Está en la página 1de 148

Analytical and Numerical Modelling of Thermal

Conductive Heating in Fractured Rock

by

Daniel Peter Baston

A thesis submitted to the

Department of Civil Engineering

in conformity with the requirements for

the degree of Master of Science (Engineering)

Queen’s University

Kingston, Ontario, Canada

April 2008

Copyright
c Daniel Peter Baston, 2008
Abstract

Analytical and numerical modelling studies were conducted to assess the performance of

thermal conductive heating (TCH) systems for the purpose of contaminated site remedia-

tion. Modelling was conducted in a fractured bedrock environment containing a system of

parallel, equally-spaced horizontal fractures.

A semi-analytical solution to the two-dimensional heat conduction equation was devel-

oped and used to study temperature distributions between two thermal wells. A sensitivity

analysis was conducted to assess the relative importance of hydrogeological parameters (hy-

draulic gradient, fracture aperture, fracture spacing) and rock material properties (density,

thermal conductivity, heat capacity). Hydrogeological parameters were far more important

than rock material properties in determining treatment zone temperature distributions.

Knowledge of the bulk groundwater influx may be sufficient to predict the temperature

within the treatment zone for low to moderate values of influx.

To further the analysis, numerical modelling was employed. A three-dimensional domain

was constructed, representing a symmetrical portion of a heater well cluster. Simulations

were run for different combinations of bulk permeability, fracture spacing, matrix perme-

ability, and matrix porosity. Flow concentration in fractures had a significant effect on

treatment zone temperature distributions when bulk permeability was high. For low values

of bulk permeability (kb ≤ 10−14 m2 ), the minimum treatment zone temperature changed

by less then 7% when modelling the fractured medium as an equivalent homogeneous porous

i
medium.

Fracture spacing significantly influenced the time needed reach complete steam satura-

tion, even in cases where it did not affect temperature distributions. A pressure rise may

occur in the matrix as water expands thermally, elevating the boiling point of water. The

magnitude of the pressure rise is affected by the distance to the nearest fracture, as well as

the matrix permeability and porosity. For a given bulk permeability, the time needed to

reach complete steam saturation will be lengthened by an increase in fracture spacing, an

increase in matrix porosity, or a decrease in matrix permeability. Of these parameters, the

matrix permeability is the most significant.

The time needed to reach complete steam saturation in the matrix cannot be predicted

if the fracture spacing, matrix permeability, and porosity are not known. Further, a clear

temperature plateau is not observed during boiling in the matrix, posing a difficulty in mon-

itoring thermal treatment, where temperature measurements may be the only information

available.

ii
Acknowledgements

The work presented in this thesis was supported through a Discovery Grant from the Natural

Sciences and Engineering Research Council of Canada (NSERC), contract ER-0715 from

the U.S. Department of Defense Environmental Security Technology Certification Program

(ESTCP), and Queen’s University through scholarships to the author.

I would like to thank my advisor, Bernie Kueper, for providing me with his guidance

and support, while giving me the independence to chart my own path in this research. Kent

Novakowski was of great help to me in the early phase of my work – I am very thankful

that there was someone out there willing to talk shop about integral transforms. Ron Falta

and Karsten Pruess both provided me with a great deal of assistance in my numerical

modeling work, the results of which would have been difficult to analyze if not for the

data management ideas given to me by Rob Harrap and Gerry Barber. And I could have

accomplished nothing on the mathematical front without the foundations given to me by

Joe Siddiqui, Pat Farrell, Duncan Innes (deceased), and Guy Kember.

I am greatly indebted to the other students in my research group. In particular, Mike

West was especially generous with his knowledge about numerical modeling, scientific pub-

lishing, and the ins and outs of grad school at Queen’s. Tom Gleeson is thanked for his

valuable insights on my research, numerous edits to my papers, and letting me raise poul-

try in his backyard. Morgan Schauerte and Stephanie Villeneuve were both a great help in

my modeling work. Everything would have been more difficult had I not had the fun and

iii
supportive environment provided by my office-mates: Luis Bayona, Stephanie Grell, John

Kozuskanich, Eric Martin, Justin Matthew, Titia Praamsma, David Rodriguez, and Shawn

Trimper.

Many thanks to my friends, my family, and my partner, Valerie, for your patience,

support, and convincing me that it would all be worth it someday.

iv
Forward

Chapters 3 and 4 in this thesis have been written as self-contained manuscripts intended

for publication in Advances in Water Resources and Ground Water, respectively. Daniel

Baston is the senior author of both publications. Bernard Kueper is a co-author of both

publications. Supporting information for these chapters is provided in the appendices.

v
Table of Contents

Abstract i

Acknowledgements iii

Forward v

Table of Contents vi

List of Tables viii

List of Figures ix

Nomenclature xi

Chapter 1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

Chapter 2: Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1 Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Thermal Properties of Rock . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Subsurface Remediation by Thermal Conductive Heating & Soil Vapour Ex-
traction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.4 Laboratory Studies of Thermal Remediation . . . . . . . . . . . . . . . . . . 28
2.5 Heat Transfer in Fractured Media: Analytical Solutions . . . . . . . . . . . 31
2.6 Heat Transfer in Fractured Media: Numerical Models . . . . . . . . . . . . 33

Chapter 3: Screening Calculations . . . . . . . . . . . . . . . . . . . . . . . 41


3.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.3 Model Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4 Outline of Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.7 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

vi
3.8 Green’s Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

Chapter 4: Numerical Modelling . . . . . . . . . . . . . . . . . . . . . . . . 63


4.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3 Numerical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.6 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

Chapter 5: Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.1 Semi-Analytical Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.2 Numerical Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.3 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

Appendix A: Semi-analytical Solutions . . . . . . . . . . . . . . . . . . . . . 97


A.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
A.2 Single Fracture Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
A.3 Fracture Set Solution (Bessel Function Solution) . . . . . . . . . . . . . . . 106
A.4 Fracture Set Solution (Improved Solution) . . . . . . . . . . . . . . . . . . 110

Appendix B: Verification of Simplified Heat Balance . . . . . . . . . . . . 119

Appendix C: Numerical Discretization . . . . . . . . . . . . . . . . . . . . . 122


C.1 Discretization of Fracture . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
C.2 Radial Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

Appendix D: Calculated Numerical Model Input Parameters . . . . . . . 128


D.1 Fracture Zone Physical Properties . . . . . . . . . . . . . . . . . . . . . . . 128
D.2 Capillary Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
D.3 Relative Permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

vii
List of Tables

3.1 Base case parameters for sensitivity analysis . . . . . . . . . . . . . . . . . 52


3.2 Summary of sensitivity testing trials . . . . . . . . . . . . . . . . . . . . . . 52
3.3 Rock material properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.4 Parameters for runs used to assess correlation between bulk influx and treat-
ment zone temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.1 Base case properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68


4.2 Parameters varied. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

A.1 Parameters used for verification against the Gringarten et al. (1975) solution 115
A.2 Numerical model material properties . . . . . . . . . . . . . . . . . . . . . . 118

B.1 Parameters used for verification of heat storage term omission . . . . . . . . 121
B.2 Comparison of solution times when heat storage is included and neglected . 121

D.1 Capillary pressure parameters used in TOUGH2 simulation . . . . . . . . . 134


D.2 Relative permeability parameters used in TOUGH2 simulations . . . . . . . 134

viii
List of Figures

2.1 Photograph of thermal wells at a field site. . . . . . . . . . . . . . . . . . . . 21

3.1 Conceptual model of fractured rock environment. . . . . . . . . . . . . . . . 45


3.2 Schematic of model domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.3 Temperature distribution resulting from heating by a single thermal well. . 54
3.4 Summary of computed one-year interwell fracture temperatures. . . . . . . 55
3.5 Transient fracture temperature profiles at the centre of the treatment zone
for high and mid-level values of groundwater influx. . . . . . . . . . . . . . 56
3.6 Early and late-time fracture temperature profiles for various rock types. . . 58
3.7 Transient fracture temperature profiles at midpoint between heater wells,
showing the influence of hydraulic gradient. . . . . . . . . . . . . . . . . . . 59
3.8 Effect of increased heat production rate on the time needed to reach a target
temperature of 100 ◦ C and total energy consumption. . . . . . . . . . . . . 60

4.1 Plan view of model domain . . . . . . . . . . . . . . . . . . . . . . . . . . . 66


4.2 Section of model domain in r - z plane . . . . . . . . . . . . . . . . . . . . . 67
4.3 Isometric view of model domain . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.4 Pressure as a function of distance from the fracture for a location just outside
the treatment zone. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.5 Temperature vs. time for a reference block in the rock matrix and fracture,
at the boundary of the treatment zone. . . . . . . . . . . . . . . . . . . . . . 70
4.6 Impact of matrix permeability (km ) on magnitude of pressure spike at centre
of rock matrix and steam saturation within treatment zone. . . . . . . . . . 72
4.7 Impact of porosity (φ) on magnitude of pressure spike at centre of rock matrix
and steam saturation within the treatment zone. . . . . . . . . . . . . . . . 73
4.8 Minimum treatment zone temperature profiles for various values of fracture
spacing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.9 Relationship between steam saturation and distance from the fracture in base
case simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.10 Sensitivity of treatment zone boiling time to bulk medium properties. . . . 76

A.1 Heat balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98


A.2 Determining an appropriate number of fracture sets . . . . . . . . . . . . . 110
A.3 Comparison of new solution with that of Carslaw and Jaeger (1959, p. 263)
for the case of zero fracture aperture. . . . . . . . . . . . . . . . . . . . . . . 114
ix
A.4 Comparison of the present solution with that of Gringarten et al. (1975) . . 115
A.5 Comparison of temperatures in fracture after one year of heating, as calcu-
lated using the De Hoog et al. (1982) and Weeks (1966) algorithms. . . . . 117
A.6 Fracture temperature after 4 months of heating, computed using semi-analytical
solution and TOUGH2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

C.1 Comparison of semi-analytical solution, discrete fracture numerical solution,


and “fracture zone” numerical solution before and after boiling. . . . . . . . 123
C.2 Comparison of CPU time and error in temperature for different values of
fracture zone thickness. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
C.3 Fracture primary variable profiles for three domain sizes . . . . . . . . . . . 126
C.4 Effect of radial discretization on pressure, steam saturation, and temperature
at two reference points. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

D.1 Outline of calculation of derived input parameters . . . . . . . . . . . . . . 129

x
Nomenclature

Latin Letters

af anisotropy factor - page 18


Anm interfacial area of elements n and m m2 page 36
B Stefan-Boltzmann constant W/m2 ·K4 page 9
c gravimetric heat capacity J/kg·K page 8
D internodal distance m page 36
e fracture aperture m page 44
F mass flux vector kg/m2 ·s page 34
F radiative configuration factor - page 10
G heat flux vector W/m2 page 35
g gravitational acceleration vector m/s2 page 11
g internal heat generation W page 8
H fracture half-spacing m page 44
Hcc Henry’s Law constant - page 23
h specific enthalpy J/kg page 35
∇h hydraulic gradient - page 51
J Jacobian matrix - page 39
K thermal conductivity W/m·K page 7
k intrinsic permeability m2 page 11
kr relative permeability - page 35
L characteristic length m page 12
M mass kg page 34
m pore-size distribution parameter - page 131
N number of volume elements - page 37
n direction vector m page 7
P fluid pressure Pa page 11
Pcap capillary pressure Pa page 131
Pd displacement pressure Pa page 132
Pe fracture entry pressure Pa page 131
Pmax maximum capillary pressure for van Pa page 131
Genuchten (1980) function
Pe Peclet number - page 12
xi
q bulk groundwater influx L/m2 ·day page 54
qh heat flux W/m2 page 7
qm mechanical flux m/s page 8
qmass mass source kg/s page 34
q̄(x0 , s) fracture-matrix heat exchange function W page 49
R residual in equation varies page 37
r radial coordinate in numerical model m page 66
s Laplace variable s page 49
Se effective liquid saturation - page 131
Sgr residual gas saturation - page 133
Sl liquid saturation - page 131
Slr residual (irreducible) liquid saturation - page 131
Sls maximum liquid saturation - page 131
T temperature ◦C page 7
T̄ Laplace-space temperature ◦C page 49
Tmax maximum interwell fracture temperature ◦C page 53
Tmin minimum interwell fracture temperature ◦C page 53
U internal energy J page 35
V volume m3 page 34
v fluid velocity m/s page 12
wj quadrature weights at point j - page 50
x coordinate in direction of groundwater flow in m page 44
fracture
y coordinate normal to fracture plane in two- m page 45
dimensional solutions
∆zf z fracture zone thickness m page 127
z coordinate normal to fracture plane in three- m page 66
dimensional model

Greek Letters

α thermal diffusivity m2 /s page 8


αvg parameter in van Genuchten (1980) equation 1/L page 131
Λ radiative reflectance - page 9
λ pore size distribution index - page 132
Θ radiative absorptance - page 9
θ angular coordinate in numerical model ◦ page 66
µ dynamic fluid viscosity Pa·s page 11
ξ radiative energy flux W/m2 page 9
ρ density kg/m3 page 8
σ interfacial tension N/m page 132
φ porosity - page 14
Ω radiative transmittance - page 9

xii
Subscripts

b property of bulk medium page 67


dry “dry” value of parameter (Sl = 0) page 130
f fluid phase page 14
f rac property of fracture page 129
fz property of fracture zone page 127
i node number page 50
k time step number page 37
m node number page 36
m property of rock matrix page 67
max maximum value of parameter page 15
min minimum value of parameter page 14
n node number page 36
p Newtonian iteration number page 38
s solid phase page 14
v vapour phase page 35
w water phase page 35
wet “wet” value of parameter (Sl = 1) page 130

xiii
Chapter 1

Introduction

The vast majority of the world’s unfrozen fresh water is in the form of groundwater. Surface

freshwater sources such as lakes and rivers contain less than 4% of the volume present in

groundwater (Heath, 2004). Despite the importance of groundwater as a natural resource,

it is only recently that concerted efforts to protect groundwater quality have emerged.

Pollutants can enter groundwater by a multitude of pathways, including septic tanks,

landfills, waste lagoons, leaking storage tanks, urban runoff, and agricultural applications

(e.g. Fetter, 1993). Particularly problematic are non-aqueous phase liquids (NAPLs) such

as chlorinated solvents. These compounds, which are largely immiscible with and insoluble

in water, constitute widespread and persistent sources of contamination. Often trapped in

place by capillary barriers, NAPLs dissolve extremely slowly and may be a source of ground-

water contamination for hundreds of years (e.g. Kueper et al., 2003). Non-aqueous phase

liquids that are more dense than water (DNAPLs) present an especially difficult challenge

for removal. DNAPLs can migrate to great depths, often entering bedrock. Fractured rock

impacted by DNAPLs is considered to be among the most difficult environments to clean

up (National Research Council, 1994).

In recognition of this difficulty, recent efforts to mitigate risks from difficult sites have

been focused on either containment or in-situ destruction of the contaminant. In the past
1
CHAPTER 1. INTRODUCTION 2

decade, a strong shift was observed towards in-situ destruction methods such as chemi-

cal oxidation, thermal remediation, and bioremediation, and away from extraction-based

methods such as pump-and-treat, surfactant flushing, and air sparging (Simon, 2006).

Recently, several techniques have been developed to use heat to assist remediation ef-

forts. Heat can be useful through its effect on a variety of physical and chemical processes.

An increase in temperature typically causes a lowering of both NAPL viscosity and interfa-

cial tension with water, allowing pooled NAPL to be more easily removed. Most commonly,

however, in-situ thermal techniques seek to increase mass transfer out of the source zone

without moving the NAPL. For most organic contaminants, the processes of diffusion, disso-

lution, vapourization, volatilization, and desorption are all enhanced at higher temperatures

(Davis, 1997). At very high temperatures, contaminants may be destroyed in-situ through

the processes of oxidation and pyrolysis (Baker and Kuhlman, 2002).

Several approaches exist to deliver heat to the subsurface. One approach is to pass

an electrical current through the subsurface, generating heat through electrical resistance.

Electrical resistive heating has been used commercially at approximately fifty sites (Beyke

and Fleming, 2005). Because the pore water provides the bulk of the electrical conductance,

practical application of electrical resistive heating is limited to temperatures at the boiling

point of water. This temperature may be sufficient to vapourize many common volatile or-

ganic contaminants such as chlorinated solvents and gasoline compounds, but is inadequate

to initiate full boiling in less volatile compounds such as polychlorinated biphenyls (PCBs)

and polycyclic aromatic hydrocarbons (PAHs).

Heating has also been performed by the injection of steam into the subsurface. Steam

injection has been used at several research sites (e.g. Newmark et al., 1998). One limitation

of steam injection is its inability to penetrate low-permeability regions, although these

regions may be heated by conduction from adjacent high-permeability regions (Gudbjerg

et al., 2004). However, recent research has shown that it may be difficult to significantly

raise the temperature of a rock mass using steam injection alone (Davis et al., 2005).
CHAPTER 1. INTRODUCTION 3

Thermal conductive heating (TCH) systems are capable of reaching high temperatures

through the direct heating of the subsurface material. Arrays of “thermal wells,” each

consisting of an electrical resistive heating element inside of a steel casing, are installed

throughout a treatment area. The heater wells may reach temperatures as high as 800 ◦ C,

creating strong thermal gradients and allowing for a high rate of heat delivery. Heating

is continued until the boiling point of the target compound is reached. The produced

vapours are collected in vacuum extraction wells and treated ex-situ. In some cases, target

compounds may be destroyed in-situ through breakdown reactions such as pyrolysis and

oxidation (Stegemeier and Vinegar, 2001).

All methods of thermal remediation are potentially limited by the flow of heat away

from the treatment area. Conductive losses occur at the boundaries of the treatment area,

as heat diffuses outward towards lower-temperature areas. Groundwater flow through the

treatment area may cause convective losses of heat, as heated water is replaced by incoming

cool groundwater, or incoming cool groundwater is boiled.

Very little research exists about the significance of these convective heat losses from the

treatment area (National Research Council, 2004). In one study, numerical modelling was

used to study the influence of groundwater influx on the rate of contaminant removal in a

saturated porous medium (Elliott et al., 2003). It was found that the success of treatment

was highly dependent on the amount of groundwater flow into the treatment area. Another

modelling study examined the effectiveness of impermeable barriers in managing ground-

water influx (Elliott et al., 2004). No similar studies have been conducted for fractured rock

environments. Groundwater flow in fractured rock has the potential to be rapid and highly

localized, relative to flow in unconsolidated deposits. For this reason it is suspected that

the influence of convective cooling in fractured rock environments may be different than

in unconsolidated deposits. Despite this, almost no research has been conducted into heat

transfer in fractured rocks at small scales where equivalent continuum and dual-continuum

models may not hold (National Research Council, 1996).


CHAPTER 1. INTRODUCTION 4

1.1 Research Objectives

The objective of this study was to use analytical and numerical modelling to determine if

there are cases in which incoming groundwater may inhibit heating of a fractured rock mass

using TCH and, if so, if this effect may be overcome through careful consideration in the

system design.

Throughout this study, a conceptual model was used wherein a body of fractured rock

was considered to have a system of parallel, equally-spaced discrete fractures. Using this

conceptual model, new semi-analytical solutions were developed to the governing equations

of heat transfer. Using one of these solutions, a sensitivity analysis was performed to

determine the relative importance of rock material properties and hydrogeological properties

on the rate of heating in the treatment zone at temperatures up to the boiling point of water.

Three material properties were studied: rock density, thermal conductivity, and specific heat

capacity, along with three hydrogeological properties: fracture aperture, fracture spacing,

and hydraulic gradient.

Using the results of this sensitivity analysis, numerical modelling was performed to

assess the performance of thermal conductive heating in fractured rock at temperatures

above the boiling point of water, where latent heat effects and two-phase flow preclude

the development of semi-analytical solutions. In this analysis, the rock matrix is no longer

considered to be completely impermeable, permitting an investigation of the importance of

matrix permeability and porosity.

1.2 Organization

In the following chapter, a literature review is presented that summarizes much of the

research to date on heat transfer in fractured rock, specifically in the context of thermal

conductive heating systems. Chapter 3 is a stand-alone manuscript intended for publication


CHAPTER 1. INTRODUCTION 5

in Advances in Water Resources and describes a sensitivity analysis of material and hydro-

geological properties, performed using a new semi-analytical solution to the heat equation.

Derivations and verifications of the new solution are provided in Appendices A and B.

Discussion of the numerical modelling is given in Chapter 4, which is a manuscript

prepared for publication in Ground Water. Supporting information for this work is included

in Appendices C and D.
Chapter 2

Background

The process of thermal conductive heating in fractured rock is a complex process, drawing

on research in a variety of fields. This chapter provides an overview of this research, which is

lumped into several categories. First, a discussion is given of the fundamental concepts and

equations used to describe heat transfer, with a particular focus to hydrogeological systems.

Next, a summary is provided of the thermal properties of rock and their relationship with

physical properties, mineralogical properties, and thermodynamic conditions. An overview

of thermal conductive heating is then provided, with attention given to the many physical

and chemical processes proposed to contribute to the removal of contaminants from the

subsurface. Finally, a summary is given of previous analytical and numerical modelling

efforts focused on heat transfer in fractured media. Although very few studies have given

explicit consideration to thermal conductive heating, relevant research is presented from

the fields of nuclear waste disposal and “hot dry rock” geothermal systems.

6
CHAPTER 2. BACKGROUND 7

2.1 Heat Transfer

2.1.1 Conduction

The transfer of heat by conduction is perhaps the simplest and most direct form of heat

transfer. Conductive heat transfer occurs at the molecular scale, by means of collisions

and interactions between molecules at different energy states. In nonmetallic solids these

interactions typically take the form of lattice vibrational waves. Heat may also be conducted

by the movement of free electrons, this behaviour being characteristic of metals (Incropera

and DeWitt, 2002).

Rigorous quantitative study of heat conduction began with Joseph Fourier’s 1822 pub-

lication of The Analytical Theory of Heat (Fourier, 1955, translation). In this work Fourier

presented the basic law concerning the rate of heat movement at a steady state, now known

as Fourier’s Law:
dT
qh ∝ (2.1)
dn

where T is the temperature, and qh is the heat flux in a direction n. Fourier also noted

that the rate of heat conduction is dependent on the material through which it travels.

Thermal conductivity, written as K in this text, can be inserted in to the above relation as

a proportionality constant, giving:


dT
qh = −K (2.2)
dn

Like specific heat, thermal conductivity is often assumed to be a constant, though in

many cases it exhibits a dependence on temperature. At very high temperatures, solids

may also conduct heat radiatively – transferring energy by radiation between the particles

of the lattice or matrix. This will cause the apparent thermal conductivity to increase.

An equation to describe transient heat conduction can be constructed by performing a

heat balance on a small control volume, using Fourier’s Law to account for heat flow through

the surfaces of the volume. The result is a second-order partial differential equation, often
CHAPTER 2. BACKGROUND 8

called the “heat diffusion equation” or simply “the heat equation”:

∂T
∇ · (K∇T ) + g = ρc (2.3)
∂t

where g represents heat generated internally, ρ is the density , and c is the gravimetric

specific heat capacity. If the heat is conducted through an isotropic medium (in which the

thermal conductivity is equal in all directions) then the equation can be expanded as follows

in three-dimensional Cartesian coordinates:

     
∂ ∂T ∂ ∂T ∂ ∂T ∂T
K + K + K + g = ρc (2.4)
∂x ∂x ∂y ∂y ∂z ∂z ∂t

If the medium is also homogeneous, having a spatially constant value of thermal con-

ductivity, then the equation may be further simplified:

∂2T ∂2T ∂2T g ρc ∂T 1 ∂T


2
+ 2
+ 2
+ = = (2.5)
∂x ∂y ∂z K K ∂t α ∂t

where α represents the thermal diffusivity, the ratio of thermal conductivity to volumetric

heat capacity:
K
α=
ρc

2.1.2 Convection

Heat may also be transferred through the movement of heated matter. For example, heat

is transferred by convection when hot water flows through a pipe. Convective heat transfer

is further classified into free and forced convection. When water is heated, its density

decreases, causing it to expand and rise. As the water rises, heat is transferred upwards by

free convection. If, in contrast, the warm water is injected through a pumping well, heat is

transferred away from the well by forced convection.

The convective heat flux is directly proportional to the mechanical flux of the fluid (qm ),
CHAPTER 2. BACKGROUND 9

and is given by qm cρT . When a fluid moves with a constant velocity, the effects of forced

convection can be modelled by the addition of this convective flux into the heat equation:

∂T
∇ · (K∇T ) − ∇ · (ρcqm T ) + g = ρc (2.6)
∂t

2.1.3 Radiation

All objects constantly emit energy as electromagnetic radiation, of an intensity determined

by the object’s temperature and surface characteristics. This energy can be absorbed by

other objects, also as a function of their temperature and surface characteristics. The

transfer of heat by electromagnetic radiation is termed radiative heat transfer.

Not all electromagnetic radiation that reaches a body is absorbed. The energy may be

either absorbed, reflected, or transmitted through the body, at fractions described by the

absorptance Θ , reflectance Λ, and transmittance Ω. These coefficients, whose sum must

equal unity, will vary depending on the wavelength of the incoming radiation.

The Stefan-Boltzmann law describes the maximum amount of energy that may be emit-

ted from a body at a given temperature. The law, which is valid for a theoretical “black

body” having Θ = 1, predicts a maximum energy flux (ξ) of ξ = BT 4 per unit emitting

area. The proportionality constant B is referred to as the Stefan-Boltzmann constant and

is equal to 5.67 × 10−8 W/m2 ·K4 .

When radiation is exchanged between two objects, the rate of heat transfer is a function

of the difference in the fourth power of their temperatures. The heat flux from object 1 to

object 2 is given by ξ(T1 ), while the heat flux from object 2 to object 1 is given by ξ(T2 ).

The net heat flux from object 1 to object 2 is thus described by ξ(T1 )−ξ(T2 ) = B T14 − T24 .


The actual rate of heat transfer from object 1 to object 2 is a function of the heat

fluxes of the two objects, the area through which heat is emitted and absorbed, and their

configuration in space. The rate can be calculated according to:


CHAPTER 2. BACKGROUND 10

QR = A1 F1−2 B T14 − T24



(2.7)

where A1 is the area from which radiation is emitted and absorbed, and F1−2 is the

configuration factor, a function of the emittances and absorptances of the two objects, as

well as their arrangement in space.

2.1.4 Heat Transfer in the Subsurface

A variety of mechanisms contribute to the transfer of heat in a porous medium, making it

a difficult system to model mathematically. Bear (1972) identifies six mechanisms of heat

transfer in a porous medium:

1. Heat transfer through the solid phase (considered as a continuum) by

conduction;

2. Heat transfer through the fluid phase (considered as a continuum) by con-

duction;

3. Heat transfer through the fluid phase (considered as a continuum) by con-

vection;

4. Heat transfer through the fluid phase by dispersion;

5. Heat transfer from the solid phase to the fluid phase;

6. Heat transfer between solid grains by radiation, when the fluid is a gas.

Local thermal equilibrium

At the pore-scale, temperature differences may exist between soil grains and pore fluids.

However, these temperature differences are typically small and require consideration only

when rapid thermal transients or highly concentrated heat sources are present in one phase
CHAPTER 2. BACKGROUND 11

but not the other (Kaviany, 1995). In porous media, where fluids move quite slowly, it is

appropriate to assume local thermal equilibrium between the fluid and solid phases. This

very useful assumption permits the definition of an effective medium thermal conductivity,

removing the need to separately examine heat conduction in the solid and liquid phases.

In fractured media, fluid is typically assumed to be in thermal equilibrium with the

fracture walls, although temperature gradients may be present within the rock matrix itself.

In some geologic systems, such as geysers and phreatic eruptions, the assumption of local

thermal equilibrium may be inappropriate (Ingebritsen et al., 2006).

