You are on page 1of 17

J187255 DOI: 10.

2118/187255-PA Date: 6-October-18 Stage: Page: 1689 Total Pages: 17

Settling Velocity of Particles in Viscoelastic


Fluids: A Comparison of the Shear-Viscosity
and Elasticity Effects
Sumanth Kumar Arnipally and Ergun Kuru, University of Alberta

Summary
The objective of this paper is to determine how fluid shear viscosity and elasticity might influence the particle-settling velocity,
and even more so to answer the question of which one of these two rheological properties is more dominant in controlling the particle-
settling velocity when viscoelastic drilling fluids are used.
The settling velocities of spherical particles (diameters: 1.18, 1.5, 2, and 3 mm) in partially hydrolyzed polyacrylamide (HPAM)
polymer fluids were measured using the particle-image-shadow graph (PIS) technique. Two sets of test fluids were formulated by mixing
three different grades of HPAM (molecular weights of 500,000, 8 million, and 20 million g/g mol) at polymer concentrations of 0.09,
0.05, and 0.03 wt%. The shear-viscosity and elasticity characteristics of test fluids were determined by performing shear-viscosity and
frequency-sweep oscillatory measurements, respectively. The first set of fluids had almost identical shear-viscosity characteristics while
showing significantly different elastic properties (quantified in terms of relaxation time). The second set of fluids had similar elastic
properties but different shear-viscosity characteristics. In addition, the effect of the particle size on the settling velocities in these test
fluids was also investigated.
The experimentally measured settling velocities were compared with the values calculated from the Shah et al. (2007) model devel-
oped for predicting the settling velocity of spherical particles in power-law (viscoinelastic) fluids as well as the values calculated from
the Malhotra and Sharma (2012) correlation developed for settling velocity in shear-thinning viscoelastic fluids in unconfined media.
Experimental results showed the following:
1. When the fluids with similar shear-viscosity profiles were used, the settling velocity of spherical particles decreased significantly
with the increasing fluid elasticity.
2. The settling-velocity values can be 14 to 50 times overestimated if the effect of the elasticity is not considered.
3. At constant elasticity, the settling velocity of spherical particles also decreased significantly when the fluid shear viscosity
was increased.
4. The spherical particle-settling velocity increased pronouncedly as particle diameter increased from 1.18 to 3 mm. However, the
magnitude of the increase in settling velocity with the increasing particle diameter is less for the samples with higher elasticity
and similar shear-viscosity characteristics.
The fluid shear viscosity and the elasticity both seem to have significant effect on the particle-settling velocity. However, from the
field operational point of view, fluids with high shear-viscosity values are not always practical to use because the high shear viscosity
increases parasitic pressure losses and potentially has a negative effect on the drilling rate. Hence, in such cases increasing the fluid
elasticity can help to reduce the particle-settling velocity even at lower shear-viscosity values.
By conducting experiments under controlled conditions, we were able to quantify the individual effects of fluid shear viscosity and
elasticity on the particle-settling velocity for the first time in drilling literature.

Introduction
Knowledge of settling behavior of particles in various types of fluids is indispensable to design and optimize numerous industrial opera-
tions, such as cuttings transport in oil- and gas-well drilling, proppant transport in hydraulic-fracturing operations, mineral transporta-
tion in mining engineering, solid separations, and solid/liquid mixing in the chemical industry.
When a particle is falling freely in a fluid because of gravity, it achieves a constant velocity once equilibrium is attained. This veloc-
ity is termed as terminal settling velocity or simply settling velocity (McCabe et al. 2004). If the settling of a particle is not affected by
the presence of other particles or by the container walls, then such settling is called free-particle settling. If the particle movement is
retarded by other particles or nearby walls, then it is classified as hindered-particle settling (McCabe et al. 2004). Hindered-particle
settling is commonly observed in most industrial processes, but particle velocity in hindered settling is reportedly related to free-
particle-settling velocity (Kelessidis and Mpandelis 2004). As a result, emphasis has historically been given to the estimation of free-
particle-settling velocity in various fluids.
The particle-settling velocity is known to be affected by fluid properties such as viscosity and density and by physical particle prop-
erties such as size, shape, and density (McCabe et al. 2004). From the practical viewpoint of field applications and design of industrial
processes, however, settling velocities of the solid particles are controlled by changing the fluid properties. Because analytical solutions
of the problem are difficult to obtain, empirical correlations are commonly used to predict the particle-settling velocity in various indus-
trial processes involving complex fluids. The solution generally requires the development of a correlation of drag coefficient (CD) and
particle Reynolds number (Rep). Terminal velocity of the particles can then be determined once the drag coefficient is known for the
respective particle Reynolds number (McCabe et al. 2004).
The classical work of Stokes (1851) provided one of the first treatises of the problem, resulting in a correlation that can be used for deter-
mining particle-settling velocity in a creeping flow of a Newtonian fluid (Rep<0.1). According to the Stokes model, the particle-settling
velocity is inversely proportional to the fluid viscosity. Several correlations were developed for Newtonian fluids to determine the drag
coefficient of spherical particles (Flemmer and Banks 1986; Turton and Levenspiel 1986; Khan and Richardson 1987) and of nonspherical

Copyright V
C 2018 Society of Petroleum Engineers

This paper (SPE 187255) was accepted for presentation at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, USA, 9–11 October 2017, and revised for publication.
Original manuscript received for review 16 June 2017. Revised manuscript received for review 7 March 2018. Paper peer approved 13 March 2018.

October 2018 SPE Journal 1689

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1690 Total Pages: 17

particles (Turton and Levenspiel 1986; Haider and Levenspiel 1989). Most industrial applications, however, require the use of non-
Newtonian fluids. Models were also developed to describe the particle-settling velocity in non-Newtonian shear-thinning fluids by simply
modifying Newtonian correlations with the introduction of apparent shear viscosity at various flow regimes (Chien 1972; Moore 1974; Lali
et al. 1989; Miura et al. 2001). Apparent viscosity is the viscosity of the fluid at the respective shear rates induced by the settling particles.
The maximum value of shear rate induced by the settling particle that is relevant to polymer fluids is given by Eq. 1 (Chhabra 2007):

2Vs
c¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ
Ds

Most of these models were developed for purely viscous non-Newtonian shear-thinning fluids where there were no considerations of
the fluid-elasticity effect. Therefore, these models are not suitable for settling-velocity prediction when the process involves fluids of
viscoelastic nature such as drilling fluids and fracturing fluids. Viscoelastic properties of the drilling fluids strongly influence gel-
strength, barite-sag, cuttings-transport, and filtration-loss characteristics of the drilling fluids (Bui et al. 2012). In hydraulic fracturing, it
was observed that not only the shear viscosity but also the elasticity of the fracturing fluid needs to be considered to optimize the
hydraulic-fluid properties for effective transport and to avoid premature settling of proppants (Gomaa et al. 2015; Hu et al. 2015). How-
ever, the fluid-elasticity properties are still not being properly considered when designing drilling and fracturing-fluid rheological prop-
erties (Bui et al. 2012). Further study is therefore needed to understand the effect of elasticity and relative importance of the shear
viscosity vs. the elasticity for the settling velocity of particles in viscoelastic fluids.
Several experimental studies were conducted previously to understand the settling behavior of the particles in viscoelastic liquids
and Boger fluids (i.e., elastic fluids of constant viscosity). Acharya et al. (1976a, b) studied the settling behavior of a sphere in shear-
thinning viscoelastic polymer solutions and reported that the fluid drag was reduced at higher Reynolds numbers (Acharya et al.
1976b). Soon after, Chhabra et al. (1980) conducted particle-settling experiments with Boger fluids. Using the Weissenberg number as
an extent of the elasticity, Chhabra et al. (1980) reported similar observations of decrease in the drag coefficient at higher values of the
Weissenberg number. Acharya (1986, 1988) dealt with viscoelastic fracturing fluids and stated that in the creeping-flow regime, the set-
tling velocities of the proppant particles were governed by the fluid-viscosity parameters but not by fluid elasticity, whereas in the inter-
mediate-Reynolds-number flow regime, proppant-settling velocities were increased because of fluid elasticity.
In contrast, van den Brule and Gheissary (1993) observed a noticeable decrease in particle-settling velocity caused by the fluid elas-
ticity in highly elastic shear-thinning fluids. Walters and Tanner (1992) described the dependence of drag on elasticity for Boger fluids.
They observed that in Boger fluids, at low Weissenberg numbers, the elasticity caused drag decline, and at higher Weissenberg num-
bers, the elasticity caused drag improvement (Walters and Tanner 1992). Malhotra and Sharma (2012) reported similar observations for
particle settling in surfactant-based shear-thinning viscoelastic fluids. These studies on Boger fluids and others that are cited here show
that there is already evidence in the literature that fluid elasticity has an appreciable effect on the settling behavior of the particles. How-
ever, as was noted by Chhabra (2007) regarding theoretical advancements of shear-thinning viscoelastic systems, it is an onerous prob-
lem to integrate individual effects of both the shear viscosity and the elasticity of the fluid on the settling velocity of the particle. The
reason for this complexity could be that the individual influences of these factors are not known conclusively because no study in the lit-
erature disengaged the effects. To do this, individual outcomes of the fluid shear viscosity and the fluid elasticity on the settling velocity
have to be investigated and their functional relationship with the particle-settling velocity should be understood. In addition, the domi-
nant factor between these two factors should be identified to optimize the settling velocity of particles in viscoelastic fluids, which are
relevant to the design of many industrial processes.
In this paper, we present the results of a comparative study conducted to determine the prominent factor between the shear-rate-
dependent fluid viscosity and the fluid elasticity that governs the particle-settling velocity in viscoelastic polymer fluids. These experi-
ments are performed in a controlled manner to dissociate the effect of the shear viscosity and the effect of the elasticity on settling
velocity of particles. In this regard, first an experimental study is performed to investigate the effect of elasticity on the settling velocity
of particles in viscoelastic fluids that possess nearly identical shear-thinning viscosity properties but different elastic properties. Then,
the experiments to investigate the effect of shear viscosity on particle-settling velocity are performed using fluids of similar elasticity
but different shear viscosity. The effect of particle size on the settling velocities in these two sets of test fluids was also examined.

