Você está na página 1de 20

SPE-184858-MS

An Experimental Investigation of the Conductivity of Unpropped Fractures


in Shales

Weiwei Wu, Pratik Kakkar, Junhao Zhou, Rodney Russell, and Mukul M. Sharma, The University of Texas at Austin

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
There is a great deal of evidence that shows that hydraulic fracturing creates a large surface area of induced
unpropped (IU) fractures, that are too small to accommodate commonly used proppants and subsequently
close during production (Sharma and Manchanda 2015). Due to their enormous surface area, IU fractures
can play an important role in hydrocarbon production if they are allowed to remain open during production.
Therefore, the conductivity of these IU fractures under different conditions of stress and when exposed to
different fracturing fluids is of great importance.
In this study, core-scale IU fractures were created with preserved shale samples from the Eagle Ford
and Utica shales. Samples with different mineralogies were selected to represent a broad cross-section of
representative samples. Great care was taken to ensure that the shale samples were preserved since we had
observed that shale desiccation results in large changes in mechanical properties. The fracture conductivities
of unpropped fractures created in each of the shale samples was measured as a function of closure stress
using nitrogen or brine. The unpropped fractures were exposed to several water-based fracturing fluids
including neutral brine, alkaline brine (pH 11-12) and acidic brine (pH < 1), with or without clay stabilizers.
The effects of fluid type, pH, clay stabilizers, shale mineralogy and cyclic stress on IU fracture conductivities
were investigated. Batch tests were also performed to study the change of mechanical properties and fines
production caused by fluid-shale interaction.
Unpropped fractures demonstrated conductivities that were 3 to 4 orders magnitude lower than propped
fractures, and were more susceptible to closure stress. Exposure to water-based fracturing fluids decreased
the unpropped conductivity by one order of magnitude. The primary mechanism for the decrease was
shale softening caused by exchange of water and ions between the native fluid of shale and the exposed
fracturing fluid. Shale softening was observed in exposure to all brines tested, regardless of their pH. In
addition to shale softening, fines generation also contributed to the reduction of unpropped conductivity
when shales were exposed to alkaline or acidic brine. Amine-based clay stabilizers improved the unpropped
conductivity by reducing the amount of clay-based fines. However, they were not as effective at stabilizing
non-clay fines. Shale mineralogy affected the unpropped conductivities in two ways: it controlled the
mechanical properties of the native preserved shale, and also impacted the fluid-shale interactions. A clear
correlation was observed between mineralogy and stress dependence. Clay-rich samples showed the most
2 SPE-184858-MS

stress sensitivity in the presence of water or brine at neutral pH, whereas the calcite-rich samples showed
less stress sensitivity. High clay content also resulted in lower restored conductivity after cyclic stress.
Mechanical properties of shale such as hardness and Young's modulus, before and after fluid exposure,
strongly correlated with the mineralogy of shales. Unpropped conductivity was more sensitive to cyclic
stress than propped conductivities, and it dropped by 80% after one cycle of closure stress between 300
and 4000 psi of closure stress. It is clearly shown that water-based fracturing fluids are able to impact
conductivities of IU fractures in shales significantly, and these impacts need to be taken into account in the
selection of fracturing fluids.

Introduction
A large number of induced unpropped (IU) fractures are created during hydraulic fracturing in shale.
Their presence has been proved by several independent sources of evidence including microseismic data
(Manchanda and Sharma 2013), leakoff analysis (Sharma and Manchanda 2015), tracer test (Manchanda
et al. 2014), production history matching (Mayerhofer et al. 2010) and inter-well pressure communication
(Gupta et al. 2012). These IU fractures induced during the process of pumping in the fracturing treatment,
however, cannot be accessed by proppants, either because their fracture width is smaller than the proppants
or because of flow constrictions in the fracture that result in proppant buildup and jamming. IU fractures
form by tensile or shear failure along planes of weakness. These include pre-existing faults and natural
fractures (Zoback et al. 2012), bedding planes (Gross et al.1995), or micro-heterogeneity such as mineral
interfaces (Ouchi et al. 2017).
The importance of IU fractures in hydraulic fracturing has been progressively recognized: IU fractures
are now thought to account for a major part of the leakoff (Fan et al.2010; Sharma and Manchanda 2015),
can influence fluid flowback (McClure 2014), interfere the growth of subsequent fractures (Manchanda et
al. 2014), lead to different fracture design and production profile (Mayerhofer et al. 2010; Manchanda et
al. 2014; Gu et al. 2015), and play critical roles in diagnostic fracture injection test (DFIT, Potocki 2012)
and after-closure analysis (ACA, Chipperfield 2006; Soliman et al. 2010). Most of the above impacts of IU
fractures are directly or indirectly related to the hydraulic conductivity of the fracture.
The conductivity of an IU fracture (which by definition has no proppant in it) decreases when it is
subjected to the complex mechanical process that it experiences when a closure stress is applied to it. As
both sides of the IU fractures are pressed against each other, only a certain fraction of the surfaces are in
contact due to the rough topography of the fracture surfaces. This uneven contact creates a localized stress
which in turn deforms the surface asperities and changes the fracture apertures and the conductivity.
Several experimental studies have been performed in the past to understand the closure and the
conductivity of unpropped fractures under stress (Nierode and Kruk 1973; Beg et al. 1998; Ruffet et al.
1998; Montemagno and Pyrak-Nolte 1999; Fredd et al. 2001; Lee and Cho 2002; Auradou et al. 2005;
Yasuhara et al. 2006; Watanabe et al. 2011; Kamenov et al. 2013; Tripathi and Pournik 2014; Janse et al.
2015). Many empirical and numerical models have also been developed to model the observed behavior
(Nierode and Kruk 1973; Gangi 1978; Tsang and Tsang 1987; Gong et al. 1999; Pyrak-Nolte and Morris
2000; Deng et al. 2012; Kamali and Pournik 2016; Wang and Cardenas 2016; Wu and Sharma 2017).
Unpropped fracture conductivities depend strongly on factors such as surface topography and the
mechanical properties of the asperities. The decline in fracture conductivity with stress, therefore, often
varies substantially among different rock types with different mineralogy (Nierode and Kruk 1973; Pyrak-
Nolte and Morris 2000). Most of the previous experimental studies were performed on carbonate cores
(Nierode and Kruk 1973; Beg et al. 1998; Ruffet et al. 1998), coal (Montemagno and Pyrak-Nolte 1999),
granite (Lee and Cho 2002; Auradou et al. 2005; Watanabe et al. 2011), marble (Lee and Cho 2002),
novaculite (Yasuhara et al. 2006), and sandstone (Fredd et al. 2001). These unpropped conductivities may
not reflect the behavior of shales. Shale outcrops were used in several recent studies (Kamenov et al. 2013;
SPE-184858-MS 3

