Вы находитесь на странице: 1из 9

ARMA 16-494

Modeling of near-wellbore fracture reorientation


using a fluid-coupled 2D XFEM algorithm
Gordeliy, E.
Schlumberger-Doll Research Center, Cambridge, MA, USA
Abbas, S.
Schlumberger, Sugar Land, TX, USA
Prioul, R.
Schlumberger-Doll Research Center, Cambridge, MA, USA

Copyright 2016 ARMA, American Rock Mechanics Association


This paper was prepared for presentation at the 50th US Rock Mechanics / Geomechanics Symposium held in Houston, Texas, USA, 26-29 June
2016. This paper was selected for presentation at the symposium by an ARMA Technical Program Committee based on a technical and critical
review of the paper by a minimum of two technical reviewers. The material, as presented, does not necessarily reflect any position of ARMA, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper for commercial purposes without the written
consent of ARMA is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 200 words; illustrations may not be
copied. The abstract must contain conspicuous acknowledgement of where and by whom the paper was presented.

ABSTRACT: We have developed a fluid-coupled algorithm for modeling reorientation of longitudinal fractures propagating from
a wellbore aligned with the intermediate or maximum far-field stress direction. A two-dimensional plane-strain model is adopted.
The rock deformation solver is implemented using the extended finite element method (XFEM). The model for the fluid flow in
hydraulic fractures is based on the Reynolds lubrication equation. The fluid pressure and the fracture opening are obtained from the
solution of a system of nonlinear equations resulting from the coupling of the rock deformation solver with the fluid flow model.
To model fracture reorientation, the direction of fracture propagation is determined based on the maximum tensile stress criterion.
We present the results of modeling of the near-wellbore fracture reorientation in the viscosity-dominated regime with this algorithm
and study the effect of the wellbore casing on the fracture opening and the fluid pressure required to initiate and propagate the
fracture.
formation exposed to the fluid pressure. For example, if
1. INTRODUCTION
perforations in a cased wellbore are misaligned with the
During the hydraulic fracturing process, a hydraulic in-situ principal stress directions, a lower misalignment
fracture initiates from an initial flaw at a wellbore and angle results in a lower viscous pressure drop near the
ultimately propagates in the preferred fracture plane wellbore and a lower pressure required to propagate the
orthogonal to the minimum compressive far-field stress fracture (Abass et al., 1994; Zhang et al., 2011).
direction. If the initial flaw is not orthogonal to the
In this work, we aim to develop a computational tool to
minimum compressive far-field stress, reorientation of
study the implications of the near-wellbore fracture
the hydraulic fracture takes place, leading to one form of
reorientation onto the fluid pressure required to
near-wellbore fracture tortuosity. In this case, the
propagate the fracture, including conditions when a
fracture opening near the wellbore is reduced, which
highly-viscous fluid is injected into a fracture at a
generates a near-wellbore pressure loss and creates a risk
relatively high injection rate. In the present work, we
of wellbore screenout due to proppant bridging (Mack
consider a rock of low permeability. In this case, the
and Warpinski, 2000).
effect of poro-elasticity does not contribute to the first-
Previous studies (Romero et al., 1995; Zhang et al., order characteristics of fracture propagation. When the
2011), as well as observed field data and experiments energy spent to drive the viscous fluid in the fracture
(Aud et al., 1994; Van de Ketterij and de Pater, 1999), dominates the energy spent for the creation of the new
have shown that using a fluid of a higher viscosity and a fracture surfaces, the fracture propagates in the
higher injection rate reduces the negative effects of this viscosity-dominated regime. To model this propagation
form of near-wellbore fracture tortuosity. However, the regime properly, the computational model must be fully
increased viscosity and rate require higher pressure to coupled with the fluid flow in the fracture and accurately
initiate and propagate the fracture. capture the behavior of the fluid pressure near the
Other factors that can influence reduced fracture width fracture tip.
and the near-wellbore pressure loss due to the near- Thus we have developed a fluid-coupled algorithm for
wellbore fracture reorientation include completion type modeling reorientation of longitudinal fractures
and the geometry of the initial defect of the rock propagating from a wellbore aligned with the
