Anda di halaman 1dari 29

SPE-184852-MS

A Model for the Conductivity and Compliance of Unpropped Fractures

Weiwei Wu 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
Fluid flow in unpropped and natural fractures is critical in many geophysical processes and engineering
applications. The flow conductivity in these fractures depends on their closure under stress, which is a
complicated mechanical process that is chanllenging to model. The chanllenges come from the deformation
interaction and the close coupling among the fracture geometry, pressure and deformation, making the
closure computationally expensive to describe. Hence, most of the previous models either use a small grid
system or disregard deformation interaction or plastic deformation.
In this study, a numerical model is developed to simulate the stress-driven closure and the conductivity for
fractures with rough surfaces. The model integrates elastoplastic deformation and deformation interaction,
and can handle contact between heterogeneous surfaces. Computation is optimized and accelterated by
employing an algorithm that combines the conjugate-gradient method and the fast Fourier transfrom
technique. Computation time is significantly reduced compared to traditional methods. For example a 5
orders of magnitude speedup is obtained for a grid size of 512 by 512. The model is validated against
analytical problems and experiments, for both elastic-only and elastoplastic scenarios.
It is shown that interaction between asperities and plastic deformation can not be ignored when modeling
fracture closure. By applying our model, roughness and yield stress are found to have a larger impact
on fracture closure and compliance than Young's modulus. Plastic deformation is a dominant contributor
to closure and can make up more than 70% of the total closure in some shales. The plastic deformation
also significantly alters the relationship between fracture stiffness and conductivity. Surfaces with reduced
correlation length produce greater conductivity due to their larger apertures, despite more fracture closure.
They have a similar fraction of area in contact as surfaces with longer fracture length, but the pattern of area
in contact is more scattered. Contact between heterogeneous surfaces leads to increased plastic deformation
and fracture closure, and results in lower fracture conductivity. Fracture compliance appears not be sensitive
to the distribution pattern of hard and soft components. Our model compares well with experimental data
for fracture closure, and can be applied to unpropped or natural fractures.

Introduction
Fractures provide effective conductive pathways for fluid flow in rocks, and they play critical roles in many
subsurface applications such as hydraulic fracturing, production through naturally factured reserovirs, CO2
2 SPE-184852-MS

sequenstration and extraction /protection of acquifers. In hydraulic fracturing, abundant unpropped fractures
with enormous surface area are created, and their importance and potential in completion and production
through their flow capability is being recognized and explored (Sharma and Manchanda 2015).
Fluid flow through fractures depends on the fracture geometry, which is the net result from fracture
surface topography, closure stress, and surface deformation. One way to model fracture closure is by
stochastic methods, in which the rough fracture surface profile is approximated by a particular probability
distribution. Vairous disitrbutions have been used. They include Gaussian distribution (Greenwood and
Williamson 1966; Walsh and Grosenbaugh 1979; Swan 1983), power-law distribution (Gangi 1978; Gong
et al 1999), exponential distribution (Brown and Scholz 1985), inverted chi-square distribuition (Brown and
Scholz 1985), normal and lognormal distribution (Tsang and Tsang 1987).
The analytical equations developed from the stochastic methods enable quick estimates of the
mean contact pressure, fracture aperture, contact area and flow conductivity. Unfortunately, several
limitations hinder the application of these stochastic methods. Most of them assume that asperities deform
independently and deformation interaction among asperities is ignored. In reality, this interaction is very
important. It is commonly observed due to the long-range nature of the elastic contact, and it intensifies
as more asperities come into contact under an increased load (Berthe and Vergne 1987). The stochastic
methods also fail to provide details such as pressure and aperture distribution andthe exact contact area. In
addition, these methods can be difficult to apply for cases where surface profiles can not be represented by
a certain distribution, such as rough surfaces with channels and pits in irregular dimensions which are often
seen in acid fracturing (Nieto et al. 2008; Wu and Sharma 2016).
As an alternative to stochatisc methods, fracture closure can be modeled numerically, where the closure
is embodied in the deformation of each indidivual grid that the surfaces are discretized into. Deformation
interaction is considered in several numerical models, such as the Hopkins model (Hopkins 1991). In
this model, the fracture surfaces are treated as asperities and half-space underneath, and the deformation
interaction is considered through the deformation of the half-space. The model has been applied to study
fracture flow in serveal rock types (Pyrak-Nolte and Morris 2000; Wang and Cardenas 2016), and it was
also extended to account for plastic deformation (Kamali and Pournik 2016).
However, two challenges impede the computation in most of the currently available numerical models
for rock fracture closure. The first challenge comes from the close coupling among the asperities in contact.
Multi-level iteration (often 3-level) with an initial guess in terms of small increments either in pressure or in
displacement is often adopted (Pyrak-Nolte and Morris 2000; Kamali and Pournik 2016). The increments
often have to be modified by introducing a certain relaxation scheme if convergence can not be achieved.
This frame work has been demonstrated to be computationally inefficient (Polonsky and Keer 1999).
As the deformation interaction occurs globally on the whole surface between any pair of grids, to calculate
the interaction is computationally expensive. The interaction increases the computation demand from order
of O(N) to O(N2), where N is the total number of grids. This increase can be substantial given a large N,
which is often preferred and necessary to describe the contact between two rough surfaces. For example, if
N equals to 512 by 512 (order of 105), N2 would be at order of 1010, 5 orders of magnitude more than N. Due
to this computational challenge, interaction between asperities is ignored (Deng et al. 2011), or considered
locally on the surfaces (Pyrak-Nolte and Morris 2000), or a grid system with small N is used (Kamali and
Pournik 2016; Wang and Cardenas 2016) in some previous models.
Besides the efficiency issue, current aspertity-based models also require asperity height as an important
input to calculate the deformation. To define the asperity height, modelers have to arbitrarily choose a
baseline in surface profile. And this arbitraty decision can potentially affect the computation results and
introduces artificial dependence.
The fracture closure model presented in this study computes deformation with interaction using the
Boussinesq equation to avoid to arbitrarily defining the asperity height. Plastic deformation is also taken
SPE-184852-MS 3

into account. The model incorporates the conjugate-gradident (CG) method and fast Fourier transform
(FFT) technique, which have been widely used in tribology for contact analysis, to optimize the modeling
algorithm and accelerate the computation. The deformation interaction is effectively calculated on the whole
fracture surface discetized with a large grid number. The model also demonstrates the ability to simulate
the closure between surfaces with heterogeneity, which is important for rock such as shale that features
with high heterogeneity. By applying the model, fracture closure is investiaged comprehensively for a wide
range of conditions to understand the effect of roughness, yield stress, Young's modulus and correlation
length in surface profile.

Model Development
Fracture Closure
The rough surfaces of two contacting fractures are discretized in the same uniformly spaced rectangular
grids, with rows and columns aligned with the x and y direction of a Cartesian coordinate system (Fig.1).
The number of the rows and columns in the grids are denoted as Mx and My, and the total grid number
N equals to MxMy. The grid spacing in x and y directions are denoted as ax and ay. An individual grid is
denoted with its row and column number, for example (i, j) or (k, l), where 1 i, k Mx , 1 j, l My.

Figure 1Discretization of composite surfaces in contact.

Figure 2Illustration of contact between surfaces. h is the initial aperture without external load;
g is the aperture after load F applied, which generates deformation u and relative approach .

As the two surfaces come into contact, their topography is assumed to match point to point in the grids,
with the aperture and surface height normal to the grids, in the z direction. The ultimate goal of the model is
to obtain the aperture distribution gij between the composite surfaces under stress, which can be described
as follows (Johnson 1985):
4 SPE-184852-MS

(1)
Where gij is the aperture at grid (i,j) after deformation; hij is the initial aperture when the two surfaces are
just in contact without external pressure; uij is the total deformation inward to the surfaces imposed by the
external pressure; is the relative approach between the two surfaces. The geometric meaning of Eq .1 is
illustrated using a simplified case as shown in Fig.2.
The deformation uij includes elastic and plastic deformation. In our model, the surfaces are assumed first
to reach elastic equilibrium (only elastic deformation occurs). The plastic deformation is then imposed after
elastic equilibrium is reached. It is important to note that the elastic deformation is reversible, and it changes
when the the external load varies; while plastic deformation is permanent, once it happens it can not be
reversed.
This model is a "dry-contact" model, and it ignores shear and the fluid pressure in fractures. It only
considers pressure caused by the external load in a direction normal to the fracture surface. Pressure exists
only in the contacting grids, and is assumed uniform within each grid. The pressure distribution among the
grids satisfies the force balance with the applied external load F:
(2)

(3)

(4)
Although the pressure is only carried by the contacting grids, the elastic deformation caused by the
pressure occurs globally in all the asperities. The elastic deformation at (i,j) is the combined result of the
stress at all the grids. This deformation with interaction is described by the Boussinesq equation, which
takes into account the deformation of both the asperity and the half space (Johnson 1985):

(5)

Where uij_E is the elastic deformation at (i,j); pkl is the pressure at (k,l); A is the computation domain;
Ki-k,j-l is the influence coefficient that describes the deformation of (i, j) caused by a unit pressure acting at
(k, l). It can be found as:

(6)

Where (xi, yj) and (xk, yl) are the coordinates of grid (i, j) and (k, l), respectively. E* is the effective Young's
modulus of the composite surfaces:

(7)

Where E1 and E2 are the Young's moduli of the two surfaces, and v1, v2 are their Poisson's ratio.
After reaching the elastic equilibrium, each grid is evaluated by comparing the stress it receives to
the yield stress Y of the material. For grids with pressure higher than yield stress, plastic deformation is
introduced to further deform these yielded asperities until there are asperities that were previously separate
but are now in contact. Since the surfaces have been modified by the plastic deformation, elastic deformation
has to be recalculated. This process is iteratively repeated until the pressure distribution converges:
SPE-184852-MS 5

(8)

Here 0 is the tolerance to control the convergence, and set as 10-5 in this study unless otherwise specified.
New plastic deformation is only introduced during a loading stage. The overall algorithm to simulate the
fracture closure is summarized in the following flow chart (Fig.3).

Figure 3Algorithm to model fracture closure.

Experiments show that yield stress Y of a material is related to its hardness H measured in the indentation
test by (Fischer-Cripps 2011):
(9)
Where C is a constant ranging from 2.8 to 3.0. In this study, C is assumed 3.0, and hardness is measured
by an indentation test, described in detail in (Wu et al. 2017).
It is important to point out that defining asperity height is not required in our model, neither in the elastic
deformation described by the Boussinesq equation (Eq.5), nor in the plastic deformation. However, it is
necessary in the traditional asperity-based models. In such models an arbitrary baseline has to be selected
in the surface profile to define the asperity heights. This can result in user dependent results. This is not
an issue in our model.

Fracture Conductivity
To calculate the fracture conductivity, the aperture distribution obtained from the contact model above is
applied to construct a 2D fracture network model (Fig.4).
6 SPE-184852-MS

Figure 42D fracture network for calculating fracture conductivity.

Local cubic law is employed to calculate the fluid flow between neighboring grids, and a harmonic mean
of the aperture is used (Brush and Thomson 2003):

(10)

Where qIJ is the volumetric flow between grid I and J3; pI and pJ are the pressure at grid I and J3; WI
3 3

and WJ are their fracture widths; ax and ay are the grid length in x and y direction; is the viscosity of
3

an incompressible fluid, for this case water is used. By solving the mass balance in the fracture network
under assigned constant inlet and outlet pressure (Eq.11), the pressure distribution in the network and the
total volumetric flow rate can be obtained. With that, fracture conductivity between the two surfaces can
be calculated with Darcy equation (Eq.12).

(11)

Where qI is the total volumetric flow rate through grid I.

(12)

Where kfW is the fracture conductivity; qt is the total volumetric flow rate through the fracture network;
L and h are the fracture length and height; p is the pressure drop applied between the inlet and outlet.

Mathematical Implementation of Fracture Closure Model


To overcome the challenge of the close coupling involved in fracture closure and to solve the elastic
contact efficiently, an algorithm based on conjugate-gradient method is used in our model. Eq.1-5 are
mathematically recognized as an optimization problem with Kuhn-Tucker constraints, which can be solved
effectively by CG methods. In fact, the contact process represented by these equations has been proved to
be equivalent to the strain energy minimization process (Ai and Sawamiphakdi 1999):

(13)

Where F is the strain energy function; K is the influence coefficient matrix. In the CG method, the contact
process is solved using just one level of iteration, and convergence is guaranteed with rigorous mathematical
proof. The details for implementing the CG method can be found in a previous study (Polonsky and Keer
1999).
FFT technique is applied to speed up the computation in calculating the deformation with interaction.
The discrete variables in Eq.5 are first converted from space to frequency domain by FFT as follows (Liu
et al. 2000):
SPE-184852-MS 7

(14)

Where pkl is the pressure in frequency domain. Conversion for other variables is done in a similar fashion.
In the frequency domain, the convolution involved in Eq.5 can then be computed directly using arithmetic
multiplication:

(15)

Where ij_E, Kij and pij are the elastic deformation, influence coefficient and pressure in frequency domain.
The deformation in space domain can be eventually obtained by applying inverse fast Fourier transform
(IFFT) to the deformation in frequency domain:

(16)

The advantage of employing FFT is that the number of operations is reduced from O(N2) to O(NlogN).
The computation is mathematically divided into log2N sections of N operations (Press et al. 1996). The
difference brought by FFT can be significant in case of large grid number N. The run time for computing
Eq.5 under different grid numbers by FFT and by direct multiplication is shown in Fig.5. When the grid
number is larger than 256 by 256, the computation time by FFT is five orders of magnitude less than by
direction multiplication. It is due to such a boost in computation efficiency that all the results presented in
this work were completed on grid systems of 128 by 128 or 512 by 512. This is a much higher number of
grids than those used traditionally. All the simulations were conducted on a personal computer.

Figure 5Run time for deformation calculation by FFT and by direct multiplication.

It is worth noting that FFT can not be directly applied to solve the contact problem, since it involves
periodic functions by definition (Eq.14). Additional mathematical treatment including cyclic extension in
convolution, zero padding and wrap-around order is required to prevent error and avoid aliasing in using
the FFT technique. The corresponding mathematical manipulation is detailed by Liu et al. (2000).

Surface Profile
Surface profiles are important inputs in the closure model, and they determine the initial aperture and the
following fracture geometry during the closure process. In analytical models, the surface profile often has
to be approximated or simplified to be explicitly expressed in terms of certain distributions such as a power-
law model (Gangi 1978), or a Gaussian distribution (Greenwood and Williamson 1966). Numerical models,
in the other hand, allow more flexible input of the surface profile, which can either be represented as a certain
distribution, or be directly measured from experiments. As the measured profile is usually at very high
8 SPE-184852-MS

resolution, for example 250,000 data points/ mm2 was used in the profile shown in Fig.6. Appropriate data
reduction has to be performed before the profile can be used in the closure model. In this study, the measured
surface profiles are found to behave very closely to a Gaussian distribution (Fig.7) and demonstrate an
isotropic spatial correlation in x and y direction. To obtain a profile that characterizes the surfaces with a
suitable resolution and fits the capacity of the closure model, synthetic profiles with Gaussian distributions
were generated based on the measured roughness and correlation length using the method developed by
Garcia and Stoll (1984). The synthetic profiles are then used as an input to model the closure process.

Figure 62D contour of surface profile measured through optical profiler.

Figure 7Distribution of the surface profile shown in Fig.6.

The two surfaces of a fracture in theory should couple to each other in geometry when they are in
contact. However, in reality this generally is not the case due to the shear and sliding they experience. In this
study, the average initial aperture between the two initially coupled surfaces is found to increase with the
relative distance they slide for the first three correlation lengths, and then stabilize with a small fluctuation
(Fig.8). This phenomenon is consistent when sliding happens in different directions. The distribution of
initial aperture does not change much when the sliding distance is beyond five correlation lengths (Fig.9).
For all the cases evaluated in this study, the initial apertures are based on fracture surfaces that slide for
10 correlation lengths.
SPE-184852-MS 9

Figure 8Average initial apertures between two contacting surfaces. Rq=20 m.

Figure 9Distribution of initial aperture. Rq=20 m.

Model Validation
Validation of the fracture closure model is conducted in 3 steps. The 1-D constant line load problem is first
used to verify the application of FFT technique in calculating the convolution between influence coefficient
and pressure in Eq.5, which is used to compute the surface deformation under pressure. Then the classic
Hertzian contact between a sphere and a flat surface is used to verify the combination of CG and FFT in
modeling a 2D elastic contact process. Eventually, the model is compared with the fracture conductivity
measured in experiments which involves both elastic and plastic deformation, and the comparison is done
for both loading and unloading stages.

Constant Line Load Problem


When a uniform pressure p is imposed between a to a in a 1D space (Fig.10), the relationship between the
dimensionless displacement and distance is known as the following analytical expression (Johnson 1985):

(17)

Where C is a constant; is the dimensionless distance, ; and z is the dimensionless displacement,


defined as:

(18)
10 SPE-184852-MS

Figure 10Constant line load problem.

Figure 11Hertzian contact between a sphere and a flat plate. F is the external load, a
is the radius of contact zone, is the relative approach between sphere and flat plate.

Alternatively, the displacement can be computed through the Flamant equation:

(19)

Eq. 19 can be rewritten into the discrete convolution form, which is similar to Eq.5:

(20)

Where N the is the total number that the line space is uniformly discretized into; ui is the deformation of
the ith element; Kj-i is the influence coefficient based on the Flamant equation between jth and ith elements;
and pj is the pressure applied on jth element.
Eq.20 is solved with the FFT techniques introduced earlier, and the results are shown to be consistent
with the analytical solution (Fig.10). The relative error is at the order of 10-17 when an element length of
a/32 is used for discretization and the computation domain is set at 12a. The FFT technique is, therefore,
shown to be reliable in solving the discrete convolution involved in the deformation interaction.

Hertzian Contact between Sphere and Flat Plate


After confidence has been established in FFT, the next step is to validate the combination of CG method
and FFT in modeling the 2D contact process. Hertzian elastic contact between a sphere and a flat surface
has been well investigated (Fig.11), and analytical solutions are available for the contact area, pressure and
relative approach with respect to external load applied (See appendix Eq. A-1 to A-5. Johnson 1985).
With the parameters in Table 1, results produced from our closure model based on CG- FFT method
match well to the analytical solutions (Fig.12). As the total grid number increases from 32x32 to 1024x1024,
SPE-184852-MS 11

the relative error of approach and deformation decreases from 10-4 to 10-6, and the relative error of pressure
decreases from 10-2 to 10-4 (Fig.13). The errors are mainly associated with the approximation of a circular
area by rectangular grids. The CG-FFT method is also shown to be highly computationally efficient, it
converges at 64 iterations and takes 45.6 s for the whole computation when using a grid system of 1024x1024
on a personal laptop (dual-core processor of 2.60 GHz, memory of16 GB, Fig.14).
The deformation interaction is also found important in simulating the contact process. Without
considering the interaction, the simulated contact area and approach deviate from the analytical solution,
with relative errors of 99.88% and 42.20%, respectively (Fig.12). Therefore, the interaction has to be
included in the contact model in order to correctly simulate the closure process.

Table 1Input Parameters for Modeling Hertzian Contact

Sphere Radius R 0.01 m


E1, E2 20 GPa
v1, v2 0.25
Tolerance 0 10-6
Computation Domain Twice as the
contact area a2
Subscript 1 Sphere
Subscript 2 Flat plate

Figure 12Comparison between analytical solutions and our model for the
Hertzian contact between a sphere and a flat plate. Gird of 256 x 256 is used.

Figure 13Relative error in modeling Hertzian contact with contact model.


12 SPE-184852-MS

Figure 14Computation efficiency of contact model in computing Hertzian contact.

Model Validation against Experiments


The contact model is further compared with experiments in terms of fracture conductivity. The fracture
conductivity was measured on a polished unpropped preserved Eagle Ford shale core at a series of closure
stresses in loading and unloading stages, where both elastic and plastic deformation occurred (Wu et al.
2017). All of the parameters used for simulation are based on experimental measurements as shown in
Table 2.

Table 2Parameters for Modeling Unpropped Fracture Conductivity

Measurement Simulation
Rq (m) 1.1-3.0* 2.8-3.2
(m) 37- 53* 50
H (MPa) 303-1059** 600-900
E (GPa) 20-80*** 20-60
Grid 128 x 128
ax, ay (m) 10

*: Wu 2017;
**: Wu et al. 2017
***: Sone and Zoback 2013; Zhou 2015

The model is capable of producing results that agree with the experiments with no data fitting parameters
(Table 2). A comparison between the experiment and one of the simulation case is shown in Fig.15. The
agreement between the simulation and experiment suggests that the contact model is able to accurately
describe both of the elastic and plastic deformation involved in the contact process, and in both loading and
unloading stages. It also implies that it is appropriate to use the workflow described earlier to prepare the
input surface profile, as well as the assumed relationship between the yield strength of the surface arperities
and the measured Brinell hardness shown in Eq. 9.
SPE-184852-MS 13

Figure 15Comparison in fracture conductivity between experimental


results and contact model. Rq=2.8 m, H= 900 MPa, E=30 GPa.

In additional to fracture conductivity, the load-displacement curves obtained from the simulation (shown
in a later section) is also in line with experiments in terms of range and shape. For surfaces with roughness
from 5 to 20 m, the simulated fracture closure ranges from 20 to 140 m, and it is consistent with
the laboratory measurement by Bandis et al. (1983). Agreement is also achieved between simulation and
experiment in the range of the fractional contact area. Bandis et al (1983) measured the area for aperture
smaller than 12 m FCA12 between two fractures at 5000 psi closure stress, and it ranged between 40 and
70%. Our simulation also demonstrates similar results, and one example with FCA12 of 46.47% under 5000
psi on fracture surfaces of 8 m in roughness is shown in Fig. 16.

Figure 16Area of aperture smaller than 12 m (in red) at 5000 psi. Rq=8 m, H= 900 MPa, E=40 GPa.

Results and Discussion


Comparison between Elastic-Only and Elastoplastic Model
The elastoplastic closure model developed in this study considers plastic deformation during the contact
process, and it predicts distinct fracture behavior from that predicted by those models that assume that
deformation is only elastic. Many popular models actually fall into this category and they include the
Greenwood and Williamson model (Greenwood and Williamson 1966), the the bed-of-nails model (Gangi
1978), and Hopkins model (Hopkins 1991). In elastoplastic contact, plastic deformation occurs in addition
to the elastic strain. Hence, more closure is caused in elastoplastic contacts than in elastic-only contacts
(Fig.17A). The evolution of aperture in the loading stage of an elastoplastic contact is shown in Fig.18.
14 SPE-184852-MS

Since the plastic deformation is irreversible, displacement hysteresis exists in elastoplastic contact; while
in elastic-only contact, the loading and unloading curves overlap (Fig.17A).

Figure 17Fracture closure behavior predicted by elastoplastic and elastic-only models: (A) load-
displacement; (B) fractional contact area at loading stage. Rq=10 m, Y= 300 MPa, E=30 GPa.

Figure 18Aperture evolution in the loading stage of an elastoplastic


closure. Gray zone indicates area in touch. Rq=10 m, Y= 300 MPa, E=30 GPa.

The elastoplastic model also predicts more fracture surface will come into contact than the elastic
model (Fig.17B). In the given example, the fractional contact area FCA rises to 24.19% as closure stress
increases to 10,000 psi in the elastoplastic model, whereas it only increases slightly to 2.05% in the elastic
model. The corresponding contact maps and pressure distribution are shown in Fig.19. In the elastoplastic
model, the pressure that an asperity can sustain is limited by its yield strength. An asperity fails when the
pressure is beyond its yield strength and further deforms until more asperities come into contact to share
the load. Indeed, the pressure on most of the asperities in the elastoplastic contact is shown to be around
their yield strength (Fig.19A). While in the elastic-only model, asperities can assume unlimited pressure
without failure, and the pressure imposed on individual asperity can be as high as 18,000 MPa (Fig.19B).
This obviously is not reasonable. In fact, the results predicted by elastic-only models in displacement
(discussed above) and contact area conflict with experiments, where hysteresis in displacement and in
fracture conductivity is always found and large contact area between fractures is also observed (Bandis
1983; Wu. et al 2017). Therefore, to accurately model the contact process, plastic deformation can not
be ignored. Without including plastic deformation, the elastic-only models also overestimate the fracture
SPE-184852-MS 15

conductivity (Fig.20). The example in Fig.20 shows the overestimation can be one order of magnitude, and
it changes case by case.

Figure 19Area in contact and pressure distribution of (A) elastoplastic and (B) elastic-only closure
at 10000 psi. Gray zone indicates no contact and no pressure. Rq=10 m, Y= 300 MPa, E=30 GPa.

Figure 20Comparison in fracture conductivity between elastic-only and elastoplastic closure.

Fracture Compliance and Conductivity


Fracture compliance plays a critical role in determining the fracture conductivity, and is subjected to many
factors, among which the topological and mechanical properties of fracture surfaces are important ones.
The influence of roughness, yield stress and Young's modulus on fracture compliance is shown in Fig.21.
Fracture closure appears to be more sensitive to roughness and yield stress, than to Young's modulus. The
fracture closure at 5000 psi grows from around 20 m to 60 m, as surface becomes rougher with roughness
from 4 m to 20 m (Fig.21A). Increasing closure from around 30 m to 50 m is also achieved, when the
yield stress reduces from 1000 MPa to 100 MPa (Fig.21B). Different from the effect caused by roughness
and yield stress, closure is found to only slightly increase as the surfaces become less stiff with Young's
modulus decreasing from 60 GPa to 20 GPa (Fig.21C).
16 SPE-184852-MS

Figure 21Effect of roughness (A), yield stress (B) and Young's modulus (C) on fracture compliance.

Roughness and yield stress also affect average fracture aperture more significantly than Young's modulus
does. Despite more closure, rougher surfaces embrace larger average aperture as the geometry of the aperture
is directly controlled of roughness (Fig.21A). The average aperture at 5000 psi increases from around 10
m to 50 m (by 4 times) when roughness rises from 4 to 20 m. Average aperture also increases with yield
stress, as the asperities are more difficult to yield. The aperture increases from 7.7 to 25 m (by slightly
over two times), when the yield stress increases from 100 MPa to 1000 MPa for a fracture roughness of
10 m (Fig.21B). When varying Young's modulus, the average aperture, however, changes very slightly.
The variation of the average aperture is within 1.4 m, (around 6%), when the Young's modulus increases
from 20 GPa to 60 GPa.
Since fracture conductivity is closely related to average fracture aperture, the impact of roughness and
yield stress on fracture conductivity is similar to that on average aperture, except that it is further amplified
(Fig.22). The fracture conductivity at 5000 psi increases by around 2 orders of magnitude when roughness
increases from 4 m to 20 m (Fig.22A). The conductivity also increases with yield stress, however to
a reduced degree: It increases by 2 orders of magnitudes when yield stress jumps from 100 MPa to 300
MPa. The degree of increase drops by 5 times when yield stress further rises to 1000 MPa (Fig.22B).
Young's modulus, again, affects conductivity only slightly, as seen for fracture closure and average aperture
(Fig.22C).
SPE-184852-MS 17

Figure 22Effect of roughness (A), yield strength (B) and Young's modulus (C) on fracture conductivity.

The above effects on fracture conductivity based on simulation are consistent with experiments and field
observation. For example, acid fracturing has been actively used to enhance production by introducing
rougher topography to fracture surfaces (Nierode and Kruk 1973; Ruffet et al. 1998). Shale softening in
terms of reduced hardness caused by exposure to water-based fluid usually leads to reduced conductvitiy
(Wu et al. 2017). Lower hardness/yield stress surfaces, fracture (as seen in clay-rich shales) are observed
to be more susceptible to pressure than carbonate-rich shales.
As fracture conductivity has been found to be less sensitive to Young's modulus, we summarize how the
fracture conductivity evolves in the loading stage by covering a wide range of roughness and yield stress.
These results can serveas a reference for future studies and applications (Fig.23).
18 SPE-184852-MS

Figure 23Fracture conductivity between surfaces of varying roughness and yield stress.

Role of Elastic and Plastic Deformation


The total deformation in elastoplastic closure comprises elastic and plastic deformation. The elastic
deformation is calculated based on the Boussinesq equation (Eq.5-7) and is mainly controlled by Young's
modulus; while the plastic deformation depends on the yield stress of the asperity as explained earlier in
the model development section. In the prior section, fracture closure process is shown to be affected more
by yield stress than by Young's modulus. The difference is mainly because the elastic deformation that
is determined by the Young's modulus has less of a contribution to the total deformation than the plastic
deformation that is controlled by the yield stress. When the Young's modulus is reduced from 60 GPa to
20 GPa, the elastic deformation does increase (Fig.24 A, B), while the total deformation remains similar,
especially at high-deformation areas (Fig.24 C, D). However, when the yield stress is reduced from 700
MPa to 100 MPa, both plastic and total deformation change to a similar degree (Fig.25). This indicates that
the total deformation is dominated by plastic deformation, and the whole closure process is more sensitive
to yield stress than to Young's modulus. The contribution of plastic closure to total closure at 5000 psi is
shown to decrease with Y/E, and appears to be higher than 50% in most of the cases (Fig.26). For shale,
with the Y/E typically varying between 1.7 x10-3 and 1.5 x10-2, the plastic closure makes up more than 70%
of the total closure. This percentage is even higher in case of rougher surfaces.
SPE-184852-MS 19

Figure 24Elastic (A, B) and total (C, D) deformation at 5000 psi for surfaces of different Young's modulus.

Figure 25Plastic (A, B) and total (C, D) deformation at 5000 psi for surfaces of different yield stresses.
20 SPE-184852-MS

Figure 26Ratio of plastic to total closure at 5000 psi.

Effect of Correlation Length


Besides roughness, correlation length is another important parameter that determines the surface profile.
However, its effect on fracture closure is often overlooked. Here, a group of fractures with the same
roughness of 20 m, but different correlation length ranging from 50 m to 200 m are created (Fig.27).
The effect of correlation length is investigated by comparing the closure of these surfaces.

Figure 27Surface of same roughness of 20 m but in different correlation lengths.

Fractures with longer correlation length are shown to have smaller initial apertures (Fig.28A). As the
correlation length extends, the spatial density of peaks and valleys on the surface generally decreases, and
their slopes also become more gentle (Fig.27). Therefore, they have more chances to approach each other
and this results in smaller initial apertures. In the case of shorter correlation lengths, surfaces are more
easily compressed by stress due to their larger apertures. Hence more fracture closure is produced (Fig.28B).
Despite more fracture closure, surfaces with shorter correlation length still sustain a greater average fracture
aperture and, therefore, a higher conductivity (Fig 28 B, C).
SPE-184852-MS 21

Figure 28Effect of correlation length on initial aperture (A), fracture compliance


(B), fracture conductivity (C), and fractional contact area (D). Rq=20 m, E=30 GPa.

Slightly more area is in contact in surfaces with longer correlation length during the closure process
(Fig.28 D). The contacts between surfaces with shorter correlation length scatter into many small isolated
areas. However, they aggregate into fewer but larger zones as the correlation length becomes longer
(Fig.29).The fractional contact area is also shown more affected by the yield strength than by correlation
length (Fig. 28 D).

Figure 29Area in contact (in red) at 5000 psi for fractures of roughness
of 20 m but in different correlation lengths. Y= 300 MPa, E=30 GPa.
22 SPE-184852-MS

Normal Fracture Stiffness


Normal fracture stiffness describes the rate of change of normal stress with respect to the normal fracture
closure, and it is defined as (Goodman et al 1968):

(21)

Stiffness involves the geometry and the mechanical properties of the fracture, (Fig.30). Since the stiffness
can be estimated based on seismic attenuation and wave velocity (Far et al.2014), attempts have been made
to seek universal correlations between fracture stiffness and fluid conductivity in the fracture (Pyrak-Nolte
and Nolte 2016; Wang and Cardenas 2016). However, the correlations obtained are from closure models
that consider only elastic deformation. Based on our elastoplastic model, the relationship between stiffness
and fracture conductivity can change significantly when plastic deformation is involved. As yield stress
decreases and more plastic deformation occurs, the originally separate stiffness-conductivity curves tend to
collapse (Fig.30). When given a low yield stress, the plastic deformation has been shown to dominate the
total deformation (Fig.26). The fracture compliance and stiffness is thus more dependent on yield stress,
rather than on roughness and Young's modulus. Given the strong effect of plastic deformation on stiffness,
special attention should be paid when applying the universal correlation between stiffness and fracture flow.
Any correlation must account for plasticity, which is generally involved in the actual contact process.

Figure 30Relationship between stiffness and fracture conductivity for cases of low (A) and high (B) yield stress.

Closure of Surfaces with Heterogeneity


One key advantage of the numerical closure models over the analytical ones is the ability to simulate
closure for heterogeneous surfaces. This ability is critical for application in shales, where heterogeneity
is commonly observed. Shales are usually composed of multiple minerals such as clay, quartz, carbonate,
and these minerals are known to have quite different mechanical properties (Kumar et al 2012; Wu and
Sharma 2016). The mineral composition and distribution also generally vary from shale to shale. Therefore,
to understand how heterogeneity affects fracture closure is important and desirable.
Here, eight synthetic shale surfaces with a simplified composition of only soft and hard components are
generated (Fig.30). Since Young's modulus has been found to have a minor impact on fracture closure, both
components are assumed to have the same Young's modulus of 40 GPa, but different yield stress, with 1000
MPa for the hard component and 300 MPa for the soft one. In the synthetic surfaces, the content of the soft
components gradually increases from 10% to 70 % (Fig.30 A to D, E to H). The relative correlation in y
direction is kept constant as 0.03; while 0.03 (Fig.30 A to D) and 0.3 (Fig.30 E to H) are used in x direction,
representing the scattered and laminated distribution pattern, respectively. The closure process of these 8
SPE-184852-MS 23

surfaces are compared with the cases of uniformly hard (H=1, S=0) and uniformly soft (H=0, S=1). All
surfaces have identical initial surface profiles with a roughness of 10 m.

Figure 31Surfaces with hetegenous distribution of soft (in orange) and hard (in dark
gray) components. Rq=10 m, Esoft= Ehard = 40 GPa, Ysoft =300 MPa, Yhard = 1000 MPa.

The fracture closure and fracture conductivity with respect to closure stress for the 8 surfaces of
heterogeneity are all bound between the results of the uniformly soft and the uniformly hard surfaces
(Fig.32). More soft component causes more plastic deformation, and leads to greater fracture closure and
lower fracture conductivity (Fig.32-33). By comparing the distribution of pressure with the distribution of
components, it is clear that the pressure limit that a grid can hold is specifically associated with its yield
stress (Fig.34)

Figure 32Fracture closure and conductivity for surfaces with


heterogeneity. Rq=10 m, Ysoft= 300 MPa, Yhard = 1000 MPa, E=40 GPa.
24 SPE-184852-MS

Figure 33Total deformation at 10,000 psi for surfaces with heterogeneity.


S: soft components. Rq=10 m, Ysoft= 300 MPa, Yhard = 1000 MPa, E=40 GPa.

Figure 34Comparison between distribution of pressure (A) with distribution of components (B) for
pattern G in Fig.31 at closure stress of 10000 psi. Gray zone in (A) indicates no contact and no pressure.

Interestingly, the fracture closure appears relatively insensitive to the distribution pattern of hard and soft
components. Surfaces with scattered and laminated patterns present very similar closure behavior (Fig.32).
The distribution pattern also seems to have little effect on the locations where most deformation occurs,
which is likely more dependent on the overall surface profile (Fig.33 A and B).

Conclusions
The closure model developed in this work is able to simulate the elastoplastic contact of unpropped fractures
and compute the resulting fracture conductivity. The model integrates elastic deformation and plastic
deformation with consideration of asperities interaction. The contact model has been validated against
analytical solutions and experiments, for both elastic-only and elastoplastic contacts. A comprehensive
SPE-184852-MS 25

investigation of the closure of unpropped fractures was conducted for a wide range of conditions with the
model. The main conclusions are as follows:
1. Our algorithm that combines a conjugate gradient method and FFT technique is effective in modeling
the contact process and computing the fracture closure process. The computation time is significantly
reduced (by five orders of magnitude) compared with the traditional method for a grid of 512 by 512;
2. Interactionbetween multiple asperities can not be ignored for an accurate description of the contact
process between rough surfaces;
3. The elastic-only model is shown to underestimate fracture closure, overpredict fracture conductivity,
and produce results that contradict experiments;
4. Roughness and yield stress have a stronger impact on fracture compliance than Young's modulus.
Fracture conductivities increase by two orders magnitude when roughness is increased from 4 m to
20 m. Increasing the yield stress of the asperities from 100 MPa to 1000 Mpa also leads to a similar
degree of increase in conductivities;
5. The total deformation in a typical fracture closure process has a greater contribution from plastic
deformation than from elastic deformation. For shales, the plastic closure can make up more than
70% of the total closure;
6. Shorter correlation lengths in the surface profile results in greater average aperture and fracture
conductivity. The fractional contact area is very similar for contact between surfaces with different
correlation lengths, however the aperture opening pattern for surfaces with shorter correlation length
is more scattered.
7. The relationship between fracture stiffness (compliance) and conductivity is strongly affected plastic
deformation.
8. The model presented in this paper can simulate closure of heterogeneous fracture surfaces. More soft
mineral components cause more plastic deformation, and this leads to more fracture closure and lower
fracture conductivity. The distribution pattern of the minerals on the surfaces has only a minor impact
on fracture compliance.

Acknowledgements
We express our gratitude to the support from companies that sponsor the Hydraulic Fracturing and Sand
Control Joint Industry Project at the University of Texas at Austin.

Nomenclature
a Radius of contact area, m
ax, ay Grid size in x or y direction, m or m
C Constant used to relate H to Y, 3.0
CG Conjugate gradient method
Relative approach or fracture closure, m or m
E1, E2 Young's modulus of the two contacting surfaces, GPa
E* Effective Young's modulus, GPa
0 Tolerance for convergence, 10-5 or 10-6
F External load, N
FCA Fractional contact area
FCA12 Fractional area for aperture smaller than 12 m
FFT Fast Fourier transform
g Fractuer aperture after deformation, m or m
h Initial fracture aperture (m) or fracture height (m or ft)
26 SPE-184852-MS

H Hardness measured from indentation test, MPa


IFFT Inverse fast Fourier transform
K Influence coefficient, m/GPa
Kn Normal stiffness, psi/ m
K Influence coefficient in frequency domain
kfW Fracture conductivity, mD.ft
L Fracture length, m or ft
Mx, My Grid number in x or y direction
Viscosity of fluid, cp or Pa.s
N Total grid number
p Pressure at specific grid or inlet/outlet pressure, Pa or psi
p0 Maximum pressure at the center of the contact, Pa
p Pressure at specific grid in frequency domain
p Pressure drop applied to the fracture network, Pa or psi
q Volumetric flow rate, m3/s
R Radius of sphere, m
Rq Roughness, m
n Normal closure stress, psi
u Total deformation, m or m
uij_E Elastic deformation, m or m
Deformation in frequency domain
Dimensionless deformation
W Aperture in the fracture network, m
x Distance from origin, m
Dimensionless distance
Y Yield stress, MPa
v1, v2 Poisson's ratio of the two contacting surfaces

References
Ai, X. and Sawamiphakdi, K., 1999. Solving elastic contact between rough surfaces as an unconstrained
strain energy minimization by using CGM and FFT techniques. Journal of tribology, 121(4):639-647. http://
dx.doi.org/10.1115/1.2834117
Bandis, S.C., Lumsden, A.C. and Barton, N.R., 1983. Fundamentals of rock joint deformation. International
Journal of Rock Mechanics and Mining Sciences & Geomechanics Abstracts, 20(6): 249-268. http://
dx.doi.org/10.1016/0148-9062(83)90595-8
Berthe, D. and Vergne, P., 1987. An elastic approach to rough contact with asperity interactions. Wear, 117(2):211-222.
http://dx.doi.org/10.1016/0043-1648(87)90256-0
Brown, S.R. and Scholz, C.H., 1985. Closure of random elastic surfaces in contact. J. geophys. Res, 90(B7):5531-5545.
http://dx.doi.org/10.1029/jb090ib07p05531
Brush, D.J. and Thomson, N.R., 2003. Fluid flow in synthetic roughwalled fractures: NavierStokes, Stokes, and local
cubic law simulations. Water Resources Research, 39(4). http://dx.doi.org/10.1029/2002wr001346
Deng, J., Hill, A.D. and Zhu, D., 2011. A theoretical study of acid-fracture conductivity under closure stress. SPE Prod
& Oper., 26(01): 9-17. SPE-124755-PA. http://dx.doi.org/10.2118/124755-pa
Far, M.E., de Figueiredo, J.J., Stewart, R.R., Castagna, J.P., Han, D.H. and Dyaur, N., 2014. Measurements of seismic
anisotropy and fracture compliances in synthetic fractured media. Geophysical Journal International, 197:1845-1857.
http://dx.doi.org/10.1093/gji/ggu101
Fischer-Cripps, A.C., 2011. Nanoindentation, 3rd edtion. New York: Springer. http://
dx.doi.org/10.1007/978-1-4419-9872-9
Gangi, A.F., 1978. Variation of whole and fractured porous rock permeability with confining pressure. Int. J. Rock Mech.
Min. Sci. Geomech. Abstr., 15(5): 249-257. http://dx.doi.org/10.1016/0148-9062(78)90957-9
SPE-184852-MS 27

Garcia, N. and Stoll, E., 1984. Monte Carlo calculation for electromagnetic-wave scattering from random rough surfaces.
Physical review letters, 52(20):1798-1801. http://dx.doi.org/10.1103/physrevlett.52.1798
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.
Goodman, R.E., Taylor, R.L. and Brekke, T.L., 1968. A model for the mechanics of jointed rocks. Journal of Soil
Mechanics & Foundations Div. Proc. Am. Soc. Civ. Engrs, 94(SM3):637-659.
Greenwood, J.A. and Williamson, J.B.P., 1966. Contact of nominally flat surfaces. Proceedings of the Royal Society A:
Mathematical, Physical and Engineering Sciences, 295(142):300-319. http://dx.doi.org/10.1098/rspa.1966.0242
Hopkins, D.L., 1991. The Effect of Surface Roughness on Joint Stiffness, Aperture, and Acoustic Wave Propagation. Ph.D.
dissertation. University of California at Berkeley, Berkeley, California.
Johnson, K.L., 1985, Contact Mechanics. Cambridge, UK: Cambridge University Press.
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
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.
Liu, S., Wang, Q. and Liu, G., 2000. A versatile method of discrete convolution and FFT (DC-FFT) for contact analyses.
Wear, 243(1):101-111. http://dx.doi.org/10.1016/s0043-1648(00)00427-0
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.
Nieto, C.M., Pournik, M. and Hill, A.D. 2008. The Texture of Acidized Fracture Surfaces: Implications for Acid Fracture
Conductivity. SPE Prod & Oper 23(3): 343-352. SPE-102167-PA. http://dx.doi.org/10.2118/102167-PA.
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.
Polonsky, I.A. and Keer, L.M., 1999. A numerical method for solving rough contact problems based on the
multi-level multi-summation and conjugate gradient techniques. Wear, 231(2):206-219. http://dx.doi.org/10.1016/
s0043-1648(99)00113-1
Press, W.H., Teukolsky, S.A., Vetterling, W.T. and Flannery, B.P., 1996. Numerical recipes in C (Vol. 2). Cambridge:
Cambridge university press.
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
Pyrak-Nolte, L.J. and Nolte, D.D., 2016. Approaching a universal scaling relationship between fracture stiffness and fluid
flow. Nature communications, 7:10663-10668. http://dx.doi.org/10.1038/ncomms10663
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.
Sone, H. and Zoback, M.D., 2013. Mechanical properties of shale-gas reservoir rocksPart 1: Static and dynamic elastic
properties and anisotropy. Geophysics, 78(5):D381-D392. http://dx.doi.org/10.1190/geo2013-0050.1
Swan, G., 1983. Determination of stiffness and other joint properties from roughness measurements. Rock Mechanics and
Rock Engineering, 16(1):19-38. http://dx.doi.org/10.1007/bf01030216
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
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.
Walsh, J.B. and Grosenbaugh, M.A., 1979. A new model for analyzing the effect of fractures on compressibility. Journal
of Geophysical Research: Solid Earth, 84(B7): 3532-3536. ttp://dx.doi.org/10.1029/jb084ib07p03532
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).
28 SPE-184852-MS

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.
Zhou, J. 2015. Interactions of organic-rich shale with water-based fluids. PhD dissertation, The University of Texas at
Austin, Austin, Texas (August 2015).
SPE-184852-MS 29

Appendix
In the Hertzian contact between a sphere and a flat plat with an external load F applied on the sphere (Fig.11),
a series of analytical equations are available to describe the contact process (Johnson 1985). The radius of
the contact area is calculated as:

(A-1)

Where R is the radius of the sphere; E* is the effective Young's modulus and defined as Eq.7. The
maximum normal pressure at center of the contact zone is obtained as:

(A-2)

Normal displacement and pressure within contact zone are calculated as:

(A-3)

(A-4)

The relative approach between the sphere and the flat surface is described as:

(A-5)

Anda mungkin juga menyukai