Anda di halaman 1dari 14

URTeC: 1934272

Efficient Completions in Anisotropic Shale Gas Formations


Saied Mighani*, Carl Sondergeld and Chandra Rai, Mewbourne School of Petroleum and
Geological Engineering, University of Oklahoma
Copyright 2014, Unconventional Resources Technology Conference (URTeC) DOI 10.15530/urtec-2014-1934272
This paper was prepared for presentation at the Unconventional Resources Technology Conference held in Denver, Colorado, USA, 25-27 August 2014.
The URTeC Technical Program Committee accepted this presentation on the basis of information contained in an abstract submitted by the author(s). The contents of this paper
have not been reviewed by URTeC and URTeC does not warrant the accuracy, reliability, or timeliness of any information herein. All information is the responsibility of, and , is
subject to corrections by the author(s). Any person or entity that relies on any information obtained from this paper does so at their own risk. The information herein does not
necessarily reflect any position of URTeC. Any reproduction, distribution, or storage of any part of this paper without the written consent of URTeC is prohibited.

Summary
Hydraulic fracturing is crucial to geothermal and hydrocarbon recovery. This process creates new fractures and
reactivates existing natural fractures forming a highly conductive Stimulated Reservoir Volume (SRV) around the
borehole. The fracturing process of anisotropic rocks such as shales is examined through this report. We divide the
rock anisotropy into two groups: a) conventional and b) unconventional (shaly) anisotropy. As the first group, we
study two extreme rock types: 1) Lyons sandstone, a brittle, low porosity and permeability, weakly anisotropic
(11%) material and 2) pyrophyllite, a strongly anisotropic (19%) metamorphic rock similar chemically and
mechanically to shale with extremely low porosity and permeability. As the second group, shale samples (18%
anisotropy) from Wolfcamp formation are studied. The calcite filled veins are observed to be mostly subparallel to
the fabric direction. Brazilian tests are carried out to observe the fracture initiation and propagation under tension.
Strain gauges and Acoustic Emission (AE) sensors record the deformation leading to and during failure. SEM
imaging and surface profilometry are employed to study the post-failure fracture system and failed surface topology.
Fracture permeability is measured as a function of effective stress. The effect of anisotropy on fracturing is also
investigated by rotating the fabric direction of the sample disks relative to the loading axis through increments of 15
degrees.
The rock microstructure, lamination, and brittleness control the activation of the layers. Lyons sandstone shows a
wide brittle fracture with larger process zone with twice as much layer activation at lower stress levels. The
fracturing process in shale is however a coupled function of rock fabric and calcite veins. The veins easily activate at
15 degrees orientation with respect to the loading axis at 30% of the original failure load. The resulting unpropped
fracture has enhanced permeability by orders of magnitude. The findings of this research bring new insights toward
an economic and efficient completion design. Fracking through a deviated well reduces the breakdown pressure
significantly and activates a large number of veins with enhanced conductivity without the need for proppant
injection.
1. Introduction
During the hydraulic fracturing job in a shale gas reservoir, the induced fracture reactivates the present natural
fractures forming a Stimulated Reservoir Volume (SRV) region around the borehole (Medeiros et al,. 2007). This
volume is the major contribution for production from these low-permeability reservoirs (Mayerhofer et al., 2010).
Natural fractures present as calcite filled veins embedded through the rock matrix. This makes a complex fracture
network which is not an easy task to predict. Clay minerals also present as one of the main constituent of the shale
gas reservoirs matrix which could be as high as 70 %. The platy nature of these minerals results in anisotropy which
makes the fracture even more complex. During this paper, we explain how the presence of natural fractures along
with rock anisotropy impacts the tensile fracture phenomenon in shale rocks. This paper extends the previous efforts
of the authors (Mighani et al., 2013) to shale rocks with detailed description. The results of this report help us better
understand the tensile fracture interactions with rock layers and present natural fractures. They also give insights
regarding the tensile strength variations with respect to rock anisotropy.

