Вы находитесь на странице: 1из 12

SPE-184829-MS

An Investigation into Proppant Dynamics in Hydraulic Fracturing

Baidurja Ray, Chris Lewis, Vladimir Martysevich, Dinesh A. Shetty, Harold G. Walters, Jie Bai, and Jianfu Ma,
Halliburton

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
Effective proppant placement during hydraulic fracturing is essential to obtain maximum stimulation
effectiveness. Understanding proppant placement requires the understanding of the time and space
dependent dynamics of proppant motion in fluids, which include the phenomena of proppant transport,
bridging, settling, and resuspension. This paper proposes a laboratory test method that can be used to
investigate any aspect of proppant dynamics in a variety of channel configurations and fracturing fluids.
3D printing technology is used to rapidly manufacture channel flow devices of various dimensions. After
a 3D printer is available, such manufacturing is extremely inexpensive with rapid turn-around times.
These channels, in conjunction with laboratory scale pumps and blenders, are used to investigate proppant
transport and bridging, settling, and resuspension in various fracturing fluids. Several different channel
configurations, ranging from uniform width to uniform tapered, are used to investigate the dynamics of
small and large diameter proppants with fluids ranging from water to linear gels. The results from these
experiments are compared with numerical models for validation, and in some cases, calibration of model
inputs, that can ultimately lead to improved fracturing treatment design and understanding. In addition, the
paper provides a comparison to existing data (Patankar et al. 2002) to validate settling and resuspension
models.

Introduction
Proppant dynamics is a critical physical (and in certain cases, chemical) process that determines the
effectiveness of the hydraulic fracturing treatment. It determines, among other things, which of the fractures
stimulated during the treatment are propped and the distribution of proppant within them. This, in turn,
determines the conductivity of the fractures and their eventual production. In this paper, experiments and
numerical model results are used to investigate three major aspects of proppant dynamics in hydraulic
fractures: proppant transport, proppant bridging, and the formation and evolution of a settled proppant bed.
For the purposes of this paper, the fractures are considered to be smooth-walled channels with widths varying
along the length, although the experimental methodology can be used with nearly any prescribed geometry.
Proppant transport mechanisms in channels have long been studied in the industry (e.g., Kern et al. 1959,
Daneshy 1978, Woodworth and Miskimins 2007, Patankar et al. 2002, Sahai et al. 2014). Previous laboratory
2 SPE-184829-MS

scale experiments on proppant dynamics have focused on the transport properties of proppant-laden fluid,
specifically as it relates to the settling of the proppant and the formation of an equilibrium settled bed. These
works considered fixed width channels (Patankar et al. 2002) and network of channels (Sahai et al. 2014),
and explored the transport of sand in Newtonian and non-Newtonian fluids. A few studies have focused on
establishing the bridging criteria for proppant-laden slurries entering channels or perforations (Smith 1991,
Gruesbeck and Collins 1982, Barree and Conway 2001).
3D printing has recently become an attractive tool to rapidly manufacture complex geometries for a
variety of applications. In the oil and gas industry, they have recently been used to investigate transport
in channels with walls of different roughness profiles to mimic fracture walls (Roy et al. 2016). In this
study, 3D printing technology is used to rapidly manufacture a number of different channel configurations to
investigate various aspects of proppant dynamics. The focus of this work includes establishing the bridging
criteria in the interior of channels, investigating settling and resuspension by observing the equilibrium bed
height for different proppant-fluid combinations under different flow conditions, and generally providing a
comprehensive testing methodology that can be used to investigate the various aspects of proppant dynamics
in different channel configurations. Another goal is to use these experimental results to calibrate and validate
numerical models that attempt to predict proppant dynamics in hydraulic fracturing.

