratory tests

Figure 2.1 Typical deformation behaviour of soils


Deviatoric stress, q

Mean effective stress, p

Primary loading
stress path Primary loading
stress path
Unload/reload
Unload/reload stress path
stress path

Axial strain, a Axial strain, a


Triaxial compression Oedometric (K0) compression

32

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

on reconstituted samples without structure and so do not include the effects of


destructuration.

Intermediate principal stress


Many constitutive models are based on the results of laboratory triaxial tests where the
intermediate principal stress s2 equals either the minor principal stress s3 (compression
tests) or the major principal stress s1 (extension tests). However, when implemented into
FE analyses, constitutive models operate in general stress space and s2 may vary between
the values of s3 and s1 stress states that may not have been tested in the original con-
stitutive model. This variation is dened by the ratio b where b = 0 corresponds with
s2 = s3 and b = 1 with s2 = s1 (as shown in Figure 2.2).

Anisotropy
Most soils are anisotropic to some extent. Assuming isotropy can over- or under-estimate
the strength and stiffness of the soil, it is not necessarily a conservative assumption.
Fabric and stress history can give an element of (inherent) anisotropy to the strength and
stiffness of soils, while stress and strain changes can increase the anisotropy (induced) or
reduce it. Soil properties generally do not vary in the plane of deposition (which is often
horizontal, but not always) but only between the plane of deposition and direction of
deposition (often vertical). Consequently, cross-anisotropic (also called transverse aniso-
tropic and orthotropic) conditions can usually be assumed for soils. Anisotropy is
expressed in terms of the angle a between the major principal stress direction and the
direction of soil deposition, as shown in Figure 2.3.

Strain-dependent stiffness
In addition to stress-dependency, soil stiffness is also strain-dependent. At small strains,
soil stiffness is high and it decays to a lower value as strains increase, as illustrated in
Figure 2.4. Note that the rate of decay is particularly high in the typical ranges of strain
occurring around geotechnical structures, so it is highly relevant to the modelling of
most structures. The stiffness also returns to higher small-strain values on stress reversal
and, to a lesser extent, after stress rotation before decaying again with increasing strain.

Figure 2.2 Example principal stress orientations

b=0 b=1 0b1


1 3
1
3 2 1 2 1 3 3

2 = 3 2 = 1 2 3
2 3 2 1 2 1 1
3

1 3 1 > 2 > 3

Triaxial Triaxial Typical application in


compression extension general stress space

33

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 2.3 Expressing anisotropy of soils

= 0 = 90
1 3

90 1 3
1 1 bedding
3 3 1 1 3 3 planes
bedding 3
1
planes active state passive state
= 0 (when = 90 (when
1 3 bedding planes bedding planes
are horizontal) are horizontal)
Triaxial Triaxial Spread foundation
compression extension

When using linear elasticity, a lot of care is needed to ensure that the stiffness value
selected is appropriate for the strain level.

Creep
Creep (or secondary compression) is deformation that continues even after excess pore
pressures have dissipated and under constant pore pressure and effective stress. Alterna-
tively, if deformations are constrained, stress relaxation will occur over time. It is a major
contributor to the deformation of soft clays, silts and peat.

No constitutive model can take account of all of these aspects of soil behaviour. Even if
one could, it would probably be too complex to implement into an FE analysis and the
parameters would be too difcult to determine. Many of these aspects of behaviour are
still to be researched in detail and understood before accurate constitutive models can be
produced and rigorously tested. Nevertheless, continuous developments are leading to
more unied models that incorporate more of the aspects described above.

Figure 2.4 Typical decay of soil stiffness with strain (redrawn from Mair (1993))
Q1

Typical strain range


around geotechnical
structures
Linear elastic
Stiffness

LEPP model

Perfectly plastic

0.0001 0.001 0.01 0.1 1 10


Strain: %

34

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

2.2.2 Which aspects of rock behaviour may need to be considered?


While the aspects of behaviour in soft rocks and hard soils, e.g. mudstone and hard clay,
are quite similar, in harder rocks the behaviour becomes rather different and constitutive
models intended specically for rocks are usually more appropriate.

The following are some of the aspects of rock behaviour that need to be considered when
selecting a constitutive model:

g Stress-dependency of strength: highly jointed or weathered rock is a frictional


material and the stress-dependency of shear strength is signicant, and is also
highly non-linear. Stiffness, on the other hand, is not very stress-dependent and a
constant stiffness can usually be assumed.
g Inuence of discontinuities: discontinuities are surfaces in a rock mass where there
is a sudden change in physical and chemical characteristics. In this book,
discontinuities will be considered as the physical type where relative movement can
occur between the rock on each side of the discontinuity surface and they include
joints, fractures, cleavage and faults. They govern rock mass behaviour in low-
stress conditions because they are signicantly weaker than the intact rock. They
introduce strength and stiffness anisotropy and pre-determined failure planes.
Somehow the mechanical properties, spacing, orientation and persistence of the
discontinuities need to be incorporated into a constitutive model of a continuum.
g Strength changes during failure: the stressstrain response of rock shows a
relatively linear elastic response initially, compared with soils, until failure occurs.
Rock failure can be brittle or ductile depending on the stress state, as dened by
the empirical Mogi line (Mogi, 1971). When the ratio of major to minor principal
stress exceeds 3.4, brittle failure is predicted. Brittle failure in rock results in strain
softening to a residual strength that is much lower than the peak strength.
Progressive failure mechanisms can also form.
g Tensile strength: rock possesses tensile strength, but not at discontinuities, which
requires specic treatment in constitutive models for rocks.

2.2.3 When can FE analysis be used for rocks?


Fractured rock is a complex assembly of intact blocks between discontinuities. Rather
like soil, forces are transmitted through the ground via normal and shear forces at
contact points, but behaviour can be more complex than soil due to the effect of the size
of the intact blocks relative to the structure being analysed. At some point it becomes
necessary to model each block explicitly because their form starts to dominate engineer-
ing behaviour and failure mechanisms.

The advantage of using FE analysis is that it is quicker and more practical than model-
ling with discontinuous media. It is also easier to compare analyses with different discon-
tinuity patterns and properties. Yet, on some occasions it will be necessary to model a
rock mass as a discontinuous medium using, for example, discrete element modelling.

Here are three different options for modelling rock masses, two as a continuum and one
as a discontinuous medium, together with descriptions of when it is appropriate to adopt

35

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

each option, as illustrated in Figure 2.5:

g As a continuum, e.g. FE analysis, with implicit modelling of discontinuities


(when present) by smearing them into a continuum with appropriately reduced
strength and stiffness and, perhaps, anisotropy. This is appropriate when the rock
mass is essentially free of discontinuities or when the discontinuity spacing is
small relative to the structure being modelled and there is no unfavourable dip in
the discontinuity that could lead to instability.
g As a continuum, e.g. FE analysis, with explicit modelling of discontinuities using
interface elements. Interface elements have zero tensile strength, a compressive
stiffness and a shear stiffness subject to a Coulomb friction criterion and are
described in more detail in Section 5.1.3. This method is appropriate when the
discontinuity spacing is similar to the size of the structure being modelled,
provided that the discontinuity geometry remains unchanged during the
construction process since nodal coordinates remain unchanged during FE
analysis. The geometry of the discontinuities can be created individually or, since
that level of detail is not usually available from geological eld mapping, using
different fracture network models to investigate the effect of different
discontinuity patterns, spacings and orientations. In continuum modelling, the
onset of failure is predicted but not the complete separation or rotation of blocks.
g As a discontinuous medium, e.g. discrete element method, where complete
separation and rotation of individual blocks are possible. This method is also
appropriate when the discontinuity spacing is similar to the size of the structure
being modelled, but also when large deformations are expected or when the
contact points between blocks are expected to change during the analysis due to
slippage, rotation or separation of blocks. This method is particularly suited to
rock slope stability problems. Relatively well-dened discontinuity patterns are
needed to model the locations of the discontinuities.

Hybrid numerical methods combining, for example, discrete element modelling with FE
analysis may be used to combine the advantages of each method.

For borderline cases, the quality of the eld data plays a role in the choice of modelling
method. If no discontinuity pattern emerges and spacings and orientations seem arbi-
trary, then continuum modelling is better because it is easier to study different discon-
tinuity patterns.

Remember that bedrock situated below a softer soil and the area of interest can be simu-
lated as a xed bottom boundary if strains in the bedrock are expected to be insignicant
(see Section 1.2.3).

2.3. Common constitutive model types


2.3.1 How do constitutive models commonly account for the elastic
behaviour of soils?
Rarely is an elastic constitutive model on its own sufcient to model soil behaviour
because soil is a comparatively weak material and irreversible strains, shear-dilatancy

36

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

Figure 2.5 Methods of modelling rock masses depending on discontinuity spacing

Essentially free of
discontinuities
Model as a continuum

Slope stability analysis may


require discontinuous medium
due to large deflections
Discontinuity spacing large
relative to structure
Model as continuum with
interface elements or as a
discontinuous medium

May require explicit modelling of


discontinuities due to unfavourable dip

Discontinuity spacing small


relative to structure
Model as continuum with
implicit discontinuities

37

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

and other behaviours associated with being near or beyond failure are not recreated by
an elastic model. Perhaps only very stiff clays at low stress levels could be simulated with
an elastic model. Therefore, soil constitutive models need both elastic and plastic behav-
iour to produce sufciently accurate predictions.

This section provides a summary of the ways that the elastic part of elastoplastic consti-
tutive models work. For more detail on specic constitutive models, readers should refer
to their software manuals and the reference papers associated with each model.

Linear elasticity
Isotropic linear elasticity, also called Hookes law, is the most basic way to treat the
elastic part of soil behaviour. It requires only two parameters (Youngs modulus E and
Poissons ratio n). Real soil stiffness is stress- and strain-dependent, so the parameters
selected for a linear elastic model must be appropriate for the expected stress and strain
level and stress path (e.g. primary loading or unloadreload path). It may not be possible
to select an appropriate value for large ranges of stress and strain (see Figure 2.6) or for
complex stress paths. An increasing stiffness with depth (a linear increase specied in the
input parameters or else separate soil layers with different stiffness) can be specied to
account for increasing in situ stress with depth, but subsequent stiffness changes due
to stress changes are not be taken into account.

Anisotropic linear elasticity


Cross-anisotropic stiffness can be added to the linear strain elastic model with ve input
parameters instead of two for the isotropic model (Clayton, 2011). The more advanced
non-linear elasticity models described below usually adopt isotropic elasticity because of
the high number of parameters that would be required to dene non-linear, anisotropic
behaviour. However, some non-linear elastic, anisotropic models are used in research
and it is possible that they will see more common use in practical problems in the future.

Non-linear elasticity and stress-path dependent stiffness


To take account of the non-linear stiffness of soils on primary loading, hyperbolic func-
tions are commonly used to t observed stressstrain curves in triaxial compression tests

Figure 2.6 Linear and non-linear elasticity

Linear elasticity
Hyperbolic function
Deviatoric stress, q

for primary loading

Real soil behaviour

Unload/reload
stiffness

Axial strain, a

38

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

up to a failure stress plateau (Figure 2.6). This was rst proposed by Kondner (1963) and
is used in the Duncan and Chang (1970) and Hardening Soil (Shanz et al., 1999) models.
On unloading, the soil does not return along the same stressstrain path but exhibits a
stiffer, linear elastic response requiring a separate unloadreload stiffness to be specied
in the constitutive model. Then the model provides good predictions of displacements
under deviatoric monotonic loading and load reversals.

Stress-dependent stiffness
As stated previously for linear elasticity, increasing stiffness with depth can be specied
to take account of its variation due to in situ stress, but this does not take account of
stress changes. Full stress-dependency is obtained if the constitutive model includes
expressions relating stiffness with conning stress. For example, a power law between
conning stress and stiffness forms the basis of the expressions used in the Duncan and
Chang and Hardening Soil models. The MCC model (Roscoe and Burland, 1968; Muir
Wood, 1991) uses a logarithmic relationship between average effective stress p and void
ratio e and therefore a linear stress-dependent stiffness which is appropriate for nor-
mally or lightly over-consolidated clays. The Lade model (Lade, 1977) uses a logarith-
mic dependency of stiffness on the stress state ( p and q). Full stress-dependency of
stiffness is important for more accurate deformation prediction where stresses change
signicantly, e.g. for settlement under a new embankment or heave under a deep
excavation.

Strain-dependent stiffness
The reduction of stiffness with strain is specied with parameters dening stiffness degra-
dation curves. The original Jardine et al. (1986) model simulated undrained behaviour
and the decay of shear modulus G only. Since then the model has been extended to
drained behaviour with two trigonometric functions to dene the decay of shear
modulus G and bulk modulus K requiring ten input parameters and upper and lower
limits to the curves. Its main disadvantage is that it does not simulate small-strain behav-
iour following stress reversals or stress rotations. The HS Small model (Benz et al., 2009)
is simpler in that only two parameters are required to t a logarithmic function to the
decay curve of G while the decay of K is calculated from G using a constant Poissons
ratio. In reality, Poissons ratio is not constant but varies with strain and real soil data
indicates that the decay curves for these two moduli are only broadly similar, so this is a
disadvantage of the model. In any case, accurate measurement of the K decay curve is
notoriously difcult. Stress reversals cause a resetting of the small-strain stiffness to its
maximum value before decaying again with strain while stress rotations of less than
1808 are also reset to an intermediate, interpolated stiffness. Since these models are an
add-on to other elastoplastic models, careful attention should be paid to the way stiff-
ness is modelled at the interface between the models.

2.3.2 How do constitutive models commonly account for the plastic


behaviour of soils?
Yield surfaces
The elastic limit in stress space is dened by a yield surface. Stresses beyond the yield
surface cause permanent and irreversible strains and the constitutive relationship is

39

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 2.7 Types of idealised plastic behaviour

Perfectly plastic Isotropic hardening


Yield surface 2 2
maintains Yield surface
original position after hardening

Stress path Stress path

3 1 3 1

Kinematic hardening Isotropic softening

2
2
Yield surface
after hardening Yield surface
after softening

Stress path

3 1 3 1

always non-linear (regardless of whether the elastic part of an elastoplastic model was
linear). Usually, plastic analysis assumes that stressstrain behaviour is independent of
strain rate, although in reality this is not quite true high strain rates result in a slightly
stiffer response and this should be considered in parameter selection (see Section 3.4.1).

Plastic behaviour can be idealised in the following ways, as also illustrated in Figure 2.7:

g perfectly plastic (or non-work hardening): the yield stress remains constant during
yield
g strain hardening (or work hardening): the yield stress increases during yield
isotropic hardening: the yield surface increases in size with increasing plastic
strain but remains the same shape and in the same position
kinematic hardening: the yield surface is translated to a new position in
response to plastic strain but does not change size or shape
g strain softening: the yield stress decreases during yield, for example due to dilation
of a dense soil during shear. Consequently, the yield surface needs to expand up
to the peak stress and then contract beyond the peak. Convergence of the
calculation needs to be monitored very carefully and outputs are highly mesh-
dependent, so such models tend to be used in research rather than in practice.

40

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

Figure 2.8 Principal stress space and the stress invariants

Space diagonal Deviatoric Space diagonal


Deviatoric plane 1 = 2 = 3 2 plane 1 = 2 = 3

2 Current
stress state
Lode
angle
p Increasing q
g
sin
cr
ea 1
In Current
stress state

3 3

View 1 View 2

Yield surfaces need to be dened in general stress space in order to be adopted in con-
stitutive models for FE analysis, and they are visualised on three axes of the principal
effective (or total) stresses (s1 , s2 and s3 ), i.e. in principal stress space as shown in
Figure 2.8, so that the graphs are unaffected by the chosen coordinate axes directions.
To describe any point in that stress space, rather than use the three values of principal
stress, it is more useful in constitutive modelling to use the following three invariants
(an invariant has the same magnitude and direction no matter which directions are
chosen for the coordinate axes):

1 mean effective stress p which is the average of the three components of normal
effective stress with no shear stress component. Changes in this stress cause
volumetric strains.
2 deviatoric stress q which is the shear component of stress remaining after
subtracting p . Changes in this stress cause deviatoric strains.

The value of p is a measure of the distance along the space diagonal where s1 = s2 = s3
(see Figure 2.8). In other elds of engineering this space diagonal is called the hydrostatic
stress axis, but this term is used less often in geotechnical engineering because hydro-
static has a different meaning concerning the behaviour of bodies of water.

Any plane perpendicular to the space diagonal is called the deviatoric plane (see
Figure 2.8) while q is a measure of the distance from the space diagonal to the current
stress state along that plane. But in which direction is the current stress state from the
space diagonal? This is dened by the third invariant:

3 Lodes angle u which is the angle between a chosen reference axis and the line
between the space diagonal and current stress state. Due to the requirement that

41

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

s1 s2 s3 , u has a range limited to 608 with the extreme ends of the range
representing triaxial compression (s2 = s3 ) and triaxial extension (s1 = s2 )
conditions.

Flow rules
When the stress state reaches the yield surface, the material undergoes plastic defor-
mation. So an elastoplastic constitutive model needs a ow rule that denes the plastic
strain increment at every stress state. There are two types of ow rules:

g Associated ow: the direction of plastic strain is the same as the outward normal
to the yield surface. This simplies calculations in FE analysis but is better suited
to the simulation of metals. In frictional materials like soil, the dilation angle for
the prediction of irreversible volume change during shear and the friction angle are
assumed the same, leading to the prediction of excessively high rates of dilation,
particularly for high friction angles such as with dense sand. Clays with lower
friction angles may be modelled with reasonable accuracy with associated ow.
g Non-associated ow: the direction of plastic strain is specied separately. This
adds complexity to the calculations but is required for frictional materials such as
soil and concrete to avoid the excessive dilation of associated ow rules. Even
though the dilation is less, it can still continue indenitely, which is unrealistic.
Dilation can be linked to plastic strain or a pre-dened cut-off to keep volumetric
strains to realistic levels.

Failure surfaces
Failure points in stress space are dened by a failure surface. Stresses inside the failure
surface are not in a state of failure while stresses on the failure surface are in a state
of failure. Stresses cannot exist outside the failure surface. The MohrCoulomb failure
surface is the most commonly used for soils (see Figure 2.9) and is an extension of
Coulombs friction law (dened by the internal friction angle w and cohesion c ) to
general stress space. Its failure predictions in drained conditions are quite good, but the
strength is obtained from the difference between the major and minor principal stresses
so the intermediate principal stress s2 is not taken into account in the prediction of
failure in general stress space. Therefore, careful selection of strength parameters is
required (see Section 3.4.1).

Sometimes, a c value above zero is required to t test data but this can give the soil an
unrealistic tensile strength. In such cases a tension cut-off should be imposed allowing
small or zero tension.

Undrained strength predictions in effective stress analyses depend heavily on accurate


excess pore water pressure predictions (see Section 4.2.5). With w set to zero, Mohr
Coulomb becomes equivalent to the Tresca failure surface for undrained shear failure
in terms of total stress. This also provides good predictions of failure provided that the
selected undrained shear strength cu value, which is not a fundamental soil parameter, is
appropriate for conditions at that time. Note that Tresca takes no account of strength
changes due to consolidation.

42

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

Figure 2.9 Common failure surfaces in principal stress space: (a) MohrCoulomb; (b) Tresca;
(c) DruckerPrager; (d) von Mises

Space diagonal
(a) 2 Space diagonal (b) 2 1 = 2 = 3
1 = 2 = 3

1 1

3 3

Space diagonal
(c) Space diagonal (d) 1 = 2 = 3
2 1 = 2 = 3 2

1 1

3 3

The hexagonal shape of the MohrCoulomb and Tresca failure surfaces creates some
difculties in their implementation into FE analysis, so versions simplied to a circular
cross-section are available in some programs, i.e. the DruckerPrager and von Mises
surfaces, respectively. Note that these simpler surfaces provide reasonably accurate
failure predictions only for quite simple stress paths (e.g. triaxial compression or exten-
sion), but for complex stress paths they can signicantly over-estimate shear strength and
should not be used.

While the hexagonal shape provides good predictions of failure for many stress paths,
it does not match perfectly with laboratory test observations. Some rounding of the
corners of the hexagon provides a more exact t with observed data, such as in the
Matsuoka and Nakai (1974) surface that is used in some constitutive models.

43

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Perfect plasticity
This is the most simple form of plasticity because the yield surface remains unchanged
during yielding and adopts the same surface as the failure surface. Consequently, it does
not include aspects of behaviour such as continuous yielding from the onset of primary
loading, or the effects of stress path (e.g. unloadreload versus primary loading). It is
sufcient for the prediction of failure mechanisms and assessing factors of safety.

Isotropic hardening single surface plasticity


Single surface plasticity marks the rst level of sophistication over perfect plasticity. It
allows continuous plastic (i.e. irreversible) strain below a failure surface. Examples
include the cap models (DeMaggio and Sandler, 1971) which have a cap yield surface
incorporating volumetric hardening and a xed failure surface due to deviatoric stress,
and MCC (Roscoe and Burland, 1968; Muir Wood, 1991) which has an elliptic yield
surface in qp stress space and is based on critical state soil mechanics. At the critical
state, no more volume change on shearing occurs, deviatoric strain becomes innitesimal
and the soil has failed. The MCC model forms the basis for many other advanced con-
stitutive models addressing different aspects of soil behaviour. There are a number of
drawbacks associated with the original model, some of which have been overcome in
various implementations of the model into FE analyses, so users of the model need to
check carefully the particular implementation in their software. Some of the drawbacks
of the original MCC model are as follows:

g The critical state line is similar to the DruckerPrager failure surface, so the same
inaccurate failure predictions are possible in some stress states, but some models
have been implemented with the MohrCoulomb failure surface for more robust
strength prediction.
g Plastic deviatoric straining is derived from an associated ow rule so it depends
on the amount of friction angle mobilised but some modications have adopted
non-associated plasticity in the deviatoric plane.
g Since only volumetric hardening is considered, deviatoric loading in heavily over-
consolidated soils remains inside the yield surface leading to an overly long and
linear elastic range and a high peak strength. Some modications have introduced
a cut-off surface to address the overly high peak strength.
g The K0 value is determined implicitly from the soil strength parameters.
Furthermore, the yield surface is based on isotropically consolidated clay whereas
K0 consolidated clays have a yield surface rotated toward the K0 line in stress
space and some models include such a modication (e.g. Sekiguchi and Ohta,
1977). Nevertheless, many eld problems have been predicted well with the MCC
model without this modication.

The MCC model is good for predicting the deformation behaviour of very soft soils
under compression (e.g. embankment construction on very soft soil). It is not so good
for unloading problems (e.g. excavations).

Isotropic hardening double surface plasticity


As well as a volumetric yield surface (often called a cap) that is pushed out toward
higher p values during compaction or compression hardening (Figure 2.10), double

44

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

Figure 2.10 Compression (cap) hardening

MohrCoulomb
Void ratio, e

Deviatoric stress, q
Primary failure surface
compression line
Expanding cap
W yield surface
X
Elastic Y Y
unload/reload Z XZ
lines W

pp Mean effective pp Mean effective


stress, p stress, p
pp = isotropic pre-consolidation stress

surface plasticity models include a deviatoric yield surface too (Figure 2.11), e.g.
Vermeer (1978), the hardening soil model (Schanz et al., 1999) and Lades double-
hardening model (Lade, 1977). The deviatoric yield surface expands during shear or
friction hardening toward a failure surface. Such plasticity models are often combined
with the non-linear and stress-path dependent stiffness elasticity type models described
earlier since they have the unloadreload and primary stiffness values to be activated
depending on whether the stress state is inside or on the yield surfaces. These models
provide more realistic displacement predictions, particularly for excavations where
primary shear loading (on the yield surface) may occur even though the volumetric yield
surface predicts elastic unloading. These types of models are becoming the most com-
monly used due to their versatility and ability to handle changes in stress-path direction.

Kinematic hardening multi-surface plasticity models or bubble models


These are the more sophisticated of the advanced models which can describe many
aspects of soil behaviour including anisotropy, destructuration and small-strain stiffness.

Figure 2.11 Friction (shear) hardening

MohrCoulomb
Deviatoric stress, q

Deviatoric stress, q

failure surface
Hyperbolic primary
loading stiffness B
B

Linear elastic
unload/reload
stiffness C
C Expanding inner
yield surface
A A
Axial strain, a Mean effective stress, p
Permanent
plastic strain

45

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 2.12 Multi-surface plasticity models

q
a b, d e Stress rotation
d
Kinematic surface c
stationary, elastic strains e

b
Kinematic a
b c, e f surface f

Kinematic surface
dragged by stress point, p
plastic strain controlled
e
by kinematic surface fac
g sur
cd ndin
Bou
Bounding surface
enlarged, plastic strain
controlled by bounding
surface

Small, kinematic (i.e. they can move about) yield surfaces or bubbles are within, typically, a
critical state bounding surface based on the MCC yield surface except that within the
surface, behaviour is elastoplastic rather than elastic. Within a smaller kinematic surface,
elastic behaviour is predicted but when the stress state reaches the surface, the whole
kinematic surface is dragged toward the bounding surface (Figure 2.12) and plastic strain
occurs in accordance with a ow rule and hardening law associated with the smaller surface.
When the smaller surface reaches the bounding surface, behaviour is controlled by the larger
surface. Examples include Al-Tabbaa and Wood (1989) and Stallebrass and Taylor (1997)
which has a second kinematic surface to take account of recent stress history. Overall, there
is potential for a growing use of such model types in the simulation of clay soils.

Stress-dependent strength
A curved failure envelope, particularly for granular materials, can be more appropriate
than the common linear envelope of the MohrCoulomb failure surface. The Lade model
(Lade, 1977) is an example of a constitutive model with a curved failure envelope. It is
useful in simulating granular materials with large stress changes, for instance in the con-
struction of a rockll embankment where the rst layers have very low conning stress
but on completion of the embankment the stresses in these layers are signicantly higher.

Destructuration
Stress changes may cause loss of structure and a change in behaviour. Some models
recreating this phenomenon that are used in research may see more common use in prac-
tical problems. These include a multi-laminate model (Schweiger et al., 2009) and those
based on modications to the MCC model (Kavvadas and Amorosi, 2000; Rouainia and
Muir Wood, 2000; Baudet and Stallebrass, 2004).

Anisotropic strength
Some models consider only inherent anisotropy while others consider both inherent and
induced and are formulated in terms of all six components of stress and strain rather
than invariants.

46

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

Model types include a rotated yield surface based on the MCC model (Wheeler et al.,
2003), or one based on the multi-laminate framework (Schweiger et al., 2009). The
MIT-E3 model (Whittle, 1993; Ganendra and Potts, 1995) is a complex model similar
to kinematic surface models in that plasticity can occur within a critical state bounding
surface but the bounding surface rotates in general stress space to take account of both
inherent and induced anisotropy. In the NGI-ADP model (Grimstad et al., 2012) for
anisotropic undrained shear strengths of clays and silts, cu is dened for active, passive
and direct simple shear stress states.

Creep
Constitutive models incorporating creep behaviour are complex because viscosity and
ageing effects need to be considered in combination with yielding. More research is
needed before models can be applied routinely in practice. One example by Yin et al.
(2002) is based on the MCC model where the equivalent time concept of Bjerrum
(1967) was developed to obtain time-dependent stressstrain behaviour based on both
MCC and visco-plasticity concepts. It predicts accelerated creep when the stress state
is near the yield surface, unloadreload behaviour, relaxation and the effects of a change
in shearing rate.

Hypoplasticity
Models based on hypoplasticity (e.g. Kolymbas, 1991; Gudehus, 1996; von Wolffers-
dorff, 1996) may see increasing use in practical problems. They take a different approach
to elastoplasticity since there is no distinction between elastic and plastic behaviour and
hence there is no explicit yield surface or hardening rules. They are relatively simple since
a single stress tensor equation is used to describe the mechanical behaviour of a soil and
the model parameters are based on fundamental properties of the soil grains. In particu-
lar, they model pressure and density coupling, dilation, contraction and variable strength
and stiffness.

2.3.3 How do constitutive models commonly account for rock


behaviour?
As described in Sections 2.2.2 and 2.2.3, discontinuities have a signicant effect on the
engineering behaviour of rock. How a constitutive model accounts for the discontinuities
is the most important consideration in the modelling of rock masses. Discontinuities
cause distributions of stress and strain that are different to those predicted by the
common elastic or elastoplastic models for soil. They inuence strength and stiffness
properties in a non-linear and anisotropic fashion.

If there are no signicant discontinuities in a hard rock mass, then it can be modelled as
an isotropic linear elastic material provided that the stresses to be applied are truly within
the elastic range.

Otherwise, discontinuities can be modelled in an implicit or explicit manner as follows:

Implicit discontinuity modelling


MohrCoulomb failure criterion
The strength parameters w and c of intact rock are reduced signicantly to take account
of discontinuities. The strength parameters need to be selected for an appropriate stress

47

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

range because the model has a linear stress-dependency of shear strength whereas rock
mass has a highly non-linear stress-dependency. Therefore, it is not adequate for large
stress ranges in rock. Another disadvantage is that the determination of equivalent
strength parameters for a discontinuous rock mass involves a lot of approximation and
uncertainty. The behaviour of soft rocks is more similar to soils and so the Mohr
Coulomb failure criterion is more suited to soft rocks.

HoekBrown model
This is an isotropic LEPP model with a non-linear shear and tensile strength criterion
intended specically for isotropic, homogeneous, weathered rock. It is suited to irregular
discontinuity patterns where no signicant anisotropy or dominant sliding directions
occur. The input parameters include an intact rock strength, a Geological Strength Index
(GSI) based on rock mass descriptions and Youngs modulus. The model has evolved
many times since its original version in 1980, with the Hoek et al. (2002) version being
the rst suited to implementation in FE analyses, as summarised by Hoek and Marinos
(2007).

Anisotropic LEPP model


This is suited to regular discontinuity patterns where shear is likely to occur along pre-
dened shear planes. Such models often use a MohrCoulomb failure criterion, so
would still have the disadvantage of a linear stress-dependency of strength. Cross-aniso-
tropy of stiffness may also be allowed in some models in order to specify a different stiff-
ness in one direction due to the effect of discontinuities in one plane. A disadvantage
with such models is that potential failure mechanisms may be missed due to the de-
nition of pre-dened failure surfaces.

Explicit discontinuity modelling


Interface elements allow separation and relative sliding of elements along pre-dened sur-
faces where the elements are installed in the mesh. When the interface is in compression,
normal and shear stresses are transferred across the element according to specied
normal and shear stiffness values. A small or zero tensile strength is specied at which
point the element simulates a gap opening. A Coulomb-type friction criterion is specied
at which the shear stiffness drops to a specied residual value. More detail on interface
elements is provided in Section 5.1.3. Setting up interface elements in an FE mesh manu-
ally can be laborious, so some programs perform this task automatically according to
various fracture network models. The intact rock between the interface elements can
be modelled with a simple isotropic linear elastic constitutive model (if failure modes
through intact rock need not be predicted), or more usually with an LEPP model with
MohrCoulomb failure criterion with appropriate parameters for the intact rock.

2.4. Typical applications


Table 2.1 provides brief guidance on the appropriate selection of the elastic and plastic
parts of elastoplastic constitutive models for common structure types in soil. Further
project-specic guidance can be obtained from FE analysis case studies of similar
projects in similar ground conditions where the selection of constitutive models may
be justied and the outputs appraised.

48

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Table 2.1 Appropriate elastoplastic constitutive models for typical applications

Application type Characteristics Elastic part Plastic part

Embedded retaining wall Unloading stress path with Isotropic linear elastic (stiffness increasing Perfect plasticity provides minimum appropriate
deflection and structural deviatoric loading. with depth) usually provides minimum degree of accuracy for design prediction of wall
forces Strain level low to medium. appropriate degree of accuracy for design structural forces (bending moment, shear force,
prediction of wall structural forces support forces).
(bending moment, shear force, support Double surface hardening required in
forces). combination with higher order stiffness for
Accuracy improved with non-linear improved predictions and for excavation heave
stress(path)-dependent stiffness, and predictions.
required to avoid over-prediction of
excavation heave.

Excavation or tunnelling- Unloading stress path. Strain-dependent stiffness combined with Double surface hardening in combination with
induced ground Strain level very low to low. non-linear stress(path)-dependent higher order elastic part.
movements stiffness.

Cut slopes Prediction of soil strength is Isotropic linear elastic for granular soils Perfect plasticity for granular soils and stiff clays.
important. The prediction of and stiff clays. Specialised analysis required to simulate softening
soil displacement in normally In normally to lightly over-consolidated and progressive failure in stiff plastic clays.
to lightly over-consolidated clays, creep displacements become Alternatively, use perfect plasticity with

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
soils can be difficult due to significant but are difficult to predict. c , 1 kPa and post-rupture or residual value of
creep effects. w for conservatism.
High strain levels. Double surface hardening for normally to lightly
over-consolidated clays to predict plastic straining
How are constitutive models selected?

49
pre-failure.
Table 2.1 Continued

50
Application type Characteristics Elastic part Plastic part

Embankment Compression loading problem Stress-dependent stiffness important for At least single surface hardening for soft clay
construction (primary loading on normally both fill and foundation due to large foundations, double surface hardening for other
consolidated foundation stress ranges. Non-linear (primary loading) foundations and fill material.
soils), with unloading stiffness for soft clay foundations. Anisotropic strength of foundation soil is
reloading if simulating Anisotropic stiffness of foundation soil is important, particularly in low-plasticity clays.
reservoir filling and important, particularly in low plasticity If using isotropic model, an average soil strength
drawdown. clays, but it may not be possible to model must be used with care to prevent over-
High strain levels. this in combination with other prediction of safe embankment heights.
Geotechnical Finite Element Analysis

Any significant anisotropy in requirements. In which case use average Stress-dependent strength important for granular
foundation soils has an isotropic stiffness and perform parametric fill.
important effect on study on stiffness.
behaviour. Stress-path dependent stiffness important
if simulating reservoir filling and
drawdown for embankment dams.
In normally to lightly over-consolidated
clay foundations, creep settlements
become significant but are difficult to
predict.

Piled foundation Compression loading stress Axial behaviour depends heavily on Axial behaviour depends heavily on pilesoil
path (unless tension pile) and pilesoil interface properties following pile interface properties following pile installation, so
deviatoric stress. installation, so simple isotropic linear simple perfect plasticity probably sufficient for
Low strain levels. elastic (increasing with depth) probably surrounding soil.
sufficient for surrounding soil. Double surface hardening required in
Non-linear, stress (path) dependent combination with higher order stiffness for
stiffness required for accurate lateral load improved predictions for lateral deflection and

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
deflection predictions. pile group effects.
Also strain-dependent stiffness required
for accurate prediction of displacements
and interactions in pile groups.
Spread foundation Compression loading stress Isotropic linear elastic (with stiffness Perfect plasticity probably sufficiently accurate for
path and some deviatoric appropriate for stress and strain level) bearing failure prediction.
stress. probably sufficient for settlement Anisotropic strength reduces bearing resistance,
Medium strain levels. prediction. requiring an anisotropic model in certain soils or
Improved settlement and raftsoil careful selection of average shear strength in an
interaction predictions with non-linear isotropic model.
stress (path) dependent stiffness.
Prediction of displacements adjacent to
foundation further improved with strain-
dependent stiffness.

Dynamic analysis Very small strain levels. Simple isotropic linear elastic model could Plasticity not required for basic wave propagation
(low strain) Earthquake analysis: greater be used for basic wave propagation studies.
range of strain levels, cyclic studies, but with small-strain stiffness For high strain dynamic and earthquake studies,
loading effects and possible (G0). permanent strains under cyclic loading may need
liquefaction. Strain-dependent stiffness would be to be considered, requiring specialised
required for larger strain ranges. constitutive models.
For high strain dynamic and earthquake
studies, volume change characteristics
under cyclic loading may need to be
considered, requiring specialised
constitutive models.

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

51
Geotechnical Finite Element Analysis

REFERENCES
Al-Tabbaa A and Wood DM (1989) An experimentally based bubble model for clay. In
Numerical Models in Geomechanics NUMOG 3 (Pietruszezak S and Pande GN (eds.)).
Elsevier Applied Science, London, pp. 9199.
Baudet B and Stallebrass S (2004) A constitutive model for structured clays. Geotechnique
54(4): 269278.
Benz T, Vermeer PA and Schwab R (2009) A small-strain overlay model. International
Journal for Numerical Methods in Geomechanics 33(1): 2544.
Bjerrum L (1967) Engineering geology of Norwegian normally-consolidated marine clays
as related to settlements of buildings. Geotechnique 17(2): 81118.
Clayton CRI (2011) Stiffness at small strain: research and practice. Geotechnique 61(1):
537.
DeMaggio FL and Sandler IS (1971) Material model for granular soils. Journal of Soil
Mechanics and Foundations Division ASCE 97: 935950.
Duncan JM and Chang CY (1970) Nonlinear analysis of stress and strain in soils. Journal
of Soil Mechanics and Foundations Division ASCE 96: 16291653.
Ganendra D and Potts DM (1995) Discussion on evaluation of constitutive model for
overconsolidated clays by AJ Whittle. Geotechnique 45(1): 169173.
Grimstad G, Andresen L and Jostad HP (2012) NGI-ADP: Anisotropic shear strength
model for clay. International Journal for Numerical and Analytical Methods in Geomecha-
nics 36(4): 483497.
Gudehus G (1996) A comprehensive constitutive equation for granular materials. Soils and
Foundations 36: 112.
Hoek E, Carranza-Torres CT and Corkum B (2002) HoekBrown failure criterion 2002
edition. Proceedings of the 5th North American Rock Mechanics Symposium, Toronto,
Canada, 1: 267273.
Hoek E and Marinos P (2007) A brief history of the development of the HoekBrown
failure criterion. Soils and Rocks 30(2): 8592.
Jardine RJ, Potts DM, Fourie AB and Burland JB (1986) Studies of the inuence of
nonlinear stressstrain characteristics in soil-structure interaction. Geotechnique 36(3):
377396.
Kavvadas M and Amorosi A (2000) A constitutive model for structured soils. Geotechnique
50(3): 263273.
Kolymbas D (1991) An outline of hypoplasticity. Archive of Applied Mechanics 61: 143
151.
Kondner RL (1963) Hyperbolic stressstrain response: cohesive soils. Journal of Soil
Mechanics and Foundations Division ASCE 89: 115143.
Lade PV (1977) Elasto-plastic stressstrain theory for cohesionless soil with curved yield
surfaces. International Journal of Solids and Structures 13: 10191035.
Matsuoka H and Nakai T (1974) Stress-deformation and strength characteristics of soil
under three different principal stresses. Proceedings Japanese Society Civil Engineering
232: 5970.
Mogi K (1971) Fracture and ow of rocks under high triaxial compression. Journal of
Geophysical Research 76(5): 12551269.
Muir Wood D (1991) Soil Behaviour and Critical State Soil Mechanics. Cambridge Univer-
sity Press, Cambridge.

52

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are constitutive models selected?

Roscoe KH and Burland JB (1968) On the generalized stressstrain behaviour of wet


clay. In: Engineering Plasticity. Cambridge University Press, Cambridge, pp. 535609.
Rouainia M and Muir Wood D (2000) A kinematic hardening constitutive model for
natural clays with loss of structure. Geotechnique 50(2): 153164.
Schanz T, Vermeer PA and Bonnier PG (1999) The hardening soil model: formulation and
verication. Beyond 2000 in Computational Geotechnics 10 Years of Plaxis, Balkema,
Rotterdam.
Schweiger HF, Wiltafsky C, Scharinger F and Galavi V (2009) A multilaminate framework
for modelling induced and inherent anisotropy of soils. Geotechnique 59(2): 87101.
Sekiguchi H and Ohta H (1977) Induced anisotropy and time dependency in clays. Consti-
tutive Equations of Soils, Proceedings Speciality Session 9, Ninth International Conference
of Soil Mechanics and Foundation Engineering, Tokyo, 1: 229238.
Stallebrass SE and Taylor RN (1997) The development and evaluation of a constitutive
model for the prediction of ground movements in overconsolidated clay. Geotechnique
47(2): 235254.
Vermeer PA (1978) A double hardening model for sand. Geotechnique 28(4): 413433.
von Wolffersdorff PA (1996) A hypoplastic relation for granular materials with a prede-
ned limit state surface. Mechanics of Frictional and Cohesive Materials 1(3): 251271.
Wheeler SJ, Naatanen A, Karstunen M and Lojander M (2003) An anisotropic elasto-
plastic model for natural soft clays. Canadian Geotechnical Journal 40(2): 403418.
Whittle AJ (1993) Evaluation of constitutive model for overconsolidated clays. Geotech-
nique 43(2): 289313.
Yin JH, Zhu JG and Graham J (2002) A new elastic viscoplastic model for time dependent
behaviour of normally and overconsolidated clays: theory and verication. Canadian
Geotechnical Journal 39(1): 157173.

53

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.055

Chapter 3
How are soil and rock parameters
obtained?

3.1. Introduction
3.1.1 Why is it difficult to obtain accurate geotechnical parameters?
Soil and rock are complex materials, often requiring complex constitutive models, a high
number of parameters and advanced testing methods. Obtaining the parameters would
be difcult enough on specied, quality-controlled, manufactured engineering materials,
such as steel or concrete, but soil and rock are not manufactured. They are already in
place under the site and have been subjected to natural and largely unknown geological
processes for, typically, millions of years. It is usually possible to access only a tiny
fraction of the ground volume inuenced by a structure, so the vast majority of it is never
seen. Yet, nite element (FE) analysis and advanced constitutive models allow the
ground to be simulated in precise detail. Bridging the gap between the uncertainty of real
site conditions and the idealised world of constitutive models and FE analysis is the main
difculty in obtaining geotechnical parameters. It requires excellent skills in the
interpretation of site information and thorough background knowledge of geology and
soil/rock mechanics.

Failure to obtain parameters to a sufciently high degree of accuracy will render the sub-
sequent FE analyses of little value, potentially leading to incorrect design decisions. This
chapter provides guidance on the stages of obtaining geotechnical parameters that are
particularly relevant to FE analysis. For further guidance, readers should refer to
specialised publications in the particular area of testing and site investigation.

3.1.2 How is parameter testing of soil planned in a site investigation?


In order to obtain soil parameters for FE analysis, they must be measured in some way in
tests. Tests can be performed either on samples in the laboratory, in which case the
parameters are usually measured directly, or in situ, in which case the parameters are
usually determined empirically or analytically from the test results. Both testing tech-
niques are used together, where possible, to achieve more reliable parameter deter-
minations. But before selecting and undertaking expensive parameter testing, it is
important to form an overview of soil conditions at a site and to characterise the soil into
zones (usually layers) a process called site characterisation (see Figure 3.1). This may
appear more laborious than doing all the parameter testing in one go during the initial
site investigation, but can be more cost-effective, particularly when advanced testing
methods are employed, because fewer, targeted parameter tests need to be carried out.

55

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 3.1 Steps to obtaining soil parameters for FE analysis

Site characterisation

Classification tests
Logging of samples Form ground
e.g. CPT, SPT, Atterberg
(visual description) model
limits, sieve analysis

Form analysis Parameter Advanced in situ and laboratory


model and derivation, selection testing on representative
assess reliability and validation ground or samples

Obtaining parameters

Site characterisation
At an early stage, the ground must be categorised into a number of discrete zones or,
usually, layers within which the soil is expected to have similar engineering behaviour.
Site characterisation is performed using both visual description of samples (ISO, 2002
and 2004a) and a large number of inexpensive index and classication tests (e.g. moisture
content, Atterberg limits, particle size distribution, standard penetration test (SPT), cone
penetration test (CPT)). It should then be possible to identify soil types with certain
visual characteristics and which fall within certain test result value ranges. Once the soil
types have been identied, the geometry of the layers can be estimated by interpolating
between the sub-surface investigation (e.g. borehole, trial pit) locations, thereby produ-
cing a ground model.

During subsequent parameter testing, soil samples and in situ test locations are selected
that are representative of each soil layer based on visual description and the same index
and classication tests. The many published correlations between index test results and
engineering parameters also allow the likely range of measured parameters to be assessed
and provide a plausibility check on the values of the measured parameters (see
Section 3.4.2).

How many parameter tests and where?


The required number of tests for each parameter for each soil layer depends on the
degree of uncertainty in each parameter and the sensitivity of the key FE analysis
outputs to each parameter. Annexes P, Q and S of Eurocode 7 Part 2 (CEN, 2007)
provide guidance on the number of tests to perform depending on parameter variability
and experience. Sampling and in situ test locations should be selected where the soil will
inuence signicantly the behaviour of the structure being studied.

In situ or laboratory testing?


In soils that are not conducive to high-quality sampling (e.g. sand, very soft clay, seabed
sediments), parameter tests on the soil in its natural state can only be performed in situ.

56

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Table 3.1 Examples of dominant stress paths and test procedures

Construction activity Dominant stress path Example test procedure

Settlement due to landfill Compressive stress path K0 consolidation in triaxial


test
Embankment construction Undrained (deviatoric) loading Consolidated undrained
on soft clay triaxial test
Excavation Unloading in mean stress, Triaxial extension test
loading in shear

In soils conducive to high-quality sampling (e.g. clay), laboratory testing is preferable


but, funds permitting, a combination of the two is certainly desirable. The advantages
and disadvantages of each type are summarised in many texts (e.g. Clayton et al.,
1995). The principal advantages of in situ testing are that larger volumes of soil are
tested, their results can be obtained earlier in the site investigation and the soil is less
disturbed. As opposed to this, the principal advantages of laboratory testing are that
boundary conditions are more precisely controlled and stresses and strains are uniform
throughout the specimen which allows non-linear soil parameters to be determined more
precisely.

Which test methods?


The critical parameters should be determinable with sufcient accuracy from the chosen
test method(s) see Section 3.3. If further selection is required, this will probably be
based on economy and the availability of the methods. Then consider the stress level and
dominant stress path in the construction sequence to be analysed, in order to specify the
correct test procedure. Some examples are given in Table 3.1.

3.1.3 How is the site information needed to perform FE analysis of rock


masses obtained?
For massive rock, only the intact rock parameters are needed, but for fractured rock,
parameters for both the intact rock and the discontinuities are required. The parameters
are then combined to form a mechanical picture of the whole fractured rock mass.

Intact rock parameters are commonly obtained from laboratory tests on carefully pre-
pared core samples, or sometimes from in situ tests provided that the volume of rock
tested is free of discontinuities. Testing discontinuities is more difcult, so parameters are
more commonly estimated from the observed characteristics of the rock mass and the
discontinuities. Consequently, the process of site characterisation plays a more direct
role in the derivation of rock mass parameters than for soil parameters. It is also possible
to test the mechanical properties of the rock mass as a whole by in situ testing, provided
that the discontinuity spacing is small compared with the dimensions of the test.

Due to the variation of the geometry and characteristics of discontinuities and the
difculty of accessing them in site investigations, it is not possible to be precise in the

57

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

denition of fractured rock parameters for FE analyses, so parametric studies play an


even greater role in rock mass modelling than in soil modelling.

Site characterisation
Requirements for the characterisation of rock are described in ISO (2003). Index and
classication tests include porosity, density, hardness and abrasiveness, as well as the
point load strength test which is used to derive a value of the tensile strength of intact
rock. In the visual descriptions, a lot of emphasis is placed on describing the discontinu-
ities. Ideally, large exposures can be examined, but this is not always possible and then
the descriptions would need to be based on cores from drillholes from which it is more
difcult to obtain a comprehensive picture of discontinuity patterns. Not only are the
characteristics of the discontinuity itself described (e.g. colour, texture, weathered state,
estimated strength) but also their geometry (e.g. orientation, spacing, persistence). Often
they are found in parallel sets, otherwise they need to be considered individually, which is
a painstaking task.

How many parameter tests and where?


As with soil, the required number of tests for each parameter for each rock type depends
on the degree of uncertainty in each parameter and the sensitivity of the key FE analysis
outputs to each parameter. Due to the stress-dependency of rock strength, Hoek (2000)
recommends about ve strength tests at different conning stresses to obtain the strength
parameters for the HoekBrown model. Coring and in situ test locations should be
selected where the rock will inuence signicantly the behaviour of the structure being
studied.

In situ or laboratory testing?


Laboratory testing of intact rock on core samples is more common than in situ testing. In
highly fractured rock, in situ testing may be more appropriate due to the difculty of
obtaining a sufciently large test specimen for laboratory testing, provided that the
dimensions of the in situ test are at least six times bigger than the discontinuity spacing
in order to obtain sufciently representative parameters for the whole rock mass.

Which test methods?


The triaxial compression test is preferred for obtaining the compressive strength of intact
rock at different conning stresses, as well as Youngs modulus. Unconned compres-
sion tests do not obtain the non-linear strength of rock due to stress-dependency. For
detailed studies of discontinuity behaviour, triaxial testing of jointed rock is possible,
as is direct shear testing on discontinuities, but more usually laboratory tests are per-
formed on intact rock specimens.

In highly fractured rock where the blocks of intact rock are too small to form laboratory
test specimens, in situ testing is more appropriate. Such tests obtain the strength and
stiffness parameters of the fractured rock mass rather than the intact rock blocks, so are
useful in validating constitutive model parameters by test simulation. In situ test
methods used for rock include the pressuremeter and plate load test, as well as the SPT
for index testing. Seismic testing is also available to obtain the very small strain stiffness.

58

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

3.2. Soil and rock sampling and groundwater measurement


3.2.1 Why is sample quality important?
If realistic parameters are to be obtained from laboratory tests on soil, the sampling must
be of the highest quality in order to ensure that the sample in the test, on reconsolidation,
represents as closely as possible its behaviour in the eld. This is particularly so for
advanced laboratory testing where the measurements are particularly sensitive to sample
disturbance. Just a minor disturbance could result in unrepresentative stressstrain
behaviour, even with proper reconsolidation techniques. Advanced sampling and testing
techniques still give data scatter and one important contributory factor is sample quality.
The goal is to minimise the causes of disturbance as far as is reasonably possible.

Similarly, when obtaining rock cores, drilling and coring methods should be selected and
executed carefully to ensure that the features observed in the core are inherent rather
than induced by the coring process itself.

3.2.2 How are high-quality samples obtained?


The least sample disturbance is achieved by block sampling from trial pits or from large
diameter boreholes (using the Sherbrooke sampler in soft clays). These methods are
usually too impractical and time-consuming for commercial use but are available to
obtain the highest quality samples (refer to Clayton et al., 1995).

The more common sampling methods to obtain the highest quality Class 1 soil samples
(as dened in Eurocode 7 (CEN, 2007)) required for the measurement of soil shear
strength and stiffness are listed in ISO (2006) and are among the Category A denoted
sampling methods. In essence, the triple-tube corebarrel for rotary drilling in (stiff to
hard) clayey or clay soils, or thin-walled open tubes or xed piston tubes in soft to stiff
cohesive soils are the only commonly used sampling methods that can obtain Class 1
samples in ne-grained soils.

In rock, the Category A methods that can obtain samples with no or only slight disturb-
ance of the rock structure in all rock types, as indicated in ISO (2006), are the triple-tube
corebarrel and wireline (double or triple) corebarrel.

The porosity of granular soils is too high for samples to remain intact when the stresses
around them are released. Specialised sampling methods do exist, e.g. pore water freezing
or resin or grout injection, but resulting volume changes may still prevent a Class 1
sample being obtained. Class 1 samples may be obtainable from mixed (clayey) soils
(i.e. silts, sands or gravels with greater than 5% clay), provided they are homogeneous
rather than layered, otherwise lower quality samples would be obtained.

Tube sampling
High-quality samples can be obtained in soft to stiff soils with hydraulically jacked thin-
walled open tubes (Shelby tubes) and thin-walled pistons. They involve pushing a tube
into the ground and then retracting it with the sample inside. The thin-walled open tubes
can be used in clays up to stiff consistency. In very stiff clays the jacking forces become
high and tubes often buckle or samples break up in the tube, so rotary coring is more

59

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

common in stiff to hard clays. For more details on these sampling methods, refer to
Clayton et al. (1995).

Rotary coring
Drilling and sampling are combined because the drilling tool is the corebarrel itself.
Double corebarrels have an outer barrel with a corebit at its end that grinds away an
annulus of soil or rock while the inner barrel is connected via a swivel to the outer barrel,
so that it does not rotate as it slides over the core. The drilling uid ows down between
the inner and outer barrels, thereby avoiding the core. Triple corebarrels have a thin wall
split tube or plastic liner inside the inner barrel that provides further protection to the
core from the drilling uid, reduces inside clearance and allows easier entry and with-
drawal of the core.

In wireline drilling the outer barrel forms a continuous rotating casing for the full depth
of the hole and the inner barrel is raised and lowered through the outer barrel on a wire
line.

For more guidance on rotary coring, refer to Clayton et al. (1995) and Binns (1998).

3.2.3 What are the common causes of soil sample disturbance and how
can they be minimised?
Tube sampling
Reduction in effective stress
Stress relief during sampling and extrusion, pore water migration and air entry all cause
s reduction, particularly in soft, low-plasticity clays. The consequences of this are
volume changes that will be difcult to reverse during reconsolidation without further
disturbance and destructuration.

To help minimise s reduction, weighted drilling mud can be used in open boreholes in
soft clay (Ladd and DeGroot, 2003) while boreholes in stiff clay should be kept dry (if
the clay is strong enough to resist bottom heave) to reduce swelling. Sampling operations
and the preparation of laboratory specimens should be performed with minimum delay.

Loss of structure
Loss of structure (see Section 2.2.1) results in a shrinking of the soils yield surface and a
reduction in stiffness. As a soil is approached by a sampler tube, it undergoes compres-
sive strain in front of the cutting shoe, followed by extensile strain as it enters the tube.
Such strains can damage the sample and cause loss of structure. The magnitude of the
strains can be reduced by using the appropriate cutting shoe geometry (refer to Hight,
2003; Clayton and Siddique, 1999).

Water content changes


Shear distortions during tube penetration create suction gradients between the periphery
and centre of the samples. Moisture will therefore migrate toward the centre in soft clays
resulting in consolidation at the periphery and swelling in the centre, while the opposite
will occur in stiff clays.

60

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

To mitigate this, samples should be extruded and the remoulded soil on the periphery
removed as soon as possible before sealing and storing the sub-samples for later testing.

Rotary coring
Provided the drilling is performed in a skilled manner, the main source of disturbance in
cored samples of clay is reduction of effective stress s , which is greater than with thin-
walled tube samples. This occurs because of the complete stress relief around the core
and the presence of drilling uid that softens the outside of the core and ssure surfaces.
Use of a triple corebarrel system helps to alleviate this. The driller should also select
drilling uids and corebits that help to avoid s reduction, and undertake careful drilling.

3.2.4 Can soil sample quality be measured?


An evaluation of strength and stiffness testing data must include an assessment of sample
quality because it may explain anomalous data or indicate which are the more reliable
results. Destructive techniques exist, which can only indicate sample quality during or
after a test, while more useful non-destructive techniques are still in development.
Non-destructive techniques would allow the selection of the highest quality samples for
testing while portable methods would even allow sample quality to be assessed on site to
aid the selection of sub-samples.

This section contains a summary of examples of both types of technique.

De/e0 measurement (destructive technique)


The ratio of the change in void ratio De over the initial void ratio e0 of a sample is
measured for reconsolidation of a sample in a triaxial cell or oedometer to its in situ ver-
tical effective stress sv . A theoretically perfect sample would behave in an undrained way
(zero volume change) on sampling and reconsolidation and the ratio De/e0 would be
zero. In reality, some disturbance always occurs and the ratio De/e0 is above zero. The
lower the value, the higher the quality of the sample and Lunne et al. (1997) assigned
quality ratings from very poor to excellent to De/e0 ratios depending on the over-conso-
lidation ratio (OCR) (for example, less than 0.03 for excellent and greater than 0.10 for
very poor for OCR of 2 to 4). Lo Presti et al. (2001) stated that the De/e0 measurement
alone is not sufcient to take account of soil type and noted a clear decrease in De/e0 with
increase in the plasticity index of Pisa clay, so they proposed lower limits than those of
Lunne et al. (1997) for high plasticity clays.

