You are on page 1of 21

International Journal of Rock Mechanics & Mining Sciences 175 (2024) 105676

Contents lists available at ScienceDirect

International Journal of Rock Mechanics and Mining Sciences


journal homepage: www.elsevier.com/locate/ijrmms

A multiscale poroelastic damage model for fracturing in permeable rocks


Jianxiong Yang a, Jianfeng Liu a, *, Wenfeng Li b, Jingjing Dai a, Fujun Xue a, Xiaoying Zhuang c, d
a
State Key Laboratory of Hydraulics and Mountain River Engineering, College of Water Resource & Hydropower, Sichuan University, Chengdu, 610042, China
b
Earth and Environmental Sciences, Los Alamos National Laboratory, Los Alamos, NM 87545, United States
c
Institute of Photonics, Faculty of Mathematics and Physics, Leibniz University Hannover, Hannover, 30167, Germany
d
Department of Geotechnical Engineering, College of Civil Engineering, Tongji University, Shanghai, 200092, China

A R T I C L E I N F O A B S T R A C T

Keywords: A new poroelastic damage model is developed in the paper to describe the macroscopic failure of rock materials
Multiscale fracturing due to microcrack nucleation and propagation based on a multiscale framework. The model is deduced from
Heterogeneous rock locally periodic microstructure with dynamically evolved microcracks in heterogeneous rock body. The ho­
Microcrack
mogenization method based on asymptotic expansions gives rise to the damage evolution law coupled with the
Damage law
Homogenization
poroelastic fracture system, which includes the fracture opening induced permeability change. The obtained
model takes into account the complex coupling between fluid pressure-deformation and hydro-mechanical (HM)
properties at the microscale, leading to the nonlinear anisotropic mechanical behavior, degradation of both
elastic stiffness and poroelastic properties at the macroscale, which is fundamental to describe the complex
fracturing behavior influenced by microcrack distribution. The homogenized coefficients are illustrated in detail
for a given set of initial material parameters, with dependence on the normalized damage variable. Results of
numerical simulations well reproduce specific experimental observations where fracturing in heterogeneous
rocks is shown to be a multiscale phenomenon that initiates from the microcrack nucleation and propagation,
while the fracture propagation direction is shown to be influenced by both external loadings and microcrack
distribution. The easy implementation in finite element framework and revealed micro-mechanism for macro­
scopic failure under strict mathematical formulations make the wide application of model in rock mechanics
problems.

1. Introduction between microcracks and elongated large-size pores is challenging to be


identified.9 The fracturing process in porous rocks may be seen as a
In the poroelastic cemented materials such as rocks, the distribution damage phenomenon that initiates from the boundary of contained
and interaction of contained natural microcracks affect the strength, microcracks or pores.10 These microcracks or elongated pores, which are
failure and hydraulic characteristics of rocks at the macroscale. When open at the initial state, then evolve and coalesce when subjected to
the microcracks are saturated by fluids and interact with the solid grains, external loading. At a certain state, they gradually grow to be larger
complex coupling between damage and hydro-mechanical (HM) in­ defects at the macroscale that are visible to the naked eyes.11 The further
teractions is occurring at multiple scales.1,2 The coupling plays an evolution and coalescence of macrocracks form a dominant fracture
important role in the damage behavior of the materials with definite plane that leads to the material strength loss. An issue forms at how to
change in poroelastic properties and fluid flow condition, which have effectively describe such a multiscale damage process starting from
been validated in the experimental studies, as recorded in.3–7 Therefore, microcracking in heterogeneous continuum media as accurate as
understanding the basic mechanism between microcrack evolution and possible.
HM coupling is fundamental to describe the fracturing process in There are two general ways to modelling the microcracking process
permeable rocks. in rocks, i.e., discrete approach and continuum approach. The former
Natural rock materials contain opening microcracks and pores at one generally treats the heterogeneous rock as a representative
different scales that influence the HM properties such as the strength, elementary volume consisting of discrete parts, e.g., particles, blocks or
compressibility, permeability, etc.,.8 In most cases, the distinction springs, which are connected by bonding force.12,13 The individual

* Corresponding author. College of Water Resource & Hydropowerm Sichuan University, NO. 24, South Section 1, Yihuan Road, Chengdu, Sichuan, 610042, China.
E-mail address: liujf@scu.edu.cn (J. Liu).

https://doi.org/10.1016/j.ijrmms.2024.105676
Received 8 July 2023; Received in revised form 31 January 2024; Accepted 15 February 2024
Available online 24 February 2024
1365-1609/© 2024 Elsevier Ltd. All rights reserved.
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

stiffness and strength of each part comprise jointly the overall properties 2. Multiscale poroelastic model
of the rock, while the local failure between discrete parts represents the
microcracking process and as a result the macroscopic damage is grad­ 2.1. Problem statement
ually formed due to multiple failure events.14,15 In the significant efforts
made during the past decades, the discrete approach has proved its We consider a porous medium consisting of a large number of peri­
capability by reproducing the complex fracture propagation behavior, e. odically distributed microcracks, in a 2D plane strain framework.
g., the wing crack evolution by interacting with pre-existed microcracks Infinitesimal deformation and isothermal conditions are adopted. Ten­
in rocks,16 hydraulic fracturing (HF) in natural fractured rock mass.17 sile strain is positively counted.
Since the mechanical behavior and fluid flow are coupled at the The porous macrostructure and the set of microcracks are denoted by
microscale, the approach is able to describe pore fluid induced micro­ B and C , respectively, as seen in Fig. 1(a). Each crack is straight of
cracking behavior. However, the adopted microstructure and scale may length l and has a constant spacing lc , which is much smaller than the
not represent the real rock properties so that the model parameters may macro-structural size Lc . The microstructure containing one straight
not correspond to the measurable physical properties, which need to be crack can be scaled to a reference cell Y, which is considered as a
determined by curve fitting.18 In addition, the computational cost be­ representative elementary volume (REV). The single crack (denoted by
comes expensive when focusing on the macroscale problems. C Y ) contained in cell Y is oriented in an angle θ with respect to hori­
For the continuum approach, the heterogeneous rock is generally zontal direction. The considered REV is made of compressible solid
idealized to be continuous body. A simple treatment is incorporating the phase and saturated by incompressible fluid.
elastic model together with damage criterion and post-failure behavior, We begin with the Biot’s theory for the deformable porous
as to represent the microcracking process, some widely-adopted elasto- macrostructure35,36:
damage models are referred to.19,20 This kind of elasto-damage models
may become more sophisticated by considering the effect of micro­
∂σ ij
= 0, with σ ij = Cijkl exkl (u) − αpf δij (1)
cracks, the porosity change and the structural damage as internal vari­ ∂xj
ables, detailed works can be seen in.21–24 As another enrichment, the
∂qi k ∂pf
microcrack-based continuum model generally adopts the + φ̇ = 0, with qi = − (2)
∂xi μf ∂xi
semi-analytical method that evaluates the microcrack evolution and the
corresponding permeability change through fracture mechanics.6,25,26
φ̇ = αėxii (u) + βṗf (3)
To overcome the model limitations that can represent microcrack
induced anisotropy and progressive damage behavior, anisotropic
where σij is total stress tensor, u is displacement tensor, Cijkl is elastic
damage variables, strength softening laws and heterogeneous HM
stiffness tensor, exij is strain tensor with respect to x-axis, α is Biot co­
properties may be incorporated.27–30 However, as the above-mentioned
efficient, pf is fluid pressure in porous matrix, qi is Darcy’s velocity, φ is
models are partly or totally based on macroscopic phenomenological
considerations to reproduce experimental observations, the physical porosity, β is Biot modulus, δij is the Kronecker delta.
mechanisms describing microstructural damage are missing. Inside the saturated cracks, fluid yields free-flow condition in a di­
To address the above limitations and represent the microcracking rection parallel to crack. To ease the difficulty in describing the coupling
process in a more physical manner, a multiscale damage model is regions between porous-flow and free-flow, the following Stokes re­
introduced that considers poroelastic coupling and stiffness degradation lations are introduced:
anisotropically due to microcrack propagation. Compared to our pre­ ∇x ⋅ σb = 0, with σb = 2μf ex (vb ) − pb I; ∇x ⋅ vb = 0 (4)
vious work in,31 the current contribution is improved from the following
aspects: (1) degradation of poroelastic properties is taken into account at where σb , pb and vb are stress tensor, fluid pressure and velocity inside
the microscale, which is fundamental to represent the real physical microcracks, respectively; ex (vb ) is rate of deformation tensor. Noted
process; (2) microcracks can have a certain arbitrary orientation, which that the crack geometry is a thin straight line which induces significant
is more aligned with the microcrack nature; (3) self-balanced unit cell anisotropy of fluid flow, where flow only takes place along the crack
problems under complex conditions are enhanced. direction.
Herein, the asymptotic homogenization upscaling scheme by32 is On the boundary of crack face, the continuity of total stress and
adopted to describe the adjacent microcrack interaction under the local pressure is reached, while the fluid pressure in matrix and crack is
periodicity assumption. Starting with a fully coupled poroelastic frac­ equal37:
ture problem at the microscopic scale, we use a micromechanics-based
approach to obtain the poroelastic damage model, which considers the σ ⋅ N = σb ⋅N, and pf = pb (5)
complex coupling between fluid pressure, deformation and HM prop­
where N is the unit vector normal to crack face.
erties. As a result, the nonlinear anisotropic mechanical behavior,
For fluid flow condition on crack face, since parallel flow is assumed
degradation of both elastic stiffness and poroelastic properties are con­
along crack direction, the fluid velocity with respect to solid deforma­
structed at the macroscale, and the change of effective HM coefficients
tion is in an equilibrium state between matrix and cracks.37,38 While at
depends on the introduced damage variable. Apart from the change of
the direction normal to crack face, there is no flux taking place. There­
poro-mechanical behaviors due to microcracking, the change of fluid
fore, we have the following relations along and normal to crack face:
flow behavior is implemented through damage induced permeability
change. q ⋅ e = (vb − u̇)⋅e (6)
In the remainder of the paper, we formulate the multiscale poroe­
lastic model for solid body containing periodically distributed micro­ qi Ni = 0 (7)
cracks in Section 2. Then the damage evolution law and the
corresponding permeability change are proposed in Section 3. The in which e is unit vector along crack direction, as seen in Fig. 2.
dependence of homogenized coefficients on damage variable is analyzed
in Section 4. Numerical examples in view of demonstrating the robust­ 2.2. Separation of scales
ness of the multiscale damage model are presented in the last section.
In order to derive the macroscopic effective equations for the above
local problem, it is assumed that the introduced crack length and crack
spacing in Fig. 1 are significantly smaller than the macroscopic

2
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 1. (a) The macroscopic porous medium with length Lc consists of periodically distributed microcracks of length l, lc is length of micro-period. (b) Basic principles
of field f along x-axis using asymptotic expansions (up to second order). (c) Reference cell with rescaled crack length d × Lc (adapted from Dascalu et al.33;
Orgeas et al.34).

