Accepted Manuscript
Combined inter- and intra-fracture free convection in fracture networks embedded in a low-permeability matrix Katharina Vujevi´c, Thomas Graf PII: DOI: Reference:
S0309-1708(15)00161-X 10.1016/j.advwatres.2015.07.014 ADWR 2426
To appear in:
Advances in Water Resources
Received date: Revised date: Accepted date:
28 October 2014 22 June 2015 23 July 2015
Please cite this article as: Katharina Vujevi´c, Thomas Graf, Combined inter- and intra-fracture free convection in fracture networks embedded in a low-permeability matrix, Advances in Water Resources (2015), doi: 10.1016/j.advwatres.2015.07.014
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Highlights • Inter- and intrafracture convection are simulated separately and in com-
CR IP T
bination • A critical Rayleigh number for interfracture convection is proposed • Interfracture convection dominates in continuous fracture circuits
• Combined inter- and intrafracture convection develops in complex frac-
AN US
ture networks
• Complex fracture networks require 3D simulation to capture free con-
AC
CE
PT
ED
M
vection
1
ACCEPTED MANUSCRIPT
AC
CE
PT
ED
M
AN US
CR IP T
Graphical Abstract
2
ACCEPTED MANUSCRIPT
CR IP T
Combined inter- and intra-fracture free convection in fracture networks embedded in a low-permeability matrix Katharina Vujevi´ca,∗, Thomas Grafa
Institute of Fluid Mechanics and Environmental Physics in Civil Engineering, Leibniz Universit¨ at Hannover. Appelstrasse 9A, 30167 Hannover, Germany
AN US
a
Abstract
M
Density-driven, haline free convection in fractured porous rock is studied in a 3-dimensional numerical model to evaluate the onset and interdependence of
ED
intra- and interfracture convection. When the rock matrix allows for solute diffusion, the most likely mode of convection in a regular fracture circuit is interfracture convection along the circuit such that regular fracture circuits can
PT
be modeled in a reduced dimensionality (2D) model. The critical Rayleigh number for the onset of interfracture convection in a regular fracture circuit
CE
is determined to be half of the critical Rayleigh number for convection in a single vertical fracture. In more complex fracture networks, vertical “dead-
AC
end fractures” and additional fracture intersections complicate convective patterns. Combined inter- and intrafracture convection is possible, such that ∗
Corresponding author Email addresses:
[email protected] (Katharina Vujevi´c),
[email protected] (Thomas Graf)
Preprint submitted to Advances in Water Resources
July 28, 2015
ACCEPTED MANUSCRIPT
convection is hard to predict.
1
CR IP T
Keywords: Groundwater; fractures; Rayleigh number; density-driven flow; 3-dimensional, HydroGeoSphere; numerical modelling 1. Introduction
In many deep groundwater aquifers, density-driven free convection is an
3
important process for solute transport [11, 36]. Free convection has to be con-
4
sidered when solute or heat influence groundwater density [31, 54]. When
5
dense brine is overlying less dense fresh water, this potentially unstable fluid
6
stratification may be dissipated by diffusion, or may lead to free convection in
7
case of dominating buoyancy forces [11, 24, 31, 54]. A number of authors em-
8
phasize the widespread importance of density-driven flow, as density-driven
9
flow transports solutes faster and farther into the aquifer then diffusion alone
10
[33, 37, 47]. Therefore, neglecting free convection will lead to a considerable
11
underestimation of solute and/or heat transport.
M
AN US
2
In low-permeability geological formations, such as shales or granites, frac-
13
tures may enable free convective flow even if the matrix is virtually imperme-
14
able to flow. Therefore, fracture networks are conjointly considered to play
15
an important role in solute and heat transport in low-permeability strata
16
and in determining the onset and patterns of free convection in such sys-
17
tems [19, 20, 26, 33, 34, 37, 43, 44]. Understanding the onset conditions of
18
free convection in fractured rock is increasingly important because fractured
AC
CE
PT
ED
12
19
rocks are widespread and gain importance as potential hosts for hazardous
20
waste disposal, heat and CO2 storage, natural gas extraction by fracking,
21
and are used for thermal energy exploitation. Furthermore, density-driven
22
flow in fractures and fault systems is known to play an important role in the 4
ACCEPTED MANUSCRIPT
23
formation of ore deposits [1, 8, 16, 49, 53, 55]. Convection in fractured low-permeability strata is possible along a circuit
25
of connected fractures (interfracture convection, “mode 1” as shown in Fig.
26
1a) or within a single fracture (intrafracture convection, “mode 2” as shown
27
in Fig. 1b) [37, 43]. Both types of convection have been examined separately
28
in a couple of theoretical studies with different considerations of fracture-
29
matrix interactions as discussed below.
b) isolated vertical fractures Intrafracture convection (mode 2B) (mode 2A)
AN US
a) 2-dimensional regular fracture circuit Interfracture convection (mode 1)
L =6 m
m
H=10 m
h=6 m
H=10 m
2B=6 m
h=6 m
L=0.2
CR IP T
24
z
y
M
x
W=10 m
W=10 m
ED
Figure 1: Conceptual models to study (a) interfracture convection and (b) intrafracture convection. Mode 2A was not considered in this study and is just shown for completeness.
Early studies on (thermal) convection within a single fracture approxi-
31
mated the fracture by a 3-dimensional fluid-filled vertical slab or by a sat-
32
urated porous box heated from below. Fracture walls were assumed to be
33
impermeable to flow and non-conducting, such that any fracture-matrix in-
34
teraction was neglected [7, 56]. In these cases, the onset of convection can be
35
predicted using the classical Rayleigh number Ra which compares buoyancy
AC
CE
PT
30
36
to diffusivity. If the critical Rayleigh number of Rac = 4π 2 is exceeded, con-
37
vection with streamlines parallel to the fracture plane (mode 2B) is expected
38
to establish [3, 9, 15, 30, 56].
39
Some studies accounted for heat-conducting fracture walls and applied a 5
ACCEPTED MANUSCRIPT
constant vertical temperature gradient to the lateral fracture walls [23, 25, 50,
41
52]. This boundary condition corresponds to assuming perfectly conducting
42
fracture wall and has a stabilizing effect on thermal convection [27]. Thus,
43
higher vertical temperature differences are required to initiate free convection
44
and the critical Rayleigh number greatly exceeds 4π 2 [8, 23]. However, heat
45
transport within the matrix was not considered [23, 25, 50, 52]. In reality,
46
this would correspond to a case where the fracture gains or looses heat along
47
its vertical walls, but the adjacent matrix does not change temperature.
CR IP T
40
More advanced models for convection within a fracture embedded in a
49
conducting matrix were not restricted to simulating processes within the
50
fracture, but also accounted for heat transport in the adjacent matrix [1, 21,
51
26, 27, 28, 44, 53]. The stabilizing effect of the matrix is less distinct in
52
this case, such that resulting critical Rayleigh numbers exceed 4π 2 , but are
53
below those determined with vertical fracture walls at constant temperature
54
gradient [27].
M
AN US
48
Interfracture convection within a fracture network was usually studied in
56
2-dimensional domains with fractures represented as 1-dimensional line ele-
57
ments such that convection within a fracture was neglected. Even without
58
considering intrafracture convection (mode 2), it was shown that fracture net-
59
works considerably affect the onset, strength and patterns of free convection
60
in fractured porous media [20, 34, 38, 43, 48]. Depending on the geometry
61
of individual fracture networks, convection was enhanced or inhibited, such
AC
CE
PT
ED
55
62
that making predictions is a complex task.
63
Very few studies link inter- and intrafracture convection. A theoretical
64
comparison of the likelihood of each mode of free (haline) convection is pro-
65
vide by Simmons et al. [37], who considered a fracture circuit embedded 6
ACCEPTED MANUSCRIPT
in an impermeable rock matrix. A comparison of critical concentration dif-
67
ferences (derived from critical Rayleigh numbers) led to the conclusion that
68
intrafracture convection parallel to the fracture plane (mode 2B) is the most
69
likely mode of convection and occurs for even very small density differences.
70
According to Simmons et al. [37], interfracture convection around an im-
71
permeable rock matrix block (mode 1) is likely as well, while intrafracture
72
convection normal to the fracture plane (mode 2A) requires large density dif-
73
ferences in combination with large fracture apertures and is hence unlikely.
CR IP T
66
Ramazanov [32] analytically determined the onset of interfracture ther-
75
mal convection (called transverse convection) along a fracture circuit and
76
intrafracture convection (called longitudinal convection) within the vertical
77
fractures of a fracture circuit embedded in an impermeable, conductive rock
78
matrix. Ramazanov’s [32] results suggest that the onset of convection in a
79
fracture circuit depends on fracture aperture and aspect ratio of the fracture
80
circuit. Interfracture convection sets in at lower Rayleigh numbers than in-
81
trafracture convection for small fracture apertures and fracture circuits where
82
the aspect ratio (height to width) is between 0.25 and 8.67.
83
Contribution of this study
PT
ED
M
AN US
74
The onset of possible free convection in fractures has been theoretically
85
studied and numerically simulated separately, either within a single fracture
86
(e.g. [26, 44]) or along a fracture network (e.g. [43, 48]). One of the reasons
87
for the lack of studies on combined intra- and interfracture convection is that
88
a combined consideration requires a fully 3-dimensional approach, which is
89
demanding especially for problems with complex geometries. Studies that
90
theoretically considered both inter- and intrafracture convection simplified
91
the fracture-matrix interaction by either averaging over fracture and
AC
CE
84
7
ACCEPTED MANUSCRIPT
matrix properties (Simmons et al. [37]) or assuming diffusion in the
93
fracture and the matrix to be identical (Ramazanov [32]). Thus, the
94
resulting probabilities for each mode of convection may not be applicable to a
95
discrete fracture circuit embedded in a porous matrix that allows for diffusive
96
transport and that has different diffusivity parameters than the fracture.
