You are on page 1of 292

Thermal Modeling with

TEMP/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” grossly overstates the case. This book is
the result of a group effort by everybody at GEO-SLOPE. My name is listed as
author primarily for referencing convenience.
At the top of the list of contributors are 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 Carola Preusser
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
TEMP/W Table of Contents

Table of Contents
1 Introduction ............................................................. 1
1.1 Typical applications ..................................................................... 1
Thermal design for roads and airstrips.................................. 1
Ground freezing for soil stabilization ..................................... 2
Insulation design for shallow buried piping ........................... 5
Thawing or freezing beneath heated or chilled structures .... 6
Freezing around chilled pipelines.......................................... 7
Convective cooling of surfaces ............................................. 8
1.2 About this book ............................................................................ 9

2 Numerical Modeling – What, Why and How ......... 13


2.1 Introduction ................................................................................ 13
2.2 What is a numerical model ........................................................ 14
2.3 Modeling in geotechnical engineering ....................................... 18
2.4 Why model ................................................................................. 21
Quantitative predictions....................................................... 21
Compare alternatives in design or modeling methodology . 24
Identify governing parameters............................................. 25
Discover and understand physical process - train our thinking
26
2.5 How to model ............................................................................. 30
Make a guess ...................................................................... 31
Simplify geometry................................................................ 32
Start simple ......................................................................... 34
Do numerical experiments .................................................. 35
Model only essential components ....................................... 37
Start with estimated material properties.............................. 37

Page i
Table of Contents TEMP/W

Interrogate the results ......................................................... 38


Evaluate results in the context of expected results............. 39
Remember the real world.................................................... 39
2.6 How not to model....................................................................... 40
2.7 Closing remarks ......................................................................... 41

3 Meshing ................................................................ 43
3.1 Introduction ................................................................................ 43
3.2 Element fundamentals ............................................................... 44
Element nodes .................................................................... 44
Field variable distribution .................................................... 45
Element and mesh compatibility ......................................... 46
Numerical integration .......................................................... 48
Secondary variables............................................................ 50
Element shapes................................................................... 50
3.3 Regions...................................................................................... 51
Region types ....................................................................... 52
Region points ...................................................................... 52
Region properties ................................................................ 54
3.4 Mesh types & patterns ............................................................... 54
Structured mesh .................................................................. 54
Unstructured mesh .............................................................. 54
Triangular regions ............................................................... 55
Transfinite meshing ............................................................. 57
Mesh with openings............................................................. 59
Surface regions ................................................................... 60
Joining regions .................................................................... 64
3.5 Structured versus unstructured meshing................................... 66
3.6 Meshing for transient analyses .................................................. 69

Page ii
TEMP/W Table of Contents

3.7 Interface elements ..................................................................... 70


3.8 Infinite elements......................................................................... 71
3.9 General guidelines for meshing ................................................. 72
Number of elements ............................................................ 73
Effect of drawing scale ........................................................ 74
Simplified geometry............................................................. 75

4 Material Properties................................................ 77
4.1 Soil heat storage........................................................................ 77
Typical values of volumetric heat capacity.......................... 79
Estimating volumetric heat capacity for soils ...................... 80
4.2 Thermal conductivity.................................................................. 83
Typical values of thermal conductivity................................. 84
Estimating thermal conductivity for soils ............................. 87
4.3 Sensitivity of results to thermal material properties and water
content in soil ............................................................................. 88
4.4 Function data application in solver ............................................ 91
Weighted splines ................................................................. 91
Best-fit splines ..................................................................... 93
4.5 Soil material database and default function estimation ............. 95

5 Boundary Conditions............................................. 97
5.1 Introduction ................................................................................ 97
5.2 Fundamentals ............................................................................ 98
5.3 Temperature boundary conditions ........................................... 100
5.4 Heat flow boundary conditions................................................. 100
5.5 Sources and sinks ................................................................... 104
5.6 Ground surface energy balance .............................................. 105
5.7 Rigorous approach to calculation of ground surface temperatures
................................................................................................. 106

Page iii
Table of Contents TEMP/W

No snow present ............................................................... 107


With snow pack ................................................................. 108
5.8 Traditional approach to computing ground temperatures........ 110
Empirical approach to estimating ground surface temperature
110
5.9 Convective heat transfer surfaces ........................................... 113
Artificial ground freezing application ................................. 113
Pyrex tile cooling application............................................. 116
Time steps for convective surface analysis....................... 117
5.10 Thermosyphon boundary conditions ....................................... 118
5.11 Boundary functions .................................................................. 121
General.............................................................................. 121
Temperature versus time .................................................. 121
Nodal and unit heat flux versus time (Q and q)................. 122
Modifier functions .............................................................. 123
5.12 Null elements ........................................................................... 125

6 Analysis Types.................................................... 127


6.1 Steady-state............................................................................. 127
Boundary condition types in steady-state ......................... 128
6.2 Transient .................................................................................. 128
Initial conditions................................................................. 128
Drawing the initial nodal temperatures.............................. 130
No initial condition ............................................................. 130
6.3 Time stepping - temporal integration ....................................... 131
Finite element temporal integration formulation................ 131
Problems with time step sizes........................................... 132
General rules for setting time steps .................................. 133
6.4 Axisymmetric............................................................................ 134

Page iv
TEMP/W Table of Contents

6.5 Plan view.................................................................................. 136

7 Numerical Issues ................................................ 139


7.1 Convergence............................................................................ 140
Vector norms ..................................................................... 141
7.2 Steep material property functions ............................................ 143
7.3 Gauss integration order ........................................................... 144
7.4 Equation solvers (direct or iterative) ........................................ 145
7.5 Time stepping .......................................................................... 146
Automatic adaptive time stepping ..................................... 147

8 Visualization of Results....................................... 151


8.1 Transient versus steady-state results...................................... 151
8.2 Node and element information................................................. 152
8.3 Equipotential lines.................................................................... 154
Projecting Gauss point values to nodes............................ 155
8.4 Contours .................................................................................. 157
8.5 Energy flow vectors and flow paths ......................................... 158
Calculating gradients and velocities.................................. 158
Energy flow vectors ........................................................... 159
Flow paths ......................................................................... 161
8.6 Flux sections ............................................................................ 161
Flux section theory ............................................................ 162
Flux section application..................................................... 164
8.7 Changes between selected times............................................ 166
8.8 Graphing .................................................................................. 168
8.9 Reporting ................................................................................. 171

9 Modeling Tips and Tricks .................................... 173


9.1 Introduction .............................................................................. 173

Page v
Table of Contents TEMP/W

9.2 Problem engineering units ....................................................... 173


9.3 Convergence Tolerances......................................................... 175
9.4 Flux section location ................................................................ 176
9.5 Source or sink flux values........................................................ 177
9.6 Unit energy flux versus total energy flux?................................ 178
9.7 Stopping and restarting an analysis ........................................ 179
9.8 Element addition and removal ................................................. 180

10 Product Integration Illustrations .......................... 181


10.1 SEEP/W generated pore-water pressures in SLOPE/W stability
analysis .................................................................................... 182
10.2 VADOSE/W generated pore pressures in SLOPE/W stability
analysis .................................................................................... 184
10.3 SEEP/W dissipation of pore pressures generated in a QUAKE/W
earth quake analysis................................................................ 189
10.4 SEEP/W velocity data in CTRAN/W contaminant transport
analysis .................................................................................... 192
10.5 VADOSE/W velocity data in CTRAN/W contaminant transport
analysis .................................................................................... 197
10.6 Density-dependent flow – salt water intrusion ......................... 198
10.7 Ground freezing and water flows (SEEP/W and TEMP/W)..... 201
10.8 Seepage-dependent embankment settlement (SEEP/W and
SIGMA/W)................................................................................ 205
10.9 Uncoupled consolidation.......................................................... 210

11 Illustrative Examples ........................................... 213


11.1 Steady-state heat flow analysis ............................................... 213
Problem definition.............................................................. 213
Solution ............................................................................. 214
11.2 Neumann thaw and freeze analyses ....................................... 217
Problem definition.............................................................. 218

Page vi
TEMP/W Table of Contents

Solution ............................................................................. 218


Depth of thawing front ....................................................... 220
Depth of freezing front....................................................... 222
11.3 Hwang finite element analysis ................................................. 224
Problem definition.............................................................. 224
Solution ............................................................................. 225
11.4 Two-dimensional permafrost analysis ..................................... 227
Problem definition.............................................................. 227
Solution ............................................................................. 229
11.5 Pipeline freezing analysis ........................................................ 232
Problem definition.............................................................. 232
Solution ............................................................................. 233
11.6 Convective tile cooling ............................................................. 235
Problem definition.............................................................. 235
Solution ............................................................................. 236

12 Theory................................................................. 239
12.1 Flow law ................................................................................... 239
12.2 Governing equations................................................................ 239
12.3 Finite element heat flow equations .......................................... 241
12.4 Temporal integration................................................................ 243
12.5 Numerical integration............................................................... 244
12.6 Thermal conductivity matrix ..................................................... 247
12.7 Heat storage matrix ................................................................. 248
12.8 Flux boundary vector ............................................................... 250
12.9 Convective heat transfer.......................................................... 254
12.10 Establishing the phase change region..................................... 254
Selecting the Gauss regions / iteration (GRI) parameter .. 257

Page vii
Table of Contents TEMP/W

13 Appendix A: Interpolating Functions ................... 259


13.1 Coordinate systems ................................................................. 259
13.2 Interpolating functions.............................................................. 261
Field variable model .......................................................... 262
Interpolation function derivatives....................................... 263
13.3 Infinite elements....................................................................... 268
Mapping functions ............................................................. 269
Pole definition.................................................................... 272

References................................................................... 275

Index ............................................................................ 278

Quick Reference Tables............................................... 281

Page viii
TEMP/W Chapter 1: Introduction

1 Introduction
TEMP/W is a finite element software product that can be used to model the
thermal changes in the ground due to environmental changes, or due to the
construction of facilities, such as buildings or pipelines. The comprehensive
formulation makes it possible to analyze both simple and highly complex
geothermal problems, with or without temperatures that result in freezing or
thawing of soil moisture.
TEMP/W cannot be used to predict frost heave at this time. While TEMP/W can
now be used with SEEP/W to model the affect of moving water on heat transfer, it
cannot deal with the very complex and little understood physical coupling between
water, ice, air and soil during heaving.
Another assumption inherent in all TEMP/W analysis is that the moisture content
does not change during the solution. The ground need not be saturated, but the total
moisture content is assumed to be fixed. This means that the summation of ice and
water always add up to the total moisture content by volume.
It is also possible to model situations where there is no moisture in the soil, which
means that TEMP/W can actually be used to model heat transfer through any type
of porous or solid material. Heat transfer in solid materials is actually a very
simplified case of what TEMP/W is capable of analyzing. While this simplified
case will be discussed periodically throughout the book, the main focus is heat
transfer through porous media containing soil, water, ice and air at the same time.

1.1 Typical applications

Thermal design for roads and airstrips


For designing roads and airstrips in areas that experience cold winters, the use of
insulation to control the depth of frost penetration may be used to eliminate
freezing of frost susceptible soils located beneath the subgrade. TEMP/W can be
used to compute the transient distribution of subsurface temperatures during
design.
Figure 1-1 shows the maximum frost penetration below a cleared highway
designed without insulation. Additional TEMP/W modeling could be used to
assess alternative designs using insulation and various depths of drained subgrade
to minimize the frost penetration directly beneath the highway.

Page 1
Chapter 1: Introduction TEMP/W

20
Bare Road Surface
Snow Covered Ground
-20
-16
15 -12
-8
-4
Depth (m)

10
0

0
0 5 10 15 20

Distance (m)

Figure 1-1 Frost Penetration beneath a Cleared Highway

Ground freezing for soil stabilization


For many geotechnical engineering projects, excavation through soft soils is
required. Ground freezing is sometimes used before excavation to provide
stabilized soil. For a successful soil freezing system, things like the appropriate
number of wells, well placement, energy flux requirements, freezing time
requirements, and coverage of frozen zones must all be estimated during design.
TEMP/W is a useful tool for obtaining these estimates, because the ground
freezing process can be modeled on a site-specific basis.
Figure 1-2 shows the results of modeling axisymmetric freezing around a single
brine pipe using TEMP/W. This particular modeling example can be used to
estimate energy flux requirements, freezing time, and freezing coverage for a
single well. An axisymmetric analysis works ok for a single freeze pipe, but
multiple freeze pipes should be modeled in plan view to show the inter-relationship
between adjacent pipes and growth of the frozen wall after adjacent frozen
columns have connected. This is discussed in more detail in the illustrative
examples chapter.

Page 2
TEMP/W Chapter 1: Introduction

Brine Pi pe
Gro und Surface

8.0 10

7.5

7.0

6.5
5
6.0
-20

5.5
-15

5.0
-10

4.5
-5

4.0

3.5
0

3.0

5
2.5

2.0

1.5

1.0

0.5

10
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Distance (m)

Figure 1-2 Frost bulb around freezing pipe


In Figure 1-3, the convective heat transfer boundary condition option has been
used to compute the actual heat removal from each freeze pipe in a mine shaft
freezing program. The heat removed is a function of ground temperature, brine

Page 3
Chapter 1: Introduction TEMP/W

temperature, brine flow rate, and pipe geometry. TEMP/W computes this value at
each time step as it changes during the active freezing period. In this figure, the
actual pipe geometry is entered into TEMP/W and then the heat flux boundary
condition is applied at a single node in end view. This way, you do not have to
create a fine mesh around the pipe and a coarse mesh far away from the pipe. The
other advantage of the convective heat transfer boundary condition approach is that
you do not have to assume the freeze pipe wall temperature, which is always less
than the actual brine temperature.

Figure 1-3 Plan view of freeze wall growth around a shaft excavation
Note: in Figure 1-3, the ground water being frozen is Sodium Chloride rich so the
actual freeze point temperature was set in the model to be minus 11 degrees
Celsius. The "frozen ground" isotherm is the light blue line. In this case, the

Page 4
TEMP/W Chapter 1: Introduction

ground water being frozen was set to freeze at zero degrees Celsius, so the frozen
ground isotherm is at the zero degree point.

Insulation design for shallow buried piping


Many buried pipes, such as water supply pipes, do not require insulation design
because they are typically placed below the frost line. Even when frost does
penetrate to buried pipes, sufficient fluid flow through the pipes carries enough
latent heat that the fluid does not freeze while flowing through the pipe. However,
in cases where there are limitations on trench depth and/or intermittent flow
through the buried pipe, a design which includes insulation to protect against
freezing may be required. TEMP/W can be used in these cases.
Figure 1-4 shows the results of an analysis performed during the design of a
shallow buried piping system for a groundwater remediation project. The pipes
were located beneath inverted-U insulation at a depth of about 1.0 m below the
ground surface. Insulation design in this case is important because of the
limitations on trench depth and the intermittent water flow within the buried pipes.
The figure below shows the maximum frost penetration, which occurs in early
spring as the thaw front is migrating downward from the ground surface.

Backfill Ground Surface

3.0
1 1
0
-0.5
Native Soil
0

2.5 -1
Depth (m)

Insulation 0
-1
2.0

-0.5

1.5

1.0

0.5

0.25

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

Distance (m)

Figure 1-4 Frost penetration around shallow buried piping

Page 5
Chapter 1: Introduction TEMP/W

Thawing or freezing beneath heated or chilled structures


Construction of heated structures on permafrost is complicated by the fact that
excessive permafrost thawing can have detrimental effects on the structure’s
foundation. Successful design of these structures therefore depends on an analysis
of the heat transfer from the structure into the underlying soil. TEMP/W can be
used to perform these analyses.
Figure 1-5 illustrates the results of an axisymmetric analysis of heat transfer below
a warm circular structure. The results show the long term thawing of the
permafrost beneath the structure. This analysis shows the permafrost thawing in
the absence of any insulation in the foundation of the structure. Additional
modeling could show the effect of installing insulation to minimize thawing.

25
W arm Circular Structure

Ground Surface

20 5
4
D th ( )

3
2

15 1 -0 .5
0 .5

10 0

5 0.5

0
0 5 10 15 20

Distance (m)

Figure 1-5 Thawing permafrost beneath a heated structure


In the next example, thermosyphons have been installed beneath the building to
help stabilize the permafrost. Outside the building, the climate data option has been

Page 6
TEMP/W Chapter 1: Introduction

used to predict the ground temperature beneath the snow or on the exposed ground
surface. The thermosyphon is a special type of boundary condition in TEMP/W.

Figure 1-6 Thermosyphons used to support permafrost beneath a


structure

Freezing around chilled pipelines


Chilled pipelines, for transporting commodities such as natural gas, typically
traverse large distances through a diversity of soil types. An important factor to
consider in the design of chilled pipelines is the size of the frost bulb that will
develop in the soil around the pipe. In frost susceptible soils, the frost action
around the pipe will cause frost heave which will lift the pipe vertically. This is
especially troublesome where the pipeline crosses the interface between a frost
susceptible soil and soil which is not very frost susceptible. In these cases,
differential frost heave will cause the pipe to bend, thus placing stresses on the
pipe. Pipeline design therefore depends on factors such as native soil type, backfill
material, pipe diameter, pipe wall thickness, pipe temperature, and amount of
insulation applied to the outside of the pipe. Central to the design is the heat
transfer from the soil to the pipe in both the transient, start-up phase, and in the
long term operational phase. TEMP/W can be used to perform the heat transfer
analyses required for pipeline design.

Page 7
Chapter 1: Introduction TEMP/W

Figure 1-7 shows the long term distribution of temperature around a chilled
pipeline. The analysis shows the depth of frost penetration below the pipe and can
be used to determine the heat flux into the pipe from the surrounding soil. In this
example, fluid flow is affecting the shape of the thermal profile. TEMP/W can
account for transient fluid flow in the heat transfer analysis. It does so my
simultaneously solving SEEP/W which provides water contents and flow
velocities.
1.7
1.6
1.5
1.4
1.3
1.2
1.1
1.0 -2
Depth (m)

0.9
0.8 0
0.7 1
0.6
0.5 2
0.4 2.5
0.3
0.2
0.1
0.0
-0.1
-0.2
-0.3
-1.6 -1.4 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

Distance (m)
Figure 1-7 Freezing around a chilled pipeline

Convective cooling of surfaces


TEMP/W has a special boundary condition that considers the heat removal from
the model due to a fluid flowing past the edge of the model. Consider Figure 1-8
which shows a model set up to compute the cooling of a fabricated 10 mm thick
Pyrex tile for two cases: still air and forced air. In this case the tile cools from
140°C to about 25°C; therefore, phase change is not an issue and the model is quite
easy to set up and solve. TEMP/W can compute the actual removal of heat from
the model surface based on the specified fluid temperature and the continually
changing model boundary temperature.

Page 8
TEMP/W Chapter 1: Introduction

Pyrex Tile
140C initial temperature
Insulated on bottom
K = 1.4W/mC, Cp=835J/kgC, Density 2225 kg/m3
20

18

16 Air, 10m/s, 25C


Convective heat transfer coefficient = 14 W/mC (still air)
Distance(m) (x 0.001)

14 Convective heat transfer coefficient=55.7 W/mC (moving air)

12

10

6
Tile

2
Insulation material
0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40

Distance(m) (x 0.001)

Figure 1-8 Heat flow vectors during air cooling of a fabricated Pyrex
tile

1.2 About this book


Modeling the flow of heat through soil (with or without water changing phases)
with a numerical solution can be very complex. Natural soil deposits are generally
highly heterogeneous and non-isotropic. In addition, boundary conditions often
change with time and cannot always be defined with certainty at the beginning of
an analysis. In fact, the correct boundary condition can sometimes be part of the
solution, as is the case for artificial ground freezing, where the amount of heat
removed from the ground through the chilled pipe depends on the difference
between the pipe fluid temperature and the ground temperature itself.
Furthermore, when a soil becomes partially frozen, the coefficient of thermal
conductivity becomes a function of the negative temperature in the soil. It is the
temperature that is the primary unknown and needs to be determined. Because of
this, an iterative numerical technique is required to match the computed
temperature and the material property, making the solution highly non-linear.
These complexities make it necessary to use some form of numerical analysis to
analyze thermal problems for all but the simplest cases. A common approach is to

Page 9
Chapter 1: Introduction TEMP/W

use finite element formulations, and TEMP/W, the subject of this book, is an
example of a numerical software tool.
While part of this document is about using TEMP/W to do thermal analyses, it is
also about general numerical modeling techniques. Numerical modeling, like most
things in life, is a skill that needs to be acquired. It is nearly impossible to pick up
a tool like TEMP/W and immediately become an effective modeler. Effective
numerical modeling requires some careful thought and planning, and it requires a
good understanding of the underlying physical fundamentals. Aspects such as
discretization of a finite element mesh and applying boundary conditions to the
problem are not entirely intuitive at first. Time and practice are required to become
comfortable with these aspects of numerical modeling.
A large portion of this book focuses on the general guidelines of how to conduct
effective numerical modeling. In fact, chapter 2 is devoted exclusively to
discussions on this topic. The general principles discussed there apply to all
numerical modeling, including thermal analysis. In this chapter, however, the
discussion pertains to seepage analyses because it is more intuitive to most people.
It is easier to mentally understand the flow of water than heat.
Broadly speaking, there are three main parts to a finite element analysis. The first
is Discretization: dividing the domain into small areas called elements. The second
part is specifying and assigning material properties. The third is specifying and
applying boundary conditions. Separate chapters have been devoted to each of
these three key components within this document.
Frozen / unfrozen numerical modeling is a highly non-linear problem that requires
iterative techniques to obtain solutions. Numerical convergence is consequently a
key issue. Also, the temporal integration scheme which is required for a transient
analysis is affected by time step size relative to element size and material
properties. These and other numerical considerations are discussed in a chapter
titled Numerical Issues.
Two chapters have been dedicated to presenting and discussing illustrative
examples. One chapter deals with examples where geotechnical solutions are
obtained by integrating more than one type of analysis, and the other chapter
presents and describes how a series of different geotechnical problems can be
solved.
A full chapter is dedicated to theoretical issues associated with the finite element
solution of the partial differential heat flow equation for frozen and unfrozen soils.

Page 10
TEMP/W Chapter 1: Introduction

Additional finite element numerical details regarding interpolating functions and


infinite elements are given in Appendix A.
The chapter entitled “Modeling Tips and Tricks” should be consulted to see if
there are simple techniques that can be used to improve your general modeling
method. You will also gain more confidence and develop a deeper understanding
of finite element methods, TEMP/W conventions and data results.
In general, this book is not a “how to use TEMP/W” manual. This is a book about
how to model. It is a book about how to engineer thermal problems using a
powerful calculator; TEMP/W. Details of how to use various program commands
and features are given in the on-line help inside the software.

Page 11
Chapter 1: Introduction TEMP/W

Page 12
TEMP/W Chapter 2: Numerical Modeling

2 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 GeoStudio 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
user’s 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, you do
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 user’s
ability to guide the analysis process that makes it a powerful tool. The spreadsheet
can do all the mathematics, but it is the user’s 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 GeoStudio.
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 13
Chapter 2: Numerical Modeling TEMP/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.

2.2 What is a numerical model


A numerical model is a mathematical simulation of a real physical process.
TEMP/W is a numerical model that can mathematically simulate the real, physical
process of heat flowing through a fully or partially saturated, frozen or thawed
particulate or solid medium. Numerical modeling is purely mathematically and in
this sense is very different than scaled physical modeling in the laboratory or full-
scaled field modeling.
Newman and Maishman (2000) presented details of an artificial ground freezing
project involving freezing a barrier wall around three sides of an underground
uranium ore body in northern Canada. The freezing was necessary in order to seal
off potential excessive inflow of high pressure water from the south, north and
west sides of the ore body as illustrated in Figure 2-1. This figure shows the
southern row of freeze pipes that were installed at 2m intervals around the region.
A freeze pipe operates circulates chilled brine which in turn removes heat from the
ground and eventually freezes all water between the pipes. This results in a frozen
barrier wall.
Figure 2-2 shows a plan view of the three freeze walls through different types of
ground after about six months of active freezing. It is interesting to note the
different rates of freeze wall thickness that develop in the different materials.
While different materials have different thermal conductivity values, other factors
come into play such as the ground water content by volume in each material and
the initial ground temperature.

Page 14
TEMP/W Chapter 2: Numerical Modeling

Athabasca
Sandstone
530m Level Development

High pressure water


Ore/Clay
Mix

South Freeze Row

Abrasive Basement
Dry
Basement
Sand and Gravel Rock
640m Level Development

Figure 2-1 End view of freeze wall around underground ore body
Figure 2-3 shows a comparison of predicted and observed ground temperature
decay over time for various ground types, and at various distances offset from the
row of freeze pipes. TEMP/W was able to accommodate the wide range of ground
types, water contents, thermal properties and initial in-situ ground temperatures in
its analysis. In particular, once it was realized that the initial ground temperature in
the ore rock itself was much higher than the surrounding ground temperature (+22
C versus +10 C), TEMP/W was able to model the actual time it would take to
freeze this ground and, therefore, provide a more reasonable estimate of when
production could begin once development of a safe water sealing barrier around
the mining region was established.
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 15
Chapter 2: Numerical Modeling TEMP/W

Q u a rtz ite

S a n d s to n e

O re Q u a rtz ite

Figure 2-2 Plan view of freeze wall growth around ore body after 6
months of active freezing (south freeze wall is at bottom of figure,
crosses mark surveyed position of freeze pipes at modeled elevation)

Page 16
TEMP/W Chapter 2: Numerical Modeling

Temperature Comparison for Different Ground Types


25.0
Measured at 2m offset from freeze row in high grade ore/clay
Predicted
20.0 Measured at 3.7m offset from freeze row in low grade sandy/clay
Predicted
15.0 Measured at 3.5m offset from freeze row in barren sandstone
Temperature (C)

Predicted
10.0

5.0

0.0

-5.0

-10.0

-15.0
0 2 4 6 8 10 12

Time (Months)

Figure 2-3 TEMP/W analysis versus measured data


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.

Page 17
Chapter 2: Numerical Modeling TEMP/W

• 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 thermal heat flow there may also be pore-water pressure 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 TEMP/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. TEMP/W is
formulated only for heat flow that follows a Darcy type conduction law with a
modification for convective heat flow when combined with SEEP/W.
The important point to remember is that the mathematical formulations
implemented in software like GeoStudio 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 2-4. This has come to be known as the Burland triangle
(Burland, 1987; Burland, 1996).

Page 18
TEMP/W Chapter 2: Numerical Modeling

Ground
profile

Empiricism,
Precedent

Soil
Modeling
behaviour

The soil mechanics triangle

Figure 2-4 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.
Of great significance is that, in Burland’s 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 2-5.
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

Page 19
Chapter 2: Numerical Modeling TEMP/W

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.
Genesis / geology

Site investigation,
Ground
ground description
Profile

Empiricism,
precedent,
experience,
risk management

Soil
Modeling
Behaviour

Lab / field testing, Idealization followed by


observation, evaluation. Conceptual
measurement or physical modeling,
analytical modeling

Figure 2-5 The enhanced Burland triangle (after Anon. 1999)

Page 20
TEMP/W Chapter 2: Numerical Modeling

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,


• 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 seepage quantity, 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 quantity of seepage,
for example, is in large part controlled by the hydraulic conductivity and, as a
result, changing the hydraulic conductivity by an order of magnitude will usually
change the computed seepage quantity by an order of magnitude. The accuracy of
quantitative prediction is directly related to the accuracy of the hydraulic
conductivity specified. Unfortunately, for a heterogeneous profile, there is not a
large amount of confidence about how precisely the hydraulic conductivity can be
specified. Sometimes defining the hydraulic conductivity within an order of

Page 21
Chapter 2: Numerical Modeling TEMP/W

magnitude is considered reasonable. The confidence you have defining the


hydraulic conductivity depends on many factors, but the general difficulty of
defining this soil parameter highlights the difficulty of undertaking modeling to
make quantitative predictions.
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 2-6 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 2-6. 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 2-6 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
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

Page 22
TEMP/W Chapter 2: Numerical Modeling

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.
Deflection (mm)
-60 -50 -40 -30 -20 -10 0 10
0

12 Depth below surface (m)

16

20

24

computed

28
measured

32
-60 -50 -40 -30 -20 -10 0 1

Figure 2-6 Comparison of predicted and measured lateral movements


of a shoring wall (after Carter et al, 2000)

Page 23
Chapter 2: Numerical Modeling TEMP/W

Compare alternatives in design or modeling methodology


Numerical modeling is useful for comparing alternatives. Keeping everything 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 frost penetration on a roadway over a winter. There are
two types of boundary conditions that could be applied at the ground surface: a
detailed climate analysis or a simplified analysis that assumes ground temperature
is equal to average monthly air temperature (which also assumes the road is kept
clear of any insulating snow cover). The detailed climate analysis considers the
daily warming and cooling of the air while setting the ground temperature to be
equal to the average daily air temperature ignores any diurnal variation or any
affect on ground temperature from the ground heat flux itself or snow cover etc. A
simple one-dimensional model can be set up as an initial test to just see how big
the impact of both boundary options is. It may be significant, it may not.
Regardless, it’s MUCH faster and easier to test these alternatives in a simple model
before you build the detailed two-dimensional model.

Temperature vs. Y
15

5.0000e+000

14

2.7500e+002

13
Y

2.8000e+002
12

2.8500e+002
11

3.6500e+002
10
-10 -5 0 5 10 15 20 25
Temperature

Figure 2-7 Near surface temperatures over time computed with daily
detailed data

Page 24
TEMP/W Chapter 2: Numerical Modeling

The actual computed values are not of significance in the context of this
discussion. What is significant is that ignoring detail in these cases produces
ground temperatures that get colder in winter and do not get warmer in summer,
based on energy balances at the surface on bare soil or soil with a snow pack. This
is an example how a simple model set up in TEMP/W can be used to quickly
compare alternatives prior to a more detailed analysis. Another thing we may want
to check at this stage of our modeling is how sensitive the results are to slight
variations in thermal conductivity.

Temperature vs. Y
15

5.0000e+000
14

2.7500e+002
13
Y

2.8000e+002
12

2.8500e+002
11

3.6500e+002
10
-20 -15 -10 -5 0 5 10
Temperature

Figure 2-8 Near surface temperature profile using ground temperature


set to average daily air temperature

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

Page 25
Chapter 2: Numerical Modeling TEMP/W

as a tornado plot, such as shown in Figure 2-9. A similar type plot may be
generated based on TEMP/W simulations to determine the optimum insulation
thickness, thermal conductivity or water content for a design that prevents freezing
temperatures from developing beneath a slab foundation, for example.
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.
Base Case

thinner
Thickness
of Growth Medium
high low bare surface
Transpiration

low high
Hydraulic Conductivity
of Compacted Layer
deep shallow
Root Depth
low high
Hydraulic Conductivity
of Growth Medium

Decreasing Increasing
Net Percolation Net Percolation

Figure 2-9 Example of a tornado plot (O’Kane, 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 2-10. 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

Page 26
TEMP/W Chapter 2: Numerical Modeling

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
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 2-11 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. For a full explanation of
this relationship, refer to the SEEP/W book. For this example, accept that at high
suctions the coarse material is less conductive than the fine material.

Fine
Coarse
Coarse Fine
OR
Material to be Material to be
protected protected

Figure 2-10 Two possible earth cover configurations

Page 27
Chapter 2: Numerical Modeling TEMP/W

1.00E-04

1.00E-05
Coarse
Fine
1.00E-06
Conductivity

1.00E-07

1.00E-08

1.00E-09

1.00E-10
1 10 100 1000

Suction

Figure 2-11 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
material over the coarse material results in infiltration being drained laterally
through the fine layer, as shown in Figure 2-12. 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 2-13. The design of fine soil over coarse soil
may work, but only in arid environments. The occasional cloudburst 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

Page 28
TEMP/W Chapter 2: Numerical Modeling

parameters. Consider the functions in Figure 2-14. 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 2-12 Flow diversion under low infiltration

Intense rainfall rates


Fine
Coarse

Figure 2-13 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.

Page 29
Chapter 2: Numerical Modeling TEMP/W

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
Coarse
Fine
1.00E-06
Conductivity

1.00E-07
Intense
Rainfall
1.00E-08
Low to Modest
Rainfall
1.00E-09

1.00E-10
1 10 100 1000

Suction

Figure 2-14 Conductivities under low and intense infiltration


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

Page 30
TEMP/W Chapter 2: Numerical Modeling

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. If there is no resemblance between what is expected and what is
computed, then either the preliminary mental picture of the situation was not right
or something has been inappropriately specified in the numerical model. Perhaps
the boundary conditions are not correct or the material properties specified are
different than intended. The difference ultimately needs 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 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
thermal distribution, or is the quantity of heat added or removed from the system?
If your main objective is to determine the temperature distribution, there is no need
to spend a lot of time on establishing the thermal conductivity; any reasonable
estimate of conductivity is adequate. If on the other hand your main objective is to
estimate heat flow quantities, then a greater effort is needed in determining the
thermal conductivity.
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

Page 31
Chapter 2: Numerical Modeling TEMP/W

good practice to simplify the problem so you can make a guess and then add
complexity in stages, such 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
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 2-15 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 2-16. 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

Page 32
TEMP/W Chapter 2: Numerical Modeling

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 2-15 Example of a stratigraphic cross section


(from National Research Report 1990)

Figure 2-16 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.
Another example of keeping the geometry simple relates to modeling actual freeze
pipe geometry, versus having a single freeze pipe be represented by a single node.

