Journal of Natural Gas Science and Engineering 44 -e364
Contents lists available at ScienceDirect
Journal of Natural Gas Science and Engineering
journal homepage: www.elsevier.com/locate/jngse
A quick and energy consistent analytical method for predicting
hydraulic fracture propagation through heterogeneous layered
media and formations with natural fractures: The use of an
effective fracture toughness
Oluwafemi Oyedokun*, Jerome Schubert
Department of Petroleum Engineering, Texas A&M University, College Station, TX, USA
a r t i c l e i n f o
a b s t r a c t
Article history:
Received 30 August 2016
Received in revised form
3 May 2017
Accepted 4 May 2017
Available online 8 May 2017
An effective fracture toughness, based on equivalent energy release-rate hypothesis is presented for
homogenizing heterogeneous layered media. Using crack closure method, the energy released when a
mode-1 fracture propagates through an equivalent homogenized layer is equated to the sum of the
energies released in the heterogeneous layered media. And from extensive numerical experiments, the
predictions of fracture tips' positions through this proposed method are of the same range of accuracy as
the known linear blend rule; the weakest link arguments technique performed poorly when compared to
the other two methods. Therefore, homogenizing heterogeneous layered media with energy consistent
approach will reduce the complexities associated with modeling fracture propagation in multi-layer
without losing accuracy. Furthermore, an effective fracture toughness, based on equivalent energy
release rate, equivalent strain energy, and modified Kachanov's damage theory is presented in this study.
The proposed approach will reduce the computation time required in predicting fracture containment
potential in formations with opened or sealed natural fractures. And comparing the proposed
phenomenological model with the rigorous solution provided by Mori-Tanaka, it was observed that the
margin of error was negligible. The benefit of the proposed phenomenological model over the MoriTanaka's effective shear modulus model is the ease of estimating the effective fracture toughness. Thus,
these models can be applied to quickly estimate the potential for hydraulic fracture containment or
broaching possibilities during oil and gas blowouts and fracturing operations.
Published by Elsevier B.V.
Keywords:
Effective fracture toughness
Heterogeneous layered media
Hydraulic fracture propagation
Crack closure method
Continuum damage mechanics
Natural fractures
1. Introduction
“Pseudo-3D” (P-3D) models are often used for quick hydraulic
fracturing design or prediction of hydraulic fracture containment in
the petroleum industry; these models are less computationally
expensive compared to the fully 3D and planar 3D models (Morales
et al., 1993; Clifton and Abou-Sayed, 1979, 1981; Hirth and Lothe,
1968; Bui, 1977). In the P-3D models, the fluid flow is assumed to be
one-dimensional along the length of the fracture, while the pressure profile in each elliptic cross sections is constrained to be linear.
And as assumed in the PKN model (Perkins and Kern, 1961;
Nordgren, 1972) that the cross sections are independent of each
* Corresponding author.
E-mail addresses:--(J. Schubert).
http://dx.doi.org/10.1016/j.jngse-/Published by Elsevier B.V.
(O.
Oyedokun),
other, so that the plane strain assumption decouples the fluid flow
and solid mechanics; similarly, this constraint is applied in the P3D
models. The linear pressure profile in the cross sections is based on
the assumption that the vertical fracture extension is slow such that
the dynamic pressure gradient in the vertical sections is negligible.
Many authors have contributed to the development of the
equilibrium height problem (P-3D), starting from the work of
Simonson et al. (1978), who developed the model for symmetric 3layer formation, with constant internal applied pressure. Ahmed
(1984), and Newberry et al. (1985) solved an asymmetric 3-layer
equilibrium height problem; while Fung et al. (1987) and
Economides (2000) extended the asymmetric equilibrium height
problem to multi-layer formations with constant and linearly
varying pressure profiles in the vertical sections respectively. But as
pointed out by Liu and Valko and Liu (2015), the model developed
by Economides and Nolte has some intrinsic errors; and sometimes
352
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
yield unacceptable results. Other notable contributions to the
development of the P-3D models include Palmer and Carroll
1983a,b, Palmer and Craig (1984), Settari and Cleary (1986),
Meyer (1986), Advani et al. (1990), and Adachi et al. (2010).
Analytical P-3D model for asymmetric multi-layer formations
with linearly varying pressure profile in each cross section is very
complicated, and difficult to tract. To reduce the complexity, an
effective fracture toughness, based on equivalent energy release
rate, developed for the upper and lower layers will condense the
multi-layer problem to the classical three-layer problem, which
algebra is tractable.
In the same vein, we also proposed in this paper the use of an
effective fracture toughness, based on the same hypothesis
mentioned earlier, which can approximately describe hydraulic
fracture propagation in formations with natural fractures. The interactions between hydraulic and natural fractures have been
studied by many authors in the past (Olson and Taleghani, 2009,
Zhou et al., 2015; Renshaw and Pollard, 1995; Dahi-Taleghani and
Olson, 2011, Gu et al., 2012; Gu and Weng, 2010; Chuprakov et al.,
2014, Zhou et al., 2008; Warpinski and Teufel, 1987; Gao and Rice,
1989; Wu, 2014). From the various studies, there are three possibilities that may occur when a hydraulic fracture intersect a natural
fracture: (1) the hydraulic fracture may cross the natural fracture
without change in propagation direction (2) the hydraulic fracture
is arrested by the natural fracture; consequently, the natural fracture, becoming part of the hydraulic fracture network and propagates from its end(s) or (3) from a weak point along its length.
In this study, the complex hydraulic fracture network created by
the interactions of the parent hydraulic fracture with ordered and
disordered natural fractures will be simplified such that an effective
fracture toughness, dependent on area of the discontinuities of a
representative volume element and the mechanical properties of
the formation matrix can suggest the containment potential of the
hydraulic fracture in the host formation.
In past studies, the use of linear blend rule (Atkins, 1979, Atkins
and Mai, 1985; Eriksson and Atkins, 1995; Eriksson, 1998) and
weakest link arguments (Landes and Shaffer, 1980; Wallin et al.,
1984; Slatcher, 1986; Iwadate et al., 1983; Beremin et al., 1983)
have been used to homogenize heterogeneous layered media.
However, Heerens et al. (1994) showed that the predictions
through the weakest link method are inaccurate.
