Sie sind auf Seite 1von 96

COMPUTING AND COMPUTER MODELLING IN

GEOTECHNICAL ENGINEERING
J.P. Carter
1
, C.S. Desai
2
, D.M. Potts
3
, H.F. Schweiger
4
and S.W. Sloan
5
1.0 ABSTRACT
A broad review is presented of the role of computing in geotechnical engineering. Included in the
discussions are the conventional deterministic techniques for numerical modelling, stochastic techniques for
dealing with uncertainty, soft-computing tools, as well as modern database software for geotechnical
applications. Considerable emphasis is given to the methods commonly used for the solution of boundary
and initial value problems. Constitutive modelling of soil and rock mass behaviour and material interfaces is
an essential component of this type of computing, and so a review of recent developments and capabilities of
constitutive models is also included. The importance of validating computer simulations and geotechnical
software is emphasised, and some methodologies for achieving this are suggested. A description of several
previously conducted validation studies is included. The paper also includes discussion of the limitations of
various numerical modelling techniques and some of the more notable pitfalls. The concepts described in
the paper are illustrated with examples taken from research and practice. In presenting these concepts and
examples, emphasis has been placed on the behaviour of soil, but it is noted that many of the models and
techniques described also have application in rock engineering.
2.0 INTRODUCTION
The desire to understand the physical world and to be able to describe it using mathematical concepts and
numbers has long been a goal of scientists and engineers. This desire has been evident since at least the time
of Pythagoras. The discipline of geotechnical engineering is no exception, as first researchers and now
practitioners routinely make use of mathematical models and computer technology in their day-to-day work,
trying to understand and predict their world of geomaterials. For example, simple numerical procedures
have been used for many years in geotechnical practice in the assessment of strength, analysis of soil
consolidation, and estimation of slope stability. With the introduction of electronic computers after World
War II came the opportunity for engineers to make much more use of numerical procedures to solve the
equations governing their practical problems. This new computing power has made possible the solution of
quite complicated non-linear, time-dependent problems, boundary and initial value problems that were once
too tedious and intractable using hand methods of calculation.
The ready availability of desktop and portable computers has meant that these numerical tools for the
solution of boundary and initial value problems are no longer the preserve of academic and research
engineers. With the spectacular improvements in hardware have come major developments in software, and
very sophisticated packages for solving geotechnical problems are now available commercially. This
includes a wide range of software for solving problems from the more routine type, such as limiting
equilibrium calculations, to the most powerful non-linear finite element analyses.
The availability of this powerful hardware and sophisticated geotechnical software has allowed
geotechnical engineers to examine many problems in much greater depth than was previously possible. In

1
Challis Professor, Department of Civil Engineering, The University of Sydney, Sydney, NSW, AUSTRALIA
2
Regents Professor, Department of Civil Engineering and Engineering Mechanics, University of Arizona, Tucson,
Arizona, USA
3
Professor of Analytical Soil Mechanics, Department of Civil and Environmental Engineering, Imperial College of
Science, Technology and Medicine, London, UK
4
Associate Professor, Institute for Soil Mechanics and Foundation Engineering, Graz University of Technology,
Graz, AUSTRIA
5
Professor of Civil Engineering, Department of Civil, Surveying and Environmental Engineering, University of
Newcastle, Newcastle, NSW, AUSTRALIA
particular, they have allowed the possibility of using numerical methods to examine the important
mechanisms that control the overall behaviour in many problems. Further, they can be used to identify the
key parameters in any problem, thus indicating areas that require more detailed and thorough investigation.
These developments have also meant that generally better quality field and laboratory data are required as
inputs to the various models. However, often precise values for some of the input data for these numerical
models will not be known, but knowledge of which parameters are most important allows judgements to be
made about what additional information should be collected, and where additional resources are best
directed in any particular problem. Furthermore, knowledge of the key model parameters together with the
results of parametric investigation of a problem can often allow engineering judgements to be made with
some confidence about the consequences of a particular decision, rather than in complete or partial
ignorance of them. Increasingly, the tools for deterministic modelling of geotechnical problems are being
coupled with statistical techniques, to provide a means of dealing with uncertainties in the key problem
parameters, and of associating probabilities with the predicted response.
The application of computer technology in geotechnical engineering has not been confined to the area of
analysis and mathematical modelling. The availability of sophisticated information technology tools has
also had noticeable impact on the way geotechnical engineers record, store, retrieve, process, visualise and
display important geotechnical data. These tools range from the ubiquitous spreadsheet packages to
database software and high-end graphics visualisation tools.
This paper provides a broad review of the role of computing and computer technology in geotechnical
engineering. A major part of the review includes discussion of some of the more common analytical and
numerical modelling techniques available to geotechnical engineers, particularly those used today to obtain
solutions on personal computers. The use of the numerical methods and the associated software is illustrated
by applications taken from geotechnical research and practice. Throughout this discussion, emphasis is
given to the role of these deterministic techniques in identifying mechanisms and providing the user with
explanations of observed behaviour, as well as the consequences of planned actions and proposed
construction activities. The examples considered involve analysis of strength and deformation of soil, and
include pre-failure behaviour as well as the development of failure mechanisms within soil bodies. The need
to validate and calibrate numerical models, particularly those now being used commonly in practice is
emphasised, and some examples of validation studies are described. Some of the potential pitfalls of
numerical analysis are also described.
Included in the paper is a brief review of the stochastic and soft computing tools that are increasingly
being applied in practice to model geotechnical problems. In particular, an application of the artificial neural
network technique is described. The role of database packages in geotechnical practice is also illustrated by
a recent application on a large-scale construction project.
3.0 GEOTECHNICAL DESIGN
Traditionally, geotechnical design has been carried out using simple analysis or empirical approaches.
The ready availability of inexpensive, but sophisticated, computer hardware and software has lead to
considerable advances in analysis and design of geotechnical structures, with much progress made recently
in the application of the modelling techniques to geotechnical structures.
In common with other branches of engineering, the design objectives in geotechnics may be identified as
follows:
Local stability of the structure and its support system as well as overall stability should be ensured.
The induced movements must be tolerable, not only for the structure being designed but also for any
neighbouring structures and services.
The design process usually consists of some form of assessment of these important aspects of ground
behaviour, and often this will include calculations to provide estimates of stability and the deformations
resulting from the proposed works. Analysis therefore provides the mathematical framework for these
calculations, and almost invariably the analytical tools are embodied in computer software allowing
numerical analysis to be efficiently and conveniently executed.
However, it is important to recognise that geotechnical design involves much more than just analysis, and
it often includes data gathering prior to analysis, as well as observation and monitoring during and following
construction. Construction issues may also be significant and should be included in the decision making
process that constitutes geotechnical design.
The analytical tools used most often in the geotechnical design process are still those based on
deterministic analysis. It is therefore appropriate to begin a review of computing in geotechnical
engineering by discussing the various deterministic analysis methods. However, as there is growing
development and use of computing tools that can deal with imprecise information, e.g., stochastic methods,
neural networks and fuzzy logic, these will also be discussed later in this paper. As already mentioned, data
gathering and processing is also important in geotechnical design, and so computer tools to assist with these
tasks are also described.
4.0 DETERMINISTIC GEOTECHNICAL ANALYSIS
Usually, geotechnical analysis involves the solution of a boundary value or initial value problem and
most often this is achieved by some form of numerical solution procedure. In all geotechnical applications
involving numerical analysis it is essential, for economic computer processing and to obtain a reliable
numerical solution, to provide a good model of the physical problem. In finite difference and finite element
analysis, for example, this normally involves at least two distinct phases, i.e.,
idealisation, and
discretisation.
Idealisation is achieved by breaking down the physical problem into its component parts, e.g., continuum
components such as elastic regions, and discrete structural components such as beams, columns and plates.
At this stage of the modelling
process, reliable knowledge of
the site geology is of paramount
importance. In addition, various
constitutive models that will be
employed in the analysis must be
determined at this stage. The
final subdivision of the problem
domain should be only as
detailed as is necessary for the
purpose at hand. Too much
detail only clutters the analysis
and may cloud important aspects
of the behaviour. The choice of
an adequate level of detail is a
matter for experience and
judgement.
In a finite element procedure,
for example, the idealised model
is then further subdivided or
discretised using an appropriate
subdivision of elements. The
aim here is usually to satisfy the
governing equations of the
problem separately within each
element. Details, such as node and element numbers, will be assigned as part of this subdivision process.
Each of these phases is illustrated schematically in Figure 1.
As indicated previously, most geotechnical analysis involves an assessment of stability and deformation.
Once the problem has been idealised, there are four fundamental conditions that should then be satisfied by
the solution of the boundary or initial value problem. These are:
equilibrium,
compatibility,
constitutive behaviour, and
boundary and initial conditions.
Unless all four conditions are satisfied (either exactly or approximately), the solution of the ideal problem is
not rigorous in the mathematical sense.
Material 1
Material 2
Fixed base
Smooth, rigid
boundary
Fill
Sand
Clay
Rock
Fill
Sand
Clay
Rock
(a) Physical problem
(b) Idealisation
Nodal points
Elements
L
c
(c) Discretisation
Figure 1 : Steps in the modelling process
5.0 METHODS OF ANALYSIS
Some of the most common methods of analysis used in geotechnical engineering to solve boundary value
problems are listed in Table 1. Included are numerical methods as well as some more traditional techniques
that may be amenable to hand calculation. The numerical methods may be classified as follows:
the finite difference method (FDM),
the finite element method (FEM),
the boundary element method (BEM), and
the discrete element method (DEM).
Table 1 also provides an indication of whether the four basic requirements of a mathematically rigorous
solution are satisfied by each of these techniques. It is clear that only the elastoplastic analyses are capable
of providing a complete solution while also satisfying (sometimes approximately) all four solution
requirements. The difficulty of obtaining closed-form elastoplastic solutions for practical problems means
that numerical methods are the only generally applicable techniques.
Detailed descriptions of each of the numerical methods listed in Table 1 may be found in a large number
of textbooks (e.g., Zienkiewicz, 1967; Desai and Abel, 1972; Britto and Gunn, 1987; Smith and Griffiths,
1988; Beer and Watson, 1992; Potts and Zdravkovic, 1999, 2000), so there is no need to duplicate such
detail here. However, it is probably worth noting that the FDM, FEM and DEM methods consider the entire
region under investigation, breaking it up, or discretising it, into a finite number of sub-regions or elements.
The governing equations of the problem are applied separately and approximately within each of these
elements, translating the governing differential equations into matrix equations for each element.
Table 1. Summary of common analysis methods
Method of Analysis
Limit Bound Theorems Elastic Elastoplastic Analysis
Equilibrium Lower Upper Analysis Closed-form Numerical
Equilibrium
Overall
Locally "
"
Compatibility
" "

(1)
Boundary
Conditions
Force only Force only Displacement
only

Constitutive
Model
Failure criterion Perfectly rigid plasticity Elastic Elastoplastic Any
(2)
Collapse
Information
"
Information
before Collapse
" " "
Comment
Simple
Safe or unsafe?
Safe estimate
of collapse
Unsafe estimate
of collapse
Closed form
solutions
available
Complicated Powerful
computer
techniques
Examples
Slip circle,
Wedge
Methods
- - Many Limited FDM, FEM,
BEM and
DEM
(1) Inherent and induced material discontinuities can be simulated.
(2) Includes perfect plasticity and models that can allow for complicated behaviour such as
discontinuous deformations, degradation (softening), and non-local effects.
Compatibility, equilibrium and the boundary conditions are enforced at the interfaces between elements and
at the boundaries of the problem.
On the other hand, in the BEM only the boundary of the body under consideration is discretised, thus
providing a computational efficiency by reducing the dimensions of the problem by one. The BEM is
particularly suited to linear problems. For this reason, and because it is well suited to modelling infinite or
semi-infinite domains, the BEM is sometimes combined with the finite element technique. In this case, a
problem involving non-linear behaviour in part of the infinite domain can be efficiently modelled by using
finite elements to represent that part of the domain in which non-linear behaviour is likely, while also
modelling accurately the infinite region by using boundary elements to represent the far field.
For all methods, an approximate set of matrix equations may be assembled for all elements in the region
considered. This usually requires storage of large systems of matrix equations, and the technique is known
as the implicit solution method. An alternative method, known as the explicit solution scheme, is also
employed in some software packages. This method usually involves solution of the full dynamic equations
of motion, even for problems that are essentially static or quasi-static. The explicit methods do not require
the storage of large systems of equations, but they are known to present difficulties in determining reliable
solutions to some problems in statics.
In the following, brief details of each of the main methods of numerical analysis are presented together
with a summary of their main advantages and disadvantages, and their suitability for various geotechnical
problems. Much of this discussion is based on suggestions proposed by Schweiger and Beer (1996).
5.1 Finite Element Method
The finite element method is still the most widely used and probably the most versatile method for
analysing boundary value problems in geotechnical engineering. The main advantages and disadvantages
for geotechnical analysis may be summarized as follows.
5.1.1 Advantages
nonlinear material behaviour can be considered for the entire domain analysed.
modelling of excavation sequences including the installation of reinforcement and structural support
systems is possible.
structural features in the soil or rock mass, such as closely spaced parallel sets of joints or fissures, can
be efficiently modelled, e.g., by applying a suitable homogenisation technique.
time-dependent material behaviour may be introduced.
the equation system is symmetric (except for non-associated flow rules in elasto-plastic problems using
tangent stiffness methods).
the conventional displacement formulation may be used for most load-path analyses.
special formulations are now available for other types of geotechnical problem, e.g., seepage analysis,
and the bound theorem solutions in plasticity theory.
the method has been extensively applied to solve practical problems and thus a lot of experience is
already available.
5.1.2 Disadvantages
The following disadvantages are particularly pronounced for 3-D analyses and are less relevant for 2-D
models.
the entire volume of the domain analysed has to be discretised, i.e., large pre- and post-processing
efforts are required.
due to large equation systems, run times and disk storage requirements may be excessive (depending on
the general structure and the implemented algorithms of the finite element code).
sophisticated algorithms are needed for strain hardening and softening constitutive models.
the method is generally not suitable for highly jointed rocks or highly fissured soils when these defects
are randomly distributed and dominate the mechanical behaviour.
5.2 Boundary Element Method
Significant advances have been made in the development of the boundary element method and as a
consequence this technique provides an alternative to the finite element method under certain circumstances,
particularly for some problems in rock engineering (Beer and Watson, 1992). The main advantages and
disadvantages may be summarized as follows.
5.2.1 Advantages
pre- and post-processing efforts are reduced by an order of magnitude (as a result of surface
discretisation rather than volume discretisation).
the surface discretisation leads to smaller equation systems and less disk storage requirements, thus
computation time is generally decreased.
distinct structural features such as faults and interfaces located in arbitrary positions can be modelled
very efficiently, and the nonlinear behaviour of the fault can be readily included in the analysis (e.g.,
Beer, 1995).
5.2.2 Disadvantages
except for interfaces and discontinuities, only elastic material behaviour can be considered with surface
discretisation.
in general, non-symmetric and often fully-populated equation systems are obtained.
a detailed modelling of excavation sequences and support measures is practically impossible.
the standard formulation is not suitable for highly jointed rocks when the joints are randomly distributed.
the method has only been used for solving a limited class of problems, e.g., tunnelling problems, and
thus less experience is available than with finite element models.
5.3 Coupled Finite Element - Boundary Element Method
It follows from the arguments given above that it should be possible to minimise the respective
disadvantages of both methods by combining them. This is in fact true and very efficient numerical models
can be obtained by discretising the soil or rock around the region of particular interest, e.g., representing the
region around a tunnel by finite elements and the far field by boundary elements (e.g., Beer and Watson,
1992; Carter and Xiao, 1993). Two disadvantages however remain, namely the cumbersome modelling of
major discontinuities intercepting the region of interest in an arbitrary direction, e.g., a tunnel axis, and the
non-symmetric equation system that is generated by the combined model. The latter problem may be
resolved by applying the principle of minimum potential energy for establishing the stiffness matrix of the
boundary element region (Beer and Watson, 1992). If this is done, then after assembling with the finite
element stiffness matrix, the resulting equation system remains symmetric.
5.4 Explicit Finite Difference Method
The finite difference method does not have a long-standing tradition in geotechnical engineering, perhaps
with the exception of analysing flow problems including those involving consolidation and contaminant
transport. However, with the development of the finite difference code FLAC (Cundall and Board, 1988),
which is based on an explicit time marching scheme using the full dynamic equations of motion, even for
static problems, an attractive alternative to the finite element method was introduced. Any disturbance of
equilibrium is propagated at a material dependent rate. This scheme is conditionally stable and small time
steps must be used to prevent propagation of information beyond neighbouring calculation points within one
time step. Artificial nodal damping is introduced for solving static problems in FLAC. The method is
comparable to the finite element method (using constant strain triangles) and therefore some of the
arguments listed above basically hold for the finite difference method as well. However, due to the explicit
algorithm employed some additional advantages and disadvantages may be identified.
5.4.1 Advantages
the explicit solution method avoids the solution of large sets of equations.
large strain plasticity, strain hardening and softening models and soil-structure interaction are generally
easier to introduce than in finite elements.
the model preparation for simple problems is very easy.
5.4.2 Disadvantages
the method is less efficient for linear or moderately nonlinear problems.
until recently, model preparation for complex 3-D structures has not been particularly efficient because
sophisticated pre-processing tools have not been as readily available, compared to finite element pre-
processors.
because the method is based on Newtons law of motion no converged solution for static problems
exists, as is the case in static finite element analysis. The decision whether or not sufficient time steps
have been performed to obtain a solution (below but close to failure) has to be taken by the user, and this
judgement may not always be easy under certain circumstances, although several checks are possible
(e.g., unbalanced forces, velocity field).
5.5 Discrete Element Method
The methods described so far are based on continuum mechanics principles and are therefore restricted to
problems where the mechanical behaviour is not governed to a large extent by the effects of joints and
cracks. If this is the case discrete element methods are much better suited for numerical solution. These
methods may be characterised as follows:
finite deformations and rotations of discrete blocks (deformable or rigid) are calculated.
blocks that are originally connected may separate during the analysis.
new contacts which develop between blocks due to displacements and rotations are detected
automatically.
Several different approaches to achieve these criteria have been developed, with probably the most
commonly used methods being the discrete element codes UDEC and 3-DEC (Lemos et al., 1985), which
both employ an explicit finite difference scheme, as in the program FLAC.
Due to the different nature of a discontinuum analysis, as compared to continuum techniques, a direct
comparison seems to be not appropriate. The major strength of the distinct element method is certainly the
fact that a large number of irregular joints can be taken into account in a physically rational way. The
drawbacks associated with the technique are that establishing the model, taking into account all relevant
construction stages, is still very time consuming, at least for 3-D analyses. In addition, a lot of experience is
necessary in determining the most appropriate values of input parameters such as joint stiffnesses. These
values are not always available from experiments and specification of inappropriate values for these
p ar amet e rs ma y le ad to c o mp ut at i on al pr ob le ms. I n ad di t io n, ru n ti mes f o r 3- D a na ly s es a re us ua l ly q ui t e hi g h.
5.6 Which Method For Which Problem?
Having discussed the main advantages and disadvantages of the most common numerical methods a
reasonable question is which method should be used for any particular problem. Of course, the answer to
this question will be very problem dependent. In many cases several methods may be appropriate and the
decision on which to use will be made simply on the basis of the experience and familiarity of the analyst
with these techniques.
However, it is possible in certain classes of problem to provide broad guidelines. One example concerns
the analysis of tunnels. This problem area has been addressed by Schweiger and Beer (1996), who
suggested guidelines for tunnelling by considering the separate problems of shallow and deep tunnels. In
making suggestions they discussed the importance of the input parameters and what can be expected from
numerical analyses. For example, they concluded that the finite element and finite difference method are
most suitable for shallow tunnels in soil, whereas the boundary element and discrete element method are
most suitable for deep tunnels in jointed and faulted rocks. They also concluded that the most promising
continuum approach is a combination of finite elements and boundary elements, because the merits of each
method can be fully exploited as appropriate.
In the following sections of this paper some of the methods of numerical analysis described above and
some of their essential components are described further. Their use is illustrated by applications from
geotechnical practice. As indicated in Table 1, not all techniques are designed to provide a solution for the
complete load-deflection behaviour of a soil or rock structure up to the point of collapse. Some address only
the pre-failure behaviour and some the ultimate condition.
The first method to be described in detail involves the use of the bound theorems of classical plasticity
theory together with special finite element formulations to assess stability. These well-known methods for
bracketing the true collapse load do not require a sophisticated constitutive model for the soil or rock mass.
They are based on the assumption of a rigid plastic model for the soil or rock and require only the definition
of a failure criterion and a plastic flow rule. With recent developments in computer technology, linear and
non-linear programming and the finite element method, these classical theorems have a renewed and
important role to play in geotechnical analysis, particularly as they may now be used routinely for three-
dimensional problems. After all, determining when collapse will occur, and avoiding that condition in
practice, is one of the fundamental requirements of geotechnical design.
6.0 LIMIT ANALYSIS USING FINITE ELEMENTS
Stability analysis in geotechnical engineering has traditionally been carried out using either slipline field
or limit equilibrium techniques. Whilst slipline methods have the advantage of being mathematically
rigorous, they are notoriously difficult to apply to problems with complex geometries or complicated
loading. A further shortcoming of these techniques is that the boundary conditions need to be treated
specifically for each problem, thus making it difficult to develop general purpose computer programs which
can analyse a broad range of cases. Despite these limitations, slipline analysis has provided many
fundamental solutions that are used routinely in geotechnical engineering practice (Sokolovskii, 1965).
Limit equilibrium methods, although less rigorous than slipline methods, can be generalised to deal with a
variety of complicated boundary conditions, soil properties and loading conditions. The accuracy of limit
equilibrium solutions is often questioned because of the assumptions that are needed to make the method
work. Nonetheless, this approach is often favoured by practising engineers because of its simplicity and
generality.
Another approach for analysing the stability of geotechnical structures is to use the upper and lower
bound limit theorems developed by Drucker et al. (1952). These theorems can be used to bracket the exact
ultimate load from above and below and are based, respectively, on the notions of a kinematically admissible
velocity field and a statically admissible stress field. A kinematically admissible velocity field is simply a
failure mechanism in which the velocities (displacement increments) satisfy both the flow rule and the
velocity boundary conditions, whilst a statically admissible stress field is one where the stresses satisfy
equilibrium, the stress boundary conditions, and the yield criterion.
The bound theorems assume the material is perfectly plastic and obeys an associated flow rule. The latter
assumption, which implies the strain increments are normal to the yield surface, is often perceived to be a
shortcoming for frictional soils as it predicts excessive dilation upon shear failure. For geotechnical
problems which are not strongly constrained in a kinematic sense (e.g., those with a freely deforming surface
and a semiinfinite domain), the use of an associated flow rule for frictional soils may in fact give good
estimates of the collapse load. This important result is discussed at length by Davis (1969) and has been
confirmed in a number of finite element studies (e.g., Sloan, 1981).
Put simply, the upper bound theorem states that the power expended by the external forces may be
equated to the power dissipated in a kinematically admissible failure mechanism to compute an
unconservative estimate of the true collapse load. In geotechnical engineering, the simplest form of upper
bound calculation is based on a mechanism comprised of rigid blocks where power is dissipated solely at the
interfaces between adjacent blocks. Once a kinematically admissible mechanism has been formulated, the
best upper bound is found by optimising the geometry of the blocks to yield the minimum dissipated power
and, hence, the corresponding collapse load. This type of calculation is very useful for undrained stability
analysis of clays where the soil can be modelled using a Tresca yield condition and deformation occurs at
constant volume. For drained loading, however, where the soil is often assumed to obey a MohrCoulomb
failure criterion, this type of computation is more difficult because of the dilation that accompanies plastic
shearing along the discontinuities.
The lower bound theorem states that the collapse load obtained from any statically admissible stress field
will underestimate the true collapse load. Generally speaking, the upper bound theorem is applied more
frequently than the lower bound theorem to predict soil behaviour, since it is usually easier to construct a
good kinematically admissible failure mechanism than it is to construct a good statically admissible stress
field. Although an upper bound solution often gives a useful estimate of the ultimate load, a lower bound
solution is more desirable in engineering practice as it results in a safe design.
The bound theorems are especially powerful when both types of solution can be computed so that the
actual collapse load can be bracketed from above and below. This feature is invaluable when an exact
solution cannot be determined, since it provides a builtin error check on the accuracy of the approximate
collapse load.
6.1 Lower Bound Limit Analysis Formulations
The use of finite elements and linear programming to compute rigorous lower bounds for
twodimensional stability problems appears to have been first proposed by Lysmer (1970). Lysmers
formulation is based on a simple threenoded triangular element with the nodal normal and shear stresses
being taken as the problem variables. Following previous studies, the linearised yield condition is obtained
by adopting an internal polyhedral approximation to the parent yield surface, so that each nonlinear
inequality is replaced by a series of linear inequalities. By assuming a linear approximation for the stress
field inside each element, it can be guaranteed that this yield condition is satisfied throughout the discretised
region. The presence of statically admissible discontinuities, which are permitted between adjacent
elements, greatly improves the accuracy of the final results. Application of the stressboundary conditions,
equilibrium equations, and linearised yield criterion generates the linear constraints on the stress field, while
the objective function, which is maximised, corresponds to the collapse load. In Lysmers original
formulation, the optimal solution to the linear programming problem, and hence the statically admissible
stress field, was isolated by using the simplex algorithm.
Though Lysmers method represented a significant advance, it had a number of shortcomings. The first
of these resulted from the choice of problem variables which, although ingenious, led to a poorly
conditioned constraint matrix. The second limitation was one of computational efficiency. With the
simplex solution technique used by Lysmer, the analyses had to be restricted to meshes with only a few
elements in order to avoid excessive computer times. This is a serious limitation and probably explains why
the technique did not achieve the prominence that it deserved. The third and final shortcoming of the
method was its inability to generate a complete stress field for a semiinfinite continuum. Semiinfinite soil
masses arise frequently in geotechnical engineering, and it is necessary to extend the stress field throughout
the entire domain for the solution to be classed as a rigorous lower bound. In Lysmers method, the solution
obtained is an incomplete one, as the resultant stress field is statically admissible only in the region limited
by the boundaries of the finite element mesh.
Following Lysmer, other investigators, including Anderheggen and Knopfel (1972), Pastor (1978), and
Bottero et al. (1980), proposed alternative twodimensional lower bound techniques which are based on the
linear programming method. These studies led to a number of key improvements, such as the development
of extension elements for analysing semiinfinite media and the use of cartesian stresses as problem
variables to simplify the formulation. In 1982, Pastor and Turgeman generalised their lower bound
technique to deal with the important case of axisymmetric loading.
Although potentially powerful, these early lower bound formulations were restricted by the performance
of their underlying linear programming (LP) solvers. Because the techniques often lead to very large
optimisation problems, their evolution has been linked closely with the development of efficient LP
algorithms. Indeed, special features of the lower bound formulation, such as the extreme sparsity of the
overall constraint matrix, must be exploited fully to avoid excessive computation times. In the late eighties,
Sloan (1988a) introduced a formulation based on an active set algorithm that permits large twodimensional
problems to be solved efficiently on a PC or workstation. Practical applications of this scheme to date
include tunnels (Sloan et al., 1990; Assadi and Sloan, 1991; Sloan and Assadi, 1991, 1992), foundations
(Utrichton et al., 1998; Merifield et al., 1999) and slopes (Yu et al., 1998).
Despite the success of LP based lower bound formulations for the solution of twodimensional and
axisymmetric problems, these approaches are unsuitable for performing general threedimensional limit
analysis. Although possible, the linearisation process for 3D yield functions inevitably generates huge
numbers of linear inequalities. This in turn, results in unacceptably long solution times for any LP solver
which conducts a vertextovertex search (such as the traditional simplex method or any active set method).
One alternative for finite element formulations of the lower bound theorem is to use nonlinear programming
(NLP) solution methods. Irrespective of the approximation chosen to represent the stress field in the
continuum, this approach assumes that the yield criterion is employed in its original nonlinear form.
Threedimensional stress fields thus present no special difficulty, apart from the extra variables involved.
One such formulation, which uses linear stress finite elements, incorporates nonlinear yield conditions
explicitly, and exploits the underlying convexity of the corresponding optimisation problem, has been
developed recently by Lyamin and Sloan (1997, 2000a) and Lyamin (1999). In this method, the lower
bound solution is found very efficiently by solving the system of nonlinear equations that define the
KuhnTucker optimality conditions. The solver used for this purpose, a variant of the one originally
developed by Zouain et al. (1993) in their study of mixed limit analysis formulations, is a twostage
quasiNewton scheme. After accounting for the nature of the lower bound optimisation problem and
developing a new deflection strategy in the solution phase, the resulting optimisation procedure is many
times faster than an equivalent LP formulation. Indeed, comparisons to date suggest that the new technique
typically offers at least a 50fold reduction in CPU time for large scale twodimensional applications. The
scheme can deal with any (convex) type of yield criterion and permits optimisation with respect to surface as
well as body forces, which can be of unknown distribution. Because of its speed, the NLP formulation of
Lyamin and Sloan (1997, 2000a) is ideally suited to threedimensional applications, where the number of
unknowns is usually very large.
6.2 Upper Bound Limit Analysis Formulations
General formulations of the upper bound theorem, based on finite elements and linear programming,
have been investigated by Anderheggen and Knopfel (1972), Maier et al. (1972) and Bottero et al. (1980).
These methods permit plastic deformation to occur throughout the continuum and inherit all of the
advantages of the finite element technique, but have tended to be computationally cumbersome due to the
large linear programming problems they generate. In the plate studies performed by Anderheggen and
Knopfel (1972), an attempt was made to address this shortcoming by suggesting various solution strategies
based on the revised simplex optimisation algorithm. This type of algorithm was also used by Bottero et al.
(1980), who generalised the method of Anderheggen and Knopfel (1972) to include velocity discontinuities
in plane strain limit analysis. In their formulation, a threenoded constant strain triangular element is used
to model the velocity field, each node has two unknown velocities, and each triangular element is associated
with a specified number of unknown plastic multiplier rates (one for each side of the linearised yield
surface). To be kinematically admissible, the velocities and plastic multiplier rates are subject to a set of
linear constraints arising from the flow rule and the velocities must match the appropriate boundary
conditions. For a given set of prescribed velocities, the finite element formulation works by choosing the set
of velocities and plastic multiplier rates which minimise the dissipated power. This power is then equated to
the power dissipated by the external loads to yield a strict upper bound on the true limit load. To ensure that
the finite element formulation leads to a LP problem, the actual yield surface is linearised by using an
external polyhedral approximation.
Turgeman and Pastor (1982), in an important generalisation, extended their linear programming scheme
to handle axisymmetric problems for the case of a Von Mises or Tresca material. Although significant, the
formulations of Bottero et al. (1980) and Turgeman and Pastor (1982) both suffer from the disadvantage that
the direction of shearing must be specified for each discontinuity a priori. This precludes the use of a large
number of discontinuities in an arbitrary arrangement, since it is generally not possible to determine these
directions so that the mode of failure is kinematically acceptable. The revised simplex optimisation
procedure used by Bottero et al. (1980) also appears to be rather slow and, indeed, they suggested that a
more efficient solution strategy needed to be found.
One effective means of solving large, sparse LP problems is the steepest edge active set scheme (Sloan,
1988b). Although it was originally proposed for the solution of problems arising from the finite element
lower bound method (Sloan, 1988a), this algorithm has also proved efficient for upper bound analysis
(Sloan, 1989). This is because the form of the dual upper bound LP problem is very similar to the form of
the lower bound LP problem, with more rows than columns in the constraint matrix and all of the variables
unrestricted in sign.
All LP formulations mentioned so far are based on the threenoded triangular element. With this simple
element it is necessary to use a special grid arrangement, in which four triangles are coalesced to form a
quadrilateral with the central node lying at the intersection of the diagonals. If this pattern is not used, then
the elements cannot provide a sufficient number of degrees of freedom to satisfy the incompressibility
condition, as discussed in detail by Nagtegaal et al. (1974). In response to this shortcoming, Yu et al. (1994)
developed a linear strain element for upper bound limit analysis. A major advantage of using this element,
rather than a constant strain one, is that the velocity field can be modelled accurately with fewer elements
and the incompressibility condition can be satisfied without resorting to a restrictive grid arrangement.
More recently, the upper bound formulation of Sloan (1989) was generalised by Sloan and Kleeman
(1995) to allow a large number of discontinuities in the velocity field. In the latter formulation, a velocity
discontinuity may occur at any edge that is shared by two adjacent triangles, and the sign of shearing is
chosen automatically during the optimisation process to give the least amount of dissipated power. Each
discontinuity is defined by four nodes and requires four unknowns to describe the tangential velocity jumps
along its length. Although it is still based on a LP solver, the Sloan and Kleeman (1995) formulation is
computationally efficient for twodimensional applications and gives good estimates of the true limit load
with a relatively coarse mesh. Moreover, it does not require the elements to be arranged in a special pattern
in order to model the incompressibility condition satisfactorily.
As the finite element formulation of the upper bound theorem is inherently nonlinear, a number of upper
bound formulations have been based on nonlinear programming (NLP) methods. Early studies focused
mainly on plates and shells and include the work of Hodge and Belytschko (1968), Biron and Chasleux
(1972), and Nguyen et al. (1977). The most commonly used solution algorithm in these formulations is the
sequential unconstrained minimisation technique (McCormick and Fiacco, 1963) with a Carroll penalty
function. Unfortunately, this scheme is not computationally attractive for large scale problems and few
practical applications have been considered in the literature. Another drawback, which also persists in more
recent upper bound formulations, such as those of Huh and Yang (1991) and Capsoni and Corradi (1997), is
that these formulations can only be used with a limited variety of yield functions.
In a completely different vein, Jiang (1994) proposed an upper bound formulation based on
viscoplasticity theory. Following the practice of others, Jiang (1994) employed an augmented Lagrangian
method to solve the resulting nonlinear optimisation problem, but implemented it in conjunction with the
algorithm of Uzawa. One year later, Jiang (1995) demonstrated that the same nonlinear programming
scheme can be applied to direct upper bound analysis. Despite its good performance for twodimensional
examples, Jiangs formulation has not been extended to deal with threedimensional problems. More
importantly, because the method does not permit discontinuities in the velocity field, it is unlikely to give
accurate results for cohesivefrictional materials.
A straightforward method for evaluating the ultimate loads of structures has been suggested recently by
de Buhan and Maghous (1995). This scheme is based upon the kinematic approach of yield design theory
and leads to the problem of minimising a function of a finite number of variables without constraints. In
their implementation, the optimisation procedure was carried out by means of the simplex method of Nelder
and Mead (1965). Though this formulation was claimed to be computationally efficient, it needed several
minutes of CPU time just to solve a plane strain problem with a grid of around one hundred triangles. These
timings suggest the technique would have difficulty in coping with large scale threedimensional
geometries, where the number of unknowns is typically much greater than twodimensional cases.
In the same year as the work of de Buhan and Maghous, Liu et al. (1995) proposed a method for
performing threedimensional upper bound limit analysis which uses a direct iterative algorithm. The basic
characteristic of this algorithm is that, at each iteration, the rigid zones are distinguished from the plastic
zones and are constrained accordingly. This involves modifying the goal function and the constraint
conditions and neatly avoids the numerical difficulties that are caused by undetermined rigid zones and an
undifferentiable objective function. In essence, the problem is reduced to one of solving a series of relevant
elastic problems. According to their paper the process is efficient, numerically stable, and can be
implemented easily in an existing displacement finite element code.
A new two and threedimensional upper bound formulation, which is based on nonlinear programming,
has very recently been proposed by Lyamin and Sloan (2000b). This scheme uses a similar solver to their
lower bound method (Lyamin and Sloan 1997, 2000a) and employs nodal velocities, element stresses, and a
set of discontinuity variables as the unknowns. Over each element, the velocity field is assumed to vary
linearly while the stresses are assumed constant. As in the formulation of Sloan and Kleeman (1995),
velocity discontinuities are permitted at shared edges between two elements, and the sign of shearing is
chosen automatically to minimise the dissipated power. The upper bound solution is found by using a
twostage quasiNewton scheme to solve the system of nonlinear equations that define the KuhnTucker
optimality conditions. This strategy is very efficient, with preliminary comparisons suggesting a 100fold
speedup over an equivalent linear programming formulation for large scale twodimensional applications.
The scheme can deal with any (convex) type of yield criterion and permits optimisation with respect to both
surface and body forces. Because of its speed, the NLP formulation of Lyamin and Sloan (2000b) is well
suited to threedimensional applications.
6.3 Applications
Some typical soil stability problems are now studied to demonstrate the efficiency of the upper and lower
bound formulations of Lyamin and Sloan (1997, 2000a, 2000b). Three different meshes have been
generated for each of the examples and these can be treated as coarse, medium and fine models of the
original problem. All runs use linear elements and the timings are for a Dell Precision 220 PC with a
Pentium III 800MHz processor.
6.3.1 Rigid Strip Footing On CohesiveFrictional Soil
The exact collapse pressure for a rigid strip footing on a weightless cohesivefrictional soil with no
surcharge is given by the Prandtl (1920) solution:
( )

