Sunteți pe pagina 1din 19

SPE 152192

Gas Shale Hydraulic Fracturing: A Numerical Evaluation of the Effect of


Geomechanical Parameters
Neal B. Nagel, Marisela Sanchez-Nagel, and Byungtark Lee, Itasca Houston, Inc.

Copyright 2012, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Hydraulic Fracturing Technology Conference held in The Woodlands, Texas, USA, 68 February 2012.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Due to the low permeability of many shale gas reservoirs, multi-stage horizontal well completions are used to provide
sufficient stimulated area to make an economic well. Furthermore, access to, and stimulation of, the natural fracture system is
often critical to an economically successful well.

During a given hydraulic fracture stimulation, the physical displacement of the fracture alters the stress field around it.
Numerous authors have suggested that this altered stress field is beneficial to the stimulation of the natural fracture system;
however, other authors have shown the potential to stabilize the natural fracture system - making it less likely to shear - due to
the presence of a created hydraulic fracture.

In this paper, we present the results of a detailed parametric evaluation of the shear failure (and, by analogy, the
microseismicity) due to the creation of a hydraulic fracture as a function of fracture length within two different fracture
networks (DFNs) using the 2D Distinct Element Model (DEM), UDEC. Simulations were conducted as a function of: 1)
fracture strength; 2) DFN orientation within the stress field; 3) stress ratio (the ratio of the maximum horizontal stress to the
minimum); 4) Poissons ratio of the shale; and 5) Youngs modulus of the shale. The results show the critical impact that
changes in the hydraulic fracture length and the DFN orientation have on the shear of the natural fracture system. In contrast,
the simulations suggest that stress ratio, Poissons ratio, and Youngs modulus have, at best, a second-order effect on the
shearing - and likely the stimulation - of the natural fracture system.

The results of the study provide a further, quantitative assessment of the critical parameters affecting shale gas completions
and aid in the understanding and optimization of hydraulic fracture stimulations in very low permeability, naturally fractured
reservoirs.

Introduction
According to a US Department of Energy-commissioned report (GWPC, 2009), the United States contains 1,744 tcf of
technically recoverable natural gas. Of this, nearly 60% of the onshore recoverable resource (Navigant Consulting, 2008) is in
unconventional plays (shale gas, tights sands, and coalbed natural gas). These statistics, and similar statistics on a global basis
(Rogner, 1997), make it clear that shale gas developments will play a major role in providing clean energy well into the 21st
century.

Shale gas became significant with the development of the Barnett shale in the late 90s (Navigant Consulting, 2008). The
Barnett developments were driven by: 1) the application of horizontal wells; 2) the application and improvements in hydraulic
fracturing; and 3) significant natural gas prices (GWPC, 2009). Hydraulic fracturing is one of the key drivers to shale gas
development because of its low to ultra-low permeability (Frantz and Jochen, 2005 and Harper, 2008). Also key to shale gas
development is the presence of natural fractures (Lancaster et al., 1996 and Chong et al., 2010) and planes of weakness that
can result in complex fracture geometries during stimulations (Palmer et al., 2010). Furthermore, the presence and ability to
open and maintain flow in both primary and secondary natural fracture systems are critical to shale gas production (King,
2010).
2 SPE 152192

If stimulating and maintaining both primary and secondary natural fracture systems are critical to shale gas developments, then
completion engineers need tools to design stimulations that can address both the geomechanics of creating a hydraulic fracture
and the geomechanics of the natural fracture system. However, as King (2010) noted, many commercial fracture simulators,
developed primarily for sandstones, assume bi-planar fracture development under reasonably constant leakoff, and they do not
properly address the influence of shale fracture leakoff into fissures and cracks. As presented by Adachi et al. (2007), pseudo-
3D fracture simulators are not only built upon the basic assumption that the created fracture is bi-planar, but leakoff is often
assumed to be one-dimensional, Carter-type in a direction orthogonal to the fracture plane. These limitations simply make
many of the commercial fracture simulators ill-suited for modeling the complex interaction of a created hydraulic fracture with
a dynamic natural fracture system as seen in many shale gas developments.

