Anda di halaman 1dari 11

REE166332 DOI: 10.

2118/166332-PA Date: 30-April-14

Stage:

Page: 233

Total Pages: 11

Characterizing Hydraulic Fracturing With a


Tendency-for-Shear-Stimulation Test
Mark McClure, SPE, University of Texas at Austin, and Roland Horne, SPE, Stanford University

Summary
The classical concept of hydraulic fracturing is that a single, planar, opening mode fracture propagates through the formation. In
recent years, there has been a growing consensus that natural fractures play an important role during stimulation in many settings.
There is not universal agreement on the mechanisms by which
natural fractures affect stimulation, and these mechanisms may
vary depending on formation properties. One potentially important mechanism is shear stimulation, in which increased fluid
pressure induces slip and permeability enhancement on pre-existing fractures. We propose a tendency-for-shear-stimulation (TSS)
test as a direct, relatively unambiguous method for determining
the degree to which shear stimulation contributes to stimulation in
a formation. In a TSS test, fluid injection is performed while
maintaining the bottomhole fluid pressure slightly less than the
minimum principal stress. Under these conditions, shear stimulation is the only possible mechanism for permeability enhancement
(except, perhaps, thermally induced tensile fracturing). A TSS test
is different from a conventional procedure because injection is
performed at a specified pressure (rather than a specified rate).
With injection at a specified rate, fluid pressure may exceed the
minimum principal stress, and it may cause tensile fractures to
propagate through the formation. If this occurs, it will be ambiguous whether stimulation was because of shear stimulation or tensile fracturing. Maintaining pressure less than the minimum
principal stress ensures that the effect of shear stimulation can be
isolated. Low-rate injectivity tests could be performed before and
after the TSS test to estimate formation permeability. An increase
in formation permeability would indicate that shear stimulation
has occurred. The flow-rate transient during injection may also be
interpreted to identify shear stimulation. Numerical simulations of
shear stimulation were performed with a discrete-fracture-network
(DFN) simulator that couples fluid flow with the stresses induced
by fracture deformation. These simulations were used to qualitatively investigate how shear stimulation and fracture connectivity
affect the results of a TSS test. Two specific field projects are discussed as examples of a TSS test, the Enhanced Geothermal Systems (EGS) projects at Desert Peak, Nevada, and Soultz-sousForets, France.
Introduction
Classically, hydraulic fracturing has been conceptualized as creating a single, planar, opening mode tensile fracture. But in low-matrix-permeability applications, such as hydrocarbon production
from shale or geothermal production from crystalline rock, the
process of hydraulic stimulation has been conceptualized as creating a complex network of newly forming fractures and/or natural
fractures that slip and open in response to injection (Pine and
Batchelor 1984; Murphy and Fehler 1986; Brown 1989; Ito 2003;
Ito and Hayashi 2003; Fisher et al. 2004; Evans et al. 2005;
Bowker 2007; Gale et al. 2007; Cipolla et al. 2008; King 2010;
Ledesert et al. 2010). The precise geometry of these networks is a
major uncertainty. The networks cannot easily be observed
directly in the subsurface; it is difficult to know how laboratory
C 2014 Society of Petroleum Engineers
Copyright V

This paper (SPE 166332) was accepted for presentation at the SPE Annual Technical
Conference and Exhibition, New Orleans, 30 September2 October 2013, and revised for
publication Original manuscript received for review 23 June 2013. Revised manuscript
received for review 25 November 2013. Paper peer approved 16 December 2013.

experiments relate to the reservoir scale; and microseismic interpretations with respect to network geometry are nonunique.
One potentially important process is the termination of propagating natural fractures against pre-existing planes of weakness.
This has been described in laboratory experiments (Blanton 1982;
Renshaw and Pollard 1995; Zhou et al. 2008; Gu et al. 2011;
Suarez-Rivera et al. 2013), mine-back experiments (Warpinski
and Teufel 1987; Warpinski et al. 1993; Mahrer 1999; Jeffrey
et al. 2009; Chuprakov et al. 2013), and computational investigations (Dahi-Taleghani and Olson 2009; Gu and Weng 2010; Fu
et al. 2012). If termination occurs, then it may be difficult for a
single, continuous fracture to propagate across the formation, and
pathways for flow through the reservoir may occur through both
new and pre-existing fractures [a process we refer to as mixedmechanism stimulation (MMS)]. This process could play an important role in generating stimulated-fracture surface area and,
therefore, increasing recovery. The MMS concept has been
applied toward the field-scale modeling of hydraulic stimulation
in shale by several authors (Damjanac et al. 2010; Weng et al.
2011; Wu et al. 2012; Section 3.3.2 of McClure and Horne
2013a).
In contrast to the MMS concept, some authors have modeled
stimulation in shale with a single, large, continuous tensile fracture per stage (Warpinski et al. 2001; Palmer et al. 2007; Rogers
et al. 2010; Nagel et al. 2011; Roussel and Sharma 2011). With
this approach, a problem arises because high production rates
from low-permeability shale formations apparently require more
stimulated-fracture surface area than can be explained by a single,
linear fracture per stage (Mayerhofer et al. 2010; Fan et al. 2010;
Cipolla et al. 2010). This problem can be resolved by postulating
that there are secondary fractures surrounding the primary fracture
that contribute significantly to production. Some modelers are
agnostic about the nature of the secondary fractures (Warpinski
et al. 2001), and others suggest that the secondary fractures may
be newly forming tensile fractures (Roussel and Sharma 2011).
Low-permeability formations may contain some fractures that are
naturally conductive (Laubach et al. 2004; Olson et al. 2009),
and, thus, hydraulic fracturing may also improve recovery by connecting these conductive fractures to the wellbore. Other authors
have modeled secondary fractures as shear stimulating (or possibly opening) of pre-existing fractures (Palmer et al. 2007; Rogers
et al. 2010; Nagel et al. 2011). We refer to this latter mechanism
as primary fracturing with shear-stimulation leakoff.
Shear stimulation occurs when slip along a fracture causes a
mismatch of asperities. This mismatch creates additional void
space between the fracture walls, increasing the fracture transmissivity, a mechanism sometimes called self-propping (Barton
et al. 1985; Esaki et al. 1999). Slip can be induced by an increase
in fluid pressure, which weakens the frictional resistance to slip.
Fractures in more-brittle formations are more likely to be selfpropping (Lutz et al. 2010), and the creep of fracture asperities
could cause time-dependent loss of transmissivity after slip.
Because of the high compressive strength of intact rock, fluid injection should not typically be expected to create the formation of
new shear fractures from intact rock (Olson et al. 2009). Instead,
the term shear stimulation refers to induced slip on pre-existing
fractures.
Projects in which hydraulic stimulation is performed for geothermal-energy production are often referred to as EGS. EGS
projects are pursued in a variety of settings, but it is most common
to perform hydraulic fracturing in deep wells drilled into the

May 2014 SPE Reservoir Evaluation & Engineering


ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

233

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

crystalline basement (3- to 5-km depth). The potential size of the


geothermal resource is very large, because temperatures at this
depth are high enough for geothermal-energy production across a
significant percentage of the Earths surface. However, the natural
permeability in the crystalline basement is typically very low, and
thus, hydraulic stimulation is needed to improve economic performance (Tester 2006). EGS is most often performed by injecting water into a long, openhole section of a vertical wellbore
without the use of proppant.
Most authors believe that, during EGS stimulation in crystalline formations, the dominant mechanism of stimulation is induced slip on pre-existing fractures, which we refer to as pure
shear stimulation (PSS). Early EGS experience indicated that flow
from the wellbores was localized at pre-existing fractures and that
microseismic hypocenter locations did not form into narrow, planar features perpendicular to the minimum principal stress (Pine
and Batchelor 1984; Murphy and Fehler 1986). Some early
authors postulated a dendritic network model of stimulation
(Murphy and Fehler 1986), and authors eventually converged on
the shear-stimulation idea, to the exclusion of new propagating
fractures (Pine and Batchelor 1984; Lanyon et al. 1993; WillisRichards et al. 1996; Ito and Hayashi 2003; Tester 2006; Cladouhos et al. 2009). The modeling workflow is to stochastically generate a realization of the pre-existing fracture network and then
simulate flow and shear stimulation in the network (Kohl and
Hopkirk 1995; Bruel 1995; Jing et al. 2000; Rahman et al. 2002;
Bruel 2007; Kohl and Megel 2007; Rachez and Gentier 2010;
Cladouhos et al. 2011; Zhou and Ghassemi 2011; Riahi and Damjanac 2013).
McClure and Horne (2013b) argued that the PSS may not be
as universal in crystalline EGS as the conventional wisdom suggests, and that, in many or most projects, MMS may occur, with
both new and pre-existing fractures playing an important role.
McClure and Horne (2013b) pointed out that many geological
conditions must be present for PSS to be possible, and these conditions cannot always be expected to be satisfied. These conditions are: (1) abundant natural fractures oriented in the local stress
state so that pressure increase causes them to slip, (2) natural fractures that are self-propping such that they become conductive after they slip, (3) a percolating fracture network (that there are
continuous pathways for flow through the network), (4) adequate
fracture-network storativity to contain most of the injected fluid
(if the formation has low matrix permeability), (5) sufficiently
high prestimulation fracture transmissivity so that fluid pressure
can enter the fractures and start the shear-stimulation process, and
(6) stimulated-fracture transmissivity that is sufficiently high so
that the fluid pressure does not build up to the minimum principal
stress and cause the propagation of new tensile fractures (even as
shear stimulation is taking place).
Because of the conditions that must be met for shear stimulation to occur, the TSS to occur in response to fluid injection could
be considered a formation property. TSS could be estimated
through indirect analyses such as Coulomb-stress (CS) analyses
of the local fracture network (Ito and Hayashi 2003; Evans 2005),
rock mechanics testing of core to evaluate the tendency for selfpropping (Lutz et al. 2010), or correlations from measurements
with well-logging tools. But fully integrated methodologies to
predict TSS do not yet exist, and efforts to develop these methodologies will benefit from having a technique to verify their
predictions.
There is a need for techniques to directly, unambiguously
diagnose shear stimulation in the field. Unfortunately, conventional hydraulic-stimulation practice does not provide unambiguous evidence of shear stimulation. When injection is performed at
a high rate into a low-permeability formation, there are two possible outcomes: (a) fluid pressure reaches the minimum principal
stress, potentially causing new tensile fractures to form and propagate or (b) shear stimulation occurs, preventing the bottomhole
pressure from reaching the minimum principal stress. In Case (a),
both shear stimulation and the propagation of new tensile fractures may occur, and it will be ambiguous regarding which had a
greater role.
234

