Vous êtes sur la page 1sur 412

Stress and Deformation

Modeling with SIGMA/W


An Engineering Methodology

John Krahn

First Edition, May 2004

Copyright 2004 by GEO-SLOPE International, Ltd.


All rights reserved. No part of this work may be reproduced or transmitted in any
form or by any means, electronic or mechanical, including photocopying,
recording, or by any information storage or retrieval system, without the prior
written permission of GEO-SLOPE International, Ltd.
Printed in Canada.
Acknowledgements
To say that this book is by John Krahn overstates the case. Without the
significant contributions and assistance by other people, this book would not exist.
At the top of the list are all the co-workers at GEO-SLOPE, especially Greg
Newman, Lori Newman and Leonard Lam.
In addition, significant contributions were made by Prof. S. Lee Barbour,
University of Saskatchewan, Saskatoon, Canada, especially with the chapter on the
What, How and Why of Numerical Modeling.
All of us who participated in creating the content are grateful to Helen Hekman
and Patricia Stooke for their valuable assistance with editing and formatting this
book.

GEO-SLOPE International Ltd


1400, 633 6th Ave SW
Calgary, Alberta, Canada T2P 2Y5
E-mail: info@geo-slope.com
Web: http://www.geo-slope.com

SIGMA/W

Table of Contents

Table of Contents
1

Introduction ............................................................. 1
1.1

Applications.................................................................................. 1
Deformation analysis............................................................. 1
Embankment/excavation construction .................................. 3
Excess pore-water pressures................................................ 4
Soil-structure interactions...................................................... 4
Consolidation analyses ......................................................... 6

Numerical Modeling: What, Why and How ............. 9


2.1

Introduction .................................................................................. 9

2.2

What is a numerical model? ...................................................... 10

2.3

Modeling in geotechnical engineering ....................................... 13

2.4

Why model? ............................................................................... 15


Quantitative predictions....................................................... 16
Compare alternatives .......................................................... 18
Identify governing parameters............................................. 20
Discover and understand physical process - train our thinking
...................................................................................... 21

2.5

How to model ............................................................................. 25


Make a guess ...................................................................... 25
Simplify geometry................................................................ 28
Start simple ......................................................................... 29
Do numerical experiments .................................................. 30
Model only essential components ....................................... 31
Start with estimated material properties.............................. 34
Interrogate the results ......................................................... 35
Evaluate results in the context of expected results............. 35

Page i

Table of Contents

SIGMA/W

Remember the real world.................................................... 35

2.6

How not to model....................................................................... 36

2.7

Closing remarks ......................................................................... 37

SIGMA/W: Fundamentals and Practical Modeling


Considerations ...................................................... 39
3.1

Introduction ................................................................................ 39

3.2

Density and unit weight.............................................................. 40

3.3

Compaction and density ............................................................ 42


Compaction and strength .................................................... 43

3.4

Plasticity..................................................................................... 44

3.5

Volume change .......................................................................... 46


Effective and total stress ..................................................... 46
Compressibility and consolidation....................................... 48
Zero change in pore-water pressure ................................... 52
Ko horizontal and vertical stresses ................................... 52

3.6

Strength parameters .................................................................. 54


Mohr stress circle ................................................................ 54
Mohr Coulomb failure criteria .............................................. 56
Triaxial testing ..................................................................... 57
Soil stiffness Youngs modulus ........................................ 59

3.7

Total stress, effective stress and pore-water pressures............ 60

3.8

Incremental formulation ............................................................. 63


Body forces ......................................................................... 64

3.9

Serviceability versus stability ..................................................... 65

3.10

Modeling progression ................................................................ 66

3.11

Units........................................................................................... 67

Meshing ................................................................ 69
4.1

Introduction ................................................................................ 69

Page ii

SIGMA/W

4.2

Table of Contents

Element fundamentals ............................................................... 70


Element nodes .................................................................... 70
Field variable distribution .................................................... 71
Element and mesh compatibility ......................................... 72
Numerical integration .......................................................... 74
Secondary variables............................................................ 76
Element shapes................................................................... 76

4.3

Regions...................................................................................... 77
Region types ....................................................................... 78
Region points ...................................................................... 78
Region properties ................................................................ 80

4.4

Mesh types and patterns ........................................................... 80


Structured mesh .................................................................. 80
Unstructured mesh .............................................................. 80
Triangular regions ............................................................... 81
Transfinite meshing ............................................................. 82
Mesh with openings............................................................. 85
Surface regions ................................................................... 86
Joining regions .................................................................... 90

4.5

Structured versus unstructured meshing................................... 92

4.6

Meshing for transient analyses .................................................. 95

4.7

Interface elements ..................................................................... 96

4.8

Infinite elements......................................................................... 97
Infinite elements in SIGMA/W ............................................. 98

4.9

General guidelines for meshing ................................................. 99


Higher order elements in SIGMA/W.................................. 100
Number of elements .......................................................... 101
Effect of drawing scale ...................................................... 101
Mesh purpose.................................................................... 102

Page iii

Table of Contents

SIGMA/W

Simplified geometry........................................................... 104

Material Properties.............................................. 107


5.1

Constitutive models overview .................................................. 108

5.2

Linear-elastic model................................................................. 110

5.3

Anisotropic elastic model ......................................................... 112

5.4

Nonlinear elastic (Hyperbolic) model....................................... 114


Initial modulus ................................................................... 116
Tangent modulus............................................................... 117
Unloading-reloading modulus ........................................... 118
Poisson's ratio ................................................................... 118
Negative pore-water pressure........................................... 119
Yield zones........................................................................ 120
Nonlinear-elastic (Hyperbolic) soil parameters ................. 121

5.5

Elastic-plastic model ................................................................ 122


Plastic matrix ..................................................................... 123
Yield criterion..................................................................... 125

5.6

Strain-softening model............................................................. 128


Yield criterion..................................................................... 129
Plastic matrix ..................................................................... 131
Evaluation of the plastic matrix ......................................... 132
Von Mises criterion............................................................ 133

5.7

Cam-clay model ....................................................................... 134


Soil parameters ................................................................. 135
Yield function..................................................................... 136
Plastic matrix ..................................................................... 139
Evaluation of the plastic matrix ......................................... 142
Initial condition................................................................... 144
Soil parameters ................................................................. 147

Page iv

SIGMA/W

Table of Contents

5.8

Modified Cam-clay model ........................................................ 148

5.9

Slip surfaces ............................................................................ 151

Boundary Conditions........................................... 155


6.1

Multiple boundary condition types ........................................... 155

6.2

Force or displacement ............................................................. 156

6.3

Body loads ............................................................................... 158

6.4

Nodal boundary conditions ...................................................... 159

6.5

Edge boundary conditions (pressures or stresses) ................. 161


Converting edge conditions to nodal equivalent forces .... 162
Fluid pressure conditions .................................................. 163
Defining a pressure boundary at internal element edges . 164

6.6

Transient boundary conditions................................................. 167

Analysis Types.................................................... 169


7.1

Problem view ........................................................................... 169

7.2

Initial in-situ stresses ............................................................... 169


In-situ 1 analysis................................................................ 169
In-situ 2 analysis................................................................ 170
Load deformation analysis for in-situ stresses.................. 171
Historic in-situ stress ......................................................... 172
In-situ stress example ....................................................... 173
Soil models and modulii for in-situ analyses ..................... 175

7.3

Load / deformation analysis..................................................... 176

7.4

Initial water table position ........................................................ 176

7.5

Specifying initial conditions using external file......................... 177

Structural Elements............................................. 179


8.1

Introduction .............................................................................. 179

8.2

Beam elements ........................................................................ 179

Page v

Table of Contents

SIGMA/W

Interpolating functions for a beam element....................... 179


Stiffness matrix for a beam element.................................. 180
8.3

Beam examples ....................................................................... 183

8.4

Bar elements............................................................................ 186


Interpolating functions for a bar element........................... 186
Stiffness matrix for a bar element ..................................... 187

8.5

10

Bar example............................................................................. 190

Fill and Excavation.............................................. 193


9.1

Simulating fill placement .......................................................... 193

9.2

Excavation elements................................................................ 196

9.3

Specifying fill / excavation elements........................................ 197

9.4

Braced excavations ................................................................. 198

Consolidation ...................................................... 201


10.1

Constitutive equation for soil structure .................................... 201

10.2

Flow equation for water phase................................................. 204

10.3

Finite element formulation for coupled analysis ...................... 205

10.4

Additional material properties for unsaturated coupled analysis210

10.5

Coupled analysis practical issues......................................... 211

10.6

SIGMA/W material models in an unsaturated coupled


consolidation analysis.............................................................. 213
H-Modulus function in a consolidation analysis ................ 213

10.7

Saturated zone only coupled analysis ..................................... 215

10.8

Time step size in a coupled consolidation analysis ................. 216

10.9

Uncoupled consolidation analysis............................................ 217

10.10

Consolidation analyses limitations........................................... 218

10.11

Examples of coupled and uncoupled consolidation................. 219


One dimensional consolidation examples......................... 219
Mendel-Cryer effect........................................................... 224

Page vi

SIGMA/W

Table of Contents

Two dimensional uncoupled consolidation example......... 226

11

Numerical Issues ................................................ 229


11.1

Nonlinear analysis ................................................................... 229

11.2

Convergence............................................................................ 233
Graphing convergence...................................................... 236

12

11.3

Limiting loads use incremental loading................................. 238

11.4

Tension zones.......................................................................... 239

11.5

Stop-restart .............................................................................. 240

11.6

Halt iteration............................................................................. 240

11.7

Solver generated data ............................................................. 241

Visualization of Results....................................... 243


12.1

Introduction .............................................................................. 243

12.2

Load steps ............................................................................... 243


Incremental difference....................................................... 243

12.3

Types of data available in Contour .......................................... 244

12.4

Viewing node and element Information ................................... 247

12.5

Projecting Gauss point values to nodes .................................. 249

12.6

Contouring data ....................................................................... 250


Contouring specific element regions ................................. 252

12.7

Graphing data .......................................................................... 253


Sum versus average graphs ............................................. 255

13

12.8

Viewing mesh displacement .................................................... 256

12.9

Viewing Mohr circles................................................................ 257

Illustrative Examples ........................................... 259


13.1

Strip footing stresses ............................................................... 259

13.2

Circular footing stresses .......................................................... 261

13.3

Nonlinear-elastic (Hyperbolic) analysis ................................... 262

Page vii

Table of Contents

SIGMA/W

In-situ stresses .................................................................. 265


Settlement analysis ........................................................... 265
13.4

Elastic-plastic material ............................................................. 269


Bearing capacity of a purely cohesive material................. 270
Bearing capacity of a purely frictional material ................. 273
In-situ analysis................................................................... 274

13.5

Strain-softening analysis.......................................................... 278

13.6

Cam-clay triaxial test ............................................................... 280


Material properties............................................................. 281
Initial conditions................................................................. 281
Hydraulic conditions .......................................................... 282
Initial yield surface............................................................. 283
Normally consolidated test ................................................ 283
Over-consolidated test ...................................................... 286

13.7

Modified Cam-clay ................................................................... 289


Modified Cam-clay results................................................. 289

13.8

Excavation in anisotropic material ........................................... 291

13.9

Embankment construction / consolidation ............................... 294

13.10

Braced excavations ................................................................. 297

13.11

Consolidation - Cryers problem .............................................. 301

13.12

Consolidation analysis of a saturated soil column................... 304

13.13

Coupled analysis of unsaturated soils ..................................... 310


Problem description........................................................... 311
Material properties............................................................. 312
Computed results .............................................................. 313
Commentary...................................................................... 315

13.14

Truss example ......................................................................... 316

13.15

Tie-back shoring wall ............................................................... 318


In-situ stress state ............................................................. 321

Page viii

SIGMA/W

Table of Contents

Simulation of construction procedure................................ 322


Results .............................................................................. 323
Lateral wall movement ...................................................... 324
Moment distributions in the sheet piling ............................ 324
Axial force in grouted length.............................................. 325
Lateral stress distributions ................................................ 326
13.16

Embankment on soft ground.................................................... 327


Problem description........................................................... 327
Clay properties .................................................................. 328
Sand properties ................................................................. 329
In-situ stresses .................................................................. 329
Loading sequence ............................................................. 331
Data specified in SEEP/W................................................. 332
Summary of the results ..................................................... 333

14

Product Integration.............................................. 339


14.1

SEEP/W generated pore-water pressures in SLOPE/W stability


analysis .................................................................................... 339

14.2

VADOSE/W generated pore pressures in SLOPE/W stability


analysis .................................................................................... 341

14.3

SEEP/W dissipation of pore pressures generated in a QUAKE/W


earth quake analysis................................................................ 346

14.4

SEEP/W velocity data in CTRAN/W contaminant transport


analysis .................................................................................... 349

14.5

VADOSE/W velocity data in CTRAN/W contaminant transport


analysis .................................................................................... 354

14.6

Density-dependent flow salt water intrusion ......................... 356

14.7

Ground freezing and water flows (SEEP/W and TEMP/W)..... 359

14.8

Seepage-dependent embankment settlement (SEEP/W and


SIGMA/W)................................................................................ 363

Page ix

Table of Contents

15

SIGMA/W

Theory................................................................. 369
15.1

Introduction .............................................................................. 369

15.2

Finite element equations.......................................................... 369


Strain-displacement matrix................................................ 371
Elastic constitutive relationship ......................................... 373
Body forces ....................................................................... 374
Forces due to boundary stresses...................................... 374
Nodal forces ...................................................................... 376

16

15.3

Numerical integration............................................................... 377

15.4

Assembly and solving of global equations............................... 380

15.5

Element stresses ..................................................................... 381

15.6

Pore pressure parameters A and B ......................................... 382

Appendix A: Interpolating Functions ................... 385


16.1

Coordinate systems ................................................................. 385

16.2

Derivatives of interpolating functions....................................... 387

References................................................................... 393
Index ............................................................................ 397

Page x

SIGMA/W

Chapter 1: Introduction

Introduction

SIGMA/W is a finite element software product that can be used to perform stress
and deformation analyses of earth structures. Its comprehensive formulation makes
it possible to analyze both simple and highly complex problems. For example, you
can perform a simple linear elastic deformation analysis or a highly sophisticated
nonlinear elastic-plastic effective stress analysis. When coupled with other GEOSLOPE software products, it can also model the pore-water pressure generation
and dissipation in a soil structure in response to external loads using either a fully
coupled or un-coupled formulation. SIGMA/W has application in the analysis and
design for geotechnical, civil, and mining engineering projects.

1.1

Applications

SIGMA/W can be used to compute stress-deformation with or without the changes


in pore-water pressures that arise from stress state changes. In addition, it is
possible to model soil structure interaction using beam or bar elements. The
following are some typical cases that can be analyzed using SIGMA/W.
Deformation analysis
The most common application of SIGMA/W is to compute deformations caused by
earthworks such as foundations, embankments, excavations and tunnels. Figure 1-1
shows a typical case of a fluid-filled tank on the ground surface. This figure
presents the deformation at an exaggerated scale as a deformed mesh. Figure 1-2
shows the associated change in vertical stress in the ground caused by the applied
load.

Page 1

Chapter 1: Introduction

SIGMA/W

30

Exaggeration 25X

28

Tank
El

26
24

ti
(

22

20

)
18
16
14
12
10
8
6
4
2
0

Figure 1-1 Deformation under a fluid-filled tank


30
28

Tank
Elevation (metres)

26
24
22

30

20
20

18
16
14

10

12
10
8

Figure 1-2 Vertical stress change under a fluid-filled tank

Page 2

SIGMA/W

Chapter 1: Introduction

Embankment/excavation construction
Finite elements can be added or removed from the finite element mesh to simulate
the construction of fill placement or excavation. You can identify elements to be
activated or deactivated at various stages, making it possible to simulate the
process over time. Figure 1-3 shows a deformed mesh after placement of one lift,
and Figure 1-4 shows the closure around a tunnel in anisotropic linear-elastic
material.

+2

+3
+3

+3

+3

+3

+3

+3
+3

+2

+2

+2

+2

+3
+2

+2

+2

+2

+2

+2

+2

+2

+2

+2

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+2
+1

+3
+2

+2
+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

Figure 1-3 Earth deformation after partial fill placement

Exaggerati on 150 00X

Figure 1-4 Closure around a tunnel

Page 3

Chapter 1: Introduction

SIGMA/W

Excess pore-water pressures


The effect of excess pore-water pressures generated during fill placement is often a
major consideration in slope stability during construction. SIGMA/W can be used
to estimate these types of pore-water pressures. Excess pore-water pressures
computed using SIGMA/W may be imported into SLOPE/W for slope stability
analysis. Figure 1-5 shows the excess pore-water pressures in the foundation
beneath an embankment immediately after construction. Note the use of infinite
elements for modelling the right and left boundaries.

40

60

80

100

120

160

140

100

120

60

80

40

140

Infinite Elements

160

Infinite Elements

200
260

Figure 1-5 Excess pore-water pressures caused by fill placement

Soil-structure interactions
SIGMA/W can accommodate soil-structure interaction problems by including
structural elements in two-dimensional plain strain analyses. These elements can
be beam elements which have flexural stiffness, or bar elements which have only
axial stiffness with no flexural stiffness. These structural elements are particularly
useful when analyzing cases such as sheet-pile walls. Figure 1-6 illustrates the
deformation for such a case, and Figure 1-7 shows the associated moment
distribution in the sheet-pile wall.

Page 4

SIGMA/W

Chapter 1: Introduction

Figure 1-6 Sheet pile wall example


Moment vs. Distance
25

Distance

20

15

10

0
-300

-200

-100

100

200

300

400

Moment

Figure 1-7 Moment distribution in a sheet pile wall

Page 5

Chapter 1: Introduction

SIGMA/W

Consolidation analyses
SIGMA/W can be used together with SEEP/W to perform a fully-coupled
consolidation analysis. When these two integrated products are run simultaneously,
SIGMA/W calculates the deformations resulting from pore-water pressure changes
while SEEP/W calculates transient pore-water pressure changes. This procedure is
used to simulate the consolidation process in both saturated and unsaturated soils.
A fully-coupled analysis is required to correctly model the pore-water pressure
response to an applied load. In certain cases, the pore-water pressure increase
under an applied load can be greater than the applied load. This phenomenon is
known as the Mendel-Cryer effect. Figure 1-8 shows a SIGMA/W analysis of a
saturated triaxial sample with an applied lateral load of 100 kPa. The initial porewater pressure before loading is zero. Figure 1-9 shows the pore-water pressure
response with time at the center of the sample. The pore-water pressure rises to
about 110 kPa, (110% of the applied load), before it gradually decreases.

Applied pressure 100 kPa


Initial pwp is zero

Figure 1-8 Tri-axial sample with applied lateral load

Page 6

SIGMA/W

Chapter 1: Introduction

Pore-Water Pressure vs. Time

Pore-Water Pressure

150

100

50

0
0

100

200

300

400

500

600

Time

Figure 1-9 Pore-water pressure change with time

Page 7

Chapter 1: Introduction

SIGMA/W

Page 8

SIGMA/W

Chapter 2: Numerical Modeling

Numerical Modeling: What, Why and How

2.1

Introduction

The unprecedented computing power now available has resulted in advanced


software products for engineering and scientific analysis. The ready availability
and ease-of-use of these products makes it possible to use powerful techniques
such as a finite element analysis in engineering practice. These analytical methods
have now moved from being research tools to application tools. This has opened a
whole new world of numerical modeling.
Software tools such as SIGMA/W do not inherently lead to good results. While the
software is an extremely powerful calculator, obtaining useful and meaningful
results from this useful tool depends on the guidance provided by the user. It is the
users understanding of the input and their ability to interpret the results that make
it such a powerful tool. In summary, the software does not do the modeling, the
user does the modeling. The software only provides the ability to do highly
complex computations that are not otherwise humanly possible. In a similar
manner, modern day spreadsheet software programs can be immensely powerful as
well, but obtaining useful results from a spreadsheet depends on the user. It is the
users ability to guide the analysis process that makes it a powerful tool. The
spreadsheet can do all the mathematics, but it is the users ability to take advantage
of the computing capability that leads to something meaningful and useful. The
same is true with finite element analysis software such as SIGMA/W.
Numerical modeling is a skill that is acquired with time and experience. Simply
acquiring a software product does not immediately make a person a proficient
modeler. Time and practice are required to understand the techniques involved and
learn how to interpret the results.
Numerical modeling as a field of practice is relatively new in geotechnical
engineering and, consequently, there is a lack of understanding about what
numerical modeling is, how modeling should be approached and what to expect
from it. A good understanding of these basic issues is fundamental to conducting
effective modeling. Basic questions such as, What is the main objective of the
analysis?, What is the main engineering question that needs to answered? and,
What is the anticipated result?, need to be decided before starting to use the
software. Using the software is only part of the modeling exercise. The associated
mental analysis is as important as clicking the, buttons in the software.

Page 9

Chapter 2: Numerical Modeling

SIGMA/W

This chapter discusses the what, why and how of the numerical modeling
process and presents guidelines on the procedures that should be followed in good
numerical modeling practice. It should be noted now that some of the examples in
this chapter pertain to stress-deformation analysis while others relate to seepage
analysis. While the actual type of analyses may differ, the concepts illustrated
apply to all engineering analyses.

2.2

What is a numerical model?

A numerical model is a mathematical simulation of a real physical process.


SIGMA/W is a numerical model that can mathematically simulate the real physical
process of ground volume change in response to self or external loading.
Numerical modeling is purely mathematical and in this sense is very different than
scaled physical modeling in the laboratory or full-scaled field modeling.
Rulon (1985) constructed a scale model of a soil slope with a less permeable layer
embedded within the slope and then sprinkled water on the crest to simulate
infiltration or precipitation. Instruments were inserted into the soil through the side
walls to measure the pore-water pressures at various points. The results of her
experiment are shown in Figure 2-1. Modeling Rulons laboratory experiment with
SEEP/W gives the results presented in Figure 2-2, which are almost identical to the
original laboratory measurements. The positions of the equipotential lines are
somewhat different, but the position of the water table is the same. In both cases
there are two seepage exit areas on the slope, which is the main important
observation in this case.

Page 10

SIGMA/W

Chapter 2: Numerical Modeling

Figure 2-1 Rulons laboratory scaled model results


1.0
0.9
0.8

Z (m)

0.7

Fine Sand

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

X (m)

Figure 2-2 SEEP/W analysis of Rulons laboratory model

The fact that mathematics can be used to simulate real physical processes is one of
the great wonders of the universe. Perhaps physical processes follow mathematical
rules, or mathematics has evolved to describe physical processes. Obviously, we
do not know which came first, nor does it really matter. Regardless of how the
relationship developed, the fact that we can use mathematics to simulate physical
processes leads to developing a deeper understanding of physical processes. It may
even allow for understanding or discovering previously unknown physical
processes.

Page 11

Chapter 2: Numerical Modeling

SIGMA/W

Numerical modeling has many advantages over physical modeling. The following
are some of the more obvious advantages:

Numerical models can be set up very quickly relative to physical


models. Physical models may take months to construct while a
numerical model can be constructed in minutes, hours or days.

A physical model is usually limited to a narrow set of conditions. A


numerical model can be used to investigate a wide variety of different
scenarios.

Numerical models have no difficulty accounting for gravity. Gravity


cannot be scaled, which is a limitation with laboratory modeling. A
centrifuge is often required to overcome this limitation.

With numerical modeling, there is no danger of physical harm to


personnel. Physical modeling sometimes involves heavy equipment
and worker safety is consequently a concern.

Numerical modeling provides information and results at any location


within the cross-section. Physical modeling only provides external
visual responses and data at discrete instrumented points.

Numerical models can accommodate a wide variety of boundary


conditions, whereas physical models are often limited in the types of
boundary conditions possible.

It would be wrong to think that numerical models do not have limitations.


Associated with seepage flow there may also be temperature changes, volume
changes and perhaps chemical changes. Including all these processes in the same
formulation is not possible, as the mathematics involved simply become too
complex. In addition, it is not possible to mathematically describe a constitutive
relationship, due to its complexity. Some of these difficulties can and will be
overcome with greater and faster computer processing power.
It is important to understand that numerical modeling products like SIGMA/W will
have limitations that are related to the current capability of hardware or integral to
the formulation of the software, since it was developed to consider specific
conditions. SIGMA/W is formulated only for small strain cases and not for post
failure deformation. In many cases, the constitutive equations used in this model
are not defined for post failure stress conditions, so it is pointless to begin an
analysis hoping to investigate those states. The important point to remember is that

Page 12

SIGMA/W

Chapter 2: Numerical Modeling

the mathematical formulations implemented in SIGMA/W result in a very


powerful and versatile means of simulating real physical processes.
A mathematical model is a replica of some real-world object or system. It is an attempt to
take our understanding of the process (conceptual model) and translate it into mathematical
terms. National Research Council Report (1990).

2.3

Modeling in geotechnical engineering

The role and significance of analysis and numerical modeling in geotechnical


engineering has been vividly illustrated by Professor John Burland, Imperial
College, London (UK). In 1987 Professor Burland presented what is known as the
Nash Lecture. The title of the lecture was The Teaching of Soil Mechanics a
Personal View. In this lecture he advocated that geotechnical engineering consists
of three fundamental components: the ground profile, the soil behavior and
modeling. He represented these components as the apexes of a triangle, as
illustrated in Figure 2-3. This has come to be known as the Burland triangle
(Burland, 1987; Burland, 1996).

Ground
profile

Empiricism,
Precedent
Soil
behaviour

Modeling

The soil mechanics triangle

Figure 2-3 The Burland triangle (after Burland 1996)

The soil behavior component includes laboratory tests, in situ tests and field
measurements. The ground profile component basically involves site
characterization: defining and describing the site conditions. Modeling may be
conceptual, analytical or physical.

Page 13

Chapter 2: Numerical Modeling

SIGMA/W

Of great significance is that, in Burlands view, all three components need to be


tied together by empiricism and precedent. This is the part inside the triangle.
The Burland triangle idea has been widely discussed and referred to by others
since it was first presented. An article on this topic was presented in an issue of
Ground Engineering (Anon. 1999). Morgenstern (2000) discussed this at some
length in his keynote address titled Common Ground at the GeoEng2000
Conference in Melbourne Australia in 2000. With all the discussion, the triangle
has been enhanced and broadened somewhat, as shown in Figure 2-4.
One important additional feature has been to consider all the connecting arrows
between the components as pointing in both directions. This simple addition
highlights the fact that each part is distinct yet related to all the other parts.
The Burland triangle vividly illustrates the importance of modeling in geotechnical
engineering. Characterizing the field conditions and making measurements of
behavior is not sufficient. Ultimately, it is necessary to do some analysis of the
field information and soil properties to complete the triangle.
As Burland pointed out, modeling may be conceptual, analytical or physical.
However, with the computing power and software tools now available, modeling
often refers to numerical modeling. Accepting that modeling primarily refers to
numerical modeling, the Burland triangle shows the importance that numerical
modeling has in geotechnical engineering.
Making measurements and characterizing site conditions is often time consuming
and expensive. This is also true with modeling, if done correctly. A common
assumption is that the numerical modeling component is only a small component
that should be undertaken at the end of a project, and that it can be done simply
and quickly. This is somewhat erroneous. Good numerical modeling, as we will
see later in the section in more detail, takes time and requires careful planning in
the same manner that it takes time and planning to collect field measurements and
adequately characterize site conditions.
Considering the importance of modeling that the Burland triangle suggests for
geotechnical engineering, it is prudent that we do the modeling carefully and with
a complete understanding of the modeling processes. This is particularly true with
numerical modeling. The purpose of this book is to assist with this aspect of
geotechnical engineering.

Page 14

SIGMA/W

Chapter 2: Numerical Modeling

Genesis / geology

Ground
Profile

Site investigation,
ground description

Empiricism,
precedent,
experience,
risk management

Soil
Behaviour

Modeling

Idealization followed by
evaluation. Conceptual
or physical modeling,
analytical modeling

Lab / field testing,


observation,
measurement

Figure 2-4 The enhanced Burland triangle (after Anon. 1999)

2.4

Why model?

The first reaction to the question, why model? seems rather obvious. The
objective is to analyze the problem. Upon more thought, the answer becomes more
complex. Without a clear understanding of the reason for modeling or identifying
what the modeling objectives are, numerical modeling can lead to a frustrating
experience and uncertain results. As we will see in more detail in the next section,
it is wrong to set up the model, calculate a solution and then try to decide what the
results mean. It is important to decide at the outset the reason for doing the
modeling. What is the main objective and what is the question that needs to be
answered?
The following points are some of the main reasons for modeling, from a broad,
high level perspective. We model to:

make quantitative predictions,

Page 15

Chapter 2: Numerical Modeling

SIGMA/W

compare alternatives,

identify governing parameters, and

understand processes and train our thinking.

Quantitative predictions
Most engineers, when asked why they want to do some modeling, will say that
they want to make a prediction. They want to predict the deformation under a
footing, for example; or the time for a contaminant to travel from the source to a
seepage discharge point; or the time required from first filling a reservoir until
steady-state seepage conditions have been established in the embankment dam.
The desire is to say something about future behavior or performance.
Making quantitative predictions is a legitimate reason for doing modeling.
Unfortunately, it is also the most difficult part of modeling, since quantitative
values are often directly related to the material properties. The extent of
deformation under a footing, for example, is in large part controlled by the
stiffness of the soil, which is itself controlled by the amount of confining stress.
While it is realistic to obtain accurate stress distributions beneath the footing, we
should realize that the accuracy of the deformation is only as good as our
knowledge and confidence in our stiffness parameters.
Carter et al. (2000) presented the results of a competition conducted by the German
Society for Geotechnics. Packages of information were distributed to consulting
engineers and university research groups. The participants were asked to predict
the lateral deflection of a tie-back shoring wall for a deep excavation in Berlin.
During construction, the actual deflection was measured with inclinometers. Later
the predictions were compared with the actual measurements. Figure 2-5 shows the
best eleven submitted predictions. Other predictions were submitted, but were
considered unreasonable and consequently not included in the summary.
There are two heavy dark lines superimposed on Figure 2-5. The dashed line on
the right represents the inclinometer measurements uncorrected for any possible
base movement. It is likely the base of the inclinometer moved together with the
base of the wall. Assuming the inclinometer base moved about 10 mm, the solid
heavy line in Figure 2-5 has been shifted to reflect the inclinometer base
movement.
At first glance one might quickly conclude that the agreement between prediction
and actual lateral movement is very poor, especially since there appears to be a

Page 16

SIGMA/W

Chapter 2: Numerical Modeling

wide scatter in the predictions. This exercise might be considered as an example of


our inability to make accurate quantitative predictions.
However, a closer look at the results reveals a picture that is not so bleak. The
depth of the excavation is 32 m. The maximum predicted lateral movement is just
over 50 mm or 5 cm. This is an extremely small amount of movement over the
length of the wall certainly not big enough to be visually noticeable.
Furthermore, the actual measurements, when corrected for base movement fall
more or less in the middle of the predictions. Most important to consider are the
trends presented by many of the predicted results. Many of them predict a
deflected shape similar to the actual measurements. In other words, the predictions
simulated the correct relative response of the wall.
Consequently, we can argue that our ability to make accurate predictions is poor,
but we can also argue that the predictions are amazingly good. The predictions fall
on either side of the measurements and the deflected shapes are correct. In the end,
the modeling provided a correct understanding of the wall behavior, which is more
than enough justification for doing the modeling, and may be the greatest benefit
of numerical modeling, as we will see in more detail later.
Numerical modeling is sometimes dismissed as being useless due to the difficulty
with defining material properties. There are, however, other reasons for doing
numerical modeling. If some of the other objectives of numerical modeling are
completed first, then quantitative predictions often have more value and meaning.
Once the physics and mechanisms are completely understood, quantitative
predictions can be made with a great deal more confidence and are not nearly as
useless as first thought, regardless of our inability to accurately define material
properties.

Page 17

Chapter 2: Numerical Modeling

SIGMA/W

Deflection (mm)
-60

-50

-40

-30

-20

-10

10
0

16

20

Depth below surface (m)

12

24
computed
28

measured

32
-60

-50

-40

-30

-20

-10

Figure 2-5 Comparison of predicted and measured lateral movements of a


shoring wall (after Carter et al, 2000)

Compare alternatives
Numerical modeling is useful for comparing alternatives. Keeping everything else
the same and changing a single parameter makes it a powerful tool to evaluate the
significance of individual parameters. For modeling alternatives and conducting
sensitivity studies it is not all that important to accurately define some material
properties. All that is of interest is the change between simulations.
Consider the example of a cut-off wall beneath a structure. With SEEP/W it is easy
to examine the benefits obtained by changing the length of the cut-off. Consider

Page 18

SIGMA/W

Chapter 2: Numerical Modeling

two cases with different cut-off depths to assess the difference in uplift pressures
underneath the structure. Figure 2-6 shows the analysis when the cutoff is 10 feet
deep. The pressure drop and uplift pressure along the base are shown in the left
graph in Figure 2-7. The drop across the cutoff is from 24 to 18 feet of pressure
head. The results for a 20-foot cutoff are shown in Figure 2-7 on the right side.
Now the drop across the cutoff is from 24 to about 15 feet of pressure head. The
uplift pressures at the downstream toe are about the same.
The actual computed values are not of significance in the context of this
discussion. It is an example of how a model such as SEEP/W can be used to
quickly compare alternatives. Secondly, this type of analysis can be done with a
rough estimate of the conductivity, since in this case the pressure distributions will
be unaffected by the conductivity assumed. There would be no value in carefully
defining the conductivity to compare the base pressure distributions.
We can also look at the change in flow quantities. The absolute flow quantity may
not be all that accurate, but the change resulting from various cut-off depths will be
of value. The total flux is 6.26 x 10-3 ft3/s for the 10-foot cutoff and 5.30 x 10-3 ft3/s
for the 20-foot cutoff, only about a 15 percent difference.

6.2592e-003

Figure 2-6 Seepage analysis with a cutoff

Page 19

Chapter 2: Numerical Modeling

SIGMA/W

Cutoff - 20 feet

25

25

20

20

Pressure Head - feet

Pressure Head - feet

Cutoff 10 feet

15
10
5
0
30

15
10
5

50

70

90

110

Distance - feet

0
30

50

70

90

110

Distance - feet

Figure 2-7 Uplift pressure distributions along base of structure

Identify governing parameters


Numerical models are useful for identifying critical parameters in a design.
Consider the performance of a soil cover over waste material. What is the most
important parameter governing the behavior of the cover? Is it the precipitation,
the wind speed, the net solar radiation, plant type, root depth or soil type? Running
a series of VADOSE/W simulations, keeping all variables constant except for one
makes it possible to identify the governing parameter. The results can be presented
as a tornado plot such as shown in Figure 2-8.
Once the key issues have been identified, further modeling to refine a design can
concentrate on the main issues. If, for example, the vegetative growth is the main
issue then efforts can be concentrated on what needs to be done to foster the plant
growth.

Page 20

SIGMA/W

Chapter 2: Numerical Modeling

Base Case
thinner

high

low

Thickness
of Growth Medium

bare surface
Transpiration

low

high
deep

Hydraulic Conductivity
of Compacted Layer

shallow
Root Depth

low

Decreasing
Net Percolation

high

Hydraulic Conductivity
of Growth Medium
Increasing
Net Percolation

Figure 2-8 Example of a tornado plot (OKane, 2004)

Discover and understand physical process - train our thinking


One of the most powerful aspects of numerical modeling is that it can help us to
understand physical processes in that it helps to train our thinking. A numerical
model can either confirm our thinking or help us to adjust our thinking if
necessary.
To illustrate this aspect of numerical modeling, consider the case of a multilayered
earth cover system such as the two possible cases shown in Figure 2-9. The
purpose of the cover is to reduce the infiltration into the underlying waste material.
The intention is to use the earth cover layers to channel any infiltration downslope
into a collection system. It is known that both a fine and a coarse soil are required
to achieve this. The question is, should the coarse soil lie on top of the fine soil or
should the fine soil overlay the coarse soil? Intuitively it would seem that the
coarse material should be on top; after all, it has the higher conductivity. Modeling
this situation with SEEP/W, which handles unsaturated flow, can answer this
question and verify if our thinking is correct.
For unsaturated flow, it is necessary to define a hydraulic conductivity function: a
function that describes how the hydraulic conductivity varies with changes in
suction (negative pore-water pressure = suction). Chapter 4, Material Properties, in

Page 21

Chapter 2: Numerical Modeling

SIGMA/W

the SEEP/W engineering book, describes in detail the nature of the hydraulic
conductivity (or permeability) functions. For this example, relative conductivity
functions such as those presented in Figure 2-10 are sufficient. At low suctions
(i.e., near saturation), the coarse material has a higher hydraulic conductivity than
the fine material, which is intuitive. At high suctions, the coarse material has the
lower conductivity, which often appears counterintuitive.

Fine

Coarse

Coarse

Fine

OR
Material to be
protected

Material to be
protected

Figure 2-9 Two possible earth cover configurations


1.00E-04

1.00E-05

Conductivity

Coarse
Fine
1.00E-06

1.00E-07

1.00E-08

1.00E-09

1.00E-10
1

10

100

1000

Suction

Figure 2-10 Hydraulic conductivity functions

After conducting various analyses and trial runs with varying rates of surface
infiltration, it becomes evident that the behavior of the cover system is dependent
on the infiltration rate. At low infiltration rates, the effect of placing the fine

Page 22

SIGMA/W

Chapter 2: Numerical Modeling

material over the coarse material results in infiltration being drained laterally
through the fine layer, as shown in Figure 2-11. This accomplishes the design
objective of the cover. If the precipitation rate becomes fairly intensive, then the
infiltration drops through the fine material and drains laterally within the lower
coarse material as shown in Figure 2-12. The design of fine soil over coarse soil
may work, but only in arid environments. The occasional cloud burst may result in
significant water infiltrating into the underlying coarse material, which may result
in increased seepage into the waste. This may be a tolerable situation for short
periods of time. If most of the time precipitation is modest, the infiltration will be
drained laterally through the upper fine layer into a collection system.
So, for an arid site the best solution is to place the fine soil on top of the coarse
soil. This is contrary to what one might expect at first. The first reaction may be
that something is wrong with the software, but it may be that our understanding of
the process and our general thinking is flawed.
A closer examination of the conductivity functions provides a logical explanation.
The software is correct and provides the correct response given the input
parameters. Consider the functions in Figure 2-13. When the infiltration rate is
large, the negative water pressures or suctions will be small. As a result, the
conductivity of the coarse material is higher than the finer material. If the
infiltration rates become small, the suctions will increase (water pressure becomes
more negative) and the unsaturated conductivity of the finer material becomes
higher than the coarse material. Consequently, under low infiltration rates it is
easier for the water to flow through the fine, upper layer soil than through the
lower more coarse soil.

Low to modest rainfall rates

Fine
Coarse

Figure 2-11 Flow diversion under low infiltration

Page 23

Chapter 2: Numerical Modeling

SIGMA/W

Intense rainfall rates

Fine
Coarse

Figure 2-12 Flow diversion under high infiltration

This type of analysis is a good example where the ability to utilize a numerical
model greatly assists our understanding of the physical process. The key is to think
in terms of unsaturated conductivity as opposed to saturated conductivities.
Numerical modeling can be crucial in leading us to the discovery and
understanding of real physical processes. In the end the model either has to
conform to our mental image and understanding or our understanding has to be
adjusted.
1.00E-04

1.00E-05

Conductivity

Coarse
Fine
1.00E-06

1.00E-07

Intense
Rainfall

1.00E-08

Low to Modest
Rainfall
1.00E-09

1.00E-10
1

10

100

1000

Suction

Figure 2-13 Conductivities under low and intense infiltration

Page 24

SIGMA/W

Chapter 2: Numerical Modeling

This is a critical lesson in modeling and the use of numerical models in particular.
The key advantage of modeling, and in particular the use of computer modeling
tools, is the capability it has to enhance engineering judgment, not the ability to
enhance our predictive capabilities. While it is true that sophisticated computer
tools greatly elevated our predictive capabilities relative to hand calculations,
graphical techniques, and closed-form analytical solutions, still, prediction is not
the most important advantage these modern tools provide. Numerical modeling is
primarily about process - not about prediction.
The attraction of ... modeling is that it combines the subtlety of human judgment with the
power of the digital computer. Anderson and Woessner (1992).

2.5

How to model

Numerical modeling involves more than just acquiring a software product.


Running and using the software is an essential ingredient, but it is a small part of
numerical modeling. This section talks about important concepts in numerical
modeling and highlights important components in good modeling practice.
Make a guess
Generally, careful planning is involved when undertaking a site characterization or
making measurements of observed behavior. The same careful planning is required
for modeling. It is inappropriate to acquire a software product, input some
parameters, obtain some results, and then decide what to do with the results or
struggle to decide what the results mean. This approach usually leads to an
unhappy experience and is often a meaningless exercise.
Good modeling practice starts with some planning. If at all possible, you should
form a mental picture of what you think the results will look like. Stated another
way, we should make a rough guess at the solution before starting to use the
software. Figure 2-14 shows a very quick hand calculation of stresses beneath a
submerged, horizontal surface.

Page 25

Chapter 2: Numerical Modeling

SIGMA/W

=0

=0

u=0

Z=2m, = 10
kN/m3
0

20

20

=-u
Z=10m, = 20 kN/
m3

220

120

100

Pressure / Stress units

Figure 2-14 Hand calculation of in situ stresses

Y-Effective Stress vs. Y


25

20

15

10
0

20

40

60

80

100

Y-Effective Stress

Figure 2-15 SIGMA/W computed effective stress profile

The hand calculation together with the SIGMA/W output can now be used to judge
the validity of the computed results. If there is no resemblance between what is
expected and what is computed with SIGMA/W then either the preliminary mental
picture of the situation was not right or something has been inappropriately
specified in the numerical model. Perhaps the pressure boundary condition was not
specified on the ground surface in addition to the definition of the water table. The
water table controls pore-water pressures in the soil, but does not add a weight load

Page 26

SIGMA/W

Chapter 2: Numerical Modeling

at the ground surface. Both must be specified and this error would be obvious by
comparing expectations to computed results. Any differences ultimately need to be
resolved in order for you to have any confidence in your modeling. If you had
never made a preliminary guess at the solution then it would be very difficult to
judge the validity of the numerical modeling results.
Another extremely important part of modeling is to clearly define, at the outset, the
primary question to be answered by the modeling process. Is the main question the
stress distribution or is it deformation? If your main objective is to determine the
stress distribution, there is no need to spend a lot of time on establishing an
advanced soil model, a more simple linear-elastic model is adequate. If on the
other hand your main objective is to estimate deformation, then a greater effort is
needed in determining the appropriate stress-strain model.
Sometimes modelers say I have no idea what the solution should look like - that is
why I am doing the modeling. The question then arises, why can you not form a
mental picture of what the solution should resemble? Maybe it is a lack of
understanding of the fundamental processes or physics, maybe it is a lack of
experience, or maybe the system is too complex. A lack of understanding of the
fundamentals can possibly be overcome by discussing the problem with more
experienced engineers or scientists, or by conducting a study of published
literature. If the system is too complex to make a preliminary estimate then it is
good practice to simplify the problem so you can make a guess and then add
complexity in stages so that at each modeling interval you can understand the
significance of the increased complexity. If you were dealing with a very
heterogenic system, you could start by defining a homogenous cross-section,
obtaining a reasonable solution and then adding heterogeneity in stages. This
approach is discussed in further detail in a subsequent section.
If you cannot form a mental picture of what the solution should look like prior to
using the software, then you may need to discover or learn about a new physical
process as discussed in the previous section.
Effective numerical modeling starts with making a guess of what the solution should look
like.

Other prominent engineers support this concept. Carter (2000) in his keynote
address at the GeoEng2000 Conference in Melbourne, Australia, when talking
about rules for modeling, stated verbally that modeling should start with an

Page 27

Chapter 2: Numerical Modeling

SIGMA/W

estimate. Prof. John Burland made a presentation at the same conference on his
work with righting the Leaning Tower of Pisa. Part of the presentation was on the
modeling that was done to evaluate alternatives and while talking about modeling
he too stressed the need to start with a guess.
Simplify geometry
Numerical models need to be a simplified abstraction of the actual field conditions.
In the field the stratigraphy may be fairly complex and boundaries may be
irregular. In a numerical model the boundaries need to become straight lines and
the stratigraphy needs to be simplified so that it is possible to obtain an
understandable solution. Remember, it is a model, not the actual conditions.
Generally, a numerical model cannot and should not include all the details that
exist in the field. If attempts are made at including all the minute details, the model
can become so complex that it is difficult and sometimes even impossible to
interpret or even obtain results.
Figure 2-16 shows a stratigraphic cross section (National Research Council Report
1990). A suitable numerical model for simulating the flow regime between the
groundwater divides is something like the one shown in Figure 2-17. The
stratigraphic boundaries are considerably simplified for the finite element analysis.
As a general rule, a model should be designed to answer specific questions. You
need to constantly ask yourself while designing a model, if this feature will
significantly affects the results. If you have doubts, you should not include it in the
model, at least not in the early stages of analysis. Always start with the simplest
model.

Figure 2-16 Example of a stratigraphic cross section


(from National Research Report 1990)

Page 28

SIGMA/W

Chapter 2: Numerical Modeling

Figure 2-17 Finite element model of stratigraphic section

The tendency of novice modelers is to make the geometry too complex. The
thinking is that everything needs to be included to get the best answer possible. In
numerical modeling this is not always true. Increased complexity does not always
lead to a better and more accurate solution. Geometric details can, for example,
even create numerical difficulties that can mask the real solution.
Start simple
One of the most common mistakes in numerical modeling is to start with a model
that is too complex. When a model is too complex, it is very difficult to judge and
interpret the results. Often the result may look totally unreasonable. Then the next
question asked is - what is causing the problem? Is it the geometry, is it the
material properties, is it the boundary conditions, or is it the time step size or
something else? The only way to resolve the issue is to make the model simpler
and simpler until the difficulty can be isolated. This happens on almost all projects.
It is much more efficient to start simple and build complexity into the model in
stages, than to start complex, then take the model apart and have to rebuild it back
up again.
A good start may be to take a homogeneous section and then add geometric
complexity in stages. For the homogeneous section it is likely easier to judge the
validity of the results. This allows you to gain confidence in the boundary
conditions and material properties specified. Once you have reached a point where
the results make sense, you can add different materials and increase the complexity
of your geometry.

Page 29

Chapter 2: Numerical Modeling

SIGMA/W

Another approach may be to start with a steady-state analysis even though you are
ultimately interested in a transient process. A steady-state analysis gives you an
idea as to where the transient analysis should end up: to define the end point. Using
this approach you can then answer the question of how does the process migrate
with time until a steady-state system has been achieved.
It is unrealistic to dump all your information into a numerical model at the start of
an analysis project and magically obtain beautiful, logical and reasonable
solutions. It is vitally important to not start with this expectation. You will likely
have a very unhappy modeling experience if you follow this approach.
Do numerical experiments
Interpreting the results of numerical models sometimes requires doing numerical
experiments. This is particularly true if you are uncertain as to whether the results
are reasonable. This approach also helps with understanding and learning how a
particular feature operates. The idea is to set up a simple problem for which you
can create a hand calculated solution.
Consider the following example taken from a seepage analysis. You are uncertain
about the results from a flux section or the meaning of a computed boundary flux.
To help satisfy this lack of understanding, you could do a numerical experiment on
a simple 1D case as shown in Figure 2-18. The total head difference is 1 m and the
conductivity is 1 m/day. The gradient under steady state conditions is the head
difference divided by the length, making the gradient 0.1. The resulting total flow
through the system is the cross sectional area times the gradient which should be
0.3 m3/day. The flux section that goes through the entire section confirms this
result. There are flux sections through Elements 16 and 18. The flow through each
element is 0.1 m3/day, which is correct since each element represents one-third of
the area.
Another way to check the computed results is to look at the node information.
When a head is specified, SEEP/W computes the corresponding nodal flux. In
SEEP/W these are referred to as boundary flux values. The computed boundary
nodal flux for the same experiment shown in Figure 2-18 on the left at the top and
bottom nodes is 0.05. For the two intermediate nodes, the nodal boundary flux is
0.1 per node. The total is 0.3, the same as computed by the flux section. Also, the
quantities are positive, indicating flow into the system. The nodal boundary values
on the right are the same as on the left, but negative. The negative sign means flow
out of the system.

Page 30

Chapter 2: Numerical Modeling

12

15

11

14

10

13

18

21

24

27

30

17

20

23

26

29

16

19

22

25

28

1.0000e-001

3.0000e-001

1.0000e-001

SIGMA/W

Figure 2-18 Horizontal flow through three element section

A simple numerical experiment takes only minutes to set up and run, but can be
invaluable in confirming to you how the software works and in helping you
interpret the results. There are many benefits: the most obvious is that it
demonstrates the software is functioning properly. You can also see the difference
between a flux section that goes through the entire problem versus a flux section
that goes through a single element. You can see how the boundary nodal fluxes are
related to the flux sections. It verifies for you the meaning of the sign on the
boundary nodal fluxes. Fully understanding and comprehending the results of a
simple example like this greatly helps increase your confidence in the
interpretation of results from more complex problems.
Conducting simple numerical experiments is a useful exercise for both novice and
experienced modelers. For novice modelers it is an effective way to understand
fundamental principles, learn how the software functions, and gain confidence in
interpreting results. For the experienced modeler it is an effective means of
refreshing and confirming ideas. It is sometimes faster and more effective than
trying to find appropriate documentation and then having to rely on the
documentation. At the very least it may enhance and clarify the intent of the
documentation.
Model only essential components
One of the powerful and attractive features of numerical modeling is the ability to
simplify the geometry and not to have to include the entire physical structure in the
model. A very common problem is the seepage flow under a concrete structure
with a cut-off as shown in Figure 2-19. To analyze the seepage through the

Page 31

Chapter 2: Numerical Modeling

SIGMA/W

foundation it is not necessary to include the dam itself or the cut-off as these
features are constructed of concrete and assumed impermeable.

Figure 2-19 Simple flow beneath a cutoff

Another common example is the downstream toe drain or horizontal under drain in
an embankment (Figure 2-20). The drain is so permeable relative to the
embankment material that the drain does not contribute to the dissipation of the
head (potential energy) loss through the structure. Physically, the drain needs to
exist in the embankment, but it does not need to be present in a numerical model. If
the drain becomes clogged with fines so that it begins to impede the seepage flow,
then the situation is different and the drain would need to be included in the
numerical model. With any material, the need to include it in the analysis should
be decided in the context of whether it contributes to the head loss.
Another example is the downstream shell of a zoned dam as illustrated in Figure
2-21. Often the core is constructed of fine-grained soil while the shells are highly
permeable coarse granular material. If there is a significant difference between
core and shell conductivities then seepage that flows through the core will drip
along the downstream side of the core (usually in granular transition zones) down
to an under drain. If this is the case, the downstream shell does not need to be
included in the seepage analysis, since the shell is not physically involved in the
dissipation of the head loss. Once again the shell needs to exist physically, but
does not need to be included in the numerical seepage model.

Page 32

SIGMA/W

Chapter 2: Numerical Modeling

10

15

20

25

30

35

40

Figure 2-20 Flow through a dam with coarse toe drain

Figure 2-21 Head loss through dam core with downstream shell

Including unnecessary features and trying to model adjacent materials with


extreme contrasts in material properties create numerical difficulties. The
conductivity difference between the core and shell of a dam may be many, many
orders of magnitude. The situation may be further complicated if unsaturated flow
is present and the conductivity function is very steep, making the solution highly
non-linear. In this type of situation it can be extremely difficult if not impossible to
obtain a good solution with the current technology.
The numerical difficulties can be eased by eliminating non-essential segments
from the numerical model. If the primary interest is the seepage through the core,
then why include the downstream shell and complicate the analysis? Omitting nonessential features from the analysis is a very useful technique, particularly during
the early stages of an analysis. During the early stages, you are simply trying to
gain an understanding of the flow regime and trying to decide what is important
and what is not important.
While deliberately leaving components out of the analysis may at first seem like a
rather strange concept, it is a very important concept to accept if you want to be an
effective numerical modeler.

Page 33

Chapter 2: Numerical Modeling

SIGMA/W

Start with estimated material properties


In the early stages of a numerical modeling project it is often good practice to start
with estimates of material properties. In SIGMA/W this means always start with a
linear-elastic soil model. It will converge quickly because of its linearity, it will let
you refine the geometry if needed, and it will let you establish the load step
sequence you need. Simple estimates of material properties and simple property
functions are more than adequate for gaining an understanding of the stress regime,
for checking that the model has been set up properly, or to verify that the boundary
conditions have been properly defined. Estimated properties are usually adequate
for determining the importance of the various properties for the situation being
modeled.
The temptation exists when you have laboratory data in hand that the data needs to
be used in its entirety and cannot be manipulated in any way. There seems to be an
inflexible view of laboratory data which can sometimes create difficulties when
using the data in a numerical model. A common statement is; I measured it in the
lab and I have full confidence in my numbers. There can be a large reality gap
that exists between laboratory determined results and actual in-situ soil behavior.
Some of the limitations arise because of how the material was collected, how it
was sampled and ultimately quantified in the lab. Was the sample collected by the
shovelful, by collecting cuttings or by utilizing a core sampler? What was the size
and number of samples collected and can they be considered representative of the
entire profile? Was the sample oven-dried, sieved and then slurried prior to the test
being performed? Were the large particles removed so the sample could be
trimmed into the measuring device? Some of these common laboratory techniques
can result in unrealistic property functions. Perhaps the amount of data collected in
the laboratory is more than is actually required in the model. Because money has
been spent collecting and measuring the data, it makes modelers reticent to
experiment with making changes to the data to see what effect it has on the
analysis.
It is good modeling practice to first obtain understandable and reasonable solutions
using estimate material properties and then later refine the analysis once you know
what the critical properties are going to be. It can even be more cost effective to
determine ahead of time what material properties control the analysis and decide
where it is appropriate to spend money obtaining laboratory data.

Page 34

SIGMA/W

Chapter 2: Numerical Modeling

Interrogate the results


Powerful numerical models such as SIGMA/W need very careful guidance from
the user. It is easy to inadvertently and unintentionally specify inappropriate
boundary conditions or incorrect material properties. Consequently, it is vitally
important to conduct spot checks on the results to ensure the constraints and
material properties are consistent with what you intended to define and the results
make sense. It is important to check, for example that the boundary condition that
appears in the results is the same as what you thought was specified when defining
the model. Is the intended property function being applied to the correct soil? Or,
are the initial conditions as you assumed?
SIGMA/W has many tools to inspect or interrogate the results. You can view node
or element details and there are a wide range of parameters that can be graphed for
the purpose of spot checking the results.
Inspecting and spot checking your results is an important and vital component in
numerical modeling. It greatly helps to increase your confidence in a solution that
is understandable and definable.
Evaluate results in the context of expected results
The fundamental question that should be asked during modeling is; Do the results
conform to the initial mental picture? If they do not, then your mental picture
needs to be fixed, there is something wrong with the model or both the model and
your concept of the problem need to be adjusted until they agree. The numerical
modeling process needs to be repeated over and over until the solution makes
perfect sense and you are able to look at the results and feel confident that you
understand the processes involved.
Remember the real world
While doing numerical modeling it is important to occasionally ask yourself how
much you really know about the input compared to the complexity of the analysis.
The following cartoon portrays an extreme situation, but underscores a problem
that exists when uneducated or inexperienced users try to use powerful software
tools.

Page 35

Chapter 2: Numerical Modeling

SIGMA/W

If we can incorporate boundary


elements rather than simple finite
elephants, enhance the statistical
evaluation of parameter generation
and stick with the fuzzy sets, I am
confident that accuracy will be
increased to at least the fourth
decimal place.

Put another shovel


full in Pat! It is fullcores theyre wanting!

Note: origins of this figure are unknown at time of printing.

