Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

SPE-199310-MS

A Concise Review of Experimental Works on Proppant Transport and Slurry


Flow

Morteza Roostaei, RGL Reservoir Management Inc.; Alireza Nouri, University of Alberta; Seyed Abolhassan
Hosseini and Mohammad Soroush, RGL Reservoir Management Inc., University of Alberta; Arian Velayati,
University of Alberta; Mahdi Mahmoudi, RGL Reservoir Management Inc.; Ali Ghalambor, Oil Center Research
International; Vahidoddin Fattahpour, RGL Reservoir Management Inc.

Copyright 2020, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE International Conference and Exhibition on Formation Damage Control held in Lafayette, Louisiana, USA, 19-21
February 2020.

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
Final proppant distribution inside hydraulic fractures which depends on particle properties, movement and
deposition highly impact wellbore productivity and consequently is crucial in modeling and design of
hydraulic fracturing. This paper presents a thorough review of laboratory scale tests performed on proppant
transport related to hydraulic fracturing treatments and governing physics behind its mechanisms.
The interaction between fluid (gas and liquid) and solid particles has been investigated in applied
mathematics and physics. In such phenomena, there is always a relative motion between particles and fluids.
In this work this relative motion during proppant movement, sedimentation and fluidization in both small-
and large-scale lab tests have been assessed in detail. Existing correlations which relate proppant particles
settling velocity to concentration of proppant particles, fracture wall and inertia effect in Newtonian and
non- Newtonian fluid are presented as well.
Lab tests show that various parameters determine the proppant particles distribution inside the fractures.
Particle settling velocity, an influential parameter in this regard, is impacted by fracture walls, inertia
and the presence of other particles. Inertia changes the relation of drag coefficient and Reynold number.
Fracture wall and particles concentration decrease settling velocity as drag force increases. At a certain level,
concentration reaches to its limit. Proppant concentration, in addition, increases the suspension viscosity,
fracture width and net pressure. However, it deceases the fracture length as more pressure loss occurs along
the fracture. As a result, well productivity is highly impacted by the proppant settling and distribution.
Many studies have been devoted to identifying different aspects of hydraulic fracturing and proppant
transport mechanisms in porous media. This study highlights the key parameters and their effects, existing
correlations and physics behind them for better understanding and management of this mechanism.
Keywords: Proppant Transport, Hydraulic Fracturing, Settling Velocity, Slurry Flow
2 SPE-199310-MS

Introduction
Basic Principles of fluid-particle (gas bubbles, liquid droplets, and solid particles) interactions have been
investigated in applied physics and mathematics for a long time. For example, several processes such as
flotation, distillation, and absorption in chemical engineering, which are based on bubbles and droplets or
fluidized catalytic processes, rely on solid particles. The hydrodynamics and operation of all these processes
are highly influenced by the motion of particles and flow medium relative to each other. This paper focuses
on the solid-liquid interactions which are valid during the transport of proppants inside hydraulic fractures
and an extensive literature review has been performed on the related experiments.
Proppants create a high permeability pack inside the created fracture which help to avoid fracture closure
after the shut-down of pump and to increase wellbore productivity. This permeable pack facilitates the fluid
flow through the well by providing higher permeability compared to the formation. Moreover, proppant
particles need to have enough compressive strength as they will be subjected to high in-situ stresses
underground and resistance to motion during the flowback. Hence, the well productivity highly relies on
the characteristics of proppant bed, transport, placement, and deposition
Accurate determination of settling velocity for proppant transport is still an area of study in the literature.
This is despite the fact that the significant aspects of particle motion like fluidization and sedimentation
have been investigated in other areas for many years. Although such research results can be applied to the
proppant application area without any compromise, they are neglected in proppant-related research studies.
The current study attempts to provide such a comprehensive review.
Throughout this paper, unless stated otherwise, it is assumed that the continuous phase (surrounding
liquid) is a Newtonian fluid. Also, the dispersed phase is assumed to be rigid particles which can resist
high normal and shearing stresses without considerable deformation. Another assumption is that the solid
particles have a uniform size. Graded proppant particles are used in practice, which can make the problem
more complicated.

Experimental Studies on Proppant Transport


In early experimental studies in the area of proppant transport, researches focused on settling velocity of
falling particles inside fractures. Kern et al. (1959) experimentally investigated sand movement in a segment
of a vertical fracture, formed between two plates. The results showed that only if the rate of injection per
foot of was high, proppant bed formation at the bottom of fracture will not take place. Additionally, large
particles should be injected during the first stage of the operation. Babcock et al. (1967) identified proppant
terminal velocity (i.e. equilibrium velocity) and formation of proppant bed as two key factors influencing
proppant distribution. Clark et al. (1977) investigated how the bank build-up is influenced by the type and
viscosity of the carrying fluid, proppant concentration, and flow rate. Their experimental slot model did not
allow the leak off of the fluid. Schols and Visser (1974) using a similar slot model and injection of low-
viscosity fracturing fluid identified three sequential stages in the formation of proppant bed. In the initial
stage, proppant bed gradually grows to an equilibrium height after which the height growth is stopped due
to bed erosion in the direction of flow. In the second stage, the height of the bed grows to this final height
developed in the first stage, over the entire bed. During last stage, eventually, proppant particles saltate
further in the flow direction.
After 1980, the effects of convective flow, momentum transfer between solid and liquid, fracture
containment and presence of multiple particles on proppants settling drew significant attention. Proppant
convection happening during the stages with heavier proppants is reported to be the dominant influencing
factor, even more important than proppant settling in uncontained vertical hydraulic fractures (Cleary and
Fonseca, 1992). Barree and Conway (1994) designed and commissioned one of the earliest experimental
facilities for slurry transport in a slot as a part of their study to develop a numerical transport simulator.
SPE-199310-MS 3

Their results showed that the convection transport of proppants generates settling velocities that are several
orders of magnitude higher compared to terminal velocity of single proppant.
Al-quraishi and Christiansen (1999) used small glass models to study different flow conditions. They
reported that even the smallest density difference may contribute to considerable convection. Clark and Zhu
(1996) proposed a dimensionless group called Buoyancy number based on the results of a series of slot
experiments which gives an indication of the dominance of convection for various rheology of fluids.
Brannon et al. (2006) studied a wide range of proppant transport parameters in large-scale slot flow tests.
They developed an empirical model that could estimate the propped fracture length. It was claimed that the
model could be used to optimize the effective facture length and well performance by altering properties
of proppant and carrying fluid.
In real field operations, the mass concentration of injected proppants can be between 2 to 3 lb/gal
(0.24-0.36 kg/l) which is equivalent to volume concentration of 0.1-0.15 Vol/Vol (Novotny, 1977). However,
during the movement of proppants, this concentration may increase to up to 11 lb/gal, i.e. a volume
concentration of around 0.5 (Adachi et al., 2007).
The transport of proppant particles is due to the flow of continuous phase and also the downward force
of gravity. Sediment transport principles have been investigated in other applications besides the petroleum
industry. Although in the study of sediment transport, different methods have been reported in the literature,
fundamental characteristics of transport due to the fluid flow stay the same. The proposed models can
predict proppant transport behavior in hydraulic fractures without loss of generality. The central goal of
this investigation is to provide a comprehensive literature survey on experimental work carried out on the
fluid-solid systems. The concepts presented in this paper are gathered in a way that can be included in any
proppant transport numerical tool to capture physical parameters that govern proppant transport and settling
inside hydraulic fractures.

Inertia Effect on Proppant Settling


The continuous phase in suspensions exert a hydrodynamic drag force on the suspending particles which
depends on the flow regime. According to the value of Re, three different flow regimes, namely "Stokes"
or "creeping flow", "intermediate", and "Newtonian" regime are distinguished (Clift et al., 1978). There are
theoretical, empirical and numerical models in the literature which describe drag force or drag coefficient
at different velocities. Theoretical models of drag coefficient only exist in the Stokes regime (Clift et al.,
1978; Khan and Richardson, 1987). However, several empirical models have been proposed for various
ranges of Reynolds numbers. This section presents the drag coefficient expressions for different flow
regimes proposed to determine drag force. We do not intend to develop a new approach nor to compare the
performance of the equations presented in the literature. The objective is only providing a simple technique
based on the existing drag coefficient correlations to easily obtain the terminal velocity of a suspended solid
particle over the entire Reynolds number range.
Navier-Stokes equations are derived by applying Newton's second law (which governs the movement
of suspending solids in fluids) to a small control volume of an incompressible fluid. An exact analytical
solution has not been provided for the Navier-Stokes equation in general form, and exact analytical
solutions are only valid under very strict conditions with simplifying assumptions. Consequently, numerical
approaches are generally used to achieve a solution. The Navier-Stokes equation is defined as follows (Clift
et al., 1978; Batchelor, 2000):
(1)
The left-hand side of the equation can be viewed as the multiplication of mass and acceleration with the
local derivative term representing changes in time at a fixed position and the convective term corresponding
to changes as a result of fluid movement (Bird et al., 1976; Schlichting, 1979). On the right-hand side,
4 SPE-199310-MS

there are terms related to gravity and surface forces exerted on the unit volume of fluid. (Bird et al., 1976;
Schlichting, 1979).
Navier-Stokes equations should be solved for velocity components and pressure. Therefore, the
continuity equation, which is the simple form of the conservation of mass for incompressible fluids, should
also be considered as follows:
(2)

Drag Force at Low Reynolds Numbers


Certain assumptions should be applied to Eqs. 1 and 2 to obtain an analytical solution. These assumptions
depend on the flow regime and the dominance of either viscosity or inertia terms which have led to, for
example, concepts of idealized inviscid flow (in which inertia terms are dominant and is applicable at high
Re) or the boundary layer theory (in which except at a very thin layer of fluid close to the boundary, viscosity
terms are ignored)
The most logical simplifying assumption in solving Eq. 1 and 2 is the creeping flow estimation which
is only valid at very low Re and has broad applications in flow in porous media. At this condition, the
convective term is negligible in the Navier-Stokes equation. Therefore, fluid inertia is completely neglected
compared to the viscosity term (Clift et al., 1978). In this section, the Stoke's flow regime estimation of
the Navier-Stokes equations is introduced and followed by more advance discussion of proppant settling
velocity correlations.
The exact solution to the partial differential equation in creeping flow regime, in terms of force of drag,
FD acting on a sphere gives (Stokes, 1851; Batchelor, 2000):
(3)
The drag coefficient, on the other hand, is defined as drag force acting in a projected cross-section
perpendicular to the fluid movement which is a circle. Therefore:

(4)

By reformatting Eq. 4, two famous dimensionless groups appear as:

(5)

Re appears on the right-hand side of the equation while the definition of dimensionless drag coefficient
is appeared on the other side of the equation.