Materials and Methods


PIS Experimental Setup. The settling velocities of the particles in various fluids were measured using the PIS technique. This tech-
nique was also used to observe the movements of particles in the fluids. The PIS technique works on the fundamental principle that
when light travels across media of different refractive indices, the formation of shadows takes place (Settles 2001). The fluid media
must be transparent and must have significantly different refractive indices for the technique to work (Castrejón-Garcı́a et al. 2011).
However, the PIS technique does not depend on physical properties of the particle, such as size, shape, and density. The basic PIS mea-
surement setup requires a light source for providing a homogeneous background (this light also passes through different-density media
to form the shadows) and an image-acquisition device to capture the images of these shadows.
Fig. 1 provides the layout of the experimental setup used for measuring the settling velocities of the particles in this study. The light
source used was a 50-mJ double-pulsed Nd:YAG laser of Class 4 with a frequency of 15 Hz. The background illumination was better
achieved by attaching a high-efficiency diffuser. The laser beam was green in color, with 532-nm wavelength and 15-Hz frequency.
The image-acquisition device used for capturing the images was a double-framed camera with 12X lens. The camera had a
13761040-pixel charge-coupled-device sensor that could capture 5 frames/sec and deliver 12-bit digital images. This was a double-
frame camera that had the capability of seizing the images at two different exposure times (as short as 500 ns) within a very short time
interval. The double-frame camera should be synchronized with the double-pulsed laser to obtain the two frames at different exposures.
The working distance of the 12X lens was 32 to 341 mm, and the field of view in the current study was 1310 mm.
An acrylic-made transparent rectangular container (111170 cm) was used as the fluid-particle column in the experimental setup.
The size and velocity of the particles were obtained using commercial data-acquisition software. Using this, the positions of the par-
ticles in every frame were determined depending on the differences in their intensities, and the particle velocity was calculated using
particle displacement in two consecutive frames captured at defined time intervals, as shown in Fig. 2. A similar experimental setup
was used in the past (Shahi 2014; Shahi and Kuru 2015, 2016; Arnipally and Kuru 2017) to measure the settling velocities of particles.

Particles Used for Settling-Velocity Measurements. Four different sizes of glass spheres (1.8, 1.5, 2.0, and 3.0 mm) with specific
gravity (SG) of 2.51 and 2.5-mm-diameter steel spheres with SG of 8.05 were used for the settling-velocity measurements.

1690 October 2018 SPE Journal

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1691 Total Pages: 17

DAQ

Fluid Medium

12X Lens Diffuser


Particles
Double-Frame
Camera
Illumination
High-Efficiency
Diffuser Double-Pulsed
Laser

Fig. 1—Experimental-setup layout to measure settling velocities of particles using the PIS technique. DAQ 5 data acquisition software.

Particle, First Frame


Initial Shift X

Initial Shift Y

Window Size Y
Particle, Second Frame

Window Size X

Fig. 2—Schematic to illustrate the calculation procedure of particle velocity in the interrogation window (Shahi 2014).

Test-Fluid Preparation. Two sets of test fluids (Sets I and II) were formulated by mixing three different grades of HPAM (molecular
weights of 500,000, 8 million, and 20 million g/g mol) at polymer concentrations of 0.09, 0.05, and 0.03 wt%. These HPAM polymers
were water-soluble anionic polymers with a degree of hydrolysis of 25 to 30%.
The objective of the Set I test fluids was to prepare a sample of nearly identical shear-viscosity characteristics while showing signifi-
cantly different elastic properties. The Set II test fluids were prepared with the objective to have samples of similar elastic properties
but significantly different shear-viscosity characteristics.
The Set I test fluids were formulated in such a way that the average molecular weight of the polymer blend in all the test fluids was
constant and the molecular-weight distribution was different. The reason behind this is that the shear viscosity of the polymer solution
is primarily determined by its molecular weight. The average molecular weight of the polymer blend was calculated using Eq. 2
(Dehghanpour 2008; Urbissinova et al. 2010; Dehghanpour and Kuru 2011; Veerabhadrappa et al. 2013a, b).
Y
n
xi
Mw;B ¼ Mw;i : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ
i¼1

Molecular-weight distribution in the polymer blends was quantified in terms of the polydispersity index (I) (Zang et al. 1987;
Veerabhadrappa et al. 2013a). The polydispersity index of a polymer blend is defined as (Dehghanpour 2008)
! !
Mw Xn X n
xi
I¼ ¼ xi Mw;i  : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ
Mn i¼1 i¼1
Mw;i

The average molecular weight of the polymer blend in the Set I test fluids was maintained to be 8 million g/g mol. The concentration
of the polymer in Test Fluid 1 was 0.045 wt% and in other test fluids was 0.05 wt%. The composition details of polymer blends of the
Set I test fluids are provided in Table 1.

wt% of HPAM Polymer of Molecular Weight (g/g mol)


Polydispersity
Sample 20 million 8 million 500,000 Index (I)
Fluid 1 0 100 0 1
Fluid 2 35.64 52.58 11.8 3.6
Fluid 3 42.13 43.95 13.9 4.26

Table 1—Polymer-blend-composition details of Set I test fluids (average molecular weight 5 8 million
g/g mol).

October 2018 SPE Journal 1691

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1692 Total Pages: 17

The Set II test fluids were formulated by maintaining constant average molecular weight and molecular-weight distribution but dif-
ferent concentration of polymer in the test fluids. The HPAM polymer with molecular weight of 8 million g/g mol was used to prepare
the Set II test fluids. The composition details of the second set of test fluids are provided in Table 2. As can be seen, Test Fluid 5 is the
same as Test Fluid 1, and they were named differently in these two sets of test fluids to avoid confusion.

Concentration of HPAM
Polymer of Molecular Weight Polydispersity
Sample 8 million g/g mol (wt%) Index (I)
Fluid 4 0.03 1
Fluid 5 0.045 1
Fluid 6 0.09 1

Table 2—Polymer-blend-composition details of Set II test fluids.

The mixing procedure of the HPAM polymer to prepare the test fluids of Sets I and II is similar to the one followed by Foshee et al.
(1976). The steps followed to prepare the test fluids were the following:
1. Deionized water of calculated amount was taken and stirred continuously at 300 rev/min.
2. Appropriate amounts (depending on wt%) of HPAM polymers of three grades were added on to the well-developed vortex
shoulder in decreasing order of their molecular weights.
3. After adding the polymers, the solution was stirred at 100 rev/min for 2 hours. The stirring speed was reduced from 300 to
100 rev/min to avoid mechanical degradation of the polymers.
4. Once the stirring was completed, the solution was allowed to stand overnight to remove air bubbles. Filtration was not required
because all the test fluids of Sets I and II were clear and transparent.

Measurements of Rheological Properties. Test fluids were categorized depending on their rheological properties. The rheological
measurements such as shear viscosity and oscillatory measurements were performed on a low-torque commercial magnetic-bearing
rheometer in controlled-stress mode. A Peltier steel 988689 concentric cylinder (cup-and-bob) geometry was used; the diameters of the
cup and bob were 30 and 28 mm, respectively, and the bob length was 42 mm. All the rheological measurements were performed at
25 C; they were conducted by loading a fresh sample and allowing it to rest for 10 minutes so that material relaxation and thermal equi-
librium were obtained. Before conducting any measurements, sample preshearing was performed at high shear stress for 5 minutes to
remove the aging effect (if any) and 5 minutes were given to reach equilibrium. Rheological measurements of every test fluid were
repeated more than three times to ensure repeatability.

Measurement of Fluid Density. The density of the test fluids was determined by measuring the weight of the test fluid using the
weighing balance of intelligent-weighing technology that could precisely measure up to 0.1 mg, and the volume of the test fluid was
determined using a commercial Class A 10-cm3 measuring jar that could accurately measure volume up to 0.1 cm3. By later dividing
the weight of the test fluid by its volume, the density of the test fluid was calculated. The average value of three density readings was
taken for every test fluid.