2.6

How not to model

As mentioned earlier in this chapter, it is completely unrealistic to expect to set up


a complex model at the start of a project and immediately obtain realistic,
understandable and meaningful results. There are far too many parameters and
issues which can influence the results, so if this is your expectation, then modeling
is going to lead to major disappointments.
For novice modelers; the initial reaction when faced with incomprehensible results
is that something must be wrong with the software. It must be a limitation of the
software that the solution is inappropriate or completely senseless. It is important
to remember that the software is very powerful; it can keep track of millions of
pieces of information and do repetitive computations which are far beyond the
capability of the human mind. Without the software it would not be possible to
make these types of analyses. The software by itself is extremely powerful

Page 36

SIGMA/W

Chapter 2: Numerical Modeling

numerically speaking, but essentially unintelligent. Conversely, the human mind


has the capability of logic and reasoning, but has significant limitations retaining
large amounts of digital data. It is the combination of the human mind together
with the capability of a computer that makes numerical modeling so immensely
powerful. Nether can do the task in isolation. The software can only be used
effectively under the careful guidance and direction of the modeler.
Sometimes it is suggested that due to a time limitation, it is not possible to start
simple and then progress slowly to a more complex analysis. A solution is needed
quickly and since the budget is limited, it is necessary to immediately start with the
ultimate simulation. This approach is seldom, if ever, successful. Usually this leads
to a lot of frustration and the need to retreat to a simpler model until the solution is
understandable and then build it up again in stages. Not following the above how
to modeling procedures generally leads to requiring more time and financial
resources than if you follow the recommended modeling concepts.
Remember, the software is only as good as your ability to guide and direct it. The
intention of this document is to assist you in providing this guidance and direction
so that you can take full advantages of the power the software can offer.

2.7

Closing remarks

As noted in the introduction, numerical modeling is a relatively new area of


practice. Most university educational curricula do not include courses on how to
approach numerical modeling and, consequently, the skill is often self-taught. As
software tools such as SIGMA/W become increasingly available at educational
institutions and educators become comfortable with these types of tools, classes
and instruction should improve with respect to numerical modeling.
When the numerical analysis software tool, SIGMA/W, is effectively utilized as it
was intended to be used, it becomes an immensely powerful tool, making it
possible to do highly complex analyses. It can even lead to new understandings
about actual physical process.
The process of modeling is a journey of discovery, a way of learning something
new about the complex behavior of our physical world. It is a process that can help
us understand highly complex, real physical process so that we can exercise our
engineering judgment with increased confidence.

Page 37

Chapter 2: Numerical Modeling

SIGMA/W

Page 38

SIGMA/W

Chapter 3: Fundamentals

SIGMA/W: Fundamentals and Practical


Modeling Considerations

3.1

Introduction

SIGMA/W is a powerful finite element method tool that can be used to model a
wide range of stress-strain problems. In its appearance and usability it is much like
other products in GeoStudio. For example, geometry and finite element meshes or
soil regions common to seepage or slope analysis can be used directly in
SIGMA/W. However, there are certain fundamental differences between
SIGMA/W and the other models. Foremost of these differences is that SIGMA/W
uses an incremental load formulation. This means that you enter a step change
condition (e.g. an applied surface load) and it computes step changes in parameters
such as strain and pore-water pressure. It does not compute the actual end
condition unless you have established in-situ conditions prior to loading.
All of the GeoStudio Engineering books have a chapter focusing on Modeling
Tips. The Modeling Tips chapters in the other books generally cover items specific
to enhancing the usefulness of the program from a model set up and data
interpretation perspective. However, due to SIGMA/Ws uniqueness, this book has
this chapter titled Fundamentals and Practical Modeling Considerations and it is
placed near the front of the book in hopes it is read and understood prior to reading
the rest of the book. This entire chapter will be most useful to novice users. More
advanced users will find the first part a good review, but not entirely necessary. All
users should ensure they read and understand the second part of this chapter.
This chapter has two main parts: the first few sections deal with an overview of
fundamental concepts related to stress, volume change and strength of materials. A
detailed theoretical discussion of various soil strength models is given in the
Theory chapter. This discussion will focus more on ensuring there is a basic
understanding of important parameters affecting volume change and soil strength
so that the reader appreciates at a deeper level why SIGMA/W was formulated a
certain way and what it is formulated to do.
The second several sections of this chapter will focus on how the fundamental
concepts have been put into the model and how the model can be used in various
unique and creative ways to solve real problems.

Page 39

Chapter 3: Fundamentals

SIGMA/W

A key concept that will be developed and stressed throughout this chapter is that
SIGMA/W is a very powerful tool for investigating the serviceability of
engineered soil systems. It is not a tool for predicting stability of these systems.
Serviceability deals with obtaining confidence that a design will function as
intended. The numerical soil models available in SIGMA/W can help show that
engineered systems exposed to various applied external or internal loads will not
fail.
Stability, on the other hand, deals with quantification of a factor of safety for an
engineered system. However, in order to obtain a factor of safety it is necessary to
have solved the numerical models to some point past failure so that the stress
condition after failure can be compared to the allowable stress condition at failure.
The problem with this is that the soil models in a small strain stress-deformation
analysis are not defined at stress conditions that exceed failure. Therefore, a safety
factor can not be computed.
It is important to realize this, so that you can begin the modeling and design
process by looking for an engineered system that works! If you generate results
that indicate failure, two things should come to mind: 1) that the results are likely
meaningless and, 2) that you are reminded the true objective is to design a working
system, not one that fails. In other words, model to design a system that is
serviceable.

3.2

Density and unit weight

For many applications in geotechnical engineering, a measurement of the in-situ


density of a soil is required in order to determine the forces exerted by its self
weight. By definition, the density is the mass of soil in a sample divided by the
volume of soil in the sample. In equation form:

m
V

where:

the bulk density of soil,

the mass of soil, and

the volume of soil.

Page 40

SIGMA/W

Chapter 3: Fundamentals

Some caution should be exercised when working with the term density. Density
relates to mass per unit volume, which should not be confused with weight
density denoted by the symbol, . The weight density is also called the unit
weight and it has units of gravitational force (N, lbf) per unit volume. Since
SIGMA/W is formulated to deal with displacement in response to forces, it makes
use of the unit weight form of density.
The bulk, or total density of a soil refers to the mass of bulk soil which includes
soil particles, water and air in a given volume. The bulk density will therefore
change throughout a domain if there is a change in water content throughout the
domain. There are other useful ways of quantifying density: the dry density,
saturated density and buoyant density.
Dry density is the mass of dry soil per unit volume and is commonly referred to
when dealing with compaction of earthworks. The dry density is related to the bulk
density of the soil according to:

d =

1+ w

where:
w

the water content by weight (in decimal form).

The saturated density, sat, describes the bulk density of a soil when all voids are
filled with water and the submerged density is simply the saturated density minus
the density of water as follows:

' = sat w
where:

the submerged density, and

the density of water.

As with the unit weight of the bulk soil (including water weight), the submerged
unit weight can be a useful parameter to use in SIGMA/W. The submerged unit
weight is simply the total unit weight minus the unit weight of water as follows:

' = w

Page 41

Chapter 3: Fundamentals

SIGMA/W

where:
w

the unit weight of water.

Further discussion on the use of submerged unit weight is given in the Analysis
Types chapter discussion on in-situ analysis, and in the chapter on Pore-water
Pressures.

3.3

Compaction and density

Soil is used as a fill material in many engineering projects and in order to obtain
satisfactory serviceability of the project, compaction (densification) of the fill is
often required. Compaction of a loose material has several advantages from a
design perspective, as it can be used to increase shear strength, decrease
compressibility, and decrease permeability. In order to know how much
compaction effort is required in order to achieve the desired soil property, it is
useful to conduct a laboratory compaction test on a representative soil sample from
the site. The objective of the test is to determine, for a given compactive effort, the
optimum moisture content that will result in the maximum dry density. The
laboratory test used most often is the Standard Proctor Test, introduced by R.R.
Proctor in 1933.
Compaction involves application of mechanical energy in order to reduce the air
voids in a soil with little or no reduction in water content. At low moisture
contents, the soil particles are surrounded by thin films of water which tend to keep
the grains separated when compacted. When more water is added, the soil grains
move together more easily which results in a reduction of the air voids and an
increase in bulk (total) density. At some point, making the soil too wet prior to
applying the compactive effort will result in a reduction in density. The water
content that corresponds with the maximum dry density is termed the optimum
water content. Figure 3-1 shows the water contents and maximum dry densities for
a single soil type subjected to different compactive efforts. Inset in the figure are
illustrations of the orientation of soil fabric at different states. Generally, soils are
quantified based on whether they are compacted dry of optimum water content, at
the optimum water content, or wet of optimum water content. The usefulness of
this classification in terms of SIGMA/W will be discussed shortly.

Page 42

SIGMA/W

Chapter 3: Fundamentals

0
10
ati
tur
sa
on

Increasing dry density,

Increasing
compactive
effort

Increasing water content, w

Figure 3-1 Dry density, water content and compactive effort (after Holtz and
Kovacs, 1981)

Notice that adding more effort does not result in elimination of all air voids in a
soil. It does result in reducing the required optimum water content and in
increasing the maximum dry density. This type of a test comparison is very useful
for helping assess the economic benefits of applying less water or with more
compaction effort or visa versa.
Compaction and strength
Knowing the compactive effort and water content applied to an engineered soil can
lead to a better understanding of its anticipated behavior under loading. For clay
type soils, it is recognized that for a fixed compactive effort, as more water is
added, the soil structure becomes increasingly oriented (refer to points A, B and C
in Figure 3-1). That is, the clay minerals become less flocculated and more
dispersed. Dry of optimum, the structure or soil fabric is considered flocculated,
while wet of optimum, it is considered dispersed, or oriented. As the compactive
effort is increased, the amount of orientation is increased (refer to points C and D
in the figure).

Page 43

Chapter 3: Fundamentals

SIGMA/W

A full discussion of the effects on soil properties due to changing water content
and compactive effort is beyond the scope of this book. In general, however, it is
important to understand that soils compacted dry of optimum are likely to have
higher strengths than those compacted wet of optimum. In addition, hydraulic
permeability at a constant compactive effort decreases with increasing water
content and reaches a minimum at about the optimum water content. If compactive
effort is increased, the permeability decreases because the void ratio decreases.
Understanding the anticipated behavior based on these observations can help to
make the model set up and definition easier and the interpretation of output data
more meaningful. For example, if you are solving a consolidation analysis and you
are getting delayed response in dissipation of pore-water pressures within an
engineered structure, you may want to adjust the applied unit weight (a function of
total density) and make a corresponding change to the hydraulic permeability
function applied in SEEP/W. This type of fine tuning soil properties is not usually
recommended. However, you should understand from this discussion that there is a
connection between how things are actually built in the field and what the soil
properties can be.

3.4

Plasticity

SIGMA/W does not require, as input, any of the Atterberg limit parameters.
However, as they are fairly simple to ascertain, it is useful to discuss their
significance from a practical perspective. If you know the soil plastic and liquid
limits, then you can anticipate the soils response and better understand what you
are asking the model to do.
The Atterberg Limits were developed in 1911 as a practical means of describing
the plastic behavior of clays. While there are several classifications, the most
relevant to this discussion are the Plastic Limit, the Liquid Limit and the Liquidity
Index.
The liquid and plastic limits are both water content by weight values expressed in
decimal form. The plastic limit, PL, is the water content at which point the soil
starts to behave plastically when loaded. The liquid limit, LL, is the water content
when the soil just starts to behave as a viscous fluid when loaded. The liquidity
index is defined as:

LI =

w PL
LL PL

Page 44

SIGMA/W

Chapter 3: Fundamentals

where:
w

the natural water content of the soil.

If the LI is between 0 and 1, the soil will behave plastically, and if it is greater than
1, the soil will behave like a viscous fluid when sheared. Figure 3-2 compares
anticipated stress-strain behavior for a range of liquid and plastic limits. It is clear
that as the plastic limit is approached, the potential for larger strains in response to
small changes is shear stress. The figure clearly shows that for soils approaching or
exceeding the liquid limit, the amount of strain in response to small load changes
can be infinite.
This is important to understand. If you are asking the model to compute strain for a
condition where the soil is at a state between its plastic and liquid limits, then there
is chance for large strains. The SIGMA/W formulation is a small strain
formulation, and often it can not converge on a reasonable solution when strains
become too large. In some cases, a solution can be reached, but it can have no
meaning. It is up to you to understand the behavior you expect from the soil and to
model a stress condition that will lead to a meaningful response. If you want a
structure to be useful in its application, then you must design (and model) it for
conditions that are within is range of serviceability.

Page 45

Shear stress

Chapter 3: Fundamentals

SIGMA/W

w < PL

w approaches PL

w approaches LL

w > LL

Strain
Figure 3-2 Stress-strain and Atterberg Limits

3.5

Volume change

Effective and total stress


It is absolutely critical from this point onwards, to have a clear understanding of
effective and total stresses and the difference between the two. Incorrect use of
these parameters will result in incorrect modeled output.
The total stress in a soil is dependent on the total unit weight of the soil above it
plus any applied surface loads (such as free water). Additionally, the total stress at
a given depth in the soil is different in the horizontal and vertical directions.
However, for now, we will consider only the stresses in the vertical direction.
Consider the element of soil illustrated in Figure 3-3 that sits 10m below surface
where the surface is covered by 2m of water. The total vertical stress in the soil
element at this point is equal to the weight of everything above it divided by the
unit cross-sectional area (assumed to be 1m2). In equation form:
n

v = i zi
i

Page 46

SIGMA/W

Chapter 3: Fundamentals

where:
z

the thickness of each material layer,

the total unit weight of each material, and

i,n

the individual and total number of materials.

So, in this case, the total stress equals (2m*10 kN/m3) + (10m*20 kN/m3), or 220
kN/m2 = 220 kPa.

Z=2m, = 10 kN/m3

Z=10m, = 20 kN/m3

Figure 3-3 Illustration of total and effective stress

The effective stress is different than the total stress because the presence of water
in between the soil grains provides a force that can either tend to push the particles
apart or pull them closer together; as would be the case if the water in the soil was
in a capillary state. In equation form, the effective stress is:

v' = v u
where:
u

the pore-water pressure.

Page 47

Chapter 3: Fundamentals

SIGMA/W

Based on the example illustrated in Figure 3-3, the vertical effective stress in the
soil element would be computed as the total stress, 220 kPa, minus the pore-water
pressure (= 12m*10 kN/m3 = 120 kPa) = 100 kPa.
The effective stress is less than the total stress which is, in effect, stating that the
water in the soil is taking up some of the weight (body) load of the soil and thereby
reducing the stress in the soil.
Compressibility and consolidation
Compressibility and consolidation deal with how the soil changes in response to a
change in its loading. When a load is first applied on a soil, it will compress, or
change volume due to:

deformation of the soil grains,

compression of the air and water in the voids, and

squeezing of water and air from the voids.

In general, we neglect the deformation of soil grains and the compressibility of air
and water because they are small compared with the third point.
In a coarse material, the squeezing of water from the voids can occur quite quickly
which results in a rapid compression of the soil under loading. However, in a fine
material, the rate of compression is directly a function of the hydraulic
permeability of the soil and, therefore, the compression is time dependent. This
phenomenon is called consolidation.
Consider Figure 3-4 where a submerged 1m2 footing with a 50 kN load is applied
to the previously illustrated soil element. Two important points can be made using
this illustration.
When the load is first applied, there is a change in total stress on the soil of 50
kN/m2 or 50 kPa. Therefore, the total stress in the soil element is now 270 kPa.
When the load is first applied, there is no time for the finer grained soil particles to
shift around in order to assume the new load so the pore-water carries the load and
its pressure increases by an amount equal to the applied new stress. Therefore, the
pore-water pressure at the soil element increases by 50 kPa and the new effective
stress is computed to be 270 kPa minus 170 kPa = 100 kPa; which is the same as
the starting effective stress. If there is no change in effective stress, there is no
change in soil volume.

Page 48

SIGMA/W

Chapter 3: Fundamentals

50 kN load

Z=2m, = 10 kN/m3

Z=10m, = 20 kN/m3

Figure 3-4 Surface load applied to example

Because this is a fine grained soil, we know that with time, the excess pore-water
pressure will dissipate back to the pre-load value. When this happens, the effective
stress will be computed as the new total stress minus the original pore-water
pressure, or 270 kPa 120 kPa = 150 kPa. The effective stress has increased which
means that the soil must change volume or consolidate.
No change in effective stress means no volume change. An increase in effective stress
means consolidation or shrinkage. A decrease in effective stress means swelling or heave.

Sometimes it is useful to plot individual effective stress parameters at the start and
end of a load step. Consider Figure 3-5 which shows water pressure response to the
loading discussed in Figure 3-4. The new total stress load here is considered to be
applied instantly which results in an equal increase in pore-water pressure at the
start of the consolidation phase.

Page 49

Chapter 3: Fundamentals

Z=2

SIGMA/W

= u

Pressure
dissipation over
time

Z=10

u end

u start

Figure 3-5 Dissipation of pressure over time

We can use this same approach and simultaneously consider the calculation of all
parameters for the two cases described above. Figure 3-6 shows the calculation of
all stresses and pressures for the two examples as a function of depth to the soil
element. The top of the figure shows the calculation of effective stress for the insitu case (prior to the load step) while the bottom of the figure shows what the
stresses are just after the footing load is applied and after consolidation has
occurred.
There are several key points to understand from this figure:

In order to get the actual stresses after consolidation, the in-situ


stresses must be known ahead of time.

The effective stress at the ground surface is zero at all times.

The step load from the footing instantly increases the pore water
pressure because the soil has not had time to dissipate the pressure.
The pressure increase is equal to the newly applied load.

There is no change in total stress over time after the loading.

There is no initial change in effective stress at the time the load is


applied.

As the pore-water pressure dissipates, the effective stress increases


towards its new value.

Page 50

SIGMA/W

Chapter 3: Fundamentals

=0
Z=2

=0

u=0
20

20

=-u

Z=10

220

120

100

Pressure / Stress units

=0
Z=2

Pressure
dissipation
over time = 0

u=0
70

70

Increase in
effective stress
with time

= 50
Z=10

=-u

270

120

170
u decreases

Pressure / Stress units

100

150

increases

Figure 3-6 Stress and pressure profiles for two example cases

A concept that should be getting more evident at this point is that we can not
generally compute the actual stresses in the ground at any given point in time
simply based on the applied loads at that time. We must consider how we have
changed the external conditions at any given time and then compute the change in
the grounds response. For modeling purposes, we can, and should, think of our
effective stress equation as:

v' = v u

Page 51

Chapter 3: Fundamentals

SIGMA/W

where the delta symbol indicates a step change in each parameter. From our earlier
examples, the change in total stress, , equals 50 kPa and the change in porewater pressure, u, is also 50 kPa. Therefore, the change in effective stress just
after loading is equal to zero, and, there is no volume change.
This concept of changes in loads, or incremental loading, is discussed in more
detail later in this chapter. It is a critical concept to understand.
Zero change in pore-water pressure
What happens if there is no pore-water pressure change during a change in total
stress? If this is the case then u = 0 and our effective stress equation becomes:

v' = v 0
or

v' = v
In this case, the change in effective stress is equal to the total stress. By always
thinking of the stresses in the soil in terms of effective stress, our rule about
volume change only occurring with a change in effective stress always applies. It
will get confusing if you start thinking of volume change in response to changes in
total stress some of the time, and in terms of effective stress at other times. Always
understand what the pore-water pressure changes are (or are not), and think in
terms of effective stress.
Ko horizontal and vertical stresses
We know that with depth the water pressure acts equally in both the horizontal and
vertical directions. This is rarely the case for soils. We can express the ratio of
horizontal to vertical stress in the ground as:

h = K v
where:
K

the earth pressure coefficient.

This is not generally a valuable parameter to use because there can be variations in
the water content and total stress in the soil which means K is not constant. A

Page 52

SIGMA/W

Chapter 3: Fundamentals

better way to relate the horizontal to vertical stresses is to think in terms of


effective stress as follows:

h' = K 0 v'
where:
KO

the coefficient of lateral earth pressure at rest.

This parameter is more useful because it relates effective stresses and is


independent of the location of the water table. So, as long as the soil layer is a
constant material, the K0 value will be constant within that layer.
K0 is a function of material type and stress history and it can be estimated in
several ways. In SIGMA/W the value is related to Poissons ratio as follows:

K0 =

where:

Poissons ration (limited to 0.49 in SIGMA/W).

A detailed discussion of lateral earth pressure coefficients is beyond the scope of


this chapter. At this time, however, it is useful to work through an exercise to show
how a hand calculation can be used to check modeled results. This is especially
useful to help determine that the model has computed the correct in-situ total stress
conditions.
We know from the previous examples that it is quite easy to determine the vertical
effective stress at any depth. If we know that the effective vertical stress is 100 kPa
and the pore-water pressure is 120 kPa, then we can use a K0 equal to 0.5 to
compute the effective horizontal stress as 50 kPa. Once the effective horizontal
stress is known, we can back calculate what the total horizontal stress is as follows:

h' = h u
h = h' + u
h = 50 + 120 = 170kPa

Page 53

Chapter 3: Fundamentals

SIGMA/W

Notice that the horizontal total stress is not equal to one half of the total vertical
stress of 220 kPa. It would be an error to apply the K0 value to the total stress. The
only way to get the correct total stress is to follow the procedure outlined above.
Having said that, the K0 hand calculation method only applies to an in-situ analysis
where the ground surface is horizontal and the vertical effective stress can readily
be computed as a function of depth. For the case where the ground is sloping and
the water table is not horizontal it is necessary to use a more advanced analysis.
This is discussed further in the Analysis Types chapter of this book.

3.6

Strength parameters

Mohr stress circle


The Mohr circle can be used to state of stress at a point in equilibrium. Consider
the soil element in Figure 3-7 with some horizontal, vertical, shear, and normal
forces acting upon it.
N= x1

H = x Sin

A = Sin

T= x1

A=

A = Cos

V = y Cos

Figure 3-7 Forces acting on a soil element

If the forces in on each face are broken into horizontal and vertical components
and converted to stresses by dividing them by the cross sectional area upon which
they act, two equations in terms of , and are developed. If these equations are
each set to zero, they can be squared and added together. The end result is a single
equation for a circle with a radius of:

Page 54

SIGMA/W

Chapter 3: Fundamentals

( x y )
2
and with its center at:

( x + y )
, 0

The vertical and horizontal planes in the soil element have no shear force acting on
them so they are, by definition, principal planes. Therefore, the forces acting on
them are principal forces. An important point to remember here is that principal
stresses act on planes where the shear strength is zero. Another point to note is that
the stress with the largest magnitude is called the major principal stress (denoted
1), and the lower value stress is the minor principal stress (denoted 3).
If the information from the summation of forces equations can be used to construct
a circle on the shear-normal stress plane as shown in Figure 3-8. Using this circle,
the applied normal and shear stress along any potential failure plane in a soil
element can be determined. In this case, the shear and normal stress point is shown
for a plane inclined at an angle, , from horizontal.

( , )

1
Figure 3-8 Mohr circle of stress

Page 55

Chapter 3: Fundamentals

SIGMA/W

Mohr Coulomb failure criteria


The Mohr stress circle can be assembled for a series of shear tests on a similar soil
using different applied normal loads and then combined with a frictional and
cohesive strength theory developed by Coulomb to arrive at the Mohr Coulomb
failure criteria. A more detailed discussion of this criterion is available in most
geotechnical engineering text books so it is omitted from this book. The purpose of
introducing it now is to help in our subsequent discussion of effective and total
strength parameters and how they are obtained and what they mean.

'

c'
3'

1'

3'

1'

Figure 3-9 Mohr Coulomb effective failure parameters

Figure 3-9 shows a Mohr Coulomb failure envelope for a clay type soil. Two Mohr
circles are plotted in this figure and a line is drawn tangent to the two circles such
that it also intersects the vertical shear stress axis. The point of intersection of the
tangent line provides the effective cohesion parameter and the angle of inclination
of the tangent line provides the effective friction angle. All of the parameters
illustrated in the figure have the apostrophe symbol, which indicates that they
are effective stress based values and not total stress based values. They are
effective stress based because the pore-water pressure in the sample was measured
during the testing and subtracted from the total stress values applied in the test.
Once the data from the shear tests has been graphed, as shown in Figure 3-9, it is
possible to create the equation that represents the data. We see that the tangent line
is a straight line with a y-intercept. Therefore, the equation of the line, termed the
Mohr Coulomb failure criteria, is given by:

f = c '+ ' f (tan ')

Page 56

SIGMA/W

Chapter 3: Fundamentals

where the subscript, f, denotes the shear and normal stress condition in the soil
element at failure.
Triaxial testing
Entire textbooks can be written on the subject of performing and interpreting data
from the various types of triaxial tests on various types of soil. This type of a
discussion is far too advanced in this context. It is important, however, to
distinguish between two main types of test that can be carried out in a triaxial cell.
Undrained tests
Undrained tests can be performed on undisturbed as well as disturbed samples of
clay and silt, and disturbed samples of sand or gravel. The sample is allowed to
consolidate under a desired confining pressure and then sheared in axial
compression.
In an un-drained test, the pore-water in the sample is not allowed dissipate, so the
pressure builds up in response to changes in total stress. The changes in stress
result from application of an applied strain to the top of the sample. So, for
example, if the applied strain results in a principal major deviator stress of 50 kPa,
the water pressure in the soil should also increase by an equal amount.
The undrained test can be performed with or without pore-water pressure
measurements. If pore-water pressure is not measured, the results must be
expressed in terms of total stress. The field conditions that a soil consolidates
under are usually quite different than the conditions in the laboratory so the results
of this test, when expressed in terms of total stress, can only be applied in a limited
way to the field.
The results of a total stress triaxial test would look like those illustrated in Figure
3-10.

Page 57

Chapter 3: Fundamentals

SIGMA/W

=0

cu

1 3

Figure 3-10 Mohr circles for undrained, total stress tests

If the pore-water pressure is measured during loading, the rate of applied strain
must be slow enough to allow the pressures to equalize and be measured. With the
measurement of pore-water pressure, the results can be expressed in terms of
effective stress and the values of c and can then be applied to a wider range of
field applications. The results of an undrained effective stress triaxial test were
illustrated previously in Figure 3-9.
The undrained test results are useful when a soil structure has had a long time to
come to equilibrium and then a sudden change in total stress is applied with
limited drainage. This may be the case for application of fill or rapid drawdown of
a reservoir.
Note, in SIGMA/W, if you use undrained total stress soil property parameters then
an estimate of the pore-water pressure response can be obtained using the porewater pressure parameters A and B. These are discussed in more detail in the
following section and in the Theory chapter.
Drained tests
A drained test can be performed on all types of samples, whether disturbed or not.
It is prepared in the same manner as an undrained test sample, but during the
loading stage, the water is allowed to drain further. The applied rate of strain must
be slow enough to prevent the development of excess pore-water pressures; in
other words, it should remain at the state it was in after initial consolidation prior
to loading.

Page 58

SIGMA/W

Chapter 3: Fundamentals

At completion of testing, the Mohr circle failure envelope will look very similar to
that shown for the undrained effective stress analysis, which makes sense because
if the soil is permitted to drain, then the change in pore-water pressure will be zero
and the change in effective stress will equal the change in total stress.
The drained test results are quite useful to use when the long term seepage
dependent consolidation is an issue. However, it is not an easy test to conduct
because it is important to ensure that pressures are not building up in the sample. In
addition, for low permeability soils, it can take a long time to allow drainage which
means a very slow rate of applied strain is required.
Soil stiffness Youngs modulus
Regardless of how the testing is carried out, the objective is to determine the
strength parameters (described above) and also the stress-strain behavior. The
stress-strain behavior is the actual constitutive model used in the solution of the
partial differential equations. The strength parameters are only useful for telling if
any given soil element has exceeded its yield point. Further discussion on the
usefulness of strength parameters in stress-deformation modeling is included in the
section on stability and serviceability below.
Consider Figure 3-11, which shows the stress-strain behavior for three confining
loads during a strength test. As the confining stress increases, so does the slope of
the deformation curve. It is the slope of these curves that relates the amount of
strain in response to an applied stress. In equation form, the slope is given by:

Ei =

or:

= Ei
which is better known as Hookes Law.
The slope, E, is known as Youngs Modulus and it is NOT constant for any given
soil. This modulus, which represents the stiffness of the soil, is dependent on the
effective confining stress, and it can change throughout the entire problem domain
depending on the stresses throughout the domain. So, the parameter we are solving
in SIGMA/W is strain, and in order to know strain we must know the stress and the
stiffness. However, the stiffness is itself a function of stress. This circular logic is

Page 59

Chapter 3: Fundamentals

SIGMA/W

Stress, (13)

what makes the solution a non-linear process, and it is why we have to iterate to
reach a converged result.

Strain,
Figure 3-11 Stress-strain behavior for various confining loads during testing

3.7

Total stress, effective stress and pore-water pressures

As we have been discussing in the previous few sections, the presence of water
under pressure can affect both how we obtain material properties and how they are
used in the solution. Following common geotechnical practice, material properties
can be specified in SIGMA/W using either effective stress parameters for analyses
of drained soils, or total stress parameters for undrained soils. In materials
specified using total stress parameters, undrained pore-water pressure changes can
be computed from total stress changes using pore-pressure parameters A and B. If
you want to model a situation using effective stress parameters with pore-water
pressure change, then you must do a fully coupled consolidation type analysis
where pore-water pressures are computed simultaneously using SEEP/W. An
uncoupled consolidation analysis does not compute pore-water pressure changes
due to changes in total stress, because the pore-water pressure changes are
computed independent of total stress changes in an external program such as
SEEP/W, SIGMA/W, VADOSE/W or QUAKE/W.

Page 60

SIGMA/W

Chapter 3: Fundamentals

When you specify the material properties as effective stress parameters, then
SIGMA/W will take the existing total stress, subtract the specified pore-water
pressure and arrive at an effective stress. This effective stress is then the value used
to obtain the soil stiffness in a non-linear soil model. The solver then computes the
change in total stress due to loading and it ADDS back on the original pore-water
pressures in order to report a new effective stress at the end of the load step. So, in
every case where effective stress parameters are specified and the analysis is not
fully coupled, the change in effective stress will always equal the change in total
stress.
In every case where effective stress parameters are specified and the analysis is not fully
coupled, the change in effective stress will always equal the change in total stress.

When you specify material properties as total stress parameters, you have the
option to specify A and B pore-water pressure parameters. If you specify valid A
and B pore-water pressure parameters, then SIGMA/W will compute a change in
pore-water pressure based on those parameters. Consider the equation for porewater pressure change using A and B parameters shown below. A full discussion
of this equation is provided in the Theory chapter.

+ 2 + 3
u = 1
+
3

( 1 2 ) + ( 2 3 ) + ( 3 1 )
2

where:

B, and

A.

From this equation, it is clear that any generated pore-water pressure is due to a
stress change consisting of two components; namely, u, which is a function of
changes in total stresses, and u, which is due to changes in deviatoric stress. It
can be quite common to neglect a change in pore-water pressure along shearing
planes so the A parameter can be set as zero.
Use of the B parameter requires a bit of thinking. Often, it is desired to have the
pore-water pressure response be equal to a change in total stress during loading.
This can be achieved by setting the B value to 1 in a fully saturated soil and then

Page 61

Chapter 3: Fundamentals

SIGMA/W

also setting Poissons ratio to a value of 0.49 (ideally 0.5, but for numerical
reasons it is fixed at 0.49 as a maximum value). With this Poissons ratio, the
average in all three direction will equal the applied and then u will also
equal .
The pore-pressure parameters A and B are defined in SIGMA/W using functions.
B is defined as a function of the pore-water pressure and A is defined as a function
of the deviatoric stress. The B function can be defined for both positive and
negative pore-water pressures, which makes it possible to analyze both saturated
and unsaturated soil conditions. Example functions of these parameters are shown
in Figure 3-12 and Figure 3-13 .
In a load-deformation analysis, materials with properties specified using effective
stress parameters and materials specified using total stress properties can be
intermixed. However, in a fully coupled consolidation analysis using SEEP/W
only materials with effective stress parameters are allowed.

Figure 3-12 Example of pore-water pressure B function

Page 62

SIGMA/W

Chapter 3: Fundamentals

Figure 3-13 Example of pore-water pressure A function

It is possible to specify a B function, but no A function. This means that your porewater pressure change will result purely from volumetric strain and not from any
deviatoric stress change as discussed above. No pore-water pressure changes will
be computed if both parameter A and parameter B are undefined in a total stress
analysis.

3.8

Incremental formulation

SIGMA/W is formulated for incremental analysis. For each time step, incremental
displacements are calculated for the incremental applied load. These incremental
values are then added to the values from the previous time step and the
accumulated values are reported in the output files. In equation form:

step 2 = step1 + new


The incremental formulation has several advantages and, if used creatively, can be
an effective way to establish unique loading or stress conditions. At this point it is
important to remember that time is not a parameter of relevance for all analysis
types except consolidation which relies on time dependent dissipation of porewater pressures. Because time is not relevant, the incremental formulation is not
always sensitive to the order of loading. What is important is that the load steps

Page 63

Chapter 3: Fundamentals

SIGMA/W

can be added or subtracted such that at the end of the loading sequence, the ground
stresses reflect the desired conditions.
Consider for example that you want to set up some initial conditions such that the
total horizontal stress at a depth of 10 m is 200 kPa , while the total vertical stress
is only 100 kPa. Assume, for now, that the unit weight of the soil is 20 kN/m3. It is
not common in reality to have the horizontal stress be greater than the vertical
stress, but this example makes a good point of illustrating how some creative
thinking using incremental load steps can work to your advantage.
There are at least two ways to establish the desired in-situ stress conditions. In the
first approach:
1. Set the Ko value to be 2 and run an In-situ 1 type of analysis. This
would get the desired result in one step.
2. However, to illustrate a point, the same conditions could be
established by the following two step approach:

Step 1: Set Ko to be 1 which results in the horizontal and vertical


stresses at a depth of 10 m being equal to 200 kPa.

Step 2: Change the unit weight to a value of -10 kN/m3 and give
Poissons ratio a value of zero. Solve a load / deformation analysis
using the results of step 1 as the initial conditions for step 2. The result
of this load step is to subtract 100 kPa from the vertical stresses and 0
kPa from the horizontal stress.

In both cases, the end result is a horizontal stress of 200 kPa and a vertical stress of
100 kPa.
Obviously, if you had a choice, you would choose the one step approach to get the
desired result. The point of this example is that the incremental formulation allows
to you add and subtract loads, such that you end up with the ground stress
conditions you think exist in the field.
Body forces
The body force is the self weight of the soil and it must be included in an analysis
if actual stresses in the ground are desired, or if a non-linear soil model requires
that the actual stresses are known so that the correct soil stiffness values can be
obtained. The body force can be left out of a linear-elastic soil model analysis,

Page 64

SIGMA/W

Chapter 3: Fundamentals

however, you must realize that any stresses reported are only incremental stresses
due to the added load, and not the load plus the weight of the soil itself.
Full details of how the body loads are computed are given in the Theory chapter.
For now, it is only necessary to realize that the body load applied in the model is
simply a function of the unit weight specified and the size of the finite element. In
other words, a finite element shape one meter square with the assumed one meter
unit thickness would have one cubic meter volume. The body load applied at each
node in the element would then have a proportion of the unit weight multiplied by
the elemental volume.
Using the incremental approach, the unit body force is generally only applied when
an element is included for the first time during an analysis. However, if there is a
numerical advantage to apply the body load over several load steps, then it is
possible to set this as an option.
If the body forces were applied in an in-situ analysis, and then the results from the
in-situ analysis are used as the starting condition for a different analysis, you must
turn off the body forces in the new analysis. It does not know they were applied
previously, so it will apply them again in the first loading step if valid body force
data exists in the file. This point is often over-looked by users. Hopefully, if you
understand how the incremental loading formulation works, you will check what
you are asking the model to do.

3.9

Serviceability versus stability

As stated in the introduction to this chapter, SIGMA/W is a very powerful tool for
investigating the serviceability of engineered soil systems. It is not a tool for
predicting stability of these systems.
Serviceability deals with obtaining confidence that a design will function as
intended. Serviceability can tell you how a soil will move towards an unstable state
(e.g. going excessively plastic), but it does not tell you, or quantify for you, what
that state of instability is. Stability, on the other hand, deals with quantification of a
factor of safety for an engineered system. It deals with applied loads and resisting
forces. It says nothing about how a soil moves from one state to the other. A limit
equilibrium tool such as SLOPE/W should be used where factors of safety are the
primary concern.
In this chapter we have discussed some fundamental soil properties and how they
can be measured, computed and interpreted in the model. One of the more

Page 65

Chapter 3: Fundamentals

SIGMA/W

important concepts introduced was that the strength properties of a soil are a
measurement of the failure shear strength under different principal stresses. This
has significant implications in a numerical stress-deformation model, because it is
very likely that there are several failure planes throughout the geometry, and that
they are at various orientations to the principal stresses. Given the highly varied
combination, and add the fact that there is no way to quantify a soil strength once it
has failed, how can a quantification of stability be made?
In a limit equilibrium analysis, the potential failure plane is specified. With this
information, it is possible to determine what the shear strength is along that plane
given principle stresses acting at that point. The mobilized shear strength can be
divided by the resisting shear strength to determine a stability factor greater or
less than 1.0. This can not be done in stress-deformation analysis.

3.10

Modeling progression

A finite element analysis consists of two steps. The first step is to model the
problem, while the second step is to formulate and solve the associated finite
element equations. Modelling involves designing the mesh, defining the material
properties, choosing the appropriate constitutive soil model, and defining the
boundary conditions. SIGMA/W can formulate and solve the finite element
equations. The modelling, however, must be done by the user; that is, the user
must design an acceptable mesh, select the applicable soil properties, and control
the boundary conditions. SIGMA/W cannot make judgments; this is your
responsibility as the user.
Good modelling techniques require practice and experience. To assist you with this
part of the analysis, this chapter presents some general modelling guidelines. The
information presented is not an exhaustive statement on the "how-to" of modelling,
but instead provides suggestions on how to model various conditions and how to
outline conditions that may lead to difficulties.
One of the most important rules to follow in finite element modelling is to progress
from the simple to the complex. When you include all the possible complexities at
the start of an analysis, it often becomes difficult to interpret the results,
particularly when the results are unrealistic. Moving from the simple to the
complex makes it easier to pinpoint the reasons for unrealistic results. It is
therefore good practice to first define a simplified version of the problem and then
add complexity in stages.

Page 66

SIGMA/W

Chapter 3: Fundamentals

In finite element modelling, it is also important that the results obtained are of a
form similar to results obtained from simple hand calculations. It is easier to make
this judgment if you start with a simplified version of the problem.
For example, a good method of moving from a simple analysis to a complex
analysis is to start with a linear-elastic model. These results provide a reference
with which to compare more complex nonlinear results, and they are also a means
of checking that the boundary conditions have been applied correctly. Once you
are satisfied that this part of the analysis is correct, you can move onto the
nonlinear analysis with more confidence.

3.11

Units

Any set of units can be used in a SIGMA/W analysis. However, the units must be
used consistently throughout the analysis. For example, if the geometry is defined
in meters, then deformation is in meters.
Units must be selected for length, force and unit weight. The unit weight of water
is set when the units of length are selected. Table 3-1 shows examples of consistent
sets of units.
Table 3-1 Examples of consistent sets of units
Property

Units

Metric

Imperial

Geometry

meters

feet

pcf

Unit Weight of Water

F/L

kN/m

kN/m

pcf

Cohesion

F/L

kPa

pcf

Pressure

F/L2

kPa

psf

Force

kN

lbs.

kPa

psf

Soil Unit Weight

E (modulus)

F/L

F/L

The unit weight of water must be entered correctly for whichever system of units
you may have chosen. This unit weight is used in SIGMA/W to convert hydraulic
heads into pressures. It is also used to calculate a limiting confining stress in some
nonlinear constitutive models. For example, if the unit weight of water is kN/m3,
the minimum principal stress is limited to 1 kPa in the Elastic-Plastic Model when
the internal friction angle is greater than 0. If the unit weight is 62.4 pcf, this

Page 67

Chapter 3: Fundamentals

SIGMA/W

limiting value is approximately 20 psf. For more information about limiting


confining stress, see Elastic-Plastic Model in the Material Properties chapter.

Page 68

SIGMA/W

Meshing

4.1

Introduction

Chapter 4: Meshing

Finite element numerical methods are based on the concept of subdividing a


continuum into small pieces, describing the behavior or actions of the individual
pieces and then reconnecting all the pieces to represent the behavior of the
continuum as a whole. This process of subdividing the continuum into smaller
pieces is known as discretization or meshing. The pieces are known as finite
elements.
Discretization or meshing is one of the three fundamental aspects of finite element
modeling. The other two are defining material properties and boundary conditions.
Discretization involves defining geometry, distance, area, and volume. It is the
component that deals with the physical dimensions of the domain.
A numerical book keeping scheme is required to keep track of all the elements and
to know how all the elements are interconnected. This requires an ordered
numbering scheme. When finite element methods were first developed, creating
the mesh numbering was very laborious. However, many computer algorithms are
now available to develop the mesh and assign the element numbering. Developing
these algorithms is in some respects more complex than solving the main finite
element equations. GeoStudio has its own system and algorithms for meshing,
which are designed specifically for the analysis of geotechnical and geoenvironmental problems.
Some human guidance is required to develop a good finite element mesh in
addition to using the powerful automatic meshing algorithms available. One of the
issues, for example, is mesh size. Computers, particularly desktop or personal
computers, have limited processing capability and therefore the size of the mesh
needs to be limited. Variable mesh density is sometimes required to obtain a
balance between computer processing time and solution requirements. Ensuring
that all the elements are connected properly is another issue. Much of this can be
done with the meshing algorithm, but it is necessary for the user to follow some
fundamental principles. In finite element terminology this is referred to as ensuring
mesh compatibility. GeoStudio ensures mesh compatibility within a region, but the
user needs to provide some guidance in ensuring compatibility between regions.

Page 69

Chapter 4: Meshing

SIGMA/W

The purpose of this chapter is to introduce some of the basic concepts inherent in
meshing and outline some procedures which must be followed when developing a
mesh. An understanding of these fundamentals is vital to proper discretization.
Much of this chapter is devoted to describing the meshing systems and the features
and capabilities available in GeoStudio. In addition, there are also discussions on
the selection, behavior and use of various element types, sizes, shapes and patterns.
A summary of practical guidelines for good meshing practice are also outlined.

4.2

Element fundamentals

Element nodes
One of the main features of a finite element is the nodes. Nodes exist at the corners
of the elements or along the edges of the elements. Figure 4-1 and Figure 4-2 show
the nodes, represented as black dots.

The nodes are required and used for the following purposes:

The positions of the nodes in a coordinate system are used to compute


the geometric characteristics of the element such as length, area or
volume.

The nodes are used to describe the distribution of the primary


unknowns within the element. In the SIGMA/W formulation, the
primary field variable is displacement.

The nodes are used to connect or join all the elements within a
domain. All elements with a common node are connected at that node.
It is the common nodes between elements that ensure compatibility,
which is discussed in further detail below.

All finite element equations are formed at the nodes. All elements common to a
single node contribute to the characteristics and coefficients that exist in the
equation at that node, but it is the equation at the node that is used to compute the
primary unknown at that node. In other words, the seepage equation is developed
for each node and the material properties which are used within the equations are
contributed from the surrounding elements.
There can be multiple finite element equations developed at each node depending
on the degrees of freedom. In a SIGMA/W analysis there are up to three degrees of
freedom at each node, and the number of finite element equations to be solved is

Page 70

SIGMA/W

Chapter 4: Meshing

equal to the number of nodes used to define the mesh multiplied by the number of
degrees of freedom at each node. In a 2D stress-deformation analysis, there are two
degrees of freedom at each node displacement x and displacement y.
Consequently, the number of equations for the whole domain is equal to two times
the number of nodes. In a coupled consolidation analysis there are three degrees of
freedom at each node displacement x, displacement y and pore-water pressure.
For a coupled consolidation analysis the total number of equations required to
solve the problem is three times the number of nodes. For a stress-deformation
analysis with structural beam elements, there is an additional rotational degree of
freedom at beam nodes only.
Since the number of finite element equations is related to the number of nodes, the
number of nodes in a problem is one of the main factors in the computing time
required to solve for the primary unknowns.
Field variable distribution
In a finite element formulation it is necessary to adopt a model describing the
distribution of the primary variable within the element (e.g., displacement). The
distribution could be linear or curved.
For a linear distribution of the primary unknown, nodes are required only at the
element corners. The two nodes (points) along an edge are sufficient to form a
linear equation. Figure 4-1 illustrates this situation. Elements with nodes existing
at the corners are referred to as first-order elements.

Figure 4-1 Primary field variable distribution in first-order elements

The derivative of the primary unknown with respect to distance is the gradient. For
a linear distribution the gradient is consequently a constant. In the context of a
stress-deformation formulation the primary unknown is the displacement. The
derivative of displacement with respect to distance is the gradient (or strain) and
the gradient is therefore constant within a first order element.

Page 71

Chapter 4: Meshing

SIGMA/W

With three nodes defined along an edge, we can write a quadratic equation
describing the distribution of the primary unknown within the element.
Consequently the distribution of the primary unknown can be curved as shown in
Figure 4-2. The derivative of the quadratic displacement distribution results in a
linear gradient distribution. Elements with three or more nodes along an edge are
referred to as higher order elements. More specifically, an element with three
nodes along an edge is known as a second-order element.

Figure 4-2 Primary field variable distribution in higher-order elements

Higher order elements are more suited to problems where the primary unknowns
are vectors as in a stress-deformation analysis (deformation x and y). When the
primary unknown is a scalar, there is often little to be gained by using higher-order
elements. Smaller first-order elements can be as effective as larger higher-order
elements. This is discussed in more detail in the meshing guidelines at the end of
this chapter.
Element and mesh compatibility
Element and mesh compatibility are fundamental to proper meshing. Elements
must have common nodes in order to be considered connected, and the distribution
of the primary unknown along an element edge must be the same for an edge
common to two elements.
Consider the illustration in Figure 4-3. Element numbers are shown in the middle
of the element and node numbers are presented beside the nodes. Even though
elements 4, 5 and 6 appear to be connected to elements 7, 8 and 9, they are actually
not connected. Physically, the elements would behave the same as the two element
groups shown with a physical separation on the right side of Figure 4-3. Common
nodes are required to connect the elements as shown in Figure 4-4. Node 11, for
example, is common to Elements 5, 6, 8 and 9.

Page 72

SIGMA/W

Chapter 4: Meshing

Mixing elements of a different order can also create incompatibility. Figure 4-5
shows 4-noded quadrilateral elements connected to 8-noded elements. Elements 1
and 2 are 8-noded elements while Elements 3 to 10 are 4-noded first-order
elements. The field variable distribution in Element 1 along edge 9 to 11 could be
curved. In Elements 3 and 4 the field variable distribution between 9 and 10 and
between 10 and 11 will be linear. This means the field variable distributions
between Elements 1 and 2 are incompatible with the field variable distributions in
Elements 3 to 6.
The meshing algorithms in GeoStudio ensure element compatibility within
regions. A special integer-based algorithm is also included to check the
compatibility between regions. This algorithm ensures that common edges between
regions have the same number of elements and nodes. Even though the software is
very powerful and seeks to ensure mesh compatibility, the user nonetheless needs
to be careful about creating adjoining regions. The illustration in Figure 4-3 can
also potentially exist at the region level. At the region level, region points need to
be common to adjoining regions to ensure compatibility.

16
15

20

14
13

19

12
11

10
9

15

23

18

31

39

22
26

17

30

43

38

23

42
19

13

21
25

16

29

46

37

22

41

33

Figure 4-3 Disconnected elements lack of compatibility

Page 73

47

34

10

17

24

20
14

48

35

11

18

44

36
21

27

7
1

32

12

8
2

40
28

9
3

24

45

Chapter 4: Meshing

SIGMA/W

8
3

12
6

7
2

11

20
12

15
8

10
4

16

19
11

14
7

18
10

13

17

Figure 4-4 Connected elements compatibility satisfied


5

13
12
11
10
9

18
6
5
4
3

17
16
15
14

23
10
9
8
7

22
21
20
19

Figure 4-5 Element incompatibility

The integer programming algorithm in GeoStudio seeks to ensure that the same
number of element divisions exist between points along a region edge. The number
of element divisions are automatically adjusted in each region until this condition
is satisfied. It is for this reason that you will often notice that the number of
divisions along a region edge is higher than what was specified. The algorithm
computes the number of divisions required to achieve region compatibility.
Numerical integration
In a finite element formulation there are many integrals to be determined, as shown
in the Theory chapter. For example, the integral to form the element characteristic
matrix is:

Page 74

SIGMA/W

Chapter 4: Meshing

[ B ] [C ][ B ] dv
t

For simple element shapes like 3-noded or 4-noded brick (rectangular) elements, it
is possible to develop closed-formed solutions to obtain the integrals, but for
higher-order and more complex shapes it is necessary to use numerical integration.
GeoStudio uses the Gauss quadrature scheme. Basically, this scheme involves
sampling the element characteristics at specific points known as Gauss points and
then adding up the sampled information. Specific details of the numerical
integration in GeoStudio are presented in the Theory Chapter.
Generally, it is not necessary for most users to have a comprehensive
understanding of the Gauss integration method, but it is necessary to understand
some of the fundamentals since there are several options in the software related to
this issue and some results are presented at the Gauss sampling points.
The following table shows the options available. Use of the defaults is
recommended except for users who are intimately familiar with numerical
integration and understand the significance of the various options. The integration
point options are part of the meshing operations in GeoStudio. In the table you will
see two comments related to special cases when using the elastic-plastic soil
model. The reasons for this are discussed in the meshing guidelines section later in
this chapter.
Element Type

Integration Points

Comments

4-noded quadrilateral

Default

8-noded quadrilateral

4 or 9

4 is the default

3-noded triangle

1 or 3

3 is the default

6-noded triangle

Default

Some finite element results are computed at the Gauss sampling points. GeoStudio
presents the results for a Gauss region, but the associated data is actually computed
at the exact Gauss integration sampling point. Even though a Gauss region is
displayed, the data is not necessarily constant within the region.
With the View Element Information command, you can click inside an element
and the nearest Gauss region is displayed together with all the associated
information. The number of Gauss regions within an element is equal to the
number of Gauss integration points used in the analysis.

Page 75

Chapter 4: Meshing

SIGMA/W

It is important to be cognizant of the impact of Gauss points on computing time


and data storage. Nine-point integration in a quadrilateral element, for example,
means that the element properties need to be sampled nine times to form the
element characteristic matrix and element data is computed and stored at nine
points. This requires more than twice the computing time and disk storage than for
four-point integration. Sometimes nine-point integration is necessary, but the
option needs to be used selectively.
Secondary variables
Earlier it was noted that finite element equations are formed at the nodes and the
primary unknowns are computed at the nodes. Once the primary unknowns have
been computed, other variables of interest can be computed such as the shear stress
within the element. Since these parameters are computed after the primary values
are known, they are called secondary variables.
Secondary quantities are computed at the Gauss integration points. GeoStudio
displays a Gauss region, but the associated values are strictly correct only at the
Gauss integration point.
For contouring and graphing, the secondary values are projected and then averaged
at the nodes. This can sometimes result in unrealistic values if the parameter
variations are excessive between Gauss points. The procedure and consequence of
the projection from Gauss points to the nodes is discussed further in the
Visualization of Results Chapter. The important point here is that it is necessary to
be aware of the fact that secondary parameters are computed at Gauss integration
points.
Element shapes
The quadrilateral and triangular elements available in GeoStudio can have almost
any shape. The performance of the elements, however, deteriorates if they deviate
too far from the ideal. Quadrilateral elements provide the optimum performance
when they are square and triangular elements when they are equilateral triangles or
isosceles right triangles. The elements can deviate from these ideal shapes and still
obtain entirely acceptable performance. Square elements can be rectangles or
trapezoids and triangular elements can be any scalene acute triangle. There are,
however, limits to the distortion permissible. The following Figure 4-6 shows a
relative performance comparison of element shapes.

Page 76

SIGMA/W

Chapter 4: Meshing

Ideal

G ood

Acceptable

Poor

Unacceptable
Figure 4-6 Element shapes and relative performance

The influence of element shapes is addressed both in the discussion below and a
later discussion on structured and unstructured meshing.

4.3

Regions

GeoStudio uses the concept of regions to define the geometry of a problem and to
facilitate the discretization of the problem. The attraction of regions is that they
replicate what we intuitively do as engineers and scientists to illustrate concepts
and draw components of a system. To draw a stratigraphic section, for example,
we intuitively draw the different soil types as regions.
The utilization of regions offers all the advantages of dividing a large domain into
smaller pieces, working and analyzing the smaller pieces, and then connecting the
smaller pieces together to obtain the behavior of the whole domain, exactly like the
concept of finite elements. Generally, all physical systems have to be broken down

Page 77

Chapter 4: Meshing

SIGMA/W

into pieces to create, manage and control the whole body. A vehicle is a good
example. It is made up of numerous pieces joined together to make a useful means
of transportation.
A collection of highly adaptive individual pieces that can be joined together makes
it possible to describe and define almost any complex domain or physical system.
Such an approach is more powerful and can be applied to a wider range of
problems than any system that attempts to describe the whole domain as a single
object.
Further evidence of the advantages of the region approach is the acceptance in
industry of object-oriented technology, which has been widely applied in software
design and development.
In GeoStudio, regions can be thought of as super elements even though they do not
behave like conventional finite elements. They are the host or parent of the smaller
conventional elements. From a usability point of view it is obviously much easier
to draw and describe a few big elements or regions than to individually draw and
describe each finite element.
Region types
Regions may be simple straight-side shapes like quadrilaterals or triangles or a
non-regular, multi-sided polygon. Figure 4-7 illustrates a domain constructed using
one quadrilateral and two triangular regions. Figure 4-8 shows a single multi-sided
polygonal region defined using 10 points.
Region points
Regions have points like elements have nodes. The points make it possible to join
regions and ensure there is continuity between the regions. Points common to
different regions means the regions are joined at these points.

Page 78

SIGMA/W

Chapter 4: Meshing

Figure 4-7 Illustration of a quadrilateral region and triangular regions


10

Figure 4-8 A multi-side polygonal region

The points can be selected and moved to modify the shape and position of regions,
which provides for great flexibility in making adjustments and alterations to a
problem definition.
15

16

14

17

13

11

12

Figure 4-9 Regions of different size

Page 79

Chapter 4: Meshing

SIGMA/W

Points are also required in order to join regions of different sizes and to control the
meshing for specific purposes. Figure 4-9 shows a homogeneous soil region with a
concrete footing region. The foundation region is made up of Points 11, 12, 13, 17
and 14. The footing region is made up of Points 14, 17, 16, and 15. Points 14 and
17 are common to both regions and therefore the two regions are properly joined
and connected along this edge. In addition, Point 17 ensures that an element node
will be created and will exist at the edge of the footing, which is required for
proper meshing.
Region properties
When a region is defined, it is restricted to:

one type of material,

one type of element meshing pattern,

one order of elements; either first- or second-order, and

one integration order.

However, after a region is generated, any individual element within that region can
be altered manually using the Draw Element Property command. If the region is
regenerated intentionally or because an adjacent region required a regeneration,
then any manually changed element properties will be set back to the default
region values.