Tripathi and Pournik 2014; Janse et al. 2015). The behavior of these unpreserved shales is expected to be
different than that of native shales. Significant changes in mechanical properties have been observed as a
shale is exposed to air (Jung et al. 2013). Outcrop shales or desiccated shales that have been exposed to air
and allowed to dry show very large changes in both elastic moduli and Brinell hardness when compared
with native, preserved shales. It is, therefore, important to use preserved shale samples to better approximate
the in-situ conditions.
A significant percentage of all fracturing fluids are water-based fluids in current hydraulic fracturing
practice. The pH and composition of the fracturing fluids can vary greatly (Gulbi and Hodge, 2000).
Additives such as friction reducers, crosslinkers and clay stabilizers are commonly used. Shale properties
can be significantly affected when it is exposed to these fluids. Shale softening is frequently observed
when a shale is exposed to brine or fresh water, leading to large changes in mechanical properties such as
hardness, Young's modulus, and compressive strength, (Zhang et al. 2006; Akrad et al. 2011; Zhou 2015).
Fines production can be also be induced by fluid exposure. Both shale softening and fines production result
in reduced conductivities in propped fractures (Pedlow and Sharma 2014; Zhang et al. 2015). The fluid-
shale interaction is not only affected by fluids, but also depends on shale properties, especially mineralogy
(Zhou 2015). So far, little is known about how the interaction between water-based fracturing fluids and
shale affects the conductivity of unpropped fractures.
In this work, we present an experimental study that investigates how the conductivity of unpropped
fractures in shales changes as it is exposed to different water-based-fluids. Experiments were conducted
with preserved, native Eagle Ford and Utica shale core samples. Effects of fracture fluid pH, clay stabilizers,
shale mineralogy and cyclic stress were investigated. Batch tests were also performed to study the impact
of fluid-shale interaction on mechanical properties of shale and on fines production.

Materials and Methods


Shale Samples
Shale samples from Eagle Ford and Utica formations were used in this study. Special procedures, described
in detail in previous studies (Chenevert and Amanullah 2001; Pedlow and Sharma 2014), were followed to
preserve the shale samples during handling and storage. These procedures have been found to be critical
for reducing the changes in the mechanical and petrophysical properties of the shale (Jung et al. 2013).

Fracture Conductivity
Shale cores with a diameter of 1 or 1.5 inches and a length of 1 to 3 inches were cut into half in the axial
direction to create the unpropped fractures. The fractured core was then loaded into a Hassler core holder
for fluid exposure and conductivity measurements. A schematic of the experimental setup is shown in Fig.1.
At each applied closure stress, fractures were first flowed with fluids of interest at 5 cc/min for one hour
and left exposed to the fluid for 24 hours. Nitrogen was then introduced for one hour to flush out the fluids
before a conductivity measurement was made.
4 SPE-184858-MS

Figure 1Schematic for experimental setup used for conductivity measurements.

During conductivity measurements, the downstream pressure was maintained constant at 500 psi.
Nitrogen stored in an accumulator pressurized to 3000 psi was pumped at a constant rate through the fracture
until a stable pressure drop was achieved. For a given closure stress, at least four flow rates were used and
their corresponding pressure drops were recorded to calculate the fracture conductivity with the Forchheimer
equation (Eq.1) (any non-Darcy flow effect was taken into account).

(1)

Where p is the pressure drop [atm], q is the volumetric velocity [cc/s], is the Forchheimer coefficient
[1/cm], is the fluid density [g/cc], L is the length of the core [cm], D is the diameter of the core or fracture
height [cm], w is the fracture width [cm], is the fluid viscosity [cp], and kf is the fracture permeability
[Darcy].
To study the effect of mineralogy, a slightly different approach was used. The fracture surfaces of the
halved cores were polished sequentially by 60-, 150- and 320- grit sandpapers to ensure a starting roughness
of around 2 m. This was done to eliminate the effect of the initial variable surface topography left by
the saw cut. The surface topography was verified with an optical profiler, described in a later section. The
polished cores were exposed to 3% KCl brine at 170 F at atmospheric pressure for one hour before being
placed into the core holder for conductivity measurements. Nitrogen was flowed through the sample at 10
to 150 cc/min until a stable inlet pressure was obtained while the outlet pressure was maintained constant
at 300 psi. The flow rate was carefully regulated by a mass flow controller to ensure a Darcy-flow regime,
and the conductivity was calculated with Eq.2.

(2)

Where qsc is nitrogen flow rate at surface condition [scc/s], kfw is the fracture conductivity [md.ft]; T and
Tsc are the test temperature and surface temperature, respectively [R]; Pin, pout and psc are the inlet, outlet
and surface pressure, respectively [atm], and Z is the compressibility factor of nitrogen at the measurement
conditions.
The unpropped fracture was exposed to the following fluids: neutral brine 3% KCl (with pH between
6-7), alkaline brine (3% KCl with pH adjusted to 11.5-12 by adding Na2CO3), acidic brine (1% HCl in 3%
SPE-184858-MS 5

KCl, pH =0.75). Clay stabilizers (0.5 gpt) were added into the neutral brine and alkaline brine in some of
the conductivity and batch tests. The clay stabilizers used were polyquaternary amines (PQA#1, PQA#2),
tetramethylammonium chloride (TMAC) and choline chloride.

Elemental Composition and Mineralogy


The elemental composition and mineralogy of shale was determined by a portable energy-dispersive x-
ray fluorescence (ED-XRF) spectrometer (Bruker, Tracer IV-SD), which allows rapid, non-destructive
elemental quantification on a wide range of rock types including shale (Marsala et al. 2012; Rowe et al. 2012;
Zhou 2015). Consistent results between ED-XRF and traditional wavelength-dispersive x-ray fluorescence
(WD-XRF) had been demonstrated by Rowe et al. (2012).
Shale samples with flat surfaces were rinsed with hexane to remove any superficial oil, and
analyzed by low-energy X-ray fluorescence (15 kv, 40 A) for 60 seconds under vacuum. The excited
fluorescence spectra associated with the major elements were recorded and quantified against the reference
concentrations through software S1calprocess. At least three measurements were performed for each
sample. The relative abundance of minerals was then estimated based on the average measured elemental
concentrations, in which elements were assigned to minerals based on the following assumptions: aluminum
(Al) only exists in clay; calcium (Ca) is found in calcite and dolomite; while silicon (Si) is present in both
quartz and clay.

Indentation Test
The shale Brinell hardness (BHN) and Young's modulus were measured using a modified indentation
test adapted from the standard method for hardness measurement on metallic materials (ASTM E10-15a).
Samples were placed on the flat platform of a load frame. Samples of preserved shale were indented by a
spherical tungsten carbide indenter 5mm in diameter, at a displacement rate of 0.003 inch/min. When the
applied load reached 50 lb, the indenter was held for 10s to allow full development of plastic deformation.
An unloading process at 0.001 inch/min was followed until the indenter was detached from the sample. The
load and displacement were closely monitored and recorded for the entire loading and unloading process.
BHN is calculated by either directly dividing the maximum load applied P (50 lb in this case) by the
projected area of indentation (Eq. 3), or by using the load-displacement curve (Eq.4).

