You are on page 1of 16

International Journal of Multiphase Flow 66 (2014) 46–61

Contents lists available at ScienceDirect

International Journal of Multiphase Flow


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / i j m u l fl o w

Solid velocity and concentration fluctuations in highly concentrated


liquid–solid (slurry) pipe flows
S.A. Hashemi ⇑, A. Sadighian, S.I.A. Shah, R.S. Sanders
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, AB T6G 2V4, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Solid velocity and concentration fluctuations were measured for concentrated sand-water mixtures
Received 13 January 2013 (20–35% solid by volume) in horizontal pipe flow using Electrical Impedance Tomography (EIT). Narrowly
Received in revised form 11 May 2014 sized sand (d50 = 100 lm) was used to prepare each slurry tested in a 52 mm (i.d.) horizontal pipe loop at
Accepted 26 June 2014
mixture velocities (2–5 m/s) that were significantly above the deposition velocity. The EIT measurements
Available online 10 July 2014
were used to obtain solid velocity and concentration fluctuation maps. Results show that the magnitude
of the local solid concentration fluctuations is greater near the pipe wall and increases as the mixture
Keywords:
velocity increases. Additionally, the concentration fluctuations are greater near the pipe invert, particu-
Solid–liquid flow
Electrical impedance tomography
larly at lower mixture velocities and/or concentrations where the solid concentration profiles are asym-
Concentration fluctuations metric. The Fast Fourier Transform (FFT) method was employed to study the power spectral density of
these fluctuations. This analysis indicates that concentration fluctuations are produced almost entirely
by particle–fluid turbulence interactions, rather than through particle–particle or particle–wall interac-
tions. Comparison of the particle diameter with the characteristic turbulent length scales shows that
the particles interact with turbulent eddies in the dissipative range, which is in accordance with the
power spectral density analysis. The findings presented here are consistent with previous studies of
fluidized beds and gravity-driven flows.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction commonly found in commercial CFD (Computational Fluid Dynam-


ics) packages. However, these models have not been widely com-
Pipeline flows of coarse-particle slurries are of great importance pared with experimental results, especially for liquid–solid flows.
in many industries, including hard-rock mining, oil sands produc- The Zenit et al. (1997) experimental study of collisional pressure
tion, and nuclear waste treatment. Industrial slurries have different in liquid–solid fluidized bed showed the unsatisfactory perfor-
particle sizes and hence the relative positions and velocities of mance of current models, which is mainly a consequence of the
these particles play an important role in pipeline design and oper- lack of understanding of the physics and mechanisms that govern
ation. An improved understanding of the complex behaviour of the behaviour of these complex systems.
these flows and the physics behind them will help to improve One of the barriers to improved understanding and model
models needed to predict, for example, frictional pressure losses, development is the difficulty associated with making the appropri-
optimal operating velocities, and pipeline wear. ate measurements (Zenit and Hunt, 1998), particularly for highly
Highly concentrated two-phase flows are generally unsteady concentrated coarse-particle slurries where the axial velocities
and previous experiments have shown that velocities and concen- are high and mixtures are opaque.
trations of both phases undergo fluctuations about their mean val- One of the most thorough studies of solid fluctuations in con-
ues (Zenit and Hunt, 2000). Parameters such as collisional particle centrated liquid–solid flow is the experimental investigation of
pressure or granular temperature are used to describe the solid Zenit and Hunt (2000). They measured the cross-sectional aver-
fluctuations. Presently, different models describing particle aged solid concentration fluctuations in a fluidized bed and for
interactions, e.g. kinetic theory (Jenkins and Savage, 1983), are gravity driven flow, for large particles with different diameters
and densities. They compared their results with the Buyevich and
Kapbasov (1994) model. In this model, the solid concentration fluc-
⇑ Corresponding author.
tuation is only a function of solid concentration. Zenit and Hunt
E-mail addresses: reza.hashemi@ualberta.ca (S.A. Hashemi), sadighia@ualberta.
ca (A. Sadighian), sshah@ualberta.ca (S.I.A. Shah), ssanders@ualberta.ca (2000) found that the averaged solid concentration fluctuations
(R.S. Sanders). are a function of both solid concentration and Stokes number with

http://dx.doi.org/10.1016/j.ijmultiphaseflow.2014.06.007
0301-9322/Ó 2014 Elsevier Ltd. All rights reserved.
S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 47