Conduction and convection

In the heat equation presented above (2.6) the effect of convection was included in the

∇ · (ρcqm T ) term. In environments where large temperature differences are not present,

flow equations can be solved to determine the fluid mechanical flux qm . This approach has

the advantage of preserving the linearity of (2.6); however, it may be a source of error if

large temperature differences are present.

If the fluid flow is assumed to follow Darcy’s Law, then the flux term qm can be expanded

as:
k
qm = − (∇P − ρg) (2.8)
µ

where k is the permeability of the medium, µ is the dynamic fluid viscosity, P is the fluid

pressure, and g is the gravitational acceleration vector. Because the fluid dynamic viscosity

and density are dependent on temperature, the rate of fluid flow will be affected by changes

in temperature. Further, the increase in fluid density that accompanies an increase in

temperature will cause an increase in fluid pressure. In high temperature environments,

these fluid property variations can not be ignored (Ingebritsen et al., 2006).

Conduction and convection are typically the dominant forms of heat transfer in porous

media, to the extent that other forms are typically neglected (Domenico and Schwartz,
CHAPTER 2. BACKGROUND 12

1998). For example, radiation is insignificant at temperatures below 600 ◦ C, and dispersion

is almost always regarded as negligible due to the rapid rates of heat diffusion in geologic

materials (Ingebritsen et al., 2006). The Peclet number, a dimensionless parameter, can be

used to describe the relative importance of conduction and convection (Bear, 1972):

Lv
Pe = (2.9)
α

where L is a characteristic length, often taken as the mean grain diameter, v is the fluid

velocity, and α is the thermal diffusivity. As Pe increases, so does the importance of heat

convection.

Heat dispersion

Differences in pore size, shape, and friction can cause microscale variations in fluid velocity.

These variations tend to cause fluid to spread out, or disperse, as it moves along a path

defined by the hydraulic gradient. Hydrodynamic dispersion plays a large role in the transfer

of solute throughout the subsurface, and the same mechanism affects heat transport through

convection. However, because heat diffusion (conduction) is so rapid relative to molecular

diffusion, the effect on heat transport is far less significant than for solute transport, and is

typically neglected (Ingebritsen et al., 2006).

Radiative heat transfer

At temperatures about 600 ◦ C, radiation can begin to play a significant role in inter-grain

heat transfer. This effect is typically modelled by adding a radiative contribution to the

thermal conductivity of the medium (Clauser and Huenges, 1995; Ingebritsen et al., 2006).
CHAPTER 2. BACKGROUND 13

2.2 Thermal Properties of Rock

As seen in equation (2.3), the rate of transient heat conduction through rock is a function of

both thermal gradient and thermal diffusivity. Though thermal diffusivity may be measured

directly in a laboratory, historical research has tended to focus on the thermal conductivity

of rocks, and to a lesser extent, their heat capacity. The following section provides a

summary of this research and its implications for heat transfer in rocks.

2.2.1 Thermal Conductivity

A large range of thermal conductivities can be observed in geologic materials. An impure

diamond may have a thermal conductivity on the order of 1000 W/m·K, while many soils

have thermal conductivities below 1 W/m·K(e.g. Abu-Hamdeh and Reeder, 2000). Com-

mon thermal conductivities for rocks range from 1 to 6 W/m·K. Most rocks have anisotropic

thermal conductivity, exhibiting a preference for heat flow in a particular direction. This is

typically due to the preferential alignment of crystals within the rock. However, rocks that

have no preferred orientation of mineral grains may behave as an isotropic body (Bunte-

barth, 1984).

Much of the early data on the thermal conductivity of rocks comes from a two-part 1940

paper by Harvard University researchers Birch and Clark (1940a). In addition to providing

a wealth of new thermal conductivity measurements, the paper provides a comprehensive

description of the mechanisms of heat transfer in crystals, as predicted by Debye theory

(Birch and Clark, 1940b).

The thermal conductivity of rocks is neither constant nor readily predicted from simple

measurements such as density. Nonetheless, a large body of research has elucidated the

major factors influencing the thermal conductivity of rocks, and a number of relationships

have been proposed to permit a degree of extrapolation from existing measurements. The

most reliable estimate of a rock’s thermal conductivity, apart from direct measurement, is
CHAPTER 2. BACKGROUND 14

given by its mode, or mineral composition (Buntebarth, 1984; Birch and Clark, 1940b).

The thermal conductivity of sedimentary rocks is largely dependent on porosity and

fluid saturation. Crystalline rocks, on the other hand, are more sensitive to mineral grain

orientation, temperature, and pressure effects. These relationships are discussed in greater

detail later in this section. More comprehensive presentations are given by by Robertson

(1988) and Clauser and Huenges (1995).

Dependence on Porosity and Fluid Saturation

Increases in fluid saturation typically result in increased thermal conductivity, as water

begins to occupy pore spaces formerly filled with air (e.g. C̆ermák and Rybach, 1982). As

a result of increasing interest in thermally enhanced oil recovery, petroleum researchers in

the 1950s began conducting extensive research into the relationships between the thermal

conductivity of porous rocks samples and their porosity and fluid saturation. These parame-

ters were found to correlate well with thermal conductivity, and a number of equations were

developed to describe this dependence. The result was a series of “mixing law” formulas

that predict the thermal conductivity of a porous rock, given the thermal conductivities of

its solid and fluid components.

In a paper outlining many of the mixing law models, Woodside and Messmer (1961)

used an analogy to electric circuits to describe the two limiting cases of the mixing law

models. The minimum value of thermal conductivity is described by a series model, where

heat passes from the solid phase, then to the liquid phase. This is mathematically described

by the weighted harmonic mean of the two conductivities, or:

Kmin = Ks Kf /[φKs + (1 − φ)Kf ] (2.10)

where φ is the porosity and the subscripts f and s denote the fluid and solid phases, re-

spectively. The maximum value of thermal conductivity corresponds to the parallel circuit
CHAPTER 2. BACKGROUND 15

model, mathematically expressed as the weighted arithmetic mean of the conductivities:

Kmax = φKf + (1 − φ)Ks (2.11)

Woodside and Messmer (1961) also proposed the use of the geometric mean to describe the

conductivities, based on the mathematical fact that Kharmonic < Kgeometric < Karithmetic :

Kgeo = Kfφ Ks1−φ (2.12)

While Woodside and Messmer (1961) do not propose a physical justification for the use

of the geometric mean thermal conductivity, it is appears to be one of the more widely

used mixing law models (Beck, 1976). Other researchers have extended the analogy of

electrical conductivity, developing equations based on Maxwell’s theoretical equation for

the electrical conductivity of a system consisting of solid spheres randomly distributed

throughout a continuous medium at sufficient spacing so that they do not interact (Woodside

and Messmer, 1961; Beck, 1976). Translated into thermal quantities, this is expressed as:

 
2φKf + (3 − 2φ)Ks
K = Kf (2.13)
(3 − φ)Kf + φKs

Kunii and Smith (1960) developed a mixing law model for porous sandstones based on a

theoretical packed bed reactor, with an empirically determined “consolidation parameter”

to account for the additional conduction between the cemented grains. In studies where

predicted and measured thermal conductivities have been compared, most of the mixing

law models have been found to be able to predict thermal conductivity to within 10-15%

accuracy, depending on the data set (Clauser and Huenges, 1995).


CHAPTER 2. BACKGROUND 16

Dependence on Temperature

Since at least the early 20th century, researchers have been aware of the significant cor-

relation between temperature and thermal conductivity measurements in rocks. However,

much of the early research into the thermal properties of rocks is of limited quantitative

value, due to poor characterizations of both the rock samples and the ambient conditions

(temperature, pressure) during testing. An extensive study published by Birch and Clark

(1940a) set a new standard for research into the thermal conductivity of rocks, tabulating

thermal conductivity values over the 0 ◦ C to 400 ◦ C range for a series of well-characterized

rock samples.

A more recent compilation of the now-extensive experimental data is provided by Clauser

and Huenges (1995). As seen in this work, which includes the data of Birch and Clark

(1940a) as well as many others, the thermal conductivity of most rocks decreases by a

factor of 1.5 to 4 as temperature is increased from 0 ◦ C to 800 ◦ C. The temperature depen-

dence of thermal conductivity is generally largest for sedimentary and metamorphic rocks,

and weakest for plutonic and volcanic rocks. For some rocks, the temperature-dependent

variation in thermal conductivity may be large enough to be significant in the modelling of

terrestrial heat flow, geothermal reservoirs, and subsurface heating for soil and groundwater

remediation. This dependence on temperature presents difficulties in the modelling of heat

conduction, as it introduces non-linearity to the heat equation (e.g. Carslaw and Jaeger,

1959).

According to C̆ermák and Rybach (1982), two separate effects explain the change in

thermal conductivity with increasing temperature. Lattice conductivity, mentioned in Sec-

tion 2.1.1 as the primary mechanism of conduction in non-metals, is inversely proportional

to T . Theory predicts that the radiative component of thermal conductivity should increase

with T 3 , although empirical evidence suggests that the increase is proportional to T rather

than T 3 and does not become significant until temperatures are high (Buntebarth, 1984).
CHAPTER 2. BACKGROUND 17

Exceptions are present to the general rule of decreasing thermal conductivity with tem-

perature. The thermal conductivities of rocks with high feldspar content are often unaf-

fected or may even increase with rising temperature. In addition, rocks with an amorphous

phase, such as fused (noncrystalline) quartz may have thermal conductivities that rise with

temperature (Ratcliffe, 1959; Abdulgatov et al., 2006; C̆ermák and Rybach, 1982). Several

researchers have proposed equations to relate the conductivity a rock at an elevated tem-

perature to its conductivity at a lower temperature. For crystalline rocks, Seipold (1998)

proposed an expression of the form:

T
K(T ) = (2.14)
F ×T +E

where E and F are constants. Vosteen and Schellschmidt (2003) proposed an a slightly

more complex relationship:


K(0)
K(T ) = a−b
(2.15)
0.99 + T K(0)

where a and b are constants.

Dependence on Pressure

Several references provide data on the thermal conductivity of rocks as a function of pres-

sure. In the data compiled by Clauser and Huenges (1995), increasing pressure generally

causes an increase in thermal conductivity up to about 20%, as fractures are closed and

fluid is forced out of the rock. Above 15 MPa, little change is observed in the thermal

conductivity. C̆ermák and Rybach (1982) suggest that the pressure effect is somewhat

less important, observing a maximum thermal conductivity increase of 10%. A more pro-

nounced effect was observed by Woodside and Messmer (1961) in tests on a sandstone with

φ = 22%. Increases in thermal conductivity were observed until approximately 20 MPa, by

which time the thermal conductivity had risen by about 35%. Combining their original data
CHAPTER 2. BACKGROUND 18

on pressure dependence with that from a compendium of previous works, Abdulgatov et al.

(2006) propose a much higher range of 50 MPa to 100 MPa for the “cross-over pressure” at

which further pressure increases cease to affect thermal conductivity significantly. In sum-

mary, the extent of the pressure effect is strongly dependent on the minerology, porosity,

and density of the rock sample. Although the value of the cross-over pressure varies widely

between data sets, there seems to be agreement that there is an upper limit to the pressure

effect on thermal conductivity.

Dependence on Orientation

Many rocks are thermally anisotropic, showing a preferred direction of thermal conductivity.

Mineral-anisotropy, or microanisotropy, refers to anisotropy in rock thermal conductivity

that arises from a dominant orientation of an anisotropic mineral. Shape-anisotropy, or

macroanisotropy, arises from bedding and foliation. For example, a metamorphic rock may

have alternating layers of quartz and feldspar, causing a dependence on the direction of

foliation (Vosteen and Schellschmidt, 2003).

C̆ermák and Rybach (1982) define the anisotropy factor as the ratio of parallel thermal

conductivity to normal thermal conductivity, or:

Kp
af =
Kn

2.2.2 Heat Capacity

The specific heat of rocks tends to rise with temperature (Robertson, 1988). Because

the thermal diffusivity of a rock is calculated as the thermal conductivity divided by the

volumetric heat capacity, an increase in c has the same effect as a decrease in thermal

conductivity. The temperature effects in both c and K generally drive the thermal diffusivity
CHAPTER 2. BACKGROUND 19

in the same direction - toward a slower conductive transfer of heat at higher temperatures.

For multi-phase systems, such as porous rocks saturated with water, aggregate specific

heat can be calculated simply from a weighted arithmetic mean of the specific heats of the

different components. Rocks with large water contents may have very high values of specific

heat, due to the very high specific heat of water.

The same reasoning can be used to calculate the specific heat of a rock from the specific

heats of its mineral components. Somerton (1958) performed a chemical analysis on several

sedimentary rocks to determine their mineral composition, then computed a value for the

specific heat of the rock based on the specific heats of the minerals. Heat capacities of the

rocks were also measured directly using a calorimeter, and the calculated and measured

values agreed within 2%. According to Robertson (1988), for most purposes the specific

heat capacity can be calculated with reasonable accuracy from an estimate of the mineral

composition determined only by hand-lens inspection of the rock.

2.3 Subsurface Remediation by Thermal Conductive Heat-

ing & Soil Vapour Extraction

In-situ Thermal Desorption, or ISTD, is a method of groundwater and soil remediation

in which a combination of conductive heating (TCH) and soil-vacuum extraction (SVE)

is used to treat subsurface contamination. Outside of the United States and Canada, the

combination of thermal conductive heating and vacuum extraction is occasionally referred

to as TEVE (thermally enhanced vacuum extraction) or THERIS (thermischen In-Situ

Sanierung). More recently, the term “thermal conductive heating” has been used to refer

to the combination of conductive heating and vacuum extraction.

In a TCH system, heat is delivered to the formation by heater elements installed in steel

casings. Heat is transfered by radiation from the heater element to the casing, from where

it can conduct into the formation. Conduction is the primary mechanism of heat transfer
CHAPTER 2. BACKGROUND 20

away from the thermal wells (Stegemeier and Vinegar, 2003). However, in formations with

significant movement of groundwater, convection may also play a role in heat transfer. Very

close to the heater wells, where the temperature may exceed 600 ◦ C, radiation may make

an additional contribution to the heat transfer away from the well.

Thermal conductive heating is relatively insensitive to formation heterogeneities, as the

thermal conductivities in the subsurface vary by a factor of less than five. By contrast, the

transfer of heat into the subsurface by steam injection is limited by the intrinsic permeability

of the soil and rock, which may vary by over ten orders of magnitude (Ingebritsen et al.,

2006).

Heater wells are typically installed in a hexagonal configuration. In some cases, the well

at the centre of the hexagon may be a heater-vacuum well, while the wells along the edges

of the hexagon are heater-only wells. In other cases, all of the wells may be heater-vacuum

wells. In another approach, the combination heater-vacuum wells may be omitted from the

design, instead allowing contaminants to migrate upwards through a permeable sandpack

installed around the heater-only wells (LaChance et al., 2006). The sandpack can be used to

direct contaminant vapours to horizontal vacuum extraction wells, where they are extracted

and directed to the treatment facility. An example TCH configuration is shown in Figure

2.1.

Spacing of the wells is determined by the amount of time available for treatment, as

well as the target temperature for treatment of the compounds of interest. At the upper

end of the treatment temperature range, semivolatile organic compounds (SVOCs) such as

polychlorinated biphenyls (PCBs) require treatment temperatures in the range of 300 ◦ C

to 400 ◦ C (Uzgiris et al., 1995). In this case a well spacing of 2-3 m is typically used.

More volatile compounds such as chlorinated solvents (CVOCs) can be treated at 100 ◦ C,

allowing for a well spacing of 5 m to 8 m (LaChance et al., 2006). Treatability studies on

polycyclic aromatic hydrocarbons have found the level of removal to be affected by both

the treatment temperature and the treatment time (Hansen et al., 1998).
CHAPTER 2. BACKGROUND 21
North Adams-MGP Site.jpg (JPEG Image, 1640x642 pixels) - Scaled (75%) http://www.terratherm.com/site%20Photos/Well%20Field%20Overvie...

Figure 2.1: Photograph of thermal wells at a field site. (TerraTherm photo)

Although remediation using thermal conductive heating is energy intensive, it applica-

tion may represent a long-term savings in energy relative to treatment methods that operate

over a longer period of time. Using a life-cycle analysis methodology, Hiester and Schenk

(2005) compared the energy use of a “cold” soil vapour extraction system with a combined

TCH-vapour extraction system. They found that, although the thermally enhanced system

required much more power to operate, it used 58% less energy than the “cold” due to its

shorter operation time.

2.3.1 Mechanisms of Remediation

Heat can aid in the removal of contaminants from the subsurface through its effect on a

number of physical and chemical processes (e.g. Davis, 1997). Physical changes resulting
1 of 1 elevated temperatures can include enhanced contaminant desorption 24/04/2008
from from soil, 12:25 PM
in-

creased dissolution, increased vapourization and volatilization, increased rates of diffusion,

and an increase in soil permeability. On a chemical level, heat contributes to contaminant

breakdown through oxidation and pyrolysis.


CHAPTER 2. BACKGROUND 22

Not all effects of added heat may be beneficial to remediation. NAPL viscosity and

interfacial tension with water both tend to decline with increasing temperature, increasing

the mobility of NAPL present. In a worst-case scenario, this could enable NAPL to migrate

downwards, spreading the source zone.

A number of these effects are reported to have significant effects on contaminant removal

during subsurface heating by TCH (Baker and Kuhlman, 2002). However, little research has

been conducted to demonstrate the relative importance of these mechanisms in a subsurface

environment.

Desorption of Contaminants from Soil

Thermal desorption has long been recognized as an effective means of extracting contami-

nants from soil, allowing for subsequent treatment of the vapours by combustion or other

means. In earlier applications of thermal desorption, soil was excavated prior to thermal

treatment in a rotary kiln or other ex-situ heating device (Lighty et al., 1988). The applica-

tion of thermal desorption in-situ is more recent. Iben et al. (1996) used heater blankets to

conductively deliver heat to a shallow area of PCB contamination, reporting a thousand-fold

reduction in PCB concentration after 24 hours of heating. The kinetics of thermal desorp-

tion are complex, with a portion of the contaminant mass being released rapidly from the

soil, and a “recalcitrant fraction” requiring a much longer treatment time (Uzgiris et al.,

1995).

Increased Dissolution

When thermal treatment is applied to a NAPL source zone, the rate of overall mass transfer

may be affected by increased partitioning from the NAPL phase into the aqueous phase.

Knauss et al. (2000) presented new data on the solubility of TCE and PCE at temperatures

ranging from 21 ◦ C to 117 ◦ C and 22 ◦ C to 161 ◦ C, respectively. For both TCE and PCE,

solubility is increased by a factor of approximately 2 when the system is heated to just


CHAPTER 2. BACKGROUND 23

below the boiling point of water.

Increase in NAPL mobility

She and Sleep (1998) measured imbibition and drainage curves for a PCE-water system at

varying temperatures and found that water-PCE capillary pressure decreased by a factor of

about two when temperature was increased from 20 ◦ C to 80 ◦ C. A variety of mechanisms

may contribute to capillary pressure lowering, including changes in NAPL-water interfacial

tension, NAPL-water contact angle, residual wetting phase saturation, and the presence of

small amounts of vapour in soil pores (She and Sleep, 1998).

Vapourization and Steam Stripping

At equilibrium, the concentration of contaminant in the vapour phase is related to the

concentration of dissolved contaminant by the Henry’s Law constant: Hcc = Cg /Cw . Heron

et al. (1998a) found that the Henry’s Law constant for TCE increased by a factor of 20

when temperature was increased from 10 ◦ C to 95 ◦ C.

Consequently, as steam is produced, the concentration of contaminant in the gas phase

will decrease, while the concentration in the aqueous phase is relatively unchanged. In

order to maintain equilibrium, the rate of volatilization will increase (e.g. Fair, 1987). This

phenomenon is known as “steam stripping.”

In-Situ Contaminant Destruction: Pyrolysis

When subjected to very high temperatures, many gaseous phase organic contaminants will

undergo pyrolysis, a unimolecular decomposition reaction. Interpretations of field studies

of TCH have suggested that pyrolysis reactions may occur in the superheated zones directly

adjacent to thermal wells (Baker and Kuhlman, 2002). The pyrolysis of chlorinated solvents

can be expressed by the following reaction (e.g. Frenklach, 1990):


CHAPTER 2. BACKGROUND 24

Cx Hy Clz → Cx Hy−1 Clz−1 + HCl

Baker and Kuhlman (2002) write the overall reaction for the complete pyrolysis of TCE

in the presence of water as follows:

C2 HCl3 + 4H2 O → 2CO2 + 3HCl + 3H2

The first equation provides a more complete description of the pyrolysis mechanism.

Dehalogenation does not occur in one step; while the original contaminant may be almost

completely removed, it will likely be transformed into a number of intermediary products

in addition to CO2 , HCl, and H2 .

Yasuhara (1993) and Yasuhara and Morita (1990) studied the pyrolysis of PCE at tem-

peratures between 310 ◦ C and 940 ◦ C and of TCE at temperatures between 300 ◦ C and

800 ◦ C. Measurable amounts of 15 compounds were detected from the pyrolysis of PCE,

while 23 were detected from the pyrolysis of TCE.

A similar study was performed by Thomson et al. (1994) on the high-temperature ox-

idation of 1,1,1-trichloroethane. Vapour phase 1,1,1-TCA gas was burned at a controlled

temperature. It was found that, at temperatures near to 530 ◦ C, significant amounts of

1,1,1-TCA were converted into phosgene (COCl2 ), a toxic gas. Phosgene concentration

increased as temperature was increased, reaching a peak at a temperature of 890 ◦ C, when

fully 29% of the 1,1,1-TCE was converted into phosgene in 0.1 s. Only at temperatures

exceeding 1200 ◦ C did phosgene cease to be a major oxidation product. These results

suggest that phosgene may be a product of in-situ combustion under thermal remediation

conditions (Costanza et al., 2003).

Conditions favourable to pyrolysis may occur in the superheated zone that occurs close

to heater wells. Costanza (2005) calculated that gaseous TCE will be 99% destroyed in

the superheated zone with a residence time of about 7 days at 500 ◦ C or with a residence
CHAPTER 2. BACKGROUND 25

time of 7 seconds at 700 ◦ C. Assuming a 1-foot superheated zone with no preferential flow

pathways, the residence time of vapours in the superheated zone was calculated to be on the

order of 30-70 seconds for typical gas extraction rates. It should be noted that treatment in

the superheated zone is not a necessity; vapours extracted from the subsurface are treated

above-ground.

While destruction of contaminant by pyrolysis reduces the need for above-ground treat-

ment, the formation of gaseous HCl can cause excessive corrosion in the vapour extraction

system. The problem may be even more severe if the HCl vapours are allowed to condense,

as HCl in the liquid form may be 20 times more corrosive than in the gaseous form (U.S.

EPA, 2002).

In its analysis of the failure of an ISTD system installed at the Rocky Mountain Arsenal

“ hex pit”, the Rocky Mountain Arsenal Remediation Venture Office describes the complete

degradation of heater elements, casings, and collection equipment as a result of corrosion

from condensed HCl vapours (U.S. EPA, 2002).

In-Situ Contaminant Destruction: Oxidation

In addition to the destructive reactions presented in the previous section, the elevated

temperatures present in thermal remediation conditions can enhance in-situ oxidation re-

actions. The following reaction for the aqueous oxidation of TCE is presented by Knauss

et al. (1999):

2C2 HCl3 (aq) + 3O2 (aq) + 2H2 O → 4CO2 (aq) + 6H+ (aq) + 6Cl− (aq)

While this reaction is thermodynamically favoured (∆Gr < 0), the reaction rate con-

stant is so low at standard groundwater temperatures that the process becomes negligible.

However, Knauss et al. (1999) report that the rate constant increases by a factor of about

2500 from 25 ◦ C to 90 ◦ C.
CHAPTER 2. BACKGROUND 26

Increase in Soil Permeability

At very high temperatures, soils may undergo structural changes that affect the intrin-

sic permeability. Vinegar et al. (1997) heated a PCB-contaminated soil for 42 days and

compared soil samples taken before and after heating. They reported an increase in soil

porosity from about 30% to about 40%. In addition, reported air permeabilities increased

from 3 × 10−3 md to 50 md (horizontal) and from 1 × 10−3 md to 30 md (vertical). The au-

thors propose that permeability was increased due to the removal of moisture from the pore

spaces, clay dessication, fracturing, and the removal of organic material. This increase in

permeability allowed vapour-phase contaminants to travel more rapidly to the SVE system.

Some data on thermally-induced structural changes of rocks are available from petroleum

research. For example, Somerton et al. (1965) heated several sandstone samples to 800 ◦ C

in an oven and allowed them to cool to room temperature, observing permeability increases

on the order of 200% to 400% and fracture index increases (the number of fractures observed

in a cross-section) of about 800%.

2.3.2 Heat Losses

Heat losses from the treatment area are an important consideration in the design of a TCH

system. To compensate for “end effects” – conductive losses of heat out the top and bottom

of the treatment zone – the heating elements are extended at least two feet above and below

the treatment area and may be designed to produce about 25% additional heat output in

these sections. This can be accomplished by reducing the cross-sectional area of the heater

element or adding an additional element in parallel (Stegemeier and Vinegar, 2001, 2003).

During thermal applications below the water table, heat may be lost due to the flow

of groundwater through the treatment area. When the temperature in the treatment zone

is below the boiling point of water, groundwater flow can cause heated water to flow out

of the treatment zone. The incoming cool water, which must be continually heated, can
CHAPTER 2. BACKGROUND 27

in some cases overwhelm the heater wells, preventing the attainment of boiling tempera-

tures (National Research Council, 2004). Above the boiling point of water, incoming cool

groundwater will be boiled, causing a loss of energy into the latent heat of vapourization.

Baker and Heron (2004) state that an injection of steam may also be used to prevent

groundwater influx from outside of the treatment zone. The injection of steam at high

pressure serves to provide a pressure barrier to groundwater flow and to reduce the relative

permeability of water in the formation. In addition, control of groundwater influx may be

possible through the use of extraction wells or impermeable barriers.

Modelling Studies

In one of the few published model studies on ISTD, Elliott et al. (2003) used the commer-

cially available STARS reservoir model to study the cooling influence of groundwater influx

to the treatment zone. The model comprised a two-dimensional grid with a NAPL source

zone spread between two heater wells spaced at 3.048 m, with a vacuum-heater well in the

centre of the space between the two wells. Using TCE, a heavy hydrocarbon, and a PCB,

the authors measured the heating time necessary for treatment (defined as 99% removal).

When the permeability of the simulated porous medium was increased, the treatment

time decreased, as vapourized contaminants were allowed to move more quickly through

the soil. However, this trend was reversed as permeability was increased beyond a certain

threshold (k = 9.86 × 10−11 m2 for the parameters of the study), as the increased water

influx from outside of the treatment zone prevented the soil inside the treatment zone from

reaching the treatment temperature.

A similar effect was observed as the hydraulic gradient across the treatment zone was

increased. Above a certain gradient, the heaters were unable to deliver sufficient energy to

counter the cooling influence of groundwater, and contamination was removed only from

the downgradient side of the treatment zone, after the incoming groundwater had been

preheated by the heater well on the upgradient side of the treatment zone. In order to
CHAPTER 2. BACKGROUND 28

ensure an adequate temperature rise throughout the treatment zone, the authors proposed

using a combination of tighter well spacings, higher heater temperatures, and impermeable

barriers.

The effectiveness of impermeable barriers was studied by Elliott et al. (2004), who used

a three-dimensional numerical model to study several impermeable barrier designs. A 20 m

× 20 m square was used as the model domain, constant head boundaries at the edges of

the square. At the centre of the domain, a hexagon of heater wells was placed, with a

single heater-extraction well at the centre of the hexagon. It was found that with a medium

permeability of k = 9.86 × 10−12 m2 , 99.8% of the TCE could be removed after 19 days of

heating. However, when the intrinsic permeability was increased to k = 9.86 × 10−11 m2 ,

TCE removal did not exceed 50%, even after 150 days of heating.