(3)

Where BHN is the Brinellhardness [MPa or kgF/mm2], P is the maximum load applied, 50 lb used in the
study, and A is the projected area of the dent [mm2].

(4)

Where S is the slope of the linear part of the load-displacement curve during the loading stage [kgF/mm],
Di is the diameter of the indenter, which is 5 mm in this case.
The projected area of indentation used in Eq.3 was measured with an optical profiler to ensure accuracy.
An example of the high- resolution image of the dent is shown in Fig.2. The dented area was then quantified
through an image processing software.
6 SPE-184858-MS

Figure 2Dent created in indentation test and imaged by a surface profiler.

The Young's modulus was also estimated based on the load-displacement curve during unloading stage
of indentation. The effective modulus Eeff is calculated as

(5)

Where hmax and hr are the maximum displacement by indentation and residual displacement after fully
unloading, respectively; is the slope of early part of unloading load-displacement curve; R is the effective
radius that can be obtained from Eq. 6.
(6)
Where R1 is the contact radius at the indented depth between indenter and sample; Ri is the radius of the
indenter, 2.5 mm. The Young's modulus of the shale sample can then be calculated using Eq.7:

(7)

where E is the Young's modulus of the shale sample; v is the and Poisson's ratio of the shale sample, assumed
as 0.3 in the calculation; Ei is the Young's modulus of the indenter, 700 GPa for tungsten carbide; and vi is
the Poisson's ratio of the indenter, 0.24.
A detailed derivation for the above equations related to interpretation of load-displacement curves
(Eq.4-7) can be found in Oliver and Pharr (2004).

Surface Topography
The surface roughness of the polished surface and the area of indentation were quantified with a 3D optical
profiler in a non-contact manner. Samples were placed on a vibration-free computer-controlled motorized
stage, and scanned under the vertical scanning interferometry (VSI) mode with back-scan distance of
100-500 m and forward-scan distance of 100-1000 m. One scan under a magnification of 10 x 0.55 times
typically covers an area of 1.2 mm by 0.9 mm, and acquires around 250,000 data points/ mm2. The profiler
achieves a lateral resolution of 0.38 m and an extremely high vertical resolution of 0.01 m enabled by
light interference. In some cases, multiple scans were stitched together with 20% area overlapped in between
scans to cover a broader area. Three to four measurements at random locations were performed for each
sample. The obtained surface topography was visualized in 2D or 3D. To calculate the roughness, the profile
data were first corrected to achieve a zero mean and remove any effect of sample tilting by plane fitting;
then a root mean squared roughness Rq was calculated as:
SPE-184858-MS 7

(8)

Where Zi is the ith corrected surface height [m].

Results and Discussion


Effect of Brine
To study the effect of brine, pH and clay stabilizers on unpropped fracture conductivities, preserved Eagle
Ford shale samples with a high clay content (around 40 wt%) were used, since clay plays important role in
fluid-shale interaction. All the unpropped fractures, except the one treated by nitrogen, were exposed to the
corresponding water-based-fluids for 24 hours at each effective stressand then dried by flowing nitrogen
for 1 hour before each conductivity measurement (all conductivities were measured with nitrogen to avoid
any multi-phase flow effects).
All the unpropped fractures, regardless of the fluid treatment received, demonstrated much lower fracture
conductivities compared to propped fractures (Fig. 3A). The unpropped fracture conductivities started at
about the same magnitude of 1 mD.ft at a closure stress of 1000 psi, and decreased to 10-2 - 10-1 mD.ft at
5000 psi (a decrease of 88- 99%). Propped fractures with a proppant loading of 1 lb/ft2 were more resistant
to closure stress, and their conductivity only decreased by 40 % when closure stress increased to 5000
psi. They maintained a conductivity of 102 mD.ft. Even higher conductivities (1000-2000 mD.ft) can be
obtained at 4000- 6000 psi in propped fractures with an increased proppant loading or with proppants of
enhanced mechanical strength such as ceramics or bauxite (Palisch et al 2007; Pedlow and Sharma 2014).
It is difficult to compare the effect of fluids when using the absolute conductivities (Fig.3A), since these
are also a strong function of surface topography. The unpropped fractures in this section were achieved by
a saw cut and no additional surface polishing was employed. The different surface topography for different
fractures is expected to lead to different low stress conductivities. However, the relative change of fracture
conductivities with respect to closure stress clearly indicates the effect of different fluids. Normalized
conductivities, i.e. the conductivities normalized to the initial conductivities measured at 1000 psi, were
used here to show the relative change in fracture conductivities (Fig.3B).

Figure 3Absolute (A) and normalized (B) conductivities of unpropped fractures of high-
clay Eagle Ford shale (40 wt% clay). PQA#1: is a poly-quaternary amine based clay stabilizer.

Unpropped fractures exposed to brine, either at neutral or high pH, always resulted in a steeper
conductivity reduction with closure stress, compared to that only exposed to nitrogen (Fig.3B). The
8 SPE-184858-MS

normalized conductivity under without exposure to brine at 5000 psi was around 0.12, while it dropped by
50%-91% to 0.01-0.06 after exposure to brine. The reason for the steeper reduction is thought to be due to
the softening of the shale caused by brine exposure. The lower hardness and Young's modulus, results in
more deformation of asperities on the face of the unpropped fractures under closure stress, and this decreases
the fracture aperture and conductivity as the stress is increased (Wu and Sharma, 2017).
When shale is exposed to a water-based fluid, exchange of water and ions occurs between the native fluid
of the shale and the bulk solution (Zhang et al. 2006; Zhou 2015). If the water activity of the native fluid
is lower than that of the bulk solution, the shale will gain more water, which often softens the shale (Zhang
et al. 2006; Zhou 2015). The Eagle Ford shale used here had a native water activity of 0.46, measured in
our earlier study (Zhou 2015), and this was much lower than the water activity of the brine solution (>0.9).
Water was measured to be moving into shale from brine, and the amount of water transferred rose with the
water activity difference between the native fluid in shale and the brine as shown in Fig.4.

Figure 4Water gained by Eagle Ford shale when exposed to brines of various water activities.

Gaining additional water has been shown to cause a reduction in both the hardness and the Young's
modulus, as demonstrated for an Eagle Ford (Fig.5A) and a Utica (Fig.5B) shale. The reduction increased
with the water activity of the brines that the shale is exposed to. Higher water activity brines transferred
more water to the shale as seen in Fig.4.

Figure 5Hardness and Young's modulus reduction in Eagle Ford (A) and Utica (B) shale caused by brine exposure.