Fig. 1. Asymmetric multilayer hydraulic fracture propagation.
having mode-1 fracture toughnessKc;1 ðzÞ. Therefore, the multilayer equilibrium-height problem reduces to the classical threelayer problem (Fig. 2).
Assuming the formations have a linearly elastic response, the
energy released per unit length of the fracture is
P¼
ZHN
ZH1
G1þ ðdzÞ þ
0
G1 dz
(2)
0
Using superposition theorem, the normal stresses at the tips of
the vertical fracture plane (in the y z plane) at arbitrary positions
z0 and z1 are
K
I
syy ðz ¼ z1 Þ ¼ syy ¼ Shmin ðz ¼ z1 Þ þ pffiffiffiffiffiffiffiffiffiffiffiffi
2pðzÞ
2. Mathematical formulations
2.1. Effective fracture toughness for layered media
Fractures tend to grow in the direction that maximizes the potential energy released. And as a result, most hydraulic fractures
propagate in mode-1; although diversion may occur at the
boundary of two formations or when intercepted by natural fractures. In this study, the hydraulic fracture is assumed to propagate
through boundaries in mode-1 only.
As the hydraulic fracture propagates through each formation,
there is an increase in the total energy released; and considering
Fig. 1, the total energy released is
PT ¼
ZHN ZL
G1 dxdz
H1
(1)
0
where G1 is the energy release rate, as defined by Griffith (1921).
Layers n þ 1 to N can be lumped together as a heterogeneous formation having a varying fracture toughnessKc;3 ðzÞ, while layers 1 to
n 1 are also lumped together as a heterogeneous formation 1,
Fig. 2. Asymmetric three-layer equilibrium height problem.
(3)
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
353
propagates into the abutting layer (lumped), is
K
Iþ
ffi
syy ðz ¼ z0 Þ ¼ syy þ ¼ Shmin ðz ¼ z0 Þ þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2pðzÞ
(4)
þ
with the extension of the upper tip byz0, new fracture surfaces are
created in 0 z z0 and the displacement of the upper-right face
is
uyy þ
k þ 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2pðz0 zÞ
¼
K
4mp Iþ
(5)
Using crack closure method (Irwin, 1958; Broek, 1991), the work
done by the normal stress in the fracture extension is equal to the
energy required to close the fracture after opening
Pþ ¼ 2
Zz0
1 þ
syy uyy þ dz
2
(6)
0
P ¼
hN1
Z þh
0
syy ðzÞ
k þ 1 þ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K ðzÞ 2pðh zÞdz
4mp Ic
(8)
Hence, the effective fracture toughness for the lumped upper layer,
þ
þ
KIC
can be estimated from the hypothesis that P ¼ Pþ .
Supposing the in-situ stresses are uniformly distributed throughout
each layer, and the upper layers are lumped as a homogeneous
þ
formation with fracture toughness, KIC , therefore, the effective
fracture toughness of the upper layer can be determined from the
algebraic expression below. It should be noted that a Heaviside
function was used to represent the uniformly distributed in-situ
stresses in each layer in Eq. (8) above. And k and m are averaged
properties (Poisson ratio function and shear modulus respectively)
of the formations; for plane strain approximation, k ¼ 3 4n and
plane stress k ¼ ð3 nÞ=ð1 þ nÞ.
#)
"
N
3
3
2 X
2
2
ðh
S
þ h hn1 Þ Shmin;n ðhN1 þ h hn Þ
3 n¼1 hmin;n N1
"
#
pffiffiffiffiffiffi
N
X
3
k þ 1 þ 2
2 2p
kn þ 1
kn þ 1
2
2
Shmin;n
KIC
ðhN1 þ hÞ ¼
K ðhn hn1 Þ þ
K
ðhn hn1 Þ
þ
8m
4mn p IC;n
8mn IC;n
3
n¼1
pffiffiffiffiffiffi k þ 1 þ
K
2p
4mp IC
(
The factor 2 in Eq.6 accounts for the right and left surfaces of the
fracture. And2uyy þ ðzÞ ¼ wþ ð0; zÞ. Therefore, the total energy
released when the upper fracture tip propagates from the bottom of
formation nþ1 to point h from the bottom of layer N is
Alternatively, the shear modulus and Poisson ratio can also be
represented with Heaviside functions, which will further make Eq.
(9) more complicated. Similarly, the effective fracture toughness for
the formations below the host formation can be estimated from
Eq. 10
"
#)
M
3
3
2 X
Shmin;n ðhM1 þ h hn1 Þ2 Shmin;n ðhM1 þ h hn Þ2
3 n¼1
"
#
pffiffiffiffiffiffi
M
X
3
k þ 1 2
2 2p
kn þ 1
k
þ
1
n
2
Shmin;n
KIC
ðhM1 þ hÞ ¼
K ðhn hn1 Þ2 þ
K
ðhn hn1 Þ
þ
8m
4mn p IC;n
8mn IC;n
3
n¼1
pffiffiffiffiffiffi k þ 1
K
2p
4mp IC
Pþ ¼
hnþ1
Z
0
(
hnþ2
Z
syy;nþ2
hnþ1
hN1
Z þh
…þ
hN1
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
knþ2 þ 1
KIc;nþ2 2pðhnþ2 hnþ1 zÞ dz þ …
4mnþ2 p
syy;N
(10)
2.2. Effective fracture toughness derivation without tensile stresses
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
k þ1
syy;nþ1 nþ1
KIc;nþ1 2pðhnþ1 zÞ dz
4mnþ1 p
þ
(9)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
kN þ 1
KIc;N 2pðh zÞdz
4mN p
As fracture propagates through each medium, the tensile forces
acting on the fracture edges, away from the tips reduces with crack
length. Thus, neglecting the tensile force contributions and
following the same procedure in Section 2.2 above, the effective
fracture toughness for the homogenized upper and lower barriers,
based on equivalent energy-release rate hypothesis are derived as
þ
(7)
Where, the Irwin (1957) criterion, KI ¼ KIC was implied in Eq. (7).