As summarized in Nagel et al. (2011), there have been relatively few critical efforts to specifically address the limitations of
the commercial fracture simulators and develop robust and theoretically sound tools for the modeling and design of shale gas
hydraulic fracture stimulations. One of the earliest efforts to recognize the significant limitations of conventional hydraulic
fracturing models for addressing hydraulic fracturing in NFRs was by Hossain et al. (2000), where the authors noted the
importance of the fracture network, the in-situ stress conditions, and fracture dilation.

More recent modeling efforts fall roughly into three camps. Several authors (Rahman, et al., 2009 and Nassir, et al., 2010)
have attempted to model the complex interactions of a created hydraulic fracture with a natural fracture system using
traditional finite element approaches with mixed results. A second camp consists of authors who have developed simplified
models based upon existing pseudo-3D models (Meyer and Bazan, 2011) or have developed simplified rule-based analytical
models for addressing the complex interactions of a created hydraulic fracture and a natural fracture system (Kresse, et al.,
2011). In the third camp are the authors who have developed and used models that have attempted to capture the fundamental
hydro-mechanical behavior of hydraulic fracturing in a naturally fractured reservoir. Damjanac et al., (2010) presented the
results of hydraulic fracture modeling in an naturally fractured reservoir with a discrete element model (DEM) in which they
highlighted the impact of fluid compressibility on hydraulic fracture geometry (a very compressible fluid created a more
complex fracture geometry). Deisman et al., (2009) presented a methodology for an adaptive approach combining continuum
fluid flow modeling with discontinuum modeling (via the Synthetic Rock Mass, SRM, technique) of a naturally fractured
reservoir. Further, Nagel et al., (2011) and Nagel and Sanchez-Nagel (2011) presented specific modeling results of hydraulic
fracture simulations using DEM tools.

Distinct Element Modeling


Hydraulic fracturing in naturally fractured (or weakness planes) shale gas plays involves significant heterogeneity and the need
to develop complex fracture systems both of which significantly exceed the capabilities of conventional hydraulic fracturing
simulators. To-date, there have been few efforts, at a complete, numerical modeling of hydraulic fracturing in shale gas plays,
that honor the proper physics of the process. Admittedly, honoring the physics of the process is neither easy nor trivial;
however, relying upon rules-of-thumb, assumptions, approximations, or guesswork too often results in less than optimal well
stimulations.

The proper physics of naturally fractured reservoirs and shale gas plays involve the solution of force-displacement
relationships (e.g., Hookes law) and strength relationships (e.g., Mohr-Coulomb) for both: (1) the solid, matrix material; and
(2) the fractures or weakness planes. Consequently, any model to simulate the behavior of naturally fractured reservoirs and
shale gas plays must be able to specifically address the behavior of both the solid matrix and the discontinuities (Cundall and
Hart, 1992, Starfield and Cundall, 1988 and Cundall, 1988).

Whereas the behavior of the solid material that constitutes the particles or blocks in a fractured reservoir may be treated as
rigid or deformable, where the deformable body is divided into internal elements over which the differential equations for
stress and displacement are solved, modeling the behavior of the fractures or weakness planes is non-trivial. In order to model
fractures or weakness planes, the model must explicitly recognize the existence of contacts or interfaces between the discrete
bodies that comprise the rock system and, ideally, allow for finite displacements and rotations of discrete bodies, including
complete detachment, and recognize new contacts automatically as the calculation progresses. A numerical approach that
achieves these goals is the discrete element method (DEM), of which a subset is the distinct element method, which has
progressed for a period of nearly 40 years beginning with the initial presentation by Cundall (1971).