Mineralogy, in particular the clay content, appeared to affect the extent of reduction in hardness and
moduli: higher clay content tended to show more reduction (Fig.5b). The mechanical properties of shale
SPE-184858-MS 9

after brine exposure were also found to be inversely correlated with the clay content. This correlation was
observed in preserved samples from other shale plays as well, including the Barnett, Haynesville and Utica
(Fig.6) shales, and agrees with observations made in our earlier study (Zhou, 2015). Unpropped conductivity
in clay-rich shale is more susceptible to brine exposure. Increased clay content also leads to more proppant
embedment in propped fractures (Alramahi and Sundberg 2012).

Figure 6Hardness of shales after exposure to 3% KCl

Effect of Brine pH (pH =11-12)


A high pH environment is often necessary to stabilize borate crosslinkers used to gel fluids to improve the
fluid proppant carrying ability (Gulbi and Hodge, 2000). We compared the unpropped conductivity under
neutral pH and alkaline pH (pH=11 -12). The alkaline brine was found to be consistently detrimental to
unpropped fracture conductivity (Fig. 3B). The conductivity of unpropped fractures treated by brines with
an alkaline pH of 11-12 always decreased more rapidly with closure stress than those treated by the same
brines but with a neutral pH. At a closure stress of 5000 psi, neutral brine with or without clay stabilizers
yielded a normalized conductivity as 3-6 times higher than shales exposed to the same fluid but at a higher
pH (above 11).
Shale softening was first thought to be responsible for the lower conductivity measured for alkaline
brine. For shales with a high clay content (37.5 wt%), exposure to alkaline brine did lead to lower hardness
(decreased by 52%) than neutral brine (decreased by 24%) after 5 hours of exposure (Fig. 7A). However, the
hardness difference progressively decreased with exposure time. The hardness values were 10% different
after 24 hours, and about the same after 40 hours, where both decreased by 58% from the initial level.
Another mechanism for a reduction in the fracture conductivity is the dispersion of clay minerals caused
by high pH solutions (discussed in the following section). This may facilitate the water and ion exchange
between the shale pore fluid and the bulk fluid resulting in a more rapid reduction in hardness than by
neutral brine. Since both shale samples exposed to neutral and alkaline brine had very similar clay content
and native water activity, their capacities to exchange with bulk fluid were expected to be close. Thus, as
both samples approached equilibrium, the difference between their hardness was reduced.
10 SPE-184858-MS

Figure 7Change in hardness for an Eagle Ford shale with high (A)
and low (B) clay content when exposed to neutral and alkaline brine.

For shales with a low clay content (17.3 wt%), hardness response under high pH was very similar to
that under neutral pH during the entire exposure process. Both decreased slightly from the initial values
(Fig.7B). Shales with low clay content experienced less softening from brine exposure than those clay-rich
ones, similar to what is shown earlier (Fig. 5B and Fig.6).
In the conductivity measurement shown in Fig.3, long exposure to brines (24 hours at each closure stress)
was imposed on the shales before each conductivity measurement. Thus, shales exposed to high pH and to
neutral pH were expected to have similar hardness values during the conductivity measurement. Hence, the
reduced conductivity under high pH was likely not mechanically driven.
Fines generation and mobilization is another common reason for conductivity loss that has been observed
in propped fractures (Pedlow and Sharma 2014; Zhang et al. 2015). To investigate fines generation in
unpropped fractures, aqueous elements in the bulk fluid that shales were exposed to were measured with
inductively coupled plasma mass spectrometry (ICP-MS) in this study. Not all the fines were detected by
ICP-MS due to its low solid tolerance. However, the aqueous elements, Al and Si in particular, are still
useful indicator for fines production. Aluminum is believed to associate with clay-based fines, while Si is
related to both clay-based and non-clay fines from quartz, feldspar or mica.
Fig. 8 shows that alkaline brine produced much more Al and Si than a solution at pH of 3, in which the
amount of Al was almost negligible. This implied that a high pH solution caused more dispersion of the clay
and produced more fines. It is well known that at a higher pH, the electrostatic surface potential becomes
more negative on clay surfaces (Sharma and Yortsos 1987). Based on the DLVO theory, the more negative
potential leads to a more repulsive surface forces and promotes the dispersion of clay particles. A sharp
decrease in the permeability has been reported during injection of a high pH solution in Berea sandstone
(Leone and Scott 1987; Vaidya and Fogler 1992). Neutral brine (pH 5-7) was believed to produce very little
fines at pH of 3, since no noticeable fines were observed for a pH below 9 in previous studies (Vaidya and
Fogler 1992). Therefore, the conductivity damage caused by alkaline brine compared to neutral brine was
probably due to clay dispersion which lead to stimulated production of fines, and not shale softening alone.
SPE-184858-MS 11

Figure 8Aqueous Al and Si in neutral and alkaline brines

Effect of Acidic Brine (pH <1)


Acid is often used at the pre-flush stage of hydraulic fracturing to remove near-wellbore perforation damage.
Recent studies show that acid fracturing in shale can potentially further enhance production by stimulating
unpropped fractures (Grieser et al 2007; Wu and Sharma 2015). It is thus also important to investigate
how unpropped conductivities behave under low pH conditions. We found that an acidic brine at pH <1
demonstrated negative effects on unpropped fracture conductivities. In the carbonate-rich Eagle Ford shale
(71.7 wt% of carbonate), shale treated by acidic brine showed a conductivity almost two orders of magnitude
lower than that treated by neutral brine at a closure stress of 3000 psi (Fig.9). The initial roughness of
fracture surfaces was polished to within 2 m to eliminate the effect of surface topography (Fig.10). The
negative effect by acidic brine on conductivity increased with carbonate content of shale (data not shown),
which was reasonable since carbonate minerals are the main component that reacts with acid. It is important
to note that, the conductivity decrease shown here was based on a uniform exposure of the fracture faces
to the acidic brine. We have shown elsewhere that non-uniform exposure to acidic brine can enhance the
fracture conductivity (Wu, 2017).

Figure 9Unpropped fracture conductivity of Eagle Ford Shale (with 71.7 wt% carbonate, 16.0 wt% clay)
exposed to neutral and acidic brine (pH=0.75). The fracture surfaces were polished before brine exposure.
12 SPE-184858-MS

Figure 10An example of polished fracture surface with an initial roughness of 1.442 m

Shale softening was one important reason for the conductivity loss caused by acidic brine. As shown
earlier, exposure to alkaline and neutral brines caused similar reduction in hardness; while hardness was
always further reduced by acidic brine compared to a neutral brine (Fig.11). The ratio of hardness brought
down by acidic brine to that by neutral brine decreased with carbonate content. Besides fluid exchange,
which was the main contributor to hardness reduction under neutral and alkaline brine, mineralogy change
and increased porosity of the roughened surface also contributed to weakening the shale in acidic brine.
More clay and less carbonate minerals were observed on shales treated by acidic brine (Fig.12), and this is
expected to soften the shale. Hardness is known be inversely correlated to clay content, and positively to
carbonate mineral content (Kumar et al. 2012; Zhou 2015).

Figure 11Comparison between hardness of shale after reaction with acidic brine vs a neutral brine.