From Fig. 2, the equivalent energy released, as the upper tip
K IC ¼
PN
3
kn þ1
2
n¼1 mn Shmin;n KIC;n ðhn hn1 Þ
n
3
3
kþ1 PN S
2
2
n¼1 hmin;n ðhN1 þ h hn1 Þ ðhN1 þ h hn Þ
m
o
(11)
354
K IC
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
3
PM kn þ1
2
n¼1 mn Shmin;n KIC;n ðhn hn1 Þ
n
o
¼
3
3
P
M
kþ1
2
2
n¼1 Shmin;n ðhM1 þ h hn1 Þ ðhM1 þ h hn Þ
m
(12)
The fracture extension pressure may not only be affected by
gravity, but the relative positions of the fracture tips can affect the
value. Considering the three different cases in Fig. 3. The critical
extension pressures for the lower and upper tips in case 1 respectively, as derived in the Appendix are
0
1
1
BK
C
ffiffiffiffiffiffiffiffi þ S2;hmin ghx rA
PFD ¼ @qIC
2
phx
2.3. Minimum fracture extension pressures
When the pressure driving the hydraulic fracture is assumed to
be constant, the fracture extension pressure was derived by
Simonson et al. (1978) as
KIC
PF ¼ S2;hmin ðzÞ þ qffiffiffiffiffiffiffiffi
phx
(13)
where, hx is the half thickness of the host formation. But when the
displacements of the upper and lower tips of the fracture are significant, the hydrostatic pressure can significantly affect the
extension potential.
0
PFU
þ
(14)
1
1
BK
C
ffiffiffiffiffiffiffiffi þ S2;hmin þ ghx rA
¼ @qIC
2
phx
(15)
Noting that S2;hmin is the minimum principal in-situ stress at the
host formation. Cases 2 and 3 are not very common; but could be
relevant to the extension of existing fractures in formations. In case
2, the critical extension pressures for the upper and lower tips are
respectively derived in the Appendix as
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
#)
1 þ k3
1 þ k3
2KIC
þ k3
þ p S2;hmin
1 k3
1 k3
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
qffiffiffiffiffi
1 þ k3
1 þ k3
þ k3
2 hx
þ 1 k3 2 þ p
1 k3
1 k3
"rffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
qffiffiffiffiffi
1 k3
4 hx S3;hmin S2;hmin arcsin
2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
þ
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
qffiffiffiffiffi
1 þ k3
1 þ k3
2
þ k3
þ 1 k3 þ p
2 hx
1 k3
1 k3
"
(
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
qffiffiffiffiffi
p þ hx ghx pr þ 2 S3;hmin 1 k3 2 þ
þ pffiffiffi
PFU
Fig. 3. Possible fracture tips positions in three-layer media.
(16)
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
PFD
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(
"
!
#)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3
1 k3
1 k3
k3 þ 1 ghx pr þ 2 S3;hmin
ð1 þ k3 Þ
k þ
þ p S2;hmin
1 þ k3
1 þ k3 3
1 þ k3
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼
!
)
(
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3
1 þ k3 þ 1 k3
1 þ k3 p 1 þ k3
2 k3
1 þ k3
1 þ k3
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
"rffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3
pffiffiffi 1 þ k3
4 k3 þ 1 S3;hmin S2;hmin arcsin
2KIC
p
hx
2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(
!
)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 k3
1 þ k3 þ 1 k3
1 þ k3 p 1 þ k3
2 k3
1 þ k3
1 þ k3
h
x
Noting that, k3 ¼ h þh
, where hu is the displacement of the
u
x
upper tip. For case 3, the critical extension pressures for the upper
and lower tips are
355
(17)
Sometimes the natural fractures may be filled with materials
that have higher modulus than the rock matrix; on the other hand
the infill materials may have lower Young's modulus than the
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
#)
1 k1
1 k1
2KIC
þ k1
p S1;hmin
1 þ k1
1 þ k1
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
qffiffiffiffiffi
1 k1
1 k1
2
k1
þ 1 k1 þ p
2 hx
1 þ k1
1 þ k1
"rffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
qffiffiffiffiffi
1 þ k1
4 hx S2;hmin S1;hmin arcsin
2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
þ
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
qffiffiffiffiffi
1 k1
1 k1
2
k1
þ 1 k1 þ p
2 hx
1 þ k1
1 þ k1
(18)
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
rffiffiffiffiffi
p
1 þ k1
1 þ k1
KIC ghx pr S2;hmin
þ k1 S2;hmin
þ S1;hmin 1 k1 2 þ pS1;hmin
hx
1 k1
1 k1
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
1 þ k1
1 þ k1
2
þ k1
þ 1 k1 þ p
1 k1
1 k1
"rffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
"rffiffiffiffiffiffiffiffiffiffiffiffiffiffi #
1 k1
1 þ k1
ghx r arcsin
þ ghx r þ 2S2;hmin 2S1;hmin arcsin
2
2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
þ
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
qffiffiffiffiffi
1 þ k1
1 þ k1
þ k1
þ 1 k1 2 þ p
2 hx
1 k1
1 k1
(19)
"
(
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
qffiffiffiffiffi
p þ hx ghx pr þ 2 S2;hmin 1 k1 2
þ pffiffiffi
PFU
PFD
And k1 ¼ h
hx
.
d þhx
2.4. Effective fracture toughness for formations with disordered
natural fractures
The growth of hydraulic fractures in formations with natural
fractures depends on many factors, including the orientation of the
micro cracks, the infill materials in the cracks, the tensile strength
of the rock, treatment pressure, fluid viscosity, and many other
factors as studied by Zhou et al., 2015. To describe the random path
of the hydraulic fracture in disordered natural fracture network will
be very complicated. But the use of damage theory with the crack
closure method (described above) can approximately suggest the
containment potential of the hydraulic fracture.
matrix. The surface density of the cracks and its mechanical
strength affect the bearing capacity of the crack-matrix system. If
the cracks are filled with higher modulus materials, they help to
reinforce the strength of the crack-matrix system; but when the
cracks are empty or filled with lower modulus materials, there is a
reduction in the bearing capacity of the crack-matrix system. This is
the crux of the effective stress introduced by Kachanov (2013); in
the classical continuum damage theory, the inclusions are empty;
thus, this study will cover both opened and filled discontinuities.
For disordered natural fracture network, the impact of the inclinations of the natural fractures to the propagating hydraulic
fracture cannot be included in the model development in a deterministic way, except by the use of random theory, which we have
decided not to pursue in this paper.