the magnitude of the fluctuations increasing with increasing for slurry flow applications (Dyakowski et al., 2000; Graham et al.,
Stokes number. This was in accord with their previous findings; 2002; Pachowko et al., 2003). Numerous studies have been con-
that is, high frequency fluctuations are mainly due to direct ducted where electrical tomography techniques are used on
collisions and the power of an immersed collision increased as fluid-particle systems, such as pneumatic conveying of granular
the Stokes number increased (Zenit and Hunt, 1999). They con- solids (Zhu et al., 2003; Azzopardi et al., 2008), flow distribution
cluded that the Buyevich and Kapbasov (1994) model relates to and velocity measurement in a fixed bed reactor (Bolton et al.,
the condition where Stokes number tends to infinity. 2004), hydraulic conveying of materials (Fangary et al., 1998),
Picciotto et al. (2005) studied the interactions between a dis- and fluidization (Azzi et al., 2010). However, only a very limited
persed solid and fluid turbulence in a turbulent channel flow. They number of studies directly related to electrical tomography and
noted that interaction between particles and coherent turbulent slurry pipe flow measurements have been published (Dyakowski
structures near the wall results in streamwise velocity fluctuations et al., 2000; Norman and Bonnecaze, 2005; Xu et al., 2009).
in solid velocity. Their results show that solid streamwise velocity Applications of electrical tomography to slurry pipeline flow
fluctuations are higher near the wall and decrease with increasing measurements have thus far been restricted to time- and spatial-
distance from the wall. They also show that the magnitude of these averaged values of concentration.
fluctuations increases with increasing particle Stokes number. The goal of this paper is to investigate the mechanisms that are
Kechroud et al. (2010) studied the dynamic behaviour of the responsible for the production of high frequency-low amplitude
continuous phase in a solid–liquid fluidized bed. They compared solid concentration and velocity fluctuations in highly concen-
their measurements of liquid velocity fluctuations with solid con- trated slurry flows. Novel high-frequency Electrical Impedance
centration fluctuations from Didwania and Homsy (1981) and Tomography (EIT) measurements for highly concentrated pipeline
Zenit and Hunt (2000). They found a high degree of similarity flows of solid–liquid mixtures are used to evaluate the validity and
between liquid velocity fluctuations and solid concentration limitations of the analysis. Specifically, solid turbulent intensity
fluctuations. and concentration fluctuation and time-averaged solid concentra-
Varaksin and Polyakov (2000) studied particle velocity fluctua- tion profiles for slurry flow in a horizontal pipe were measured
tions in an air- solid turbulent pipe flow. They classified the mech- using Electrical Impedance Tomography. The results were also
anisms for solid velocity fluctuations into four main categories: (1) compared to the models available in the literature to evaluate their
solid–fluid turbulence interaction, (2) presence of particles with capabilities and limitations. Measurements of this type are needed
different sizes, i.e. not truly monosized particles, (3) particle–parti- to develop and/or validate numerical simulations of slurry flows.
cle and particle–wall collisions and (4) migration of particles to
regions with different velocities (streaming mechanism). These
four phenomena could also be considered as the main mechanisms 2. Experimental detail and analysis
producing solid concentration fluctuations.
One of the difficulties associated with experiments involving 2.1. Experiments
concentrated multiphase flows is the lack of viable measurement
techniques. Common single phase and dilute multiphase flow mea- A 52 mm (i.d.) horizontal pipeline loop located at the Saskatch-
suring techniques, such as LDV (Laser Doppler Velocimetry), PIV ewan Research Council (SRC) Pipe Flow Technology Centre, SK,
(Particle Image Velocimetry), and PTV (Particle Tracking Velocime- Canada, was used to perform the experiments. The schematic lay-
try) are widely used for transparent flows where the solid concen- out of the loop is shown in Fig. 1. The loop includes a centrifugal
tration is low. However, their capabilities in highly concentrated pump to circulate the slurry at different velocities. Operating volu-
and opaque flows, such as dense slurry flows, are debatable metric flow rates were measured using a Foxboro 2802-SABA-TS
(Graham et al., 2002). magnetic flow meter. Two heat exchangers were used to keep
Advances in measurement techniques in recent years, the operating temperature constant during experiments. Pressure
especially electrical tomography methods, have opened a new drop along a test section was measured using a Valydyne DP15 dif-
window in the experimental study of multiphase flows. The most ferential pressure transducer. The transparent observation section
important advantage of these methods is their ability to perform was used to ensure that air, which sometimes enters loop during
measurements in concentrated and opaque systems. It also the preparation of the slurries, was completely removed before
appears that these methods are sufficiently fast and robust enough measurements were taken. The slurries were prepared using

Fig. 1. Schematic of 52 mm horizontal pipe-loop.


48 S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61

The frequency of injecting AC current was set at 80 kHz. Data were


collected at a frequency of 820 frames per second (fps) per plane
(Hashemi, 2013). The instrument measures the resistivity distribu-
tion map within the sensor plane. The resistivity map is then con-
verted to solid concentration using the Maxwell equation
(Dyakowski et al., 2000):

2r1 þ r2  2rm  rmr1r2


Cs ¼ ð1Þ
rm  rr21 rm þ 2ðr1  r2 Þ
where r1 ; r1 and rm are conductivity of continuous phase, conduc-
tivity of dispersed phase and reconstructed measured conductivity,
respectively and C s is the dispersed phase volume concentration.
Fig. 3 shows the EIT reconstruction grid. The map divides the
pipe cross section into 316 pixels of equal area. The solid concen-
tration within each pixel is measured. Pixel to pixel cross-correla-
tion of concentration maps between two planes was used to obtain
Fig. 2. Particle size distribution of sand particles used for experiments. the solid velocity distribution map (Datta et al., 2007).
For each set of experiments, 8000 conductivity maps were col-
municipal tap water as the carrier fluid and sand (Lane Mountain lected from each sensor plane for every test, i.e. at different mix-
LM125) with d50 = 100 lm as dispersed phase. The particle size ture flow rates and sand concentrations. The equipment was
distribution of the sand is shown in Fig. 2. Experiments were per- calibrated using tap water before performing experiments. The
formed at solid volumetric concentrations of 20–35% and mixture resulting concentration maps were used to calculate solid phase
velocities of 2–5 m/s. concentration fluctuations, velocities and turbulent intensities.
An Industrial Tomography Systems (ITS) Z8000 Electrical Before any slurry tests were conducted, EIT measurements were
Impedance Tomography (EIT) data acquisition system along with collected for single phase (water only) runs. Fig. 4 shows conduc-
a dual-plane sensor was employed to measure solid velocity and tivity measurements and a reconstructed concentration map for
concentration distributions. Each sensor plane consists of 16 elec- single phase flow. The conductivity values given have units of
trodes which are arranged so that they are uniformly spaced mS/cm. The solid phase concentration fluctuations of Fig. 4(b)
around the circumference of the pipe. Each electrode is identical should of course be zero in single phase flow. Any nonzero values
in construction, and is 10 mm in width and 10 mm in length. More represent measurement noise. Comparison of actual solid-phase
detailed information on the construction of the test spool and the concentration fluctuation measurements with the baseline mea-
electrodes themselves can be found elsewhere (Hashemi, 2013). surements (noise) made for water flow illustrates the ability of
The injection current range was up to 50 mA and the measured the instrument to provide meaningful data. Fig. 5 shows concentra-
voltage between each electrode pair was between 10 to 10 V. tion fluctuation profiles measured for flow of sand slurry with

Fig. 3. Schematic diagram of EIT reconstruction grid.


S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 49

Fig. 4. Single phase flow measurements at 2 m/s: (a) measured conductivity map; (b) reconstructed concentration fluctuation map, i.e. signal noise.

Fig. 5. Comparison between raw concentration fluctuations (including noise) for 100 lm sand flowing in a 52 mm pipeline loop at C s = 25% and measurement noise for single
phase water flow: (a) V = 2 m/s; (b) V = 5 m/s.