4.4

Mesh types and patterns

Structured mesh
Figure 4-10 presents what is known as a structured mesh. The elements are ordered
in a consistent pattern. This illustration has five regions. The same mesh could
have been created with two regions, but the same number of points would have
been required. All rectangular regions could be one region, but intermediate points
(4, 5 and 8) would be required to achieve the same mesh.
Unstructured mesh
The diagram in Figure 4-11 shows the same section as in Figure 4-10, but this time
with an unstructured mesh. In this case the mesh is automatically created using
Delaunay triangulation techniques. Only one region is required, but now the only

Page 80

SIGMA/W

Chapter 4: Meshing

user control is the number of divisions along edges of the region. The shape and
positions of the elements within the region is controlled by the algorithm.
Changing the number of division along any region edge can result in a completely
different mesh.
One of the great attractions of unstructured meshing is that almost any odd-shaped
region can be meshed. This meshing simplicity, however, has some numerical and
interpretation consequences as discussed in more detail in the Structured versus
Unstructured Section below. Automatic mesh generation with unstructured
elements is NOT the answer to all meshing problems!
9

10

Figure 4-10 Structured mesh


26

25

24

21

23

22

Figure 4-11 Unstructured mesh

Triangular regions
GeoStudio has a special structured pattern for triangular regions. The next figure
(Figure 4-12) shows a typical triangular region with a structured mesh. The

Page 81

Chapter 4: Meshing

SIGMA/W

elements are a mixture of squares, rectangles, trapezoids and triangles. The use of
this pattern is fairly general, but it does have some limitations and restrictions.

2
Figure 4-12 Triangular region

It is useful to think of the region as having three sides: a short side, an intermediate
length side and a long side. The algorithm attempts to sort the edges so that the
sides go from the shortest to the longest in a counter-clockwise direction. In this
example, the shortest side is 3-1, the intermediate 1-2 and the longest 2-3.
The meshing algorithm works best when the number of divisions is controlled on
the shortest and intermediate sides. To retain the even pattern shown in Figure
4-12, the number of divisions should be defined on the shortest side first and then
on the intermediate side. The number of divisions on the intermediate side can be
an even multiple of the number on the shortest side. In the above example, the
shortest side has 5 divisions and the intermediate side can have 10, or 2 times that
of the shortest side. The algorithm works best and gives the best structured mesh if
the numbers of divisions on the longest side are left undefined allowing the
algorithm to compute the appropriate number of divisions.
If a triangular region is mixed in with other more general regions, GeoStudio will
attempt to ensure mesh compatibility. Sometimes, however, it may not be possible
to adhere to the requirements for generating a structured mesh in a triangular
region and then GeoStudio will substitute an unstructured mesh.
Generally, the triangular structured pattern functions best when it is used with
other structured mesh regions, although this is not a strict requirement.
Transfinite meshing
Much of the structured meshing in GeoStudio is based on transfinite mapping
techniques. Without going into detail, this mapping technique creates a mesh that

Page 82

SIGMA/W

Chapter 4: Meshing

reflects the perimeter of the region as illustrated in Figure 4-13. This is a


quadrilateral region with one curved side. Note how the mesh reflects the curvature
of the top side of the quadrilateral.

Figure 4-13 Transfinite mapping pattern

The transfinite mapping technique provides for great flexibility in that the sides
can have almost any curved or jagged line-segmented shape. However, the system
needs what in GeoStudio are referred to as corners.
Quadrilateral regions must have four corners and triangular regions must have
three corners. The mesh in Figure 4-14 has four corners defined, even though there
are 10 points used to define the region. In this case the corners are at the natural
corners of the rectangle, but this is not always necessary as we will see a little later.
The second essential requirement for transfinite mapping is that the number of
element edges between corners must be the same on opposing sides of the region.
In this example, the number of element edges between Points 13 and 16 must be
the same as between Points 11 and 17, and the number of element edges between
Points 11 and 13 must be equal to the number of edges between Points 16 and 17.

Page 83

Chapter 4: Meshing

11

SIGMA/W

19

18

17

12
20

13

14

15

16

Figure 4-14 Region corners in structured mesh

In Figure 4-14 the number of divisions along the bottom horizontal and top curved
sides have the same number of divisions and the two vertical sides have the same
number of divisions.
Figure 4-15 further illustrates transfinite mapping. In this case the corners are
Points 1, 4, 5 and 8. Edge 1-8 is opposite edge 4-5 and must have the same number
of divisions. Edge 1-2-3-4 is opposite edge 5-6-7-8. The total number of divisions
along 1-2-3-4 must be the same as along 5-6-7-8.
GeoStudio makes a best guess at the corner locations when you draw the regions,
but the algorithm does not always identify the user-intended corner positions. One
of the features of regions is that the corner positions can be moved using the mouse
as necessary to achieve the intended mesh configuration.
1

Figure 4-15 Transfinite mesh in a u-shaped region

Page 84

SIGMA/W

Chapter 4: Meshing

Mesh with openings


Regions can have openings as shown in Figure 4-16 to simulate drains or tunnels.
The radius is a user-specified variable and the exact position of the center of the
opening can be specified. Boundary conditions can be specified along the
perimeter of the opening just like at any other node or element edge.
It is sometimes useful to construct an opening with a ring of elements around the
opening as shown in Figure 4-17. The ring of uniform elements can be especially
useful when simulating the presence of a liner. This feature is intended more for
use in a stress-deformation analysis, but it can be used in any application where it
is desirable to have a uniform element shape around the opening.
4

Figure 4-16 Mesh with an opening


4

Figure 4-17 Meshing opening with ring of elements around the opening

Page 85

Chapter 4: Meshing

SIGMA/W

Regions with openings have various parameters to control the opening and mesh
characteristics. Descriptions of all the parameters are in the software online help.
The important point here is that GeoStudio has special regions to assist with
meshing special conditions.
Surface regions
At the ground surface conditions change in response to the climate and climatic
conditions can change dramatically over short periods of time. For example, the
ground maybe highly desiccated near the surface on a hot day before a
thunderstorm. In a short period of time, the soil changes from being very dry to
being saturated. Another example may be penetration of frost from the ground
surface. To numerically deal with rapid and dramatic boundary changes it is
necessary to have fine discretization near the ground surface. GeoStudio has a
special procedure for constructing a surface mesh. Figure 4-18 illustrates a surface
mesh placed over the surface of a larger region. The surface mesh capability is also
invaluable for discretizing features such as engineered soil covers over waste
material, which may consist of several relatively thin layers of soil.
The ability to construct a surface mesh is available in VADOSE/W, SEEP/W and
TEMP/W. In SEEP/W the surface mesh is used to tell the solver that it should
track seepage face flows and infiltration events for any unit flux boundary
condition. As a result, water that does not immediately infiltrate the ground is not
considered lost from the analysis, but is allowed to pond and build up a positive
pressure head in any user-defined low points along the surface. In VADOSE/W
and TEMP/W, the surface region is use for application of climate boundary
conditions. The other GeoStudio modules cannot be used to construct a surface
mesh, but once the surface mesh has been created it will exist in all the other
modules. Consequently, if a surface mesh has been created for a particular
analysis, the surface mesh will also be part of all other analyses, since GeoStudio
uses only one mesh definition within a single data file.
Once the main soil profile has been meshed, a special Draw Surface Region
command can be used to build up a single or multi layer region along all or part of
a ground surface. Parameters such as the soil type and individual layer geometry
are defined and a quadrilateral element mesh with vertically oriented nodes is
automatically built on top of the existing ground region. The structure of the mesh
will ensure optimum numerical stability during the solution.
Quadrilateral elements are much better for modeling ground surface processes
because the primary unknown gradients are usually steeper in a direction

Page 86

SIGMA/W

Chapter 4: Meshing

perpendicular to the surface. The presence of triangular elements in thin layers


near the surface causes excessive fluctuation in the computed results relative to the
orientation of the triangular elements. Also, dealing with plant root zones in the
VADOSE/W model necessitates that element nodes in the surface region all fall on
vertical lines. Moreover, using quadrilaterals greatly reduces the number of
elements required, an important consideration when dealing with situations that
will be very computationally intensive.

Figure 4-18 Illustration of a surface region mesh

Surface mesh regions have special viewing options. Consider the two meshes
illustrated in Figure 4-19. The left diagram shows a surface region mesh without
all the cluttering details as illustrated on the right. When many thin elements are
located in a close proximity to each other, they can appear unclear when viewed
from a far away scale. By optionally turning off the surface mesh details a clearer
image of the structure of the near surface soil layers can be viewed.

Figure 4-19 Surface region mesh with details off (left) and on (right)

Page 87

Chapter 4: Meshing

SIGMA/W

Figure 4-20 is another illustration of this optional viewing concept. The left
diagram in the figure shows the detailed mesh and soil layers across the 0.75m
thick surface region and the right diagram leaves the details out, but still shows the
layer colors. A couple of additional key points can be made in regards to the
figure. Notice that bottom two elements of the left diagram are the same soil type
as the main underlying soil. This is a good mesh design strategy that being to
have the bottom most layer of the surface mesh be made of the same soil as the
existing ground. Consider if the bottom layer of the surface soil was VERY
different from the underlying soil. If a fine spaced mesh were placed directly on
top of the different underlying soil then the numerical integration of material
properties at the common mesh node between the two soils would be less accurate
because of the influence of the large element area from the material below the
common nodal point. By having the bottom layer of the surface region the same as
the underlying soil, the element shapes are very similar in size and aspect at the
common nodal point between the two very contrasting soils.
The second point to node from Figure 4-20 is that in the right diagram the nodes
that are located at the interface between two soils are still viewable even though
the main mesh details are not. This is intentional so that you can easily see and
graph data at nodes that are used for automatic tracking of interlayer fluxes in the
VADOSE/W model.

Figure 4-20 Close up of surface details on and off


(note inter-layer nodes still visible in figure on right)

Boundary flux modeling with rainfall infiltration, runoff, snow melt etc. can be
very numerically demanding from a convergence perspective. Potential problems
can be made worse if the shape of the surface mesh is not realistic. Consider the
two meshes illustrated in Figure 4-21 and Figure 4-22. In the first figure, the
ground profile has rounded corners which are much more natural and much more
numerically friendly. In the second figure, changes in slope angle are represented
by a sharp break. This sharp break is not only un-natural, but the shape of the
individual elements right at the transition points creates numerical problems if

Page 88

SIGMA/W

Chapter 4: Meshing

there are large changes in boundary condition type at different nodes within the
same element. This would be the case when the corner node at the bottom of the
slope becomes a seepage face point while the next node up slope is still an
infiltration node. Basically, it is better to build the mesh to look somewhat natural.
g

Figure 4-21 GeoStudio region mesh showing rounded surface slope breaks

Figure 4-22 Version 5 mesh showing angular surface slope breaks

In order to create a surface mesh with more rounded features it is necessary to


build the underlying soil mesh with the same rounded profile. This is easily
accomplished in GeoStudio by adding additional region points near a slope break
such that the region points can be moved slightly to create a rounded profile. This
is the case in Figure 4-23 below where three region points are used at both the toe

Page 89

Chapter 4: Meshing

SIGMA/W

and crest of the slope. Also notice that three region points are used on the bottom
of the mesh beneath the toe and crest location. This is a useful tip to remember.
When you want to have more control over the trans-finite element mesh you
should add region points on opposite sides of the mesh from where you need the
detail. As a final note, adding region points can be done at any time even after
the surface mesh is created. When the region beneath a surface mesh is changed,
the surface mesh above it will be automatically regenerated to ensure mesh
compatibility with the region below.

Figure 4-23 Region mesh with region corner points viewed and surface
details not viewed

Joining regions
Compatibility must be maintained between regions to ensure the regions are
connected. Regions must be joined at the region Points and Points must be
common to adjoining regions for the regions to be properly connected. GeoStudio
has a number of features to assist in achieving region compatibility.
The following are some of the main characteristics:

If the cross-hair symbol moves close to an existing Point, the symbol


will snap to the existing point.

A new Point will be created if the cursor is on the perimeter of an


existing region. The new Point will then be common to the new region
and to the existing region.

Page 90

SIGMA/W

Chapter 4: Meshing

Points in between selected points are automatically selected along an


existing region edge unless the Ctrl key is held down.

Consider the diagram in Figure 4-24. Region 1 is drawn first and Region 2 can be
drawn by clicking on Points 7, 3, 9 and 8. Points 4, 5 and 6 are automatically
added to Region 2.
9

2
7

3
5
6

4
1

Figure 4-24 Regions joined along jagged line


9

7
5

6
4

1
1

Figure 4-25 Adjoining regions with an open space

Sometimes it may be desirable to create an open area in a mesh and then it is


necessary to hold down the Ctrl key when going from Point 7 to 3 or 3 to 7. Doing
this results in a mesh as in Figure 4-25. In this case the Ctrl key was held down
after clicking on Point 7, but before clicking on Point 3. Once again, other details
on joining regions are presented in the on-line help.

Page 91

Chapter 4: Meshing

4.5

SIGMA/W

Structured versus unstructured meshing

There is a large volume of published literature on unstructured triangular meshing


and structured quad meshing. Both approaches have advantages and disadvantages.
The purpose of this document is not to go into all the technical details of one
method versus the other. The purpose here is to briefly comment on the two
different approaches from a practical applications point of view in GeoStudio.
Unstructured meshing in some ways is easier and faster than structured meshing.
In unstructured meshing, there is no need to give any thought as to what the mesh
should look like, since the positioning and shape of the elements is completely
controlled by the meshing algorithm. The mesh just appears. The only control the
user has is to specify the number of element divisions along a region edge.
However, structured meshing requires a little more forethought and guidance from
the user. Figure 4-26 shows the same section with an unstructured and a structured
mesh. The unstructured mesh can be created very quickly by simply drawing one
region. To create the nice structured mesh shown on the right in the figure, it is
necessary to do a little planning. The best approach is to first draw the triangular
region and then draw the other rectangular regions. Being able to quickly create a
mesh without forward-thinking as to the mesh layout is one of the attractive
features of unstructured meshing, but this procedural ease has some numerical and
presentation consequences that will be discussed further.

Figure 4-26 Unstructured and structured meshes

Generally, quadrilateral elements offer better behavior and performance than


triangular elements. However, this depends to some degree on the type of problem
being solved and the order of the elements relative to the order of the underlying
partial differential equation. For example, the performance of first-order constant
gradient triangular elements is rather poor in a stress-deformation analysis where

Page 92

SIGMA/W

Chapter 4: Meshing

the primary unknowns are displacement vectors. In a seepage analysis, however,


where the primary unknowns are scalar values, the same triangular elements can
generally offer quite acceptable performance. This may not be the case where there
is a sharp contrast in material properties in adjoining layered soils (e.g. in
computation of seepage across a capillary break).
Figure 4-27 shows the results of a seepage analysis of leakage from a clay-lined
pond. Note the direction of the flow vectors. In the structured mesh they follow a
nice uniform direction. In the unstructured mesh, direction of the flow vectors is
influenced by the geometric orientation of the triangular elements. This is typical
of triangular elements. The computed primary unknowns at the nodes (head,
displacement, etc) are often not significantly influenced by the orientation of the
triangular elements, but the secondary quantities inside the elements can be
affected by the geometric orientation of these elements.

Figure 4-27 Comparison of flow vectors in an un-structured and a structured


mesh

The systematic layout of a structured mesh offers advantages when it comes to


visualizing and interpreting results, especially if you want to look at a particular
variable along vertical profiles through the mesh. In a structured mesh it is easy to
pick up all the nodes on a vertical profile as illustrated in Figure 4-28. This is more
complicated in an unstructured mesh. It is not possible to pick up a group of nodes

Page 93

Chapter 4: Meshing

SIGMA/W

along a vertical line. In this case it is necessary to find all the elements intersected
by a cut-line and then to project the results onto the cut-line from the nearest
nodes. This can be done, but this means that interpolated values are displayed, as
opposed to looking at the exact values computed at the nodes.

Vertical profiles

Figure 4-28 Vertical profiles through a mesh

It is easier to spot check the computed results against hand-calculations when using
quadrilateral elements in a structured mesh than it is with triangular elements at
some random orientation. Take for example the case of determining in-situ
stresses. In a structured mesh it is fairly straightforward to spot check the stress in
an element. The vertical stress is directly related to the vertical distance from the
surface and since the element is aligned with the x and y directions, the hand
calculation is trivial. It is not as straightforward for triangular elements, which are
all at different orientations.
Unstructured meshing techniques are clearly better for highly irregular geometric
shapes. Care is required in geotechnical numerical modeling to not make the
geometry unnecessarily complicated as discussed elsewhere in this chapter and in
the previous chapters. Just because it is easy to mesh complicated shapes is not
sufficient reason to make the geometry unnecessarily complicated.
Another advantageous feature of unstructured meshing is that the method has no
restrictions on the number of element divisions along region edges. In transfinite
mapping, as discussed above, it is necessary to define corners and opposing sides
between the corners must have the same number of divisions. Unstructured

Page 94

SIGMA/W

Chapter 4: Meshing

meshing has no such restrictions. This makes it easy to transition from a fine mesh
to a course mesh.
Unstructured meshing also opens the door to adaptive meshing in the future. Mesh
refinement and adaptive meshing is possible with a structured mesh pattern, but it
is more restrictive than with triangular unstructured meshing. Adaptive meshing
during a solve process does not necessarily address all numerical problems related
to finite element analysis. For example, adaptive meshing can create many small
elements in areas of high gradients, but solving the equations in these areas
requires smaller time steps in a transient analysis. This is fine in principle, except
that the time steps are common to the entire geometry and can therefore become
too small for other regions that have larger element sizes.
Generally, some extra effort and care up front in creating a mesh pays dividends
when it comes to solving, interpreting, visualizing, discussing and presenting the
results. Often there is a trade off in effort between defining the problem and
dealing with the results. Extra effort is either required at the start or at the end of
the analysis. Our experience is that it is better to spend the extra effort defining the
problem and in the mesh creation, than struggle to interpret and justify results at
the end of the analysis.
The ability to construct either structured or unstructured meshes in GeoStudio
offers the best of both approaches. Users can select and use whatever they wish
based on personal preference or as dictated by problem requirements.

4.6

Meshing for transient analyses

Modeling transient processes requires a procedure to march forward in time


increments. The time increments are referred to in GeoStudio as time steps.
Selecting and controlling the time step sequence is a topic in itself and will be dealt
with later. Obtaining acceptable transient solutions is not only influenced by the
time steps, but also by the element size. In a contaminant transport advectiondispersion analysis (CTRAN/W) it is necessary to have a time step sufficiently
large to allow an imaginary contaminant particle to move a significant distance
relative to the element size, while at the same time not have the time step size be so
large as to allow the particle to jump across several elements. The particle should,
so to speak, make at least one stop in each element. In CTRAN/W this is
controlled by the Peclet and Courant criteria.

Page 95

Chapter 4: Meshing

SIGMA/W

In a simulation of consolidation, the time step size for the first time step needs to
be sufficiently large so that the element next to the drainage face consolidates by at
least 50 percent. Achieving this is related to the element size; the larger the
element the greater the required initial time step. If the time step size is too small,
the computed pore-water pressures may be unrealistic.
The important point in this section on meshing is to realize that meshing, more
particularly element sizes, comes into play in a transient analysis. Rules and
guidelines for selecting appropriate time stepping are discussed elsewhere with
reference to particular types of analysis.

4.7

Interface elements

In a finite element formulation, generally all the elements are connected at the
nodes and remain connected at the nodes. Sometimes it is desirable, however, to
allow elements to move relative to each other, such as along a joint between two
rock blocks or along the contact surface between soil and structure. Another
example might be the potential slippage that may exist between a geo-synthetic
and surrounding soil. In a seepage analysis, a wick drain will open up the flow
capacity in the axial direction of the drain. Flow can in essence follow the contact
between elements. In a thermal analysis, an interface element could be used to
simulate a thin layer of insulation for the case where actually drawing the thin
layer is hard to do because of the thickness relative to the scale of the entire model.
Features and behavior like this can be simulated with what is known as interface or
contact elements.
In the current version of GeoStudio, only SIGMA/W has a form of an interface
element which can be used to simulate the slippage between neighboring elements.
They are called slip elements. The slip elements are fairly primitive and in need
of enhancement.
A more general interface formulation with a wider range of applications has been
formulated for GeoStudio and will be available in a future release.
In GeoStudio it is important to recognize that if the interface element is present in
one analysis, it will be present in all the other products within a particular data file.

Page 96

SIGMA/W

4.8

Chapter 4: Meshing

Infinite elements

There are many problems in geotechnical engineering where there is no distinct


boundary within a reasonable distance from the main area of interest. An example
might be modeling a dam embankment on top of a flat topography that extends
laterally over a significant distance. In a numerical model it is necessary to define a
boundary at some practical distance from the dam. Ideally, the boundary should be
far enough away from the main area of interest so as to not significantly influence
the results. One solution is to make a mesh that extends over a large distance, but
this can lead to a mesh that is awkward to work with. Another solution is to use
what are called infinite elements. These types of elements make it possible to
greatly extend the position at which the boundary conditions are effective without
actually extending the mesh. The conditions far from the main area of interest are
referred to as far field boundary conditions.
The details of the infinite element formulation are presented in the appendix
chapter on interpolating functions. Only a conceptual overview and some
guidelines for using these elements are presented here.
The infinite elements are formed on the basis of a decay shape function. Three
points are required to describe this decay function. The formulation in GeoStudio
uses some of the element nodes together with what is known as the pole point to
create the three points. Figure 4-29 illustrates the concept.
X2

X1

Pole

Far field nodes

Figure 4-29 Pole position relative to an infinite element

When infinite elements are used, the x-coordinates of Nodes 6 and 8 are modified
so that the distance between Node 1 and 8 is the same as between Node 1 and the
pole. That is, the x-coordinates of Nodes 6 and 8 are modified so that X2 in Figure
4-29 is equal to X1. Next the decay function is created using the x-coordinate of
the pole, the x-coordinate of Node 1 (or 2 or 5) and the modified x-coordinate of

Page 97

Chapter 4: Meshing

SIGMA/W

Node 8 (or 6). Nodes 3, 4 and 7 are considered to be at infinity and are not used in
creating the shape functions.
Infinite elements need to be 8-noded quadrilaterals. The secondary nodes (6 and 8)
are required to form the decay function. Triangular elements cannot be used for
this purpose.
Infinite elements are a convenient way of extending the far field of a problem, but
they need to be used with some caution. The infinite elements work best if the
specified boundary conditions at the far field are the primary unknowns in the
analysis; that is, head, displacement, temperature, and so forth. Specifying
gradients (for example, unit rates of flow or stress) along the element edge
representing the far field will work, but the effect is less certain than specifying
primary field variables.
Boundary conditions should not under any conditions be specified on the element
edges that are parallel to the infinity direction. In the element in Figure 4-29,
boundary conditions should not be applied along edge 1-8-4 and edge 2-6-3. You
cannot apply a boundary condition that is a function of edge length when the edge
length is considered to be infinite.
The distance between the pole and the inner edge of the infinite element obviously
influences the coefficients of the decay function and the modification of the
secondary node x-coordinates. The pole position consequently does have some
influence on the behavior of the elements. Fortunately, the results are not all that
sensitive to the pole position. Generally, the objective is to make the distance
between the pole and the inner edge of the element as large as conveniently
possible, but not beyond the extents of the overall mesh.
We recommend that you always complete an analysis without infinite elements
first. Then after you have obtained an acceptable solution, save the problem to a
new file and then add the infinite elements to refine your analysis. Having
solutions with and without infinite elements is a useful way of understanding the
effect of extending the far field boundary and understanding the computed
response of these elements.
Infinite elements in SIGMA/W
Infinite elements can be drawn in both the x-direction and y-direction in
SIGMA/W so that the mesh will be compatible with other GeoStudio models.
However, caution should be used when specifying infinite elements in the y-

Page 98

SIGMA/W

Chapter 4: Meshing

direction because they do not make sense in a stress-deformation analysis. We


know that stresses increase with depth, but it is meaningless to try and have the
model compute what they are infinitely deep in the ground. Also, we cannot
compute displacement over an infinite element in the y-direction. X-infinity
elements will work in SIGMA/W because the increase in stress with depth through
the element is not in the direction of infinity within that element.
We recommend that you always complete an analysis without infinite elements first.
Having solutions with and without infinite elements is a useful way of understanding the
effect of extending the far field boundary and understanding the computed response of
these elements.

4.9

General guidelines for meshing

Meshing, like numerical modeling, is an acquired skill. It takes practice and


experience to create the ideal mesh. Experience leads to an understanding as to
how the mesh is related to the solution and visa versa. It is when you can anticipate
an approximation of the solution that you will be more proficient at meshing.
The attraction of the GeoStudio system is that a mesh can quickly be created with
relative ease and later modified with relative ease. This makes it convenient to try
various configurations and observe how the meshing influences the results.
An appropriate finite element mesh is problem-dependent and, consequently, there
are no hard and fast rules for how to create a mesh. In addition, the type of mesh
created for a particular problem will depend on the experience and creativity of the
user. However, there are some broad guidelines that are useful to follow. They are
as follows:

Use as few elements as possible at the start of an analysis. Seldom is it


necessary to use more than 1000 elements to verify concepts and get a
first approximate solution.

All elements should be visible to the naked eye when the mesh is
printed at a zoom factor of 100 % and when the horizontal and vertical
scales are the same. The exception to this guideline is a surface mesh.

The mesh should be designed to answer a specific question, and it


should do not include features that do not significantly influence the
system behavior.

Page 99

Chapter 4: Meshing

SIGMA/W

The mesh should represent a simplified abstraction of the actual


complex geometric field configuration.

Higher order elements in SIGMA/W


When an element has secondary nodes at the mid-points between the corner nodes,
the element is known as a "higher order" element, since the equations describing
the deformation within the element are of a higher order than when there are no
secondary nodes.
Higher order elements exhibit better behavior than the ordinary 4-noded
quadrilateral and 3-noded triangular elements. However, the higher order elements
also greatly increase the processing time and memory requirements due to the
additional nodes in the elements. Considerable computing efficiency can be
achieved by being selective in choosing element types. Always selecting higher
order elements can lead to a large amount of unnecessary computations.
As a general rule, ordinary elements are adequate for the linear-elastic model.
While the ordinary elements result in a poor stress distribution within a single
element, the stress distribution is reasonable when averaged to the nodes for
contouring. Also, for a linear-elastic model, the material property is not a function
of the computed stress; consequently, the poor stress distribution has little effect on
the computed deformations. If linear-elastic deformation is the primary objective
of the analysis, then ordinary 4-noded or 3-noded elements are usually adequate.
For the nonlinear constitutive soil models except the elastic-plastic model, the
material properties are a function of the computed stresses. Therefore, a reasonable
stress distribution within the element is essential, and higher order elements should
be used. This is particularly true for elements which are subject to bending, such as
elements located at the edge of a footing or in a retaining wall.
The elastic-plastic model is a special case. Recent studies have revealed that first
order triangular elements with constant gradients give the most reliable results.
Alternately, 4-noded quadrilateral elements can be used, however, the solver will
compute the gradient at the mid-point of the element, assume it is constant
throughout the element, and assign it to all Gauss points within the element. If you
set up a quadrilateral element with secondary nodes, then the solver will not make
any special consideration and convergence may be difficult or impossible to attain.
Higher order elements are not better suited to solving the elastic-plastic behavior.

Page 100

SIGMA/W

Chapter 4: Meshing

In summary, higher order elements are essential for the all nonlinear soil models
(except elastic-plastic), particularly in regions where the stress and strain changes
are high. In regions where the stress and strain changes are small, or the material is
assigned a linear-elastic model, the ordinary 4-noded and 3-noded elements may be
adequate.
Number of elements
Based on many years of responding to GEO-SLOPE user support questions, most
users start with a mesh that is too complex, containing too many elements for the
objective of the analysis. The thinking when users first start doing finite element
analyses seems to be the more elements, the better; that a large number of elements
will somehow improve the accuracy of the solution. This is not necessarily true. If
the mesh is too large, the time required to obtain a solution can become
unattainable. Sometimes it also becomes very difficult to interpret the results,
particularly if the solutions appear to be unreasonable. The effort required to
determine the reason for an unreasonable solution increases dramatically with
mesh size.
We highly recommend that you try and create a mesh with less than 1000
elements, particularly at the start of an analysis. Our experience is that most
geotechnical problems can be modeled with 1000 elements or less. Obviously there
are exceptions, but this number is a good goal to strive for. Later, once you have a
good first understanding of the critical mechanisms in your problem, you can
increase the mesh density to refine the analysis.
Effect of drawing scale
Another good guideline is that all elements should be visible to the naked eye
when the mesh is printed or viewed at a 100% zoom factor. Groups of elements
that appear as a solid or nearly solid black smudge on the drawing are too small.
This means a suitable element size is related to the drawing scale. A drawing at a
scale of 1:100 can have much smaller elements than a drawing at a scale of 1:2000.
In other words, if it is necessary to zoom in on an area of the drawing to
distinguish the elements, the elements may be unnecessarily small.
All elements should be readily distinguishable when a drawing is viewed when the
vertical scale is equal to the horizontal scale. It is possible to draw a nice looking
mesh at a vertical exaggerated scale, but when viewed at a true vertical scale the
mesh appears as a wide black line. This effect is illustrated in Figure 4-30. The top
part of the figure shows a nice mesh at 10V:100H, a 10 times vertical

Page 101

Chapter 4: Meshing

SIGMA/W

exaggeration. The same mesh at a scale of 100V:100H appears at the bottom of


Figure 4-30. At an exaggerated scale the elements appear suitable, but at a true
scale they are not appropriate.
It is important to remember that the main processor which solves the finite element
equations sees the elements only at the true scale. The vertical exaggeration is used
only in DEFINE and CONTOUR for presentation purposes.
A good rule to follow is to always view the mesh at a true scale before solving the
problem to check that the mesh is reasonable for the purpose of the analysis.

Figure 4-30 Mesh at an exaggerated scale (upper) and at a true scale (lower)

Mesh purpose
The How To modeling chapter notes that in good numerical modeling practice it
is important to form a mental imagine of what the solution may possibly look like
and to clearly define the purpose of the model before trying to create a model.
Meshing is closely tied to this guideline. The mesh should be designed to answer
specific questions. Trying to include all possible details in a mesh makes meshing
unnecessarily time consuming and can sometimes make it difficult to interpret the
results.

Page 102

SIGMA/W

Chapter 4: Meshing

Let us assume that we are interested in estimating the seepage though the clay core
of a zoned dam with rock shells. Figure 4-31 shows a typical case. The rock shells
are considered to be many orders of magnitude more permeable than the core. In
addition, the granular drain filter layers between the clay and the rock are clean
and can easily handle any seepage though the core without impeding the drainage.
In other words, the granular filter layers and rock shells make no contribution to
dissipating the hydraulic head on the upstream side of the core. If this
consideration is true, then there is nothing to be gained by including the highly
permeable materials in the analysis. A mesh such as in Figure 4-31 is adequate to
analyze the seepage though the core.
1

Figure 4-31 Modeling core of zoned dam

Figure 4-32 shows the total head contours (equipotential lines) in the core. From
this the seepage quantities through the core can be computed.

Figure 4-32 Equipotential lines in core of dam

Sometimes a mesh may be required to include the shells in the analysis for other
reasons, such as a stress-deformation analysis. In such a case, the mesh can exist,
but does not need to be included in the analysis. This is accomplished using null
elements as shown in Figure 4-33. Elements in GeoStudio can be null (not active)
by leaving a key material property undefined. In SEEP/W the elements are null if
there is no specified conductivity function for the material. In the example in
Figure 4-33 the rock shells have no conductivity function assigned to the material.

Page 103

Chapter 4: Meshing

SIGMA/W

Figure 4-33 Mesh with null elements in shells of dam

One of the attractions inherent to numerical modeling is that the geometry and
finite element mesh do not necessarily have to conform strictly to the physical
conditions. As in Figure 4-31, the core can be analyzed in isolation. This would not
be possible in physical modeling. The dam with a toe drain in Figure 4-34 is
another good example. The toe drain does not have to be included in the numerical
analysis. This, of course, would not be possible in a physical model.
8

10

11

Figure 4-34 Dam with a toe drain

Simplified geometry
A numerical model needs to be a simplified abstraction of the actual field
conditions. This is particularly true when it comes to the geometry. Including all
surface irregularities is unnecessary in most situations. Geometric irregularities can
cause numerical irregularities in the results, which distract from the main overall
solution. The main message can be lost in the numerical noise.
Simplifying the geometry as much as possible is particularly important at the start
of an analysis. Later, once the main processes involved are clear, the geometry can
be altered to determine if the geometric details are important to the main
conclusions.
The situation is different if the main objective of the analysis is to study the effects
of surface irregularities. Then the irregularities of course need to be included. So,

Page 104

SIGMA/W

Chapter 4: Meshing

once again, the degree of geometric complexity depends on the objectives of the
analysis.
Also, the level of geometric detail that needs to be included in the problem must be
evaluated in light of the certainty with which other factors such as the boundary
conditions and material properties are known. There is little to be gained by
defining a very detailed geometry if the material properties are just a rough
estimate. A simplified geometry is more than adequate if the material properties
are rough estimates. There needs to be a balance in complexity between all the
aspects of a finite element analysis, including the geometry.
Over-complicating the geometry is a tendency when users first get into numerical
modeling. Then, as modelers gain more experience, they tend to use more simple
geometries. This obviously comes from understanding how the mesh can influence
the results and what level of complexity is required. The situation should be the
reverse. It is the modelers with limited experience who should use simplified
geometries.
The main message to remember when starting to model is to keep the problem as
simple as possible until the main engineering issues are well understood.

Page 105

Chapter 4: Meshing

SIGMA/W

Page 106

SIGMA/W

Chapter 5: Material Properties

Material Properties

SIGMA/W includes eight different soil constitutive models. To an inexperienced


user, it may be difficult to decide which model to select for a particular application.
While the ultimate choice is your responsibility, there are some points that you
should consider: primarily, the material stiffness, tolerable displacement, and
stability.
Consider the case of a heavy industrial structure founded on highly overconsolidated soil. Settlement is often the main design criterion, and the settlement
must be fairly small. To ensure that the settlement is small, the applied loads are
kept low relative to the ultimate capacity of the soil. The load-displacement
response therefore is likely linear elastic along the lower initial portion of the
stress-strain curve. A simple linear elastic analysis is adequate, and little would be
gained by using a nonlinear analysis.
In the case of the construction of an embankment, considerable yielding and
deformation can perhaps be tolerated without affecting the serviceability of the
structure. A nonlinear analysis is required to obtain a realistic estimate of the
potential displacements; thus, a simple linear-elastic analysis could considerably
underestimate the displacements.
Placing fill for an embankment on soft soil can generate excess pore pressures to
the point where the stability is affected. In such a case, and especially if you are
interested in how pore-water pressure is being dissipated, it may be necessary to
use one of the more sophisticated effective stress models, such as the Cam-clay
model, together with a coupled or uncoupled consolidation analysis.
It is important to remember that each soil model is not necessarily applicable to all
soil conditions. For example, the Cam-clay and Modified Cam-clay models are
best suited for use with slightly over-consolidated soils, not heavily overconsolidated soils. A linear-elastic model can give more realistic results for heavily
over-consolidated soils than the Cam-clay models.
In summary, some thought must be given to selecting a soil model that is
consistent with the soil conditions and the objective of the analysis.

Page 107

Chapter 5: Material Properties

5.1

SIGMA/W

Constitutive models overview

SIGMA/W is formulated for several elastic and elasto-plastic constitutive soil


models. All models may be applied to two-dimensional plane strain and
axisymmetric problems. The following models are supported:

Linear-elastic

Anisotropic Linear-elastic

Page 108

SIGMA/W

Chapter 5: Material Properties

Nonlinear Elastic (Hyperbolic)

Elastic-plastic (Mohr-Coulomb or Tresca)

Strain-Softening

Page 109

Chapter 5: Material Properties

SIGMA/W

Cam-clay (Critical State)

Modified Cam-clay (Critical State)

Each of these is discussed in greater detail below.

5.2

Linear-elastic model

The simplest SIGMA/W soil model is the linear elastic model for which stresses
are directly proportional to the strains. The proportionality constants are Young's
Modulus, E, and Poisson's Ratio, . The stresses and strain are related by the
equation:

Page 110

SIGMA/W

Chapter 5: Material Properties

x

E
y
=
z (1 + )(1 2 )
xy

0
0
0

0
x
0
y
0
z
1 2
xy
2

For a two-dimensional plane strain analysis, z is zero.


It is noteworthy that when approaches 0.5, the term (1 2v ) / 2 approaches
zero and the term (1 v ) approaches . This means that the stresses and strains
are directly related by a constant, which is representative of pure volumetric strain.
Furthermore, the term E / (1 + v )(1 2v ) tends towards infinity as (1 2v )
approaches zero. Physically, this means that the volumetric strain tends towards
zero as Poissons ratio, , approaches 0.5.

For computational purposes, can never be 0.5. Even values greater than 0.49 can
cause numerical problems. Consequently, SIGMA/W limits the maximum value
for Poissons ratio, , to 0.49.
Table 5-1 Linear-elastic properties
Edit Box Label
E Modulus

Property
Young's Modulus

Poissons Ratio

constant value

Cohesion

constant value

Friction angle

degrees

Data for the linear-elastic model includes the cohesion and friction angle. This
information is not used in the solution, but it is used in the Contour program to
help illustrate regions of the soil where the computed stresses have exceeded the
yield strength. Recall that for a linear-elastic model, there is no yield value defined
and computed strains may be very unrealistic. Using cohesion and friction angle
along with the Mohr-Coulomb failure criterion will enable the computed shear
stress to be compared visually with the theoretical yield stresses.

Page 111

Chapter 5: Material Properties

5.3

SIGMA/W

Anisotropic elastic model

Natural ground deposits are often stratified and inclined. Therefore, it is desirable
to consider the possibility of having different stiffness values in two orthogonal
directions. Consider the case illustrated in Figure 5-1. The soil strata are crossanisotropic in the local orthogonal directions x and y with x, making an angle
with the global x-axis. The sign convention for can be described as follows:
when counter-clockwise from the x-axis, is positive. The anisotropic elastic
parameters in the local directions are defined by the following parameters:

in the x-direction: Ex' , x'

in the y-direction: E y'

coupling between x and y: Gxy ' , yx'

The parameter vxy ' is the Poissons Ratio of horizontal (x) strain to vertical (y)
strain caused by a stress change in the y-direction. These parameters must satisfy
the following restrictions (Pickering, 1970):

Ex , E y , and Gxy > 0,


1 < x < 1,

and,

1 x > 2 ( Ex E y ) yx
The constitutive matrix [C] for anisotropic conditions is symmetric and can be
defined as (Britto and Gunn (1987):

c11 c12
c
c
[C ] = c0 c21 c22
31
32

0
0

c13
c23
c33
0

0
0
0

c44

where:

Page 112

SIGMA/W

c0 =

Chapter 5: Material Properties

Ey

(1 + x ) (1 x 2 ( Ex

E y ) yx2

c11 = ( Ex E y ) 1 ( Ex E y ) 2yx

c12 = ( Ex E y ) yx (1 + x )

c13 = ( Ex E y ) x + ( Ex E y ) 2yx

c21 = c12

c22 = (1 2x )
c23 = c12
c31 = c13
c32 = c23
c33 = c11
c44 = Gxy c0

Figure 5-1 Anisotropic material properties

Page 113

Chapter 5: Material Properties

SIGMA/W

When the local and global axes do not coincide (when is non-zero), it is
necessary to transform the [C] matrix from the local x-y coordinate system to the
global x-y coordinate system. The transformation of [C] to [C] is given by the
following equation (Zienkiewicz and Taylor, 1989):

[C ] = [T ][C ][T ]

where:

cos 2

sin 2

=
T
[ ]
0

sin cos

sin 2
cos
2

0
sin cos

2sin cos

0 2sin cos

1
0

0 cos 2 sin 2
0

Table 5-2 Anisotropic elastic properties


Edit Box Label

Property

E Modulus (1)

Youngs Modulus in x-direction, (Ex)

P. Ratio (1)

Poissons ratio in x-direction, ( vx ' )

E Modulus (2)

Youngs modulus in y-direction, (Ey)

P. Ratio (2)

Poissons ratio, ratio of x-strain to y-strain caused by y-stress, ( v y ' )

G Modulus (2)

Shear modulus, ( Gxy ' )

Angle

Inclination of the strata in degrees from x-axis, ( )

5.4

Nonlinear elastic (Hyperbolic) model

The stress-strain behavior of soil becomes nonlinear, particularly as failure


conditions are approached (see Figure 5-2). A procedure for modeling this soil
behavior by varying the soil modulus is addressed in the following sections.

Page 114

SIGMA/W

Chapter 5: Material Properties

SIGMA/W uses the formulation presented by Duncan and Chang (1970) to


compute the soil modulus. In this formulation, the stress-strain curve is hyperbolic
and the soil modulus is a function of the confining stress and the shear stress that a
soil is experiencing. This nonlinear material model is attractive since it requires
soil properties can be obtained quite readily from triaxial tests or the literature (for
example, Duncan et al. , 1980).
Duncan and Changs non-linear stress-strain curve is a hyperbola in the shear
stress, (1 - 3), versus axial strain space. Depending on the stress state and stress
path, three soil moduli are required; namely, the initial modulus, Ei the tangential
modulus, Et, and the unloading-reloading modulus, Eur (see the following figure).

Figure 5-2 Non-Linear stress-strain behavior

Page 115

Chapter 5: Material Properties

SIGMA/W

Initial modulus
When a soil is subjected to zero shear stress (i.e. when ( 1 3 ) = 0), its stressstrain behavior is modeled using the initial modulus, Ei . This initial tangent
modulus is controlled by the confining stress, 3 and is calculated as follows:
Equation 5-1


Ei = K L Pa 3
Pa

where:

Ei

initial tangent modulus as a function of confining stress,

KL

loading modulus number

pa

atmospheric pressure (used as a normalizing parameter)

confining stress

exponent for defining the influence of the confining pressure on the


initial modulus

When the exponent n is equal to 1.0, the initial tangent modulus, Ei , is directly
proportional to the confining stress. When n is equal to 0, Ei is independent of the
confining stress.
Two different methods are used in SIGMA/W to calculate the initial soil modulus.
When K L is equal to zero (i.e., not defined by the user), the initial tangent
modulus, Ei , is set to the user-specified value for E and remains constant
throughout an analysis. If K L is nonzero (i.e., defined by the user), Ei is computed
using the equation above.
When the confining stress is zero or negative (i.e., the soil is in a state of tension),
it is possible for the initial modulus, Ei , to become zero or negative. In order to

Page 116

SIGMA/W

Chapter 5: Material Properties

avoid this problem when computing Ei , SIGMA/W imposes a lower limit of

( 0.01 pa ) on the confining stress, 3 .

Tangent modulus
A soil is said to be following a loading path when it is subjected to a shear stress
higher than it has previously experienced, for example, from point O to point A in
Figure 5-2. Along this loading path, its constitutive behavior is governed by the
tangent modulus, Et . This tangent modulus is defined in the Duncan and Chang
model as a function of soil properties, triaxial deviatoric stress, ( 1 3 ) , and

confining stress, 3 using the following equation:

R ( 1 3 )(1 sin )
Et = 1 f
Ei
c

+
2
cos
2
sin
(
)
3

where:

Ei

initial tangent modulus,

Et

tangent modulus,

friction angle of soil,

cohesive strength of soil,

Rf

ratio between the asymptote to the hyperbolic curve and the maximum
shear strength. This value is usually between 0. 75 and 1. 0,

major principal stress, and

minor principal stress.

Et can be limited to a minimum value. The default is equal to Pa . This userspecified lower limit is called Emin in the material properties dialog box. If Et
becomes too small it causes convergence difficulties.

Page 117

Chapter 5: Material Properties

SIGMA/W

Unloading-reloading modulus
When a soil is unloaded from a higher shear stress state, (for example, going from
point B to point C in Figure 5-2), the non-linear model uses the unload-reloading
modulus, Eur .
The unload-reloading modulus, Eur , is computed in a manner similar to the
computation of the initial modulus, Ei except that the unloading-reloading
modulus number, K ur replaces K L in Equation 5-1. Thus, the unloading-reloading
modulus is calculated as:


Eur = K ur Pa 3
Pa

Unlike the tangent modulus, this unloading-reloading modulus is unaffected by the


shear stress level.
If the unloading-reloading modulus number, K ur , is not defined by a user, it is
assigned the value of the loading modulus number, K L , and the soil will unload or
reload along the hyperbolic curve.
Poisson's ratio
The Poisson's ratio of the non-linear elastic model can either be specified as a
constant, which is independent of stress state, or it can be computed from the soil
bulk modulus, which depends on the confining stress. For the latter case, the bulk
modulus is given by:
Equation 5-2


Bm = K b Pa 3
Pa

where:

Bm

bulk modulus,

Kb

modulus number,

Page 118

SIGMA/W

Chapter 5: Material Properties

Pa

atmospheric pressure, and

bulk modulus exponent.

The relationship of the bulk modulus to the Poisson's ratio can be defined as
follows from theory of elasticity.

Bm =

E
3 (1 2 )

Letting the Poissons ratio, v, be equal to zero in the above equation gives a bulk
modulus, Bm , equal to E/3. When letting the Poissons ratio, v, be equal to 0.49,

Bm has a value of 17E. SIGMA/W imposes both upper and lower limits on the
computed value of Poissons ratio. The upper limit is 0.49. The lower limit is
either the user-defined constant value or 0.01.
If K b is not defined (i. e. zero), SIGMA/W assigns the user input constant value to
the Poissons ratio. Otherwise it is calculated using Equation 5-2.
Negative pore-water pressure
The cohesive component of a soil can consist of two components; namely,
effective cohesion and cohesion due to matric suction (Fredlund and Rahardjo,
1993):

c = c + (ua uw ) tan b
where:
c

(ua - uw) =

total cohesion intercept on the Mohr-Coulomb failure envelope,


matric suction (i. e. , negative pore-water pressure referenced to the poreair pressure), and

an angle relating the increase in shear strength of a soil to an increase in


matric suction.

Page 119

Chapter 5: Material Properties

SIGMA/W

SIGMA/W allows a b value to be specified. This value is then used to compute


the total cohesion, c, for the soil. The soil strength is consequently a function of
soil suction.
Yield zones
From a theoretical standpoint, yield conditions cannot be defined when using a
non-linear elastic model. In order to show high shear stress zones in a non-linear
elastic material, SIGMA/W identifies such zones as "yielded" when the following
criterion is satisfied:
Equation 5-3

1 3 1 + 3

sin R f c cos

In Duncan and Changs formulation of the hyperbolic model, Rf, the failure ratio,
is used in the following manner:

( 1 3 ) f

= R f ( 1 3 )ult

The ultimate strength term, (1 -3)ult, represents the asymptote which the
hyperbolic stress-strain curve will approach at large strains. The (1 -3)f is the
deviatoric stress at failure. From a Mohrs diagram, it can be seen that:

( 1 3 )ult ( 1 + 3 )ult
2

sin c cos

When multiplied by Rf and substituting in the second equation above, this equation
can be written as follows:

( 1 3 ) f ( 1 + 3 ) f
2

sin R f c cos

Comparing this equation with Equation 5-3, it is seen that the inequality given in
Equation 5-3 provides an indicator as to how close the stress state is to the failure
state.

Page 120

SIGMA/W

Chapter 5: Material Properties

Nonlinear-elastic (Hyperbolic) soil parameters


The modulus and exponent values for the Nonlinear-Elastic model can be
determined by plotting triaxial test results on a log vs. log plot, as shown in Figure
5-3. The stiffness values E/Pa and Bm/Pa are plotted versus 3 / Pa . The slope of a
straight line through the data points gives the exponent. The modulus number K is
equal to the value on the vertical scale, where 3 / Pa is equal to 1.0.

Figure 5-3 Determination of non-linear soil properties

The bulk modulus Bm is defined as,

Bm =

( 1 + 2 + 3 ) / 3

where:

change in principal stress values, and

change in volumetric strain.

Page 121

Chapter 5: Material Properties

SIGMA/W

Further details of the hyperbolic model can be found in the report by Duncan et al.
(1980). This report also contains a list of hyperbolic parameters for some 135
different soils.
Table 5-3 Non linear elastic (hyperbolic) properties
Edit Box Label
E Modulus

Property
Initial soil stiffness, it is used when K(load) is zero.

Poissons Ratio

Constant value used when K(bulk) is zero.

K(load)

Modulus number describing the soil stiffness.

n(exponent)

A value describing the rate of change of the soil stiffness as a function


of the confining stress.

K(ur)

Modulus number used during unloading and reloading. When K(ur) is


zero, the material will unload or reload along the nonlinear stressstrain curve.

K(bulk)

Modulus number used to compute the bulk modulus. The bulk


modulus is used in turn to compute the Poisson's ratio.

m(exponent)

A value describing the rate of change of the bulk modulus as a


function of the confining stress.

Min. P. Ratio*

A minimum value for Poissons Ratio when it is computed from the


bulk modulus.

Pa

Atmospheric pressure (units must be consistent with the units of


length and force used throughout the problem).

Rf

Ratio between the asymptote to the hyperbolic curve and the


maximum shear strength (the ratio is usually between 0. 75 and 1. 0).

Cohesion

Cohesive strength of the soil.

Phi

Soil friction angle in degrees

Phi B

A value

(b ) used to make the cohesive strength a function of soil

suction (negative pore-water pressure).


* The computed Poisson's ratio is limited to a maximum of 0. 49 and a minimum of 0.01.

5.5

Elastic-plastic model

The Elastic-Plastic model in SIGMA/W describes an elastic, perfectly-plastic


relationship. A typical stress-strain curve for this model is shown in Figure 5-4.
Stresses are directly proportional to strains until the yield point is reached. Beyond
the yield point, the stress-strain curve is perfectly horizontal.

Page 122

SIGMA/W

Chapter 5: Material Properties

Figure 5-4 Elastic-perfectly plastic constitutive relationship

Plastic matrix
In SIGMA/W, soil plasticity is formulated using the theory of incremental
plasticity (Hill, 1950). Once an elastic-plastic material begins to yield, an
incremental strain can be divided into an elastic and a plastic component.

{d } = {d e } + {d p } ,
{d e } = {d } {d p }

or

Only elastic strain increments, de, will cause stress changes. As a result, stress
increments can be written as follows.
Equation 5-4

{d } = [Ce ]{d e } ,

or

{d } = [Ce ] ({d } {d p })

A function which describes the locus of the yield point is called the yield function
and is defined using the symbol, F. In the Elastic-Plastic model of SIGMA/W, the

Page 123

Chapter 5: Material Properties

SIGMA/W

yield point depends only on the stress state. Consequently, the yield function can
be written as follows in equation form.

F = F ( x , y , z , xy )
An incremental change in the yield function is given by:

dF =

F
F
F
F
d x +
d y +
d z +
d
x
y
z
xy xy

Alternatively, this equation can be written in the following matrix form.

dF =

{d }

The theory of incremental plasticity dictates that the yield function, F < 0, and,
when the stress state is on the yield surface, dF is zero. This latter condition is
termed the neutral loading condition, and, can be written mathematically as:

dF =

{d } = 0

The plastic strain is postulated to be:


Equation 5-5

{d } = G
p

where:
G

plastic potential function, and

plastic scaling factor.

Substituting the plastic strain from Equation 5-5 into the incremental stress
equation (Equation 5-4) gives:

{d } = [Ce ]{d } [Ce ]

Page 124

SIGMA/W

Chapter 5: Material Properties

Substituting the stress vector, {d}, into the neutral loading condition, the
following expression for the plastic scaling factor, , can be derived.

[Ce ]

G
[Ce ]

{d }

From the previous two equations, a relationship between stress increments and
strain increments can be obtained.

{d } = ([Ce ] C p ) {d }
where:

G F
[Ce ]


G
F
[Ce ]

[Ce ]
C p =

To evaluate the plastic matrix, [Cp], the yield function, F , and the plastic potential
function, G , need to be specified.
Yield criterion
SIGMA/W uses the Mohr-Coulomb yield criterion as the yield function for the
Elastic-Plastic model. The following equation provides a common form of the
Mohr-Coulomb criterion expressed in terms of principal stresses.

J
I

F = J 2 sin + 2 cos + sin 1 sin c cos


3
3
3
3

The Mohr-Coulomb criterion can also be written in terms of the stress invariants
I1 , I 2 and . The yield function, F, can then be written as follows (Chen and
Zhang, 1991).

Page 125

Chapter 5: Material Properties

SIGMA/W

J
I

F = J 2 sin + 2 cos + sin 1 sin c cos


3
3
3
3

where:

J2 =

2
2
1
2
x y ) + ( y z ) + ( z x ) + xy2
(

the second

deviatoric stress invariant,

3 3 J3
32

2 J2

1
3

= cos 1

the lode angle,

J 3 = xd yd zd zd xy2

the third deviatoric stress invariant,

I1 = x + y + z

the first stress invariant,

angle of internal friction, and

cohesion of the soil.

The deviatoric stress id in the ith-direction can be defined as:

id = i

I1
3

where i = x, y or z.

When the angel of internal friction, , is equal to zero, the Mohr-Coulomb yield
criterion becomes the Tresca criterion (Smith and Griffiths, 1988):

F = J 2 sin + c
3

The plastic potential function, G, used in SIGMA/W has the same form as the yield
function, F (i.e. G = F) except the internal friction angle, , is replaced by the
dilation angle, . Thus, the potential function is given by:

Page 126

SIGMA/W

Chapter 5: Material Properties

J
I

F = J 2 sin + 2 cos + sin 1 sin c cos


3
3
3
3

The derivatives of the yield function in terms of the stress invariants are computed
using the chain rule of differentiation.

F
F I1
F J2
F
=
+
+

I1
J 2

Derivatives of the Mohr-Coulomb yield function, with respect to the stress
invariants, can be written as follows:

F
sin
=
3
I1
F
1

1

=
sin cos +
sin + +
J2 2 J2
3
3
3

Equation 5-6

J
F

= J 2 cos + + 2 sin sin +

3
3
3

The derivatives of the stress invariants with respect to the stresses are:

I1
= 1 1 1 0

J2
= xd yd zd

2 xy

J3
3J J 2
3

=
3

2 J 2

2 J 2 sin 3
J3
J
J
J
= yd zd + 2 xd zd + 2 xd yd + 2 xy2

3
3
3
3

2 zd xy

Similarly, the derivatives of the potential function can be obtained by substituting


for in Equation 5-6.

Page 127

Chapter 5: Material Properties

SIGMA/W

The material properties required for this model are given in Table 5-4.
Table 5-4 Elastic-plastic material properties
Edit Box Text

Property

E Modulus

Initial linear-elastic stiffness of the soil

Poissons Ratio

Constant value

Cohesion

Cohesive strength of the soil

Phi

Soil internal friction angle

Dilation Angle

Soil dilation angle in degrees

in degrees

(0 ) .

If a value is not specified, the dilation angle is considered to be the


same as the internal friction angle.
Phi B

A value

(b ) used to make the cohesive strength a function of soil

suction (negative pore-water pressure).

5.6

Strain-softening model

A strain-softening model with a stress-strain curve as shown in Figure 5-5 is


included in SIGMA/W. The stress-strain curve is made up of three linear portions;
namely: an elastic portion up to the peak shear strength, a softening portion in
which the shear strength reduces from peak to residual, and, finally, a constant
residual shear strength portion. This model is similar to that proposed by
Chan (1986), who used a hyperbolic curve to describe the post-peak behavior.

Page 128

SIGMA/W

Chapter 5: Material Properties

Figure 5-5 Strain-softening constitutive relationship in SIGMA/W

Yield criterion
The strain-softening model in SIGMA/W is an elastic-softening plastic model. Its
yield criterion is a function of the stress state and the equivalent plastic strain. The
yield function, F, for the strain-softening model can be written in terms of shear
stress, q, and shear strength, cu.

F = F ( , p )
= q 3 cu
Under plane strain condition, it can be shown that at failure:

2 =

1 + 3
2

which gives at failure a shear strength, cu, equal to (1 - 3 )/2.