Figure 12Mineralogy of Eagle Ford shale changed after exposure to acidic brine (1% HCl in 3% KCl, pH <1)
SPE-184858-MS 13

Acidic brine also promoted fines production which can damage the unpropped fracture conductivity. Our
previous microstructural study showed that as carbonate was dissolved by acidic brine, the neighboring
minerals such as quartz and clay could be dislodged and released as fines into the bulk solution (Wu and
Sharma, 2015). Increased Al and Si, associated with clay-based and non-clay fines, were found in the
aqueous phase.

Effect of Clay Stabilizers


Unpropped conductivities were increased slightly by adding 0.5 gpt clay stabilizer (poly quaternary amine
(PQA#1)) into brines of neutral or alkaline brines (Fig.3). The conductivity increase at closure stress of
4000-5000 psi was more significant in high pH (by about 3 times) than in neutral pH (by 18%).
The conductivity increase can mainly be attributed to fines inhibition. Clay stabilizers are known to attach
to clay platelets to inhibit swelling and fines production. Both alkaline and acidic brines, as discussed earlier,
dispersed and suspended clays and fines into the flow stream. By introducing PQA clay stabilizers into a
high pH solution, the aqueous aluminum content of the brine (indicating clay-type fines) decreased (Fig.13).
Temporary amine-based clay stabilizers such as TMAC and choline chloride also demonstrated an ability to
inhibit release of clay-types fines. In neutral brine, fewer fines were produced compared to high pH (Fig.8).
This explains why clay stabilizers were more effective in alkaline brine than in neutral brine.

Figure 13Aqueous Al and Si in alkaline brines with and without clay stabilizers

It is worth noting that, the amount of aqueous Si in alkaline brine with clay stabilizers was only slightly
reduced or still comparable to (69-100% of) the case without clay stabilizers (Fig.13). Aqueous Si is present
in both clay-based and non-clay fines. The reduction in Si was likely due to a decrease in clay-based fines,
suggested by the reduced Al. Most of the remaining Si after adding clay stabilizers may be from non-clay
fines, and it implies that clay stabilizers were not effective in controlling some non-clay fines.
Clay stabilizers were not able to mitigate shale softening. In spite of adding clay stabilizers, hardness
was at about the same level as without stabilizers (Fig.14). Hence, the conductivity increase seen when clay
stabilizers were used was not due to any alleviation of shale softening.
14 SPE-184858-MS

Figure 14Hardness (A) and normalized hardness (B) of shale


after exposure to alkaline brines with and without clay stabilizers

Effect of Shale Mineralogy


Mineralogy plays an important role in unpropped fracture conductivities, since it affects the mechanical
properties of fractures as well as shale-fluid interactions. Before exposure to fracturing fluids, the
mechanical properties of any shale are strongly correlated with its mineralogy. Hardness and Young's
modulus of fractures generally increase when there is less clay and more carbonate or quartz in a shale
(Kumar et al. 2012; Zhou 2015). Mineralogy thus directly affects the deformation of asperities on the surface
of unpropped fractures and, therefore, its conductivity.
When shales are in contact with fracturing fluids, their mechanical properties after fluid exposure are
strongly affected by mineralogy. When exposed to neutral and alkaline brine, where no chemical reactions
were involved, shales with higher clay content often underwent more fluid exchange and softening (Fig. 5B
and Fig.7), and they were expected to provide lower conductivity.
By polishing the fracture surfaces to ensure the same starting surface roughness of around 2 m (Fig.10),
we were able to eliminate the effect of surface topography and demonstrated a clear correlation between
mineralogy of unpropped fractures and their conductivities after exposure to neutral brine (Fig. 15). In both
loading and unloading stages, a higher conductivity was achieved with less clay and more carbonate (which
has a higher hardness) (Fig. 16). In addition to shale softening, clay-rich shale under high pH is expected to
produce more fines and resulted in even lower conductivities. When exposed to acidic brines, where acid
reacts with and dissolves carbonate minerals, the mineralogy effect can be even more significant. Carbonate-
rich shales tended to be softer and generated less fracture conductivity (Fig. 9 and Fig. 11).

Figure 15Unpropped fracture conductivities of Eagle Ford shale samples with different
mineralogy exposed to neutral brine: (A) absolute conductivity, (B) normalized conductivity.
SPE-184858-MS 15

Figure 16Correlation between fracture conductivity and mineralogy.

Mineralogy can also influence rock-fluid interactions by changing the mineral surface charge properties.
For example, in pure carbonate minerals, the surface potential becomes more positive with pH (Chen et al.
2014); while in a limestone formation with 18 wt% of quartz and clay minerals, surface charge tended to
be more negative at higher pH (Alotaibi et al. 2011). By changing the surface potential, mineralogy can
directly influence the generation of fines and, therefore, have a significant effect on fracture conductivity.

Effect of Cyclic Stress


Fractures often have to experience cyclic closure stress from fracturing to production, from production
to shut-in, and from shut-in to production again. It is believed that fracture conductivity can significantly
decrease during cyclic stress due to the irreversible plastic deformation and fines generation induced by
stress cycling. It is important to consider the effect of cyclic stress on fracture conductivity in fracturing
design and production prediction.
Unpropped conductivity appeared much more sensitive to cyclic stress than propped conductivity. The
unpropped conductivity dropped by around 80% just after one cycle between 500 psi and 4000 psi in a
preserved Eagle Ford shale sample when exposed to neutral brine at neutral pH (Fig.15B). A similar decrease
was also found in unpropped fractures in preserved Utica shale samples with low and high clay content
(Fig.17). More than 97% of the initial conductivity was lost in several unpropped fractures of preserved
Eagle Ford shale samples examined that went through series of stages of loading-unloading-loading between
500psi and 4000 psi (Fig.18).

Figure 17Normalized unpropped fracture conductivities of Utica shale under cyclic closure stress.
16 SPE-184858-MS

Figure 18Normalized conductivities of an Eagle Ford unpropped


fracture that went through loading-unloading-loading stages.

Propped fracture conductivity also decreases during cyclic stress, however, to a smaller degree compared
to unpropped fractures. Palisch et.al. (2007) observed a 30-70% reduction in propped fracture conductivity
after 25 cycles of closure stress between 4000 psi and 8000 psi. The higher susceptibility of unpropped
conductivity to cyclic stress may be due to their extremely small apertures. The apertures in unpropped
fractures are much smaller, as can be seen from their conductivities ranging from 10 3 to 101 mD.ft compared
to 102-104 mD.ft in propped fractures. Therefore, the same level of plastic deformation of asperities, or fines
produced imparts a much greater damage to unpropped fractures.