(6)

As stated before, Eq. 6 is valid only in the Stokes region where the Reynolds number is 0.2 or less, and
the error is approximately 4% (Clift et al., 1978).
In proppant movement in hydraulic fractures, the drag force exerted on particles is of lesser importance
compared to their settling velocity and therefore, drag coefficient expressions should somehow be expressed
in terms of settling velocity. To this purpose, the general 1D momentum equation in terms of force per unit
volume can be written for proppants as: (Wallis, 1969):
(7)
where b and f embody any sort of body force or surface force, p represents pressure, ρp is proppant density
and v is terminal velocity. Assuming steady conditions, v will be constant and therefore, the left side of Eq.
7 will be zero. Gravity is the only active body force and hydrodynamic drag is the only involved surface
SPE-199310-MS 5

force since for suspended particles no contact force exists. Applying these assumptions and Multiplying
Eq. 7 by the proppant volume, one can obtain the force balance equation of a single suspended particle at
steady state condition as:
(8)
where Vp represents volume of proppant particle and FD is the hydrodynamic drag. Under equilibrium
conditions, the pressure gradient along the proppant carrying fluid will be:

(9)

After Substituting Eq. 9 into the momentum equation (Eq. 8):

(10)

Finally, the hydrodynamic drag force, FD, per unit projected area (circular cross-section) at 90° angle to
the direction of fluid motion exerted on a spherical proppant particle when it is moving at terminal velocity
is:

(11)

The implication of Eq. 11 is that the drag force should balance out the difference between weight of the
particle and buoyancy exerted on the particle. Therefore, based on Eq. 6:

(12)

The corresponding terminal velocity in the viscous regime is obtained by substituting Eq. 12 into Eq. 6.

(13)

Obtaining the settling velocity from the relationship between CD and Re is not always this simple, and
most of the time, highly non-linear equations are used for CD-Re relationship. In the current study, this topic
will be discussed in detail, and another dimensionless group will be introduced, which is more useful than
the drag coefficient for proppant transport applications.
To obtain the relationship between CD and Re, which is expressed in Eq. 6, it was assumed that in the
Navier-Stokes equation, viscosity terms are dominant over inertia terms. In this approach, a substantial
problem, as first addressed by Oseen (1910), is that the aforementioned approximation is not valid at large
distances from the proppant where inertia terms cannot be neglected, and this is true for any value of
Reynolds number. The validity of Eq.6 is only for distance in the order of dp/Re away from the particle.
To overcome this problem, Oseen (1910) provided a solution by simplifying the Navier-Stokes problem
instead of ignoring inertia terms and gained larger drag forces compared to the one obtained from Stokes
law value, applicable for Re < 0.1. Oseen's expression is as follows:
(14)
To expand the validity of Oseen's solution to larger Re, various expressions were proposed to ignore
complexities of the Navier-Stokes equation, including one series solution with twenty-four terms (Dyke,
1970). As one more instance, Goldstein (1929) suggested an equation with Re up to the power of 5:
6 SPE-199310-MS

(15)
Although the proposed solutions represent the Oseen's drag precisely, the Oseen's solution to drag is an
estimate to the actual drag. It is noticeable that as Re increases, all the Oseen's based or Goldstein based
analytical equations deviate from the experimental data. Lewis and Carrier (1949) used a parameter named
CL in Oseen's equation that is a weak function of Re, to modify the solution:
(16)
Proudman and Pearson (1957) developed a stream function by employing Stokes' estimation in the
vicinity of the particle and Oseen's solution far from the particle to consecutively approximate Navier-
Stokes equation:
(17)
Chester et al. (1969) modified the approximation by decreasing the order of error:
(18)
where γ is 0.5772157.
Using a similar methodology, Chester et al. (1969) suggested below equation, as well:
(19)
where:

(20)

Liao's (2002) analytical expression approximates the solution to Navier-Stokes equations up to 10th order
and is valid for steady, viscous flow over a spherical particle for Reynolds numbers up to 30, which is wider
than the previous theoretical solutions.
The proposed equations were plotted in Fig. 1, except for Eq. 19, which does not provide CD in terms
of Re explicitly and requires trial and error.

Figure 1—Comparing proposed equation for drag coefficient in creeping flow regime
SPE-199310-MS 7

Figure 1 shows all these expressions do provide a significant improvement over Stokes equation to
determine the drag coefficient. Moreover, they diverge quickly at higher Re>1. Therefore, Empirical
correlations should be used for drag coefficient calculation for flow around particles at higher Reynold
numbers.

Drag at High Reynolds Number


When the Reynolds number of the system is more than 1, the solid-flow interaction described by analytical
techniques deteriorates. Depending on the Reynolds number and flow regime, either experimental or
numerical simulation of flow field are more valid. Numerical models of steady, axisymmetric flow are very
popular for intermediate Reynolds numbers (1 <Re <1000) (Clift et al., 1978).
Various types of expressions obtained through regression on experimental data of terminal falling
velocities have been proposed (Lapple and Shepherd, 1940; Tanaka et al., 1970; Vlajinac and Covert, 1972).
Founded on the experimental work of Allen (1900), Wieselsberger (1922), Prandtl and Tietjens (1931), and
Schiller (1932), the standard drag curve which presents CD vs Re on logarithmic axes, has been well put
in place (Fig. 2).

Figure 2—CD vs Re in different flow regimes

As mentioned earlier, only three theoretical methods have been presented to calculate CD: Stokes, Oseen
and the Goldstein laws. All of these expressions are only accurate at low values of Re. Several empirical
expressions have been provided in the literature based on this graph with different complexities and
various constant parameters. Most of these empirical equations are listed in Table 1. Clift et al. (1978), by
subdividing the Reynolds interval shown in Fig. 2 into 10 subintervals, proposed a correlation which agree
very well with the experimental data of Fig. 2.
8 SPE-199310-MS

Table 1—Empirical equations of CD


SPE-199310-MS 9
10 SPE-199310-MS
SPE-199310-MS 11

Even though the development of advanced computers has contributed to the expansion of several
complicated empirical models with close predictions to experimental data, most of such models are rooted
in the three aforementioned theoretical methods. As some examples of equations based on Stokes' law we
can mention the work of Flemmer and Banks (1986), Turton and Levenspiel (1986), Brown and Lawler
(2003), and Cheng (2009). Also, Dou's (1981) correlation is developed based on Oseen's law.
12 SPE-199310-MS

Yang (2015) classified the empirical equations used to calculate CD according to the form of the equations
into multiplication, addition or rational type. If the Stokes or Oseen's equations are multiplied by a power
or exponential function, the multiplication equations are developed. In case the equations are added by an
increasing function, the addition type expressions are attained. Rational type relationships are according to
Goldstein law, but to keep the manuscript concise, they are not described here.
The multiplication expressions based on Stokes and Oseen's laws can be represented as:
(21)

(22)
And the addition type correlations based on Stokes and Oseen's laws can be represented as:
(23)

(24)
where f1 (Re), f2 (Re) and f3 (Re) are experimentally determined functions. Table 2 presents Yang's (2015)
expressions proposed for Re < 2×105.

Table 2—Yang's (2015) equations to determine the drag coefficient


SPE-199310-MS 13

Non-Newtonian Fluids
The hydrodynamic knowledge about the particles in Newtonian liquids has reached an adequate level.
It is now possible to provide predictions about the viscous, inertial, and surface tension forces for
dispersed particulate objects in Newtonian fluids. However, for non-Newtonian fluids, there is no certain
intuition, and consequently, many efforts were made to investigate the impact of non-Newtonian rheological
characteristics on the motion of solid particles in liquid (Zenit and Feng, 2018). In this section, some
studies on the fundamental understanding of hydrodynamic behavior of non-Newtonian fluids have been
summarized. Most of the researches used an experimental approach to investigate the behavior of non-
Newtonian fluids because of the complexity of the problem (Zenit and Feng, 2018).
Various models can be employed to represent the performance of non-Newtonian liquids. Nevertheless,
the most common models are the Power Law and Bingham Plastic models.
Based on the Bingham Plastic model, shear stress is a function of yield stress and plastic viscosity as:
(25)
where τy is the yield stress, is the shear rate and κ is the plastic viscosity.
Also, the Power Law model is represented by the following equation:
(26)
with K being the consistency index and n isbeing the power-law index.
Metzner and Reed (1955) defined the Reynolds number for Power Law fluid over a sphere as:

(27)

Chhabra, 2006 proposed that the same relationship between CD and Re that exists for Newtonian fluids
can be extended to Non-Newtonian fluids by including a correction component Y which depends on index n:
(28)
Hirose and Moo-Young (1969) and Acharya et al. (1976) provided Eq. 29 to determine the correction
factor Y:

(29)

Later, Kawase and Moo-Young (1986) studied this model again and presented the following equation to
determine the drag correction factor:

(30)

Betancourt et al. (2015) represented the following model which was based on the data of Tripathi et al.
(1994) and Dazhi and Tanner (1985) for the creeping flow over a sphere:
(31)
At higher Reynolds number, drag coefficient can be determined from Reynolds number (Eq. 32) based
on the experiments of the settling of spheres in non-Newtonian fluids (Chhabra, 2006).
(32)
Chhabra and Peri (1991) developed an explicit model to determine the non-Newtonian Archimedes
number (Ar), as a function of the CD and Re:
14 SPE-199310-MS

(33)

where Re = A(Ar)B, and .


Kelessidis (2004) proposed a non-dimensional particle velocity based on CD, Re and dimensionless
particle diameter.

(34)

where , K1 = 0.428 and K2 = 0.824.


Shah et al. (2007) developed an explicit equation based on a dimensionless term as a function of Reynolds
number which obtained by regression of 391 experimental points.
(35)
where α = 6.9148 n2 − 24.838 n + 22.642 and β = −0.5067 n2 + 1.3234 n − 0.1744
Overall, there are many variations of Drag coefficient versus Reynolds number in the literature which
are ignored in this study to keep the manuscript concise.