Suction measurement (triaxial) (destructive technique)



Prior to sampling, a soil element is subjected to in situ effective stresses sv0 and sh0 . On
sampling, total stresses drop to zero and in a perfect sample the suctions, or residual
effective stress pr , would equal the mean in situ effective stress pm

= (sv0
+ 2sh0 )/3. In

reality, however, pr is much less than pm due to sample disturbance. The sample quality
can be assessed by comparing the values of pr and pm
. The drawback of this method is
the need to estimate or measure K0 . Indeed, this method is also used to estimate K0 , so
one can be left with a single measurement and two unknowns, unless an alternative
method of determining K0 is used. pr is also affected by pore water migration within
samples and moisture loss, so, on its own, it is an insufcient indicator of sample quality.

61

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

To perform the measurement, the conning stress in the triaxial cell is increased iso-
tropically during the saturation stage of a test. The value of cell pressure when the pore
pressure readings become positive corresponds with the suction pr in the sample. The
De/e0 measurement procedure described above could then follow during reconsolidation
of the sample.

Suction measurement (direct) (non-destructive technique)


Direct measurements of suction pr can be made on unconned samples using tensi-
ometers (low suctions), pressure plates or with a portable suction probe (Ridley and
Burland, 1993). The disturbed outer zone of the sample should be trimmed away before
taking readings.

Shear wave velocity (non-destructive technique)


Shear wave velocity vs can be measured both in situ (see Section 3.3.5) and in the labora-
tory using bender elements (see Section 3.3.1). Differences between vs in the in situ state
and in the sample will depend on changes in effective stress s and in the soil skeleton, so
they provide an indication of sample quality. vs can be measured on unconned samples
to allow sample selection, or can be measured on reconsolidated samples where the stress
state matches approximately with the in situ stress state. Recent research suggests that a
ratio of unconned vs over in situ vs in excess of about 0.60.8 is indicative of a very good
sample.

The disadvantage with vs measurement on reconsolidated samples is that it does not


allow sample selection because the laboratory test has already started. To overcome this
problem, vs measurement on unconned samples can be combined with residual effec-
tive stress pr measurement using a suction probe in order to apply a correction to the
vs value for comparison with the in situ vs value. This allows more informed sample
selection but vs measurement should still be performed subsequently on reconsolidated
samples to provide the most accurate comparison with the in situ vs value and assess-
ment of sample quality. For more guidance on the shear wave velocity method of sample
quality assessment refer to Hight et al. (2003), Landon et al. (2007) and Sukolrat et al.
(2008).

3.2.5 How is groundwater pressure measured in the field?


Piezometric pressure is measured with piezometers and this section includes a brief
description of suitable types and installation methods. Comprehensive guidance in this
area is provided by Dunnicliff (1993). Piezometers can be constructed as either open or
closed types, as shown in Figure 3.2.

Open piezometers
These are the traditional standpipe or Casagrande piezometers composed of a plastic
pipe through which water rises via a lter. The piezometric level is determined by
measuring the depth to the water level in the plastic pipe. They are cheap and reliable
but suffer from slow response times so are less suited to rapid changes in groundwater
pressure (e.g. tidal variation, reservoir lling/drawdown, pumping tests) and cannot
measure suctions.

62

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Figure 3.2 Common piezometer types

Grout seal

Standpipe

Bentonite seal

Diaphragm
Filter transducer, e.g.
vibrating wire
Sand

Open type Closed type Multiple piezometers


piezometer piezometer in a fully grouted
borehole

Closed (or diaphragm) piezometers


These use a transducer for direct measurement of water pressure via a diaphragm in
contact with the pore water. Measurement methods include pneumatic, electrical resist-
ance and vibrating wire strain gauge transducers. They have rapid response times and
can measure suctions but are more expensive and more likely to malfunction than the
open type.

Installation
A porous lter element (or well point) is installed in a sand zone in a borehole, sealed
from the rest of the borehole using bentonite clay or a bentonite-cement grout, such that
the groundwater pressure is measured in the sand zone only. The lter element can be
connected to a plastic pipe (open system) or to a pressure measuring device (closed
system).

A more recent alternative installation method for closed type piezometers involves back-
lling the borehole entirely with bentonite-cement grout, without a sand zone since
diaphragm type piezometers require only a small change in water volume to equalise
pressures. This is called a fully grouted piezometer and has the advantages of economy
and simplicity, particularly for multiple piezometers, as well as allowing other monitor-
ing instruments to be placed in the same borehole. Further guidance is available in
Contreras et al. (2008), Dunnicliff (2009), McKenna (1995) and Mikkelsen and Green
(2003).

63

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Multipoint piezometers
It can be wrong to assume that groundwater pressures above and below a single point
reading are hydrostatic, so piezometers should be installed at different depths, either
in individual boreholes or in the same borehole. Multiple piezometers in single boreholes
can be cost-effective and may be achieved with fully grouted piezometers, by careful
installation of a pair of sand zones separated by a bentonite seal or by dedicated multi-
point piezometers (Dunnicliff, 1993).

3.3. Parameter testing


This section provides a summary of the laboratory and in situ test methods commonly
used to obtain soil and rock parameters.

3.3.1 How is the triaxial cell used to obtain soil and rock parameters?
The conventional triaxial cell apparatus is described widely in textbooks and test stan-
dards (e.g. ISO, 2004b). It is suitable for measuring the shear strength and large strain
stiffness of soil and soft rock specimens. Hard rock specimens may require a different
set-up with steel cell walls instead of perspex and hydraulic oil instead of cell water
in order to apply the high conning stresses required to obtain the non-linear stress-
dependent failure envelopes (e.g. Hoek and Franklin, 1968).

For improved accuracy, the ram force should be measured inside the cell with a water-
proof load cell and pore pressure should be measured at both the base pedestal and at
mid-height with a pore pressure probe in effective stress triaxial tests to monitor the
equalisation of pore pressures in the specimen. Also, the most accurate volume change
measurements are obtained using electronic, screw-driven combined pressure/volume
controllers where volume is measured to a precision of 1 mm3 by the stepper motor.
Larger diameter (75100 mm) specimens are preferable in order to include more of the
fabric of a soil and to reduce the effects of disturbance during specimen preparation.

Stress path triaxial cell with local strain measurement


Stress path triaxial cells (or hydraulic triaxial cells) with computer control of cell, ram
and pore pressure allow the stress state in a soil specimen to be manipulated along any
stress path. Local strain measurement is performed with transducers attached to the
specimen itself inside the cell of the triaxial apparatus, usually on the middle third of the
specimen, remote from its ends. Local axial strain measurement eliminates the compli-
ance errors associated with external instrumentation while an additional transducer
attached to a radial belt around the specimen allows radial strains to be measured, as
illustrated in Figure 3.3.

Local strain measurement on hard rock specimens is achieved by xing strain gauges
directly to the specimen at mid-height for strain measurement in the axial and radial
directions.

Triaxial cell with bender elements


Bender elements are piezoelectric plates that can either produce a bending motion in
response to an input voltage or produce an output voltage in response to a bending motion.

64

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Figure 3.3 Common triaxial cell apparatus variations

External vertical
strain measurement

Internal Bender element


load cell Local vertical
and radial strain pairs on up to
measurement 3 orientations
(for assessment of
anisotropy of
very small strain
stiffness)

Mid-height pore
pressure probe
Computer
control

Conventional Stress path triaxial cell with With bender elements


triaxial cell local strain measurement

When pairs of bender elements are placed on opposite sides of a soil or rock specimen
(Figure 3.3), one transmits a shear wave while the other detects its arrival and the
velocity vs of the shear wave propagated through the specimen is determined from the
time interval between transmission and reception. Shear wave velocity vs is converted
to the very small strain shear modulus G0 using the equation

G0 = rv2s (3.1)

where r is the bulk mass density of the soil.

There are several benets to the use of bender elements in a triaxial test:

g sample quality can be assessed by comparing vs in the specimen with vs in situ


(see Section 3.2.4)
g soil stiffness at very small strains (, 0.002%) can be obtained directly from vs in
order to determine the elastic plateau (G0 ) at the upper end of the decay curve of
stiffness with strain (Figure 3.4)
g anisotropy of stiffness can be assessed from the measurement of vs between
bender element pairs mounted in three orientations on the specimen (as shown in
Figure 3.3) and using polarised wave sources. Note that this will still not provide
enough parameters even for a cross-anisotropic stiffness constitutive model
(refer to Lings et al., 2000, for further guidance).

65

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 3.4 Typical approximate decay of soil secant stiffness with strain and instrument
measurement ranges

Typical strains around


E, G or K
geotechnical structures

E0, G0 or K0
Large
strain
Very small Small
strain strain External strain
measurement
Bender Local strain
elements measurement

~0.002% ~0.1% f log

More guidance on the use of local strain measurement and bender elements in triaxial
testing is provided in Clayton (2011).

Reconsolidation stage
It is well known that stress history, stress state and stress path all inuence the engineer-
ing behaviour of soil. For the accurate measurement of soil parameters, particularly stiff-
ness, these must be recreated in the triaxial test as accurately as possible, recognising that
the void ratio will differ slightly from the in situ state (depending on sample quality). In
conventional triaxial cells, only isotropic reconsolidation is possible. Soils in the eld
overlain by at ground experience an anisotropic (K0 ) stress history of consolidation and
swelling (K0 conditions means vertical strain with zero horizontal strain).

To recreate eld conditions accurately, the soil specimen must be reconsolidated along the
same stress path as in the eld and this requires anisotropic (or K0 ) consolidation. This can
be achieved in stress path triaxial cells by changing the ram pressure while maintaining the
cell pressure constant (stress control) or preferably with zero radial strain as measured by the
radial belt around the specimen (strain control). Strain control mimics the oedometer test
and, since radial stress and pore pressure are measured, allows measurement of the K0 value
in normally consolidated soils and Poissons ratio. Tests with anisotropic reconsolidation
are denoted CAU or CK0U for undrained and CAD or CK0D for drained shear.

Volumetric strains should be minimised during reconsolidation so that further disturb-


ance to the specimen is minimised. This is achieved by avoiding excessive stress excur-
sions and plastic strains during reconsolidation. It is normal practice to reconsolidate
a specimen initially isotropically and then anisotropically to simulate the most recent
stress path experienced by the soil. The reconsolidation stress path must be specied
by the engineer and requires a good grounding in geology in order to estimate the green-
eld stress history of a soil and, in built-up areas, a knowledge of the stress history
resulting from construction (e.g. loading, excavation or dewatering). Figure 3.5 shows
a typical greeneld stress path from initial deposition of a heavily over-consolidated clay:

66

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Figure 3.5 Typical stress path and reconsolidation of a greenfield heavily over-consolidated clay

q
K0 consolidation due to increasing
overburden from geological
deposition
K0 swelling due to decreasing
overburden from geological erosion
B
Field stress path
A E p
Sampling stress
path (approx.)
C K0 reconsolidation due to recent
D geological deposition to current Reconsolidation
in situ stress state at E stress path

K0 consolidation in a normally consolidated state followed by K0 unloading to D to form


a heavily over-consolidated clay, with some K0 reconsolidation to point E resulting from
more recent deposition of supercial deposits. On sampling, the clay is likely to suffer
some effective stress reduction to point A, for instance. The reconsolidation task is to
bring a specimen of the clay back to point E, but if reconsolidated directly back to E
(along path AE), this would recreate the correct stress state but not the stress path and
recent stress history. The recent stress history of the specimen should be followed, but
without large stress excursions that would cause excessive disturbance. For example, the
specimen could be reconsolidated initially isotropically to B, then anisotropically by
reducing q to point C, followed by K0 unloading to D and K0 reloading to the in situ
state at E, thereby recreating the geologically recent stress reversal at point D.

Similarly, Figure 3.6 shows an appropriate reconsolidation stress path for a typical
greeneld lightly over-consolidated clay. The specimen could be reconsolidated initially
isotropically to B, then under axial compression to point C, followed by K0 consolida-
tion to D and K0 unloading to the in situ stress state at E. Note that, due to the inevitable
sampling disturbance of a lightly over-consolidated clay, while the stress state may be

Figure 3.6 Typical stress path and reconsolidation of a greenfield lightly over-consolidated clay

Field stress path


Sampling stress
path (approx.)
q D Reconsolidation
K0 consolidation due to increasing C stress path
overburden from geological K0 swelling due to
deposition decreasing overburden
from geological erosion
E
A

B p

67

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

restored, the void ratio will have reduced, resulting in a higher apparent OCR than for
the in situ soil.

On completion of reconsolidation, creep strains (or secondary compression), which can


affect small strain measurements, may continue to develop for some time, particularly in
high-quality samples. These should be monitored and the shear stage started only when
creep strains have fallen to an acceptable rate, e.g. axial strain less than 0.05% per day as
suggested by Jardine et al. (1991), ideally at the end of each stress path stage but certainly
at the end of the whole reconsolidation stage. Clayton and Heymann (2001) adopted a
tighter criterion of 0.01% per day and found that, provided reconsolidation had not led
to yielding, the measured stiffness of Bothkennar Clay and London Clay were unaffected
by the recent stress history. Consequently, more attention should be focused on allowing
specimens to creep (perhaps for a period of weeks) than on following complex drained
loading paths during reconsolidation. The added benet of allowing creep to complete is
that it can mitigate, to some extent, the effects of earlier sample disturbance.

Shear stage
The strain rate should be much lower than simply to ensure equalisation of pore
pressures, particularly in the early part of the shear stage when the small strain behaviour
of the specimen needs to be captured. An initial axial strain rate as low as 0.05% per
hour is typical, increasing to 0.2% per hour once a vertical strain of 0.2% has been
passed. Such low strain rates allow the effects of creep to become signicant hence, the
importance of minimising creep at the consolidation stage. An alternative is to employ
stress rather than strain control in the early part of the test.

Stiffness measurement
In the externally instrumented conventional triaxial cell, the axial strain will only become
reasonably accurate at strains larger than about 0.5% (less for soft soils) due to compli-
ance effects. With local strain measurement, stiffness measurement is more accurate in
the 0.002 to 0.1% strain range (see Figure 3.4), where stiffness is particularly sensitive
to strain level, therefore most of the stiffness decay curve is obtained in this way.
However, to locate the upper plateau of the curve at strains below 10 (0.002% in clays,
0.0001% in sands), bender element testing needs to be included. Obtaining the com-
plete stiffness decay curve is not only a requirement for using strain-dependent stiffness
constitutive models, but also allows appropriate stiffness values for particular strain
levels to be selected for non-strain-dependent stiffness models.

It is very useful to include an unloadreload cycle in the shear stage in order to obtain
the elastic unload/reload stiffness, as well as Poissons ratio in drained tests.

Anisotropic stiffness
Provided the bedding planes in a specimen are horizontal, three different shear moduli G
can be determined with three pairs of bender elements (refer to Pennington et al., 1997,
and Lings et al., 2000):

g Gvh : vertical shear waves oscillating in the horizontal direction between the bender
elements at the specimen ends

68

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

g Ghv : horizontal shear waves oscillating in the vertical direction between a bender
element pair placed on radially opposite sides of the specimen
g Ghh : horizontal shear waves oscillating in the horizontal direction between a
bender element pair placed on radially opposite sides of the specimen (offset from
the other radially opposite pair).

Isotropic consolidation test


Consolidation properties can be measured during the consolidation stage of triaxial tests
or by dedicated isotropic consolidation testing on samples with a lower aspect ratio of
one (to shorten consolidation times). In conventional triaxial cells the consolidation is
isotropic (in terms of stress, but not necessarily in terms of strain) as opposed to one-
dimensional (K0 ). So, the soil properties derived are for isotropic rather than K0
consolidation.

Bulk modulus K can be obtained with precise pressure control and volume change
measurement (although its decay with strain is very difcult to measure accurately), as
can the Modied Cam Clay parameters k and l, provided the specimen is consolidated
along both an unload/reload line and the normal compression line, but all more accu-
rately with local strain measurement.

Anisotropic consolidation test


Stress path triaxial cells can perform anisotropic consolidation tests by stress or strain
control, including K0 consolidation which is equivalent to an oedometer test. There
are several advantages of consolidation testing in a triaxial cell over a standard
oedometer:

g less specimen disturbance than that associated with inserting an oedometer


conning ring into clay
g no boundary effects associated with specimen bedding into a conning ring
g degree of saturation of the specimen in the apparatus can be measured
g measurement of radial stress and pore pressure for complete determination of
stress state
g larger specimens include more of the soils fabric.

Permeability test
Direct measurement of permeability is also possible in low to intermediate permeability
soils in a triaxial cell, with the advantages over the permeameter that total stresses and
pore pressures equivalent to those in the eld can be applied to the specimen and satur-
ation of the specimen can be veried. The effective stress can be changed in stages to
assess the stress- or void ratio-dependency of permeability. The test can even be per-
formed as an additional stage in a standard triaxial compression test, between the
consolidation and shear stages, although it will take longer for steady-state conditions
to be established in a 2 : 1 aspect ratio specimen rather than a short specimen specically
prepared for permeability testing.

The derivation of permeability values is covered at the end of Section 3.3.7.

69

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
70
Table 3.2 Summary of the parameters obtainable from common laboratory tests

Laboratory apparatus Specimen Parameters obtained


Geotechnical Finite Element Analysis

Relatively accurately Approximately Estimated

1. Triaxial test (external strain Clay Undrained compression: cu; w , c Undrained compression: Eu Measurement of sample
measurement) (with pore pressure measurement) Drained compression: E suction during saturation
For shear strength testing Drained compression: w , c , c Isotropic consolidation: stage: K0
Isotropic consolidation: l, k (Cam Clay), pre-consolidation stress sp Isotropic consolidation: k
Bulk modulus K
Permeability test: k
Sand Drained compression: w , c , c Drained compression: E
Rock w , c , uniaxial compressive strength sci E

2. Stress path triaxial test with Clay Undrained compression: as above + Eu Drained compression: Measurement of sample
local strain measurement or Drained compression: as above + E Poissons ratio n suction during saturation

standard triaxial test with Anisotropic consolidation: E oed Anisotropic consolidation: stage: K0
strain gauges on hard rock K0 , sp , n Anisotropic consolidation: k
specimens

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
For accurate stiffness testing Sand Drained compression: w , c , c, E Drained compression: n
Rock As above + E n
3. Triaxial test with bender All As above + G0
elements
For complete strain-dependent
stiffness

4. Oedometer test Clay E oed , sp E, k


For economic (K0) stiffness Sand E oed E
testing Rock E

5. Direct shear test (shear box Clay Residual strength wr , cr w , c , c, cu


and ring shear) Interface strength d
For residual and interface shear Sand d w , c , c
strength testing Rock d w , c , discontinuity shear
strength


The parameters obtained from these laboratory tests may achieve the indicated degree of accuracy for the specimen tested, but remember that the specimen may not be
representative of the soil in the field. This is particularly the case for sand, which is very difficult to sample to a high quality (see Section 3.2.2) or to reconstitute to the field
condition.

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

71
Geotechnical Finite Element Analysis

3.3.2 How can the other standard laboratory tests be used to obtain
soil and rock parameters?
Triaxial cells are the most versatile and accurate of the common laboratory test methods.
Other methods include the oedometer and direct shear test which are generally less accu-
rate but often more economical and can provide useful parameters in some instances, as
described in this section. A summary of the parameters that can be obtained accurately
from all these common laboratory test methods is provided in Table 3.2.

Oedometer test
The standard oedometer with incremental loading provides limited information on the
stressstrain behaviour of soil specimens due to the small number of equilibrium data
points. Alternatively, constant rate of strain (CRS) oedometers (described in ASTM,
2012) apply constant rates of vertical strain to the specimen, slow enough to be com-
parable with strain rates in the eld. The pre-consolidation stress sp of a specimen can
be identied more accurately because a continuous plot of void ratio/ln p is obtained
rather than an incremental plot.

There are several advantages of the triaxial cell over the oedometer for measuring K0-
consoldiation stiffness, as described in Section 3.3.1, but these advantages are likely to
be less signicant in rm to stiff clays. Permeability values derived from oedometer tests
are generally too low and as much as three orders of magnitude in error.

Many of the drawbacks of the Casagrande oedometer can be overcome by using the
hydraulic oedometer (described in BSI, 1990) for either consolidation or permeability
testing. It is generally more expensive than using a triaxial cell but its advantages over
the triaxial cell are control of the drainage direction and the possibility of using larger
specimens (refer to Head and Epps, 2014). The oedometer can also be used to measure
the compressibility of intact rock specimens.

Direct shear test


Since neither pore pressures nor principal stresses can be determined in standard equip-
ment, the direct shear test should not be used for the general determination of shear
strength parameters for FE analysis. However, measurement of the following parameters
is particularly suited to the direct shear test:

g interface shear strength d between soil or rock and other structural materials
g residual shear strength of rock or soil
g shear strength of rock discontinuities
g shear strength of coarse-grained soils, provided that the in situ density (for peak
strengths) and particle size distribution can be recreated.

3.3.3 How are parameters obtained from in situ tests?


In laboratory testing, elements of soil or rock are tested where stresses are uniform and
can be represented by a single stress point or a single stressstrain relationship, whereas
with in situ testing different locations around the test experience different stress paths
and stress levels.

72

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Consequently, the non-uniform stress state, combined with material stiffness and
strength, inuence test results and prevent the direct measurement of parameters. Indi-
vidual parameters must be derived by analytical or empirical means, making assump-
tions about the other parameters and stress state.

Table 3.3 summarises the parameters that can be determined from common in situ test
methods. The parameter interpretation methods can be classied as analytical (stresses
and strains in the soil can be calculated with some assumptions), semi-analytical
(approximate analysis with broad assumptions) and empirical (direct comparison with
structure performance or laboratory test results). The analytical interpretations are
generally more accurate but some have not been applied widely, so should not be relied
on alone for parameter determination. Even the analytical interpretations make use of
assumptions about geometry, boundary conditions and soil behaviour. It is important
to know these assumptions when determining parameters. In particular, most interpret-
ations assume undrained soil behaviour in tests in clay and drained behaviour in tests on
sands. In intermediate soils (e.g. silts, laminated soils, mixed soils), such interpretations
should only be applied with great care and engineering judgement.

Empirical interpretations are soil and rock-type dependent, so are reasonably accurate
usually only for the types or even specic locations and loadings from which they were
derived. Site-specic empirical correlations still represent the most reliable method of
interpreting in situ test data and they should be derived wherever possible. Analytical
interpretations provide an alternative and they have the advantage of allowing the
engineer to assess the impact of the different assumptions on the derived parameter.

3.3.4 How are parameters obtained from the pressuremeter test?


A pressuremeter is a cylindrical probe usually installed vertically into the ground such
that it applies a uniform horizontal pressure to the ground via a exible rubber mem-
brane (Figure 3.7). The radial pressure and deformation of the expanding cavity in the
ground are recorded which allows interpretations of the stiffness and, in weaker soils, the
strength of the ground.

Calibration procedures are essential and are described in test standards and Clarke
(1995). For instance, the measurement of unload/reload shear modulus requires very
precise measurements, so small inaccuracies in the measuring system will result in large
errors.

There are three main types of pressuremeter, depending on their method of installation,
which are described later in this section. Further guidance on each type can be found in
Clarke (1995).

Pressuremeter tests (PMTs) can be performed using strain and/or stress control, depend-
ing on the type. Stress control is better in the initial stiff elastic phase of the test and then
strain control (typically 1% per minute) during the plastic phase in order to record a
good number of data points on the stressstrain curve. The pressure at soil failure (when
cavity strain increases at constant pressure) is called the limit pressure pL if it is not

73

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Table 3.3 Summary of parameters obtainable relatively accurately from common in situ tests

In situ test method Soil Parameters obtainable relatively accurately1


type
Analytically Semi- Site-
analytically specific2

1. Pressuremeters (SBP, HPD and CPMT) Clay sh , cu , G


For more accurate testing of strength, Sand sh (CPMT only),
stiffness and sh . w , c, G
SBP is particularly recommended Rock G

2. Seismic test All G0


For accurate measurement of very small
strain stiffness

3. Piezocone penetration test (CPTu) Clay cu , OCR


For profiling and more approximate Sand w
strength testing of soil using site-specific
correlations

4. Flat plate dilatometer (DMT) Clay OCR, K0 , cu


For profiling and more approximate Sand w
testing of strength, OCR and K0 in
soils

5. Standard penetration test (SPT) Clay cu


For more approximate strength testing Sand w
using site-specific correlations in soils Rock sc
and weak rock

6. Plate load test (PLT) Clay cu , E u


For approximate strength and stiffness Sand w , E
testing Rock E

7. Permeability tests (all types) Clay k (SBP)


Good permeability testing is notoriously Sand k (pumping test)
difficult but, with care, SBP, packer and Rock k (packer test)
pumping tests are the most accurate.

1
The parameters listed in this table are those considered to be relatively accurately obtainable for FE analysis in
some cases. It is difficult to grade the relative accuracy of each test method and parameter as this depends on
many other factors. Many other approximate relationships exist between these test results and other parameters,
but these should not generally be considered as providing accurate input parameters for FE analysis, but more as
other sources of parameters for parametric study and parameter validation.
2
Site-specific empirical correlations are the most reliable and should also be used where possible to validate the
analytical relationships. For parameters not listed above, site-specific correlations may be less reliable because test
results are significantly influenced by secondary soil and rock properties.

74

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Figure 3.7 Pressuremeter apparatus and typical test data

Pressuremeter curve:

Unloadreload loops on
loading portion of curve
or, on
unloading

Pressure p
Borehole portion
Loading

Probe
Unloading

Pressure p Cavity strain c

Test pocket
Lift-off pressure

Cavity strain c

reached, it can be estimated by extrapolating the stressstrain curve ( pL in the Menard


pressuremeter is different and is dened as the pressure required to double the volume
of the cavity). Failure cannot normally be achieved in PMTs in rock.

Two or three unload/reload loops are performed to measure the shear modulus G.
Unloadreload loops on the unloading portion of the test have become more common
because there is less creep and so less interference on the main pressuremeter curve.
However, similar results should be obtained whether loops are performed during loading
or unloading. G is determined from the slope of the unloadreload loop. Strain-
dependent stiffness can also be derived from the curvature of good quality loops (see
example in Section 8.2.3).

Creep characteristics can also be measured using the pressuremeter. If the pressure is
held constant at any stage during the test, the probe continues to expand. The rate of
creep decreases with time, and a plot of creep against log(time) becomes a straight line
after about a minute. Therefore, the pressure needs to be held constant for only 2 to 3
minutes in order to determine the creep rate.

Results are presented as a pressuremeter curve (i.e. volumetric expansion against


pressure). With direct cavity displacement measurement (as in self-boring pressuremeter
(SBP), high-pressure dilatometer (HPD) and cone pressuremeter (CPMT)), results are
presented as pressure against cavity strain 1c (also called radial strain 1r ), derived from
(r r0 )/r0 , where r is the current radius of the cavity and r0 is the original radius of the

75

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

cavity at in situ state. Note that this is not necessarily the same as the radius of the
pressuremeter probe or the cavity radius at the start of the test. r0 can be approximated
as the probe radius and this is often done initially, but once the in situ lateral stress sh0
has been interpreted from the data, the new r0 value at the point where the in situ lateral
stress was restored is used to recalculate strains. The strain at r0 is called the reference
strain.

Of all the in situ test methods, the pressuremeter test has the most well-dened boundary
conditions, which allows more analytical interpretations of soil parameters to be derived
and permits its straightforward simulation by FE analysis (see Section 3.4.2).

Pre-bored pressuremeters (PBP): high-pressure dilatometer (HPD) and Menard


pressuremeter (MPM)
In a conventional borehole a test pocket, usually at least 2 metres in length, is formed
with smooth, vertical sides at the location of each test. Therefore, the PBP is only suited
to soils where the walls of the test pocket will not collapse prior to installation of the probe.
In the MPM, radial displacement is determined from the change in volume of a water-
lled cell and only stress control is possible. The HPD (also called exible dilatometer) was
developed to test weak rocks but can be used in dense sands (with the advantage that the
test can be performed to beyond peak failure) and rm to hard clays that are too stiff for
other pressuremeters. It measures displacement in a similar way to the SBP, using six
strain-gauged feeler arms, allowing both stress and strain control to be employed.

The MPM is more commonly used for direct design methods using empirical corre-
lations rather than for parameter determination, and can also be used as a proling tool.
It is, however, possible to obtain more fundamental soil parameters from analysis of its
eld curve, particularly if the test procedure is modied.

The main drawbacks of the PBP are the ground disturbance resulting from forming the
borehole and the likelihood of forming a geometrically imperfect test pocket. The hori-
zontal total stress sh at the start of the test is virtually zero, so relatively large cavity
strains are needed to bring the pressure back to its in situ value, which makes the
measurement of in situ sh rather uncertain.

Self-boring pressuremeter (SBP)


To minimise the problem of borehole disturbance, the SBP was developed, which is the
superior of the pressuremeter types. Versions with a cutter can be used in all soils up to
very stiff consistency and soft rock, but not in gravels and stony soils, while versions with
a drilling bit can be used in rock. There are British and French versions of the device and
they differ in several respects. The most signicant difference is that radial displacements
are measured with three or six strain-gauged feeler arms equi-spaced at mid-height in the
British device which allow strain control testing, while the French SBP uses volume
change measurement as in the Menard device.

The horizontal total stress sh at the start of a test in clay should be approximately equal
to the in situ value if the SBP has been installed with only minimal disturbance, which

76

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

makes the SBP the only pressuremeter type that can make reasonably accurate measure-
ments of in situ horizontal total stress sh in clay. In sands, signicant disturbance is
usually caused by installation of the SBP so the initial sh may not be the in situ value.

The SBP is the most accurate of all the pressuremeters because the ground disturbance
should be small and recoverable. But the amount of disturbance is heavily dependent on
the skill and experience of the operator, so the test should be performed by specialist sub-
contractors.

Cone pressuremeter (CPMT)


The cone pressuremeter (CPMT), or full displacement pressuremeter, comprises a pres-
suremeter module mounted behind a standard CPT 15 cm2 cone and friction sleeve.
Radial displacement is measured by three feeler arms. The CPMT is pushed into soil
as part of the CPT operation and, at the required depth, the cone is halted and the pres-
suremeter test undertaken, thus providing a means of interpreting stiffness and strength
parameters from the pressuremeter test as well as strength parameters and proling from
the CPT results.

The horizontal total stress sh at the start of the test can be higher than the in situ value
due to the effects of cone penetration. However, a method of interpreting sh0 in sand has
been derived (as described later in this section) which puts the CPMT at an advantage
over the other pressuremeters in this respect because sh0 cannot be determined reliably
in sands using the other pressuremeter types.

It is the most practical and economical of the pressuremeters, but the initial penetration
of the cone causes soil disturbance which must be accounted for in CPMT-specic
interpretations of the data.

Soil parameters in clay


The in situ earth pressure coefcient K0 is determined from the in situ total horizontal
stress sh0 which is most commonly interpreted using the lift-off method, Marsland and
Randolph (1977) method, or by reconstruction of the pressuremeter curve. The lift-off
is a break in the initial slope of the cavity strainpressure curve (Figure 3.7), so it can only
be interpreted from pressuremeters with direct cavity displacement measurement (i.e. SBP,
HPD, CPMT). The interpretation is more accurate in the SBP because the initial pressure
is close to the in situ sh , while the interpretation is rather uncertain in the HPD and
CPMT due to soil disturbance. Even with the SBP, as the stiffness of the soil increases the
method becomes more uncertain because the deformations of the instrument become
similar to the initial deformations of the soil. Whittle et al. (1995) describe a method for
the six-arm SBP which also allows ground disturbance and anisotropy to be estimated.
Houlsby and Withers (1988) proposed a method of determining sh0 from CPMTs which,
with the application of a correction to account for the nite length of the pressuremeter,
has been found to produce results similar to those measured with an SBP (Yu, 2004).

Undrained shear strength cu can be interpreted from plane strain analytical solutions
of both the loading and unloading stages of pressuremeter tests. Generally, cu values

77

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

obtained from the unloading stage are more reliable because the effects of soil disturb-
ance are less signicant. However, due to high strain rates and the assumption of innite
pressuremeter length, cu values interpreted from pressuremeter tests tend to be higher
than those measured in other in situ and laboratory tests (Yu and Collins, 1998;
Prapaharan et al., 1989). In contrast, the total stress approach of the analytical methods
tends to under-estimate cu for heavily over-consolidated clays (Yu and Collins, 1998).
Therefore, overall, cu values interpreted from pressuremeter tests should not be used
in isolation. A different interpretation method for cu is needed for the CPMT as derived
by Houlsby and Withers (1988). It is a method that has not been applied extensively in
practice, so should be used with caution. Yu (2004) found that cu was over-estimated
by as much as 10%, at least in part due to the assumption of innite pressuremeter
length.

The shear modulus G is obtained directly as half the slope of the unloadreload loops on
the cavity strainpressure plots. The average slope of the unloadreload loop is deter-
mined from a line drawn between the two apexes of the loop, in order to obtain average
G over the range of shear strain. In accurate pressuremeter tests, the strain-dependency
of G can be assessed from the curvature of the unloadreload loops, as shown by
example in Section 8.2.3, between shear strains of about 0.01% and 1%. By combining
PMT-derived G values with seismic test measurements of G0 , the complete decay curve
of stiffness with strain can be estimated. The corresponding mean effective stress p
should be estimated (in order to normalise G for stress). Note that p is assumed to
remain constant and equal to the in situ value in undrained tests, but will increase during
drained pressuremeter tests.

The derived stiffness is the shear modulus in the horizontal plane Gh , which is directly
applicable to, for example, the analysis of retaining walls. For situations involving an
element of vertical deformation, e.g. under a spread foundation or around an axially
loaded pile, Gv is applicable and the value of this parameter would need to be deter-
mined taking into account any anisotropy in the soil.

As with all in situ tests, the pressuremeter does not measure an element stiffness as in
a laboratory test on a specimen of soil, because the extent of soil that inuences the
stiffness value is unknown. Strains undergone by the soil also vary strongly with radial
distance from the probe, so a reference shear strain must be arbitrarily selected as
representative of G for the soil. This reference value is often taken as the strain
measured by the pressuremeter (i.e. at the pressuremeter surface) and Houlsby
(2001) justies this choice by demonstrating that G is strongly inuenced by the soil
stiffness close to the pressuremeter. However, the measured stiffness will be a little high
because of the higher (small strain) stiffness of the soil further away from the
pressuremeter.

Note that, in clay, the tangent modulus G from a pressuremeter curve corresponds theor-
etically with the secant modulus (Figure 3.8) in a laboratory test (Muir Wood, 1990),
although Clarke (1995) found some variation in secant stiffness between triaxial and
SBP derived values in London Clay.

78

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Figure 3.8 Secant and tangent stiffness

Tangent dq
q stiffness da

Secant q
stiffness a

If a constitutive model requires equivalent laboratory tangent stiffness values, then the
pressuremeter values would need to be estimated using the equation:

dGs
Gt = Gs + g (3.2)
dg
(Muir Wood, 1990)

where Gt and Gs are the tangent and secant shear moduli, respectively, and g is the shear
strain.

Soil parameters in sand


Shear strength w can be interpreted from plane strain analytical solutions of both the
loading and unloading stages of pressuremeter tests (Hughes et al., 1977; Houlsby
et al., 1986), as demonstrated by example in Section 8.2.3. Both methods require esti-

mation or, preferably, separate measurement of the critical state shear strength wcv of the
sand. The method based on the loading stage also interprets a value of dilation angle c
but is sensitive to installation disturbance. Both methods use the assumption of zero
elastic strain and a linear relationship between volumetric and shear strains which leads
to under-estimation of w in medium-dense and loose sands. The assumption of innite
pressuremeter length over-estimates w but this has a lesser effect when deriving w from
the unloading stage due to the small cavity contraction (Yu, 2004).

A semi-analytical interpretation of the CPMT based on large laboratory calibration


chamber tests was derived by Schnaid and Houlsby (1992) to obtain peak shear strength
wp and in situ horizontal effective stress sh0

. Determining sh0 in sands from the other
pressuremeter types is difcult and rarely can the lift-off method be employed, so test

simulation probably remains the most reliable method to estimate sh0 in these cases.

The shear modulus G is obtained in the same way as described above for clays, but note
that p cannot be assumed to be constant in drained conditions and its variation must be

79

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

estimated in order to take account of the stress-dependency of stiffness. sv remains


constant while Bellotti et al. (1989) proposed the following equation to estimate sh
during a pressuremeter test in sand:

average sh = sh0

+ 0.2( pu sh0

) (3.3)

where pu is the maximum pressuremeter pressure at the start of unloading. Note however
that sh around a CPMT is likely to be much higher than around other pressuremeter
types.

Soil parameters in rock


The pressuremeter can measure the shear modulus G in the same fashion as pressure-
meters in soil. It is also particularly suited to the measurement of the stressstrain behav-
iour of fractured rock masses for validation of constitutive model parameters by
simulation of the test. This is provided that the pressuremeter probe diameter is at least
six times the discontinuity spacing.

3.3.5 How can stiffness be measured using in situ seismic testing?


As described in Section 3.3.1 for using bender elements in laboratory triaxial testing,
shear wave velocity vs provides a direct determination of the very small strain shear
modulus G0 . Measured G0 values can be combined with pressuremeter measurements
to derive the full stiffness decay curve. Seismic tests can be undertaken by several differ-
ent means as shown in Figure 3.9. Anisotropy of stiffness can be assessed by measuring
vertically propagating, horizontally oscillating shear wave velocity vvh using the down-
hole or uphole technique, together with horizontal shear wave velocity oscillating in each
direction vhv and vhh using the crosshole technique (Fioravante et al., 1998).

Figure 3.9 Common types of in situ seismic testing for G0 determination

Twin receivers for


true interval velocity

Downhole Uphole Crosshole

Source
Receiver

Surface (Rayleigh) wave

Seismic CPT or DMT


(Downhole)

80

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

A useful overview of the seismic method is given by Clayton (2011) and guidance on the
use of the seismic CPT is available in Butcher et al. (2005).

Boreholes must have plastic linings grouted into them for good mechanical coupling
with the ground, and, for crosshole testing, be surveyed for verticality with an inclin-
ometer to verify the distances between them. The crosshole method requires typically
three in-line boreholes and possibly a fourth orthogonal borehole for anisotropy assess-
ment, with spacings of about 57 metres.

Pressure wave velocity vp can also be measured at the same time. This is of little use in
soils and fractured rock because vp is governed by the low compressibility of pore water
(vp 1500 m/s), but in relatively unfractured and unweathered rock, the compressibility
may be lower than water and vp can be used in combination with vs to obtain two iso-
tropic linear elastic parameters. For instance:

G0 = rv2s (3.1 ibid.)


 
v2p
1
2v2s
Poissons ratio n =  2  (3.4)
vp
1
v2s

An alternative to using holes is the surface wave technique where the velocity vr of
Rayleigh waves travelling at the ground surface is measured (shear wave velocity vs
1.09 vr assuming Poissons ratio n = 0.25). vr depends on frequency as well as ground
stiffness, so a range of frequencies must be measured. The values of G0 obtained from
the various Rayleigh wavelengths are combined into a stiffnessdepth prole. Due to the
interpretation needed, these surface methods have a greater uncertainty than the sub-
surface methods and the uncertainty increases with depth.

3.3.6 What parameters can be obtained from other in situ test


methods?
Piezocone penetration test (CPTu)
Common additions to the CPT are seismic wave receivers to form a seismic cone (SCPT)
for shear wave velocity measurement and a pressuremeter to form a cone pressuremeter
(CPMT). These tests have the signicant advantages of effectively combining two test
methods in one while providing perhaps the most cost-effective means of performing
a seismic or a pressuremeter test on site.

A variant of the CPT for soft soils is the full-ow penetrometer where a T-bar or ball
replaces the cone. This has the advantage that penetration resistance is not affected by
soil stiffness or stress anisotropy, so more precise correlations with soil shear strength
may be obtained. Additionally, if full-ow penetration resistances are compared with
cone resistances, soil stiffness or anisotropy may be estimated more accurately than with
the CPT alone. Full-ow penetrometers are often used in offshore site investigations.

81

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

For more information and details of a study of a worldwide database of full-ow


penetration measurements, refer to Low et al. (2011).

The undrained shear strength cu of clays can be estimated using an empirical cone factor
dependent on soil type, soil stiffness, cone type, OCR and anisotropy that is best deter-
mined from a site-specic correlation. In sands, both analytical and empirical methods
exist to estimate shear strength w values that should be reasonably accurate in silica
sands, but slightly under-estimated in compressible calcareous sands.

Flat dilatometer test (DMT)


Use of the DMT also called Marchetti dilatometer is increasing in geotechnical prac-
tice. It comprises a stainless steel blade with a at 60 mm diameter thin steel membrane
mounted on one face. The blade is driven vertically into soils and halted every 20 cm for
the membrane to be inated and the pressure to be measured.

Reasonably accurate interpretations of OCR, K0 and cu can be made from the DMT in
most clays, except heavily over-consolidated clays, although it is not as accurate as the
SBP. Numerical analyses by Yu (2004) showed that interpretations of w from the
DMT in sand are heavily dependent on the rigidity index (G0/p0 ) which leads to much
uncertainty in the direct interpretation of w . The recent development of a seismic dilat-
ometer (SDMT) will help to overcome this difculty by allowing G0 to be measured. Soil
stiffness cannot be determined reliably from the DMT because the expansion of the
membrane by only 1.1 mm does not reach beyond the disturbed soil around the blade
resulting from initial penetration.

Standard penetration test (SPT)


The SPT remains the most widely used in situ testing technique and many design corre-
lations and charts have been derived. These are all based on purely empirical correlations
due to the difculties of interpreting the test by analytical means and taking account of
borehole disturbance and variations in apparatus and procedures. Corrections must be
applied to the N value for overburden stress, energy delivered to the rods and rod length,
as described in ISO (2005) and Clayton (1995).

Quite consistent results can be obtained and reliable shear strength parameters derived
using site-specic empirical correlations with the same test and drilling methods and
equipment. One of the advantages of the SPT is its large database of results in different
soils and fractured rocks to facilitate the derivation of correlations. However, non-site-
specic correlations for the SPT are only approximate and are not appropriate for FE
analysis.

Plate load test (PLT)


Use of the PLT in soil is less common because it is expensive compared with alternatives,
but in fractured rock it is one of the only methods of determining stiffness in a vertical
orientation. Rather like the pressuremeter, it is better suited to simulation of the test to
validate model parameters for fractured rock masses rather than deriving parameters
directly. This is provided that the plate width or diameter is at least six times the discon-
tinuity spacing.

82

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

3.3.7 How is permeability measured in situ?


Laboratory determination of permeability, particularly in granular soils, can be highly
inaccurate. More reliable measurements can be made in situ since a greater extent of the
ground is tested meaning that the effects of large-scale heterogeneities are included in the
determination of permeability.

There are several methods available, using single boreholes, multiple boreholes or the
dedicated self-boring permeameter, as shown in Figure 3.10. The most accurate tech-
nique is the pumping test with multiple boreholes, although in clays this test may take
weeks and a self-boring permeameter would be more appropriate. In rocks, the packer
test is commonly used. The various techniques are described briey below. Further
guidance is given by Clayton et al. (1995) and Cashman and Preene (2012).

Variable head tests (rising or falling)


These tests are used in relatively permeable soils in a cased borehole or open standpipe
piezometer. For a falling head test, clean water is added to the borehole or standpipe to
raise the water level and then the water level is recorded regularly to determine the rate
of fall until the natural level is restored. The rising head method is very similar but
involves removing water to lower the water level and measuring its rate of rise.
Permeability is determined using Hvorslevs method, based on the time taken for water
levels to return to equilibrium, as described in Clayton et al. (1995) and Cashman and
Preene (2012).

Figure 3.10 Common methods of in situ permeability measurement

Constant
Water Water head supply Pressured supply
Ground level removed added t= 0
Falling head
Groundwater
level Packer
Piezometer or Rising head
standpipe tube t=0
Packer
Piezometer tip
or sand filter

Rising head Falling head Constant head Packer test


test test test

Constant, Water supplied to cavity Water supplied


Test measured at controlled rates with as described
Observation wells well for modified SBP
pumping rate constant or variable head
Borehole, Borehole,
Drawdown if required if required

Probe
After pressuremeter test,
probe is retracted a short Perforated
distance to form a test cavity metal tube
in the base of the test pocket
Test cavity
Pumping test

Modified self-boring Self-boring


pressuremeter (SBP) permeameter

83

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Constant head tests


The effective stress changes during an in situ permeability test will cause large volume
changes in a plastic clay. Consequently, constant rather than variable head testing
should be undertaken in such soils. The Gibson method is used to extrapolate a plotted
curve of changing ow rate, caused by the clay consolidating or swelling, to a constant
steady-state ow rate. The permeability can then be determined from the steady-state
ow rate using Hvorslevs equations as described by Clayton et al. (1995). It is better
to perform the test in a standpipe piezometer rather than a cased borehole because small
leakages through any gaps between the casing and the soil will render the measurements
useless.

Pumping tests
Pumping tests involve a larger volume of ground than the single borehole tests
described above so they are more representative of ground conditions and are less
affected by borehole disturbance. A borehole (or test well) is sunk into the stratum
to be measured and water is pumped from it at a constant, measured rate. By measuring
the resulting drawdown (i.e. fall in piezometric level) around the test well in a series of
observation boreholes, the permeability of the stratum can be estimated. Permeability
should be determined using transient ow (or non-steady-state) techniques because they
allow data to be analysed in real time and reduce the period of pumping required. There
is a range of complex manual techniques using different assumptions as summarised in
Cashman and Preene (2012). Alternatively, simulation of pumping tests using FE
analysis capable of transient ow analysis can be employed to determine permeability.
This is particularly useful where anisotropic permeability exists. Axisymmetric analyses
are sufcient where ow patterns have such symmetry, otherwise 3D analyses may be
necessary.

Packer test
This is also called the Lugeon test and is intended for permeability testing in rock.
Packers are inated to seal the top and bottom of a test section in a drillhole. Water
is supplied to the test section at different pressures and the permeability calculated from
the measured ow rate.

Self-boring devices
Under continuing development are self-boring devices based on the SBP. There are two
devices:

1 Self-boring permeameter consisting of a perforated metal tube with an internal


membrane that is inated to seal the tube during installation and is then deated
to allow the permeability test to start using a constant ow system to supply water
down to the tube (Chandler et al., 1990).
2 Modied self-boring pressuremeter consisting of a conventional SBP combined
with the constant ow system for permeability measurement. To perform a
permeability test the SBP is retracted a short distance, leaving a well-dened
cavity in the ground. The length of the cavity can be varied to assess the
anisotropy of permeability (Ratnam et al., 2005).

84

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Constant or variable head tests can be performed in both types by supplying water to the
tube at controlled rates and measuring the injection pressure required to achieve a par-
ticular ow rate. Both methods can minimise soil disturbance by careful installation but,
as with single borehole tests, the disadvantage is that only a small volume of soil is tested.

Deriving permeability values


For consolidation and groundwater ow analyses (see Chapter 4) it is necessary to specify
ground permeability in the input parameters in order to calculate dissipation times and
ow rates, respectively. Permeability in most soils varies with void ratio and therefore
with effective stress. A variation of permeability with depth can be entered to take
account of in situ variations in void ratio, but this will not take account of void ratio
changes during an analysis so is only suitable where insignicant changes in void ratio are
expected. Alternatively, non-linear relationships between permeability and void ratio or
effective stress can be adopted that vary the permeability in response to changes in these
variables. This is useful in situations with large stress changes in soft soils where void
ratio, and hence permeability, changes occur, such as in the construction of embankments
on soft clay as demonstrated in the example in Section 8.4. Where permeability relation-
ships are not available, carefully selected average permeability values should be adopted.

Permeability is often anisotropic due to, for example, soil fabric or laminations, with
higher horizontal than vertical permeability. Therefore, most groundwater ow analyses
allow different permeabilities to be specied for each global axis direction.

Large permeability differences can cause ill-conditioning of the groundwater ow


matrix. Aim to keep the difference to an order of magnitude less than 105.

The permeability of intact rock is normally so small compared with the permeability of
discontinuities that it can be ignored, except in high porosity rocks such as sandstone.
With implicit modelling of discontinuities, the permeability of the discontinuities needs
to be smeared across the domain according to the formula k = kdisc w/s where w is the
width of each discontinuity, s the discontinuity spacing and kdisc the permeability of the
discontinuity. With explicit modelling of discontinuities, the permeability of each discon-
tinuity can be specied and the permeability of the intact rock normally ignored.

Permeability is notoriously variable and difcult to measure accurately. The variation or


measurement error could easily be an order of magnitude. Therefore, it is important to
perform parametric studies of permeability in order to estimate a range of plausible
outputs rather than rely on a single deterministic value.

3.4. Parameter derivation and validation


3.4.1 How are parameters derived from test results?
As described in Eurocode 7 (CEN, 2004) and its guidebooks (Bond and Harris, 2008;
Frank et al., 2004; Orr and Farrell, 2011), parameter derivation is divided into two steps:

Step 1: establish a derived value from each test measured value. The value should be
appropriate for the analysis situation (e.g. axisymmetric/plane strain, stress
level, strain level, etc.).

85

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Step 2: from all the derived values, select a characteristic value that is appropriate to
model the soil or rock layer in the analysis.

These two steps are described in this section.

Derived values
There are many factors to consider in order to determine appropriate derived values,
including aspects of ground behaviour, the test type and conditions, the expected stresses
and strains during construction and the assumptions of the FE model and constitutive
models. Some of the important factors are described here:

g Stress level: this affects most model parameters, so derived values must be
normalised for stress or else be appropriate for the stress levels in the structure to
be analysed.
g Mode of deformation: rarely does the mode of deformation or stress path in
laboratory or in situ tests match that around real structures in the eld, so
judgement is required during the derivation of parameters and assessment of the
reliability of input parameters becomes important. On major projects, the results
of full-scale eld trials or centrifuge tests are useful in this respect.
g Strain level: stiffness at small strains is much larger than the stiffness at large
strains. What was the strain level in the parameter tests and what is the expected
strain level in the ground during construction?
g Specimen volumes: only small volumes of soil and rock are tested in situ and in
the laboratory which may not take proper account of features such as
discontinuities, laminations or large particles. Corrections may need to be applied
to measured values to take account of these features by comparing, for example,
laboratory and in situ shear wave velocity, studying detailed sample descriptions
(e.g. ssure spacing) and back-analysis of case studies in similar ground. On
major projects the results of full-scale eld trials are useful in such cases.
g Brittleness or ductility: plastic clays and rocks tend to be brittle (i.e. shear strength
falls rapidly post-peak) at low conning stresses, while all soils and rocks tend to
be ductile (i.e. a small post-peak drop in shear strength) at high conning stresses.
g Strain rate effects: strain rates in soil tests are typically far higher than in the eld,
so they exclude creep effects. Soil deformations in the eld may continue for some
time following stress change due to the effects of creep and this is an area of
ongoing research.
g Soil ageing: mechanical, biological and chemical processes that are not, as yet,
fully understood lead to improved strength and stiffness with time which might be
quite rapid in relatively recent deposits, e.g. man-made earthwork structures.
g Drainage conditions: most soil tests and subsequent parameter derivations create
or assume wholly drained or undrained conditions but this may not actually be
the case in the test or in the eld.
g Sample disturbance: this should be assessed as described in Section 3.2.4.
g Accuracy of parameter derivation method: this is summarised in Tables 3.2 and
3.3 for certain methods. Many other methods exist, some of which are more
approximate.

86

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Here are seven issues that may need particular attention when determining derived
values of geotechnical parameters for FE analysis:

Intermediate principal stress s2 (including plane strain parameters)


In the laboratory, most strength and stiffness properties are measured using a conven-
tional triaxial compression test where the intermediate principal stress s2 equals the
minor principal stress s3 . In plane strain, axisymmetric and 3D analyses, however, s2
may vary between the values of s3 and the major principal stress s1 . The variation of
s2 between the values of s3 and s1 is dened by the ratio b (as was shown in Figure 2.2).
In plane strain compression, 0.15 b 0.35 approximately (Potts and Zdravkovic,
1999), while in other stress states and in axisymmetric and 3D geometries, b will depend
on the particular stress and strain conditions across the model. But what corrections are
needed to soil parameters determined from triaxial tests where b = 0 (compression) or
b = 1 (extension) in order to apply them in a plane strain or other analysis types with
different b values?


A factor wps /wtc = 1.1 for converting triaxial compression test-derived friction angle to a
plane strain value is often quoted, which has been found to give reasonable results in
bearing capacity calculations (Bolton and Lau, 1993; Oh and Vanapalli, 2008). In the
absence of any other information, this would appear to provide a reasonable, conserva-
tive estimate. Note that pressuremeter tests are analysed with the assumption of plane
strain deformation, thereby deriving plane strain soil properties directly.

Advanced laboratory testing techniques (e.g. hollow cylinder apparatus, true triaxial
testing) allow the independent inuences of b and anisotropy on soil parameters to be
investigated fully. Nishimura et al. (2007) found that failure states in London Clay with
b = 1 or 0 provided lower bound MohrCoulomb failure lines with somewhat higher
strengths observed at intermediate values of b. Vaid and Campanella (1974) measured
higher cu values in normally consolidated Haney clay in plane strain compression com-
pared with triaxial compression (b = 0) or extension (b = 1).

In Cumbria sand, Ochiai and Lade (1983) noted a large increase in w of up to 98 from
b = 0 to b = 1. In other similar studies on sands, including Symes (1983) on Ham
River sand and Sayao and Vaid (1996) on Ottawa sand, strength and stiffness were
noted to increase with b up to about 0.6, and then remain constant or decrease slightly
to b = 1.

From this small number of tests, it appears that adopting triaxial compression or exten-
sion strength values for clay is adequately conservative for other b values between 0 and
1 occurring in FE analysis. Triaxial testing of sands is less common commercially, but it
appears that both triaxial extension and compression tests should be performed before
determining an appropriate w value for FE analysis, since relying on compression tests
alone may be overly conservative.

Some failure surfaces, such as Matsuoka and Nakai (1974) and Lade (1977) offer the
possibility of varying strength with b.

87

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Anisotropy
As described in Chapter 2, some of the commonly used advanced constitutive models for
soils assume isotropy to avoid the additional complexity of anisotropic stiffness and
strength. In which case, it is common practice to measure the anisotropic properties
of a soil in order to enter some average values of strength and stiffness into an isotropic
model that includes the more critical elements of soil behaviour. Anisotropy of small
strain stiffness can be measured using laboratory or in situ seismic testing, and the
assumption is often made that the same degree of anisotropy applies to the larger strain
stiffness. Alternatively, anisotropy of both stiffness and strength can be assessed by com-
paring the results of triaxial compression and extension tests or direct shear tests, but
results may also be affected by the variation of s2 .

Anisotropy is expressed in terms of the angle a, as was shown in Figure 2.3. As for the
study of s2 described above, advanced laboratory testing techniques (e.g. hollow cylin-
der apparatus) are required for independent control of the a and b values. From these, it
appears that cu for clays decreases by as much as 50% from a = 08 to a = 908 for nor-
mally to lightly over-consolidated reconstituted Boston Blue clay (ONeill, 1985; Seah,
1990), while maxima occur at a = 08 and a = 908 with values about 40% lower at
a = 458 in heavily over-consolidated London Clay (Nishimura et al., 2007). Clays also
possess anisotropic stiffness (the higher value can be in the vertical or horizontal direc-
tion) and depends on strain level, clay type, structure and stress history, while w only
changes marginally up or down with a.

In similar studies on sands (Symes (1983) on medium-loose Ham River sand, Wong and
Arthur (1985) on dense Leighton Buzzard sand, and Oda et al. (1978) on dense Toyura

sand), wpeak was observed to decrease by 58 to 128 from a = 08 to a = 908 and stiffness
was also observed to decrease.