Using this value of intrinsic permeability, three barrier configurations were modelled to

test their effectiveness in preventing influx cooling. When a barrier was placed along one

edge of the treatment zone, the TCE removal rate was increased to slightly above 60%.

With barriers on two sides of the treatment zone, the removal rate reached 80%. When

the treatment zone was entirely enclosed by impermeable barriers, the removal rate reached

97%.

2.4 Laboratory Studies of Thermal Remediation

A number of laboratory studies have been conducted to show the ability of heat to remove

contaminants from soil, using an oven to heat a contaminated soil sample (e.g. Merino

and Bucalá, 2007; Burghardt, 2007). However, there have been very few laboratory-scale

implementations of the technologies most commonly used for in-situ thermal treatment:

steam injection, electrical resistive heating, and thermal conductive heating. Research on

these studies has primarily been conducted at the field-scale, where only limited monitoring

of the treatment area is possible.


CHAPTER 2. BACKGROUND 29

Nonetheless, a few laboratory-scale studies have been conducted. Heron et al. (1998b)

applied resistive heating to a two-dimensional soil tank filled with a TCE-contaminated

silty soil. After 37 days of heating, 99.8% of the TCE mass was removed through vapour

extraction. Gudbjerg et al. (2004) simulated steam injection into a two-dimensional soil

tank with a fine and a course layer. The injected steam was not able to enter the fine layer;

however, the fine layer was heated by heat conduction from the coarse layer, albeit at a

much longer time scale than the heating of the coarse layer.

A few laboratory studies have been conducted of thermal conductive heating and are

discussed at length in the following sections. Kunkel et al. (2006) studied the applicability

of TCH to removal of elemental mercury from soil. Hiester et al. (2004, 2006) conducted a

laboratory-scale study of TCH for the removal of trimethylbenzene from a partially satu-

rated soil.

2.4.1 University of Texas Column Studies

While field studies have quantified the removal potential of thermal conductive heating

for organic DNAPLs such as halogenated solvents and PCBs, little data is available for

other DNAPLs, such as elemental mercury. In an effort to fill this gap, researchers at the

University of Texas performed a series of thermal well experiments in a 5 cm diameter

column of dry Ottawa sand (Kunkel et al., 2006). In one set of experiments, soil in the

column was contaminated with a known mass of perfluorodecalin, a surrogate compound for

Hg(0). Air was passed through the column and the mass of contaminant in the effluent was

measured. The experiments were conducted at varying temperature and air flow rate, and

it was concluded that temperature had a far greater impact on clean-up time than did flow

rate. In addition, three trials were performed using elemental mercury as a contaminant,

and similar results were achieved. Both sets of experiments were modelled numerically using

the STARS simulator, and good agreement was seen between the predicted and measured

results. In the case of both contaminants, rapid removal was achieved at temperatures well
CHAPTER 2. BACKGROUND 30

below the boiling point.

2.4.2 VEGAS Laboratory Studies

Some larger scale laboratory studies of ISTD are ongoing at the VEGAS Research Facility

for Subsurface Remediation at the University of Stuttgart. In a recent study, an ISTD

system was used to remove 30 kg of trimethylbenzene, a semivolatile organic compound,

which had been released into a 6 m×6 m×4.5 m container filled with a partially saturated

soil (Hiester et al., 2004).

Two vacuum extraction wells were installed in the soil, on one side of the TMB source

zone. Two air inlets were installed on the opposite side of the source zone. For two months,

vapours were extracted at a rate of 35 m3 per hour. Concentrations of TMB in the vapour

remained nearly constant for the two months over which the system was run.

After two months, four heater wells installed in the periphery of the source zone were

turned on. The temperature of the heater wells remained constant, at about 500 ◦ C. After

20 days of operation, the centre area between the wells had reached a temperature of slightly

under 100 ◦ C, and all of the remaining TMB had been removed. The authors calculated that

had the contaminant been removed by SVE alone, remediation time would have exceeded 8

months. The use of heater wells thus improved remediation time significantly, and resulted

in large energy savings.

Since the initial study at VEGAS, additional studies in a modified container have been

performed (Baker et al., 2006; Hiester et al., 2006). Recent experiments have been conducted

using a constant head boundary along the perimeter of the container, permitting a study

of ISTD behaviour in the saturated zone. Currently, no contaminant has been added to

the modified container, but researchers have activated the thermal wells and monitored

the temperature distribution throughout the container. The results from the experiment,

yet unpublished, are being used to calibrate the STARS model for use in designing future

experiments.
CHAPTER 2. BACKGROUND 31

2.5 Heat Transfer in Fractured Media: Analytical Solutions

Heat transfer in fractured media is difficult to model analytically. Separate governing equa-

tions are used for the matrix, in which there is no convection, and the fracture, where heat

is both conducted and convected. The two domains are coupled by a Cauchy boundary

condition (simultaneous Type I and II) that assumes thermal equilibrium and conservation

of energy between the two domains. In this section, an overview is given of previous analyt-

ical solutions to the heat equation in this environment. Consideration is given both to fully

analytical (closed-form solutions) as well as semi-analytical solutions (in which an inverse

Laplace transform is performed numerically, etc.)

Although solution of these equations is tractable only for very simple geometries, analyti-

cal solutions allow rapid and exact calculations of temperature in these cases. Consequently,

analytical solutions have an important application in sensitivity analysis and are indispens-

able for determining the governing parameters in a particular system. Further, they provide

a reliable solution against which numerical solutions can be verified.

Several analytical solutions have been developed for the problem of heat exchange be-

tween water in a single fracture and an impermeable matrix. These solutions do not consider

heat generation within the rock matrix; the system under consideration is typically cold wa-

ter flowing through hot rock.

For the case of a single discrete fracture, solutions are provided in 2D Cartesian coor-

dinates by Lauwerier (1955), Carslaw and Jaeger (1959, p. 396), Bodvarsson (1969), Yang

et al. (1998), and Kocabas (2004). Some 3D Cartesian solutions exist, though few are

published in English. Some examples are the approximate solution of Alishaev (1979) and

the semi-analytical solution of Heuer et al. (1991), who developed solutions for an arbi-

trary flow field in a planar fracture. In addition, solutions can be readily adapted from the

equivalent problem of matrix solute diffusion; such solutions are provided by Grisak and

Pickens (1981) and Tang et al. (1981). Solutions to the heat transport problem were found
CHAPTER 2. BACKGROUND 32

in radial coordinates by Bodvarsson (1972) and to the equivalent solute transport problem

by Feenstra et al. (1984).

Extension of the problem to a series of equally-spaced parallel fractures was presented

first in the geothermal literature by Gringarten et al. (1975) and Lowell (1976), who devel-

oped Laplace-space solutions to the problem using Cartesian coordinates in two dimensions.

Bodvarsson and Tsang (1982) solved the problem in Laplace space using radial coordinates.

Sudicky and Frind (1982) presented a solution to the equivalent solute problem, using

Cartesian coordinates.

All of the aforementioned solutions assume that diffusion of heat or solute in the rock

matrix occurs only in the direction normal to the fracture plane. Even in the geothermal

applications for which many of these solutions were intended, numerical modelling has

shown that multidimensional conduction in the rock matrix can play a significant role in

determining the final temperature distribution (Kolditz, 1995).

There are no published closed-form analytical solutions that incorporate two-dimensional

diffusion in the rock matrix. However, some relevant approximate and semi-analytical solu-

tions exist in the literature. Cotta et al. (2003) used a radial lumping procedure to develop

an approximate analytical solution to the equivalent solute transport problem. Cheng et al.

(2001) developed a Laplace-space semi-analytical solution for multidimensional conduction

from a single fracture by formulating the problem as an integral equation which can then

be solved numerically. In a later paper (Ghassemi et al., 2003) this was extended to include

three-dimensional conduction from a single fracture, although a finite difference procedure

must first be used to calculate the temperature on the fracture plane. Although the inte-

gral equation formulations allow for rapid calculation of the temperature at any point after

the temperature in the fracture is known, accuracy is generally poor at points close to the

fracture-matrix boundary.

Analytical solutions are very useful for developing an understanding of the important
CHAPTER 2. BACKGROUND 33

parameters governing heat transfer in a particular environment. For problems where ge-

ometry is simple or little information is known about heterogeneity, they can be an ideal

method of solution. Yet, as the geometry of a problem becomes more and more complex, the

applicability of analytical solutions becomes limited. For a true representation of the mul-

tidimensional heat flow effects present when groundwater flows in to a thermal treatment

area, it is necessary to solve the three-dimensional heat conduction-convection equation. At

the present time, numerical models present the only means of solving these equations.

2.6 Heat Transfer in Fractured Media: Numerical Models

Several numerical models are available to model the concurrent flow of heat and groundwater

in fractured media. However, few are designed around a discrete fracture conceptual model.

Many codes have been developed to model the single-phase coupled transfer of heat and

fluid in the subsurface (e.g. Molson and Frind, 2002; Kipp, 1997; Therrien et al., 2007).

Other codes allow for multiphase flow, including unsaturated flow and the boiling of pore

water (e.g. Pruess et al., 1999; White and Oostrom, 2006).

Of the aforementioned models, only HydroGeoSphere (Therrien et al., 2007) is designed

to work with a discrete fracture conceptual model. The code uses a finite element dis-

cretization to efficiently model planar fractures. However, the model is presently not able

to handle the boiling of water.

TOUGH2 (Pruess et al., 1999) uses a finite-difference discretization. Using finite dif-

ferences, a discrete fracture approach is possible, but the very small gridblocks that result

impose a limitation on the range of possible timesteps, lengthening the simulation time con-

siderably (Pruess et al., 1990a). Consequently, modelling of fractured media in TOUGH2 is

typically conducted using a dual-porosity or multiple interacting continua approach (Pruess

and Narasimhan, 1982; Pruess, 1992). Support does not exist in released versions of the

code for hydrodynamic dispersion or temperature-dependent thermal conductivity.


CHAPTER 2. BACKGROUND 34

2.6.1 The TOUGH Family of Codes

The name “TOUGH” refers to “transport of unsaturated groundwater and heat,” a code

developed at Lawrence Berkeley National Laboratory in the 1980s. The code was developed

primarily to model transport processes in the unsaturated formations at Yucca Mountain,

then (and still) under consideration for development as a high-level nuclear waste disposal

site. Several more recent models have been developed using TOUGH as a base. The most

widely used of these is TOUGH2, which is applied to problems such as geothermal reservoir

modelling, nuclear waste storage, and geologic CO2 sequestration (Pruess, 2004).

Governing Equations

TOUGH2 numerically solves the governing equations for fluid and heat flow using a so-

called “integral finite difference” (IFD) approach (Narasimhan and Witherspoon, 1976),

commonly referred to in other implementations as a “finite volume” method (Ferziger and

Perić, 1996; Kolditz, 2002). In an IFD or finite volume approach, the integral forms of the

conservation laws are used as the governing equations, rather than the more-commonly used

differential forms. One advantage of IFD over traditional finite differences is its flexibility in

problem geometry; any grid satisfying the Voronoi conditions can be used with IFD (Palagi

and Aziz, 1994).

In integral form, the governing equation for fluid flow is given for a one-component

system by (Pruess and Narasimhan, 1982):

Z Z Z
d
M dVn = F · ndΓn + qmass dVn (2.16)
dt Vn Γn Vn

where M is the mass contained in volume element Vn , F is the mass flux vector, and

qmass is a mass source term. For heat transport, the governing equation is (Pruess and
CHAPTER 2. BACKGROUND 35

Narasimhan, 1982):

Z Z Z
d
U dVn = G · ndΓn + QdVn (2.17)
dt Vn Γn Vn

where U is the internal energy in volume element Vn , G is the heat flux vector, and Q is

a heat source term. The mass flux F is given by Darcy’s Law (Pruess and Narasimhan,

1982):
kkrw kkrv
F =− ρw (∇P − ρw g) − ρv (∇p − ρv g) (2.18)
µw µv

where the subscripts w and v denote the liquid water and vapour phases, respectively. Each

phase has an absolute permeability k, relative permeability kr , dynamic viscosity µ, density

ρ, and pressure P . Although it is assumed in this formulation that only one liquid phase

(water) is present, extended versions of TOUGH2 have support for both single-component

and multi-component non-aqueous phase liquids (Falta et al., 1995; Pruess and Battistelli,

2002).

The heat flux G is given by the conduction-convection equation:

G = −K∇T + hw F w + hv F v (2.19)

where K is the thermal conductivity of the rock-fluid mixture, T is the temperature, hw is

the specific enthalpy of the water phase, and hv is the specific enthalpy if the vapour phase.

Discretization

In order to solve for the primary variables of temperature and pressure throughout the

solution domain, it is assumed that the volume integrands within equations (2.16) and

(2.17) are constant throughout each volume element. Further, the surface Γn is broken

down into discretized surfaces between adjacent volume elements Vn and Vn , having an
CHAPTER 2. BACKGROUND 36

area of Anm . With this simplification, the mass conservation equation can be rewritten as

follows:
d X
V n Mn = Anm Fnm + qVn (2.20)
dt

A similar treatment is given to the heat conservation equation. Individual mass flux

terms are approximated using a first-order finite difference. For example, the flow of water

is approximated as (Pruess et al., 1999):

   
krw ρw Pwn − Pwm
Fw = −knm − ρw gnm (2.21)
µw nm Dnm

In this equation, the subscripts n and m refer to nodes n and m; Dnm is the distance

between these two nodes. The value of quantities such as permeability, density, and relative

permeability (mobility) must be approximated at the element interface using an average

between their defined values at nodes n and m. Several weighting schemes may be used, in-

cluding the arithmetic mean, harmonic mean, and upstream weighting. The appropriateness

of the various schemes has been a subject of considerable study. Earlier references (Peace-

man, 1977; Aziz and Settari, 1979) generally recommend upstream weighting of relative

permeabilities to minimize error. Still, these authors consider that absolute permeabilities

may be weighted with a simple arithmetic or harmonic mean. More recent references (e.g.

Pruess et al., 1999) recommend the use of upstream weighting for both absolute and relative

permeabilities in two-phase flow simulations.

Solution of Discretized Equations

The time derivatives in equations (2.16) and (2.17) are also approximated by a first-order

finite difference. TOUGH2 uses a backward difference expression taken at time step k + 1,

which gives the following series of equations:


CHAPTER 2. BACKGROUND 37

( )
∆t X
Mnk+1 = Mnk − k+1
Anm Fnm + Vn qnk+1 for mass flow (2.22)
Vn m
( )
∆t X
Unk+1 k
= Un − k+1
Anm Gnm + Vn Qn k+1
for heat flow (2.23)
Vn m

Because the backward difference is taken from time step k + 1, these time-discretized

equations are written entirely in terms of the flux, source, and sink terms at the “new” time

step k + 1. This “fully implicit” treatment causes the numerical solution to be uncondition-

ally stable, meaning that it will approach the exact solution as the discretization is refined

(Jaluria and Torrance, 2003).

Because equations (2.22) and (2.23) can be written for each of the N volume elements,

a 2N × 2N system of nonlinear algebraic equations can be developed. The 2N unknowns

in the system are the “primary variables,” a set of variables which can completely define

the state of the flow system. In a water-only system in TOUGH2, the primary variables X1

and X2 are pressure and temperature for single-phase conditions, and gas-phase pressure

and gas saturation for two-phase conditions (Pruess et al., 1999). The system is nonlinear

because the flux terms Fnm and Gnm are themselves functions of the primary variables.

An alternative way to write equations (2.22) and (2.23) is in terms of residuals Rnk+1 :

( )
∆t X
Rnk+1 = Mnk+1− Mnk
− k+1 k+1
Anm Fnm + Vn qn , n = 1...N for mass flow (2.24)
Vn m
( )
∆t X
Rnk+1 k+1 k
= Un − Un − k+1 k+1
Anm Gnm + Vn Qn , n = N + 1...2N for heat flow
Vn m

(2.25)

When the equations are written in this way, the solution can be seen as a root-finding

problem; for a “perfect” solution, Rnk+1 would be equal to zero for all n. The strong
CHAPTER 2. BACKGROUND 38

nonlinearity in equations (2.24) and (2.25) makes their solution difficult. The approach of

TOUGH2 is to use Newton-Raphson iteration to linearize these equations. Newton-Raphson

iteration is based on the idea that, given x0 an estimated value of the root of f (x), a better

estimate of the root can be obtained by considering the first two terms of a Taylor series

expansion of f (x) about x0 :

f (x) ≈ f (x0 ) + f 0 (x0 )(x − x0 ) (2.26)

where x0 is the initial estimate of the root of f (x). When f (x) is set equal to zero, equation

(2.26) can be written as:


f (xp−1 )
xp = xp−1 − (2.27)
f 0 (xp−1 )

where xp−1 is the previous estimate of the root, and xp is the improved estimate. The very

same process can be used to find an approximate solution to equations (2.24) and (2.25);

only now, a multi-variable Taylor series must be used (e.g. Ferziger and Perić, 1996). The

multi-variable equivalent to equation (2.27) is given by:

Rnp,k+1 (xp1 , xp2 , ..., xp2N ) = Rnp−1,k+1 (xp−1 p−1 p−1


1 , x2 , ...x2N )
2n
∂Rnp−1,k+1 (x1p−1 , xp−1 p−1
2 , ...x2n )
(xpj − xp−1
X
+ j ) (2.28)
j=1 ∂xp−1
j

where p refers to the pth Newtonian iteration. Equation (2.28) provides a 2N × 2N system

of linear algebraic equations. The system can be written in a concise matrix form as:

J p−1 d = Rp−1 (2.29)

where d = xp − xp−1 and:

∂Ri
Jij = , i = 1...2N, j = 1...2N (2.30)
∂xj
CHAPTER 2. BACKGROUND 39

Solution of the linear system of equations (2.29) provides a new estimate of the primary

variables at the new time step, xp,k+1


i . Using the new values of the primary variables, the

residuals Rnp,k+1 are again calculated. If the residuals are considered to be small enough,

no further Newton-Raphson iterations are conducted, and work advances to the next time

step. Specific convergence criteria are discussed in the TOUGH2 manual.

The number of Newtonian iterations needed to achieve convergence is of importance.

If p is very small, then the values of the primary variables are changing slowly. This is an

indication that the time step can probably be safely increased. In TOUGH2, the time step

is doubled when Newton-Raphson convergence is reached with p ≤ 3. On the other hand,

if the residuals are still too large at higher values of p (p > 8 in TOUGH2), the Newton-

Raphson iteration is considered to have failed to converge. In this case, the timestep is

reduced, and a new set of Newtonian iterations is initiated (Pruess et al., 1999).

In TOUGH2, equation (2.29) may be solved using either sparse direct solvers or iterative

matrix solvers. Iterative solvers, while less reliable than direct solvers, are far more efficient

in terms of both memory and computational effort; for this reason, they are preferred

(Moridis and Pruess, 1997). A maximum number of iterations is specified as a fraction of

the total number of equations 2N . Each solver has its own convergence criterion, which can

be controlled through the input parameter CLOSUR (Moridis and Pruess, 1995). Most of

the computational work of TOUGH2 is spent in the computation of the Jacobian matrix

and the solution of the resulting set of linear equations. The Jacobian matrix at each

Newtonian iteration can be included in the TOUGH2 output stream by specifying MOP(6)

= 7 in the input file. Software such as vismatrix, or the spy() function in MATLAB, may

then be used to view a graphical representation of the matrix.

Applications

TOUGH2 has been primarily applied to problems in geothermal reservoir modelling and

nuclear waste disposal (Pruess, 2004). Two derivatives of the code, T2VOC and TMVOC
CHAPTER 2. BACKGROUND 40

(Falta et al., 1995; Pruess and Battistelli, 2002), are able to model the behaviour of non-

aqueous phase liquids and have been used to model environmental remediation technologies.

Published applications of these codes have focused on air sparging at the bench scale (Hein

et al., 1997) and field scale (McCray and Falta, 1996), and steam injection at the bench

scale (Gudbjerg et al., 2004) and field scale (Gudbjerg et al., 2005; Kling et al., 2004).

Although no published studies show the use of TOUGH2 to model thermal conductive

heating systems, heater wells are easily handled through the definition of heat source terms

in the gridblocks.

TOUGH2 is often applied to fractured rock environments and provides a numerical

multi-continuum framework to handle this (Pruess and Narasimhan, 1982). Although

TOUGH2 may be used in conjunction with a discrete-fracture conceptualization, such ap-

plications of the code are rare. Pruess et al. (1984, 1990a,b) used discrete fractures to

model unsaturated flow in vertical fractures near heat-generating nuclear waste canisters.

In this particular context, it was concluded that the fluid and heat flow could be adequately

modelled by considering an equivalent continuum with specially constructed relative per-

meability and capillary pressure functions. In a study of tracer transport, Seol et al. (2005)

used a discrete fracture network model as a “base case,” the results of which the authors

then attempted to match with a modified dual-continuum model.

An alternative to discrete fracture models and equivalent continuum models is the “chan-

nelized flow” or “fracture zone” concept. Pruess and Tsang (1994) reasoned that, when

detailed information about fractures is not available, it is appropriate to approximate their

effect by introducing a porous zone of high permeability. This approach also permits the

use of much larger gridblocks, and is expected to be in between a discrete fracture and

equivalent continuum model in terms of computational demands.


Chapter 3

Thermal Conductive Heating in


Fractured Bedrock: Screening
Calculations to Assess the Effect of
Groundwater Influx

3.1 Abstract

A two-dimensional semi-analytical heat transfer solution is developed and a parameter sen-

sitivity analysis performed to determine the relative importance of rock material properties

(density, thermal conductivity and heat capacity) and hydrogeological properties (hydraulic

gradient, fracture aperture, fracture spacing) on the ability to heat fractured rock using ther-

mal conductive heating (TCH). The solution is developed using a Green’s function approach

in which an integral equation is constructed for the temperature in the fracture. Subsurface

temperature distributions are far more sensitive to hydrogeological properties than material

properties. The bulk groundwater influx (q) can provide a good estimate of the extent of

41
CHAPTER 3. SCREENING CALCULATIONS 42

influx cooling when influx is low to moderate, allowing the prediction of treatment zone

temperatures without specific knowledge of the aperture and spacing of fractures. Target

temperatures may not be reached or may be significantly delayed when the groundwater

influx is large.

3.2 Introduction

Many of the chemical and physical properties of organic chemicals frequently encountered

at hazardous waste sites exhibit a functional dependence on temperature. Elevated temper-

atures often bring about a decrease in non-aqueous phase liquid (NAPL) viscosity resulting

in an increase in NAPL mobility, a decrease in the organic carbon partition coefficient re-

sulting in decreased sorption, an increase in vapour pressure resulting in increased NAPL-air

mass transfer (vapourization), and an increase in the Henry’s constant leading to increased

water-air mass transfer (volatilization) (e.g. U.S. EPA, 2001; Sleep and McClure, 2001; Na-

tional Research Council, 2004). At high temperatures (> 100 ◦ C), heat may also stimulate

processes such as aqueous oxidation and pyrolysis that destroy contaminants in-situ and

reduce the need for above-ground treatment (Baker and Kuhlman, 2002). Although these

in-situ destruction mechanisms are typically significant only at the high temperatures used

to treat non-volatile compounds such as polychlorinated biphenyls (PCBs), they may alone

provide 95-99% of the removal in these cases (Baker and Kuhlman, 2002).

Heat can be delivered to the subsurface using several different approaches. Steam-

enhanced extraction (SEE), originally developed by the petroleum industry, has been ap-

plied at both pilot and full scales in unconsolidated deposits (e.g. Newmark et al., 1998).

However, field pilot testing of steam injection in fractured rock has demonstrated the dif-

ficulty of achieving large temperature increases throughout a treatment area (Davis et al.,

2005). Electrical resistive heating (ERH) achieves heating by passing an electrical current
CHAPTER 3. SCREENING CALCULATIONS 43

between electrodes inserted in-situ throughout the treatment area (e.g. Beyke and Flem-

ing, 2005). The amount of resistive heat produced is relatively uniform throughout the

treatment area, providing heat to low-permeability areas that may be by-passed by injected

steam. Because ERH relies on pore water to conduct electrical current, it only generates

temperatures below and at the boiling point of water. Thermal conductive heating (TCH)

systems employ arrays of wells containing resistive heating elements to provide heat to the

treatment area (e.g. Stegemeier and Vinegar, 2001). The resistive heating elements radiate

heat to the well casing, from where it is transferred away by conduction. One principal

difference between TCH and both SEE and ERH is the ability to heat to temperatures

of up to approximately 800 ◦ C, which allows the technology to target higher boiling point

compounds such as PCBs.

Several mechanisms may cause a loss of heat from the treatment area during thermal

applications. Strong vertical temperature gradients may cause heat to be lost through

conduction; in the case of TCH, wells are typically extended a minimum of two feet (0.6 m)

beyond the limit of the treatment zone to mitigate this effect (Stegemeier and Vinegar,

2001). In addition, insulating blankets may be placed on the ground surface above the

treatment area. The influx of cool groundwater may present another source of heat loss.

When the temperature of the treatment area is below 100 ◦ C, groundwater flow may cause

heated water to be carried out of the treatment area, representing a loss of energy. At

higher temperatures, cool incoming water must be boiled, causing a delay in the attainment

of target temperatures. In the presence of large groundwater influxes, the cooling influence

may be lessened by steam injection or the installation of an impermeable barrier at the

periphery of the treatment zone (Baker and Heron, 2004). Alternatively, an extra row of

heater wells could be used to preheat incoming groundwater before it enters the treatment

area.

Although the cooling effect of incoming groundwater may be a critical parameter in the

design of TCH systems, few published studies have quantitatively examined its importance
CHAPTER 3. SCREENING CALCULATIONS 44

in porous media, and none has done so in fractured rock. Elliott et al. (2003) used a commer-

cial reservoir simulator to study the cooling influence of groundwater in saturated porous

media. They found that the remediation time was largely governed by soil permeability and

hydraulic gradient; when these parameters were increased above certain threshold values,

treatment temperatures were not reached. A second modelling study (Elliott et al., 2004)

examined the effectiveness of several impermeable barrier designs in managing groundwater

influx.

The objective of this study is to present a screening-level model that can be used to

assess the effect of inflowing groundwater on the ability to heat a treatment zone in frac-

tured rock using TCH. A new semi-analytical solution is developed and used to model

different scenarios in a sensitivity analysis. A base case was established from which six

properties were varied to assess their relative importance to treatment time: hydraulic gra-

dient, fracture aperture, fracture spacing, rock density, rock thermal conductivity, and rock

heat capacity. Hydrogeological parameters were varied independently; rock properties were

varied as a group using measured values from literature. The potential for the mitigation

of groundwater influx cooling by both installation of an upgradient preheating well and

increases in thermal well power was also assessed.

3.3 Model Development

The fractured rock environment is conceptually modelled using a discrete fracture approach

whereby the location and aperture of fractures are specified directly. Fractures, which

have an aperture of e, are assumed to be parallel and evenly spaced by a distance of

2H. A schematic of the conceptual model is shown in Figure 3.1. The screening model

simulates heat transfer within a two-dimensional vertical cross section (x - y plane) oriented

perpendicular to the fractures and in line with the direction of groundwater flow (x) which

occurs at a specified velocity (v) in the fractures only. Thermal wells (line sources of heat)
121  location and aperture of fractures are specified directly.  Fractures, which have an aperture of e, are 
122  assumed to be parallel and evenly spaced by a distance of 2H.  A schematic of the conceptual model is 
123  shown in Figure 1.  The screening model simulates heat transfer within a two‐dimensional vertical cross 
124  section (x – y plane) oriented perpendicular to the fractures and in line with the direction of 
CHAPTER 3. SCREENING CALCULATIONS 45
125  groundwater flow (x) which occurs at a specified velocity (v) in the fractures only.  Thermal wells (line 
126  sources of heat) are placed within the cross section at a specified spacing.  The screening model does 
are placed within the cross section at a specified spacing. The screening model does not
127  not simulate heat losses from the top or bottom of the overall target zone, implying that lines of 
simulate heat losses from the top or bottom of the overall target zone, implying that lines
128  symmetry exist at a distance H above and below each fracture.  
of symmetry exist at a distance H above and below each fracture.