Settling of Single Particle in Infinite Medium


During the movement of proppants, particle settling velocity is one of the most important parameters that
need to be determined. However, almost all the suggested relationships are between Reynolds number (Re)
and drag coefficient (CD) in the studies of fluidized bed expansions (e.g. Clift et al., 1978; Liao, 2002; Yang
et al., 2015). Terminal velocity of particles cannot be easily obtained from such relationships since settling
velocity exists in the definition of both Re and CD. Moreover, in certain ranges of Reynolds number a simple,
explicit equation between these two parameters does not exist.
In this section, it is attempted to write the proposed equations in terms of terminal velocity. At Re<0.1,
as mentioned earlier:
(36)
and

(37)

The terminal velocity in this range of Re was given in Eq. (13). However, for 750 Re < 3.5 × 105, CD is
constant (CD = 0.445 according to Clift et al. 1978) and, thus, terminal velocity can be calculated by:

(38)

However, such relationship does not exist at higher Reynolds number. In other words, in the intermediate
region an iterative procedure is needed to obtain terminal velocity of particles. This suggests the idea
of defining or employing a dimensionless number which does not include terminal velocity. Such
dimensionless number exists and is called Galileo Number (Ga). According to Clift et al. (1978):

(39)
SPE-199310-MS 15

Based on Eq. 39, only physical properties of slurry system such as proppant density and diameter and
fluid density and viscosity are required to obtain Ga. Hence, for the proppant transport-related problems, it
would be more convenient to use equations expressed in terms of Ga and Re rather than CD vs Re:
(40)
(41)
In the literature there are limited equations that express the relationship between Ga and Re, however,
most of these relationships are so complex that does not provide settling velocity explicitly. Table 3 presents
some of these relationships.

Table 3—Relationships between Re and Ga


16 SPE-199310-MS
SPE-199310-MS 17

Per Table 3, Ga-Re equations are complex. A plot of Re versus Ga (Fig. 3) was generated, which can be
used to obtain settling velocity at any Reynolds number.

Figure 3—Re vs Ga
18 SPE-199310-MS

Effect of Slurry Containment on Settling


Confining fracture walls around proppant particles have a retardation impact on the particle motion and
settling in a liquid. During particle settling, the liquid medium exerts a drag force on the particles. When
the liquid medium is confined, the exerted drag force increases due to the backflow of liquid through the
gap between particles and the confining wall. The smaller the gap the higher the drag force will be. This
increase in drag force lead to smaller terminal velocity compared to the case of infinite medium. A second
effect of settling in confined condition is the modification of flow regimes boundaries due to the fact that
the presence of wall alter the onset of flow separation (start of intermediate regime).
The presence of walls can be accounted for in the analytical solutions by modifying the far field boundary
conditions and therefore, the distance between the wall and settling particle is of importance. The research
work in the literature in this regard can be categorized into settling of spheres in Newtonian fluids (Miyamura
et al., 1981; Machac and Lecjaks, 1995; Balaramakrishna and Chhabra, 1992), settling in non-Newtonian
Fluids (Lali et al., 1989; Madhav and Chhabra, 1994; Chhabra, 1996; Ataide et al., 1999; Chhabra eta l.,
1980; Song et al., 2009) or selling of irregularly shaped particles (Chhabra, 1995). Additionally, the shape
of the confining walls is considered to be influential and wall geometries other than the simplest case of
long cylindrical tube, such as parallel plates (Staben, 2003) or triangular and square cylinders (Miyamura
et al., 1981) have been investigated. All the expressions presented in this study are for spherical particles
settling in a Newtonian, incompressible fluid bounded by a long cylindrical tube.
The impact of the walls on particle settling is generally termed retardation effect or wall factor effect.
Retardation effect can be expressed in terms of drag force change (at identical terminal velocities) or in
terms of velocity change (for identical particles) or in terms of viscosity change, according to Eqs. 42, 43
and 44, respectively (Latto et al., 1973; Duduković, and Končar-Djurdjević, 1981; Iwaoka and Ishii, 1979).

(42)

(43)

(44)
Unbounded fluid viscosity in Eq. 44 can be described in terms of stokes velocity:

(45)

In some publications, the inverse of Eq. 42 to 44 is used to define the wall factor. These three ratios are
equivalent in creeping flow regime while at higher Re numbers, the relationships are complex. Defining
wall factor (fw) in terms of velocity ratio is the most relevant method for proppant transport:
(46)
where uw and u∞ are the sphere settling velocity with and without the existence of the walls, respectively.
For a long container, dimensional analysis shows that end effect is negligible (which is a valid assumption
for hydraulic fracture case). Thus, fw is only a function Re and λ (the ratio of diameters of sphere or proppant
(dp) and container (Dc)):

(47)

It should be noted that container diameter is twice the hydraulic fracture width. The boundary conditions
are: 1) at λ = 0 the settling velocity should be equal to infinite medium settling velocity and 2) at λ = 1
settling velocity must be zero.
SPE-199310-MS 19

Beyond the creeping flow region, there are only limited analytical works while the experimental studies
in this regard are numerous (Clift et al., 1978; Di Felice, 1996; Chhabra et al., 2003). According to these
experimental investigations, at very low and very high Re, retardation factor due to the walls only depends
on λ while in the intermediate regime, both λ and Re affect this parameter.
It should be emphasized that Re boundaries for fluid regimes are different for unconfined and confined
media which is a function of λ. To differentiate these boundaries, Clift et al. (1978) based on his experiments,
concluded that the transition boundary between regimes increases due to the presence of the walls.
Coutanceau (1972) (mentioned in Clift et al., 1978) reported that recirculatory wakes attached to particles
are formed with delay and consequently, the intermediate region starts when Reynolds equals:
(48)
And the upper limit of Stokes flow defined by:
(49)
As mentioned earlier, analytical solutions are only applicable in the creeping flow regime due to the
large effects of inertial effects beyond this regime. For instance, Faxen (1923) (mentioned in Clift et al.,
1978) drag force predictions using Oseen's linearization are not reliable at high Re (Fidleris and Whitmore,
1961; Happel and Brenner, 1973; Sutterby, 1973). Beyond creeping flow regime, abundance of experimental
works for free settling of spheres (McNown et al., 1948; Fidleris and Whitmore, 1961; Sutterby 1973) as
well as spheres fixed in a flowing fluid (McNown et al., 1948; Achenbach, 1974) has opened a space for
empirical equations which are indecent agreement with experiments. In Table 4 some of the fw correlations
in the literature with their range of applicability are presented.

Table 4—Wall retardation factor correlations


20 SPE-199310-MS
SPE-199310-MS 21

Chhabra et al. (2003) by reviewing a large number of experimental data of spherical particle settling in
confined Newtonian fluids suggested using Haberman and Sayre (1958) equation in Stokes flow region,
Newton (1678) expression in the turbulent region and Di Felice's (1996) expression in the intermediate
region with a total error of around 12.5% and a sharp deviation by increasing λ and/or the Re. Due to the
limited data availability for λ>0.88, Di Felice (1996) recommended to use the equations with extra caution.
In most of the experiments, zigzag or spiral movement and rotation of the particles were ignored.
The theoretical equation proposed by Haberman and Sayre (1958) was initially acceptable for λ<=0.8,
however, its applicability was extended to λ=0.9 according to later numerical investigations (Paine et al.,
1975; Bowen and Sharif, 1994; Higdon and Muldowney, 1995).
According to vast number of studies on circular and square cross-section containers, Newton's equation
(Eq. 50) is found as the most reliable one at high Re. It is usually assumed that this equation is applicable
for the entire range of λ (0<λ <1) and general cases other than cylindrical tubes.

(50)

where A in Eq. 50 is the container cross-sectional area. It should be noted that Newton's equation results
are slightly different than Munroe's Equation (1889) results.
Miyamura et al. (1981) proposed a 19th order equation for various geometries based on several
experiments:

(51)

where the coefficients are presented in Table 5.


22 SPE-199310-MS

Table 5—Coefficients used in Eq. 51 (Miyamura et al., 1981)

Fidleris and Whitmore (1961) performed a comprehensive investigation on the data of more than 3,000
velocity measurements and represented f graphically, which is re-plotted here (Fig. 4). Based on this figure,
f is the lowest at high λ and low Re values.
SPE-199310-MS 23

Figure 4—Graphical technique for wall factor determination (Fidleris and Whitmore, 1961)

Effect of Concentration on Particle Settling


There are numerous studies on settling velocity in suspensions that follow similar physics as to particle
transport in fractures and therefore, can be applied to this concept without the loss of generality. The presence
of multiple particles in a system not only increase the viscosity of the medium compared to single particle
system, but also creates a "return flow" of the medium which consequently modify the drag force and
terminal velocity of each particle. Related empirical correlations are described and classified based on their
method of derivation and also their graphical shape.
In fluidization and sedimentation, the ratio of modified terminal drag force of multi-particle system (FDt)
to that of single particle system (FDs), at the same settling velocity, is defined as concentration function:

(52)

f(c) depends on particle concentration and drag forces should be evaluated at the same velocity. In other
research areas, concentration function may be defined in other forms to calculate certain parameters related
24 SPE-199310-MS

to that specific area. For instance, in sedimentation the pressure drop through particle bed is of importance
and concentration function tries to calculate this parameter while in our investigation the focus will be on
hindered settling velocity. Di Felice (1995), by employing drag force and momentum equations, expressed
the concentration function in terms of velocity as:

(53)

where for Rep < 0.1 (Stokes regime):

(54)

Consequently:
(55)
On the other hand, for 500 <Rep <200,000 (turbulent regime):
(56)
Then:

(57)

And by substitution:
(58)
Concentration function can be estimated using theoretical, empirical and semi-empirical approaches.
In the theoretical approach, the solution of Navier-Stokes equation is obtained by including fluid-solid
interaction for very low Reynolds numbers or dilute slurries. For instance, Barnea and Mizrahi (1973)
obtained concentration function for spatially stationary particle configurations as (Eq. 59):
(59)
where 1<K<2, while Batchelor (1972) stated it for random particle arrangements (Eq. 60):
(60)
In addition, Happel (1958) suggested the following equation:

(61)

Some researchers also used theoretical approaches and extended the equations to higher Re with limited
applicability (LeClair and Hamielec, 1968; El-Kaissy and Homsy, 1973). In this regard, experimental
expressions covering the full Reynolds number are abundant; as an example, we can mention Richardson-
Zaki, (1954) correlation:
(62)
n=4.65 Ret <0.2, n=2.4 500 <Ret
Thus, for low Reynolds regime:
(63)
And for high Reynolds regime:
SPE-199310-MS 25