CR IP T
92
The purpose of this study is to examine intra- and interfracture haline
98
convection separately (Fig. 1) and in combination for simple fracture network
99
geometries shown in Fig. 2 and to determine under which conditions which
100
mode dominates. The effect of the adjacent matrix will be accounted for by
101
allowing for diffusion within the rock matrix, within the fractures, and across
102
the fracture-matrix interface. Thus, we want to contribute to a profound un-
103
derstanding of the interdependence of intra- and interfracture convection,
104
the onset of combined convection as well as the convective patterns which
105
can be expected in fractured low-permeability media. We further aim at
106
encompassing in which type of fracture networks a fully 3-dimensional repre-
107
sentation is required or when a reduced dimensionality representation for the
108
simulation of free convection yields reasonable results. We focus our study
109
on haline free convection on the basis of findings from Graf and Therrien
110
[21], Sharp et al. [33] and Alt-Epping and Zhao [1], who stated
111
that in low-permeability shale layers and geological fault zones,
112
salinity gradients are the dominating driving force for free convec-
113
tion (in contrast to temperature gradients), because geothermal
AC
CE
PT
ED
M
AN US
97
114
convection requires much higher fracture and matrix permeabili-
115
ties to establish than haline convection due to the large stabilizing
116
thermal diffusivity. This is the first numerical study of free convection in
117
fractured-porous rock that considers a combination of intra- and interfracture 8
ACCEPTED MANUSCRIPT
convection and allows for matrix diffusion. a) 3-dimensional regular fracture circuit Intra-, interfracture or combined convection (mode 1 or mode 2B or mode 12B)
H=10 m
h=6 m
H=10 m
2B=6 m
h=4 m
L=4 m
L=6 m
CR IP T
b) 3-dimensional complex fracture circuit Intra-, interfracture or combined convection (mode 1 or mode 2B or mode 12B)
circuit height
118
W=10 m
AN US
W=10 m
AC
CE
PT
ED
M
Figure 2: Conceptual models to study combined inter- and intrafracture convection. a) regular fracture circuit consisting of 4 fractures. Arrows show convection patterns of combined inter- and intrafracture convection (mode 12B) b) more complex fracture circuit consisting of 6 fractures.
9
ACCEPTED MANUSCRIPT
119
2. Methodology
120
2.1. Mathematical model In density-driven flow and transport problems, flow and trans-
122
port are coupled by fluid density which varies with concentration
123
and affects flow by adding a buoyancy component to the flow equa-
124
tion. In fully saturated porous media, the coupled governing equa-
125
tions are the Darcy equation and mass conservation equations for
126
fluid and solute mass. The non-linear system of equations is com-
127
pleted by the constitutive equations that define porous medium
128
and fracture freshwater hydraulic conductivity and relationships
129
between fluid density and salinity. The governing and constitutive
130
equations can be found in the Appendix.
AN US
CR IP T
121
Free convection in fractured porous rock is investigated numerically using
132
the variable-density groundwater flow and transport model HydroGeoSphere
133
[39] which applies the control volume finite element method to the flow equa-
134
tion and the Galerkin finite element method to the transport equation. The
135
primary unknown for flow is equivalent fresh water head h0 [L]
136
defined as [14]:
ED
PT
CE
h0 =
p +z ρ0 g
(1)
where p [ML−1 T−2 ] is fluid pressure, ρ0 [ML−3 ] is reference fluid density,
AC
137
M
131
138
g [LT−2 ] is gravitational acceleration and z [L] is elevation. The primary
139
unknown for transport is relative concentration c [-].
140
Fractures are represented as discrete 2-dimensional planes that
141
share common nodes with 3-dimensional elements of the rock ma10
ACCEPTED MANUSCRIPT
trix. The common node approach ensures continuity of the un-
143
knowns at the fracture-matrix interface. By superimposing the
144
contributions from fracture and matrix at common nodes, fluid and
145
solute mass exchange do not have to be explicitly calculated in the
146
locally mass conservative discretized flow and transport equations.
147
A linear relationship between fluid density and relative concentra-
148
tion is assumed as described in the Appendix. Further details on
149
the HydroGeoSphere model can be found in the Appendix, in Graf
150
and Therrien [20], and in Therrien et al. [39].
151
2.2. Dimensionless numbers
AN US
CR IP T
142
The Rayleigh number is used to predict the onset of convection [35, 42].
153
The Sherwood number is used to characterize the strength of free convection
154
[4, 43].
155
Rayleigh Number
ED
M
152
The Rayleigh number Ra [-] relates buoyancy that promotes free con-
157
vection to diffusion that dissipates free convection [11, 35, 42, 43, 45]. The
158
classical Rayleigh number Ra for salinity-driven free convection in homoge-
159
neous porous media reads:
CE
H 2 kgρ0 β ∆c H Ra = Dm µ
(2)
where H [L] is the height of the domain, k [L2 ] is the permeability of
AC
160
PT
156
161
the porous medium, g [L T−2 ] is gravitational acceleration, µ [M L−1 T−1 ] is
162
fluid dynamic viscosity, and Dm [L2 T−1 ] is the effective diffusion coefficient
163
in the porous matrix. If Ra is smaller than the critical Rayleigh number
164
Rac , any small initial perturbation (trigger) decays with time, such that the 11
ACCEPTED MANUSCRIPT
system becomes purely diffusive and stable [15, 31, 54]. If Ra is larger than
166
Rac , the system is dominated by buoyancy-induced free convection, and it is
167
called unstable [15, 31, 54]. The value of the critical Rayleigh number Rac
168
depends on the applied flow and transport boundary conditions [30]. The
169
critical Rayleigh number is Rac = 4π 2 for a closed box filled with saturated
170
homogeneous porous media, with fixed concentrations cmax , cmin at the top
171
and bottom of the domain, respectively.
CR IP T
165
When a single 2-dimensional open vertical fracture is considered instead
173
of a box of porous medium, the permeability k has to be replaced by kfr , the
174
diffusion coefficient in the fracture corresponds to D0 in Eq. (22), and the
175
height of the fracture h [L] is introduced, such that Rafr reads [44]:
AN US
172
176
(3)
M
h2 kfr gρ0 β ∆c H Rafr = D0 µ
Using linear stability analysis, Malkovsky and Pek [26] and Wang et al.
[44] determined a critical Rayleigh number Rac-fr
178
for convection in a single vertical fracture embedded in a fully
179
3-dimensional fracture-matrix system. The matrix was assumed to
180
be impermeable, but heat-conducting, and upper and lower bound-
181
aries of the setup were impermeable and of constant temperature.
182
Malkovsky and Pek [26] and Wang et al. [44] pointed out that
183
a conducting matrix has a stabilizing effect, due to heat transfer
184
along the fracture-matrix interface. The magnitude of heat ex-
185
change depends on the ratio of fracture aperture (2b) to fracture
186
height h, and on the thermal conductivity ratio between matrix
187
and fracture [44]. Consequentially, the critical Rayleigh number
188
Rac-fr can be formulate as a function of those two ratios, where
AC
CE
PT
ED
177
12
ACCEPTED MANUSCRIPT
Malkovsky and Pek [26] determined Rac-fr = 16.38·h/(2b) and Wang
190
et al. [44] derived Rac-fr = 16.32 · h/(2b). The formulation of the critical
191
Rayleigh number determined by Wang et al. [44] is adapted here for haline
192
convection by replacing the thermal conductivity ratio between matrix and
193
fracture by the solutal diffusivity ratio rD = Dm /D0 : Rac-fr = 16.32rD
h (2b)
CR IP T
189
(4)
We combine Eq. (3) with Eq. (4) and replace kfr by (2b)2 /12 [5, 20] to
195
obtain a critical fracture aperture (2b)c in Eq. (5), which is the parameter
196
that is usually unknown. With parameters given in Table 1 that are typical
197
for limestone, claystone or shale [6, 29], (2b)c becomes:
3
rD µ ∆c β H hρg
· 12 · 16.32 = 1.74 · 10−5 m
M
(2b)c =
r
AN US
194
(5)
A critical Rayleigh number Rac-circ for convection along a fracture circuit
199
has been formulated by Ramazanov [32]. As that formulation is quite com-
200
plex and does not account for diffusivity differences between fracture and
201
matrix, it is not shown here. A new formulation is proposed in this paper in
202
section 3.1.
203
Sherwood Number
PT
The dimensionless Sherwood number Sh is used as an indicator for the
strength of convection [4, 43]:
AC
205
CE
204
ED
198
206
Sh =
j Dm ρ0 ∆c H
(6)
where j [M T−1 L−2 ] is the total mass flux through the top layer of the 13
ACCEPTED MANUSCRIPT
Table 1: Simulation parameters that are typical for a claystone or shale
9.81 m s−2 0.1 m−1 0.07 1000 kg m−3 1.1×10−3 kg m−1 s−1 1.0×10−18 m2 0.1 0.1 10−9 m2 s−1 10−11 m2 s−1 0.01 0.1 m 0.005 m
CR IP T
g ∆c/H β ρ0 µ km θ τ D0 Dm rD αL αT
AN US
Gravitational acceleration Concentration gradient Expansion coefficient Reference fluid density (fresh water) Fluid dynamic viscosity Rock matrix permeability Rock matrix porosity Rock matrix tortuosity Free solution diffusion coefficient Diffusion coefficient in the matrix (D0 θτ ) Diffusivity ratio matrix to fracture Longitudinal dispersivity Transverse dispersivity
simulation domain, and the denominator describes the expected mass flux
208
through this layer for a purely diffusive regime. Therefore, Sh is 1 in the
209
steady state of stable and purely diffusive scenarios, and Sh is greater than
210
1 once free convection begins [17].
211
2.3. Conceptual models of simulation scenarios
ED
M
207
Figs. 1 and 2 depict the basic conceptual models we examined, where Fig.