It is unusual for anisotropy in soils to govern FE model accuracy over other aspects of
soil behaviour, such as stress- and strain-dependency of stiffness, which is why constitu-
tive models considering these other aspects of behaviour are more common. However,
situations where anisotropy may be particularly important include surface loads (e.g.
from shallow foundations or embankments) on soft low-plasticity clays where the use
of isotropic shear strength could lead to over-prediction of bearing resistance (Potts and
Zdravkovic, 2001; Zdravkovic et al., 2002), and the pull-out resistance of bucket foun-
dations in soft cohesive soils may be over-predicted with an isotropic shear strength
(Potts and Zdravkovic, 2001).

Dilation angle c
The dilation angle c of soil becomes signicant in dense granular soils at relatively low
conning stresses and can inuence FE analysis outputs. It can be measured (refer to
Lees, 2012) or estimated from the equation:

c = w 308 (3.5)
for quartz sands, c is generally taken as zero for clays and calcareous sands. In undrained
effective stress analyses and conned problems (e.g. pile analysis), specication of a

88

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

non-zero dilation angle can allow unrealistic increased effective stresses to be predicted
that delay failure and which would be non-conservative. Therefore, in such analyses, c
should also be set to zero.

Initial state parameters, e.g. stress ratio K0 and pre-consolidation stress


Initial state parameters are not fundamental soil parameters but are used to set up initial
conditions in an FE analysis, such as stress state and stress history (e.g. location of a
yield surface). Consequently, they have a direct inuence on FE analysis outputs and
their importance should not be under-estimated. It should not be assumed that they are
constant in each stratum. Indeed, they often vary with depth, particularly near the
ground surface.

The in situ stress ratio K0 (= sh /sv ) is an important initial state parameter used in the
setting up of initial stresses. It has a particularly signicant inuence on FE analyses
of retaining walls, cut slope stability in clay and shallow foundations in drained con-
ditions. Careful measurement of K0 is required, or it can be estimated by correlation with
other parameters. One common equation (and often the default equation used in pro-
grams) is Jakys equation:

K0 = 1 sin w (3.6)

Note, however, that this equation is appropriate only for normally consolidated soils
while over-consolidated soils are likely to have signicantly higher values of K0 (see
Appendix 3.1 for approximate equations). Compacted soils may also possess high K0
values due to compaction pressures, whose effect may be signicant on retaining walls.
Unfortunately, there is no straightforward way of estimating compaction pressures or of
simulating them in FE analysis. Clayton and Symons (1992) and Clayton et al. (2013)
describe approaches for estimating compaction pressures.

The pre-consolidation stress denes the stress history and is the maximum effective stress
that a soil has previously experienced. It denes the starting location of yield surfaces in
some constitutive models (e.g. Modied Cam Clay, cap hardening models). It can be
determined from consolidation tests (e.g. triaxial consolidation test or CRS oedometer)
where stresses pass from the reloading line to the primary loading line. In heavily over-
consolidated clays, very high applied pressures may be required to reach primary load-
ing. When the pre-consolidation stress cannot be measured, it needs to be estimated
using knowledge of the geological history of the site.

Note that the reloading behaviour and transition to primary loading predicted by
advanced constitutive models is inuenced by the pre-consolidation stress, K0 and
Poissons ratio. Therefore, it is important to validate that all three parameters provide
a sufciently accurate simulation of such behaviour.

Drained parameters from undrained tests


Tests on clays are generally undertaken in undrained conditions but many constitutive
models require (drained) effective stress parameters, e.g. drained stiffness. For elastic

89

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

(unload/reload) Youngs modulus, the conversion from undrained to drained value can
be made using the equation:
 
2 1 + n
E = Eu (3.7)
3

To obtain other parameters and to validate selections, simulation of the undrained tests
should be performed with the effective stress parameters and adjustments made as
necessary to improve the agreement between the real test results and the simulation
outputs.

Poissons ratio, n
Drained Poissons ratio n is rather difcult to measure accurately (Lees, 2012) but most
FE analyses are not particularly sensitive to n values, in which case it is acceptable to
estimate values. In advanced constitutive models, n is usually a true elastic parameter
and a value in the range 0.1 to 0.25 is usually appropriate. In basic models, such as
LEPP models, stiffness behaviour is more sensitive to n values and n differs between
primary loading and elastic unloading/reloading. In the former case, n should be in the
range 0.25 to 0.4 to include behaviour that is otherwise covered by hardening plasticity in
the more advanced models, whereas in the latter case n is a true elastic parameter as in
the advanced models. Loose sands tend to be at the lower end of these ranges, clays mid
to upper and dense sands at the upper end.

When using the gravity switch-on method to establish initial stresses and simulate ll
placement in subsequent construction stages (see Section 1.4.1), the n values are set
to establish appropriate initial K0 values in LEPP models.

In undrained conditions (zero volume change), n equals 0.5 but, since such a value would
create a singularity in the stiffness matrix, n is set close to 0.5 (usually 0.495) in FE
analyses of undrained soil in terms of total stress (see Chapter 4).

Cohesion c
A low c value is sometimes specied in order for a failure criterion to t observed data,
but remember that this gives soil a tensile strength at zero conning stress which is
unrealistic. Therefore, a tension cut-off should be invoked in the constitutive model to
keep tensile stress at very low values or zero.

It is good practice to input a nominal value (e.g. 0.1 kPa) for c to ensure that initial zero
stress states at ground surfaces remain within the yield surface and hence help to avoid
calculation problems.

Characteristic values
Statistical determination of the characteristic value is possible, but rarely are there suf-
cient data available and constitutive model parameters can be rather complex for this
approach. Alternatively, characteristic values can be determined by eye from plots of
derived values and by using engineering judgement. Eurocode 7 (CEN, 2004) provides
an alternative denition of the characteristic value for this purpose, namely a cautious

90

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Table 3.4 Factors affecting the degree of caution exercised when establishing characteristic
values

Narrow caution margin Wide caution margin

Data has small scatter (uniform properties) Data has large scatter (variable ground
and/or sampling/test effects)
Large number of data points, including from Small number of data points from a single
different test types and other sources (e.g. other test type and no other data sources
local tests, published studies, databases)

Experience of soil type No experience of soil type

Large volume of ground involved in the limit state Limit state could occur in small volume of
ground

Structure is strong and stiff and able to Structure is weak and/or flexible and
redistribute loads unable to redistribute loads

No risk of pre-existing failure surfaces Risk of pre-existing failure surfaces

estimate of the value affecting the occurrence of the limit state. Caution means how far
below (or, in rare cases, above) the mean of the derived values is taken as the character-
istic value. The limit state refers to the need to select a characteristic value that is appro-
priate for the limit state to be analysed. For example, in a serviceability limit state (SLS)
analysis, peak soil strength is often appropriate while in ultimate limit state (ULS)
analyses, a post-peak critical state strength is safer. The following procedure is recom-
mended for establishing characteristic values by eye:

1 Plot derived values in an appropriate way (often against depth).


2 Draw by eye or using regression methods a best t line through the data.
3 Draw the characteristic line with an appropriate caution margin depending on the
factors listed in Table 3.4.
4 Assess the range of permissible values for parametric studies (see Section 7.3.3).

Figure 3.11 illustrates the determination of characteristic values by example for an


arbitrary set of data. It can be seen that between 0 and 4 metres depth, there are only
a few data points from a single test method and they have a wide scatter. Therefore, the
characteristic line was drawn with a wide caution margin from the best t line due to the
uncertainty in the value of this parameter. However, below 4 metres depth, many more
data points were obtained from three different test methods and the scatter was small.
Therefore, the characteristic line was drawn with a narrow caution margin from the best
t line because there was more condence in the data.

3.4.2 How are parameters assessed for accuracy?


There is considerable uncertainty in the characterisation of soil, so it could be dangerous
to adopt any derived parameter as representative of the real soil to be simulated without

91

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 3.11 Determining characteristic values for an arbitrary data set

Parameter
0 10 20 30 40 50
0

WIDE CAUTION MARGIN


3

4
Depth: m

8 Test method 1 NARROW CAUTION MARGIN


Test method 2
9 Test method 3

10

Best fit line drawn by eye Characteristic value line drawn with
or regression analysis suitable margin from best fit line

appropriate validation. There are several methods of parameter validation, as described


here, and all should be employed to some extent at the end of the parameter derivation
process.

Soil test simulation


Some soil tests can be simulated straightforwardly by FE analysis and this will check
that the derived parameters and constitutive model represent the soil behaviour
recorded in soil tests. This is particularly important in tests where soil behaviour tran-
sitions from reloading to primary loading behaviour to check that the constitutive
model and input parameters predict this transition realistically. Soil test simulation also
provides an opportunity to adjust the model parameters to achieve a better t with the
test data.

Among the laboratory tests, triaxial and oedometer tests can be simulated straight-
forwardly, either with simple FE models of unit length and only one or two elements,
or even with single-point algorithms, since uniform stresses can be assumed. Idealised
single-element models are shown in Figure 3.12. Single-point algorithms cannot be used
for time-dependent consolidation properties because a drainage path length is required.

92

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Figure 3.12 Idealised laboratory test simulations

Applied stress
or displacement
Centre-line
Applied stress
or displacement
Centre-line Arbitrary specimen
dimensions unless
simulating
time-dependent Cell
consolidation pressure

Zero soil density

Axisymmetric simulation Axisymmetric simulation


of oedometer test of triaxial test

For more detailed studies, multi-element models that simulate end-effects and other test
conditions more accurately may be preferable.

Of the in situ tests, since penetration tests are rather complex to simulate, only PMTs and
PLTs can be simulated in a routine manner. FE models of these tests can be set up in an
idealised way as shown in Figure 3.13. With advanced constitutive models or for greater
accuracy, full simulation of the pressuremeter is necessary, as described in Figure 3.14.

In situ permeability tests are also well suited to simulation by FE analysis. For tests in
boreholes, 2D axisymmetric analyses can be performed provided that ground strata are
horizontal and that permeability is transversely isotropic (i.e. the same in all horizontal
directions but can be different from the vertical direction).

In test simulations, the precision of the analysis needs to be signicantly higher than used
typically in other analyses. Specify a maximum equilibrium error of 0.01% or less, rather
than the more typical 1%. With automatic step-sizing this will also increase the number
of data points, thereby giving a well-dened curve on graphical plots for comparison
with test data.

Plausibility check
While test simulations may verify that the constitutive model and its parameters are
representative of the soil test conditions, what if the test results themselves were in error?
A check should be performed to see whether the derived parameters are plausible, i.e.
that they lie within permissible ranges. Typical values that can be referred to, based
on soil descriptions and characterisation tests, have been published widely (e.g.
Brinkgreve et al., 2010; Day, 2000; Look, 2007; Schnaid, 2009 (Chapter 7)).

93

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 3.13 Idealised in situ test simulations

Centre-line

p Unit height

r Distance to remote boundary ~30r Set initial


stresses
Probe radius r
Axisymmetric simulation of a pressuremeter test
(length/diameter ratio higher than 6)

Applied
displacement

Centre-line

r
Plate Distance to remote
radius r boundaries >6r

Axisymmetric simulation of a plate load test

3.4.3 Are there other sources of parameters?


Much of this chapter has described the most accurate but commercially available
methods for obtaining geotechnical parameters for FE analysis. Nevertheless, FE analy-
sis often needs to be performed without an adequate site investigation. This section
summarises some reasons for a lack of site investigation information and possible means
of overcoming them, before describing other potential sources of parameters that may
provide sufciently accurate model parameters for FE analysis. These techniques are
also useful in providing further validation of model parameters.

Possible reasons for insufficient site investigation information


Highly variable or gravelly ground
Highly variable soil is difcult to divide into identiable layers for parameter testing
while the presence of gravels or cobbles prevents high-quality sampling and the use of
many in situ and laboratory testing techniques.

94

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Figure 3.14 Full pressuremeter test simulation

Centre-line
Interface elements may be
required at the corners of
the test pocket to allow
deformation away from
Test pocket will need to be the fixed vertical boundary
supported by applied
equivalent in situ horizontal
stress until test is simulated
Probe
radius

Detail
p p

Remote model
Pressuremeter boundaries or, for
calibration chamber
test
simulation, at
Axisymmetric chamber boundaries Axisymmetric simulation
simulation (detail)

To overcome this, large volumes of the ground can be tested in situ, for example by seis-
mic testing, large diameter PLT or by trial excavations. Monitored structures in similar
ground conditions can be back-analysed by FE analysis to obtain macro properties for
variable soils.

Insufficient funds available or site investigation completed earlier


Advanced site investigation techniques can be expensive and may be unaffordable on
some projects. Also, the site investigation may have been completed earlier without
having FE analysis in mind and only basic test results and low-quality samples are
available. Of course, such a paucity of information would require very conservative
characteristic values of parameters, bringing into question the value of performing FE
analysis at all. Alternatively, by back-analysing similar case studies or simply estimating
advanced parameters, preliminary FE analyses could be used to demonstrate the poten-
tial benet of performing further advanced parameter testing. If the potential savings in
construction costs outweigh the initial costs of further investigation, then it may be
possible to secure funds for additional testing.

Other sources of parameters


Having taken all possible steps to obtain geotechnical parameters by high-quality site
investigation even in difcult circumstances, the alternative is to use other means to
obtain parameters, as described here. These sources should be used in combination, not
in isolation, to derive parameters by different means, thereby improving their reliability.

Site-specific empirical relationships


After performing a small number of advanced tests to obtain accurate parameters,
deriving site-specic empirical relationships with more basic tests allows a high number
of reasonably accurate parameters to be obtained economically.

95

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

For example, a small number of SBP tests and stress path triaxial tests may be under-
taken to derive accurate strength and stiffness parameters. By comparing these param-
eters with more basic test results, such as SPT N values, adjacent to SBP test or
triaxial test sampling locations, existing empirical relationships between N and derived
parameters can be rened for the specic site. Provided the same equipment, procedures
and personnel are employed for all the other SPTs in that soil layer, reasonably accurate
derivations of parameters could be made from a large number of inexpensive SPTs
across the whole site.

Case study parameters


It is likely that high-quality site investigation data or derived parameters for ground
strata beneath major towns and cities have been reported in case studies somewhere.
These provide a valuable alternative source of parameters, but remember that param-
eters can vary signicantly even within a geological stratum, and different stress states,
stress paths and stress histories also affect test results. Therefore, care is needed to
identify subdivisions within strata. Parameters should not simply be adopted from a
case study without considering likely variations, perhaps based on characterisation
tests, and their effects on FE analysis results. For example, Pantelidou and Simpson
(2007) noted a consistent vertical variation of soil parameters of the London Clay
across central London, closely following its geological subdivisions. However, vari-
ations were also noted within subdivisions which could be categorised to a certain
extent on the basis of Atterberg limits. This shows that even when using a case study
from the same geological subdivision, it cannot be assumed that the parameters will
be the same.

Case study monitoring data


Case studies providing monitoring data for structures built in similar ground conditions
also provide a valuable source of parameters. Back-analysis of a case study can be used
to obtain and validate model parameters, but the case study needs to contain sufcient
detail on the design and construction sequence in order to simulate the case study con-
ditions sufciently accurately.

Databases of soil parameters


Several parameter databases have been developed and they are growing in number and
size all the time. Most contain particular parameters for a specic geological stratum,
city/region or sampling/test method, such as the Norwegian Geotechnical Institute
(NGI) block sample database (Karlsrud et al., 2005), and many are integrated into a
geographical information system (GIS) to identify site investigation locations. Some are
geared toward specic constitutive models, e.g. Duncan et al. (1980) for the Duncan and
Chang model. Databases can give an indication of the permissible range of soil param-
eters but not a single, accurate value. Ideally, detailed information on the source and
method of obtaining each parameter, as well as visual descriptions and characterisation
test results should be available in a database, so that the user can assess its reliability and
suitability for a specic project.

96

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Appendix 3.1
Some equations that may be useful in the validation of model or initial state parameters.

Equations for estimating K0


For normally consolidated soils, Jakys equation:

K0 = 1 sin w (3.6 ibid.)

provides quite a reliable stress ratio. For over-consolidated soils, other, more approxi-
mate equations have been proposed to estimate K0 , such as (from Wroth, 1975):

  n
K0oc = OCR 1 sin w (OCR 1) (3.8)
1 n
       
3 1 Knc 3 1 K0 OCR 1 + 2Knc
m = ln (3.9)
1 + 2Knc 1 + 2K0 1 + 2K0


where m = 0.0022875PI + 1.22, OCR = sv,max/sv and Knc is obtained from Equation
3.6.

Equation 3.8 has provided a reasonable prediction for a number of soils up to an OCR of
about 5 and provided n lies within an acceptable range of 0.1 to 0.25, while Equation 3.9
was proposed for OCR values above 5 and requires an iterative solution.

There are also the following empirical relationships:




0 = (1 sin w )OCR
K oc sin w
(3.10)
(Mayne and Kulhawy, 1982)
 oc 2
K0
= OCR (3.11)
K0nc
(Schmidt, 1966)

Correlations between cu and drained strength or pre-consolidation stress of clays


   
1 + K0
cu = sin w c cot w + sv (3.12)
2

for normally consolidated clays

cu
= 0.23OCR0.8 (3.13)
sv0
(Jamiolkowski et al., 1985)

cu = 0.22sp (3.14)
(Mesri, 1975)

where sp is the maximum pre-consolidation stress.

97

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Elastic relationships
E
G= (3.15)
2(1 + n )
E
K = (3.16)
3(1 2n )
G = Gu since pore water has no shear stiffness, therefore:
1.5E
Eu = (3.17)
(1 + n )
 
1 n E
E oed = (3.18)
(1 2n )(1 + n )

Correlations between G0 and soil states


(2.17 e)2 0.5
G0 = 70 p (MPa) for rounded sands (3.19)
( 1 + e)
(2.97 e)2 0.5
G0 = 33 p (MPa) for angular sands (3.20)
( 1 + e)
(both from Richart et al., 1970)
OCRk 
G0 = 625 pa p (3.21)
0.3 + 0.7e2
for clays where k is obtained from the graph in Figure 3.15 (Hardin, 1978)


p
Gv 0 = Cp (1 + e)3 (MPa) (3.22)
pa

where Cp is a constant generally between 300 and 600 MPa. From a survey of resonant
column tests on clays and sands (Clayton, 2011).
Figure 3.15 k value for Equation 3.21

0.5

0.4

0.3
k

0.2

0.1

0
0 20 40 60 80 100
Plasticity index Ip

98

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

REFERENCES
ASTM (2012) ASTM D4186/D4186M-12e1, Standard test method for one-dimensional
consolidation properties of saturated cohesive soils using controlled-strain loading.
ASTM International, West Conshohocken, PA.
Bellotti R, Ghionna V, Jamiolkowski M, Robertson PK and Peterson RW (1989)
Interpretation of moduli from self-boring pressuremeter tests in sand. Geotechnique
39(2): 269292.
Binns A (1998) Rotary coring in soils and soft rocks for geotechnical engineering. Proceed-
ings of the Institution of Civil Engineers Geotechnical Engineering 131(2): 6374.
Bolton MD and Lau CK (1993) Vertical bearing capacity factors for circular and strip
footings on MohrCoulomb soil. Canadian Geotechnical Journal 30(6): 10241033.
Bond A and Harris A (2008) Decoding Eurocode 7. CRC Press, Abingdon.
Brinkgreve RBJ, Engin E and Engin HK (2010) Validation of empirical formulas to derive
model parameters for sands. In Numerical Methods in Geotechnical Engineering (Benz
and Nordal (eds.)), Taylor & Francis, London, pp. 137142.
BSI (1990) BS 1377-6:1990. Methods of test for soils for civil engineering purposes, Part 6:
Consolidation and permeability tests in hydraulic cells with pore pressure measurement.
BSI, London.
Butcher AP, Campanella RG, Kaynia AM and Massarsch KR (2005) Seismic cone down-
hole procedure to measure shear wave velocity a guideline, ISSMGE TC10: Geophysi-
cal Testing in Geotechnical Engineering.
Cashman PM and Preene M (2012) Groundwater Lowering in Construction, A Practical
Guide to Dewatering, 2nd edn. CRC Press, Boca Raton, FL.
CEN (2004) EN 1997-1 Eurocode 7: Geotechnical design Part 1: General rules. CEN,
Brussels.
CEN (2007) EN 1997-2 Eurocode 7: Geotechnical design Part 2: Ground investigation
and testing. CEN, Brussels.
Chandler RJ, Leroueil S and Trenter NA (1990) Measurements of permeability of London
Clay using a self-boring permeameter. Geotechnique 40(1): 113124.
Clarke BG (1995) Pressuremeters in Geotechnical Design. Blackie Academic and Profes-
sional, Glasgow.
Clayton CRI (1995) The Standard Penetration Test (SPT): Methods and Use, CIRIA
Report 143. CIRIA, London.
Clayton CRI (2011) Stiffness at small strain: research and practice. Geotechnique 61(1):
537.
Clayton CRI and Heymann G (2001) Stiffness of geomaterials at very small strains.
Geotechnique 51(3): 245255.
Clayton CRI and Siddique A (1999) Tube sampling disturbance forgotten truths and new
perspectives. Proceedings of the Institution of Civil Engineers Geotechnical Engineering
137(3): 127135.
Clayton CRI and Symons IF (1992) The pressure of compacted ll on retaining walls.
Geotechnique 42(1): 127130.
Clayton CRI, Matthews MC and Simons NE (1995) Site Investigation, 2nd edn. Blackwell
Science, Oxford.
Clayton CRI, Woods RI, Bond AJ and Milititsky J (2013) Earth Pressure and Earth-
Retaining Structures, 3rd edn. CRC Press, Abingdon.

99

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Contreras IA, Grosser AT and Ver Strate RH (2008) The use of the fully-grouted method
for piezometer installation. Geotechnical News, June, pp. 3037.
Day RW (2000) Geotechnical Engineers Portable Handbook. McGraw-Hill, New York, NY.
Duncan JM, Byrne PM, Wang KS and Mabry P (1980) Strength, stressstrain and bulk
modulus parameters for nite element analysis of stresses and movements in soil masses.
Geotechnical Engineering Research Report No. UCB/GT/80-01, University of Califor-
nia, Berkeley, CA.
Dunnicliff J (1993) Geotechnical Instrumentation for Monitoring Field Performance,
2nd edn. Wiley, New York, NY.
Dunnicliff J (2009) More on fully-grouted piezometers. Geotechnical News, June, p. 32.
Fioravante V, Jamiolkowski M, Lo Presti DCF, Manfredini G and Pedroni S (1998)
Assessment of the coefcient of the earth pressure at rest from shear wave velocity
measurements. Geotechnique 48(5): 657666.
Frank R, Bauduin C, Driscoll RMC, Kavvadas M, Krebs Ovesen N, Orr TLL and
Schuppener B (2004) Designers Guide to EN 1997-1 Eurocode 7: Geotechnical Design
General Rules. ICE Publishing, London.
Hardin BO (1978) The nature of stressstrain behaviour of soils. ASCE Speciality Confer-
ence on Earthquake Engineering Soil Dynamics 1: 390.
Head KH and Epps R (2014) Manual of Soil Laboratory Testing: Volume 3: Effective
Stress Tests, 3rd edn. Whittles Publishing, Caithness.
Hight DW (2003) Sampling effects in soft clay: an update on Ladd and Lambe (1963) In
Soil Behavior and Soft Ground Construction (Germaine, Sheahan and Whitman (eds.)).
Geotechnical Special Publication No. 119, ASCE, pp. 86121.
Hight DW, McMillan F, Powell JJM, Jardine RJ and Allenou CP (2003) Some character-
istics of London clay. Proceedings of the International Workshop on Characterisation and
Engineering Properties of Natural Soils, Singapore, 2: 851907.
Hoek E (2000) Practical Rock Engineering, Chapter 11 Rock mass properties, downloaded
from www.rocscience.com
Hoek E and Franklin JA (1968) Simple triaxial cell for eld or laboratory testing of rock.
Transactions of the Institutions of Mining and Metallurgy 77: A2226.
Houslby GT (2001) In situ tests and the pre-failure deformation behaviour of soils. In Pre-
Failure Deformation Characteristics of Geomaterials (Jamiolkowski, Lancellotta and Lo
Presti (eds.)). Swets & Zeitlinger, Lisse, pp. 13191324.
Houlsby GT and Withers NJ (1988) Analysis of the cone pressuremeter test in clays. 38(4):
575587.
Houlsby GT, Wroth CP and Clarke BG (1986) Analysis of the unloading of a pressure-
meter in sand. Proceedings of the 2nd International Symposium on Pressuremeter and its
Marine Applications. ASTM, SPT950, pp. 245262.
Hughes JMO, Wroth CP and Windle D (1977) Pressuremeter tests in sands. Geotechnique
27(4): 455477.
ISO (2002) ISO 14688-1:2002. Geotechnical investigation and testing Identication and
classication of soil Part 1: Identication and description. ISO, Geneva.
ISO (2003) ISO 14689-1:2003. Geotechnical investigation and testing Identication and
classication of rock Part 1: Identication and description. ISO, Geneva.
ISO (2004a) ISO 14688-2:2004. Geotechnical investigation and testing Identication and
classication of soil Part 2: Principles for a classication. ISO, Geneva.

100

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

ISO (2004b) ISO/TS 17892-9:2004. Geotechnical investigation and testing Laboratory


testing of soil Part 9: Consolidated triaxial compression tests on water-saturated soils.
ISO, Geneva.
ISO (2005) ISO 22476-3:2005. Geotechnical investigation and testing Field testing
Part 3: Standard penetration test. ISO, Geneva.
ISO (2006) ISO 22475-1:2006. Geotechnical investigation and testing Sampling methods
and groundwater measurements Part 1: Technical principles for execution. ISO, Geneva.
Jamiolkowski M, Ladd CC, Germaine JT and Lancellotta R (1985) New developments in
eld and laboratory testing of soils. 11th International Conference of Soil Mechanics and
Foundation Engineering, San Francisco 1: 57153.
Jardine RJ, St John HD, Hight DW and Potts DM (1991) Some practical applications of a
non-linear ground model. Proceedings of the 10th European Conference on Soil Mechanics
and Foundation Engineering, Florence, Vol. 1, Balkema, Rotterdam, pp. 223228.
Karlsrud K, Lunne T, Kort DA and Strandvik S (2005) CPTU correlations for clays.
Proceedings of the 16th International Conference on Soil Mechanics and Foundation
Engineering, Osaka, Vol. 2, pp. 693702.
Ladd CC and DeGroot DJ (2003) Arthur Casagrande Lecture: Recommended practice for
soft ground site characterization. 12th Panamerican Conference on Soil Mechanics and
Geotechnical Engineering, Boston, MA, Vol. 1, pp. 357.
Lade PV (1977) Elasto-plastic stressstrain theory for cohesionless soil with curved yield
surfaces. International Journal of Solids and Structures 13: 10191035.
Landon MM, DeGroot DJ and Sheahan TC (2007) Nondestructive sample quality assess-
ment of a soft clay using shear wave velocity. Journal of Geotechnical and Geoenviron-
mental Engineering 133(4): 424432.
Lees AS (2012) Obtaining Parameters for Geotechnical Analysis. Glasgow, NAFEMS.
Lings ML, Pennington DS and Nash DFT (2000) Anisotropic stiffness parameters and
their measurement in a stiff natural clay. Geotechnique 50(2): 109125.
Look B (2007) Handbook of Geotechnical Investigation and Design Tables. Taylor &
Francis, London.
Lo Presti DCF, Shibuya S and Rix GJ (2001) Innovation in soil testing. In Pre-Failure
Deformation Characteristics of Geomaterials (Jamiolkowski, Lancellotta and Lo Presti
(eds.)). Swets & Zeitlinger, Lisse, 2: 10271076.
Low HE, Randolph MF, Lunne T, Andersen KH and Sjursen MA (2011) Effect of soil
characteristics on relative values of piezocone, T-bar and ball penetration resistances.
Geotechnique 61(8): 651664.
Lunne T, Berre T and Strandvik S (1997) Sample disturbance effects in soft low plastic
Norwegian clay. Symposium on Recent Developments in Soil and Pavement Mechanics,
Rio de Janeiro, pp. 81102.
Marsland A and Randolph MF (1977) Comparison of the results from pressuremeter tests
and large insitu plate tests in London Clay. Geotechnique 27(2): 217243.
Matsuoka H and Nakai T (1974) Stress-deformation and strength characteristics of soil
under three different principal stresses. Proceedings of the Japan Society of Civil
Engineers 232: 5970.
Mayne PW and Kulhawy M (1982) K0-OCR relationships in soil. Proceedings of the Amer-
ican Society of Civil Engineers: Journal of the Geotechnical Engineering Division 108:
851872.

101

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

McKenna GT (1995) Grouted-in installation of piezometers in boreholes. Canadian Geo-


technical Journal 32(2): 355363.
Mesri G (1975) Discussion: new design procedure for stability on soft clays by Ladd and
Foott. Journal of Geotechnical Engineering Division ASCE 101(4): 409412.
Mikkelsen PE and Green GE (2003) Piezometers in fully grouted boreholes. Symposium on
Field Measurements in Geomechanics, FMGM 2003, Oslo, September.
Muir Wood D (1990) Stain-dependent moduli and pressuremeter tests. Geotechnique 40(3):
509512.
Nishimura S, Minh NA and Jardine RJ (2007) Shear strength anisotropy of natural
London clay. Geotechnique 57(1): 4962.
Ochiai H and Lade PV (1983) Three-dimensional behaviour of sand with anisotropic
fabric. Journal of Geotechnical Engineering 109(10): 13131328.
Oda M, Koshikawa I and Higuchi T (1978) Experimental study of anisotropic shear
strength of sand by plane strain test. Soils and Foundations 18(1): 2538.
Oh WT and Vanapalli SK (2008) Modelling the stress versus settlement behaviour of
model footings in saturated and unsaturated sandy soils. 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics
(IACMAG), Goa, pp. 21262137.
ONeill DA (1985) Undrained strength anisotropy of an overconsolidated thixotropic clay,
MSc thesis, Department of Civil Engineering, MIT, Cambridge, MA.
Orr TLL and Farrell ER (2011) Geotechnical Design to Eurocode 7. Springer-Verlag, London.
Pantelidou H and Simpson B (2007) Geotechnical variation of London Clay across central
London. Geotechnique 57(1): 101112.
Pennington DS, Nash DFT and Lings ML (1997) Anisotropy of G0 shear stiffness in
Gault clay. Geotechnique 47(3): 391398.
Potts DM and Zdravkovic L (1999) Finite Element Analysis in Geotechnical Engineering:
Theory. Thomas Telford, London.
Potts DM and Zdravkovic L (2001) Finite Element Analysis in Geotechnical Engineering:
Application. Thomas Telford, London.
Prapaharan S, Chameau JL and Holtz RD (1989) Effect of strain rate on undrained
strength derived from pressuremeter tests. Geotechnique 39(4): 615624.
Ratnam S, Soga K and Whittle RW (2005) A eld permeability measurement technique
using a conventional self-boring pressuremeter. Geotechnique 55(7): 527537.
Richart Jr FE, Hall Jr JR and Woods RD (1970) Vibrations of soils and foundations.
Prentice-Hall, Englewood Cliffs, NJ.
Ridley AM and Burland JB (1993) A new instrument for the measurement of soil moisture
suction. Geotechnique 43(2): 321324.
Sayao A and Vaid YP (1996) Effect of intermediate principal stress on the deformation
response of sand. Canadian Geotechnical Journal 33(5): 822828.
Schmidt B (1966) Discussion: Earth pressures at rest related to stress history. Canadian
Geotechnical Journal 3(4): 239242.
Schnaid F (2009) In Situ Testing in Geomechanics. Taylor and Francis, Oxford.
Schnaid F and Houlsby GT (1992) Measurement of the properties of sand in a calibration
chamber by cone pressuremeter tests. Geotechnique 42(4): 578601.
Seah TH (1990) Anisotropy of normally consolidated Boston Blue Clay. ScD thesis,
Massachusetts Institute of Technology, Cambridge, MA.

102

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are soil and rock parameters obtained?

Sukolrat J, Nash DFT and Benahmed N (2008) The use of bender elements in the assess-
ment of disturbance of soft clay samples. In Geotechnical and Geophysical Site Charac-
terisation (Huang and Mayne (eds.)). Taylor and Francis, London, pp. 14891495.
Symes MJ (1983) Rotation of principal stresses in sand, PhD thesis, Imperial College of
Science, Technology and Medicine, University of London.
Vaid P and Campanella RG (1974) Triaxial and plane strain behaviour of natural clay.
Journal of Geotechnical Engineering 100(3): 207224.
Whittle RW, Hawkins PG and Dalton JCP (1995) The view from the other side Lift-off
stress and the six arm self boring pressuremeter. In The Pressuremeter and its New
Avenues (Ballivy (ed.)). Balkema, Rotterdam, pp. 379386.
Wong RKS and Arthur JRF (1985) Induced and inherent anisotropy in sand. Geotech-
nique 35(4): 471481.
Wroth CP (1975) In situ measurement of initial stresses and deformation characteristics.
Proceedings of the Speciality Conference in In Situ Measurement of Soil Properties,
ASCE, Rayleigh, NC, pp. 181230.
Yu HS (2004) James K. Mitchell Lecture. In situ soil testing: from mechanics to interpret-
ation. Proceedings ISC-2 on Geotechnical and Geophysical Site Characterization (Viana
da Fonseca and Mayne (eds.)), Millpress, Rotterdam, pp. 338.
Yu HS and Collins IF (1998) Analysis of self-boring pressuremeter tests in overconsoli-
dated clays. Geotechnique 48(5): 689693.
Zdravkovic L, Potts DM and Hight DW (2002) The effect of strength anisotropy on the
behaviour of embankments on soft ground. Geotechnique 52(6): 447457.

103

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.105

Chapter 4
How are groundwater effects taken
into account?

4.1. Introduction
4.1.1 How are saturated, partially saturated and dry soils modelled?
In practical geotechnical FE analysis the pore water and solid particles are never
modelled separately, each with their own constitutive relationship as in uidstructure
interaction analyses. Rather, the effect of the groundwater is included in the analysis
using some assumptions. The soil mass of solid particles, water and air is modelled as
a single continuum with a constitutive model that represents the engineering properties
of this combined mass. In effective stress analyses, the stresses within the soil mass are
divided into pore pressure and effective stress.

The modelling of saturated soils is the topic of this whole chapter. Consideration is
required of:

g groundwater pressure and its direct inuence on effective stress


g the inuence of any groundwater ow on groundwater pressure
g any change in the volume of a soil, perhaps due to loading or unloading, because
this requires water to ow in or out of voids so that they can change volume
g temporal effects in low-permeability soils where incompressible groundwater
supports load changes in the short term until they dissipate and load is
transferred to the soil skeleton.

How the FE analysis takes into account these effects is covered in the following sections
of this chapter.

Dry, granular soils are relatively straightforward to model in this respect because there
are no groundwater pressures or groundwater ow to consider. The soil is modelled in
terms of effective stress and the pore pressure is zero. Note that natural dry-looking clays
are never truly dry and cannot be modelled as dry soils. They always retain some
moisture, so are partially saturated.

The behaviour of partially saturated soils is signicantly more complicated than that of
saturated or dry soils. The soil is composed of three phases (solid, liquid and gas) with
the interaction between the liquid water and water vapour/air phases being particularly
complex.

105

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Due to the difculty of simulating partially saturated soil behaviour, it is common


practice, and conservative, to assume the soil above groundwater level is dry with pore
pressure zero and to adopt the same soil parameters as for the saturated soil. This
neglects the elevated effective stress arising from pore suctions that otherwise appear
to give the soil higher strength and stiffness.

Alternatively, partially saturated clay can be modelled in terms of total stress (undrained
Method C see Section 4.2.4) and the stiffness and undrained shear strength set to
higher values as appropriate for the partially saturated clay. However, always remember
that any change in moisture content would cause signicant changes in the stiffness and
undrained shear strength and these would not be taken into account when modelling the
clay in this simplied way.

Also, neither of these simplied approaches would predict volume changes arising from
changes in the moisture content of partially saturated ne-grained soils. Consequently,
simulation of foundation heave on an expansive clay, for example, by FE analysis
requires specialised approaches not covered by this book. Readers could refer to, for
example, Fredlund et al. (2012) and Gens et al. (2006).

4.1.2 What do the different pore pressure terms mean?


In FE analysis it is useful to divide pore pressure into different terms because there are
different options for the way each is calculated.

Steady-state or at rest pore pressure


This is the pore pressure arising from constant hydraulic boundary conditions (e.g.
stationary groundwater level or constant extraction rate from a well). This part of the
pore pressure does not change with time during a displacement or consolidation
analysis. However, it is possible to change the hydraulic boundary conditions in an
analysis stage in order to establish a new steady-state pore pressure distribution (see
Figure 4.1).

This pore pressure can be specied directly for horizontal phreatic surfaces when setting
up the initial stresses and in subsequent analysis stages, or it can be calculated by a
steady-state groundwater ow (or seepage) analysis (see Section 4.3). Horizontal phrea-
tic surfaces occur in hydrostatic conditions, where there is negligible groundwater ow,
and can be used to enter the steady-state pore pressure in other situations where hydrau-
lic gradients are not very high. A typical example may be the steady-state pore pressure
around a retaining wall to a basement with a dewatering scheme, where it may be suf-
ciently accurate, provided hydraulic gradients are not too high and changes in ground
density are taken into account, to specify phreatic surfaces on each side of the retaining
wall and assume a linear variation of steady-state pore pressure toward the wall toe
where the pressures should be equal. The use of groundwater ow analysis to calculate
rather than specify the steady-state pore pressure would be necessary in more complex
cases, e.g. strata with different permeabilities or void ratio-dependent permeability, or
where hydraulic gradients are high.

106

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

Figure 4.1 Generic contours of pore pressure in typical problems in low-permeability soil

Steady state Excess pore Total


pore pressure pressure, ue pore pressure

In situ ue = 0
conditions

Impermeable
wall
Rapid change
in hydraulic Positive ue
boundary
conditions

End of
consolidation
ue = 0

Excavation

Excavation Positive ue
and load Negative
application (suction) ue

End of
consolidation
ue = 0

Excess or non-equilibrium pore pressure


This is the change in pore pressure from the steady-state value caused by loading,
unloading or a rapid change in the hydraulic boundary conditions in undrained or con-
solidation conditions. It can also be generated by deviatoric stress changes where
increased excess pore pressure is generated in normally and lightly over-consolidated
soils and decreased (suction) excess pore pressure in heavily over-consolidated soils
during undrained shear, if the constitutive model incorporates such behaviour. In
general terms, hardening models can predict positive excess pore pressure in normally
and lightly over-consolidated clays, while entering a non-zero dilation angle causes

107

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

negative (suction) excess pore pressures to be generated for heavily over-consolidated


soils, although the latter is not recommended because it can lead to an over-prediction
of undrained strength.

Excess pore pressure can be calculated in a displacement analysis in undrained con-


ditions (Method A or B see Section 4.2.4), or in a coupled consolidation analysis
(Section 4.4) when prediction of the variation of pore pressure with time is needed. It
is generated in the analysis model by small volumetric strains in the virtually incompres-
sible pore water but the accuracy of excess pore pressure predictions depends to a large
extent on the accuracy of the constitutive model (see Section 4.2.5). At typical loading
rates in the eld, excess pore pressures occur only in low-permeability soils (e.g. clays)
and would always be zero in free-draining soils (e.g. sand and gravel). Only at very high
loading rates, e.g. seismic loading, might excess pore pressure occur in high-permeability
soils such as sand.

Excess pore pressures dissipate to zero during consolidation until steady-state conditions
are restored (see Figure 4.1).

Transient pore pressure


Sometimes there are not steady-state conditions, for example with a tidal variation
of groundwater level, or during a pumping test or rapid drawdown of a reservoir.
Consequently, the groundwater pressure is changing with time due to changing hydraulic
boundary conditions and the steady-state pore pressure becomes transient pore pressure.
This is distinct from the temporal dissipation of excess pore pressure occurring during
consolidation as described in the preceding paragraph.

To handle transient pore pressures in an FE analysis, either extreme steady states can be
considered (e.g. in tidal conditions adopt hydrostatic pore pressure with the phreatic
level at the highest and lowest astronomical tides) or time-dependent changes in ground-
water ow and pore pressure in response to time-dependent hydraulic boundary
conditions can be predicted using transient groundwater ow analysis. The predicted
values are therefore transient rather than steady-state pore pressures. It is preferable
to perform such analyses in a separate stage to displacement analyses to avoid unnecess-
ary complexity, but in exceptional cases a transient groundwater ow analysis can be
coupled with a displacement and consolidation analysis in order to predict the time-
dependent ground displacement caused by temporal changes in hydraulic boundary
conditions.

Total or active pore pressure


This is the actual value of pore pressure at a point at a particular time and is the sum of
the steady-state and excess pore pressure (see Figure 4.1). In the case of a transient
groundwater ow analysis, the total pore pressure would be the sum of the transient pore
pressure and excess pore pressure. Where possible, only one of these pore pressure terms
should be changing in each analysis stage, to avoid unnecessary complexity. In excep-
tional cases, they may vary simultaneously.

108

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

4.2. Drained and undrained analyses


4.2.1 What do the terms drained and undrained mean?
Undrained means no water ow, so the volume of voids in a saturated soil remains
unchanged and the soil density remains constant. Since water is essentially incompres-
sible in comparison with the soil skeleton, any imposed loading is transferred directly
to the pore water (as excess pore pressure) and the effective stress remains constant, while
any attempt to impose volume change is resisted by excess pore pressure changes.

In an undrained analysis the excess pore pressure is not allowed to dissipate. Although,
in reality, consolidation and dissipation of excess pore pressure starts immediately on
loading, it is sufciently slow in some cases for undrained conditions to be a reason-
able assumption for eld problems. These cases are where the rate of loading is high
relative to the soils permeability, which occurs in most construction activities in stiff
clays, or during particularly high rates of loading (e.g. earthquake accelerations) in any
soil.

Drained means that water is free to ow through the voids of a soil such that void
volume changes can occur and no excess pore pressures develop. Any loadings result
in total stress changes that equal effective stress changes. These conditions occur when
the rate of loading is low relative to the soils permeability, which occurs in most con-
struction activities in sands and gravels.

Modelling undrained conditions in FE analysis is not without its pitfalls, as described in


Section 4.2.4. Modelling drained conditions is more straightforward.

4.2.2 When are drained or undrained assumptions appropriate?


In this section, rate of loading is taken to also include rates of unloading, hydraulic
boundary condition change and shear strain. If the rate of loading is sufciently slow
relative to soil permeability that no signicant excess pore pressures are generated, then
a drained analysis is appropriate, as summarised in Figure 4.2. For most practical cases,
this encompasses construction activities in granular soils such as sand and gravel. It does
not matter whether short-term or long-term conditions are needed because the output
would be the same.

Vermeer and Meier (1998) proposed that U . 70% or T . 0.40 in order to adopt the
drained assumption, where U = degree of consolidation, T = time factor:

kEoed
T= t (4.1)
gw d 2

where k = permeability, Eoed = one-dimensional stiffness, gw = weight density of water,


d = drainage path length, t = construction time being simulated.

If the rate of loading is sufciently high relative to soil permeability that no signicant dis-
sipation of excess pore pressure occurs during the loading itself, then an undrained analysis
is appropriate. For most practical cases, this includes construction activities in stiff clays.

109

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 4.2 Selection of drained, undrained or consolidation analyses

Drained Undrained construction Partial consolidation


construction during construction
Short-term Time-dependent Long-term
outputs outputs outputs

Drained Undrained Undrained Consolidation


analysis analysis analysis analysis
(method A)*
For all time
periods
Consolidation
analysis

* Method A should only be used with an appropriate, advanced constitutive model.


Refer to Section 4.2.5.

An undrained analysis provides outputs for the short-term case immediately on


completion of loading, unloading, hydraulic boundary condition change or shear
straining.

Vermeer and Meier (1998) proposed that an undrained assumption may be appropriate
when U , 10% or T , 0.10 during construction. It is wrong to assume that all clay soils
behave in an undrained manner. In particular, normally or lightly over-consolidated
clays may behave in an almost drained manner since often they are quite thin deposits
and have sandy or silty laminations that shorten drainage path lengths.

If long-term outputs were required from a soil that behaves in an undrained way during
construction, it would be wrong to perform only a drained analysis in an attempt to
obtain the long-term outputs directly. The reason for this is illustrated by the example
of triaxial compression test simulation outputs shown in Figure 4.3. An undrained triax-
ial compression test on a lightly over-consolidated clay under stress control was simu-
lated, which was followed by consolidation to allow excess pore pressures to dissipate.
Then a drained triaxial compression test was simulated on the same clay directly to the
same stress state. On the graph of deviatoric stress against axial strain, the undrained and
drained lines are initially quite close, with the undrained case being slightly stiffer.
However, once yield starts in the undrained specimen as it approaches the failure line,
the lines diverge considerably. This simple example demonstrates the importance of
simulating the correct stress path rather than establishing the nal stress state by a direct
path, and why long-term conditions following undrained or partially drained construc-
tion conditions cannot be simulated simply by performing a drained analysis. Doing so
introduces errors to the outputs and, worse still, could miss potential failure states in
undrained conditions.

110

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

Figure 4.3 Stress paths in the simulation of long-term conditions in low-permeability soils

Deviatoric stress, q
Consolidation
Small difference
e with stress control
e lin
i lur
Fa
Undrained Drained
compression compression

Mean effective stress, p


Deviatoric stress, q

Undrained
compression Consolidation

Large difference
Drained due to yielding in
compression undrained case

Axial strain, a

Therefore, to obtain long-term outputs following undrained construction conditions, con-


struction must be simulated in undrained conditions using Method A (see Section 4.2.4)
such that excess pore pressures are generated, followed by a consolidation analysis (see
Section 4.4) to dissipate excess pore pressures for the long-term case, as shown in
Figure 4.3. Note that the accuracy of Method A predictions is heavily dependent on the
constitutive model adopted for the soil.

In cases where only adequate safety against geotechnical failure of a structure constructed
in undrained or partially drained conditions is being assessed, it might be acceptable to
perform an undrained (Method C) and/or drained analysis, depending on which is critical
(see Figure 4.4), with basic but appropriate constitutive models to check for failure,
remembering that displacement and structural force predictions may be inaccurate.

In cases where the rate of loading falls between the limits of drained and undrained
behaviour, i.e. excess pore pressures develop and then partially dissipate during con-
struction, a consolidation analysis is necessary (see Section 4.4). This tends to occur
when simulating construction activities in normally or lightly over-consolidated clays,

111

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

silts and any mixed soils with a signicant clay content. The rate of loading will need to
be entered into the FE analysis as will the rate of excess pore pressure dissipation (using
soil permeability and hydraulic boundary conditions). Note that a consolidation analysis
would normally include a parametric study of soil permeability due to the high degree of
uncertainty in this parameter.

Where intermediate cases are nearly drained or nearly undrained during construction,
it may be acceptable to undertake a simpler drained or undrained analysis, respectively,
where it is conservative, as described in Figure 4.4 and the following paragraphs.

Unloading (e.g. excavation, cut slope) causes an immediate negative (suction) excess
pore pressure. With time, the soil swells as the excess pore pressure dissipates, effective
stress reduces and failure is approached. Therefore, the long-term drained assumption is
safety critical. For short-term stability during construction, stiff, heavily over-consoli-
dated clays can be assumed to be undrained, but normally consolidated or lightly
over-consolidated clays should usually be assumed as drained even in short-term cases.

Loading (e.g. embankment construction on a clay foundation) causes an immediate


increase in excess pore pressure. With time, the soil consolidates as the excess pore
pressure dissipates, effective stress increases and the soil moves away from failure. There-
fore, stability increases with time and undrained short-term stability is critical.

However, clay ll for embankments may have negative (suction) excess pore pressure on
placement, particularly in plastic clays. The effective stress would decrease with time and
the soil would approach failure. Therefore, the long-term drained case may be critical for
the stability of clay ll in embankments.

These unloading and loading cases so far refer to mean total stress changes. Deviatoric
stress changes also cause excess pore pressure, and beware of cases where the deviatoric
stress governs the sense of excess pore pressure. Normally consolidated and lightly

Figure 4.4 Safety-critical drainage conditions for constructions in low-permeability soils

Unloading problems
Long-term (drained) case
critical except perhaps
in very soft normally
consolidated clays

e.g. basement excavation e.g. cut slope

Loading problems
Short-term (undrained)
case critical except
perhaps in heavily
over-consolidated clays
e.g. foundation e.g. embankment construction

112

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

over-consolidated clays generate positive excess pore pressure during undrained shear
while heavily over-consolidated clays generate negative (suction) excess pore pressure
and there is a sliding scale between these extremes of over-consolidation. Therefore, very
soft normally consolidated clay may generate sufcient positive excess pore pressure
from undrained shear due to unloading to overcome the negative excess pore pressure
generated from the reduced mean total stress. In such a case, the factor of safety on
stability may actually be lower in the short term. Similarly, a heavily over-consolidated
clay may generate sufcient negative excess pore pressure from undrained shear during
loading to overcome the positive excess pore pressure arising from the increase in mean
total stress. In such a case, stability would be more critical in the long term.

How well an FE model can predict excess pore pressure in the face of these conicting
inuences depends to a large extent on the constitutive model adopted for the soil. In
most cases there will be a high degree of uncertainty regarding excess pore pressures,
so careful parametric studies and other validation exercises are required to assess the
reliability of outputs.

Where a ground model contains some layers requiring drained analysis and some layers
undrained analysis, then clearly the FE analysis should be run with the appropriate
assumption for each layer concurrently. Where some layers require consolidation analy-
sis and some require drained or undrained analysis, there are two options:

1 The layers for drained or undrained analysis are composed of non-consolidating


elements (displacement degrees of freedom only) while the layers for consolidation
analysis are composed of consolidating elements (displacement and pore pressure
degrees of freedom) with appropriate hydraulic boundary conditions at interfaces
with other non-consolidating layers (see Section 4.3.3).
2 Perform a consolidation analysis for all soil layers with high permeability for the
drained layers and low permeability for the undrained layers.

4.2.3 How is drained analysis performed?


In drained analysis, the pore water does not contribute to soil stiffness because it is
assumed free owing. The engineering behaviour of the soil is governed only by the
mechanical properties of the soil skeleton (Figure 4.5). Consequently, no excess pore
pressure is generated and total stress changes equal effective stress changes. Since the
pore water contributes no stiffness, a drained analysis is also appropriate for dry soils.

Alternatively, drained behaviour can be simulated in a consolidation analysis (see


Section 4.4) by allowing excess pore pressures to dissipate to insignicant levels.

Constitutive models should be formulated in terms of effective stress using drained


parameters.

4.2.4 How is undrained analysis performed?


There is a higher potential for error with undrained analyses, so increased caution and
checking of results is required.

113

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 4.5 Methods of simulating drained and undrained soil behaviour

Dry or drained soil Undrained (A & B) Undrained (C)


V V V
V V V
(small) (small)

u = 0, = and u calculated = 0, = u
V V V
=
V K V Kw V Ku
K = bulk modulus Kw = bulk modulus of Ku = undrained
of soil skeleton water 100 to 1000K bulk modulus
Constitutive model in Constitutive model in Constitutive model in
terms of effective stress terms of effective stress terms of total stress
Common input: Common input: Common input: Eu, vu
E, v, E, v, Kw, or cu (= 0.495 to 0.499), cu

There are three methods of simulating undrained soil behaviour (not including consoli-
dation analysis with a short time interval):

Method A (effective stress analysis)


A high value of bulk modulus Kw for the pore water is added into the stiffness of the soil
so that volumetric strains are small and excess pore pressures are generated (Figure 4.5).
Kw is either entered manually or calculated automatically depending on the software,
while all other model parameters, including shear strength, are entered as drained effec-
tive stress values. This method has the advantage of providing outputs of excess pore
pressure, but these are only likely to be reasonably accurate when using appropriate,
advanced constitutive models. It should also take account of changes in soil behaviour,
in particular undrained strength, due to preceding construction stages because undrained
strength is continuously formulated in terms of effective stress. However, the accuracy of
the formulated undrained strength is dependent on the computation of excess pore
pressure and is often detrimentally affected by a lack of effective stress testing data (see
Section 4.2.5).

Method B (effective stress analysis, specified undrained strength)


This works in the same way as Method A, except that the shear stress is limited by the
specied undrained shear strength (and drained shear strength is no longer an input
parameter). This removes the potential for a dangerous over-prediction of undrained
shear strength when using Method A, which is more likely in basic models such as the
LEPP MohrCoulomb model, for which Method B can be more appropriate. However,
excess pore pressure predictions may become highly inaccurate, so Method B should not

114

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

be followed with a consolidation analysis, and advanced constitutive models may lose
some of their features when using Method B. Furthermore, changes in undrained shear
strength due, for example, to consolidation would not be taken into account unless the
undrained shear strength were re-specied in a new model.

Method C (total stress analysis)


Undrained soil parameters are entered directly into the model with undrained Youngs
modulus Eu and undrained Poissons ratio nu being the common stiffness parameters.
Theoretically, nu should be 0.5 for the bulk modulus Ku to become innitesimal, but
to avoid numerical problems a value slightly less than 0.5 is adopted, typically nu =
0.495 to 0.499. There is no separate term for the bulk modulus of pore water, so no
excess pore pressure is calculated. The undrained shear strength is also entered directly.

Method C is appropriate for basic constitutive models where unrealistic conditions


might otherwise be predicted with Methods A or B. It is not suited to advanced soil
models, except those formulated in terms of total stress. One disadvantage is that any
undrained strength changes occurring due to consolidation would not be accounted for
because consolidation is not modelled, so these would need to be specied by the user.

Consolidation analysis with short time interval


An alternative method of simulating undrained behaviour is a coupled consolidation
analysis (see Section 4.4) with a time interval short enough for the dissipation of excess
pore pressure to be insignicant. This is equivalent to Method A or B, depending on
whether a drained or undrained shear strength is specied in the consolidation analysis.
Remember, however, that specifying an undrained shear strength in a consolidation
analysis has the drawback that the strength is not updated automatically due to the
effects of consolidation and the prediction of excess pore pressure is likely to be
inaccurate.

Bulk modulus K
The bulk modulus K of soil grains is about 30 times greater than K of water, so the
change in volume of soil grains is assumed to be zero. This is distinct from K of the soil
skeleton which arises from rearrangement of the soil grains. In drained soil, pore water is
free to ow so it possesses zero bulk modulus, so K governs volumetric strain and stres-
ses are carried by the soil skeleton.

The bulk modulus of undrained saturated soil is governed by Kw of the water phase only,
because the pore water cannot ow and because Kw is signicantly higher than K of the
soil skeleton. Therefore, in Methods A and B and in coupled consolidation and ground-
water ow analyses, Kw is set to a value signicantly higher than K for the accurate
simulation of undrained behaviour but less than the true Kw of water to avoid numerical
problems. Some programs set Kw automatically, otherwise Kw needs to be set by the user.
A value of 100 to 1000 times K is appropriate and outputs are not particularly sensitive
to values within this range.

In Method C, Kw is set to zero and the high bulk stiffness of the undrained soil mass is
set indirectly by specifying a Poissons ratio of 0.495 to 0.499.

115

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Undrained shear strength cu


cu (or su ) is not a fundamental parameter but depends on stress state and stress history
and typically changes during construction activities. In Method A it is calculated by the
constitutive model and is not an input parameter, so can take account of changes in the
stress state. In Methods B and C it is an input parameter that needs to be appropriate for
the stress history and stress state at any point during the analysis.

The dilation angle should be set to zero when using Method A otherwise failure can be
prevented due to the continuous generation of negative (suction) excess pore pressure,
and hence increased effective stress, during shear.