As mentioned in the previous section, the effective fracture
356
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
Fig. 4. Region map of the tips positions when the pressure inside the fracture is (a) 53.09 MPa (7700 psi), (b) 53.78 MPa (7800 psi), and (c) 54.47MPa (7900 psi) for 11-EHP.
toughness that describes the containment potential of the hydraulic fracture, propagating in the rock with natural fractures, is
based on the use of equivalent energy released rate and equivalent
strain energy (by compressional loading) hypotheses. In the
equivalent system, the hydraulic fracture network is represented by
a mode-1 fracture propagating in a homogeneous medium; and
having the same energy released as the fracture propagates to the
same height or length. In the equivalent system, the energy
released due to mode-2 propagation of the hydraulic fracture is
disregarded: when hydraulic fracture is arrested by a natural fracture and consequently the natural fracture propagates from its
ends, if the natural fracture is inclined to the direction of maximum
in-situ stress in the plane, the propagation will be mixed mode.
The model development are based on the following
assumptions:
Assumption 1: The surface discontinuities do not evolve with
the applied loadings. This suggest that there is no need to
develop an evolution equation for the damage variable, which is
the ratio of the effective surface area of the discontinuities in a
representative volume element (RVE) and surface area of the
RVE.
Assumption 2: The surface discontinuities are isotopically
distributed.
Assumption 3: The rock behaves as a linearly elastic material.
Assumption 4: The natural fractures are filled with a material
that also has a linearly elastic response.
Table 1
Description of the formation properties and in-situ stress profiles for the 11- eleven
problem.
Layer
-
Top
Thickness
Stress
KIC
E
m.
MPa
pffiffiffiffiffi
MPa m
n
m.
MPa
e
e
-
-
-
-
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
-
Shale
Shale
Sand
Shale
Sand
Shale
Sand
Shale
Sand
Shale
Shale
Lithology
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
357
Fig. 5. Region map of the tips positions when the pressure inside the fracture is (a) 53.09 MPa (7700 psi), (b) 53.78 MPa (7800 psi), and (c) 54.47MPa (7900 psi) for 3ER-EHP.
Table 2
Description of the formation properties and in-situ stress profiles for the equivalent three-layer problem.
358
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
Fig. 6. Region map of the tips positions when the pressure inside the fracture is
54.47MPa (7900 psi) for 3BR-EHP.
The effective stress tensor defined by Kachanov for a material
body having empty inclusions is given as
s0 ¼
s
(20)
1f
But assuming the filling of the discontinuities reduces its
effective area relative to the matrix, then, the effective area of the
discontinuities is proposed as
Sf;e ¼ Sf ð1 lÞ
b
Fig. 8. Region map of the tips positions when the pressure inside the fracture is
54.47MPa (7900 psi) for 11-EHP, Case 2.
material has higher elastic modulus than the matrix, then the
overall elastic modulus of the inclusion-matrix system will be
higher; a reinforcement of the matrix. But the works by Eshelby
(1957), Hashin and Shtrikman (1963), Hill (1965), Mori and
Tanaka (1973), Willis (1981), Walpole (1981), Weng (1984), and
Chen et al. (1992) provide more rigorous studies on the impact of
inclusions on the elastic modulus of the host matrix. However, the
computation of the effective fracture toughness is demanding (Li
and Zhou, 2013).
As stated in Assumption 1, the damage variable is defined as
(21)
where l is the ratio of elastic moduli of the infill material and the
matrix, and b is a constant parameter. In a case where the infill
Fig. 7. Region map of the tips positions when the pressure inside the fracture is
54.47MPa (7900 psi) for 3WL-EHP.
D¼
Sf;e
¼ fð1 lÞb
S
(22)
Fig. 9. Region map of the tips positions when the pressure inside the fracture is
54.47MPa (7900 psi) for 3ER-EHP, Case 2.
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
359
Fig. 10. Region map of the tips positions when the pressure inside the fracture is
54.47MPa (7900 psi) for 3BR-EHP, Case 2.
Fig. 12. Region map of the tips positions when the pressure inside the fracture is
54.47MPa (7900 psi) for 3EER-EHP, Case 2.
and using the crack closure method, the energy released in the
damaged rock material is equivalent to the energy released in a
virgin rock with the normal stress acting on the face of the fracture
being the effective stress, i.e.
u0yy ¼
k0 þ 1 0 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2pðz0 zÞ
K
4m0 p IC
(24)
and,
K0
IC
s0yy ¼ S0hmin þ pffiffiffiffiffiffiffiffi
2pz
P¼2
Zz0
0
1
syy uyy dz ¼ 2
2
Zz0
1 0 0
s u dz
2 yy yy
(23)
uyy ¼
0
Where,
(25)
k þ 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K
2pðz0 zÞ
4mp IC
(26)
K
IC
syy ¼ Shmin þ pffiffiffiffiffiffiffiffi
2pz
(27)
by substituting equations 22e25 into (21), the effective fracture
toughness can be determined from the algebraic expression
K 0IC p
3
k0 þ 1 0
2
z0
K IC S0hmin z0 2 þ pffiffiffiffiffiffi
0
3
m
2p 2
3
kþ1
2
KIC p
z0
¼
KIC Shmin z0 2 þ pffiffiffiffiffiffi
3
m
2p 2
=
=
(28)
Table 3
Description of the formation properties and in-situ stress profiles for the 11-EHP,
Case 2.
Layer
Fig. 11. Region map of the tips positions when the pressure inside the fracture is
54.47MPa (7900 psi) for 3WL-EHP, Case 2.
-
Top
Thickness
Stress
KIC
E
m.
MPa
pffiffiffiffiffi
MPa m
n
m.
MPa
e
e
-
-
-
-
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
29,510
-
Shale
Sand
Shale
Shale
Shale
Shale
Sand
Shale
Sand
Shale
Shale
Lithology
360
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
2.5. Effective fracture toughness for formations with ordered
natural fractures
When the natural fractures are oriented in the same direction,
the path of the main hydraulic fracture stem can be easily traced
compared to the disordered pattern. In this case the energy
released as the fracture propagates in both modes 1 and 2 is
P¼2
Z
c
1
syy uyy ds þ
2
Z
c2
1
2N sxy uxx ds2 ¼ 2
2
Zz0
1 0 0
s u dz
2 yy yy
0
(32)
Fig. 13. Variations of the normalized shear modulusmD , and normalized fracture
toughness with damage in any material displaying a linear elastic response. l ¼ 0 in
this case.