213
2 shows the scenarios that are novel. We assume a 10 m×6 m×10 m (width×
214
length× height) matrix block, in which fractures of height h = 6 m (vertical
215
fractures) and width w = 6 m (horizontal fractures) are embedded. Thus,
216
the fractures do not traverse the entire height nor width of the block, but
217
they traverse the entire length L of the block in y-direction. The matrix has
AC
CE
PT
212
218
a very low hydraulic permeability (10−18 m2 ), but allows for diffusive mass
219
transport. As a reference, we separately consider each mode of convection:
220
interfracture convection mode 1 (Fig. 1a) and intrafracture convection mode
221
2B (Fig. 1b) as described below. Fig. 1a represents a 2-dimensional cross
14
ACCEPTED MANUSCRIPT
section of the fracture network, which is a common approach used to simulate
223
free convection in fractured rock: The continuous fracture circuit represents
224
the most fundamental building block (as of [37]) of an orthogonal fracture
225
network. Fig. 1b refers to free convection in an isolated vertical fracture,
226
e.g. in fractured rock where almost uniform fracture orientations avoid
227
inter-connection of fractures. As a novelty, we then consider a 3-dimensional
228
fracture circuit (Fig. 2a), where both modes of convection are theoretically
229
possible. Finally, we show how our results can be applied to other fracture
230
network geometries and discuss the predictability of convection in a more
231
complex 3-dimensional fracture network (Fig. 2b). Thus, we approach the
232
reality of fractured rocks with a variety of geometries of connected fracture
233
clusters.
AN US
CR IP T
222
For each geometry (Fig. 1a, Fig. 1b, Fig. 2a, Fig. 2b), we modify
235
the fracture aperture (2b) to numerically determine the critical (2b) for the
236
onset of convection. The concentration gradient along the vertical axis and
237
all other parameters will be kept constant as listed in Table 1 unless other-
238
wise stated. The onset of convection is determined calculating Sh for each
239
scenario, were Sh = 1 indicates the absence of convection, and Sh > 1 indi-
240
cates the existence of free convection. We determine the convective patterns
241
visually by evaluating the shape of the streamlines and isochlors.
ED
PT
It should be made clear that we distinguish between the geometry of a
fracture network as shown in Figs. 1 and 2 and the modes of convection
AC
243
CE
242
M
234
244
defined by the occurring flow patterns. The 2-dimensional fracture circuit
245
(Fig. 1a) only allows for interfracture convection (named mode 1). Isolated
246
vertical fractures (Fig. 1b) only allow for intrafracture convection (named
247
mode 2B). Only the 3-dimensional fracture circuits (Figs. 2a, b) allow for 15
ACCEPTED MANUSCRIPT
248
either interfracture convection (mode 1) or intrafracture convection (mode
249
2B) or for combined convection (named mode 12B). It should be acknowledged that we do not study intrafracture convection
251
perpendicular to the fracture plane (mode 2A as of Simmons et al. [37] or
252
2-D slender circle flow as of Zhao et al. [50]). According to Simmons
253
et al. [37] and several earlier studies [3, 9, 15, 56], convection parallel to the
254
fracture plane is the preferred mode within a finite vertical fracture, such
255
that neglecting intrafracture convection perpendicular to the fracture plane
256
is acceptable without loss of generality of this study. In some studies where
257
conduction/diffusion across the fracture walls was allowed [1, 26, 27, 44, 50],
258
intrafracture convection cells parallel to the fracture plane are described as
259
(weakly) 3-dimensional: Assuming a linear vertical temperature dis-
260
tribution along the impermeable vertical fracture walls, Kassoy and
261
Cotte [23] and Zhao et al. [50, 54] theoretically showed that con-
262
vection cells with rotation axes parallel to the fault width (fracture
263
aperture) dominate thermal convective patterns. The convective
264
pattern was described as 3-dimensional (called “standard convec-
265
tive mode” or “finger-like convective mode”) because the fracture
266
walls have a dampening effect on the flow field, such that flow ve-
267
locities are highest in the central vertical plane of the fault. In
268
contrast to the studies of Kassoy and Cotte [23] and Zhao et al.
269
[50, 54], concentration at vertical fracture walls in our study is not
AC
CE
PT
ED
M
AN US
CR IP T
250
270
fixed, but transport in the rock matrix surrounding the vertical
271
fracture is fully considered, similar to the theoretical analyses by
272
Wang et al. [44] and Malkovsky and Pek [26, 27]. Wang et al.
273
[44] and Malkovsky and Pek [26, 27] showed that convection ap16
ACCEPTED MANUSCRIPT
proaches a 2-dimensional pattern for low ratios of matrix diffusiv-
275
ity to fracture diffusivity, and for low ratios of fracture aperture to
276
fracture height, which are both given in our study. Thus, averaging
277
over the fracture width seems a valid approach here, even if that
278
2-dimensional approach does not resolve the flow field within the
279
fracture in the third dimension. The main advantage of the present
280
2-dimensional fracture approximation is that computational efforts
281
are significantly reduced. As the first step of dealing with compli-
282
cated geoscience problems, such an approximation is commonly
283
acceptable for numerical models [55]. Thus, in the present study,
284
the term “intrafracture convection” corresponds to convective flow
285
with streamlines parallel to the fracture plane (mode 2B).
286
Boundary and initial conditions
M
AN US
CR IP T
274
All domain boundaries are impermeable to flow. The initial state is in hy-
288
drostatic equilibrium. We assign constant concentration boundary conditions
289
to the bottom and top of the domain, while lateral boundaries are assigned
290
a zero-mass flux boundary condition. Concentration at the top is cmax = 1,
291
and bottom concentration is cmin = 0. Initially, solute is distributed linearly
292
between upper and lower boundary. A physical trigger was applied, where
293
the central node of the vertical fracture was assigned a 10% higher concentra-
294
tion relative to the ambient fluid. In linear stability analyses, a system
295
is defined as stable, when a small initial perturbation (trigger) de-
296
cays with time [15, 31]. Using numerical simulations to determine
297
the onset and mode of convection limits the number of possible
298
initial perturbations that may be tested. We varied the magnitude
299
of the applied trigger as well as the location of the trigger with-
AC
CE
PT
ED
287
17
ACCEPTED MANUSCRIPT
out observing any change in the critical Rayleigh number and the
301
patterns of free convection. The fact that results for the onset of
302
convection in a single vertical fracture are in good agreement with
303
the results of linear stability analyses obtained by Wang et al. [44]
304
and Malkovsky and Pek [26] further substantiates the reliability of
305
our numerical simulations. The advantage of using numerical sim-
306
ulations to determine onset and patterns of free convection is that
307
this approach allows to test more complex scenarios than linear
308
stability analysis.
309
Spatial and temporal discretization
AN US
CR IP T
300
The domain is discretized using regular blocks with an edge length of 0.2
311
m. Finer uniform grids (minimum edge length 0.05 m) as well as refined grids
312
close to fractures as proposed by Weatherill [46] were tested, too (results not
313
shown). The resulting convective pattern using all 3 different grids is visually
314
identical and the steady-state Sh deviate by less than 0.1%. Thus, the grid
315
consisting of 0.2 m blocks is sufficient.
ED
M
310
An adaptive time stepping scheme is used, where the time step size is
317
calculated based on the maximum concentration change during a time step
318
[12, 13, 39]. A maximum concentration change of 0.1 at any node is allowed.
319
Reducing the tolerance criteria to a maximum concentration change of 0.01
320
did not affect the results. Simulations were run until a steady state was
CE
PT
316
reached.
AC 321
18
ACCEPTED MANUSCRIPT
322
3. Results and discussion
323
3.1. Interfracture convection (mode 1) in a 2-dimensional regular fracture circuit
CR IP T
324
Interfracture convection is examined in a regular fracture circuit as shown
326
in Fig. 1a. We define a “regular fracture circuit” to consist of 4 fractures
327
that form an isolated continuous rectangle without any dead-end fractures.
328
A 2-dimensional representation is sufficient here, and the domain thickness
329
in y-direction can be reduced to one element (0.2 m). Thereby, intrafracture
330
convection (mode 2B) is deliberately precluded. Critical fracture aperture
331
for the onset of interfracture free convection (mode 1) in a 2-dimensional,
332
square regular fracture circuit as well as the strength of convection (Sh)
333
for various fracture apertures are evaluated numerically. The results are
334
later compared to those obtained from a 3-dimensional representation in
335
section 3.3. Table 2 shows the required fracture aperture (2b) for the onset
336
of interfracture convection (mode 1). Convection was detected at the critical
337
fracture aperture (2b)c = 1.4 · 10−5 m.
AC
CE
PT
ED
M
AN US
325
19
ACCEPTED MANUSCRIPT
Table 2: Onset and patterns of inter- and intrafracture convection
3d fracture circuit (Fig. 2a) −
CR IP T
3d isolated fracture (Fig. 1b) − − − − − (1) (2) (2) (2) (4)
AN US
(2b) 2d fracture circuit [10−5 m] (Fig. 1a) 1.0...1.3 − 1.4 1.5 1.6 1.7 1.8 1.9 2.0 2.5 3.0...5.0
ED
M
- : no convection : interfracture convection (mode 1) along the fracture circuit as shown in e.g. Figs. 3 and 6, (1): intrafracture convection mode 2B and number of convection cells per vertical fracture in brackets, as e.g. Fig. 5.
No convection occurs for (2b) < 1.4 · 10−5 m (Fig. 3a). For (2b) ≥
338
1.4 · 10−5 m however, circular flow along the fracture circuit occurs, shown
340
in Figs. 3b and 3c. With increasing fracture aperture, fracture velocity and
341
strength of convection increase. Therefore, solute is transported farther along
342
the fracture circuit, thereby increasing the concentration difference between
343
fracture and matrix at a transient stage of the simulation (transient results
AC
CE
PT
339
344
not shown). Thus, diffusion along the fracture-matrix interface increases. As
345
a result, the steady-state concentration along the fracture circuit is almost
346
uniform for high fracture apertures of (2b) = 5.0 · 10−5 m as shown in Fig.
347
3c. 20
ACCEPTED MANUSCRIPT
b)
c)
CR IP T
a)
AN US
Figure 3: Interfracture convection (mode 1): Solute distribution in the matrix is indicated by transparent colors (red high concentration, blue zero concentration) and grey isochlors. a) - no convection for (2b) < 1.4 · 10−5 m b) Clockwise interfracture convection for (2b) = 1.4 · 10−5 m c) Clockwise interfracture convection for (2b) = 5.0 · 10−5 m 348
Critical Rayleigh number for convection in a regular fracture circuit (mode
349
1)
The critical Rayleigh number for the onset of interfracture convection
351
(mode 1) in a 2-dimensional fracture circuit of variable size and aspect ra-
352
tio was determined numerically. A variety of regular fracture circuits with
353
variable circuit height h (distance between top and bottom fracture of the
354
circuit) and circuit width w (lateral distance between vertical fractures of
355
the circuit) were tested such that circuit area (h × w) and circuit aspect ratio
356
(w/h) varied. Fig. 4 shows the strength of convection (Sh) depending on the
357
number Rafr × (2b)/(hDr), where Rafr is calculated with Eq. (3) and where
PT
ED
M
350
fracture aperture is varied from (2b) = 5 · 10−6 m to 6 · 10−5 m. Varying
359
the fracture aperture of a 2-dimensional regular fracture circuit, all scenarios
360
below the critical number of 8.16 are stable, while convection was found in
361
all scenarios with a critical number larger than 8.16.