C s = 25% at mixture velocities of 2 and 5 m/s. The measurement


noise for single phase (water) flow at the same velocities is also Table 1
shown in these figures. Note that y is the distance from the bottom Calculated deposition velocities for 100 lm sand-in-water mixtures in a 52 mm pipe.
of the pipe, and that this terminology is used for all figures in this
Solid concentration, C s (%) Deposition velocity (m/s)
paper where measurements made at numerous positions over
20 0.8
the flow domain are reported. The results clearly show the effect
25 0.9
of the dispersed solid phase on both the magnitude and the 30 1.0
shape of the concentration fluctuation profiles. To obtain more 35 1.1
accurate concentration fluctuation values, the noise effects were
removed from the concentration fluctuation signals by decomposi-
tion of signal variance. The same procedure was followed for all corresponds to the transit time between planes 1 and 2. Since the
other velocities and concentrations reported here. separation distance between the two planes, Ls , is known, the tran-
sit velocity can be obtained using:
2.2. Analysis
Ls
If Sp1 and Sp2 are signals obtained from planes 1 and 2, respec- u¼ ð3Þ
smax
tively, at time t, the cross correlation function is defined as:
Z T where smax is the time delay at the maximum value of the cross-cor-
1
RðsÞ ¼ lim Sp1 ðtÞSp2 ðt þ sÞdt ð2Þ relation coefficient, Rmax .
T!1 T 0
Pixel-to-pixel cross correlation can be used to determine the
where s is the time delay between Sp1 and Sp2 and T is the observa- local solid velocity between corresponding pixels in the two
tion time. The time delay at the maximum value of R value planes. If we assume that C p1m and C p2m are concentration signals
50 S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61

Fig. 6. Solid concentration fluctuation before (a, c, e, g) and after (b, d, f, h) applying digital filter for 100 lm sand flowing in a 52 mm pipeline loop at C s = 20%: (a) and (b)
V = 2 m/s; (c) and (d) V = 3 m/s; (e) and (f) V = 4 m/s; (g) and (h) V = 5 m/s.
S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 51

obtained for pixel m, in planes 1 and 2, the cross-correlation fluctuation signals before and after the digital filter was applied.
coefficient is obtained by: The magnitude of RMS solid concentration fluctuations decreased
slightly in all cases after filtering due to the removal of higher
X
N1
R¼ C p1m ½kC p2m ½k þ s ð4Þ amplitude, lower frequency phenomena. Longer measurement
k¼0 times are required to accurately capture the low frequency fluctu-
ations that are mainly due to the bulk motion of the flow (Zenit and
where s is the time delay, k is the image number and N is the num- Hunt, 2000).
ber of images. The time delay at maximum R; smax , can be used to The power spectral density of the solid concentration and veloc-
obtain the local solid velocity in each pixel. ity fluctuation signals was also obtained so that energy distribution
Instantaneous solid velocities can be obtained by using shorter at different frequencies could be studied. The Welch method,
time windows i.e. shorter T values. The windows were used to which is a technique for estimating averaged power spectra with
select shorter time sequences and the cross-correlation between data windowing (Oppenheim et al., 1999), was used to estimate
these sections used to determine the solid-phase instantaneous power spectrums. In the Welch method, the original signal is
velocity in each pixel. The time resolution of instantaneous veloc- divided into a number of sections each of length L and the averaged
ities was set to be 0.01s. The instantaneous velocities were then power spectrum is calculated using:
used to calculate RMS solid velocity fluctuations and solid turbu-
lent intensities. The RMS velocity concentration fluctuation was 1XK 1

calculated from: Ið f Þ ¼ Ir ð f Þ ð10Þ


K r¼0
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 XN 2
where K is the number of sections and Ir ðf Þ is the power spectrum
V s ¼ ðV s  V si Þ ð5Þ
N i¼1
estimation for each section. In the current study, the original signal
where N; V  s ; V s and V si are the number of measurements, RMS solid is segmented into eight sections ðK ¼ 8Þ with 50% overlap, to min-
velocity fluctuations, average solid velocity and instantaneous solid imise the spectral variance and obtain the optimum results. The
velocity, respectively. The time-averaged solid velocity was calcu- segmented signals along with Hamming windows of the same
lated by cross correlation of the entire signal using Eq. (3). Solid tur- length as the segment (Press et al., 1988) were used to estimate
bulent intensity was also calculated by dividing the RMS solid the average power spectrums.
velocity fluctuations by the time-averaged solid velocity. The RMS
solid concentration fluctuation was also computed from: 3. Results and discussion
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
XN
s ¼ 1
C ðC s  C si Þ
2
ð6Þ 3.1. Solid velocity and turbulent intensity
N i¼1

 s ; C s and C si are the number of measurements, RMS solid


where N; C The main mechanisms producing velocity fluctuations in the
concentration fluctuations, average solid concentration and instan- centre of the flow are particle–particle interactions and solid–fluid
taneous solid concentration, respectively. The average solid concen- turbulent interactions (Varaksin and Polyakov, 2000). Here, we
tration was also calculated using: analyse the turbulent intensity profiles and power spectra of solid
velocity fluctuations to determine the relative importance of the
1X N
two mechanisms.
Cs ¼ C si ð7Þ
N i¼1 Solid axial turbulent intensity profiles are provided in Fig. 7. The
experimental results show that the solid turbulent intensities are
The confidence interval for measurements was estimated as higher near the wall, i.e. wall peaked. A similar trend was observed
(Walpole et al., 2007): by Kulick et al. (1994) for 50 and 90 lm glass beads and 70 lm
C:I: ¼ 1:96  SE ð8Þ copper particles in a wind tunnel. Alajbegovic et al. (1994)
observed wall-peaked solid axial turbulent intensities for the
where SE is the Standard Error and is defined as: upward flow of ceramic particles in a vertical pipe. The main
reason for the higher solid turbulence intensities near the wall is
S
SE ¼ pffiffiffiffi ð9Þ the particle–wall interactions that occur in that region. Another
N important mechanism is the off-the-wall migration of particles to
Here, S and N are standard deviation and number of regions with higher velocities (Wilson and Sellgren, 2003; Wilson
measurements, respectively. et al., 2010). The results presented in Fig. 7 also illustrate that
All measurements were made at operating velocities that the magnitude of the solid turbulent intensities are roughly
exceeded the predicted deposition velocity for each condition. uniform at the pipe centre, particularly at mixture velocities of
Table 1 shows predicted values of the deposition velocity for 3 m/s or greater. Fig. 8 shows sample turbulent intensity maps
100 lm sand at different concentrations (Shook et al., 2002). The obtained from tomography measurements.
highest deposition velocity, V c = 1.1 m/s, is obtained at C s = 35%. The presence of the dispersed solid phase affects the fluid tur-
The minimum velocity at which experiments were performed bulent structure (Truesdell and Elghobashi, 1994; Lightstone and
was 2 m/s to ensure that the formation of a stationary deposit Hodgson, 2004) and is generally referred to as turbulence modula-
was avoided. Additionally, the averaged solid concentration and tion. The study of turbulence modulation in both experimental and
frictional pressure gradient were monitored to ensure that they theoretical frameworks is presently an active area of research. The
were constant during the period of time that EIT data were ideal situation is to study the fluid-particle interactions by
collected. measuring solid and fluid velocity fluctuations simultaneously.
As the main focus of this work is to study the low amplitude- Unfortunately, the measurement techniques for measuring fluid
high frequency fluctuations, the high amplitude-low frequency velocity fluctuations are only applicable to very low solid concen-
fluctuations were filtered using a 4th order, high pass Butterworth trations and are not suitable for concentrated mixtures of the type
digital filter algorithm. A cut-off frequency of 3 Hz was chosen by studied here. The studies and models of turbulence modulation
examining the frequency domain signals, which is the same as that found in the literature are limited to very low solid concentrations
chosen by Zenit and Hunt (2000). Fig. 6 shows the concentration and are far from being applicable to concentrated systems such as
52 S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61