(64)
As noticed in Eqs. 48 and 49, both forms have almost similar power (∼-3.7) which brings this
conclusion that concentration function is independent of Re. Similar equations can be found in other
studies for intermediate flow regimes (Richardson and Jeronimo, 1979; Foscolo et al., 1983; Kmiec, 1976).
However, Khan and Richardson (1990) observed inconsistency between the mentioned generalization in
the intermediate regime.
Richardson and Zaki (1954) suggested that hindered settling velocity can be obtained from:
(65)
Where n can be estimated using Rowe (1987) correlation:
(66)
Di Felice (1995) also proposed a similar equation in which the exponent was a function of Re:
(67)
where β is estimated as follows:
(68)
Eventually, for semi-empirical approaches, certain parameters of previously developed theoretical
models are modified. Density and viscosity are among the most popular modifications in so called "pseudo-
fluid" concept. The assumption behind such modification is that by changing the viscosity and density of
slurry, the same fluid dynamic process happening in single particles systems can be applied to multi-particle
systems (Robinson, 1926; Steinour, 1944). As another example, Hawksley (1951) considered pseudo-fluid
density as suspension density with viscosity estimated by Vand equation (mentioned in Table 7), and
suggested the following equation which is only applicable for creeping flow regime:
(69)
Using a similar method, Barnea and Mizrahi (1973) stated concentration function for the entire flow
regime, where the explicit equations for low and high Re are:

(70)

(71)

Letan (1974), Ishii and Zuber (1979), Patwardhan and Chi Tien (1985a) reported other examples of
the semi-empirical approach. In should be highlighted that semi-empirical methods results are in good
agreement with experimental data for a wide flow regime range. In addition, the pressure drop-concentration
relationship in pipes was used for analogy in some research studies (Dullien, 1975; Barnea and Mednick,
1978).
26 SPE-199310-MS

Table 6—Equations relating settling velocity to particle concentration


SPE-199310-MS 27
28 SPE-199310-MS
SPE-199310-MS 29
30 SPE-199310-MS

Table 7—Equations of viscosity evolution of suspensions


SPE-199310-MS 31

Table 6 presents a summary of the equations relating the settling velocity to the concentration along with
their range of applicability.
Di Felice (1996) used a systematic approach based on the shape of u/ut versus voidage (1-c) and classified
the settling velocity correlations into four categories (Fig. 5). Figure 5 shows Case I, where the curve is
linear, and single-particle velocity is estimated at c = 0. The plot in Case 2 has two regions of different
slopes, and the extrapolation of the higher slope line approaches the single-particle velocity. In Case III, the
slope of the second line is decreasing, whereas, for Case IV, the non-linear part advances toward single-
particle terminal velocity.
32 SPE-199310-MS

Figure 5—Different categories of velocity vs concentration correlations (after Di Felice, 1996)

Random and Maximum Close Packing Concentration


Random Close Packing (RCP) is defined as the highest volume fraction of randomly packed particles.
Random close packing concentration (cRCP) is determined at a geometric point where proppant densely
packs, but not at the crystalline state. Although exact definition of randomness, crystallinity starting point,
and dependency to shear history are debatable (Torquato et al., 2000; Mewis and Wagner, 2012), some
theoretical models to determine cRCP from size of particles exist (Phan et al., 1998; Kansal et al., 2002; Farr
and Groot, 2009; Brouwers, 2011). Moreover, by increasing size dispersity of the particle arrangement,
cRCP value raises (Chong et al., 1971; Farr and Groot, 2009; Mewis and Wagner, 2012). cRCP is the most
"geometrical" consolidated packing obtained by vibrating or tapping of an arrangement of spheres (Mewis
and Wagner, 2012). It is considered a property of particles as its experimental value is greatly repeatable.
Depending on the particle sizes, cRCP value in the literature varies; for instance, a cRCP value of 0.52 (Shapiro
and Probstein, 1992) or 0.61 (Chong et al., 1971) have been measured. Koos (2009) has provided a complete
review of random close packing concentration values.
SPE-199310-MS 33

A new approach for cRCP determination has been presented by Desmond and Weeks (2014). They
presented an empirical model to determine the random close packing density as a function of polydispersity
and skewness. It shows that the skewness may affect the packing concentration significantly, and in some
cases, it has a larger effect than polydispersity. Kurita and Weeks (2010) experimentally studied the RCP
of colloidal particles and found that the fraction of crystalline particles was between 0.01% and 3%. Kang
et al. (2019) used the Computational Fluid Dynamics (CFD) approach to study random close packing
and the effect of the geometry of packing elements. They developed a packing generation procedure that
can estimate the porosity of the actual packing. Also, the experimental results were collected using a lab-
scale packed column for the validation of CFD results. The random close packing concentration, which
was obtained from CFD model, was found to be close to the experimental data. The results show that
the CFD model can predict the essential hydrodynamic characteristics of the gas-liquid flow. Gan and
Yu (2019) reported a numerical study on the effect of particle sizes and vibration on packing density by
using a Discrete Element Method (DEM) approach. They presented that larger particle sizes increase the
packing density. Also, since RCP increases when packing solid particles under one-dimensional vibration
in a cylindrical container, vibration in contrast to poured packing conditions, influence the packing structure
and characteristics. Al-Raoush and Alsaleh (2007) presented a model that mimics the random packing
concentration of polydisperse particles to represent porous medium. The result shows that the RCP for the
monodisperse and polydisperse packings were 0.5895 and 0.6212, respectively.
Two other packing parameters also exist in the literature: maximum packing concentration (cm), which
is a thermodynamic term defined as the point where suspensions become highly viscous; and random loose
packing (concentration range of 0.55-0.61), which signifies the loosest state of in-contact particles. The
values of cm are close to cRCP for the mono-disperse suspensions.

Slurries Viscosity Evolution


Suspension bulk viscosity increases as a function of particle concentration or phase volume fraction
(especially when the suspension is sheared) since each particle has an impact on the momentum transfer
between fluid and solid particles.
In low-viscosity liquids, particle diameter has a small effect on the slurry velocity. The focus of this
section is the rheology of proppant carrying suspensions, whose viscosity growth changes its mobility inside
fractures. The following assumptions are used in this study: large particles, high solid fraction (applicable to
a range of proppant properties in hydraulic fracturing), arbitrary arrangement of particles with no segregation
and highly concentrated suspensions. The assumption is that phenomena such as flocculation and colloidal
aggregation are negligible for proppant laden slurries.
Rutgers (1962) and Horri et al. (2011) have comprehensively reviewed the theoretical and semi-empirical
expressions of slurry viscosity increase as a function of solid concentration. Almost all of these relationships
share a term called "relative viscosity" defined as the ratio of slurry viscosity to that of solid-free fluid:

(72)

Einstein (1906) theoretically developed viscosity equation based on dissipated energy for dilute
suspensions of rigid particles in liquid:
(73)
where K is intrinsic viscosity equal to 2 for spherical particles. This assumption is applicable for very dilute
slurries containing microscopic size spheres. For dense dispersion systems, in the portion of slurry occupied
by the solids no energy is dissipated, and theoretical calculation overestimates the dissipated energy. Rutgers
(1962) stated that Einstein equation is valid for concentration around 0.01-0.02. Higher K values (say 4.5
or 5.5) has been reported in the work by Hatschek (1913) or Happel (1957).
34 SPE-199310-MS

Rutgers (1962) classified viscosity evolution expressions as a function of concentration, based on their
mathematical forms (only more popular expressions are presented in Table 7).
According to some equations the viscosity of the suspension approaches infinity at certain concentrations,
giving rise to the term "maximum solid fraction" (Fig. 6). This term can be misleading since particle
packs under compaction may possess higher concentration than maximum solid fraction. Properties of the
proppant pack are influenced by inter-particle interactions and the packing micro-structure. The behavior
of the packing can be characterized by such parameters as Suspension Modulus (G'). Further discussion of
the topic can be found in Shewan (2015) but is beyond the scope of this study.

Figure 6—Suspensions rheological behavior

Conclusion
In this study, a comprehensive review of experimental studies on hydrodynamics of the deposition of
solid particles was performed and the terminal velocity of particles in liquids was investigated. In general,
parameters such as containment of suspension medium, velocity of fluid, and concentration of suspending
particles are the most important controlling elements of terminal settling velocity. In summary, the following
conclusions have been made in this study:

• For a system consisting of particle and fluid, inertia impacts the drag force. Based on the existing
experimental data, we described the standard drag curve which relates drag coefficient to Reynolds
number.
• To avoid iteration in obtaining settling velocity and improve the applicability of the graph, we
graphed Galileo number vs Reynolds number. It should be noted that Galileo number can be
explicitly estimated using particle and fluid properties.
SPE-199310-MS 35

• The walls of fractures decrease settling velocity. This is more pronounced when particle sizes are
larger and close to fracture width.
• The concentration of proppant particles will decrease settling velocity. There is a maximum level
at which settling velocity approaches zero and the system is fully packed.
• At the maximum level of packing, the system behaves like porous media.

• By increasing concentration, the viscosity of the proppant slurry highly increases, which in turn
reduces the mobility and, therefore, increases the pressure drop inside the fracture.

Acknowledgements
The authors would like to acknowledge RGL Reservoir Management Inc. for the permission to publish this
paper. The financial support provided by BP through the CRD program and Mitacs through the Accelerate
program is also acknowledged.

