geometric and hydrodynamic modelling of fluid-structural

203
Geometric and hydrodynamic modelling of fluid-structural relationships in the southeastern Athabasca Basin, Saskatchewan, Canada, with implications for uranium ore genesis A Thesis Submitted to the Faculty of Graduate Studies and Research In Partial Fulfillment of the Requirements For the Degree of Doctor of Philosophy in Geology University of Regina by Zenghua Li Regina, Saskatchewan March, 2016 © 2016: Z. Li

Upload: others

Post on 02-Jun-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Geometric and hydrodynamic modelling of fluid-structural

Geometric and hydrodynamic modelling of fluid-structural relationships in the

southeastern Athabasca Basin, Saskatchewan, Canada, with implications for

uranium ore genesis

A Thesis

Submitted to the Faculty of Graduate Studies and Research

In Partial Fulfillment of the Requirements

For the Degree of

Doctor of Philosophy

in

Geology

University of Regina

by

Zenghua Li

Regina, Saskatchewan

March, 2016

© 2016: Z. Li

Page 2: Geometric and hydrodynamic modelling of fluid-structural

UNIVERSITY OF REGINA

FACULTY OF GRADUATE STUDIES AND RESEARCH

SUPERVISORY AND EXAMINING COMMITTEE

Zenghua Li, candidate for the degree of Doctor of Philosophy in Geology, has presented a thesis titled, Geometric and hydrodynamic modelling of fluid-structural relationships in the southeastern Athabasca Basin, Saskatchewan, Canada, with implications for uranium ore genesis, in an oral examination held on March 1, 2016. The following committee members have found the thesis acceptable in form and content, and that the candidate demonstrated satisfactory knowledge of the subject material. External Examiner: Dr. Jianwen Yang, University of Windsor

Co-Supervisor: Dr. Kathryn Bethune, Department of Geology

Co-Supervisor Dr. Guoxiang Chi, Department of Geology

Committee Member: Dr. Kenneth Ashton, Adjunct, Geology

Committee Member: Dr. Ian Coulson, Department of Geology

Committee Member: Dr. Tsilavo Raharimahefa, Department of Geology

Committee Member: Dr. Yee-Chung Jin, Faculty of Engineering & Applied Science

Chair of Defense: Dr. Taehan Bae, Department of Mathematics & Statistics

Page 3: Geometric and hydrodynamic modelling of fluid-structural

i

Abstract

The high-grade unconformity-related uranium deposits in the Athabasca Basin in

northern Saskatchewan, Canada, generally show close spatial relationships with

reactivated basement faults. However, the nature of this relationship has not been well

understood. This thesis aims to address this problem through two interrelated components

of research. The first phase of the research involved 3-dimensional (3D) geometric

modelling of the unconformity surface in the southeastern Athabasca Basin, in

conjunction with basement structures interpreted from regional-scale geophysical data.

The second phase was accomplished through 2-dimensional (2D) numerical modelling of

fluid flow involving basement faults, with thermal convection and faulting-related

deformation as the driving forces, respectively.

The detailed 3D model of the sub-Athabasca unconformity surface, constructed

from drill hole data with GoCAD, reveals numerous dominantly NE-trending ridges and

valleys in the unconformity surface. The most prominent ridge, up to 320 m high and

extending northeast semi-continuously from the Key Lake deposit in the south to

McArthur River deposit in the north, is interpreted to result from displacement along a

major basement fault zone. It coincides with a regional, northeast-trending alteration

corridor that hosts several large uranium deposits and a number of prospects.

Numerical modelling of free thermal convection involving basement faults using

FLAC3D indicates that the location, spacing, orientation and thermal conductivities of

basement faults exert influence on the size and location of the convection cells. For

models with an isolated basement fault zone, an up-flow centre coincides with the fault

zone and the dip angle of the fault zone does not affect the fluid flow pattern. In the case

Page 4: Geometric and hydrodynamic modelling of fluid-structural

ii

of two fault zones, however, the upwelling plume between two convection cells may

either coincide with each fault zone or be located between the two fault zones, depending

on fault spacing. When the permeability of the basement is less than two orders of

magnitude lower than that of the overlying sandstone, the basement faults can act as fluid

conduits of either ingress or egress flow, depending on their thermal conductivities and

relative locations in the models.

Numerical modelling of fluid flow in relation to deformation in a compressional

stress regime indicates that the fluid flow pattern is sensitive to the degree of bulk

shortening. At a low bulk shortening stage, fluid is driven up along the fault zone into the

sandstone in the basin, whereas at a relatively high degree of bulk shortening, fluid tends

to flow down into the fault zone and the basement. The results demonstrate that varying

the dip angle and pre-existing offset along the fault has a negligible effect on the strain

distribution and fluid flow patterns. These modelling results imply that both sandstone-

hosted and basement-hosted orebodies may be generated at different stages of

deformation within the same fault system under a unified compressional stress regime.

The results of this research confirm the importance of pre-existing basement

faults in the formation of unconformity-related uranium deposits in the Athabasca Basin.

The fluid circulation in the basin and basement may be characterized by the dominance of

thermally-driven fluid convection during periods of tectonic quiescence, and by

deformation-driven fluid flow during relatively active periods, perhaps related to far-field

tectonic events.

Keywords: 3D modelling, numerical modelling, fault interpretation, thermal convection,

fault reactivation, fluid flow, unconformity, uranium deposit, Athabasca Basin

Page 5: Geometric and hydrodynamic modelling of fluid-structural

iii

Acknowledgements

I would like to extend my sincere gratitude to my supervisors, Dr. Guoxiang Chi

and Dr. Kathryn Bethune, for providing guidance, funding and supervision for this thesis.

This research was financially supported partly by the TGI-4 (Targeted Geoscience

Initiative Phase 4) uranium ore systems project from Natural Resources of Canada, and

partly by Cameco Corp., through grants to Drs. Guoxiang Chi and Kathryn Bethune. It

was also funded by a Saskatchewan Innovation and Opportunity Graduate Scholarship, a

Faculty of Graduate Studies and Research Graduate Scholarship, and a Graduate Student

Travel Award from the University of Regina. Access to the GoCAD licenses was kindly

provided by the Saskatchewan Geological Survey.

I also extend special thanks to Colin Card and Sean Bosman from Saskatchewan

Geological Survey, for providing the drill hole data for building the 3D model and the

geophysical data used for structural interpretation; they also assisted me with using

GoCAD and provided constructive advice in academic writing. I would also like to thank

Dr. Kenneth Ashton (Saskatchewan Geological Survey) and Dr. Tsilavo Raharimahefa

(University of Regina) for valuable comments and discussions during seminars and thesis

writing. I acknowledge Dr. Yee-Chung Jin, Dr. Ian Coulson and Dr. Osman Salad Hersi

from my supervisory committee for their meaningful and constructive advice on thesis

writing and publishing.

Page 6: Geometric and hydrodynamic modelling of fluid-structural

iv

Post-defense acknowledgements

I am very grateful to my external reviewer Dr. Jianwen Yang (University of

Windsor), for his thorough examination of my thesis and for his many constructive

suggestions and comments.

Page 7: Geometric and hydrodynamic modelling of fluid-structural

v

Preface

This thesis contains chapters modified from articles published in, submitted to,

and to be submitted to peer-reviewed journals. These articles are first-authored by the

author of this thesis and co-authored by people who have made various contributions in

the research and writing of the articles. The primary contributions, including key ideas,

experimental designs, and data collection and interpretation, however, were mainly from

the author of this thesis. The details of contributions from co-authors are as follows:

Chapter 2 of this thesis is modified from the paper “Li, Z., Bethune, K.M., Chi, G.,

Bosman, S.A., Card, C.D., 2015. Topographic features of the sub-Athabasca Group

unconformity surface in the southeastern Athabasca Basin and their relationship to

uranium ore deposits. Canadian Journal of Earth Sciences 52 (10), 903-920.

http://dx.doi.org/10.1139/cjes-2015-0048”. I was responsible for the design of the models,

the modelling, and the interpretation of the modelling results, as well as the writing of the

manuscript. Bethune and Chi, as co-supervisors of the thesis, participated in the design of

the project, interpretation of the results, and editing of the paper. Card and Bosman

provided some of the data for modelling and were involved with software use and data

interpretation. The material has been reproduced with permission from NRC Research

Press.

Chapter 3 is modified from a paper submitted to the journal Geofluids: “Li, Z.,

Chi, G., and Bethune, K.M., 2015. The effects of basement faults on thermal convection

and implications for the formation of unconformity-related uranium deposits in the

Athabasca Basin”. I was responsible for the design of the models, the modelling, and

interpretation of the results, as well as the writing of the manuscript. Chi and Bethune, as

Page 8: Geometric and hydrodynamic modelling of fluid-structural

vi

co-supervisors of the thesis, participated in the design of the models, interpretation of the

results, and editing of the paper.

Chapter 4 is modified from a paper to be submitted to the journal Tectonophysics:

“Li, Z., Chi, G., Bethune, K.M., Thomas, D., Zaluski, G. and Kotzer, T., 2016. Structural

controls on fluid flow during compressional reactivation of basement faults: insights from

numerical modelling for the formation of unconformity-related uranium deposits in the

Athabasca Basin, Canada”. I was responsible for the design of the models, the modelling,

and interpretation of the results, as well as the writing of the manuscript. Chi and Bethune,

as co-supervisors of the thesis, participated in the design of the models, interpretation of

the results, and editing of the paper. Thomas, Zaluski, and Kotzer provided some of the

parameters related to rock properties of the Athabasca Basin and its basement used for

the modelling, and participated in the design of the models as well as the interpretation of

the results.

Page 9: Geometric and hydrodynamic modelling of fluid-structural

vii

Table of Contents

Abstract ............................................................................................................................... i

Acknowledgements .......................................................................................................... iii

Post-defense acknowledgements ..................................................................................... iv

Preface ................................................................................................................................ v

Table of Contents ............................................................................................................ vii

List of Tables ..................................................................................................................... x

List of Figures ................................................................................................................... xi

Chapter 1 Introduction..................................................................................................... 1

1.1 Research background ................................................................................................ 1

1.2 Research objectives ................................................................................................... 5

1.3 Previous work and literature review .......................................................................... 6

1.3.1 Three-dimensional (3D) geological modelling of the Athabasca Basin............. 6

1.3.2 Numerical modelling of fluid flow in the Athabasca Basin ............................... 7

1.4 Organization of the thesis ........................................................................................ 12

Chapter 2 Topographic features of the sub-Athabasca Group unconformity surface

in the southeastern Athabasca Basin............................................................................. 14

2.1 Introduction ............................................................................................................. 14

2.2 Geological framework ............................................................................................. 19

2.2.1 Pre-Athabasca basement rocks ......................................................................... 19

2.2.2 The Athabasca Group ....................................................................................... 21

2.2.3 Younger intrusive rocks.................................................................................... 22

2.2.4 Reactivated faults, alteration and uranium mineralization ............................... 24

2.3 Current study ........................................................................................................... 25

2.3.1 Structural (fault) analysis.................................................................................. 27

2.3.2 Three-dimensional (3D) modelling .................................................................. 30

2.4 Interpretation and discussion ................................................................................... 42

2.4.1 Evaluating the factors at play ........................................................................... 42

2.4.2 Conceptual model and relevance to uranium exploration/ore systems ............ 49

2.5 Conclusions ............................................................................................................. 53

2.6 Acknowledgements ................................................................................................. 56

Chapter 3 The effects of basement faults on thermal convection ............................... 57

Page 10: Geometric and hydrodynamic modelling of fluid-structural

viii

3.1 Introduction ............................................................................................................. 57

3.2 Geological background ........................................................................................... 61

3.2.1 Crystalline basement, the Athabasca Group and younger intrusive rocks ....... 61

3.2.2 Uranium deposits and structures....................................................................... 65

3.3 Study methods and inputs ....................................................................................... 67

3.3.1 Numerical modelling background .................................................................... 67

3.3.2 Physical models ................................................................................................ 70

3.3.3 Initial and boundary conditions ........................................................................ 78

3.4 Modelling results ..................................................................................................... 78

3.4.1 Scenario 1 – convection with no fault (base model) ........................................ 78

3.4.2 Scenario 2 – pure conduction with various conductivities for the fault ........... 81

3.4.3 Scenario 3 – convection with one vertical fault ............................................... 84

3.4.4 Scenario 4 – convection with one inclined fault ............................................... 87

3.4.5 Scenario 5 – convection with two vertical faults .............................................. 87

3.4.6 Scenario 6 – convection with two inclined faults ............................................. 94

3.4.7 Scenario 7 – convection with two vertical faults with different thermal

conductivities ............................................................................................................. 97

3.4.8 Scenario 8 – convection with permeable upper basement and vertical faults of

equal thermal conductivities ...................................................................................... 98

3.4.9 Scenario 9 – convection with permeable upper basement and vertical faults of

different thermal conductivities ............................................................................... 102

3.4.10 Scenario 10 – convection with elevated thermal gradients .......................... 104

3.5 Discussion ............................................................................................................. 105

3.5.1 Fluid convection related to faults ................................................................... 105

3.5.2 Implications for the formation of unconformity-related deposits .................. 107

3.6 Conclusions ........................................................................................................... 110

3.7 Acknowledgements ............................................................................................... 112

Chapter 4 Structural controls on fluid flow during compressional reactivation of

basement faults .............................................................................................................. 113

4.1 Introduction ........................................................................................................... 113

4.2 Geological setting .................................................................................................. 115

4.2.1 Regional geology ............................................................................................ 115

4.2.2 Structural controls on unconformity-related uranium deposits ...................... 117

Page 11: Geometric and hydrodynamic modelling of fluid-structural

ix

4.3 Principles, physical models and initial and boundary conditions of numerical

modelling ..................................................................................................................... 122

4.3.1 Simulation principles and methods................................................................. 122

4.3.2 Model setup .................................................................................................... 123

4.3.3 Initial and boundary conditions ...................................................................... 126

4.4 Modelling results ................................................................................................... 130

4.4.1 Effects of fault dip angle variation and degree of bulk shortening ................ 130

4.4.2 Effects of pre-existing fault offset .................................................................. 138

4.4.3 Effects of episodic compressional deformation .............................................. 140

4.5 Discussion ............................................................................................................. 140

4.5.1 Fluid flow during compressional reactivation of the basement faults ............ 140

4.5.2 Implication for uranium mineralization .......................................................... 143

4.6 Conclusions ........................................................................................................... 148

4.7 Acknowledgements ............................................................................................... 149

Chapter 5 Conclusions and recommendations for future work ............................... 150

5.1 Conclusions ........................................................................................................... 150

5.2 Future research ...................................................................................................... 153

References ...................................................................................................................... 155

Appendix ........................................................................................................................ 188

Page 12: Geometric and hydrodynamic modelling of fluid-structural

x

List of Tables

Table 3.1 Input parameters of various hydrological units used in model setups .............. 72

Table 3.2 Various scenarios conducted in this study ........................................................ 77

Table 4.1 Input parameters of various hydrological units used in model setups. ........... 129

Page 13: Geometric and hydrodynamic modelling of fluid-structural

xi

List of Figures

Figure 1.1 Simplified geological map of Saskatchewan ..................................................... 4

Figure 2.1 Precambrian geological map of Saskatchewan ............................................... 18

Figure 2.2 Geological framework of the Athabasca Basin ............................................... 23

Figure 2.3 Lithogeochemical map of the southeastern Athabasca Basin ......................... 26

Figure 2.4 Distribution of basement faults in the study area ............................................ 31

Figure 2.5 Rose diagram of 109 basement faults .............................................................. 32

Figure 2.6 Locations of drill holes used in this study ....................................................... 33

Figure 2.7 Contour maps showing the elevations of the sub-Athabasca unconformity ... 34

Figure 2.8 3D view of the sub-Athabasca unconformity .................................................. 38

Figure 2.9 Calculated slope of the unconformity surface in the study area ...................... 39

Figure 2.10 Axes of topographic features of the unconformity ........................................ 40

Figure 2.11 Three cross-sections derived from the 3D model .......................................... 41

Figure 2.12 Simplified cross-sections of the three deposits discussed in text .................. 43

Figure 2.13 Schematic conceptual evolutionary model of the sub-Athabasca

unconformity ..................................................................................................................... 54

Figure 3.1 Location and regional geologic framework of the Athabasca Basin ............... 62

Figure 3.2 Representative sections of three well-known unconformity-related uranium

deposits ............................................................................................................................. 63

Figure 3.3 Cross-sectional view of the physical model .................................................... 74

Figure 3.4 Modelling result for Scenario 1 ....................................................................... 80

Figure 3.5 Modelling results for Scenario 2 ..................................................................... 83

Figure 3.6 Modelling results for Scenario 3 ..................................................................... 86

Page 14: Geometric and hydrodynamic modelling of fluid-structural

xii

Figure 3.7 Modelling results for Scenario 4 ..................................................................... 91

Figure 3.8 Modelling results for Scenario 5 ..................................................................... 92

Figure 3.9 Expanded views of the encompassing fault zone on the left-hand side of the

models in Figures 3.8A – C .............................................................................................. 93

Figure 3.10 Modelling results for Scenario 6 ................................................................... 95

Figure 3.11 Expanded views of the corresponding fault in Figures 3.10A – C ................ 96

Figure 3.12 Modelling results for Scenario 7 ................................................................... 99

Figure 3.13 Modelling results for Scenario 8 ................................................................. 101

Figure 3.14 Scenario 9 .................................................................................................... 103

Figure 3.15 A simplistic, schematic illustration of an idealized genetic model ............. 111

Figure 4.1 Location and regional geologic framework of the Athabasca Basin ............. 118

Figure 4.2 Representative sections of four well-studied unconformity-related uranium

deposits ........................................................................................................................... 121

Figure 4.3 Cross-sectional view of the physical model .................................................. 128

Figure 4.4 Models with a fault dipping 45° .................................................................... 132

Figure 4.5 Plot of depth of the unconformity surfaces ................................................... 133

Figure 4.6 Models with a 60° dipping fault .................................................................... 135

Figure 4.7 Models with a 30° dipping fault .................................................................... 137

Figure 4.8 Modelling results showing the model geometry with a pre-existing vertical

offset ............................................................................................................................... 139

Figure 4.9 Schematic diagram illustrating an idealized genetic model .......................... 146

Page 15: Geometric and hydrodynamic modelling of fluid-structural

1

Chapter 1 Introduction

1.1 Research background

Unconformity-related uranium deposits are pods, veins, and semimassive

replacements consisting of mainly uraninite, and located close to basal unconformities

between basins and metamorphosed basement rocks (Jefferson et al., 2007). They

contribute approximately one-third of the western world’s uranium resources, and are

mainly located in Paleoproterozoic basins in Canada and Australia1. The Athabasca Basin

in northern Saskatchewan and Alberta, Canada (Figure 1.1), hosts the world's largest

known high-grade unconformity-related uranium deposits (Jefferson et al., 2007). The

average grade for all mined deposits is 1.97% U, which is about five times the average

grade of 0.44% for the Australian counterparts (Jefferson et al., 2007).

Uranium deposits in the Athabasca Basin may be hosted in the sandstone of the

Athabasca Group, in the underlying basement, or along the unconformity, and are

associated with reactivated basement faults that cut and variably displace the

unconformity, and are commonly but not exclusively graphite bearing (Jefferson et al.,

2007). A generally accepted model for the formation of unconformity-related uranium

deposits involves circulation of basin-derived, oxidizing, uranium-bearing fluids in fault

zones crosscutting the unconformity, where uraninite was precipitated through fluid-rock

interaction and/or mixing with basement-derived reducing fluids (see Jefferson et al.,

2007; Kyser and Cuney, 2009, for review of genetic models).

1 Geology of Uranium Deposits, World Nuclear Association, February 2015. http://world-

nuclear.org/info/Nuclear-Fuel-Cycle/Uranium-Resources/Geology-of-Uranium-Deposits/. Accessed

January 22, 2016.

Page 16: Geometric and hydrodynamic modelling of fluid-structural

2

The formation of the high-grade uranium deposits requires the circulation of a

large volume of fluids, and various models have been proposed for the migration of fluids

in the Athabasca Basin (Chi et al., 2013). These models include fluid flow induced by

tectonic deformation (Schaubs et al., 2005; Cui et al., 2012b), thermally-driven fluid

convection (Raffensperger and Garven, 1995; Boiron et al., 2010; Cui et al., 2012a, b), as

well as other mechanisms (see review by Chi et al., 2013). However, the exact

mechanism(s) controlling fluid flow related to mineralization in the basin remains poorly

understood. In particular, while it is generally thought that fluid flow was either

downward from the basin into basement fault zones (ingress flow), or upward into the

Athabasca Group and outward from such zones (egress flow), it is not clear how this fluid

flow is related to the fault zones (Jefferson et al., 2007).

The spatial association of the unconformity-related uranium deposits with

reactivated basement faults suggests that the fluid flow responsible for mineralization

may have been related to the deformational processes during fault reactivation, which

were perhaps related to certain far-field tectonic events (e.g., Davidson, 2008; Alexandre

et al., 2009). However, there is the question of whether ingress flow and egress flow were

associated with alternating compressional and extensional stress fields (Cui et al., 2012b),

or alternatively whether egress flow and ingress flow can both occur under a unified

compressional stress regime.

On the other hand, in an intracratonic basin like the Athabasca Basin, one would

not expect deformation to occur frequently and/or as a continuous process. The basin was

most likely tectonically quiet most of the time, as implied by the flat-lying,

unmetamorphosed sedimentary rocks within it (Ramaekers et al., 2007). In addition,

Page 17: Geometric and hydrodynamic modelling of fluid-structural

3

volcaniclastic rocks with a U-Pb zircon age of 1644 ± 13 Ma (Rainbird et al., 2007) are

present in the Wolverine Point Formation of the upper Athabasca Group, and basalt with

a U-Pb baddeleyite age of 1540 ± 30 Ma (Chamberlain et al., 2010) is developed in the

Kuungmi Formation of the geologically similar Thelon Basin, suggesting that these

intracratonic basins experienced some thermal events during their evolution. Hoeve and

Quirt (1984) suggested that fluid flow associated with uranium mineralization may have

been related to fluid convection driven by heat anomalies due to the relatively high heat

conductivities of graphitic fault zones. Raffensperger and Garven (1995a) and Cui et al.

(2012a, b) further numerically demonstrated that thermally-driven fluid convection is

possible in the Athabasca Basin; however, it remains unclear what roles basement faults

may have played on thermally-driven fluid flow during periods of tectonic quiescence.

Answering these questions is of economic significance because it will help to

determine which fault-fluid-lithology combinations are favourable for uranium

mineralization, not only in the Athabasca Basin, but also in other prospective Proterozoic

basins. The relationship between faults and fluid flow is therefore of universal concern in

the exploration for unconformity-related uranium deposits as well as other deposit types

that show a clear structural control. A general approach to address these questions is to

establish 3-dimensional (3D) geometric models integrating the most important ore-

controlling geological features, and to carry out numerical modelling of the fluid flow

responsible for mineralization (e.g., Feltrin et al., 2009; Zhang et al., 2011).

Page 18: Geometric and hydrodynamic modelling of fluid-structural

4

Figure 1.1 Simplified geological map of Saskatchewan (modified from Card et al., 2007). The rectangle

outlines the study area. Black squares represent the major uranium deposits. AB = Athabasca Basin.

Page 19: Geometric and hydrodynamic modelling of fluid-structural

5

1.2 Research objectives

A 100 km by 60 km zone within the southeastern Athabasca Basin was selected to

study the relationship between structures, fluid flow and uranium mineralization (Figure

1.1). The first stage of this study involved analysis of the surface/subsurface regional

geology and construction of a 3D model to image the fundamental lithological, structural

and alteration features. This was then followed by numerical modelling of fluid flow to

test the effects of various combinations of lithology, orientations of structures, stress

fields and thermal conditions.

The objectives of this study are to better understand the inter-relationship between

basement structures, unconformity surface topography and uranium mineralization. In

more detail, the aims are: 1) to create a 3D model to better determine the physical

relationships among these geological features; 2) to undertake numerical modelling to: i)

determine how the orientation and spacing of basement faults affect thermal convection;

and ii) examine how the basement faults with various dip angles and offsets affect fluid

flow during compressional deformation. The overall goal is to determine the main

physical and dynamic factors controlling fluid flow and the related implications for the

formation of unconformity-related uranium deposits.

These objectives are fulfilled through three components of research, i.e., 3D

geometric modelling of the unconformity surface in conjunction with basement structures,

2-dimensional (2D) modelling of fluid flow with thermal convection as the driving force,

and 2D modelling of fluid flow with deformation as the driving force.

Page 20: Geometric and hydrodynamic modelling of fluid-structural

6

1.3 Previous work and literature review

1.3.1 Three-dimensional (3D) geological modelling of the Athabasca Basin

Three-dimensional (3D) modelling is used extensively today in a range of

applications. 3D models are, for instance, used in geostatistical simulations, numerical

modelling of fluid flow and deformation, and to image geological bodies at depth (e.g.,

Rawling et al., 2006; Feltrin et al., 2008; Heise et al., 2008; Zanchi et al., 2009; Blöcher

et al., 2010; Skytta, 2012).

Traditionally, geoscience information from the Athabasca region has been

presented in a series of geologic maps and cross sections (e.g., Ramaekers et al., 2007).

However, the same data presented in the 3D environment provides a new perspective that

is sometimes more easily interpreted because visualisation of the intricacies in 2D is not

always intuitive (Bosman et al., 2011). The geometry of geologic units, their intersections

with structural features such as faults, for example, are all more easily interpreted in 3D

space. The ongoing Athabasca Basin ore-systems project (Bosman et al., 2010, 2012,

2014), launched by Saskatchewan Geological Survey (SGS), presents data in the 3D

environment using Mira Geoscience’s ParadigmTM GOCAD® Mining Suite (GOCAD) in

order to gain a better understanding of the Athabasca Basin’s uranium deposits. A best-fit

model for the major geologic features of the Athabasca Basin, including the

unconformity and the various lithostratigraphic units of the Athabasca Group, has been

established (Bosman et al., 2012, 2014).

More detailed modelling, however, coupled with structural analysis, is clearly

required to investigate subtle variations in unconformity surface elevation and its

associated controls. The model can serve as a basis for exploring the relationship between

Page 21: Geometric and hydrodynamic modelling of fluid-structural

7

various topographic features, the structures, and known uranium deposits and occurrences.

It also serves as the basis for the following coupled thermal-mechanical fluid flow

modelling (FLAC3D) to more rigorously test how thermal conditions and deformation

associated with fault zones influence fluid flow patterns and related alteration-

mineralization.

1.3.2 Numerical modelling of fluid flow in the Athabasca Basin

Various fluid flow models have been proposed for the migration of fluids in the

Athabasca Basin (Chi et al., 2013). These models include: 1) topography-driven flow

(Derome et al., 2005; Alexandre and Kyser, 2012); 2) compaction-driven (Hiatt and

Kyser, 2007) or 3) overpressure-driven flow (Chi et al., 2013, 2014); 4) flow induced by

tectonic deformation (Schaubs et al., 2005; Cui et al., 2012b); 5) buoyancy-driven flow,

including thermal convection (Raffensperger and Garven, 1995; Boiron et al., 2010; Cui

et al., 2012b) and thermohaline convection due to salinity and temperature differences

(Cui et al., 2012a). However, it remains poorly understood exactly which mechanism(s)

controls the fluid flow related to mineralization in the Athabasca Basin (Chi et al., 2013,

2014), and particularly, how ingress flow and egress flow interacted along single or

multiple fault zones (Jefferson et al., 2007).

Numerical modelling can provide considerable insights into what the controlling

factors might be in the process of fluid flow and the formation of uranium deposits, and

how these factors interact (e.g., Chi and Xue, 2011). A review of numerical modelling of

fluid flow related to thermal convection and tectonic deformation, including some carried

Page 22: Geometric and hydrodynamic modelling of fluid-structural

8

out for the Athabasca Basin (Raffensperger and Garven, 1995a, b; Schaubs et al., 2005,

Cui et al., 2012a, b), is provided below.

1.3.2.1 Thermal convection

Thermal convection has been advocated by many researchers as an important

mechanism for mass and heat transport and related mineralization in sedimentary basins

(e.g., Raffensperger and Garven, 1995; Raffensperger and Vlassopoulos, 1999; Garven et

al., 2001; Simms and Garven, 2004; Yang et al., 2004b; Large et al., 2005; Oliver et al.,

2006; Guillou-Frottier et al., 2013; Malkovsky and Pek, 2015). Some researchers have

proposed that large-scale free thermal convection may be responsible for uranium uptake

and transport in the Athabasca Basin (Hoeve and Quirt, 1984; Boiron et al., 2010).

Chi et al. (2013) carried out a numerical modelling study of the development of

fluid overpressure in the Athabasca Basin due to disequilibrium sediment compaction,

and the results suggested that no significant fluid overpressure was developed in the basin

throughout its depositional history. Fluid flow related to sediment compaction was very

slow and the temperature profile was undisturbed, implying that if compaction-driven

flow was responsible for mineralization, the sites of mineralization would not show a

thermal anomaly. The development of a near-hydrostatic pressure regime in the

Athabasca Basin may have facilitated circulation of oxidizing fluids from the shallow

part of the basin into the deeper part through thermally-driven fluid convection (Chi et al.,

2013; Chu and Chi, 2016). It is unknown, however, whether there was significant fluid

overpressure in the basement due to factors like tectonic deformation (e.g., Cui et al.,

2012b). Based on a follow-up study, Chi et al. (2014) also suggested that the hydrocarbon

Page 23: Geometric and hydrodynamic modelling of fluid-structural

9

generation processes in the Douglas Formation of the Athabasca Group contributed little

to the development of fluid overpressure. This implies that the pressure regime was

favourable for the development of free thermal convection induced by the thermal

gradient; in contrast, strong fluid overpressure would have tended to suppress such fluid

flow (Zhao et al., 2012).

Raffensperger and Garven (1995) carried out a finite element modelling of

coupled groundwater flow and heat transport at the basin scale. They examined the role

of thermally-driven free convection as a mechanism for driving regional groundwater

flow systems and uranium ore formation in a 2D model constrained by some common

features of the Athabasca (Canada) and McArthur (Australia) basins. The results

suggested that free thermal convection is possible in the Athabasca basin assuming a

reasonable thermal gradient and hydrostatic fluid pressure. Based on numerical studies,

free thermal convection was also inferred in the Thelon Basin (Cui et al., 2010),

considered the counterpart of the Athabasca Basin to the north (Hoffman, 1988), at

geothermal gradients of 25 to 35 °C/km. Furthermore, Cui et al. (2012a) conducted finite

element modelling to investigate the potential mechanisms driving fluid interaction

across a basement/cover unconformity. Using a simplified conceptual model constructed

by integrating geological features shared by the Athabasca, Thelon and Kombolgie basins,

they designed various numerical scenarios to examine buoyancy-driven fluid flow

patterns and the corresponding solute transport. The results showed that thermohaline

convection may have penetrated into the low-permeability basement for up to 1 to 2 km

below the unconformity at extremely low flow rates.

Page 24: Geometric and hydrodynamic modelling of fluid-structural

10

Most of these studies, however, were focused on the pervasive convective fluid

flow in the high-permeability basinal sequence, and fluid flow in the basement has not

been paid much attention, except for the generic basin-basement model of Cui et al.

(2012a). In addition, the study conducted by Raffensperger and Garven (1995) suggested

that the existence of conductive graphitic zones (1-5 km wide) does not affect the

regional flow patterns at basin scale (~800 km long), whereas other researchers proposed

that graphitic units give rise to thermal anomalies that may drive fluid convection (e.g.,

Hoeve and Quirt, 1984; Schaubs et al., 2002). The question is whether such graphitic

fault zones are conductive enough to create thermal anomalies and thus affect the flow

patterns on a smaller scale (compared to the whole basin).

More importantly, empirical observation suggests that there may be a systematic

spatial relationship between ore deposits and structures and none of these studies fully

examined the role of faults in thermally-driven fluid convection. In addition, some studies

were based on highly simplified conceptual basin-scale models with a highly exaggerated

vertical scale. This in turn exaggerated the horizontal size of individual convection cells

and made incorporation of faults in the models difficult. It is of interest to investigate the

roles faults may have played in fluid flow, especially during periods of tectonic

quiescence.

1.3.2.2 Tectonic deformation and fluid flow

Many researchers have proposed tectonic deformation as one of the possible

mechanisms responsible for fluid flow related to mineralization in various geologic

settings (e.g., Cox et al., 2005; Oliver et al., 2006). Fault zones have been generally

Page 25: Geometric and hydrodynamic modelling of fluid-structural

11

considered one of the most important permeable fluid pathways for hydrothermal ore-

forming systems (e.g., Cox, 2005). Numerous studies have been conducted to examine

the interaction of faults and fluid flow in various ore systems (e.g., Garven et al., 2001;

Matthäi et al., 2004; Simms and Garven, 2004; Yang et al., 2004b; Oliver et al., 2006;

Zhang et al., 2007; Feltrin et al., 2009; Schaubs and Fisher, 2009). For the Athabasca

Basin, the intimate spatial relationship between reactivated basement faults and uranium

deposits has been well illustrated at the deposit scale (e.g., Hoeve and Sibbald, 1978;

McGill et al., 1993; Harvey and Bethune, 2007; Jefferson et al., 2007; Long, 2007;

Tourigny et al., 2007; Kerr, 2010). Brittle reactivated fault zones have also been

considered to be one of the main controls on fluid flow in the vicinity of deposits

(Jefferson et al., 2007). However, to date only a few numerical studies have been

conducted to examine how the basement faults affect fluid flow during tectonic

deformation in the Athabasca Basin.

