You are on page 1of 13

Proceedings of the ASME 2012 Fluids Engineering Summer Meeting

FEDSM2012
July 8-12, 2012, Rio Grande, Puerto Rico

FEDSM2012-72379

MODELLING CONCENTRATED SLURRY PIPELINE FLOWS

Charlene L. Antaya, Kofi Freeman K. Adane and R. Sean Sanders*


Department of Chemical and Materials Engineering
University of Alberta
Edmonton, Alberta, Canada

ABSTRACT injection and dispersion of polymer flocculant into a fine


A numerical investigation of two-phase solid-liquid (slurry) tailings mixture during pipeline transport. In such a situation,
flow in horizontal pipes has been carried out. Simulations of the interactions among the solid particles, fluid and pipeline
concentrated slurry flows in pipes 0.0515 m and 0.15 m in systems need to be well understood. These advancements are
diameter were performed using the two-fluid approach required in the resource extraction industries world-wide to
implemented in the commercial CFD code, ANSYS CFX. move to the next level of process intensification, to reduce
Mixtures of monosized and bimodal particle sizes were tested. operating costs, to implement advanced process control
Several test cases were investigated to predict particle velocity- schemes, and to develop effective preventative maintenance
and concentration-distributions and frictional pressure programs.
gradients. The effects of turbulence model selection, dispersed With an ever decreasing cost-to-computing power
phase wall boundary conditions, and interphase force terms on ratio, computational fluid dynamics (CFD) has become more
model performance were evaluated. The selection of turbulence attractive for the design of new processes, to predict the effects
model had a significant impact on the dispersed phase velocity of manipulating process parameters on existing systems, or to
and concentration distributions. Comparison of simulations with evaluate the performance of existing systems. Although CFD
benchmark experimental data shows clearly that for the has its own challenges, its ability to (i) provide detailed local
relatively small particle sizes (~ 100 microns), poor solids information on flow parameters and (ii) allow a user to conduct
concentration profile predictions are obtained if the turbulent parametric studies with ease and at limited expense makes this
dispersion force is not included. In general, very good tool very attractive. This is particularly true since experimental
agreement between numerical and experimental results was measurement of local parameters is difficult, time-consuming
observed. and expensive. CFD analysis is therefore needed, but few
studies comparing CFD results to experiments have been
INTRODUCTION
published. Additionally, CFD analysis of highly concentrated
Slurry pipeline flows are an important part of many
slurry flows requires the use of closure equations, which are, to
industrial processes, including oil sands mining and extraction,
some degree, empirical or semi-empirical in nature (which is
mining and mineral processing and pulp and paper production.
described in greater detail in the following paragraph).
Presently, good phenomenological models (such as the SRC
Therefore, the appropriate place to begin is to conduct CFD
Two-Layer model) exist and these are able to predict bulk
analyses for slurry flows in relatively simple straight-pipe
parameters needed for engineering design, i.e. frictional
geometry, for which some benchmark experimental data are
pressure loss and deposition velocity, which is the minimum
available.
mixture velocity required to ensure that a stationary deposit of
For turbulent multiphase flows, such as the slurry
particles does not form. However, engineering models of this
flows studied here, models are used to describe interactions
type provide almost no information about variations of particle
among phases. A number of these models are flow-specific and
concentration and velocity within the flow domain. Knowledge
their general applicability is thought to be an issue. For
of time-averaged, local phasic information is necessary to, for
example, kinetic theory, which is now commonly used in
example, advance the science of pipe wear rate predictions and
multiphase flow simulations, was originally developed for gas-
to characterize the flow in more complex geometries, e.g.
solid flows to account for particle-particle interactions. One
pumps and pipe fittings. In addition, it is crucially important
would expect that the effects of fluid density and viscosity are
that pipelines are treated as “process units”: for example, the