,
_


,
_

cot 1
2 4
tan tan exp
2
c
q
(1)
where c and are, respectively, the effective cohesion and the effective friction angle. For a soil with a
friction angle of =35

this equation gives q/c=46.14. The boundary conditions and material properties
used in the lower and upper bound analyses, together with some typical meshes, are shown in Figure 2 and
Figure 3.
Figure 3 : Upper bound mesh for strip footing
Figure 2 : Lower bound mesh for strip footing
The lower bound results presented in Table 2 compare the performance of a recent nonlinear
programming formulation (Lyamin and Sloan 1997, 2000a) with that of an equivalent linear programming
formulation (Sloan 1988a, 1988b). The newer procedure demonstrates fast convergence to the optimum
solution and, most importantly, the number of iterations required is essentially independent of the problem
size. For the coarsest mesh, the nonlinear formulation is nearly four times faster than the linear
programming formulation and gives a lower bound limit load that is 2% better. For the finest mesh the
speed advantage of the nonlinear procedure is more dramatic, with a 54fold reduction in the CPU time. In
this case, the lower bound limit load is 1.3% below the exact limit load and the analysis uses just 3.6 seconds
of CPU time. Because the number of iterations with the new algorithm is essentially constant for all cases,
its CPU time grows only linearly with the problem size. This is in contrast to the linear programming
formulation, where the iterations and CPU time grow at a much faster rate, and the CPU time savings from
the nonlinear method become larger as the problem size increases.
The upper bound results presented in Table 3 compare the performance of a new nonlinear programming
formulation (Lyamin and Sloan, 2000b) with that of an equivalent linear programming formulation (Sloan
1989, 1988b; Sloan and Kleeman, 1995). As in the lower bound analyses, the new procedure demonstrates
fast convergence to the optimum solution with an iteration count that is essentially independent of the
problem size. For the coarsest mesh, the nonlinear formulation is nearly four times faster than the linear
programming formulation and gives an upper bound limit load which is 1% better. For the finest mesh the
speed advantage of the nonlinear procedure is more dramatic, with a 155 fold reduction in the CPU time. In
this case, the upper bound limit load is 2.5% above the exact limit load and the analysis uses just 13.6
seconds of CPU time.
Comparing the results in Table 2 and Table 3 it may be seen that, for the fine meshes, the nonlinear
programming formulations bracket the true collapse load to within 3.8%. The solution times required to
achieve this level of accuracy are modest, and the methods can be run easily on a desktop PC. For a similar
Table 2. Lower bounds for smooth strip footing on weightless cohesivefrictional soil.
Mesh Linear programming
NSID=24
Nonlinear programming Ratio of LP/NLP
values
q/c
Error
*
(%)
CPU
(Sec)
Iter. q/c
Error
*
(%)
CPU
(Sec)
Iter. Collapse
pressure
CPU Iter.
coarse 36.70 -20.0 1.19 288 37.64 -18.4 0.31 21 0.98 3.8 13.7
medium 42.52 -7.8 25.9 1845 43.79 -5.1 1.72 33 0.97 15 56
fine 43.99 -4.6 197 5552 45.53 -1.3 3.63 29 0.97 54 191
* With respect to exact Prandtl (1920) solution q
exact
= 46.12c
NSID = number of sides in linearised yield surface.
Table 3. Upper bounds for smooth strip footing on weightless cohesivefrictional soil
Mesh Linear programming
NSID=24
Nonlinear programming Ratio of LP/NLP
values
q/c
Error
*
(%)
CPU
(Sec)
Iter. q/c
Error
*
(%)
CPU
(Sec)
Iter. Collapse
pressure
CPU Iter.
coarse 50.48 +9.4 6.76 1396 49.82 +8.0 1.82 28 1.01 3.7 49.9
medium 48.77 +5.7 133.7 6875 48.00 +4.0 5.5 29 1.02 24.3 237
fine 48.03 +4.1 2105 16771 47.30 +2.5 13.6 30 1.02 155 559
* With respect to exact Prandtl (1920) solution q
exact
= 46.12c
discretisation, the lower bound scheme is generally more accurate than the upper bound scheme, especially
for analyses involving frictional soils with high friction angles.
6.3.2 Square And Rectangular Footings On Weightless CohesiveFrictional Soil
Rigorous solutions for the bearing capacity of rectangular footings are difficult to derive and
approximations are usually adopted in practice. The most common approach is to apply empirical shape
factors to the classical Terzaghi formula for a strip footing (see, for example, Terzaghi, 1943 and Vesic,
1973, 1975).
It is of much interest to compare the lower bound bearing capacity values for square and rectangular
footings with approximations that are commonly used in practice. Many empirical and semiempirical
engineering solutions for the bearing capacity q have been proposed, but focus here is placed on two that use
the modified Terzaghi (1943) equation:

BN N q N c q
i d s q s qi qd qs c ci cd cs

,
_

+ +
2
1
(2)
where B is the footing width, q
s
is the ground surcharge, is the soil unit weight, N
c
, N
q
, N

are bearing
capacity factors,
cs
,
qs
,
s
are shape factors,
cd
,
qd
,
d
are embedment factors and
ci
,
qi
,
i
are
inclination factors. In the case of a surface footing on a weightless soil with vertical loading, Equation (2)
reduces to the simple form:
c cs
N c q (3)
In his original study, Terzaghi (1943) presented a table of values for N
c
and suggested the use of
cs
=1.3
for square and circular footings. Thirty years later Vesic (1973), on the basis of experimental evidence,
advocated the shape factors
cs
=1+(B/L)(N
q
/N
c
) for rectangular footings and
cs
=1+(N
q
/N
c
) for square
footings. In these approximations, N
c
and N
q
are the Prandtl values N
c
=(N
q
1)cot and
N
q
=exp(

tan)tan
2
(/4+

/2).
Table 4 summarises various bearing capacity values for a rough square footing and two types of rough
rectangular footing. These results were obtained using the boundary conditions and mesh layout shown in
Figure 4. All of the cases assume a cohesivefrictional soil with zero selfweight and results are presented
for the lower bound method, the original Terzaghi (1943) method, and the Vesic (1973) method. The
Terzaghi estimates exceed the square footing lower bounds by an average of around 19%, and would appear
Table 4. Predicted bearing capacities q/c=
cs
N
c
for rough square footing and rough rectangular footings
on weightless cohesivefrictional soil.
() Square Rectangular (L/B=2) Rectangular (L/B=5)
Lower
bound
Terzaghi
(1943)
Vesic
(1973)
Lower
bound
Vesic
(1973)
Lower
bound
Vesic
(1973)
0 5.54 7.41 (25%)

cs
=1.3
N
c
=5.7
6.14 (10%)

cs
=1.19
N
c
=5.14
5.35 5.64 (5%)

cs
=1.10
N
c
=5.14
5.17 5.34 (3%)

cs
=1.04
N
c
=5.14
10 9.83 12.48 (21%)

cs
=1.3
N
c
=9.6
10.82 (9%)

cs
=1.30
N
c
=8.35
9.16 9.59 (4%)

cs
=1.15
N
c
=8.35
8.55 8.84 (3%)

cs
=1.06
N
c
=8.35
20 20.33 23.14 (12%)

cs
=1.3
N
c
=17.8
21.23 (4%)

cs
=1.43
N
c
=14.83
17.56 18.03 (3%)

cs
=1.22
N
c
=14.83
15.58 16.11 (3%)

cs
=1.09
N
c
=14.83
to be unsafe. The Vesic (1973) approximations, on the other hand, are within 5% of the rigorous lower
bound results for rectangular footings of all shapes, and their accuracy increases with increasing friction
angle.
The average number of nodes for the finite element models considered here was around 12000, while the
largest mesh analysed had 13392 nodes, 3348 elements and 6444 discontinuities.
Bearing in mind that each
of these nodes has six unknown stresses, these grids generate very large optimisation problems and it comes
as no surprise that the method is computationally demanding. On average, the CPU time for the analyses
shown in Table 4 was 4500 set for a Dell Precision 220 PC with a Pentium III 8OOMHz processor.
Although large, these timings are certainly competitive with those that would be needed for a three-
dimensional incremental analysis with the displacement finite element method. Moreover, the solutions
have the advantage that the limit load, which is obtained explicitly and does not have to be inferred from a
load-deformation plot, is a rigorous lower bound on the true collapse load.
regular mesh
Mesh
nodes 1 13392
c z
#f z ;t200
elements 3348
y=o
I discontinuities 6444
Figure 4 : Smooth rigid rectangular footing on cohesive-frictional soil: (a) general view with applied
boundary conditions and (b) typical generated mesh
CONSTI TUTI VE MODELS FOR GEOMATERI ALS AND I NTERFACES
As identified previously, one of the primary aims of geotechnical design is to ensure an adequate margin
of safety, and this normally involves some form of stability analysis. In the previous section, it was
demonstrated how modern computing methods could be combined with the classical limit theorems of
plasticity to conduct very accurate and very sophisticated stability analyses of geotechnical problems.
Of
course, these analyses provide no information about the behaviour of the soil or rock mass prior to collapse.
They are not designed to do so. In order to predict the response prior to collapse a complete constitutive
model is required for the soil or rock material. Such models must be capable of predicting not only the onset
of failure, but also the complete stress-strain response leading up to failure.
In many cases the ability to
predict the post-failme (post-peak) behaviour is also desirable. For this type of calculation a complete
constitutive model is required.
It is often said that the developments in computer solution methods are far ahead of those related to the
characterisation of the mechanical behaviour of geomaterials and interfaces or joints.
Fortunately, the recent
research emphasis placed on the development and application of constitutive models, so as to account for
important factors that were not included in previous empirical and simplified models, has been well directed.
Unless the materials and interfaces in geotechnical systems are characterised in such a way that they
realistically account for the important factors that influence the material behaviour, results from
sophisticated computer methods will have only limited validity, if at all.
The behaviour of geomaterials and interfaces is affected significantly by factors such as the state of stress
or strain, the initial or i n si t u conditions, the applied stress path, the volume change response, and
environmental factors such as temperature and chemicals, as well as time dependent effects such as creep
15
and the type of loading (static, repetitive and dynamic). In addition, natural geomaterials are often
multiphase, and include fluids (water) and gas as well as solid components. The pursuit of the development
of constitutive models for these complex materials has involved progressive consideration of the foregoing
factors. Most often, the models that have been developed are designed to allow for a limited number of
factors only. Recently, efforts have been made to develop more unified models that can allow inclusion of a
larger number of factors in an integrated mathematical framework. A review of some of the recent
developments is presented below, together with comments on their capabilities and limitations. In view of
space limitations and the large number of available publications, such a review can only be brief and
somewhat selective.
7.1 Elasticity Models
The theory of linear elasticity has a long history of application to geotechnical problems. Initially,
emphasis was placed on the use of linear elasticity for a homogeneous isotropic medium. With the
development of numerical solution techniques, the behaviour of non-homogeneous and anisotropic media
was addressed.
Nonlinear elastic or piecewise linear elastic models were subsequently developed to account mainly for
the influence of the state of stress or strain on the material behaviour. These include the models based on the
functional representation of one or more observed stress-strain and volumetric response curves. The
hyperbolic representation (Kondner, 1963; Duncan and Chang, 1970; Desai and Siriwardane, 1984; Fahey
and Carter, 1993) has been often used for static and quasi-static behaviour, and can provide satisfactory
prediction of the load-displacement behaviour under monotonic loading. However, such functional
simulation of curves is not capable of accounting for other factors such as stress path, volume change, and
repetitive and dynamic loading. Furthermore, they are usually not appropriate when evaluation of stresses
and deformations (strains) are required in local zones in geotechnical systems.
Over the past few decades, great emphasis has been placed on research into the small strain behaviour of
geomaterials. This has been seen as a fundamental issue for many problems in geotechnical engineering
(e.g., Burland, 1989; Atkinson et al., 1990; Atkinson, 1993; Ng et al., 1995; Puzrin and Burland, 1998) In
parallel with detailed experimental work to investigate the small strain response (e.g., Shibuya et al., 1995;
Jamiolkowski et al., 1999), several constitutive models that include the influence of strain level on the
stress-strain response have been developed. There are also examples of their application to boundary value
problems documented in the literature (e.g., Jardine et al., 1986; Powrie et al., 1998; Addenbrooke et al.,
1997; Schweiger et al., 1999).
7.2 Hyperelasticity And Hypoelasticity
Hyperelastic or higher order elastic models have been considered for geological materials. However,
they do not allow for factors such as stress path and irreversible deformations. Hence, their applications to
geomaterials have been quite limited. However, it has been demonstrated that hypoelasticity and its
combination with plasticity can provide useful models for some geomaterials (e.g., Desai and Siriwardane,
1984; Kolymbas, 1988; Desai, 1999).
7.3 Plasticity Models
The available elastoplastic models can be divided into two categories: (a) classical plasticity models, and
(b) continuous yielding or hardening plasticity models. In the classical models, such as those based on the
von Mises, Mohr-Coulomb and Drucker-Prager criteria, plastic yielding occurs after the elastic response,
when a specific yield stress or criterion is exceeded. These models are capable of predicting ultimate or
failure stresses and loads. However, they cannot allow properly for factors such as continuous yielding that
in many cases occurs from the very beginning of loading, volume change response and the effect of stress
path on the failure strength.
Models based on the critical state concept (Roscoe et al., 1958), cap models (e.g., DeMaggio and
Sandler, 1971) and their various modifications were proposed largely to allow for the continuous yielding
r es po ns e e xh i bi te d b y ma n y ge ol o gi ca l mat er i al s. As a r es ul t , th ey ha ve be en o f te n u se d in co mp u te r methods.
At the same time, it should be recognised that they do suffer from some limitations, including the following:
they cannot predict dilative volume change before the peak stress,
they do not allow for different strengths mobilised under different stress paths,
they do not include the effect of deviatoric plastic strains on the yielding behaviour, and
they are usually based on associative plasticity, and hence they do not always adequately include the
effects of friction.
In order to overcome these limitations, a number of advanced hardening models have been proposed with
continuous yield surfaces. These include the hierarchical single surface (HISS) model (Desai et al., 1986;
Desai, 1995; 1999), as well as those proposed by other investigators (e.g., Lade and Duncan, 1975; Kim and
Lade, 1988; Nova, 1988; Lagioia and Nova, 1995; Whittle, et al., 1994).
Kinematic and anisotropic hardening models have been proposed to account for cyclic behaviour under
dynamic loading (Mroz, 1967; Dafalias, 1979; Mroz et al., 1978; Prevost, 1978; Somasundaram and Desai,
1988; Wathugala and Desai, 1993). Most of these models involve moving yield surfaces and may involve
associated computational difficulties. Models based on the unified disturbed state concept (DSC), do not
involve moving surfaces, and generally they are computationally more efficient than the more advanced
hardening and softening models (Katti and Desai 1995; Shao and Desai, 2000).
7.4 Hypoplasticity
Of course, other possibilities of describing the mechanical behaviour of soils have been proposed in the
literature such as hypoplastic formulations (e.g., Kolymbas, 1991; Niemunis and Herle, 1997). Contrary to
elastoplasticity, in hypoplastic models no distinction is made between elastic and plastic deformation and no
yield and plastic potential surfaces or hardening rules are needed. The aim in developing hypoplastic
models for soils is to discover a single tensorial equation that can adequately describe many important
features of the mechanical behaviour of the material. It is also desirable that the parameters of the model for
a granular material depend directly on the properties of the grains.
There are several useful treatments in the literature of hypoplastic models applied to soils (e.g.,
Kolymbas, 1991; Gudehus, 1996; von Wolffersdorff, 1996; Wu et al., 1996; Bauer, 1996; Herle and
Gudehus, 1999). Despite their apparent simplicity some indeed are able to reproduce features of real soils,
such as pressure and density coupling, dilatancy, contractancy, variable friction angle and stiffness.
Although some of the features of these models seem to be very attractive, so far they have been applied only
to sands. More experience in the application of these models is still required.
7.5 Degradation And Softening
Plasticity based models that allow for softening or degradation have also been proposed. In these models
the yield surfaces expand during loading up to the peak stress, and then during softening they are allowed to
contract. This approach may suffer from non-uniqueness of computer solutions because the discontinuous
nature of the material is not adequately taken into account.
The classical continuum damage models (Kachanov, 1986) allow for degradation in the material stiffness
and strength due to microcracking and resulting damage. However, since the coupling between the
undamaged and damaged parts is not accounted for, these models are essentially local and can entail
spurious mesh dependence when used in numerical solutions. In order to render damage models to be non-
local, various enrichments have been proposed. These include consideration of microcrack interaction on
the observed response (Bazant, 1994), and gradient and Cosserat theories (de Borst et al., 1993; Mhlhaus,
1995). Such enhanced models can provide computations that are independent of the mesh layouts.
However, they can be complex and may entail other computational difficulties. The DSC also allows for the
coupling within its framework, and hence, leads to non-local models that are free from spurious mesh
dependence (Desai et al., 1997; Desai, 1999).
The problem of strain localisation that accompanies softening and degradation has received a great deal
of attention in the literature, ever since the developments in numerical methods made possible the solution of
boundary value problems with such material features. It is in problems such as these that the inherent
material response, i.e., softening, and the numerical solution strategy have a strong interaction. Numerical
procedures that deal inappropriately with post-peak material response are likely to provide spurious answers.
It is therefore worth providing a brief summary of the main approaches that have been developed to deal
with the problem of strain localisation. It should be noted here that to date the methods discussed below
have been applied almost exclusively to two-dimensional (plane strain) problems.
7.5.1 Enhanced Finite Element Models
In this approach the domain undergoing localisation is divided into two regions: the localised zone in
which large displacements take place and the non-localised zone in which the strains remain small. The
actual boundary of these two zones is not very well defined, but provided that the dimension of the
representative domain under consideration is large compared to the thickness of the localised zone, the shear
band may be treated, at least conceptually, as a plane of discontinuity in displacement.
Two conceptually different approaches have been proposed. One way is to use special finite elements
with appropriate discontinuous or modified shape functions (Belytschko et al., 1988; Ortiz et al., 1987). In
this approach, firstly, a bifurcation analysis is carried out at the element level. When the onset of
localisation is detected, the element interpolation is extended by adding to it suitably defined shape functions
that reproduce the localised deformation modes.
The other approach is to introduce the effect of the shear band after the onset of localisation through a
constitutive framework (Pietruszczak and Mrz, 1981), which is generally known as the homogenisation
technique.
7.5.2 Non-Local Models
A fully non-local model is obtained by introducing a relationship between the average stresses and the
average strains, which complies with the relationship between macroscopic and microscopic stress for
granular bodies. The macroscopic stress can be thought of as some average of the more rapidly varying
microscopic stress (Brinkgreve and Vermeer, 1995).
In most applications, the non-local formulation is restricted to a specific part of the constitutive equation.
Bazant and Lin (1988) averaged the invariant of plastic shear strains, but alternatively an average of the
plastic strain rate tensor could be taken and then the invariant formulated. Either approach leads to a set of
rate equations, which remain elliptic after the onset of localisation.
A modified non-local model overcoming some of the numerical problems generally associated with non-
local models has been presented by Vermeer and Brinkgreve (1994) and Brinkgreve and Vermeer (1995)
with the considerable advantage of a relatively easy implementation into existing finite element codes. The
numerical results for a purely cohesive material presented by Brinkgreve and Vermeer (1995) showed that
the results are mesh independent.
7.5.3 Gradient Plasticity Models
Gradient dependent models can be viewed as an approximation of fully non-local models. The spatial
derivatives of the inelastic state variables enter the continuum description of the material in addition to the
inelastic state variables. The gradient terms can be introduced either into the flow rule (Vardoulakis and
Aifantis, 1991) or the dilatancy constraint (Vardoulakis and Aifantis, 1989). Due to the gradient term the
tangent stiffness matrix becomes non-symmetric even for associated plasticity. Although gradient plasticity
theories are highly versatile for describing localisation of deformation in a continuum, a disadvantage of the
approach is the introduction of an additional variable at global level in addition to the conventional
displacement degrees of freedom.
7.5.4 Micropolar Continua
In this approach rotational degrees of freedom are added to the conventional translational degrees of
freedom. Additional strain quantities (relative rotation and micro-curvatures) and additional stress quantities
(couple stresses) enter into the continuum description (Mhlhaus and Vardoulakis, 1987; de Borst, 1991).
Since these static and kinematic components are related through a bending modulus, which has the
dimension of length incorporated, an internal length scale is introduced, even in the elastic regime.
The determination of the material properties from test data has been given relatively little attention so far.
A further disadvantage is that the rotations emerge as additional degrees-of-freedom at global level, which
increases the computational effort.
More recently Cosserat continua have been successfully applied to model strain localisation within the
framework of hypoplasticity (Tejchman and Bauer, 1996) using the mean grain diameter as the characteristic
length. Although results look promising the required fineness of the finite element mesh in order to avoid
mesh dependence seems to prohibit practical applications without significant further enhancements.
Cosserat-type models have also been adapted to the solution of boundary value problems in rock mechanics,
for cases where the rock mass has a distinct structure, e.g., foliation (Adhikary et al., 1995, 1996)
7.5.5 Viscoplastic Models
In viscoplasticity the strain rate included in the constitutive equation prevents the set of equations
describing the dynamic motion of the softening solid from becoming elliptic. Failure modes are usually
accompanied by high strain rates and therefore the inclusion of the strain rate into the constitutive equation
seems natural (Perzyna, 1992).
This kind of approach is purely mathematical in the sense that the additional parameters needed are not
directly derivable from experiments and may have little physical meaning. Therefore a semi-inverse
approach is needed (Sluys and de Borst, 1992). A basic drawback of the use of viscosity to restore well-
posedness of the governing equations is the fact that the regularising effect gradually vanishes for the rate-
independent limit or a very slow process, i.e., the method is generally not suitable for static loading. Some
recently published results by Oka et al. (1994) however suggest that by introducing additional material
functions the viscoplastic approach still offers some potential for analysing quasi-static strain localisation
problems.
7.5.6 Mesh Adaptivity Techniques
Mesh adaptation is a numerical approach to capture the localisation pattern by refining the mesh in the
regions where the strain gradients are highest and vice versa for regions with low strain gradients. The total
number of elements may also be increased so that the global error associated with the mesh meets certain
accuracy requirements (Hicks, 1995).
In the adaptive procedure the first step is to define the remeshing criterion. This is closely linked to error
estimation (Zienkiewicz and Zhu, 1987) and the mesh is redefined when the specified error has been
exceeded. It is also possible to redefine the mesh at fixed intervals, but an error analysis is still needed in
forming the new mesh.
The second step is to generate the new mesh by increasing the mesh in the areas where the error is
highest. The third step is to map the current state variables, i.e., stresses, strains and strain hardening or
softening parameters, from the old mesh to the new one, in a process which is called smoothing. For both
these steps several possibilities exist and the final accuracy of the analysis is dependent on the choices made.
With standard finite elements, the mesh geometry may induce the formation of unrealistic and wrong
failure mechanisms - especially when low order triangles are used (Pastor et al., 1992); this is overcome by
using adaptive meshes. However one problem still remains: as the mesh is refined, the shear band thickness
will decrease to zero leading to zero energy dissipation. Therefore, other of limitations need to be imposed,
such as the minimum element size should be limited to a finite proportion (e.g., 1/3) of the actual shear band
width (Zienkiewicz et al., 1995).
7.6 Saturated And Unsaturated Materials
The development of models for saturated geomaterials has been the subject of research for a long time.
Terzaghis effective stress concept (Terzaghi, 1943) is considered to be the earliest model for saturated soils
and forms the basis of most of the stress-strain models developed since then.
Constitutive models capable of describing at least some aspects of the complex behaviour of partially
saturated soils were first explored seriously in relation to foundation problems associated with shrinking and
swelling clay soils (e.g., Aitchison, 1956, 1961; Richards, 1992). The recent emphasis on geoenvironmental
engineering has spurred considerable renewed interest in the characterisation of constitutive models for
partially or unsaturated materials. The literature on the subject is now wide in scope and a review is
available in the text by Desai (1999). Some of the recent models based on the critical state concept,
plasticity theory and the DSC are given in publications by Alonso, et al. (1990), Bolzon et al. (1996),
Pietruszczak and Pande (1996), and Geiser et al. (1997).
7.7 Liquefaction
Liquefaction represents the phenomenon of instability of the saturated geomaterials microstructure at
certain critical values of stress, pore water pressure, (plastic) strains and plastic work or dissipated energy, as
affected by the initial conditions, e.g., initial effective mean pressure and physical state (density).
A number of empirical approaches, based on experimentally observed states of pore water pressure and
stress at which liquefaction can occur, and index properties have often been used to predict the onset of
liquefaction (Casagrande, 1976; Castro and Poulos, 1977; Seed, 1979; National Research Council, 1985;
Ishihara, 1993). Although such conventional approaches have provided useful results, they are not based on
mechanistic considerations that can account for the microstructural modifications in the deforming material.
Dissipated energy has been proposed as a criterion for the identification of liquefaction potential (Nemat-
Nasser and Shokooh, 1979; Davis and Berrill 1982, 1988; Figueroa et al., 1994; Desai, 2000). Recently, the
DSC has been proposed as a mechanistic procedure that is fundamental yet simple enough for practical
application to liquefaction problems (Desai et al., 1998; Park and Desai 1999; Desai, 2000).
7.8 Comments
Most of the constitutive models in the past have addressed a limited number of factors at any time. As a
result, separate models are often warranted for different behavioural features of the same material. This can
lead to complexities such as a greater number of parameters, which is the sum of the parameters for each
model, for given characteristic(s). Hence, it is desirable and useful to develop unified models that can
permit consideration of the behavioural features as special cases. A unified approach can lead to compact
models that are significantly simplified, involve fewer parameters, and are easier to implement in computer
procedures. Such a unified concept, called the disturbed state concept (DSC), has been developed and found
to be successful for numerous
geological materials and interfaces.
The DSC allows implicitly for the
nonl ocal effect s and t he
characteristic dimension and as a
result does not suffer from spurious
mesh dependence (Desai, 1999).
Because of its potentially unifying
qualities, a brief summary of the
DSC is provided in the following
section.
7.9 Disturbed State Concept
(DSC)
The recently developed DSC
represents a unified approach for
const i t ut i ve model l i ng of
geomaterials and interfaces and
joints. It allows for various factors
such as elastic, plastic and creep
strains, microcracking leading to
degradation or damage, and
stiffening or healing under
thermomechanical loading. Its
hierarchical framework permits the
user to adopt specialised version(s),
e.g., including elasticity, plasticity,
viscoplasticity, with disturbance
(degradation and stiffening),
Figure 5 : Schematic illustration of stress-strain behaviour and
material disturbance assumed in DSC models
depending upon the material and application need. Details of the DSC are given in various publications
(Desai, 1995, 2000; Desai et al., 1986, 1991, 1995, 1997, 1998; Desai and Salami, 1987; Desai and
Varadarajan, 1987; Desai and Fishman, 1991; Desai and Ma, 1992; Desai and Rigby, 1997; Desai and Toth,
1996; Liu et al., 2000; Park and Desai, 1999; Shao and Desai, 2000; Wathugala and Desai, 1993) and a
textbook (Desai, 1999).
The essential idea of the DSC is that the response of a material to variations in both internal and external
conditions can be described in terms of the responses of the constituent material parts in two reference states,
the relatively intact state (RI) and the fully adjusted state (FA). The overall response of the material,
i.e., the observed response, is then related to the responses at the reference states through a disturbance
function, which provides a coupling and interpolation mechanism between the responses of the material in
the RI and FA states. A summary of the formulation of DSC models is provided below.
7.9.1 Incremental Equations
The incremental constitutive equations for the DSC are given by:
( ) ( ) d d 1
i c c i a
dD D D d + + (4)
or
( ) ( ) d C d C 1
c i c c i i a
dD D D d + + (5)
where and are the stress and strain vectors, respectively, and superscripts a, i, and c denote observed,
relatively intact (RI), and fully adjusted (FA) states, respectively. D is the disturbance assumed to be scalar
function (it can also be treated as a tensiorial quantity), dD denotes increment or rate of disturbance, and C
denotes the constitutive matrix.
In the DSC, a material element during deformation is considered to be composed of material parts in
different reference states, as indicated in Figure 5. For example, in the case of a dry material, the initial
continuum state is considered to be a reference state and is called the relative intact (RI) state. During
deformation, the material transforms to a fully adjusted (FA) state at disturbed locations, as a consequence of
the changes in its microstructure due to relative particle motions, microcracking or stiffening. The observed
or average response of the material (a) is then expressed in terms of the responses of the material parts in the
RI (i) and FA (c) states. The disturbance, D, acts as the coupling and interpolation mechanism between the
RI and FA states so as to yield the observed response.
The RI response can be characterised by using continuum models such as those based on elasticity,
plasticity or elastoviscoplasticity. For instance, the RI response can be characterised by using the
hierarchical single surface (HISS) plasticity models. For the associative case, the yield surface, F, in the
HISS-
0
model is given by:
( )( ) 0 1
5 . 0 2
1 1 2
+

r
n
D
S J J J F
(6)
where J
1
is the first invariant of the stress tensor,
ij
,
J
2D
is the second invariant of the deviatoric stress tensor, S
ij
,
S
r
is the stress ratio defined as
2 3
2 3
2
27
/
D D
J J

, and
J
3D
is the third invariant of S
ij
.
The superior bar denotes a quantity non-dimensionalised with respect to atmospheric pressure (p
a
). J
1
= J
1
+ 3R where 3R is the bonding (tensile or cohesive strength), and are ultimate (failure) parameters, n is
the phase change parameter associated with the state at which volume change from contraction to dilation
occurs, and is the hardening or growth function, which, in a simple form, is expressed as:
1
1

(7)
where ( ) ( )
1
]
1

1/2

p
T
p
d d is the trajectory of total plastic strains, and a
1
and
1
are the hardening
parameters.
The FA response can be characterised in different ways. If the FA (microcracked or damaged) material
is assumed to carry no stress at all, i.e., it acts as a void as in the classical damage model, Equation (4) will
specialise to that for the classical damage model (Kachanov, 1986). This characterisation is considered to be
inappropriate, as it does not include the coupling between the RI and FA states. If the FA material is
assumed to carry hydrostatic stress and no shear stress, its behaviour (C
c
) can be characterised on the basis
of its bulk response. In a general method, the FA response can be characterised by using the critical state
concept (Roscoe et al., 1958), in which the FA material continues to carry the shear stress and deform in
shear without change in its volume, under a given mean pressure. The critical state equations to characterise
the FA response are given by (Roscoe et al., 1958; Desai, 1999):
c c
D
J m J
1 2

(8)
( ) 3 ln
1 1 a
c c
o
c
p / J e e (9)
where e is the voids ratio and is a material parameter.
7.9.2 Disturbance
Disturbance is expressed as the ratio of the volume of the material in the FA state to the total volume. In
a phenomenological context, D is expressed on the basis of observed response (e.g., stress-strain, volumetric
or void ratio, effective stress or pore water pressure and nondestructive properties such as P- or S-wave
velocities), and in terms of the deviatoric plastic strain trajectory or plastic work (Desai, 1999). For
example:
c i
a i
D

(10)
and
1
,
_



Z
D
A
u
e D D

(11)
where is an appropriate stress measure, i.e.,
1
,
1
-
3
or D
J
2 .
D
is the deviatoric plastic strain
trajectory, and A, Z and D
u
are disturbance parameters.
7.9.3 Specialisations
Equation (4) includes various continuum and damage models as special cases. For example, if D=0, the
following results:
i i i
d C d (12)
where C
i
can be based on elasticity, plasticity or elastoviscoplasticity theories. If the FA part cannot carry
any stress at all, Equation (4) gives the classical damage model:
( )
i i i a
dD d C D d 1
(13)
which does not include the effect of the microcracked or damage parts on the observed behaviour.
7.9.4 Interfaces And Joints
An interface or joint in a geological system constitutes a junction between two (similar or dissimilar)
materials and represents a discontinuity where relative motions (slip, debonding or separation, rebonding
and interpenetration) can occur. In order to allow for the relative motions, it becomes necessary to develop
constitutive models based on appropriate and specialised laboratory tests. The force and kinematic
conditions at the interface are introduced by developing various models: springs to represent normal and
shear response, special joint elements to allow for relative displacements (Goodman et al., 1968; Ghaboussi
et al., 1973) and force constraints (Katona, 1983). A number of models have been proposed as
modifications or variations of the foregoing approaches. An alternative, called the thin-layer element
approach, has been proposed (Desai et al., 1984). In this approach the smeared interface zone is treated as
a solid element, and its stiffness properties are evaluated in the same manner as the other solid elements in
the neighbouring regions. However, the material parameters are obtained from special shear tests, such as
simple shear or direct shear tests (Desai and Rigby, 1997).
The DSC can also be applied for the characterisation of interfaces and joints. The same mathematical
framework, Equation (4), is applicable. As a result, the models for geological materials and interfaces are
essentially the same. This eliminates the inconsistency in previous approaches in which different models
were used for soils and interfaces. For example, if the HISS-
0
model is used to define the RI behaviour, the
yield function is given by (Desai and Fishman, 1991; Desai and Ma, 1992):
0
2
+
q
n
n
n
F
(14)
where and
n
are the shear and
normal stresses in a two-
dimensional interface, respectively,
is the ultimate or failure
parameter, n is the phase change
parameter, q allows a curved
ultimate (failure) envelope, and is
the hardening function expressed in
terms of the plastic relative shear
and normal displacements. The
disturbance is defined on the basis
of results from tests involving the
application of shear and normal
loads.
7.9.5 Parameters
The material parameters in the
DSC have physical meanings in that
they are related to physical states
during deformation. Their number
is usually smaller than that in other
available models of comparable
capabilities. For instance, with two
elastic (E, ), and five plasticity (,
, a
1
,
1
, R) parameters, the HISS-