References
Abraham, F. F. 1970. Functional Dependence of Drag Coefficient of a Sphere on Reynolds Number. Phys. Fluids 13 (8):
2194–2195. https://doi.org/10.1063/1.1693218.
Acharya, A., Mashelkar, R. A., and Ulbrecht, J. 1976. Flow of Inelastic and Viscoelastic Fluids Past a Sphere. Rheol. Acta
15 (9): 471–478. https://doi.org/10.1007/BF01530349.
Achenbach, E. 1974. The Effects of Surface Roughness and Tunnel Blockage on the Flow Past Spheres. J. Fluid Mech.
65 (1): 113–125. https://doi.org/10.1017/S0022112074001285.
Adachi, J., Siebrits, E., Peirce, A. et al 2007. Computer Simulation of Hydraulic Fractures. Int. J. Rock Mech. Min. Sci.
44 (5): 739–757. https://doi.org/10.1016/j.ijrmms.2006.11.006.
Al-quraishi, A. A. and Christiansen, R. L. 1999. Dimensionless Groups for Interpreting Proppant Transport in Hydraulic
Fractures. Presented at the Middle East Oil Show and Conference, Bahrain, 20-23 February. SPE-53262-MS. https://
doi.org/10.2118/53262-MS
Al-Raoush, R. and Alsaleh, M. 2007. Simulation of Random Packing of Polydisperse Particles. Powder Technol. 176 (1):
47–55. https://doi.org/10.1016/j.powtec.2007.02.007.
Al-Salim, Q. A. W. and Geldart, D. 1969. A New Method for the Calculation of Large Numbers of Terminal Velocities.
Powder Technol. 3 (1): 251–253. https://doi.org/10.1016/0032-5910(69)80085-9.
Allen, H. S. 1900. L. The Motion of a Sphere in a Viscous Fluid. London, Edinburgh, Dublin Philos. Mag. J. Sci. 50 (306):
519–534. https://doi.org/10.1080/14786440009463941.
Almedeij, J. 2008. Drag Coefficient of Flow around a Sphere: Matching Asymptotically the Wide Trend. Powder Technol.
186 (3): 218–223. https://doi.org/10.1016/j.powtec.2007.12.006.
Arsenijević, Z. L., Grbavčić, Ž. B., Garić-Grulović, R. V. et al 2010. Wall Effects on the Velocities of a Single Sphere
Settling in a Stagnant and Counter-Current Fluid and Rising in a Co-Current Fluid. Powder Technol. 203 (2): 237–242.
https://doi.org/10.1016/j.powtec.2010.05.013.
Ataide, C. H., Pereira, F. A. R., and Barrozo, M. A. S. 1999. Wall Effects on the Terminal Velocity of Spherical
Particles in Newtonian and Non-Newtonian Fluids. Brazilian J. Chem. Eng. 16 (4): 387–394. https://doi.org/10.1590/
S0104-66321999000400007.
Babcock, R. E., Prokop, C. L. and Kehle, R. O. 1967. Distribution of Propping Agents in Vertical Fractures. Presented at
the Drilling and Production Practice, New York, New York, 1 January. API-67-207.
Baker, F. 1913. CLXXX.—The Viscosity of Cellulose Nitrate Solutions. J. Chem. Soc., Trans. 103: 1653–1675. https://
doi.org/10.1039/CT9130301653.
Balaramakrishna, P. V. and Chhabra, R. P. 1992. Sedimentation of a Sphere along the Axis of a Long Square Duet Filled
with Non-newtonian Liquids. Can. J. Chem. Eng. 70 (4): 803–807. https://doi.org/10.1002/cjce.5450700427.
Ballesteros, R. L., Riba, J. P., and Couderc, J. P. 1982. Dissolution of Non Spherical Particles in Solid-Liquid Fluidization.
Chem. Eng. Sci. 37 (11): 1639–1644. https://doi.org/10.1016/0009-2509(82)80034-1.
Barnea, E. and Mednick, R. L. 1978. A Generalized Approach to the Fluid Dynamics of Particulate Systems Part III:
General Correlation for the Pressure Drop through Fixed Beds of Spherical Particles. Chem. Eng. J. 15 (2): 215–227.
https://doi.org/10.1016/0300-9467(78)85015-1.
Barnea, E. and Mizrahi, J. 1973. A Generalized Approach to the Fluid Dynamics of Particulate Systems. Part 1. General
Correlation for Fluidization and Sedimentation in Solid Multiparticle Systems. Chem. Eng. J. 5 (2): 171–189. https://
doi.org/10.1016/0300-9467(73)80008-5.
36 SPE-199310-MS

Barree, R. D. and Conway, M. W. 1994. Experimental and Numerical Modeling of Convective Proppant Transport. In
SPE Annu. Tech. Conf. Exhib., 47:216–222. Society of Petroleum Engineers.
Batchelor, G. K. 1972. Sedimentation in a Dilute Dispersion of Spheres. J. Fluid Mech. 52 (2): 245–268. https://
doi.org/10.1017/S0022112072001399.
Batchelor, G. K. 1977. The Effect of Brownian Motion on the Bulk Stress in a Suspension of Spherical Particles. J. Fluid
Mech. 83 (1): 97–117. https://doi.org/10.1017/S0022112077001062.
Batchelor, G. K. 2000. An Introduction to Fluid Dynamics. Cambridge, UK: Cambridge University Press.
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. Int. J. Miner. Process. 143: 125–130. https://
doi.org/10.1016/j.minpro.2015.07.005.
Bird, R. B., Stewart, W. E., and Lightfoot, E. N. 2006. Transport Phenomena, 2nd edition. John Wiley & Sons, Inc.
Bowen, W. R. and Sharif, A. O. 1994. Transport through Microfiltration Membranes—Particle Hydrodynamics and Flux
Reduction. J. Colloid Interface Sci. 168 (2): 414–421. https://doi.org/10.1006/jcis.1994.1437.
Brady, J. F. 1993. The Rheological Behavior of Concentrated Colloidal Dispersions. J. Chem. Phys. 99 (1): 567–581.
https://doi.org/10.1063/1.465782.
Brannon, H. D., Wood, W. D., and Wheeler, R. S. 2006. Large-Scale Laboratory Investigation of the Effects of Proppant
and Fracturing-Fluid Properties on Transport. Presented at the SPE International Symposium and Exhibition on
Formation Damage Control, Lafayette, Louisiana, 15-17 February. SPE-98005-MS. https://doi.org/10.2118/98005-
MS.
Brauer, H. and Mewes, D. 1972. Strömungswiderstand Sowie Stationärer Und Instationärer Stoff- Und Wärmeübergang
an Kugeln. Chemie Ing. Tech. 44 (13): 865–868. https://doi.org/10.1002/cite.330441314.
Brinkman, H. 1947. A Calculation of the Viscosity and the Sedimentation Constant for Solutions of Large Chain Molecules
Taking into Account the Hampered Flow of the Solvent through These Molecules. Physica 13 (8): 447–448. https://
doi.org/10.1016/0031-8914(47)90030-X.
Brouwers, H. J. H. 2010. Viscosity of a Concentrated Suspension of Rigid Monosized Particles. Phys. Rev. E 81 (5):
051402. https://doi.org/10.1103/PhysRevE.81.051402.
Brouwers, H. J. H. 2011. Packing Fraction of Geometric Random Packings of Discretely Sized Particles. Phys. Rev. E 84
(4): 042301. https://doi.org/10.1103/PhysRevE.84.042301.
Brown, P. P. and Lawler, D. F. 2003. Sphere Drag and Settling Velocity Revisited. J. Environ. Eng. 129 (3): 222–231.
https://doi.org/10.1061/(ASCE)0733-9372(2003)129:3(222).
Ceylan, K., Altunbaṣ, A., and Kelbaliyev, G. 2001. A New Model for Estimation of Drag Force in the Flow of
Newtonian Fluids around Rigid or Deformable Particles. Powder Technol. 119 (2–3): 250–256. https://doi.org/10.1016/
S0032-5910(01)00261-3.
Cheng, N. S. 2009. Comparison of Formulas for Drag Coefficient and Settling Velocity of Spherical Particles. Powder
Technol. 189 (3): 395–398. https://doi.org/10.1016/j.powtec.2008.07.006.
Chester, W., Breach, D. R., and Proudman, I. 1969. On the Flow Past a Sphere at Low Reynolds Number. J. Fluid Mech.
37 (4): 751–760. https://doi.org/10.1017/S0022112069000851.
Chhabra, R. P. 1995. Wall Effects on Free-Settling Velocity of Non-Spherical Particles in Viscous Media in Cylindrical
Tubes. Powder Technol. 85 (1): 83–90. https://doi.org/10.1016/0032-5910(95)03012-X.
Chhabra, R. P. 1996. Wall Effects on Terminal Velocity of Non-Spherical Particles in Non-Newtonian Polymer Solutions.
Powder Technol. 88 (1): 39–44. https://doi.org/10.1016/0032-5910(96)03100-2.
Chhabra, R. P. 2006. Bubbles, Drops, and Particles in Non-Newtonian Fluids, 2nd edition. CRC press.
Chhabra, R. P., Agarwal, S., and Chaudhary, K. 2003. A Note on Wall Effect on the Terminal Falling Velocity of a Sphere
in Quiescent Newtonian Media in Cylindrical Tubes. Powder Technol. 129 (1–3): 53–58. https://doi.org/10.1016/
S0032-5910(02)00164-X.
Chhabra, R. P. and Peri, S. S. 1991. Simple Method for the Estimation of Free-Fall Velocity of Spherical Particles in Power
Law Liquids. Powder Technol. 67 (3): 287–290. https://doi.org/10.1016/0032-5910(91)80110-5.
Chhabra, R. P., Uhlherr, P. H. T., and Boger, D. V. 1980. The Influence of Fluid Elasticity on the Drag
Coefficient for Creeping Flow around a Sphere. J. Nonnewton. Fluid Mech. 6 (3–4): 187–199. https://
doi.org/10.1016/0377-0257(80)80002-4.
Chong, J. S., Christiansen, E. B., and Baer, A. D. 1971. Rheology of Concentrated Suspensions. J. Appl. Polym. Sci. 15
(8): 2007–2021. https://doi.org/10.1002/app.1971.070150818.
Clark, P. E., Harkin, M. W., Wahl, H. A. et al 1977. Design Of A Large Vertical Prop Transport Model. Presented at
the SPE Annual Fall Technical Conference and Exhibition, Denver, Colorado, 9-12 October. SPE-6814-MS. https://
doi.org/10.2118/6814-MS.
Clark, P. E. and Zhu, Q. 1996. Convective Transport of Propping Agents During Hydraulic Fracturing. Presented at the
SPE Eastern Regional Meeting, Columbus, Ohio, 23-25 October. SPE-37358-MS. https://doi.org/10.2118/37358-MS.
SPE-199310-MS 37