Fig. 7. Solid axial turbulent intensity profiles for 100 lm sand flowing in a 52 mm pipeline loop at different solid concentrations: (a) V = 2 m/s; (b) V = 3 m/s; (c) V = 4 m/s;
(d) V = 5 m/s.

the system of interest in this study. Another approach that has  Models describing the effect of the dispersed solid phase on
been used in the past to account for the presence of the solid fluid turbulence have been developed (and validated) for very
particles is to use a mixture viscosity model, such as that proposed dilute flows only and thus are unlikely to accurately describe
by Thomas (1965), to account for the effect of fluid-particle inter- the modulating effects that occur at high solid concentrations.
actions. Although the application of mixture viscosity models
might be reasonable for homogenous mixtures of spherical As a consequence, it seemed logical to ignore the effect that the
particles at low solid concentrations, their ability to accurately particles may have on the fluid turbulence and compare instead
describe fluid-particle interactions or any other flow characteris- the well-known single phase fluid turbulence characteristics with
tics deteriorates badly as the solid concentration and particle size our solid-phase fluctuating measurements. The result is a clearly
and/or angularity increase (Schaan, 2001). Hence, mixture viscos- qualitative analysis, but one that is unfettered by questions that
ity models are not applicable to the slurries of interest here, where would naturally arise if, for example, turbulence modulation mod-
solid concentrations are high and particles are angular. els meant for dilute mixtures were applied here. Additionally,
In the present study, we choose to compare qualitatively the when the comparisons with single phase flow are taken in concert
dispersed solid phase velocity fluctuation measurements with sin- with the results obtained by other researchers (e.g. Zenit and Hunt,
gle phase flow fluid turbulence characteristics. The motivation for 2000), a clear picture of the relative importance of fluid-particle
this choice can be stated as: interactions (over particle–particle collisions) in the production
of solid phase fluctuations emerges.
 Relatively simple approaches, such as mixture viscosity models, The energy spectrum of single phase turbulent motion can be
do not properly or accurately describe energy dissipation mech- divided into 3 main ranges: ð1Þ energy containing range, ð2Þ
anisms for concentrated slurry flows of the type studied here. inertial sub-range and ð3Þ dissipation range. Fig. 9 shows these
 Models describing single phase fluid turbulence characteristics three ranges and the relevant length scale associated with each
are well known and accepted. (Pope, 2008).
 Measurements of time-averaged solid concentration profiles in The energy containing range includes large scale turbulent
dense slurry pipeline flows were accurately predicted using a eddies whose motion is roughly independent of viscosity. These
solid-phase turbulent diffusion coefficient equivalent to the large-scale structures contain the bulk of the turbulent energy. In
eddy kinematic viscosity calculated for turbulent, single phase the Inertial sub-range, the turbulent energy transfer from larger
flow (Gillies and Shook, 1994). length scales to smaller length scales occurs through a cascading
S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 53

Fig. 8. Solid turbulent intensity maps for 100 lm sand flowing in a 52 mm pipeline loop at different solid concentrations and mixture velocities: (a) V = 2 m/s and C s = 20%;
(b) V = 3 m/s and C s = 35%; (c) V = 4 m/s and C s = 30%; (d) V = 5 m/s and C s = 25%.

(water) with the same velocity as the mixture were calculated.


Table 2 shows the calculated Kolmogorov length scales and lDI for
single phase flow of water in a 52 mm pipe. The rate of energy dis-
sipation was estimated using (Hughmark, 1977):

4u2 V
e¼ ð12Þ
D
where u ; V and D are the shear velocity, mixture velocity and pipe
Fig. 9. Turbulent length scales and the turbulent energy spectrum (Pope, 2008).
diameter, respectively. Shear velocity is defined as:
rffiffiffiffiffi
fw
process. The turbulent energy dissipates in the dissipation range u ¼ V ð13Þ
8
due to the viscosity of the fluid. The dissipation range includes
lengths larger than the Kolmogorov length scale and smaller than where fw is the Moody friction factor at the mixture velocity. Any
lDI . The Kolmogorov length scale ðgÞ is defined as: turbulent length scale that falls between g and lDI is in the
dissipation range.
 3 14
m
g¼ ð11Þ Table 2
e Kolmogorov length scale for the flow of water in a 52 mm pipe at different mixture
velocities.
where m and e, respectively, are the kinematic viscosity of the fluid
and the rate of energy dissipation. The Kolmogorov length scale is Velocity (m/s) g (lm) lDI (lm)

the smallest turbulence length scale. For many turbulent flows, 2 28 1690
lDI  60g (Pope, 2008). 3 21 1270
4 17 1040
To analyse the power spectrum of solid velocity fluctuations,
5 15 890
turbulence length scales for single phase flow of the carrier fluid
54 S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61

A model for energy-spectrum function is given as (Pope, 2008): Table 3


Calculated average power spectrum slope for 100 lm sand-in-water mixtures in a
2=3 5=3
EðjÞ ¼ C e j fL ðjLÞfg ðjgÞ ð14Þ 52 mm pipe at C s = 30%.