Stage:

Page: 234

Total Pages: 11

If shear stimulation is frequently effective, we might expect


that Case b would frequently occur, that injection pressure is prevented from reaching the minimum principal stress. Our review of
the EGS literature found only a single example in which stimulation at a specified rate did not cause the bottomhole fluid pressure
to reach the minimum principal stressthe June 2000 stimulation
of the Well GPK2 at Soultz, as discussed later in the text (Weidler
et al. 2002; Valley and Evans 2007). According to published values, the bottomhole pressure exceeded the minimum principal
stress at the projects at Hijiori, Fjallbacka, Le Mayet, Rosemanowes, Fenton Hill (Willis-Richards et al. 1995), Ogachi (assuming a vertical stress gradient of 25 MPa/km) (Kaieda et al. 2010),
Desert Peak (Chabora et al. 2012), Soultz (except GPK2) (Valley
and Evans 2007), Cooper Basin (Baisch et al. 2006), and Grob
Schonebeck (Zimmermann et al. 2008). The assessments of the
minimum principal stress for the Basel and Newberry projects are
not available. Well injectivity increased in response to injection in
all cases listed, but because fluid pressure exceeded the minimum
principal stress in every case, it is ambiguous whether the stimulation was caused by shear stimulation or the propagation of new
tensile fractures.
We propose that a TSS test could be used in the field to evaluate the importance of shear stimulation practically and relatively
unambiguously. In a TSS test, fluid is injected into an openhole
section of the wellbore such that the bottomhole fluid pressure is
maintained slightly less than the minimum principal stress. Under
these conditions, shear stimulation is the only possible mechanism
of stimulation (except possible thermal fracturing if injection is
performed for a long duration of time, but this would be confined
relatively close to the wellbore). The formation permeability
would be measured with conventional injection/falloff tests before
and after the stimulation treatment; the change in permeability
could be used to assess TSS. In an injection/falloff test, injection
is performed at a constant rate, and then the well is shut in. The
resulting pressure transient is analyzed to infer reservoir permeability (Horne 1995; Kamal 2009). The injectivity tests would
need to be performed at a rate low enough that the tests cause neither shear stimulation nor tensile fracturing. If the permeability is
determined to be higher after the TSS test, this would indicate
that shear stimulation has occurred. The behavior of the injection
rate over time during the TSS test could also be used to diagnose
TSS and possibly gather some additional information about the
properties of the formation. The spreading of microseismicity a
significant distance from the wellbore during the TSS test may
also indicate that shear stimulation is occurring.
Our review of the EGS literature suggests that the bottomhole
fluid pressure has roughly equaled or exceeded the minimum principal stress during stimulation in almost every case. There is one
case in the literature in which bottomhole pressure was intentionally maintained at less than the minimum principal stressthe
Desert Peak EGS project (discussed later in the text), and the
results of this test suggest that shear stimulation was only modestly effective.
To investigate the properties of TSS tests, three simulations
were performed with CFRAC, a 2D discrete fracture-network
simulator that fully couples fluid flow and the stresses induced by
fracture deformation (McClure and Horne 2013a). The simulations were configured to loosely resemble the reservoir parameters
at the Desert Peak EGS project, in which fluid injection was intentionally performed with a bottomhole pressure less than the minimum principal stress (Chabora et al. 2012; Benato et al. 2013;
Dempsey et al. 2013), which makes the project a prototype example of a TSS test. During injection into Well GPK2 at the EGS
project at Soultz-sous-Forets, shear stimulation occurred during
injection at fluid pressure less than the minimum principal stress,
though this was not intentional. The results of both projects are
discussed in this paper.
Several simplifications are made by CFRAC (most importantly, it neglects matrix flow and is 2D), which limit its ability to
be used for forward modeling of realistic reservoir transients
(because these transients are affected by leakoff and the dimensionality of the problem). These simplifications do not affect the
May 2014 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

results qualitatively, but they do affect the specific shapes of the


plots of injection rate vs. time. In future work, we plan to repeat
these simulations with a 3D model that includes matrix flow to
perform forward modeling of the pressure and rate transients
before, during, and after the TSS test. CFRAC was used rather
than a more conventional code because the character of the overall shear-stimulation process is significantly affected by both the
idiosyncrasies of flow in the fracture networks and the stresses
induced by deformation.
If shear stimulation contributes significantly to production in
shale, this would have important implications for stimulation design and formation assessment. Geologists need to know whether
to incorporate the variables affecting TSS into their resource assessments and in their search for sweet spots for production. Modelers need to know whether to incorporate shear stimulation into their
simulations. Concepts about shear stimulation shape the physical
intuition of engineers who make design decisions.
In geothermal engineering, most EGS modeling neglects the
propagation of new fractures, and projects are designed to optimize the process of shear stimulation. If tensile fracturing (or
MMS) was reconsidered as a viable mechanism, it would open up
new possibilities for overall system design (such as the use of
proppant), would change thinking about optimal geological conditions, and would lead to different modeling approaches.
Chipperfield et al. (2007) suggested that a step-rate test could
be used to diagnose shear stimulation. In a step-rate test, injection
is performed at a series of increasing rates. At the beginning of
the step-rate test, the injection pressure should be roughly proportional to the injection rate. But if shear stimulation occurs, the
injection pressure should increase by less with each progressive
increase in rate. This approach is conceptually similar to the TSS
test that we propose in this paper, and, in practice, these approaches may turn out to be quite similar. The difference is that
we propose to inject at a bottomhole pressure slightly less than
the minimum principal stress, which would not necessarily occur
during a step-rate test. With a step-rate test, the optimal injection
rates may not be known in advance. If injection-rate steps are too
large, the pressure will build up higher than the minimum principal stress, and it will be ambiguous whether stimulation has
occurred as a result of newly forming fractures or shear stimulation. If injection rates are too low, pressure will not build up
enough to cause slip on the pre-existing fractures, and it will not
be possible to identify shear stimulation. Therefore, the key
requirement for the successful diagnosis of shear stimulation is
injection at bottomhole pressure slightly less than the minimum
principal stress (which we define as a TSS test). A step-rate test
can be used to identify shear stimulation, but only if injection-rate
steps are chosen appropriately. Another issue is that conventional
step-rate tests are performed for brief durations with small fluid
volumes. But, as demonstrated by Simulation B in this paper
(shown later in this text), nonpercolating natural-fracture networks could experience shear stimulation initially, but this effect
could be confined close to the wellbore. It would be ideal to perform hours or several days of pumping during a TSS test to appropriately sample the behavior of the reservoir.
Chipperfield et al. (2007) used a continuum-based model to
describe shear stimulation. In this paper, we demonstrate how
continuum-based models of shear stimulation can lead to different
results than DFN models. In the Discussion section later in this
paper, we discuss the differences between the shear-stimulation
behavior observed in the models in this study and the behavior
described by Chipperfield et al. (2007).
Methodology
Computational Model. The full details of CFRAC were summarized by McClure and Horne (2013a), and only a brief overview is given in this section. The model assumes single-phase
liquid water (no proppant), isothermal conditions, Darcy flow in
the fractures, and zero permeability in the matrix around the
fractures, and gravity is neglected. Stresses induced by fracture
deformation are calculated with the Shou and Crouch (1995) dis-

Stage:

Page: 235

Total Pages: 11

placement-discontinuity (boundary element) method that uses


quadratic-basis functions and assumes homogeneous, isotropic,
and linear elastic deformation. A code called Hmmvp (Bradley
2011) is used to very accurately and efficiently approximate the
matrices of interaction coefficients that arise from the displacement-discontinuity method, which increases the efficiency of the
code significantly. When creating the matrix approximations for
this study, a relative error etol equal to 106 was used (defined in
Section 3.5 of McClure and Horne 2013a).
Fractures may be open, walls out of contact, or closed,
walls in contact. Stresses induced by normal displacement discontinuities of closed fractures (described by Eq. 3) are neglected.
For a crack, these displacements are caused by fracture stiffness
and are quite small. For a thicker, fault-zone-like feature, displacements may be bigger, but they are spread along a larger volume, reducing stress and strain (see Section 2.3.3 in McClure and
Horne 2013a for more discussion).
The simulations are 2D and should be interpreted as showing
normal faults from the side, viewed in the direction of the maximum horizontal stress. The out-of-plane thickness of the formation, h, was assumed to be 100 m. In this paper, the vertical
direction is referred to as the y-axis direction, and the horizontal
direction is referred to as the x-axis direction. The Shou and
Crouch (1995) method assumes plane-strain deformation, implying infinite fracture size in the out-of-plane dimension, but, in this
study, the Olson (2004) adjustment was used to approximate the
effect of a finite-sized out-of-plane dimension.
Flow is not upscaled to an effective continuum model.
Because flow in the matrix is neglected and the boundary element
method is used to calculate stresses induced by deformation, it is
only necessary to discretize the fractures, not the matrix. Implicit
timestepping is used. During every timestep, the fluid pressure
and (if the element is opening and/or sliding) opening and sliding
displacement discontinuities are calculated to satisfy simultaneously the unsteady-state mass-balance equation and appropriate
stress conditions at each element. Elements may be closed (walls
in contact) or open (walls out of contact), depending on whether
the fluid pressure has reached the normal stress. If the walls are in
contact, Coulombs law with a constant coefficient of friction is
used to determine whether the fracture should slide, and if so, displacements are calculated so that Coulombs law is satisfied. If
walls are out of contact, displacements are calculated so that the
walls bear zero shear stress. A radiation-damping term (Rice
1993; Segall 2010) is included for shear stress to approximate the
effect of inertia at a high slipping velocity (though a high slipping
velocity is uncommon if a constant coefficient of friction is used)
and to prevent sliding from happening instantaneously.
The Coulomb-failure criterion with a radiation-damping term
is (Segall 2010)
js  gvj lf r0n S0 ; . . . . . . . . . . . . . . . . . . . . . . . 1
where s is the shear stress, g is the radiation-damping coefficient,
v is the sliding velocity of the fracture, lf is the coefficient of friction (assumed constant in this study), S0 is fracture cohesion, and
r0n is the effective normal stress, defined as (Segall 2010)
r0n rn  P; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
where P is fluid pressure and compressive stresses are taken to be
positive. For fractures with shear stress less than the frictional resistance to slip, shear deformation is assumed to be negligible.
Nonlinear relationships are used between fracture stress, fluid
pressure, opening displacement discontinuity, sliding displacement discontinuity, hydraulic aperture, and void aperture (modified from Willis-Richards et al. 1996):
E