Numerical Modeling of Hydraulic Fracturing in Naturally Fractured Reservoirs


In this paper, a series of simulations were conducted with the two-dimensional distinct element model UDEC. UDEC was used
to evaluate the influence of: (1) the mechanical properties of the rock (Youngs modulus and Poissons ratio); (2) the ratio of
the maximum total horizontal principal stress (SHmax) to the minimum total horizontal principal stress (Shmin) the
stress ratio; (3) the friction angle of the natural fractures; (4) the DFN orientation within the stress field; and (5) the effect of
hydraulic fracture half-length.
SPE 152192 3

UDEC Model Capabilities


UDEC (Univeral Distinct Element Code) is a two-dimensional numerical program based on the distinct element method for
discontinuum modeling (Itasca, 2011a). UDEC simulates the response of discontinuous media (such as naturally fractured
reservoirs and shale gas plays) subjected to either static or dynamic loading. The discontinuous medium is represented as an
assemblage of discrete blocks. The discontinuities are treated as boundary conditions between blocks, and large displacements
along discontinuities and rotations of blocks are allowed. Individual blocks behave as either rigid or deformable material.
Deformable blocks are subdivided into a mesh of finite-difference elements, and each element responds according to a
prescribed linear or nonlinear stress-strain law. The relative motion of the discontinuities is also governed by linear or
nonlinear force-displacement relations for movement in both the normal and shear directions. UDEC has several built-in
material behavior models, for both the intact blocks and the discontinuities, which permit the simulation of response
representative of discontinuous geologic, or similar, materials. UDEC is based on a Lagrangian calculation scheme that is
well-suited to model the large movements and deformations of a blocky system.

UDEC DEM Simulations of Hydraulic Fracturing


A series of 13 simulation cases, with 11 simulation phases each, was conducted in order to evaluate the influence of the
various parameters on the natural fracture shearing caused by a growing single, hydraulic fracture. The 13 simulation cases
evaluated two stress ratios (5% and 15%, where stress ratio is defined here as the increase in SHmax over the Shmin value),
four values of Youngs modulus (equal to 27.6 GPa/4 MMpsi, 41.4 GPa/6 MMpsi, 68.9 GPa/10 MMpsi, and 6.9 GPa/1
MMpsi), three values of Poissons ratio (equal to 0.10, 0.25, and 0.40), two DFN orientations (180, which was orthogonal to
the created hydraulic fracture, and 135, which was oriented at 45 to the created hydraulic fracture), and three fracture friction
angles (5, 15, and 30).

Within each of the 13 simulation cases, a solution was found for stress equilibrium and then for ten lengths of a simulated,
single hydraulic fracture from 30m to 300m in half-length. Table 1 shows the simulation matrix for the simulations run.

Table 1 Simulation Matrix for UDEC DEM Evaluations

Simulation Case YM (GPa) PR Stress Ratio DFN Angle() Friction Angle()


Case A 27.6 0.25 5% 180 5
Case B 27.6 0.25 5% 135 15
Case C 27.6 0.25 5% 180 15
Case D 27.6 0.25 15% 135 30
Case E 27.6 0.25 15% 180 15
Case F 41.4 0.10 5% 135 15
Case G 41.4 0.10 5% 180 5
Case H 41.4 0.40 5% 180 15
Case I 41.4 0.40 5% 135 15
Case J 68.9 0.25 5% 135 15
Case K 6.9 0.25 5% 135 15
Case L 68.9 0.25 5% 180 5
Case M 6.9 0.25 5% 180 5

DEM Model Setup, Fracture Network, and Boundary Conditions


The base, two-dimensional model was constructed along the horizontal centerline of the hydraulic fracture and was 600m x
600m (600m in the x-direction, SHmax-direction along the length of the hydraulic fracture and 600m behind the fracture, the
Shmin-direction, representing the spacing between fractures). The two fracture networks used for the simulations were created
from the UDEC, built-in fracture generator.

Each network consisted of two fracture sets. In the 180 network (per Table 1), the first set had a strike of 180 with a standard
deviation, std, of 2; a trace length of 75m and std of 20m; a gap length between fracture traces of 5m; a spacing of 20m and
std of 5m; and an origin in the model at (300m,100m). The second set had a strike of 90 and std of 2; a trace length of 75m
and std of 20m; a gap length between fracture traces of 5m; a spacing of 20m and std of 5m; and an origin in the model at
(300m,100m). Figure 1 shows the 180, 600m x 600m model with the natural fracture system in place. In the 135 network
(per Table 1), the first fracture set had a strike of 135 with a standard deviation, std, of 2; a trace length of 75m and std of
20m; a gap length between fracture traces of 5m; a spacing of 20m and std of 5m; and an origin in the model at (300m,100m).
The second set had a strike of 45 and std of 2; a trace length of 75m and std of 20m; a gap length between fracture traces of
5m; a spacing of 20m and std of 5m; and an origin in the model at (300m,100m). Figure 2 shows the 135, 600m x 600m
model with the natural fracture system in place.
4 SPE 152192

Figure 1: 180 UDEC fracture network model.