Page 33
Chapter 2: Numerical Modeling TEMP/W

Consider the illustrations in Figure 2-17, where on the left the actual pipe geometry
is built into the mesh, and on the right the freeze pipe is represented by a single
nodal boundary condition. Depending on the boundary condition applied in the
analysis, the mesh on the right will give equally reliable results and it is much
simpler to set up. This is especially true if you need to model many freeze pipes in
a complex pattern to assess the freezing that develops over the region.
When a small geometry freeze pipe is modeled (e.g. several inch diameter),
TEMP/W has the ability to apply a special convective heat transfer boundary
condition that includes the actual pipe surface area in the boundary value. With this
option, the actual heat removed from the ground will be almost identical in both
models and there is no need to spend time building very small detail into the actual
finite element mesh.

Figure 2-17 Detailed pipe geometry versus simplified geometry

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.

Page 34
TEMP/W Chapter 2: Numerical Modeling

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 build 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.
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 the process migrates 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 that 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. 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 2D case as shown
in Figure 2-18. The total temperature difference is 1 degree and the thermal
conductivity is 1 W/mC. The gradient under steady-state conditions is the
temperature difference divided by the length, and is this case the gradient is 0.1.
The resulting total heat flow through the system is the cross-sectional area
multiplied by the gradient, which should be 0.3 W. The flux section that goes
through the entire section confirms this result. There are flux sections through
Elements 16 & 18. The flow through each element is 0.1 W, which is correct since
each element represents one-third of the area.

Page 35
Chapter 2: Numerical Modeling TEMP/W

Another way to check the computed results is to look at the node information.
When a temperature is specified, TEMP/W computes the corresponding nodal flux.
In TEMP/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.

1.0000e-001
3 6 9 12 15 18 21 24 27 30
3.0000e-001

2 5 8 11 14 17 20 23 26 29

1 4 7 10 13 16 19 22 25 28
1.0000e-001

Figure 2-18 Horizontal heat 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 that 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

Page 36
TEMP/W Chapter 2: Numerical Modeling

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.
Including unnecessary features and trying to model adjacent materials with
extreme contrasts in material properties create numerical difficulties. The hydraulic
conductivity difference between the core and shell of a dam, for example, 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 non-
essential 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.

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. Simple estimates of material properties and
simple property functions are more than adequate for gaining an understanding of
the flow 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 more than adequate for determining the importance of the various
properties for the situation being modeled.

Page 37
Chapter 2: Numerical Modeling TEMP/W

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.

Interrogate the results


Powerful numerical models, such as TEMP/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 defining the model. Is the
intended property function being applied to the correct soil? Or, are the initial
conditions as you assumed?
TEMP/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.

Page 38
TEMP/W Chapter 2: Numerical Modeling

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.

“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 full-
cores they’re wanting!”

** Origins of this figure are unknown at time of printing

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.

Page 39
Chapter 2: Numerical Modeling TEMP/W

The following cartoon portrays an extreme situation, but underscores a problem


that exists when uneducated or inexperienced users try to use powerful software
tools.

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
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 advantage of the power the software can offer.

Page 40
TEMP/W Chapter 2: Numerical Modeling

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 GeoStudio, 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, GeoStudio, 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 41
Chapter 2: Numerical Modeling TEMP/W

Page 42
TEMP/W Chapter 3: Meshing

3 Meshing

3.1 Introduction
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 geo-
environmental 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 43
Chapter 3: Meshing TEMP/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.

3.2 Element fundamentals

Element nodes
One of the main features of a finite element are the nodes. Nodes exist at the
corners of the elements or along the edges of the elements. Figure 3-1 and Figure
3-2 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 TEMP/W formulation, the primary field variable is temperature.
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 thermal 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 thermal analysis there is only one degree of
freedom at each node, which is the temperature. The number of finite element
equations to be solved is equal to the number of nodes used to define the mesh. In

Page 44
TEMP/W Chapter 3: Meshing

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-pressure. For a coupled consolidation
analysis the total number of equations required to solve the problem is three times
the number of nodes.
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., temperature). 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 3-1 illustrates this situation. Elements with nodes existing
at the corners are referred to as first-order elements.

Figure 3-1 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
thermal formulation the primary unknown is the temperature. The derivative of
temperature with respect to distance is the thermal gradient and the gradient is
therefore constant within a first order element.
With three nodes defined along an edge, we can write a quadratic equation
describing the distribution of the primary unknown within the element.

Page 45
Chapter 3: Meshing TEMP/W

Consequently the distribution of the primary unknown can be curved as shown in


Figure 3-2. The derivative of the quadratic temperature 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 3-2 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 value as in a thermal formulation, 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 3-3. 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 3-3. Common
nodes are required to connect the elements as shown in Figure 3-4. Node 11, for
example, is common to Elements 5, 6, 8 and 9.
Mixing elements of a different order can also create incompatibility. Figure 3-5
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

Page 46
TEMP/W Chapter 3: Meshing

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 careful about creating adjoining regions. The illustration in Figure 3-3 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 20 24 40 44 48
4 8 15 28 32 36
9 12 21 24
3 6 15 18
14 19 23 39 43 47
3 7 13 27 31 35
8 11 20 23
2 5 14 17
12 18 22 38 42 46
2 6 11 26 30 34
7 10 19 22
1 4 13 16
10 17 21 37 41 45
1 5 9 25 29 33

Figure 3-3 Disconnected elements – lack of compatibility

Page 47
Chapter 3: Meshing TEMP/W

4 8 12 16 20

3 6 9 12
3 7 11 15 19

2 5 8 11
2 6 10 14 18

1 4 7 10
1 5 9 13 17

Figure 3-4 Connected elements – compatibility satisfied


5 8 13 18 23
6 10
4 12 17 22
2
5 9
3 7 11 16 21
4 8
2 10 15 20
1
3 7
1 6 9 14 19

Figure 3-5 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 48
TEMP/W Chapter 3: 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.

Element Type Integration Points Comments


4-noded quadrilateral 4 Default
8-noded quadrilateral 4 or 9 4 is the default
3-noded triangle 1 or 3 3 is the default
6-noded triangle 3 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.
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

Page 49
Chapter 3: Meshing TEMP/W

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. Again, in a thermal formulation the
primary unknowns are the temperatures at the nodes. Once the primary unknowns
have been computed, other variables of interest can be computed, such as the
thermal gradients 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 3-6 shows a
relative performance comparison of element shapes.

Page 50
TEMP/W Chapter 3: Meshing

Ideal

Good

Acceptable

Poor

Unacceptable

Figure 3-6 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.

3.3 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

Page 51
Chapter 3: Meshing TEMP/W

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
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 3-7 illustrates a domain constructed using
one quadrilateral and two triangular regions. Figure 3-8 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 52
TEMP/W Chapter 3: Meshing

1 4

2 3

Figure 3-7 Illustration of a quadrilateral region and triangular regions


10 9

8 7

6 5

4 3

1 2

Figure 3-8 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 3-9 Regions of different size

Page 53
Chapter 3: Meshing TEMP/W

Points are also required in order to join regions of different sizes and to control the
meshing for specific purposes. Figure 3-9 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 regeneration, then
any manually changed element properties will be set back to the default region
values.

3.4 Mesh types & patterns

Structured mesh
Figure 3-10 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 3-11 shows the same section as in Figure 3-10, 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 54
TEMP/W Chapter 3: 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 divisions 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 1

8 2 3 7

10 4 5 6

Figure 3-10 Structured mesh


26 25

24 23

21 22

Figure 3-11 Unstructured mesh

Triangular regions
GeoStudio has a special structured pattern for triangular regions. The next figure
(Figure 3-12) shows a typical triangular region with a structured mesh. The

Page 55
Chapter 3: Meshing TEMP/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.
3

1 2

Figure 3-12 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
3-12, 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 the best when it is used with
other structured mesh regions, although this is not a strict requirement.

Page 56
TEMP/W Chapter 3: Meshing

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
reflects the perimeter of the region as illustrated in Figure 3-13. This is a
quadrilateral region with one curved side. Note how the mesh reflects the curvature
of the top side of the quadrilateral.

Figure 3-13 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 3-14 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 57
Chapter 3: Meshing TEMP/W

11 19 18 17

12
20

13 14 15 16

Figure 3-14 Region corners in structured mesh


In Figure 3-14 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 3-15 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 8 5 4

7 6

2 3

Figure 3-15 Transfinite mesh in a u-shaped region

Page 58
TEMP/W Chapter 3: Meshing

Mesh with openings


Regions can have openings, as shown in Figure 3-16 to simulate drains or tunnels.
The radius is a user-specified variable and the exact position of the centre 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 3-17. 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 3

1 2

Figure 3-16 Mesh with an opening


4 3

1 2

Figure 3-17 Meshing opening with ring of elements around the


opening

Page 59
Chapter 3: Meshing TEMP/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 goes 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 3-18 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 TEMP/W, a climate boundary condition can only be applied on a
surface region mesh. 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. 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 TEMP/W analysis, the surface mesh will
also be part of a SLOPE/W analysis, since GeoStudio uses only one mesh
definition within a single data file.
Once the main soil profile has been meshed, a special Draw Surface Mesh
command can be used to build up a single or multi-layer regions 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 shape. 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 extreme in a direction

Page 60
TEMP/W Chapter 3: 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 3-18 Illustration of a surface region mesh


Surface mesh regions have special viewing options. Consider the two meshes
illustrated in Figure 3-19. 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 3-20 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 regard to the figure.
Notice that the 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 made of the same soil as the existing
ground. Consider if the bottom layer of the surface soil were 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

Page 61
Chapter 3: Meshing TEMP/W

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 note from Figure 3-20, 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 3-19 Surface region mesh with details off (left) and on (right)

Figure 3-20 Close up of cover 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 3-21 and Figure 3-22. 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 unnatural, but the shape of the

Page 62
TEMP/W Chapter 3: Meshing

individual elements right at the transition points creates numerical problems if


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 upslope is still an
infiltration node. Basically, it is better to build the mesh to look somewhat natural.
g

Figure 3-21 GeoStudio region mesh showing rounded surface slope


breaks

Figure 3-22 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

Page 63
Chapter 3: Meshing TEMP/W

such that the region points can be moved slightly to create a rounded profile. This
is the case in Figure 3-23 below, where three region points are used at both the toe
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 3-23 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.

Page 64
TEMP/W Chapter 3: Meshing

• 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.
• 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 3-24. 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 8

2
7 3
5
6 4

1 2

Figure 3-24 Regions joined along jagged line


9 8

2
7

5 3
6
4
1
1 2

Figure 3-25 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

Page 65
Chapter 3: Meshing TEMP/W

this results in a mesh like in Figure 3-25. 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.

3.5 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 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 you
have 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 3-26 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 3-26 Unstructured and structured meshes

Page 66
TEMP/W Chapter 3: Meshing

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
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 3-27 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.
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 3-28. This is more
complicated in an unstructured mesh. It is not possible to pick up a group of nodes
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.
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.

Page 67
Chapter 3: Meshing TEMP/W

Figure 3-27 Comparison of flow vectors in an un-structured and a


structured mesh
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.

Vertical profiles

Figure 3-28 Vertical profiles through a mesh


Another advantageous feature of the 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.

Page 68
TEMP/W Chapter 3: Meshing

Unstructured meshing has no such restrictions. This makes it easy to transition


from a fine mesh to a coarse 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. You can select and use whatever you wish
based on personal preference or as dictated by problem requirements.

3.6 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 advection-
dispersion 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. The same issues can arise when using
TEMP/W with SEEP/W where seepage velocities are used in the thermal model to
compute the heat moving with the water. This part of the heat transfer equation is a

Page 69
Chapter 3: Meshing TEMP/W

first order term and can result in overshoot of the computed temperatures if the
time steps are too large or the elements too small.
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-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.

3.7 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 70
TEMP/W Chapter 3: Meshing

3.8 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 the long-term average ground temperature at some large distance from an
artificial ground freezing site; or the constant thermal gradient with depth deep in
the soil. In a numerical model it is necessary to define a boundary at some practical
distance from the area of interest. 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 3-29 illustrates the concept.

X1 X2
Far field nodes

1 8 4

5 7

Pole 2 6 3

Figure 3-29 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
3-29 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 71
Chapter 3: Meshing TEMP/W

Node 8 (or 6). Nodes 3, 4 and 7 are considered to be at infinity and are not used in
the 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
side edges that are parallel to the infinity direction. In the element in Figure 3-29,
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.

3.9 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 of how

Page 72
TEMP/W Chapter 3: Meshing

the mesh is related to the solution and vise 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.
The mesh should represent a simplified abstraction of the actual complex
geometric field configuration.

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

Page 73
Chapter 3: Meshing TEMP/W

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 3-30. The top
part of the figure shows a nice mesh at 10V:100H, a 10 times vertical
exaggeration. The same mesh at a scale of 100V:100H appears at the bottom of
Figure 3-30. 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 that 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.

Page 74
TEMP/W Chapter 3: Meshing

Figure 3-30 Mesh at an exaggerated (upper) and at a true scale (lower)

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,
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

Page 75
Chapter 3: Meshing TEMP/W

are rough estimates. There needs to be a balance in complexity between all the
aspects of a finite element analysis, including the geometry.
Overcomplicating 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 an understanding as to 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 76
TEMP/W Chapter 4: Material Properties

4 Material Properties
This chapter describes the various soil thermal properties that are required in the
solution of the TEMP/W partial differential equation. It is important to have a clear
understanding of what the soil properties mean and what influence they have on
the type of results generated. This chapter is not meant to be an all inclusive
discussion of these issues. It is meant to highlight the importance of various
parameters and the implications associated with not defining them adequately.
Well defined soil properties can be critical to obtaining an efficient solution of the
finite element equations. When is it ok to guess at a function and when must you
very carefully define one? This chapter will address these issues.

4.1 Soil heat storage


Fundamental to the formulation of a general freeze-thaw thermal analysis of soil
systems is an understanding of how energy stored in the soil varies as the soil
temperature changes. An example of this relationship is shown in Figure 4-1. The
function represents the relative energy required for the soil medium to sustain a
certain temperature. The steep part of the function in the region of the freeze-thaw
front represents the latent heat absorbed or released by a soil due to phase changes
of the soil water. The slope of the function away from the freeze-thaw zone
represents the volumetric heat capacity of the frozen and unfrozen zones.
Except in the case of a pure water medium, the water within the soil changes from
liquid to ice or from ice to water over a temperature range. In other words, not all
of the water within the soil experiences a phase change at a single temperature.
The percentage of the soil water volume that remains unfrozen at a certain
temperature is referred to as the unfrozen water content.
The unfrozen water content characteristic of a soil can be expressed as a function
of temperature. An example of the unfrozen water content function for a coarse
material is illustrated in Figure 4-2. In TEMP/W, the unfrozen water content is
expressed as a value between 0 and 1.0. An unfrozen water content of 1.0 means
that 100% of the water in the soil medium is unfrozen. Similarly, an unfrozen
water content of 0.0 means that 0% of the water in the soil medium is unfrozen.

Page 77
Chapter 4: Material Properties TEMP/W

Figure 4-1 An example of an energy storage function

1.0

0.8
Unfrozen W. C.

0.6

0.4

0.2

0.0
-300 -250 -200 -150 -100 -50

Temperature (x 0.001)

Figure 4-2 Unfrozen water content function for coarse material

Page 78
TEMP/W Chapter 4: Material Properties

In general, above the freezing point (generally around 0 °C or 32 °F), all of the
water within the soil medium is unfrozen. Below the freezing point, the portion of
water that remains unfrozen gradually decreases as the temperature decreases.
Finally, below some negative temperature (e.g., -10 °C), all water within the soil
medium is frozen.
TEMP/W uses the unfrozen water content function of a soil to estimate the latent
heat absorbed or released by the soil medium due to the phase changes of the soil
water. The slope of the curve represents the rate of change of thermal energy in the
soil.
When a soil medium is completely frozen (i.e., unfrozen water content is zero), the
energy storage capability of the soil is represented by a constant volumetric heat
capacity of the frozen soil. Similarly, when a soil medium is completely thawed or
unfrozen (i.e., unfrozen water content is 1.0), the energy storage capability of the
soil is represented by a constant volumetric heat capacity of the unfrozen soil.
However, when a soil medium is undergoing phase change (i.e., unfrozen water
content is larger than zero, but smaller than 1.0), the slope of the unfrozen water
content function, the volumetric water content of the soil medium, and the latent
heat of water must be considered in defining the energy storage capability of the
soil.

Typical values of volumetric heat capacity


Table 4-1 provides typical values the specific and volumetric heat capacity of
various materials, (extracted from Johnston, Ladanyi, Morgenstern and
Penner, 1981 and Harlan and Nixon, 1978):

Page 79
Chapter 4: Material Properties TEMP/W

Table 4-1 Specific and volumetric heat capacities of various materials

Material Specific Heat Specific Heat Volumetric Volumetric


Capacity Capacity Heat Heat
(Btu/lb/F) (kJ/kg/C) Capacity Capacity
3
(Btu/ft /F) (kJ/m3/C)
Water 1.00 4.187 62.4 4187
Ice 0.50 2.094 28.1 1880
Air 0.24 1.0 0.0187 1.25
Soil minerals 0.17 0.71 28.0 1875
Organic soil minerals 0.40 1.674 37.5 2520
Extruded polystyrene 0.24 1.0 0.65 43.5
insulation
Concrete 0.21 0.895 30.0 2010
Asphalt 0.40 1.674 37.5 2520
Snow, fresh 3.11 209
Snow, drifted and 7.80 523.5
compacted
Granite 37.1 2490
Limestone 0.29 1.2 48.9 3285
Dolomite 0.21 0.88 37.4 2510
Sandstone 37.4 2510
Shale 27.4 1840
Glass 26.2 1760
Steel 0.11 0.46 56.0 3890
Wood 0.19 0.8 7.79 523

Estimating volumetric heat capacity for soils


The heat capacity of a material is defined as the quantity of heat required to raise
the temperature of the material by a unit degree. When expressed on a per unit
weight basis, this quantity of heat is referred to as the specific heat capacity; when
expressed on a unit volume basis, the quantity is known as the volumetric heat
capacity. The units for specific heat capacity are J/(kg.°C) and Btu/(lb.°F), and the
units for volumetric heat capacity are J/(m3.°C) and Btu/(ft3.°F).

Page 80
TEMP/W Chapter 4: Material Properties

TEMP/W uses the volumetric heat capacity in its formulation. The volumetric heat
capacity of a soil can be approximated by the dry density of the soil and the sum of
the specific heat capacities of its different constituents (namely, soil particles,
water, ice, and air). The air component is very small and generally is neglected. A
general equation for estimating unfrozen and frozen volumetric heat capacity is:
(see Johnston, Ladanyi, Morgenstern and Penner, 1981, pp. 73-147).

where:
C = volumetric heat capacity of the soil,
c = specific heat capacity of the soil,
γ = bulk density of the soil,
gd = dry density of the soil,
cs = specific heat capacity of soil particle,
cw = specific heat capacity of water,
ci = specific heat capacity of ice,
wu = unfrozen water content expressed in % of dry weight of the soil, and
wf = frozen water content expressed in % of dry weight of the soil.

Examples
The following examples illustrate how to estimate the volumetric heat capacity for
various soils:
1. For a dry mineral soil with dry density of 120 lb/ft3 or 1923 kg/m3, the
volumetric heat capacity of the soil can be estimated as:
C = 120 lb/ft3 * 0.17 Btu/(lb. °F)
= 20.4 Btu/(ft3. °F)
or in SI units as:
C = 1923 kg/m3 * 0.71 kJ/(kg. °C)
= 1365 kJ/(m3. °C)

Page 81
Chapter 4: Material Properties TEMP/W

2. For an unfrozen mineral soil with dry density of 120 lb/ft3 or 1923 kg/m3, and
unfrozen water content of 0.25, the volumetric heat capacity of the soil can be
estimated as:
C = 120 lb/ft3 * [0.17 Btu/(lb. °F) + 1.0 Btu/(lb. °F) * 0.25]
= 50.4 Btu/(ft3. °F)
or in SI units as:
C = 1923 kg/m3 * [0.71 kJ/(kg. °C) + 4.187 kJ/(kg. °C) * 0.25]
= 3378 kJ/(m3. °C)
3. For a frozen mineral soil with dry density of 120 lb/ft3 or 1923 kg/m3 and frozen
water content of 0.25, the volumetric heat capacity of the soil can be estimated as:
C = 120 lb/ft3 * [0.17 Btu/(lb. °F) + 0.5 Btu/(lb. °F) * 0.25]
= 35.4 Btu/(ft3. °F)
or in SI units as:
C = 1923 kg/m3 * [0.71 kJ/(kg. °C) + 2.094 kJ/(kg. °C) * 0.25]
= 2372 kJ/(m3. °C)
4. For a frozen organic soil with dry density of 80 lb/ft3 or 1282 kg/m3 and frozen
water content of 0.65, the volumetric heat capacity of the soil can be estimated as:
C = 80 lb/ft3 * [0.40 Btu/(lb. °F) + 0.5 Btu/(lb. °F) * 0.65]
= 58.0 Btu/(ft3. °F)
or in SI units as:
C = 1282 kg/m3 * [1.674 kJ/(kg. °C) + 2.094 kJ/(kg. °C) * 0.65]
= 3891 kJ/(m3. °C)
The estimation of the volumetric heat capacity is not as simple when a phase
change is involved. In such a case, the latent heat absorbed or released during the
ice-water phase transformation must be included in estimating the volumetric
water content. This involves the knowledge of the unfrozen water content function

Page 82
TEMP/W Chapter 4: Material Properties

of a soil. TEMP/W calculates the additional capacity due to phase change based on
the user-defined unfrozen water content function.

4.2 Thermal conductivity


The thermal conductivity function characterizes the ability of a soil medium to
transmit heat by conduction. For an unfrozen soil medium, the thermal
conductivity of the soil medium can be assumed constant without appreciable error
in most engineering applications. However, for frozen soil, the thermal
conductivity is highly dependent on the unfrozen water content of the soil medium,
which is also a function of temperature. Since the thermal conductivity of ice is
much higher than that of water, the thermal conductivity of a soil increases as the
soil freezes and decreases as the soil thaws.
The thermal conductivity can be expressed as a function of temperature, as
illustrated in Figure 4-3. TEMP/W uses this type of general function to
accommodate the variability in thermal conductivity.

Figure 4-3 A typical thermal conductivity function for soil

Page 83
Chapter 4: Material Properties TEMP/W

At temperatures above zero (0) °C, the thermal conductivity tends to increase with
temperature. However, the increase is small and can be ignored without significant
error in most engineering applications. (see Harlan and Nixon, 1978, pp. 103-163)
At temperatures below zero (0)°C, the thermal conductivity is usually greater than
that of an unfrozen soil, since the thermal conductivity of ice is much higher than
that of water. The thermal conductivity of frozen soil is highly dependent on the
unfrozen water content within the soil.
As the volumetric water content approaches 1 (or 100%), the thermal conductivity
of an unfrozen soil approaches that of water, while the thermal conductivity of a
frozen soil approaches that of ice. On the other hand, as the volumetric water
content approaches zero, the thermal conductivity of the soil approaches the
average thermal conductivity of its mineral constituents.
Furthermore, coarse-grained soil’s are commonly dominated by quartz, which has
a relatively large thermal conductivity; fine-grained soil’s are dominated by clayey
minerals, which generally have a substantially lower thermal conductivity.

Typical values of thermal conductivity


Table 4-2 provides typical values of thermal conductivity for various materials,
(extracted from Johnston, Ladanyi, Morgenstern, and Penner, 1981):

Page 84
TEMP/W Chapter 4: Material Properties

Table 4-2 Unit weight and thermal conductivity of various materials

Material Unit Weight Conductivity Conductivity


3
(lb/ft ) Btu/(hr.ft.°F) J/(sec.m.°C)
Water 62.4 0.35 0.605
Ice 57 1.29 2.23
Air (dry, still) 0.014 0.024
Snow
loose, new 0.05 0.086
on ground 0.07 0.121
dense, compacted 0.20 0.340
Soil and Rock Minerals
shale 0.9 1.5
evaporites 3.1 5.4
limestone 168 0.75-2.9 1.3-5.0
dolomite 178 2.9 5.0
sandstone 1.1-2.4 1.8-4.2
schist 0.90 1.6
gneiss 1.4 2.5
greenstone 1.9 3.3
slate 2.2 3.8
argillite 1.9 3.3
quartzite 2.6-4.1 4.5-7.1
granite 1.0-2.3 1.7-4.0
diabase 1.2 2.1
gabbro 1.4 2.5
grandiorite 1.5 2.6
Metal
steel 490 20-30 35-52
cast iron 29 50
aluminum 90-110 156-190
copper 223 386
Building Brick 0.40 0.69
Concrete

Page 85
Chapter 4: Material Properties TEMP/W

Material Unit Weight Conductivity Conductivity


3
(lb/ft ) Btu/(hr.ft.°F) J/(sec.m.°C)
with sand and gravel 140 0.75-1.0 1.3-1.7
aggregate
with lightweight 120 0.43 0.74
aggregate
with polystyrene beads 40 0.18-0.21 0.31-0.36
Concrete, asphalt 131-138 0.61-0.88 1.05-1.52
Dry Wood
fir, pine, similar softwoods 32 0.066 0.12
maple, oak, similar 45 0.094 0.16
hardwoods
Building Boards
asbestos-cement board 120 0.33 0.57
plywood 34 0.067 0.12
wood fibreboard 26-33 0.035-0.046 0.061-0.080
wood fibre hardboard type 65 0.12 0.21
Insulating Materials
asbestos fiber blanket 9-12 0.04 0.07
mineral wool blanket 1.5-4.0 0.022 0.038
wood fiber blanket 3.2-3.6 0.021 0.036
cellular glass board 7.0-9.5 0.032 0.055
corkboard 6-9 0.022 0.038
glass fiber board 9.5-11.0 0.021 0.036
wood or cane fiber board 15 0.029 0.050
polyurethane foam board 2.0 0.009-0.022 0.016-0.038
polystyrene board 1.3-3.0 0.019-0.022 0.033-0.038
compressed straw board 23 0.05 0.09
sulphur foam board 4-30 0.02-0.04 0.03-0.07
Loose Fill
asbestos fibers (bulk) 20-30 0.05-0.13 0.09-0.22
cork (granules) 5-12 0.02-0.03 0.035-0.052
expanded clay (powder) 37-41 0.07 0.12
expanded perlite (powder) 3-4 0.025 0.043

Page 86
TEMP/W Chapter 4: Material Properties

Material Unit Weight Conductivity Conductivity


3
(lb/ft ) Btu/(hr.ft.°F) J/(sec.m.°C)
glass fibers 2-12 0.02 0.035
glass, cellular (pellets) 6 0.03 0.05
mineral wool 2-5 0.025 0.043
sawdust or shavings 8-15 0.037 0.064
slag, foamed (granules) 38-45 0.07 0.12
straw fibers 7-8 0.025 0.043
vermiculite (expanded) 7.0-8.2 0.04 0.069
wood fibers 2.0-3.5 0.025 0.043

Estimating thermal conductivity for soils


Thermal conductivity is defined as the quantity of heat that will flow through a unit
area of a soil medium of unit thickness in unit time under a unit temperature
gradient. Thermal conductivity units commonly are J/(sec.m.°C) and Btu/(hr.ft.°F).
Various empirical or semi-empirical methods have been developed for estimating
the thermal conductivity of soil’s. These methods have been evaluated in detail by
Farouki, (1981), who concluded that different types of soil’s may require different
estimation methods. The method developed by Johansen, (1975), however, appears
to be the most general method. Since it is beyond the scope of this chapter to
present the various estimation methods, only Johansen’s method is presented in
this section.
For dry, natural soil’s, the thermal conductivity kdry can be estimated based on its
dry density using the following equation:

where the dry density γd is in kg/m3 and the unit weight of soil particles is taken as
2700 kg/m3.
For dry crushed rock materials, the thermal conductivity kdry can be estimated
based on its porosity n using the following equation:

Page 87
Chapter 4: Material Properties TEMP/W

For a saturated unfrozen soil, the thermal conductivity ksat is estimated based on the
thermal conductivities of its components and their respective volume fractions.

where, ks is the thermal conductivity of the soil particles, and kw is the thermal
conductivity of the pore-water.
For a saturated frozen soil containing some unfrozen water content, wu, the thermal
conductivity ksat becomes:

where ki is the thermal conductivity of ice.


For an unsaturated soil, the thermal conductivity kunsat is estimated based on its
saturated conductivity, dry conductivity and degree of saturation S using the
following equation:

where:
Ke = 0.7 LogS + 1.0 for unfrozen coarse grained soil,
Ke = LogS + 1.0 for unfrozen fine grained soil, and
Ke = S for frozen soil.

The above equations may be used for a rough, general estimation of the thermal
conductivity of a soil. It is your responsibility to ensure the applicability of the
above equations to the respective soils.

4.3 Sensitivity of results to thermal material properties and


water content in soil
In seepage analyses where the hydraulic conductivity values can change many
orders of magnitude a sensitivity study usually considers multiplying or dividing
values by a factor of 10 in order to see a change in output data. With thermal
properties, the change in values (be they thermal conductivity or heat storage)
generally occur within a 1/3 order of magnitude. In addition, if freezing is taking

Page 88
TEMP/W Chapter 4: Material Properties

place, the amount of unfrozen water that freezes and releases latent heat can be a
significant factor.
In this discussion, the attempt to calibrate a thermal model to real site data is used
for illustrative purposes. The geometry of the model was established and default
material properties were used in the initial simulations based on previous modeling
for the site. At iterative approach was then used to adjust material properties until a
reasonable agreement between computed and actual observed temperatures was
obtained. The calibrated model was needed for future site planning.
There are three primary material properties that can be adjusted. These are listed
below with an explanation of the affect they have on the computed temperature
trends. The comparison between actual measured and best fit calibrated data is
provided in Figure 4-4.

• Thermal Conductivity: Increasing this value increases the rate at


which the temperature drops to the freezing point and then after the
phase change has occurred. It is evident in the slope of the curves in
the figure at temperatures above freezing and at temperatures below
about –4 C when most of the pore-water is frozen.
• Volumetric Water Content: Adjusting the water content affects the
length of time the temperature profile “hovers” around the freezing
point before it starts to drop off rapidly again. This length of time is
directly related to the latent heat that is released so the more water
there is, the more latent heat, and the longer it takes to be removed.
• Slope of Unfrozen Water Content Function: This parameter controls
the rate at which the water is frozen at temperatures just below
freezing. If the function is steep, all the water freezes just below 0 C
and the temperature decay curve drops off quickly at a rate almost
equal to the rate before the freezing point was reached. If the function
is shallow and unfrozen water is still present at temperatures around –2
to –4 C, then the temperature decay curve shows a slower drop in the
temperature range just below 0 C. Steep functions are typical of very
coarse-grained material and shallow functions are typical of fine-
grained materials.

Page 89
Chapter 4: Material Properties TEMP/W

Test Freezing Measured and Calibrated Ground Temperatures


10
Notes:
Sensor # Position Unfrozen Cond (W/mC) Frozen Cond (W/mC) Vol. W/C Freeze Temp.
2 Top Zone 6.6 7.5 5% 0 to -2.5 C
5 3 Bot. Zone 4.4 4.9 20% -1 to -3.5 C
4 Below 6.6 7.5 7% -1 to -3.5 C

0
Temperature (C)

-5

-10

-15 Sensor-2 Field


Sensor-3 Field
Sensor-4 Field

-20 Sensor-2 Calibrated


Sensor-3 Calibrated
Sensor-4 Calibrated

-25
Dec-99 Feb-00 Apr-00 May-00 Jul-00 Sep-00
Date

Figure 4-4 Comparison between actual and calibrated ground


temperatures
When the model was initially run using assumed material properties the predicted
freezing times were much longer than those observed. In fact, the temperatures
were just starting to pass through the 0 C point by the end of the test freezing
period. The slopes of the decay curves from the initial temperature of +8 C to the 0
C point indicated that the material thermal conductivity was too low. Sensors 2 and
3 were assumed to be in a clay/ore type ground with very low thermal conductivity
and sensor 4 was assumed to be in a broken sandy type ground. The model was re-
run a few times with slight increases to the thermal conductivity on each
subsequent iteration. When the drop in temperature from +8 C to 0 C agreed
somewhat with the measured data the thermal conductivity was held constant.
Once the thermal conductivity was adjusted it was necessary to adjust the
volumetric water contents in order to match the time the temperature profile
“hovered” near the freezing point. The final calibrated water contents are likely
lower than one may expect, but even a slight increase in the calibrated values
resulted in several more weeks of freezing delay near the freezing point. The final
calibrated values are within the ranges originally reported by the project owner.