The shear stress state, q, is related to the second deviatoric stress invariant, J2, such
that:

q = 3J 2

Page 129

Chapter 5: Material Properties

SIGMA/W

The shear strength, cu, is dependent on the total equivalent plastic strain, p, as
given by the following equation:

Cu

cu = Cu R p
C
r

when p = 0
when 0 < p pr
when p pr

where:

( 1 3 )

Cu

peak strength and is equal to

Cr

residual cohesive shear strength,

total equivalent plastic strain,

pr

plastic strain where the softening line intersects the residual line, and

rate of softening.

2 at peak conditions,

p is the softening parameter for this model and is a measure of the total
(accumulated) deviatoric plastic strain given by:

p = d p
d p =

2
2
2
2
2 p
4
dex deyp ) + ( deyp dezp ) + ( dezp dexp ) + ( dexy )
(

3
9

where:

1
deip = d ip d vp , i = x , y , or z component
3
1
dexyp = d xyp
2
p
d v = ( d xp + d yp + d zp )

Page 130

SIGMA/W

Chapter 5: Material Properties

Plastic matrix
An incremental change in the yield function, F, is given by the following total
differential:

dF =

{d } +

F cu
d
cu p p

When the stress state is on the yield surface, the incremental change in yield
function, dF, is zero.

dF = 0 =

Equation 5-7

{d } +

F cu
d
cu p p

The incremental plastic strain vector {dp} is related to the derivative of the plastic
potential function G through the plastic scaling factor, .:

{d } = G
p

Since only elastic strains can cause stress changes, the constitutive relationship can
be written as:

{d } = [Ce ] ({d } {d p })
Equation 5-8

G
= [Ce ] {d }

Substituting the above expression for incremental stresses, {d}, into Equation 5-7
gives:

[Ce ]{d } =

F F cu
d p

cu p

[Ce ]

The associated flow rule states that plastic potential function is the same as the
yield function (i.e. G =F). Applying the associated flow rule and using the
definition of incremental plastic strain, the incremental equivalent plastic strain can
be expressed as follows.

Page 131

Chapter 5: Material Properties

SIGMA/W

2
2
2
2 q q q q q q 4 q
d p =

+
+
+
9 x y y z z x 3 xy

From these two equations, an expression for the plastic scaling factor, , can be
derived:

F
[Ce ]

=
{d }
F
F F cu
d p
[Ce ]

cu p
Finally, the incremental stress corresponding to a given incremental strain is
obtained by substituting into Equation 5-8.

{d } = ([Ce ] C p ) {d }
The plastic matrix [Cp] is calculated as follows:

F F
[Ce ]

C p =
F
F
F cu
d p
[Ce ]

cu p

[Ce ]

The plastic matrix [Cp] is the same as that used for the elastic-plastic model except
for the additional second term in the denominator.
Evaluation of the plastic matrix
To evaluate the plastic matrix, [Cp], it is necessary to compute the derivatives of

F cu
F
,
, and the
, and other derivatives terms
cu p

incremental equivalent plastic strain, d p .
the yield function,

Page 132

SIGMA/W

Chapter 5: Material Properties

The yield function for the strain-softening model can be written as:

F = q 3 cu
Since the shear strength, cu, is independent of stresses, the derivatives of the yield
function, F, with respect to stresses is given by:

F q

=

Considering that the shear stress, q , is equal to q = 3 J 2 , this equation can be
rewritten as follows.

q 1
=
2

3 J2

J 2

The derivatives of the stress invariant, J2, are given in above.


Finally, if it is further assumed that the soil strength, cu, remains unchanged within
a load increment, the following derivatives for the yield function, F, and the shear
strength, cu, can be obtained.

F
= 3
cu
cu
=0
p
These derivatives are used to evaluate [Cp], as given previously.
Von Mises criterion
When the rate of softening, R, is zero and the residual strength, Cr, is the same as
the peak strength, Cu, the stress-strain curve takes the form of an elastic-rigid
plastic curve. Under these conditions, the yield criterion for the strain-softening
model becomes the von Mises criterion and is given by the following equation:

F = q 3 Cu

Page 133

Chapter 5: Material Properties

SIGMA/W

where:
Cu

(undrained) shear strength.

Table 5-5 summarizes the material properties required in the strain-softening


model.
Table 5-5 Strain softening properties
Edit Box Label

5.7

Property

E Modulus

Initial linear-elastic stiffness of the soil

Poissons Ratio

Constant value

C(peak)

Peak cohesive strength of the soil

C(residual)

Residual cohesive strength of the soil

Softening Rate

Slope of softening portion of curve

Cam-clay model

The Cam-clay model is a critical state model as well as an elastic, hardening


plastic model. Its formulation in SIGMA/W is based on presentations by Atkinson
and Bransby (1978), and Britto and Gunn (1987). More detailed background and
theory of the Cam-clay model can be found in these references. In this section,
only the information required to explain the SIGMA/W formulation is presented.
The Cam-clay model uses effective stress parameters. In the following discussion,
effective stresses are denoted by a superscript ().
Figure 5-6 (a) schematically shows volume change versus pressure plots for a soil
comprising of a normal consolidation line and an over-consolidation line. The
over-consolidation line is also known as the swelling line. Consider a stress state
on the over-consolidation line. An increase in applied stress will cause the stress
state to move along the over-consolidation line towards the normal consolidation
line. Once past the intersection of the two lines, any further stress increase will
cause the stress state to move down the normal consolidation line.
When Figure 5-6 (a) is rotated counterclockwise through 90, the overconsolidation and normal consolidation lines show the characteristic of an elastichardening plastic stress-strain curve, illustrated in Figure 5-6 (b). The overconsolidation line is analogous to the initial linear elastic portion, while the normal

Page 134

SIGMA/W

Chapter 5: Material Properties

consolidation line is analogous to the hardening plastic portion of the stress-strain


relationship.

Figure 5-6 Analogy between volume-pressure and stress-strain relationships

Soil parameters
The Cam-clay model is an effective stress model which requires the following soil
properties:

Slope of the critical state line in the p ' q plane

Specific volume at the critical state when p ' is 1.0 (or, ln ( p ' ) is
0)

Slope of the isotropic over-consolidation (swelling) line

Slope of the isotropic normal consolidation line

v Specific volume

These five parameters are illustrated in Figure 5-7. The critical state line shown in
the p ' q plane is the locus of critical states projected onto that plane. The critical
state line has a slope of, , which is related to the angle of internal friction of the
soil. For the case of triaxial compression, can be expressed as:

Page 135

Chapter 5: Material Properties

SIGMA/W

6sin '
3 sin '

By definition, the specific volume, v , is related to the void ratio, e , through the
following expression:

v = 1+ e

Figure 5-7 Definition of model parameters for Cam-clay

Yield function
The yield function for Cam-clay, as illustrated in Figure 5-8, can be expressed in
terms of stress invariants p' and q as follows (Atkinson and Bransby, 1978):

F=

p
q
+ ln 1
Mp
px

where:
p'x

the peak mean stress and is the value of p at the critical state line or t.

The peak mean stress, p'x, is related to the pre-consolidation pressure, p'c, by:

Page 136

SIGMA/W

Chapter 5: Material Properties

ln px = ln pc 1
or
px = p 2.71828

Figure 5-8 Yield curve for the Cam-clay model

The stress invariants p' and q are alternative forms of the first stress invariant, I1',
and the second deviatoric stress invariant, J2, respectively.

p =

I1'
3

and:

q = 3J 2
By definition, the first stress invariant, I1', can be expressed in terms of effective
stresses as:

I1' =

1
( x + y + z )
3

Page 137

Chapter 5: Material Properties

SIGMA/W

and the second deviatoric stress invariant, J2, can be written as:

J2 =

2
2
1
2

+ ( z x ) + xy2

(
)
(
)
x
y
y
z

The second deviatoric stress invariant, J2, is the same for either total or effective
stresses.
Equations for the over-consolidation line and the critical state line can be used to
calculate the peak mean stress, p'x. As illustrated inFigure 5-8, the specific volume
at critical state, vx, for a particular over-consolidation line can be written as:

vx = v + ln p ln px
From the critical state line, the same specific volume, vx, can also be calculated
using:

vx = ln px
Eliminating the specific volume, vx, from these two equations gives the following
expression for the peak mean stress, p'x.
Equation 5-9

v ln p
px = exp

Page 138

SIGMA/W

Chapter 5: Material Properties

Figure 5-9 Definition of soil properties for Cam-clay model

Plastic matrix
The plastic matrix [Cp] is developed using the yield function, F, in a manner
similar to the development used for the Elastic-Plastic and Strain-Softening
models. In the following discussion, subscripts e and p denote elastic and plastic
state respectively.
As shown in Equation 5-9, the peak mean stress, p'x, is a function of the specific
volume, v. If it is postulated that a change in the peak mean stress, p'x, can only be
caused by a plastic change in the specific volume, an incremental change in the
yield function, F, is given by:
Equation 5-10

dF =

{d } +

F px p
dv = 0
px v p

Page 139

Chapter 5: Material Properties

SIGMA/W

From a given set of normal and over-consolidation curves, an incremental plastic


change in specific volume can be expressed as:

dv p = dv dv e = ( )

dp
p

The incremental plastic specific volume is related to the incremental plastic


volumetric strain through:

dv p = v0 d vp
where:
dvp

incremental plastic specific volume,

v0

initial specific volume, and

dv

incremental plastic volumetric strain.

Since only elastic strains can cause stresses to change, the incremental effective
stress constitutive relationship can be written as follows:

{d } = [Ce ] ({d } {d p })

G
= [Ce ] {d } p

where:
[C'e]

the effective elastic matrix,

the plastic potential function, and

the (plastic) scaling factor.

Substituting {d'}, from the previous equation into Equation 5-10 gives:
Equation 5-11

F
F
G F px p
Ce ]{d } =
Ce ] p
dv
[
[

p


px v

Page 140

SIGMA/W

Chapter 5: Material Properties

The plastic increment in specific volume, dvp, can be calculated from the original
specific volume and the incremental plastic volumetric strain, d vp .

dv p = vo d vp = vo p (

G G G
+
+
)
d x d y d z

Applying the associated flow rule (i.e. the plastic potential function, G, is the same
as the yield function, F), the following expression is obtained for the plastic scaling
factor, p.

F
[Ce ]

p =
{d }
F
F F px

[Ce ] p v0 Fm

px v
where:

Fm =

G G G F F F
+
+
=
+
+
d x d y d z d x d y d z

A relationship between incremental stress and incremental strain is obtained by


substituting p into Equation 5-11.

{d } = ([Ce ] C p ) {d }
where:

F F
[Ce ]

C p =
F
F F px
Ce ]
vF
[

p 0 m

px v

[Ce ]

[C'p] is termed the plastic matrix.

Page 141

Chapter 5: Material Properties

SIGMA/W

Evaluation of the plastic matrix


To evaluate the plastic matrix, C ' p , for the Cam-clay model, it is necessary to
compute the elastic matrix [C 'e ] , the vector of derivatives of the yield function,

p
F px
F
,
, v0 , and Fm , px and Fm , and the

, the derivative terms


p
px v
v

initial specific volume, v0 .
For given values of the effective stress, Youngs modulus, E ' , and Poissons ratio,
v ' , the elastic matrix, [C 'e ] , is the same as that for a linear elastic model. The
Poisson's ratio, v ' , is a constant value input by the user, while the modulus, E ' , is
calculated from the slope of the over-consolidation line, .
The slope of the over-consolidation line, , can be written as:

dv
dv
= p
d ln p
dp

A change in volumetric strain is related to a change in specific volume through the


following equation:

d v =

dV
de
dv
=
=
1+ e
V
v

where:

total volume

void ratio of the soil

Combining the previous two equations, the following expression for the
incremental volumetric strain is obtained.

d v = (

vp

)dp

By definition, the effective bulk modulus, K ' , is given by:

Page 142

SIGMA/W

K =

Chapter 5: Material Properties

dp
d v

The bulk modulus, K ' , can also be expressed in terms of .

K =

vp

(1 + e ) p

In turn, the following expression is obtained for the effective stress modulus E ' .

E = 3K (1 2 ) =

3 p

(1 + e )(1 2 )

The vector of derivatives of yield function with respect to stresses is:

F
F

=
x

F
y

F
z

F
xy

These derivatives can be expressed as derivatives with respect to stress invariants


I1 ' and J 2 using the chain rule of differentiation.

F F p F q
=
+
i p i q i
F I1
F J2
=
+ 3
I1 i
q i
where:
i

stress components x, y, z and xy

By definition, q = 3 J 2 . The derivatives of these stress invariants are given


previously. The derivatives of the yield function F with respect to the stress
invariants are expressed as follows:

Page 143

Chapter 5: Material Properties

SIGMA/W

F
q
1
=
+
2
p
p
p
1
F
=
q p
The remaining terms in the plastic matrix C p ,

F px
,
, v , and Fm , and
px v p 0

Fm , are described in the following equations:

F
1
=
px
px
px
p
= x
p
v

Fm =

F F F
+
+
x y z

F p p p F q q q
+
+
+
+

p x y z q x y z
F
F
1) +
(0)
=
(
p
q
=

1
q
+
2
p
p

Initial condition
The initial soil conditions are defined by the initial specific volume, v0, together
with the initial stresses. In an analysis using the Cam-clay model, the initial
stresses must be defined. It is necessary to know the initial pre-consolidation
pressure, p'c, in order to compute v0. The relationship between the initial void ratio,
v0, and the pre-consolidation pressure, p'c, is illustrated in Figure 5-9.
A user can either specify the pre-consolidation pressure, p'c, directly when defining
the model parameters or allow SIGMA/W to approximate the pre-consolidation

Page 144

SIGMA/W

Chapter 5: Material Properties

pressure, p'c, from the in-situ stresses, the over-consolidation ratio, OCR, and the
coefficient of earth pressure at rest, K0.
If the pre-consolidation pressure is not specified when defining a Cam-clay soil,
SIGMA/W estimates p'c in the following manner. In an analysis using the Camclay model, the initial stresses represent the present in-situ stress state. In equation
form, the present in-situ stress state is

x 0

{ 0 } = y 0
z0
xy 0
The maximum vertical stress that the soil has experienced in the past is usually
determined from an oedometer test. The ratio of the maximum vertical stress in the
past, 'vmax, to that at present, 'v, is known as the over-consolidation ratio, OCR.
In SIGMA/W, OCR is a user-specified parameter for each soil type. The
relationship between the maximum vertical stress, 'vm, and maximum horizontal
stress, 'hm, is approximated as follows:

K 0 = 1 sin
= K 0 vmax

hmax
Assuming that the shear stress is zero, the maximum stress vector, {'m}, can be
written as:

' y 0 K 0 OCR
xmax
' OCR

y0

{ m } ymax
=

zmax ' y 0 K 0 OCR


0

0
The historic past maximum mean stress, p'm, and maximum shear stress, qm, are
respectively:

pm =

1
( xm + ym + zm )
3

Page 145

Chapter 5: Material Properties

=
qmax

1
2

xmax

SIGMA/W

) + ( zmax
xmax
)
ymax ) + ( ymax zmax
2

The pre-consolidation pressure, p'c, occurs in the p'-q space where the shear stress,
q, is zero. Therefore, from the yield function of Cam clay, the following expression
can be obtained:

p
ln c = 1 , or, ln px = ln pc 1
px
Substituting the natural logarithm of the pre-consolidation pressure, ln p'x, the yield
function for Cam-clay can be rewritten as:

q
+ ln p ln pc = 0
p
Therefore, the pre-consolidation pressure, p'c, is given by:

exp max
pc = pmax

M pmax
Once the pre-consolidation pressure, p'c, is obtained either as a user-specified value
or from the above equation, SIGMA/W proceeds to calculate the initial specific
volume, v0. First, N, the intersection of the normal consolidation line and the
vertical axis (i.e., when ln p = 0), is calculated from the following expression:

= + ( )
From the normal consolidation line passing through the pre-consolidation pressure,
p'c, the specific volume at the critical state is given by:

vc = N ln pc
Also, from the over-consolidation line passing through the pre-consolidation
pressure, p'c, the initial specific volume, v0, can be expressed as:

p
v0 = vc + ln c
p0

Page 146

SIGMA/W

Chapter 5: Material Properties

In this expression, the initial mean stress, p'0 , can be calculated from the initial
stresses using the equation:

p0 =

1
( x 0 + y 0 + z0 )
3

This completes the computation for parameters defining a Cam-clay model in


SIGMA/W.
Soil parameters
The Cam-clay models use the parameters; M , , , and . These material
properties can be obtained from some more common parameters such as ' and
the compression index Cc . This section presents some useful relationships when
you are working with the Cam-clay models.
The slope of the critical state line, M, obtained from a triaxial compression test is
related to the friction angle ' through:

M=

6sin
3 sin

The compression versus vertical pressure characteristics of soil are usually


obtained from a one-dimensional consolidation test, and it is common practice to
obtain the compression index Cc from a plot of void ratio (e) versus log10(p). The
compression indices are related to the slope and by the relationship:

Cc
2.303
C
= s
2.303

Alternatively, and can be obtained from a plot of void ratio versus ln(p)
instead of void ratio versus log10(p).
The specific volume v is equal to 1 plus the void ratio (e).

Page 147

Chapter 5: Material Properties

SIGMA/W

The parameter N can be estimated by projecting the normal consolidation line on


an e versus ln(p) plot onto the vertical line where p equals 1. 0. Once N is known,
the parameter can be computed by the following equation:

= N +

for Cam-Clay

= N ( + ) ln 2

for Modified Cam-Clay

The parameter is defined as the intersection point of the critical state line with
the vertical p = 1.0 line.
Another vitally important item of data required for a Cam-clay analysis is the
present (initial) in-situ stress state and the initial yield surface. Establishing the
present in-situ stress state is relatively simple and is the same procedure as for any
other SIGMA/W nonlinear analysis. Establishing the initial yield surface can be
done by specifying the pre-consolidation pressure (pc) as a material property, or by
computing pc from the initial in-situ stresses together with a specified overconsolidation ratio OCR. The SIGMA/W procedure for using this second
alternative is described in the chapter on Boundary and Initial conditions.
Table 5-6 summarizes the material properties required in the Cam-clay model.
Table 5-6 Cam clay / modified cam clay properties
Edit Box Label
O. C. Ratio

Property
Over-consolidation ratio

Poissons Ratio

Constant value

Lambda

Slope of normal consolidation line

Kappa

Slope of over-consolidation (swelling) line)

Gamma

Specific volume when p' is equal to 1.0

Mu

Slope of critical state line

Pc'

Pre-consolidation pressure. When Pc' is not specified, it is


calculated using the initial stress state and the O. C. Ratio

5.8

Modified Cam-clay model

The modified Cam-clay model is similar to the Cam-clay model except that the
yield function is in the shape of an ellipse instead of a tear drop. The yield curve
for the modified Cam-clay model is illustrated in Figure 5-10.

Page 148

SIGMA/W

Chapter 5: Material Properties

Figure 5-10 Yield function for the modified Cam-clay

The yield function for the Modified Cam-clay model is given by the following
equation (Britto and Gunn, 1987):

q 2 = 2 ppc 2 p2
where:
p'c

pre-consolidation pressure.

The parameters used to define the Modified Cam-clay model are: M , , , and v .
These parameters are the same as those used for the Cam-clay model as discussed
in the previous section.
Similar to the Cam-clay model, SIGMA/W uses the peak mean stress, p 'x , to
determine the size of the yield locus for the Modified Cam-clay model. The peak
mean stress, p 'x , is the isotropic pressure when a soil reaches its critical state. At
the critical state, the shear stress, q, is given by the following equation:

Page 149

Chapter 5: Material Properties

SIGMA/W

q = px
Substituting this value of shear stress, q, into the modified yield function equation
gives:

pc = 2 px
Therefore, the yield function, F, for the Modified Cam-clay model can be written
as:

F=

q2
+ 2 p 2 2 px
p

The derivatives of the yield function, F, with respect to stress invariants p ' and q
are respectively:

F
q2
2
= 2
p
p
F 2q
=
q p
F 2q
=
q p
px
q
are the same as those used for the
, [Ce ] , and
vp

Cam-clay model. The remaining variables in the plastic matrix, C ' p , for
p
,

The terms

modified Cam-clay are:

q2
Fm = M 2
p
2

Page 150

SIGMA/W

Chapter 5: Material Properties

F
= 2 M 2

px
Fm = M 2

q2
p 2

Except for the differences noted in the following discussion, the procedure to
define the initial condition of the Modified Cam-clay model is the same as that for
the Cam-clay model described in the previous section. If a pre-consolidation
pressure is not defined by the user, it is calculated using:

pc =

1
2
2
+ M 2 pmax
( qmax
)

M pmax
2

The interception of the isotropic normal consolidation line with the vertical axis
(i.e. ln p = 0 ), is given by:

= + ( ) ln 2
Using these values of the pre-consolidation pressure, pc , and the vertical
intercept, , the initial specific volume, v0 , can be computed using equations
outlined in the Cam-clay model section of this chapter.

5.9

Slip surfaces

SIGMA/W can simulate a slip surface using quadrilateral elements with Slip
Surface as its soil model. Using the Slip Surface material model, a user can specify
stiffness in two orthogonal directions. These directions are parallel and
perpendicular to the long side of the element. Figure 5-11 shows a typical slip
surface element with its long (x-) axis inclined at angle from the global x-axis
with nodes numbered as 1, 2, 3 and 4.

Page 151

Chapter 5: Material Properties

SIGMA/W

Figure 5-11 Typical slip surface element

SIGMA/W treats a slip surface element as a combination of four bar elements. For
example, for the element illustrated in the figure, there are two bar elements with
stiffness, Kn, in the local normal y direction, 1-4 and 2-3; and two bar elements in
the local tangential x-direction, 1-2 and 4-3, with stiffness Kt. Kt and Kn are the
user-specified tangential and normal stiffness for the element.
When formulating the element stiffness matrix [K], instead of evaluating the
integral

[ B ] [C ][ B ] dv , SIGMA/W first forms the local stiffness matrix [K'] by


T

assigning stiffness values Kt or Kn directly to the appropriate location in the matrix.


For the element illustrated above, its local stiffness matrix, [K'], in the local (x' - y')
coordinate system is:

Kt
0

Kt

0
0

0
0

0
Kn

Kt
0

0
0

0
0

0
0

0
0

Kt

Kn

Kn

Kt

Kt

Kn

Kn

0
Kn

0
0

0
0

Kt
0

0
0

Kt
0

Page 152

0
K n
0

0
0

0
0

K n

SIGMA/W

Chapter 5: Material Properties

[K'] is then rotated to the global coordinate system to obtain [K] using the
following transformation:

[ K ] = [T ][ K ][T ]

where:

cos 2

sin 2
[T ] =
0

sin cos

sin 2
cos
0
2

sin cos

2sin cos

0 2sin cos

1
0

0 cos 2 sin 2
0

When the normal stress on a slip is zero or tensile (negative), the stiffness values
are reduced by a factor of 0. 001.
Table 5-7 summarizes the slip surface element material properties.
Table 5-7 Slip surface properties
Edit Box Label

Property

K Normal

Stiffness normal to the slip surface

K Shear

Stiffness parallel to the slip surface

In order to model a slip surface, a slip element should be rectangular and relatively
thin. However, it is highly recommended that you do not make the slip elements
too thin. The slip elements should have sufficient thickness such that they are
readily visible in the finite element mesh when the mesh is displayed at 100% and
the problem scale is the same in both horizontal and vertical directions. SIGMA/W
treats the longer side of a slip element as length and the shorter side as thickness.
Stresses and pore-water pressures in a slip element are obtained from the
neighboring elements above and below. In these neighboring elements, stresses are
first extrapolated from the Gauss points to the interface nodes. Then, from these
nodes, SIGMA/W interpolates for values at the Gauss point using shape functions.
The slip surface elements are treated as having drained behavior. Therefore, porewater pressures are not computed on the basis of stress changes in the element.

Page 153

Chapter 5: Material Properties

SIGMA/W

A slip surface element is linear elastic and has no maximum shear resistance. If
you wish to model the slip material as having some maximum shear strength, it is
best to select a different soil model (e.g., the Elastic-plastic model).

Page 154

SIGMA/W

Chapter 6: Boundary Conditions

Boundary Conditions

This chapter discusses the various types of boundary conditions that can be applied
in SIGMA/W analyses. There are many ways to apply forces and displacements to
soil and structural elements in order to replicate real world loading conditions.
Before any loading is applied, it is important to know what the model will do with
the data you input. A discussion of the incremental finite element formulation has
been on-going throughout the book (see Chapter 3 and the Theory chapter). It is
critical to understand that a load should only be applied on one step in the model
and then removed. Sometimes you must remove a load manually (such as turning
off body loads after the in-situ conditions are solved) and sometimes SIGMA/W
will turn them off automatically (such as during fill placement when body loads
are only counted on the step the fill is activated).
This chapter shows what boundary condition options are available and how the
model uses the data. The following chapter on Analysis Types discusses the
differences between in-situ analysis and load-deformation analysis and when
various boundary conditions may be used. Therefore, the present discussion makes
limited reference to when different types of boundary conditions are applied.

6.1

Multiple boundary condition types

Multiple boundary condition types are implemented to support virtually all loaddeformation modelling scenarios. In SIGMA/W, displacement, force or spring
boundary conditions may be applied to nodes, and stress and fluid pressure
boundary conditions may be applied to edges. Stress boundary conditions can be
applied to external element edges and may be specified as a normal and tangential
stress pair, or as an x-stress and y-stress pair.
An example of the unique way in which boundary conditions can be used is given
in the following figure, where struts used in a braced excavation can be modeled
by using a spring boundary condition at the location of the struts.

Page 155

Chapter 6: Boundary Conditions

SIGMA/W

Figure 6-1 Example of spring boundary condition representing a strut

6.2

Force or displacement

Fundamentally, there are only two types of boundary conditions that can be
applied in a stress-deformation model: force or displacement. How these
conditions are interpreted and applied is a different matter, because there are many
options depending on what your objectives are.
In all stress-deformation models it is critical to bound the problem. This means
that you must define some parts of the geometry as zero displacement boundary
conditions. If there is no bounding to the problem, how can there be any
reactionary force to the applied loading?
In general, it is common to bound the left and right sides of a problem along with
the bottom edge. A typical example of this is given in Figure 6-2, where the
bottom and side boundary nodes are specified as zero x- and zero y-displacement.
It is quite easy to see why there should be zero movement at the base of the model
in both directions, but there can be some question about when to also set the
vertical movement along the sides to zero. The answer to this latter question
depends on what you are trying to model. If you are modeling a larger domain
where you are saying the soil at the far sides is solidly attached to other soil even
farther away, then both x- and y-displacement should be zero. If you are modeling
a triaxial cell that is symmetric about the vertical axis, the axis of symmetry should
be fixed in the horizontal direction, and the bottom should be free to move in the x, but not y-direction as shown in Figure 6-3.

Page 156

SIGMA/W

Chapter 6: Boundary Conditions

+1

+2
+1

+6
+5

+6

+6

+5

+5

+6
+5

+4

+4

+4

+4

+5
+4

+3

+3

+3

+3

+3

+3

+4
+3

+2

+2

+2

+2

+2

+2

+2

+2

+3
+2

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+2
+1

+1

+3
+2

+4
+3

+5
+4

Figure 6-2 Bounding the model with zero displacement along edges

Figure 6-3 Symmetrical axis displacement boundary conditions

Page 157

Chapter 6: Boundary Conditions

6.3

SIGMA/W

Body loads

Body forces are included in an analysis by specifying a non-zero load (or weight)
per unit volume. For each element, SIGMA/W computes the volume of the
element, multiplies the volume by the specified unit load of the material, and
applies the total element body load as forces at the nodes of the elements.
Gravity body loads are included by setting the vertical body load to the unit weight
of the soil. Horizontal body loads can be specified to simulate pseudo-static
seismic forces. The direction of the body load is controlled by making the load
positive or negative, as described in Table 6-1.
Table 6-1 Body load sign convention
Force

Direction

Sign

Gravity

Vertical

Positive

Uplift

Vertical

Negative

Applied to the right

Horizontal

Positive

Applied to the left

Horizontal

Negative

The K0 value is only used in an in-situ type of analysis, as described in the


following chapter and in Chapter 3. A K0 value greater than 1.0 means that the
horizontal stresses are greater than the vertical stresses. In SIGMA/W a graphical
feedback lets you know if a body load is applied to an element. Any horizontal or
vertical body loads are displayed as a hatched pattern within each element as show
below in Figure 6-4. If you have applied a body load in an in-situ analysis and you
see the hatching is still visible in the load-deformation analysis, you should reset
the body load values to zero so that they are not counted on each load step in the
new analysis.

Page 158

SIGMA/W

Chapter 6: Boundary Conditions

Figure 6-4 Hatch pattern to show body load applied to element

6.4

Nodal boundary conditions

In SIGMA/W, boundary conditions can be defined at nodes or along element


edges. While some types of boundary conditions can be drawn on either the edge
or node to get the same desired effect, others cannot. A fluid pressure boundary
condition must be drawn on an edge for example.
The Draw Node Boundary Conditions command allows you to specify
displacement, force, or spring boundary conditions at nodes. Additionally, rotation
or moment boundary conditions can be specified on nodes that also belong to
structural beam elements. All of these descriptions are unique to different types of
real life conditions, but numerically, they all reduce to either forces or
displacements.
The on-line help provides details of how to apply the boundary condition in the
model. In general, however, node boundary conditions are defined in two stages.
First, specify the boundary conditions, such as the boundary type, action, and
boundary function number. Secondly, select all nodes that are to have these
boundary conditions.
Table 6-2 lists the various nodal boundary conditions available in SIGMA/W and
the shape and color of the symbol that will appear once they are specified. Nodes
can be changed from active boundary nodes to regular nodes by setting the x- and
y-boundary types to (none) and then clicking on the boundary nodes. The boundary
symbols are removed from the selected nodes.

Page 159

Chapter 6: Boundary Conditions

SIGMA/W

Table 6-2 Symbols used for each type of boundary condition


Type

Value

Symbol
X-Direction

Symbol
Y-Direction

Description

(none)

--------

--------

--------

none

Displacement

positive (+)

Displacement

negative (-)

Displacement

zero (0)

Force

positive (+)

Force

negative (-)

Force

zero (0)

Spring

not applicable

right
up
left
down

hollow red
arrow
hollow red
arrow
hollow red
triangle
solid red arrow

right
up

solid red arrow

left
--------

down
--------

none
solid green
arrows

left or right
up or down
Spring

zero (0)

Rotation

--------

--------

none

zero (0)

--------

hollow red circle

Moment

positive (+)

clockwise
rotation

hollow blue
circle

Moment

negative (-)

counterclockwise

hollow blue
circle

Boundary nodes have a negative x- or y-value point in the negative x- or y-direction, while
boundary nodes that have an x- or y-value that is positive or equal to zero point in the
positive x- or y-direction.

Page 160

SIGMA/W

6.5

Chapter 6: Boundary Conditions

Edge boundary conditions (pressures or stresses)

In SIGMA/W, boundary conditions can be defined along element edges or at


nodes. Edge pressure boundary conditions are defined in two stages. First, specify
the boundary conditions, such as the pressure boundary type, action, and boundary
function number. Secondly, select all edges that are to have these boundary
conditions. Edges can be changed from pressure boundary edges back to regular
edges by setting the pressure boundary type to (none) and then specifying the
boundary edges.
Table 6-3 shows the various types of edge boundary conditions that can be
specified along with their associated symbol and data conventions. An easy way to
distinguish between positive (+) and negative(-) normal stress is to remember that
compressive stress is positive. For tangential stresses, going around the element in
a counter-clockwise direction is positive. You can also check the direction of the
arrow displayed at a normal/tangential pressure boundary.

Table 6-3 Symbols used for edge boundary conditions


Type

Value

Symbol

Description

(none)

--------

none

--------

Normal Stress

positive (+)

blue line along edge, blue arrow


perpendicular towards edge

Normal Stress

negative (+)

blue line along edge, blue arrow


perpendicular away from edge

Normal Stress

zero (0)

blue line along edge

Tangential Stress

positive (+)

blue line along edge along edge,


blue arrow in counter-clockwise
direction around an element

Tangential Stress

negative (-)

blue line along edge along edge,


blue arrow in clockwise direction
around an element

Page 161

Chapter 6: Boundary Conditions

SIGMA/W

Tangential Stress

zero (0)

blue line along edge

X-Stress

positive (+)

blue line along edge, blue arrow


pointing right

X-Stress

negative (-)

blue line along edge, blue arrow


pointing left

X-Stress

zero (0)

blue line along edge

Y-Stress

positive (+)

blue line along edge, blue arrow


pointing up

Y-Stress

negative (-)

blue line along edge, blue arrow


pointing down

Y-Stress

zero (0)

blue line along edge

Fluid Elevation

greater than
min. edge
y-coordinate

blue line along edge, blue arrow


perpendicular towards edge

Fluid Elevation

less than min.


edge
y-coordinate

blue line along edge

NOTE: Edge boundary arrow symbols are only drawn for pressure boundary edges that are
on the perimeter of the mesh. Edge boundaries along internal element edges do not have
any arrow symbols shown, since the net resultant force on internal pressure boundaries is
always zero.

Converting edge conditions to nodal equivalent forces


Pressures along the edge of an element are a form of nodal boundary forces. The
specified pressure multiplied by the length of the element edge gives a total force.
The force is then proportionately divided among the nodes along the element edge.

Page 162

SIGMA/W

Chapter 6: Boundary Conditions

As illustrated in Figure 6-5, the force computed from the specified pressure is
equally divided between the two corner nodes for a 4-noded element. For a higherorder 8-noded element, 1/6 of the force is assigned to each of the corner nodes and
4/6 is assigned to the middle intermediate node.

F/2

4/6

F/2
constant pressure

F/6

F/6

4-noded

8-noded

Figure 6-5 Pressure equivalent nodal forces

The nodal forces are actually computed numerically by integrating along the edge
of each element (see the Theory chapter). This generalizes the scheme and makes it
possible to consider a variable pressure distribution along the edge of the element.
It is useful to remember that specified boundary pressures cannot be used directly in a
finite element analysis - they must always be converted to nodal forces.

Fluid pressure conditions


A fluid pressure boundary is a special type of normal pressure boundary. The
pressure is defined by specifying a water surface elevation as illustrated in Figure
6-6. The pressure is computed from the distance between the specified water
surface elevation and the y-coordinate of the boundary node. All the element edges
between 2-3 and 3-4 are flagged as fluid pressure boundaries.

Page 163

Chapter 6: Boundary Conditions

SIGMA/W

Specified elevation

Fluid pressure boundary

Figure 6-6 Illustration of fluid pressure boundary

When you apply the fluid pressure type to an element edge, SIGMA/W internally
calculates equivalent force applied at each of the nodes defining the edge. The
nodal pressure is obtained by subtracting the nodal y-coordinate from the fluid
elevation and multiplying it by the fluid self-weight. SIGMA/W will only consider
positive fluid pressures. Therefore, for the fluid pressure boundary condition to
have an effect, the selected edge must have a y-coordinate less than the fluid
elevation.
NOTE: drawing a water table above the mesh does not generate a pressure
boundary condition on the mesh. It allows pore-water pressures in the soil to be
known, but it does not apply a total force (load) condition due to the weight of the
water on the ground surface. You must specify both water table and fluid pressure
if you need soil water pressures and surface water forces.
Defining a pressure boundary at internal element edges
By design, SIGMA/W allows you to define a pressure boundary only at external
element edges, which are the edges belonging to a single active element. If you
wish to define a pressure boundary at an internal element edge, you can use the
equivalent nodal force approach as described below.
Along an internal element edge, the equivalent nodal forces can be calculated by
multiplying the applied pressure with the corresponding contributing areas and
then applying those nodal forces as nodal boundary conditions. The contributing
areas are dependent on the analysis type (either two-dimensional or axisymmetric)
and on whether there is a mid-side node. Figure 6-7 and Figure 6-8 show the
contributing areas computed for corner nodes, a1 and a2 ; and the mid-side node,

Page 164

SIGMA/W

Chapter 6: Boundary Conditions

a4 . For an axisymmetric analysis, it is important to note that the contributing areas


are specified per unit radian when the element thickness is 1.0 and in units of
2 radians when the element thickness is 2 .

Figure 6-7 Contributing area for two dimensional elements of unit thickness

Page 165

Chapter 6: Boundary Conditions

SIGMA/W

Figure 6-8 Contributing area for axisymmetric elements over 1 radian

Page 166

SIGMA/W

6.6

Chapter 6: Boundary Conditions

Transient boundary conditions

A boundary condition function may be associated with each boundary condition


instead of specifying a constant value. This feature is useful for specifying
boundary conditions that vary with time or load step. In addition, transient
boundary conditions may be cycled, allowing specification of transient boundary
conditions that repeat themselves with some frequency.
SIGMA/W uses the incremental change in a function as the boundary condition for
each load step. In the case of a force function, the incremental applied force is:

F = Fi Fi 1
where i is the time step number.
SIGMA/W obtains Fi and Fi-1 from the boundary function. The difference is
applied as the boundary condition for the i-th time step.
As a result, it is necessary to start defining the function at Time 0. The time
increment for the first step then is:

F1 = F1 F0
When you create a boundary function such as the one illustrated in Figure 6-9, you
should think of the Y-value as a cumulative value. If you want 10 lbs force applied
at load step 2, but at no other steps, then the function should look like the one
below. For all steps past step 2, there is no change in load so no load is applied. If
you had the function drop back to zero for all additional steps, then on the second
step, the program would use a change in load of 0 100 = -100 lbs and you would
undo what you tried to accomplish in the first step.

Page 167

Chapter 6: Boundary Conditions

SIGMA/W

10

Y-Force

0
0

Time

Figure 6-9 Step load of 10 on step 2 and zero on all other steps

Page 168

SIGMA/W

Chapter 7: Analysis Types

Analysis Types

SIGMA/W is capable of solving different types of analyses and in two different


views. This chapter discusses the options available. The type of analysis you
specify will depend, in many cases, on which aspect of a real life project you are
focusing on. It is not unreasonable to use several of the analysis types on a single
problem. For example, you may set up initial soil stress conditions using an in-situ
analysis, then do a load/deformation analysis to apply fill and form and
embankment, then use a coupled or uncoupled consolidation analysis to dissipate
the excess pore-water pressures and determine the volume change in the
foundation soil due to embankment loading.
Consolidation is listed as a special analysis type, but it is discussed in detail in a
dedicated chapter later in this book.

7.1

Problem view

A two-dimensional view analyzes a vertical cross-section with a unit thickness as


specified when you create the region geometry. An axisymmetric view analyzes a
problem that is symmetrical around a vertical axis (e.g., a right cylindrical tank or
a laboratory tri-axial test). In SIGMA/W, the axis of symmetry is fixed at the
vertical axis (coordinate r = 0). By default, the axisymmetric analysis is formulated
for a sector of one radian.

7.2

Initial in-situ stresses

There are different ways to determine the in-situ stresses prior to a loaddeformation or consolidation analysis. SIGMA/W has two in-situ analysis types,
plus the load-deformation analysis can also be used.
In-situ 1 analysis
An In-situ 1 analysis is used to establish the initial in-situ stress conditions induced
by the self-weight of the soil. In this analysis type, you must specify the body load
(unit weight) and the relationship between horizontal and vertical effective stresses
(K0). This analysis only uses the linear-elastic soil model and it only applies to
horizontal ground surface where there is no shear stress in the ground.

Page 169

Chapter 7: Analysis Types

SIGMA/W

SIGMA/W computes the vertical stress y on the basis of the gravity load (unit

weight). The horizontal stress x is computed as y u multiplied by the userspecified K0 plus the pore-water pressure u. In equation form:

x = ( y u ) K 0 + u
The shear stress is arbitrarily set to zero.
This procedure is shown in Figure 7-1.

Figure 7-1 In-situ 1 analysis to establish in-situ stress conditions

In-situ 2 analysis
An In-situ 2 analysis establishes the initial in-situ stress conditions using the
submerged weight of the soil. The initial pore-water pressure conditions are
obtained from the specified initial drawn water table or the selected external porewater pressure file. A discussion on drawing the water table is given later in this
chapter.
In an In-situ 2 analysis, SIGMA/W does the following: for any element that is
below the specified water table (or any positive pore-water pressure at a Gauss
point), SIGMA/W computes the submerged unit soil weight. The submerged unit

Page 170

SIGMA/W

Chapter 7: Analysis Types

weight is taken as the specified soil unit weight (specified as a vertical body load)
minus the water unit weight. With the modified soil unit weight(s), SIGMA/W
runs a gravity turn-on analysis (the applied load comes from the soil weight). This
step gives the in-situ effective stresses below the water table and the total stresses
above the water table. Next, below the water table the pore-water pressure is added
to the effective stresses to get the total stresses, and above the water table the
negative pore-water pressures are added to the total stress to get the effective
stress.
This procedure has the effect of applying K0 to the effective stresses below the
water table and applying K0 to the total stresses above the water table.
Load deformation analysis for in-situ stresses
A load/deformation analysis can also be used to establish the initial in-situ stress if
K 0 is less than 1.0. The reason for this limitation is that when there is no lateral
movement, K 0 is related to Poissons ratio by the equation:

K0 =

Since v is limited to 0. 49 in SIGMA/W, K 0 cannot be greater than 1.0.


The advantage of this method is that the ground surface can have any shape. The
limitation is that the method can only be used for a particular value of K 0 .
To apply this method, the boundary conditions at the ends of the problem must be
as shown in Figure 7-2, and the soil must be assigned a gravity load (vertical body
load) equal to the soil unit weight. Then you can use the resultant stress output file
as the initial stress condition file for your analysis.

Page 171

Chapter 7: Analysis Types

SIGMA/W

Figure 7-2 Load deformation analysis to compute in-situ stress

Historic in-situ stress


For the Cam-clay and Modified Cam-clay models, it is necessary to define the
initial yield locus that corresponds to conditions that existed when the soil was at
its normally consolidated state. This initial yield locus is controlled by the preconsolidation pressure, pc. This parameter can be user-specified or calculated
using an input over-consolidation ratio (OCR) and the initial stresses.
When specifying the over-consolidation ratio (OCR), consideration must be given
to the relationship between K nc , K 0 , OCR, and (Poissons ratio), where K nc
is the K 0 value when the soil was at its normal consolidation state. The
relationship between these variables follows the documentation by Britto and
Gunn (1987).
The value of K nc is related to the soil property by the equation:

K 0 = 1 sin
Two equations relating K nc , K 0 , OCR, and have been proposed by
Wroth (1975). The first equation is:

Page 172

SIGMA/W

Chapter 7: Analysis Types

K 0 = OCR K nc

v'
( OCR 1)
1 v '

and the second equation is:

3 (1 K
m
1 + 2K

) 3 (1 K 0 ) = ln OCR (1 + 2 K ) .

1 + 2K0

1 + 2K0

The value of m is an empirical constant that can be estimated from the plasticity
index (PI) using:
m

0. 022875 PI+1.22

For example, when ' = 30 , OCR =3, and v ' = 0.3 , K 0 must be 0.643.
The above equations are not used in SIGMA/W, but are presented here to help you
select an appropriate value for K 0 when you are establishing the current in-situ
stress state for a subsequent Cam-clay or Modified Cam-clay analysis.
In-situ stress example
Figure 7-3 shows a simple example we can use to illustrate how to spot check that
the correct initial in-situ stresses have been computed. The ground surface
elevation is 10 m. The surface of the free standing water is at EL. 12. The soil unit
weight is 20 kN/m3 and the water unit weight is 10 kN/m3 (rounded for
convenience here).
The effective stress at the ground surface should be zero. The total stress at the
ground surface should be equal to the weight of the free standing water (10 x 2 =
20 kPa). At the base of the problem, the effective stress v is (20 10 kN/m3) x 10
m = 100 kPa. Possions ratio is 0.334 and therefore Ko is 0.5
The effective horizontal stress is h = v Ko = 100 x 0.5 = 50 kPa
The total stresses are the effective stresses plus the pore-pressure which is 120 kPa
at the base of the problem: h = 50 + 120 = 170, v =100 + 120 = 220 kPa
The following graphs (Figure 7-4 and Figure 7-5) from SIGMA/W confirm and
correspond with these computed values. Note that the effective stresses at the
ground surface are zero and that the total stresses are 20 kPa.

Page 173

Chapter 7: Analysis Types

SIGMA/W

16
15
14
13
12
11
10
9
8
7
6
5
4
3
2
1
0

10

10

8
Vertical distance

Vertical distance

Figure 7-3 Insitu analysis example

6
4
2

6
4
2

0
0

20

40

60

80

100

Y-Effective Stress

10

20

30

40

X-Effective Stress

Figure 7-4 Effective stress profiles

Page 174

50

60

Chapter 7: Analysis Types

10

10

8
Vertical distance

Vertical distance

SIGMA/W

6
4
2

6
4
2

0
0

50

100

150

200

250

Y-Total Stress

50

100

150

200

X-Total Stress

Figure 7-5 Total stress profiles

Soil models and modulii for in-situ analyses


In any finite element displacement formulation such as SIGMA/W, it is necessary
to compute some strain to compute stress changes. The strains and deformation
computed as part of an in-situ analysis, however, have no meaning and should be
ignored. Since we are not interested in strains and deformations, the actual soil
stiffness properties are not all that important. Similar stress distributions can be
obtained for a range of elastic modulii
A second point is that contrasts in soil stiffness for different materials can create
stress concentrations. Such stress concentrations likely do not exist in the field. To
avoid such stress concentrations it is necessary sometimes to use artificial soil
modulii for the in-situ analysis.
You can likely get the most realistic in-situ stress distribution by treating all the
materials as linear elastic and by assigning all the materials the same elastic
modulus. In other words, treat the materials as uniform and homogeneous.
The best and most realistic in-situ stresses are usually obtained from an analysis of a
homogeneous section.

It is good practice to always spot check your in-situ stresses using the Graphing
function in CONTOUR. For example, you can plot total and effective, vertical (y)

Page 175

Chapter 7: Analysis Types

SIGMA/W

and horizontal (x) stresses along a vertical profile through your section. It is
particularly important to make sure you are getting realistic horizontal effective
stresses. This was illustrated in the example above.

7.3

Load / deformation analysis

The Load/Deformation Analysis type is used whenever you want to apply loads
and find the resulting stress changes and displacements, including simulation of fill
placement and excavation construction procedures. In the case of a fill placement
analysis, the weight of the fill is added to your model on the first load step that
each fill layer is activated. In the case of an excavation analysis, SIGMA/W
calculates the resultant forces associated with removing the excavated elements
and applies these forces as negative values at the nodes along the excavation face.
The amount of the excavated forces are written to file so you can check them in the
Contour program.
If you check Adjust Fill to Design Elevation, then, after each time step, SIGMA/W
will adjust the top surface of any filled elements (that have settled) back to their
original design elevation. This simulates the continued placement of fill (while the
fill settles) until the fill reaches the prescribed elevation. This adjustment option is
available only for Load/Deformation analysis and is disabled for other analysis
types. By default, the fill adjustment option is not selected. More details of fill and
excavation are given in a separate chapter on the subject.

7.4

Initial water table position

An initial water table is specified in two stages. First, the maximum negative
pressure head (i.e. the capillary rise) allowed in the initial condition is specified.
Secondly, the initial water table is drawn by clicking at each point in the initial
water table.
This feature is available for all analysis types in SIGMA/W. When you define an
initial water table, the initial pore-water pressure at each node is computed
proportionally to the vertical distance between the node and the defined water
table. The effect is that the pore-water pressure varies hydrostatically with distance
above and below the water table. Above the water table, the negative pore-water
pressure can be set to a limit to produce a pressure distribution such as shown in
Figure 7-6. The calculation of pressure at each point in the figure is given by:

Page 176

SIGMA/W

Chapter 7: Analysis Types

u a = ( y w - yr ) w
ub = ( yw - yb ) w
uc = ( yw - yc ) w
The pore-pressure at A is the specified maximum negative pressure head times the
unit weight of water.
The pore-pressure a B is the elevation of the water table minus the elevation of
Point B times the unit weight of water. This will be a negative number since
elevation B is higher than elevation of the water table.
The pore-pressure a C is the elevation of the water table minus the elevation of
Point C times the unit weight of water.

A
Max negative
pressure head

Figure 7-6 Calculation of pore water pressures using water tables

7.5

Specifying initial conditions using external file

A series of initial conditions can be defined at the start of an analysis or after restarting the current analysis. SIGMA/W will add the computed incremental
changes to the user-specified initial condition if any are specified. Consequently,
the output files always contain the sum of the initial conditions and the incremental
changes.
The initial conditions that can be specified are:

Page 177

Chapter 7: Analysis Types

SIGMA/W

Nodal deformation

Total stress state

Strain

Property (stop/restart conditions for the various soil models)

Pore-water pressure (PWP) or head

The type and objective of a particular analysis will determine which initial
conditions need to be defined. All or none of the initial conditions can be specified.
The initial conditions files must have the same number of nodes and elements as
the DEFINE data file, as well as the same number of integration points. In other
words, the initial conditions files must be compatible with the DEFINE data file
and previously computed results must be obtained from a compatible mesh.
For a linear-elastic analysis where the primary interest is settlement and no
excavation is involved, there is no need to open any initial conditions file. For
analyses using Cam-clay or Modified Cam-clay materials, it is necessary to define
the initial in-situ stress conditions, since the soil properties depend on the stress
state. For nonlinear elastic and the elastic-plastic materials, an initial stress
condition file is desirable, but not mandatory.
When the initial stress conditions have been established with an in-situ analysis,
the deformations and strains from such an analysis have no meaning.
Consequently, the initial deformation and strain should be zero, which is
accomplished by not opening an initial displacement or strain file.
Your can define the initial pore-water pressure conditions in two different ways:

use the Draw Initial Water Table Command in SIGMA/W DEFINE; or

use a pore-water pressure output file from a previous SIGMA/W,


SEEP/W, VADOSE/W or QUAKE/W analysis.

Page 178

SIGMA/W

Chapter 8: Structural Elements

Structural Elements

8.1

Introduction

This chapter presents the formulation used in SIGMA/W for structural elements. In
SIGMA/W, a structural element can be either a bar element, which is capable of
resisting axial force only, or a beam element, which is capable of resisting both
bending and axial force. Structural elements can be used only in a two-dimensional
plane strain load-deformation analysis.

8.2

Beam elements

In SIGMA/W, a beam element is formulated using the conventional (Bernoulli)


beam theory (Hinton and Owen, 1979). This beam element requires the slope as
well as the lateral displacement to be continuous within the element. Consequently,
cubic Hermitian shape functions are used. Each node associated with a beam
element is given a rotational degree-of-freedom in addition to the two
displacement degrees-of-freedom.
In version 6.0 of SIGMA/W, beam elements must be drawn along the edges of
existing, non null, soil elements. However, we are currently planning to have beam
elements be free objects and independent of soil geometry.
Interpolating functions for a beam element
Figure 8-1 shows that for a one-dimensional element with two nodes at r = 1
and r = 1 , the lateral displacement w can be expressed in terms of the nodal
lateral displacements, a , using the following:

da
da
w = N1a1 + N1 + N 2 a2 + N 2
dr 1
dr 2
where:

Ni , Ni

the Hermitian interpolating functions evaluated at node i

The n -th derivatives in the local r and global x-coordinate systems are related
by:

Page 179

Chapter 8: Structural Elements

SIGMA/W

d nw l d nw
= n
n
dr 2 dx

Equation 8-1

where:

the length of the beam element.

Figure 8-1 A 2-noded beam element

The Hermitian interpolation functions are tabulated as follows:


Table 8-1 Interpolating functions for beams
Interpolating Function

N1 = (2 + r )(1 r ) 2 4
N1 = (1 + r )(1 r ) 2 4
N 2 = (2 r )(1 + r ) 2 4
N 2 = (1 r )(1 + r ) 2 4
Stiffness matrix for a beam element
For a beam of flexural rigidity EI and length l , the strain energy b due to
bending is given by:
l

b =
0

EI d 2 w

dx
2 dx2

where:

Page 180

SIGMA/W

Chapter 8: Structural Elements

lateral displacement of the beam

distance along the beam

A beam element in SIGMA/W can be either 3-noded or 2-noded. However, a 3noded beam element is treated as a combination of two 2-noded elements. A 2noded element is formulated using cubic Hermitian interpolation functions. For
this element the lateral displacement can be expressed in terms of the Hermitian
interpolating functions as follows:
e

dw
dw
e
w = N1w + N1
+ N 2 w2 + N 2

dr 1
dr 2
e

e
1

The superscript e denotes that the variable is evaluated within an element.


Substituting in Equation 8-1, the previous equation can be written as:
e

l e dw
l e dw
e
w = N1w + N1
+ N 2 w2 + N 2
2 dx 1
2 dx 2
e

e
1

where:

le

length of the beam element.

Figure 8-2 Local and global coordinates for a structural element

Page 181

Chapter 8: Structural Elements

SIGMA/W

Again using Equation 8-1 and neglecting axial displacements, the curvature at any
point within the beam element can be expressed as:

d 2w
4
2 = e 2 0
dx l

d 2 N1
dr 2

d 2 N1 l e
dr 2 2

d 2 N2
dr 2

= [ B ]{a}

u1e

e
w1

d 2 N 2 l e ( dw dx )1

u2e
dr 2 2

w2e

e
( dw dx )2

where:
[B]

the strain matrix for the beam element,

{a}

vector of nodal field variables consisting of nodal displacements u, w and


nodal rotation dw/dx,

nodal axial displacement, and

local coordinate ( 1 r 1) .

The strain energy due to bending in a beam element now can be written as:
le

= {a} [ B ] C [ B ]{a} dx =
T

e
b

{a} [ B ]
T

C [ B ]{a}

le
dr
2

where:
C

EI, the flexural rigidity of the beam.

The stiffness matrix for the beam element [ K ] is given by:


e

[K ]

le
2

[ B ] C [ B ] dr
T

Page 182

SIGMA/W

Chapter 8: Structural Elements

The formulation is for a coordinate system ( x, y ) with the x-axis coinciding


with the longitudinal axis of the beam. This integral is evaluated numerically in
SIGMA/W. Before the stiffness matrix can be assembled into the global finite
element equation , it needs to be rotated into the global Cartesian coordinate
system through the following transformation:

[ K ]g = [T ] [ K ] [T ]
e

For a beam element inclining at angle to the global x-axis, the transformation
matrix is:

cos
sin

0
[T ] =
0
0

sin
cos
0
0
0
0

0
0
0
0
1
0
0 cos
0 sin
0
0

0
0
0
sin
cos
0

0
0
0

0
0

In order to define a beam in SIGMA/W it is necessary to specify the E modulus,


the cross-sectional area, the moment of inertia and the activation step. If the
moment of inertia is set to zero, a beam will behave as a bar element with axial
forces only.

8.3

Beam examples

Figure 8-3 shows two examples of the beam formulation applied in SIGMA/W.
The upper beam is fixed at the left end and on a roller at the right. A unit load has
been applied all along the top of the beam and the computed deflection is
illustrated. A closed form solution for this case was used to verify that the
SIGMA/W results are exact.
The lower beam in the figure is a cantilever beam with an applied moment
boundary condition on the right side. Again, the vertical displacement matches the
closed form solution.
In Version 6.0 of SIMGA/W it is necessary to define beams along the edges of
existing, non null, elements. In these examples, a single row of elements were

Page 183

Chapter 8: Structural Elements

SIGMA/W

specified and given almost zero strength. This way, the strength of the beam is
taking all the loading. In a future version of the software, it will be possible to
draw free beams that are not attached to two dimensional elements. This is possible
because the cross-sectional area of the beam is entered as a beam property and
does not need to be defined true to scale in the mesh.