Figure 2: 135 UDEC fracture network model.

Fracture normal and shear stiffness were set at 20 GPa, fracture cohesion was set at 0 MPa and the initial fracture friction
angle was set at 5 for each of the two models. The models were initialized using the appropriate mechanical properties (see
Table 1) and stresses. In each case, the vertical stress was initialized to 90.5 MPa/13125 psi, pore pressure was initialized to
77.91 MPa/11300 psi (in both the matrix and within the fracture system) and Shmin was initialized to 86.2 MPa/12500 psi. For
the 5% stress ratio cases, SHmax was initialized to 90.5 MPa/13125 psi. For the 15% stress ratio cases, SHmax was initialized
to 99.1 MPa/14380 psi. Once the models were initialized, fixed boundary conditions (roller boundaries) were applied and the
models were stepped to equilibrium prior to the application of pressure associated with the hydraulic fractures.

As shown in Figure 3, the fracture was simulated by applying a pressure along the lower left, bottom edge of the model equal
to Shmin plus the fracture net pressure. For all the cases presented in Table 1, the net pressure was equal to 4.3MPa/625 psi.
SPE 152192 5

The initial fracture was 30m long (half-length). For the subsequent simulations for each of the cases in Table 1, the fracture
half-length was increased by 30m until a total half-length of 300m was reached.

Figure 3: Example UDEC simulation result showing the hydraulic fracture location and
current shear in the natural fracture system.

DEM Model Simulation Results


Figures 4 through 8 show the simulation results for each of the 10 fracture half-lengths for Case A. Note that the red portion
of the fracture system represents the fractures that are currently in shear at that particular fracture half-length. As the hydraulic
fracture growths in length, new fractures reach shear conditions (and, thus, may emit microseismicity) and previous shear
regions become stable and are no longer in shear. As shown by Nagel and Sanchez-Nagel (2011), this occurs because
formation shear occurs just behind the tip of a growing hydraulic fracture and the region far behind the tip is stabilized due to
the increase in the Shmin stress.

Figure 4: Natural fracture shear region for Case 'A' half-lengths of 30m and 60m.
6 SPE 152192

Figure 5: Natural fracture shear region for Case 'A' half-lengths of 90m and 120m.

Figure 6: Natural fracture shear region for Case 'A' half-lengths of 150m and 180m.
SPE 152192 7

Figure 7: Natural fracture shear region for Case 'A' half-lengths of 210m and 240m.

Figure 8: Natural fracture shear region for Case 'A' half-lengths of 270m and 300m.

A number of significant observations can be made from the Case A simulations alone:

The shear region within the natural fracture system is small and modest at a 30m fracture half-length, but it reaches
nearly 300m behind the fracture (that is, into the formation perpendicular to the vertical hydraulic fracture plane) by a
120m fracture half-length. By a 300m half-length, shear can be seen to nearly 500m into the formation behind the
hydraulic fracture.
Beginning at a fracture half-length of 150m and very evident at a half-length of 180m natural fracture shear is
occurring that is clearly disconnected from the main hydraulic fracture. It is possible that these shear regions, and
their associated microseismicity, are not in hydraulic communication with the wellbore.
8 SPE 152192

From a half-length of 240m to 300m, the shear region is significant in size. The region is 100 to 150m wide (parallel
to the hydraulic fracture) and more than 400m into the formation.
In many of the simulations, there are random shear areas (again, possible microseismicity), which occur mostly at
fracture intersections, that are several hundreds of meters behind the hydraulic fracture tip, into the formation, and
even ahead of the fracture tip.