4.2.5 Why is the prediction of cu in Method A often inaccurate?


The prediction of excess pore pressure is critical to predicting accurate effective stress
paths and failure at realistic values of cu . The problem is that only advanced models
appropriate for the soil and model conditions are capable of producing reasonably
accurate predictions of excess pore pressure in response to changes in mean total stress
and, in particular, deviatoric stress. Basic models can generate highly inaccurate stress
paths, leading to the calculation of erroneous cu values based on effective stress
parameters.

Figure 4.6 shows the prediction of undrained shear strength in a lightly over-consoli-
dated clay by Method A for a purely deviatoric loading. The LEPP model with
MohrCoulomb failure criterion is a basic model with no hardening properties, so
predicts a vertical stress path in undrained deviatoric loading. No excess pore pressure
is generated, so effective stress is over-estimated leading to dangerous over-prediction
of the undrained shear strength when the stress path reaches the effective stress failure
line.

Figure 4.6 The prediction of cu in a lightly over-consolidated clay using Method A (Mansikkamaki,
2015)

Deviatoric stress, q
Failure line
(effective stress)

Range of predicted
undrained strengths by Actual undrained strength
LEPP MohrCoulomb
different hardening model stress path
models and parameters

Undrained domain
of different
hardening models
and parameters

Mean effective stress, p

116

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

Hardening models predict the excess pore pressure that lowers the estimate of undrained
shear strength, but this is not an exact science. Different models and parameters will
produce a range of possible stress paths around the true stress path, as indicated by the
grey shaded area.

Method A also predicts cu changes due to consolidation, but these might also be wrong if
the wrong stress path is predicted. A further potential source of error is the unavailability
of effective stress parameters for low-permeability soils due to the expense of obtaining
them. Simulation of the undrained tests (see Section 3.4.2) can be used to back-calculate
effective stress parameters. Further test simulations should be performed to check the
generation of cu for similar stress state, stress history and stress path to the conditions
simulated in the main analysis model, e.g. plane strain, axisymmetric, loading/unload-
ing, primary loading, etc.

The dilation angle should always be set to zero in undrained Method A analyses to avoid
the negative (suction) excess pore pressures being generated that can cause undrained
shear strength to be over-estimated.

To help identify any errors, always check that the output of deviatoric stress is less than
the estimated value of prevailing undrained shear strength:

s1 s3 2cu (4.2)

as demonstrated for generic foundation and retaining wall examples in Figure 4.7 and in
the example in Section 8.3.4. In both examples, the shaded areas indicate where a higher
shear strength has been mobilised in the FE analysis than should be available according
to the strength prole shown. In which case, the constitutive model parameters, or the
constitutive model itself, should be revised until the mobilised strength everywhere is less
than or equal to the strength prole. The strength prole should be based on site inves-
tigation data, but should also take account of any consolidation or hydraulic boundary
condition changes since the site investigation was undertaken. It may be a best estimate
of undrained shear strength, or a characteristic or design line, as appropriate for the aims
of the FE analysis.

Figure 4.7 Checking mobilised undrained shear strength in undrained (Method A) analysis

0 10 20 30 40 50 cu: kPa

>2cu 40
40
>2cu
60 40
30
Contours of 1 3 Best estimate,
characteristic Contours of 1 3
or design line
Depth

117

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

4.3. Groundwater ow analyses


4.3.1 What types of groundwater flow analysis are performed?
In groundwater ow (or seepage) analyses, the soil skeleton is assumed rigid so only pore
pressure degrees of freedom are considered at the nodes and seepage equations are used
in the FE analysis. Hydraulic boundary conditions are used to dene the problem (see
Section 4.3.3).

The continuity equation must be satised at all locations, which means that volumetric
ow rates into an element must be the same as volumetric ow rates out of an element
(plus any sources or sinks). The relationships between ow rates in saturated soil and
permeability in the three axis directions (i.e. Darcys law) are substituted into the
continuity equation. In cases of the same permeability in all directions, this equation
becomes Laplaces equation as used in other elds of engineering.

Steady-state analysis
The hydraulic head and soil permeability remain constant everywhere with time.

Transient analysis
The hydraulic head changes with respect to time in order to model, for example, seasonal
or tidal variations, establishing steady-state conditions (e.g. initiating dewatering or a
pumping test, rapid drawdown). Boundary conditions are dened as a function of time.

Coupled flow and displacement


To avoid unnecessary complexity, it is usually preferable to perform a groundwater ow
analysis to establish the steady-state pore pressure distribution separate to a subsequent
analysis stage of displacement or consolidation analysis. In some cases, more usually
involving transient groundwater ow analysis, it is required to predict both displacement
and pore pressure changes due to time-dependent changes in hydraulic boundary con-
ditions, e.g. to predict displacement and stability of a reservoir embankment during
rapid drawdown. In such cases, a groundwater analysis and displacement or consolida-
tion analysis are coupled in the same analysis stage. In such cases, the total pore pressure
is calculated directly, then excess pore pressures can be calculated at the end from the
steady-state pore pressure. In some programs it may also be possible to include a void
ratio or effective stress-dependent permeability in order to simulate the effect of volume
change on permeability and, in turn, its effect on groundwater ow.

4.3.2 Why is it difficult to simulate unconfined flow?


Unconned ow occurs where the phreatic surface forms a owline. A common example
is seepage through an embankment dam. These cases are more difcult because the FE
analysis needs to predict the position of the phreatic surface. Above the phreatic surface,
permeability reduces because the soil becomes partially saturated and the water has a
smaller volume through which to ow. So, one method employed in some programs
involves reducing the permeability of the soil when specied threshold pore pressures
(typically zero or small suction values) are reached. On the compressive side of the
threshold, saturated permeability is adopted in the analysis, while on the tensile side a
signicantly reduced permeability is adopted. Some programs include unsaturated

118

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

groundwater ow analysis, in which case the soil-water characteristic curve (SWCC)


describing the relationship between saturation ratio and suction needs to be obtained,
as does the relationship between permeability and saturation ratio. Such relationships
are not straightforward to obtain accurately for soils, so some parametric study is necess-
ary to assess the likely variation of outputs within permissible ranges of seepage analysis
input parameters. Refer to, for example, Fredlund et al. (2012).

4.3.3 What do the different hydraulic boundary conditions mean?


Hydraulic boundary conditions affect excess pore pressure in consolidation analyses (see
Section 4.4) and steady-state or transient pore pressure in groundwater ow analyses.
Either prescribed ows or changes in pore pressure can be specied and they can be
constant (for steady-state groundwater ow analysis and consolidation analysis) or
time-dependent (for transient groundwater ow analysis). They are not used in the
drained and undrained analyses described in Section 4.2. Most programs adopt default
boundary conditions at mesh boundaries, so the user must be aware of these in order to
decide where user-dened conditions would be more appropriate.

Closed, no-flow or impermeable boundary


Closed boundaries allow zero ow across the boundary in either direction and no dis-
sipation of excess pore pressure. Interface elements may also form this boundary at
impermeable structures within a mesh when structural elements are otherwise assumed
permeable by default.

A closed boundary is usually adopted at the bottom boundary of a mesh, unless perhaps to
allow dissipation of excess pore pressure to a hard but permeable layer represented by the
bottom boundary. Also, a vertical boundary to a mesh that forms an axis of symmetry
should be closed since no horizontal ow should occur across the axis of symmetry.

Open, permeable, free-draining or seepage boundary


An open boundary allows unrestricted ow across a boundary in either direction. This is
the usual ow boundary condition at the vertical and top mesh boundaries (except axes
of symmetry). If the top surface of a mesh is submerged, then the water pressure at an
open boundary is determined by the depth of the overlying external water. If a top
surface is above the phreatic level, then the open boundary becomes a seepage surface
allowing water to ow out through this boundary if the water level coincides with this
open downstream boundary during the analysis. In which case, water pressure would
be zero but the elevation head would vary in non-horizontal surfaces, so a seepage
boundary is not necessarily an equipotential or a streamline.

In consolidation analyses, excess pore pressure is zero at an open boundary and the total
pore pressure equals the steady-state value.

Infiltration/extraction boundary
An inowing or outowing unit volume per unit time (per metre width in the out-of-
plane direction in plane strain analysis) is specied on an inltration or extraction line
boundary specied in the geometry, e.g. a vertical line for an extraction well or the

119

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

ground surface to simulate rainfall inltration (where soil has sufcient permeability to
absorb the water continuously, otherwise a conditional rainfall/evaporation boundary
would be required). Rather like distributed loads, these ow rates are converted by soft-
ware into equivalent nodal ows.

Drain
A drain is like a seepage boundary except that it is placed inside the mesh. In con-
solidation analyses, the excess pore pressure is zero and the total pore pressure equals
a specied head value or the steady-state pore pressure. Such boundary conditions
are commonly used to simulate vertical wick drains installed in a soft clay foundation to
hasten consolidation during embankment construction, as illustrated in Section 8.4.

Sources/sinks
A source (inow) or sink (outow) is a ow rate applied at a node. In 3D FE analysis,
this represents a point source or sink while in 2D FE analysis it simulates ow in or out
of an innite line perpendicular to the plane of the analysis.

Precipitation/evaporation
This is a conditional boundary condition. For precipitation, specied inow occurs up to
a specied threshold total pore pressure at which this prescribed pressure takes over and
no further ow occurs. If the pore pressure is more tensile than the specied threshold,
then the specied inow occurs. Similarly, for evaporation, the specied outow occurs
until the pore pressure becomes more tensile than a specied threshold total pore
pressure, at which point the pore pressure is xed at the threshold value.

Prescribed head or pressure


Total pore pressure, head or excess pore pressure can be the nodal degree of freedom in
FE analysis software and the boundary conditions need to match. Programs allow users
to prescribe head or pressure on lines, or pressure gradients on areas and volumes, and
the software interpolates the values at nodes. Normally, the user can choose between
specifying an incremental change in pressure or the accumulated value.

Some examples of hydraulic boundary conditions applied to common groundwater ow


and consolidation analyses are shown in Figure 4.8. Note that where a phreatic surface is
used to dene the steady-state pore pressure, hydrostatic conditions are assumed below
this surface and no groundwater ow analysis is performed. However, some programs
also use the phreatic surface as a tool for dening the prescribed head on model bound-
aries for groundwater ow analyses. In such cases, check that the prescribed head or pore
pressure on the model boundaries has been dened correctly.

4.4. Consolidation analysis


4.4.1 When is a consolidation analysis necessary?
In low-permeability soils, excess pore pressure dissipates slowly. Therefore, soil volume
change and effective stress change occur over time, even when loadings and hydraulic
boundary conditions are constant. This means that the strength and stiffness of the soil
are also changing with time.

120

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

Figure 4.8 Typical hydraulic boundary conditions in common applications

Closed
(impermeable wall)
Open
Open with prescribed Open
head or pressure
Open with Open with
Open with prescribed prescribed head prescribed head
head or pressure, or or pressure Open with or pressure
closed if plane/line prescribed head
of symmetry or pressure
Closed* Closed*

Supported excavation Embankment dam

Closed
(impermeable wall)
Open with
Extraction line
prescribed head
or pressure
Open
Closed
(plane/line of Open Open
symmetry) Vertical drains
Closed* Closed*

Cofferdam Embankment construction


(consolidation analysis)

Line/plane of
symmetry
Open
*Can be set to open when significant flow or
dissipation of excess pore pressure is expected
Closed Open across the bottom boundary, e.g. to a porous
rock forming the bottom boundary.

Closed*

Shallow foundation
(consolidation analysis)

Where construction activities occur in undrained conditions, they can be simulated with
an undrained analysis, but a consolidation analysis is required to simulate the sub-
sequent dissipation of excess pore pressure in order to predict longer term behaviour.
Where construction activities occur in partially drained conditions, a consolidation
analysis is required at all stages, as described in Figure 4.2. Typical problems where a
consolidation analysis may be necessary (all in low-permeability soils) include:

g deformation of soft soil


g to compare outputs with site monitoring data inuenced by consolidation (e.g. the
redistribution of forces in retaining walls for excavations in clay)
g progressive failure
g embankment stability during construction (see example in Section 8.4)
g long-term predictions following construction in undrained conditions (to dissipate
excess pore pressure following an undrained Method A analysis) (see example in
Section 8.3)
g construction activities in partially drained conditions, i.e. excess pore pressures are both
generated and dissipate signicantly during construction time, typically in normally
and lightly over-consolidated clays or mixed soils with a signicant clay content.

121

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

The dissipation of excess pore pressure depends on soil permeability, rate of loading
(when loading and consolidation occur concurrently see Section 4.4.3) and hydraulic
boundary conditions. Soil permeability is always an uncertain parameter, so where a pre-
diction of consolidation time is required, a parametric study of permeability is necessary
in order to estimate a permissible range of consolidation times.

The types of hydraulic boundary conditions for consolidation analysis are described in
Section 4.3.3. Their location relative to areas of excess pore pressure also affects dissipa-
tion rates since this determines drainage path lengths. Pore pressure xities are specied
in terms of steady-state, excess or total pore pressure, depending on the software.

4.4.2 What is coupled consolidation analysis?


Many of the assumptions of undrained analysis (Methods A and B see Figure 4.5)
apply in a consolidation analysis too. The pore water is assumed virtually incompressible
by adding a high Kw into the soil stiffness, the soil particles are assumed incompressible
while the K denes the bulk modulus of the soil skeleton.

Then the set of equations describing groundwater ow (Section 4.3.1) and the set of
equations describing equilibrium as used in displacement analysis are integrated in a
coupled fashion to simulate time-dependent ow of water and volume change. Nodes
have both displacement and pore pressure degrees of freedom with both hydraulic and
displacement boundary conditions required to dene a problem.

Construction stages can be dened in terms of a period of time over which the dissipa-
tion of excess pore pressure is predicted by the FE analysis. This is particularly impor-
tant when dening loading rates in a construction stage. Alternatively, a target minimum
excess pore pressure or degree of consolidation can be specied in order to calculate the
time taken to reach the target.

To integrate the equations over time, consolidation is broken down into individual time
steps. These time steps need to be small enough for an accurate solution but if too small,
large uctuations in the calculated excess pore pressures can result. Suitable time steps
vary from a few seconds for laboratory test simulations to several days for eld obser-
vations. Many programs have automatic time step control. Where only manual setting
is available, an initial choice is made and outputs of excess pore pressure examined. If
large uctuations occur, time steps can be changed by an order of magnitude up and
down until satisfactory outputs are obtained. As time passes, dissipation rates slow so
time steps are progressively increased.

4.4.3 Can loading and consolidation be performed in the same stage?


Generally, it is preferred to simulate the development of excess pore pressure during
loading, unloading or a change in hydraulic boundary conditions in one analysis stage
and then to simulate the dissipation of excess pore pressure in a subsequent analysis
stage with the appropriate pore pressure boundary conditions to avoid the complexity
of simultaneous application of both types of boundary condition. In such a case, the rst
stage must be an undrained analysis (Method A) and not Method B because excess pore

122

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are groundwater effects taken into account?

pressure predictions can be highly inaccurate, and not Method C either because no
excess pore pressures are determined.

However, both loading and consolidation can be performed in the same stage (where
steady-state pore pressure remains unchanged but excess pore pressure varies). This is
needed to simulate construction activities where excess pore pressures are generated but
also partially dissipate within the construction time. In this case, the time period needs to
match the construction time so that the loading rate is approximately the same as that
occurring in the eld. In some cases, it may be conservative to assume undrained con-
ditions during construction (see Figure 4.4) if this is preferred.

It is also possible for groundwater ow, consolidation and displacement all to be


analysed together (in which case volume changes are based on the change in total pore
pressure) but it may not be possible to distinguish accurately between excess and steady-
state pore pressure changes. Such coupling is useful, for instance, in the analysis of rapid
drawdown of a reservoir to predict deformation due to consolidation and stability.
However, near failure, convergence of the calculation may be difcult to achieve.

REFERENCES
Fredlund DG, Rahardjo H and Fredlund MD (2012) Unsaturated Soil Mechanics in
Engineering Practice. Wiley, Hoboken, NJ.
Gens A, Sanchez M and Sheng D (2006) On constitutive modelling of unsaturated soils.
Acta Geotechnica 1(3): 137147.
Mansikkamaki J (2015) Effective stress nite element stability analysis of an old railway
embankment on soft clay, PhD thesis, Tampere University of Technology, Finland.
Vermeer PA and Meier CP (1998) Stability and deformations in deep excavations in cohe-
sive soils. Proceedings of the International Conference on Soil-Structure Interaction in
Urban Civil Engineering, Darmstadt, October.

123

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.125

Chapter 5
How are geotechnical structures
modelled?

5.1. Structural geometry


5.1.1 What types of elements are used for structures?
There are several element types available to model structures, each of which works in a
different way (see Table 5.1). It is important to understand how each element type works
in order to select the most appropriate element. In this section, a brief description of the
common element types is provided together with typical applications. The pros and cons
of using continuum elements for structures are covered in Section 5.1.2.

Spring element
The most basic type is the spring element, which connects one node to a xed point
thereby providing the node with some stiffness against translation in a particular
direction, or connects two nodes in the model together providing an axial stiffness
between them. It can extend or compress according to the displacement of the nodes
connected to it and generates a compressive or tensile axial force according to the
spring stiffness specied for it. For a linear spring, the force simply equals the spring
stiffness k times spring extension or compression ( f = ku ). The k value for a strut or
cable, which has units force/displacement (typically kN/mm), can be calculated from
Equation 5.1.

EA
k= (5.1)
L

where E is the Youngs modulus of the material, A is the cross-sectional area of the strut,
cable, etc. and L is the length of the strut, cable, etc.

The spring element is used for structures that act in axial tension or compression but
require no interaction with the model except at each end. Typical examples would be
a strut supporting a retaining wall or the free length of a ground anchor. If the strut
were supported by other means not included in the analysis model, by a shear wall
for example, then the spring could be connected to a node xed in space and an appro-
priate spring stiffness assigned, as described under Supports to retaining walls in
Section 5.1.4.

125

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Table 5.1 Summary of element types and structural applications

Characteristics Applications

2D/3D spring element Provides an axial stiffness For linear structures without
between two nodes. interaction with the ground
along their length, e.g.
ground anchor free length,
strut support.

2D bar/membrane element Translational degrees of For membrane-like structures


freedom at nodes. Axial interacting with the ground in
stiffness only. Zero bending 2D models, e.g. geotextiles.
stiffness. Approximate modelling of
closely spaced linear
structures, e.g. ground anchor
fixed length, soil nails and
rock bolts.

3D bar element Translational degrees of For linear structures in 3D


freedom at nodes. Axial models where bending
stiffness only. Zero bending stiffness is neglected (but
stiffness. interfaces with the ground are
poorly modelled without
modifications to the model),
e.g. ground anchor fixed
length, soil nails, rock bolts.

3D beam element Translational and rotational For linear structures in 3D


degrees of freedom at models with bending stiffness,
nodes. Axial and bending e.g. capping beam.
stiffness.

2D plate element Translational and rotational For planar structures


degrees of freedom at orientated in the out-of-plane
nodes. Axial and bending direction in 2D models, e.g.
stiffness. embedded retaining wall,
tunnel lining, raft foundation.
3D membrane element Translational degrees of For membrane-like structures
freedom at nodes. Axial in 3D models, e.g. geotextiles.
stiffness only. Zero bending
stiffness.

126

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Table 5.1 Continued

Characteristics Applications

3D plate/shell element Translational and rotational For planar structures in 3D


degrees of freedom at models, e.g. embedded
nodes. Axial and bending retaining wall, tunnel lining,
stiffness. raft foundation, shear wall.

2D continuum element Translational degrees of For all structures interacting


freedom at nodes. with the ground orientated in
the out-of-plane direction in
2D models, e.g. embedded
retaining wall, tunnel lining,
raft foundation, also closely-
spaced structures in the out-
of-plane direction.
3D continuum element Translational degrees of For all structures interacting
freedom at nodes. with the ground (see Section
5.1.2).

Bar element
The bar element (also called the truss element when there are no mid-side nodes), rather
like the spring element, can only sustain axial load. However, it can be curved and con-
nected to the mesh at other nodes along its length so that it changes shape and deforms
in response to deformations around it. The axial stiffness is dened in terms of the material
Youngs modulus and section area EA. Since it has no rotational degrees of freedom and
stress and strain are assumed constant across the section, it can have no bending stiffness
and forms pinned connections with other elements. It can only resist bending to the extent
that any bending mobilises tension in the bar element (with updated coordinates see
Section 1.4.3). Bar elements could be used to represent linear ground structures, e.g.
ground anchors (xed length), soil nails and rock bolts, in 3D analyses, if the bending
stiffness of these structures need not be considered. The drawback is that the element has
no surface area so it models groundstructure interfaces poorly, although some programs
have the facility to establish an elastic zone of equivalent structure volume in the ground
to obtain more realistic interface behaviour (see Section 5.1.2). Linear ground structures
can also be modelled using bar elements in 2D analyses but only if the structures are
closely spaced and, even then, only to an approximate degree (see Section 5.1.5).

127

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Bar elements can be used to represent struts and cables, as can spring elements, but bar
elements have the advantage that their engineering behaviour is specied using a con-
stitutive model that can incorporate elastic-plastic behaviour and, in 2D axisymmetric
analyses, the signicant resistance provided by hoop forces is taken into account. The
use of bar elements to represent planar structures, e.g. geotextiles, is described under
Membrane element.

Beam element
The beam element is similar to the bar element except that it has rotational as well as dis-
placement degrees of freedom and stress and strain can vary across the section. Therefore,
it can model the rotational, or bending, stiffness of a structure as well as axial stiffness,
dened as EI and EA, respectively, where I is the second moment of area of the section.

The line of the beam element itself represents the central axis, or locus of centroids of
transverse cross-sections. The neutral axis (where bending stresses are zero) coincides
with the central axis in straight beams but not in curved beams. There are two common
beam bending assumptions that may be adopted for beam elements:

g EulerBernoulli theory (equivalent to Kirchhoff theory for plates and shells):


transverse planes remain normal and at after deformation, so no transverse shear
deformation occurs. Suitable for cross-sectional areas very small relative to beam
length.
g Timoshenko theory (equivalent to Mindlin theory for plates and shells):
transverse planes remain at but can rotate away from the normal during
deformation, so transverse shear deformation can occur. This is suitable for larger
cross-sectional areas relative to beam length.

For more complex behaviour, continuum elements should be used.

In 3D analyses, beam elements may be used to represent linear structures such as


capping and waling beams to a retaining wall, or other linear structural features. The use
of beam elements in 2D analyses is described under plate and shell elements.

Membrane element
Membrane elements are the 3D equivalents of bar elements and, as such, have three
coordinates per node, are planar and have no volume. They have translational but no
rotational degrees of freedom, so can deform, extend and sustain tension and com-
pression (if allowed) in directions within their plane but have no bending stiffness. Their
most common use is in simulating planar soil reinforcement such as steel strips in
reinforced earth walls and geotextiles, in which case they are set to resist only tensile
forces. In 2D plane strain and axisymmetric analyses, the membrane is equivalent to the
bar element, resisting forces in its plane, including in the out-of-plane direction.

Plate and shell elements


Plate and shell elements are planar like membrane elements, but they have rotational as
well as displacement degrees of freedom at their nodes so that they possess bending

128

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

stiffness dened in terms of EI. The difference between plates and shells is rather subtle
and there are different denitions in the literature. A plate element is sometimes dened
as having bending stiffness but no axial stiffness (i.e. no membrane action) while the shell
element has in-plane membrane behaviour coupled with bending behaviour so called
because they are intended for shell structures where membrane behaviour dominates
over bending. Nonetheless, some programs do contain plate elements with both axial
and bending stiffness. A plate element is sometimes considered as being at while a shell
element is curved but, again, this denition does not hold everywhere. For the modelling
of geotechnical structures it is essential to have both axial and bending stiffness. Plate
and shell elements are used to represent structural walls, such as embedded retaining
walls and shear walls, tunnel linings and spread foundations where the bending stiffness
needs to be simulated.

There are some assumptions associated with plate and shell elements which should be
valid whenever they are used. They include:

g thickness small compared with length and width


g bending deection small compared with thickness
g the mid-surface is the neutral surface during bending
g stresses normal to the mid-surface are small compared with the bending stresses
and are assumed constant or zero through the thickness (plane stress condition)
g linear strain distribution with depth.

Two common plate bending theories exist which differ in the way they calculate out-of-
plane displacement:

g Kirchhoff theory: out-of-plane normals remain straight and normal to the surface,
so any point on the mid-surface during bending only deects in the normal
direction to the undeformed surface and no shear deformation occurs. This is
suitable for very thin plates.
g Mindlin theory: out-of-plane normals remain straight but can rotate relative to
the surface, so shear deformation can occur. This is suitable for thicker plates,
which includes most geotechnical structures. It can also be used for thin plates
but accuracy deteriorates when plates become very thin due to numerical ill-
conditioning.

For more complex behaviour, continuum elements should be used, particularly for thick
structures and for detailed analysis around connections.

When changing the material properties of a plate or shell element already installed in an
earlier construction stage, the same ratio of bending to axial stiffness EI/EA must be
maintained, otherwise the resulting change in effective section depth would cause an
unrealistic change in bending moment.

Plate and shell elements in 2D analyses are equivalent to beam elements except that they
are formulated in terms of plane strain and axisymmetric versions of 3D plate theory.

129

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Stresses are generated in the out-of-plane direction (due to the prevention of strains in
this direction) which inuence the behaviour of plate and shell elements within the
analysis plane via the Poissons ratio. Hence, for structures that are relatively exible
in the out-of-plane direction, e.g. sheet pile walls, Poissons ratio should be set to zero
in isotropic material models. Anisotropic (orthotropic) models allow the stiffness in each
direction to be specied explicitly (see Section 5.1.6).

Continuum elements
Continuum elements are the area or domain elements in 2D or the volume or solid
elements in 3D. This is the element type normally used to represent the soil or rock but
they can be used to represent structures too and there are advantages to this, as covered
in Section 5.1.2. The 3D version is the fundamental continuum element because the three
displacement degrees of freedom are equally weighted in each direction, while the 2D
version results from a simplication to the FE formulation in the out-of-plane direction.
The 2D continuum element can be used in plane strain or axisymmetric analyses to
represent structures with thickness, e.g. an embedded retaining wall, tunnel lining, raft
foundation or, in an axisymmetric analysis, a single pile. The 3D continuum element can
be used to simulate the same structure types in 3D meshes.

5.1.2 Should continuum or non-continuum elements be used for


structures?
Structures that have signicant volume in reality, such as piles, diaphragm walls, piled
walls, raft foundations and tunnel linings, can be modelled either with continuum elements
(volume elements in 3D, area elements in 2D) or with non-continuum (line or surface)
elements, e.g. membrane, shell. The pros and cons of each are summarised in Table 5.2.

Due to their ease of use and tendency to be more conservative in some soilstructure
interaction problems, non-continuum elements are used more widely than continuum
elements in routine analysis work. However, non-continuum elements should only be
used when the structure can be represented in this way with sufcient accuracy. They
have certain assumptions (see Section 5.1.1) and are intended for thin-walled structures,
so as structures become thicker, the non-continuum elements become less suitable. Step
changes in thickness, voids and connections may also become difcult to model.

Structures with only a small volume, such as sheet pile walls, are more suited to simula-
tion with non-continuum elements. Note that moment-reducing effects (Figure 5.2) still
occur in steel sheet piles to some extent due to their corrugated section but will not be
taken into account with non-continuum elements. Refer also to Section 5.1.6 regarding
the anisotropy of sheet pile walls (and other structure types).

5.1.3 How are groundstructure interfaces modelled?


At interfaces between the ground and structures, relative movement can occur, e.g.
settlement of soil immediately behind a deected retaining wall. Relative movement can
also occur at discontinuities in rock. However, the requirement for compatibility of
displacements in an FE analysis prevents separation, overlap or slippage along planes
in a mesh. Therefore, interfaces need to be modelled with dedicated elements, which has

130

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Table 5.2 Factors to consider in the use of continuum or non-continuum structural elements

Continuum elements Non-continuum elements

Setting up Takes longer because full geometry Easier set-up because only planar
needs to be defined and material geometry of structure needs to be
parameters are less straightforward, defined and section properties are
particularly for 2D models. input directly as material
parameters.

Element size The true thickness of structures is The geometry of non-continuum


often small compared with the elements is suited to structures that
overall geometry, which can require are thin relative to the overall
a large number of elements in order geometry.
to avoid high aspect ratios.

Self-weight Use true density. The true volume is not represented,


so the specified self-weight of
structures in the ground should be
the net additional weight of the
structure over and above the
ground weight occupying the true
structure volume in the FE model
(see Figure 5.1).

General accuracy Should be more accurate, Should be reasonably accurate for


particularly in cases of complex simple geometries and where the
geometry or where the assumptions inherent assumptions of the
of non-continuum elements do not elements hold. Often more
hold. The moment-reducing effects conservative than continuum
of interface friction are included (see elements in soilstructure
Figure 5.2). interaction problems because the
moment-reducing effects of
interface friction are not taken into
account (see Figure 5.2).

End bearing The end bearing of axially loaded Negligible end bearing is obtained
structures, e.g. piles and retaining because the element has no contact
walls, is simulated appropriately area with the ground. Plate or shell
because the true contact area with elements may be added at the end
the ground is defined (see to simulate the width of the
Figure 5.3). structure or some programs have
the facility to establish an elastic
zone of equivalent structure width
in the ground to obtain more
realistic end bearing failure loads,
but the input parameters should be
validated carefully (see Figure 5.3).

131

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Table 5.2 Continued

Continuum elements Non-continuum elements

Interface stresses The contact area between structure In 2D analysis, the contact area
on linear structures and ground is accurately becomes continuous in the out-of-
represented so interface friction is plane direction, so specifying
modelled more accurately, interface element friction properties
particularly in axially loaded accurately is more difficult. In 3D
prismatic linear structures such as analysis, beam and bar elements
piles, soil nails, rock bolts, ground have zero contact area, so would
anchor fixed lengths. not simulate interface friction
correctly. Some programs have the
facility to establish an elastic zone
of equivalent structure volume in
the ground to obtain more realistic
interface behaviour with the
ground, but the input parameters
should be validated carefully.

Permeability The permeability can be defined as Structures are treated as permeable


for soil and rock elements. because the ground on each side of
a structure has common nodes.
Interface elements are required
around the structure to separate the
ground nodes on each side and
render the structure impermeable.

Output Obtaining outputs of structural Direct output of structural forces.


forces involves some manipulation
of the output data, e.g. integrating
the output stresses on sections to
obtain axial force, shear force and
bending moment (see example in
Section 8.2.4).

the advantage of allowing the specic properties of the interface to be assigned to those
elements rather than to the elements on either side of the interface.

The interface can be modelled with thin continuum elements, as shown in Figure 5.4,
with appropriate stiffness and strength properties but this requires a lot of mesh rene-
ment to avoid unacceptably high aspect ratios.

Interface (or slip) elements of zero thickness are the element types most commonly used
to simulate groundstructure interfaces. They have the advantage of allowing innitely
thin elements without signicant renement to the mesh.

132

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Figure 5.1 Specifying structure self-weight for continuum and non-continuum elements

Continuum elements Non-continuum


and reality elements

Use true Structure Specify full Structure


External structure self-weight structure self-weight
structure density weight

Structure
element

Internal Use true Specify net


structure structure structure Ground occupying
density Structure weight true structure volume
self-weight

The nodes on each side of an interface element are coincident because, in reality, the
structure and the ground are in contact. Relative elastic shear and normal displacements
are governed by the shear stiffness (Ks ) and normal stiffness (Kn ) of the interface. Deter-
mining these stiffness values is not straightforward because they have different units
(kN/m3) to standard material stiffness values and no parameter testing methods are
available.

Figure 5.2 Influence of interface friction on structures and comparison of modelling approaches

Continuum elements and reality Non-continuum elements

Interface Interface
friction friction

Lever
arm No lever
arm

M0 M=0

133

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 5.3 Accounting for end bearing in structural elements

Continuum elements Non-continuum element options


and reality

Base Structure
elements volume

End No end End End


bearing bearing bearing bearing

Structure elements Added base Automatic stiffening


only elements of continuum elements
within structure volume

Figure 5.4 Options for modelling groundstructure interfaces

No interface modelling Thin continuum Interface


(rigid interface) elements elements

Structure Ground Structure Ground Structure Ground

Nodes
coincident
at interface

Structure Structure Structure


Ground Ground Ground

No relative displacement Thin continuum element Interface elements allow


at interface deforms allowing relative separation and sliding
displacement at interface at interface

134

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

If they are too low, excessive deformation at the interface will be calculated, too high and
the large difference in stiffness between adjacent elements may cause numerical ill-con-
ditioning leading to stress oscillations in the output. However, provided that the order of
magnitude of these values is in the range of the material stiffness either side of the inter-
face, then outputs may not be too sensitive to the values adopted. This should be
checked by a parametric study.

The element thickness also plays an important role in the stiffness of the element.
Geometrically, the element has zero thickness, but a virtual value is required as a
material parameter for the calculation of elastic deformations. High values result in large
elastic deformations, so the value should be small, particularly for large normal stresses.
However, if the value is too low, then numerical ill-conditioning may occur. Appropriate
values typically lie in the range 1 to 10% of the interface element length but a parametric
study allows appropriate values to be identied.

A Coulomb friction criterion (Equation 5.2) denes the change from elastic to plastic
behaviour in terms of effective stress, while a simple limiting shear stress is used for
undrained analyses in terms of total stress. If the calculated shear stress is less than the
failure stress, then the interface element binds the ground and structure together while
allowing some relative elastic deformation. If the shear stress reaches the failure stress,
then permanent slippage occurs at the interface. Also, where any calculated tensile stress
at the interface exceeds the tensile strength (c tan w ) or a specied tension cut-off, the
interface element allows separation between the ground and the structure and Kn and
Ks reduce essentially to zero. If gap closure occurs in a subsequent analysis stage, the
program needs to record the amount of separation so that compressive stresses are not
restored at the interface until the same amount of separation is reversed and contact at
the interface element is re-established.

t = c + sn tan w (5.2)

The interface shear strength can be measured by, for example, a laboratory direct shear
test. In the absence of test data, interface friction is typically adopted between a half and
the full internal shear strength of soil, depending on the characteristics of the structure
(e.g. material, installation method). If a smooth interface is required, perhaps when
simulating certain laboratory tests, interface elements with very low or zero Ks and shear
strength can be used.

Where relative movements at interfaces are quite small, sufciently accurate outputs may
be obtained without interface elements. This can be checked by running an analysis with
and without interface elements and comparing the outputs. Where relative movements
are quite large, interface elements will certainly be needed and renement of the mesh
around the interface may be required too due to the high stress and strain gradients in
these regions.

Interface elements introduce additional complexity to FE analysis models, so it is worth-


while running an analysis model prior to installing interface elements to check that it

135

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

completes all the analysis stages satisfactorily. In setting up interface element geometry,
particular care is required at connections between structures to ensure that the interface
elements do not change the intended connection type.

5.1.4 How are the common geotechnical structure types modelled?


Much of the guidance on how to model a specic problem is common to different struc-
ture types, such as about using continuum or non-continuum elements, and is provided
throughout this chapter and, indeed, throughout this book. This section provides some
useful tips that are mostly exclusive to specic structure types.

More guidance on particular structure types can be obtained from published case studies
by others who have applied FE analysis to similar problems.

Supports to retaining walls (including ground anchors)


The common supports to retaining walls include struts, slabs and ground anchors. The
stiffness of these supports has a signicant and non-linear effect on outputs of retaining
wall deection and bending moment. Support stiffness is often an uncertain parameter
due to thermal effects, concrete shrinkage and connection details (see Section 5.1.7), so a
parametric study of support stiffness is often required in FE analyses of supported
retaining walls.

Linear supports can be modelled with spring or bar elements and slabs with plate or shell
elements, as described in Section 5.1.1. Where supports interact directly with the ground,
as shown in Figure 5.5, the entire support should be modelled. When a slab or strut
spans between two identical retaining walls such that an axis of symmetry can be used
as a model boundary (see Section 1.2.4), then only half the slab or strut length need
be modelled reacting against a xed point on the axis of symmetry. In many cases, a slab
or strut is supported by other structural members, such as shear walls and bracing,
before loads are eventually transferred back to the ground through foundations remote
from the retaining wall being simulated (Figure 5.5). Either these additional structural
members and foundations are included in the model or they can be substituted for spring
elements with an equivalent stiffness. The spring element option is more straightforward
for the geotechnical modeller when the structural supports are complex but requires
some analysis by the structural engineer in order to provide appropriate spring stiffness
values. Further iterations between the geotechnical and structural models would also
be required until the deections and loads at the interface between the two models match.

The xed lengths of ground anchors need to be modelled with continuum elements in
order to simulate the anchorground interface more accurately. Alternatively, some
programs adopt non-continuum elements with a function that automatically sets the
ground within the specied volume to the properties of the grout material. This is
acceptable provided that the input parameters have been validated carefully.

A pre-stress in an anchor or strut is imposed as illustrated in Figure 5.6. Axial forces are
applied at the node at each end (or at one end if the strut intersects a model boundary on
an axis of symmetry) in the direction that the pre-stress will impose itself on the

136

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Figure 5.5 Support types

Supports interacting directly with the ground

Full simulation of support structure or equivalent spring supports

More complex structural support

connected structure or ground before activating the anchor or strut. In the subsequent
analysis stage the anchor or strut is activated and the axial forces removed. The rebound
of the connected structure or ground will cause the pre-stress to be transferred to the
anchor or strut elements. The outputs of anchor or strut force should approximately
match the initial applied loads and the applied loads can be adjusted in subsequent
re-runs until the desired pre-stress is achieved. The outputs of anchor or strut force in
later stages are likely to be somewhat above or below the pre-stress due to the effects
of the other construction activities being simulated.

Geosynthetics and reinforced soil/earth walls


Membrane elements are commonly used to represent geosynthetics and the steel strips of
reinforced earth walls, with interface elements placed between the membrane elements
and soil. This method works quite well for steel strips but not so well for geosynthetics,
particularly geogrids. This is because soil sits within the apertures of a geogrid leading to
a different type of interaction between soil and geogrid than simulated by membrane and
interface elements. It is also difcult to select appropriate stiffness and strength values
for polymers because they are heavily dependent on rate effects and creep.

137

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 5.6 Applying pre-stress to ground anchors and struts

1. Install fixed length and 2. Install free length 3. Release loads


apply pre-stress loads (pre-stress transferred to anchor)

1. Apply pre-stress loads 2. Install strut 3. Release loads


to strut connections (pre-stress transferred to strut)

As with all geotechnical FE analyses, the construction sequence needs to be simulated to


obtain realistic stress distributions which, for multi-layered reinforced soil walls, can
mean many construction stages in order to place each layer, although some simplica-
tion may be possible as described in Section 1.4.2.

Simulating the stabilising effect of geogrids is a particular challenge, as described by Lees


(2017). In order to simulate membrane action in geosynthetics, it is essential to use
updated coordinates in the FE analysis, as described in Section 1.4.3.

Spread foundations
Usually a 3D analysis is necessary when simulating spread foundations although 2D
plane strain or axisymmetric models are appropriate for certain geometries and loadings.
The axisymmetric assumption is suitable for circular foundations or as an acceptable
approximation of a square foundation (with the radius set to achieve the same equivalent
foundation area as the square foundation) but only with a vertical, concentric applied
load. The plane strain assumption is suitable for long strip foundations with closely
spaced or continuous uniform loading along their length. Inclined or eccentric loads and
moments are permitted but only acting in the plane of the analysis.

Analyses involve applying either load or displacement to foundations. The former tends
to be used for simulating foundations in service while the latter tends to be used in order
to obtain a failure load (or resistance) for the foundation from the output.

138

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Figure 5.7 Options for idealised simulation of spread foundations

Flexible foundation (load control):


Smooth Rough
FOUNDATION FOUNDATION
Prescribed vertical
displacement
Prescribed
vertical load
Horizontal fixity

Vertical displacement
tied between nodes

Rigid foundation (displacement control):


Smooth Rough
FOUNDATION FOUNDATION

Rigid foundation (load control):


Smooth Rough
FOUNDATION FOUNDATION

When certain assumptions are appropriate (see Section 5.3.1), it is not necessary to
model the foundation itself. If a foundation is assumed perfectly rigid, a uniform dis-
placement can be imposed on the soil nodes where the foundation base would be, or the
nodes tied to displace uniformly in the case of applying a load, as shown in Figure 5.7. If
assumed perfectly exible, the foundation load is applied at the soil nodes where the
foundation base would be. Similarly, a perfectly rough soilfoundation interface can
be simulated without including the foundation itself by constraining the horizontal
displacement of the soil nodes coincident with the base of the foundation while the nodes
at the sides of the foundation can be constrained to displace with the foundation base.
For a perfectly smooth interface, the same nodes would be unconstrained.

139

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Piles
Particular care is needed with axially loaded piles when specifying the properties for the
interface elements between piles and the ground. Pile installation disturbs the ground
immediately around piles such that its properties differ from those of the undisturbed
ground encountered during site investigations. Therefore, input parameters must be
validated by back-analysing pile load tests and the test pile types, loads and ground
conditions must be similar to the piles being simulated in the main model.

Interface elements are also required for laterally loaded piles to avoid tension being
generated where the pile deects away from the ground.

Piles need to be modelled with continuum elements in order to simulate the pileground
interface more accurately, although some programs adopt special beam elements that
take an arbitrary orientation through the 3D solid elements representing the ground.
Interface line elements between the beam nodes and virtual nodes within the surround-
ing solid elements model interaction with the soil and spring elements dene the base
resistance, while plasticity in the ground is disabled around the beam elements within the
specied pile radius (refer to Tschuchnigg and Schweiger, 2015). This method is accep-
table provided that the input parameters have been validated carefully, particularly for
pile groups.

For single piles, a 2D axisymmetric analysis is possible only in the specic case of a
circular section (or a square section approximately if the radius is set to achieve the same
surface area in contact with the soil) pile with vertical orientation, a vertical, concentric
applied load and horizontal soil layers and groundwater level. A pile with a non-vertical
applied load or a moment needs to be simulated with a 3D model.

The constrained conditions created by a loaded pile in an FE analysis can result in


unrealistic raised effective stresses and over-predicted failure loads if the dilation angle
of the soil is set above zero. Therefore, in FE analyses of piles, a dilation angle of zero
or a carefully set dilation cut-off are required.

Clearly 3D models are required for the analysis of pile groups and piled rafts. A consti-
tutive model for the soil that accounts for non-linear strain-dependent stiffness becomes
important for more accurate simulation of pilepile interaction. The validation of pile
group analyses is more difcult because load tests on pile groups are rare, so published
case studies of similar conditions and other analysis methods (e.g. boundary element
method) are more common sources of validation data.

Ground improvement and grouting


Ground improvement can be divided into diffuse methods (where a homogeneous
material is formed, e.g. compaction, consolidation and grouting) and discrete methods
(where inclusions are created that remain separated from the ground, e.g. stone and
concrete columns, deep mixing and jet grouting). Depending on the size and spacing
of inclusions relative to the dimensions of the problem, either each inclusion can be
modelled by FE analysis in a similar fashion to piles, or the benecial effects of the

140

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

inclusions can be smeared across the ground and dened in terms of the improved
properties of the ground, as for diffuse methods.

With the smeared approach, rather than attempt to simulate the ground improvement
activity itself, it is common practice simply to substitute the in situ material for one with
appropriately enhanced density, strength and stiffness. Ground improvement may also
have an impact on ground permeability that would need to be considered in consolida-
tion and groundwater ow analyses.

Obtaining constitutive model parameters for improved ground prior to improvement


can, of course, be difcult. Ideally, eld trials will be performed so that parameters can
be obtained from those. Case studies of similar ground improvement projects also
provide a valuable source of parameters. Post-improvement testing provides a means
to validate input parameters, but only after completion of the construction activity.

Jet grouting, permeation grouting and compaction grouting in soil and consolidation
grouting in rock are all forms of ground improvement whose benecial effects can be
simulated in the same way (i.e. material substitution). Jet grouted structures are brittle
in tension which may require special consideration in the constitutive model (see
Section 5.2.1) in detailed studies of such structures.

Compensation grouting is a form of hydrofracture or soil-fracture grouting intended to


control ground movements. It is commonly used between an advancing tunnel and the
ground surface below existing structures to compensate for stress relief and ground loss
and to control tunnelling-induced settlements. There are two main approaches to simu-
lating compensation grouting: prescribed strain or prescribed pressure, as summarised
by Wisser et al. (2005). The latter and Addenbrooke et al. (2002) describe in detail
prescribed pressure approaches using interface elements that open out to represent the
growing grout body. Soga et al. (2000) adopted a similar approach using continuum
elements to represent the grout body.

Tunnels
A distinction can be made between soft ground tunnelling and rock tunnelling. In soft
ground tunnelling signicant stresses are taken by the tunnel lining and there are larger
deformations, including ground surface settlements in the case of relatively shallow tun-
nelling. In rock tunnelling a greater proportion of tunnel-induced stresses are supported
by the ground and design and simulation is more focused on stress relief, rock quality
deterioration and local instability. The remainder of this section on tunnels will focus
on soft ground tunnelling.

In addition to tunnel lining deformation, the main cause of tunnelling-induced ground


deformations is ground loss. Ground loss is the extra volume of soil that is excavated over
and above the tunnel volume due to stress relief and partial closure of the excavation. It
is expressed as a percentage of the tunnel volume and typical values are 23% for
conventional tunnelling methods without adequate support and 0.5% for modern earth
pressure balance (EPB) tunnel boring machines (TBM), although this can increase to

141

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

about 1% in more difcult ground conditions. Therefore, simulation of both the tunnel
lining and the tunnelling process are critical elements to obtaining sufciently accurate
predictions of ground deformation.

There are three commonly used methods of simulating tunnelling. The rst two
described in the following paragraphs are idealised methods intended for 2D plane strain
models or 3D models where the tunnel is activated in its entirety in a single stage. The
third method is a full 3D simulation of the advancing tunnel.

1 Volume loss control method: soil elements inside the tunnel are deactivated and
the lining elements activated in the same analysis stage. In the next stage, the
volume loss is prescribed, usually by specifying a circumferential contraction of
the tunnel lining. For circular linings the percentage contraction is approximately
half the percentage volume loss owing to the geometry of a circle. The actual
volume loss achieved should then be checked in the outputs.
Since volume loss is an input parameter, this method is better suited to cases
where the volume loss can be determined for a particular tunnelling method and
ground conditions. This method is used more commonly in the simulation of
TBM tunnels.
2 Load reduction method: this is also called the b-method or a-method (Panet and
Guenot, 1982). The soil elements to be excavated are deactivated and an articial
support pressure is applied to the faces of the excavation in the same analysis
stage prior to installation of the lining. The pressure is a proportion (b) of the in
situ stress s0 prior to tunnel excavation. Either the stress from the excavated soil
elements is completely deactivated and the pressure bs0 is applied or (1 b)s0 of
the excavated soil elements stress is deactivated, as is possible in some programs.
Further support pressure reductions can be imposed, if required, when simulating
time effects (e.g. consolidation, creep). In a subsequent stage the tunnel lining is
activated and the remaining support pressure removed.
Here the proportion of unloading prior to lining construction is prescribed and
the volume loss is calculated. The b value is not straightforward to determine it
is based on experience of the tunnelling method in particular ground conditions
and experience of applying the method in FE analysis. Trial values can be used
and the outputs compared with expected outcomes. Higher b values (typically up
to about 0.7 for conservative predictions of structural forces) are appropriate for
earlier lining installation, higher lining forces and less ground deformation. Lower
b values (typically down to about 0.2 for conservative predictions of ground
movements) are appropriate for later lining installation, lower lining forces and
more ground deformation. This method is commonly used in the simulation of
sprayed concrete lining (SCL) tunnels, as well as TBM tunnels.
3 Step-by-step method: this method involves a full simulation of the tunnel
excavation and lining installation steps in a 3D FE analysis, without the idealised
input parameters of the previous two methods. Many analysis stages are required
in order to simulate an advancing tunnel so it is very time-consuming to set up
and to run. Therefore, this method is normally used only for detailed studies
rather than more routine analyses.

142

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

When simulating tunnelling in soft clay, it may be necessary to include grout pressures in
TBM tunnelling (from the injection of grout to ll the annular space between the bore
and the lining the lining would need to be removed temporarily to apply this pressure)
or other support pressures, e.g. compressed air, to prevent ground movements and heave
being over-predicted. These additional pressures can be adjusted to achieve the expected
ground loss or lining forces.

Tunnel linings may be simulated with continuum or non-continuum elements with the
associated advantages and disadvantages described in Section 5.1.2. Jointed circular
tunnel linings can be simulated as continuous linings, but with an effective second
moment of area Ie calculated according to Equation 5.3 (Muir Wood, 1975) to help
account for the additional exibility introduced by the joints.
 2
4
Ie = Ij + I (Ie  I, n . 4) (5.3)
n

where I is the second moment of area of a segment section, n is the number of segments
and Ij is the effective second moment of area at a joint (which is difcult to determine
because it depends on hoop thrust and joint geometry so is often assumed as zero or
close to zero). This approach should obtain a reasonably accurate compressible stiffness
of the lining and, hence, outputs of hoop thrust, and reasonably accurate predictions of
ground movement. The prediction of lining bending moment may be inaccurate because
the effect of the joints is smeared around the lining rather than being considered expli-
citly. Hinges can be considered to simulate joints and obtain more representative lining
bending moment distributions while more accurate outputs can be obtained with springs
placed at joints between segments simulated using shell elements to model the specic
characteristics of the lining joints (refer to, for example, Li et al., 2015, for cast iron
bolted tunnels and Wang et al., 2012, for concrete segmental linings).

The main difculty in simulating SCLs is in the constitutive behaviour. The lining is
loaded immediately on application and as the sprayed concrete (shotcrete) is still curing,
it gains strength and stiffness while performing its supporting role. Shotcrete also
exhibits creep behaviour and post-peak softening in both compression and tension. The
simplest method to model the lining is to use a linear elastic model with an articially low
stiffness that is increased in subsequent stages to account for shotcrete curing (e.g.
Moller and Vermeer, 2005). However, this can over-predict structural forces in the lining
due to high tensile stresses. More advanced constitutive models have been developed to
account for more aspects of shotcrete behaviour, such as Schadlich and Schweiger (2014)
that includes hardening/softening plasticity, time-dependent strength and stiffness,
creep and shrinkage. Further description of the modelling of concrete is given in
Section 5.2.1.

5.1.5 How are structures modelled in 2D with the plane strain or


axisymmetric assumption?
Judging whether a 2D plane strain or axisymmetric assumption is appropriate for a par-
ticular problem was covered in Section 1.2.1. As far as structures are concerned, certain

143

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

geometries are suited to these assumptions. Structures with a long, straight horizontal
dimension and a uniform prismatic section (e.g. a long strip foundation, oor slab or
diaphragm wall) are suited to the plane strain assumption, while similar structures whose
long dimension is circular around a vertical axis of symmetry (e.g. a vertical shaft or
single pile) are suited to the axisymmetric assumption. Deriving the input parameters
and interpreting output for such structures is relatively straightforward.

Similar structures can be composed of a series of identical, closely spaced discrete struc-
tures (e.g. contiguous piles) arranged in a straight line (quasi-plane strain) or in a circle
about a vertical axis of symmetry (quasi-axisymmetric). Deriving the input parameters
and interpreting output for such structures is less straightforward and the primary aim
of this section is to provide guidance in this area.

As the spacing between structures in the out-of-plane direction increases, it becomes


more difcult to justify adopting the plane strain or axisymmetric assumption. With a
small spacing, 2D models are acceptable for predicting deformations and stability at a
global level. With a larger spacing and for detailed studies of such structures at any
spacing, a 3D model is necessary. For instance, in a simulation of a long, straight
reinforced earth wall with closely spaced steel strip reinforcement, a 2D plane strain
FE analysis study of overall stability or surface settlement may be acceptable but a
verication of adequate safety against reinforcement sliding would require a 3D analysis
because a plane strain analysis would not simulate the detailed soilreinforcement inter-
action correctly.

If closely spaced structures are connected by beams orientated in the out-of-plane direc-
tion, this helps to maintain the plane strain or axisymmetric condition. Common
examples are capping and waling beams used in embedded retaining walls to distribute
load from supports and to connect wall elements together.

Input and output in 2D plane strain models


Figure 5.8 shows a true plane strain case with a retaining wall and supporting slab of
continuous section. If using continuum elements to model the structures, then the geo-
metry is simply a 2D section and the structural elements have the same thickness and
material properties as the actual structure. If using line elements, their geometry should
coincide with the mid-plane of the structures they represent and the section properties
are specied in the input parameters for the structural materials. Note that outputs of
structural forces such as axial force, bending moment and shear force from both
methods will be provided per unit length in the out-of-plane direction.

Figure 5.9 shows a quasi-plane strain case with closely spaced piles forming a contiguous
piled wall and struts at a larger spacing connected in the out-of-plane direction via a
capping beam. First, the structural section properties need to be converted to their
equivalent plane strain values by dividing them by their spacing. The values will be
expressed per unit length in the out-of-plane direction. The conversion for the piles is
shown in Table 5.3. Note that the pile spacing is the distance between each pile, typically
centre-to-centre, and not the size of the gap between them.

144

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Figure 5.8 Example true plane strain structure

Slab
thickness ds

Section properties
entered directly

Wall thickness dw
Slab thickness ds
Wall thickness dw
3D 2D plane strain 2D plane strain
(continuum elements) (non-continuum elements)

Figure 5.9 Example quasi-plane strain structure

Spring stiffness Spring stiffness


Strut k = 4.2 kN/mm/m k = 4.2 kN/mm/m
spacing 12 m
Strut stiffness Equivalent wall
k = 50 kN/mm section properties
per unit length
Wall EI entered directly
Contiguous piled wall: thickness d = 12
0.6 m diameter piles at EA
0.9 m spacing

3D 2D plane strain 2D plane strain


(continuum elements) (non-continuum elements)

Table 5.3 Example wall properties per pile and their plane strain equivalent

Single pile properties Equivalent plane strain properties

Diameter B = 0.6 m Divide single pile properties by spacing (0.9 m)


Area A = 0.283 m2 EA = 6.29 106 kN/m
Second moment of area I = 6.36 10 3 m4 EI = 141 103 kNm2/m
Youngs modulus E = 20 106 kN/m2 Wall weight 7.9 kN per m length per m run
EA = 5.66 106 kN
EI = 127 103 kNm2
Weight density g = 25 kN/m3
Pile weight = 7.1 kN per m length

145

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

If using plate or shell elements to model the wall, the plane strain properties on the right
of Table 5.3 can be entered directly as input for the material model. If using continuum
elements to model the wall, an equivalent thickness d of the 2D continuum element needs
to be derived. In order to model both the axial and bending stiffness correctly, there is
only one unique combination of d and E values. From the ratio EI/EA, a unique value of
d can be obtained from Equation 5.4.

EI
d = 12 (5.4)
EA
Therefore, in this example, d = 0.52 m and E can only be 10.9 106 kN/m2. Also, the
weight density for the wall material should be g = 15.2 kN/m3 in order to obtain the
same equivalent wall weight.

To convert the structural force outputs per unit length in the out-of-plane direction to
values per pile, they are simply multiplied by the spacing. For example, an axial load
of 150 kN/m would be equivalent to 135 kN per pile in this example, while a bending
moment of 400 kNm/m would be equivalent to 360 kNm per pile.

In this example the struts will be represented by spring elements and the structural
engineer provided a spring stiffness value of k = 50 kN/mm per strut to represent the
support from the structure. At a spacing of 12 m, the equivalent plane strain value is
k = 4.2 kN/mm per m run. Depending on the software being used, the k value may be
entered directly or the parameters E, A, L and s may need to be entered, which the
program uses to calculate k according to Equation 5.5.
EA
k per unit length = (5.5)
Ls
where L is the member length and s is the spacing. In this example, A, L and s could be set
articially to 1 and then an E value of 4200 kN/m2 entered to obtain the required k value.

In the output, a strut force of 80 kN/m, for example, would be equivalent to a strut force
of 960 kN per strut.