Cleary, M. P. and Fonseca, A. 1992. Proppant Convection and Encapsulation in Hydraulic Fracturing: Practical
Implications of Computer and Laboratory Simulations. Presented at the SPE Annual Technical Conference and
Exhibition, Washington, D.C., 4-7 October. SPE-24825-MS. https://doi.org/10.2118/24825-MS.
Clift, R., Grace, J.R., and Weber, M. E. 1978. Bubbles, Drops, and Particles. New York, NY: Academic Press.
Concha, F. and Almendra, E. R. 1979. Settling Velocities of Particulate Systems, 1. Settling Velocities of Individual
Spherical Particles. Int. J. Miner. Process. 5 (4): 349–367. https://doi.org/10.1016/0301-7516(79)90044-9.
Concha, F. and Barrientos, A. 1982. Settling Velocities of Particulate Systems, 3. Power Series Expansion for the
Drag Coefficient of a Sphere and Prediction of the Settling Velocity. Int. J. Miner. Process. 9 (2): 167–172. https://
doi.org/10.1016/0301-7516(82)90025-4.
Dabak, T. and Yucel, O. 1986. Shear Viscosity Behavior of Highly Concentrated Suspensions at Low and High Shear-
Rates. Rheol. Acta 25 (5): 527–533. https://doi.org/10.1007/BF01774404.
Dallavalle, J. M. 1948. Micromeritics: The Technology of Fine Particles. Londan, UK: Pitman.
Dazhi, G. and Tanner, R. I. 1985. The Drag on a Sphere in a Power-Law Fluid. J. Nonnewton. Fluid Mech. 17 (1): 1–12.
https://doi.org/10.1016/0377-0257(85)80001-X.
Di Felice, R. 1995. Hydrodynamics of Liquid Fluidisation. Chem. Eng. Sci. 50 (8): 1213–1245. https://
doi.org/10.1016/0009-2509(95)98838-6.
Di Felice, R. 1996. A Relationship for the Wall Effect on the Settling Velocity of a Sphere at Any Flow Regime. Int. J.
Multiph. Flow 22 (3): 527–533. https://doi.org/10.1016/0301-9322(96)00004-3.
Dou, G. R. 1981. Turbulence Mechanics. Beijing: People's Education Press.
Duduković, A. P. and Končar- Djurdjević, S. K. 1981. The Effect of Tube Walls on Drag Coefficients of Coaxially Placed
Objects. AIChE J. 27 (5): 837–840. https://doi.org/10.1002/aic.690270519.
Dullien, F. A. L. 1975. Single Phase Flow Through Porous Media and Pore Structure. Chem. Eng. J. Biochem. Eng. J.
10 (1): 1–34.
Dyke, M. V. 1970. Extension of Goldstein's Series for the Oseen Drag of a Sphere. J. Fluid Mech. 44 (02): 365. https://
doi.org/10.1017/S0022112070001878.
Eilers, H. 1941. The Viscosity of the Emulsion of Highly Viscous Substances as Function of Concentration. Kolloid-
Zeitschrift 97 (3): 313–321.
Einstein, A. 1906. A New Determination of Molecular Dimensions. Ann. Phys. 19 (2): 289–306.
El-Kaissy, M. M. and Homsy, G. M. 1973. A Theoretical Study of Pressure Drop and Transport in Packed Beds at
Intermediate Reynolds Numbers. Ind. Eng. Chem. Fundam. 12 (1): 82–90. https://doi.org/10.1021/i160045a014.
Fair, G. M., Geyer, J. C., and Morris, J. C. 1954. Water Supply and Waste-Water Disposal. New York, NY:Wiley.
Farr, R. S. and Groot, R. D. 2009. Close Packing Density of Polydisperse Hard Spheres. J. Chem. Phys. 131 (24): 244104.
https://doi.org/10.1063/1.3276799.
Fidleris, V. and Whitmore, R. L. 1961. Experimental Determination of the Wall Effect for Spheres Falling Axially in
Cylindrical Vessels. Br. J. Appl. Phys. 12 (9): 490–494. https://doi.org/10.1088/0508-3443/12/9/311.
Flemmer, R. L. C. and Banks, C. L. 1986. On the Drag Coefficient of a Sphere. Powder Technol. 48 (3): 217–221. https://
doi.org/10.1016/0032-5910(86)80044-4.
Foscolo, P. U., Gibilaro, L. G., and Waldram, S. P. 1983. A Unified Model for Particulate Expansion of Fluidised Beds
and Flow in Fixed Porous Media. Chem. Eng. Sci. 38 (8): 1251–1260. https://doi.org/10.1016/0009-2509(83)80045-1.
Fouda, A. E. and Capes, C. E. 1976. Calculation of Large Numbers of Terminal Velocities or Equivalent Particle
Diameters Using Polynomial Equations Fitted to the Heywood Tables. Powder Technol. 13 (2): 291–293. https://
doi.org/10.1016/0032-5910(76)85016-4.
Francis, A. W. 1933. Wall Effect in Falling Ball Method for Viscosity. J. Appl. Phys. 4 (11): 403–406. https://
doi.org/10.1063/1.1745151.
Frankel, N. A. and Acrivos, A. 1967. On the Viscosity of a Concentrated Suspension of Solid Spheres. Chem. Eng. Sci.
22 (6): 847–853. https://doi.org/10.1016/0009-2509(67)80149-0.
Ganguly, U. P. 1980. Direct Method for the Prediction of Expanded Bed Height in Liquid-solid Fluidization. Can. J. Chem.
Eng. 58 (5): 559–563. https://doi.org/10.1002/cjce.5450580502.
Garslde, J. and Al-Dibouni, M. R. 1977. Velocity-Voidage Relationships for Fluidization and Sedimentation in Solid-
Liquid Systems. Ind. Eng. Chem. Process Des. Dev. 16 (2): 206–214. https://doi.org/10.1021/i260062a008.
Gel'perin, N. I., Ainshtein, V. G., and Toskubaev, I. N. 1972. Heat Exchange between Finned Tubes and a Fluidized Bed
of Granular Material. Chem. Technol. Fuels Oils 8 (9): 697–699. https://doi.org/10.1007/BF00717975.
Goldstein, S. 1929. The Steady Flow of Viscous Fluid Past a Fixed Spherical Obstacle at Small Reynolds Numbers. Proc.
R. Soc. A Math. Phys. Eng. Sci. 123 (791): 225–235. https://doi.org/10.1098/rspa.1929.0067.
Guth, E. and Simha, R. 1936. Investigations of Viscosity of Suspensions and Solutions: Part 3. The Viscosity of Spherical
Suspensions. Kolloid 74: 266–275.
38 SPE-199310-MS

Haberman, W. L. and Sayre, R. M. 1958. David Taylor Model Basin. Report No. 1143, US Navy Department, Washington
D.C.
Haider, A. and Levenspiel, O. 1989. Drag Coefficient and Terminal Velocity of Spherical and Nonspherical Particles.
Powder Technol. 58 (1): 63–70. https://doi.org/10.1016/0032-5910(89)80008-7.
Happel, J. 1957. Viscosity of Suspensions of Uniform Spheres. J. Appl. Phys. 28 (11): 1288–1292. https://
doi.org/10.1063/1.1722635.
Happel, J. 1958. Viscous Flow in Multiparticle Systems: Slow Motion of Fluids Relative to Beds of Spherical Particles.
AIChE J. 4 (2): 197–201. https://doi.org/10.1002/aic.690040214.
Happel, J. and Bart, E. 1974. The Settling of a Sphere along the Axis of a Long Square Duct at Low Reynolds' Number.
Appl. Sci. Res. 29 (1): 241–258. https://doi.org/10.1007/BF00384149.
Happel, J. and Brenner, H. 1973. Low Reynolds Number Hydrodynamics. Leyden, Netherland: Noordhoff Int. Publishing.
Hatschek, E. 1913. The General Theory of Viscosity of Two-Phase Systems. Trans. Faraday Soc. 9: 80–92. https://
doi.org/10.1039/tf9130900080.
Hawksley, P. G. W. 1951. The Effect of Concentration on the Settling of Suspensions and Flow through Porous Media. In
Some Aspects of Fluid Flow (4): 114–135., Edward Arnold Co., London, UK.
Hesketh, R. P., Etchells, A. W., and Russell, T. W. F. 1991. Bubble Breakage in Pipeline Flow. Chem. Eng. Sci. 46 (1):
1–9. https://doi.org/10.1016/0009-2509(91)80110-K.
Higdon, J. J. L. and Muldowney, G. P. 1995. Resistance Functions for Spherical Particles, Droplets and Bubbles in
Cylindrical Tubes. J. Fluid Mech. 298: 193–210. https://doi.org/10.1017/S0022112095003272.
Hirata, A. and Bulos, F. B. 1990. Predicting Bed Voidage in Solid-Liquid Fluidization. J. Chem. Eng. Japan 23 (5):
599–604. https://doi.org/10.1252/jcej.23.599.
Hirose, T. and Moo-Young, M. 1969. Bubble Drag and Mass Transfer in Non-Newtonian Fluids: Creeping Flow with
Power-Law Fluids. Can. J. Chem. Eng. 47 (3): 265–267. https://doi.org/10.1002/cjce.5450470318.
Horri, B. A., Ranganathan, P., Selomulya, C. et al 2011. A New Empirical Viscosity Model for Ceramic Suspensions.
Chem. Eng. Sci. 66 (12): 2798–2806. https://doi.org/10.1016/j.ces.2011.03.040.
Ihme, F., Schmidtt, H., and Brauer, H. 1972. Theoretical Studies on Mass Transfer at and Flow Past Spheres. Chemie
Ing. Tech. 44 (5): 306–313.
Ishii, M. and Zuber, N. 1979. Drag Coefficient and Relative Velocity in Bubbly, Droplet or Particulate Flows. AIChE J.
25 (5): 843–855. https://doi.org/10.1002/aic.690250513.
Iwaoka, M. and Ishii, T. 1979. Experimental Wall Correction Factors of Single Solid Spheres In Circular Cylinders. J.
Chem. Eng. Japan 12 (3): 239–242. https://doi.org/10.1252/jcej.12.239.
Jean, R.-H., Tang, W.-T., and Fan, L.-S. 1989. The Sedimentation-Dispersion Model for Slurry Bubble Columns. AIChE
J. 35 (4): 662–665. https://doi.org/10.1002/aic.690350418.
Jottrand, R. 1952. An Experimental Study of the Mechanism of Fluidization. J. Appl. Chem. 2: S17–S26.
Kansal, A. R., Torquato, S., and Stillinger, F. H. 2002. Computer Generation of Dense Polydisperse Sphere Packings. J.
Chem. Phys. 117 (18): 8212–8218. https://doi.org/10.1063/1.1511510.
Kawase, Y. and Moo-Young, M. 1986. Approximate Solutions for Power-Law Fluid Flow Past a Particle at Low Reynolds
Numbers. J. Nonnewton. Fluid Mech. 21 (2): 167–177. https://doi.org/10.1016/0377-0257(86)80034-9.
Kehlenbeck, R. and Felice, R. D. 1999. Empirical Relationships for the Terminal Settling Velocity of Spheres in Cylindrical
Columns. Chem. Eng. Technol. 22 (4): 303–308. https://doi.org/10.1002/(SICI)1521-4125(199904)22:4303::AID-
CEAT303>3.0.CO;2-8.
Kelessidis, V. C. 2004. An Explicit Equation for the Terminal Velocity of Solid Spheres Falling in Pseudoplastic Liquids.
Chem. Eng. Sci. 59 (21): 4437–4447. https://doi.org/10.1016/j.ces.2004.07.008.
Kern, L. R., Perkins, T. K., and Wyant, R. E. 1959. The Mechanics of Sand Movement in Fracturing. J. Pet. Technol. 11
(07): 55–57. https://doi.org/10.2118/1108-G.
Khan, A. R. and Richardson, J. F. 1987. The Resistance to Motion of a Solid Sphere in a Fluid. Chem. Eng. Commun. 62
(1–6): 135–150. https://doi.org/10.1080/00986448708912056.
Khan, A. R. and Richardson, J. F. 1990. Pressure Gradient and Friction Factor for Sedimentation and Fluidisation of
Uniform Spheres in Liquids. Chem. Eng. Sci. 45 (1): 255–265. https://doi.org/10.1016/0009-2509(90)87097-C.
Kmieć, A. 1976. Some Remarks on the Richardson-Zaki Equation. Chem. Eng. J. 11 (3): 237–238. https://
doi.org/10.1016/0300-9467(76)80046-9.
Koos, E. 2009. Rheological Measurements in Liquid-Solid Flows. PhD Dissertation, California Institute of Technology,
Pasadena, California (March 2009).
Krieger, I. M. and Dougherty, T. J. 1959. A Mechanism for Non-Newtonian Flow in Suspensions of Rigid Spheres. Trans.
Soc. Rheol. 3 (1): 137–152. https://doi.org/10.1122/1.548848.
Krigbaum, W. R. and Flory, P. J. 1953. Molecular Weight Dependence of the Intrinsic Viscosity of Polymer Solutions. II.
J. Polym. Sci. 11 (1): 37–51. https://doi.org/10.1002/pol.1953.120110103.
SPE-199310-MS 39