E0
/Edil
DE;eff tan
; . . . . . . 3
1 9r0n =rn;Eref
1 9r0n =rn;Eref

where E0, rn,Eref, and /Edil are specified constants. DE,eff is


defined as equal to D (cumulative sliding displacement

May 2014 SPE Reservoir Evaluation & Engineering


ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

235

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

Stage:

Page: 236

Total Pages: 11

TABLE 1SIMULATION SETTINGS USED IN SIMULATIONS A, B, AND C


De,eff,max
E0
G (shear modulus)

0.005 m
.005 m
15 GPa

h
Pinit (initial fluid pressure)
S0
Mechtol [error tolerance used in solving
shear-stress equations, Section 2.3.5 in
McClure and Horne (2013a)]
Itertol [error tolerance used in
iterative coupling scheme, Section 2.3.1
in McClure and Horne (2013a)]
/Edil
tp (Poissons ratio)

3 MPa/(m/s)
0.5 MPa

100 m
8.7 MPa
0.5 MPa
.003 MPa

gtarg [for timestepping, Section 2.3.9 in


McClure and Horne (2013a)]
lf
rn,Eref
rn,eref
rxx (initial x-direction principal stress at y 0)

0.6
20 MPa
20 MPa
14.5 MPa

0.01 MPa

ryy (initial y-direction principal stress at y 0)

23.8 MPa

0

rxx,trend (trend in initial rxx with distance in


y-axis direction)
ryy,trend (trend in initial ryy]- with distance in the
y-axis direction)

5 MPa/km

0.25

discontinuity) if D < DE,eff,max, and equal to DE,eff,max otherwise.


The constants are allowed to be different for hydraulic aperture e
and void aperture E. Nonzero /Edil corresponds to pore-volume
dilation with slip, and nonzero /edil corresponds to transmissivity
enhancement with slip.
Hydraulic aperture is equal to void aperture between two
smooth plates but is lower than void aperture for rough surfaces
such as rock fractures (Liu 2005). A fracture in a DFN model
may represent a crack, but it may also represent a more-complex
feature such as a fault zone (Section 2.3.3 in McClure and Horne
2013a). In the latter case, the void aperture may be much larger
than the hydraulic aperture, which is the reason that the model
allows e and E to be different.
The cubic law for fracture transmissivity (the product of permeability and hydraulic aperture) is (Witherspoon et al. 1980)
T ke

e3
; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
12

where T is transmissivity and k is permeability. In the simulations


in this paper, it was assumed that no new fractures could form.
Even though the injection pressure was less than the minimum
principal stress, it was a simplification to neglect fracture propagation because concentrations of tensile stress could locally develop from sliding fractures, causing tensile fractures. This type
of fracture is sometimes called a splay crack or wing crack (Mutlu
and Pollard 2008).
Details of the Simulations. Three simulations (A, B, and C)
were performed to investigate the behavior of a TSS test under
different conditions. Simulation A was the baseline simulation in
which the conditions needed for shear stimulation were metthe
coupling of slip to transmissivity enhancement, adequate initial
transmissivity, and a percolating fracture network. In Simulation
B, a nonpercolating fracture network was used. Shear stimulation
was possible, but fluid was unable to propagate far from the wellbore because of the lack of continuous pathways for flow. The
total number of fractures intersecting the wellbore was roughly
the same in Simulations A and B, but Simulation B had a lower
average fracture length. Simulation C used the same fracture network as Simulation A, but there was no coupling between slip and
transmissivity (/edil equal to zero), making shear stimulation
impossible.
In Simulations A and B, the initial fracture transmissivity was
very low, causing the fracture transmissivity to increase very substantially during slip, from approximately 1018 m3 to approximately 5  1014 m3. The initial fracture transmissivity was set to
be higher in Simulation C than in Simulations A and B (approximately 1016 m3) because, otherwise, very little flow would have
occurred (because there was no shear stimulation to increase
transmissivity).
236

14 MPa/km

The simulations were designed to loosely resemble the shear


stimulations performed in Well 27-15 at the Desert Peak EGS project. The simulations represent a vertical cross section viewed from
the side looking in the direction of the maximum principal stress.
At Desert Peak, the openhole section was approximately 150 m
long and centered at a depth of 990.5 m. According to the stress
profile of Hickman and Davatzes (2010), at the center of the openhole section, at a depth of 990.5 m, rhmin 14.5 MPa, rv 23.8
MPa, and P 8.7 MPa. The gradients of effective rhmin and rv
with depth (stress minus in-situ fluid pressure) were 5 and 14 MPa/
km, respectively. In the simulations, the initial stress was initialized
with a vertical gradient in rxx and ryy of 5 and 14 MPa/km, respectively, and with rxx and ryy equal to 14.5 and 23.8 MPa at y 0.
During the low-rate stimulations of Well 27-15, injection was
performed at wellhead pressures between 1.7 and 3.8 MPa
(approximately 11.6 to 13.7 MPa at 990.5 m). In the simulations,
injection was performed at a constant pressure of 12.4 MPa.
Because the simulations were 2D and neglected matrix flow, no
attempt was made to match the injection-rate trends from the Desert Peak data.
The fracture networks in the simulations were generated stochastically, with fractures at random and obeying specified orientation statistics. To enable shear stimulation, the networks contained
abundant fractures that were well-oriented to slip. The fractures
were oriented approximately 30 clockwise/counterclockwise
from the y-axis. These fracture orientations were roughly consistent with the fracture orientations from Well 27-15, in which wellbore-imaging logs indicated that the dominant fracture orientation
was dipping approximately 60 and striking roughly parallel to the
maximum principal stress. The average fracture spacing in Well
27-15 was approximately 3 m, suggesting there were approximately fifty fractures intersecting the wellbore along the 150-m
openhole section, but only eight were observed to be flowing
before injection (Benato et al. 2013). Because fracture flow tends
to be dominated by a few, major fracture features, the number of
hydraulically important fractures that need to be included in a
model is likely to be closer to eight than fifty. In Simulations A and
C, the wellbore intersected eleven fractures, and in Simulation B,
the wellbore intersected sixteen fractures.
The simulations were performed for 7 days. Table 1 gives the
settings that were common to all the simulations. Table 2 gives
the settings that were different between the simulations.
Results
Figs. 1 through 3 show the final pressure distribution at the end of
simulations for Simulations A, B, and C. Figs. 4 and 5 show contours of induced change in CS (defined in Eq. 5) and shear stress,
respectively, at the end of Simulation A, for a fracture oriented 30
May 2014 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

Stage:

Page: 237

Total Pages: 11

Seconds Elapsed = 604800, Max: 12.4407, Min: 7.4216

TABLE 2SIMULATION SETTINGS SPECIFIC TO


SIMULATIONS A, B, AND C
B

150

2.5
.00001 m

2.5
.00001 m

0
.00005 m

100

Seconds Elapsed = 604800, Max: 12.4019, Min: 8.173


12.5
100

11.5
11

50
0

10.5

50

10

100

9.5

12

80
60

11

20
0

10.5

20

10

40

Fluid Pressure (MPa)

11.5

40

Distance (m)

12

Fluid Pressure (MPa)

Distance (m)

/edil
e0

12.5
200

9.5

60
80

150

200
200

100

100

200

8.5

Distance (m)
Fig. 1The final pressure distribution for Simulation A. The
black line represents the wellbore. The blue/red lines are natural fractures, with color proportional to fluid pressure. The figure can be viewed as showing a vertical cross section of the
formation.

100
100

50

50

100

8.5

Distance (m)
Fig. 2The final pressure distribution for Simulation B. The
black line represents the wellbore. The blue/red lines are natural fractures, with color proportional to fluid pressure. The figure can be viewed as showing a vertical cross section of the
formation.

counterclockwise from the y-axis direction. Fig. 6 shows the normalized flow rate vs. time for Simulations A, B, and C.
Discussion
Overall Behavior. In Simulation A, the final fracture network
was more spatially dispersed and sparsely connected than in Simulation C (Figs. 1 and 3), even though the simulations used the

same fracture network. This is expected because simulations of


shear stimulation that couple flow, stresses induced by shear, and
transmissivity coupling to shear typically predict spatially dispersed and sparsely connected networks [McClure and Horne
(2010a); Section 3.4.2.2 of McClure (2012)]. In continuum simulations of EGS stimulation, fluid pressure and stimulation tend to
spread uniformly in a way that is consistent with a diffusive
spreading of fluid pressure (e.g., Dempsey et al. 2013). In Simulation C, in which there was no coupling of slip to transmissivity
enhancement, the spreading of fluid pressure occurred in a diffusive pattern, similar to results from continuum simulations. In
Simulations A and B, the spreading of pressure was controlled by
coupled, nonlinear processes that are not diffusive in nature, and
the pattern of pressure spreading was quite different.
Sparsely connected networks arise from shear stimulation
because of a process that we refer to as crack-like shear stimulation
(CSS) [McClure and Horne (2011); Section 3.4.2.2 in McClure

Seconds Elapsed = 604800, Max: 12.4032, Min: 8.6831


12.5

200

12

150

1.5

11.5

100

11

50

10.5

0
50

10

100
150

200

100

100

200

0.5
50
0
0
0.5
50

9.5

100

150

200
8.5

Distance (m)
Fig. 3The final pressure distribution for Simulation C. The
black line represents the wellbore. The blue/red lines are natural fractures, with color proportional to fluid pressure. The figure can be viewed as showing a vertical cross section of the
formation.