Among these, Schaubs et al. (2005) proposed that fluid can flow into a basement

fault (e.g., the McArthur River deposit) or out of the fault (e.g., the Cigar Lake deposit) in

an overall compressional stress regime, depending on the attitude of the faults. For the

unconformity-related uranium deposits in the Alligator Rivers region of the Northern

Territory (Australia), Schaubs and Fisher (2009) reported that during deformation,

shallowly dipping faults allow sandstone-derived fluid to flow deep into the basement,

whereas steeply dipping faults only allow for the interaction of sandstone and basement-

derived fluids in close proximity to the unconformity. However, the research was

published in the form of abstracts and details regarding the interactivity between

structures and fluid flow are unclear.

Page 26: Geometric and hydrodynamic modelling of fluid-structural

12

The numerical studies by Cui et al. (2012b) involving coupled thermo-hydro-

mechanical modelling suggested that during periods of tectonic quiescence thermally-

driven free convection may develop in the basin sequence, whereas deformation

suppressed free convection and led to deformation-dominated fluid flow or mixed

convection, depending on strain rates. Moreover, it was found that during compressive

deformation, fluids in the basement flow upwards along the fault zone into the sandstone,

whereas during extensional deformation fluids in the sandstone flow downwards into the

basement. In other words, ingress flow and egress flow were found to be associated with

alternating compressional and extensional stress fields.

However, considering the fact that most of the ore-bounding structures

demonstrate obvious reverse displacements, it is reasonable to speculate that the stress

responsible for the fluid flow related to mineralization was likely in a (trans-)

compressional regime, although the possibility of intermittent extensional stress regimes

during the mineralization periods cannot be excluded. It is therefore interesting to

investigate how various faults with different dip angles and offsets interact with fluid

flow during compressional deformation. In particular, the question of whether or not both

ingress and egress flow can be generated in a compressional environment needs to be

answered.

1.4 Organization of the thesis

This thesis is composed of five chapters. Chapter 1 introduces the research

background, scientific problems to be addressed, and the objectives of this research.

Chapter 2 mainly deals with the topographic features of the unconformity surface in the

Page 27: Geometric and hydrodynamic modelling of fluid-structural

13

southeastern Athabasca Basin and their relationships with structures based on 3D

geomodelling. This work has been published in the Canadian Journal of Earth Sciences.

Chapter 3 uses 2D numerical modelling to examine how basement faults affect the

thermal convection and the implications for the formation of the unconformity-related

uranium deposits. This work has been submitted to Geofluids and is currently in review.

Chapter 4 discusses how basement faults with various geometries (i.e., different dip

angles and pre-existing offsets) affect fluid flow during compressional deformation based

on 2D numerical modelling. The work will be submitted to Tectonophysics. Chapter 5

summarizes the conclusions of this study and provides recommendations for future study.

Page 28: Geometric and hydrodynamic modelling of fluid-structural

14

Chapter 2 Topographic features of the sub-Athabasca Group unconformity surface

in the southeastern Athabasca Basin2

2.1 Introduction

Unconformity-related uranium deposits, best developed in Canada and Australia,

are typically situated close to the unconformities between Archean – Palaeoproterozoic

metamorphic and granitic basement rocks and Late Paleo- to Mesoproterozoic sandstone

basins (Jefferson et al., 2007). In the Athabasca Basin in northern Saskatchewan and

Alberta, unconformity-related uranium ore deposits are spatially and genetically related

to reactivated basement faults that cross-cut the sub-Athabasca unconformity (Jefferson

et al., 2007). Topographic features on the sub-Athabasca Group unconformity surface,

such as paleovalleys, topographic highs and fault scarps, have been documented in some

previous studies, and have been linked to localization of the mineral systems in the basin

(Jefferson et al., 2007). Wallis et al. (1985) noticed that paleovalleys are spatially

associated with the McClean Lake uranium deposits, and Macdonald (1980) elaborated

on how protoliths may influence weathering and hence paleotopography, noting that

metapelite and metasemi-pelite units tend to develop deeper regoliths. Harvey and

Bethune (2007; cf. Harvey, 2004) documented a paleovalley in the Deilmann pit at Key

Lake based on sedimentological, isopach, and stratigraphic analysis, as reflected by

onlapping of the basal strata of the Manitou Falls Formation over the unconformity

surface; however, they also recognized that this topography had been accentuated by

2 This chapter has been published in Canadian Journal of Earth Sciences, "Li, Z., Bethune, K.M., Chi, G.,

Bosman, S.A., Card, C.D., 2015. Topographic features of the sub-Athabasca Group unconformity surface

in the southeastern Athabasca Basin and their relationship to uranium ore deposits. Canadian Journal of

Earth Sciences 52 (10), 903-920. http://dx.doi.org/10.1139/cjes-2015-0048”. The material has been

reproduced with permission from NRC Research Press.

Page 29: Geometric and hydrodynamic modelling of fluid-structural

15

post-Athabasca faulting. In the case of the Sue C pit at McClean Lake, Long (2001, 2007)

similarly suggested that initial deposition of the Athabasca Group was strongly controlled

by basement paleotopography with early deposits confined between distinct valley walls,

although this primary relief was also recognized to have been modified by subsequent

faulting (Tourigny et al., 2007). Similar observations of primary relief on the

unconformity were made by Ramaekers et al. (2007) in the Key Lake and Sue C mine

areas, where local N- to NNE-trending paleovalleys were marked by paleotalus and

conglomerate wedges at the bases of the MFb member of the Manitou Falls and Read

formations. Substantial topographic highs associated with basement units (so-called

‘quartzite ridges’ of Earle and Sopuck (1989) and Marlatt et al. (1992) have also been

documented in the McArthur River deposit, and in the Phoenix deposit in the Wheeler

River area (Marlatt et al., 1992; Györfi et al., 2007; Jefferson et al., 2007; Ramaekers et

al., 2007; Yeo and Delaney, 2007; Kerr, 2010). Two prominent quartzite ridges occurring

in the McArthur River area were interpreted as compressional pop-up structures bounded

by outwardly-dipping faults (Jefferson et al., 2007), based on seismic reflection data and

geological study (Györfi et al., 2007). In general, the quartzite ridges are bounded by

weaker graphitic pelitic units, with post-Athabasca faults localized in the pelitic rocks

proximal to, or right along, the contact (Kerr, 2010). This ‘quartzite ridge’ concept

(coupled with spatially associated conductors coinciding with adjacent graphite-rich units)

has proven to be a useful model in uranium exploration in the eastern Athabasca Basin

and contributed to the discovery of the Phoenix deposit (Kerr, 2010).

While these relationships have been known for some time, the mechanisms by

which such topographic features were generated, their size and distribution at the regional

Page 30: Geometric and hydrodynamic modelling of fluid-structural

16

scale, as well as their relationship to mineralization, are still poorly understood. Three-

dimensional (3D) modelling represents a highly useful technique to determine the spatial

configuration and attributes of such topographic features. In the past decade, such

modelling has been routinely employed in sedimentary basin analysis, providing insights

into petroleum systems (e.g., Batycky et al., 1997; Blöcher et al., 2010), and has also

been used to gain understanding of mineral deposit systems (e.g., Rawling et al., 2006;

Feltrin et al., 2008; De Veslud et al., 2009; Skytta, 2012). Using public domain data,

Bosman et al. (2010, 2012, 2014) produced a basin-wide 3D model of the major geologic

features of the Athabasca Basin, including the sub-Athabasca unconformity and

lithostratigraphic units of the Athabasca Group. However, more detailed modelling,

coupled with structural analysis, is clearly required to investigate subtler variations in

unconformity surface elevation and its associated controls. Accordingly, in this study, we

selected a 100 km by 60 km area of the southeastern Athabasca Basin (Figure 2.1) to

examine relationships between: 1) topographic features of the unconformity; 2)

geological features, in particular all major fault systems; and 3) uranium mineralization.

To optimize geological and structural information, the rectangular area was oriented with

the long dimension parallel to regional structural trends (Figure 2.1). This area straddles

the Wollaston-Mudjatik domain boundary and encompasses the well-studied Key Lake

(e.g., Carl et al., 1988; Long, 2001; Harvey, 2004; Harvey and Bethune, 2007) and

McArthur River (e.g., Marlatt et al., 1992; McGill et al., 1993; Györfi et al., 2007; Hajnal

et al., 2007; White et al., 2007; Adlakha et al., 2013) deposits, as well as the intervening

Russell Lake area, where several new deposits have been discovered more recently,

including Millennium (Robertson, 2006) and Phoenix (Kerr, 2010).

Page 31: Geometric and hydrodynamic modelling of fluid-structural

17

In this paper, which builds upon initial work published in a Geological Survey of

Canada Open File report (Li et al., 2013), we present a detailed 3D model of the

unconformity surface in the southeastern Athabasca Basin using the most up to date

publicly available geological and drill-hole data. This model and associated structural

analysis then serve as a basis for exploring the relationship between various topographic

features, the geology, and known uranium deposits and occurrences. In particular, we will

address questions such as: to what extent are these topographic features primary,

reflecting primary changes in relief related to pre-Athabasca Group processes (e.g.,

basement structure and differential erosion), or an effect of syn- to post-Athabasca Group

faulting and differential uplift? In addition, what factors controlled the development of

the topography and ultimately controlled the mineralization process? In posing these

questions, we will present and discuss our own data, as well as referring to specific, well-

known deposit areas in order to more fully evaluate relationships between the processes

involved.

Page 32: Geometric and hydrodynamic modelling of fluid-structural

18

Figure 2.1 Precambrian geological map of Saskatchewan (after Card, 2007a with permission from the

editors). The rectangle outlines the study area. Black squares represent major uranium deposits. AB =

Athabasca Basin.

Page 33: Geometric and hydrodynamic modelling of fluid-structural

19

2.2 Geological framework

The study area encompasses a 6000 km2 area of the southeastern Athabasca Basin

in Saskatchewan. The Late Paleoproterozoic (<1.75 Ga), dominantly siliciclastic rocks of

this basin unconformably overlie older Precambrian rocks of the Wollaston and Mudjatik

domains of the Hearne Province, part of the larger Churchill Province of the Canadian

Shield (Figure 2.1) (Lewry and Sibbald, 1980; Hoffman, 1988). To the west, the Hearne

Province is separated from the Rae Province by the Snowbird tectonic zone, whereas to

the east, it is bounded by the Reindeer Zone of the Trans-Hudson orogen (Figure 2.1)

(Card et al., 2007a, b).

2.2.1 Pre-Athabasca basement rocks

The oldest rocks in the Hearne Province are Archean granitoid gneisses and

granites. The U-Pb zircon ages of these rocks are locally as old as ca. 2.9-2.8 Ga (Orrell

et al., 1999), but are mainly in the range 2.58-2.56 Ga (Annesley et al., 1997). The

Archean granitoid gneisses are unconformably overlain by Early Paleoproterozoic (< 2.1

Ga; Ansdell et al., 2000) supracrustal rocks of the Wollaston Supergroup (Yeo and

Delaney, 2007). Broadly speaking, the Wollaston Supergroup contains a record of rift to

passive margin, through to foreland basin sedimentation, thought to reflect a complete

Wilson cycle (Yeo and Delaney, 2007).

Starting at ca. 1.9 Ga, the Archean basement rocks and overlying Wollaston

Supergroup were intensely deformed during the Trans-Hudson orogeny (Lewry and

Sibbald, 1980; Portella and Annesley, 2000; Tran, 2001). The initial stages of

deformation (D1/2) involved, WNW-directed folding and thrusting of cover rocks over

Page 34: Geometric and hydrodynamic modelling of fluid-structural

20

basement, with formation of a strong ductile transposition foliation, and were superseded

by development of close to tight, NE-trending, upright folds (D3) (Tran, 2001). The

intensity of D3 folding was highest in the Wollaston domain, whose boundary with the

Mudjatik domain to the west is marked by a transition to more open D3 structures,

overprinted by a younger orthogonal fold set (D4) (Tran, 2001), giving rise to a

characteristic dome-and-basin fold interference pattern (Lewry and Sibbald, 1977). The

Mudjatik-Wollaston domain boundary is considered by most researchers to be

transitional with regard to observed lithological units, with an increase in the proportion

of Wollaston Supergroup metasedimentary rocks toward the east (e.g., Annesley et al.,

2005; Tran, 2001; Yeo and Delaney, 2007).

In the study area, the quartzose rocks of the Wollaston Supergroup have

traditionally been interpreted to represent orthoquartzites of sedimentary derivation

(Annesley et al., 2005). However, recent subsurface study by Card (2012, 2013) indicates

that the quartz-rich rocks (that typically give rise to prominent ridges) are at least in part

the result of intense silicification and related pegmatite emplacement, suggesting an in

part hydrothermal, rather than solely metamorphic, origin for these rocks. In addition, the

fact that coarse-grained graphite and pyrite are commonly observed to crosscut regional

metamorphic fabrics, along with the presence of macroscopic, cm-scale graphite veins,

led Card (2012, 2013) to propose that the copious graphite in pelitic gneisses was

precipitated from carbon-bearing hydrothermal fluids. These assertions are supported by

recent detailed work in the Phoenix deposit area by Wang et al. (2015), where quartzite

and adjacent pelitic gneisses are invaded by numerous metre-scale, lit-par-lit sheets of

massive quartz–pegmatite, locally containing interstitial, coarse-grained graphite.

Page 35: Geometric and hydrodynamic modelling of fluid-structural

21

2.2.2 The Athabasca Group

The sub-Athabasca unconformity is characterized by an intense zone of

paleoweathering that penetrates the basement to depths of as much as 50 m (Hoeve and

Sibbald, 1978). The overlying basin is filled predominantly of quartz arenite and quartz-

rich sandstone with subordinate conglomerate and siltstone ranging from Statherian to

Calymmian in age (Ramaekers et al., 2007; Gradstein et al., 2012; Bosman and

Ramaekers, 2013). The preserved sedimentary record includes the following formations,

in order from oldest to youngest: the Fair Point, Smart, Read, Manitou Falls, Lazenby

Lake, Wolverine Point, Locker Lake, Otherside, Douglas, and Carswell formations

(Figure 2.2) (Ramaekers et al., 2007). Recent stratigraphic analysis indicates that the

larger basin actually comprises a series of smaller, stacked basins of different

configuration and age; the Jackfish, Cree and Mirror basins (Ramaekers et al., 2007), the

easternmost Cree Basin being most relevant to this study. The latter is dominated by

fluviatile sandstone and conglomerate of the Read and Manitou Falls formations,

representing a preserved aggregate thickness of up to 1100 m (Ramaekers et al., 2007). A

maximum depositional age of ca. 1803 Ma has been determined for the Read Formation

based on U-Pb zircon ages of pegmatites from the Wollaston domain (Annesley et al.,

1997). A somewhat younger maximum age of ca. 1750 Ma is implied by 207Pb/206Pb and

U–Pb rutile ages from metapelitic rocks in the Mudjatik domain (Orrell et al., 1999).

40Ar/39Ar dating of muscovite in metamorphic basement rocks also indicates that

significant post-Hudsonian metamorphic cooling had taken place by 1750 Ma, before or

near the onset of basin development (Alexandre et al., 2009). U-Pb dating of detrital

zircons of the Athabasca Group by Rainbird et al. (2007) provides broadly similar age

Page 36: Geometric and hydrodynamic modelling of fluid-structural

22

constraints on basin development. The Athabasca Group reaches a maximum thickness of

about 1500 m in the central part of the basin, although most researchers concede that a

considerable portion of the original strata has been eroded (Ramaekers et al., 2007). The

estimated thickness of the basin by the Calymmian was five to six km on the basis of the

paleogeothermal gradients deduced from fluid inclusion studies (Pagel et al., 1980).

There are nevertheless significant uncertainties regarding the thickness of erosion due to

ambiguity of the fluid inclusion data and their interpretations (Chi et al., 2015).

2.2.3 Younger intrusive rocks

Rocks of the Athabasca Group and its basement are cut by NW-trending mafic

dykes of the McKenzie swarm (Cumming and Krstic, 1992). These sub-vertical dykes

range in width from 1 m to several hundred metres and have been dated at 1267 ± 2 Ma

(LeCheminant and Heaman, 1989). Another suite of olivine diabasic and gabbroic rocks,

belonging to the 1.11 Ga (MacDougall and Heaman, 2002) Moore Lakes Complex

(Figures 2.2 and 2.3), intruded the eastern Hearne craton basement and the Athabasca

Group (MacDougall and William, 1993) along the southeastern perimeter of the basin,

forming a complex of sill-like intrusions.

Page 37: Geometric and hydrodynamic modelling of fluid-structural

23

Figure 2.2 Geological framework of the Athabasca Basin (after Ramaekers et al., 2007 and Geological

Atlas of Saskatchewan), with location of the 60 km wide, 100 km long and 1.7 km deep volume selected

for detailed 3D modelling in this study in the southeast corner of the basin. Dd = diabase, C = Carswell, D

= Douglas, O = Otherside (a = Archibald, b = Birkbeck), LL = Locker Lake (m = upper pebbly, b =

conglomerate, s = lower pebbly), W = Wolverine Point, LZ = Lazenby Lake (h = Hodge, c = Clampitt, s =

shiels, l = Larter, d = Dowler), MF = Manitou Falls (members: b = Bird (l = lower, u = upper) c = Collins

(p = pebbly), d = Dunlop (p = pebbly), r = Raibl (up =upper pebbly, cr = clay-intraclast-rich, ps = pebbly),

w = Warnes (up = upper pebbly, cr = clay intraclast rich, s = quartz arenite, lp = lower pebbly)), S = Smart,

RD = Read, FP = Fair Point. Original in colour.

Page 38: Geometric and hydrodynamic modelling of fluid-structural

24

2.2.4 Reactivated faults, alteration and uranium mineralization

The Athabasca Basin is cut by numerous late brittle faults that demonstrably

offset the unconformity and overlying sandstone and commonly represent reactivation of

older, ductile basement structures (e.g., Card et al., 2007a, b; Hoeve and Sibbald, 1978;

Hoeve et al., 1980; Jefferson et al., 2007; McGill et al., 1993; Tourigny et al., 2007). The

faults are generally rooted in sheared graphitic basement rocks and have offsets on the

order of tens to hundreds of metres. The largest known vertical displacement (200 m) of

the unconformity has been identified along the Dufferin Lake fault, which represents a

reactivation of earlier ductile structures of the Snowbird tectonic zone (Figure 2.1) (Card

et al., 2007b). The linkage between such reactivated basement faults, graphitic pelitic

units and uranium deposits is well established (e.g., Hoeve and Sibbald, 1978; Hoeve et

al., 1980; McGill et al., 1993; Baudemont and Pacquet, 1996; Bernier, 2004; Harvey and

Bethune, 2007; Long, 2007; Tourigny et al., 2007; Yeo and Delaney, 2007). It has also

served to inform exploration programs, in terms of targeting electromagnetic (EM)

conductors that are proxies for faults rooted in graphitic pelitic units (Jefferson et al.,

2007). There is also a strong spatial association between faults and clay mineral alteration

(e.g., Hoeve and Quirt, 1984). In particular, early work by Earle and Sopuck (1989)

established that district-scale clay alteration features are superimposed on the regional

diagenetic clays (primarily dickite). As shown in Figure 2.3, they defined a relatively

linear alteration corridor, up to 30 km wide, that is dominated by illite and extends for

more than 100 km northeastward from Key Lake to Cigar Lake. The illite alteration is

accompanied by subparallel linear zones of anomalous chlorite and dravite (Earle and

Sopuck, 1989), and altogether, this alteration system closely adheres to an anastomosing

Page 39: Geometric and hydrodynamic modelling of fluid-structural

25

system of NE-trending faults, as inferred from the distribution of ground EM conductors.

This major structural-alteration corridor also encompasses the most significant uranium

deposits and prospects in this part of the basin (Figure 2.3) (Jefferson et al., 2007).

2.3 Current study

As mentioned in the introduction to this paper, previous work indicates that

variations in the unconformity surface elevation may be primary (i.e., paleotopographic

relief) caused by pre- and syn-depositional faulting, or secondary, i.e., these variations

were accentuated and/or new ones were developed by post-depositional faulting. This

study reinforces those conceptual advances while focusing on secondary fault

reactivations and, for the first time in the public domain, places such fault reactivations

and related alteration into a regional 3D geological framework with respect to uranium

deposits and prospects.

Page 40: Geometric and hydrodynamic modelling of fluid-structural

26

Figure 2.3 Lithogeochemical map of the southeastern Athabasca Basin, showing regional illite, chlorite,

and dravite anomalies in the surficial material and outcrops of the Athabasca Group (after Earle and

Sopuck, 1989). Ground electromagnetic (EM) conductors and uranium deposits, occurrences and prospects

are superimposed on the map, data from Geological Atlas of Saskatchewan. dd = diabase, for other symbols

see Figure 2.2. Original in colour.

Page 41: Geometric and hydrodynamic modelling of fluid-structural

27

2.3.1 Structural (fault) analysis

In order to explore these relationships more fully and interpret the 3D

configuration of the unconformity surface in a more complete context, a regional-scale

analysis of faults systems within the study area was considered essential. The Athabasca

Group is essentially non-magnetic (Thomas and McHardy, 2007), therefore airborne

magnetic surveys provide the most effective method of mapping basement geology and

structure in a continuous fashion (e.g., Pilkington, 1989; Pilkington and Roest, 1992;

Madore et al., 2000; Portella and Annesley, 2000; Thomas and McHardy, 2007). This,

along with the fact that pre- and post-Athabasca faults commonly coincide, means that

geophysical analysis provides a first-order means of identifying post-Athabasca faults

(Portella and Annesley, 2000; Annesley et al., 2005). This study utilized Geotiffs of

aeromagnetic data from Saskatchewan Geological Survey Open File 2010-46 (Card et al.,

2010). This dataset includes maps of: 1) reduced to pole Total Magnetic Intensity (TMI);

2) First Vertical Derivative (1VD); 3) Automatic Gain Control; and 4) Tilt Derivative and

Horizontal Tilt Derivative, from the region including and surrounding the Athabasca

Basin. In addition, using Geosoft Oasis Montaj software, the Saskatchewan Geological

Survey clipped TMI and 1VD data from two recent aeromagnetic surveys (Buckle et al.,

2009, 2010) to the boundaries of NTS-74H in order to bring out the detail of magnetic

patterns local to the study area.

All these datasets were incorporated into an ArcGIS environment for subsequent

fault analysis. The regional Total Magnetic Intensity (TMI) map was selected to show the

results of fault interpretation (Figure 2.4); the outline of the study area and positions of

major lithostructural domain boundaries below Athabasca Group cover were also

Page 42: Geometric and hydrodynamic modelling of fluid-structural

28

projected onto this map. It is notable that, at the regional scale, the transitional zone

between Mudjatik and Wollaston domains is a pronounced NE-trending zone of rocks

with overall subdued aeromagnetic signature and that most of the major deposits are

situated within this zone (Figure 2.4) (Annesley et al., 2005). Fault lineaments were

interpreted from these aeromagnetic data based on the visual identification of sharp

contrasts in magnetic anomaly patterns and the displacement of continuous marker

horizons, such as a magnetic lithological boundary (Tschirhart et al., 2013). To

complement this map-based interpretation, a statistical analysis of fault trends was

undertaken; results are shown on the rose diagram in Figure 2.5. Three dominant sets of

sub-vertical faults were identified: NE-trending, NNW-trending and NW-trending. The

largest population of NE-trending fault lineaments can be separated into two subsets of

045° and 070°, whose respective directions are readily evident on the map (Figure 2.4), in

spite of the apparent continuity on the rose diagram (Figure 2.5). A set of less prolific,

NNW-trending lineaments have strikes ranging from 330° to 010°, with an average of

355°, whereas the smallest grouping of NW-trending lineaments encompasses strikes

ranging from 290° to 330° (average ~315°).

The orientations of the lineament arrays identified herein are similar to known

orientations of faults mapped in the Athabasca region by the Saskatchewan Geological

Survey, among others (e.g., Bosman et al., 2014). The most dominant, NE-trending faults

are interpreted to be the oldest based on the fact that their strong NE-oriented magnetic

grain is locally crosscut by the NNW-trending structures and locally also by the NW-

trending faults (Figure 2.4), although the latter relationship is not as obvious. Notably, the

oldest, NE-trending faults, which dip steeply to sub-vertically to the SE (Saskatchewan

Page 43: Geometric and hydrodynamic modelling of fluid-structural

29

Geological Survey, 2003; Hajnal et al., 2007), closely follow the structural grain of the

Wollaston and Mudjatik domain basement rocks, indicating that they most likely

represent reactivated Hudsonian structures (Annesley et al., 2005). The subset of 045°-

trending structures is comparable in orientation to the Snowbird tectonic zone and Needle

Falls shear zone (Figure 2.1), whereas the 070°-trending subset is an analogue for the

Key Lake Fault (e.g., Harvey and Bethune 2007). As suggested by Card et al. (2006), the

overall pattern invites comparison with the Riedel model, whereby 070°-trending

structures represent R shears in an overall dextral transcurrent system (at least at late

Trans-Hudson time; Portella and Annesley, 2000; Annesley et al., 2005). As discussed in

more detail in sections below, these NE-trending structures also appear to have been the

principal control on basement topography. The NW-trending structures appear to be the

second oldest fault group; however, there is only local evidence for cross-cutting

relationships and/or truncations, so one cannot rule out that they developed

synchronously with the NNW-trending faults (Portella and Annesley, 2000). The NNW-

trending faults are the youngest in the series, and likely represent the continuation of the

Tabbernor fault system (Portella and Annesley, 2000; Annesley et al., 2005; Card et al.,

2006), which is better developed east of the Athabasca Basin in the Hearne (Figure 2.1)

(Card et al., 2006). Field studies of exposed faults (Saskatchewan Geological Survey,

2003, and references therein) indicate that they are mainly sub-vertical. On the

aeromagnetic map (Figure 2.4), the lineaments associated with these faults clearly cut the

NE-trending faults, with evidence of apparent sinistral displacement of the related

magnetic signature in the area just north of the Peter Lake domain (see also Annesley et

al., 2005).

Page 44: Geometric and hydrodynamic modelling of fluid-structural

30

2.3.2 Three-dimensional (3D) modelling

The ParadigmTM GoCAD® Suite and Mira Geoscience GoCAD® Mining Suite

were used to develop 3D models of geologic surfaces in the study area from drill-hole

data. The Discreet Smooth Interpolation approach (DSI) implemented in GoCAD was

employed to create surfaces from discrete data. The DSI algorithm uses a spline function

and finite difference approximation to fit a generic surface (a triangulated mesh) to a set

of constraints imposed by imported geological datasets (Mallet, 2002). The construction

of such surfaces also takes into account the geologist’s interpretations, wherever

applicable, providing an additional human element to the interpretation (Mallet, 2002).

The data used in this study contain the X-Y-Z coordinates of geologic layers

intersected by 1206 drill holes (Figure 2.6) and relevant surface topographic information

(Digital Elevation Model, DEM) extracted from the Geological Atlas of Saskatchewan.

The unconformity map is mainly based on the 1206 drill holes; in addition, 1540 of the

total number of intersection points were incorporated from the basin-wide 3D model

created by Bosman et al. (2010) to further constrain the unconformity surface. For

reference, the average topographic elevation of the Athabasca Basin region is ca. 500 m

above mean sea level and all elevation information in this study is given in metres

relative to mean sea level. The Z dimension of the 3D model volume was set to range

from −1000 to +700 m, which enables visualization to a reasonable distance below the

unconformity surface (Figure 2.2).

Page 45: Geometric and hydrodynamic modelling of fluid-structural

31

Figure 2.4 Distribution of basement faults in the study area as interpreted from the aeromagnetic maps from

Card et al. (2010). Northeast faults (white), northwest faults (black) and north-northwest faults (blue) are

presented. Original in colour.

Page 46: Geometric and hydrodynamic modelling of fluid-structural

32

Figure 2.5 Rose diagram of 109 basement faults illustrating three dominant orientations: NE set, which can

be divided into two subsets: 045° and 070°, NNW set trends approximately 355° and NW set trends

approximately 315°. The horizontal axis and related semicircles display the frequency of the measurements

in a given petal.

Page 47: Geometric and hydrodynamic modelling of fluid-structural

33

Figure 2.6 Locations of drill holes used in this study.

Page 48: Geometric and hydrodynamic modelling of fluid-structural

34

Figure 2.7 Contour maps showing the elevations of the sub-Athabasca unconformity in the study area. Rose

diagram (upper left) shows that various topographic features (ridges, valleys) have a dominantly NE-SW

trend. See text for explanation of various topographic features (letters ‘A’ to ‘L’) on map. The locations of

three cross-sections shown in Figure 3.11 are also marked. Faults interpreted on the basis of aeromagnetic

data are also superimposed on the contoured unconformity surface. Original in colour.

Page 49: Geometric and hydrodynamic modelling of fluid-structural

35

Figure 2.7 displays a 2D contour map of the study area with interpreted fault sets

reported above, whereas Figure 2.8 displays the associated 3D image of the unconformity

surface, which, to enhance visualization, was constructed with a vertical exaggeration of

20 times. As shown on the contour map (Figure 2.7), the unconformity surface generally

increases in depth radially inward from basin margins. The elevation generally decreases

from circa +500 m (datum, mean sea level) at the southeastern margin to about −250 m at

the northwestern edge, with a highest elevation of +543 m and lowest elevation of −267

m, for a total relief 810 m. Thanks to the 3D calculation tool of GoCAD, the slope can

also be determined anywhere on the unconformity surface. Figure 2.9 presents a contour

map showing variation in slope of the unconformity surface across the study area. Overall,

the slope is very low, being on average 2.1°, with 90% of the values below 4.25°,

reflecting the regional basinal topographic variations described above. More pronounced

gradients, however, reaching a maximum of 39.6° are locally observed, especially

proximal to major faults, as discussed below. It should be kept in mind that slope

calculations are highly dependent on the density of observation points, and one must be

aware that ‘spurious’ slopes may appear. However, even with the non-systematic

distribution of observation points in our model (Figure 2.9), this slope data map still

provides useful first-order information on the regional scale.

More significantly, the model reveals a number of topographic features, including

ridges and valleys, in the unconformity surface, that have dominantly NE-SW trends

(Figure 2.7). These topographic features, delineated by thick dashed lines labelled ‘A’

through ‘L’, are evident on the contour map (Figure 2.7) and corresponding 3D view

(Figure 2.8). They also stand out prominently on the accompanying slope map (Figure

Page 50: Geometric and hydrodynamic modelling of fluid-structural

36

2.9). Most of these features correlate with interpreted NE-trending basement faults and

exhibit a close spatial relationship to uranium deposits and prospects (Figure 2.7). The

most notable of these is a semi-continuous ridge (‘A’ in Figure 2.7) situated next to the

Phoenix and allied deposits in the north-central part of the area. This ridge can be traced

to the NNE, where it splays into two shorter segments (‘B’ and ‘C’), and to the south,

west of Key Lake (‘D’). To the northeast, the McArthur River deposit is situated between

two prominent ridges (‘E’ and ‘F’), and to the east the BJ zone is located alongside an

elongate dome (‘G’). Along ridge segment ‘A’, near the Phoenix deposit, an abrupt NW

to SE change in elevation of the unconformity surface is recorded, from an elevation of

circa 0 m on the NW side and +120 m on the SE side, to +320 m on the top of the ridge,

for a maximum relief of 320 m. As discussed in more detail below, this semi-continuous

ridge coincides, along much of its length, with a prominent basement quartzite unit,

suggesting that contrasting rock competencies in the sub-Athabasca basement played a

role in its development. It also stands out on the aeromagnetic map (Figure 2.4), sharply

delineating the NW boundary of the prominent low in the Mudjatik-Wollaston

transitional zone. In addition, in the south, in the vicinity of Key Lake, many short ridges

and valleys are developed with NE and ENE trends (‘H’, ‘I’ and ‘J’ in Figure 2.7). A

dense array of similarly oriented basement faults was identified in this area and must

have controlled these features (Figure 2.4). However, it is notable that not all of the

topographic features trend NE; for example, the Millennium deposit is situated next to a

north-trending lineament (‘K’) that deviates from typical trends. Finally, a prominent

break in slope about 15 km inboard of the basin margin represented by the curvilinear

trace of ‘L’ on Figures 2.7–2.9, continuously mantles the basin margin at about the +300

Page 51: Geometric and hydrodynamic modelling of fluid-structural

37

m contour. We suggest that this break in slope, well-imaged in cross-sections presented

below, likely represents an early basin margin, and/or possibly the locus of an early

normal fault that accommodated subsidence (Ramaekers and Catuneanu, 2004;

Ramaekers et al., 2007) (see Figure 2.11).