Page 90
TEMP/W Chapter 4: Material Properties

Finally, the shape of the unfrozen water content function was adjusted in order to
approximate the rate of decay of the temperatures after they have passed through
the freeze point. No extreme changes were made to the material properties
(compared to earlier modeling), but they were adjusted somewhat to approximate a
more fine-grained, than coarse-grained material.
In carrying out all of the above changes it was important to ensure that the final
combination of material properties for any given soil made sense for that soil. In
other words, care was taken to not have water contents typical of clay, combined
with a thermal conductivity obtained from a solid rock core sample, applied to a
soil with an unfrozen water content function found in beach sand.

4.4 Function data application in solver


TEMP/W uses spline interpolation techniques to create smooth, continuous
conductivity, and unfrozen water content functions. Once the spline fit is created
and adjusted to the user’s satisfaction, the solver is capable of looking up any “y”
value on the splined function for any input “x” value. A typical example would be
for the solver to obtain a water content value from the spline data for any value of
soil water pressure. In all GEO-SLOPE products, the spline fit you observe when
you set up the function is identical to the spline data used in the solver. What you
see is what you get, so it is important to view the spline function to ensure that it is
smooth, defined over a full range of “x” values and not discontinuous.

Weighted splines
Smooth curves were produced in the past using mechanical means. This involved
the use of a thin, flexible strip of wood or metal held in place with weights. The
flexible strip would bend in such a way that the internal energy due to bending was
at a minimum (see Lancaster and Salkauskas, 1986).
Such a curve can be described mathematically by defining the x-y coordinates of
the points (weights) and then computing the curvature at the points that minimizes
the internal energy term in the equations. Mathematically, this is referred to as a
natural spline (see Lancaster and Salkauskas, 1986).
A natural spline can have many undesirable humps and hollows when the data
points are not near the natural maximum curvature positions. Figure 4-5a illustrates
this behavior.

Page 91
Chapter 4: Material Properties TEMP/W

Salkauskas, 1974, and Lancaster and Salkauskas, 1986, have developed a


procedure for controlling the undesirable humps and hollows. They called the
resulting curve a weighted spline. Figure 4-5b illustrates the effect on the shape of
the curve using the weighted spline interpolation technique.
TEMP/W utilizes the weighted spline technique to create smooth conductivity,
unfrozen water content, boundary, and modifier functions. In addition to defining a
continuous and smooth function, spline interpolation also provides first derivatives
of the curve at any point. The first derivative of the unfrozen water content
function is the rate of release of phase change latent heat and is needed in the
solution of the transient heat transfer equation.

Figure 4-5 Natural and weighted splines

Page 92
TEMP/W Chapter 4: Material Properties

Best-fit splines
Since all functions are approximations of real-world behavior, it is often
convenient to use measured data values for the definition of a function. These
values, however, usually do not lie along a smooth, continuous curve. A spline
function that is fit to these values will appear jagged and will not accurately reflect
the measured data. To overcome this problem, TEMP/W allows you to define a
"best-fit" spline through the data points, as illustrated in the figures below.
The input function commands provide a means of controlling how the spline curve
is fit to the data points. For each function, you can assign a "Fit Curve to Data
percentage" and a "Curve Segments percentage" between 0% and 100%.
When the curve is fit exactly (100%) to the data points, the spline passes through
each data point. As the curve fitting is reduced, the spline shape approaches a
straight line that passes close to each data point. This is useful when you want to
approximate a spline through laboratory-measured data points without moving any
of the data points.
When the curve segments are curved (100%) between data points, the curve is
defined as a natural spline. As the curve segments are made straighter, the curve
segments approach a straight line between data points. Straightening the curve
segments helps to prevent "spline overshoots" (extreme peaks or valleys in the
spline). It also allows you to define "step" functions that have straight line
segments between each data point.
When specifying a best-fit percentage, it is best to experiment with different values
until you obtain a smooth spline that still passes close to the data points.

Page 93
Chapter 4: Material Properties TEMP/W

185

180

175
Conductivity

170

165

160

155
-10 -8 -6 -4 -2 0

Temperature

Figure 4-6 Spline fit to data using a 30% "fit" value

185

180

175
Conductivity

170

165

160

155
-10 -8 -6 -4 -2 0

Temperature

Figure 4-7 Spline fit to data using an exact “fit” value

Page 94
TEMP/W Chapter 4: Material Properties

4.5 Soil material database and default function estimation


TEMP/W has the ability to estimate the shape of the thermal conductivity function
given the user selected soil type and the user input frozen and unfrozen thermal
conductivity values. Figure 4-8 shows the built in unfrozen water content functions
(normalized to range between zero and 100% unfrozen) for various soil types. The
user can select from any of the default soil’s and the function will be generated.
Once generated in the program, the user can change the shape to any desired value.
The importance of using one of these default functions is that they have a well
defined shape with no sharp breaks in the curve. The actual water content used in
the solver is the percentage value obtained from this function multiplied by the
user input soil water content.
Figure 4-9 shows estimated thermal conductivity functions for the various default
soil types. The range of thermal conductivity in these functions as they will appear
in the software will depend on the user input values for frozen and unfrozen
thermal conductivity. The program will scale the curves to fall within the specified
range while keeping the negative temperature values fixed.

1.20

Clay
Silty Clay
1.00
Unfrozen Water Content
Silt
Silty Sand
Sand
0.80
Rock

0.60

0.40

0.20

0.00
-6 -5 -4 -3 -2 -1 0

Temperature

Figure 4-8 Default unfrozen water content functions for various soil
types

Page 95
Chapter 4: Material Properties TEMP/W

2.20

2.00

1.80

Conductivity
1.60

Clay
Silty Clay 1.40

Silt
Silty Sand
1.20
Sand
Rock

1.00
-6 -5 -4 -3 -2 -1 0

Temperature

Figure 4-9 Estimated thermal conductivity using Kf=2 and Ku=1 for
various soil’s
The importance of fixing the temperature values is that this function must agree
with the unfrozen water content function. The thermal conductivity increases as ice
replaces water in the soil pores so the increase in conductivity must coincide with a
decrease in water content at each temperature.

Page 96
TEMP/W Chapter 5: Boundary Conditions

5 Boundary Conditions

5.1 Introduction
Specifying conditions on the boundaries of a problem is one of the key
components of a numerical analysis. This is why these types of problems are often
referred to as “boundary-valued” problems. Being able to control the conditions on
the boundaries is also what makes numerical analyses so powerful.
Solutions to numerical problems are a direct response to the boundary conditions.
Without boundary conditions it is not possible to obtain a solution. The boundary
conditions are in essence the driving force. What causes heat to flow? It is the
temperature difference between two points or some specified rate of heat flow into
or out of the system. The solution is the response inside the problem domain to the
specified conditions on the boundary.
Sometimes specifying conditions is fairly straight forward, such as defining the
temperature or heat flux conditions that exist on a year round basis at some known
depth in the soil. Many times, however, specifying boundary conditions is complex
and requires some careful thought and planning. Sometimes, the correct boundary
conditions may even have to be determined through an iterative process since the
boundary conditions themselves are part of the solution, as for instance the heat
removal into a freeze pipe where the amount of heat removed depends on the
temperature difference between the fluid in the pipe and the soil. The resulting heat
removal changes the soil temperature which in turn changes the heat removal.
Furthermore, the conditions on the boundaries may change with time during a
transient analysis, which can also add to the complexity.
Due to the extreme importance of boundary conditions, it is essential to have a
thorough understanding of this aspect of numerical modeling in order to obtain
meaningful results. Most importantly, it is essential to have a clear understanding
of the physical significance of the various boundary condition types. Without a
good understanding it can sometimes be difficult to interpret the analysis results.
To assist the user with this aspect of an analysis, TEMP/W has tools which make it
possible to verify that the results match the specified conditions. In other words, do
the results reflect the specified or anticipated conditions on the boundary?
Verifying that this is the case is fundamental to confidence in the solution.
This Chapter is completely devoted to discussions on boundary conditions.
Included are explanations of some fundamentals, comments on techniques for

Page 97
Chapter 5: Boundary Conditions TEMP/W

applying boundary conditions and illustrations of boundary condition types


applicable for various conditions.

5.2 Fundamentals
All finite element equations just prior to solving for the unknowns untimely boil
down to:

[ K ]{ X } = { A}
where:
[K] = a matrix of coefficients related to geometry and materials properties,
{X} = a vector of unknowns which are often called the field variables, and
{A} = a vector of actions at the nodes.

For a thermal analysis the equation is:

Equation 5-1 [ K ]{T } = {Q}


where:
{T} = a vector of the temperature at the nodes, and
{Q} = a vector of the heat flow quantities at the node.

The prime objective is to solve for the primary unknowns which in a thermal
analysis are the temperatures at each node. The unknowns will be computed
relative to the T values specified at some nodes and/or the specified Q values at
some other nodes. Without specifying either T or Q at some nodes, a solution
cannot be obtained for the finite element equation. In a steady-state analysis, at
least one node in the entire mesh must have a specified T condition. The specified
T or Q values are the boundary conditions.
A very important point to note here is that boundary conditions can only be one of
two options. We can only specify either the T or the Q at a node. It is very useful
to keep this in mind when specifying boundary condition. You should always ask
yourself the question: “What do I know? Is it the T or the heat flow, Q?” Realizing
that it can be only one or the other and how these two variables fit into the basic

Page 98
TEMP/W Chapter 5: Boundary Conditions

finite element equation is a useful concept to keep in mind when you specify
boundary conditions.
As we will see later in this Chapter, flow across a boundary can also be specified
as a gradient or a rate per unit area. Such specified flow boundary conditions are
actually converted into nodal Q values. So, even when we specify a gradient, the
ultimate boundary condition options still are either T or Q.
Remember! When specifying thermal boundary conditions, you only have one of
two fundamental options – you can specify T or Q. These are the only options
available, but they can be applied in various ways.
Another very important concept you need to fully understand is that when you
specify T, the solution to the finite element Equation 5-1 will provide Q.
Alternatively, when you specify Q, the solution will provide T. The equation
always needs to be in balance. So when a T is specified at a node, the computed Q
is the Q that is required to maintain the specified T; you cannot control the Q as it
is computed. When Q is specified, the computed T is the T that is required to
maintain the specified flow Q.
Recognizing the relationship between a specified nodal value and the
corresponding computed value is useful when interpreting results. Let’s say you
know the specified flow across a surface boundary. Later when you check the
corresponding computed temperature at that node you may find that it is
unreasonably high or low. You would use your knowledge of the problem to assess
if the heat flow applied was reasonable. The T values are computed based on Q
and the soil properties so it must be one of three things. Knowing what to look for
helps you to judge whether that is reasonable or not.
TEMP/W always provides the corresponding alternative when conditions are
specified at a node. When T is specified, Q is provided, and when Q is specified, T
is provided. The computed Q values at nodes where a temperature is specified are
referred to as Boundary Flux values with units of heat per time (e.g. J/sec). These
Boundary Flux values are listed with all the other information provided at nodes.
A third important fundamental behavior that you need to fully understand is that
when neither T nor Q is specified at a node, then the computed Q is zero.
Physically, what it means is that the heat flow coming towards a node is the same
as the flow leaving the node. Another way to look at this is that no flow is entering
or leaving the system at these nodes. Heat leaves or enters the system only at nodes
where T, or a non-zero Q, has been specified. At all nodes for no specified

Page 99
Chapter 5: Boundary Conditions TEMP/W

condition, Q is always zero. This, as we will see later in this chapter, has important
implications when simulating such conditions as freeze pipes at a single node.
The temperatures in a thermal analysis are the primary unknowns or field
variables. A boundary condition that specifies the field variable (T) at a node is
sometimes referred to as a Type One or a Dirichlet boundary condition. Flow
gradient (flux) boundary conditions are often referred to as Type Two or Neumann
boundary conditions. You will encounter these alternate names in the literature, but
they are not used here. This document on thermal modeling simply refers to
boundary conditions as temperature (T) or heat flux (Q) boundary conditions. Later
we will differentiate between nodal flux Q and specified gradients (rates of flow
per unit area) across an element edge.

5.3 Temperature boundary conditions


Temperature boundary conditions are the easiest to understand. Unlike seepage
where the total head is a combination of pressure and elevation head, in thermal
modeling the desired boundary temperature is applied as a fixed (or time varying)
number, and it is used as is in the solution of the equations.
TEMP/W will allows a temperature boundary condition to be applied in various
ways: as a fixed value; as a time varying function; or even as a value that is
modified based on the existing ground temperature, as would be the case for
modifying an air temperature versus time function in order to apply a ground
surface temperature boundary condition. This latter example is discussed in more
detail later in this chapter.
One thing to keep in mind when applying a temperature boundary condition in a
transient analysis is that the applied temperature should agree with the initial
condition temperature at the node or else a large heat flux will be generated to
transition the temperature from the initial condition value to the new boundary
value. Initial conditions are discussed in more detail in the Numerical Issues
chapter.

5.4 Heat flow boundary conditions


The second of two options for specifying boundary conditions is to specify a rate
of heat flow across the edge of an element. An example may be the rate of heat
flow from a pipe or across the ground surface. This is often referred to as a flux
boundary.

Page 100
TEMP/W Chapter 5: Boundary Conditions

Thinking back to the discussion on fundamentals in Section 5.2, when it comes to


solving the unknowns in the finite element equations, it is necessary to specify or
compute the flow at the nodes. Consequently, when specifying a unit rate of flow
across the edge of an element, it is necessary to integrate along the edge of the
element and convert the unit rate of heat flow (q) into nodal total heat flows (Q).
Even though TEMP/W automatically does the integration, it is nonetheless useful
to have an understanding of how the integration is done. Understanding the
relationship between unit rates of flow across an element edge and nodal flows can
be useful in understanding and verifying results. Sometimes in unusual situations,
it can also be useful to know how to manually compute the nodal Q flows from
unit rates of flow (q) and then specify Q at the node instead of q on the edge of the
element.
First of all, TEMP/W only handles uniform rates of flow across the edge of an
element. The uniform rate of flow can be set to vary with time, but not with
distance across an edge of any individual element. Since the flow across the edge
of an element is uniform, the total flow across the edge is simply the flow rate (q)
times the length of the element edge. The manner in which the total flow across the
edge gets distributed to the nodes depends on the number of nodes on the edge. In
TEMP/W the edge can have either two or three nodes.
When the edge of an element has only two nodes, the total specified flow across
the edge is evenly divided between the end nodes. Half goes to one end node the
other half goes to the other end node. This is graphically illustrated in Figure 5-1.
When two or more edges have a common node, the contributions from each edge
accumulate at the common node as illustrated in Figure 5-1. These illustrations
refer to seepage quantities, but the same rates and concepts can apply to thermal
models.
Higher-order elements with three nodes along the edge of an element are usually
not required in a thermal analysis. Sometimes, however, it is desirable to do a
SIGMA/W or QUAKE/W analysis on the same mesh and for these types of
analyses, higher order elements are required. Consequently, higher order elements
are used in a thermal and seepage analysis not because they improve the results,
but because mesh consistency is required when integrating with other types of
analyses.

Page 101
Chapter 5: Boundary Conditions TEMP/W

Seepage boundary with


water table initially at s urface
1.00

0.75
2.0000e+000
0.50

0.25
5.0

01
1.0000e+000
0

-0
00

0e
0.00
e

00
-00

5.0
1

q = 1 m ^3 / tim e / meter edge length


unit flux boundary

Figure 5-1 Unit and total nodal flux boundary condition relationships
When there are three nodes along the edge of an element, the total flow across the
edge is distributed as 1/6 to each corner node and 4/6 to the middle node. This is
graphically illustrated in Figure 5-2. Once again, when one node is common to
more than one edge the 1/6 end contribution from each edge is accumulated at the
common node. The distributions of the nodal fluxes come directly from the
numerical integration of a variable, such as a gradient along the edge of an
element. Further details on this are presented later in the Theory chapter and the
Visualization of Results chapter.

Seepage boundary with


water table initially at s urface
1.00

0.75
2.0000e+000
0.50

Nodal flux: 0.66667


0.25
1.6

1
0

3.3333e-001
6

-0
67

7e

0.00
e

66
-00

1.6
1

q = 1 m ^3 / tim e / meter edge length


unit flux boundary

Figure 5-2 Distribution of unit flux in element with secondary nodes

Page 102
TEMP/W Chapter 5: Boundary Conditions

For an axisymmetric case the Q boundary flux values are a function of the radial
distance from the vertical symmetric axis. For elements with nodes located only at
the corners, as shown in Figure 5-3, the nodal boundary flux Q is the contributing
area from each element times the radial distance from the symmetric axis times the
element thickness which is expressed in terms of radians. The nodes further away
from the axis of symmetry have more contributing area so their flux values are
greater. In the case of the middle node, it has a total flux value of 1.0, but this is
comprised of two flux quantities of 0.5, each contributed by the element edges on
either side of it. The node on the right has a total flux of 0.833333, but only half
the contributing area as the middle node. As the distance from the axis of
symmetry increases, so does the nodal flux quantity.

Seepage boundary with


water table initially at surface
1.00

0.75
2.0000e+000
0.50

0.25
1.6

1
1.0000e+000 00
66

e-
7e

0.00
33
-0

33
01

8.

q = 1 m^3 / time / meter edge length


unit flux boundary

Figure 5-3 Distribution of unit flux in axisymmetric analysis without


secondary nodes
The situation is not quite as straight forward for higher order elements where three
nodes exist along the element edge as illustrated in Figure 5-4. The detailed
formulas for these cases are provided in the Theory Chapter. The case described in
Figure 5-3 is fairly straightforward and easy to remember for quick spot checking
the computed nodal boundary flux values.
Conversion of specified flow rates (gradients) across element edges is dependent
on the specified element thickness. For a two-dimensional vertical section the
thickness by default is one unit, although the thickness can be some other specified
value. For axisymmetric cases, the thickness is in terms of radians. The default is 1
radian. Sometimes, it is useful and convenient to specify the thickness as 2*Pi
radian so that the computed nodal boundary flux values are for the entire

Page 103
Chapter 5: Boundary Conditions TEMP/W

circumferential area. Specifying the thickness is simply a user preference. It is


important, however, to remember the element thickness when interpreting the
computed Q values.

Seepage boundary with


water table initially at surface
1.00

0.75
2.0000e+000
0.50

Nodal flux: 0.3333 Nodal flux: 1.0000


0.25
1.3

1
00
3.3333e-001
78

e-
5e

0.00

33
-

33
01

3.
q = 1 m^3 / time / meter edge length
5

unit flux boundary

Figure 5-4 Distribution of unit flux in axisymmetric analysis with


secondary nodes
In TEMP/W, the specified in or out flow rates (gradients) are always deemed to be
normal to the element edges. In addition, flow into the system is positive, while
flow out of the system is negative.
Another important concept to keep in mind when specifying and interpreting flow
quantities is that nodal boundary flux values Q are scalar values. Nodal Q values
have no specified or computed direction. The direction of flow can only be inferred
from the computed gradients or unit flow vectors inside the elements. TEMP/W
can display heat flow vectors in each element that are a graphical representation of
flow rates and direction.

5.5 Sources and sinks


There is another type of boundary condition called a source or a sink. These
boundary conditions are sometimes referred to as a Type 3 boundary condition. A
typical sink might represent a drain at some point inside a mesh. A source could be
hot buried pipe while a sink may be a buried freeze pipe or thermosyphon pipe.
The important concept about sinks and sources is that they represent flow into or
out of the system.

Page 104
TEMP/W Chapter 5: Boundary Conditions

In TEMP/W, flux boundary conditions can be applied along outside edges or nodes
of the mesh or along inside edges or nodes. There is no difference in how the
equations are solved in either case. The only thing to watch for is that you have an
understanding of what the area the heat flows across is.
There are a couple special types of heat sink boundary conditions that have been
formulated directly in TEMP/W: the convective heat transfer boundary, and the
thermosyphon boundary. These are fancy ways of setting up and applying a simple
flux boundary sink at a node or group of nodes. They are discussed in more detail
below.

5.6 Ground surface energy balance


In nature, the ground surface boundary condition is the net heat flux arising from
absorbed solar and long wave radiation and from sensible and latent heat transfer
between the ground surface and the overlying air. Schematically, this net flux can
be expressed by the energy balance equation as follows: (Goodrich and
Gold, 1981)

where:
QG = ground surface heat flux,
k = thermal conductivity at the ground surface,

= heat gradient normal to and evaluated at the ground surface,


QSW = net flux of solar radiation,
QLW = net flux of long wave radiation,
QH = net flux of sensible heat, and
QE = flux of latent heat associated with evaporation of moisture from the
surface

Page 105
Chapter 5: Boundary Conditions TEMP/W

5.7 Rigorous approach to calculation of ground surface


temperatures
Because the individual heat flux components cannot be measured or estimated with
great precision, large errors in the determination of QG are possible. Recently
however, iterative techniques have been developed that can use readily available
climate data so that a prediction of the ground surface temperature on bare soil or
beneath a snow pack can be estimated.
In order to use actual climate data for computing the ground surface temperature
you must gather and format the necessary data so that it can be pasted into the
Climate Data Set of the DEFINE program. There is also an option to import
existing data from one or more of the 40+ data sets for various global sites. More
than one data set can be applied in a single simulation if there are variations in
topography which result in different boundary conditions along the surface.
The units of your problem geometry will determine the units of the weather data
you must enter. The table below specifies the type of daily weather you must
compile along with the appropriate units for metric and non-metric geometries.
The distance or time units specified below do not have to agree exactly with your
model units. However, the distinction must be made between imperial and metric
units. The solvers will covert the climate data to the necessary format to match
your problem geometry.

Unit Type Max Daily Min. Daily Max. Daily Min. Daily Average Total Daily
Air Temp. Air Temp. Relative Relative Daily Wind Precip. (or
Humidity Humidity Speed snow water
equivalent)

Metric C C % (0-100) % (0-100) m/s mm


Imperial F F % (0-100) % (0-100) ft/s inches

TEMP/W requires that you enter a complete table of weather data for each day of
the simulation. Because of the diurnal nature of all weather data, you should also
enter time steps in the transient analysis so that the maximum time step is in the
order of 1/8 to 1/4 a day. The use of larger time steps will result in poor averaging
of daily climate values and may result in the application of unrealistically
high temperature values over the entire day. Remember also that the thermal
property "time" units must match the user-specified time step units.
Due to possible extreme changes in climate data from day to day or within any
given day, the user-specified time steps may not be small enough to handle a large
change in temperature. An adaptive time stepping scheme has been added to the

Page 106
TEMP/W Chapter 5: Boundary Conditions

program that will insert extra time steps if necessary. Extra time steps will be
added if the percent change in temperature from one time step to the next exceeds
the user-specified "tolerance" value as entered in the Key In Time Increments box.
Adaptive time stepping should be used in addition to appropriate user established
time steps. You should not specify one day time steps without adaptive stepping,
and hope the solver follows the diurnal nature of climate data within the one large
daily time step. More information on adaptive time stepping is provided later in
this chapter.

No snow present
Temperatures within the soil profile are required for the solution of the moisture
and heat flow equations. The surface temperature may be estimated (for conditions
where no snow pack is present) with the following relationship (Wilson, 1990):

1
Ts = Ta + (Q − E )
υ f (u )
where:
Ts = temperature at the soil surface (oC),
Ta = temperature of the air above the soil surface (oC),
υ = psychrometric constant,
Q = net radiant energy (minus transpiration energy) available at the surface
(mm/day), and
AE = actual vertical evaporative flux (mm/day),

This equation is basically stating that the ground temperature is increased by


energy that was not consumed in the evaporation or transpiration process. When
calculating the actual evaporation component of the above equation, TEMP/W
assumes that the soil is about 80% saturated and that the actual evaporation rate
will approach the potential evaporation rate. The VADOSE/W model computes
actual evaporation and ground temperature based on actual soil moisture stress
conditions; however, TEMP/W does not consider the water stress state.

Page 107
Chapter 5: Boundary Conditions TEMP/W

With snow pack


If there is snow cover, then snow accumulation / melt and ground temperatures are
based on an energy balance approach adopted from various methods presented in
the following references: SNTHERM89 model, Bras (1990), Liang et al (1994),
and Flerchinger (1989).
Full details of how GeoStudio determines snow accumulation and melt, as well as
ground temperatures will not be given here, as there are many physical and
empirical equations and an entire book could be written on the subject. In general,
the snow routine in GeoStudio should balance user supplied precipitation data with
applied snow melt during the freshet. The computed snow depths and ground
temperatures are affected by many variables, consequently, exact correlation with
field data may be hard to obtain. However, the intent of the snow routine in
GeoStudio is not to predict snow drifting or frost heaving. The intent is to enable
year round analysis that does not force the user to neglect what happens during
winter months or during snow melt when it comes to applying a temperature
boundary at the surface.
When the climate routine is called, the following activities occur:
The relative humidity, air temperature and precipitation are computed at the
elapsed time of the day. If a sinusoidal climate distribution option is selected, the
air temperature is assumed to be at its minimum value at sunrise and at its
maximum value at 12:00 pm (noon). The relative humidity is assumed to be at its
maximum at sunrise and its minimum at 12:00 pm. Both the air temperature and
relative humidity are distributed sinusoidally between their minimum and
maximum times. The daily precipitation is assumed to follow a sinusoidal
distribution over your specified hours of the day, with the peak rainfall (or
snowfall) occurring at the mid-point of the specified interval.
If you select to have climate data applied as a constant value over the day then the
temperature and relative humidity are fixed for the day and the amount of
precipitation will be the daily total, divided by the duration of the day, multiplied
by the current solver time step. If you select to have climate data be averaged from
day to day, then the previous days and next day’s average values are assumed to
occur at mid day and then a function is created to look up any value based on the
elapsed time since the previous mid day.
If the there is precipitation and the air temperature is cold enough, then the albedo
of the ground surface is set to be that for fresh snow. If there is no new snow or

Page 108
TEMP/W Chapter 5: Boundary Conditions

there is old snow, the albedo is computed accordingly based on techniques in


the references cited above.
The net radiation for the time of day is computed using the albedo and other
climate parameters as discussed in the references. The net radiation is assumed to
be at its peak value at 12:00 pm and is assumed to be zero prior to sunrise or after
sunset. It is possible to have a maximum radiation value that is negative. This
occurs if the incoming long and short wave radiation is less than the reflected and
surface emitted radiation.
Using the data calculated above, the snow accumulation or melt routine is called.
The heat storage capacity of any existing snowpack and any fresh snow is
determined.
The snow surface temperature is assumed to be at the freezing point and then an
energy balance is carried out to determine if there is a net surplus of energy or an
energy deficit. Ground heat flux, sensible heat flux, and latent heat flux are
included in the energy balance, as is the ground surface temperature at the last time
step.
If the net energy balance across the snow is positive, then some snow is melted to
balance the equation. If the net energy is negative, then an iterative scheme is
called to compute the correct snow surface temperature to balance the equation. If
the energy balance returns zero, then the assumed ground surface temperature is
correct.
Once the snow surface temperature is calculated, the snow water equivalent and
snow heat storage capacities are updated.
Using the new snow parameters, the snow density is updated as is the snow depth.
The snow density of the main snowpack is assumed to be a function of time and is
set to increase as the snow season increases or as the duration since the last snow
fall increases. The maximum snow density is assumed as 500 kg per cubic meter.
The snow depth is a function of the snow density and the snow water equivalent.
Once the snow layer parameters are established, the temperature beneath the
snowpack is computed by using the fixed snow surface temperature and iteratively
solving for a new ground surface temperature such that the ground heat flux into
the bottom layer of the snowpack will balance with the heat storage capacity of the
snowpack.

Page 109
Chapter 5: Boundary Conditions TEMP/W

5.8 Traditional approach to computing ground temperatures


TEMP/W enables you to choose from a detailed climate data boundary condition
option or an empirical approach that correlates ground temperatures with air
temperatures. The empirical approach is discussed next.

Empirical approach to estimating ground surface temperature


In practice, ground surface boundary conditions are often correlated with the air
temperature using an empirically determined coefficient called the “n-factor”. The
n-factor (nf and nt) is defined as the ratio of the surface freezing or thawing index
(Isf and Ist) to the air freezing or thawing index (Iaf and Iat) as represented in the
following equations:

The magnitude of freezing and thawing n-factors depends on the climatic


conditions, as well as on the type of ground surface. Approximate values of the n-
factor for several different surfaces have been suggested in the literature and are
given in Table 5-1, (Goodrich and Gold, 1981).
In TEMP/W, the boundary conditions at the ground surface can be simulated with
a combination of a temperature function and a modifier function. The temperature
function describes the seasonal variation of the air temperature, while the modifier
function describes the variation of the n factor as a function of the ground
temperature at the surface. Example functions are presented in Figure 5-5 and
Figure 5-6.

Page 110
TEMP/W Chapter 5: Boundary Conditions

Table 5-1 N-factor values for various surfaces

Surface Type Freezing Thawing


(nf) (nt)
Soil surface - spruce trees, brush, 0.29 (under snow) 0.37
moss over peat
Soil surface - brush, moss over peat 0.25 (under snow) 0.73
Turf 0.5 (under snow) 1.0
Snow 1.0 ———
Gravel 0.6 - 1.0 1.3 - 2
probable range for northern 0.9 - 0.95
conditions
Asphalt pavement 0.29 - 1.0 or greater 1.4 - 2.3
probable range for northern 0.9 - 0.95
conditions
Concrete pavement 0.25 - 0.95 1.3 - 2.1
probable range for northern 0.7 - 0.9
conditions

In TEMP/W, at the beginning of a new time step, the air temperature (Tair) at the
elapsed time is obtained from the temperature function, and the n-factor is obtained
from the modifier function based on the computed ground surface temperature
from the previous time step. The phase change temperature (Tphase) is specified
by you. The ground surface temperature (Tground) at the new time step is
calculated as follows:
Tground = (n - factor) (Tair - Tphase) + Tphase

When no modifier function is specified, TEMP/W assumes the n-factor value to be


1.0 at all temperatures.

Page 111
Chapter 5: Boundary Conditions TEMP/W

Figure 5-5 Example of T vs time boundary function

Figure 5-6 Example of modifier function for simulating n-factors

Page 112
TEMP/W Chapter 5: Boundary Conditions

5.9 Convective heat transfer surfaces


TEMP/W can model convective heat transfer boundary conditions for transient
analysis where the boundary is either an element edge or one or more individual
nodes. Convective heat transfer is the process that removes heat from a surface
when that surface is exposed to fluid (liquid or gas) of a different temperature
flowing over it. The general rate equation for this type of heat transfer is:

where:
Q = the total heat flux,
h = the convective heat transfer coefficient,
A = the area in contact with the flowing fluid,
Tg = the surface temperature, and
Tfluid = the bulk temperature of the flowing fluid.

The units for convective heat transfer coefficient will depend on the user-specified
heat and time units of the problem, but are in general heat/length/temperature (e.g.
W/m/C).

Artificial ground freezing application


This type of boundary condition is particularly useful in TEMP/W for determining
the ground temperature adjacent to a buried pipe - whether it is a heated pipe or a
freeze pipe used for artificial ground freezing. Traditionally, ground freezing
analysis has made an assumption about the actual temperature of the outside of the
freeze pipe over time, and often the assumption is that the wall temperature is
equal to or very close to the brine temperature inside the pipe. This is a very poor
assumption, since if the two temperatures were close, then the amount of heat
transfer would be quite low due to the small thermal gradient. A more accurate
approach is to specify the heat transfer characteristics and let the model determine
the heat removal rates and changing ground temperature.
In a typical chilled brine ground freezing application the flow of brine inside the
pipe is laminar, the Nusselt number is therefore constant at a value of 4.1, and the
convective heat transfer coefficient can be computed from:

Page 113
Chapter 5: Boundary Conditions TEMP/W

where:
Nu = the Nusselt number,
k = is the thermal conductivity of the brine, and
Dh = is the hydraulic diameter (equal to diameter of outer pipe minus diameter
of inner pipe).

For other applications, the convective heat transfer coefficient should be


determined by consulting an appropriate heat transfer text book or by searching for
methods to calculate it on the internet. The convective heat transfer coefficient is a
function of fluid temperature, fluid velocity, mass flow rate, and thermal
conductivity. Its value depends on whether the flow is laminar, turbulent or in
transition. It is calculated based on knowledge of the Reynolds’ number, the
Nusselt number, and the Prantl number. It matters if the flow is forced convection
or free convection.
The convective heat transfer for brine freezing applications typically ranges
between 25 and 75 W/m2 C.
As discussed in the Chapter on meshing, if the freeze pipe diameter is known, then
the actual boundary condition for the freeze pipe can be applied to a single node.
The heat removed at that node will be a function of the pipe surface area, the heat
transfer coefficient, and the difference in temperature between the brine and the
ground. When this type of boundary condition is set up in TEMP/W, then you
must enter the fluid temperature (as a constant value of function of time), the pipe
perimeter length, and the convective heat transfer value. If the pipe is to be located
at a mesh corner or edge node, the exposed pipe perimeter would be 1/2 or 1/4 of
the whole pipe perimeter (i.e., 1/4 * 3.1416 * diameter).
The following is an example of a freeze pipe boundary condition applied to a shaft
freezing project. Each individual pipe is represented by a big Q convective surface
boundary condition, as illustrated by the blue triangles at various nodes in Figure
5-7. Notice that there are two defined convective surface boundaries in this mesh
depending on whether the nodes are internal or along the edge. Obviously, the pipe
perimeter at the edge nodes is 1/2 of those at internal mesh nodes.