1
1.5

150 100

50

50

100

150

Change in Coulomb stress (MPa)

Distance (m)

100

Distance (m)

150

Fluid Pressure (MPa)

200

Distance (m)
Fig. 4The final distribution of pressure and induced change
in CS (Eq. 5) for Simulation A. Gold/green contours in the background show the change in CS for fractures oriented 30 counterclockwise from the y-axis. The black line represents the
wellbore, and the blue/red lines represent natural fractures,
with color proportional to fluid pressure. The figure can be
viewed as showing a vertical cross section of the formation.

May 2014 SPE Reservoir Evaluation & Engineering


ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

237

150

1.5
1

Distance (m)

100

0.5
50
0
0
0.5
50

100

1.5

150

150 100

50

50

100

150

Distance (m)
Fig. 5The final distribution of pressure and induced change in
the absolute value of shear stress (Eq. 5) for Simulation A. Gold/
green contours in the background show the change in absolute
value of shear stress for fractures oriented 30 counterclockwise
from the y-axis. The black line represents the wellbore, and the
blue/red lines represent natural fractures, with color proportional
to fluid pressure. The figure can be viewed as showing a vertical
cross section of the formation.

Shear
Displacement (m)

(2012)]. Fig. 7 illustrates the CSS mechanism. Fig. 7 shows the distribution of fluid pressure, shear stress, frictional strength, shear displacement discontinuity, and transmissivity along a particular fault
in a network. Fluid is flowing from right to left. The fluid-pres0.06
0.04
0.02
0
0

50

100

150

200

250

Pressure (MPa)

Distance along Fracture (m)


60
50
40
30

3
Transmissivity (m ) Shear Stress (MPa), Blue
Friction (MPa), Red

10

10

10

50

100

150

200

250

Distance along Fracture (m)

10
5
0
50

100

150

200

250

Distance along Fracture (m)


10

15

20

50

100

150

200

250

Distance along Fracture (m)


Fig. 7Illustration of CSS. Reproduced from Fig. 3-33 in McClure
(2012). The distribution of transmissivity, shear stress, frictional
strength, pressure, and shear displacement discontinuity is
shown along a particular fracture in a fracture-network simulation
involving shear stimulation. The fluid is flowing from right to left.
238

Stage:

Page: 238

Total Pages: 11

1.00E+01

Simulation A
Simulation B
Simulation C

1.00E+00

1.00E01

1.00E02

1.00E+00

1.00E+01

1.00E+02

1.00E+03

Time elapsed (hours)


Fig. 6Normalized flow rate vs. time for Simulations A (red), B
(blue), and C (green). Flow rates for each simulation are normalized by dividing by their flow rate after 1 hour, 0.164 kg/s, 0.114
kg/s, and 1.077 kg/s, respectively.

sure front is at approximately 80 m, and the front of shear displacement discontinuity and transmissivity enhancement is at
approximately 40 m. The transmissivity at the fluid-pressure
front is much higher than the initial transmissivity (which was
extremely low in this particular simulation and is seen at the far
left side of the fracture).
CSS arises because of the interaction between induced stresses,
transmissivity enhancement, and fluid flow. Along a particular
fault, a concentration of shear stress develops at the boundary
between where slip has occurred and slip has not occurred (effectively, this is a shear crack tip). The concentration of shear stress
causes slip and transmissivity enhancement to propagate ahead of
the fluid-pressure front. This allows the fluid-pressure front to propagate along the fracture at a rate related to the stimulated, not the
initial, fracture transmissivity. The maximum magnitude of shear
stress that can be borne by the fracture in the region of stress concentration is limited by its frictional strength.
CSS can help stimulation propagate along a fracture, but cannot cause the first patch of slip to occur on a fracture. The initiation of slip on a fracture requires fluid to flow into it at the initial
transmissivity. If the initial transmissivity is much lower than the
stimulated transmissivity, slip will propagate along fractures
much more rapidly than slip can start on fractures. This creates an
episodic behavior in which individual fractures are stimulated in
relatively short durations of time, separated by periods of quiescence [Section 3.4.2.2 of McClure (2012)].
Change in CS, DCS, can be calculated to estimate whether
induced stresses, Ds and Drn0 , on a fracture will encourage or discourage slip (positive changes encourage slip):
DCS jDsj  lDr0n : . . . . . . . . . . . . . . . . . . . . . . . . 5

15

Normalized flow rate (unitless)

200

Change in abs() (Mpa)

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

Fig. 4 shows contours of induced changes in CS for l equal to


0.6 on a fracture oriented 30 counterclockwise from the y-axis
direction (equivalent to a fault dipping at 60 ), which is roughly
the orientation of one of the two fracture sets in Simulations A, B,
and C. The distributions of induced stresses for fractures orientated 30 clockwise from the y-axis direction are similar (but not
identical) and, for brevity, are not shown.
The stresses in Fig. 4 were calculated by assuming that there
were no fluid-pressure changes in the fractures caused by the
induced stresses. But, in fact, there was a poroelastic response to
changes in normal stress, Drn. The poroelastic response occurred
because void aperture E is proportional to effective normal stress
(Eq. 3) and water is only slightly compressible. To conserve
mass, the fluid pressure had to change in response to the induced
normal stresses to keep the effective normal stress nearly constant
(because the response was initially close to undrained). With liquid water filling the fractures, which would occur in most EGS
applications, the poroelastic response would counteract the
induced normal stress. Over time, the fluid pressure would reequilibrate with the surroundings, and induced change in effective
normal stress would become equal to the induced change in
May 2014 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

normal stress. This process has been proposed as one explanation


for the time delay in earthquake aftershocks (King 2007). The
magnitude of the poroelastic response is determined by the relative compressibility of fluid density and the void aperture. If the
fractures were full of gas, as in a gas-shale reservoir (in fractures
not yet reached by injection water), the much greater compressibility of the gas would make the poroelastic response on fluid
pressure less significant. Fig. 5 shows contours of induced
changes in the absolute value of shear stress, which is equal to the
change in CS in the limiting case of the undrained response for an
incompressible fluid (which reasonably approximates water).
Figs. 4 and 5 show that when the fractures slip, they reduce CS
and shear stress along their sides, creating a stress shadow that
inhibits their neighbors from slipping. The CS and shear-stress
changes are positive ahead of the crack tips, encouraging further
propagation of slip along the slipping fractures (and also potentially
encouraging slip on neighboring fractures ahead of the crack tip).
Shear fractures should not be expected to propagate farther in their
own plane (Segall and Pollard 1983), because the strength of intact
rock is much greater than the strength of pre-existing flaws. In other
words, the effective shear crack tip can propagate along a fault, but
after it reaches the tip of the fault, it will not propagate farther. Tensile splay fractures might initiate near the tips of faults because of
local stress concentration, but in Simulations A, B, and C, tensile
fractures were not permitted to form as a model assumption.
The result of these processes is that stimulation tends to localize into a smaller number of fractures, but propagate farther from
the wellbore than would be expected from a continuum model or
a DFN model that does not include shear stimulation, transient
fluid flow, and induced stresses (such as Simulation C, which
neglected shear stimulation). The tendency for localization to
occur depends on the average fracture length because, with
shorter fractures, it is necessary for flow to start on new fractures
more frequently. In Simulation A, individual fractures were quite
long, and the difference between the initial and stimulated transmissivity was quite high, both factors that encouraged the localization of stimulation.
In Simulation B, shear stimulation occurred readily, as in Simulation A. However, stimulation remained confined close to the
wellbore because the fracture network was not percolating (Fig.
2). If injection had been performed at a constant rate, rather than a
constant pressure, the fluid pressure would have eventually been
forced to rise, eventually leading to the propagation of new fractures through the formation and the generation of continuous pathways for flow. Because of the constant pressure injection, the
injection rate tended to zero as the fluid pressure in the fractures
connected to the wellbore became nearly equal to the injection
pressure.
Simulation B may have somewhat overestimated the importance of the natural-fracture-network percolation. As a simulation
assumption, no new fractures could form in the network. In reality, slip on pre-existing fractures could create local concentrations
of tensile stress that could cause new splay fractures to form
and propagate, even with fluid pressure less than the minimum
principal stress. These fractures could propagate only a limited
distance, however, confined to the region of local stress perturbation, unless fluid pressure exceeded the minimum principal stress.
Chipperfield et al. (2007) used a continuum-based model of
shear stimulation and made several observations that are not consistent with the observations that come from shear-stimulation
models run in CFRAC. Chipperfield et al. (2007) reported that the
rate of spreading of shear stimulation was higher if the initial fracture transmissivity was lower, and that the amount that permeability increased in response to shear did not affect the rate of
spreading. Models of propagation of shear stimulation (along a
single fracture) that couple fluid flow, induced stress, and shear
stimulation show that the rate of propagation is unrelated to the
initial formation permeability. This occurs because of the CSS
mechanism described previously. McClure and Horne (2010b)
derived an equation for the rate of spreading of shear stimulation
along a single, 1D fault, and matched it with numerical-simulation
data.