*Corresponding author: ssanders@ualberta.ca 1 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


much more important in slurry flows than in gas-solid flows. In d mass median particle diameter (mm)
addition, a variant of the k-H turbulence model is widely used to g acceleration due to gravity (m/s2)
model turbulent multiphase flows, even those involving k turbulent kinetic energy (m2/s2)
curvature effects, flow separations, etc. While success has been L pipe length (m)
achieved in the use of k-H based models, it is worth noting that M interphase force (kg/m·s2)
some inherent shortcomings, especially over-estimation of the p thermodynamic pressure (Pa)
eddy viscosity in the near-wall region and early onset flow Px frictional pressure gradient (kPa/m)
separation, still persists. TM turbulent modulation (%)
Very few attempts to compare slurry pipeline flow uc bulk velocity at the pipe inlet (m/s)
experiments with CFD simulations can be found in the Us normalized streamwise solids velocity (= us / uc)
literature. Notable exceptions include the studies of Krampa- us streamwise solids velocity (m/s)
Morlu et al [1], Lin and Ebadian [2], Chen et al [3], Ekambara y transverse coordinates (m)
et al [4], Bossio et al [5], and Lahiri and Ghanta [6]. Subscripts
Interestingly, none of these works evaluated the importance of l liquid or continuous phase
turbulence model selection, nor did they evaluate the s sand or dispersed phase
importance of different interphase force terms on model Greek
performance. I local concentration or volume fraction
For two-phase flows of the type studied here, two main ρ density (kg/m3)
numerical approaches are usually used: namely, the Eulerian- σ turbulent Prandtl number
Lagrangian and Eulerian-Eulerian methods. The Eulerian- H dissipation rate of k (m2/s3)
Lagrangian approach is typically used for very low solids P effective viscosity (Pa·s)
concentrations and is not considered in the present study. The Pm dynamic viscosity (Pa·s)
Eulerian-Eulerian approach, also referred to as the two-fluid Pt eddy viscosity (Pa·s)
model, is widely used to model high solids concentration flows
because inter-particle interactions can more easily be accounted PREVIOUS WORKS
for. The two-fluid model considers both the liquid (continuous) Many engineering models currently exist for slurry
and solid (dispersed) phases as two interpenetrating continuum, pipeline flow; however, most of these models focus only on the
and the mathematical governing equations, i.e. the continuity prediction of frictional pressure drop and minimum operating
and momentum (Reynolds Average Navier-Stokes, RANS) velocity, or deposition velocity, required to prevent solids
equations, are solved for both phases. The two-fluid approach accumulation in the pipeline [9]. Typically, these models are
requires that the turbulence and momentum exchange between empirically derived.
phases be modelled. The momentum exchange between phases There are also several mechanistic-based models
is considered in the form of forces, such as drag, lift, turbulence available that can accurately predict bulk (averaged)
dispersion, turbulence enhancement, wall lubrications, etc. For parameters. For example, the SRC Two-Layer model (Gillies et
dense slurry flows, turbulence dispersion is the dominant al [10]) is known for its ability to accurately predict the
interphase force [7] and is therefore the focus of this work. In frictional pressure gradient and deposition velocity in slurry
this study, two turbulence models, the k-epsilon (k-H) and shear pipelines. Predictions are obtained using pipe diameter, mixture
stress transport (SST), were implemented and the effect each flow rate, solids properties (size, density, concentration) and
had on model performance was evaluated. fluid properties (density and viscosity) as input parameters. This
The main objective of this study is to investigate the model is widely used in the mining and mineral processing
performance of a commercial CFD code, ANSYS CFX, in industry [9]. One of the major shortcomings of this model is its
predicting slurry pipeline flows. This was done by comparing inability to predict local quantities, such as fluid velocity
the computed results with experimental data of Hashemi et al fluctuations, local solids concentrations, or local time-averaged
[7] and Matousek [8]. The Hashemi et al [7] study presents velocities of either phase. Additionally, the model cannot be
measurements of local concentration distribution and local used for complex geometries (e.g. elbows, noncircular conduits
solids velocity whereas the Matousek [8] work contains or pumps). It should be mentioned that the model employs a
information on local solids concentration and frictional pressure limited number of semi-empirical coefficients obtained from
gradient. These experimentally-determined quantities were controlled experiments.
compared with those obtained from computations. Many previous slurry pipeline flow studies are not
particularly useful in testing or validating CFD models because
NOMENCLATURE small pipelines were used, low solids volume concentrations
C bulk concentration or volume fraction at pipe inlet were tested, or only averaged parameters (e.g. deposition
CTD turbulent dispersion coefficient velocity and frictional pressure gradient) were reported. Thus,
D pipe diameter (m) one of the focus areas for the research group at the University

