Accepted Manuscript Analyzing wellbore stability in chemically-active anisotropic formations under thermal, hydraulic, mechanical and chemical loadings Majed F. Kanfar, Z. Chen, S.S. Rahman PII:
S1875-5100(17)30051-3
DOI:
10.1016/j.jngse.2017.02.006
Reference:
JNGSE 2059
To appear in:
Journal of Natural Gas Science and Engineering
Received Date: 31 March 2016 Revised Date:
29 January 2017
Accepted Date: 1 February 2017
Please cite this article as: Kanfar, M.F., Chen, Z., Rahman, S.S., Analyzing wellbore stability in chemically-active anisotropic formations under thermal, hydraulic, mechanical and chemical loadings, Journal of Natural Gas Science & Engineering (2017), doi: 10.1016/j.jngse.2017.02.006. 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 Title: Analyzing Wellbore Stability in Chemically-Active Anisotropic Formations under
2
Thermal, Hydraulic, Mechanical and Chemical Loadings
3
Authors Order: Majed F. Kanfar *, Z. Chen, S.S. Rahman,
4
School of Petroleum Engineering, University of New South Wales, Australia
5
*Corresponding author: Email address:
[email protected] (Majed Kanfar)
6
Keywords: anisotropy; porochemothermoelasticity; THMC; thermo-chemo-hydro-
7
mechanical; wellbore stability; shale
8
Abstract
RI PT
1
Several factors influence the stress state in subsurface rocks. In addition to pore pressure
10
and far-field in-situ stresses, thermal and chemical gradients have substantial bearings on the
11
in-situ stress state during and after drilling operations. Drilling inclined boreholes through
12
laminated formations such as shaley sands presents several challenges. Due to their laminated
13
nature, shales demonstrate high degree of elastic anisotropy, and are often chemically-
14
reactive to drilling fluid. The majority of models utilized in the industry assume a
15
homogeneous isotropic elastic static model that fails to give an accurate depiction of the
16
observed borehole failure.
17
porochemothermoelasticity are developed to simulate the drilling of an inclined borehole
18
problem in a transversely isotropic rock. Using the developed constitutive and transport
19
equations, a finite element method based numerical model is constructed to estimate the pore
20
pressure, temperature, solute concentration and stress distribution. Nonlinear conductive-
21
convective heat transfer and diffusive-advective solute transfer models are considered in this
22
work to address both high and low permeability formations. Finally, the model is used to
23
assess time-dependent wellbore stability during and after drilling. The novel pseudo-3D
24
analysis developed in this work is advantageous for real-time operations due to its
25
computational speed and stability.
M AN U
TE D
In this work, governing equations for anisotropic
EP
AC C
26
SC
9
ACCEPTED MANUSCRIPT 27
1. Introduction With skyrocketing demand for energy supply, the oil and gas industry has been
29
aggressively pursuing more ambitious ventures. Recent technological developments enabled
30
the exploitation of oil and gas prospects previously deemed unattainable, such as
31
unconventional shale reservoirs and deep water horizons. Drilling applications extend outside
32
of the fossil fuel industry, including for CO2 sequestration operations and harvesting
33
alternative sources of energy such as geothermal. These recent complex applications require
34
vigorous knowledge of the reservoir fluid and rock properties to accurately simulate coupled
35
dynamic problems. These rocks are chemically active and the physico-chemical nature should
36
be carefully analyzed. The borehole-reservoir system in most cases is a non-isothermal one.
37
Bespoke simulators are crucial to account for the solute concentration and temperature
38
gradient, and their effect on pore pressure and in-situ stresses. These coupled phenomena
39
produce complex unpredictable environments during drilling operations if they are not
40
properly modeled.
M AN U
SC
RI PT
28
Most simulators that couple fluid flow with geomechanics assume an isotropic rock
42
despite the fact that only 10% of subsurface formations exhibit true isotropic behavior (Ong,
43
1994). More than 30% of rocks classified as anisotropic have an anisotropy ratio of 1.5 for
44
Young’s modulus (Batugin and Nirenburg, 1972). The majority of shales are deemed
45
anisotropic. Shales make up approximately 75% of the drilled section and are the culprit for
46
90% of borehole instability problems (Mody and Hale, 1993). Several additional factors
47
affect shales behavior under stress such as permeability variation under different confining
48
pressures (Vishal et al., 2015, 2013), and its brittleness which is dependent on its mechanical
49
properties (Zhang et al., 2016).
EP
TE D
41
It is crucial to model the dynamic nature of a drilling problem during and after drilling.
51
Static elastic models (Aadnoy and Chenevert, 1987; McLean and Addis, 1990) fall short as
52
they are not time-dependent. Modelling the transient pore pressure behavior and its influence
53
on effective stresses were addressed by several authors using poroelasticity (Abousleiman et
54
al., 1995; Detournay and Cheng, 1988). The generalized anisotropic poroelasticity was
55
developed by Biot (1955). The majority of subsurface rocks exhibit a transversely isotropic
56
(TI) behavior caused by compaction and gravitational forces (Biot, 1955). Analytical solution
57
for pore pressure and stress distribution around a borehole in a poroelastic TI material was
58
developed by Abousleiman et al. (1995). They assume that the plane of isotropy of the TI
59
material is always perpendicular to the borehole axis.
AC C
50
ACCEPTED MANUSCRIPT Several publications have extended Biot's poroelasticity to a fully coupled
61
porothermoelastic model (Bear and Corapcioglu, 1981; Coussy, 1989; Kurashige, 1989;
62
Schiffman, 1971). A fully coupled isotropic conductive-convective porothermoelastic heat
63
and fluid flow model for boreholes with constant pressure and temperature was developed by
64
Wang and Dusseault (2003). A special formulation for isotropic porothermoelasticity for
65
swelling shale has been developed by Ghassemi and Diek (2002) to account for thermal
66
osmosis. Abousleiman and Ekbote (2005) developed an analytical solution for inclined
67
boreholes in transversely isotropic media. They assume that the plane of isotropy of the TI
68
material is always perpendicular to the borehole axis, and heat transfer is by conduction only.
69
Kanfar et al. (2016b) developed a fully coupled anisotropic porothermoelastic model using
70
the generalized plane strain (GPS) assumption that removes the abovementioned assumption
71
made by Abousleiman and Ekbote (2005).
SC
RI PT
60
Increased borehole collapse problems are observed during drilling in shale formations.
73
This is attributed to the chemically active nature of clays that are ubiquitous in shales and
74
shaley formations (Mody and Hale, 1993). Coupling this chemical phenomenon with Biot's
75
poroelastic model and its effect on chemically active rocks was first discussed by Sherwood
76
(1993). A chemo-mechanical model for wellbore stability that accounts for osmotic and
77
poroelastic effects was developed by Ghassemi and Wolfe (1999). Specialized physico-
78
chemical analytical solution for swelling shales were developed for pore pressure, stresses
79
and solute mass fraction around a vertical isotropic well (Ghassemi and Diek, 2003). Ekbote
80
and Abousleiman (2006) derived analytical solutions for porochemoelasticity in inclined
81
boreholes drilled in TI formations by assuming that the plane of isotropy is always
82
perpendicular to the borehole axis. The aforementioned authors assume solute transfer by
83
diffusion governed by Fick’s law only. A diffusion-advection solute transfer was developed
84
by Bader and Kooi (2005) that also accounts for membrane efficiency. The electrochemical
85
effect on the chemoelastic model was investigated in literature (Nguyen and Abousleiman,
86
2010; Tran and Abousleiman, 2013) and it is out of the scope of this work. Kanfar et al.
87
(2016a) investigated time-dependent wellbore stability by developing a porochemoelastic
88
model for anisotropic rocks that accounts for Fick’s diffusion and solute transfer by
89
advection. The aforementioned models do not consider the effect of the thermal gradient.
