Sie sind auf Seite 1von 15

SPE-184854-MS

Fully Coupled 3-D Hydraulic Fracture Growth in the Presence of Weak


Horizontal Interfaces

Ghazal Izadi, Daniel Moos, Leonardo Cruz, Michael Gaither, and Laura Chiaramonte, Baker Hughes; Scott
Johnson, GeoNumerical Solutions

Copyright 2017, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Hydraulic Fracturing Technology Conference and Exhibition held in The Woodlands, Texas, USA, 24-26 January
2017.

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
Weak bedding planes create a unique mechanism for hydraulic fracture height containment, providing
one possible explanation for unusual patterns of height growth in shale formations. This paper describes
an investigation into how bedding planes modify the interactions between multiple, simultaneously
propagating hydraulic fractures in a formation with weak horizontal interfaces with laterally varying
properties. The investigation used a 3-D simulator that fully coupled geomechanics, fracture mechanics,
and fluid behavior. Three equally spaced fractures were simulated along a horizontal trajectory. Fluid was
injected simultaneously into all three locations, and partitioned according to maintain a specified total
injection rate. Variations in perforation spacing, fluid viscosity and injection rate are modeled. The four
designs investigated were: 1) 10-cp fluid, 20b-pm injection rate with 30-m cluster spacing. 2) 100-cp fluid
and 20-bpm fluid injection rate with 30-m cluster spacing. 3) 10-cp fluid and 40-bpm injection rate with
30-m cluster spacing. 4) 10-cp fluid, 20-bpm and 45-m cluster spacing. Results showed how these changes
affected fracture area and shape. The propped surface area for each scenario was also estimated. The results
suggested that the presence of laterally varying weak interfaces can significantly affect fracture interference.

Introduction
For several decades the process of hydraulic fracturing has been widely used for the stimulation of low-
permeability unconventional source rocks such as shales. Many uncertainties still exist about hydraulic
fracture propagation during stimulation, including interaction between multiple fractures (Delaney and
Pollard 1981, Germanovich and Astakhov, 2004), effects of fluid rheology (Huang and Desroches, 2004;
Garagash, 2006), and proppant transport (Hammond, 1995; Economides and Nolte, 2000). Hydraulic
fractures interact with joints (Lacazette and Engelder, 1992), veins (Srivastava and Engelder, 1991; Al-
Aasm et al., 1995) and weakly cohesive frictional discontinuity (Chuprakov et al., 2013). Because of
material property heterogeneities, rock structure and in-situ stress state, the hydraulic fracturing process is
highly complex (Germanovich et al., 1997). As a result, it is difficult to measure and predict the behavior
of hydraulic fractures in field conditions. Modeling of hydraulic fracture propagation is important to the
treatment design and the ability to evaluate post-treatment production. Several theoretical, numerical, and
2 SPE-184854-MS

experimental studies have focused on this task (Daneshy 1974, Blanton 1982, Zhang 2007, Chuprakov et
al. 2011, Olson 2004, Meyer and Bazan 2011, Bunger 2013, Pierce and Bunger 2014). These modeling
approaches captured the essential physical mechanisms that determine the impacts on hydraulic fracture
growth of varying engineering design decisions such as fracture spacing, entry characteristics, and pumping
schedule. However, most of these models have been used to study the growth of fractures only in 1-D
layered media with no lateral variations in properties or stresses. Recent investigations demonstrate the
effect of weak bedding planes on fracture height growth of single fractures, which, as for example shown
by Gu et al. 2008, also affect fracture width and fracture pressure. Gu et al. did not investigate the effects of
interference between multiple fractures in the presence of weak interfaces. This paper presents a numerical
study of fracture interference in the presence of lateral variations in reservoir properties, specifically, lateral
variations in the strength of weak bedding planes.