Figure 8-3 Beam bending under constant load (upper), cantilever moment
load (lower)

Shear Force vs. Distance


0.0020
0.0015

Shear Force

0.0010
0.0005
0.0000
-0.0005
-0.0010
-0.0015
-0.0020
0

10

15

20

Distance

Figure 8-4 Shear force diagram for cantilever beam

Page 184

SIGMA/W

Chapter 8: Structural Elements

Figure 8-4 shows a SIGMA/W generated shear force diagram for the cantilever
beam example. In addition to shear force, it is possible to generate graphs of axial
force, moment, and rotation in any structural member.
Another example of an application of a structural beam is given in Figure 8-5
where a series of beams are tied back during excavation. Full details of how this
analysis is set up and carried out are given in the Illustrative Examples chapter. For
now, it is useful to see how a beam can be used.
Figure 8-6 shows the bending moment diagram for the beam in the tie-back
example. It is clear from this figure that the additional anchoring of the beam by
bar elements is helping minimize the magnitude of the bending moment within the
beam.

20
18

Elevation - metres

16
14
12
10
8
6
4
2
0
0

10

15

20

25

30

35

Distance - metres

Figure 8-5 Tie back of structural beam during excavation

Page 185

40

Chapter 8: Structural Elements

SIGMA/W

Moment vs. Distance


150

Moment

100

50

-50
0

10

15

Distance

Figure 8-6 Moment diagram for tie back beam example

8.4

Bar elements

In SIGMA/W, a structural element defined with a zero flexural rigidity (EI) is


considered to be a bar element. A bar element is capable of resisting axial forces
only. Thus, for the nodes of a bar element, rotational degree-of-freedom is not
required. A bar element is not restricted to being drawn along a soil element edge
so it can be drawn between any two mesh nodes and its stiffness is subsequently
only applied at the points of contact with the nodes.
Interpolating functions for a bar element
In SIGMA/W, a bar element can be either 2-noded or 3-noded. The node ordering
is the same as in a beam element. Interpolating functions similar to those used for
two-dimensional solid elements are used in formulating a bar element. The
interpolating functions in terms of local coordinate r (1 r 1) are listed in the
following table.

Page 186

SIGMA/W

Chapter 8: Structural Elements

Table 8-2 Interpolating functions for bars


Function

Include in function if node 3


is present

1
N1 = (1 r )
2

1
(1 r 2 )
2

1
(1 + r )
2

1
(1 r 2 )
2

N2 =

N3 = (1 r 2 )

___

Stiffness matrix for a bar element


For a bar of axial rigidity EA and length l, the strain energy a due to axial
deformation is given by:
l

a =
0

EA du

dx
2 dx

where:
u

axial displacement along the beam, and

distance along the beam.

In a bar element, axial displacements can be expressed in terms of the nodal


displacement using the interpolating functions:
n

u e = N i uie
i =1

where:

2 or 3, the number of nodes used in the bar element.

The superscript e denotes that the variable is evaluated within an element.

Page 187

Chapter 8: Structural Elements

SIGMA/W

The derivative e can also be expressed in terms of the derivatives of the


interpolating functions as follows:
e

n
dN i e
du
=
ui
dx
i =1 dx

dN1
dx

dN1
dx

dN 2 u1e
, for a 2-noded element
dx u2e
u1e
dN 2 dN 3 e
u2 , for a 3-noded element
dx dx e
u3

where:
{n}

vector of nodal axial displacements, and

2 or 3, the number of nodes used in the bar element.

The derivative of the interpolating function with respect to x at node i can be


evaluated using:
e

dN i dr dN i
=
dx dx dr

1 dN i
=J

dr

where:
r

local coordinate ( 1 r 1)

Jacobian operator = dr dx

After substituting, the expression for the derivative du dx becomes:

Page 188

SIGMA/W

Chapter 8: Structural Elements

du
1 dN1
= J
dr
dx
= J 1

dN1
dr

e
dN1 u1
, for a 2-noded element
dr u2e

dN 2
dr

dN 3
dr

= J 1 [ B ]{a}

u1e
e
u2 , for a 3-noded element
u e
3

where:
[B]

strain matrix for a bar element

{a}

vector of nodal axial displacements

The following expression for the strain energy due to axial deformation for a bar
element is obtained:
le

= {a} [ B ] C [ B ]{a} dx =
T

e
a

{a} [ B ] C [ B ]{a} J
T

dr

where:
C

EA, the axial stiffness of the bar element.

The stiffness matrix for the bar element [ K ] is given by:


e

[K ] =
e

[ B] C [ B] J
T

dr

The formulation is for a coordinate system ( x, y ) with x coinciding with the


longitudinal axis of the bar. This integral is evaluated numerically in SIGMA/W.
Before this stiffness matrix can be assembled into the global finite element
equation , it needs to be rotated into the global Cartesian coordinate system in a
procedure similar to that for a beam element.

Page 189

Chapter 8: Structural Elements

SIGMA/W

Bar elements can be set to become active at different load steps. If the active step
is defined as zero, the bar element is considered to be an active element from step
1. If the active step is nonzero, the bar element is considered to be an active
element from that active step.
Enter tension axial force in a bar element as a negative value and compression
force as a positive value. The reaction forces at the two end nodes will be shown as
inward arrows for tension axial force and outward arrows for compression axial
force those are the nodal forces that will be actually applied on the rest of the mesh
at the step defined. A pre-axial force can be applied before activating a bar
element.

8.5

Bar example

Figure 8-7 shows a structure composed of bar elements prior to and after loading.
Notice that the bars do not have to be drawn relative to any existing soil elements.
By definition, a bar can only support an axial load, so it can be drawn anywhere in
the domain as long as it is connected to valid nodes.
Another example of an application of a bar was given previously in the tie back
example (see Figure 8-5 above). In that example, the bar was use to tie back a
structural beam into the soil to help minimize bending moments within the beam
during excavation.
In order to define a bar, it is necessary to know the E modulus and the crosssectional area as well as any applied pre-axial force. Finally, you can also specify
what loading step in the analysis the bar becomes active. This is useful in a braced
excavation analysis, for example, because the bar does not become installed until
after the excavation has moved below the intended bar support elevation.
Another example of the use of a bar is given in the braced excavation file
discussed in more detail in the Illustrative Examples chapter. Figure 8-8 shows the
application of a strut (bar member) during excavation. Notice that the entire
excavation is included in the model and not just the left or right half. This is
necessary because the at this time, it is not possible to draw the fixed boundary
condition on the mid point of the bar while it is adjacent to a null element (as
would be the case if only half the geometry were modeled).

Page 190

SIGMA/W

Chapter 8: Structural Elements

90 deg.

metres

20 kN

30 kN
-1
-1

10

distance - metres

Figure 8-7 Bar structure before / after loading

Excavation with a Horizontal Strut


Strut

Figure 8-8 Application of a strut (bar) during open excavation

Page 191

11

Chapter 8: Structural Elements

SIGMA/W

Page 192

SIGMA/W

Chapter 9: Fill and Excavation

Fill and Excavation

Many classes of stress-deformation problems involve adding or removing soil.


Adding soil involves adding mesh elements that have a self weight (body load)
plus apply external load to the soils below. Removing soil involves unloading
stress at the excavation face. In either case, it is important to think about what the
model is experiencing.
Another issue dealing with excavation is how to integrate structural elements that
serve to support near vertical faces and prevent excessive deformation or collapse.
While SIGMA/W is not the model to use to predict what collapse may look like, it
is a powerful design tool to use in order to place adequate support such that
collapse is never imminent.
This chapter discusses some of these issues.

9.1

Simulating fill placement

During a sequential fill placement simulation, forces due to the self-weight of each
lift are applied to the finite element grid as the lift is being placed. The forces are
computed by knowing the unit weight and the volume of each fill element. This
total force is then divided up and applied as a negative Y-force at each element
node. The advantage of doing this is that the fill element not only applies a load to
the soil below, it has some self-compression as well.
When a fill element becomes active, it must have the body load for its material
specified. In other types of analysis it is important to ensure that the body load is
only applied on one load step and it is your responsibility to make sure you do not
apply it in both an in-situ analysis and any subsequent analyses. Because you will
generally do a fill analysis over several load steps, it is not convenient to stop the
analysis just to turn off the body load force. SIGMA/W will recognize the special
nature of fill elements and only apply the body load on the first step that the fill
element becomes active.
SIGMA/W can optionally compensate for the settlement caused by self
compression of newly added fill elements. For each lift, SIGMA/W adds the
current displacement to the element nodal coordinates to compute the element
weight and volume, which in turn are used to compute the gravitational nodal
forces to simulate the fill placement. Adding the displacements gives the element

Page 193

Chapter 9: Fill and Excavation

SIGMA/W

additional mass which represents the additional fill placed to compensate for
settlement during construction.
Consider the Embankment example in Figure 9-1 which shows the final deformed
mesh after placing six lifts. Note that the final crest is close to the design elevation
while there is considerable settlement at the foundation-embankment contact.
40
35

Height (metres)

30
25
20
15
10
5
0
0

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Distance (metres)

Figure 9-1 Deformed mesh after placing six lifts

This effect is further illustrated by Figure 9-2 which shows the vertical settlement
along the dam and foundation centerline. The maximum settlement is near the base
of the dam, as it should be.
The settlement at the crest is 0.0647m from the placement of the last lift. This can
be minimized by making the last lift small.
Figure 9-3 shows a one-step analysis with gravity turned on. The crest now has
settled significantly. By comparing the two kinds of analyses, you can get an
estimate of the amount of fill that was placed to compensate for the settlement.
This example illustrates another important point. If entire embankment is applied
as a single load, then the maximum displacement is at the top of the dam, which is
not correct. In reality, the maximum vertical displacement is at the base of the dam
which is modeled correctly if several lifts are used to simulate construction.

Page 194

SIGMA/W

Chapter 9: Fill and Excavation

Y-Displacement vs. Y
40

30

20

10

0
-0.30

-0.25

-0.20

-0.15

-0.10

-0.05

0.00

Y-Displacement

Figure 9-2 Vertical settlement along vertical mid point of dam

35
30
25
20
15
10
5
0
0

10

20

30

40

50

60

70

80

90

100

Figure 9-3 Deformed mesh after a single load step embankment addition

Page 195

Chapter 9: Fill and Excavation

9.2

SIGMA/W

Excavation elements

Excavation elements work in a similar way to fill elements in that they can be
turned off at various load steps to simulate the process of digging an excavation
from the top down (or in the case of a tunnel, from the center to the liner). When
an element is removed, a force that pulls outward on the remaining soil must be
applied. Unlike the fill element where the forces are known by the weight of the
soil being added, an excavation element uses the existing total stresses at the
element edges adjacent to the elements being removed. In this way, the actual total
stresses acting in the x- and y-directions can be used to arrive at a nodal force load
for the element edge nodes.
Some caution should be used when doing an excavation analysis where pore-water
pressures are included prior to excavating. In general, when an excavation is made
it is assumed that the any water in the excavation will be pumped out. This means
that on the soil side of the excavation the total stress at the excavation face is
comprised of the effective stress and the pore-water pressure. When an excavation
element is removed, the total stress existing prior to removal of the element is used
to determine the unloading forces. If the total stresses are correct prior to
excavation (such that they already account for water pressures), then the
excavation analysis should be carried out without defining a water table or
importing pore-water pressure data from another source.
If the pre-excavation analysis did not account for any pore-water pressures, it
would still be wrong to then apply pore-water pressures simultaneously with
excavation. The reason for this is the total horizontal stress on the excavation face
will not be the same in both cases. Consider the following example where we wish
to know the excavation forces at a depth of 10m with the water table assumed to be
at surface, the unit weight of soil = 20, and K0 = 0.5.
For the following analysis we will use the total stress equation:

x = ( y u ) K 0 + u
where total stress equals effective stress plus pore-water pressure.
Case 1: pore-water pressure applied in pre-excavation analyses
In this case, the total stress is computed from the previous analysis step to
be:

Page 196

SIGMA/W

Chapter 9: Fill and Excavation

x = (10* 20 100 ) 0.5 + 100 = 150 .


Case 2: pore-water pressure applied during excavation analyses
In this case, the total stress equals the effective stress from the previous
analysis plus the current pore-water pressure and is computed to be:

x = (10* 20 0 ) 0.5 + 0 = 100 plus 100 = 200


In the second case, the unloading forces are 33% greater which is not correct.
This example illustrates two things: first, it is wrong to apply the pore-water
pressure during the excavation step; and secondly, it is always a good idea to doe
some simple hand calculations to check your thinking.

9.3

Specifying fill / excavation elements

In general, this engineering book is not about how to use the software commands.
However, in this case, it is useful to see how fill and excavation elements are
designated and how they look once specified.
The Draw Fill/Excavation Elements command allows you to specify the time step
at which to add or remove selected elements to simulate fill placement and
excavation construction. For example, Figure 9-4 illustrates layers of elements
that are added incrementally at load times 1, 2, and 3 respectively:
+3

+3

+3

+3

+3

+3

+3

+3

+3

+2

+2

+2

+2

+2

+2

+2

+2

+2

+2

+2

+2

+2

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+3
+3
+2
+2
+1
+1

+1

Figure 9-4 Specification of fill elements

Figure 9-5 shows part of an excavation problem in which elements are excavated
(removed) incrementally at load times 1 and 2. Elements with time 0 are
permanent elements which are present throughout an analysis.

Page 197

Chapter 9: Fill and Excavation

SIGMA/W

0
0 -1 -1
0 -1 -1
-1

-1 -10 0
-1 -1 -10

-1 -1

-1

-1

-1 -1

-2 -2

-2

-2

-2 -2

-2
0 -2 -2
0 -2 -2
0

-2

-2 -20
-2 -20 0
0

Figure 9-5 Specification of excavation elements

Fill elements are added to the problem at the selected time step, and do not exist
before this time step; excavation elements are removed from the problem at the
selected time step, and do not exist after this time step; permanent elements exist
throughout the entire problem, and therefore do not require a selected time step.
NOTE: An initial stress file is required when performing an excavation analysis. Without
any prior stress conditions, it would not be possible to compute the unloading forces caused
by removal of stress. In addition, excavation analysis should use correctly computed total
stress conditions which means do not apply the pore-water pressure load at the same time
as the excavation loads.

9.4

Braced excavations

Often it is necessary to use some mechanical device to support an open excavation.


This may include piling, bracing or a combination of both. One typical approach is
to use a bar structural member to prop open the excavation. SIGMA/W can
accommodate this in two ways.
Full details of beam and bar elements are discussed in a previous chapter, therefore
no additional details are given here. What is provided is a discussion of the
similarity between using an actual bar element, or applying a spring boundary
condition as a force on the face of the excavation.
The spring constant of a strut in a braced excavation can be estimated from the
relationship,

Page 198

SIGMA/W

Ks =

Chapter 9: Fill and Excavation

F EA
=
l
L

where:
F

force on the member,

change in length,

Young's Modulus,

cross-sectional area of the member, and

length of the member.

Some displacement is usually required before a strut is fully seated and acts like a
pure spring. You can simulate this by defining a spring boundary function using
the Boundary command in DEFINE. The spring function must be specified in
terms of the potential movement.
NOTE: The actual force acting at the node is given in the displacement file and can
be displayed in CONTOUR using the View Node Information command and
checking the Boundary Forces. You can use this force to check or design the size
of the strut. Furthermore, the force presented in the file is per unit width of wall (or
per unit radian in the case of an axisymmetric analysis).
A better alternative to using the spring node boundary condition is to use a bar
element. The bar has the advantage in that its stiffness is a function of its length
and the length is included as an actual value because it is drawn as part of the
problem geometry. In addition, the bar can be set to become active at different time
steps so you can define all the bars ahead of time and they will only become active
as the excavation passes their installation elevation.
The chapter on Illustrative Examples includes a model that uses structural supports
in an excavation analysis.

Page 199

Chapter 9: Fill and Excavation

SIGMA/W

Page 200

SIGMA/W

10

Chapter 10: Consolidation

Consolidation

SIGMA/W is formulated to solve soil consolidation problems using a fully coupled


or any of several un-coupled options. A fully coupled analysis requires that both
the stress-deformation and seepage dissipation equations be solved simultaneously.
The theoretical formulation for this rigorous method along with some practical
limitations are discussed below. With the coupled analysis, it is necessary to have
SEEP/W as well as SIGMA/W.
When coupled, both SIGMA/W and SEEP/W contribute to forming a common
global characteristic (stiffness) matrix. Three equations are created for each node
in the finite element mesh. Two are equilibrium (displacement) equations formed
by SIGMA/W and the third is a continuity (flow) equation formed by SEEP/W.
Solving all three equations simultaneously gives both displacement and pore-water
pressure changes
Of perhaps more use in many real situations is the un-coupled consolidation
formulation in SIGMA/W. In this analysis, the seepage analysis is solved
independently of the volume change analysis. The incremental change in porewater pressures from the seepage solutions are used at each load step in the stressdeformation calculation in order to determine the change in effective stresses. With
the un-coupled analysis, the changes in pore-water pressures can be obtained from
any other SIGMA/W, SEEP/W, VADOSE/W or QUAKE/W analysis and for any
two or multiple time steps in those analyses.
Details of both types of consolidation analyses are discussed below along with
several examples to illustrate the use and interpretation of this powerful option.
The first part of this chapter discusses general theoretical issues related to
consolidation and the fully coupled formulation. With this developed, it is more
obvious what is happening when the formulation is un-coupled.

10.1

Constitutive equation for soil structure

The incremental strain-stress relationship for an unsaturated soil medium can be


written as follows (Fredlund and Rahardjo, 1993):

Page 201

Chapter 10: Consolidation

x
1


x

x 1

xy

E 0
yz
0

zx
0

SIGMA/W

0
0
0

1
1

1
1
+
0
H

( x ua )

u
0
0
0

( y a )
( z ua )
1
0
0
0

0
2(1 + )
0
0 xy

0
0
2(1 + )
0 yz

0
0
0
2(1 + )

zx

( ua uw )
( u u )
a
w

( ua uw )

( ua uw )
( ua uw )
0

0
( ua uw )

where:

normal strain,

engineering shear strain,

normal stress,

shear stress,

ua

pore-air pressure,

uw

pore-water pressure,

elastic modulus for soil structure,

unsaturated soil modulus for soil structure with respect to matrix suction
(ua - uw), and

Poissons ratio.

This equation is similar in form to the constitutive equation presented by Biot


(1941).

Page 202

SIGMA/W

Chapter 10: Consolidation

This relationship can be re-written in the following stress-strain form for a twodimensional space.

1 0 0
( x ua )

1 0
E (1 )
( y u a )

=
( z ua ) (1 + )(1 2 )

xy


u u
x a w

H
0

0
ua u w
y

H
0

1 2
u u
z a w
2 (1 + )
H

xy

Alternatively, this incremental stress-strain relationship can be written as:

[ ]{ } [ D ]{mH }( ua uw ) + {ua }

Equation 10-1 { } = D

where:
[D]

{mH }

drained constitutive matrix

1 1 1
0
H H H

If it can be further assumed that air pressure remains atmospheric at all times,
Equation 10-1 becomes:

{ } = [ D ]{ } + [ D ]{mH } uw
On the other hand, for a soil element which is fully saturated, the total stress on the
soil structure is given by:

{ } = {D}{ } + {m} uw
where:
{m}

unit isotropic tensor, < 1 1 1 0 >.

Page 203

Chapter 10: Consolidation

SIGMA/W

Comparing these last two equations, it can be seen that, when the soil is fully
saturated (i.e. S = 100%):

[ D ]{mH } = {m}
For a linearly elastic material, this condition is satisfied when:

E
H =

1 2v
10.2

Flow equation for water phase

The two-dimensional flow of pore-water through an elemental volume of soil is


given by Darcys equation:

k x 2uw k y 2uw w
+
+
=0
w x2 w y2 t
where:
kx, ky

the hydraulic conductivity in x and y direction, respectively,

uw

seepage velocity,

the unit weight of water,

the volumetric water content, and

time.

The volumetric water content for an elastic material is given by the following
expressions (Dakshanamurthy et al. , 1984):
Equation 10-2

w =

v uw

and:

Page 204

SIGMA/W

Chapter 10: Consolidation

E
1
, and,
H (1 2 )
Equation 10-3
1
=
R H

where:
KB

bulk modulus,

a modulus relating the change in volumetric water content with change in


matric suction.

Since a soil-water characteristic curve is a graph showing the change of volumetric


water content corresponding a change in matrix suction, (ua - uw), the parameter R
can be obtained from the inverse of the slope of the soil-water characteristic curve.
Assuming material properties remain unchanged during an increment, Equation
10-2 can be written in the following incremental form.
Equation 10-4 w = v uw

At full saturation, the change in volumetric water content, w, is equal to the


change in volumetric strain, v. This condition is satisfied in Equation 10-3 when
is equal to zero.

10.3

Finite element formulation for coupled analysis

In a coupled consolidation analysis, both equilibrium and flow equations are


solved simultaneously. When uncoupled, the field variables in SIGMA/W are the
incremental displacements at nodes, and in SEEP/W the field variables are the
hydraulic heads at nodes.
In SIGMA/W, the finite element equilibrium equations are formulated using the
principle of virtual work, which states that for a system in equilibrium, the total
internal virtual work is equal to the external virtual work. In the simple case when
only external point loads {F} are applied, the virtual work equation can be written
as:

{ } { } dV = { } {F } dV
*

Page 205

Chapter 10: Consolidation

SIGMA/W

where:
{*}

virtual displacements,

{*}

virtual strains, and

{}

internal stresses.

Substituting Equation 10-4 into this equation and applying numerical integration, it
can be shown that the finite element equations that SIGMA/W solves are given by:

[ B ] [ D ][ B ]{ } + [ B ] [ D ]{m }
T

{uw } = F ,

[ K ] = [ B ] [ D ][ B ] ,
T

[ Ld ] = [ B ] [ D ]{mH }
T

{mH }

1
H

1
H

N , and

1
H

where:
[B]

gradient matrix, (also called the strain matrix),

[D]

drained constitutive matrix,

[K]

stiffness matrix,

[Ld]

coupling matrix,

{}

incremental displacement vector, and

uw

incremental pore-water pressure vector.

For a fully saturated soil, the coupling matrix, [Ld], can be written as:

[ Ld ] = [ B ] {m}
T

N ,

with {m} = <1, 1, 1, 0>


T

The flow equation can similarly be formulated for finite element analysis using the
principle of virtual work in terms of pore-water pressure and volumetric strains. If
virtual pore-water pressures, u*w, are applied to the Flow Equation and integrated
over the volume, the following virtual work equation can be obtained.

Page 206

SIGMA/W

Chapter 10: Consolidation

*
k y 2uw w
* k x uw
u
+
w w x 2 w y 2 + t dV = 0

Applying integration by parts to this equation gives:

k uw* uw k y uw* uw
* w
dV = uw* vn dA
x
+
dV + uw
x
x
y
y
t

w
w

where:
Vn

boundary flux.

Substituting in the expression for the volumetric water content, w, gives:

k x uw* uw k y uw* uw
* ( v u w )
dV = uw* vn dA

+
dV + uw
t
w x x w y y
Using finite element approximations, this equation can be written as:

[ B ] [ K w ][ B ]{uw } dV
T

( )
dV
t

{m} [ B ]
T

(uw )
N
+
t

= N vn dV
T

where:

K f = [ B ] [ K w ][ B ] dV ,
T

[M N ] =

L f = N

N , and
T

{m} [ B ] dV
T

with:
[B]

gradient matrix,

[Kw]

hydraulic conductivity matrix,

Page 207

Chapter 10: Consolidation

SIGMA/W

[Kf]

element stiffness matrix,

<N>

row vector of shape functions,

[MN]

mass matrix,

[Lf]

coupling matrix for flow,

{m}

isotropic unit tensor, < 1 1 1 0 >, and

nodal displacement.

Integrating this equation from time t to time t + t gives:


t +t

K {u } dt
w f w

t +t
(uw )
( )
L f
dt
+

dt

t
t
t

t +t

[M N ]
t

t +t

vn dAdt

Applying the time differencing technique using as the time stepping factor to this
equation, the following finite element equation is obtained.

( K

{uw }

L f { }

t +t

t +t
t

+ (1 ) K f {uw }

= t N

( v

n t +t

) [M

] (uw ) tt +t +

+ (1 ) vn

) dA

When the backward (fully implicit) time-stepping scheme is used (by setting = 1)
and assuming that and remain constant within a time increment, this equation
becomes:

K {u }
w f w

t +t

[ M N ]{uw } + L f { } = t {Q}

where:
{Q}

the flow at boundary nodes.

Page 208

t +t

SIGMA/W

Chapter 10: Consolidation

In order to obtain an equation involving an incremental pore-water pressure only,


the first term,

K {u } , is added to both sides of the equation. The


w f w t

resultant equation describing the flow of pore-water is:

K f + [ M N ] {uw } = t {Q}
w

L f { }

t +t

K f {uw } t
w

A coupled consolidation analysis for saturated/unsaturated soils is thus formulated


using incremental displacement and incremental pore-water pressure as field
variables.
In summary, the coupled equations for finite element analysis are rewritten in the
following form.

[ K ]{ } + [ Ld ]{uw } = {F } , and
t

K f + [ M N ] {uw } = t {Q}
w

L f { }
where:

[ K ] = [ B ] [ D ][ B ]
T
[ Ld ] = [ B ] [ D ]{mH }
T

{mH } =

1
H

1
H

1
H

N
0

K f = [ B ] [ K w ][ B ]
T

[M N ] =

L f = N

{m} [ B ]

Also, as given by Equation 10-3:

Page 209

t +t

K f {uw } t
w

Chapter 10: Consolidation

E
1
3K B
=
H (1 2v )
H

1 3

R H

SIGMA/W

In order for these equations to model the fully saturated case, the following
conditions must be satisfied:

1,

0, and

[Lf]

[Ld]T

10.4

Additional material properties for unsaturated coupled


analysis

For a coupled analysis involving unsaturated soils, two additional material


properties H and R need to be defined. H is a modulus relating to the change of
volumetric strain in the soil structure to a change in suction. R is another modulus
relating the change in volumetric water content to suction; therefore, it is given by
the inverse of the slope of the soil water characteristic curve.
In this section, a procedure to obtain the H modulus parameter from the slope of a
void ratio (e) versus matric suction (ua - uw) curve is described. For a soil element,
a change in its volume can be decomposed into two parts:

dV = dVs + dVv
where:
dVs

the change in volume of the soil particles, and

dVv

the change in the volume of voids.

If the volume change of the soil particles, dVs, is small and thus neglected, the
volumetric strain can be approximated as follows.

v =

dV dVv

V
V

Page 210

SIGMA/W

Chapter 10: Consolidation

From the definition of void ratio, e, a change in void ratio, de, is given by:

V dV
dVv
d v
=
de = d v = v =
(1 n )V (1 n )
Vs Vs
where:
n

the porosity of the soil.

The slope of a void ratio versus matrix suction curve can be written as:
Equation 10-5

d v
de
=
d ( ua uw ) (1 n ) d ( ua uw )

In an unsaturated soil element, when only a change in matric suction occurs, the
incremental volumetric strain, dv, can be written as:

d v = d x + d y + d z =

3d ( ua uw )
H

or:

d v
3
=
d ( ua u w ) H
After substituting this into Equation 10-5, it can be seen that the slope of a void
ratio versus matric suction curve is:

3
(1 n ) H

10.5

Coupled analysis practical issues

You can do a consolidation analysis by coupling SIGMA/W and SEEP/W. In a


consolidation analysis, we are interested in how displacement and pore-water
pressure change simultaneously. SIGMA/W computes displacements and stresses
while SEEP/W computes the changes in pore-water pressure with time.

Page 211

Chapter 10: Consolidation

SIGMA/W

When doing a coupled analysis, it is essential to recognize that all equilibrium


(force and displacement) conditions are defined in SIGMA/W and all hydraulic
(flow) conditions are specified in SEEP/W. In SIGMA/W, you have to specify the
usual force and displacement boundary conditions together with soil properties
defined using effective stress parameters. In SEEP/W, you have to specify the head
and flow boundary conditions together with hydraulic conductivity and volumetric
water content functions. In essence, you have to set up both a complete SIGMA/W
problem and a partial complementary SEEP/W problem. In SEEP/W you will have
to define soil hydraulic properties and boundary conditions.
When undertaking a fully coupled analysis, you have to be particularly aware of
the following:
SIGMA/W is the master in a coupled analysis while SEEP/W is the slave.
SIGMA/W controls the processing.
The time sequence defined in SIGMA/W controls the time increments. Any time
sequence defined in SEEP/W is ignored.
The processing can only be started, stopped or halted in the SIGMA/W SOLVE
window. The Start, Stop and Stop Iteration, buttons in SEEP/W are disabled.
You can only select a starting time step using the Set Start Time command in
SIGMA/W SOLVE; you cannot use the Set Start Time command in SEEP/W
SOLVE.
Even for a fully saturated consolidation analysis, it is mandatory to define a
volumetric water content function for each soil type. The function may be defined
only for positive pore-water pressures and not for negative pore-water pressures,
but the function must exist nonetheless.
Convergence criteria must be specified independently in SIGMA/W and SEEP/W.
Convergence is only achieved when the convergence criteria are satisfied in both
SIGMA/W and SEEP/W.
You can view the results of a consolidation analysis using SIGMA/W CONTOUR
and SEEP/W CONTOUR, as you would for any uncoupled analysis.

Page 212

SIGMA/W

10.6

Chapter 10: Consolidation

SIGMA/W material models in an unsaturated coupled


consolidation analysis

Only Linear-elastic and Nonlinear-elastic (Hyperbolic) material models can be


used in an unsaturated consolidation analysis. The Elastic-plastic and Cam-clay
material models are not formulated in SIGMA/W for use in unsaturated soil
consolidation.
For elements using the Elastic-plastic and Cam-clay models, SIGMA/W limits
(beta) to 1. 0 and (omega) to 0. 0 during a consolidation analysis. This means
that the pore-water pressure change is equal to the volumetric strain, which is
incorrect for unsaturated soils. Consequently, these two models should not be used
in regions where the soil is unsaturated. For saturated consolidation analyses, you
can use all material models, including Linear-elastic and Cam-clay.
NOTE: When attempting any type of consolidation analysis, you should always
begin by modelling the problem using Linear-elastic materials. Once you are fully
confident in the validity of the results from the Linear-elastic consolidation
analysis, you can then attempt to use other material models. This is applicable even
for saturated consolidation analyses. Selecting the appropriate soil properties and
boundary conditions for a consolidation analysis is a complex process; therefore,
you should be confident this has been done correctly before you add the
complexity of nonlinear material models.
H-Modulus function in a consolidation analysis

The H is the unsaturated modulus that relates the volumetric strain of the soil to a
change in negative pore-water pressure or change in suction. The H-modulus may
be defined as a function of negative pore-water pressure. At saturation, H is related
to the elastic constants E and v by the equation:

E
H =

1 2v
Therefore, H must be set to E / (1 2v ) at zero pore-water pressure when defining
an H-Modulus vs. pore-water pressure function. As a soil dries and the pore-water
pressure becomes highly negative, the soil becomes very stiff. This increase in
stiffness can be represented by an increase in H. Figure 10-1 illustrates a potential
increase in H as a function of the negative pore-water pressure.

Page 213

Chapter 10: Consolidation

SIGMA/W

The H-Modulus cannot be specified less than E / (1 2v ) . If you define an


H-Modulus function with an H value less than E / (1 2v ) , SIGMA/W will
automatically set H to E / (1 2v ) during the analysis. Consequently, when you
define an H-Modulus function, the lowest H value should be E / (1 2v ) at the
point where the pore-water pressure is zero.

Figure 10-1 H-modulus as a function of pore-water pressure

Since H is a relatively unfamiliar parameter, you should start a consolidation


analysis by not defining this function. Once you are satisfied with the
consolidation results, you may wish to add the H-Modulus function to see what
effect it has on your results. If you are doing a fully saturated consolidation
analysis, you do not need to define the H-Modulus function, since H is always set
to E / (1 2v ) under saturated conditions.
It is not essential to define an H-Modulus function. In fact, it is best not to define this
function if you are uncertain about the value of the H-Modulus. SIGMA/W will
automatically set H to E / (1 2v ) if the H-Modulus function is not defined.

Page 214

SIGMA/W

10.7

Chapter 10: Consolidation

Saturated zone only coupled analysis

The formulation for saturated / unsaturated fully coupled consolidation has been
applied to a one dimensional model where results matched with expectations. This
model is discussed towards the end of this chapter. The application of fully
coupled saturated/unsaturated analysis in two dimensions is questionable because
of our inability to fully understand and define the relationship between the slope of
the soil water characteristic curve (Mv) and the soil stiffness (E) during variably
saturated conditions.
There are times, however, where you can successfully use a fully coupled analysis
in two dimensions. SIGMA/W has a powerful feature to only apply the fully
coupled formulation to the saturated zone within a model geometry that contains
both saturated and unsaturated zones. The assumption inherent in consolidating
within the saturated zones only is that there is sufficient water storage capacity
within the surrounding unsaturated zones to take in any dissipated water from
consolidating saturated zones without undergoing volume change. This is a
reasonable assumption in many cases.
If you choose the saturated zone consolidation analysis, SIGMA/W will solve the
fully coupled equations for all nodes in the mesh, however, it will fix the porewater pressures in the unsaturated zones to be equal to their value at the start of the
analysis. In other words, there will be no incremental change in pressure
throughout the analysis. The pore-water pressure in the saturated zones will be
dissipated according to the permeability defined in the seepage portion of the
analysis. This means that the rate of dissipation is likely controlled by the
permeability value associated with the unsaturated soils surrounding any saturated
zones experiencing consolidation. The unsaturated zones likely have a lower K
value than the saturated conductivity so they will control dissipation of pressures.
When this type of analysis is chosen, it is important to define the volumetric water
content function in the SEEP/W program. At a minimum, it is necessary to define
the Mv value of the function. This is necessary because SIGMA/W will set the
Poissons ratio to 1/3 and then compute the elastic modulus, E, from the Mv value
as follows:

1 3
E
H =
for a saturated soil (from Equation 10-2),
combined with =
R H
1 2v
results in

Page 215

Chapter 10: Consolidation

E=R=

SIGMA/W

1
Mv

This is a valid approach because for a saturated soil, the change in water content is
equal to the change in volumetric strain. The Poissons ratio is set to be 1/3 which
validates the above assumptions for a linear-elastically behaved soil. This
combination of parameters will ensure that there is compatibility between
hydraulic and deformation moduli in the saturated only analysis; it conforms to the
theoretical coupled formulation in this zone; and it means the analysis can be
carried out without a full definition of the volumetric water content function (e.g.,
only the Mv is required).

10.8

Time step size in a coupled consolidation analysis

In formulating the finite element equations for coupled consolidation, the


backward difference scheme is used for time integration. Consequently, the
solution is unconditionally stable for any size of time steps (Smith, 1971).
However, numerical oscillation may still occur. To eliminate this oscillation, Britto
and Gunn (1987) have provided a guideline to the minimum time step size as given
by the following equation:

tmin

l2
l 2 w (1 + )(1 2 )
=
12cv 12 k
E (1 )

where:
l

distance between nodes, and

permeability or hydraulic conductivity.

You can use this formula at a couple of typical locations in your problem to help
you select appropriate time step sizes. In most cases, the best approach is to apply
the formula to areas where you expect the largest hydraulic gradients.
Another method of computing time step sizes is to do a simple, hand-calculated 1D consolidation analysis using text-book, closed form solutions. A quick, rough
estimate of the time required to reach, for example, 10% and 50% consolidation
can help a lot in selecting a time-increment sequence for a SIGMA/W SEEP/W
consolidation analysis.

Page 216

SIGMA/W

Chapter 10: Consolidation

Time step sizes can be too small or too large. If the time step sizes are too small,
your solution will be mostly determined by numerical noise, making the results
look totally unrealistic. If the time step sizes are too large, the computed
deformation can be in considerable error.
Time step sizes can vary over a fairly wide range, but they must be reasonable.
Unfortunately, an appropriate time-step sequence is problem dependent;
consequently, several trial runs must be made in the early stages of analyzing a
new problem.

10.9

Uncoupled consolidation analysis

The validity of fully coupled saturated / unsaturated consolidation analyses applied


in two dimensions is questionable because of uncertainty in defining and
measuring the relationship between water storage and elasticity. The one
dimensional formulation appears to be valid, but in this case there is no lateral
stress redistribution and the stress is the same in all elements (e.g. in a triaxial cell).
Obtaining this level of confidence in two dimensional analyses has not yet been
demonstrated in the literature. We allow the two dimensional fully coupled
analysis to be used in SIGMA/W because it is hoped that current researchers can
use it to learn more about the formulation and the material properties required to
make it a useful tool.
The fully coupled formulation can work quite well in two dimensions for both
saturated and unsaturated cases if the change in pore-water pressure at each node is
not solved for in the analysis, but is input as a boundary condition at all nodes. In
other words, the seepage solution is removed from the coupled equations because
the change in pore-water pressure is known ahead of time. This change in porewater pressure, u , will still result in a change in effective stress and subsequent
volumetric strain when the coupled equations are solved. It is just not part of the
solution.
The validity of the assumption that the changes in pore-water pressure can be
computed independent of the fully coupled solution becomes evident in the
following sections where comparisons are made between coupled and uncoupled
solutions. For most cases, the uncoupled formulation gives almost identical results.
The exception is the Cryers Ball example in which the Mendel-Cryer effect is not
accounted for in the uncoupled analysis. The Mendel-Cryer effect is when the
change in pore-water pressure exceeds the change in total stress which can occur in
certain special, not typical, cases.

Page 217

Chapter 10: Consolidation

SIGMA/W

In order to do an uncoupled analysis in SIGMA/W, you must specify where the


change in pore-water pressure data will come from. There are a lot of options
within the GeoStudio family of tools for this data. SIGMA/W requires two porewater pressure conditions for the same geometry, so that it can take the difference
between the two states and compute the u. The two pore-water pressure
conditions can come from a combination of user-defined water table and / or
solved analysis in SEEP/W, VADOSE/W, QUAKE/W, or SIGMA/W. The data
can be the start and end conditions from one or more of these files, or it can be all
time steps within any single file.
If only a start and end pore-water pressure condition are specified, SIGMA/W will
create a single load step analysis and computed the consolidation (volume change)
resulting from the specified change in pore-water pressure. If a file with multiple
seepage time steps is selected, SIGMA/W will generate a multiple load step
analysis with load steps that equal the elapsed time in the seepage solution for
times when seepage data is written to an output file. The actual real time from
the seepage solution does not have meaning in SIGMA/W, but SIGMA/W will
report the elapsed times of each consolidation step such that they correspond with
the seepage solution.

10.10 Consolidation analyses limitations


In an uncoupled consolidation analysis you should not apply any other loading in
the analysis. For example, you should not add a surface load that will cause a
change in pore-water pressure because it will not be accounted for correctly. The
pore-water pressures are supplied by the seepage solution. Applying any load that
results in a change in total stress may lead to a meaningless solution.
In a saturated only coupled consolidation analysis, you should not have seepage
boundary conditions specified that will result in an appreciable change in size of
the saturated zone. This means you should not use this option if the seepage
solution is for seepage through a dam during filling or draining of a reservoir, for
example. The saturated zone only option tells SIGMA/W to fix the pressures in the
unsaturated zone at the value they are prior to the load step. This means that they
will be fixed in the seepage solution as well which is not quite valid if there are
boundary conditions that want to change the size of this zone during the current
loading step.
The consolidation analysis (couple or uncoupled) should be restricted to observing
and understanding the dissipation of excess pore-water pressures, or to calculating

Page 218

SIGMA/W

Chapter 10: Consolidation

swelling or shrinking due to changes in pore-water pressure. It should not be


combined with a load deformation analysis. Keep in mind, however, that with the
new uncoupled option, two different SIGMA/W load deformation pore-water
pressure conditions can be used in a separate analysis in order to observe the
consolidation resulting from the load application.
Two different SIGMA/W load deformation analysis pore-water pressure distributions can
be used in an uncoupled consolidation analysis to observe dissipation of pressure due to
changes in total stress.

10.11 Examples of coupled and uncoupled consolidation


There are many examples that can be used to illustrate the power of coupled and
un-coupled consolidation analysis in SIGMA/W. This section will not show any
step-by-step procedures for setting up these examples, but it will show the use of
different types of consolidation options in order to compare them with each other
where applicable. Full details of several of these examples are given in the
Illustrative Examples chapter later in this book.
One dimensional consolidation examples

Figure 10-2 shows a simple 1D mesh that can be used to illustrate several
components of coupled and un-coupled consolidation. In this 1m tall saturated
column problem, a 1000 kPa load has been applied to the surface. There is zero
drainage on the sides and base of the sample so the entire 1000 kPa total stress load
is transferred to the pore-water and the pressure instantly jumps up to 1000 kPa in
the entire sample. Over time, drainage out the top of the column permits this
pressure to dissipate which increases the effective stress and results in shrinkage.

Page 219

Chapter 10: Consolidation

SIGMA/W

1.1
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0

100

200

300

400

500

600

(x 0.001)

Figure 10-2 Mesh used in 1D consolidation examples

SEEP/W SIGMA/W coupled solution

The problem of dissipating the excess pore-water pressure can be modeled fully
coupled using SEEP/W and SIGMA/W formulations simultaneously. Figure 10-3
shows the dissipation of excess pore-water pressure over time as computed using
the coupled formulation. As expected, the pressures want to revert back to the preload condition. Figure 10-4 shows the surface node displacement over time during
consolidation. As shown, the maximum displacement is about -0.8m over the time
modeled.
You may question how a 1m tall column can displace by 0.8m. In reality it cannot,
but numerically it can. In this analysis we have used a linear-elastic soil model
which has no yield point. What this means is that it can displace as much as it
needs to in order to take up the stress. This is a good example to illustrate that it is
important to know what the different soil models mean. In this example, the
computed stresses are correct, but the computed strains given by the linear model
are not realistic.

Page 220

SIGMA/W

Chapter 10: Consolidation

Pore-Water Pressure vs. Y


1.0
0.0000e+000
1.0000e+002

0.8

Elevation

2.0000e+002
3.0000e+002

0.6

4.0000e+002
6.0000e+002
0.4

8.0000e+002
1.0000e+003

0.2

1.2000e+003
1.4000e+003
1.6000e+003

0.0
0

200

400

600

800

1000

Pore-Water Pressure

Figure 10-3 Dissipation of pressure (fully coupled)


Y-Displacement vs. Time
0.0

Y-Displacement

-0.2

-0.4

-0.6

-0.8
0

500

1000

1500

2000

Time

Figure 10-4 Coupled analysis surface displacement with time

Page 221

Chapter 10: Consolidation

SIGMA/W

SIGMA/W SIGMA/W uncoupled solution

The output from the coupled solution can be put back into SIGMA/W to test the
uncoupled formulation. In other words, we can take the coupled pore-water
pressures from the end of the first load step in the coupled solution and use it as the
starting condition for an uncoupled solution. The ending condition for the
uncoupled solution can be set as the initial condition from the coupled solution. So,
we can simply test the un-coupled formulation by solving the coupled solution
backwards. Figure 10-5 shows the two pore-water pressure conditions applied in
the uncoupled analysis. These two conditions are directly obtained from the
previously solved SIGMA/W files. These are the two pressure conditions that will
give SIGMA/W the u at every node so that it can solve the consolidation
equations independently of a seepage solution. Figure 10-6 shows the computed ydisplacement at the end of the single load step un-coupled consolidation. The
maximum displacement at the surface is about -0.7 m which is not quite the same
as the -0.8 m computed in the fully coupled formulation.
The difference between the two displacements is explainable. In the coupled
formulation, the computed pressure written to file at the end of the first load step is
not quite equal to the total applied stress of 1000 kPa. In other words, during this
first load step, the load is applied and some of it has consolidated. So, when we
take the end of the first load step in the coupled formulation and use it as the start
of the un-coupled consolidation, we are missing a small amount of pressure
dissipation and this is giving some small difference in computed displacement.

Page 222

SIGMA/W

Chapter 10: Consolidation

Pore-Water Pressure vs. Y


1.0

Elevation

0.8

0.0000e+000

0.6

0.4

0.2

1.0000e+000

0.0
0

200

400

600

800

1000

Pore-Water Pressure

Figure 10-5 Single step uncoupled consolidation

This example has shown how output pressures from a previous SIGMA/W analysis
can be used to solve an un-coupled consolidation analysis without using the
SEEP/W program. In fact, SIGMA/W can use pore-water pressures generated by
itself, SEEP/W, QUAKE/W or VADOSE/W in its un-coupled formulation. There
is no need to specify hydraulic properties because the change in pore-water
pressures were previously solved in the external programs. SIGMA/W simply uses
these pressures to obtain the change in effective stress in order to determine the
volume change.

Page 223

Chapter 10: Consolidation

SIGMA/W

Y-Displacement vs. Y
1.0

Elevation

0.8

0.0000e+000

0.6

0.4

0.2

0.0
-0.8

1.0000e+000

-0.6

-0.4

-0.2

0.0

Y-Displacement

Figure 10-6 Single load step un-coupled consolidation surface displacement

Mendel-Cryer effect

The Mendel-Cryer effect involves pore-pressures that respond to loading such that
the pressure peaks at a higher value than the applied total stress. This can happen
when a total stress is applied to the surface of a sphere as illustrated in Figure 10-7.
In this case, a total stress of 100 kPa is applied on the first step and the pressures
are then allowed to dissipate with time. The seepage boundary condition is
specified in the SEEP/W program and it allows drainage out of the surface of the
sphere. Figure 10-8 shows the Mendel-Cryer effect where the pressure at the center
of the sphere slowly increases as the external pore-water pressure impact reaches
the center. After the peak, the excess pressures are allowed to dissipate back to the
pre-load conditions. Along with this increase in effective stress (e.g., dissipation of
excess pressure) comes shrinkage of the sphere as shown magnified 10x in Figure
10-7.

Page 224

SIGMA/W

Chapter 10: Consolidation

Figure 10-7 Mendel-Cryer sphere and deformed sphere (symmetric about left
axis)
Pore-Water Pressure vs. Time

Pore-Water Pressure

150

100

50

0
0

10

Time

Figure 10-8 Mendel-Cryer effect: pressure increase before dissipation

Page 225

Chapter 10: Consolidation

SIGMA/W

The same analysis can be carried out in an uncoupled formulation. The excess
pressures of 100 kPa can be input into SEEP/W and then allowed to dissipate over
the same time period. When this happens, the Mendle-Cyer effect of the pressure
increase in the center of the sphere is not reproduced and the change in volume of
the sphere will be slightly different.
Two dimensional uncoupled consolidation example

The real strength of un-coupled consolidation comes in analysis of two


dimensional saturated / unsaturated flow problems. Consider the following
saturated example of dissipation of excess pore-water pressure resulting from
placement of an embankment on the ground. The fill analysis was carried out in
stages with SIGMA/W using a total stress analysis with pore-water pressure
parameters in order to generate the pressure response to loading. The final
condition after loading is shown in Figure 10-9.
Using the un-coupled consolidation analysis option (no seepage solution required)
it is possible to use the illustrated SIMGA/w generated pore-water pressure
condition as a starting condition and then specify zero pressure as an ending
condition in order to compute the volume change resulting from dissipation of
these excess pressures.

40
35

25
20
15
160

100

80

10

120

160

Height (metres)

30

5
0
-5
-5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

Distance (metres)

Figure 10-9 Excess pressure beneath a fill embankment prior to


consolidation

Page 226

95

100

SIGMA/W

Chapter 10: Consolidation

Figure 10-10 shows the deformed mesh generated in a single load step using the
change in pore-water pressure between the post construction condition (previous
figure) and pre-construction condition. The maximum displacement is computed to
be about 8cm at the base of the fill in the center of the embankment.
40
35

Height (metres)

30
25
20
15
10
5
0
-5
-5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Distance (metres)

Figure 10-10 Deformed mesh after dissipation of construction excess


pressures

The most useful aspect to the un-coupled formulation is that you can use two
pressure conditions from any other GeoStudio analysis as start and end conditions
for a volume change (consolidation) analysis in SIGMA/W; and you obtain very
reasonable results without excessive computational effort or convergence issues.
NOTE: if you are dissipating an excessive pressure, and you have a non-linear soil
model, it may be useful to use the un-coupled option that applies the formulation
over successive time steps instead of a single load step. SIGMA/W can read
multiple pore-water pressure conditions and create the necessary loading sequence
to dissipate the pressures in smaller steps.

Page 227

Chapter 10: Consolidation

SIGMA/W

Page 228

SIGMA/W

11

Chapter 11: Numerical Issues

Numerical Issues

This chapter discusses several numerical issues related to solving the partial
differential equation for stress-deformation. It deals with non-linearity of soil
constitutive models and how they are dealt with. The chapter also discusses some
features in the SIGMA/W program that can help you assess convergence of your
solution such as graphing convergence data during the solve process or stopping
and restarting an analysis in order to make adjustments without having to re-start
an entire solve process.

11.1

Nonlinear analysis

SIGMA/W is capable of performing analysis involving nonlinear material


properties. The finite element equilibrium is derived for a linear static case. For
nonlinear material properties, the global stiff matrix , [K], is no longer linear and
iterations are required to achieve an acceptable solution. Additional information on
the procedure for nonlinear analysis can be found in finite element text-books such
as Bathe (1982), and Zienkiewicz and Taylor (1991). This section details some
general aspects of nonlinear incremental analyses relating to convergence criteria
and using SIGMA/W to estimate limiting loads.
In general, the basic problem in a nonlinear analysis is to find the state of
equilibrium of a body corresponding to the externally applied loads. If these loads
are a function of time, at time t the nonlinear equilibrium problem can be
formulated as a solution of the following equation:
Equation 11-1

{ } = { R ( a )} { F } = 0
t

where:

nodal unbalanced loads,

externally applied nodal loads,

nodal loads due to element stresses, and,

nodal displacements.

It is assumed that the solution will start from an equilibrium (or, at least, near
equilibrium) state such that (t-1) is approximately zero. The vector

Page 229

Chapter 11: Numerical Issues

{R} = [ B ]{ } dv

SIGMA/W

is calculated based on the internal element stresses as

described previously in this chapter. In SIGMA/W, external applied loads are


described using continuous spline functions, which describe the total load being
applied on the system versus time. Therefore, for a particular time increment t,
{tF} is an incremental load obtained by the difference between the value of the
load functions evaluated at time t and at time t=1.
SIGMA/W solves Equation 11-1 iteratively using the Newton Raphson technique
which approximates this equation as:

Equation 11-2

i +1

} = { } + a { a }
i

where i is the iteration counter and the superscript t has been omitted for clarity. To
start the iteration process for time step t, displacements at the end of the previous
time step (i. e. , time t-1) are used as the initial estimate:
t

a1 = t 1a

If the derivative term in Equation 11-2 is written as:

R
a = a = [ KT ]

where:
[KT]

the tangential stiffness matrix.

The iterative correction applied to the nodal displacements can be calculated as:

KT i { a i } = { i }
or:

{ a } = K
i

i
T

{ }
i

A series of successive iterations gives the following result:

Page 230

SIGMA/W

Chapter 11: Numerical Issues

a i +1} = { t 1a} + { a i}
k =1

In the Newton Raphson method, the stiffness matrix is tangential to the loaddisplacement curve and is updated every iteration. This process is illustrated in
Figure 11-1(a). For an analysis involving strain-softening materials, this process is
modified such that the stiffness matrix is not updated, but the stiffness matrix used
in the first iteration is retained throughout the time step, as illustrated in Figure
11-1(b). In order to achieve stability in SIGMA/W, the initial modulus is always
calculated based on elastic behavior.
It should be noted that the load term is the externally applied load for the first
iteration in a load step; and it is the unbalanced load for subsequent iterations.
Incremental displacements resulting from increment load {tF} at time t are
obtained upon solving Equation 11-1. These incremental displacements are added
to the displacements at the beginning of the time increment. In the displacement
output files, SIGMA/W reports the total displacements at each node since the
beginning of an analysis at Time Step 1.

Page 231

Chapter 11: Numerical Issues

SIGMA/W

Figure 11-1 Iteration schemes for nonlinear analysis

Page 232

SIGMA/W

11.2

Chapter 11: Numerical Issues

Convergence

In all non-linear analyses, it is necessary to use iterative techniques to compute


acceptable solutions. Non-linear analyses exist when a soil property is dependent
on the computed results. For example, the soil stiffness modulus E is dependent on
the stress state in the ground, but the stresses are dependent on the soil stiffness.
This means that the analysis has to be done many times until there is a reasonable
match between the soil properties and the computed stresses. When the analysis
has produced an acceptable match, the solution is deemed to have converged. As
the user, you must decide what constitutes convergence.
Two convergence criteria, the displacement convergence criterion and the
unbalanced load criterion, can be used in SIGMA/W to control the iteration
process. These are described in more detail below. When a convergence criterion is
satisfied, the analysis is considered to have converged (or achieved convergence)
and SIGMA/W will stop the iteration process and go to the next time step. As a
general guideline, a convergence criterion of 1% to 2% and a maximum number of
iterations of 25 are adequate for most problems.
Convergence parameter definitions
Maximum # of iterations - This parameter limits the number of iterations SOLVE
will execute in an attempt to obtain a solution. Execution will come to a halt or
move onto the next step if the iteration number reaches the maximum specified.
Tolerance (%): displacement norm - This parameter is the desired percentage
change in the norm of the current nodal (incremental) displacement vector with
reference to the total displacement vector in the load step. The iteration process
stops if the percentage difference is less than the specified tolerance. If the
percentage difference is greater than the tolerance, the iteration process continues
until it reaches the maximum number of iterations.
Tolerance (%): unbalanced load norm - This parameter is the desired percentage
change in the norm of the unbalanced load vector when compared to the norm of
the applied load vector. The iteration process stops if the percentage difference is
less than the specified tolerance. If the percentage difference is greater than the
tolerance, the iteration process continues until it reaches the maximum number of
iterations. This tolerance is only applicable when you are applying stress or load. It
is especially useful when nonlinear elastic or elasto-plastic materials are used in
your finite element model.

Page 233

Chapter 11: Numerical Issues

SIGMA/W

Both criteria are expressed as the Euclidean norm of a vector quantity. The
(Euclidean) norm of a vector is a measure of its magnitude and is defined as:

|| D || = d 2
i =1

where:

|| D || =

the norm of a vector quantity d,

a vector quantity, for example, incremental nodal displacement for a

particular iteration, and


n

the size of that vector.

For example, the norm of nodal displacements, a , is given by:


N

( a

a =

n =1

2
xn

+ a yn 2 )

where:
N

total number of nodes,

axn

incremental x-displacement at node n, and

a yn

incremental y-displacement at node n.

The displacement convergence criterion compares the norm of incremental


displacements in an iteration to the norm of the accumulated incremental
displacements in a time step. This criterion is expressed as a ratio in percentages,
for the i-th iteration, it can be expressed as:

ai
i 1

100%
k

k =1

where:

Page 234

SIGMA/W

Chapter 11: Numerical Issues

a user-defined tolerance number,

accumulated nodal displacements in the load step


i 1

, and

k =1

the vector of nodal increment displacements for the i-th iteration.

The unbalanced load convergence criterion compares the norm of unbalanced load
in an iteration to the norm of the externally applied load in a time step. This
criterion is also expressed as a ratio in percentages, for the i-th iteration, it can be
expressed as:

i
F

100% f

where:

a user-defined tolerance number,

nodal unbalanced load for the i-th iteration, and

applied increment load for the load step.

For an analysis without any externally applied load (i. e. when only displacement
boundary conditions are applied), SIGMA/W ignores the unbalanced load
criterion. For an analysis involving only linear elastic materials (the Linear-elastic
model and slip elements), SIGMA/W skips the iteration process because the
elasticity is not a function of confining stress and is therefore constant for all
elements at all times.
NOTE: You may select one or both convergence criterion; if you have selected
both criteria, iteration will only be stopped after both criteria are satisfied or when
the maximum number of iterations has been exceeded. As a general rule, the
unbalanced load convergence criterion is more useful when you are approaching a
limit load; otherwise, the displacement convergence criterion is usually sufficient.