Page 114
TEMP/W Chapter 5: Boundary Conditions

Figure 5-7 shows the temperature contours after two years of freezing. The starting
temperature of the ground in this case was +25 C and the phase change
temperature of the ground water was set at -11 C due to the high sodium chloride
content in the water. This is the reason the phase change (blue) contour is not at the
zero degree Celsius line.

Figure 5-7 Ground freezing of a mine shaft to be excavated


Figure 5-8 shows the solved pipe surface temperature versus time (days) for the
problem above. Observe that there is a smooth temperature decay at the pipe. Also
observe that the pipe surface temperature is gradually approaching the brine fluid
temperature (in this case -25 C). It is important to note that the pipe surface
temperature will never equal the brine fluid temperature for if it did, there would
be no temperature gradient across the surface and the heat flow would equal zero.

Page 115
Chapter 5: Boundary Conditions TEMP/W

If you do not want to model the pipe as a single node, there is the option to specify
the boundary condition along element edges and the model will compute the area
across which the heat flows.

Figure 5-8 Computed pipe surface temperature for ground freezing


model

Pyrex tile cooling application


The convective heat transfer boundary condition type is not only suited to buried
hot or cold pipes. Consider the example in Figure 5-9 where a hot Pyrex tile 10mm
thick is rapidly cooled from 140 degrees to 25 degrees after coming out of the
oven. In this example, the model was set up to compare the cooling rates with
either still air over the tile or air moving at 10 m/s. The only difference in the
applied boundary condition is the computed heat transfer coefficient for the still or
moving air.

Page 116
TEMP/W Chapter 5: Boundary Conditions

In this example, the convective surface is not a single node, but the entire top and
right edges of the tile. TEMP/W will compute the distances between nodes along
these edges and apply the appropriate Q heat flux in the solution of the equations.
Pyrex Tile
140C initial temperature
Insulated on bottom
K = 1.4W/mC, Cp=835J/kgC, Density 2225 kg/m3
20

18

16 Air, 10m/s, 25C


Convective heat transfer coefficient = 14 W/mC (still air)
Distance(m) (x 0.001)

14 Convective heat transfer coefficient = 55.7 W/mC (moving air)

12

10

6
Tile

2
Insulation material
0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40

Distance(m) (x 0.001)

Figure 5-9 Convective cooling of a hot Pyrex tile after fabrication

Time steps for convective surface analysis


Depending on the starting temperature difference between the ground surface and
the fluid, it may be necessary to specify very small starting time steps in order to
get the problem to converge properly. If a brine temperature was -30 C and the
starting ground temperature was +20 C, then it would not be uncommon for the
first time step in the analysis to be in the order of 1/2 an hour (or its equivalent in
whichever time units you have specified). If, after the analysis is complete and the
temperature at the pipe surface is plotted versus time, you notice a large jump or
drop in the temperature at a given time step, then the analysis will have to be re-
solved with small time steps. Likewise, if you notice that the ground temperature
increases slightly above its starting value when it is supposed to be dropping, then
the first time step is likely too small and should be increased somewhat.

Page 117
Chapter 5: Boundary Conditions TEMP/W

5.10 Thermosyphon boundary conditions


TEMP/W has the ability to let you apply a thermosyphon boundary condition at
individual nodes. Each node can represent a single thermosyphon as if you were
looking at the end of the pipe. A thermosyphon cannot be represented along the
edge of an element because it is a point source (per meter depth) and not a line
source (which would make it a plane into the page).
In general, a thermosyphon is comprised of a sealed pipe which may contain a gas,
such as Carbon Dioxide. When the gas is warmed, it has a lower density and it will
rise to the top of the pipe. If the affect of the cold air blowing across a heat
dissipating radiator mounted to the top of the pipe cools the gas, then it will
condense and flow down the inside of the pipe wall. If the ground is warm enough
at the bottom of the pipe, it will add heat to the condensed fluid, which will cause
it to evaporate, become less dense, and rise to the top. The cycle repeats itself as
long as the air temperature is less than the ground temperature.

Figure 5-10 Schematic of thermosyphon operation

Page 118
TEMP/W Chapter 5: Boundary Conditions

Figure 5-10 is a schematic of a thermosyphon installed in the ground. Heat flows


into the thermosyphon at the base and is removed by the combination of wind
speed, air temperature, and radiator surface area at the top. There are several
different heat transfer processes occurring in this cycle. There is conduction and
convection at the base of the pipe, convection along the length of the pipe, and
convection between the surface radiator and the wind. These individual processes
can be combined into an overall heat transfer process with the following rate
equation:

where:
P = the performance characteristic of the thermosyphon,
Tg = the ground temperature, and
Tair = the air temperature.

The key to using the thermosyphon boundary condition option is determining the
performance characteristic, P, in the above equation.
Experimentation carried out by Artic Foundations Inc. (personal communication)
has lead to the following relationship for determining P as a function of wind
speed and radiator surface area. This equation is valid for thermosyphons installed
at angles of inclination greater than five degrees (i.e., five degrees from horizontal
to vertical).

where:
P = the performance for the entire length of pipe in W / oC,
wind = the wind speed in m/s, and
area = the radiator surface area in m2.

This rate equation gives the maximum heat extraction capability of the
thermosyphon over its entire length. Since TEMP/W only models thermosyphons
per unit length, this value must be divided by the actual constructed length of pipe
in the field.

Page 119
Chapter 5: Boundary Conditions TEMP/W

Figure 5-11 is the thermosyphon data entry dialogue for TEMP/W. You must enter
the actual length of pipe in the units of your mesh geometry. You must also enter a
maximum functioning air temperature and the minimum temperature difference
between ground and air that is required for continuous operation of the
thermosyphon.
In terms of heat extraction performance (P), you can enter your own fixed value or
you can have TEMP/W compute the value based on the performance equation
above and on the radiator size. If you specify a fixed value, the only unknown in
the above rate equation is the air temperature and this is obtained from the climate
data set that you must enter. Your fixed value must be in units of heat per time per
degree temperature. If you use the radiator size option, then the rate equation will
become a function of air temperature and wind speed, as well as the actual surface
area of the radiator.

Figure 5-11 Data required for thermosyphon boundary

Page 120
TEMP/W Chapter 5: Boundary Conditions

5.11 Boundary functions


TEMP/W is formulated to accommodate a very wide range of boundary
conditions. In a steady-state analysis all of the boundary conditions are either fixed
temperatures or fixed flux values. In a transient analysis however, the boundary
conditions can also be functions of time, or a response to flow amounts exiting or
entering the flow regime. TEMP/W accommodates a series of different boundary
functions. Each one is discussed in this section.

General
A non-fixed boundary condition must be entered as a user-defined function. In
TEMP/W, all functions are defined using a combination of manually entered or cut
and paste data points and all functions can be customized to suite your exact needs.
In certain cases, it may be desirable to have a stepped function. The additional
functionality of automatically fitting the data with a step function has been added
to the program. There is also an option to have a cyclic function repeat itself over
time which saves you the task of defining it repeatedly.
In general, all functions are comprised of a series of x and y data points that are fit
by a spline curve. A spline curve is a mathematical trick to fit a curved shape
between a series of points. The simplest way to fit a series of data points is to draw
a straight line between the points. This is often a very poor way to represent a non-
linear function. The advantage of the spline is that is joins all data points with a
continuous smooth curve.
During the solution process, the solver uses the “y” value along the spline curve for
any required “x” value. It is therefore important to make sure the spline fit, not
your original data, portrays how you want the boundary condition applied.

Temperature versus time


A very useful boundary function is user-specified Temperature versus Time.
Consider the case of a ground freezing simulation where the very cold brine
temperature is introduced into the much warmer ground. Applying a sudden
change in boundary temperature can be numerically difficult to solve, but also
physically incorrect. In reality, when the refrigeration plant is turned on, it takes a
few days for the brine to cool to its operating temperature. This fact can be applied
in the model as illustrated in Figure 5-12, where it takes about 12 days for the brine
to reach its minimum operating temperature. This function can be used directly at a

Page 121
Chapter 5: Boundary Conditions TEMP/W

boundary node or in combination with the convective flux boundary option as


discussed above.

10

0
T

-10

-20

-30
0 2 4 6 8 10 12

Time

Figure 5-12 Temperature versus time brine cool down function


There are occasions where the boundary function may cycle, or repeat itself over
time. This may be the case if a typical year of air temperature data were to be used
in a single multi-year analysis. Instead of defining the cyclic nature of the function
from start to finish, you have the option to define it over the first “period” of time
and then have it repeat itself for as long as you defined time steps to solve.
In any “time” function, make sure the units of time you specify in the function
match those used to define the conductivity function (e.g. seconds, hours, days
etc.) For long transient solutions, it is often useful to define the conductivity in
terms of kJ/m/C/day so that all time steps and functions can be set up with “day”
units. This avoids having to input time values of 864000 seconds to represent 10
days.

Nodal and unit heat flux versus time (Q and q)


As discussed previously, there are only two fundamental types of boundary
conditions: either a temperature or flux boundary. As with temperature boundaries,

Page 122
TEMP/W Chapter 5: Boundary Conditions

flux boundaries can be applied as fixed values of total or unit flux or a time-
dependent function of either.
In many cases it is useful to apply a total nodal flux (Q) boundary that is a function
of time. If the function “y” value is positive, this indicates a heat source at the node
and if the “y” value is negative, this would indicate a nodal heat sink flux. When
you apply a total nodal flux function at a node, you are not taking into account any
mesh geometry in the flux quantity. You are simply stating that you are
introducing a set amount of heat flow at a single point in the mesh, and the amount
that you are introducing changes with time.
During the solver process, the solver will take the elapsed time of the solution and
look up from the defined function what the total flux rate of heat should be at that
time. It will then multiply the function value by the current time step duration to
obtain a total amount of heat to add at that time step. Be careful when using
functions like this. It is very important to make sure the time steps you set up in the
solver are small enough to follow the desired shape of the function. Using too large
a time step will result in the look up heat rate value from the function being applied
over a longer time period, which will mean too much heat is applied as a boundary.

Modifier functions
The previous two types of functions allow you to apply a set total or unit flux at
element edges or mesh nodes. They are functions that will let the solver apply the
appropriate action value for the stated elapsed time.
A modifier function is a powerful way of having the applied flux or applied
temperature be dependent on the current temperature in the soil. Consider the
process of convective heat transfer where the heat removed from the ground
depends on the fluid temperature flowing over the ground and the ground
temperature itself (the difference in these two temperatures establishes the thermal
gradient causing heat to flow from the ground to the passing fluid). A modifier
function can be set up which establishes what the heat flow should be based on the
changing ground temperature. This modifier function is then multiplied by a Q or q
boundary function containing a value of -1 for all time. The -1 value simply forces
the modifier function to become a sink value that will remove heat from the
ground. Consider the function illustrated in Figure 5-13. This says that if the
ground temperature is +10, the heat removed from the soil should be 3000*(Q=-
1)=-3000. As the ground temperature drops due to this heat removal, the future
heat removal drops. Thus, if the ground temperature gets to be -10, the heat
removal would be 1000*(Q=-1)=-1000. The Q=-1 notation is the main boundary

Page 123
Chapter 5: Boundary Conditions TEMP/W

function that the modifier function is multiplied with to get the desired heat
removal rate.
As a closing comment to this example, TEMP/W has a built in unique boundary
condition option to deal with convective heat transfer so that it is not necessary to
use a Q function with a modifier in this manner. This example is just an illustration
of how the modifier function can be used.

3
Modifier Factor (x 1000)

1
-10 0 10

Temperature

Figure 5-13 Thermal modifier function


Another example of modifier function usage is to adjust a temperature versus time
air temperature function so that ground temperatures become a function of both
ground and air temperature. This is discussed in detail earlier in this chapter.
One thing to note about the thermal modifier function, it can have any value
greater than 0. The seepage modifier function is a percentage modifier, but the
thermal modifier is not restricted in that way.

Page 124
TEMP/W Chapter 5: Boundary Conditions

5.12 Null elements


There are many situations where only a portion of the mesh is required. Parts of the
mesh that are not required in a TEMP/W analysis can be flagged as null elements.
Consider a case where you want to use a similar mesh in different GeoStudio
analyses. In the case of a cutoff wall to prevent water flow, you could set those
elements as NULL in the seepage analysis. This would treat the elements as if they
were not a part of the mesh and water would not flow around them. However, if
you now wanted to model the heat flow in this scenario, you would turn the
elements back on in TEMP/W and assign them valid properties.
Elements present, but not required in a particular analysis, can be assigned a
material type (number) with a conductivity function number of zero. Elements with
this uncharacterized material are treated as null elements. As far as the main solver
is concerned, these elements do not exist.
Null elements are not boundary conditions in the traditional sense, but they do alter
the actual solved domain and the locations where boundary conditions can be
applied. The boundary between active elements and null elements is just like any
other perimeter boundary where conditions can be specified.

Page 125
Chapter 5: Boundary Conditions TEMP/W

Page 126
TEMP/W Chapter 6: Analysis Types

6 Analysis Types
There are two fundamental types of finite element thermal analyses: these are
steady-state and transient. A description of each type and the implications
associated with each type are discussed in this chapter.

6.1 Steady-state
Think about the meaning of the words “steady-state.” They mean that the state of
the model is steady and not changing. In a thermal analysis, the “state” means the
temperature and heat flow rates. If they have reached a steady value throughout the
entire geometry, it means that they will be in that state forever. In many cases
where the geotechnical problem is exposed to nature with its cyclical conditions, it
is possible that a steady-state will never be reached. Heat flow beneath an artificial
skating rink surface may come close to steady-state if the ground thermal
conditions are held fairly constant over time. Ground surface temperatures on an
artic runway covered by snow will likely never reach a constant value. The model
only knows what you tell it and if you make the assumption the boundary
conditions are constant over time, then the model can compute the long term
ground conditions in response to your assumption. When the analysis is specified
as steady-state, the model can compute these long term conditions without
marching forward over multiple time steps. It can solve right to the end “steady-
state” condition.
This type of analysis does not consider how long it takes to get to a steady
condition and you have to understand that. The model will reach a solved set of
temperature and flow conditions for the given set of unique boundary conditions
applied to it and that’s it.
That sounds a bit blunt, but it’s very important to understand that when you are
doing a steady-state analysis you are not making any estimation of how long it
takes the final condition to develop nor how long it will last. You are only
predicting what the ground will look like for a given set of boundary values, and it
is implied that you are pretending the boundary values have been in place forever
and will be in place forever.
Because steady-state analyses are taking out the “time” component of the problem
it greatly simplifies the equations being solved. However, at the same time it can
make convergence somewhat harder to achieve – depending on the degree of non-
linearity of your soil property functions. The steady-state thermal equation leaves

Page 127
Chapter 6: Analysis Types TEMP/W

out the actual “time” variable and omits the entire unfrozen water content function
– which means the release of latent heat due to the phase change is not considered.
They are not needed in the solution. The unfrozen water content function gives an
indication of the rate of release of latent heat as the temperature drops and turns to
ice. In a steady-state analysis, the rate of heat release is not considered as any
generated latent heat is assumed to have long since dissipated by the time the
steady conditions are reached.

Boundary condition types in steady-state


In a steady-state analysis there are two choices of boundary conditions: a constant
temperature and a constant heat flux rate. For convenience the heat flux rate can be
specified as a total nodal flux or a unit flux applied to an element edge, but the end
result applied to the equations is identical. Convective heat transfer, thermosyphon
and climate boundary conditions are not applicable to steady-state analysis, as all
of these are time dependent processes.

6.2 Transient
A transient analysis by definition means one that is always changing. It is changing
because it considers how long the soil takes to respond to the user boundary
conditions. Examples of transient analyses include predicting the time it takes the
frost to reach a certain depth, or the time it takes to cool a heated ceramic tile etc.
In order to move forward in time during a transient analysis you must tell the
solver what the soil temperature conditions are at the start of the time period in
question. You cannot move ahead in time if you don’t have a starting point. In
other words, you must provide initial conditions as well as current or future
boundary conditions.

Initial conditions
For a transient analysis it is essential to define the initial (or starting) temperature
at all nodes. TEMP/W allows you to specify the initial conditions by either reading
the data from an initial conditions file created in a separate analysis, or by drawing
the initial nodal temperatures at all nodes. It is important to recognize that the
initial conditions for a transient analysis can have a significant affect on the
solution. Unrealistic initial conditions will lead to unrealistic solutions which may
be difficult to interpret, especially in the early stage of the transient analysis.

Page 128
TEMP/W Chapter 6: Analysis Types

When you can specify initial conditions by instructing the solver to use data from a
previously completed analysis, the initial conditions file must be identical in
geometry to the current file and can be from one of the following sources:

• A temperature file created by a steady-state thermal analysis,


• A temperature file created by a transient thermal analysis at a specific
time step for the same mesh geometry, or
• A temperature file created by the current analysis at a previously saved
time step to that which the current analysis is starting.
Consider the following: you know that the in-situ temperature is -2°C at the ground
surface and 0°C at a depth of 5 m. This initial condition can be established by
running a steady-state analysis that has a surface boundary temperature of -2°C
and a boundary temperature along the bottom of the problem of 0°C as illustrated
in Figure 6-1. The computed steady-state temperature file can then be used as the
initial conditions. The learn example in the getting started manual uses this
approach.

Figure 6-1 Using a steady-state analysis for initial conditions

Page 129
Chapter 6: Analysis Types TEMP/W

Drawing the initial nodal temperatures


In some cases, it is easier to specify initial nodal temperatures by simply defining
them at each node. You can do this in GeoStudio by simply specifying the value
and then using the mouse to select individual or groups of nodes. This is quite
simple to do in the cases where the ground is a fairly uniform temperature, but is
not so convenient to use where there is a varied thermal profile prior to the
transient process. In the latter case, it is easier to solve a steady-state solution to
obtain the varied starting conditions for the main transient analysis.
An example of every node defined by drawing an initial temperature is given in
Figure 6-2. The temperatures do not have to be the same at all nodes and nodes can
be grouped using the mouse in order to apply the desired value to many nodes at
one time.

Pipe

Figure 6-2 Defined initial temperatures at every node

No initial condition
In the event no initial condition is specified, TEMP/W will still solve the model.
However, it makes the assumption that the ground temperature is uniform at a
value of zero. This value of zero does not consider whether the problem thermal

Page 130
TEMP/W Chapter 6: Analysis Types

units are degrees Celsius or Fahrenheit, so it is not wise to omit leaving out a user-
specified initial condition.

6.3 Time stepping - temporal integration


An incremental time sequence is required for all transient analyses and the
appropriate time sequence is problem dependent. In most cases it will likely be
necessary to try a reasonable sequence and then adjust the sequence as necessary in
response to the computed results. For example, if the migration of the freezing
front is too rapid, the time steps need to be decreased; if the migration is too slow,
the time steps may need to be increased.
The accuracy of the computed results is dependent to some extent on the size of
the time step. Over the period of one time increment, the process is considered to
be linear. Each time step analysis is equivalent to a mini steady-state analysis. The
incremental stepping forward in time is in reality an approximation of the
nonlinear process. For the same rate of change, large time steps lead to more of an
approximation than small time steps. It follows that when the rate of change is
high, the time steps should be small, and when the rate of change is low, the time
steps should be large.
Many thermal processes related to the ground cooling follow an exponential form.
The dissipation of heat is rapid at first due to higher thermal gradients and then
decreases with time. A typical example is artificial ground freezing. To model this
situation, the time step sequence should approximately follow an exponential form.
The time steps should be small at first and then progressively increase.

Finite element temporal integration formulation


The following discussion is presented here and again in more detail in the Theory
chapter where the full development of the equations is given. It is important now to
just show the equation so that a key point can be highlighted that will enhance your
understanding of the transient finite element method.
The finite element solution for a transient analysis has temperature changes being a
function of time as indicated by the {T},t term in the finite element equation
below. In order to solve for the temperature at some point in the future, time
integration can be performed by a finite difference approximation scheme.
Writing the finite element equation in terms of a “Backward Difference” finite
difference form leads to the following equation (Segerlind, 1984):

Page 131
Chapter 6: Analysis Types TEMP/W

( ∆t [ K ] + [ M ]) {T } = ∆t {Q } + [ M ]{T }
1 1 0

Equation 6-1 or
∆t {Q1} + [ M ]{T0 }
{T1} =
∆t [ K ] + [ M ]

where:
dt = the time increment,
T1 = the temperature at end of time increment,
T0 = the temperature at start of time increment,
{Q1} = the nodal flux at end of time increment,
[K] = the element thermal conductivity matrix, and
[M] = the element heat storage matrix.

Look at this equation carefully for a moment. You can even ignore the braces and
brackets as they just indicate a grouping of node and element information with
some geometry tied in. The thing to focus on is that in order to solve for the new
temperatures at the end of the time increment, it is necessary to know the
temperatures at the start of the increment along with the average material
properties calculated at the average of the new and old temperatures. If you do not
have reasonable values for starting temperatures, then you make the equation
difficult to solve because you use these starting temperatures directly in the
equation AND also in the calculation of the average material properties.

Problems with time step sizes


The detailed discussion of how to compute optimum time steps is very advanced
and is left to those with advanced understanding of finite element mathematics
(Segerlind, 1984). For our purposes, let us just point out some key issues and make
some safe conclusions.
As discussed directly above, the shape and size of an element is tied into the
assembly of the [K] and [M] matrices. It is also plain to see that changing the time
step magnitude can change the solution of the computed temperatures. Small time
steps can cause overshoot in the calculation of the new temperatures and time steps
that are too big can result in poor averaging of material properties at the mid point

Page 132
TEMP/W Chapter 6: Analysis Types

of the time step. Elements that are too big can also result in poor material property
averaging while elements that are too small can lead to over shoot problems as
well.
The mathematical formulations for time step sizes generally take into account the
soil’s heat storage capability, its thermal conductivity, and the size of the element.
In a homogeneous, non-freezing analysis, the optimum time step can be readily
estimated. However, the problem in a 2D freezing analysis with multiple soil types
quickly becomes apparent. What may be an optimum time step for one region of
the problem, can also be a very poor choice of time step for another region; and,
unfortunately, the entire problem must be solved with the same time step.
Some programs have attempted to use adaptive meshing and adaptive time
stepping to deal with these issues, but the bottom line is that they cannot be
applicable to all points in a mesh at all times. Consequently, you are going to have
to try some different things and use some common sense.

General rules for setting time steps


So, now that we have made the issue of time steps somewhat unclear, what do we
recommend are some methods to deal with it? Some of the following points have a
strong theoretical basis and some are just common sense based on years of
experience.
The finite element shape is important. Triangular elements should not have any
interior angles greater than 900. Square elements can have double the time step size
as triangular elements.
Since the element size is directly proportional to time step size, doubling the
element size means you can double the time step. The corollary of this is that
decreasing the element size and not decreasing the time steps accordingly will not
improve the calculated results.
If you have a “instantaneous loading” of the system, such as instantaneous
application of a cold temperature you should set the initial time step size to be
approximately equal in order of magnitude to the unfrozen volumetric specific heat
value divided by the unfrozen thermal conductivity value of the soil that is
instantaneously loaded. This approach of dividing the storage rate by the
conductivity has relevance in consolidation theory, but can also be used as a
guideline in this situation.

Page 133
Chapter 6: Analysis Types TEMP/W

If you are modeling a rapidly cooling boundary condition, a better approach is to


use a Temperature versus Time boundary function to more accurately simulate the
initial cooling over a short, finite time period.
The best option is to try some time steps pattern and turn on the adaptive time
stepping scheme which is discussed and in the Numerical Issues chapter. TEMP/W
will permit you to activate an adaptive time stepping routine that will insert extra
time steps between your specified time steps in the event the solution is not
meeting the user’s specified time stepping or convergence criteria.
Just to get started, take the total time you need to model and divide it by five. Set
the model to have five equal time steps and run it. Watch how long it takes to
converge and if it even does converge. If it does not converge and you suspect it is
a time step issue, choose 10 smaller time steps. Once the model solves all right, as
a test, drop the time steps some more and see if the converged results change at all.
The bottom line is… the ONLY way to do time steps properly is by trial and error.
Not too advanced an approach, we know, but it’s the truth!

6.4 Axisymmetric
An axisymmetric analysis can be used to simulate three-dimensional problems
with symmetry about a vertical axis of rotation. The problem is defined in two
dimensions, but for analysis it is as if the section is rotated about a vertical central
axis. A typical example of an axisymmetric analysis is the flow into a single freeze
pipe installed in the ground as illustrated in Figure 6-3. In TEMP/W the vertical
symmetric axis of rotation is always at x-coordinate equal to zero. The x
coordinates in an axisymmetric finite element mesh must, therefore, all be greater
or equal to zero.

Page 134
TEMP/W Chapter 6: Analysis Types

Brine Pi pe
Gro und Surface

8.0 10

7.5

7.0

6.5
5
6.0
-20

5.5
-15

5.0
-10

4.5
-5

4.0

3.5
0

3.0

5
2.5

2.0

1.5

1.0

0.5

10
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Distance (m)

Figure 6-3 Frost bulb around freezing pipe


For an axisymmetric analysis, the computed flux is per unit radian if the element
thickness is specified as 1.0. If you want the computed flux value for the entire
circumferential area, you must either specify the element thickness as 6.2832 (i.e.,
2 pi radians) before you do the analysis, or simply multiply the above value by 2 pi
after the solution is finished. You can change the element thickness for the entire
mesh with the Draw Element Properties command.
Infinite elements may be used for the outside far field edge of the axisymmetric
mesh (right side away from the axis of symmetry). However, the application of

Page 135
Chapter 6: Analysis Types TEMP/W

non-zero q (unit flux) boundary conditions along the infinite element edges is not
allowed. This is because the nodal contributing area is dependent on the distance of
the node from the rotation axis, and since the far nodes of the infinite elements are
at infinity, the nodal contributing area becomes undefined. Even though TEMP/W
may still compute a solution to the problem in some simple cases, the solution
becomes suspect, and the use of q-type boundary conditions in this case is
therefore not recommended.
It is also not relevant to apply a small “q” unit flux boundary to the left edge of the
mesh if it is at an “x” coordinate of zero because then there is no area (no radius) to
compute an area over which the flux should be applied. You can, however, apply a
big “Q” flux because when you do this you are including the area of flow inside
the Q value you specify.

6.5 Plan view


A plan view analysis views the finite element mesh as lying on its side instead of
standing upright in a vertical plane. In TEMP/W the only difference between a
plan view and a 2D view analysis is the way the area of small “q” heat flux
boundary is computed.
In a 2D view, the area the “q” heat flux rate is multiplied by comes from the length
of the element edges that the boundary condition is specified on. This length is
then multiplied by the specified thickness of the element (typically unit thickness
into the screen) to obtain a total flux, Q, to apply at the relevant nodes.
In a plan view, the area the “q” heat flux rate is multiplied by comes from the plan
view areas of each element that contribute data to the “q” specified nodes. The
solver will compute the element areas and apply the total heat flux, Q, to the nodes.
In general, you should NOT need to use a plan view analysis. Even if you want to
model a plan view of a given elevation of the soil, you will typically use a 2D
section so that the flux boundary conditions consider the edge length of the
elements and not the elemental area of the element when computing the correct
total flux to apply.
You can use infinite elements in a Plan View analysis. However, the application of
a non-zero q (unit flux) to infinite elements is not recommended. The outer edge of
infinite elements is theoretically projected to infinity and consequently the
contributing areas for the outer edge nodes are not well defined. The resulting
temperatures may or may not be reasonable. If you are going to apply a surface

Page 136
TEMP/W Chapter 6: Analysis Types

heat flux to infinite elements, you will have to make a careful assessment of the
results to make sure they are reasonable. You can apply T and Q type boundaries
without difficulty.

Page 137
Chapter 6: Analysis Types TEMP/W

Page 138
TEMP/W Chapter 7: Numerical Issues

7 Numerical Issues
Entire textbooks can be written on numerical issues related to finite element
analysis. While modern computers and powerful graphics can make defining an
analysis quite fast and easy, they cannot necessarily deal with some of the intricate
issues related to the concept of taking a natural process and breaking it down into
finite time and special domains (i.e. individual elements within a soil geometry).
There are various ways to deal with many of the numerical issues, but the
unfortunate part is that there is no single method, approach, or technique that can
deal with all problems. Some numerical issues relate to restrictions in computer
hardware, such as rounding off of non-integer variables during math operations;
some issues relate to non-linearity of soil properties; some issues relate to the fact
the physical equations being solved do not apply to all cases (for example,
TEMP/W does not account for ice lensing or frost heave); some issues relate to our
inability to discretize a domain to small enough element sizes; and other issues
relate to the fact we have made the elements too small!
There are numerical solvers that make use of adaptive meshing or adaptive time
stepping or both in attempts to be more suited to a wider range of problems. All of
these, however, have their limitations from a sound mathematical perspective –
regardless of what the software developer will tell you. If you don’t know what the
limitations are, it becomes somewhat risky to rely on a solver that claims to
“handle it all.”
Some finite element solutions attempt to march forward in time by considering soil
property values taken from the last, the current or the mid-time step average. Some
solvers simply make assumptions that limit their applicability to real life in order
to get a solution, such as moving the mesh to find the water table in a seepage
program instead of solving for the physics of flow above and below the water
table. Finally, some solvers may only work if you put in an initial guess of the
solution for the dependent variable being solved that is close to the desired
solution, in other words, start off the solution by pointing it in the right direction.
While it may appear all hope is lost, this is far from the case. Sound judgment and
common sense can usually overcome most of these challenges and result in
meaningful interpretations of how the soil will respond to changes in various
parameters.

Page 139
Chapter 7: Numerical Issues TEMP/W

It is not always possible to get an exact solution in many challenging cases, so you
should not necessarily be seeking an exact solution. If the problem is so difficult
that it is not solving reasonably, then it is very likely that either mistakes have been
made in the input, or, that you are pushing the envelope of the physical theory
applied in the model. This chapter looks at some of these issues as they pertain to
TEMP/W.

7.1 Convergence
The objective of solving the finite element equations is to compute the total
temperature at each node. For linear analyses when the material properties are
constant, the nodal temperature can be computed directly. However, in the cases of
nonlinear analyses, when the material thermal conductivity is a function of
temperature, the correct material properties are not known at the start of the
analysis; consequently, an iterative scheme is required to solve the equations.
TEMP/W uses a repeated substitution technique in the iterative process. For the
first iteration, the user-specified initial temperatures are used to define the material
properties. The material properties are updated in the subsequent iteration using the
computed temperature from the previous iteration. For transient analyses, the
temperature at the mid-point of the time interval is used to define the material
properties; that is, the material properties are defined at the average of the past and
current computed temperatures. The iterative process continues until the iteration
number reaches the maximum number specified or until the results satisfy the
convergence criteria.
Convergence means repeated solving of the nodal thermal equations until the
computed solution does not change by more than a specified amount on successive
iterations. If the equations want to solve to the same results over and over, then this
means that the soil property values used in the solution agree with the computed
temperature values at each position in the mesh and that the influence of the
boundary conditions (i.e., adding or removing heat) also are reflected in the correct
soil properties and temperatures. Everything must be just right to balance both
sides of the thermal equation.
There are various ways to determine convergence from a mathematical
perspective, but not all of the ways are suited to Thermal analysis. A problem can
occur in Thermal equations because the variable we are solving, temperature, can
be negative, zero, or positive. Different logical considerations are necessary,

Page 140
TEMP/W Chapter 7: Numerical Issues

depending on what the temperature is, and the wrong logic can lead to wasted time
and no solutions. This is discussed in the next section.

Vector norms
TEMP/W does not use individual nodal temperatures to make the comparison
between successive iterations. Instead, it uses the percent change in the vector
norm – a process that considers all nodal temperatures simultaneously. The vector
norm is computed as the square root of the sum of each nodal temperature value
squared, or in equation form:

n
N= ∑T
i =1
i
2

where:

N = the vector norm,


i = a counter,
n = the total number of nodes, and
Ti = the individual nodal temperatures.