Since damage prior to loading is under consideration and not
damage induced by the loading; thus, considering a compression
loading of the RVE. The elastic strain energy (compressional
loading) and the effective elastic energy are equivalent, i.e.
1
2
Z
sεdv ¼
B
1
2
Z
s0 ε0 dv0
(29)
B0
Hence, the relationship between the damaged and virgin shear
moduli is
h
i2
m ¼ m0 1 fð1 lÞb
(30)
The constant parameterb, can be determined by curve-fitting
the effective shear modulus in Eq. (30) with Mori-Tanaka's model
or from experiment.
When the contributions of the tensile stresses along the fracture
edges are disregarded, the damaged fracture toughness of the formation becomes
KIC ¼ K 0IC
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
b
1 fð1 lÞ
(31)
Where, c is the path of the main hydraulic fracture, and c2 is the
distance the fracture propagates in mixed mode. N is the number of
repeated pattern of the natural fractures; in other words, the
number of rows. Eq. (32) is derived based on the assumption that
the natural fracture will propagate from one of its ends, when it
arrests the hydraulic fracture.
uxx ¼
k þ 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2pðz0 zÞ
K
4mp IIC
K
IIC
sxy ¼ pffiffiffiffiffiffiffiffi
2pz
(33)
(34)
Depending on the projected path, the steps in Section 2.3 can be
followed in estimating the effective fracture toughness from
equation (32).
3. Numerical examples
A hydraulic fracture stands in layer 6, with its tips at the top and
bottom of the formation, which is 9 m (30 ft) thick. Using the full
multi-layer equilibrium height model, the tips displacements at
different internal pressures are shown in Fig. 4a b, and c. Table 1
shows the description of the layers. While the use of the energyreleased-based effective fracture toughness for the same range of
pressures are shown in Fig. 5a, b, and c.
The equivalent three-layer system, based on equivalent energy
release rate hypothesis, (3ER-EHP) to the eleven-layer problem (11EHP) is shown in Table 2 below. Comparing the region maps in
Fig. 4a,b and c and Fig. 5a and b, and 5c, it is evident that the two
solutions are equivalent. Points A, B, and C in Fig. 4a, b, 5a, and 5b
Fig. 14. (a) Variation of the normalized shear modulusmD , considering different infill materials (b) Variation of normalized fracture toughness, KIC;D , with damage considering
different infill materials.
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
Table 4
Elastic properties of the matrix and inclusions.
Case
1
2
3
Matrix
Poisson
Ratio
Infill
Poisson
Ratio
Matrix
Young's
Mod.
Infill
Young's
Mod.
GPa.
GPa
e
e
-
0
5.0
25.0
-
-
are not the practical solutions to the nonlinear equation; at these
pressure values, the initial upper and lower tips are stable. In
Figs. 4c and 5c Point B is the pair solution to the practical solution A.
From the map, the upper tip is located at depth 2790 m, while the
lower tip is located at depth 3005 m.
Using the linear blend rule and weakest link arguments, the
effective fracture toughness for the upper and lower barriers are
1.11 MPa and 1.1 MPa respectively. The locations of the tips in these
equivalent systems are similar to the 3ER-EHP (Fig. 6 and 7); for
shorthand writing, the equivalent three layer problem based on
blend rule is 3BR-EHP, and that based on weakest link arguments is
3WL-EHP. In this case, both 3ER-EHP and 3WL-EHP yielded similar
results, while 3BR-EHP and 3ERR-EHP gave more accurate positions
of the tips relative to the other homogenization methods
(Figs. 8e12).
From these examples, it is evident that the use of energy-release
361
rate hypothesis for homogenizing multilayer problems may not be
accurate (to a small degree), except when the contributions of the
tensile stresses are neglected.
Table 3 shows the description of eleven layers with significant
modulus contrasts. And the effective fracture toughness for the
upper and lower barriers using the equivalent energy release rate
hypothesis are 1.44MPa and 1.59 MPa respectively. While the
effective fracture toughness for the upper and lower homogenized
barriers when using the linear blend rule are 4.17 MPa and
5.46 MPa respectively; and using the weakest point arguments, the
effective fracture toughness are 1.6MPa and 3.28MPa respectively.
In the same vein, the effective fracture toughness using the reduced
equivalent energy release rate hypothesis (3ERR-EHP) are 3.38MPa
and 4.15 MPa respectively.
Alternatively, the use of equivalent strain energy hypothesis,
which is mathematically tedious, may be used in lieu of the energy
released rate; this comparison will not be pursued in this paper.
The tensile stress is greatest at the tips of the fracture, and reduces along the length of the fracture. Along the edges of the
fracture, away from the tips, the Irwin criterion, KI ¼ KIC , does not
apply; but KI < KIC . Thus, neglecting the tensile stress contribution
in the crack closure energy function will yield more accurate results; and this is evident in the two example cases.
Fig. 13 shows that the impact of damage on the elastic properties
of the formation matrix do not follow the same path. The Young's
and shear moduli are more severely damaged than the fracture
toughness. And based on isotropic damage assumption, the Poisson
Fig. 15. Comparing the performance of the proposed model for effective shear modulus of a material body having (a) empty or open natural fractures, Case 1, (b) micro-fractures
filled with lower modulus material, Case 2, and (c) micro-fractures filled with higher modulus material, Case 3.
362
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
Appendix. Derivation of minimum fracture propagation for
three-layer problem
ratio will not be affected by the presence of the discontinuities
(damage), based on the continuum damage theory; the impact of
damage on other elastic properties are beyond the scope of this
article. And Fig. 14a and b, show the variations of the effective shear
modulus and fracture toughness when the discontinuities (inclusions, or natural fractures) are filled with material having lower
or higher Young's modulus.
To compare the predictions of the proposed effective shear
modulus for the damaged/reinforced material with Mori-Tanaka's
model, Table 4 shows the three cases:
As observed in Fig. 15a, the prediction from the proposed
phenomenological model mimics the rigorous Mori-Tanaka model.
In the second and third cases, the values of b are 2.5 and 1.2
respectively.
Sneddon and Elliott (1946) derived the stress intensity factor at
the tip of a fracture subjected to a constant internal pressure.