2H

129   

Figure Figure 1: Conceptual model of fractured rock environment.  Groundwater flows at velocity v within equally 
130  3.1: Conceptual model of fractured rock environment. Groundwater flows at velocity
v within equally spaced (2H) fractures of aperture e. The model plane (x - y) is oriented
131  spaced (2H) fractures of aperture e.  The model plane (x – y) is oriented perpendicular to the fractures and in line 
perpendicular to the fractures and in line with v.
132  with v.   

Two-dimensional heat transfer within the model plane may be described by two coupled
133  Two‐dimensional heat transfer within the model plane may be described by two coupled differential 
differential
134  equations. In the rock matrix, the heat conduction equation is written as (e.g.
equations.  In the rock matrix, the heat conduction equation is written as (e.g., Özişik 1980): 
Özişik, 1980):
∂ 2T ∂ 2T g ( x, y) ∂ 21Tm ∂T ∂ 2 Tm g(x, y) 1 ∂Tm
135  + + = +   +    =         (3.1)  (1) 
∂x 2 ∂y 2 Kr ∂xα r ∂t ∂y
2 2 Kr αr ∂t

where Tm is the temperature in the rock matrix, Kr is the thermal conductivity of the rock

[W/m·K], αr is the thermal diffusivity of the rock [m2 /s] and g is the strength of energy Baston 6 

generation
  [W/m3 ] at the point (x, y). Performing a heat balance on a control volume of

water in a fracture in the x-direction gives the following equation for the temperature in
CHAPTER 3. SCREENING CALCULATIONS 46

the fracture (Appendix A):


∂Tf ∂Tf 2Kr ∂Tf
ρw cw = −vρw cw + (3.2)
∂t ∂x e ∂y y=e/2

where ρw is the density of water [kg/m3 ], cw is the specific heat of water [J/kg·K], v is the

average linear velocity of groundwater in the fracture [m/s], and e is the aperture of the

fracture [m]. Heat exchange between the fracture and the matrix is handled through the

g(x, y) term as follows (Cheng et al., 2001):


Kr ∂Tm
g(x, y) = (3.3)
e/2 ∂y y=e/2

For the dimensions and velocities typical of flow in fractured rock, it has been shown

that the heat storage term may be omitted from (3.2) without significant error (Appendix

B). This allows the governing equation for the fracture to be simplified to:


∂Tf 2Kr ∂Tm
= (3.4)
∂x ρw cw ve ∂y y=e/2

Because the fracture and the edge of the rock matrix are assumed to be in thermal

equilibrium, Tm (x, 0, t) is equal to Tf (x, t) for all x, t. Consequently, no further distinction

is made between Tm and Tf in this derivation.

When heat conduction in the rock matrix occurs primarily in the direction normal to the

fracture plane, the one-dimensional form of the heat conduction equation may be substituted

in place of (3.1), and the system is reduced to a more tractable system of ordinary differential

equations. Using this assumption, Lauwerier (1955) developed a solution applicable to heat

transfer between a body of rock and a single fracture; Yang et al. (1998) published a similar

solution that includes the effect of longitudinal conduction in the fracture. Gringarten et al.

(1975) developed a Laplace-space solution for heat transfer between a body of rock and a

set of parallel fractures. Lowell (1976) simplified that solution by showing that, for the
CHAPTER 3. SCREENING CALCULATIONS 47

modelling of hot dry rock geothermal systems where fracture spacing is typically very large,

little error is introduced by considering only a single fracture.

When the rock is heated directly, as is the case in TCH, a solution must consider multi-

dimensional conduction and heat generation within the matrix. To the authors’ knowledge,

the semi-analytical solution of Cheng et al. (2001) is the only solution to consider multidi-

mensional heat conduction in the matrix. However, there does not appear to be a published

solution to the case where heat is generated within the rock matrix.

Several features distinguish the present solution from previous works. First is the explicit

modelling of multiple parallel fractures with two-dimensional heat conduction in the rock

matrix. Although previous solutions have modelled parallel fractures or multidimensional

heat conduction, no solution has included both. Second, the present solution provides for

heat generation within the matrix, allowing the inclusion of an unlimited number of heater

wells, located at arbitrary coordinates. Third, the solution is given in terms of elementary

functions rather than special functions such as Bessel functions. This reduces computation

time and provides improved accuracy when evaluating the temperature at points inside

the rock matrix. The solution is not capable of modelling boiling within the fracture, or

thermally-induced changes in rock properties. Although the two-dimensional formulation

is adequate for an exploration of the important parameters in this system, use of a three-

dimensional numerical model may be preferable for design purposes.

A schematic of the two-dimensional model domain is presented in Figure 3.2. The

domain comprises a two-dimensional strip of rock of infinite dimension in x and of finite

width (H) in y. At y = 0, water flows through a fracture of aperture e at an average

linear velocity determined from a specified hydraulic gradient using the cubic law (e.g.

Witherspoon et al., 1980). The rock matrix is assumed to be impervious to the flow of

groundwater.

Type I zero-temperature boundary conditions are assigned at x = ±∞, and homoge-

neous Type II (no-flux) boundaries are assigned at y = 0 and y = H to represent lines of


CHAPTER 3. SCREENING CALCULATIONS 48

Symmetry boundary

y y=H
heater well
locations, x = Wi
x
V
y = 0• • • • • • • • • • • • • • • • • • • • • •
discretized
Symmetry boundary fracture,
x=0 to x=L

Figure 3.2: Schematic of model domain. Fractures are spaced at distance 2H, groundwater
flows through the fracture at velocity v, and heater wells are located at x = Wi .

symmetry. Prior to heating, the temperature is equal to zero throughout the domain. It

is important to note that the temperature rise will be computed; therefore, the computed

temperature rise can be added to any desired uniform initial temperature. At t = 0, heater

wells located at x = Wi begin generating heat at a constant rate of g [W/m]. Although the

rate of heat generation is assumed here to be constant, it will be seen that any function

g(t) may be used, provided that its Laplace transform is known.

The solution to (3.1) is given in terms of Green’s functions as (Beck et al., 1992):

Z t Z H Z ∞
αr
T (x, y, t) = G(x − x0 , y − y 0 , t − τ )g(x0 , y 0 , τ )dx0 dy 0 dτ (3.5)
Kr τ =0 y 0 =0 x0 =−∞

where G(x−x0 , y −y 0 , t−τ ) is a Green’s function corresponding to the domain and boundary

conditions described above and shown in Figure 3.2, and g(x0 , y 0 , τ ) is the strength of an

instantaneous point source. A more complete discussion of the Green’s function used in

this solution is found in the appendix. The Laplace transform of (3.5) can be taken using

the convolution property, giving:

Z H Z ∞
αr
T̄ (x, y, s) = Ḡ(x − x0 , y − y 0 , s)ḡ(x0 , y 0 , s)dx0 dy 0 (3.6)
Kr y 0 =0 x0 =−∞
CHAPTER 3. SCREENING CALCULATIONS 49

where s is the Laplace variable, and the overbar denotes a transformed quantity. Two

vehicles of heat transfer to/from the rock matrix are present in this problem: heat generated

by the thermal wells and heat exchange with groundwater in the fracture. The source

function ḡ(x0 , y 0 , s) can therefore be expanded as:

2
X gw
ḡ(x0 , y 0 , s) = q̄(x0 , s)δ(y 0 ) + δ(Wi − x) (3.7)
s
i=1

where q̄(x0 , s) is an as yet unknown function of x0 representing heat exchange between the

rock matrix and groundwater in the fracture, and gw is the constant strength of the thermal

wells, expressed in W/m. Because the two heat source terms are geometrically linear and

orthogonal, the two integrals of (3.7) may be separated, giving:

∞ 2 Z
gw αr X H
Z
αr 0 0 0
T̄ (x, y, s) = q̄(x , s)Ḡ(x − x , y − 0, s)dx + Ḡ(W − x, y − y 0 , s)dy 0
Kr 0
x =−∞ K r s 0
i=1 y =0
(3.8)

By substituting the fracture heat balance (3.3) into the first integral of (3.8), integrating

by parts, and evaluating at y = 0, the problem is converted into a Fredholm integral equation

of the second kind:

Z ∞
veρw cw αr ∂ Ḡ
T̄ (x, 0, s) = 0
(x − x0 , 0, s)dx0 + T̄ (−∞, 0, s)Ḡ(x, 0, s)
T̄ (x0 , 0, s)
2 Kr 0
x =−∞ ∂x
2 Z
gw αr X H

+ T̄ (∞, 0, s)Ḡ(x, 0, s) + Ḡ(W − x, y 0 , s)dy 0 (3.9)
Kr s y 0 =0
i=1

The derivative of the Green’s function, ∂ Ḡ/∂x0 , is singular as (x0 − x) → 0. In order

to provide accurate results, (3.9) must be regularized. A subtraction method is used (e.g.
CHAPTER 3. SCREENING CALCULATIONS 50

Delves and Mohamed, 1985). The regularized form of (3.9) is given by:

Z ∞
veρw cw αr  ∂ Ḡ
T̄ (x0 , 0, s) − T̄ (x, 0, s) (x − x0 , 0, s)dx0

T̄ (x, 0, s) = 0
2 Kr 0
x =−∞ ∂x

+ T̄ (x, 0, s)Ḡ(∞, 0, s) − T̄ (x, 0, s)Ḡ(−∞, 0, s) − T̄ (−∞, 0, s)Ḡ(−∞, 0, s)


2 Z
gw αr X H

+ T̄ (∞, 0, s)Ḡ(∞, 0, s) + Ḡ(W − x, y 0 , s)dy 0 (3.10)
Kr s 0
y =0
i=1

A wide array of techniques may be used to solve (3.10) numerically (e.g. Delves and

Mohamed, 1985). Like Cheng et al. (2001), our approach is to use a quadrature method

to approximate the first integral. The result is an n × n system of linear inhomogeneous

equations. The ith equation is given by:



n
veρw cw αr X ∂ Ḡ
Ti = wj [Tj − Ti ] 0 (xi − xj , 0, s) + T1 Ḡ(xi , 0, s) − Tn Ḡ(xn − xi , 0, s)
2 Kr  ∂x
j=1
2 Z
)
gw αr X H
+ Ti Ḡ(xn − xi , 0, s) + Ti (xi , 0, s) + Ḡ(W − x, y 0 , s)dy 0 , i = 1...n
Kr s 0
y =0
i=1

(3.11)

where Ti is the Laplace-space temperature at xi , and wj are the values of the nodal weights.

The values of x and w are dependent on the choice of quadrature; the formula is sufficiently

general to allow the choice of any x, w pair. A quadrature method designed for infinite

integration limits, such as Gauss-Hermite quadrature, may be used, but this requires that

the temperature be scaled by a weighting function. An alternative is to evaluate only a

finite portion of the integral, from x0 = 0 to L. This approach is valid provided that that T

is equal to zero everywhere outside the evaluated portion. This condition will be satisfied if

L is chosen to be sufficiently large, and the condition may be easily checked by examining

the computed values of the temperature at the endpoints. Once the temperature on the
CHAPTER 3. SCREENING CALCULATIONS 51

boundary is known, the temperature at any point in the domain may be determined by:

n
veρw cw αr X   ∂ Ḡ
T̄ (x, y, s) = wj Tj − T̄ (x, 0, s) (x − xj , 0, s)
2 Kr  ∂x0
j=1
)
+ T1 (x, y, s) + Tn Ḡ(L, y, s) + T̄ (x, 0, s)Ḡ(L − x, 0, s) + T̄ (x, 0, s)Ḡ(x, 0, s)

2 Z
gw αr X H
+ Ḡ(W − x, y − y 0 , s)dy 0 , i = 1...n (3.12)
Kr s y 0 =0
i=1

A number of numerical Laplace inversion routines may be used to determine time-domain

temperatures from the values calculated in (3.11) and (3.12). The authors have had success

using the algorithms proposed by Weeks (1966) and De Hoog et al. (1982). The special case

of zero fracture aperture was verified against an analytical solution by Carslaw and Jaeger

(1959, p. 263). In addition, the numerical simulator TOUGH2 (Pruess et al., 1999) was

used to model the base case scenario (Table 3.1). The maximum difference in computed

fracture temperatures between the semi-analytical solution and the numerical solution was

6% (Appendix A).

3.4 Outline of Simulations

The base case scenario (Table 3.1) consists of shale with 500 µm horizontal fractures spaced

at 1 m. Groundwater flows through the fractures subject to a hydraulic gradient (∇h)

of -0.005 , resulting in an average linear velocity of 67 m/day (computed using µ = 1.31 ×

10−3 Pa·s). Heater wells, located at x = 30 m and x = 33 m, each provide a constant

heat output (gw ) of 100 W/m in the y direction. This output is equivalent to the spatially

averaged heat flux generated by a row of heater wells perpendicular to the model plane,

spaced at 3 m, each providing 300 W/m - well within the range attainable by the heater

elements in current use (Stegemeier and Vinegar, 2001). Use of an analytical solution

(Carslaw and Jaeger, 1959, p. 263) shows that, in the absence of cooling from fractures,
CHAPTER 3. SCREENING CALCULATIONS 52

a target temperature of 100 ◦ C would be reached throughout the interwell zone after 17

weeks of heating.

Table 3.1: Base case parameters for sensitivity analysis


Parameter Unit Value
Rock Type - Shale
Heater Well Locations m x = 30, 33
Heater Well Power (gw ) W/m 100
Initial Temperature ◦C 10
Fracture Aperture (e) µm 500
Fracture Spacing (2H) m 1
Hydraulic Gradient (∇h) - -0.005
Influent Temperature ◦C 10

Parameter sensitivity was assessed through four sets of trials (Table 3.2). Three sets

involved the variation of a hydrogeological parameter: hydraulic gradient, fracture aperture,

or fracture spacing. In the fourth set, the host rock type was changed. The hydraulic

gradient values selected range from those representative of a natural gradient (0.0001) to

those representative of pumping conditions (0.05). Values for the thermal conductivity,

density, and specific heat capacity of each rock type used (Table 3.3) were taken from

literature (C̆ermák and Rybach, 1982). Although the thermal conductivity of sedimentary

rocks tends to decline with temperature (e.g. Clauser and Huenges, 1995), all rock properties

were assumed to remain constant with temperature in order to preserve the linearity of the

governing equations (3.1) and (3.4).

Table 3.2: Summary of sensitivity testing trials


Parameter Range of Values Bulk Influx Range (L/m2 ·day)
Hydraulic Gradient (∇h) 0.0001 to 0.05 0.674 to 337
Fracture Aperture (e) 10 µm to 2000 µm 2.69 × 10−4 to 2160
Fracture Spacing (2H) 0.25 m to 4 m 8.42 to 135
Rock Type shale, limestone,
dolomite, sandstone
CHAPTER 3. SCREENING CALCULATIONS 53

Table 3.3: Rock material properties (C̆ermák and Rybach, 1982)


Rock Type Thermal Conductivity Density Specific Heat Capacity
(W/m·K) (kg/m3 ) (J/kg·K)
Shale 2.98 2757 1180
Sandstone 3.03 2391 960
Limestone 2.40 2520 890
Dolomite 2.87 2536 920

3.5 Results and Discussion

In order to facilitate direct comparison of the various trials, the temperature distribution

in the system is summarized by two values: the minimum (Tmin ) and maximum (Tmax )

temperatures in the fracture between the two thermal wells after one year of heating. This

time was chosen as a point for comparison because transient results (discussed later) show

that the effect of changes in some parameters may not be apparent during the earlier part

of the heating period.

Although groundwater influx generally skews the distribution of temperature contours

near thermal wells, the contours remain primarily normal to the fracture plane. This is

a consequence of the rapid thermal diffusion that occurs in the direction normal to the

fracture plane. Figure 3.3(a), for example, illustrates the two-dimensional distribution of

temperatures after one year of heating in the vicinity of a single heater well for the case

of e = 1 mm, 2H = 1 m, and v = 20 m/day. Figure 3.3(b) illustrates the temperature

distribution in the absence of groundwater flow. Because the temperature difference between

the fracture and the centre of the matrix is typically less than one degree, Tmin and Tmax

provide good indicators of the overall temperature distribution in the treatment zone.

3.5.1 Sensitivity to Hydrogeological Parameters

Minimum and maximum interwell fracture temperatures (Tmin ) and Tmax ) after one year

of heating are plotted in Figure 3.4(a) for the variation of hydraulic gradient between
CHAPTER 3. SCREENING CALCULATIONS 54

0.5 100

Distance from Fracture Plane (m)


0.4 80

0.3 60

0.2 40

0.1 20

0
0 10 20 30 40 50 60 70 T (°C)
(a) Distance Along Fracture Plane (m)

0.5 100
Distance from Fracture Plane (m)

0.4 80

0.3 60

0.2 40

0.1 20

0
0 10 20 30 40 50 60 70 T (°C)
(b) Distance Along Fracture Plane(m)

Figure 3.3: Temperature distribution resulting from heating by a single thermal well, show-
ing the general effect of groundwater influx. (a) v = 20 m/day, 2H = 1 m, e = 1 mm,
gw = 100 W/m (b) no flow, gw = 100 W/m.

−10−4 and −5 × 10−2 . The semi-analytical solution is not capable modelling the boiling

of water, so temperatures above 120 ◦ C are not shown. Computed temperatures for the

hydraulic aperture trials and fracture spacing trials are found in Figures 3.4(b) and 3.4(c),

respectively. It can be observed that a high hydraulic gradient, or the presence of large-

aperture or closely-spaced fractures can significantly inhibit heating in the treatment zone.

A variation in any of the hydrogeological parameters will have an impact on the bulk

groundwater influx (q). Since the matrix is considered to be impermeable, q can be calcu-

lated as:
ρge3 (∇h)
q= (3.13)
24Hµ

where µ is the dynamic viscosity of water. For each of the hydrogeological parameter

variation trials plotted in Figures 3.4(a-c), the bulk groundwater influx was calculated.

Figure 3.4(d) presents Tmin and Tmax plotted against the calculated bulk groundwater

influx values.
CHAPTER 3. SCREENING CALCULATIONS 55

120 120

100 100 Tmin


Tmin
Temperature (°C)

Tmax

Temperature (°C)
80 80
Tmax
60 60

40 40

20 20

0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0 500 1000 1500 2000 2500

Hydraulic Gradient Fracture Aperture (μm)

(a) Hydraulic Gradient (b) Aperture

120 120

100 100
Tmin

Temperature (°C)
Temperature (°C)

80 80 Tmax

60 60
Tmin
40 Tmax 40

20 20

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0 50 100 150 200 250 300 350 400

Fracture Spacing (m) Ground Water Influx (L/m2∙day)

(c) Fracture Spacing (d) Bulk Influx

Figure 3.4: Summary of computed one-year interwell fracture temperatures.

It is apparent from the temperatures plotted in Figure 3.4(d) that there is some degree of

correlation between bulk influx and treatment zone temperatures. To test the robustness of

this correlation, a series of tests was performed in which several combinations and/or values

of hydrogeological parameters were employed to arrive at a single value of bulk groundwater

influx. The semi-analytical solution was applied by assigning each combination of influx

and spacing presented in Table 3.4 and adjusting the aperture accordingly, with a gradient

fixed at -0.005. Results for the cases of q = 33.7 L/m2 ·day and q = 3.23 L/m2 ·day are

summarized in Figure 3.5, where transient temperature profiles are shown for the point in

the fracture at the centre (x = 31.5 m) of the treatment zone (not necessarily the location

of Tmin or Tmax ).

For the case of q = 33.7 L/m2 ·day, the temperature profiles are in good agreement
CHAPTER 3. SCREENING CALCULATIONS 56

Table 3.4: Parameters for runs used to assess correlation between bulk influx and treatment
zone temperature
Parameter Unit Values
Bulk Influx L/m2 ·day 33.7, 3.23, 0.323, 0.0323
Fracture spacing m 0.25, 0.5, 1.0, 2.5, 5.0, 10.0

120 120
q = 33.7 L/m2∙day q = 3.23 L/m2∙day
100 100

erture (°C)
erture (°C)

80 80

60 0.25 m 60 0.25 m
0 50 m
0.50 m
Tempe

Tempe
0 50 m
0.50 m
40 1.0 m 40 1.0 m
2.5 m 2.5 m
20 5.0 m 20 5.0 m
10 m 10 m
0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Time (years) Time (years)
(a) high influx (b) mid-level influx

Figure 3.5: Transient fracture temperature profiles at the centre of the treatment zone
for high (a) and mid-level (b) values of groundwater influx. Curves correspond to various
fracture spacings.

for fracture spacings of 2.5 m or less, but temperatures drop off when fracture spacing is

increased beyond this threshold. Similarly, for the case of q = 323 L/ m2 ·day, deviation

was observed for fracture spacings of 5 m and greater. When the bulk influx is decreased

to 3.23 L/m2 ·day (Figure 3.5(b)), fracture spacing does not appear to play a large role

in determining the temperature distribution. Results are not shown for the cases of lower

influx, where there is no significant difference between the temperature profiles at any

time during heating. Therefore, while the correlation between influx and Tmin and Tmax

appears to be strong in Figure 3.4(d), the temperature profiles in Figure 3.5(a) show that

the correlation becomes weaker when more extreme combinations of parameters are used.

For a screening-level estimate of the extent of fracture cooling at sub-boiling tempera-

tures for conditions of low to moderate influx, it may be necessary only to know the influx
CHAPTER 3. SCREENING CALCULATIONS 57

through the treatment area and not have knowledge of the particular combination of spac-

ing and aperture. This forms a distinction between the problem of heat extraction from hot

dry rock and subsurface heating using thermal wells. For example, Gringarten et al. (1975)

found that the number and size of fractures had a strong effect on heat extraction from a

hot dry rock reservoir, even when the total flow rate was kept constant. This difference in

behaviour may be due to the different time scales of the two problems; it is after a reser-

voir has been in operation for several years that the predicted outlet temperature becomes

dependent on the nature of the fractures carrying the flow.

3.5.2 Sensitivity to Host Rock Type

Using the four sets of rock properties (Table 3.3) and the base case properties (Table 3.1),

the temperature in the fracture between the two thermal wells was computed. Figure 3.6

illustrates the fracture temperature profiles for each rock type after one month and one year

of heating. Compared to the hydrogeological parameters, host rock material properties play

a relatively minor role in determining temperature distributions throughout the treatment

zone. This behaviour is not surprising, as the range of material properties is far smaller

than the range of hydrogeological parameters. Heating in rocks with low thermal diffusivity

will progress more slowly than in rocks with high thermal diffusivity; yet, this variation

does little to affect the shape of the steady-state temperature profile.

3.5.3 Transient Behaviour

As an example of transient behaviour during heating, the temperature in the fracture at

the midpoint between the two heater wells is plotted in Figure 3.7 for eight different values

of hydraulic gradient, using the base-case parameters of 500 µm fractures spaced by 1 m.

The transient temperature plot shows a correlation between the groundwater influx (in this

case determined by a varying hydraulic gradient) and the time needed to reach a steady-

state temperature profile. Similar results are seen when the influx is varied by a change in
CHAPTER 3. SCREENING CALCULATIONS 58

120 120
Shale
100 Limestone 100
erature (°C)

erature (°C)
80
Dolomite
80
Sandstone
60 60
Shale
Tempe

Tempe
40 40 Limestone
Dolomite
20 20
Sandstone
0 0
30.0 30.5 31.0 31.5 32.0 32.5 33.0 30.0 30.5 31.0 31.5 32.0 32.5 33.0

Distance (m) Distance (m)

(a) t = 1 month (b) t = 1 year

Figure 3.6: Early (a) and late-time (b) fracture temperature profiles for various rock types.

aperture or fracture spacing.

For a given value of groundwater influx, a temperature threshold exists below which

the fractures have little effect. When groundwater influx is very small (due to low-aperture

or widely spaced fractures, shallow gradient, etc.) the cooling effect is negligible until

temperatures approach 100 ◦ C or higher. For a mid-level groundwater influx, such as the

base case in this study, the cooling effect is negligible until a temperature of about 50 ◦ C,

where heating begins to lag before reaching a steady state. When the groundwater influx

is very high, the threshold temperature is so low that a steady state is reached almost

immediately, before any significant heating occurs.

3.5.4 Mitigation of Cooling Effect

The semi-analytical solution was employed to model two methods of overcoming the ground-

water influx cooling effect: the installation of a pre-heating well, and an increase in the

thermal well heat production. Using the base case parameters (Table 3.1), the time to

reach a target temperature of 100 ◦ C was calculated for thermal well heat production rates

between 70 W/m and 1000 W/m. The amount of energy required to achieve this target was

calculated by the product of the heat generation rate [W/m2 ] and the time [days] to reach
CHAPTER 3. SCREENING CALCULATIONS 59

120
‐0.025
100
‐0.005

erature (°C)
80

60 ‐0.0075
Tempe
‐0.010
40
‐0.015
20 ‐0.025

0
0 1 2 3
Heating Time (years)

Figure 3.7: Transient fracture temperature profiles at midpoint between heater wells, show-
ing the influence of hydraulic gradient.

the target. Computed values for the time to reach the target and energy consumption are

shown in Figure 3.8.

With these parameters, the target is not reached within three years at heat production

rates below 140 W/m. However, the time required to reach the target decreases sharply

above 140 W/m, reaching a minimum at approximately two weeks for heat fluxes above

900 W/m. Because the total power consumption is correlated to the heating time, it appears

to be generally advantageous to generate heat at a high rate, although for the parameters

used in this study, efficiency peaks at approximately 300 W/m. At peak efficiency, electrical

costs total approximately 94 dollars per well, per metre depth (assuming an electrical cost

of 15 cents / kW·h.)

Mitigation using a preheating well was also modelled (results not shown). Using the base

case parameters, a third thermal well was placed 3 m upgradient of the first well (x = 27 m),

and the temperature in the original interwell zone was monitored through time. Although

the preheating well caused a rise in the interwell temperatures, it was not sufficient to reach

the target temperature throughout this entire zone.


CHAPTER 3. SCREENING CALCULATIONS 60

350 2500

mption (kW‐h/m2)
Time to Reach 100 °C (days)
300 Power Consumption (kW‐h)
2000
250

e (days)
200 1500

Total Energy Consum
Time 150 1000
100
500
50

0 0
0 200 400 600 800 1000
Thermal Well Heat Production (W/m)

Figure 3.8: Effect of increased heat production rate on the time needed to reach a target
temperature of 100 ◦ C and total energy consumption.

3.6 Conclusions

Groundwater influx may prevent or delay the heating of fractured rock during application

of thermal conductive heating (TCH). When bulk groundwater influx is high, temperatures

in the fractures are influenced by the aperture and spacing of fractures. For medium and

low values of influx, fracture properties do not appear to be important in determining the

temperature in fractures. In these cases, it appears not to be important to characterize dis-

crete fracture features in the treatment zone; only a quantification of the total groundwater

influx through the treatment zone is necessary.

Variations in material properties (rock density, rock thermal conductivity, and rock heat

capacity) amongst rock types do have a small effect on the early-time temperature distri-

bution in the rock, but on the whole are less significant than variations in hydrogeological

parameters (hydraulic gradient, fracture aperture, and fracture spacing). It is noted that the

range of variation in material properties is much smaller than the range of hydrogeological

properties, which may vary by several orders of magnitude.

Transient analysis shows that influx cooling affects treatment zone temperatures only
CHAPTER 3. SCREENING CALCULATIONS 61

once a certain temperature threshold has been passed during heating. It is possible that, if

target treatment temperatures are low, influx cooling may not pose a problem.

One solution to the problem of groundwater influx cooling is to simply increase the power

delivered to the thermal wells. In the case where this may not be done due to equipment

limitations or other concerns, preheating wells installed outside of the treatment zone may

be used to partially mitigate the cooling effects.