In Figure 2.10, uranium deposits, prospects and occurrences are projected onto the

unconformity surface. Using the GIS proximal query function, it was determined that 90%

of these uranium showings occur within 1 km of various topographic features and

lineaments, with many being situated either right on or adjacent to them. In addition,

these topographic features and lineaments tie directly to EM conductors (Figure 2.10),

suggesting that the vast majority of the conductors are fault controlled. Another apparent

spatial association is the overlap between the illite alteration corridor and the identified

topographic features (Figure 2.10). Moreover, the most significant (first order) NE-

trending ridges – ‘A’, ‘B’, ‘C’, ‘D’, ‘E’, ‘F’ and ‘G’ – coincide roughly with the regional

dravite and chlorite alteration corridors (Figure 2.10), as well as several major uranium

deposits (McArthur River and Phoenix) and occurrences (Gryphon zone and Mann Lake),

implying that these ridges (and related faults) were fluid conduits. This highlights the

important coincidence of topographic features of the sub-Athabasca unconformity and

faults in the circulation of ore-forming fluids.

Page 52: Geometric and hydrodynamic modelling of fluid-structural

38

Figure 2.8 3D view of the sub-Athabasca unconformity in the study area. See text for explanation of

various topographic features (letters ‘A’ to ‘L’) on map. Original in colour.

Page 53: Geometric and hydrodynamic modelling of fluid-structural

39

Figure 2.9 Calculated slope of the unconformity surface in the study area. See text for explanation of

various topographic features (letters ‘A’ to ‘L’) on map. Original in colour.

Page 54: Geometric and hydrodynamic modelling of fluid-structural

40

Figure 2.10 Axes of topographic features of the unconformity, interpreted faults, ground electromagnetic

(EM) conductors and uranium deposits, occurrences and prospects are superimposed onto the regional clay

alteration map. For letters from ‘A’ to ‘L’, see text for explanation. Original in colour.

Page 55: Geometric and hydrodynamic modelling of fluid-structural

41

Figure 2.11 Three cross-sections derived from the 3D model, showing main units of Athabasca Group, the

basement and the locations of the McArthur River, Phoenix and Key Lake deposits. See Figure 2.7 for

locations of the cross-sections. Original in colour.

Page 56: Geometric and hydrodynamic modelling of fluid-structural

42

2.4 Interpretation and discussion

2.4.1 Evaluating the factors at play

The structural analysis and associated 3D modelling undertaken in this study

clearly demonstrate a strong spatial relationship between linear topographic features of

the unconformity, faults and uranium mineralization (and related alteration). As discussed

in the introductory section of this paper, however, previous work has documented

primary (or paleo-) topographic relief at several deposits, including Key Lake and

McClean Lake, which was ascribed to lithological contrasts and related pre-Athabasca

Group differential weathering. In these cases, it is also clear that post-Athabasca fault

reactivation contributed further to the development of topography. In order to more fully

evaluate the relative influence of these factors, three NW-SE trending sections illustrating

unconformity relief in relation to the main lithological and structural features were

prepared from the 3D model (Figures. 2.7–2.9 and 2.11). Sections X-X’ and Y-Y’

traverse the prominent ridges proximal to the McArthur River and Phoenix deposits,

respectively, whereas Section Z-Z’, in the south, traverses the Key Lake deposit. To

complement these regional sections, and serve to further guide discussion, three more

detailed sections of deposit-scale relationships are presented in Figure 2.12.

Page 57: Geometric and hydrodynamic modelling of fluid-structural

43

Figure 2.12 Simplified cross-sections of the three deposits discussed in text. (a) Sectional view of the

McArthur River deposit (after White et al., 2007; Bronkhorst et al., 2012); (b) sectional view of the

Phoenix deposit (after Kerr, 2010); (c) sectional view of the Deilmann orebody, Key Lake deposit (after

Harvey and Bethune, 2007 with permission from the editors). Original in colour.

Page 58: Geometric and hydrodynamic modelling of fluid-structural

44

2.4.1.1 McArthur River deposit

The first section, X-X’ (Figure 2.11a), traverses the relatively well-studied

McArthur River deposit in the northern part of the study area (Figure 2.7), known for a

strong structural control associated with the P2 fault (McGill et al., 1993). The basement

at McArthur River comprises moderately SE-dipping panels of Archean gneiss and

Paleoproterozoic Wollaston Supergroup metasedimentary rocks, mainly pelitic gneiss

and quartzite (Marlatt et al., 1992). This deposit is situated along the unconformity where

it is intersected by the P2 fault (Figures 2.11a and 2.12a). The latter is localized along the

contact between a thick quartzite unit, underlying topographic ridge ‘E’ in the 3D model

(Figures 2.7-2.10), and graphitic gneisses, and demonstrably offsets the unconformity

(Figure 2.12a; Bronkhorst et al., 2012), as indicated by the sharp break in slope between

ridge segments ‘E’ and ‘F’ (Figures 2.8 and 2.9). Seismic reflection results (Györfi et al.,

2007; Hajnal et al., 2007; White et al., 2007) demonstrated the P2 fault has listric

geometry with a steep dip at the surface that shallows at depth (Figure 2.11a). Based

mainly on seismic data, Györfi et al. (2007) also interpreted some ridges flanking the east

side of P2 fault as back-thrusts antithetic to the main reverse fault; one of these is

incorporated in section X-X’, albeit in a simplified way. According to McGill et al.

(1993), and later Bernier (2002, 2004), the P2 fault zone preserves a record of multiple

reactivations with a subtle to 80 m net reverse offset of the unconformity that placed the

basement rocks above the basal conglomerate and sandstone units of the Athabasca

Group. In the basement, the P2 fault zone strikes ~050° and is hosted within graphite-

bearing pelitic gneiss, whereas in the sandstone the P2 splays into multiple faults that run

Page 59: Geometric and hydrodynamic modelling of fluid-structural

45

parallel to bedding (Figure 2.12a) (McGill et al., 1993) such that formation contacts in

the sandstone are not displaced (Figure 2.11a).

The regional section (Figure 2.11a) highlights two important features that were

initially modelled seismically in a small 3D area by Györfi et al. (2007). First, the P2

fault clearly coincides with a significant topographic depression in the unconformity

surface, bounded by the two prominent ridges, ‘E’ underlain by quartzite, and ‘F-G’, a

composite ridge comprising hanging wall pelitic gneisses (‘F’) and an overlying,

topographically higher quartzite unit (‘G’). The base of the valley right next to the fault is

at −30 m, whereas the crest of ridges (‘E’ and ‘F’) reach +130 and +150 m, respectively,

translating to a total relief of 180 m. Second, the basal unit of the Athabasca Group, the

Read Formation, is attenuated and pinched out against the ridge (or elongate dome, as

viewed in 3D) near the BJ zone (‘G’ in Figures 2.7–2.9 and 2.11a), a relationship

originally documented by Bernier (2004), whereas the overlying MFb member is severely

thinned (Figure 2.11a). This onlapping relationship indicates that this more southeasterly

ridge/dome must have been a topographically high-standing feature at the time of

deposition. Similar paleotopographic relief and restricted distribution of the Read

Formation was noted by Long (2007) in the Sue C pit, by Jefferson et al. (2001) across

the Wheeler River ridge, and by Yeo et al. (2007) beside a ridge west of McArthur River,

where DDH-RL92 (Figure 2.11a) intersected thick angular paleo-talus deposits of the

basal Read Formation flanking a fault scarp. In contrast, the other ridges, which are

blanketed by both the Read Formation and MFb member, may have been relatively lower

during deposition, then subsequently uplifted. Regarding the quartzite along this section,

it is worth noting that Card (2012, 2013) studied the core from drill hole MAC-167,

Page 60: Geometric and hydrodynamic modelling of fluid-structural

46

which intersects one of the pinnacles on ‘G’ (Figures 2.8 and 2.11a), in detail, and

discovered a zone of intense silicification, as was also observed at the Phoenix deposit

(Wang et al., 2015). In terms of the factors at play, this substantiates the idea that these

three quartzite ridges may represent silicification zones related to pre-Athabasca Basin

fault systems (Card 2012, 2013), and thus that the early topography was generated at least

in part by faulting and related hydrothermal alteration. It is also considered possible that

the alteration was superposed on pre-existing orthoquartzite units (of metamorphic

origin), resulting in severe modification of original textures, but this question requires

further detailed investigation, which is beyond the scope of this paper. Subsequent pre-

Athabasca preferential erosion between the more competent quartzite and incompetent

graphite-rich and pelitic units laid the foundation for the development of the topographic

features. Nevertheless, post-Athabasca compressional activity, represented by reverse

brittle displacement along the P2 fault, also played a role, generating additional relief and

creating the basement-wedge structures that host the high-grade uranium mineralization

(McGill et al. 1993; Figure 2.12a). This reactivation phase clearly must have accentuated

the already existing topographic low between ridges ‘E’ and ‘F-G’ (Figure 2.11a),

creating a more substantial depression that may have served to focus the flow of

uraniferous fluids.

2.4.1.2 Phoenix deposit

As seen in Figures 2.11b and 2.12b, the Phoenix deposit shares similar geological

features with McArthur River, being hosted by graphitic pelitic rocks next to a prominent

quartzite ridge. The quartzite unit at Phoenix strikes northeast (~055°) and dips

Page 61: Geometric and hydrodynamic modelling of fluid-structural

47

moderately (45°) to steeply (75°) to the southeast (Kerr, 2010). Kerr (2010) reported that

the geological units at Phoenix form a southeast-dipping homoclinal panel. He postulated

that the quartzite ridge may have initially developed as a result of preferential erosion of

pre-Athabasca fault zones localized along the contact between the mechanically

competent quartzite unit and relatively incompetent graphitic pelite. In terms of the origin

of the quartzite and graphite, and as suggested for the McArthur deposit and in the

introductory section of this paper, recent studies by both Card (2012, 2013) and Wang et

al. (2015) advocate that the quartzite and graphite may partly be of hydrothermal origin,

implying that both the strong/hard (quartzite) and weak/soft (graphitic pelite) rocks

attained their character in part from fluids moving along early ductile faults, before

development of the Athabasca Basin. This pre-Athabasca differential weathering was

superseded by syn- to post-Athabasca fault reactivation. The main ‘WS shear zone’,

which lies within or along the base of the graphitic pelite unit controlling the Phoenix

deposit, creates a small scarp in the graphitic pelitic unit with a vertical displacement of

less than 10 m (Figure 2.12b); however, other drill holes show that the unconformity has

been displaced as much as 60 m vertically by the reverse shear zones localized along the

NW contact of this quartzite unit, which appears to have controlled the WR-204

mineralization zone (Figure 2.11b) (Kerr 2010).

The regional section Y-Y’ (Figure 2.11b) illustrates that, like the McArthur River

deposit, there is obvious thinning of basal Athabasca Group units over the quartzite ridge

in the Phoenix deposit area, confirming earlier observations by Jefferson et al. (2001).

Moreover, in the case of Phoenix a distinctive sedimentary breccia, interpreted as a fault-

scarp talus deposit, was identified along the western margin of the ridge (Bosman and

Page 62: Geometric and hydrodynamic modelling of fluid-structural

48

Korness, 2007), indicating a pre-depositional high-standing feature that contributed

detritus while Read Formation was being deposited.

2.4.1.3 Key Lake deposit

Much farther south, near the margin of the basin, the Key Lake deposit is

characterized by a paleovalley, where it is intersected by the Key Lake fault, as shown in

section Z-Z’ (Figures 2.11c and 2.12c). The Key Lake deposit is quite similar to both

McArthur and Phoenix, the difference being that the fault is localized along the contact of

competent Archean granitoid unit and Wollaston Supergroup graphitic metasedimentary

rocks (Figure 2.12c) (Harvey and Bethune, 2007), rather than being proximal to a

quartzite ridge. The Key Lake fault also dips steeply northwest, the dip being governed

by the attitude of pre-Athabasca faults, which were localized along the steeply NW-

dipping limb of a regional-scale, anticline cored by the Archean rocks (Harvey, 2004;

Harvey and Bethune, 2007).

The sub-Athabasca paleovalley identified near the Deilmann orebody corresponds

to an east- to northeast-trending, asymmetric valley that plunges and widens towards the

east, representing a vertical change of 15 m over a lateral distance of 100 m (Figure 2.12c)

(Harvey, 2004; Harvey and Bethune, 2007). The paleovalley is a primary feature, as

evidenced by onlapping of the basal Manitou Falls Formation onto the unconformity

surface (Figure 2.12c), a feature first noted in well-exposed faces of the Deilmann pit

(Collier and Yeo, 2001; Harvey, 2004; Harvey and Bethune, 2007). Similar to the

McArthur and Phoenix deposits, Harvey and Bethune (2007) reported possible pre-

Athabasca graphite enrichment in the pelitic basement rocks. These sheared graphite-

Page 63: Geometric and hydrodynamic modelling of fluid-structural

49

enriched rocks were clearly of lower resistance and were preferentially eroded to form

the valley systems. The fact that the break in the unconformity slope parallels the strike

of the basement pelitic units provides added evidence that the latter directly controlled

paleovalley development. As at the McArthur and Phoenix, these pre-Athabasca

deformation zones were later reactivated. For example, the ore-controlling Key Lake

fault, which was initially formed during late-D3 (pre-Athabasca) deformation, was

reactivated during D5 (post-Athabasca), offsetting the unconformity surface by up to 12

m (Figure 2.12c) (Harvey and Bethune, 2007). These D5 faults created conduits for fluid

flow resulting in post-Athabasca alteration and mineralization, as manifested by

silicification zones, a dravite halo, and kaolinite alteration in the sandstone coincident

with D5 faults and associated breccia zones (Harvey, 2004; Harvey and Bethune, 2007).

2.4.2 Conceptual model and relevance to uranium exploration/ore systems

Based on this study and other geological information available in the literature, a

conceptual evolutionary model of the sub-Athabasca Group unconformity surface is

proposed (Figure 2.13). The early stages of Hudsonian deformation (D1/2) were

characterised by ductile shearing/thrusting localized along the contact between Archean

basement rocks and Paleoproterozoic cover rocks of the Wollaston Supergroup (Tran,

2001). Thrust faults (and related recumbent folds) were also developed above this

regional décollement (Figure 2.13a) (Tran, 2001). These early structures, including a

regionally extensive ductile transposition foliation (composite S1/2), were subsequently

tightly folded about NE-trending axes (D3), giving rise to steeper structures (Tran, 2001).

As D3 deformation continued, further post-peak metamorphic shortening was

Page 64: Geometric and hydrodynamic modelling of fluid-structural

50

accommodated by shear displacement of fold limbs (Figure 2.13b) (Card, 2006). Some of

the steep, reverse shear zones may have served as fluid movers, possibly giving rise to

silicification and graphite enrichment (Card, 2012, 2013), as well as contributing to the

differences in permeability and rock strength (Figure 2.13b). During subsequent uplift

and erosion these basement units were differentially weathered (Figure 2.13c). Low-lying

areas were created by preferential erosion of weaker graphitic rocks that were generally

coincident with early fault zones, whereas high ridges were developed over stronger

quartzite (± hydrothermal quartz?) and granitoid units of the Archean inliers (Figure

2.13c). Finally, during Athabasca Group deposition, many faults were reactivated, once

again becoming fluid conduits. This dual history is indicated by the near-coincidence of

post-Athabasca silicification and clay alteration in the sandstone column and underlying

(pre-Athabasca) faults and metasomatic rocks (Earle and Sopuck, 1989; Jefferson et al.,

2007) (Figure 2.13c). At this stage, many primary topographic features inherited from

differential weathering were accentuated and some new topography was created (e.g., P2

fault) (Figure 2.13c).

It is well known that syn- to post-Athabasca Group reactivated faults/shear zones

have acted as conduits and barriers to fluid flow, mainly vertically along the fault

conduits causing basement alteration and mineralization (Jefferson et al., 2007). The

importance of topographic features, however, has not attracted enough attention. As

suggested by Harvey and Bethune (2007; cf. Harvey, 2004) and Long et al. (2007),

basement paleotopography may have played an important role in controlling, and perhaps

localizing, Athabasca Group sedimentation and fluid flow related to mineralization.

Topographic features may have exerted some control on fluid flow related to alteration

Page 65: Geometric and hydrodynamic modelling of fluid-structural

51

and mineralization both in terms of their morphology and orientation. Furthermore, one

can envisage that uraniferous fluid flow may have slowed down in togographically low

areas due to increased ‘accommodation space’, thus facilitating chemical reactions to

precipitate uraninite. In addition, some rock types, such as quartzite, would have been

relatively impermeable and would have acted as physical barriers (or ‘aquitards’; Hiatt

and Kyser, 2007; Jefferson et al., 2007) to fluid flow (Kerr, 2010). This may have been

the case in the Phoenix deposit, where uraniferous fluid flow was limited between the

relatively high-standing quartzite ridge to the NW and the small (~10 m high) fault scarp

created by reverse displacement along the WS shear to the SE (Figure 2.12b) (Kerr,

2010), a comparable setting to that of the mineralization at McArthur River (Marlatt,

1992). In other cases, paleovalley walls may have acted as vertical barriers based on

porosity changes in the sediments, limiting the horizontal flow to within the structure.

This was likely the case for the uranium ore bodies at Key Lake, which occupied

topographic depressions along pre-existing fault trends (Harvey, 2004; Harvey and

Bethune, 2007).

In summary, although the unconformity topographic features immediately

adjacent to uranium deposits may be positive (ridges) or negative (valleys), most are

ultimately related to faults. Most of these faults were developed before the basin formed,

and clearly contributed to shaping the paleotopography of the unconformity surface that

controlled early sedimentation. Some of the faults were reactivated after deposition of the

Athabasca Group, further modifying the topography of the unconformity surface. The

formation of uranium deposits requires circulation of large amounts of fluids through

localized, high-permeability areas (Raffensperger and Garven, 1995; Cui et al., 2012b;

Page 66: Geometric and hydrodynamic modelling of fluid-structural

52

Chi et al., 2013). Pre-Athabasca ductile faulting and related hydrothermal alteration may

have selectively enhanced the permeability of some basement rocks, setting the stage for

large fluid fluxes and related uranium mineralization during the reactivation stage.

Alternatively, pre-Athabasca alteration may have decreased the permeability of some

units rendering them barriers to fluid flow. This may apply to the basement quartzite

units, which are arguably metamorphic in origin, but are locally intensely silicified (Card,

2012, 2013; Wang et al., 2015). These altered quartzites may in fact have had a dual role,

being competent bodies that served to localize post-Athabasca faulting and at the same

time acting as aquitards, focussing fluid flow related to mineralization. As discussed in

detail by Hiatt and Kyser (2007), fluids were likely also limited in a vertical sense, within

the sandstone column, being channelled horizontally through the relatively permeable

lower members of the Manitou Falls and Read formations. While the unconformity (and

related rocks immediately above it) may have been a significant aquifer, it seems

probable that high-standing basement features, in some cases being impermeable

themselves (e.g., quartzite), but also producing subtle facies variations and related

porosity/permeability characteristics in the Athabasca Group strata above, would have

exerted some influence on such lateral flow (Mwenifumbo et al., 2007). Whatever the

case, the fundamental nature of the spatial relationship between unconformity topography

and uranium deposits is that both of them are tightly related to faults with multi-stage

histories, as suggested by the spatial coincidence of pre-Athabasca fault-related basement

alteration (silica and graphite enrichment) and syn- to post-Athabasca fault-reactivation-

related alteration (silicification, clay mineral alteration, dravite alteration), coupled with

Page 67: Geometric and hydrodynamic modelling of fluid-structural

53

uranium enrichment in the overlying sandstone column (e.g., Hoeve and Quirt, 1984;

Earle and Sopuck, 1989; Jefferson et al., 2007; Card, 2012, 2013).

2.5 Conclusions

The 3D modelling and structural analysis of the southeastern Athabasca Basin

undertaken in this study yields the following conclusions:

1. The 3D model of the sub-Athabasca unconformity reveals numerous,

dominantly NE-trending ridges and valleys.

2. These topographic features are the products of the combined action of three

main factors: 1) pre-Athabasca group ductile-brittle faulting and alteration (i.e.,

silicification); 2) differential weathering and erosion; and 3) syn- to post-Athabasca

ductile-brittle reactivation of pre-existing graphite-rich, ductile shear zones.

3. Regional-scale structural analysis identified three dominant sets of faults: NE,

NW, and NNW, in chronological order from oldest to youngest. The oldest, NE-trending

faults represent reactivated Hudsonian structures and were the principal control on

basement topography and uranium mineralization.

4. The 3D model illustrates that the most significant, NE-trending fault zone is

delineated by a prominent (up to 320 m high) ridge, and is coincident with a change in

magnetic signature that extends semi-continuously from the Key Lake mine in the south

to McArthur River in the north, with numerous subsidiary faults and splays. The fault

zone is encompassed by a northeast-elongate alteration corridor that hosts several large

deposits (Key Lake, Phoenix, McArthur), as well as a number of promising new

prospects (e.g., Gryphon zone, Mann Lake).

Page 68: Geometric and hydrodynamic modelling of fluid-structural

54

Figure 2.13 Schematic conceptual evolutionary model of the sub-Athabasca unconformity surface. (a) NW-

SE cross-sectional view of the Wollaston fold-thrust belt at earlier stages of its development. Development

of regional foliation, basal décollement, thrusting; (b) Later stages of its development. Structure dominated

by D3 close to tight upright folds, late-D3 reverse faults offset fold limbs. Faults serve as hydrothermal

fluid conduits; (c) Geological scenario following uplift, erosion, and Athabasca Group deposition. Syn- to

post-Athabasca fault reactivation and enhancement of primary topography. Faults are re-used as fluid

conduits. Original in colour.

Page 69: Geometric and hydrodynamic modelling of fluid-structural

55

5. The spatial association of deposits with this NE-trending fault zone and related

basement topography indicates that both played critical roles in mineralization. Pre-

Athabasca ductile faulting and related hydrothermal alteration probably enhanced the

permeability of some basement rocks, but decreased others (e.g., quartzite), preparing the

ground for large fluid fluxes, and controlling the related fluid pathways, during the

reactivation stage. Both the fault zones themselves, as well as the primary (pre-Athabasca)

and secondary topography, would have also influenced the circulation of ore-forming

fluids, controlling alteration patterns and mineralization. The broad aeromagnetic low

characterizing the Mudjatik-Wollaston transitional zone may in part reflect such high

fluid fluxes.

6. This study indicates that NE-trending faults, and their related topographic

lineaments, are the most highly prospective for mineralization. It also supports notions

that the highest ridges, recording the largest magnitudes of unconformity offset, represent

the most promising exploration targets.

7. Finally, this paper has drawn attention to the close spatial relationship between

multi-stage faults and topographically high-standing ‘quartzite ridges’ in the basement.

This spatial association, coupled with the observation that many quartzite units record

strong pre-Athabasca hydrothermal alteration, could imply an important, heretofore

unrecognized genetic linkage. While admittedly contentious, the assertion that such

‘quartzites’ are potentially hydrothermal in origin and related to pre-Athabasca faulting

(e.g., Card, 2012, 2013) is a testable hypothesis that clearly requires further in-depth

investigation.

Page 70: Geometric and hydrodynamic modelling of fluid-structural

56

This contribution represents part of an ongoing Ph.D. study. Using the 3D model

presented in this paper as a guide, work is continuing with coupled thermal-mechanical

fluid flow modelling (FLAC3D) to more rigorously test how topography, enhanced

permeability, and deformation associated with fault zones influenced fluid flow patterns

and related alteration-mineralization in this part of the basin.

2.6 Acknowledgements

This research was supported by the TGI-4 (Targeted Geoscience Initiative 4)

uranium ore systems program led by the Geological Survey of Canada (Natural

Resources Canada), through a grant to K. Bethune and G. Chi. Access to the GoCAD

licenses was kindly provided by the Saskatchewan Geological Survey. Constructive

reviews by William A. Morris, Charles W. Jefferson and Bruno Lafrance greatly

improved the work presented herein.

Page 71: Geometric and hydrodynamic modelling of fluid-structural

57

Chapter 3 The effects of basement faults on thermal convection3

3.1 Introduction

The Athabasca Basin in northern Saskatchewan and Alberta, Canada (Figure 1)

hosts the richest unconformity-related uranium deposits in the world, which are generally

located at the intersection of reactivated basement faults and the basin-basement

unconformity (Jefferson et al., 2007). Early genetic models of unconformity-related

uranium mineralization invoked pre-Athabasca weathering at the top of the basement as

the main mechanism of ore enrichment (e.g., Knipping, 1974; Langford, 1978), but the

observation that the ore-controlling faults and mineralization-associated alteration

crosscut the unconformity indicates that the mineralization is syn- to post-Athabasca

Basin, which, together with the fluid inclusion data indicating high fluid temperatures

(~200 °C) and salinities both in the basin and in the deposits, led to the widely adopted

“diagenetic-hydrothermal” model (e.g., Pagel, 1975; Hoeve and Sibbald, 1978; Hoeve et

al., 1980; Hoeve and Quirt, 1984; Kotzer and Kyser, 1995; Fayek and Kyser, 1997;

Thomas et al., 2000). The original diagenetic-hydrothermal model proposed by Hoeve

and Sibbald (1978) suggested that unconformity-related uranium mineralization resulted

from mixing of an oxidizing, uranium-bearing, basin-derived fluid with a reducing,

hydrocarbon-bearing (derived from fluid-graphite reactions) fluid from the basement.

More recent geochemical studies of fluid inclusions and boron isotopes of tourmaline

associated with mineralization support the idea that the mineralizing fluids were derived

from the basin (Richard et al. 2010, 2011, 2013, 2014; Mercadier et al. 2012), but the

3 This chapter has been submitted to Geofluids and is under peer review, “Li, Z., Chi, G., and Bethune,

K.M., 2015. The effects of basement faults on thermal convection and implications for the formation of

unconformity-related uranium deposits in the Athabasca Basin.”

Page 72: Geometric and hydrodynamic modelling of fluid-structural

58

source of uranium is still under debate, i.e., sediment-derived (e.g., Hoeve and Sibbald,

1978; Hoeve et al., 1980; Kotzer and Kyser, 1995; Fayek and Kyser, 1997; Kyser et al.,

2000) versus basement-derived (e.g., Hecht and Cuney, 2000; Cuney et al., 2003; Richard

et al., 2010).

Thermal convection has been advocated by many researchers as a potential

mechanism for mass and heat transport and related mineralization in sedimentary basins

(e.g., Raffensperger and Garven, 1995; Raffensperger and Vlassopoulos, 1999; Garven et

al., 2001; Simms and Garven, 2004; Yang et al., 2004b; Large et al., 2005; Oliver et al.,

2006; Cui et al., 2012a; Guillou-Frottier et al., 2013) and various numerical studies have

been conducted to test how the faults affect thermal convection inside the basinal

sequence (e.g., Simms and Garven, 2004; Yang et al., 2004b; Yang, 2006; Souche et al.,

2014) or in permeable rock masses (e.g., Bächler et al., 2003; Kühn et al., 2006;

Malkovsky and Pek, 2015). For example, Simms and Garven (2004) demonstrated that in

a faulted rift basin, model geometry, including faults and boundaries, can control

convection cell spacing and structure. Malkovsky and Pek (2015) further showed that

high fault-zone permeability and hydrostatic fluid pressure favour the development of the

fault-bounded convective flow in homogeneous permeable rocks. For the Athabasca

Basin, it has been proposed that convective fluid flow driven by a thermal gradient may

have developed in the high-permeability basinal sequence at the basin scale

(Raffensperger and Garven 1995), and that large-scale free thermal convection may be

responsible for uranium uptake and transport (Hoeve and Quirt, 1984; Kyser and Cuney,

2009). More recent diagenetic studies based on fluid inclusions and illite

geothermometers also suggest very steep isotherms in the lower part of the basin

Page 73: Geometric and hydrodynamic modelling of fluid-structural

59

indicating that it may have experienced extensive thermally-driven fluid convection (Chi

et al., 2015; Chu and Chi, 2016). Free thermal convection was also inferred in the Thelon

Basin (Cui et al., 2010), considered a counterpart of the Athabasca Basin in northern

Canada (Figure 3.1), based on numerical studies at geothermal gradients of 25 to

35 °C/km. Furthermore, numerical modelling by Cui et al. (2012a) showed that

buoyancy-driven thermohaline convection may have penetrated into the low-permeability

basement at extremely low velocities. However, none of these numerical studies fully

examined the role of faults in fluid convection, and some studies were based on highly

simplified conceptual basin-scale models with a highly exaggerated vertical scale, which

in turn exaggerated the horizontal size of individual convection cells and made

incorporation of faults in the models difficult.

Fault zones commonly exert primary controls on fluid flow (Caine et al., 1996)

and have been generally considered as one of the most important permeable fluid

pathways for hydrothermal ore-forming systems (e.g., Cox, 2005). Numerous studies

have been conducted to examine the interaction of faults and fluid flow in various ore

systems (e.g., Garven et al., 2001; Matthäi et al., 2004; Simms and Garven, 2004; Yang

et al., 2004b; Oliver et al., 2006; Zhang et al., 2007; Feltrin et al., 2009; Schaubs and

Fisher, 2009). For the Athabasca Basin, the linkage between basement faults and uranium

deposits has been well established (Hoeve and Sibbald, 1978; Jefferson et al., 2007;

Kyser and Cuney, 2009). Based on numerical dynamic modelling, Schaubs et al. (2005)

proposed that the dip of the basement faults and strike with respect to the maximum

shortening direction may be an important factor in controlling the site of potential fluid

mixing and uranium mineralization in the Athabasca Basin. Cui et al. (2012b) further

Page 74: Geometric and hydrodynamic modelling of fluid-structural

60

suggested that ingress flow and egress flow were associated with alternating extensional

and compressional stress fields. Numerical studies done by Li et al. (2015b) also

confirmed that fluid migrates up a fault during compressional deformation, and the

models with low dip angle faults and those with fault offsets of the unconformity show

slightly greater flow rates than the other models.

However, deformation along the faults is typically episodic in nature, and not all

unconformity-related uranium deposits demonstrate active faulting during mineralization

(e.g., Cigar Lake), which is not unusual considering the overall non-deformed nature of

the Athabasca Basin. It is, therefore, of interest to investigate the roles faults may have

played in fluid flow during periods of tectonic quiescence. In addition, some researchers

have proposed that basement graphitic units, being conductors, may have given rise to

thermal anomalies that drove fluid convection (e.g., Hoeve and Quirt 1984), whereas

others have suggested that the more conductive graphitic zones appear to have no impact

on the regional flow patterns at the basin scale (Raffensperger and Garven, 1995). The

question is: can a minor thermal anomaly induced by elevated thermal conductivity drive

flow convection at a smaller scale? If so, how will the distribution of the high-

conductivity fault zones influence convection? In particular, we want to determine

whether or not: 1) individual faults can affect the location and sizes of the convection

cells; 2) the distance between two faults influences the convection pattern; 3) some fault

configurations favour outward (egress) fluid flow, whereas others favour inward (ingress)

flow; 4) fluid can flow into a fault and out the same fault in a fluid convection framework;

and 5) a change of the thermal gradient can affect the interaction between faults and

thermal convection. In order to address these questions, a series of numerical modelling

Page 75: Geometric and hydrodynamic modelling of fluid-structural

61

experiments with equal horizontal and vertical scales and various combinations of fault

geometries and spacing, permeabilities and heat conductivities were carried out. The

implications of the results for the genesis of unconformity-related uranium mineralization

are discussed.

3.2 Geological background

3.2.1 Crystalline basement, the Athabasca Group and younger intrusive rocks

The late Paleoproterozoic to Mesoproterozoic Athabasca Basin of northern

Saskatchewan and Alberta, Canada (Figure 3.1), unconformably overlies a crystalline

basement complex of Archean to Paleoproterozoic rocks (Jefferson et al., 2007). From

west to east, these rocks comprise mainly highly-deformed, medium- to high-grade

metamorphic granitoid gneisses, metasedimentary rocks and mafic to felsic plutons,

belonging to the Taltson magmatic zone, the Rae Province, and the Hearne Province,

respectively (Card et al., 2007a, b) (Figure 3.1). The Rae and Hearne provinces, forming

the larger Churchill Province of the Canadian Shield and separated by the northeast-

trending Snowbird tectonic zone (Hoffman, 1988), are divided into several domains

(Card et al., 2007a, b) (Figure 3.1).

Page 76: Geometric and hydrodynamic modelling of fluid-structural

62

Figure 3.1 Location and regional geologic framework of the Athabasca Basin (after Card et al., 2007a, b;

Ramaekers et al., 2007; Card et al., 2014). FP = Fair Point; S = Smart; RD = Read; MF = Manitou Falls;

LZ = Lazenby Lake; W= Wolverine Point; LL = Locker Lake; O = Otherside; D = Douglas; C = Carswell.

AB = Athabasca Basin. Th = Thelon Basin. Red stars represent major uranium deposits. Original in colour.

Page 77: Geometric and hydrodynamic modelling of fluid-structural

63

Figure 3.2 Representative sections of three well-known unconformity-related uranium deposits of the

eastern Athabasca Basin showing the strong spatial association of the deposits with the intersection of

basement-rooted fault zones and the unconformity surface. (A) Cigar Lake deposit, consisting of

predominantly unconformity ore and perched ore in the overlying sandstone (after Jefferson et al., 2007).

(B) Deilmann pit, Key Lake deposit, including both basement-hosted and unconformity ore, controlled by

the Key Lake fault (after Harvey and Bethune 2007). (C) Eagle Point deposit, mostly basement-hosted ore,

controlled by the Collins Bay thrust and Eagle Point fault (after Marlatt et al., 1992). MF = Manitou Falls

Formation. Vertical scale = horizontal scale. Original in colour.