The percent change in the vector norm is simply the change in the value N
between two successive iterations. If the change in N were exactly zero, then
there would be no change in solution on successive iterations. In general, there are
always very small differences between iterations, so it is not reasonable to expect a
change in vector norm of zero.
You may wonder why we use the vector norm. It is a powerful way of considering
the entire problem in a single comparison, which prevents any individual nodal
temperature from potentially failing the convergence test in a solution that really
should be considered converged.
When the dependent variable (e.g. temperature in a thermal analysis) gets close to
zero, very small changes in a single variable between successive iterations can
appear to be very big when converted to a percent difference comparison. Consider
if a single nodal temperature is changing from 0.01 to 0.02 degrees between
successive iterations. This is a change of 100%, but likely totally irrelevant from a
physical perspective. It’s a failure in the logic of checking convergence by

Page 141
Chapter 7: Numerical Issues TEMP/W

individual nodal percent difference because it lets irrelevant physical changes


become grossly significant mathematical changes.
One way around this is to not make the comparison on a percent difference basis
for each node, but to make it based on percent difference of all nodes (e.g. the
vector norm). However, this logic also breaks down for the case where all the
temperatures are very close to zero as in the case of a permafrost analysis with
most of the region near the phase change point and only a few boundary nodes at
higher specified temperatures. When considering the vector norm, the higher
temperatures have more weight in the vector norm comparison, while all of the
changes in ground temperature are occurring near a value of zero. For this unique
case, it is a good idea to specify a very low percent difference so that small
changes in the vector norm caused by temperatures near zero are included in the
comparison between iterations.
During the iterative process, TEMP/W calculates the percent difference in vector
norm which represents the percent difference in the temperature of all nodes
between two consecutive iterations. In a fully unfrozen problem with an assumed
horizontal conductivity function the percent difference will be almost zero in the
second iteration. This indicates that the solution converges in two iterations, and
the temperature values of all nodes between the first and second iterations are
basically identical. However, in the case of a frozen problem where the
conductivity changes with temperature, the vector norm may be large in the early
iterations and gradually reduced to a small number.
If the solution is non-linear and the convergence process takes some time, you can
use the SOLVE Graph command to watch the convergence process graphically in
order to decide when convergence has been reached. Figure 7-1 presents a typical
vector norm versus Iteration plot of a converging process, the solution appears to
level out after about 22 iterations at which point it converges, makes adjustments
for the phase change, and repeats the solve. On the second attempt at a solution for
time step 2, the solver converges again by the 42 iteration (or another 20 iterations
after the phase change location was determined). While not clear in the figure, the
percent difference between the last two data points is less than the user-specified
tolerance.
It is interesting to note that the value of the vector norm has dropped slightly from
its starting value. This is a sign that the overall temperature of all the nodes has
dropped also. This is a good early indication that the solution is proceeding in the
intended direction. However, having stated that, if the majority of nodal

Page 142
TEMP/W Chapter 7: Numerical Issues

temperatures start to solve to more negative values, the vector norm will increase
again because it is the square of the temperature and the negative sign does not
affect the final vector norm.
Finally, if you see the vector norm value jump significantly up or down, you
should stop the analysis and determine if adjustments need to be made to the time
steps, soil properties or element sizes.

Step Number 2
85
Vector Norm

80

75

70
0 10 20 30 40 50
Iteration #

Figure 7-1 Vector norm versus iteration count prior with a phase
change

7.2 Steep material property functions


The solutions to frozen thermal problems can be highly nonlinear due to the
variability in conductivity as water changes to ice and in the rate at which heat is
released or retained during a transient process. The degree of nonlinearity is
dependent on the steepness of the conductivity function and unfrozen water
content function. Difficulties with convergence can be encountered when either of
these material property functions becomes relatively steep.
How steep is steep? There is no single answer. In general, you can think of
steepness in terms of two factors: how much the heat content or conductivity will

Page 143
Chapter 7: Numerical Issues TEMP/W

change for a small change in temperature; and, how fast will that temperature be
changing? A steep looking soil property function can be used quite readily, but
generally this requires smaller time steps in a transient analysis and smaller mesh
element sizes. In an extreme case of non convergence in a steady-state analysis,
you can try to use a long term transient analysis to simulate the steady-state. The
transient analysis has the advantage in that you have to specify initial conditions
and the solution can then march forward in time towards the steady-state solution.
Coarse-grained materials can sustain a relatively small capillary zone and tend to
have all their water change phase at small negative temperatures. This behavior
manifests itself in a very steep thermal conductivity function; that is, a function
where the thermal conductivity varies greatly with small changes in negative
temperature. If this is the case, consider dropping the time steps to allow the solver
to capture the phase change process more accurately.

7.3 Gauss integration order


The details of numerical integration are provided in the appendices, along with a
discussion of how different integration orders can affect results for various types of
elements. Part of this discussion is repeated here as it pertains to improving
solution convergence.
The appropriate integration order is a function of the presence of secondary nodes.
When secondary nodes are present, the interpolating functions are nonlinear and
consequently a higher integration order is required. Table 7-1 gives the acceptable
integration orders.

Table 7-1 Acceptable element integration orders

Element Type Secondary Nodes Integration Order


Quadrilateral no 4
Quadrilateral yes 9
Triangular no 1
Triangular yes 3

It is also acceptable to use four-point integration for quadrilateral elements that


have secondary nodes. This is called a reduced integration order (see Bathe, 1982).
Acceptable results can be obtained with reduced integration. For example, reduced
integration is useful in unfrozen zones where the thermal gradient is low and the

Page 144
TEMP/W Chapter 7: Numerical Issues

thermal conductivity is constant. Selective use of reduced integration can greatly


reduce the required number of computations.
It is also possible to use three-point and nine-point integration with elements that
have no secondary nodes. However, the benefits of this are marginal, particularly
for quadrilateral elements. Nine-point integration for a quadrilateral element
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. One-point integration
means the material properties and flow gradients are constant within the element.
This can lead to poor performance of the element, particularly if the element is in a
frozen zone where the thermal conductivity varies sharply with changes in pore-
water pressure. 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 for triangular elements 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.
In general, it is sufficient to use three-point integration for triangular elements and
four-point integration for quadrilateral elements. In situations where there is frozen
zone with thermal conductivity that varies sharply within an element, it is best to
use quadrilateral elements with secondary nodes together with nine-point
integration.

7.4 Equation solvers (direct or iterative)


TEMP/W has two types of equation solvers built into it; a direct equation solver
and an iterative equation solver. Both offer certain advantages.
Select the direct equation solver option if you want the system equations to be
solved using a Gauss elimination skyline direct solver. The processing speed of the
direct solver is bandwidth (the maximum node number difference of all the
elements in a domain) dependent. In other words, the direct solver is very fast
when solving simple problems with small bandwidth, but it can be quite slow
when solving more complex problems with a large bandwidth. TEMP/W

Page 145
Chapter 7: Numerical Issues TEMP/W

automatically sorts the nodes so that the bandwidth is the smallest possible value
which helps the solution solve faster using the direct solver. By default, the direct
equation solver is selected.
Select the iterative equation solver option if you want the system equations to be
solved using a preconditioned Bi-Conjugate Gradient (BiCG) iterative solver. The
processing speed of the iterative solver is independent of bandwidth (the maximum
node number difference of all the elements in a domain). The iterative solver can
be slower than the direct solver when solving a simple problem with small
bandwidth, but it can be much faster than the direct solver when solving more
complex problems with a large bandwidth. If you choose the iterative equation
solver, the maximum number of iterations the solver can attempt is set equal to the
maximum number of degrees of freedom in the problem domain.
The iterative solver tolerance parameter is different from the finite element
solution tolerance although they appear similar. The iterative solver tolerance is the
desired residual relative to the norm of the temperature vector ||T|| as a solution for
the global finite element matrices is obtained. The finite element solution
tolerance, on the other hand, is related to how the solution of all the equations
changes between two trials based on attempts to obtain the correct material
properties for the given boundary conditions.
If, for example, you choose an iterative solver tolerance of 10-6 it means that you
consider ||T|| to have errors in the range of 10-6. The iterative solver tolerances have
been defaulted to values that will work in most cases within a reasonable
computational effort. In the case where a high level of accuracy is required or
where the solution does not seem reasonable, you may need to tighten the solver
tolerance (e.g., to 10-8).

7.5 Time stepping


An incremental time sequence is required for all transient analyses and the
appropriate time sequence is problem dependent. In most cases it will likely be
necessary to try a reasonable sequence and then adjust the sequence as necessary in
response to the computed results. For example, if the migration of the wetting front
is too rapid, the time steps need to be decreased; if the migration is too slow, the
time steps need to be increased.
The accuracy of the computed results is dependent to some extent on the size of
the time step. Over the period of one time increment, the process is considered to

Page 146
TEMP/W Chapter 7: Numerical Issues

be linear. Each time step analysis is equivalent to a mini steady-state analysis. The
incremental stepping forward in time is in reality an approximation of the
nonlinear process. For the same rate of change, large time steps lead to more of an
approximation than small time steps. It follows that when the rate of change is
high, the time steps should be small, and when the rate of change is low, the time
steps should be large.
Many thermal processes related to the freezing of soils follow an exponential form.
The heat is rapid at first and then decreases with time. A typical example is the
artificial ground freezing. To model this situation, the time step sequence should
approximately follow an exponential form. The time steps should be small at first
and then progressively increase.

Automatic adaptive time stepping


TEMP/W will permit you to activate an adaptive time stepping routine that will
insert extra time steps between your specified time steps. There are three methods
of calculating adaptive time steps, plus a fourth scenario should the solution reach
its maximum allowable iteration count without reaching convergence.
In the first two methods, the change in nodal temperatures is used to determine an
increase or decrease in time step between the allowable ranges specified. The first
method checks each node to see if the temperatures between successive time steps
are changing by more than your percentage. If the allowable percentage
temperature change at any given node is too high, then the time steps will be
reduced such that the percent change is upheld. In the second option, the vector
norm of nodal temperatures is used as the criteria. The vector norm considers all
temperatures simultaneously. Experience should show that the vector norm
approach is faster for a large mesh with two dimensional heat flows. Analysis of a
column study or a mesh with heat flow primarily in one direction is better solved
with the individual nodal temperature comparison. The time stepping scheme that
compares temperature changes as adopted in TEMP/W is that proposed by Milly
(1982). As a general rule, nodal temperatures should not be allowed to vary more
than 5 or 10 percent over any given time step. Keep in mind that this criterion
depends somewhat on the actual temperatures being solved. For example, 5% of 1
degree is 0.05 degrees. 5% of 100 degrees is 5 degrees. A tighter tolerance may be
necessary for problems with an overall higher temperature range.
Regardless of the option to choose nodal temperatures or vector norm of nodal
temperatures, the equation used to adjust the time step up or down is:

Page 147
Chapter 7: Numerical Issues TEMP/W

Max Step
MinMultFactor =
Last Step
⎛ tolerance × T0 ⎞
MultFactor = abs ⎜ ⎟
⎝ T1 − T0 ⎠

The actual time step multiplication factor used to modify the existing time step is
the lesser of the two values calculated in the above equation.
The third adaptive time stepping control routine does not consider nodal
temperature change percentage. It continually tracks the total iteration count
required to achieve convergence. If the iteration count exceeds 5, then time step
adjustments are assessed. If the iteration count is less than 5, then no time step
adjustment is made for the next solve step. When deciding how much to increase
or decrease time steps, the solver compares the last converged iteration count with
the current average iteration count. So, if the average iterations to converge are 8
and the last time step took 10 iterations, then the solver will reduce the time step
by 8/10. Likewise, if the average count is 8 and the most recent count is 6, then the
time steps will increase by 8/6.
In the adaptive scheme, the time steps are reduced for the first 2 iterations and are
then held constant until convergence is reached. If the maximum number of
iterations is reached without convergence, then the solver checks to see if the
current time step is larger than the user-specified minimum allowable step. If the
current step is larger then the solver will reduce the current step by half and repeat
the time step. If there is no room to reduce the step by half without reducing it
below the user’s minimum value, then the minimum value will be applied and the
step repeated.
This multi faceted time stepping logic can be quite powerful. Often you will find it
quite useful to specify a lower number of allowable iterations to reach convergence
rather than a higher number, because if the solver gets stuck on any given time
step, it will reach the maximum allowable iterations more quickly and then try
reducing the time step in half. The trade-off becomes repeating a time step or
allowing more iterations. It is worth a simple trial and error test early on in a more
lengthy analysis to see which option is better in the long run.
The adaptive scheme will always insert just enough time steps to return to the
increments established by you. This way, no write out data steps are missed and

Page 148
TEMP/W Chapter 7: Numerical Issues

you still have control over the general time stepping of the solution. In order to
activate adaptive time stepping you must set up their time steps with preferred
"save" time steps. Then you check the adaptive time stepping box and time step
criteria as well as the maximum allowable change in nodal temperature per time
step, the maximum adaptive step, and the minimum allowable time step. The time
steps are entered in the same time units as the thermal conductivity.

Page 149
Chapter 7: Numerical Issues TEMP/W

Page 150
TEMP/W Chapter 8: Visualization of Results

8 Visualization of Results
When you get to the visualization of results stage of a finite element analysis, you
can congratulate yourself for having completed the hardest parts – setting up the
geometry, defining meaningful soil property functions, and applying appropriate
boundary conditions to the mesh. If, at this point, you do not have the tools or the
understanding of how to interpret the massive amount of data that may have been
generated by the solver, then you have wasted your time.
This chapter describes the various types of output data that are computed by the
solver, and it attempts to get you thinking about what the data is trying to tell you.
For example, did the solution solve properly? Did the boundary conditions I
applied get reflected in the actual solution? Did the soil respond how I thought it
would respond? If not, how do I methodically determine what to check next?
The chapter is structured to explain what type of data is available for visualization.
In the various sections, comments are provided that relate the type of result data in
question to how is should be used in the overall thought process. It’s a good idea to
read this entire chapter.

8.1 Transient versus steady-state results


The type of data you can view is somewhat dependent on the type of analysis you
have completed. For example, if you do a steady-state analysis, you cannot
interpret any data related to changes in time. You can view instantaneous flux
values that have units of heat per time: however, you cannot see how these values
may change with time because steady-state, by nature, means things do not change
with time.
In a steady-state analysis, you are not required to define an unfrozen water content
function, because it is the slope of this function that is needed in the solution of the
transient, not steady-state, finite element equation. If you do not define an unfrozen
content function, then it stands to reason that you cannot view unfrozen contents.
Water contents can only be viewed in a steady-state solution if you have defined an
unfrozen content function that the solver can access to report water contents based
on solved steady-state temperatures. No water content function… no reported
unfrozen water contents!

Page 151
Chapter 8: Visualization of Results TEMP/W

In a transient analysis, you can look at how all of the various output data values
change with respect to time and / or position; whereas, in steady-state, you can
only graph how the data changes with position.
Finally, in a transient analysis, you cannot plot the heat flow path across the entire
geometry. This is because in a transient process there are no single flow lines.
Sure, heat is flowing, but a single source of heat at one point in a mesh has an
infinite number of possible places to flow with time, and so it is not known what
the path will be. You can plot the heat flow vector within any element at any point
in time ,because by the simple nature of the “finite element” we are assuming the
flow within any single element is known at all points in time. In a steady-state
analysis, the flow line is useful to follow the path of a heat flow from its entrance
into the geometry to its exit. The entire path is known,because once the flow is
established it is at a steady value and therefore is fixed at that position. There is a
more detailed discussion on flow lines and flow vectors later in this chapter.

8.2 Node and element information


In order to understand what type of information can be viewed as results output, it
helps a bit to know how the data is obtained. So, recapping… you set up the
problem geometry, define material properties, and apply boundary conditions of
either known temperature or flux. The solver assembles the soil property and
geometry information for every Gauss point in every element and applies it to the
heat flow equation that is written for every node. Therefore, at each node we have
some applied boundary data, some interpolated soil property data and geometry
data. The solver then computes the unknown value in the equation for each node –
the unknown value being either temperature or flux. It is the Gauss point data that
is used to set up the nodal equations, so the Gauss point data written to the output
file is the actual data used in the solver.
Figure 8-1 is an illustration of the type of information that can be viewed for each
node in the finite element mesh. You can view the types of data in the list to see
that there is a combination of temperatures, fluxes, velocities, gradients,
conductivities, heat capacity and unfrozen water contents. There is also a summary
of the position of the node within the problem domain. In effect, the node
information is a summary of the problem geometry, the soil material properties,
and the boundary conditions – the three main parts of any finite element analysis.
One key point to note in the figure below is that the nodal Boundary Flux quantity
is exactly zero. This is an important point to understand because it can help with

Page 152
TEMP/W Chapter 8: Visualization of Results

your overall interpretation of results. This boundary flux is computed by summing


the contributing fluxes from each of the four Gauss points that surround this node.
So, if heat is flowing out of one Gauss region, it HAS TO be flowing into an
adjacent Gauss region. For all internal nodes with no user boundary applied to
them, the sum of all the heat fluxes at a node should equal zero in a properly
converged solution. In other words, there is heat balance at the node.
If the node being viewed is a boundary condition node (not necessarily at the edge
of the geometry, but with an allowed influx or outflux) then the summation of all
the heat fluxes at that node will not be zero, because heat is either gained or lost at
that point.

Figure 8-1 Visualization of node information


Figure 8-2 is the corresponding Gauss point information for the Gauss point
located just below and to the right of the node illustrated in the previous figure.
The shaded region in the figure shows the contributing area of that Gauss point and
in this case because the element is rectangular and has secondary nodes, this Gauss

Page 153
Chapter 8: Visualization of Results TEMP/W

area is equal to one ninth of the total area of the element. The inset in the figure
below shows the type of data that can be viewed at each Gauss point. If you
consider the unfrozen water content value of 0.226%, for example, you should
realize that this water content is assumed to exist throughout the Gauss point area
displayed; and you should next realize that if the element size is increased, the
estimate of the water content becomes less accurate, as we are averaging it over a
larger area. The real trick to getting good finite element analysis results is to create
a finite element mesh with just the right sized elements that are not too big or too
small, that can represent the highly non-linear soil properties within them, and that
can handle the potentially extreme boundary conditions you apply. It’s not always
easy and there is no sure quick or automatic method to make that happen.

Figure 8-2 Visualization of element Gauss point information

8.3 Equipotential lines


Equipotential lines are lines connecting all points of equal energy potential. The
energy in this case is represented by temperature and the equipotential lines are

Page 154
TEMP/W Chapter 8: Visualization of Results

called isotherms. In a one-dimensional analysis with heat flow in the vertical


direction, the equipotential lines are exactly horizontal. In a two-dimensional
analysis, the equipotential lines can exist in an infinite number of directions, as
shown in Figure 8-3.

Projecting Gauss point values to nodes


TEMP/W performs contouring calculations based on parameter values at the
nodes. Since the primary parameter, temperature, is computed at the nodes, this
parameter can be contoured directly. However, secondary parameters, (velocity,
gradient, conductivity, and unfrozen water content), are computed at the element
Gauss points 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).

4
-4.5

-2.5
3
metres

-0.5

2 0.5

1.5
1

2.5

0
0 1 2 3 4 5 6 7

metres

Figure 8-3 Illustration of isotherms generated in TEMP/W program

Page 155
Chapter 8: Visualization of Results TEMP/W

In quadrilateral elements, the Gauss point values are projected using the
interpolating functions. (For more information about interpolating functions, see
the appendix). In equation form:

x= N {X }
where:
x = the projected value outside the Gauss points at a local coordinate greater
than 1.0,
<N> = a matrix of interpolating functions, and
{X} = the 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 8-4 is an
example of the local coordinates at the element corner nodes when projecting
outward from the four Gauss points in the element. The value of 1.7320 is the
reciprocal of the Gauss point coordinate 0.57735.
This projection technique can result in some over-shoot at the corner nodes when
variation in the parameter values at the Gauss points is large. For example,
consider that we wish to contour unfrozen water content and that in some elements
the water content at the Gauss points varies over the complete range of the water
content function. Projecting such a large variation to the nodes can result in
projected nodal water contents beyond the range of the unfrozen water content
function.
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.

Page 156
TEMP/W Chapter 8: Visualization of Results

Figure 8-4 Local coordinates at the corner nodes of an element with


four integration points

8.4 Contours
The power of using advanced graphical interfaces with finite element analysis is
that the computer can quickly convert thousands of pieces of data into meaningful
pictures. In the section above we introduced isotherms. We can use a picture of the
isotherms to tell us something about what is going on in the soil. In particular, if
we consider how close adjacent isotherms are to each other, we are in effect,
considering how steep the thermal gradient is. If we recall that the amount of heat
flow is equal to the thermal gradient multiplied by the conductivity, we then have a
fast and clear picture of where the areas of high heat flow are in the domain we
have modeled.
TEMP/W is a powerful tool in that it will let you contour many different
parameters, such as temperatures, gradients, heat fluxes, water contents and
thermal conductivity. Figure 8-5 is a contour of the actual heat flux values for the
temperature isotherms illustrated above. As we already determined, the isotherms
were close together near the surface and we therefore realized the heat flux was
higher in this region also. Figure 8-5 confirms our expectations.

Page 157
Chapter 8: Visualization of Results TEMP/W

39.811
630.96
3
metres

398.11

158.49
2

100
1

0
0 1 2 3 4 5 6 7

metres

Figure 8-5 TEMP/W computed heat flux contours

8.5 Energy flow vectors and flow paths

Calculating gradients and velocities


Once the solution has converged and the nodal temperatures are known, TEMP/W
computes the thermal gradients and heat flow velocities at each of the integration
points within each element. The gradient at each Gauss or integration point is
computed from the equation:

⎧ ix ⎫
⎨ ⎬ = [ B ]{T }
⎩i y ⎭

where:

ix = the gradient in x direction,

iy = the gradient in y direction,

Page 158
TEMP/W Chapter 8: Visualization of Results

[B] = the gradient matrix as defined in the theory chapter, and


{T} = the vector of temperature at the nodes.

The heat flow velocities at each Gauss point are computed from the equation:

⎧ vx ⎫
⎨ ⎬ = [C ][ B ]{T }
⎩v y ⎭
where:

vx = the velocity in x direction,

vy = the velocity in y direction, and

[C] = the thermal conductivity matrix.

TEMP/W stores in an array the thermal conductivity at each Gauss point used in
the formulation of the finite element equations. The same thermal conductivity
values are later used to compute the velocities.

Energy flow vectors


Energy flow vectors are a useful way of seeing not only where the heat flow is
occurring, but how much flow there is relative to other regions of the domain.
Examples of such energy flow vectors are given in Figure 8-6. TEMP/W uses the
magnitude of the actual energy flow in its calculation of how large to display the
vector, so that you have a visual representation of where the energy flow rates are
high or low. For each element, the average x-energy flow rate and average y-
energy flow rate from the Gauss point flow rate values are computed and then
vectorally summed to obtain an average energy flow vector for the element. This
average vector is plotted with the tail of the vector at the center of the element.

Page 159
Chapter 8: Visualization of Results TEMP/W

3
metres

0
0 1 2 3 4 5 6 7

metres

Figure 8-6 Heat flow vectors


When displaying vectors, TEMP/W finds the maximum vector and draws it at the
length specified in the Draw Vectors dialog box. All other vectors are drawn in
proportion to the element energy flow rate relative to the maximum flow rate. For
example, if the element flow rate is one quarter of the maximum flow rate for the
domain, then the length of the vector is one-quarter of the length specified in the
Draw Vectors dialog box.
You have the option of controlling the size of the vectors as they are displayed by
controlling the magnification scale when you issue the command to create the
vectors. Specifying a magnification value allows you to control the scale at which
all vectors are 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
vector will be drawn. You can control the vector length either by specifying a
magnification value or by specifying a maximum display length value. If you
specify a length, the magnification value is computed by dividing the maximum
length by the maximum flow rate and adjusting the value for the scale of the page
and engineering units.

Page 160
TEMP/W Chapter 8: Visualization of Results

Vectors are only drawn in element regions being viewed. Choose View Element
Regions if you wish to view different materials or to not view infinite elements. If
no elements are viewed, an error message appears when you choose Draw Vectors.

Flow paths
The TEMP/W flow paths are not truly flow lines or stream lines as in a traditional
seepage flow net sense of the word. In many cases the flow paths are a very good
approximation of stream lines, but they are not the same. The flow paths are
simply a line based on energy flow rate vectors in an element that a source of heat
would follow under steady-state conditions; and slight variations between a flow
path and a flow line should not be of concern, because they are computed in
entirely different ways. At best, the TEMP/W flow path should be viewed as a
reasonable approximation of the flow lines within a flow net.
The flow paths will always be the most realistic in high thermal gradient zones
where there is significant energy flow rate. In zones where there is little or no flow,
the TEMP/W flow paths may not be realistic. When you create a flow path, a
message will be displayed if you attempt to draw it in an area where there is little
or no flow. After accepting the message, the flow path will be drawn, but it may
not be complete; that is, the path will end inside the flow regime and not at an
external mesh boundary. This message will also be displayed if the flow path
encounters a no-flow perimeter boundary.
You can select any point to draw a flow path by clicking on a point within the flow
domain. The path is drawn strictly on the basis of the energy flow rate vectors
within each element. The path is projected forward and backward incrementally
within each element until the path encounters a boundary. The flow path is simply
a graphical representation of the route a source of heat would move under steady-
state conditions from the entrance to exit point within the flow regime.
Flow paths can only be drawn for steady-state conditions. Flow paths based on
energy flow rate vectors at an instant in time during a transient process have no
physical meaning. For this reason, TEMP/W does not permit you to draw flow
paths for transient conditions.

8.6 Flux sections


TEMP/W has the ability to compute the instantaneous total energy flow rate across
a user-defined section for either a steady-state or transient analysis. This is a very

Page 161
Chapter 8: Visualization of Results TEMP/W

useful tool for isolating quantity of heat to or from specific regions of interest, and
it can save you manually adding up individual nodal flows in the case of a heat
source or sink that is comprised of many nodes.

Flux section theory


This value can be computed from the nodal temperatures and the coefficients of
the finite element equation. For example, consider a mesh with only one element,
as illustrated in Figure 8-7. The objective is to compute the total flow across a
vertical section of the element.
In the Theory chapter, the finite element form of the heat flow equation is
presented. It can be re-written with the flux value isolated on one side as follows:

∆T
[ K ]{T } + [ M ] = {Q}
∆t
∆T
In a steady-state analysis, the storage term [ M ] becomes zero, and the
∆t
equation can be reduced to:

[ K ]{T } = {Q}

Figure 8-7 Illustration of a flux section across a single element


The global set of finite equations for one element is as follows:

Page 162
TEMP/W Chapter 8: Visualization of Results

⎡ c11 c12 c13 c14 ⎤ ⎧T1 ⎫ ⎧ Q1 ⎫


⎢c ⎪ ⎪
⎢ 21 c22 c23 c24 ⎥⎥ ⎪T2 ⎪⎪ ⎪Q2 ⎪⎪
Equation 8-1 ⎨ ⎬=⎨ ⎬
⎢ c31 c32 c33 c34 ⎥ ⎪T3 ⎪ ⎪Q3 ⎪
⎢ ⎥
⎣ c41 c42 c43 c44 ⎦ ⎪⎩T4 ⎪⎭ ⎩⎪Q4 ⎪⎭

From the basic heat flow equation, the total flow between two points is:

∆T
Equation 8-2 Q=k A
l
KA
The coefficients, c, in Equation 8-1 are a representation of in Equation 8-2
l
Therefore, the flow from Node i to Node j is:

Qij = cij ( Ti − T j )

In a transient analysis, because of heat storage, the calculation of the total flow
quantity must include the storage effect. The change in flow quantity due to the
storage term can be expressed as:

⎡ m11 m12 m13 m14 ⎤ ⎧ ∆T1 ⎫ ⎧ Q1 ⎫


⎢m ⎪ ⎪
1 ⎢ 21 m22 m23 m24 ⎥⎥ ⎪∆T2 ⎪⎪ ⎪Q2 ⎪⎪
⎨ ⎬=⎨ ⎬
∆t ⎢ m31 m32 m33 m34 ⎥ ⎪ ∆T3 ⎪ ⎪Q3 ⎪
⎢ ⎥
⎣ m41 m42 m43 m44 ⎦ ⎪⎩∆T4 ⎪⎭ ⎪⎩Q4 ⎪⎭

where ∆T1,2,3,4 etc. are the changes of temperature at the various nodes between
the start and the end of a time step. In general, the average change of temperature
from Node i to Node j can be expressed as:

∆Ti + ∆T j
∆Tij =
2
Therefore, the change in flow quantity from Node i to Node j due to a change in
storage is:

Page 163
Chapter 8: Visualization of Results TEMP/W

∆Tij
Qij = mij
∆t
The total flow quantity from Node i to Node j for a transient analysis then
becomes:

∆Tij
Qij = cij (Ti − T j ) + mij
∆t
The total flow quantity through the flux section shown in Figure 8-7 is:

Q = Q21+ + Q24 + Q31 + Q34

The imaginary flow lines from one side of the section to the other side are known
as subsections. TEMP/W identifies all subsections across a user-defined flux
section, computes the flow for each subsection, and then sums the subsection flows
to obtain the total heat flow across the flux section.

Flux section application


Flux sections can be used in many ways because they can be drawn any place you
want to know the flux across. You many want to check if an influx is equal to an
outflux, such as illustrated in Figure 8-8. In this case the values are NOT the same,
which is to be expected because the problem is a transient analysis and there is a
change in the amount of heat storage over time. If this solution were allowed to
solve for a much longer time frame, then the two fluxes would eventually come to
a similar value, which would indicate that at that point in time, the solution has
reached the equivalent of a steady-state condition.

Page 164
TEMP/W Chapter 8: Visualization of Results

4
3.4474e+003

3
metres

9.1957e+002
0
0 1 2 3 4 5 6 7

metres
Figure 8-8 Flux section used to check balance of inflow and outflow
Flux sections do not have to be drawn as single straight lines. They can be made of
continuous attached segments, as illustrated in both figures above. When a
multiple segmented flux section is drawn, the value of flux reported for the section
applies to the entire section, not any individual segment.
The key point to note when defining a flux section is to make the flux section cross
the sides of the elements and not the nodes of the elements. Also, if you want to
check the flux around a closed loop, as illustrated for the SEEP/W seepage drain
nodes in Figure 8-9, make sure the end of the flux section crosses over the tail of
the first segment of the flux section.
Two words of caution: flux sections MUST be defined before you solve the
problem, because the program needs to calculate the values during the solution
sequence, not afterward. In addition, all flux values are reported as positive, which
means direction is not taken into account. This is required because the sign of the
flux value will depend on which way you draw the section. To avoid any
misinterpretation, all flux section values are reported as positive and then you can

Page 165
Chapter 8: Visualization of Results TEMP/W

plot flux vectors in order to determine the direction of flow, if it is not obvious
based on your problem definition.

No flow boundary with


water table initially at surface
1.00 Sum of each nodal Q:
(3 x 0) = 0 plus
(2 x 0.15522) = 0.31044 plus
(2 x 0.3919) = 1.09424 plus
(1 x 0.40576) = 1.50000
0.75

Seepage
0.50 boundary

1.5000e+000

0.25

0.00
Q=0.5 m^3/time nodal flux boundary

Figure 8-9 Flux section used in SEEP/W around series of drain nodes
to check flow

8.7 Changes between selected times


It can often be useful to view how certain data changes with time and position.
TEMP/W will let you simultaneously view the change in the freeze-thaw line for
any or all of the time steps in your model, as illustrated in Figure 8-10. In this case,
the freeze-thaw line is shown to be moving downward in the soil beneath the
skating rink. The numerical value shown on the actual freeze thaw line is an
integer value representing the computed time step, not the actual elapsed time.

Page 166
TEMP/W Chapter 8: Visualization of Results

4
2
4

6
3
metres

8
2

0
0 1 2 3 4 5 6 7

metres

Figure 8-10 Transient position of freeze-thaw line during frost


penetration beneath a skating rink
5

3
metres

0
0 1 2 3 4 5 6 7

metres
Figure 8-11 Freeze-thaw line and temperature contours after time step
2

Page 167
Chapter 8: Visualization of Results TEMP/W

3
metres

0
0 1 2 3 4 5 6 7

metres

Figure 8-12 Freeze-thaw line and temperature contours after time step
6
Most parameters cannot be contoured for more than a single time step at a time
because, unlike plotting the change in freeze-thaw line, most contour values vary
over the entire domain. In the event you want to view contours at different time
steps, you must select the time step in question and re-plot the contour as shown in
the next two figures. In the event you want more specific data output as a function
of time and position, you can use the graphing feature, which is discussed next.