AC C
EP
TE D
M AN U
72
90
Due to the complexity of the problem, few authors have coupled the influence of
91
temperature gradient and chemical potential with Biot's poroelasticity (Ekbote and
92
Abousleiman, 2005; Ghassemi et al., 2009; Zhou and Ghassemi, 2009). Using the thermo-
93
chemo-hydro-mechanical (THMC) model, a numerical experiment that investigates CO2
ACCEPTED MANUSCRIPT sequestration by saturating it in water, and injecting the mixture into a carbonate aquifer was
95
conducted by Yin et al. (2011). Yasuhara et al. (2016) investigated the evolution of rock
96
permeability under THMC coupling. While these numerical and analytical models account
97
for the thermo-chemo-hydro-mechanical effects, none of which account for complete
98
material anisotropy to properly simulate case scenarios routinely encountered in field
99
applications. The effect of solute advection and thermal convection are also overlooked.
RI PT
94
In this work, a numerical model is developed to simulate fully coupled
101
porochemothermoelasticity for anisotropic formations. First, constitutive and transport
102
equations are developed. A numerical model using the finite element method (FEM) is built
103
using the governing equations. The GPS assumption is used for full three dimensional
104
analysis and stability of the numerical computation (Kanfar et al., 2015a). An inclined
105
borehole problem drilled in in shaley sand TI formation is mainly discussed in this work, but
106
the model can be applied to different facets of rock mechanics. The model is not exclusive to
107
shales or shaley sand, but can be generalized to other rock types as long as the material
108
properties are representative. It is assumed that the rock is homogeneous anisotropic and
109
fully-saturated with formation water. The model can be extended to other single phase liquid
110
fluids such as dead oil. Finally, time-dependent wellbore stability after drilling is analyzed
111
using the developed multi-physics anisotropic chemo-thermo-hydro-mechanical model.
112
2. Governing Equations for Porochemothermoelasticity
113
2.1. Constitutive Equations
EP
TE D
M AN U
SC
100
The constitutive equations for porochemothermoelasticity are governed by the stress-
115
strain relation, and the variation of the species content that comprise the solution of the
116
chemical system. The classical stress-strain relation is modified to account for thermal
117
expansion, and change in stresses due to the solute mass fraction perturbation. The variation
118
of fluid content per unit referential volume ( ) is also modified to incorporate indirect flows
119
of solute (
120
The constitutive equations are given by (Ekbote and Abousleiman, 2005; Ghassemi et al.,
121
2009):
AC C
114
) and diluent (
=
=
) by chemical osmosis, and fluid mixture thermal expansion.
+
− 1
+
+
∆
+ Ξ ∆
− Λ ∆
− Ω ∆
(1) (2)
ACCEPTED MANUSCRIPT 122
where
is the total stress,
is the total strain,
is the temperature,
is Biot’s modulus and
123
is the pore pressure. A repeated subscript refers to the Einstein summation. The elastic,
124
thermal and chemical coefficients presented in the constitutive equations (1) and (2) are
125
defined as follows:
̅
=$ −
=
=
# = +1 1 1 Ξ=
̅
(
.
(3)
3&' (
)*
0 0
/1 −
̅
̅
0
1 − #2
0-.
= #. + 1 − #2- / Λ = 3' + # /
4(
̅
̅
0
̅ !
TE D
4( Ω = 3 ! + #. + 1 − #2- / 3! =
3' = '
+1
!
'
−
'
!
(10) (11) (12) (13)
where
127
coefficient (chemoelastic coupling term),
128
universal gas constant,
129
temperature and 5 is the rock porosity. It is assumed that the swelling parameters for the
'
is the fluid density,
(
126
EP
is Biot’s effective stress coefficient vector,
(6)
(9)
0
25
(5)
(8)
0
̅
(4)
(7)
M AN U
=
∙#
̅ !
RI PT
−
SC
=
is the swelling
is the molar mass of the solute, R is the
is the solid thermal expansion coefficient vector,
!
is the fluid
131
solute (
132
value and the transpose of the matrix, respectively. The drained elastic tensor ( ) and the
133
compliance tensor ( ) in the directions of the orthotropic material are defined as:
AC C
130
thermal expansion coefficient, 4( is the fluid specific entropy measured at the average system ) and diluent (
) are equal. The overbar and the superscript
denote the average
ACCEPTED MANUSCRIPT
134 135
where ;, ? and B are the Young’s modulus, Poisson’s ratio and shear modulus, respectively,
and the subscripts refer to the coordinate systems described in Fig. 1. A detailed explanation
136
on the parameters for anisotropic poroelasticity is discussed by Cheng (1997). The strain can
137
be related to the solid displacement (F) as follows:
1 = HF , + F , J 2
0 0 0 0 1⁄B
0 E 0 D 0 D 0 D D 0 D 1⁄B<@= C
67
(14)
RI PT
0 0 0 1⁄B@A= 0 0
SC
−?A<= ⁄;A= −?A@= ⁄;A= 1⁄;A= 0 0 0
67
1⁄;<= : 9 −?@<= ⁄;@= 9−?A<= ⁄;A= =9 0 9 0 9 0 8
(15)
where a subscript followed by a comma denotes the spatial differentiation.
M AN U
138
−?@<= ⁄;@= 1⁄;@= −?A@= ⁄;A= 0 0 0
=
139
It is important to note that all material property tensors and stress tensor are rotated to the
140
borehole coordinate systems. This is achieved by defining three coordinate systems that are
141
referenced to the North-East-Vertical (NEV) global coordinates: 1. The far-field stresses
143 144
coordinates 1K' , L' , M' 2 that are controlled by angles N' and O' , 2. The intrinsic rock
1KP , LP , MP 2 that are controlled by angles NP and OP and 3. The
borehole coordinates 1KQ , LQ , MQ 2 that are controlled by angles NQ and OQ . Figure 1
properties coordinates
TE D
142
147
borehole 1R2 are referenced to the bottom of the hole (BOH), which always coincides with
148
the KQ axis. A schematic of the finite element method mesh is shown in Fig. 2, where the red lines indicate the cross-sectional plane of the FEM mesh. Property rotation is discussed
149
thoroughly in literature (Aadnoy and Chenevert, 1987; Amadei, 1983; Lekhnitskii, 1981).
EP
146
summarizes the coordinate systems and the associated rotation angles. Angles around the
AC C
145
150
The model formulation is simplified for a binary solution consisting of a single solute
151
(CS) and a diluent (CD) or solvent. It can be easily modified to include multiple solute species
152
as follows:
=S T7
153
= 1−
where k is the number of solute species.
(16)
ACCEPTED MANUSCRIPT 2.2. Transport Equations
155 156
A general conduction law or transport equation (Onsager’s relation) can be written as (Groot, 1951):
U = V K,
(17)
158
where U is the flux, V
159
law, chemical osmosis and thermal osmosis is given by:
161
162
gradient. The transport equation for fluid flux due to hydraulic potential governed by Darcy’s U = − ̅! WX !
,
− V X
− Y.
,
,
Z
(18)
where Y . is the thermal osmosis coefficient tensor, X is mobility tensor and V is the osmotic
pressure and are given by:
X =
V =
Y [
SC
160
is the phenomenological coefficient tensor and K, is the potential
) * !̅ ℜ ̅ ̅
M AN U
157
RI PT
154
(19) (20)
where Y is the rock permeability tensor, [ is the fluid viscosity measured at in-situ conditions
164
and ℜ is the reflection or membrane efficiency coefficient. The reflection coefficient can be
165
Kooi, 2005). Similarly, the solute transport equation due to solute diffusion governed by a
166
modified Fick’s law, thermal filtration and fluid advection is given by:
169
170
TE D
168
U ' = − ̅! 11 − ℜ2W
,
+
.
,
! ̅ −U .
Z
(21)
Fick’s law is modified by introducing the multiplier 11 − ℜ2 to account for the membrane
where
is the solute diffusion coefficient tensor and
EP
167
defined as the ratio of the hydraulic gradient and the osmotic pressure gradient (Bader and
is the thermal filtration coefficient.
efficiency (Bader and Kooi, 2005). Finally, the heat transfer equation is given by: U] = −^Q
,
+ ̅! ! U
where ^Q is the thermal conductivity and
AC C
163
!
