Sie sind auf Seite 1von 31

Fluid flow through anisotropic and deformable double porosity media with

ultra-low matrix permeability: A continuum framework

Qi Zhanga , Ronaldo I. Borjaa,∗


a
Department of Civil and Environmental Engineering, Stanford University, Stanford, CA 94305, USA
arXiv:2002.08180v4 [physics.geo-ph] 2 Apr 2020

Abstract
Fractured porous media or double porosity media are common in nature. At the same time, accurate mod-
eling remains a significant challenge due to bi-modal pore size distribution, anisotropy, multi-field coupling
and various flow patterns. The purpose of this study is to formulate a comprehensive coupled flow and ge-
omechanics model of anisotropic and deformable double porosity media with ultra-low matrix permeability.
Fluid in fissures is modeled with the generalized Darcy’s law with an equivalent permeability upscaled from
the detailed geological characterizations while the liquid in much less permeable matrix follows a low veloc-
ity non-Darcy flow characterized by threshold values and non-linearity, and fluid mass transfer is dependent
on the shape factor, phase pressure difference, and interface permeability. The geomechanics relies on a
thermodynamically consistent effective stress derived from the energy balance equation, and it is modeled
following poroelastic theory. Scaling analysis is performed to drop negligible force density terms under
reasonable parameters’ ranges which also guarantee no violation of entropy inequality. The discussion re-
volves around generic double porosity media. Numerical simulation of the initial boundary value problem
reveals the capability of this framework to capture the crucial role of coupling, anisotropy and ultra-low
matrix permeability in dictating the pressure and displacement fields.
Keywords: Double porosity, Geomechanics, Upscaling, Anisotropy, Ultra-low matrix permeability

1. Introduction

In our natural environment, the real reservoirs tend to be very heterogeneous in both porosity and perme-
ability characteristics due to the existence of porous constituents at various length scales [3]. The accurate
simulation of real reservoirs remains a significant challenge. Instead, people will idealize the actual reser-
voir as an aggregate of different geological regions (e.g., host rock, fracture, fault, compaction band, and
so on), among which the fractured porous media is widely adopted in engineering practice. Sometimes
fractured porous media is also called double (dual) porosity media. Although researchers in reservoir en-
gineering firstly proposed the concept of fractured porous media, nowadays, it has been widely used in
other branches of earth sciences, with different terminologies to describe fracture and porous matrix. In
the context of aggregated soils, they are usually termed macropores and micropores [32], whereas in the
context for sedimentary rocks, we shall call them micro-cracks and nanopores [124]. It must be noted that
the macro-fractures which can be seen with the naked eye constitute another (much larger) porosity scale
[126], which is not covered in the present discussion.


Corresponding author
Email address: borja@stanford.edu (Ronaldo I. Borja)

Preprint submitted to Advances in Water Resources April 6, 2020


In fractured porous media, due to significant differences in the pore sizes, fluid transmission through
the fissures and the rock matrix could be very distinctive. In most cases, fissures serve as the primary
conduit for fluid flow, whereas the smaller pores in the rock matrix generally play a supporting role in the
fluid transmission, and for the most part, simply discharge fluid into the fissures [115]. Generally speaking,
there are two classes of methods used for modeling fractured porous media: discrete (explicit) methods and
continuum (implicit) methods [3, 13, 43, 120]. In discrete methods, we have different grids for fracture and
matrix, therefore the fracture flow and interplay between fracture and surrounding matrix can be modeled
accurately [13, 43, 58, 106, 120]. However, the computational cost could become a big issue when the
fracture network is very involved in the reservoir scale [3]. In contrast, in continuum methods, there is no
geometric distinction between fracture and matrix [74, 115], which means the reservoir model is represented
by two overlapping continua [12, 103], and the interaction between these two continua is modeled through
a transfer function [74], also called the shape factor [63, 110]. As a result, the computational cost can be
reduced. However, the existence of representative elementary volume (REV) and accuracy of homogenized
parameters such as equivalent permeability and shape factor cannot be guaranteed [13]. Some work has
been done to derive the continuum model from the fine scale model. For example, Arbogast et al. [2]
provided equations of the same form as the standard dual porosity model using homogenization theory.
However, these results are available only under very specific assumptions on geometry. In Prof. Durlofsky’s
research group, they have applied the upscaling technique to construct a dual porosity model from a discrete
fracture matrix (DFM) model [51, 59, 61, 78]. However, their upscaled model is still in a discrete sense.
Besides these efforts, numerous papers have been published to improve the constant, time-independent
shape factor used in [12, 110]. For instance, the improvements could be incorporation of time (history)
dependence [25, 95, 98, 100], higher-order models [41, 72], a series of first-order transfer terms with varying
shape factors [46, 52], or a series of multiple interacting continua (MINC) that replaces the original matrix
continuum [58, 71, 94, 115, 116, 117]. While these approaches improve the accuracy, their implementation
tends to be inelegant [13].
Darcy’s law is the most fundamental equation to describe fluid flow in conventional reservoirs or micro-
cracks [89, 101, 124]. However, for unconventional and tight reservoirs with ultra-low matrix permeability
(assuming isotropic), Darcy’s law will overestimate the flow rate of liquid due to the interaction between
fluid particles and the solid pore wall [38, 108, 118, 119, 123]. The result of this interaction is the formation
of a boundary layer on which the liquid exhibits higher viscosity [108]. Although this boundary layer effect
becomes weaker when |∂p/∂xi | is high [109] due to destruction of interacting forces [123], in any case,
the average flow rate tends to be lower than the value calculated from Darcy’s law [73]. In the literature,
this phenomenon is known as low velocity non-Darcy flow of liquid (lower limit of Darcy’s law [38]) or
nonlinear percolation [109]. It has been argued in the literature that when |∂p/∂xi | is extremely small,
say, lower than a scalar called threshold pressure gradient (TPG) [54], the boundary layer will prevent the
fluid from flowing [73], and above this TPG, the flow curve in each direction shows a certain level of non-
linearity [109], followed by a straight line [79]. In other words, Darcy’s law should be corrected for the
effect of the TPG [93], and the resulting flow velocity qi is a nonlinear function of the pressure gradient
∂p/∂xi [73, 123]. In addition to the TPG, a scalar called pseudo threshold pressure gradient (PTPG) has
also been defined in the literature as the intersection between the reverse extension of the linear flow plot
qi vs. ∂p/∂xi and the horizontal axis [79]. However, there also exist some opposite opinions about TPG
[108], which asserted such threshold does not exist or it is a misinterpretation of experimental data, and
they chose an alternative nonlinear model [108]. To sum it up, the abstract form of many non-Darcy flow
models (single phase, single porosity, no gravity, isotropic permeability k) that are prevalent in the literature

2
is given as follows [53, 79, 102, 104, 108, 119, 124]
 
q = F k, µ f , ∇p, ã, b̃, c̃, d̃, . . . , (1)

where F is either a piecewise vectorial function [79] or a single nonlinear vectorial function [108], and
the scalar parameters ã, b̃, c̃, d̃, . . . could represent TPG, PTPG, any other critical values [79], factor of
adjustment [73] and even some fitting parameters [108] that could also be a function of ∂p/∂xi [109] and
pore radius [109]. In conventional reservoir or under large |∂p/∂xi |, above terms should become trivial.
Besides, the whole flow curve in each direction is required to be continuous.
In addition to extensive investigations on flow problem, the tightly coupled hydromechanical behavior
(also known as the phenomenon of geomechanics) is central to the performance of many subsurface systems
and is critical for assessing environmental impacts [24]. This strong coupling is always modeled as a two
way coupled process, which is described by the well-established poromechanical theory for conventional
single porosity media. For fractured porous media, a fair amount of papers have been published to consider
geomechanics in both continuous and discrete senses. For continuum modeling, Wilson et al. [114] made
the first attempt to consider geomechanics in double porosity media by introducing generalized notions of
deformation. Over the last 30 years, efforts along this line have resulted in seven main modeling approaches
that are individually developed, and they differ in terms of assumptions and parameters. The representative
authors for each modeling approach are (in alphabetical order of their last names): Abousleiman [84, 85, 86,
91], Bai and Elsworth and Roegiers [5, 6, 7, 8, 9, 10, 42, 125], Berryman and Wang [14, 15, 16, 75], Borja
[21, 22, 32, 34, 124], Chen [27, 28], Khalili [64, 65, 66, 67, 68, 69, 80, 82, 105], Lewis [50, 76, 77, 83, 90,
92]. Almost all the discrepancies come from the actual modeling of two porosity change and this remains,
to the best of current authors’ knowledge, an open question. For discrete modeling, Garipov et al. [45]
combined DFM with fracture contact problem, leading to a fully implicit formulation of coupled flow and
geomechanics for fractured three-dimensional subsurface formations. Jiang et al. [57] further considered
the compaction and embedment of proppant in the fracture contact problem of a shale gas reservoir, and
a splitting-node technique was used to deal with the discontinuities in the displacement field across the
fracture interface. Yan et al. [121, 122] applied embedded discrete fracture matrix (EDFM) method to model
macro-fractures and they also developed a stabilized extended finite element method (XFEM) to eliminate
displacement oscillation along macro-fracture boundaries. We can see the difficulty for discrete modeling
lies in the accurate characterizations of fracture behaviors and their impacts on flow and deformation.
This paper aims to develop a mathematically consistent framework for fluid flow through anisotropic
and deformable double porosity media with ultra-low matrix permeability. Since dual continuum models
can provide practical insight into the global behaviors of the multiscale system with much less computa-
tional cost [3], I will focus on the continuum method while at the same time try to connect with discrete
descriptions of the fissures. All the derivations are presented under the assumption of single phase flow
and infinitesimal deformation. The paper is organized as follows: detailed mathematical formulations are
first presented which adopt mixture poromechanics theory, but this time I will reveal from the most general
perspective which to the authors’ knowledge, has never been proposed in previous papers. Model simplifi-
cations including scaling analysis are then performed under reasonable assumptions of material properties.
Subsequently, a simple upscaling approach is proposed to fill the gap between discrete geological descrip-
tions and double porosity flow parameters. A 3D numerical example is given in the end. In this paper, I will
use subscript 1 and 2 to represent primary porosity (porous blocks or micropores) and secondary porosity
(fissures or macropores), respectively.