Experimental Procedure. Once the test fluid was prepared, it was transferred into the rectangular fluid/particle column and the air
bubbles in the fluid were removed by leaving the fluid undisturbed overnight. The presence of any air bubbles in the test fluid would
result in noise in the captured images. The fluid/particle column was placed in such a way that the double-pulsed laser and the double-
frame camera were on the opposite sides of the column. The camera was connected to the computer that had data-acquisition software
running. A calibration sheet that had grids of dimensions 1.51.5 mm was inserted into the fluid/particle column. The 12X lens that
was mounted on the camera was adjusted so that the grid pattern of the calibration sheet was in focus in the 1310 mm field of view.
The double-pulsed laser was switched on when the camera temperature was less than –11 C and the laser power was adjusted to
20% for the first pulse (1A) and to 1% for the second pulse (1B). Depending on the particle sizes, the time interval between the two
images was decided to be within the range of 500 to 1,000 ms. For scaling the imaging system, the image of the grid pattern of the cali-
bration sheet was recorded and two points were chosen on the recorded image. Because the distance between the two points was known,
the mapping from pixel space to physical space was performed. Before removing the calibration sheet from the fluid column, its posi-
tion was marked so that it would be helpful to later drop the particles in focus.
A sufficient number of images of the fluid medium without particles were captured and the average of those recorded images was
used as a reference image. Spherical particles were then dropped in the test column from the earlier marked position and experimental
images were recorded. Because of differences in the refractive indices, particles appeared as dark spots on the display screen when they
passed through the plane of focus. Images were recorded only when the particles were in focus. Later, using the PIS operation in the
software, all the experimental images were processed to determine the particle size and settling velocity.
Once the first particle was dropped in the test fluid, it was found that a 20-minute waiting time was required before the second parti-
cle was released. This waiting time allows the fluid to recover from any disturbance caused by the first settling particle. All the velocity
values were confirmed to be constant terminal settling velocities by performing the measurements at two different heights (20 and
26 cm) from the air/liquid interface. Accuracy and repeatability of the velocity values were ensured by taking a sufficient number of
measurements (at least 10) and reporting them along with the error bars.

Calibration of PIS Measurements. Calibration of the PIS technique was performed by measuring the diameters and the settling veloc-
ities of four different glass spheres in water. The comparison of measured sphere diameters against the manufacturer specifications is
provided in Table 3.
The settling velocities of these four glass spheres measured by the PIS technique were compared with the settling-velocity values
reported by Shahi (2014), and the details are given in Table 4.

1692 October 2018 SPE Journal

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1693 Total Pages: 17

Diameter Provided by Diameter Obtained from


Manufacturer (mm) PIS Measurements (mm)
0.71±0.02 0.718±0.017
1.18±0.02 1.186±0.02
1.5±0.03 1.59±0.03
2.00±0.04 2.01±0.03

Table 3—Comparison of glass-sphere diameters obtained from PIS


measurements with manufacturer specifications.

Settling Velocities (m/s)


Glass-Sphere From PIS Reported Values
Diameter (mm) Measurements (Shahi 2014) Deviation (%)
0.71 0.11 0.1141 3.6
1.18 0.19 0.1845 3.0
1.50 0.23 0.2347 2.0
2.00 0.27 0.2740 1.5

Table 4—Comparison of glass-sphere settling velocities obtained from PIS measurements with
reported values in the literature.

PIS measurements were further substantiated by calculating the drag coefficients for four glass spheres using Eq. 4 (McCabe et al.
2004) and plotting those values on the correlation graph of universal drag coefficient vs. particle Reynolds number, as shown in Fig. 3.
The universal drag coefficient vs. Rep correlation is given in Eq. 5 (Morrison 2013) and the particle Reynolds number is determined
using Eq. 6 (McCabe et al. 2004).
4gDs ðqs  qf Þ
CD ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð4Þ
3qf Vs2
   
Rep Rep 7:94 !
2:6 0:411
24 5:0 26; 300 Re0:80
p
CD ¼ þ   þ   þ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð5Þ
Rep Rep 1:52 Rep 8:00 461; 000
1þ 1þ
5:0 26; 300
Dp V s q f
Rep ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð6Þ
lf

All the drag-coefficient values calculated from experimentally measured settling-velocity values using the PIS technique were
within 5% deviation from the theoretical universal drag coefficient values.

102
Universal-drag coefficients for spheres
CD values obtained from PIS measurements
Drag Coefficient (CD)

101

100

10–1

100 101 102 103 104 105 106


Particle Reynolds Number (Rep)

Fig. 3—PIS measured values of CD plotted on the curve of universal drag coefficient vs. particle Reynolds number (Rep).

Results and Discussion


The densities of all the test fluids are provided in Table 5. All the fluids seemed to have practically the same densities.

Rheological Characterization of the Set I Test Fluids. All the test fluids of Set I were rheologically characterized using their viscous
and elastic properties. Figs. 4 and 5 provide the shear-stress vs. shear-rate profiles and the shear-viscosity vs. shear-rate profiles for all

October 2018 SPE Journal 1693

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1694 Total Pages: 17

the Set I test fluids, respectively. Within the studied range of shear rate (from 1 to 200 s1), all the test fluids were found to exhibit the
shear-thinning behavior as their shear viscosities were decreasing with increasing shear rate. The data shown in Figs. 4 and 5 confirmed
that Test Fluids 1, 2, and 3 had similar viscous characteristics.

3
Test Fluid Density (kg/m )
Fluid 1 998
Fluid 2 997
Fluid 3 997
Fluid 4 998
Fluid 5 998
Fluid 6 1005

Table 5—Densities of test fluids of Sets I and II.

101
Fluid 1
Fluid 2
Fluid 3
Shear Stress (Pa)

100

10–1
10–1 100 101 102 103
Shear Rate (1/seconds)

Fig. 4—Shear-stress vs. shear-rate profiles of Set I test fluids.

100
Fluid 1
Fluid 2
Fluid 3
Viscosity (Pa.s)

10–1

10–2

10–3
10–1 100 101 102 103
Shear Rate (1/seconds)

Fig. 5—Shear-viscosity vs. shear-rate profiles of Set I test fluids.

To further verify the similarity, a power-law model was fitted to shear-stress vs. shear-rate data of all the test fluids. The values of
consistency index (K) and the flow-behavior index (n) are provided in Table 6. The values of K and n also substantiate the fact that the
Set I fluids had nearly identical viscous characteristics.

n
K (Pa·s ) R
2
Test Fluid n
Fluid 1 0.27 0.35 0.99
Fluid 2 0.25 0.36 0.98
Fluid 3 0.26 0.35 0.98

Table 6—Power-law parameters of K and n of Set I test fluids. R2 is


the least-squares regression correlation coefficient.

1694 October 2018 SPE Journal

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1695 Total Pages: 17

Oscillatory frequency-sweep measurements were performed to study the viscoelastic properties of the Set I test fluids. The storage-
modulus (G0 ) and the loss-modulus (G00 ) profiles as a function of angular frequency for the Set I test fluids (Fluids 1, 2, and 3) are shown
in Fig. 6. The storage modulus (G0 ) represents the solid-like behavior of the test fluid, whereas the liquid-like behavior is represented
by the loss modulus (G00 ). As presented in Fig. 6, initially the storage moduli of the test fluids were lower than their loss moduli. For all
three test fluids, both the storage and the loss moduli were increasing as the angular frequency increases. Nevertheless, the increase in
the storage modulus was more prompt compared with the increase in the loss modulus.

100 100
Storage modulus, G′ Storage modulus, G′
Loss modulus, G″ Loss modulus, G″
Storage and Loss Moduli (Pa)

Storage and Loss Moduli (Pa)


10–1 10–1

10–2 10–2

10–3 10–3
10–3 10–2 10–1 100 101 10–3 10–2 10–1 100 101
Angular Frequency (rad/s) Angular Frequency (rad/s)
(a) Fluid 1 (b) Fluid 2

100
Storage modulus, G′
Loss modulus, G″
Storage and Loss Moduli (Pa)

10–1

10–2

10–3
10–3 10–2 10–1 100 101
Angular Frequency (rad/s)
(c) Fluid 3

Fig. 6—Oscillatory frequency sweep data for Set I test fluids.

The frequency at which the storage modulus crosses over the loss modulus is called the crossover frequency. The inverse of the
crossover frequencies provides the longest characteristic relaxation times of the test fluids (Larson 1998; Sunthar 2010). Elasticity of a
fluid can be quantified by its relaxation time because it is the time needed for any deformed material to regain its original structure
(Choi 2008). Therefore, in this study, elastic properties of the test fluids were measured in terms of their relaxation times. The higher
the relaxation time, higher is the elastic property of the test fluid compared with other test fluids.
Table 7 presents the longest relaxation times of Set I test fluids, and as expected, polydispersity seemed to be controlling the elastic
properties of these polymer solutions. It was found that Fluid 3 (having higher polydispersity) had the highest relaxation time among
the three test fluids. Hence, Test Fluid 3 was more elastic, followed by Test Fluid 2 and then Test Fluid 1. Based on the previous rheo-
logical characterization, it was concluded that all the test fluids in Set I had nearly identical shear-thinning viscous properties (on
average, K ¼ 0.26 Pasn and n ¼ 0.35) but significantly different elasticity properties.