Fig. 2. Scaling of a micro-period to a reference cell (modified from39).

characteristic length. Such condition of scale separation is directly In the microstructure containing one straight crack that evolves sym­
correlated to a scale parameter ε, which represents the difference be­ metrically with respect to the center point, the normalized crack length
tween the microscale and macroscale, expressed as: can be defined between the crack length and microstructural length:
lc
ε= ≪1 (8) d=
l
(9)
Lc lc

3
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

where d is the normalized crack length at microscale, also the damage Introducing such expansions (18)–(20) in problem (13)–(17) yields
variable at macroscale. to a set of boundary value problems corresponding to different orders of
Since the heterogeneous medium is assumed to consist of periodi­ ε over the period Y. After the analysis of each problem (more details are
cally distributed microstructure, the material response also varies peri­ referred to Appendix A), we can obtain the macroscopic governing
odically over the macroscale domain. The damage evolution process at equations in the following form:
macroscale can be understood in a way that the microcrack length ∂ hm ( hm )
gradually propagates from small length to large length until completely Σ = 0, with Σhm ∗
ij = Cijkl exkl u − α∗ij phm (21)
∂xj ij f

partitions the microstructure.


∂Qhm ∂exij (uhm ) ∂phm kij∗ ∂phm
(22)
f f
i
+ γ ∗ij + β∗ = 0, with Qhm
i = −
2.3. Microscale description ∂xi ∂t ∂t μf ∂xj

Following the methodology developed in,34 we introduce in Eqs. (1)– where the superscript ’ hm’ denotes the homogenized medium, Σhm hm
ij , Qi ,
(5) the following microscopic coordinates: uhm , phm
f are the macroscopic total stress tensor, Darcy’s velocity vector,

yi =
xi
(10) displacement vector, and fluid pressure in the homogenized medium,
ε respectively. The homogenized coefficients, e.g., stiffness tensor C∗ijkl ,
where xi and yi are the macroscopic and microscopic coordinates, Biot coefficients α∗ij and γ∗ij , Biot modulus β∗ and intrinsic permeability k∗ij
respectively. Introducing different coordinates in the domain allows us are defined by:
to scale the micro-period to a reference cell, which contains the rescaled ( )

Cijkl = 〈Cijkl + Cijmn eymn ξkl 〉 (23)
crack, as seen in Fig. 2.
By means of the above coordinate transformation and the adopted
α∗ij = αδij + 〈(1 − α)Cijkl eykl (κ)〉 (24)
local periodicity condition, all the involved fields in Eqs. (1)–(4), i.e., σ ij ,
u, pf , qi , φ, σ bij , pb and vb depend on ε. To underline this dependence, they ( )
γ ∗ij = αδij − (1 − α)〈eykk ξij 〉 (25)
are denoted by σεij , uε , pεf , qεj , φε , pεb and vεb . Hence, the material properties
and responses are Y-periodic and the field f on microscale can be
β∗ = β + (1 − α)2 〈eyii (κ)〉 (26)
expressed as:
( ) ∂ζ j
f ε (xi ) ≅ f (xi , yi ) = f xi , yj + nYj n = 1, 2, … (11) kij∗ = kA∗ij with A∗ij = 〈δij + 〉 (27)
∂yi
where f ε (xi ) is the material property or response, the superscript ε de­
notes that the function f ε (xi ) depends on ε, which can be defined with where the symbol 〈 ⋅〉 stands for volume average. The proof of γ∗ij = α∗ij
respect to both macroscopic and microscopic coordinates in the can be found in.37
following form32: The Y-periodic vector ξpq is the solution of the cell problem:

df ε ∂f 1 ∂f ∂ ( )
= + (12) C e (ξpq ) = 0 in Y (28)
dxi ∂xi ε ∂yi ∂yj ijkl ykl

Therefore, it is possible to rewrite the considered macroscopic ( )


Cijkl δkp δlq + eykl (ξpq ) Nj = 0 on C Y (29)
problem (1)-(5), (7) at the microscale, and reorganize the formulations
as follows: And the Y-periodic vector κ is the solution of
ε
∂σ ij ∂ ( )
= 0, with σ εij = Cijkl exkl (uε ) − αpεf δij (13) C e (κ) = 0 in Y (30)
∂xj ∂yj ijkl ykl

∂qεi k ∂pεf Cijkl eykl (κ)Nj = Ni on C Y (31)


+ αėxii (uε ) + βṗεf = 0, with qεi = − (14)
∂xi μf ∂xi Note that the tensor ξ and vector κ are similar to the ones obtained for
)( microporous medium containing one vug size in,2 such a proof is
∇x ⋅ σεb = 0, with σεb = 2ε4 μf ex vεb − pεb I inside the crack (15) recalled here:
∫ kl ∫
σε ⋅ N = σεb ⋅N and pεf = pεb on crack face C Y (16) ∂ξi
dΩ = − Cijkl eyij (κ)dΩ (32)
∂yi
Y Y
qεi Ni = 0 normal to crack face C Y (17)
The Y-periodic vector ζ is the solution of
( )
2.4. Homogenized governing equations ∂ ∂ζ
δjm + m = 0 in Y and (33)
∂yj ∂yj
By following the asymptotic expansions, we look for the involved
fields uε , pεf and pεb in the form: ( )
∂ζm
δjm + Nj = 0 on C Y (34)
uε = u(0) (x, y, t) + εu(1) (x, y, t) + ε2 u(2) (x, y, t) + … (18) ∂yj

It is noted that the first-order macroscopic model presents the clas­


pεf = p(0) (19)
(1) 2 (2)
f (x, y, t) + εpf (x, y, t) + ε pf (x, y, t) + … sical symmetry and the poroelastic problem is fully coupled with
normalized crack length, such that α∗ij ∕= αδij , β∗ ∕
= β when damage oc­
pεb = p(0) (20)
(1) 2 (2)
b (x, y, t) + εpb (x, y, t) + ε pb (x, y, t) + … curs. The equality only holds for undamaged HM problem. The influence
of damage on homogenized poroelastic coefficients will be further
where u(i) (x,y,t), pf (x,y,t), pb (x,y,t), x ∈ B , y ∈ Y are smooth functions
(i) (i)
investigated in Section 5.
and Y-periodic in y.

4
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

3. Damage equations macroscopic damage energy per unit volume.


Substitution of Eqs. (37) and (38) in the subcritical law (35) and
To fulfill the poroelastic fracture problem with evolving microcracks, considering both the damage variable (9) and normalized damage var­
the macroscopic equations need to be supplemented with damage evo­ iable d∗ = d × max(|cos θ|, |sin θ|), the following macroscopic time-
lution law. Besides, the fracture opening induced permeability change dependent damage evolution law can be defined:
needs to be incorporated in the homogenized permeability equation (27) (√̅̅̅̅̅̅̅̅̅̅̅̅̅ / )n
to represent the transition of fluid flow process. dd∗ v0
=
lc EYd
K0 max(|cos θ|, |sin θ|) (39)
dt lc 1 − υ2
3.1. Damage evolution law It can be seen from Eqs. (38) and (39) that the energy dissipation D d
related to damage evolution is given by:
In brittle elastic materials, crack evolves under tensile loading con­
ditions once the mode-I stress intensity factor reaches the critical value. (40)