3
2. Mathematical formulations
2.1. Fluid flow in anisotropic double porosity media with ultra-low matrix permeability
In continuum modeling of double porosity media, three mass conservation equations are built on the
same overlapping domain, distinguished by Eulerian volume fractions φ1 = V1 /Vb , φ2 = V2 /Vb and 1 −
φ1 − φ2 = V s /Vb where Vb = V1 + V2 + V s is the total (bulk) volume of the mixture. Following Eulerian
description, above three mass conservation equations are given as [36, 44, 70]
∂ 
ρ s (1 − φ) + ∇ · ρ s (1 − φ) v s = 0 ,
  
(2)
∂t
∂ h i h i
ρ1 f φ1 + ∇ · ρ1 f φ1 v̂1 = c1 , (3)
∂t
∂ h i h i
ρ2 f φ2 + ∇ · ρ2 f φ2 v̂2 = c2 , (4)
∂t
where φ = φ1 + φ2 , ρ s is the solid density, ρ1 f and ρ2 f are the fluid densities, v s is the velocity of the solid
skeleton, v̂1 and v̂2 are the interstitial velocities or tracer velocities, c1 and c2 represent the source terms
(Unit: c1 = c2 = ML−3 T −1 ). If there is only mass exchange between domain 1 and domain 2, then


c1 + c2 = 0.
The remaining task is to rewrite Eqs. (3)(4) into equivalent forms. To begin with, we need to introduce
the material time derivative following a particular constituent (s or 1 f or 2 f ). The definitions are given here
d s (·) ∂ (·)
= + ∇ (·) · v s , (5)
dt ∂t
d1 f (·) ∂ (·)
= + ∇ (·) · v̂1 , (6)
dt ∂t
d2 f (·) ∂ (·)
= + ∇ (·) · v̂2 . (7)
dt ∂t
In the following text, since we will prefer material time derivative following the solid phase motion when-
ever possible, I will drop the superscript s in d s (·) /dt. For Eq. (3) and Eq. (4), the following processes are
basically the same except for the change of the subscript. Thus I will deal with Eq. (3) in detail and provide
the final form of Eq. (4) at the end of my derivation. By expanding derivatives (time and space) in Eq. (3),
we will get
∂φ1 ∂ρ1 f
ρ1 f + φ1 + φ1 v̂1 · ∇ρ1 f + ρ1 f ∇ · φ1 v̂1 = c1 .
 
(8)
∂t ∂t
In Eq. (8), after extraction of φ1 , the combination of the second and the third terms on LHS is exactly the
material time derivative of constituent 1 f . Adopting this definition and further dividing ρ1 f on both sides,
we will obtain
∂φ1 φ1 d1 f ρ1 f c1
+ + ∇ · φ1 v̂1 = .
 
(9)
∂t ρ1 f dt ρ1 f
Now we need to define superficial velocity q1 in our context which is given as follows

q1 = φ1 (v̂1 − v s ) . (10)

Then we can further rewrite Eq. (9) using q1 and we will have

∂φ1 φ1 d1 f ρ1 f c1
+ + ∇ · q1 + ∇ · φ1 v s = .
 
(11)
∂t ρ1 f dt ρ1 f
4
We can further combine the first and last terms on LHS in Eq. (11), see below
∂φ1  ∂φ1 dφ1
+ ∇ · φ1 v s = + v s · ∇φ1 + φ1 ∇ · v s = + φ1 ∇ · v s .

(12)
∂t ∂t dt
By substituting the result of Eq. (12) into Eq. (11), we would have

φ1 d1 f ρ1 f dφ1 c1
+ ∇ · q1 + + φ1 ∇ · v s = . (13)
ρ1 f dt dt ρ1 f
Note in the literature you might see alternative expressions of the first two terms in Eq. (13) [27]. In fact,
we can prove following identical relation easily
h i
φ1 dρ1 f ∇ · ρ1 f q1 φ1 d1 f ρ1 f
+ = + ∇ · q1 . (14)
ρ1 f dt ρ1 f ρ1 f dt
For the detailed proof, see Appendix A. If we further define the fluid bulk modulus K1 f (also known as the
PVT relationship), which can be expressed mathematically as

1 d1 f ρ1 f 1 d1 f p1 1 dρ1 f 1 d p1 1 ∂ρ1 f 1 ∂ p1
= , = , = , (15)
ρ1 f dt K1 f dt ρ1 f dt K1 f dt ρ1 f ∂t K1 f ∂t
where p1 is the fluid pressure. It is easy to verify following relation between ρ1 f and p1 satisfies Eq. (15)

p1 − p0
!
ρ1 f = ρ f exp , (16)
K1 f

where ρ f is the reference fluid density and p0 is the reference pressure. Equation (16) will be used in this
paper. Combining Eq. (15) and Eq. (13) gives

φ1 d1 f p1 dφ1 c1
+ ∇ · q1 + + φ1 ∇ · v s = . (17)
K1 f dt dt ρ1 f
Exactly the same logic also applies to Eq. (4). As a result, we will obtain

φ2 d2 f p2 dφ2 c2
+ ∇ · q2 + + φ2 ∇ · v s = . (18)
K2 f dt dt ρ2 f
Until now, except for the assumption of the basic double porosity model which admits two overlapping
continua and PVT relationships for 1 f and 2 f , no additional assumption has been made in the preceding
derivations to Eqs. (17)(18). In other words, we can regard Eqs. (17)(18) as the starting point to introduce all
kinds of specific constitutive relations (approximations), among them the modeling of dφ1 /dt and dφ2 /dt is
still an open question, and that is why we will have at least seven different modeling approaches mentioned
in Sec. 1. In this paper, we will try to move one step further by proposing a new expression inspired by
[3, 29, 30, 70] that could incorporate all the other existing modeling approaches. The expression is given as
follows
(α1 − φ1 1) .. d
 dφ1  " #  d p1  ( )
A11 A12 
= + ,
 dt 
  dt  dt
(α2 − φ2 1) .. d
(19)
 
 dφ2 
  A21 A22  ddtp2 
 
dt dt
h i
where Ai j is the 2 × 2 coefficient matrix, α1 and α2 are the symmetric Biot tensors, 1 is the second order
identity tensor,  = (∇u + u∇) /2 is the infinitesimal strain tensor and u is the solid displacement vector
5
which satisfies v s = du/dt. Equation (19) will be used in Sec. 2.2 to manipulate linear momentum and
energy balance equations.
Besides dφ1 /dt and dφ2 /dt, we still need to provide expressions for q1 , q2 , c1 and c2 to close our mass
balance formulation. For q1 , we assume it follows the low velocity non-Darcy flow of liquid that can be
expressed as a nonlinear function of ϕ = ∇p1 − ρ1 f g

ϕ j < λmin ,



 0

ξ

  
k1 ϕ j − λmin



  
ϕ λ ϕ j ≤ λmax ,

−
q1 j =  sgn ≤
 
 µ1 f ξ δ ξ−1
j min (20)


ξ−1
 !
k1

  
− µ ϕ j − λmin − ξ δ sgn ϕ j ϕ j > λmax ,





1f

where g is the gravity acceleration vector (the last element and also the only non-zero element is −g),
j = 1, · · · , ndim where ndim is the space dimension, k1 is the Darcy permeability of the matrix, µ1 f is the
fluid viscosity, ξ ≥ 1 is the exponent parameter [127], δ = λmax − λmin , λmin has the the physical meaning
of threshold gradient and λmax has the physical meaning of critical gradient [127], the pseudo gradient of
this model can be calculated analytically as λmin + δ (ξ − 1) /ξ. In this paper, we assume ξ, λmin and λmax
are constants. From Eq. (20), we can see that this non-Darcy flow model is a combination of a no flow
part, a nonlinear flow part and a linear flow part, separated by λmin and λmax . In addition, if we set ξ = 1
and λmin = 0, isotropic Darcy’s law is automatically recovered. Thus Eq. (20) is more general than similar
equations in [53, 79, 102, 124]. For q2 , we assume it follows the generalized Darcy’s law, which is given
by the relation
k2  
q2 = − · ∇p2 − ρ2 f g . (21)
µ2 f
Due to this full tensor permeability k2 , the direction of q2 depends on both pressure gradient and the
principal directions of anisotropy [89]. Finally, for c1 and c2 , we adopt prototype in [48, 49, 110] with
minor modifications
ρ f σsh k̄
c1 = (p2 − p1 ) + MS1 , (22)
µf
ρ f σsh k̄
c2 = (p1 − p2 ) + MS2 , (23)
µf
where σsh is the shape factor whose mathematical models are summarized in [95, 96, 98], k̄ is the interface
permeability whose mathematical model will be elaborated in the next paragraph, µ f is the reference fluid
viscosity, MS1 and MS2 represent other mechanisms besides mass transfer.
In double-porosity models, the interface permeability k̄ also depends on the apparent permeability
∂q1 /∂ϕ of the matrix [68]. In other words, in double porosity media the term “non-Darcy flow” pertains
not only to the flow through nearly impermeable matrix, but also to the discharge of fluid from the matrix
into the fissures. Therefore, we propose following form for k̄ that is analogous to the variations of aquifer
properties in space [52]

ϕmax < λmin ,





 k̄min

 k̄max + k̄min k̄min − k̄max ϕmax − λmin ξ


 "   #
k̄ =  + cos π λmin ≤ ϕmax ≤ λmax ,

(24)



 2 2 δ

ϕmax > λmax ,

k̄max

6
where k̄max and k̄min are the maximum and minimum interface permeabilities, respectively, and

ϕmax = max ϕ j . (25)
j∈{1,··· ,ndim }

When k̄max = k̄min , the interface permeability k̄ becomes a trivial constant, whereas when k̄max > k̄min and
as ϕmax decays, the decrease in k̄ due to the boundary layer effect is analogous to that which occurs when
a small value of diffusion property is specified for clay layers or clay pods (immobile zone) [52]. Later we
shall show that this variation in k̄ also affects the entire fluid flow patterns. In addition, since in general k̄
will not be a constant, this composite non-linear mass transfer term can be regarded as a supplement for the
coupling terms proposed in [41, 72].

2.2. Geomechanics
To address the issue of geomechanics, we need to define several “partial” quantities. First of all, we
need to define partial stress tensor σa for every constituent a = s, 1 f and 2 f . By using Cauchy’s traction
theorem, the σa is defined such that it linearly transforms a normal vector n to a partial traction vector t a
that equals the force δ f a acting on constituent a divided by the total area δA, i.e.,
δ fa
σa · n = t a = lim . (26)
δA→0 δA
Secondly we need to define the partial density ρa as the mass of constituent a divided by the total volume,
and we can relate ρa with ρa easily using porosity as

ρ s = ρ s (1 − φ) , (27)

ρ1 f = ρ1 f φ1 , (28)
ρ 2f
= ρ2 f φ2 . (29)
Thirdly, we use ha to represent the internal drag force on the constituent a by the surrounding other con-
stituents divided by the total volume and it has the property that
X
ha = 0 . (30)
a

Finally, v̂a represents the interstitial velocity for constituent a and it is just v s , v̂1 and v̂2 for constituents s,
1 f and 2 f , respectively.
The balance of linear momentum for constituent a in integral form is given as follows
da
Z Z Z Z
σ · n dA +
a
ρ g dV +
a
h dV =
a
ρa v̂a dV . (31)
A V V dt V
Note for the RHS of Eq. (31), we cannot interchange the time derivative and integral sign due to a moving
control volume V. As a result, we need to use Reynold transport theorem to simplify RHS as
da ∂ρ
Z Z " a # Z
ρ v̂a dV =
a
+ ∇ · ρ v̂a v̂a dV +
a 
ρa aa dV , (32)
dt V V ∂t V

where aa = da v̂a /dt. Surprisingly, the term inside the square bracket is exactly the LHS of Eqs. (2)(3)(4).
Therefore, we can further simplify Eq. (32) as
da
Z Z Z
ρ v̂a dV =
a
ca v̂a dV + ρa aa dV , (33)
dt V V V
7
where we assume c s = 0, c1 f = c1 and c2 f = c2 . For LHS of Eq. (31), we just use divergence theorem and
we will obtain Z Z Z
∇ · σa + ρa g + ha = ca v̂a dV + ρa aa dV .
 