Page 78: Geometric and hydrodynamic modelling of fluid-structural

64

The sedimentary rocks of the Athabasca Basin, known as the Athabasca Group,

unconformably overlie the crystalline basement. Sediment deposition commenced at ca.

1750 Ma (Orrell et al., 1999; Alexandre et al., 2009) and lasted until ca. 1500 Ma

(Ramaekers et al., 2007). In detail, the Athabasca Basin consists of three smaller,

northeast-trending, stacked sub-basins of different configuration and age, i.e., the

Jackfish Sub-basin, the Cree Sub-basin and the Mirror Sub-basin (Ramaekers et al.,

2007). The present thickness of the Athabasca Group in the central part of the basin is

about 1,500 m (Ramaekers et al., 2007); however, the original thickness is thought to

have reached 5 to 6 km by early Mesoproterozoic time on the basis of the P-T estimates

from fluid inclusion studies (Pagel, 1975). There are nevertheless significant uncertainties

regarding the thickness of eroded strata due to ambiguity of the fluid inclusion data and

their interpretations (Chi et al., 2015). The preserved sedimentary record of the

Athabasca Group consists of the following formations including, from the base to the top:

quartz-rich sandstones of the Fair Point, Read, Smart and Manitou Falls formations;

sandstones, siltstones and mudstones of the Lazenby Lake and Wolverine Point

formations; sandstones of the Locker Lake and Otherside formations; shales of the

Douglas Formation; and stromatolitic carbonates of the Carswell Formation (Ramaekers

et al., 2007) (Figure 3.1). It is worth mentioning that volcaniclastic rocks with a zircon U-

Pb age of 1644 ± 13 Ma (Rainbird et al., 2007) are present in the Wolverine Point

Formation, and basalt is present in the Kuungmi Formation of the Thelon Basin, a

geological counterpart of the Athabasca Basin (see Figure 3.1 for location), and has

yielded a U-Pb baddeleyite age of 1540 ± 30 Ma (Chamberlain et al., 2010). This is close

to the Re-Os isochron age of 1541 ± 13 Ma for carbonaceous shales in the Douglas

Page 79: Geometric and hydrodynamic modelling of fluid-structural

65

Formation of the Athabasca Basin (Creaser and Stasiuk, 2007), suggesting that both the

Athabasca and Thelon basins experienced some thermal events during their evolution

history.

Rocks of the Athabasca Group and its basement are cut by the northwest-trending

McKenzie diabase dykes dated at 1267 ± 2 Ma (LeCheminant and Heaman, 1989). The

eastern Hearne Craton basement and the Athabasca Group are intruded by another suite

of olivine diabasic and gabbroic rocks, belonging to the 1.11 Ga Moore Lakes Complex,

along the southeastern perimeter of the basin (Figure 3.1; MacDougall and William, 1993;

MacDougall and Heaman, 2002).

3.2.2 Uranium deposits and structures

Uranium deposits in the Athabasca Basin are generally classified into two groups:

sandstone-hosted and basement-hosted (Fayek and Kyser, 1997; Jefferson et al., 2007).

The sandstone-hosted deposits primarily occur in the conglomerate and sandstone of the

Athabasca Group immediately above the unconformity (e.g., Cigar Lake deposit; Figure

3.2A) (Hoeve and Quirt, 1984). These sandstone-hosted deposits (also known as ‘egress’

type) are characterized by flattened, elongate orebodies, and by polymetallic (complex)

ores containing significant amounts of Ni, Co, Cu, Pb, Zn, and Mo (Jefferson et al., 2007).

The basement-hosted deposits (e.g., Eagle Point deposit; Figure 3.2C), also known as

‘ingress’ type (Hoeve and Quirt 1984; Fayek and Kyser 1997; Jefferson et al., 2007), are

generally controlled by steeply to moderately dipping brittle fault zones, and the uranium

mineralization is generally monometallic (simple) containing mainly U and Cu (Jefferson

et al., 2007). However, many deposits appear to be a hybrid of the two groups, containing

Page 80: Geometric and hydrodynamic modelling of fluid-structural

66

both sandstone-hosted ore and basement-hosted ore (e.g., Key Lake deposit; Figure 3.2B)

(Jefferson et al., 2007). A ca. 1590 Ma age was proposed to represent the most important

primary uranium mineralization event, and other ages were taken to represent

remobilization events (Alexandre et al., 2009). However, it cannot be ruled out that some

of the young ages represent new mineralization event rather than remobilization.

The Athabasca Basin and underlying basement rocks are cut by numerous late

brittle faults of variable orientation commonly representing reactivated pre-Athabasca

ductile basement structures (e.g., McGill et al., 1993; Card et al., 2007a, b; Jefferson et al.,

2007; Tourigny et al., 2007). The faults are generally rooted in sheared graphitic pelitic

basement rocks and offset the unconformity and overlying sandstone on the order of tens

up to hundreds of metres, commonly reflected by abrupt change in the unconformity

topography, and likely have served as conduits focusing mineralizing fluids (Jefferson et

al., 2007; Li et al., 2015a). Recent analysis of fault systems using aeromagnetic data and

3D modelling, suggests that the northeast-trending faults represent reactivated Hudsonian

structures and may be the principal control on basement topography and uranium

mineralization in the southeastern Athabasca Basin (Li et al., 2015a). In fact, the close

association of the uranium deposits with faulting has been observed at different deposits

(Figure 3.2). For instance, the orebody in the Deilmann pit at Key Lake is mainly

controlled by the Key Lake fault, an east-northeast-trending reverse structure (Harvey

and Bethune, 2007) (Figure 3.2B), whereas the ore zones at Eagle Point deposit are

related to the northeast-trending graphite-bearing Collins Bay thrust and Eagle Point fault

(Marlatt et al., 1992) (Figure 3.2C). The ore-controlling structures are typically

enveloped by alteration haloes (Figure 3.2), mainly illite, chlorite, dravite, kaolinite,

Page 81: Geometric and hydrodynamic modelling of fluid-structural

67

silicification or desilicification, with a dimension of up to 400 m wide at the basal

unconformity and a thousand metres in strike length (Hoeve and Quirt, 1984).

3.3 Study methods and inputs

3.3.1 Numerical modelling background

Due to the large scale (hundreds of kilometres) and long duration (on order of

millions of years) of the fluid flow processes involved in mineralization, it is impossible

to physically model them. Computational modelling has provided an effective way to test

the factors related to mineralization and has been successfully used in a variety of studies

(e.g., Chi and Xue, 2011, and references therein; Ingebritsen and Appold, 2012, and

references therein; Zhao et al., 2012). Computational modelling uses numerical methods

to solve a series of partial differential equations pertaining to a geoscience system (Zhao

et al., 2009, 2012). In this investigation FLAC3D (Fast Lagrangian Analysis of Continua

3D version; Itasca 2012) was chosen to examine the effects of basement faults on

thermal-driven fluid convection. FLAC3D is a three-dimensional (3D) explicit finite

difference code designed by Itasca for simulating the interaction of fluid flow,

deformation, and heat transport of a continuous porous medium (Itasca, 2012). It has

been applied successfully to, for example, the fields of structural geology and fluid flow

related to various types of deposits (e.g., Oliver et al., 1999, 2006; McLellan et al., 2004;

Vos et al., 2007; Feltrin et al., 2009; Sheldon, 2009; Zhang et al., 2011; Cui et al., 2012b;

Upton and Craw, 2014; Li et al., 2015b).

As explained in the introduction, the purpose of this study is to examine the effect

of faults on thermal convection without invoking rock deformation in the fault zones. We

Page 82: Geometric and hydrodynamic modelling of fluid-structural

68

only consider free convection related to density differences due to temperature variations.

The theory of free thermal convection in porous media has been reviewed by many

people (e.g., Kühn et al., 2006; Oliver et al., 2006; Turcotte and Schubert, 2014). The

occurrence of free thermal Rayleigh-Benard convection in fluid saturated, homogeneous,

isotropic, horizontal porous media can be estimated with the dimensionless Rayleigh

number (Ra) (Turcotte and Schubert, 2014), which is defined as

Ra=𝜌𝑓

2gα𝑐𝑝𝑓kH∆T

μλ𝑚 (3.1)

Where 𝜌𝑓 is the density of liquid water (kg/m3), g is the gravitational acceleration (m/s2),

α is the thermal expansion coefficient of water (K-1), 𝑐𝑝𝑓 is the isochoric heat capacity

(J/(kg ∙ K)), k is the permeability (m2), H is the characteristic height of the system (m),

∆T is the temperature difference across this height (K), μ is the dynamic viscosity of

water (kg/(m ∙ s)), and λ𝑚 is the thermal conductivity of the porous medium (J/(s ∙ m ∙

K)).

Ra represents the ratio of the buoyant forces, which drives convective fluid flow,

to the viscous forces inhibiting fluid movement, and can be used to measure the tendency

toward free convection (Ingebritsen and Sanford, 2006). When Ra exceeds a critical

value, Rayleigh convection occurs (Ingebritsen and Sanford, 2006). The critical Ra, and

the corresponding critical wave number (related to the width of a convection cell, as

explained in Turcotte and Schubert, 2014), are sensitive to the boundary conditions

exerted on the porous medium (Simms and Garven, 2004). The classical Ra for an

infinite horizontal layer with impermeable boundaries is 4π2 (Lapwood, 1948; Turcotte

and Schubert, 2014) and in an ideal case the number of convection cells can be calculated

from Ra and the length of the system based on analytic solution (Turcotte and Schubert,

Page 83: Geometric and hydrodynamic modelling of fluid-structural

69

2014). Some studies, however, have suggested that convection may occur at relatively

low Ra values due to complex geological geometries, variable thermal and fluid

conditions, and variable-viscosity and non-Boussinesq effects (e.g., Straus and Schubert,

1977; Raffensperger and Garven, 1995; Raffensperger and Vlassopoulos, 1999; Kühn et

al., 2006) ; in these cases, due to the complicated model configuration and physical

properties the number and width of convection cells may only be calculated via

numerical solution (Simms and Garven, 2004).

Thermally-driven free convection for porous media is governed by a series of

partial differential equations, including Darcy’s law, Fourier’s law, mass conservation

and energy conservation (Itasca, 2012), as described in some published studies (e.g.,

Kühn et al., 2006; Oliver et al., 2006). The fluid density 𝜌𝑤 is related to temperature

changes by the linear equation of Boussinesq approximation

𝜌𝑤 = 𝜌0[1 − 𝛽𝑓(𝑇 − 𝑇0)] (3.2)

where 𝛽𝑓 is the thermal volumetric expansion of the fluid (°C-1), and 𝑇0 is the reference

temperature (°C).

Besides temperature, the density of the pore fluid can also be affected by salinity,

which in turn affects the buoyance force driving fluid flow (Batzle and Wang, 1992).

Therefore, salinity plays an important role in controlling fluid migration (e.g., Yang et al.,

2004a; Koziy et al., 2009; Park et al., 2009). However, considering the uncertainties of

salinity distribution at the time of mineralization in the Athabasca Basin (e.g., Richard et

al., 2015; Chu and Chi, 2016), we have elected to exclude the effects of salinity and focus

on pure thermal convection in order to reach more general conclusions about the

Page 84: Geometric and hydrodynamic modelling of fluid-structural

70

relationships between thermal convection and basement faults without creating

unnecessary ambiguities.

The temperature dependence of fluid viscosity was incorporated in this study,

with an empirical equation linking dynamic viscosity 𝜇𝑤 with temperature (Kestin et al.,

1978; Rabinowicz et al., 1998), as follows:

𝜇𝑤 = 2.414 × 10−5 × 10(247.8

𝑇−140) (3.3)

where T is temperature in degrees K and 𝜇𝑤 in Pa∙s.

FLAC3D solves the coupled fluid flow and heat transfer equations on a Cartesian

grid, with an iterative process involving sequential stepping of fluid and thermal modules

(Itasca, 2012). The initial temperatures for a coupled simulation are established by

running a pure thermal conduction calculation. These initial temperatures are then used to

calculate the fluid densities and viscosities involved in the fluid flow equation, which is

computed for fluid velocity. The fluid flow velocity values are then used to compute the

temperature distribution by solving the thermal energy conservation equation during

thermal calculation, and the derived temperatures are used to calculate the fluid flow

velocities once again.

3.3.2 Physical models

The relatively mud-rich and fine-grained Wolverine Point Formation in the

Athabasca Basin generally has been considered to be a physical barrier (an aquitard) to

fluid flow (Hiatt and Kyser, 2007; Jefferson et al., 2007). Our preliminary numerical

modelling (Li et al., 2015b) has confirmed that when a low permeability was assigned to

the Wolverine Point Formation, the convection cells are only developed in the relatively

Page 85: Geometric and hydrodynamic modelling of fluid-structural

71

more permeable lower part of the basin. This permeable aquifer, confined by the

overlying Wolverine Point Formation and by the metamorphosed basement below,

consists of quartz-rich sandstone with subordinate conglomerate with a maximum

aggregate thickness of about 1.5 km (Ramaekers et al., 2007). As discussed earlier, the

thickness of the eroded strata is uncertain, and therefore, as shown in Figure 3.3, we

arbitrarily assigned a thickness of 1.5 km to the basal aquifer, a thickness of 4.5 km to the

overlying aquitard (including the Wolverine Point Formation and the strata above it), and

a thermal gradient of 35 °C/km to the physical model. This configuration is consistent

with the burial model of Pagel (1975) and that used by many other models (e.g., Cui et al.,

2012a, b; Chi et al., 2013, 2014). However, in order to focus on the mineralization zones

around the unconformity, the topmost 3 km was not included in the modelling (Figure

3.3). Instead, the top of the physical model was assigned a temperature of 125 °C based

on a thermal gradient of 35 °C/km and a surface temperature of 20 °C. The

metamorphosed basement below the permeable sandstone layer has an arbitrary thickness

of 2 km. The model has a horizontal dimension of 10 km. Although FLAC3D is capable

of dealing with fluid flow and heat transport in porous media in 3D, we are considering

only 2D in this study, i.e., the physical model has only one zone (element) in the out-of-

plane direction due to the limit of computation time.

Page 86: Geometric and hydrodynamic modelling of fluid-structural

72

Table 3.1 Input parameters of various hydrological units used in model setups (after Oliver et al., 2006; Cui

et al., 2012b).

* Some of these parameters were adjusted in some models, see text for details.

Units /

properties

Density

(kg/m3) Porosity

Permeability

(m2)

Thermal

conductivity

(W/(m ∙ °C))

Thermal

expansivity

(°C–1)

Specific heat capacity

(J/(kg ∙ °C))

Cover 2,400 0.1 1.0×10-17 2.5 8.0×10-6 803.0

Sandstone 2,500 0.2 5.0×10-14 3.5 10.0×10-6 803.0

Basement 2,650 0.1 1.0×10-17 * 2.5 8.0×10-6 803.0

Fault zone 2,400 0.2 1.0×10-13 4.0 * 10.0×10-6 803.0

Water 1,000 0.6 1.85×10-3 4,185.0

Page 87: Geometric and hydrodynamic modelling of fluid-structural

73

Table 3.1 shows the physical properties of the basin sediments, fault zones and

basement rocks, including density, porosity, permeability and thermal properties, which

were assigned to each element of the model grid. These values were adopted from those

used in similar numerical modelling and published compilations (e.g., Raffensperger and

Garven, 1995; McLellan et al., 2004; Yang et al., 2004b; Oliver et al., 2006; Cui et al.,

2012a, b; Chi et al., 2013, 2014). As in previous studies, the rock properties are assumed

to be homogeneous and isotropic within individual units (e.g., Kühn et al., 2006; Oliver

et al., 2006; Cui et al., 2012b). Such an approach was adopted in consideration of the

computation time required, the uncertainties about what natural heterogeneities and

anisotropies may exist in various rock units, and the observation that the flow pattern and

temperature distribution in stratigraphic units with homogeneous permeabilities are not

significantly different from those with heterogeneous permeabilities (but the same

average permeability) (Kühn et al., 2006).

A high permeability of 5.0×10-14 m2 was assigned to the aquifer, and a

significantly lower value (1.0×10-17 m2) was assigned to the cover aquitard. The

permeability of continental crystalline rocks generally varies widely, and generally

decreases with depth (e.g., Manning and Ingebritsen, 1999; Stober and Bucher, 2007).

Considering the small thickness of the basement used in this study, a fixed low

permeability (1.0×10-17 m2) was used in most models. However, we have also considered

scenarios with increased permeabilities for the upper part of the basement to account for

the possible effects of paleo-weathering and fracture abundance on permeability (Hoeve

and Quirt, 1984; Macdonald, 1985).

Page 88: Geometric and hydrodynamic modelling of fluid-structural

74

Figure 3.3 Cross-sectional view of the physical model forming the foundation of all numerical modelling

carried out in this study. Original in colour.

Page 89: Geometric and hydrodynamic modelling of fluid-structural

75

Faults are generally considered one of the most important fluid conduits for

hydrothermal ore-forming fluids, implying that they have relatively high permeabilities

(e.g., Cox, 2005). For a typical fault, it generally has a narrow, low-permeability core

zone that is flanked by a wider, high-permeability damage zone (Caine et al., 1996). The

close association of unconformity-related uranium deposits with reactivated basement

faults supports the concept that the faults served as the fluid conduits (Jefferson et al.,

2007). Therefore, the fault zones in our models were all treated as homogeneous zones

with high permeabilities (Table 3.1) as in many other studies (e.g., Yang et al., 2004b;

Oliver et al., 2006; Cui et al., 2012b). The dimensions of different ore-bounding fault

zones, which formed as a result of brittle reactivation of older syn- to late-Hudsonian

structures in the Athabasca Basin, vary significantly depending on the local tectonic

setting (Jefferson et al., 2007), ranging from tens of metres to hundreds of metres in

width (e.g., the orebody in the Cigar Lake deposit has a maximum width of 100 m,

Bruneton, 1993) and up to 2 kilometres in depth (e.g., the P2 fault in the McArthur River

deposit, Hajnal et al., 2007). A representative dimension of 100 m wide by 400 m high

was therefore used in this study. In addition, fault offset of the unconformity surface is

common in many deposits (e.g., McGill et al., 1993; Harvey and Bethune, 2007; Kerr,

2010), with displacements on the order of tens to hundreds of meters (Jefferson et al.,

2007); however, this phenomenon is not universal, for example, there is little evidence of

unconformity offset along graphitic basement units that underlie the ore zone in the Cigar

Lake deposit (Andrade, 2002). Whatever the case, in this study the effect of fault offsets

on fluid convection was not considered as only thermally-driven fluid convection, and

not deformation-driven flow, was modelled.

Page 90: Geometric and hydrodynamic modelling of fluid-structural

76

Many of the ore-controlling faults in the Athabasca Basin contain abundant

graphite, which likely gave rise to higher thermal conductivities than most basement

rocks (an average of 2.5 W/(m ∙ °C) for basement rocks; Fowler et al., 2005), and it was

proposed that the high heat conductivity of graphitic faults may have been responsible for

fluid convection related to uranium mineralization (Hoeve and Quirt, 1984). Therefore, in

our models we have also assigned a relatively high conductivity value to the fault zones

to account for the presence of graphite (Table 3.1). The thermal conductivities in Table

3.1 are mainly adapted from Raffensperger and Garven (1995) and Cui et al. (2012b).

In summary, our physical models consist of a high permeability aquifer

sandwiched between a low permeability cover and a low permeability basement, with

faults of high permeability and heat conductivity straddling the unconformity (Figure 3.3).

Different combinations (scenarios) of permeability, heat conductivity and fault spacing

and orientation (Table 3.2) were modelled to evaluate the factors that affect the fluid-flow

pattern.

Page 91: Geometric and hydrodynamic modelling of fluid-structural

77

Table 3.2 Various scenarios conducted in this study.

Scenarios Number of fault

zones incorporated

in each scenario

Description

1 0 Base model; Figure 3.4

2 1 One vertical fault zone, conduction only; Figure 3.5

3 1 One vertical fault zone; Figure 3.6

4 1 One inclined fault zone (dipping 45°, 60° respectively); Figure

3.7

5 2 Two vertical fault zones with various spacing; Figures 3.8 and

3.9

6 2 Two inclined fault zones (45°) with a fixed spacing; Figures

3.10 and 3.11

7 2 Like scenario 5, but one fault zone has a higher thermal

conductivity, other one has a same thermal conductivity as the

sandstone; Figure 3.12

8 0, 1, 2 Increase of the permeability of the uppermost basement; Figure

3.13

9 2 Like scenario 8, but one fault zone has a higher thermal

conductivity, other one has a same thermal conductivity as the

sandstone; Figure 3.14

10 0, 1, 2 Change of the thermal gradient

Page 92: Geometric and hydrodynamic modelling of fluid-structural

78

3.3.3 Initial and boundary conditions

Numerical modelling of sediment compaction as well as hydrocarbon generation

(Chi et al., 2013, 2014) suggests that the fluid pressure within the basin was near the

hydrostatic regime during sedimentation. As such, hydrostatic pressure was used as the

starting condition of the system, as was also done by Raffensperger and Garven (1995)

and Cui et al. (2012a, b). For fluid transport, the initial fluid velocity was set to zero over

the whole simulation domain and all the boundaries (sides, top and bottom) were

assumed to be impermeable to fluid flow.

The temperature at the top of the model (3 km deep) was maintained at 125 °C,

and that at the bottom (8 km deep) was fixed at 300 °C. The two side boundaries were

configured to be insulated to heat transport. The initial temperature distribution was

determined assuming purely conductive heat transport and was calculated prior to each

model run.

3.4 Modelling results

3.4.1 Scenario 1 – convection with no fault (base model)

The first model consists of the three hydrostratigraphic units as described above

and does not have a fault. It serves as a reference for the other scenarios to test the effects

of faults on fluid convection, and is referred to hereafter as the “base model”. All the

parameters are the same as described in Table 3.1, and the initial and boundary conditions

are the same as presented in the previous section. These physical parameters and

conditions remain unchanged in most of the following models unless otherwise specified.

Page 93: Geometric and hydrodynamic modelling of fluid-structural

79

Two observation points (Figure 3.4A) were placed in the model to monitor the

change of temperature, which was used to determine if the thermal convection system

reaches the steady state. The model demonstrates that free convection takes about 0.5

million years to approach steady-state (Figure 3.4B) and that at this stage seven regularly

spaced convection cells are developed in the permeable sandstone layer, each with an

approximately equal width of 1.4 km (Figure 3.4A). The temperature field is

characterized by alternating patterns of warm and cool thermal anomalies, associated

with the upwelling and downwelling centres, respectively (Figure 3.4A). Here and in the

following models, the temperature profile is used as an indicator to demonstrate the

degree of convection in each specific model. Direction of the fluid flow is shown by the

arrows and the intensity of fluid flow is reflected by the length of the arrows. Fluid

velocities are greatest at the boundaries between the sandstone and the confining layers

(cover and basement), as well as in the middle centre of upwelling and downwelling

zones, with a maximum value of 0.15 m/yr (Figure 3.4A). In an individual convection

cell, the velocity decreases gradually towards the centre of the cell, located approximately

in the middle of the sandstone sequence. The convective circulation does not penetrate

into the underlying basement, however, the temperature field of the basement is affected

by the overlying convection cells and the temperature field in the sandstone (Figure 3.4A).

The same situation is also observed for the cover. Fluid velocities in the basement and in

the cover are 6×10-6 m/yr or less, which are not shown in all the figures due to scaling

factors.

Page 94: Geometric and hydrodynamic modelling of fluid-structural

80

Figure 3.4 Modelling result for Scenario 1 showing fluid flow patterns and related thermal field in the base

model (A). Variations of measured temperatures (solid lines) with time at each observation point. OP =

Observation Point (B). Original in colour.

Page 95: Geometric and hydrodynamic modelling of fluid-structural

81

3.4.2 Scenario 2 – pure conduction with various conductivities for the fault

The models in this scenario have the same geometry as the base model, the only

difference being that one vertical fault zone (100 m by 400 m) straddling the

unconformity was introduced in the central part of the model (Figure 3.5). Previous

modelling conducted by Raffensperger and Garven (1995) suggests that the existence of

conductive graphitic zones (1-5 km wide) does not affect the regional flow patterns at the

basin scale (~800 km long); however, it is not clear whether at a much smaller scale

(relative to the basin as a whole) the graphitic fault zones are conductive enough to create

thermal anomalies and thus affect fluid-flow patterns at the local scale. Modelling by

simple heat conduction (no fluid flow) was therefore carried out to test if a significant

heat anomaly can be created by increasing the conductivity of the fault without changing

the conductivities of the other units. Three cases were tested with successively higher

thermal conductivities for the fault but holding all of the other parameters constant.

Using the thermal conductivity values listed in Table 3.1, i.e., 4.0 W °C-1 m-1 for

the fault, 3.5 W °C-1 m-1 for the aquifer, and 2.5 W °C-1 m-1 for the aquitard and basement,

one can see (Figure 3.5A) that the isotherms near the fault above the unconformity bend

slightly upwards, whereas those near the fault below the unconformity bend slightly

downwards. The temperature disturbance above and below the fault is less than 1 °C

compared to that of the base model. When the thermal conductivity of the fault is

increased to 7 W °C-1 m-1 (Figure 3.5B), which is twice as large as that of the sandstone,

the thermal disturbance above and below the fault is raised to about 1 °C, and the overall

thermal field is similar to the previous case. If the thermal conductivity of the fault is

further increased to 35 W °C-1 m-1 (Figure 3.5C), which is 10 times that of the

Page 96: Geometric and hydrodynamic modelling of fluid-structural

82

surrounding rocks, the temperature disturbance above and below the fault is increased to

about 3 °C compared to that of the base model.

These results indicate that the thermal disturbances caused by high heat

conduction along graphitic faults are likely relatively small. The small thermal anomalies

created by such increased local heat conductivities may nevertheless be sufficient to

stimulate convection, which in turn may produce more significant thermal disturbances

(See also Yang et al., 2004b).

Page 97: Geometric and hydrodynamic modelling of fluid-structural

83

Figure 3.5 Modelling results for Scenario 2 showing the initial temperature distribution when the thermal

conductivity of the fault zone is equal to (A) 4 W °C-1 m-1; (B) 7 W °C-1 m-1; (C) 35 W °C-1 m-1. Heat

transfers by simple conduction. Note that the thermal anomaly developed adjacent to the fault zone became

more pronounced in each case. See text for details.

Page 98: Geometric and hydrodynamic modelling of fluid-structural

84

3.4.3 Scenario 3 – convection with one vertical fault

Two simulations were carried out to examine the behavior of convection cells in

response to one vertical fault at different locations in the model. In the first simulation,

the fault was placed at 4.3 km from the left boundary of the model (Figure 3.6A). The

results indicate that, after the system reaches the steady state, the temperature profile and

fluid flow pattern are broadly similar to that in the base model (compare Figure 3.4A and

Figure 3.6A). The maximum fluid velocity is 0.16 m/yr, slightly greater than that in the

base model (0.15 m/yr), probably due to the presence of the more permeable fault. The

size of the convection cells (~1.4 km in average) is also the same as in the base model

and therefore unaffected by the incorporation of the fault. However, the location of the

fault, which used to be centred on a downwelling plume between convection cells in the

base model (Figure 3.4A, below point OP1), is now the site of up-flow (Figure 3.6A).

Consequently, the fluid flow directions in all the other convection cells are reversed. For

example, the upward flow on the left side of base model (Figure 3.4A) has become a

locus of downward flow (Figure 3.6A).

In the second simulation (Figure 3.6B), the fault is moved to the right for a

distance of the width of one convection cell (i.e., located at 5.7 km from the left

boundary). The modelling results show that the location of the fault, which used to be the

site of descending flow in the first simulation (Figure 3.6A), becomes the site of

upwelling flow. In other words, when the fault is moved from the left to the right, the

upwelling plume shifts accordingly. The results also indicate that the size of the cells

does not change.

Page 99: Geometric and hydrodynamic modelling of fluid-structural

85

The association of the fault zone with upward flow is likely related to the heat

anomaly induced by the higher thermal conductivity of the fault zone, as demonstrated by

simulations in scenario 2 (Figure 3.5). In fact, this type of initial temperature distribution

gives rise to up-flow immediately above the fault zone (e.g., Yang et al., 2004b). As a

result, the fault zone becomes the focus of egress flow, and no fluid flows from the basin

into the basement (Figure 3.6C).

Page 100: Geometric and hydrodynamic modelling of fluid-structural

86

Figure 3.6 Modelling results for Scenario 3 showing fluid convection patterns and related thermal fields,

(A) with the presence of one fault zone located at 4.2 km from the left boundary, coincide with a

downwelling plume in the base model (Figure 3.5A); (B) when the fault zone is moved to the right for a

distance of 1.4 km corresponding to the width of one convection cell in the base model (Figure 3.5A). (C)

The expanded view of the model at locations in A and B showing that the fluid flow migrates upward from

basement along the fault zones. Original in colour.

Page 101: Geometric and hydrodynamic modelling of fluid-structural

87

3.4.4 Scenario 4 – convection with one inclined fault

Two simulations incorporating a fault zone with dip angles of 45° and 60°,

respectively, were conducted to examine whether or not and how the attitude of the fault

zone may influence the convection patterns. In both simulations the fault zone is located

4.3 km from the left boundary (Figures 3.7A and B) for comparison with the model with

a vertical fault in Scenario 3 (Figure 3.6A). Other parameters are exactly the same as in

Scenario 3.

Comparing Figure 3.7A and Figure 3.7B with Figure 3.6A, respectively, indicates

that the temperature profile, fluid flow patterns and distribution of convection cells are

almost the same as the ones with the vertical fault. In all simulations, the maximum fluid

velocity is also around 0.16 m/yr. In addition, the location of the fault zone, regardless of

its dip angle, coincides with an upwelling plume (Figures 3.7A and B), as in the vertical-

fault model. The sizes of the convection cells are also the same as in the base model and

the vertical-fault model.

Therefore, the simulation results suggest that the dip angle of the fault zone does

not affect the overall flow pattern. However, unlike the vertical-fault model where the

fluid flow is parallel to the fault, the fluid flow in the inclined-fault models are oblique to

the fault (Figures 3.7C and D), although the overall direction of flow is still upward, i.e.,

egress, and no fluid flows into the basement.

3.4.5 Scenario 5 – convection with two vertical faults

This scenario is similar to scenario 3, except that two vertical faults instead of one

were incorporated into the model. The purpose is to investigate how the spacing of two

Page 102: Geometric and hydrodynamic modelling of fluid-structural

88

fault zones affects fluid convection and if any fluid penetrates into the basement along

fault zones. Three spacings were tested: 1.4 km or the width of one convection cell in the

base model (Figure 3.8A), 2.1 km or 1.5 times the width of a convection cell (Figure

3.8B), and 2.8 km or double the width of a convection cell (Figure 3.8C). To facilitate

comparison between simulations, the fault zone on the left was fixed at 4.3 km from the

left-hand boundary, whereas the fault zone on the right was shifted according to the

spacing assigned above (Figure 3.8).

The simulations indicate that for each of the fault spacing configurations, fluid

flows upward along each fault zone at the beginning of the simulation, which is probably

related to the heat anomaly induced by the relatively high heat conductivity of the fault

zones, as is also observed in the single-fault models. With continuation of the modelling

toward steady state, however, the fluid flow patterns changed depending on the spacing

of the fault zones.

Where the spacing of the two fault zones is equal to the width of a convection cell

in the base model (Figure 3.8A), 6 convection cells instead of 7 are developed in the

model, with an average width of 1.7 km rather than 1.4 km as in the base model and the

single-fault models. It is noticeable that this increase in width is slightly more

pronounced for the two convection cells above the two faults, although the widths of the

other cells are also affected (Figure 3.8A). Additional simulations with the model being

extended laterally indicate that the convection cells away from the two faults maintain the

same width as in the base model, i.e., 1.4 km, but the two convection cells immediately

above the faults remain enlarged as in Figure 3.8A. Unlike the single-fault model, the

faults in this model are overlain by those parts of two convection cells with horizontal

Page 103: Geometric and hydrodynamic modelling of fluid-structural

89

flow rather than coinciding with individual upwelling plumes at the boundary between

cells (Figure 3.8A), and this situation does not change with the model boundaries being

extended laterally. This is probably because the two fault zones are too close to allow

development of two separate pairs of convection cells to accommodate upward flow

above the fault zones and a common downward flow between the fault zones, and yet

they are too far from each other to accommodate a common upward flow.

When the spacing between the two fault zones is increased to 2.1 km (1.5 times

the width of a convection cell), seven convection cells are developed in the model (Figure

3.8B) as in the base and single-fault models. Unlike the 1.4-km spacing model (Figure

3.8A), the two fault zones in this model are close to (but not directly underneath) two

separate upwelling plumes, and there is a common downwelling flow roughly halfway

between the two fault zones (Figure 3.8B).

When the spacing between the two fault zones is increased to 2.8 km (double the

width of a convection cell), seven convection cells are developed in the model (Figure

3.8C). The two faults coincide with two separate upwelling plumes, again with a common

downwelling plume between them (Figure 3.8C).

In the last model described above, only egress flow was observed in the fault

zones (Figures 3.8C and 3.9C) as in the single-fault models. In the other two models

(Figures 3.8A and B), however, minor amounts of fluid flow into and out of the basement

along the fault zones (Figures 3.9A and B). There also appears to be a convective

circulation inside each fault zone with down-flow on one side and up-flow on the other

side (Figures 3.9A and B). The lateral flow across the fault zone in the sandstone is split

into two parts: one continues to travel in the sandstone, and the other follows the fault

Page 104: Geometric and hydrodynamic modelling of fluid-structural

90

zone into the basement. The fluid flowing into the fault zone eventually flows out from

the same fault zone because the low permeability of the basement prevents the fluid from

going down further.