Conclusions
The present work undertook a detailed study of the effect of fluid type, pH, clay stabilizers, shale mineralogy
and cyclic stress on the conductivity of unpropped fractures in preserved shales. The mechanical properties
(Young's modulus and Brinell hardness) and fines production were also measured to better understand the
results of the fracture conductivity tests. The main conclusions obtained are the following:
1. As expected, unpropped fractures in shales have conductivities that are 3 to 4 orders of magnitude
lower than propped fractures at a closure stress of 5000 psi. In addition, unpropped fracture
conductivities decrease more rapidly with closure stress.
2. Exposure to water-based fracturing fluids decreases the unpropped fracture conductivities by an order
of magnitude. This is mainly due to shale softening induced by contact with water. A lower water
activity in the shale and a higher water activity in the fracturing fluid results in more transfer of water
to the shale. A higher water uptake by the shale results in greater shale softening.
3. An alkaline fracturing fluid (pH of 11.5-12) results in a larger reduction in fracture conductivity than
a neutral brine. The primary mechanism of this is shown to be dispersion of clays and fines, not shale
softening.
4. Acidic brine (pH <1) results in removal of the calcite in the shale and this softens the shale. This change
in mineralogy also promotes fines generation, and this damages the conductivity of the unpropped
fracture.
5. Amine-based clay stabilizers improve unpropped conductivity by reducing the creation of clay-based
fines. However, they are not as effective at reducing the production of non-clay fines.
6. Shale mineralogy impacts unpropped fracture conductivities in two ways: it controls the mechanical
properties of the native preserved shale, and also affects fluid-shale interactions.
7. Higher clay content and lower quartz and carbonate content reduce the conductivity of unpropped
fractures and increase the stress and water sensitivity.
SPE-184858-MS 17

8. Unpropped fracture conductivity drops by 80% after one cycle of closure stress, and is more sensitive
to cyclic stress than the conductivity of propped fractures.

Acknowledgments
We thank Drs. Daniel Stockli and Lisa Stockli for the generous help in the use of the optical profiler. We
are also grateful for the support from companies that sponsor the Hydraulic Fracturing and Sand Control
Joint Industry Project at the University of Texas at Austin.

Nomenclature
A Projected area of dent, mm2
AwB Water activity of brine
BHN Brinell hardness, kgF/mm2 or MPa
D Diameter of the core, cm
Di Diameter of the indenter, 5mm
E Young's modulus of shale sample, GPa
EB Young's modulus of shale exposed to brine, GPa
Eeff Effective Young's modulus, GPa
Ei Young's modulus of indenter, 700 GPa
HB Brinell hardness of shale exposed to brine, MPa
HB' Normalized Brinell hardness of shale exposed to brine
hmax Maximum displacement by indentation, mm
hr Residual displacement after fully unloading, mm
kf Fracture permeabiliy, D or mD
kfw Fracture conductivity, mD.ft
Normalized fracture conductivity
L Length of the core, cm
P Maximum load applied in indentation test, lb
pin Inlet pressure of the core, psi or atm
psc Pressure at surface condition, 14.72 psi
pout Outlet pressure of the core, psi or atm
p Pressure drop across unpropped fracture, psi or atm
dP / dh Slope of early part of unloading load-displacement curve, kgf/mm
q Volumetric velocity, cc/s
qsc Nitrogen flow rate at surface condition, scc/s
R Effective radius, mm
R1 Contact radius at the indented depth between indenter and sample
Ri Radius of the indenter, 2.5 mm
Rq Root mean square roughness, m
S Slope of the linear part of the load-displacement curve during the loading stage, kgf/
mm
T Temperature at test condition, R
Tsc Temperature at surface condition, 520 R
w Fracture width, cm
Z Compressibility factor
zi Corrected surface height, m
18 SPE-184858-MS

Forchheimer coefficient, 1/cm


Fluid density, g/cc
Fluid viscosity, cp
v Poisson's ratio of shale sample, 0.3
vi Poisson's ratio of the indenter, 0.24

References
Alotaibi, M., Nasralla, R. A. and Nasr-El-Din, H. A. 2010. Wettability Challenges in Carbonate Reservoirs. Presented
at Society of Petroleum Engineers Improved Oil Recovery Conference, Tulsa, 2428 April. SPE-129972-MS. http://
dx.doi.org./10.2118/129972-MS.
Alramahi, B. and Sundberg, M.I. 2012. Proppant Embedment and Conductivity of Hydraulic Fractures in Shales. Presented
at the 46th US Rock Mechanics / Geomechanics Symposium, Chicago, 24-27 Jun. ARMA 12-291.
Akrad, O. M., Miskimins, J. L., and Prasad, M. 2011. The Effects of Fracturing Fluids on Shale Rock Mechanical
Properties and Proppant Embedment. Presented at the SPE Annual Technical Conference and Exhibition, Denver,
Colorado, 30 October-2 November. SPE-146658-MS. http://dx.doi.org/10.2118/146658-MS.
ASTM E10-15a. 2015. Standard Test Method for Brinell Hardness of Metallic Materials, ASTM International, West
Conshohocken, PA. http://dx.doi.org/10.1520/E0010-15A.
Auradou, H., Drazer, G., Hulin, J.P. and Koplik, J. 2005. Permeability anisotropy induced by the shear displacement of
rough fracture walls. Water Resources Research, 41(9). http://dx.doi.org/10.1029/2005wr003938.
Beg, M.S., Kunak, A.O., Gong, M., Zhu, D., and Hill, A.D. 1998. A Systematic Experimental Study of Acid Fracture
Conductivity. SPE Production & Facilities. 13(4): 267-271. SPE-52402-PA doi: 10.2118/52402-PA.
Chen, L., Zhang, G., Wang, L., Wu, W. and Ge, J. 2014. Zeta potential of limestone in a large range of salinity. Colloids and
Surfaces A: Physicochemical and Engineering Aspects, 450: 1-8. http://dx.doi.org/10.1016/j.colsurfa.2014.03.006.
Chenevert, M.E. and Amanullah, M. 2001. Shale Preservation and Testing Techniques for Borehole-Stability Studies. SPE
Drill & Compl 16(3): 146-149. SPE-73191-PA. http://dx.doi.org/10.2118/73191-PA.
Chipperfield, S. 2006. After-Closure Analysis to Identify Naturally Fractured Reservoirs. SPE Form Eval 9 (1):50-60.
SPE-90002-PA. http://dx.doi.org/10.2118/90002-PA.
Deng, J., Mou, J., Hill, A. D., and Zhu, D. 2012. A New Correlation of Acid-Fracture Conductivity Subject to Closure
Stress. SPE Prod & Oper 27(02):158-169. SPE-140402-PA. http://dx.doi.org/10.2118/140402-PA.
Fan, L., Thompson, J.W., and Robinson, J.R. 2010. Understanding Gas Production Mechanism and Effectiveness of
Well Stimulation in the Haynesville Shale Through Reservoir Simulation. Presented at the Canadian Unconventional
Resources and International Petroleum Conference, Calgary, Alberta, Canada, 19-21 October. SPE-136696-MS. http://
dx.doi.org/10.2118/136696-MS.
Fredd, C.N., McConnell, S.B., Boney, C.L. and England, K.W. 2001. Experimental study of fracture conductivity
for water-fracturing and conventional fracturing applications. SPE J, 6(03); 288-298.SPE-74138-PA. http://
dx.doi.org/10.2118/74138-PA.
Gangi, A.F. 1978. Variation of whole and fractured porous rock permeability with confining pressure. In
International Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts 15(5):249-257. http://
dx.doi.org/10.1016/0148-9062(79)90824-6.
Gong, M., Lacote, S. and Hill, A.D. 1999. New Model of Acid-Fracture Conductivity Based on Deformation of Surface
Asperities. SPE J. 4(3): 206-214. SPE-57017-PA. http://dx.doi.org/10.2118/57017-PA.
Grieser, B., Wheaton, R., Magness, B., Blauch, M. and Loghry, R. 2007. Surface Reactive Fluids Effect on Shale.
Presented at the SPE Production and Operations Symposium, Oklahoma City, Oklahoma, USA, 31 March1 April.
SPE-106815-MS. http://dx.doi.org/10.2118/106815-MS.
Gross, M., Fischer, M., Engelder, T. and GreenField, R. 1995. Factors controlling joint spacing in interbedded sedimentary
rocks; integrating numerical models with field observations from the Monterey Formation, USA. In: Ameen,
M.S. (Ed.). Fractography. Geological Society Special Publication, Geological Society, London: 215-233. http://
dx.doi.org/10.1144/GSL.SP.1995.092.01.12.
Gu, M., Kulkarni, P., Rafiee, M., Ivarrud, E. and Mohanty, K. 2015. Optimum Fracture Conductivity for Naturally
Fractured Shale and Tight Reservoirs. SPE Prod & Oper. (In press; posted October 2015). SPE-171648-PA. http://
dx.doi.org/10.2118/171648-PA.
Gulbis, J, Hodge, R.M. 2000. Fracturing Fluid Chemistry and Proppants. In Reservoir Stimulation. 3rd edition,
Economides M.J. and Nolte, K.G., Chap.7. New York, NY: Wiley.
Gupta, J.K., Zielonka, M.G., Albert, R.A., El-Rabaa, A.M., Burnham, H.A. and Choi, N.H. 2012. Integrated Methodology
for Optimizing Development of Unconventional Gas Resources. Presented at the SPE Hydraulic Fracturing
SPE-184858-MS 19