(34)
V V V
Now we can apply the localization theorem to drop the integral sign and get partial differential equation as

∇ · σa + ρa g + ha = ca v̂a + ρa aa . (35)

The deformation equations for the entire mixture under quasi-static condition is obtained by summing
Eq. (35) for all constituents, using Eq. (30) and setting aa = 0, this yields

∇ · σ + ρ g = c̄ , (36)

where σ = a σa is the total Cauchy stress tensor, ρ = a ρa is the bulk density, c̄ = c1 v̂1 + c2 v̂2 is the
P P
additional momentum supply exerted by c1 and c2 [3].
Constitutive theory for granular materials is usually expressed in terms of the effective stress, so defining
its mathematical form as a function of the total stress and pore pressure is crucial for constitutive modeling
[18]. The expression for effective stress can be established with the aid of the first law of thermodynamics
subject to no violation of entropy inequality (second law of thermodynamics). To proceed, we denote by e
the internal energy per unit total mass of the mixture and η the entropy per unit total mass of the mixture.
Without loss of generality we shall assume an isothermal process, i.e., we consider only mechanical terms.
Next we define two additional variables E, H such that they satisfy
Z Z
dI d dE
= ρe dV = dV , (37)
dt dt V V dt
Z Z
dH d dH
= ρη dV = dV . (38)
dt dt V V dt
The first law of thermodynamics says
Z
dE X da Z ρa v̂a · v̂a
dV = P −
ME
dV . (39)
V dt a
dt V 2

If we follow the same logic as shown in Eqs. (32)(33), we can rewrite above equation as
Z X Z ca v̂a · v̂a XZ
dE
dV = P −
ME
dV − ρa v̂a · aa dV . (40)
V dt a V 2 a V

The mechanical power P ME is the sum of the powers of t a , ρa g and ha for all constituents, which is shown
here
X "Z Z Z #
P =
ME
σ · n · v̂a dA +
a
ρ g · v̂a dV +
a a

h · v̂a dV (41)
a A V V

Applying divergence theorem on the first term of RHS yields


Z Z Z
σ · n · v̂a dA =
a
∇ · σ · v̂a dV +
a
σa .. la dV ,

(42)
A V V

8
where la = (∇v̂a + v̂a ∇) /2 is the infinitesimal strain rate tensor of constituent a. Thus, we can rewrite
Eq. (41) as
X "Z Z Z Z #
a ..
P =
ME
∇ · σ · v̂a dV +
a
σ la dV + ρ g · v̂a dV +
a
h · v̂a dV .
a
(43)
a V V V V

Combing Eqs. (40)(43) yields


Z
dE X "Z Z Z
ca v̂a · v̂a
#
a ..
dV = ∇ · σ + ρ g + h − ρ aa · v̂a dV +
a a a a 
σ la dV − dV . (44)
V dt a V V V 2

Now we need to use balance of linear momentum Eq. (35) to replace ∇ · σa + ρa g + ha − ρa aa , the result is
Z
dE X "Z ca v̂a · v̂a Z #
dV = dV + σa .. la dV . (45)
V dt a V 2 V

Again, if we apply localization theorem to above equation, we will reach our first milestone
dE 1 X
σa .. la .
X
= ca v̂a · v̂a + (46)
dt 2 a a

However, there are still something else that we can do with a σa .. la in Eq. (46). Let us expand it
P

σa .. la = σ s .. l s + σ1 f .. l1 f + σ2 f .. l2 f .
X
(47)
a

In the following text I will use l to represent l s since this is the quantity that we are interested in. From the
definition of total Cauchy stress tensor σ, we know σ s = σ − σ1 f − σ2 f and if we substitute this relation
into Eq. (47), we will get

σa .. la = σ .. l + σ1 f .. l1 f − l + σ2 f .. l2 f − l .
X    
(48)
a

Now we will assume σ1 f and σ1 f are isotropic andcan be represented as σ1 f = −φ1 p1 1 and  σ = −φ2 p2 1.
2f

From the definition of la , we can easily prove 1 .. l1 f − l = ∇ · (v̂1 − v s ) and 1 .. l2 f − l = ∇ · (v̂2 − v s ).


Therefore,
σa .. la = σ .. l − φ1 p1 ∇ · (v̂1 − v s ) − φ2 p2 ∇ · (v̂2 − v s ) .
X
(49)
a
From the definition of q1 and q2 , we know

∇ · q1 = ∇ · φ1 (v̂1 − v s ) ' φ1 ∇ · (v̂1 − v s )


 
(50)

and
∇ · q2 = ∇ · φ2 (v̂2 − v s ) ' φ2 ∇ · (v̂2 − v s )
 
(51)
given (v̂1 − v s ) · ∇φ1 and (v̂2 − v s ) · ∇φ2 are relatively small. At the same time, ∇ · q1 and ∇ · q2 also appear in
Eqs. (17)(18), respectively. As a result, we can represent φ1 p1 ∇ · (v̂1 − v s ) and φ2 p2 ∇ · (v̂2 − v s ) in Eq. (49)
in terms of other items using Eqs. (17)(18)(50)(51)

φ1 d1 f p1 dφ1
" #
c1
φ1 p1 ∇ · (v̂1 − v s ) = p1 ∇ · q1 = p1 − − − φ1 ∇ · v s , (52)
ρ1 f K1 f dt dt
9
φ2 d2 f p2 dφ2
" #
c2
φ2 p2 ∇ · (v̂2 − v s ) = p2 ∇ · q2 = p2 − − − φ2 ∇ · v s . (53)
ρ2 f K2 f dt dt
Then we can insert Eqs. (52)(53) into Eq. (49) and further substitute the new Eq. (49) and abstract form
Eq. (19) into Eq. (46), we will reach our second milestone

= σ .. l +
dE c1 v̂1 · v̂1 c2 v̂2 · v̂2
+
dt 2 2
φ1 d p1
1 f
" #
+ p1 + A11
d p1
+ A12
d p2
+ α1 .. d

c1
K1 f dt dt dt dt ρ1 f (54)
φ2 d2 f p2
" #
+ α1 ..
d p1 d p2 d c2
+ p2 + A21 + A22 − .
K2 f dt dt dt dt ρ2 f

Now we just need to use the definition l = d/dt and we can rewrite above Eq. (54) into an equivalent form

= (σ + α1 p1 + α2 p2 ) .. l +
dE c1 v̂1 · v̂1 c2 v̂2 · v̂2
+
dt 2 2 #
φ1 d p1
1 f
"
d p1 d p2 c1
+ p1 + A11 + A12 − (55)
K1 f dt dt dt ρ1 f
φ2 d2 f p2
" #
d p1 d p2 c2
+ p2 + A21 + A22 − .
K2 f dt dt dt ρ2 f

This is our final destination. In this Eq. (55)

σ̄ = σ + α1 p1 + α2 p2 (56)

is identified as the effective Cauchy stress tensor. Here we assume vertically transversely isotropic (VTI)
linear elasticity [128], which is given by the constitutive relation in Voigt notation
  1 − νEhhh − νEvhv
 
   σ̄ x 

x  Eh
− νhh − νEvhv
 
1
    
y  σ̄

 
  

 
  Eh Eh   
 y



 νhv − νEhvh
  
1
 z   σ̄z 
    
 − Eh

  
 
= Ev
.
    
(57)

2 xy  2(1+νhh )
σ̄ xy 

      

 
  Eh    

σ̄ xz 
   

 2

  1    


 xz 

  Gvh

 
 


 2    1  σ̄ 
 

yz yz
Gvh
| {z }
S

From displacement reciprocal theorem, we have


νvh νhv
= , (58)
Ev Eh
which guarantees the symmetry of the compliant matrix S. The elastic Poisson’s ratio νvh should be under-
stood as the stress applied in the vertical (bed-normal BN) direction and strain measured in the horizontal
(bed-parallel BP) direction, νhv should be interpreted as the stress applied in the horizontal (bed-parallel
BP) direction and strain measured in the vertical (bed-normal BN) direction, Gvh is the shear modulus along
the BN direction, Eh and Ev are two Young’s moduli. Note all the elastic constants are drained quantities.
If Eh = Ev = E, νvh = νhv = ν and Gvh = G = 0.5E/ (1 + ν), then we restore the isotropic linear elasticity.
10
Note both VTI linear elasticity and isotropic linear elasticity are just special cases of the anisotropic linear
elasticity.
For entropy inequality, we need to define free energy density Ψ per unit current volume of the mixture
[17, 18, 22] as Ψ = E − T H, where T > 0 is the absolute temperature. The Clausius-Duhem inequality for
isothermal process reads
dH dE dΨ
D =T = − ≥ 0. (59)
dt dt dt
We will recap this equation in Sec. 3.

3. Model simplifications

Although Eq. (19) brings additional physical insight into the properties of double porosity material,
the resulting material system (Ai j , α1 and α2 ) is too cumbersome to be practically used. Approximate
expressions are often introduced to make the model more tractable. Appendix B and Appendix C discuss
several different approximations for single porosity media (modeling of dφ/dt) and double porosity media
(modeling of dφ1 /dt and dφ2 /dt). Even though single porosity media is not our focus, it is worthwhile to
mention this interesting discrepancy. As a result, we assume α1 = α1 1, α2 = α2 1 and all the coefficients
in Eq. (19) are constants, which means porosities changes and inelasticity are excluded from coefficients
calculations. Besides, from the last paragraph of Appendix C, we know Ai j ≡ 0 is not a bad idea. Also,
for soils and some rocks, the drained bulk modulus Kb is a few orders of magnitude lower than the fluid
bulk modulus, so we will also assume K1 f = K2 f = ∞, which means ρ1 f = ρ2 f = ρ f from Eq. (16). For
water, its viscosity can be regarded as a constant in a certain range, i.e., µ1 f = µ2 f = µ f . Under all of these
assumptions, we can simplify Eqs. (17)(18)(55), which is shown in turn

σsh k̄ (p2 − p1 )
∇ · q1 + α1 ∇ · v s = , (60)
µf

σsh k̄ (p1 − p2 )
∇ · q2 + α2 ∇ · v s = , (61)
µf
dE c1 v̂1 · v̂1 c2 v̂2 · v̂2 σsh k̄ (p1 − p2 )2
= σ̄ .. l + + + . (62)
dt 2 2 µf
Note in Eqs. (60)(61)(62), we have already used Eqs. (22)(23) and assumed only mass transfer is incorpo-
rated as a source or a sink term.
In Eq. (36), there is no direct constitutive formula to calculate c̄. However, we will show through scaling
analysis given in Appendix D, we can drop this term safely and Eq. (36) will become

∇ · σ̄ − α1 ∇p1 − α2 ∇p2 + ρg = 0 . (63)

Now let us come back to entropy inequality, substituting Eq. (62) into Eq. (59) and further assuming Ψ =
 .. S−1 .. /2 under elasticity [17, 22] yields
dΨ  −1 .  . d
= S . . = σ̄ .. l , (64)
dt dt
dE dΨ σsh k̄ (p1 − p2 )2 c1 v̂1 · v̂1 c2 v̂2 · v̂2
D= − = + + ≥ 0, (65)
dt dt µf 2 2
11
where S−1 is S−1 written in 4th order tensorial form. Again, in Appendix D, we have shown through scaling
analysis, the term c1 v̂1 · v̂1 /2 is not dominant. Thus we are left with the first term and the last term. In most
cases such as consolidation, the fluid will always transfer from matrix to fissures, which makes c2 > 0. As
a result, the entropy inequality is satisfied and we can conclude that our constitutive relations are valid.