Additional simulations with various fault spacings indicate that when the ratio of

the spacing of the two fault zones to the width of the cell is less than 1, the two fault

zones will be overlain by two separate convection cells or share one upwelling plume,

whereas when the ratio of the spacing of the two fault zones to the width of the cell

ranges between 1 and 2, the two fault zones are close to two separate upwelling plumes,

with a downwelling plume situated between them. In either case, minor convective

circulation can occur inside each fault zone. In all simulations, the maximum fluid

velocity is around 0.16 m/yr.

Page 105: Geometric and hydrodynamic modelling of fluid-structural

91

Figure 3.7 Modelling results for Scenario 4 showing fluid convection patterns and related thermal fields

with the presence of one 45° dipping fault (A) and one 60° dipping fault (B), both situated at 4.2 km from

the left boundary. Corresponding view of more detailed flow vectors around the fault (C and D). Original in

colour.

Page 106: Geometric and hydrodynamic modelling of fluid-structural

92

Figure 3.8 Modelling results for Scenario 5 showing fluid convection patterns and related thermal fields

with two vertical fault zones spaced at a distance of, (A) 1.4 km, equal to of the width of one convection

cell in the base model, (B) 2.1 km, equal to one and a half times the width of one convection cell in the

base model, (C) 2.8 km, two times the width of one convection cell in the base model. Original in colour.

Page 107: Geometric and hydrodynamic modelling of fluid-structural

93

Figure 3.9 Expanded views of the encompassing fault zone on the left-hand side of the models in Figures

3.8A – C, showing more detailed flow patterns. Original in colour.

Page 108: Geometric and hydrodynamic modelling of fluid-structural

94

3.4.6 Scenario 6 – convection with two inclined faults

Two inclined faults, with a dip angle of 45° and a spacing of 2.1 km, are

considered in this scenario. Three combinations of dip directions are considered, i.e.,

faults dipping towards each other (inwardly dipping; Figure 3.10A), faults dipping away

from each other (outwardly dipping; Figure 3.10B) and faults dipping parallel to one

another (Figure 3.10C), which are probably more representative of the fault attitudes in

the basement of the Athabasca Basin than the vertical-faults models.

Overall the thermal field and fluid patterns obtained from the three simulations

are similar to each other (Figure 3.10) and similar to those of the vertical-faults

counterpart. Each fault zone is close to the upwelling plume of two separate convection

cells, and the two fault zones share one downwelling plume, situated about halfway

between them (Figure 3.10). The flow patterns inside each of the fault zones are also

similar to the vertical-fault model, i.e., weak convective circulation is developed inside

each fault with down-flow on one side and up-flow on the other side. However, detailed

flow patterns differ among the three different dip combinations. Specifically, for the

inwardly-dipping model, the overall flow trend within the faults is upward despite the

internal convective flow (Figures 3.11A and B), whereas for the outwardly-dipping

model the overall flow is downward (Figures 3.11C and D). For the parallel-dipping

model, the left fault shows an overall ingress flow pattern (Figure 3.11E), whereas the

right one shows an egress flow pattern (Figure 3.11F). In all models, the maximum flow

velocity is close to each other (around 0.16 m/yr), with flow velocity larger inside the

fault zones within the sandstone than in the surrounding area due to the higher

permeability, which is similar to previous models (Figure 3.11).

Page 109: Geometric and hydrodynamic modelling of fluid-structural

95

Figure 3.10 Modelling results for Scenario 6 showing fluid convection patterns and thermal field with two

inclined fault zones spaced at a distance of 2.1 km (equal to 1.5 times the width of a convection cell in the

base model), (A) dipping inwardly, (B) dipping outwardly, (C) dipping in the same direction. See text for

explanation. Original in colour.

Page 110: Geometric and hydrodynamic modelling of fluid-structural

96

Figure 3.11 Expanded views of the corresponding fault in Figures 3.10A – C showing that weak convective

circulation occurs inside the fault zone, with one side going down and the other side going up. Note the

overall flow pattern in each fault, see text for explanation. Original in colour.

Page 111: Geometric and hydrodynamic modelling of fluid-structural

97

3.4.7 Scenario 7 – convection with two vertical faults with different thermal

conductivities

For all the models up till now, both of the fault zones are assigned equal thermal

conductivities that are slightly higher than the surrounding rock units. In the present

scenario, the left fault zone is assigned a higher thermal conductivity (4.0 W °C-1 m-1) as

was done in the previous models, whereas the right fault zone is given the same thermal

conductivity as the sandstone (3.5 W °C-1 m-1). The purpose is to examine the potential

fluid flow patterns in the case where one fault is graphitic (thus high thermal conductivity)

and a neighboring one is not. Three fault spacings, i.e., 1.4 km, 2.1 km and 2.8 km are

simulated as in scenario 5.

The modelling results (Figure 3.12) indicate that the overall relationships between

fluid patterns and the fault zones are significantly different from the models in which

both the two fault zones have the same thermal conductivities (Figure 3.8). Regardless of

the spacing of the two fault zones, the left fault zone with a higher thermal conductivity

always corresponds to an upward-flow centre (Figure 3.12). The right fault zone with a

lower thermal conductivity is roughly centred with a down-flow plume when the spacing

of the two fault zones is equal to one cell size (Figure 3.12A), or is within a convection

cell when the spacing of the two fault zones is one and half times of the width of a

convection cell (Figure 3.12B). In both cases, weak convective flow (not shown) is

observed inside the right fault zone which is similar to scenario 5 (see Figures 3.9A and

B). When the spacing is equal to twice the width of a convection cell, the right fault zone

is centred with an up-flow plume (Figure 3.12C), the same as the situation in which both

Page 112: Geometric and hydrodynamic modelling of fluid-structural

98

the fault zones have high thermal conductivities (Figure 3.8C). The maximum flow

velocity in each simulation is close to each other (around 0.16 m/yr).

3.4.8 Scenario 8 – convection with permeable upper basement and vertical faults of equal

thermal conductivities

This scenario considers the possibility that the upper part of the basement has a

significantly higher permeability than the rest of the basement. Such a possibility has

been invoked based on the presence of a regolith below the unconformity and

development of fault networks and microfractures (e.g., Jefferson et al., 2007; Mercadier

et al., 2010). The paleoweathered zone immediately below the Athabasca Group ranges

from a few centimetres up to 220 metres thick (Macdonald, 1980). In the models of this

scenario, a permeability of 5.0×10-16 m2, which is 50 times higher than that of the rest of

the basement (1.0×10-17 m2) but two orders of magnitude lower than the sandstone unit

(1.0×10-14 m2), was assigned to the uppermost 220 m of the basement. All the fault zones

incorporated in this scenario have a higher thermal conductivity (4.0 W °C-1 m-1) than the

surrounding rocks.

The first model without a fault zone (Figure 3.13A) reveals a general fluid flow

pattern and temperature profile similar to those in the base model (Figure 3.4A), with

development of 7 convection cells averaging 1.4 km in size (Figure 3.13A). However,

unlike the base model, the convection cells invade into the permeable basement (Figure

3.13A), due to the increased permeability, which is consistent with the models

Page 113: Geometric and hydrodynamic modelling of fluid-structural

99

Figure 3.12 Modelling results for Scenario 7 showing fluid convection patterns and thermal fields for

model setup same as Figure 3.8. The only difference from Figure 3.8 is that the left fault was assigned with

a higher thermal conductivity while the right one with a thermal conductivity the same as the sandstone.

Note that the left fault is always centred on an up-going flow. See text for explanation. Original in colour.

Page 114: Geometric and hydrodynamic modelling of fluid-structural

100

investigated by Oliver et al. (2006), although the overall fluid velocities in the permeable

basement are about one order of magnitude smaller (maximum 0.011 m/yr) than the ones

in the sandstone unit (maximum 0.15 m/yr).

When a single fault zone is introduced into the model, similar to previous

scenarios (Figures 3.6 and 3.7), the fault zone sets up convection cells with up-flow

immediately above the fault zone, due to the thermal anomaly induced by the fault.

Consequently, the fault zone corresponds to an upwelling flow centre after the system

reaches the steady state (Figure 3.13B). The flow pattern in the relatively permeable

basement is similar to the model without a fault zone (Figure 3.13A).

When two fault zones are introduced (Figure 3.13C), the fluid flow patterns are

similar to those without the permeable top in the basement (Figures 3.8 and 3.10), in

which each of the fault zones initially coincides with an upward plume, and then evolves

to be located under a convection cell or near an upward plume, depending on the spacing

of the fault zones (Figure 3.8). For instance, when the two fault zones are spaced by 1.4

km, they are under two separate convection cells (Figure 3.13C). Unlike previous

simulations without a permeable basement, in which a minor convection flow is

developed within the fault zone in the basement (Figures 3.9A and B), the fault zones in

this model allow fluid to pass through without creating a secondary convective flow in

the fault zones.

Page 115: Geometric and hydrodynamic modelling of fluid-structural

101

Figure 3.13 Modelling results for Scenario 8 showing that (A) convection cells can invade the relatively

permeable uppermost basement without the fault in the model; (B) single fault coinciding with an

upwelling centre in the model; (C) two faults separated by one cell width sharing one upwelling centre. See

text for more explanation. Original in colour.

Page 116: Geometric and hydrodynamic modelling of fluid-structural

102

3.4.9 Scenario 9 – convection with permeable upper basement and vertical faults of

different thermal conductivities

This scenario has the same configuration as scenario 8 (Figure 3.13C) except the

two fault zones have different thermal conductivities. The right fault zone has the same

thermal conductivity as the sandstone (3.5 W °C-1 m-1), whereas the left fault zone has a

higher thermal conductivity as in other models (4.0 W °C-1 m-1).

The modelling result (Figure 3.14) is distinctively different from the one in

Scenario 8 in which both fault zones have high thermal conductivities (Figure 3.13C).

One can see that the left fault zone with a higher thermal conductivity is centred with an

up-flow plume, whereas the right one coincides with a down-flow plume, and neither of

them has convective flow inside it (Figure 3.14). The fluid flow pattern is similar to

scenario 7 (Figure 3.12A), except for the presence of a permeable top layer of the

basement. Because of the permeable top of the basement, fluid can flow into the

basement from the right fault zone and out of the basement from the left one (Figure

3.14). This scenario is comparable to the fault-bounded convection cell model of Simms

and Garven (2004), in which the down-flow zone is centred on one fault and the up-flow

zone is located at the other fault in a sedimentary basin.

Additional simulations with various fault spacings indicate that the fault zone

with higher thermal conductivity generally corresponds to an up-flow centre, whereas the

fault zone with a thermal conductivity equal to that of the sandstone can correspond to a

down-flow centre when the spacing is equal to one-cell width, an up-flow centre when

the spacing is equal to two-cell width, or is located within a convection cell when the

spacing is in between one-cell width and two-cell widths.

Page 117: Geometric and hydrodynamic modelling of fluid-structural

103

Figure 3.14 Scenario 9. Same model geometry as Figure 3.13C but with the left fault assigned with a higher

thermal conductivity whereas the right one with a thermal conductivity same as sandstone. Note the

significant difference from Figure 3.13C that the left fault (higher thermal conductivity) coincides to an

upwelling centre whereas the right one is centred with down-going flow. See text for explanation. Original

in colour.

Page 118: Geometric and hydrodynamic modelling of fluid-structural

104

3.4.10 Scenario 10 – convection with elevated thermal gradients

This scenario is designed to investigate whether or not increasing the thermal

gradient can change the fluid flow patterns observed in the various scenarios described

above, considering the uncertainties in the maximum burial or the depth of mineralization

in the Athabasca Basin (Chi et al., 2015). The thermal gradient is increased from

35°C/km to 45°C/km, and the physical model is changed as follows: the thickness of the

top aquitard is reduced from 4.5 km to 3.5 km, whereas the thicknesses of the sandstone

layer and the basement remain 1.5 km and 2 km respectively. Similar to previous

scenarios, the topmost 3 km is not included in the modelling to focus on the

mineralization zones around the unconformity, and such that the top and bottom

boundaries are fixed at a temperature of 155 °C and 335 °C respectively based on a

thermal gradient of 45 °C/km and a surface temperature of 20 °C. Other parameters and

boundary conditions remain unchanged.

Various models similar to previous scenarios were tested, however, the modelling

results (not shown) indicate that the mode of fluid flow and the corresponding

temperature field are similar to the previous scenarios in which a thermal gradient of

35 °C/km was used. This is because the increase in the thermal gradient, which increases

the temperature difference for the effective thickness (i.e., the permeable sandstone below

the Wolverine Point Formation, 1.5 km), does not affect the Rayleigh number very much.

A similar relationship was also observed in the basin-wide modelling of the Thelon Basin

conducted by Cui et al. (2010). As noted previously, however, due to the existence of the

thermal anomaly induced by the fault zone, the fault zone still controls the initiation of

the convection system. In other words, the interactivity between fault zones and thermal

Page 119: Geometric and hydrodynamic modelling of fluid-structural

105

convection is the same as previous models with a lower thermal gradient. This implies

that changes in the geothermal gradient do not appear to fundamentally alter the overall

fluid flow patterns. However, the maximum flow velocity in these models is about 0.22

m/yr, slightly higher than those in previous models (0.16 m/yr).

3.5 Discussion

3.5.1 Fluid convection related to faults

The modelling results suggest that pervasive convective flow may have developed

in the more permeable lower sandstone sequence of the Athabasca Group, which is

consistent with previous studies (Raffensperger and Garven, 1995; Cui et al., 2010; Cui et

al., 2012b). It is suggested that thermal convection may have provided an effective way

to leach uranium (Hoeve and Quirt 1984; Raffensperger and Garven 1995; Borion et al.,

2010; Cui et al., 2012b) in the long period between the formation of the basin at ca. 1750

Ma and uranium mineralization at ca. 1590 Ma (Alexandre et al., 2009) or later.

Numerous studies have shown that various factors, including the model setup and

permeability, can affect the onset of the convection and the size of the convection cells in

a range of geological settings (e.g., Bächler et al., 2003; Simms and Garven, 2004; Yang

et al., 2004b; Kühn et al., 2006; Yang, 2006). In this study, various models with different

configurations of fault zones (location, spacing and orientation) that straddle the

unconformity between the permeable sandstone and relatively impermeable basement,

indicate that if other hydraulic conditions remain unchanged, the presence of fault zone(s)

can have significant effects on the patterns of the convective circulation, especially the

flow adjacent to the fault zone(s).

Page 120: Geometric and hydrodynamic modelling of fluid-structural

106

Our modelling results indicate that the slightly higher thermal conductivity

associated with a graphite-rich fault zone can induce a subtle heat anomaly, where the

temperature can be raised by 1 – 3 °C depending on the thermal conductivity. The

inference that temperature is raised in such zones is supported by the observation of

quartz dissolution and residual enrichment of clay within host rock alteration haloes, both

immediately below the unconformity and above it in the sandstone (Hoeve and Quirt,

1984). More importantly, although the heat anomaly is small, it is sufficient to promote

the onset of thermal convection in the more permeable sandstone. This suggests that the

graphite-rich units, as the sites of reactivated fault zones and key sources of reductant

(Jefferson et al., 2007), may have driven thermal convection which is critical to uranium

uptake and precipitation (Hoeve and Quirt, 1984; Boiron et al., 2010). On the other hand,

graphite may not be the only material in a fault zone that can raise the thermal

conductivity, which may explain why some deposits are not associated with graphitic

units, e.g., the Kiggavik deposits in the Thelon Basin and the Cluff Lake deposits in the

Athabasca Basin (Jefferson et al., 2007).

The modelling results also indicate that the fluid flow patterns can be substantially

affected by the spacing and thermal conductivities of the fault zones, but appear to be less

sensitive to the attitudes of the fault zones and changes in the geothermal gradient. In the

one-fault simulations, the sizes of the convection cells are not affected by the position

and dip angle of the fault zone, whereas for the models with two fault zones, when the

spacing between the two fault zones is less than two times of the width of a convection

cell, the widths of the convection cells close to the fault zones are slightly larger due to

Page 121: Geometric and hydrodynamic modelling of fluid-structural

107

the heterogeneity in physical properties resulting from the distribution of permeable fault

zones.

The incorporation of a relatively permeable basement gives rise to the invasion of

convection cells into the basement, which is in good agreement with previous studies

(Oliver et al., 2006; Cui et al., 2012a). When fault zones are introduced into the models,

the fluid flow patterns are similar to those without the permeable basement lid. In

addition, some fault configurations favour outward (egress) fluid flow, whereas others

favour inward (ingress) flow (Figure 3.14). More importantly, under certain

configurations, fluid can flow into a fault and out the same fault in a fluid convection

framework (Figures 3.9 and 3.11). These different scenarios appear to represent an

effective way to exchange fluids between the basin and the basement (e.g., Hoeve and

Sibbald 1978), or to drive basinal fluid into the basement and leach uranium through

fluid-rock interaction (e.g., Derome et al., 2005; Boiron et al., 2010; Richard et al., 2010).

3.5.2 Implications for the formation of unconformity-related deposits

As discussed in the introduction, although the origin of the uranium is still under

debate, the generally accepted generic model is that uranium was transported by

oxidizing, basinal fluids, which promoted uranium dissolution. These fluids subsequently

made their way to the unconformity where they interacted with reducing fluids coming

out of reactivated basement shear zones, or directly with basement rocks, and uranium

precipitated as uraninite (e.g., Hoeve and Sibbald, 1978; Hoeve and Quirt, 1984). Stable

isotopic compositions of H and O for fluids associated with pre- and syn-ore alteration

minerals both in the unconformity-hosted and basement-hosted deposits, as well as 87Sr/

Page 122: Geometric and hydrodynamic modelling of fluid-structural

108

86Sr ratios of chlorite, illite and dravite in the basin and basement rocks (Kotzer and

Kyser, 1995), suggest a mixing of isotopically- and chemically-distinct basinal and

basement fluids (Wilson and Kyser, 1987; Kotzer and Kyser, 1995). More recently,

analysis of individual fluid inclusions based on microthermometry, Raman spectroscopy

and laser-induced breakdown spectroscopy (LIBS) (Derome et al., 2005) and laser

ablation–inductively coupled plasma–mass spectrometry (LA-ICP-MS) (Richard et al.,

2010), suggests mixing between a Na-rich brine and a Ca-rich brine, which both likely

originate from evaporated seawater, at the sites of uranium mineralization (Richard et al.,

2011, 2013). Thermal convection has been proposed to be responsible for the penetration

of the Na-rich basinal brine into basement (Boiron et al., 2010), where it interacts with

Ca-rich rocks and gains U and Ca becoming a Ca-rich brine (Derome et al., 2005;

Richard et al., 2010); uranium precipitation accompanies the mixing of brines in the

vicinity of major faults (Richard et al., 2010).

Our numerical modelling provides a conceivable explanation for the circulation of

basinal fluid, its penetration into basement, and uptake of uranium from uranium-rich

basement lithologies. The modelling results suggest that during periods of tectonic

quiescence, thermal convection may be the dominant mechanism driving fluid flow, and

the permeable and thermoconductive fault zones may play a critical role in the ore-

forming fluid circulation system. The oxidizing uraniferous basinal fluids may migrate

down into the basement along the fault zones, where they may interact with reducing

lithologies and/or fluids, precipitating uranium and forming the basement-hosted deposits

(location A in Figure 3.15). On the other hand, if the fault zones are not associated with

reducing lithologies or fluids, the basinal brines may react with Ca- and U-rich basement

Page 123: Geometric and hydrodynamic modelling of fluid-structural

109

units, leach uranium and form Ca-rich fluids, and return to the unconformity and basin,

where they mix with the Na-rich basinal fluids and precipitate uranium to form

sandstone-hosted deposits (location B in Figure 3.15). The paleo-regolith and/or

existence of microfractures probably favour the pervasive infiltration of basinal Na-rich

fluids through the basement (locations C and D in Figure 3.15) (Boiron et al., 2010;

Mercadier et al., 2010), where they may either interact with reducing lithologies and/or

fluids to form the basement-hosted deposits (location C in Figure 3.15), or interact with

Ca- and U-rich basement rocks to extract U and evolve towards Ca-rich brines, and then

migrate upwardly along the highway of the more permeable fault zones into the

sandstone where they mix with Na-rich fluids to form sandstone-hosted deposits (location

D in Figure 3.15).

These modelling-based predictions substantiate the statement made by Jefferson

et al. (2007), that each fault system had the potential to generate both basement-hosted

deposits and/or sediment-hosted ore deposits (p. 297, Jefferson et al., 2007), and may also

explain why both basement-hosted and sandstone-hosted ore occur in the same deposit.

The formation of the exceptionally high grade uranium deposits may have benefited from

a situation of prolonged stationary hydrodynamic conditions, in a system with continuous

replenishment of ore constituents (Hoeve and Quirt 1984). This system may have

developed during periods of tectonic quiescence, in which the graphite-rich basement

faults promoted the onset of thermal convection in the permeable sandstone unit, and

interacted with the convection framework by controlling the location of ascending flow,

or by acting as fluid conduits of either ingress or egress flow, such that basinal fluids

circulated and penetrated into basement through microfractures and fault zones, leached

Page 124: Geometric and hydrodynamic modelling of fluid-structural

110

uranium from uranium-rich basement lithologies and formed Ca-rich fluids, and then

migrated upwards to mix with basinal fluids to precipitate uranium in the vicinity of

major faults (Boiron et al., 2010; Richard et al., 2010).

3.6 Conclusions

This study used numerical experiments to test the effects of basement faults on

thermal convection, and the modelling results yield the following conclusions:

1. Free thermal convection may have developed in the more permeable Athabasca

Group. At given hydraulic parameters, the size of convection cells varies slightly due to

the perturbation induced by the presence of fault zones.

2. The relatively higher thermal conductivity of the graphitic faults can induce

subtle thermal anomalies adjacent to them, which can promote the onset of the thermal

convection.

3. When isolated fault zones are introduced into the model, they generally

coincide with upwelling flow, and the location and dip angle of the fault do not seem to

affect the fluid flow pattern. In addition, the presence of a single fault zone does not

change the overall thermal profile nor the width of convection cells.

4. In cases when there are two fault zones, they may coincide with upwelling

plumes, or be situated within convection cells, depending on fault spacing and orientation.

In the latter case, weak convective circulation tends to develop inside each fault zone.

The convection cells immediately on either side of the fault zone in such models may

increase in width slightly, depending on the spacing and orientation of the two fault zones.

Page 125: Geometric and hydrodynamic modelling of fluid-structural

111

Figure 3.15 A simplistic, schematic illustration of an idealized genetic model related to thermal convection

for unconformity-related uranium deposits. The proposed fluid circulation paths, alteration and

mineralization zones are indicated. Not to scale. See text for explanation. Original in colour.

Page 126: Geometric and hydrodynamic modelling of fluid-structural

112

5. Convection can penetrate into the uppermost basement when its permeability is

less than two orders of magnitude lower than that of the sandstone. Under such

conditions, the basement faults are not only centred on an up-flow, but can also passively

act as fluid conduits, depending on the thermal conductivity and the relative location

within the model.

6. The interaction between thermal convection and basement faults provides

conditions for fluid exchange between the basement and the overlying sandstone aquifer,

facilitating uranium transportation and precipitation. The modelling results provide

plausible explanations for the genesis of both basement-hosted and sandstone-hosted

uranium deposits related to thermal convection in the vicinity of basement faults.

3.7 Acknowledgements

The research was funded by a grant to Chi and Bethune from the Natural

Resources of Canada through the TGI-4 (Targeted Geoscience Initiative Phase 4),

uranium ore systems project.

Page 127: Geometric and hydrodynamic modelling of fluid-structural

113

Chapter 4 Structural controls on fluid flow during compressional reactivation of

basement faults4

4.1 Introduction

The Athabasca Basin in northern Saskatchewan and Alberta, Canada (Figure 4.1),

hosts a number of high-grade unconformity-related uranium deposits, which are generally

located at the intersection of reactivated basement faults and the basin-basement

unconformity (Jefferson et al., 2007; Kyser and Cuney, 2009). Some of the deposits are

situated in the basement rocks below the unconformity, whereas others are hosted in

sandstones of the Athabasca Group above the unconformity; yet others are observed to

straddle the unconformity surface (Jefferson et al., 2007; Kyser and Cuney, 2009).

Although the source of uranium is still under debate, i.e., sediment-derived (e.g., Hoeve

and Sibbald, 1978; Hoeve et al., 1980; Kotzer and Kyser, 1995; Fayek and Kyser, 1997)

versus basement-derived (e.g., Hecht and Cuney, 2000; Cuney et al., 2003; Derome et al.,

2003; Richard et al., 2010; Mercadier et al., 2013), the most widely accepted model for

the formation of unconformity-related uranium deposits involves circulation of basin-

derived, oxidizing, uranium-bearing fluids in fault zones crosscutting the unconformity,

where uraninite was precipitated through fluid-rock interaction and/or mixing with

basement-derived reducing fluids (see Jefferson et al., 2007; Kyser and Cuney, 2009, for

review of genetic models). Alteration and mineralization patterns suggest that the

sandstone-hosted deposits were produced by fluid flow from the basement into the basin,

4 This chapter is currently in preparation, and will be submitted to Tectonophysics for peer review. “Li, Z.,

Chi, G., Bethune, K.M., Thomas, D., Zaluski, G. and Kotzer, T., 2016. Structural controls on fluid flow

during compressional reactivation of basement faults: insights from numerical modelling for the formation

of unconformity-related uranium deposits in the Athabasca Basin, Canada.”

Page 128: Geometric and hydrodynamic modelling of fluid-structural

114

or egress flow, whereas the basement-hosted deposits were formed by fluid flow from the

basin into the basement, or ingress flow (Fayek and Kyser, 1997).

Tectonic deformation has been advocated by many researchers as a potential

mechanism for the circulation of ore-forming fluids in various geologic settings (e.g.,

Cox, 2005; Zhang et al., 2011). Fault zones have been generally considered as one of the

most important permeable fluid pathways for hydrothermal ore-forming systems (e.g.,

Cox, 2005). Numerous studies have been conducted to examine the interaction between

deformation in fault zones and fluid flow in various ore systems (e.g., Ord and Oliver,

1997; Garven et al., 2001; Cox, 2005; McLellan et al., 2006; Oliver et al., 2006; Schaubs

et al., 2006; Ford et al., 2009; Sheldon, 2009; Liu et al., 2012; Leader et al., 2013; Zhang

et al., 2013; Zhu et al., 2014), which have greatly contributed to our understanding of the

genesis of these ore deposits.

For the Athabasca Basin, many deposit-scale studies have illustrated the intimate

spatial relationship between basement faults and uranium deposits, as reviewed by

Jefferson et al. (2007) and Li et al. (2015a). These faults, which were originally ductile

and underwent repeated brittle reactivation during and after the formation of the

Athabasca Basin, are considered to have played a key role in controlling fluid flow and

ore deposition (Davies, 1998; Annesley et al., 2005; Jefferson et al., 2007; Li et al.,

2015a). However, the exact mechanisms at play are still not well understood. An early

numerical modelling study of fluid flow associated with deformation by Schaubs et al.

(2005) suggested that fluid can flow into a basement fault (e.g., the McArthur River

deposit) or out of the fault (e.g., the Cigar Lake deposit) in an overall compressional

stress regime, depending on the attitude of the fault. However, the information provided

Page 129: Geometric and hydrodynamic modelling of fluid-structural

115

was limited (in an abstract), and the driving forces for the fluid flow patterns, as

described by Schaubs et al. (2005), were not clear. On the other hand, numerical

modelling by Cui et al. (2012b) demonstrated that ingress flow and egress flow were

associated with alternating extensional and compressional stress fields, respectively.

Considering the observation that most of the unconformity-related ore-bounding

structures demonstrate obvious reverse displacements, it is reasonable to speculate that

fluid flow related to mineralization took place in a (trans-) compressional regime,

although the possibility of intermittent extensional stress regimes during the

mineralization periods cannot be excluded. However, it remains unclear whether or not

both ingress and egress flows can be generated in a compressional environment, and

particularly how these may be related to the dip angle and pre-existing fault offset, as

well as the degree of the compressional deformation.

This study was formulated to more fully explore the relationships by undertaking

numerical modelling of fluid flow in relation to reactivation of basement faults within a

compressional stress regime. Various two-dimensional models were tested to

systematically investigate the effects of dip angle, pre-existing offset, and volumetric

strain, on fluid flow patterns. The effects of episodic reactivation of the basement faults

on fluid flow and the significance for unconformity-related uranium mineralization are

discussed in light of the numerical modelling results.

4.2 Geological setting

4.2.1 Regional geology

Page 130: Geometric and hydrodynamic modelling of fluid-structural

116

The late Paleoproterozoic to Mesoproterozoic Athabasca Basin unconformably

overlies Archean to Paleoproterozoic basement, belonging to the Taltson magmatic zone,

the Rae Province, and the Hearne Province from west to east (Card et al., 2007a, b)

(Figure 4.1). The Rae Province is separated from the Hearne Province by the Snowbird

tectonic zone (Figure 4.1) (Hoffman, 1988). The basement rocks comprise mainly

granitoid gneisses, metasedimentary rocks and mafic to felsic intrusions, and have been

highly deformed and metamorphosed (Card et al., 2007a, b). In the eastern Athabasca

Basin, many of the known uranium deposits are located close to the transitional zone

between the Wollaston and the Mudjatik domains of the Hearne Province (Figure 4.1)

(Annesley et al., 2005).

The sedimentation of the Athabasca Group started from ca. 1750 Ma (Orrell et al.,

1999; Ramaekers et al., 2007; Alexandre et al., 2009). The preserved Athabasca Group is

grouped into four basin-wide unconformity-bounded sequences (Ramaekers et al., 2007).

Sequence 1 (Fair Point Formation) and Sequence 2 (Read, Smart and Manitou Falls

formations) mainly comprise quartz-rich conglomerates and sandstones; Sequence 3

consists of sandstones of the Lazenby Lake Formation, and siltstones and mudstones of

the Wolverine Point Formation. Sequence 4 includes sandstones of the Locker Lake and

Otherside formations, shales of the Douglas Formation and stromatolitic carbonates of

the Carswell Formation (Ramaekers et al., 2007) (Figure 4.1). The paleogeothermal

gradients deduced from fluid inclusion studies suggest that the basinal cover could have

reached 5 to 6 km (Pagel, 1975; Derome et al., 2005), although much of the strata have

been eroded; the current maximum thickness in the central part of the basin is about 1.5

km (Ramaekers et al., 2007). There is a great deal of uncertainty, however, about the

Page 131: Geometric and hydrodynamic modelling of fluid-structural

117

amount of material that has been eroded and the related geothermal gradient based on

alternative interpretations of fluid inclusion data (Chi et al., 2015; Chu and Chi, 2016).

The strata of the Athabasca Basin are commonly cut by the northwest-trending McKenzie

diabase dykes, which have been dated at 1267 ± 2 Ma (LeCheminant and Heaman, 1989).

4.2.2 Structural controls on unconformity-related uranium deposits

Uranium deposits in the Athabasca Basin are generally located near the sub-

Athabasca unconformity, and are commonly spatially associated with reactivated

basement faults that crosscut the unconformity (Jefferson et al., 2007). The uranium

mineralization may be sandstone-hosted, located in the conglomerate and sandstone of

the Athabasca Group immediately above the unconformity (e.g., Cigar Lake deposit;

Figure 4.2A), or basement-hosted (e.g., Eagle Point deposit; Figure 4.2D) (Hoeve and

Quirt, 1984; Fayek and Kyser, 1997). However, many deposits appear to be a

combination of the two mineralization styles, containing both sandstone-hosted and

basement-hosted orebodies (e.g., Key Lake and McArthur River deposits; Figures 4.2B

and C). Sandstone-hosted mineralization can occur up to 40 m above the unconformity,

whereas basement-hosted mineralization can be found up to 400 m below the

unconformity (Jefferson et al., 2007).

Page 132: Geometric and hydrodynamic modelling of fluid-structural

118

Figure 4.1 Location and regional geologic framework of the Athabasca Basin (after Card et al., 2007a, b;

Jefferson et al., 2007; Ramaekers et al., 2007; Card et al., 2014). FP = Fair Point; S = Smart; RD = Read;

MF = Manitou Falls; LZ = Lazenby Lake; W= Wolverine Point; LL = Locker Lake; O = Otherside; D =

Douglas; C = Carswell. AB = Athabasca Basin. Red stars represent major uranium deposits. Original in

colour.

Page 133: Geometric and hydrodynamic modelling of fluid-structural

119

Numerous basement faults, commonly cutting the Athabasca Group and

underlying basement rocks, have offset the unconformity and overlying sandstone on the

order of tens up to hundreds of metres, reflected by abrupt elevation change in the

unconformity topography (Jefferson et al., 2007; Li et al., 2015a). They are generally

thought to represent reactivated pre-Athabasca ductile basement structures and have

undergone repeated brittle reactivation (Jefferson et al., 2007). For example, a vertical

displacement of up to 250 m has been estimated for the Dufferin Lake thrust fault, which

represents a brittle reactivation of the Virgin River shear zone (Card et al., 2007a) and is

associated with the Centennial uranium deposit (Figure 4.1) (Reid et al., 2014). Recent

3D modelling of the sub-Athabasca unconformity surface and analysis of fault systems

using regional aeromagnetic data, also suggested that the reactivated northeast-trending

faults may be the principal control on basement topography and uranium mineralization

(Li et al., 2015a). However, the detailed reactivation history of these structures,

particularly the tectonic stress status at the time of mineralization remains unclear.