Mixture velocity (m/s) Power spectrum slope


where j is the wave number and defined as:
2 2.16
2p 3 2.37
j¼ ð15Þ
l 4 1.90
5 1.80
Here, l is the characteristic length scale. In Eq. (14), fL and fg are
non-dimensional functions which determine the shape of spec-
trum in the energy-containing and dissipation ranges, respectively. velocities at a solid volume concentration of 30%. Although the tur-
The dissipation function fg is usually expressed as an exponential bulence power spectra are a function of wave number (spatial
function (Pope, 2008): variable), the frequency spectra are usually measured during
fg ðjgÞ ¼ expðb0 jgÞ ð16Þ experiments. Taylor’s hypothesis or the frozen turbulence approx-
imation (Pope, 2008) allows one to approximate wave numbers by
where b0 = 2.094. An alternative model for fg is the Pao spectrum frequencies. The rate of energy decay in the dissipation range is not
  constant and increases with decreasing turbulent length scale. The
3 4
fg ðjgÞ ¼ exp  CðjgÞ3 ð17Þ averaged slopes of the experimental power spectra of Fig. 10 are
2
shown in Table 3. The calculation of a single value to represent
where C is constant and is equal to 1.5 (Pope, 2008). the slope of fluid turbulent energy decay in the dissipation range
Based on the calculated values of g and lDI shown in Table 2, it for comparison with the experimental values reported here is not
can be concluded that the particles used in this study practical, simply because the particles are not interacting with a
ðd ¼ 100 lmÞ interact with turbulent eddies in the energy dissi- specific turbulent length scale; rather a range of length scales are
pation range, since particles typically interact with turbulent involved. Since the rate of energy decay in this region is sensitive
eddies whose length scales are of the same order of magnitude to turbulent length scale, increasing with decreasing l, it is not pos-
as the particle diameter. sible to calculate a single slope which represents the turbulent
Fig. 10 shows the power spectrum of solid velocity fluctuations energy decay rate. The only valid statement is that the rate of
(solid lines) and turbulent energy spectrum decay rate for single turbulent energy decay should be higher than the value for the
phase flow from Eq. (14) (dotted lines) for different mixture sub-range region. Table 3 shows that the experimental slope values

Fig. 10. Power spectrum for cross sectional averaged solid velocity fluctuations of 100 lm sand flowing in a 52 mm pipeline loop at C s = 30%: (a) V = 2 m/s; (b) V = 3 m/s;
(c) V = 4 m/s; (d) V = 5 m/s. Dotted lines are the model spectrum (Eq. (14)) decay rates.
S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 55

Fig. 11. Axial velocity profiles for 100 lm sand flowing in a 52 mm pipeline loop at different mixture velocities: (a) C s = 20%; (b) C s = 25%; (c) C s = 30%; (d) C s = 35%.

are greater than 5=3 which is the energy spectrum slope for the solid phase is mainly affected by fluid turbulence, the solid velocity
inertial sub-range and in accordance with the dissipation range defect should be comparable to that of the fluid.
where the slope is greater than for the inertial sub-range. The cal- The velocity defect ðumax  uÞ for a single phase turbulent flow is
culated slopes decrease with increasing mixture velocity which is (Longwell, 1966):
also in accordance with the single phase turbulent energy decay  
ðumax  uÞ ymax
law. The results indicate that solid-turbulent interactions are pri- ¼ 2:71 ln ð18Þ
u y
marily responsible for the production of solid velocity fluctuations,
particularly in the central core of the flow. Similar trends were
observed for other mixture velocities and solid concentrations.
Experimental values of fluid velocity fluctuations in highly concen-
trated mixtures are required for further validation as the effect of
solid particles on fluid turbulent in neglected in this analysis.
Fig. 11 shows solid velocity profiles at different mixture veloc-
ities and average in situ solid concentrations. The velocity values
plotted here are chord-averaged values. The velocity profiles are
nearly axisymmetric for the flow conditions tested here. In
Fig. 12, we compare velocity profiles obtained here using cross-cor-
related EIT measurements with those obtained with a resistivity
probe for nearly identical slurries (100 lm Lane Mountain sand
in water, 30% solid by volume) at SRC in the same test loop
(Unpublished results from SRC). The velocity profiles are in good
agreement.
Time averaged solid velocity profiles in horizontal slurry flows
are usually asymmetric. The location at which the maximum solid
velocity occurs varies depending on particle properties and flow
conditions. Shook et al. (2002) point out that the velocity defect
ðumax  uÞ can be used to evaluate the solid velocity profiles and Fig. 12. Comparison of velocity profiles measured in the present study with
the effect of fluid turbulence on solid particles. When the dispersed unpublished SRC results for 100 lm sand at C s = 30%.
56 S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61

Fig. 13. Velocity defect for solid velocity profiles of 100 lm sand flowing in a 52 mm pipeline loop at different concentrations: (a) V = 2 m/s; (b) V = 5 m/s.

Fig. 14. Concentration fluctuation profiles for 100 lm sand flowing in a 52 mm pipeline loop at different mixture velocities: (a) C s = 20%; (b) C s = 25%; (c) C s = 30%; (d)
C s = 35%.

where umax is the maximum velocity and ymax is the position at


Table 4 which the maximum velocity occurs. The dimensionless velocity
þ
d for sand particles at different mixture velocities.
defect ðumax
u
uÞ
for y < ymax is calculated using the solid velocity pro-
Mixture velocity (m/s) Shear velocity (m/s) þ
d files obtained here (see Fig. 11) and the result is compared to the
2 0.09 8.5 known velocity defect distribution for single phase turbulent flow.
3 0.13 13.3 Fig. 13 shows the comparison of the solid velocity defect with that
4 0.17 17.2
expected for single phase turbulent flow. The results show that in
5 0.20 21.1
the core of the flow, the solid velocity defect is in relatively good
S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 57

Fig. 15. Concentration profiles for 100 lm sand flowing in a 52 mm pipeline loop at different concentrations: (a) V = 2 m/s; (b) V = 3 m/s.