2 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


of Alberta is to provide benchmark data that can be used to their CFD results tended to underestimate solids concentration
validate numerical models. Some of the data collected by this at the top of the pipe while overestimating it at the pipe invert.
group are reported here. Chen et al [3] considered the bimodal version of the
Recently, Krampa-Morlu et al [1] investigated the model formulation of Ekambara et al [4] on relatively smaller
capabilities of a commercial CFD code, ANSYS CFX 4.4 to pipe sizes (D < 0.055 m) for low density ratio, water-coal
predict flow parameters for coarse particle slurry pipeline flow slurries. Their simulations employed the RNG k-Hturbulence
in a vertical pipe. For this study, the k-H turbulence model was model. The predicted values of frictional pressure gradient were
used for each phase, and the effect of particle diameter, mean compared with experimental data. A reasonable agreement was
particle concentration and solid viscosity were investigated. observed when treating the flows as bimodal mixtures rather
Particle-particle interactions were modeled using empirical than monosize ones. Since no experimental measurements of
correlations which are now known to be inadequate. The only local quantities were made, their comparison was between the
interphase force included in their simulations was the drag monosize mixture setting and the bimodal one. Their conclusion
force. Krampa-Morlu et al [1] reported that the capability of was that the coal was predicted to be more uniformly distributed
CFX was very limited with poor agreement between the CFD over the pipe cross section in the bimodal than in the monosize
predictions and the experimental results, particularly for slurries mixture. Again, no experimental data were available to support
comprised of particles greater than 1.7 mm in diameter. Their this conclusion.
evaluation showed that prediction of (i) particle velocity and The recent work by Bossio et al [5] is also very similar
concentration profiles and (ii) the effect of solids concentration to that of Ekambara et al [4], except that in their case the
on frictional pressure gradient were highly unsatisfactory. continuous phase was a non-Newtonian fluid. Also, the bulk
Ekambara et al [4] conducted a numerical study on concentrations were much lower, i.e. less than 10% by volume.
horizontal slurry pipeline flows using ANSYS-CFX, version Comparison was only made for frictional pressure gradient, for
10.0. In this study, the effect of bulk concentration and velocity, which reasonable agreement was observed. Bossio et al [5] also
particle size and pipe diameter on local phasic concentrations concluded that turbulent kinetic energy influences the local
and velocities were modeled. The two-fluid approach was used solids concentration.
with the continuous (liquid) phase considered to be turbulent In other studies (Lin and Ebadian [2]; Ling et al [12]),
and the dispersed phase treated as a laminar flow. The k-H the algebraic slip model was employed for horizontal slurry
turbulence model was used for the continuous phase. The flow. In this model, it is assumed that the fluid and solid phases
particle-particle interactions were modeled using kinetic theory share the same flow field, which is different from the two-fluid
for granular flow, which represents one of a number of model approach.
differences between the Ekambara et al [4] study and the To the best knowledge of the authors, none of the
previous one conducted by Krampa-Morlu et al [1]. The other previous studies in the open literature has examined the effect
significant differences were that the Ekambara et al [4] study of different turbulent dispersion force and/or dispersed phase
involved horizontal flows in 50 to 500 mm diameters whereas turbulence models in CFD codes for predicting slurry pipeline
Krampa-Morlu et al [1] studied vertical flows in pipes of 25.8 flows. This study investigates the model performance for four
and 40 mm. Additionally, interphase force terms were different dispersed phase turbulence models, as well as various
accounted for. Ekambara et al [4] reported that both lift and interphase forces.
wall-lubrications forces were insignificant in horizontal slurry
pipelines. Similarly, neither of the drag force models they tested NUMERICAL PROCEDURE
provided superior predictions. The Favre average of the A commercial CFD code, ANSYS CFX was used to
interphase drag force (FAD) was used to model the turbulence solve mass and momentum conservation equations for each
dispersion force. Turbulence induced by the particle on the phase. Here, liquid is considered as the continuous phase and
liquid phase was also modeled. Comparison of previously the particles are regarded as the dispersed phase. Both phases
reported experimental data with the results of their numerical are assumed to be incompressible, Newtonian, isothermal
simulations was reported. Solids concentration and liquid fluids. Particle-particle interactions are modeled using kinetic
velocity distributions and frictional pressure drops were theory. The drag (Gidaspow [13]), lift (0.5 as lift coefficient),
compared. Satisfactory agreement between measured and near-wall lubrication (Antal et al [14]), and turbulence
predicted flow parameters was obtained for particles smaller dispersion forces were implemented. Virtual mass transfer was
than 270 Pm, especially in the core region of the pipes. ignored since the dispersed phase is solid. The Favre average
Discrepancies reported between the numerical results and drag (Burns et al [15]) and Lopez de Bertodano [16] models
experimental measurements for larger particles was ascribed to were used for the turbulence dispersion force, hereby referred to
a near-wall lift force (Wilson et al [11]) which may be different as FAD and Lopez, respectively. Another force that was ignored
from the interphase lift force implemented in ANSYS CFX. is the particle induced turbulence on the continuous phase. This
Although Ekambara et al [4] did not specifically mention it, force results from a wake forming behind the particles, which
can be significant for large particles. The full mathematical
details are given in the Appendix. In addition, Appendix