Modeling Methodology
A fully coupled 3D hydraulic fracturing simulator to carry out the analyses is presented here. Finite element
method (FEM) is used to calculate the rock solid's stress and deformation under external loading and fracture
fluid pressure, and a Finite Volume method (FV) to compute fluid flow and proppant transport. A stress
intensity factor-based criterion is employed to handle the propagation of fractures based on linear elastic
fracture mechanics (LEFM). The modeling approach is detailed in Fu et al. 2013 and Settgast et al. 2014.
Fractures propagate along a mesh consisting of the faces of the FEM solid elements. Although this is often
considered problematic because the fracture path is largely predefined, there is only a minor amount of
fracture curvature due to stress changes resulting from fracture opening. This is because the net pressure
never exceeds 400 psi even for the highest viscosity fluid, and is otherwise less than 150 psi (see below). A
large horizontal stress contrast exsist (more than 10 MPa or 1450 psi, see Figure 1 and below). Consequently,
the constraint that specifies fractures must propagate along mesh faces has minimal impact on the results.

Figure 1Stress profile, poisson's ratio, and young modulus vs. depth. The stress magnitudes shown here are
reduced by subtracting the first stress invariant. The rock properties are anisotropic, as discussed in the text.
SPE-184854-MS 3

The vertical distribution of the three in-situ principal stress components, namely, the minimum horizontal
stress (Shmin), the maximum horizontal stress (Shmax), and the vertical stress (Sv), are shown in Figure
1. The actual stresses in the modeled reservoir are much higher because the fluidand solid properties are
essentially independent of pressure or stress. The results are presented in terms of the net pressure which
is the same regardless of the magnitude of the first stress invariant. The computation was accelerated by
using the reduced stresses shown in Figure 1. A strike-slip stress state is modeled here, as Shmin < Sv <
SHmax. Consequently, hydraulic fractures are expected to be vertical and to propagate perpendicular to the
Shmin direction.
Dipole acoustic logs recorded in local wells indicate a moderate amount of stiffness anisotropy. Therefore
a nine-parameter (i.e., three Young's moduli (E), three Poisson's ratios (v) and three shear moduli) generic
orthotropic elasticity model was used in which Ex/Ey=1 and vxz/vyz=1 (Ez/Ex=0.9, vxz/vxy=0.95).
Horizontal frictional interfaces are distributed at three different depths above below and within the
fracturing zone. The depths of these interfaces are shown in Figure 1, and the lateral distributions are shown
in Figure 2.

Figure 2Figure showing the reservoir model, and the locations of the three initiation points of competing hydraulic
fractures spaced 30 meters apart along a horizontal trajectory which is perpendicular to the least stress. Also shown
are the locations of weak patches (color indicates a weak zone; absence of color indicates higher strength) on four
bedding planes located above and below the perforations. The depths of these zones are indicated in Figure 1.

In this model fractures originate at three cluster locations and propagate simultaneously, driven by a
constant injection rate. The complexities of fracture initiation are ignored, and it is assumed that the fractures
all reach the same finite size at the same time and initiate as equally sized planes perpendicular to the least
stress, as indicated in Figure 2. The locations in the base case are as shown in this figure, for which the
distance between clusters is 30 meters. Weak horizontal bedding planes are at fixed depths above and below
the fracturing zone. Their strength is randomly distributed laterally. Because the hydraulic fractures are
4 SPE-184854-MS

vertical and bedding is horizontal, the angle between the hydraulic fracture plane and the weak bedding
planes is always 90 degrees.
The first investigation included the final geometry of hydraulic fractures in a case without horizontal
weak bedding planes. This case is one in which 10-cP fluid is injected at a constant injection rate of 20
bbls per minute (bpm); fractures initiate 30 meters apart at the same depth along a trajectory perpendicular
to the least stress.
Figure 3 shows the fracture geometries for the base case after 30 minutes of injection. The fractures
initiate as radial fractures; upward growth approximately preserves the radial geometry. However, a small
stress barrier, at the depth indicated in Figure 1, truncates downward growth. The exterior fractures are
identical, and are longer and taller than the interior fracture, which also is much narrower. The pressure
profile showing higher pressure towards the base of the fractures reveals the effect of the hydrostatic pressure
variation due to gravity. Note that downward growth is truncated against a relatively small stress barrier;
there is no indication of truncation in upwards growth at this stage in the injection.