Fig. 16. Sample solid concentration fluctuations maps for 100 lm sand flowing in a 52 mm pipeline loop: (a) V = 5 m/s and C s = 25%; (b) V = 5 m/s and C s = 35%.

agreement with the single phase turbulent flow velocity defect. As Sellgren, 2003). Wilson and Sellgren (2003) showed that at certain
y ! 0, the solid velocity defect starts to deviate from that calculated conditions, near-wall lift plays an important role in promoting par-
for single phase turbulent flow. This is probably due to particle–wall ticle suspension. The near-wall lift force is important when the
interactions (Shook et al., 2002), but a more detailed study is particles are partially enclosed in the boundary layer. Wilson
þ
required to address this issue more precisely. Fig. 13 illustrates that et al. (2010) showed that near-wall lift is important when d is
þ
the dispersed solid phase is primarily affected by fluid turbulence, between 9 and 27 where d is defined as:
particularly at the pipe axis.
þ qf dp u
3.2. Solid concentration fluctuations d ¼ ð19Þ
lf
Fig. 14 shows solid concentration fluctuation profiles obtained Here dp and lf are particle diameter and fluid viscosity, respec-
þ
at different mixture concentrations and velocities. At lower bulk tively. At d < 9, particles are fully enclosed in the boundary layer
þ
concentration and velocities, concentration fluctuations in the and at d > 27 particles are fully unenclosed and they are not
lower section of the pipe are greater in magnitude than those mea- experiencing near-wall lift force (Wilson et al., 2010). Values of
þ
sured in the upper section. This is due to asymmetrical concentra- d for the 100 lm sand used in the present study, at different
tion profiles at lower mixture velocities and concentrations. These velocities, are shown in Table 4. The results show that at 2 m/s,
asymmetrical fluctuation profiles become more symmetrical as these particles are fully enclosed in the boundary layer and parti-
solid concentration and mixture velocity increase. At higher mix- cles are not experiencing near-wall lift force. At higher velocities,
þ
ture velocities, turbulent dispersion forces are much stronger and d > 9 and particles are experiencing lift force. Fig. 15 shows con-
promote particle suspension, which produces a more uniform solid centration profiles for sand particles at 2 and 3 m/s at different
concentration distribution. The more uniform concentration pro- concentrations. It is evident that the concentration profiles are
files will result in more uniform concentration fluctuation profiles. much more uniform at 3 m/s than at 2 m/s.
More uniform concentration profiles are also partially attributable The energy level of a high energy particle will decrease when it
to Saffman (Saffman, 1965) and near-wall lift forces (Wilson and collides with a particle of lower energy level due to energy transfer
58 S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61

Fig. 17. Time and frequency domain plots of cross sectional averaged solid concentration fluctuations for 100 lm sand flowing in a 52 mm pipeline loop at C s = 20%: (a) and
(b) V = 2 m/s; (c) and (d) V = 3 m/s; (e) and (f) V = 4 m/s; (g) and (h) V = 5 m/s. Dotted lines are the model spectrum (Eq. (14)) decay rates.
S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 59

during collision. An increase in the solid concentration will shows nearly a flat portion at high frequencies (see, for example,
increase the frequency of particle collisions and consequently will Fig. 15(b) and (d)). This is a result of the low signal-to-noise ratio
increase the rate of momentum exchange between particles. This inherent in the measurements at low mixture velocities. As the
will result in an equalisation of the energy distribution among mixture velocity increases, the Kolmogrov length scale decreases
the particles and should therefore produce more uniform solid and extends the turbulent dissipation range to smaller length
concentration fluctuation profiles. scales (higher frequencies). This produces stronger signals at high
Fig. 16 shows typical solid concentration fluctuation maps. In all frequencies and increases the signal-to-noise ratio.
cases, solid concentration fluctuations in the near-wall region were Zenit and Hunt (2000) studied solid concentration fluctuations
greater than those measured in the core of the flow, mainly in fluidized beds and gravity driven flows. Based on their experi-
because of solid-wall collisions and the high shear zone near the mental results, they concluded that the most important parame-
wall. ters in determination of concentration fluctuations are solid
Cross-sectional averaged concentration fluctuations and the concentration and Stokes number. They suggested that solid con-
resulting power spectra are illustrated in Fig. 17. The dotted line centration fluctuations increased with increasing Stokes number
of Fig. 17 represents the fluid turbulence energy decay model spec- at a constant solid concentration. They calculated Stokes number
trum in the inertial subrange. The results show that the energy of based on the particle terminal settling velocity, defining it as
concentration fluctuations of 100 lm sand decays roughly with the (Zenit and Hunt, 2000):
same slope of solid velocity fluctuations energy, which suggests  
that solid–fluid turbulence interaction is an important mechanism
qs Ret
Stt ¼ ð20Þ
for the production of concentration fluctuations. The energy qf 9
spectrum for concentration fluctuations at lower mixture velocities
where Ret is the particle Reynolds number based on the terminal
settling velocity. In multiphase flows, mean quantities such as con-
Table 5
Stokes number for sand particles at different mixture velocities. centration fluctuations, granular temperature and pressure gradient
could be expressed as a function of mean concentration and mix-
Mixture velocity (m/s) g (lm) St s
ture velocity. However, in a fluidized bed, these mean values could
2 28.2 1.8 be expressed as a function of either mean concentration or mean
3 21.2 3.3
velocity as these two parameters are related and are not indepen-
4 17.3 4.9
5 14.8 6.7
dent values, i.e. any given mean concentration corresponds to a spe-
cific fluidization velocity. In slurry pipe flow, the mean
concentration and level of turbulence are independent variables
and both affect any mean quantity. As a consequence, the Stokes
(a) number based on terminal settling velocity is not applicable to
slurry pipe flows. In order to investigate the applicability of Zenit
and Hunt’s (Zenit and Hunt, 2000) finding to slurry pipe flows, a
new definition for Stokes number for slurry pipe flows was
employed. The particle Stokes number is defined as (Collins and
Keswani, 2004):
 2
1 qs dp
Sts ¼ ð21Þ
18 qf g