2
URTeC 1934272
The laboratory measurements show the significance of tensile strength for breakdown pressure calculations. In other
words, it is not an insignificant variable for breakdown pressure calculations. In Figure 1 we plot the breakdown
pressure vs. tensile strength values drawn from experimental observations of Damani (2013), Chitarala (2011), and
Schmitt and Zoback (1992).
7000

2:1

1:1

Breakdown pressure, psi

6000
5000
4000
3000

Limestone, Uniaxial hydraulic fracturing **


pyrophyllite, Uniaxial hydraulic fracturing **
pyrophyllite, Triaxial hydraulic fracturing *

2000

Lyons sandstone, Uniaxial hydraulic fracturing**

Lyons sandstone, Triaxial hydraulic fracturing *

1000

Westerly granite, Unjacketed hollow cylinder ***


Westerly granite, Jacketed hollow cylinder ***

0
0

1000

2000

3000

4000

5000

6000

7000

Tensile strength, psi


Figure 1: Breakdown pressure vs. tensile strength in laboratory scale experiments. The tensile strength of the rock is not insignificant for
breakdown pressure calculations. (The experimental results are obtained from *Damani (2013), **Chitarala (2011), and ***Schmitt and Zoback
(1992)).

Therefore, an attempt for finding the minimum value of the tensile strength could be helpful for an efficient drilling
and completion job. In this paper, first we explain the experimental procedure we have selected for this study. The
samples are then described based on their properties. The observations of the experiment are then demonstrated and
analyzed. Finally, we conclude the takeaways of the experiments for field scale operation.
2. Experimental Procedure
To investigate and analyze the mechanism of tensile failure, the indirect tensile strength measurement or Brazilian
test is employed. Brazilian test, proposed by Carneiro (1943), is currently the standard method for indirect
measurement of the tensile strength. The facilitated sample preparation and test procedure are the main advantages
of this method. It includes the compression of a cylindrical disk via two steel jaws. The compression induces a
maximum horizontal tension at the center of the disk. The disk fails when this tension exceeds its intrinsic tensile
strength resulting in a vertical fracture. The details of the stress distribution in this test are elaborated by Hondros
(1959). The tested samples in our study are cylinders with 1.5 inch diameter and 0.75 inch length. The thickness to
diameter ratio is then 0.5 according to ISRM standards (ISRM, 1978). The disks are loaded with a constant load rate
of 270 lb/min. The constant load rate test simulates the actual hydraulic fracturing process in which the fluid is
injected under constant flow rate conditions. For preparation of the shale samples, special consideration is followed.
Shale samples are drilled and cut using air. There are no observable cracks in the samples during and after the
preparation. Vertical samples are used to analyze the impact of rock properties on microscale behavior of the tensile
fracture. For the vertical samples, strain gauges are mounted on the disk surface to measure both axial and lateral
strain simultaneously. The tolerance and resistance for these series of strain gauge are +/- 0.40 % and 300 ohms
respectively. One Acoustic Emission (AE) piezoelectric sensors is also mounted on the other surface of the disk to
listen to the micro cracks forming during the failure process. Horizontal samples are then used to investigate the
impact of anisotropy on the tensile failure. Anisotropy direction is determined measuring the compressional velocity
across the disk which is explained in the next section. The fabric direction of the disks is rotated relative to the
vertical axis with increments of 15 degrees. The rock failure for these samples is being recorded using a high speed

3
URTeC 1934272

Frame force, kip

camera. Figure 2 depicts an example of the disk of rock placed into the steel jaws. The anisotropy direction is drawn
on the surface with a pencil mark as shown in Figure 2. The sample is loaded with the defined loading rate until the
failure. The failure load is detected from the sharp drop in the continuously recorded frame force as shown in Figure
2.
2

Failure

1.5
1
0.5
0
0

100

200

300

400

500

Bedding orientation (HorTime, seconds


45)
Figure 2: The configuration of Brazilian test experiment. In this figure, Lyons sandstone is placed under the Brazilian test steel jaws. The bedding
orientation is illustrated as the line drawn on the surface of the disk. Hor-45 is a horizontal sample with a bedding orientation of 45 degrees
relative to the loading axis. The failure load is detected from the sharp drop in the frame force as shown in the right picture.