AC
CE
358
21
1.3
1.05
1.25
1.04
1.2 1.15 1.1 1.05
1.03 1.02 1.01 1
1 10
20
30
40
50
Rafr x (2b)/(hθτ)
0.99
60
5
CR IP T
Sherwood number
Sherwood number
ACCEPTED MANUSCRIPT
8.16
10
Rafr x (2b)/(hθτ)
15
362
AN US
Figure 4: Onset of interfracture convection mode 1: Strength of convection depending on Rayleigh number constant (Rafr × (2b)/(hrD )). Data points show circuits of different aspect ratio (width/height) and area (width×height). Black points represent circuits with large aspect ratio, white points represent circuits with small aspect ratio. Area is indicated by datapoint size.
Interestingly, the numerically determined critical number Rafr ×(2b)/(hDr) for the onset of interfracture convection (mode 1) is equal to half of the critical
364
number of 16.32 for the onset of intrafracture convection (mode 2B) as of Eq.
365
(4) determined by Wang et al. [44]. The critical Rayleigh number Rac-circ for
366
the onset of interfracture free convection (mode 1) in a 2-dimensional regular
367
fracture circuit can therefore be formulated as:
PT
ED
M
363
370
h (2b)
(7)
The corresponding formulation of the critical fracture aperture found here
is therefore:
AC
369
CE
368
Rac-circ = 8.16rD
(2b)c-circ =
s 3
rD D0 µ · 12 · 8.16 = 1.38 · 10−5 m β ∆c hρg H
(8)
That number is in excellent agreement with the numerically determined 22
ACCEPTED MANUSCRIPT
value of 1.4 · 10−5 m given in Tab. 2. Eqs. (7) and (8) are valid for a
372
regular fracture circuit embedded in a low permeable matrix, which inhibits
373
fluid flow but allows for diffusive solute transport, where upper and lower
374
boundaries of the matrix block are of fixed concentration.
CR IP T
371
Compared to previous studies which examined the onset of con-
376
vection in fracture circuits, our results are similar to those obtained
377
by Ramazanov [32] who gave a ratio between the critical Rayleigh
378
numbers for intra- and interfracture convection (mode 1) in a
379
square fracture circuit of approximately 2. While Ramazanov [32]
380
assumed that the thermal conductivity of fracture-filling material
381
and surrounding rock matrix are of the same order of magnitude,
382
his formulation of the critical Rayleigh number does not include
383
the thermal equivalent of the diffusivity ratio rD . For the special
384
case of a square fracture circuit with thin fractures, Ramazanov’s
385
[32] formulation can be approximated by Rac-circ = 8h/(2b).
M
Simmons et al.
[37] applied the classical formulation of the
ED
386
AN US
375
Rayleigh number for convection in a homogeneous porous medium
388
using the average permeability of the fractured medium and com-
389
pared it to the critical Rayleigh number of 4π 2 to determine the
390
onset of interfracture convection (mode 1).
391
that the Rayleigh number formulation corresponds to Eq. (3), the
392
critical Rayleigh number proposed by Simmons et al. [37] would
Reformulated such
AC
CE
PT
387
393
be almost 5 times higher than what was determined here, and
394
would thus underestimate the occurrence of interfracture convec-
395
tion. However, using the average permeability is appropriate to
396
determine the onset of free convection under the assumption of 23
ACCEPTED MANUSCRIPT
large fracture density and uniformly distributed fractures. For less
398
dense fracture networks, it was numerically shown by Vujevi´ c et
399
al. [43] and Yang et al. [48] that convection may occur in a frac-
400
tured medium containing interconnected fractures such as contin-
401
uous fracture circuits even if the Rayleigh number (Ra) averaged
402
over the entire medium is below the critical value [43, 48].
403
3.2. Intrafracture convection (mode 2B) in isolated vertical fractures
CR IP T
397
Intrafracture convection (mode 2B) is studied in two isolated parallel
405
vertical fractures embedded in a 3-dimensional matrix (Fig. 1b). The vertical
406
fractures are 6 m high, 6 m long in the y-direction and 6 m apart from each
407
other such that they are hydraulically disconnected by the low permeable
408
matrix. Fracture aperture is identical in both fractures and was varied in
409
order to determine the critical fracture aperture.
M
AN US
404
Onset of convection was detected at (2b) = 1.8·10−5 m (Table 2), which is
411
in very good agreement with the critical fracture aperture of (2b) = 1.74·10−5
412
m predicted by Eq. (5) based on the results by Wang et al. [44]. The
413
critical Rayleigh number Rac-fr as of Eq. (4) and the corresponding critical
414
fracture aperture for intrafracture convection (mode 2B) in Eq. (5), were
415
originally formulated for impermeable upper and lower fracture boundaries
416
of constant temperature (equivalent to constant concentration for haline con-
417
vection). This is to say, the vertical fracture is in contact with both, upper
418
and lower boundary. A new result obtained here is that Eqs. (4) and (5)
419
are also applicable if the upper and lower end of the fracture is separated
420
from the constant concentration boundary condition by an impermeable rock
421
matrix block that allows for solute diffusion. As a further result, it can be
422
expected that Eqs. (4) and (5) are also applicable to thermal convection
AC
CE
PT
ED
410
24
ACCEPTED MANUSCRIPT
423
in vertical fractures separated from the boundary conditions, because the
424
equations were originally formulated for a heat convection problem. The critical fracture aperture determined here for a single verti-
426
cal fracture can be compared to results obtained theoretically with
427
different considerations of the fracture-matrix solute exchange us-
428
ing (critical) Rayleigh numbers determined in previous studies.
429
Applying the classical Rayleigh criterion (Rac = 4π 2 ) using only
430
the fracture properties neglects the stabilizing effect of the ma-
431
trix and would considerably underestimate the required fracture
432
aperture for the onset of convection as demonstrated by Graf and
433
Therrien [21]. Thus, applying the closed box model by Beck [3]
434
leads to a decrease of the critical fracture aperture by almost two
435
orders of magnitude.
436
no-transport boundary conditions at fracture walls in combination
437
with the lower diffusion coefficient of the matrix (Dm ) in the formu-
438
lation of the Rayleigh number. Thus, the critical fracture aperture
439
further decreases, such that intrafracture convection seems even
440
more likely. When, on the other hand, solute loss along the vertical
441
fracture walls is considered by applying a Dirichlet fixed concen-
442
tration at the vertical fracture walls which decreases with depth
443
as e.g. by Kassoy and Cotte [23], the critical fracture aperture for
444
intrafracture convection increases by the factor of 4.
AN US
CR IP T
425
[37] used these no-flow,
AC
CE
PT
ED
M
Simmons et al.
25
ACCEPTED MANUSCRIPT
b)
c)
d)
CR IP T
a)
AN US
Figure 5: Intrafracture convection mode 2B: solute distribution and flow path within the fracture. a) − no convection (2b) < 1.8 · 10−5 m b) (1) single convection cell per vertical fracture (2b) = 1.8 · 10−5 m c) (2) two convection cells per vertical fracture (2b) = 2.0 · 10−5 m d) (4) four convection cells per vertical fracture (2b) = 5.0 · 10−5 m
The simulated convective patterns of intrafracture convection (mode 2B)
446
are displayed in Fig. 5 and listed in Table 2. The convective pattern just
447
above the critical Rayleigh number is a single convection cell in
448
each vertical fracture (Fig. 5b). This is in agreement with findings
449
by Wang et al. [44], who predicted that a single cell is the most
450
likely mode of convection in a square fracture when the length ratio
451
between fracture aperture and adjacent matrix block width as well
452
as the conductivity ratio between fracture and matrix are small,
453
both of which is the case in the scenario displayed in Fig. 5b. Sim-
454
ilarly, Tournier et al. [41] numerically showed that a single cell is
455
the most likely steady-state convective pattern in a 2-dimensional
456
vertical fracture embedded in a heat-conducting matrix. This is
457
because conduction into and within the matrix dampens higher or-
458
der convective modes (many vertically elongated side-by-side cells)
459
such that higher Rayleigh numbers are required to increase the
460
number of convection cells. This is also demonstrated in Figs. 5c
AC
CE
PT
ED
M
445
26
ACCEPTED MANUSCRIPT
461
and 5d where the convective patterns for larger fracture apertures
462
and thus increasing Rayleigh numbers are shown, producing up to
463
four vertically elongated side-by-side convection cells (Fig. 5d). For brevity, the effect of fracture inclination is neglected in this
465
study. Theoretical investigations by Zhao et al. [51, 54] of con-
466
vection in an inclined fracture with constant temperature gradient
467
along the inclined fracture walls suggest that the critical Rayleigh
468
number of a vertical fracture is remarkably different from that of
469
an inclined fracture [54]. On the other hand, Tournier et al. [41],
470
who considered a conducting matrix adjacent to the fracture, as-
471
sumed that the effect of fracture inclination is smooth for moderate
472
inclinations. However, the effect of the fracture inclination on the
473
intrafracture convection mode should be considered in future nu-
474
merical models.
475
3.3. Convection in a 3-dimensional regular fracture circuit
ED
M
AN US
CR IP T
464
We combine geometries shown in Figs. 1a and 1b in a 3-dimensional
477
regular fracture circuit (Fig. 2a). This is the first numerical 3-dimensional
478
combined convection scenario that allows for interfracture convection mode
479
1, intrafracture convection mode 2B, and for a combination of both modes
480
(named “mode 12B”).