Figure 3Simultaneous growth of multiple interacting hydraulic fractures without weak bedding
planes. Left) Fluid pressure distribution. Right) Fracture width distribution after 30mins pumping.

Simultaneous propagation of multiple hydraulic fractures in the existence of


weak bedding planes
The first case investigated to study the effects of the weak bedding plans shown in Figure 2 is otherwise
identical to the case shown in Figure 3. This "base case" uses a 10-cp fluid and a 20-bpm injection rate. For
comparison, a high-viscosity case of using 100-cP fluid, and a high-rate case using 40-bpm rather than 20-
bpm was considered. For these three cases the fracture locations are identical and the spacing is 30 meters.
A fourth case is discussed later in which the outer locations are translated along the well such that they
are 40 meters from the central fracture, which is always at the same location. During the discussion that
follows, the leftmost fracture and the rightmost fracture were defined as Fracture 1 and Fracture 3. The
middle fracture is Fracture 2.
Fracture geometries and fluid pressure distributions for the three cases with 30-meter fracture spacing
are shown in Figure 4. The models are compared at the same pumped volume, which means for the base
case and the high-viscosity case (10 times greater than the base case) the pumping time is 30 minutes (1,800
sec.), and in the high flow rate case (twice as high as the base case) the pumping time is 15 minutes (900
sec.). Figure 4-A, B and C show the pressure in the three fractures for the base, high-flow rate and high-
SPE-184854-MS 5

viscosity cases, respectively. Sub plots a, b and c show the pressure distributions of fractures 1, 2 and 3
at the indicated times.

Figure 4Final fracture geometries and pressure distribution after 30 minutes injection. A) 10-cp, 20-bpm- B)
10-cp, 40-bpm- C) 100-cp, 20-bpm. Subplots a, b, and c show the pressure distribution of fractures 1, 2 and 3.

In all cases the fractures are highly asymmetric. Fracture 3 also grows farther from the midpoint than
Fracture 1, for example, has less height on the back side and grows further downward on the front side. In
fact, Figure 4 shows by comparison to Figure 3 in which no weak bedding planes are present, that there is
downward growth through the lower stress barrier in all three cases along front limb of Fracture 3. Fracture
1 has a more "radial" upward growth than Fracture 3, but it also penetrates the lower stress barrier. Details
of the fracture shape predicted for the higher viscosity model are distinctly different than for the lower-
viscosity models. For example, there is considerably more asymmetry in the fracture shapes (Fractures 1
and 3 are much more different). Even though the outer fractures are similar when varying the flow rate, the
higher flow rate model predicts a slightly larger interior fracture than the lower rate model.
6 SPE-184854-MS

Figure 5 shows net pressure vs. time at the inlet of Fracture 3 for the base, high-rate and high-viscosity
cases. As discussed above, the net pressure is a fraction of the differential horizontal stress, even in the
high-viscosity case for which it is more than twice as high as for the high-rate case. The low-rate case has
slightly lower inlet pressure than the high-rate case. Pumping time to achieve the same total volume is half
as long for the high-rate case, hence the line showing net pressure for the high-rate case ends at 900 seconds
vs. 1800 seconds for the other cases.

Figure 5Net pressure vs. time at inlet. Green, blue and red curves represent base, high-
rate (twice as high as the base case) and high-viscosity (10 times greater than the base case).