Kunitz, M. 1926. An Empirical Formula For The Relation Between Viscosity Of Solution And Volume Of Solute. J. Gen.
Physiol. 9 (6): 715–725. https://doi.org/10.1085/jgp.9.6.715.
Lali, A. M., Khare, A. S., Joshi, J. B. et al 1989. Behaviour of Solid Particles in Viscous Non-Newtonian Solutions:
Settling Velocity, Wall Effects and Bed Expansion in Solid-Liquid Fluidized Beds. Powder Technol. 57 (1): 39–50.
https://doi.org/10.1016/0032-5910(89)80102-0.
Langmuir, I. and Blodgett, K. B. 1946. A Mathematical Investigation of Water Droplet Trajectories. US Army Air Force
Technical Report No. 5418.
Lapple, C. E. 1951. Fluid and Particle Mechanics. Newark, DE: University of Delaware Publ.
Lapple, C. E. and Shepherd, C. B. 1940. Calculation of Particle Trajectories. Ind. Eng. Chem. 32 (5): 605–617. https://
doi.org/10.1021/ie50365a007.
Latto, B., Round, G. F., and Anzenavs, R. 1973. Drag Coefficients and Pressure Drops for Hydrodynamically Suspended
Spheres in a Vertical Tube with and without Polymer Addition. Can. J. Chem. Eng. 51 (5): 536–541. https://
doi.org/10.1002/cjce.5450510502.
LeClair, B. P. and Hamielec, A. E. 1968. Viscous Flow through Particle Assemblages at Intermediate Reynolds Numbers.
Steady-State Solutions for Flow through Assemblages of Spheres. Ind. Eng. Chem. Fundam. 7 (4): 542–549. https://
doi.org/10.1021/i160028a003.
Letan, R. 1974. On Vertical Dispersed Two-Phase Flow. Chem. Eng. Sci. 29 (2): 621–624. https://
doi.org/10.1016/0009-2509(74)80073-4.
Lewis, E. W. and Bowerman, E. W. 1952. Fluidization of Solid Particles in Liquids. Chem. Eng. Sci. 48 (12): 603–610.
Lewis, J. A. and Carrier, G. F. 1949. Some Remarks on the Flat Plate Boundary Layer. Q. Appl. Math. 7 (2): 228–234.
https://doi.org/10.1090/qam/34160.
Liao, S. J. 2002. An Analytic Approximation of the Drag Coefficient for the Viscous Flow Past a Sphere. Int. J. Non.
Linear. Mech. 37 (1): 1–18. https://doi.org/10.1016/S0020-7462(00)00092-5.
Loeffler, A. L. and Ruth, B. F. 1959. Particulate Fluidization and Sedimentation of Spheres. AIChE J. 5 (3): 310–314.
https://doi.org/10.1002/aic.690050312.
Lunnon, R. G. 1928. Fluid Resistance to Moving Spheres. Proc. R. Soc. A Math. Phys. Eng. Sci. 118 (780): 680–694.
https://doi.org/10.1098/rspa.1928.0077.
Machac, I. and Lecjaks, Z. 1995. Wall Effect for a Sphere Falling through a Non-Newtonian Fluid in a Rectangular Duct.
Chem. Eng. Sci. 50 (1): 143–148. https://doi.org/10.1016/0009-2509(94)00211-9.
Madhav, G. V. and Chhabra, R. P. 1994. Settling Velocities of Non-Spherical Particles in Non-Newtonian Polymer
Solutions. Powder Technol. 78 (1): 77–83. https://doi.org/10.1016/0032-5910(93)02761-X.
Maron, S. H. and Pierce, P. E. 1956. Application of Ree-Eyring Generalized Flow Theory to Suspensions of Spherical
Particles. J. Colloid Sci. 11 (1): 80–95. https://doi.org/10.1016/0095-8522(56)90023-X.
McGauhey, P. H. 1956. Theory of Sedimentation. J. Am. Water Works Assoc. 48 (4): 437–448.
McNown, J. S., Lee, H. M., McPherson, M. et al 1948. Influence of Boundary Proximity on the Drag of Spheres. Proc.,
7th nternational Congress for Applied Mechanics.
Mendoza, C. I. and Santamaría-Holek, I. 2009. The Rheology of Hard Sphere Suspensions at Arbitrary Volume Fractions:
An Improved Differential Viscosity Model. J. Chem. Phys. 130 (4): 044904. https://doi.org/10.1063/1.3063120.
Metzner, A. B. and Reed, J. C. 1955. Flow of Non-Newtonian Fluids—Correlation of the Laminar, Transition, and
Turbulent-Flow Regions. AIChE J. 1 (4): 434–440. https://doi.org/10.1002/aic.690010409.
Mewis, J. and Wagner, N. J. 2012. Colloidal Suspension Rheology. Cambridge, UK: Cambridge University Press.
Mikhailov, M. D. and Freire, A. P. S. 2013. The Drag Coefficient of a Sphere: An Approximation Using Shanks Transform.
Powder Technol. 237 (March): 432–435. https://doi.org/10.1016/j.powtec.2012.12.033.
Miyamura, A., Iwasaki, S., and Ishii, T. 1981. Experimental Wall Correction Factors of Single Solid Spheres
in Triangular and Square Cylinders, and Parallel Plates. Int. J. Multiph. Flow 7 (1): 41–46. https://
doi.org/10.1016/0301-9322(81)90013-6.
Mooney, M. 1951. The Viscosity of a Concentrated Suspension of Spherical Particles. J. Colloid Sci. 6 (2): 162–170.
https://doi.org/10.1016/0095-8522(51)90036-0.
Morrison, F. A. 2013. An Introduction to Fluid Mechanics. Cambridge, UK: Cambridge University Press.
Morsi, S. A. and Alexander, A. J. 1972. An Investigation of Particle Trajectories in Two-Phase Flow Systems. J. Fluid
Mech. 55 (2): 193–208. https://doi.org/10.1017/S0022112072001806.
Munroe, H. S. 1889. The English Versus the Continental System of Jigging: Is Close Sizing Advantageous? Trans. Am.
Inst. Min. Engrs. New York, NY.
Newton, I. 1687. Principia. Lib. II, Prop. XXXIX, Theorem XXXI. Cambridge, UK:Cambridge University Press.
Nicodemo, L. and Nicolais, L. 1974. Viscosity of Concentrated Fiber Suspensions. Chem. Eng. J. 8 (2): 155–156. https://
doi.org/10.1016/0300-9467(74)85018-5.
40 SPE-199310-MS