Modifying the equation by including the effect of gravity, the stress
intensity factor at the bottom tip, when the fracture is contained in
formation n only:
KI;D
1
¼ pffiffiffiffiffi
pl
Let z* ¼
4. Summary
Using equivalent energy release-rate hypothesis, an effective
fracture toughness was derived to homogenize heterogeneous
layered media. Homogenizing heterogeneous layered media into a
single layer can reduce multi-layer equilibrium-height problem to
the classical three-layer equilibrium-height problem. And the
reduction of the multi-layer problem to the classical three-layer
problem reduces the model complexity. The predictions from the
proposed model have the same range of accuracy as the wellknown linear blend rule; while the predictions from the weakest
link arguments method were also observed to be inaccurate.
Furthermore, an effective fracture toughness formulation for
predicting the growth of hydraulic fracture in a formation with
micro-cracks or any other form of ordered or disordered inclusions
was developed. The formulation is based on the use of damage
theory, equivalent energy-release rate, and equivalent strainenergy hypotheses. The presence of infill materials in the microcracks (closed natural fractures) will affect the overall elastic
properties of the damaged/reinforced rock-fracture system. And
using a phenomenological approach, which is based on the use of
an effective area, the proposed model performs very well as the
rigorous Mori-Tanaka model, when the fractures are open. Then
using the Mori-Tanaka model for the effective shear modulus, as
reference, the constant parameter in the proposed model can be
determined; subsequently, estimating the effective fracture
toughness for the rock-closed fracture system. The benefit of the
proposed model over the Mori-Tanaka is its simplicity in computing
the effective fracture toughness.
And at
Zz
l
rffiffiffiffiffiffiffiffiffiffi
lþz
pðz; tÞ sn
dz
lz
z
l
(A.1)
(A.2)
z ¼ z; z* ¼ z*
z ¼ l; z* ¼ 1
dz* 1
¼
dz
l
(A.3)
(A.4)
Substituting (A.3) and (A.4) into (A.1),
KI;D ¼
rffiffiffi Zz*
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ z*
l
dz*
pQ ðtÞ þ rglz* sn
1 z*
p
(A.5)
1
Solving (A.5) and substituting (A.2) back yields
n
pffiffiffiffiffiffiffiffiffiffi
1
KI;D ¼ pffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðl zÞ l þ zðgð2l þ zÞr 2sn Þ
2 p lðl zÞ
0qffiffiffiffiffiffi1
lþz
pffiffiffiffiffiffiffiffiffiffi
l
1 @
pffiffiffi A
þ 2l l z ðglr 2sn Þsin
2
0
0qffiffiffiffiffiffi119
lþz
=
pffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffi
l
1
þ pQ @2ðl þ zÞ l þ z þ 4l l z sin @ pffiffiffi AA
;
2
(A.6)
Acknowledgement
lim KI;D ¼
z/l
This research is primarily supported by the Blowout Risk
Assessment Joint Industry Projects.
pffiffiffiffiffi
lp
glr þ 2pQ 2sn
2
(A.7)
Therefore, the critical pressure at the mid position Q to cause
the lower tip to move is
Nomenclature
D
GIþ
GI
hu
hd
KIþ
KI
K IC
K IC
PFU
þ
damage variable
energy release rate in upper barriers
energy release rate in lower barriers
displacement of the upper tip
displacement of the lower tip
stress intensity factor at upper tip
stress intensity factor at lower tip
effective fracture toughness for upper homogenized upper barriers
mn
mu
md
l
effective fracture toughness for homogenized lower barriers
f
minimum fracture extension pressure for upper tip
PFD
Sh;min
uyy þ
uyy
minimum fracture extension pressure for lower tip
minimum horizontal in-situ stress
fracture half width in the upper barriers
fracture half width in the lower barriers
formation, n, shear modulus
Averaged shear modulus for upper barriers
Averaged shear modulus for lower barriers
ratio of inclusion and undamaged matrix elastic moduli
Ratio of the surface area of discontinuity to surface area of representative
volume element
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
pQ ¼ PFD ¼
KIC;D
1
pffiffiffiffiffi þ sn glr
2
pl
(A.8)
k3 approaches unity, the fracture extension pressure for the upper
tip corresponds to case 3 in Fig. 3.
Similarly, the stress intensity factor at the lower tip is
At the upper tip, the stress intensity factor is
KI;U
1
¼ pffiffiffiffiffi
pl
Zz
l
sffiffiffiffiffiffiffiffiffiffi
lz
dz
pðz; tÞ sn
lþz
KI;D
(A.9)
And at
(A.10)
(A.11)
(A.12)
Substituting (A.11) and (A.12) into (A.9),
rffiffiffi Z1
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ z*
l
dz*
pQ ðtÞ rglz* sn
1 z*
p
(A.13)
z*
Solving (A.13) and substituting (A.2) back yields
1
KI;U ¼ pffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 p lðl zÞ
ðl zÞ
qffiffiffiffiffiffiffiffiffiffiffiffiffi
ðl þ zÞðgð2l þ zÞr þ 2sn Þ
0qffiffiffiffiffiffi1
lþz
qffiffiffiffiffiffiffiffiffiffiffiffiffi
l
1 @
pffiffiffi A
2l ðl zÞ ðglr þ 2sn Þcos
2
0
0qffiffiffiffiffiffi119
lþz
qffiffiffiffiffiffiffiffiffiffiffiffiffi
qffiffiffiffiffiffiffiffiffiffiffiffiffi
=
l
þ 2pQ @ðl zÞ ðl þ zÞ þ 2l ðl zÞ cos1 @ pffiffiffi AA
;
2
(A.14)
lim KI;U ¼
z/l
pffiffiffiffiffi
lp
glr þ 2pQ 2sn
2
(A.15)
Therefore, the critical pressure at midpoint Q to cause the upper
tip to move is
KIC;U
1
pffiffiffiffiffi þ sn þ glr
2
pl
(A.16)
Noting that l ¼ hx . Hence, extending the derivation to three
layer: the stress intensity factor at the upper tip is
KI;U
8
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
> Z1
1 þ z*
1 <
dz*
¼ pffiffiffiffiffi
pQ ðtÞ rglz* s3
1 z*
pl >
:
k3
Zk3
þ
k1
Zk1
þ
z*
k3
ZZ
z ¼ z; z* ¼ z*
z ¼ l; z* ¼ 1
pQ ¼ PFU ¼
1
þ
dz*
1
¼
dz
l
KI;U ¼
8
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
> Zk3
1 þ z*
1 <
dz*
¼ pffiffiffiffiffi
pQ ðtÞ þ rglz* s3
1 z*
pl >
:
Zk1
z
Let z* ¼
l
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ z*
dz*
pQ ðtÞ rglz* s2
1 z*
(A.17)
9
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
>
=
1 þ z*
dz*
pQ ðtÞ rglz* s1
>
1 z*
;
z/l
Therefore, the minimum fracture extension pressure for the
upper tip at any location z is pQ in (A.17); since the algebraic
expression is complicated, it will not be displayed in this paper. In
the same vein, when k1 approaches unity, the fracture extension
pressure for the upper tip corresponds to case 2 in Fig. 3. And when
363
þ
k1
pQ ðtÞ þ rglz* s2
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ z*
dz*
1 z*
(A.18)
9
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
>
=
1 þ z*
dz*
pQ ðtÞ þ rglz* s1
>
1 z*
;
z/l
and when k1 approaches unity, the fracture extension pressure for
the lower tip corresponds to case 3 in Fig. 3. While when k3 approaches unity, the fracture extension pressure for the lower tip
corresponds to case 2 in Fig. 3.