CE
481
PT
476
The onset of convection in this 3-dimensional regular fracture circuit was
determined numerically. Table 2 shows that, similar to interfracture convec-
483
tion (mode 1) in a 2-dimensional regular fracture circuit, the critical fracture
484
aperture for the onset of convection in a 3-dimensional regular fracture circuit
485
is 1.4 · 10−5 m. As the critical fracture aperture for intrafracture convection
AC 482
27
ACCEPTED MANUSCRIPT
486
(mode 2B) is much higher (1.8 · 10−5 m) than that for interfracture convec-
tion (1.4 · 10−5 m), it can be concluded that interfracture convection (mode
488
1) will dominate over intrafracture convection (mode 2B). It can further be
489
concluded that the formulation of the new critical Rayleigh number as of Eq.
490
(7) found for convection in a 2-dimensional fracture circuit is also valid for a
491
regular 3-dimensional fracture circuit. b)
c)
d)
AN US
a)
CR IP T
487
ED
M
Figure 6: Convection in combined fracture scenario: solute distribution and flow path in the fracture circuit. a) - no convection for (2b) < 1.4 · 10−5 m. b) to d) interfracture convection only for (2b) = 1.4 · 10−5 m, 2.0 · 10−5 m, 5.0 · 10−5 m, respectively.
Fig. 6 shows convective patterns in 3-dimensional regular fracture cir-
493
cuits. Streamlines within each fracture are straight with no component in
494
the y-direction. This indicates that the regarded convection problems are
495
purely 2-dimensional and convective patterns correspond to those of purely
496
interfracture convection (mode 1).
CE
497
PT
492
Convective patterns remained to be 2-dimensional interfracture convec-
tion (mode 1) if the aspect ratio of the fracture circuit is modified by mul-
499
tiplying and dividing (one at a time) the extent of the fracture circuit in
500
each dimension by up to 1.5. Different scenarios and geometries of the re-
501
sulting fracture circuit dimensions with the highest and lowest aspect ratios
502
in each dimension are listed in Table 3. The range of aspect ratios given
AC
498
28
ACCEPTED MANUSCRIPT
503
in Table 3 covers the critical wave lengths for the lowest critical
506
number given in Eq. (4), which corresponds to a single intrafrac-
507
ture (mode 2B) convection cell in a vertical fracture with a fracture
508
length to fracture height ratio (L/h) of 0.7. However, the critical
509
Rayleigh numbers for larger aspect ratios (including aspect ratio
510
1 for a square fracture as used before) hardly deviate and do not
511
change the critical fracture apertures determined in section 3.2.
512
Ramazanov [32] determined the lowest critical Rayleigh number
513
for interfracture convection (mode 1) around a fracture circuit of
514
aspect ratio (2B/h) of 1.2. Thus, allowing both modes of convection
515
(mode 1 and mode 2B) in the potentially most favourable fracture
516
circuit geometries for convection ensures that no mode is favoured
517
by the geometry.
M
AN US
CR IP T
505
Rayleigh numbers in previous studies. Wang et al. [44] give a mini√ mum critical wave length of 2π for the minimum critical Rayleigh
504
Table 3 confirms that the strength of convection (Sh) increases with in-
519
creasing height, increasing area, decreasing width and decreasing aspect ratio
520
of the fracture circuit as previously shown by Vujevi´c et al. [43]. A new find-
521
ing is that the strength of convection is invariant to changes of the length of
522
a fracture circuit, such that Sh is identical for the 2-dimensional represen-
523
tation and for the 3-dimensional fracture circuit. Therefore, a 2-dimensional
524
representation captures all relevant processes for a regular fracture circuit,
AC
CE
PT
ED
518
525
because intrafracture convection appears to be absent in all cases.
29
ACCEPTED MANUSCRIPT
Table 3: Convection in a fracture circuit with varying geometry
strength of convection Sh 1.28 1.28 1.28 1.28 1.16 1.00 1.93 1.39 1.35
CR IP T
fracture circuit convective aspect ratio area pattern 2 (2B/h) m 1 36 1 36 1 36 1 36 0.67 24 0.67 24 0.75 48 1.33 48 1.2 48
AN US
Fracture circuit fracture width×height×length aspect ratio in [m] (L/h) 6 × 6 × 0.2 (2D) 0.03 6 × 6 × 6 (3D) 1 6 × 6 × 4 (3D) 0.67 6 × 6 × 8 (3D) 1.5 4 × 6 × 6 (3D) 1 6 × 4 × 6 (3D) 1.5 6 × 8 × 6 (3D) 0.75 8 × 6 × 6 (3D) 1 7.2 × 6 × 4.2(3D) 0.7
ED
M
Strength of convection Sh depending on fracture circuit geometry for a 3-dimensional regular fracture circuit. Fracture aperture is 1.5 · 10−5 m. The 2-dimensional fracture circuit like in Fig. 1a is listed as reference.
Fig. 7 compares the strength of interfracture convection (mode 1) in a
527
2-dimensional regular fracture circuit, intrafracture convection (mode 2B) in
528
isolated vertical fractures, and convection in a 3-dimensional regular fracture
529
circuit using the Sherwood number Sh (Eq. (6)). Fig. 7 shows that the
530
Sh curve of the 3-dimensional regular fracture circuit is identical to that of
531
the 2-dimensional circuit. This indicates that all features of interfracture
532
convection in a regular fracture circuit can be captured with a 2-dimensional
AC
CE
PT
526
533
representation. Intrafracture convection (mode 2B) in two isolated vertical
534
fractures is weaker (smaller Sh) and therefore of subordinate importance.
30
1.3
1.05
1.25
1.04
1.2 1.15 1.1 1.05
1.03 1.02 1.01 1
1 10
20
30
40
Rafr x (2b)/(h rD)
50
60
0.99
8.16
5
10
Rafr x (2b)/(h rD)
CR IP T
Sherwood number
Sherwood number
ACCEPTED MANUSCRIPT
15
535
536
3.4. Effect of matrix permeability
AN US
Figure 7: Strength of convection (Sherwood number Sh) for inter- and intrafracture convection separately and for the combined convection scenario. Colored symbols indicate convective patterns in each case as in Tab. 2.
In all cases described above, the rock matrix permeability is km = 1 ·
10−18 m2 such that the matrix is essentially impermeable. We increase the
538
permeability of the matrix around the 3-dimensional regular fracture circuit
539
(Fig. 2a) and its 2-dimensional representation (Fig. 1a) to km = 1 · 10−17
M
537
m2 , km = 5 · 10−17 m2 and km = 1 · 10−16 m2 , respectively, to demonstrate
541
the effect of matrix permeability.
ED
540
Fig. 8 shows the strength of convection (Sh) depending on km and (2b).
543
Higher matrix permeability enhances convection because convective flow is
544
no longer restricted to fractures and to the fracture circuit, but can also
545
develop in the matrix. Fig. 8 shows that critical fracture aperture remains
546
at 1.4 · 10−5 m for km = 1 · 10−17 m2 and decreases to 0.9 · 10−5 m for
CE
PT
542
km = 5 · 10−17 m2 . For km = 1 · 10−16 m2 , the Rayleigh number for the
548
matrix is above the critical Rayleigh number, such that a critical fracture
549
aperture can not be determined as the occurrence of free convection does
550
not depend on the presence of fractures. In the last case of high matrix
551
permeability, the classical Rayleigh criterion may be applied to
552
determine the onset of convection in fractured rock using the bulk
AC
547
31
ACCEPTED MANUSCRIPT
permeability as proposed by Simmons et al. [37]. This is only
554
possible in a system where fracture permeability is small relative
555
to matrix permeability such that flow is predominately controlled
556
by matrix permeability as stated by Graf and Therrien [21].
CR IP T
553
2.5
2
AN US
Sherwood number (Sh)
3
-16
km=1·10 m² -17
km=5·10 m²
1.5
-17
km=1·10 m² -18
km=1·10 m²
onset of convection 1
1.5
2
2.5
3
3.5
4
fracture aperture (2b)
ED
0.5
M
1
4.5
5 5.5 -5 · 10 m
558
3.5. More complex scenarios In addition to studying a regular fracture circuit, we also give an example
CE
557
PT
Figure 8: Strength of convection (Sh) in the combined convection scenario depending on km and (2b).
of a more complex geometry. We thereto reduce the extent of the fractures
560
to 4 m and add two more fractures, such that the fracture network consists
AC
559
561
of 6 instead of 4 fractures (Fig. 2b). The 6 fractures are all connected and
562
form a fracture circuit with some “dead-end fractures” at the corners of the
563
fracture network. All fractures traverse the entire domain in y-direction. The
564
location of the center of each fracture is given in Table 4. 32
ACCEPTED MANUSCRIPT
Table 4: Fracture location in the complex fracture circuit consisting of 6 fractures
CR IP T
Fracture x-coordinate z-coordinate orientation of fracture center [m] vertical 2.2 6.6 vertical 4.6 4.6 vertical 5.8 7.2 horizontal 3.0 5.0 horizontal 6.0 6.0 horizontal* 4.0 8.0/9.0
AN US
* the vertical position of this fracture is variable and is 8.0 m in the 3 m high fracture circuit and 9.0 m in the 4 m high fracture circuit.
Based on the observations from interfracture convection (mode 1) and in-
566
trafracture convection (mode 2B), we hypothesize that the critical Rayleigh
567
number depends on either the fracture height or the height of the fracture
568
circuit. Studying a case of constant fracture height and varying circuit height
569
allows us to evaluate whether (i) the fracture circuit is the dominating struc-
570
ture in the fracture network such that predicting the onset of convection is
571
possible using a 2-dimensional representation, or (ii) the vertical fractures
572
in the network promote interfracture convection such that a 3-dimensional
573
representation is required. To answer this question, the location of the up-
574
permost horizontal fracture is variable (red arrow in Fig. 2b), such that the
575
fracture circuit is either 3 m high (Fig. 9a), or 4 m high (Fig. 9b).
AC
CE
PT
ED
M
565
576
If the circuit height is only 3 m and thus smaller than the height of each
577
vertical fracture, intrafracture convection (mode 2B) develops (Fig. 9a). The
578
resulting convection pattern is combined inter- and intrafracture convection
579
(mode 12B) where intrafracture convection cells in the vertical fractures are 33
ACCEPTED MANUSCRIPT
hydraulically connected along the horizontal fractures as shown in Fig. 9a.