3 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


includes the mathematical equations describing the turbulence computed here are shown in Table 2. In addition, Matousek’s
models used. For the present study, the continuous phase is [8] results offer the ability to validate the numerical results for
assumed to be turbulent and therefore, the effective viscosity is larger pipe diameters and also for bimodal particle distributions.
computed using one of the following two-equation turbulence As reported in Table 2, the fine particle-to-medium particle
models: the standard linear-eddy-viscosity k-H (referred to as k- concentration ratio for Matousek’s [8] tests is 0.72. The
H) and shear stress transport (SST) models. bimodal mixture tested by Chen et al [3] had a fine particle-to-
ANSYS CFX has the following options for the coarse concentration ratio of 3. It should be mentioned that for
dispersed phase: all flow conditions considered here, one would expect
i. Laminar interesting and complex combinations of flow physics to exist,
ii. Zero-Equation model including fluid-particle and particle-particle interactions, and
iii. Dispersed Phase Zero-Equation (DPZ) model: turbulent dispersion, thus which make them good datasets for
P st ( U s Plt ) /( UlV ) ; V is turbulent Prandtl number, ρ is CFD validations.
density, μt is eddy viscosity, and s and l denote solid and Table 1: List of computed test cases
liquid, respectively
Case Name Dispersed Phase Forces
iv. Homogeneous turbulence
The Zero-Equation model is based on the Prandtl
k-e-LamNS Laminar
mixing length and the user must specify both eddy length and
velocity scales. This model is known to be inadequate for k-e-DisNS1 DPZ, V = 1.0
capturing the physics of turbulent flows. Also, because k-e-DisNS2 DPZ, V = 2.65
reasonable length and velocity scales are not known a priori, SST-DisNS1 DPZ, V = 1.0 FAD
this model is not considered in the present work. In the DPZ
k-e-DisFS1 DPZ, V = 1.0
model, the value of the turbulent Prandtl number must be
specified by the user. Here, values of 1 and 2.65 were used SST-DisFS1 DPZ, V = 1.0
since the particle relaxation time is approximately of the same SST-Homo
order of magnitude as the turbulence dissipation time scale. The
SST-HomoL015 Lopez, CTD = 0.15
premise of the turbulence homogeneous approach is that both
phases share the same turbulence flow field and therefore bulk SST-HomoL1 Homogeneous Lopez, CTD = 1.0
turbulence transport equations are solved using bulk density and SST-HomoL10 Lopez, CTD = 10.0
viscosity rather than solving individual phasic turbulence SST-HomoL50 Lopez, CTD = 50.0
transport equations. It should be stressed that these bulk
quantities are computed locally by summing the product of SST-HomoWL FAD, WL + Lift
local volume fractions and respective phasic properties. The Table 2: Experimental test case studied here
disadvantage of this approach is that one cannot specify phasic
boundary conditions at the wall. In certain situations, this may Case D(m) d(micron) uc(m/s) C
not be an issue, especially if one ensures that the first node from 1 0.2
the wall is within the logarithmic zone and also that the particle 2 2 0.3
diameter is greater than the viscous sublayer thickness. Table 1 3 0.0515 100 3 0.2
shows the test cases considered. Each test case name contains 4 0.3
the turbulence model used for both phases and the wall 5 0.15 370 6 0.43
boundary condition selected for the dispersed phase: for 6 [120, 370] [0.18, 0.25]
example, “k-e-LamNS” means that the k-H model was used for
Fluid properties and particle densities are given in
the continuous phase, the dispersed phase was assumed to be
Table 3. These values, where available, were taken from the
laminar and the no-slip condition was applied to the solid
respective experimental work. Matousek [8] did not provide any
particles at the wall. It should be noted that the drag force is a
data on particle density; however, his studies involve sand and
default interphase force included in all the test cases.
therefore, the value obtained from Hashemi et al [7] was used.
Problem Definition: Experimental works by Hashemi et al For both experimental studies mentioned (Hashemi et al [7] and
[7] and Matousek [8] were used for validation. Hashemi et al Matousek [8]), the test section is oriented horizontally.
[7] reported experimental data for bulk velocity, uc and Although, Hashemi et al [7] did not provide the length of the
concentration, C in the range of 2-5 m/s and 0.2 – 0.35. For this test loop it is known that measurements were made a distance
validation, any of the experimental conditions are appropriate (L) of 5 m (L/D ~ 97) from the nearest (upstream) bend. The
and therefore, only selected conditions are computed, see Table total length of Matousek’s loop was 65m including vertical
2. Since the results obtained for Cases 1 through 4 are similar, sections, elbows, etc., but measurements were made at 6.55 m
the results of Case 1 are presented here. Similarly, Matousek [8] (L/D ~ 44) downstream of the bend in the horizontal section. In
reported measurements for several flow conditions and those