Input and output in 2D axisymmetric models


Figure 5.10 shows a true axisymmetric case with a circular shaft supported by a retaining
wall of uniform section and a solid slab near the top of uniform thickness. The spokes
drawn on the gure divide the structure into sectors of angle one radian each to help illus-
trate how input and output data are sometimes expressed in axisymmetric analyses, i.e.
per radian.

If using continuum elements to model the structures, then the geometry is simply a 2D
section from the central axis of symmetry and the structural elements have the same
thickness and material properties as the actual structure. If using line elements, their
geometry should coincide with the mid-plane of the structures they represent and the
section properties are specied in the input parameters for the structural materials.

146

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Figure 5.10 Example of a true axisymmetric structure

Wall thickness dw

Slab
thickness ds

Section
Slab thickness ds properties
entered
directly

Wall
thickness
dw

3D 2D plane strain 2D plane strain


(continuum elements) (non-continuum elements)

Figure 5.11 shows a quasi-axisymmetric case with closely spaced piles forming a
contiguous piled wall supported by ground anchors at a spacing of 368 connected in the
out-of-plane direction via a waling beam. First, the structural section properties need to
be converted to their equivalent axisymmetric values. For continuous structures in the
out-of-plane direction, properties are expressed per unit hoop length as shown for the
piles in Table 5.4.

If using plate or shell elements to model the wall, the properties on the right of Table 5.4
can be entered directly as input for the material model. If using continuum elements to
model the wall, an equivalent thickness d of the 2D continuum element needs to be
derived using Equation 5.4. Note that it would be essential to adopt an anisotropic
model for a contiguous piled wall to avoid unrealistic hoop forces being generated (see
Section 5.1.6). In this example, d = 0.65 m was obtained. E can be calculated from the
equivalent EA value, where A is simply d times the unit arc length of the wall centreline
as illustrated in Figure 5.12, giving E = 13.6 106 kN/m2. Also, the weight density
for the wall material should be g = 5.1 kN/m3 in order to obtain the same equivalent
wall weight.

To convert the structural force outputs per metre in the out-of-plane direction to values
per pile, they are simply multiplied by the spacing. For example, an axial load of
125 kN/rad would be equivalent to 150 kN per pile, while a bending moment of
230 kNm/m would be equivalent to 276 kNm per pile.

The equivalent properties of spoke-like structures such as the ground anchors are
expressed per radian as shown in Table 5.5. An output of anchor force of 390 kN/rad,
for example, would be equivalent to an anchor force of 246 kN per anchor.

147

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 5.11 Example of a quasi-axisymmetric structure

Spring stiffness
k = 23.8 kN/m/rad

Wall EI
thickness d = 12
EA

2D axisymmetric
(continuum elements)

Spring stiffness
k = 23.8 kN/m/rad

Anchor Contiguous piled wall:


spacing 36 0.75 m diameter piles
at 11.6 or
1.2 m hoop spacing Equivalent wall
section properties
per unit length
entered directly

3D 2D axisymmetric
(non-continuum elements)

5.1.6 How is geometrical anisotropy in structures handled?


Geometrical anisotropy occurs most often in embedded retaining walls, but can occur in
any structure type where structural sections are discontinuous, non-uniform or proled
in one direction. Embedded retaining walls are continuous in the vertical direction where
the maximum bending resistance is required, but have discontinuities in the horizontal
direction in the plane of the wall. Piled walls have gaps or softer inll between them and
even diaphragm walls have joints that reduce their axial and bending stiffness in the

Table 5.4 Example wall properties per pile and their axisymmetric equivalent

Single pile properties Equivalent axisymmetric properties

Diameter B = 0.75 m Divide single pile properties by spacing (1.2 m)


Area A = 0.442 m2 EA = 8.84 106 kN/m
Second moment of area I = 15.5 10 3 m4 EI = 310 103 kNm2/m
Youngs modulus E = 24 106 kN/m2 Wall weight 3.3 kN per m length per m run
EA = 10.6 106 kN
EI = 37.2 103 kNm2
Weight density g = 25 kN/m3
Pile weight = 4.0 kN per m length

148

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Figure 5.12 Calculation of example wall section area for a unit length sector

Centreline length 1 m

A = 0.65 1.0 = 0.65 m2

d = 0.65 m
Internal shaft
radius 5.68 m

horizontal direction. Sheet piles have a prole that gives them signicantly lower stiffness
in the horizontal direction than the vertical.

In 2D plane strain analyses of wall sections, this does not present a problem (provided
the plane strain assumption is appropriate, e.g. away from the corners of an excavation)
because strains in the out-of-plane horizontal direction along the wall are zero in any
case, so isotropic material properties can be assumed adopting the bending and axial
stiffness in the vertical direction. However, in 2D axisymmetric analyses, the generation
of hoop forces in isotropic wall elements would lead to wall deection being governed by
hoop stress rather than wall bending and a signicant under-prediction of wall deection
and bending, as illustrated by example in Section 8.3.4. Therefore, in 2D axisymmetric
FE analyses, it is essential that anisotropic (orthotropic) constitutive models are adopted
in situations where hoop stresses are generated in anisotropic structures.

Similarly, in 3D analyses of approximately plane strain deformations (e.g. near the


centre of a long retaining wall), isotropic structural properties should be acceptable, but
in non-plane strain situations, e.g. near corners of excavations, anisotropy should be
included in the model. For piled walls, the individual piles can be modelled with gaps
between them or, when using shell elements, appropriate anisotropic axial and bending

Table 5.5 Example anchor properties per anchor and their axisymmetric equivalent

Single anchor properties Equivalent axisymmetric properties

Area A = 600 10 6 m2 Divide single anchor properties by spacing (0.63 rad)


Youngs modulus E = 200 106 kN/m2 EA = 190 103 kN/rad
EA = 120 103 kN Free length stays the same at 8 m
Free length 8 m k = 23.81 103 kN/m/rad
k = 15 103 kN/m Pre-stress 317 kN/rad
Pre-stress 200 kN

149

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

stiffness parameters should be selected. Where only isotropic material behaviour is


available in an FE analysis program, the rotational degrees of freedom in the nodes
of shell elements at excavation corners can be released to reduce the amount of error.
More information on the importance of modelling structural anisotropy is provided
by Zdravkovic et al. (2005).

When specifying anisotropic structural properties, the local axis directions of the struc-
tural elements must be known and particular care is needed to ensure that the local axes
are set up in a consistent way to help avoid errors in the correspondence of material
parameters and axis directions. The difference in stiffness between two perpendicular
directions should be limited to a factor of about 20 to avoid ill-conditioning, even if the
true difference in stiffness is greater.

5.1.7 How are structural connections modelled?


In modelling terms there are three main types of structural connection available, namely
roller (simple), pinned and full (xed) and methods of simulating these connection types
are illustrated in Figure 5.13. Actual structural connections are likely to fall somewhere

Figure 5.13 Modelling idealised connection types

Roller (simple) Pinned Full (fixed)


connection connection connection

Continuum/
continuum
elements

Nodes tied in
one direction

Beam or shell
element extended
into continuum
Continuum/ element
non-continuum
elements Nodes tied in
one direction

One side must be


bar or membrane Beam or shell
Non-continuum/ element elements only
non-continuum
elements Nodes tied in or hinge specified
one direction between beam or
shell elements

150

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

between these idealised categories, e.g. a full connection is likely to allow a small degree
of movement or play that is difcult to simulate. The most appropriate of the idealised
connection types needs to be selected and, where there is uncertainty, connection types
can be changed in order to assess their effect on the key outputs. The choice of con-
nection type is important because it has a large effect on the outputs of structural forces
and deections. Some structural connections may transfer only tensile or compressive
loads and this also needs to be reected in the setting up of these connections in the
FE model.

Connections also inuence the overall stiffness of structures. For instance, the axial
stiffness of a strut support to a retaining wall may be calculated from the section and
material properties of the strut, but exibility at the connection between the strut and
retaining wall, due to packing, for example, could signicantly reduce the true stiffness
of the strut that may not be reected in the FE model due to the assumption of a pinned
or full connection. This could result in an over-prediction of strut force and an under-
prediction of wall deection at the strut. Connection stiffness is a common uncertainty
that needs to be considered when interpreting analysis outputs.

Where elements are joined at a single node, only connections between beam and plate/
shell elements transfer moment and model fully xed connections. When joining a plate/
shell or beam element to a continuum element, no moment is transferred. A full connec-
tion can be simulated approximately by extending the non-continuum element two or
three elements into the continuum, as shown in Figure 5.13.

There are certain assumptions inherent in non-continuum elements (see Section 5.1.1)
that often become invalid near connections, so detailed outputs from connections may
be unreliable. Furthermore, there are assumptions regarding the geometry of connec-
tions inherent in the use of non-continuum elements. Connections are assumed to
occur at the central axis of structural members, as shown in Figure 5.14, which may
not be the case in reality and which affects the prediction of structural forces and
deections.

Figure 5.14 Typical connection between non-continuum elements

Central axes
of members
Spring element

Shell element

FE model Inherent assumptions Reality

151

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

5.1.8 How are distributed loads applied in an FE model?


Distributed loads can be specied in an FE analysis program (e.g. line loads in units of
force/distance and area loads in units of force/area), and although they may be displayed
in the graphical interface as such, remember that loads can only be applied at element
nodes. Load arrangements are converted (automatically in most programs) to equivalent
point loads at nodes. However, the resulting equivalent loads, particularly in a coarse
mesh, may not represent the intended load adequately. So it is important to check how
distributed loads have been represented by equivalent point loads in the outputs, and to
make renements to the specied loading or the mesh geometry appropriately. This is
another reason to have ner meshes in loaded areas, in addition to where there are steep
gradients of stress and strain, as covered in Section 1.3.2.

5.1.9 What are singularities?


A singularity is a point where articially high or low output is calculated due to some
assumptions in the model. A typical example would be very high values of shear force
at a point load obtained from an elastic shell element representing a spread foundation
(as illustrated in Section 8.2.4). In reality, the point load would be distributed over an
area rather than concentrated at a single point. Rather than being innitely thin, the
foundation would have thickness allowing greater redistribution of stress. Also, the
reinforced concrete of the foundation would not be elastic but would crack or yield at
high stresses and the stresses would be redistributed.

Other potential singularity sources include corners with small or large angles, pinned
supports and connections and folds in plate or shell elements. Some singularities can
be removed relatively straightforwardly, for example by replacing a concentrated load
with a more realistic distributed load, as shown by example in Section 8.2.4. Others are
more difcult to remove.

With experience, singularities in outputs can be recognised as such. This is important to


avoid designing structures to resist articially high outputs of stress.

5.2. Structural materials


5.2.1 Can linear elastic models be used for concrete and grouted
structures?
Concrete and grout, both reinforced and unreinforced, are very commonly used
materials in geotechnical structures. Rather like soil and rock, they are complex
materials, but since they are much stronger and stiffer than soils and soft rocks, geo-
technical material characteristics are more likely to govern soilstructure interaction
behaviour and it is usually possible to simplify the concrete or grout constitutive model
to a simple, homogeneous, isotropic, linear elastic model. However, to recognise when
this is not an appropriate assumption, it is necessary to understand the aspects of
concrete and grout behaviour not included in the linear elastic model. Therefore, as well
as providing guidance on the use of linear elastic models, this section also summarises
the simulation of other aspects of concrete and grout behaviour and non-linear models.
More comprehensive guidance on FE analysis of concrete structures is provided by
Rombach (2011) and Brookes (2016).

152

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

For the remainder of this section, the term concrete is taken to mean both concrete and
grout.

Linear elastic model


Concrete is very stiff compared with soil and soft rocks so the strains in concrete are
likely to be small and within the approximately linear initial portion of the stressstrain
curve (Figure 5.15). Therefore, in most cases, a linear elastic model is sufcient to model
concrete structures in soilstructure interaction analyses. In a uniaxial compressive stress
state, normal concrete behaves in an approximately linear elastic way up to a stress of
about 40% of the compressive strength according to Eurocode 2 (CEN, 2004).

Exceptions to this would be when concrete stresses and strains are likely to lie outside the
linear portion of the curve. This can be checked by running an FE analysis with a linear
elastic model for the concrete initially and identifying any parts of the model where
non-linear behaviour would be expected. Also, there are certain cases, such as sprayed
concrete (shotcrete), where signicant stresses can be experienced by the concrete before
it has cured, or unreinforced jet grouted structures, where signicant tensile strain may
occur, that require more advanced constitutive models. These cases are covered later in
this section.

When adopting an isotropic linear elastic model, only two stiffness parameters are
required, namely Poissons ratio n and Youngs modulus E. n is usually between zero,
which is appropriate for concrete in tension, and 0.2, which is appropriate for concrete
in compression. For bending, which has both tension and compression, the n value is
often not critical but outputs of structural forces can be checked to see if they are inu-
enced signicantly.

Figure 5.15 Typical stressstrain curve for concrete in compression

Initial linear
c stiffness

153

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

There are a number of factors affecting the E value. As well as the particular concrete
mix, concrete stiffness is inuenced by age, shrinkage, creep, temperature, cracking and
bedding-in effects. Under sustained stress the concrete will creep and it is common prac-
tice to reduce the E value by about 50% in long-term cases to help take account of
creep and cracking. Like soil and rock, concrete properties are also affected to a certain
extent by the level of connement. A further important factor to consider is shrinkage
strains in concrete which could lead to signicant changes to the dimensions of large
concrete structures, or large stress changes if the structures are prevented from shrink-
ing. Dong et al. (2016) used a relatively simple thermal strain approach to obtain satis-
factory predictions of shrinkage effects in concrete slabs interacting with a diaphragm
wall.

Reinforcement
There are two approaches to modelling reinforcement bars in concrete. Usually, a
smeared approach is taken where the properties of the steel are smeared across the
section to create a single material with the combined strength and stiffness of the
concrete and steel parts of the section, and this is clearly more straightforward to model.

Alternatively, for very detailed studies of soilstructure interaction or connection details,


an explicit approach is taken where the concrete and steel materials are discretised as in
the real structure (the reinforcement as bar elements within the continuum or shell
elements for the concrete). A rigid bond is normally assumed between the concrete and
the reinforcement but this may not actually be the case in anchorage regions or at laps
between bars.

Cracking
A particular feature of concrete behaviour is cracking as the material goes into tension.
This affects the stress distribution in the concrete as well as the stiffness so is important in
detailed studies of concrete behaviour and makes the study of plastic behaviour in
reinforced concrete rather complicated. The stiffness of cracked concrete sections is
inuenced by the reinforcement arrangement and orientation of the cracks and cannot
be calculated easily. One option is to attempt to model the cracking and there are
essentially two approaches.

The discrete crack approach requires contact or gap elements to be set up in the model at
the start, and crack opening and propagation are predicted based on pre-dened criteria.
The drawbacks of this approach are that the location and crack direction are pre-
determined in the analysis and the material parameters are difcult to measure.

The smeared crack approach is more common and is used in advanced models for
sprayed concrete and jet grout. Here a tensile failure surface is included in the constitu-
tive model, which causes the material to soften on crack formation.

Non-linear models
There are some occasions when non-linear geometrical behaviour may inuence struc-
tural behaviour, such as in the buckling of slender columns and struts. Such instances

154

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

are rare in geotechnical structures but may need to be considered for slender piles with
signicant lengths exposed above ground level.

More commonly, a non-linear constitutive model may be required when concrete


cracking needs to be considered, for instance in SCL or in some jet grouted structures
subjected to tensile stresses (e.g. a jet grouted slab subjected to uplift pressure). As
described under Tunnels in Section 5.1.4, advanced constitutive models have been
developed for sprayed concrete (shotcrete) and grout, e.g. Schadlich and Schweiger
(2014), that include tension softening to take account of concrete cracking.

Sprayed concrete models should also include time-dependent strength and stiffness to
take account of curing following load application. Concrete loaded early in its curing
process is also subject to signicantly more creep than concrete loaded when curing is
substantially complete. To obtain input parameters, compression tests on shotcrete are
needed at different ages. Time-dependent properties are not so relevant for jet grouting
because these structures tend to be stressed only after curing.

Fluid concrete
As described in Section 1.4.2, approximate installation effects can be considered and, for
concrete structures, this may require the simulation of static, uid concrete. Since uid
concrete has no strength, it is modelled simply by applying its self-weight to the model
and, if necessary, its uid pressure to any non-horizontal surfaces in contact with the
concrete.

As well as bored piles and diaphragm walls, the installation effect of a thick concrete raft
may need to be considered, for instance. The ground response to the weight of uid
concrete in a freshly cast raft differs somewhat from that under the same weight of stiff,
hardened concrete.

5.2.2 Can linear elastic models be used for steel structures?


Steel structures are also quite commonplace in the ground. Examples include tubular or
pipe piles for foundations and sheet pile walls. Steel is also used in ground anchors, soil
nails, reinforced earth and rock bolts.

Steel material behaviour


The typical stressstrain curve for steel in uniaxial tension in Figure 5.16 shows that the
elastic portion of the curve below the yield stress is quite linear, so the modelling of steel
material within its elastic range is well suited to a simple linear elastic model.

If the plastic behaviour of steel needs to be modelled, the yield criterion that best relates
multi-axial stress states to the uniaxial behaviour shown in Figure 5.16 is generally con-
sidered to be the von Mises criterion, although the Tresca criterion can also be used.
Often, a simple LEPP model is used to model steel because it captures the initial linear
elastic portion of behaviour well and then the steel deforms plastically at constant yield
stress which is conservative for material behaviour. Alternatively, a hardening model
could be adopted in order to simulate the hardening of the steel material up to its
ultimate strength.

155

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 5.16 Typical stressstrain curve for steel in tension

Strain hardening

Yield stress

Elastic

Sheet pile walls


One of the advantages of FE analysis over limit equilibrium methods is that exibility is
included in the calculations, which can lead to economies in design. As described in
Section 5.3.1, while a exible wall has higher deections, the bending moment is actually
lower and cost savings can be made in the structural sections required to resist the bend-
ing moment, provided that the higher deection is acceptable.

If performing plastic design of sheet piles, which is permitted for some pile sections in
Eurocode 3, remember to consider geometric effects. While sheet piles will resist higher
bending moments beyond the initial yield moment (and therefore sustain permanent
deformations), only certain sections will be able to develop the full plastic moment resist-
ance, and possibly even slightly higher due to strain hardening, and these are called high
rotation capacity sections. Low rotation capacity sections suffer local buckling before
the full plastic moment resistance is developed, and all sections will eventually soften due
to geometric buckling effects. Such softening behaviour can be incorporated into the
material model for the plate or shell elements. Alternatively, when modelling high
rotation capacity sections, plasticity can be allowed only up to the rst occurrence of
plastic moment so that the softening portion need not be modelled. More information
on the FE analysis of sheet pile walls with plastic hinges can be found in Bourne-
Webb et al. (2011).

5.3. Soilstructure interaction


5.3.1 How does relative soil/structure stiffness influence outputs?
The engineering behaviour of two materials of very different stiffness (e.g. soil and con-
crete) interacting with each other is signicantly more complicated than that of one
material alone. For instance, an elastic beam on xed supports can be assumed to have
a deection under a certain load that is inversely proportional to the beams bending
stiffness EI, while the bending moment remains essentially unchanged in spite of vari-
ations in EI (Figure 5.17).

156

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Figure 5.17 Elastic behaviour of a simple structure

Displacement
Low EI:
High deflection

High EI:
Low deflection
Bending moment unchanged
EI

When a structural member transfers load to or from soil, the situation is more compli-
cated, as illustrated in Figure 5.18. While increasing structural stiffness relative to the soil
leads generally to less deection of the structure and soil, the bending moment is not
constant but increases due to the greater ability of the member to distribute loads. As
relative structural stiffness decreases, the soil redistributes more stress and the bending
moment in the structure decreases. The relationships between relative stiffness and
deection or bending moment are non-linear, as shown for the range of typical graphs
in Figure 5.19. Even against the log of EI, bending moment and displacement usually
plot as curves, whose shape depends on the particular characteristics of each application.

It is possible to make simplifying assumptions in certain cases of soilstructure inter-


action analysis, as shown in Figure 5.20. If a structure is sufciently stiff relative to the
ground, it can be assumed perfectly rigid and represented as a rigid body (linear elastic
material with high stiffness) or by prescribed displacements (see Figure 5.7 for spread
foundations) in the FE model. If a structure and its foundations have sufciently low
stiffness relative to the soil, the structure can be assumed perfectly exible. The structural
loads are then independent of settlement because the structure has no ability to transfer

Figure 5.18 Soilstructure interaction behaviour

High relative structure stiffness:


Lower deflection
Higher bending moment

Low relative structure stiffness:


Higher deflection
Lower bending moment

157

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 5.19 Influence of relative structural stiffness on soilstructure interaction outputs

Bending Displacement
moment

Log EI Log EI

loads, so ground deformation can be predicted merely by applying the structural loads to
the ground and ignoring the structure.

In most cases, the stiffness of the structure relative to the ground will lie well within these
extremes so that the structure will deform with the ground and it will have the ability to
redistribute some of the loads. The structural loads on the ground then depend on defor-
mation and both the structure and ground stiffness will need to be modelled correctly in
the FE analysis. In complex cases, the ground and geotechnical structures are normally
simulated in a separate FE analysis model from the superstructure. Therefore, some
iterations between the separate models are needed until deformations and loads are in
sufcient agreement. The structural model may use a simpler method to simulate the
ground, such as springs, as discussed in Section 5.3.2.

Figure 5.20 Idealised assumptions in soilstructure interaction

Perfectly rigid structure Perfectly flexible structure

Settlement on a plane Structure loads independent of settlement

Perfectly rigid ground Stiff structure

Load transfer Load transfer

Structure loads dependent on settlement

158

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Finally, there are cases, usually on hard rock, where ground deformations are sufciently
small to be ignored so that the structure can be assumed to have xed supports.

Calculating the relative stiffness of a structure compared with the ground in order to
decide whether any simplifying assumptions can be made is rather difcult without
setting up an FE analysis model of the structure and ground anyway. An approximate
assessment can be made using, for example, Equation 5.6 from Annex G of Eurocode 2
Part 1-1 (CEN, 2004).
(EJ )s
KR = (5.6)
El 3
where KR = relative stiffness, (EJ )s is the approximate value of the exural rigidity per
unit width of the building structure under consideration, obtained by summing the
exural rigidity of the foundation, of each framed member and any shear wall, E is the
Youngs modulus of the ground and l is the length of the foundation.

A KR value higher than 0.5 is said to be indicative of a rigid structural system.

As with most assumptions taken in designing an FE model, the best way to check
whether the assumption is valid is to perform FE analyses with and without the assump-
tion and to compare the outputs from each analysis.

5.3.2 How are coefficients of subgrade reaction determined for


beam-spring models?
This section has been included in this book on FE analysis because the outputs of FE
analysis can provide a reliable source of coefcients of subgrade reaction.

The coefcient of subgrade reaction k (also called modulus of subgrade reaction and
bedding modulus) is derived from the bearing pressure (or subgrade reaction) on soil
divided by the resulting ground deection. Hence its units are pressure/distance
(typically MN/m3). Clearly this method is not intended to simulate soil but merely to
simulate the inuence of soil on the engineering behaviour of structures.

Typically, the coefcient of subgrade reaction is discretised into individual springs with
spring constants, xed at one node and attached to beam or plate elements representing
the structure at the other. The main advantage of this method is its simplicity, particu-
larly in programs dedicated to more complex analyses of structures. However, one
signicant disadvantage is that the shear stiffness of soil is not taken into account, lead-
ing to unrealistic deections in some cases, particularly at the edges of structures where
the transfer of stress to soil outside the structure is not modelled. Often, the k value at
edges is increased to help take account of this and to obtain more realistic deections.

Unfortunately, k is sometimes mistaken for a soil property, like stiffness, as though there
is a soil test available to measure k. It is not a soil parameter but rather an interaction
parameter applicable to a particular combination of circumstances in the ground and
in the structure.

159

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

The only way to measure k directly is to measure the deection of a full-scale structure
with known loads, which is clearly impractical, so a second signicant disadvantage of
the method is the difculty of determining k. This is because a lot of inuences, including
the following, need to be included in a single parameter:
g stiffness of the ground, and its variation with depth, appropriate for the stress and
strain level
g exural stiffness of the structure
g geometry of the structure and the soil (e.g. soil layers, proximity to bedrock)
g loading conditions (e.g. point loads, distributed loads).
The most accurate way to estimate k is to perform an FE analysis of the problem that
takes explicit account of all these inuences and then obtain k from the outputs of bear-
ing pressure and deection on an appropriately spaced grid, as illustrated by example in
Section 8.2.4. Such a grid of k values could then provide the input to a structural analysis
program modelling the inuence of the ground in this way. Further iterations of applied
load and deection may be required until the load distribution in the geotechnical and
structural analyses are sufciently close.

REFERENCES
Addenbrooke TI, Ong JCW and Potts DM (2002) Finite element analysis of a compen-
sation grouting eld trial in soft clay. Proceedings of the Institution of Civil Engineers
Geotechnical Engineering 155(1): 4758.
Bourne-Webb PJ, Potts DM, Konig D and Rowbottom D (2011) Analysis of model sheet
pile walls with plastic hinges. Geotechnique 61(6): 487499.
Brookes CL (2016) How to Model Structural Concrete Using Finite Element Analysis.
NAFEMS, Hamilton.
CEN (2004) EN 1992-1-1 Eurocode 2: Design of concrete structures Part 1-1: General
rules and rules for buildings. CEN, Brussels.
Dong YP, Burd HJ and Houlsby GT (2016) Finite-element analysis of a deep excavation
case history. Geotechnique 66(1): 115.
Lees AS (2017) The simulation of geogrid-stabilised soil by nite element analysis. Pro-
ceedings of the 19th International Conference on Soil Mechanics and Geotechnical
Engineering, Seoul, Korea, 1722 September.
Li Z, Soga K and Wright P (2015) Behaviour of cast-iron bolted tunnels and their model-
ling. Tunnelling and Underground Space Technology 50: 250269.
Moller SC and Vermeer PA (2005) On design analyses of NATM-tunnels. In Underground
Space Use: Analysis of the Past and Lessons for the Future (Erdem and Solak (eds.)).
Taylor & Francis Group, London, pp. 233238.
Muir Wood AM (1975) The circular tunnel in elastic ground. Geotechnique 25(1): 115127.
Panet M and Guenot A (1982) Analysis of convergence behind the face of a tunnel.
Proceedings Tunnelling 82, London. Institution of Mining and Metallurgy, pp. 197204.
Rombach GA (2011) Finite-element Design of Concrete Structures, 2nd edn. ICE Publishing,
London.
Schadlich B and Schweiger HF (2014) A new constitutive model for shotcrete. In
Numerical Methods in Geotechnical Engineering (Hicks, Brinkgreve and Rohe (eds.)).
Taylor & Francis Group, London, pp. 103108.

160

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How are geotechnical structures modelled?

Soga K, Bolton MD, Au SKA, Komiya K, Hamelin JP, van Cotthem A, Buchet G and
Michel JP (2000) Development of compensation grouting modelling and control system.
In Geotechnical Aspects of Underground Construction in Soft Ground (Kusakabe, Fujita
and Miyazaki (eds.)). Balkema, Rotterdam, pp. 425430.
Tschuchnigg F and Schweiger HF (2015) The embedded pile concept Verication of an
efcient tool for modelling complex deep foundations. Computers and Geotechnics 63:
244254.
Wang F, Huang H, Soga K, Li Z, Zhang D and Tsuno K (2012) Deformation analysis of
a tunnel with concrete segmental lining subjected to ground surface loading using novel
joint model. Proceedings of the World Tunnel Congress 2012, Thailand, pp. 364366.
Wisser C, Augarde CE and Burd HJ (2005) Numerical modelling of compensation grout-
ing above shallow tunnels. International Journal for Numerical and Analytical Methods in
Geomechanics 29: 443471.
Zdravkovic L, Potts DM and St John HD (2005) Modelling of a 3D excavation in nite
element analysis. Geotechnique 55(7): 497513.

161

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.163

Chapter 6
Can FE analysis be used with
design codes?

6.1. Introduction
In this chapter, only the limit state design concept will be considered, since this has been
adopted in the majority of geotechnical design codes. This concept uses design criteria to
dene limits within which structures are safe and t for use.

The ultimate limit state (ULS) concerns safety and occurs when a structure or the ground
suffers from a loss of stability. The serviceability limit state (SLS) concerns the proper
functioning and appearance of a structure in service and occurs when a structure or the
ground experiences, for example, deformations that are perceptible, cause damage (e.g.
cracking) or prevent functioning of machines (e.g. lifts).

A safety margin against a ULS occurring is introduced by statistical analysis, by direct


selection of conservative parameters or, more commonly, by applying safety factors pre-
scribed by a design code. The safety factor may be a single (global) value applied once in
the calculation or else distributed among partial factors applied to individual parameters
such as loads, material strengths (e.g. undrained shear strength) and resistances (e.g.
bearing resistance). Except for Section 6.1.3, where statistical methods are considered
briey, it is assumed throughout this chapter that the more commonplace partial factor
method is being used.

6.1.1 Why perform geotechnical design with FE analysis instead of


conventional methods?
Design codes are written primarily with conventional design methods in mind, so
performing design by FE analysis in accordance with a design code is not always
straightforward. Indeed, the exibility of FE analysis means that no geotechnical design
code could provide a set of rules to cover all possible applications of FE analysis. Only a
relatively small set of quite fundamental rules may be provided, if at all, leaving the user
responsible for interpreting the code for each specic problem and making the correct
decisions to ensure that designs comply with the code.

Consequently, FE analysis should only be used to perform design in accordance with


a design code when the additional workload can be justied. Some of the advantages
of the FE method were covered in Section 1.1.1. In addition to these, some specic
advantages associated with design codes include:

163

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

g simultaneous checking for limit states in multiple forms


g checking SLS and ULS with one analysis model
g taking into account the exibility of structures and soilstructure interaction
effects when checking SLS and ULS.

FE analysis can be applied to SLS design more straightforwardly, as described in


Section 6.2. ULS design is also possible by FE analysis, but there are more aspects to
consider, as described in Section 6.3.

6.1.2 What influences the occurrence of a limit state apart from ground
strength (for ULS) and stiffness (for SLS)?
Due to the complexity of FE analysis, there are many inuences on the prediction of
limit states. Certainly much more than considering merely loads, ground strength (and
its partial factor) for ULS design and ground stiffness for SLS design.

The sensitivity of limit state predictions to these other inuences should be considered,
including the following.

Discretisation of geometry
Stresses are calculated only at the integration points, so the distribution of stress is
known only approximately. As a mesh becomes ner in critical areas, the resolution of
stress values improves and the exact solution of a ULS can be approached. If the mesh
is too coarse, FE analysis tends to under-predict displacements (SLS) and over-predict
failure loads (ULS). This is a particular problem in 3D analysis where lower order
elements are normally used and a very high number of elements are required to simulate
failure states accurately in 3D compared with 2D. Whether a particular mesh is sufciently
ne in critical areas can only be conrmed by performing a sensitivity analysis with
progressively ner meshes until there is no signicant change in the predicted limit state.
Alternatively, adaptive meshing provides an automated approach (see Section 1.3.2).

The choice of element type can also inuence the occurrence of a limit state. Some higher
order elements with full integration, such as the 15-noded triangle (Sloan, 2013) in 2D
analyses, have been found to provide reasonably accurate failure state predictions (see
Section 1.3.1). Lower order elements that were used more commonly in the past were
prone to locking (i.e. an over-stiff response due to incompressible constraints and insuf-
cient degrees of freedom) in undrained analyses, particularly in 2D axisymmetric
models, and in some drained analyses with dilation, but this happened less when reduced
integration was used. The choice of element type for structures (e.g. continuum or non-
continuum) as covered in Section 5.1.1 also has a signicant inuence on the prediction
of limit states in soilstructure interaction problems.

Initial stress state


The initial stress state and the stress ratio K0 have a signicant inuence on the prediction
of limit states in subsequent analysis stages. K0 inuences the prediction of structural
forces, particularly when displacements are small and active and passive pressures have
not been mobilised as well as in heavily over-consolidated soils with high K0 values.

164

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

The earth pressure on stiff structures, such as tunnel linings, is usually higher at SLS than
at ULS.

Horizontal in situ stresses or K0 are difcult to measure or estimate, as described in


Sections 1.4.1 and 3.4.1. Upper and lower bound values should be considered where
there is uncertainty in order to understand the inuence of K0 on key outputs and to
judge which values should govern the design.

Equations used to estimate K0 (e.g. Equations 3.6, 3.8 to 3.11) generally contain the
friction angle w but this does not necessarily mean that K0 should be factored along with
shear strength. Efforts should focus on obtaining a realistic initial stress state as well as
performing a parametric study on K0 where there is uncertainty.

Preceding construction stages


The simulated construction activities cause changes to the stress state, stress path and
stress history which, particularly with advanced constitutive models, inuence the pre-
diction of limit states in subsequent stages.

Boundary conditions
Concentrated rather than distributed loads are a convenient simplication but can cause
articial local failures and unrealistically high outputs of structural forces at singularities
(see Section 5.1.9).

When determining failure loads or resistances by FE analysis, imposing displacements


and obtaining the failure load from the output (strain control) tends to be more success-
ful than imposing an increasing load to failure (stress control).

Assumptions at soilstructure interfaces, such as perfectly smooth, perfectly rough or


adopting interface elements clearly have an impact on the prediction of limit states, as
they do in conventional design. Similarly, the assumptions of rigid, stiff or exible struc-
tures (see Section 5.3.1) affect the distribution of stresses imposed on the ground and,
hence, the prediction of limit states.

When model boundaries are too close to the area of interest, the imposed xities intro-
duce boundary effects to the analysis and could cause failure loads to be over-predicted
(see Section 1.2.3).

Drainage conditions
An assumption of wholly drained or undrained conditions has an important effect on
the prediction of limit states effects that are conservative only in certain cases (see
Section 4.2.2). In the intermediate cases of partial drainage, a coupled consolidation
analysis may be required if adopting the drained or undrained assumption would be too
inaccurate. Consolidation analyses allow temporal predictions of deformation as well as
taking account of strength changes in soil due to consolidation. Such predictions are
heavily dependent on soil permeability and drainage path lengths which are often highly
uncertain parameters, so should be subject to a parametric study in order to understand
more fully the likelihood of a limit state occurring.

165

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Constitutive behaviour
A constitutive model denes the relationship between stress and strain in an FE model.
Clearly, the accuracy of limit state prediction is dependent on the selection of an appro-
priate constitutive model, as covered in Chapter 2. In general, more advanced models are
required for accurate deformation (SLS) and structural force (ULS) prediction than
ground failure (ULS) prediction, but there are exceptions. For instance, when modelling
undrained behaviour in terms of effective stress (Method A see Section 4.2.4), param-
eters such as stiffness have a direct inuence on the prediction of excess pore pressures
and hence undrained shear strength.

Flow rules inuence limit state predictions. An associated ow rule where the dilation
and friction angles are equal can lead to dangerously over-predicted failure loads in
undrained and conned problems. In most cases, a non-associated ow rule should be
used and dilation set to zero to obtain realistic failure loads or to help ensure that they
err on the safe side.

Predictions of deformation (and whether deformations are reversible) for the verication
of SLS are inuenced by any yield of the ground and hence by strength-related param-
eters. In turn, the prediction of structural forces for ULS verication is inuenced by the
stiffness of both the ground and structure (see Section 5.3.1).

Analysis options
The accuracy of limit state predictions is affected by choices made by the user or auto-
matically by the program in the setting up of FE analysis models, such as the number of
load steps, the integration scheme and tolerances. Stringent tolerances should be
adopted to help obtain accurate predictions (see Section 1.4.3).

6.1.3 How can the reliability of designs by FE analysis be checked?


It is difcult to achieve high accuracy in geotechnical design because there are uncertain-
ties in ground properties and site conditions, because all design methods have elements
of approximation and assumption and because errors can occur. One of the fundamental
roles of a design code is to introduce reliability into design, i.e. to reduce the probability
of limit state occurrences to acceptable levels, taking into account these uncertainties.
Design codes introduce reliability into designs primarily in two ways:

1 by requiring the study or selection of parameters to obtain a prescribed degree of


conservatism (see Section 3.4.1 for geotechnical material parameters)
2 by requiring additional reliability for ULS design by, for example, a partial factor
method or statistical analysis.

There are a number of different approaches to try to achieve the level of reliability in FE
analysis required by a design code, as described in the following paragraphs.

Deterministic approach
Single parameter values are selected according to the requirements of the design code
(e.g. moderately conservative values) and the design calculation is performed once with

166

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

the selected parameters together with any partial factors. There are two main drawbacks
with this approach. First, it provides no indication of the effect of the variability in
ground properties on the reliability of the design, thereby placing a high importance
on the selected value. Second, single values of partial factors provided in a design code
cannot possibly t all cases and provide a consistent level of reliability.

Probabilistic or stochastic approach


Rather than selecting single values for calculation, all available data are used to dene
random probability distributions allowing a problem to be fully analysed statistically.
This approach is called non-deterministic because random patterns of parameters rather
than specic values are used in calculations. Multiple calculations can be performed (as
in a Monte Carlo simulation) leading to a direct calculation of reliability (i.e. probability
of failure). This approach can be applied in FE analysis using the Random Finite
Element Method (Fenton and Grifths, 2008) where random, spatially varying param-
eter distributions are generated. Multiple analyses are then performed and reliability
determined from the outputs.

Sensitivity analysis and parametric study


Sensitivity analysis involves varying the input parameters to an FE analysis in order to
determine which have the most inuence on the key outputs from the model. This is then
often followed by a parametric study where a smaller number of the critical input param-
eters identied in the sensitivity analysis are varied between permissible ranges in order
to determine the permissible ranges of critical outputs. These are commonly performed
in FE analyses and form an intermediate approach between the deterministic and
probabilistic approaches described previously. Rather than relying on the selection of
single values without considering the effect of parameter variability, as in the determi-
nistic approach, different values across a range are studied and the outputs used to assess
reliability. Simple probabilistic methods can be introduced by considering the mean and
coefcient of variation of parameters in the selection of input values for parametric
studies and then reliability values determined from the outputs, as outlined in
Section 7.3.3.

6.2. Serviceability limit state (SLS)


6.2.1 How is the SLS verified using FE analysis?
The prediction of SLSs is well suited to FE analysis because the displacement nite
element method is intended primarily, as the name suggests, to predict displacements.
It is also suited to FE analysis because the SLS can occur in the conditions experienced
by a structure in service, whereas a ULS requires unrealistic conditions due to the low
probability of its occurrence.

However, it is worth making the distinction between verifying that an SLS is sufciently
unlikely to occur and predicting actual behaviour as accurately as possible. In the
former, due to uncertainties in the properties of the ground, cautious estimates of input
parameters, initial stresses and groundwater level, etc. may be adopted in an FE model.
Other simplifying assumptions and decisions in setting up the model (such as drainage
conditions, ignoring anisotropic behaviour, adopting non-continuum elements for

167

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

structures, etc.) should also err on the conservative side. Specied loads, particularly the
variable (live) loads, are necessarily signicantly higher than day-to-day actual loadings,
even when unfactored. Consequently, should outputs show that the SLS was not
reached, there is a high degree of condence that the SLS will not be exceeded in the real
structure because of all the cautious assumptions and selections. If the outputs were then
compared with site monitoring data, it should come as no surprise if the predictions were
well on the safe side of the actual measurements.

If a most probable estimate of deformations were required that would be expected to be


closer to measured deformations, then best estimate rather than cautious estimates of
parameters should be adopted. These include loadings, drainage conditions, ground-
water level, etc., as well as constitutive model input parameters.

Where verication of both ULS and SLS is required, it is possible in some instances to
verify both by simulating the main construction sequence only once since partial factors
in the ULS case are applied either to the outputs in output factoring or in separate ULS
stages for input factoring, as described in Section 6.3.2. This leaves unfactored parameters
(except possibly for a small factor on variable loads) which are suited to SLS verication in
the main construction sequence. This is not always possible because sometimes different
parameters are appropriate for SLS and ULS, e.g. a peak strength may be appropriate for
realistic deformation predictions in the SLS case while a safer post-peak strength may be
appropriate for the ULS case. Also, a more conservative view may be taken of some
geometries, such as excavation depth, in the ULS case compared with the SLS case.

6.3. Geotechnical ultimate limit state (ULS)


6.3.1 How is the ULS verified using FE analysis?
Simulating a failure (or ULS) is less straightforward than for an SLS because it involves
more complicated material behaviour (yielding and failure) and because an event is
being simulated that should be highly unlikely, so the FE model needs to be manipulated
in order to simulate an unrealistic situation. It is primarily for this second reason that
verifying ULS by FE analysis is not straightforward, particularly for complex models.

There are essentially two main approaches to introducing partial factors in geotechnical
limit state design, as described in this section and summarised in Figure 6.1. Combining
these into a dual factoring approach is the most consistent for FE analysis, as described
at the end of this section.

Input factoring (or material factoring approach)


Partial factors are applied on input parameters in an FE analysis at the sources of uncer-
tainty, i.e. on variable loads and ground strength parameters. They can also be applied
on permanent (dead) structural loads, depending on the requirements of the design code.
Shear strengths at groundstructure interfaces are also factored. Water levels (and hence
pressures) should be set to their worst case or design levels as dened by the design code.

Input factoring is suited to problems involving equilibrium of the ground (e.g. retaining
walls, embankments, cut slopes), as shown in Figure 6.2, because factoring the

168

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

Figure 6.1 Input and output factoring approaches

INPUT OUTPUT

Worst case water levels YES 


Any geotechnical failure?
Factored NO
Ground
strength Calculation Structural forces Already factored
Factored
Structural and/or
variable loads
Structural ULS check

Ground Unfactored
self-weight

Input factoring approach

INPUT OUTPUT
Characteristic
water levels
Geotechnical loads, e.g. Apply load
Ground Unfactored
anchor load, pile load (effect) factor
strength
Small factor Calculation Geotechnical resistance check
Variable
(live) loads
Structural forces, e.g. Apply load
retaining wall bending (effect) factor
Other loads,
including ground moment
Unfactored
self-weight
Structural ULS check

Output factoring approach

self-weight of the ground (which is very difcult to do) is not required and because these
failures are governed by the shear strength of the ground mass. If, having factored the
ground strength parameters, geotechnical failure does not occur, the geotechnical ULS
can be said to have been veried (while not forgetting the other inuences on limit state
prediction described in Section 6.1.2). The factors should be applied at dedicated stages
during the analysis to avoid running the whole analysis with unrealistic factored input
parameters (see Section 6.3.2). The ground strength can be further reduced until failure
occurs in order to identify the most critical failure mechanism and obtain a factor of
safety at each stage. The strength reduction can be performed by a one-step reduction
for basic constitutive models or by a stepwise strength-reduction procedure, as described
in Section 6.3.5.

Input factoring is less suited to problems where predominantly interface rather than
ground failure occurs (e.g. pile foundations and ground anchors). This is because failure
is governed by the particular properties of the interface which differ from those of the
undisturbed ground mass. Installation of such structures results in remoulding, mixing
and changes in density of the ground immediately surrounding the structure, so the
parameters of the undisturbed ground (and their partial factors) become less relevant

169

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 6.2 Suitable methods for verifying ULS of geotechnical structures

Internal loads External loads External loads


Ground failure Ground failure Interface failure

Overall stability Piled foundations

Spread foundations

Suited to either

Retaining walls Ground anchors,


soil nails, rock bolts, etc.

Suited to Suited to
input factoring output factoring

in local failures of these structures. Furthermore, their performance is dened in terms of


a directly measured resistance, e.g. pile compressive resistance or anchor pull-out load,
and resistance is often either determined by load testing or calculated by direct design
methods from in situ tests. This contrasts with a retaining wall, for example, whose
performance is dened in terms of the parameters of the ground around it.

Input factoring is also less suited to providing factored values of structural forces with a
consistent degree of conservatism (although it does provide a valuable additional check
as described in Dual approach later in this section). Factoring ground strength usually
transfers more stress to structural elements, causing structural forces to increase, but the
relationship can be non-linear. Stiff structures with small ground displacements (e.g. a
multi-propped embedded retaining wall in stiff soil) result in no yield in the ground,
so factoring the shear strength of the ground can have no effect on structural forces and
so introduces no safety margin. Therefore, it should be viewed as an additional check on
structural forces in cases where weaker than expected ground has a particularly strong
inuence on structural forces, e.g. in support structures to marginally stable slopes (see
Output factoring).

170

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

Output factoring (or load (effect) and resistance factoring approach)


With output factoring, FE analyses are performed with essentially unfactored input
parameters (except perhaps for a small factor on variable or live loads to take account
of the greater uncertainty on variable loads see later in this section) and characteristic
water levels (as dened by the design code). Then partial factors are applied on load
effects in the output of FE analyses. Partial factors are also applied on resistances
usually obtained from a parallel calculation by either FE analysis (see Section 6.3.6)
or other calculation methods, and the factored outputs are compared with the factored
resistance. If the factored resistance (e.g. ground anchor pull-out resistance) is not less
than the factored load (e.g. ground anchor load), then that particular ULS for that
construction stage has been veried.

It is suited to cases where the load effect and resistance to be compared are both well
dened and largely independent of each other. For instance:

g pile axial load and compressive or tensile resistance


g anchor, nail or rock bolt load and pull-out resistance
g applied load on a spread foundation and bearing or sliding resistance.

The last of those three examples refers more to pad foundations with simple loadings
(for which an FE analysis would not normally be required). In raft foundations, soil
structure interaction effects become signicant and checking for geotechnical failure
by output factoring is less straightforward. Also, once inclined or eccentric loads are
introduced, combined horizontal and vertical failures may be predicted by FE analysis,
which are more difcult to compare with calculated resistances.

Output factoring is less suited to cases where an output load is not well dened (e.g. in
overall stability) or where the output acting on a structure comes from the ground, e.g.
the pressure from retained ground acting on a retaining wall. This is because the output
interacts with the wall and, in turn, with the resistance also acting on the structure. The
output and resistance are therefore not independent and the difference between them is
often too small (particularly when the K0 value is high) to demonstrate that passive
resistance failure is sufciently unlikely to occur (Lees, 2013).

Therefore, for verifying geotechnical ULS by FE analysis, output factoring is more


suited to interface failure type structures with external loads, e.g. pile foundation,
ground anchor, soil nail and rock bolt, where the loads applied to them are outputs from
the FE analysis arising from other interactions within the model, for example:

g raft foundation supported on piles


g embedded retaining wall supported by ground anchors
g slope supported by soil nails
g tunnel supported by rock bolts.

For cases where a pre-dened external load is applied directly to a structure, e.g. axial
load applied to a foundation pile, the ULS check is a simple comparison between applied

171

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

load and resistance and an FE analysis is not needed. The exception to this is where the
resistance is being calculated by FE analysis, which is covered in Section 6.3.6.

Output factoring provides factored values of structural forces with a consistent degree of
conservatism in most cases as dened by the partial factor value. The factored values are
then compared against structural resistances as described in Section 6.4.1. However,
there are some instances where output factoring may not provide an adequate safety
margin. A common example is when structures (e.g. retaining wall or ground anchors)
are used to improve the margin of safety on the stability of a slope. With unfactored
input parameters, the slope is stable (just) and structural force outputs would be very
low or zero. Applying a partial factor to a near-zero output value would not provide
an adequate safety margin. When the ground strength is factored (i.e. input factoring),
the slope becomes unstable and structural forces increase to values higher than those
obtained from output factoring. In this case, input factoring would provide a more
appropriate output for verication of ULS and this is one advantage of the dual factor-
ing approach described later in this section.

Design codes generally have different partial factors for permanent (dead) and variable
(live) loads. Since it is not possible to differentiate between permanent and variable loads
in outputs, the input values of variable load should be factored by the ratio between the
variable and permanent load factors. For example, the Structural Eurocodes, at the time
of writing, had partial factors of 1.35 and 1.50 on permanent and variable loads, respect-
ively. Therefore, the input values of variable load would be factored by 1.5/1.35 = 1.1
and then the outputs factored by 1.35 so that the effects of all the loads are factored
in accordance with the code.

A disadvantage of verifying geotechnical ULS by resistance factoring is that it requires


prior selection of failure mechanisms for which resistance factors are provided in the
design code. The FE methods advantage of determining the most critical of all failure
forms is therefore lost.

Some programs allow a pile or anchor resistance (factored or unfactored) to be specied


in the input parameters. In effect, the resistance is still determined in a parallel calcu-
lation (by FE analysis or other method) or from a load test but the comparison between
the output (e.g. anchor load) and resistance (e.g. anchor pull-out resistance) is made
automatically by the program. When the output load reaches the specied limiting force,
any additional load would need to be redistributed elsewhere in the FE model. Such a
program feature is potentially useful for automatic verication of ULS of pile and
anchor-type structures as well as to check for potential combined failure mechanisms.

Dual approach
Both the input and output factoring approaches have their advantages and disadvantages
when applied in FE analysis. Taking only one approach risks missing a critical ULS due to
shortcomings in one or other of the two approaches. Therefore, both approaches should
be employed to help verify that all possible ULSs are sufciently unlikely to occur and to
gain from the advantages of each approach, as summarised in Table 6.1.

172

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

Table 6.1 Summary of the advantages combined in the dual factoring approach

Input factoring Output factoring

g Suited to verifying ULS ground failures g Suited to verifying geotechnical ULS of


g Further strength reduction identifies most piles, anchors, soil nails, rock bolts, etc.
critical of all ground failures g Factored structural forces obtained with
g Failure prediction not constrained to consistent degree of conservatism in most
particular failure forms cases
g Obtains more onerous structural forces in g Calculations are performed with unfactored
cases where weaker than expected ground parameters (except for variable load) so
has a significant effect should be more realistic

In particular, input factoring is suited to checking for possible failures in the ground
while output factoring is suited to verifying adequate resistance of foundation piles,
ground anchors, soil nails, rock bolts and similar structures, as well as providing factored
values of structural forces.

Structural forces resulting from both input and output factoring should be obtained and
the most onerous values from each used to verify structural resistance. This is to cover
cases where weaker than expected ground would have a particularly signicant effect
on structural forces, such as in support structures to marginally stable slopes.

6.3.2 How are partial factors applied in FE analysis?


A well-executed FE analysis should provide a reasonably accurate simulation of a real-
world geotechnical structure. A ULS should be sufciently unlikely to occur that to
simulate it would not be an accurate representation of the real geotechnical structure.
The advantage of output factoring is that the FE analysis remains realistic while the
check on whether a ULS is sufciently unlikely is performed only on the outputs.
However, as described in Section 6.3.1, input factoring also needs to be performed and
this is more difcult because the FE model begins to depart from reality.

There are essentially two approaches to applying input factors. Factored input param-
eters can be used from the start and throughout every stage of the analysis, as shown
in Figure 6.3, or the main construction sequence can be simulated with unfactored input
parameters and then loads factored and ground strengths reduced in adjunct analysis
stages separated from the main construction sequence (see Figure 6.4).

Input factoring from the start is the easier of the two approaches and can be applied
in any software without modication. However, the whole FE analysis becomes less
realistic because the loads and ground strengths have been factored. The initial stresses
may not be realistic (as demonstrated by example in Section 8.4.2) and, with successive
analysis stages, the FE model may depart further from reality, particularly for highly
non-linear constitutive models and in situations close to failure, and not necessarily in
a conservative direction. Potts and Zdravkovic (2012) obtained greater errors when using

173

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 6.3 Dual factoring approach with input factored at start

Ground strength and


Small factor on variable external loads factored
OUTPUT load in input for all in input for all INPUT
FACTORING stages, characteristic stages, worst case FACTORING
water levels water levels

Initial state

Factor outputs for geotechnical Any geotechnical failure?


(piles, anchors, etc.) and Construction stage 1
structural resistance checks Structural force outputs for
structural resistance check

Non-critical stage,
Construction stage 2 no ULS check

Factor outputs for geotechnical Any geotechnical failure?


(piles, anchors, etc.) and Construction stage 3
structural resistance checks Structural force outputs for
structural resistance check

Figure 6.4 Dual factoring approach using strength reduction

Ground strength unfactored,


small factor on variable load,
characteristic water levels
OUTPUT INPUT
FACTORING FACTORING
Initial state

Factor external
loads, worst case
Factor outputs
water levels Any geotechnical failure?
for geotechnical
Construction stage 1 ULS stage
(piles, anchors, etc.) and Reduce ground Structural force outputs for
structural resistance checks strength structural resistance check

Non-critical stage,
Construction stage 2 no ULS check

Factor external
Factor outputs loads, worst case
water levels Any geotechnical failure?
for geotechnical
(piles, anchors, etc.) and
Construction stage 3 ULS stage
Reduce ground Structural force outputs for
structural resistance checks structural resistance check
strength

174

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

this approach with advanced constitutive models in bearing capacity analyses, for
example. This approach also requires the complete construction sequence to be simu-
lated twice when undertaking dual factoring: once with unfactored input parameters for
the output factoring approach and then again with factored input parameters for the
input factoring approach, as shown in Figure 6.3.

Strength reduction has the advantage of allowing a dual factoring approach by simu-
lating the complete construction sequence only once (see Figure 6.4) because the input
factoring is performed on adjunct stages. This leaves the main construction sequence
to be simulated with unfactored parameters (except for the small factor on variable loads
see Section 6.3.1), so it may be possible to verify the SLS as well (see Section 6.2.1).
Consequently, each construction stage is arrived at with realistic parameters and,
hopefully, realistic stress states. Non-critical stages, such as adding additional support
to a retaining wall without excavation, do not need an adjunct strength-reduction stage.

The main disadvantage of strength reduction is that a robust procedure is needed to


perform this and there is no agreed, unique way of doing this. This is discussed in
Section 6.3.5.

6.3.3 What values should the partial factors have?


Design codes provide values of partial factors but these are unlikely to have been cali-
brated on or intended for FE analysis. Design codes are primarily intended for conven-
tional design methods which often have degrees of conservatism inherent within them.

Codes differ in their requirements for selecting geotechnical parameters. Terms such as
most probable, characteristic, moderately conservative and worst credible are used,
each with different meanings which should be reected both in the selection of input
parameters and in the interpretation of the partial factors stated in the code.

Given the many inuences on limit state prediction by FE analysis described in


Section 6.1.2, the mere application of partial factors in an FE analysis should not be
regarded as a satisfactory verication of ULS. All the inuences should be taken into
account. The validation exercises described in Chapter 7 should include an assessment
of whether an FE analysis has achieved the expected safety margin when performing
ULS design.

Where a design code provides partial factors on ground strength, these are likely to be on
the commonly used w , c and cu shear strength parameters. The majority of designs by
FE analysis are still performed with these basic parameters but the strength derived by a
constitutive model also depends on other inuences (e.g. stress state, discretisation, stiff-
ness and dilatancy). Furthermore, some advanced models have strength dened by other
parameters. Judgement is clearly needed in order to select appropriate partial factors on
ground strength. This can be helped by simulating tests, such as triaxial compression
tests, with appropriate stress paths to assess the partial factors needed to achieve the
appropriate safety margin on shear strength, as demonstrated under Construction stages
in Section 8.4.2.

175

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

6.3.4 Should ground stiffness be factored as well as strength?


It is true that if ground strength were lower than expected, the ground stiffness would
probably be lower than expected too. However, this does not necessarily mean that the
ground stiffness should be factored along with ground strength.

In many cases, ground stiffness does not actually have much inuence on failure loads.
Differences in stiffness between ground layers could have a critical inuence on stress
distributions and factoring stiffness may actually reduce this inuence.

Some design codes and guidance recommend a reduction of ground stiffness in ULS
calculations in cases where ground strength is not a critical parameter, for example to
obtain structural forces for an embedded retaining wall in stiff clay in short-term
situations. If employing the dual factoring approach described in Section 6.3.1, such
cases should be accounted for because the output factoring provides a consistent degree
of conservatism on structural forces in most cases.

