You are on page 1of 12

SPE-174046-MS

Conductivity of Proppant-Packs under Variable Stress Conditions: An


Integrated 3D Discrete Element and Lattice Boltzman Method Approach
M.M Mollanouri Shamsi, Sh. Farhadi Nia, and K. Jessen, University of Southern California

Copyright 2015, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Western Regional Meeting held in Garden Grove, California, USA, 2730 April 2015.
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
The conductivity of the proppant-pack plays a critical role in the productivity of a hydraulically fractured
well. Here we combine the Discrete Element Method (DEM) and the Lattice Boltzmann Model (LBM)
to investigate the conductivity of proppant-packs under variable confining stress. Three different proppant-packs including well graded, uniformly graded and poorly graded proppants are used to generate
loose packing of spherical rigid particles with no initial contact. We then use a DEM algorithm, with
specified confining stress, to compact the proppant-packs. A representative elementary volume of the pack
is subsequently used for fluid flow simulations. LBM is used to calculate a detailed flow field at the pore
scale and the Darcy-scale permeability is then calculated from the average pore-scale velocity and
pressure information. The results show, as expected, that the porosity and permeability of the proppantpack decreases with increasing confining pressure. The permeability of well graded proppant-packs is
more sensitive to the stress while uniformly graded packs are more permeable than poorly graded
proppant-packs. The calculated permabilities are in good agreement with available experimental data for
a range of confining pressures. In addition, the permeability estimated from the Kozeny-Carman equation
based on calculated porosities compares well with calculated permeabilities.

Introduction
To achieve economic production rates from low permeability shale formations, the reservoir rocks must
be stimulated/fractured. To induce fractures, large volumes of water are pumped into the subsurface.
During hydraulic stimulation, the induced fracture extends until the rate of fluid loss into the formation
exceeds the pumping rate. Once pumping is stopped and pressure decreases below the fracture closure
pressure, the fracture may close rapidly resulting in loss of production from the induced fracture area. To
avoid this, a propping agent (sand or ceramic spheres) is added to the injected fluid to ensure that the flow
paths remain open after hydraulic fracturing (Economids and Nolte, 2000; Reinicke et al., 2010):
Proppants are injected to keep the fractures open after the pumping stops. The proppant particles
accumulate between fracture faces as a compacted pack with a sufficient permeability to enable continued
oil/gas production. At the production stage, fluids flow from the rock matrix and micro fractures into the
larger fractures and move to the wellbore through the proppant-packed fractures.

SPE-174046-MS

Long-term proppant testing data in the literature show a general trend of a rapid conductivity/
permeability decline of the fracture network during the first couple of months followed by little decline
during the rest of test (Montgomery, C.T. and Steanson, R.E., 1985; B. W. McDaniel, 1986; Hahn
Drilling, 1986; Cobb and Farrell, 1986; Jim Weaver et al., 2007; P. J. Handren and Terrence T. Palisch,
2007; Subhash N. Shah et al., 2010; M. C. Vincent, 2010a, 2011b). The initial fast decline is the results
of several factors including proppant consolidation and reorientation (Cobb and Farrell, 1986).
The pack of proppants is formed and subjected to compressive forces from the rock faces and shear
forces due to the pressure differential over the time (Kulkarni and Ochoa, 2012). The shear forces crush
the proppant-pack and change the porocity and finally the productivity. Also, transverse forces due to
differential stress/pressure tend to dislodge the proppant from the pack which may result in erosion of the
well which interrupt production (Kulkarni and Ochoa, 2012). Fines production that results from crushing
and erosion also has the potential to block the pore space between proppants and reduce the productivity
over the time.
Mechanical interactions between the reservoir and hydraulic fractures under stress and formationfracture damage processes results from several mechanisms. These include physical, biological, chemical,
and thermal interactions between formation and fluid and can significantly influence the performance of
propped areas (Reinicke et al., 2010). Significant efforts have been dedicated via laboratory, field and
theoretical studies to delineate fracture damage mechanisms (Bishop, 1997; Lynn and Nasr-El-Din, 1998;
Behr et al., 2002; Civan, 2000; Fredd et al., 2000; Moghadasi et al., 2002; Nasr-El-Din, 2003; Wen et al.,
2006; Reinicke et al., 2010).
To have a better evaluation of the problems such as oil-proppants in which the interfacial physics along
the solid-fluid interface dominates the performance of the system, a coupled system that consists of fluid
mechanics and solid mechanics must be utilized. To simulate the behavior of a proppant-pack, Discrete
Element Method (DEM) as discussed by Cundall and strack (1979), is very effective in modeling
quasi-static and dynamics of granular materials deformation. On the fluid mechanics side, the lattice
Boltzmann Method (LBM) consists of a set of kinetic equations capable of recovering Navier-Stokes
equations for slightly compressible fluids at the incompressible limit with constraints of low Mach and
Knudsen numbers (Chen et al, 1992).
This paper presents an integrated 3D DEM-LBM study of three different proppant-packs under relevant
conditions. Three different grain size distributions of proppant-packs, including well graded, uniformly
graded and poorly graded are generated and compacted in DEM under variable confining stress. The
conductivity of the different proppant packs was then calculated via LBM simulations. A representative
elementary volume of the pack was used for fluid flow simulations. LBM was used to calculate a detailed
flow field at the pore scale and the Darcy-scale permeability was then calculated using the pore-scale
velocity and pressure information. To verify the model, the results are compared with previous theoretical
and experimental studies. The proposed workflow provides an effective approach for studying multiphase
characteristics of proppant-packs, including optimization of proppant size distributions for maximum
conductivity at a given the stress state or stress profile over time.

Modelling Approach
Any reduction in pore pressure, and the associated increase in net stress, results in the compaction of the
proppant pack and reduces its conductivity. A schematic of a proppant pack in a crack created through
hydraulic fracturing is shown in Fig. 1. For modeling purposes, it is assumed that the proppants are rigid
spherical particles in a cubic space. The particles are selected from a distribution for diameter sizes and
are randomly distributed in space with no initial contact. The Discrete Element Method is then used to
compress a proppant-pack under a specified confining pressure. Fluid flow simulation using the Lattice
Boltzmann Method is then performed to calculate the permeability of a given proppant-pack state.

SPE-174046-MS

Figure 1(a) A hydraulically fractured opening filled with proppants, (b) Magnified view of a proppant-pack, (c) Numerical model
representative of a proppant-pack

Proppant-Pack Compaction: Discrete Element Method (DEM)


To model an assembly of proppants, 1000 rigid spherical particles are randomly distributed in a loose
cubic packing without any initial contacts. The diameter of particles is selected from a distribution
function to represent the gradation of the pack. The program divides the cubic space into several boxes
for convenience of contact detection. At this stage, the program assigns a box or boxes for every particle
(Shamsi and Mirghasemi, 2012). By applying boundary loads, particles can move based on the Newtons
second law as follow:
(1)

(2)
In eq. (1), i and ui denotes acceleration and velocity of a particle, respectively. Fi is the sum of forces
acting on each particle, mi is the mass of a particle and gi is the gravity acceleration. In eq. (2), and
represent angular acceleration and velocity, respectively. Mi is total moment acting on a particle and Ii
is the moment of inertia. in both equations is the viscous (mass proportion) damping constant. Eqs. (1)
and (2) describe the linear movement and rotation movement of particles, respectively. A time step should
be selected small enough to achieve a stable numerical solution. The procedure is repeated with the
updated particle position in each time step via the expressions for the force-displacement relation.
Table 1 reports the DEM parameters that were used in the simulation. Parameters are chosen to ensure
numerical stability.

SPE-174046-MS

Table 1Parameters for DEM simulation


Parameters

values

Normal contact stiffness (N)


Shear contact stiffness (S)
Inter particle friction coefficient ()
Inter particles cohesion (c)
Density of particles ()
Damping coefficient (D)

2x1010 (N/m)
2x1010 (N/m)
0.5
0.0
2x103 (N/m3)
5x10-5

Proppant-pack conductivity: Lattice Boltzamnn Method (LBM)


The conductivity of a propant pack is calculated using the Lattice Boltzmann Method (LBM). The Lattice
Boltzmann Method (Benzi et al., 1992, Succi, 2001) is a mesoscopic computational fluid dynamics
method and represents a discrete form of the Boltzmann equation near equilibrium. In this method, the
fluid is modeled with a distribution function of particles that can move between specified lattice sites. For
each lattice time step, particles move to their neighboring sites and exchange momentum through
collisions while under the influence of external forces and satisfying the imposed boundary conditions.
LBM can be derived directly from the Boltzmann equation (He and Lou, 1997), however historically it
is the evolved version of the Lattice Gas Automata (LGA).
To calculate the permeability in a given direction (e.g. x), the inlet and outlet faces are open and all
the other faces are closed with no-flow boundaries (approximates the experimental configuration used for
permeability measurements). A periodic boundary condition with a body force is then applied in the flow
direction. The LBM simulation results in a flow field (pressure and velocity) that we average to calculate
the permeability from Darcys equation as follow:
(3)
Here, U is the sum of the x-direction components of velocity along a constant x-plane, which due to
mass conservation for incompressible systems is the same for all planes perpendicular to the x axis. is the
viscosity of the fluid, is the length, and is the pressure difference.

Results and Observations


Propped fractures greater than 0.25 in (6.35 mm) are very difficult to achieve, especially at the deeper
setings, and there is not a unique particle size selection for injection of proppants trough the fracture
(Montgomery et al., 1985). To reach a good distribution of proppants in a pack, five different types with
equivalent diameters ranging from 0.1 to 1.5 mm were generated for each assembly. The equivalent
diameter is the diameter of the sphere which circumscribes each particle. For this study, as mentioned
earlier, three different ensample of proppants with different size distribution were considered, as shown
in Fig. 2 and Fig. 3.

SPE-174046-MS

Figure 2Different grain size distribution of proppant-packs including well graded, uniformly graded and poorly graded.

Figure 3Different proppant-packs, (a) well graded, (b) uniformly graded and (c) poorly graded.

Figure 4(a) shows the initial condition of a propant pack prior to any compression and with no initial
contact. The particles are subjected to a hydrostatic strain rate equal to 1.010-7 in each step. Figure 4
(b), (c) and (d) show the propant pack under 1, 10 and 25 MPa of confining stress, respectively.

SPE-174046-MS

Figure 4 DEM simulation (a) initially generated particles, (b), (c) and (d) compacted particles in 1 MPa, 10 MPa, and 25 MPa confining
pressure, respectively.

After the compaction process in DEM, we arrive at a proppant-pack in a 200x200x200 domain and e.g.
a confining pressure that equals 1 MPa (See Figure 5(a) and 5(b) below). To decrease computational cost,
an element of 50x50x50 from the DEM simulation was extracted for LBM analyses (Fig. 5(c)). To ensure
that we extract a representative volume, several elements from different parts of the DEM simulation were
tested and the permeability was computed with LBM. Table 2 reports the permeability in the different
areas of the larger domain and we observe that they are in good agreement. Therefore, we assume that one
subelement can represent the overall DEM ensample.

SPE-174046-MS

Figure 5(a) Schematic view of cubic sample (DEM, 200x200x200) and LBM sample (50x50x50), (b) Compacted particles after DEM
simulation, (c) A proppant-pack for LBM analyses
Table 2LBM permeability in different parts of main sample (DEM simulation sample)
Subvolume (x, y, z)
Permeability (Darcy)

(40,40,40)
2.24e02

(70,70,70)
2.12e02

(100,100,100)
2.18e02

(130,130,130)
2.21e02

Constant pressure boundary conditions were applied at the inlet and outlet faces (P1.1 at X 1 and
P1.0 at X50). Figure 6 shows the LBM simulation setup. All four sides orthogonal to the flow
direction (X) have no-flow boundaries, defined as solid points.

Figure 6 LBM simulation setup for flow in the X-direction

SPE-174046-MS

Permeability has the dimension of length squared, and in the LBM approach the length dimension is
denoted lu (length unit) which is, in fact, the resolution of the input image.
(4)
To reach an acceptable resolution and decrease the CPU cost of LBM simulations, different meshes
were considered. Finally, we decided to discretize our model in LBM simulation with the same resolution
as the DEM simulation.After running the LBM simulation, the permeability can be calculated from
Darcys equation (equation 3).
Figure 7 shows the porosity of proppants over a range of confining pressures. As we expect, by
increasing the confining pressure the porosity of samples decreases. We also can see that the porosity of
well graded sample is lower in comparison with uniformly graded and poorly graded proppant packs.

Figure 7Porosity versus confining pressure for different proppant-packs

Figure 8 shows the permeability of the proppant-packs as a function of its porosity. The permeability
of the samples decreases with decreasing porosity.

SPE-174046-MS

Figure 8 Permeability versus porosity of different proppant-packs

Comparison with Experimental Data and the Kozeny-Carman Equation


In order to validate our numerical calculations, we compare our results with experimental work and/or
analytical methods. To this end, we start by comparing our results with experimental data. Zou et al.,
(2013) conducted an experimental study to simulate coal fines migration trough a proppant-pack. Figure
9 compares the results from the experimental works without any fine coals with the current numerical
study.

Figure 9 Permeability results from LBM-DEM method (current study) and experimental study

As observed from Fig. 9, the permeabilities obtained from our numerical calculations are larger than
the experimental data. In our numerical calculations we have not applied a differential stress and we
expect to achieve a better agreement between clculations and experimental observationsas we include
differential stress. Works is currently underway to apply differential stress in our calculation framwork.

10

SPE-174046-MS

As a second validation, we consider the Kozeny-Carman correlation. The Kozeny-Carman correlation


relates the permeability of porous medium to the specific surface area and the porosity as follow (Wyllie
and Gregory, 1955):
(5)
where, is the porosity of the sample, S0 is the surface area per unit volume of the solid phase, and
C is the Kozeny-Carman constant. For equal sized spheres, the specific surface area can be expressed
where r is the radius of the uniform spheres (Jin. G et al., 2012). The
through the grain diameter
constant C is approximately equal to 5 for flow through unconsolidated sand (Wyllie and Gregory, 1955).
Figure 10 compares the permeability of proppant packs as a function of confining pressure as obtained
from our modeling framework and the Kozeny-Carman correlation. We observe a good agreement
between numerical calculations and Kozeny-Carman correlation over the full range of porosities that is
generated during the DEM compaction calculations.

Figure 10 comparing permeability versus confining pressure between LBM-DEM simulation and Kozeny-Carman method

Discussions and Conclusions


To enhance the evaluation of productivity from hydraulically fractured reservoirs, a coupled DEM-LBM
approach was applied to model the conductivity a proppant-packed fracture. In this preliminary work
different packs of proppant with different grain size distribution were considered. The conductivity of
each ensample under variable (uniform) confining stress was then calculated. Our results show that
uniformly graded proppants provide for a better conductivity relative to poorly/well graded proppant
packs in a compacted stage. We note that this preliminary work only considers uniform confining stress.
The results also confirm that a poorly graded proppant-pack has more porosity/permeability in comparison
with a well graded proppant-pack at a constant confining stress. The numerical results wres compared with
experimental observations and the Kozeny-Carman correlation. While some departure from the experi-

SPE-174046-MS

11

mental data was observed due to the omission of differential stress, the values and trends are in good
agreement with estimates from the Kozeny-Carman correlation. An improved accuracy of the proposed
work flow, relative to experimental observations, is expected by inclusion of differential stress and is
currently underway.

References
Andreas Reinicke, Erik Rybacki, Sergei Stanchits, Ernst Huenges, Georg Dresen, 2010. Hydraulic
fracturing stimulation techniques and formation damage mechanisms-Implications from laboratory
testing of tight sandstoneproppant systems, Chemie der Erde 70 S3, 107117.
Behr, A., Mtchedlishvili, G., Friedel, T., Haefner, F., 2002. Consideration of damage zone in tight gas
reservoir model with hydraulically fractured well. SPE 82298.
Benzi, R., Succi, S., and Vergassola, M., 1992. The lattice Boltzmann equation: theory and applications. Physics Reports, 222(3), 145197.
Bishop, S.R., 1997. The experimental investigation of formation damage due to the induced flocculation of clays within a sandstone pore structure by a high salinity brine. SPE 38156.
B. W. McDaniel, 1986. Conductivity testing of proppant at high temperature and stress, 56th Candaian
Regional Meeting of the SPE in Oakland.
Chen, S., & Doolen, G. D., 1998. Lattice Boltzmann method for fluid flows. Annual review of fluid
mechanics, 30(1), 329 364.
Chen H, Chen S, Matthaeus W.H, 1992. Recovery of the NavierStokes equations using a lattice-gas
Boltzmann method, Phys. Rev. A. 45:5339 5342.
Civan, F., 2000. Reservoir Formation Damage. Gulf Publishing Company, Houston, TX, USA.
Economides, M.J., Nolte, K.G., 2000. Reservoir Stimulation 3rd ed Wiley and Sons Ltd., United
Kingdom
Fredd, C.N., McConnell, S.B., Boney, C.L., England, K.W., 2000. Experimental study of hydraulic
fracture conductivity demonstrates the benefits of using proppants. SPE 60326.
Guodong Jin, Tad W. Patzek, Dmitriy B. Silin, 2012. Modeling the impact of rock formation history
on the evolution of absolute permeability, Journal of Petroleum Science and Engineering 100
153161.
Jim Weaver, Mark Parker, Diederik van Batenburg, Philip Nguyen, 2007. Fracture-related diagenesis
may impact conductivity, International Symposium and Exhibition on Formation Damage Control,
Lafayette, L.A., USA.
Lynn, J.D., Hisham, A., Nasr-El-Din, H.A., 1998. Evaluation of formation damage due to frac
stimulation of a Saudi Arabian clastic reservoir. J. Petrol. Sci. Eng. 21, 179 201.
M. C. Kulkarni and O. O. Ochoa, 2012. Mechanics of Light Weight Proppants: A Discrete Approach,
Composites Science and Technology, Vol. 72, pp. 879 885.
M. C. Vincent, 2010. Refracs-why do they work, and why do they fail in 100 published field studies,
SPE Annual Technical Conference and Exhibition held in Florance, Italy.
M. C. Vincent, 2010. Restimulation of unconventional reservoirs: when are refracs beneficial,
Canadian Unconventional Resources and International Petroleum Conference, Calgary, Canada.
M. M. M. Shamsi, A. A., 2012. Mirghasemi, Numerical simulation of 3D semi-real-shaped granular
particle assembly, Powder Technol., 221, pp. 431446
Moghadasi, J., Jamialahmadi, M., Muller-Steinhagen, H., Sharif, A., Izadpanah, M.R., 2002. Formation damage in Iranian oil fields. SPE73781.
Montgomery, C.T. and Steanson, R.E. 1985. Proppant Selection: The Key to Successful Fracture
Stimulation (Revised). J. Pet Tech 37 (12): 21632172. SPE-12616-PA., 10.2118/12616-PA.
Nasr-El-Din, H.A., 2003. New mechanisms of formation damage: lab studies and case histories. SPE
82253.

12

SPE-174046-MS

P. A. Cundall, O.L. Strack, 1979. A discrete numerical model for granular assemblies, Geotechnique
29 (1) 4765.
Patrick Joseph Handren and Terrence T. Palisch, 2007. Successful Hybrid Slickwater Fracture Design
Evolution--An East Texas Cotton Valley Taylor Case History, SPE Annual Technical Conference
and Exhibition, Anaheim, California, U.S.A, SPE 110451, http://dx.doi.org/10.2118/110451-MS.
Reinicke, A., 2010. Mechanical and hydraulic aspects of rockproppant systems laboratory experiments and modelling approaches. Doctoral Thesis, University at Potsdam, Germany.
S. L. Cobb, J.J. Farrell, 1986. Evaluation of long-term proppant stability, International Meeting on
Petroleum Engineering, 17-20 March, Beijing, China, SPE 14133, 10.2118/14133-MS.
Subhash N. Shah, M. C. Vincent, Robert Rodriquez, Terry Palisch, 2010. Fracture orientation and
proppant selection for optimizing production in horizontal wells, SPE Oil and Gas India Conference & Exhibition held in Miumbai, India.
Succi, S., 2001. The lattice Boltzmann equation: for fluid dynamics and beyond. Oxford university
press.
Wen, Q., Zhang, S., Wang, L., Liu, Y., Li, X., 2006. The effect of proppant embedment upon the
long-term conductivity of fractures. J. Petrol. Sci. Eng. 55, 221227.
Wyllie, M. R. J., Gregory, A. R., 1955. Fluid flow through unconsolidated porous aggregates: effect
of porosity and particle shape on Kozeny-Carman constants. Ind. Eng. Chem. 47 (7), 1379.
Y. S. Zou, S. C. Zhang, J. Zhang, 2013. Experimental method to simulate coal fines migration and coal
fines aggregation prevention in the hydraulic fracture, Transp Porous Med, 101:1734.
He, X., & Luo, L. S., 1997. Lattice Boltzmann Model for the Incompressible Navier-Stokes Equation,
Journal of Statistical Physics, Vol. 88, Nos. .
He, X., & Luo, L. S., 1997. Theory of the lattice Boltzmann method: From the Boltzmann equation
to the lattice Boltzmann equation. Physical Review E, 56(6).

You might also like