0
model can include the effects of
the factors previously mentioned,
that are not present in the critical
state models (with 6 parameters)
and Cap models (with 10
parameters).
The parameters in the DSC
model can be found from laboratory
Figure 6 : Finite element mesh used for field pile test
uniaxial, shear, triaxial or multiaxial tests. The standard triaxial compression (and extension) test is suitable
to quantify the parameters defining the elasticity, plasticity and disturbance. For viscous behaviour, creep or
relaxation tests are required. Details of procedures for finding the parameters are available in various
publications (e.g., Desai and Ma 1992; Desai et al., 1995; Desai, 1999).
7.9.6 Validations and Applications
The DSC model and its specialised versions have been calibrated with respect to test data for a wide
range of geological materials and interfaces, and other materials such as concrete ceramics, metals and
alloys (Desai, 1995, 1999; Desai et al., 1984, 1991, 1995, 1997, 1998; Desai and Salami, 1987; Desai and
Fishman, 1991; Desai and Ma, 1992; Desai and Toth, 1996; Desai and Rigby, 1997). They have been
validated with respect to tests used to find the parameters and independent tests not used in finding them.
The DSC has been implemented in nonlinear finite element codes for static and dynamic analysis
including dry and saturated materials. The codes are used to predict observed behaviour of a number of
practical problems in geotechnical engineering. These include two- and three-dimensional analysis of
footings and pile foundations, reinforced retaining walls, dams and excavations, underground works
(tunnels), anchor-soil systems, landslides, consolidation problems, and seepage and liquefaction in shake
table tests. Two typical examples are given below to illustrate some of the capabilities of these sophisticated
models.
7.9.7 Example - Cyclic Analysis Of Piles In Marine Clay
Figure 6 shows the finite element mesh for a displacement-controlled field load test performed by Earth
Technology Corp. (1986) on a 76.2 mm diameter pile segment. The numerical calculations involved
simulation of the in situ stress conditions, driving effects, consolidation, tension tests and finally, axial cyclic
loading (Shao and Desai 2000).
Figure 7 : Comparison between field measurements and predictions from DSC and HISS models: one-way
cyclic load tests
The material properties for the DSC with the HISS-
0
model for the RI behaviour of the marine clay were
determined from static and cyclic triaxial tests on cylindrical specimens and multiaxial tests on cubical
specimens (Katti and Desai, 1995; Shao and Desai 2000; Wathugala and Desai, 1993).
The parameters for the interface between the steel pile and the clay were determined from laboratory tests
using the cyclic multi degree-of-freedom (CYMDOF) shear device, including measurements of pore water
pressures. Details of the parameters are given elsewhere (Katti and Desai, 1995; Shao and Desai 2000;
Wathugala and Desai, 1993).
The finite element analyses were performed by using the DSC model and the kinematic hardening HISS-

0
* model (Wathugala and Desai, 1993) implemented in the DSC-DYN2D computer code. The latter allows
for cyclic degradation and interface effects.
Figures 7 and 8 show comparisons between the computed results using the DSC and
0
* model and the
observed data for one-way and two-way cyclic load tests, respectively. These figures also include the
predicted growth of disturbance during the loading. It can be seen that the DSC model provides very good
correlation with the observed results. Furthermore, it shows improved predictions compared to those from
the HISS-
0
* plasticity model because the DSC model allows for cyclic degradation and interface effects.
7.9.8 Example - Dynamic Analysis Of Shake Table Test And Liquefaction
Figure 9(a) shows the shake table test configuration (Akiyoshi et al., 1996). The applied loading
involved specifying horizontal displacement, X, at the bottom nodes, given by the following function:
( ) t f u X
x
2 sin (15)
Figure 8 : Comparison between field measurements and predictions from DSC and HISS models: two-way
cyclic load tests
where u
x
is the amplitude (= 0.0013 m), f is the frequency (= 5 Hz) and t is time.
Fuji river sand was used in the shake table test. Because various physical properties of the Fuji river sand
and the Ottawa sand are similar, e.g., the grain-size distribution, Figure 9(b), the multiaxial cyclic test data
for the Ottawa sand were used in the finite element analysis (Park and Desai 1999).
Shaking direction Shaking machine
1
0
0
3
0
0
7
0
0
8
0
0 1
0
0
0
Sand
1500
1700
Accelerometer
Pore water pressure meter
Unit: mm
0.01 0.1 1 10
(b) Grain Size Distribution Curves of Ottawa Sand and Fuji River Sand
0
20
40
60
80
100
Figure 5 Shake Table Setup and Grain Size Distributions for Sands
Gradation Curve
Fuji Sand
Ottawa Sand
(a) Shake Table Test Set-Up (Akiyoshi, et al., 1996)
%
F
i
n
e
r
Particle size (mm)
Figure 9 : Shake table test setup and grain size distribution for sands
The finite element mesh is
shown in Figure 10. The steel
in the test box was assumed to
be linear elastic, while the
DSC model was used to
represent the behaviour of the
sand. The idea of the
repeating side boundaries was
employed, in which the
displacements of the side
boundary nodes on the same
hori zont al pl anes were
assumed to be the same. The
analysis used the computer
code DSC-DYN2D.
Fi gur e 11 s hows
comparisons between the
measured and computed
excess pore water pressures
with time, at the point (depth
= 300 mm) shown as a solid
dot in Figure 12. The test data
indicate liquefaction after
about 2.0 secs when the pore
water pressure equals the
initial effective stress. It can
be seen from Figure 11 that the
finite element predictions
compare well with the test
results.
Figures 12(a) to (d) show
growth of disturbance in the
mesh at typical times = 0.5,
1.0, 2.0 and 10.0 secs,
respectively; the plot of
computed disturbance at depth
= 300 mm is shown in Figure
12(e). Laboratory tests on the
Ottawa sand showed that the
liquefaction initiates at an
average value of the critical
disturbance, D
c
=0.84 (Desai
et al., 1998). At time t = 0.5
sec, the computed disturbance
in the sand is well below the
critical value in all elements.
At t

=1.0sec, the disturbance


has grown, and its value is
between 0.50 and 0.70 in the
elements in the middle zone,
which is below D
c
= 0.84. At
t

2.0

secs, the disturbance has reached values higher than 0.80 at and below the depth of 300 mm; this
indicates that liquefaction initiates at about 2.0 secs. At t=10secs, the disturbance in about 80% of the test
box has grown to a value equal to or greater than the critical value, indicating that the soil has liquefied and
failed.
Figure 11 : Excess pore pressures at a depth of 300 mm
Figure 10 : Finite element mesh used for simulation of shake table test
These and other results show that the DSC model can provide a fundamental and simplified approach for
the evaluation of liquefaction potential (Desai, 2000; Desai et al., 1998).
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
(a) Time = 0.5 sec (b) Time = 1.0 sec
(d) Time = 10.0 secs (c) Time = 2.0 secs
0 1 2 3 4 5 6 7 8 9 10
Time (sec)
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
D
i
s
t
u
r
b
a
n
c
e
(
D
)
(e) Disturbance vs. time at depth = 300mm
Figure 8 Growth of Disturbance in Sand and at Depth = 300mm
Figure 12 : Growth of the disturbed zone in sand
7.10 Structured Soils
Much of the research on constitutive models conducted to date has focused on the behaviour of soil
samples reconstituted in the laboratory. The emphasis on reconstituted materials arose because of both
convenience and the need to maintain control over test and sample conditions, if models truly reflecting the
response to changing applied stresses were to be developed. However, natural soils (and rock masses) differ
significantly from laboratory prepared specimens, and these differences need to be accounted for if the
developed models are to be applicable to materials found in nature. For example, there still exists a need to
develop constitutive models that can take into account the structure and fabric of most natural materials.
The important influence of structure on the mechanical properties of soil has long been recognised (e.g.,
Casagrande, 1932; Mitchell, 1976). The term soil structure is used here to mean the arrangement and
bonding of the soil constituents, and for simplicity it encompasses all features of a soil that cause its
mechanical behaviour to be different from that of the corresponding reconstituted soil. The removal of soil
structure is referred to as destructuring, and destructuring is usually a progressive process.
In recent years, there have been several studies in which a theoretical framework for describing the
behaviour of structured soils has been formulated (e.g., Burland, 1990; Leroueil and Vaughan, 1990; Gens
and Nova, 1993; Cotecchia and Chandler, 1997; Liu and Carter, 1999). It is rational to study the behaviour
of natural soils by using knowledge of the corresponding reconstituted soils as a frame of reference
(Burland, 1990, Liu and Carter, 1999). The disturbed state concept theory (DSC), described previously,
provides a convenient framework to predict the difference in structured soil behaviour from that of the
reconstituted soil, and it may also have the potential to describe the destructuring of a natural soil with
loading. One such DSC model, suitable for isotropic and one-dimensional compression, has been
formulated by Liu et al. (2000), based on a particular disturbance function. Simulations of the proposed
DSC model for both structured soil and reconstituted soil have been made and compared to experimental
data. An illustration of the accuracy of this method for incorporating the effects of soil structure is included
here.
The disturbed state concept allows flexibility in choosing ways to define the two reference states
according to the practical problem of interest and the knowledge available, including its response in
laboratory and field tests. A special selection of the reference states for quantifying the influence of soil
structure on soil behaviour has been suggested by Liu et al. (2000). The fully adjusted state (FA) was
chosen to be the corresponding reconstituted state. The fully adjusted state for the virgin compression of a
structured soil is therefore based on the following two assumptions: (1) the material is reconstituted and has
the same mineralogy as the structured soil, and (2) the soil is in a state of virgin compression and the stress
state is the same as that applied to the structured soil. The relative intact state (RI) was chosen to be the
zero state, i.e., the state with no response to stress (a perfectly rigid material).
As indicated previously, a key step in building a DSC model lies in finding a disturbance function.
Based on a study of a large body of experimental data on the compression behaviour of structured clays and
other soils (e.g., Burland, 1990; Leroueil and Vaughan, 1990; Smith et al., 1992), the following disturbance
function, D
v
, for the compression behaviour of naturally structured soils was proposed by Liu et al. (2000).

,
_

p
p
b D
i , y
v
1
(16)
where b is the disturbance index for compression, p is the mean effective, and p
y,i
represent the mean
effective stress at which virgin yielding first occurs. Further details of the mathematical formulation of this
approach can be found in Liu et al. (2000).
An example of the application of the DSC to the destructuring of clay during one-dimensional
compression is shown in Figure 13 for Winnipeg clay (after Graham and Li, 1985). Simulations using the
DSC for Winnipeg clay and six other soils over a range of mean effective stresses from 10kPa to 2,000kPa
have been made by Liu et al. (2000). For one soil (Mexico City clay) the magnitude of volumetric strain
reached was as high as 120%. It was found that the proposed DSC model can describe successfully the
compression behaviour of both naturally structured and artificially structured soils, no matter if the
behaviour of the corresponding reconstituted soil is linear or non-linear in the e-lnp space. A complete
t hree-di mensi onal model ,
capable of quantifying the
shearing response of structured
soil using the DSC is now
required before this approach
can be adopted in situations
other than one-dimensional
compression.
7.11 Which Constitutive
Model Shoul d Be
Chosen?
Choosi ng a sui t abl e
constitutive model for any given
soil for use in the solution of a
particular boundary value
p r o b l e m c a n p r o v e
problematical. Many theoretical models have been published in the literature and each has its own particular
strengths and all have their limitations. Determining the key model parameters is often a major difficulty,
either because some model parameters may have no real physical meaning, or inadequate test data may be
available. Considerable experience and judgement are therefore required in order to make a sensible,
meaningful selection of the most appropriate model and the most suitable values of its parameters.
It is unlikely that a universal constitutive model, capable of providing accurate predictions of soil
behaviour for all soil types, all loading and all drainage conditions, will be proposed, at least in the
foreseeable future. Although the DSC seems to offer a framework for unifying many of these models, it
seems likely, at least in the short to medium term, that progress will continue to be made through the
development and use of models of limited applicability, but nevertheless capable of accurate predictions.
Knowing the limitations of such models is as important as knowing their strengths.
7.11.1 Example Foundations On Carbonate Sands
The issue of selecting the most suitable stress-strain model and its parameter values comes into sharp
focus for the case of shallow foundations resting on structured carbonate soils of the seabed. This is because
natural samples of these materials are often lightly to moderately cemented and the degree of cementation
can by highly variable. There are difficulties associated with obtaining high quality samples of these natural
soils. In order to avoid these difficulties, and as a means of initiating meaningful studies of cemented soils,
laboratory test samples have been prepared by artificially cementing reconstituted samples of carbonate
sands recovered from the sea floor (e.g., Huang, 1994; Huang and Airey, 1998; Carter and Airey, 1994).
The results of a sequence of related studies on artificially cemented calcareous soils (Huang, 1994; Yeoh,
1996; Pan, 1999) and attempts to model them numerically, are described briefly here.
7.11.2 Typical Behaviour
The volumetric behaviour of the artificially cemented sands is illustrated in Figure 14. Typical behaviour
during drained triaxial compression is indicated in Figures 15 and 16.
Figure 14 includes results for samples of cemented and uncemented soils, prepared with different initial
densities. All specimens exhibit consolidation behaviour resembling overconsolidated clay. Initially the
response is stiff, but eventually at a sufficiently high stress level the rate of volume change with stress level
exhibits a marked change and the samples become much more compressible. The transition is a result of
breakdown in the cement bonding accompanied by some particle crushing and perhaps particle
rearrangement. High volume compressibility is eventually exhibited by all specimens, and the presence of
an initial cementation only slightly affects the position of the consolidation curve.
Figures 15 and 16 indicate that the cemented material exhibits relatively brittle behaviour at low
confining pressures, accompanied by a tendency to dilate. However, even at relatively moderate confining
0
5
10
15
20
25
0 100 200 300 400 500
Vertical effective stress s'
v
(kPa)
V
o
l
u
m
e
t
r
i
c

s
t
r
a
i
n

e
v

(
%
)
Winnipeg clay
s'
vy,I
= 200 kPa
b
v
= 0.28
Structured
Reconstituted
Structured
Reconstituted
simul.
test
Figure 13 : Comparison of DSC predictions and test results for a
structured clay
st resses sheari ng i nduces
significant volume reduction in the
specimens and the behaviour
appears to be much more ductile,
although there is still some
brittleness due to cementation.
7.11.3 Stress-Strain Models
A selection of constitutive
models was investigated to
determine their suitability to
represent t he st ress-st rai n
behaviour of the artificially
cemented carbonate soil in both
single element (i.e., triaxial)
tests and boundary value problems
(Islam, 1999). The models
considered are listed in Table 5,
together with the number of
parameters required to describe
completely the constitutive
behaviour. None employs the
DSC.
All models are elastoplastic
and involve either strain hardening
or softening. All have only one
yield surface, except the
Molenkamp (1981) model (also
referred to as Monot), that has
two. Some adopt associated
plastic flow, while others do not.
None includes the effects of
cementation directly by including
a finite cohesion and tensile
strength. Rather, they incorporate
its effects indirectly by regarding a
c e me nt e d s oi l a s a n
overconsolidated material.
The model developed by
Lagioia and Nova (1995) was
designed specifically for a
cemented carbonate soil, though
not the soil specifically considered
here. The Molenkamp model has
been used previously in a
comprehensive study of the
behaviour of foundations on
calcarenite on the North-West
Shelf of Australia (Smith et al.,
1988), as well as in studies of
other granular media (e.g., Hicks,
1992). Although formally it requires specification of a relatively large number of input parameters (23),
many of these parameters may be assigned standard values.
1.0
1.2
1.4
1.6
1.8
2.0
2.2
0.1 1 10 100
Mean effective stress (MPa)
S
p
e
c
i
f
i
c

v
o
l
u
m
e

(
v
=
1
+
e
)
Uncemented
20% Cement
Figure 14 : Isotropic compression of carbonate sands
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
0 5 10 15 20 25 30 35
Axial strain (%)
D
e
v
i
a
t
o
r

s
t
r
e
s
s

(
M
P
a
)
1.2
0.6
0.3
0.1
Effective confining stress (MPa)
Figure 15 : Typical stress-strain behaviour of carbonate sand
-2
0
2
4
6
8
10
12
0 5 10 15 20 25 30 35
Axial strain (%)
V
o
l
u
m
e
t
r
i
c

s
t
r
a
i
n

(
%
)
0.1
0.3
0.6
1.2
Effective confining stress (MPa)
Figure 16 : Typical volumetric behaviour of carbonate sand
The model labelled SU2 is based very
closely on the well-known Modified Cam
Clay (MCC) model, but with one important
distinction (Islam, 1999). It does not use the
MCC ellipse as a yield surface, however it
does use the original ellipse as a plastic
potential. In SU2 the yield surface in p-q
space is a flatter ellipse, since this shape better
matches the experimental data for the
uncemented carbonate sand. However, it does underpredict the peak strength of the cemented soil. The new
flattened ellipse corresponds to an increased separation of the isotropic consolidation and critical state lines
in voids ratio effective stress space. Whereas in Modified Cam clay this separation corresponds to a ratio
of mean effective stresses of 2, in the model SU2 a ratio of 5 has been found to fit the data for carbonate
soils more accurately. These details are illustrated graphically in Figure 17. It is evident from this figure
that the flow law resulting from the adoption of the MCC ellipse as a plastic potential produces good
predictions of the volumetric response of both the cemented and uncemented soil.
Values for the parameters of each model have been selected to provide a good fit to the triaxial test data.
The fitting process was conducted for a comprehensive series of test results, using data from both drained
and undrained tests, in order to obtain the best possible overall agreement between model predictions and
measured behaviour. A typical comparison between the model predictions and the experimental data is
given in Figures 18 and 19 for the case of a drained triaxial test conducted at a confining pressure of
300kPa.
Figure 18 indicates that most of the models considered predict the strength of the cemented soil at large
strains and under-predict the shear strength at smaller values of axial strain. The models can reasonably
predict the final critical state shear strength, but not the larger peak strengths due to the contribution of
cementation. In this case the Modified Cam Clay model provides the best prediction of the stress-strain
Table 5. Stress-strain models
Model Parameters
Modified Cam Clay 5
Lagioia & Nova 6
SU2 6
Molenkamp 23
0.0
0.2
0.4
0.6
0.8
1.0
1.2
0.0 0.2 0.4 0.6 0.8 1.0
Normalized mean stress
N
o
r
m
a
l
i
s
e
d

d
e
v
i
a
t
o
r
i
c

s
t
r
e
s
s Modified Cam Clay yield locus
SU2 yield locus
Critical State Line
Experimental data
Cemented soil
0.0
0.2
0.4
0.6
0.8
1.0
1.2
0.0 0.2 0.4 0.6 0.8 1.0
Normalized mean stress
N
o
r
m
a
l
i
s
e
d

d
e
v
i
a
t
o
r
i
c

s
t
r
e
s
s
Modified Cam Clay yield locus
SU2 yield locus
Critical State Line
Experimental data
Uncemented soil
-1.0
0.0
1.0
2.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Stress ratio q/p
'
D
i
l
a
t
a
n
c
y
Cemented soil
-1.0
0.0
1.0
2.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Stress ratio q/p
'
D
i
l
a
t
a
n
c
y
Uncemented soil
Figure 17 : Details of the model SU2 for uncemented (left) and artificially cemented (right) carbonate soils
curve, consistent with the fact that
it was also able to predict
reasonably accurately the peak
strengths (Figure 17). However,
Figure 19 reveals that the MCC
model is a poor predictor of the
volumetric behaviour during
drained shearing. It predicts
dilation, while the test data
indicate that the sample contracted
during shearing. All other models
provide reasonable predictions of
the volume reduction. Clearly the
adoption of an associated flow
rule in the MCC model is
inappropriate for this material.
7.11.4 Model Footing Tests
Vertical Loading
Model footing tests have been
carried out on 300mm diameter
samples of the artificially
cement ed carbonat e sand
described previously (Yeoh,
1996). An isotropic effective
confining pressure of 300kPa was
applied to the sample and a
50mm diameter rigid footing was
pushed into its surface at a
constant rate, slow enough for
fully drained conditions to be
achieved. A typical set of results
from this series of tests is
presented in Figure 20, which also
indicates the model predictions
obtained using the same set of
model parameters obtained by
fitting each model to the triaxial
test data. It is evident that most
models provide reasonable
predictions of the performance of
the footing throughout the entire
range of footing displacements
considered, i.e., up to a settlement
equal to 30% of the footing
diameter. However, the MCC
model generally overpredicts the
stiffness of the footing. This is
not unexpected, since for this case
MCC predicts dilation during
shearing rather than the observed
compression.
7.11.5 Model Footing Tests Inclined Loading
0
2
4
6
8
10
12
0 5 10 15 20 25 30
Normalised displacement d/B (%)
B
e
a
r
i
n
g

p
r
e
s
s
u
r
e

(
M
P
a
)
Modified Cam Clay
Molenkamp
Nova
Sydney Uni. 2
Experimental data
Finite element simulation of axisymmetric
model footing on artificially cemented
carbonate soil
Figure 20 : Predictions of model footing tests
0
500
1000
1500
0 10 20 30 40 50
Axial strain (%)
D
e
v
i
a
t
o
r

s
t
r
e
s
s

(
k
P
a
)
Modified Cam Clay
Molenkamp
Nova
Sydney Uni. 2
Experimental data
Carbonate sand, North West shelf, Australia
Triaxial test, Cell pressure : 300 kPa
Density : 13 kN/m
3
, Cement content : 20 %
Figure 18 : Comparison of stress-strain predictions
-5.0
0.0
5.0
0 10 20 30 40 50
Axial strain (%)
V
o
l
u
m
e

s
t
r
a
i
n

(
%
)
Modified Cam Clay
Molenkamp
Nova
Sydney Uni. 2
Experimental data
Carbonate sand, North West shelf, Australia
Triaxial test, Cell pressure : 300 kPa
Density : 13 kN/m
3
, Cement content : 20 %
Figure 19 : Comparison of volumetric-strain predictions
Dilation
Predictions of the SU2 model for
cases of circular footings subjected to
inclined loading (Islam, 1999) have
also been compared with experimental
measurements (Pan, 1999) made at 1g
in the laboratory. These comparisons
are shown in Figure 21, where
satisfactory agreement can be
observed over a large range of values
of t he nor mal i sed f oot i ng
displacement.
7.11.6 Centrifuge Tests
Further validation of the SU2
model in predicting the behaviour of
footings on carbonate sand is
demonstrated in Figure 22 (Islam,
1999), where model predictions are
compared with centrifuge test results
obtained by Finnie (1993). These
tests involved the vertical loading of a
rigid circular footing resting on a
finite layer of cemented carbonate
sand, overlying carbonate silt. The
failure mechanism in this case
involved a punching mechanism with
the region of cemented soil
immediately under the footing
penetrating the underling silt.
7.11.7 Discussion
It may be concluded from this study of stress-strain behaviour that the Modified Cam Clay model
matches well the yield surface of the artificially cemented soil but is a poor predictor of the volume change
that accompanies shearing. This has important consequences when the model is used in the prediction of a
boundary value problem, as demonstrated. The other models considered in this study provide reasonable
matches to the single element behaviour, although they are generally incapable of predicting accurately the
peak strengths. However, these same models provide quite reasonable predictions of the behaviour of model
footings, especially at very large footing displacements. This indicates that the peak strengths of the
materials have only a small influence on the overall response of the footings. Excluding MCC, it would
seem that there is little to choose between the models for use in analysing the behaviour of vertically loaded
footings on this type of material. However, the model SU2 has the attraction of being closely based on the
well-known MCC model, with only one simple but significant change, and it has relatively few parameters
to be quantified, all of which have a clear physical meaning. Again, it is emphasised that none of the models
considered here incorporates explicitly the effects of cementation or the break down of that cementation due
to increasing isotropic and deviatoric stress. For cases where these factors may be important, different
constitutive models will be required, and the disturbed state concept described previously, is one method that
shows promise for capturing these effects.
8.0 SOME LIMITATIONS AND PITFALLS OF NUMERICAL ANALYSIS
8.1 Introduction
When using numerical analysis it is necessary to discretise the geometry of the boundary value problem
into a finite number of sub-regions. For example, when using the finite element method the geometry is
0
2500
5000
7500
10000
0 5 10 15 20 25 30
Normalised displacement d/B (%)
A
v
e
r
a
g
e

t
r
a
c
t
i
o
n

(
k
P
a
)
Load inclination, degrees
0
30
20
10
0
30
20
10
Predictions
Data
Figure 21 : Circular footing on cemented sand
0
600
1200
1800
0 5 10 15 20 25 30
Normalised vertical displacement d/B (%)
A
v
e
r
a
g
e

b
e
a
r
i
n
g

p
r
e
s
s
u
r
e

q
B : 4m
t/B : 0.5
Strong crust
Weak crust
Medium crust
Prediction
Data
Figure 22 : Circular footing on cemented layer over silt
divided into an assemblage of finite elements, whereas in the finite difference approach the geometry is
represented by a grid of points. Approximations are then made as to how the primary variables, usually
displacements, vary within these sub-regions. Clearly this introduces approximations and to obtain accurate
results it is necessary to have sufficiently small sub-regions in areas where there are high gradients of stress
and strain. This problem is fundamental to all numerical analysis and occurs no matter what constitutive
model is being used, although the use of a nonlinear constitutive model can complicate the issue. The other
main source of error involved in nonlinear numerical analysis is associated with the integration of the
constitutive equations. Approximations must be made for this to be achieved and many solution strategies
exist for performing this task.
Other approximations that are involved in numerical analysis are those arising from the idealisations
made when reducing the real problem to a form which can be analysed. This usually involves geometric
approximations and idealisations as to material behaviour. A potential source of error is associated with a
lack of in depth understanding of the constitutive models employed to represent soil behaviour. This is a
common source of error, due to the complexities of many of the constitutive models currently available.
To illustrate the potential pitfalls associated with the above approximations a short discussion on
nonlinear solution strategies and two common problems that occur with constitutive models are presented
below. An in depth discussion of many of the restrictions and pitfalls involved in numerical analysis of
geotechnical problems is given by Potts and Zdravkovic (1999, 2000).
8.2 Nonlinear Numerical Analysis
When analysing any boundary value problem, four basic solution requirements need to be satisfied:
equilibrium, compatibility, constitutive behaviour and the boundary conditions. Nonlinearity introduced by
the constitutive behaviour causes the governing equations to be reduced to an incremental form. For
example for the finite element method the equations take the form:
[ ]{ } { }
i
G
i
nG
i
G
R d K
(17)
where [K
G
]
i
is the incremental global system stiffness matrix, {d}
i
nG
is the vector of incremental nodal
displacements, {R
G
}
i
is the vector of incremental nodal forces and i is the increment number. To obtain a
solution to a boundary value problem, the change in boundary conditions is applied in a series of increments
and for each increment Equation (17) must be solved. The final solution is obtained by summing the results
of each increment. Due to the nonlinear constitutive behaviour, the incremental global stiffness matrix [K
G
]
i
is dependent on the current stress and strain levels and therefore is not constant, but varies over an
increment. Unless a very large number of small increments are used this variation should be accounted for.
Hence, the solution of Equation (17) is not straightforward and different solution strategies exist. The
objective of all such strategies is the solution of Equation (17), ensuring satisfaction of the four basic
solution requirements listed above.
This strategy is a key component of a nonlinear analysis, as it can strongly influence the accuracy of the
results and the computer resources required to obtain them. Many different solution strategies exist but few
comparative studies have been performed to establish their merits for geotechnical analysis. To illustrate the
limitations and pitfalls that can arise three different categories (namely, tangent stiffness, visco-plastic and
modified Newton-Raphson) of solution algorithm are considered and compared.
8.3 Tangent Stiffness Method
8.3.1 Introduction
The tangent stiffness method, sometimes called the variable stiffness method, is the simplest solution
strategy. This is the method implemented in a variety of computer codes including CRISP (Britto and Gunn,
1987), which is widely used in engineering practice.
In this approach, the incremental stiffness matrix [K
G
]
i
in Equation (17) is assumed to be constant over
each increment and is calculated using the current stress state at the beginning of each increment. This is
equivalent to making a piece-wise linear approximation to the nonlinear constitutive behaviour. To illustrate
the application of this approach, the simple problem of a uniaxially
loaded bar of nonlinear material is considered, see Figure 23. If
this bar is loaded, the true load displacement response is shown in
Figure 24. This might represent the behaviour of a strain
hardening plastic material which has a very small initial elastic
domain.
8.3.2 Numerical Implementation
In the tangent stiffness approach the applied load is split into a
sequence of increments. In Figure 24 three increments of load are
shown as R
1
, R
2
and R
3
. The analysis starts with the
application of R
1
. The incremental global stiffness matrix [K
G
]
1
for this increment is evaluated based on the unstressed state of the
bar corresponding to point a. For an elasto-plastic material this
might be constructed using the elastic constitutive matrix [D].
Equation (17) is then solved to determine the nodal displacements
{d}
1
nG
. As the material stiffness is assumed to remain constant,
the load displacement curve follows the straight line
ab on Figure 24. In reality, the stiffness of the
material does not remain constant during this
loading increment and the true solution is
represented by the curved path ab. There is
therefore an error in the predicted displacement
equal to the distance bb, however in the tangent
stiffness approach this error is neglected. The
second increment of load, R
2
, is then applied, with
the incremental global stiffness matrix [K
G
]
2
evaluated using the stresses and strains appropriate
to the end of increment 1, i.e., point b on Figure
24. Solution of Equation (17) then gives the nodal
displacements {d}
2
nG
. The load displacement
curve follows the straight path bc on Figure 24.
This deviates further from the true solution, the
error in the displacements now being equal to the
distance cc. A similar procedure now occurs
when R
3
is applied. The stiffness matrix [K
G
]
3
is
evaluated using the stresses and strains appropriate
to the end of increment 2, i.e., point c on Figure 24. The load displacement curve moves to point d and
again drifts further from the true solution. Clearly, the accuracy of the solution depends on the size of the
load increments. For example, if the increment size were reduced so that more increments were needed to
reach the same accumulated load, the tangent stiffness solution would be nearer to the true solution.
From the above simple example it may be concluded that in order to obtain accurate solutions to strongly
nonlinear problems many small solution increments are required. The results obtained using this method can
drift from the true solution and the stresses can fail to satisfy the constitutive relations. Thus the basic
solution requirements may not be fulfilled. As shown later in this paper, the magnitude of the error is
problem dependent and is affected by the degree of material nonlinearity, the geometry of the problem and
the size of the solution increments used. Unfortunately, in general, it is impossible to predetermine the size
of solution increment required to achieve an acceptable error.
The tangent stiffness method can give particularly inaccurate results when soil behaviour changes from
elastic to plastic or vice versa. For instance, if an element is in an elastic state at the beginning of an
increment, it is assumed to behave elastically over the whole increment. This is incorrect if during the
increment the behaviour becomes plastic and results in an illegal stress state that violates the constitutive
model. Such illegal stress states can also occur for plastic elements if the increment size used is too large,
for example a tensile stress state could be predicted for a constitutive model that cannot sustain tension.
Figure 24 : Application of the tangent stiffness
algorithm to the uniaxial loading of a bar of
nonlinear material
Figure 23 : Uniaxial loading of a bar
This can be a major problem with critical state type models, such as modified Cam clay, which employ a v-
ln

p relationship (v = specific volume, p = mean effective stress), since a tensile value of p cannot be
accommodated. In that case, either the analysis has to be aborted or the stress state has to be modified in
some arbitrary way, which would cause the solution to violate the equilibrium condition and the constitutive
model.
8.3.3 Uniform Compression Of A Mohr-Coulomb Soil
To illustrate some of the above deficiencies, drained
one-dimensional loading of a soil element (i.e., an ideal
oedometer test) is considered. This is shown graphically in
Figure 25. Lateral movements are restrained and the soil
sample is loaded vertically by specifying vertical
movements along its top surface. No side friction is
assumed and therefore the soil experiences uniform stresses
and strains. Consequently, to model this using finite
elements, only a single element is needed. There is no
discretisation error, the finite element program is
essentially used only to integrate the constitutive model over the loading path.
Firstly, it is assumed that the soil behaves according to a linear elastic perfectly plastic Mohr-Coulomb
model with Youngs modulus, E=10000kPa, Poissons ratio, =0.2, cohesion, c=0, angle of shearing
resistance =30
o
and angle of dilation, =30
o
. As the angles of dilation, , and shearing resistance, ,
are the same, the model is associated. For this analysis the yield function and plastic potential are given by:
{ }{ } ( )
( )
0 1
tan

,
_

g p
c
J
k , F
(18)
where
( )
3
sin sin
cos
sin

g , (19)
mean effective stress
( ) 3 /
3 2 1
+ + p
, (20)
deviatoric stress
( ) ( ) ( )
2
1 3
2
3 2
2
2 1
6
1
+ + J
, (21)
Lodes angle
( )
( )
1
1
]
1

,
_


1 2
3
1
tan
3 1
3 2 1


, (22)
and {k} is a vector of hardening or softening parameters, which in turn may depend on the state parameters.
For this material the values of the vector {k} are all zero.
It is also assumed that the soil sample has an initial isotropic stress
v
=
h
=50kPa and that loading is
always sufficiently slow to ensure drained conditions.
Figure 26 shows the stress path in J-p space predicted by a tangent stiffness analysis in which equal
increments of displacement were applied to the top of the sample. Each increment gave an incremental axial
strain
a
=3%. Also shown on the figure is the true solution. This was obtained by noting that initially the
soil is elastic and that it only becomes elasto-plastic when it reaches the Mohr-Coulomb yield curve. In J-p
space it can be shown that the elastic stress path is given by:
Figure 25 : Uniform one-dimensional
compression
( )
( )
( ) ( ) 50 866 . 0
1 3
2 1 3