D d = Yd ḋ ≥ 0
In case of subcritical microcracking, the time effect and size effect
Since fracture healing is not considered in the study, the damage
cannot be ignored and the crack may propagate for stress intensity factor
should never decrease: ḋ ≥ 0. To satisfy the non-negativity of energy

lower than the critical value.40,41 Specifically, the subcritical fracture
properties have been experimentally investigated for shales and other dissipation (40), a restriction should be applied on the damage energy
release rate, such that: damage does not propagate (ḋ = 0) whenever

brittle rocks.40 This is partly due to the clay-rich characteristics of
shales, once contacting with water leads to a reduction in fracture Yd < 0. To avoid the physically unrealistic fluid driven fracturing pro­
toughness and as a result enhances the subcritical fracture growth. cess in compression, only the positive part of fracture driving force is
A well-known function describing the relation between subcritical considered in the study and a stress-based criterion47,48 to define the
crack growth and mode-I stress intensity factor, is the Charles subcritical tensile deformation is adopted to initiate the damage evolution, as
law42: follows:
( )n ( ( ))
dl KI ̂ ft σ′ij Yd
Yd+ = H (41)
= v0 (35)
dt K0
( ) ∑ 3 ( / )2
where KI is the mode-I stress intensity factor, v0 is a reference constant ft σ′ij = 〈σ′pa 〉+ σcr − 1 (42)
for crack propagation velocity; K0 is a reference stress intensity corre­ a=1

sponding to reference crack velocity v0 , and the value of K0 is about


10%–20% of critical stress intensity factor,40 which can be determined where Yd+ is the positive part of damage energy release rate, ft is a
by fracture toughness test. Symbol n is the crack growth index. It should function with respect to effective stress, {σ′pa }a=1,2,3 are the three prin­
be noted that K0 , v0 and n are material parameters that are influenced by cipal effective stresses, σ cr is the critical stress to identify the crack onset
rock type, environmental humidity and room temperature. in uniaxial tension, the term H(f ̂ t (σ′ )) is a stress-based crack onset cri­
ij
The relation (35) needs to be completed with the reduced dissipation terion to avoid fracture propagation due to compressive load, which is
inequality in the following form: defined by:
D f = G ε l˙ ≥ 0 (36) ⎧ ( )

⎨ 1, ft σ′ij ≥ 0
̂
H(x) = ( ) (43)
where D f is the energy dissipation caused by crack propagation, G ε is ⎪
⎩ 0, ft σ′ij < 0
the energy release rate for microcracks, l is the length of microcrack, as
seen in Fig. 1(a), while l˙ represents the crack velocity. The above re­
In order to formulate the damage model within the framework of irre­
lations (35)–(36) constitute the general criterion of time-dependent
versible deformation, the positive part of fracture driving force is
microcrack propagation. For brittle/quasi-brittle fracture problems in 47,49
substituted by its historical maximum value H +
M , defined by :
rock materials, the Griffith criterion is generally considered: l˙ = 0 when
{ }
G ε < G f and l˙ ≥ 0 when G ε = G f , where G f is the critical fracture
+ +
H M = max Yd (x, s) (44)
s∈[0,t]
energy. This kind of fracture propagation problems has been studied
in.43,44 For more complex fracture propagation problems, G f may be Making use of Eqs. (41)–(44) in Eq. (39) is allowed to obtain the final
damage evolution law:
expressed in terms of microcrack length l and crack velocity l˙ .45 In the
case of subcritical law, Eq. (35) is associated with the fracture energy (√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ / )n
dd∗ v0 lc EH +
that depends on the crack velocity l˙ = v(G ε ). = M
K0 ⋅max(|cos θ|, |sin θ|) (45)
dt lc 1 − υ2
Then we need to link the fracture energy G ε and mode-I stress in­
tensity factor KI in the following form: Due to the damage evolution depends on the microstructural size lc
1− υ 2 2 and time, the damage equation (45) accounts not only for the rate effect
Gε= KI (37) but also for the size effect, which is consistent with the theory in classic
E
fracture mechanics.
in which E and υ are Young’s modulus and Poisson’s ratio, respectively. It should be recognized that the obtained damage law is applicable
It was proved in46 that for evolving microcracks, the microscopic and for a straight microcrack trajectory and a traction-free opening mode of
macroscopic damage energy release rate have the following relation: crack lips, as consistent with the original development in.44 For kinked
microcrack propagating along the direction that maximizes the energy
Gε ∗
1 dCijkl (d) ( (0) ) ( (0) ) dα∗ij (d) ( (0) ) 1 2 dβ∗ (d)
release rate, it is recommended to follow the theoretical development
= Yd = − exkl u exij u + pf exij u + pf
lc 2 dd dd 2 dd and computation algorithm in,50 where the crack direction needs to be
(38) determined in each iteration step.

in which Yd is known as the macroscopic damage energy release rate.


The scaling law with microstructural length lc makes the link between
dissipated fracture surface energy during microcrack propagation and

5
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

3.2. Permeability change

Considering the fact that fluid pressure is acting on the impermeable


crack face to open the crack, significant hydraulic anisotropy is induced
at the crack tip, which is well represented by the hydraulic anisotropic
factors A∗ij in Eq. (27). This mathematical derivation is consistent with
the experimental results observed in the fluid injection tests, where the
pressure evolution at different lateral locations is not uniform along the
circular direction.51 For the poroelastic material containing isolated
cracks, fluid flow in the media generally includes flow in the porous
matrix as well as in the fractures.52,53 Instead of simulating fluid flow in
separate medium, a continuity relation of damage induced permeability
is considered in the study.
To provide a reasonable representation of crack transmissivity and
ease the difficulty in modelling fracture flow, fluid flow in fractures may
be described by laminar flow through a slot, where the fracture
permeability is related to crack opening. Fluid flow in the matrix is a
background process that contributes to the global flow. Therefore, the
global permeability is generally composed of two parts, one from matrix
Fig. 3. Shape of calibration function for permeability.
permeability and the other one from fracture permeability.
The model developed by Carman54 is commonly used to describe the
permeability evolution of a porous matrix due to porosity change. In damage is about to reach the unity value, the permeability increases at a
case of fluid driven fracturing, the pore size of matrix is significantly fast rate. Besides, the corrector factor βt represents the scaled magnitude
smaller than the fracture aperture. Therefore, any contribution to of permeability.
permeability resulting from porosity change is disregarded in this study,
only changes in permeability induced by fracture propagation are 4. Homogenized coefficients
considered. The cubic law by55 is typically used to describe fracture
permeability parallel to the flow direction when the fracture opening The effective coefficients are derived with formulae (23) and (25)-
size is known. However, determining the accurate fracture opening size (27) using the functions ξij , κ and ζ that are solutions of the cell problems
is challenging due to the dynamic and unstable nature of fluid flow (28)–(31) and (33)–(34), respectively. For a set of normalized micro­
process. Additionally, an idealized formulation of permeability evolu­ crack length d∗ ∈ [0, 1] (d∗ = d × max(|cos θ|, |sin θ|) being the normal­
tion may not fully capture the experimental phenomena observed in the ized damage variable) and microcrack orientation θ ∈ [0∘ , 180∘ ], these
hydraulic fracturing tests as described in.56,57 To balance the need for an functions are computed numerically with the finite element software
accurate representation of fracture propagation and the physical phe­ COMSOL Multiphysics. Then the derived values of effective coefficients
nomena of preferential fluid flow, an empirical relation will be adopted are interpolated linearly between continuous data points. The values of
to implicitly represent the permeability change induced by fracture material properties in the computation are extracted from the literature
opening, which is described by damage evolution that is included in the in,53,63 where the thermo-hydro-mechanical behaviors of argillite were
formula. measured in.64 The basic parameters are summarized in Table 1.
Similar mathematical treatment for preferential fluid flow in rock
materials can be found in the literature, see e.g., the empirical expo­ 4.1. Mechanical coefficients
nential relation developed in,58 the piecewise relation based on cubic
law constructed in,59 the complicated S-shape function in,60 et al. A The homogenized mechanical coefficients C∗ijkl are presented in Fig. 4
detailed review for permeability evolution model of fractured porous with respect to normalized damage variable d∗ and crack orientation θ.
media can be found in,61 where the empirical relation is proposed with As we assumed the isotropic porous matrix at the microscale, corre­
respect to different variables depending on specific research objectives. sponding to the virgin material with coefficients C∗ijkl (0), the presence of
In the permeability model by Chen et al.,62 the relation between microcracks introduces the anisotropic effective response, but the usual
permeability growth and deviatoric effective stress was represented by a symmetries still hold true for the coefficients: C∗ijkl = C∗jikl = C∗ijlk = C∗jilk .
modified logistic function, which effectively captures the flow behavior It is also noted that the homogenized mechanical coefficients have a
from the elastic to post-failure state. Motivated by the work in,62 we nonlinear dependence on the normalized damage variable, while d∗ = 1
adopted a modified logistic function considering the normalized crack corresponds to a total damaged state that the microstructural cell is
length that links the microscopic to macroscopic scale, as follows: completely separated by the crack.
(
βt
) For horizontal and vertical crack orientation (e.g., θ = 0∘ , 180∘ ), we
k = k0 + k0 d⋅exp (46) have the same values for the mechanical coefficients except that C∗1111 is
⏟⏞⏞⏟ exp(1 − d)
matrix ⏟̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅⏞⏞̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅⏟ replaced by C∗2222 and reciprocally. The vanishing response of mechan­
fracture
ical coefficients at total failure is only observed for θ = 0∘ and 180∘ as the
where k0 represents the matrix permeability, βt is a corrector factor that crack overlaps with the coordinate axes. For other crack orientations,
calibrates the homogenized permeability. It is seen from Eq. (46) that the material still has residual rigidity.
the permeability of fractured porous material includes the matrix
permeability and fracture permeability. When the material is undam­
aged (d = 0), the permeability k is equal to the matrix permeability.
While the permeability increases nonlinearly with the damage Table 1
Material properties of the isotropic porous matrix.
evolution.
To provide a straightforward understanding of equation (46), we plot Young’s modulus E Poisson’s ratio Biot coefficient Initial porosity
the function with respect to damage variable in Fig. 3. The permeability (MPa) υ α φ0

increases very slowly when damage initiates from a small value. When 2000 0.3 0.6 0.146

6
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 4. Evolution of homogenized mechanical coefficients with respect to normalized damage variable d∗ and crack orientation θ for elastic parameters E = 2 GPa
and υ = 0.3.