Novotny, E. J. 1977. Proppant Transport. Presented at the SPE Annual Fall Technical Conference and Exhibition, Denver,
Colorado, 9-12 October. SPE-6813-MS. https://doi.org/10.2118/6813-MS.
Oliver, D. R. 1961. The Sedimentation of Suspensions of Closely-Sized Spherical Particles. Chem. Eng. Sci. 15 (3–4):
230–242. https://doi.org/10.1016/0009-2509(61)85026-4.
Oliver, D. R. and Ward, S. G. 1953. Relationship between Relative Viscosity and Volume Concentration of Stable
Suspensions of Spherical Particles. Nature 171 (4348): 396–397. https://doi.org/10.1038/171396b0.
Oseen, C. W. 1910. Stokes' Formula and a Related Theorem in Hydrodynamics. Ark. Mat. Astron. Fysik. 6: 20.
Paine, P. L. and Scherr, P. 1975. Drag Coefficients for the Movement of Rigid Spheres through Liquid-Filled Cylindrical
Pores. Biophys. J. 15 (10): 1087–1091. https://doi.org/10.1016/S0006-3495(75)85884-X.
Patwardhan, V. S. and Tien, C. 1985. Sedimentation and Fluidization in Solid-Liquid Systems: A Simple Approach. AIChE
J. 31 (1): 146–149. https://doi.org/10.1002/aic.690310117.
Phan, S.-E., Russel, W. B., Zhu, J. et al 1998. Effects of Polydispersity on Hard Sphere Crystals. J. Chem. Phys. 108 (23):
9789–9795. https://doi.org/10.1063/1.476453.
Prandtl, L. and Tietjens, O. 1931. Hydro- Und Aeromechanik Nach Vorlesungen von L. Prandtl. J. Springer. 2.
Proudman, I. and Pearson, J. R. A. 1957. Expansions at Small Reynolds Numbers for the Flow Past a Sphere and a Circular
Cylinder. J. Fluid Mech. 2 (3): 237–262. https://doi.org/10.1017/S0022112057000105.
Quemada, D. and Berli, C. 2002. Energy of Interaction in Colloids and Its Implications in Rheological Modeling. Adv.
Colloid Interface Sci. 98 (1): 51–85. https://doi.org/10.1016/S0001-8686(01)00093-8.
Ramamurthy, K. and Subbaraju, K. 1973. Bed Expansion Characteristics of Annular Liquid-Fluidized Beds. Ind. Eng.
Chem. Process Des. Dev. 12 (2): 184–189. https://doi.org/10.1021/i260046a010.
Riba, J. P. and Couderc, J. P. 1977. Expansion de Couches Fluidisées Par Des Liquides. Can. J. Chem. Eng. 55 (2):
118–121. https://doi.org/10.1002/cjce.5450550202.
Richardson, J. F. and Jerónimo, M. A. S. 1979. Velocity-Voidage Relations for Sedimentation and Fluidisation. Chem.
Eng. Sci. 34 (12): 1419–1422. https://doi.org/10.1016/0009-2509(79)85167-2.
Richardson, J. F. and Zaki, W. N. 1954. The Sedimentation of a Suspension of Uniform Spheres under Conditions of
Viscous Flow. Chem. Eng. Sci. 3 (2): 65–73. https://doi.org/10.1016/0009-2509(54)85015-9.
Robinson, C. D. 1926. Some Factors Influencing Sedimentation. Ind. Eng. Chem. 18 (8): 869–871. https://doi.org/10.1021/
ie50200a036.
Roscoe, R. 1952. The Viscosity of Suspensions of Rigid Spheres. Br. J. Appl. Phys. 3 (8): 267–269. https://
doi.org/10.1088/0508-3443/3/8/306.
Rouse, H. 1946. Elementary Mechanics of Fluids. New York, NY: John Wiley & Sons Inc.
Rowe, P. N. 1987. A Convenient Empirical Equation for Estimation of the Richardson-Zaki Exponent. Chem. Eng. Sci.
42 (11): 2795–2796. https://doi.org/10.1016/0009-2509(87)87035-5.
Rutgers, I. R. 1962. Relative Viscosity and Concentration. Rheol. Acta 2 (4): 305–348. https://doi.org/10.1007/
BF01976051.
Schiller, L. 1932. Fallversuche Mit Kugeln Und Scheiben. Handb. Der Exp. 4 (339–398).
Schiller, L. and Naumann, A. 1933. Fundamental Calculations in Gravitational Processing. Zeitschrift Des Vereines Dtsch.
Ingenieure 77 (318–220).
Schlichting, H. and Kestin, J. 1979. Boundary- Layer Theory. New York, NY: McGraw-Hill.
Schols, R. S. and Visser, W. 1974. Proppant Bank Buildup in a Vertical Fracture Without Fluid Loss. Presented at the SPE
European Spring Meeting, Amsterdam, Netherlands, 29-30 May. SPE-4834-MS. https://doi.org/10.2118/4834-MS
Shah, S. N., Fadili, Y. E., and Chhabra, R. P. 2007. New Model for Single Spherical Particle Settling Velocity in Power Law
(Visco-Inelastic) Fluids. Int. J. Multiph. Flow 33 (1): 51–66. https://doi.org/10.1016/j.ijmultiphaseflow.2006.06.006.
Shapiro, A. P. and Probstein, R. F. 1992. Random Packings of Spheres and Fluidity Limits of Monodisperse and Bidisperse
Suspensions. Phys. Rev. Lett. 68 (9): 1422–1425. https://doi.org/10.1103/PhysRevLett.68.1422.
Shewan, H. M. 2015. Rheology of Soft Particle Suspensions. B. Tech. Thesis, The University of Queensland, Brisbane,
Australia (April 2015).
Simha, R. 1952. A Treatment of the Viscosity of Concentrated Suspensions. J. Appl. Phys. 23 (9): 1020–1024. https://
doi.org/10.1063/1.1702338.
Slot, R. E. 1984. TERMINAL VELOCITY FORMULA FOR OBJECTS IN A VISCOUS FLUID. J. Hydraul. Res. 22
(4): 235–243. https://doi.org/10.1080/00221688409499381.
Song, D., Gupta, R. K., and Chhabra, R. P. 2009. Wall Effects on a Sphere Falling in Quiescent Power Law Fluids in
Cylindrical Tubes. Ind. Eng. Chem. Res. 48 (12): 5845–5856. https://doi.org/10.1021/ie900176y.
Staben, M. E., Zinchenko, A. Z., and Davis, R. H. 2003. Motion of a Particle between Two Parallel Plane Walls in Low-
Reynolds-Number Poiseuille Flow. Phys. Fluids 15 (6): 1711–1733. https://doi.org/10.1063/1.1568341.
Steinour, H. H. 1944. Rate of Sedimentation. Nonflocculated Suspensions of Uniform Spheres. Ind. Eng. Chem. 36 (7):
618–624. https://doi.org/10.1021/ie50415a005.
SPE-199310-MS 41

Stokes, G. G. 1851. On the Effect of the Internal Friction of Fluids on the Motion of Pendulums - Section III. Trans.
Cambridge Philos. Soc. 9 (2): 8–106.
Sutterby, J. L. 1973. Falling Sphere Viscometry. I. Wall and Inertial Corrections to Stokes' Law in Long Tubes. Trans.
Soc. Rheol. 17 (4): 559–573. https://doi.org/10.1122/1.549308.
Swamee, P. K. and Ojha, C. S. P. 1991. Drag Coefficient and Fall Velocity of Nonspherical Particles. J. Hydraul. Eng.
117 (5): 660–667. https://doi.org/10.1061/(ASCE)0733-9429(1991)117:5(660).
Tanaka, N., Makino, K., and Iinoya, K. 1973. Theoretical Analysis of Dust Cleaning Operarion in Multi-Compartment
Bag Filter. Chem. Eng. 37 (7): 718–724,a1. https://doi.org/10.1252/kakoronbunshu1953.37.718.
Terfous, A., Hazzab, A., and Ghenaim, A. 2013. Predicting the Drag Coefficient and Settling Velocity of Spherical
Particles. Powder Technol. 239: 12–20. https://doi.org/10.1016/j.powtec.2013.01.052.
Thomas, D. G. 1965. Transport Characteristics of Suspension: VIII. A Note on the Viscosity of Newtonian Suspensions
of Uniform Spherical Particles. J. Colloid Sci. 20 (3): 267–277. https://doi.org/10.1016/0095-8522(65)90016-4.
Torobin, L. B. and Gauvin, W. H. 1959. Fundamental Aspects of Solids-Gas Flow: Part I: Introductory Concepts
and Idealised Sphere Motion in Viscous Regime. Can. J. Chem. Eng. 37 (4): 129–141. https://doi.org/10.1002/
cjce.5450370401.
Torquato, S., Truskett, T. M., and Debenedetti, P. G. 2000. Is Random Close Packing of Spheres Well Defined? Phys. Rev.
Lett. 84 (10): 2064–2067. https://doi.org/10.1103/PhysRevLett.84.2064.
Tripathi, A., Chhabra, R. P., and Sundararajan, T. 1994. Power Law Fluid Flow over Spheroidal Particles. Ind. Eng. Chem.
Res. 33 (2): 403–410. https://doi.org/10.1021/ie00026a035.
Turian, R. M., Yuan, T.-F., and Mauri, G. 1971. Pressure Drop Correlation for Pipeline Flow of Solid-Liquid Suspensions.
AIChE J. 17 (4): 809–817. https://doi.org/10.1002/aic.690170409.
Turton, R. and Clark, N. N. 1987. An Explicit Relationship to Predict Spherical Particle Terminal Velocity. Powder
Technol. 53 (2): 127–129. https://doi.org/10.1016/0032-5910(87)85007-6.
Turton, R. and Levenspiel, O. 1986. A Short Note on the Drag Correlation for Spheres. Powder Technol. 47 (1): 83–86.
https://doi.org/10.1016/0032-5910(86)80012-2.
Vand, V. 1948. Viscosity of Solutions and Suspensions. II. Experimental Determination of the Viscosity–Concentration
Function of Spherical Suspensions. J. Phys. Colloid Chem. 52 (2): 300–314. https://doi.org/10.1021/j150458a002.
Vlajinac, M. and Covert, E. E. 1972. Sting-Free Measurements of Sphere Drag in Laminar Flow. J. Fluid Mech. 54 (3):
385–392. https://doi.org/10.1017/S0022112072000746.
Wadell, H. 1934. The Coefficient of Resistance as a Function of Reynolds Number for Solids of Various Shapes. J.
Franklin Inst. 217 (4): 459–490. https://doi.org/10.1016/S0016-0032(34)90508-1.
Wallis, G. B. 1969. One-Dimensional Two-Phase Flow. New York, NY: McGraw-Hill.
Weissberg, S. G., Simha, R., and Rothman, S. 1951. Viscosity of Dilute and Moderately Concentrated Polymer Solutions.
J. Res. Natl. Bur. Stand. (1934). 47 (4): 298. https://doi.org/10.6028/jres.047.038.
Wen, C. Y. and Fan, L. S. 1974. A Comparison of Recent Axial Dispersion Correlations in Liquid-Solid Fluidized Beds.
Can. J. Chem. Eng. 52 (5): 673–675. https://doi.org/10.1002/cjce.5450520523.
Wen, C. Y. and Yu, Y. H. 1966. A Generalized Method for Predicting the Minimum Fluidization Velocity. AIChE J. 12
(3): 610–612. https://doi.org/10.1002/aic.690120343.
Wham, R. M., Basaran, O. A., and Byers, C. H. 1996. Wall Effects on Flow Past Solid Spheres at Finite Reynolds Number
†. Ind. Eng. Chem. Res. 35 (3): 864–874. https://doi.org/10.1021/ie950354c.
Wieselsberger, C. 1922. Further Information on the Laws of Fluid Resistance. National Advisory Committee for
Aeronautics. Washington D.C.
Yang, H., Fan, M., Liu, A. et al 2015. General Formulas for Drag Coefficient and Settling Velocity of Sphere Based on
Theoretical Law. Int. J. Min. Sci. Technol. 25 (2): 219–223. https://doi.org/10.1016/j.ijmst.2015.02.009.
Zenit, R. and Feng, J. J. 2018. Hydrodynamic Interactions Among Bubbles, Drops, and Particles in Non-Newtonian
Liquids. Annu. Rev. Fluid Mech. 50 (1): 505–534. https://doi.org/10.1146/annurev-fluid-122316-045114.
Zigrang, D. J. and Sylvester, N. D. 1981. An Explicit Equation for Particle Settling Velocities in Solid-liquid Systems.
AIChE J. 27 (6): 1043–1044. https://doi.org/10.1002/aic.690270629.

You might also like