Single porosity model Our model is capable of reproducing the conventional single porosity model
by assigning α2 = 0, α1 = 1 and isotropic k2 on the whole domain, assigning homogeneous boundary
condition p2 = 0, and dropping mass transfer terms c1 and c2 . Under this condition, the PDE for p2 will
become trivial (∇ · q2 = 0) and the solution is p2 ≡ 0, while the Eqs. (60)(63) will become

∇ · q1 + ∇ · v s = 0 , (66)

∇ · σ̄ − ∇p1 + ρ g = 0 , (67)
which correspond exactly to what we use in [111, 112] if we drop the subscript 1.

Steady state flow model The steady state flow model is derived by setting ∇ · v s = 0 in Eq. (60)(61),
in other words, we assume the solid skeleton to be rigid. As a result, a disturbance at a point is felt instan-
taneously over the entire domain.

4. Computing equivalent permeability k2

In the double porosity model, one of the most challenging tasks is to give an accurate estimate of
homogenized or equivalent k2 for each sub-region. In this section, we will provide a feasible way to
calculate k2 . The method will follow similar procedures as those describedin Sec. 4.1 of [40]. Here, we
will illustrate this method using a purely local 2D Cartesian grid l x × ly × 1 with unit thickness shown in
Fig. 1, and we require it should contain at least one fracture. Since dual continuum models are especially
appropriate for reservoirs with a large number of highly connected, small-scale fractures [59, 61, 87] such
as Fig. 3 in [87] and Fig. 1 in [91], this requirement is pretty trivial. Extensions to 3D and extended
local upscaling (in order to include the effects of neighboring regions in these calculations) could be done
analogously.
We will consider two solutions of incompressible (both solid skeleton and fluid) single phase flow
problems. Note here we must deal with fissures explicitly either through DFM combined with local grid
refinement (LGR) [26, 58, 60] or EDFM [39], but the good thing is that we just need to solve a single PDE,
i.e., ∇ · q = 0. Both solutions will use linear pressure gradient boundary conditions [40]. In the first solution
we set
p (0, y) = 1 , (68)
p (l x , y) = 0 , (69)
  x
p (x, 0) = p x, ly = 1 − , (70)
lx
and in the second solution we set
p (x, 0) = 1 , (71)
 
p x, ly = 0 , (72)
y
p (0, y) = p (l x , y) = 1 − . (73)
ly
12
Figure 2

n2 ∂D2
ly (lx, ly)

∂D4

n3 n4

∂D3

x
O lx
∂D1 n1

Figure 1: Schematic of a rectangular local solution domain with explicit fractures, the blue triangle represents the pressure
boundary for the first solution and the green triangle represents the pressure boundary for the second solution.

From these two solutions, we can compute total flow rates through the faces of the region, and we denote
st st nd nd
these flow rates as Q1∂D4 , Q1∂D2 , Q2∂D4 and Q2∂D2 where the subscript represents the face and the superscript
represents the solution number. Then, the equivalent permeability k2 can be computed by equating the total
flow rate from the local fine scale solution with the flow rate that would result in a double porosity region
subject to the same boundary conditions. This gives
st st nd 
 µ f lx Q1∂D4 µ f Q1∂D +µ f Q2∂D 

2 4
− k1
k2 =  1st 2  .
 ly 
nd nd (74)
 µ f Q∂D2 +µ f Q∂D4 µ f ly Q∂D2
2 2 
2 lx − k1
Note k1 can be estimated from the geometric mean of original local fine scale matrix permeabilities.
A layered system allows us to evaluate Eq. (74) analytically, see Fig. 2. Here the layered system means
all fractures (from 1 to n) are oriented in the horizontal direction, with possibly different apertures w f ,
and we evaluate the permeability of each fracture as w2f /12 according to the cubic law [81]. In [35], they
mentioned a large w f is controlled by the highly permeable clay-free layers and a small w f is due to the
shale content of less permeable silty layers. We further assume the matrix properties are isotropic and
layered
homogeneous. The analytical result of k2 is given as a diagonal matrix
 P w3  P
  
  n f i −k1 n w f i 
i=1 12 i=1
layered
 
k2 =  H  . (75)
 P H H−Pn w − k1 
f i
i=1 w f i
n
+
12
k1
i=1

In Eq. (75), the (1,1) component is always larger than the (2,2) component because from basic algebra, we
know that the harmonic mean is less or equal to the arithmetic mean. This result is consistent with the
13
Figure 2

wfn

……
H Matrix k1

wf 2
wf 1
x
O

Figure 2: Schematic of a layered system, we have n horizontal fractures with apertures w f 1 to w f n .

findings of kV /kH in [4, 101] and R in [35]. If above layered system is rotated through an angle θ, we can
approximate the new k#2 as [26, 62, 88, 89, 97]

cos θ − sin θ layered cos θ sin θ


" # " #
#
k2 ≈ k . (76)
sin θ cos θ 2 − sin θ cos θ

Alternative methods are also available in the literature. For instance, in [101], they use Oda’s method to
calculate equivalent hydraulic conductivity, which is prevalent in in-situ tests.

5. Numerical example

In this paper, we use a specific u/p1 /p2 finite element (FE) formulation that satisfies LBB condition
[20, 32, 111] to solve Eqs. (60)(61)(63) numerically. Although this discretization scheme cannot guarantee
element-wise mass conservation compared with finite volume (FV) formulation, mixed p/q formulation,
discontinuous Galerkin method and enriched Galerkin method [31, 33], it satisfies dual-grid mass conser-
vation and it is still applicable to a homogeneous system or the system with mild heterogeneity to bene-
fit from more desirable accuracy. Besides, the finite element method deals with full tensor permeability
very efficiently, which has an advantage over messy expressions in multi-point flux approximation (MPFA)
[88, 89, 97]. The derivations of residual vectors R (X) and tangent operators dR/dX are similar to those in
[32, 124] except for the explicit treatment of k̄ due to the sharp edge (non-differentiability) in the surface of
ϕmax as a function of ϕ (see Fig. 3 for the illustration), i.e., k̄ = k̄tn is calculated from ϕtn = ∇pt1n − ρ f g.
The code used for this study, Geocentric, relies heavily on the deal.II Finite Element Library [11] for
discretization functionality, p4est mesh handling library [23] and the Trilinos project [55]. In addition, due
to the large size of the linear system stemming from a three-field formulation, we have utilized the 3 × 3
block preconditioned Newton-Krylov solver built into the code, see [32, 112]. For code verification, we use
the third example in [24] which is strip load on infinite half space truncated at L. Here, we slightly modify
14
Figure 3: A simple illustration of the sharp edge in the surface of ϕmax as a function of ϕ = (ϕ1 , ϕ2 ) in 2D.

the original drainage boundary condition Eqs. (A.7e)(A.7f) in [24] to be p = 0 on Γt2 × I and −q · n = 0 on
(Γt1 ∪ Γl ∪ Γb ∪ Γr ) × I, respectively. The parameters are shown in Table 1 A and we compare result of our
simplified model with that of ANSYS, see Fig. 4. Note in Fig. 4, the dimensionless time T is defined as
E (1 − ν) kt
T= . (77)
(1 + ν) (1 − 2ν) µ f a2

Figure 4: Result comparison of ANSYS and Geocentric. We choose to plot dimensionless pressure p/w and dimensionless
settlement Guy /w/a vs. T at 1 m below the center of the strip load. The two isolated points on the left represent initial values after
normal consolidation which imply instantaneous changes in the pressure and displacement fields [24]. A perfect match is achieved
here.

A 3D synthetic model is generated to contain superimposed porosity domains and horizontal planes
of isotropy. The configuration of the problem and boundary (loading) conditions are shown in Fig. 5.
A
Isotropic elasticity. Darcy’s law is applied to calculate q1 = q and since φ2 = 0, k2 , k̄ and σsh will not be used.

15
Table 1: Parameter table of the verification example.

Parameter Value Unit


Gravity acceleration constant g 9.81 m/s2
Model size L [24] 5 m
Half strip load length a [24] 1 m
External mechanical load w [24] 150 kPa
Unjacketed bulk modulus K s ∞ kPa
Solid grain density ρ s 2600 kg/m3
Young’s modulus E 6300 kPa
Poisson’s ratio ν 0.15 -
Porosity φ1 0.1 -
Porosity φ2 0 -
Permeability k1 = k 10−10 m2
Viscosity µ f 1 cP
Fluid bulk modulus K f ∞ kPa
Fluid density ρ f 1000 kg/m3

This model could be recast into the famous Cryer’s problem in the limit of single porosity and isotropy
[1, 32, 113]. The parameters are summarized in Table 2 and we have meshed this 3D synthetic model into
3456 hexahedron elements. The simulation begins with a time increment of 0.05 [s] and with subsequent
increments magnified by a factor of 1.125, i.e., ∆tn+1 = 1.125∆tn . The total number of simulation time steps
is 75.