Here, the Kolmogrov length scale is used to account for the level
of fluid turbulence. Table 5 shows the calculated Stokes number for
sand particles at different mixture velocities. The results show that
Stokes numbers are relatively small for 100 lm sand particles,
meaning that the particle’s response time is relatively small com-
pared to the characteristic flow time scale.
Aguilar-Corona et al. (2011) mention that at Stokes numbers
(b) below the critical value of 10  5, collisions are dampened by the
interstitial fluid and direct particle–particle collisions are unlikely
to happen. At low Stokes numbers, fluctuations in particle motion
are caused by particle–fluid interactions, which is in accord with
the experimental results presented here.
Filtered cross-sectional averaged concentration fluctuations and
Stokes numbers calculated using Eq. (21) were compared with
Zenit and Hunt’s (Zenit and Hunt, 2000) experimental data and
the Buyevich and Kapbasov (1994) models for high frequency solid
concentration fluctuations. The results show increases in filtered
RMS concentration fluctuation with increasing concentration.
Due to experimental limitations, it was not possible to run the
experiments at concentrations beyond 35%. More experimental
results at higher solid concentrations are needed to comment on
the concentration at which the maximum RMS fluctuation occurs.
These results are shown in Fig. 18(a). Comparison of our RMS solid
Fig. 18. (a) Cross-sectional averaged concentration fluctuations at different con-
concentration fluctuations with those obtained by Zenit and Hunt
centrations and velocities: (b) Comparison of the results of the present study with is illustrated in Fig. 18(b). Note that the effect of fluid turbulence in
results reported by Zenit and Hunt (2000). pipe flows is implicitly accounted for in the Stokes number and is
60 S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61

not explicitly expressed in the figure. Concentration fluctuations Finally, the cross-sectional averaged concentration fluctuations
are generally smaller in our experiments, which is expected were compared to experimental data previously reported for fluid-
because Zenit and Hunt (2000) showed that the solid concentra- ized bed experiments and for gravity driven flows. The present
tion fluctuations decreased with decreasing Stokes number. results were consistent with the findings previously reported: that
Buyevich and Kapbasov (1994) obtained the following mathemat- is, that the magnitude of the concentration fluctuations increases
ical models for calculation of random small scale fluctuations with in situ concentration and should be expected to increase with
based on the Carnahan-Sterling and Enskog models. Their model increasing Stokes number. We have demonstrated that the solid
based on the Carnahan-Sterling model is concentration fluctuation results obtained for turbulent slurry flow
" #1 can be compared to those collected using fluidized beds if an
 2 4  Cs appropriate Stokes number is defined.
C s ¼ 1  2C s ð22Þ
ð1  C s Þ4

where C s and C s are RMS solid concentration fluctuations and aver-


References
age solid concentration, respectively. They mentioned that the
application of this model shown above to dense mixtures might Aguilar-Corona, A., Zenit, R., Masbernat, O., 2011. Collisions in a liquid fluidized bed.
not hold since it is based on chaotic states of a gas. To address this Int. J. Multiphase Flow 37 (7).
Alajbegovic, A., Assad, A., Bonetto, F., Lahey, R., 1994. Phase distribution and
concern, they introduced an alternative model, based on the Enskog
turbulence structure for solid/fluid upflow in a pipe. Int. J. Multiphase Flow 20
model for dense gases, which is expressed as: (3), 453–479.
2 0  1=2 131 Azzi, A., Azzopardi, B.J., Abdulkareem, N.H., Hunt, A., May 30–June 4 2010. Study of
"  1=3 # Cs fluidization using electrical capacitance tomography. In: 7th International
2 ¼ C2 Cs 6 1 B Cmax C7 Conference on Multiphase Flow, Tampa, FL, USA.
C s s 1 41 þ @  1=2 A5 ð23Þ Azzopardi, B., Jackson, K., Robinson, J., Kaji, R., Byars, M., Hunt, A., 2008. Fluctuations
C max 3
1  CCmax
s in dense phase pneumatic conveying of pulverised coal measured using
electrical capacitance tomography. Chem. Eng. Sci. 63 (9), 2548–2558.
Bolton, G., Hooper, C., Mann, R., Stitt, E., 2004. Flow distribution and velocity
where C max is the maximum packing concentration for the solid measurement in a radial flow fixed bed reactor using electrical resistance
which can be taken as solid volumetric concentration in the packed tomography. Chem. Eng. Sci. 59 (10), 1989–1997.
bed state. Both the above equations were compared in the present Buyevich, Y., Kapbasov, S., 1994. Random fluctuation in a fluidized bed. Chem. Eng.
Sci. 49 (8), 1229–1243.
work.
Collins, L.R., Keswani, A., 2004. Reynolds number scaling of particle clustering in
Fig. 18(b) shows the comparison of the present results with the turbulent aerosols. New J. Phys. 6, 119–135.
previous work by Zenit and Hunt (2000) and the theoretical mod- Datta, U., Dyakowski, T., Mylvaganam, S., 2007. Estimation of particulate velocity
components in pneumatic transport using pixel based correlation with dual
els (Eqs. (22) and (23)). In Fig. 18(b), the solid and dashed lines rep-
plane ect. Chem. Eng. J. 130 (2–3), 87–99.
resent these two theoretical models. Not surprisingly, both models Didwania, A.K., Homsy, G.M., 1981. Flow regimes and flow transitions in liquid
provide unsatisfactory predictions, as was previously observed by fluidized beds. Int. J. Multiphase Flow 7, 580–593.
Zenit and Hunt. The models predict only the limiting case where Dyakowski, T., Jeanmeure, L.F.C., Jaworski, A.J., 2000. Application of electrical
tomography for gas–solids and liquid–solids flows – a review. Powder Technol.
St ! 1. The comparison suggests the relationship between con- 112, 174–192.
centration fluctuation, solid concentration and Stokes number is Fangary, Y., Williams, R., Neil, W., Bond, J., Faulks, I., 1998. Application of electrical
not only valid for fluidized beds and gravity driven flow, but also resistance tomography to detect deposition in hydraulic conveying systems.
Powder Technol. 95 (1), 61–66.
holds for pipe flows provided that an appropriate definition of Gillies, R.G., Shook, C.A., 1994. Concentration distribution of sand slurries in
the Stokes number is employed. horizontal pipeflows. Particulate Sci. Technol. 12, 45–69.
Graham, L., Hamilton, R., Rudman, M., 2002. Coarse solids concentration profiles in
laminar pipe flows. In: Hydrotransport, BHR Group, pp. 149–158.
4. Conclusions Hashemi, S.A., 2013. Velocity and concentration fluctuations in concentrated solid–
liquid flows. Ph.D. thesis, University of Alberta.
Hughmark, G.A., 1977. Turbulent properties in the core of pipe flow. Indus. Eng.
The RMS solid concentration fluctuations of high-concentration Chem. Fundam. 16, 307–308.
sand slurries ðd50 ¼ 100 lmÞ in horizontal pipe flow were obtained Jenkins, J.T., Savage, S.B., 1983. Theory for the rapid flow of identical, smooth, nearly
using high-speed Electrical Impedance Tomography (EIT) for a elastic, spherical particles. J. Fluid Mech. 130, 187–202.
Kechroud, N., Brahimi, M., Djati, A., 2010. Characterization of dynamic behaviour of
range of in situ solid concentration and mixture velocities. In gen-
continuous phase in liquid fluidized bed. Powder Technol. 200, 149–157.
eral, the solid concentrations fluctuations and solid axial turbulent Kulick, J.D., Fessler, J.R., Eaton, J.K., 1994. Particle response and turbulence
intensities are higher near the wall because of particle–wall modification in fully developed channel flow. J. Fluid Mech. 277, 109–134.
collisions and because of the presence of the high shear zone near Lightstone, M.F., Hodgson, S.M., 2004. Turbulence modulation in gas-particle flows:
a comparison of selected models. Can. J. Chem. Eng. 82 (2), 209–219.
the wall. At constant solid concentrations, the magnitude of the con- Longwell, P., 1966. Mechanics of Fluid Flows. McGraw- Hill, New York.
centration fluctuations increases with increasing mixture velocity. Norman, J.T., Bonnecaze, R.T., 2005. Measurement of solids distribution in
Time-series power spectra of concentration- and velocity- fluc- suspension flows using electrical resistance tomography. Can. J. Chem. Eng.
83 (1), 24–36.
tuations (averaged over the pipe cross section) were obtained Oppenheim, A.V., Schafer, R.W., Buck, J.R., 1999. Discrete-Time Signal Processing.
using Fast Fourier Transform (FFT). This analysis showed that the Prentice Hall, Upper Saddle River, NJ, USA.
rate of decay in solid fluctuations is closely related to the turbulent Pachowko, A.D., Wang, M., Poole, C., Rhodes, D., September 2–5, 2003. The use of
electrical resistance tomography (ERT) to monitor flow patterns in horizontal
energy decay in the dissipation range for single-phase flow, slurry transport pipelines. In: 3rd Congress on Industrial Process Tomography,
suggesting that solid concentration fluctuations are produced Banff, AB, Canada.
primarily through particle–fluid turbulence interactions. Compari- Picciotto, M., Marchioli, C., Reeks, M.W., Soldati, A., 2005. Statistics of velocity and
preferential accumulation of micro-particles in boundary layer turbulence.
son of the calculated turbulence length scale with the solid length Nucl. Eng. Des. 235 (10–12), 1239–1249.
scale (i.e. particle diameter) illustrated that, for these slurries, the Pope, S., 2008. Turbulent Flows, fifth ed. Cambridge University Press, Cambridge,
particles interact with turbulent eddies in the dissipative range UK.
Press, W.A., Flannery, B.P., Teukolsky, S.A., Vetterling, W.A., 1988. Numerical Recipes
of the turbulent energy spectrum. Additionally, measurements of
in C. Cambridge University Press, Cambridge, UK.
the local, time-averaged solid velocity distributions were used to Saffman, P.G., 1965. The lift on a small sphere in shear flow. J. Fluid Mech. 22,
calculate velocity defect profiles for these flows. The velocity defect 385–400.
profiles were very similar to those expected for single phase Schaan, J.J., 2001. Pipeline flow of newtonian fine particle slurries. Master’s thesis,
University of Saskatchewan.
turbulent flow (at the same flow conditions), particularly in the Shook, C.A., Gillies, R.G., Sanders, R.S., 2002. Pipeline Hydrotransport with
core of the flow. Applications in the Oil Sand Industry. SRC Publications, Saskatoon, SK, CA.
S.A. Hashemi et al. / International Journal of Multiphase Flow 66 (2014) 46–61 61