4.2. Poroelastic coefficients normal to crack face. This result is consistent with the numerical anal­
ysis of Biot’s coefficient with respect to the rock microstructure in.65
The homogenized Biot coefficients α∗ij and Biot modulus β∗ are pre­
sented in Fig. 5 with respect to normalized damage variable d∗ and crack
4.3. Hydraulic coefficients
orientation θ. The usual symmetries hold for the coefficients: α∗ij = α∗ji .
With an increase of damage value, the upper and lower parts of the cell The homogenized hydraulic coefficients A∗ij are presented in Fig. 6
are progressively partitioned by a traction free and impermeable crack
with respect to normalized damage variable d∗ and crack orientation θ.
face. The deterioration degree of material with respect to the virgin state
The usual symmetries hold for the coefficients: A∗ij = A∗ji . The presence of
is becoming stronger and as a result the values of Biot coefficients in­
microcracks leads to anisotropic hydraulic response and the hydraulic
crease gradually.
coefficients have a nonlinear dependence on the normalized damage
By comparing Figs. 4 and 5, it is noted that the principal axes of Biot
variable. When the porous matrix is totally partitioned by the crack,
coefficients tensor correspond to the direction of extreme material ri­
fluid flow only occurs along the crack face and flow within the porous
gidity degradation. The associated principal values of α∗ij represent the
matrix perpendicular to the crack plane is interrupted.
extreme deterioration degree of the elastic stiffness. Therefore, the fast
increase in Biot coefficients correspond to the abrupt decrease in elastic
5. Numerical examples
coefficients, as seen from the evolution of α∗22 and C∗2222 for horizontal
crack.
To get a deeper understanding of the multiscale model, the local
As concerns the Biot modulus, the virgin material (d∗ = 0) has the
response analysis is generally needed to be conducted. This part of ex­
minimum value β∗ (0) = (α − φ0 ) /Ks , where Ks is the bulk modulus of
amples has been investigated in detail in our previous works, see e.
solid grain. While the homogenized Biot modulus is increasing for a
g.,31,46 where the rate effect and size effect of the damage model are
gradually damaged porous matrix. In addition, since we focus on fluid
discussed. Thus it will not be explored in the current study, our objective
pressure acting to open crack face, it is obvious from Fig. 5 that the
is to evaluate the model performance for fracturing in heterogeneous
obtained Biot’s coefficient tensor shows significant anisotropy along and
rocks due to microcrack nucleation and propagation. In the following

7
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 5. Evolution of homogenized poroelastic coefficients with respect to normalized damage variable d∗ and crack orientation θ.

Fig. 6. Evolution of homogenized hydraulic coefficients with respect to normalized damage variable d∗ and crack orientation θ.

8
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

part, we study the macroscopic behavior predicted by the multiscale


damage model.

5.1. Direct tension test

5.1.1. Conceptual failure theory


The direct tension tests on clay-rich rock materials have been con­
ducted by many authors, e.g.,.66–70 It is found that the final failure mode
has a clear relation with the foliation plane. Typically, the rock materials
with small dipping angles (≤ 45◦ ) generally fail along the foliation
plane,70 which indicates that the failure is initiated from the tip of
pre-existing microcracks along the foliation. For the rock materials with
large dipping angle, e.g., 75◦ , the failure plane exhibits a sawtoothed
shape partly along and partly through the foliation plane.70 For the rock
materials with vertical foliation plane, the failure is almost horizontal.71
The theory of single plane weakness was originally introduced by
Jaeger72 on shear failure, and has drawn much attention in the area of
rock mechanics. This theory may be extended to analyze the tensile
failure behavior by assuming that each foliation plane belongs to a same
family of weakness planes and has the same tensile strength. Since the
strength of foliation plane is generally lower than the intact material, the
tensile fracturing is more likely to occur along this weakness plane.
Based on the experimentally observed phenomena in70 and the theory of
single plane weakness,72 the failure behavior in the direct tension test
may be represented by the plane shown in Fig. 7.
Fig. 8. Geometry and boundary conditions for the direct tension test.

5.1.2. Mesh objectivity


A pre-existing crack is inserted into the geometry, with a position
It is known that the rate-dependent damage model avoids the issue of
located at the mid-height of the specimen. Two meshing schemes are
mesh dependence.73 The damage evolution law (45) belongs to this type
considered in the simulation: element size of 0.2 mm and 0.4 mm in the
and we expect the model has such an advantage. Therefore, the objec­
refined zone, respectively, as seen in Fig. 9. The evolution of stress-strain
tivity of the multiscale damage model with respect to mesh indepen­
relation is illustrated in Fig. 10, while the damage band along crack
dence is addressed in the following part.
propagation direction is presented in Fig. 11. The axial stress increases
Consider a direct tension test for the configuration described in
linearly with the applied strain in the first stage, and reaches the
Fig. 8. The upper boundary x2 = 50 and lower boundary x2 = 0 are set to
maximum value with damage evolution. When the specimen is gradu­
be rigid that are kept to be parallel in horizontal direction. A
ally partitioned by the crack, stress goes down quickly. Comparison of
displacement-controlled boundary is applied at the upper boundary
the stress-strain curve and damage band shows very small sensitivity for
through constant rate ėx22 = 10− 6 s− 1 , while the lower boundary is fixed
different meshes. This result is consistent with the demonstration in rate-
in vertical direction. The basic mechanical properties of the rock spec­
dependent damage models74 where a regularizing effect is introduced to
imen are referred to Table 1. The reference stress intensity factor and
avoid mesh dependency.
subcritical index are given as K0 = 2.31 × 104 Pa•m0.5 and n = 3,
respectively, which are extracted from the rock fracture properties in.31
5.1.3. Model validation
For the parameters associated with damage evolution, larger micro­
It should be noted that due to the complexity of the developed
structural size and reference crack velocity lead to lower material
multiscale damage model, the analytical solution to verify the model is
strength, as seen in the local response analysis in.46 Considering an
not constructed yet. Instead, we use the experimental results recorded
appropriate tensile strength using the specific microstructural size and
by Hashiba and Fukui68 in a direct tension test to validate the multiscale
reference velocity, e.g., lc = 1.0 × 10− 4 m and v0 = 1.0 × 10− 5 m/s, are
damage model.
chosen for the simulation of mesh objectivity. For the direction of
The experimental geometry is that presented in Fig. 8, and the
microcrack distribution, we consider a horizontal crack orientation in
loading rate applied at the upper boundary is ėx22 = 10− 6 s− 1 . The basic
the test, which corresponds to the curve of θ = 0◦ in Fig. 4. All of the
rock properties have been considered (E = 2 GPa; υ = 0.3; K0 = 2.31 ×
adopted parameters are summarized in Table 2.
104 Pa•m0.5). Considering the bedding planes inclined at 45 ◦ to the
axial direction of Wakkanai siliceous mudstone specimen,68 and the fact
that bedding may be regarded as a same family of microcracks, thus we
assume the microcrack oriented at a same direction to the bedding,
which is a common type of microcracks identified in sedimentary rocks.9
Based on the information of mechanical parameters and microcrack
orientation, the corresponding homogenized mechanical coefficients
can be calculated as a function of damage variable, corresponding to the
curve of θ = 45◦ in Fig. 4.
Due to the lack of experimental data on the rock microstructural
characteristics, as well as the heterogeneous nature of rock material, the
stress-strain behavior and failure characteristic in a same borehole may
exhibit a considerable difference. In view of the above considerations,
the determination of micro-structural size lc and damage parameter v0
Fig. 7. Conceptual failure plane in the direction tension test (modified from71). has been done by calibration in order to get a good agreement with the

9
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Table 2
Parameters for simulation of direct tension test31.
E (MPa) υ (− ) α (− ) φ0 (− ) θ (◦ ) K0 [Pa•m0.5] n [− ] v0 [m/s] lc [m]
4 − 5 4
2000 0.3 0.6 0.146 0 2.31×10 3 1×10 1×10−

model to reproduce a certain damaged behavior. The mismatch in strain


hardening/softening behavior is mainly caused by sample stiffness,
bounding surface evolution and testing system, etc., which is not our
focus in the multiscale damage model. In addition, it is worth
mentioning that the reproductivity of simulated results in Figs. 12 and
13 consist mainly in calibration rather than blind prediction. The me­
chanical behavior influenced by the micro-structural size and damage
parameters can be referred to the local response analysis in.31,46 After
getting a close response to the actual behaviors in the local model, the
parameters can be used to simulate the observed experimental phe­
nomena, thus to get a good agreement in an efficient manner.

5.2. Borehole failure

In this subsection, we numerically investigate the propagating


behavior of hydraulic fracture in a wellbore system. The example re­
produces the procedure of hydraulic fracturing (HF) where fracturing
fluid is injected into the rock formation rapidly, which leads to the
growth and evolution of microcracks. The hydraulic fractures are
created to substantially improve hydraulic conductivity and produc­
ibility for tight reservoirs, such as source shales and hot dry rocks. Our
interest here is to investigate how the multiscale poroelastic damage
model predicts fracture initiation from localized deformation area and
demonstrates the fundamental mechanism of hydraulic fracturing in
Fig. 9. (a) Mesh size of 0.4 mm and (b) mesh size of 0.2 mm around the central tight rocks considering the anisotropic increase of Biot coefficients due
horizontal line. to damage accumulation.