3.7 Acknowledgements

The authors would like to thank the U.S. Department of Defence Environmental Security

Technology Certification Program (Contract ER-0715), the Natural Sciences and Engineer-

ing Research Council of Canada (Discovery Grant), and Queen’s University for financial

support of this work. Tom Gleeson is thanked for many helpful conversations and sugges-

tions.

3.8 Green’s Function

The Green’s function used in the presented solution is determined by the product of two

one-dimensional Green’s functions: GX00 (x, t|x0 , τ ), for an infinite domain in x, where the

temperature vanishes as x → ±∞; and GY 22 (y, t, y 0 , τ ), for finite domain in y of length H,

with Type-II boundary conditions at y = 0 and y = H. These functions are given by Beck

et al. (1992) as:


!
(x − x0 )2
 
0 1
GX00 (x, t|x , τ ) = p exp − (3.14)
2 πα(t − τ ) 4α(t − τ )

" #
2 π 2 α(t − τ ) 0
  
1 X −m  mπy  mπy
GY 22 (y, t|y 0 , τ ) = 1+2 exp cos cos (3.15)
H H2 H H
m=1
CHAPTER 3. SCREENING CALCULATIONS 62

The effect of a point source function g(x0 , y 0 , τ ) at (x, y, t) is given by:

Z t Z a2 Z b2
α
T (x, y, t) = GXIJ (x, t|x0 , τ )GY M N (y, t|y 0 , τ )g(x0 , y 0 , τ )dy 0 dx0 dτ
K τ =0 x0 =a1 y 0 =b1
(3.16)

where G is the product GX00 GY 22 . From the convolution property of the Laplace transform,

the Laplace transform of (3.16) is given by:

Z a2 Z b2
α
T̄ (x, y, s) = L[GXIJ (x, t|x0 , τ )GY M N (y, t|y 0 , τ )]L[g(x0 , y 0 )]dy 0 dx0 (3.17)
K x0 =a1 y 0 =b 1

where Ḡ, the Laplace transform of GX00 GY 22 , is given by:

 r 
0 0 1 0 s
Ḡ(x, y|x , y ) = √
exp −|x − x |
2H sα α
 q   
0 s m2 π 2 mπy  mπy 0

X exp −|x − x | α + H cos H cos H
+ √ (3.18)
m=1
H 2 s + m2 π 2 α
Chapter 4

Numerical Modelling of Thermal


Conductive Heating in Fractured
Bedrock

4.1 Abstract

Numerical modelling was employed to study the performance of thermal conductive heating

(TCH) systems in a fractured shale under a variety of hydrogeological conditions. Model

results show that the effect of concentrated flow in fractures does not significantly affect

treatment zone temperature distributions, except near the beginning of heating or when

groundwater influx is high. Excluding these scenarios, there is little difference in the ability

of discrete fracture and equivalent porous media (EPM) simulations to model tempera-

ture distributions. However, fracture and rock matrix properties can significantly influence

the time necessary to reach complete steam saturation in the treatment area. A low-

permeability matrix, large porosity, or high fracture spacing can contribute to boiling point

elevation in the rock matrix. Consequently, knowledge of these properties is important for

63
CHAPTER 4. NUMERICAL MODELLING 64

the estimation of treatment times in these environments. Because of the variability in boil-

ing point throughout a fractured rock treatment zone, it may be difficult to monitor the

progress of thermal treatment using temperature measurements alone.

4.2 Introduction

Thermal methods of in-situ remediation attempt to take advantage of enhanced mass trans-

fer processes occurring at elevated temperatures in order to remove organic contaminants

from the subsurface, or destroy them in-situ. For many organic chemicals, elevated tem-

peratures bring about enhanced dissolution, vapourization, volatilization, desorption and

non-aqueous phase liquid (NAPL) mobility (e.g. National Research Council, 2004). A va-

riety of methods have evolved for heating the subsurface; in this study, consideration is

given to thermal conductive heating (TCH). TCH systems employ electrical heater wells

to directly heat the subsurface to temperatures of up to 800 ◦ C. The technology has been

applied at several field sites in porous media, but has been used only recently in fractured

rock environments.

The design of thermal conductive heating systems may be significantly affected by the

flow of cool groundwater into the treatment zone. This cooling effect has been examined in

modelling studies conducted in porous media (Elliott et al., 2003, 2004) and fractured media

(Baston et al., 2007). For high values of intrinsic permeability or hydraulic gradient, target

temperatures may not be attained unless the influx cooling is accommodated in the heating

system design through the use of components such as preheating wells and impermeable

barriers. Although the effect of influx on boiling has been studied in porous media (Elliott

et al., 2003, 2004), no published study has examined how boiling during thermal treatment

may be affected by the high degree of heterogeneity present in fractured rock environments.

Using a two-dimensional semi-analytical solution, Baston et al. (2007) found that for the

prediction of treatment zone temperatures, the size and location of fractures are important
CHAPTER 4. NUMERICAL MODELLING 65

when groundwater influx is high. However, the semi-analytical solution used in that study

was not able to model the boiling of water.

The goal of the present study is to analyze the performance of a TCH system in a

three-dimensional fractured shale environment where boiling of the water is considered. A

numerical model is utilized to compute temperature distributions and steam saturations

in a circular treatment zone heated by seven thermal wells. The importance of matrix

permeability, matrix porosity, bulk permeability, and fracture spacing is examined.

4.3 Numerical Model

4.3.1 Model Domain and Boundary Conditions

Numerical modelling was conducted using the TOUGH2 simulator, which is capable of

modelling nonisothermal multiphase flows (Pruess et al., 1999). The numerical simulations

were designed to model seven heater wells arranged in a hexagon, with a heater-extraction

well at the centre (Figure 4.1). This geometry permits the use of partial radial symmetry,

whereby only a slice of the hexagon is modelled. The rock is a fractured shale, with equal-

aperture fractures occurring in the horizontal plane only and at regular spacing. The shape

of the model domain in the horizontal plane is shown in Figure 4.1. In the vertical direction,

the domain comprises one half of a fracture and adjacent matrix block, making use of vertical

symmetry (Figure 4.2).

The control volume formulation of TOUGH2 permits the use of irregularly shaped grids

such as those used in this study (Figure 4.3). In order to represent the flow that would

occur in fractures, a row of 1 mm thick “fracture zone” cells with high permeability was

created along the z = 0 plane (Figure 4.2). It was shown through a grid sensitivity study

and comparison with a semi-analytical discrete fracture solution that this approach results

in temperature distributions that are very similar to those predicted by a discrete fracture

model, while achieving much more rapid convergence due to the larger cell sizes (Appendix
horizontal plane only and at regular spacing. The shape of the model domain in the horizontal

plane is shown in Figure 1. In the vertical direction, the domain comprises one half of a fracture

and adjacent matrix block, making use of vertical symmetry (Figure 2).
CHAPTER 4. NUMERICAL MODELLING 66

symmetry boundary y 
heater wells 
(6)  x

flow

symmetry boundary
T = To 
P = P0 
heater‐extraction well 
(1)  θ = 30°
3 m well spacing  model boundary
 
Figure 4.1: Plan view of model domain
Figure 1: Plan view of model domain

TheC).
control volume formulation of TOUGH2 permits the use of irregularly shaped grids such as

Initially, thestudy
entire(Figure
domain3).is at ◦
those used in this In 10 C.toThe
order initial pressure
represent the flow distribution throughout
that would occur the a
in fractures,
model domain is hydrostatic, with the pressure at z = 0 corresponding to a depth of 15 m
row of 1 mm thick “fracture zone” cells with high permeability was created along the z = 0 plane
below the water table. Pressures in the extraction well, which were held constant through-
(Figure 2). It was shown through a grid sensitivity study and comparison with a semianalytical
out the simulation, were based on 1 m of drawdown from the initial pressure condition. At
discrete fracture
the outer edgesolution that this
of the domain (r approach results
= 25m), both in temperature
temperature distributions
and pressure that are
were held nearly
constant.

A gridto
identical dependence analysis
those predicted by showed that
a discrete increasing
fracture thewhile
model, size ofachieving
the model domain
much moreinrapid
the ra-

dial direction did not significantly affect temperature or pressure distributions within the
convergence due to the larger cell sizes (Baston, 2008).
heated region (Appendix C). The two thermal wells located within the model domain, both

assumed to produce heat at a steady rate of 800 W/m, were represented by line sources.

The equations of van Genuchten (1980) were used for capillary pressure and relative per-

meability functions. All properties of water and steam were computed using the EOS1

module of TOUGH2, which provides an implementation of the 1967 International Formu-

lation Committee steam table equations (e.g. ASME, 1979). The model does not consider

processes such as fracture creation or dilation that may occur under elevated pressures.
CHAPTER 4. NUMERICAL MODELLING 67

symmetry boundary 
symmetry boundary 

symmetry boundary 
symmetry boundary 
extraction well 
extraction well  T = To T = To 
P = P0 – 9806 Pa 
P = P0 – 9806 Pa  P = P0 P = P0
z  z 
1 mm fracture zone 
1 mm fracture zone 

r  r 
   
Figure 2: Section of model domain in rz plane 
Figure 2: Section of model domain in rz plane  Figure 3: Isometric view of model domain 
Figure 3: Isometric view of model domain 
   
Figure 4.2: Section of model domain in r - Figure 4.3: Isometric view of model do-
 
z plane   main

4.3.2 Rock Properties and Outline of Simulations


Initially, the domain
entire domain is °C. °C. initial
at 10The The initial pressure distribution throughout the model
Initially, the entire is at 10 pressure distribution throughout the model
A series
domainof
issimulations
domain wasthe
is hydrostatic,
hydrostatic, with conducted
with attoz =examine
the pressure
pressure the effect
at0 zcorresponding to aof
= 0 corresponding to matrix
of 15ofpermeability
a depth
depth m15below the the(km ),
m below

matrix porosity (φ),Pressures


water table. bulk permeability (kb ),
in the extraction andwhich
well, fracture spacing
were held (2H)
constant on the temperature
throughout the
water table. Pressures in the extraction well, which were held constant throughout the
and time simulation,
necessarywere
to reach
based complete steam saturation
on 1 m of drawdown within
from the initial the treatment
pressure zone,
condition. At defined
the outer
simulation, were based on 1 m of drawdown from the initial pressure condition. At the outer
as the area within
edge of the adomain
3 m radius of the
(r = 25 m), bothcentre well.and pressure were held constant. A grid
temperature
edge of the domain (r = 25 m), both temperature and pressure were held constant. A grid
A base case set analysis
dependence of rock showed
matrixthat andincreasing
bulk medium
the size properties
of the model is defined
domain in radial
in the Tabledirection
4.1. To
dependence analysis showed that increasing the size of the model domain in the radial direction
evaluate the sensitivity to rock matrix properties, km and φ were varied independently over
did not significantly affect temperature or pressure distributions within the heated region
did not significantly affect temperature or pressure distributions within the heated region
the ranges in Table 4.2. The range of rock matrix permeabilities, 10−18 m2 to 10−22 m2 ,
(Baston, 2008). The two thermal wells located within the model domain, both assumed to
(Baston, 2008). The two thermal wells located within the model domain, both assumed to
is representative of literature values for the permeability of shale (e.g. Keith and Rimstidt,
produce heat at a steady rate of 800 W/m, were represented by line sources. The equations of
produce heat at a1986;
1985; de Marsily, steadyHart
rate of
et 800
al., W/m,
2006).were
Rockrepresented
thermal by line sources.
properties, The equations
variations of have
in which
van Genuchten (1980) were used for capillary pressure and relative permeability functions. All
beenvanshown
Genuchten (1980)
to play were used
a minor roleforincapillary pressure
this context and relative
(Baston permeability
et al., 2007) werefunctions. All
held constant
properties of water and steam were computed using the EOS1 module of TOUGH2 was used,
throughout
propertiesthe simulations.
of water and steam were computed using the EOS1 module of TOUGH2 was used,

To evaluate the sensitivity to bulk medium properties, a simulation was conducted for

each combination of bulk permeability and fracture spacing in Table 4.2. In addition to

the five values of fracture spacing used, a simulation was run for each value of kb using an

equivalent porous medium (EPM) approximation.


CHAPTER 4. NUMERICAL MODELLING 68

For a given value of kb , an increase in the fracture spacing is associated with increased

fracture permeability and therefore a higher concentration of flow in the fracture zone cells.

The fracture zone permeability was calculated by considering the bulk permeability to be

a weighted average of the matrix and fracture zone permeabilities.

Table 4.1: Base case properties. Density, specific heat capacity, and thermal conductivity
from C̆ermák and Rybach (1982).
Property Unit Value
Rock Matrix
Permeability m2 10−18
Porosity - 0.03
Density kg/m3 2757
Specific heat capacity J/kg·K 1100
Thermal conductivity W/m·K 2.98

Bulk Medium
Fracture spacing m 2.5
Bulk permeability m2 10−13

Table 4.2: Parameters varied.


Parameter Unit Values Tested
Rock Matrix
Permeability m2 10−17 , 10−18 , 10−19 ,10−20 ,10−22
Porosity - 0.005, 0.010, 0.050, 0.100

Bulk Medium
Bulk permeability m2 1.0 × 10−11 , 1.0 × 10−12 , 1.0 × 10−13 , 7.5 × 10−14 ,
5.0 × 10−14 , 1.0 × 10−14 , 1.0 × 10−15 , 1.0 × 10−16
Fracture spacing m 0 (EPM), 1.0, 2.5, 5.0, 7.5, 10.0

4.4 Results and Discussion

The fluid pressure response to the heating of a saturated rock mass depends on the relative

influence of the thermal and hydraulic diffusivities of the rock (Palciauskas and Domenico,
CHAPTER 4. NUMERICAL MODELLING 69

1982). When the hydraulic diffusivity is much larger than the thermal diffusivity, the

thermal expansion occurs at near-constant pressure and the system responds as a drained

medium, with water being easily expulsed from the heated region. When the hydraulic

diffusivity is less than the thermal diffusivity, the system behaves as an undrained medium.

In this case, the heated water is not able to escape and a significant pressure rise occurs

(Palciauskas and Domenico, 1982).

The thermal diffusivity of the rock matrix is less than or approximately equal to the

hydraulic diffusivity for the rock considered in this study. In the fracture, however, the

hydraulic diffusivity is far in excess of the thermal diffusivity. Consequently, the fractures

act as drains for the fluid expansion in the rock matrix pores.

As boiling progresses in the treatment zone, a steam front propagates radially outward.

When a gridblock on the steam front is heated, water within the pores expands. The

mobility of this water is limited; it cannot flow radially inwards as a result of expanding

steam drive. Due to the low permeability of the rock matrix, the water can flow outward

only slowly; consequently, the pressure in the gridblock will rise. Because the pressure rise

causes the boiling point of water to increase, the process is self-promoting: the boiling point

elevation causes the water to expand more prior to boiling, with a concomitant increase in

pressure. Finally, the water does boil and the pressure is relieved as the less viscous steam

flows toward the extraction well. Still, some time is necessary for pressures in the rock

matrix to return to their “background” levels.

This pressure response is shown in Figure 4.4 for a location just outside the treatment

zone (r = 3.88 m, θ = 28.5◦ ), using the base case parameters (Table 4.1). The pressure dis-

tribution is very similar to that seen in a low-permeability medium subjected to an external

loading. For example, a similar pore pressure distribution was calculated by Nogami and Li

(2003), using an analytical solution for consolidation in a system of alternating horizontal

sand and clay layers. The magnitude of this pressure spike, which effectively determines

the time necessary to reach complete steam saturation in the matrix, is determined in part
CHAPTER 4. NUMERICAL MODELLING 70

by the fracture spacing, matrix permeability, and matrix porosity.

Because boiling in the rock matrix does not necessarily take place at constant pressure,

the temperature may not remain constant throughout the boiling period. Consequently,

a temperature plateau may not be observed in the rock matrix, as it typically is in the

fracture. In Figure 4.5, temperature is plotted against time for a reference block in the rock

matrix (r = 2.96 m, θ = 28.5◦ , z = 1.17 m) and in the fracture zone (r = 2.96 m, θ = 28.5◦ ,

z = 0.5 mm) for the base case parameters (Table 4.1). Although a clear temperature plateau

is observed during the boiling period in the fracture, the boiling period in the rock matrix

can not be discerned from temperature alone.

1.3
292 285 350
226 Matrix Fracture
Distance Frrom Fracture (m)

168 314
1.0 300
139 343
37 start of boiling 
perature (°C)

250
0.8 in fracture
end of boiling 
200
in matrix
0.5 150 start of boiling 
start of boiling
Temp

in matrix
100
0.3 end of boiling 
50 in fracture
0.0 0
0 5 10 15 20 0 50 100 150 200 250 300 350
Pressure (bar) Heating Time (days)

Figure 4.4: Pressure as a function of dis- Figure 4.5: Temperature vs. time for a
tance from the fracture for a location just reference block in the rock matrix (dot-
outside the treatment zone, computed us- ted line) and fracture (solid line), at the
ing the base case parameters. Heating time boundary of the treatment zone.
(days) is indicated for each line. Pressure
profiles prior to the beginning of boiling
are indicated with a solid line; profiles af-
ter the end of boiling are indicated with a
dashed line.
CHAPTER 4. NUMERICAL MODELLING 71

4.4.1 Influence of Matrix Permeability

The lower the matrix permeability, the more slowly the heated water will move outwards.

Consequently, the pressure rise becomes larger as the matrix permeability is decreased,

as shown in Figure 4.6(a) for a reference gridblock at the centre of the rock matrix and

edge of the treatment zone (r = 2.96 m, θ = 28.5◦ , z = 1.17 m). The degree of boiling

point elevation can be quite significant in a very low-permeability rock matrix; even at

treatment zone temperatures in excess of 300 ◦ C, water in the matrix pores may still be

in the liquid phase. It is conceivable that stresses resulting from the high pore pressure in

these cases could cause failure or microfracturing of the rock, thus attenuating the pressure

spike and reducing the boiling point elevation (Palciauskas and Domenico, 1982; Horseman

and McEwen, 1996).

In cases where the boiling point elevation is large, the time necessary to bring about

complete steam saturation throughout the treatment zone will increase significantly. In

Figure 4.6(b), the volume fraction of the treatment zone that has reached complete steam

saturation is shown for several values of matrix permeability. The significance of the boiling

point elevation may depend on the context of the thermal treatment. A boiling point of

300 ◦ C may not affect the treatment of low boiling-point compounds such as PCBs, where

the treatment temperature would need to exceed 300 ◦ C even in the absence of boiling point

elevation. However, boiling point elevation may significantly delay the treatment of volatile

compounds, which would be treated near 100 ◦ C in a porous medium (LaChance et al.,

2006).

4.4.2 Influence of Matrix Porosity

Figure 4.7(a) presents the transient pressure response for a reference block near the centre

of the rock matrix, at the outer radius of the treatment zone. The time necessary to reach

complete steam saturation is influenced by the matrix porosity, which not only determines
CHAPTER 4. NUMERICAL MODELLING 72

10000 1.2

Averagge Steam Saturation
1.0

in TTreatment Zone
1000
Pressure (bar)

0.8
100 0.6
0.4
10
02
0.2
1 0.0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350

Heating Time (days) Heating Time (days)

km = 1e‐17 km = 1e‐18 km = 1e‐19 km = 1e‐17 km = 1e‐18 km = 1e‐19


km = 1e‐20 km = 1e‐22 km = 1e‐20 km = 1e‐22

(a) (b)

Figure 4.6: Impact of matrix permeability (km ) on magnitude of pressure spike at centre of
rock matrix (a) and steam saturation within treatment zone (b).

the quantity of water to be boiled, but also has an effect on the pressure distribution. For

a given value of intrinsic permeability, an increase in the porosity causes a larger volume of

water to undergo thermal expansion, without increasing the ability of the expanded water

to flow outward. Consequently, the magnitude of the pressure spike will increase. For the

range of porosities tested, the significance of matrix porosity is much less than permeability

in determining the magnitude of the pressure spike.

The effect of porosity on steam saturation in the treatment zone is shown in Figure

4.7(b). In addition to boiling point elevation due to an increase in pore pressures, an

increase in matrix porosity causes a larger amount of energy to be expended to overcome

the latent heat of vapourization, thereby resulting in prolonged treatment times.

4.4.3 Influence of bulk medium and fracture properties

The importance of bulk medium and fracture properties was assessed through an exami-

nation of treatment zones temperatures and the time necessary to reach complete steam

saturation within the treatment zone. For each combination of bulk medium properties in
CHAPTER 4. NUMERICAL MODELLING 73

35 1.2
φ = 0.005

am Saturation in
30 1.0
φ = 0.01
25

ment Zone
sure (bar)

φ = 0.03 0.8
20 φ = 0.05
0.6 φ = 0.005

Treatm
Average Stea
1
15
Press

φ = 0.10 φ = 0.01
0.4 φ = 0.03
10
φ = 0.05
5 0.2
φ = 0.10
0 0.0
0 100 200 300 400 500 0 50 100 150 200 250 300 350
Heating Time (days) Heating Time (days)

(a) (b)

Figure 4.7: Impact of porosity (φ) on magnitude of pressure spike at centre of rock matrix
(a) and steam saturation within the treatment zone (b).

Table 4.2, the minimum temperature and the average steam saturation in the treatment

zone were recorded at fixed time intervals. It is important to note that the location of

the minimum treatment zone temperature may vary throughout time; consequently, the

temperatures plotted do not correspond to one single location in space.

Treatment Zone Temperatures

Figure 4.8 plots the minimum treatment zone temperature versus time for four values of bulk

permeability. For each value of bulk permeability, the temperature is plotted as calculated

using four different values of fracture spacing, as well an equivalent porous medium. It is

important to note that the location of the minimum treatment zone temperature may vary

throughout time; consequently, the temperatures plotted do not correspond to one single

location in space.

At high values of bulk permeability, the temperature profiles are very dependent on the

heterogeneity of the system. When kb is equal to 10−12 m2 , the treatment zone does not

reach boiling temperatures and comes to a steady state with a minimum temperature of

approximately 50 ◦ C (Figure 4.8(a)). The time needed to reach this steady state varies by
CHAPTER 4. NUMERICAL MODELLING 74

a factor of approximately five, depending on the fracture spacing. For each case of fracture

spacing, the final temperature is the same, illustrating that flow heterogeneity has an effect

only at early and mid-time in this scenario.

When the bulk permeability is decreased to 10−13 m2 , the effect of fracture spacing is

somewhat different. Although the temperature profiles are closely spaced at early time,

the profiles diverge as boiling is initiated in the fracture, and begin to converge some time

after boiling has finished (Figure 4.8(b)). Once again, the differences between the trials

are diminished at late-time. After 150 days of heating, only the temperature profile for

the 10 m fracture spacing case is significantly different. This gap is gradually closed as

the simulation time increases. For the cases of high fracture spacing in Figure 4.8(b), a

temperature plateau can be observed. In these cases, the coldest point in the treatment

zone is located in the fracture, where a plateau does develop during boiling.

At smaller values of bulk permeability (kb ≤ 10−14 m2 ), the differences between the

trials become less significant. For both kb = 10−14 m2 and kb = 10−15 m2 (Figures 4.8(c-

d)), the difference in predicted temperatures between the simulations rises gradually to a

maximum of 7% at 226 days of heating, before declining.


CHAPTER 4. NUMERICAL MODELLING 75

350 350
EPM
300 300
1 m
e (°C)

e (°C)
250 250
2.5 m
perature

perature
200 200 EPM
5 m
1 m
150
50 10 m
10 m 150
50
Temp

Temp
2.5 m
100 100
5 m
50 50
10 m
0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Heating Time (days) Heating Time (days)
(a) (b)

350 350

300 300
e (°C)

e (°C)

250 250
perature

perature

200 EPM 200 EPM


1 m 1 m
150
50 150
50
Temp
Temp

2.5 m 2.5 m
100 100
5 m 5 m
50 10 m
50
10 m
0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Heating Time (days) Heating Time (days)
(c) (d)

Figure 4.8: Minimum treatment zone temperature profiles for various values of fracture
spacing, with (a) kb = 10−12 m2 , (b) kb = 10−13 m2 , (c) kb = 10−14 m2 , (d) kb = 10−15 m2
CHAPTER 4. NUMERICAL MODELLING 76

Treatment Zone Boiling

Fracture properties can have a large impact on boiling times within the treatment zone, even

in cases where temperatures seem to be little affected by flow heterogeneity. In the fracture,

where there is no significant pressure spike, boiling will begin at a lower temperature than

in the rock matrix. Consequently, for all parameters used in this study, boiling occurred in

the fracture at a significantly earlier time than in the matrix. An example of this behaviour

is shown in Figure 4.9, which shows the relationship between steam saturation and distance

from the fracture plane in the treatment zone, for the base case simulation (Table 4.1).

Although the entire fracture is saturated with steam after 117 days, the matrix does not

reach complete steam saturation until 263 days of heating.

Although fracture spacing and bulk permeability may not have a significant effect on the

amount of time needed to reach a target temperature, these factors can have a significant

impact on the amount of time necessary to bring about complete steam saturation in the

treatment zone, as seen in Figure 4.10. For a given bulk permeability, the treatment period

is prolonged more by an increase in fracture spacing than an increase in bulk permeability.

2.5 300
Distance From Fracture Plane (m)

263
Boiling in 

190 250
e (days))

2.0 153
117
200
ent Zone

80
me to EEnd of B

1.5 k = 1.0e‐15
44 150
k = 1.0e‐14
TTreatme

1.0
100 k = 5.0e‐14
Tim

0.5 50 k 7 5 14
k = 7.5e‐14

k = 1.0e‐13
0.0 0
0.0 0.2 0.4 0.6 0.8 1.0 0 2 4 6 8 10 12
Average Steam Saturation in Treatment Zone Fracture Spacing (m)

Figure 4.9: Relationship between steam Figure 4.10: Sensitivity of treatment zone
saturation and distance from the fracture boiling time to bulk medium properties.
in base case simulation. The heating time
(days) is indicated for each curve.
CHAPTER 4. NUMERICAL MODELLING 77

When the fracture spacing is large, it is more difficult for the heated water in the

matrix to drain into a fracture. Consequently, the pressure rise is larger and the boiling

point more elevated. There appears to be a limit to this behaviour, however; as seen in

Figure 4.10, incremental increases in fracture spacing have a progressively smaller impact

on the time needed to reach complete steam saturation. Also apparent in Figure 4.10 is

the significant difference in boiling times between the EPM and 1 m fracture spacing cases.

Although there is no significant difference in computed temperatures between an EPM and

1 m-spaced fractures in any of the simulations conducted (Figure 4.8), the difference in

time needed to reach complete steam saturation may be over 50%. The large variability in

the time needed to reach complete steam saturation for a single value of bulk permeability

suggests a difficulty in predicting TCH treatment times in fractured bedrock, if complete

steam saturation is the goal of treatment.

4.5 Conclusions

The performance of thermal conductive heating in fractured rock environments may be

strongly dependent on the hydraulic properties of the rock matrix (permeability, porosity)

and the aperture and spacing of fractures. If complete steam saturation is the goal of thermal

treatment, treatment time will be strongly governed by the magnitude of the pressure spike

that occurs in the rock matrix during heating. When the rock matrix has a low permeability,

high porosity, or sparse fracturing, this pressure rise may be enough to significantly raise the

boiling point of water in the matrix, thus delaying treatment. Because a clear temperature

plateau may not be observed in the matrix during boiling, it may be difficult to determine

if boiling has occurred throughout a treatment area from temperature measurements alone.

Due to the importance of fracture spacing in determining the pressure rise in the matrix,

a discrete fracture model is most appropriate for modelling boiling in this context. However,

treatment zone temperatures are only moderately affected by the location of fractures,
CHAPTER 4. NUMERICAL MODELLING 78

for a given value of bulk permeability. An equivalent porous medium (EPM) model may

provide an adequate estimation of treatment zone temperatures, especially when the bulk

permeability is low or heating times are long.