Clearly the observations in Figures 4 through 8 show massive shear regions (and massive amounts of microseismicity) that are
not often seen in field cases. One important reason for this, and a reason that is commonly overlooked, is that shear occurs due
to a change in the initiating force (from the net pressure of the hydraulic fracture) relative to the resistant force (the strength of
the rock or fracture). If we assume that fracture strength can be represented by the Mohr-Coulomb failure criterion:

(1)

where: = shear strength of the fracture;


So = cohesion of fracture;
= coefficient of friction along the fracture; and
= normal stress acting on the fracture.

Shear occurs whenever the shear stress exceeds the shear strength of the fracture. This can occur if the normal stress, , acting
on the fracture plane is reduced (possibly by changes in pressure within the fracture system) or if the coefficient of friction is
low (or reduced by some means). Figure 9 shows the influence of increasing the coefficient of friction (which is the tangent of
the friction angle).

Figure 9: Comparison of natural fracture shear for fracture half-length of 300m for (left) 5
friction angle and (right) 15 friction angle.

Because the fracture half-length in both plots in Figure 9 is 300m, and the fracture net pressure was the same in each, the
change in stress in the two simulations is the same. Consequently, the sole cause of the difference in shear between the two
simulations is the increase in the friction angle for the Case C simulation. This result highlights, for example, that even if the
stimulation treatment were identical between two wells, and even if the natural fracture system was exactly the same, if the
strength properties of the natural fracture system (either cohesion or friction angle) change, then the formation shear (and
microseismicity) will differ as well as the effectiveness of the stimulation.

Stress ratio (where stress ratio is defined here as the increase in SHmax over the Shmin value) also plays a role in the
generation of natural fracture shear. Figure 10 shows a comparison between a 5% and a 15% stress ratio (for the cases with
fracture friction angle of 15).
SPE 152192 9

Figure 10: Comparison of natural fracture shear for fracture half-length of 300m for (left)
5% stress ratio and (right) a 15% stress ratio.

The influence of stress ratio is to focus the natural fracture shear along horizontal fractures (i.e., those parallel to the created
hydraulic fracture) potentially leading to additional sheared fractures that are disconnected from the wellbore and not in
hydraulic communication. This is more readily seen in Figure 11 (from Nagel and Sanchez-Nagel, 2011).

Figure 11: Natural fracture shear regions with a high stress ratio. Note that shear occurs
dominantly along horizontal fractures.

Natural fracture orientation within the stress field also has a consequential influence on the shear of the fractures during a
stimulation. Figures 12 through 16 show the shear results for each of the 10 fracture half-lengths for Case B.
10 SPE 152192

Figure 12: Natural fracture shear region for Case 'B' half-lengths of 30m and 60m.

Figure 13: Natural fracture shear region for Case 'B' half-lengths of 90m and 120m.
SPE 152192 11

Figure 14: Natural fracture shear region for Case 'B' half-lengths of 150m and 180m.

Figure 15: Natural fracture shear region for Case 'B' half-lengths of 210m and 240m.
12 SPE 152192

Figure 16: Natural fracture shear region for Case 'B' half-lengths of 270m and 300m.

As for the Case A simulations, a number of significant observations can be made from the Case B simulations:

Most critically, the Case B simulations were conducted at a fracture friction angle of 15. This was done because the
model was not stable at a 5 friction angle. That is, even at a stress ratio of only 5%, the stress conditions along the
fracture system resulted in unstable fracture shear suggesting that the combination of in-situ stress and fracture
orientation creates a lower-limit on the possible fracture friction angle (otherwise the region would be seismically-
active due to on-going natural fracture shear).
Like the Case A simulations, the shear region within the natural fracture system is small and modest at a 30m
fracture half-length. Unlike the Case A simulations, the Case B simulations showed shear behind the fracture (that
is, into the formation perpendicular to the vertical hydraulic fracture plane) to approximately 150m for a 120m
fracture half-length and only to 200-220m for a 300m half-length.
Unlike the Case A results, the Case B results show that the shear region is ahead of the fracture tip and little, if
any, shear occurs behind the tip. This would suggest, for example, that a comparison of microseismic data relative to
the known (or expected) hydraulic fracture tip could provide some indication of the natural fracture system
orientation.
There appears to be less disconnected shear regions in the Case B results than in the Case A results.
The shear region is still significant in size, but smaller than in the Case A simulations.
The random shear areas (again, possible microseismicity), which occur mostly at fracture intersections, are focused
well ahead of the fracture tip and few occur behind the tip.