5.2.1. Fracturing along preferred stress direction


The geometry and boundary conditions of the numerical domain are
presented in Fig. 14 where a borehole of radius R = 10 mm is drilled in
the center of rock formation. The domain is 200 mm in width and 100
mm in height. To improve the computational efficiency and due to the
symmetrical problem, only half of the borehole is considered in the
simulation. All the boundaries are impermeable except for the borehole
boundary, where an injection rate of 10 mL/min is applied on the
borehole. The bottom boundary is set to be roller while the other
boundaries are subjected to in-situ stress loadings.
The properties of rock material are same to that used in Table 1.
Water is adopted to be an incompressible fracturing fluid in the simu­
lation. Other parameters used in the simulation of borehole failure are
given in Table 4.
The initial pore pressure is set to zero in the reservoir (pf0 = 0). The
(0)

rock formation is subjected to in-situ stress of σ h = 1 MPa and σ H = 6


MPa. The water pressure on the borehole is applied by setting an in­
jection rate of 10 mL/min on the borehole boundary. An initial crack of
2 mm length is inserted into the geometry around the borehole to initiate
the non-homogeneous response, as seen from the red line in Fig. 14. The
intrinsic permeability of crack along the opening direction is set to be
Fig. 10. Stress-strain relation in direct tension test for different mesh­
much higher than the matrix permeability, e.g., 1.0×10− 14 m2 in the
ing schemes.
example. The problem is solved with time step of Δt = 0.5 s until failure
is completely throughout the rock formation.
experimental observations, e.g., v0 = 1 × 10− 4 m/s and lc = 5× 10− 4 m As water is injected through borehole at a constant rate, the evolu­
in the simulation. The adopted parameters have been summarized in tion of borehole water pressure and water pressure along the cutting line
Table 3. are shown in Fig. 15. The water pressure along the horizontal line in­
Using the parameters and boundary conditions, the obtained stress- creases gradually with the process of water injection, and the curve
strain curve and failure characteristic are presented in Figs. 12 and 13, changes from inverted-V shape (t = 20 s) to M-shape (t = 60 s, 100 s)
respectively. As seen from the figures, the simulated results for both with time elapsed. When the hydraulic fracture front has passed through
peak stress and brittle failure characteristics agree well with the the center part, water pressure in the center goes up again and the curve
experimental phenomena, which shows the capability of the multiscale becomes to be inverted-V shape, as seen from t = 195.5 s in Fig. 15(b).

10
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 11. Damage band in refined mesh zone with respect to different meshing schemes.

different time in Fig. 16. Due to the heterogeneous material properties


Table 3
between intact part and crack, the fluid is more likely to be transported
Material properties for the direct tension test in the validation.
along the initial crack area, as seen from Fig. 16, thus leading to the
E υ α φ0 θ K0 lc (m) v0 (m/s) localized deformation near the borehole. Accompanied by water flowing
(MPa) (− ) (− ) (− ) (◦ ) (Pa•m0.5)
into the porous rock formation, borehole water pressure increases
2000 0.3 0.6 0.146 45 0.6 5.0 × 1.0 × nonlinearly until reaching a peak value at t = 14 s, see Fig. 15(a). Each
10− 4 10− 4 drop in water pressure is closely related to the fracture propagation in
the formation, represented by the increase of fracture area. It can be
validated from Fig. 15(a) that the most significant drop in water pressure
is accompanied by the sharpest increase of fracture area, approximately
at t = 16 s. This modelling result is consistent with the fracturing process
in other materials using the phase field method by Guo and Fall.47,60
After the largest drop in borehole water pressure, the pressure value
undergoes an oscillating behavior until the fracture completely

Fig. 12. Comparison of simulated stress-strain curve and experimental data.


Fig. 14. Geometry and boundary conditions of the borehole failure problem.

The lower water pressure value in the center part of domain is closely
associated with fracture propagation, where a slight stress concentration Table 4
is formed ahead of fracture front and moving with hydraulic fracture Parameters used in the simulation of borehole failure.
correspondingly.
K0 [Pa•m0.5] n [− ] v0 [m/s] lc [m] k0 [m− 2] at [− ] θ [◦ ]
To validate the relation between water pressure change and fracture
4 − 5 − 4 − 18
propagation, we present the damage distribution in the domain at 2.31×10 3 1×10 1×10 5×10 3.5 90

Fig. 13. Failure characteristics in the simulation and experimental images reproduced from Hashiba and Fukui.68 Note: Red line represents the failure plane. (For
interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

11
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 15. Example of borehole failure along preferred stress direction: (a) evolution of borehole water pressure, (b) water pressure along the horizontal line at the
height of 50 mm.

Fig. 16. Time-serial damage distribution.

penetrates the rock formation. It is clearly seen from Fig. 16 that the respect to the initial stress, see the stress profile of t = 20 s in Fig. 17(c).
hydraulic fracture has formed in a way parallel to the direction of the As the hydraulic fracture front almost reaches the cutting line, see
maximum in-situ stress. The fracture area acts as preferential pathways Fig. 17(a), the applied water pressure acts to increase the effective stress
for fluid flow directly from borehole into the formation. in both directions. Under the condition of stress difference, the hydraulic
It should be clarified here that the multiple fractures formed around fracture preferably propagates along the maximum principal stress di­
the borehole may be closely associated with the effect of injection rate, rection, which induces the anisotropic distribution of Biot coefficients
the viscosity of fracturing fluid, HM properties of rock, etc. For the around the crack front. This is validated by the different Biot coefficients
multiscale model predicted single crack propagation under other con­ of cracked elements along horizontal and vertical directions, see the
ditions, readers are referred to our previous works in.46 Current re­ curve of t = 195.5 s in Fig. 17(d). As a result, the effective stress in
searches mainly relate multiple crack propagation with fracture horizontal direction increases more rapidly than that in the vertical di­
spacings,75 sequencing effect,76 natural fracture distribution,77 etc. The rection, see the stress profile of t = 100 s in Fig. 17(c). This result is
effect of injection rate on multiple crack propagation needs to be further consistent with the experimentally observed anisotropic damage in the
investigated in the future work. fractured rock material.6
To study the fundamental hydraulic fracturing mechanism along When the hydraulic fracture passes the cutting line and gets to the
preferred stress direction, we present water pressure distribution in the impermeable boundary, as seen in Fig. 17(b), the no flow condition leads
fracturing area (defined by the area of d > 0.9) and effective stress to water flow back. Due to the microcracking mechanism along vertical
around the fracturing area in Fig. 17(a)-(b). When the hydraulic fracture direction, the fracturing area presents a higher permeability with sig­
front is far from arriving the cutting line (t = 20 s), the effective stresses nificant anisotropy in different directions, see Fig. 17(e). Thus, the
in horizontal and vertical direction only have a minor variation with flowback water is more likely to be transported along the existed

12
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 17. Simulation results of hydraulic fracture propagates along preferred stress direction: (a) water pressure in the fracturing area (d > 0.9) and effective stress
distribution at t = 100 s, (b) water pressure and effective stress distribution at t = 195.5 s, (c) effective stress σ‘11 and σ‘22 along the line at different time, (d) effective
Biot coefficients along the cutting line at different time, (e) components of effective permeability tensor along the line at different time.

13
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

hydraulic fractures in a preferred manner. As water is continuously multiple hydraulic fractures evolve towards the wall, as shown in
injected in the borehole, the flowback water is interacting with the Fig. 20.
injected water, which leads to the significant decrease of effective To study the fundamental hydraulic fracturing mechanism along
stresses in vertical direction that is even lower than the initial stress preferred microcrack direction, we present the simulation results in
value, see the stress profile of σ‘22 (t = 195.5 s) in Fig. 17(c). Fig. 21. When the hydraulic fracture front is far from arriving the cutting
The fluid pressure at failure and dominant flow streamlines in the line (t = 20 s), the effective normal stress and shear stress only have a
domain are shown in Fig. 18. Comparison of damage distribution and minor variation with respect to the initial stress, see the stress profile of
flow streamline reflects that the developed fractures provide clear paths t = 20 s in Fig. 21(c). As the hydraulic fracture front almost reaches the
for fluid directly flow from the borehole into the rock formation. The cutting line at t = 90 s, see Fig. 21(a), the applied water pressure acts to
modelling results on fracture pattern and fluid flow qualitatively agree increase the effective stresses in normal and shear directions to the crack
with the experimental description in.57 It is noted that a negative pore face, as seen from the stress profile of t = 90 s in Fig. 21(c). Due to the
pressure is occurring ahead of the fracture tip, which is caused by the cracking mechanism of microstructure degraded by the fluid filled
water flowing back towards the borehole due to impermeable boundary cracks, the hydraulic fracture preferably propagates along the micro­
conditions. Similar results have also been reported by other researchers, crack direction.
see e.g., Eghbalian et al..11 When the hydraulic fracture passes the cutting line and moves close
to the impermeable boundary in Fig. 21(b), the redistribution of pore
5.2.2. Fracturing along preferred microcrack direction pressure due to boundary effect acts to preventing the further propa­
To evaluate the model performance to reproduce the borehole failure gation of hydraulic fracture as well as leading to water flowback, and as
along preferred microcrack direction, an example of similar hydraulic a result injected water is hardly transported along the cracking path,
fracturing test is conducted in the following part. In this example, the which can be validated by the negative pore water pressure in Fig. 22.
geometry and boundary conditions are the same as that in Fig. 14, Besides, the interaction between flowback water and continuously
except that σh = σH = 6 MPa and the set of microcrack is oriented in an injected water gives rise to the significant decrease in the crack normal
angle of 45∘ with respect to the horizontal direction. An initial crack of 2 stress in the hydraulic fractures, which is much lower than the initial
mm length is inserted into the geometry around the borehole at a di­ stress value, see the stress profile of σ‘n (t = 160 s) in Fig. 17(c).
rection of 45◦ with respect to x-axis, and the crack permeability is To illustrate the poroelastic behavior of the multiscale damage
1.0×10− 14 m2. The modelling parameters used in the test are referred to model and the dominant flow path, we present the evolution of Biot
Table 4 except that θ = 45∘◦ . coefficients along the cutting line and the distribution of water pressure
In the numerical example of borehole failure along preferred with fluid velocity in Fig. 22. By comparing Fig. 22(a) with damage
microcrack direction, the evolution of borehole water pressure and distribution in Fig. 20, the abrupt increase of Biot coefficients occurs in
water pressure along the horizontal line at the height of 50 mm are the damaged area, while the values in the non-damaged area are kept
shown in Fig. 19. The drop in borehole water pressure is due to the water constant, which shows the highly localized HM behavior. For the rock
flowing into the porous medium through induced fracture pathways, as material with low porosity, the increase of Biot coefficients in the HF is
validated by the damage distribution in Fig. 20. Water pressure along accompanied by the local drop in pore water pressure, as validated by
the horizontal line increases gradually with the process of water injec­ the decreased trend of pore water pressure distribution along the
tion. The pressure value around the damaged area presents an opposite cracking direction in Fig. 22(b), which is caused by the pore dilation
trend, where the positive pressure value is existed in the area away from around the fracture tip. As a consequence, the maximum water pressure
the hydraulic fractures, while the negative value is found in the area may be lower than the confining pressure, see Fig. 19(a), but the HF can
close to the hydraulic fractures, as seen from Fig. 19(b). still propagate along the preferred microcrack direction. Under this
The damage distribution in the domain is shown in Fig. 20 with circumstance, the distribution of water pressure shows a significant
respect to different time, in which the damage propagates along the anisotropic behavior, and negative pressure value occurs ahead of the
preferred microcrack direction. Due to the existence of initial crack, fracture tip due to the impermeable boundary effect, as seen in Fig. 22
more fluid penetrates this area that causes the tensile deformation of (b).
rock material. When the principal effective stress reaches the critical This example shows the dominant effect of microcrack distribution
stress that provides a driving force for fracture evolution, damage on fracture propagation behavior obviously, which is able to explain the
propagates along the initial crack area, as seen from damage distribution complex behavior of hydraulic propagation in naturally fractured rock
at t = 20 s. As constant water injection rate is applied continuously, materials, where the fracture propagation is influenced by both the
maximum stress and the microcrack distribution.