4.6 Acknowledgements

The authors would like to thank the U.S. Department of Defense Environmental Security

Technology Certification Program (Contract ER-0715), the Natural Sciences and Engineer-

ing Research Council of Canada (Discovery Grant), and Queen’s University for financial

support of this work. Ron Falta is thanked for sharing his expertise with TOUGH2.
Chapter 5

Conclusions

5.1 Semi-Analytical Modelling

Results from semi-analytical modelling indicate that treatment zone target temperatures

may not be reached using thermal conductive heating if the cooling effect of groundwater

influx is not considered in the design of the treatment system. These results are in good

agreement with Elliott et al. (2003) and Elliott et al. (2004), who used numerical modelling

to study the cooling problem in porous media.

It was determined using a sensitivity analysis that, for the prediction of treatment

zone temperatures, the characterization of hydrogeological parameters (fracture aperture,

fracture spacing, hydraulic gradient) is more important than the characterization of rock

material properties (density, specific heat capacity, thermal conductivity). However, when

influx is low to moderate, predicted temperatures have only a minor dependence on the

particular combination of hydrogeological parameters used to arrive at a given bulk influx.

Two design options were shown to partially offset the cooling effect: increasing the power

delivered to the thermal wells or installing upgradient preheating wells. It was shown to be

more effective to increase the power of existing thermal wells before installing preheating

wells.

79
CHAPTER 5. CONCLUSIONS 80

5.2 Numerical Modelling

The treatment zone temperatures predicted using numerical modelling results were in agree-

ment with those of the semi-analytical solution. As with the results of the semi-analytical

modelling, a minor dependence of temperature on hydrogeological parameters was observed

for cases of low to moderate influx. Unlike the semi-analytical solution, the numerical model

was able to simulate boiling processes within the treatment zone.

Predictions of boiling in the treatment zone suggest that fracture spacing is important

in predicting the time needed to reach complete steam saturation, even in cases where the

temperature distribution is not affected by fracture spacing. It was shown that a pressure

spike can occur when a low-permeability rock matrix is heated, elevating the boiling point of

water. The magnitude of this pressure spike determines the time needed to reach complete

steam saturation in the rock matrix, which is typically the last part of the treatment zone

to boil. Matrix permeability, matrix porosity, and fracture spacing were all shown to play a

role in determining the magnitude of the pressure spike. For the ranges of properties tested,

the magnitude of the pressure spike was most sensitive to changes in matrix permeability.

Consequently, knowledge of these properties is important in the prediction of treatment

times.

Because boiling in the rock matrix does not occur at constant pressure, a temperature

plateau is not observed during boiling. Consequently, it may be difficult to determine

when the entire rock matrix has boiled, if temperature measurements are the only source

of information available.

5.3 Recommendations

Boiling in the a fractured rock environment is a very complex process, and a number of

simplifications were made in the modelling presented in this thesis. The removal of some of

these simplifications presents a number of opportunities for future work. In particular:


CHAPTER 5. CONCLUSIONS 81

• In this study, no consideration was given to the behaviour of NAPL in fractured rock

during heating. It would be useful to apply the TMVOC code (Pruess and Battistelli,

2002) to observe how NAPL boiling occurs in these environments.

• The thermal diffusivity of rocks tends to decline as temperature is increased (Clauser

and Huenges, 1995). In addition, anisotropy in thermal conductivity (C̆ermák and

Rybach, 1982) may affect temperature distributions during heating in fractured rock.

Investigation of these effects could provide useful insights into the design of thermal

wells. The TOUGH2 code would need to be modified in order to permit changes in

these properties (Pruess et al., 1999).

• When the magnitude of the pressure spike is large, it is possible that pressure-induced

fracturing may occur. In addition, thermally-induced expansion of the rock matrix and

pressure-induced compression may have an impact on the magnitude of the pressure

spike. In order to quantify the pressure response in the rock matrix when matrix

permeability is very low (e.g. 10−22 m2 ), it would be useful to apply a fully-coupled

thermohydromechanical (THM) model.


References

Abdulgatov, I. M., Emirov, S. N., Abdulagatova, Z. Z., and Askerov, S. Y. (2006). Effect of

pressure and temperature on the thermal conductivity of rocks. Journal of Engineering

& Chemistry Data, 51 , 22–33.

Abu-Hamdeh, N. H., and Reeder, R. C. (2000). Soil thermal conductivity: Effects of

density, moisture, salt concentration, and organic matter. Soil Science Society of America

Journal , 64 , 1285–1290.

Alishaev, M. (1979). Calculation of the temperature field of a porous stratum when fluid

is injected for plane flow. Fluid Dynamics, 14 (1), 49–56.

ASME (1979). ASME Steam Tables. American Society of Mechanical Engineers.

Aziz, K., and Settari, A. (1979). Petroleum Reservoir Simulation. London: Applied Science

Publishers Ltd.

Baker, R., and Heron, G. (2004). In-situ delivery of heat by thermal conduction and steam

injection for improved DNAPL remediation.

Baker, R., Heron, G., Hiester, U., Koschitzky, H.-P., Trötschler, O., Färber, A., and

Kuhlman, M. (2006). DNAPL removal from the saturated zone using thermal wells. In

B. M. Sass (Ed.) Remediation of Chlorinated and Recalcitrant Compounds: Proceedings

of the Fifth International Conference. Battelle, Columbus, OH.

82
REFERENCES 83

Baker, R., and Kuhlman, M. (2002). A description of the mechanics of in-situ thermal

destruction (ISTD) reactions. In H. El-Akabi (Ed.) Proceedings of the 2nd International

Conf. on Oxidation and Reduction Technologies for Soil and Groundwater, ORTs-2 .

Baston, D. P., Heron, G., and Kueper, B. (2007). Assessing the influence of ground water

inflow on thermal conductive heating in fractured rock. In Proceedings of the U.S. EPA

/ NGWA Fractured Rock Conference. Westerville, OH: NGWA Press.

Bear, J. (1972). Dynamics of Fluids in Porous Media. New York: Dover Publications, Inc.

Beck, A. E. (1976). An improved method of computing the thermal conductivity of fluid-

filled sedimentary rocks. Geophysics, 41 (1), 133–144.

Beck, J., Cole, K., Haji-Sheikh, A., and Litkouhi, B. (1992). Heat Conduction Using Green’s

Functions. Hemisphere.

Beyke, G., and Fleming, D. (2005). In situ thermal remediation of DNAPL and LNAPL

using electrical resistive heating. Remediation Journal , 15 (3), 5–22.

Birch, F., and Clark, H. (1940a). The thermal conductivity of rocks and its dependence upon

temperature and composition (Part I). American Journal of Science, 238 (8), 529–558.

Birch, F., and Clark, H. (1940b). The thermal conductivity of rocks and its dependence

upon temperature and composition (Part II). American Journal of Science, 238 (9),

613–635.

Bodvarsson, G. (1969). On the temperature of water flowing through fractures. Journal of

Geophysical Research, 74 (8), 1987–1992.

Bodvarsson, G. (1972). Thermal problems in the siting of reinjection wells. Geothermics,

1 (2), 63–66.
REFERENCES 84

Bodvarsson, G. S., and Tsang, C. F. (1982). Injection and thermal breakthrough in fractured

geothermal reservoirs. Journal of Geophysical Research, 87 (B2), 1031–1048.

Buntebarth, G. (1984). Geothermics. New York: Springer-Verlag.

Burghardt, J. M. (2007). Laboratory Study Evaluating Thermal Remediation of Tetra-

chloroethyene Impacted Soil . Master’s thesis, Queen’s University.

Carslaw, H., and Jaeger, J. (1959). Conduction of Heat in Solids. New York: Oxford

University Press, 2nd ed.

C̆ermák, V., and Rybach, L. (1982). Thermal conductivity and specific heat of minerals

and rocks. In G. Angenheister (Ed.) Landolt-Börnstein: Numerical Data and Functional

Relationships in Science and Tehcnology, Group V (Geophysics and Space Research),

Volume 1a (Physical Properties of Rocks), (pp. 305–343). Berlin-Heidelberg: Springer.

Cheng, A.-D., Ghassemi, A., and Detournay, E. (2001). Integral equation solution of heat

extraction from a fracture in hot dry rock. International Journal for Numerical and

Analytical Methods in Geomechanics, 25 , 1327–1338.

Clauser, C., and Huenges, E. (1995). Thermal conductivity of rocks and minerals. In

Rock Physics and Phase Relations – A Handbook of Physical Constants, (pp. 105–126).

American Geophysical Union.

Costanza, J. (2005). Degradation of Tetrachloroethylene and Trichloroethylene Under Ther-

mal Remediation Conditions. Ph.D. thesis, Georgia Institute of Technology.

Costanza, J., Pennell, K. D., and Mulholland, J. A. (2003). Laboratory study on the

transformation of trichloroethylene under thermal source zone remediation conditions.

In K. J. Hatcher (Ed.) Proceedings of the 2003 Georgia Water Resources Conference.

Atlanta, GA.
REFERENCES 85

Cotta, R., Ungs, M., and Mikhailov, M. (2003). Contaminant transport in a finite fractured

porous medium: integral transforms and lumped-differential formulations. Annals of

Nuclear Energy, 30 , 261–285.

Davis, E., Akladiss, N., Hoey, R., Brandon, B., Nalipinski, M., Carroll, S., Herom, G.,

Novakowski, K., and Udell, K. (2005). Steam enhanced remediation research for DNAPL

in fractured rock. Report EPA/540/R-05/010, Environmental Protection Agency.

Davis, E. L. (1997). How heat can enhance in-situ soil and aquifer remediation: Important

chemical properties and guidance on choosing the appropriate technique. Ground Water

Issue EPA/540/S-97/502, Environmental Protection Agency.

De Hoog, F., Knight, J., and Stokes, A. (1982). An improved method for the numerical

inversion of Laplace transforms. SIAM Journal on Scientific and Statistical Computing,

3 (3), 357–366.

de Marsily, G. (1986). Quantitative Hydrogeology. New York: Academic Press, Inc.

Delves, L., and Mohamed, J. (1985). Computational methods for integral equations. Cam-

bridge University Press.

Domenico, P. A., and Schwartz, F. W. (1998). Physical and Chemical Hydrogeology. New

York: Wiley, 2nd ed.

Elliott, L. J., Pope, G. A., and Johns, R. T. (2003). In-situ thermal remediation of con-

taminant below the water table. SPE 81204.

Elliott, L. J., Pope, G. A., and Johns, R. T. (2004). Multidimensional numerical reservoir

simulation of thermal remediation of contaminants below the water table. In A. Gavaskar,

and A. Chen (Eds.) Proceedings of the Fourth International Conference on Remediation

of Chlorinated and Recalcitrant Compounds. Monterey, CA: Battelle Press.


REFERENCES 86

Fair, J. R. (1987). Distillation. In R. W. Rousseau (Ed.) Handbook of Separation Process

Technology, (pp. 229–335). John Wiley & Sons.

Falta, R. W., Pruess, K., Finsterle, S., and Battistelli, A. (1995). T2VOC user’s guide.

Report LBL-36400, Lawrence Berkeley Laboratory.

Falta, R. W., Pruess, K., Javandel, I., and Witherspoon, P. A. (1992). Numerical modeling

of steam injection for the removal of nonaqueous phase liquids from the subsurface. 2.

Code validation and application. Water Resources Research, 28 (2), 451–465.

Feenstra, S., Cherry, J., Sudicky, E., and Haq, Z. (1984). Matrix diffusion effects on con-

taminant migration from an injection well in fractured sandstone. Ground Water , 22 (3),

307–316.

Ferziger, J. H., and Perić, M. (1996). Computational Methods for Fluid Dynamics. Berlin-

Heidelberg: Springer-Verlag.

Fetter, C. (1993). Contaminant Hydrogeology. New York: Macmillan.

Fourier, J. (1955). The Analytical Theory of Heat. New York: Dover.

Frenklach, M. (1990). Production of polycyclic aromatic hydrocarbons in chlorine containing

environments. Combustion Science and Technology, 74 , 283–296.

Ghassemi, A., Tarasovs, S., and Cheng, A.-D. (2003). An integral equation solution for

three-dimensional heat extraction from planar fracture in hot dry rock. International

Journal for Numerical and Analytical Methods in Geomechanics, 27 , 989–1004.

Gringarten, A., Witherspoon, P., and Ohnishi, Y. (1975). Theory of heat extraction from

fractured hot dry rock. Journal of Geophysical Research, 80 (8), 1120–1124.

Grisak, G., and Pickens, J. (1981). An analytical solution for solute transport through

fractured media with matrix diffusion. Journal of Hydrology, 52 , 47–57.


REFERENCES 87

Gudbjerg, J., Heron, T., Sonnenborg, T. O., and Jensen, K. H. (2005). Three-dimensional

numerical modeling of steam override observed at a full-scale remediation of an unconfined

aquifer. Ground Water Monitoring & Remediation, 25 (3), 116–127.

Gudbjerg, J., Sonnenborg, T., and Jensen, K. (2004). Remediation of NAPL below the

water table by steam-induced heat conduction. Journal of Contaminant Hydrology, 72 ,

207–225.

Hansen, K. S., Conley, D. M., Vinegar, H. J., Coles, J. M., Menotti, J. L., and Stegemeier,

G. L. (1998). In situ thermal desorption of coal tar. In IGT/GRI International Symposium

on Environmental Biotechnologies and Site Remediation Technologies. Orlando, Florida.

Hart, D. J., Bradbury, K. R., and Feinstein, D. T. (2006). The vertical hydraulic conduc-

tivity of an aquitard at two spatial scales. Ground Water , 44 (2), 201–211.

Heath, R. C. (2004). Basic ground-water hydrology. Water-Supply Paper 2220, United

States Geological Survey.

Hein, G. L., Gierke, J. S., Hutzler, N. J., and Falta, R. W. (1997). Three-dimensional

experimental testing of a two-phase flow-modeling approach for air sparging. Ground

Water Monitoring & Remediation, 17 (3), 222–230.

Heron, G., Christensen, T. H., and Enfield, C. G. (1998a). Henry’s law constant for

trichloroethylene between 10 and 95 . Environmental Science & Technology, 32 (10),

1433–1437.

Heron, G., van Zutphen, M., Christensen, T., and Enfield, C. (1998b). Soil heatinf for

enhanced remediation of chlorinated solvents: A laboratory study on resistive heating

and vapor extraction in a silty, low-permeable soil contaminated with trichloroethylene.

Environmental Science & Technology, 32 (10), 1474–1481.


REFERENCES 88

Heuer, N., Küpper, T., and Windelberg, D. (1991). Mathematical model of a hot dry rock

system. Geophysical Journal International , 105 , 659–664.

Hiester, U., Gand, A. K., Färber, A., and Koschitzky, H.-P. (2004). Quantitative com-

parison of the remediation efficiency between conventional SVE and additional thermal

well application. In Petroleum Hydrocarbons and Organic Chemicals in Ground Water:

Prevention, Assessment and Remediation Conference. Costa Mesa, CA.

Hiester, U., Koschitzky, H.-P., Trötscler, O., Färber, A., Baker, R. S., LaChance, J. C.,

Heron, G., and Kuhlman, M. (2006). Thermal well operation in the saturated zone – new

options for DNAPL remediation. Land Contamination & Reclamation, 14 (2), 615–619.

Hiester, U., and Schenk, V. (2005). In-situ thermal remediation: ecological and economic

advantages of the TUBA and THERIS methods. In ConSoil 2005 Proceedings, (pp.

1581–1587). Bordeaux.

Horseman, S., and McEwen, T. (1996). Thermal constraints on disposal of heat-emitting

waste in argillaceous rocks. Engineering Geology, 41 , 5–16.

Iben, I., Edelstein, W., Sheldon, R., Shapiro, A., Uzgiris, E., Scatena, C., Blaha, S., Sil-

verstein, W., Brown, G., Stegemeier, G., and Vinegar, H. (1996). Thermal blanket for

in-situ remediation of surficial contamination: A pilot test. Environmental Science &

Technology, 30 (11).

Incropera, F. P., and DeWitt, D. P. (2002). Fundamentals of Heat and Mass Transfer . New

York: John Wiley & Sons, 5th ed.

Ingebritsen, S., Sanford, W., and Neuzil, C. (2006). Groundwater in Geologic Processes.

Cambridge University Press.

Jaluria, Y., and Torrance, K. E. (2003). Computational Heat Transfer . London: Taylor &

Francis.
REFERENCES 89

Kaviany, M. (1995). Principles of Heat Transfer in Porous Media. Berlin-Heidelberg:

Springer, 2nd ed.

Keith, L., and Rimstidt, J. (1985). A numerical compaction model of overpressuring in

shales. Mathematical Geology, 17 (2), 115–135.

Kipp, K. L., Jr. (1997). Guide to the revised heat and solute transport simulator: HST3D

– version 2. Water Resources Investigation Report 97-4157, United States Geological

Survey.

Kling, T., Korkealaakso, J., and Saarenpää, J. (2004). Application of nonisothermal mul-

tiphase modeling to in situ soil remediation in Söderkulla. Vadose Zone Journal , 3 ,

901–908.

Knauss, K. G., Dibley, M. J., Leif, R. N., Mew, D. A., and Aines, R. D. (1999). Aqueous

oxidation of trichloroethene (TCE): a kinetic analysis. Applied Geochemistry, 14 , 531–

541.

Knauss, K. G., Dibley, M. J., Leif, R. N., Mew, D. A., and Aines, R. D. (2000). The

aqueous solubility of trichloroethene (TCE) and tetrachloroethene (PCE) asa function of

temperature. Applied Geochemistry, 15 , 501–502.

Kocabas, I. (2004). Thermal transients during nonisothermal fluid injection into oil reser-

voirs. Journal of Petroleum Science and Engineering, 42 , 133–144.

Kolditz, O. (1995). Modelling flow and heat transfer in fractured rocks: Dimensional effect

of matrix heat diffusion. Geothermics, 24 (3), 421–237.

Kolditz, O. (2002). Computational Methods in Environmental Fluid Mechanics. Berlin:

Springer.
REFERENCES 90

Kueper, B., Wealthall, G., Smith, J., Leharne, S., and Lerner, D. (2003). An illustrated

handbook of DNAPL transport and fate in the subsurface. R&D Publication 133, U.K.

Environment Agency.

Kueper, B. H., and McWhorter, D. B. (1991). The behavior of dense, nonaqueous phase

liquids in fractured clay and rock. Ground Water , 29 (5), 716–728.

Kunii, D., and Smith, J. (1960). Heat transfer characteristics of porous rocks. American

Institute of Chemical Engineers Journal , 6 (1), 71–78.

Kunkel, A. M., Seibert, J. J., Elliott, L. J., Kelley, R., Katz, L. E., and Pope, G. A. (2006).

Remediation of elemental mercury using in situ thermal desorption ISTD. Environmental

Science & Technology, 40 (7), 2384–2389.

LaChance, J., Heron, G., and Baker, R. (2006). Verification of an improved approach for

implementing in-situ thermal desorption for the remediation of chlorinated solvents. In

B. M. Sass (Ed.) Remediation of Chlorinated and Recalcitrant Compounds: Proceedings

of the Fifth International Conference. Colombus, OH: Battelle Press.

Lauwerier, H. (1955). The transport of heat in an oil layer caused by the injection of hot

fluid. Applied Scientific Research A, 5 (2–3), 145–150.

Leverett, M. (1941). Capillary behavior in porous solids. Trans. AIME, Petrol. Div., 142 ,

152–169.

Lighty, J. S., Pershing, D. W., Cundy, V. A., and Linz, D. G. (1988). Characterization

of thermal desorption phenomena for the cleanup of contaminated soil. Nuclear and

Chemical Waste Management, 8 , 225–237.

Lowell, R. (1976). Comments on ‘Theory of heat extraction from fractured hot dry rock’ by

A.C. Gringarten, P.A. Witherspoon, and Yuzo Ohnishi. Journal of Geophysical Research,

81 (2), 359.
REFERENCES 91

McCray, J. E., and Falta, R. W. (1996). Defining the air sparging radius of influence for

groundwater remediation. Journal of Contaminant Hydrology, 24 , 25–52.

Merino, J., and Bucalá, V. (2007). Effect of temperature on the release of hexadecane from

soil by thermal treatment. Journal of Hazardous Materials, 143 , 455–461.

Molson, J., and Frind, E. (2002). HEATFLOW version 2.0 user guide. Tech. rep., Depart-

ment of Earth Sciences, University of Waterloo.

Morel-Seytoux, H. J., Meyer, P. D., Nachabe, M., Touma, J., and van Genuchten, M. (1996).

Parameter equivalence for the Brooks-Corey and van Genuchten soil characteristics: Pre-

serving the effective capillary drive. Water Resources Research, 32 , 1251–1258.

Moridis, G., and Pruess, K. (1995). Flow and transport simulations using T2CG1, a pack-

age of conjugate gradient solvers for the TOUGH2 family of codes. Report LBL-36235,

Lawrence Berkeley Laboratory.

Moridis, G. J., and Pruess, K. (1997). T2SOLV: An enhanced package of solvers for the

TOUGH2 family of reservoir simulation codes. Report LBNL-40933, Lawrence Berkeley

National Laboratory.

Narasimhan, T., and Witherspoon, P. (1976). An integrated finite difference method for

analyzing fluid flow in porous media. Water Resources Research, 12 (1), 57–64.

National Research Council (1994). Alternatives for Ground Water Cleanup. Washington:

National Academy Press.

National Research Council (1996). Rock Fractures and Fluid Flow: Contemporary Under-

standing and Applications. Washington: National Academy Press.

National Research Council (2004). Contaminants in the Subsurface: Source Zone Assess-

ment and Remediation. Washington: National Academy Press.


REFERENCES 92

Newmark, R., Aines, R., Knauss, K., Leif, R., Chiarappa, M., Hudson, B., Carrigan, C.,

Thompson, A., Richards, J., Eaker, C., Wiedner, R., and Sciarotta, T. (1998). In-situ

destruction of contaminants via hydrous pyrolysis / oxidation: Visalia field test. Report

UCRL-ID-132671, Lawrence Livermore National Laboratory.

Nogami, T., and Li, M. (2003). Consolidation of clay with a system of vertical and horizontal

drains. Journal of Geotechnical and Geoenvironmental Engineering, 129 (9), 838–848.

Özişik, M. N. (1980). Heat Conduction. New York: Wiley.

Palagi, C. L., and Aziz, K. (1994). Use of Voronoi grid in reservoir simulation. SPE

Advanced Technology Series, 2 (2), 69–77.

Palciauskas, V., and Domenico, P. (1982). Characterization of drained and undrained

response of thermally loaded repository rocks. Water Resources Research, 18 (2), 281–

290.

Peaceman, D. W. (1977). Fundamentals of Numerical Reservoir Simulation. Amsterdam:

Elsevier.

Pruess, K. (1992). Brief guide to the MINC-method for modeling flow and transport in

fractured media. Report LBL-32195, Lawrence Berkeley Laboratory.

Pruess, K. (2004). The TOUGH codes – a family of simulation tools for multiphase flow

and transport processes in permeable media. Vadose Zone Journal , 3 , 738–746.

Pruess, K., and Battistelli, A. (2002). TMVOC, a numerical simulator for three-phase

non-isothermal flows of multicomponent hydrocarbon mixtures in saturated-unsaturated

heterogeneous media. Report LBNL-49375, Lawrence Berkeley National Laboratory.

Pruess, K., and Narasimhan, T. (1982). A practical method for modeling fluid and heat

flow in fractured porous media. Report LBL-13487, Lawrence Berkeley Laboratory.


REFERENCES 93

Pruess, K., Oldenburg, C., and Moridis, G. (1999). TOUGH2 User’s Guide, Version 2.0.

Report LBNL-43134, Lawrence Berkeley National Laboratory.

Pruess, K., and Tsang, Y. (1994). Thermal modeling for a potential high-level nuclear

waste repository at Yucca Mountain, Nevada. Report LBL-35381, Lawrence Berkeley

Laboratory.

Pruess, K., Tsang, Y., and Wang, J. (1984). Numerical studies of fluid and heat flow near

high-level nuclear waste packages emplaced in partially saturated fractured tuff. Report

LBL-18552, Lawrence Berkeley Laboratory.

Pruess, K., Wang, J., and Tsang, Y. (1990a). On thermohydrologic conditions near high-

level nuclear wastes emplaced in partially saturated fractured tuff: 1. Simulation studies

with explicit consideration of fracture effects. Water Resources Research, 26 (6), 1235–

1248.

Pruess, K., Wang, J., and Tsang, Y. (1990b). On thermohydrologic conditions near high-

level nuclear wastes emplaced in partially saturated fractured tuff: 2. Effective continuum

approximation. Water Resources Research, 26 (6), 1249–1261.

Ratcliffe, E. (1959). Thermal conductivities of fused and crystalline quartz. British Journal

of Applied Physics, 10 , 22–25.

Reynolds, D. A., and Kueper, B. H. (2003). Multiphase flow and transport through fractured

heterogeneous porous media. Journal of Contaminant Hydrology, 71 , 89–110.

Robertson, E. C. (1988). Thermal properties of rocks. Open-File Report 88-441, United

States Geological Survey.

Seipold, U. (1998). Temperature dependence of thermal transport propreties of crystalline

rocks – a general law. Tectonophysics, 291 , 161–171.


REFERENCES 94

Seol, Y., Kneafsey, T. J., and Ito, K. (2005). An evaluation of the active fracture concept

with modeling unsaturated flow and transport in a fractured meter-sized block of rock.

Report LBNL-52818, Lawrence Berkeley National Laboratory.

She, H. Y., and Sleep, B. E. (1998). The effect of temperature on capillary pressure-

saturation relationships for air-water and perchloroethylene-water systems. Water Re-

sources Research, 34 (10), 2587–2597.

Simon, J. A. (2006). Editor’s perspective – an analysis of the Battelle remediation confer-

ences as a bellwether for treatment technology trends – part 2. Remediation Journal ,

(pp. 1–3).

Sleep, B. E., and McClure, P. D. (2001). The effect of temperature on adsorption of organic

compounds to soils. Canadian Geotechnical Journal , 38 , 46–52.

Somerton, W., Mehta, M., and Dean, G. (1965). Thermal alteration of sandstones. Journal

of Petroleum Technology, 17 (5), 589–593.

Somerton, W. H. (1958). Some thermal characteristics of porous rocks. Petroleum Trans-

actions, AIME , 213 , 375–378.

Stegemeier, G., and Vinegar, H. (2001). Thermal conduction heating for in-situ thermal

desorption of soils. In C. H. Oh (Ed.) Hazardous & Radioactive Waste Treatment Tech-

nologies Handbook , chap. 4.6-1. Boca Raton, Florida: CRC Press.

Stegemeier, G., and Vinegar, H. (2003). Heater element for use in an in situ thermal

desorption soil remediation system. U.S. Patent3.

Sudicky, E., and Frind, E. (1982). Contaminant transport in fractured porous media:

Analytical solutions for a system of parallel fractures. Water Resources Research, 18 (4),

1634–1642.
REFERENCES 95

Tang, D., Frind, E., and Sudicky, E. (1981). Contaminant transport in fractured porous

media: Analytical solution for a single fracture. Water Resources Research, 17 (3), 555–

564.

Therrien, R., McLaren, R., Sudicky, E., and Panday, S. (2007). HydroGeoSphere: A three-

dimensional numerical model describing fully-integrated subsurface and surface flow and

solute transport. Tech. rep.

Thomson, M. J., Higgins, B. S., Lucas, D., Koshland, C. P., and Sawyer, R. F. (1994).

Phosgene formation from 1,1,1-trichloroethane oxidation. Combustion and Flame, 98 ,

350–360.

U.S. EPA (2001). Correcting the Henrys Law constant for soil temperature. Tech. rep.

U.S. EPA (2002). Hex pit remediation project in-situ thermal desorption (ISTD) remedy

failure assessment report. Tech. rep., Rocky Mountain Arsenal Remediation Venture

Office.

Uzgiris, E. E., Edelstein, W. A., Philipp, H. R., and Iben, I. T. (1995). Complex thermal

desorption of PCBs from soil. Chemosphere, 30 (2), 377–387.

van Genuchten, M. (1980). A closed-form equation for predicting the hydraulic conductivity

of unsaturated soils. Soil Science Society of America Journal , 44 , 892–898.