+

p p p J
i

(23)
The Mohr-Coulomb yield curve is given by Equation (18), which, with the parameter values listed above,
gives:
p J 693 . 0 (24)
Combining Equations (23) and (24)
gives the stress state at which the stress
path reaches the yield surface. This
occurs when J =173kPa and
p=250kPa. It can be shown that this
occurs when the applied axial strain

a
=3.6%. Consequently, the true
solution follows the path abc, where
ab is given by Equation (23) and bc
is given by Equation (24).
Inspection of Figure 26 indicates a
discrepancy between the tangent
stiffness and the true solution, with the
former lying above the latter and
indicating a higher angle of shearing
resistance, . The reason for this
discrepancy can be explained as
follows.
For the first increment of loading the
material constitutive matrix is assumed
to be elastic and the predicted stress
path follows the path ab. Because the applied incremental axial strain is only
a
=3%, this is less than

a
=3.6% which is required to bring the soil to yield at point b. As the soil is assumed to be linear elastic,
the solution for this increment is therefore correct. For the second increment of loading the incremental
global stiffness matrix [K
G
]
2
is based on the stress state at the end of increment 1 (i.e. point b). Since the
soil is elastic here, the elastic constitutive matrix [D] is used again. The stress path now moves to point c.
As the applied strain
a
=(
a
1
+
a
2
)=6% is greater than
a
=3.6%, which is required to bring the soil to
yield at point b, the stress state now lies above the Mohr-Coulomb yield surface. The tangent stiffness
algorithm has overshot the yield surface. For increment three the algorithm realises the soil is plastic at
point c and forms the incremental global stiffness matrix [K
G
]
3
based on the elasto-plastic matrix, [D
ep
],
consistent with the stress state at c. The stress path then moves to point d. For subsequent increments
the algorithm uses the elasto-plastic constitutive matrix, [D
ep
], and traces the stress path de.
The reason why this part of the curve is straight, with an inclination greater than the correct solution, path
bc, can be found by inspecting the elasto-plastic constitutive matrix, [D
ep
], defined by Equation (25). In
this expression F is the yield surface, P is the plastic potential, {m} is a vector of state parameters, the values
of which are immaterial, and the parameter A reflects the influence of hardening or softening on the
incremental stress-strain response. For perfect plasticity A=0.
[ ] [ ]
[ ]
{ }{ } ( ) { }{} ( )
[ ]
{ }{} ( )
[ ]
{ }{ } ( )
A
m P
D
k F
D
k F m P
D
D D
T
T
ep
+

'

'

'

'

, ,
, ,
(25)
For the current model the elastic [D] matrix is constant, and as the yield and plastic potential functions
are both assumed to be given by Equation (18), the variation of [D
ep
] depends on the values of the partial
differentials of the yield function with respect to the stress components. In this respect, it can be shown that
the gradient of the stress path in J-p space is given by:
Figure 26 : Oedometer stress path predicted by the tangent
stiffness algorithm
{ }{ } ( )
{ }{ } ( ) p
J
J
k F
p
k F

'

'



,
,

(26)
As this ratio is first evaluated at point
c, which is above the Mohr-Coulomb
yield curve, the stress path sets off at the
wrong gradient for increment three. This
error remains for all subsequent
increments.
The error in the tangent stiffness
approach can therefore be associated with
the overshoot at increment 2. If the
increment sizes had been selected such that
at the end of an increment the stress path
just reached the yield surface (i.e., at point
b), the tangent stiffness algorithm would
then give the correct solution. This is
shown in Figure 27 where, in analysis
labelled A, the first increment was selected
such that
a
=3.6%. After this increment
the stress state was at point b, which is
correct, and for increment 2 and
subsequent increments the correct elasto-
plastic constitutive matrix [D
ep
] was used to obtain the incremental stiffness matrix [K
G
]
i
. As the matrix
[D
ep
] remains constant along the stress path bc, the solution is independent of the size of the increments
from point b onwards. In analysis labelled B in Figure 27, a much larger first increment,
a
=10%, was
applied. This causes a large overshoot on the first increment and results in a significant divergence from the
true solution even if the subsequent increments are reduced to 1%. As noted above, once the stress state has
overshot, making subsequent increments smaller does not improve the solution.
It can be concluded that, for this particular problem, the tangent stiffness algorithm is always in error,
unless the increment size is such that at the end of an increment the stress state happens to be at point b.
Because the solution to this simple one dimensional problem is known, it can be arranged for this to occur,
as for analysis A in Figure 27. However, in general multi-axis boundary value problems, the answers to
which are not known, it is impossible to choose the correct increment sizes so that overshoot never occurs.
The only solution is to use a very large number of small load increments and hope for the best.
Another source of error arising from the way the tangent stiffness method works is that the answers
depend on the way the yield function is implemented. While it is perfectly acceptable, from a mathematical
point of view, to write the yield surface in either of the forms shown in Equations (27) or (18), the
predictions from the tangent stiffness algorithm will differ if, as is usually the case, overshoot occurs.
{ }{} ( ) ( ) 0
tan
,

,
_

g p
c
J k F
(27)
For the simple oedometer situation, this can be seen by calculating the partial differentials in Equation
(26), for the yield function given by Equation (27). This gives:
{ }{ } ( )
{ }{ } ( )
( )

g
J
k F
p
k F

'

'



,
,
(28)
Figure 27 : Effect of the first increment size on a tangent
stiffness prediction of an oedometer stress path
As noted above, this equation gives the
gradient of the resulting stress path in J-p space.
Whereas Equation (26) indicates that the
inclination depends on the amount of overshoot,
Equation (28) indicates that the inclination is
constant and equal to the gradient of the Mohr-
Coulomb yield curve. The two results are
compared in Figure 28. The stress path based on
the yield function written in the form of Equation
(18) appears to pass through the origin of stress
space, but to have an incorrect slope, indicating a
value of that is too high, but the correct value of
c. In contrast, the stress path based on the yield
function written in the form of Equation (27) is
parallel to the true solution, but does not pass
through the origin of stress space, indicating that
the material has a fictitious c, but the correct .
Clearly, if there is no overshoot, both formulations give the same result, which for this problem agrees with
the true solution. The reason for this inconsistency is that, in theory, the differentials of the yield function
are only valid if the stress state is on the yield surface, i.e., F( {},{k})=0. If it is not, it is then
theoretically incorrect to use the differentials and inconsistencies will arise. The implications for practice
are self-evident. Two different pieces of software which purport to use the same Mohr-Coulomb condition
can give very different results, depending on the finer details of their implementation. This is clearly yet
another draw back with the tangent stiffness algorithm for nonlinear analysis.
The analysis labelled A in Figure 27 was performed with a first increment of axial strain of
a
=3.6%
and subsequent increments of
a
=1%. As the first increment just brought the stress path to the yield
surface, the results from this analysis are in
agreement with the true solution. The situation is
now considered where, after being loaded to point
c, see Figure 27, the soil sample is unloaded with
two increments of
a
=-1%. The results of this
analysis are shown in Figure 29. The predicted
stress path on unloading is given by path cde,
which indicates that the soil remains plastic and the
stress path stays on the yield surface. This is
clearly incorrect as such behaviour violates the
basic postulates of elasto-plastic theory. When
unloaded, the soil sample should become purely
elastic, and the correct stress path is marked as path
cf on Figure 29. Because the soil has constant
elastic parameters, this path is parallel to the initial
elastic loading path ab. The reason for the error
in the tangent stiffness analysis arises from the fact
that when the first increment of unloading occurs,
the stress state is plastic, i.e., point c. The algorithm does not know that unloading is going to occur, so
when it forms the incremental global stiffness matrix, it uses the elasto-plastic constitutive matrix [D
ep
]. The
result is that the stress path remains on the yield surface after application of the unloading increment. Since
the soil is still on the yield surface, the same procedure occurs for the second increment of unloading.
8.3.4 Uniform Compression Of A Modified Cam Clay Soil
The above one-dimensional loading problem is now repeated with the soil represented by a simplified
form of the modified Cam clay model. The soil parameters are listed in Table 6.
Because a constant value of the critical state stress ratio in J-p space, M
J
, has been used, the yield (and
plastic potential) surface plots as a circle in the deviatoric plane. A further simplification has been made for
Figure 29 : Example of an unloading stress path
using the tangent stiffness algorithm
Figure 28 : Effect of yield function implementation
on errors associated with the tangent stiffness
algorithm
(18)
(27)
Table 6. Properties for modified Cam clay model
Specific volume at unit pressure on virgin
consolidation line, v
1
1.788
Slope of virgin consolidation line in v-lnp space,
0.066
Slope of swelling line in v-lnp space, 0.0077
Slope of critical state line in J-p space, M
J
0.693
the present analysis. Instead of
using the slope of the swelling
line to calculate the elastic bulk
modulus, constant elastic
parameters, E=50000kPa and
=0.26, have been used. This
simplification has been made to
be consistent with results
presented in the next section of
this paper. For the present
investigation, it does not
si gni fi cant l y affect soi l
behaviour and therefore any conclusions reached are valid for the full model. Again the initial stresses are

v
=
h
=50kPa, and the soil is assumed to be normally consolidated. This later assumption implies that
the initial isotropic stress state is on the yield surface.
Three tangent stiffness analyses, with displacement controlled loading increments equivalent to

a
=0.1%,
a
=0.4% and
a
=1% respectively, have been performed. The predicted stress paths are
shown in Figure 30. Also shown in this figure is the true solution. Consider the analysis with the smallest
increment size,
a
=0.1%. Apart from the very first increment the results of this analysis agree with the
true solution. This is not so for the other two analyses. For the analysis with
a
=0.4% the stress path is in
considerable error for the first three
increments. Subsequently the stress
path is parallel to the true solution,
however, there is still a substantial
error. Matters are even worse for the
analysis with the largest increment
size,
a
=1%. This has very large
errors initially.
The reason for the errors in these
analyses is the same as that explained
above for the Mohr-Coulomb analysis.
That is the yield (and plastic potential)
derivatives are evaluated in illegal
stress space, i.e., with stress values
which do not satisfy the yield (or
plastic potential) function. This is
mathematically wrong and leads to
incorrect elasto-plastic constitutive
matrices. The reason why the errors
are much greater than for the Mohr-
Coulomb analyses is that the yield
(and plastic potential) derivatives are not constant on the yield (or plastic potential) surface, as they are with
the Mohr-Coulomb model, but vary. Matters are also not helped by the fact that the model is strain
hardening/softening and, once the analysis goes wrong, incorrect plastic strains and hardening/softening
parameters are subsequently calculated.
The comments made above for the Mohr-Coulomb model on implementation of the yield function and on
unloading also apply here. In fact, they apply to any constitutive model because they are caused by flaws in
the tangent stiffness algorithm itself.
Figure 30 : Effect of increment size on the tangent stiffness
prediction of an oedometer stress path
8.4 Visco-plastic Method
8.4.1 Introduction
This method uses the equations of visco-plastic behaviour and
time as an artifice to calculate the behaviour of nonlinear, elasto-
plastic, time independent materials (Owen and Hinton, 1980;
Zienkiewicz and Cormeau, 1974). The method was originally
developed for linear elastic visco-plastic (i.e., time dependent)
material behaviour. Such a material can be represented by a
network of the simple rheological units shown in Figure 31. Each
unit consists of an elastic and a visco-plastic component connected
in series. The elastic component is represented by a spring and the
visco-plastic component by a slider and dashpot connected in
parallel. If a load is applied to the network, then one of two
situations occurs in each individual unit. If the load is such that the
induced stress in the unit does not cause yielding, the slider remains
rigid and all the deformation occurs in the spring. This represents
elastic behaviour. Alternatively, if the induced stress causes
yielding, the slider becomes free and the dashpot is activated. As
the dashpot takes time to react, initially all deformation occurs in
the spring. However, with time the dashpot moves. The rate of
movement of the dashpot depends on the stress it supports and its
fluidity. With time progressing, the dashpot moves at a decreasing rate, because some of the stress the unit
is carrying is dissipated to adjacent units in the network, which as a result suffer further movements
themselves. This represents visco-plastic behaviour. Eventually, a stationary condition is reached where all
the dashpots in the network stop moving and are no longer sustaining stresses. This occurs when the stress
in each unit drops below the yield surface and the slider becomes rigid. The external load is now supported
purely by the springs within the network, but, importantly, straining of the system has occurred not only due
to compression or extension of the springs, but also due to movement of the dashpots. If the load was now
removed, only the displacements (strains) occurring in the springs would be recoverable, the dashpot
displacements (strains) being permanent.
8.4.2 Numerical Implementation
Application to finite element analysis of elasto-plastic materials can be summarised as follows. On
application of a solution increment the system is assumed to instantaneously behave linear elastically. If the
resulting stress state lies within the yield surface, the incremental behaviour is elastic and the calculated
displacements are correct. If the resulting stress state violates yield, the stress state can only be sustained
momentarily and visco-plastic straining occurs. The magnitude of the visco-plastic strain rate is determined
by the value of the yield function, which is a measure of the degree by which the current stress state exceeds
the yield condition. The visco-plastic strains increase with time, causing the material to relax with a
reduction in the yield function and hence the visco-plastic strain rate. A marching technique is used to step
forward in time until the visco-plastic strain rate is insignificant. At this point, the accumulated visco-plastic
strain and the associated stress change are equal to the incremental plastic strain and stress change
respectively. This process is illustrated for the simple problem of a uniaxially loaded bar of nonlinear
material in Figure 32.
For genuine visco-plastic materials the visco-plastic strain rate is given by:
{ } { }{} ( ) { }{ } ( )
{ }

,
_

m P
F
k F
f
t
o
vp
, ,
(29)
where is the dashpot fluidity parameter and F
o
is a stress scalar to non-dimensionalise F({},{k})
(Zienkiewicz and Cormeau, 1974). When the method is applied to time independent elasto-plastic materials,
both and F
o
can be assumed to be unity (Griffiths, 1980).
Figure 31 : Rheological model for
visco-plastic material
In order to use the procedure described above,
a suitable time step, t, must be selected. If t is
small many iterations are required to obtain an
accurate solution. However, if t is too large
numerical instability can occur. The most
economical choice for t is the largest value that
can be tolerated without causing such instability.
An estimate for this critical time step is suggested
by Stolle and Higgins (1989).
Due to its simplicity, the visco-plastic
algorithm has been widely used. However, the
method has severe limitations for geotechnical
analysis. Firstly, the algorithm relies on the fact
that for each increment the elastic parameters
remain constant. The simple algorithm cannot
accommodate elastic parameters that vary during
the increment because, for such cases, it cannot
determine the true elastic stress changes
associated with the incremental elastic strains.
The best that can be done is to use the elastic
parameters associated with the accumulated
stresses and strains at the beginning of the increment to calculate the elastic constitutive matrix, [D], and
assume that this remains constant for the increment. Such a procedure only yields accurate results if the
increments are small or the elastic nonlinearity is not great.
A more severe limitation of the method arises when the algorithm is used as an artifice to solve problems
involving non-viscous material (i.e., elasto-plastic materials). As noted above, the visco-plastic strains are
calculated using Equations (29). In Equation (29) the partial differentials of the plastic potential are
evaluated at an illegal stress state {}
t
, which lies outside the yield surface, i.e., F({},{k})>0. As noted for
the tangent stiffness method,
this is theoretically incorrect
and results in failure to satisfy
the constitutive equations.
The magnitude of the error
depends on the constitutive
model and in particular on
how sensitive the partial
derivatives are to the stress
state. This is now illustrated
by applying the visco-plastic
algorithm to the one-
dimensional loading problem
(i.e., ideal oedometer test)
considered above for the
tangent stiffness method.
8.4.3 Uniform Compression
Of A Mohr-Coulomb
Soil
As with the tangent
stiffness method, the problem shown graphically in Figure 25, with the soil properties given in the previous
Mohr-Coulomb example is considered. Figure 33 shows the stress path in J-p space predicted by a visco-
plastic analysis in which equal increments of vertical displacement were applied to the top of the sample.
Each increment gave an axial strain
a
=3% and therefore the predictions in Figure 33 are directly
comparable to those for the tangent stiffness method given in Figure 26. The results were obtained using the
critical time step. It can be seen that the visco-plastic predictions are in remarkably good agreement with the
Figure 32 : Application of the visco-plastic algorithm
to the uniaxial loading of a bar of a nonlinear material
Figure 33 : Oedometer stress path predicted by the visco-plastic algorithm
true solution. Even when the increment size was doubled (i.e.,
a
=6%), the predictions did not change
significantly. Due to the problem highlighted above, concerning evaluation of the plastic potential
differentials in illegal stress space, there were some small differences, but these only caused changes in the
fourth significant figure for both stress and plastic strains. Predictions were also insensitive to values of the
time step as long as it was not greater than the critical value.
The results were therefore not significantly dependent on either the solution increment size or the time
step. The algorithm was also able to deal accurately with the change from purely elastic to elasto-plastic
behaviour and vice versa. In these respects the algorithm behaved much better than the tangent stiffness
method.
It can therefore be concluded that the visco-plastic algorithm works well for this one-dimensional loading
problem with the Mohr-Coulomb model. It has also been found that it works well for other boundary value
problems, involving either the Tresca or the Mohr-Coulomb model.
8.4.4 Uniform Compression Of A Modified Cam Clay Soil
The one-dimensional loading problem was repeated with the soil represented by the simplified modified
Cam clay model described previously. This model has linear elastic behaviour and therefore the problem of
dealing with nonlinear elasticity is not relevant. In fact, it was because of this deficiency in the visco-plastic
algorithm that the model was simplified. The soil properties are given in Table 6 and the initial conditions
are discussed above.
Results from four visco-plastic analyses, with displacement controlled loading increments equivalent to

a
=0.01%,
a
=0.1%,
a
=0.4% and
a
=1%, are compared with the true solution in Figure 34. The
results have been obtained using the critical time step and the convergence criteria was set such that the
iteration process stopped when there was no change in the fourth significant figure of the incremental
stresses and incremental plastic strains.
Only the solution with the smallest increment size (i.e.,
a
=0.01%) agrees with the true solution. It is
instructive to compare these results with those given in Figure 30 for the tangent stiffness method. In view
of the accuracy of the analysis with the Mohr-Coulomb model, it is, perhaps, surprising that the visco-plastic
algorithm requires smaller increments than the tangent stiffness method to obtain an accurate solution. It is
also of interest to note that when the increment size is too large, the tangent stiffness predictions lie above
the true solution, whereas for the visco-plastic analyses the opposite occurs, with the predictions lying below
the true solution. The visco-plastic solutions are particularly in error during the early stages of loading, see
Figure 34b.
To explain why the visco-plastic solutions are in error, consider the results shown in Figure 35. The true
solution is marked as a dashed line on this plot. A visco-plastic analysis consisting of a single increment,
equivalent to
a
=1%, is performed starting from point a, which is on the true stress path. To do this in
the analysis, the initial stresses are set appropriate to point a:
v
=535.7kPa and
h
=343.8kPa. This
loading increment should move the stress path from point a to point e. The line ae therefore represents
Figure 34 : Effect of increment size on the visco-plastic prediction of an oedometer stress path
the true solution to which the visco-plastic
analysis can be compared. However, the visco-
plastic analysis actually moves the stress path
from point a to point d, thus incurring a
substantial error.
To see how such an error arises, the
intermediate steps involved in the visco-plastic
algorithm are plotted in Figure 35. These can
be explained as follows. Initially, on the first
iteration, the visco-plastic strains are zero and
the stress change is assumed to be entirely
elastic. This is represented by the stress state at
point b. This stress state is used to evaluate
the first contribution to the incremental visco-
plastic strains. These strains are therefore
based on the normal to the plastic potential
function at b. This normal is shown on Figure 35 and should be compared to that shown for point a,
which provides the correct solution. As the directions of the normals differ significantly, the resulting
contribution to the visco-plastic strains is in error (note: along path ae of the true solution, the normal to the
plastic potential does not change significantly, being very similar to that at point a). This contribution to
the visco-plastic strains is used to calculate a correction vector. They are also used to update the
hardening/softening parameter for the constitutive model. A second iteration is performed which, due to the
correction vector, gives different incremental displacements and incremental total strains. The incremental
stresses are also now different as they depend on these new incremental total strains and the visco-plastic
strains calculated for iteration 1. The stress state is now represented by point c. A second contribution to
the visco-plastic strains is calculated based on the plastic potential at point c. Again, this is in error because
this is an illegal stress state. The error is related to the difference in direction of the normals to the plastic
potential surfaces at points a and c. This second contribution to the incremental plastic strains is used to
obtain an additional correction vector. A third iteration is then performed which brings the stress state to
point d on Figure 35. Subsequent iterations cause only very small changes to the visco-plastic strains and
the incremental stresses and therefore the stress state remains at point d. At the end of the iterative
process, the incremental plastic strains are equated to the visco-plastic strains. As the visco-plastic strains
are the sum of the contributions obtained from each iteration, and as each of these contributions has been
calculated using the incorrect plastic potential differentials (i.e., wrong direction of the normal), the
incremental plastic strains are in error. This is evident from Figure 36, which compares the predicted and
true incremental plastic strains. Since the
hardening parameter for the model is calculated
from the plastic strains, this is also incorrect. It
is therefore not surprising that the algorithm
ends up giving the wrong stress state
represented by point d in Figure 35.
If the soil sample is unloaded at any stage,
the analysis indicates elastic behaviour and
therefore behaves correctly.
It is concluded that for complex critical state
constitutive models the visco-plastic algorithm
can involve severe errors. The magnitude of
these errors depends on the finer details of the
model and, in particular, on how rapidly the
plastic potential differentials vary with changes
in stress state. The problems associated with
the implementation of a particular constitutive
model, as discussed for the tangent stiffness
method, also apply here. As the plastic strains
are calculated from plastic potential differentials evaluated in illegal stress space, the answers depend on the
Figure 36 : Comparison of incremental plastic strains
from a single increment of a visco-plastic analysis and
the true solution
Figure 35 : A single increment of a visco-plastic
analysis
finer details of how the model is implemented in the software. Again, two pieces of software which purport
to use the same equations could give different results.
The above conclusion is perhaps surprising as the visco-plastic algorithm appears to work well for simple
constitutive models of the Tresca and Mohr-Coulomb types. However, as noted previously, in these simpler
models the differentials of the plastic potential do not vary by a great amount when the stress state moves
into illegal stress space.
8.5 Modified Newton-Raphson Method
8.5.1 Introduction
The previous discussion of both the tangent stiffness and visco-plastic algorithms has demonstrated that
errors can arise when the constitutive behaviour is based on illegal stress states. The modified Newton-
Raphson (MNR) algorithm described in this section attempts to rectify this problem by only evaluating the
constitutive behaviour in, or very near to, legal stress space.
The MNR method uses an iterative technique to solve Equation (17). The first iteration is essentially the
same as the tangent stiffness method. However, it is recognised that the solution is likely to be in error and
the predicted incremental displacements are used to calculate the residual load, a measure of the error in the
analysis. Equation (17) is then solved again with this residual load, { }, forming the incremental right
hand side vector. Equation (17) can be rewritten as:
[ ] { } ( ) { }
1

j
j
i
nG
i
G
d K
(30)
The superscript j refers to the iteration number and
{ }
0
={R
G
}
i
. This process is repeated until the
residual load is small. The incremental displacements
are equal to the sum of the iterative displacements.
This approach is illustrated in Figure 37 for the simple
problem of a uniaxially loaded bar of nonlinear
material. In principle, the iterative scheme ensures that
for each solution increment the analysis satisfies all
solution requirements.
A key step in this calculation process is determining
the residual load vector. At the end of each iteration the
current estimate of the incremental displacements is
calculated and used to evaluate the incremental strains
at each integration point. The constitutive model is
then integrated along the incremental strain paths to
obtain an estimate of the stress changes. These stress
changes are added to the stresses at the beginning of the
increment and used to evaluate consistent equivalent
nodal forces. The difference between these forces and
the externally applied loads (from the boundary
conditions) gives the residual load vector. A difference
arises because a constant incremental global stiffness
matrix [K
G
]
i
is assumed over the increment. Due to the
nonlinear material behaviour, [K
G
]
i
is not constant but varies with the incremental stress and strain changes.
Since the constitutive behaviour changes over the increment, care must be taken when integrating the
constitutive equations to obtain the stress change. Methods of performing this integration are termed stress
point algorithms and both explicit and implicit approaches have been proposed in the literature. There are
many of these algorithms in use and, as they control the accuracy of the final solution, users must verify the
approach used in their software. Two of the most accurate stress point algorithms are described
subsequently.
The process described above is called a Newton-Raphson scheme if the incremental global stiffness
matrix [K
G
]
i
is recalculated and inverted for each iteration, based on the latest estimate of the stresses and
Figure 37 : Application of the modified Newton-
Raphson algorithm to the uniaxial loading of a
bar of linear material
strains obtained from the previous iteration. To reduce the amount of computation, the modified Newton-
Raphson method only calculates and inverts the stiffness matrix at the beginning of the increment and uses it
for all iterations within the increment. Sometimes the incremental global stiffness matrix is calculated using
the elastic constitutive matrix, [D], rather than the elasto-plastic matrix, [D
ep
]. Clearly, there are several
options here and many software packages allow the user to specify how the MNR algorithm should work. In
addition, an acceleration technique is often applied during the iteration process (Thomas, 1984).
8.5.2 Stress Point Algorithms
Two classes of stress point algorithms are considered. The substepping algorithm is essentially explicit,
whereas the return algorithm is implicit. In both the substepping and return algorithms, the objective is to
integrate the constitutive equations along an incremental strain path. While the magnitudes of the strain
increment are known, the manner in which they vary during the increment is not. It is therefore not possible
to integrate the constitutive equations without making an additional assumption. Each stress point algorithm
makes a different assumption and this influences the accuracy of the solution obtained.
The schemes presented by Wissman and Hauck (1983) and Sloan (1987) are examples of substepping
stress point algorithms. In this approach, the incremental strains are divided into a number of substeps. It is
assumed that in each substep the strains {
ss
} are a proportion, T, of the incremental strains {
inc
}. This
can be expressed as:
{ } { }
inc ss
T
(31)
It should be noted that in each substep, the ratio between the strain components is the same as that for the
incremental strains and hence the strains are said to vary proportionally over the increment. The constitutive
equations are then integrated numerically over each substep using either an Euler, modified Euler or Runge-
Kutta scheme. The size of each substep (i.e., T) can vary and, in the more sophisticated schemes, is
determined by setting an error tolerance on the numerical integration. This allows control of errors resulting
from the numerical integration procedure and ensures that they are negligible.
The basic assumption in these substepping approaches is therefore that the strains vary in a proportional
manner over the increment. In some boundary value problems, this assumption is correct and consequently
the solutions are extremely accurate. However, in general, this may not be true and an error can be
introduced. The magnitude of the error is dependent on the size of the solution increment.
The schemes presented by Borja and Lee (1990) and Borja (1991) are examples of one-step implicit type
return algorithms. In this approach, the plastic strains over the increment are calculated from the stress
conditions corresponding to the end of the increment. The problem, of course, is that these stress conditions
are not known, hence the implicit nature of the scheme. Most formulations involve some form of elastic
predictor to give a first estimate of the stress changes, coupled with a sophisticated iterative sub-algorithm to
transfer from this stress state back to the yield surface. The objective of the iterative sub-algorithm is to
ensure that, on convergence, the constitutive behaviour is satisfied, albeit with the assumption that the plastic
strains over the increment are based on the plastic potential at the end of the increment. Many different
iterative sub-algorithms have been proposed in the literature. In view of the findings presented previously, it
is important that the final converged solution does not depend on quantities evaluated in illegal stress space.
In this respect some of the earlier return algorithms broke this rule and are therefore inaccurate.
The basic assumption in these approaches is therefore that the plastic strains over the increment can be
calculated from the stress state at the end of the increment. This is theoretically incorrect as the plastic
response, and in particular the plastic flow direction, is a function of the current stress state. The plastic flow
direction should be consistent with the stress state at the beginning of the solution increment and should
evolve as a function of the changing stress state, such that at the end of the increment it is consistent with the
final stress state. This type of behaviour is exemplified by the substepping approach. If the plastic flow
direction does not change over an increment, the return algorithm solutions are accurate. Invariably,
however, this is not the case and an error is introduced. The magnitude of any error is dependent on the size
of the solution increment.
Potts and Ganendra (1994) performed a fundamental comparison of these two types of stress point
algorithm. They conclude that both algorithms give accurate results, but, of the two, the substepping
algorithm is better.
Another advantage of the substepping approach is that it is extremely robust and can easily deal with
constitutive models in which two or more yield surfaces are active simultaneously and for which the elastic
portion of the model is highly nonlinear. In fact, most of the software required to program the algorithm is
common to any constitutive model. This is not so for the return algorithm, which, although in theory can
accommodate such complex constitutive models, involves some extremely complicated mathematics. The
software to deal with the algorithm is also constitutive model dependent. This means considerable effort is
required to include a new or modified model.
As the MNR method involves iterations for each solution increment, convergence criteria must be set.
This usually involves setting limits to the size of both the iterative displacements, ({d}
i
nG
)
j
, and the residual
loads, {

}
j
. As both these quantities are vectors, it is normal to express their size in terms of the scalar
norms.
8.5.3 Uniform Compression Of Mohr-Coulomb And Modified Cam Clay Soils
The MNR method using a substepping stress point algorithm has been used to analyse the simple one
dimensional oedometer problem, considered previously for both the tangent stiffness and visco-plastic
approaches. Results are presented in Figures 38a and 38b for the Mohr-Coulomb and modified Cam clay
soils, respectively. To be consistent with the analyses performed with the tangent and visco-plastic
algorithm, the analysis for the Mohr-Coulomb soil involved displacement increments which gave
incremental strains
a
=3%, whereas for the modified Cam clay analysis the increment size was equivalent
to
a
=1%.
The predictions are in excellent agreement with the true solution. An unload-reload loop is shown in
each figure, indicating that the MNR approach can accurately deal with changes in stress path direction. For
the modified Cam clay analysis it should be noted that at the beginning of the test the soil sample was
normally consolidated, with an isotropic stress p=50kPa. The initial stress path is therefore elasto-plastic
and not elastic. Consequently, it is not parallel to the unload/reload path. Additional analysis, performed
with different sizes of solution increment, indicate that the predictions, for all practical purposes, are
independent of increment size.
These results clearly show that, for this simple problem, the MNR approach does not suffer from the
inaccuracies inherent in both the tangent stiffness and visco-plastic approaches. To investigate how the
different methods perform for more complex boundary value problems, a small parametric study has been
performed, see Potts and Zdravkovic (1999). Some of the main findings of this study are presented next.
Figure 38 : Oedometer stress paths predicted by the MNR algorithm: a) Mohr-Coulomb and b) modified
Cam Clay models
Table 7. Modified Cam clay properties for drained triaxial tests
Overconsolidation ratio 1.0
Specific volume at unit pressure on virgin consolidation
line, v
1
1.788
Slope of virgin consolidation line in v-lnp space, 0.066
Slope of swelling line in v-lnp space, 0.0077
Slope of critical state line in J-p space, M
J
0.693
Elastic shear modulus, G/Preconsolidation pressure, p
0
100
8.6 Comparison Of Solution Strategies
8.6.1 Introduction
A comparison of the three solution strategies presented above suggests the following. The tangent
stiffness method is the simplest, but its accuracy is influenced by increment size. The accuracy of the visco-
plastic approach is also influenced by increment size, if complex constitutive models are used. The MNR
method is potentially the most accurate and is likely to be the least sensitive to increment size. However,
considering the computer resources required for each solution increment, the MNR method is likely to be the
most expensive, the tangent stiffness method the cheapest and the visco-plastic method is probably
somewhere in between. It may be possible though, to use larger and therefore fewer increments with the
MNR method to obtain a similar accuracy. Thus, it is not obvious which solution strategy is the most
economic for a particular solution accuracy.
All three solution algorithms have been incorporated into the single computer program, ICFEP.
Consequently, much of the computer code is common to all analyses and any difference in the results can be
attributed to the different solution strategies. The code has been extensively tested against available
analytical solutions and with other computer codes, where applicable. The program was used to compare the
relative performance of each of the three schemes in the analysis of two simple idealised laboratory tests and
three more complex boundary value problems. Analyses of the laboratory tests were carried out using a
single four-noded isoparametric element with a single integration point, whereas eight-noded isoparametric
elements, with reduced integration, were employed for the analyses of the boundary value problems.
As already shown, the errors in the solution algorithms are more pronounced for critical state type models
than for the simpler linear elastic perfectly plastic models (i.e., Mohr-Coulomb and Tresca). Hence, in the
comparative study the soil has been modelled with the modified Cam clay model. To account for the
nonlinear elasticity that is present in this model, the visco-plastic algorithm was modified to incorporate an
additional stress correction based on an explicit stress point algorithm, similar to that used in the MNR
method, at each time step, see Potts and Zdravkovic (1999).
8.6.2 Ideal Drained Triaxial
Test
Idealised drained triaxial
compr essi on t est s wer e
considered. A cylindrical
sample was assumed to be
i s o t r o p i c a l l y n o r ma l l y
consolidated to a mean effective
stress, p=200kPa, with zero
pore water pressure. The soil
parameters used for the analyses
are shown in Table 7.
Increments of compressive
axial strain were applied to the
sample until the axial strain reached 20%, while maintaining a constant radial stress and zero pore water
pressure. The results are presented as plots of volumetric strain and deviatoric stress, q, versus axial strain.
Deviatoric stress q is defined as:
( ) ( ) ( ) [ ]
conditions triaxial for
2
1
3
3 1
2
1 3
2
3 2
2
2 1