Stage:

Page: 239

Total Pages: 11

Flow Rate. In Fig. 6, flow rate during the simulations is


shown. Injection rates are normalized by the flow rate at 1 hour,
which was 0.11 kg/s in Simulation A, 0.17 kg/s in Simulation B,
and 1.08 kg/s in Simulation C. The qualitative behavior of flow
rate with time was different for the three models. In practice, it
may be possible to use flow-rate behavior over time to diagnose
the effectiveness of shear stimulation during a TSS test. However,
a fully realistic simulation of these transients would require a 3D
simulation that included flow in the matrix.
In Simulation C (no coupling of slip and transmissivity), the
injection rate had a negative one-half slope on the log-log plot
(Fig. 6), indicating decay with the inverse of the square root of
time. This behavior is consistent with the 1D solution to the diffusivity equation for a constant-pressure boundary adjacent to an infinite half-space (Bird et al. 2006). This is the same behavior that
would have been expected in a continuum pressure-transient
model, in analogy to the matrix linear-, ellipsoidal-, and lateradial-flow-regime sequence expected from a hydraulically fractured well connected to an infinite-conductivity fracture [Chapter
11 of Kamal (2009)].
In Simulations A and B, shear stimulation occurred, arresting
the decay in injection rate. The shape of the normalized rate curve
for these two simulations started identically. In Simulation B, the
boundary of the reservoir (the fractures connected to the wellbore through a percolating flow pathway) began to be felt after a
few hours, causing a downward deviation in the flow-rate curve.
After approximately 100 hours, the rate dropped much more rapidly than in Simulation C.
In practice, a comparison of the observed flow rate vs. time
curve to the flow rate vs. time curve expected from a standard
continuum model (assuming constant transmissivity) could be
used to diagnose the presence of shear stimulation. In two dimensions, the continuum result was linear (for early time), and in
three dimensions, the continuum result would be radial. The welltest solution for radial constant-pressure injection was given by
Ehlig-Economides (1979). In practice, this solution would probably need to include dual-porosity behavior. Shear stimulation
would be indicated by an upward deviation in rate higher than the
expected result. An anomalously rapid decay in rate could be an
indication of a poorly connected network. These trends in rate
could be caused by other reservoir effects, such as boundary
effects; thus, it would be useful to perform an injectivity test
some period after the TSS test (with a constant pressure sufficiently low to avoid shear stimulation) and to compare between
the two transients.
Thermal Stresses. Thermal stresses have the potential to be
important if a long-term injection of colder fluid is performed.
Thermal contraction could cause the formation and propagation
of tensile fractures through the formation. If tensile fractures
form, it could increase the injectivity of the well, arresting the
expected decline in injection rate and creating an effect that could
be mistaken for shear stimulation. Presumably, this increase in
permeability would be reversible. This behavior needs to be
investigated in future work.
Fracture Creep. Even if fractures experience shear stimulation, the asperities in the fracture walls could creep over time and
cause a gradual reduction in transmissivity. This phenomenon
would be particularly likely in clay-rich formations. For this reason, it would be best to perform a post-TSS-test injectivity test
several weeks later.

Combining TSS Tests With Standard Well Tests


The simplest way to interpret a TSS test would be to perform
standard pressure-transient tests, such as injection/falloff tests,
before and after the TSS test. Shear stimulation creates a permanent increase in formation permeability, and this could be
observed by comparing the permeabilities estimated from these
transients. The falloff that occurs at the end of the TSS test probably could not be used because the redistribution of fluid pressure
after injection could cause shear stimulation to occur at the

May 2014 SPE Reservoir Evaluation & Engineering


ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

239

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

periphery of the stimulated region, creating nonlinearity that is


not included in standard well-test models.
The injection/falloff test performed before the TSS test would
need to be performed with a bottomhole pressure sufficiently low
to avoid causing shear stimulation to occur. The injection/falloff
test performed after the TSS test would not be expected to cause
further shear stimulation as long as pressures and volumes were
kept less than the values used during the TSS test because fractures
that have already slipped in response to a particular fluid pressure
will not slip again when returned to that same fluid pressure.
A production/buildup transient test could be useful after the
TSS test because the production of fluid from the formation would
tend to reheat the region of the formation that was cooled by
injection, which would reduce any potential impact from thermal
stresses.
Desert Peak EGS Project. There were two low-rate stimulations
attempted at the Desert Peak project, both carried out with downhole fluid pressure at less than the minimum principal stress. At
the beginning of the first stimulation, the injectivity was estimated
at 0.092 (L/s)/MPa. At the end of the first injection period, after
10 days of injection, the injectivity was estimated at 1.2 (L/s)/
MPa. At the beginning of the second injection period, injectivity
was estimated at 0.55 (L/s)/MPa, and at the end of the period,
injectivity was estimated at 1.4 (L/s)/MPa. After a period of shutin, an injectivity test was performed that indicated that the injectivity was 0.37 (L/s)/MPa (Chabora et al. 2012). Therefore, injectivity tests performed after both of the stimulation periods
indicate that the injectivity gains were partially reversible.
The partial reversibility of the injectivity increase may indicate
that thermal fracturing, rather than shear stimulation, was responsible for the increase in injectivity during the test. Thermal fracturing would be expected to be reversible because the reheating of
the formation would be expected after shut-in, which would
reclose some of the fractures unloaded by thermal stresses. Injection was performed for 10 days for the first stimulation and 9 days
for the second, and the formation was very hot, approximately
180 C (Dempsey et al. 2013), supporting the hypothesis that thermal stimulation accounted for some or most of the observed
stimulation.
To explain the partial reversibility of the injectivity gain,
Dempsey et al. (2013) said, this may be due to poroelastic or
thermoelastic fracture closure due to thermal recovery and pressure dissipation, chemical sealing or creep closure of shear
fractures. Processes such as chemical sealing or creep closure
cannot be ruled out but are difficult to characterize. It could have
been possible to distinguish between thermal effects and at least
some of the other possible effects by performing a subsequent
constant-pressure injection test with a pressure equal to or lower
than the previous pressures used. Under these conditions, fracture slip would not have been expected (because a fracture
should not reslip when exposed to a fluid pressure it has already
experienced), and, thus, if the injectivity had been subsequently
recovered, this would have suggested thermal effects were
responsible.
Regardless of the reason for the reversibility of the injectivity
gains, the results at the Desert Peak project indicate that the formation had modest TSS. Low-rate injections increased the injectivity, but much of the gain was lost during subsequent injectivity
tests. The overall increase between the initial injectivity and the
injectivity measured during a subsequent test was approximately
a factor of four, but this was from a low initial injectivity. The
overall gain in injectivity from weeks of low-rate injection was
(0.37 0.092) 0.28 (L/s)/MPa. Because of the possible role of
thermal fracturing in the reversibility of injectivity gains, it is possible that the injectivity gains would have further reversed with
greater shut-in time. The low TSS was recognized by the project
managers, who subsequently injected at a high rate with the intention of performing tensile fracturing, and achieved a much more
substantial increase in well injectivity (Chabora et al. 2012).
It was unique that the operators of this project committed to
injecting at a bottomhole fluid pressure less than the minimum
240

Stage:

Page: 240

Total Pages: 11

principal stress, effectively performing what we propose to call a


TSS test. This decision made it possible to unambiguously determine that the formations TSS was modest. In all other EGS
projects, injection has been performed at a specified rate. At a
specified rate, if shear stimulation is ineffective, injection pressure
will rise until the fluid pressure reaches the minimum principal
stress and potentially causes tensile fracturing. If this occurs, the
relative importance of new and pre-existing fractures in contributing to permeability enhancement will be ambiguous. Therefore,
from an experimental design point of view, the Desert Peak project was the first EGS project (we are aware of) in which it was
even theoretically possible to demonstrate unambiguously that
shear stimulation had limited effectiveness.
Soultz-sous-Forets EGS Project. The EGS project at Soultzsous-Forets involved several wells drilled and stimulated to
depths of 3.5 and 5 km in granite. All wells were stimulated by
injecting water with no proppant into long openhole wellbore sections. During the June 2000 stimulation of the Well GPK2, fluid
was injected at rates up to 50 L/s (Weidler et al. 2002) without the
fluid pressure reaching the minimum principal stress (Valley and
Evans 2007). Hydraulic tests before the stimulation estimated an
injectivity of approximately 0.2 (L/s)/MPa. Hydraulic tests immediately after the test did not give conclusive estimates of injectivity (Weidler 2000), but a test in 2003 (with no intervening
stimulations having been performed) showed a peak overpressure
of 5 MPa after several days of injection at 15 L/s, equivalent to an
injectivity of 3.0 (L/s)/MPa (Hettkamp et al. 2004). Wellbore
observations in GPK2 were not possible because of a wellbore
obstruction, but in neighboring GPK3, production logs confirmed
that flow from the wellbore was concentrated at pre-existing,
large-scale fault zones, and critical-stress analysis and caliper logs
confirmed that these faults had failed in shear as a result of injection (Evans 2005; Evans et al. 2005).
Overall, the results at the Soultz project indicate that the formation had a high TSS. Injection at pressure less than the minimum principal stress resulted in a sustained, large increase in the
well injectivity, and wellbore observations confirmed that this
increase was a result of induced slip on the pre-existing faults.
The increase in injectivity was approximately 2.8 (L/s)/MPa, ten
times greater than the injectivity increase at the Desert Peak
project.
Unlike at the Desert Peak project, the fluid pressure was not
kept less than the minimum principal stress intentionally. Instead,
shear stimulation was so effective at increasing injectivity that the
specified-rate injection was unable to increase fluid pressure sufficiently. Probably, the difference in TSS occurred because at
Soultz, large, thick fault zones were present (Evans 2005), and at
Desert Peak, fractures and faults intersecting the wellbore were no
thicker than a few centimeters (Benato et al. 2013). McClure and
Horne (2013b) argued that in very-low-matrix-permeability settings, greater fault-zone thickness should increase TSS. McClure
and Horne (2012) reviewed historical EGS projects and found
that there was a strong correlation between the thickness of
observed fault zones and the magnitude of induced seismicity.
TSS Tests in Shale
We are not aware of a project in shale in which injection was intentionally carried out at a fluid pressure less than the minimum principal stress. We believe these tests could be very beneficial because
they would resolve the fundamental uncertainty about whether slip
on pre-existing fractures contributes significantly to production. The
results of these tests may depend on the formation properties, and it
would be valuable to perform tests in different wells and seek correlations between the results and geological/geophysical parameters.
Conclusions
Shear stimulation is a potentially important process in the hydraulic stimulation of shale formations and in EGS. However, in practice, the role of shear stimulation is inferred and subject to
interpretation. The TSS highly depends on local geological factors
May 2014 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