8.8 Graphing
The Draw Graph command allows you to plot a graph containing any of the
following computed parameter values: temperature, nodal boundary flux, x-energy
flow rate, y- energy flow rate, x-y energy flow rate, x gradient, y gradient, x-y
gradient, x conductivity, y conductivity, heat capacity, and unfrozen water content.
The addition of a climate data boundary condition means the following additional
parameters can be viewed for any node using a climate boundary condition option:
net solar radiation, snow depth, snow water equivalent, snow surface temperature,

Page 168
TEMP/W Chapter 8: Visualization of Results

surface latent heat flux, and surface sensible heat flux. All the latter heat flux
values are cumulative since the previous write out time step.
The above-listed parameters 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, and the distance between nodes (starting
at the first selected node). If your problem is a transient analysis, you can also plot
a dependent variable versus time for one or more of the time steps.
The independent graph variable that you choose affects how the selected nodes and
time steps are used in the graph. 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. These two cases are illustrated in the Figure 8-13 and Figure 8-14
where the unfrozen water content value beneath the rink is compared versus
position and time respectively. The nodes down the left edge of the above-
illustrated examples are chosen for plotting data in the following figures. Note in
these figures that there is some “overshoot” of graphed values (e.g. the negative
water content). The reasons for this are discussed above in the section on
projection of Gauss point values to nodes.
These illustrations show the default graph template used in the model. There is,
however, the ability to customize the scales, labels, and symbols so that the plots
can be used directly within your reports. In addition, the above figures were simply
copied from the screen and pasted into this document using the COPY command
inside TEMP/W.

Page 169
Chapter 8: Visualization of Results TEMP/W

Unfrozen W.C. vs. Y


4

3.0000e+000

1.5000e+001
Y

6.3000e+001
1

2.2700e+002
0
-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6
Unfrozen W.C.

Figure 8-13 Left side unfrozen water contents as a function of y-


coordinate for each time

Unfrozen W.C. vs. Time


0.6

Node 1
0.5 Node 20
Node 30
Node 49
0.4 Node 59
Node 78
Unfrozen W.C.

Node 88
Node 107
0.3
Node 117
Node 136
Node 146
0.2 Node 165
Node 175
Node 194
0.1 Node 204
Node 223
Node 233
0.0 Node 252
Node 262
Node 281
-0.1 Node 291
0 50 100 150 200 250
Time

Figure 8-14 Left side unfrozen water contents as a function of time for
each y-coordinate

Page 170
TEMP/W Chapter 8: Visualization of Results

8.9 Reporting
In the event you need to further customize your output data, TEMP/W will let you
view, then COPY data to many other programs that, such as Microsoft Excel or
Word or other graphing packages. Once you decide what you want to graph, and
perhaps you have even viewed the graph on screen, you can select the option to
create a table of all of the data. The data in Table 8-1 shows the unfrozen water
content versus time for node 211. The data was copied directly from the TEMP/W
generated graph and pasted as unformatted text in this Word document.
This data can also be obtained directly from the actual output files created by the
solver. While this is not the general recommendation, if you are adventurous, you
can simply open the output data files at any given time step and manually extract
the necessary data. The very adventurous can even program macros within
Microsoft Excel to open files, get the data and close the files. This may be useful if
you want to automatically track the volume of water over a unique region of the
mesh at multiple time steps. Having said this, you should know that GEO-SLOPE
will not be helping you with any macros you want to write in Excel or any other
program. We just mention it here because it is do-able.

Table 8-1 Data copied from TEMP/W and pasted directly into Word

Point # Time Unfrozen W.C. 211


1 +3.0000e+000 +5.0000e-001
2 +1.5000e+001 +5.0000e-001
3 +6.3000e+001 +4.9999e-001
4 +2.2700e+002 +1.9033e-001

Page 171
Chapter 8: Visualization of Results TEMP/W

Page 172
TEMP/W Chapter 9: Modeling Tips and Tricks

9 Modeling Tips and Tricks

9.1 Introduction
This chapter contains many useful hints about using the software and
understanding what it does. READ THIS CHAPTER!
There have been many occasions where GEO-SLOPE has been contacted by
clients with questions about how the model behaves in response to changes in
various parameters. If we do not know the answer, we conduct a numerical
experiment to test what will happen. The first few sections of this chapter illustrate
a few common examples of numerical experiments. You are strongly encouraged
to learn why these types of simple tests are so powerful in testing how the program
computes results, but ALSO in enhancing your understanding of how the physical
mechanisms of flow through porous medium occurs.
A numerical experiment is carried out by making a very simple finite element
problem. It is useful to use a mesh that is one distance unit wide and one distance
unit high. This makes hand-calculating flux values very simple and they can easily
be checked against the computed flux values. The following discussion illustrates
how some simple numerical experiments have been carried out to test some
simple, yet valid, questions.
When setting up these experiments, it is a good idea to input simple soil property
functions. In most cases, two data points are sufficient to define the conductivity
and storage function. Just as a reminder, give both functions some slope – don’t
make them horizontal!

9.2 Problem engineering units


Any system of units can be used for a TEMP/W analysis; the only requirement is
that you must be consistent. Fundamentally, you must select the units for length
(geometry), time, heat and temperature. Once you have selected units for these
parameters, all other units must be consistent. Table 9-1 to Table 9-4 present some
typical sets of consistent units.

Page 173
Chapter 9: Modeling Tips and Tricks TEMP/W

Table 9-1 Consistent Set of Units in Meters, Seconds, Joules and Degrees
Celsius

Parameter Symbol Units


Length L meters (m)
Time t seconds (sec)
Heat H Joules (J)
Temperature T °C
3
Latent heat of water H/L J/m3
Thermal conductivity H/(t × L × T) J/(sec × m × °C)
Volumetric heat capacity H/(L × T)
3
J/(m3 × °C)
Total flux (Q) H/t J/sec
Unit heat flux (q) H/(t × L ) 2
J/(sec × m2)

Table 9-2 Consistent Set of Units in Meters, Days, Kilojoules Degrees Celsius

Parameter Symbol Units


Length L meters (m)
Time t days (day)
Heat H kilojoules (kJ)
Temperature T °C
3
Latent heat of water H/L kJ/m3
Thermal conductivity H/(t × L × T) kJ/(day × m × °C)
Volumetric heat capacity H/(L × T)
3
kJ/(m3 × °C)
Total flux (Q) H/t kJ/day
Unit heat flux (q) H/(t × L ) 2
kJ/(day × m2)

Table 9-3 Consistent Set of Units in Feet, Hours, Btu and Degrees Fahrenheit

Parameter Symbol Units


Length L feet (ft)
Time t hours (hr)
Heat H Btu (Btu)
Temperature T °F
Latent heat of water H/L3 Btu/ft3

Page 174
TEMP/W Chapter 9: Modeling Tips and Tricks

Thermal conductivity H/(t × L × T) Btu/(hr × ft × °F)


Volumetric heat capacity H/(L × T)
3
Btu/(ft3 × °F)
Total flux (Q) H/t Btu/hr
Unit heat flux (q) H/(t × L ) 2
Btu/(hr × ft2)

Table 9-4 Consistent Set of Units in Meters, Hours, Kilocalories and Degrees
Celsius

Parameter Symbol Units


Length L meters (m)
Time t hours (hr)
Heat H kilocalories (kcal)
Temperature T °C
3
Latent heat of water H/L kcal/m3
Thermal conductivity H/(t × L × T) kcal/(hr × m × °C)
Volumetric heat capacity H/(L × T)
3
kcal/(m3 × °C)
Total flux (Q) H/t kcal/hr
Unit heat flux (q) H/(t × L ) 2
kcal/(hr × m2)

NOTE: The unit of J/sec is sometimes expressed as Watts, (1 J/sec = 1 W).


The latent heat of water expressed in the above four example units are:
Latent heat of water = 334 x 106 J/m3 = 334 x 103 kJ/m3 = 8975 Btu/ft3 = 79760
kcal/m3
In summary, many systems of units can be used in TEMP/W. The key requirement
is that the system of units be consistent. Generally, all units are established once
you select the units for thermal conductivity. For example, if the thermal
conductivity is chosen to be in kJ/day/m/°C, then length must be in meters, heat
energy must be in kJ, time must be in days and temperature must be in °C.

9.3 Convergence Tolerances


Once the phase change region has been established for a particular time step,
TEMP/W continues the iteration process until the user-specified tolerance is
achieved. Tolerance is the allowable maximum percentage difference between two
consecutive vector norms. The vector norm is the normal of the computed

Page 175
Chapter 9: Modeling Tips and Tricks TEMP/W

temperature vector. Tolerance is specified in DEFINE and entering a percentage


value in the Tolerance edit box. For more information about convergence
tolerance, please refer to the chapter on Numerical Issues.
Similar to the Gauss Regions Iterations (GRI) parameters (as discussed in the
Theory chapter), a very small tolerance (e.g., 0.001%) may result in excessive
number of iterations required to reach a solution. Conversely, making the tolerance
too big (e.g., 5%) may result in false convergence, thereby making the solution
unrealistic. The following discusses some factors that need to be considered when
selecting a suitable tolerance.
For a highly nonlinear problem (i.e., steep conductivity function or steep unfrozen
water content function), the vector norm tends to vary considerably between
iterations, therefore, using a small tolerance may results in many iterations. In such
cases, it is helpful to start the analysis with a bigger tolerance (say 0.5%). The
tolerance can then be lowered in stages until a further reduction in the tolerance
has no significant effect on the computed temperature distribution.
The selection of a suitable tolerance is related to the specified GRI parameters. In
general, when a small GRI value is used, many iterations are required to establish
the phase change. As a result, the solution is less sensitive to the specified
tolerance, and a larger tolerance may be used without significantly affecting the
solution. Similarly, when a large GRI value is used, the solution is more sensitive
to the specified tolerance, and a smaller tolerance should be used.
You may use the Graph feature in TEMP/W SOLVE to view the convergence of
the vector norm as a function of iteration. For more information about graphing the
vector norm, refer again to the Numerical Issues chapter.

9.4 Flux section location


Question: Does the location of a flux section within an element have any influence
on the computed flux value?
Answer: No. The flux section value will be the same regardless of whether the
section is drawn near the element edge or element middle. Figure 9-1 shows this to
be the case, and it is true for a transient and steady-state solution.

Page 176
TEMP/W Chapter 9: Modeling Tips and Tricks

2.6869e+002

2.6869e+002

Figure 9-1 Test to check flux section locations

9.5 Source or sink flux values


Question: What is the best way to compute a total heat sink flux?
Answer: You can sum each individual nodal flux, or, draw a flux section around
the nodes.
The figure below shows a total flux into the system through the three bottom nodes
of -500 kJ per time unit (each node contributing a total nodal flux of -125 kJ). This
example also shows that with no other place for the heat to flow except the sink,
the flux into the base equals the flux out of the sink nodes.

Page 177
Chapter 9: Modeling Tips and Tricks TEMP/W

Distance (m)
5.0000e+002

0
5.0000e+002
5 6

Distance (m)
Figure 9-2 Test to see best method for observing sink flux

9.6 Unit energy flux versus total energy flux?


There are many people who are unsure of the difference between a unit flux and a
total nodal flux. Do a simple test if you are unsure.
Question: How is a unit flux related to a total nodal flux in a 2D analysis?
Answer: The total nodal flux should be exactly equal to the unit flux multiplied by
the total length of the element edges that contribute to that node.
In the figure below, a unit flux of -500 kJ / time / meter edge length has been
applied to the top of the element. The top is a heat sink face which will let the
energy out. The flux sections drawn in the element confirm that the total edge flux
of -500 kJ / time has been converted by the solver into two equal total nodal fluxes
of -250 kJ / time each. For such a simple mesh, it is also possible to use the View
Node information option and click on each node to see the computed total energy
flux at each node. The sum of the individual total nodal fluxes is the total energy
flux across the element edge.

Page 178
TEMP/W Chapter 9: Modeling Tips and Tricks

1 2
0 2.
+0 50
00
0 0e e+
50 00
2. 2
Distance (m)

5.0000e+002

0
5 6

Distance (m)

Figure 9-3 Test to compare unit flux and total flux

9.7 Stopping and restarting an analysis


TEMP/W has a powerful Stop-Restart feature. The processing may be stopped at
any time step and then restarted after making changes in the problem. For example,
consider a transient problem with 10 time steps, and that you click the Stop, button
after the 5th time step is complete. You can then make changes with DEFINE and
re-save the problem.
Typical changes you may want to consider when you stop and restart an analysis
include:

• addition or removal of an element as discussed below,


• changing adaptive time stepping parameters,
• changing convergence parameters,
• adjusting material properties to account for time dependent changes, or

Page 179
Chapter 9: Modeling Tips and Tricks TEMP/W

• adding or deleting a boundary condition, such as a freeze pipe or heat


source

9.8 Element addition and removal


Any element can be considered to be nonexistent by assigning the conductivity
function number a value of zero. This feature makes it possible to simulate the
construction of embankments and excavations or the deposition of hot waste
material on existing frozen ground where the waste is placed in lifts.
It is necessary at the start of the problem to define the elements you anticipate
adding or removing. Defining the complete mesh at the beginning of an analysis
will maintain compatibility between files. The total number of elements and
number of nodes must not change as you progress from one stage of the analysis to
another. If the number of nodes or elements changes, it is no longer possible to use
previously computed results as initial conditions for a subsequent step, because the
files will be incompatible.
The Draw Element Properties command in DEFINE is useful for applying a null
material type to specific elements as you are simulating the construction of
embankments or excavations. A null material type may be defined by using the
Key In Material Properties command to create a material type with a thermal K-
Function number of zero.
If you add an element using the stop-restart feature, you must consider that the
initial conditions of the element may not be realistic. By default, the program will
consider the newly added elements to have a temperature of zero, which is not
likely true, especially if the solver units are in degrees Fahrenheit. There is a
roundabout way to correct this problem in the short term, and GEO-SLOPE
International Ltd. is working on a better solution for a future release.
In the short term, when the new elements are made active, a one second time step
can be solved where each node in the new elements is set at the correct
temperature using a T boundary condition. After the onr second analysis, stop the
solver, remove the T boundary, and continue solving until the next desired element
addition elapsed time or until the completion of the simulation.

Page 180
TEMP/W Chapter 10: Product Integration Illustrations

10 Product Integration Illustrations


GeoStudio is a unique geotechnical engineering tool that has been designed to
allow seamless integration between various types of geotechnical engineering
analyses. It is now possible to define problem geometry one time and then apply
various types of boundary conditions in order to solve various partial differential
equations.
It is possible, for example, to use SEEP/W or VADOSE/W generated pore-water
pressures in a SLOPE/W stability or QUAKE/W analysis. It is possible to account
for flowing water in a CTRAN/W contaminant transport process or TEMP/W heat
transfer processes related to natural or artificial ground freezing. It is possible to
specify different pore pressure conditions over time in order to carry out coupled
or uncoupled seepage consolidation and seepage influenced stress–deformation
analyses using SIGMA/W.
This chapter provides several examples that illustrate how GeoStudio facilitates the
smooth integration between various analytical tools.
The following examples are included as a sampling of what can be carried out
using GeoStudio.

• SEEP/W generated pore pressures in SLOPE/w stability analysis


• VADOSE/W generated pore pressures in SLOPE/W stability analysis
• SEEP/W dissipation of pore pressures generated in a QUAKE/W earth
quake analysis
• SEEP/W velocity data in a CTRAN/W contaminant transport analysis
• VADOSE/W climate influenced velocity data in a CTRAN/W
contaminant transport analysis
• SEEP/W with CTRAN/W density dependent flow – salt water
intrusion
• TEMP/W ground freezing including water flow affects computed from
SEEP/W
• SEEP/W seepage dependent embankment settlement coupling with
SIGMA/W

Page 181
Chapter 10: Product Integration Illustrations TEMP/W

• Uncoupled consolidation using seepage and pressure data from


SEEP/W, VADOSE/W, QUAKE/W

10.1 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 pore-
water 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 10-1. 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 182
TEMP/W Chapter 10: Product Integration Illustrations

26

24

22
Infiltration 5.0e-5 m/day
20

18 K(sat) = 1e-3 m/day 1


16 Seepage Face
K(sat) = 1e-5 m/day 2
14

12
K(sat) = 1e-3 m/day Pressure = zero
10
3
8

4
0 10 20 30 40 50 60

Figure 10-1 Seepage problem definition


The soil property information for the SLOPE/W portion of the analysis is given in
Table 10-1.

Table 10-1 Slope soil information

Layer Unit Weight Phi Cohesion Unit Weight Phi B


above WT
1 18 25 5 18 15
2 19 20 5 19 15
3 20 30 10 20 15

Figure 10-2 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 10-3 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 183
Chapter 10: Product Integration Illustrations TEMP/W

44

42
11
40

38

36

34

32 1
30
1.324
2
28 12 13

26

24
3
22
1 Infiltration 2
20

18
8 9
16
7 10
14

12
3 4
10
14
8

6 6 5

4
0 10 20 30 40 50 60

Figure 10-2 SLOPE/W solution showing perched water table from


SEEP/W
Pore-Water Pressure vs. Distance
30

25

20

15
Pore-Water Pressure

10

-5

-10

-15

-20
0 10 20 30 40

Distance

Figure 10-3 SEEP/W pore pressures at base of slices in SLOPE/W

10.2 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 slope’s stability.

Page 184
TEMP/W Chapter 10: Product Integration Illustrations

The VADOSE/W analysis mesh as well as the SLOPE/W grid and radius are
illustrated in Figure 10-1. 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 10-1. 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 10-2 Slope soil information

Layer Unit Weight Phi Cohesion Unit Weight Phi B


above WT
1 18 25 5 18 15

Figure 10-4 Stability problem definition (pwp from end of day 2 in


VADOSE/W)

Page 185
Chapter 10: Product Integration Illustrations TEMP/W

Figure 10-2 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 pore-
water pressures at the base of each slice.
Figure 10-6 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 10-7 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 10-8 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 186
TEMP/W Chapter 10: Product Integration Illustrations

1.159

Figure 10-5 SLOPE/W solution using water pressures from


VADOSE/W (solution including suction effect on strength)

0.999

Figure 10-6 SLOPE/W solution using water pressures from


VADOSE/W (solution NOT including suction effect on strength)

Page 187
Chapter 10: Product Integration Illustrations TEMP/W

1.739

Figure 10-7 Factor of safety after long drying period (e.g. day 30 pore-
water pressure data from VADOSE/W)

Pore-Water Pressure vs.


Distance
10
Pore-Water Pressure

-10

-20

-30

-40
0 1 2 3 4 5 6
Distance

Figure 10-8 Day 30 VADOSE/W pore pressures at base of slices in


SLOPE/W (for simulation with suction effects on strength)

Page 188
TEMP/W Chapter 10: Product Integration Illustrations

10.3 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 10-9 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 10-10. Note the duration of the
earthquake was only 40 seconds, yet during this time the pore pressures at point
“A” in Figure 10-11 increased significantly. The computed pore-water pressures at
this point during the entire earthquake are illustrated in Figure 10-12. In general,
the shaking almost doubled the pressure over pre-quake hydrostatic conditions.

1.14
1.11
1.08
(x 1000)

1.05
1.02
0.99
0.96
0.93
0.90
-500 -400 -300 -200 -100 0 100 200 300 400 500 600

Distance - feet
Figure 10-9 Dam pore-water pressure prior to earth quake

Page 189
Chapter 10: Product Integration Illustrations TEMP/W

Pacoima dam earthquake


0.6
0.5
0.4
0.3
Acceleration ( g )

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 10-10 Horizontal acceleration record

Figure 10-11 Pore-water pressure contours after earthquake. Note


point A

Page 190
TEMP/W Chapter 10: Product Integration Illustrations

Pore-Water Pressure vs. Time


9000

8000

Pore-Water Pressure

7000

6000

5000
0 5 10 15
Time

Figure 10-12 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
1.11 Seepage face (Q=0)
1.08 Head BC = 1110 ft
(x 1000)

1.05
A P=0
1.02
0.99
0.96
0.93
0.90
-500 -400 -300 -200 -100 0 100 200 300 400 500 600

Distance - feet

Figure 10-13 Model set up for SEEP/W pressure dissipation analysis

Page 191
Chapter 10: Product Integration Illustrations TEMP/W

Figure 10-13 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
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 10-14 below.

Pressure vs. Time


9000

8000
Pressure

7000

6000

5000
0 100000 200000 300000 400000 500000
Time

Figure 10-14 Dissipation of excess pore pressure at point “A” over 5


days following quake

10.4 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

Page 192
TEMP/W Chapter 10: Product Integration Illustrations

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”.
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 Θ is defined as a constant value of 0.5. The average linear velocity is:

ν =U /Θ
= 0.05 / 0.5
= 0.1 m/s

The dispersivity α L is set to 4 m, and the molecular diffusion coefficient D* is set


to zero. The resulting hydrodynamic dispersion coefficient is:

D = α 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:

Cs ⎡ ⎛ x - vt ⎞ ⎛ vx ⎞ ⎛ x - vt ⎞ ⎛ v( x +vt ) ⎞ ⎤
C = ⎢ erfc ⎜ ⎟ − exp ⎜ D ⎟ erfc ⎜ ⎟ ⎜1 + D ⎟ ⎥
2 ⎣ ⎝ 2 Dt ⎠ ⎝ ⎠ ⎝ 2 Dt ⎠ ⎝ ⎠⎦
v t ⎛ ( x − vt ) 2 ⎞
+ exp ⎜ − ⎟
πD ⎝ 4 Dt ⎠

Page 193
Chapter 10: Product Integration Illustrations TEMP/W

where:
C = concentration,
Cs = specified concentration of the source,
D = hydrodynamic dispersion coefficient,
v = average linear velocity,
t = elapsed time,
x = distance from the source boundary, and
erfc = complementary error function.

In this verification example, no adsorption and no decay are considered. Figure


10-15 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 10-16. 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 10-17 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 10-18 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 194
TEMP/W Chapter 10: Product Integration Illustrations

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

4.8000e+002
0.0
0 10 20 30 40
Distance

Figure 10-15 CTRAN/W solution of concentration versus distance at


various times using one free exit boundary

1.0
CTRAN/W Solution
(Exit Qd > 0)
0.9

0.8 Analytical Solution

0.7
Concentration

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 10 20 30 40

Distance

Figure 10-16 Comparison with analytical solutions for one exit


boundary condition

Page 195
Chapter 10: Product Integration Illustrations TEMP/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

4.8000e+002
0.0
0 10 20 30 40
Distance

Figure 10-17 CTRAN/W solution with two zero dispersive mass flux
exit boundary conditions

1.0
CTRAN/W Solution
(Exit Qd > 0)
0.9

0.8 CTRAN/W Solution


(Exit Qd = 0)

0.7
Concentration

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 10 20 30 40

Distance

Figure 10-18 Comparison with analytical solution for two exit


boundary conditions

Page 196
TEMP/W Chapter 10: Product Integration Illustrations

10.5 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
is assumed that a source of contaminant is also present with the water at the ground
surface.
Figure 10-19 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 10-19 Concentration contours without ground-climate coupling

Page 197
Chapter 10: Product Integration Illustrations TEMP/W

In Figure 10-20, 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.

Figure 10-20 Concentration contours with ground-climate coupling

10.6 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 O’Sullivan, (1995), noted that none of the published numerical
model comparisons with Henry’s solution that they examined were able to match
Henry’s solution to a great extent. In addition to outlining the possible reasons for
the discrepancies, Croucher and O’Sullivan, 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 10-21. It consists of a 2.0m long
section of a 1.0m thick aquifer where the right boundary is in direct contact with

Page 198
TEMP/W Chapter 10: Product Integration Illustrations

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.
Henry's Problem for Seawater Intrusion

Impermeable Top Boundary


1.0

0.9

0.8

0.7

Left Boundary 0.6 Right Boundary


Freshwater Seawater
Qf = 6.6E-5 m2/s 0.5
Hs = 1.0m
C=0.0 C = 1.0
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 Seawater Density Axes lengths in metres.


Ks = 1E-2 m/s SG=1.025 @ Cref=1.0 Element thickness = 1.0m
n = 0.35
D = 1.89E-5 m2/s

Figure 10-21 Henry's problem definition


The above Henry’s 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 10-22 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

Page 199
Chapter 10: Product Integration Illustrations TEMP/W

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 10-22 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
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 Seawater
C=0 C=1
1.00

0.75

0.50 2
0.
4
0.

6
0.

0.25
0.

0.00
0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00

Figure 10-22 Henry's problem verification - computed concentration


contours
Comparison of the CTRAN/W computed results with the highly accurate
numerical solution of Croucher and O’Sullivan is given in Figure 10-23. 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
O’Sullivan.

Page 200
TEMP/W Chapter 10: Product Integration Illustrations

0.9
Croucher and O'Sullivan, 1995

0.8 CTRAN/W

0.7

0.6
Elevation (m)

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 2
X-Coordinate (m)

Figure 10-23 Comparison of 0.5 concentration isochlors

10.7 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 10-24 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 201
Chapter 10: Product Integration Illustrations TEMP/W

1.6
2.5
1.4

1.2
Pipe

1
0
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 10-24 Temperature contours around freezing pipe without


flowing water
Figure 10-25 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 10-26 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 10-27 for a typical shaft sinking artificial ground freezing project.

Page 202
TEMP/W Chapter 10: Product Integration Illustrations

1.6
2.5
1.4

1.5
0
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 10-25 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:

v∆t
C=
∆x
where:
v = the Darcian velocity,
∆t = the time step, and
∆x = the element edge length.

Page 203
Chapter 10: Product Integration Illustrations TEMP/W

1.6
1.5
1.4
1.3
1.2
1.1 Pipe
1.0
0.9

1.6045
1.6035
1.6025
0.8

1.601
1.6005
0.7
0.6

5
0.5
0.4
0.3
0.2
0.1
0.0
-0.1
-0.2
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

Figure 10-26 Water flow vectors and head contours around frozen
pipe

Figure 10-27 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 204
TEMP/W Chapter 10: Product Integration Illustrations

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.

10.8 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 10-28 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 205
Chapter 10: Product Integration Illustrations TEMP/W

15

14
3
13

12 1
11
Elevation - metres

10
9

8
7

6
5

4
3

2
1

0
0 5 10 15 20 25 30 35 40

metres

Figure 10-28 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
10-3. 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
ν = 0.36
k = 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 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.

Page 206
TEMP/W Chapter 10: Product Integration Illustrations

Table 10-3 Summary of clay material properties

Property Value
Cc (compression index) 0.46
λ (slope of normal consolidation line) 0.20
κ (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
φ′ (effective friction angle) 26 degrees
Μ (slope of critical state line) 1.0
Ko (coefficient of earth pressure at rest) 0.56
ν (Poisson’s Ratio) 0.36
OCR (over consolidation ratio) 1.2
Γ (specific volume) 2.2

The analysis will be run for 40 days. Figure 10-29 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
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Figure 10-29 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 207
Chapter 10: Product Integration Illustrations TEMP/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 10-30, 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 10-31 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 pore-
water 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 208
TEMP/W Chapter 10: Product Integration Illustrations

Figure 10-30 SEEP/W boundary conditions


Figure 10-32shows 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 10-33. In addition,
the toe of the fill has been lifted up slightly.

Pore-Water Pressure vs. Time


110

90
Pore-Water Pressure

70

50

30

10
0 10 20 30 40
Time

Figure 10-31 Pore-water pressure changes beneath embankment


versus time

Page 209
Chapter 10: Product Integration Illustrations TEMP/W

Y-Displacement vs. X
0.2

0.0000e+000
0.0

Y-Displacement
7.0000e+000
-0.2

1.3000e+001
-0.4

1.9000e+001
-0.6

2.5000e+001
-0.8
0 5 10 15 20 25 30
X

Figure 10-32 Ground surface settlement profiles at various times

30
25

25

30

Figure 10-33 Horizontal effective stress after second lift placed

10.9 Uncoupled consolidation


SIGMA/W can be used with SEEP/W, VADOSE/W, QUAKE/W and other
SIGMA/W projects in order to computed uncoupled consolidation. In this type of
analysis, the change in pore-water pressure at each time step is obtained from the

Page 210
TEMP/W Chapter 10: Product Integration Illustrations

external analysis and used in the solution of the consolidation equations inside
SIGMA/W.
More details of this type of analyses will be incorporated into this book in the next
edition. Specific questions can be addressed to support@geo-slope.com in the
interim.

Page 211
Chapter 10: Product Integration Illustrations TEMP/W

Page 212
TEMP/W Chapter 11: Illustrative Examples

11 Illustrative Examples
This chapter presents a variety of verification and typical real life example
problems analyzed with TEMP/W. The real life examples are not case studies as
you would find them in a journal, but they can be used as a starting point for
setting up your own models.
The verification examples compare TEMP/W solutions with benchmark references
that show that the software is functioning properly. The real life examples help to
illustrate how some of the unique features in TEMP/W are applied to practical
situations.
A point to keep in mind while reviewing these examples is that they are set up and
solved using techniques and parameters that help illustrate the use of the software
in applying the theory. It is possible, likely, and recommended, that you will need
to adjust your unique parameters, in order to fine tune your own analysis and gain
confidence in the results and your understanding of the processes being modeled.

11.1 Steady-state heat flow analysis


The first verification example is a simple one-dimensional, steady-state heat flow
analysis problem. This example problem can be used to test various fundamental
aspects of the heat flow equation implemented in TEMP/W.
Closed-form solutions for this type of simple analysis can be obtained using the
basic flow law. The example problem is first solved numerically with TEMP/W
and then analytically using spreadsheet (Microsoft Excel) calculations.

Problem definition
Figure 11-1 illustrates the geometry and the finite element discretization of the
verification problem. The soil column is 10 meters long and 1 meter wide. The
column is discretized using 10 rectangular elements and 53 nodes. The soil
material is homogeneous and has a constant thermal conductivity of
1.0 J/(s•m•°C). The unfrozen water content function is not required in a steady-
state analysis and therefore is not defined for the analysis.
The boundary conditions are as follows:

Page 213
Chapter 11: Illustrative Examples TEMP/W

Left and right boundaries, constant temperature, T = 5.0°C and T = -5.0°C,


respectively.
Top and bottom boundaries, no flow, Q = 0.

2 Q=0
3 5 8 10 13 15 18 20 23 25 28 30 33 35 38 40 43 45 48 50 53
1
o
T=-5 o C
2 7 12 17 22 27 32 37 42 47 52
T=5 C 1
1
4 6
2
9 11
3
14 16
4
19 21
5
24 26
6
29 31
7
34 36
8
39 41
9
44 46
10
49 51
0

Q=0
-1

-2
-2 -1 0 1 2 3 4 5 6 7 8 9 10 11 12

Distance
Figure 11-1 One-dimensional steady-state heat flow problem

Solution
Based on the one-dimensional flow law for heat transfer by conduction, the heat
flux through the soil column at steady-state can be expressed as:

∆T
Q = −k A
∆x
where:
Q = heat flux,
k = thermal conductivity,
T = temperature,
x = distance, and
A = flow area.

The negative sign in the equation indicates that heat flux is flowing from a high
temperature to a low temperature (i.e., from left to right in this example).

Page 214
TEMP/W Chapter 11: Illustrative Examples

From the problem definition, the parameters on the right hand side of the equation
are as follows:
k = 1.0 J/(s•m•°C),
∆T = -10.0°C,
∆x = 10.0 m, and
A = 1.0 m2

Since the above quantities are constants, the resulting heat flux across the soil is
also a constant and is calculated to be 1.0 J/s using the heat flow equation. The
constant heat flux is consistent with the principle of conservation of energy (i.e.,
that the heat flux in and out of the soil column must be the same at steady-state).
Figure 11-2 presents the computed solution of the flow problem using TEMP/W
(contained in the VERIFY files). The direction of the flow vectors in the figure
indicates that heat flux is flowing from left to right. The sizes (magnitude) of the
vectors are identical, indicating that the heat flux is a constant through out the
entire soil column. Furthermore, the heat flux across the soil column is computed
to be 1.0 J/s which is exactly the same as the analytical solution.
1.0000e+000

2
Freeze-Thaw Line
1
T=5 oC T=-5 o C
0

-1

-2
-2 -1 0 1 2 3 4 5 6 7 8 9 10 11 12

Distance
Figure 11-2 TEMP/W solution of the one-dimensional heat flow
problem
Rearranging the heat flow equation, the steady-state temperature profile of the one-
dimensional flow problem can be expressed as:

Page 215
Chapter 11: Illustrative Examples TEMP/W

Qx
T = 5−
kA
Since Q, k, and A are all equal to 1.0, the equation can be simplified as:

T = 5− x
This suggests that the temperature profile along the column decreases linearly from
left to right.
Figure 11-3 presents the computed temperature profile using TEMP/W. The
computed temperature profile is consistent with the results from the closed form
equation, as illustrated in Figure 11-4.

Temperature vs. Distance


5

-1

-2

-3

-4

-5
0 2 4 6 8 10
Distance

Figure 11-3 Temperature profile from TEMP/W

Page 216
TEMP/W Chapter 11: Illustrative Examples

4
A n a ly tic a l S o lu tio n
3
T E M P /W S o lu tio n
2

1
Temperature

-1

-2

-3

-4

-5
0 2 4 6 8 10

D is ta n c e

Figure 11-4 Comparison of analytical and TEMP/W solution

11.2 Neumann thaw and freeze analyses


The second verification example is the thawing and freezing analyses of a one-
dimensional soil column. Graphical and closed-form solutions for these types of
problems are available when a step temperature is applied at the surface. The thaw
and freeze depths can be computed using a Neumann solution (Nixon and
McRoberts, 1973), when the following parameters are defined:
Ku = thermal conductivity of unfrozen soil,
Kf = thermal conductivity of frozen soil,
Cu = volumetric heat capacity of unfrozen soil,
Cf = volumetric heat capacity of frozen soil,
L = volumetric latent heat,
Tg = uniform initial ground temperature, and
Ts = applied constant surface temperature.