5.1. Sensitivity analysis


Sensitivity analysis is conducted to investigate the impacts of k2V , k̄min and F on the hydromechanical
responses. Twelve analysis cases are specified in Table 3, which are comprised of three values for k2V , two
values of k̄min and two values of F.
The pressure responses are shown in Fig. 6 and Fig. 7. For every three sub-figures in each row, we could
conclude the effect of k2V . In this example, the decrease in k2V leads to a slight increase in p2 and almost no
change in p1 . This phenomenon seems to contradict with what we have observed in [124], but instead, if we
consider the combined effects of principal directions of anisotropy and drainage boundary, we can resolve
this contradiction. In Sec. 5.1 of [124], the rotation angle θ is π/18, which means the principal direction
with maximum eigenvalue is approximately perpendicular to lateral boundaries that are impermeable. Thus
in [124], the magnitude of k2V will control the pattern of p2 . In contrast, in this 3D example, the plane of
isotropy intersects with the drainage boundary, which leads to a horizontal preferential fluid flow direction
and the magnitude of k2V will not control the pattern of p2 . In the next section, the 3D pressure distribution
will further confirm our arguments.
Another interesting finding is the non-monotone characteristic of p2 from Case 1 to Case 3, which will
not appear when k̄ is a constant or the magnitude of F is small (Case 4 to Case 12), as shown in the second
row of Fig. 6 and the whole Fig. 7. This difference is because when F is small, the magnitude of ϕ is also

16
Figure 5: 3D compression problem of the 1/8 saturated poroelastic double porosity sphere. The gray horizontal planes indicate
a VTI medium [128]. Radial load F is applied on the spherical surface with rainbow color, which also serves as a drainage
boundary but only for fissures. The other three surfaces that coincide with coordinate planes are impermeable with zero normal
displacements. Gravity is not included in the analysis for this problem, i.e., excess form.

small, and from Eq. (24), a small ϕmax will lead to k̄ ≡ k̄min , so we would expect the same behaviors as
k̄max = k̄min , i.e, constant k̄. In one word, The non-monotone characteristic of p2 is related to a changing k̄
and a broad range of ϕmax compared with λmax .

5.2. Pressure distribution analysis


The 3D pressure distribution profiles on the deformed domain of Case 3 are depicted in Fig. 8. For
the first column, we can see that the early time response of p1 is dependent on the spatial distribution of
F since it is the fluid in the primary porosity that supports a large portion of the load in the beginning.
Note here in mixture poromechanics theory, the secondary porosity is all void space, which means the
underlying fracture has no intrinsic stiffness, so we cannot think of excess pressure accumulation in terms
of the relative magnitude of intrinsic stiffness. For the second column, we can see that a permeability
tensor with a high anisotropy ratio may not skew the pressure distribution, which confirms our arguments in
Sec. 5.1. By comparing these two columns, we may conclude that in this case, non-equilibrium/preferential
flow [47, 56, 99] appears as soon as when we apply the load F, and it becomes weaker at a later stage. In
the next section, we will analyze the reasons for non-equilibrium/preferential flow in detail.
From the last time step, since p1 ≈ p2 ≈ 0, we could obtain the largest amounts of compression in all
x = −0.071 [m], uy = −0.09792 [m] and uz = −0.1551 [m]. This
three directions, they are u∞ ∞ ∞
anisotropic

response is due to both the external load F and the mechanical properties. In other words, u ∞
x < u ∞
y is

because F is monotonically decreasing with x ∈ [0, 1] [m], and uy < uz is because of Ev < Eh .
∞ ∞

5.3. Impact of the drainage boundary on the secondary compression


In this section, we will assume F = 1 [MPa] to focus on the impact of the drainage boundary on the
secondary compression. We compare the pressure changes in origin for a fully drained boundary p2 = p̄ ≡ 0
17
Table 2: Parameter table of the synthetic model.

Parameter Value Unit


External mechanical load F Table 3 MPa or kPa
Unjacketed bulk modulus K s ∞ kPa
Solid grain density ρ s 2600 kg/m3
Young’s modulus Ev 5000 kPa
Young’s modulus Eh 7500 kPa
Poisson’s ratio νvh 0.25 -
Poisson’s ratio νhh 0.15 -
Shear modulus Gvh 3000 kPa
Biot coefficient α1 0.8 -
Biot coefficient α2 0.2 -
Permeability k1 10−15 m2
Critical gradient λmax [54, 127] 0.2306 MPa/m
Threshold gradient λmin [79] 0 MPa/m
Exponent parameter ξ [79, 127] 2 -
Equivalent permeability k2H 2.5 × 10−13 m2
Equivalent permeability k2V Table 3 m2
Maximum interface permeability k̄max 10−15 m2
Minimum interface permeability k̄min Table 3 m2
Shape factor σsh 2000 m−2
Viscosity µ f 1 cP
Fluid bulk modulus K f ∞ kPa
Fluid density ρ f 1000 kg/m3

and a time-dependent drainage boundary p2 = p̄ = exp (t/50) [MPa] on r = x2 + y2 + z2 = 1 [m]. Fig. 9


p

portrays the results. Since the secondary porosity always has a low storage capacity, its behavior will
be similar to that of the elliptic equation. Therefore, in Fig. 9a, the sudden pressure drop will promptly
propagate to the origin, and the non-equilibrium/preferential flow is immediately established. In contrast,
in Fig. 9b, since there is an inherent time scale in the boundary pressure, the primary compression and the
secondary compression will take place sequentially, i.e., a primary compression in which p2 decreases fast,
followed by a secondary compression in which p1 decreases faster while p2 decreases slower.

6. Conclusions

We have presented a comprehensive continuum framework for anisotropic and deformable porous ma-
terials exhibiting two dominant porosity scales with ultra-low matrix permeability. Through mathematical
formulations, we have identified challenges in modeling dφ1 /dt and dφ2 /dt. Then we revisited and reviewed

18
Table 3: Specification of analysis cases.

Case k2V k̄min F


1 0.5k2H 0.75k̄max 1 + 0.5 cos x [MPa]
2 0.1k2H 0.75k̄max 1 + 0.5 cos x [MPa]
3 0.02k2H 0.75k̄max 1 + 0.5 cos x [MPa]
4 0.5k2H k̄max 1 + 0.5 cos x [MPa]
5 0.1k2H k̄max 1 + 0.5 cos x [MPa]
6 0.02k2H k̄max 1 + 0.5 cos x [MPa]
7 0.5k2H 0.75k̄max 10 + 5 cos x [kPa]
8 0.1k2H 0.75k̄max 10 + 5 cos x [kPa]
9 0.02k2H 0.75k̄max 10 + 5 cos x [kPa]
10 0.5k2H k̄max 10 + 5 cos x [kPa]
11 0.1k2H k̄max 10 + 5 cos x [kPa]
12 0.02k2H k̄max 10 + 5 cos x [kPa]

several main modeling approaches, which differ in the formulas used for the calculation of constitutive co-
efficients Ai j , α1 and α2 . Fortunately, under reasonable ranges of material properties, we can drop higher-
order terms through scaling analysis and propose a simplified hydromechanical framework proven to be
valid for no violation of entropy inequality. A simple upscaling approach is proposed in this work to fill
the gap between discrete geological descriptions and double porosity flow parameters. The finite element
method is applied for the discretization of fluid flow and geomechanics in our framework. Finally, the 3D
numerical example demonstrates several facts:
1. A permeability tensor with a high anisotropy ratio may not skew the pressure distribution due to the
combined effects of the principal directions of anisotropy and the drainage boundary.
2. The non-monotone characteristic of p2 is related to a changing k̄ and a broad range of ϕmax compared
with λmax .
3. The early time response of p1 is dependent on the spatial distribution of F, while the late time re-
sponses of both scales tend to be the same.
4. The drainage boundary could significantly impact the instantaneous fluid pressure responses.
Incorporation of anisotropic plastic deformation [19] into our framework will be addressed in the companion
paper: Hydromechanical modeling of solid deformation and fluid flow in transversely isotropic rocks with
double porosity.

Acknowledgments

This material is based upon work supported by the U.S. Department of Energy, Office of Science, Office
of Basic Energy Sciences, Geosciences Research Program, under Award Number DE-FG02-03ER15454.
Support for materials and additional student hours were provided by the National Science Foundation under
Award Number CMMI-1462231.

19
(a) Case 1 (b) Case 2 (c) Case 3

(d) Case 4 (e) Case 5 (f) Case 6

Figure 6: Sensitivity analysis results under F = 1 + 0.5 cos x [MPa], the pressure is dimensionless.

Appendix A. Proof on the equivalence of derivatives in the mass balance equation

Here we will give a detailed proof of Eq. (14). By expanding the LHS, we will get
h i
φ1 dρ 1f ∇ · ρ 1 f q 1 φ1 dρ1 f q1 · ∇ρ1 f
+ = + ∇ · q1 + . (A.1)
ρ1 f dt ρ1 f ρ1 f dt ρ1 f

Next we replace q1 in the last term with its definition Eq. (10), we will have
h i
φ1 dρ1 f ∇ · ρ1 f q1 φ1 dρ1 f
" #
+ = + (v̂1 − v s ) · ∇ρ1 f + ∇ · q1 . (A.2)
ρ1 f dt ρ1 f ρ1 f dt

Now if we subtract Eq. (5) from Eq. (6), we will get

d1 f (·) d(·)
= + ∇ (·) · (v̂1 − v s ) . (A.3)
dt dt
Replacing (·) with ρ1 f and substituting Eq. (A.3) into Eq. (A.2) gives
h i
φ1 dρ1 f ∇ · ρ1 f q1 φ1 d1 f ρ1 f
+ = + ∇ · q1 , (A.4)
ρ1 f dt ρ1 f ρ1 f dt

and we finish the proof.

20
(a) Case 7 (b) Case 8 (c) Case 9

(d) Case 10 (e) Case 11 (f) Case 12

Figure 7: Sensitivity analysis results under F = 10 + 5 cos x [kPa], the pressure is dimensionless.

Appendix B. Constitutive law for the change of porosity in the single porosity media
For single porosity media, Eqs. (17)(18) will collapse into one equation (no mass source or sink)
φ df p dφ
+∇·q+ + φ∇ · v s = 0 , (B.1)
K f dt dt
where φ is the Eulerian porosity. To handle dφ/dt, following constitutive prototype is proposed
φ
" #
dφ dp
= Kb ∇ · v s + (β − b) − φ∇ · v s , (B.2)
dt Kp dt
where 1/K p is the jacketed pore compressibility, 1/Kb is the drained jacketed bulk compressibility, b =
1− Kb /K s is the bulk volume effective stress coefficient and 1/K s is known as the unjacketed bulk compress-
ibility, β is the pore volume effective stress coefficient. People have proven that 1/K p is not an independent
quantity, instead, 1/K p is related to Kb and K s through following equation [27, 28, 107]
1 b
= . (B.3)
K p φKb
Therefore, the only undetermined quantity is β. However, inconsistency still exists in the literature and
different choices of β will lead to slightly different governing equations but they must coincide with each
other when K s = ∞, i.e., Eq. (66). Here I will give two prevalent choices.
If we let β = b and use Eq. (B.3), Eq. (B.2) can be simplified to

= (b − φ) ∇ · v s , (B.4)
dt
21
Time p1 (kPa) p2 (kPa)
step

#2

# 20

# 30

# 40

# 50

# 55

Figure 8: 3D pressure distribution of Case 3.

and Eq. (B.1) will become


φ df p
+ ∇ · q + b∇ · v s = 0 . (B.5)
K f dt
Eq. (B.5) is widely used in Prof. Borja’s community [20].
There are some other researchers who use a different β = 1− K p /K s , see [30, 107]. Using this definition,
we can easily prove following identity with purely algebraic manipulations
φ b−φ
(β − b) = . (B.6)
Kp Ks

Using above identity, Eq. (B.2) can be simplified to


dφ b − φ dp
= (b − φ) ∇ · v s + , (B.7)
dt K s dt
22
(a) (b)

Figure 9: Pressure responses of the double porosity media for (a) a fully drained boundary and (b) a time-dependent drainage
boundary.

and we can rewrite Eq. (B.1) as


φ df p b − φ dp
+ ∇ · q + b∇ · v s + = 0. (B.8)
K f dt K s dt
Now if we make approximations that d/dt ≈ ∂/∂t and d f /dt ≈ ∂/∂t, Equation (B.8) could be written in a
more concise form
1 ∂p
+ ∇ · q + b∇ · v s = 0 , (B.9)
M ∂t
where
1 φ b−φ
= + (B.10)
M Kf Ks
is known as the constrained specific storage. Equation (B.9) is even more widely used in numerical mod-
eling. From the author’s point of view, for most porous media, the K s is always very large. Therefore both
Eq. (B.5) and Eq. (B.9) will give accurate results even though b is not exactly 1. However, for soft materials,
it would be better to compare the numerical simulation results with experimental data to determine which
equation to use.