Experimental Setup
This section describes the experimental procedure. The channel flow experimental setup consisted of a set
of syringe pumps, a differential pressure transducer, a fluid reservoir, a blender, the channel with centered
entry points, and a high speed camera, as shown in Fig. 1 and Fig. 2. The channel was mounted and held in
place by transparent acrylic sheets for effective visualization. Fluid (or slurry) was pulled from the reservoir
by the pump, and flowed through the channel at different gel concentrations (for linear gel fracturing fluids)
and flow rates. Proppants of varying sizes and size distributions were tested. Proppant-fluid mixture at the
specified concentration was poured into the reservoir that was mounted on a laboratory scale blender and
continuously churned at an appropriate rate to maintain the required concentration when pulled into the
channel by the pump. Such a set-up ensures unidirectional flow throughout the duration of the experiments,
and no issues with reverse flow were encountered. The camera was used to record the propagation of the
slurry through the channel, and the images were processed to measure the necessary quantities, such as
bridging location and settled bed height. The differential pressure across the inlet and outlet was recorded
throughout the duration of the test.

Figure 1Conceptual schematic of the laboratory scale experimental


setup used to investigate proppant dynamics in channels.
SPE-184829-MS 3

Figure 2Fully assembled experimental setup example showing the various components.

For each test, the mixture is pumped through the channel until the maximum extension of the pump is
reached or until bridging occurs and a maximum pressure differential is reached, based on the pressure
rating of the transducer. As previously described, the mixture is under constant shear in the blender to avoid
premature settling while pumping and to maintain the required concentration. Although the setup shown in
Fig. 1 shows one channel, any geometric configuration can easily be used. Future work will include the use
of multiple channels that meet one another under various configurations to investigate fluid and proppant
dynamics in fracture networks. The setup is at the laboratory scale; therefore, careful consideration must
be given to scaling up the results.
Stereolithograpy (SLA) 3D printing was used to manufacture several channel flow devices. SLA was
chosen because of its high printing resolution and optical properties for flow visualization. The 3D printing
of these devices enabled inexpensive, high throughput manufacture and rapid turnaround of designs with
different sizes, geometries, and textures. To remove the burden of transparency from the channel walls, the
channels were printed with only one wall. The channel can then be mounted between highly transparent
acrylic sheets for good visualization. This process enables the appropriate material to be used for printing
the channels without concern about its optical properties. Uniform width and tapered width channels are
used in the test results reported in this paper. Fig. 3 shows a tapered channel immediately after printing.
The inlet and outlet of the channels are suitably designed to minimize turbulence and flow separation as
the slurry enters and exits the channels through pipes connected to the center. For the tapered channels
with very small widths at one end that were used to investigate bridging, an additional widening ramp is
provided (Fig. 3) to ensure smooth exit of the slurry from the channel. Fig. 2 provides an example of a
fully assembled experimental setup. As previously described, the workflow begins with printing a channel
of the desired configuration in the 3D printer, as shown in Fig. 2(a). For better visualization, the channels
are printed with one wall only, but this restriction can be removed by using a more transparent material for
printing the channels. The channel is then mounted on transparent acrylic blocks for ease of visualization,
4 SPE-184829-MS

as shown in Fig. 2(b). After the channel is ready, it is mounted appropriately for viewing with a camera,
and the pumps, blenders, and pressure transducer are attached, as shown in Fig. 2(c).
Results from these experiments are shown for three different fluids: water, a thin linear gel G1 (base gel
viscosity 6.5 cp at 511 s1 with power-law index n'= 0.692, and consistency index K' =0.038 Pa-sn') and a
slightly thicker linear gel G2 (base gel viscosity 11 cp at 511 s1 with n' =0.585,k' = 0.128 Pa-sn').

Figure 33D printed channel right after completion of printing. Notice that it has only one
wall; the inner wall is tapered from a small width on the left to a larger width on the right.