intermediate or maximum compressive far-field stress the tension-positive stress convention throughout this
direction. A two-dimensional plane-strain model is paper. The principal (confining) stresses are chosen to be
adopted. The rock deformation solver is based on linear xx and yy , such that xx yy . The initial defect is
elasticity and is implemented using the extended finite
element method (XFEM). The XFEM is a finite element inclined at an angle with respect to the maximum
based method that can simulate fractures propagating principal compressive stress xx . The Youngs modulus,
through a predefined finite element mesh. It represents Poissons ratio and the fracture toughness of rock are
the discontinuous and singular elastic fields, associated denoted, correspondingly, E, and K Ic . The hydraulic
with fractures, using an enriched shape function space
that includes discontinuous and singular functions. The fracture propagates under the injection of a fluid of
model for the fluid flow in hydraulic fractures is based viscosity at the injection rate Qo . We assume that the
on the Reynolds lubrication equation. The fluid pressure fronts of the fluid and the fracture coincide.
and the fracture opening are obtained from the solution
of a system of nonlinear equations resulting from the Throughout the paper, we will use the following scaled
coupling of the rock deformation solver with the fluid material parameters:
flow model. To model fracture reorientation, the E
direction of fracture propagation is determined based on K 32
K Ic , E , 12
1 2
the maximum tensile stress criterion.
Numerical studies of the near-wellbore tortuosity in 2D
were previously performed by means of different
methods, e.g., models based on the displacement
discontinuity method (DDM) (Zhang et al., 2011) and
the finite differences method (Cherny et al., 2009a,
2009b). The XFEM was used for modeling the near-
wellbore fracture tortuosity by Sepehri et al. (2015);
however, their model does not include coupling between
the rock deformation solution, provided by the XFEM,
and the fluid flow. Such model can only be applied in
the toughness-dominated regime, where the energy spent
on the creation of the fracture surfaces dominates the
energy required to drive the viscous fluid through the
fracture. A fluid-coupled XFEM model, based on the Fig. 1. Wellbore in a rock formation with initial defects
finite element software Abaqus, was used by Wang misaligned with the maximum principal compressive stress
(2015) to model near-wellbore fracture reorientation. xx .
Their work focused on poro-elastic and plastic effects on
the fracture reorientation and did not consider viscosity-
dominated propagation. 2.2. XFEM solver for rock deformation
The rock deformation solver is based on linear elasticity
The present XFEM model is fully coupled with the fluid and is implemented using the extended finite element
flow in the fracture and can address toughness- method (XFEM). The XFEM is a finite element based
dominated and viscosity-dominated regimes of fracture method that can simulate fractures propagating through a
propagation in low-permeability rocks. We present the predefined finite element mesh (Mos et al., 1999;
results of modeling of the near-wellbore fracture Stolarska et al., 2001; Fries and Belytschko, 2010) and
reorientation in the viscosity-dominated regime with this thus does not require remeshing. It represents the
algorithm and study some aspects of the effect of the discontinuous and singular elastic fields associated with
wellbore casing on the fracture opening and the fluid fractures, using an enriched shape function space that
pressure required to propagate the fracture. includes discontinuous and singular functions.

2. COUPLED MODEL For the present model, we use the mixed XFEM
formulation developed by Gordeliy and Peirce (2013)
2.1. Problem formulation for the solution of hydraulic fracture problems in
We consider a two-dimensional plane-strain model of a different propagation regimes. The discontinuous and
longitudinal hydraulic fracture propagating from a singular displacement and stress fields associated with a
circular wellbore of radius R. The fracture is initiated at hydraulic fracture are represented by linear combinations
an initial defect of the rock formation, such as a of standard finite-element shape functions, discontinuous
perforation tunnel or a slot or at a pair of symmetric sign enrichment functions and the singular crack-tip
defects for a bi-wing fracture (see Figure 1). We adopt enrichment that takes into account the specific power
law singularity associated with the viscosity-dominated wellbore pressure pw . For a bi-wing fracture, this is
regime of fracture propagation. expressed via
For the examples in this paper, we consider fracture 2