Appendix C. Constitutive laws for the changes of porosities in the double porosity media
In double porosity media, it is evident that two bulk volume effective stress coefficients (b1 and b2 ), two
pore volume effective stress coefficients (β1 and β2 ) and two pore compressibilities (1/K p1 and 1/K p2 ) are
needed (some of them may not be used explicitly). All of these quantities should depend on 1/Kb , 1/K s
and b = 1 − Kb /K s of the double porosity media and 1/K fb , 1/K b = 1−K
fs and e fb /K
fs of the corresponding
non-fractured single porosity media with the same total porosity φ. Many choices are possible, and here I
will address three of them.
The first choice is analogous to that in Appendix B, i.e., we assume β1 = b1 , β2 = b2 and 1/K p1 =
1/K p2 = 1/K p = b/ (φKb ), as a result, we will have our first instance of Eq. (19)
dφ1
= (ψ1 b − φ1 ) ∇ · v s , (C.1)
dt
23
dφ2
= (ψ2 b − φ2 ) ∇ · v s , (C.2)
dt
where ψ1 = φ1 /φ and ψ2 = φ2 /φ are known as the pore fractions and they satisfy ψ1 + ψ2 = 1. Note in this
choice, we tend not to use 1/Kfb , 1/K
fs and e
b and the result is very concise.
The second choice is based on [27, 28] and we refer interested readers to look at these references to see
their theoretical arguments. The second instance of Eq. (19) is shown as follows
dφ1 β1 − b1 d p1 β2 − b2 d p2
= (b1 − φ1 ) ∇ · v s + + , (C.3)
dt K p1 dt K p1 dt
dφ2 β1 − b1 d p1 β2 − b2 d p2
= (b2 − φ2 ) ∇ · v s + + , (C.4)
dt K p2 dt K p2 dt
where
1 b
e
= , (C.5)
fp φK
K fb
Kp
β=1− , (C.6)
Ks
K
fp
β=1−
e , (C.7)
K
fs
bKb
e
b1 = , (C.8)
K
fb
b2 = b − b1 , (C.9)
βK p
e
β1 = , (C.10)
K
fp
β2 = β − β1 , (C.11)
1 φ
= , (C.12)
K p1 Kfp
1 φ 1
= − . (C.13)
K p2 K p K p1
We can see that this choice uses 1/K fb , 1/K
fs and eb quite a lot and the result is pretty messy.
The last choice is based on [66] and the third instance of Eq. (19) is given as follows

b2 − φ2 b1 − φ1 − b2 + φ2 d p1 b2 − φ2 b2 − φ2 d p2
" # " #
dφ1
= (b1 − φ1 ) ∇ · v s + + + − , (C.14)
dt K
fb Ks dt Ks K
fb dt

b2 − φ2 b2 − φ2 d p1 b2 − φ2 d p2
" #
dφ2
= (b2 − φ2 ) ∇ · v s + − + , (C.15)
dt Ks K
fb dt K
fb dt
where
Kb Kb
b1 = − , (C.16)
fb K s
K
Kb
b2 = 1 − . (C.17)
K
fb
24
You can easily establish the unique mapping from each of the three choices to A11 , A12 , A21 , A22 , α1 and
α2 in Eq. (19), respectively. Fig. C.10 compares these three choices under Kb = 5000 [kPa], φ1 = 0.2 and
φ2 = 0.05 of typical fissured clays. Fig. C.11 compares these three choices under Kb = 5 [GPa], φ1 = 0.1
and φ2 = 0.005 of typical fissured rocks. As you can see from these figures, the reciprocals of Ai j are a
few orders of magnitude larger than Kb , thus if we assume the characteristic quantity of the rate of the fluid
pressure change is approximately equal to the characteristic quantity of the rate of the mean effective stress
change, it is not a bad idea to ignore 2j=1 Ai j dp j /dt comapred with ∇ · v s and use the simplified equations
P
shown in Sec. 3.

(a) (b)

Figure C.10: Values of double porosity coefficients Ai j , α1 and α2 for different theoretical models [27, 28, 66] of typical fissured
clays, note for the first choice, Ai j ≡ 0 while α1 and α2 remain the same.

(a) (b)

Figure C.11: Values of double porosity coefficients Ai j , α1 and α2 for different theoretical models [27, 28, 66] of typical fissured
rocks, note for the first choice, Ai j ≡ 0 while α1 and α2 remain the same.

25
Appendix D. Scaling analysis

In this appendix, we will quantify the relative impacts of various terms through scaling analysis. To
achieve it, we need to define some characteristic quantities as follows.
For ∇p2 and p1 − p2 , we assume O (∇p2 ) = ∆P/L and O (p1 − p2 ) = ∆Pd where ∆P and ∆Pd represent
the characteristic pressure variations, L is the characteristic length. Thus O (∇p1 ) = ∆P/L + ∆Pd /L. For
α1 , α2 and porosities, we have O (α1 ) = O (α2 ) = O (φ1 ) = O (φ2 ) = 1, therefore, the force density term
α1 ∇p1 + α2 ∇p2 in Eq. (63) will have the order of 2∆P/L + ∆Pd /L. For  v̂1 and v̂2 , we quantify them
using Darcy’s law (drop the v s term), as a result, O (v̂1 ) = k1 (∆P + ∆Pd ) / µ f L and O (v̂2 ) = k2 ∆P/ µ f L
where k1 , k2 and µ f are the characteristic quantities. Finally, for c1 and c2 , we can use Eqs. (22)(23) to get
O (c1 ) = O (c2 ) = ρ f σsh k1 ∆Pd /µ f where ρ f and σsh are also the characteristic quantities and we further
adopt O k̄ = k1 .
To argue that we can drop c̄ and c1 v̂1 · v̂1 /2 in Eqs. (36)(65) safely, we will calculate two dimensionless
quantities
O (c1 v̂1 + c2 v̂2 )
ND = (D.1)
O (α1 ∇p1 + α2 ∇p2 )
and
O (c1 v̂1 · v̂1 ) µ f
NDD = . (D.2)
σsh k1 ∆P2d
h i h i h i
Here we adopt following values: k1 = 10−15 m2 , k2 = 10−10 m2 , σsh = 1000 m−2 , µ f = 1 [cP] and
h i
ρ f = 1000 kg/m3 . We plot ND and NDD as functions of ∆P and ∆Pd , both under the same reasonable
range from 1 [kPa] to 1 [GPa] (Reynolds number cannot be infinite). Since ND is independent of L while
NDD increases with the decrease of L, we choose the core scale to test our hypothesis, i.e., L = 0.1 [m].
The results are shown in Fig. D.12 which support our arguments and similar assertions can also be found in
[3, 71] (without mathematical proofs).

(a) (b)

Figure D.12: Scaling analysis results: (a) Quantify c̄ vs. α1 ∇p1 + α2 ∇p2 . (b) Quantify c1 v̂1 · v̂1 /2 vs. σsh k̄ (p1 − p2 )2 /µ f .

26
References
[1] Y. Abousleiman, A. H.-D. Cheng, L. Cui, E. Detournay, J. C. Roegiers, Mandel’s problem revisited, Géotechnique 46 (1996)
187–195.
[2] T. Arbogast, J. Jr. Douglas, U. Hornung, Derivation of the double porosity model of single phase flow via homogenization
theory, SIAM Journal on Mathematical Analysis 21 (1990) 823–836.
[3] M. Ashworth, F. Doster, Foundations and their practical implications for the constitutive coefficients of poromechanical
dual-continuum models, Transport in Porous Media (2019).
[4] C. Ayan, N. Colley, G. Cowan, et al., Measuring permeability anisotropy: The latest approach, Oilfield Review 6 (1994)
24–35.
[5] M. Bai, On equivalence of dual-porosity poroelastic parameters, Journal of Geophysical Research 104 (1999) 461–466.
[6] M. Bai, D. Elsworth, J. C. Roegiers, Modeling of naturally fractured reservoirs using deformation dependent flow mecha-
nism, International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts 30 (1993) 1185–1191.
[7] M. Bai, D. Elsworth, J. C. Roegiers, Multiporosity/multipermeability approach to the simulation of naturally fractured
reservoirs, Water Resources Research 29 (1993) 1621–1633.
[8] M. Bai, F. Meng, D. Elsworth, Y. Abousleiman, J. C. Roegiers, Numerical modeling of coupled flow and deformation in
fractured rock specimens, International Journal for Numerical and Analytical Methods in Geomechanics 23 (1999) 141–
160.
[9] M. Bai, J. C. Roegiers, Fluid flow and heat flow in deformable fractured porous media, International Journal of Engineering
Science 32 (1994) 1615–1633.
[10] M. Bai, J. C. Roegiers, D. Elsworth, Poromechanical response of fractured-porous rock masses, Journal of Petroleum
Science and Engineering 13 (1995) 155–168.
[11] W. Bangerth, R. Hartmann, G. Kanschat, Deal.II—A general-purpose object-oriented finite element library, ACM Transac-
tions on Mathematical Software 33 (2007).
[12] G. I. Barenblatt, Iu. P. Zheltov, I. N. Kochina, Basic concepts in the theory of seepage of homogeneous liquids in fissured
rocks, Journal of applied mathematics and mechanics 24 (1960) 852–864.
[13] I. Berre, F. Doster, E. Keilegavlen, Flow in fractured porous media: A review of conceptual models and discretization
approaches, Transport in Porous Media 130 (2019) 215–236.
[14] J. G. Berryman, Extension of poroelastic analysis to double-porosity materials: new technique in microgeomechanics,
Journal of Engineering Mechanics 128 (2002) 840–847.
[15] J. G. Berryman, S. R. Pride, Models for computing geomechanical constants of double-porosity materials from the con-
stituents’ properties, Journal of Geophysical Research 107 (2002) 1–15.
[16] J. G. Berryman, H. F. Wang, The elastic coefficients of double porosity models for fluid transport in jointed rock, Journal of
Geophysical Research 100 (1995) 611–627.
[17] R. I. Borja, Cam-Clay plasticity, Part V: A mathematical framework for three-phase deformation and strain localization
analyses of partially saturated porous media, Computer Methods in Applied Mechanics and Engineering 193 (2004) 5301–
5338.
[18] R. I. Borja, On the mechanical energy and effective stress in saturated and unsaturated porous continua, International
Journal of Solids and Structures 43 (2006) 1764–1786.
[19] R. I. Borja, Plasticity, Springer (2013).
[20] R. I. Borja, Computational poromechanics (Stanford University CEE 294) (2018).
[21] R. I. Borja, J. Choo, Cam-Clay plasticity, Part VIII: A constitutive framework for porous materials with evolving internal
structure, Computer Methods in Applied Mechanics and Engineering 309 (2016) 653–679.
[22] R. I. Borja, A. Koliji, On the effective stress in unsaturated porous continua with double porosity, Journal of the Mechanics
and Physics of Solids 57 (2009) 1182–1193.
[23] C. Burstedde, L. C. Wilcox, O. Ghattas, p4est: Scalable algorithms for parallel adaptive mesh refinement on forests of
octrees, SIAM Journal on Scientific Computing 33 (2011) 1103–1133.
[24] N. Castelletto, J. A. White, H. A. Tchelepi, Accuracy and convergence properties of the fixed-stress iterative solution of
two-way coupled poromechanics, International Journal for Numerical and Analytical Methods in Geomechanics 39 (2015)
1593–1618.
[25] M. M. Chang, Deriving the shape factor of a fractured rock matrix, Topical Report, National Institute for Petroleum and
Energy Research, Bartlesville, OK, USA (1993).
[26] T. Chen, C. Clauser, G. Marquart, K. Willbrand, D. Mottaghy, A new upscaling method for fractured porous media, Ad-
vances in Water Resources 80 (2015) 60–68.
[27] H. Y. Chen, L. W. Teufel, Coupling fluid flow and geomechanics in dual porosity modeling of naturally fractured reservoirs,
SPE Annual Technical Conference and Exhibition (1997).