5.3. Hydraulic fracturing in a three-hole rock specimen

The last example in the following part is to examine the multiscale


damage in reproducing localized fracturing behavior under more com­
plex stress conditions. For the example simulated here, the hydraulic
fracturing experiment56 on a square rock specimen with three holes is
numerically reproduced. The square specimen has a length of 152 mm
and three holes with radius of 3 mm are arranged equilaterally at same
distances. A small notch with rectangular dimensions of 1 × 6 mm is
inserted to the top hole, in order to initiate the fracturing process. The
four boundaries of specimen are constrained mechanically, while at­
mospheric pressure is applied that fluid can flow freely through the
boundaries. The detailed geometry and boundary conditions are illus­
trated in Fig. 23.
The experiment was conducted in two steps56: in the first step, fluid
Fig. 18. Water pressure distribution and dominant fluid flow (represented by is injected through the left bottom hole at a constant pressure condition
black arrow line) in the domain. pf = 1.4 MPa, while no external pressure are applied on the other two
(0)

14
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 19. Example of borehole failure along preferred microcrack direction: (a) Evolution of borehole water pressure. (b) Water pressure along the horizontal line at
the height of 50 mm.

Fig. 20. Damage distribution along preferred microcrack direction in the domain.

holes. The fluid injection process continues for sufficient time so that damage model. The performed numerical analysis provides fundamental
fluid completely permeates the rock specimen. In this way, a insight into rock fracture behavior and may have a certain guidance to
non-homogeneous distribution of fluid pressure will be generated in the field application. In the shale gas exploitation with multiple fracturing
notch and through the specimen. In the second step, fluid is injected wells, the hydraulic fractures are not only influenced by the distribution
through the top hole at a constant rate q = 5 mL/min until fracture of natural cracks, but also influenced by the pore pressure gradient field
occurs and the fracture type is affected by the pore pressure distributed before fracture operation is started.
in the first step.
The parameters of rock specimen used in the simulation are listed in 6. Conclusions
Table 5. Water is chosen as the fracturing fluid and assumed to be
incompressible for simplicity. The phenomenon of macroscopic fracturing in rock materials has
The final pore fluid pressure distribution in two steps is presented in been numerically studied by incorporating the dynamically evolved
Fig. 24. The fracture path in the experiment and in the model are shown microcracks in the modelling framework through scale separation. The
in Fig. 25. It can be seen from the figures that the simulated pore fluid present approach constructing the coupled poroelastic damage model
pressure and damage distribution agree well with that in the experi­ appears to be the first one that proposes a multiscale description of fluid
mental observation. When the rock material is subjected to non-uniform flow-deformation due to distributed damage nucleation and propaga­
pore pressure field and in the absence of stress difference conditions, the tion, which is fundamental to describe the complex fracturing behavior
generated fluid pressure gradient between the notch and left bottom influenced by microcrack distribution.
hole affects the fracture path, which is moving towards the region of Starting with a fully coupled poroelastic fracture problem at the
higher local pore pressure field. Two fracture propagation mechanisms microscale, a periodically distributed microstructure has been assumed
are observed during the simulation: one fracture path is affected by the and asymptotic homogenization method has been adopted to formulate
combined effect of microcrack distribution and local pore pressure dis­ the multiscale damage model. The model has considered the complex
tribution, in which the fracture initiates from the notch tip. The second coupling between fluid pressure-deformation and HM properties at the
fracture is initiating from the injection port towards the top boundary, microscale, giving rise to the nonlinear anisotropic mechanical
where the path is dominated by the microcrack distribution at vertical behavior, degradation of both elastic stiffness and poroelastic properties
direction. at the macroscale. The microstructural size has been included in the
This example clearly shows the effect of local pore fluid pressure field damage evolution law, which is able to describe the rate effect and size
on fracture propagation, which is well captured by the multiscale effect. In addition, a permeability model has been proposed accounting

15
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

Fig. 21. Simulation results of hydraulic fracture propagates along preferred microcrack direction: (a) water pressure in the fracturing area (d > 0.9) and effective
stress distribution at t = 90 s, (b) water pressure and effective stress distribution at t = 160 s, (c) effective normal stress σ‘n and shear stress σ ‘s along the line at
different time.

Fig. 22. (a) Biot coefficients along the cutting line at t = 160 s, (b) water pressure with fluid velocity (represented by black arrow line) in the domain at t = 160 s.

for crack opening induced permeability change. performed for examples of direct tension test, borehole failure at the lab
The homogenized coefficients have been illustrated in detail for a scale and hydraulic fracturing in a three-hole rock specimen. The frac­
given set of initial material parameters, with dependence on the turing process in heterogeneous rock materials has been shown to be a
normalized damage variable. Numerical simulations have been multiscale phenomenon that initiates from the microcrack nucleation

16
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

and propagation. The fracture propagation direction has been proved to


be influenced by both the external loadings and the microcrack
distribution.
The strictly math-based upscaling method with proposed damage
equations has allowed the model to uniquely describe the preferred
microcracking process in rocks. In some complicated engineering sce­
narios that the in-situ stress states dominantly determine the orientation
of microcracks, the developed model has great values to investigate the
impact of microcrack orientation on the macroscopic rock failure. The
easy implementation within finite element framework and revealed
micro-mechanism for macroscopic failure underscore its wide applica­
tion in rock mechanics problems.

CRediT authorship contribution statement

Jianxiong Yang: Conceptualization, Formal analysis, Investigation,


Methodology, Validation, Visualization, Writing – original draft. Jian­
feng Liu: Funding acquisition, Resources, Supervision, Writing – review
& editing. Wenfeng Li: Writing – review & editing. Jingjing Dai: Data
curation, Visualization. Fujun Xue: Visualization. Xiaoying Zhuang:
Resources.

Fig. 23. The geometry and boundary conditions in the test of hydraulic frac­ Declaration of competing interest
turing in a three-hole specimen.
The authors declare that they have no known competing financial

Table 5
Parameters in the example of HF in a three-hole rock specimen.
E (MPa) υ α φ0 K0 [Pa•m0.5] n v0 [m/s] lc [m] k0 [m− 2] at θ [◦ ]

2000 0.3 0.6 0.146 2.31×104 3 1×10− 5


1×10− 4
5×10− 20
3.5 90

Fig. 24. The final pore fluid pressure distribution in (a) first step, (b) second step. Black arrow line denotes the major fluid flow path.

Fig. 25. (a) Fracture pattern in the experiment,56 (b) simulated fracture path.

17
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

interests or personal relationships that could have appeared to influence Acknowledgements


the work reported in this paper.
This research was financially supported by the National Natural
Data availability Science Foundation of China (Grant No. 12302503, U20A20266) and
the Scientific and technological research projects in Sichuan province
Data will be made available on request. (Grant No. 2023ZYD0154, 2024GJHZ0035).