As with the Case A simulations (Figure 9), the shear region for the Case B simulations is also sensitive to the fracture
friction angle. Figure 17 shows a comparison of the shear region for 15 and 30 fracture friction angles for a 300m fracture
half-length.

The influence of formation mechanical properties was also evaluated. In particular, the influence of Youngs modulus and
Poissons ratio were simulated. Figures 18 and 19 show the change in shear region, at 150m and 300m fracture half-lengths,
respectively, for simulations with Youngs modulus set at 41.4 GPa /6 MMpsi and Poissons ratio values of 0.1 and 0.4 for the
180 fracture network. Figures 20 and 21 show similar results for the 135 fracture network.
SPE 152192 13

Figure 17: Comparison of natural fracture shear for fracture half-length of 300m for (left)
15 friction angle and (right) 30 friction angle.

Figure 18: Comparison of natural fracture shear for fracture half-length of 150m and 180
fracture network for (left) Poisson's ratio equal to 0.10 and (right) equal to 0.40.
14 SPE 152192

Figure 19: Comparison of natural fracture shear for fracture half-length of 300m and 180
fracture network for (left) Poisson's ratio equal to 0.10 and (right) equal to 0.40.

Figure 20: Comparison of natural fracture shear for fracture half-length of 150m and 135
fracture network for (left) Poisson's ratio equal to 0.10 and (right) equal to 0.40.
SPE 152192 15

Figure 21: Comparison of natural fracture shear for fracture half-length of 300m and 135
fracture network for (left) Poisson's ratio equal to 0.10 and (right) equal to 0.40.

The effect of changing Poissons ratio, as shown in Figures 18 through 21, can be summarized as:

For the 135 fracture network (Figures 20 and 21), there is no discernible influence of Poissons ratio on fracture
shear regardless of hydraulic fracture half-length.
For the 180 fracture network with a hydraulic fracture half-length of 300m (Figure 19), the lower Poissons ratio
simulation caused only a slight increase in the amount of natural fracture shear.
For the 180 fracture network with a hydraulic fracture half-length of 150m (Figure 18), the greater Poissons ratio
simulation had considerably more natural fracture shear.

The reason for the apparently inconsistent, yet minor, influence of Poissons ratio is not clear. The results are not believed to
be simulation artifacts but rather suggest that under some limited conditions (likely defined by the character of the natural
fracture network), Poissons ratio may influence the shear region (and associated microseismicity) caused by a growing
hydraulic fracture.

Figures 22 and 23 show the change in shear region, at 150m and 300m fracture half-lengths, respectively, for simulations with
Poissons ratio set at 0.25 and Youngs modulus varying between 68.9 GPa /10 MMpsi and 6.9 GPa /1 MMpsi for the 180
fracture network. Figures 24 and 25 show similar results for the 135 fracture network.
16 SPE 152192

Figure 22: Comparison of natural fracture shear for fracture half-length of 150m and 180
fracture network for (left) Young's modulus equal to 68.9 GPa and (right) equal to 6.9 GPa.

Figure 23: Comparison of natural fracture shear for fracture half-length of 300m and 180
fracture network for (left) Young's modulus equal to 68.9 GPa and (right) equal to 6.9 GPa.
SPE 152192 17

Figure 24: Comparison of natural fracture shear for fracture half-length of 150m and 135
fracture network for (left) Young's modulus equal to 68.9 GPa and (right) equal to 6.9 GPa.

Figure 25: Comparison of natural fracture shear for fracture half-length of 300m and 135
fracture network for (left) Young's modulus equal to 68.9 GPa and (right) equal to 6.9 GPa.