27
[28] H. Y. Chen, L. W. Teufel, Coupling fluid flow and geomechanics in dual porosity modeling of naturally fractured reservoirs:
model description and comparison, SPE International Petroleum Conference and Exhibition in Mexico (2000).
[29] A. H.-D. Cheng, Material coefficients of anisotropic poroelasticity, International Journal of Rock Mechanics and Mining
Sciences 34 (1997) 199–205.
[30] A. H.-D. Cheng, Poroelasticity, Springer (2016).
[31] J. Choo, Large deformation poromechanics with local mass conservation: An enriched Galerkin finite element framework,
International Journal for Numerical Methods in Engineering 116 (2018) 66–90.
[32] J. Choo, R. I. Borja, Stabilized mixed finite elements for deformable porous media with double porosity, Computer Methods
in Applied Mechanics and Engineering 293 (2015) 131–154.
[33] J. Choo, S. Lee, Enriched Galerkin finite elements for coupled poromechanics with local mass conservation, Computer
Methods in Applied Mechanics and Engineering 341 (2018) 311-332.
[34] J. Choo, J. A. White, R. I. Borja, Hydromechanical modeling of unsaturated flow in double porosity media, International
Journal of Geomechanics 16 (2016) D4016002.
[35] J. B. Clavaud, A. Maineult, M. Zamora, P. Rasolofosaon, C. Schlitter, Permeability anisotropy and its relations with porous
medium structure, Journal of Geophysical Research 113 (2008) 1–10.
[36] O. Coussy, Poromechanics, John Wiley and Sons (2004).
[37] C. W. Cryer, A comparison of the three-dimensional consolidation theories of biot and terzaghi, The Quarterly Journal of
Mechanics and Applied Mathematics 16 (1963) 401–412.
[38] N. M. Dmitriyev, V. M. Maksimov, Non-linear laws of fluid flow through anisotropic porous media, Journal of Applied
Mathematics and Mechanics 65 (2001) 935–940.
[39] Z. Dong, W. Li, G. Lei, H. Wang, C. Wang, Embedded discrete fracture modeling as a method to upscale permeability for
fractured reservoirs, Energies 12 (2019) 1–15.
[40] L. J. Durlofsky, Upscaling and gridding of fine scale geological models for flow simulation, International Forum on Reser-
voir Simulation (2005).
[41] R. C. Dykhuizen, A new coupling term for dual-porosity models, Water Resources Research 26 (1990) 351–356.
[42] D. Elsworth, M. Bai, Flow-deformation response of dual-porosity media, Journal of Geotechnical Engineering 118 (1992)
107–124.
[43] B. Flemisch, I. Berre, W. Boon, A. Fumagalli, N. Schwenck, A. Scotti, I. Stefansson, A. Tatomir, Benchmarks for single-
phase flow in fractured porous media, Advances in Water Resources 111 (2018) 239–258.
[44] X. Gai, A coupled geomechanics and reservoir flow model on parallel computers, Ph.D. thesis, The University of Texas at
Austin (2004).
[45] T. T. Garipov, M. Karimi-Fard, H. A. Tchelepi, Discrete fracture model for coupled flow and geomechanics, Computational
Geosciences 20 (2016) 149–160.
[46] S. Geiger, M. Dentz, I. Neuweiler, A novel multi-rate dual-porosity model for improved simulation of fractured and multi-
porosity reservoirs, SPE Journal 18 (2013) 670–684.
[47] H. H. Gerke, Preferential flow descriptions for structured soils, Journal of Plant Nutrition and Soil Science 169 (2006)
382–400.
[48] H. H. Gerke, M. T. van Genuchten, A dual-porosity model for simulating the preferential movement of water and solutes in
structured porous media, Water Resources Research 29 (1993) 305–319.
[49] H. H. Gerke, M. T. van Genuchten, Evaluation of a first-order water transfer term for variably saturated dual-porosity flow
models, Water Resources Research 29 (1993) 1225–1238.
[50] Hamid R. Ghafouri, R. W. Lewis, A finite element double porosity model for heterogeneous deformable porous media,
International Journal for Numerical and Analytical Methods in Geomechanics 20 (1996) 831–844.
[51] B. Gong, M. Karimi-Fard, L. J. Durlofsky, An upscaling procedure for constructing generalized dual-porosity/dual-
permeability models from discrete fracture characterizations, SPE Annual Technical Conference and Exhibition (2006).
[52] R. Haggerty, S. M. Gorelick, Multiple-rate mass transfer for modeling diffusion and surface reactions in media with pore-
scale heterogeneity, Water Resources Research 31 (1995) 2383–2400.
[53] S. Hansbo, Aspects of vertical drain design: Darcian or non-Darcian flow, Géotechnique 47 (1997) 983–992.
[54] F. Hao, L. S. Cheng, O. Hassan, J. Hou, C. Z. Liu, J. D. Feng, Threshold pressure gradient in ultra-low permeability
reservoirs, Petroleum Science and Technology 26 (2008) 1024–1035.
[55] M. A. Heroux, J. M. Willenbring, A new overview of the Trilinos project, Scientific Programming 20 (2012) 83–88.
[56] N. J. Jarvis, A review of non-equilibrium water flow and solute transport in soil macropores: principles, controlling factors
and consequences for water quality, European Journal of Soil Science 58 (2007) 523–546.
[57] J. Jiang, J. Yang, Coupled fluid flow and geomechanics modeling of stress-sensitive production behavior in fractured shale
gas reservoirs, International Journal of Rock Mechanics and Mining Sciences 101 (2018) 1–12.
[58] J. Jiang, R. M. Younis, Hybrid coupled discrete fracture/matrix and multicontinuum models for unconventional reservoir

28
simulation, SPE Journal 21 (2016) 1009–1027.
[59] M. Karimi-Fard, L. J. Durlofsky, A general gridding, discretization, and coarsening methodology for modeling flow in
porous formations with discrete geological features, Advances in Water Resources 96 (2016) 354–372.
[60] M. Karimi-Fard, L. J. Durlofsky, K. Aziz, An efficient discrete fracture model applicable for general purpose reservoir
simulators, SPE Reservoir Simulation Symposium (2003).
[61] M. Karimi-Fard, B. Gong, L. J. Durlofsky, Generation of coarse-scale continuum flow models from detailed fracture char-
acterizations, Water Resources Research 42 (2006) 1–13.
[62] T. Karmakar, G. P. Raja Sekhar, Effect of anisotropic permeability on fluid flow through composite porous channel, Journal
of Engineering Mathematics 100 (2016) 33–51.
[63] H. Kazemi, L. S. Merrill, K. L. Porterfield, P. R. Zeman, Numerical simulation of water-oil flow in naturally fractured
reservoirs, SPE Journal 16 (1976) 317–326.
[64] N. Khalili, Coupling effects in double porosity media with deformable matrix, Geophysical Research Letters 30 (2003) 1–3.
[65] N. Khalili, Two-phase fluid flow through fractured porous media with deformable matrix, Water Resources Research 44
(2008) 1–12.
[66] N. Khalili, A. P. S. Selvadurai, A fully coupled constitutive model for thermo-hydro-mechanical analysis in elastic media
with double porosity, Geophysical Research Letters 30 (2003) 1–5.
[67] N. Khalili, S. Valliappan, Flow through fissured porous media with deformable matrix: Implicit Formulation, Water Re-
sources Research 27 (1991) 1703–1709.
[68] N. Khalili, S. Valliappan, C. F. Wan, Consolidation of fissured clays, Géotechnique 49 (1999) 75–89.
[69] N. Khalili, R. Witt, L. Laloui, L. Vulliet, A. Koliji, Effective stress in double porosity media with two immiscible fluids,
Geophysical Research Letters 32 (2005) 1–5.
[70] J. Kim, Sequential methods for coupled geomechanics and multiphase flow, Ph.D. thesis, Stanford University (2010).
[71] J. Kim, E. L. Sonnenthal, J. Rutqvist, Formulation and sequential numerical algorithms of coupled fluid/heat flow and
geomechanics for multiple porosity materials, International Journal for Numerical Methods in Engineering 92 (2012) 425–
456.
[72] J. M. Köhne, Binayak P. Mohanty, Jirka Simunek, H. H. Gerke, Numerical evaluation of a second-order water transfer term
for variably saturated dual-permeability models, Water Resources Research 40 (2004) 1–14.
[73] Q. Lei, W. Xiong, J. Yuan, S. Gao, Y. S. Wu, Behavior of flow through low-permeability reservoirs, SPE Europec/EAGE
Annual Conference and Exhibition (2008).
[74] P. Lemonnier, B. Bourbiaux, Simulation of naturally fractured reservoirs. State of the art. Part 2 matrix-fracture transfers
and typical features of numerical studies, Oil & Gas Science and Technology 65 (2010) 263–286.
[75] K. T. Lewallen, H. F. Wang, Consolidation of a double-porosity medium, International Journal of Solids and Structures 35
(1998) 4845–4867.
[76] R. W. Lewis, Hamid R. Ghafouri, A novel finite double porosity model for multiphase flow through deformable fractured
porous media, International Journal for Numerical and Analytical Methods in Geomechanics 21 (1997) 789–816.
[77] R. W. Lewis, W. K. S. Pao, Numerical simulation of three phase flow in deforming fractured reservoirs, Oil & Gas Science
and Technology 57 (2002) 499–514.
[78] J. Li, Z. Lei, G. Qin, B. Gong, Effective local-global upscaling of fractured reservoirs under discrete fractured discretization,
Energies 8 (2015) 10178–10197.
[79] D. Li, W. Zha, S. Liu, L. Wang, D. Lu, Pressure transient analysis of low permeability reservoir with pseudo threshold
pressure gradient, Journal of Petroleum Science and Engineering 147 (2016) 308–316.
[80] B. Loret, N. Khalili, A three phase model for unsaturated soils, International Journal for Numerical and Analytical Methods
in Geomechanics 24 (2000) 893–927.
[81] J. Luo, K. Wang, H. Liu, Z. Chen, Coupled geomechanics and fluid flow modeling in naturally fractured reservoirs, SPE
Low Perm Symposium (2016).
[82] J. Ma, G. Zhao, N. Khalili, A fully coupled flow deformation model for elasto-plastic damage analysis in saturated fractured
porous media, International Journal of Plasticity 76 (2016) 29–50.
[83] I. Masters, W. K. S. Pao, R. W. Lewis, Coupling temperature to a double-porosity model of deformable porous media,
International Journal of Numerical Methods in Engineering 49 (2000) 421–438.
[84] A. Mehrabian, Y. N. Abousleiman, Generalized biot’s theory and mandel’s problem of multiple-porosity and multiple-
permeability poroelasticity, Journal of Geophysical Research: Solid Earth 119 (2014) 2745–2763.
[85] A. Mehrabian, Y. N. Abousleiman, Gassmann equations and the constitutive relations for multiple-porosity and multiple-
permeability poroelasticity with applications to oil and gas shale, International Journal for Numerical and Analytical Meth-
ods in Geomechanics 39 (2015) 1547–1569.
[86] A. Mehrabian, Y. N. Abousleiman, Theory and analytical solution to Cryer’s problem of N porosity and N permeability
poroelasticity, Journal of the Mechanics and Physics of Solids 118 (2018) 218–227.