!
(22)
is the fluid specific heat. The first term in the
171
heat transport equation accounts for bulk conduction by Fourier’s law, and the second for
172
heat transfer by fluid convection. The thermal conductivity (^Q ) and bulk heat capacity
173
(
Q Q)
are defined as:
and
^Q = 5^! + 11 − 52^'
Q Q
= 5 !̅
!
+ 11 − 52
' '
(23)
are the density and specific heat, respectively. The subscripts _, 4 and ` refer
174
where
175
to the bulk, solid matrix and fluid, respectively.
ACCEPTED MANUSCRIPT
177 178
179
2.3. Conservation Laws The conservation laws presented in this section ignore all body forces such as gravity and inertial forces. The momentum balance equation yields the equilibrium equation: ,
=0
(24)
A mass balance for the fluid and solute are given by: 1a ⁄ab2 + U !, = 0
5c1a
⁄ab2 + U ', = 0
RI PT
176
(25) (26)
181
where c is the solute retardation coefficient, which depends on the rock susceptibility to
182
assigned to this coefficient to neglect its effect on the analysis. Finally, the energy balance
183
equation is given by: Q Q
184
where
185
2.4. Field Equations
Q Q 1a
is the bulk heat capacity.
SC
absorb the chemical species and cation exchange (Bader and Kooi, 2005). A value of 1 can be
⁄ab2 + U], = 0
M AN U
180
(27)
The presented constitutive and transport equations are substituted into the conservation
187
laws to yield the field equations. The Navier-type equation for displacement in the x-y plane,
188
using the generalized plane strain assumption, is obtained by substituting the stress-strain
189
relation into the momentum balance equation as follows:
1 W1 2
77
−
7d 2F ,
TE D
186
+1
77
+
7d 2F ,
Z−
,
+
,
−Λ
,
=0
(28)
Incorporating the constitutive equation for the variation of fluid content, and the fluid
191
transport equation into the mass balance yields the local continuity equation:
1
+
AC C
192
1∇Ff 2 +
EP
190
5c f + 11 − ℜ2 1U '
f + Ξ f − Ω f − X∇d + V ∇d
+ Y . ∇d = 0
̅ ∇d = 0
(29)
Similarly, the field equation for the solute mass fraction is given by: 2, − 11 − ℜ2
− 11 − ℜ2
∇d
.
(30)
193
Substituting the heat flux equation into the energy balance relation yields the field equation
194
for the temperature:
195 196 197
Q Q
f − ^Q ∇d + ̅!
! HU
!
J, = 0
(31)
It is important to note the nonlinearity in equations (30) and (31) due to their dependence on
the fluid flux term 1U ! 2, which is responsible for the solute transfer by advection and heat
flux by convection, respectively. In this work, a homogenous anisotropic rock is assumed.
ACCEPTED MANUSCRIPT However, heterogeneity in rock properties such as permeability and porosity can be easily
199
incorporated into the model by varying them in each element of the FEM mesh. All rock
200
pores are assumed to be interconnected and fully saturated. In addition, no mass loss from the
201
rock matrix is allowed by the model by means such as sanding. It is also assumed that the
202
effluent from the borehole is fully miscible with the reservoir fluid.
203
3. Finite Element Formulae
RI PT
198
This section presents the anisotropic porochemothermoelastic formulae using the finite
205
element method. The generalized plane strain assumption will be used to allow for 3D
206
analysis and incorporation of material anisotropy (Amadei, 1983; Lekhnitskii, 1981). This
207
pseudo-3D analysis utilizes a 2D mesh, which is computationally far less intensive and
208
consumes significantly lower memory (Kaliakin, 2002). The classical plane strain assumption
209
only allows a 2D stress tensor, where the axial stress is always perpendicular to the mesh
210
plane and no anti-plane shear stresses can be applied. Moreover, the classical plane strain
211
assumption does not permit the use of material anisotropy. The GPS assumption and how
212
material anisotropy is incorporated was discussed thoroughly in a recent publication by the
213
authors (Kanfar et al., 2016b, 2015a). Quadratic eight node quadrilateral elements are used in
214
the examples presented in this work, and a simplified coarse prototype is shown in Fig. 3. An
215
inclined borehole problem drilled in a transversely isotropic shaley sand that is fully-saturated
216
with water will be mainly analyzed in this work. The unknown variables of the FEM system,
218
M AN U
TE D
F, ,
and , are given by:
F = gh Fi , = gj i,
= gk l m , = g. n
(32)
where gh , gj , gk l and g. are the shape functions of solid displacement, pore pressure,
EP
217
SC
204
solute mass fraction and temperature, respectively, and the tilde denotes the nodal values.
220
Using the Galerkin weighted residuals method for spatial discretization and the implicit finite
221
difference method for temporal discretization, the finite element formulation of the field
222
equations is given by:
AC C
219
oh Fif + oj if + ok m f + o. nf = `f
o.j Fif + ℙj if + ℙjq i + ℙk m f + ℙkq m + ℙ. nf + ℙ.q n = 0 ℂk m f + ℂkq m + ℂ.q n = 0 s . nf + s .q n = 0
(33)
ACCEPTED MANUSCRIPT
225
oh : . 9oj 9 0 9 8 0
oj
−ok
o.
in the appendix. The incremental form of the system of equations is given by: −Hℙj + ∆b ℙjq J −1ℙk − ∆b ℙkq 2 0 −1ℂk + ∆b ℂkq 2 0 0
E ∆Fi 1ℙ. − ∆b ℙ.q 2 D ∆ i D t∆ m u −∆b ℂ.q D −1s . + ∆b s .q 2C ∆ n
∆` : E m n 9∆b ℙjq i1vw 672 − ∆b ℙkq 1vw 672 + ∆b ℙ.q 1vw 672 D =9 D ∆b ℂkq m 1vw 672 + ∆b ℂ.q n1vw 672 9 D ∆b s .q n1v 672 8 C w
RI PT
224
The matrices oh , oj , ok , o. , ℙj , ℙjq , ℙk , ℙkq , ℙ. , ℙ.q , ℂk , ℂkq , ℂ.q , s . and s .q are defined
(34)
where ∆` is the change of the external traction forces, ∆b = bx − bx67 , bx and bx67 indicate
SC
223
the current and the previous time steps, respectively. First, the temperature equation is solved
227
separately. Then, the displacement, pore pressure and solute mass fraction are solved
228
simultaneously. However, the system of equation will then become nonsymmetrical. The
229
and s .q contain the fluid discharge term which drives the advection and convection
M AN U
226
230
symmetry is restored by employing a partitioned solution (Schrefler, 1985). The matrices ℂkq
231
contributions, respectively. An iterative scheme is developed to handle the nonlinearity in
232
equation (34). The solution converges when the following criteria are met: y∆ ivw − ∆ i1vw 672 y y∆ i1vw 672 y
< {,
y∆ mvw − ∆ m1vw 672 y y∆ nvw − ∆ n1vw 672 y < {, <{ y∆ n1v 672 y y∆ m1v 672 y
TE D
233
w
w
(35)
235
where { is the maximum error allowed by the iterative solver. In this work, the maximum
236
Along the borehole boundary, a constant pressure, solute concentration and temperature
237
are applied. Similarly, the in-situ far-field stresses are applied to the reservoir outer boundary
238
as traction forces. The FEM program boundary conditions (BC) are summarized in Table 1. It
239
is assumed throughout this paper that the borehole wall is permeable. An impermeable
240
borehole wall boundary condition can be easily implemented for cases where a “perfect”
241
filter cake is developed. In such cases, the pore pressure and solute concentration are not
242
influenced by the borehole boundary conditions (Fjaer, 2008). Moreover, the wellbore BC
243
will only affect the traction force acting on the borehole wall and heat transfer by conduction.