Page 235

Chapter 11: Numerical Issues

SIGMA/W

Graphing convergence

SIGMA/W SOLVE allows you to view the convergence process graphically.


Whenever a large number of iterations is required to reach a solution, a graphical
representation will give you a much better appreciation of the convergence process
than if you strictly rely on the numerical data in the SOLVE window. Generally,
the convergence parameters should tend toward reasonable, stable, constant values.
Viewing the convergence parameters graphically makes it easier to judge whether
this is the case.
A graph of the displacement vector norm or the unbalanced load vector norm
versus the iteration number can be displayed by clicking the Graph, button. This
feature is particularly useful when you are analyzing a problem approaching limit
equilibrium. You can either view the graph as during an analysis, or you can open
a data file and display the graph for a previously-computed problem.
The displacement vector norm or unbalanced load vector norm may oscillate
substantially at the beginning of the iteration process. The convergence plot makes
it possible to graphically watch the convergence process and visually judge the
convergence.
Figure 11-2 shows the convergence graph for a SIGMA/W example analyzing a
bearing capacity problem in a cohesionless soil.
Step Number 24

Displacement Norm

0.0025

0.0020

0.0015

0.0010

0.0005

0.0000
0

10

15

20

25

Iteration #

Figure 11-2 Convergence graph showing displacement norms during load


step 2

Page 236

SIGMA/W

Chapter 11: Numerical Issues

Plotting displacement norms, the graph in Figure 11-2 seems to indicate that the
solution is converging quite nicely. Figure 11-3 shows unbalanced load norms for
the same problem. It shows significant fluctuations in unbalanced load indicating
that the analysis may be experiencing convergence difficulties.
Step Number 24

Unbalanced Load Norm

1
0

10

15

20

25

Iteration #

Figure 11-3 Convergence graph showing unbalanced load norms at step 24

The convergence graph for the following step (Step 25) shows an unacceptable
increase in the unbalanced load norm (See Figure 11-4). The analysis needs to be
restarted at Step 25 after reducing the load step size in SIGMA/W DEFINE. A
graph of unbalanced load norms may provide a better indicator for convergence
and it is especially useful when performing a non-linear analysis.

Page 237

Chapter 11: Numerical Issues

SIGMA/W

Step Number 25

Unbalanced Load Norm

150

100

50

0
0

20

40

60

80

100

Iteration #

Figure 11-4 Convergence graph showing unbalanced load norms at step 25

When a problem involving sequential fill placement with fill height adjustment is
analyzed, the convergence record is reset to Iteration 1 after the fill height is
adjusted. The convergence record for the entire time step can be viewed once the
fill height adjustment process has been completed.
A trick that is sometimes useful in judging convergence is to set the maximum
number of iterations to a very high number and also set the convergence tolerance
to a very low number. This will force the solution far past an acceptable
convergence, but it can give you confidence that you have indeed reached an
acceptable, stable solution. You can use this knowledge to select an acceptable
criterion. In some cases, you may be able to relax the criterion while in other cases,
you may have to tighten the criterion.

11.3

Limiting loads use incremental loading

The SIGMA/W finite element equations are a set of equilibrium equations. The
driving (applied) forces are in equilibrium with the resisting forces that arise from
the soil strength. The load that can be applied is therefore limited by the ultimate
resistance of the system. In the terminology of SLOPE/W (another GEO-SLOPE
product), the limiting load is the load that results in a factor of safety equal to 1. 0.

Page 238

SIGMA/W

Chapter 11: Numerical Issues

The modeling consequence of applying loads in excess of the limiting load is


usually numerical instability; that is, a lack of convergence and widely varying
oscillation in the solution. Unfortunately, the limiting load is not always known.
Lack of convergence after a series of successive load steps has been applied may
be a hint that you have reached the limiting load.
In certain cases, it may be of interest to carry the analysis very close to the point of
limiting equilibrium. A technique that can be used is to specify the displacement
instead of the applied load. In the case of a footing analysis, you could apply a load
sequence until you approach the limiting load, stop the processing, specify the
boundary condition as a specified displacement, and then restart the processing.
Specifying displacement instead of force is likely mandatory for the strainsoftening model if you wish to carry on an analysis until the ultimate strength of
your material is mobilized. When the applied load approaches the ultimate value,
you may experience difficulties with convergence unless you use
strain-controlled loading. This is illustrated in the Strain- Softening analysis in
the Illustrative Examples chapter.
Another technique for obtaining a solution when limiting equilibrium is
approached is to reduce the size of the load increment for this particular load step
(and subsequent load steps) using SIGMA/W DEFINE and then restart the
analysis. This method is illustrated in the Bearing Capacity analysis in the
Illustrative Examples chapter.
It is also important to remember that SIGMA/W is not formulated for large
displacements or large strains. Therefore, carrying an analysis well past the point
of limiting equilibrium may produce unrealistic results.

11.4

Tension zones

The effect of tension zones can be handled indirectly with the Nonlinear Elastic
(Hyperbolic) soil model. Consider the case of a steep slope with tension in the
crest area. If the material has tensile strength, energy will be stored in the system,
preventing the soil from deforming to the point where cracks develop. However, if
the soil has no tensile capacity, cracks will form and the equivalent stored energy
will contribute to additional deformation. To prevent the soil from storing energy
representative of the tensile strength, the soil can be made very soft by assigning it
a low E modulus.

Page 239

Chapter 11: Numerical Issues

SIGMA/W

In the Nonlinear Elastic (Hyperbolic) model, the soil stiffness can be a function of
the confining stress ( 3 ) . In a tension zone, the minor principal stress will tend
towards zero or even become negative. In response to this low confining stress, the
soil is assigned a very low E modulus value.
You can specify both Displacement Norm Tolerance and Unbalanced Load Norm
Tolerance. In this case, iterations will continue until both criteria are satisfied or
until the maximum number of iterations is exceeded.
In most modelling problems, only the displacement norm criterion is required. The
unbalanced load norm criterion is especially useful in an elasto-plastic analysis
when approaching the collapse load.
In summary, the Nonlinear Elastic model can be used to make the soil soft in a
zone of tension, essentially eliminating the energy representative of tensile
strength.

11.5

Stop-restart

SIGMA/W has a powerful Stop-Restart feature. The processing may be stopped at


any time step and then restarted after making changes in the finite element model.
For example, consider a load/deformation analysis with 60 time steps. If you click
the Stop, button after the 31st time step, you can then modify the material
properties or reduce the size of the applied load using the DEFINE function and resave the problem.
NOTE: When you set the starting time step in SIGMA/W for a consolidation
analysis, SEEP/W SOLVE will automatically select the head and property files for
the previous step, Step 31, as the SEEP/W initial conditions files.

11.6

Halt iteration

The iteration process for the current time step can be halted by clicking the Halt
Iteration, button. SOLVE uses the results from the last iteration to create the output
files (if necessary) for the current time step, and then continues analyzing the next
time step.
To halt the analysis at the current time step, you should press the Halt Iteration,
button and wait until the next time step is being analyzed; then, press the Stop,
button. This allows SOLVE to create the necessary output files for the current time

Page 240

SIGMA/W

Chapter 11: Numerical Issues

step. If you press the Stop, button before the next time step is processed, the
analysis may be stopped before SOLVE can create the output files.
If you stop an analysis and wish to restart at the next time step, you must select the
latest displacement, stress, strain, pore-water pressure and property files as your
initial condition files as described in the Stop-Restart section.
The iteration number is reset to 1 when the next time step is started.

11.7

Solver generated data

The solver program reads all input geometry, material properties, initial and
boundary conditions and then computes results for all user specified output load
steps. All computed solver output is written to text files that are then available to
the Contour program for visualization. The various output files are discussed
below. The letter designation associated with each file is the first letter in the
output file extension and is followed by an integer numeral that designates which
load step the file is associated with. Therefore myfile.d01 is the filename of the
displacement file generated for load step 1.
The displacement (D) files contain the displacement, the applied and computed
reaction forces at the nodes, pressure boundary conditions applied at the edges, and
axial and shear forces at structural element edges. The extension of each
displacement file name begins with D. The displacement file information can be
contoured, plotted, and viewed at each node.
The stress (S) files contain the x-stress, y-stress, z-stress, and x-y shear stress at
each element integration point. The extension of each stress file name begins with
S. The stress file information can be contoured, plotted, displayed as a Mohr
Circle, and viewed at each element Gauss region or approximated at each node.
The strain (E) files contain the x-strain, y-strain, z-strain, and x-y shear strain at
each element integration point. The extension of each strain file name begins with
E. The strain file information can be contoured, plotted, displayed as a Mohr
Circle, and viewed at each element Gauss region or approximated at each node.
The pore-water pressure (U) files contain the pore-water pressure at each element
integration point. The extension of each pore-water pressure file name begins with
U. The file information can be contoured, plotted, and viewed at each element
Gauss region or approximated at each node.

Page 241

Chapter 11: Numerical Issues

SIGMA/W

The material property (L) files contain the yield conditions, deviator stress, and
material properties at each element integration point. The extension of each
material property file name begins with L. The various material properties can be
contoured, plotted, and viewed at each element Gauss region or approximated at
each node.
While it is useful to know the file formats, it is not necessary to ever access the
data using a text editor. It is possible to have the Contour program present all data
in graphical or tabular format. This is discussed in detail in the following chapter
on Visualization of Results.

Page 242

SIGMA/W

Chapter 12: Visualization of Results

12

Visualization of Results

12.1

Introduction

An attractive feature of SIGMA/W is the varied yet flexible ways that the results
that can be viewed, evaluated and visualized. This is important for any finite
element analysis due to the large amount of data involved. Much of the data can
only be interpreted through graphical visualization using contours and graphs. And
yet at the same time it is often necessary to look at details by examining the digital
data at a particular node or at a Gauss integration point in an element.
This chapter describes and summaries the capabilities available in SIGMA/W for
viewing, evaluating and visualizing the results.

12.2

Load steps

Some SIGMA/W analyses such as in-situ or single step un-coupled consolidation


are solved in one load step. Others, however, such as sequential fill placement,
require many small loading steps. The Contour program is capable of letting you
view data from individual load steps or multiple load steps. If you have a single
load step data set active, then you can visualize data such as the deformed mesh or
contours of pressures, stresses etc. If you have multiple load step data active in
Contour, then you can not view a contour of pressures, but you can then graph how
any parameters change over the loading sequence. In either case, the data you are
viewing is the actual condition in the soil at the time (or load step) you view it.
Unlike the solver, which computes incremental changes between each load step,
the Contour program presents the end result of adding up all the incremental
changes up. The option to view only incremental differences is discussed next.
Incremental difference

Often it is of interest to look at the difference between two time steps. Say for
example that we want to examine only the excess generated pore-pressures. This
can be done by subtracting the initial static conditions from the results for any time
step or all the time steps.
There is an option to identify and select the load step that should be used for
comparison with all other steps. The results data conditions in this identified file
are subtracted from all the other time increments selected for viewing. If this file

Page 243

Chapter 12: Visualization of Results

SIGMA/W

represents the initial static conditions, then the initial static conditions will be
subtracted from the other time increments selected for viewing. If for example, you
are looking at pore-pressures, the pore-pressures will be the pore-pressures at each
selected time increment less the initial static conditions.

12.3

Types of data available in Contour

All computed parameters except displacement, boundary force, and pore-water


pressure are computed as values at element Gauss points. The Gauss values are
then projected from the Gauss points to the nodes in order to obtain nodal values
for viewing and graphing. A discussion on the implications of projecting Gauss
point data to nodes is provided in the following section. Keep in mind that if nodal
values appear inconsistent, you can always view the actual computed Gauss point
data to verify any concerns.
Some of the computed parameters are obtained directly from the problem data
files, while others are calculated from the values in the output files. The following
three tables show all the types of post processor computed stress and strain data
available for visualization as well as how they are computed. Parameters computed
directly by the solver are not included in these tables, even though they are
accessible for visualization.
Table 12-1 Force and displacement parameters computed in Contour
Parameter
XY-Displacement
XY-Force

(F )
xy

Equation

( )
xy

xy = x2 + y2
Fxy = Fx2 + Fy2

Page 244

SIGMA/W

Chapter 12: Visualization of Results

Table 12-2 Strain parameters computed in Contour


Parameter

( max )

Maximum Strain

Minimum Strain

Equation

( min )

Maximum Shear Strain

Volumetric Strain

( v )

Deviatoric Strain

( q )

( max )

max

(
=

min

(
=

+ x )
2

y + x )

( y x ) xy 2
+
+

2
2

( y x ) xy 2

+

2
2

( y x ) xy 2

2
2

max =

v = x + y + z

y ) + ( y z ) + ( z x )
2

1
q =
2 + 3 2
xy
2
x

Page 245

Chapter 12: Visualization of Results

SIGMA/W

Table 12-3 Stress parameters computed in Contour


Parameter
XY-Displacement

( xy )

XY-Force

(F )
xy

Equation

xy = x2 + y2
Fxy = Fx2 + Fy2

Maximum Total
Stress

( max )

max

(
=

min

(
=

Minimum Total Stress

( min )

Mean Total Stress (p)

p=
Normal Total Stress

( n )

Tangential Total
Stress

( t )

n =
t =

+x )
2
+x )
2

( y x )

+ xy2
2

x + y
2

( x y )
2

Y-Effective Stress

y' = ( y u )

( x y )

Z-Effective Stress

z' = ( z u )

Maximum Effective

'
max
= max u

Stress

( 'max )

Minimum Effective
Stress

( 'min )

cos 2 + xy sin 2

sin 2 xy cos 2

( ' z )

x' = ( x u )

( ' )

+ y +z )

X-Effective Stress

( 't )

( y x )
+
+ xy2
2

'
min
= min u

Page 246

SIGMA/W

Chapter 12: Visualization of Results

Mean Effective Stress

(p )

p' =

'

Normal Effective
Stress

( 'n )

Tangential Effective
Stress

=
'
n

( 't )

t' =

xy max

Deviatoric Stress (q)

12.4

+ y +z )

x' + y'
2

( x' y' )

Maximum Shear
Stress

( x' y' )
2

cos 2 + xy sin 2

sin 2 xy cos 2

( y x )

+ xy2 = 1 3
2

xy max =

1 ( x y ) + ( y z ) + ( z x )
q=
2 +6 xy2
2

Viewing node and element Information

The Contour viewing node and edge information differs from the similar command
in Define. In Contour, this command displays the boundary pressure applied by
SOLVE along a particular element edge or at a node, while in Define, this
command displays all element edges along which the user has specified a
boundary pressure. In particular, if you have specified a pressure boundary
condition using a function, Contour will show you the pressure applied by SOLVE
at the time step you are viewing, while SOLVE will only show you the function
number.
Figure 12-1 and Figure 12-2 show the process of viewing node and element
information for a node and one of its adjacent Gauss points at the center and base
of a dam after consolidation of the excess pore-water pressure caused by placement
of the fill. You can see that there is a similarity between Gauss values and nodal
values, but that they are not always the same. The reason for this is given in the
next section.
In general, SIGMA/W has many ways to access and view results data; at the core
of these options is the ability to see what is happening at each point in the define
mesh.

Page 247

Chapter 12: Visualization of Results

SIGMA/W

Figure 12-1 Viewing node information

Figure 12-2 Viewing element information

Page 248

SIGMA/W

12.5

Chapter 12: Visualization of Results

Projecting Gauss point values to nodes

SIGMA/W contours parameter values calculated at nodes. Since displacement and


boundary force are computed and stored at the nodes, these parameters can be
contoured directly. All other parameters are stored at the element Gauss points,
however, and must therefore be projected to the nodes for contouring purposes.
In triangular elements, the Gauss point values are projected on the basis of a plane
that passes through the three Gauss points. For one-point integration, the value at
the Gauss point is also taken to be the value at the nodes (i.e., the Gauss point
value is constant within the element).
In quadrilateral elements, the Gauss point values are projected using the
interpolating functions described in Interpolating Functions the appendix. In
equation form:

x= N

{X }

where:
x

projected value outside the Gauss points at a local coordinate greater than
1. 0

<N>

matrix of interpolating functions

{X}

value of Gauss point variable

The local coordinates at the element nodes are the reciprocal of the Gauss point
local coordinates when forming the element characteristic matrix. Figure 12-3 is an
example of the local coordinates at the element corner nodes when projecting
outwards from the four Gauss points in the element. The value of 1.7320 is the
reciprocal of the Gauss point coordinate 0.57735.

Page 249

Chapter 12: Visualization of Results

SIGMA/W

Figure 12-3 Local coordinates at the corner nodes of an element with four
integration points

This projection technique can result in some over-shoot at the corner nodes when
variation in the Gauss point values is large. For example, consider that we wish to
contour stress in a highly deformed element which, consequently, has large
variations in stresses at the Gauss points. Projecting such large variation in stresses
can result in unrealistic stress at the nodes.
Extreme changes in the parameter values at the Gauss points within an element
often indicate numerical difficulties (the over-shoot at the nodes being just a
symptom of the problem). This over-shoot can potentially be reduced by a finer
mesh discretization. Smaller elements within the same region will result in a
smaller variation of parameter values within each element, therefore lowering the
potential for encountering unrealistic projections.

12.6

Contouring data

There is a long list of computed values that can be contoured, and it generally
includes any of the information that can be viewed at a node as illustrated above.
Figure 12-4, for example, is a contour plot of excess (final minus initial) porepressures in a selected zone in a dam.

Page 250

SIGMA/W

Chapter 12: Visualization of Results

0
1 50
0

50

Figure 12-4 Contours of excess pore-pressure

The contour dialog box, illustrated in Figure 12-5, shows the data range for the
selected parameter. In this example, the pore-pressure range is -418 to 4214 psf.
From this, SIGMA/W computes a starting contour value, a contour interval and the
number of contours. These are default values. Seldom do these default contour
parameters give a nice picture. The problem usually is some numeric noise in the
data at the extremities of the data range. Usually the contour parameters need some
adjustment to produce meaningful and publishable contour plots. The default
values are good for a quick glance at the results, but not adequate for presentation
purposes.
For the illustration in Figure 12-5, the range starts at -418, but the starting value
has been modified to be zero. The number of contours was adjusted to be 10 so the
last contour is at 4500, just greater than the maximum in the data range
It usually takes several iterations to obtain a contour plot that is meaningful and
presents the intended message. The contour coloring and shading scheme is
described in the software on-line help.

Page 251

Chapter 12: Visualization of Results

SIGMA/W

Figure 12-5 Contour dialog box

Some of the minor numerical irregularities can be ignored by simply selecting a


contour range that excludes these values or by contour data specific to certain
element regions as discussed next.
Contouring specific element regions

Sometimes it is desirable to look at the results separately for a particular region.


Contours are created from conditions at the element nodes. Nodes common to two
different regions with different soil properties represent averages for the two types
of soils. This can lead to a confusing picture of the results, particularly if there is a
sharp contrast in properties or conditions in the adjoining regions. This difficulty
can be overcome by viewing the results only for a selected region. Figure 12-6

Page 252

SIGMA/W

Chapter 12: Visualization of Results

illustrates a situation where excess pore-pressures are contoured in a selected


region; in this case, only in the foundation soil beneath an embankment fill.
40
35

25
20
0
18

80

60

80

10

160

15

120

Height (metres)

30

180

5
0
-5
-5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Distance (metres)

Figure 12-6 Contours in a selected region

12.7

Graphing data

The graphing options in SIGMA/W are very powerful. In most cases, any type of
data you need to access and graph can be done right inside the program. In rare
cases, you can get tabular access to the raw graph data and paste it directly into an
external program such as Excel for additional graphing or manipulation.
At selected nodes, the Draw Graph command allows you to plot a graph containing
any of the computed nodal parameter values listed in Table 12-4.

Page 253

Chapter 12: Visualization of Results

SIGMA/W

Table 12-4 Data available for graphing


XY-Displacement

X-Boundary Force

X-Displacement

Y-Displacement

Y-Boundary Force

XY-Boundary Force

X-Total Stress

Y-Total Stress

Z-Total Stress

Maximum Total Stress

Minimum Total Stress

Mean Total Stress (p)

Normal Total Stress

Tangential Total Stress

X-Effective Stress

Y-Effective Stress

Z-Effective Stress

Maximum Effective Stress

Minimum Effective Stress

Mean Effective Stress (p')

Normal Effective Stress

Tangential Effective Stress

X-Y Shear Stress

Maximum Shear Stress

Deviatoric Stress (q)

Pore-Water Pressure

X-Strain

Y-Strain

Z-Strain

X-Y Shear Strain

Maximum Strain

Minimum Strain

Maximum Shear Strain

Volumetric Strain

Deviatoric Strain

Poisson's Ratio

Undrained Shear Strength

Void Ratio

Tangential Modulus (E)


Moment (beam element)

Rotation (beam element)

Axial Force (beam / bar element)

Shear Force (beam element)

The parameters in this table are the dependent variables of the graph. Any of the
dependent variables can be plotted versus the following independent variables:
nodal x-coordinates, nodal y-coordinates, or distance between nodes (starting at the
first selected node).
The following independent variables are also available if more than one time
increment is being viewed: time, X-Displacement, Y-Displacement, X-Strain, YStrain, X-Y Shear Strain.
The independent graph variable that you choose affects how the selected nodes and
time steps are used in the graph:

Page 254

SIGMA/W

Chapter 12: Visualization of Results

If the graph independent variable is x-coordinate, y-coordinate, or distance, then


the parameter value at each selected node is plotted versus the nodal coordinate or
the distance between nodes. Each selected time step is plotted as a separate line on
the graph.
If the graph independent variable is time, then the parameter value at each selected
node is plotted versus the elapsed time for each of the selected time steps. Each
selected node is plotted as a separate line on the graph.
Sum versus average graphs

A new graphing option that is quite powerful in some cases is the option to graph
the sum of all selected nodal dependent variables versus the average of all selected
independent variables. Consider for example that you want to create a loaddeformation graph for designing a footing. In the past, you would have extract
force and displacement data for each node at the base of the footing and then take
the data into Excel to sum the forces and average the displacement. This can be
done automatically now as illustrated in Figure 12-7. In this figure, all the nodes
beneath the footing are selected and then the graphing routine sums the forces and
plots them against the average displacement.
This graphing option can be used in many scenarios throughout the GeoStudio
analysis tools.

Page 255

Chapter 12: Visualization of Results

SIGMA/W

Figure 12-7 Load-displacement graph for footing design

12.8

Viewing mesh displacement

Contour draws the displacement at each node in the mesh. Most displacements
cannot be drawn un-scaled, since they would be too small to be visible. Specifying
a magnification value allows you to control the scale at which the displacement is
drawn. When you type a value in the magnification edit box, the maximum length
edit box is updated to display the length at which the maximum displacement will

Page 256

SIGMA/W

Chapter 12: Visualization of Results

be drawn. You can control the displacement length either by specifying a


magnification value or by specifying a maximum length value. The magnification
value is computed according to the following general relationship:

Magnification=

Max.Length Eng. Scale


Unit Conversion
Max. Deformation

Displacement is only drawn in element regions being viewed. As described in the


previous section, sometimes it is desirable to view specific regions. If a sub-region
of the domain is specified, CONTOUR finds the maximum displacement in the
selected region and uses it to scale the displaced mesh.

12.9

Viewing Mohr circles

A complete picture of the stress state at a point can be viewed with a Mohr circle
diagram such as in Figure 12-8.
Effective Stress at Node 655

40

85.488

30

-24.559
36.421

sx

20

Shear

10
0

95.668
-10
-20

26.241

sy

-30
-40
20

30

40

50

60

70

80

90

100

Normal

Figure 12-8 Mohr circle for element stresses

Mohr circles can be viewed at nodes or Gauss regions in an element. Mohr circles
can also be generated and viewed in terms of total stresses or effective stresses. In
addition, strain Mohr circles can be viewed as well as stress Mohr circles.

Page 257

Chapter 12: Visualization of Results

SIGMA/W

Page 258

SIGMA/W

13

Chapter 13: Illustrative Examples

Illustrative Examples

This chapter presents the analyses of some common stress and deformation
example problems. These examples are presented in order to:

Verify the SIGMA/W formulation, where possible, by comparing the


SIGMA/W solutions with closed form or published solutions;

Illustrate how to model various stress and displacement conditions;


and

Demonstrate how to access SIGMA/W analysis results using the


CONTOUR function.

The GeoStudio project files for each problem are included with the SIGMA/W
software. The files can be used to re-run each analysis and to verify that you can
obtain the same results as presented in this chapter.

13.1

Strip footing stresses

Included files

Strip footing.gsz.

Contours of the vertical stress beneath a footing form what is commonly known as
a pressure bulb. The shape of the pressure bulb in a homogeneous linear elastic soil
can be used to verify the SIGMA/W formulation.
In this example, SIGMA/W is used to analyze a strip footing using plane strain
analysis. Drained (effective stress) material parameters are used. They are: E' =
100 kPa and v' = 0.45. In a homogenous elastic material, these parameters will
affect only the displacements, but not the stress distribution. The footing load is
applied as a Y pressure boundary with a pressure of unity (1.0) so that the resulting
contours can be presented as a ratio of the applied pressure.
Figure 13-1 shows the finite element mesh. Note the combination of element sizes.
Near the footing, smaller 8-noded elements are used while further away, the
elements increase in size. The boundary conditions on the bottom and right
indicate zero x- and y-displacement which assumes there is no influence from the
footing at these locations. The left side boundary condition is zero x-displacement

Page 259

Chapter 13: Illustrative Examples

SIGMA/W

which will allow vertical deformation only. Figure 13-2 presents the resulting
vertical stress contours.
B/2

1B

Strip Footing

2B

3B

Figure 13-1 Finite element mesh for strip footing


B/2

Strip Footing
0.9

2B

0.3

0.

0.

0.7

1B

0.2

3B

Figure 13-2 Vertical stress contours beneath a strip footing

Page 260

SIGMA/W

Chapter 13: Illustrative Examples

The contours have the correct general shape and are essentially the same as the
values predicted by closed form solution. Table 13-1 shows a comparison between
the position of the vertical stress contours along the center-line beneath the footing
obtained from a closed form solution and from SIGMA/W using the CONTOUR
function. The closed-form solution is obtained from the pressure bulb charts
presented in Lambe and Whitman (1979). The location of the two sets of contours
is similar with the exception of the stress near the bottom boundary of the problem.
Table 13-1 Comparison of SIGMA/W and closed form results for strip footing

13.2

Stress Contour Ratio

SIGMA/W

Closed Form Charts

0. 8

0. 52 B

0. 56 B

0. 6

0. 88 B

0. 89 B

0. 5

1. 10 B

1. 11 B

0. 4

1. 51 B

1. 48 B

0. 3

2. 25 B

2. 04 B

Circular footing stresses

Included files

Round footing.gsz

In this example, the vertical stress distribution under a circular footing is modeled
using SIGMA/W. The vertical stress contours beneath a circular footing have the
same general shape as for a strip footing. However, the size of the pressure bulb is
much smaller.
Reflecting the circular geometry of the footing, this problem is modeled using an
axisymmetric view. The foundation material is considered to be homogenous
linear elastic with E' = 100 kPa and ' = 0. 45. In this material, the choice of these
parameters will affect only the displacement results, but not the stress distribution.
The footing load is applied as a y-pressure boundary with a downward pressure of
1.0 kPa. The details of the problem may be obtained by viewing the input file
using the DEFINE function.
Figure 13-3 shows the SIGMA/W computed vertical stresses beneath a circular
(round) footing.

Page 261

Chapter 13: Illustrative Examples

SIGMA/W

B/2

0.9

1B

2
0.

1
0.

2B
Figure 13-3 Vertical stress contours beneath a round footing

As in the strip footing example, contours of vertical stress obtain by SIGMA/W


have the correct general shape and are at similar locations along the center-line
beneath the footing as those obtained from charts presented in Winterkorn and
Fang (1975). These locations are summarized in Table 13-2.
Table 13-2 Comparison of SIGMA/W and closed form results for circular
footing

13.3

Stress Contour Ratio

SIGMA/W

Closed Form Charts

0. 9

0. 28 B

0. 26 B

0. 7

0. 43 B

0. 45 B

0. 5

0. 63 B

0. 64 B

0. 3

0. 94 B

0. 97 B

0. 2

1. 20 B

1. 26 B

Nonlinear-elastic (Hyperbolic) analysis

Included files

Hyperbolic_i.gsz

Page 262

SIGMA/W

Chapter 13: Illustrative Examples

Hyperbolic.gsz

This example involves a settlement analysis of a rigid circular footing on clay, as


shown in Figure 13-4. The clay is modeled using the Non-Linear-elastic model in
SIGMA/W. The radius of the footing is 4 feet. The footing is included in the finite
element mesh and is given an elastic modulus much higher than the soil. In this
way, the rigidity of the footing is taken into account.

Figure 13-4 Rigid circular footing on clay soil

The foundation clay is modeled as a non-linear elastic (hyperbolic) material with


the parameters given in Table 13-3.
Table 13-3 Non linear-elastic (hyperbolic) parameters
Parameter

Value

cohesive strength, c

1000 psf

internal friction angle,

Page 263

Chapter 13: Illustrative Examples

SIGMA/W

loading modulus number, KL

47

loading exponent, n

unloading modulus number, KUR

47

failure ratio, Rf

0.9

Poissons ratio,

0.48

Minimum stress, S3

200

For an explanation of these parameters, see the non-linear elastic section of chapter
on Material Properties. In this example, the stiffness of the clay is assumed to be
independent of the confining stress (i.e., depth). This is simulated by setting the
hyperbolic exponent, n, to zero. The Poisson's ratio is assumed to remain constant
as the footing load is applied. The clay has an un-drained strength of 1000 psf; that
is, c equals 1000, and equals zero.
The initial tangent modulus applied in the analysis is 1x105 psf, and the
atmospheric pressure is set to 2116 psf. In this example, the initial soil stiffness is
assumed to be independent of the confining stress 3; that is, the exponent n is 0.
Consequently, Ei is equal to KL multiplied by 2116 and the modulus number KL is
47. In equation form:

A minimum allowable stress value, S3, of 200 has been applied to the model so
that there is some stiffness to the soil when the confining stresses approach zero.
This reduces convergence problems because it prevents the soil stiffness from
behaving like fluid clay which we do not anticipate in this case.
Before proceeding with the actual settlement analysis, the in-situ stress state needs
to be defined. The in-situ stresses are necessary when the soil stiffness under
loading is a function of the confining stress present in the ground prior to loading.
This is the case for the Hyperbolic soil model.

Page 264

SIGMA/W

Chapter 13: Illustrative Examples

In-situ stresses

A SIGMA/W in-situ analysis can be used to establish the initial stress conditions.
The analysis type is set to In-situ 1 because we have a horizontal ground surface
and we are not using submerged unit weight. It is assumed that the clay is uniform,
its bulk unit weight is 110 pcf, and the at rest earth pressure coefficient, K0, is 0.
92.
The self-weight of the footing is ignored in the in-situ analysis because it does not
exist at the time the stresses are developed in the in-situ soil. This is achieved by
removing the footing elements from the finite element grid either by setting the
footing material to be a null element, or by assigning its self-weight to zero.
Examination of the computed results confirms that, as expected, the vertical total
stress at a point is equal to the self-weight times the depth. In equation form:

where:
d

the depth below the ground surface.

The horizontal total stresses x and v are equal to y multiplied by 0. 92.


NOTE: When contouring the vertical stress (Y-Total Stress), you may notice that
some contours are not perfectly horizontal. This comes about because of the
different element characteristics (number of nodes, shape, size and integration
order) and the projection of Gauss point stresses to the nodes. Stresses are exact at
Gauss points and, therefore, the results are correct even though the contours are not
perfect. A better vertical stress distribution is obtained when you are using 8-nodes
elements as compared to 4-noded elements.
Settlement analysis

After the initial stresses are established and verified using hand calculations if
possible, a settlement analysis can be carried out using the non-linear elastic
model.
When developing the finite element model for this settlement analysis, the footpound (force) unit system is used. Finite elements representing the rigid concrete
footing are simulated as linear elastic materials with a very stiff modulus. The

Page 265

Chapter 13: Illustrative Examples

SIGMA/W

foundation clay is modeled as a non-linear elastic material. The footing load is


applied using a boundary function as illustrated in Figure 13-5 such that the
applied load increases gradually throughout the loading sequence. The time label
on the x-axis in the figure is not real time, it is a loading step. We are saying in this
analysis that we want to gradually load the footing so that we do not shock load the
analysis and give ourselves unnecessary convergence problems.

0
-1

Y-Stress (x 1000)

-2
-3
-4
-5
-6
-7
-8
-9
0

10

15

20

Time

Figure 13-5 Y-pressure function for application of footing load

When solving SIGMA/W, we set the output stress file from the in-situ analysis as
the initial condition file for stresses.
After solving, SIGMA/W Contour is used to process the results of the analysis. At
any given time step, nodal displacements at individual nodes can be obtained using
the View Node Information command. A nodal displacements graph as a function
of time can also be generated. In this case, it is useful create a total load versus ydisplacement graph for all nodes at the base of the footing. The graph Sum versus
Average option is selected so that an average footing displacement versus total
loading function can be obtained. In previous versions of SIGMA/W, this data
would have to be gathered into Excel and plotted there.

Page 266

SIGMA/W

Chapter 13: Illustrative Examples

Figure 13-6 shows a plot of the total applied load versus the average vertical
displacement. The shape of this load-displacement curve clearly shows the nonlinear nature of the problem.

Y-Boundary Force vs.


Y-Displacement

Y-Boundary Force

-20000

-40000

-60000

-80000
-2.5

-2.0

-1.5

-1.0

-0.5

0.0

Y-Displacement

Figure 13-6 Non-Linear-elastic load-displacement curve for a footing on clay

The ultimate footing pressure can be evaluated in terms of the Terzaghi bearing
capacity formula for a circular footing (Terzaghi, 1943). For a round footing with a
rough base, the ultimate bearing capacity is given as:

This approximate value is somewhat higher than the 6000 psf "ultimate" load
applied using the boundary stress function in the SIGMA/W analysis.

Page 267

Chapter 13: Illustrative Examples

SIGMA/W

Figure 13-7 shows the tangent modulii at the end of the loading. The initial
modulus E is about 100,000 psf. Note how under the footing the tangent modulus
has decreased to below 20,000.

80000

60
00

40
00
0

00
20

Figure 13-7 Contour of tangent modulii at end of loading

When modeled as a non-linear elastic material, a soil is considered past the yield
point when the stress state is within Rf of the soil strength. This consideration
provides an arbitrary means of showing zones that are stressed close to the soil
strength. Figure 13-8 shows the "yield zone" at Load Step 11. For a rigid footing,
this analysis indicates that the soil has not "yielded" below the center of the
footing. This is consistent with the known stress distribution beneath a rigid
footing.
This figure also shows contours of maximum shear stress. In this example, the
shear strength of the soil is 1000 psf, and Rf is 0.9, making the maximum 900 psf.
As shown in the figure, the 900 psf contour is at the edges of the yellow area which
is the area where the strength has reached 900 psf.

Page 268

SIGMA/W

Chapter 13: Illustrative Examples

900

70
0

0
90

30
0

0
50

Figure 13-8 Maximum shear stress contour values and yield zone (step 11)

It should also be noted that the shear stress contours are drawn only for the
foundation material. In this example, the footing material is much stiffer than the
foundation material. Consequently, the stress distribution in the two zones are also
very different. To present a true picture of the stress distribution, only the finite
elements in the foundation are selected when plotting the shear stress contours by
using the View Element Regions command in CONTOUR.

13.4

Elastic-plastic material

Included files

EP undrained.gsz

Page 269

Chapter 13: Illustrative Examples

SIGMA/W

Bearing capacity of a purely cohesive material

This example demonstrates the analysis of a smooth, flexible, strip footing on a


purely cohesive clay. The cohesive strength of the clay, c, is 100 kPa and its
frictional angle, , is zero. Its constitutive behavior can be characterized by an
elastic-plastic stress-strain curve, as shown in Figure 13-9. Since is chosen to be
zero, this is an analysis using the Tresca failure criterion. Also, for a purely
cohesive material, the behavior is independent of confining stresses and the initial
stress state in the soil does not need to be defined. No initial conditions file is
therefore required.

Figure 13-9 Stress-strain curve for an elastic-plastic material

For this analysis, only the displacement convergence criterion is specified and is
set to 1%. The footing load is applied as normal pressure along element edges
using a boundary function. Computation is continued until Step 35 when it is no
longer possible to meet the convergence criterion. It is important to watch the
convergence information during solving of non-linear soil models because it is a
sign that the soil is approaching a failure condition. Once past this point, the
solutions are not reliable. The reasons for this are discussed in the section of
Chapter 3 called Stability and Serviceability.

Page 270

SIGMA/W

Chapter 13: Illustrative Examples

Figure 13-10 shows vertical displacement under center of the footing as a function
of the time step number. The geometry of the example is similar to the previous
footing examples. The maximum displacement at this point is 0.091370m or
91.37mm. This being a flexible footing, the displacement, of course, varies under
the footing. At the outer edge the displacement is only 25.729 mm. With the
application of the last load, the displacement is beginning to increase rapidly
indicating the load is near the ultimate collapse load.

Y-Displacement vs. Time


0.00

Y-Displacement

-0.02

-0.04

-0.06

-0.08

-0.10
0

10

20

30

40

Time
Figure 13-10 Settlement (meters) under center of footing

Page 271

Chapter 13: Illustrative Examples

SIGMA/W

Y-Boundary Force vs.


Y-Displacement

Y-Boundary Force

-500

-1000

-1500
-0.08

-0.06

-0.04

-0.02

0.00

Y-Displacement

Figure 13-11 Load versus deformation plot for footing

Figure 13-11 is the load versus deformation plot for this footing. In this example,
the maximum load at the end is about 1050 kN.
For a footing with a smooth base, which is the case in this problem, the Terzaghi
(1943) formula gives an ultimate bearing capacity of 5.14c (=514 kPa in this
problem). The footing is 2m wide so the ultimate load would be 1020 kN. This is
nearly identical to the SIGMA/W value of 1050 kN obtained for the last converged
solution.
Figure 13-12 shows the contours of deviatoric stress q and the yield zone at Load
Step 21. The deviatoric stress at yield for the Tresca yield criterion is given by:

As one would expect, the yield zone in this figure is approximately bounded by the
170 kPa contour.

Page 272

SIGMA/W

Chapter 13: Illustrative Examples

Figure 13-12 Deviatoric stress contours and yield zone at step 21

This example illustrates that the elastic-plastic model gives reasonable results for a
total stress, un-drained analysis (e.g., when the friction angle phi is zero). The next
example looks at the case of a frictional material.
Bearing capacity of a purely frictional material
Included files

EP Frictional.gsz

EP Frictional_i.gsz

This example presents an analysis of a strip footing on a purely frictional soil (e.g.,
cohesion is zero and phi is some value greater than zero). This could be a sand soil
type, for example, and it is modeled in this case as an elastic-plastic material. The

Page 273

Chapter 13: Illustrative Examples

SIGMA/W

objective here is to show the nonlinearity of a load deformation curve and discuss
some considerations unique to this type of analysis. A secondary objective is to
compare the SIGMA/W results with a closed-form bearing capacity solution.
In this problem the soil friction angle (phi) is set to 30 degrees and the cohesion is
set to zero. Two different cases regarding the soil dilation angle are considered.
One is with a dilation angle of 30 degrees (the same as the friction angle) and the
other is with a dilation angle of zero. When the dilation angle is equal to the
friction angle, an associated flow rule is used in the analysis. When these two
parameters are not equal, the flow rule is said to be non-associated.
For all non-linear constitutive soil models for frictional soils, the soil properties
depend on the stress state in the ground. In general, higher confining stresses mean
the soil is stronger and stiffer. For an analysis such as this, it is therefore necessary
to establish the in-situ stress conditions before applying the footing load. This
needs to be done as a separate preliminary step. The results from this in-situ
analysis will become the initial conditions for the subsequent loading part of the
analysis.
In-situ analysis

The in-situ stresses are established by first running an in-situ analysis. An In-situ 1
type of analysis is suitable here since the ground surface is horizontal. The unit
self-weight is 20 kN/m3, and a coefficient of soil pressure at rest, Ko, of 0.67 is
assumed (this is analogous to a Poissons ratio of 0.4).
It is always good practice to verify that the correct starting stresses have been
established by inspecting the stresses along a vertical profile through the section
with the Draw Graph command. The most crucial variable to check is that the
correct vertical and horizontal effective stresses have been established.
The finite element mesh for these analyses is shown in Figure 13-13. The mesh is
made up of six soil regions and each region is automatically meshed with
unstructured elements. It is worth noting the types of elements used. Experience
has shown that triangular elements with constant stress gradients give the most
reasonable results from a computation perspective when the frictional elasticplastic model is used. Using the automatic meshing feature with full control of
mesh density allows you to concentrate the density of elements where the stresses
are most likely to reach peak values.

Page 274

SIGMA/W

10

11

12

Chapter 13: Illustrative Examples

Figure 13-13 Finite element mesh for footing analysis on a purely frictional
material

After the initial stresses are established, the problem can be saved as a new name
and the loading analysis can be carried out.
For the loading part of the analysis the body load needs to be set to zero, the
analysis type needs to be set to a Load Deformation Analysis, and the footing
pressures need to be defined and applied. The details can be viewed by opening the
respective files to see the parameters used in the analysis.
In this case, the load is applied in increments of 10 kPa.. Since SIGMA/W is based
on an incremental formulation, the fixed action value of 10 kPa will be applied
each time step and accumulated during the loading sequence. There will be 88 load
steps which will result in a final applied pressure of 880 kPa. The results are saved
every fifth step plus the 88th step.
A structural element has been added across footing. This structural element is
stiffer than the soil (E = 2.4e8 kPa, I = 3.5e-5 m4, cross sectional area = 2e-2 m2 )

Page 275

Chapter 13: Illustrative Examples

SIGMA/W

and helps reduce excessive uneven footing deformations by pushing the footing
nodes down more evenly.
Figure 13-14 shows the extent of the yield zone at the end of the loading. The yield
zone under the footing is reasonable as one would expect from pressure bulb
shapes below a footing load and from bearing capacity theory. While not
illustrated in this figure, the yield zone just below the ground surface and well to
the right of the footing is likely not consistent with reality. The reason for the
indicated yield in this area is the low confining stress. The soil has no cohesion,
and consequently as the confining stress approaches zero, the elastic-plastic
relationship indicates yielding even under small changes in shear stress or if there
is a tendency for some tensile stresses. This side effect can often be reduced by
giving the soil a small cohesion which is likely more representative of actual field
conditions.

Figure 13-14 Yield zone at the end of loading

Figure 13-15 shows the load versus deformation curve generated in the Contour
program. Here we see a maximum load force on the entire footing of about 1700

Page 276

SIGMA/W

Chapter 13: Illustrative Examples

kN. If we divide this value by the 2m square surface area of the footing, we have a
bearing capacity on the footing of about 850 kPa. From closed form solutions the
ultimate bearing capacity for a four-meter wide strip footing is 800 kPa (based on a
bearing capacity coefficient of 20). The SIGMA/W computed results have
converged up to and slightly past this theoretical maximum bearing capacity.

Y-Boundary Force

-500

-1000

-1500

-2000
-0.10

-0.08

-0.06

-0.04

-0.02

0.00

Y-Displacement
Figure 13-15 Load versus displacement curve

One thing to note is the curvature of the load-deformation curves. They start out on
a straight line representing linear-elastic conditions, but then begin to bend over
and deviate from the linear elastic condition as they should. These data indicate
that the SIGMA/W formulation provides acceptable results. The trends are
reasonable and logically consistent with a traditional text-book understanding of
soil behavior, and the ultimate loads available are consistent with expectations
from bearing capacity formula.

Page 277

Chapter 13: Illustrative Examples

13.5

SIGMA/W

Strain-softening analysis

Included files

StrainSoft peak.gsz

StrainSoft residual.gsz

StrainSoft.gsz

Figure 13-16 shows a strain-softening stress-strain relationship. The peak and


residual un-drained shear strength are 100 and 50 kPa, respectively. The linear
elastic stiffness is 10,000 kPa, and the slope of the softening portion of the curve is
500 kPa per unit strain.
The stress-strain relationship is used to analyze the load/deformation behavior of a
smooth flexible strip footing 4.0 m wide. The footing load is applied as nodal
displacements. Taking advantage of the geometric symmetry, only half of the
footing is analyzed. This is similar in geometry to the previous footing examples.

Figure 13-16 Strain-softening model stress-strain curve

Three different analyses are carried out. First, the residual strength is set equal to
the peak strength (100 kPa in this example) to obtain the load/displacement curve
for the case with no softening. This in effect makes the analysis similar to an
elastic-plastic analysis. Second, the analysis is repeated with the peak equal to the

Page 278

SIGMA/W

Chapter 13: Illustrative Examples

residual shear strength (50 kPa in this example). Again, the analysis will behave
elastic-plastic, but in this way, it is possible to bracket the range of possible
solutions for the actual strain-softening simulation. The analysis is then repeated a
third time using the strain-softening parameters.
Results of these analyses are presented in Figure 13-17. The upper and lower
curves are essentially elastic-plastic analyses using the von Mises failure criterion.
The center curve is the result for the strain-softening analysis case. It is noted that
the maximum load is between the peak and the residual. The footing pressure is
calculated by summing the Y-boundary force at the footing nodes and then
dividing this sum by the width of the modeled footing (i.e. 2 m). Both Y
displacement and Y boundary force data is extracted by CONTOUR using the
Draw Graph command. The data is then charted using Excel and the graph is
imported into this document.

Figure 13-17 Footing pressure versus displacement curve

The pressure versus displacement results obtained are similar to results obtained by
Chan (1986), who showed that the load-displacement curve for a footing on a

Page 279

Chapter 13: Illustrative Examples

SIGMA/W

strain-softening clay reaches a maximum between the peak and residual values and
remains fairly constant until the displacement becomes very large. The pressuredisplacement curve eventually bends down towards the residual line, but only after
significant footing displacement. In this example, the displacement is
approximately 0.8 m for a 4.0 m wide footing.
According to the closed-form solution by Prandtl (Smith & Griffiths, 1988), the
ultimate bearing capacity is given by:

The ultimate footing pressures of 257 and 514 kPa are calculated for cohesive
strength values of 50 and 100 kPa respectively. When peak and residual values are
used without invoking softening, SIGMA/W computes ultimate values which are 8
to 10% higher than the closed-form solution.
Another point worth noting is that it takes a very large number of small
displacements to achieve the curve for the strain-softening material. Each solid
black square on the curve represents a displacement increment. In addition, the
stress distributions under the footing are quite irregular. However, when we take
the nodal boundary forces along the footing and sum them for the entire footing
load, the results are reasonably good as in the above figure. Stated another way,
overall the results are reasonable, but locally at individual points they can be fairly
erratic.

13.6

Cam-clay triaxial test

Included Files

CamClay ini.gsz

CamClay OCR1.gsz

CamClay OCR5.gsz

This example simulates a triaxial test on Cam-clay soil. The main purpose here is
to verify the SIGMA/W formulation.
The analysis simulates a consolidated-undrained test; that is, the sample is
consolidated by applying a confining pressure and then loaded without any pore-

Page 280

SIGMA/W

Chapter 13: Illustrative Examples

pressure dissipation. The loading is applied in steps at an equal specified vertical


displacement. This is equivalent to a strain-controlled test.
An un-drained test is in essence equivalent to a coupled consolidation analysis, but
with a no-flow boundary around the perimeter. The modeling therefore has to be
performed as a fully coupled consolidation analysis.
Also, to simulate a triaxial test the problem is run as an axisymmetric analysis.
The sample is one meter high and has a half-meter radius (one meter diameter).
Material properties

The material properties are:


= 0.3
= 4.0
= 0.03
= 0.05
= 1.0 ( = 25.4)
Initial conditions

The first step is to establish the consolidated stress state. This is done by applying
a normal pressure of 100 kPa on the top and right side of the mesh as shown in
Figure 13-18 (arrows towards the mesh). The left side of the mesh is the symmetric
axis.
Note the boundary conditions on the left and bottom which represent rollers.

Page 281

Chapter 13: Illustrative Examples

SIGMA/W

1.0

0.8

metres

0.6

0.4

0.2

0.0
0

200

400

600

metres (x 0.001)
Figure 13-18 Triaxial test mesh

The mesh consists of only four elements. They are higher order 8-noded elements
with 9-point integration. Four elements are sufficient for a test like this since the
stress is uniform throughout.
At the end of consolidation there is in essence no excess pore-pressure and there is
no significant hydrostatic pore-pressure. For this numerical analysis the initial
pore-pressures therefore can be zero.
This part of the analysis induces a stress of 100 kPa in each element and becomes
the initial conditions for the loading phase.
Hydraulic conditions

As noted earlier, to simulate an undrained test it is necessary to do a fully coupled


analysis, and for a coupled analysis it is necessary to define some hydraulic
properties. These are defined inside SEEP/W. The two required properties are:

Page 282

SIGMA/W

Chapter 13: Illustrative Examples

Hydraulic conductivity = 1.0 x 10-7 m/sec


Coefficient of volume compressibility mv = 2.0 x 10-3 1/kPa
Mv is the slope of the volumetric water content function, which for this case can be
specified as constant. The absolute water content is not important here; it is the
slope of the function that is important.
This is an undrained test so the hydraulic (SEEP/W) boundary condition around
the entire sample needs to be a no-flow boundary.
The Cam-clay model in SIGMA/W is formulated only for fully saturated conditions.

Initial yield surface

An important issue in an analysis using Cam-clay is the initial yield surface. The
existing or pre-load yield surface must be defined. SIGMA/W has two methods of
doing this. The Over Consolidation Ratio (OCR) option is used here.
The initial yield surface is computed from the following stress states:

y max = y OCR
x max =

y max

(1 sin )

The term (1-sin ) is Ko .


So for a confining stress of 100 and an OCR of 1.0:
ymax = 100
xmax = 57.1
The parameters v0 and p'c can be computed from these stress values. Sufficient
information is then available to draw the initial yield locus.
Normally consolidated test

The results for a test with OCR equal to 1.0 (normally consolidated) are
summarized in Figure 13-19. As it properly should be, the total stress path is a
straight line at a slope of 3V:1H. The effective stress path is vertical until it reaches

Page 283

Chapter 13: Illustrative Examples

SIGMA/W

the initial yield locus. The effective stress then gradually moves up and to the left
until it arrives at the critical state line. The value of deviatoric stress parameter (q)
at the critical state is computed by SIGMA/W to be 54.1 kPa. The corresponding
hand calculated value is 54.6 kPa. Figure 13-19 was plotted using Excel using data
extracted using the Graph in CONTOUR. The initial yield locus was calculated
using Excel.

Figure 13-19 Triaxial test for normally consolidated Cam-clay

The effective stress path can be viewed directly in SIGMA/W as shown in Figure
13-20. The critical state line in this case is a 45-degree line which passes through
the end-point of the stress path.

Page 284

SIGMA/W

Chapter 13: Illustrative Examples

110
100

Deviatoric Stress (q)

90
80
70
60
50
40
30
20
10
0
0

10

20

30

40

50

60

70

80

90

100

110

Mean Eff. Stress (p')

Figure 13-20 Effective stress path

Two other plots of interest are presented in Figure 13-21 and Figure 13-22. Both
the deviatoric stress q and the pore-water pressure u reach a maximum and then
remain essentially constant as they should for a normally consolidated soil.
60

Deviatoric Stress (q)

50
40
30
20
10
0
0.00

0.02

0.04

0.06

0.08

Y-Strain

Figure 13-21 Deviatoric stress versus vertical strain

Page 285

Chapter 13: Illustrative Examples

SIGMA/W

Pore-Water Pressure

80

60

40

20

0
0.00

0.02

0.04

0.06

0.08

Y-Strain
Figure 13-22 Pore-pressure versus vertical strain

Over-consolidated test

Specifying OCR as 5 to simulate an over-consolidated soil gives the following


comparable results (Figure 13-23, Figure 13-24 and Figure 13-25).
Again, the effective stress path is vertical until it meets the yield locus. It then
moves up and to the right, indicating a shrinking yield surface. The SIGMA/W
value of deviatoric stress parameter q at which the effective stress path meets the
critical state line is 206.1 kPa.

Page 286

SIGMA/W

Chapter 13: Illustrative Examples

Figure 13-23 Triaxial test for normally over-consolidated Cam-clay

The shear stress q reaches a maximum of 209.7 kPa and then drops slightly. The
pore-water pressure increases until the effective stress reaches the initial yield
locus. Then it decreases to zero and becomes negative as the effective stress path
moves towards the critical state line. All of these responses are in accordance with
the actual behavior of over-consolidated clay.

Deviatoric Stress (q)

250

200

150

100

50

0
0

50

100

150

200

250

Mean Eff. Stress (p')

Figure 13-24 Effective stress for over-consolidated case

Page 287

Chapter 13: Illustrative Examples

SIGMA/W

Deviatoric Stress (q)

250

200

150

100

50

0
0.00

0.05

0.10

0.15

Y-Strain

Figure 13-25 Deviatoric stress versus vertical strain


60

Pore-Water Pressure

50
40
30
20
10
0
-10
-20
-30
-40
0.00

0.05

0.10

0.15

Y-Strain

Figure 13-26 Pore-pressure versus vertical strain

Page 288

SIGMA/W

13.7

Chapter 13: Illustrative Examples

Modified Cam-clay

Included files

MCamClay ini.gsz

MCamClay OCR1.gsz (OCR is actually 1.25)

MCamClay OCR5.gsz

This example is essentially a repeat of the previous one except that the triaxial tests
simulate a Modified Cam-clay model. The two models are of course very similar
except for the shape of the yield surface. The initial yield locus for the Modified
Cam-clay is elliptical.
Much of the preliminary discussion on setting up initial conditions and so forth are
the same and are consequently not repeated here.
Modified Cam-clay results

A summary of the results for this simulation of a triaxial test based on the
Modified Cam-clay model are presented in the following four figures. The results
have the shape and form as they should.
Also, the point at which the effective stress path meets the critical state line is in
good agreement with the theoretical values as summarized in Table 13-4.
Table 13-4 Intersection of effective stress path and critical state line
Test

Theoretical
(kPa)

SIGMA/W
(kPa)

OCR 1

65.9

66.7

OCR 5

209.1

206.1

Page 289

Chapter 13: Illustrative Examples

SIGMA/W

Figure 13-27 Modified Cam-clay test results for OCR 1.25

Figure 13-28 Modified Cam-clay test results for OCR 5.0

Page 290

SIGMA/W

Chapter 13: Illustrative Examples

250

80

Deviatoric Stress (q)

Deviatoric Stress (q)

70
60
50
40
30
20

200

150

100

50

10

0
0

10

20

30

40

50

60

70

80

90

100

110

50

100

150

200

250

0.20

0.25

Mean Eff. Stress (p')

Mean Eff. Stress (p')

60

80

50

60
Pore-Water Pressure

Pore-Water Pressure

Figure 13-29 Effective stress paths for OCR 1 and OCR 5

40
30
20
10

40
20
0
-20

0
0.00

0.05

0.10

0.15

0.20

-40
0.00

Y-Strain

0.05

0.10

0.15

Y-Strain

Figure 13-30 Pore-pressures for OCR 1 and OCR 5

13.8

Excavation in anisotropic material

Included files

Anisotropic Tunnel.gsz

Anisotropic Tunnel_i.gsz

SIGMA/W can simulate the construction of excavations by removing (nullifying)


elements. An excavation analysis is carried out as a load-deformation analysis, and

Page 291

Chapter 13: Illustrative Examples

SIGMA/W

the elements to be excavated at specific load steps are identified using the Draw
Fill/Excavation Elements command in SIGMA/W Define. The SIGMA/W solver
then calculates the nodal force induced by the stresses inside the elements to be
excavated and applies an equal, but opposite nodal force along the excavation face.
An initial stress state should be specified if elements are to be excavated at the
starting step of an analysis. It is very important to make sure that the initial stresses
are correct. In particular, make sure the initial total stress has any added pore-water
pressure effects prior to excavating. A discussion of the reasons for this is given in
the chapter on Fill and Excavation.
This example shows a tunnel excavation in stratified rock. The rock strata is
assumed to be inclined at 30 to the horizontal. The longitudinal modulus is set to
60 GPa and the transverse modulus is set to 20 GPa. The stratified rock is modeled
using the anisotropic elastic material model.
The first step in an excavation analysis is to establish the in-situ stress conditions.
For the in-situ analysis, the self-weight of the rock is assumed to be 25 kN/m3 and
the coefficient of earth pressure at rest, Ko, is 1. Therefore, the initial stress state is
isotropic. The in-situ stress analysis output file is used as the initial stress file for
the excavation analysis.
In this example, the excavation is carried out in four steps. The number of time
steps is specified using the KeyIn Time Increments command in SIGMA/W.
Elements to be excavated are identified using the Draw Fill/Excavation Elements
command and they appear on screen with an integer number in each element that
shows the load step the element is to be activated (fill) or removed (excavated).
After choosing element type as excavation and entering the time step to excavate,
the desired elements can be selected by clicking and dragging the mouse. By
default, all elements are given a step number of 0, indicating that they will always
be included in an analysis.
The View Preferences command can be used to show the excavation sequence at
each time step. Elements to be excavated are identified with a negative step
number. Figure 13-31 shows the tunnel part the finite element grid used in the
excavation example where the excavation is carried out in four steps. The back
(top portion) of the tunnel is excavated in the first load step, followed by the
remainder of the tunnel in subsequent steps.