The tensile strength can then be calculated using the ISRM standard equation (ISRM, 1978):
(1)
Where P is the load at failure, is the half of the contact angle between disk specimen and steel jaws, R is the disk
radius and t is the disk thickness. The contact angle between the disk and jaws is 35 degrees.
3. Sample Characterization
The candidate rocks are evaluated based on their microstructure and petrophysical properties. This helps for better
interpretation of pre-during-post failure mechanism of the tensile fracture. Several measurements evaluate these
rocks including: porosity and permeability measurements, Fourier Transform-Infrared (FT-IR) mineralogy, thin
section, and SEM analysis. We divide the rock anisotropy into two distinct groups: a) conventional and b)
unconventional (shaly) anisotropy. Here, we describe the properties of candidate rocks we selected for these two
groups.
3.1. Conventional Anisotropy
Conventional anisotropy group includes the layered rocks in which the anisotropy is a consequence of mineral
alignment. Lyons sandstone and pyrophyllite are the two rocktypes studied as the first group. Table 1 describes these
rocks based on their measured properties. Figure 3 also depicts the thin section analysis of these rocks. The
sandstone contains a high concentration of quartz minerals. Pyrophyllite (metamorphic rock) consists of Aluminum
Silica minerals. The grain size in this rock is less than 5 micrometers. The grains in Lyons sandstone are about 190
micrometers. The mineralogy analysis suggests the sandstone to be more brittle compared to pyrophyllite. Uniaxial
compression results also show a higher Youngs modulus and lesser Poissons ratio for Lyons sandstone. The higher
ratio of Youngs modulus to Poissons ratio suggests the Lyons sandstone to have a higher index of fracability.
Lyons sandstone

pyrophyllite

1mm

1mm

Figure 3: Thin section analysis of Lyons sandstone and pyrophyllite. The left picture depicts the Lyons sandstone grains with an average
dimension of 190 m. The right picture depicts the pyrophyllite grains with an average dimension of less than 5 m (Damani, 2013).

4
URTeC 1934272
Rock type
Character
Mineralogy (wt %)

Table 1. Characterization of different rock types.


Lyons sandstone
Quartz (85), mixed clay (10), Dolomite (5)

Pyrophyllite

Aluminum silicate hydroxide

Porosity (%)
8
5
Permeability (nD)
15e3
100
Grain size (m)
190
<5
Uniaxial compression results
UCS (Mpsi)
UCSh =14, UCSv =12.6
UCSh =15.7, UCSv =13.9
E (Mpsi),
Eh=5.2, Ev=3.6, h=0.14, v=0.14
Eh=4.8, Ev=2.2, h=0.22, v=0.23
Anisotropy (h/v)
UCS=1.11, E=1.4
UCS=1.13, E=2.2
*
UCS: Uniaxial compressive strength, E: Youngs modulus, : Poissons ratio, h: horizontal, v: vertical

To investigate the degree of elastic anisotropy, the compressional velocity variation around the diameter of a
cylindrical sample is measured. This method is called Circumferential Velocity Analysis (CVA). These sets of
measurements are conducted on both horizontal and vertical samples cored from the same cubic block. Figure 4
shows the variation of the velocity values as a function of rotational azimuth for Lyons sandstone and pyrophyllite.
The measured anisotropy on horizontal samples illustrates 11% anisotropy for Lyons sandstone and 19% for
pyrophyllite. The low degree of anisotropy (less than 5%) observed for the vertical core ensures the Transverse
Isotropy (TI) as an appropriate assumption for both rock types. Revisiting the uniaxial compression results presented
in Table 1, pyrophyllite shows also higher strength and stiffness anisotropy which is in accordance with observations
from the velocity anisotropy analysis.

Lyons sandstone

pyrophyllite

Figure 4: Circumferential Velocity Analysis (CVA) results for Lyons sandstone (left) and pyrophyllite (right). Lyons sandstone shows a weak
anisotropy (11%). Pyrophyllite shows a strong anisotropy (19%). The Transverse Isotropy (TI) is an appropriate assumption for both rock types.

X-ray Computed Tomography (CT) helps us look into the layers of these rocks. The scanned images for these two
rocktypes are demonstrated in Figure 5. The resolution for the micro CT scan is 1 m. The rock fabrics for the
sandstone are visible through micro CT. On the other hand, pyrophyllite does not exhibit any visible mineral
alignment. Therefore, the fabrics for pyrophyllite are submicron.

Figure 5: Micro CT scan of the Lyons sandstone (left) and pyrophyllite (right). The micro-CT with a resolution of 1 m reveals the fabrics for
Lyons sandstone; yet unable to detect the submicron fabrics for pyrophyllite.

Based on the mentioned observations, the nature of anisotropy for Lyons sandstone is due to the depositional
history. There is a friction resistance between these layers. On the other side, the nature of anisotropy for
pyrophyllite is due to the platy nature of pyrophyllite clays. The elastic anisotropy here is complex and strong.
Revisiting the CVA measurements on vertical samples for pyrophyllite there is a weak sinusoidal velocity variation

5
URTeC 1934272
around the sample. This indicates a probable weak orthorhombic anisotropy resulting from the clay complexity. The
pyrophyllite is similar chemically and mechanically to shale.
3.2. Unconventional (Shaly) Anisotropy
The distinguishing characteristic of this group is the existence of calcite filled veins. It brings more complexity to
the anisotropic behavior of layered shales. In this study, the experiments are focused on shale samples from
Wolfcamp formation. Figure 6 depicts an SEM example of these veins present in the Wolfcamp shale formation.
The veins are filled mostly with calcite. They are weakly bonded to the matrix and potential to activate under stress.