Rather like strength factoring, there is a danger that stiffness factoring would provide a
false impression that ULSs have been adequately veried simply because the factoring
was carried out. A better approach, particularly because there is generally more uncer-
tainty on ground stiffness than shear strength, would be to perform a parametric study as
described in Section 7.3.3. This would identify whether stiffness (or other parameters)
has a signicant effect on the occurrence of a ULS and, if so, allow an appropriate safety
margin to be introduced into the design.

6.3.5 How can strength reduction be performed?


As described in Section 6.3.1, input factoring of ground strength is best performed in
dedicated, adjunct ULS analysis stages during the construction sequence rather than
starting the analysis at the initial state with unrealistic, factored ground strength.
Consequently, this requires a reduction of ground strength in the ULS analysis stages
from realistic (e.g. characteristic in Eurocode) values to factored values.

Strength reduction can be performed in two ways:

g By a one-step reduction, i.e. substituting the material for another with the same
parameters except for a lower, factored shear strength. Where the stress state then
violates the failure criterion, stresses are resolved in an iterative manner with the
same stress point algorithm in the software used for a normal elastic-plastic
analysis. A one-step reduction is usually acceptable for basic constitutive models
but the stress paths must be checked to see whether they are credible (see Section
8.4.4). On reducing the strength, the outputs are checked to see whether any
geotechnical failure has occurred. One disadvantage of this approach is that it
does not provide a factor of safety on geotechnical failure.
g By a stepwise strength-reduction procedure included in the analysis software.
Usually, in the rst step the strength is reduced by a user-specied factor on
tan w and c or cu and then the shear strength is successively reduced by an
automatic procedure until a specied number of steps or target factor of safety

176

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

has been achieved. As with the one-step reduction, stresses that violate the failure
criterion need to be redistributed. This can be performed by the same stress point
algorithm used by the software for a normal elastic-plastic analysis, which works
for basic failure criteria such as MohrCoulomb. More advanced failure criteria
need a more complex algorithm, such as that suggested by Potts and Zdravkovic
(2012). Different programs perform the stepwise reduction in different ways and
users should be aware of the method and its shortcomings. One advantage of the
stepwise reduction is that a target strength factor can be reached in order to
obtain factored outputs and then the shear strength further reduced in order to
obtain a factor of safety on ground strength for the rst failure mechanism
occurring during the strength reduction.

Non-convergence of the solution means that ground failure may have occurred, but this
must be checked in the output. Similarly, successful completion of the stepwise strength
reduction does not necessarily mean that ground failure has occurred and this needs to
be checked. There are several ways to identify a failure mechanism in the outputs, as
demonstrated in Section 8.4.4, including the following:

g Plot a graph of displacement at key nodes against strength factor. Failure has
occurred (and the factor of safety value can be determined) at the point where
displacements continue to develop toward high values without further reduction
in ground strength).
g Plot contours or vectors of incremental displacement at the end of the strength-
reduction analysis stage. This often allows failure mechanisms to be identied
more easily than when plotting total displacements for one stage or for all
previous stages.
g Display the yielding elements. Where they join up to form a mechanism, failure has
occurred. Note that the appearance of yielding elements surrounded by elements
still deforming elastically means that a complete failure mechanism has yet to form.

As mentioned under Constitutive behaviour in Section 6.1.2, dilatancy in conned and


undrained problems can lead to a signicant over-prediction of soil strength, so the
dilation angle should be set to zero in such cases. However, Tschuchnigg et al. (2015)
also noted oscillations in outputs of factor of safety and a strong inuence of the dilation
angle even in unconned problems, such as slope stability, when there was a large
difference between friction and dilation angles. Clearly, the dilation angle has a strong
inuence on factors of safety obtained from FE analyses of a range of problems, so its
effect on outputs should always be considered. In particular, stepwise strength-reduction
procedures in different programs may treat the reduction of dilation angle in different
ways. In general, the dilation angle should be reduced with the friction angle until it
reaches zero, then reduction of the friction angle only should continue.

6.3.6 Can geotechnical resistances be calculated by FE analysis?


A geotechnical resistance is the capacity of the ground to withstand a load or displace-
ment without mechanical failure, e.g. bearing resistance, passive resistance, anchor
pull-out resistance.

177

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

As described in Section 6.3.1, specic forms of geotechnical ULS (e.g. bearing resistance
or anchor pull-out) can be veried by comparing factored outputs of load (e.g. anchor
load) with the corresponding factored resistance. This was part of the output factoring
approach.

The resistance is normally calculated in a parallel calculation, either by FE analysis,


other calculation methods or direct testing. It is not normally calculated in the main
FE model simulating the construction sequence because of the additional complexity
that this would involve.

Calculating resistances by FE analysis requires a particular structure to be brought to


failure in a specic way without factoring the strength parameters. This is easier in cases
where the resistance is dened in terms of an externally applied load. For example, the
bearing resistance of a spread foundation is dened as the applied load that causes bear-
ing failure. Therefore, it can be determined by imposing vertical displacement in an FE
analysis (see Figure 5.7) and plotting the output of displacement and corresponding
calculated force, rather like the graph from a plate load test. The bearing resistance is the
value of load where displacement continues to develop without further increase in load.

Output factoring is most suited to structures such as piles, ground anchors, soil nails and
rock bolts whose resistance is also dened in terms of an external load but which is
dependent on groundstructure interface strength rather than ground mass strength.
Since resistances of such structures are heavily dependent on the interface properties,
their determination is not well suited to FE analysis. This is because the remoulding,
mixing and density changes associated with installation effects are not typically
measured in order to provide input parameters for the interfaces in the FE model. The
resistance is measured directly in load tests or determined by direct design methods
calibrated against load tests. Consequently, the only practical way to simulate failure
by FE analysis is to adjust interface properties until outputs match load test results, in
which case the resistance may as well be obtained directly from the load test results.

For structures where failure is caused by geotechnical (internal) loads, such as earth
pressures on a retaining wall, the calculation of resistance by FE analysis is far from
straightforward. Geotechnical loads cannot be arbitrarily increased like external loads
because they depend on the self-weight of the ground. Perturbing forces can be applied
to induce failure, as described by Smith and Gilbert (2011a, 2011b), but in complex cases
it may not be clear where to apply such forces. Hence, when verifying the passive resist-
ance of retaining walls by output factoring, the passive resistance tends to be calculated
by a separate method such as stress eld analysis. This has the signicant drawback of
combining two different calculation methods in an interaction problem and in some
cases it is not possible to verify the ULS (Lees, 2013).

6.3.7 What partial factors are applied to undrained shear strength in


FE analysis?
As described in Section 4.2.4, there are different methods of modelling undrained soil
behaviour. While Methods B and C have their disadvantages, such as not taking into

178

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

account changes in undrained strength due to, for example, consolidation, applying the
partial factor on undrained shear strength cu recommended in a design code is relatively
straightforward.

Method A has some advantages, including taking into account undrained shear strength
changes and its compatibility with advanced constitutive models, but it has the
important difference that cu is calculated by the FE analysis rather than being an input
parameter. As well as potential inaccuracies in the calculation of cu (as described in
Section 4.2.5), this means that the partial factor on cu recommended in a design code
cannot be applied directly to cu . In many constitutive models, shear strength is specied
typically in terms of the effective stress parameters w and c . While design codes provide
partial factors on these parameters, they are intended for drained behaviour where there
is less uncertainty on shear strength. Recommended partial factors for undrained
strength tend to be higher due to the greater uncertainty. However, this higher factor
should not simply be applied to the effective stress shear strength parameters because
(a) there are other inuences on the predicted shear strength (e.g. excess pore pressure)
and (b) the factors of safety in terms of effective and total stresses work in slightly differ-
ent ways, as highlighted by Tschuchnigg et al. (2015) and illustrated in Figure 6.5.

Engineering judgement is needed to select appropriate partial factors on effective stress


strength parameters when using Method A in order to obtain the safety margin intended
by the design code. Then, as described in Section 4.2.5, plots of deviatoric stress in the
outputs should be checked to see that mobilised strength does not exceed the expected
(factored) value of cu according to Equation 4.2.

Figure 6.5 Strength reduction in terms of effective and total stress


tan i
tan i ci ss
FoS = = stre ction
tan f ctive u
cf Effe gth red
n
stre
tan f
cu,i
Total stress
strength reduction
cu,f
cu,i
FoS =
cu,f

ci
cf

Mohrs circle
at failure

179

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

6.3.8 Can ultimate limit states be verified in rock masses using


FE analysis?
The modelling of rock masses by FE analysis was described in Sections 2.2.2 and 2.2.3.
The process of verifying ULSs in soft rocks is similar to that for soils described elsewhere
in this chapter. In hard rocks, the strength of the rock mass is heavily inuenced by
the mechanical properties, spacing, orientation and persistence of the discontinuities,
particularly in low-stress conditions such as in slopes and surface excavations. Therefore,
it is more difcult to take account of the uncertainty in rock mass strength simply by
factoring a strength parameter. Variability in the geometry of discontinuities also needs
to be considered, for instance by adopting different fracture network models in FE
models with explicit discontinuity modelling (see Section 2.2.3).

Hammah et al. (2007) applied FE analysis with explicit modelling of discontinuities


using interface elements to the prediction of factors of safety for various rock slopes.
By employing a strength-reduction approach on the MohrCoulomb friction properties
of both the intact blocks and the interfaces, they obtained results that compared well
with values obtained from discrete element analyses of the same rock slopes.
Although the FE analysis could not simulate very large displacements and detachment
of blocks, they found that factors of safety could be obtained with sufcient accuracy by
identifying only the onset of large displacements.

Since the failure criterion of the HoekBrown model (described in Section 2.3.3) is not
MohrCoulomb, some modication to the yield function is necessary in order to per-
form strength reduction in an equivalent way to MohrCoulomb-based models, as
described by Benz et al. (2007).

6.4. Structural limit states


6.4.1 How are structural limit states verified in soilstructure interaction
problems?
Factored outputs of structural forces (e.g. bending moment, shear force and axial force)
are obtained for structural members from input and output factoring approaches as
described in Section 6.3.1. When employing the dual factoring approach, the most
onerous value from the input and output factoring is used to verify that a structural ULS
is sufciently unlikely to occur. This value is compared with the corresponding structural
resistance calculated and factored in accordance with the relevant structural design code.
If it is less than or equal to the factored resistance, the particular structural ULS has been
veried for that member in that construction stage.

Linear elastic material models provide sufciently accurate predictions of structural


forces in most cases, as described in Section 5.2. It is less likely that predictions of stress
would be accurate, particularly for reinforced concrete where uncracked sections are
assumed, so it is important that structural forces rather than stresses are used to verify
ULS.

The sections used in geotechnical structures tend to be directly comparable with struc-
tural design codes for checking ULS, e.g. rectangular sections for reinforced concrete

180

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Can FE analysis be used with design codes?

and standard sections for steel members, so verication of ULS should be straight-
forward in most cases. With unusual sections, it is usually necessary to divide them
up into regular sections in order to obtain output compatible with structural design
codes.

Some FE analysis programs allow limits on structural forces to be specied for structural
members which are equivalent to resistances. In effect, the program is performing the
comparison between output and resistance automatically. When the output reaches the
limit, any additional load needs to be redistributed. This is a useful feature for checking
combined ground/structure failure mechanisms, e.g. active failure of soil combined with
a plastic hinge in an embedded retaining wall. Similarly, advanced non-linear con-
stitutive models can be adopted for structural materials to simulate yielding and other
features of material behaviour as summarised in Section 5.2.

However, real structures have a limited capacity for yield and could even fail in a brittle
way. The rotation capacity of reinforced concrete sections is limited and depends on the
amount of reinforcement, while only certain steel sections have high rotation capacity, as
described in Section 5.2.2 for sheet piles. Therefore, limits on plastic strains should be
checked carefully rather than allowing unlimited yield.

It is possible to apply factors to structural material strength parameters but note that the
effective stiffness of the structure would be reduced and so structural forces may be
under-estimated due to soilstructure interaction effects (see Section 5.3.1). The output
factoring approach of comparing structural forces with structural resistances is really
intended for linear structural behaviour rather than the highly non-linear behaviour
inherent in many advanced structural constitutive models.

When performing ground strength reduction as part of the input factoring approach (see
Section 6.3.5), it is advantageous to incorporate structural material strength reduction
concurrently in order to identify critical failure mechanisms of combined geotechnical
and structural failures. In addition, structural resistance should still be veried by the
output factoring approach.

REFERENCES
Benz T, Schwab R, Kauther RA and Vermeer PA (2007) A HoekBrown criterion with
intrinsic material strength factorization. International Journal of Rock Mechanics and
Mining Sciences 45(2): 210222.
Fenton GA and Grifths DV (2008) Risk Assessment in Geotechnical Engineering. Wiley,
Hoboken, NJ.
Hammah RE, Yacoub TE, Corkum B, Wibowo F and Curran JH (2007) Analysis of
blocky rock slopes with nite element shear strength reduction analysis. Proceedings of
the 1st Canada-U.S. Rock Mechanics Symposium, Vancouver, Canada, 2731 May
(Eberhardt, Stead and Morrison (eds.)). CRC Press, Boca Raton, FL, pp. 329334.
Lees A (2013) Using numerical analysis with geotechnical design codes. In Modern Geo-
technical Design Codes of Practice (Arnold, Fenton, Hicks, Schweckendiek and Simpson
(eds.)). IOS Press, Amsterdam.

181

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Potts DM and Zdravkovic L (2012) Accounting for partial material factors in numerical
analysis. Geotechnique 62(12): 10531065.
Sloan SW (2013) Geotechnical stability analysis. Geotechnique 63(7): 531572.
Smith C and Gilbert M (2011a) Ultimate limit state design to Eurocode 7 using numerical
methods, Part 1: methodology and theory. Ground Engineering 44(10) October: 2530.
Smith C and Gilbert M (2011b) Ultimate limit state design to Eurocode 7 using numerical
methods, Part 2: proposed design procedure and application. Ground Engineering 44(11)
November: 2429.
Tschuchnigg F, Schweiger HF, Sloan SW, Lyamin AV and Raissakis I (2015) Comparison
of nite-element limit analysis and strength reduction techniques. Geotechnique 65(4):
249257.

182

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.183

Chapter 7
How is the accuracy of outputs
assessed?

7.1. Introduction
7.1.1 Does the accuracy of every geotechnical FE analysis need to be
checked?
In many elds of engineering, once an FE analysis model has been validated for one
application, even for quite complex physical behaviour, it can be applied again and again
in similar applications without the need for re-validation. However, once a geotechnical
FE analysis model has been validated for one application, it does not become auto-
matically valid for other similar applications. This is because each application in ground
engineering is unique. Ground conditions can be very variable and engineering behav-
iour too complicated to be represented by a single, calibrated model. Therefore, each and
every geotechnical FE analysis model needs to be validated. The process to achieve this is
described in this chapter.

Failing to validate an FE analysis model and providing output to a design team that is
inaccurate could lead to incorrect design decisions being made. If the errors were overly
pessimistic, this could lead to the unnecessary costs and delays associated with an un-
economic design. If the errors were overly optimistic, this could lead to an unserviceable
structure (e.g. cracks in walls or malfunctioning machinery), a reduced margin of safety
or even failure, as well as the delays and additional costs associated with overcoming
these problems.

7.1.2 Who is responsible for the accuracy of an FE model?


In any design calculation, the engineering team is responsible for the correctness and
appropriateness of the calculation. The same applies for any geotechnical design
calculations performed by FE analysis. The engineers have ultimate responsibility for the
calculation but should receive some support in the different tasks from other roles, as
described in this section. However, the engineers still have the responsibility to ascertain
whether all the necessary tasks have been completed.

Engineer or user of FE analysis software


The engineer is responsible for the way a program is used. This includes verifying that a
program works properly on the computer and operating system version being used (see
Section 7.2.1). It also includes selecting the program options, constitutive models and
input parameters and creating the geometry in order to create an FE model that provides

183

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

the critical outputs with sufcient accuracy. The engineer must also perform a proper
validation of the model and assess its reliability, as described in this chapter.

The engineer should also report to his/her management or the client when there is
insufcient information, e.g. regarding ground conditions or a proposed structure, if this
prevents an adequate FE model being created.

The engineer may also take on some developer roles when creating user-dened sub-
routines or constitutive models see Software developer in this section.

Engineering manager
The engineering manager has overall responsibility for the reliability of the FE
model and seeing that proper analysis practices have been followed. Managers without
sufcient experience in FE analysis would need to consult some guidance in this
area. NAFEMS, for example, provide documentation including guidebooks on the
validation of geotechnical numerical models (Brinkgreve, 2013) and quality assurance
procedures (e.g. Chillery, 2014). However, any engineering manager should be able to
assess the quality of FE model outputs using his/her wider experience of geotechnical
engineering.

The engineering manager also has the important responsibility of ensuring that the
engineer has sufcient competency to perform particular FE analysis tasks, including
background knowledge of the geotechnical materials and problems being simulated.
As well as maintaining existing levels of competency and keeping engineers up to date
with developments, this requires mentoring and training to be arranged for engineers
to gain new competences in order to undertake more FE analysis tasks.

Client
The client has the responsibility for providing sufcient resources and allowing sufcient
time to complete FE analysis tasks properly, in order to benet from the potential con-
struction cost savings brought by more advanced analysis. Additional investment in site
investigations is also required to provide the more detailed ground information needed
by FE analysis compared with conventional design methods.

Software developer
The software developer is responsible for programming the code of the FE analysis
program, checking that the code works correctly and that the theory behind the program
has been correctly applied (see Section 7.2.1). He or she must also properly document
the models and methods implemented in the software and make this documentation
available to the user.

Program users also take on some developer responsibility when user-developed sub-
routines or constitutive models are implemented into a program. They need to check that
the added feature works correctly and also interacts properly with other parts of the
program.

184

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How is the accuracy of outputs assessed?

7.1.3 What are the potential sources of error in an FE analysis?


The following is a summary of some of the potential sources of error in any FE analysis.
Readers should refer to Brinkgreve (2013) for a more comprehensive description of the
potential sources of error and for a very useful checklist.

The power and exibility of FE analysis brings with it a high potential for things to go
wrong. The techniques to avoid many of the potential errors listed here are described in
the preceding chapters. Indeed, the likelihood of error is clearly very much reduced when
software users have the competency to perform the particular analysis task in hand.

Operating systems and hardware


With the almost innite possible combinations of hardware, operating systems and
other software packages running simultaneously with the FE analysis program, small
differences in the calculation process that cannot be foreseen may occur. These small
differences can become magnied later in the calculation to such an extent that signi-
cant errors may be introduced. This can be checked by running the same analysis on
different combinations of hardware and operating systems.

Misuse of methods
Errors can occur when FE analysis is applied to cases to which it is not well suited or
when the user misunderstands the theory behind the software and sets up an analysis
model in an incorrect way. For example, using FE analysis of a continuum with small
deformation theory to model the effects of large deformations (see Section 1.4.3) or
toppling mechanisms in rock masses (see Section 2.2.3) could lead to large errors in the
outputs because the analysis methods are fundamentally inappropriate. Such errors can
be avoided when users of software understand the methods being employed as well as the
limitations of the methods.

Bugs
In spite of a software developers best efforts to identify and solve bugs in a program,
they may still exist, particularly in rarely used parts of the program or with unusual
combinations of conditions. Therefore, bugs are always a potential source of error and
users should beware.

Tweaks
Tweaks are known features in a program that introduce error but are needed to improve
the robustness of software. For example, the MohrCoulomb failure criterion in princi-
pal stress space is formed of a set of yield surfaces that meet at corners. The corners
create numerical difculties because two yield surfaces can be active at the same time
when stress states are at a corner. Programs overcome this in different ways, such as
rounding off the corners, which leads to slightly different results being obtained from
different programs.

Therefore, it is important to know well the analysis software being used in order to be
aware of its tweaks and approximations that may become signicant in certain
applications.

185

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Calculation approximations
Non-linear problems require the governing FE equations to be solved in incremental
form. A number of solution strategies are available and the choice of strategy has a
strong inuence on the accuracy of the outputs. None of the strategies are exact but
require a criterion (usually a tolerance on equilibrium error in static loading problems)
dening when the solution is accepted. The specied tolerance is a compromise between
accuracy and efciency. A tight tolerance leads to higher accuracy but longer compu-
tation times, while a loose tolerance leads to lower accuracy but shorter computation
times. The tolerance value should not be viewed as a direct measure of calculation accu-
racy because other factors play a role. For instance, an ill-conditioned stiffness matrix
(due to, for example, large stiffness or permeability ratios) could have a detrimental
effect on calculation accuracy compared with a conditioned stiffness matrix with the
same tolerance value. Furthermore, the different stress point algorithms used to calcu-
late stress changes in increments have a strong inuence on accuracy.

Tolerances that may or may not be controlled by the user are often used in other parts of
FE analysis software for greater robustness, for instance to deal with zero stress states in
stress-dependent stiffness models. It is important that users understand the tolerances
and calculation methods used by the software in order to assess whether they achieve the
required degree of accuracy for each analysis task.

Input errors
FE analyses require a large amount of input data which increases the likelihood of some
data simply being entered incorrectly. To help avoid such errors, the input data should
be checked. Checking can be facilitated when the software contains features to tabulate
input data and displays warning messages when data fall outside of credible ranges.

Constitutive model
As described in Chapter 2, no constitutive model captures all aspects of ground behaviour,
so one is selected that includes the important aspects but ignores others. This introduces
some approximation.

Most models have been calibrated only over limited stress space (e.g. triaxial, hollow
cylinder), while on implementation in FE analysis they may be used in general stress
space where they have not been calibrated and where some tweaks to the model may have
been necessary in order to extend them to general stress space.

Linear elastic models are often adopted for structural materials, which is acceptable in
many cases (see Section 5.2) but is an approximation of real structural material behav-
iour. Further simplifying assumptions are also made at groundstructure interfaces
(Section 5.1.3).

Obtaining geotechnical parameters


Even with the best efforts to minimise errors during sampling, testing and interpretation,
they will still occur, as described in Chapter 3. The parameter validation procedures
described in Section 3.4.2 should identify any signicant errors but there are still further

186

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How is the accuracy of outputs assessed?

sources of error in the adoption of model parameters. Stress paths during testing are
unlikely to match those occurring in the ground during construction, so the derived
parameters may not be totally representative of eld behaviour. Parameters derived from
triaxial test conditions, for example, may need some manipulation for plane strain
conditions to take account of intermediate effective stress values (as described in
Section 3.4.1), which introduces further uncertainty.

There is also natural variation in ground parameters (aleatoric errors) and the locations
of ground layers, but only tiny fractions of the ground are investigated. Consequently,
simplied geometries are adopted for ground layers and parameter variation within
ground layers tends to be ignored. This leads to the conservative selection of model
parameters and the assumption of homogeneous ground layers. However, the effect
of the natural variation on key outputs can be assessed in parametric studies (see
Section 7.3.3).

Geometrical simplifications
As described in Section 1.2, simplications are introduced to the geometry to make the
modelling process more manageable and efcient. Simplications include adopting 2D
axisymmetric or plane strain assumptions, eliminating details from the geometry,
particularly from the periphery of the area of interest, that are considered unimportant
and locating the boundaries to the FE model such that the ground far away from the
area of interest can be excluded. If properly judged, such simplications should not
introduce signicant error to the key outputs, as described in Section 1.2.

Discretising the geometry into a mesh of elements also involves approximation but pro-
ducing an efcient mesh can help to minimise the error, as described in Section 1.3.2.
Adopting non-continuum elements for structures is a common simplication that
introduces some error, as summarised in Table 5.2. Structural connections are also
usually idealised in some way, as described in Section 5.1.7.

Construction process simplifications


The rst stage of any FE analysis involves establishing the initial stresses in the ground.
Due to the same uncertainties in the ground parameters, this involves some approxi-
mation of ground layer geometry, density, stress ratio and groundwater conditions.
Where historical man-made activities have affected in situ stresses, further approxi-
mation is introduced in simulating these. Advanced constitutive models require
additional initial state parameters, such as pre-consolidation stress, that are often
difcult to estimate or determine. The errors introduced at the initial stage are likely
to inuence all the subsequent stages of the analysis, so it is very important to keep these
to a minimum.

Simulation of construction processes involves some approximation since, while real


construction activities are essentially continuous, simulations require such processes to
be discretised into large steps in order to be manageable. Processes tend to be simplied,
with installation effects, for example, often being ignored (see Section 1.4.2).

187

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Loadings are often idealised into point or line loads that do not exist in reality. Even
distributed loads are converted to equivalent nodal loads whose accuracy depends on the
density of nodes in the mesh (see Section 5.1.8).

Groundwater conditions
Groundwater levels, both internal and external, vary in reality due to tidal, seasonal and
other effects but assumptions are usually made, often placing them at maximum or
minimum levels for conservatism, which introduces error. Hydrostatic conditions are
regularly assumed although, in reality, they may only be close to hydrostatic. Ground-
water ow and consolidation analyses are always subject to the high degree of uncertainty
over the permeability of the ground.

An assumption of undrained behaviour is an approximation because even low-


permeability soils experience partial drainage at normal loading rates, while the methods
of simulating undrained behaviour (see Section 4.2.4) also involve some approximation.

7.2. Assessing accuracy


7.2.1 What is the difference between verification and validation?
Verication essentially means checking that a program works correctly. It involves
checking that the underlying mathematical model has been implemented in the computer
program correctly and this is normally carried out by reproducing a theoretical solution
until sufcient accuracy has been achieved. It also involves identifying and solving bugs
in the software code. The software developer is responsible for verication (except in the
cases of user-dened sub-routines and constitutive models where the user has responsi-
bility for verication). Nevertheless, some bugs may still remain within the program and
the user should beware of these. Verication also involves checking that a program
works correctly on a particular computer and operating system version and this is the
responsibility of the user. This can be performed by running an FE analysis of a previous
project or case study with the new computer or operating system version and comparing
its output with that from a known compatible computer and operating system.

Validation means checking whether an FE model represents reality sufciently accu-


rately. Therefore, while a mathematical model may have been implemented correctly
(as checked by verication), the mathematical method may still be inappropriate for the
problem being simulated (misuse of method) or incorrect input parameters may have
been specied. The purpose of validation is to check whether the mathematical methods,
input parameters and set-up of the FE model have produced sufciently accurate
outputs. This is the responsibility of the user and the techniques for doing this are
described in the next section.

7.2.2 How are outputs checked for accuracy?


The full output from a typical FE analysis is too voluminous and the calculations too
complicated to be checked explicitly as in conventional calculations. Therefore, agree-
ment needs to be reached at an early stage between project stakeholders on an acceptable
level of checking and documentation, which is likely to differ from the requirements for
conventional design calculations.

188

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How is the accuracy of outputs assessed?

This section contains a summary of the common methods used to check FE analysis
outputs. They generally involve comparing outputs with expected outcomes (based on
experience or published case histories), other calculation methods or ongoing site moni-
toring. Above all, a check of FE analysis outputs should be based on a combination of
these methods rather than relying on just one.

More detail and further references on each method can be found in the guidebook on
validation by Brinkgreve (2013).

Initial check
Signicant errors can be identied readily from a quick glance at the outputs by
engineers with relevant experience in the structure types being simulated. Any values that
are very different from what would be expected based on experience indicate that there
is a signicant error in the FE analysis. The following list contains outputs that are
commonly viewed for each analysis stage as part of an initial check:

g deformed mesh, using both true and exaggerated scales, to check that
deformations look reasonable
g steady-state and excess pore pressure contour plots to check that the imposed
groundwater ow boundary conditions and drainage conditions worked as expected
g vector plots of principal stress and strain directions to check that they are
consistent with applied loads
g stress paths for key integration points to check that they are as expected and that
constitutive model and input parameter selections were appropriate
g contour plots of normal and shear strain to check that they seem reasonable and
to help highlight any problem areas in the FE model
g plot state parameters to check that advanced constitutive models are performing
as expected, e.g. stiffness in strain-dependent models.

On identifying an issue, further investigation of the output and input data should reveal
the source of the error.

With 3D models, output for locations within the model is not immediately apparent. It is
important to display data on sections that step sequentially through the model in order
to reveal the output in the heart of the model.

Comparison with known solutions


For each of the common geotechnical structure types there are alternative analytical
methods to FE analysis. They are usually less sophisticated than FE analysis, so typically
provide a single output (e.g. displacement or failure load) for a single structure type (e.g.
strip foundation or slip circle). Many of the methods are well known, reliable and simple
to use. Therefore, they provide an alternative calculation method whose output should
be reasonably close to the output from an FE analysis of the same problem. Comparing
the two outputs with each other provides a means of validating the FE analysis. How
close the outputs should be depends on the appropriateness of each method for the
problem being analysed.

189

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

The following is a list of commonly used alternative analytical methods:

g Analytical (or closed-form) solutions: for relatively simple problems, a set of


equations describing behaviour can be solved exactly by symbolic manipulation.
These tend to be available for linear elastic problems such as the settlement
beneath a rigid strip foundation or for beam bending in structures, or for steady
ow in conned and unconned porous media. These are generally more useful
for verication than validation because of their limited scope.
g Limit equilibrium and stress eld methods: these provide simplied predictions of
failure loads on pre-dened mechanisms. For example, slope stability solutions,
CaquotKerisel equations for active and passive pressure with soilwall friction,
Prandtls solution for bearing capacity on cohesive soil.
g Design charts and empirical design methods: largely based on case study
observations, a number of design charts and methods have been published for
most geotechnical structures, particularly for deformations where, in the past,
practical calculation methods did not exist. For example, Clough and ORourke
(1990) published design charts for wall movements and ground surface settlements
around retaining walls. Such methods are rather broad and approximate but
provide an indication of the order of magnitude of outputs to be expected from
FE analyses of similar problems.

As an example, the validation of outputs of ground surface settlement behind an


embedded retaining wall supported at the top is shown in Figure 7.1. The empirical
prole was obtained from Clough and ORourke (1990) and this is compared with
outputs from two FE analysis models (A and B) that adopted different constitutive
models for the soil.

The settlement output from analysis A was in the opposite direction and plotted as a
different shape to the empirical prole. Such instances suggest that the analysis had a

Figure 7.1 Example validation of ground surface settlement outputs

15
FE analysis A
10

5 Horizontal distance behind wall: m


Settlement: mm

0 10 20 30 40 50
0

5
Empirical profile
10

FE analysis B
15

190

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How is the accuracy of outputs assessed?

serious error and some aspect, such as the selection of constitutive model, was in error.
The output from analysis B showed a settlement prole with the same shape and direc-
tion as the empirical prole but with a small difference in the values. In such cases the
ground and structure are probably being modelled in the correct way while some minor
adjustment to the input parameters would achieve a closer match with the validation
data, if that was considered necessary.

In analysis A, an LEPP model was adopted for the soil which resulted in an over-
prediction of heave on excavating the soil in front of the wall. Therefore, the constitutive
model was inappropriate as proved by this validation exercise. In analysis B, a more
appropriate non-linear strain-dependent stiffness model was adopted, hence the better
match with the validation data. The slight difference between the analysis output and the
empirical prole was considered to be due to the effect of the wall stiffness and, since the
empirical prole is only approximate, it would probably be unnecessary to try to achieve
a closer match with the empirical prole in this case.

Comparison with site monitoring data


Monitoring data collected during construction should be compared with the outputs
from FE analyses of the same construction activities. This will help validate the FE analy-
sis and allow changes to be made in order to improve the reliability of predictions for the
remainder of construction. Such observational approaches to design are discussed further
in Section 7.3.1. Note that a close match between FE analysis outputs and monitoring
data should only be expected if most probable estimates of input parameters, loadings,
etc. have been adopted. In many FE models undertaken for design purposes, more
conservative estimates of parameters and assumptions are usually taken which would
normally result in more onerous outputs compared with the monitoring data.

Comparison with other numerical analysis software


When additional software is available, it is very useful to perform the same simulation
on another veried FE analysis program and/or a program using a different numerical
modelling technique, such as the nite difference method, beam-spring method or limit
analysis (Sloan, 2013), also on different hardware. The outputs are likely to differ due to
the different models and methods employed in each program. However, if the outputs
are reasonably close, perhaps within 10%, then this would provide some validation of
the original FE analysis, provided that the same errors were not repeated in each analysis
model.

Case histories
Published case histories of similar geotechnical structures in similar ground conditions
can provide a useful source of validation data. They must be well documented in order
to be useful, including full details on the construction sequence and ground properties.
Ideally, the case histories would include FE analysis simulations in order to learn from
the methods, models and parameters used successfully in similar projects. Detailed case
histories with monitoring data can also be useful since these can be simulated and
outputs compared with the monitoring data in order to validate the modelling approach
to a particular problem.

191

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

7.3. Managing errors


7.3.1 If inaccuracy exists, how should this be managed?
A judgement needs to be made about an acceptable level of accuracy. Clearly, an exact
match between FE model and reality will never be achieved and nor is it necessary. For
design, it is important that errors cause the critical outputs to err on the conservative
side. After that, the appropriate degree of accuracy depends on the individual needs
of each project. Where the potential cost savings in construction can justify additional
effort on FE analysis at the design stage, then the FE analysis can be improved with
more detailed study, perhaps by a specialist, and possibly additional site investigation.
In other cases, a less accurate but conservative FE analysis may be sufcient to meet the
needs of the project.

Sensitivity analyses are required to identify the parameters with greatest inuence on
the key outputs, followed by parametric studies in order to assess the reliability of FE
analysis outputs, as described in Sections 7.3.2 and 7.3.3.

The design should be able to withstand, by adequate margins, the range of outputs
arising from the parametric study. Additionally, an observational approach can be
employed where, if monitoring during the early stages of construction is found to lie
within the more favourable ranges of output from the parametric study, then a more
cost-effective design can be introduced during construction. This requires less control
than the observational method (which can bring greater cost savings in some situations
see Section 7.3.4) because if anybody neglects to study the monitoring data, then the
more pessimistic design prevails.

7.3.2 What is the difference between sensitivity analysis and


parametric study?
Both involve evaluating the effect of input parameters, which can include constitutive
model parameters, boundary conditions (e.g. imposed loads or displacements) or
geometrical values (e.g. excavation depth), on the critical outputs required from the
FE analysis.

A sensitivity analysis involves varying all or certain input parameters to determine which
affect the outputs the most.

A parametric study involves varying input parameters between certain ranges in order to
determine the dependency of key outputs on input parameter uncertainty. To study all
the parameters in this way can be very time-consuming, so more attention is paid to the
critical parameters as identied in the sensitivity analysis. Typically, three values of each
parameter are assessed: one at each end of a range (as discussed in Section 7.3.3) and the
deterministic value.

On completion of the parametric studies, as well as gaining an improved understanding


of the behaviour of the structure, a range of plausible responses of the structure will be
obtained, allowing the reliability of the design to be assessed.

192

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How is the accuracy of outputs assessed?

7.3.3 Over what range of parameters should a parametric study be


undertaken?
Where geotechnical parameters are considered to vary randomly across a site, a para-
metric study could be undertaken where each parameter is varied between the maximum
and minimum values interpreted from the site investigation data. However, this may
result in an exaggerated range of outputs because it is highly unlikely where parameters
vary randomly that the maximum and minimum parameter values would ever exist in a
large enough volume of ground to govern the outputs in this way.

It is more appropriate to vary parameters over a narrower range, but one that still takes
account of the variability of a parameter about its mean. The statistical measure of
this variability is the standard deviation s which represents the range of a parameter
each side of the mean within which about 68% of values would lie assuming a normal
distribution. In most cases it would be appropriate to vary a parameter across a range of
about 1s above and below the mean.

But how can s be calculated? Rarely is there sufcient data available from a site
investigation for it to be calculated directly from the data using a formula. So, it can
be estimated using the 3s rule as illustrated in Figure 7.2 and described by Duncan
(2000). Since 99.7% of values in a normal distribution lie within +3s of the mean, for
practical purposes +3s can be considered the upper and lower bound to all the data and
this is probably easier to estimate than the +s bounds. As shown in Figure 7.2, the
mean, highest conceivable and lowest conceivable lines are drawn on the data.

Figure 7.2 The 3s rule and range for parametric study of arbitrary data set

Arbitrary parameter
0 20 40 60 80
0

4
Depth

able

Hi

6
g
he
ceiv

st
co
con

nc
Mean

eiv

8
est

ab
le
Low

10 3 + +3
Suitable range for
parametric study

193

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Table 7.1 Typical coefficient of variation values for geotechnical


parameters (Duncan, 2000)

Parameter Coefficient of variation V: %

Weight density, g 37
Buoyant weight density, gb 010
Friction angle, w 213
Undrained shear strength, cu 1340
Undrained strength ratio, cu /sv 515
Compression index, Cc 1037
Pre-consolidation pressure, pc 1035
Permeability k of saturated clay 6890
Permeability k of partially saturated clay 130240

Since all the data should lie within the outer lines, these often need to be placed further
apart than might intuitively be judged based on a limited amount of data. However, any
data considered to be anomalous should be ignored because these are not representative.
The +s lines can then be added one-third of the distance between the mean and +3s
lines and these represent a suitable range for parametric study in most cases.

Duncan (2000) collected published values of standard deviation for a number of


different geotechnical parameters expressed as coefcients of variation V which can be
used to validate derived values. A selection of the values is shown in Table 7.1. s can
be determined from V simply by multiplying the mean of a particular data set by V.

A parametric study may then be performed using the following procedure. The FE
analysis is rst performed with deterministic values of parameters. These can be the
average (in situations where most probable predictions are required) or more conserva-
tive values if the design is in accordance with a design code. The parameters to be
included in the parametric study (possibly identied from a sensitivity analysis) are then
changed to the top or bottom of their range (mean +s) such that each combination is
used as input data to a separate FE analysis. All other parameters not included in the
parametric study remain at their deterministic values.

For example, in a simple FE analysis of an embedded wall supported at the top in a


granular soil simulated with a LEPP model with a MohrCoulomb failure criterion and
stiffness increasing with depth, a parametric study was undertaken as follows. Two
parameters were studied: the friction angle w and Youngs modulus E of the soil for
which coefcients of variation of 10% and 40%, respectively, were determined. The
FE analysis was run ve times with the combinations of parameters shown in Table 7.2,
starting with the deterministic values (which were the mean values in this case) and then
with each parameter either at the top end of the range (mean + one standard deviation)
or at the bottom end of the range (mean one standard deviation). All other parameters
remained constant at their deterministic values.

194

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How is the accuracy of outputs assessed?

Table 7.2 Example parametric study combinations

Name w value E value

Det. wdet E det


Comb. 1 wmean + sw E mean + sE
Comb. 2 wmean sw E mean sE
Comb. 3 wmean + sw E mean sE
Comb. 4 wmean sw E mean + sE

Outputs of retaining wall deection and bending moment obtained from these com-
binations are shown in Figures 7.3 and 7.4, respectively. From these it is immediately
apparent how much the outputs change in response to the two soil parameters varying
within plausible ranges of values around the mean. It facilitates greatly in assessing the
reliability of the design should these soil parameters be higher or lower than the inter-
preted mean value.

Also added to the bending moment diagram in Figure 7.4 are the maximum and
minimum values that might be determined using a partial factoring approach. Output
factoring (see Section 6.3.1) was applied to the outputs of maximum and minimum
bending moment from the deterministic analysis by multiplying them by 1.35 which is
the recommended permanent load (effect) factor in Eurocode 7 (CEN, 2004), for
example. The factored minimum value was about the same as that obtained from the
parametric study, while the factored maximum value was somewhat higher.

Figure 7.3 Parametric study outputs retaining wall deflection

Wall deflection: mm
0 10 20 30 40
0

2
Det.
Comb.1
4
Comb.2
Comb.3
Depth: m

6 Comb.4

10

12

14

195

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 7.4 Parametric study outputs retaining wall bending moment

Wall bending moment: kNm/m


300 200 100 0 100 200
0

det. min. 1.35 2


Det.
4 Comb.1
Comb.2
Comb.3
Depth: m

6
Comb.4

10

12
det. max. 1.35

14

This suggests that the 1.35 factor would have been appropriate to take account of
uncertainty in the soil parameters in the upper part of the wall but rather high for the
lower part. This highlights the advantage of performing even a simple reliability analysis
compared with using the one size ts all partial factoring approach.

However, such a simple probabilistic approach considers only randomly variable


parameters. If there were clear zoning of parameters with softer ground encountered
in some areas more than others, then this would either need to be considered explicitly
in the FE model or more advanced stochastic methods would be required that consider
the spatial variability of input parameters (refer to Fenton and Grifths, 2008, for
example).

7.3.4 What is the observational method?


There is more to the observational method than simply observing the implementation of
a design. It is a carefully considered approach to design, well suited to numerical analy-
sis. It was developed by Peck (1969) and is also summarised by Clayton et al. (1995).
Some key elements to the method are as follows:

g sufcient site investigation to determine a ground model and parameters, but not
necessarily in detail
g interpretation of the most probable and the permissible range of ground
conditions
g numerical analyses with parametric studies to determine most probable and
permissible ranges of outputs
g selection of parameters to be monitored during construction

196

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
How is the accuracy of outputs assessed?

g documentation of most probable and permissible range of monitored parameters


as obtained from the analysis output
g pre-determination of design or construction sequence modications for every
foreseeable deviation of monitored parameters from most probable values
g start of construction and monitoring
g implementation of design changes and construction sequence modications in
response to actual observed behaviour.

There is a greater potential for cost savings than with the more simple observational
approach (see Section 7.3.1) because pre-determined design options are selected during
construction in response to better or worse than expected monitored performance.
However, the observational method requires strict control with clearly dened responsi-
bilities for studying and interpreting the monitoring data since, if the data were not
studied in a timely manner, a worse than expected performance may go unnoticed, lead-
ing to a serious failure. Ideally, the engineers who performed the parametric studies
should also study and interpret the monitoring data since they would understand most
the implications of the data on the design options.

REFERENCES
Brinkgreve RBJ (2013) Validating Numerical Modelling in Geotechnical Engineering.
NAFEMS, Hamilton.
CEN (2004) EN 1997-1 Eurocode 7: Geotechnical design Part 1: General rules. CEN,
Brussels.
Chillery M (2014) NAFEMS Simulation Handbook Quality Management. NAFEMS,
Hamilton.
Clayton CRI, Matthews MC and Simons NE (1995) Site Investigation, 2nd edn. Blackwell
Science, Oxford.
Clough GW and ORourke TD (1990) Construction induced movements of insitu walls.
Proceedings ASCE Conference on Design and Performance of Earth Retaining Structures,
Cornell, ASCE Pub. no. 25, pp. 439470.
Duncan JM (2000) Factors of safety and reliability in geotechnical engineering. Journal of
Geotechnical and Geoenvironmental Engineering 126(4): 307316.
Fenton GA and Grifths DV (2008) Risk Assessment in Geotechnical Engineering. Wiley,
Hoboken, NJ.
Peck RB (1969) Advantages and limitations of the observational method in applied soil
mechanics. Geotechnique 19(2): 171187.
Sloan SW (2013) Geotechnical stability analysis. Geotechnique 63(7): 531572.

197

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.199

Chapter 8
Examples

8.1. Introduction
Three examples are presented in this chapter to help illustrate some of the points made in
the preceding chapters. The relevant sections from Chapter 1 on setting up an FE
analysis model are shown in brackets (e.g. 1.2.3) in the second section of each example,
where more information to justify a particular decision can be found.

The examples have been kept relatively simple in order to illustrate the decision-making
process, parameter determination and validation of outputs as described in the preced-
ing chapters clearly without being burdened by the level of detail associated with real
projects. They are not intended to be benchmarks or to be used for validation purposes,
nor should the outputs or ndings from these examples necessarily be expected to be
applicable to other structures of the same type. Readers should refer to published case
histories relevant to the ground conditions and structure types in their particular projects
for more guidance and for sources of validation data.

In an effort to keep this book entirely software-neutral, the specic software and, some-
times, the specic constitutive models employed in these examples are not mentioned.
However, enough information is provided for readers to prepare similar models using
their preferred software and to compare outputs with those presented here for each
example.

8.2. Raft foundation with settlement-reducing piles


example
8.2.1 Summary
This example concerns the construction of a raft foundation to support a multi-storey
building. The ground was composed of a medium-dense, lightly over-consolidated sand
and because settlements were likely to be large, consideration was given to the use of
settlement-reducing piles.

The particular features of this example include:

g 3D model with a plane of symmetry


g K0 initial stress procedure
g double-hardening, hyperbolic constitutive model with stress-dependent stiffness
g obtaining geotechnical parameters from pressuremeter and pile load test data
g drained displacement analysis

199

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

g both continuum and non-continuum elements were used for the raft and piles and
the outputs compared
g validation using alternative analysis methods.

8.2.2 Setting up the FE analysis model


Justification for using FE analysis (1.1.1)
The alternative analysis methods (e.g. boundary element method) adopt assumptions
such as a single size of pile or that the piles are distributed on a regular grid across the
entire area of the raft. In this example, a small number of piles of different lengths placed
near the centre of the raft needed to be considered. Such irregular geometry was more
suited to FE analysis.

Aims of the model (1.1.2)


1 to predict the settlement distribution across the raft with and without the
proposed settlement-reducing piles in order to select the appropriate option or
assess alternatives
2 to predict the loads in the settlement-reducing piles in order to check their design
3 to provide appropriate coefcients of subgrade reaction for the raft to be used in
the structural design of the raft
4 to predict bending moment in the raft in order to provide validation data for the
structural analysis model.

Geometrical simplifications (1.2.1 and 1.2.2)


This site was surrounded by roads on all four sides, as shown in Figure 8.1, so there was
no signicant interaction with nearby buildings and these were omitted from the model.
Similarly, interactions between the road, its associated underground services and the raft
were not considered signicant. Therefore, the geometrical features included in the FE
analysis model could be simplied to those only of the proposed raft foundation and
piles. The layout of the settlement-reducing piles is shown in Figure 8.2.

Regarding the superstructure, while its stiffness would have a signicant effect on the
behaviour of the raft foundation (as described in Section 5.3.1), its inclusion in the geo-
technical FE analysis model was not necessary in order to take this into account. Rather,
coefcients of subgrade reaction obtained from the output of the geotechnical FE model
were adopted in the springs representing the soil in the structural FE model. Revised
foundation loads from the structural model were then applied back into the geotechnical
FE model. Although not presented in this example, in practice this iterative procedure
between the geotechnical and structural models until the foundation loads and co-
efcients of subgrade reaction in both models are in approximate agreement provides
an efcient means of taking account of superstructure stiffness effects on raft foundation
behaviour. The core walls had a signicant stiffening effect on the raft so these were
included in the geotechnical FE analysis model, but rather than the full height of the core
walls, they were included to one-storey height (3.5 m).

The square shape of the raft foundation and inclusion of piles precluded the use of a 2D
plane strain or axisymmetric assumption, so a 3D model was necessary.

200

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.1 Raft foundation example geometry and loads

All columns 1 m 1 MN
square, on 10 m Flat ground level
grid except at core 0.0 m 2 MN
level ROA
D
2 MN
2 MN
1 MN
5 MN
3m 1m
4m 1m
3m

5m
2 MN
AD

2 MN
RO

5m
5 MN

AD
RO
6m
1 MN
2 MN
2 MN

2 MN Plane of
symmetry
Raft at ground level, 1 MN
32 m square, 1.2 m-thick
reinforced concrete Core walls: 300 mm-thick
ROA reinforced concrete,
D
250 kN/m vertical line
load on all walls

Figure 8.2 Plan of raft showing settlement-reducing pile locations

5m 5m 5m 8.5 m

5 6
5m

Pile no. 1 2 3 4 2.5 m


2.5 m

5m

All piles 600 mm dia.


Pile with toe level 15 m
Pile with toe level 20 m

201

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

However, a plane of symmetry existed in the structure (see Figure 8.1) and the ground
conditions were uniform, so only half of the geometry needed to be modelled.

When continuum (volume) elements were used to represent the raft, the geometry of the
raft in the model matched reality, i.e. the top surface of the raft was at ground level and
the raft volume was embedded within the ground. However, this was not possible when
using non-continuum (shell) elements. If the shell elements were placed at ground level,
the raft would not benet from the surrounding overburden and unrealistic, local failures
would result at the raft edges. Furthermore, the tops of the piles would be at the wrong
level. If the shell elements were placed at formation level such that they were embedded
in the ground, vertical slopes would be left in the ground around the raft that would
collapse. To overcome these problems, the ground level in the model was set at forma-
tion level (1.2 m) with a surcharge placed on the ground surface to represent the weight
of the omitted ground. Such an assumption omitted the strength and stiffness of the
ground above formation level, but was conservative.

Model boundaries and fixities (1.2.3 and 1.2.4)


Initially based on rules of thumb (Figure 1.7), the boundaries to the model were placed at
90 m (3B) from the raft. It was found that the vertical boundaries could be moved
inwards to about 60 m (1.8B) and 75 m (2.3B) from the raft edge in the X- and Y-axis
directions, respectively, as shown in Figure 8.3(a), without introducing any signicant
boundary effects on the key outputs (except on the plane of symmetry).

The bottom boundary was xed in all three axis directions and the vertical boundaries
(including the plane of symmetry) were xed only in the horizontal direction perpendicu-
lar to the boundary plane. When non-continuum (shell) elements were used for the raft,
its rotation at the plane of symmetry perpendicular to the boundary was xed in order to
simulate the rotational restraint from the raft on the other side of the plane of symmetry
excluded from the model.

Finite element mesh (1.3)


Second order 10-node tetrahedral elements with three degrees of freedom per node were
used to model the ground and structures (when solid elements were used).

When non-continuum elements were used to represent the structures, second order
6-node triangular shell elements were used to represent the raft foundation and core
walls. These elements had ve degrees of freedom per node (axial displacement (1), trans-
verse displacement (2) and rotation (2)). Since the raft was quite thick (1.2 m), it was
probably around the limit of what could normally be modelled reasonably accurately
with shell elements, hence why both continuum and non-continuum elements were
adopted in this case. The piles were modelled with special 3-node beam elements with
six degrees of freedom per node (three translational and three rotational), as described
in Section 5.1.4.

Since the core walls were relatively thin (0.3 m thickness), these were modelled using shell
elements in all the analyses and continuum elements were not used. Where shell elements

202

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.3 (a) FE mesh with solid elements for raft; (b) Close-up of FE mesh with solid elements
for raft

150 m Side fixed in


90 m Y direction
No fixity on
top surface

Side fixed in
X direction
90 m

Side fixed in Side fixed in


Y direction X direction

Y
Base fixed in X,
Y and Z directions
X
(a)

Predefined 1 m wide Column locations with node at


prism through full centre and elements defining
length of raft 1 m2 area for distributed load

Shell elements on plane of symmetry had


material properties of half-thickness wall Z
(150 mm) and half line load (125 kN/m)
Y
Shell elements for core walls
extended into solid raft elements
to simulate fixed connections X
(b)

were used for the raft, a fully xed connection with the core walls was automatically
formed. Where solid elements were used for the raft, a pinned connection would be
formed with the core walls composed of shell elements. Since concentric, vertical loads
were applied to the core walls, a pinned connection would probably be adequate in this
case, but in order to be consistent with the shellshell connection in the other case, the
core wall shell elements were extended 0.5 m into the solid raft elements in order to
simulate a fully xed connection, as described in Section 5.1.7.

203

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.3 (c) FE mesh with solid elements for raft and piles (front elements hidden); (d) Close-up
of FE mesh with shell elements for raft

(c)

Surcharge applied to top surface


around raft to represent weight
of ground omitted from mesh Column locations with node at
centre and elements defining
1 m2 area for distributed load

Shell elements and


top surface at
Z
1.2 m level
Shell elements on plane of symmetry had Y
material properties of half-thickness wall Shell elements at
(150 mm) and half line load (125 kN/m) plane of symmetry
fixed against rotation X
(d)

Interface elements were adopted between the piles and ground when solid elements were
used for the piles, while special interface elements were used between the piles and
ground when special beam elements were used. A perfectly rough interface (no interface
elements) was assumed between the raft and the ground because no signicant slippage
or uplift was expected.

The meshes were generated as shown in Figure 8.3 with smaller elements located where
steep stress and strain gradients were expected. During simulations of pile load tests, it
was found that outputs were particularly sensitive to mesh renement when using either
solid or special beam elements to represent the piles. Hence, the mesh was made particu-
larly ne immediately around each pile, as shown (with the elements in front removed) in
Figure 8.3(c). A further FE analysis was performed with a ner mesh and it was found

204

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

that the critical outputs (raft settlement and pile load) were not affected signicantly.
Therefore, the adopted mesh was considered acceptable.

Initial stresses (1.4.1)


The initial stresses were specied with vertical effective stress based on a uniform ground
density of 20 kN/m3, horizontal ground surface and hydrostatic pore water pressure with
a groundwater level at 1.2 m (coincident with formation level of raft) and zero pore
pressure above this level. A K0 value of 0.5 was derived by simulation of pressuremeter
tests as described in Section 8.2.3.

Construction stages (1.4.2)


No installation effects were considered because the weight of the raft was approximately
equal to the weight of the ground excavated during its construction while the effects of
pile installation were taken into account in the back-analysis of the pile load tests.
Drained soil conditions were assumed throughout.

1 Establish in situ stresses


As described earlier, with ground level at 0.0 m and groundwater level at 1.2 m
with hydrostatic conditions.
2 Lower ground to formation level (shell raft elements only)
The ground level was lowered from 0.0 m to 1.2 m level and a uniform
surcharge of 24 kPa applied at the same time to represent the weight of the
removed ground. This stage prepared the geometry for those cases where shell
elements were used to represent the raft.
3 Install raft foundation and core walls
In the case of solid elements, the constitutive model for the elements in the raft
volume was changed from the soil model to the concrete model. This resulted in a
moderate increase in weight due to the increased density (from 20 to 24 kN/m3).
In the case of shell elements, the shell elements were added to the top surface of
the model. The surcharge representing the ground was removed from the raft area
and the weight of the raft was included in the material properties of the raft.
Alternatively, the surcharge could have been increased to 28.8 kPa and a zero
density specied for the shell elements.
The core walls were added as shell elements in all the analyses. They had zero
specied weight because the weight of the core walls was included in the line
loads.
The displacements were reset to zero at the beginning of this stage so that outputs
of displacement from this stage onwards were those due solely to the self-weight
of the raft and piles (if any) and the applied loads.
4 Install piles (if used)
In the case of solid elements, the method was the same as for the raft.
In the case of special beam elements, the beam elements were added to the model
along the pile axes with no removal of ground elements. An additional weight of
1.13 kN per metre pile length was included to take account of the additional pile
weight over and above the ground weight occupying the pile volume in the FE
model.

205

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

5 Apply structural loads


The loads shown in Figure 8.1 were applied concurrently in one step. This was
considered a reasonable assumption because the multi-storey structure was to be
constructed in complete oors such that the loading on the foundations would
increase approximately uniformly. Were one side to be constructed before the
other, consideration would have been given to different loadings during the
construction phases.
Since the core walls were represented by shell elements, the loading could only be
applied as a line load (and not an area load), but this was considered a reasonable
assumption since the core walls were only 0.3 m thick in reality. However, the
columns were 1.0 m square and applying the column loads as point loads might
have been unrealistic. In order to assess the validity of such an assumption,
analyses were performed with point loads and area loads and the outputs
compared. Each 1.0 m2 area was represented by one side of four tetrahedral
elements, so the area load was converted into 13 equivalent nodal loads
signicantly more than a single nodal load as described in Section 5.1.8.

Calculation options (1.4.3)


The Modied NewtonRaphson solution scheme was adopted with arc length control.
Automatic step-sizing was utilised and the maximum equilibrium error was set at 1%.
The small deformation (Total Lagrangian) formulation was adopted in all analysis stages.

8.2.3 Obtaining parameters and constitutive model features


Constitutive model selection
The soil at this site was a medium-dense, lightly over-consolidated sand. The important
aspects of its behaviour in this case were hardening behaviour under deviatoric stress and
compression and a non-linear stressstrain response during stress changes from low,
near-surface stresses to high stresses under the loaded foundation. No signicant aniso-
tropy was expected in the soil. Therefore, a double-hardening model with hyperbolic
stressstrain relationship and stress-dependent stiffness was selected.