The effect of changing Youngs modulus, as shown in Figures 22 through 25, can be summarized as:

For the 135 fracture network (Figures 24 and 25), there is no discernible influence of Youngs modulus on the main
region of fracture shear regardless of hydraulic fracture half-length; however, there is significantly greater far-field,
random, disconnected shear for the simulations with a lower Youngs modulus. This is unlikely to influence the
production of the well, but may reveal itself as a more dispersed microseismic cloud.
For the 180 fracture network with a hydraulic fracture half-length of 300m (Figure 23), the lower Youngs modulus
simulation showed only a slight increase in the amount of natural fracture shear; however, shear was evident well
behind the fracture tip that was not present in other simulations.
18 SPE 152192

For the 180 fracture network with a hydraulic fracture half-length of 150m (Figure 22), the lower Youngs modulus
simulation had considerably greater natural fracture shear.

Again, the reason for the apparently inconsistent, yet minor, influence of Youngs modulus is not clear. The results are not
believed to be simulation artifacts but rather suggest that under limited conditions (likely defined by the character of the
natural fracture network), Youngs modulus may influence the shear region (and associated microseismicity) caused by a
growing hydraulic fracture.

Conclusions

1. As expected from geomechanical principles, the degree of fracture shear (and MS events) is directly linked to fracture
friction angle. Fracture friction angle is a critical parameter for relating field microseismicity back to stimulated
volume and well production.
2. The 135 fracture network simulations required a minimum fracture friction angle of 15 to achieve stability. The
180 fracture network simulations were stable at 5. This highlights the importance of understanding the orientation
of the fracture network within the in-situ stress field.
3. For the simulations performed, stress ratio (5% vs. 15%) did not cause a significant change in the amount of shear;
however, the shear that did occur suggested more disconnected shear (and associated microseismicity) from the
wellbore.
4. The 135 fracture network showed no influence of Poissons ratio on the amount of fracture shear.
5. For the 180 fracture network, a higher Poissons ratio caused significantly more shear at 150m, but slightly less at
300m.
6. Overall, the influence of Poissons ratio on natural fracture shear was minimal and, likely, inconsequential except in
isolated cases.
7. A lower formation stiffness (Youngs modulus) caused more far-field shear (but only modestly) but little additional
main shear for the 135 fracture network.
8. A lower stiffness caused significantly more shear for the 180 fracture network at 150m but only modestly more at
300m. However, at 300m, shear was seen well behind the fracture tip for the lower stiffness case.
9. Overall, the influence of Youngs modulus on natural fracture shear was only modest and, in some cases, essentially
inconsequential.

References

Adachi, J., Siebrits, E., Peirce, A. and Desroches, J., 2006, Computer simulation of hydraulic fractures, Int. J. Rock Mech.
Min. Sci. and Geomech. Abstr., 44, p739-757.

Chong, K.K., Grieser, W.V., Passman, A., Tamayo, C.H., Modeland, N., and Burke, B., 2010, A completions Guide Book to
Shale-Play Development: A Review of Successful Approaches Towards Shale-Play Stimulation in the Last Two Decades,
Paper CSUG/SPE 133874 presented at the Canadian Unconventional Resources & International Petroleum Conference,
Calgary, Alberta, Canada, 19-21 October.

Cundall, P. A., 1971, A Computer Model for Simulating Progressive Large Scale Movements in Blocky Rock Systems, in
Proceedings of the Symposium of the International Society for Rock Mechanics (Nancy, France, 1971), Vol. 1, Paper No. II-8.

Cundall, P. A., 1988, Formulation of a Three-Dimensional Distinct Element ModelPart I: A Scheme to Detect and
Represent Contacts in a System Composed of Many Polyhedral Blocks, Int. J. Rock Mech., Min. Sci. & Geomech. Abstr., 25,
107-116.

Cundall, P. A., and Hart, R. D., 1992, Numerical Modeling of Discontinua, Engr. Comp., 9(2), 101-113.

Damjanac, B., Gil, I., Pierce, M., and Sanchez, M., 2010, A New Approach to Hydraulic Fracturing Modeling in Naturally
Fractured Reservoirs, Paper ARMA 10-400 presented at the 44th US Rock Mechanics Symposium, Salt Lake City, Utah,
USA, 27-30 June.

Deisman, N., and Chalaturnyk, R. J., 2009, An Adaptive Continuum/Discontinuum Coupled Reservoir Geomechanics
Simulation Approach for Fractured Reservoirs, Paper SPE 119254 presented at the SPE Reservoir Simulation Symposium,
The Woodlands, Texas, USA, 2-4 February.