Polydispersity Longest Relaxation


Test Fluid Index (I) Time (seconds)
Fluid 1 1 12
Fluid 2 3.6 50
Fluid 3 4.26 110

Table 7—Longest relaxation times of the Set I test fluids.

October 2018 SPE Journal 1695

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1696 Total Pages: 17

Rheological Characterization of the Set II Test Fluids. Figs. 7 and 8 provide the shear-stress vs. shear-rate profiles and the shear-
viscosity vs. shear-ate profiles for all the Set II test fluids. The rheology data of Fluid 6 are also included in Arnipally and Kuru (2017).

101
Fluid 4
Fluid 5
Fluid 6

Shear Stress (Pa)


100

10–1
10–1 100 101 102 103
Shear Rate (1/seconds)

Fig. 7—Shear-stress vs. shear-rate profiles of Set II test fluids (longest relaxation time 5 12 seconds).

101
Fluid 4
Fluid 5
Fluid 6
100
Viscosity (Pa.s)

10–1

10–2

10–3
10–1 100 101 102 103
Shear Rate (1/seconds)

Fig. 8—Shear-viscosity vs. shear-rate profiles of Set II test fluids (longest relaxation time 5 12 seconds).

Within the studied range of shear rates (from 1 to 200 s1), all the test fluids were found to exhibit shear-thinning behavior as their
shear viscosities were decreasing with the increasing shear rate. However, among the three, Fluid 6 seemed to have the highest shear
viscosity, Fluid 4 to have the lowest shear viscosity, and the Fluid 5 shear viscosity was in between the other two.
The consistency-index (K) and the flow-behavior-index (n) values of the test fluids were obtained by curve fitting the power-law
model to the shear-stress vs. shear-rate data. The data provided in Table 8 show that all three test fluids had shear-thinning characteris-
tics. The flow-behavior-index values were approximately the same for the Set II test fluids. The consistency index of Fluid 6 was, how-
ever, greater than that of Fluid 5, followed by Fluid 4. Fluid 6 was found to be nearly two times more viscous than Fluid 5 and was also
nearly 3.5 times more viscous than Fluid 4.

n
K (Pa·s ) R
2
Test Fluid n
Fluid 4 0.16 0.38 0.99
Fluid 5 0.27 0.35 0.99
Fluid 6 0.53 0.38 0.96

Table 8—Power-law parameters of K and n of the Set II test fluids.


R2 is the least-squares regression correlation coefficient.

Oscillatory frequency sweep measurements were performed to study the viscoelastic properties of the Set II test fluids. The storage-
modulus (G0 ) and the loss-modulus (G00 ) profiles as a function of the angular frequency for the Set II test fluids (Fluids 4, 5, and 6) are
shown in Fig. 9. For all three test fluids of Set II, both the storage and the loss moduli increased as the angular frequency increased.
The longest relaxation times of the test fluids were determined by taking the inverse of the crossover frequency. As shown in Table 9,
Set II test fluids had the same longest relaxation times of 12 seconds. Therefore, Fluids 4, 5, and 6 were similar in terms of elasticity.
From the results given in Tables 8 and 9, it could be concluded that Fluids 4, 5, and 6 of Set II had similar elasticity but very differ-
ent shear viscosity.

1696 October 2018 SPE Journal

ID: jaganm Time: 11:05 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1697 Total Pages: 17

100 100
Storage modulus, G′ Storage modulus, G′
Loss modulus, G″ Loss modulus, G″

Storage and Loss Moduli (Pa)

Storage and Loss Moduli (Pa)


10–1 10–1

10–2 10–2

10–3 10–3
10–3 10–2 10–1 100 101 10–3 10–2 10–1 100 101
Angular Frequency (rad/s) Angular Frequency (rad/s)
(a) Fluid 4 (b) Fluid 5

101
Storage modulus, G′
Loss modulus, G″
Storage and Loss Moduli (Pa)

100

10–1

10–2
10–3 10–2 10–1 100 101
Angular Frequency (rad/s)
(c) Fluid 6

Fig. 9—Oscillatory frequency sweep data for Set II test fluids.

Polydispersity Longest Relaxation


Test Fluid Index (I) Time (seconds)
Fluid 4 1 12
Fluid 5 1 12
Fluid 6 1 12

Table 9—Longest relaxation times of Set II test fluids.

Effect of Elasticity on Particle-Settling Velocities. The settling velocities of spherical glass particles of 2-mm diameter in Set I test
fluids with similar shear viscosity (K ¼ 0.26 Pasn and n ¼ 0.35) and different elasticity were measured using the PIS technique. The
settling-velocity experimental results along with standard deviations are summarized in Table 10. The results indicated that the
settling-velocity value of 2-mm spheres in Fluid 1 was higher than that of the values in Fluid 2, followed by Fluid 3.

Relaxation
Time, λ Settling Standard
Test Fluid (seconds) Velocity (m/s) Deviation
–4
Fluid 1 12 0.0307 3.60×10
–4
Fluid 2 50 0.0144 3.40×10
–4
Fluid 3 110 0.0084 4.50×10

Table 10—Settling velocities of spherical glass particles of 2-mm


diameter in Set I test fluids of K 5 0.26 Pasn and n 5 0.35.

The hinderance effect caused by container walls was negligible because the ratio of sphere diameter to the test-column width is very
low (less than 0.02). Because the shear-thinning-viscosity behavior of all the fluids was approximately identical, the variation in the
settling-velocity values was predominantly caused by the differences in their elastic properties. Therefore, the settling-velocity values

October 2018 SPE Journal 1697

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1698 Total Pages: 17

of 2-mm spheres are plotted against the fluid-relaxation times in Fig. 10 to show the effect of elasticity. The model proposed by Shah
et al. (2007) for predicting settling velocity of spheres in viscoinelastic power-law-type fluids was used to determine the settling veloc-
ity of 2-mm spheres in the fluid with K ¼ 0.26 Pasn and n ¼ 0.35 without any elasticity. This value is plotted in Fig. 10 as the
settling velocity at zero relaxation time. The details of the settling-velocity calculations using the Shah et al. (2007) model are given in
Appendix A.

0.5

Settling Velocity (m/s)


0.4

0.02

0.00
0 20 40 60 80 100 120
Relaxation Time (seconds)

Fig. 10—Settling velocity of glass spheres of 2-mm diameter vs. relaxation time of the Set I test fluids with similar viscosity but dif-
ferent elasticity.

Results shown in Fig. 10 demonstrate that the particle-settling velocity is a strong function of the fluid-relaxation time. The settling
velocity of the glass spheres decreased significantly as the relaxation time (and, hence, the elasticity) of the fluid increased. As the relax-
ation time increased from 12 to 50 seconds, the settling velocity decreased by approximately two times. The reduction in settling veloc-
ity was three times when the relaxation time was increased from 12 to 110 seconds.
The settling velocities of spherical steel particles (2.5-mm diameter and SG of 8.05) in Set I test fluids with similar shear viscosity
(K ¼ 0.26 Pasn and n ¼ 0.35) and different elasticity were also measured using the PIS technique. The measured settling-velocity values
along with standard deviations are summarized in Table 11. Results have shown that the settling velocity of high-density steel particles
were also reduced significantly with the increasing relaxation times.

Relaxation
Time, λ Settling Standard
Test Fluid (seconds) Velocity (m/s) Deviation
–2
Fluid 1 12 0.18 1.20×10
–3
Fluid 2 50 0.06 7.80×10
–3
Fluid 3 110 0.04 3.80×10

Table 11—Settling velocities of spherical steel particles (2.5-mm


diameter and SG 5 8.05) in Set I test fluids of K 5 0.26 Pasn and
n 5 0.35.