581
However, if the fracture circuit height equals 4 m like the individual fracture
582
height, interfracture convection (mode 1) dominates (Fig. 9b). Intrafrac-
583
ture convection (mode 2B) is only observed in the dead-end fractures when
584
the fracture aperture exceeds the critical value for intrafracture convection
585
(2b)c = 2.0 · 10−5 m as of Eq. (5). b)
M
AN US
a)
CR IP T
580
PT
ED
Figure 9: More complex fracture network with 6 fractures: solute distribution and flow paths in the fracture circuit. Fracture length is 4 m, fracture aperture (2b) = 2.0 · 10−5 m. a), (1) Fracture circuit height is 3 m, and lower than the height of individual vertical fractures, combined intra- and interfracture convection (mode 12B) develops. b) Fracture circuit height is 4 m, identical to the height of individual vertical fractures, interfracture convection mode 1 dominates.
For both heights (3 and 4 m) of the fracture circuit, a 2-dimensional cross-
587
section of the 3-dimensional network is simulated in order to compare the
588
results with respect to convective pattern and strength of convection (Sh).
589
The results are summarized in Table 5. If fracture and circuit height are not
AC
CE
586
590
identical (4 m vs. 3 m), predictions based on a 2-dimensional representation
591
are critical because the 2-dimensional model neglects intrafracture convec-
592
tion (mode 2B) in dead-end fractures and combined convection (mode 12B)
593
such that the occurrence and the strength of convection are underestimated. 34
ACCEPTED MANUSCRIPT
This is because in the 3 m high fracture circuit with combined convection
595
(mode 12B), the dead-end fractures vertically elongate the fracture circuit
596
and promote convection by increasing the vertical density contrast. If, how-
597
ever, fracture and circuit height are identical (both 4 m), the 2-dimensional
598
cross-section is an appropriate simplification showing the same convective
599
pattern and the same Sherwood number for (2b) ≤ 2 · 10−5 m because there
600
is no convection in the dead-end fractures.
CR IP T
594
AN US
Table 5: Observed convective patterns in a fracture circuit consisting of 6 fractures
3-dimensional network deviance in Sh fracture circuit height between 2d and 3d 3m 4m 3m 4m 0% , (1) 4% 0% , (1) 5% 0% , (1) * 6% 1%
ED
M
Fracture 2-dimensional network aperture fracture circuit height ·10−5 [m] 3 m 4m 1.5 a 1.6 1.7 b 1.8 1.9 c 2.0 a
onset of mode 1 convection in a 4 m high circuit predicted by Eq. (7). onset of mode 1 convection in a 3 m high circuit predicted by Eq. (7). c onset of mode 2B convection in a 4 m high fracture predicted by Eq. (4). : interfracture convection along the fracture circuit (mode 1). *: only interfracture convection in the circuit, but intrafracture convection in the dead-end fractures. , (1): combination of interfracture and intrafracture convection (mode 12B).
AC
CE
PT
b
601
Results in Table 5 show that predicting the onset of convection in com-
602
plex fracture networks using theoretical considerations (Rayleigh criterion)
603
are difficult. Beside the effect of long vertical fractures promoting convection,
604
additional fracture intersections along the fracture circuit disturb interfrac35
ACCEPTED MANUSCRIPT
ture flow and thus inhibit convection, such that complex fracture net-
606
works impede the prediction of free convective patterns and vigor
607
of convection as already stated by Shikaze et al. [34].
608
4. Summary and conclusions
CR IP T
605
Continuing prior work on convective flow modes in fractured rock by
610
Wang et al. [44], Malkovsky and Pek [26], Ramazanov [32], or Simmons et
611
al. [37], this study examines inter- and intrafracture haline free convection
612
and their combination. The onset of convection, convective patterns and the
613
dominating mode of convection are determined. The key new findings of this
614
study are:
AN US
609
1. The critical Rayleigh number for the onset of convection in a 2- and
616
3-dimensional regular fracture circuit is determined numerically to be
617
half of the critical Rayleigh number for the onset of convection in a
618
single vertical fracture determined by Wang et al. [44].
ED
M
615
2. The most likely mode of convection in a regular fracture circuit is inter-
620
fracture convection along the circuit (mode 1). When both inter- and
621
intrafracture convection are theoretically possible, the onset of con-
622
vection is marked by the onset of interfracture convection (mode 1). Convective patterns correspond to streamlines along the circuit.
3. Interfracture convection (mode 1) is stronger in terms of mass transport
AC
624
CE
623
PT
619
625
than intrafracture convection (mode 2B). Sherwood numbers for in-
626
trafracture convection in two vertical fractures are considerably smaller
627
than for interfracture convection in a regular fracture circuit of the same
628
height. 36
ACCEPTED MANUSCRIPT
4. In more complex fracture networks, convection is hard to be theoret-
630
ically predicted using Rayleigh criteria. Vertical “dead-end fractures”
631
exceeding the fracture circuit promote convection, whereas additional
632
fracture intersections along the fracture circuit constrain convection.
CR IP T
629
5. A 2-dimensional representation of a 3-dimensional fracture network
634
only yields good results in predicting the onset and strength of con-
635
vection if the network is composed of regular fracture circuits, where
636
the circuit height equals the individual fracture height. Otherwise, 2-
637
dimensional representations tend to underestimate the occurrence and
638
strength of convection, as intrafracture convection (mode 2B) and com-
639
bined convection (mode 12B) are neglected.
AN US
633
Considering the easier onset and the higher mass transport rates of in-
641
terfracture convection (mode 1) compared to intrafracture convection (mode
642
2B), interfracture convection along a fracture circuit (mode 1) is the most
643
likely mode of convection in a fracture network consisting of regular frac-
644
ture circuits. However, more complex fracture networks, which not exclu-
645
sively consist of regular fracture circuits, challenge this rule and require 3-
646
dimensional representations. A fully 3-dimensional modelling approach of
647
free convection in fractured rock remains challenging in terms of computa-
648
tional power. It took up to 10 times more CPU time to run a 3-dimensional
649
complex scenario relative to its 2-dimensional representation. A series of
AC
CE
PT
ED
M
640
650
2-dimensional cross-sections perpendicular to the fracture planes of a frac-
651
ture network (covers mode 1) combined with a theoretical evaluation of the
652
existence of intrafracture free convection in isolated vertical fractures using
653
Eq. (4) might be a practical solution to at least approximate the probability 37
ACCEPTED MANUSCRIPT
654
of inter- and intrafracture convection in fractured rock. Nevertheless, this
655
approach will not be able to examine combined convection (mode 12). This study focusses on simple deterministic fracture networks,
657
consisting of few fractures and therefore represents an academic
658
configuration rather than a real-world field setting. However, the
659
new findings presented may help to better assess the probability
660
of free convection in various fracture networks. Furthermore, the
661
conclusions drawn in this study lay the foundation for further re-
662
search on more complex fracture networks.
AN US
CR IP T
656
On the basin scale, interfracture convection enables convective
664
flow and transport over larger distances than intrafracture convec-
665
tion, especially if fracture length is small compared to the consid-
666
ered scale. Cui et al. [8] stated that isolated vertical faults do not
667
affect basin-scale free convection, and Yang et al. [49] showed that
668
open vertical faults may determine convective patterns in a sedi-
669
mentary basin only if they are hydraulically connected via highly
670
permeable horizontal sediment layers. Thus, further investigation
671
of free convection in complex fracture networks is required.
PT
ED
M
663
In field studies, the exact geometry of fracture networks is often unknown.
673
Parameters like average and maximum vertical extent and permeability of
674
fractures should be estimated to be able to calculate a (critical) fracture
675
Rayleigh number (Eqs. (3) and (4)), such that the existence of intrafracture
AC
CE
672
676
convection (mode 2) can be assessed. Assessment of interfracture convection
677
(mode 1) is more difficult, because the existence and geometry of fracture
678
circuits is harder to detect than the existence of fractures. An elementary
679
relationship between the strength of interfracture convection and average 38
ACCEPTED MANUSCRIPT
fracture length and fracture density is given in Vujevi´c et al. [43]. How-
681
ever, more statistically oriented research is required to better quantify these
682
relationships.
AC
CE
PT
ED
M
AN US
CR IP T
680
39
ACCEPTED MANUSCRIPT
683
Acknowledgement This research was supported by the Deutsche Forschungsgemeinschaft
685
(DFG) under grant number GR 3463/2-1. We thank the editorial board of
686
AWR (D. Andrew Barry) for handling our manuscript. We also thank four
687
anonymous reviewers whose constructive comments have helped to improve
688
the manuscript.