Technology Conference, The Woodlands, Texas, 6-8 February. SPE-152224-MS. http://dx.doi.org/10.2118/152224-


MS.
Janse, T., Zhu, D. and Hill, A.D. 2015. The effect of Rock Mechanical Properties on Fracture Conductivity for Shale
Formations. Presented at SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 3-5
February. SPE-173347-MS. http://dx.doi.org/10.2118/173347-MS.
Jung, C.M., Zhou, J., Chenevert, M.E. and Sharma, M.M. 2013, The impact of shale preservation on the petrophysical
properties of organic-rich shales. Presented at SPE Annual Technical Conference and Exhibition, New Orleans,
Louisiana, 30 September-2 October. SPE-166419-MS. http://dx.doi.org/10.2118/166419-MS.
Kamali, A. and Pournik, M. 2016. Fracture closure and conductivity decline modelingApplication in unpropped
and acid etched fractures. Journal of Unconventional Oil and Gas Resources, 14:44-55. http://dx.doi.org/10.1016/
j.juogr.2016.02.001.
Kamenov, A., Zhu, D., Hill, A. D. and Zhang, J. 2013. Laboratory Measurement of Hydraulic Fracture Conductivities
in the Barnett Shale. Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 4
6 February. SPE-163839-MS. http://dx.doi.org/10.2118/163839-MS.
Kumar, V., Sondergeld, C., and Rai, C. 2012. Nano to Macro Mechanical Characterization of Shale. Presented
at SPE Annual Technical Conference and Exhibition, San Antonio, 8-10 October. SPE-159804-MS http://
doi.org/10.2118/159804-MS.
Lee, H.S. and Cho, T.F. 2002. Hydraulic characteristics of rough fractures in linear flow under normal and shear load.
Rock Mechanics and Rock Engineering, 35(4): 299-318. http://dx.doi.org/10.1007/s00603-002-0028-y.
Leone, J.A. and Scott, E.M. 1987, January. Characterization and control of formation damage during waterflooding of a
high-clay-content reservoir. SPE Res Eng, 3(04): 1279-1286. SPE-16234-PA. http://dx.doi.org/10.2118/16234-PA
Manchanda, R., and Sharma, M.M. 2013. Time-Delayed Fracturing: A New Strategy in Multi-Stage, Multi-Well Pad
Fracturing. Presented at the SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, 30 September
2 October. SPE-166489-MS. http://dx.doi.org/10.2118/166489-MS.
Manchanda, R., Sharma, M.M. and Holzhauser, S. 2014. Time-Dependent Fracture-Interference Effects in Pad Wells. SPE
Prod & Oper, 29(04): 274287. SPE-164534-PA. http://dx.doi.org/10.2118/164534-PA.
Marsala, A.F., Loermans, T., Shen, S., Scheibe C., and Zereik, R. 2012. Portable Energy-Dispersive X-Ray Fluorescence
Integrates Mineralogy and Chemostratigraphy into Real-Time Formation Evaluation. Petrophysics, 53(2): 102-109.
Mayerhofer, M.J., Lolon, E.P. and Warpinski, N.R. 2010. What is Stimulated Reservoir Volume? SPE Prod and Oper,
25(1): 16-18. SPE-119890-PA. http://dx.doi.org/10.2118/119890-PA.
McClure, M.W. 2014. The Potential Effect of Network Complexity on Recovery of Injected Fluid Following Hydraulic
Fracturing. Presented at the SPE Unconventional Resources Conference, The Woodlands, Texas, USA, 13 April.
SPE-168991-MS. http://dx.doi.org/10.2118/168991-MS.
Montemagno, C.D. and Pyrak-Nolte, L.J. 1999. Fracture network versus single fractures: measurement of fracture
geometry with X-ray tomography. Physics and Chemistry of the Earth, Part A: Solid Earth and Geodesy,
24(7):575-579. http://dx.doi.org/10.1016/s1464-1895(99)00082-4.
Nierode, D. and Kruk, K. 1973. An Evaluation of Acid Fluid Loss Additives Retarded Acids and Acidized Fracture
Conductivity. Presented at Fall Meeting of the Society of Petroleum Engineers of AIME, Las Vegas, Nevada, 30
September-3 October. SPE-4549-MS. http://dx.doi.org/10.2118/4549-MS.
Oliver, W.C. and Pharr, G.M. 2004. Measurement of hardness and elastic modulus by instrumented indentation:
Advances in understanding and refinements to methodology. Journal of materials research, 19(01): 3-20. http://
dx.doi.org/10.1557/jmr.2004.0002.
Ouchi, H., Agrawal, S., Foster, J.T. and Sharma, M.M. 2017. Effect of Small Scale Heterogeneity on the Growth of
Hydraulic Fractures. Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 24-26
January. SPE- 184873-MS. http://dx.doi.org10.2118/184873-MS.
Palisch, T., Duenckel, R., Bazan, L., Heidt, H.J., and Turk, G. 2007. Determining Realistic Fracture Conductivity
and Understanding Its Impact on Well Performance--Theory and Field Examples. Presented at the SPE
Hydraulic Fracturing Technology Conference, College Station, Texas, USA, 29-31 January. SPE-106301-MS. http://
dx.doi.org/10.2118/106301-MS.
Pedlow, J. and Sharma, M. 2014. Changes in shale fracture conductivity due to interactions with water-based fluids.
Presented at the SPE Hydraulic Fracturing Technology Conference, Woodlands, TX, 4 - 6 Feb. SPE 168586-MS. http://
dx.doi.org/10.2118/168586-MS.
Potocki, D. J. 2012. Understanding Induced Fracture Complexity in Different Geological Settings Using DFIT Net
Fracture Pressure. Presented at the SPE Canadian Unconventional Resources Conference, Calgary, 30 October1
November. SPE-162814-MS. http://dx.doi.org/10.2118/162814-MS.
20 SPE-184858-MS