244
An impermeable BC was previously discussed in literature (Cui et al., 1998; Fjaer, 2008) and
245
it is out of the scope of the paper, though it can be easily implemented in the FEM code.
EP
error is set to 0.5%.
AC C
234
ACCEPTED MANUSCRIPT 246
4. Verification of Numerical Model In this section, the anisotropic porochemothermoelastic numerical model is validated
248
using published analytical solutions. The GPS assumption, anisotropic elasticity and
249
poroelasticity models were benchmarked in a previous publication by the authors (Kanfar et
250
al., 2015a). The modified Terzaghi or Biot’s effective stress concept (Nur and Byerlee, 1971)
251
is used throughout, and given by:
252
where
=
is the effective stress vector and
−
is the pore pressure.
RI PT
247
(36)
The porochemothermoelasticity is verified for a case where an inclined borehole is drilled
254
in a transversely isotropic formation. The analytical solution assumes that the plane of elastic
255
isotropy (x-y plane) is always perpendicular to the borehole axis (Ekbote and Abousleiman,
256
2005). This assumption clearly alienates most cases encountered in everyday drilling
257
operations, but will be enforced on the numerical model for the purpose of validating against
258
the aforementioned analytical solution. This type of anisotropy will be referred to as
259
“conditional anisotropy” from this point forward. For the purpose of this work, we define an
260
anisotropy ratio as
261
laboratory measurements are required to properly describe the anisotropic medium for real
262
field applications.
263
transformations parameters, far-field stresses, rock and fluid parameters are summarized in
264
Tables 2-4. The input parameters are representative of a low permeability shaley-sand in the
265
Middle East with structured clay that is fully-saturated with formation water, chemically-
266
active and has a membrane-like behavior, where osmotic flow is observed. The thermal
267
parameters
268
porochemothermoelastic model is solved for 0.001 day, 0.01 day and 0.1 day time steps. The
269
results of the pore pressure, solute mass fraction, temperature and total stresses are presented
270
in Figs. 4 and 5. The analytical and numerical solutions are represented by solid lines and
271
markers, respectively. The numerical FEM model demonstrates close agreement with the
272
analytical solution.
273
5. Sensitivity Analysis
|~=
=
•}€=
•}~=
=
‚ •} = ‚ •~ =
to estimate the properties in the MP -direction. Proper
EP
TE D
In this example, the anisotropy ratio is set to 2.0. The coordinate
slightly
modified
from
Ghassemi
and
Diek
(2002).
The
AC C
are
|}=
M AN U
SC
253
274
Several sensitivity analyses are conducted in this section on the porochemothermoelastic
275
model, which are used to emphasize the implications of different aspects of the model. First,
276
the combined effects of elastic and thermic anisotropy are discussed. The stresses are
ACCEPTED MANUSCRIPT 277
calculated for heating and cooling cases using different solute concentrations. The influence
278
of thermal convection and solute advection are accentuated with additional examples. Finally,
279
pore pressure and effective stresses are calculated using different constitutive models, and are
280
juxtaposed to illustrate the individual and overall combined effects of each model.
281
5.1. Effect of Rock Anisotropy The rock anisotropy only affects the elastic and thermic parameters. The chemical
283
parameters are mainly properties of the exchanged fluids, and the anisotropy only affects the
284
coupling chemoelastic coefficients. In the following example, the same input data are used,
285
and the anisotropy ratio is varied to 1.0 (isotropic), 1.5 and 2.0. The effective tangential and
286
axial stresses are calculated using the “conditional” and “complete” anisotropy cases, where
287
the plane of isotropy of the TI material is not necessarily perpendicular to the borehole axis.
288
The radial distributions of effective tangential and axial stresses are summarized in Fig. 6,
289
where the solid and dashed lines represent the conditional and complete anisotropy,
290
respectively. It is shown that the degree of anisotropy has a major influence on the effective
291
stresses around and away from the wellbore region. A noticeable divergence is also observed
292
between the results obtained using the conditional and complete anisotropy cases. Figure 7
293
closely examines the effective axial stress around the borehole wall. The isotropic results are
294
mainly tensile (negative stress). Moreover, the conditional anisotropy assumption
295
consistently overestimates the actual stresses when compared to the complete anisotropy
296
results. It is important to note that these observations should not be generalized, and should
297
be examined on a case-by-case basis. Consequently, assuming isotropy or conditional
298
anisotropy could have detrimental effects on wellbore stability by underestimating borehole
299
collapse pressure, or underestimating the fracture initiation pressure. The effect of anisotropy
300
on pore pressure is not presented in this example, as the effect was trivial.
301
5.2. Effect of Thermal Gradient and Solute Concentration
AC C
EP
TE D
M AN U
SC
RI PT
282
302
The effect of heating and cooling is explored in this subsection by varying the solute
303
mass fraction of the wellbore and the reservoir fluids. This will accentuate individual and
304
combined effects of the porochemothermoelastic model. The temperature differential
306
(∆ =
307
These values are reversed or equalized for the respective cases. The sensitivity analysis for
308
pore pressure, solute concentration, total radial and tangential stresses are presented in Figs.
305
ƒ
−
!)
is set to -50 and 50°C for the cooling and heating scenarios, respectively. In
addition, the solute mass fraction is set to 0.1 and 0.2 for the rock (
!
) and the borehole (
ƒ ).
ACCEPTED MANUSCRIPT 309
8-10. The solid lines represent the heating cases, and the dashed lines represent the cooling
310
ones. The sensitivity analysis for pore pressure, only, is done using two different
312
permeabilities, 1 mD and 1x10-5 mD. In Fig. 8a, the lower permeability makes the pore
313
pressure largely dominated by temperature. Heating the formation will result in thermal
314
expansion, which will subsequently increase the pore pressure. The inverse case, wherein
315
shrinkage occurs, will results in lowering the pore pressure and total stress. The lower
316
permeability allows the fluid exposed to the transient temperature region to be volumetrically
317
enhanced before it diffuses. On the other hand, Fig. 8b, where high permeability is used,
318
illustrates that pore pressure is heavily dominated by the chemical osmosis process. This is
319
evident when the effect of chemical osmotic flow is neglected (
320
becomes mainly a porothermoelastic one. The effect of temperature on pore pressure is
321
minute when high intrinsic permeability is present. A higher solute concentration in the
322
wellbore creates an osmotic backflow that lowers the pore pressure significantly in the area
323
around the borehole.
RI PT
311
=
ƒ)
and the problem
M AN U
SC
!
The total radial and tangential stresses summarized in Figs. 9 and 10. Heating shows a
325
significant increasing effect on the total stresses. On the other hand, a higher solute
326
concentration in the borehole has a lowering effect on the total stresses in the region around
327
the borehole. Injecting a cooler fluid with a higher solute concentration creates a considerable
328
stabilizing effect on the near wellbore region by lowering the pore pressure and stresses. This
329
effect can be exploited in wellbore stability or hydraulic fracturing problems.
330
5.3. Effect of Heat Transfer by Convection and Solute Transfer by Advection
EP
TE D
324
Most conventional subsurface rocks such as sandstone and carbonate reservoirs, exhibit a
332
high intrinsic permeability in the range of 1 mD to tens of Darcys. In which case, the heat
333
transfer by convection and solute transport by advection carried by the fluid plays a
334
significant role and should not be overlooked. In the following example, the effect of heat
335
convection and solute transport by advection are investigated. The process of estimating the
336
convection and advection contributions is a non-linear one as discussed in the Finite
337
Element Formulae section. The results of the conductive-advective model, which will be
338
referred to as “porochemothermoelastic C.A.,” are contrasted against the linear
339
porochemothermoelastic model in Figs. 11 and 12.
AC C
331
340
The pore pressure, solute mass fraction and temperature radial distributions shown in Fig.