such as fracture orientation and ability to self-prop and could be


considered a formation property.
Understanding the importance of shear stimulation will have
important implications for modeling, stimulation design, and resource assessment in shale and geothermal resources. The TSS test
is a simple, relatively unambiguous way to directly measure the
effect of shear stimulation in situ. By the use of modeling, we
showed qualitatively how a TSS test might be interpreted, how geological parameters could affect the results, and how shear stimulation can affect the process of fluid propagation in a fracture
network. We compared two projects as examples of TSS teststhe
EGS projects at Desert Peak and Soultzand discussed how their
results demonstrate the difference between high and low TSS.
TSS tests could be interpreted by performing pressure-transient tests before and after the TSS test and comparing the interpreted formation permeability. It also may be possible to interpret
TSS tests directly by analyzing the trends in rate over time, but
developing this technique in future work will require 3D simulations that include matrix flow.
Nomenclature
D cumulative fracture-sliding displacement
discontinuity, m
DE,eff, De,eff effective cumulative fracture-sliding displacement discontinuity, m
DE,eff,max,, De,eff,max maximum effective cumulative sliding
displacement discontinuity, m
e hydraulic aperture, m
E void aperture, m
E0, e0 reference aperture, m
h out-of-plane dimension in the 2D simulations, m
k permeability, m2
P fluid pressure, MPa
S0 fracture cohesion, MPa
T transmissivity, m3
v fracture sliding velocity, m/s
g radiation-damping coefficient, MPa/(m/s)
lf coefficient of friction, unitless
rn normal stress, MPa
r0n effective normal stress, MPa
rn,Eref, rn,eref 90% fracture-closure stress, MPa
s shear stress, MPa
Up Poissons ratio, unitless
/Edil, /edil shear-dilation angle, degrees

Acknowledgments
Thank you the Precourt Institute for Energy for supporting this
research from 2009 to 2012. Thank you to three anonymous
reviewers and to the editorial staff for their excellent comments
and revisions.
References
Baisch, Stefan, Weidler, Ralph, Voros, Robert et al. 2006. Induced Seismicity During the Stimulation of a Geothermal HFR Reservoir in the
Cooper Basin, Australia. Bull. Seismological Soc. Am. 96 (6):
22422256. http://dx.doi.org/10.1785/0120050255-PA.
Barton, N., Bandis, S., and Bakhtar, K. 1985. Strength, Deformation, and
Conductivity Coupling of Rock Joints. International J. Rock Mechanics and Mining Sci. & Geomechanics Abstracts 22 (3): 121140.
http://dx.doi.org/10.1016/0148-9062(85)93227-9-PA.
Benato, S., Reeves, D.M., Parashar, R. et al. 2013. Computational Investigation of Hydro-Mechanical Effects on Transmissivity Enhancement
During the Initial Injection Phases at the Desert Peak EGS Project,
NV. Paper presented at the Thirty-Eighth Workshop on Geothermal
Reservoir Engineering, Stanford, California.
Bird, R. Byron, Stewart, Warren E., and Lightfoot, Edwin N. 2006. Transport Phenomena, second edition, New York: John Wiley & Sons, Inc.

Stage:

Page: 241

Total Pages: 11

Blanton, Thomas. 1982. An Experimental Study of Interaction Between


Hydraulically Induced and Pre-Existing Fractures. Paper SPE 10847
presented at the SPE Unconventional Gas Recovery Symposium, Pittsburgh, Pennsylvania, 1618 May. http://dx.doi.org/10.2118/10847MS.
Bowker, Kent A. 2007. Barnett Shale Gas Production, Fort Worth Basin:
Issues and Discussion. AAPG Bull. 91 (4): 523533. http://dx.doi.org/
10.1306/06190606018-PA.
Bradley, Andrew M. 2011. H-matrix and block error tolerances.
arXiv:1110.2807, source code available at https://pangea.stanford.edu/
research/CDFM/software/index.html, paper available at http://arxiv.org/abs/1110.2807.
Brown, D.W. 1989. The Potential for Large Errors in the Inferred Minimum Earth Stress When Using Incomplete Hydraulic Fracturing
Results. International J. Rock Mechanics and Mining Sci. & Geomechanics Abstracts 26 (6): 573577. http://dx.doi.org/10.1016/01489062(89)91437-X.
Bruel, D. 1995. Heat Extraction Modelling From Forced Fluid Flow
Through Stimulated Fractured Rock Masses: Application to the Rosemanowes Hot Dry Rock Reservoir. Geothermics 24 (3): 361374.
http://dx.doi.org/10.1016/0375-6505(95)00014-H.
Bruel, Dominique. 2007. Using the Migration of the Induced Seismicity as
a Constraint for Fractured Hot Dry Rock Reservoir Modelling. International J. Rock Mechanics and Mining Sci. 44 (8): 11061117. http://
dx.doi.org/10.1016/j.ijrmms.2007.07.001.
Chabora, Ethan, Zemach, Ezra, Spielman, Paul et al. 2012. Hydraulic
Stimulation of Well 27-15, Desert Peak Geothermal Field, Nevada,
USA. Paper presented at the Thirty-Seventh Workshop on Geothermal
Reservoir Engineering, Stanford University, Stanford, California.
Chipperfield, S.T., Wong, J.R., Warner, D.S. et al. 2007. Shear Dilation
Diagnostics: A New Approach for Evaluating Tight Gas Stimulation
Treatments. Paper SPE 106289 presented at the SPE Hydraulic Fracturing Technology Conference, College Station, Texas, 2931 January.
http://dx.doi.org/10.2118/106289-MS.
Chuprakov, Dimitry, Olga Melchaeva, Romain Prioul. 2013. Hydraulic
fracture propagation across a weak discontinuity controlled by fluid
injection. Presented at the ISRM International Conference for Effective and Sustainable Hydraulic Fracturing, 2022 May, Brisbane, Australia. ISRM-ICHF-2013-008.
Cipolla, Craig, Warpinski, Norman, Mayerhofer, Michael et al. 2008. The
Relationship Between Fracture Complexity, Reservoir Properties, and
Fracture Treatment Design. Paper SPE 115769 presented at the SPE
Annual Technical Conference and Exhibition, Denver, Colorado,
2124 September. http://dx.doi.org/10.2118/115769-MS.
Cipolla, C.L., Lolon, E.P., Erdle, J.C. et al. 2010. Reservoir Modeling in
Shale-Gas Reservoirs. SPE Res Eval & Eng 13 (4): 638-653. http://
dx.doi.org/ 10.2118/125530-PA.
Cladouhos, Trenton, Petty, Susan, Larson, Ben et al. 2009. Towards More
Efficient Heat Mining: A Planned Enhanced Geothermal System Demonstration Project. GRC Trans. 33: 165-170.
Cladouhos, Trenton T., Clyne, Matthew, and Nichols, Maisie et al. 2011.
Newberry Volcano EGS Demonstration Stimulation Modeling. Geothermal Resources Council Trans. 35: 317-322.
Dahi-Taleghani, Arash and Olson, Jon. 2009. Modeling Simultaneous
Growth of Multiple Hydraulic Fractures and Their Interaction With
Natural Fractures. Paper SPE 119739 presented at the SPE Hydraulic
Fracturing Technology Conference, The Woodlands, Texas, 1921
January. http://dx.doi.org/10.2118/119739-MS.
Damjanac, Branko, Gil, Ivan, Pierce, Matt et al. 2010. A New Approach
to Hydraulic Fracturing Modeling in Naturally Fractured Reservoirs.
Paper presented at the 44th US Rock Mechanics Symposium and 5th
US-Canada Rock Mechanics Symposium, Salt Lake City, Utah.
Dempsey, D., Kelkar, S., Lewis, K. et al. 2013. Modeling Shear Stimulation of the Desert Peak EGS Well 27-15 Using a Coupled ThermalHydrological-Mechanical Simulator. Paper presented at the 47th
United States Rock Mechanics/Geomechanics Association Symposium, San Francisco, California.
Ehlig-Economides, Christine. 1979. Well Test Analysis for Wells Produced
at a Constant Pressure. PhD Thesis, Stanford University (June 1979).
Esaki, T., Du, S., Mitani, Y. et al. 1999. Development of a Shear-Flow
Test Apparatus and Determination of Coupled Properties for a Single

May 2014 SPE Reservoir Evaluation & Engineering


ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

241

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

Rock Joint. International J. Rock Mechanics and Mining Sci. 36 (5):