A linear elastic model was selected for the reinforced concrete structures because stresses
in these elements were expected to be relatively low and within the linear range of stress
strain behaviour. No consideration of concrete cracking was considered necessary and
the properties of the reinforcement were smeared across the section.

Obtaining parameters
In situ rather than laboratory parameter testing was the only option due to the difculty
of obtaining undisturbed samples from sand. The self-boring pressuremeter (as
described in Section 3.3.4) provided the most accurate means of obtaining stiffness
parameters, and it was also used to obtain shear strength parameters.

A sample pressuremeter plot from the sand at 7 m depth is shown in Figure 8.4 from
which shear strength and stiffness parameters were derived. The shear strength was
derived from both the loading and unloading portions of the pressuremeter curve. The
loading portion of the curve was replotted on natural logarithm scales, as shown in

206

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.4 Sample pressuremeter test data

500

400
Loading Unloading

Unloadreload loop
Pressure p: kPa

300

200

100 Test data


FEA

0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14
Cavity strain c

Figure 8.5. The curve reached a constant maximum slope S = 0.415 which was used to

derive values of peak internal friction angle wpeak and dilatancy angle c according to the

Hughes et al. (1977) method and with Rowes stress dilatancy law. wpeak = 388 and c =
58 were obtained from Equations 8.1 and 8.2, respectively.

Figure 8.5 Derivation of shear strength from loading portion

S
ln p: kPa

4 1

0
8 7 6 5 4 3 2
ln c

207

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis


A constant volume value wcv for the sand was required in these equations which was

obtained from direct shear tests in a laboratory (wcv = 348). c was later set to zero in the
FE analysis of the raft foundation due to the risk of over-predicting soil strength in
conned problems such as piled foundations.
S
sin w = (8.1)
1 + (S 1) sin wcv

sin c = S + (S 1) sin wcv (8.2)

The unloading portion of the curve (ignoring the last horizontal leg of the curve) was
plotted as ln p against the natural log of the difference between the current strain and
maximum strain (ln(1max 1c )), as shown in Figure 8.6. The plot approached an

asymptote with slope Sd of 2.60 and a wpeak of 358 was obtained from Equation 8.3
according to Houlsby et al. (1986). This is comparable with the value obtained from the
loading portion. The value can be rened during simulation of the pressuremeter test as
described later.
  
 
1 + sin wcv 1 + sin wcv 2
sin w = sin wcv + sin wcv + 1 (8.3)
Sd Sd

The elastic ground stiffness can be determined from unloadreload loops in the pressure-
meter test and one of these is shown in close-up in Figure 8.7. The average shear
modulus G of this loop is calculated simply by drawing a chord through the apexes of
the loop, the slope of which equals 2G. In this case, G = 56 MPa was obtained which was

Figure 8.6 Derivation of shear strength from unloading portion

6
ln p: kPa

4 Sd

0
2 3 4 5 6 7 8
ln (max c)

208

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.7 Unloadreload loop and derivation of elastic shear modulus G

550

500
Pressure p: kPa
450

400
2G

350
1
300
0.0960 0.0965 0.0970 0.0975 0.0980
Cavity strain c

appropriate for the strain level of the unloadreload loop (approximately 0.04%)
whereas slightly higher strain levels were expected in the raft analysis. The full strain-
dependency of stiffness was determined from the unloadreload loop, as demonstrated
in the following paragraph.

To determine the strain-dependency of stiffness, the reloading portion of the loop was
plotted on natural log axes of p against shear strain g according to Equation 8.4, with
local values (i.e. p and 1c reset to zero at the start of the reload portion) as shown in
Figure 8.8. A straight line was obtained with a slope b of 0.576. The intercept of this line
at ln local dA/A = 0 was h = 8.76, or 10.8 MPa when the log was reversed.

dA 1
g = area ratio =1 2 (8.4)
A 1 + 1c

The secant and tangent pressuremeter shear moduli were calculated from Equations 8.5
and 8.6, respectively, as described by Whittle (1999). The plots of both moduli are shown
in Figure 8.9. Practically, the minimum strain for elastic stiffness from a pressuremeter is
about g = 0.01% depending on the resolution of the instrument and the maximum is
about g = 1% depending on the soil because plastic strains begin to develop. The stiff-
ness at smaller strains could be obtained by seismic testing (see Section 3.3.5). For an
estimated average strain level beneath the raft of 0.2%, a pressuremeter tangent elastic
shear modulus G pt of 44 MPa was obtained. Muir Wood (1990) showed that (in clay) the
tangent modulus (G pt ) obtained in the pressuremeter test is equal to the secant modulus
G s obtained from laboratory testing. For this test in sand, G s was assumed approxi-
mately equivalent to G pt and then the value adjusted as necessary during pressuremeter
test simulations.

G ps = hbg b 1 (8.5)
G pt =hb2g b 1 (8.6)

209

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.8 Derivation of strain-dependent elastic shear modulus G

ln local p: kPa


4
1

Line intercept

3
11 10 9 8 7
ln local c

The same strength and stiffness parameters were derived from the other pressuremeter
tests (not presented here) in order to select appropriate characteristic values for input
into the model. In order to account for the stress-dependency of stiffness, the effective
stress at the start of each unloadreload loop was estimated using Equation 3.3. In the
unloadreload loop analysed in Figure 8.7, pu was 477 kPa and sv0
and sh0 were esti-

mated to be 100 and 50 kPa, respectively, giving sh = 135 kPa. The stress-dependency

of stiffness was determined in terms of the minor principal effective stress sv0 which
in this case was 100 kPa because the horizontal stress had become the major principal
stress.

Figure 8.9 Decay of elastic stiffness with strain

300

200

Secant pressuremeter shear modulus Gsp


G: MPa

100

Tangent pressuremeter shear modulus Gtp

0
0.01 0.1 1
Shear strain : %

210

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.10 Stress-dependency of elastic unloadreload stiffness

150

100
Eur: MPa

50
Model
Test data

0
0 50 100 150 200 250
3: kPa

The elastic unloadreload Youngs modulus Eur was calculated from the standard elastic
relationships provided in Appendix 3.1 (assuming a Poissons ratio of 0.2) from all the
pressuremeter tests and plotted against the corresponding s3 values in Figure 8.10. The
stress-dependency was expressed as an exponential function (Equation 8.7) as used by
Duncan and Chang (1970) and in some other constitutive models.
 n
s3
Eur = Kur pa (8.7)
pa

where pa is atmospheric pressure (100 kPa). A close t with the data was achieved
with Kur = 900 and n = 0.5. From this relationship it was possible to estimate similar
parameters for the primary loading stiffness as described in the next paragraph. These
parameters were then checked and revised in test simulations as described later.

In order to account for the non-linear behaviour of the sand on primary loading, the
Kondner (1963) hyperbolic stressstrain relationship as described by Duncan and
Chang (1970) was adopted as shown in Figure 8.11. The primary loading stiffness
can be expressed as the initial tangent modulus Ei or the secant modulus E50 at 50%
of the deviatoric stress at failure (qf ). In many soils, Ei is about 0.6Eur and E50 is about
one-third of Eur . Therefore, using similar equations to Equation 8.7 for the stress-
dependency of stiffness, as shown in Equations 8.8 and 8.9, Ki = 545 or K50 = 300 were
adopted while n was kept at 0.5 because this tends to remain the same for the different
stiffness moduli.
 n
s3
E i = Ki p a (8.8)
pa
 n
s3
E50 = K50 pa (8.9)
pa

Rf is the ratio between qf and the asymptote to the hyperbolic curve which was estimated
as 0.9 as typically used for most soils.

211

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.11 Hyperbolic stressstrain relationship (Kondner, 1963)

q
Asymptote
qa
Failure
qf

E50
Ei
1
1

0.5qf
Eur

There are uncertainties inherent in all the parameter derivations from the pressuremeter
described in the preceding paragraphs. The pressuremeter test is well suited to simulation
due to its axisymmetry and the full pressuremeter test was simulated (using the method
described in Figure 3.12) in a 2D axisymmetric FE analysis. The w value was adjusted to
368 in order to improve the match between the test simulation outputs and the test data
as shown in Figure 8.4. c was kept at 58 during this test simulation but taken as zero in
the FE analysis of the raft foundation and piles due to the likelihood of over-predicting
soil strength when c . 08 in conned problems such as piled foundations. The initial
state parameters, which were difcult to determine directly from the test data, were also
estimated and adjusted until a good match with the test data was achieved. It was found
that a K0 value of 0.5 was appropriate and the pre-consolidation stress (dening the
location of the cap yield surface for compression hardening) was assumed equal to the
in situ stress, i.e. normally consolidated conditions.


The peak shear strength wpeak was adopted in order to obtain more realistic predictions of
deformation. However, if ultimate limit states needed to be predicted, it would be
appropriate to adopt a lower, post-peak shear strength. Furthermore, vertical pressure-
meters provide strength and stiffness parameters in the horizontal direction whereas
stress changes in the raft problem were orientated predominantly in the vertical
direction. Nevertheless, the degree of anisotropy in the sand was judged to be low,
so the pressuremeter-derived parameters were considered appropriate for this raft
problem.

A Youngs modulus E of 30 GPa was adopted for the reinforced concrete with a
Poissons ratio of 0.15. The weight density of the concrete was taken as 24 kN/m3.

212

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.12 Back-analysis of pile load test

350

300
Applied axial load: kN

250

200
Modified beam
150
Solid elements
100 Test data

50

0
0 1 2 3 4 5
Pile head settlement: mm

With the soil and concrete parameters set, the only simple way to parameterise ground
pile interaction behaviour was in the properties of the interface between them. Interface
elements with a Coulomb friction criterion were installed between the pile and soil solid
elements. The friction properties of the interface were determined by back-analysis of
load tests on similar pile types and sizes to those proposed to be included in the raft
analysis model. The output from one such back-analysis on a 0.6 m diameter, 10 m long
pile is shown with the test data in Figure 8.12. A reasonable match with the test data was
achieved with a shear strength at the interface of 80% of the internal shear strength of
the sand. In the case of the special beam elements for the piles, the variation of shaft
friction along the pile and the end bearing were specied separately. This was slightly
more complicated than with the solid elements but a reasonable match with the test data
was also achieved as shown in Figure 8.12.

8.2.4 Outputs
The outputs of vertical displacement in the raft at the end of the nal construction
stage are presented in Figure 8.13 for the cases with and without piles and with either
continuum or non-continuum structural elements (a positive value denotes settlement).
Comparing the outputs without piles, it is apparent that when shell elements were used
for the raft, about 1015% higher settlement was predicted and the maximum deected
slope was higher. This was possibly due to the additional support provided to the solid
elements by the interaction with the ground at the sides and by the moment-reducing
effect of interface friction (see Section 5.1.2). As expected, the inclusion of piles resulted
in smaller settlements with a signicant reduction in the deected slope (or differential
settlement) which is the main function of settlement-reducing piles. The piles exhibited
a stiffer response as solid elements than they did as special beam elements, in spite of the
parameters for both being obtained from the same pile load test.

The outputs of bearing pressure (vertical total stress at formation level) and pile load in
the last construction stage are shown in Figure 8.14 and Table 8.1, respectively (pile

213

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.13 Output of vertical displacement in raft in final construction stage: (a) raft only, solid
elements; (b) raft with piles, solid elements

16

19

22

25

28

31

34

37

40 mm

(a)
16

19

22

25

28

31

34

37

40 mm
(b)

numbers are shown in Figure 8.2). While the bearing pressure plots without piles are
similar, except for more stress concentrations at the edges of the raft composed of shell
elements, the plots with piles show clearly that a higher bearing pressure was predicted
with the shell elements compared with the solid elements, with a greater share of the load
being taken by the piles when using solid elements, as was also evident in the pile loads
(12% of the total foundation load taken by the piles with solid elements compared with
8% with shell and special beam elements). In this respect, the example illustrates well
the effect of decisions taken at the analysis planning stage on the outputs obtained at
the end.

Obtaining outputs of bending moment, axial force and other structural forces from the
solid elements is less straightforward than from the shell elements (see Table 5.2). The
axial forces at the tops of the piles shown in Table 8.1 were obtained by integrating
the vertical normal stress over the cross-sectional area of each pile. In this case, the area

214

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.13 (c) raft only, shell elements; (d) raft with piles, shell and special beam elements

16

19

22

25

28

31

34

37

40 mm
(c)

16

19

22

25

28

31

34

37

40 mm

(d)

to be integrated was well dened, but in the case of the raft, rather than integrate over the
entire cross-sectional area (which would be of limited use in design), sections needed to
be pre-dened in the mesh. As shown in Figure 8.3(b), a 1.0 m wide prism was pre-
dened in the mesh running in the X-axis direction, such that bending moment per metre
width could be obtained for the raft for any section along that prism. The bending
moment was obtained by integrating the normal stresses at integration points multiplied
by corresponding lever arms about the neutral axis for sections along the prism. Some
programs perform these calculations automatically which signicantly simplies the
process. The resulting bending moment is shown in Figure 8.15 along with the direct out-
put from the equivalent shell element analysis along the same section line (a positive
value denotes sagging moment). The plots were very similar, except the shell element
output was more spiky at the column locations (X = 61, 71 and 91 m) while the solid
element output was more spiky at the connections with the core walls (X = 76 to 86 m)

215

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.14 Output of bearing pressure at formation level beneath raft: (a) raft only, solid
elements; (b) raft with piles, solid elements

40

65

90

(a)
115

140

165

190 kPa

(b)

where the shell elements were extended into the solid elements. The addition of settlement-
reducing piles appeared to cause a greater reduction in bending moment in the core wall
area in the solid element case compared with the shell element case.

The bending moment outputs presented in Figure 8.15 were from analyses where the
column loads were distributed over the 1 m2 area of the columns. Analyses were also per-
formed with point loads to represent the column loads and the bending moment outputs
from each case are compared in Figure 8.16. Clearly, the output in between the column
locations was unaffected by the loading assumption but was signicantly affected
immediately around each column location. With solid raft elements, the local maxima
under point loads were about 50% higher at the small column loads (X = 61 and
91 m) and about 10% higher at the large column load (X = 71 m). With shell raft
elements the effect was even greater at about 80% and 30% higher at the small and large
loads, respectively. Therefore, in this example, the use of point loads resulted in an

216

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.14 (c) raft only, shell elements; (d) raft with piles, shell and special beam elements

40

65

90

(c)
115

140

165

190 kPa

(d)

Table 8.1 Output of axial load taken by settlement-reducing piles

Axial load at top of pile: kN

Pile no. Solid elements Special beam elements

1 686 430
2 1117 635
3 1132 717
4 745 457
5 681 482
6 686 482

Total 5047 (12%) 3203 (8%)

217

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.15 Outputs of bending moment in raft along section on X-axis; (a) raft only; (b) raft with
piles

2000

1500
Bending moment: kNm/m

1000
Shell elements
Solid elements
500

0
60 70 80 90
X: m
500

(a)

2000

1500
Bending moment: kNm/m

1000
Shell elements
Solid elements
500

0
60 70 80 90
X: m
500
(b)

unrealistic over-prediction of local bending moment at the column locations, particu-


larly when using shell elements for the raft.

A key advantage of using non-continuum elements for structures is the direct output
of structural forces. The distribution of bending moment across the entire raft could be
plotted straightforwardly, as shown in Figure 8.17 for distributed column loads. Mx
denotes bending orientated along the X-axis for rotation about the Y-axis while My
denotes bending orientated along the Y-axis for rotation about the X-axis and a
positive value denotes sagging moment. It was apparent that the inclusion of piles
reduced the bending moment generally in the centre of the raft where they were
located.

218

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.16 Outputs of bending moment in raft along section on X-axis with point and distributed
loads: (a) raft only, shell elements; (b) raft only, solid elements

2000

1500
Bending moment: kNm/m

1000
Point loads
Area loads
500

0
60 70 80 90
X: m
500

(a)

2000

1500
Bending moment: kNm/m

1000
Point loads
Area loads
500

0
60 70 80 90
X: m
500
(b)

The outputs of vertical displacement and bearing pressure shown in Figures 8.13 and
8.14 were processed to derive coefcients of subgrade reaction k which could be used
in other structural analysis models where groundstructure interaction effects were
simulated approximately by spring elements. k is simply the bearing pressure divided
by the settlement, but an element of approximation is introduced into the calculation
because displacement output is provided at the nodes while stress output is provided
at integration points and they are not coincident. Consequently, direct output is not
normally available in an FE analysis program and the user needs to process the data.
In this example, the raft was divided into a grid 1 m wide at the edge and 3 m square
internally as shown in Figure 8.18. By inspecting the output data, approximate average
values of settlement and bearing pressure for each square were obtained. From these, k

219

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.17 Bending moment output from shell elements: (a) Mx , raft only; (b) Mx , raft with piles

2000

1600

1200

(a) 800

400

400 kNm/m

(b)

was determined as shown in Figure 8.18. The values obtained were quite consistent
across each raft, increasing generally toward the edges where settlement was lower.
Higher values were obtained when solid elements were used for the raft because lower
settlements were predicted in these cases. In the cases with piles, spring stiffness values
for each pile are also shown which were obtained simply by dividing the pile axial force
(Table 8.1) by the vertical displacement (Figure 8.13). Springs representing the settle-
ment-reducing piles would need to be added to any beam-spring type model in addition
to the springs representing the ground mass.

8.2.5 Validation
In order to validate the critical outputs, they were compared with the results of alter-
native analysis methods. When comparing different analysis methods, many of the
assumptions inherent in each are different so an exact match between outputs should not
be expected. The goal was to achieve outputs that were reasonably close in order to have
greater condence that the FE analysis model represented reality sufciently accurately.

220

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.17 (c) My, raft only; (d) My, raft with piles

2000

1600

1200

(c) 800

400

400 kNm/m

(d)

First, the one-dimensional settlement under the centre of the raft was calculated assum-
ing a uniformly distributed load over the raft area. The ground beneath the raft was
divided into a number of layers with a one-dimensional stiffness E 0 assigned to each
layer and the Boussinesq solution was used to calculate stress changes at the centre of
each layer. The compression of each layer was summed in order to obtain a settlement
value. The total applied load (including self-weight of the raft) distributed uniformly
over the raft area gave a surcharge q = 60.5 kPa. To take account of the increasing stress
and decreasing strain with depth, the E 0 value was assumed to increase from 20 MPa
near the surface to 140 MPa at 64 m (2B) depth which was assumed to be the depth
of inuence of the foundation. From these values a maximum settlement of 44 mm was
obtained. This compared very favourably with the values of 35 mm and 39 mm obtained
in the FE analysis with solid and shell elements, respectively, representing the raft. Since
the Boussinesq type solution would be expected to be conservative, obtaining somewhat
lower values in the FE analysis was considered reasonable.

221

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.18 Coefficients of subgrade reaction and pile spring stiffness values derived from outputs:
(a) raft only, solid elements; (b) raft only, shell elements

3.3 3.1 2.9 3.1 3.2 2.9 2.9 3.1 2.9 2.9 3.1 2.9

2.9 3.1 2.9 2.7 2.8 2.8 2.7 2.8 2.8 2.8 3.0 3.2

2.8 2.9 2.7 2.6 2.6 2.6 2.5 2.5 2.6 2.7 2.8 3.0

3.0 2.8 2.6 2.6 2.5 2.5 2.5 2.5 2.5 2.6 2.7 3.0

2.8 2.8 2.6 2.5 2.5 2.4 2.4 2.4 2.5 2.5 2.6 2.9

2.9 2.8 2.6 2.4 2.4 2.4 2.4 2.4 2.4 2.4 2.6 2.9

(a)

4.1 3.7 3.4 3.3 3.1 3.2 3.2 3.1 3.2 3.5 3.7 4.3

3.6 2.8 2.6 2.4 2.4 2.4 2.4 2.4 2.5 2.6 2.9 3.8

3.4 2.6 2.3 2.3 2.3 2.2 2.2 2.2 2.3 2.3 2.6 3.4

3.4 2.5 2.2 2.2 2.2 2.2 2.1 2.2 2.2 2.3 2.5 3.3

3.3 2.5 2.2 2.2 2.1 2.2 2.2 2.2 2.2 2.3 2.5 3.0

3.3 2.5 2.3 2.1 2.2 2.2 2.2 2.1 2.2 2.2 2.4 2.9

(b) Y
3
Coefficients of subgrade reaction (MN/m )

In order to take account of the raft stiffness, a second elastic analysis solution was
studied. Results from a uniformly loaded square raft in frictionless contact with a homo-
geneous isotropic half-space were used, as derived by Fraser and Wardle (1976) and
described in Hemsley (1998).

For the case with the settlement-reducing piles, there was no readily available, simple,
alternative analysis method. While piled raft analysis methods exist, they tend to assume
piles are distributed across the whole area of the raft, rather than just in the central area
as in this case. Therefore, the same elastic raft analysis method was used but with a crude

222

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.18 (c) raft with piles, solid elements; (d) raft with piles, shell and special beam elements

3.3 3.2 3.0 3.1 3.1 3.0 3.1 3.1 3.0 2.9 3.2 3.2

3.2 3.0 2.8 2.7 2.7 2.5 2.6 2.6 2.7 2.8 3.1 3.2

3.0 2.9 2.7 2.8 2.5 2.4 2.4 2.5 2.6 2.6 2.8 3.0

2.9 2.8 2.6 2.4 2.4 2.4 2.3 2.4 2.5 2.5 2.6 3.0
26 26

2.8 2.6 2.5 2.5 2.4 2.4 2.3 2.4 2.4 2.4 2.5 2.8

27 42 42 27
2.8 2.6 2.4 2.4 2.3 2.3 2.2 2.3 2.3 2.3 2.5 2.7

(c)

3.9 3.5 3.2 3.2 3.0 3.0 3.0 3.0 3.1 3.2 3.4 3.8

3.4 2.8 2.6 2.5 2.3 2.4 2.4 2.4 2.4 2.5 2.9 3.5

3.2 2.6 2.4 2.2 2.2 2.1 2.1 2.2 2.2 2.2 2.6 3.2

3.0 2.5 2.2 2.2 2.1 2.0 2.0 2.1 2.1 2.2 2.6 3.2
15 15

2.9 2.4 2.2 2.1 2.11 2.2 2.1 2.0 2.2 2.3 2.5 2.9

14 19 21 14
2.9 2.4 2.1 2.0 2.0 2.0 2.0 2.0 2.1 2.3 2.5 2.8

(d) Y
Coefficients of subgrade reaction (MN/m3)

Pile spring stiffness (kN/mm)


X

average stiffness of the soil and piles taken for the half-space (the 12 piles occupied
0.08% of the ground volume under the raft to a depth of 64 m (2B)).

The raft dimensions were B = L = 32 m, the thickness t1 = 1.2 m, Youngs modulus


E1 = 30 000 MPa and Poissons ratio n1 = 0.15. The uniform surcharge was q =
60.5 kPa.

For the half-space, a single Youngs modulus needed to be selected taking into account
the stress state and strain level in the soil. This value was judged to be E2 = 30 MPa.

223

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Table 8.2 Comparison between elastic analysis and FE analysis outputs for raft (without piles)

DwAB : mm DwAC : mm M : kNm/m

Elastic analysis 12 23 620

FE analysis (solid structural elements) 9 17 500 (away from columns)


FE analysis (non-continuum structural 10 18 500 (away from columns)
elements)

In the case with piles, 0.08% of the soil was assumed to be composed of concrete
which raised the E2 value to 54 MPa. The Poissons ratio in both cases was taken as
n2 = 0.2.

These values gave relative raft stiffness values of Kr = 0.052 in the raft only case and
0.029 for the raft with piles. The elastic analysis solutions provided outputs of the
differential vertical displacement between the centre and mid-side (DwAB ) and between
the centre and corner (DwAC ) as well as the maximum bending moment M . The results
from this elastic analysis are compared with the FE analysis outputs in Tables 8.2
and 8.3.

As before, the FE analysis outputs compared very favourably with the elastic analysis
results. The values from the FE analyses were slightly lower, which was expected. The
bending moment outputs shown were the maximum values away from the column loads
because the elastic analysis assumed a uniformly distributed load so the values at the
column loads would not be expected to be close to the values obtained from the elastic
analysis.

The comparisons between the outputs of the FE analyses and the two elastic analyses
showed a reasonably close match and gave more condence in the accuracy of the FE
analysis model. Further techniques of validation should be undertaken, as described
in Section 7.2.2, in order to improve condence in the outputs further.

Table 8.3 Comparison between elastic analysis and FE analysis outputs for raft with settlement-
reducing piles

DwAB : mm DwAC : mm M : kNm/m

Elastic analysis 8 15 415

FE analysis (solid structural elements) 5 8 400 (away from columns)


FE analysis (non-continuum structural 6 11 300 (away from columns)
elements)

224

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

8.3. Shaft excavation example


8.3.1 Summary
This example concerns the excavation of a shaft in rm over-consolidated clay supported
by a sheet pile wall. The sheet pile wall will be supported by hoop forces in a capping
beam and by passive resistance from the ground inside the shaft. The groundwater level
in the shaft will be permanently lowered. A buried sewer that is sensitive to deection
exists near to the shaft, requiring an accurate prediction of excavation-induced ground
movement.

The particular features of this example include:

g 2D axisymmetric model
g K0 initial stress procedure
g double-hardening, stress- and strain-dependent stiffness constitutive model
g obtaining drained geotechnical parameters from advanced, undrained laboratory
triaxial tests
g groundwater ow analysis and hydraulic failure
g undrained A conditions followed by consolidation and drained conditions
g structural non-continuum elements with anisotropic stiffness properties
g dual factoring ULS check with stepwise strength reduction
g validation using monitoring data from a similar shaft excavation.

8.3.2 Setting up the FE analysis model


Justification for using FE analysis (1.1.1)
The alternative analysis methods, such as beam-spring models, do not consider 2D axi-
symmetric conditions (e.g. the contribution of hoop forces and orthotropic structures) or
else consider them in an approximate way.

In particular, the requirement to predict excavation-induced ground movement around


the shaft was justication for using FE analysis. While empirical methods of predicting
ground surface settlement behind retaining walls exist, these tend to be for straight walls
rather than shafts.

Furthermore, by using an advanced constitutive model, there was scope for obtaining a
more economical design.

Aims of the model (1.1.2)


1 to predict the ground movement at the location of a buried sewer near the
proposed shaft
2 to perform a ULS design to select the appropriate sheet pile section and
embedment depth
3 to predict groundwater head due to permanent groundwater lowering and check
for hydraulic failure.

Geometrical simplifications (1.2.1 and 1.2.2)


This site had no signicant structural features except for the proposed shaft and the
sewer nearby, as shown in Figure 8.19. Including the sewer in the model would require

225

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

a full 3D geometry. However, by ignoring the contribution of the sewer to ground


stiffness (which was conservative), it was possible to obtain approximate ground
deections along the line of the sewer using a 2D axisymmetric model, as described in
Section 8.3.4.

Furthermore, the ground conditions were quite uniform and with no inclined strata
which also allowed a 2D axisymmetric geometrical assumption to be adopted. Although
a concrete base to the shaft and some infrastructure would be constructed later, no input
to the design of these was required from this FE analysis. Therefore, for conservatism in
the shaft wall design, these structures and their associated loads were omitted from the
model.

Model boundaries and fixities (1.2.3 and 1.2.4)


Initially based on rules of thumb (Figure 1.7), the remote vertical boundary to the model
was placed 45 m (3B) from the sheet pile wall and the bottom boundary 30 m (2B) below
the excavation level. It was found that the remote vertical boundary could be moved
inwards to about 43 m (2.8B) from the sheet pile wall and the bottom boundary to
25 m (1.7B) below the excavation level, as shown in Figure 8.20, without introducing any
signicant boundary effects on the key outputs (except on the plane of symmetry). These
locations were governed by the groundwater ow analysis rather than the displacement
analysis.

The bottom boundary was xed in both axis directions and the vertical boundaries
(including the axis of symmetry) were xed only in the horizontal direction. The hydrau-
lic boundary conditions are described under Construction stages.

Figure 8.19 Shaft excavation example geometry and groundwater levels

1 m square rc
capping beam

Water levels: ground level 0 m 3m


Sewer
Worst case 0 m
Characteristic 1 m
1 m
Excavation level 5 m

Lowered to 6 m
inside shaft
Wall toe level 10 m
Sheet pile wall

Soil: firm over-consolidated clay


15 m diameter

226

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.20 FE mesh for shaft excavation example

Axis of
symmetry 10 m
0m 7.5 m

5 m

10 m

Side fixed in X direction


Side fixed in X direction

30 m Base fixed in X and Y directions


50 m

Finite element mesh (1.3)


Cubic strain 15-node triangular elements with three degrees of freedom per node were
used to model the ground. The use of such higher order elements is particularly
important for the prediction of failure states in axisymmetric models as described in
Section 1.3.1.

The sheet pile wall and capping beam were modelled with 5-node line elements forming
2D shell elements with three degrees of freedom (two translational and one rotational)
per node. The elements were located at the centreline radius of the shaft wall. The
benets of using continuum elements (see Section 5.1.2) were considered minor for a thin
sheet pile wall compared with the benets of their easier use.

Interface elements were used between the shell elements and the area elements for the
ground to take account of the reduced friction between the wall and ground as well as
to create a closed hydraulic boundary condition for the groundwater ow analysis.

The mesh was generated as shown in Figure 8.20 with smaller elements created near the
wall where steep stress and strain gradients were expected.

227

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

A further FE analysis was performed with a ner mesh and it was found that the critical
outputs were not affected signicantly. Therefore, the adopted mesh was considered
acceptable.

Initial stresses (1.4.1)


The initial stresses were specied with vertical effective stress based on a uniform ground
density of 20 kN/m3 and a horizontal ground surface. Pore water pressures were assumed
hydrostatic below a groundwater level at 1 m depth and zero above this level. A K0 value
of 1.0 (obtained as described in Section 8.3.3) was used to calculate horizontal effective stress.

Construction stages (1.4.2)


The installation effects on the clay soil associated with the proposed sheet pile wall were
expected to be minor so were not considered in the FE analysis.

Undrained A conditions (see Section 4.2.4) were simulated during construction followed
by dissipation of excess pore pressures in a consolidation analysis to simulate long-term
conditions. Drained conditions were assumed in the subsequent ULS stages since the
drained long-term condition was known to be the most onerous case. The importance
of following the correct stress path (i.e. undrained construction then consolidation to
long-term drained conditions) rather than analysing only drained conditions was high-
lighted in Section 4.2.2.

The sequencing of the construction stages is illustrated in Figure 8.21, while each stage is
described in detail as follows:

1 Establish in situ stresses


As described earlier, with ground level at 0.0 m and groundwater level at 1.0 m
with hydrostatic conditions.
2 Install sheet pile wall and capping beam (undrained)
The shell elements of the sheet piled wall and capping beam as well as the
interface elements were activated.
3 Excavate shaft (undrained)
The ground elements and attached interface elements within the shaft to 5.0 m
depth were deactivated. The steady-state pore water pressure remained unchanged
but no external water pressures were applied within the shaft (in effect, the
groundwater within the shaft was removed with the soil).
4 Apply surcharge (undrained)
A uniformly distributed load of 10 kPa was applied at the ground surface on a
10 m wide strip immediately behind the wall to represent construction trafc.
5 Long-term, drained conditions (groundwater ow then consolidation analysis)
It was proposed to install permanent dewatering measures in the shaft to lower
the groundwater level to 1.0 m below the base of the shaft. Therefore, a
groundwater ow analysis was performed in order to establish the new steady-
state pore pressures with the hydraulic boundary conditions shown in Figure 8.22
(with the characteristic water level of 1 m on the right-hand boundary).
Then a consolidation analysis was performed until the excess pore pressures

228

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.21 Sequence of analysis stages

1. Establish in situ stresses

2. Install sheet pile wall

3. Excavate shaft

4. Apply surcharge

5. Long-term conditions

ULS output ULS input


factoring factoring

6. Factor surcharge and 7. Higher groundwater level,


over-dig factor surcharge and over-dig

Factor structural force outputs 8. Ground strength reduction by


factor required by code

9. Continue ground strength


reduction to failure

resulting from the excavation, surcharge and new steady-state pore pressure
dissipated to less than 1 kPa.
The outputs from this stage represented the long-term, drained conditions
following undrained construction conditions and were expected to be more
onerous in terms of ground movement, wall stability and wall structural forces
since this was primarily an unloading problem (refer to Section 4.2.2).
6 ULS output factoring (drained)
Continuing from Stage 5, the surcharge was increased by the ratio of partial
factors variable to permanent as given in the design code (1.5/1.35 1.1 in this
example) to 11.1 kPa. The excavation depth was increased to 5.5 m by
deactivating an additional layer of elements and the attached interface elements to
allow for over-dig as required by the design code. The steady-state pore pressure
remained the same. Outputs of structural force in the sheet pile wall elements were
factored for structural ULS checks as described in Section 6.3.1.
7 ULS input factoring (groundwater ow analysis then drained analysis)
Continuing from Stage 5, a groundwater ow analysis was performed with the

229

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

groundwater level outside the shaft raised by 1 m to the ground surface which was
judged to be the worst case water level. The groundwater level within the shaft
remained at 6.0 m level because this was controlled by the dewatering measures.
Since the clay had low permeability, it was not considered necessary to consider
the case of a temporary failure of the dewatering measures. This generated the
ULS steady-state pore pressure.
The surcharge was increased by the partial factor required by the design code
(1.3 in this case) to 13 kPa. The excavation depth was increased to 5.5 m by
deactivating an additional layer of elements and the attached interface elements to
allow for over-dig as required by the design code. This stage prepared the analysis
for strength reduction in the subsequent stages.
8 ULS input factoring strength reduction by required partial factor
Continuing from Stage 7, the shear strength of the ground was reduced in a
stepwise fashion until the partial factor on drained strength was reached. For this
process, the constitutive model was simplied to a LEPP model with Mohr
Coulomb failure criterion in order to use the stepwise strength reduction feature
of the FE analysis program. The outputs were then checked for any geotechnical
failure while the outputs of structural force in the sheet pile wall from this stage
were compared with those from Stage 6 and the more onerous of the two was
used in the structural ULS check, as described in Section 6.3.1.
9 ULS input factoring continued strength reduction
Strength reduction was continued in the same way until equilibrium could not be
established and the most critical ground failure mechanism was identied. The
sheet pile wall constitutive model remained linear elastic without a strength
criterion, but it would have been an option to adopt an elastic-plastic model to
study the effect of reduced bending resistance in the sheet piles.

Figure 8.22 Hydraulic boundary conditions

Open boundary 0m
1 m
Closed boundary

6 m Specified head on right-


hand boundary, worst case
and characteristic levels
Open boundary
(axis of symmetry)
Closed boundary

Specified head on left-hand boundary

Closed boundary

230

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Calculation options (1.4.3)


The Modied NewtonRaphson solution scheme was adopted with arc length control.
Automatic step-sizing was utilised for both displacement and consolidation analysis and
the maximum equilibrium error was set at 1%. The small deformation (Total Lagran-
gian) formulation was adopted in all analysis stages.

8.3.3 Obtaining parameters and constitutive model features


Constitutive model selection
The clay was over-consolidated so was likely to exhibit initially elastic behaviour in
deviatoric loading followed by yielding behaviour when its stress state reached the yield
surface. Large stress changes would be caused by the excavation of the shaft, so a stress-
dependent soil stiffness was important in order not to over-estimate the heave of the base
of the excavation. While the critical case for deformations and ULS design was likely to
be the long-term drained case in this unloading problem, construction needed to be
simulated in undrained conditions followed by dissipation of excess pore pressures in
order to follow a more accurate stress path (as explained in Section 4.2.2). This required
a non-linear stiffness constitutive model capable of reasonably accurate excess pore
pressure prediction using the Undrained A approach. Furthermore, accurate prediction
of excavation-induced ground movements around the shaft required a model with strain-
dependent stiffness. No signicant anisotropy was expected in the clay, so an isotropic,
double-hardening model with hyperbolic stressstrain relationship and strain-dependent
stiffness was selected.

Stresses within the steel sheet piles and reinforced concrete capping beam were expected
to be relatively low and within the linear elastic range of these materials. Therefore,
linear elastic models were adopted for these materials. However, orthotropic elasticity
was introduced in the steel sheet pile model due to the geometrical anisotropy of the
proled sheets (see Section 5.1.6).

Obtaining parameters
High-quality samples of the clay from various locations within the zone of inuence of
the proposed shaft were tested in stress path triaxial cells with local strain measurement
and bender elements at the top and bottom of each specimen. This allowed the full
strain-dependency of stiffness to be derived. Each specimen was saturated and then
reconsolidated to their in situ stress state along their most recent stress path, as described
under Reconsolidation stage in Section 3.3.1. Then each specimen was unloaded axially
in extension until shear failure which matched approximately the stress path in this
unloading problem.

The difculty with obtaining parameters for the constitutive model was that the model
parameters were dened in terms of drained triaxial behaviour whereas performing
drained triaxial tests on low-permeability clays would have been very time-consuming
and expensive. Therefore, undrained shear stages were employed and the model param-
eters were obtained by back-analysis of the undrained tests. K0-consolidation tests
(equivalent to oedometer tests) were also undertaken using a stress path triaxial cell
on specimens of reduced height in order to reduce consolidation times.

231

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.23 Simulation of triaxial extension test

100

80

60
q: kPa

Lab test data


Single-point algorithm
40

20

0
0 1 2 3 4 5 6
a: %

a : %
0 1 2 3 4 5 6
0

Lab test data


10 Single-point algorithm
ue : kPa

15

20

25

Some FE analysis programs have single-point algorithms for simulating simple laboratory
tests and these were used to back-analyse the triaxial extension and K0-consolidation tests.
The outputs from these are compared with the test results for one set of specimens in
Figures 8.23 and 8.24. The specimens were obtained from where the in situ vertical effec-
tive stress was 100 kPa, with an assumed K0 value of 1.0 and vertical pre-consolidation
stress of 1100 kPa. The single-point algorithm outputs were obtained using the model

232

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.24 Simulation of K0-consolidation test

v: kPa
0 50 100 150 200
0

Lab test data


0.1 Single-point algorithm

0.2
v : %

0.3

0.4

0.5

parameters shown in Table 8.4 and as dened in Section 8.2.3 for the raft foundation
example. The Koed parameter refers to the soil stiffness in one-dimensional (K0 ) com-
pression. A lower, post-peak shear strength was adopted since this FE model would
be used for ULS verications. For more accurate pre-failure deformation predictions,
a peak strength could have been used, but the strength softening post-peak would not
have been recreated with the adopted constitutive model. The stress-dependency of stiff-
ness was determined by comparing the stiffness values obtained from different tests at
different reconsolidation stresses.

The in situ stress ratio K0 is an important initial state parameter in retaining wall design
and a value of 1.0 was estimated from suction measurements undertaken on high-quality
samples as soon as they had been brought to the surface. The vertical pre-consolidation
stress was estimated to be 1000 kPa higher than the existing in situ stress, which was
found to give reasonably accurate predictions in all the test simulations.

From the same triaxial extension test studied earlier in this example, a plot of secant
shear modulus G s against shear strain is shown in Figure 8.25. The precision of the local
strain instrumentation allowed shear strain to be measured from a value of about 0.01%.
A very small strain (below about 0.002%) G0 of 150 MPa was obtained from the bender
element test on the reconsolidated specimen with an average normal effective stress p of
100 kPa.

Table 8.4 Model parameters for the clay soil

Kur n n Ki or K50 Koed Rf w C c gsat gunsat

700 0.9 0.2 420 or 230 180 0.9 238 5 kPa 08 20 kN/m3 18 kN/m3

233

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.25 Measured non-linear stiffness and Benz et al. model

160
G0 = 150 MPa
140

120
Lab test data
100 Benz et al. model
Gs: MPa

80

60

40

20
0.7 = 0.012%
0
0.0001 0.001 0.01 0.1 1 10
: %

Two strain-dependent stiffness models are commonly used, namely those of Benz et al.
(2009) and Jardine et al. (1986), and the derivation of parameters for these is shown
in Figures 8.25 and 8.26, respectively. For the Benz et al. model, g0.7 = 0.012% was
obtained with G0 = 150 MPa. The stress-dependency was taken into account in the same
way as for the large strain stiffness values using the n value of 0.9. For the Jardine et al.
model, the parameters shown in Table 8.5 were obtained.

Figure 8.26 Measured non-linear stiffness and Jardine et al. model

1600

1400

1200 Lab test data


Jardine et al. model
1000
Gs/p

800

600

400

200

0
0.0001 0.001 0.01 0.1 1 10
: %

234

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Table 8.5 Jardine et al. model parameters

A B C a g

1100 1200 8.0 10 5% 1.2 0.6

Table 8.6 Sheet pile section properties

d t I A M
t
d

280 mm 7.5 mm 10 830 cm4/m 103.3 cm2/m 81.1 kg/m

The PU 8R section was proposed in the design for the sheet piles, and its section prop-
erties as provided by the manufacturer are shown in Table 8.6. Shell model parameters
are specied for an equivalent shell of uniform thickness but when representing a
proled structure, such as a sheet pile wall, it is not possible to dene parameters that
represent both the bending and axial stiffness accurately. Consequently, priority was
given to the bending stiffness of the sheet piles in the vertical orientation. Furthermore,
the bending and equivalent axial stiffness in the out-of-plane (hoop) direction were
orders of magnitude lower than in the vertical direction due to the proled section which
needed to be taken into account in order not to over-predict hoop stresses (the effect of
this over-prediction is demonstrated in the output to this example). However, in order to
avoid an ill-conditioned stiffness matrix, the difference between the bending stiffness in
the vertical and horizontal directions in the model parameters was restricted to 20 times.
The Poissons ratio was set to zero in order to further reduce the generation of hoop
stresses. The model parameters are shown in Table 8.7. A1 refers to the equivalent
sectional area in the vertical direction calculated to satisfy Equation 5.4. A2 refers to the
equivalent sectional area in the horizontal (hoop) direction calculated to satisfy
Equation 5.4 with I2 = 0.05I1 .

The reinforced concrete capping beam was represented by shell elements with isotropic
linear elastic material properties because its 1 m square section was uniform in the
out-of-plane direction such that it could sustain hoop stresses effectively. A Youngs
modulus E of 30 GPa was adopted with a Poissons ratio of 0.15. The weight density
of the concrete was taken as 24 kN/m3, giving a net weight of 4 kN/m3 over the soil that
occupied the actual area of the capping beam in the model.

Table 8.7 Sheet pile model parameters for shell elements

E I1 d A1 A2 w n

207 GPa 1.08 10 4 m4/m 0.28 m 0.0165 m2/m 8.26 10 4 m2/m 0.8 kN/m/m 0

235

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Interface elements with a Coulomb friction criterion were installed between the shell
elements and the area elements of the soil. The shear strength at the interface was taken
as 70% of the internal shear strength of the clay but the outputs were not particularly
sensitive to variations in this value.

8.3.4 Output
The output of sheet pile wall horizontal deection in the long-term case following dissi-
pation of excess pore pressures in construction Stage 5 is shown in Figure 8.27. Only
minor deection occurred at the top of the wall due to the support of the capping beam
and the maximum value of 7.6 mm occurred approximately at the excavation level. The
output from an analysis with isotopic properties for the sheet pile wall is also shown.
Clearly, such an assumption signicantly under-estimates deection because large,
unrealistic hoop forces were generated in the shell elements of the wall.

The predicted variation of horizontal (radial) and vertical deection of the ground at 1 m
depth with distance from the sheet pile wall is shown in Figure 8.28. The shapes of the
proles correspond with those of the wall deection, with less deection adjacent to the
wall where the capping beam provided support, increasing to a maximum at about 2 m
from the wall due to the deection of the wall at lower elevations. The vertical deection
reduced to an insignicant level of less than 0.5 mm at a distance of 10 m (equivalent to
two times the excavation depth) from the wall which agreed with typical rules of thumb
for the zone of inuence of a retaining wall.

By calculating the radial distance from the wall corresponding with the linear distance a
along the buried sewer at 1 m depth as illustrated in Figure 8.29, it was possible to derive
excavation-induced deection plots for the sewer as also shown in Figure 8.29. This
involved some approximation seeing that the stiffness of the sewer was ignored, and also

Figure 8.27 Output of sheet pile wall horizontal deflection

Wall deflection: mm
2 0 2 4 6 8
0

2
Elevation: m

Orthotropic wall
8
Isotropic wall

10

236

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.28 Output of ground movement at 1 m depth behind sheet pile wall

Radial distance from wall: m


0 10 20 30 40
0.0

0.5

1.0
Deflection: mm

1.5

2.0

2.5
Vertical deflection
Horizontal (radial) deflection
3.0

3.5

because the horizontal displacement ux direction varied from being perpendicular to the
sewer at a = 0 to a more longitudinal direction as a increased or decreased away from
zero. Nevertheless, these plots could be used to estimate approximate distortions in the
sewer, and to compare them with allowable values.

The ULS values of sheet pile wall bending moment are shown in Figure 8.30. The output
factoring (OF) values were obtained by multiplying the output from Stage 6 by a load
effect factor of 1.35 (required by the design code) and the input factoring (IF) values
were obtained directly from the output of Stage 8. The maximum absolute value of
66 kNm/m occurred at the connection with the capping beam at 1 m elevation and
happened to be about the same in both the OF and IF cases. This maximum value should
be used to check whether the sheet pile wall has adequate bending moment resistance.

Figure 8.29 Derived output of ground movement along line of sewer

Distance along sewer a: m


30 20 10 0 10 20 30
0.0
ux
Analysis
0.5
planes
Sewer deflection: mm

ux
a 1.0
Shaft ux
axis 7.5 m 3 m 1.5

Sewer 2.0

PLAN VIEW 2.5 Vertical deflection


Horizontal deflection ux
3.0

237

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.30 Factored output of bending moment in sheet pile wall

Wall bending moment: kNm/m


100 50 0 50 100
0

2
Elevation: m

6
Orthotropic wall OF
Orthotropic wall IF
8 Isotropic wall OF
Isotropic wall IF

10

The output from an identical FE model with isotropic sheet pile wall properties is also
shown and clearly the bending moment was signicantly under-predicted due to the hoop
forces generated in the shell elements.

In Stage 8 with factored ground strength, no failure was apparent in the output. The
strength reduction was continued in Stage 9 and the plot of wall toe deection against
strength factor in Figure 8.31 shows that failure was predicted at a strength factor of
about 2.1. The plot of incremental displacement vectors from the end of Stage 9 in
Figure 8.32 shows that the predicted critical failure mechanism was a bearing failure
beneath the toe of the sheet pile wall.

Figure 8.31 Output of wall toe deflection during strength reduction

2.2

2.0
Ground strength factor

1.8

1.6

1.4

1.2

1.0
0 200 400 600 800
Wall toe deflection: mm

238

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.32 Vectors of incremental displacement at end of strength-reduction stage

In order to check that the undrained shear strength of the clay was not over-predicted
during the undrained analysis stages, a contour plot of mobilised shear strength in terms
of half the deviatoric stress in Stage 4 was plotted adjacent to the measured prole of in
situ undrained shear strength (which was not expected to change signicantly in the
short-term construction case) as shown in Figure 8.33. The maximum mobilised shear
strength of about 40 kPa beneath the toe of the sheet pile wall was well below the
approximate measured shear strength of 130 kPa at this depth, so the FE model did not
appear to be over-predicting undrained shear strength with the Undrained A approach
(see Section 4.2.5).

Hydraulic gradients will be generated by the groundwater lowering measures, so it was


necessary to check that hydraulic failure beneath the base of the shaft was sufciently
unlikely to occur. Since the weight density of the saturated ground and groundwater were
about 20 and 10 kN/m3, respectively, the critical hydraulic gradient for hydraulic failure
was 1.0. Depending on the requirements of the design code, an adequate safety margin
(a safety factor of 1.5, say) would require a hydraulic gradient of less than 0.67 (= 1/1.5).
The plot of steady-state pore pressure with the raised groundwater level in Stage 7 in

239

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Figure 8.33 Output of 0.5 deviatoric stress compared with cu profile determined from site investigation

240
40 kPa
cu: kPa
0 50 100 150 200 250
0
35

5 30
Geotechnical Finite Element Analysis

25
10

Elevation: m
20

15

15

20
cu profile from site investigation 1 3 10
2

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
0
Examples

Figure 8.34 Steady-state pore pressure with 10 m wall toe level

20 kPa

40 kPa

60 kPa

Hydraulic
3.1 m

gradient, 80 kPa
i 0.94

100 kPa

120 kPa

Figure 8.34 shows a maximum hydraulic gradient of about 0.94 which was clearly
unacceptable. Note that the proximity to hydraulic failure was not immediately apparent
from the output in Stage 7, nor during the subsequent strength reduction which showed
adequate safety against geotechnical failure. Verifying adequate safety against hydraulic
failure requires a specic, separate check of the groundwater ow analysis outputs.

A further groundwater ow analysis was performed with a deeper wall toe at 13 m


elevation and the outputs of steady-state pore pressure are shown in Figure 8.35. The
maximum hydraulic gradient was about 0.62, which was acceptable. Therefore, the sheet
pile wall needed to be installed to a toe level of 13 m in order to ensure adequate safety
against hydraulic failure. The FE analysis should then be re-run with the modied toe
level in order to obtain updated outputs of deection, bending moment, etc. (not pre-
sented here).

8.3.5 Validation
Validation in this case was rather difcult because, although alternative analysis
methods, empirical design methods and extensive case study data exist for straight-sided
excavations, there is a paucity of similar sources of validation data for circular shafts. As
mentioned in the previous section, the output of ground settlement behind the retaining
wall in Figure 8.28 appeared credible since its shape matched inversely with the wall
deection prole in Figure 8.27 in accordance with the behaviour observed in many case
studies of straight-sided excavations, e.g. Clough and ORourke (1990).

241

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.35 Steady-state pore pressure with 13 m wall toe level

20 kPa

40 kPa

60 kPa

Hydraulic 80 kPa
3.7 m

gradient,
i 0.62

100 kPa

120 kPa

Moderate uncertainty still existed in the predicted values of ground movement and hence
sewer deection due to the absence of any case study data or alternative analysis
methods which should be taken into account in assessing the reliability of the sewer
deection predictions. Consideration could be given to an observational approach invol-
ving monitoring of ground settlements with contingency measures if ground movements
were found to be larger than expected.

To help validate the outputs from the sheet pile wall, inclinometer readings from a similar
shaft in similar ground and groundwater conditions were available. The excavation depth
of this shaft was 8 m instead of 5 m and the diameter was 18 m instead of 15 m. There was
also a waling beam at 4 m below ground level in addition to the capping beam to provide
additional support to the sheet piles. Although the geometry differed somewhat, if the
deection of the sheet pile wall could be predicted using a modied version of the FE
analysis model, then there would be increased condence in the predictions of the similar
original FE analysis model. Inclinometer data only during construction were available,
so only the short-term undrained conditions were simulated. The FE analysis output and
monitoring data are compared in Figure 8.36 with the toe of the inclinometer set to
match the analysis output. The agreement between the plots was very good which gave
condence in the prediction of sheet pile wall behaviour in undrained conditions.

A judgement would then have to be made on the reliability of the predictions, particularly
in the drained case where no case study data were found. An observational approach
or design changes might be options if the margin of safety was considered too small.

242

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.36 FEA output and monitoring data of sheet pile wall deflection for a similar shaft

Wall horizontal deflection: mm


1 0 1 2 3 4 5 6 7
0

5
Elevation: m

10
FEA output
Monitoring

15

As indicated in Figure 8.31, the margin of safety on geotechnical failure obtained from
the strength reduction was quite high, but what about on hydraulic failure?

8.4. Embankment construction example


8.4.1 Summary
This example concerns the construction of a highway embankment on soft, normally
consolidated clay. The design of the embankment relied on the soft clay gaining strength
due to consolidation under the growing weight of the embankment during construction,
so the construction sequence needed to be timed and monitored carefully.

The particular features of this example include:

g 2D plane strain model


g gravity switch-on initial stress procedure
g coupled consolidation analysis
g double-hardening, stress-dependent stiffness constitutive model for embankment ll
g Modied Cam Clay (MCC) model for clay foundation
g obtaining MCC model parameters from triaxial tests
g modelling vertical wick drains in a plane strain model
g large deformation (Updated Lagrangian) formulation
g ULS check by one-step strength reduction
g validation using basic analysis methods.

8.4.2 Setting up the FE analysis model


Justification for using FE analysis (1.1.1)
An important element of the design of this embankment was the improvement of the clay
foundations strength due to consolidation under the weight of the embankment during
construction. The consolidation time required for the foundation to gain sufcient

243

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

strength for continued embankment construction impacted directly on the construction


programme. Excess pore pressures due to both normal and shear stress changes in the
foundation needed to be predicted, along with their dissipation during construction and
in any intermediate consolidation periods. Such complex behaviour can only really be
predicted using advanced constitutive models implemented using a numerical analysis
technique such as FE analysis.

In addition, the sloping geometry of the clay foundation layer further precluded the use
of simpler analysis methods.

Aims of the model (1.1.2)


1 to predict the timing of the embankment construction sequence necessary to
ensure adequate stability
2 to predict the settlement of the embankment crest on completion of consolidation.

Geometrical simplifications (1.2.1 and 1.2.2)


There were no signicant man-made structural features at the site except for the embank-
ment itself. As shown in Figure 8.37, the proposed embankment had a long, prismatic
geometry that was well suited to the plane strain assumption. The structural geology was
also quite uniform in the same long axis as the embankment, which allowed a 2D plane
strain assumption to be adopted.

From the information available, both the ground surface and clay layer appeared to be
inclined at approximately uniform slopes of about 1 : 10 which were adopted in the FE
model. The sloping ground level and layers meant that no axis of symmetry was present
to allow only part of the plane strain section geometry to be analysed. If the ground
layers were horizontal then an axis of symmetry would have existed down the centre
of the embankment and only half the geometry would have needed to have been
analysed. The embankment was assumed to have a horizontal crest coincident with the
highest proposed carriageway level at +4.0 m level and side-slopes of 1 : 3 on the down-
slope side and 1 : 2.85 on the upslope side. Minor details such as carriageway cambering
and the shoulder were excluded from the FE model.

The embankment would be constructed in a progressive fashion in layers of about 0.3 m


thickness. To include the progression would require a 3D analysis but it was considered
sufciently accurate to assume innitely long and full-width layers in a 2D plane strain
analysis. Adding 0.3 m-thick layers in the FE model would also have been unnecessarily
complicated, so it was considered sufciently accurate to construct the embankment in
three layers of up to 2 m thickness and this was checked by also running an analysis with
thinner layers and comparing the key outputs. Since each layer was constructed in a
coupled consolidation analysis over a time period, the gradual build-up of each 2 m-
thick layer was taken into account in the FE analysis to some extent.

Vertical band drains on a 2.3 m triangular grid were to be installed to the full depth of
the clay layer across the embankment footprint, as shown in Figure 8.37, to hasten
consolidation times. However, installing a vertical drain in a 2D plane strain analysis

244

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.37 Geometry of embankment construction example

10 8m
8m
1 Embankment fill placed
in 3 layers in FE model
Soft clay
+0.5 m +4.0 m
+2.0 m
+0.0 m
10
1
16.5 m Vertical drains on
2.3 m triangular grid

Dense sand

would simulate a continuous trench drain running in the out-of-plane direction which
would cause consolidation times to be under-predicted and potentially lead to an unsafe
design. To account for this, the horizontal permeability of the clay between the vertical
drains in the plane strain model was reduced, as described by Hird et al. (1995).