Numerical Models
This section briefly describes the numerical models that are mentioned in this paper. The experimental
results are used to calibrate and validate the proprietary numerical models. These models simulate the time-
dependent multiphase proppant-laden flow in channels by solving for conservation of mass and momentum
of fluid and proppant in a fully coupled manner. To capture the evolution of the proppant bed, two competing
mechanisms are considered: proppant settling, which causes the proppants to fall out of the fluid vertically
and settle at the bottom of the channel, and proppant resuspension, which represents the erosion of an
existing proppant bed driven by shear at the bed-slurry interface. A large number of particle settling models
have been proposed in the literature with varying ranges of validity (Gadde et al. 2004, Malhotra and Sharma
2011, Blyton et al. 2015, Liu 2006, Wu et al. 2014). The results presented in this paper use one of the most
comprehensive models of settling of a single particle in infinite domain (Betancourt et al. 2015), which
provides a unified formulation for both Newtonian and power-law fluids. In confined domains with many
particles, the settling velocity needs to be augmented by accounting for the effect of proppant concentration
(Clark and Quadir 1985, Dunand and Soucemarianadin 1985, Liu 2006), and for the effect of fracture walls
(Missirlis et al. 2001, Song et al. 2009, Liu 2006, Machac and Lecjaks 1995) for both Newtonian and non-
Newtonian fluids. Modeling proppant resuspension in a computationally tractable manner is a challenging
issue because of the complexity of the process. The sediment pick-up functions (Van Rijn 1984, Nakagawa
and Tsujimoto 1980) provide an efficient framework for predicting the resuspension velocity as a function
of shear velocity at the slurry-bed interface.
SPE-184829-MS 5

Results

Bridging and Bridging Criteria. Proppant bridging inside hydraulic fractures is investigated using the
experimental setup previously described. The results presented in this paper pertain to bridging inside the
fractures, as opposed to at the perforation (Gruesbeck and Collins 1982, Smith 1991) or at the junctions
between fractures (Barree and Conway 2001). Gruesbeck and Collins (1982) and Smith (1991) investigated
the bridging criteria (defined as the ratio of local flow aperture (w) to mean particle diameter (dp) at
which bridging occurs, referred to henceforth as Bc) for sand flowing through perforations and obtained
an empirical estimate that ranged from Bc~1 at very small proppant cocentrations to as high as Bc~6
at very large concentrations. Barree and Conway (2001) considered the problem of proppant bridging
in smooth wall slots and concluded that the bridging phenomenon is fundamentally different in slots or
channels (fractures), as compared to circular openings, such as perforations. Their experiments supported
the hypothesis that a stable bridge is formed in a channel junction only if the particle diameter is equal to or
larger than the channel width, irrespective of the proppant concentration. They showed that when proppant
encounters a right angle turn to a fracture, although a bridge can be formed at w > dp, it is stable only if w =
dp. However, it is a still fairly common practice to use values of BC~3 as an input in fracture modeling. This
work attempts to establish the bridging criteria for smooth wall bridging within a single fracture. For this
purpose, a tapered channel is used that can represent the tapered geometry of a fracture. In this work, several
experiments are performed with different proppant fluid combinations, and all the data support a bridging
criteria of BC~1. In other words, particles will begin bridging in a fracture when it encounters an aperture
that is equal to its diameter. Fig. 4 shows that in a uniformly tapered channel with maximum width of 3.98
mm (left end of the channel) and a minimum width of 0.25 mm (right end of the channel), a 40/50 mesh
ceramic proppant in a linear gel begins bridging at BC~1, based on the mean diameter. The small variability
in the location of onset of bridging is attributable to a size distribution present in the proppant sample (which
was intentionally chosen to be narrow). This behavior occurs regardless of the inlet proppant concentration
and fluid type, and the same Bc is observed for 80/100 mesh sand. Numerical models often accept Bc as
an input, and these experiments provide a means of determining it under different conditions. In principle,
any kind of channel (for example, with roughness or non-uniform height) can be used to investigate BC and
consequently, to inform relevant numerical models.

Figure 4A channel flow experiment showing the bridging of 40/50 mesh ceramic sand
at Bc. The flow is from left to right and the channel is uniformly tapered from left to right
with the following dimensions: L=222 mm, H=20.1 mm, WL=3.98 mm, and WR=0.25 mm.