Figure 6 shows the fracture width at the end of the injection period for comparison with Figure 5, which
shows internal fluid pressure. The columns (A, B, and C) show the base, high-flow rate, and high-viscosity
results, respectively; the characteristics of the individual fractures are shown beneath the corresponding
plots. The shapes are identical to those shown in Figure 4, however the widths are functions not only of the
net pressure but also of the least horizontal stress, which varies with depth and between individual layers
in the model. The widest point in each fracture is at the injection point, however width also is somewhat
asymmetric. To illustrate the impact of the specific distribution of bedding plane strength on each fracture,
we also show in Figure 6-a, b, c the weak points (in dark blue) along the laminations above and below
each fracture initiation point where vertical growth might be truncated by bedding. Not surprisingly, where
bedding is strong (gaps in the dark blue colors) the fracture can grow upwards or downwards. It is interesting
to note that once past a lamination, the fracture will propagate around it and begin to fill in the gap introduced
by the barrier. This impacts the flow pattern discussed below. It also highlights the importance of using
a fully 3D model that allows such bypass to happen (see also Fu et al. 2015, for a discussion of fracture
bypass of vertical fractures).
SPE-184854-MS 7

Figure 6Final distribution of fracture width. A) 10-cp, 20-bpm- B) 10-


cp, 40-bpm- C) 100-cp, 20-bpm. Subset a, b, c represent Fracture 1, 2 and 3.

Total surface area is shown below each perspective view in Figure 6. The largest total area was created
by injection of 10-cp fluid at 20-bpm. By increasing the fluid viscosity to 100-cP, aperture increases and
since the injected volume is the same, the total surface area is correspondingly reduced by ~24%. The high-
rate model predicts a surface area slightly less than for the lower rate; this also likely due to the slightly
larger width resulting from the pressure needed to maintain the higher flow rate.
Fracture width at the injection point of each fracture is plotted vs. time in Figure 7. The widths of Fractures
1 and 3 are different due to the difference in how the weak points in the bedding planes interact with
each fracture. This would not be the case if lateral variations were absent and indicates the effect of the
laminations on width. Widths reach a maximum and then decline slowly in every case, however the timing
of the maximum and the amount of decline varies depending on the rate and viscosity. Comparing the net
pressure in Fracture 3 shown in Figure 5 to the widths shown in Figure 7, there is, at most, a weak correlation
of net pressure to width, suggesting that width is also affected by confinement due to the ability of the
fracture to grow uniformly away from the well, and, perhaps, by stress changes induced by the innermost
fracture's opening.
8 SPE-184854-MS

Figure 7Fracture width vs. time at the inlet of each fracture. Green, blue and red curves represent base,
high-rate and high viscosity cases. Square dot, dash dot and solid line represents fracture 1, 2 and 3.

If primary hydrocarbon recovery scales with productive surface area, and if to be productive it is
necessary to place proppant in the fracture, then it would be useful to investigate the propped area of
these fractures. Although this fully coupled 3D hydraulic fracturing simulator provides a means to compute
the distribution of proppant using a simple transport model, a potentially useful but less rigorous way of
investigating propped area is to compare the width during pumping to the nominal width required to allow
proppant to pass through the fracture. A typical rule of thumb is that the fracture must be at least 3 to 5
times wider than the nominal proppant diameter to allow entry (King at al. 2009 and 2010).
Figure 8-A, B, C, D illustrate how much of the fracture surface which was created and shown in Figure 6
could be propped using for illustration a 20/40 mesh size. Using a threshold width equal to 5x the proppant
diameter for an assumed 0.02 inch (0.0005m) proppant diameter, results in a minimum width required to
allow proppant to enter the fracture of 0.1inch (0.0025m). The orange bars show total area; the blue bars
show total propped area, and the ratio is shown as a percentage above each blue bar.
SPE-184854-MS 9

Figure 8A, B, C illustrates the final propped area. Color scale on the fracture represents fracture
width. (Red color-wider width). D) Comparing the total area vs. propped area for base (10-cp, 20-bpm),
high-rate (10-cp, 40-bpm) and high-viscosity (100-cp, 20-bpm). Orange and blue bars show the total
area and propped area in this figure. The ratio of these is shown as a percentage above the blue bars.