Frantz, J.K. and Jochen, V., 2005, Shale Gas White Paper, Schlumberger, 05OF299.
SPE 152192 19

Ground Water Protection Council and ALL Consulting, 2009, Modern Shale Gas Development in the United States: A
Primer, prepared for the US DOE, Office of Fossil Energy, DE-FG26-04NT15444.

Harper, J., 2008, The Marcellus Shale An Old New Gas Reservoir in Pennsylvania, V28, N1, Bureau of Topographic
and Geologic Survey, Pennsylvania Department of Conservation and Natural Resources, Pennsylvania Geology.

Hossain, M.M., Rahman, M.K., and Rahman, S.S., 2000, Volumetric Growth and Hydraulic Conductivity of Naturally
Fractured Reservoirs During Hydraulic Fracturing: A Case Study Using Australian Conditions, Paper SPE 63173 presented at
the SPE Annual Technical Conference and Exhibition, Dallas, Texas, USA, 1-4 October.

Itasca Consulting Group, Inc. UDEC (Universal Distinct Element Code), Version 5.0. Minneapolis: ICG, 2011.

King, G.E., 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, 19-22 September.

Kresse, O., Cohen, C., Weng, X., Wu, R., and Gu, H., 2011, Numerical Modeling of Hydraulic Fracturing in Naturally
Fractured Formations, Paper ARMA 11-363 presented at the 45th US Rock Mechanics Symposium, San Francisco,
California, USA, 26-29 June.

Lancaster, D.E., Cain, B.A., Hill, D.G., Jochen, V.A., and Parks, W.M., eds, 1996, Production from Fractured Shales,
Reprint Series, 45, Society of Petroleum Engineers, Richardson, Texas, USA.

Meyer, Bruce R., and Bazan, Lucas W., 2011, A Discrete Fracture Network Model for Hydraulically Induced Fractures
Theory, Parametric and Case Studies, Paper SPE 140514 presented at the SPE Hydraulic Fracturing Technology Conference
and Exhibition, The Woodlands, Texas, USA, 24-26 January.

Nagel, N., and Sanchez-Nagel, M., 2011, Stress Shadowing and Microseismic Events: A Numerical Evaluation, Paper SPE
147363 presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, USA, 30 October-2 November.

Nagel, N., Gil, I., Sanchez-Nagel, M., and Damjanac, B., 2011, Simulating Hydraulic Fracturing in Real Fractured Rock
Overcoming the Limits of Pseudo3D Models, Paper SPE 140480 presented at the SPE Hydraulic Fracturing Technology
Conference and Exhibition, The Woodlands, Texas, USA, 24-26 January.

Nassir, M., Settari, A., and Wan, R., 2010, Modeling Shear Dominated Hydraulic Fracturing as a Coupled Fluid-solid
Interaction, Paper SPE 131736 presented at the International Oil and Gas Conference and Exhibition, Beijing, China, 8-10
June.

Navigant Consulting, 2008, North American Natural Gas Supply Assessment, prepared for American Clean Skies
Foundation.

Palmer, I., and Moschovidis, Z., 2010, New Method to Diagnose and Improve Shale Gas Completions, Paper SPE 134669
presented at the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19-22 September.

Rahman, M. M., Aghighi, Ali, and Rahman, Sheik S., 2009, Interaction Between Induced Hydraulic Fracture and Pre-
Existing Natural Fracture in a Poro-Elastic Environment: Effect of Pore Pressure Change and the Orientation of a Natural
Fractures, Paper SPE 122574 presented at the Asia Pacific Oil and Gas Conference & Exhibition, Jakarta, Indonesia, 4-6
August.

Rogner, H. H., 1997, An Assessment of World Hydrocarbon Resources, Annual Review of Energy and Environment.

Starfield, A. M., and Cundall, P. A., 1988, Towards a Methodology for Rock Mechanics Modelling, Int. J. Rock Mech. Min.
Sci. & Geomech. Abstr., 25, 99-106.

S-ar putea să vă placă și