341
11 illustrate that the linear porochemothermoelastic results significantly overestimates the
ACCEPTED MANUSCRIPT reservoir pore pressure, solute mass fraction and temperature. At longer time steps, the heat
343
transfer by convection and solute transport by advection eclipse the contributions by
344
conduction and diffusion as shown by the large separation between the two models. The
345
added solute contributed by advection increases the counteracting chemical osmotic
346
backflow. Subsequently, the incremental change in pore pressure translates into the effective
347
stresses.
RI PT
342
The total radial, tangential and axial stresses are summarized in Fig. 12. The linear
349
porochemothermoelastic model, shown in solid lines, consistently overestimates the stresses.
350
This is attributed to the underestimation of the cooling effect carried by fluid convection,
351
which consequently decreases the thermal stresses. In addition, the overestimation of solute
352
mass increases the effect of swelling, and consequently increases the total stresses. The effect
353
of advection and convection can be neglected if the borehole wall is impermeable, where the
354
borehole BC only affects the traction force and heat transfer by conduction. Advection and
355
convection problems are computationally intensive due to the non-linearity of the solution.
356
Kanfar et al. (2016a, 2016b) concluded that advection and convection contributions can be
357
neglected when the rock permeability is lower than 1x10−4 mD and 1x10−1 mD, respectively.
358
5.4. Effect of Different Constitutive Models
M AN U
SC
348
In this section, a comparison between the different constitutive models is investigated.
360
The poroelastic, porothermoelastic, porochemoelastic, porochemothermoelastic and the
361
convective-advective (C.A.) porochemothermoelastic models are used to calculate the pore
362
pressure, effective radial, tangential and axial stresses presented in Figs. 13-16. The same set
363
of inputs is used with the exception of rock permeability, which is changed to 1x10-4 mD to
364
magnify the effect of each model and avoid permeability bias as previously discussed.
EP
TE D
359
It was previously demonstrated that the thermal effect on pore pressure diminishes in
366
high permeability rocks. This can be attributed to the short time of exposure of the pore fluid
367
for it to thermally expand or shrink before it mobilizes outside of the radius of the thermally
368
enhanced area. In contrast, Fig. 13 presents the results for pore pressure away from the
369
borehole in a low permeability rock. Caused by cooling, the pore pressure estimated by the
370
porothermoelastic model shows significant divergence from the isothermal poroelastic model.
371
Similarly, the pore pressure obtained by the porochemothermoelastic model diverges from
372
the porochemoelastic model caused by the temperature enhancement which lowers the pore
373
pressure and the solute concentration.
AC C
365
ACCEPTED MANUSCRIPT The enhancement of the different models on the effective radial stresses in Fig. 14 is less
375
pronounced when compared to the tangential and axial stresses in Figs. 15 and 16,
376
respectively. It is clear that there is virtually no added effect by the conductive-advective
377
model when compared to the linear porochemothermoelastic model. The ultra-low
378
permeability makes solute concentration diffusion dominated, and the heat transfer mainly a
379
conduction problem. Consequently, the convection and advection contribution can be
380
neglected at lower permeabilities.
381
6. Wellbore Stability
RI PT
374
In this section, wellbore stability is analyzed using different constitutive models. The
383
chemical and thermal effects are mainly discussed. The effect of elastic anisotropy and rock
384
strength anisotropy on time-dependent wellbore stability was discussed previously by the
385
authors (Kanfar et al., 2015a). An improved mud window estimation for inclined boreholes
386
drilled in highly-stressed anisotropic formations to prevent shear and tensile failures using a
387
semi-analytical model was developed by the authors (Kanfar et al., 2015b).
M AN U
SC
382
Figure 17 summarizes the pore pressure, effective radial and tangential stress
389
distributions around the borehole using different constitutive models after 0.1 day from
390
drilling. The input data used in this section are summarized in Tables 1-4. Borehole shear
391
failure is assessed using the Mohr-Coulomb criterion (Jaeger et al., 2009). A linearly elastic-
392
brittle model is assumed, where shear failure occurs when the compressive strength of the
393
rock is exceeded. It is important to note that the failure analysis does not use a remeshing
394
technique or post-failure analysis between time steps to account for the failed elements. The
395
analysis provides a practical tool to prognosticate whether the borehole will stabilize or
396
worsen as time progresses. The strength parameters used for the Mohr-Coulomb criterion are
397
4.14 MPa and 20 degrees for the cohesive strength and angle of internal friction, respectively.
398
The corresponding borehole failure is presented in Fig. 18, where the failure zone is shown in
399
blue. If we consider the results of the poroelastic model (Fig. 18a) as a base case, the results
400
of the porothermoelastic model (Fig. 18b) and the porochemoelastic model (Fig. 18c) show a
401
more stable wellbore. This stabilizing effect is caused by temperature cooling and the higher
402
solute mass injected, respectively. The combined effects of the porochemothermoelastic
403
model (Fig. 18d) show a significant added improvement to the stability of the wellbore when
404
juxtaposed against the poroelastic model. As shown in Fig. 17d, the effective tangential stress
405
is considerably lower for the porochemothermoelastic model, which creates a stable
AC C
EP
TE D
388
ACCEPTED MANUSCRIPT 406
environment in the near wellbore region. Similar concepts can be used during hydraulic
407
fracture operations to create a more favorable environment to induce a tensile failure.
408
7. Conclusions In this work, constitutive and transport equations for fully coupled chemo-thermo-hydro-
410
mechanical processes are developed for homogenous anisotropic formations. A numerical
411
model using the FEM is developed to simulate the drilling of an inclined borehole in
412
transversely isotopic shaley sand. It was assumed that the formation pores are fully-saturated
413
with water, and the rock has a membrane-like characteristics, where chemical osmosis is
414
applicable. A novel pseudo 3D analysis using the generalized plane strain concept is used for
415
more efficient and stable computing without compromising the accuracy of the results.
417
•
418 419
SC
Several conclusions are drawn from the work presented in this paper as follows:
M AN U
416
RI PT
409
Anisotropy should be properly characterized by acquiring core data in the formation of interest to assess the degree of elastic anisotropy.
•
The presented examples demonstrated that assuming isotropy or conditional
420
anisotropy could potentially lead to large errors in the estimated stresses if high
421
degree of elastic anisotropy is present, especially in the region around borehole. •
It is assumed that the rock is fully saturated with formation brine, but the model
TE D
422 423
can be generalized for live oil or gas-filled formations for qualitative estimates of
424
optimum borehole pressure selection to avoid borehole collapse. •
426
428 429 430 431
model inputs are representative of the rock in question. •
Coupling heat flow is especially relevant in deep wells, where there is a significant temperature differential between the drilling fluid and the original
AC C
427
The model is inclusive of all rock types (clastics or carbonates) as long as the
EP
425
geothermal gradient of the formation.
•
The presented examples demonstrated that heating the formation increases the
pore pressure and the total stresses, and the opposite is observed when the
432
formation is cooled. Assuming an isothermal model will result in significant
433
errors in the estimated stresses.
434
•
Coupling solute transport, chemical potential and formation swelling should be
435
carefully investigated as the model is not always applicable. The model should be
436
used when the formations exhibits a membrane like behavior where the above
437
mentioned elements are relevant.
ACCEPTED MANUSCRIPT 438
•
It is observed from the presented examples, a higher solute concentration in the borehole fluid creates a higher chemical potential in the reservoir fluid. This
440
results in a reverse osmosis and subsequently a drop in pore pressure.
441
Concurrently, the effective radial stress increased, and the effective tangential
442
stress decreased. This has a stabilizing effect on the borehole wall, which is in
443
line with the drilling industry common practices.
444
•
RI PT
439
Advection and convection are computationally intensive and should only be used when the permeability thresholds are reached (1x10−4 mD and 1x10−1 mD,
446
respectively).Injecting a cooler fluid with a higher salinity than the intrinsic
447
formation fluid adds a significant stabilizing effect on the near wellbore region by
448
lowering the effective tangential stress.