These results showing the effect of relaxation time (and hence, the elasticity) on the settling velocity of particles were in agreement
with Gomaa et al. (2015), who observed an increase in the fluids’ ability to suspend proppant particles when the crossover point of G0
and G00 in the frequency sweep of the fluid was decreased.
We have also calculated the ratio of the settling velocity of 2-mm glass spheres and 2.5-mm steel spheres in viscoinelastic shear-
thinning fluids [using the Shah et al. (2007) model; K ¼ 0.26, n ¼ 0.35, k ¼ 0) to the settling velocities of the same particles measured in
fluids with different elasticities (k ¼ 12, 50, and 110 seconds). As shown in Fig. 11, for 2-mm glass spheres, when the relaxation time of
the fluid was 12 seconds, the settling velocity was 14 times lower than that of the ones in viscoinelastic fluid with nearly identical shear
viscosity. Similarly, when we used the fluids with relaxation times of 50 and 110 seconds, the measured settling velocities of the glass
particles were approximately 29 times and 50 times lower than that of the case with viscoinelastic fluid having the similar shear viscos-
ity, respectively. The settling velocities of the steel spheres measured in viscoelastic fluids of different elasticities (k ¼ 12, 50, and
110 seconds; K ¼ 0.26 and n ¼ 0.35) were also significantly lower than the ones calculated using a model given for viscoinelastic fluids
(Shah et al. 2007) (K ¼ 0.26, n ¼ 0.35, k ¼ 0). However, the reduction in the settling velocity of spherical steel particles with increasing
elasticity was much more pronounced (varying from 140 to 630 times within the current experimental conditions) than that of the case
with glass beads (varying from 14 to 50 times).
These results clearly indicate that settling-velocity values can be significantly overpredicted if they are calculated using models that
do not consider the elasticity effect (Shah et al. 2007) and the overprediction can be more pronounced when higher-density particles are
involved. These results also substantiate the fact that the settling velocity of the particle can be reduced significantly by enhancing the
elasticity of the fluid and, therefore, this vital elasticity effect on the particle-settling velocity should be taken into consideration to
develop more-realistic models of particle-settling velocity in viscoelastic fluids.

1698 October 2018 SPE Journal

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1699 Total Pages: 17

2-mm glass spheres (SG = 2.51)


600 2.5-mm steel spheres (SG = 8.05)

400

Vinelastic/Velastic
200

40

20

0 20 40 60 80 100 120
Relaxation Time (seconds)

Fig. 11—Overestimation of settling velocities if the effect of elasticity is not considered.

Effect of Shear Viscosity on the Particle-Settling Velocity. The effect of the shear viscosity of the fluid on the settling velocity of
particles was studied by evaluating the settling velocities of 2-mm spheres in Set II fluids with relaxation time of 12 seconds. The appa-
rent shear viscosity of the fluids was approximated by (Lali et al. 1989)

la ¼ Kcn1 : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð7Þ

In Eq. 7, the maximum value of shear rate c induced by the settling particle in polymer fluids was calculated using Eq. 1 (Chhabra
2007). The measured values of settling-velocity values were used to evaluate the shear rates (Eq. 1), which were then used to determine
the apparent shear viscosity. The settling-velocity results along with apparent shear viscosities and other rheological properties (K, n) of
the fluids are shown in Table 12.

Consistency Index Flow-Behavior Apparent Shear Settling Velocity


n
Test Fluid K (Pa·s ) Index n Viscosity (Pa·s) (m/s) Standard Deviation
–3
Fluid 4 0.16 0.38 0.0167 0.039 1.70×10
–4
Fluid 5 0.27 0.35 0.0281 0.0307 3.60×10
–4
Fluid 6 0.53 0.38 0.1198 0.011 5.90×10

Table 12—Settling velocities of spherical glass particles of 2-mm diameter in Set II test fluids along with apparent shear viscosities and other
rheological properties (K, n).

Note that measured settling-velocity values consider the effects of the shear viscosity as well as the elasticity. Therefore, shear rates
estimated using Eq. 1 are actually lower than what would have been observed in fluids with the same shear viscosity but no elasticity;
as a result, the apparent viscosity values calculated using Eq. 7 (and reported in Table 12) are actually on the high side.
The settling velocity of the particles was the highest in the test fluid that had the lowest apparent shear viscosity (Test Fluid 4) and
was the lowest in the test fluid that had the highest apparent shear viscosity (Test Fluid 6). Because the elasticities of all the test fluids
in Set II were the same, the disparity in the settling-velocity results was attributed to the variation in shear-viscosity characteristics of
the fluids. Fig. 12 shows the correlation between the particle-settling velocity and the apparent shear viscosity of the fluids.
The particle-settling velocity also correlated well with the consistency index K. As the consistency index K of the fluid increased,
the particle-settling velocity reduced noticeably. Because the flow-behavior indices n of the fluids were nearly the same, we could say
that there was strong correlation between the drop in the settling velocity and the increase in the consistency index. When K changed
from 0.16 to 0.53 Pasn, the settling velocity decreased by nearly 3.5 times. The correlation between the particle-settling velocity and
the consistency index K of the fluids is shown in Fig. 13.

Effect of Particle Size on Settling Velocities. In general, the size of the particle has a pivotal role in regulating the particle-settling
velocity in any type of fluid. To assess the effect of particle size on the settling velocity of particles in shear-thinning viscoelastic fluids,
glass spheres of varying diameters (1.18 to 3.0 mm) were dropped in Set I and Set II test fluids and the corresponding settling velocities
were measured. The settling velocities of spherical glass particles of various diameters in Set I test fluids with similar shear-thinning
behavior (K ¼ 0.26 Pasn and n ¼ 0.35) and significantly different elastic properties are provided along with the Weissenberg numbers
in Table 13; that of in Set II test fluids with the same elastic property (k ¼ 12 seconds) and different shear viscosities are provided along
with the Weissenberg numbers in Table 14.
In Fig. 14, the settling-velocity values are also plotted against the particle diameter for comparison purpose. In Set I fluids, the settling
velocities of all particles in Fluid 1 (k ¼ 12 seconds) were the highest, followed by Fluid 2 (k ¼ 50 seconds) and then Fluid 3 (k ¼ 110 sec-
onds). In Set II fluids, the settling velocities of all particles in Fluid 4 (K ¼ 0.16 Pasn) were highest, followed by Fluid 5 (K ¼ 0.26 Pasn)
and then Fluid 6 (K ¼ 0.53 Pasn). This reasserts the observation that when the fluid shear viscosity is constant, particle-settling velocity
will be lower in the fluids with the higher elasticity (Fluid 3), and when the fluid elasticity is constant, particle-settling velocity will be
lower in the fluids with the higher shear viscosity (Fluid 6). In addition, the settling-velocity values were overall in an increasing trend
with respect to the increase in particle diameter for all test fluids in Sets I and II. It was, however, noticeable that the magnitude of the

October 2018 SPE Journal 1699

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1700 Total Pages: 17

change in settling-velocity values was not the same for all the test fluids. The magnitude of the increase in settling-velocity values (with
increasing particle size) seemed to occur at a lower rate when the relaxation time of the fluid was increased, as shown in Fig. 14a, than
that of the case when the fluid consistency index (as well as the fluid shear viscosity) was increased, as shown in Fig. 14b.

0.045

0.040

0.035

Settling Velocity (m/s)


0.030

0.025

0.020

0.015

0.010

0.00 0.02 0.04 0.06 0.08 0.10 0.12


Apparent Shear Viscosity (Pa.s)

Fig. 12—Settling velocity of spherical glass particles of 2-mm diameter vs. apparent shear viscosity of the Set II test fluids.

0.040

0.035
Settling Velocity (m/s)

0.030

0.025

0.020

0.015

0.010

0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55


Consistency Index K (Pa.sn)

Fig. 13—Settling velocity of glass spheres of 2-mm diameter in the Set II test fluids with similar elasticity (k 5 12 seconds) and dif-
ferent consistency index (K).

Diameter of
Relaxation Time Spherical Particles Settling Velocity
Test Fluid (seconds) (mm) (m/s) Standard Deviation Weissenberg Number
–4
1.18 0.0052 4.52×10 105.2
–4
1.5 0.0163 4.43×10 260.6
Fluid 1 12 –4
2.0 0.0307 3.60×10 368.6
–3
3.0 0.0788 2.74×10 630.1
–5
1.18 0.0034 9.42×10 291.0
–4
1.5 0.0093 5.58×10 621.7
Fluid 2 50 –4
2.0 0.0144 3.40×10 721.7
–4
3.0 0.0353 8.45×10 1175.0
–4
1.18 0.0025 1.11×10 459.6
–4
1.5 0.0058 4.03×10 849.1
Fluid 3 110 –4
2.0 0.0084 4.50×10 920.3
–3
3.0 0.0220 2.57×10 1616.0

Table 13—Settling velocities of spherical glass particles of various diameters in Set I test fluids with different elasticity and similar shear
viscosity (K 5 0.26 Pasn and n 5 0.35).

1700 October 2018 SPE Journal

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1701 Total Pages: 17

Consistency Diameter of Spherical Settling Velocity


n
Test Fluid Index K (Pa·s ) Particles (mm) (m/s) Standard Deviation Weissenberg Number
–4
1.18 0.0114 5.65×10 231.9
–4
1.5 0.0245 4.50×10 392.0
Fluid 4 0.16 –3
2.0 0.039 1.70×10 468.0
–3
3.0 0.0916 1.86×10 732.8
–4
1.18 0.0052 4.52×10 105.2
–4
1.5 0.0163 4.43×10 260.6
Fluid 5 0.27 –4
2.0 0.0307 3.60×10 368.6
–3
3.0 0.0788 2.74×10 630.1
–4
1.18 0.0017 1.08×10 34.6
–4
1.5 0.0049 2.15×10 78.4
Fluid 6 0.53 –4
2.0 0.0110 5.90×10 129.6
–4
3.0 0.0294 5.29×10 235.2

Table 14—Settling velocities of spherical glass particles of various diameters in Set II test fluids with different shear viscosities and similar
elasticity (k 5 12 seconds).