Pyrak-Nolte, L.J. and Morris, J.P. 2000. Single fractures under normal stress: The relation between fracture specific
stiffness and fluid flow. International Journal of Rock Mechanics and Mining Sciences, 37(1): 245-262. http://
dx.doi.org/10.1016/s1365-1609(99)00104-5.
Rowe, H., Hughes, N. and Robinson, K. 2012. The Quantification and Application of Handheld Energy-Dispersive
X-ray Fluorescence in Mudrock Chemostratigraphy and Geochemistry. Chem. Geol., 324: 122-131. http://
dx.doi.org/10.1016/j.chemgeo.2011.12.023.
Ruffet, C., Fery, J.J. and Onaisi, A. 1998. Acid Fracturing Treatment: A Surface-Topography Analysis of
Acid-Etched Fractures to Determine Residual Conductivity. SPE J., 3(2): 155-162. SPE-38175-PA. http://
dx.doi.org/10.2118/38175-PA.
Sharma, M.M. and Manchanda, R. 2015. The Role of Induced Un-propped (IU) Fractures in Unconventional Oil and Gas
Wells. Presented at SPE Annual Technical Conference and Exhibition Houston, Texas, 28-30 September. SPE-174946-
MS. http://dx.doi.org/10.2118/174946-MS.
Sharma, M.M. and Yortsos, Y.C. 1987. Fines migration in porous media. AIChE Journal, 33(10):1654-1662. http://
dx.doi.org/10.1002/aic.690331009.
Soliman, M.Y., Miranda, C., and Wang, H.M. 2010. Application of After-Closure Analysis to a Dual-Porosity
Formation, to CBM, and to a Fractured Horizontal Well. SPE Prod & Oper, 25(4): 472-483. SPE-124135-PA. http://
dx.doi.org/10.2118/124135-PA.
Tsang, Y.W. and Tsang, C.F. 1987. Channel model of flow through fractured media. Water Resources Research, 23(3):
467-479. http://dx.doi.org/10.1029/wr023i003p00467.
Tripathi, D. and Pournik, M. 2014. Effect of Acid on Productivity of Fractured Shale Reservoirs. Presented at SPE/AAPG/
SEG Unconventional Resources Technology Conference, Denver, Colorado, USA, 25-27 August. SPE-2014-1922960-
MS. http://dx.doi.org/10.15530/urtec-2014-1922960.
Vaidya, R. N. and Fogler, H. S. 1992. Fines Migration and Formation Damage: Influence of pH and Ion Exchange. SPE
Prod Eng, 7(4): 325-330. SPE-19413-PA. http://dx.doi.org/10.2118/19413-PA.
Wang, L. and Cardenas, M.B. 2016. Development of an empirical model relating permeability and specific stiffness
for rough fractures from numerical deformation experiments. Journal of Geophysical Research: Solid Earth, 121(7):
4977-4989. http://dx.doi.org/10.1002/2016jb013004.
Watanabe, N., Ishibashi, T., Hirano, N., Ohsaki, Y., Tsuchiya, Y., Tamagawa, T., Okabe, H. and Tsuchiya, N. 2011. Precise
3D numerical modeling of fracture flow coupled with X-ray computed tomography for reservoir core samples. SPE
J, 16(03): 683-691. SPE-146643-PA. http://dx.doi.org/10.2118/146643-PA.
Wu, W. and Sharma, M.M. 2016. Acid Fracturing Shales: Effect of Dilute Acid on Properties and Pore Structure of Shale.
SPE Prod & Oper. (In press; posted April 2016). SPE-173390-PA. http://dx.doi.org/10.2118/173390-PA.
Wu, W. 2017. Hydraulic Conductivity and Mechanical Behavior of Unpropped Fractures in Shale. PhD dissertation. The
University of Texas at Austin, Austin, Texas (May 2017).
Wu, W. and Sharma, M.M. 2017. A Model for the Conductivity and Compliance of Unpropped Fractures. Presented at the
SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 24-26 January. SPE- 184852-MS. http://
dx.doi.org10.2118/184852-MS.
Yasuhara, H., Polak, A., Mitani, Y., Grader, A.S., Halleck, P.M. and Elsworth, D. 2006. Evolution of fracture permeability
through fluidrock reaction under hydrothermal conditions. Earth and Planetary Science Letters, 244(1): 186-200.
http://dx.doi.org/10.1016/j.epsl.2006.01.046.
Zhang, J., Al-Bazali, T., Chenevert, M., Sharma, M. M., Clark, D., Benaissa, S. and Ong, S. 2006. Compressive strength
and acoustic properties changes in shale with exposure to water-based fluids. Presented at the 41st U.S. Symposium
on Rock Mechanics (USRMS), Golden, Colorado, 17-21 June. ARMA-06-900.
Zhang, J., Zhu, D., and Hill, A. D. 2015. Water-Induced Damage to Propped-Fracture Conductivity in Shale Formations.
SPE Prod & Oper, 31(02):147-156. SPE-173346-PA. http://dx.doi.org/10.2118/173346-PA.
Zhou, J. 2015. Interactions of organic-rich shale with water-based fluids. PhD dissertation, The University of Texas at
Austin, Austin, Texas. (August 2015).
Zoback, M. D., Kohli, A., Das, I. and Mcclure, M.W. 2012. The Importance of Slow Slip on Faults During
Hydraulic Fracturing Stimulation of Shale Gas Reservoirs. Presented at the SPE Americas Unconventional Resources
Conference, Pittsburgh, Pennsylvania, 5-7 June. SPE-155476-MS. http://dx.doi.org/10.2118/155476-MS.

Você também pode gostar