The intimate association of the uranium deposits with reactivated basement faults

has been manifested at different deposits (Figure 4.2). For instance, the orebody in the

Deilmann pit at Key Lake mine is mainly controlled by the west-southwest striking and

moderately to steeply dipping Key Lake fault, that is rooted in interlayers of granitic

pegmatite and graphitic pelitic gneiss and has reversely offset the unconformity by up to

12 m (Harvey and Bethune 2007) (Figure 4.2B). The ore zones at the McArthur River

deposit are related to the P2 reverse fault (striking 050° and dipping 40° to 45° SE) which

is localized along the contact between a quartzite unit and graphitic gneisses, and shows a

reverse displacement of the unconformity by up to 80 m (McGill et al., 1993; Bernier,

Page 134: Geometric and hydrodynamic modelling of fluid-structural

120

2004) (Figure 4.2C). Stratigraphic studies of the Manitou Falls Formation suggested that

the P2 structure preserves a record of multiple reverse reactivations with occasional

normal faulting in an overall contractional regime (Bernier, 2004), although it is unclear

whether or not all these reactivation events are related to mineralization. The Rabbit Lake

thrust at the Rabbit Lake deposit dips south-southeast at about 30° and has reversely

offset the unconformity with a vertical displacement of about 75 m (Hoeve and Sibbald,

1978). These ore-controlling structures are typically enveloped by alteration haloes

(Figure 4.2), mainly illite, chlorite, dravite, kaolinite, silicification or desilicification,

with dimensions of up to 400 m wide at the basal unconformity and a thousand metres in

strike length (Hoeve and Quirt, 1984; Jefferson et al., 2007).

The primary uranium mineralization age is around ca 1590 Ma (Alexandre et al.,

2009), although a spectrum of younger ages has been reported, suggesting multiple

episodes of uranium mineralization and/or remobilization events (Alexandre et al., 2009;

Cloutier et al., 2011; Mercadier et al., 2011). The main uranium mineralization event may

reflect far-field, continent-wide tectonic events such as the Mazatzal Orogeny (ca 1.6 Ga)

(Davidson, 2008; Alexandre et al., 2009). Such events could be responsible for brittle

reactivation of the pre-Athabasca ductile structures and stimulation of fluid flow related

to uranium mineralization (e.g., Cui et al., 2012b). Other younger mineralization events

that have been detected may reflect subsequent far-field tectonic events (Alexandre et al.,

2009).

Page 135: Geometric and hydrodynamic modelling of fluid-structural

121

Figure 4.2 Representative sections of four well-studied unconformity-related uranium deposits of the

eastern Athabasca Basin showing the strong spatial association of the deposits with the intersection of

basement-rooted fault zones and the unconformity surface. (A) Cigar Lake deposit, consisting of

predominantly unconformity ore and perched ore in the overlying sandstone (after Jefferson et al., 2007).

(B) Deilmann pit, Key Lake deposit, including both basement-hosted and unconformity ore, controlled by

the Key Lake fault (after Harvey and Bethune, 2007). (C) McArthur River deposit, mostly basement-hosted

ore, controlled by the P2 fault (after Marlatt et al., 1992; Bronkhorst et al., 2012). (D) Eagle Point deposit,

mostly basement-hosted ore, controlled by the Collins Bay thrust and Eagle Point fault (after Marlatt et al.,

1992). Vertical scale = horizontal scale. Original in colour.

Page 136: Geometric and hydrodynamic modelling of fluid-structural

122

4.3 Principles, physical models and initial and boundary conditions of numerical

modelling

4.3.1 Simulation principles and methods

The FLAC3D (Fast Lagrangian Analysis of Continua 3-dimensional version)

software (Itasca, 2012), which is based on the explicit finite difference method, was used

in this study. This code is capable of simulating the interaction between fluid flow and

tectonic deformation, and has been successfully applied to various ore-forming systems

(e.g., Oliver et al., 2006; Vos et al., 2007; Cui et al., 2012b; Leader et al., 2013; Zhang et

al., 2013). The principles, governing equations and numerical methods have been

reviewed in many previous studies (e.g., Oliver et al., 1999, 2006; Zhang et al., 2008,

2009, 2011; Cui et al., 2012b; Liu et al., 2012; Leader et al., 2013), therefore only an

overview is given here.

In FLAC3D, a 3D mesh consisting of hexahedral elements is built first to

represent the geometry of the materials to be modelled, and each element is modelled as

isotropic elastic–plastic Mohr–Coulomb material (Itasca, 2012). The theory underlying

this constitutive model has been summarized by Vermeer and de Borst (1984). Once

mechanical and hydraulic properties are assigned to each element, they will behave

according to the Mohr–Coulomb constitutive law in response to the applied boundary and

initial conditions.

FLAC3D uses an iterative process to model the interaction between deformation

and fluid flow. First, in the mechanical calculation step, the material initially responds to

the applied stress in an elastic manner; after the stress exceeds the yield point defined by

the Mohr–Coulomb yield criteria, plastic deformation (shear failure or tensile failure)

Page 137: Geometric and hydrodynamic modelling of fluid-structural

123

occurs. Such plastic deformation can cause volume changes, characterized by the dilation

angle, which leads to changes in pore pressure (volume increase or dilation giving rise to

a decrease in pore fluid pressure). Second, in the fluid flow calculation step, these

changes in fluid pressure contribute to changes in the hydraulic head that drives fluid

flow according to Darcy’s law. In turn, the changes in pore pressure cause changes in the

effective stresses, which further affect the response of the solid (e.g., a reduction in

effective stress may induce plastic yield) in the next mechanical calculation step. This

interaction between deformation and fluid flow is fully coupled through iteration at each

time step, as also adopted by other studies (e.g., McLellan et al., 2004; Schaubs et al.,

2006).

It needs to be mentioned that FLAC3D does not simulate explicitly the variation

of porosities and permeabilities during the computing cycle. Instead, FLAC3D uses the

concept of the dilation angle as a dynamic permeability and porosity modifier to Mohr –

Coulomb rocks (Oliver et al., 2006). More detailed descriptions of deformation and fluid

flow coupling in FLAC3D can be found in Ord and Oliver (1997), McLellan et al. (2004)

and Oliver et al. (2006).

4.3.2 Model setup

A conceptual model has been constructed with a dimension of 5 km in the

horizontal direction, 7.5 km in the vertical direction, and 50 m in the direction

perpendicular to the cross-section (see below for explanation). The model consists of a

permeable sandstone aquifer sandwiched between the relatively impermeable cover and

Page 138: Geometric and hydrodynamic modelling of fluid-structural

124

basement (Figure 4.3), reflecting the basic geological features of the host rocks

surrounding an unconformity-related uranium deposit (e.g., Jefferson et al., 2007).

The relatively mud-rich and fine-grained Wolverine Point Formation has been

considered to be a physical barrier (an aquitard) to fluid flow (Hiatt and Kyser, 2007).

Therefore, the cover in our models includes the Wolverine Point Formation, the

preserved strata above it, and the eroded strata. As discussed earlier, the thickness of the

eroded strata is uncertain, and therefore we assigned a thickness of 5 km to the cover, in

line with the estimation of Pagel (1975). The quartz-rich sandstone and conglomerate

below the Wolverine Point Formation has a maximum aggregate thickness of about 1.5

km (Ramaekers et al., 2007), and accordingly the permeable aquifer in our models was

also assigned a thickness of 1.5 km.

A fault zone of varying dip straddling the unconformity is incorporated into the

model. The fault zone used in this study has a width of 50 m and extends from the

unconformity to the bottom of the model. As mentioned previously, a reverse fault offset

of the unconformity has been observed in many deposits (e.g., McGill et al., 1993;

Harvey and Bethune, 2007; Kerr, 2010); offsets are on the order of tens to hundreds of

meters (Jefferson et al. 2007). For this study, we start with models without an offset, and

then gradually increase the offset to a maximum of 100 m in order to test the effects of

variable offsets on fluid flow during deformation. Our models have considered only the

simplest fault geometry, i.e., a straight fault (Figure 4.3), although it is understood that

faults can have more complex geometries in nature (e.g., curviplanar surfaces) and hence

more complicated interaction between deformation and fluid flow. This simplified

Page 139: Geometric and hydrodynamic modelling of fluid-structural

125

approach is necessary in order to grasp the fundamental relationships between

compressional deformation and fluid flow.

The deformed and metamorphosed basement below the permeable sandstone

layer has an arbitrary thickness of 0.5 km. Instead of a uniform basement, however, a

footwall unit and a hangingwall unit adjacent to the fault zone, with equal widths of 200

m (Figure 4.3), are assigned with different mechanical properties to simulate the

behaviour of different basement lithologies (Table 1). For example, the more competent

footwall unit may represent quartzite or granitoids, whereas the relatively weak hanging-

wall unit may represent foliated psammitic or pelitic gneisses, which is a common

configuration of basement units in relationship to faults (Figures 4.2B to D) (e.g., Harvey

and Bethune, 2007; Kerr, 2010).

This overall configuration is consistent with the models adopted by many

previous studies (e.g., Cui et al., 2012a, b; Chi et al., 2013, 2014). However, in order to

focus on the mineralization zones around the unconformity, the topmost 5 km was not

included in the modelling. It should be noted that in order to save computation time, our

3D models only have one element in the direction normal to the section with widths of 50

m. For this reason, the current models have not incorporated the variation of the fault

strikes, and only considered the situation in which the strike of the fault is perpendicular

to the maximum compressional direction. In this respect, our models are 2D in nature.

The mesh of each model has 2640 elements (zones), with the sizes ranging from a

minimum of 10 m (fault zone) to a maximum of about 170 m (near model edges).

The constitutive parameters, including shear modulus, bulk modulus, cohesion,

tensile strength, friction angle and dilation angle, permeability and porosity (Table 1),

Page 140: Geometric and hydrodynamic modelling of fluid-structural

126

were provided by Cameco (unpublished data). The faults in the current models, as in

many other studies (e.g., Oliver et al., 2006; Cui et al., 2012b; Zhang et al., 2013), are

treated as weak zones with higher permeability and lower mechanical properties than

other rocks (Table 1). All the units in the models are considered to be homogeneous and

isotropic with respect to permeability.

4.3.3 Initial and boundary conditions

Numerical modelling of sediment compaction as well as hydrocarbon generation

(Chi et al. 2013, 2014) suggest that the fluid pressure within the Athabasca Basin was

near hydrostatic regime during sedimentation. As such, hydrostatic pressure was used as

the starting condition of the system, as previously adopted in similar studies (e.g., Cui et

al. 2012b). Fluid is free to flow out of the model at the upper boundary, whereas the

lower and vertical boundaries are impervious to fluid flow.

The top boundary is subject to an applied normal stress representing 5 km of

overburden (5000 m × 9.8 m/s2 × 2400 kg/m3). Prior to the application of compression,

the models are brought to initial mechanical equilibrium to simulate the in situ stress

balance at a depth of 5 km. After equilibrium is reached, a constant convergent

displacement rate of 1.0×10−8 m/s is applied normal to the left and right boundaries of the

models. This induces a compressive bulk strain rate on the order of 10−10 to 10−11 s−1,

which is much higher than the typical geologic active convergence rates of 10−12 to 10−16

s−1 (Pfiffner and Ramsay, 1982; Campbell-Stone, 2002), but is required in order to obtain

the desired strain in FLAC3D within a reasonable time frame (i.e., days or weeks rather

Page 141: Geometric and hydrodynamic modelling of fluid-structural

127

than months). Nevertheless, one simulation with a lower compressive bulk strain rate on

the order of 10−13 to 10−14 s−1 was also tested for sensitivity study purposes (see below).

The lower boundary of the models is fixed in the vertical direction but can move

horizontally, whereas the other boundaries are free to move in any direction. The entire

model deforms according to its mechanical properties and boundary conditions. All the

models in this study were deformed to a maximum 2% bulk shortening.

Page 142: Geometric and hydrodynamic modelling of fluid-structural

128

Figure 4.3 Cross-sectional view of the physical model forming the foundation of all numerical modelling

carried out in this study. Original in colour.

Page 143: Geometric and hydrodynamic modelling of fluid-structural

129

Table 4.1 Input parameters of various hydrological units used in model setups.

Property Sandstone Basement Fault

zone Hangingwall Footwall

Density (kg/m3) 2,500 2,650 2,400 2,650 2,650

Bulk modulus (Pa) 3.20E+10 4.95E+10 9.50E+09 4.95E+10 4.95E+10

Shear modulus (Pa) 4.00E+09 2.90E+10 3.00E+07 2.90E+10 2.90E+10

Cohesion (Pa) 3.00E+06 4.00E+06 3.00E+03 4.00E+06 4.00E+06

Tensile strength (Pa) 1.20E+06 5.00E+06 6.00E+05 1.50E+06 6.00E+06

Friction angle 30° 30° 15° 20° 30°

Dilation angle 4° 3° 5° 4° 3°

Porosity 0.2 0.1 0.2 0.1 0.1

Permeability (m2) 3.00E-13 3.00E-16 1.00E-12 3.00E-16 1.00E-17

Page 144: Geometric and hydrodynamic modelling of fluid-structural

130

4.4 Modelling results

4.4.1 Effects of fault dip angle variation and degree of bulk shortening

4.4.1.1 Case 1—Fault dipping 45°

The first model incorporates a fault zone with a dip angle of 45° (Figure 4.4). All

the parameters are the same as described in Table 1, and the initial and boundary

conditions are the same as presented in the previous section. These physical parameters

and conditions remain unchanged in the following models unless otherwise specified.

After a bulk shortening of 0.2% is reached, the unconformity surface is not

affected by the compression (Figure 4.4A). The fault zone stands out as a zone of

intensive shear strain accumulation (Figure 4.4B), and the whole model, especially the

fault zone, is under contraction, although subtle taper-shaped dilation (pore volume

increase) is developed in the area above the fault tip in the sandstone (Figure 4.4C). At

this stage, due to the fact that pore pressure increases at a relatively faster rate in the low

permeability basement, as well as the fact that the fault zone is relatively more

permeable, the fluid flows up the fault into the sandstone (Figure 4.4C).

After the deformation reaches a stage of bulk shortening of 2%, the unconformity

surface on the hangingwall and footwall sides has been uplifted 32.3 m and 1.2 m,

respectively (Figures 4.4D and 4.5). Reverse displacement along the fault has resulted in

a vertical displacement (i.e., fault throw) of 31.1 m (Figure 4.5). The fault zone shows the

highest shear strain localization, which extends up into the sandstone wedge above the

footwall (hereinafter termed the ‘footwall sandstone wedge’) as well as into the hanging

wall of the basement (Figure 4.4E), in a symmetrical, flared pattern. More importantly,

the fault zone becomes a region of dilation, which correlates closely with the distribution

Page 145: Geometric and hydrodynamic modelling of fluid-structural

131

of the shear strain (Figures 4.4E and F). The dilation, however, is not limited to inside the

fault zone, but is also localized along the contact area between the hangingwall unit and

the fault zone and widens at depth, similar to the shear strain pattern (Figure 4.4F). As a

result of the dilation, pore pressures are decreased and fluid within the sandstone units is

drawn first horizontally towards the region of dilation within the footwall sandstone

wedge and then downwards into the fault zone (Figure 4.4F). In fact, a series of snapshots

of strain patterns and fluid migration from a continuously run experiment for different

degrees of strain indicate that it is at around 0.6 % bulk shortening that the dilation occurs

in the fault zone, which approximately corresponds to the onset of the reverse

displacement along the fault (Figure 4.5).

The maximum fluid flow rates at the two compression stages occur inside the

fault zone, with a rate of 3.7e-7 m/sec and 6.6e-7 m/sec, respectively (Figures 4.4C and

F). In contrast, the flow rate in the basement outside the fault zone (including the hanging

wall and footwall) is about two orders of magnitude lower than that in the fault zone due

to the low permeability.

To better understand whether or not the phenomena observed above are related to

the strain rate, a slower strain rate on the order of 10−13 to 10−14 was tested. The

modelling results (not shown) indicate that the distribution of shear and volumetric strain

and associated fluid flow patterns are the same as those shown above, except for a longer

time to reach desired strain.

Page 146: Geometric and hydrodynamic modelling of fluid-structural

132

Figure 4.4 Models with a fault dipping 45°. Modelling results showing the model geometry after

deformation (A), shear strain distribution (B), volumetric strain distribution and fluid flow patterns (C) at

the 0.2% bulk shortening stage. D, E and F are the counterparts at the 2% bulk shortening stage. Note that

the arrows are not to scale and only used to show the direction of the flow (the maximum flow velocity is

provided). Note also that positive volumetric strain represents dilation (increase in volume). See text for

more explanation. Original in colour.

Page 147: Geometric and hydrodynamic modelling of fluid-structural

133

Figure 4.5 Plot of depth of the unconformity surfaces both on the hangingwall and footwall sides for each

model showing the vertical displacements (fault throw) at different strain stages. See text for more details.

Original in colour.

Page 148: Geometric and hydrodynamic modelling of fluid-structural

134

4.4.1.2 Case 2—Fault dipping 60°

For the second case, the dip of the fault zone was increased to 60° (Figure 4.5A),

and all other parameters were kept the same as the previous model. The modelling results

are similar to those of the 45° fault model. At the earlier stage (0.2% bulk shortening), the

fault zone is the main locus of shear strain (Figure 4.5B) and contraction (Figure 4.5C),

and fluid flows up along the fault zone into the sandstone (Figure 4.5C). At the 2% bulk

shortening stage, the top part of the fault zone as well as the nearby footwall sandstone

wedge represent the localization of most significant shear strain and dilation (Figures

4.5E and F). The shear strain also extends farther into the hanging wall than in the 45°

fault model. Broad zones of high shear strain and coincident dilation extend upward into

the sandstone and downward into the basement hanging wall domain, counter-clockwise

to and inclined at a shallower angle (~45°) than the fault zone. The amount of fault throw

after 2% bulk shortening also differs from the one of the previous simulation. In this case

the unconformity surface on the hanging wall and footwall sides is uplifted 29.5 m and

5.7 m, respectively, with a fault throw of 23.8 m (Figure 4.5). Like the 45° fault model,

the fluid flows down into the fault zone, and reaches out into the hanging wall towards

the dilation zone at this stage of shortening.

The maximum flow rates all occur inside the fault zone, with a rate of 3.4e-7

m/sec at the 0.2% bulk shortening stage, and 3.0e-7 m/sec at the 2% bulk shortening

stage, which are slightly slower than the ones in the 45° fault model.

Page 149: Geometric and hydrodynamic modelling of fluid-structural

135

Figure 4.6 Models with a 60° dipping fault. Modelling results showing the model geometry after

deformation (A), shear strain distribution (B), volumetric strain distribution and fluid flow patterns (C) at

the 0.2% bulk shortening stage. D, E and F are the counterparts at the 2% bulk shortening stage. Note that

the arrows are not to scale and only used to show the direction of the flow (the maximum flow velocity is

provided). Note also that positive volumetric strain represents dilation (increase in volume). See text for

more explanation. Original in colour.

Page 150: Geometric and hydrodynamic modelling of fluid-structural

136

4.4.1.3 Case 3—Fault dipping 30°

The third simulation incorporates a fault zone dipping 30° (Figure 4.7). The

overall distribution of the shear strain and volumetric strain, and related flow patterns at

different stages of bulk shortening (Figure 4.7) are similar to those of the previous two

simulations, characterized by upward flow in the early stage and downward flow in the

later stage. However, some differences are notable. For example, in the late stage of

deformation, the high strain zone (both shear strain and dilation) is much more limited

within the fault zone than in the previous models (Figures 4.7E and F). With the decrease

of fault dip angle from 60° to 30°, the dilation zone shows a tendency to migrate from a

focused area in the upper part of the fault zone to the whole fault zone (compare Figures

4.6E and F, 4.4E and F, and 4.7E and F).

The amount of fault throw after 2% bulk shortening is about 29.0 m, which is

slightly less than that for the 45° model (31.1 m) but greater than the one for the 60°

model (23.8 m) (Figure 4.5). The maximum flow rates in this case are the highest among

the models with a rate of 1.6e-6 m/sec and 1.2e-6 m/sec at the two stages of compression,

respectively.

Page 151: Geometric and hydrodynamic modelling of fluid-structural

137

Figure 4.7 Models with a 30° dipping fault. Modelling results showing the model geometry after

deformation (A), shear strain distribution (B), volumetric strain distribution and fluid flow patterns (C) at

the 0.2% bulk shortening stage. D, E and F are the counterparts at the 2% bulk shortening stage. Note that

the arrows are not to scale and only used to show the direction of the flow (the maximum flow velocity is

provided). Note also that positive volumetric strain represents dilation (increase in volume). See text for

more explanation. Original in colour.

Page 152: Geometric and hydrodynamic modelling of fluid-structural

138

4.4.2 Effects of pre-existing fault offset

Two cases were considered in this scenario. The models have the same geometry

as the 45° fault model, the only difference being that the unconformity surface on the

hangingwall side was maintained at a depth of -6560 m (as in previous models), whereas

the one on the footwall side was lowered to a depth of -6610 m and -6660 m to simulate

pre-existing vertical offsets of 50 m and 100 m, respectively (Figure 4.6).

Generally, the distribution of shear and volumetric strain and related fluid patterns

obtained for these two cases are similar to each other (Figure 4.8) and are similar to those

of the no-offset counterparts (Figure 4.4). At the early stage of compression (not shown),

the fault zones are locations of contraction, and fluid migrates upward into the sandstone

along the fault zone, whereas at the late stage of compression, the fault zones are

locations of dilation, and fluid flows into the fault zones (Figure 4.8).

After 2% bulk shortening, there is a fairly significant increase in vertical offset

(i.e., fault throw) in the two models, from 50 m to 83.7 m and from 100 m to 131.1 m,

respectively. In each case, most of this increase results from uplift of the unconformity

surface on the hangingwall side, whereas the unconformity surface on the footwall side is

only slightly elevated (less than 1 m) in both models (Figures 4.8B and F).

Like the models without a pre-existing offset (Figure 4.4F), in both cases the

dilation is mainly localized inside the fault zone, but it also extends into the sandstone on

the footwall side of the fault as well as into the basement on the hangingwall side, in a

symmetrical fashion (Figures 4.8D and H). There is little difference in maximum flow

rate in both cases, with rates of 6.5e-7 m/sec and 6.1e-7 m/sec, respectively.

Page 153: Geometric and hydrodynamic modelling of fluid-structural

139

Figure 4.8 Modelling results showing the model geometry with a pre-existing vertical offset of 50 m before

compression (A) and after 2% bulk shortening (B), shear strain distribution (C), volumetric strain

distribution and fluid flow patterns (D). E, F, G and H are the counterparts for the model with a pre-existing

vertical offset of 100 m. Note that the arrows are not to scale and only used to show the direction of the

flow (the maximum flow velocity is provided). Note also that positive volumetric strain represents dilation

(increase in volume). See text for more explanation. Original in colour.

Page 154: Geometric and hydrodynamic modelling of fluid-structural

140

4.4.3 Effects of episodic compressional deformation

The effects of episodic compressional deformation were simulated by pausing of

compression after a certain degree of deformation, and then resuming it again. The model

with a 45° dip angle of the fault zone was again used for this purpose.

If the deformation ceases after 0.2% bulk shortening, the fluid continues to flow

upwards into the sandstone until the pore pressure in the basement equilibrates back to

the hydrostatic level. If the compression is applied again afterwards, the flow patterns

will depend on the bulk shortening as described above for the models without pre-

existing offsets.

If the deformation ceases after 2% bulk shortening, the fluid continues to flow

downwards into the basement along the fault zone until the pore pressure evolves back to

the hydrostatic regime. Similarly, if the deformation starts over again, the fault zone will

be reactivated and the fluid flow patterns will be similar to the models with a pre-existing

offset as discussed above.

4.5 Discussion

4.5.1 Fluid flow during compressional reactivation of the basement faults

The models in this study have been used to explore the effects of varying the fault

geometry (dip angles and offsets) on fluid flow in a compressional stress regime. The

modelling results demonstrate that compressional reactivation of the basement faults and

related fluid flow patterns are sensitive to the degree of bulk shortening. At a low bulk

shortening stage, the fault zone is reactivated without significant displacement and

represents an area of contraction. At this stage, fluid is driven up along the fault zone into

Page 155: Geometric and hydrodynamic modelling of fluid-structural

141

the sandstone, which is consistent with the observation by Cui et al. (2012b). This is

interpreted to indicate that compression results in overpressures in the whole system, but

pore pressure increase is faster in the low permeability basement than in the other parts of

the model (see also Cui et al., 2012b).

Dilation (pore volume increase) can occur because of Mohr–Coulomb plasticity in

which shear induces dilatancy (Vermeer and de Borst, 1984; Ord and Oliver, 1997;

Oliver et al., 2006). Shear strain and related dilation in fault zones during compressional

deformation has been observed in many other studies (e.g., Ford et al., 2009; Zhang et al.,

2006, 2013). The sites of shear strain and dilation, which tend to localize fluid flux, have

been considered to be favourable places for mineral precipitation in hydrothermal

systems (e.g., Oliver et al., 2006; Ford et al., 2009; Zhang et al., 2013). Dilation is also

observed in our models. At the low bulk shortening stage, a subtle tapper-shaped dilation

zone occurs above the fault tip (Figure 4.4C). This dilation zone may represent the initial

stage of the fault growth, in which shear microfractures predominately form near the fault

tip as a result of the fault-tip stress concentration and eventually allow a fault to grow

(e.g., Anders and Wiltschko, 1994; Vermilye and Scholz, 1998; d'Alessio and Martel,

2004).

More extensive dilation is observed at a relatively high bulk shortening stage

(great than ~ 0.6% bulk shortening), which corresponds to the reverse displacement along

the faults. Under a compressional stress regime, the dilation may represent the formation

of extensional fractures, extensional-shear fractures, or shear fractures (Cox et al., 2001;

Sibson, 2001), according to the Mohr diagram and Griffith-Coulomb failure criterion (see

Figure 4 in Sibson, 1996). This has been exemplified by the Sue C deposit in the eastern

Page 156: Geometric and hydrodynamic modelling of fluid-structural

142

Athabasca Basin (Figure 1), where uraninite is hosted in sub-horizontal extensional veins,

steeply-dipping shear veins, dilational jogs and fault breccias, all developed within a

reverse fault zone rooted in the basement (Tourigny et al., 2007). This process may also

be compatible with the development of an earthquake in a reverse-fault system as

explained by Doglioni et al. (2011, 2014), in which the hanging wall is over-compressed

during the interseismic stage, and uplifted and dilated at the coseismic stage

accompanying fracture opening and decrease of fluid pressure (see Figure 1 in Doglioni

et al., 2011).

The resultant fault throws in individual models varied depending on the dip angles

and the pre-existing offsets of the faults (Figures 4.5 and 4.8). These dilation zones

mainly develop in the footwall sandstone wedge as well as within the fault zone, but can

also extend into the hangingwall basement. More importantly, the dilation gives rise to a

drop in pore pressure and fluid in the sandstone is sucked into the dilation zones.

In all the models, the maximum flow rates occur within the dilated fault zones,

due to their relatively higher permeability. However, there is a difference in the

maximum flow rates for different individual models, with the shallowly dipping fault

exhibiting the highest flow rate among the models, likely due to the larger associated

vertical displacement and the fact that the fault is closer in orientation to the direction of

maximum compression (horizontal).

The simulation results also demonstrate that neither varying the dip angle nor the

magnitude of pre-existing offset exerts any significant effect on the interactivity between

fault zones, shear and volumetric strain distribution and related fluid flow patterns during

compressional deformation. These factors may therefore be less critical than the degree

Page 157: Geometric and hydrodynamic modelling of fluid-structural

143

of strain (bulk shortening) in controlling fluid flow patterns. The results of monitoring

episodic compression suggest that if certain fault zones were subjected to episodic

periods with low degrees of compression (i.e., events with low bulk shortening), fluid

could be continually pumped up into the sandstone along the fault zones from basement.

In contrast, if they were subjected to episodic periods with high degrees of compression

(i.e., events with high bulk shortening), there could be alternating egress flow at an early

stage and ingress flow at a later stage for each event.

4.5.2 Implication for uranium mineralization

In an intracratonic basin like the Athabasca Basin, one would not expect

deformation, in the form of faulting, to occur frequently and/or as a continuous process.

The most likely scenario is that the basin was tectonically quiet most of the time, as

implied by the flat-lying, unmetamorphosed sedimentary rocks within it (Ramaekers et

al., 2007), although early growth (syndepositional) normal faulting may have facilitated

subsidence (Ramaekers, 2004). Whatever the case, this relatively quiescent situation was

likely interrupted by discrete periods of tectonic activity, perhaps related to certain far-

field tectonic events (e.g., Davidson, 2008), during which the pre-Athabasca ductile

structures, and/or earlier formed syn-Athabasca faults were episodically reactivated

(Jefferson et al., 2007). During periods of tectonic quiescence, large-scale thermal

convection may have been pervasively developed in the basin sequence (Raffensperger

and Garven, 1995; Cui et al., 2012a; Li et al., 2015b), which may have facilitated

infiltration of the basinal fluid into the basement along pre-existing microfractures and

fault zones to leach uranium (Derome et al., 2005; Boiron et al., 2010; Richard et al.,

Page 158: Geometric and hydrodynamic modelling of fluid-structural

144

2010; Mercadier et al., 2011). During tectonically active periods, however, deformation

was most likely the dominant mechanism controlling fluid flow, as thermal convection is

generally thought to be suppressed by heightened tectonic deformation (e.g., Oliver et al.,

2006; Cui et al., 2012b).

Under the conditions of far-field compressional deformation, pre-existing shear

zones/faults within the basin would be prone to variable degrees of reactivation, in some

cases with offset of the unconformity, and in others not, depending on the local

geological setting. Under low degrees of bulk shortening, and limited fault offsets, our

modelling results suggest that fluids from the basement (relatively reducing) could be

driven up along fault zones into the sandstone, where they would potentially interact with

oxidizing basinal fluids and (under chemically favourable conditions) precipitate

uranium. Sandstone-hosted mineralization may have developed in this manner (location

A in Figure 4.9). This may be exemplified by the Cigar Lake deposit (Figure 4.2A),

where no obvious offset along the faults has been observed (implying a relatively low

degree of bulk shortening). It is possible that the partly sandstone-hosted ore in the

McArthur River deposit may also have developed due to this scenario.

If the deformation continued and more shortening took place, certain fault zones

initially under low degrees of compression could theoretically accommodate additional

shortening by substantial reverse displacements. During this process, our modelling

indicates that the faults would become loci of dilation and that these dilation zones would

extend into the footwall sandstone wedge, as well as at depth in the hanging wall of the

basement (Figures 4.4F, 4.6F and 4.7F). The dilation zones would suck the basinal

oxidizing fluids within the sandstone into the faults and basement, where they could

Page 159: Geometric and hydrodynamic modelling of fluid-structural

145

interact with reducing fluids and/or reduced basement lithologies (e.g., graphitic pelite) to

precipitate uranium. Basement-hosted mineralization could be generated under these

conditions (location B in Figure 4.9). The geological observation that the main orebody

in the McArthur River deposit is located around the footwall sandstone wedge of the P2

fault zone (Figure 4.2C), correlates well with our modelling results (Figure 4.4F). A

similar relationship is also seen in the Key Lake deposit (Figure 4.2B), where the orebody

mainly occurs at the intersection between the unconformity surface and the Key Lake

fault, which has offset the former by 12 m (Harvey and Bethune, 2007). The fact that

fluid can flow to significant depth along the fault zone, may also explain the occurrence

of individual lenses of high-grade ore at depth in the basement (e.g., at Key Lake, Figure

4.2B), as well as why some orebodies are found at depths up to 400 m below the

unconformity (e.g., Eagle Point; Figure 4.2D).

Page 160: Geometric and hydrodynamic modelling of fluid-structural

146

Figure 4.9 Schematic diagram illustrating an idealized genetic model for unconformity-related uranium

deposits illustrating that both basement-hosted and sandstone-hosted mineralization may be formed under a

unified compressional stress regime. Arrows (blue) demonstrate fluid flow directions. See text for

explanation. Original in colour.

Page 161: Geometric and hydrodynamic modelling of fluid-structural

147

The observation that egress flow and ingress flow are strongly sensitive to the

degree of the compression (bulk shortening), and that variable dip angles and pre-existing

fault offsets have little effect on fluid flow patterns, explains the observation that major

uranium deposits are associated with faults having widely ranging attitudes and degrees

of displacement. In addition, our results indicate that repeated/episodic compressional

events produce similar fluid flow patterns that may lead to repeated, discrete

mineralization events. This supports the statement proposed by Alexandre et al. (2009)

that the younger uranium mineralization and/or remobilization events may be associated

with some far-field tectonic events.