Cases with closely sieved proppants with narrow size distributions, such as those shown in Fig. 4,
are well-suited for model calibration and validation studies; in reality, however, proppants are most often
pumped with a much broader size distribution. Fig. 5 shows a case with a bimodal distribution (with
the modes approximately at 60 mesh and 40 mesh) of proppant pumped with the linear gel G1. The
multiple bridging locations are immediately obvious between the larger mode diameter and the smaller mode
diameter, which also corresponds to the minimum fracture width (in this case, 0.25 mm). A comparison
with Fig. 4 shows the effect of size distribution on the bridging location. The differential pressures can also
be obtained between any two points by using a pressure transducer. Fig. 6 shows the differential pressure
between the inlet and outlet of this channel. It shows that the onset of bridging causes a sharp increase
6 SPE-184829-MS

in pressure relative to the unbridged state, and this indicates a shock-like characteristic of the bridging
phenomena, which is also evident in the formulation and results of existing bridging models (Dontsov and
Peirce 2015; Shiozawa and McClure 2016). In the current experiments, it was observed that the sharp rise in
pressure drop coincides with the onset of the first bridging event. The smaller size proppants that bridge at
the minimum width (right end of the channel, blue dotted line in Fig. 5) are transported faster, and the first
bridge is observed at that location. The differential pressure increase in this case is thus controlled by the
smaller size proppants that first reach the width representing its diameter. The blue and green dotted lines in
Fig. 5 and Fig. 6 indicate that when larger proppant grains reach their equivalent width and begin bridging
(green dotted line), the pressure had already begun to sharply increase because of the earlier bridging of
the smaller proppants. This indicates that numerical models that only consider the mean proppant diameter,
rather than a size distribution, may not able to predict the correct pressure response. The pump is turned off
sometime after the bridge begins propagating upstream, which leads to the eventual decay of the pressure
shown in Fig. 6, in which the slope of the decay is dictated by the compressibility of the system. Overall,
the experimental data supports the hypothesis that BC~1 is the correct bridging criteria inside a fracture, and
that the size distribution has a major effect on bridging and the associated pressure response.

Figure 5A channel flow experiment showing the bridging of the bimodal proppant distribution at two different time points.
The large diameter mode corresponds to 40 mesh and the small diameter mode corresponds to 60 mesh. Notice the two
primary bridging locations approximately correspond to the two modes(green dotted line indicating the large diameter mode
and the blue dotted line indicating the small diameter mode) because of the size distribution of proppant. The channel is
uniformly tapered from left to right with the following dimensions: L=222 mm, H=20.1 mm, WL=3.98 mm, and WR=0.25 mm.

Figure 6Differential pressure response to bridging across the inlet and outlet
of the tapered channel shown in Fig. 5 indicating the two time points in Fig. 5.
SPE-184829-MS 7