+ + J q
(32)
Results of these analyses are presented in Figures 39, 40 and 41. The label associated with each line in
these plots indicates the magnitude of axial strain applied at each increment of that analysis. The tests were
deemed ideal as the end effects at the top and bottom of the sample were considered negligible and the stress
and strain conditions were uniform throughout. Analytical solutions are also provided on the plots for
comparison purposes.
Results from the MNR analyses are compared with
the analytical solution in Figure 39. The results are
not sensitive to increment size and agree well with the
analytical solution. The tangent stiffness results are
presented in Figure 40. The results are sensitive to
increment size, giving very large errors for the larger
increment sizes. The deviatoric stress, q at failure
(20% axial strain) is over predicted. The results of the
visco-plastic analyses are shown in Figure 41.
Inspection of this figure indicates that the solution is
also sensitive to increment size. Even the results from
the analyses with the smallest increment size of 0.1%
are in considerable error.
It is of interest to note that results from the tangent
stiffness analyses over predict the deviatoric stress at
any particular value of axial strain. The opposite is
true for the visco-plastic analysis, where q is under
predicted in all cases. This is similar to the
observations made earlier for the simple oedometer
test.
8.6.3 Footing Problem
A smooth rigid strip footing subjected to vertical loading, as depicted in Figure 42, has been analysed.
The same soil constitutive model and parameters as used for the idealised triaxial test analyses, see Table 7,
have been employed to model the soil which, in this case, was assumed to behave undrained. The finite
element mesh is shown in Figure 43. Note that due to symmetry about the vertical line through the centre of
the footing, only half of the problem needs to be considered in the finite element analysis. Plane strain
conditions are assumed. Before loading the footing, the coefficient of earth pressure at rest, K
o
, was
assumed to be unity and the vertical effective stress and pore water pressure were calculated using a
saturated bulk unit weight of the soil of 20kN/m
3
and a static water table at the ground surface. The footing
Figure 41 : Visco-plastic: Drained triaxial test
Figure 40 : Tangent stiffness: Drained triaxial test Figure 39 : Modified Newton-Raphson: Drained
triaxial test
was loaded by applying a series of equally sized increments of vertical displacement until the total
displacement was 25mm.
The load-displacement curves for the tangent stiffness, visco-plastic and MNR analyses are presented in
Figure 44. For the MNR method, analyses were performed using 1, 2, 5, 10, 25, 50 and 500 increments to
reach a footing settlement of 25mm. With the exception of the analysis performed with only a single
increment, all analyses gave very similar results and plot as a single curve, marked MNR on this figure. The
MNR results are therefore insensitive to increment size and show a well-defined collapse load of 2.8kN/m.
For the tangent stiffness approach, analyses using 25, 50, 100, 200, 500 and 1000 increments have been
carried out. Analyses with a smaller number of increments were also attempted, but illegal stresses
(negative mean effective stresses, p) were predicted. As the constitutive model is not defined for such
stresses, the analyses had to be aborted. Some finite element packages overcome this problem by arbitrarily
resetting the offending negative p values. There is no theoretical basis for this and it leads to violation of
both the equilibrium and the constitutive conditions. Although such adjustments enable an analysis to be
completed, the final solution is in error.
Results from the tangent stiffness analyses are shown in Figure 44. When plotted, the curve from the
analysis with 1000 increments is indistinguishable from those of the MNR analyses. The tangent stiffness
results are strongly influenced by increment size, with the ultimate footing load decreasing from 7.5kN/m to
2.8kN/m with reduction in the
size of the applied displacement
increment. There is also a
t endency for t he l oad-
displacement curve to continue
to rise and not reach a well-
defined ultimate failure load for
the analysis with large applied
displacement increments. The
results are unconservative, over
predicting the ultimate footing
load. There is also no indication
from the shape of the tangent
stiffness load-displacement
curves as to whether the solution
is accurate, since all the curves
have similar shapes. Figure 44 : Footing load-displacement curves
Figure 43 : Finite element mesh for footing analysis
Figure 42 : Geometry of
footing problem
Visco-plastic analyses with 10, 25, 50, 100 and
500 increments were performed. The 10 increment
analysis had convergence problems in the iteration
process, which would initially converge, but then
diverge. Similar behaviour was encountered for
analyses using still fewer increments. Results from
the analyses with 25 and 500 increments are shown
in Figure 44. The solutions are sensitive to
increment size, but to a lesser degree than the
tangent stiffness approach.
The load on the footing at a settlement of 25mm
is plotted against number of increments, for all
tangent stiffness, visco-plastic and MNR analyses,
in Figure 45. The insensitivity of the MNR analyses
to increment size is clearly shown. In these
analyses the ultimate footing load only changed
from 2.83kN/m to 2.79kN/m as the number of
increments increased from 2 to 500. Even for the
MNR analyses performed with a single increment,
the resulting ultimate footing load of 3.13kN/m is
still reasonable and is more accurate than the value of 3.67kN/m obtained from the tangent stiffness analysis
with 200 increments. Both the tangent stiffness and visco-plastic analyses produce ultimate failure loads
which approach 2.79kN/m as the number of increments increase. However, tangent stiffness analyses
approach this value from above and therefore over predict, while visco-plastic analyses approach this value
from below and therefore under predict. This trend is consistent with the results from the triaxial test, where
the ultimate value of q was over predicted by the tangent stiffness and under predicted by the visco-plastic
approach.
It can be shown that, for the material properties and initial stress conditions adopted, the undrained
strength of the soil, s
u
, varies linearly with depth below the ground surface. Davis and Booker (1973)
provide approximate solutions for the bearing capacity of footings on soils with such an undrained strength
profile. For the present situation their charts give a collapse load of 1.91kN/m and it can be seen (Figure 45)
that all the finite element predictions exceed this value. This occurs because the analytical failure zone is
very localised near the soil surface. Further analyses have been carried out using a refined mesh in which
the thickness of the elements immediately below the footing has been reduced from 0.1m to 0.03m. The
MNR analyses with this mesh predict an ultimate load of 2.1kN/m. If the mesh is further refined the Davis
and Booker solution will be recovered. For this refined mesh even smaller applied displacement increment
sizes were required for the tangent stiffness analyses to obtain an accurate solution. Analyses with an
increment size of 0.125mm displacement (equivalent to the analysis with 200 increments described above)
or greater, yielded negative values of p in the elements below the corner of the footing and therefore these
analyses could not be completed.
8.6.4 Comments
Results from the tangent stiffness analyses of both the idealised triaxial tests and the more complex
boundary value problems are strongly dependent on increment size. The error associated with the tangent
stiffness analyses usually results in unconservative predictions of failure loads and displacements in most
geotechnical problems. For the footing problem large over predictions of failure loads are obtained, unless a
very large number of increments (1000) is employed. Inaccurate analyses based on too large an increment
size produced ostensible plausible load displacement curves. Analytical solutions are not available for most
problems requiring a finite element analysis. Therefore it is difficult to judge whether a tangent stiffness
analysis is accurate on the basis of its results. Several analyses must be carried out using different increment
sizes to establish the likely accuracy of any predictions. This could be a very costly exercise, especially if
there was little experience in the problem being analysed and no indication of the optimum increment size.
Results from the visco-plastic analyses are also dependent on increment size. For boundary value
problems involving undrained soil behaviour these analyses were more accurate than tangent stiffness
analyses with the same increment size. However, if soil behaviour was drained, visco-plastic analyses were
Figure 45 : Ultimate footing load against number of
increments
only accurate if many small solution increments were used. In general, the visco-plastic analyses used more
computer resources than both the tangent stiffness and MNR approaches. For the triaxial tests, footing and
pile problems (not presented here), the visco-plastic analyses under-predicted failure loads if insufficient
increments were used and were therefore conservative in this context.
For both the tangent stiffness and visco-plastic analyses the number of increments to obtain an accurate
solution is problem dependent. For example, for the footing problem the tangent stiffness approach required
over 1000 increments and the visco-plastic method over 500 increments, whereas for an excavation problem
(not presented here) the former required only 100 and the later 10 increments. Close inspection of the
results from the visco-plastic and tangent stiffness analyses indicated that a major reason for their poor
performance was their failure to satisfy the constitutive laws. This problem is largely eliminated in the
MNR approach, where a much tighter constraint on the constitutive conditions is enforced.
The results from the MNR analyses are accurate and essentially independent of increment size. For the
boundary value problems considered, the tangent stiffness method required considerably more CPU time
than the MNR method to obtain results of similar accuracy, e.g., over seven times more for the foundation
problem and over three and a half times more for the excavation problem. Similar comparisons can be
found between the MNR and visco-plastic solutions. Thus not withstanding the potentially very large
computer resources required to find the optimum tangent stiffness or visco-plastic increment size, the
tangent stiffness or visco-plastic method with an optimum increment size is still likely to require more
computer resources than an MNR analysis of the same accuracy. Though it may be possible to obtain
tangent stiffness or visco-plastic results using less computer resources than with the MNR approach, this is
usually at the expense of the accuracy of the results. Alternatively, for a given amount of computing
resources, a MNR analysis produces a more accurate solution than either the tangent stiffness or visco-
plastic approaches.
The study has shown that the MNR method appears to be the most efficient solution strategy for
obtaining an accurate solution to problems using critical state type constitutive models for soil behaviour.
The large errors in the results from the tangent stiffness and visco-plastic algorithms in the present study
emphasise the importance of checking the sensitivity of the results of any finite element analysis to
increment size.
8.7 Using The Mohr-Coulomb Model In Constrained Problems
The Mohr Coulomb model can be used with a dilation angle ranging from =0 to ='. This
parameter controls the magnitude of the plastic dilation (plastic volume expansion) and remains constant
once the stress state of the soil is on the yield surface. This implies that the soil will continue to dilate
indefinitely if shearing continues. Clearly such behaviour is not realistic, as most soils will eventually reach
a critical state condition, after which they will deform at constant volume if sheared any further. While such
unrealistic behaviour does not have a great influence on boundary value problems that are unrestrained (e.g.,
the drained surface footing problem), it can have a major effect on problems which are constrained (e.g.,
drained cavity expansion, end bearing of a deeply embedded pile), due to the restrictions on volume change
imposed by the boundary conditions. In particular, unexpected results can be obtained in undrained analysis
in which there is a severe constraint imposed by the zero total volume change restriction associated with
undrained soil behaviour. To illustrate this problem two examples will now be presented.
The first example considers ideal (no end effects) undrained triaxial compression (
v
>0,
h
=0) tests
on a linear elastic Mohr Coulomb plastic soil with parameters E'=10000kPa, =0.3, c'=0, '=24
o
. As
there are no end effects, a single finite element is used to model the triaxial test with the appropriate
boundary conditions. The samples were assumed to be initially isotopically consolidated with p'=200kPa
and zero pore water pressure. A series of finite element runs were then made, each with a different angle of
dilation, , in which the samples were sheared undrained. Undrained conditions were enforced by setting
the bulk modulus of water to be 1000 times larger than the effective elastic bulk modulus of the soil
skeleton, K'. The results are shown in Figure 46a and 46b in the form of J-p' and J-
z
plots. It can be seen
that in terms of J-p' all analyses follow the same stress path. However, the rate at which the stress state
moves up the Mohr Coulomb failure line differs for each analysis. This can be seen from Figure 46b. The
analysis with zero plastic dilation, =0, remains at a constant J and p' when it reaches the failure line.
However, all other analyses move up the failure line, with those with the larger dilation moving up more
rapidly. They continue to move up this failure line indefinitely with continued shearing. Consequently, the
only analysis that indicates failure (i.e., a limiting value of J) is the analysis performed with zero plastic
dilation.
The second example considers the
undrained loading of a smooth rigid strip
footing. The soil was assumed to have the
same parameters as those used for the triaxial
tests above. The initial stresses in the soil
were calculated on the basis of a saturated
bulk unit weight of 20kN/m
3
, a ground water
table at the soil surface and a K
o
=1-sin'.
The footing was loaded by applying
increments of vertical displacement and
undrained conditions were again enforced by
setting the bulk modulus of the pore water to
be 1000 times K'. The results of two analyses,
one with =0
o
and the other with =', are
shown in Figure 47. The difference is quite
staggering: while the analysis with =0
o
reaches a limit load, the analysis with ='
shows a continuing increase in load with
displacement. As with the triaxial tests, a
limit load is only obtained if =0
o
.
It can be concluded from these two examples that a limit load will only be obtained if =0
o
.
Consequently great care must be exercised when using the Mohr Coulomb model in undrained analysis. It
could be argued that the model should not be used with >0 for such analysis. However, reality is not that
simple and often a finite element analysis involves both an undrained and a drained phase (i.e., undrained
excavation followed by drained dissipation). Consequently it may be necessary to adjust the value of
between the two phases of the analysis. Alternatively, a more complex constitutive model which better
represents soil behaviour may have to be employed.
Figure 46 : Prediction of dilation a) stress paths and b) stress-strain curves in triaxial compression, using
Mohr-Coulomb model with different angles (Note: in this figure)
Figure 47 : Load-displacement curves for strip footing,
Mohr Coulomb model with different angles of dilation

8.8 Influence Of The Shape Of The Yield And Plastic Potential Surfaces
As noted by Potts and Gens (1984) and Potts and Zdravkovic (1999) the shape of the plastic potential in
the deviatoric plane can affect the Lodes angle at failure in plane strain analyses. This implies that it will
affect the value of the soil strength that can be mobilised. In many commercial software packages, the user
has little control over the shape of the plastic
potential and it is therefore important that its
implications are understood. This phenomenon is
investigated by considering the modified Cam-
clay constitutive model.
Many software packages assume that both the
yield and plastic potential surfaces plot as circles
in the deviatoric plane. This is defined by
specifying a constant value of the parameter M
J
(i.e., the slope of the critical state line in J-p
stress space). Such an assumption implies that the
angle of shearing resistance, ', varies with the
Lodes angle, . By equating M
J
to the
expression for g() given by Equation (19) and
re-arranging, gives the following expression for '
in terms of M
J
and :

,
_


3
sin
1
cos

sin
=
1

M

M
J
J
(33)
From this equation it is possible to express M
J
in terms of the angle of shearing resistance, '
TC
, in triaxial
compression (=-30
o
):
TC
TC
TC J M

sin 3
sin 3 2
= ) (
(34)
In Figure 48 the variation of ' with , given by Equation (33), are plotted for three values of M
J
. The
values of M
J
have been determined from Equation (34) using '
TC
=20
o
, 25
o
and 30
o
. If the plastic potential
is circular in the deviatoric plane, it can be shown that plane strain failure occurs when the Lode's angle
=0
o
. Inspection of Figure 48
indicates that for all values of M
J
there
is a large change in ' with . For
example if ' is set to give '
TC
=25
o
,
then under plane strain conditions the
mobilised ' value is '
PS
=34.6
o
. This
difference is considerable and much
larger than indicated by careful
laboratory testing. The differences
between '
TC
and '
PS
becomes greater
the larger the value of M
J
.
The effect of on undrained
strength, s
u
, for the constant M
J
formulation is shown in Figure 49. The
variation has been calculated for
OCR=1, g()=M
J
, K
o
=1-sin'
TC
and
/=0.1. The equivalent variation
based on the formulation which
Figure 49 : Effect of on s
u
, for the constant M
J
formulation
Figure 48 : Variation of with for constant M
J
assumes a constant ', instead of a constant M
J
, is given in Figure 50. This is also based on the above
parameters, except that g() is now given by Equation (18). The variation shown in this figure is in much
better agreement with the available experimental data than the trends shown in Figure 49.
To investigate the effect of the
plastic potential in a boundary
value problem two analyses of a
rough rigid strip footing have been
performed. The finite element
mesh was similar to the one shown
in Figure 43. The modified Cam
clay model was used to represent
the soil which had the following
material parameters, OCR=6,
v
1
=2.848, =0.161, =0.0322
and =0.2. In one analysis the
yield and plastic potential surfaces
were assumed to be circular in the
deviatoric plane. A value of
M
J
=0.5187 was used for this
analysis, this is equivalent to
'
TC
=23
o
. In the second analysis a
constant value of '=23
o
was used
giving a Mohr Coulomb hexagon
for the yield surface in the
deviatoric plane. However, the
plastic potential still gave a circle
in the deviatoric plane and
therefore plane strain failure
occurred at =0
o
, as for the first
analysis. In both analysis the
initial stress conditions in the soil
were based on a saturated bulk unit
weight of 18kN/m
3
, a ground water
table at a depth of 2.5m and a
K
o
=1.227. Above the ground
water table the soil was assumed to
be saturated and able to sustain
pore water pressure suctions.
Coupled consolidation analyses
wer e per f or med but t he
permeability and time steps were
chosen such that undrained
conditions occurred. Loading of
the footing was simulated by
imposing increments of vertical displacement.
In summary, the input to both analysis is identical, except that in the first, the strength parameter M
J
is
specified, whereas in the second, ' is input. In both analyses '
TC
=23
o
and therefore any analyses in
triaxial compression would give identical results. However, the strip footing problem is plane strain and
therefore differences are expected. The resulting load displacement curves are given in Figure 51. The
analysis with a constant M
J
gave a collapse load some 58% larger than the analysis with a constant '. The
implications for practice are clear, if a user is not aware of the plastic potential problem and is not fully
conversant with the constitutive model implemented in the software being used, they could easily base the
input on '
TC
=23
o
. If the model uses a constant M
J
formulation, this would then imply a '
PS
=31.2
o
, which
in turn leads to a large error in the prediction of any collapse load.
Figure 51 : Load-displacement curves for two different approaches
Figure 50 : Effect of on s
u
, for the constant formulation
9.0 VALIDATION AND CALIBRATION OF COMPUTER SIMULATIONS
To date, less attention has been paid in the literature to the important issue of validation and reliability of
numerical models in general, and specific software in particular, than has been paid to the development of
the methods themselves. The work by Schweiger (1991) is one of the limited studies on the subject of
model validation. There is now a strong need to define procedures and guidelines to arrive at reliable
numerical methods and, more importantly, input parameters which represent accurately the strength and
stiffness properties of the ground in situ.
As should be evident from previous discussion, benchmarking is of great importance in geotechnical
engineering, probably more so than in other engineering disciplines, such as structural engineering. Reasons
for the importance of benchmarking may be summarized as follows:
the domain to be analysed is often not clearly defined by the structure,
it is not always clear whether continuum or discontinuum models are more appropriate for the problem
at hand,
a wide variety of constitutive models exists in the literature, but there is no approved model for each
type of soil,
in most cases construction details cannot be modelled very accurately in time and space (e.g., the
excavation sequence, pre-stressing of anchors, etc.), at least not from a practical point of view,
soil-structure interaction is often important and may lead to numerical problems (e.g., with certain types
of interface elements),
the implementation details and solution procedures may have a significant influence on the results of
certain problems, but may not be important for others, and
there are no approved implementation and solution procedures for commercial codes (such as implicit
versus explicit solution strategies, return algorithms, etc.).
Obviously, there is currently considerable scope for developers and users of numerical models to exercise
their personal preferences when tackling geotechnical problems. From a practical point of view, it is
therefore very difficult to prove the validity of many calculated results because of the numerous modelling
assumptions required. So far, no clear guidelines exist, and thus results for a particular problem may vary
significantly if analysed by different users, even for reasonably well-defined working load conditions.
Difficult issues such as these have been addressed by various groups, including a working group of the
German Society for Geotechnics, viz., 1.6 Numerical Methods in Geotechnics. It is the aim of this group
to provide recommendations for numerical analyses in geotechnical engineering. So far the group has
published general recommendations (Meissner, 1991), recommendations for numerical simulations in
tunnelling (Meissner, 1996) and it is expected that recommendations for deep excavations will also be
published shortly. In addition, benchmark examples have been specified and the results obtained by various
users employing different software have been compared. Some of the work presented by this group is
summarised here. The efforts of the working group may be seen as a first step towards greater objectivity in
numerical analyses of geotechnical problems in practice.
To date, three example problems have been specified by the working group, and these have been
discussed in two separate workshops. The first two examples, involving a tunnel excavation and a deep
open excavation problem, have been rather idealised problems, with a very tight specification so that little
room for interpretation was left to the analysts. Despite the simplicity of the examples and the rather strict
specifications, significant differences in the results were obtained, even in cases where the same software
has been utilised by different users. Further details of these examples are given below.
The third example, which represents an actual application (a tied back diaphragm wall in Berlin sand),
which was only slightly modified in order to reduce the computational effort, will also be presented.
Limited field measurements are available for this example, providing information on the order of magnitude
of the deformations to be expected. Whereas in the first two examples, the constitutive model and the
parameters were pre-specified, in the latter example the choice of constitutive model has been left to the user
and the parameter values had to be selected either from the literature, on the basis of personal experience, or
determined from laboratory tests which were made available to the analysts.
9.1 Specifications For Benchmark Examples
Keeping in mind the purpose of benchmarking, from a practical point of view the following requirements
for benchmark examples are suggested:
no analytical solution is available but commercial codes are capable of solving the problem,
the actual practical problem should be addressed, and simplified in such a way that the solution can be
obtained with reasonable computational effort,
no calibration of laboratory tests has been performed (this is done extensively in research and is of minor
interest to engineers in practice),
preferably the examples should be set in such a way that, in addition to global results, specific aspects
can also be checked (e.g., handling of initial stresses, dilation behaviour, excavation procedures, etc.),
the influence of different constitutive models on the predicted results should become apparent, and
the problem of parameter identification for various constitutive models should be addressed.
Once a series of examples has been designed and solutions are available they could serve as:
a check of commercial codes,
learning aids for young geotechnical engineers to help them to become familiar with numerical analysis,
and
verification examples for proving competence
in numerical analysis of geotechnical problems.
In addition, these examples will identify
limitations of the present state of the art in
numerical modelling in practice, provide the
possibility to show alternative modelling
assumptions, and highlight the importance of
appropriate constitutive models.
9.2 Tunnel Excavation Example
Figure 52 depicts the geometry of the first
validation example, a tunnel excavation problem,
and Table 8 contains a list of the material
parameters given to all participants in this exercise.
Additional specifications are as follows.
9.2.1 General Assumptions
plane strain conditions apply,
a linear elastic - perfectly plastic analysis with
the Mohr-Coulomb failure criterion was
required,
perfect bonding existed between the shotcrete
and the ground,
the shotcrete lining should be represented by
beam or continuum elements, with 2 rows of elements over the cross-section if continuum elements with
quadratic shape function are used, and

60 m
layer 2
50 m
30 m
5 m
II
ground surface
1
2
.
0
5

m
layer 1
II
I
I
Figure 52 : Geometry of tunnel excavation example
Table 8. Material parameters for the tunnel excavation example
E (kN/m
2
)
(
o
) c (kN/m
2
) K
o
(kN/m
3
)
Layer 1 50000 0.3 28 20 0.5 21
Layer 2 200000 0.25 40 50 0.6 23
Shotcrete (d = 250 mm): linear elastic - E
1
= 5 000 MPa, E
2
= 15 000 MPa, = 0.15
to account for deformations occurring ahead of the face (pre-relaxation), the load reduction method or a
similar approach should be adopted.
9.2.2 Computational Steps
The following computational steps had to be performed by the analysts:
the initial stress state was set to
v
= H,
h
= K
o
H, and subsequently the deformations were set to zero,
the pre-relaxation factors given are valid for the load reduction method,
construction stage1: 40% pre-relaxation of the full cross section, and
construction stage2: excavation of full cross section, installation of shotcrete with E = E
2
.
In addition to a full face excavation, excavation of a top heading and bench was also considered, but
these results will not be discussed here (see Schweiger, 1997; 1998).
9.2.3 Selected Results
In the following, some of the most
interesting results are presented. In
Figure 53, surface settlements obtained
from 10 different analyses are compared.
50% of the calculations predict 52 or
53mm as the maximum settlement and
most of the others were within
approximately 20% of these values.
However, the calculations identified as
TL1A and TL10 show significantly lower
settlements. In both cases the reason was
that it was not possible to apply the load
reduction method correctly and therefore
other methods have been used. TL10
used the stiffness reduction method and it
is known that it is difficult to match these
two methods (Schweiger et al., 1997).
TL9 also employed the stiffness
reduction method but obtained larger
settlements, which is rather unusual.

distance from tunnel axis [m]
0 5 10 15 20
s
u
r
f
a
c
e

s
e
t
t
l
e
m
e
n
t

[
c
m
]
0
1
2
3
4
5
6
TL 1A
TL 2
TL 4
TL 5
TL 6
TL 7
TL 8
TL 9
TL 10
TL11
Figure 53 : Surface settlements for tunnel excavation in one step

TL 1: -99
TL 4: -176
TL 9: -317
TL 10: -236
TL 1: 74
TL 2: -178
TL 8: -165
TL 2: 133
TL 4: 144
TL 5: 185
TL 8: 108
TL 10: 200
TL 5: -234
TL 9: 113
[kNm/m]
calculation
TL1 TL2 TL4 TL5 TL8 TL9 TL10
m
a
x
.

n
o
r
m
a
l

f
o
r
c
e

[
k
N
]
0
1000
2000
3000
4000
5000
6000
Figure 54 : Comparison of maximum normal forces and bending moments in tunnellining
Figure 54 shows the calculated normal forces and bending moments in the shotcrete lining. Reasonable
agreement is observed for normal forces, with the exception of TL1 and TL10, but a wide scatter is obtained
for bending moments. Figure 54 indicates also the location of maximum bending moments and the
significant differences in magnitude (approximately 300%) and location are obvious. Even if TL1, TL9 and
TL10 are excluded because they did not refer exactly to the problem specification, the variation is still 70%.
Unfortunately, it was not possible from the information available to identify clearly the reasons for these
discrepancies but most probably they are due to differences in modelling the lining and in evaluation of the
internal forces.
9.3 Deep Excavation
Example
Figure 55 illustrates the
ge ome t r y a nd t he
excavation stages analysed
in this problem, and Table 9
lists the relevant material
parameters. Additional
speci fi cat i ons are as
follows.
9.3.1 General
Assumptions
plane strain conditions
apply,
a linear elastic -
pe r f e c t l y pl a s t i c
analysis with the Mohr-
Coul omb f a i l ur e
criterion was required,
perfect bonding was to
be assumed between the diaphragm wall and the ground,
struts used in the excavation could be modelled as rigid members (i.e., the horizontal degree of freedom
was fixed),
any influence of the diaphragm wall construction could be neglected, i.e., the initial stresses were
established without the wall, and then the wall was wished-in-place, and
the diaphragm wall was modelled using either beam or continuum elements, with 2 rows of elements
over the cross section if continuum elements with quadratic shape function were adopted.
9.3.2 Computational Steps

layer 1
layer 2
diaphragm
wall
strut No. 1
excavation step 1
strut No. 2
excavation step 2
-7.0
-8.0
-4.0
-3.0
-12.0
final excavation
-20.0
4
0

m
60 m
15.00
0.80
ground surface = 0.0
-26.0
layer 3
x
y
x'
Figure 55 : Geometry of the deep excavation example
Table 9. Material parameters for deep excavation example
E (kN/m
2
) (
o
) c (kN/m
2
) K
o
(kN/m
3
)
Layer 1 20000 0.3 35 2.0 0.5 21
Layer 2 12000 0.4 26 10.0 0.65 19
Layer 3 80000 0.4 26 10.0 0.65 19
Diaphragm wall (d = 800 mm): linear elastic E = 21 000 MPa, = 0.15, = 22 kN/m
3
The following computational steps had to be performed by the various analysts:
the initial stress state was set to G = J&T, 4 = K&Y?,
all deformations were set to zero and then the wall was wished-in-place,
construction stage 1:
excavation step 1 to a level of
-4.0 m,
construction stage 2:
excavation step 2 to a level of
-8.0 m, and strut 1 installed at
-3.0 m, and
construction stage 3: final
excavation to a level of -12.0
m, and strut 2 installed at -7.0
m.
Comparison of Results
It is worth mentioning that 5
out of the 12 calculations
submitted for comparison were
made by different analysts using
the same computer program.
Figure 56 compares surface
displacements for construction
stage 1 and shows 2 groups of
results. The lower values for the
heave fi-om calculations BGl and
BG2 may be explained because
of the use of interface elements,
which were used despite the fact
that the specification did not
require them. The results of BG3
and BG12 could not be explained
in any detail. There were
indications though that for the
particular program used a
significant difference in vertical
displacements was observed
depending whether beam or
continuum elements were used for
modelling the diapbragm wall.
This emphasises the significant
influence of different modelling
assumptions and the need for
evaluating the validity of these
models under defmed conditions.
It may be worth mentioning that
this effect was not observed to the
same extent in the other programs
used. Figure 57 shows the same
displacements for the final
excavation stage, and the results
are now almost evenly distributed
between the limiting values.
3.0
2.5
E 2.0
is
j 1.5
St
5 1.0
Y
5 IO 15 20 25
distance from wall [m]
Figure 56 : Vertical displacement of surface behind wall:
construction stage 1
6
distance from wall [m]
Figure 57 : Vertical displacement of surface behind wall: final
construction stage
61
It is apparent from
Figures 56 and 57 that
elasto-perfectly plastic
constitutive models are
not well suited for
analysing the
displacement pattern
around deep excavations,
especially for the surface
behind the wall because
the predicted heave is
certainly not realistic.
However, it was not the
aim of this exercise to
compare results with
actual field observation,
but merely to see what
differences are obtained
when using slightly
different modelling
assumptions within a
rather tight problem
specification.
It is interesting to
compare predictions of
the horizontal
_ 0.50
6
z 0.25
6
E
$
0.00
a -0.25
g
3 -0.50
c
0 .w -0.75
z
c -1.00
-1.25
BGI BG2 BG3 BG4 BG5 BG6 BG7 BG8 BG9 BGlOBGll BG12
calculation
Figure 58 : Horizontal displacement of head of wall
displacement of the head of the wall for excavation step 1. Only 50% of the analyses predict displacements
towards the excavation (+ve displacement in Figure 58) whereas the other 50% predict movements towards
the soil, which seems to be not very realistic for a cantilever situation. Significant differences were not
found in the predictions for the horizontal displacements of the bottom of the wall, the heave inside the
excavation and the earth pressure distributions. Calculated bending moments varied within 30% and strut
forces for excavation step 2 varied from 155 to 232 kN/m. A more detailed examination of this example can
be found in Schweiger (1997, 1998).
Tied-Back Deep Excavation Example
The final example presented here is closely related to an actual project in Berlin. Slight modifications
have been introduced in modelling of the construction sequence, in particular the groundwater lowering,
which was performed in various steps in the field, but was modelled in one step prior to excavation.
In this example the constitutive model to be used was not prescribed but the choice was left to the
analysts. Some basic material parameters have been taken from the literature and additional results from
one-dimensional compression tests on loose and dense samples were given to the participants, together with
the results of niaxial tests on dense samples. Thus the exercise represents closely the situation that is ofkn
faced in practice. Inclimometer measurements made during construction provided information of the actual
behaviour in situ, although due to the simplifications mentioned above a one-to-one comparison is not
possible. Of course, the measurements were not disclosed to the the participants prior to them submitting
their predictions.
It is the aim of this section to demonstrate the necessity of performing exercises of this kind. However,
due to space limitations only the most relevant aspects of the problem specification will be given here
(Figure 59). For the same reason only a limited number of results will be presented. They indicate however
the wide scatter of results that can be produced due to different interpretations of the available data. A much
more comprehensive discussion of this validation problem may be found in Schweiger (2000).
Some reference values for stifmess and strength parameters, obtained from the literature and frequently
used in the design of excavations in Berlin sand, are given below.
62
- 3 2 . O O m = b a s e o f d i a p h r a g m w a l l
S p e c i f i c a t i o n f o r a n c h o r s :
p r e s t r e s s e d a n c h o r f o r c e : 1 . r o w : 7 6 8 K N
2 . r o w : 9 4 5 K N
3 . r o w : 9 8 0 K N
d i s t a n c e o f a n c h o r s : 1 . r o w : 2 . 3 O m
2 . r o w : 1 . 3 5 m
3 . r o w : I . 3 5 m
c r o s s s e c t i o n a r e a : 1 5 c m 2
Y o u n g s m o d u k ~ s E = 2 . 1 e 8 k N / m 2
F i g u r e 5 9 : G e o m e t r y a n d e x c a v a t i o n s t a g e s f o r t h e t i e d b a c k e x c a v a t i o n
ES
= 2 0 0 0 0 & k P a , f o r 0 < z 2 2 0 m ( z = d e p t h b e l o w s u r f a c e ) ,
J %
= 6 0 0 0 0 & k P a , f o r z > 2 0 m ,
4 = 3 s ( m e d i u m d e n s e )
Y = 1 9 k N / m 3 ,
Y /
= 1 0 k N / m 3 , a n d
K , , = 1 - s i n + !
T h e m o d u l i o b t a i n e d f r o m o e d o m e t e r t e s t s c o u l d a l s o b e u s e d t o e s t i m a t e a p p r o x i m a t e v a l u e s f o r t h e
Y o u n g s m o d u l e f o r c a l c u l a t i o n s w i t h l i n e a r e l a s t i c - p e r f e c t l y p l a s t i c m a t e r i a l m o d e l s ( M o h r - C o u l o m b )
a s s u m i n g a n a p p r o p r i a t e P o i s s o n s r a t i o .
I n a d d i t i o n t o t h e s e v a l u e s , r e s u l t s f r o m o e d o m e t e r t e s t s o n l o o s e a n d d e n s e s a m p l e s a n d t r i a x i a l t e s t s
w i t h C T ; = 1 0 0 , 2 0 0 a n d 3 0 0 k P a w e r e p r o v i d e d t o p a r t i c i p a n t s . I t w a s n o t p o s s i b l e t o i n c l u d e a s i g n i f i c a n t l y
l a r g e n u m b e r o f t e s t r e s u l t s a n d t h u s t h e q u e s t i o n a r o s e l a t e r w h e t h e r t h e s t i f f n e s s v a l u e s o b t a i n e d f r o m t h e
o e d o m e t e r t e s t s w e r e a c t u a l l y r e s p r e s e n t a t i v e . I n p a r t i c u l a r , i f t h e c o n s t i t u t i v e m o d e l r e q u i r e d a s a n i n p u t
p a r a m e t e r a n o e d o m e t e r s t i f f h e s s a t a r e f e r e n c e p r e s s u r e o f 1 0 0 k P a , a v a l u e o f o n l y E s = 1 2 0 0 0 k P a w a s
f o u n d i n t h e d a t a p r o v i d e d , a n d t h i s w a s c o n s i d e r e d t o b e t o o l o w b y m a n y a u t h o r s . H o w e v e r , a g a i n t h i s
r e p r e s e n t s a s i t u a t i o n t h a t i s c o m m o n i n p r a c t i c e , a n d l e f t r o o m f o r i n t e r p r e t a t i o n b y t h e i n d i v i d u a l a n a l y s t s .
T h e p r o p e r t i e s o f t h e d i a p h r a g m w a l l w e r e s p e c i f i e d a s f o l l o w s .
E = 3 0 0 0 0 M P a
V = 0 . 1 5
Y
= 2 4 k N / m 3
6 3
The angle of wall friction was specified as /2.
9.4.1 General Assumptions
Additional specifications for this example are as follows:
plane strain conditions could be assumed,
any influence of the diaphragm wall construction could be neglected, i.e., the initial stresses were
established without the wall, and then the wall was wished-in-place and its different unit weight
incorporated appropriately,
the diaphragm wall could be modelled using either beam or continuum elements,
interface elements existed between the wall and the soil,
the domain to be analysed was as suggested in Figure 59,
the horizontal hydraulic cut-off that existed at a depth of 30.00 m was not to be considered as structural
support, and
the pre-stressing anchor forces were given as design loads.
9.4.2 Computational Steps
The following computational steps had to be performed by the various analysts:
the initial stress state was given by
v
= H,
h
= K
o
H,
the wall was wished-in-place and the deformations reset to zero,
construction stage 1: groundwater-lowering to 17.90 m,
construction stage 2: excavation step 1 (to level -4.80 m),
construction stage 3: activation of anchor 1 at level 4.30 m and prestressing,
construction stage 4: excavation step 2 (to level -9.30 m),
construction stage 5: activation of anchor 2 at level 8.80 m and prestressing,
construction stage 6: excavation step 3 (to level -14.35 m),
construction stage 7: activation of anchor 3 at level 13.85 m and prestressing, and
construction stage 8: excavation step 4 (to level 16.80 m).
The length of the anchors and their prestressing loads are indicated in Figure 59.
9.4.3 Brief Summary Of Assumptions Of Submitted Analyses
In Tables 10 and 11 the main features of all analyses submitted have been summarised in order to
highlight the different assumptions made, according to the personal preferences and experience of the
participants.
It follows from Tables 10 and 11 that a wide variety of programs and constitutive models has been
employed to solve this problem. Only a limited number utilised the laboratory test results provided in the
specification to calibrate their models. Most of the analysts used data from the literature for Berlin sand or
their own experience to arrive at input parameters for their analysis. Close inspection of Tables 10 and 11
reveals that only marginal differences exist in the assumptions made about the strength parameters for the
sand (everybody believed the experiments in this respect), and the angle of internal friction was taken as
36 or 37 and a small cohesion was assumed by many authors to increase numerical stability. A significant
variation is observed however in the assumption of the dilatancy angle , with values ranging from 0 to 15.
An even more significant scatter is observed in the assumption of the soil stiffness parameters. Although
most analysts assumed an increase with depth, either by introducing some sort of power law, similar to the
formulation presented by Ohde (1951), which in turn corresponds to the formulation by Janbu (1963), or by
defining different layers with different Youngs moduli. Additional variation is introduced by different
formulations for the interface elements, element types, domains analysed and modelling of the prestressed
anchors. Some computer codes and possibly some analysts may have had problems in modelling the
prestressing of the ground anchors, and actually part of the force developed due to deformations occurring in
the ground appears lost. Where this is the case a remark has been included in Table 10. The constitutive
model Hardening Soil corresponds to a shear and volumetric hardening plasticity model provided in the
commercially available finite element code PLAXIS. It features also a stress dependent stiffness, different
for loading and unloading or reloading paths.
9.4.4 Results
A total of 15 organisations (comprising University Institutes and Consulting Companies from Germany,
Austria, Switzerland and Italy), referred to as B1 to B15 in the following, submitted predictions. Figure 60
shows the deflection curves of the diaphragm wall for all entries. It is obvious from the figure that the
results scatter over a very wide range, which is unsatisfactory and probably unacceptable to most critical
observers of this important validation exercise. For example, the predicted horizontal displacement of the
top of the wall varied between 229 mm and +33 mm (-ve means displacement towards the excavation).
Looking into more detail in Figure 60, it can be observed that entries B2, B3, B9a and B7 are well out of the
mainstream of results. These are the ones which derived their input parameters mainly from the
oedometer tests provided to all analysts, but it should be remembered that these tests showed very low
stiffnesses as compared to values given in the literature. Some others had small errors in the specific weight,
but these discrepancies alone cannot account for the large differences in predictions.
As mentioned previously, field measurements are available for this project and although the example here
has been slightly modified in order to facilitate the calculations, the order of magnitude of displacements is
known. Figure 61 shows the measured wall deflections for the final construction stage together with the
calculated results. Only those calculations which are considered to be near the measured values are
included. The scatter is still significant. It should be mentioned that measurements have been taken by
inclinometer, but unfortunately no geodetic survey of the wall head is available. It is very likely that the
base of the wall does not remain fixed, as was assumed in the interpretation of the inclinometer
measurements, and that a parallel shift of the measurement of about 5 to 10 mm would probably reflect the
in situ behaviour more accurately. This has been confirmed by other measurements under similar conditions
in Berlin. If this is true a maximum horizontal displacement of about 30mm can be assumed and all entries
that are within 100% difference (i.e., up to 60mm) have been considered in the diagram. The predicted
maximum horizontal wall displacements still varied between 7 and 57 mm, and the shapes of the predicted
curves are also quite different from the measured shape. Some of the differences between prediction and
measurements can be attributed to the fact that the lowering of the groundwater table inside the excavation
has been modelled in one step whereas in reality a stepwise drawdown was performed (the same has been
assumed by calculation B15). Thus the analyses overpredict horizontal displacements, the amount being
strongly dependent on the constitutive model employed, as was revealed in further studies. In addition, it
can be assumed that the details of the formulation of the interface element have a significant influence on the
lateral deflections of the wall, and arguments similar to those discussed in the previous section for
implementing constitutive laws also hold, i.e., no general guidelines and recommendations are currently
available. A need for them is clearly evident from this exercise.
Figure 62 depicts the calculated surface settlements, again only for the same solutions that are presented
in Figure 61. These key displacement predictions vary from settlements of up to approximately 50 mm to
surface heaves of about 15 mm. Considering the fact that calculation of surface settlements is one of the
main goals of such an analysis, this lack of agreement is disappointing. It also highlights the pressing need
for recommendations and guidelines that are capable of minimising the unrealistic modelling assumptions
that have been adopted and consequently the unrealistic predictions that have been obtained. The
importance of developing such guidelines should be obvious.
Figure 63 shows predictions of the development of anchor forces for the upper layer of anchors.
Maximum anchor forces for the final excavation stage range from 106 to 634kN/m. As mentioned
previously, some of the analyses did not correctly model the prestressing of the anchors because they do not
show the specified prestressing force in the appropriate construction step. Predicted bending moments,
important from a design perspective, also differ significantly from 500 to 1350kNm/m.
Taking into account the information presented in Figures 60 to 63 and Table 10, it is interesting to note
that no definitive conclusions are possible with respect to the constitutive model or assumptions concerning
element types and so on. It is worth mentioning that even with the same finite element code (PLAXIS) and
the same constitutive model (Hardening Soil Model) significant differences in the predicted results are
observed. Clearly these differences depend entirely on the personal interpretation of the stiffness parameters
from the information available. Again, it is noted that a more comprehensive coverage of this exercise is
beyond the scope of this paper but further details may be found in Schweiger (2000).
Table 10. Summary of analyses submitted for the tied back excavation problem
No. Code
Constitutive
Model