It is immediately apparent from the results shown in the perspective plots A, B, C of Figure 8 that none of
the pumping scenarios allow any proppant to enter the innermost fracture it will be ineffective in all three
cases. Furthermore, only a relatively small percentage of the total area can be propped under this somewhat
optimistic set of assumptions in any of these cases.
Comparison of propped area vs. total area shown in Figure 8-D provides an insight into the importance
of fluid properties and fluid injection rate on propped area. The base case (10-cP, 20b-pm) delivered 8000
square meters total surface area which is the largest among all the cases however the propped area is
only 30% of the total area. Increasing viscosity to 100-cp delivers larger propped area (65% of total area)
compared to other cases and proppant placement in the exterior fractures is more uniform (Figure 8-C).
Fracture 3 accepts more proppant in the target zone as its width is larger; this effect is more pronounced
in the lower-viscosity scenarios.
To further evaluate the likely locations where it will be possible to place proppant, the vectors of the fluid
velocity are plotted in Figure 9. These are shown only for Fractures 1 and 3, because Fracture 2 doesn't take
any proppant. The total fracture surface area is illustrated with yellow color. However, the flow vectors are
shown only where the aperture is large enough to accept proppant (i.e., the width is larger than 0.1inch).
10 SPE-184854-MS

Figure 9In this figure we illustrated the total fracture surface area (yellow color) when the fluid velocity vector
(arrows on the fractures) is imposed only on the portion of fracture that takes proppant (width larger than
0.1inch). A, B) 10-cp, 20-bpm- C, D) 10-cp, 40-bpm- E,F) 100-cp,20-bpm. Fracture 1 (A, C, E) and fracture 3 (B, D, F)

Areas with higher fluid velocity will be the most likely paths for proppant transport. These tend to be
somewhat asymmetric, suggesting that proppant placement will also be asymmetric. The asymmetry, at
least in the base case, does not mirror the asymmetry in fracture shape. On the other hand, flow vectors are
somewhat more symmetrical for the high injection rate case. However, the difference between the overall
shape asymmetry and the flow vector asymmetry suggests that it is inappropriate to assume, that fracture
aperture alone is a good proxy for propped area or proppant distribution / propped width. This highlights
once again that lateral variations in reservoir properties can have surprisingly complex effects on all aspects
of fracturing behavior, and, that geometrical asymmetries cannot be predicted solely from those variations;
3D physics-based models must be employed to capture the multiple effects.

Impact of changing cluster spacing and locations


Field observations suggest that the hydraulic fracture growth also may have been affected by completion
parameters including cluster spacing. This section investigates the effects of the distance between the
clusters on the propagation of multiple hydraulic fractures. In the 45-m spacing configuration, the location
of the middle fracture (fracture 2) remains stationary and the exterior fractures are moved out 15-m (in x-
axis) compared to the base case (30-m spacing). Consequently, the fracture 2 in both cases has the same
interaction with interfaces, but fractures 1 and 3 are placed in different locations. This models explain the
importance of variation of weak bedding planes on the hydraulic fractures growth.
Figure 10 shows the pressure distribution when the perforation cluster is equal to 30m (left side) and
45m (right side).
SPE-184854-MS 11

Figure 10Fluid pressure profile after 30 minutes injection. A) 30 meters cluster spacing. B) 45 meters cluster spacing.

By comparing fracture 3 in small and large spacing cases (Figure 10), observed differences in pressure
are a result of the interface variations. In the case that cluster spacing is 30-m, fracture 3 intersects with
the larger portion of interfaces and pressure increases at inlet due to height containment. By increasing the
spacing to 45 meters, there is no indication of truncation in downwards growth and larger upward growth
on the back side of fracture 3. Also vertical growth is truncated by weak bedding planes in the back side
of fracture 1.
Fracture width distribution is shown in Figure 11, and sub plots a, b and c show the aperture distributions
of fractures 1, 2 and 3 at the indicated times. The weak points (in dark blue) along the laminations above
and below each fracture initiation point where vertical growth might be truncated by bedding is illustrated
in sub plots a, b and c.
12 SPE-184854-MS

Figure 11A, B) Aperture profile of 30-m and 45-m cluster spacing. Sub plots a, b, c represent fracture 1, 2, 3. This figure
illustrates that increasing the spacing reduces interference and provided wider and larger width of middle fracture.