AC
CE
PT
ED
M
AN US
CR IP T
684
40
ACCEPTED MANUSCRIPT
APPENDIX Nomenclature
AC
CR IP T
fracture aperture critical fracture aperture in a single vertical fracture critical fracture aperture in the fracture circuit fracture spacing relative solute concentration hydrodynamic dispersion tensor in the matrix hydrodynamic dispersion tensor in the fracture effective diffusion coefficient in the matrix free solution diffusion coefficient unit vector in vertical direction gravitational acceleration height of fracture domain height equivalent fresh water head identity matrix permeability of (isotropic) porous medium fracture permeability matrix permeability freshwater hydraulic conductivity tensor of porous medium freshwater hydraulic conductivity tensor of porous medium length of the domain (y-direction) fluid pressure Darcy flux vector in the porous matrix Darcy flux vector in the fracture diffusivity ratio between matrix and fracture Rayleigh number (homogeneous porous medium) Rayleigh number in the fracture critical Rayleigh number critical Rayleigh number in a regular fracture circuit critical Rayleigh number in a single vertical fracture
PT
ED
[L] [L] [L] [L] [-] [L2 T−1 ] [L2 T−1 ] [L2 T−1 ] [L2 T−1 ] [-] [LT−2 ] [L] [L] [L] [-] [L2 ] [L2 ] [L2 ] [LT−1 ] [LT−1 ] [L] [ML−1 T−2 ] [LT−1 ] [LT−1 ] [-] [-] [-] [-] [-] [-]
CE
691
(2b) (2b)c (2b)c,circ (2B) c D Dfr Dm D0 ez g h H h0 I k kfr km K0 K0,fr L p q qfr rD Ra Rafr Rac Rac,circ Rac,fr
AN US
690
M
689
41
ACCEPTED MANUSCRIPT
CR IP T
AN US
M
Governing equations
AC
694
CE
693
specific storage of porous medium specific storage of the fracture Sherwood number time total solute mass flux through top of the domain width of the domain in x-direction horizontal coordinates vertical coordinate longitudinal dispersion coefficient longitudinal dispersion coefficient in the fracture transverse dispersion coefficient transverse dispersion coefficient in the fracture solutal expansion coefficient fluid compressibility matrix compressibility matrix porosity dynamic viscosity of the fluid fluid density reference fluid density tortuosity incline of fracture (ϕ for a vertical fracture is 0) exchange term for fluid in the porous matrix exchange term for fluid in the fracture exchange term for solute in the porous matrix exchange term for solute in the fracture
ED
[L−1 ] [L−1 ] [-] [T] [ML−2 T−1 ] [L] [L] [L] [L] [L] [L] [L] [-] [LT2 M−1 ] [LT2 M−1 ] [-] [ML−1 T−1 ] [ML−3 ] [ML−3 ] [-] [-] [T−1 ] [LT−1 ] [T−1 ] [LT−1 ]
PT
692
SS SS,fr Sh t j W x, y z αL αL,fr αT αT,fr β γ fl γm θ µ ρ ρ0 τ ϕ Ωfluid Ωfluid,fr Ωsolute Ωsolute,fr
695
The Darcy equation describes the Darcy flux in a porous medium as a
696
function of both, the equivalent freshwater head h0 and the relative fluid
42
ACCEPTED MANUSCRIPT
density (ρ − ρ0 )/ρ0 [-] [2]: q = −K0
ρ(c) − ρ0 ez ∇h0 + ρ0
(9)
CR IP T
697
698
Assuming Darcy flow in the fracture, the Darcy equation for density-driven
699
flow in the fracture reads [21]:
700 701
AN US
ρ(c) − ρ0 qfr = −K0,fr ∇h0 + ez cosϕ ρ0
The conservation equations for fluid mass in the 3-dimensional porous medium (11) and in the 2-dimensional fracture (12) are [10, 21]:
ρ(c) − ρ0 ∂h0 ∇ · K0 (∇h0 + ez ) − SS = Ωfluid ρ0 ∂t
(11)
M
ρ(c) − ρ0 ∂h0 (2b) ∇ · K0,fr (∇h0 + ez cosγ) − SS,fr = (2b)Ωfluid,fr (12) ρ0 ∂t The specific storage in matrix (SS ) and fracture (SS,fr ) are defined as [14, 21]:
ED
703
PT
702
(10)
(13)
SS,fr = ρ0 gγ fl
(14)
CE
704
SS = ρ0 g(γ m + θγ fl )
The conservation equations for solute mass in the porous medium (15) and
706
in the fracture (16) are [10, 21]:
AC
705
707
∇ · (θD∇c − qc) −
∂(θc) = Ωsolute ∂t
∂c (2b) ∇ · (Dfr ∇c − qfr c) − = (2b)Ωsolute,fr ∂t 43
(15)
(16)
ACCEPTED MANUSCRIPT
708
The hydrodynamic dispersion tensor in the matrix (D) and in the fracture
709
(Dfr ) are defined as [2, 20]:
710
and Dfr = (αL,fr − αT,fr )
(17)
qfr qT fr + αT,fr |qfr |I + D0 I |qfr |
(18)
where the effective diffusion coefficient in the porous matrix is [39]:
AN US
711
qqT + αT |q|I + Dm I |q|
CR IP T
θD = (αL − αT )
Dm = θτ D0
713 714
Constitutive equations
Porous medium freshwater hydraulic conductivity tensor in the matrix K0 and in the fracture K0,fr are given by [2, 5, 20]:
M
712
ED
K0 =
715
km Iρ0 g µ
(20)
ρ0 g µ
(21)
PT
K0,fr = kfr I
where fracture permeability kfr for an open fracture is defined as [2, 5] kfr =
(2b)2 12
(22)
AC
CE
716
(19)
717 718
Fluid density ρ is determined by relative concentration c and is calculated
using a linear density function [11, 37]: ρ(c) = ρ0 (1 + βc)
44
(23)
ACCEPTED MANUSCRIPT
The dependency of viscosity on salinity is neglected as it has been shown that
720
the effect is negligible for density-dependent flow and transport compared to
721
the density effect of salinity [18]. Thus, fluid viscosity is assumed to be
722
constant in accordance with most studied applying Rayleigh theory [26, 32,
723
34, 37, 44, 50].
AC
CE
PT
ED
M
AN US
CR IP T
719
45
ACCEPTED MANUSCRIPT
724
References [1] P. Alt-Epping, C. Zhao, Reactive mass transport modelling of a three-
726
dimensional vertical fault zone with a finger-like convective flow regime,
727
J. Geochemical Explor.(2010),106(1-3):8-23.
728
http://dx.doi.org/10.1016/j.gexplo.2009.12.007.
730
731
[2] J. Bear, Dynamics of fluids in porous media, American Elsevier Publishing Co., New York, 1972.
AN US
729
CR IP T
725
[3] J.L. Beck, Convection in a box of porous material saturated with fluid,
732
Phys. Fluids(1972),15:1377-1383.
733
http://dx.doi.org/10.1063/1.1694096
[4] A. Bejan, K.R. Khairy, Heat and mass transfer by natural convection
735
in a porous medium, Int. J. Heat Mass Tran.(1985),28(5):909-918.
736
http://dx.doi.org/10.1016/0017-9310(85)90272-8.
ED
737
M
734
[5] B. Berkowitz, Characterizing flow and transport in fractured geological media: A review, Adv. Water Resour.(2002),25:861-884.
739
http://dx.doi.org/10.1016/S0309-1708(02)00042-8.
741
rocks: Correlation to porosity and hydraulic conductivity, J. Contam. Hydrol.(2001),53:85100.
AC
742
[6] T.B. Boving, P. Grathwohl, Tracer diffusion coefficients in sedimentary
CE
740
PT
738
743
hhtp://dx.doi.org/10.1016/S0169-7722(01)00138-3.
744
[7] J.P. Caltagirone, Convection in a porous medium, in: J. Zierep, H.
745
Oertel jr. (Eds.), Convective transport and instability phenomena, G.
746
Braun, Karlsruhe(1982), pp.199-232. 46
ACCEPTED MANUSCRIPT
[8] T. Cui, J. Yang, I.M. Samson, Numerical modeling of hydrothermal
748
fluid flow in the Paleoproterozoic Thelon Basin, Nunavut, Canada, J.
749
Geochem. Explor.(2010),106,69-76.
750
http://dx.doi.org/10.1016/j.gexplo.2009.12.008.
751 752
CR IP T
747
[9] S.H. Davis, Convection in a box: linear theory, J. Fluid Mech.(1967),30:465478. http://dx.doi.org/10.1017/S0022112067001545
[10] H.-J.G. Diersch, O. Kolditz, Coupled groundwater flow and trans-
754
port: 2. Thermohaline and 3D convection systems, Adv. Water Re-
755
sour.(1998),21:401425. http://dx.doi.org/10.1016/S0309-1708(97)00003-
756
1.
757
AN US
753
[11] H.-J.G. Diersch, O. Kolditz, Variable-density flow and transport in porous media: approaches and challenges, Adv. Water Resour.(2002),25:899-
759
944.
760
http://dx.doi.org/10.1016/S0309-1708(02)00063-5.
ED
M
758
[12] P.A. Forsyth, P.H. Sammon, Practical considerations for adaptive im-
762
plicit methods in reservoir simulation, J. Comput. Phys.(1986),62:265-
763
281.
764
http://dx.doi.org/10.1016/0021-9991(86)90127-0.
CE
[13] P.A. Forsyth, M.C. Kropinski, Monotonicity considerations for saturated-
AC
765
PT
761
766
unsaturated subsurface flow, SIAM J. Sci. Comp.(1997),18:1328-1354.
767
http://dx.doi.org/10.1137/S1064827594265824.
768 769
[14] E.O. Frind, Simulation of long-term transient density-dependent transport in groundwater, Adv. Water Resour.(1982),5:7388. http://dx.doi.org/10.1016/030947
ACCEPTED MANUSCRIPT
770
1708(82)90049-5. [15] G.Z. Gershuni, E.M. Zhukhovitskii, Convective stability of incompress-
772
ible fluids, translated from Russian by D. Louvish, Keter Publishing
773
House Jerusalem Ltd., Jerusalem, 1976.
CR IP T
771
[16] G. Garven, S.W. Bull, R.R. Large, Hydrothermal fluid flow models
775
of stratiform ore genesis in the McArthur Basin, Northern Territory,
776
Australia, Geofluids(2001),1:289-311.
777
http://dx.doi.org/10.1046/j.1468-8123.2001.00021.x.
AN US
774
[17] T. Graf, Modeling coupled thermohaline flow and reactive transport in
779
discretely-fractured porous media, PhD thesis, Universit´e Laval(2005),
780
209pp. Available at
.
782
[18] T. Graf, M.C. Boufadel, Effect of viscosity, capillarity and grid spacing on thermal variable-density flow, J. Hydrol.(2011),400:4157.
ED
781
M
778
[19] T. Graf, R. Therrien, Variable-density groundwater flow and solute
784
transport in porous media containing nonuniform discrete fractures,
785
Adv. Water Resour.(2005),28:1351-1367.
786
http://dx.doi.org/10.1016/j.advwatres.2005.04.011. [20] T. Graf, R. Therrien, Variable-density groundwater flow and solute transport in irregular 2D fracture networks, Adv. Water Resour.(2007),
AC
788
CE
787
PT
783
789
790 791
30:455-468. http://dx.doi.org/10.1016/j.advwatres.2006.05.003.