Thomas, D.G., 1965. Transport characteristics of suspensions: Viii. A note on the Xu, J., Wu, Y., Zheng, Z., Wang, M., Munir, B., Oluwadarey, H.I., Schlaberg, H.I.,
viscosity of Newtonian suspensions of uniform spherical particles. J. Colloid Sci. Williams, R.A., 2009. Measurement of solid slurry flow via correlation of
20 (3), 267–277. electromagnetic flow meter, electrical resistance tomography and mechanistic
Truesdell, G.C., Elghobashi, S., 1994. On the two-way interaction between modelling. J. Hydrodynam. 21 (4), 557–563.
homogeneous turbulence and dispersed solid particles. Part ii: Particle Zenit, R., Hunt, M.L., 1998. The impulsive motion of a liquid resulting from a particle
dispersion. Phys. Fluids 6 (3), 1405–1407. collision. J. Fluid Mech. 375, 345–361.
Varaksin, A.Y., Polyakov, A.F., 2000. Experimental study of fluctuations of particle Zenit, R., Hunt, M.L., 1999. Mechanics of immersed particle collisions. J. Fluids Eng.
velocity in turbulent flow of air in a pipe. High Temp. 38 (5), 764–770. Trans. ASME 121 (1), 179–184.
Walpole, R.E., Myers, R.H., Myers, S.L., Ye, K., 2007. Probability and Statistics for Zenit, R., Hunt, M.L., 2000. Solid fraction fluctuations in solid–liquid flows. Int. J.
Engineers and Scientists, eighth ed. Pearson Prentice Hall, Upper Saddle River, Multiphase Flow 26 (5), 763–781.
NJ, USA. Zenit, R., Hunt, M.L., Brennen, C.E., 1997. Collisional particle pressure measurements
Wilson, K.C., Sellgren, A., 2003. Interaction of particles and near-wall lift in slurry in solid–liquid flows. J. Fluid Mech. 353, 261–283.
pipelines. J. Hydraul. Eng. 129 (1), 73–76. Zhu, K., Madhusudana Rao, S., Wang, C., Sundaresan, S., 2003. Electrical capacitance
Wilson, K.C., Sanders, R.S., Gillies, R.G., Shook, C.A., 2010. Verification of the near- tomography measurements on vertical and inclined pneumatic conveying of
wall model for slurry flow. Powder Technol. 197 (3), 247–253. granular solids. Chem. Eng. Sci. 58 (18), 4225–4245.

You might also like