Figure 12 shows the fracture width vs. time at the inlet. Increasing the spacing to 45-m, reduces
interference and middle fracture can achieve larger fluid flow. The benefit of larger (33%) cluster spacing
is not only that interior fracture length/height is larger, but also that fracture width is larger (50%).

Figure 12Fracture width vs. time at inlet. Green and red curves represent base (30-m
spacing) and 45-m spacing. Square dot, dash dot and solid line represents fracture 1, 2 and 3.

Increasing the spacing reduces interference and provided larger propped surface area (+7%) which can
improve the proppant placement (Figure 13) with 10% less required pressure.
SPE-184854-MS 13

Figure 13Comparing the total area vs. propped area for base (30-m spacing), and 45-m spacing. Orange and blue bars
show the total area and propped area in this figure. The ratio of these is shown as a percentage above the blue bars.

Conclusion
To make investments in unconventionals pay off, more than the conventional approach is required.
Understanding where proppant goes and how fractures propagate is needed when considering reservoir
complexities. To date, making some assumptions about the reservoir using highly idealized, over-simplified
mathematical models is common, but these models are usually two-dimensional and don't accurately
reflect reservoir reality. Current numerical simulators generally assume 2D or pseudo-3D planar fracture
geometry and weak coupling behavior. This approach cannot account for laterally varying properties
that lead to laterally variable interactions between fractures. In this paper, a 3D fully coupled flow and
mechanical simulator to model a complex fracture propagation in the existence of weak bedding interfaces
and capturing some of the fundamental effects which influence fracture growth during stimulation. In this
study interactions were modeled among three clusters in a horizontal well. Height growth is contained
where fractures intersect weak portions of bedding surfaces, leading to different but asymmetric fracture
length, height and width. Total surface area, pressure and width are investigated for each scenarios. Then
the propped area was estimated, because without proppant, the fracture will close so an unpropped area
won't contribute to production.
Four different scenarios were investigated: 1) 10-cp fluid, 20-bpm injection rate with 30-m cluster
spacing. 2) 100-cp fluid and 20-bpm fluid injection rate with 30-m cluster spacing. 3) 10-cp fluid and 40-
bpm injection rate with 30-m cluster spacing. 4) 10-cp fluid, 20-bpm and 45-m cluster spacing.
A number of observations are summarized below:

In the case with no weak bedding planes, exterior fractures grow with identical pressure and
aperture profile.
Total fracture surface area is largest when using thin fluid, but fracture width is smaller than the
high-viscosity fluid and only 30% of fracture is propped.
In the 100-cp case, 65% of total area is propped and it requires larger pressure (2 times
greater) compare to 10-cp fluid. The difference between pressures is important when converted to
horsepower and directly impacts operational cost. Higher surface pressure means more cost.
In a case with higher injection rate (40-bpm) compare to 20-bpm, for the given volume, pressure
is slightly higher (~45 psi) but propped area is 10% larger. This means with about the same cost,
40-bpm has 10% more propped area.
14 SPE-184854-MS

Increasing cluster spacing allows middle fracture to grow larger in length, height and width. By
optimizing the perforation location (45-m), total propped area could be increased up to 10%.
These fully 3D modeling results reveal the importance of considering such variability, leading to
improved understanding of the risk associated with ignoring it, and allow determination of the value
of technologies which provide realistic subsurface images that allow deterministic application of full
3D simulation. Capturing such subsurface complexities and utilizing in modeling of hydraulic fracture
propagation helps us to improve treatment designs, operational costs and ultimately hydrocarbon recovery.

Acknowledgements
The authors gratefully acknowledge Baker Hughes for permission to publish this paper. We have also
benefitted from the insights of others, especially Pengcheng Fu, Lawrence Livermore National Laboratory.