SC
445
The combined effects of the thermo-chemo-hydro-mechanical model produce different
450
results from the individual thermoelastic or chemoelastic models. The different elements of
451
the anisotropic porochemothermoelastic model are interwoven with one another, and each
452
component should not be considered in isolation. In this work, the model was applied for an
453
inclined borehole problem, but can be generalized to a wide range of applications.
454
References
455
Aadnoy, B., Chenevert, M., 1987. Stability of highly inclined boreholes. SPE Drill. Eng.
456 457
Abousleiman, Y., Ekbote, S., 2005. Solutions for the Inclined Borehole in a Porothermoelastic Transversely Isotropic Medium. J. Appl. Mech. 72, 102–114.
458 459 460
Abousleiman, Y., Roegiers, J., Cui, L., Cheng, A., 1995. Poroelastic solution of an inclined borehole in a transversely isotropic medium, in: Rock Mechanics: Proceedings of the 35th US Symposium on Rock Mechanics. Taylor & Francis Group.
461 462
Amadei, B., 1983. Rock anisotropy and the theory of stress measurements. Springer-Verlag, Berlin; New York.
463 464
Bader, S., Kooi, H., 2005. Modelling of solute and water transport in semi-permeable clay membranes: comparison with experiments. Adv. Water Resour. 28, 203–214.
465 466
Batugin, S.A., Nirenburg, R.K., 1972. Approximate relation between the elastic constants of anisotropic rocks and the anisotropy parameters. Sov. Min. Sci. Sov. Min. Sci. 8, 5–9.
467 468
Bear, J., Corapcioglu, M.Y., 1981. A mathematical model for consolidation in a thermoelastic aquifer due to hot water injection or pumping. Water Resour. Res. 17, 723–736.
469 470
Biot, M., 1955. Theory of Elasticity and Consolidation for a Porous Anisotropic Solid. J. Appl. Phys. 26, 182–185.
471 472
Cheng, A., 1997. Material coefficients of anisotropic poroelasticity. Int. J. Rock Mech. Min. Sci. 34, 199–205.
AC C
EP
TE D
M AN U
449
ACCEPTED MANUSCRIPT Coussy, O., 1989. A general theory of thermoporoelastoplasticity for saturated porous materials. Transp. Porous Media 4, 281–293.
475 476 477
Cui, L., Ekbote, S., Abousleiman, Y., Zaman, M., Roegiers, J., 1998. Borehole stability analyses in fluid saturated formations with impermeable walls. Int. J. Rock Mech. Min. Sci. 35, 582–583.
478 479
Detournay, E., Cheng, A., 1988. Poroelastic response of a borehole in a non-hydrostatic stress field. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 25, 171–182.
480 481
Ekbote, S., Abousleiman, Y., 2005. Porochemothermoelastic Solution for an Inclined Borehole in a Transversely Isotropic Formation. J. Eng. Mech. 131, 522–533.
482 483
Ekbote, S., Abousleiman, Y., 2006. Porochemoelastic Solution for an Inclined Borehole in a Transversely Isotropic Formation. J. Eng. Mech. 132, 754–763.
484
Fjaer, E., 2008. Petroleum related rock mechanics. Elsevier, Amsterdam; Boston.
485 486
Ghassemi, A., Diek, A., 2002. Porothermoelasticity for swelling shales. J. Pet. Sci. Eng. 34, 123–135.
487 488
Ghassemi, A., Diek, A., 2003. Linear chemo-poroelasticity for swelling shales: theory and application. J. Pet. Sci. Eng. 38, 199–212.
489 490 491
Ghassemi, A., Tao, Q., Diek, A., 2009. Influence of coupled chemo-poro-thermoelastic processes on pore pressure and stress distributions around a wellbore in swelling shale. J. Pet. Sci. Eng. 67, 57–64.
492 493
Ghassemi, A., Wolfe, G., 1999. A chemo-mechanical model for borehole stability analyses, in: USRMS. American Rock Mechanics Association, pp. 239–246.
494 495
Groot, S.R. de, 1951. Thermodynamics of Irreversible Processes. North-holland Pub. Co.; [sole distributors for U.S.A.: Interscience Publishers, New York], Amsterdam.
496 497
Jaeger, J., Cook, N., Zimmerman, R., 2009. Fundamentals of rock mechanics. Chapman and Hall ; Wiley, London; New York.
498 499
Kaliakin, V.N., 2002. Introduction to approximate solution techniques, numerical modeling, and finite element methods. Marcel Dekker, New York.
500 501
Kanfar, M.F., Chen, Z., Rahman, S., 2016a. Anisotropic Diffusive-Advective Porochemoelasticity Modeling for Inclined Boreholes. Int. J. Geomech. 6016025.
502 503
Kanfar, M.F., Chen, Z., Rahman, S.S., 2015a. Effect of material anisotropy on timedependent wellbore stability. Int. J. Rock Mech. Min. Sci. 78, 36–45.
504 505
Kanfar, M.F., Chen, Z., Rahman, S.S., 2015b. Risk-controlled wellbore stability analysis in anisotropic Formations. J. Pet. Sci. Eng. 134, 214–222.
506 507 508
Kanfar, M.F., Chen, Z., Rahman, S.S., 2016b. Fully coupled 3D anisotropic conductiveconvective porothermoelasticity modeling for inclined boreholes. Geothermics 61, 135– 148.
509 510
Kurashige, M., 1989. A thermoelastic theory of fluid-filled porous materials. Int. J. Solids Struct. 25, 1039–1052.
511
Lekhnitskii, S.G., 1981. Theory of elasticity of an anisotropic body. Mir Publishers, Moscow.
512
McLean, M.R., Addis, M.A., 1990. Wellbore Stability: The Effect of Strength Criteria on
AC C
EP
TE D
M AN U
SC
RI PT
473 474
ACCEPTED MANUSCRIPT Mud Weight Recommendations, in: SPE Annual Technical Conference and Exhibition. Society of Petroleum Engineers.
515 516
Mody, F.K., Hale, A., 1993. Borehole-stability model to couple the mechanics and chemistry of drilling-fluid/shale interactions. J. Pet. Technol. 45.
517 518
Nguyen, V.X., Abousleiman, Y.N., 2010. Incorporating electrokinetic effects in the porochemoelastic inclined wellbore formulation and solution. An. Acad. Bras. Cienc.
519 520
Nur, A., Byerlee, J.D., 1971. An exact effective stress law for elastic deformation of rock with fluids. J. Geophys. Res. 76, 6414–6419.
521
Ong, S.H., 1994. Borehole Stability, Ph.D. Thesis. The University of Oklahoma.
522 523
Schiffman, R., 1971. A thermoelastic theory of consolidation. Environ. Geophys. Heat Transf. 78–84.
524 525
Schrefler, B., 1985. A partitioned solution procedure for geothermal reservoir analysis. Commun. Appl. Numer. methods 1, 53–56.
526 527
Sherwood, J.D., 1993. Biot Poroelasticity of a Chemically Active Shale. procmathphysscie Proc. Math. Phys. Sci. 440, 365–377.
528 529
Tran, M.H., Abousleiman, Y., 2013. Anisotropic Porochemoelectroelastic Solution for an Inclined Wellbore Drilled in Shale. J. Appl. Mech. 80, 20912.
530 531 532
Vishal, V., Jain, N., Singh, T.N., 2015. Three dimensional modelling of propagation of hydraulic fractures in shale at different injection pressures. Sustain. Environ. Res. 25, 217–225.
533 534
Vishal, V., Ranjith, P.G., Singh, T.N., 2013. CO2 permeability of Indian bituminous coals: Implications for carbon sequestration. Int. J. Coal Geol. 105, 36–47.
535 536
Wang, Y., Dusseault, M.B., 2003. A coupled conductive convective thermo poroelastic solution and implications for wellbore stability. J. Pet. Sci. Eng. 38, 187–198.
537 538 539
Yasuhara, H., Kinoshita, N., Ogata, S., Cheon, D.-S., Kishida, K., 2016. Coupled thermohydro-mechanical-chemical modeling by incorporating pressure solution for estimating the evolution of rock permeability. Int. J. Rock Mech. Min. Sci. 86, 104–114.