Settling, Resuspension and Proppant Bed. The formation of sand beds in channels has been investigated
extensively during the last few decades (Daneshy 1978, Kern et al. 1959, Woodworth and Miskimins 2007).
One of the most important quantities that determines the settling behavior of proppants is the equilibrium
bed height under a given flow condition, which represents the proppant bed height (hb,eq) under the condition
in which the settling and resuspension processes balance one another. In this work, experiments are used
to calibrate and validate the proprietary numerical models for settling and resuspension in both Newtonian
(e.g., water) and power-law (e.g., linear gels) fluids for both sand and ceramic proppants. A comprehensive
set of experiments of the settling of sand and ceramic proppants in water and slickwater was performed by
Patankar et al. (2002); they reported their measurement of equilibrium bed height, which is the condition at
which settling and resuspension balance one another, and this bed height remains constant under a constant
flow rate. This study builds on that data set using the current setup and obtains equilibrium bed height data
for two different linear gels (G1, G2), in addition to water. Although this paper shows a small number of
representative data for linear gels, the goal is to build a database of equilibrium bed heights in power-law
fluids against which numerical models can be calibrated and/or validated. In the following sections, data
from two different experimental set-ups are used to compare against numerical model results. First, we use
some selected data from the experiments by Patankar et al. (2002), which considers proppant dynamics in
water and slickwater. Next, we use the experiments described in this paper and consider the equilibrium bed
height of proppant in linear gels G1 and G2, in addition to water.
This section shows the performance of the settling and resuspension model against the existing database
provided by Patankar et al. (2002). In their experiments, proppant and fluid were separately pumped into
a slot with constant flow rates (denoted by Qp and Qf respectively), and the equilibrium bed height was
recorded. Patankar et al. (2002) reports two equilibrium heights: one for the immobile proppant bed, and
another for the mobile bed that forms on top of the immobile bed. For the purposes of comparison with
our model, we have chosen to use the average of those two heights from that paper. They also used water,
slickwater, and high temperature water for their experiments, and their reported data for the density (f)
and viscosity(f) are used to run our model (Table 1 and Table 2). They used a variety of proppants, and
their reported proppant properties (diameter dp, density p) are used. Table 1 reproduces some data points
selected from that paper; it also shows the results from our numerical model under identical conditions. Fig.
7 demonstrates that an excellent overall agreement with the data was obtained. Table 2 reproduces some
data points for the case of pure bed erosion (i.e., only fluid pumped over an existing proppant bed), and
Fig. 8 shows the corresponding comparison with our model. In general, very good agreement was obtained
with both of these data sets.

Table 1List of experimental cases with bed erosion and deposition taken from Patankar
et al. (2002) (proppant trade names excluded) and used for the numerical model validation
3
shown in Fig. 7. All of these cases use water with density f=1000 kg/m and viscosity f=1 cp.

Proppant Name Case Number Qf(m3/s) dP(m) P(kg/m3) QP(m3/s) (Model)

16/30 Ceramic 1 306.6E-6 9E-4 2710 22.3E-6 0.926 0.922


16/30 Ceramic 2 306E-6 9E-4 2710 5.6E-6 0.903 0.885
16/30 Ceramic 3 316E-6 9E-4 2710 4.2E-6 0.895 0.868
16/30 Ceramic 4 309.7E-6 9E-4 2710 2.8E-6 0.882 0.855
20/40 Sand 1 314.2E-6 6E-4 2650 5.7E-6 0.895 0.88
20/40 Sand 2 312.9E-6 6E-4 2650 1.4E-6 0.834 0.805
20/40 Sand 3 314.8E-6 6E-4 2650 11.4E-6 0.911 0.906
20/40 Sand 4 311.6E-6 6E-4 2650 0.4E-6 0.811 0.731
8 SPE-184829-MS

Table 2List of experimental cases with bed erosion only taken from Patankar
et al. (2002) and used for the numerical model validation shown in Fig. 8.

Proppant Size f(kg/m3) f(kg/m-s) Qf(m3/s) dp(m) p(kg/m3) (Model)

60/40 999 1.115E-3 133.295E-6 3.42E-4 2650 0.816 0.706


20/40 999 1.115E-3 227.542E-6 5.6E-4 2650 0.731 0.625
16/20 998 1E-3 258.642E-6 9.49E-4 2730 0.675 0.701
16/20 972 0.378E-3 232.588E-6 9.49E-4 2730 0.744 0.772
16/30 999 1.115E-3 100.681E-6 8.84E-4 3450 0.885 0.899
12/20 998 1.015E-3 155.185E-6 10.9E-4 2650 0.810 0.83

Figure 7Comparison of numerical model predictions with selected data from Patankar et al. (2002) listed in Table 1.

Figure 8Comparison of numerical model predictions with selected data from Patankar et al. (2002) shown in Table 2.