Page 217
Chapter 11: Illustrative Examples TEMP/W

Problem definition
Figure 11-5 illustrates the problem and presents the parameters selected for
verifying TEMP/W.

Figure 11-5 One dimensional thaw and freeze problem


In this example, the thermal conductivity, (K), and volumetric heat capacity, (C),
are the same for the frozen and unfrozen zones, and the volumetric water content
is 1.0. The latent heat, (L), is therefore equal to the latent heat of water.
The closed-form solution assumes that all the latent heat is absorbed or liberated at
a temperature of 0°C. To simulate such a stepped function in TEMP/W, the
unfrozen water content function is defined as a near-perfect stepped function. For
temperatures below -0.001°C, Θ r = 0 ; for temperatures above 0°C, Θ r = 1 . This
is an extremely steep function; consequently, the numerical solution to this
problem may require a large number of iterations.

Solution
The closed-form solution for the advance of the thawing and freezing front for this
problem was derived by Neumann about 1860 and is given by Carslaw and Jaeger
(1947) as:

Page 218
TEMP/W Chapter 11: Illustrative Examples

X =α t

where:
X = depth of the thawing or freezing fronts,
t = elapsed time, and
α = a constant which is a function of material properties and boundary
conditions.

The relationship of α to the various parameters is:

⎧ ⎡ TK ⎛k
1/ 2
⎞ ⎤⎫
α ⎪ ⎢− g f ⎥ ⎪⎬
= f ⎨ ste, ⎜⎜ u ⎟⎟
2(ku )1/ 2 ⎢ Ts K u ⎝ kf ⎥⎪
⎩⎪ ⎣ ⎠ ⎦⎭
where Ste, the Stefan number, is defined as the ratio of sensible heat to latent heat
by:

Cu Ts
Ste =
L
The parameters ku and kf are the diffusivities of unfrozen and frozen soil, and are
defined as the thermal conductivity divided by the volumetric heat capacity. In
equation form:

Ku
ku =
Cu
Kf
kf =
Cf

Nixon and McRoberts have presented a graphical solution to the Neumann


equation, as shown in Figure 11-6.

Page 219
Chapter 11: Illustrative Examples TEMP/W

Figure 11-6 Solution of the Neumann Equation (after Nixon and


McRoberts)

Depth of thawing front


For this example, since ku equals kf and Ku equals Kf, the parameter
1/ 2
T K ⎛k ⎞ Tg
− g f ⎜ u ⎟⎟ is reduced to a ratio of − . In the case of thawing, Ts = 5°C
Ts K u ⎜⎝ k f ⎠ Ts
and Tg = -3°C. Therefore, the ratio is:

Tg − ( −3)
− = = 0.6
Ts 5

The variable Ste is calculated as:

Page 220
TEMP/W Chapter 11: Illustrative Examples

Cu Ts ( 2.0 )( 5 )
Ste = = = 0.03
L 334
α
Using these two known variables, the normalized thaw parameter can be
2 ku
obtained from the graph in Figure 11-6 as 0.1. We can then calculate α as:

α = 2 ku ( 0.1) = 2 0.05 ( 0.1) = 0.045

With α known, the depth of the thawing front can be computed.

This problem is simulated in TEMP/W with 35 elements and 178 nodes. A


TEMP/W solution to the problem is also obtained by using an infinite element at
the bottom boundary. The initial temperature of the entire column is defined at -
3°C using the Draw Initial Conditions command in DEFINE. The ground surface
is specified as a constant temperature boundary, T = 5°C. The associated files for
this example are named thaw and thaw_infinite.
Figure 11-7 compares the analytical solution, (Neumann solution), with the
TEMP/W computed results. The solid line is the analytical solution and the data
points represent the thaw depth computed by TEMP/W. As indicated in the
comparison, excellent agreement exists between the graphical Neumann solution
and the TEMP/W results.
The TEMP/W solution using the infinite element appears to compare even closer
to the analytical solution than the solution without the infinite element, especially
after a large elapsed time. This is consistent with the fact that the analytical
solution is developed for a semi-infinite column, which can be better simulated
with an infinite element at the bottom boundary.

Page 221
Chapter 11: Illustrative Examples TEMP/W

1 .6

1 .4

1 .2
THAW DEPTH (m)

0 .8

0 .6
A n a ly t ic a l S o lu t io n

0 .4
T E M P / W S o lu t io n
( N o I n f in it e E le m e n t )

0 .2 T E M P / W S o lu t io n
( W it h I n f in it e E le m e n t )

0
0 200 400 600 800 1000 1200

T IM E ( D a y s )

Figure 11-7 Comparison of the Neumann and TEMP/W results for thaw
depth vs time

Depth of freezing front


The one-dimensional thaw problem in the previous section is repeated here with
the initial ground temperature at +3°C and the initial surface temperature at -5°C.
In all other respects, this problem is identical to the thaw problem in the previous
section.
In the case of freezing, Ts = -5°C and Tg = 3°C. Therefore, the -Tg/Ts ratio is:

Tg − ( 3)
− = = 0.6
Ts −5

The variable Ste is calculated as:

Cu Ts ( 2.0 )( −5 )
Ste = = = 0.03
L −334

Page 222
TEMP/W Chapter 11: Illustrative Examples

α
Using these two known variables, the normalized thaw parameter can be
2 ku
obtained from the graph in Figure 11-6 as 0.1. We can then calculate α as:

α = 2 ku ( 0.1) = 2 0.05 ( 0.1) = 0.045

With α known, the depth of the thawing front can be computed.

The problem is also simulated in TEMP/W with 35 elements and 178 nodes. The
TEMP/W solution to the problem is also obtained using an infinite element along
the bottom boundary. The initial temperature of the entire column is defined at 3°C
using the Draw Initial Conditions command in DEFINE. The ground surface is
specified with a constant temperature boundary, T = -5°C. The associated files for
this example are named freeze and freeze_infinite.
The solutions for freezing and thawing must be identical when the material
properties for both frozen and unfrozen soils are the same. Figure 11-8 shows the
frozen depth versus time functions for the Neumann and TEMP/W solutions. A
comparison of Figure 11-7 and Figure 11-8 reveals that the computed solutions are
identical for both the thawing and freezing cases. This verifies that TEMP/W is
working properly for both thawing and freezing simulations.

Page 223
Chapter 11: Illustrative Examples TEMP/W

1 .6

1 .4

1 .2
FREEZE DEPTH (m)

0 .8

0 .6
A n a ly tic a l S o lu tio n

0 .4
T E M P / W S o lu t io n
( N o In fin ite E le m e n t)

0 .2 T E M P / W S o lu t io n
( W it h I n f in it e E le m e n t )

0
0 200 400 600 800 1000 1200

T IM E (D a y s )

Figure 11-8 Comparison of Neumann and TEMP/W results for freeze


depth vs time

11.3 Hwang finite element analysis


Hwang, Murray, and Brooke, (1972), were among the first to develop a geothermal
finite element model that included the effects of latent heat. The third verification
example problem compares Hwang's finite element solution with the TEMP/W
results for a one-dimensional thaw analysis similar to the Neumann thaw and
freeze problem in this chapter.

Problem definition
Hwang used a finite element mesh that was 5 meters high with an initial ground
temperature of -2°C and an initial surface temperature of +5°C. The material
properties used in the analysis are given in Table 11-1.

Page 224
TEMP/W Chapter 11: Illustrative Examples

Table 11-1 Material properties used in Hwang analysis

Material Property Value


3
Thermal conductivity 1x10 cal/m/h/°C
Specific heat capacity 1 cal/gm/°C
Unit weight (density) 1x103 kg/m3
Volumetric heat capacity 1x106 cal/m3/°C
Latent heat 5x107 cal/m3
Volumetric water content 1.0 (100%)

Hwang adopted the same parameters for both the frozen and the unfrozen zones.
The thaw depth at different times are tabulated in Table 11-2.

Solution
The thaw depth of the Hwang example problem can also be calculated using the
graphical solution presented by Nixon and McRoberts in Figure 11-6. Since Ts =
5°C and Tg = -2°C, the (-Tg/Ts) ratio is therefore:

Tg − ( 5)
− = = 2.5
Ts −2

The variable Ste is calculated as:

Cu Ts (1x10 ) ( 5 )
6

Ste = = = 0.1
L 5 x107
α
Using these two known variables, the normalized thaw parameter can be
2 ku
obtained from the graph in Figure 11-6 as 0.2. We can then calculate α as:

⎛ 103 ⎞
α = 2 ku ( 0.2 ) = 2 ⎜ 6 ⎟ (
0.2 ) = 0.0126
⎝ 10 ⎠

With α known, the depth of the thawing front can be computed.

Page 225
Chapter 11: Illustrative Examples TEMP/W

The example problem is simulated in TEMP/W using 27 elements and 138 nodes.
The initial temperature of the entire column is defined at -2°C using the Draw
Initial Conditions command. The ground surface is specified with a constant
temperature boundary (T) of 5°C. Table 11-2 compares the thaw depths calculated
from the analytical solution, Hwang’s finite element solution, and the TEMP/W
solution for three different times. The associated files for this example are named
Hwang.

Table 11-2 Comparison of calculated thaw depth from different procedures

Elapsed Time Thaw Depth (m)

In Hours In Days Analytical Hwang TEMP/W


126 5 0.14 0.1 0.14
1580 66 0.50 0.4 0.49
6310 263 1.00 0.9 0.92

The thaw depths computed by TEMP/W compare satisfactorily with both the
analytical and Hwang solutions. Hwang’s thaw depths are only available
graphically and consequently are approximated to only one-tenth of a meter.
Figure 11-9 compares the temperature profiles obtained by Hwang and by
TEMP/W. Good agreement is achieved between Hwang’s solution and TEMP/W
solution. Except for the small differences in the temperature profiles for the last
time step (6310 hours), the temperature distribution is essentially the same for both
finite element formulations.
TEMP/W gives the same results as Hwang obtained using a finite element
analysis. This is important because the procedure that Hwang used to formulate the
latent heat component is different than the procedure used by TEMP/W. Hwang
used the rate of change of frozen area to account for the latent heat, while
TEMP/W uses the slope of the unfrozen water content function. The unfrozen
water content function formulation is considered more appropriate, since it is more
general.
The TEMP/W approach is preferred for practical field problems, since it can
accommodate any gradual function; the Hwang approach can only accommodate
the equivalent of a vertical stepped unfrozen water content function.

Page 226
TEMP/W Chapter 11: Illustrative Examples

Figure 11-9 Comparison of Hwang and TEMP/W temperature profiles

11.4 Two-dimensional permafrost analysis


Hwang, Murray, and Brooke, (1972), presented an example to show the
degradation of permafrost beneath a warm building. The same example is analyzed
using TEMP/W.

Problem definition
The in-situ ground temperature varies from -2°C at the surface to 0°C at a depth of
60 m, as shown in Figure 11-10. The building is 80 m wide; only half of the

Page 227
Chapter 11: Illustrative Examples TEMP/W

building (40 m) is used in the analysis. The building temperature is 15.5°C and the
average surface temperature outside the building is -1.94°C.

Figure 11-10 Definition of permafrost degradation analysis


The material properties used in the analysis are listed in Table 11-3:

Page 228
TEMP/W Chapter 11: Illustrative Examples

Table 11-3 Material properties used in permafrost degradation analysis

Material Property Frozen Unfrozen Units


3 3
Thermal Conductivity 2x10 1x10 cal/(hr•m•°C)
Specific Heat 0.29 0.48 cal/(gm•°C)
Specific Heat 290 480 cal/(kg•°C)
Unit Weight 1310 1310 kg/m3
Volumetric Heat Capacity 3.8x105 6.3x105 cal/(m3•°C)
Latent Heat 8x107 – cal/m3
Volumetric Water Content
0.18 0.18 –
(assumed)

Since Hwang does not record the volumetric water content and handles the latent
heat (due to phase changes) differently than TEMP/W, it is necessary to make
some assumptions. Setting the latent heat to the latent heat of water and setting the
volumetric water content to 0.18 produces results that are similar to Hwang's
solution. With these values, the thaw depth at the center line is similar for the two
finite element analyses. Since the thaw depths at the center line are the same, the
shape of the freeze-thaw interface and the subsurface ground temperature profiles
can be compared.

Solution
The problem is simulated in TEMP/W with 250 elements and 661 nodes. The
initial temperature of the material is established by running a steady-state analysis
with the in-situ boundary condition as illustrated in Figure 11-10. The associated
files for this example are named perma_init and perma. Figure 11-11 illustrate the
TEMP/W computed temperature profile along the centerline of the problem for 5
time steps.

Page 229
Chapter 11: Illustrative Examples TEMP/W

Temperature vs. Depth


60

50 0.0000e+000

40 1.0000e+004

Depth (m)
30 1.0000e+005

20 2.2000e+005

10 2.5000e+005

0 4.0000e+005
-5 0 5 10 15 20
Temperature

Figure 11-11 Computed temperature profile along center line


Table 11-4 compares the computed thaw depths at the centerline of the problem for
the two finite element methods. Good agreement is obtained.

Table 11-4 Comparison of computed thaw depths

Elapsed Time Thaw Depth at Centerline (m)

In Hours In Years Hwang TEMP/W


4
1.0x10 1.1 4 3.8
1.0x105 11 13 13.0
2.2x105 25 19 19.3
5
2.5x10 29 20 20.4
4.0x105 46 24 26.0

Page 230
TEMP/W Chapter 11: Illustrative Examples

Figure 11-12 compares the computed freeze-thaw interface lines, and Figure
11-13compares the computed temperature profiles at the center line of the
problem. As illustrated in these figures, the computed results from TEMP/W give
satisfactory comparison to those computed independently by Hwang.

Figure 11-12 Comparison of Hwang and TEMP/W freest-thaw interface


lines

Figure 11-13 Comparison of Hwang and TEMP/W ground


temperatures along centerline

Page 231
Chapter 11: Illustrative Examples TEMP/W

11.5 Pipeline freezing analysis


Coutts and Konrad (1994) presented the solution of a freezing analysis of a buried
pipeline using the “node state” finite element method. This last verification
example attempts to compare the TEMP/W solution with Coutts and Konrad’s
solution.

Problem definition
In this example problem, a vertical cross-section perpendicular to a 30 cm pipe
buried 30 cm below the ground surface is considered. Figure 11-14 presents the
TEMP/W finite element discretization of the vertical cross-section. Symmetry is
assumed and the soil domain is modeled using 322 elements and 327 nodes.

1.7
Centerline Ground Surface
1.6
1.5
1.4
1.3
1.2 o
1.1 Pipe T initial = 3 C
1.0
T ground surface = 3 oC
Depth (m)

0.9
0.8
0.7 o
0.6
T pipe = -2 C
0.5
0.4
0.3
0.2
0.1
0.0
-0.1
-0.2
-0.3
-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

Distance (m)

Figure 11-14 Problem definition for pipeline freezing analysis

Page 232
TEMP/W Chapter 11: Illustrative Examples

Table 11-5 Material properties used in pipeline freezing analysis

Material Property Frozen Unfrozen Units


Thermal Conductivity 0.15552 0.12960 MJ/(sec.m.°C)
Volumetric Heat Capacity 1.95 1.95 MJ/(m3.°C)
Phase Change Temperature Range 0.2 – °C
Phase Change Temperature 0 – °C
Latent Heat 334 – MJ/m3
Volumetric Water Content 0.3772 0.3772 –

The in-situ ground temperature throughout the cross-section is assumed to be at


+3°C. During the transient process, the pipe boundary and the ground surface are
maintained at constant temperatures of -2°C and +3°C respectively. All other
boundaries are set as zero flux boundary conditions.
Table 11-5 shows the material properties that are used in the analysis:
The phase change temperature range is the total change in temperature required for
the unfrozen water content to decrease from 1.0 to 0. In other words, with the
phase change temperature at 0°C and a phase change temperature range of 0.2°C,
the unfrozen water contents are 0.0 and 1.0 for temperatures below -0.1°C and
above 0.1°C, respectively.

Solution
The problem is simulated in TEMP/W for up to 730 days, (2 years). The initial
temperature of the ground is defined at 3°C using the Draw Initial Conditions
command in DEFINE. The ground surface is specified with a constant temperature
boundary, T = 3°C and the pipe is specified with a constant temperature of T = -
2°C. The Table 11-6 compares the computed positions of the freezing front after
730 days for the two finite element methods. Excellent agreement is obtained.

Page 233
Chapter 11: Illustrative Examples TEMP/W

Table 11-6 Comparison of freezing front after 730 days

Position Freezing Front from Pipe (m)

Coutts and Konrad TEMP/W


Below pipe 0.60 0.58
Right side of pipe 0.23 0.22

Figure 11-15 shows the computed temperature profile of the soil medium after
730 days, as presented by Coutts and Konrad. Figure 11-16 shows the TEMP/W
computed temperature profile of the soil medium after 730 days. As is evident in
these figures, the two finite element methods give essentially the same results.

Figure 11-15 Temperature profile after 2 years (Coutts and Konrad,


1994)

Page 234
TEMP/W Chapter 11: Illustrative Examples

-2
-1

1
0

Figure 11-16 TEMP/W profile after 2 years

11.6 Convective tile cooling


The convective heat transfer boundary condition option can be used for many types
of thermal analysis. Originally implemented for use in artificial ground freezing, it
is used in this example to illustrate the rate of cooling of a heated Pyrex tile.

Problem definition
A tile of Pyrex material that is 200 mm square and 10mm thick emerges from a
curing process at a uniform temperature of 140°C. The back side of the tile is
insulated while the upper surface is exposed to ambient air temperature at 25°C.
The ambient air is moving over the tile at a rate of 10 m/s, but in this example, still
air is assumed. The objective is to determine how long it takes the tile to cool to
40°C.

Page 235
Chapter 11: Illustrative Examples TEMP/W

Figure 11-17 Tile cooling problem definition and material properties

Solution
The problem is simulated in TEMP/W for 3440 seconds. The data file for this
example is named Tile_cooling.gsz. The initial temperature of the tile is defined at
140°C using the Draw Initial Conditions command in DEFINE.
The adaptive time stepping routine was selected, as it is assumed there will be a
high thermal gradient as cooling progresses. This option will limit the time steps if
the ones specified are too large for any given time step.
The comparison of TEMP/W with a closed form analytical solution is presented in
Figure 11-18. The TEMP/W solution was obtained for node 128 of the mesh,
which is at the base of the tile where it is in contact with the insulation.

Page 236
TEMP/W Chapter 11: Illustrative Examples

Figure 11-18 Comparison of TEMP/W with closed form analytical


solution

Page 237
Chapter 11: Illustrative Examples TEMP/W

Page 238
TEMP/W Chapter 12: Theory

12 Theory
This chapter describes the theoretical engineering basis for the TEMP/W program.
More specifically, it deals with the fundamental flow laws for steady-state and
transient flow and it shows how these laws are represented in numerical form.

12.1 Flow law


The principal mechanism for heat flow in soils in most engineering applications
involving freezing and thawing is conduction. (see Harlan and Nixon, 1978)
Conduction is the flow of heat by the passage of energy from one soil particle to
another or through soil pore fluids. An option for convective heat transfer is also
available in TEMP/W and is discussed towards the end of this chapter.
TEMP/W assumes that heat flux due to conduction is governed by the following
equation:

∂T
Equation 12-1 q = −k
∂x
where:
q = the heat flux,
k = the thermal conductivity,
T = the temperature, and
x = distance.

Equation 12-1 shows that heat flow due to conduction is directly dependent on the
thermal conductivity of the soil medium and temperature gradient. The negative
sign indicates that the temperature decreases in the direction of increasing x; that
is, the heat flows in the direction from high temperature to low temperature.

12.2 Governing equations


The governing differential equation used in the formulation of TEMP/W is:

Page 239
Chapter 12: Theory TEMP/W

∂ ⎛ ∂T ⎞ ∂ ⎛ ∂T ⎞ ∂T
⎜ kx ⎟+ ⎜ ky ⎟ +Q=λ
∂x ⎝ ∂x ⎠ ∂y ⎝ ∂y ⎠ ∂t

where:
T = temperature,
kx = thermal conductivity in the x-direction,
ky = thermal conductivity in the y-direction,
Q = applied boundary flux,
λ = capacity for heat storage, and
t = time.

This equation states that the difference between the heat flux entering and leaving
an elemental volume of soil at a point in time is equal to the change in the stored
heat energy.
Under steady-state conditions, the flux entering and leaving an elemental volume is
the same at all times. The right side of the equation vanishes and the equation
reduces to:

∂ ⎛ ∂T ⎞ ∂ ⎛ ∂T ⎞
⎜ kx ⎟+ ⎜ ky ⎟ +Q=0
∂x ⎝ ∂x ⎠ ∂y ⎝ ∂y ⎠

The capacity to store heat is composed of two parts. The first part is the volumetric
heat capacity of the material (either frozen or unfrozen) and the second part is the
latent heat associated with the phase change. In equation form:

∂wu
λ =c+L
∂T
where:
c = volumetric heat capacity (material property),
L = latent heat of water,
Θu = total unfrozen volumetric water content,
T = temperature, and

Page 240
TEMP/W Chapter 12: Theory

λ = capacity for heat storage.

The volumetric heat capacity c is the slope of the energy curve in the frozen and
∂wu
unfrozen zones while the term L represents the rate of change of the latent
∂T
heat component.
When the unfrozen water content function of a soil is defined, the total unfrozen
volumetric water content can be expressed as:

wu = Wu w

where:
Θr = unfrozen water content (0 ≤ Θr ≤ 1), and
Θ = volumetric water content of the soil.

Substituting for Θu for λ in the main thermal equation leads to the complete
differential equation:

∂ ⎛ ∂T ⎞ ∂ ⎛ ∂T ⎞ ⎛ ∂Wu ⎞ ∂T
⎜ kx ⎟+ ⎜ ky ⎟ + Q = ⎜ c +Lw
∂T ⎟⎠ ∂t
Equation 12-2
∂x ⎝ ∂x ⎠ ∂y ⎝ ∂y ⎠ ⎝

12.3 Finite element heat flow equations


Applying the Galerkin method of weighed residual to the governing differential
equation, the finite element for two-dimensional thermal equation can be derived
as:
Equation 12-3

τ∫ ([B] [C ] [B]) dA {T } + τ ∫ ( λ <N >


A
T
A
T
<N > ) dA {T },t

= qτ ∫ ( <N > ) dL T
L

where:
[B] = the gradient matrix,

Page 241
Chapter 12: Theory TEMP/W

[C] = the element thermal conductivity matrix,


{T} = the vector of nodal temperatures,
<N> = the vector of interpolating function,
q = the unit flux across the edge of an element,
τ = the thickness of an element,
t = time,
λ = the capacity for heat storage,
A = a designation for summation over the area of an element, and
L = a designation for summation over the edge of an element.

In an axisymmetric analysis the equivalent element thickness is the circumferential


distance at different radii, R about the symmetric axis. The complete
circumferential distance is 2π radian times R , since TEMP/W is formulated for
one radian, the equivalent thickness is R . Therefore, the finite element equation
for the axisymmetric case is:

∫ ([B] [C ] [B]R ) dA {T } + ∫ ( λ <N > <N >R ) dA {T },t


T T
A A

= q ∫ ( <N > R ) dL
T
L

Note that the radial distance R is not a constant within an element as in the case of
the thickness τ in the two-dimensional analysis, consequently, R is a variable
inside the integral.
In an abbreviated form, the finite element thermal equation can be expressed as:

Equation 12-4 [ K ]{T } + [ M ]{T }, t = {Q}

where:
[K] = the element characteristic matrix,
[M] = the element heat matrix, and
{Q} = the element applied flux vector.

Page 242
TEMP/W Chapter 12: Theory

Equation 12-4 is the general finite element equation for a transient thermal
analysis. For a steady-state analysis, the head is not a function of time and
consequently the term {T }, t vanishes, reducing the finite element equation to:

[ K ]{T } = {Q}

which is the abbreviated finite element form of the fundamental flow equation.

12.4 Temporal integration


The finite element solution for a transient analysis is a function of time as indicated
by the {T }, t term in the finite element equation. The time integration can be
performed by a finite difference approximation scheme. Writing the finite element
equation in terms of finite differences leads to the following equation (see
Segerlind, 1984, pp. 183-185):

(ω∆t [ K ] + [ M ]) {T } =
1

∆t ( (1 − ω ){Q } + ω {Q } ) + ([ M ] − (1 − ω ) ∆t [ K ]) {T }
0 1 0

where:
t = the time increment,
ω = a ratio between 0 and 1,

T1 = the temperature at end of time increment,

T0 = the temperature at start of time increment,

Q1 = the nodal flux at end of time increment, and

Q0 = the nodal flux at start of time increment.

TEMP/W uses the Backward Difference Method, a method that sets ω to 1.0, the
finite element equation is then simplified to:

Equation 12-5 ( ∆t [ K ] + [ M ]) {T } = ∆t {Q } + [ M ]{T }


1 1 0

As indicated by Equation 12-5, in order to solve for the new temperature at the end
of the time increment, it is necessary to know the temperature at the start of the

Page 243
Chapter 12: Theory TEMP/W

increment. Stated in general terms, the initial conditions must be known in order to
perform a transient analysis.

12.5 Numerical integration


TEMP/W uses Gaussian numerical integration to evaluate the element
characteristic matrix [ K ] and the heat matrix [ M ] . The integrals are evaluated by
sampling the element properties at specifically defined points and then summed
together for the entire element.

Using the characteristic matrix [ K ] as an example, the following integral (from


Equation 12-3):

[K ] = τ ∫
A
([B]
T
[C ] [B ]) dA

can be replaced by:


n
[ K ] = τ ∑ [B j ]T [C j ] det |J j | W1 jW2 j
j =1

where:
j = an integration point,
n = the number of integration points,
[Cj] = the element thermal conductivity matrix at the integration point,
[Bj] = the element matrix at the integration point,
det|Jj| = the determinant of the Jacobian matrix, and
W1j = a weighting factor.

The number of sample (integration) points required in an element depends on the


number of nodes and the shape of the elements. The tables below contain the
number and location of sampling points that are used by TEMP/W.

Page 244
TEMP/W Chapter 12: Theory

Table 12-1 Sample point locations and weightings for four-point quadrilateral
element

Point r s w1 w2
1 +0.57735 +0.57735 1.0 1.0
2 -0.57735 +0.57735 1.0 1.0
3 -0.57735 -0.57735 1.0 1.0
4 +0.57735 -0.57735 1.0 1.0

Table 12-2 Sample point locations and weightings for nine-point quadrilateral
elements

Point r s w1 w2
1 +0.77459 +0.77459 5/9 5/9
2 -0.77459 +0.77459 5/9 5/9
3 -0.77459 -0.77459 5/9 5/9
4 +0.77459 -0.77459 5/9 5/9
5 0.00000 +0.77459 8/9 5/9
6 -0.77459 0.00000 5/9 8/9
7 0.00000 -0.77459 8/9 5/9
8 +0.77459 0.00000 5/9 8/9
9 0.00000 0.00000 8/9 8/9

Table 12-3 Sample point locations and weighting for one-point triangular
element

Point r s w1 w2
1 0.33333 0.33333 1.0 0.5

Page 245
Chapter 12: Theory TEMP/W

Table 12-4 Sample point locations and weightings for three-point triangular
element

Point r s w1 w2
1 0.16666 0.16666 1/3 1/2
2 0.66666 0.16666 1/3 1/2
3 0.16666 0.66666 1/3 1/2

One-point integration for a triangular element results in a constant gradient


throughout the element. The number of integration points is denoted as the
integration order. The appropriate integration order is a function of the presence of
secondary nodes. When secondary nodes are present, the interpolating functions
are nonlinear and consequently a higher integration order is required. Table 12-5
gives the acceptable integration orders.

Table 12-5 Acceptable element integration orders

Element Type Secondary Nodes Integration Order


Quadrilateral no 4
Quadrilateral yes 4 or 9
Triangular no 1 or 3
Triangular yes 3

It is also acceptable to use four-point integration for quadrilateral elements that


have secondary nodes. This is called a reduced integration order (see Bathe, 1982,
p. 282). Acceptable results can be obtained with reduced integration. For example,
reduced integration is useful in unfrozen zones where the thermal conductivity is
constant. Selective use of reduced integration can greatly reduce the required
number of computations.
It is also possible to use three-point and nine-point integration with elements that
have no secondary nodes. However, the benefits of this are marginal, particularly
for quadrilateral elements. Nine-point integration for a quadrilateral element
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. One-point integration
means the material properties and flow gradients are constant within the element.

Page 246
TEMP/W Chapter 12: Theory

This can lead to poor performance of the element, particularly if the element is in a
frozen zone where the thermal conductivity varies sharply with changes in
temperature. 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 one-point
integration in triangular elements with secondary nodes is not acceptable.
In general, it is sufficient to use three-point integration for triangular elements and
four-point integration for quadrilateral elements. In situations where there is
unsaturated zone with thermal conductivity varies sharply within an element, it is
best to use quadrilateral elements with secondary nodes together with nine-point
integration.

12.6 Thermal conductivity matrix


The general form of the TEMP/W element thermal conductivity matrix is:

⎡C C12 ⎤
[C ] = ⎢ 11 ⎥
⎣C21 C22 ⎦
where:

C11 = kx cos 2 α + ky sin 2 α

C22 = kx sin 2 α + ky cos 2 α


C12 = kx sin α cos α + ky sin α cos α
C21 = C12

The parameters k x , k y and α are defined in Figure 12-1.

Page 247
Chapter 12: Theory TEMP/W

Figure 12-1 Definition of thermal conductivity matrix parameters

When α is zero, [C ] is reduced to:

⎡kx 0⎤
[C ] = ⎢
⎣0 k y ⎥⎦

The parametric k x is always determined from the thermal conductivity function.


Parameter k y is then computed as k x multiplied by k Ratio . In equation form:

k y =k x × k Ratio

12.7 Heat storage matrix


As first presented in Equation 12-3 the element heat (or storage) matrix for a two-
dimensional analysis is defined as:

[M ] = τ ∫
A
(λ < N > T
)
< N > dA

Similar to the element characteristic matrix, the heat storage matrix is also
evaluated by numerical integration as shown below:

Page 248
TEMP/W Chapter 12: Theory

n
[ M ] = τ ∑ λ j <N>T <N> det |J j | W1 jW2 j
j =1

TEMP/W uses a consistent formulation, as opposed to a lumped formulation, to


establish the capacitance matrix. The two methods are presented and discussed by
Segerlind (1984). The adoption of the consistent formulation is necessitated by the
generality of TEMP/W which allows the elements to have all, some, or none of the
secondary nodes.
The capacitance matrix is formed by numerical integration as discussed in this
chapter. This has the advantage of making it possible for the parameter λ to vary
throughout the element. At each integration point, TEMP/W obtains a value of λ
based on the latent heat of water, the volumetric water content, the volumetric heat
capacity, and the slope of the unfrozen water content function.
TEMP/W computes the slope of the unfrozen water content function from a
straight line between the previously-computed temperature and the newly-
computed temperature at a Gauss point, as illustrated in Figure 12-2.
The slope of this straight line can be viewed as the average rate of change during
one increment of time. This is considered to be a more realistic value than taking
the derivative of the function at a specific point.
An exception to this procedure is when the old and new temperature values are
nearly identical. In this case, TEMP/W computes the slope by calculating the
derivative of the function at the average of the two temperature values.

Page 249
Chapter 12: Theory TEMP/W

Figure 12-2 Computation of heat storage term

12.8 Flux boundary vector


The nodal flux boundary vector {Q} for a two-dimensional analysis is defined as:

{Q} = qτ ∫ ( <N > ) dL


T
L

or for an axisymmetric analysis as:

{Q} = q ∫ ( <N > R ) dL


T
L

and for a plan view analysis as:

{Q} = q ∫ ( <N > < N > ) dA


T
A

where:
q = the unit flux across the side of an element,
τ = the element thickness,
A = the area of the element in plan view, and

Page 250
TEMP/W Chapter 12: Theory

R = the radial distance from the symmetric axis to the element corner nodes.