Ca Si Al

170 m

500 m

Easy to activate

100 m

500 m

Figure 6: SEM image of a calcite filled vein in Wolfcamp shale formation. The vein filler is mostly calcite. The weak bond between the calcite
layer and matrix is potentially easy to activate.

CVA measurements around horizontal and vertical samples from the Wolfcamp formation are shown in Figure 7.
The constant circumferential velocity for the vertical sample tells the TI anisotropy is an appropriate assumption for
this shale. For the horizontal samples, the transmitted compressional wave cannot be picked (dissipates) for angles
between 60-120 and 240-300 degrees with respect to the bedding orientation. The reason is the presence of calcite
filled veins. Figure 8 illustrates the samples studied in this work. The bedding orientation shown in the figure is
determined using CVA analysis. It is observed in this figure that most of the veins lay mostly subparallel to the
bedding orientation. For the vertical sample, there is no observable calcite vein.

Unable to pick

Anisotropy: 18 %

Unable to pick
Horizontal

Anisotropy: 1 %

Vertical

Figure 7: CVA analysis of the Wolfcamp shale samples. For the horizontal samples, the transmitted compressional wave dissipates in the
direction perpendicular to the bedding orientation intersecting the maximum number of veins.

Therefore, the transmitted compressional wave dissipates when it is mostly perpendicular to the bedding orientation.
This is where it intersects the maximum number of subparallel veins. It can be then rigorously concluded that the
veins are mostly subparallel to the bedding orientation.

6
URTeC 1934272

Bedding orientation

Hor-0

Hor-15

Hor-30

Hor-45

Hor-60

Hor-75

Hor-90

Vert

Figure 8: Wolfcamp shale formation samples. The white line indicates the bedding orientation determined from CVA analysis. The veins appear
to be mostly subparallel to the bedding orientation. The vertical sample does not show any visible vein.

4. Report on Experiment Results


4.1. Conventional Anisotropy
4.1.1. Vertical Disks
The vertical disks in which the fabrics are parallel to the disk surface act isotropic under the Brazilian test loading
condition. They carry information on the isotropic behavior of rock in the absence of anisotropic layers.

200
150
100
50

Lyons sandstone

200
180
160
140
120
100
80
60
40
20
0

0
0

0.5

1.5

2.5

3.5

5,000

5,000

4,000

4,000

3,000

3,000
Micro crack coalescence

2,000
1,000

1,000

4.5

0
0.0

0.5

1.0

1.5

Load (kip)

Load (kip)

1.5
1
0.5
0
0.000

0.001

0.002

-0.0020

-0.0010

Strain (in/in)
Axial

Lateral

2.5

3.0

3.5

4.0

4.5

pyrophyllite

Lyons sandstone

-0.001

2.0

Load, kip

Load, kip

-0.002

2,000

pyrophyllite

Cumulative number of AE

pyrophyllite

Lyons sandstone

Number of AE

4
3.5
3
2.5
2
1.5
1
0.5
0
0.0000

0.0010

0.0020

Strain (in/in)
Axial

Lateral

Figure 9: The acoustic emission and strain response of the Brazilian test on vertical samples of Lyons sandstone and pyrophyllite.

Therefore, they can be useful for analysis of the impact of rock properties on microscale observations of the tensile
fracture. The strain gauges and also Acoustic Emission (AE) sensors mounted on the sample explain the differences
in the damage accumulation mechanism before the main tensile failure. The linear trend for the axial strain behavior
indicates an elastic behavior until failure. The non-linear strain trend in pyrophyllite is however an indication of