Although our modelling results provide plausible explanations for the formation

of uranium deposits in the Athabasca Basin, considering the protracted history of the

basin, this does not preclude other mechanisms such as thermal convection

(Raffensperger and Garven, 1995; Cui et al., 2012a; Li et al., 2015b), which may have

dominated fluid flow during periods of tectonic quiescence. Importantly, this study

indicates that an extensional environment is not essential to draw fluids down into the

basement; fluids can also flow into the basement during compression. However, this also

does not preclude the possibility of extensional deformation, which may also be

associated with the reactivation of some basement faults at certain stages of basin

evolution (e.g., Cui et al., 2012b). Due to the unclear details of deposit-scale stress states

at the time of mineralization, detailed investigation, e.g., study of fluid inclusion planes

(FIP) (e.g., Liu et al., 2011), is required to better understand the relationship between

deformation, fluid flow and the dynamic regime. Nevertheless, the observation that the

fluid can flow up and down the same fault system at different stages of the compressional

Page 162: Geometric and hydrodynamic modelling of fluid-structural

148

history, may explain why many of the unconformity-related uranium deposits have both

sandstone-hosted and basement-hosted orebodies. The basement faults with complex

histories of repeated, episodic reactivation, may have played critical roles in the

formation of the giant high-grade uranium deposits in the Athabasca Basin.

4.6 Conclusions

The present 2D numerical modelling study has investigated the effects of varying

fault dip angles and pre-existing offsets in reverse fault systems on the shear and

volumetric strain and fluid flow patterns during compressional reactivation of basement

faults in the Athabasca Basin. The modelling results yield the following conclusions:

1. During compressional deformation, fault zones act as fluid conduits, attracting

up and outward (egress) flow, or down and inward (ingress) flow, depending on the

distribution of strain (contraction or dilation), which is in turn related to the degree of

compression. At relatively low degrees of bulk shortening (<0.6%), the fault zones are

dominantly under contraction and egress flow is favoured; at higher degrees of bulk

shortening, dilation occurs within the fault zone as well as the adjacent footwall

sandstone wedge and ingress flow into the basement is favoured.

2. The variation of the dip angle and amount of pre-existing fault offset has little

effect on the strain distribution and related fluid flow patterns.

3. Our modelling results allow that both sandstone-hosted and basement-hosted

mineralization may be formed in the same fault system at different stages of bulk

shortening within a unified compressional stress regime.

Page 163: Geometric and hydrodynamic modelling of fluid-structural

149

4. Episodic reactivation of a given fault in response to different far-field

deformational events may result in a complex combination of sandstone-hosted and

basement-hosted orebodies within a deposit, and/or the superposition of orebodies

formed at different times.

4.7 Acknowledgements

This research is financially supported partly by the TGI-4 (Targeted Geoscience

Initiative Phase 4) uranium ore systems project from Natural Resources of Canada, and

partly by Cameco Corp., through grants to Chi and Bethune. The authors are indebted to

Cameco Corp. for providing the modelling parameters and the permission to publish the

results.

Page 164: Geometric and hydrodynamic modelling of fluid-structural

150

Chapter 5 Conclusions and recommendations for future work

5.1 Conclusions

This thesis represents an investigation of the structural and related fluid controls

on uranium mineralization in the southeastern Athabasca Basin. It involves two

interrelated components of research. The first phase of research deals with the gross

geological architecture of the southeastern part of the basin and the structural setting of

uranium deposits within it. This was accomplished through 3D modelling and related

study of aeromagnetic data to evaluate the spatial configuration of the unconformity, the

nature of faults as well as associated basement topographic features. The second phase of

research is composed of numerical modelling of fluid flow related to the basement faults,

with thermal convection and compressional deformation as the driving forces,

respectively. The purpose of the research is to examine whether the reactivated basement

faults crosscutting the sub-Athabasca Basin may be reflected by the topography of the

unconformity, and to understand how these faults and the unconformity surface may have

controlled the fluid flow and uranium mineralization. The main conclusions from this

research are summarized as follows:

1. Numerous, dominantly NE-trending ridges and valleys have been revealed in

the 3D model of the sub-Athabasca unconformity in the southeastern Athabasca Basin.

2. Factors including pre-Athabasca group ductile-brittle faulting and alteration,

differential weathering and erosion, and syn- to post-Athabasca ductile-brittle

reactivation of pre-existing ductile structures, have contributed to the formation of these

topographic features.

Page 165: Geometric and hydrodynamic modelling of fluid-structural

151

3. The identified NE-trending faults based on aeromagnetic data represent

reactivated pre-Athabasca structures and were the principal control on basement

topography and uranium mineralization.

4. The most prominent ridge, which is up to 320 m high and interpreted to

represent a significant NE-trending fault zone, extends semi-continuously from the Key

Lake mine in the south to McArthur River in the north. Many large deposits and

promising prospects are located adjacent to this ridge/fault zone, which is covered by a

northeast-trending alteration corridor.

5. The NE-trending fault zones and related basement topographic features are

spatially associated with uranium deposits, indicating that both of them may have

influenced the circulation of ore-forming fluids and thus controlled alteration patterns and

mineralization.

6. During periods of tectonic quiescence, free thermal convection may have

developed in the more permeable Athabasca Group. The average width of individual

convection cells is 1.4 km, which is close to the thickness of the aquifer (1.5 km). The

width of the convection cells varies only slightly due to perturbations induced by the

presence of basement faults.

7. Fault zones with higher thermal conductivity than surrounding units can induce

subtle thermal anomalies adjacent to them, which can promote the onset of thermal

convection.

8. For models with one fault zone, an upwelling flow centre generally coincides

with the fault zone, and the location and dip angle of the fault zone has negligible effect

Page 166: Geometric and hydrodynamic modelling of fluid-structural

152

on the overall fluid flow pattern. The presence of a single fault zone does not change the

overall thermal profile nor the width of convection cells.

9. For models with two fault zones, each fault zone may coincide with an

upwelling plume, or be situated within a convection cell, respectively, depending on fault

spacing and orientation. In the latter case, the fluid can flow down into the fault zone

along one side and then flow upwards back into the sandstone along the other side. The

width of the convection cells coinciding with fault zones may increase slightly depending

on the spacing and orientation of the two fault zones.

10. Thermal convection can penetrate into the uppermost basement where its

permeability is less than two orders of magnitude lower than that of the sandstone. Under

such conditions, the fault zone may be centred on an up-flow (from the basement to the

basin), but can also correspond to a down-flow (from the basin to the basement),

depending on the thermal conductivity and their relative location within the model.

11. The interaction between thermal convection and basement faults provides

conditions for fluid exchange between the basement and the overlying sandstone,

facilitating uranium transportation and precipitation. Both basement-hosted and

sandstone-hosted uranium mineralization may be generated in the vicinity of basement

faults.

12. Under the conditions of compressional deformation, the fluid flow patterns are

sensitive to the degree of bulk shortening. At relatively low degrees of bulk shortening,

fluid flows upwards into the sandstone along the fault zone, whereas at high degrees of

bulk shortening, fluid flows downwards into the basement along the fault zone.

Page 167: Geometric and hydrodynamic modelling of fluid-structural

153

13. The variation of the dip angle and amount of pre-existing fault offset has little

effect on the strain distribution and related fluid flow patterns during compressional

deformation.

14. During tectonically active periods, both sandstone-hosted and basement-

hosted mineralization may be potentially formed in a single fault system at different

stages of bulk shortening in a unified compressional stress regime. Alternatively, the

sandstone-hosted and basement-hosted orebodies in a single deposit and/or the

superposition of orebodies, may be generated at different times, in relation to multiple,

discrete, far-field tectonic events.

15. The modelling results presented in this study suggest that formation of

unconformity-related uranium deposits in the Athabasca Basin may have been related to

alternately thermally-driven and deformation-driven fluid flow, both closely associated

with the reactivated basement structures.

5.2 Future research

Although this study draws some important conclusions regarding the fluid-

structural relationships and the formation of unconformity-related uranium deposits in the

Athabasca Basin, further in-depth investigation of certain aspects is still needed to better

understand the interaction between basement faults and fluid flow.

1. Coupled modelling of deformation, heat transport and fluid flow: Due to the

limits of both research time and calculation speed of computers, this study has not

included simulations coupling deformation-driven fluid flow and heat transfer. Coupled

Page 168: Geometric and hydrodynamic modelling of fluid-structural

154

modelling of deformation, heat transport and fluid flow may provide a more realistic

understanding of how structures affect fluid flow and thus the mineralization.

2. 3D modelling of fluid flow: Again due to time limits, this study has not

incorporated variation in the strikes of faults. In the future, more realistic results could be

provided by varying the strikes of fault zones with respect to the direction of maximum

compression and thereby modelling within a 3D framework.

3. Deposit-scale 3D modelling: Coupled 3D modelling of deformation, heat

transport and fluid flow for a specific deposit would determine whether the general

findings based on the 2D conceptual models are applicable to the detailed deposit-scale

modelling.

4. The physical 3D model of the southeastern part of the basin provided a good

conceptual understanding/confirmation of the close spatial association between structures

(faults), fluids and mineralization. With further exploration and as more drill-hole data

becomes available, models with this level of detail could be constructed for currently

lesser known areas of the basin, which could greatly assist future exploration strategies in

those areas.

Page 169: Geometric and hydrodynamic modelling of fluid-structural

155

References

Adlakha, E., Hattori, K., Zaluski, G., Kotzer, T., Potter, E.G., 2013. Alteration within the

basement rocks associated with the P2 fault and the McArthur River uranium

deposit, Athabasca Basin. Geological Survey of Canada Open File Report 7462,

33.

Alexandre, P., Kyser, K., Thomas, D., Polito, P., Marlat, J., 2009. Geochronology of

unconformity-related uranium deposits in the Athabasca Basin, Saskatchewan,

Canada and their integration in the evolution of the basin. Mineralium Deposita

44 (1), 41–59.

Alexandre, P., Kyser, T.K., 2012. Modeling of the fluid flow involved in the formation of

Athabasca basin unconformity-type uranium deposits, Geological Association of

Canada - Mineralogical Association of Canada Annual Conference, St. John's, p.

3.

Andrade, N., 2002. Geology of the Cigar Lake uranium deposit, in: Andrade, N., Breton,

G., Jefferson, C.W., Thomas, D.J., Tourigny, G., Wilson, W., Yeo, G.M. (Eds.),

The Eastern Athabasca Basin and Its Uranium Deposits: Field Trip A-1

Guidebook. Geological Association of Canada and Mineralogical Association of

Canada, Saskatoon, pp. 53–72.

Annesley, I.R., Madore, C., Shi, R., Krogh, T.E., 1997. U-Pb geochronology of

thermotectonic events in the Wollaston Lake area, Wollaston Domain: a summary

of 1994–1996 results. Summary of Investigations 1997. Saskatchewan Geological

Survey, Saskatchewan Energy and Mines, Miscellaneous Report 97-4. 162–173.

Page 170: Geometric and hydrodynamic modelling of fluid-structural

156

Annesley, I.R., Madore, C., Portella, P., 2005. Geology and thermotectonic evolution of

the western margin of the Trans-Hudson Orogen: evidence from the eastern sub-

Athabasca basement, Saskatchewan. Canadian Journal of Earth Sciences 42 (4),

573–597.

Ansdell, K., MacNeil, A., Delaney, G., Hamilton, M., 2000. Rifting and development of

the Hearne craton passive margin: age constraint from the Cook Lake area,

Wollaston Domain, Trans-Hudson Orogen, Saskatchewan, GeoCanada 2000

Conference, Extended abstract.

Bächler, D., Kohl, T., Rybach, L., 2003. Impact of graben-parallel faults on hydrothermal

convection––Rhine Graben case study. Physics and Chemistry of the Earth, Parts

A/B/C 28 (9–11), 431–441.

Batycky, R., Blunt, M.J., Thiele, M.R., 1997. A 3D field-scale streamline-based reservoir

simulator. SPE Reservoir Engineering 12 (04), 246–254.

Batzle, M.L., Wang, Z., 1992. Seismic properties of pore fluids. Geophysics 57 (11),

1396-1408.

Baudemont, D., Pacquet, A., 1996. The Sue D and E uranium deposits, northern

Saskatchewan: Evidence for structurally controlled fluid circulation in the

Athabasca Basin, in: Ashton, K.E., Harper, C.T. (Eds.), MinExpo' 96

Symposium-Advances in Saskatchewan Geology and Mineral Exploration:

Saskatchewan Geological Society, Special Publication No. 14, pp. 85–94.

Bernier, S., 2002. Statigraphy of the Manitou Falls Formation along the McArthur River

high-resolution seismic survey B-B', Athabasca Basin, Saskatchewan; in

Page 171: Geometric and hydrodynamic modelling of fluid-structural

157

Summary of Investigations 2002, Volume 2, Saskatchewan Geological Survey,

Sask. Industry Resources, Misc Rep. 2002-4.2, CD-ROM, Paper D-8, 11p.

Bernier, S., 2004, Stratigraphy of the Late Paleoproterozoic Manitou Falls Formation, in

the Vicinity of the McArthur River Uranium Deposit, Athabasca Basin,

Saskatchewan, Canada. M.Sc. thesis, Laurentian University, Sudbury, Ontario.

184.

Blöcher, M.G., Cacace, M., Lewerenz, B., Zimmermann, G., 2010. Three dimensional

modelling of fractured and faulted reservoirs: Framework and implementation.

Chemie der Erde - Geochemistry 70, 145–153.

Boiron, M.C., Cathelineau, M., Richard, A., 2010. Fluid flows and metal deposition near

basement /cover unconformity: lessons and analogies from Pb-Zn-F-Ba systems

for the understanding of Proterozoic U deposits. Geofluids 10, 270–292.

Bosman, S.A., Korness, J., 2007. Building Athabasca Stratigraphy: Revising, Redefining,

and Repositioning. Summary of Investigations 2007, Volume 2, Saskatchewan

Geological Survey, Saskatchewan Ministry of Energy and Resources, Misc. Rep.

2007-4.2, CD-ROM, Paper A-8, 29.

Bosman, S.A., Ramaekers, P., 2013. The lower Athabasca Group stratigraphy in the

Patterson Lake region, Saskatchewan Geological Open House 2013 Abstract

Volume, Saskatoon, p. 3.

Bosman, S.A., Card, C.D., Slimmon, W.L., Delaney, G., Gouthas, G., Heath, P., Katona,

L., Fairclough, M., 2010. Three-dimensional model of the Athabasca Basin,

Saskatchewan. Saskatchewan Ministry of Energy and Resources, Open File 2010-

49.

Page 172: Geometric and hydrodynamic modelling of fluid-structural

158

Bosman, S.A., Card, C.D., Gouthas, G., Zmetana, D., 2012. Athabasca Basin 3-D model

(NTS 64E, 63L, 63M, and 74E to 74P); Sask. Geol. Survey, Sask. Ministry of the

Economy, 3-D File 2012-1.

Bosman, S.A., Card, C.D., Gouthas, G., Zmetana, D., 2014. Athabasca Basin 3-D model

(NTS 64E, 63L, 63M, and 74E to 74P); Sask. Geol. Survey, Sask. Ministry of the

Economy, 3-D File 2014-1.

Bronkhorst, D., Mainville, A.G., Murdock, G.M., Yesnik, L.D., 2012. McArthur River

Operation Northern Saskatchewan, Canada: National Instrument 43-101

Technical Report.

Bruneton, P., 1993. Geological environment of the Cigar Lake uranium deposit. Canadian

Journal of Earth Sciences 30 (4), 653–673.

Buckle, J.L., Coyle, M., Carson, J.M., Harvey, B.J.A., Delaney, G., 2009. Southern

Athabasca Basin Geophysical Survey, Saskatchewan. 20 printed and digital

colour maps (2 1:250,000 scale open files, 10 sheets each) and 210 digital colour

maps (21 1:50,000 scale open files, 10 sheets each).

Buckle, J.L., Coyle, M., Kiss, F., Carson, J.M., Delaney, G., Hefford, S.W., 2010.

Eastern Athabasca Basin Geophysical Survey, Saskatchewan. 30 printed and

digital colour maps (3 open files at 1:250 000 scale, 10 sheets each) and 400

digital colour maps (40 open files at 1:50 000 scale, 10 sheets each).

Caine, J.S., Evans, J.P., Forster, C.B., 1996. Fault zone architecture and permeability

structure. Geology 24 (11), 1025–1028.

Campbell-Stone, E., 2002. Realistic geologic strain rates, 2002 GSA Annual Conference,

Denver, p. 374.

Page 173: Geometric and hydrodynamic modelling of fluid-structural

159

Card, C.D., 2012. The Origins of Anomalously Graphitic Rocks and Quartzite Ridges in

the Basement to the Southeastern Athabasca Basin, Summary of Investigations

2012, Volume 2, Saskatchewan Geological Survey, Sask. Ministry of the

Economy, Misc. Rep. 2012-4.2, Paper A-6, p. 15.

Card, C.D., 2013. Altered pelitic gneisses and associated "quartzite ridges" beneath the

southeastern Athabasca Basin: alteration facies and their relationship to uranium

deposits along the Wollaston-Mudjatik transition, Summary of Investigations

2013, Volume 2, Saskatchewan Geological Survey, Sask. Ministry of the

Economy, Misc. Rep. 2013-4.2, Paper A-4, p. 23.

Card, C.D., Harper, C.T., Barsi, N., Lesperance, J., Smith, J.S., 2006. Investigation of the

Wollaston-Mudjatik transition, Charcoal Lake and Cochrane River (parts of NTS

64L-9, -10, -11, -14, -15, and -16), Summary of Investigations 2006, Volume 2,

Saskatchewan Geological Survey, Sask. Industry Resources, Misc. Rep. 2006-4.2,

CD-ROM, Paper A-5, p. 16.

Card, C.D., Pan, D., Stern, R.A., Rayner, N., 2007a. New insights into the geological

history of the basement rocks to the southwestern Athabasca Basin, Saskatchewan

and Alberta, in: Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV: Geology and

Uranium EXploration TECHnology of the Proterozoic Athabasca Basin,

Saskatchewan and Alberta: Geological Survey of Canada, Bulletin 588, (also

Geological Association of Canada, Mineral Deposits Division, Special

Publication 4; Saskatchewan Geological Society, Special Publication 18), pp.

119–134.

Page 174: Geometric and hydrodynamic modelling of fluid-structural

160

Card, C.D., Pana, D., Portella, P., Thomas, D.J., Annesley, I.R., 2007b. Basement rocks

to the Athabasca Basin, Saskatchewan and Alberta, in: Jefferson, C.W., Delaney,

G. (Eds.), EXTECH IV: Geology and Uranium EXploration TECHnology of the

Proterozoic Athabasca Basin, Saskatchewan and Alberta: Geological Survey of

Canada, Bulletin 588, (also Geological Association of Canada, Mineral Deposits

Division, Special Publication 4; Saskatchewan Geological Society, Special

Publication 18), pp. 69–87.

Card, C.D., Bosman, S.A., Slimmon, W.L., Delaney, G., Heath, P., Gouthas, G.,

Fairclough, M., 2010. Enhanced geophysical images and multi-scale edge (worm)

analysis for the Athabasca region; Sask. Ministry of Energy and Resources, Open

File 2010-46.

Carl, C., Ho hndorf, A., Pechmann, E.V., Strnad, J.G., Ruhrmann, G., 1988.

Geochronology of the key lake uranium deposit, Saskatchewan, Canada.

Chemical Geology 70 (1–2), 133.

Chamberlain, K.R., Schmitt, A.K., Swapp, S.M., Harrison, T.M., Swoboda-Colberg, N.,

Bleeker, W., Peterson, T.D., Jefferson, C.W., Khudoley, A.K., 2010. In situ U–Pb

SIMS (IN-SIMS) micro-baddeleyite dating of mafic rocks: Method with examples.

Precambrian Res 183 (3), 379–387.

Chi, G., Xue, C., 2011. An overview of hydrodynamic studies of mineralization.

Geoscience Frontiers 2 (3), 423–438.

Chi, G., Bosman, S.A., Card, C.D., 2013. Numerical modeling of fluid pressure regime in

the Athabasca basin and implications for fluid flow models related to the

Page 175: Geometric and hydrodynamic modelling of fluid-structural

161

unconformity-type uranium mineralization. Journal of Geochemical Exploration

125, 8–19.

Chi, G., Li, Z., Bethune, K.M., 2014. Numerical modeling of hydrocarbon generation in

the Douglas Formation of the Athabasca basin (Canada) and implications for

unconformity-related uranium mineralization. Journal of Geochemical

Exploration 144 (Part A), 37–48.

Chi, G., Chu, H., Scott, R., Li, Z., 2015. Basin-scale hydrodynamic and fluid P-T-X

characterization of the Athabasca Basin (Canada) and significance for

unconformity-related U mineralization, The 13th SGA meeting 2015. Society for

Geology Applied to Mineral Deposits, Abstract Volume, Nancy, France, 24-27

August 2015.

Chu, H., Chi, G., 2016. Thermal profiles inferred from fluid inclusion and illite

geothermometry from sandstones of the Athabasca basin: Implications for fluid

flow and unconformity-related uranium mineralization. Ore Geology Reviews 75,

284-303.

Cloutier, J., Kyser, K., Olivo, G.R., Brisbin, D., 2011. Geochemical, isotopic, and

geochronlologic constraints on the formation of the Eagle Point basement-hosted

uranium deposit, Athabasca Basin, Saskatchewan, Canada and recent

remobilization of primary uraninite in secondary structures. Mineralium Deposita

46 (1), 35–56.

Collier, B., Yeo, G.M., 2001. Stratigraphy of the Paleoproterozoic Manitou Falls B

member at the Deilmann Pit, Key Lake, Saskatchewan, Summary of

Page 176: Geometric and hydrodynamic modelling of fluid-structural

162

Investigations, 2001, Volume 2: Saskatchewan Geological Survey, Saskatchewan

Energy and Mines, Report 2001-4.2, pp. 297–305.

Cox, S.F., 2005. Coupling between deformation, fluid pressures, and fluid flow in ore-

producing hydrothermal systems at depth in the crust. Society of Economic

Geologists : Littleton, CO, United States, United States, pp. 39–75.

Cox, S.F., Knackstedt, M.A., Braun, J., 2001. Principles of Structural Control on

Permeability and Fluid Flow in Hydrothermal Systems. Reviews in Economic

Geology 14, 1-24.

Creaser, R.A., Stasiuk, L.D., 2007. Depositional age of the Douglas Formation, northern

Saskatchewan, determined by Re–Os geochronology. EXTECH IV: Geology and

Uranium EXploration TECHnology of the Proterozoic Athabasca Basin,

Saskatchewan and Alberta: Geological Survey of Canada, Bulletin 588, (also

Geological Association of Canada, Mineral Deposits Division, Special

Publication 4; Saskatchewan Geological Society, Special Publication 18), 341–

346.

Cui, T., Yang, J., Samson, I.M., 2010. Numerical modeling of hydrothermal fluid flow in

the Paleoproterozoic Thelon Basin, Nunavut, Canada. Journal of Geochemical

Exploration 106 (1–3), 69–76.

Cui, T., Yang, J.W., Samson, I.M., 2012a. Solute transport across basement/cover

interfaces by buoyancy-driven thermohaline convection: Implications for the

formation of unconformity-related uranium deposits. American Journal of Science

312 (9), 994–1027.

Page 177: Geometric and hydrodynamic modelling of fluid-structural

163

Cui, T., Yang, J.W., Samson, I.M., 2012b. Tectonic deformation and fluid flow:

Implications for the formation of unconformity-related uranium deposits.

Economic Geology 107 (1), 147–163.

Cumming, G.L., Krstic, D., 1992. The age of unconformity-related uranium

mineralization in the Athabasca Basin, northern Saskatchewan. Canadian Journal

of Earth Sciences 29 (8), 1623–1639.

Cundall, P., 1988. A microcomputer program for modelling large-strain plasticity

problems, Proc. 6th Int Conf on Numerical Methods in Geomechanics, Innsbruck,

Austria, 1988.

Cuney, M., Brouand, M., Cathelineau, M., Derome, D., Freiberger, R., Hecht, L., Kister,

P., Lobaev, V., Lorilleux, G., Peiffert, C., 2003. What parameters control the high

grade-large tonnage of the Proterozoic unconformity related uranium deposits,

Uranium Geochemistry 2003, International Conference, Nancy, France, pp. 123–

126.

Davidson, A., 2008. Late Paleoproterozoic to mid-Neoproterozoic history of northern

Laurentia: An overview of central Rodinia. Precambrian Research 160 (1–2), 5–

22.

Davies, J., 1998, The origin, structural style, and reactivation history of the Tabbernor

Fault Zone,. Saskatchewan, Canada. McGill University, Montreal.

De Veslud, C.L.C., Cuney, M., Lorilleux, G., Royer, J.-J., Jébrak, M., 2009. 3D modeling

of uranium-bearing solution-collapse breccias in Proterozoic sandstones

(Athabasca Basin, Canada)—Metallogenic interpretations. Computers &

Geosciences 35 (1), 92–107.

Page 178: Geometric and hydrodynamic modelling of fluid-structural

164

Derome, D., Cuney, M., Cathelineau, M., Fabre, C., Dubessy, J., Bruneton, P., Hubert, A.,

2003. A detailed fluid inclusion study in silicified breccias from the Kombolgie

sandstones (Northern Territory, Australia): inferences for the genesis of middle-

Proterozoic unconformity-type uranium deposits. Journal of Geochemical

Exploration 80 (2–3), 259–275.

Derome, D., Cathelineau, M., Cuney, M., Fabre, C., Lhomme, T., Banks, D.A., 2005.

Mixing of sodic and calcic brines and uranium deposition at McArthur River,

Saskatchewan, Canada: A Raman and laser-induced breakdown spectroscopic

study of fluid inclusions. Economic Geology 100 (8), 1529–1545.

Doglioni, C., Barba, S., Carminati, E., Riguzzi, F., 2011. Role of the brittle–ductile

transition on fault activation. Physics of the Earth and Planetary Interiors 184 (3–

4), 160–171.

Doglioni, C., Barba, S., Carminati, E., Riguzzi, F., 2014. Fault on–off versus coseismic

fluids reaction. Geoscience Frontiers 5 (6), 767–780.

Earle, S.A.M., Sopuck, V.J., 1989. Regional lithogeochemistry of the eastern part of the

Athabasca Basin uranium province, Saskatchewan, Canada. Uranium Resources

and Geology of North America: International Atomic Energy Agency, Technical

Document TECDOC-500, 263–296.

Fayek, M., Kyser, T.K., 1997. Characterization of multiple fluid-flow events and rare-

earth-element mobility associated with formation of unconformity-type uranium

deposits in the Athabasca Basin, Saskatchewan. Canadian Mineralogist 35, 627–

658.

Page 179: Geometric and hydrodynamic modelling of fluid-structural

165

Feltrin, L., Baker, T., Oliver, N., Scott, M., Wilkinson, K., Fitzell, M., Dixon, O., Bertelli,

M., 2008. Using Geomodelling and Geophysical Inversion to Evaluate the

Geological Controls on Low‐Sulphidation Epithermal Au‐Ag mineralisation in

the Drummond and Bowen Basins, Australia, AIP Conference Proceedings, p.

123.

Feltrin, L., McLellan, J.G., Oliver, N.H.S., 2009. Modelling the giant, Zn–Pb–Ag

Century deposit, Queensland, Australia. Computers & Geosciences 35 (1), 108–

133.

Ford, A., Blenkinsop, T.G., McLellan, J.G., 2009. Factors affecting fluid flow in strike–

slip fault systems: coupled deformation and fluid flow modelling with application

to the western Mount Isa Inlier, Australia. Geofluids 9 (1), 2–23.

Fowler, C.M., Stead, D., Pandit, B.I., Janser, B.W., Nisbet, E.G., Nover, G., 2005. A

database of physical properties of rocks from the Trans-Hudson Orogen, Canada.

Canadian Journal of Earth Sciences 42 (4), 555–572.

Garven, G., Bull, S., Large, R., 2001. Hydrothermal fluid flow models of stratiform ore

genesis in the McArthur Basin, Northern Territory, Australia. Geofluids 1 (4),

289–311.

Gradstein, F.M., Ogg, J.G., Schmitz, M., Ogg, G., 2012. The Geologic Time Scale 2012

2-Volume Set. elsevier.

Guillou-Frottier, L., Carrė, C., Bourgine, B., Bouchot, V., Genter, A., 2013. Structure of

hydrothermal convection in the Upper Rhine Graben as inferred from corrected

temperature data and basin-scale numerical models. Journal of Volcanology and

Geothermal Research 256, 29–49.

Page 180: Geometric and hydrodynamic modelling of fluid-structural

166

Györfi, I., Hajnal, Z., White, D.J., Takács, E., Reilkoff, B., Annesley, I.R., Powell, B.,

Koch, R., 2007. High-resolution seismic survey from the McArthur River region :

contributions to mapping of the complex P2 uranium ore zone , Athabasca Basin ,

Saskatchewan, in: Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV: Geology

and Uranium EXploration TECHnology of the Proterozoic Athabasca Basin,

Saskatchewan and Alberta: Geological Survey of Canada, Bulletin 588, (also

Geological Association of Canada, Mineral Deposits Division, Special

Publication 4; Saskatchewan Geological Society, Special Publication 18), pp.

397-412.

Hajnal, Z., Takács, E., White, D.J., Györfi, I., Powell, B., Koch, R., 2007. Regional

seismic signature of the basement and crust beneath the McArthur River mine

district, Athabasca Basin, Saskatchewan, in: Jefferson, C.W., Delaney, G. (Eds.),

EXTECH IV: Geology and Uranium EXploration TECHnology of the Proterozoic

Athabasca Basin, Saskatchewan and Alberta: Geological Survey of Canada,

Bulletin 588, (also Geological Association of Canada, Mineral Deposits Division,

Special Publication 4; Saskatchewan Geological Society, Special Publication 18),

pp. 389–396.

Harvey, S.E., 2004, Structural geology and its relationship to paleotopography, alteration,

and uranium mineralization in the Deilmann orebody, Key Lake Mine,

Saskatchewan. M.Sc. thesis, University of Regina, Regina, Saskatchewan. 168 p.

Harvey, S.E., Bethune, K.M., 2007. Context of the Deilmann orebody, Key Lake mine,

Saskatchewan, in: Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV: Geology

and Uranium EXploration TECHnology of the Proterozoic Athabasca Basin,

Page 181: Geometric and hydrodynamic modelling of fluid-structural

167

Saskatchewan and Alberta: Geological Survey of Canada, Bulletin 588, (also

Geological Association of Canada, Mineral Deposits Division, Special

Publication 4; Saskatchewan Geological Society, Special Publication 18), pp.

249–266.

Hecht, L., Cuney, M., 2000. Hydrothermal alteration of monazite in the Precambrian

crystalline basement of the Athabasca Basin (Saskatchewan, Canada):

implications for the formation of unconformity-related uranium deposits.

Mineralium Deposita 35 (8), 791–795.

Heise, W., Caldwell, T.G., Bibby, H.M., Bannister, S.C., 2008. Three-dimensional

modelling of magnetotelluric data from the Rotokawa geothermal field, Taupo

Volcanic Zone, New Zealand. Geophysical Journal International 173 (2), 740–750.

Hiatt, E.E., Kyser, T.K., 2007. Sequence stratigraphy, hydrostratigraphy, and

mineralizing fluid flow in the Proterozoic Manitou Falls Formation, eastern

Athabasca Basin, Saskatchewan. 489–506.

Hoeve, J., Quirt, D., 1984. Mineralization and host rock alteration in relation to clay

mineral diagenesis and evolution of the Middle-Proterozoic, Athabasca Basin,

northern Saskatchewan, Canada. Saskatchewan Research Council, p. 187.

Hoeve, J., Sibbald, T., 1978. On the genesis of Rabbit Lake and other unconformity-type

uranium deposits in northern Saskatchewan, Canada. Economic Geology 73,

1450–1473.

Hoeve, J., Sibbald, T., Ramaekers, P., Lewry, J., 1980. Athabasca Basin unconformity-

type uranium deposits. A special class of sandstone-type deposits, in: Ferguson, J.,

Goleby, A.B. (Eds.), Uranium in the Pine Creek Geosyncline, pp. 575–594.

Page 182: Geometric and hydrodynamic modelling of fluid-structural

168

Hoffman, P.F., 1988. United Plates of America, the Birth of a Craton - Early Proterozoic

Assembly and Growth of Laurentia. Annual Review of Earth and Planetary

Sciences 16, 543–603.

Ingebritsen, S.E., Appold, M.S., 2012. The Physical Hydrogeology of Ore Deposits.

Economic Geology 107 (4), 559-584.

Ingebritsen, S.E., Sanford, W.E., 2006. Groundwater in geologic processes, Second

Edition ed. Cambridge University Press, Cambridge, UK.

Itasca, 2012. FLAC3D: Fast Lagrangian Analysis of Continua in 3 Dimensions, User

Manual, Version 5.0, Fifth ed. Itasca Consulting Group Inc., Minneapolis,

Minnesota.

Jefferson, C., Percival, J.B., Bernier, S., Cutts, C., Drever, G., Jiricka, D., Long, D.,

McHardy, S., Quirt, D., Ramaekers, P., Wasyliuk, K., Yeo, G.M., 2001.

Lithostratigraphy and mineralogy in the Eastern Athabasca Basin, northern

Saskatchewan - progress in year 2 of EXTECH IV, Summary of Investigations

2001, Volume 2, Part B, EXTECH IV Athabasca Uranium Multidisciplinary

Study: Saskatchewan Geological Survey, Saskatchewan Energy and Mines,

Miscellaneous Report 2001-4.2b (CD-ROM), pp. 272–290.