Fluid 1 (λ = 12 seconds) Fluid 4 (K = 0.16)


0.08 0.10 Fluid 5 (K = 0.27)
Fluid 2 (λ = 50 seconds)
Fluid 3 (λ = 110 seconds) Fluid 6 (K = 0.53)
Settling Velocity (m/s)

0.08

Settling Velocity (m/s)


0.06
0.06
0.04
0.04

0.02
0.02

0.00 0.00

1.0 1.5 2.0 2.5 3.0 1.0 1.5 2.0 2.5 3.0
Particle Diameter (mm) Particle Diameter (mm)
(a) (b)

Fig. 14—Settling velocity vs. diameter of glass spheres in (a) Set I test fluids with similar shear viscosity (K 5 0.26 Pasn and
n 5 0.35) but different elasticity; (b) Set II test fluids with similar elasticity (k 5 12 seconds) but different shear viscosity.

To identify the change in the magnitude of the settling velocities caused by particle-size increase, we have defined an “increase
factor” as the ratios of settling velocities of 1.5-, 2-, and 3-mm particles to that of 1.18-mm particles. The increase factors for different
particle sizes and fluid combinations were calculated and tabulated in Table 15.

Set I Fluids Set II Fluids


Ratio of Velocities of Fluid 1 Fluid 2 Fluid 3 Fluid 4 Fluid 5 Fluid 6
Spheres of Diameter (λ = 12 seconds) (λ = 50 seconds) (λ = 110 seconds) (K = 0.16) (K = 0.26) (K = 0.53)
1.5/1.18 mm 3.2 2.7 2.3 2.2 3.2 2.9
2/1.18 mm 5.9 4.2 3.4 3.4 5.9 6.4
3/1.18 mm 15.2 10.3 8.9 8.0 15.2 17.2

Table 15—Increase factor of settling velocities of glass spheres in Set I and Set II test fluids.

For the Set I fluids, the increase factor in particle-settling velocity (caused by increasing particle size) was found to be decreasing
from Fluid 1 to Fluid 3 as the relaxation times (and hence the elasticity) of the fluids increased from 12 to 110 seconds. That is, larger
particles settle down at a slower rate as the elasticity of the fluid increases. In other words, as the particle size increases, the increase in
particle-settling velocity occurs at a relatively lower rate as the fluid elasticity increases. From the practical point of field operations,
these results imply that increasing the fluid elasticity may be a good solution for controlling particle-settling velocity of large-size
drilled cuttings.
For the Set II fluids, the magnitude of the increase in the settling velocity of the particles (caused by increasing particle size) was
observed at a higher rate as the shear viscosity of the fluid increased. In other words, the effect of particle-size increase on the particle-
settling velocity becomes more dominant at the high-shear-viscosity values. From the practical point of field operations, these results
imply that increasing the shear viscosity may not be the most-effective solution for controlling particle-settling velocity of large-size
drilled cuttings.
Using these experimental results, it can be concluded that increasing the fluid elasticity (rather than the fluid shear viscosity) would
be a more-effective way of controlling the particle-settling velocity when dealing with large-size drilled cuttings. However, consequen-
tial effects of increasing fluid elasticity on the other operational parameters are still uncertain and yet to be investigated. One area of

October 2018 SPE Journal 1701

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1702 Total Pages: 17

interest, for example, would be to determine how the increasing fluid elasticity would influence the frictional-pressure losses while drill-
ing deep vertical wells, and/or long horizontal and extended-reach wells.
An experimental study of measuring settling velocity of spherical particles in viscoelastic surfactant systems was performed by
Malhotra and Sharma (2012), in which they calculated the settling velocity of particles in inelastic fluid (VIE ) using a correlation pro-
Velastic
vided by Renaud et al. (2004). They calculated the velocity ratio of and plotted against the Weissenberg number. For comparison
VIE
purpose, a similar graph has been plotted in Fig. 15 using our experimental data and calculating VIE using the same correlation used by
Malhotra and Sharma (2012). It should be noted that the Reynolds-number values in the current study were more than 0.1, so they were
in the noncreeping-flow regime.

1.0

0.9

0.8

0.7

0.6
Velastic/VIE

0.5

0.4

0.3

0.2

0.1

0.0
0 200 400 600 800 1,000 1,200 1,400 1,600 1,800
Weissenberg Number

Fig. 15—Particle-settling-velocity ratios vs. Weissenberg numbers observed in the current experimental study.

 
Velastic
The results shown in Fig. 15 indicate that the velocity ratio initially increases with the increasing Weissenberg number
VIE
within the lower ranges of Weissenberg numbers, and later the velocity ratio decreases with the increasing Weissenberg number within
the higher ranges of Weissenberg numbers. Although there is some scatter in the data, these observations were in agreement with those
reported by Malhotra and Sharma (2012). However, it should be noted that the Weissenberg-number values observed in the current
study were significantly higher than those reported by Malhotra and Sharma (2012). Because the wide range of Weissenberg numbers is
common in the practical-field conditions, measurements at broader Weissenberg-number limits would be instructive.
Malhotra and Sharma (2012) proposed a correlation for the settling velocity in shear-thinning viscoelastic fluids in unconfined
media. Using the Malhotra and Sharma (2012) correlation [i.e., Eq. 17 in Malhotra and Sharma (2012)], the ratio of settling-velocity
values was calculated and plotted along with the experimental settling-velocity-ratio values vs. the inelastic Weissenberg number
(WiIE ), as shown in Fig. 16. The Malhotra and Sharma (2012) correlation was developed for the following ranges of the Weissenberg
number, the Reynolds number, and the n value: 0.15 < Wi < 50; 3.8104 < Re < 37; and 0.259 < n < 0.9, respectively, where Wi and
Re were calculated using the inelastic settling-velocity correlation given by Renaud et al. (2004).

1.0
PIS experimental results
Malhotra and Sharma (2012) correlation
0.8

0.6

0.4

0.2
Velastic/VIE

0.025
0.020
0.015
0.010
0.005
0.000
102 103 104
WiIE

Fig. 16—Comparison of the experimental settling-velocity ratios with those calculated from the Malhotra and Sharma
(2012) correlation.

1702 October 2018 SPE Journal

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1703 Total Pages: 17

The results showed that the settling-velocity-ratio values obtained from the current experimental measurements were significantly
higher than those calculated from the Malhotra and Sharma (2012) correlation. The pronounced differences in the settling-velocity val-
ues could be because the Malhotra and Sharma (2012) correlation was valid for the inelastic Weissenberg numbers ranging from 0.15
to 50. However, the inelastic Weissenberg numbers observed in the current experimental study (320 to 7,700) were significantly higher
than the ranges of Weissenberg numbers applied in the Malhotra and Sharma (2012) study. In addition, the Malhotra and Sharma
(2012) model was developed using the experimental measurements of particle-settling velocities in viscoelastic surfactant systems,
whereas the measurements in the current study were conducted using polymer-based fluids. The surfactant-based fluid systems (because
of their hydrophobic or hydrophilic characteristics) might alter the drag on the settling particles, thereby affecting the settling velocity
of particles. Therefore, the differences in the nature of fluid systems (i.e., surfactant-based vs polymer-based fluid systems) would also
be one of the causes of the observed variations in the values of settling-velocity ratios.

Recommendations for Future Work


In this study, we have used spherical particles to minimize the influence of the any other particle-related effects (such as particle shape)
on the particle-settling velocity because we were solely focusing on how to determine the effects of the fluid elasticity and the shear vis-
cosity. It is known that the particle shape also affects the particle-settling velocity. The term “sphericity” was introduced previously to
account for the particle-shape effect on the settling velocity (Chien 1994). We recommend and plan to use natural-quartz sand and other
available natural-rock cuttings (e.g., limestone and marble) in similar tests in the future.
The fluids used in this study are shear-thinning power-law fluids. Future studies should also include Bingham-plastic and yield-
power-law fluids. In particular, investigating the effect of yield stress vs. elasticity on the particle-settling velocity would be of high
interest because the yield stress is also commonly said to have a strong influence on the particle-settling velocity.
The thixotropy is a characteristic of a fluid to form a gelled structure over the time when not subject to shearing, and then to liquefy
when agitated. The viscosity of a thixotropic fluid varies with time under constant shear rate until it reaches an equilibrium. Most drill-
ing fluids exhibit thixotropy, which is desired for efficient cuttings lifting and suspension of weighting material when drilling-fluid cir-
culation stops. In this study, we did not consider the effect of thixotropy on the settling velocity of cuttings. We recommend that
investigating the effect of the drilling-fluid thixotropy on the settling velocity of cuttings would be a good subject for a future study to
complement the results presented here.