The volume of soil drained by each vertical drain is equivalent to a cylinder of diameter
1.05 times the drain spacing on an equilateral triangular grid (or 1.13 times the spacing
on a square grid), giving an equivalent radius R of the drained zone for each drain of
1.2 m. The vertical drains were installed in the mesh at 2R (= 2.4 m) centres, as shown
in Figure 8.38. The horizontal permeability in the drained zone of the clay layer was then
corrected according to Equation 8.10 to kpl = 1.8 10 9 m/s, assuming ow rates into
the drains were sufciently high for drain resistance to be neglected and ignoring the
effects of installation-induced smear on soil permeability.

kpl 2
=          (8.10)
kax R k r 3
3 ln + ax ln s
rs ks rw 4 (Hird et al., 1995)

245

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

where rw is the radius of the drain (the band drains had an effective value of 30 mm), rs
the radius of the shear zone (assumed equal to rw , i.e. smear ignored).

Model boundaries and fixities (1.2.3 and 1.2.4)


No rules of thumb on locating boundaries were known for an unusual geometry such as
this so, after some trial and error, it was found that the vertical boundaries to the mesh
needed to be placed about 40 to 50 m from the embankment toe in order to eliminate
signicant boundary effects. A greater distance was needed on the downslope side, as
shown in Figure 8.38, because more deformation of the clay foundation occurred on
this side. The bottom boundary could be placed closer because the dense sand layer was
signicantly stiffer than the soft clay.

The bottom boundary was xed in both axis directions and the vertical boundaries were
xed only in the horizontal direction.

Finite element mesh (1.3)


Cubic strain 15-node triangular elements with three degrees of freedom per node were
used throughout the mesh. The mesh was generated as shown in Figure 8.38 with smaller
elements created at the embankment foundation where steep stress, pore pressure and
strain gradients were expected. A further FE analysis was performed with a ner mesh
and it was found that the critical outputs were not affected signicantly. Therefore, the
adopted mesh was considered acceptable.

Initial stresses (1.4.1)


Since the ground surface and soil layers were inclined, the initial stress prole was not
uniform across the model and so could not be specied simply with pore pressures,
ground density and K0 values. The initial stresses were generated by activating the
self-weight of the ground (gravity switch-on).

Steady state pore pressures were generated in a groundwater ow analysis with the
groundwater level at each side of the mesh assumed at 1 m below ground level, vertical
boundaries were assumed open and the bottom boundary closed. Pore pressures were
assumed zero above the groundwater level. The clay density was taken as 18 kN/m3
above groundwater level while the saturated clay and sand densities were assumed to
be 20 and 18 kN/m3, respectively.

In order to establish in situ stresses, a basic LEPP MohrCoulomb model was adopted
for both strata (and remained so throughout all stages for the sand layer). In accordance
with Equation 1.2, Poissons ratios of 0.38 and 0.33 were adopted for the clay and sand
to obtain target K0 values of 0.6 and 0.5, respectively. The self-weight of the soil layers
were activated in drained conditions. While attempting to obtain a realistic initial stress
state in subsequent analysis stages with the MCC model, it was found that the initial
Poissons ratio for the clay in the LEPP model needed to be changed to 0.34. At the end
of the analysis stage activating the grounds self-weight, no plastic points were evident in
the output.

246

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Figure 8.38 FE mesh for embankment construction example

41 m 6m 16 m 18 m 49 m

Node B Nodes and integration points


selected for continuous output
Node C Node A

Int. pt. Z
2m

Soft clay

28 m
Drains
21 m

Side fixed in X direction


Dense sand

8m
Side fixed in X direction

Base fixed in X and Y directions


Y

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

247
Geotechnical Finite Element Analysis

Following establishment of initial stresses with the basic constitutive models, the clay
model was changed to the MCC model. Some further manipulation was necessary to
obtain what was considered to be a realistic initial stress state and yield surface. This can
often be achieved with temporarily enhanced ground densities or by applying vertical
stress to the ground surface and then releasing it to simulate pre-loading due to depo-
sition or groundwater lowering. Due to the variable thickness of the clay layer in this
example, it was found that more credible results were obtained by applying a uniform
0.3 m vertical displacement downwards to the ground surface and then releasing it. This
was performed in drained conditions and achieved the stress and over-consolidation
ratio (OCR) proles in the centre of the mesh shown in Figure 8.39, which were con-
sidered to be realistic based on the site investigation information available. The analysis
outputs were quite sensitive to the initial stress prole so a parametric study (not shown
here) was undertaken to assess the reliability of the outputs.

As described under Construction stages in this section, strength reduction was performed
in dedicated ULS stages in order to check the stability of the embankment during
construction. Another option, as described in Section 6.3.2, is to adopt factored strength
parameters for the ground from the start and throughout all analysis stages. The draw-
back of this approach is that the stress state and predicted behaviour of the model can
become unrealistic. To illustrate this, the same procedure described previously to estab-
lish the initial stress state was repeated but with the shear strength factored, as described
later in this section, from the start. The output of the initial stress state is shown in
Figure 8.40 for comparison with that shown in Figure 8.39. Clearly, a very different
initial stress state was obtained in the soft clay and, if the analysis were continued
through its subsequent stages, very different and probably inaccurate outputs would
be obtained throughout.

Figure 8.39 Initial K0 and OCR profiles established in soil layers in FE analysis

K0 or OCR
0.0 0.5 1.0 1.5 2.0 2.5
0

10
Elevation: m

Soft clay

15
Dense sand
20

K0
25
OCR

30

248

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.40 Initial K0 and OCR profiles established in soil layers in FE analysis (factored strength
parameters)

K0 or OCR
0.0 0.5 1.0 1.5 2.0 2.5
0

10
Elevation: m

Soft clay

15
Dense sand
20

K0
25
OCR

30

Construction stages (1.4.2)


The following stages were run in the FE analysis which, apart from the drained stages to
establish the initial stresses, were all coupled consolidation stages. The ULS stages were
consolidation stages with zero time which were equivalent to Undrained A analyses
using the active pore pressure distribution from the most recent consolidation stage.

The sequencing of the construction stages is illustrated in Figure 8.41 while each stage is
described in detail as follows:

1 Steady-state pore pressures


Since the groundwater regime was non-hydrostatic, steady-state pore pressures
were established by groundwater ow analysis with groundwater level set at 1 m
below ground level at left and right model boundaries.
2 Gravity switch-on
Activate in situ ground elements in drained conditions with basic LEPP models.
3 MCC model for clay
Change constitutive model for clay stratum to MCC model.
4 Pre-load
Apply 0.3 m vertical displacement to ground surface under drained conditions to
simulate pre-loading due to deposition or groundwater lowering.
5 Release pre-load
Release displacement under drained conditions to create OCR prole.
6 Install drains (0 days)
The vertical drains as shown in Figure 8.38 were installed which set excess pore
pressure as zero along their length (the steady-state pore pressure remained
unchanged). The horizontal permeability of the clay between and within 1.2 m of
the drains was reduced to the corrected value as described under Geometrical
simplications in this section. This and all following stages were performed as
coupled consolidation analyses with the time interval indicated in brackets.

249

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.41 Sequence of analysis stages

1. Steady-state pore pressure

2. Gravity switch-on

3. MCC model for clay

4. Pre-load

5. Release pre-load Establishing initial state

6. Install drains Construction stages

7. Embankment layer 1 8. ULS strength reduction

9. Intermediate consolidation

10. Embankment layer 2 11. ULS strength reduction

12. Intermediate consolidation

13. Embankment layer 3 14. ULS strength reduction

15. Complete consolidation

7 Install embankment layer 1 (5 days)


Activate layer 1 embankment ll elements. In this stage, a construction period of
5 days to install layer 1 was assumed. Nodal displacements were reset to zero at
the beginning of this stage. In the analyses where these were used, the large
deformation (Updated Lagrangian) formulation and updated water pressures
were activated from this stage onwards.
8 Layer 1 ULS strength reduction (0 days)
Ground shear strength in all strata reduced in one step.

250

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

9 Intermediate consolidation (90 days)


Continuing from Stage 7, dissipation of excess pore pressures was allowed for
90 days with no other changes in the model.
10 Install embankment layer 2 (5 days)
Activate layer 2 embankment ll elements. A construction period of 5 days was
assumed.
11 Layer 2 ULS strength reduction (0 days)
Ground shear strength in all strata reduced in one step.
12 Intermediate consolidation (80 days)
Continuing from Stage 10, dissipation of excess pore pressures was allowed for
80 days with no other changes in the model.
13 Install embankment layer 3 (5 days)
Activate layer 3 embankment ll elements. A construction period of 5 days was
assumed.
14 Layer 3 ULS strength reduction (0 days)
Ground shear strength in all strata reduced in one step.
15 Complete consolidation (to 1 kPa maximum excess pore pressure)
Continuing from Stage 13, dissipation of excess pore pressures was allowed until
the maximum value anywhere in the clay stratum was 1 kPa.


The groundwater level was left unchanged in the ULS stages because signicant level
changes were highly unlikely in a low-permeability soil over short construction time-
scales. Input factoring (strength reduction) was performed. Output factoring was not
necessary because only ground stability was being considered.

Regarding strength reduction, several issues needed to be overcome:

1 The strength factors required by design codes usually differ between drained and
undrained strengths whereas this example involved drained embankment ll and a
partially drained (consolidating) clay foundation. So, each required a different
strength factor that could be applied either individually to each layer or else a
uniform, stepwise strength reduction could have been performed and then the
output viewed to identify through which soil layers the failure mechanism passed
and which strength factor requirement was appropriate.
2 Design codes usually provide partial factors to be applied to w , c and cu whereas
the MCC model does not use these parameters directly to dene the failure
criterion. Furthermore, the FE analysis software did not have a rigorous strength-
reduction routine for constitutive models with failure criteria other than those
dened by w , c and cu .
3 The prediction of shear strength in undrained or partially drained conditions
(using the Undrained A approach) is not straightforward and is heavily
dependent on the prediction of excess pore pressure. Simply applying partial
factors from design codes is not an adequate approach.

To help overcome Issues 2 and 3, a large number of CAU triaxial compression test
simulations were performed using a single stress point algorithm covering the range of

251

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

stress states in the clay foundation. Initially, the simulations were performed with the
characteristic model parameters described in Section 8.4.3 (including M = 1.07), then
the simulations were repeated with the M parameter reduced until the output of shear
strength from each test simulation was reduced by the factor required in the design code
(in this case 1.4 for undrained strength). It was found that the computed undrained shear
strength was reduced by at least a factor of 1.4 in most cases when M was 0.72. However,
the factor was signicantly less when K0 and OCR were low, but such stress states were
located deeper in the clay foundation, below where failure mechanisms were expected to
occur.

No stepwise strength-reduction procedure was available for the MCC model in the FE
analysis software, so the shear strength parameters were reduced to their factored values
directly in the model parameters in the ULS analysis stages leaving the programs stress
point algorithm to correct any stress states violating a yield surface. Stress path outputs
were checked as described in Section 8.4.4. This method of strength reduction also over-
came Issue 1 because the shear strengths of each soil layer could be reduced by different
factors according to the requirement on drained (for embankment ll and dense sand
layer) and undrained (for soft clay layer) shear strength.

Calculation options (1.4.3)


The Modied NewtonRaphson solution scheme was adopted with arc length control.
Automatic step-sizing was utilised for both displacement and consolidation analysis and
the maximum equilibrium error was set at 1%.

The small deformation (Total Lagrangian) formulation was adopted in all analysis
stages in the rst runs of the analysis. Up to 1.2 m settlement of the embankment was
predicted which was large enough to cause the small deformation assumption where the
original mesh geometry is retained in calculations to become unrealistic. In particular,
highly deforming ground below or near the groundwater level would, in reality, experi-
ence signicant steady-state pore pressure change due to its displacement relative to the
stationary groundwater level. Therefore, an additional analysis run was performed with
Updated Lagrangian formulation adopted from Stage 7 onwards and then the outputs
were compared as described in Section 8.4.4.

8.4.3 Obtaining parameters and constitutive model features


Constitutive model selection
Normally and lightly over-consolidated clays are highly compressible which is behaviour
particularly suited to the MCC model (described under Isotropic hardening single surface
plasticity in Section 2.3.2) which has stress-dependent primary and unloadreload stiff-
ness dened in terms of logarithmic isotropic compression curves. The MCC model also
has a yield surface dened in terms of a pre-consolidation stress in order to simulate the
effects of over-consolidation that existed near the ground surface. This model was
implemented with a MohrCoulomb failure surface for more robust strength prediction.

Signicant anisotropy of shear strength and stiffness is common in lightly over-


consolidated clays, particularly those of low plasticity, which can have a signicant effect

252

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

on the analysis of embankment construction (see Zdravkovic et al., 2002). In the soft
clay in this example, no signicant anisotropy was recorded in the site investigation and
it was considered acceptable to adopt an isotropic MCC model.

Due to the large range of stresses in the embankment ll material from near-zero at
placement to high stresses as subsequent layers are placed on top, and because the
material is essentially normally consolidated when placed, highly non-linear elastic and
work hardening behaviour should be expected. Therefore, a double-hardening, hyper-
bolic model was adopted with stress-dependent stiffness. A simple MohrCoulomb
failure criterion was adopted because the strength of the soft clay foundation was
expected to govern the stability of the embankment during construction.

The in situ, dense sand layer was expected to undergo insignicant stress and strain
changes, so a simple LEPP MohrCoulomb model was adopted for the dense sand.

Obtaining parameters
The parameters l, k and N for the MCC model of the soft clay were obtained from iso-
tropic consolidation tests on specimens of the soft clay in a triaxial cell as shown in the
plot of specic volume v against ln p in Figure 8.42. Even on a log plot, both the
unloadreload and normal compression lines are curved rather than linear, so tangents
to these lines at stress levels appropriate for the analysis were drawn for deriving the
parameters. There was an unusually large difference between l and k, probably due
to the clay being structured.

For the derivation of M, triaxial test stress paths may be plotted in qp space and the
slope of the critical state line determined directly. Alternatively, as in this example, M

can be derived from the critical state friction angle wcs . The stress paths from four CAU
triaxial tests (see Section 3.3.1) are shown in ts space in Figure 8.43 [s = (s1 +s3 )/2

and t = (s1 s3 )/2]. A best t line was drawn from the origin through the end points

Figure 8.42 Isotropic consolidation test data and derivation of MCC model parameters

N0 = 4.31
3.0

2.6 = 0.028
v

2.2

= 0.37

1.8
0 2 4 6 8
ln p: kPa

253

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.43 CAU triaxial test stress paths and derivation of M

30

20
t : kPa

10 sin cs

1 Reconsolidation
stress path
0
0 10 20 30 40 50 60
s: kPa

of the stress paths (where they became stationary and where the pore pressure had stabil-

ised). The slope sin wcs was 0.453, giving wcs = 278. The slope of the critical state line M
can then be determined using one of the following equations:

6 sin wcs
triaxial compression: M = (8.11)
3 sin wcs

6 sin wcs
triaxial extension: M = (8.12)
3 + sin wcs

In this example, for triaxial compression, M = 1.07. For a plane strain analysis the M
value could be raised slightly to take account of the intermediate principal stress (as
described in Section 3.4.1) but for conservatism in the ULS check, the M value remained
unchanged in this example.

The permeability of the clay was derived from the isotropic consolidation tests and in
situ constant head tests. Vertical and horizontal permeabilities of 1 10 9 and 1
10 8 m/s, respectively, were obtained. Higher horizontal permeability occurred due to
the presence of silty laminations in the clay.

The variation of permeability in both directions with changes in void ratio e was taken
into account in the model according to Equation 8.13
 
k De
log = (8.13)
k0 1.2

where k and k0 are the active and initial permeabilities, respectively.

As in most cases, there was considerable uncertainty in the permeability values so a


parametric study (not presented here) of permeability to assess the reliability of predic-
tions inuenced by permeability would need to be made.

254

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Table 8.8 Derived model parameters for the soft clay

M l k N gsat gunsat kv kh

1.07 0.37 0.028 4.31 20 kN/m3 18 kN/m3 1 10 9 m/s 1 10 8 m/s

Table 8.9 Model parameters for the embankment fill

Kur n n Ki or K50 Koed Rf w c c g

750 0.5 0.2 455 or 250 250 0.9 358 0.1 kPa 08 16 kN/m3

Table 8.10 Model parameters for the dense sand

E n w c c g

200 MPa 0.25 368 0.1 kPa 08 18 kN/m3

In summary, the parameters shown in Table 8.8 were derived for the soft clay in this
example.

Obtaining parameters for proposed earthworks structures is obviously difcult because


they are yet to be installed. If the material to be used can be obtained, it is possible to
prepare trial structures in situ or trial specimens in the laboratory for parameter testing.
Similar materials may have been used in similar, existing structures that can be tested or
back-analysed to obtain parameters, or case study data may exist. For example, Duncan
et al. (1980) summarised parameters for a range of ll materials for the hyperbolic
Duncan and Chang (1970) model.

In this example, the embankment ll parameters were obtained based on experience of


using the same material in other earthworks structures. The model parameters are shown
in Table 8.9 and the symbols have the meanings described in Section 8.2.3.

The model parameters adopted for the LEPP MohrCoulomb model for the dense sand
are shown in Table 8.10. A high linear stiffness was adopted to reect the high conning
stress and expected low strains in this layer.

8.4.4 Output
The vertical settlement of the three nodes selected for continuous output (see Figure 8.38)
is shown plotted against time in Figure 8.44. The settlement resulting from the placement
of each layer of the embankment is clearly visible. Node C experienced only minor settle-
ment due to the placement of Layer 1 because it was located at the edge of Layer 1 where
the wedge of ll material was very thin. Note that the displacement output for Node B at
the crest of the embankment needed to be corrected back to zero because it already had

255

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.44 Output of settlement (small deformation formulation)

Time: days
0 200 400 600 800
0.0
Node A
0.2 Node B
Node C
0.4
Settlement: m

0.6

0.8

1.0

1.2
Layer Layer Layer
1 2 3

106 mm settlement caused by the settlement of lower layers associated with it when it
was activated on placing Layer 3. The total construction time calculated for all excess
pore pressures to fall below 1 kPa was 812 days.

The same output is shown in Figure 8.45 but from the analysis with large deformation
(Updated Lagrangian) formulation and updated water pressures. Signicantly less settle-
ment and shorter consolidation times (605 days total construction time for excess pore
pressures to dissipate to below 1 kPa) were predicted, so using this large deformation
formulation could potentially lead to a more economic design. The ULS checks were
also more favourable with the large deformation formulation, but experience shows that

Figure 8.45 Output of settlement (large deformation formulation)

0.2 Time: days


0 200 400 600 800
0.0

0.2
Settlement: m

0.4 Node A
Node B
0.6 Node C

0.8

1.0

1.2
Layer Layer Layer
1 2 3

256

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.46 Output of Node A displacement due to strength reduction

100

Analysis stage progression: %


80

60

40 Layer 1
Layer 2
Layer 3
20

0
0 10 20 30 40
Node A displacement change: mm

reasonably accurate failure predictions are often obtained with the small deformation
(Total Lagrangian) formulation in spite of the large deformations. Therefore, the large
deformation formulation should not normally be relied upon for ULS verications
and in this example the ULS check was performed using the more conservative small
deformation formulation.

One of the main aims of this analysis was to predict the intermediate consolidation times
necessary in order to construct the embankment with adequate stability. To check this,
the shear strength of both the embankment ll and the clay foundation were factored
by changing the model parameters in Stages 8, 11 and 14 immediately after placing each
layer of the embankment. The maximum displacement of Node A during these stages
was plotted against the progression of each stage as shown in Figure 8.46. The reduc-
tion in shear strength after placing each layer was completed without any onset of
failure apparent at Node A, suggesting that the intermediate consolidation times were
adequate.

Rather than rely on the displacement output of a single node, it is of course important to
view the displacement of the whole model when the shear strength has been reduced, to
check for failure mechanisms elsewhere. Total displacements are useful but shear failures
may be masked by the deformations occurring prior to failure. It is a good idea to plot
shear strains or, in this case, vectors of displacement in the last increment of the strength-
reduction stages, as shown for Stages 11 and 14 in Figures 8.47(a) and (b). When
reducing strength, failure is most likely at the end of the stage when the strength is at its
minimum value, so plotting the change in displacement in the last increment of the stage
is more likely to highlight any failure mechanisms and omits the displacements occurring
earlier in the stage. The relative magnitude of the incremental displacements (as
represented by the size of the arrows) at different nodes is more important than the
actual magnitude of each nodal displacement in this case. Failure mechanisms are often
indicated by a sudden change in magnitude across a shear plane. In these cases, no
failure was evident. However, in the plots shown in Figures 8.47(c) and (d) at the same

257

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.47 Vectors of incremental displacement at end of strength-reduction stages: (a) Layer 2
no failure; (b) Layer 3 no failure; (c) Layer 2 (inadequate consolidation time) failure; (d) Layer 3
(inadequate consolidation time) failure

(a)

(b)

(c)

(d)

258

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

Figure 8.48 Output of Node A displacement due to strength reduction (with inadequate
consolidation times)

100

Analysis stage progression: %


80

60

40 Layer 1
Layer 2
Layer 3
20

0
0 10 20 30 40
Node A displacement change: mm

stages but with inadequate intermediate consolidation times, the sudden change in
displacement magnitude across each shear plane is clearly evident.

The Node A displacements from the same analysis with inadequate intermediate consoli-
dation times are also plotted in Figure 8.48 for comparison with those in Figure 8.46.
The intermediate consolidation times were reduced to 60 and 50 days following place-
ment of Layers 1 and 2, respectively. The strength-reduction stage following placement
of Layer 2 reached 94% completion before failing to converge, while the stage following
placement of Layer 3 reached only 90% completion.

Particularly when using advanced constitutive models, stress paths should be checked for
their response during strength reduction. This is because strength reduction can have
unexpected and potentially non-conservative effects on complex models. In this
example, because a robust strength-reduction procedure was not available in the soft-
ware for the MCC model, the shear strength was factored in one step and the softwares
stress point algorithm left to correct any stress states lying outside a yield surface. The
stress path of deviatoric stress q against average normal effective stress p at Integration
Point Z (Figure 8.38) from the analysis with adequate intermediate consolidation times is
shown in Figure 8.49. The stress path during the ULS stages, shown in black, appeared
to be credible for the decrease of strength with no unexpected stress path excursions
which gave more condence in the strength-reduction operation. Stress paths at other
integration points in the vicinity of credible failure mechanisms in the clay foundation
were also checked in a similar fashion.

8.4.5 Validation
The total settlement of the original ground surface beneath the centre of the embankment
was estimated using Terzaghis one-dimensional consolidation theory. Since the average
thickness (16 m) of the clay layer was signicantly less than the width of the embankment
(40 m), the stress change in the clay was assumed uniform and equal to the surface
surcharge imposed by the embankment.

259

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Figure 8.49 Stress path at Integration Point Z

25
Consolidation

20 Layer 3
construction
Layer 2
15 construction Consolidation
q: kPa

10 Layer 1
construction Consolidation

0
0 10 20 30 40 50
p: kPa

With a coefcient of compressibility mv of 0.9 MPa 1, which was considered appropri-


ate for the stress level in the centre of the clay layer, a consolidation settlement of 0.46 m
under Layers 1 and 2 (average surcharge 32 kPa) was obtained. Under all three layers,
the average surcharge of 60 kPa gave a consolidation settlement of 0.86 m. Typically
in soft clays, consolidation settlement accounts for about 90% of total settlement.
Taking this into account, the total settlement became about 0.51 m and 0.96 m under
Layer 1 + 2 and all three layers, respectively. The settlement output for Node C at the
original ground surface beneath the centre of the embankment shown in Figure 8.45 with
large deformation formulation reached 0.80 m under all three layers. Extrapolating the
curve after placement of Layer 2 suggests that the total settlement would have been
about 0.40 m. These values are about 80% of those obtained from one-dimensional
consolidation theory which, given the assumptions and conservatism of this simplied
method, would be about the level expected, and so provided more condence in the
outputs of settlement from the FE model. In fact, the values obtained from one-dimen-
sional consolidation theory were more similar to the outputs of settlement from the FE
analysis with small deformation formulation (Figure 8.44).

By making the reasonable assumption that drainage from the clay layer occurred only
in the horizontal direction toward the vertical drains, it was possible to estimate the
consolidation time of the clay foundation simply from the Kjellman (1948) equation for
consolidation by radial drainage into vertical drains (Equation 8.14).
 
D2 D 1
t= ln 0.75 ln (8.14)
8ch d 1 U

where t = time, ch = horizontal coefcient of consolidation, D = zone of inuence of


drain, d = equivalent diameter of drain and U = average degree of consolidation.

260

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Examples

For the arrangement of vertical drains in this example, D was 2.4 m and d was 0.06 m.
An average ch value of 15 m2/year was estimated based on laboratory test data. For 95%
consolidation, Equation 8.14 gave a consolidation time t of 154 days. Assuming full
consolidation was achieved when excess pore pressure dissipated below 1 kPa, then
95% consolidation occurred at a settlement of Node C of 0.76 m which corresponded
with a total consolidation time of about 320 days (Figure 8.45). Taking into account the
intermediate consolidation periods, it was estimated from Figure 8.45 that a theoretical
instantaneous loading by all three embankment layers would have taken about 180 days
to reach 95% consolidation. Since the FE model took into account the decreasing
permeability of the clay due to consolidation, it was considered reasonable that the
FE model predicted a slightly longer consolidation period.

Although these validation exercises gave more condence in the accuracy of the FE
analysis outputs, continuous site monitoring of embankment settlement should still be
recommended due to the residual uncertainty and the safety implications of judging the
necessary consolidation times between placing layers of ll. Site monitoring could also
bring the economic benet of hastening construction times if consolidation of the clay
foundation were recorded to occur more rapidly than predicted by the FE model.

REFERENCES
Bellotti R, Ghionna V, Jamiolkowski M, Robertson PK and Peterson RW (1989) Inter-
pretation of moduli from self-boring pressuremeter tests in sand. Geotechnique 39(2):
269292.
Benz T, Vermeer PA and Schwab R (2009) A small-strain overlay model. International
Journal for Numerical Methods in Geomechanics 33(1): 2544.
Clough GW and ORourke TD (1990) Construction induced movements of insitu walls.
Proceedings ASCE Conference on Design and Performance of Earth Retaining Structures,
Cornell, ASCE Pub. no. 25, pp. 439470.
Duncan JM and Chang YC (1970) Nonlinear analysis of stress and strain in soils. ASCE
SM5 96: 16291653.
Duncan JM, Byrne PM, Wang KS and Mabry P (1980) Strength, stressstrain and bulk
modulus parameters for nite element analysis of stresses and movements in soil masses.
Geotechnical Engineering Research Report No. UCB/GT/80-01, University of Califor-
nia, Berkeley, CA.
Fraser RA and Wardle LJ (1976) Numerical analysis of rectangular rafts on layered foun-
dations. Geotechnique 26(4): 613630.
Hemsley JA (1998) Elastic Analysis of Raft Foundations, 1st edn. Thomas Telford,
London.
Hird CC, Pyrah IC, Russell D and Cinicioglu F (1995) Modelling the effect of vertical
drains in two-dimensional nite element analyses of embankments on soft ground.
Canadian Geotechnical Journal 32(5): 795807.
Houlsby GT, Wroth CP and Clarke BG (1986) Analysis of the unloading of a pressure-
meter in sand. Proceedings of the 2nd International Symposium on Pressuremeter and its
Marine Applications. ASTM, SPT950, 245262.
Hughes JMO, Wroth CP and Windle D (1977) Pressuremeter tests in sands. Geotechnique
27(4): 455477.

261

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis

Jardine RJ, Potts DM, Fourie AB and Burland JB (1986) Studies of the inuence of non-
linear stressstrain characteristics in soil-structure interaction. Geotechnique 36(3): 377
396.
Kjellman W (1948) Consolidation of ne-grained soils by drain wells. Transactions ASCE
113: Contribution to the discussion.
Kondner RL (1963) Hyperbolic stressstrain response: cohesive soils. ASCE SM1 82: 115
143.
Muir Wood D (1990) Stain-dependent moduli and pressuremeter tests. Geotechnique 40(3):
509512.
Whittle RW (1999) Using non-linear elasticity to obtain the engineering properties of clay
a new solution for the self boring pressuremeter. Ground Engineering 32(5): 3034.
Zdravkovic L, Potts DM and Hight DW (2002) The effect of strength anisotropy on the
behaviour of embankments on soft ground. Geotechnique 52(6): 447457.

262

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Geotechnical Finite Element Analysis
ISBN 978-0-7277-6087-6

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/gfea.60876.263

Index

Page locators in italics refer to gures separate from the corresponding text.

2D axisymmetric models, concepts 9, 11 approximations and error potential 186


2D continuum elements 127, 130136 assessment of accuracy 188191
2D vs. 3D models 7 associated ow 42
2D plane strain models 710, 143146 at rest pore pressure 106107
3D continuum elements 127, 130132 axes of symmetry 12, 16
3s rule 193194 axisymmetric simulations
2D 9, 11
accuracy 183197 oedometer tests 9293
assessment 188191 plate load tests 93, 94
initial checks 189 pressuremeter tests 93, 95
limit states 164166 structural representations 143144, 146148
modelling elements 131 triaxial cells 9293
observational method 196197
output comparisons 189191 bar elements 125128
parameter acquisition 8587, 9198, 186 beam elements 126, 128
parametric studies 192196 bearing resistance 178
reliability testing 166167 bender elements for triaxial cells 6466
responsibilities 183184 bending moment outputs
sensitivity analysis 167, 192 raft and pile example 214220
sources of error 185188 shaft excavation example 237, 238
acquisition see parameter acquisition BIM see building information modelling
active pore pressure 108 block sampling, soils 59
analysis planning 17 bonding, soils 3233, 9798
aims 2 boundaries
design integration 57 dening 1215
information gathering 24 examples 202, 203204, 226, 246, 247
software packages 45 xities 1516, 202, 203204, 226, 246, 247
utility 12 threshold positions 14
analytical solutions 190 boundary effects 12-16, 165
anisotropy bubble models 4546
consolidation tests 69 building information modelling (BIM) 4
derivation 8889 bulk modulus 115
geometrical 148150
linear elasticity 38, 48 calculation options, setup 2426
rocks 48 cap hardening 4445
soils 33, 34, 38, 4647 case histories 191
stiffness tests 6869, 8081 case studies 96

263

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

characteristic value derivation 8687, 91 destructuration 46


city centres 12 elasticity 3639
clays embankment construction 252253
embankments 112, 243261 errors 186
at plate dilatometer tests 74, 82 failure surfaces 4243
over-consolidated 116117, 225243 ow rules 42
partially saturated 106 ground behaviour 3136, 4243
permeability tests 74, 8485, 253255 hypoplasticity 47
piezocone penetration test 74, 8182 limit states 166
plate load tests 74 plasticity 3947
pre-consolidated stress 97 principal stresses 33
pressuremeter tests 7779 raft and pile example 206
seismic testing 74, 8081 rocks 3537, 4748
shaft excavation procedure 225243 selection effects 2930
shear strength 98 soil behaviour 3134
standard penetration tests 82 types 3648
tunnelling models 143 yield surfaces 3942
closed-form solutions 190 construction stages
closed hydraulic boundaries 119 consolidation analysis 120123
closed piezometers 63 dual factoring 173175
coefcient of variation 194 embankment construction 249252
cofferdam hydraulic boundaries 121 groundwater analyses 109113
cohesion, derivation 91 limit states 165
comparisons for validation raft and pile example 205206
case histories 191 setup 2124
known solutions 189191 shaft excavation example 228231
numerical analysis methods 191 simplication errors 187188
site monitoring data 191 strength reduction 174179, 181
concrete 24, 51, 142143, 152155, 181 continuum elements
cone pressuremeters (CPMT) 77 connection 150151
consolidation properties 127, 130136
coupled analysis 122123 raft and pile model 202
embankment construction example 243261 continuum models, rocks 3637, 4748
groundwater effects 107111, 115, 120123 conventional methodologies versus FE
limit states 165 analysis 12
low-permeability soils 120122 Coulomb friction criterion 135, 213
oedometer tests 72 coupled consolidation analysis 122123, 165
parameters 27 coupled groundwater ows 118
pore pressure 107111 CPMT see cone pressuremeters
Terzaghis one-dimensional theory 259261 CPTu see piezocone penetration tests
triaxial cell tests 6668 cracking, concrete 154
constant head tests 83, 84 creep 34, 47, 68, 75, 154155
constant rate of strain (CRS) oedometers 72 CRS see constant rate of strain
constitutive models 2953 cubic strain elements, concepts 17
anisotropy 3334, 38, 4647 cut slopes, constitutive models 49
applications 4851
appropriateness 3031, 4951 deep tunnels 15
choice in setup 26 degrees of freedom 125130, 131132
concepts 29 derivation
creep 34, 47 anisotropy 8889

264

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

characteristic values 8687, 91 parameter derivation from undrained tests


cohesion 91 90
dilation angles 89 parameters 27
initial state parameters 8990 drains 120, 244247, 245246
intermediate principal stress 8788 driven piles
parameters 8591 constitutive models 50
permeability 85 installation effects 23
Poissons ratio 9091 output factoring 171172
pre-consolidation stress 8990 DruckerPrager failure surface 43
stress ratio 8990, 97 dry, granular soils, modelling 105
useful equations 9799 dual factoring 172175
design dynamic analysis 51
analysis planning 57
integration 6 earth walls 137138
design charts 190 effective stress
design codes 163182 groundwater effects 114115, 116117
rock limit states 180 sample disturbance 60
serviceability limit states 163, 164, 167168 strength reduction 179, 238239
structural limit states 180181 elasticity
ultimate limit states 163, 164, 168181 anisotropic linear 38, 48
destructuration 46, 60 constitutive models 3639
deterministic reliability testing 166167 equations 98
deviatoric stress 41, 112113 linear 38, 155156, 180184
diaphragm piezometers 63 materials 152156, 180181
diaphragm wall installation 2223 non-linear 3839
dilation angles 89, 107108, 166, 177 structural limit states 180181
direct shear tests 72 elements
discontinuities in rocks 3536, 37, 4748 choice 130
discrete crack approach 154 connections 150151
discretisation of geometry 164, 187 construction stages 2124
displacement piles continuum-type 127, 130136, 150151
constitutive models 50 end bearing 131, 134, 150151
installation effects 23 hierarchy 17
output factoring 171172 initial stresses 1821
distributed load modelling 152 interface-type 132136, 204, 227, 236
disturbance of soil samples 6061 limit states 164
DMT see at plate dilatometers non-continuum 125134, 150151
double corebarrels 60 partial factors 181
double surface plasticity 4445 raft and pile model 202205
drainage self-weight 131, 133
embankment construction 245247, 249, size 131
260261 stiffness 130, 135
low-permeability soils 112 types 125132
model setup 27, 120, 244247 embankments
drained, denition 109 constitutive models 50
drained analysis construction stages 249251
assumptions 109113 example procedure 243261
concepts 109 hydraulic boundary conditions 121
limit states 165166 model setup 243252, 245246
methodology 113 model validation 259261

265

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

embankments (continued ) uid concrete 155


outputs 255259 uid support 23 MERGE
parameter acquisition 252255 foundations
pore pressure 112 constitutive models 5051
strength reduction 248252, 257259 example procedure 199224
test procedures 57 hydraulic boundary conditions 121
unconned groundwater ows 118119 input factoring 168170
empirical design methods 190 modelling 138139
empirical relationships, stress ratio 97 output factoring 170, 171173, 174
end bearing 131, 134, 150151 fractured rocks 3536, 37, 4748, 58, 82
engineering managers 184 free-draining boundaries 119
engineers responsibility 183 friction angles 166
equilibrium errors 25 friction hardening 45
errors see also shear hardening
management 192197 full connections 150151
potential sources 185188 full-ow penetrometers 8182
EulerBernoulli theory 128
evaporation 120 gathering information 24
excavations Geological Strength Index (GSI) 48
constitutive models 49 geometry
hydraulic boundary conditions 121 2D axisymmetric assumptions 9, 11
pore pressure 107, 112, 121 2D plane strain assumption 710, 143146
shaft modelling example 225243 anisotropy 3334, 148150
test procedures 57 boundaries
excess pore pressure 107108, 112113, 243247 examples 202, 203204, 226, 246, 247
existing structures 34 xities 1516
explicit discontinuity modelling 48 location 1215
extension 3334 discretisation 164, 187
external load input factoring 168170 elements 125136
extraction boundaries 119120 principal stress 33
extrusion, 2D plane strain models 910 rock behaviour 3637
setup 716
failure surfaces, constitutive models 4243 simplication
falling head tests 83 detail needed 9, 1112, 13
nite element analysis embankment construction example
aims 2 244247, 246
utility 12, 163164 errors 187
xed connections 150151 raft foundation with piles example
xities 1516, 202, 203204, 226, 246, 247 200202
at plate dilatometers (DMT) 74, 82 shaft excavation example 225226
ows structures 125152
groundwater geosynthetics, modelling 137138
analysis 118120 geotechnical parameters
embankment construction 245247, 249, errors 186
260261 resistance calculations 177178
limit states 166 see also parameter acquisition
meshing 17 gravelly ground, in situ testing 9495
permeability 85, 253255 gravity switch-on 2021, 247
shaft excavation example 239241 greeneld conditions 19, 6668
Sichardts empirical formula 15 ground anchors 136137, 137138

266

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

ground behaviour HoekBrown model 48


modelling 3136, 130136, 156160 horizontal effective stress 1920
strength factoring 174179, 181, 238239 HPD see high-pressure dilatometers
ultimate limit states 163, 164, 168181 HS Small model 39
ground improvement 22, 140141 hydraulic boundary conditions 119121, 121,
ground information requirements 3 230
groundwater 105123 hydraulic triaxial cells 6466
assumptions 109113 hyperbolic stressstrain relationships, sand
boundary conditions 119121 211, 212
bulk modulus 115 hypoplasticity 47
consolidation analysis 115, 120123
coupled ows 118 impermeable boundaries 119
drainage 27, 112, 120, 244247 implicit discontinuity modelling 4748
drained analysis 90, 113 inltration boundaries 119120
effective stress analysis 114115, 116117 information gathering 24
errors 188 see also parameter acquisition
ows infrastructure, information requirements 34
analysis 118120 initial checks for accuracy 189
embankment construction 245247, 249, initial stresses
260261 derivation 8990
limit states 166 direct specication 20
meshing 17 embankment construction example 247
permeability 85 gravity switch-on 2021, 247
shaft excavation example 239241 limit states 164165
Sichardts empirical formula 15 raft and pile example 205
initial stresses 19 setup 1821
large deformations 26 shaft excavation example 228
level 19, 27, 83, 84, 120 input factoring 168171, 173174
limit states 165166 in situ testing 7285
model rigour 27 concepts 7273
pore pressure terms 106108 at plate dilatometers 74, 82
pressure measurement 6264 gravelly ground 9495
rates of loading 109113 highly variable ground 9495
sample disturbance 6061 parameters obtainable 74
shaft excavation example 239241 permeability 74, 8385
sources/sinks 120 piezocone penetration 74, 8182
steady-state ows 19, 106107, 118, 239241 plate load tests 74, 82, 93, 94
total stress analysis 115 pressuremeters 7380, 75
transient ows 118 raft and pile example 206213
unconned ows 118119 rock characterisation 58, 74, 8082, 8485
undrained analysis 113117, 231235, 239 seismic testing 74, 8081
see also pore water pressure soil characterisation 5657
grouting 2324, 140141, 152155 standard penetration tests 74, 82
GSI see Geological Strength Index installation effects 2224
installation of piezometers 6364
hexahedral elements 17 interface elements 132136, 204, 227, 236
hierarchy of element types 17 interface friction 130, 133
highly variable ground testing 9495 interface stresses 132136
high-pressure dilatometers (HPD) 76 intermediate principal stress 33, 8788
historical information requirements 3 intrusive investigation 34

267

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

isotropic consolidation 69 linearly elastic, perfectly plastic (LEPP)


isotropic hardening 40, 4445 models 20, 48, 194196, 247
isotropic linear elastic perfectly plastic models 48 linear strain elements 17
isotropic softening 40 linear structures
iterations, parameter setup 24 2D plane strain models 9
modelling elements 17, 126127, 132
Jakys equation 97 linings of tunnels 143
load (effect) and resistance factoring (LRFA)
K0 see stress ratio 171175
key outputs loadings
aims 2 coupled consolidation analysis 122123
presentation 57 existing structures 4
see also outputs see also construction stages; initial stresses;
kinematic hardening 40, 4546 structures
Kirchhoff theory 129 load reduction method, tunnels 142
known solution comparisons 189191 local strain measurement 6466
Kondner hyperbolic stressstrain relationship Lodes angle 41
211, 212 low-permeability soils 107, 112, 120122
LRFA see load (effect) and resistance
laboratory testing 5758, 6472 factoring
planning 5657 Lugeon test 83, 84
direct shear measurement 72
oedometers 72, 9293 managing errors 192197
parameters obtainable 7071 Marchetti dilatometers 82
simulations 9295 material factoring approach (MFA) 168171,
triaxial cells 6469 173174
Lade model 39, 46 materials, modelling 152156, 181
Lagrangian formulations 26 Modied Cam Clay (MCC) models 39, 44, 47,
landll 57 247248, 253255
large deformations 2526 mean effective stress, soils 41
LEPP see linearly elastic, perfectly plastic membrane elements 126, 128
models Menard pressuremeters (MPM) 76
level of geometric detail requirements 9, 1112 meshing 1718, 2526
level of groundwater 19, 27, 83, 84, 120 embankment construction example 246
limit equilibrium methods 190 raft and pile foundation example 202205
limit pressures 73, 75 shaft excavation example 227228
limit states MFA see material factoring approach
accuracy 164166 Mindlin theory 129
dual factoring 172175 MIT-E3 model 47
input factoring 168171, 173174 modelling
output factoring 169, 170, 171175 2D axisymmetric 9, 11, 143-144, 146-148
rock 180 boundaries 1216
serviceability 163, 164, 167168 choice of elements 130
strength reduction 174177, 179, 238239 connections 150151
ultimate 163, 164, 168181 constitutive model, choice in setup 26
linear elasticity construction stages 2124
concrete and grouting 152154 continuum elements 127, 130136, 150151
soils 38 distributed loads 152
steel structures 155156 dual factoring 172175
structural limit states 180181 earth/soil walls 137138

268

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

embankment construction example 243261 numerical analysis methods 191


xities 1516, 202, 203204, 226, 246, 247
geometrical anisotropy 148150 observational method 6, 196197
geometry 125152 oedometers 72, 9293
geosynthetics 137138 open hydraulic boundaries 119
ground improvement 22, 140141 open piezometers 6263
groundstructure interactions 130136, output factoring 169, 170, 171175, 178,
156160 237238
groundwater parameters 27 outputs
initial stresses 1821 accuracy assessment 188191
input factoring 168171, 173174 aims 2
installation effects 2224 element readouts 132
iterations 24 embankment construction 255259
materials 152156 presentation 57
noncontinuum elements 125134, 150151 raft and pile example 213220
output factoring 169, 170, 171175, 178 shaft excavation example 236241
parametric studies 192196 sources of error 185188
partial factor input 173177 over-consolidation
piles 140 clays 116117, 225243
plane strain methods 710, 143146 excess pore pressure 112113, 116117
raft with piles example 199224 stress ratio 97
reliability testing 166167
retaining wall supports 136137, 137138 packer test 83, 84
rigour 185187 parameter acquisition 55103
sensitivity analysis 167, 192 accuracy 8587, 9198, 186
shaft excavation example 225243 anisotropy 8889
singularities 152 case studies 96
soil tests 9293 characteristic values 8687, 91
spread foundations 138139 common laboratory tests 7071
strength reduction 174177, 179, 238239 databases 96
structural limit states 180181 derivation 8591
submerged surfaces 27 dilation angle 89
threshold boundary positions 14 direct shear test 72
tunnels 141143 drained parameters from undrained tests 90
MohrCoulomb failure criterion 43, 4748, embankment example 252255
177, 179 errors 186
monitoring data, case studies 96 groundwater pressure 6264
MPM see Menard pressuremeters initial state parameters 8990
multipoint piezometers 64 intermediate principal stress 8788
multi-surface plasticity models 4546 laboratory tests 6472
oedometer tests 72
NGI-ADP model 47 other sources 9496
no-ow boundaries 119 permeability 74, 8385, 253255
non-associated ow 42 planning 5558
non-continuum elements 125134, 150151, plausibility checks 93
202204 Poissons ratio derivation 9091
non-equilibrium pore pressure 107108 pressuremeter tests 7380, 75
non-linear elasticity of soils 3839 axisymmetric simulations 9395
non-linear models of concrete 154155 clay 74, 7779
normally consolidated soils, stress ratio 97 cone-type 77

269

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

parameter acquisition (continued ) permeable boundaries 119


pre-bored 76 piezocone penetration tests (CPTu) 74,
rock 74, 80 8182
sands 74, 7980 piles
self-boring 7677 constitutive models 50
quality analysis of samples 6162 installation effects 2223
raft and pile example 206213 modelling 140, 155
sampling 5962 output factoring 171172
shaft excavation example 231236 raft foundation example 199224
site characterisation 5558 setup 140, 155, 200206, 201, 203204
site-specic empirical relationships 9596 pinned connections 150151
in situ tests 7285 plane strain models 710, 143146
concepts 7273 planning, parameter acquisition 5558
at plate dilatometers 74, 82 plasticity
obtainable parameters 74 idealised behaviours 40
permeability 74, 8385 materials 155156, 181
piezocone penetration 74, 8182 soils
plate load tests 74, 82, 93, 94 anisotropic strength 4647
pressuremeters 7380 constitutive models 3947
seismic testing 74, 8081 creep 34, 47, 68, 75
standard penetration tests 74, 82 destructuration 46
test simulations 9295 failure surfaces 4243
triaxial cells 6471, 231235 ow rules 42
anisotropic consolidation 69 hypoplasticity 47
isotropic consolidation 69 stress-dependent strength 46
permeability 69 types 40, 4446
reconsolidation 6668, 231234 yield surfaces 3942
shear stage 68 steel structures 155156
stiffness measurement 6869 plate elements, properties 127, 128130
stress paths 6466 plate load tests (PLT) 74, 82, 93, 94
useful equations 9799 plausibility checks, parameter acquisition 93
validation 9198 PLT see plate load tests
parametric studies 167 Poissons ratio, derivation 9091
partial factors pore pressure
dual approach 172175 consolidation 107111, 120123
input factoring 168171, 173174 embankments 112
materials 181 excavations 107, 112
models 173177 groundwater simulation 120
output factoring 169, 170, 171175, 178 initial stresses 19
undrained shear strength 178179 large deformations 26
values 175 limit states 166
partially drained conditions 105106, measurement 6264
121122 parameters 27
PBP see pre-bored pressuremeters terms 106108
perfect plasticity 40, 44 see also groundwater
permeability pre-bored pressuremeters (PBP) 76
elements 132 precipitation 120
in situ testing 74, 8385 pre-consolidation stress derivation 8990
triaxial cell tests 69, 253255 prescribed heads 120
value derivation 85 presentation of key outputs 57

270

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

pressuremeter tests (PMT) 7380, 75 site characterisation 58


axisymmetric simulations 93, 95 standard penetration tests 82
clay 74, 7779 triaxial cell tests 64
cone-type 77 ultimate limit states 180
pre-bored 76 roller connections 150151
raft and pile example 206213 rotary coring 60, 61
rock 74, 80 rounded sands, shear strength 98
sands 74, 7980
self-boring 7677 samples
pressure wave velocity 81 acquisition 5962
principal stresses 33, 3943 disturbance 6061
probabilistic reliability testing 167 quality evaluation 6162
proposed structures, requirements 4 testing 6485
pumping tests 83, 84 sands
at plate dilatometer tests 74, 82
quadrilateral elements 17 Kondner hyperbolic stressstrain
quality evaluation, samples 6162 relationship 211, 212
permeability tests 74, 84
raft foundations with settlement-reducing piles piezocone penetration test 74, 8182
example 199224 plate load tests 74
constitutive model selection 206 pressuremeter tests 7980
outputs 213220 raft with pile example 199224
parameter acquisition 206213 seismic testing 74, 8081
setup 200206, 201, 203204 shear strength 82, 98
validation 220224 standard penetration tests 82
rafts, modelling 155 saturation modelling 105106
rates of loading 109113 SBP see self-boring pressuremeters
reconsolidation 6668, 231234 SCL see sprayed concrete lining
reinforcement bars 154 SCPT see seismic cone penetration tests
reliability testing 166167 secondary compression tests 68
resistance calculations 177178 see also creep
responsibility for model accuracy 183184 seepage boundaries 119
retaining walls seismic cone penetration tests (SCPT) 81
constitutive models 49 seismic testing 74, 8081
input factoring 168170 selection of constitutive models 2953
output factoring 171172 self-boring permeameters 83, 8485
supports 136137, 137138 self-boring pressuremeters (SBP) 7677,
surface settlement 190191 206213
rigour of models 185187 self-weight, modelling elements 131, 133
rising head tests 83 sensitivity analysis 167, 192
robustness 185186 serviceability limit states (SLS) 163, 164,
rocks 167168
constitutive models 3537, 4748 settlement, test procedures 57
direct shear tests 72 settlement-reducing piles with raft foundations
discontinuity modelling 3536, 37, 4748 199224
MohrCoulomb failure criterion 4748 constitutive model selection 206
permeability 74, 84, 85 outputs 213220
plate load tests 74, 82 parameter acquisition 206213
pressuremeter tests 80 setup 200206, 201, 203204
seismic tests 81 validation 220224

271

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

setup 128 detail needed 9, 1112, 13


analysis planning 17 embankment construction example
analysis stages 1826 244247, 246
boundaries 1216 errors 187
calculation options 2426 raft foundation with piles example
constitutive model choice 26 200202
construction stages 2124 shaft excavation example 225226
embankment construction 243252, simulations see modelling
245246 single surface plasticity 44
xities 1516, 202, 203204, 226, 246, 247 singularities, denition 152
geometry 716 sinks, groundwater 120
ground improvement 22 site characterisation 5558, 9495
groundwater parameters 27 site monitoring data 191, 196197
initial stresses 1821 site-specic empirical relationships 9596
installation effects 2224 slabs 24, 51
large deformations 2526 slip elements 132136
meshing 1718, 2526 slopes 49
raft foundation with piles 200206, 201, SLS see serviceability limit states
203204 smeared crack approach 154
shaft excavation 225231, 226227 soft soils, full-ow penetrometers 8182
submerged surfaces 27 software packages 45, 184
shaft excavation, example procedure 225243 soils
model setup 225231, 226227 anisotropy 33, 34, 38, 4647, 65
outputs 236241 bonding and structure 3233
parameter acquisition 231236 bulk modulus 115
validation 241243 constitutive models 3134, 3647
shear hardening 45 creep 34, 47, 68, 75
see also friction hardening databases 96
shear stage, triaxial cells 68 destructuration 46, 60
shear strength deviatoric stress 41, 112113
changes 31 elasticity 3639, 98
clays 98 failure surfaces 4243
direct tests 72 hypoplasticity 47
full-ow penetrometers 8182 intermediate principal stress 33
interfaces 135 isotropic hardening 40, 4445
partial factors 178179 isotropic softening 40
pressuremeter tests 75 kinematic hardening 40, 4546
raft and pile example 206208 linear elasticity 38
rocks 64, 72 mean effective stress 41
sands 82, 98 non-linear elasticity 3839
standard penetration tests 82 perfect plasticity 40, 44
undrained 116117, 178179 permeability 69, 74, 8385, 132
shear wave velocity 62, 65 plasticity 34, 3947, 68, 75
sheet pile walls 156, 237, 238 sampling 5962
shell elements 127, 128130, 202204, 235 seismic tests 8081
Sichardts empirical formula 15 site characterisation 5557
3s rule 193194 stiffness 32, 3839, 65, 6869, 9798, 176
simple connections 150151 strain-dependency 3334, 39, 209210,
simplication 234235
geometry strength 3132, 4647

272

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

stress-dependency 32, 39, 211, 234 strain-dependency 3334, 39, 209210,


stress-path dependency 32, 3839 234235
structure modelling 3233 strain hardening 40, 4445
test simulations 9293 strain softening 40
triaxial cell tests 6471 stratum boundaries 12
undrained shear strength 116117, 178179, strength
206208 rocks 35, 4748
yield surfaces 3942 soils 3132, 4647
see also individual soil types . . . strength reduction 174177, 179, 181
soilstructure interaction embankment construction example 248252,
modelling 130136, 156160 257259
strength reduction 174177, 179, 238239 shaft excavation example 238239
structural limit states 180181 stress-dependency
ultimate limit states 163, 164, 168181 rocks 35, 4748, 58
reinforced soil walls, modelling 137138 soils 32, 39, 46, 211, 234
solution schemes 2425 stress eld methods 190
sources stress paths
errors 185188 stiffness 32, 3839
groundwater 120 test procedures 57
sprayed concrete lining (SCL) tunnels triaxial cell tests 57, 6466, 253254
142143, 155 stress ratio (K0 ) 1921, 44, 8990, 97, 164165
spread foundations structural connections 150151
bearing resistance 178 structural limit states 180181
constitutive models 51 structures 125161
input factoring 168170 axisymmetric models 143144, 146148
modelling 138139, 139 connections 150151
output factoring 171172 distributed load modelling 152
spring elements 125, 126 elements 125136
SPT see standard penetration test geometrical anisotropy 148150
standard penetration test (SPT) 74, 82 geometry 125152
steady-state pore pressure 19, 106107, 118, ground interactions 130136, 156160
239241 materials 152156
steel structures 155156 plane strain models 710, 143146
step-by-step method, tunnel models 142 subgrade reaction 159160, 220224
step size, parameter setup 24 submerged surfaces 27
stiffness suction measurement 6162
factoring 176 supports to retaining walls 136137, 137138
parameter validation 9798
plate load tests 82 tensile strength
in situ seismic testing 74, 8081 rocks 35, 4748
strain-dependency 3334, 39, 209210, soils 3132
234235 Terzaghis one-dimensional consolidation
stress-dependency 32, 39, 211, 234 theory 259261
stress-path dependency 32, 3839 tetrahedral elements 17, 202
structural connections 151 threshold position boundary effects 14
triaxial cell tests 6566, 6869 Timoshenko theory 128
stochastic approaches, reliability testing 167 tolerances 186
strain total pore pressure 107, 108, 120
2D modelling 711, 143146 total stress 115, 179
measurement 6468 transient groundwater ows 118

273

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.
Index

transient pore pressure 107, 108 methodology 113117


Tresca failure surface 43 parameters 27
triangular elements 17, 202, 203204, 227 shaft excavation example 239241
triaxial cells 6471, 65 undrained shear strength 116117, 178179,
anisotropy 3334, 6869 231235, 239
with bender elements 6466
extension tests 3334, 231235 validation
isotropic consolidation 69 acquired parameters 9198
permeability 69, 253255 concepts 188
principal stress 33 embankment construction example 259261
reconsolidation stage 6668 output comparisons 188191
rock testing methods 58 parametric studies 192196
shear stage 68 raft and pile example 220224
simulations 9293, 232233 sensitivity analysis 167, 192
tube sampling 5961 shaft excavation example 241243
tunnel boring machines (TBM) 141143 values, partial factors 175
tunnels 8, 49, 141143, 155 variable head tests 83
tweaks 185 verication 183184, 188191
see also validation
ultimate limit states (ULS) 163, 164, 168181 vertical boundary xities 1516
dual factoring 172175 vertical effective stress 19
input factoring 168171, 173174 void ratio changes 61, 85
output factoring 169, 170, 171175, 178 volume loss control method 142
rock 180 von Mises failure surface 43
strength reduction 174177, 179, 238239
unconned groundwater ows 118119 water see groundwater
undrained, denition 109 wireline drilling 60
undrained analysis wished in place 23
assumptions 109113
concepts 109 yield limits 181
drained parameter derivation 90 yield surfaces 3942
limit states 165166 Youngs modulus 125

274

Downloaded by [ University College London] on [16/11/16]. Copyright ICE Publishing, all rights reserved.

S-ar putea să vă placă și