7
URTeC 1934272
ductile (plastic) behavior. Based on the strain behavior, a brittle fracture with a large process zone is expected for the
sandstone. On the other hand, the failure for pyrophyllite is expected to be ductile with a narrow process zone. In
accordance with the strain behavior, Figure 9 shows comparably higher number of acoustic events for Lyons
sandstone compared with the one for pyrophyllite. The cumulative number of AE events states also a basic
difference between the fracturing processes of these two rocks. The plot shows a change is slope, a point at which
the micro cracks start to connect to make the macro scale failure. The pyrophyllite however flows with a plastic
deformation until the failure. The evidences of AE and strain behavior findings can be observed in the process zone
analysis around the main tensile fracture. Figure 10 shows the process zone and the microcracks generated around
the main tensile fracture during its propagation. The fracture path for pyrophyllite is not straight. It seems to contain
a wavy path with a particular frequency. The tensile fracture is clean with few number of microcracks and narrow
process zone. Even though this rock breaks at higher level of stress (higher energy) the resulted fracture is narrow
with low process zone. On the other hand, the fracture path in Lyons sandstone is close to a straight line with high
process zone and large number of grains crushed due to high energy content of fracture propagation. The tensile
fracture for Lyons sandstone is two times wider. The process zone is also eight times larger for the brittle rock.
Lyons sandstone - Average fracture width: 75 m, Tensile strength: 1350 psi, process zone: 400 m

Intact rock

Micro crack

Propagation

Crushed grains

Micro crack

pyrophyllite - Average fracture width: 40 m, Tensile strength: 2280 psi, process zone: 50 m
Figure 10: The fracture process zone around the main fracture. The tensile fracture for Lyons sandstone is two times wider. An extensive process
zone is also observed for Lyons sandstone eight times larger than the one for pyrophyllite.

The topography of the resulted fractures is also shown in Figure 11. The topography is measured on a 5 5 mm
section of the fracture surface using NANOVEA ST-400 optical profilometer. The root mean square roughness of
the surface is also calculated. The Mode I or opening mode fracture has a high roughness. It also should not include
a high amount of grains broken in the surface. The Mode II or shear mode fracture however has a low value of
roughness and high number of broken grains on the surface. Therefore, a higher value of roughness is an indicator of
pure tensile failure. It is worth note that this value depends highly on the grain size of the rock. In theory, the pure
tensile fracture should have a roughness value close to the average size of the grains. The measured roughness for
the Lyons sandstone shows a roughness value 5 times greater than the one for pyrophyllite. The fracture
permeability measurement as a function of confining pressure tells us some information regarding the fracture
conductivity after the fracture asperities come together due to the overburden. The measurement results are shown in
Figure 12.

8
URTeC 1934272
mm
0.7

pyrophyllite
Rq (m): 4.8

0.6

Lyons sandstone
Rq (m): 30.6

0.5
0.4
0.3
0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7

mm
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7

Figure 11: Topography of the tensile fracture surface. Lyons sandstone illustrates a fracture surface with higher roughness value compared with
pyrophyllite.

10

0.1

Native Lyons sandstone


Fractured Lyons sandstone

0.01
0.001

0.0001

Permeability, md

10

Dislodged grain

Permeability, md

The fracture permeability measurements show enhanced permeability by orders of magnitude. This permeability is
still maintained under high confining pressure. The reason could be the placement of the dislodged rock particle
during the fracturing on the path of the fracture. Therefore, the fracture does not close and the particle play the role
as a natural proppant. This permeability depends on the breaking pattern of the rock. So, a higher drop in the
permeability is observed for pyrophyllite.
Native pyrophyllite
Fractured pyrophyllite
0.1
0.01
0.001
0.0001
0

1000

2000

3000

4000

Confining pressure, psi

5000

1000

2000

3000

4000

5000

Confining pressure, psi

Figure 12: Fracture permeability measurements under confinement. The fracture shows enhanced permeability in both rocks by orders of
magnitude. The reason for this is the emplacement of dislodged grains as a natural proppant on the path of fracture. Lyons sandstone shows less
reduction with increase in confining pressure.

4.1.2. Horizontal Disks


The Horizontal samples are investigated to evaluate the impact of anisotropy on the tensile failure. Anisotropy
direction is determined measuring the compressional velocity across the disk as explained in the previous section.
The fabric direction of the disks is rotated relative to the vertical axis with increments of 15 degrees. In the presence
of rock fabric, the tensile failure turns into a combination of different modes of fracture. The fracture initiation and
propagation process is recorded using a high speed camera. The impact of anisotropy on the tensile failure is
obvious in Figure 13. For the Berea sandstone which is an isotropic rock, the fracture initiates at the center of the
disk and propagates vertically toward the loading points. However, for an anisotropic rock, i.e. Lyons sandstone, the
fracture can have different patterns depending on the orientation of the rock fabric. The modes of fractures can be
divided into two distinct mode. The first mode is the Mode I or opening mode fracture. This fracture initiates from
the center of the disk and propagates vertically toward the loading points. On the other hand, Mode II fracture is a
shear fracture resulting from the rock fabric activation. This mode of fracture results from the sliding of weakness
bedding planes. Figure 14 shows these modes of fracture. It also shows how the two faces of fracture dislocate with
respect to each other for these modes of fracture. The observed vertical fracture is a Mode I fracture provided it
happens in the disk center within 10 % of the disk diameter length.