Conclusions
Two sets of test fluids were prepared by mixing HPAM polymer of three different molecular weights. Set I test fluids were prepared by
maintaining average molecular weight of the polymer blend at 8 million g/g mol and varying the polydispersity index. Set II fluids were
prepared using HPAM polymer of molecular weight of 8 million g/g mol at different polymer concentrations.
Results of rheological measurements indicated that Set I test fluids had nearly identical shear viscosity but significantly different
elasticity properties and Set II fluids had similar elasticity but different shear viscosities.
The settling velocities of glass spheres in these two sets of test fluids were measured using the PIS technique, and the experimental
results showed the following:
1. In fluids of nearly identical shear viscosity, the settling velocity of spherical particles decreased significantly with the increasing
fluid elasticity.
2. In fluids of constant elasticity, the settling velocity of spherical particles also decreased significantly when the fluid shear viscosity
was increased.
3. The settling velocity of spherical particles increased significantly in both sets of test fluids as the particle diameter increased from
1.18 to 3 mm. However, the magnitude of the increase in settling velocity with the increasing particle diameter was considerably
less for the fluids with higher elasticity and similar shear-viscosity characteristics.
4. The experimentally measured settling velocities of spherical glass particles (SG ¼ 2.51) in fluids of nearly identical shear viscosity
but different elasticity were compared with the settling-velocity values calculated from the Shah et al. (2007) model developed for
predicting the settling velocity of spherical particles in viscoinelastic power-law-type fluids. Results showed that the settling-velocity
values of glass spheres could be overpredicted by a range of 14 to 50 times if the effect of the fluid elasticity was not considered.
5. The experimentally measured settling velocities of spherical steel particles (SG ¼ 8.05) in the fluids of nearly identical shear viscos-
ity but different elasticity were compared with the settling-velocity values calculated from the Shah et al. (2007) model developed
for predicting the settling velocity of spherical particles in viscoinelastic power-law-type fluids. Results showed that the settling-
velocity values of glass spheres could be overpredicted by a range of 140 to 630 times if the effect of the fluid elasticity was
not considered.
In conclusion, the fluid shear viscosity and elasticity both seem to have significant effects on the particle-settling velocity. However,
from the field operational point of view, fluids with high-shear-viscosity values may not always be practical to use because the high-
shear-viscosity values increase the parasitic pressure losses. In such cases, increasing the fluid elasticity may help to reduce the particle-
settling velocity even when using fluids with low-shear-viscosity values, especially when large-sized drill cuttings need to be trans-
ported. Although practical problems associated with the increasing elasticity of fluids are yet to be investigated, leveraging the effect of
elasticity to reduce the settling velocity of particles appears to be beneficial.

Nomenclature
CD ¼ drag coefficient
Dp ¼ characteristic length of a particle, m
Ds ¼ diameter of the sphere, m
g ¼ acceleration caused by gravity, m/s2
G0 ¼ elastic modulus, Pa
G00 ¼ viscous modulus, Pa
I ¼ polydispersity index
K ¼ flow consistency index, Pasn
Mn ¼ number-average molecular weight, g/g mol
Mw ¼ weighted-average molecular weight, g/g mol
Mw;B ¼ average molecular weight of the polymer blend, g/g mol

October 2018 SPE Journal 1703

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1704 Total Pages: 17

Mw;i ¼ molecular weight of polymer grade i, g/g mol


n ¼ flow-behavior index
Rep ¼ particle Reynolds number
Vs ¼ settling velocity, m/s
VIE ¼ settling velocity of particles in inelastic fluid calculated using a correlation provided by Renaud et al. (2004), m/s
Vinelastic ¼ settling velocity of particles in inelastic fluid calculated using a correlation provided by Shah et al. (2007), m/s
Wi ¼ Weissenberg number calculated using experimentally measured velocities
WiIE ¼ Weissenberg number calculated using VIE
c ¼ shear rate, s–1
k ¼ relaxation time, seconds
la ¼ apparent shear viscosity, Pas
lf ¼ viscosity of a fluid, kg/ms
qf ¼ density of fluid, kg/m3
qs ¼ density of sphere, kg/m3
x ¼ angular frequency, rad/s
xi ¼ weight fraction of polymer grade i

Acknowledgments
This research is financially supported through the funds available from the Natural Sciences and Engineering Research Council of
Canada (NSERC RGPIN-2016-04647 KURU). The authors wish to thank SNF Floerger Company for providing the polymers used in
this study.

References
Acharya, A., Mashelkar, R. A., and Ulbrecht, J. 1976a. Flow of Inelastic and Viscoelastic Fluids Past a Sphere. Part 1: Drag Coefficient in Creeping and
Boundary-Layer Flows. Rheol. Acta 15 (9): 454–470. https://doi.org/10.1007/bf01530348.
Acharya, A., Mashelkar, R. A., and Ulbrecht, J. 1976b. Flow of Inelastic and Viscoelastic Fluids Past a Sphere. Part 2: Anomalous Separation in the
Viscoelastic Fluid Flow. Rheol. Acta 15 (9): 471–478. https://doi.org/10.1007/bf01530349.
Acharya, A. R. 1986. Particle Transport in Viscous and Viscoelastic Fracturing Fluids. SPE Prod Eng 1 (2): 104–110. SPE-13179-PA. https://doi.org/
10.2118/13179-PA.
Acharya, A. R. 1988. Viscoelasticity of Crosslinked Fracturing Fluids and Proppant Transport. SPE Prod Eng 3: 483–488. SPE-15937-PA. https://
doi.org/10.2118/15937-PA.
Arnipally, S. K. and Kuru, E. 2017. Effect of Elastic Properties of the Fluids on the Particle Settling Velocity. Proc., ASME 36th International Confer-
ence on Ocean, Offshore and Arctic Engineering, Trondheim, Norway, 25–30 June, Vol. 8. https://doi.org/10.1115/OMAE2017-61192.
Bui, B., Saasen, A., Maxey, J. et al. 2012. Viscoelastic Properties of Oil Based Drilling Fluids. Annual Trans. Nordic Rheol. Soc. 20: 33–47.
Castrejón-Garcı́a, R., Castrejón-Pita, J. R., Martin, G. D. et al. 2011. The Shadowgraph Imaging Technique and Its Modern Application to Fluid Jets and
Drops. Revista Mexicana de Fisica 57 (3): 266–275.
Chhabra, R. P. 2007. Bubbles, Drops, and Particles in Non-Newtonian Fluids, second edition. Boca Raton, Florida: CRC Press.
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. Non-Newton. Fluid Mech. 6 (3–4): 187–199. https://doi.org/10.1016/0377-0257(80)80002-4.
Chien, S.-F. 1972. Annular Velocity for Rotary Drilling Operations. Int. J. Rock Mech. Min. 9 (3): 403–416. https://doi.org/10.1016/0148-
9062(72)90005-8.
Chien S.-F. 1994. Settling Velocity of Irregularly Shaped Particles. SPE Drill & Compl 9 (4): 281–289. SPE-26121-PA. https://doi.org/10.2118/
26121-PA.
Choi, S. K. 2008. pH Sensitive Polymers for Novel Conformance Control and Polymer Flooding Applications. PhD dissertation, University of Texas,
Austin, Texas (August 2008).
Dehghanpour, H. and Kuru, E. 2011. Effect of Viscoelasticity on the Filtration Loss Characteristics of Aqueous Polymer Solutions. J. Pet. Sci. Eng. 76
(1–2): 12–20. https://doi.org/10.1016/j.petrol.2010.12.005.
Dehghanpour, H. A. H. 2008. Investigation of Viscoelastic Properties of Polymer Based Fluids as a Possible Mechanism of Internal Filter Cake Forma-
tion. Master’s thesis, University of Alberta, Edmonton, Canada.
Flemmer, R. L. C. and Banks, C. L. 1986. On the Drag Coefficient of a Sphere. Powder Tech. 48 (3): 217–221. https://doi.org/10.1016/0032-
5910(86)80044-4.
Foshee, W. C., Jennings, R. R., and West, T. J. 1976. Preparation and Testing of Partially Hydrolyzed Polyacrylamide Solutions. Presented at the SPE
Annual Fall Technical Conference and Exhibition, New Orleans, 3–6 October. SPE-6202-MS. https://doi.org/10.2118/6202-MS.
Gomaa, A. M., Gupta, D. V. S. V., and Carman, P. S. 2015. Proppant Transport? Viscosity Is Not All It’s Cracked up to Be. Presented at SPE Hydraulic
Fracturing Technology Conference, The Woodlands, Texas, 3–5 February. SPE-173323-MS. https://doi.org/10.2118/173323-MS.
Haider, A. and Levenspiel, O. 1989. Drag Coefficient and Terminal Velocity of Spherical and Nonspherical Particles. Powder Tech. 58 (1): 63–70.
https://doi.org/10.1016/0032-5910(89)80008-7.
Hu, Y. T., Chung, H.-S., and Maxey, J. E. 2015. What is More Important for Proppant Transport, Viscosity or Elasticity? Presented at SPE Hydraulic
Fracturing Technology Conference, The Woodlands, Texas, 3–5 February. SPE-173339-MS. https://doi.org/10.2118/173339-MS.
Kelessidis, V. C. and Mpandelis, G. 2004. Measurements and Prediction of Terminal Velocity of Solid Spheres Falling Through Stagnant Pseudoplastic
Liquids. Powder Tech. 147 (1–3): 117–125. https://doi.org/10.1016/j.powtec.2004.09.034.
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.
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 Tech. 57 (1): 39–50. https://doi.org/10.1016/0032-5910(89)80102-0.
Larson, R. G. 1998. The Structure and Rheology of Complex Fluids. New York City: Oxford University Press.
Malhotra, S. and Sharma, M. M. 2012. Settling of Spherical Particles in Unbounded and Confined Surfactant-Based Shear Thinning Viscoelastic Fluids:
An Experimental Study. Chem. Eng. Sci. 84 (24 December): 646–655. https://doi.org/10.1016/j.ces.2012.09.010.
McCabe, W. L., Smith, J. C., and Harriott, P. 2004. Unit Operations of Chemical Engineering, New York City: McGraw-Hill.