29
[87] A. Moinfar, A. Varavei, K. Sepehrnoori, R. T. Johns, Development of coupled dual continuum and discrete fracture model
for the simulation of unconventional reservoirs, SPE Reservoir Simulation Symposium (2013).
[88] A. Negara, A. Salama, S. Sun, Multiphase flow simulation with gravity effect in anisotropic porous media using multi-point
flux approximation, Computers and Fluids 114 (2015) 66–74.
[89] A. Negara, A. Salama, S. Sun, M. Elgassier, Y. S. Wu, Numerical simulation of natural gas flow in anisotropic shale
reservoirs, Abu Dhabi International Petroleum Exhibition and Conference (2015).
[90] S. K. Ngien, N. A. Rahman, R. W. Lewis, K. Ahmad, Numerical modeling of multiphase immiscible flow in double porosity
featured groundwater systems, International Journal for Numerical and Analytical Methods in Geomechanics 36 (2012)
1330–1349.
[91] V. X. Nguyen, Y. N. Abousleiman, Poromechanics solutions to plane strain and axisymmetric mandel-type problems in
dual-porosity and dual-permeability medium, Journal of Applied Mechanics 77 (2010) 1–18.
[92] W. K. S. Pao, R. W. Lewis, Three-dimensional finite element simulation of three-phase flow in a deforming fissured reservoir,
Computer Methods in Applied Mechanics and Engineering 191 (2002) 2631–2659.
[93] A. Prada, F. Civan, Modification of Darcy’s law for the threshold pressure gradient, Journal of Petroleum Science and
Engineering 22 (1999) 237–240.
[94] K. Pruess, T. N. Narasimhan, A practical method for modeling fluid and heat flow in fractured porous media, SPE Journal
25 (1985) 14–26.
[95] E. R. Rangel-German, Anthony R. Kovscek, Time-dependent matrix-fracture shape factors for partially and completely
immersed fractures, Journal of Petroleum Science and Engineering 54 (2006) 149–163.
[96] E. Ranjbar, H. Hassanzadeh, MatrixâĂŞfracture transfer shape factor for modeling flow of a compressible fluid in dual-
porosity media, Advances in Water Resources 34 (2011) 627–639.
[97] A. Salama, S. Sun, M. F. Wheeler, Solving global problem by considering multitude of local problems: Application to
fluid flow in anisotropic porous media using the multi-point flux approximation, Journal of Computational and Applied
Mathematics 267 (2014) 117–130.
[98] P. Sarma, K. Aziz, New transfer functions for simulation of naturally fractured reservoirs with dual-porosity models, SPE
Journal 11 (2006) 328–240.
[99] J. Simunek, N. J. Jarvis, M. T. van Genuchten, A. Gärdenäs, Review and comparison of models for describing non-
equilibrium and preferential flow and transport in the vadose zone, Journal of Hydrology 272 (2003) 14–35.
[100] S. Su, O. Gosselin, H. Parvizi, M. A. Giddins, Dynamic matrix-fracture transfer behavior in dual-porosity models, EAGE
Annual Conference & Exhibition incorporating SPE Europec (2013).
[101] J. Sun, Z. Zhao, Effects of anisotropic permeability of fractured rock masses on underground oil storage caverns, Tunnelling
and Underground Space Technology 25 (2010) 629–637.
[102] D. Swartzendruber, Non-Darcy flow behavior in liquid-saturated porous media, Journal of Geophysical Research 67 (1962)
5205–5213.
[103] Y. Tao, B. Guo, K. W. Bandilla, M. A. Celia, Vertically integrated dual-continuum models for CO2 injection in fractured
geological formations, Computational Geosciences (2019) 1–12.
[104] W. Tian, A. Li, X. Ren, Yapcheptoyek Josephine, The threshold pressure gradient effect in the tight sandstone gas reservoirs
with high water saturation, Fuel 226 (2018) 221–229.
[105] S. Vallippan, N. Khalili, Flow through fissured porous media with deformable matrix, International Journal for Numerical
Methods in Engineering 29 (1990) 1079–1994.
[106] M. N. Vu, A. Pouya, D. M. Seyedi, Theoretical and numerical study of the steady-state flow through finite fractured porous
media, International Journal for Numerical and Analytical Methods in Geomechanics 38 (2014) 221–235.
[107] H. F. Wang, Theory of linear poroelasticity with applications to geomechanics and hydrogeology, Princeton University Press
(2000).
[108] X. Wang, J. J. Sheng, Effect of low-velocity non-Darcy flow on well production performance in shale and tight oil reservoirs,
Fuel 190 (2017) 41–46.
[109] X. Wang, Z. Yang, Y. Sun, X. Liu, Experimental and theoretical investigation of nonlinear flow in low permeability reservoir,
Procedia Environmental Sciences 11 (2011) 1392–1399.
[110] J. E. Warren, P. J. Root, The Behavior of Naturally Fractured Reservoirs, SPE Journal 3 (1963) 245–255.
[111] J. A. White, R. I. Borja, Stabilized low-order finite elements for coupled solid deformation/fluid diffusion and their applica-
tion to fault zone transients, Computer Methods in Applied Mechanics and Engineering 197 (2008) 4353–4366.
[112] J. A. White, R. I. Borja, Block-preconditioned Newton-Krylov solvers for fully coupled flow and geomechanics, Computa-
tional Geosciences 15 (2011) 647–659.
[113] J. A. White, N. Castelletto, H. A. Tchelepi, Block-partitioned solvers for coupled poromechanics: A unified framework,
Computer Methods in Applied Mechanics and Engineering 303 (2016) 55–74.
[114] R. K. Wilson, Elias C. Aifantis, On the theory of consolidation with double porosity, International Journal of Engineering

30
Science 20 (1982) 1009–1035.
[115] Y. S. Wu, Numerical simulation of single-phase and multiphase non-Darcy flow in porous and fractured reservoirs, Transport
in Porous Media 49 (2002) 209–240.
[116] Y. S. Wu, J. Li, D. Ding, C. Wang, Y. Di, A generalized framework model for the simulation of gas production in unconven-
tional gas reservoirs, SPE Journal 19 (2014) 845–857.
[117] Y. S. Wu, K. Pruess, A multiple-porosity method for simulation of naturally fractured reservoirs, SPE Reservoir Engineering
(1988) 327–336.
[118] W. Xiong, Q. Lei, S. Gao, Z. Hu, H. Xue, Pseudo threshold pressure gradient to flow for low permeability reservoirs,
Petroleum Exploration and Development 36 (2009) 232–236.
[119] Y. Xiong, J. Yu, H. Sun, J. Yuan, Z. Huang, Y. S. Wu, A new non-Darcy flow model for low-velocity multiphase flow in
tight reservoirs, Transport in Porous Media 117 (2017) 367–383.
[120] X. Yan, Z. Huang, J. Yao, Y. Li, D. Fan, An efficient embedded discrete fracture model based on mimetic finite difference
method, Journal of Petroleum Science and Engineering 145 (2016) 11–21.
[121] X. Yan, Z. Huang, J. Yao, Y. Li, D. Fan, H. Sun, K. Zhang, An efficient numerical hybrid model for multiphase flow in
deformable fractured-shale reservoirs, SPE Journal 23 (2018) 1–26.
[122] X. Yan, Z. Huang, J. Yao, Y. Li, D. Fan, K. Zhang, An efficient hydro-mechanical model for coupled multi-porosity and
discrete fracture porous media, Computational Mechanics 62 (2018) 943–962.
[123] B. Zeng, L. Cheng, C. Li, Low velocity non-linear flow in ultra-low permeability reservoir, Journal of Petroleum Science
and Engineering 80 (2011) 1–6.
[124] Q. Zhang, J. Choo, R. I. Borja, On the preferential flow patterns induced by transverse isotropy and non-Darcy flow in
double porosity media, Computer Methods in Applied Mechanics and Engineering 353 (2019) 570–592.
[125] J. Zhang, J. C. Roegiers, Double porosity finite element method for borehole modeling, Rock Mechanics and Rock Engi-
neering 38 (2005) 217–242.
[126] Q. Zhang, H. Zhu, Collaborative 3D geological modeling analysis based on multi-source data standard, Engineering Geol-
ogy 246 (2018) 233–244.
[127] Y. Zhang, D. Zhang, Q. Fang, L. Xiong, L. Yu, M. Zhou, Analytical solutions of non-Darcy seepage of grouted subsea
tunnels, Tunneling and Underground Space Technology 96 (2020).
[128] R. Villamor Lora, E. Ghazanfari, E. Asanza Izquierdo, Geomechanical Characterization of Marcellus Shale, Rock Mechanics
and Rock Engineering 49 (2016) 3403–3424.

31

Das könnte Ihnen auch gefallen