propagation in the viscosity-dominated regime. The q


i 1
i Qo , p f , i pw , i 1, 2 (4)
mixed XFEM formulation directly takes into account the
asymptotic behavior of the crack width consistent with
the viscosity-dominated propagation. The boundary The boundary conditions at the fracture tip prescribe
conditions along the crack prescribe the fluid pressure in vanishing crack width and fluid flux.
the so-called channel region far away from the fracture
tips and prescribe the crack width near the crack tip The above fluid flow equations are discretized for the
according to the viscous crack-tip asymptote. nodal values of the crack width and the fluid pressure
using a finite volume scheme as it is done in Gordeliy
The boundary conditions for the elasticity problem and Peirce (2013). The fracture is split into the channel
within the XFEM solver include the traction boundary and the fracture tip region. The boundary conditions at
condition along the wellbore wall. For an openhole the fracture tip for the fluid flux and the crack width are
completion, the normal traction component n on the satisfied by using the viscous crack-tip asymptote for the
wellbore wall is equal to the negative of the wellbore computation of the fracture tip volume.
pressure:
For a given fracture configuration and a given fluid
n pw (1) pressure in the channel, the XFEM solver provides the
corresponding crack width in the channel. Thus, the
To account for the pressurized wellbore, we decompose discretized nonlinear fluid flow equations can be
the total stress field into a superposition of (i) the stress reformulated in terms of the nodal fluid pressures by
wb, due to the pressurized wellbore when no crack is eliminating the crack width. These equations for nodal
present, and (ii) the stress due to the crack in the rock pressures are solved iteratively using either a Newtons
formation with a traction-free borehole. For an openhole algorithm or a fixed-point iteration.
wellbore, wb, is given by the classical Kirsch solution
2.4. Fracture propagation
(Kirsch, 1898).
To model fracture propagation, we use the implicit level
Coupling of the XFEM with the fluid flow solver is done set algorithm (ILSA) developed by Peirce and
similarly as in Gordeliy and Peirce (2013). Detournay (2008). This algorithm makes use of the
fracture tip asymptotics: for each time step, it finds the
fracture extension consistent with the crack width
2.3. Fluid flow model asymptote. We apply this approach as in the coupled
The flow of the fracturing fluid is governed by the XFEM model in Gordeliy and Peirce (2013) and Abbas
Reynolds lubrication theory. Accordingly, the fluid flux et al. (2014). We use the viscous tip asymptote for the
q is given by Poiseuilles law: examples in this paper.

w3 p f For a plane strain (KGD) hydraulic fracture, the


q (2) viscosity-dominated propagation regime is relevant
s
when K * 0.5 , where K * is the dimensionless
and the conservation law is given by
toughness defined by
w q
0 (3) K
t s K* (5)
E 3Qo
1/4
where w and p f are, correspondingly, the crack width
and the fluid pressure, and s and t are, correspondingly,
In this case, the fracture tip width is governed by the
the curvilinear coordinate along the crack and the time.
viscous asymptote (Desroches et al., 1994):
The boundary conditions at the wellbore prescribe that
V 2/3
1/3