Page 292

SIGMA/W

Chapter 13: Illustrative Examples

0
0 0
0

0
0
0

0
0

0
0

-3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3

0
0
0
0

0
0

0
0

0 0
0

0
0
0
0

-3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3

0
-3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3
0
0
-3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3
-3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3 -3
0
0
0
0
-4-4
-4-4
-4-4
-4-4
0
0
-4 -4
-4 -4
0
0
00
-4 -4
-4 -4
00
-4
-4
-4
-4
0
0
0
0
0
0
0
0
0
0
0 0
0
0
0
0 0
0
0
0 0
0
0
0
0 0
0
0
0
0
0
0
0 0
0
0
0
0
0 0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
-1 -1 -1 -1 -1
0
-1 -1
-1 -1
0
-1 -1
-1-1
0
-1-1
-1-1
-1-1
-1
0
-2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2
-2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2
0
-2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2
0
-2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2
0
-2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2 -2

0
0
0
0

0
0

0
0

Figure 13-31 Zoom in image of tunnel excavation sequence

The deformed mesh after complete excavation is shown in Figure 13-32. After
excavation, the circular opening is deformed into an ellipse with its axis coinciding
with the orientation of the anisotropic strata.

Figure 13-32 Deformed tunnel opening after excavation

Page 293

Chapter 13: Illustrative Examples

13.9

SIGMA/W

Embankment construction / consolidation

Included files

Embankment construction.gsz

Embankment consolidation.gsz

The construction of an embankment can be simulated in a manner similar to


constructing an excavation. The difference is that elements are added (activated)
instead of removed (nullified). The load step number assigned to added elements is
positive instead of negative.
The Draw Fill/Excavation Elements command is used to fill elements and
associate them with a step number. This load step number corresponds to the step
at which the element will first become active (i.e., included in the analysis grid).
Figure 13-33 shows a simple finite element mesh where an embankment is
constructed in six placement lifts. The example is also a total stress un-drained
analysis in which pore-water pressures are calculated using the A & B porepressure parameters. Infinite elements are used to model a foundation of infinite
horizontal extent. The B parameter is 1. 0 and the A parameter is 1/3. Making A
equal to 1/3 makes alpha () zero, which means the pore-pressure change is not
dependant on the deviatoric stress changes, but only on the principal stress
changes.
The weight of the fill is specified in using the Body Load command. The load is
applied by assigning a bulk unit weight to the fill material. The foundation material
is assumed to be weightless in this example; which means that only the stress
changes resulting from the fill placement will be available, and consequently only
the pore-pressures change due to embankment loading will be available.
Figure 13-34 plots the deformation along a vertical profile at the center of the
embankment. Notice how the maximum settlement is near the bottom of the
embankment. In this case the assumption is that the fill placement will always be
up to a specified elevation. In other words, enough fill is placed to compensate for
settlement during construction. Only the last lift shows some crest settlement.
SIGMA/W does not correct for the settlement of the last lift. The crest settlement
can be mitigated by using thin layers at the top if this is considered important.

Page 294

SIGMA/W

Chapter 13: Illustrative Examples

40

Embankment Construction

35
+6
+5

+6

+6

+5

+5

+6
+5

+4

+4

+4

+4

+5
+4

+3

+3

+3

+3

+3

+3

+4
+3

+2

+2

+2

+2

+2

+2

+2

+2

+3
+2

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+2
+1

+1

Height (metres)

30
+3
+2

25
20
15

+1

+2
+1

+4
+3

+5
+4

10
5
0
-5
-5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Distance (metres)

Figure 13-33 FEM mesh for embankment construction

SIGMA/W can optionally adjust the weight of each lift to compensate for
settlement from the placement of previous lifts. You can select or disable this
option using the analysis settings command.
40
1.0000e+000

30

2.0000e+000

3.0000e+000
20

4.0000e+000
10
5.0000e+000

0
-0.30

6.0000e+000
-0.25

-0.20

-0.15

-0.10

-0.05

0.00

Y-Displacement

Figure 13-34 Vertical displacement vesus elevation for each load step

In this example, the material properties have been defined using total stress
parameters. If A and / or B pore-water pressure parameters are also specified, it is

Page 295

Chapter 13: Illustrative Examples

SIGMA/W

possible to estimate the pore-water pressure change response due to changes in


total stresses.
Contours of pore-water pressure in the foundation after the completion of the
embankment are shown in Figure 13-35. These contours are limited to the
foundation region by using the View Element Regions command. It should also be
noted that, in this figure, contouring in the infinite elements has been disabled.
Infinite elements are mathematically stretched to infinity. Consequently, when
infinite elements are drawn in their natural scale (as shown in the finite element
mesh), the contours inside these elements will not represent the actual distribution.

Figure 13-35 Contours of excess pore-water after construction phase

The previous analysis outlines the construction phase of the embankment and the
build up of excess pore-water pressure in the foundation soil. It is now useful to
predict the amount of volume change that will occur during consolidation (as
excess pore-pressures are dissipated).
The consolidation volume change can be estimated using the un-coupled
consolidation analysis option in SIGMA/W where the start and ending pore-water
pressure conditions are specified. No estimate of the time to reach a fully
consolidated state is made using this option. If that information is necessary, a
coupled consolidation analysis could be carried out using SEEP/W. The ability to
estimate the volume change can be done with pore-pressure information generated
solely by SIGMA/W.
Since we are interested in estimating the volume change due to dissipation of the
excess pore-water pressure, we are in effect saying that our ending pore-water

Page 296

SIGMA/W

Chapter 13: Illustrative Examples

pressure condition is the same as the condition prior to placing any fill. We also
know that the starting pressure condition is that which exists at the end of the
placement of the last lift. We can obtain both of these conditions directly from the
SIGMA/W project file.
After completing the embankment construction analysis, the file was saved as a
new name, the analysis type changed to un-coupled consolidation, the body load
was turned off, all fill elements were set to permanent, the infinite elements were
set to regular elements, and the load steps set to one. When selecting the pressure
conditions, the ending conditions as PWP=0 and the starting condition as the load
step 6 pressure condition from the construction project file. The PWP=0 option is
valid because we are only interested in volume change due to excess pressure
dissipation and we did not establish any pore-water pressure condition in the
ground prior to placing any fill. So, the u values needed by the solver will be the
excess pressure minus 0.
Figure 13-36 shows the volume change computed by the un-coupled consolidation
analysis where the initial and final pore-water pressure data was obtained from a
previously solved SIGMA/W analysis.
Embankment Consolidation
Vertical settlement: 6.8 cm

40
35

Height (metres)

30
25
20
15
10
5
0
-5
-5

10

15

20

25

30

35

40

45

50

55

60

65

70

75

80

85

90

95

100

Distance (metres)

Figure 13-36 Vertical settlement after embankment consolidation

13.10 Braced excavations


Included files

Excavation with strut.gsz

Page 297

Chapter 13: Illustrative Examples

SIGMA/W

Excavation with strut_i.gsz

This simple example illustrates using a truss element to simulate a crossexcavation brace or strut.
The first step is to do an analysis to establish the in-situ stresses. In a finite element
analysis, an excavation is simulated by applying forces on the excavation face
which represent the stresses that were in the ground before the excavation is made.
To know what forces to apply on the excavation face, we need to know the total
stresses in the ground prior to the excavation. This is the reason for first
establishing the in-situ stresses. This is done as a separate analysis, and the results
become the initial conditions for the excavation part of the problem.
26
24
22
20
18
16
14
12
10
8
6
4
2
0
-2
-5

10

15

20

25

30

35

40

45

50

55

60

65

70

Figure 13-37 Total in-situ Y stress contours and mesh used in strut example

Figure 13-37 shows the computed total vertical stress for the in-situ conditions.
Also shown is the finite element mesh used in the analysis. It is important to note
the use of both structured and unstructured finite elements. The region to be
excavated is made of a structured mesh which is very advantageous because we
want to excavate in even lifts. The remaining mesh can be structured or
unstructured. Also important to note is that even while the problem appears
symmetrical, the entire region is modeled. This is necessary because once the strut
becomes active it must be attached at both sides. If only half the problem was
modeled, the mid point of the strut would not have a valid boundary condition
attached to it because it would be adjacent to Null element nodes.
For this simple example, we'll select some arbitrary values for the brace properties.
E is 1 x 108 kPa and the cross-sectional area is 1 x 10-3 m2. The brace is 10 m long
and there is no pre-force in the member.

Page 298

SIGMA/W

Chapter 13: Illustrative Examples

The problem is set up to excavate the trench in five steps with the strut to become
active on the second step. Figure 13-38 shows the deformed mesh at the end of the
excavation with the strut in place. The deformed mesh is shown at 5x
magnification. It is clear that the strut is holding back deformation into the
opening.
Excavation with a Horizontal Strut
Strut

26
24
22
20
18
16
14
12
10
8
6
4
2
0
-2
-5

10

15

20

25

30

35

40

45

50

55

60

65

70

metres

Figure 13-38 Deformed mesh (5x magnification) at end of excavation

Figure 13-39 shows the horizontal displacement of the left side of the excavation
during the sequence of excavating. The strut becomes active at step two and
virtually limits all further lateral displacement from this time onwards. Figure
13-40 shows the x boundary force at the end of the strut and it confirms a force
was added at step two. The force is not present at step 3 or 4 etc. because it was
already added as a boundary force once and should not be added again.

Page 299

Chapter 13: Illustrative Examples

SIGMA/W

25
0.0000e+000

1.0000e+000

20
Y

2.0000e+000

3.0000e+000

15
4.0000e+000

5.0000e+000

10
-0.02

0.00

0.02

0.04

0.06

0.08

0.10

X-Displacement

Figure 13-39 X displacement versus Y coordinate and load step


30

X-Boundary Force

25

20

15

10

0
0

Time

Figure 13-40 X boundary force versus load step

Page 300

SIGMA/W

Chapter 13: Illustrative Examples

13.11 Consolidation - Cryers problem


Included files

Cryers Ball.gsz

One of the classic examples for the consolidation process is Cryers problem in
which a sphere of saturated clay is subjected to a sudden increase in all-round
confining pressure. Drainage is permitted on the surface of the sphere. Gibson et
al., (1963) have graphed the analytical solution to this problem showing the
variation of the pore-water pressure at the center of the sphere with time for
various Poissons ratios.
A fully coupled consolidation analysis of Cryers problem can be done with
SIGMA/W and the results can be compared with the published information.
Figure 13-41 shows the problem as defined for the SIGMA/W analysis. It is an
axisymmetric analysis and therefore represents a sphere. The sphere is 1 m in
diameter.
The coefficient of volume compressibility mv is 1 x 10-4 kPa-1. The problem is
analyzed as saturated only. This means that E is computed from mv.
E = 1/mv = 1 / 1 x 10-4 = 10,000 kPa.
Poissons ratio v is set to 0.334.
A normal pressure of 100 kPa is applied to the surface of the sphere. The soil is
free to move in the vertical direction along the central vertical axis and is free to
move horizontally along the horizontal central axis.
The hydraulic conductivity is 4.91 x 10-5 m/sec.
The mesh is an unstructured mesh with 6-noded triangular elements. The mesh has
two regions. The outer curved edge of the region has 23 Points. The x-y
coordinates of these points where computed in EXCEL and then pasted into the
Point dialog box in SIGMA/W.

Page 301

Chapter 13: Illustrative Examples

SIGMA/W

Figure 13-41 Sphere used for analysis of Cryers problem

Figure 13-42 shows the pore-pressure response and dissipation at the center of the
sphere. The point of interest is that the pore-pressure rises above the applied
surface pressure and then dissipates. The peak is 118 kPa.
The effect is less pronounced towards the outer surface, but nonetheless exists as
shown in Figure 13-43. The peak is 106, but disappears very quickly.

Page 302

SIGMA/W

Chapter 13: Illustrative Examples

Pore-Water Pressure

150

100

50

0
0

10

Time

Figure 13-42 Pore-pressure response at center of sphere

Pore-Water Pressure

150

100

50

0
0.0

0.1

0.2

0.3

0.4

Time

Figure 13-43 Pore-pressure response close to outer surface

The SIGMA/W solution is compared with a published close form solution in


Figure 13-44, and it is evident from this figure the SIGMA/W solution is
essentially the same as the analytical solution.

Page 303

Chapter 13: Illustrative Examples

SIGMA/W

Figure 13-44 Comparison of SIGMA/W with an analytical solution

In Figure 13-44 the pore-pressure is normalized to the initial applied pressure and
T is the usual consolidation related time factor defined as:

T=

cv t
kE (1 )
t
=
2
w (1 + )(1 2 ) r 2
r

This example helps to confirm that this aspect of SIGMA/W has been formulated
and coded correctly.

13.12 Consolidation analysis of a saturated soil column


Included file

Consolidation 1D.gsz

This example looks at some aspects of fully coupled saturated consolidation.


The problem is a 1m high column as shown in Figure 13-45. The water table is at
the top.

Page 304

SIGMA/W

Chapter 13: Illustrative Examples

1.0
0.9
0.8
0.7

Metres

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Figure 13-45 Problem for one-dimensional consolidation

A 1000 kPa load is applied in Step 1 and then the excess pore-pressure is allowed
to dissipate. The load is applied as a boundary function (Figure 13-46). Recall that
SIGMA/W is an incremental formulation and loads are therefore applied as
increments. The change in the function between Time 0.0 and 1.0 is 1000. For all
other time steps the difference is zero and therefore no further load is applied.

Page 305

Chapter 13: Illustrative Examples

SIGMA/W

0.0

Y-Stress (x 1000)

-0.2

-0.4

-0.6

-0.8

-1.0
0

Time

Figure 13-46 Surface boundary function

The hydraulic conductivity is a constant 5.0 x 10-5 m/min.


The coefficient of volume compressibility is 1 x 10-3. The constrained modulus D
therefore is 1/mv = 1000 kPa. The vertical sides are fixed in the x-direction making
this the equivalent to a constrained test.
From the theory of elasticity the relationship between D and E is:

E=

D (1 + )(1 2 )

(1 )

2
D
3

E therefore is 667 kPa.


An appropriate time stepping sequence is related to mv and the hydraulic
conductivity K. A starting time step can be estimated from:

t =

w mv A
4 K sat

A is the area. The gradient will be the highest during the first time step at the first
element at the surface. Substituting into this equation gives:

Page 306

SIGMA/W

t =

Chapter 13: Illustrative Examples

w mv A 10 (1 e 3) 0.05 0.5
4 K sat

4 (5 e 6)

= 12.5

So a starting time step around 10 minutes will give a reasonable solution. Through
several runs it was learned that an initial time step of 5 minutes actually gives nice
results. This estimation formula is only approximate, but very useful to get within
the right order of magnitude. Time steps that are too small can lead to jagged
unrealistic pore-pressure dissipation curves near the surface. Often several trial
runs are required to establish an appropriate time stepping sequence.
The time steps increase exponentially up to a total elapsed time of 5115 minutes
(about 85 hours).
The dissipation of the excess pore-pressure is shown in Figure 13-47. The porepressure increase is 1000 kPa which is equal to the applied load. There is small
dissipation in the upper 0.2 m (20 cm) during the first load step. The dissipation is
also reflected in the vertical strain as illustrated in Figure 13-48.
1.0
0.0000e+000
5.0000e+000

0.8

1.5000e+001
3.5000e+001

0.6

7.5000e+001
1.5500e+002
0.4
3.1500e+002
6.3500e+002
0.2

1.2750e+003
2.5550e+003
5.1150e+003

0.0
0

200

400

600

800

1000

1200

Pore-Water Pressure

Figure 13-47 Dissipation of excess pore-pressure

Page 307

Chapter 13: Illustrative Examples

SIGMA/W

1.0

0.8

0.6

0.4

0.2

0.0
0.0

0.2

0.4

0.6

0.8

1.0

Y-Strain

Figure 13-48 Vertical strain with time

The same results can be obtained from a SEEP/W analysis if the only interest is to
dissipate the excess pore-pressure. Say that some kind of an analysis has resulted
in an induced excess pore-pressure of 1000 kPa. The excess pressure can be
dissipated in SEEP/W, and all else being the same, the time to dissipate the excess
pressure is the same as from a coupled analysis. The graph Figure 13-49 is from a
SEEP/W analysis and is identical to the coupled results in Figure 13-47. In the
SEEP/W analysis the excess pore-pressure comes from another analysis while in
the coupled analysis the excess pore-pressure comes from the applied load.

Page 308

SIGMA/W

Chapter 13: Illustrative Examples

1.0
0.0000e+000
5.0000e+000

0.8

1.5000e+001
3.5000e+001

0.6
Y

7.5000e+001
1.5500e+002
0.4

3.1500e+002
6.3500e+002

0.2

1.2750e+003
2.5550e+003
5.1150e+003

0.0
0

200

400

600

800

1000

1200

Pressure

Figure 13-49 Dissipation of excess pore-pressure from SEEP/W alone

Water is not always completely incompressible due to included air and


consequently the pore-pressure response is not necessarily 100% of the applied
load. Pore-pressure responses less than 100% can be accommodated by
considering the relationship between mv and E. The reciprocal of mv is referred to
as R. For a fully saturated case when Poissons ratio is 1/3, R is equal to E for a
general 2-D situation. Figure 13-50 shows the pore-pressure responses for a range
of E/R ratios at mid-height of the 1-D example under consideration here. Of
significance is the fact that the pore-pressure response drops off very fast when the
stiffness of the grain structure becomes larger than the stiffness of the water. With
E just 10 percent greater than R, the response already drops to about 87 percent.

Page 309

Chapter 13: Illustrative Examples

SIGMA/W

100

Percent pore-pressure response

90

80

70

60

50

40

30
1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

E/R ratio

Figure 13-50 Pore-pressure response versus E/R ratio

An R/E ratio (reciprocal of E/R) is somewhat analogous to the Skempton porepressure parameter B. An R/E of 0.9 is roughly equivalent to a B value of 0.9.
Controlling the stiffness of the soil grain structure (E) relative to the water stiffness
(mv) makes it possible to compute a pore-pressure response relative to the loading.

13.13 Coupled analysis of unsaturated soils


Included file

Sat-Unsat Consolidation.gsz

The consolidation component in SIGMA/W is formulated for general saturatedunsaturated consolidation. This example looks at a simple 1D problem that has
both a saturated and an unsaturated zone to illustrate the issues and concepts.
SIGMA/W is based on the formulation published by Wong, Fredlund and Krahn
(1998).
A key relationship in the saturated-unsaturated formulation is:

w = v ( ua uw )

Page 310

SIGMA/W

Chapter 13: Illustrative Examples

where w is the volumetric water content, v is the volumetric strain and (ua-uw) is
the soil suction or negative pore-pressure. The volumetric water content w can
change as a result of strain and or a change in suction. The constants and are
related to the soil stiffness and the degree of saturation.
If the soil is saturated, is zero and is 1.0. The change in volumetric water
content then is equal to the volumetric strain. In equation form:

w = v
Saturated consolidation analyses can consequently be performed through the
definition and control of and .
Problem description

This example consists of a 2-m high column (Figure 13-51). The water table is at
mid-height and the initial pore-pressure is disturbed hydrostatically both above and
below water table.
The column is free to move up and down, but is restrained laterally. This is like a
conventional consolidation test in a steel ring.
A 10 kPa load is applied to the surface. This is done through a boundary function
so that the load is applied only in Step 1.
This is like an undrained test and the hydraulic boundary condition around the
perimeter is consequently a no flow boundary (nodal flux is zero; default value).

Page 311

Chapter 13: Illustrative Examples

SIGMA/W

2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0.0

0.2

0.4

0.6

0.8

1.0

Figure 13-51 Problem for saturated-unsaturated consolidation

Material properties

Recall that mv (coefficient of volume compressibility) is the slope of the


volumetric water content function. More correctly, mv is the slope of the
volumetric water content function for positive pore-pressures (saturated
conditions). The slope of the volumetric water content function in the unsaturated
zones is m2w.
For this example, mv is 1 x 10-4 kPa-1. The corresponding constrained modulus D
is, 1/mv which is equal to 10,000 kPa. Youngs modulus E then is 6700 from the
equation:

E=

D (1 + )(1 2 )

(1 )

2
D
3

Page 312

SIGMA/W

Chapter 13: Illustrative Examples

when v is 1/3 as in this example.


The slope of the volumetric water content function for the unsaturated condition is
2.5 x 10-3 and is defined as in Figure 13-52 (0.1 / 40 = 2.5 x 10-3).

Vol. Water Content (x 0.001)

200

180

160

140

120

100
-40

-30

-20

-10

Pressure

Figure 13-52 Volumetric water content function for unsaturated conditions

Further information on the relationship between E and mv is provided in the


Consolidation chapter.
For discussion convenience the unit weight of water is 10 kN/m3.
Computed results

Figure 13-53 shows the pore-pressure immediately after applying the surface
pressure of 10 kPa. Below the water table in the saturated zone the response is 100
percent; that is the pore-pressure increase is equal 10 kPa. In the unsaturated zone
above the water table the response is much less; towards the top there is almost no
response.
One way of looking at this is that below the water table the load is taken up by the
water since water is near incompressible; or the water is much stiffer than the soil
grain structure. At the top the soil grain structure absorbs the load since the voids
are not full of water. Just above the water table there is a smooth transition from
the water absorbing the load to the soil grain structure absorbing the load.

Page 313

Chapter 13: Illustrative Examples

SIGMA/W

2.0

1.5

1.0

0.5

0.0
-15

-10

-5

10

15

20

Pore-Water Pressure kPa

Figure 13-53 Pore-pressure response due to applied load

The excess pore-pressure then dissipates as shown in Figure 13-54. Of interest is


the response that the water pressure decreases below the water table, but increases
(becomes less negative) above the water table.
2.0
0.0000e+000
1.0000e+000

1.5

3.0000e+000

7.0000e+000
1.5000e+001

1.0

3.1000e+001
6.3000e+001

0.5

1.2700e+002
2.5500e+002
5.1100e+002

0.0
-15 -10 -5

5 10 15 20

1.0230e+003

Pore-Water Pressure kPa

Figure 13-54 Dissipation of excess pore-pressure

Page 314

SIGMA/W

Chapter 13: Illustrative Examples

The picture is clear by looking at the pore-pressure at the end relative to the porepressure immediately after the load is applied as in Figure 13-55. The porepressure distribution again becomes hydrostatic, but the position where the new
hydrostatic line crosses the zero pressure line is now higher. This means the water
table now is higher.
2.0
0.0000e+000

1.5

1.0

1.0000e+000

0.5
1.0230e+003

0.0
-15 -10 -5

10 15 20

Pore-Water Pressure kPa


Figure 13-55 Pore-pressures at first and last step

Intuitively this behavior is entirely consistent with what one might expect.
Commentary

At the heart of saturated-unsaturated consolidation is the relationship between the


slope of the volumetric water content function and the soil grain structure stiffness
defined through the E modulus. Stated another way, it is the relative stiffness of the
water phase and the soil grain structure. In the saturated zone the water is stiffer, so
to speak, and in the unsaturated zone the soil grain structure is stiffer.
Conceptually, the saturated-unsaturated consolidation formulation is correct.
Unfortunately, the relationship between m2w and E is yet not well established for
unsaturated conditions. This currently limits the use of this SIGMA/W feature in

Page 315

Chapter 13: Illustrative Examples

SIGMA/W

practice. The feature at this stage is useful for understanding the mechanism, but
no so much for making actual predictions of saturated-unsaturated consolidation
magnitudes.
Moreover, the formulation as shown here gives logical and reasonable results for a
1D analysis. The same is not true for general 2D analyses. More research and
development is required before this feature can be applied to 2D problems.
Uncoupled consolidation analyses should be used when both saturated and
unsaturated zones are present in a 2D field problem.

13.14 Truss example


Included files

Truss.gsz

SIGMA/W has what are called truss elements. They are basically 1D elements
with loads or displacements at the ends. Loads or displacements cannot be
specified along the edge of the truss elements. Physically, these elements can only
have axial loads.
A positive force in a truss element means compression; a negative force means
tension. This is consistent with the sign convention in SIGMA/W for stresses in
ordinary elements.
Two of the main applications of truss elements are cross-excavation braces or the
free length of a tie-back anchor. One of the key features of truss elements is that
they can span across other ordinary elements.
This example is provided to demonstrate that the truss elements function properly.
This can be done by analyzing a simple truss. SIGMA/W is not intended as a tool
to analyze trusses, but such a structure can be analyzed to verify the software. The
following diagram shows a simple truss. This example is taken from the book by J.
L. Meriam, entitled Engineering Mechanics - Statics and Dynamics (SI Version)
published by John Wiley & Sons. The example is on page 134. The forces in each
member of the truss can actually be hand computed by the method of joints
(summation of moments and forces about certain joints).
For the SIGMA/W analysis, the members are all given the same E modulus and
cross-sectional area. The actual magnitude of these values is arbitrary, as long as
they are the same for all the members. The length of each truss member is 5 m.

Page 316

SIGMA/W

Chapter 13: Illustrative Examples

As just one example of a hand calculation, let us compute the force in 2-3. Taking
moments about 1,
5F - (-20*5) - (-30*10) = 0 (The applied forces are negative since they
point in the negative y-direction).
F = (-100 - 300) / 5 = -80 kN (negative means tension).
The SIGMA/W result can be viewed with the View Truss Element Information
command in CONTOUR. The SIGMA/W computed value is -80 kN. The closedform and SIGMA/W forces for all the members are given in Table 13-5 below.
Table 13-5 Truss verification data
Member

Closed-form

SIGMA/W

2-3

-80.00

-80.000

5-6

-34.64

-34.641

4-6

-17.32

-17.318

4-5

+34.64

+34.641

3-5

-34.64

-34.644

1-4

+63.51

+63.509

1-3

+11.55

+11.543

The SIGMA/W values are the same as the closed-form hand calculations
indicating SIGMA/W is computing these values correctly.

Page 317

Chapter 13: Illustrative Examples

SIGMA/W

Figure 13-56 Simple truss example for verification of algorithm

13.15 Tie-back shoring wall


Included files

Tie Back.gsz

Tie Back_ini.gsz

The objective here is to analyze the stability of a temporary shoring tied-back wall
for the proposed design shown in the following diagram. The excavation depth is 9
m and the shoring system consists of sheet piling with two rows of pre-stressed
anchors. The sheet piling will be driven to 3 m below the design excavation base,
and the anchors will be at 3 m and 6 m from the top of the wall.

Page 318

SIGMA/W

Chapter 13: Illustrative Examples

Figure 13-57 Tie back wall geometry

The construction plan is as follows:

Excavate down 4 m to El 15 after the sheet piling has been put in place

Install top row of anchors and stress to the design working load

Excavate another 3 m down to El 12

Install second row of anchors and stress to the design working load

Excavate the remaining 2 m down to the final excavation base

The sheet piling is 12 m long, driven into the ground 3 m below the base of the
excavation. The properties of the sheet piling are:

I = moment of inertia = 5 x 10-4 m4 / m of wall

S = section modulus = 2. 5 x 10-3 m3 / m of wall

A = cross sectional area = 0. 02 m2 / m of wall

E = elastic modulus = 200,000 MPa = 2 x 108 kPa

Anchors will be installed at 2 m spacing along the wall with a working design load
of 200 kN per anchor. Since this is a 2D analysis, the anchor loads must be

Page 319

Chapter 13: Illustrative Examples

SIGMA/W

specified per unit length of wall. This makes the anchor loads 100 kN per meter of
wall in the analysis.
The anchors will be installed at an inclination of approximately 17 degrees. The
top anchors have a bond length of 4.7 m and a free length of 3 m; 7 m for a total
length of 8.4 m. The bottom anchors have a bond length of 4.7 m and a free length
of 2.6 m for a total length of 7. 3 m.
The anchors will consist of 40 mm diameter steel bars, pressure grouted into
150 mm diameter boreholes. The other anchor properties are:

Bar section area 1256 mm2 = 1.26 x 10-3 m2

Steel yield stress 250 MPa

E modulus 200 GPa = 2 x 108 kPa

I (moment of inertia) 2.5 x 10-5 m4 (nail moment of inertia includes the


grout)

Grout skin friction 150 kPa

Unit pull-out resistance 70 kN/m

Design pull-out 47 kN/m (applying a safety factor of 1.5)

Total design pull-out 220 kN (for 4.7 grouted section)

The soil will be modeled as a linear elastic material with E = 5000 kPa and
Poissons ratio equal to 0.334. In a linear-elastic 2D analysis, a of 0.334
equates to a Ko of 0. 5. Recall that:
Ko = / (1- ) = 0.334 /(1 0.334) = 0.5
This value is important when we establish the initial in-situ stress state.
The unit weight of the soil is 20 kN/m3. This is specified in SIGMA/W as a body
load for computing the in-situ stresses. The friction angle of the soil is 32 degrees,
and for design purposes the cohesion will be considered to be zero even though the
soil likely has some cohesive strength component. (The soil strength actually does
not come into play in this linear-elastic analysis; it is used only when the
SIGMA/W stresses are used to look at stability).

Page 320

SIGMA/W

Chapter 13: Illustrative Examples

In-situ stress state

The first step is to establish the stress state in the ground before the sheet piling is
installed and before excavation starts. In this case, it is best done with an Insitu
type of analysis. The original ground surface is level so we can use the Insitu1
option. The soil is assigned a Body Load of 20 kN/m3 , and a Ko of 0. 5. The
boundary along the base is fixed (zero x and y displacements). Along both ends,
vertical movement is allowed, but lateral movement is not allowed. The ycondition is specified as none which means the node is free to move up or down.
Eight-noded elements are used throughout, since they give better stress
distributions within the elements than 4-noded elements. Four-point Gauss
numerical integration is adequate in this case, as opposed to the nine-point
integration.
The following figure shows the vertical stresses for a portion of the section.
Elevation 3 m is 16 m below the ground surface. The y-stress is 320 kPa which
correctly matches H which is 16 m x 20 kN/m3 = 320 kPa.

Figure 13-58 Initial in situ vertical stresses

You can use the Draw Mohr Circle feature in CONTOUR and click on any node to
spot check the results. You will see that the horizontal stresses are half the vertical

Page 321

Chapter 13: Illustrative Examples

SIGMA/W

stresses as they should be for Ko equal 0. 5. The x-y shear stresses are essentially
zero making the vertical and horizontal stresses the principal stresses as they
correctly should be. It is vitally important that you ensure that the horizontal
stresses are reasonable, since these stresses come into play on the wall when the
ground is excavated.
Simulation of construction procedure

The construction will be simulated in five steps. This means we need to define a
time sequence of 5 steps.
The next task is to flag the elements that will be removed to simulate the
excavation stages. This is done with the Draw Fill/Excavation Elements command.
Looking at the problem definition, when you open the file, you will notice some
elements marked with a 0, -1, -3, and -5. The 0 elements are permanent. The
elements with -1 are removed in Step 1, the ones with -3 are removed in Step 3 and
the elements with -5 are removed in Step 5.
The grouted portions of the anchors are simulated as beam elements. The grouted
portion of the upper anchor will become active in Step 2.
In finite element modelling terms, the pre-stressing of the anchor is achieved by
applying boundary forces at the ends of the free lengths of the anchors. SIGMA/W
applies these forces based on the pre-stress force specified as part of the bar
element definition to simulate the anchor free (unbonded) length. A very important
point is that the pre-stress force must not be applied at the same time as the bar
element becomes active. The pre-stress forces cannot act directly on the bar
element. One way to think of this is that the pre-stress forces are applied first and
then they are locked into the bar element in the next step. In this example, the
upper pre-stress force is applied in Step 2 and the bar element representing the free
length becomes active in Step 3. With this background explanation, the simulation
activities in each of the steps are:

Step 1 excavate upper 4 m

Step 2 activate upper grouted bond length and apply upper anchor
pre-force

Step 3 activate upper free length (bar element) and excavate next
3m

Page 322

SIGMA/W

Chapter 13: Illustrative Examples

Step 4 activate lower grouted bond length and apply lower anchor
pre-force

Step 5 activate lower free length (bar element) and excavate final
2m

You can check the specified values and activation times with the View Edge
Information command to look at the grouted (beam) lengths and the View Bar
Element Information to look at the bar elements.
Results

There are many, many ways of looking at the results. Only some are presented here
to illustrate what can be done and to provide a reference for checking that the
software is working properly.
Figure 13-59 shows the force in the free length of the upper anchor. The force is
zero until Step 2 when the 100 kN pre-force is applied. Then the force increases to
150 kN as the second lift is excavated; then the force decreases when the lower
anchor is pre-stressed and finally the force again increases when the last lift is
excavated.

Total Axial Force

-50

-100

-150
0

Time

Figure 13-59 Force in free length of upper anchor

Page 323

Chapter 13: Illustrative Examples

SIGMA/W

Lateral wall movement

The following diagram (Figure 13-60) shows the lateral displacement on a vertical
profile through the wall. The displacement is in meters.
Notice how the wall gets pulled back with the application of the upper pre-stress
load (Step 2) and then moves back out with the next excavation stage (Step 3).
Applying the lower pre-stress has the same effect, but not nearly as significant.

20
1.0000e+000

15

2.0000e+000

10
3.0000e+000

4.0000e+000

0
0.00 0.02 0.04 0.06 0.08 0.10

5.0000e+000

X-Displacement
Figure 13-60 Lateral wall movement

Moment distributions in the sheet piling

The next diagram (Figure 13-61) shows the moment distribution in the sheet piling
at the five different construction stages. The maximum moment is about
125 kN-m.

Page 324

SIGMA/W

Chapter 13: Illustrative Examples

15

Distance

1.0000e+000

10

2.0000e+000

3.0000e+000

5
4.0000e+000

0
-50

5.0000e+000

50

100

150

Moment
Figure 13-61 Moment distributions in the sheet piling

Axial force in grouted length

The final axial force in the grouted portion of the upper anchor is shown in Figure
13-62. Worth noting is that the maximum axial load (about 120 kN) is less than the
140 kN in the free length portion. The reason for this is the interaction between the
grouted length and the soil. The soil is absorbing some of the energy.
The axial force is shown as leveling off at the ends. This is due to a numeric
boundary effect in the beam formulation. The formulation is based on a constant
strain in each element of the beam. For the inner structural elements is not
noticeable due to averaging to the nodes. It is noticeable only for the end elements.
It is the same as four 4-noded elements when insitu stresses are computed. The
first row at the surface and the last row at the bottom have constant stresses, which
cause a stress profile to deflect at the top and at the bottom. Without these end
effects the maximum axial load is 140 kN as in the free length.

Page 325

Chapter 13: Illustrative Examples

SIGMA/W

-20

Axial Force

-40

-60

-80

-100

-120
0

Distance

Figure 13-62 Axial force in the upper grout length of the anchor

Lateral stress distributions

The final diagram (Figure 13-63) here shows the lateral (x) stress distribution at
the end of construction. Notice how the stress contours are affected by the presence
of the anchors.

40

60

80

40
60
80

Figure 13-63 Lateral stress distribution at the end of construction

Page 326

SIGMA/W

Chapter 13: Illustrative Examples

13.16 Embankment on soft ground


This example is about the construction of an embankment on soft ground. The
embankment is constructed in stages with periods between the stages for some of
the excess pore-pressure to dissipate. This requires a fully-coupled consolidation
analysis.
Included Files

Soft Ground_ini.gsz

Soft Ground.gsz

Problem description

The subsurface clay stratum is 9 m deep and the water table is 1 m below the
ground surface. The 1 m layer above the water table is highly weathered,
desiccated and fissured making its behavior similar to a fine granular soil. The
embankment to be analyzed is shown in the following figure and is constructed of
relatively sandy soil. The height of the embankment is 5 m with 3h:1v side slopes
and a 10 m crest width. Due to symmetry about the center, only half of the cross
section is used in the analysis.
15
14

13

Elevation - metres

12
11
10
9
8
7
6
5
4
3
2
1
0
0

10

15

20

25

30

35

40

metres

Figure 13-64 Embankment of soft consolidating ground

The objective of this example is to outline the correct modeling techniques for
determining the embankment stability and settlement.
The construction of an embankment can be simulated in a manner similar to
constructing an excavation. The difference is that elements are added (activated)

Page 327

Chapter 13: Illustrative Examples

SIGMA/W

instead of removed (nullified). The time step number assigned to added elements is
positive instead of negative.
Clay properties

The clay is modeled as a Modified Cam-clay (MCC) soil. It is highly plastic with a
liquid limit (LL) of 61%. The compression index Cc can be estimated from the well
known equation:
Cc = 0. 009 (L.L. - 10 percent)
= 0. 009 (61-10)
= 0. 46
The slope of the normal consolidation line can be computed from Cc / 2. 303);
this makes equal to 0.20.

The ratio of to (slope of the swelling line) is 5, making equal to


0.04.

The coefficient of volume compressibility mv is 3 x 10-4 (1/kPa).

The saturated coefficient of permeability or hydraulic conductivity is


5.8 x 10-6 cm/sec or 5 x 10-3 m/day.

The effective friction angle is 26 degrees. The slope of the critical


state line is therefore equal to 1.0.

With = 26 degrees, Ko is 0.56, which is estimated from Ko = (1 sin


).

For linear-elasticity, a Ko of 0.56 equates to a Poissons Ratio, , of


0.36. In a MCC model, this value is required for elastic conditions
under the yield surface.

The over-consolidation ratio (OCR) is 1.2.

The specific volume is 2.2.

A summary of the properties above is provided in the following table.

Page 328

SIGMA/W

Chapter 13: Illustrative Examples

Table 13-6 Summary of clay properties


Property

Value

Cc (compression index)

0. 46

(slope of normal consolidation line)

0. 20

( slope of swelling line)

0. 04

mv ( coefficient of volume compressibility

3 x 10-4 (1/kPa)

k ( hydraulic conductivity)

5 x 10-3 m/day

( effective friction angle)

26 degrees

( slope of critical state line)

1. 0

Ko (coefficient of earth pressure at rest)

0. 56

( Poissons Ratio)

0. 36

OCR (over consolidation ratio)

1. 2

(specific volume)

2. 2

Sand properties

The embankment and upper meter of subsoil will be modeled as highly permeable
linear-elastic material, which is considered adequate since the major settlement
issue arises in the underlying compressible clay. The properties adopted for the
sand are:
E = 2000 kPa
= 0.36
k = 1.0 m/day
mv = 3 x 10-4 kPa-1
The k and mv values are actually arbitrary for the sand, since the pore-water
pressure conditions will be specified, meaning there is no change with time. In a
transient (consolidation) analysis, the software requires these values be specified,
but the actual values do not affect the results.
In-situ stresses

It is essential when using the Modified Cam-Clay (MCC) soil model to establish
the initial yield surfaces. Any changes due to the embankment loading are relative

Page 329

Chapter 13: Illustrative Examples

SIGMA/W

to the yield surfaces that exist prior to the loading. The initial yield surface is
related to the initial insitu stress and the over-consolidation ratio (OCR). The OCR
is specified as a soil property, but the insitu stress must be computed in a separate
step.
The finite element mesh must be complete at the start, but the elements
representing the embankment are ignored during the insitu stress analysis by
giving the embankment elements a soil model set to "none".
The two soils representing the subsurface are given a body load equal to the unit
weight of 20 kN/m3, and a Ko equal to 0.56. The water table is specified as a
horizontal line 1 m below the ground surface.
The unit weight of water has been set to 10 kN/m3. This makes it convenient to
check that the correct total and effective stresses and pore-water pressures have
been computed correctly. In real field problems, the unit weight of water would of
course be 9.81 kN/m3.
The bottom of the problem is fixed, while both vertical ends are allowed to move
vertically, but not horizontally.
When establishing the insitu stresses, it is important to select "Insitu" as the
analysis type. Selecting a Load/Deformation analysis type would result in wrong
horizontal effective stresses. The following figures show the total and effective
horizontal stress profiles in the subsoil before starting the embankment fill
placement.
The effective horizontal stress should be:
h

(t x 1m) Ko + ((t - w) Hw) Ko

(20 x 1)0.56 + ((20 -10) 8)) 0.56

56.0 kPa

The total horizontal stress is computed as h + u corresponding to 56 + 80 =


136 kPa.
The horizontal stress plots below (Figure 13-65) confirm these values.

Page 330

SIGMA/W

Chapter 13: Illustrative Examples

6
Y

10

10

0
0

50

100

150

X-Total Stress

10

20

30

40

50

60

X-Effective Stress

Figure 13-65 Horizontal in situ stress profiles

Note that in the Modified Cam-Clay definition dialog box that pc is not defined.
When this is value undefined, SIGMA/W uses the OCR together with the insitu
stresses to compute pc at each Gauss point in each element. The initial yield
surface is consequently defined at each Gauss stress Loading Sequence.
Loading sequence

The embankment fill is going to be placed in five stages as follows:


Day

Fill Lift

1st

2nd

13

3rd

19

4th

25

5th

The analysis will be run for 40 days. Figure 13-66 figure shows the embankment
with the numbers in each row of elements indicating at what time (day) the
elements become active. The clay will consolidate during the time between each
lift placement and for 15 days after the last lift is placed.

Page 331

Chapter 13: Illustrative Examples

SIGMA/W

+25 +25 +25 +25 +25 +25 +25


+25
+19 +19 +19 +19 +19 +19 +19 +19 +19 +19
+19
+13 +13 +13 +13 +13 +13 +13 +13 +13 +13 +13 +13
+13 +13
+7 +7 +7 +7 +7 +7 +7 +7 +7 +7 +7 +7 +7 +7 +7 +7
+7
+1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1 +1
+1
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Figure 13-66 Time sequence for fill loads

The time sequence in a consolidation analysis is controlled by SIGMA/W. In any


transient analysis it is important to start with a reasonable time step, which can be
estimated from:

t =

w mv A
4 K sat

where:
A

the element area.

During the initial time step, the size of the element area (A) where the
consolidation will be the greatest (just below the water table) is roughly one square
meter. Correspondingly, the initial time step should be approximately 1 to 2 days.
A value of one day is used as the initial time step for this problem.
A useful thing to remember is that a reasonable starting time step is roughly mv divided by
Ksat.

Data specified in SEEP/W

Important parameters and conditions to specify in SEEP/W include the coefficient


of volume compressibility mv, the hydraulic conductivity and the hydraulic
boundary conditions.
In SEEP/W the slope of the volumetric water content function is mv. In this
example, consolidation of the clay will occur under saturated conditions, so the
volumetric water content function can be defined as a straight line with two points

Page 332

SIGMA/W

Chapter 13: Illustrative Examples

resulting in a line with a slope of 3 x 10-4 (kPa-1). Alternatively, mv can be entered


as a constant.
The hydraulic conductivity of the clay will be constant and equal to the saturated
hydraulic conductivity, defined by a horizontal function at 5 x 10-3 m/day.
The problem is analyzed with the Saturated Zone Only option selected. This
means that consolidation is computed in the saturated zone below the water table.
When this option is selected, SIGMA/W automatically sets the hydraulic boundary
condition in the unsaturated zone to a zero water pressure condition. SIGMA/W
does this by setting the boundary condition equal to the elevation (y-coordinate) at
each node in the unsaturated zone.
Under the water table the boundary condition along the perimeter is a no-flow
boundary. The no-flow boundary does not necessarily have to be specified since
this is the default in any seepage analysis.
The effect of specifing a head at all nodes in the unsaturated zone is illustrated in
Figure 13-67. Physically, the significance of specifying the head boundary is that
there will be no volume change due to a pore-pressure change.

Figure 13-67 Effect of saturation-only option

Summary of the results

The computed results can be viewed in several different ways. Only a few
approaches are presented here to illustrate some of the highlights. Many additional
graphs and contour plots can be viewed by using the CONTOUR option.
Figure 13-68 shows the pore-water pressure changes with time at a point one meter
to the right of the centerline and two meters below the water table. The pore-water

Page 333

Chapter 13: Illustrative Examples

SIGMA/W

pressure immediately rises after the placement of a lift and then falls until the next
lift is placed. The pore-water pressure drops from a high of 90 kPa down to about
55 kPa after the final lift has been placed. Significant excess pore-water, however,
remains at the end of 40 days indicating that additional long-term settlement will
occur.
Figure 13-69 shows the settlement that occurs at the center-line at the water table
elevation. Note how some settlement occurs immediately as the lifts are placed and
additional settlement occurs gradually during the consolidation phases. To be
precise, there is also some settlement due to consolidation during the one-day
interval that the lift is placed, but most of it is due to the additional load.

Pore-Water Pressure

100

80

60

40

20
0

10

20

30

Time
Figure 13-68 Change in pore-pressure with time

Page 334

40

SIGMA/W

Chapter 13: Illustrative Examples

Y-Displacement

0.0

-0.2

-0.4

-0.6

-0.8
0

10

20

30

40

Time
Figure 13-69 Immediate and consolidation settlement with time

Settlement profiles along the original ground surface on Days 0, 7, 13, 19, and 25
are shown in Figure 13-70. These are the days that the fill was placed. Of
particular interest is the location of maximum settlement. Note that the maximum
does not occur along the centerline, but is located more toward the outer edge of
the fill in the early stages of loading. At the end of 40 days the maximum is under
the center of the embankment, but not during the construction.
The fact that the maximum settlement is not under the center of the embankment
during the early stages of construction is due to the two-dimensional dissipation of
the excess pore-pressure. There is some lateral flow towards the toe area which
does not exist under the center. The excess pore-pressure relative to the load is also
the highest in the maximum settlement area as is evident from the low effective
horizontal stresses depicted in Figure 13-71. They are the lowest in this area.
In addition, the settlement towards the toe area is aided by the uplift outside the
toe.

Page 335

Chapter 13: Illustrative Examples

SIGMA/W

0.1

Y-Displacement

0.0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
0

10

15

20

25

X distance
Figure 13-70 Settlement profiles along the original ground surface

25

40
50

Figure 13-71 Horizontal effective stresses at Step 7

Page 336

SIGMA/W

Chapter 13: Illustrative Examples

The results clearly show the advantages of the staged construction. Without the
staged construction, the excess pore-pressure would be much higher and could
potentially lead to failure during construction. This could be easily demonstrated
by rerunning the included files after all the fill elements have been assigned a
Step 1 activation flag.

Page 337

Chapter 13: Illustrative Examples

SIGMA/W

Page 338

SIGMA/W

Chapter 14: Product Integration

14

Product Integration

14.1

SEEP/W generated pore-water pressures in SLOPE/W


stability analysis

The objective of this illustration is to observe how to use finite element pore-water
pressure results in a stability analysis. Including or deliberately ignoring negative
pore-water pressures can be critical to understanding and interpreting a slope
stability analysis.
In particular, the objectives of this illustration are to:

Set up and solve a steady-state finite element SEEP/W simulation. In


CONTOUR show the positive pressure heads that develop.

Set up a slope stability problem in SLOPE/W based on the SEEP/W


finite element mesh and computed pore-water pressure; determine the
critical slip surface and the factor of safety using the SEEP/W porewater pressure; and graph the pore-water pressure and strength along
the slip surface.

Repeat the analysis, but remove the advanced parameters from the soil
property information. Graph the pore-water pressure and strength
along the slip surface and note how the negative pore-water pressures
have been ignored).

Repeat the stability analysis, using the piezometric line option in


SLOPE/W to reflect a water table at an elevation of 10 m. In this case
we will not use advanced parameters.

The seepage portion of the analysis is illustrated in Figure 14-1. In general, the
slope is comprised of multiple layers with a finer, lower permeability layer located
half way up the slope face. Note that more coarse soil soils in region 1 and 3 have
the same hydraulic properties (Ksat 1x10-3 m/day). In addition, a steady-state
infiltration rate of q = 5.0x10-5 m/day is applied along the top and the slope with a
pressure equals zero condition on the downstream surface. A potential seepage
review has been applied along the face of the slope.

Page 339

Chapter 14: Product Integration

SIGMA/W

26
24
22

Infiltration 5.0e-5 m/day

20

K(sat) = 1e-3 m/day 1

18

Seepage Face

16

K(sat) = 1e-5 m/day

14
12

K(sat) = 1e-3 m/day

Pressure = zero

10

8
6
4
0

10

20

30

40

50

60

Figure 14-1 Seepage problem definition

The soil property information for the SLOPE/W portion of the analysis is given in
Table 14-1.
Table 14-1 Slope soil information
Layer

Unit Weight

Phi

Cohesion

Unit Weight
above WT

Phi B

18

25

18

15

19

20

19

15

20

30

10

20

15

Figure 14-2 shows the location of grid and radius points applied in the stability
analysis, as well as the SEEP/W computed perched water table and the computed
factor of safety. SLOPE/W was able to read the seepage results directly from
SEEP/W in order to compute the actual pore-water pressures at the base of each
slice. Figure 14-3 shows the actual pore-water pressures applied on each slice.
Notice how the pore pressures on the slices change from negative to positive to
negative and back to positive as the slice number increases from left to right. This
type of pore-water pressure condition would not have been possible to accurately
establish without the use of a rigorous saturate-unsaturated seepage flow model.
Note in the SLOPE/W results that the finite element mesh used in the seepage
analysis is superimposed on the solution. The finite element mesh is not actually
used by SLOPE/W, but the data from the mesh at all locations can be used to
determine pore pressures on the base of any slice geometry.

Page 340

SIGMA/W

Chapter 14: Product Integration

44
42
11

40
38
36
34
32

1.324

30

2
28

12

13

26
24

3
22
1

20

Infiltration

18
8

16

14

10

12
3

10

14

8
6

4
0

10

20

30

40

50

60

Figure 14-2 SLOPE/W solution showing perched water table from SEEP/W
Pore-Water Pressure vs. Distance
30
25
20

Pore-Water Pressure

15
10
5

0
-5
-10
-15

-20
0

10

20

30

40

Distance

Figure 14-3 SEEP/W pore pressures at base of slices in SLOPE/W

14.2

VADOSE/W generated pore pressures in SLOPE/W


stability analysis

The objective of this illustration is to observe how to use climate coupled finite
element pore-water pressures results in a stability analysis. You will see how
important including negative pore pressure can be to a slopes stability.

Page 341

Chapter 14: Product Integration

SIGMA/W

The VADOSE/W analysis mesh as well as the SLOPE/W grid and radius are
illustrated in Figure 14-1. In general, a large rainfall event was applied over the
undulating ground profile on the first day of a 30 day long transient climate
coupled VADOSE/W analysis. This heavy rainfall created ponded water
conditions in the low point near the center of the mesh. The rainfall day was
followed by 29 days of infiltration from the pond and evaporation through the
climate-ground boundary. Stability results are compared for the day just after the
rainfall (e.g., day 2 from VADOSE/W) as well as after a long drying period (e.g.
day 30). A comparison is also made between using and not using soil strength
parameters that account for increased strength with increased negative water
pressure (e.g. Phi-B).
The soil property information for the SLOPE/W portion of the analysis is given in
Table 14-1. Two stability simulations were carried out: one with a Phi B parameter
defined to account for the effect of soil water suction on strength and one without
Phi B.
Table 14-2 Slope soil information
Layer

Unit Weight

Phi

Cohesion

Unit Weight
above WT

Phi B

18

25

18

15

Figure 14-4 Stability problem definition (pwp from end of day 2 in


VADOSE/W)

Page 342

SIGMA/W

Chapter 14: Product Integration

Figure 14-2 shows a minimum factor of safety of 1.16 in the stability analysis
based on advanced pore-water pressure parameters (e.g., soil suction strength
effects) as well as the VADOSE/W computed water. SLOPE/W was able to read
the seepage results directly from VADOSE/W in order to compute the actual porewater pressures at the base of each slice.
Figure 14-6 shows the computed factor of safety after the heavy rainfall day based
on soil property strength data that does not include the effect of suction on the
strength of the soil. In this analysis the factor of safety is reduced to less than 1.0.
One final analysis was carried out using advanced strength parameters to
determine the factor of safety after an extended drying period. Figure 14-7 shows a
factor of 1.74 after 29 days of infiltration from the pond (e.g. no rainfall) and
surface evaporation on the ground profile. It is clear that from this example that the
factor of safety is very dependent on ground-climate influences and on the
application of advanced soil strength parameters that take into account increased
strength due to negative water pressure in the soil (e.g. soil water suction).
Finally, Figure 14-8 shows the actual pore-water pressures applied on each slice in
the stability analysis for day 30 of the simulation. Notice how the pore pressures
on the slices change from positive to negative as the slice number increases from
left to right. This type of pore-water pressure condition would not have been
possible to accurately establish without the use of a rigorous saturate-unsaturated
climate coupled seepage flow model.
Note in the SLOPE/W results that the finite element mesh used in the seepage
analysis is superimposed on the solution. The finite element mesh is not actually
used by SLOPE/W, but the data from the mesh at all locations can be used to
determine pore-water pressures on the base of any slice geometry.

Page 343

Chapter 14: Product Integration

SIGMA/W

1.159

Figure 14-5 SLOPE/W solution using water pressures from VADOSE/W


(solution including suction effect on strength)

0.999

Figure 14-6 SLOPE/W solution using water pressures from VADOSE/W


(solution NOT including suction effect on strength)

Page 344

SIGMA/W

Chapter 14: Product Integration

1.739

Figure 14-7 Factor of safety after long drying period (e.g. day 30 pore-water
pressure data from VADOSE/W)

Pore-Water Pressure vs.