References
Adachi, J.I., Detournay, E., Peirce, A.P., 2010. Analysis of the classical pseudo-3D
model for hydraulic fracture with equilibrium height growth across stress
barriers. Int. J. Rock Mech. Min. Sci. 47 (4), 625e639.
Advani, S.H., Lee, T.S., Lee, J.K., 1990. Three-dimensional modeling of hydraulic
fractures in layered media: part Idfinite element formulations. J. Energy
Resour. Technol. 112 (1), 1e9.
Ahmed, U., 1984, January. A practical hydraulic fracturing model simulating
necessary fracture geometry, fluid flow and leakoff, and proppant transport. In:
SPE Unconventional Gas Recovery Symposium. Society of Petroleum Engineers.
Atkins, A.G., 1979. Cracking of layered structures: the suppression of yielding by cast
hardening. Mech. Behav. Mater. 3, 341e349.
Atkins, A.G., Mai, Y.W., 1985. Elastic and Plastic Fracture: Metals, Polymers, Ceramics, Composites, Biological Materials. Halsted Press, Ellis Horwood.
Beremin, F.M., Pineau, A., Mudry, F., Devaux, J.C., D'Escatha, Y., Ledermann, P., 1983.
A local criterion for cleavage fracture of a nuclear pressure vessel steel. Metall.
Trans. A 14 (11), 2277e2287.
Broek, D., 1991. Elementary Engineering Fracture Mechanics, 4th Revised Edition.
Kluwer Academic Publishers, Dordrecht, Netherlands.
Bui, H.D., 1977. An integral equations method for solving the problem of a plane
crack of arbitrary shape. J. Mech. Phys. Solids 25 (1), 29e39.
Chen, T., Dvorak, G.J., Benveniste, Y., 1992. Mori-Tanaka estimates of the overall
elastic moduli of certain composite materials. J. Appl. Mech. 59 (3), 539e546.
Chuprakov, D., Melchaeva, O., Prioul, R., 2014. Injection-sensitive mechanics of
hydraulic fracture interaction with discontinuities. Rock Mech. Rock Eng. 47 (5),
1625e1640.
Clifton, R.J., Abou-Sayed, A.S., 1979, January. On the computation of the threedimensional geometry of hydraulic fractures. In: Symposium on Low Permeability Gas Reservoirs. Society of Petroleum Engineers.
Clifton, R.J., Abou-Sayed, A.S., 1981, January. A variational approach to the prediction
of the three-dimensional geometry of hydraulic fractures. In: SPE/DOE Low
Permeability Gas Reservoirs Symposium. Society of Petroleum Engineers.
Dahi-Taleghani, A., Olson, J.E., 2011. Numerical modeling of multistrandedhydraulic-fracture propagation: accounting for the interaction between
induced and natural fractures. SPE J. 16 (03), 575e581.
Economides, M.J. (Ed.), 2000. Reservoir Stimulation, vol. 18. Wiley, Chichester.
Eriksson, K., 1998. The effective fracture toughness of structural components obtained with the blend rule. Nucl. Eng. Des. 182 (2), 123e129.
Eriksson, K., Atkins, A.G., 1995. The effective through crack fracture toughness of
structural components of older inhomogeneous steels. Mater. Ageing Compon.
Life Ext. 1, 147e154.
Eshelby, J.D., 1957, August. The determination of the elastic field of an ellipsoidal
inclusion, and related problems. In: Proceedings of the Royal Society of London
a: Mathematical, Physical and Engineering Sciences, vol. 241. The Royal Society,
pp. 376e396. No. 1226.
Fung, R.L., Vilayakumar, S., Cormack, D.E., 1987. Calculation of vertical fracture
containment in layered formations. SPE Form. Eval. 2 (04), 518e522.
Gao, H., Rice, J.R., 1989. A first-order perturbation analysis of crack trapping by
arrays of obstacles. J. Appl. Mech. 56 (4), 828e836.
Griffith, A.A., 1921. The phenomena of rupture and flow in solids. Philosophical
Trans. R. Soc. Lond. Ser. A, Contain. Pap. a Math. or Phys. character 221, 163e198.
Gu, H., Weng, X., 2010, January. Criterion for fractures crossing frictional interfaces
at non-orthogonal angles. In: 44th US Rock Mechanics Symposium and 5th USCanada Rock Mechanics Symposium. American Rock Mechanics Association.
Gu, H., Weng, X., Lund, J.B., Mack, M.G., Ganguly, U., Suarez-Rivera, R., 2012. Hydraulic fracture crossing natural fracture at nonorthogonal angles: a criterion
364
O. Oyedokun, J. Schubert / Journal of Natural Gas Science and Engineering 44 -e364
and its validation. SPE Prod. Operations 27 (01), 20e26.
Hashin, Z., Shtrikman, S., 1963. A variational approach to the theory of the elastic
behaviour of multiphase materials. J. Mech. Phys. Solids 11 (2), 127e140.
Heerens, J., Zerbst, U., Bauschke, H.M., Schwalbe, K.H., 1994. On the Application of
the Weakest Link Model in the Lower Shelf and Transition, 2013, February. In
ECF10, Berlin.