These comparisons can be augmented with data from the experimental setup described in this paper.
The slurry is pumped at a certain proppant concentration with a certain flow rate, and the formation of
proppant bed is observed until it reaches equilibrium. After proppant enters the channel, the entry effect
SPE-184829-MS 9

creates a recirculation zone in which velocities are higher, and the proppant bed builds up gradually. After
the entrance length is exceeded, the proppant bed builds up until settling and resuspension balance one
another, and equilibrium bed height is reached, as shown in Fig. 9. At this point, all incoming proppant is
transported further into the channel where it begins to settle and build the bed. Before reaching equilibrium,
the farthest distance that proppant can be transported is determined by the steepest angle of the proppant
bed that the carrier fluid can support. This is similar to the "angle of repose," which is the steepest angle
that a particle bed can support if it is allowed to fall freely in a fluid, although in this case, it is a dynamic,
rather than a static, quantity. Typically, numerical models ignore such effects, which can underestimate the
transport length of proppant.

Figure 9Formation of proppant bed up to equilibrium in a channel of dimensions: L=25 mm, H=20 mm,
wL=wR=2.6 mm with flow going from left to right. This is the case with linear gel G1 and 50 mesh ceramic proppant.

The equilibrium bed height obtained from the experiments can be measured and compared with model
predictions. Table 3 shows the data from the experiments, along with the model predictions. Fig. 10 plots the
data from Table 3 separately for water and for linear gels G1 and G2, showing very good agreement between
the model and experimental data. This is further quantified in Fig. 11 in which the data are compared to the
ideal case of perfect match (a "y = x" line), and show a R2 = 0.62.

Table 3List of experimental cases from the current work and the corresponding
numerical model results. The comparison is shown graphically in Fig. 10.

Fluid Name + Proppant Name Case (fkg/m3) f(kg/m-s) dp(m) P(kg/m3) Qs(m3/s) P(kg/m3) (Model)
Number

Water + 100 mesh sand 1 1E3 1E-3 1.5E-4 265 0 2E-6 240 0.9285 0.930
Water + 100 mesh sand 2 1E3 1E-3 1.5E-4 2650 2.67E-6 240 0.898 0.915
Water + 100 mesh sand 3 1E3 1E-3 1.5E-4 2650 3.33E-6 240 0.880 0.901
Linear gel G1 + 100 mesh sand 1 1E3 - 1.5E-4 2650 1E-6 240 0.901 0.900
Linear gel G1 + 100 mesh sand 2 1E3 - 1.5E-4 2650 2E-6 240 0.810 0.860
Linear gel G1 + 50 mesh ceramic 3 1E3 - 2.97E-4 2710 2E-6 240 0.850 0.904
Linear gel G2 + 100 mesh sand 4 1E3 - 1.5E-4 2650 0.67E-6 240 0.803 0.810
Linear gel G2 + 50 mesh ceramic 5 1E3 - 2.97E-4 2710 1E-6 240 0.841 0.845
Linear gel G2 + 50 mesh ceramic 6 1E3 - 2.97E-4 2710 2E-6 240 0.801 0.775
10 SPE-184829-MS

Figure 10Results for equilibrium bed height compared with


model predictions for the experimental cases described in Table 3.

Figure 11Evaluation of the agreement between experimental data and model


results. The green line is the ideal case where they would match perfectly.

Conclusions
This paper proposes an experimental method that can be used to generate high fidelity data on various
aspects of proppant dynamics in hydraulic fracturing. 3D printing technology was successfully used
to rapidly manufacture a variety of channel configurations. These channels are then used, along with
laboratory-scale blenders and pumps, to visualize and quantify the dynamics of proppants in fracturing
fluids. This paper demonstrates that with uniform width and tapered channels, the experimental data can be
used to investigate proppant transport, bridging, settling and resuspension. The experimental data supports
a bridging criteria of BC~1 inside smooth wall fractures and shows the critical role of the proppant size
distribution in determining the pressure response attributable to bridging. Apart from gaining valuable
insight about proppant placement, the data obtained from these experiments can be used to calibrate and
validate numerical models for predicting proppant dynamics. An existing data set, as well as the data
generated from the current experiments, were used to validate proprietary numerical models pertaining
to proppant settling and resuspension. The model results show good agreement, not only in water and
slickwater, but also for linear gels, indicating the validity of the model for both Newtonian and non-
Newtonian fluids. From the initial sets of experiments presented in this paper, it is clear that the proposed
SPE-184829-MS 11