the sum of the fluid fluxes for all fractures propagating w m s (6)
from the wellbore (i.e., two fractures for a bi-wing E
fracture, one fracture for a single-wing fracture) is equal where s is the distance from the fracture tip, V is the
to the total injection rate Qo , and that the fluid pressure fracture tip velocity, and m 21/335/6 . We use this tip
at the wellbore inlet of each fracture is equal to the asymptote within the ILSA approach to determine the
fracture extension for the examples in this paper. Note
that this asymptote does not involve the fracture
toughness K Ic , i.e., K Ic does not influence fracture discretized into a circumferential finite element mesh
propagation in the viscosity-dominated regime. refined near the wellbore to accurately capture the
fracture reorientation (Figure 3). Infinite elements are
To model fracture reorientation, the direction of fracture distributed along the external boundary, to model an
propagation is determined according to the maximum infinite plane.
tensile stress criterion or the zero shear stress criterion
(Erdogan and Sih, 1963). At each time step, the stress Figure 4 (top) shows a comparison of the obtained bi-
field is computed ahead of the fracture tip, and the wing fracture trajectories (only one of the two symmetric
fracture deflection direction for the next time step is fracture wings is shown), and Figure 4 (bottom) shows
determined according to the zero shear stress. the corresponding wellbore pressure evolution with time.
The results from the two models are in good agreement
for the fracture paths and the wellbore pressure at long
3. NUMERICAL EXAMPLES time ( t 0.025 s ). The values of the wellbore pressure
3.1. Validation from the XFEM at early time ( t 0.025 s ) are
In this section, we compare the results of the coupled significantly above the DDM result of Zhang et al.
XFEM model described above to the results obtained by (2011), this shift in pressure may be due to a difference
Zhang et al. (2011). Zhang et al. (2011) modeled the in the modeling of fracture initiation at the wellbore in
near-wellbore reorientation of hydraulic fractures by the two models and is still under investigation.
using a coupled algorithm based on the DDM. To model
fracture extension and deviation, they used the
maximum tensile stress criterion (Erdogan and Sih,
1963) based on the computed mode I and II stress
intensity factors. The fracture front was updated
according to the propagation condition K I K Ic .
However, the parameters of the simulations that we refer
to in this paper all correspond to the viscosity-dominated
regime of propagation ( K * 0.5 ).

First, we consider propagation of a symmetric bi-wing


fracture initiated at either 0 or 30 with respect
to the direction of the maximum principal compressive
stress xx (Fig. 7 in Zhang et al., 2011). The values of Fig. 2. Trajectory of a bi-wing fracture (blue, 30 )
within the computational XFEM domain between the wellbore
the principal confining stresses are xx 80 MPa and
wall at R 0.1 m and the external boundary at
yy 50 MPa , and the properties of the rock are Rmax 0.3 m .
Youngs modulus E 65 GPa , Poissons ratio
0.25 , and the fracture toughness
K Ic 1.35 MPa m . The viscosity of the injected fluid
and the injection rate are, correspondingly,
0.01 Pa s and Qo 0.0004 m2 /s . According to
Eq. (5), the dimensionless toughness in this example is
K * 0.38 . The wellbore radius is R 0.1 m and the
length of the initial crack (or formation defect at the
wellbore wall) is a = 0.006 m. The boundary condition
along the wellbore wall is the wellbore fluid pressure,
corresponding to an openhole completion.

To model the rock formation around the wellbore, we


use an annular computational domain between the
wellbore wall and an external circular boundary of Fig. 3. Detailed view of the finite element mesh and the
radius Rmax , see Figure 2. For the present test case, trajectory of a fracture initiated at the wellbore wall for
Rmax 0.3 m . The domain within these boundaries is 30 .
For a cased pressurized wellbore, the stress field wb,
in the rock is given by the analytical solution of Li
(1991). This solution is obtained under the assumption
that cement behind the casing has a similar elastic
behavior to that of the rock. The borehole radius, R , is
taken to be equal to the outer radius of the casing. The
stress field wb, depends on the wellbore pressure, the
far-field stresses, the elastic moduli of casing (Youngs
modulus Ec , Poissons ratio c ) and rock ( E , ), and
inner and outer radii of the casing (Li, 1991).
We consider a test case of a single fracture propagating
from a defect inclined to the x-axis at 80 . The
problem parameters are set to:
xx 58.5 MPa , yy 30 MPa ,
E 40 GPa , 0.25 , K Ic 1 MPa m ,
0.01 Pa s , Qo 0.05 m2 /s .
The corresponding value of the dimensionless toughness
is found from Eq. (5) as K * 0.12 . The wellbore
radius is R 0.1 m , and the length of the initial crack in
the rock is a = 0.02 m. This test case is similar to the one
studied for an openhole wellbore by Zhang et al. (2011,
their Figure 6). As shown by Zhang et al. (2011), the
fracture reorientation trajectory is governed in this
problem by the following dimensionless parameter:

Fig. 4. Top: trajectory of a bi-wing fracture for 30 .


xx yy R
(8)
Bottom: wellbore pressure evolution for 0 and 30 . 'Q E '
o
3 1/4

3.2. Effect of casing In the present example the value of ( 0.345 ) is


Now we apply the coupled XFEM model to study the the same as considered by Zhang et al. (2011, in their
effect of the wellbore casing. Similarly as for the open Figure 6a with 0.01 Pa s and Qin 0.001 m2 /s ).
hole completion, we decompose the total stress field in However, we consider sufficiently high confining
the rock surrounding a pressurized wellbore into a stresses so that the fluid lag can be neglected. We
superposition of (i) the stress wb, due to the verified that the XFEM results for the fracture trajectory
pressurized wellbore and the in-situ far-field stresses for the open hole in this test are in good agreement with
when no crack is present, and (ii) the stress due to the the results from Zhang et al., 2011 (see Fig. 5 below).
crack in the rock formation with a traction-free borehole
and zero far-field stress. The traction boundary condition We use the XFEM model to compare fracture
on the crack is updated accordingly to the superposition. propagation for the cases of an open hole (OH) and a
In particular, the net pressure acting on the crack faces is cased hole (CH). For CH, we consider the steel casing
given by with the Youngs modulus Ec 200 GPa and
p p f nwb, (7) Poissons ratio c 0.3 ; the inner radius of the casing
Above, tensile stress is positive; p f is the fluid is Rc 0.09 m , and the outer radius of the casing is
equal to the wellbore radius R 0.1 m .
pressure, and nwb, is the normal traction component of
the stress field wb, on the crack surface. For an Figure 5 shows a comparison of the obtained fracture
openhole wellbore, wb,
is given by the Kirsch trajectories for OH and CH using the XFEM. Figure 6
(top, middle) shows the evolution of the corresponding
solution (Kirsch, 1898).
crack width at the wellbore inlet and the wellbore
pressure. Figure 6 (bottom) shows the relative difference
between the wellbore pressure for CH ( pwCH ) and OH
( pwOH ), computed as
pwCH pwOH
100% (9)
pwOH
It is seen that the fracture trajectory reorients to the
direction of the maximum principal compressive stress
(x-axis) within a shorter distance for CH. The near-
wellbore crack width is restricted for both OH and CH
within the simulation time.

Note that the wellbore pressure required to propagate the


fracture and keep it open is smaller for CH. This effect
corresponds to the difference in the stress field wb, for
OH and CH. Indeed, Fig. 7 shows the hoop stress
wb,

computed along the direction of the initial crack 80


for the conditions of the present test and a wellbore
pressure of 110 MPa. Along the initial crack, this stress
represents the normal stress nwb, in Eq. (7). This
negative (compressive) stress needs to be overcome by
the fluid pressure in order to open the crack. It is seen
that for the present test case, the compressive stress
acting on the initial crack is larger for OH than for CH.
Therefore the fluid pressure required to open and
propagate the fracture is larger for OH. This effect can
be reverse for different conditions, such as in the next
example.

Fig. 6. Top: evolution of crack width at the wellbore inlet.


Middle: evolution of wellbore pressure. Bottom: relative
difference in wellbore pressure for OH and CH completions.
The results correspond to an initial crack inclined at 80 .

Fig. 5. Computational annular domain and fracture trajectories


for openhole (blue) and cased hole (red) completions, for an
initial crack inclined at 80 . Fracture trajectory for open
hole from Zhang et al. (2011) is shown by black dots.