540 541
Yin, S., Dusseault, M.B., Rothenburg, L., 2011. Coupled THMC modeling of CO2 injection by finite element methods. J. Pet. Sci. Eng. 80, 53–60.
542 543 544
Zhang, D., Ranjith, P.G., Perera, M.S.A., 2016. The brittleness indices used in rock mechanics and their application in shale hydraulic fracturing: A review. J. Pet. Sci. Eng. 143, 158–170.
545 546
Zhou, X., Ghassemi, A., 2009. Finite element analysis of coupled effects around a wellbore in swelling shale. Int. J. Rock Mech. Min. Sci. 46, 769–778.
547 548
AC C
EP
TE D
M AN U
SC
RI PT
513 514
ACCEPTED MANUSCRIPT 549
Appendix
550
The finite element matrices are defined as follows: oh = „ … . … †Ω ‡
oj = „ … . ‡
(A.1)
gj †Ω
(A.2)
RI PT
ok = „ … . gk l †Ω
(A.3)
‡
o. = „ … . Λ g. †Ω ℙj = „ gj. ‡
1
(A.4)
+
gj †Ω
SC
‡
‡
.
M AN U
ℙjq = „ H∇gj J X H∇gj J †Ω ℙk = „ gj. Ξ gk l †Ω ‡
ℙkq = „ H∇gj J V H∇gk l J †Ω ℙ. = „ gj. Ω g. †Ω ‡
TE D
‡
.
‡
‡
ℂkq = „ H∇gk l J 11 − ℜ2
AC C
.
‡
ℂ.q = „ H∇gk l J 11 − ℜ2 ‡
s . = „ g.. ‡
551
.
Q Q g.
†Ω
1∇g. 2 †Ω
s .q = „ 1∇g. 2. ^Q 1∇g. 2 †Ω − 1∇g. 2. !̅ ‡
(A.8)
(A.11)
H∇gk l J †Ω − „ H∇gk l J 11 − ℜ2 U ! Hgk l J †Ω .
(A.7)
(A.10)
EP
ℂk = „ gk. l 5c gk l †Ω
(A.6)
(A.9)
ℙ.q = „ H∇gj J Y . 1∇g. 2 †Ω .
(A.5)
.
‡
(A.13) (A.13) (A.14)
! U
!
1g. 2 †Ω
(A.15)
ACCEPTED MANUSCRIPT Fig. 1 Coordinate transformation systems (Kanfar et al., 2016). Fig. 2 The green shapes represent the well trajectory, and the red shapes represent the FEM mesh. The formation beddings are shown in blue (Kanfar et al., 2016). Fig. 3 Eight-node quadrilateral FEM mesh (Kanfar et al., 2016).
RI PT
Fig. 4 Comparison between numerical and analytical results for pore pressure, solute mass fraction and temperature at different time steps. Fig. 5 Comparison between numerical and analytical results for total radial, tangential and axial stresses at different time steps.
SC
Fig. 6 Effect of elastic and thermal anisotropy on effective tangential and axial stresses with radial distance (t = 0.01 day).
M AN U
Fig. 7 Effect of elastic and thermal anisotropy on effective axial stress around the borehole wall (t = 0.01 day). Fig. 8 Effect of reservoir heating and cooling on pore pressure using different solute gradients (t = 0.01 day). Fig. 9 Effect of reservoir heating and cooling on total radial stresses using different solute gradients (t = 0.01 day).
TE D
Fig. 10 Effect of reservoir heating and cooling on total tangential stresses using different solute gradients (t = 0.01 day). Fig. 11 Comparison between conductive and conductive-convective heat transfer models on pore pressure and temperature.
EP
Fig. 12 Comparison between conductive and conductive-convective models on effective radial, tangential and axial stresses. Fig. 13 Pore pressure calculation using different models (t = 0.01 day).
AC C
Fig. 14 Effective radial stress calculation using different models (t = 0.01 day). Fig. 15 Effective tangential calculation using different models (t = 0.01 day). Fig. 16 Effective axial calculation using different models (t = 0.01 day).
Fig. 17 Pore pressure, effective radial and tangential stress (in MPa units) distribution using a) poroelasticity, b) porothermoelasticity, c) porochemoelasticity and d) porochemothermoelasticity (t = 0. 1 day). Fig. 18 Wellbore shear failure after 0.1 day using a) poroelasticity, b) porothermoelasticity, c) porochemoelasticity and d) porochemothermoelasticity (t = 0. 1 day). Failure zone is shown in blue.
ACCEPTED MANUSCRIPT
10 15 25 20 30
MPa MPa MPa MPa MPa
AC C
EP
TE D
Table 4 Rock and fluid properties. Porosity, Permeability, Fluid viscosity, Matrix bulk modulus, Fluid bulk modulus, Initial reservoir temperature, Wellbore temperature, Matrix thermal conductivity, Fluid thermal conductivity, Matrix density, Fluid density, ̅ Solid matrix specific heat, Fluid specific heat, Matrix thermal expansion coefficient, Fluid thermal expansion coefficient, Young's modulus, , Poisson's ratio, Specific entropy, Reflection coefficient, ℜ Retardation coefficient, " Thermal filtration coefficient, # $ Formation solute mass fraction,
Degree Degree Degree Degree Degree Degree
SC
Table 3 Inputs for far-field stresses. Pore pressure, Well pressure, Maximum horizontal stress, Minimum horizontal stress, Vertical (overburden) stress,
30 60 50 30 70 7
M AN U
Table 2 Inputs for stress and property rotations. Well azimuth, Well inclination, Rock dip azimuth, Rock dipping, Maximum horizontal stress azimuth, Overburden stress inclination,
RI PT
Table 1 FEM program boundary conditions. Type Well Nodes Outer Boundary Nodes Boundary Type Traction force Well pressure Far-field stresses Neumann Pore pressure Well pressure Initial pore pressure Dirichlet Temperature Well temperature Initial temperature Dirichlet = Solute mass fraction Dirichlet =
,
0.10 1 3 x 10-4 70 2.30 394 344 1.30 0.580 2640 1113 768 4181 1.8 x 10-5 3 x 10-4 24.14 0.30 3686 0.07 1 6 x 10-12 0.1
mD Pa ∙ s GPa GPa K K W/m/K W/m/K Kg/m3 Kg/m3 J/Kg/K J/Kg/K 1/K 1/K GPa J/Kg/K m2/s/K -
ACCEPTED MANUSCRIPT 0.2 5 x 10-9 0.0585 1.50
m2/s Kg/mol MPa
AC C
EP
TE D
M AN U
SC
RI PT
Wellbore solute mass fraction, Solute diffusion coefficient, # Solute molar mass, % Osmotic (swelling) coefficient, &
ACCEPTED MANUSCRIPT
Zs
Zb
Zr
V γs
Ys
V γr
Yr
γb
Yb Xb
Xr
TE D
M AN U
SC
Xs
RI PT
V
AC C
EP
E ψs N
Far-field stresses
N
E
E
ψr
Rock Properties
ψb N
Borehole
ACCEPTED MANUSCRIPT
M AN U
SC
RI PT
𝝈𝒁𝒔
Zb
AC C
𝝈𝑿𝒔
EP
TE D
Yb
Θ
BOH
Xb
𝝈𝒀𝒔
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
10
5
RI PT
Pore Pressure [MPa]
15
150
SC Analytical solution Numerical solution t = 0.001 Day t = 0.01 Day t = 0.1 Day
100
50 0.1
1
EP
0.05 0.1
TE D
0.15
200
Temperature [C]
M AN U
0.2
0.1
1
AC C
Solute Mass Fraction [-]