methodology can be used to create a database, for not only investigating any phenomena related to proppant
dynamics, but also to calibrate and validate numerical models to be used for hydraulic fracturing stimulation
design.

Acknowledgments
The authors gratefully acknowledge Halliburton Energy Services, Inc. for the permission to publish this
work. We also thank Travis Larsen and Chris Parton for assistance with the experimental set-up and 3D
printing.

Nomenclature
hb,eq:
Equilibrium bed height
L:Length of channel
H: Height of channel
WL: Width at left end of channel
WR: Width at right end of channel
hf:Fracture height
Pf:Fluid density
Pp: Proppant density
Qf: Fluid volume flow rate
Qp: Proppant volume flow rate
Qs: Slurry volume flow rate
p: Proppant concentration
dp:Proppant diameter
G1: Guar-based linear gel with base gel viscosity of 6.5 cp (@511 s1, and power-law
parameters n' = 0.692,k' =0.038 Pa-sn'.
G2: Guar-based linear gel with base gel viscosity of 11 cp (@511 s1, and power-law
parameters n' = 0.585,k' =0.128 Pa-sn'

References
Barree, R.D. and Conway, M.W. 1994. Experimental and Numerical Modeling of Convective Proppant Transport.
Presented at the SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, USA, 25-28 September.
SPE-28564-MS. doi:10.2118/28564-MS.
Barree, R.D. and Conway, M.W. 2001. Proppant Holdup, Bridging, and Screenout Behavior in Naturally Fractured
Reservoirs. Presented at the SPE Production and Operations Symposium, Oklahoma City, Oklahoma, USA, 24-27
March. SPE-67298-MS. doi:10.2118/67298-MS.
Betancourt, F., Concha, F., and Uribe, L. 2015. Settling Velocities of Particulate Systems Part 17. Settling Velocities of
Individual Spherical Particles in Power-Law Non-Newtonian Fluids. International Journal of Mineral Processing,
143: 125-130.
Blyton, C.A., Gala, D.P., and Sharma, M.M. 2015. A Comprehensive Study of Proppant Transport in a Hydraulic
Fracture. Presented at the APE Annual Technical Conference and Exhibition, Houston, Texas, USA, 28-30 September.
SPE-174973-MS. doi:10.2118/174973-MS.
Clark, P.E. and Quadir, J.A. 1981. Prop Transport In Hydraulic Fractures: A Critical Review Of Particle Settling Velocity
Equations. Presented at the SPE/DOE Low Permeability Gas Reservoirs Symposium, Denver Colorado, USA, 27-29
May. SPE-9866-MS. doi:10.2118/9866-MS.
Daneshy, A.A. 1978. Numerical Solution of Sand Transport in Hydraulic Fracturing. JPT 30 (01): 132-140. SPE-5636-
PA. doi:10.2118/5636-PA.
Dontsov, E.V. and Peirce, A.P. 2015. Proppant Transport in Hydraulic Fracturing: Crack Tip Screen-Out in KGD and P3D
Models. International Journal of Solids and Structures, 63 (15): 206-218.
12 SPE-184829-MS