Fig. 7. Hoop stress wb, in the rock, for 80 and


wellbore pressure 110 MPa.
3.3. Effect of viscosity and injection rate, and
reverse effect of casing
In the third example, we consider propagation of a bi-
wing fracture initiated from the symmetric defects
inclined to the x-axis, the direction of the maximum
principal compressive stress, at 30 . We keep the
following parameters same as in the example in Section
3.1:

xx 80 MPa , yy 50 MPa ,
E 65 GPa , 0.25 , K Ic 1.35 MPa m ,
Fig. 8. Computational annular domain and bi-wing fracture
R 0.1 m . trajectories for openhole (blue) and cased hole (red)
However, now we increase the viscosity of the injected completions, for initial symmetric defects inclined at 30 .
fluid and the injection rate to 0.1 Pa s and
Qo 0.05 m2 /s . (The corresponding value of the
dimensionless toughness is K * 0.06 .) We look at the
results for OH and CH completions for the length of the
initial defects set to a = 0.02 m. For CH, we consider the
steel casing with the same elastic moduli and dimensions
as in the example in Section 3.2.

Figures 8 and 9 (top, middle) show a comparison of the


obtained bi-wing fracture trajectories and the
corresponding inlet crack width and the wellbore
pressure evolution with time. Figure 9 (bottom) shows
the relative difference between the wellbore pressure for
CH and OH, computed according to Eq. (9).
For the case of OH, a comparison of the results with
those in Section 3.1 shows that the increased viscosity
and injection rate result in a significantly larger pressure
required to propagate the fracture. A comparison of the
results for OH and CH completions shows that the
fracture trajectory is not significantly different for the
two completion types. Unlike the example in Section
3.2, in this example a larger fluid pressure is required to
propagate the fracture for CH completion. Indeed, Fig.
10 shows the hoop stress wb,
along the direction of
the initial crack 30 , for the conditions of the
present test and a wellbore pressure of 140 MPa. This
stress represents the tensile stress nwb, in Eq. (7),
which acts in the same direction as the opening fluid Fig. 9. Top: evolution of crack width at the wellbore inlet.
pressure. It is seen that for this case, the near-wellbore Middle: evolution of wellbore pressure. Bottom: relative
tensile stress along the direction of the initial crack is difference in wellbore pressure for OH and CH completions.
larger for OH than for CH. Correspondingly, the fluid The results correspond to initial symmetric defects inclined at
pressure required to open and propagate the fracture is 30 .
smaller for OH.
fracture offsets: An XFEM Application. In Proceedings
of the SPE Hydraulic Fracturing Technology
Conference, The Woodlands, paper SPE 168622.
3. Aud, W.W., T.B. Wright, C.L. Cipolla, J.D. Harkrider,
and J.T. Hansen. 1994. The effect of viscosity on near-
wellbore tortuosity and premature screenouts. In
Proceedings of the SPE Annual Technical Conference
and Exhibition, New Orleans, paper SPE 28492.
4. Cherny, S., D. Chirkov, V. Lapin, A. Muranov, D.
Bannikov, M. Miller, D. Willberg, O. Medvedev, and
O. Alekseenko. 2009. Two-dimensional modelling of
the near-wellbore fracture tortuosity effect. Int J Rock
Fig. 10. Hoop stress wb, in the rock, for 30 and Mech Min Sci., 46, 992-1000.
wellbore pressure 140 MPa. 5. Cherny, S., V. Lapin, D. Chirkov, O. Alekseenko, and
O. Medvedev. 2009. 2D modeling of hydraulic fracture
initiating at a wellbore with or without microannulus.
In Proceedings of the SPE Hydraulic Fracturing
4. CONCLUSIONS Technology Conference, The Woodlands, paper SPE
119352.
We have developed a fluid-coupled algorithm for
modeling reorientation of longitudinal fractures 6. Desroches, J., B. Lenoach, P. Papanastasiou, J.R.A.
propagating from a wellbore aligned with the Pearson, M. Thiercelin, and A. Cheng. 1994. The crack
intermediate or maximum far-field stress direction. A tip region in hydraulic fracturing. Proc. R. Soc. London,
Ser. A, 447: 3948.
two-dimensional plane-strain model is adopted. The rock
deformation solver is implemented using the extended 7. Erdogan, F. and G. Sih. 1963. On the crack extension in
finite element method (XFEM). The model for the fluid plates under plate loading and transverse shear. J. Basic
flow in hydraulic fractures is based on the Reynolds Eng., 85: 519527.
lubrication equation. The fluid pressure and the fracture 8. Fries, T.-P. and T. Belytschko. 2010. The
opening are obtained from the solution of a system of extended/generalized finite element method: An
nonlinear equations resulting from the coupling of the overview of the method and its applications. Internat. J.
rock deformation solver with the fluid flow model. To Numer. Methods Engrg., 85: 253304.
model fracture reorientation, the direction of fracture 9. Gordeliy, E. and A. Peirce. 2013. Implicit level set
propagation is determined based on the maximum tensile schemes for modeling hydraulic fractures using the
stress criterion. In this novel approach, the XFEM has XFEM. Comput. Methods Appl. Mech. Engrg., 266:
been used in a fully coupled manner for the first time in 125143.
viscosity-dominated regime to model the effect of near- 10. Kirsch, E.G. 1898. Die theorie der elastizitt und die
wellbore fracture reorientation on the wellbore pressure. bedrfnisse der festigkeitslehre. Zeitschrift des Vereines
We study the effect of the wellbore casing on the deutscher Ingenieure, Vol. 42, pp. 797-807.
fracture opening and the fluid pressure required to
initiate and propagate the fracture. 11. Li, Y. 1991. On initiation and propagation of fractures
from deviated wellbores. The University of Texas at
Austin. PhD thesis.
12. Mack, M.G., and N.R. Warpinski. 2000. Mechanics of
ACKNOWLEDGMENT hydraulic fracturing. In: Economides, M.J., and K.G.
The authors would like to thank Schlumberger for Nolte (eds), Reservoir stimulation, 3rd ed. John Willey
& Sons; 2000.
permission to publish this work.
13. Mos, N., J. Dolbow, and T. Belytschko. 1999. A finite
element method for crack growth without remeshing.
REFERENCES Int. J. Numer. Methods Eng., 46, 131150.