Hill, R., 1965. A self-consistent mechanics of composite materials. J. Mech. Phys.
Solids 13 (4), 213e222.
Hirth, J.P., Lothe, J., 1968. Theory of Dislocations, 780 pp.
Irwin, G.R., 1957. Analysis of stresses and strains near the end of a crack traversing a
plate. Spie Milest. Ser. MS 137 (167e170), 16.
Irwin, G.R., 1958. Fracture in “Handbuch der Physik,”, vol. V.
Iwadate, T., Tanaka, Y., Ono, S., Watanabe, J., 1983, January. An analysis of elasticplastic fracture toughness behavior for J IC measurement in the transition region. In: Elastic-plastic Fracture: Second Symposium, Volume II Fracture
Resistance Curves and Engineering Applications. ASTM International.
Kachanov, L., 2013. Introduction to Continuum Damage Mechanics, vol. 10. Springer
Science & Business Media.
Landes, J.D., Shaffer, D.H., 1980. Statistical characterization of fracture in the transition region. In: Fracture Mechanics. ASTM International.
Li, Y., Zhou, M., 2013. Prediction of fracture toughness of ceramic composites as
function of microstructure: I. Numerical simulations. J. Mech. Phys. Solids 61
(2), 472e488.
Meyer, B.R., 1986, January. Design formulae for 2-D and 3-D vertical hydraulic
fractures: model comparison and parametric studies. In: SPE Unconventional
Gas Technology Symposium. Society of Petroleum Engineers.
Morales, R.H., Brady, B.H., Ingraffea, A.R., 1993, January. Three-dimensional analysis
and visualization of the wellbore and the fracturing process in inclined wells.
In: Low Permeability Reservoirs Symposium. Society of Petroleum Engineers.
Mori, T., Tanaka, K., 1973. Average stress in matrix and average elastic energy of
materials with misfitting inclusions. Acta metall. 21 (5), 571e574.
Newberry, B.M., Nelson, R.F., Ahmed, U., 1985, January. Prediction of vertical hydraulic fracture migration using compressional and shear wave slowness. In:
SPE/DOE Low Permeability Gas Reservoirs Symposium. Society of Petroleum
Engineers.
Nordgren, R.P., 1972. Propagation of a vertical hydraulic fracture. Soc. Petroleum
Eng. J. 12 (04), 306e314.
Olson, J.E., Taleghani, A.D., 2009, January. Modeling simultaneous growth of multiple hydraulic fractures and their interaction with natural fractures. In: SPE
Hydraulic Fracturing Technology Conference. Society of Petroleum Engineers.
Palmer, I.D., Carroll Jr., H.B., 1983a. Three-dimensional hydraulic fracture propagation in the presence of stress variations. Soc. Petroleum Eng. J. 23 (06),
870e878.
Palmer, I.D., Carroll Jr., H.B., 1983b, January. Numerical solution for height and
elongated hydraulic fractures. In: SPE/DOE Low Permeability Gas Reservoirs
Symposium. Society of Petroleum Engineers.
Palmer, I.D., Craig, H.R., 1984, January. Modeling of asymmetric vertical growth in
elongated hydraulic fractures and application to first MWX stimulation. In: SPE
Unconventional Gas Recovery Symposium. Society of Petroleum Engineers.
Perkins, T.K., Kern, L.R., 1961. Widths of hydraulic fractures. J. Petroleum Technol. 13
(09), 937e949.
Renshaw, C.E., Pollard, D.D., 1995, April. An experimentally verified criterion for
propagation across unbounded frictional interfaces in brittle, linear elastic
materials. Int J. Rock Mechanics. Min Sci. Geomechanics. Abstr 32 (No. 3),
237e249. Pergamon.
Settari, A., Cleary, M.P., 1986. Development and testing of a pseudo-threedimensional model of hydraulic fracture geometry. SPE Prod. Eng. 1 (06),
449e466.
Simonson, E.R., Abou-Sayed, A.S., Clifton, R.J., 1978. Containment of massive hydraulic fractures. Soc. Petroleum Eng. J. 18 (01), 27e32.
Slatcher, S., 1986. A probabilistic model for lower-shelf fracture toughnessdtheory
and application. Fatigue & Fract. Eng. Mater. Struct. 9 (4), 275e289.
Sneddon, I.N., Elliott, H.A., 1946. The opening of a Griffith crack under internal
pressure. Q. Appl. Math. 4 (3), 262e267.
Valko, P.P., Liu, S., 2015. February. An improved equilibrium-height model for predicting hydraulic fracture height migration in multi-layered formations. In: SPE
Hydraulic Fracturing Technology Conference. Society of Petroleum Engineers.
€rro
€nen, K., 1984. Statistical model for carbide induced brittle
Wallin, K., Saario, T., To
fracture in steel. Metal Sci. 18 (1), 13e16.
Walpole, L.J., 1981. Elastic behavior of composite materials: theoretical foundations.
Adv. Appl. Mech. 21, 169e242.
Warpinski, N.R., Teufel, L.W., 1987. Influence of geologic discontinuities on hydraulic
fracture propagation (includes associated papers 17011 and 17074). J. Petroleum
Technol. 39 (02), 209e220.
Weng, G.J., 1984. Some elastic properties of reinforced solids, with special reference
to isotropic ones containing spherical inclusions. Int. J. Eng. Sci. 22 (7),
845e856.
Willis, J.R., 1981. Variational and related methods for the overall properties of
composites. Adv. Appl. Mech. 21, 1e78.
Wu, K., 2014. Mechanics Analysis of Interaction between Hydraulic and Natural
Fractures in Shale Reservoirs. Unconventional Resources Technology Conference (URTEC).
Zhou, J., Chen, M., Jin, Y., Zhang, G.Q., 2008. Analysis of fracture propagation
behavior and fracture geometry using a tri-axial fracturing system in naturally
fractured reservoirs. Int. J. Rock Mech. Min. Sci. 45 (7), 1143e1152 (SS).
Zhou, J., Huang, H., Deo, M., 2015, November. Modeling the interaction between
hydraulic and natural fractures using dual-lattice discrete element method. In:
49th US Rock Mechanics/Geomechanics Symposium. American Rock Mechanics
Association.