Appendix A. Homogenization upscaling

In this appendix, the asymptotic homogenization method32 is adopted to perform the upscaling process.
Eq. (13) at the order ε− 2 and Eq. (16) at the order ε− 1 allow us to obtain the following boundary value problem:
∂ ( ( ))
C e u(0) = 0 in Y and (A.1)
∂yj ijkl ykl
( ( ))
Cijkl eykl u(0) Nj = 0 on C Y (A.2)

with periodicity condition applied on the opposite exterior boundary of cell Y. To solve problem (A.1)-(A.2), it is easily found that u = u (x,t). (0) (0)

To solve the balanced problem in the medium with fluid pressure in both pores and cracks, we first present the order ε− 1 term of Eq. (15) in the
following form:

∂p(0)
(A.3)
b (0) (0)
= 0, thus pb = pb (x, t)
∂xi
Use of the condition on crack Eq. (16), and comparing p(0)
f = p(0)
f (x, t) with Eq. (A.3), it is deduced that:

(A.4)
(0) (0)
pf = p b

Eq. (13) at the order ε− 1 and Eq. (16) at the order ε0 give a boundary value problem for the function u(1) considering that u(0) only depends on x-axis,
as follows:
∂ ( ( ))
Cijkl eykl u(1) = 0 in Y and (A.5)
∂yj
( ( ) ( ))
Cijkl exkl u(0) + eykl u(1) Nj = − (1 − α)p(0)
f Ni on C
Y
(A.6)
37
and with periodicity boundary conditions on the external boundary of the cell. The function u (1)
can be looked for in the form :
( )
(A.7)
(1) (0)
ui (x, y, t) = ξpq
i (y)expq u(0) − κi (y)(1 − α)pf

Eq. (14) at the order ε− 2


and Eq. (17) at the order ε− 1
yield to:
( (0)
)
∂ k ∂pf
= 0 in Y and (A.8)
∂yi μf ∂yi

∂p(0)
(A.9)
f
N = 0 on C Y
∂yi i

with periodicity condition applied on the opposite exterior boundary of cell Y. We can easily deduce that p(0)
f = p(0)
f (x, t).
1
Introducing Eq. (14) at the order ε− and Eq. (17) at the order ε0 with the consideration of pf = pf (x,t), we can derive the following boundary
(0) (0)

value problem:

∂2 p(1)
(A.10)
f
= 0 in Y
∂y2i
( )
∂p(0) ∂p(1)
(A.11)
f f
+ Ni = 0 on C Y
∂xi ∂yi

This gives pressure p(1)


f in the following form31:

∂p(0)
f (x, t)
p(1)
f (x, y, t) = ζ i (y) (A.12)
∂xi
The displacement problem at the order ε0 is deduced from Eqs. (13) and (16):
∂ ( ( ( ) ( )) )
(0)
Cijkl exkl u(0) + eykl u(1) − αpf δij +
∂xj
in Y (A.13)
∂ ( ( ( (1) ) ( )) )
Cijkl exkl u + eykl u(2) − αp(1)
f δij = 0
∂yj

18
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

( ( ( ) ( )) )
(A.14)
(1) (1)
Cijkl exkl u(1) + eykl u(2) − αpf δij Nj = − pb Ni on C Y

In order to obtain the effective equations, we introduce the mean value operator as follows:

1
〈⋅〉= ⋅dy (A.15)
|Y|
Y

where |Y| = L2c is the area of Y. Let us integrate Eq. (A.13) over the domain Y with the help of Eqs. (A.14)-(A.15), the following relation can be
obtained:

( ( ) ( )) 1
〈Cijkl exkl u(0) + eykl u(1) − αp(0)
f δij 〉 = p(1)
b Ni dΓ (A.16)
|Y|
CY

To solve the right side member of Eq. (A.16), we look for the expression of fluid pressure pεb in microcracks at the order ε0 . Using Eq. (15) at the
order ε0 by taking into account of Eq. (A.3) is allowed to derive the following relation:

∂p(0) ∂p(1)
b
+ b =0 (A.17)
∂xi ∂yi

Considering the relation (16) on crack face at the order ε0 , it is known that p(1)
f = p(1)
b . As we consider small thinness crack that occupies no volume,
integrating the above equation (A.17) inside the crack with the help of divergence theorem enables us to derive the right side of Eq. (A.16):

p(1)
b Ni dΓ = 0 (A.18)
CY

Lastly, making use of Eq. (A.7) and (A.18) in (A.16) gives the effective macroscopic equilibrium equation in the form:
∂ (0)
Σ =0 (A.19)
∂xj ij

where Σij = 〈σij 〉 = C∗ijkl exkl (u(0) ) − α∗ij pf is the macroscopic total stress tensor in the microporous medium with cracks. The homogenized stiffness
(0) (0) (0)

tensor C∗ijkl and Biot coefficient tensor α∗ij are defined by:
( )

Cijkl = 〈Cijkl + Cijmn eymn ξkl 〉 (A.20)

α∗ij = αδij + 〈(1 − α)Cijkl eykl (κ)〉 (A.21)

The microscale mass balance (14) and Eq. (17) at the order ε0 writes:
( ( ) ( ))
∂q(0) ∂q(1) ∂exii u(0) ∂eyii u(1) ∂p(0)
(A.22)
i f
+ i +α + +β = 0 in Y
∂xi ∂yi ∂t ∂t ∂t

q(1)
i Ni = 0 on C
Y
(A.23)
Integrating Eq. (A.22) over the cell Y and applying the divergence theorem with the help of boundary condition (A.23), we deduce the macroscopic
mass conservation law:
( )
∂Q(0) ∂exij u(0) ∂p(0)
(A.24)
i f
+ γ∗ij + β∗ =0
∂xi ∂t ∂t

k∗ij ∂pf
(0)

where Qi is the macroscopic Darcy’s velocity, the homogenized coefficients γ∗ij , Biot modulus β∗ and intrinsic permeability k∗ij are
(0) (0)
= 〈qi 〉 = − μf ∂xj
given:
( )
γ ∗ij = αδij − (1 − α)〈eykk ξij 〉 (A.25)

β∗ = β + (1 − α)2 〈eyii (κ)〉 (A.26)

∂ζj
kij∗ = kA∗ij with A∗ij = 〈δij + 〉 (A.27)
∂yi

Finally, we put the derived quantities together and substitute superscript ’ hm’ for superscript ’(0)’, thus obtaining the multiscale poroelastic model in
homogenized manner, as represented by Eqs. (21) and (22).