9
URTeC 1934272

Figure 13: High speed photography of the fracture initiation and propagation for different rocks. The isotropic rock (Berea sandstone) fails only
on the vertical diameter which is a tensile fracture. The anisotropic rock (Lyons sandstone) however can have different failure patterns depending
on the anisotropy direction. The anisotropy direction for Lyons sandstone is drawn with pencil mark.

Tavallali et al., (2007) categorized different modes of fracture in the same manner under the Brazilian test. He
named the Mode I and Mode II fracture as central fracture and layer activation respectively.

Mode I

Mode II

Figure 14: Different modes of fracture. Mode I fracture (opening mode) initiates from the center of the disk and propagates vertically toward the
loading points. Mode II fracture is a shear fracture resulting from the bedding activation.

The tests are carried out on both rock types which are depicted in Figure 15. It can be observed that the fracture
pattern is a function of bedding orientation.

Figure 15: High speed camera images for Lyons sandstone and pyrophyllite disks after failure with different orientation of bedding. The
mechanism of failure is a function of bedding orientation.

10
URTeC 1934272
Now, we categorize the observed fractures based on the fracture modes we explained earlier. Figure 16 shows the
modes of fracture obtained from the high speed camera images (Figure 15).

pyrophyllite

Lyons sandstone

Figure 16: Fracture mode categorization using the high speed camera images for Lyons sandstone and pyrophyllite Brazilian test samples after
failure at different bedding orientations.

The tensile strength variation with respect to the orientation of beddings for these rock types is also illustrated in
Figure 17. In a vertical sample where the beddings are parallel to the disk surface, the rock acts isotropic under the
diametric loading. Hence, the highest strength is observed for vertical sample. For horizontal samples, a similar
variation in the tensile strength is observed as a function of bedding orientation. The maximum tensile strength
occurs at 75 degrees; while the minimum is at 15 degrees. The observed strength anisotropy (maximum to minimum
strength ratio) for horizontal samples is about 1.7 for Lyons sandstone and 1.5 for pyrophyllite. Even though the
elastic anisotropy for pyrophyllite is stronger, the tensile strength anisotropy is stronger for Lyons sandstone. So,
there is a fundamental difference between the anisotropy behavior of these two phenomena. If we revisit the Figure
15, the failure at 15 degrees is almost a pure shear failure. This angle seems to be the critical orientation for the
activation of the beddings. This is where the minimum tensile strength is also observed.
2500
2300

pyrophyllite:

)= 1.5

2100

Tensile strength, psi

1900
1700
1500

Lyons sandstone:

1300

)= 1.7

1100
900
700
500
-15

Vertical

15

30

45

60

75

90

Angle, Degrees
Horizontal

Figure 17: Tensile strength variation of different fabric orientation with respect to load for Lyons sandstone and pyrophyllite. Vertical sample has
the maximum strength. The minimum strength occurs at 15 degrees.

We can also quantify the length for different fracture modes by counting the fractures in Figure 16. Figure 18 shows
the length variation of fracture modes with respect to bedding orientation for both rock types. The layer activation or
Mode II fracture can be attributed to be a layered rock property. In contrast to the other properties of rock such as
fracture toughness, this is a direction dependent property (Tavallali and Vervoort, 2010, He and Hutchinson 1989).
It seems that the Mode II fracture is easy to open for orientations below 30 degrees. So, it is predominant for these
orientations. At 15 degrees where we observe the minimum tensile strength, the Mode II fracture length has its

11
URTeC 1934272
maximum value. The observed activated layers for Lyons sandstone is in general two times at lower stress levels
compared with pyrophyllite. Also, for the orientations over 45 degrees the tensile fracture propagates with the
minimum number of activated layers.
Mode II

Mode I

Figure 18: The Mode I and Mode II fracture modes length variation with respect to bedding orientation for Lyons sandstone and pyrophyllite. At
15 degrees where we have minimum tensile strength, the Mode II fracture length has its maximum value.