641650. http://dx.doi.org/10.1016/S0148-9062(99)00044-3.
Evans, Keith F. 2005. Permeability Creation and Damage Due To Massive
Fluid Injections Into Granite At 3.5 Km At Soultz: 2. Critical Stress
and Fracture Strength. J. Geophysical Research 110 (B4). http://
dx.doi.org/10.1029/2004JB003169.
Evans, Keith F., Genter, Albert and Sausse, Judith. 2005. Permeability
Creation and Damage Due To Massive Fluid Injections Into Granite at
3.5 Km at Soultz: 1. Borehole Observations. J. Geophysical Research
110 (B4). http://dx.doi.org/10.1029/2004JB003168.
Fan, Li, Thompson, John, and Robinson, John. 2010. Understanding Gas
Production Mechanism and Effectiveness of Well Stimulation in the
Haynesville Shale Through Reservoir Simulation. Paper SPE 136696
presented at the Canadian Unconventional Resources and International
Petroleum Conference, Calgary, Alberta, Canada, 1921 October.
http://dx.doi.org/10.2118/136696-MS.
Fisher, M.K., Heinze, J.R., Harris, C.D. et al. 2004. Optimizing Horizontal
Completion Techniques in the Barnett Shale Using Microseismic Fracture Mapping. Paper SPE 90051 presented at the SPE Annual Technical Conference and Exhibition, Houston, Texas, 2629 September.
http://dx.doi.org/10.2118/90051-MS.
Fu, Pengcheng, Johnson, Scott M., and Carrigan, Charles R. 2012. An Explicitly Coupled Hydro-Geomechanical Model for Simulating Hydraulic Fracturing in Arbitrary Discrete Fracture Networks. International
J. Numerical and Analytical Methods in Geomechanics. http://
dx.doi.org/10.1002/nag.2135.
Gale, Julia F.W., Reed, Robert M. and Holder, Jon. 2007. Natural Fractures in
the Barnett Shale and Their Importance for Hydraulic Fracture Treatments.
AAPG Bull. 91 (4): 603622. http://dx.doi.org/10.1306/11010606061.
Gu, H. and Weng, X. 2010. Criterion for Fractures Crossing Frictional
Interfaces at Non-Orthogonal Angles. Paper presented at the 44th US
Rock Mechanics Symposium and 5th US-Canada Rock Mechanics
Symposium, Salt Lake City, Utah.
Gu, Hongren, Weng, Xiaowei, Lund, Jeffrey et al. 2011. Hydraulic Fracture Crossing Natural Fracture at Non-Orthogonal Angles, a Criterion,
Its Validation and Applications. Paper SPE 139984 presented at the
SPE Hydraulic Fracturing Technology Conference, The Woodlands,
Texas, 2426 January. http://dx.doi.org/10.2118/139984-MS.
Hettkamp, T., Baumgartner, J. Baria, R. et al. 2004. Electricity Production
From Hot Rocks. Paper presented at the Twenty-Ninth Workshop on
Geothermal Reservoir Engineering, Stanford University.
Hickman, S.H. and Davatzes, N.C. 2010. In-situ stress and fracture characterization for planning of an EGS stimulation in the Desert Peak Geothermal Field, Nevada. Paper presented at the Thirty-Fifth Workshop on
Geothermal Reservoir Engineering, Stanford University, Stanford, CA.
Horne, Roland N. 1995. Modern Well Test Analysis: A Computer-Aided
Approach, second edition. Palo Alto, California: Petroway, Inc.
Ito, Hisatoshi. 2003. Inferred Role of Natural Fractures, Veins, and Breccias in Development of the Artificial Geothermal Reservoir at the Ogachi Hot Dry Rock Site, Japan. J. Geophysical Research 108 (B9).
http://dx.doi.org/10.1029/2001JB001671.
Ito, T. and Hayashi, K. 2003. Role of Stress-Controlled Flow Pathways in
HDR Geothermal Reservoirs. Pure and Applied Geophysics 160 (5):
11031124. http://dx.doi.org/10.1007/PL00012563.
Ispas, I., Stanchits, S., Burghardt, J., et al. 2013. Understanding the effect
of rock fabric on fracture complexity for improving completion design
and well performance. Presented at the International Petroleum Technology Conference, 2628 March, Beijing, China. IPTC-17018-MS.
http://dx.doi.org/10.2523/17018-MS.
Jeffrey, Robert, Zhang, Xi, and Thiercelin, Marc. 2009. Hydraulic Fracture Offsetting in Naturally Fractured Reservoirs: Quantifying a LongRecognized Process. Paper SPE 119351 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 1921
January. http://dx.doi.org/10.2118/119351-MS.
Jing, Z., Willis-Richards, J. Watanabe, K. et al. 2000. A Three-Dimensional Stochastic Rock Mechanics Model of Engineered Geothermal
Systems in Fractured Crystalline Rock. J. Geophysical Research 105
(B10): 2366323679. http://dx.doi.org/10.1029/2000JB900202.
Kaieda, H., Sasaki, S., and Wyborn, D. 2010. Comparison of Characteristics of Micro-Earthquakes Observed During Hydraulic Stimulation
242

Stage:

Page: 242

Total Pages: 11

Operations in Ogachi, Hijiori, and Cooper Basin HDR Projects. Paper


presented at the World Geothermal Congress, Bali, Indonesia.
Kamal, Medhat M. ed. 2009. Transient Well Testing, Vol. Monograph Series Vol. 23, Society of Petroleum Engineers.
King, G.C.P. 2007. 4.08Fault Interaction, Earthquake Stress Changes,
and the Evolution of Seismicity. In Treatise on Geophysics, ed. G.
Schubert 225255. Amsterdam: Elsevier.
King, George. 2010. Thirty Years of Gas Shale Fracturing: What Have
We Learned? Paper SPE 133456 presented at the SPE Annual Technical Conference and Exhibition, Florence, Italy, 1922 September.
http://dx.doi.org/10.2118/133456-MS.
Kohl, T. and Hopkirk, R.J. 1995. FRACTUREA Simulation Code for
Forced Fluid Flow and Transport in Fractured, Porous Rock. Geothermics 24 (3): 333343. http://dx.doi.org/10.1016/0375-6505(95)00012-F.
Kohl, T. and Megel, T. 2007. Predictive Modeling of Reservoir Response
to Hydraulic Stimulations at the European EGS Site Soultz-SousForets. International J. Rock Mechanics and Mining Sci. 44 (8):
11181131. http://dx.doi.org/10.1016/j.ijrmms.2007.07.022.
Lanyon, G.W., Batchelor, A.S., and Ledingham, P. 1993. Results From a
Discrete Fracture Network Model of a Hot Dry Rock System. Paper
presented at the Eighteenth Workshop on Geothermal Reservoir Engineering, Stanford University.
Laubach, Stephen E., Olson , Jon E., and Gale, Julia F.W. 2004. Are Open
Fracture Necessary Aligned With Maximum Horizontal Stress? Earth
and Planetary Science Lett. 222: 191195. http://dx.doi.org/10.1016/
j.epsl.2004.02.019.
Ledesert, Beatrice, Hebert, Ronan, Genter, Albert et al. 2010. Fractures,
Hydrothermal Alterations and Permeability in the Soultz Enhanced
Geothermal System. Comptes Rendus Geosci. 342 (78): 607615.
http://dx.doi.org/10.1016/j.crte.2009.09.011.
Liu, Enru. 2005. Effects of Fracture Aperture and Roughness on Hydraulic
and Mechanical Properties of Rocks: Implication of Seismic Characterization of Fractured Reservoirs. J. Geophysics and Eng. 2 (1):
3847. http://dx.doi.org/10.1088/1742-2132/2/1/006.
Lutz, S.J., Hickman, S. Davatzes, N. et al. 2010. Rock Mechanical Testing
and Petrological Analysis in Support of Well Stimulation Activities at
the Desert Peak Geothermal Field, Nevada. Paper presented at the
Thirty-Fifth Workshop on Geothermal Reservoir Engineering, Stanford University.
Mahrer, Kenneth D. 1999. A review and perspective on far-field hydraulic
fracture geometry studies. J. Petrol. Sci. and Eng. 24 (1): 1328.
http://dx.doi.org/10.1016/S0920-4105(99)00020-0.
Mayerhofer, Michael, Lolon, Elyezer, and Warpinski, Norman et al. 2010.
What Is Stimulated Reservoir Volume? SPE Prod & Oper 25 (1):
8998. http://dx.doi.org/10.2118/119890-PA.
McClure, M.W. 2012. Modeling and Characterization of Hydraulic Stimulation and Induced Seismicity in Geothermal and Shale Gas Reservoirs.
PhD Thesis, Stanford University, Stanford, California (December 2012).
McClure, M.W. and Horne, R.N. 2010a. Discrete Fracture Modeling of
Hydraulic Stimulation in Enhanced Geothermal Systems. Paper presented at the Thirty-Fifth Workshop on Geothermal Reservoir Engineering, Stanford University.
McClure, Mark W. and Horne, Roland N. 2010b. Numerical and Analytical Modeling of the Mechanisms of Induced Seismicity During Fluid
Injection. Geothermal Resources Council Trans. 34: 381396.
McClure, M.W. and Horne, Roland N. 2011. Investigation of InjectionInduced Seismicity Using a Coupled Fluid Flow and Rate/State Friction Model. Geophysics 76 (6): WC181WC198. http://dx.doi.org/
10.1190/geo2011-0064.1.
McClure, M.W. and Horne, R.N. 2012. The Effect of Fault Zone Development on Induced Seismicity. Paper presented at the Thirty-Seventh
Workshop on Geothermal Reservoir Engineering, Stanford, California.
McClure, Mark W. and Horne, Roland N. 2013a. Discrete Fracture Network Modeling of Hydraulic Stimulation: Coupling Flow and Geomechanics: Springer Briefs in Earth Sciences, Springer.
McClure, M.W. and Horne, R.N. 2013b. Is Pure Shear Stimulation Always
the Mechanism of Stimulation in EGS? Paper presented at the ThirtyEighth Workshop on Geothermal Reservoir Engineering, Stanford,
California.
Murphy, H.D. and Fehler, M.C. 1986. Hydraulic Fracturing of Jointed Formations. Paper SPE 14088 presented at the International Meeting on
May 2014 SPE Reservoir Evaluation & Engineering

ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

REE166332 DOI: 10.2118/166332-PA Date: 30-April-14

Petroleum Engineering, Beijing, China, 1720 March. http://dx.doi.org/


10.2118/14088-MS.
Mutlu, O. and Pollard, D.D. 2008. On the Patterns of Wing Cracks Along in
Outcrop Scale Flaw: A Numerical Modeling Approach Using Complementarity. J. Geophysical Research 113 (B6). http://dx.doi.org/10.1029/
2007JB005284.
Nagel, Neal, Gil, Ivan, Sanchez-Nagel, Marisela et al. 2011. Simulating
Hydraulic Fracturing in Real Fractured RocksOvercoming the Limits of Pseudo3D Models. Paper SPE 140480 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas,
2426 January. http://dx.doi.org/10.2118/140480-MS.
Olson, Jon E. 2004. Predicting Fracture SwarmsThe Influence of Subcritical Crack Growth and the Crack-Tip Process Zone on Joint Spacing in Rock. In The Initiation, Propagation, and Arrest of Joints and
Other Fractures, ed. J.W. Cosgrove and T. Engelder, 7388. London:
Geological Society, Special Publications, Geological Society of
London.
Olson, Jon E., Laubach, Stephen E., and Lander, Robert H. 2009. Natural
Fracture Characterization in Tight Gas Sandstones: Integrating
Mechanics and Diagenesis. Bull. Am. Assoc. Petrol. Geolog. 93 (11):
15351549. http://dx.doi.org/10.1306/08110909100.
Palmer, Ian, Moschovidis, Zissis, and Cameron, John. 2007. Modeling
Shear Failure and Stimulation on the Barnett Shale After Hydraulic
Fracturing. Paper SPE 106113 presented at the SPE Hydraulic Fracturing Technology Conference, College Station, Texas, 2931 January.
http://dx.doi.org/10.2118/106113-MS.
Pine, R.J. and Batchelor, A.S. 1984. Downward Migration of Shearing in
Jointed Rock During Hydraulic Injections. International J. Rock
Mechanics and Mining Sci. & Geomechanics Abstracts 21 (5): 249
263. http://dx.doi.org/10.1016/0148-9062(84)92681-0.
Rachez, X. and Gentier, S. 2010. 3D-Hydromechanical Behavior of a
Stimulated Fractured Rock Mass. Paper presented at the World Geothermal Congress, Bali, Indonesia.
Rahman, M.K., Hossain, M.M., and Rahman, S.S. 2002. A Shear-DilationBased Model for Evaluation of Hydraulically Stimulated Naturally Fractured Reservoirs. International J. for Numerical and Analytical Methods in
Geomechanics 26 (5): 469497. http://dx.doi.org/10.2118/10.1002/nag.208.
Renshaw, C.E. and Pollard, D.D. 1995. An Experimentally Verified Criterion for Propagation Across Unbounded Frictional Interfaces in Brittle,
Linear Elastic Materials. International J. Rock Mechanics and Mining
Sci. & Geomechanics Abstracts 32 (3): 237249. http://dx.doi.org/
10.1016/0148-9062(94)00037-4.
Riahi, Azadeh and Damjanac, Branko. 2013. Numerical Study of HydroShearing in Geothermal Reservoirs With a Pre-Existing Discrete Fracture Network. Paper presented at the Thirty-Eighth Workshop on Geothermal Reservoir Engineering, Stanford, California.
Rice, James R. 1993. Spatio-Temporal Complexity of Slip on a Fault. J.
Geophysical Research 98 (B6): 98859907. http://dx.doi.org/10.1029/
93JB00191.
Rogers, Stephen, Elmo, Davide, Dunphy, Rory et al. 2010. Understanding
Hydraulic Fracture Geometry and Interactions in the Horn River Basin
Through DFN and Numerical Modeling. Paper SPE 137488 presented
at the Canadian Unconventional Resources and International Petroleum Conference, Calgary, Alberta, Canada, 1921 October. http://
dx.doi.org/10.2118/137488-MS.
Roussel, Nicolas and Sharma, Mukul. 2011. Strategies To Minimize Frac
Spacing and Stimulate Natural Fractures in Horizontal Completions.
Paper SPE 146104 presented at the SPE Annual Technical Conference
and Exhibition, Denver, Colorado, 30 October2 November. http://
dx.doi.org/10.2118/146104-MS.
Segall, Paul. 2010. Earthquake and Volcano Deformation. Princeton, New
Jersey: Princeton University Press.
Segall, Paul and Pollard, David D. 1983. Nucleation and Growth of Strike
Slip Faults in Granite. J. Geophysical Research 88 (B1): 555568.
http://dx.doi.org/ 10.1029/JB088iB01p00555.
Shou, K.J. and Crouch, S.L. 1995. A Higher Order Displacement Discontinuity Method for Analysis of Crack Problems. International J. Rock
Mechanics and Mining Sci. & Geomechanics Abstracts 32 (1): 4955.
http://dx.doi.org/10.1016/0148-9062(94)00016-V.

Stage:

Page: 243

Total Pages: 11

Tester, J. ed. 2006. The Future of Geothermal Energy: Impact of Enhanced


Geothermal Systems (EGS) on the United States in the 21st Century,
Massachusetts Institute of Technology.
Valley, B. and Evans, K. 2007. Stress State at Soultz-Sous-Forets to 5 Km
Depth From Wellbore Failure and Hydraulic Observations. Paper presented at the Thirty-Second Workshop on Geothermal Reservoir Engineering, Stanford University.
Warpinski, N.R., Lorenz, J.C., Branagan, P.T. et al. 1993. Examination of
a Cored Hydraulic Fracture in a Deep Gas Well. SPE Prod & Fac 8
(3): 150158. http://dx.doi.org/10.2118/22876-PA.
Warpinski, N.R. and Teufel, L.W.. 1987. Influence of Geologic Discontinuities on Hydraulic Fracture Propagation. J. Pet Tech 39 (2):
209220. http://dx.doi.org/10.2118/13224-PA.
Warpinski, N.R., Wolhart, S.L., and Wright, C.A. 2001. Analysis and Prediction of Microseismicity Induced by Hydraulic Fracturing. Paper
SPE 71649 presented at the SPE Annual Technical Conference and
Exhibition, New Orleans, Louisiana, 30 September3 October. http://
dx.doi.org/10.2118/71649-MS.
Weidler, R. 2000. Hydraulic Stimulation of the 5 km Deep Well GPK-2,
Bureau de Recherches Geologiques et Minie`res (BRGM).
Weidler, R., Gerard, A., Baria, A. et al. 2002. Hydraulic and Micro-Seismic Results of a Massive Stimulation Test at the 5 Km Depth at the
European Hot-Dry Rock Test Site Soultz, France. Paper presented at
the Twenty-Seventh Workshop on Geothermal Reservoir Engineering,
Stanford University.
Weng, Xiaowei, Kresse, O., Cohen, C.-E. et al. 2011. Modeling of Hydraulic-Fracture-Network Propagation in a Naturally Fractured Formation. SPE Prod & Oper 26 (4). http://dx.doi.org/10.2118/140253-PA.
Willis-Richards, J., Green, A.S.P., and Jupe, A.J. 1995. A Comparison of
HDR Geothermal Sites. Paper presented at the World Geothermal
Congress, Florence, Italy.
Willis-Richards, J., Watanabe, K., and Takahashi, H. 1996. Progress Toward a Stochastic Rock Mechanics Model of Engineered Geothermal
Systems. J. Geophysical Research 101 (B8): 1748117496. http://
dx.doi.org/10.1029/96JB00882.
Witherspoon, P.A., Wang, J.S.Y., Iwai, K. et al. 1980. Validity of Cubic Law
for Fluid Flow in a Deformable Rock Fracture. Water Resources Research
16 (6): 10161024. http://dx.doi.org/10.1029/WR016i006p01016.
Wu, R., Kresse, O., Weng, X. et al. 2012. Modeling of Interaction of Hydraulic Fractures in Complex Fracture Networks. Paper SPE 152052
presented at the SPE Hydraulic Fracturing Technology Conference, The
Woodlands, Texas, 68 February. http://dx.doi.org/10.2118/152052MS.
Zhou, Jian, Chen, Mian, Jin, Yan et al. 2008. Analysis of Fracture Propagation Behavior and Fracture Geometry Using a Tri-Axial Fracturing
System in Naturally Fractured Reservoirs. International J. Rock
Mechanics and Mining Sci. 45 (7): 11431152. http://dx.doi.org/
10.1016/j.ijrmms.2008.01.001.
Zhou, Xiaoxian and Ghassemi, Ahmad. 2011. Three-Dimensional Poroelastic Analysis of a Pressurized Natural Fracture. International J. Rock
Mechanics and Mining Sci. 48 (4): 527534. http://dx.doi.org/10.1016/
j.ijrmms.2011.02.002.
Zimmermann, G., Reinicke, A., Brandt, W. et al. 2008. Results of Stimulation Treatments at the Geothermal Research Wells in Grob Schonebeck/Germany. Paper presented at the Thirty-Third Workshop on
Geothermal Reservoir Engineering, Stanford University.
Roland N. Horne is the Thomas Davies Barrow Professor of Earth
Sciences at Stanford University, and was the Chair of the Department of Petroleum Engineering from 1995 to 2006. He holds BE,
PhD, and DSc degrees from the University of Auckland, New
Zealand, all in engineering science. Horne has been an SPE Distinguished Lecturer, and has been awarded the SPE Distinguished Achievement Award for Petroleum Engineering Faculty,
the Lester C. Uren Award, and the John Franklin Carl Award. He
is a member of the US National Academy of Engineering and is
also an SPE Honorary Member. Horne is also the 20102013 President of the International Geothermal Association.
Mark McClure is an assistant professor in the Department of
Petroleum and Geosystems Engineering at the University of
Texas at Austin. He holds a PhD degree in energy resources engineering, an MS degree in petroleum engineering, and a BS
degree in chemical engineering, all from Stanford University.
McClure is also an SPE member.

May 2014 SPE Reservoir Evaluation & Engineering


ID: jaganm Time: 16:28 I Path: S:/3B2/REE#/Vol00000/140004/APPFile/SA-REE#140004

243

Anda mungkin juga menyukai