19
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

References 37. Lewandowska J, Auriault JL. Extension of Biot theory to the problem of saturated
microporous elastic media with isolated cracks or/and vugs. Int J Numer Anal
Methods GeoMech. 2013;37:2611–2628.
1. Li W, Frash LP, Lei Z, et al. Investigating poromechanical causes for hydraulic
38. Yan X, Huang ZQ, Yao J, et al. Numerical simulation of hydro-mechanical coupling
fracture complexity using a 3D coupled hydro-mechanical model. J Mech Phys Solid.
in fractured vuggy porous media using the equivalent continuum model and
2022;169, 105062.
embedded discrete fracture model. Adv Water Resour. 2019;126:137–154.
2. Lydzba D, Shao JF. Study of poroelasticity material coefficients as response of
39. Wrzesniak A, Dascalu C, Bésuelle P. A two-scale time-dependent model of damage:
microstructure. Mech Cohesive-Frict Mater. 2000;5:149–171.
influence of micro-cracks friction. Eur J Mech Solid. 2015;49:345–361.
3. Hu DW, Zhou H, Zhang F, Shao JF. Evolution of poroelastic properties and
40. Atkinson BK, Meredith PG. The Theory of Subcritical Crack Growth with Applications to
permeability in damaged sandstone. Int J Rock Mech Min Sci. 2010;47:962–973.
Minerals and Rocks. New York: Academic Press; 1987.
4. Liu J, Qiu X, Yang J, Liang C, Dai J, Bian Y. Failure transition of shear-to-dilation
41. Swanson PL. Subcritical crack growth and other time- and environment-dependent
band of rock salt under triaxial stresses. J Rock Mech Geotech Eng. 2023.
behavior in crustal rocks. J Geophys Res. 1984;89:4137–4152.
5. Liu J, He X, Huang H, et al. Predicting gas flow rate in fractured shale reservoirs
42. Charles R. Dynamic fatigue of glass. J Appl Phys. 1958;29:1657–1662.
using discrete fracture model and GA-BP neural network method. Eng Anal Bound
43. Dascalu C. A two-scale damage model with material length. Compt Rendus Mec.
Elem. 2024;159:315–330.
2009;337:645–652.
6. Shao JF. Poroelastic behaviour of brittle rock materials with anisotropic damage.
44. Dascalu C, Bilbie G, Agiasofitou EK. Damage and size effects in elastic solids: a
Mech Mater. 1998;30:41–53.
homogenization approach. Int J Solid Struct. 2008;45:409–430.
7. Tan X, Konietzky H, Fruehwirt T. Experimental and numerical study on evolution of
45. Dascalu C, Gbetchi K. Dynamic evolution of damage by microcracking with heat
Biot’s coefficient during failure process for brittle rocks. Rock Mech Rock Eng. 2015;
dissipation. Int J Solid Struct. 2019;174:128–144.
48:1289–1296.
46. Yang J, Fall M. A two-scale hydro-mechanical-damage model for simulation of
8. Kranz RL. Microcracks in rocks - a review. Tectonophysics. 1983;100:449–480.
preferential gas flow in saturated clayey host rocks for nuclear repository. Comput
9. Ougier-Simonin A, Renard F, Boehm C, Vidal-Gilbert S. Microfracturing and
Geotech. 2021;138, 104365.
microporosity in shales. Earth Sci Rev. 2016;162:198–226.
47. Guo G, Fall M. A thermodynamically consistent phase field model for gas transport
10. Sprunt ES, Brace WF. Direct observation of microcavities in crystalline rocks. Int J
in saturated bentonite accounting for initial stress state. Transport Porous Media.
Rock Mech Min Sci. 1974;11:139–150.
2021;137:157–194.
11. Eghbalian M, Wan R, Pouragha M, Fung LS. Numerical modeling of fracturing in
48. You T, Zhu QZ, Li PF, Shao JF. Incorporation of tension-compression asymmetry into
permeable rocks via a micromechanical continuum model. Int J Numer Anal Methods
plastic damage phase-field modeling of quasi brittle geomaterials. Int J Plast. 2020;
GeoMech. 2019;43:1885–1915.
124:71–95.
12. Jain AK, Juanes R. Preferential Mode of gas invasion in sediments: grain-scale
49. Guo G, Fall M. Modelling of preferential gas flow in heterogeneous and saturated
mechanistic model of coupled multiphase fluid flow and sediment mechanics.
bentonite based on phase field method. Comput Geotech. 2019;116, 103206.
J Geophys Res Solid Earth. 2009;114(19), B08101.
50. François B, Dascalu C. A two-scale time-dependent damage model based on non-
13. Vassal JP, Orgeas L, Favier D, Auriault JL, Le Corre S. Upscaling the diffusion
planar growth of micro-cracks. J Mech Phys Solid. 2010;58:1928–1946.
equations in particulate media made of highly conductive particles. II. Application
51. Cuss R, Harrington J, Giot R, Auvray C. Experimental observations of mechanical
to fibrous materials. Phys Rev E. 2008;77, 011303.
dilation at the onset of gas flow in Callovo-Oxfordian claystone. Geol Soc Spec Publ.
14. Hazzard JF, Young RP, Maxwell SC. Micromechanical modeling of cracking and
2014;400:507–519.
failure in brittle rocks. J Geophys Res Solid Earth. 2000;105:16683–16697.
52. Yang J, Fall M. Coupled hydro-mechanical modelling of dilatancy controlled gas
15. Potyondy DO, Cundall PA. A bonded-particle model for rock. Int J Rock Mech Min
flow and gas induced fracturing in saturated claystone. Int J Rock Mech Min Sci.
Sci. 2004;41:1329–1364.
2021;138, 104584.
16. Duriez J, Scholtes L, Donze FV. Micromechanics of wing crack propagation for
53. Yang J, Fall M. A dual porosity poroelastic model for simulation of gas flow in
different flaw properties. Eng Fract Mech. 2016;153:378–398.
saturated claystone as a potential host rock for deep geological repositories. Tunn
17. Fu P, Johnson SM, Carrigan CR. An explicitly coupled hydro-geomechanical model
Undergr Space Technol. 2021;115, 104049.
for simulating hydraulic fracturing in arbitrary discrete fracture networks. Int J
54. Carman PC. Fluid flow through granular beds. Trans Inst Chem Eng. 1937;15:
Numer Anal Methods GeoMech. 2013;37:2278–2300.
150–166.
18. Yuan SC, Harrison JP. A review of the state of the art in modelling progressive
55. Witherspoon PA, Wang JSY, Iwai K, Gale JE. Validity of Cubic Law for fluid flow in a
mechanical breakdown and associated fluid flow in intact heterogeneous rocks. Int J
deformable rock fracture. Water Resour Res. 1980;16:1016–1024.
Rock Mech Min Sci. 2006;43:1001–1022.
56. Bruno MS, Nakagawa FM. Pore pressure influence on tensile fracture propagation in
19. Tang CA, Tham LG, Lee PKK, Yang TH, Li LC. Coupled analysis of flow, stress and
sedimentary-rock. Int J Rock Mech Min Sci Geomech Abstr. 1991;28:261–273.
damage (FSD) in rock failure. Int J Rock Mech Min Sci. 2002;39:477–489.
57. Zoback MD, Rummel F, Jung R, Raleigh CB. Laboratory hydraulic fracturing
20. Zhu WC, Tang CA. Micromechanical model for simulating the fracture process of
experiments in intact and pre-fractured rock. Int J Rock Mech Min Sci. 1977;14:
rock. Rock Mech Rock Eng. 2004;37:25–56.
49–58.
21. Basista M, Gross D. The sliding crack model of brittle deformation: an internal
58. Fall M, Nasir O, Nguyen TS. A coupled hydro-mechanical model for simulation of
variable approach. Int J Solid Struct. 1998;35:487–509.
gas migration in host sedimentary rocks for nuclear waste repositories. Eng Geol.
22. Krajcinovic D, Fonseka GU. The continuous damage theory of brittle materials. I.
2014;176:24–44.
General theory. Transactions of the ASME. J Appl Mech. 1981;48:809–815.
59. Olivella S, Alonso EE. Gas flow through clay barriers. Geotechnique. 2008;58:
23. Lyakhovsky V. Scaling of fracture length and distributed damage. Geophys J Int.
157–176.
2001;144:114–122.
60. Guo G, Fall M. Modelling of dilatancy-controlled gas flow in saturated bentonite
24. Tao J, Shi A-C, Li H-T, Zhou J-W, Yang X-G, Lu G-D. Thermal-mechanical modelling
with double porosity and double effective stress concepts. Eng Geol. 2018;243:
of rock response and damage evolution during excavation in prestressed geothermal
253–271.
deposits. Int J Rock Mech Min Sci. 2021;147, 104913.
61. Ma J. Review of permeability evolution model for fractured porous media. J Rock
25. Chiarelli AS, Shao JF, Hoteit N. Modeling of elastoplastic damage behavior of a
Mech Geotech Eng. 2015;7:351–357.
claystone. Int J Plast. 2003;19:23–45. Pii s0749-6419(01)00017-1.
62. Chen D, Pan Z, Shi JQ, Si G, Ye Z, Zhang J. A novel approach for modelling coal
26. Shao JF, Zhou H, Chau KT. Coupling between anisotropic damage and permeability
permeability during transition from elastic to post-failure state using a modified
variation in brittle rocks. Int J Numer Anal Methods GeoMech. 2005;29:1231–1247.
logistic growth function. Int J Coal Geol. 2016;163:132–139.
27. Lu YL, Elsworth D, Wang LG. Microcrack-based coupled damage and flow modeling
63. Mahjoub M, Rouabhi A, Tijani M, et al. Numerical study of callovo-oxfordian
of fracturing evolution in permeable brittle rocks. Comput Geotech. 2013;49:
argillite expansion due to gas injection. Int J GeoMech. 2018;18, 04017134.
226–244.
64. ANDRA. Référentiel du comportement THM des formations sur le site de Meuse/Haute-
28. Shen WQ, Shao JF. A micromechanical model of inherently anisotropic rocks.
Marne. France: Chatenay-Malabry; 2012.
Comput Geotech. 2015;65:73–79.
65. Tan X, Konietzky H. Numerical study of variation in Biot’s coefficient with respect to
29. Xie N, Zhu QZ, Shao JF, Xu LH. Micromechanical analysis of damage in saturated
microstructure of rocks. Tectonophysics. 2014;610:159–171.
quasi brittle materials. Int J Solid Struct. 2012;49:919–928.
66. Dai J, Liu J, Zhou L, He X. Crack pattern recognition based on acoustic emission
30. Zhu QZ, Kondo D, Shao JF. Micromechanical analysis of coupling between
waveform features. Rock Mech Rock Eng. 2022;56:1063–1076.
anisotropic damage and friction in quasi brittle materials: role of the
67. Gao Q, Tao JL, Hu JY, Yu X. Laboratory study on the mechanical behaviors of an
homogenization scheme. Int J Solid Struct. 2008;45:1385–1405.
anisotropic shale rock. J Rock Mech Geotech Eng. 2015;7:213–219.
31. Yang J, Fall M. A two-scale time dependent damage model for preferential gas flow
68. Hashiba K, Fukui K. Effect of water on the deformation and failure of rock in
in clayey rock materials. Mech Mater. 2021;158, 103853.
uniaxial tension. Rock Mech Rock Eng. 2015;48:1751–1761.
32. Sánchez-Palencia E. Non-homogeneous Media and Vibration Theory. Berlin: Springer;
69. Jin ZF, Li WX, Jin CR, Hambleton J, Cusatis G. Anisotropic elastic, strength, and
1980.
fracture properties of Marcellus shale. Int J Rock Mech Min Sci. 2018;109:124–137.
33. Dascalu C, François B, Keita O. A two-scale model for subcritical damage
70. Liao JJ, Yang MT, Hsieh HY. Direct tensile behavior of a transversely isotropic rock.
propagation. Int J Solid Struct. 2010;47:493–502.
Int J Rock Mech Min Sci. 1997;34:837–849.
34. Orgeas L, Geindreaua C, Auriault JL, Bloch JF. Upscaling the flow of generalised
71. Lee YK, Pietruszczak S. Tensile failure criterion for transversely isotropic rocks. Int J
Newtonian fluids through anisotropic porous media. J Non-Newtonian Fluid Mech.
Rock Mech Min Sci. 2015;79:205–215.
2007;145:15–29.
72. Jaeger JC. Shear failure of anistropic rocks. Geol Mag. 1960;97:65–72.
35. Biot MA. General theory of three-dimensional consolidation. J Appl Phys. 1941;12:
73. Bhat HS, Rosakis AJ, Sammis CG. A micromechanics based constitutive model for
155–164.
brittle failure at high strain rates. J Appl Mech-Trans ASME. 2012;79(12), 031016.
36. Coussy O. Poromechanics. Chichester, England: John Wiley & Sons; 2004.
74. Atiezo MK, Dascalu C. Antiplane two-scale model for dynamic failure. Int J Fract.
2017;206:195–214.

20
J. Yang et al. International Journal of Rock Mechanics and Mining Sciences 175 (2024) 105676

75. Gunaydin D, Peirce AP, Bunger AP. Laboratory experiments contrasting growth of 76. Wang H. Numerical investigation of fracture spacing and sequencing effects on
uniformly and nonuniformly spaced hydraulic fractures. J Geophys Res Solid Earth. multiple hydraulic fracture interference and coalescence in brittle and ductile
2021;126, e2020JB020107. reservoir rocks. Eng Fract Mech. 2016;157:107–124.
77. Wang H. Hydraulic fracture propagation in naturally fractured reservoirs: complex
fracture or fracture networks. J Nat Gas Sci Eng. 2019;68, 102911.

21

You might also like