References
Abass, H. H., Soliman, M. Y., Tahini, A. M., Surjaatmadja, J., Meadows, D. L., and Sierra, L. (2009). Oriented fracturing:
A new technique to hydraulically fracture an openhole horizontal well. Proceedings SPE Annual Technical Conference
and Exhibition, New Orleans, LA, USA. SPE 124483.
Al-Aasm, I. S., M. Coniglio, and A. Desrochers (1995), Formation of complex fibrous calcite veins in Upper Triassic
strata of Wrangellia Terrain, British Columbia, Canada, Sediment. Geol., 100, 8395.
Blanton TL. An Experimental Study of Interaction Between Hydraulically Induced and Pre-Existing Fractures. SPE
Unconventional Gas Recovery Symposium; Pittsburgh, Pennsylvania, 1982.
Bunger, A. P. (2013). Analysis of the power input needed to propagate multiple hydraulic fractures. Int. J. Solids Struct.,
50, 15381549.
Chuprakov, D., Melchaeva, O., & Prioul, R. (2013, May 20). Hydraulic Fracture Propagation Across a Weak Discontinuity
Controlled by Fluid Injection. International Society for Rock Mechanics.
Chuprakov DA, Akulich AV, Siebrits E, Thiercelin M. Hydraulic-Fracture Propagation in a Naturally Fractured Reservoir.
SPE Production & Operations. 2011; 26(1).
Cipolla, C., Weng, X., Onda, H., Nadaraja, T., Ganguly, U., and Malpani, R. (2011). New algorithms and integrated
workflow for tight gas and shale completions. Proceedings SPE Annual Technology Conference and Exhibition,
Denver, Colorado, USA. SPE 146872.
Delaney, P. T., and D. D. Pollard (1981), Deformation of host rocks and flow of magma during growth of minette dikes
and breccia-bearing intrusionsnear ship rock, New Mexico, U.S. Geol. Surv. Prof. Pap., 1202.
Daneshy AA. Hydraulic Fracture Propagation in the Presence of Planes of Weakness. SPE European Spring Meeting;
Amsterdam, Netherlands, 1974.
Economides, M. J., and K. G. Nolte (2000), Reservoir Stimulation, 3rd ed., John Wiley, Hoboken, N. J.
Fisher, M. K., Heinze, J. R., Harris, C. D., Davidson, B. M., Wright, C. A., and Dunn, K. P. (2004). Optimizing horizontal
completion techniques in the barnett shale using microseismic fracture mapping. Proceedings SPE Annual Technology
Conference and Exhibition, Houston, Texas, USA. SPE 90051.
Fu, P., L. Cruz, D. Moos, R. R. Settgast, and F. J. Ryerson. "Numerical Investigation of a Hydraulic Fracture Bypassing
a Natural Fracture in 3D." In 49th US Rock Mechanics/Geomechanics Symposium. American Rock Mechanics
Association, 2015.
Fu, P., S.M. Johnson, and C.R. Carrigan. 2013. An explicitly coupled hydro-geomechanical model for simulating hydraulic
fracturing in complex discrete fracture networks. Int J Numer Anal Methods Geomech, 34(14): 2278-2300.
Garagash, D. I. (2006),Transient solution for a plane-strain fracture driven by a shear-thinning, power-law fluid, Int. J.
Numer. Anal. Methods Geomech., 30, 14391475.
Germanovich, L. N., Ring, L. M., Astakhov, D. K., Shlyopobersky, J., and Mayerhofer, M. J. (1997). Hydraulic fracture
with multiple segments II: Modeling. Int. J. Rock Mech. Min. Sci., 34(3-4), 472.
Germanovich, L. N., and D. Astakhov (2004), Fracture closure in extension and mechanical interaction of parallel joints,
J. Geophys. Res., 109, B02208, doi:10.1029/2002JB002131.
Gu, H., Siebrits, E., & Sabourov, A. (2008, January 1). Hydraulic Fracture Modeling With Bedding Plane Interfacial Slip.
Society of Petroleum Engineers. doi:10.2118/117445-MS
Hammond, P. S. (1995), Settling and slumping in a Newtonian slurry, and implications for proppant placement during
hydraulic fracturing of gas wells, Chem. Eng. Sci., 50, 32473260.
SPE-184854-MS 15