Solutions to the integrals are dependent on the analysis type and on the presence of
secondary nodes. For two-dimensional (i.e., vertical section) and axisymmetric
analyses, the integrals are solved by closed-form solutions, as illustrated in Figure
12-3 and Figure 12-4. However, for plan view analysis, the contributing area of a
node is computed by numerical integration in the same way as forming the heat
matrix. In other words, the contributing area per node in a plan view depends on
the partial contributing areas of all elements surrounding that node and this must
be computed using the gauss region area integration scheme for each element, not
standard shape relationships as discussed here.
Two types of flux boundaries may be specified in TEMP/W namely: a nodal flux
boundary (Q) and a unit flux boundary (q). A nodal flux boundary (Q) can be
specified directly on the boundary nodes. A unit flux boundary (q) must be
specified along the boundary edges of the elements, except for a plan view analysis
where the unit flux is applied per unit area on the plan view. When you set up a
boundary condition, you identify the edges of the elements across which a q
boundary should be applied. Based on this specific element edge information, the
solver performs the integration and determines the applied flux Q at the nodes. The
solver needs Q, not q, to solve the finite element equations.
For two-dimensional (i.e., vertical section or top view section in 2D) and
axisymmetric analyses, the nodal flux Q computed by the solver is dependent on
the specified element thickness. For plan view analysis, since the surface area is
independent of the element thickness, the nodal flux Q is also independent of the
element thickness.

Page 251
Chapter 12: Theory TEMP/W

Figure 12-3 Contributing area for section view elements with unit
thickness

Page 252
TEMP/W Chapter 12: Theory

Figure 12-4 Contributing area for axisymmetric elements over unit


Radian thickness

Page 253
Chapter 12: Theory TEMP/W

12.9 Convective heat transfer


TEMP/W has the ability to integrate with SEEP/W in order to take into account the
convective heat transfer that occurs due to flowing water. In order to include
convective heat transfer effects, TEMP/W must obtain the water content and water
velocity at every Gauss point for every time step.
With this information, the partial differential equation for heat flow (Equation
12-2) is modified to:

∂ ⎛ ∂T ⎞ ∂ ⎛ ∂T ⎞ ∂T ∂T
⎜ kx ⎟+ ⎜ ky ⎟ + ρ cVx + ρ cVy +Q
∂x ⎝ ∂x ⎠ ∂y ⎝ ∂y ⎠ ∂x ∂y
⎛ ∂W ⎞ ∂T
= ⎜ c +Lw u ⎟
⎝ ∂T ⎠ ∂t
where:
ρc = volumetric specific heat of water, and
Vxy = the Darcy water velocity in x and y directions.

When the convective analysis is used, the water contents specified in the TEMP/W
program are over-written by those obtained from SEEP/W at the onset of freezing.
Once freezing is initiated at a Gauss region, the total water content is assumed to
not change. Therefore, there is no frost heave affect.

12.10 Establishing the phase change region


The rate at which the phase change region develops is controlled by the Gauss
Regions/Iteration (GRI) parameter. The value of GRI is specified in the Define
module.
Figure 12-5 to Figure 12-7 show how TEMP/W uses the GRI parameter to develop
the phase change region for a one-dimensional example. This example problem
contains 7 elements with 4 Gauss regions in each element (the dotted lines in each
element show the Gauss regions) and has a GRI value of 2. After a period of time,
a frozen zone develops at the top of the mesh, as shown by the shaded area in the
figures.
The procedure TEMP/W follows for the subsequent time step is as follows:

Page 254
TEMP/W Chapter 12: Theory

Based on the existing temperature profile and the new boundary conditions, the
temperature distribution for the new time step is computed without taking into
account the latent heat. This process may take a few iterations to converge,
depending on the degree of non-linearity of the thermal conductivity functions.
Once it has converged, a potential phase change zone is established.
SOLVE identifies the two Gauss regions that have undergone a phase change and
have temperatures closest to the phase change temperature. Latent heat is applied
to these two Gauss regions, and the iteration process continues until convergence is
achieved. This process causes the potential phase change zone to shrink.
SOLVE again identifies the two Gauss regions that have undergone a phase change
and have temperature closest to the phase change temperature. Now there are a
total of four such Gauss regions. During the next iteration, latent heat is applied to
these four Gauss regions and the iteration process continues until convergence is
achieved. This process causes the potential phase change zone to shrink back
further.
The process continues until there are no more Gauss regions with a potential for
undergoing a phase change. The phase change region has now been established.

Figure 12-5 First attempt to find phase change region

Page 255
Chapter 12: Theory TEMP/W

Figure 12-6 Continual refinement of phase change region

Page 256
TEMP/W Chapter 12: Theory

Figure 12-7 Converged phase change region

Selecting the Gauss regions / iteration (GRI) parameter


In general, decreasing the GRI parameter will increase the number of iterations
required to reach a solution. Conversely, making the GRI parameter too high can
result in phase change zones that are too large, thereby making the solution
unrealistic. This section discusses some factors that need to be considered when
selecting a suitable GRI parameter.
When you select a value for GRI, it is important to consider the size and nature of
the problem. For example, if you have a problem with a large number of elements
and the freeze/thaw movement is primarily vertical (e.g., under a wide building),
you can use a larger value for GRI. This is because the phase change region will
consist of a large number of Gauss regions at the same level, even if the
freeze/thaw movement is fairly small. If the freeze/thaw interface is curved, then a
lower value may be necessary, since the direction in which the freeze/thaw
interface will move is more uncertain.

Page 257
Chapter 12: Theory TEMP/W

Another factor to consider in selecting a value for GRI is the element integration
order. For example, if you use 9-point Gauss integration for 8-noded quadrilateral
elements, the Gauss regions are smaller than when you use 4-point Gauss
integration for 4-noded quadrilateral element, provided that the elements are the
same size. Therefore, the value of GRI potentially can be higher for 9-point Gauss
integration than for 4-point Gauss integration.
A value for GRI that is a multiple of the Gauss integration order provides a useful
guideline. For example, if the integration order is 9, then a value for GRI might be
9, 18, 24, 36, or 45. While this is not a necessary requirement, it sometimes helps
the freeze/thaw interface to develop in a more uniform direction.
The value of GRI can also be set too low, especially in a one-dimensional analysis.
Consider a problem with a column of 8-noded elements that use 9-point Gauss
integration. In this case, the value of GRI should be 3, since all 3 Gauss regions at
the same level must change phase at the same time in order to keep the freeze-thaw
interface horizontal. Setting the value of GRI below 3 may result in a curved
freeze-thaw interface, which is not possible in a one-dimensional analysis.
In summary, there are many factors involved in selecting acceptable values for
GRI. Generally, it is best to begin the analysis with a reasonably high GRI value.
The value can then be lowered in stages until a further reduction in the value of
GRI has no significant effect on the computed temperature distribution. While this
procedure requires running many analyses, it will help you to gain confidence in
the acceptability of the computed results.

Page 258
TEMP/W Appendix A: Interpolating Functions

13 Appendix A: Interpolating Functions

13.1 Coordinate systems


The global coordinate system used in the formulation of TEMP/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 13-1. Presented as well in Figure 13-1 is the local element
node numbering system. The local coordinates for each of the nodes are given in
Table 13-1.

Table 13-1 Local element node numbering system

Element Type Node r s


Quadrilateral 1 +1 +1
2 -1 +1
3 -1 -1
4 +1 -1
5 0 +1
6 -1 0
7 0 -1
8 1 0
Triangular 1 0 0
2 1 0
3 1 1
4 – –
5 ½ 0
6 ½ ½
7 0 ½
8 – –

TEMP/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 259
Appendix A: Interpolating Functions TEMP/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. TEMP/W uses the same functions for relating the coordinate systems as
for describing the variation of the field variable (temperature) within the element.
The elements are consequently isoperimetric elements.
y 1 (1, 1)
s
5
2 (1, -1)
8 r

6 4 (-1, 1)

Local coordinates (r, s)


3 (-1, 1)
Global coordinates (x, y)

x
A) Quadrilateral Element

3 (0, 1)
6
7 2 (1, 0)
r

5
Local coordinates (r, s)
1 (0, 0) Global coordinates (x, y)

x
B) Triangular Element

Figure 13-1 Global and local coordinate system


The x and y coordinates anywhere in the element are related to the local
coordinates and to the x y coordinates of the nodes by the following equations:

Page 260
TEMP/W Appendix A: Interpolating Functions

x= N {X }
Equation 13-1
y= N {Y }
where:
<N> = is a vector of interpolating shape functions,
{X} = the global x coordinates of the element nodes, and
{Y} = the global y coordinates of the element node.

The interpolating functions are expressed in terms of local coordinates. Therefore,


once a set of local coordinates (r,s) have been specified, the corresponding global
coordinates can be determined by Equation 13-1.

13.2 Interpolating functions


TEMP/W uses a general set of interpolating functions presented by Bathe (1982,
pp. 200, 230). These general functions are suitable for elements which have none,
some, or all of the secondary nodes defined. This allows for considerable
versatility in the types of elements that can be used.
The interpolating functions in terms of local coordinates r and s for quadrilateral
and triangular elements are given in Table 13-2 and Table 13-3, respectively.
The functions represent a linear equation when the secondary nodes are missing
and a quadratic (nonlinear) equation when the secondary nodes are included.

Page 261
Appendix A: Interpolating Functions TEMP/W

Table 13-2 Interpolation functions for quadrilateral elements

Function Include in function if node is present

5 6 7 8
N1 = ¼(1+r)(1+s) -½N5 — — -½N8
N2 = ¼(1-r)(1+s) -½N5 -½N6 — —
N3 = ¼(1-r)(1-s) — -½N6 -½N7 —
N4 = ¼(1+r)(1-s) — — -½N7 -½N8
N5 = ½(1-r2)(1+s) — — — —
N6 = ½(1-s2)(1-r) — — — —
N7 = ½(1-r2)(1-s) — — — —
N8 = ½(1-s2)(1+r) — — — —

Table 13-3 Interpolation functions for triangular elements

Function Include in function if node is present

5 6 7
N1= 1-r-s -½N5 — -½N7
N2 = r -½N5 -½N6 —
N3 = s — -½N6 -½N7
N5 = 4r (1-s) — — —
N6 = 4rs — — —
N7 = 4s(1-r-s) — — —

Field variable model


To formulate a finite element analysis it is necessary to adopt a model for the
distribution of the field variable within the element. Since the field variable in the
thermal analysis is the temperature (T), it is necessary to adopt a model for the
distribution of T within the element.
TEMP/W assumes that the temperature distribution within the element follows the
adopted interpolating functions. This means that the temperature distribution is
linear when the secondary nodes are missing, and the temperature distribution is
nonlinear when the secondary nodes are present.

Page 262
TEMP/W Appendix A: Interpolating Functions

In equation form the temperature distribution model is:

Equation 13-2 t = N {T }

where:
t = the temperature at any local coordinate,
<N> = a vector of interpolation function, and
{T} = a vector of temperatures at the nodes.

Interpolation function derivatives


The constitutive relationship for a conductive thermal analysis is Darcy's type flow
law. The equation is:

q = ki

The gradient i is one of the key parameters required in the finite element
formulation. The following presents the procedure used by TEMP/W to compute
the gradient.
From the adopted temperature distribution model, the temperature at any point
within the element in terms of the nodal temperatures is obtained from Equation
13-2.
The gradients in the x and y directions are then known by:

∂t ∂ N
ix = = {T }
∂x ∂x
Equation 13-3
∂t ∂ N
iy = = {T }
∂y ∂y

The interpolating functions are written in terms of r and s and not in terms of x and
y. The derivatives must consequently be determined by the chain rule of
differentiation, as follows:

∂ N ∂ N ∂x ∂ N ∂y
= +
∂s ∂x ∂s ∂y ∂s

Page 263
Appendix A: Interpolating Functions TEMP/W

This can be written as:

⎧∂ N ⎫ ⎧∂ N ⎫
⎪⎪ ⎪⎪ ⎪ ⎪
∂r ⎪ ∂x ⎪
⎨ ⎬ [ ]⎨
= J ⎬
⎪∂ N ⎪ ⎪∂ N ⎪
⎪⎩ ∂s ⎪⎭ ⎪⎩ ∂y ⎪⎭

where [J] is the Jacobian matrix and is defined as:

⎡ ∂x ∂y ⎤
⎢ ∂r ⎥
Equation 13-4 [ J ] = ⎢ ∂∂xr ∂y ⎥
⎢ ⎥
⎣ ∂s ∂s ⎦
The derivatives of the interpolation function with respect to x and y is called the B
matrix and can be determined by inverting the Jacobian matrix and rewriting the
equations as:

⎧∂ N ⎫ ⎧∂ N ⎫
⎪ ⎪ ⎪ ⎪⎪
[ B ] = ⎪⎨ ∂ ∂Nx ⎪ −1 ⎪ ∂r
⎬ = [ J ] ⎨ ⎬
⎪ ⎪ ⎪∂ N ⎪
⎩⎪ ∂y ⎭⎪ ⎩⎪ ∂s ⎭⎪

Recalling Equation 13-1 and taking their derivative with respect to s and r as
follows:

∂x ∂ N
= {X }
∂r ∂r
∂x ∂ N
= {X }
∂s ∂s
∂y ∂ N
= {Y }
∂r ∂r
∂x ∂ N
= {Y }
∂s ∂s

Page 264
TEMP/W Appendix A: Interpolating Functions

then substituting these values into Equation 13-4, the Jacobian matrix becomes:

⎡ ∂ ( N1 , N 2 ...N8 ) ⎤ ⎢⎡ 1
X Y1 ⎤
⎢ ⎥ X Y2 ⎥⎥
[ J ] = ⎢ ∂( N , ∂Nr ...N ) ⎥ ⎢⎢ M 2 M ⎥
⎢ 1 2 8 ⎥
⎣ ∂s ⎦ ⎢⎣ X 8 Y8 ⎦

The derivatives of the interpolation functions with respect to r and s are required to
compute the Jacobian matrix and to compute the flow gradients (Equation 13-3
The derivatives of the interpolation functions with respect to r and s used by
TEMP/W for quadrilateral and triangular elements are given in Table 13-4 and
Table 13-5 respectively.

Page 265
Appendix A: Interpolating Functions TEMP/W

Table 13-4 Interpolation function derivatives for quadrilateral elements

Derivative of Function Include in derivative if node is present

5 6 7 8
N1,r = ¼(1+s) -½(N5,r) — — `
N2,r = -¼(1+s) -½(N5,r) -½(N6,r) — —
N3,r = -¼(1-s) — -½(N6,r) -½(N7,r) —
N4,r = ¼(1-s) — — -½(N7,r) -½(N8,r)
N5,r = -½(2r+2sr) — — — —
N6,r = -½(1-s2) — — — —
N7,r = -½(2r-2sr) — — — —
N8,r = ½(1-s2) — — — —
N1,s = ¼(1+r) - ½(N5,s) — — -½(N8,s)
N2,s = ¼(1-r) -½(N5,s) -½(N6,s) — —
N3,s = -¼(1-r) — -½(N6,s) -½(N7,s) —
N4,s = -¼(1+r) — — -½(N7,s) -½(N8,s)
N5,s = ½(1-r2) — — — —
N6,s = -½(2s-2sr) — — — —
N7,s = -½(1-r2) — — — —
N8,s = -½(2s+2sr) — — — —

Page 266
TEMP/W Appendix A: Interpolating Functions

Table 13-5 Interpolation function derivatives for triangular elements

Derivative of Function Include in derivative if node is present

5 6 7
N1,r = -1.0 -½(N5,r) — —
N2,r = 1.0 -½(N5,r) -½(N6,r) —
N3,r = 0.0 — -½(N6,r) -½(N7,r)
N5,r = (4-8r-4s) — — —
N6,r = 4s — — —
N7,r = -4s — — —
N1,s = -1.0 -½(N5,s) — —
N2,s = 0.0 -½(N5,s) -½(N6,s) —
N3,s = 1.0 — -½(N6,s) -½(N7,s)
N5,s = -4r — — —
N6,s = 4r — — —
N7,s = (4-4r-8s) — — —

The following notation is used in the preceding tables:

∂N i
Ni , r =
∂r
∂N
Ni , s = i
∂s
The Jacobian matrix is a 2x2 matrix:

⎡J J12 ⎤
[ J ] = ⎢ J 11 J 22 ⎥⎦
⎣ 21

The inverse of [J] is:

⎡J − J12 ⎤
[J ]
−1
= ⎢ 22
⎣ − J 21 J11 ⎥⎦

The determinant of [J] is:

Page 267
Appendix A: Interpolating Functions TEMP/W

det [ J ] = J11 × J 22 − J 21 × J12

13.3 Infinite elements


Many thermal problems can be classified as unbounded. Consider the case
illustrated in Figure 13-2. The condition along the bottom horizontal boundary
cannot be correctly defined as a temperature boundary or as a flux boundary. The
problem is said to be unbounded on the bottom. The correct boundary condition is
only known at some large (i.e., infinite) distance from the bottom of the problem.
These types of unbounded problems can be analyzed with infinite elements.
TEMP/W follows the infinite element formulation presented by Bettess, 1992, pp.
53-85 and is formulated only for eight-noded quadrilateral elements.

Figure 13-2 An unbounded flow problem

Page 268
TEMP/W Appendix A: Interpolating Functions

Mapping functions
For standard elements, TEMP/W uses the same interpolating functions both to
describe the distribution of the field variable within an element and to relate the
local and global coordinate system (i.e., the elements are isoperimetric). For
infinite elements, the relationship between the local and global coordinate system
must be described by a special set of shape functions. The interpolating functions
for describing the field variable distribution is the same for all elements, but the
geometric shape functions are different.
TEMP/W uses the “serendipity” family of mapping functions presented by Bettess,
1992, pp. 53-85, to relate the local and global coordinate systems for infinite
elements. The unique feature about these functions is that they are zero at the
nodes deemed to be at infinity.

Figure 13-3 Node numbering scheme for infinite elements


Figure 13-3 presents the different types of elements which extend to infinity. The
mapping or shape functions for one-directional and two-directional, (corner),
infinite elements are given in Table 13-6. In the table, r and s are the local
coordinates within the element and M1 through M8 are the mapping functions for
infinite elements shown Figure 13-3.

Page 269
Appendix A: Interpolating Functions TEMP/W

The derivatives of the mapping functions are required to compute the Jacobian
matrix and the flow gradients for infinite elements. The derivatives of the mapping
functions for one-directional and two-directional infinite elements with respect to r
and s are given in Table 13-7 and Table 13-8 respectively.

Table 13-6 Mapping functions for infinite elements

1-D Infinite Elements 2-D Infinite Elements


M1 = 0 M1 = 0

M2 =
(− 1 − r − rs + s ) 2
M2 = 0
(1 − r )

M3 =
( −1 − r + rs + s )2

M3 =
−4 (1 + r + s )
(1 − r ) (1 − r )(1 − s )
M4 = 0 M4 = 0

(1 + r )(1 + s )
M5 = M5 = 0
2 (1 − r )

2 (1 − s 2 ) 2(1 + s )
(1 − r )(1 − s )
M6 = M6 =
(1 − r )
(1 + r )(1 − s ) 2 (1 + r )
M7 = M7 =
2 (1 − r ) (1 − r )(1 − s )
M8 = 0 M8 = 0

Page 270
TEMP/W Appendix A: Interpolating Functions

Table 13-7 Derivatives of mapping functions for one-directional infinite


elements

Derivative in r-Directions Derivative in s-Direction


M1,r = 0 M1,s = 0

( −2 − s + s ) 2
M2,s =
( −r + 2s )
M2,r =
(1 − r )
2
(1 − r )
( −2 + s + s ) 2
M3,s =
( r + 2s )
M3,r =
(1 − r )
2
(1 − r )
M4,r = 0 M4,s = 0

M5,r =
(1 + s ) M5,s =
(1 + r )
(1 − r )
2
2 (1 − r )

2 (1 − s 2 ) M6,s =
−4 s
M6,r =
(1 − r )
2 (1 − r )

M7,r =
(1 − s ) M7,s =
− (1 + r )
(1 − r )
2
2 (1 − r )
M8,r = 0 M8,s = 0

Page 271
Appendix A: Interpolating Functions TEMP/W

Table 13-8 Derivatives of mapping functions for two-directional elements

Derivative in r-Direction Derivative in s-Direction


M1,r = 0 M1,s = 0
M2,r = 0 M2,s = 0

−4 ( s + 2 ) −4 ( r + 2 )
M3,r = M3,s =
(1 − r ) (1 − s ) (1 − r )(1 − s )
2 2

M4,r = 0 M4,s = 0
M5,r = 0 M5,s = 0

2 (1 + s ) 4
M6,r = M6,s =
(1 − r ) (1 − s ) (1 − r )(1 − s )
2 2

4 2 (1 + r )
M7,r = M7,s =
(1 − r ) (1 − s ) (1 − r )(1 − s )
2 2

M8,r = 0 M8,s = 0

Once these shape functions and derivatives have been defined, the numerical
integration scheme used to form the element characteristic matrix for the infinite
elements is the same as for the standard elements. These special shape functions
must also be used when formulating the heat matrix and nodal action (i.e., heat
flux) vector for axisymmetric cases.

Pole definition
In order to project the outer edge of the infinite element to infinity, it is necessary
to define a pole position. This is simply a point at some position on the opposite
side of the infinite edge of the element. Figure 13-3 shows the infinite element pole
at the center of the finite element mesh.
For a one-directional infinite element, the nodes on the infinite edge are taken to be
at infinity. Secondary Nodes 5 and 7 must be extended out towards infinity.
TEMP/W does this according to the equations below.
For infinity in the x-direction only:

Page 272
TEMP/W Appendix A: Interpolating Functions

x5 = x2 + ( x2 − x p )
x7 = x3 + ( x3 − x p )

For infinity in the y-direction only:

y5 = y2 + ( y2 − y p )
y7 = y3 + ( y3 − y p )

If the x- and y-directions of infinity are both positive or both negative, there is no
adjustment to x6 and y7. However:

x7 = x3 + ( x3 − x p )
y6 = y3 + ( y3 − y p )

If one direction of infinity is positive and the other is negative, there is no


adjustment to x7 and y6. However:

x6 = x3 + ( x3 − x p )
y7 = y3 + ( y3 − y p )

Generally, the TEMP/W solutions are not sensitive to the position of the pole,
provided the pole is in a reasonable position. As a broad guideline, the pole must
either be positioned at the origin of the flow as it migrates towards infinity or
positioned at the exit point of the flow if the flow originates at infinity. Further
discussion on infinite elements and their application is given in the Meshing
chapter.

Page 273
TEMP/W References

References
Bathe, K-J., 1982. Finite Element Procedures in Engineering Analysis. Prentice-
Hall.

Bettess, P., 1992. Infinite Elements. Penshaw Press.

Black, P. and Allen Tice, 1989. comparison of Soil Freezing Curve and Soil Water
Curve Data for Windsor Sandy Loam. Water Resources Research, Vol. 25,
No. 10. pp. 2205-2210.

Burland, J.B. 1987. Nash Lecture: The Teaching of Soil Mechanics – a Personal
View. Proceedings, 9th ECSMFE, Dublin, Vol. 3, pp 1427-1447.

Burland J.B. 1996. Closing Session Discussions, Proceedings of the First


International Conference on Unsaturated Soils Conference, Paris, Vol. 3,
p. 1562. (proceedings published by A.A. Balkema; Editors, E.E, Alonso
and P. Delage).

Carter, J.P., Desai, C.S, Potts., D.M., Schweiger, H.F, Sloan, S.W. 2000.
Computing and Computer Modeling in Geotechnical Engineering, Invited
Paper, Conference Proceedings, GeoEng2000, Melbourne, Australia.

Carslaw, H.S, and Jaeger, J.C., 1947. Conduction of Heat in Solids. Clarendon
Press, Oxford

Coutts, R.J and Konrad, J.M., 1994. Finite Element Modeling of Transient Non-
Linear Heat Flow Using the Node State Method. Intl. Ground Freezing
Conf., France, November.

Edlefsen, N.E. & Anderson, A.B.C., 1943. Thermodynamics of soil moisture.


Hilgardia 15: 2: pp. 31-298.

Farouki, O.T. 1981. Thermal Properties of Soils. CRREL Monograph 81-1. U.S.
Army Corps of Engineers, Cold Regions Research and Engineering
Laboratory, Hanover, New Hampshire.

Page 275
References TEMP/W

Flerchinger, G.N. and K.E. Saxton. 1989. Simultaneous heat and water model of a
freezing snow-residue-soil system I. Theory and development. Trans. of
ASAE, 32(2), 565-571.

Goodrich, L.E., and Gold, L.W., 1981. Ground Thermal Analysis: Permafrost
Engineering Design and Construction. Edited by Johnston, G.H. Ch. 4.
John Wiley & Sons.

Harlan, R.L., and Nixon, J.F., 1978. Ground Thermal Regime. Geotechnical
Engineering For Cold Regions, Edited by O.B. Andersland and D.M.
Anderson. McGraw-Hill Inc.

Hwang, C.T., Murray, D.W., and Brooker, E.W, 1972. A Thermal Analysis for
Structures on Permafrost. Canadian Geotechnical Journal, Vol. 9, No. 1.

Johansen, O., 1975. Thermal Conductivity of Soils. Ph.D. thesis. Trondheim,


Norway. (CRREL Draft Translation 637, 1977), ADA044002.

Johnston, G.H., Ladanyi, B., Morgenstern, N.R., and Penner, E., 1981.
Engineering Characteristics of Frozen and Thawing Soils. Permafrost
Engineering Design and Construction. Edited by Johnston, G.H. John
Wiley & Sons.

Liang, X., D. P. Lettenmaier, E. F. Wood, and S. J. Burges, A Simple


hydrologically Based Model of Land Surface Water and Energy Fluxes for
GSMs, J. Geophys. Res., 99(D7), 14,415-14,428, 1994.

Lancaster, P., and Salkauskas, K., 1986. Curve and Surface Fitting: An
Introduction. Academic Press.

Milly, P.C.D, 1982. Moisture and heat transport in hysteretic, inhomogeneous


porous media: A matric head based formulation and a numerical model.
Water Resources Research, 18(3): 489-498.

Morgenstern , N.R. 2000. Common Ground, Invited Paper, Conference


Proceedings, GeoEng2000, Melbourne, Australia.

National Research Council, 1990. Groundwater Models: Scientific and regulatory


applications. Water Science and Technology Board, Committee on Ground
Water Modeling Assessment. Commission on Physical Sciences,

Page 276
TEMP/W References

Mathematics and Resources. National Research Council, National


Academy Press, Washington, D.C.

Newman, G. P. 1995. Heat and Mass Transfer of Unsaturated Soils during


Freezing. M.Sc., Thesis, University of Saskatchewan, Saskatoon SK,
Canada.

Newman, G. P. and G. W. Wilson (1997) “Heat and Mass Transfer In Unsaturated


Soils during Freezing.” Canadian Geotechnical Journal, 34: 63-70.

Newman, G. and D. Maishman. Underground Freezing at a High Grade Uranium


Mine. Ground Freezing 2000: Frost Action in Soils Proceedings of the 9th
International Symposium, Louvain-La-Neuve, Belgium, 11-13 September
2000.

O’Kane Consultants Inc. 2004. Personal communication.

Nixon, J.F., and McRoberts, E.C., 1973. A Study of Some Factors Affecting the
Thawing of Frozen Soils. Canadian Geotechnical Journal, Vol. 10.

Salkauskas, K., 1974. C' Splines For Interpolation of Rapidly Varying Data. Rocky
Mountain Journal of Mathematics, Vol. 14, No. 1.

Segerlind, L.J., 1984. Applied Finite Element Analysis. John Wiley and Sons.

Wilson, G.W., 1990. Soil Evaporative Fluxes for Geotechnical Engineering


Problems. Ph.D. Thesis, University of Saskatchewan, Saskatoon, Canada.

Page 277
Index TEMP/W

Index
adaptive time stepping .................147 Element shapes ..............................50
Analysis Types ............................127 Energy flow vectors.....................159
Artificial ground freezing application engineering units..........................173
.................................................113
Equation solvers (direct or iterative)
Axisymmetric ..............................134 .................................................145
Best-fit splines ...............................93 Equipotential lines .......................154
Boundary Conditions .....................97 estimated material properties.........37
Boundary functions......................121 Estimating thermal conductivity....87
buried piping....................................5 Estimating volumetric heat capacity
...................................................80
Burland ..........................................19
Field variable distribution..............45
chilled pipelines ...............................7
Finite element heat flow equations
chilled structures..............................6 .................................................241
Compare alternatives .....................24 Flow law ......................................239
Contours.......................................157 Flow paths....................................161
Convective cooling of surfaces........8 Flux boundary vector...................250
Convective heat transfer ..............254 Flux section application...............164
Convective heat transfer surfaces 113 Flux section location....................176
Convergence ................................140 Flux section theory ......................162
Convergence Tolerances..............175 Flux sections ................................161
Element addition and removal .....180 function estimation ........................95
Element and mesh compatibility ...46 Gauss integration .........................144
Element fundamentals ...................44 Gauss regions / iIteration .............257

Page 278
TEMP/W Index

Governing equations....................239 Modeling Tips and Tricks............173


gradients and velocities................158 Modifier functions .......................123
Graphing ......................................168 Nodal and unit heat flux versus time
.................................................122
Ground surface temperatures .......105
Node and element information ....152
Heat flow boundary conditions....100
Null elements...............................125
heat storage ....................................77
numerical experiments...................35
Heat storage matrix......................248
Numerical integration ............48, 244
How not to model ..........................40
Numerical Issues..........................139
How to model ................................30
Numerical Modeling – What, Why
Illustrative Examples ...................213 and How.....................................13
Infinite elements ....................71, 268 phase change region.....................254
ing regions .....................................64 Plan view .....................................136
Initial conditions ..........................128 Pole definition..............................272
Interface elements..........................70 Product Integration Illustrations ..181
Interpolating functions.................261 Pyrex tile cooling application ......116
Interpolating Functions................259 Quantitative predictions.................21
Make a guess .................................31 Region points.................................52
Mapping functions .......................269 Region properties...........................54
material data base ..........................95 Region types ..................................52
Material Properties.........................77 Regions ..........................................51
Mesh types & patterns ...................54 Reporting .....................................171
Mesh with openings.......................59 roads and airstrips............................1
Meshing .........................................43 Secondary variables .......................50

Page 279
Index TEMP/W

Simplify geometry .........................32 Thermal conductivity.....................83


soil stabilization ...............................2 Thermal conductivity matrix .......247
Source or sink flux values............177 Thermosyphon boundary conditions
.................................................118
Sources and sinks.........................104
Time stepping ......................131, 146
Start simple ....................................34
train our thinking ...........................26
Steady state ..................................127
Transfinite meshing .......................57
Steady state heat flow analysis ....213
Transient ......................................128
Steep material property functions 143
Triangular regions..........................55
Stopping and restarting ................179
Unit energy flux versus total energy
Structured mesh .............................54 flux? .........................................178
Structured versus unstructured Unstructured mesh .........................54
meshing......................................66
Vector norms ...............................141
Surface regions ..............................60
Visualization of Results...............151
Temperature boundary conditions100
volumetric heat capacity ................79
Temperature versus time..............121
Weighted splines............................91
Temporal integration ...................243
Why model ....................................21
Theory..........................................239

Page 280
TEMP/W Quick Reference Tables

Quick Reference Tables


Consistent Set of Units

Parameter Symbol Units Units Units Units

Length L meters (m) feet (ft) meters (m) meters (m)

Time t days (day) hours (hr) hours (hr) seconds (sec)

Heat H kilojoules (kJ) Btu (Btu) kilocalories (kcal) Joules (J)

Temperature T °C °F °C °C

Latent heat of
H/L3 kJ/m3 Btu/ft3 kcal/m3 J/m3
water
Thermal
H/(t × L × T) kJ/(day × m × °C) Btu/(hr × ft × °F) kcal/(hr × m × °C) J/(sec × m × °C)
conductivity
Volumetric heat
H/(L3 × T) kJ/(m3 × °C) Btu/(ft3 × °F) kcal/(m3 × °C) J/(m3 × °C)
capacity

Total flux (Q) H/t kJ/day Btu/hr kcal/hr J/sec

Unit heat flux (q) H/(t × L2) kJ/(day × m2) Btu/(hr × ft2) kcal/(hr × m2) J/(sec × m2)

NOTE: The unit of J/sec is sometimes expressed as Watts, (1 J/sec = 1 W).

Page 281
Quick Reference Tables TEMP/W

Consistent Set of Units

Parameter Symbol Units Units Units Units

Length L meters (m) feet (ft) meters (m) meters (m)

Time t days (day) hours (hr) hours (hr) seconds (sec)

Heat H kilojoules (kJ) Btu (Btu) kilocalories (kcal) Joules (J)

Temperature T °C °F °C °C

Latent heat of
H/L3 kJ/m3 Btu/ft3 kcal/m3 J/m3
water
Thermal
H/(t × L × T) kJ/(day × m × °C) Btu/(hr × ft × °F) kcal/(hr × m × °C) J/(sec × m × °C)
conductivity
Volumetric heat
H/(L3 × T) kJ/(m3 × °C) Btu/(ft3 × °F) kcal/(m3 × °C) J/(m3 × °C)
capacity

Total flux (Q) H/t kJ/day Btu/hr kcal/hr J/sec

Unit heat flux (q) H/(t × L2) kJ/(day × m2) Btu/(hr × ft2) kcal/(hr × m2) J/(sec × m2)

NOTE: The unit of J/sec is sometimes expressed as Watts, (1 J/sec = 1 W).

Page 282

You might also like