1704 October 2018 SPE Journal

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038


J187255 DOI: 10.2118/187255-PA Date: 6-October-18 Stage: Page: 1705 Total Pages: 17

Miura, H., Takahashi, T., Ichikawa, J. et al. 2001. Bed Expansion in Liquid–Solid Two-Phase Fluidized Beds With Newtonian and Non-Newtonian Flu-
ids Over the Wide Range of Reynolds Numbers. Powder Tech. 117 (3): 239–246. https://doi.org/10.1016/S0032-5910(00)00375-2.
Moore, P. L. 1974. Drilling Practices Manual. Tulsa: Petroleum Publishing Company.
Morrison, F. A. 2013. Data Correlation for Drag Coefficient for Sphere. Report, Department of Chemical Engineering, Michigan Technological Univer-
sity, Houghton, Michigan.
Renaud, M., Mauret, E., and Chhabra, R. P. 2004. Power-Law Fluid Flow Over a Sphere: Average Shear Rate and Drag Coefficient. Can. J. Chem. Eng.
82 (5): 1066–1070. https://doi.org/10.1002/cjce.5450820524.
Settles, G. S. 2001. Schlieren and Shadowgraph Techniques: Visualizing Phenomena in Transparent Media, 25–38; 143–164. Berlin: Springer-Verlag
GmbH.
Shah, S. N., El Fadili, Y., and Chhabra, R. P. 2007. New Model for Single Spherical Particle Settling Velocity in Power Law (Visco-Inelastic) Fluids.
Int. J. Multiphas. Flow 33 (1): 51–66. https://doi.org/10.1016/j.ijmultiphaseflow.2006.06.006.
Shahi, S. 2014. An Experimental Investigation of Settling Velocity of Spherical and Industrial Sand Particles in Newtonian and Non-Newtonian Fluids
Using Particle Image Shadowgraph. Master’s thesis, University of Alberta, Edmonton, Canada (June 2014).
Shahi, S. and Kuru, E. 2015. An Experimental Investigation of Settling Velocity of Natural Sands in Water Using Particle Image Shadowgraph. Powder
Tech. 281 (September): 184–192. https://doi.org/10.1016/j.powtec.2015.04.065.
Shahi, S. and Kuru, E. 2016. Experimental Investigation of the Settling Velocity of Spherical Particles in Power-Law Fluids Using Particle Image Shadow-
graph Technique. Int. J. Miner. Process. 153 (10 August): 60–65. https://doi.org/10.1016/j.minpro.2016.06.002.
Stokes, G. G. 1851. On the Effect of the Internal Friction of Fluids on the Motion of Pendulums. Cambridge, UK: Pitt Press.
Sunthar, P. 2010. Polymer Rheology. In Rheology of Complex Fluids, ed. A. P. Deshpande, J. M. Krishnan, and S. Kumar, Chap. 8, 171–191. New York
City: Springer.
Turton, R. and Levenspiel, O. 1986. A Short Note on the Drag Correlation for Spheres. Powder Tech. 47 (1): 83–86. https://doi.org/10.1016/0032-
5910(86)80012-2.
Urbissinova, T. S., Trivedi, J., and Kuru, E. 2010. Effect of Elasticity During Viscoelastic Polymer Flooding: A Possible Mechanism of Increasing the
Sweep Efficiency. J Can Pet Technol 49 (12): 49–56. SPE-133471-PA. https://doi.org/10.2118/133471-PA.
van den Brule, B. H. A. A. and Gheissary, G. 1993. Effects of Fluid Elasticity on the Static and Dynamic Settling of a Spherical Particle. J. Non-Newton.
Fluid 49 (1): 123–132. https://doi.org/10.1016/0377-0257(93)85026-7.
Veerabhadrappa, S. K., Doda, A., Trivedi, J. J. et al. 2013a. On the Effect of Polymer Elasticity on Secondary and Tertiary Oil Recovery. Ind. Eng.
Chem. Res. 52 (51): 18421–18428. https://doi.org/10.1021/ie4026456.
Veerabhadrappa, S. K., Trivedi, J. J., and Kuru, E. 2013b. Visual Confirmation of the Elasticity Dependence of Unstable Secondary Polymer Floods.
Ind. Eng. Chem. Res. 52 (18): 6234–6241. https://doi.org/10.1021/ie303241b.
Walters, K. and Tanner, R. 1992. The Motion of a Sphere Through an Elastic Fluid. In Transport Processes in Bubbles, Drops and Particles, ed. R. P.
Chhabra and D. De Kee, Chap. 3. New York City: Hemisphere.
Zang, Y. H., Muller, R., and Froelich, D. 1987. Influence of Molecular Weight Distribution on Viscoelastic Constants of Polymer Melts in the Terminal
Zone. New Blending Law and Comparison With Experimental Data. Polymer 28 (9): 1577–1582. https://doi.org/10.1016/0032-3861(87)90362-4.

Appendix A
The settling velocity of a spherical particle (Vinelastic ) with diameter of 2 mm in a shear-thinning fluid with zero relaxation time (k ¼
0 seconds) is calculated using the model developed for predicting the settling velocity of spheres in a viscoinelastic power-law-type
fluid (Shah et al. 2007) as follows, with K ¼ 0.26 Pasn, n ¼ 0.35, qf ¼ 998 kg/m3, qs ¼ 2510 kg/m3, and dp ¼ 0.002 m.
A ¼ 6:9148n2  24:838n þ 22:642 ¼ 14:7957;
Step 1 is:
B ¼ 0:5067n2 þ 1:3234n  0:1744 ¼ 0:2267:
" #1
1 13:082n dpnþ2 qnf ðqp  qf Þ2n 2
Step 2 is: ðC2n 2 2
D Re Þ ¼ 2ðn1Þ
¼ 47:7074.
2 K2
" 1
#1
ðC2n
D Re 2 2 B
Þ
Step 3 is: Re ¼ ¼ 175.
A
! 1
2n1 KRe 2  n
Step 4 is: Vs ¼ ¼ 0:4374 m=s.
dpn qf

Therefore, the terminal settling velocity of 2-mm spherical particles is 0.4374 m/s in a viscoinelastic fluid with the power-law parame-
ters K ¼ 0.26 Pasn and n ¼ 0.35.

Sumanth Kumar Arnipally is a research assistant at the University of Alberta, and worked previously as a research associate for
4 years at Hindustan Unilever Limited, Bangalore, India. His research interests include fluid/particle systems, rheology, and surface
interfacial phenomena. Arnipally has authored or coauthored more than five technical papers and holds eight product-
formulation-based patents. He holds a bachelor’s degree in chemical engineering from Andhra University, India, and a master’s
degree in petroleum engineering from the University of Alberta.
Ergun Kuru is a professor and Director of the School of Mining and Petroleum Engineering at the University of Alberta. His current
research areas of interest include the development of effective hole-cleaning strategies for oil- and gas-well applications,
design and development of nondamaging fluids for oil/gas-well drilling, completion and stimulation applications, and under-
standing and mitigating leakage pathways in oil- and gas-well cements. Kuru holds a bachelor’s degree from Middle East Tech-
nical University, Turkey, and master’s and PhD degrees from Louisiana State University, all in petroleum engineering. He has
served on several SPE committees in the past, including the SPE Annual Technical Conference and Exhibition Drilling Engineering
Program Committee, the SPE Global Training Committee, and the SPE Education and Accreditation Committee; he was an Edi-
torial Review Board member for SPE Drilling & Completion; and he served as an associate editor for Journal of Canadian Petro-
leum Technology. Kuru was the recipient of the 2017 SPE Canada Region Distinguished Achievement Award for Petroleum
Engineering Faculty.

October 2018 SPE Journal 1705

ID: jaganm Time: 11:06 I Path: S:/J###/Vol00000/180038/Comp/APPFile/SA-J###180038

You might also like