Distance
Pore-Water Pressure

10
0
-10
-20
-30
-40

Distance

Figure 14-8 Day 30 VADOSE/W pore pressures at base of slices in SLOPE/W


(for simulation with suction effects on strength)

Page 345

Chapter 14: Product Integration

14.3

SIGMA/W

SEEP/W dissipation of pore pressures generated in a


QUAKE/W earth quake analysis

QUAKE/W can be used to analyze the transient build up of pore-water pressure


during an earthquake event. Consider the illustration in Figure 14-9 that shows a
phreatic surface in a dam with hydraulic fill on two sides of a core. The horizontal
acceleration earthquake record for this example is taken from the Pacoima dam
earthquake event and is illustrated in Figure 14-10. Note the duration of the
earthquake was only 40 seconds, yet during this time the pore pressures at point
A in Figure 14-11 increased significantly. The computed pore-water pressures at
this point during the entire earthquake are illustrated in Figure 14-12. In general,
the shaking almost doubled the pressure over pre-quake hydrostatic conditions.
1.14

(x 1000)

1.11
1.08
1.05
1.02
0.99
0.96
0.93
0.90
-500

-400

-300

-200

-100

100

200

300

400

Distance - feet
Figure 14-9 Dam pore-water pressure prior to earthquake

Page 346

500

600

SIGMA/W

Chapter 14: Product Integration

Pacoima dam earthquake


0.6
0.5
0.4

Acceleration ( g )

0.3
0.2
0.1
0.0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
0

10

20

30

40

50

Time (sec)

Figure 14-10 Horizontal acceleration record

Figure 14-11 Pore-water pressure contours after earthquake. Note point A

Page 347

Chapter 14: Product Integration

SIGMA/W

Pore-Water Pressure vs. Time

Pore-Water Pressure

9000

8000

7000

6000

5000
0

10

15

Time

Figure 14-12 Pore-water pressure (psf) build up at point A during quake

The QUAKE/W generated pore-water pressures can be used directly in SEEP/W as


the initial conditions in order to determine the time it takes for the quake generated
pressures to return to hydrostatic conditions. The same pore-water pressures
illustrated above can also be used directly in SLOPE/W in order to determine the
factor of safety versus time throughout the quake event. This is discussed in more
detail in the SLOPE/W engineering book.
Initial pressures to come from QUAKE/W

1.14

(x 1000)

1.11
1.08

Seepage face (Q=0)

Head BC = 1110 ft

1.05

P=0

1.02
0.99
0.96
0.93
0.90
-500

-400

-300

-200

-100

100

200

300

400

500

600

Distance - feet

Figure 14-13 Model set up for SEEP/W pressure dissipation analysis

Figure 14-13 shows the model set up for the SEEP/W analysis that will compute
the dissipation of the excess pore-water pressure generated during the earth quake.
The boundary conditions for the seepage analysis are illustrated in the figure. In

Page 348

SIGMA/W

Chapter 14: Product Integration

this analysis it is assumed that the reservoir remains full with a total head of 1110
ft (or 102 ft pressure head). The seepage analysis was carried out to simulate 5
days and the change in pore-water pressure at point A over time is illustrated in
Figure 14-14 below.
Pressure vs. Time
9000

Pressure

8000

7000

6000

5000
0

100000

200000

300000

400000

500000

Time

Figure 14-14 Dissipation of excess pore pressure at point A over 5 days


following quake

14.4

SEEP/W velocity data in CTRAN/W contaminant transport


analysis

Seepage data from a steady state or transient SEEP/W analysis can be used in the
GeoStudio CTRAN/W module to predict contaminant transport with or without
adsorption, diffusion and decay. This is a one-dimensional contaminant transport
problem verification example analysis with a free exit boundary. The CTRAN/W
results are compared with closed form analytical and published solutions.
To generate a one-dimensional steady-state flow system, a one-row finite element
mesh is created using SEEP/W DEFINE. The finite element mesh is 1 m high and
40 m long and consists of a total of 30 elements and 62 nodes. The SEEP/W files
containing this example are named EXIT.GSZ.

Page 349

Chapter 14: Product Integration

SIGMA/W

The head differential and hydraulic conductivity are selected to produce a constant
seepage velocity U of 0.05 m/s in the positive x-direction. The volumetric water
content is defined as a constant value of 0.5. The average linear velocity is:

=U /
= 0.05 / 0.5
= 0.1 m/s
The dispersivity L is set to 4 m, and the molecular diffusion coefficient D* is set
to zero. The resulting hydrodynamic dispersion coefficient is:

D = Lv + D *
= 4 0.1 + 0.0
= 0.4 m 2 /s
The time step sequence consists of 48 steps. Results are presented for Time Steps
8, 16, 24, 32, 40 and 48 with total elapsed times of 80 s, 160 s, 240 s, 320 s, 400 s
and 480 s respectively. Computed results for these six time steps are included with
the CTRAN/W software.
The boundary condition at the left end of the problem is assumed to be a source
boundary with concentration of the source Cs specified as 1.0 unit/m3. The
boundary condition at the right end is specified as a free exit boundary (Qd > 0).
The initial concentration of the flow system is set to 0.0.
Frind, (1988) has presented an analytical solution to the transport equation with a
free exit boundary. The solution is approximated by the analytical solution for
transport in a semi-infinite medium. The concentration C as a function of x and t is
expressed as:

C =

Cs
2

x - vt
erfc
2 Dt

x - vt v( x +vt )
vx
exp D erfc
1 + D

2 Dt

( x vt ) 2
v t
+
exp

4 Dt
D

where:

Page 350

SIGMA/W

Chapter 14: Product Integration

concentration,

Cs

specified concentration of the source,

hydrodynamic dispersion coefficient,

average linear velocity,

elapsed time,

distance from the source boundary, and

erfc

complementary error function.

In this verification example, no adsorption and no decay are considered. Figure


14-15 presents the CTRAN/W solutions using the central difference time
integration scheme (contained in the EXIT1 files). A comparison of the
CTRAN/W solution with the analytical solution using a free exit boundary is
presented in Figure 14-16. The Excel comparison indicates that the CTRAN/W
solution is in excellent agreement with the analytical solution. Almost identical
solutions are obtained for all three elapsed times in the first 35 m, while there are
only slight differences between 35 m and 40 m. The differences are likely due to
the approximation of the free exit boundary using a semi-infinite medium in
developing the closed form solution.
Figure 14-17 presents the CTRAN/W solution using the central difference time
integration scheme for the case with a zero dispersive mass flux exit boundary
condition (contained in the EXIT2 files). Figure 14-18 presents a comparison of
the CTRAN/W solution using a free exit boundary, with the CTRAN/W solution
using an exit boundary of zero dispersive mass flux (i.e., Type II boundary). As
expected, the concentration profile in the upstream portion is not sensitive to the
exit boundary conditions. However, there are significant differences in the
concentration profiles near the exit boundary. Since a zero dispersive mass flux
implies zero concentration gradient at the exit boundary, the nodal concentration at
the exit boundary has been forced to be the same as the nodes immediately to the
left of the exit boundary.
Both Figures below show that the concentrations at the entrance boundary are
increasing with time until the concentration is equal to the specified concentration
of the source. This is consistent with the physical relevancy of the entrance
boundary condition when the concentration of the source rather than the
concentration of the boundary nodes are specified.

Page 351

Chapter 14: Product Integration

SIGMA/W

1.0
8.0000e+001

0.8
1.6000e+002

0.6

2.4000e+002

C
0.4

3.2000e+002

0.2

4.0000e+002

0.0

4.8000e+002

10

20

30

40

Distance

Figure 14-15 CTRAN/W solution of concentration versus distance at various


times using one free exit boundary

Page 352

SIGMA/W

Chapter 14: Product Integration

1.0
CTRAN/W Solution
(Exit Qd > 0)

0.9
0.8

Analytical Solution

Concentration

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0

10

20

30

40

Distance

Figure 14-16 Comparison with analytical solutions for one exit boundary
condition
1.0
8.0000e+001

0.8
1.6000e+002

0.6

2.4000e+002

C
0.4

3.2000e+002

0.2

4.0000e+002

0.0

4.8000e+002

10

20

30

40

Distance

Figure 14-17 CTRAN/W solution with two zero dispersive mass flux exit
boundary conditions

Page 353

Chapter 14: Product Integration

SIGMA/W

1.0
CTRAN/W Solution
(Exit Qd > 0)

0.9
0.8

CTRAN/W Solution
(Exit Qd = 0)

Concentration

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0

10

20

30

40

Distance

Figure 14-18 Comparison with analytical solution for two exit boundary
conditions

14.5

VADOSE/W velocity data in CTRAN/W contaminant


transport analysis

Seepage velocity and water content data from a steady state or transient
VADOSE/W analysis can be used in CTRAN/W to predict contaminant transport
with or without adsorption, diffusion and decay. The purpose of this example is to
show the influence of including climate (or vegetation) effects on the movement of
contaminants in soils.
In this example a simple advection dispersion analysis is carried out in
CTRAN/W after solving a transient VADOSE/W analysis. The CTRAN/W
program reads the transient seepage velocity and water content data at different
time steps and uses it in the solution of the advection dispersion equation.
The problem is kept quite simple for illustration purposes. In the seepage analysis,
a pressure equal to zero boundary condition is applied at the base of the low point
near the center of the mesh. This pressure boundary condition is saying there is a
small source of water at this location just enough to keep the ground saturated. It

Page 354

SIGMA/W

Chapter 14: Product Integration

is assumed that a source of contaminant is also present with the water at the ground
surface.
Figure 14-19 shows the position of the water table and the concentration contours
for the case where the model only allowed infiltration at the source of contaminant.
For this case, there is no evaporation demand along the rest of the ground surface
and therefore the contaminant is limited in its migration to a somewhat radial
pattern beneath the source.

Figure 14-19 Concentration contours without ground-climate coupling

In Figure 14-20, the same concentration is applied in the pond location along with
the same source of water, however, the climate boundary condition is active at all
other ground surface nodes. The climate boundary in this case is removing about
5 mm per day of evaporation.
It is clear for the case with evaporation, that there is a lot more spreading of the
contaminant. The results show that the contaminant is being pulled up towards the
drying ground surface where it is either exiting with the vapor flow or being
deposited or both depending on the exit boundary condition specified in the
contaminant analysis.

Page 355

Chapter 14: Product Integration

SIGMA/W

Figure 14-20 Concentration contours with ground-climate coupling

14.6

Density-dependent flow salt water intrusion

Henry (1964), developed an analytic solution for a simplified sea water intrusion
problem. The Henry problem has since become a benchmark verification
example for many numerical models of density-dependent flow. However,
Croucher and OSullivan, (1995), noted that none of the published numerical
model comparisons with Henrys solution that they examined were able to match
Henrys solution to a great extent. In addition to outlining the possible reasons for
the discrepancies, Croucher and OSullivan, presented a new, highly accurate
numerical solution to the problem. Their numerical solution is used here for
comparison with the results from CTRAN/W.
The system being modeled is shown in Figure 14-21. It consists of a 2.0m long
section of a 1.0m thick aquifer where the right boundary is in direct contact with
sea water and the left boundary has a constant influx of freshwater. The sea water
has a relative density, (specific gravity), of 1.025 at a reference concentration of
1.0. The concentration of sea water is fixed at 1.0 along the sea water boundary
and a fixed freshwater inflow rate of 6.6x10-5 m3/s is specified along the freshwater
boundary. The top and bottom boundaries are both impermeable. The aquifer is
homogeneous and isotropic and has a saturated hydraulic conductivity Ksat =
1x10-2 m/s, a porosity n = 0.35 and a velocity independent dispersion coefficient of
D = 1.89x10-5 m2/s. The aquifer is discretized using 0.05m square elements and the
solution is sought at steady state. The SEEP/W example file containing this
problem is called HENRY.GSZ, and the full definition of the problem may be
viewed using SEEP/W and CTRAN/W module DEFINE views.

Page 356

SIGMA/W

Chapter 14: Product Integration

Henry's Problem for Seawater Intrusion


Impermeable Top Boundary
1.0
0.9
0.8
0.7

Left Boundary
Freshwater
Qf = 6.6E-5 m2/s
C=0.0

Right Boundary
Seawater
Hs = 1.0m
C = 1.0

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

Impermeable Bottom Boundary

Aquifer Properties
Ks = 1E-2 m/s
n = 0.35
D = 1.89E-5 m2/s

Seawater Density
SG=1.025 @ Cref=1.0

Axes lengths in metres.


Element thickness = 1.0m

Figure 14-21 Henry's problem definition

The above Henrys problem has been analyzed with CTRAN/W, and a steady state
solution is obtained at an elapsed time of more than 11,000 seconds (at time step
35). Figure 14-22 shows the computed seawater concentration contours at steady
state along with the water flow velocity vectors. At steady state, seawater enters
the aquifer across the lower portion of the sea water boundary via density induced
gradients and mixes with freshwater flowing in the opposite direction. The
constant influx of freshwater from the left freshwater boundary causes the diluted
seawater to exit the system across the upper portion of the sea water boundary. In
this way, a seawater flow cell is established in which the seawater toe migration
towards the freshwater boundary is controlled by the rate of freshwater flow, the
density of the seawater, and the degree of mixing between the seawater and
freshwater. The degree of mixing is controlled by the dispersion coefficient used in
the modelling, which in this case is velocity independent.
It should be noted that in Figure 14-22 near the upper left of the aquifer,
CTRAN/W computed a few small negative concentration values. This slight
numerical oscillation is a direct result of the Peclet and Courant numbers being
exceeded in these areas because of the relatively high water velocity and relatively
coarse mesh and time step discretization. It is possible to eliminate the negative
concentration by reducing both the mesh size and the time step size, however,
since we are more interested in the solution in the lower portion of the flow system

Page 357

Chapter 14: Product Integration

SIGMA/W

and we only use the 0.5 concentration contours in the comparison, refinement to
the finite element mesh and time steps were deemed unnecessary in this case.
Freshwater
C=0

Seawater
C=1

1.00

0.75

0.50

0.
8

0.
6

0.
4

2
0.

0.25

0.00
0.00

0.25

0.50

0.75

1.00

1.25

1.50

1.75

2.00

Figure 14-22 Henry's problem verification - computed concentration


contours

Comparison of the CTRAN/W computed results with the highly accurate


numerical solution of Croucher and OSullivan is given in Figure 14-23. The
figure compares the 0.5 seawater concentration isochlors at steady state. It can be
seen that the results from CTRAN/W are almost identical to those of Croucher and
OSullivan.

Page 358

SIGMA/W

Chapter 14: Product Integration

1
0.9

Croucher and O'Sullivan, 1995

0.8

CTRAN/W

0.7

Elevation (m)

0.6
0.5
0.4
0.3
0.2
0.1
0
1

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

X-Coordinate (m)

Figure 14-23 Comparison of 0.5 concentration isochlors

14.7

Ground freezing and water flows (SEEP/W and TEMP/W)

The GeoStudio module TEMP/W can be used to model natural and artificial
ground freezing without or with the addition of heat added by flowing water.
Consider the frozen ground that forms around a buried pipe with a pipe wall
boundary condition of -2 degrees Celsius. Figure 14-24 shows the ground
temperature profile after 2 years of freezing. It is clear from this figure that a
region of frozen ground has formed around the pipe and that the ground freezes to
a deeper depth beneath the pipe where there is less influence from the warmer
yearly average ground surface temperature. Details of this analysis and comparison
of the results for this case to other published data can be found in the TEMP/W
engineering book. The SEEP/W example file containing this problem is called
PIPE WITH FLOWING WATER.GSZ, and the full definition of the problem
may be viewed using SEEP/W and TEMP/W module DEFINE views.

Page 359

Chapter 14: Product Integration

SIGMA/W

1.6
2.5

1.4

Pipe

1.2
1.0
0.8
0.6
0.4
0.2
0.0

-0.2
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Figure 14-24 Temperature contours around freezing pipe without flowing


water

Figure 14-25 shows the temperature profile for same analysis except in this case
the influence of flowing water is considered. The SEEP/W module was set up such
that a hydraulic gradient of 0.27% was established across the region from right to
left with a saturated hydraulic conductivity of 0.1 m/day. It is clear from this figure
that the extents of the frozen ground are far less than for the static water condition
and that the shape of the frozen ground is more strongly influenced by the flowing
water than the warmer air temperature.
Figure 14-26 shows the actual computed water velocity vectors and total head
contours across the region after two years of analysis. This figure illustrates clearly
the diversion of water around the frozen ground region. Careful examination of the
figure shows that the velocity of water increases as it moves around the frozen
ground. This is because the cross sectional area available for flow is reduced due to
ice formation. For cases where several freezing pipes are installed to create a
frozen barrier wall, the increase in velocity of water between adjacent freeze pipes
can result in a situation where more heat is added by the water than can be
removed by the freeze pipes. When this occurs, closure of the frozen wall (due to
adjacent frozen regions growing together) will not happen. This effect can be seen
in Figure 14-27 for a typical shaft sinking artificial ground freezing project.

Page 360

SIGMA/W

Chapter 14: Product Integration

1.6
2.5

1.4
1.5

1.2

Pipe

1.0
0.8
0.6
0.4
0.2
0.0
-0.2
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Figure 14-25 Temperature contours around freezing pipe with flowing water

When carrying out this type of analysis, it is very important to have well-defined
material properties and to use a fine mesh discretization and small time steps. The
seepage velocity generated by SEEP/W and added into the TEMP/W finite element
partial differential equation is a linear term, and as such, the computed results are
quite sensitive to numerically appropriate element size and time steps. A first guess
at time step sizes can be made by computing the Courant number and ensuring that
it is less than a value of 1. The Courant number is given by:

C=

vt
x

where:

the Darcian velocity,

the time step, and

the element edge length.

Page 361

Chapter 14: Product Integration

0.4

0.6

1.6045

0.2

1.6035

1.6025

5
1.601

1.6005

1.6
1.5
1.4
1.3
1.2
1.1 Pipe
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
-0.1
-0.2
-0.2 0.0

SIGMA/W

0.8

1.0

1.2

1.4

1.6

Figure 14-26 Water flow vectors and head contours around frozen pipe

Figure 14-27 Non closing "window" in shaft ground freezing

It is suggested when modeling convective heat transfer to use the adaptive time
stepping routine and set the maximum time step by first setting the Courant
number to 1 in the above equation and computing the permissible time step to
ensure numerical stability.

Page 362

SIGMA/W

Chapter 14: Product Integration

Finally, while this example shows how water can influence freezing in ground, the
same combination of TEMP/W and SEEP/W can be used to study the influence of
freezing ground on the transient seepage of water when the parts of the ground are
subjected to cold temperatures. This might be the case for dam performance in
climates with both summer and winter seasons.
In both cases, you must set up a complete SEEP/W and TEMP/W analysis and
then start the solver process from the TEMP/W program. TEMP/W will launch
SEEP/W, pass it the ground temperatures and wait for SEEP/W to pass back water
content and computed seepage velocities. In the event that SEEP/W is passed a
ground temperature that is below the freezing point of water, it will compute a
reduced hydraulic conductivity corresponding to the magnitude of the temperature
below the phase change point. The calculation of the frozen ground conductivity is
discussed in the Material Properties chapter of this book.

14.8

Seepage-dependent embankment settlement (SEEP/W


and SIGMA/W)

The construction of an embankment on slightly over-consolidated clay is a


relatively complex modeling problem that requires a modeling approach
combining stress and seepage analyses.
In the following example, the subsurface clay stratum is 9 m deep and the water
table is 1 m below the ground surface. The 1m layer above the water table is highly
weathered, desiccated and fissured, making its behavior similar to a fine granular
soil. The embankment to be analyzed is shown in Figure 14-28 and is constructed
of relatively sandy soil. The height of the embankment is 5 m (El. 14) with 3:1 side
slopes and a 10 m crest width. Due to symmetry about the center, only half of the
cross section is used in the analysis.
The objective of this example is to determine the embankment settlement. Full
details of this example problem are given in the technical paper a351 available
on the GEO-SLOPE International Ltd. website.

Page 363

Chapter 14: Product Integration

SIGMA/W

15
14

13

12

Elevation - metres

11
10
9
8
7
6
5
4
3
2
1
0
0

10

15

20

25

30

35

40

metres

Figure 14-28 Embankment problem configuration

The clay will be modeled as a Modified Cam-clay (MCC) soil. It is highly plastic
with a liquid limit (LL) of 61%. A summary of all clay properties is given in Table
14-3. The embankment and upper meter of subsoil will be modeled as highly
permeable linear-elastic material, which is considered adequate since the major
settlement issue arises in the underlying compressible clay. The properties adopted
for the sand are:
E

2000 kPa

0.36

1.0 m/day

mv

3 x 10-4 (1/kPa)

The k and mv values are actually arbitrary for the sand, since steady-state porewater pressure conditions will be specified, meaning there is no change with time.
In a transient (consolidation) analysis, the software requires these values be
specified, but the actual values do not affect the results.

Page 364

SIGMA/W

Chapter 14: Product Integration

Table 14-3 Summary of clay material properties


Property

Value

Cc (compression index)

0.46

(slope of normal consolidation line)

0.20

(slope of swelling line)

0.04

mv (coefficient of volume compressibility)

3 x 10-4 (1/kPa)

k (hydraulic conductivity)

5 x 10-3 m/day

(effective friction angle)

26 degrees

(slope of critical state line)

1.0

Ko (coefficient of earth pressure at rest)

0.56

(Poissons Ratio)

0.36

OCR (over consolidation ratio)

1.2

(specific volume)

2.2

The analysis will be run for 40 days. Figure 14-29 shows the embankment with the
numbers in each row of elements indicating at what time (day) the elements
become active. The clay will consolidate during the time between each lift
placement and for 15 days after the last lift is placed.
+25 +25 +25 +25 +25 +25
+25 +25
+19 +19 +19 +19 +19 +19 +19 +19 +19
+19 +19
+13 +13 +13 +13 +13 +13 +13 +13 +13 +13 +13 +13
+13 +13
+7

+7

+7

+7

+7

+7

+7

+7

+7

+7

+7

+7

+7

+7

+7

+7 +7

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1

+1 +1

Figure 14-29 Fill placement sequence specified in SIGMA/W

Important parameters and conditions to specify in SEEP/W include the coefficient


of volume compressibility mv , the hydraulic conductivity and the hydraulic
boundary conditions.
In SEEP/W the slope of the volumetric water content function is mv . In this
example, consolidation of the clay will occur under saturated conditions so the

Page 365

Chapter 14: Product Integration

SIGMA/W

volumetric water content function can be defined as a straight line with two points
resulting in a line with a slope of 3 x 10-4 (1/kPa). The slope is (90 x 0.001)/300 =
0.0003.
The hydraulic conductivity of the clay will be constant and equal to the saturated
hydraulic conductivity, defined by a horizontal function at 5 x 10-3 m/day.
The pore-water pressure along the water table is zero and the total head is therefore
equal to the elevation. As a result, the boundary condition is defined as H (P=0).
Changes in pore-water pressure are not considered within the upper meter of clay
or the embankment fill. As a result, deformations in these zones due to changes in
pore-water pressure are not computed. It is important to note that pore-water
pressure changes within these zones are not allowed. Setting the boundary
conditions in these zones to be H (P=0) as illustrated in Figure 14-30, ensures this
to be the case. While the actual pore-water pressures may not be correct, the effect
has been removed, which is an objective of this analysis.
All the other boundary conditions can be left undefined, resulting in a zero flow
boundary at the vertical ends of the clay located below the water table and along
the bottom.
Figure 14-31 shows the pore-water pressure changes with time at a specified
location one meter to the right of the centerline and two meters below the water
table. The pore-water pressure immediately rises after the construction of a lift and
then falls until the next lift is placed. After the final lift has been placed, the porewater pressure at the specified location drops from a high of 90 kPa to about 55
kPa. Significant excess pore-water remains at the end of 40 days, indicating that
additional long-term settlement will occur.

Page 366

SIGMA/W

Chapter 14: Product Integration

Figure 14-30 SEEP/W boundary conditions

Figure 14-32 shows the settlement profiles along the original ground surface on
days 0, 7, 13, 19, and 25. Of particular interest is the location of maximum
settlement, which does not occur along the centerline, but is located more toward
the outer edge of the fill in the early stages of loading. This result is due to the
zone of lower effective horizontal stresses as shown in Figure 14-33. In addition,
the toe of the fill has been lifted up slightly.

Pore-Water Pressure vs. Time

Pore-Water Pressure

110

90

70

50

30

10
0

10

20

30

40

Time

Figure 14-31 Pore-water pressure changes beneath embankment versus time

Page 367

Chapter 14: Product Integration

SIGMA/W

Y-Displacement vs. X
0.2
0.0000e+000

Y-Displacement

0.0
7.0000e+000

-0.2
1.3000e+001

-0.4
1.9000e+001

-0.6

2.5000e+001

-0.8
0

10

15

20

25

30

Figure 14-32 Ground surface settlement profiles at various times

25

30

25
30

Figure 14-33 Horizontal effective stress after second lift placed

Page 368

SIGMA/W

Chapter 15: Theory

15

Theory

15.1

Introduction

This chapter presents the methods, equations, procedures, and techniques used in
the formulation and development of the SIGMA/W SOLVE function. It is of value
to be familiar with this information when using the software. An understanding of
these concepts will be of great benefit in applying the software, resolving
difficulties, and judging the acceptability of the results.
The development of the finite element equations for stress/deformation analysis
using potential energy, weighted residuals, or variational methods is well
documented in standard textbooks, and consequently is not duplicated in this
Users Guide. (See Bathe, 1982, Smith and Griffiths, 1988, Segerlind, 1984 and
Zienkiewicz and Taylor, 1989 for further information on the development of finite
element equations).
SIGMA/W is formulated for either two-dimensional plane strain or axisymmetric
problems using small displacement, small strain theory. Conforming to
conventional geotechnical engineering practice, the "compression is positive" sign
convention is used. In this chapter, the bracket sets < >, { }, and [ ] are used to
denote a row vector, a column vector and a matrix, respectively.

15.2

Finite element equations

The finite element equation used in the SIGMA/W formulation for a given time
increment is:

[ B ] [C ][ B ] dv {a}
T

= b < N >T dv + p < N >T dA + { Fn }


v

where:

[ B]

strain-displacement matrix,

[C ]

constitutive matrix,

{a}

column vector of nodal incremental x- and y-displacements,

Page 369

Chapter 15: Theory

SIGMA/W

< N> =

row vector of interpolating functions,

area along the boundary of an element,

volume of an element,

unit body force intensity,

incremental surface pressure, and

{Fn }

concentrated nodal incremental loads.

Summation of this equation over all elements is implied. It should be noted the
SIGMA/W is formulated for incremental analysis. For each time step, incremental
displacements are calculated for the incremental applied load. These incremental
values are then added to the values from the previous time step. The accumulated
values are reported in the output files. Using this incremental approach, the unit
body force is only applied when an element is included for the first time during an
analysis.
For a two-dimensional plane strain analysis, SIGMA/W considers all elements to
be of unit thickness. For constant element thickness, t, the above equation can be
written as:
Equation 15-1

t [ B ] [C ][ B ] dA {a} = bt < N >T dA + pt < N >T dL


T

However, in an axisymmetric analysis, the equivalent element thickness is the


circumferential distance about the axis of symmetry. Although the complete
circumferential distance is 2 radians times the radial distance, R, SIGMA/W is
formulated for one radian (unity). Consequently, the equivalent thickness is R and
the finite element equation for the axisymmetric case becomes:

([ B ] [C ][ B ] R ) dA {a}
T

= b ( R < N >T ) dA + p ( R < N >T ) dL + { Fn }


A

Unlike the thickness, t, in a two-dimensional analysis, this radial distance, R, is not


a constant within an element. Consequently, R needs to be evaluated inside the
integral.
In an abbreviated form, the finite element equation is:
Equation 15-2

[ K ]{a} = {F } = {Fb } + {Fs } + {Fn }


Page 370

SIGMA/W

Chapter 15: Theory

where:
[K]

element characteristic (or stiffness) matrix,

([ B] [C ][ B]) dA
= ([ B ] [C ][ B ] R ) dA
= t

(for plane strain), or

(for axisymmetric)

{a}

nodal incremental displacements,

{F}

applied nodal incremental force which is made up of the following:

{Fb}

incremental body forces,

{Fs}

force due to surface boundary incremental pressures,

= pt ( < N >T ) dL, for two-dimensional analysis, or


L

= p ( R < N >T ) dL for axisymmetric analysis , and


L

{Fn}

concentrated nodal incremental forces.

SIGMA/W solves this finite element equation for each time step to obtain
incremental displacements and calculates the resultant incremental stresses and
strains. It then sums all these increments since the first time step and reports the
summed values in the output files.
Strain-displacement matrix

SIGMA/W uses engineering shear strain in defining the strain vector:

x

{ } = y
z
xy
The field variable of a stress/deformation problem is displacement which is related
to the strain vector through:

Page 371

Chapter 15: Theory

Equation 15-3

SIGMA/W

{ } = [ B ]

where:
[B]

strain matrix,

u, v

nodal displacement in x- and y-directions, respectively.

SIGMA/W is restricted to performing infinitesimal strain analyses. For a twodimensional plane strain problem, Z is zero and the strain matrix is defined as:

N1
x

0
[ B ] =
0

N1
x

...

N8
x

N1
...
y
0

N1
y

...

0
N8
...
x

N8
y
0

N8
y

For an axisymmetric problem, the strain matrix can be written as:

u
x

x v

y y

=
{ } =

z u

xy

u + v
y x
The associated strain matrix [ B ] is then:

Page 372

SIGMA/W

N1
x

0
[ B] =
N1
R

N1
x

Chapter 15: Theory

...

N1
...
y
0

N1
y

N8
x
0

N8
R
N8
...
x
...

N8
y

N8
y

The derivatives of the interpolating functions are given in the Derivatives of


Interpolating Functions section in this chapter.
Elastic constitutive relationship

Stresses are related to strains as follows, within the theory of elasticity:

{ } = [C ]{ }
where [C ] is the constitutive (element property) matrix and is given by:

0
1

1
0

E
[C ] =

1 v
0

(1 + )(1 2 )

1 2
0
0
0

2
where:
E

Youngs modulus

Poisson's ratio

The [C ] matrix is the same for both the two-dimensional plane strain and the
axisymmetric cases.

Page 373

Chapter 15: Theory

SIGMA/W

Body forces

SIGMA/W can model body forces applied in both the vertical and the horizontal
directions. These forces are applied to all elements when they first become active.
The body force in the vertical direction, bv, is due to gravity acting on an element.
For a given material, the unit body force intensity in the vertical direction is given
by its unit weight, s, which is in turn related to its mass density, :

s = g
where g is the gravitational constant. When the unit weight s is non-zero,
SIGMA/W evaluates the integral s

( < N > ) dv by numerical integration and


T

applies a vertically downward (negative) force at each node of the element.


Similarly, when the unit body force intensity in the horizontal direction, bh , is
nonzero, nodal forces in the horizontal direction are computed using
bh < N >T dv .

(
v

Forces due to boundary stresses

The term p

( < N > ) dA in Equation 15-2 represents the nodal forces caused


T

by externally applied pressure along the boundary of the element. SIGMA/W


evaluates this integral using numerical integration in the SOLVE function.
Three loading types are available in SIGMA/W; namely, normal and tangential
pressure; x- and y-stress; and fluid pressure. The procedure used to derive the
equivalent nodal forces for each of the pressure boundary type is described below.
Consider the element edge subjected to normal and tangential pressure loading as
shown in Figure 15-1. In order to calculate the equivalent nodal forces acting along
that edge, the elemental loads must be resolved into x- and y- components. The
normal and tangential pressure, pn and pt, acting on an elemental length, dS of the
loaded edge, result in elemental forces, dPx and dPy, in the x- and y-direction.
These forces can be written as follows:

dPx = ( pt cos pn sin )dS .t = ( pt dx pn dy )t


dPy = ( pt sin + pn cos )dS .t = ( pn dx + pt dy )t

Page 374

SIGMA/W

Chapter 15: Theory

where t is the thickness of the element.


The total force can be obtained by integrating along the element edge (i.e., along
the local coordinate r). The differentials dx and dy can be expressed in terms of r:

x
dr
r
y
dy =
dr.
r
dx =

Substituting these differentials into the incremental force equation and applying the
finite element approximation, the following equations can be derived for the
equivalent nodal forces at node i along an element edge:

x
y
pn
)dr
r
r
S
x
y
+ pt
)dr.
Pyi = N i t ( pn
r
r
S

Pxi = N i t ( pt

The integration is carried out numerically in SIGMA/W SOLVE using Gaussian


quadrature. The number of integration points used corresponds to the number of
nodes along this element edge.
Resolution of elemental forces will not be necessary when x- and y-stresses being
applied along an edge of the element. The equations for the equivalent nodal forces
become:

x
)dr

r
S
y
)dr.
Pyi = N i t ( p y

r
S
Pxi = N i t ( px

The fluid pressure boundary is as a particular case of normal and tangential case in
which no tangential pressure is applied. During numerical integration, fluid
pressure is evaluated at each integration point, say the i-th integration point, using:

pni = f ( y f yi ),

( y f yi ) > 0

Page 375

Chapter 15: Theory

SIGMA/W

where:

the self-weight of the fluid,

yf

the elevation of the fluid, and,

yi

the y-coordinate at the integration point.

The fluid pressure is only computed when the fluid elevation exceeds the ycoordinate of an integration point.

Figure 15-1 Normal and tangential pressure along an element edge

Nodal forces

In SIGMA/W, nodal forces are included in the finite element formulation,


Equation 15-2, under two conditions. The nodal force can either be specified as
boundary conditions or SIGMA/W can calculate them internally when elements
are first excavated.
For each element, nodal forces are computed using:

Page 376

SIGMA/W

Chapter 15: Theory

{F } = v [ B ] { } dv
T

where:
{}

the vector of element stresses,

[B]

the transpose of the strain-displacement matrix, and

dv

elemental volume.

The resultant nodal forces are accumulated at each node. To simulate the removal
of soils, as in an excavation, the signs on the nodal forces are reversed before these
forces are incorporated into the finite element equation.
In the displacement output files, SIGMA/W reports the nodal forces at all nodes
where displacement or spring boundary conditions are specified. These nodal
forces are calculated in the manner previously described. However, there is now no
change of signs. The nodal forces represent boundary forces which may be used,
for example, to estimate the loads supported by struts in a braced excavation.

15.3

Numerical integration

SIGMA/W uses Gauss-Legendre numerical integration (also termed quadrature) to


form the element characteristic (or stiffness) matrix [K]. The variables are first
evaluated at specific points within an element. These points are called integration
points or Gauss points. These values are then summed for all the Gauss points
within an element. This mathematical procedure is as described in the following.
To carry out numerical integration, SIGMA/W replaces the following integral from
Equation 15-1:

[ B ] [C ][ B ] dA
T

with the following equation:


n

B
j =1

C j
B j det J j W1 jW1 j

where:

Page 377

Chapter 15: Theory

SIGMA/W

integration point,

total number of integration points or integration order,

det J j

determinant of the Jacobian matrix, and

W1 j , W2 j

weighting factors

Table 15-1 to Table 15-4 show the numbering scheme and location of integration
points used in SIGMA/W for various element types.
Table 15-1 Sample point locations and weightings for four point quadrilateral
element
Point

+0.57735

-0.57735

-0.57735

+0.57735

w1

w2

+0.57735

1.0

1.0

+0.57735

1.0

1.0

-0.57735

1.0

1.0

-0.57735

1.0

1.0

Page 378

SIGMA/W

Chapter 15: Theory

Table 15-2 Sample point locations and weightings for nine point quadrilateral
element
Point

w1

w2

+0.77459

+0.77459

5/9

5/9

-0.77459

+0.77459

5/9

5/9

-0.77459

-0.77459

5/9

5/9

+0.77459

-0.77459

5/9

5/9

0.00000

+0.77459

8/9

5/9

-0.77459

0.00000

5/9

8/9

0.00000

-0.77459

8/9

5/9

+0.77459

0.00000

5/9

8/9

0.00000

0.00000

8/9

8/9

Table 15-3 Sample point location and weighting for one point triangular
element
Point

w1

w2

0.33333

0.33333

1.0

0.5

Table 15-4 Sample point locations and weightings for three point triangular
element
Point

w1

w2

0.16666

0.16666

1/3

1/2

0.66666

0.16666

1/3

1/2

0.16666

0.66666

1/3

1/2

The appropriate integration order is dependent on the number of secondary nodes.


When secondary nodes are present, the interpolating functions are non-linear and
consequently a higher integration order is required. Table 15-5 gives the acceptable
integration order for various element types.

Page 379

Chapter 15: Theory

SIGMA/W

Table 15-5 Acceptable element integration orders


Element Type

Secondary Nodes

Integration Order

Quadrilateral

no

Quadrilateral

yes

Triangular

no

Triangular

yes

Under certain conditions, it is acceptable to use four-point integration for


quadrilateral elements which have secondary nodes. This procedure is called
reduced integration and is described in Bathe (1982), and Zienkiewicz and Taylor
(1989). For example, reduced integration may yield more accurate results in nearly
incompressible elasticity and in elements under flexure. In addition, selective use
of reduced integration can greatly reduce the required number of computations.
It is also possible to use higher order (three-point and nine-point) integration with
elements that have no secondary nodes. However, in this case, the benefits of using
higher order integration are marginal, particularly for quadrilateral elements. Nine
point integration for quadrilateral elements involves substantially more computing
than four point integration, and there is little to be gained from the additional
computations. As a general rule, quadrilateral elements should have secondary
nodes to achieve significant benefits from the nine point integration.
The situation is slightly different for triangular elements. Using one-point
integration implies that the material properties and strains are constant within the
element. This can lead to poor performance of the element, particularly if the
element is in a region of large stress gradients. Using three point integration, even
without using secondary nodes, can improve the performance, since material
properties and gradients within the elements are distributed in a more realistic
manner. The use of three point integration in triangular elements with no
secondary nodes is considered acceptable in a mesh that has predominantly
quadrilateral elements. This approach is not recommended if the mesh consists
primarily of triangular elements with no secondary nodes.

15.4

Assembly and solving of global equations

The element matrix for each element in the discretized finite element mesh can be
formed and assembled into a global system of simultaneous equations. The finite
element solution requires the solving of the system of simultaneous equations.

Page 380

SIGMA/W

Chapter 15: Theory

SIGMA/W stores the coefficients of the global system equations using a


Compressed Row Storage scheme (Barrett et. al, 1994). This is a general scheme
that make no assumptions about the sparsity structure of the matrix. Instead of a
full matrix with many zero elements, the Compressed Row Storage scheme stores
the matrix with 3 vectors: one for the non-zero elements, one for the column index
and one for the row pointers. As a result, it provides significant saving in storage
memory particularly in very large finite element meshes.
SIGMA/W utilizes a preconditioned Bi-Conjugate Gradient (BiCG) iterative solver
in solving the system equations. The BiCG solver is adopted from IML++
(Iterative Methods Library) made available freely by the National Institute of
Standards and Technology, Oak Ridge National Laboratory, University of
Tennessee, Knoxville, U. S. A. The BiCG solver works effectively with the
compressed row storage scheme and is suitable for both symmetric and nonsymmetric system of equations.
The equation solver can accommodate missing elements in the array. This feature
makes it possible to add and delete elements from a finite element mesh without
renumbering the nodes and elements. Unlike the Gauss elimination solver used in
previous versions of SIGMA/W, sorting the node number vertically or horizontally
will have no effect to the computational time and accuracy of the solution with this
improved iterative solver.
As a result of the iterative solver, the solution of the system of simultaneous
equations is to a large degree dependent on the convergence tolerance and the
number of iterations. In most cases, the default tolerance and number of iterations
will be adequate for a solution.
The Gauss Elimination Skyline Direct Solver is also included in the program. The
computational speed of a direct solver is depending on the bandwidth of the finite
element mesh. The direct solver is fast for simple problems, but it can be slower
than the iterative solver in complex problems with large bandwidths. You may like
to try both the direct and iterative solvers for a few iterations first before solving a
large problem with many iterations and time steps.

15.5

Element stresses

SIGMA/W computes the stresses and strains at each integration point within each
element once the nodal displacements have been obtained. Strains are computed
from nodal displacements using Equation 15-3.

Page 381

Chapter 15: Theory

SIGMA/W

Stresses are computed at each Gauss point using the constitutive matrix [C] in the
following manner:

x
x


y
y
= [C ]
z
z
xy
xy

15.6

Pore pressure parameters A and B

Pore-water pressures generated due to a change in total stress can be estimated by


the following equation (Henkel, 1960):

Equation 15-4

+ 2 + 3
u = 1
+
3

( 1 2 ) + ( 2 3 ) + ( 3 1 )
2

where and are empirical pore pressure parameters.


The values of and are related to Skemptons A and B pore pressure
parameters, which can be determined from triaxial tests. In a triaxial test, the pore
pressure generated due to a change in the principal stresses is given by:
Equation 15-5

u = B 3 + a ( 1 3 )
= ( B AB ) 3 + AB 1

After substituting the triaxial stress condition, 2 = 3, into Equation 15-4, it can be
written as:

Equation 15-6

u =
2 3 + + 2 1
3

Comparing the coefficients of 1 and 3 in Equation 15-5 and Equation 15-6,


the following equation can be derived:

Page 382

SIGMA/W

Chapter 15: Theory

2
3

Equation 15-7

( B AB ) =

Equation 15-8

AB = + 2
3

Solving Equation 15-7 and Equation 15-8 simultaneously leads to the following
expressions:

=B
=

B
1
A
3
2

Pore pressure parameters A and B are commonly measured using triaxial tests in
the laboratory. Consequently, SIGMA/W uses A and B as input data in an analysis
and internally calculates and .
Equation 15-4 is presented in terms of principal stresses. For computational
purpose, it is more convenient to express the same equation in terms of general
stress components, x, y, y, and xy. When the pore-water pressure equation is
written in terms of these stress components, it can be expressed as follows.

u = u + u

Equation 15-9

where:

u =

u =

+ y + z )

y ) + ( y z ) + ( z x ) + 6 xy2
2

The pore-water pressure generated due to a stress change consists of two


components; namely, u , which is a function of changes in isotropic stress, and

u , which is due to changes in deviatoric stress.


A discussion pertaining to the use of pore-water pressure parameters is included in
Chapter 3.

Page 383

Chapter 15: Theory

SIGMA/W

Page 384

SIGMA/W

Appendix A: Interpolating Functions

16

Appendix A: Interpolating Functions

16.1

Coordinate systems

The global coordinate system used in the formulation of SIGMA/W is the first
quadrant of a conventional x y Cartesian system.
The local coordinate system used in the formulation of element matrices is
presented in Figure 16-1. Presented as well in Figure 16-1 is the local element
node numbering system. The local coordinates for each of the nodes are given in
Table 16-1.
Table 16-1 Local element node numbering system
Element Type
Quadrilateral

Triangular

Node

+1

+1

-1

+1

-1

-1

+1

-1

+1

-1

-1

SIGMA/W uses the fourth node to distinguish between triangular and quadrilateral
elements. If the fourth node number is zero, the element is triangular. If the fourth
node number is not zero, the element is quadrilateral.

Page 385

Appendix A: Interpolating Functions

SIGMA/W

In the case of quadrilateral elements, Nodes 5, 6, 7, and 8 are secondary nodes. In


the case of triangular elements, Nodes 5, 6, and 7 are secondary nodes.
The local and global coordinate systems are related by a set of interpolation
functions. SIGMA/W uses the same functions for relating the coordinate systems
as for describing the variation of the field variable (displacement) within the
element. The elements are consequently isoperimetric elements.
y

1 (1, 1)

s
5
2 (1, -1)

4 (-1, 1)
7
Local coordinates (r, s)
Global coordinates (x, y)

3 (-1, 1)

x
A) Quadrilateral Element
y

s
3 (0, 1)

6
2 (1, 0)

r
5
Local coordinates (r, s)
Global coordinates (x, y)

1 (0, 0)

x
B) Triangular Element

Figure 16-1 Global and local coordinate system

To formulate a finite element analysis it is necessary to adopt a model for the


distribution of the field variable within each finite element. The field variable in a
stress/deformation analysis is nodal displacement.

Page 386

SIGMA/W

Appendix A: Interpolating Functions

SIGMA/W assumes that the displacement distribution within the element follows
the interpolating functions described previously in this chapter. This means that the
displacement distribution is linear when secondary nodes are not used, and the
displacement distribution is quadratic when the secondary nodes are used.
The displacement distribution model at any given location inside a finite element is
given by the following set of equations:
Equation 16-1 u = N

{U }

Equation 16-2 v = N

{V }

where:
u

x-displacement at the given location

y-displacement at the given location

{U }

x-displacement at the nodes of the element

{V }

y-displacement at the nodes of the element

interpolation functions evaluated at the given point

16.2

Derivatives of interpolating functions

The fundamental constitutive relationship used in the formulation of SIGMA/W


relates stress, , to strain, , using the stiffness , E , of the material. In equation
form:

= E
In a two-dimensional plane strain problem, there are three basic strain components:
longitudinal strain in the x-direction, x , longitudinal strain in the y-direction, y ,
and shear strain in the x-y plane, xy . SIGMA/W is formulated for small
displacement, small strain problems. The strain components are related to x- and y
displacements, u and v, as follows:

Page 387

Appendix A: Interpolating Functions

SIGMA/W

u
x
v
y =
y
u v
xy =
+
y x
x =

At any point within a finite element, displacements u and v are related to the nodal
displacement vectors {U } and {V } by Equation 16-1 and Equation 16-2. Strains,
when expressed in terms of nodal displacements, can be written as follows:

x =

u
N
=
x
x

{U }

y =

v
N
=
y
y

{V }

xy =

u v
N
+
=
y x
y

{U } +

N
x

{V }

The above equation shows that, in order to calculate strains, it is necessary to


differentiate the interpolating functions with respect to x and y. The derivatives of
the interpolating functions in the local and global coordinate systems are given by
the chain rule:

Equation 16-3

r
N
s

x
r
=
x
s

y
s

N
s

in which:

Page 388

SIGMA/W

Appendix A: Interpolating Functions

x
r
Equation 16-4
x
s

y
r = J , the Jacobian matrix.
[ ]
y
s

Thus, the derivative of the interpolation functions, with respect to x and y can be
determined by inverting Equation 16-3:

1
= [J ]
N

where [ J ]

N
s

is the inverse of the Jacobian.

The Jacobian matrix can be obtained by substituting Equation 16-1 and Equation
16-2 into Equation 16-4.

N1
r

[J ] =

N1
s

N8
N2
K
X Y
r
r 1 1
X Y
N2
s

N8
K
s

2
M

X8

Y8
2

As can be seen from Equation 16-3 and Equation 16-4, derivatives of the
interpolating functions are also required for calculating strains and the Jacobian.
The derivatives of the interpolation functions with respect to r and s used by
SIGMA/W for quadrilateral and triangular elements are given in Table 16-2
and

Page 389

Appendix A: Interpolating Functions

SIGMA/W

Table 16-3, respectively.


Derivatives of the mapping functions with respect to r and s for an one-directional
and two-directional infinite elements are given in Table 16-4.
The following notation has been used in the preceding tables to represent the
derivative of a given function Ni with respect to variable r:

N i ,r =

Ni
r

Table 16-2 Interpolation function derivatives for quadrilateral elements


Derivative of Function

Include in derivative if node is present


5

N1,r

(1+s)

-(Nv,r)

-(N8,r)

N2,r

-(1+s)

-(N5,r)

-(N6,r)

N3,r

-(1-s)

-(N6,r)

-(N7,r)

N4,r

(1-s)

-(N7,r)

-(N8,r)

N5,r

-(2r+2sr)

N6,r

-(1-s2)

N7,r

-(2r-2sr)

N8,r

(1-s2)

N1,s

(1+r)

- (N5,s)

-(N8,s)

N2,s

(1-r)

-(N5,s)

-(N6,s)

N3,s

-(1-r)

-(N6,s)

-(N7,s)

N4,s

-(1+r)

-(N7,s)

-(N8,s)

N5,s

(1-r2)

N6,s

-(2s-2sr)

N7,s

-(1-r2)

N8,s

-(2s+2sr)

Page 390

SIGMA/W

Appendix A: Interpolating Functions

Table 16-3 Interpolation function derivatives for triangular elements


Derivative of Function

Include in derivative if node is


present
5

N1,r

-1. 0

-(N5,r)

N2,r

1. 0

-(N5,r)

-(N6,r)

N3,r

0. 0

-(N6,r)

-(N7,r)

N5,r

(4-8r-4s)

N6,r

4s

N7,r

-4s

N1,s

-1. 0

-(N5,s)

N2,s

0. 0

-(N5,s)

-(N6,s)

N3,s

1. 0

-(N6,s)

-(N7,s)

N5,s

-4r

N6,s

4r

N7,s

(4-4r-8s)

Page 391

Appendix A: Interpolating Functions

SIGMA/W

Table 16-4 Mapping function derivatives for infinite elements


Mapping Function Derivatives
1-D Infinite Elements

2-D Infinite Elements

M1,r

M1,r

M2,r

(-2-s+s2)/(1-r)2

M2,r

M3,r

(-2+s+s2)/(1-r)2

M3,r

-4(2+s)/(1-r)2(1-s)

M4,r

M4,r

M5,r

(1+s)/(1-r)2

M5,r

M6,r

2(1-s2)/(1-r)

M6,r

2(1+s)/(1-r)2(1-s)

M7,r

(1-s)/ (1-r)2

M7,r

4/(1-r)2(1-s)

M8,r

M8,r

M1,s

M1,s

M2,s

(-r+2s)/(1-r)

M2,s

M3,s

(r+2s)/(1-r)

M3,s

-4(2+r)/(1-r) (1-s)2

M4,s

M4,s

M5,s

(1+r)/(1-r)

M5,s

M6,s

-4s/(1-r)

M6,s

4/(1-r) (1-s)2

M7,s

-(1+r)/(1-r)

M7,s

2(1+r)/(1-r)(1-s)2

M8,s

M8,s

Page 392

SIGMA/W

References

References
Atkinson, J.H. and Bransby, P.L., 1978. The Mechanics of Soils: An Introduction
to Critical State Soil Mechanics. McGraw-Hill.
Bathe, K-J., 1982. Finite Element Procedures in Engineering Analysis. PrenticeHall.
Bettess, Peter., 1992. Infinite Elements. Penshaw Press.
Biot, M.A. , 1941. General theory of three-dimensional consolidation. Journal of
Applied Physics, Volume 12, February, pp. 155-164.
Britto, A.M. and Gunn, M.J., 1987. Critical State Soil Mechanics via Finite
Elements. John Wiley & Sons, Inc.
Chan, D. H.-K., 1986. Finite Element Analysis of Strain Softening Material. Ph.D.
Thesis, Civil Engineering Department, University of Alberta, Edmonton,
Alberta, Canada.
Chen, W.F. and Zhang, H., 1991. Structural Plasticity: Theory, Problems, and
CAE Software. Springer-Verlag.
Dakshanamurthy, V., Fredlund, D.G, and Rahardjo, H., 1984. Coupled threedimensional consolidation theory of unsaturated porous media.
Proceedings of the Fifth International Conference on Expansive Soils,
Adelaide, Australia, pp. 99 -103.
Duncan, J.M. and Chang, C.Y., 1970. Nonlinear analysis of stress and strain in
soils. Journal of the Soil Mechanics and Foundations Division, ASCE, vol.
96, no. SM5, pp. 1629-1654.
Duncan, J.M., Byrne, P., Wong, K.S., and Mabry, P., 1980. Strength, Stress-Strain
and Bulk Modulus Parameters for Finite Element Analyses of Stresses and
Movements in Soil Masses. Report No. UBC/GT/80-01. Department of
Civil Engineering, University of California, Berkeley, California.

Page 393

References

SIGMA/W

Fredlund, D.G. and Morgenstern, N.R., 1976. Constitutive relations for volume
change in unsaturated soil. Canadian Geotechnical Journal, Vol. 13, No. 3,
pp.261 276
Fredlund, D.G. and Rahardjo, H., 1993. Soil Mechanics for Unsaturated Soils.
John Wiley & Sons, Inc.
Gibson, R.E., Knight, K. and Taylor, P.W., 1963. A critical experiment to examine
theories of three-dimensional consolidation. Proceedings from the
European Conference on Soil Mechanics and Foundation Engineering,
Weisbaden, Germany.
Henkel, D.J., 1960. The shear strength of saturated remoulded Clay. Proceedings
of the ASCE Research Conference on Shear Strength of Cohesive Soil,
Boulder, Colorado, USA, pp.533-544.
Hill, R., 1950. The Mathematical Theory of Plasticity. Oxford University Press.
Hinton, E. and Owen, D.R.J., 1979 An Introduction to Finite Element
Computations, Pineridge Press Ltd., U.K.
Lambe, T.W. and Whitman, R.V., 1969. Soil Mechanics. John Wiley and Sons,
Inc.
Lambe, T.W. and Whitman, R.V., 1979. Soil Mechanics, SI Version. John Wiley
and Sons, Inc.
Lancaster, P. and Salkauskas, K., 1986. Curve and Surface Fitting: An
Introduction. Academic Press. pp.101-106
Pickering, D.J., 1970. Anisotropic Elastic Properties for Soils. Geotechnique, Vol.
20, No.3, pp.271-276.
Salkauskas, K., 1974. C splines for interpolation of rapidly varying data. Rocky
Mountain Journal of Mathematics, Vol. 14, No. 1, pp.239-250.
Segerlind, L.J., 1984. Applied Finite Element Analysis, 2nd Ed. John Wiley and
Sons, Inc.

Page 394

SIGMA/W

References

Smith, G.D., 1971. Numerical Solution of Partial Differential Equations, Oxford


University Press, Oxford.
Smith, I.M. and Griffiths, D.V., 1988. Programming the Finite Element Method,
2nd Ed. John Wiley and Sons, Inc.
Terzaghi, K., 1943. Theoretical Soil Mechanics. John Wiley & Sons, Inc.
Winterkorn, H.F. and Fang, H.-Y., 1975. Foundation Engineering Handbook. Van
Nostrand Reinhold Company, p.429.
Wroth, C.P. (1975). In-situ measurement of initial stresses and deformation
characteristics, Proceedings of the Specialty Conference in In-situ
Measurement of Soil Properties, ASCE, Raleigh, N.C., pp. 181-230.
Zienkiewicz, O.C. and Taylor, R.L., 1989. The Finite Element Method, 4th Ed.,
Vol. 1. McGraw-Hill.

Page 395

References

SIGMA/W

Page 396

SIGMA/W

Index

Index
Anisotropic elastic model ............112

Element shapes ..............................76

Braced excavations ......................198

Evaluate results in the context of


expected results ..........................35

Burland triangle .............................13

Field variable distribution..............71

Cam-clay
soil parameters..........................135

File
open initial conditions ..............177

Compare alternatives .....................18


Consolidation analysis .................211

Fill placement
surface settlements ...................193

Contouring
Gauss point values....................249
problems when contouring Gauss
point values ...........................250

General guidelines for meshing .....99


Halt iteration ................................240
How not to model ..........................36

Contributing area .........................164

How to model ................................25

Convergence

Hyperbolic soil parameters ..........121

displacement norm ...................236

Identify governing parameters.......20

unbalanced load norm ..............236

Incremental Difference ................243

viewing the graph .....................236

Infinite elements ............................97

Do numerical experiments.............30

Initial conditions ..........................177

Draw

Interface elements..........................96

initial water table ......................176

Interrogate the results ....................35

Element and mesh compatibility ...72

Joining regions...............................90

Element fundamentals ...................70

Linear elastic model.....................110

Element nodes................................70

Mesh with openings.......................85

Page 397

Index

SIGMA/W

Meshing .........................................69

Structured mesh .............................80

Meshing for transient analyses ......95

Structured versus unstructured


meshing ......................................92

Model only essential components..31


Numerical integration ....................74
Numerical modeling ........................9
Quantitative predictions.................16
Secondary variables .......................76
Simplify geometry .........................28
Start simple ....................................29

Surface mesh..................................86
Tension zones ..............................239
Transfinite meshing .......................82
Triangular regions..........................81
understand physical process ..........21
Units
consistent set ..............................67

Start with estimated material


properties ....................................34

Unstructured mesh .........................80

Stop-Restart .................................240

Why model?...................................15

Page 398

Examples of consistent sets of units


Property

Units

Metric

Imperial

Geometry

meters

feet

Unit Weight of Water

F/L

kN/m

pcf

Soil Unit Weight

F/L3

kN/m3

pcf

Cohesion

F/L

kPa

pcf

Pressure

F/L2

kPa

psf

Force

kN

lbs

kPa

psf

E (modulus)

F/L

Examples of consistent sets of units


Property

Units

Metric

Imperial

Geometry

meters

feet

Unit Weight of Water

F/L

kN/m

pcf

Soil Unit Weight

F/L3

kN/m3

pcf

Cohesion

F/L

kPa

pcf

Pressure

F/L2

kPa

psf

Force

kN

lbs

E (modulus)

F/L2

kPa

psf

Vous aimerez peut-être aussi