Vinegar, H., Rosen, R., Stegemeier, G., Bonn, M., Conley, D., Phillips, S., Hirsch, J., Carl,

F., Steed, J., Arrington, D., Brunette, P., Mueller, W., and Siedhoff, T. (1997). In Situ

thermal desorption (ISTD) of PCBs. In Proceedings of the HazWaste / World Superfund

XVIII Conference. Washington, D.C.

Vosteen, H.-D., and Schellschmidt, R. (2003). Influence of temperature on thermal conduc-

tivity, thermal capacity, and thermal diffusivity for different types of rock. Physics and

Chemistry of the Earth, 28 , 499–509.


REFERENCES 96

Washburn, E. (2003). International Critical Tables of Numerical Data, Physics, Chemistry

and Technology. Knovel, first online ed.

Weeks, W. (1966). Numerical inversion of Laplace transforms using Laguerre functions.

Journal of the Association for Computing Machinery, 13 (3), 419–426.

White, M., and Oostrom, M. (2006). STOMP: Subsurface Transport Over Multiple Phases,

Version 4.0 User’s Guide. Report PNNL-15782, Pacific Northwest National Laboratory.

Witherspoon, P., Wang, J., Iwai, K., and Gale, J. (1980). Validity of cubic law for fluid

flow in a deformable rock fracture. Water Resources Research, 16 (6), 1016–1024.

Woodside, W., and Messmer, J. (1961). Thermal conductivity of porous media. Journal of

Applied Physics, 32 (9), 1688–1706.

Yang, J., Latychev, K., and Edwards, R. (1998). Numerical computation of hydrothermal

fluid circulation in fractured Earth structures. Geophysics Journal International , 135 ,

627–649.

Yasuhara, A. (1993). Thermal decomposition of tetrachloroethylene. Chemosphere, 26 (8),

1507–1512.

Yasuhara, A., and Morita, M. (1990). Formation of chlorinated compounds in pyrolysis of

trichloroethylene. Chemosphere, 21 (4–5), 479–486.


Appendix A

Semi-analytical Solutions

A.1 Background

Heat transport in fractured rock can be described by a series of coupled partial differential

equations. When heat conduction in the rock is considered to be one-dimensional, the gov-

erning equations are reduced to a more easily solvable set of ordinary differential equations.

While this is a reasonable approximation for many applications, such as the analysis of

heat extraction from a geothermal reservoir, the one-dimensional approximation can not be

used when heat is generated within the rock matrix. In order to model thermal conductive

heating, multidimensional conduction must be included.

To the author’s knowledge, the semi-analytical solution of Cheng et al. (2001) is the only

solution to the problem that includes two-dimensional conduction in the rock without using

a lumping procedure. In this section, the Cheng et al. (2001) solution will be extended to

include a heat source from a thermal well, as well as a set of parallel fractures.

97
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 98

A.1.1 Governing Equations

Transient heat conduction in a two-dimensional block of rock is described by the heat

conduction equation:
∂2T ∂2T g 1 ∂T
∇2 T = + =− + (A.1)
∂x2 ∂y 2 Kr αr ∂t

An appropriate differential equation for heat transfer in the fracture can be developed

by considering a small control volume of a fracture. As shown in Figure A.1, the upper half

of a fracture element is considered, having x, y, z dimensions of ∆x, 2e , 1. The midpoint of

the fracture (y = − 2e ), which forms the lower boundary of the control volume, is a symmetry

boundary, across which no heat is transferred.


 
x  2 V 

Δ  
 

Figure A.1: Heat balance

Because energy is conserved within the control volume,

 e  ∂T
ρw cw ∆x = rate of heat accumulation − rate of heat loss (A.2)
2 ∂t

For simplicity, axial conduction within the fracture is neglected, and transverse conduc-

tion is assumed to occur instantaneously, due to the small scale of 2e . Therefore, only two
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 99

sources contribute to heat accumulation in the control volume:

rate of heat accumulation = convection into CV + conduction from matrix into CV


e
∂T
= ρw cw v T (x) + Kr ∆x (A.3)
2 ∂y y=e/2

Similarly,

rate of heat loss = convection from CV


e
= ρw cw v (A.4)
2

Substituting (A.3) and (A.4) into the original statement of conservation (A.2) gives,

upon rearranging:


∂T T (x) − T (x + ∆x) 2Kr ∂T
= ρw cw v + (A.5)
∂t ∆x e ∂y y=e/2

Taking the limit as ∆x → 0:


∂T ∂T 2Kr ∂T
= −ρw cw v + (A.6)
∂t ∂x e ∂y y=e/2

Equation (A.6) is equivalent to the heat balance developed by Gringarten et al. (1975),

among others. Cheng et al. (2001) showed that the time partial derivative term may be

neglected from (A.6) without large error. A verification of this conclusion for the conditions

of thermal conductive heating may be found in Appendix B. Eliminating this term and

rearranging gives:

∂T 2Kr ∂T
= (A.7)
∂x ρw cw ve ∂y y=e/2

It should be noted that, although the time partial derivative has been eliminated, the
∂T ∂T
temporal dependence of the problem is retained through ∂x and ∂y , both of which vary
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 100

with time.

A.1.2 Method of Solution

All of the following solutions are developed using Green’s functions. Green’s functions

provide a modular solution framework through which problem parameters such as boundary

conditions and heat generation rates may be easily modified without changing the structure

of the solution. A general form for a solution to (A.1) using Green’s functions is given by

Beck et al. (1992, p. 55):

Z a2 Z b2
T (x, y, t) = GXIJ (x, t|x0 , 0)GY M N (y, t|y 0 , 0)F (x0 , y 0 )dx0 dy 0
x0 =a1 y 0 =b1
Z t Z a2 Z b2
α
+ GXIJ (x, t|x0 , τ )GY M N (y, t|y 0 , τ )g(x0 , y 0 , τ )dy 0 dx0 dτ
K τ =0 x0 =a1 y 0 =b1
+ Ix0 =a1 + Ix0 =a2 + Iy0 =b1 + Iy0 =b2 (A.8)

where GXIJ and GY M N are one-dimensional Green’s functions determined by the choice

of boundary conditions in the x and y directions, F (x0 , y 0 ) is a spatially-variable initial

condition, g(x0 , y 0 , τ ) is a point heat source, and the I-terms represent boundary condition

effects. To simplify the solution, we consider the case of homogeneous boundary conditions

(zero temperature or zero flux) and a uniform initial temperature of zero. Under these

conditions, the solution will have the form:

Z t Z a2 Z b2
α
T (x, y, t) = GXIJ (x, t|x0 , τ )GY M N (y, t|y 0 , τ )g(x0 , y 0 , τ )dy 0 dx0 dτ (A.9)
K τ =0 x0 =a1 y 0 =b1

The time integral may be removed by recognizing that (A.9) is the convolution of the

Green’s function G(x, y, t|x0 , y 0 , τ ) = GXIJ (x, t|x0 , τ )GY M N (y, t|y 0 , τ ) and the source func-

tion g(x0 , y 0 , τ ) with respect to time. The convolution property of the Laplace transform

states that:
Z t
F̄ (s)Ḡ(s) = F (τ )G(t − τ )dτ (A.10)
0
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 101

where s is the Laplace variable and the overbar denotes a Laplace-transformed quantity.

Applying this rule to (A.9) gives:

Z a2 Z b2
α
T̄ (x, y, s) = L[GXIJ (x, t|x0 , τ )GY M N (y, t|y 0 , τ )]L[g(x0 , y 0 )]dy 0 dx0 (A.11)
K x0 =a1 y 0 =b1

A.2 Single Fracture Solution

A.2.1 Problem Formulation

Using the Green’s function solution framework, one simple solution that can be constructed

is that of a single infinite-length fracture at y = 0 intersected by a heater well at x = W

in an infinite two-dimensional body of rock. In this case, the Green’s functions GXIJ and

GY M N in are equivalent:
!
(z − z 0 )2
 
1
GXIJ = GY M N = p exp − , z = x, y (A.12)
2 παr (t − τ ) 4αr (t − τ )

The product GXIJ GY M N is then given in the Laplace domain by:

r 
0 0 1 s 0 2 0 2
Gs (x, y|x , y ) = L [GXIJ GY M N ] = K0 [(x − x ) + (y − y ) ] (A.13)
2παr s αr

where αr is the thermal diffusivity of the rock, s is the Laplace variable, and K0 is the

zero-order modified Bessel function of the second kind. The effect the fracture can be math-

ematically introduced using the source/sink function ḡ(x0 , y 0 , s). If q̄(x0 , s) is the strength

of the source/sink along the fracture, the source function ḡ(x0 , y 0 , s) can be written as:

ḡ(x0 , y 0 , s) = q̄(x0 , s)δ(y 0 ) (A.14)

where q̄(x0 , s) is the strength of the fracture source/sink. Substituting the Laplace-space

source function (A.14) and Green’s function (A.13) into the general solution form (A.11)
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 102

gives:
Z L r 
1 0 s
T̄ (x, y, s) = q̄(x , s)K0 [(x − x ) + y ] dx0
0 2 2 (A.15)
2πKr 0 αr

This is the formulation used by Cheng et al. (2001) to describe heat extraction from

a hot dry rock geothermal reservoir. Although the domain in x extends from −∞ to ∞,

integration limits of 0 and L have been used to facilitate evaluation of the integral. Note that

the sources are modelled as a line and not a strip; consequently, the fracture is effectively

of zero aperture and occupies no physical space. However, the influence of the fracture

aperture on heat transfer is retained through the function q̄(x0 , s).

The effect of the thermal well is introduced by adding a term to the source function, a

line source of magnitude gw located at x = W and of infinite length in y:

gw
ḡ(x0 , y 0 , s) = q̄(x0 , s)δ(y 0 ) + δ(W − x) (A.16)
s

Substituting the revised source function (A.16) into the general solution (A.11) gives:

Z L r 
1 0 s
T̄ (x, y, s) = q̄(x , s)K0 [(x − x ) + y ] dx0
0 2 2
2πKr 0 αr
Z ∞ r 
g s
+ K0 [(W − x) + (y − y )] dy 0
2 0 (A.17)
2πKr s −∞ αr

Using this formulation, several thermal wells may be incorporated simply by adding

additional terms to (A.17). The unknown source function q̄(x, s) in the above equations

can be written in terms of the heat flux into the fracture:


∂ T̄ (x, y, s)
q̄(x, s) = −2Kr (A.18)
∂y
y=0

Making use of the heat balance in the fracture (A.7), the above can be written in terms
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 103

of the gradient in the x-direction:

∂ T̄ (x, 0, s)
q̄(x, s) = −veρw cw (A.19)
∂x

Substituting this into (A.17) gives:

−veρw cw L ∂ T̄ (x0 , 0, s)
Z r 
s
T̄ (x, y, s) = K0 [(x − x ) + y ] dx0
0 2 2
2πKr 0 ∂x0 αr
Z ∞ r 
gw s
+ K0 [(W − x) + (y − y )] dy 0
2 0 (A.20)
2πKr s −∞ αr

A.2.2 Solution Along Fracture/Matrix Interface

The first integral can be expanded by parts:

L
∂ T̄ (x0 , 0, s)
Z r 
s
K0 [(x − x ) + y ] dx0 =
0 2 2
0 ∂x0 αr
 r L
0 s 0 2 2
T̄ (x , 0, s)K0 [(x − x ) + y ]
αr 0
r Z L 0
r 
s 0 x − x s
− T̄ (x , 0, s) p K1 [(x − x ) + y ] dx0
0 2 2
αr 0 y 2 + (x − x0 )2 αr
(A.21)

Substitution of the above into (A.17) gives:

 r  r 
veρw cw s 2 2
s 2 2
T̄ (x, y, s) = − T̄ (L, 0, s)K0 [(L − x) + y ] − T̄ (0, 0, s)K0 [x + y ]
2πKr αr αr
#
x − x0
r Z L r 
s 0 s
− T̄ (x , 0, s) p K1 [(x − x ) + y ] dx0
0 2 2
αr 0 y 2 + (x − x0 )2 αr
Z ∞ r 
gw s
+ K0 [(W − x)2 + (y − y 0 )] dy 0 (A.22)
2πKr s −∞ αr

The temperature in the fracture (y = 0) is thus described by the following integral


APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 104

equation:

r  r 
veρw cw s veρw cw s
T̄ (x, 0, s) = − T̄ (L, 0, s)K0 (L − x) + T̄ (0, 0, s)K0 x
2πKr αr 2πKr αr
x − x0
r Z L r 
veρw cw s 0 s
+ T̄ (x , 0, s) K1 |x − x | dx0
0
2πKr αr 0 |x − x0 | αr
Z ∞ r 
gw s
+ K0 [(W − x) + (y ) ] dy 0
2 0 2 (A.23)
2πKr s −∞ αr

To remove the singularity in the third term, a subtraction method is used (e.g. Delves

and Mohamed, 1985, p. 268). The following similar integral is evaluated analytically:

L
x − x0
r Z r 
s s
T̄ (x, 0, s) K1 |x − x | dx0 =
0
αr 0 |x − x0 | αr
 r  r 
s s
T̄ (x, 0, s) K0 (L − x) − K0 x (A.24)
αr αr

2πKr
Multiplying (A.23) through by veρw cw and subtracting (A.24) gives:

Z ∞ r 
gw s
K0 [(W − x) + (y ) dy 0 =
2 0 2
veρw cw s −∞ α r
r  r 
s s
T̄ (L, 0, s)K0 (L − x) − T̄ (0, 0, s)K0 x
αr αr
 r  r 
2πKr s s
+ − K0 (L − x) + K0 x T̄ (x, 0, s)
veρw cw αr αr
x − x0
r Z L r 
s 0 s
− [T̄ (x , 0, s) − T̄ (x, 0, s)] K1 |x − x | dx0
0
(A.25)
αr 0 |x − x0 | αr

Equation (A.25) can be numerically solved to determine the Laplace-space temperature

along the fracture-matrix boundary. When the final integral in (A.25) is approximated by a

numerical quadrature rule, a n × n system of nonhomogeneous linear equations will result,

with the Laplace temperatures at n discrete points along the fracture as the unknowns.
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 105

The system of equations is described by:

Z ∞ r  r  r 
gw s 2 0 2 0 s s
K0 [(W − xi ) + (y ) dy = Tn K0 (L − xi ) − T1 K0 xi
veρw cw s −∞ αr αr αr
 r  r 
2πKr s s
+ − K0 (L − xi ) + K0 xi T̄ (x, 0, s)
veρw cw αr αr
n r 
xi − xj
r X
s s
− [Tj − Ti ]wj K1 |xi − xj | , i = 0..n (A.26)
αr |xi − xj αr
j=1

Once the Laplace-space temperature is known, a numerical Laplace inversion algorithm

may be used to calculate the real-space temperature at any time.

A.2.3 Calculation of Temperature Within the Rock Matrix

Once the Laplace-space temperature along the fracture/matrix boundary is known from

solution of equation (A.26), temperatures within the rock may be calculated directly using

a discretized version of equation (A.22):

 r  r 
veρw cw s 2 2
s 2 2
T̄ (x, y, s) = − Tn K0 [(L − x) + y ] − T1 K0 [x + y ]
2πKr αr αr

n r 
x − xj
r X
s s
− Tj wj p K1 [(x − xj )2 + y 2 ] 
αr y 2 + (x − x j ) 2 αr
j=1
Z ∞ r 
gw s
+ K0 [(W − x) + (y − y )] dy 0
2 0 (A.27)
2πKr s −∞ αr

A.2.4 Implementation Details

The reader is referred to Section A.3.4 for a discussion of the influence of time units on the

accuracy of calculated temperatures within the rock matrix near the fracture plane.
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 106

A.3 Fracture Set Solution (Bessel Function Solution)

A.3.1 Problem Formulation

A useful modification of the preceding solution would allow for an infinite set of evenly-

spaced parallel, discrete fractures. One approach is to consider a large finite number of

fractures and explicitly model the effect of each fracture: 2m + 1 fracture pairs are assumed

to exist, located at y = ±kH, n = 0...m. As m is increased, the computed temperatures

near the central fracture approach those predicted by an infinite set of fractures. Using this

approach, a solution can be constructed from:

m Z L r 
1 X
0 s p
T̄ (x, y, s) = q¯k (x , s)K0 (x − x ) + (y − kH) dx0
0 2 2
2πKr α r
k=−m 0
Z ∞ r 
gw s
+ K0 [(W − x) + (y − y ) ] dy 0
2 0 2 (A.28)
2πKr s −∞ αr

where q̄k (x0 , s) represent the heat sink created by each of the 2m+1 fractures. The distances
p
from the heat sinks are given by rk = (x − x0 )2 + (y − kH)2 .

A.3.2 Solution Along Fracture/Matrix Interface

The solution follows the same procedure as was used for the single fracture. Substituting

in the fracture heat balances (A.7) and integrating by parts gives:

m  r 
veρw cw X s
T̄ (x, y, s) = − T̄ (L, kH, s)K0 [(L − x)2 + (kH − y)2 ]
2πKr αr
k=−m
x − x0
r Z L r 
s 0 s
+ T̄ (x , kH, s) p K1 [(kH − y) + (x − x ) ] dx0
2 0 2
αr 0 2
(kH − y) + (x − x ) 0 2 α r
r 
s 2
− T̄ (0, kH, s)K0 [x + (kH − y)2 ]
αr
Z ∞ r 
gw s
+ K0 [(W − x) + (y − y ) ] dy 0
2 0 2 (A.29)
2πKr s −∞ αr
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 107

If an adequate number of fractures is chosen, it can be assumed that the temperature

in all fractures is equal (T̄ (x0 , kH, s) = T̄ (x0 , 0, s)). Making this substitution and evaluating

at y = 0 gives the following:

m  r 
veρw cw X s 2 2
T̄ (x, 0, s) = − T̄ (L, 0, s)K0 [(L − x) + (kH) ]
2πKr αr
k=−m
x − x0
r Z L r 
s 0 s
+ T̄ (x , 0, s) p K1 [(x − x ) + (kH) ] dx0
0 2 2
αr 0 (kH)2 + (x − x0 )2 αr
r 
s 2
− T̄ (0, kH, s)K0 [x + (kH)2 ]
αr
Z ∞ r 
gw s
+ K0 [(W − x) + (y ) ] dy 0
2 0 2 (A.30)
2πKr s −∞ αr

As in the single fracture solution, regularization needs to be performed on the integral

on the second line of (A.30). When k = 0, the integral will reach a singularity at x0 = x.

Equation (A.24) is multiplied by veρw cw /2πKr and added to (A.30), giving:

2πKr
T̄ (x, 0, s) =
veρw cw
r  r 
s s
T̄ (0, 0, s)K0 x − T̄ (L, 0, s)K0 (L − x)
αr αr
 r  r 
s s
− T̄ (x, 0, s) K0 (L − x) − K0 x
αr αr
x − x0
r Z L r 
s 0 s
− [T̄ (x , 0, s) − T̄ (x, 0, s)] K1 k(x − x ) k dx0
0 2
αr 0 |x − x0 | αr
m  r 
X s 2 2
− T̄ (L, 0, s)K0 [(L − x) + (kH) ]
αr
k=−m,k6=0

x − x0
r Z L r 
s 0 s
+ T̄ (x , 0, s) p K1 [(kH) + (x − x ) ] dx0
2 0 2
αr 0 (kH)2 + (x − x0 )2 αr
r 
s 2 2
− T̄ (0, kH, s)K0 [x + (kH) ]
αr
Z ∞ r 
gw s
+ K0 [(W − x) + (y ) ] dy 0
2 0 2 (A.31)
veρw cw s −∞ αr

A numerical quadrature is used to express the boundary integrals in (A.31) in terms of


APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 108

summations. The following is a generalized discretization of the fracture-rock interface:

 r  r 
2πKr s s
Ti + K0 (L − xi ) − K0 xi =
veρw cw αr αr
m r  m r 
X s 2 X s
T1 K0 [x + (kH)2 ] − Tn K0 [(L − xi )2 + (kH)2 ]
αr i αr
k=−m k=−m
n r 
xi − xj
r X
s s
+ wj [Tj − Ti ] K1 |xi − xj |
αr |xi − xj | αr
j=1
m n q 
xi − xj
r
s X X
2 2
+ wj Tj p K1 (kH) + (xi − xj )
αr (kH)2 + (xi − xj )2
k=−m,k6=0 j=1
Z ∞ r 
gw s p
+ K0 (W − x) + (y ) dy 0
2 0 2 (A.32)
veρw cw s −∞ αr

Values for the abscissas (xi ) and weights (wj ) may be found in a handbook of quadrature

tables.

A.3.3 Calculation of Temperature Within the Rock Matrix

Once the temperature along the discretized fracture-rock boundary has been calculated, the

temperature at any point within the rock matrix may be rapidly calculated. Since y > 0,

the singularity in (A.29) is avoided, although error stemming from the evaluation of K0 (x)

near its singuarities at y = ±kH may be unacceptable. Applying the same quadrature rule

to (A.29) gives:

m  r 
veρw cw X s 2 2
T̄ (x, y, s) = − Tn K0 [(L − x) + (kH − y) ]
2πKr αr
k=−m
n r 
x − xj
r X
s s
+ Tj wj p K1 [(kH − y)2 + (x − xj )2 ]
αr (kH − y)2 + (x − xj )2 αr
j=1
r 
s 2
− T0 K0 [x + (kH − y)2 ]
αr
Z ∞ r 
gw s
+ K0 [(W − x) + (y − y ) ] dy 0
2 0 2 (A.33)
2πKr s −∞ αr
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 109

The temperature at points along the fracture plane should be evaluated by interpolating

between the previously calculated Ti rather then using the above equation.

A.3.4 Implementation Details

Number of Modelled Fracture Pairs

An important parameter of the solution is the number of fracture pairs to explicitly model.

If too few pairs are modelled, temperatures in the area of interest will be erroneously high.

However, the modelling of too many pairs will slow computations needlessly. In order to

determine an appropriate number of fracture pairs, a sensitivity analysis was conducted on

the number of fracture pairs. The temperature after heating at five points along the fracture

plane was modelled, using an increasing number N of fracture pairs (up to N = 40).
TN
The normalized temperatures TN =40 at each of the five points are plotted in Figure A.2.

From these data, it is clear that at least 15 pairs of fractures should be explicitly modelled

in order to represent an infinite system of fractures.

Choice of Units

Although the regularization allows (A.32) to be evaluated exactly along y = 0, interior

temperatures calculated from evaluating (A.33) near y = ±kH will have an unacceptable

error. This error is caused by the evaluation of the modified Bessel functions K0 (x) and

K1 (x) near their singularity at x = 0. A judicious choice of units will increase the magnitude

of the arguments, thereby reducing the error.

The following is a typical call of K0 as y approaches 0 or kH:

r 
s
K0 (x − xj )2 (A.34)
α

It can be shown that the dimensions of the argument are [T]. Therefore, a change

of time units can increase the magnitude of the function arguments, thus increasing the
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 110

1.4

1.2

1
Normalized Temperature

0.8

0.6
x = 100
x = 105
0.4 x = 110
x = 115
x = 120
0.2

0
0 5 10 15 20 25 30 35 40 45 50
Number of Fracture Sets Modelled

Figure A.2: Determining an appropriate number of fracture sets

distance from the singularity. In the De Hoog et al. (1982) algorithm, values of s are

inversely proportional to log t. When years or centuries are used as the time units, values

of t are small and values of s become large, thus avoiding the singularity and improving

accuracy. Even with these corrections, however, temperatures may be inaccurate near the

fracture-matrix boundary.

A.4 Fracture Set Solution (Improved Solution)

A.4.1 Problem Formulation

As noted in Section (A.3.3), the accuracy of the free space (infinite domain) Green’s function

solution may be unacceptable for interior points close to the fracture. By avoiding the use of
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 111

the Bessel functions, the following solution improves the accuracy of points calculated near

a singularity of the Green’s function. Further, since nonelementary functions are avoided,

computational time is reduced. The following Green’s functions are used Beck et al. (1992):
!
(x − x0 )2
 
0 1
GX00 (x, t|x , τ ) = p exp − (A.35)
2 πα(t − τ ) 4α(t − τ )

" #
−m2 π 2 α(t − τ ) mπy 0
  
0 1 X  mπy 
GY 22 (y, t|y , τ ) = 1+2 exp cos cos (A.36)
H H2 H H
m=1

where the subscript X00 indicates that the x-domain is bounded by two zeroth-type bound-

ary conditions, and Y 22 indicates that the y-domain is bounded by two second-type condi-

tions. Beck et al. (1992) recommend that (A.36) be used for “large” values of dimensionless
αt
time, when the Fourier number Fo = H2
exceeds 0.25. For the times to be considered in

thermal heating for fractured rock, Fo should always be large enough to used the above

function; however, this should be verified prior to computation. If Fo is small, the solution

may still be used, but a large number of terms should be evaluated to ensure convergence.

The Laplace transform of the product of these two functions is given by:

 r 
0 0 1 0 s
Gs (x, y|x , y ) = √
exp −|x − x |
2H sα α
 q   
0 s m2 π 2 mπy  mπy 0
X∞ exp −|x − x | α + H cos H cos H
+ √ (A.37)
m=1
H 2 s + m2 π 2 α

whose derivative with respect to x0 is equal to:

sgn(x − x0 )
 r 
∂Gs 0 0 0 s
(x, y|x , y ) = exp −|x − x |
∂x0 2Hα α
 q   
0 0| s m2 π 2 mπy  mπy 0
X∞ sgn(x − x ) exp −|x − x α + H cos H cos H
+ (A.38)

m=1
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 112

A.4.2 Solution Along Fracture/Matrix Interface

A solution can be constructed using (A.37) in exactly the same manner as the Bessel function

solution of Section (A.3). The form of the Green’s function solution from which the integral

equation is constructed is nearly identical (to A.17); only the limits of integration and the

Green’s function have been changed.

Z L Z H
1 g
T̄ (x, y, s) = q̄(x0 , s)Gs (x, y|x0 , 0)dx0 + Gs (x, y|W, y 0 )dy 0 (A.39)
2πKr 0 2πKr s 0

Like the Bessel function-based Green’s function, the derivative (A.38) is singular as

x0 → x; again, this singularity can be removed by subtraction. The solution is calculated

by solving the n × n system of equations described by:



n
veρw cw αr  X ∂Gs
Ti = − Tn Gs (xi , 0|L, 0) − T1 Gs (xi , 0|0, 0) − [Ti − Tj ]wj (xi , 0|xj , 0)
Kr  ∂x0
j=1

gw αr H
Z
− Ti [Gs (xi , 0|L, 0) + Gs (x, y|0, 0)]} + Gs (xi , y|W, y 0 )dy 0 (A.40)
Kr s 0

A.4.3 Calculation of Temperature Within the Rock Matrix

Interior points may then be calculated using:

(
veρw cw αr
T̄ (x, y, s) = − Tn Gs (x, y|L, 0) − T1 Gs (x, y|0, 0)
Kr
n
)
X ∂Gs
− [Tj − T̄ (x, 0, s)]wj (x, y|xj , 0) − T̄ (x, 0, s)[Gs (x, y|L, 0) + Gs (x, y|0, 0)]
∂x0
j=1

gw αr H
Z
+ Gs (x, y|W, y 0 )dy 0 (A.41)
Kr s 0

An excellent “perk” of this formulation is that the heater well terms can be evaluated
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 113

analytically:

H
Z r  r 
gw αr gw
0 0 αr s
Gs (x, y|W, y )dy = exp −|W − x| (A.42)
sKr 0 Kr s αr

A.4.4 Verifications

Disappearing Fractures

As the aperture of the fractures approaches zero, their importance should become negligible

and the computed temperature field should approach the solution for a continuous plane

source, as provided by Carslaw and Jaeger (1959, p. 263):

" 1 #
−(x − x0 )2 |x − x0 | |x − x0 |
  
gw t 2
T (x) = exp − erfc √ (A.43)
ρw cw πα 4αt 2α 2 αt

A test case was prepared with two heater wells, located at x = 30 m and x = 33 m. Each

well produced 40 W/m of heat for a period of 1 year. Figure A.3 shows the temperature

profile along along x for six values of time.