(-)

(
o
)

(
o
)
K
o
Domain
analysed
width x height
(m)
Element
type -
soil
Element
type - wall
Element type -
anchor / grout
body
Element
type -
Interface
Remark
B1 Tunnel Mohr-Coulomb 0.3 35 5 100 x 64
9 noded
continuum
bar /
membrane
prestress
force ?
B2 Plaxis
Hardening
(z < 40 m)
Mohr-Coulomb
(z > 40 m)
0.2 36 6 0.41 100 x 100 quadratic beam
bar /
membrane
R
inter
= .5
Einter > EBoden
B2a Plaxis
Hardening
(z < 32 m)
Mohr-Coulomb
(z > 32 m)
0.2 36 6 0.41 100 x 100 quadratic beam
bar /
membrane
R
inter
= .5
B3 Abaqus
Hypoplastic
without
intergranular
strains
0.42 161 x 162 linear
4 noded
continuum
bar / bar = 20
prestress
force ?
B3a Abaqus
Hypoplastic
with
intergranular
strains
0.42 161 x 162 linear
4 noded
continuum
bar / bar = 20
prestress
force ?
B4 Ansys Drucker Prager 0.3 35 0 0.43 105 x 107 linear
4 noded
continuum
bar / bar
prestress
force ?
B5 Sofistik Mohr-Coulomb 0.3 35 15 80 x 60 beam spring
no
interface
prestress
force ?
B6 Z_Soil Mohr-Coulomb 0.3 35 15 122 x 90
prestress
force ?
B7 Feerepgt
Zienkiewicz/
Pande/Schad
0.26 40.5 13.5 0.35 90 x 60 quadratic continuum
Continuum /
continuum
after GE,
=20.25
B8 Plaxis
Hardening
2 layers
(z < 20 m)
(z > 20 m)
0.2 35 10 0.43 90 x 70 quadratic beam
bar /
membrane
Plaxis
B9 Plaxis Mohr-Coulomb 0.35 35 5 0.43 150 x 100 quadratic beam
bar /
membrane
R
inter
=0.5
B9a Plaxis Hardening 0.2 35 10 0.43 150 x 100 quadratic beam
bar /
membrane
R
inter
=0.5
B10 Plaxis Hardening 0.2 36 6 0.41 100 x 72 quadratic beam
bar /
membrane
R
inter
=0.61
B11 Plaxis
Hardening
2 layers
(z < 20 m)
(z > 20 m)
0.3 35 0 0.45 150 x 120 quadratic beam
bar /
membrane
Rigid ?
prestress
force ?
B12 Plaxis Mohr-Coulomb 0.3 35 4 90 x 92 quadratic beam
bar /
membrane
R
inter
=0.5
prestress
force ?
B13 Abaqus
Hypoplastic
with
intergranular
strains
100 x 100
linear +
internal
node
beam bar = 15.5
anchors
fixed at
boundy
B14 Befe
Shear Hardening
Model with
small strain
stiffness
0.2 35 5 0.43 120 x 100 quadratic
8 noded
continuum
bar / bar
Dicke = 0
= 17.5
B15 Plaxis
Hardening
2 layers
(z < 20 m)
(z > 20 m)
0.2
0.2
35
41
7
14
0.43
0.34
96 x 50 quadratic
beam +
continuum
bar /
membrane
R
inter
=0.8
9.5 Conclusions Of Validation Studies
Some results from three benchmark examples suggested by the working group 1.6 of the German Society
for Geotechnics have been presented. These problems have allowed a comparison of results obtained from
various methods of analysis, geotechnical models and different analysts. Although quite tight specifications
have been given for the first two examples, significant differences in predicted results have been observed
even in these rather simple cases.
The third example was by far the most difficult of the three problems described, because the choice of
constitutive law and parameter identification was left entirely to the analyst. Perhaps, as might have been
expected, this freedom of choice lead to large differences in results. The reasons for the differences could
not be fully identified except for the fact that information given on the soil properties leads to quite different
input parameters, depending on personal interpretation. However, it is believed that this exercise closely
represents the situation in practice and reveals once again that numerical modelling in geotechnics is a very
difficult task. A lot of experience is required in order to make sensible assumptions for material and model
parameters, which are not explicitly known beforehand, in order to arrive at a reasonable model of the real
situation in the field.
The exercise presented here very clearly indicates the need for guidelines and training for numerical
analysis in geotechnical engineering in order to achieve reliable solutions for practical problems. However,
at the same time these examples demonstrate the power of numerical modelling techniques provided
experienced users apply them. A full analysis of the benchmark example No. 3 will be given in a
forthcoming report of the working group of the German Society, and the problem will be further dealt with
also by Technical Committee 12 (TC 12 Validation of Computer Simulations) of the International Society
for Soil Mechanics and Geotechnical Engineering (ISSMGE).
Table 11. Summary of stiffness parameters used in tied back excavation analyses
No. Constitutive Model E-dependence
Number of
layers
Power
E
ref,loading
or E
min
(kN/m
2
)
E
ref,unloading
or E
max
(kN/m
2
)
B1 Mohr-Coulomb similar to Ohde 2 0.5
z<20m: 14 900
z>20m: 44 700
B2
Hardening Soil
Mohr-Coulomb
similar to Ohde
linear increase
2 0.85
z<40 m: 15 000
z=40 m: 253 000
39 000
B2a
Hardening Soil
Mohr-Coulomb
similar to Ohde
linear increase
2 0.5
z<32 m: 60 000
z=32 m: 226 300
180 000
B4 Drucker Prager Layers 43 - z=1m: 10 500 z = 105m: 457 000
B5 Mohr-Coulomb Layers 6 - z=4.8m: 32 640 z = 60m: 303 200
B6 Mohr-Coulomb Layers ? -
z<20m: 20 000 z
z>20m: 44 700 z
B7
Zienkiewicz/
Pande/Schad
- - - 60 000
B8 Hardening Soil similar to Ohde 2 0.5
z<20m: 20 000
z>20m: 60 000
z<20m: 74 400
z>20m: 120 000
B9 Mohr-Coulomb Layers 3 - z<20m: 39 400 z>40m: 310 000
B9a Hardening Soil similar to Ohde - 0.65 25 000 100 000
B10 Hardening Soil similar to Ohde - 0.5 60 000 180 000
B11 Hardening Soil similar to Ohde 2 0.5
z<20m: 20 000
z>20m: 60 000
z<20m: 100 000
z>20m: 300 000
B12 Mohr-Coulomb Layers 9 - z=0-4.8m: 23 000 z = 42-92m: 365 000
B14
Shear Hardening
Model with "small
strain stiffness"
similar to Ohde - 0.5
G
min
= 72 000
G
max
= 576 000 (at
very small strains)
G
entl
= 288 000
G
entl
= 2 304 000 (at very
small strains)
B15
Hardening Soil
Hardening Soil
similar to Ohde
similar to Ohde
2 0.5
z<20 m: 32 000
z>20 m: 96 000
192 000
384 000

horizontal displacement [mm]
-250 -225 -200 -175 -150 -125 -100 -75 -50 -25 0 25 50
-250 -225 -200 -175 -150 -125 -100 -75 -50 -25 0 25 50
d
e
p
t
h

b
e
l
o
w

s
u
r
f
a
c
e

[
m
]
0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32
B1
B2
B2a
B3
B3a
B4
B5
B6
B7
B8
B9
B9a
B10
B11
B12
B13
B14
B15
Figure 60 : Wall deflection all solutions submitted: tied back excavation

horizontal displacement [mm]
-60 -50 -40 -30 -20 -10 0 10
-60 -50 -40 -30 -20 -10 0 10
d
e
p
t
h

b
e
l
o
w

s
u
r
f
a
c
e

[
m
]
0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32
B2a
B4
B5
B6
B8
dB9
B11
B12
B13
B14
B15
measurements
measurement
Figure 61 : Wall deflection for selected solutions and inclinometer measurements: tied back
excavation
Outlook And Further Validation Work
As mentioned earlier, the
working group 1.6 of the
German Society for
Geotechnics will continue its
efforts in this area. In addition,
the Working Group A of the
COST-Action of the European
Community is also making
efforts to provide guidelines
for applications of numerical
models in geotechnical
engineering. Finally, ERTC7
and in particular TC12 of
ISSMGE have started working
in this area. The terms of
refaences of TC12 are as
follows:
l to promote cooperation
and exchange of
information on the
development and
0 ! /; 1
1 2 3 4 5 6 7 8
pr est r ess f or c e spec i f i i computational step
i n st ep 3 = 334 k N/m
application of computer
simulations in geotechnical
engineering.
Figure 63 : Forces in fast layer of anchors for selected solutions:
tied back excavation
l to foster the development of recommendations and guidelines for the application of numerical analysis
in geotechnical engineering.
l to organise workshops devoted to the themes pertinent to the work of TC12 during general and specialty
international conferences, e.g., ICSMGE, IACMAG and NUMOG.
DISCRETE ELEmNT MODELS
In the 1970s Cundall developed a computer program known as BALL. This development heralded the
birth of a revolutionary computational technique in geotechnics, which is now termed the Discrete Element
Method (DEM). It has also been termed the distinct element method. The aim of the original model was to
describe the fundamental behaviour of granular materials en musse by focusing on their basic constituents,
i.e., the grains themselves and their interactions. The technique is able to deal with both non-cohesive grains
and cohesive grains. The method was validated by Cundall and Strack (1979) by comparing force vector
plots obtained ti-om the computer program BALL, with the corresponding plots obtained from photoelastic
analysis conducted by de Jossehn de Jong and Verruijt (1969). The correspondence between these plots was
sufficiently good to mark the distinct element method as a valid tool for fundamental research into the
behaviour of particulate materials.
Since the fast proposal of this numerical method, various researchers have developed and adapted it to
\study the behaviour of a variety of materials composed of individual particles or blocks. Geotechnical
problems where the mechanical behaviour of a soil or rock mass is influenced strongly by the effects of
joints and cracks are well suited to analysis by the discrete element methods. These methods may be
characterised by a number of distinguishing capabilities, such as their ability to deal with finite deformations
and rotations of discrete blocks (deformable or rigid), blocks that are originally connected may separate
during the analysis, and the possibility that new contacts may develop between blocks due to displacements
and rotations.
Several different approaches to achieve these criteria have been developed, with probably the most
commonly used methods now being the distinct element codes UDEC and 3-DEC (Lemos et d., 1985),
which both employ an explicit finite difference scheme to solve the governing equations. These codes solve
70
which both employ an explicit finite difference scheme to solve the governing equations. These codes solve
the equations of motion explicitly and material damping must be included to allow approximations of quasi-
static behaviour of granular assemblages or blocky rock masses. With this method stability cannot be
guaranteed for the solutions of problems in statics. An alternative implicit equation solution technique has
also been proposed for the discrete element method, e.g., van Baars (1996), Hocking (1977). The implicit
method is able to provide a stable solution to quasi-static problems. A useful summary of DEM theory is
presented in the text by Pande et al. (1990).
These techniques are particularly useful for revealing the dominant mechanisms of deformation of
materials that posses a definite structure. Examples of this usage may be seen in Figures 64 to 66.
10.1 Biaxial Compression Of Granular Material
Figure 64 shows the simulation of a biaxial test of sandstone (van Baars, 1996). In this simulation the
cohesive-granular sandstone has been modelled by a relatively large number of discrete (approximately
circular) particles. The grains are initially cemented together so that the contacts between the particles have
cohesive and frictional properties. Figure 64 shows representations of the particle assemblage at various
stages of the biaxial loading, and the sequence of diagrams reveals the failure mechanism. If cemented
contacts are broken, then a thick line has been drawn perpendicular to the contact. The horizontal surfaces
of the loading platens are also shown as thick lines. The vertical boundaries are subjected to a confining
pressure.
Numerical studies such as those shown in Figure 64 can produce interesting findings. van Baars (1996)
suggested the following failure
mechanism.
In Phase A, some of the
contact forces become tensile,
even in the presence of a
compressive confining pressure.
As a result, micro-cracks form
and particles separate. In phase
B the crack growth continues
and weakens the surrounding
region and a failure surface
eventually forms. It is worth
noting that although the surface
formed by adjacent cracks is
diagonal, the microcracks are
formed by the separation of
particles along approximately
horizontal contacts. As the
vertical load applied at the
specimen boundary is increased,
the grains with broken contacts
act as rollers (phase C), and the
overall response of the specimen
involves reduction in the vertical
load (softening).
10.2 Bearing Response Of Jointed Rock
In a study by Singh et al. (1996) the bearing capacity of blocky media was investigated using both
numerical and experimental techniques. In the experiments, a physical model consisting of prismatic blocks
of a commercially available engineering plastic DELRIN (polyacetal) was constructed. The model was
confined within a rigid box. In this plane strain problem a rigid footing was placed on the surface of the

Figure 64 : Failure mechanism of a cohesive-granular material in
biaxial compression (after van Baars, 1996)
blocky medium. A vertical load was applied to the footing, and as the load increased the deformation of the
blocky medium was observed.
Numerical modelling of the same
problem has also been carried out using
the UDEC discrete element program.
During both the physical and numerical
model tests, the average contact stress
generated under the footing was logged
against the vertical displacement of the
footing. A typical plot of the contact
stress against the settlement S, normalised
by the footing width B, is shown in
Figure 65. Both numerical and physical
test results are shown on this figure. The
numerical results plot as a curve with
three distinct stages. Stage 1 starts at the
origin of the plot and is limited by the
first peak in the load-displacement curve.
This peak can be regarded as the bearing
capacity of the blocky medium, because the
numerical model reveals that for the
geometry considered, it corresponds to
opening of some of the joints and upward
buckling of the most highly stressed inclined
layer of blocks in the medium, generally a
layer that is in direct contact with the
footing. The development of this failure
mechanism in the numerical model can be
seen in Figure 66, and this was also evident
in the physical experiments. The second
stage of the load-displacement curve starts at
the first peak and extends down to the lowest
trough in the curve. This corresponds to a
stress-relieving phase. In the physical model
the drop in the load is very sudden and is
associated with the rapid release of stored
energy as a column of blocks buckles. As
the footing is pushed further into the blocky
medium the load rises again and this is
termed stage 3 in the load-displacement
curve. The load continues to increase until
another layer of blocks undergoes an upward
buckling failure.
The ability and power of the distinct
element method to give a clear and
unequivocal indication of the likely failure
mechanism is very evident in this problem.
When the blocks themselves are very strong,
failure can occur by sliding on the joints or
elastic buckling of layers of blocks. In the
example considered here, the numerical
technique indicated clearly that failure is
dominated by buckling, with only a minor
amount of energy dissipation in shearing
along joint planes in the wedge-shape region of blocks above the buckled layers - see Figure 66. Buckling
Figure 66 : (a) Initial arrangement of blocks, (b) blocks
deformed by footing penetration
Figure 65 : Load-settlement curves for a blocky mass
of inclined layers of blocks was confirmed as the dominant failure mechanism in the physical experiments.
Further research work is required to determine if such methods can also give quantitative predictions of
failure loads that are in close agreement with those measured in the laboratory, but there can be no doubting
its ability to capture the important mechanisms.
10.3 STOCHASTIC TECHNIQUES
In the past two decades the importance of reliability analyses and the stochastic nature of soil properties
has been recognised in practical geotechnical engineering. Although the advantages of probabilistic tools
are now generally accepted, there is still only a limited number of applications of these tools to practical
problems documented in the literature. The reason for this situation is probably that closed form solutions
are not available for most practical cases and numerical methods involve large computational efforts.
However recent achievements in numerical analysis and the fast development of hardware have significantly
reduced computer run times and pre- and post-processing efforts so that a probabilistic approach based on
deterministic finite element analysis is now a feasible technique, worthy of further investigation. Several
researchers have already moved in this direction
(e.g., Thurner and Schweiger, 2000; Kaggwa,
2000).
In general, three major sources of uncertainty
in soil profile modelling can be distinguished
(Vanmarcke, 1977). These are the natural
heterogeneity, the limited availability of
information on subsurface conditions and
measurements errors. In practice, most
geotechnical problems are still solved using a
deterministic approach employing soil parameters
which are judged to be on the safe side. In order to assess the behaviour of a geotechnical structure or the
change of the deterministic factor of safety with respect to variable parameters such as shear strength for
example, sensitivity analyses are usually carried out.
The uncertainty in the real value of each soil property at a given point can be defined as a random
variable described by a mean value, coefficient of variation (COV) and a probability distribution function.
Common probability distribution functions are the normal and the lognormal distribution; the latter is
preferred for variables which cannot have negative values. Although more sophisticated distributions like
the Gumbel or Beta distributions may improve results for special cases, applications are rare because they
require much more computational effort.
To determine the distribution function for a particular soil property more than 40 data points are needed,
but it is usually not feasible to obtained this amount of data in most cases. Consequently, a coefficient of
variation has to be obtained from other projects in similar soil conditions or from the literature (e.g., Lacasse
and Nadim, 1997; Rackwitz and Peintinger, 1981; Cherubini et al., 1993). For example, typical coefficients
of variation for the (drained) friction angle and the cohesion (undrained shear strength) c are shown in
Table 12.
A simple but effective procedure to combine different sources of information is presented in Rackwitz
and Peintinger (1981). Another helpful tool to take different sources of information into account is the
program CombInfo (1999).
In addition to the variability of the shear strength parameters (c and ) a weak negative correlation
should be taken into account. The assumption of independent parameters may lead to conservative results
(Mostyn and Li, 1993). Another important aspect that needs to be considered is autocorrelation (e.g.,
Lacasse and Nadim, 1997; Mostyn and Li, 1993; Wickremesinghe and Campanella, 1993; Rackwitz and
Peintinger, 1981).
Table 12. Typical coefficients of variation
Soil COV
(%)
COV c
(%)
Gravel 2 - 15 -
Sand 2 - 15 -
Silt 5 20 - 30
Clay 10 15 - 25
10.4 Safety Factor Or Reliability Analysis Of Geotechnical Structures?
10.4.1 Deterministic Analysis
In geotechnical engineering the traditional factor of safety, F
s
, for a slope, for example, is usually
determined using deterministic methods and can be defined as (e.g., Brinkgreve and Bakker, 1991):
c c
s
c
c
F


tan
tan
+
+
(35)
where c is the cohesion, is the friction angle and is the effective normal stress. c and are the peak or
ultimate strength parameters and the subscript c indicates the mobilised strength parameter, just high
enough to ensure equilibrium. For slope stability problems this definition is used in limit-equilibrium
methods as well as in non-linear finite element analyses. Some results applying analytical methods have
been published by Fredlund (1984) and Duncan (1996). It can be shown that for homogeneous conditions
the results obtained from common methods of limit equilibrium deviate at most by t6%. These types of
solutions are comparable with special finite element formulations (Yu et al., 1998; Jiang and Magnan,
1997).
The above definition of the factor of safety (Equation 35) has been implemented into the finite element
code PLAXIS (Brinkgreve and Vermeer, 1998) where an automatic reduction of strength parameters is
performed until equilibrium can be no longer satisfied thus indicating the global factor of safety. For
homogeneous conditions this procedure works very well, but for more complicated structures reliable results
may be as difficult to achieve as by methods of slices.
As shown by Li et al. (1993), F
s
is a variant index because it depends on the interpretation of resisting
and driving forces. In combination with the uncertainties of the soil properties and the fact that geotechnical
engineers tend to use conservative design values, this may lead to very unrealistic reliability levels. For
these and other reasons, attempts have been made to employ more rational reliability analyses for
geotechnical problems. These are based soundly on probability theory.
10.4.2 Probabilistic Analysis
The key feature of a probabilistic approach is the definition of a so called performance function G(X)
which can be written in general form as:
( ) ( ) ( ) X S X R X G
(36)
where R(X) is the resistance, S(X) is the action, and X is the collection of random input parameters. For
G(X)<0 failure is implied, while G(X)>0 indicates stable behaviour. The boundary defined by G(X)=0
separating the stable and unstable state is called the limit state boundary (Mostyn and Li, 1993).
Determination or definition of the resistance R(X) is generally not straightforward, especially when R(X)
is represented by a threshold value for serviceability states or when different failure modes are possible.
Th e pro bab il ity of f ail ure p
f
i s def ine d as:
( ) [ ] ( )
( )


0
0
X G
f
dx X f X G p p
(37)
where f(X) is the common probability density function of the vector formed by the variables X, which is
usually unknown.
To solve this integral many different possibilities have been proposed in the literature (e.g., Li, 1992;
Zhou and Nowak, 1988). Thurner and Schweiger (2000) suggest the use of numerical integration to
determine the kth moment of G (E[G
k
(X)]), due to the fact that simulation methods like FORM (First Order
Reliability Method) or SORM (Second Order Reliability Method) require the partial derivatives of the
performance function. In the context of numerical methods these approaches involve significant
modifications to techniques such as standard finite element formulations, and these do not seem to be
feasible for practical application in the near future.
10.4.3 Transformation And Numerical Integration
The method proposed by Zhou and Nowak (1988) can be used to integrate Equation (37). It is a
numerical procedure for computing the statistical parameters of a function G(X) of multiple random
variables. The sample of basic variables is obtained by transforming a priori the selected points from the
standard normal space to the basic variable space. Thus this method differs from a standard point estimate
method in the sense that points and weights are predetermined in the standard normal space.
If the basic variables are standard normal distributed the performance function in Equation (37) can be
written as:
( ) ( )
n
Z Z Z G Z G , , ,
2 1
K (38)
The exact kth moment of G, E[G
k
(Z)], may be obtained by evaluating the integral:
( ) [ ] ( ) ( )

+

dz z G z Z G E
k k
(39)
where (z) is the cumulative distribution function of standard normal variable Z. The integration in
Equation (39) can be approximated using the formula:
( ) [ ] ( )

m
j
j
k
j
k
z G w Z G E
1
(40)
where m is the number of points considered, w
j
are the weights and z
j
are the typical normal points.
10.5 Finite Element Methods Applied To Reliability Analysis
The combination of the point estimate method described above with the well known finite element
analysis has been demonstrated by a number of authors, e.g., Thurner and Schweiger (2000), Kaggwa
(2000). The advantages of this combined approach compared to say limit equilibrium methods or other
similar methods are: on one hand that all features of numerical modelling remain, and on the other hand that
with finite element calculations usually a number of relevant system parameters are readily obtained. These
parameters are the basis for estimation of the performance function, which is evaluated using the numerical
procedure described previously. Thurner and Schweiger (2000) illustrate this approach by taking a
tunnelling problem as their example.
Kaggwa (2000) has adopted a similar approach and shown how probabilistic methods may be applied to
a variety of problems, to determine the significance of spatial variability on the key aspects of the problem.
These examples included an examination, using the finite element methods, of the influence of spatial
variability of soil parameters on prediction of displacements and pore pressures in soft clay under the Muar
test embankment in Malaysia (Brand and Premchitt, 1989). Kaggwa showed how the field measurements
should be viewed in the light of observed spatial variation of the soil parameters. From this study he
concluded that the incorporation of spatial variability in finite element analyses should be seen as a natural
progression of the discretisation process. In addition to geometric discretisation, the values of material
parameters can be varied randomly within a soil layer to simulate better the natural variability of soils. It is
noted that with sufficient repetitions, information on the range of predictions and their statistics can be
evaluated.
11.0 SOFT COMPUTING TOOLS
In the recent past, a new field of soft computing has emerged for solving decision-making, modelling,
and control problems. This approach parallels the remarkable ability of the human mind to reason and learn
in an environment of uncertainty and imprecision. Soft computing consists of many complementary tools:
fuzzy logic, neuro-computing, probabilistic reasoning, genetic algorithms, and others. A brief description of
one of these tools and its application in geotechnical engineering are described in more detail below in order
to illustrate the power of these methods.
An artificial neural network (ANN) is a computational mechanism able to acquire, represent, and
compute a mapping from multivariate space of information to another, given a set of data representing that
mapping (Garret, 1994). The basic architecture of neural networks has been covered widely (Rumelhart
and McClelland, 1986; Lippman, 1987; Flood and Kartam, 1994). A typical artificial neural network is
shown in Figure 67. The most frequently used neural-network paradigm is the back-propagation learning
algorithm (Rumelhart et al., 1986). Neural networks are trained by the presentation of a set of examples of
associated input and output (target) values. The hidden and output layer neurons process their inputs by
multiplying each of their inputs by the corresponding weights, summing the product, and then processing the
sum using a nonlinear transfer function to produce a result. The S-shaped sigmoid function is commonly
used as the transfer function. The neural-network learns by adjusting the weights between the neurons in
response to the errors between actual output values and target output values. At the end of this training
phase, the neural network represents a model, which should be able to predict a target value given the input
value.
Artificial neural network models are adaptive and capable of generalisation. They can handle imperfect
or incomplete data, and can capture nonlinear and complex interactions among variables of a system.
Because of these strengths, the artificial neural network method is emerging as a powerful tool for
modelling. Of course, their major limitation is that they contain no information about the underlying
mechanics of the problem being modelled.
11.1 Artificial Neural Network Architecture
A typical back-propagation artificial neural network is shown in Figure 67. A network can have several
layers. The outputs of each intermediate layer are the inputs to the following layer. Each layer has a weight
matrix W, a bias vector b, and an output vector a. Each element of the input vector p is connected to each
neuron input through the weight matrix W. The i-th neuron has a summer that gathers its weighted inputs
and bias to form its own scalar output n(i). The various n(i) taken together form an S-element vector n.
Finally, the neuron layer outputs form a column vector a. The layers of a multilayer network play different
roles. The layer that produces the network output is called an output layer. The layer that gets the inputs is
called the input layer. All other layers are called hidden layers. It is common for the number of inputs to a





P(1)
P(2)
P(3)
P(4)
P(R)
W1(1,1)
W1(S1,R)
n(1)
n(2)
n(S1)
b1(1)
b1(2)
b1(S1)
.
.
.
.
.
.
.
.
.
.
.
W2(1,1)
W2(1, S2)
W1
b1


b2(1)
W2
b2
P
R1 S1R
S11
S1R


n1

S1

1

n2
S21
S2S1
S21
a1
S1
1

a2(1)
a1(1)
1
1
S1
a1= tansig (W1*P+b1) a2= logsig (W2*a1+b2)
Input Neuron Hidden Layer Output Layer
Where
R = number of input variables
S1 = number of hidden nodes
S2 = number of output nodes


b2(S2)

a1(2)
a1(S1)
a2(S2)
.
.
S2
R
Figure 67 : Schematic representation of an artificial neural network (ANN)
layer to be different from the number of neurons. The network shown in Figure 67 has R inputs (R neurons
in the input layer), S1 neurons in the hidden layer, and S2 neurons in the output layer. The number of hidden
layers can be varied based on the application. A constant input value of 1 is fed to the biases for each
neuron.
11.2 Learning Rule
The learning rule refers to the mechanism that is used to adjust the weights and biases of the networks to
achieve some desired network behaviour. Several learning rules have been developed by different authors.
The back-propagation learning rule is the most popular one that has been used to train nonlinear,
multilayered networks to perform function approximation and pattern classification. The back-propagation
learning rule can be used to adjust the weights and biases of networks in order to minimise the sum-squared
error of the network. This is done by continually changing the values of the network weights and biases in
the direction of steepest descent with respect to error. Derivatives of the error vector are calculated for the
networks output layer and then back-propagated through the network until derivatives of error are available
for each hidden layer.
11.3 Training
Training refers to the process that repeatedly applies input vectors to the network and calculates errors
with respect to the target vectors and then finds new weights and biases with the learning rule. It repeats this
cycle until the sum-squared error falls beneath an error goal, or a maximum number of epochs have
occurred. Training a feed-forward network with the back-propagation learning rule is most frequently used
in function approximation and pattern recognition. The training parameters specify the number of epochs
between displaying progress, the maximum number of epochs to train, the sum-squared error goal, and the
learning rate. The size of changes that are made in the weights and biases at each epoch are specified with
the learning rate. Training continues until the specified error goal is met, or the minimum error gradient
occurs, or the maximum number of epochs has finished. The implementation of theses training techniques is
now provided in commercially available software such as MATLAB (e.g., Demuth and Beale, 1996).
11.4 Example - Predicting Uplift Capacity Of Suction Caissons
Several applications of the neural network method to civil and geotechnical engineering problems have
beeen reported already in the literature (e.g., Flood and Kartam, 1994; Garrett, 1994; Goh, 1996; Ghaboussi
and Sidarta, 1998; Rahman et al., 2000). Considered here is just one of these applications, in which the
artificial neural network method was applied to the prediction of the uplift capacity of suction caisson
foundations in a variety of soils (Rahman et al., 2000). The problem is illustrated schematically in Figure
68.
A database consisting of the
results of 60 individual test data
sets from 12 independent
experimental studies was
available both to train and
evaluate the artificial neural
network. These data have been
obtained from both laboratory
and field-scale experiments. The
measured results have been used
separately in two stages. 50 data
sets were used in the training
stage and the remaining 10 data
sets were used in the testing
stage. The data available in the
literature are mostly for clays.
Few data are available for sands

d
L
P


D
L = the length of the caisson
D = the diameter of the caisson in plan
D = the depth of the load application point from the soil surface
P = load applied to caisson
= angle of load inclination
Figure 68 : Definition of suction caisson uplift problem
and therefore sands have not included in this example.
The training data were used to develop the model. The testing data were used for testing the
generalization capability of the artificial neural network system. The 10 testing data sets were chosen
randomly from the database. However, the L/d values of the testing data set cover the same range as those
used in the training stage. Therefore the testing result is expected to reflect accurately the prediction ability
0
50
100
150
200
250
300
350
400
0 50 100 150 200 250 300 350 400
Measured Uplift Capacity P (kPa)
C
a
l
c
u
l
a
t
e
d

U
p
l
i
f
t

C
a
p
a
c
i
t
y

P

(
k
P
a
)
FEM
Meau = Cal.
Meau. = 2 * Cal.
Meas. = 0.5 * Cal.
ANN
Figure 70 : Comparison of results of ANN and FEM Modelling for testing data
0
50
100
150
200
250
300
350
400
0 50 100 150 200 250 300 350 400
Measured Uplift Capacity P (kPa)
C
a
l
c
u
l
a
t
e
d