0 0.1
Radial Distance [m]
(at angle = 90 degrees)
1
ACCEPTED MANUSCRIPT
22 20
RI PT
18 16
M AN U
40 35
30 25
Analytical solution Numerical solution t = 0.001 Day t = 0.01 Day t = 0.1 Day
20 15 10 0.1
1
EP
20 0.1
TE D
30 25
1
SC
14 0.1
AC C
Total Radial Stress [MPa] Total Tangential Stress [MPa] Total Axial Stress [MPa]
24
Radial Distance [m]
(at angle = 90 degrees)
1
ACCEPTED MANUSCRIPT
RI PT
25
M AN U
SC
20
5
0
EP
10
TE D
15
AC C
Effective Tangential and Axial Stresses [MPa]
30
Conditional Anisotropy Complete Anisotropy Isotropic
EXr / EZr EXr / EZr
vXYr / vXZr vXYr / vXZr
-5 0.1
1.5 2.0 1
Radial Distance [m]
(at angle = 90 degrees)
ACCEPTED MANUSCRIPT
4
RI PT SC
0
M AN U
-2
TE D
-4
EP
-6 -8 -10 -12
Conditional Anisotropy Complete Anisotropy Isotropic
AC C
Effective Axial Stress [MPa]
2
0
60
120
EXr / EZr EXr / EZr 180
240
vXYr / vXZr vXYr / vXZr 300
Angle Around the Borehole Wall [Degree]
1.5 2.0 360
18
18
ACCEPTED MANUSCRIPT
14
16
14
12
RI PT
12
10
8
M AN U
6
10
8
4 0.1
Radial Distance [m]
(at angle = 90 degrees)
EP
12
4
16
14
AC C
14
1
18
TE D
18
16
6
Permeability = 1x10-5 mD
(a)
Pore Pressure [MPa]
50 C ‐50 C
SC
Pore Pressure [MPa]
16
Heating (∆T Cooling (∆T CSw > CSf CSw = CSf CSw < CSf
12
10
10
8
8
6
6
Permeability = 1 mD
(b) 4 0.1
1
Radial Distance [m]
10
(at angle = 90 degrees)
4 100
ACCEPTED MANUSCRIPT
25
RI PT
25
M AN U
SC
23
21
TE D
21
19
EP
19
17
15 0.1
Heating (∆T Cooling (∆T CSw > CSf CSw = CSf CSw < CSf
AC C
Total Radial Stress [MPa]
23
1
Radial Distance [m]
10
(at angle = 90 degrees)
50 C ‐50 C
17
15 100
ACCEPTED MANUSCRIPT
50 C ‐50 C 54
SC
RI PT
54
47
TE D
M AN U
47
40
EP
40
61
Heating (∆T Cooling (∆T CSw > CSf CSw = CSf CSw < CSf
33
26 0.1
AC C
Total Tangential Stress [MPa]
61
Radial Distance [m]
33
(at angle = 90 degrees)
1
26
ACCEPTED MANUSCRIPT
16 14 12 10
SC
0.2
150
1
10
EP
0.05 0.1
TE D
0.15 0.1
10
M AN U
0.25
200
Temperature [C]
1
AC C
Solute Mass Fraction [-]
8 0.1
RI PT
Pore Pressure [MPa]
18
Porochemothermoelastic Porochemothermoelastic (C.A.) t = 0.001 Day t = 0.01 Day t = 0.1 Day
100
50 0.1
1
Radial Distance [m]
(at angle = 90 degrees)
10
ACCEPTED MANUSCRIPT
20
10 0.1
M AN U
50 40
40 30
EP
10 0.1
TE D
30 20
10
SC
1
RI PT
15
1
10
AC C
Total Radial Stress [MPa] Total Tangential Stress [MPa] Total Axial Stress [MPa]
25
Porochemothermoelastic Porochemothermoelastic (C.A.) t = 0.001 Day t = 0.01 Day t = 0.1 Day
20 10 0 0.1
1
Radial Distance [m]
(at angle = 90 degrees)
10
ACCEPTED MANUSCRIPT
15
M AN U
13
10
9 0.1
EP
TE D
12
11
RI PT SC
+ x
AC C
Pore Pressure [MPa]
14
Poroelastic Porothermoelastic Porochemoelastic Porochemothermoelastic Porochemothermoelastic (C.A.)
Radial Distance [m]
(at angle = 90 degrees)
1
ACCEPTED MANUSCRIPT
16
RI PT M AN U
SC
12
6
4
2 0.1
EP
8
TE D
10
AC C
Effective Radial Stress [MPa]
14
+ x Radial Distance [m]
Poroelastic Porothermoelastic Porochemoelastic Porochemothermoelastic Porochemothermoelastic (C.A.)
(at angle = 90 degrees)
1
ACCEPTED MANUSCRIPT
38
SC M AN U
32
24
EP
26
TE D
30 28
RI PT
+ x
34
Poroelastic Porothermoelastic Porochemoelastic Porochemothermoelastic Porochemothermoelastic (C.A.)
AC C
Effective Tangential Stress [MPa]
36
22 20 0.1
Radial Distance [m]
(at angle = 90 degrees)
1
ACCEPTED MANUSCRIPT
18
RI PT
17
M AN U
SC
15 14
11 10
EP
12
TE D
13
AC C
Effective Axial Stress [MPa]
16
+ x
9 8 0.1
Radial Distance [m]
Poroelastic Porothermoelastic Porochemoelastic Porochemothermoelastic Porochemothermoelastic (C.A.)
(at angle = 90 degrees)
1
Effective Tangential Stress
SC
RI PT
(a)
EffectiveMANUSCRIPT Radial Stress ACCEPTED
0.40 m
Pore Pressure
TE D EP AC C
(c)
M AN U
(b)
(d)
RI PT
(a)
EP
TE D
M AN U
SC
(b)
AC C
0.30 m
ACCEPTED MANUSCRIPT
(c)
(d)
ACCEPTED MANUSCRIPT •
New numerical model for simulating anisotropic formations under THMC loadings.
•
The model addresses chemically-active, anisotropic non-isothermal formations.
•
The model can be used to assess time dependent wellbore stability.
•
The model addresses inclined boreholes drilled in arbitrarily oriented rock
EP
TE D
M AN U
SC
A novel pseudo-3D analysis is used to models 3D problems using 2D FEM mesh.
AC C
•
RI PT
laminations.
ACCEPTED MANUSCRIPT
Λ
Unit Pa Pa 1/Pa Pa Pa Pa/K K Pa 1/K mD Pa ∙ s Pa Pa W/m/K W/m/K Kg/m3 Kg/m3 J/Kg/K J/Kg/K 1/K 1/K Pa Pa J/Kg/K m2/s/K m2/s Kg/mol Pa Degree Degree Degree Degree Degree Degree Degree Degree Degree Pa Pa Pa Pa Pa
! "# $# "% $% " $ "# $# "% &' &( )* +* ,*
TE D
EP
ℜ
AC C
̅
M AN U
SC
Ξ Ω
Definition Stress tensor Drained stress-strain stiffness tensor Drained stress-strain Compliance tensor Strain tensor Pore pressure Chemo-mechanical coupling term Solute mass fraction Thermo-mechanical coupling term Temperature Biot’s effective stress coefficient Variation of fluid content per unit referential volume Biot’s modulus Hydro-chemical coupling term Hydro-thermal coupling term Porosity Permeability Fluid viscosity Matrix bulk modulus Fluid bulk modulus Matrix thermal conductivity Fluid thermal conductivity Matrix density Fluid density Solid matrix specific heat Fluid specific heat Matrix thermal expansion coefficient Fluid thermal expansion coefficient Young's modulus Poisson's ratio Shear modulus Specific entropy Reflection coefficient Retardation coefficient Thermal filtration coefficient Solute diffusion coefficient Solute molar mass Osmotic (swelling) coefficient Well azimuth Well inclination Rock dip azimuth Rock dipping Maximum horizontal stress azimuth Overburden stress inclination Well azimuth Well inclination Rock dip azimuth Pore pressure Well pressure Maximum horizontal stress Minimum horizontal stress Vertical (overburden) stress
RI PT
Symbol
Non-Business Use
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT