Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

SPE-189856-MS

Analysis and Modeling of Proppant Transport in Inclined Hydraulic


Fractures

R. Kou, G. J. Moridis, and T. A. Blasingame, Texas A&M University

Copyright 2018, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Hydraulic Fracturing Technology Conference & Exhibition held in The Woodlands, Texas, USA, 23-25 January 2018.

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 understanding of proppant transport in fractures plays a critical role in estimating propped fracture
dimensions and performance. Modeling of such processes is challenging because of the complex interactions
between fluid, proppant particles and fracture geometry. Existing models generally assume vertical planar
fracture geometry, whereas the reality in the subsurface maybe much more complex. In this study we use
discrete element method and computational fluid dynamics simulations to demonstrate that interactions
between proppant particles and fracture side walls play an important role in proppant transport efficiency.
To calibrate our numerical model, we conducted two validation simulations that describe particle settling
tests and laboratory proppant transport experiments. Through scoping calculations, we determined the
correct drag force model and matched both analytical solutions and experimental data for a wide range of
flow regimes. We then constructed two hydraulic fracture simulation domains, one with a vertical planar
fracture (as a base case), and the other with an inclined planar hydraulic fracture. In the main component
of our study, we conducted proppant transport simulations using our benchmarked models in both domains
and compared the proppant distribution results.
By analyzing the velocity and trajectory of proppant particles during transport, we identified two different
stages of the proppant transport process - a "suspension" stage and a "settling" stage. During the suspension
stage, fluid drag and gravitational forces dominate, driving proppants further into the fracture. When
the proppants collide with existing proppants dunes, the proppant particles lose momentum and become
"settled." Our results show that settled proppant are more difficult to be mobilized by fluid drag force, for
the reason that kinetic energy is easily dissipated inside proppant dunes. Finally, we observed that proppant
settles slower in inclined planar fractures, due to the supporting force from fracture side walls cancelling
part of the gravitational force which act on the particle. This leads to a better proppant placement efficiency
in inclined fractures.
Objectives: The objective of this work is to provide a better understanding of proppant transport behavior in
inclined planar fractures by means of numerical simulation. A significant difference from prior (laboratory)
experiments is the capability of our model to simulate proppant transport at field-scale flowrates. Such
acapability is critical in understanding proppant transport and appropriately designing it by ensuring that
the correct Reynolds number and flow regime are used in the design calculations.
2 SPE-189856-MS

Introduction
Low permeability reservoirs (shale and tight reservoirs) have been contributing a significant portion of oil
and gas production, especially in the United States. Multi-stage hydraulic fracturing is commonly used to
stimulate low permeability reservoirs. A successful stimulation aims to create long hydraulic fractures that
reach deep into the reservoir. To achieve optimal hydraulic fracturing design and operation, understanding
proppant transport plays a critical role. Most existing models are limited in the geometry and orientation
of fractures that are capable of describing. Typical proppant transport models assume planar (vertical)
hydraulic fracture geometry. The subsurface reality may be far removed from such a simple model. A
combination of depositional and tectonic processes have often resulted in heterogeneous and anisotropic
media involving complex stress distributions and, consequently, natural fractures of complex geometry
and orientation. This results in complex fracture geometries and multi-scale transport mechanism (Akkutlu
2016, 2017). Many recent efforts have been focusing on characterizing (Killough et al., 2017, Valko et al.,
2016, King et al. 2017) and modeling (Schechter et al., 2017, Sun et al. 2016, Cao et al. 2017) complex
fracture networks. However, the development of proppant transport models associated with such complex
geometries has been slow. This is because proppants are discrete particles, whereas the majority of transport
models describe flow processes that involve phase continuity and are represented by Eulerian methods.
The Eulerian-Eulerian two-fluid method (TFM) treats both solid particles and fluid as continua, which
cannot capture the discrete character of the solid phase of the proppant particles. The limitation of TFM
can be overcome by Lagrangian approaches such as the Discrete Element Method (DEM), in which solid
particles are tracked individually (Chen et al. 2014). DEM calculates particle-particle and particle-wall
collisions explicitly, and its strong conceptual underpinnings make it a better method to simulate solids
transport in fractutres of complex geometry.
Its theoretical and conceptual advantages notwithstanding, DEM is more computational demanding than
TFM schemes. A DEM simulation of field-scale proppant transport scenarios could require tracking of
millions of Lagrangian particles. This amount of computation has only recently become possible because of
the advent of powerful supercomputer platforms and efficient parallel algorithms. Several parallelized open
source packages are available for DEM simulations. In this work, we performed DEM simulation using
LIGGGHTS (Kloss et al. 2012), which is an open source C++ simulator built on the LAMMPS (Plimpton,
1995) platform and parallelized with MPI and CUDA.

The Numerical Model Formulation and Validation


The Discrete Element Method
The Discrete Element Method (DEM) is a numerical technique that predicts the behavior of bulk solids
by modeling particle-particle interactions and particle-boundaries interactions. It is a physics-based method
developed from first principles. DEM takes into account the forces acting on each individual particle
using Newton's law of motion, and integrates the forces over time to calculate the particle velocity and
position. Compared to previous methods, in which solid phase transport is mostly modeled as a continuous
phase, DEM does not require such an assumption. Therefore, DEM has become a powerful tool with
increasing application to many industrial processes, including: granular material compaction, powder
mixing, pneumatic conveying and drill cuttings transport.
In the DEM, particles can be considered either deformable or rigid. These two types are mostly referred to
as soft-sphere (SS) and hard-sphere (HS) formulation, respectively. For a dilute system (in which multibody
collisions are limited), the HS formulation is appropriate and is computationally efficient. In a dense system,
in which multi-body collisions between particles are a frequent occurrence, the SS formulation is more
appropriate (Norouzi, et al., 2016). In this paper, we use the SS formulation, in which the contact forces
SPE-189856-MS 3

are calculated from over lapping distance between deformable particles. In Eq. (1), we show the DEM
governing equation for Lagragian particles using Newton's law of motion.

(1)

in which mi is mass of particle i in kg. xi is location of particle i in meter. vt is velocity of particle i in m/s.
fij is force from particle j to particle i in N/m. g is gravitational constant in m/s2.
Eq. (1) indicates that the velocity changes of a particle multiplied by its mass equals to the summation
of all forces acting on the particle. On the right hand side of Eq. (1), the summation term represents the
pairwise contact force between paricles i and j, with j representing all particles in the contact list of particle
i. Integrating Eq. (1) with regard to time yields the particle's velocity at a new time step. Double integration
of Eq. (1) with regard to time yields the new position of particle i.
The pairwise contact force in Eq. 1 is calculated using force-displacement laws. One of the earliest contact
force models was the Linear Spring-Dashpot (LSD) model by Cundall and Strack (1979), in which the
contact force equals the spring stiffness of linear spring multiplied by the displacement. Wellmann et al.,
(2008) shows that even though the LSD models can match some experimental results, there are cases in
which non-linear force-displacement models result in higher accuracy, especially when particle-particle
interactions are dominant (Norouzi, et al.,2016).
In the simulation of proppant transport with slick water as the fracturing fluid, our hypothesis is that
particle-particle interactions play an important role. Therefore, we did not consider the LSD model and used
the Hertz contact model (Hertz, 1882) instead. This is described by

(2)

where kn, kt are elastic constant on normal, tangential direction, in N/m. δn,δt are overlap distance on normal,
tangential direction, in m. γn, γt are viscoelastic damping constant on normal, tangential direction, in N-s/m.
The Hertz contact force model is a nonlinear viscoelastic model that considers contact force in both the
normal and the tangential directions (Eq. 2). The force in each direction consists of two components, elastic
and viscous forces. The elastic constant (kn, kt) represents the force due to elastic deformation, while the
viscoelastic damping constant (γn, γt) controls the energy dissipation. and are the overlap in the
normal and tangential direction, respectively. and are the relative velocities in the normal and
the tangential direction, respectively. The difference between the Hertz model and the LSD model is that,
in the Hertz model, the elastic constant and viscoelastic damping constant are not constant but functions of
the overlap distance between particles. Detailed relationships are shown in Eq. 3-9 (Kloss, et al., 2012).

(3)

(4)

(5)

(6)
4 SPE-189856-MS

(7)

(8)

(9)

where Y is the Young's modulus of elasticity of the granular material, in Pa, G is the shear modulus, in Pa,
υ is the Poisson ratio (dimensionless) and e is the coefficient of restitution (dimensionless).

Coupled CFD-DEM Formulation


With the DEM formulation explained in the above section, we now introduce the coupled CFD-DEM
formulation. We first add the fluid-particle force on the right hand side of Eq. 1, yielding

(10)

One of the intuitive ways to calculate the fluid-particle force (in Newton) is to integrate the
pressure on the particle surface. This approached is called the Resolved Surface Method (RSM) and requires
knowledge of the fluid pressure at each point of the particle surface. RSM requires the computational fluid
dynamics (CFD) mesh to be much smaller than the particle size, which limits the number of particles in the
simulation if the capacity of the high-performance computing platform is isufficient. To simulate transport
behavior of a large number of particles, the Unresolved Surface Model (USM) is more practical. Instead of
integrating the pressure on the particle surfaces, in the USM the fluid-particle force is defined as:
(11)

where (in Newton) is the buoyancy force, (in Newton) is the drag force based on the difference
between the cell-based ensemble averaged particle velocity <up> and the fluid velocity (m/s). Many
unresolved drag force correlations have been proposed (Di Felice, 1994; Schiller et al., 1935; Kafui et al.,
2002). We use the Di Felice model (Eq. 12-15) because it is valid for both dense and dilute flow of particles
(Norouzi, et al., 2016), i.e., for both the SS and HS formulations. According to this model:

(12)

(13)

(14)

(15)

where Rei is the particle Reynolds number, vf is kinematic viscosity of the fluid in m2/s. dp is particle's
diameter in meters. εf is volume fraction of the fluid. uf is fluid velocity in m/s. <up> is cell averaged particle
velocity in m/s.
In the Di Felice (1994) model, the particle Reynolds number (Rei) is calculated using the kinematic
viscosity of the fluid (vf), particle the diameter (dp) and the velocity difference between the fluid and the
SPE-189856-MS 5

particle ensemble. The drag force is calculated using the drag coefficient (Cd), the particle Reynolds number
(Rei) and the volume fraction of the fluid phase (εf). To show the performance of the Di Felice (1994)
drag force model, we plotted its associated drag coefficient based and that obtained from experimental data
(summarized by Duan et al. 2015) in Figure 1. The result shows that the Di Felice (1994) model matches
the experimental data for a wide range of Reynolds numbers.

Figure 1—Drag coefficient estimated from the Di Felice drag force model, calculated from the
Stokes drag model, and determined from experimental data (summarized by Duan et al. 2015).

Finally, we added the fluid-particle force term (Eq. 11) to the fluid phase momentum conservation
equation. Considering the fluid (slick water) as a Newtonian fluid, the governing equations describing the
fluid phase flow are represented by Eq. 16-17. As:

(16)

(17)

where ρf is the fluid density in kg/m3. is the fluid velocity in m/s. μf is the fluid viscosity in poise. kv
(dimensionless) is the number of particles in the corresponding fluid cell.
The momentum conservation equations for the particles (Eq. 10) and the fluid (Eq. 17) are coupled
through the fluid-particle force term . In the numerical model, time is discretized using a first-order
forward model and, for unconditional numerical stability, the equations are solved using a fully explicit
coupling scheme. This means that all variables and parameters, including the coupling term , are
calculated with the information at the current time step. The underlying assumption here is that the position
of particles does not change dramatically during the DEM iterations.
6 SPE-189856-MS

Model Validation
To calibrate our numerical formulation, we first conducted a simulation that described a laboratory test of
a settling proppant particle. A detailed description of the problem and of the associated simulation set up
is included in Appendix A. The simulation study involved tracking a spherical particle with a diameter
of 0.512 mm (representing 50/70 mesh sand) that is dropped in a water container. The evolution of the
velocity of the descending (settling) particle is shown in Figure 2. The particle initially accelerated under
the influence of the gravity force, and eventually it reached terminal velocity.

Figure 2—50-70 mesh sand settling velocity in water

To calculate the terminal velocity, one of the commonly used analytical equations is the Stokes (Stewart
et al., 2002) equation (Eq. 18), which assumes that viscous effects dominats during settling. Therefore, it is
applicable for scenarios characterized by a particle Reynolds number (Eq. 19) smaller than 1. This is also
shown in Fig. 1.

(18)

(19)

Knowing the limitation of Stokes equation, in order to calibrate our model over the full range of Reynolds
number, we conducted laboratory settling tests using 3 different-sized particles (20/30 mesh, 30/40 mesh
and 50/70 mesh) and two types of fluids (water and oil). The recorded experiment data are included in
Tables 1-3 in Appendix A. Figure 3 shows the comparison of the terminal velocity as determined from the
laboratory experiment, Stokes analytical equation and from our simulation of the settling test. The results
indicate that the terminal velocity estimated from the the coupled CFD-DEM simulation matches well the
experimental data, but the agreement with the analytical solution deteriorates as the particle size increases.
This is particularly evident in the case of the of 20/30 mesh sand, in which the terminal velocity estimate
from the Stokes equation is inaccurate because the related Reynolds number being larger than 1.
SPE-189856-MS 7

Table 1—50-70 Mesh Sand Settling in Water

Distance, cm time 1 time 2 time 3 average time terminal velocity-1, m/s

40 11.01 10.18 10.66 10.62 0.0377


15 4.34 4.74 4.68 4.59 0.0327

Table 2—20-30 Mesh Sand Settling in Water

Distance, cm time 1 time 2 time 3 average time terminal velocity-1, m/s

40 4.21 4.28 4.03 4.17 0.0958


15 1.61 1.54 1.58 1.58 0.0951

Table 3—30-40 Mesh Sand Settling in Oil

Distance, cm time 1 time 2 time 3 average time terminal velocity-1, m/s

70 77.32 79.1 102.28 86.23 0.0081


40 39.7 33.65 36.55 36.63 0.0109

Figure 3—Comparison of simulated terminal velocity to those from a lab experiment and from the Stokes equation.

The settling test and the Stokes equation are considered 2-way coupling problems in which the fluid
to particle drag force and the gravity force dominate. The proppant transport process is considered as a
dense particle transport problem, in which particle-particle forces also play an important role. Therefore, it
is necessary to calibrate our model in a problem involving a dense particle system. This was accomplished
by using the study of Tran et al. (2017), who conducted lab scale experiments to study the effect of PGA
fiber on proppant transport. The simulation for our second validation model used the same input parameters
as those described in the experiment of Tran et al. (2017).
The lab scale experiment set up consists of 2 transparent proppant transport slots and a slick water
injection port located on the left hand side of the apparatus (Fig. 4). The dimensions of the rectangular panel
were 30 cm × 120cm. The experiment was performed by injecting 30/60 mesh proppant in slick water (with
a viscosity of 3.5 cp) at a rate of 650 ml/min. Tran et al. (2017) observed that the proppant settled near the
injection port and formed a "dune". The dune front was observed to advance when the dune front slope
8 SPE-189856-MS

exceeded a 40 degree angle. As shown in Fig. 5, our simulation result captured the same settling behavior. A
sand dune with the same frontal slope was formed near the injection port. In Fig. 5, each proppant is color-
coded according to its velocity. The proppants entering the planar fracture have a relatively high velocity
(shown in red color) because of the fluid drag force. Then, proppants move along the top of the existing dune
and gradually lose momentum. Finally, the proppants at nearzero velocity (shown in dark blue color) settle,
forming a part of the new dune. This simulated behavior is in agreement with many other recent experimental
studies (Mohanty et al., 2017, Sahai et al. 2014). Even though laboratory experiments can provide valuable
insights on the subject of the settling behavior of proppants, they are limited by the magnitude of the fluid
velocity that they can accommodate. Eq. 15 shows that the fluid velocity determines the particle Reynolds
number, which has an impact on all three components of the drag force equation (Eq. 12). Because the fluid
drag force is the only driving force that transports proppant into the fracture, it is necessary to simulate
proppant transport with realistic fluid velocities that correspond to field-level pumping rates.

Figure 4—Lab scale proppant transport experiment (From Tran et al. 2017)

Figure 5—Lab scale proppant transport simulation using the coupled CFD-DEM model

Field Scale Simulation Cases


Case 1—Vertical Fractures
Following the calibration of our model calibrated at both the particle and the lab scale, we conducted
field-scale simulation studies of proppant transport. Detailed information on the meshing (i.e., spatial
discretization) used in these studies is provided in Appendix B. To convert the surface pumping rate (barrel
SPE-189856-MS 9

per minute) to fluid velocity in the fracture, we assumed a 7-inch production casing and a maximum fracture
width of 0.3 inch near wellbore. We estimated that a surface pumping rate of 30 bpm corresponds to an inlet
velocity of 5 m/s, and that a 60 bpm surface pumping rate corresponds to a 10m/s inlet velocity. Here we
define inlet velocity as the fluid velocity at the intersection between the horizontal wellbore and the planar
fracture. For the same surface pumping rate, the corresponding fracture inlet velocity could vary from case
to case, as it can be influenced by factors such as the number of clusters per stage and the actual number
of fractures per stage.
The dimension of our simulated hydraulic fracture is 2m × 10m. The simulation described the injection
of slick water (with a viscosity of 3.5 cp) that carried 30/60 mesh proppant at an inlet velocity of 5m/s. Fig.
5-8 show the proppant distribution from early time to screen out. For clarity, the proppants are color-coded
according to their velocity. Fig. 6 shows simulated proppant distribution at 50 seconds of the simulation
time. Since we performed 2-dimentional simulation, the simulation time doesnot represent physical time in
the actual fracturing operation. At 50 seconds, one can observe that the majoriy of particles in the fracture
are in suspension-stage (light blue). We sampled the cell-averaged particle velocity and the fluid velocity
near the inlet (along the horizontal axis at Y = 0 in Fig. 6). We plotted the results in Figure 9. It shows
that momentum is being transferred from the fluid to the particles near the inlet, causing an increase in the
particle velocity and a decrease in the fluid velocity. While particles are in suspension-stage, the amount
of particle-particle collision (which dissipates momemtum energy) is minimum. Thus, the fluid drag force
was able to accelerate the particles to the same velocity as the fluid, leading to a high efficiency in proppant
transport.

Figure 6—Proppants distribution in vertical fracture at 50 seconds

Figure 7—Proppants distribution in vertical fracture at 150 seconds


10 SPE-189856-MS

Figure 8—Proppants distribution in vertical fracture at 240 seconds

Figure 9—Proppants screen out in vertical fracture at 300 seconds

Once the dune is formed Figs. 6 and 7 show that it mostly grows in height (y direction); its frontal edge (in
the x direction) advances, but proppant accumulation near the advancing front (the height of the proppant
distribution along the Y-axis) is limited and concentrated near the base of the fracture. This is because a large
fraction of suspension-stage proppants does not have enough momentum to travel over the dune, causing
most of the proppants to settle on top of the dune. We again sampled the cell averaged particle velocity and
the fluid velocity near the the dune top (along the Y axis at X = 6 m in Fig. 7) at 240 seconds and plotted the
result in Figure 10. In contrast to the suspension stage, particle-particle collisions and particle-wall friction
dominate in the settling stage. Even though the momentum energy is being transferred from the fluid to
particles on the dune top, it is dissipated due to the particle-particle collsion. The hence it is difficult to
create proppant momentum using the fluid drag force.

Figure 10—Particle and fluid velocity near the inlet region sampled at 50 seconds
SPE-189856-MS 11

Fig. 9 shows the final proppant distribution at 300 seconds. At screen out, the midpoint of the proppant
front is at 5.5 m in the x direction, leaving 45% of the fracture area unpropped. To increase the efficiency
of proppant placement, an intuitive way is to increase the inlet velocity. To test this hypothesis, we repeated
the simulation with a higher inlet velocity (10m/s, corresponding to a 60 bpm surface pumping rate).
Comparison of the proppant distribution at 150 seconds is shown in Fig. 12. Compared to the previous case,
the higher inlet velocity provides sufficient momentum to proppants during the suspension stage to travel
over the dune, thus avoiding settling on top of the dune. Thus, higher inlet veolicity results in proppant
transport advancing deeper (along the x-direction) into the fracture and in a larger propped area.

Figure 11—Particle and fluid velocity near the dune top at 240 seconds

Figure 12—Comparison between 5m/s inlet velocity (top) and 10m/s inlet velocity (bottom) at 150 seconds

Case 2—Inclined Fractures


In general, hydraulic fractures are assumed to be vertical in deep reservoirs. This is based on the assumption
that the least principal stress is horizontal. However, field studies have shown that hydraulic fractures are
rarely, if ever, perfectly vertical (Wright et al. 1998, Dinh et al. 2009). Daneshy (1972) showed that hydraulic
fractures appears vertical near the wellbore, but change orientation (tending toward the least compressive
in-situ principal stress) as they extend away from the wellbore. Wright et al. (1995) provided field examples
showing that the Chevron Lost Hills fracture inclination changed from 82 degrees from the horizontal to 45
degrees between two fracturing treatments over a long production period.
12 SPE-189856-MS

In simulation Case 2, we described proppant transport in a fracture plane with 30 degree inclination from
the vertical. A detailed description of the simulation set up and of the mesh used in this study can be found in
Appendix B. The dimensions of the inclined fracture are the same as in Case 1. As in that case, the simulation
involved the injection of slick water at inlet velocity of 5m/s, carrying a 30/60 mesh proppant (sand).
To compare the settling behaviors of proppants in vertical and inclined fractures, we sampled the velocity
of the suspension stage proppants (i.e., near the inlet) in both the vertical and inclined fracture. The vertical
components of the proppant velocity, sampled at two different times (10 seconds and 20 seconds), are
shown in Fig. 13 and 14. The results indicate that, during the suspension stage, proppants settle slower in
inclined fractures. The slower settling is attributed to the extra support force exerted by the inclined plane.
As proppants settle on the inclined plane, the contact force from the fracture sidewalls has an upwards
vertical component. This cancels out part of the gravitational force (illustrated in Fig 15). Consequently,
it takes longer for proppants to reach the bottom of fracture and accumulate as dunes. Assuming the same
amount of drag force acting on the proppants, proppants in inclined fractures travel further into the fracture
before the onset of the settlingstage. A comparison of proppant placement in inclined fractures (top) and
vertical fractures in Fig. 16 shows that, for the same pumping rate, the propped area in the inclined fracture
is 15% larger than in the vertical fracture.

Figure 13—Suspension stage particle velocity vertical component, sampled at 10 seconds

Figure 14—Suspension stage particle velocity vertical component, sampled at 20 seconds


SPE-189856-MS 13

Figure 15—Illustration of force acting on downward moving (settling) proppants (left) and upward moving proppants (right).

Figure 16—Comparison of final proppant placement in inclined fracture plane (top) and vertical fracture plane (bottom)

During the design of the hydraulic fracturing operation, one of the goals is to place stronger and better
proppants near the wellbore, in order to maximize the near-wellbore permeability. This is usually done by
pumping small amount of expensive proppants near the end the fracturing operation. Fig. 17 depicts the
particle residence time in the fracture. Our results show that the earliest-injecrted proppants (identified by
the dark blue color) occupy the bottom section of the fracture, while the latest-injected ones (identified by the
red color) occupy the top section of the proppant distribution in the fracture. This observation is significantly
different from results from earlier proppant transport models, which usually show that proppants pumped
last stay near the wellbore. Contrary to that, our results show that during slick water fracturing, near-wellbore
proppants are a combination of both the early- and late-stage proppants. This could explain why, in some
cases, proppants injected at the early/initial stages of hydraulic fracturing process are recoved first during
flow back operations.
14 SPE-189856-MS

Figure 17—Proppant residence time at screen out

Conclusions
In this paper, we focus on modeling and analyzing the transport of proppants in vertical and inclined
fractures. Using a parallelized code based on the discrete element method and high performance computing
facilities, we were able to conduct large-scale simulations involving realistic fluid injection velocities that
exceed (and cannot be reached) those attained in past laboratory experiments. The results allowed significant
insights into the proppant transport process and led to the follwong conclusions:
1. Suspension-stage proppants can be easily accelerated by the fluid drag force. This transfer of
momentum happens near the inlet region of hydraulic fractures. Low inlet fluid velocity will cause
proppants to settle on top of existing dunes during the initial suspension stage, leading to early screen
out and inefficient proppant placement.
2. Proppants settle slower in inclined fractures because the contact force from fracture side walls partially
offsets gravitational effects. Assuming the same fluid velocity at the inlet, proppants in inclined
fractures will be transported further into inclibed fractures compared to those in vertical fractures.
3. For slick water hydraulic fracturing operations, our results indicate that near wellbore proppants
consists of both early- and late-stage injected proppants. This observation differs significantly from
the results of previous models, which tend to indicate that proppants near the wellbore are those
last injected. The observations from our study could provide the basis for an alternative approach in
designing proppant pumping schemes.

Acknowledgements
This work was supported by the Crisman Institute for Petroleum Engineering Research at Texas A&M
University. The authors gratefully acknowledge their support. All simulations were performed using the
Ada cluster of Texas A&M University Supercomputing Facilities and supported by the High Performance
Research Computing (HPRC) group.

Nomenclature
Cd = dimensionless drag coefficient
dp = particle diameter, m
e= coefficient of restitution
= contact force from particle j to particle I, N
G= Shear modulus, Pa
kn, kt = elastic constant on normal, tangential direction, N/m
mi = mass of particle i, kg
SPE-189856-MS 15

p= Pressure, Pa
Rei = Particle Reynolds number for particle i, dimensionless
u∞ = terminal velocity, m/s
= fluid velocity, m/s
vi = velocity of particle i, m/s
Vcell = volume of CFD cell
xi = location of particle i, m
Y= Young's modulus, Pa
δn, δt = overlap distance on normal, tangential direction, m
εf = volume fraction of fluid phase
ρp = particle density
ρf = fluid density
υ= Poisson ratio
vf = kinematic viscosity of fluid, m2/s

References
Bird R. B., Stewart W. E., and Lightfoot E. N., Transport phenomena, 2nd end, John Wiley and Sons, Inc., Hoboken
Chen, Xizhong, and Junwu Wang. "A comparison of two-fluid model, dense discrete particle model and CFD-DEM
method for modeling impinging gas-solid flows." Powder Technology254 (2014): 94–102.
Christoph Kloss, Christoph Goniva, Alice Hager, Stefan Amberger, Stefan Pirker, "Models, algorithms and validation
for opensource DEM and CFD-DEM," Progress in Computational Fluid Dynamics, An Int. J. 2012 - Vol. 12, No.2/3
pp. 140 – 152
C. Wellmann, C. Lillie, P. Wriggers, Comparison of the macroscopic behavior of granular materials modeled by different
constitutive equations on the microscale, (2008) Finite Elements in Analysis and Design, Volume 44, Issue 5, March
2008, Pages 259–271
Cundall, Peter A., and Otto DL Strack. "A discrete numerical model for granular assemblies." geotechnique 29.1 (1979):
4765.
Daneshy, A. A. (1973, April 1). A Study of Inclined Hydraulic Fractures. Society of Petroleum Engineers.
doi:10.2118/4062-PA
Di Felice, R. (1994) The voidage function for fluid-particle interaction systems. International Journal of Multiphase Flow,
20, 153–159.
Dinh, A. V., & Tiab, D. (2009, January 1). Transient-Pressure Analysis of a Well With an Inclined Hydraulic Fracture
Using Type Curve Matching. Society of Petroleum Engineers. doi:10.2118/120540-MS
Duan, Zhipeng, Boshu He, and Yuanyuan Duan. "Sphere drag and heat transfer." Scientific reports 5 (2015).
Hamid Reza Norouzi, Reza Zarghami, Rahmat Sotudeh-Gharebagh, Navid Mostoufi. Coupled CFD-DEM Modeling:
Formulation, Implementation and Application to Multiphase Flows Wiley
Hertz, H. (1882) Ueber die Beruhrung fester elastischer Korper. Journal fur die reine und angewandte Mathematik, 92,
156171
Kafui, K.D., Thornton, C., and Adams, M.J. (2002) Discrete particle-continuum fluid modelling of gas-solid fluidised
beds. Chemical Engineering Science, 57, 2395–2410.
Kou, Rui, Saad FK Alafnan, and Yucel Akkutlu.I. "Multi-scale analysis of gas transport mechanisms in kerogen."
Transport in Porous Media 116.2 (2017): 493–519.
Kou, R., Alafnan, S. F. K., & Akkutlu, I. Y. (2016, May 30). Coupling of Darcy's Equation with Molecular Transport and
its Application to Upscaling Kerogen Permeability. Society of Petroleum Engineers. doi:10.2118/180112-MS
Plimpton S., Fast Parallel Algorithms for Short-Range Molecular Dynamics, J Comp Phys, 117, 1–19 (1995)
Sahai, R., Miskimins, J. L., & Olson, K. E. (2014, February 4). Laboratory Results of Proppant Transport in Complex
Fracture Systems. Society of Petroleum Engineers. doi:10.2118/168579-MS
Schiller, L. and Nauman, A. (1935) A drag coefficient correlation. VDI Zeitung, 77, 318–320.
Tang, H., Killough, J. E., Heidari, Z., & Sun, Z. (2017, August 1). A New Technique To Characterize Fracture Density by
Use of Neutron Porosity Logs Enhanced by Electrically Transported Contrast Agents. Society of Petroleum Engineers.
doi:10.2118/181509-PA
Tong, S., Singh, R., & Mohanty, K. K. (2017, October 9). Proppant Transport in Fractures with Foam-Based Fracturing
Fluids. Society of Petroleum Engineers. doi:10.2118/187376-MS
16 SPE-189856-MS

Tran, T., Kim, J. Y., Morita, N., & Yoshimura, K. (2017, August 28). Application of PGA Fiber and Fluid-Loss Materials
to Slick Water Fracturing. American Rock Mechanics Association.
Wang, Z., Li, C., & King, M. (2017, February 20). Validation and Extension of Asymptotic Solutions of Diffusivity
Equation and Their Applications to Synthetic Cases. Society of Petroleum Engineers. doi:10.2118/182716-MS
Wright, C. A., & Conant, R. A. (1995, January 1). Hydraulic Fracture Reorientation in Primary and Secondary Recovery
from Low-Permeability Reservoirs. Society of Petroleum Engineers. doi:10.2118/30484-MS
Wright, C. A., Davis, E. J., Minner, W. A., Ward, J. F., Weijers, L., Schell, E. J., & Hunter, S. P. (1998, January 1).
Surface Tiltmeter Fracture Mapping Reaches New Depths - 10,000 Feet and Beyond? Society of Petroleum Engineers.
doi:10.2118/39919-MS
Zhou, Y., Heidari, Z., Zhu, D., & Hill, A. D. (2017, October 4). Petrophysical Rock Classification, Permeability Estimation,
and Elastic Moduli Assessment in Tight Carbonate Reservoirs: A Case Study in Tarim Field, China. Society of
Petroleum Engineers. doi:10.2118/187516-MS
SPE-189856-MS 17

Appendix A
Particle Settling Simulation Setup and Recorded Experiment Data

Figure 1—CFD meshing and DEM particle for the particle settling simulation
18 SPE-189856-MS

Appendix B
Vertical Fracture Inclined Fracture Meshing

Figure 1—Mesh used for the vertical fracture simulation

Figure 2—Mesh used for the inclined fracture simulation.

You might also like