Dunand, A. and Soucemarianadin, A. 1985. Concentration Effects on the Settling Velocities of Proppant Slurries.
Presented at the SPE Annual Technical Conference and Exhibition, Las Vegas, Nevada, USA, 22-26 September.
SPE-14259-MS. doi:10.2118/14259-MS.
Gadde, P.B., Liu, Y., Norman, J., Bonnecaze, R. et al. 2004. Modeling Proppant Settling in Water-Fracs. Presented at
the SPE Annual Technical Conference and Exhibition, Houston, Texas, USA, 26-29 September. SPE-89875-MS.
doi:10.2118/89875-MS.
Gruesbeck, C. and Collins, R.E. 1982. Particle Transport Through Perforations. SPE J 22 (06): 857-865. SPE-7006-PA.
doi:10.2118/7006-PA.
Kern, L.R., Perkins, T.K., and Wyant, R.E. 1959. The Mechanics of Sand Movement in Fracturing. JPT 11 (07): 55-57.
SPE-1108-G. doi:10.2118/1108-G.
Liu, Y. 2006. Settling and Hydrodynamic Retardation of Proppants in Hydraulic. PhD dissertation, The University of Texas
at Austin, Austin, Texas, USA (August 2006).
Machac, I. and Lecjaks, Z. 1995. Wall Effect for a Sphere Falling through a Non-Newtonian Fluid in a Rectangular Duct.
Chemical Engineering Science, 50 (1): 143-148.
Malhotra, S. and Sharma, M.M. 2011. A General Correlation for Proppant Settling in VES Fluids. Presented at the
SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 24-26 January. SPE-139581-MS.
doi:10.2118/139581-MS.
Missirlis, K.A., Assimacopoulos, D., Mitsoulis, E., and Chhabra, R.P. 2001. Wall Effects for Motion of Spheres in Power-
Law fluids. J. Non-Newtonian Fluid Mech., 96 (3):459471.
Nakagawa, H. and Tsujimoto, T. 1980. Sand Bed Instability Due to Bed Load Motion. ASCE Journal of the Hydraulics
Division, 106 (12): 2029-2051.
Patankar, N.A., Joseph, D., Wang, J., Barree, R., et al. 2002. Power Law Correlations for Sediment Transport in Pressure
Driven Channel Flows. International Journal of Multiphase Flow, 28 (8): 1269-1292.
Roy, S., Du Frane W. L., Kanarska, Y. et al. 2016. Numerical and Experimental Studies of Particle Settling in Real Fracture
Geometries. Rock Mechanics and Rock Engineering:1-13. http://dx.doi.org/10.1007/s00603-016-1100-3.
Shiozawa, S. and McClure, M. 2016. Simulation of Proppant Transport with Gravitational Settling and Fracture Closure
in a Three-Dimensional Hydraulic Fracturing Simulator. Journal of Petroleum Science and Engineering 138: 298-314.
Sahai, R., Miskimins, J.L., and Olson, K.E. 2014. Laboratory Results of Proppant Transport in Complex Fracture Systems.
Presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, USA, 4-6 February.
SPE-168579-MS. doi:10.2118/168579-MS.
Smith, J. 1991. Predicting Proppant Bridging Improves Fracture Designs. World Oil, 7 (4): 51-59.
Song, D., Gupta, R.K., and Chhabra, R.P. 2009. Wall Effects on a Sphere Falling in Quiescent Power Law Fluids in
Cylindrical. Ind. Eng. Chem. Res. 48 (12): 58455856.
Van Rijn, L. 1984. Sediment Pick-Up Functions. Journal of Hydraulic Engineering, 110 (10): 1494-1502.
Woodworth, T.R. and Miskimins, J.L. 2007. Extrapolation of Laboratory Proppant Placement Behavior to the Field
in Slickwater Fracturing Applications. Presented at the SPE Hydraulic Fracturing Technology Conference, College
Station, Texas, USA, 29-31 January. SPE-106089-MS. doi:10.2118/106089-MS.
Wu, H., Madasu, S., and Lin, A. 2014. A Computational Model for Simulating Proppant Transport in Wellbore
and Fractures for Unconventional Treatments. Presented at the Abu Dhabi International Petroleum Exhibition and
Conference, Abu Dhabi, UAE, 10-13 November. SPE-171739-MS. doi:10.2118/171739-MS.

Вам также может понравиться