Jefferson, C., Thomas, D., Gandhi, S., Ramaekers, P., Delaney, G., Brisbin, D., Cutts, C.,

Portella, P., Olson, R., 2007. Unconformity-associated uranium deposits of the

Athabasca Basin, Saskatchewan and Alberta, in: Jefferson, C.W., Delaney, G.

(Eds.), EXTECH IV: Geology and Uranium EXploration TECHnology of the

Proterozoic Athabasca Basin, Saskatchewan and Alberta: Geological Survey of

Canada, Bulletin 588, (also Geological Association of Canada, Mineral Deposits

Page 183: Geometric and hydrodynamic modelling of fluid-structural

169

Division, Special Publication 4; Saskatchewan Geological Society, Special

Publication 18), pp. 23–67.

Kerr, W.C., 2010. The discovery of the Phoenix Deposit: a new high-grade, Athabasca

Basin unconformity-type uranium deposit, Saskatchewan, Canada. Special

Publication [Society of Economic Geologists [U. S.] 15, Vol. II, 703–728.

Kestin, J., Khalifa, H.E., Abe, Y., Grimes, C.E., Sookiazian, H., Wakeham, W.A., 1978.

Effect of pressure on the viscosity of aqueous sodium chloride solutions in the

temperature range 20-150. degree. C. Journal of Chemical and Engineering Data

23 (4), 328-336.

Knipping, H., 1974. The concepts of supergene versus hypogene emplacement of

uranium at Rabbit Lake, Saskatchewan, Canada, in: Parslow, G.R. (Ed.),

Saskatchewan Geological Society Special Publication Number 2: FUELS: a

geological appraisal. Proceedings of a Symposium, 7-8 November, 1974, pp.

119–124.

Kotzer, T.G., Kyser, T.K., 1995. Petrogenesis of the Proterozoic Athabasca Basin,

northern Saskatchewan, Canada, and its relation to diagenesis, hydrothermal

uranium mineralization and paleohydrogeology. Chemical Geology 120, 45–89.

Koziy, L., Bull, S., Large, R., Selley, D., 2009. Salt as a fluid driver, and basement as a

metal source, for stratiform sediment-hosted copper deposits. Geology 37 (12),

1107–1110.

Kühn, M., Dobert, F., Gessner, K., 2006. Numerical investigation of the effect of

heterogeneous permeability distributions on free convection in the hydrothermal

Page 184: Geometric and hydrodynamic modelling of fluid-structural

170

system at Mount Isa, Australia. Earth and Planetary Science Letters 244 (3–4),

655–671.

Kyser, K., Cuney, M., 2009. Unconformity-related uranium deposits, in: Cuney, M.,

Kyser, K. (Eds.), Recent and Not-So-Recent Developments in Uranium Deposits

and Implications for Exploration. Mineralogical Association of Canada, Short

Course Series, pp. 161–219.

Kyser, K., Hiatt, E., Renac, C., Durocher, K., Holk, G., Deckart, K., 2000. Diagenetic

fluids in Paleo- and Mesoproterozoic sedimentary basins and their implications

for long protracted fluid histories, in: Kyser, K. (Ed.), Fluids and Basin Evolution:

Mineralogical Association of Canada Short Course, pp. 225–262.

Langford, F., 1978. Origin of unconformity-type pitchblende deposits in the Athabasca

Basin of Saskatchewan, in: Kimberly, M.M. (Ed.), Uranium Deposits: Their

Mineralogy and Origin, Short Course. Mineralogical Association of Canada,

University of Toronto, pp. 485–499.

Lapwood, E., 1948. Convection of a fluid in a porous medium. Proceedings of the

Cambridge 44, 508–521.

Large, R., Bull, S., McGoldrick, P., Walters, S., Derrick, G., Carr, G., 2005. Stratiform

and strata-bound Zn-Pb-Ag deposits in Proterozoic sedimentary basins, Northern

Australia. Economic Geology 100 (Anniversary Volume), 931–963.

Leader, L.D., Wilson, C.J.L., Robinson, J.A., 2013. Structural Constraints and Numerical

Simulation of Strain Localization in the Bendigo Goldfield, Victoria, Australia.

Economic Geology 108 (2), 279–307.

Page 185: Geometric and hydrodynamic modelling of fluid-structural

171

LeCheminant, A.N., Heaman, L.M., 1989. Mackenzie Igneous Events, Canada - Middle

Proterozoic Hotspot Magmatism Associated with Ocean Opening. Earth and

Planetary Science Letters 96 (1–2), 38–48.

Lewry, J.F., Sibbald, T., 1977. Variation in lithology and tectonometamorphic

relationships in the Precambrian basement of northern Saskatchewan. Canadian

Journal of Earth Sciences 14, 1453–1467.

Lewry, J.F., Sibbald, T.I., 1980. Thermotectonic evolution of the Churchill Province in

northern Saskatchewan. Tectonophysics 68 (1), 45–82.

Li, Z., Bethune, K.M., Chi, G., Bosman, S.A., Card, C.D., 2013. Preliminary 3D

modelling and structural interpretation of southeastern Athabasca Basin.

Geological Survey of Canada, Open File 7426.

Li, Z., Bethune, K.M., Chi, G., Bosman, S.A., Card, C.D., 2015a. Topographic features

of the sub-Athabasca Group unconformity surface in the southeastern Athabasca

Basin and their relationship to uranium ore deposits. Canadian Journal of Earth

Sciences 52 (10), 903–920.

Li, Z., Chi, G., Bethune, K.M., Bosman, S.A., Card, C.D., 2015b. Geometric and

Hydrodynamic Modelling and Fluid-structural Relationships in the Southeastern

Athabasca Basin and Significance for Uranium Mineralization, in: Potter, E.G.,

Wright, D.M. (Eds.), Targeted Geoscience Initiative 4: unconformity-related

uranium systems. Geological Survey of Canada, Open File 7791, pp. 103–114.

Liu, L., Zhao, Y., Sun, T., 2012. 3D computational shape- and cooling process-modeling

of magmatic intrusion and its implication for genesis and exploration of intrusion-

Page 186: Geometric and hydrodynamic modelling of fluid-structural

172

related ore deposits: An example from the Yueshan intrusion in Anqing, China.

Tectonophysics 526–529, 110–123.

Liu, Y., Chi, G., Bethune, K.M., Dube, B., 2011. Fluid dynamics and fluid-structural

relationships in the Red Lake mine trend, Red Lake greenstone belt, Ontario,

Canada. Geofluids 11, 260–279.

Long, D., 2001. Architecture of Late Proterozoic fluvial strata in the Manitou Falls

Formation at McClean Lake and Key Lake, eastern Athabasca Group, Northern

Saskatchewan. Summary of Investigation 2001, Volume 2; Saskatchewan

Geological Survey, Saskatchewan Energy and Mines, Miscellaneous Report

2001-4.2, 306–312.

Long, D.G.F., 2007. Topographic influences on the sedimentology of the Manitou Falls

Formation, eastern Athabasca Basin, Saskatchewan, in: Jefferson, C.W., Delaney,

G. (Eds.), EXTECH IV: Geology and Uranium EXploration TECHnology of the

Proterozoic Athabasca Basin, Saskatchewan and Alberta: Geological Survey of

Canada, Bulletin 588, (also Geological Association of Canada, Mineral Deposits

Division, Special Publication 4; Saskatchewan Geological Society, Special

Publication 18), pp. 267–280.

Macdonald, C.C., 1980, Mineralogy and geochemistry of a precambrian regolith in the

Athabasca Basin. M.Sc. thesis, University of Saskatchewan, Saskatoon,

Saskatchewan. p. 151.

Macdonald, C.C., 1985. Mineralogy and geochemistry of the sub-Athabasca regolith near

Wollaston Lake. Geology of Uranium Deposits: Canadian Institute of Mining and

Metallurgy Special 32, 155–158.

Page 187: Geometric and hydrodynamic modelling of fluid-structural

173

MacDougall, D.G., Heaman, L.M., 2002. Diabase sill complexes in Northern

Saskatchewan: Analogues for the LITHOPROBE Trans-Hudson Orogen Transect

(THOT) Wollaston Lake S2b Reflector, Geological Association of Canada–

Mineralogical Association of Canada, Joint Annual Meeting, Program with

Abstracts, Saskatoon, p. 70.

MacDougall, D.G., William, D.H., 1993. The Moore Lakes Complex, Neohelikian

Olivine Diabase Lopoliths in the Athabasca Group (Part of NTS 74H-6 and -7).

Summary of Investigations, Saskatchewan Geological Survey, Saskatchewan

Energy Mines Miscellaneous Report 93-4.

Madore, C., Annesley, I., Wheatley, K., 2000. Petrogenesis, age, and uranium fertility of

peraluminous leucogranites and pegmatites of the McClean Lake / Sue and Key

Lake / P-Patch deposit areas, Saskatchewan. GAC-MAC Program with Abstracts

25 (1041), 4.

Malkovsky, V., Pek, A., 2015. Onset of Fault-Bounded Free Thermal Convection in a

Fluid-Saturated Horizontal Permeable Porous Layer. Transport in Porous Media

100(1), 1–15.

Mallet, J.L., 1992. GOCAD: A Computer Aided Design Program for Geological

Applications, in: Turner, A.K. (Ed.), Three-Dimensional Modeling with

Geoscientific Information Systems. Springer Netherlands, pp. 123-141.

Mallet, J.L., 2002. Geomodeling. Oxford University Press, New York.

Manning, C.E., Ingebritsen, S.E., 1999. Permeability of the continental crust:

Implications of geothermal data and metamorphic systems. Reviews of

Geophysics 37 (1), 127–150.

Page 188: Geometric and hydrodynamic modelling of fluid-structural

174

Marlatt, J., McGill, B., Matthews, R., Sopuck, V., Pollock, G., 1992. The discovery of the

McArthur River uranium deposit, Saskatchewan, Canada. new developments in

uranium exploration, resources, production and demand: International Atomic

Energy Agency - Nuclear Energy Agency of the OECD, Vienna, 26-29 August,

1991, Technical Committee Meeting, Proceedings, IAEA TECDOC-650, 118–

127.

Matthäi, S.K., Heinrich, C.A., Driesner, T., 2004. Is the Mount Isa copper deposit the

product of forced brine convection in the footwall of a major reverse fault?

Geology 32 (4), 357–360.

McGill, B.D., Marlatt, J.L., Matthews, R.B., Sopuck, V.J., Homeniuk, L.A., Hubregtse,

J.J., 1993. The P2 North uranium deposit, Saskatchewan, Canada. Exploration

and Mining Geology 2 (4), 321–331.

McLellan, J.G., Oliver, N.H.S., Schaubs, P.M., 2004. Fluid flow in extensional

environments; numerical modelling with an application to Hamersley iron ores.

Journal of Structural Geology 26 (6–7), 1157–1171.

McLellan, J.G., Oliver, N.H.S., Hobbs, B.E., 2006. The relative effects of deformation

and thermal advection on fluid pathways in basin-related mineralization. Journal

of Geochemical Exploration 89 (1–3), 271–275.

Mercadier, J., Richard, A., Boiron, M.C., Cathelineau, M., Cuney, M., 2010. Migration of

brines in the basement rocks of the Athabasca Basin through microfracture

networks (P-Patch U deposit, Canada). Lithos 115 (1–4), 121-136.

Mercadier, J., Cuney, M., Cathelineau, M., Lacorde, M., 2011. U redox fronts and

kaolinisation in basement-hosted unconformity-related U ores of the Athabasca

Page 189: Geometric and hydrodynamic modelling of fluid-structural

175

Basin (Canada): late U remobilisation by meteoric fluids. Mineralium Deposita 46

(2), 105–135.

Mercadier, J., Richard, A., Cathelineau, M., 2012. Boron- and magnesium-rich marine

brines at the origin of giant unconformity-related uranium deposits: δ11B evidence

from Mg-tourmalines. Geology 40 (3), 231–234.

Mercadier, J., Annesley, I.R., McKechnie, C.L., Bogdan, T.S., Creighton, S., 2013.

Magmatic and Metamorphic Uraninite Mineralization in the Western Margin of

the Trans-Hudson Orogen (Saskatchewan, Canada): A Uranium Source for

Unconformity-Related Uranium Deposits? Economic Geology 108 (5), 1037–

1065.

Mwenifumbo, C.J., Percival, J.B., Bernius, G., Elliott, B., Jefferson, C.W., Wasyliuk, K.,

Drever, G., 2007. Comparison of geophysical, mineralogical and stratigraphic

attributes in drill holes MAC-218 and RL-88, McArthur River uranium camp,

Athabasca Basin, Saskatchewan. In Jefferson, C.W., and Delaney, G., eds.,

EXTECH IV: Geology and Uranium EXploration TECHnology of the Proterozoic

Athabasca Basin, Saskatchewan and Alberta: Geological Survey of Canada,

Bulletin 588, (also Geological Association of Canada, Mineral Deposits Division,

Special Publication 4; Saskatchewan Geological Society, Special Publication 18)

p. 507–520.

Oliver, N.H.S., Pearson, P.J., Holcombe, R.J., Ord, A., 1999. Mary Kathleen

metamorphic–hydrothermal uranium – rare-earth element deposit: ore genesis and

numerical model of coupled deformation and fluid flow. Australian Journal of

Earth Sciences 46 (3), 467–483.

Page 190: Geometric and hydrodynamic modelling of fluid-structural

176

Oliver, N.H., McLellan, J.G., Hobbs, B.E., Cleverley, J.S., Ord, A., Feltrin, L., 2006.

Numerical models of extensional deformation, heat transfer, and fluid flows

across basement-cover interfaces during basin-related mineralization. Economic

Geology 101, 1–31.

Ord, A., Oliver, N., 1997. Mechanical controls on fluid flow during regional

metamorphism: some numerical models. Journal of Metamorphic Geology 15 (3),

345–359.

Orrell, S.E., Bickford, M.E., Lewry, J.F., 1999. Crustal evolution and age of

thermotectonic reworking in the western hinterland of the Trans-Hudson Orogen,

northern Saskatchewan. Precambrian Research 95 (3–4), 187–223.

Pagel, M., 1975. Détermination des conditions physico-chimiques de la silicification

diagénétique des grès Athabasca (Canada) au moyen des inclusions fluids.

Comptes Rendus Série D 280, 2301–2304.

Pagel, M., Poty, B., Sheppard, S.M.F., 1980. Contribution to some Saskatchewan

uranium deposits mainly from fluid inclusion and isotopic data, in: Ferguson, S.,

Goleby, A. (Eds.), Uranium in the Pine Creek Geosyncline. International Atomic

Energy Agency (IAEA), Vienna, pp. 639–654.

Park, Y.J., Sudicky, E.A., Sykes, J.F., 2009. Effects of shield brine on the safe disposal of

waste in deep geologic environments. Advances in Water Resources 32 (8),

1352–1358.

Pfiffner, O., Ramsay, J., 1982. Constraints on geological strain rates: arguments from

finite strain states of naturally deformed rocks. Journal of Geophysical Research:

Solid Earth 87 (B1), 311–321.

Page 191: Geometric and hydrodynamic modelling of fluid-structural

177

Pilkington, M., 1989. Variable-Depth Magnetization Mapping - Application to the

Athabasca Basin, Northern Alberta and Saskatchewan, Canada. Geophysics 54 (9),

1164–1173.

Pilkington, M., Roest, W., 1992. Draping Aeromagnetic Data in Areas of Rugged

Topography. Journal of Applied Geophysics 29 (2), 135–142.

Portella, P., Annesley, I.R., 2000. Paleoprotcrozoic thermotcctonic evolution of the

eastern sub-Athabasca basement, northern Saskatehcwan: Integrated geophysical

and geological data. , Summary of Investigations 2000. Volume 2. Saskatchewan

Geological Survey, Saskatchewan Energy Mines. Misc. Rep. 2000-4.2.

Rabinowicz, M., Boulègue, J., Genthon, P., 1998. Two‐and three‐dimensional modeling

of hydrothermal convection in the sedimented Middle Valley segment, Juan de

Fuca Ridge. Journal of Geophysical Research: Solid Earth (1978–2012) 103

(B10), 24045-24065.

Raffensperger, J.P., Garven, G., 1995. The Formation of Unconformity-Type Uranium

Ore-Deposits. 1. Coupled Groundwater-Flow and Heat-Transport Modeling.

American Journal of Science 295 (5), 581–636.

Raffensperger, J.P., Vlassopoulos, D., 1999. The potential for free and mixed convection

in sedimentary basins. Hydrogeology Journal 7 (6), 505–520.

Rainbird, R.H., Stern, R.A., Rayner, N., Jefferson, C.W., 2007. Age, provenance, and

regional correlation of the Athabasca Group, Saskatchewan and Alberta,

constrained by igneous and detrital zircon geochronology, in: Jefferson, C.W.,

Delaney, G. (Eds.), EXTECH IV: Geology and Uranium EXploration

TECHnology of the Proterozoic Athabasca Basin, Saskatchewan and Alberta:

Page 192: Geometric and hydrodynamic modelling of fluid-structural

178

Geological Survey of Canada, Bulletin 588, (also Geological Association of

Canada, Mineral Deposits Division, Special Publication 4; Saskatchewan

Geological Society, Special Publication 18), pp. 193-209.

Ramaekers, P., 2004. Development, stratigraphy and summary diagenetic history of the

Athabasca Basin, early Proterozoic of Alberta and its relation to uranium potential.

Alberta Geological Survey, Alberta Energy and Utilities Board, AGS Special

Report 62, p. 85.

Ramaekers, P., Catuneanu, O., 2004. Development and sequences of the Athabasca basin,

early Proterozoic, Saskatchewan and Alberta, Canada. The Precambrian Earth:

Tempos and Events. Developments in Precambrian Geology 12, 705–723.

Ramaekers, P., Jefferson, C.W., Yeo, G.M., Collier, B., Long, D.G.F., Drever, G.,

Mchardy, S., Jiricka, D., Cutts, C., Wheatley, K., Catuneanu, O., Bernier, S., 2007.

Revised geological map and stratigraphy of the Athabasca Group, Saskatchewan

and Alberta, in: Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV: Geology and

Uranium EXploration TECHnology of the Proterozoic Athabasca Basin,

Saskatchewan and Alberta: Geological Survey of Canada, Bulletin 588, (also

Geological Association of Canada, Mineral Deposits Division, Special

Publication 4; Saskatchewan Geological Society, Special Publication 18), pp.

155–192.

Rawling, T.J., Schaubs, P.M., Dugdale, L.J., Wilson, C.J.L., Murphy, F.C., 2006.

Application of 3D models and numerical simulations as a predictive exploration

tool in western Victoria. Australian Journal of Earth Sciences 53 (5), 825–839.

Page 193: Geometric and hydrodynamic modelling of fluid-structural

179

Reid, K.D., Ansdell, K., Jiricka, D., Witt, G., Card, C., 2014. Regional Setting, Geology,

and Paragenesis of the Centennial Unconformity-Related Uranium Deposit,

Athabasca Basin, Saskatchewan, Canada. Economic Geology 109 (3), 539–566.

Richard, A., Pettke, T., Cathelineau, M., Boiron, M.C., Mercadier, J., Cuney, M., Derome,

D., 2010. Brine-rock interaction in the Athabasca basement (McArthur River U

deposit, Canada): consequences for fluid chemistry and uranium uptake. Terra

Nova 22 (4), 303–308.

Richard, A., Banks, D.A., Mercadier, J., Boiron, M.-C., Cuney, M., Cathelineau, M.,

2011. An evaporated seawater origin for the ore-forming brines in unconformity-

related uranium deposits (Athabasca Basin, Canada): Cl/Br and δ37Cl analysis of

fluid inclusions. Geochimica et Cosmochimica Acta 75 (10), 2792–2810.

Richard, A., Boulvais, P., Mercadier, J., Boiron, M.-C., Cathelineau, M., Cuney, M.,

France-Lanord, C., 2013. From evaporated seawater to uranium-mineralizing

brines: Isotopic and trace element study of quartz–dolomite veins in the

Athabasca system. Geochimica et Cosmochimica Acta 113, 38–59.

Richard, A., Kendrick, M.A., Cathelineau, M., 2014. Noble gases (Ar, Kr, Xe) and

halogens (Cl, Br, I) in fluid inclusions from the Athabasca Basin (Canada):

Implications for unconformity-related U deposits. Precambrian Res 247, 110–125.

Richard, A., Cathelineau, M., Boiron, M.-C., Mercadier, J., Banks, D.A., Cuney, M.,

2015. Metal-rich fluid inclusions provide new insights into unconformity-related

U deposits (Athabasca Basin and Basement, Canada). Mineralium Deposita 51 (2),

249-270.

Page 194: Geometric and hydrodynamic modelling of fluid-structural

180

Robertson, R., 2006. Millennium: Cameco’s newest uranium discovery. The Northern

Miner 91, 1–14.

Schaubs, P., Fisher, L., 2009. The effects of deformation and permeability variation in

controlling the formation of uranium deposits in the Alligator Rivers region,

Northern Territory, Australia. Journal of Geochemical Exploration 101 (1), 89.

Schaubs, P.M., Ord, A., Hobbs, B.E., Annesley, I.R., Madore, C., Quirt, D., Fox, N.P.G.,

Thomas, D., Portella, P., 2002. Some thermal considerations on the formation of

unconformity related uranium deposits of the Athabasca basin, Canada based on

numerical models. GAC-MAC Abstracts 27, p. 105.

Schaubs, P.M., Ord, A., Annesley, I., Madore, C., Quirt, D., Thomas, D., Portella, P.,

2005. A comparison of McArthur River and Cigar Lake unconformity-related

uranium deposits, Saskatchewan, Canada: insight from deformation-fluid flow

models, in: Hancock, H. (Ed.), STOMP 2005, Structure, Tectonics and Ore

Mineralisation Processes: Abstract Volume. Economic Geology Research Unit,

James Cook University, Townsville, Qld, p. 121.

Schaubs, P., Rawling, T., Dugdale, J., Wilson, C., 2006. Using numerical deformation–

fluid-flow simulations to understand and target gold mineralisation around basalt

domes in the Stawell Zone, Central Victoria, Australia. Journal of Geochemical

Exploration 89 (1–3), 351–354.

Sheldon, H.A., 2009. Simulation of magmatic and metamorphic fluid production coupled

with deformation, fluid flow and heat transport. Computers & Geosciences 35

(11), 2275–2281.

Page 195: Geometric and hydrodynamic modelling of fluid-structural

181

Sibson, R.H., 1996. Structural permeability of fluid-driven fault-fracture meshes. Journal

of Structural Geology 18 (8), 1031–1042.

Sibson, R.H., 2001. Seismogenic framework for hydrothermal transport and ore

deposition. Reviews in Economic Geology 14, 25–50.

Simms, M.A., Garven, G., 2004. Thermal convection in faulted extensional sedimentary

basins: theoretical results from finite-element modeling. Geofluids 4 (2), 109–130.

Skytta, P., 2012. Crustal evolution of an ore district illustrated-4D-animation from the

Skellefte district, Sweden. Computers & Geosciences 48, 157–161.

Souche, A., Dabrowski, M., Andersen, T.B., 2014. Modeling thermal convection in

supradetachment basins: example from western Norway. Geofluids 14 (1), 58–74.

Stober, I., Bucher, K., 2007. Hydraulic properties of the crystalline basement.

Hydrogeology Journal 15 (2), 213–224.

Straus, J.M., Schubert, G., 1977. Thermal convection of water in a porous medium:

Effects of temperature- and pressure-dependent thermodynamic and transport

properties. Journal of Geophysical Research 82 (2), 325–333.

Survey, S.G., 2003. Geology, and Mineral and Petroleum Resources of Saskatchewan.

Miscellaneous Report 2003-7.

Thomas, D.J., Matthews, R.B., Sopuck, V., 2000. Athabasca Basin (Canada)

unconformity-type uranium deposits: Exploration model, current mine

developments and exploration directions. Geology and Ore deposits, 1–23.

Thomas, M.D., McHardy, S., 2007. Magnetic insights into basement geology in the area

of McArthur River uranium deposit, Athabasca Basin, Saskatchewan, in:

Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV: Geology and Uranium

Page 196: Geometric and hydrodynamic modelling of fluid-structural

182

EXploration TECHnology of the Proterozoic Athabasca Basin, Saskatchewan and

Alberta: Geological Survey of Canada, Bulletin 588, (also Geological Association

of Canada, Mineral Deposits Division, Special Publication 4; Saskatchewan

Geological Society, Special Publication 18), pp. 425–440.

Tourigny, G., Quirt, D.H., Wilson, N.S.F., Wilson, S., Breton, G., Portella, P., 2007.

Geological and structural features of the Sue C uranium deposit, McClean Lake

area, Saskatchewan, in: Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV:

Geology and Uranium EXploration TECHnology of the Proterozoic Athabasca

Basin, Saskatchewan and Alberta: Geological Survey of Canada, Bulletin 588,

(also Geological Association of Canada, Mineral Deposits Division, Special

Publication 4; Saskatchewan Geological Society, Special Publication 18), pp.

229–247.

Tran, H.T., 2001, Tectonic evolution of the Paleoproterozoic Wollaston Group in the

Cree Lake Zone, northern Saskatchewan, Canada. Ph.D, University of Regina, p.

458.

Tschirhart, V., Morris, W.A., Jefferson, C.W., 2013. Faults affecting the northeast Thelon

Basin: improved basement constraints from source edge processing of

aeromagnetic data, in: Potter, E.G., Jefferson, C.W., Quirt, D.H. (Eds.), Uranium

in Canada: Geological Environments and Exploration Developments, pp. 105–113.

Turcotte, D.L., Schubert, G., 2014. Geodynamics, 3rd ed ed. Cambridge University Press,

Cambridge, UK.

Page 197: Geometric and hydrodynamic modelling of fluid-structural

183

Upton, P., Craw, D., 2014. Extension and gold mineralisation in the hanging walls of

active convergent continental shear zones. Journal of Structural Geology 64, 135–

148.

Vermeer, P.A., de Borst, R., 1984. Non-Associated Plasticity for Soils, Concrete and

Rock. Heron 29 (3).

Vos, I.M.A., Potma, W., Bierlein, F.P., Sheldon, H.A., 2007. Numerical modelling of the

western Hodgkinson Province, northeast Queensland: implications for gold

mineralisation. Australian Journal of Earth Sciences 54 (1), 27–47.

Wallis, R.H., Saracoglu, N., Brummer, J.J., Golightly, J.P., 1985. The geology of the

McClean uranium deposits, northern Saskatchewan. Special Volume - Canadian

Institute of Mining and Metallurgy 32, 101–131.

Wang, K., Chi, G., Bethune, K.M., Li, Z., Card, C.D., 2015. Characterization of

mineralizing fluids and fluid-structural relationships in the Phoenix U deposit, SE

Athabasca Basin, northern Saskatchewan: preliminary core and petrographic

studies. Geological Survey of Canada, Open File 7709, 1 sheet,

doi:10.4095/295853.

White, D.J., Hajnal, Z., Roberts, B., Gyorfi, I., Reilkoff, B., Bellefleur, G., Mueller, C.,

Woelz, S., Mwenifumbo, C.J., Takacs, E., Schmitt, D.R., Brisbin, D., Jefferson,

C.W., Koch, R., Powell, B., Annesley, I.R., 2007. Seismic methods for uranium

exploration; an overview of EXTECH IV seismic studies at the McArthur River

mining camp, Athabasca Basin, Saskatchewan, in: Jefferson, C.W., Delaney, G.

(Eds.), EXTECH IV: Geology and Uranium EXploration TECHnology of the

Proterozoic Athabasca Basin, Saskatchewan and Alberta: Geological Survey of

Page 198: Geometric and hydrodynamic modelling of fluid-structural

184

Canada, Bulletin 588, (also Geological Association of Canada, Mineral Deposits

Division, Special Publication 4; Saskatchewan Geological Society, Special

Publication 18), pp. 363–388.

Wilson, M., Kyser, T., 1987. Stable isotope geochemistry of alteration associated with

the Key Lake uranium deposit, Canada. Economic Geology 82 (6), 1540–1557.

Yang, J., 2006. Full 3-D numerical simulation of hydrothermal fluid flow in faulted

sedimentary basins: Example of the McArthur Basin, Northern Australia. Journal

of Geochemical Exploration 89 (1–3), 440–444.

Yang, J., Large, R., Bull, S., 2001. The importance of salinity in controlling ore-forming

fluid migration in SEDEX ore systems, Geological Society of America Abstracts

with Programs, pp. A-271.

Yang, J., Bull, S., Large, R., 2004a. Numerical investigation of salinity in controlling ore-

forming fluid transport in sedimentary basins: example of the HYC deposit,

Northern Australia. Mineralium Deposita 39 (5–6), 622–631.

Yang, J., Large, R.R., Bull, S.W., 2004b. Factors controlling free thermal convection in

faults in sedimentary basins: implications for the formation of zinc–lead mineral

deposits. Geofluids 4 (3), 237–247.

Yeo, G.M., Delaney, G., 2007. The Wollaston Supergroup, stratigraphy and metallogeny

of a Paleoproterozoic Wilson cycle in the Trans-Hudson Orogen, Saskatchewan,

in: Jefferson, C.W., Delaney, G. (Eds.), EXTECH IV: Geology and Uranium

EXploration TECHnology of the Proterozoic Athabasca Basin, Saskatchewan and

Alberta: Geological Survey of Canada, Bulletin 588, (also Geological Association

Page 199: Geometric and hydrodynamic modelling of fluid-structural

185

of Canada, Mineral Deposits Division, Special Publication 4; Saskatchewan

Geological Society, Special Publication 18), pp. 89–117.

Yeo, G.M., Percival, J.B., Jefferson, C.W., Ickert, R., Hunt, P., 2007. Environmental

significance of oncoids and crypto-microbial laminites from the Late

Paleoproterozoic Athabasca Group, Saskatchewan and Alberta, in: Jefferson,

C.W., Delaney, G. (Eds.), EXTECH IV: Geology and Uranium EXploration

TECHnology of the Proterozoic Athabasca Basin, Saskatchewan and Alberta:

Geological Survey of Canada, Bulletin 588, (also Geological Association of

Canada, Mineral Deposits Division, Special Publication 4; Saskatchewan

Geological Society, Special Publication 18), pp. 315–323.

Zanchi, A., Francesca, S., Stefano, Z., Simone, S., Graziano, G., 2009. 3D reconstruction

of complex geological bodies: Examples from the Alps. Computers &

Geosciences 35 (1), 49–69.

Zhang, Y., Ord, A., Roberts, P.A., Sorjonen-Ward, P., Lin, G., Wang, Y.J., 2005.

Numerical modelling of coupled deformation and fluid flow in mineralisation

processes, Mineral Deposit Research: Meeting the Global Challenge. Springer

Berlin Heidelberg, pp. 1509–1512.

Zhang, Y., Sorjonen-Ward, P., Ord, A., 2006. Modelling fluid transport associated with

mineralization and deformation in the Outokumpu Cu–Zn–Co deposit, Finland.

Journal of Geochemical Exploration 89 (1–3), 465–469.

Zhang, Y., Lin, G., Roberts, P., Ord, A., 2007. Numerical modelling of deformation and

fluid flow in the Shuikoushan district, Hunan Province, South China. Ore

Geology Reviews 31 (1–4), 261–278.

Page 200: Geometric and hydrodynamic modelling of fluid-structural

186

Zhang, Y., Schaubs, P.M., Zhao, C., Ord, A., Hobbs, B.E., Barnicoat, A.C., 2008. Fault-

related dilation, permeability enhancement, fluid flow and mineral precipitation

patterns: numerical models. Geological Society, London, Special Publications 299

(1), 239–255.

Zhang, Y., Gartrell, A., Underschultz, J.R., Dewhurst, D.N., 2009. Numerical modelling

of strain localisation and fluid flow during extensional fault reactivation:

Implications for hydrocarbon preservation. Journal of Structural Geology 31 (3),

315–327.

Zhang, Y., Robinson, J., Schaubs, P.M., 2011. Numerical modelling of structural controls

on fluid flow and mineralization. Geoscience Frontiers 2 (3), 449–461.

Zhang, Y., Schaubs, P.M., Sheldon, H.A., Poulet, T., Karrech, A., 2013. Modelling fault

reactivation and fluid flow around a fault restraining step-over structure in the

Laverton gold region, Yilgarn Craton, Western Australia. Geofluids 13 (2), 127–

139.

Zhao, C., Hobbs, B.E., Ord, A., 2009. Fundamentals of computational geoscience:

numerical methods and algorithms. Springer Science & Business Media.

Zhao, C., Reid, L.B., Regenauer-Lieb, K., 2012. Some fundamental issues in

computational hydrodynamics of mineralization: A review. Journal of

Geochemical Exploration 112, 21–34.

Zhu, J., Li, Z., Lin, G., Zeng, Q., Zhou, Y., Yi, J., Gong, G., Chen, G., 2014. Numerical

simulation of mylonitization and structural controls on fluid flow and

mineralization of the Hetai gold deposit, west Guangdong, China. Geofluids 14

(2), 221–233.

Page 201: Geometric and hydrodynamic modelling of fluid-structural

187

Page 202: Geometric and hydrodynamic modelling of fluid-structural

188

Appendix

Page 203: Geometric and hydrodynamic modelling of fluid-structural

189