U
p
l
i
f
t

C
a
p
a
c
i
t
y

P

(
k
P
a
)
FEM
Meas = Cal.
Meas. = 2 * Cal.
Meas. = 0.5 * Cal.
ANN
Figure 69 : Comparison of results of ANN and FEM modelling for training data
of the model that is developed with the training data set. The testing data have the same format as the
training data in the artificial neural network models, but values of variables are different from those of the
training data. It is emphasised that in general, the aim of this example was not to look for an artificial neural
network system that can best fit the training data. Rather, the aim was to look for a system trained on the
training data that could subsequently respond to the testing data in a satisfactory manner.
11.4.1 Model Evaluation
A comparison of the experimental measurements of uplift capacity results and the outcome of the training
phase of the ANN is provided in Figure 69. An indication of the ability of the ANN to provide predictions
after appropriate training is provided in Figure 70, which shows a comparison of the ANN predictions and
the remaining, independent experimental data for caisson uplift capacity.
In order to assess whether the ANN model could provide improved predictions of the uplift capacity of
suction caissons, the outcome of the ANN model was also compared with the results obtained independently
from finite element modelling (FEM) of the caisson uplift problem (Deng and Carter, 1999), and this
comparison is also provided in Figure 69. The problems considered here all correspond to cases for which
there also exists experimental data, including measured uplift capacities. The values of the undrained shear
strengths of the soils considered in this study were taken as those quoted in the original source papers. A
suitably large value of the loading rate was adopted in the finite element predictions to simulate the
undrained response.
It can be observed in Figures 69 and 70 that the results from the ANN model are closer to the observed
values for the high range of uplift capacities, and the FEM model predictions are closer to the observed
values for the lower uplift capacities. The reason is that the high values of uplift capacity have the largest
effect on the basic measure of performance of the ANN model used in this example, so it will try to fit the
high uplift capacity cases as much as possible in order to maximise performance. However, for the finite
element model, the minimum value of the sum-squared error cannot be constrained, and therefore the error
will be fairly uniformly distributed.
This example demonstrates that the predictions from the ANN are as good or better than those obtained
from FEM-based solutions. However, the development of the ANN model is much easier as compared to
the FEM-based models.
As a soft-computation tool, the artificial neural network is quite robust in nonlinear relationship
modelling. However, the underlying assumption is that the input parameters are inherently reliable. This is
not always the case, particularly where uncertainties exist about the soil strength characteristics. Since fuzzy
logic can provide a systematic method to deal with imprecise and incomplete information, some authors are
in the process of developing fuzzy neural network models for the prediction in geotechnical engineering,
including the prediction of the uplift capacity of suction caissons. Others pursue the combination of
deterministic and stochastic methods of analysis, as described in a previous section of this paper.
12.0 DATABASE SOFTWARE
Most practical geotechnics involves the accumulation of observations and measurements. Obvious
examples are the results of site investigations and project monitoring, and typically include data such as
borehole logs, in situ test data, laboratory test data and the results of engineering calculations. This
information may take the form of text, numerical data or graphical images, and frequently includes all types
for any given project. Storing this data in a form that makes its retrieval quick and convenient is a
requirement of modern geotechnical practice. In most professions, digital data storage is now the norm, and
a plethora of commercial and non-commercial software has been developed to meet the growing need for
storage and retrieval of information. Geotechnical engineering is no exception. There exist a number of
commercial software packages to assist with these tasks, and many geotechnical firms have developed their
own in-house solutions to digital storage of their vital information.
12.1 Case Study - Hong Kong International Airport
A striking example of just what can be done, and the efficiency that is achievable, in the storage and
transmission of geotechnical information via a digital database can be seen in the case study provided by the
construction of the Hong Kong International Airport. Vast amounts of geotechnical data were generated on
this project and shared electronically between interested parties. It is probably one of the biggest, if not the
biggest, geotechnical data management project of its type to date in the world.
The platform on which the new Hong Kong International Airport has been built was created in just two
and a half years, commencing at the end of 1992, and its construction involved a unique combination of
dredging, mining and seawall operations. Three-quarters of the airport platform was reclaimed from the sea
and the remaining quarter is the result of the excavation of two islands Chek Lap Kok and its smaller
neighbour, Lam Chau. The construction project is one of the largest civil engineering endeavours of this
type ever undertaken. Coordination and control of the construction project was vested in the Airport
Authority, Hogn Kong, a body established by an act of the Hong Kong legislature in 1995. A detailed
account of the design and construction of the airport platform has been provided in the book by Plant, Covil
and Hughes (1998).
Extensive site investigation was carried out prior to the reclamation works for the airport, but more was
required during the course of the Site Preparation Contract. Investigations to assess the performance of the
reclamation were particularly important. In addition, investigations were required to refine construction
procedures, to assess the density of the as-placed fill materials, and for quality control. Additional
information was required for the design of subsequent works, such as tunnel alignments, runways, buildings
and services. Instrumentation was installed to monitor the settlement of the reclamation and to assess the
ground treatment measures used on site.
The investigations carried out at the various stages were numerous and varied. They included drill holes,
trial pits, vibrocores, cone penetration tests (CPTs), in situ field vane tests, seismic surveys, Spectral
Analysis of Surface Waves (SASW), and associated laboratory testing. All factual data collected during site
investigations, and some of the interpretative reports produced using this data, were made available by the
Airport Authority to relevant contractors. The information comprised factual reports, interpretative reports,
raw data, drawings, seismic surveys, digital data and photographic records.
According to Plant et al. (1998), the [Airport] Authority realised early in the project that vast amounts of
ground investigation and instrumentation would be generated during site preparation and this corpus of data
could become unwieldy unless efficiently managed. The Authoritys strategy for managing this body of
data was to establish and maintain a comprehensive database for airport ground investigations.
The database was designed to incorporate the following functions:
a repository of digital ground investigation data,
stratigraphic correlation,
geological correlation with instrument installations,
correlations between geology, field tests, laboratory tests and stratigraphy, and
comparison of geotechnical, survey and engineering information.
According to Plant et al. (1998) the contents of over 400 site investigation reports were translated or
input as digital data. The database contains an accessible master copy of all available geotechnical
exploration point data, including boreholes, prebores, CPTs, instrument installations, laboratory test results
and in situ tests.
As indicated in the list above, one of the important uses of the database appears to have been in the
correlation of data of various types, thus providing a degree of confidence in the various geological and
geotechnical models developed. In particular, geophysical methods were used to assess the stratigraphy and
these data were correlated with the geological descriptions from boreholes and with the CPT data.
In May 1993, the Authority purchased the software program gINT (Geotechnical Computer
Applications, 1991) to provide electronic database facilities. In common with a number of the available
database packages, the gINT program organises data into a tiered system. In this particular case the three
tiers consisted of project level data, hole level data, and depth level data. In common with the architecture of
most electronic databases, the data are stored in specified fields. Access to the data as output is via a set of
standard templates, created to produce output of the desired form. The data sets used within the database for
the Hong Kong International Airport are indicated in Table 13, which provides an indication of the scale of
this data management exercise. An example of the graphical output providing a representation of a small
portion of this data is presented in Figure 71.
The need for a standard format for interchanging geotechnical data should be self-evident. Its desirability
has been well documented, e.g., Greenwood and Rathery (1992), Nicholls et al. (1996). A standard
interchange format enables different organisations who have different database management systems or who
have different system configurations to pass data from one to another in a standard way.
Plant et al. (1998) note that the United Kingdom-based Association of Geotechnical Specialists (AGS)
was probably the first organisation to attempt standardisation of an interchange format for ground
investigation data (AGS, 1992, 1994). Contractors working on the airport project were obliged to present
factual data digitally in the specified AGS format. The AGS system comprises a set of rules for setting up
electronic files and a data dictionary. As noted by Plant et al., it is necessary sometimes to establish
alternative or additional data dictionaries to suit individual projects.
Significant use was made of the gINT system to produce output in the form of cross-sections and
interpretation of seismic reflectors. The output was enhanced by the use of CADD software for production
of final drawings, e.g., Figure 71.
The Hong Kong International Airport case study provides a very convincing illustration of the
importance of database software in the large-scale construction projects, particularly where vast amounts of
geotechnical data need to be stored, processed, retrieved, interpreted, visualised and analysed, by a variety of
interested parties. It is difficult to imagine coping with such quantities of data and such heavy processing
demands without the availability of modern computer technology and database software. This particular
case study also demonstrates the importance of standards, such as the one developed by AGS, for the
interchange of geotechnical information. The need for these standards should be self-evident.
Table 13. Data sets in the Hong Kong International Airport database (after Plant et al., 1998)
Dataset Description Contents
No. of
exploration
points
PAA Main database All the more important platform-wide data,
including instrumentation, laboratory testing,
deep CPTs and investigation penetrating into the
natural geology below the base of the fill or
marine (previous) layers.
4499
CPT Shallow CPT
database
Platform-wide dredge level CPTS. 3086
QCPT Quality control
database
CPTs through the Type C, sand, fill in the
middle and northern parts of the platform,
carried out to demonstrate the density of the
sand fill.
522
TB Passenger
terminal building
foundations
All the prebores carried out for the foundation
works of the passenger terminal building.
792
VIBRO Vibrocompaction
database
CPTs carried out in the Sand, Fill Type C, and in
SPC Works Areas HI/6 and H8/1 where
vibrocompaction ground treatment was carried
out.
748
302 302 Foundations Prebores for the area of Contract 302. 280
303 303 Foundations Prebores for the area of Contract 303. 118
INST Instrumentation
database
Duplicate set of exploratory hole data of
Instrumentation holes only, rationalised for
specific plots of instrument data against drilling
records.
399
While the scale of the
Hong Kong ai r por t
construction project may
have demanded the use of
mo d e r n c o mp u t e r
technology to handle all the
geotechnical data, it has
also pointed the way
forward for projects of more
modest scale. The benefits
of convenient data retrieval
would seem to be well
worth the investment in the
technology, at least for all
but the smallest projects. At
the very least, in most
geotechnical consultancies,
being able to conveniently
locate and retrieve data
from a history of similar or
related projects would be a
boon on many jobs.
13.0 EXPERT
SYSTEMS
Expert system technology started to emerge from the Artificial Intelligence research laboratories during
the 1980s. However, even now, more than a decade later, there are few expert systems that have been
developed for geotechnical work. The difficulties with the wider adoption of this technology are probably
related to both technical and psychological factors. These events are probably not dissimilar to the inertia
that apparently existed when computers were first introduced into common usage.
Expert systems are computer programs that collect human experience in solving problems by using
heuristics (rules of thumb). Forsyth (1986) proposed that expert systems should:
be able to generate conclusions or advice from uncertain, incomplete, or conflicting information or
knowledge,
must be able to explain their reasoning processes and conclusions,
must be able to separate stored knowledge and inference mechanisms so that they can be modified or
edited without affecting each other,
be able to modify, delete or enrich the stored knowledge incrementally,
deliver advice, usually non-numerical, as its output.
Expert systems are typically rule-based, containing numerous rules-of-thumb that express the judgement
processes of experts. Generally they do not involve the use of procedural algorithms, as are common in
deterministic numerical analysis.
There are several reasons for developing special expert systems software in geotechnical engineering,
and these include possibly the need to compensate for a lack of available experts, the need to save the cost of
employing experts, the need to provide additional assistance to experts, and the need to train future experts.
A detailed exposition on these needs, and the techniques adopted recently to fulfill them, has been provided
by Wong (1990).
Many of the major problems confronted in geotechnical engineering are caused by or related to the
complicated engineering behaviour of soils and rock masses. This behaviour depends on the complete
geological history of the deposit, as represented by the size, shape, mineral composition and packing of the
particles, the stress history that has been experienced, the pore fluid and other factors (Wroth and Houlsby,
1985). In addition, many of these properties also vary with time and the applied stress level. Success in the
practice of geotechnical engineering has usually required the accumulation and appropriate application of
experience, sometimes hard-won experience. Experience and knowledge of soil or rock mechanics and
Figure 71 : Example of CADD output Hong Kong International Airport
(Plant et al., 1998) [Supplied by courtesy of Airport Authority Hong Kong].
geology may be obtained by formal training, but this does not guarantee that reliable practice of geotechnical
engineering will follow. There are several problems in applying what amounts to second-hand experience,
including the following:
difficulties in identifying potential problems,
problems with interpretation of existing knowledge,
lack of appreciation of the assumptions on which underlying theories are based,
difficulties in justifying various assumptions for a given site,
difficulties in keeping up-to-date with the latest knowledge,
difficulties in learning how various facets of knowledge are integrated into practice, and
difficulties in focusing on critical information when a problem is encountered for the first time.
Assessing and assimilating all
the necessary factors involved in
any given geotechnical problem,
and being able to assess their
relative importance, requires
considerable skill and experience.
Of course, it is possible that
geotechnical engineers will learn
from their own mistakes, but
generally this is not an efficient or
economical means of gaining
experience. More efficient ways
of passing on experience for the
future need to be found.
A comparison of the way in
which humans conventionally
solve problems of this type, and
the way in which an expert
system is usually designed to
solve such problems has been
indicated by Wong (1990), see
Figure 72. It is apparent from this
comparison that expert system
technology is well-suited to assist
in the task of making the sort of
judgements that are often required in geotechnical practice. It is perhaps curious then that so few examples
exist today in geotechnical engineering practice: most of those that do exist were developed in the 1980s. It
is also curious that there seems to have been little progress in the application of this technology to
geotechnical practice in recent years. This is possibly related to the difficulty and costs associated with their
development. Examples of those that have been developed include:
SOILIDEN (Alim and Munro, 1987) an expert system to assist with the identification of soil on the
basis of field observations and soil characteristics as described in the British code of practice,
CP20004:1972.
CONE (Mullarkey, 1985; as extracted from Siller, 1987) a knowledge-based system to classify soils
and assess their shear strength based on cone penetrometer data
RETWALL (Hutchinson, 1985) a knowledge-based system to evaluate whether a retaining wall is
required in a given situation and to evaluate which wall type is most appropriate.
SOILCON (Wharry and Ashley, 1986; as extracted from Siller, 1987) an expert system designed to
advise the user about the level of geotechnical investigation necessary in order to reduce the associated
risk.
LOGS (Lok, 1987; as extracted from Adams et al., 1989) a system developed to treat information form
several boring logs and to provide the user with two-dimensional profiles.
FOOTER (Shukula, 1988) a system designed to propose a footing design for each column or shear
wall in a structure.
CONVENTIONAL
APPROACH
EXPERT SYSTEM
APPROACH
1. Understand the problem 1-Provide information need
by the expert system
2. Divide the problem into
smaller sub-problems if
necessary
2. Solve the problem by
drawing conclusions from
stored knowledge and
given information
3. Solve the problem or sub-
problems by:
-Using own knowledge and
experience,
-Getting help from
experienced people,
-Getting help from published
literature.
3. User justifies the solution
4. Justify the solution
Figure 72 : Comparison of problem solving techniques
(after Wong, 1990)
PILE (Santamarina and Chameau, 1987) - a system to assist users in selecting the appropriate type of
pile foundation according to aspects such as soil characteristics, pile loads, installation conditions, and
other factors.
SUPILE (Wong, 1990) a system designed to assist foundation engineers with pile design.
SITECLAS (Wong et al., 1989) an expert system designed to assist with site classification.
14.0 COMPUTER VISUALISATION IN GEOTECHNICS
Largely because of the complexities in geometry of many geotechnical problems, and the often vast
quantities of data required to describe such problems, computer visualisation tools are being used
increasingly in geotechnical research and practice. It is now possible to gain a comprehensive and relatively
rapid appreciation of the geometric complexity of many geotechnical challenges with the aid of modern
computing technology.
An excellent example of this use has been the use of Geographic Information Systems (GIS) for the
mapping of geology and landslides and representation and assessment of the risk of landsliding in an area of
New South Wales, near the city of Wollongong, south of Sydney, Australia. This study is being conducted
by the University of Wollongong with the collaboration of the Wollongong City Council, the Rail Services
Australia and the Australian Geological Survey Organisation (Flentje, 2000). Illustrations of the power of
computer visualisation tools being utilised on this project may been seen in Figures 73, which show
isometric views of the geomorphology of the study area, the slopes in the area, as well as a composite
representation of the geology superimposed on known landslide areas. The ability of these graphical
techniques to represent clearly the salient aspects of such a study are obvious.
There can be no doubt that the declining cost of both hardware and graphics software tools, and the
increase in user friendliness of these tools will lead to much greater use of them in the future. The
attraction of presenting such vivid and meaningful representations of this type of data should ensure that in
the near future these graphics tools will become as common as the geologist pick in the tool bag of most
geotechnical specialists.
Figure 73 : Example of computer visualisation (Flentje, 2000)
15.0 THE FUTURE
Like most branches of engineering, geotechnics will continue to benefit in the future from the on-going
developments in computer hardware and software. The ability to model real problems using mathematical
techniques will need to be enhanced. Computer hardware will become more affordable and developments in
software, though understandably often lagging behind those in hardware, will also make the process of
modelling more convenient and cost-effective. Potentially, numerical modelling could become routine in
the very near future for a large range of geotechnical problems. What then are the future requirements, and
what are the impediments, if any, to this advance? Are there areas where major research effort is required?
In the past, most of the geomechanics applications software was developed in universities and research
institutions, or their spin-off organisations. When suitable hardware platforms became ubiquitous, as a
result of the fall in hardware prices, a number of these software packages, developed first as research or
teaching tools, were commercialised. This development has occurred almost entirely as a cottage industry,
labour intensive and with only a relatively small niche market. It has meant that the unit price of many
software packages has remained relatively high. It is likely that this trend will continue in the future.
Because of limited demand, it is unlikely that major software houses will take up the development of highly
specialised geotechnical software. Possible exceptions include the general-purpose finite element packages
now reaching the market, some of which contain a variety of useful soil models, together with their other
multifarious capabilities.
It is possible to identify some areas in which further development is required, and these are:
constitutive models for many geomaterials, interfaces and joints and their critical evaluation, including:
constitutive models for cyclic behaviour and fundamental approaches for liquefaction (instability),
constitutive models for natural soils (with structure), e.g., sands, as opposed to reconstituted soils,
constitutive models for multi-phase materials, and constitutive models for composite material systems in
reinforced soil construction,
controlled laboratory and field tests to validate numerical and constitutive models,
micromechanical modelling of assemblages of particles using the distinct element model or other
suitable approaches to infer macroscopic behaviour,
appropriate test devices, particularly for interfaces including measurement of pore water pressures,
robust and reliable numerical procedures to obtain unique and realistic solutions for problems involving
discontinuities, degradation and softening,
robust and reliable numerical procedures based on mesh adaptation,
more comprehensive expert systems for geotechnical practice, and
numerical procedures to model accurately in situ penetration tests, e.g., the CPT test.
The analyses required for penetration tests must involve very large strains, rupture and moving
boundaries. The work of Hu and Randolph (1996) shows great promise in this area.
The recent development of two and three-dimensional finite element procedures for limit analysis and
load-path analysis, and software encoding sophisticated constitutive models, for instance the heirachical
DSC models, now provide the means to solve a wide range of geotechnical problems involving static,
repetitive, dynamic and time-dependent loading and material response.
It is clear from the above list, which is by no means complete, that much work is still required in the field
of constitutive modelling. Probably the greatest impediment to the use of many models proposed already is
the difficulty in assessing what values should be assigned to the input parameters. Many models in the
literature have a number of parameters that are difficult to evaluate directly from either field or laboratory
tests. If model parameters have a physical basis, then there is greater chance of parameter measurement, and
hence the eventual adoption of the model in practice. Obviously, developments in this area must go hand-in-
hand with parallel developments in experimental techniques. It has not always been possible to arrange this
simultaneous advance.
Calibration of new and existing models is also an essential requirement if genuine advances are to be
made. In this respect, the recent efforts made by various workers (a) to calibrate their models against the
results of field and laboratory tests, especially centrifuge data, and (b) to compare the predictions of various
numerical tools for the same set of control problems, is to be applauded. Such comparisons have always
been an essential component of previous advances, and they continue to point the direction forward.
As demonstrated in this paper, there is a pressing need for more and comprehensive validation studies for
numerical simulations. Ultimately these should lead to sound guidelines for geotechnical practitioners to
follow.
The future should also see the advancement of the so-called soft computing tools. The need to develop
expert systems for geotechnical practice is an obvious example.
16.0 CONCLUSIONS
This paper has provided a description of the use of computers and numerical methods in geotechnical
practice and research. It has been intentionally and necessarily selective, reflecting the experience of the
authors and their colleagues.
Sophisticated numerical models have been used regularly in geotechnical research for the past thirty
years or more, but they have now moved beyond the boundaries of research institutions. It should be clear
from the limited set of examples presented in this paper that the use of relatively sophisticated non-linear
and time-dependent numerical models is now a common occurrence in geotechnical practice. The
availability of economical personal computers, the release commercially of a number of robust software
packages, the training of a new generation of engineers, familiar and comfortable with the world of
electronic computing (including many older colleagues willing to embrace the new technology), have all
played their part in this evolution. Practitioners are now using these tools in their day-to-day work.
Numerical models have their place in the toolkit of the modern geotechnical engineer. That place may be to
examine possible behavioural mechanisms, or to throw light on the dominant parameters of a problem, or to
help understand field performance.
A discussion has also been presented about some limitations of the current techniques and a few of the
developments required in the future. In particular, constitutive models that capture the essential features of
soil or rock behaviour, and are based on physically meaningful and simple to measure parameters, are
needed. Models for cyclic loading and for natural (structured) soils are examples. The incorporation of
these deterministic constitutive models in a probabilistic assessment of future behaviour also requires
development work.
Further research and development work is also required in the field of soft computing as applied to
geotechnical problems. Much greater effort is required to validate and calibrate numerical models and the
associated computer software.
17.0 ACKNOWLEDGEMENTS
Some of the results presented in this paper were the outcome of research supported by various grants
from the Australian Research Council and the National Science Foundation, Washington, D.C. The authors
wish to acknowledge the following individuals for kindly providing some of the material presented in this
paper and for valuable assistance in its preparation: Dr H. Taiebat, Dr M. Rahmann, Mr K. Islam, Dr M. Liu,
Mr W. Deng, Mr C. Covil, Dr P. Flentje. The permission of Dr P. Flentje to reproduce the image taken from
the Wollongong Landslide Hazard and Risk study conducted jointly by the Australian Geological Survey
Organisation and the University of Wollongong is gratefully acknowledged. The permission of Mr Craig
Covil to publish the CADD image from the Hong Kong International Airport project, supplied by courtesy
of the Airport Authority Hong Kong, is also gratefully acknowledged.
18.0 REFERENCES
Adams, T.M., Christiano, P. and Hendrickson, C. (1989). Some expert system applications in geotechnical
engineering, Proc. Conf. On Foundation Engineering: Current Principles and Practices. American
Society of Civil Engineers, New York, NY, USA.
Addenbrooke, T.I., Potts, D.M., and Puzrin, A.M. (1997). The influence of pre-failure soil stiffness on the
numerical analysis of tunnel construction, Gotechnique, Vol. 47, No.3, pp. 693-712.
Adhikary, D.P., Dyskin A.V. and Jewell, R.J. (1995). Modelling of flexural toppling failures of rock
slopes. Proc. 8
th
Int. Congress on Rock Mechanics, Tokyo, Japan. pp. 25-29.
Adhikary, D.P., Dyskin, A.V. and Jewell, R.J. (1996). Numerical modelling of the flexural deformation of
foliated rock slopes. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr. Vol. 33, No. 6, pp. 611616.
Aitchison, G.D. (1956). The circumstances of unsaturation in soils with particular reference to the
Australian environment, Proc. 2
nd
. Aust.-NZ Conf. Soil Mech & Foundn. Eng., pp. 173-191.
Aitchison, G.D. (1961). Relationships of moisture stress and effective stress functions in unsaturated soils,
Pore pressure and suction in soils, London, pp. 47-52.
Akiyoshi, T., Fang, H.L., Fuchida, K. and Ha, Matsumoto (1996). A nonlinear seismic response analysis
method for saturated soil-structure systems with absorbing boundary, Int. J. Num. Analyt. Meth.
Geomech., Vol. 20, pp. 307-329.
Alim, S. and Munro, J. (1987). PROLOG-based expert systems in civil engineering, Proceedings,
Institution of Civil Engineers, Part2, pp. 1-14.
Alonso, E.E., Gens, A. and Josa, A. (1990). A constitutive model for partially saturated soils,
Gotechnique, Vol. 40, pp. 405-430.
Anderheggen, E. and Knopfel, H. (1972). Finite element limit analysis using linear programming,
International Journal of Solids and Structures, Vol. 8, pp. 14131431.
Assadi, A. and Sloan, S.W. (1991). Undrained stability of a shallow square tunnel, Journal of the
Geotechnical Division, ASCE, Vol. 117, pp. 11521173.
Association of Geotechnical Specialists (1992). Electronic transfer of geotechnical data from ground
investigations, March 1992.
Association of Geotechnical Specialists (1994). Electronic transfer of geotechnical data from ground
investigations, 2
nd
edition, July 1994.
Atkinson, J.H. (1993). A note on modelling small strain stiffness in Cam clay, Predictive soil mechanics,
Thomas Telford, London.
Atkinson, J.H., Richardson, D., and Stallebrass, S.E. (1990). Effect of stress history on the stiffness of
overconsolidated soil, Gotechnique, Vol 40, No. 4, pp. 531-540.
Bauer, E. (1996). Calibration of a comprehensive hypoplastic model for granular materials, Soils and
Foundations, Vol. 36, No. 1, pp. 13-26.
Bazant, Z.P. (1994). Nonlocal damage theory based on micromechanics of crack interactions, J. of Eng.
Mech., ASCE, Vol. 120, pp. 593-617.
Bazant, Z.P. and Lin, F.B. (1988). "Non-local yield limit degradation", Int. J. Num. Meth. Eng. Vol. 26, pp.
1805-1823.
Beer, G. (1995) Modelling of inelastic behaviour of jointed media with the boundary element method,
Proc. 4
th
Int.Conf. Computational Plasticity, Swansea (Pineridge Press), pp. 1761-1772.
Beer, G. and Watson, J.O. (1992) Introduction to Finite and Boundary Element Methods for Engineers, John
Wiley & Sons, New York.
Belytschko, T., Fish, J. and Englemann, B.E. (1988). A finite element with embedded localization zones,
Comp. Methods Appl. Mech. Eng. Vol. 70, pp. 59-89.
Biron, A. and Chasleux, G. (1972). Limit analysis of axisymmetric pressure vessel intersection of arbitrary
shape, International Journal of Mechanical Sciences, Vol. 14, pp. 2541.
Bolzon, G., Schrefler, B.A. and Zienkiewicz, O.C. (1996). Elastoplastic soil constitutive laws generalized
to partially saturated states, Gotechnique, Vol. 46, pp. 279-289.
Borja, R.I. (1991), Cam clay plasticity, part II: Implicit integration of constitutive equations based on
nonlinear elastic stress prediction, Comput. Meth. Appl. Mech. Eng., Vol. 88, pp. 225-240.
Borja, R.I. and Lee, S.R. (1990). Cam clay plasticity, part I: Implicit integration of constitutive relations,
Comput. Meth. Appl. Mech. Eng., Vol. 78, pp. 49-72.
Bottero, A., Negre, R., Pastor, J. and Turgeman, S. (1980). Finite element method and limit analysis theory
for soil mechanics problems, Computer Methods in Applied Mechanics and Engineering, Vol. 22, pp.
131149.
Brand, E.W. and Premchitt, J. (1989). Moderators report for the predicted performance of the Muar test
embankment", Proc. Int. Symp. Trial embankments on Malaysian marine clays, November 1989. Kuala
Lumpur.
Brinkgreve, R.B.J. and Bakker, H.L. (1991). Non-linear finite element analysis of safety factors,
Proceedings 7
th
International Conference on Computer Methods and Advances in Geomechanics, Cairns,
May, G. Beer, J.R. Booker and J.P. Carter Eds., Vol. 2, pp. 1117-1122.
Brinkgreve, R.B.J. and Vermeer, P.A. (1995). A new approach to softening plasticity, Proc. 5
th
Int. Symp.
Num. Models Geomech., Balkema, Rotterdam, pp. 193-202.
Brinkgreve, R.B.J. and Vermeer, P.A. (1998). PLAXIS Finite Element Code for Soil and Rock Analysis.
Version 7.1.
Britto A.M. and Gunn M.J. (1987), Critical state soil mechanics via finite elements, Ellis Horwood Ltd.,
Chichester, U.K.
Burland, J.B. (1989). Small is beautiful - the stiffness of soils at small strains, Ninth Laurits Bjerrum
Lecture, Canadian Geotechnical Journal, Vol.26, pp. 499-516.
Burland, J.B. (1990). On the compressibility and shear strength of natural soils, Gotechnique, Vol. 40,
pp. 329-378.
Capsoni, A. and Corradi, L. (1997). A finite element formulation of the rigidplastic limit analysis