Figure 19 shows the surface profile of the Mode I and Mode II fractures for these rock types. The Mode I fracture
for pyrophyllite with smaller grains is smoother and close to Mode II fracture. On the other hand, Lyons sandstone
with larger grain size has a Mode I fracture with more surface roughness (six times larger). The size of process zone
for Lyons sandstone was also earlier observed to be close to eight times larger.
Lyons sandstone

pyrophyllite
mm

Mode I
Rq (m): 30.6

Mode I
Rq (m): 4.8

Mode II
Rq (m): 5.9

Mode II
Rq (m): 5.8

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7

Figure 19: Fracture surface roughness comparison for Lyons sandstone and pyrophyllite. Red represents high surface and blue represents low
surface. Lyons sandstone with larger grain size shows a fracture with more surface roughness.

4.1.3. Discussion on the Observations


The candidate rocks represent the extremes of rocks properties, i.e. brittleness, connectivity, grain size, mineralogy,
and anisotropy characteristics. Based on the observations, the tensile fracture propagation is highly dependent on the
rock fabric characteristics and orientation. This resembles the hydraulic fracturing propagation within an anisotropic
formation. For orientation below 30 degrees, the tensile fracture activates a large number of layers with less amount
of needed stress. For orientations after 45 degrees, the tensile fracture propagates without activating the layers. The
needed stress to propagate this fracture is also high. The fracture roughness, permeability, width, and also process
zone analysis suggest comparably higher fracture efficiency for the brittle rock. Therefore, for an efficient fracturing
job, we definitely need a high amount of layer activations. Otherwise, the fracturing job is not successful (The
observed process zone for pyrophyllite). In the next section, we observe the role of veins in the shale tensile failure
mechanism.

12
URTeC 1934272

4.2. Unconventional (Shaly) Anisotropy


As mentioned and explained in the previous sections, the main difference between the anisotropy behavior for shale
is the presence of calcite veins. Therefore, the veins can be activated similar to rock beddings under critical
circumstances. Using the explained procedure, we have conducted the Brazilian tests on the Wolfcamp shales shown
in Figure 8. Figure 20 shows the failed samples with different orientations of bedding.

Figure 20: High speed camera images for Wolfcamp shale samples after failure with different orientation of bedding. The calcite filled veins open
easily for bedding orientations under 30 degrees before the central fracture.

The calcite filled veins activate easily for loading directions under 30 degrees before the central fracture. This
critical orientation is similar to the critical orientation for the rock bedding as observed for the conventional
anisotropy group rocks. The variation of tensile strength values with different orientations of bedding is shown in
Figure 21. It can be observed here again that the minimum tensile strength occurs at 15 degrees fabric orientation.

Strength anisotropy:

) = 2.8

Figure 21: Tensile strength variation for different bedding orientations with respect to vertical axis for Wolfcamp shale. The minimum strength
occurs at 15 degrees.

Meticulous comparison between the conventional and shaly anisotropy behavior reveals an interesting trend for
shale rock anisotropic behavior under tensile failure. The comparison can be seen in Figure 22. The elastic
anisotropy for shale is close to the one for pyrophyllite. The tensile strength anisotropy is however two times larger
for shale. A hypothetical trend curve is fitted on the strength values. The general anisotropic trend is similar.
However, due to the presence of calcite filled veins, a shift in the strength of the rock can be observed. This shift is
intensified for orientations below 30 degrees. This explains again the veins weakening impact for shale especially
for orientations below 30 degrees. This weakening effect is dependent on the number of the veins, their aperture, and
their orientation (In this report they are parallel to the bedding orientation).

13
URTeC 1934272
2500

Tensile strength, psi

2300

pyrophyllite:

Strength anisotropy:

)= 1.5

) = 2.8

2100
1900
1700
1500

Lyons sandstone:

)= 1.7

1300
1100
900
700
500
0

15

30

45

60

75

90

Angle, Degrees
Figure 22: The shifts in anisotropic strength values due to vein presence. Veins have a weakening impact on the anisotropic tensile strength for
shales.

4.2.1. Discussion on the Observations


During the hydraulic fracturing job, the hydraulic fracture reactivates and opens up the calcite filled veins. The
roughness measurement results show resemblance between the tensile fracture and activated vein (Figure 23, these
values are in the order of measured roughness values for pyrophyllite). Therefore, the activated vein can be as
important as a hydraulic fracture. As discussed also in the previous sections, the opened fracture has enhanced
permeability due to naturally formed proppants. Another efficacy is the misregistration of asperity of fractures.
Kassis and Sondergeld (2010) showed an increase by three orders of magnitude by only 0.02 inches asperities
misregistration.

Figure 23: Fracture surface roughness measurement results show resemblance between the tensile fracture and activated vein.

The fracturing in a deviated well with 15 degrees inclination resembles the Brazilian test on a horizontal disk with
bedding orientation of 15 degrees. The tensile strength values show a reduction by amount of 70 % which could be
then attributed to a significant reduction in breakdown pressure. A high number of veins can also activate at this
inclination. Therefore, fracturing through a deviated well could be an efficient option for completion operations in
shale reservoirs.
Conclusions
The tensile strength is not an insignificant variable for breakdown pressure designs. The Brazilian test on the
candidate anisotropic rocks including shales reveal important findings for an efficient completion in shale reservoirs.
1- Fracturing through a deviated well could be an efficient option for completion operations in shale
reservoirs. It can reduce the breakdown pressure significantly and activated a high number of veins.
2- The unpropped fracture has enhanced permeability by orders of magnitude. Therefore, the amount of
proppants currently used in field operations might be more than necessary.

14
URTeC 1934272
Acknowledgement
The authors would like to acknowledge Unconventional Shale Gas Consortium for supporting this research. Also,
Gary Stowe, Bruce Spears and Jeremie Jernigen for their laboratory help and support. Mr. Russ Lessmeier from
Vision Research is also acknowledged for his help in acquiring the high speed images.
References
Carneiro, F. L.L.B. 1943. A New Method to Determine the Tensile Strength of Concrete. Proceedings of the Fifth
Meeting of the Brazilian Association for Technical Rules. Third Section. ADINA R&D.
Chitrala, Y. 2011. Laboratory Study of Fluid Induced Hydraulic Fractures Hypocenter Locations, Source
Mechanism, Frequency Analysis and Microscopic Observations. MS thesis, The University of Oklahoma,
Norman, Oklahoma.
Damani, A. 2013. Acoustic Mapping and Fractography of Laboratory Induced Hydraulic Fractures. MS Thesis, The
University of Oklahoma, Norman, Oklahoma.
He, M. Y., and Hutchinson, J.W. 1989. Crack Deflection at an Interface between Dissimilar Elastic
Materials. International Journal of Solids And Structures. 25(9): 1053-1067.
Hondros, G. 1959. The Evaluation of Poissons Ratio and the Modulus of Materials Of a Low Tensile Resistance by
the Brazilian (Indirect Tensile) Test with Particular Reference to Concrete. Aust J Appl Sci 10(3):243-268.
ISRM. 1978. Suggested Methods for Determining Tensile Strength of Rock Materials. International Journal of Rock
Mechanics and Mining Sciences and Geomechanics Abstracts. 15: 99-103.
Kassis, S., and Sondergeld, C. H. 2010. Fracture Permeability of Gas Shale: Effects of Roughness, Fracture Offset,
Proppant, and Effective Stress. Society of Petroleum Engineers Journal 131376: 1-17.
Mayerhofer, M.J. Lolon, E.P., Warpinski, N.R. Cipolla, C.L., Walser, D. and Rightmire, C.M. 2010. What is
Stimulated Reservoir Volume?. SPE Production & Operations. 25(1): 89-98.
Medeiros, F., Kurtoglu, B. Ozkan, E. and Kazemi, H. 2007. Pressure-Transient Performances of Hydraulically
Fractured Horizontal Wells in Locally and Globally Naturally Fractured Formations. SPE paper 11781-MS
presented at International Petroleum Technology Conference, Dubai, U.A.E, December 4-6 December.
Mighani, S., Sondergeld, C., and Rai, C. 2013. Influence of Rock Layering on Fracturing Mechanism under Tensile
Failure. SEG Technical Program Expanded Abstracts 2013: pp. 2922-2926. doi: 10.1190/segam2013-1368.1
Schmitt, Douglas R., and Mark D. Zoback. 1992. Diminished Pore Pressure in Low Porosity Crystalline Rock under
Tensional Failure: Apparent Strengthening by Dilatancy. Journal of Geophysical Research: Solid Earth 97.B1:
273-288.
Tavallali, A., B. Debecker, and A. Vervoort. 2007. Evaluation of Brazilian Tensile Strength in Transversely
Isotropic Sandstone. 11th congress of the international society for rock mechanics, Proceedings, 269-272.
Tavallali, A., and Vervoort, A. 2010. Failure of Layered Sandstone under Brazilian Test Conditions: Effect of
Micro-Scale Parameters on Macro-Scale Behavior. Rock mechanics and rock engineering 43.5: 641-653.s.

Anda mungkin juga menyukai