No Heater Well; Incoming Water is Heated

An alternative scenario exists in which the rock is heated not by a thermal well, but rather

by the injection of hot fluid. In this case, the solution should behave similarly to that of

Gringarten et al. (1975), whose analytical solution models one-dimensional heat conduction

from a set of equally-spaced parallel fractures. Previous numerical modelling work (Kolditz,

1995; Cheng et al., 2001) suggests that the solutions should differ at large time, where the

multidimensional solution should predict a lower temperature.

In notation consistent with this solution, Gringarten et al. (1975)’s solution can be
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 114

100
t = 33 days
90 t = 100 days
t = 166 days
80 t = 232 days
t = 365 days
70 Analytical Solution
erature (°C)

60

50
Tempe

40

30

20

10

0
20 22 24 26 28 30 32 34 36 38 40
Distance (m)

Figure A.3: Comparison of new solution with that of Carslaw and Jaeger (1959, p. 263) for
the case of zero fracture aperture.

written in Laplace space as:


!
1 ρw cw bvHs1/2
T̄ (x, 0, s) = exp xs1/2 tanh (A.44)
s Kr

Equation (A.44) can be inverted numerically, with the values of s calculated from the

dimensionless time tD :
(veρw cw )2  x
tD = t− (A.45)
Kr ρr cr v

Multidimensional conduction effects are far more pronounced for very low Peclet num-

bers, defined here as Pe = ve/αr . In order to facilitate comparison of the two solutions, the

hydrogeological parameters used by Cheng et al. (2001) are chosen, resulting in a Peclet

number of 19.18 (Table A.1).

After 100 days, the temperature profiles were calculated using the two solutions (Figure

A.4). It is important to note the sudden drop in temperature after x = 0 in the integral
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 115

Table A.1: Parameters used for verification against the Gringarten et al. (1975) solution
Parameter Value
v (m/day) 432
e (mm) 3
H (m) 10
ρr (kg/m3 ) 2750
cr (J/kg·K) 1000
Kr (W/m·K) 2.15

equation solution. Because this solution permits multidimensional conduction in the matrix,

heat may be lost to the area directly to the left of the injection point. This is not possible

with unidirectional conduction in the matrix, so it is to be expected that the solution of

Gringarten et al. (1975) should predict less heat loss from the fracture at small x.

1
0.9 Semianalytical Solution
0.8
Gringarten et al. (1975)
perature (°C)

0.7
0.6
0.5
04
0.4
Temp

0.3
0.2
0.1
0
0 25 50 75 100 125 150
Distance (m)

Figure A.4: Comparison of the present solution with that of Gringarten et al. (1975), t =100
days
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 116

Energy Balance

Another simple verification can be performed by adding up the heat contained everywhere

in a domain and comparing it to the heat added to the domain through the heater wells.

As long as the parameters and time scale are chosen such that no heat has left the modelled

area at x = L, the total heat in the domain can be calculated by:

n−1
X m−1
X  e  n−1
X
∆x∆yρr cr (Ti,j + Ti+1,j+1 )/2 + ∆xρw cw (Ti,1 + Ti+1,1 ) (A.46)
2
i=1 j=1 i=1

The parameters used for the energy balance are as follows. A single heater well at x = 30

provided heat a constant rate of 100 W/m. The modelled domain consisted of a set of 1

mm fractures spaced at 1 m; water flowed at a constant velocity of 20 m/day. After 1 year,

the heat content of the area x = 0 to x = 70 and y = 0 to y = H/2 − e/2 was calculated

using (A.46). Grid spacings were 0.2341 m in x and 0.0172 m in y. The heat balance ratio

was calculated to be:


heat found in domain
= 1.0004255. (A.47)
heat provided to domain

Laplace Inversion Algorithm

In order to verify the results of the numerical Laplace inversion, a fracture temperature

profile was calculated using two different Laplace inversion algorithm. Figure A.5 shows

the fracture temperature after 1 year as calculated using the algorithms proposed by De

Hoog et al. (1982) and by Weeks (1966).

Comparison to Numerical Solution

In order to further verify the analytical solution, a numerical model was used to run an

equivalent simulation. A Cartesian domain was created in TOUGH2 consisting of a 200 m

long strip, 1 m in depth and 0.5 m in height. In order to model the fracture, the domain

was divided into two materials, with properties given in Table A.2. The model also required
APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 117

100
90
De Hoog et al. (1980)
80
perature (°C) Weeks (1966)
70
60
50
40
Temp

30
20
10
0
0 10 20 30 40 50 60 70 80 90 100
Distance (m)

Figure A.5: Comparison of temperatures in fracture after one year of heating, as calculated
using the De Hoog et al. (1982) and Weeks (1966) algorithms.

capillary pressure and relative permeability functions for both materials, but these have no

influence at the comparison time, before the onset of boiling.

Initially, the domain was discretized using 182 non-uniform blocks in the x-direction

(parallel to the direction of flow), 1 block in the y-direction, and 15 non-uniform blocks in

the z-direction (normal to the fracture plane). Improved agreement with the semi-analytical

solution was achieved by progressively refining the discretization in the direction of flow.

Figure A.6, created using 309 blocks in the x-direction, shows the predicted temperature

profile in the fracture after 4 months of heating.


APPENDIX A. SEMI-ANALYTICAL SOLUTIONS 118

Table A.2: Numerical model material properties


Material
Property Units Variable Name ROCK FRACT
ρ kg/m3 DROK 2757 1000
φ - POR 0.01 0.99
kx m2 PER(1) 10−14 2.0833 × 10−8
ky m2 PER(2) 10−14 2.0833 × 10−8
kz m2 PER(3) 10−14 2.0833 × 10−8
Kwet W/m·K CWET 2.98 0.616
c J/kg·K SPHT 1180 4184

100
90 Semianalytical

80 TOUGH2 Discrete Fracture

70
ure (°C)

60
Temperatu

50
40
30
20
10
0
20 25 30 35 40 45 50 55 60

Distance (m)

Figure A.6: Fracture temperature after 4 months of heating, computed using semi-analytical
solution and TOUGH2.
Appendix B

Verification of Simplified Heat


Balance

In the fracture heat balance described in Section A.1.1, the heat storage term is omitted,

following the approach of Cheng et al. (2001). Although these authors verified the validity

of this simplification using numerical modelling, it is prudent to independently verify its

appropriateness for the conditions of subsurface heating using thermal conductive heating.

If the heat storage term is retained in the heat balance, a solution may be derived

using the method of Section A.4 without any modification. It will be shown here that the

temperature difference introduced by neglecting the heat storage term is negligible (on the

order of 10−4 degrees); yet, the increase in computing time is significant.

A heat balance for a control volume of a fracture is given by equation (A.6). Applying

the Laplace transformation to (A.6) gives:


∂ T̄ 2Kr ∂ T̄
sT̄ = −ρw cw v + (B.1)
∂x e ∂y y=e/2

Equation (A.39) is again used as the basis for the solution. When the heat storage

term is retained, the fracture-matrix heat exchange term q̄(x0 , s) takes on a somewhat more

119
APPENDIX B. VERIFICATION OF SIMPLIFIED HEAT BALANCE 120

complex form:
e e ∂ T̄
q̄(x0 , s) = −s T̄ − ρw cw v (B.2)
2 2 ∂x0

Substitution of (B.2) into (A.39) gives:

−se L
Z
T̄ (x, y, s) = T̄ (x0 , y, s)Gs (x, y|x0 , 0)dx0
2 0
ρw cw ve H ∂ T̄
Z
− G (x, y|x0 , 0)dx0
0 s
2 0 ∂x
gw αr H
Z
+ Gs (x, y|W, y 0 )dy 0 (B.3)
Kr s 0

At this point we set y = 0 to solve for the fracture temperature. The second integral is

then integrated by parts to remove the spatial derivative from its kernel:

αr se L
Z
T̄ (x, y, s) = − T̄ (x0 , 0, s)Gs (x, 0|x0 , 0)dx0
2Kr 0
 Z L
ρw cw veαr ∂Gs
− − T̄ (x0 , 0, s) 0 (x, 0|x0 , 0)dx0
2Kr 0 ∂x

0 0
+ T̄ (L, 0, s)Gs (L, 0|x , 0) − T̄ (0, 0, s)Gs (0, 0|x , 0)
Z H
gw αr
+ Gs (x, y|W, y 0 )dy 0 (B.4)
Kr s 0

Subtracting the singularity at |x0 − x| → 0 gives:

Z L
αr se
T̄ (x, y, s) = − [T̄ (x0 , 0, s) − T̄ (x, 0, s)]Gs (x, 0|x0 , 0)dx0
2Kr 0
Z L 
+ T̄ (x, 0, s) Gs (x, 0|x0 , 0)dx0
0
ρw cw veαr 
− T̄ (L, 0, s)Gs (L, 0|x0 , 0) − T̄ (0, 0, s)Gs (0, 0|x0 , 0)
2Kr
Z L
∂Gs
+ [T̄ (x0 , 0, s) − T̄ (x, 0, s)] 0 (x, 0|x0 , 0)dx0
0 ∂x
− T̄ (x0 , 0, s) [Gs (x, 0|L, 0) − Gs (x, 0|0, 0)]

gw αr H
Z
+ Gs (x, y|W, y 0 )dy 0 (B.5)
Kr s 0
APPENDIX B. VERIFICATION OF SIMPLIFIED HEAT BALANCE 121

In discretized form, the above equation is given as:


 
n Z L
−αr se  X 
Ti = [Tj − Ti ]wj Gs (xi , 0|xj , 0) + Ti Gs (xi , 0|x0 , 0)dx0
2Kr  0 
j=1

n
veρw cw αr  X ∂G
− Tn G(xi , 0|L, 0) − T1 GF (xi , 0|0, 0) − [Ti − Tj ]wj 0 (xi , 0|xj , 0)
Kr  ∂x
j=1
)
gw αr H
Z
− Ti [G(xi , 0|L, 0) + G(x, y|0, 0)] + G(xi , y|W, y 0 )dy 0 (B.6)
Kr s 0

Equations (A.40) and (B.6) differ only by the first line of terms, which result from the

inclusion of the heat storage term. The code used to solve the simplified version (A.40)

was modified to include the extra terms. The parameters for the base case, shown in Table

B.1 were used to determine the temperature in the fracture after 1 year. The maximum

computed temperature difference between the two solutions at any point during the model

period was 0.000349 ◦ C. Runtime is nearly doubled when the heat storage term is included

(see Table B.2). It is thus concluded that inclusion of the heat storage term produces a

negligible difference in predicted temperatures, while significantly increasing computational

time.

Table B.1: Parameters used for verification of heat storage term omission
Parameter Value Parameter Value Parameter Value
e (µm) 500 Kr (W/m·K) 2.59 gw (W/m) 100
H (m) 1 ρr (kg/m3 ) 2650 Wi (m) x = 30, 33
∇h -0.005 cr 1046 J/kg-K ∆x 0.2 m

Table B.2: Comparison of solution times when heat storage is included and neglected
Simulation Runtime
Heat storage term included (equation B.6) 6.16 hours
Heat storage term neglected (equation A.40) 3.27 hours
Appendix C

Numerical Discretization

C.1 Discretization of Fracture

C.1.1 Two-Dimensional Simulations

Although it is desirable to model fractures using a discrete fracture approach, the very small

gridblocks make numerical solution especially challenging. For most of the simulations in

this thesis, a “fracture zone” approach is used, whereby a high-permeability zone is used to

approximate the effect of a single fracture. Further discussion of this approach is found in

Section 2.6.1.

Using a 1 m fracture spacing, the base case from the semi-analytical parameter sensitivity

analysis (Table 3.1) is reproduced in TOUGH2 using both a discrete fracture and 10 cm

fracture zone discretization. Using these three solution methods, the temperature in the

fracture is computed.

After 122 days of heating, prior to the onset of boiling, the maximum percent difference

between the discrete fracture numerical model and the semi-analytical solution is 6.02%;

between the fracture zone numerical model and the semi-analytical solution it is 5.75%

(Figure C.1(a)). After 214 days of heating, the maximum percent diference between the

122
APPENDIX C. NUMERICAL DISCRETIZATION 123

fracture zone numerical model and the discrete fracture numerical model is 1.62 % (Figure

C.1(b)).

120 120
Semianalytical
110 110
TOUGH2 Discrete Fracture TOUGH2 Discrete Fracture
100 100
TOUGH2 10 cm Fracture Zone TOUGH2 10 cm Fracture Zone
90 90

erature (°C)
Temperature (°C)

80 80
70 70
60 60

Tempe
50 50
40 40
30 30
20 20
10 10
0 0
20 25 30 35 40 45 50 55 60 65 70 75 80 20 25 30 35 40 45 50 55 60 65 70 75 80
Distance (m) Distance (m)

(a) t = 122 days (b) t = 214 days

Figure C.1: Comparison of semi-analytical solution, discrete fracture numerical solution,


and “fracture zone” numerical solution before and after boiling.

Accurate representation of fractures is believed to be more difficult for larger values of

fracture spacing, where flow heterogeneity is higher and the effect of an individual fracture

is more significant. Using a 5 m fracture spacing, the discrete fracture numerical model

failed to converge at the onset of boiling. Just prior to this point, the maximum percent

difference between the fracture zone model and the discrete fracture model was calculated

to be 6.27%.

C.1.2 Simulations in Radial Geometry

The fracture zone approach was also used for numerical simulations in a radial geometry

(Chapter 4). Here, the importance of the fracture zone thickness is explored for the param-

eters corresponding to Run G5 (kb = 10−13 m2 , km = 10−18 m2 , fracture spacing = 2.5 m).

In this case, the fracture aperture was computed to be 144 µm. Therefore, for a discrete

fracture simulation, the first layer of cells should have a thickness of 72 µm.

Using a fracture layer cell thickness of 72 µm, the model failed to converge at the onset
APPENDIX C. NUMERICAL DISCRETIZATION 124

of boiling (Run G55). The model also failed to converge when the fracture layer thickness

was increased to 100 µm (Run G61). However, convergence was obtained using fracture

layer thicknesses of 500µm and above (Runs G56-G60).

Using a variety of fracture zone thicknesses, the temperature in the fracture at the along

the line at the upper edge of the domain (θ = 30◦ ). For each simulation, the maximum

percent temperature difference was calculated using Run G60 (the converging run with the

smallest fracture zone thickness) as the “true value”. These differences, along with the CPU

time for each run, are plotted in Figure C.2.

3.5 7.00%

perature Difference
CPU Time
3 6.00%
% Difference
CPU Time (hours)

2.5 5.00%

2 4.00%

15
1.5 3 00%
3.00%

Maximum Temp
1 2.00%

0.5 1.00%

0 0.00%
0 0.02 0.04 0.06 0.08 0.1 0.12
Thickness of Fracture Zone (m)

Figure C.2: Comparison of CPU time and error in temperature for different values of
fracture zone thickness.

It can be seen that some error is incurred in the use of a 10 cm fracture zone. Therefore,

all simulation runs were conducted using the smallest fracture zone thickness possible.

Most runs were able to converge with a fracture zone thickness of 500 µm, although it was

necessary to use a 1 mm fracture zone in some cases.


APPENDIX C. NUMERICAL DISCRETIZATION 125

C.2 Radial Discretization

C.2.1 Size of Domain

In order to avoid the influence of the outer boundary condition on the temperature and

pressure in the treatment zone, a domain size dependence study was conducted. Using the

base case run from the numerical study (Table 4.1), the value of the primary variables in a

the fracture were compared for three different values of rmax , the domain size in the radial

direction (Figure C.3).

From these figures it appears that there is little difference in any of the output variables

between a radial grid size of 25 m and 50 m. Consequently, a 25 m grid was used in all

simulations.

C.2.2 Radial Grid Refinement

The effect of grid refinement in the radial direction is shown in Figure C.4, where the

base case parameters from the numerical study (Table 4.1) are used to compute pressure,

temperature, and steam saturation against time in two reference gridblocks: one near the

center of the rock matrix, and one located in the fracture. For all of the simulations discussed

in Chapter 4, where the steam saturation is of greatest interest, the medium-level radial

discretization (Nr = 90) is used.

If the radial discretization is too coarse, the pressure spike that occurs prior to boiling

will be relayed radially inward as each cell boils during outward movement of the boiling

front. This will cause pressure oscillations in cells within the boiled zone (Figure C.4(a)). A

similar effect was observed by (Falta et al., 1992) in their study of steam injection into the

subsurface. When the discretization in the R-direction is refined, the magnitude of further

pressure oscillations is reduced.


APPENDIX C. NUMERICAL DISCRETIZATION 126

500
450
400

erature (°C)
350
Rmax = 10
300
250 Rmax = 25
Tempe 200 Rmax = 50
150
100
50
0
0 5 10 15 20 25 30 35 40 45 50
Radial Distance (m)
(a) Temperature Profiles

2.5

2.48
Presssure (bar)

2.46 Rmax = 10

2.44 Rmax = 25
Rmax = 50
2.42

2.4

2.38
0 5 10 15 20 25 30 35 40 45 50
Radial Distance (m)
(b) Pressure Profiles

1.2

1
Steam Saturation

0.8 Rmax = 10

0.6 Rmax = 25
Rmax = 50
0.4

0.2

0
0 5 10 15 20 25 30 35 40 45 50
Radial Distance (m)
(c) Gas Saturation Profiles

Figure C.3: Fracture primary variable profiles for three domain sizes
APPENDIX C. NUMERICAL DISCRETIZATION 127

700 249
Nr = 50 248
600
Nr = 90 247
500 Nr = 180
246
Pressure (kPa)

Pressure (kPa)
400 245
244
300 243
200 242 Nr = 50
241 Nr = 90
100 Nr = 180
240
0 239
0 100 200 300 0 100 200 300
Heating Time (days) Heating Time (days)

(a) Pressure near matrix center (b) Pressure in fracture

1.2 1.2

1 Steam Saaturation 1
Steam Saaturation

0.8 0.8

0.6 0.6

0.4 Nr = 50 0.4 Nr = 50


Nr = 90 Nr = 90
0.2 Nr = 180 0.2 Nr = 180

0 0
0 100 200 300 0 100 200 300
Heating Time (days) Heating Time (days)

(c) Steam saturation near matrix center (d) Steam saturation in fracture

400 400

350 350

300 300
Temperaature (°C)
Temperaature (°C)

250 250

200 200

150 150
Nr = 50
100 100 Nr = 50
Nr = 90
Nr = 90
50 Nr = 180 50
Nr = 180
0 0
0 100 200 300 0 100 200 300
Heating Time (days) Heating Time (days)

(e) Temperature near matrix center (f) Temperature in fracture

Figure C.4: Effect of radial discretization on pressure, steam saturation, and temperature
at two reference points.
Appendix D

Calculated Numerical Model Input


Parameters

A number of model input parameters are not treated as independent variables, but rather

are derived from other input parameters. An outline of all calculated variables is provided

in this appendix.

D.1 Fracture Zone Physical Properties

In all simulations, a “fracture zone” approach has been used, whereby a layer of gridblocks

of thickness ∆zf z is assigned properties reflective of a fracture of aperture e and an amount

of matrix material of thickness (∆zf z − e/2). In this way, many of the benefits of a discrete

fracture model are retained, while avoiding the numerical difficulties associated with very

small gridblocks. Some calculation is necessary to derive appropriate properties for the

gridblocks in the fracture zone.

128
APPENDIX D. CALCULATED NUMERICAL MODEL INPUT PARAMETERS 129

λ  kbulk  kmatrix  φmatrix  Krock 

Equation (C.2)

kfz  Equation (C.6)


Equation (C.18)  Equation (C.16) 

Equation (C.14)
Equation (C.7)

m  Pd‐mat  Pe‐frac  φfz 


Equations 
(C.10) & (C.11) 
Equation (C.19) 

Kwet‐mat 
αmat  Kdry‐mat 
Equation (C.19)
Equations 
(C.10) & (C.11) 
αfz 

Kwet‐fz 
Kdry‐fz 
 

Figure D.1: Outline of calculation of derived input parameters

D.1.1 Fracture Zone Permeability

For most simulations, the matrix permeability and bulk permeabilty are specified. However,

the fracture zone permeability must be calculated. The bulk permeability kb of the fractured

medium is equal to the weighted arithmetic mean of the fracture zone permeability kf z and

the matrix permeability km :

kf z ∆zf z + km ∆zm
kb = (D.1)
∆zf z + ∆zm

where ∆zf z is the thickness of the fracture zone and ∆zmat is the thickness of the

modeled portion of the matrix. Equation (D.1) can be rearranged to solve for the fracture

zone permeability kf z :
kb (∆zm + ∆zf z ) − km (∆zm )
kf z = (D.2)
∆zf z
APPENDIX D. CALCULATED NUMERICAL MODEL INPUT PARAMETERS 130

D.1.2 Fracture Aperture

The fracture aperture e is not used as a model input. However, it is necessary to calculate

the fracture aperture in order to calculate several model inputs, such as the fracture zone

porosity and capillary pressure parameters.

The permeability of the fracture zone is given by the weighted arithmetic mean of the

fracture permeability (kf rac ) and the permeability of the matrix material in the fracture

zone gridblocks (km ):


e e
 
kf rac 2 + km ∆zf z − 2
kf z = (D.3)
∆zf z

The permeability of the fracture itself is given by:

e2
kf rac = (D.4)
12

Substituting (D.4) into (D.3) gives:

 3
e e

24 + km ∆zf z − 2
kf z = (D.5)
∆zf z

The term km 2e is relatively small and can be ignored. The above equation can then be

simplified to solve explicitly for e:

q
e = 3 24∆zf z (kf z − km ) (D.6)

D.1.3 Fracture Zone Porosity

The porosity of a fracture is 1.0, while the porosity the rock matrix is given by φmat . The

porosity of the fracture zone is calculated from the weighted arithmetic mean of the fracture

porosity and the porosity of the matrix material in the fracture zone gridblocks:
APPENDIX D. CALCULATED NUMERICAL MODEL INPUT PARAMETERS 131

e e
 
1 2 + φm ∆zf z − 2
φf z = (D.7)
∆zf z

Since the fracture aperture is not an input to the model, it is back-calculated from the

fracture zone permeability, as shown in Section D.1.2. The fracture zone porosity is only

appreciably larger than the matrix porosity when the fracture aperture is large, i.e. e > 100

µm.

D.1.4 Thermal Conductivity

Thermal conductivity in TOUGH2 is calculated as a function of liquid saturation, using

(Pruess et al., 1999):

p
K(Sl ) = Kdry + Sl (Kwet − Kdry ) (D.8)

The wet and dry thermal conductivities Kwet and Kdry can be calculated using the geo-

metric mean thermal conductivity (Woodside and Messmer, 1961):

Kgeo = Kfφ Ks1−φ (D.9)

where Kf is the thermal conductivity of the pore fluid, and Ks is the thermal conductiv-

ity of the matrix material. When the pores are fully saturated with water (K = Kwet ), Kf

is the thermal conductivity of water, 0.587 W/m·K (Washburn, 2003). When the pores are

fully saturated with steam, Kf is taken to be 2.17 × 10−4 W/m·K (Washburn, 2003). Using

these values, the wet and dry thermal conductivities were calculated for all simulations as:

Kwet = (0.587)φ Kr1−φ (D.10)



Kdry = 2.17 × 10−4 Kr1−φ (D.11)
APPENDIX D. CALCULATED NUMERICAL MODEL INPUT PARAMETERS 132

D.2 Capillary Pressure

In both the rock matrix and the fractures, the capillary pressure-saturation relationship

of van Genuchten (1980) is used. In TOUGH2, the setting ICP = 7 implements this

relationship as (Pruess et al., 1999):

 1−m
Pcap = −P0 Se−1/m − 1 (D.12)

where m is a parameter related to the pore-size distribution, P0 = αvg /ρw g, and the

effective saturation Se is defined as:

Sl − Slr
Se = (D.13)
Sls − Slr

where Sl is the water saturation, Slr is the residual (irreducible) water saturation (assumed

to be 0.18 in the matrix and 0.08 in the fracture zone), and Sls is the maximum liquid

saturation (assumed to be 1.00). In this implementation, the capillary pressure is allowed

only to be within the range:

−Pmax ≤ Pcap ≤ 0

Pmax is assumed to be equal to 107 Pa, following the example problems in Pruess et al.

(1999).

The fracture entry pressure Pe is calculated from the fracture aperture as (Kueper and

McWhorter, 1991, eq. 2):


2σ cos θ
Pe = (D.14)
e

where the contact angle θ is assumed to be 0◦ , and the interfacial tension between

water and steam (σws ) is taken to be 0.055 N/m (Washburn, 2003). In the matrix, the

displacement pressure Pd is estimated from the rock matrix permeability (Reynolds and
APPENDIX D. CALCULATED NUMERICAL MODEL INPUT PARAMETERS 133

Kueper, 2003, eq. 2):

Pd = 10(−0.227 log(km )+0.889) (D.15)

This regression was created for an interfacial tension of 0.04 N/m. Again taking the

interfacial tension between water and steam to be 0.055 N/m, the entry pressure can be

scaled using the Leverett function (Leverett, 1941):

σ  
ws
 0.055
Pd−mat = Pd = 10(−0.227 log(km )+0.889) (D.16)
0.04 0.04

Assuming a Brooks-Corey pore size distribution index (λ) of 2.5, the Brooks-Corey p is

calculated as (Morel-Seytoux et al., 1996, eq. 8a):

2 + 3λ
p= = 3.8 (D.17)
λ

The corresponding value of m for the (van Genuchten, 1980) capillary pressure equation

(D.12) is (Morel-Seytoux et al., 1996, eq. 16a):

2
m= = 0.714 (D.18)
p−1

The value of αvg corresponding to the entry pressure calculated in (D.14) is given by

(Morel-Seytoux et al., 1996, eq. 18):

55.6 + 7.4p + p2
 
1 Pe [2p (p − 1)]
= (D.19)
αvg ρg p+3 147.8 + 8.1p + 0.092p2

A summary of the capillary pressure parameters is given in Table D.1.


APPENDIX D. CALCULATED NUMERICAL MODEL INPUT PARAMETERS 134

Table D.1: Capillary pressure parameters used in TOUGH2 simulation


TOUGH2 Parameter Symbol Value in Fracture Zone Value in Matrix
ICP - 7 7
CP(1) m computed from (D.18) computed from (D.18)
CP(2) Slr 0.18 (assumed) 0.08 (assumed)
CP(3) 1/P0 = ρw g/αvg computed from (D.19) computed from (D.19)
CP(4) Pmax 107 Pa (assumed) 107 Pa (assumed)
CP(5) Sls 1.00 (assumed) 1.00 (assumed)

D.3 Relative Permeability

In both the rock matrix and the fractures, the relative permeability-saturation relationship

of van Genuchten (1980) is used. In TOUGH2, the setting IRP = 7 implements this

relationship (Pruess et al., 1999). The liquid relative permeability is calculated as:

h  m i 2
Se 1 − 1 − Se1/m

 if Sl < Sls
krl = (D.20)


 1 if Sl ≥ Sls

where Se is defined by equation (D.13). Assuming the residual gas phase saturation (Sgr )

to be zero, the gas relative permeability is calculated as:

krg = 1 − krl (D.21)

A summary of the relative permeability parameters used is given in Table D.2.

Table D.2: Relative permeability parameters used in TOUGH2 simulations


TOUGH2 Parameter Symbol Value in Fracture Zone Value in Matrix
IRP - 7 7
RP(1) m computed from (D.18) computed from (D.18)
RP(2) Slr 0.20 (assumed) 0.10 (assumed)
RP(3) Sls 1.00 (assumed) 1.00 (assumed)
RP(4) Sgr 0.00 (assumed) 0.00 (assumed)

También podría gustarte