problem, International Journal for Numerical Methods in Engineering, Vol. 40, pp. 20632086.
Carter, J.P. and Airey, D.W. (1994). The engineering behaviour of cemented marine carbonate soils,
Geotechnical Engineering: Emerging Trends in Designs and Practice, Saxena (ed), pp. 65-101.
Carter, J.P. and Xiao, B. (1993) Coupled Finite Element and Boundary Element Method for the Analysis of
Anisotropic Rock Masses, Proc. International Symposium on Application of Computer Methods in Rock
Mechanics and Engineering, Xian, China, Vol. 1, pp. 249-258.
Casagrande, A. (1932). The structure of clay and its importance in foundation engineering, J. Boston
Society of Civil Engineers, Vol. 19, pp. 168-209.
Casagrande, A. (1976). Liquefaction and cyclic deformation of sands a critical review, Harvard Soil
Mech. Series No. 88, Harvard Univ. Press, Cambridge, MA, USA.
Castro, G. and Poulos, S.J. (1977). Factors affecting liquefaction and cyclic mobility, J. of Geotech. Eng.,
ASCE, Vol. 103, pp. 501-516.
Cherubini, C., Giasi, C.I. and Rethati, L. (1993). The coefficients of variation of some geotechnical
parameters, Proceedings Conference on Probabilistic Methods in Geotechnical Engineering, Canberra,
February, K.S. Li and S-C.R. Lo Eds., pp. 179-183.
CombInfo 1.0, (1999). Computer Program, Institute of Structural Engineering, Swiss Federal Institute of
Technology ETH, Switzerland
Cotecchia, F. and Chandler, R.J. (1997). The influence of structure on the pre-failure behaviour of a natural
clay, Gotechnique. Vol. 47, pp.523-544.
Cundall, P.A. and Board, M. (1988) A microcomputer program for modelling large-strain plasticity
problems, Proc. Int. Conf. Numerical Methods in Geomechanics, Balkema, Rotterdam, pp. 2101-2108.
Cundall, P.A. and Strack, O.D.L. (1979). A discrete numerical model for granular assemblies,
Gotechnique, Vol. 29, pp. 47-65.
Dafalias, Y.F. (1979). A bounding surface plasticity model, Proc., 7
th
Can. Eng. Appl. Mech., Sherbrooke,
Canada.
Davies, E.H. and Booker, J.R. (1973). The effect of increasing strength with depth on the bearing capacity
of clays, Gotechnique, Vol. 23, No. 4, pp. 551-563.
Davis, E.H. (1969). Theories of plasticity and the failure of soil masses, Soil Mechanics Selected Topics,
I.K. Lee (ed), Butterworth. London, pp. 341354.
Davis, R.O. and Berrill, J.B. (1982). Energy dissipation and seismic liquefaction in sands, Earthquake
Eng. and Struct. Dyn., Vol. 19, pp. 59-68.
Davis, R.O. and Berrill, J.B. (1988). Energy dissipation and liquefaction at Port Island, Kobe, Bulletin of
New Zealand Nat. Soc. Earthquake Eng., Vol. 31, pp. 31-50.
de Borst, R. (1991). Simulation of strain localization: A reappraisal of the Cosserat continuum, Eng.
Comp., Vol. 8, pp. 317-332.
de Borst, R., Sluys, L.T., Mhlhaus, H.B. and Pamin, J. (1993). Fundamental issues in finite element
analyses of deformation, Int. J. Eng. Computation, Vol. 10, pp. 99-121.
de Buhan, P. and Maghous, S. (1995). A straightforward numerical method for evaluating the ultimate
loads of structures, European Journal of Mechanics, A/Solids, Vol. 14, No. 2, pp. 309328.
de Josselin de Jong, de G., and Verruijt, A. (1969). Etude photo-lastique dun empilement de disques.
Cahiers du GroupFranais Rhologie, Janvier, Vol. II, pp. 73-86.
DeMaggio, F.L. and Sandler, I.S. (1971). Material model for granular soils. J. Eng. Mech. Div., ASCE,
Vol. 97, pp. 935-950.
Demuth, H. and Beale, M. (1996), Neural Network Toolbox For Use with MATLAB, The MATH WORK
Inc.
Deng, W. and Carter J.P. (1999). Predictions of the Vertical Pullout Behaviour of Suction Caissons,
Research Report R797, Centre for Geotechnical Research, The University of Sydney.
Desai, C.S. (1995). Constitutive modelling using the disturbed state as microstructure self-adjustment
concept, Continuum Models for Materials with Microstructure. Mhlhaus, H.B. (ed), pp. 239-296.
Desai, C.S. (1999). Mechanics of Materials and Interfaces: The Disturbed State Concept, CRC Press, Boca
Raton, FL, USA, in press.
Desai, C.S. (2000). Liquefaction using disturbance and energy approaches, J. Geotech. and Geoenv. Eng.,
Vol. 127, No. 7, pp. 618-630.
Desai, C.S. and Abel, J.F. (1972). Introduction to the finite element method. Van Nostrand Reinhold, New
York.
Desai, C.S. and Fishman, K.L. (1991). Plasticity based constitutive model with associated testing for
joints, Int. J. Rock Mech. and Min. Sc., Vol. 4, pp. 15-26.
Desai, C.S. and Ma, Y. (1992). Modelling of joints and interfaces using the disturbed state concept, Int. J.
Num. Analyt. Meth. Geom., Vol. 16, pp. 623-653.
Desai, C.S. and Rigby, D.B. (1997). Cyclic interface and joint shear device including pore pressure
effects, J. Geotech. and Geoenv. Eng., ASCE, Vol. 123, pp. 568-579.
Desai, C.S. and Salami, M.R. (1987). A constitutive model and associated testing for soft rock, Int. J.
Rock Mech. Min. Sc., Vol. 24, pp. 299-307.
Desai, C.S. and Siriwardane, H.J. (1984). Constitutive Laws for Engineering Materials with Emphasis on
Geologic Materials, Prentice-Hall, Englewood Cliffs, NJ, USA.
Desai, C.S. and Toth, J. (1996). Disturbed state constitutive modelling based on stress-strain and non-
destructive behavior, Int. J. Solids and Structures, Vol. 33, pp. 1619-1650.
Desai, C.S. and Varadarajan, A. (1987). A constitutive model for quasistatic behavior of rock salt, J.
Geophys. Res., Vol. 92, pp. 11445-11456.
Desai, C.S., Basaran, C. and Zhang, W. (1997). Numerical algorithms and mesh dependence in the
disturbed state concept, Int. J. Num. Meth. in Eng., Vol. 40, pp. 3059-3083.
Desai, C.S., Park, I.J. and Shao, C. (1998). Fundamental yet simplified model for liquefaction instability,
Int. J. Num. Analyt. Meth. Geomech., Vol. 22, pp. 721-748.
Desai, C.S., Samtani, N.C. and Vulliet, L. (1995). Constitutive modeling and analysis of creeping slopes,
J. Geotech. Eng., ASCE, Vol. 121, pp. 43-56.
Desai, C.S., Sharma, K.G., Wathugala, G.W. and Rigby, D.B. (1991). Implementation of hierarchical single
surface
0
and
1
models in finite element procedure, Int. J. Num. Analyt. Meth. Geomech., Vol. 15, pp.
649-680.
Desai, C.S., Somasundaram, S. and Frantziskonis, G. (1986). A hierarchical approach for constitutive
modeling of geologic materials, Int. J. Num. Analyt. Meth. Geom., Vol. 10, pp. 225-257.
Desai, C.S., Zaman, M.M., Lightner, J.G. and Siriwardane, H.J. (1984). Thin-layer element for interfaces
and joints, Int. J. Num. Analyt. Meth. Geomech., Vol. 8, pp. 19-43.
Drucker, D. C., Greenberg H. J. and Prager, W. (1952). Extended limit design theorems for continuous
media, Quarterly Journal of Applied Mathematics, Vol. 9, pp. 381389.
Duncan, J.M. (1996). State of the Art: Limit Equilibrium and Finite-Element Analysis of Slopes, Journal
of Geotechnical Engineering, Vol. 122, pp. 577-596.
Duncan, J.M. and Chang, C.Y. (1970). Nonlinear analysis of stress and strain in soils, J. Soil Mech.
Found. Div., ASCE, Vol. 96, pp. 1629-1653.
Earth Technology Corp. (1986). Pile segment tests Sabine Pass, ETC Report No. 85-007, The Earth
Tech. Corp., Long Beach, CA, USA.
Fahey, M. and Carter, J.P. (1993) A finite element study of the pressuremeter test in sand using a non-linear
elastic plastic soil model, Canadian Geotechnical Journal, Vol. 30, pp. 348-362.
Figueroa, J.L., Saada, A.S., Liang, L. and Dahisaria, M.N. (1994). Evaluation of soil liquefaction by energy
principles, J. Geotech. Eng., ASCE, Vol. 120, pp. 1554-1569.
Finnie, I.M. (1993). Performance of shallow foundations in calcareous soil. PhD Thesis, University of
Western Australia.
Flentje, P. (2000) Private communication.
Flood, I., and Kartam, N. (1994). Neural networks in civil engineering I: Principles and understanding, J.
Computing in Civ. Engrg., ASCE, Vol. 8, No. 2, pp. 131-148.
Forsyth, R. (1986). The anatomy of expert systems, Artificial Intelligence: Principles and Applications,
Yazdani, M., Chapman and Hall (eds).
Fredlund, D.G. (1984). Analytical Methods for Slope Stability Analysis, Proc. 4
t h
Landslides -
International Symposium, Toronto, September, pp. 229-250.
Garrett, J. H. Jr. (1994). Where and why artificial neural networks are applicable in civil engineering, J.
Computing in Civ. Engrg. ASCE, Vol. 8, No. 2, pp. 129-130.
Geiser, F., Laloui, L., Vulliet, L. and Desai, C.S. (1997). Disturbed state concept for partially saturated
soils, Proc. Conf. on Num. Models in Geomech., Montreal, Canada.
Gens, A. and Nova, R. (1993). Conceptual bases for a constitutive model for bonded soils and weak rocks,
Geotechnical Engineering of Hard Soils Soft Rocks, Anagnostopoulos et al. (eds), Vol. 1, pp. 485-494.
Geotechnical Computer Applications (1991). gINT, gEOTECHNICAL INTegrator software, Version 3.2
Documentation, 8 July 1991.
Ghaboussi, J. and Sidarta, D.E. (1998). New nested adaptive neural networks (NANN) for constitutive
modelling, Computers and Geotechnics, Vol. 22, No. 1, pp. 29-52.
Ghaboussi, J., Wilson, E.L. and Isenberg, J. (1973). Finite element for rock joints and interfaces, J. of Soil
Mech. and Found. Eng., ASCE, Vol. 99, pp. 833-848.
Goh, A.T.C. (1996). Pile driving records reanalyzed using neural networks, J. Geotechnical Engineering,
ASCE, Vol. 122, No. 6, pp.492-495.
Goodman, R.E., Taylor, R.L. and Brekke, T.L. (1968). A model for the mechanics of jointed rock, J. of
Soil Mech. and Found. Eng., ASCE, Vol. 94, pp. 637-659.
Graham, J. and Li, C.C. (1985). Comparison of natural and remoulded plastic clay, J. Geotechnical
Engineering, ASCE, Vol. 111, pp. 865-881.
Greenwood, J.R. and Rathery, M.G.W. (1992). Collection, transfer, storage and presentation of ground
investigation data, Proc. Int. Conference on Geotechnics and Computers, Paris, Sept. 1992.
Griffiths, D.V. (1980). Finite element analyses of walls, footings and slopes, PhD thesis, University of
Manchester.
Gudehus, G. (1996). A comprehensive constitutive equation for granular materials, Soils and Foundations,
Vol. 36, No. 1, pp. 1-12.
Herle, I. and Gudehus, G. (1999). Determination of parameters of hypoplastic constitutive model from
properties of grain assemblies, Mechanics of Cohesive-Frictional Materials, Vol. 4, No. 5, pp. 461-486.
Hicks, M. (1992). Numerical implementation of a double-hardening soil model for use in boundary value
problems, Proc. 7
th
Int. Conf. Computer Methods and Advances in Geomechanics. Cairns. Balkema.
Rotterdam. Vol. 3, pp. 1705-1710.
Hicks, M.A. (1995). Computation of localisation in undrained soil using adaptive mesh refinement, Proc.
5
th
Int. Symp. Num. Models Geomech., Balkema, Rotterdam, pp. 203-208.
Hocking, G. (1977). Development and application of the boundary integral and rigid block methods for
geotechnics, PhD thesis, Imperial College, UK.
Hodge, P.G., Jr. and Belytschko, T. (1968). Numerical methods for the limit analysis of plates, Journal of
Applied Mechanics, Transactions of the ASME, Series E, Vol. 35, pp. 796802.
Hu, Y. and Randolph, M.F. (1996). A practical numerical approach for large deformation problems in soil,
Int. J. for Numerical and Analytical Methods in Geomechanics, Vol. 22, pp. 327-350.
Huang J.T. and Airey, D.W. (1998). Properties of an artificially cemented carbonate sand, J. Geotechnical
and Geoenvironmental Engineering. ASCE. Vol. 124, pp. 492-499.
Huang, J.T. (1994). The effects of density and cementation of soils. PhD Thesis, University of Sydney.
Huh, H. and Yang, W.H. (1991). A general algorithm for limit solutions of plane stress problems,
International Journal of Solids and Structures, Vol. 28, No. 6, pp. 727738.
Hutchinson, P.J. (1985). An expert system for the selection of earth retaining structures, Masterss Thesis,
University of Sydney, Australia.
Ishihara, K. (1993). Liquefaction and flow failure during earthquakes, Gotechnique, Vol. 43, pp. 351-
415.
Islam, M.K. (1999). Modelling the behaviour of cemented carbonate soils, PhD Thesis, University of
Sydney.
Jamiolkowski, M., Lancellotta, R. and Lo Presti, D. (1999). Pre-failure deformation characteristics of
geomaterials, Vol. 1, Balkema, Rotterdam.
Janbu, N. (1963). Soil compressibility as determined by oedometer and triaxial tests, European Conf. On
Soil Mechanics and Foundation Engineering, Wiesbaden, Germany, pp. 19-25.
Jardine, R.J., Potts, D.M., Fourie, A.B., and Burland, J.B. (1986). Studies of the influence of non-linear
stress-strain characteristics in soil-structure interaction, Gotechnique, Vol. 36, No. 3, pp. 377-396.
Jiang, G.L. (1994). Regularised method in limit analysis, Journal of the Engineering Mechanics Division,
Proceedings of the ASCM, Vol. 120, pp. 11791197.
Jiang, G.L. (1995). Nonlinear finite element formulation of kinematic limit analysis, International
Journal for Numerical Methods in Engineering, Vol 38, pp. 27752807.
Jiang, G.-L. and Magnan, J.P. (1997). Stability analysis of embankments: comparison of limit analysis with
methods of slices, Gotechnique, Vol. 47, pp. 857-872.
Kachanov, L.M. (1986). Introduction to Continuum Damage Mechanics. Martinus Nijhuft Publishers,
Dordrecht, The Netherlands.
Kaggwa, W. (2000). Determination of the spatial variability of soil design parameters and its significance
in geotechnical engineering analyses, Proc. Booker Memorial Symposium, Sydney. Balkema,
Rotterdam. In press.
Katona, M.G. (1983). A simple contact-friction interface element with applications to buried culverts, Int.
J. Num. Analyt. Meth. Geomech., Vol. 7, pp. 371-384.
Katti, D.R. and Desai, C.S. (1995). Modeling and testing of cohesive soil using disturbed state concept, J.
of Eng. Mech., ASCE, Vol. 121, pp. 648-658.
Kim, M.K. and Lade, P.V. (1988). Single hardening constitutive model for frictional materials: I. Plastic
potential function, Computers and Geotechnics, Vol. 5, pp. 307-324.
Kolymbas, D. (1988). Generalized hypoelastic constitutive equation, in Constitutive equations for
granular non-cohesive soils, Saada and Bianchini eds, Balkema, Rotterdam, pp. 349-366.
Kolymbas, D. (1991). An outline of hypoplasticity, Archive of Applied Mechanics, Vol. 61, pp. 143-151.
Kondner, R.L. (1963). Hyperbolic stress-strain response: cohesive soils, J. Soil Mech. Found. Div., ASCE,
Vol. 89, pp. 115-143.
Lacasse, S. and Nadim, F. (1997). Uncertainties in characterising soil properties, NGI-Publication 201.
Lade, P.V. and Duncan, J.M. (1975). Elastoplastic stress-strain theory for cohesionless soil, J. Geotech.
Eng., ASCE, Vol. 101, pp. 1037-1053.
Lagioia, R. and Nova, R. (1995). An experimental and theoretical study of the behaviour of a calcarenite in
triaxial compression, Gotechnique. Vol. 45, pp. 633-648.
Lemos, J.V., Hart, R.D. and Cundall, P.A. (1985) A Generalized Distinct Element Programm for Modeling
Jointed Rock Mass, Proc. Int. Symp. on Fundamentals of Rock Joints, Centek Publishers, Lulea,
Sweden, pp. 335-343.
Leroueil, S. and Vaughan, P.R. (1990). The general and congruent effects of structure in natural soils and
weak rocks, Gotechnique, Vol 40, pp. 467-488.
Li, K.S. (1992). Point estimate method for calculating statistical moments. Journal for Engineering
Mechanics, Vol. 118, pp. 1506-1511.
Li, K.S., Lee, I.K. and Lo, S.C.R. (1993). Limit state design in geotechnics, Proc. Conf. Probabilistic
Methods in Geotechnical Engineering, Canberra, February, K.S. Li and S.C.R. Lo (Eds), pp. 29-42.
Lippmann, R.P. (1987). An introduction to computing with neural nets, IEEE Acoustics Speech and Signal
Processing, Vol. 4, No. 2, pp. 4-22.
Liu, M.D. and Carter, J.P. (1999). Virgin compression of structured soils, Gotechnique, Vol. 49, pp. 43-
57.
Liu, M.D. and Carter, J.P. (2000). On the volumetric deformation of reconstituted soils, Int. J. for
Numerical and Analytical Methods in Geomechanics, Vol. 24, pp. 101-133.
Liu, M.D., Carter, J.P., Desai, C.S. and Xu, K.J. (2000). Analysis of the compression of structured soils
using the disturbed state concept, Int. J. Num. Analyt. Meth. Geomech., in press.
Liu, Y.H., Cen, Z.Z. and Xu, B.Y. (1995). A numerical method for plastic limit analysis of 3D structures,
International Journal of Solids and Structures, Vol. 32, No. 12, pp. 16451658.
Lok, M. (1987). LOGS: A prototype expert system to determine stratification profile from multiple boring
logs, Masters Thesis, Dept. of Civil Engineering, Carnegie-Mellon University, USA.
Lyamin, A.V. (1999). ThreeDimensional Lower Bound Limit Analysis Using Nonlinear Programming,
PhD Thesis, University of Newcastle, Australia.
Lyamin, A.V. and Sloan, S.W. (1997). A comparison of linear and nonlinear programming formulations for
lower bound limit analysis, Proceedings 6
th
International Symposium on Numerical Models in
Geomechanics (NUMOG 6), Montreal, Canada, pp. 367373.
Lyamin, A.V. and Sloan, S.W. (2000a). Lower bound limit analysis using nonlinear programming,
International Journal for Numerical Methods in Engineering, submitted.
Lyamin, A.V. and Sloan, S.W. (2000b). Upper bound limit analysis using linear finite elements and
nonlinear programming, 5
th
International Conference on Computational Structures Technology, Leuven,
Belgium, to appear.
Lysmer, J. (1970). Limit analysis of plane problems in solid mechanics, Journal of the Soil Mechanics and
Foundations Division, Proceedings Of the American Society of Civil Engineers, Vol. 96, pp. 13111334.
Maier, G., ZavelaniRossi, A. and Benedetti, D. (1972). A finite element approach to optimal design of
plastic structures in plane stress, Int. J. for Numerical Methods in Engineering, Vol. 4, pp. 455473.
McCormick, G.P. and Fiacco, A.V. (1963). Programming under nonlinear constraints by unconstrained
minimisation: A primaldual method, Research Analysis Corp., McLean, Va., RACTP96.
Meissner, H. (1991). Empfehlungen des Arbeitskreises 1.6 Numerik in der Geotechnik, Abschnitt 1,
Allgemeine Empfehlungen. Geotechnik. Vol. 14, pp. 1-10, (in German).
Meissner, H. (1996). Tunnelbau unter Tage Empfehlungen des Arbeitskreises 1.6 Numerik in der
Geotechnik, Abschnitt 2. Geotechnik, Vol. 19, pp. 99-108, (in German).
Merifield, R.S., Sloan, S.W. and Yu, H.S. (1999). Rigorous solutions for the bearing capacity of two
layered clay soils, Gotechnique, Vol. 49, No. 4, , 471490.
Mitchell, J.K. (1976). Fundamentals of Soil Behaviour. New York, Wiley.
Molenkamp, F. (1981). Elasto-plastic double hardening model Monot. LGM Report CO-218595: Delft
Geotechnics.
Mostyn, G.R. and Li, K.S. (1993). Probabilistic slope analysis State-of-play. Proceedings Conference on
Probabilistic Methods in Geotechnical Engineering, Canberra, February, K.S. Li and S.C.R. Lo (eds), pp.
89-109.
Mroz, Z. (1967). On the description of anisotropic work hardening, J. Mech. Phys. Solids, Vol. 15, pp.
163-175.
Mroz, Z., Norris, V.A. and Zienkiewicz, O.C. (1978). An anisotropic hardening model for soils and its
application to cyclic loading, Int. J. Num. Analyt. Meth. Geomech., Vol. 2, pp. 203-221.
Mhlhaus, H.B. (1995). Continuum Models for Materials with Microstructure, John Wiley, Chichester, UK.
Mhlhaus, H.-B. and Vardoulakis, I. (1987). The thickness of shear bands in granular materials,
Gotechnique, Vol. 37, pp. 271-283.
Mullarkey, P.W. (1985). CONE An expert system for interpretation of geotechnical characterization data
from cone penetrometers, PhD Thesis, Carnegie-Mellon University, USA.
Nagtegaal, J.C., Parks, D.M. and Rice, J.R. (1974). On numerically accurate finite element solutions in the
fully plastic range, Computer Methods in Applied Mechanics and Engineering, Vol. 4, pp. 153177.
National Research Council (1985). Liquefaction of soils during earthquake, Report No. CETS EE01,
National Academic Press, Washington, DC, USA.
Nelder, J.A. and Mead, R. (1965). A simplex method for function minimization, The Computer Journal,
Vol. 7, pp. 308315.
Nemat-Nasser, S. and Shokooh, A. (1979). A unified approach to densification and liquefaction of
cohesionless sand in cyclic shearing, Can. Geotech. J., Vol 16, pp. 659-678.
Ng, C., Bolton, M., and Dasari, G. (1995). The small strain stiffness of a carbonate stiff clay, Soils and
Foundations, Vol. 35, No. 4, pp. 109-114.
Nguyen, D.H., Trapletti, M. and Ransart, D. (1977). Quasilower bounds and upper bounds of collapse
loads of shells of revolution by the finite element method and by nonlinear programming, International
Journal of Nonlinear Mechanics, Vol. 13, pp. 79102, (in French).
Nicholls, R.A., Pycroft, A.S., Palmer, M.J. and Frame, J.A. (1996). From paper to the silicon chip the
growing art of computerized data management, Advances in Site Investigation Practice, Thomas
Telford, London, pp. 186-197.
Niemunis, A. and Herle, I. (1997). Hypoplastic model for cohesionless soil with elastic strain range,
Mechanics of Cohesive-Frictional Materials, Vol. 2, No. 4, pp. 279-299.
Nova, R. (1988). Sinfonietta classica: An exercise on classical soil modeling. Proc. Int. Symposium on
Constitutive Equations for Granular Non-Cohesive Soils, pp. 501-520. Rotterdam, Balkema.
Ohde, J. (1951). Grundbaumechanik, Huette, BD. III, 27, Auflage, (in German).
Oka, F., Yashima, A. Kohara, I. and Adachi, T. (1994). Instability of a viscoplastic model for clay and
numerical study of strain localization, Proc. 3
rd
Int. Woskshop Localisation and Bifurcation Theory for
Soils and Rocks, Balkema, Rotterdam, pp. 237-247.
Ortiz, M., Leroy, Y. and Needleman, A. (1987). A finite element method for localized failure analysis,
Comp. Meth. Appl. Mech. Eng., Vol. 61, pp. 189-214.
Owen D.R.J. and Hinton E. (1980), Finite elements in plasticity: Theory and practice, Pineridge Press,
Swansea, U.K.
Pan, J.P. (1999). The behaviour of shallow foundations on calcareous soil subjected to inclined load, PhD
Thesis, The University of Sydney.
Pande, G.N., Beer, G. and Williams, J.R. (1990). Numerical methods in rock mechanics, John Wiley &
Sons, Chichester.
Park, I.J. and Desai, C.S. (1999). Cyclic behavior and liquefaction of sand using disturbed state concept, J.
Geotech. and Geoenv. Eng., ASCE, in press.
Pastor, J. and Turgeman, S. (1982). Limit analysis in axisymmetrical problems: numerical determination of
complete statical solutions, International Journal of Mechanical Sciences, Vol. 24, No. 2, pp. 95117.
Pastor, J. (1978). Analyse limits: dtermination numrique des solution statiques compltes. Application au
talus vertical, Journal de Mcanique applique, Vol. 2, No. 2, pp. 167196, (in French).
Pastor, M., Rubio, C. Mira, P., Peraire, J. Vilotte, J.P. and Zienkiewicz, O.C. (1992). Numerical analysis of
localization, Proc. 4
th
Int. Symp. Num. Models Geomech., Balkema, Rotterdam, pp. 339-348.
Perzyna, P. (1992). Instability phenomena in thermoplastic flow processes: Foundations and computational
aspects, Proc. 3
rd
Int. Conf. Computational Plasticity, Pineridge Press, Swansea, pp. 27-57.
Pietruszczak, S. and Pande, G.N. (1996). Constitutive behavior of partially saturated soils containing gas
inclusions, J. of Geotech. Eng., ASCE, Vol. 122, pp. 50-59.
Pietruszczak, S.T. and Mrz, Z. (1981). Finite element analysis of deformation of strain-softening
materials, Int. J. Num. Anal. Meth. Geomech. Vol. 5, pp. 327-324.
Plant, G.W., Covil, C.S. and Hughes, R.A. (1998). Site preparation for the new Hong Kong International
Airport, Thomas Telford Publishing Limited. 576p.
Potts, D.M. and Ganendra, D. (1994). An evaluation of substepping and implicit stress point algorithms,
Comput. Meth. Appl. Mech. Eng., Vol. 119, pp. 341-354.
Potts, D.M. and Gens, A. (1984). The effect of the plastic potential in boundary value problems involving
plane strain deformation, Int. Jnl. Num. Anal. Meth. Geomech., Vol. 8, pp 259-286.
Potts, D.M. and Zdravkovic, L. (1999). Finite element analysis in geotechnical engineering: theory, Thomas
Telford, London.
Potts, D.M. and Zdravkovic, L. (2000). Finite element analysis in geotechnical engineering: applications,
Thomas Telford, London.
Powrie, W., Pantelidou, H. and Stallebrass, S.E. (1998). Soil stiffness in stress paths relevant to diaphragm
walls in clay, Gotechnique, Vol. 48, No. 4, pp. 483-494.
Prandtl, L. (1920). ber die Hrte plastischer Krper, Nachr. Ges. Wiss. Gtt., Math.Phys. Kl., Vol. 12,
pp. 7485.
Prevost, J.H. (1978). Plasticity theory of soil stress-strain behavior, J. Eng. Mech., ASCE, Vol. 104, pp.
1177-1194.
Puzrin, A.M. and Burland, J.B. (1998). Non-linear model of small-strain behaviour of soils, Gotechnique,
Vol. 48, No.2, pp. 217-233.
Rackwitz, R. and Peintinger, B. (1981). Ein wirklichkeitsnahes stochastisches Bodenmodell mit unsicheren
Parametern und Anwendung auf die Stabilittsuntersuchung von Bschungen, Bauingenieur, Vol. 56,
pp. 215 221.
Rahman, M.S., Wang, J., Deng, W. and Carter, J.P. (2000). A neural network model for the uplift capacity
of suction caissons, Computers and Geotechnics, in press.
Richards, B. (1992). Modelling interactive load-deformation and flow processes in soils, including
unsaturated and swelling soils, Proc. 6
th
Aust.-NZ Conference on Geomechanics, pp. 18-37.
Roscoe, K.H., Schofield, A. and Wroth, C.P. (1958). On the yielding of soils, Gotechnique, Vol. 8, pp.
22-53.
Rumelhart, D.E. and McClelland, J.L. (1986), Parallel distributed processing-explorations in the micro-
structure of cognition, Vol. 1 and 2, Cambridge: Massachusetts Institute of Technology Press.
Rumelhart, D.E., Hinton, G.E. and Williams R.J. (1986), Learning Internal Representation by Error
Propagation in Parallel Distributed Processing, Vol. 1. Cambridge: Massachusetts Institute of
Technology Press.
Santamarina, J.C. and Chameau, J.L. (1987). Expert systems for geotechnical engineers, Journal of
Computing in Civil Engineering, Vol. 1, No. 4.
Schweiger, H.F. (1991). Benchmark Problem No. 1: Results, Computers and Geotechnics, Vol. 11, pp.
331-341.
Schweiger, H.F. (1997). Berechnungsbeispiele des AK 1.6 der DGGT - Vergleich der Ergebnisse fr
Beispiel 1 (Tunnel) und 2 (Baugrube). Tagungsband Workshop Numerik in der Geotechnik,
DGGT/AK 1.6, pp. 1-29, (in German).
Schweiger, H.F. (1998). Results from two geotechnical benchmark problems. Proc. 4
th
European Conf.
Numerical Methods in Geotechnical Engineering. Cividini, A. (ed.), pp. 645-654.
Schweiger, H.F. (2000). Ergebnisse des Berechnungsbeispieles Nr. 3 3-fach verankerte Baugrube.
Tagungsband Workshop Verformungsprognose fr tiefe Baugruben, DGGT/AK 1.6, pp. 7-67, (in
German).
Schweiger, H.F. and Beer, G. (1996). Numerical simulation in tunneling an overview of different
computational models with emphasis on practical ability, Felsbau, Vol. 14, pp. 87-92.
Schweiger, H.F. Kofler, M. and Schuller, H. (1999). Some recent developments in the finite element
analysis of shallow tunnels, Felsbau, Vol. 17, pp. 426-431.
Schweiger, H.F., Schuller, H. and Pttler, R. (1997). Some remarks on 2-D-models for numerical
simulation of underground constructions with complex cross sections, Proc. 9
th
Int. Conf. Computer
Methods and Advances in Geomechanics, Wuhan, pp. 1303-1308.
Seed, H.B. (1979). Soil liquefaction and cyclic mobility evaluation of level ground during earthquakes, J.
of Geotech. Eng., ASCE, Vol. 105, pp. 201-255.
Shao, C. and Desai, C.S. (2000). Implementation of DSC model and application for analysis of field pile
tests under cyclic loading, Int. J. Num. Analyt. Meth. Geomech., Vol. 24, pp. 601-624.
Shibuya, S. Mitachi, T. and Miura, S. (1995). (Editors). Pre-Failure deformation of geomaterials,
Proceedings of the international symposium on pre-failure deformation characteristics of geomaterials,
Sapporo, Japan, 12-14 September 1994, Balkema, Rotterdam.
Shukula, S. (1988). EDESYN: A KBES shell for preliminary engineering design, Masters Thesis, Dept. of
Civil Engineering, Carnegie-Mellon University, USA.
Siller, T.J. (1987). Expert systems in geotechnical engineering, Expert systems for civil engineers:
Technology and application, Maher, M.L. (ed.), American Society of Civil Engineers, New York, NY,
USA.
Singh, J., Airey, D.W., Booker, J.R. and Carter, J.P. (1996). Model studies of the bearing capacity of an
orthogonally jointed medium, Proc. Int. Conf. On Distinct Element Methods. Davis, USA.
Sloan, S.W. (1981). Numerical Analysis of Incompressible and Plastic Solids Using Finite Elements, PhD
Thesis, University of Cambridge.
Sloan, S.W. (1987). Substepping schemes for numerical integration of elasto-plastic stress-strain relations,
Int. Jnl. Num. Meth. Eng., Vol. 24, pp. 893-911.
Sloan, S.W. (1988a). Lower bound limit analysis using finite elements and linear programming,
International Journal for Numerical and Analytical Methods in Geomechanics, Vol. 12, pp. 6167.
Sloan, S.W. (1988b). A steepest edge active set algorithm for solving large sparse linear programming
problems, International Journal for Numerical Methods in Engineering, Vol. 26, pp. 26712685.
Sloan, S.W. (1989). Upper bound limit analysis using finite elements and linear programming,
International Journal for Numerical and Analytical Methods in Geomechanics, Vol. 13, pp. 263282.
Sloan, S.W. and Assadi, A. (1991). Undrained stability of a square tunnel in a soil whose strength increases
linearly with depth, Computers and Geotechnics, Vol. 12, pp. 321346.
Sloan, S.W. and Assadi, A. (1992). The stability of tunnels in soft ground, Proceedings of Peter Wroth
Memorial Symposium on Predictive Soil Mechanics, Oxford, pp. 644663.
Sloan, S.W. and Kleeman, P.W. (1995). Upper bound limit analysis with discontinuous velocity fields,
Computer Methods in Applied Mechanics and Engineering, Vol. 127, pp. 293314.
Sloan, S.W., Assadi, A. and Purushothaman, N. (1990). Undrained stability of a trapdoor, Gotechnique,
Vol. 40, pp. 4562.
Sluys, L.J. and de Borst, R. (1992). Wave propagation and localization in a rate-dependent cracked medium
- model formulation and one-dimensional examples, Int. J. Solids Structures Vol. 29, pp. 2945-2958.
Smith, I.M. and Griffiths, D.V. (1988) Programming the finite element method (2
nd
edition). John Wiley &
Sons Ltd. Chichester, UK.
Smith, I.M., Hicks, M.A., Kay, S. and Cuckson, J. (1988). Undrained and partially drained behaviour of
end bearing piles and bells founded in untreated calcarenite, The Mechanics of Calcareous Soils R.J.
Jewell and M. Khorshid (eds), Balkema. Rotterdam. Vol. 2, pp. 663-680.
Smith, P.R., Jardine, R.J. and Hight, D.W. (1992). On the yielding of Bothkennar clay, Gotechnique,
Vol. 42, pp. 257-274.
Sokolovskii, V. V. (1965). Statics of Granular Media, Pergammon, New York.
Somasundaram, S. and Desai, C.S. (1988). Modelling and testing for anisotropic behavior of soils, J. Eng.
Mech., ASCE, Vol. 114, pp. 1177-1194.
Stolle, D.F. and Higgins, J.E. (1989). Viscoplasticity and plasticity - numerical stability revisited, Int.
Conf. Num. Models Geomech., NUMOG III, pp. 431-438.
Tejchman, J. and Bauer, E. (1996). Numerical simulation of shear band formation with a polar hypoplastic
constitutive model, Computers and Geotechnics, Vol. 19, pp. 221-244.
Terzaghi, K.S. (1943). Theoretical Soil Mechanics, John Wiley, New York, NY, USA.
Thomas, J.N. (1984), An improved accelerated initial stress procedure for elasto-plastic finite element
analysis, Int. J. Num. Meth. Geomech., Vol. 8, pp. 359-379.
Thurner, R. and Schweiger, H.F. (2000). Reliability analysis for geotechnical problems via finite elements -
a practical application, Proc. GEOENG 2000, Melbourne. Accepted for publication.
Turgeman, S. and Pastor, J. (1982). Limit analysis: a linear formulation of the kinematic approach for
axisymmetric mechanic problems, Int. J. for Num. Anal. Methods in Geomech., Vol. 6, pp. 109128.
Ukrichton, B., Whittle, A.J. and Sloan, S.W. (1998). Undrained limit analyses for combined loading of strip
footings on clay, J. Geotech. Geoenvironmental Engineering, ASCE, Vol. 124, No. 3, pp. 265276.
van Baars, S. (1996). Discrete element analysis of granular materials, Proefschrift Technische Univeriteit
Delft.
Vanmarcke, E.H. (1977). Probabilistic Modelling of Soil Profiles, Journal of the Geotechnical
Engineering Division, Vol. 103, pp. 1227-1246.
Vardoulakis, I. and Aifantis, E.C. (1989). Gradient dependent dilatancy and its implications in shear
banding and liquefaction, Ingenieur Archiv, Vol. 59, pp. 197-208.
Vardoulakis, I. and Aifantis, E.C. (1991). A gradient flow theory of plasticity for granular materials, Acta
Mechanica, Vol. 87, pp. 197-217.
Vermeer, P.A. and Brinkgreve, R.B.J. (1994). A new effective non-local strain-measure for softening
plasticity, Proc. 3
rd
Int. Woskshop Localisation and Bifurcation Theory for Soils and Rocks, Balkema,
Rotterdam, pp. 89-100.
Vesic, A.S. (1973). Analysis of ultimate loads of shallow foundations, J. Soil Mech. Foundation Division,
ASCE, Vol. 99, SM1, pp. 4573.
Vesic, A.S. (1975). Bearing capacity of shallow foundations, Foundation Engineering Handbook,
Winterkorn, H.F. and Fang, Y.H. (eds). Van Nostrand Reinhold, pp. 121144.
von Wolffersdorff, P.A. (1996). A hypoplastic relation for granular materials with a predefined limit state
surface, Mechanics of Cohesive-Frictional Materials, Vol. 1, pp. 251-271.
Wathugala, G.W. and Desai, C.S. (1993). Constitutive model for cyclic behavior of cohesive soils, I:
Theory, J. Geotech. Eng., ASCE, Vol. 119, pp. 714-729.
Wharry, M.B. and Ashley, D.B. (1986). Resolving subsurface risk in construction using an expert system,
Technical Report UTCEPM-86-1, University of Texas at Austin, USA.
Whittle, A.J., DeGroot, D.J., Ladd, C.C. and Seah, T.h. (1994). Model prediction of the anisotropic
behaviour of Boston Blue Clay, J. Geotechnical Engineering, ASCE, Vol. 120, No. 1, pp. 199-224.
Wickremesinghe, D. and Campanella, R.G. (1993). Scale of fluctuation as a descriptor of soil variability,
Proceedings Conference on Probabilistic Methods in Geotechnical Engineering, Canberra, February,
K.S. Li and S.C.R. Lo (eds), pp. 233-240.
Wissman, J.W. and Hauck, C. (1983). Efficient elasto-plastic finite element analysis with higher order
stress point algorithms, Comput. Struct., Vol. 17, pp. 89-95.
Wong, K.C. (1990). Expert systems for foundation design, Ph.D Thesis, University of Sydney, Australia.
Wong, K.C., Poulos, H.G. and Thorne, C.P. (1989). Site classification by expert systems, Computers and
Geotechnics, Vol. 8, pp. 133-156.
Wroth, C.P. and Houlsby, G.T. (1985). Soil mechanics Property characterization and analysis
procedures, Proc. 11
th
Int. Conf. Soil Mechanics and Foundation Engineering, San Francisco, USA.
Wu, W., Bauer, E. and Kolymbas, D. (1996). Hypoplastic constitutive model with critical state for granular
materials, Mech. Mater., Vol. 23, pp. 45-69.
Yeoh, C.K. (1996). Cyclic response of cemented soils and foundations, PhD Thesis. The University of
Sydney.
Yu, H.S., Salgado, R., Sloan, S.W. and Kim, J.M. (1998). Limit analysis versus limit equilibrium for slope
stability, J. of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 124, No. 1, pp. 111.
Yu, H.S., Sloan, S.W. and Kleeman, P.W. (1994). A quadratic element for upper bound limit analysis,
Engineering Computations, Vol. 11, pp. 195212.
Zhou, J. and Nowak, A.S. (1988). Integration formulas to evaluate functions of random variables,
Structural safety, Vol. 5, pp. 267-284.
Zienkiewicz, O.C. (1967). The finite element method, McGraw-Hill Publishing Company, New York (1
st
edition 1967, 3
rd
edition 1977).
Zienkiewicz, O.C. and Cormeau, I.C. (1974). Visco-plasticity, plasticity and creep in elastic solids - a
unified numerical solution approach, Int. Jnl. Num. Meth. Eng., Vol. 8, pp. 821-845.
Zienkiewicz, O.C. and Zhu, J.Z. (1987). A simple error estimator and adaptive procedure for practical
engineering analysis, Int. J. Num. Meth. Engng. Vol. 24, pp. 337-357.
Zienkiewicz, O.C., M. Huang and Pastor, M. (1995). Localization problems in plasticity using finite
elements with adaptive remeshing, Int. J. Num. Anal. Meth. Geomech. Vol. 19, pp. 127-148.
Zouain, N., Herskovits, J., Borges, L.A. and Feijo, R.A. (1993). An iterative algorithm for limit analysis
with nonlinear yield functions, Int. J. Solids and Structures, Vol. 30, No. 10, pp. 13971417.

Das könnte Ihnen auch gefallen