1. Abass, H.H., D.L. Meadows, J.L. Brumley, S. 14. Peirce, A. and E. Detournay. 2008. An implicit level set
Hedayati, and J.J. Venditto. 1994. Oriented perforations method for modeling hydraulically driven fractures.
- A rock mechanics view. In Proceedings of the SPE Comput. Methods Appl. Mech. Engrg., 197: 2858-2885.
Annual Technical Conference and Exhibition, New 15. Romero, J., M.G. Mack, and J.L. Elbel. 1995.
Orleans, paper SPE 28555. Theoretical model and numerical investigation of near-
2. Abbas, S., E. Gordeliy, A. Peirce, B. Lecampion, D. wellbore effects in hydraulic fracturing. In Proceedings
Chuprakov, and R. Prioul. 2014. Limited height growth of the SPE Annual Technical Conference and
and reduced opening of hydraulic fractures due to Exhibition, Dallas, paper SPE 30506.
16. Sepehri, J., M.Y. Soliman, and S.M. Morse. 2015.
Application of Extended Finite Element Method
(XFEM) to simulate hydraulic fracture propagation
from oriented perforations. In Proceedings of the SPE
Hydraulic Fracturing Technology Conference, The
Woodlands, paper SPE-173342.
17. Stolarska, M., D. Chopp, N. Mos, and T. Belytschko.
2001. Modeling crack growth by level sets and the
extended finite element method, Int. J. Numer. Methods
Eng. 51, 943960.
18. Van de Ketterij, R. G., and C. J. de Pater. 1999. Impact
of perforations on hydraulic fracture tortuosity. SPE
Production & Facilities, 14, 131-138. Paper SPE-
56193-PA.
19. Wang, H. 2015. Numerical modeling of non-planar
hydraulic fracture propagation in brittle and ductile
rocks using XFEM with cohesive zone method. Journal
of Petroleum Science and Engineering, 135, 127140.
20. Zhang, Z., R.G. Jeffrey, A.P. Bunger, and M.
Thiercelin. 2011. Initiation and growth of a hydraulic
fracture from a circular wellbore. Int. J. Rock Mech.
Min. Sci., 48, 984-995.

Вам также может понравиться