[21] T. Graf, R. Therrien, Stable-unstable flow of geothermal fluids in fractured rock, Geofluids(2009),9:138-152. 48
ACCEPTED MANUSCRIPT
792
http://dx.doi.org/10.1111/j.1468-8123.2008.00233.x. [22] C.W. Horton, F.T. Rogers, Convection currents in a porous medium,
794
J. Appl. Phys.(1945),16:367370. http://dx.doi.org/10.1063/1.1707601.
CR IP T
793
795
796
[23] D.R. Kassoy, B. Cotte, The effect of sidewall heat loss on convection in a saturated porous vertical slab, J. Fluid Mech.(1985),152:361-378.
798
http://dx.doi.org/10.1017/S0022112085000738.
799
AN US
797
[24] E.R. Lapwood, Convection of a fluid in a porous media, Proc. Cam-
800
bridge Philos. Soc.(1948),44(4):508-521.
801
http://dx.doi.org/10.1017/S030500410002452X.
[25] R.P. Lowell, C.-T. Shyu, On the onset of convection in a water-saturated
803
porous box: effect of conducting walls, Lett. heat mass trans.(1978),5:
804
371-378. http://dx.doi.org/10.1016/0094-4548(78)90046-2.
ED
805
M
802
[26] V.I. Malkovsky, A.A. Pek, Conditions for the onset of thermal convection of a homogeneous fluid in a vertical fault, Petrology+(1997),5(4):381-
807
387.
809
fluid in open vertical faults, Izvestiya, Physics of the Solid Earth(2004),40(8):672679.
AC
810
[27] V.I. Malkovsky, A.A. Pek, Onset of thermal convection of single-phase
CE
808
PT
806
811
[28] H.D. Murphy, Convective instabilities in vertical fractures and faults,
812
J. Geophys. Res.(1979),84:6121-6130.
813
http://dx.doi.org/10.1029/JB084iB11p06121. 49
ACCEPTED MANUSCRIPT
[29] C.E. Neuzil, How permeable are clays and shales?, Water Resour.
815
Res.(1994),30:145150.
816
http://dx.doi.org/10.1029/93WR02930.
817
[30] D.A. Nield, Onset of thermohaline convection in a porous medium,
818
Water Resour. Res.(1968),4(3):553-560.
819
http://dx.doi.org/10.1029/WR004i003p00553.
821
822
[31] D.A. Nield, A. Bejan, Convection in porous media, Springer, New York,
AN US
820
CR IP T
814
2013.
[32] M. Ramazanov, Convective stability of a fluid in two hydrodynamically connected vertical faults, Izvestiya, Physics of the Solid Earth(2003),39(12):1014-
824
1020.
M
823
[33] J.M. Sharp, Jr., T.R. Fenstemaker, C.T. Simmons, T.E. McKenna, J.K.
826
Dickinson, Potential thermohaline convection in a shale-rich sedimen-
827
tary basin: Example from the Gulf of Mexico Basin in South Texas.
828
AAPG Bull.(2001),85:2089-2110.
PT
ED
825
[34] S.G. Shikaze, E.A. Sudicky, F.W. Schwartz, Density-dependent solute
830
transport in discretely-fractured geological media: is prediction possi-
831
ble? J. Contam. Hydrol.(1998),34:273-291. http://dx.doi.org/10.1016/S0169-7722(98)00080-1.
AC
832
CE
829
833
[35] C.T. Simmons, K.A. Narayan, R.A. Wodding, On a test case for density-
834
dependent groundwater flow and solute transport models: The salt lake
835
problem, Water Resour. Res.(1999),35:3607-3620.
836
http://dx.doi.org/10.1029/1999WR900254. 50
ACCEPTED MANUSCRIPT
837
[36] C.T. Simmons, Variable density groundwater flow: From current challenges to future possibilities, Hydrogeol. J.(2005),13:116-119.
839
http://dx.doi.org/10.1007/s10040-004-0408-3
840
CR IP T
838
[37] C.T. Simmons, J.M. Sharp, D.A. Nield, Modes of free convection in
841
fractured low-permeability media, Water Resour. Res.(2008),44:W03431.
842
http://dx.doi.org/10.1029/2007WR006551.
[38] M.A. Simms, G. Garven, Thermal convection in faulted extensional
844
sedimentary basins: theoretical results from finite-element modeling,
845
Geofluids(2004),4,109-130.
846
http://dx.doi.org/10.1111/j.1468-8115.2004.00069.x.
847
AN US
843
[39] R. Therrien, R.G. McLaren, E.A. Sudicky, S.M. Panday, HydroGeoSphere - A Three-dimensional numerical model describing fully-integrated
849
subsurface and surface flow and solute transport, University of Wa-
850
terloo, Canada: Groundwater Simulations Group, 2010. Available at
851
http://www.ggl.ulaval.ca/fileadmin/ggl/documents/rtherrien/hydrogeosphere.pdf.
ED
M
848
PT
852
[40] R. Thiesen, K. Scheel, H. Diesselhorst, Untersuchungen u ¨ber die ther-
854
mische Ausdehnung von festen und tropfbar fl¨ ussigen K¨orpern - Bestim-
855
mung der Ausdehnung des Wassers f¨ ur die zwischen 0 und 40 liegenden Temperaturen, Wiss. Abhandlungen Der Phys. Reichsanst.(1900),3:170.
AC
856
CE
853
857
858
[41] C. Tournier, P. Genthon, M. Rabinowicz, The onset of natural con-
859
vection in vertical fault planes:consequences for the thermal regime in
51
ACCEPTED MANUSCRIPT
crystalline basements and for heat recovery experiments, Geophys. J.
861
Int.(2000),140:500-508.
862
http://dx.doi.org/10.1046/j.1365-246X.2000.00041.x.
CR IP T
860
[42] C.I. Voss, C.T. Simmons, N.I. Robinson, Three-dimensional bench-
864
mark for variable-density flow and transport simulation: matching
865
semi-analytical stability modes for steady unstable convection in an
866
inclined porous box, Hydrogeol. J.(2009),18:5-23.
867
http://dx.doi.org/10.1007/s10040-009-0556-6.
AN US
863
868
[43] K. Vujevi´c, T. Graf, C.T. Simmons, A.D. Werner, Impact of frac-
869
ture network geometry on free convective flow patterns, Adv. Water
870
Res.(2014),71:65-80. http://dx.doi.org/10.1016/j.advwatres.2014.06.001.
M
871
[44] M. Wang, D.R. Kassoy, P.D. Weidman, Onset of convection in a vertical
873
slab of saturated porous media between two conducting impermeable
874
blocks, Int. J. Heat Mass Tran.(1987),30(7):1331-1341.
875
http://dx.doi.org/10.1016/0017-9310(87)90165-7.
877 878
PT
dependent groundwater models: two-dimensional steady state unstable convection in infinite, finite and inclined porous layers, Adv. Water Resour.(2004),27:547-562. http://dx.doi.org/10.1016/j.advwatres.2004.01.003.
AC
879
[45] D. Weatherill, C.T. Simmons, C.I. Voss, N.T. Robinson, Testing density-
CE
876
ED
872
880
881
[46] D. Weatherill, T. Graf, C.T. Simmons, P.G. Cook, R. Therrien, D.A.
882
Reynolds, Discretizing the fracture-matrix interface to simulate solute
52
ACCEPTED MANUSCRIPT
883
transport. Ground Water(2008),46(4):606-615.
884
http://dx.doi.org/10.1111/j.1745-6584.2007.00430.x. [47] Y. Xie, C.T. Simmons, A.D. Werner, Speed of free convective fingering
CR IP T
885 886
in porous media, Water Resour. Res.(2011),47:W11501,16. http://dx.doi.org/
887
10.1029/2011WR010555.
[48] J. Yang, K. Latychev, R.N. Edwards, Numerical computation of hy-
889
drothermal fluid circulation in fractured earth structures, Geophys. J.
890
Int.(1998),135:627-649.
891
http://dx.doi.org/10.1046/j.1365-246X.1998.00669.x
AN US
888
[49] J. Yang, R.R. Large, S.W. Bull, Factors controlling free thermal con-
893
vection in faults in sedimentary basins: implications for the formation
894
of zinc-lead mineral deposits, Geofluids(2001),4:237-247.
895
http://dx.doi.org/10.1111/j.1468-8123.2004.00084.x
ED
M
892
[50] C. Zhao, B.E. Hobbs, H. M¨ uhlhaus, A. Ord, G. Lin, Convective insta-
897
bility of 3-D fluid-saturated geological fault zones heated from below,
898
Geophys. J. Int.(2003),155:213-220. http://dx.doi.org/10.1046/j.1365-
899
246X.2003.02032.x
[51] C. Zhao, B.E. Hobbs, A. Ord, S. Peng, H.B. M¨ uhlhaus, L. Liu,
AC
901
CE
900
PT
896
902
Theoretical investigation of convective instability in inclined
903
and fluid-saturated three-dimensional fault zones, Tectono-
904
physics (2004),387:47-64. http://dx.doi.org/10.1016/j.tecto.2004.06.007.
905
53
ACCEPTED MANUSCRIPT
906
[52] C. Zhao, B.E. Hobbs, A. Ord, S. Peng, H. M¨ uhlhaus, L.Liu, Doublediffusive-driven convective instability of three-dimensional fluid-saturated
908
geological fault zones heated from below, Math. Geol.(2005),37(4):373-
909
391. http://dx.doi.org/10.1007/s11004-005-5954-2
CR IP T
907
[53] C. Zhao, B.E. Hobbs, A. Ord, M.K¨ uhn, H. M¨ uhlhaus, S. Peng, Nu-
911
merical simulation of double-diffusive driven convective flow and rock
912
alteration in three-dimensional fluid-saturated geological fault zones,
913
Comp. Methods Appl. Mech. Engrg.(2006),195:2816-2840.
914
http://dx.doi.org/10.1016/j.cma.2005.07.008
916
917
[54] C. Zhao, B.E. Hobbs, A. Ord, Convective and advective heat transfer in geological systems, Springer, Berlin, 2008. [55] C. Zhao, B.E. Hobbs, A. Ord, Fundamentals of computational
M
915
AN US
910
geoscience: Numerical methods and algorithms, Springer, Berlin,
919
2009.
920
ED
918
[56] A. Zebib, D.R. Kassoy, Onset of natural convection in a box watersaturated porous media with large temperature variation, Phys. Fluids(1977),20:4-
922
9. http://dx.doi.org/10.1063/1.861695.
AC
CE
PT
921
54