Huang, H., and J. Desroches (2004), A PkN hydraulic fracturing model with piecewise fluid rheology, paper presented at
Gulf Rocks 2004, 6th NARMS: Rock Mechanics Across Borders and Disciplines, Am. Rock Mech. Assoc., Houston,
Tex., 5 9 June.
King, George E., Leonard, Dick.: "Utilizing Fluid and Proppant Tracer Results to Analyze Multi-Fractured Well Flow
Back in Shales: A Framework for Optimizing Fracture Design and Application," SPE 140105, prepared for the SPE
Hydraulic Fracturing Conference and Exhibition, The Woodlands, TX, USA, 24-26 January, 2011
King, G.E.: "Thirty Years of Gas Shale Fracturing: What Have We Learned?" SPE 133456. presented at the 2010 SPE
Annual Technical Conference and Exhibition, Florence, Italy, September 19-22, 2010.
Lacazette, A., and T. Engelder (1992), Fluid-driven cyclic propagation of a joint in the Ithaca siltstone, Appalachian Basin,
New York, in Fluid Mechanics and Transport Properties of Rocks, edited by B. Evans, and T.-F. Wong, pp. 297324,
Elsevier, New York.
McClure, M. W. and Horne, R. N. (2013). Discrete Fracture Network Modeling of Hydraulic Stimulation: Coupling Flow
and Geomechanics. Springer Briefs in Earth Sciences. Springer, New York.
Meyer, B. and Bazan, L. (2011). A discrete fracture network model for hydraulically induced fractures-theory, parametric
and case studies. Proceedings SPE Hydraulic Fracturing Technology Conference and Exhibition, The Woodlands,
Texas, USA. SPE 140514.
Miller, C. and Waters, G. (2011). Evaluation of production log data from horizontal wells drilled in organic shales.
Proceedings SPE North American Unconventional Gas Conference and Exhibition, The Woodlands, Texas, USA. SPE
144326.
Nagel, N., Gil, I., Sanchez-Nagel, M., and Damjanac, B. (2011). Simulating hydraulic fracturing in real fractured rocks
- overcoming the limits of pseudo3D models. Proceedings SPE Hydraulic Fracturing Technology Conference and
Exhibition, The Woodlands, Texas, USA. SPE 140480.
Olson, J. E. (2004). Predicting fracture swarms the influence of subcritical crack growth and the crack-tip process zone
on joint spacing in rock. The initiation, Page 15 propagation, and arrest of joints and other fractures, J. W. Cosgrove
and T. Engelder, eds., Geological Society, London, vol. 231, 7387.
Olson, J. E. and Dahi-Taleghani, A. (2009). Modeling simultaneous growth of multiple hydraulic fractures and their
interaction with natural fractures. Proceedings SPE Hydraulic Fracturing Technology Conference and Exhibition, The
Woodlands, Texas, USA. SPE 119739.
Peirce, A., & Bunger, A. (2015, April 1). Interference Fracturing: Nonuniform Distributions of Perforation Clusters That
Promote Simultaneous Growth of Multiple Hydraulic Fractures. Society of Petroleum Engineers. doi:10.2118/172500-
PA
Settgast, R. R., Johnson, S. M., Fu, P., & Walsh, S. D. C. (2014). Simulation of Hydraulic Fracture Networks in Three
Dimensions Utilizing Massively Parallel Computing Resources. Unconventional Resources Technology Conference.
doi:10.15530/URTEC-2014-1923299
Srivastava, D. C., and T. Engelder (1991), Fluid evolution history of brittle-ductile shear zones on the hanging wall of
Yellow Spring Thrust, Valley and Ridge Province, Pennsylvania, USA, Tectonophysics, 198, 23-34.
Zhang X, Jeffrey RG, Thiercelin M. Deflection and propagation of fluid-driven fractures at frictional bedding interfaces:
A numerical investigation. J Struct Geol. 2007; 29(3):396-410.

Das könnte Ihnen auch gefallen