4 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


the present simulations, a straight pipe section 12.5 m in length adaptation of earlier techniques by Rhie and Chow [17]. A
(L/D ~ 200) was used to ensure a fully developed flow. standard finite element derivative approximation via shape
Meanwhile, the data presented here were extracted at L/D = 80 functions is used for the diffusion terms. The high resolution
and 190 for Hashemi et al [7] and Matousek [8], respectively. It scheme (HRS), which is based on the formulation of Barth and
is worth mentioning that these locations are in the fully Jesperson [18], was used for both advection terms and
developed section and free from the influence of the outlet turbulence model equations. It should be mentioned that HRS is
boundary condition. a blend of both first and second order schemes with the
blending factor being a function that is computed locally.
Table 3: Fluid properties and particle density
The discretized mass, momentum and volume fraction
Water Sand
conservation equations were fully coupled and solved
Hashemi et al [7] simultaneously using additive correction multigrid to accelerate
Density (kg/m3) 998.2 2650 convergence. Double precision was used in the computations
Dynamic viscosity (mPa•s) 0.89 - and solutions were considered converged when the normalized
sum of the absolute dimensionless residuals of the discretized
Matousek [8] equations was less than 10-6. Although this criterion was used in
Density (kg/m3) 998.2 2650 some cases, much lower values were also observed in others
Dynamic viscosity (mPa•s) 1.002 - because convergence was actually based on monitoring the
Hashemi et al [7] used electrical impedance volume fractions at several radial locations in the fully
tomography to obtain both local distribution of the particle developed region. This procedure was adopted because the
velocity and concentration in the fully developed section. authors’ experience showed that the particle concentration
Details of this instrument and operation can be found in converged much slowly than the other parameters.
Hashemi et al [7]. They reported that the uncertainty of the
concentration distribution measurements was ±2.2%. A
radiometric (nuclear) density meter and differential pressure
transmitter were used by Matousek [8] to obtain the chord-
averaged particle concentration at various horizontal positions
through the flow domain and frictional pressure drop,
respectively. Matousek [8] reported the uncertainty as 4.7% for
frictional pressure drop. The uncertainty for concentration was
7.6%. In the subsequent presentation of the results, these
uncertainties will be shown. A maximum particle packing value
of 0.64 was used for Hashemi et al [7] and 0.62 was used for Figure 1: Sample of a computational mesh
Matousek [8], since neither provided it. This value has to be A five-block structured computational mesh
determined experimentally for any mixture as it is an important representing the entire domain (see Fig. 1 above) was created
input to any pressure loss prediction. using ANSYS ICEMCFD and then imported into ANSYS CFX.
For all the test cases, a turbulent intensity of 5% and The grid is made up of O-grids and the mesh contracts
eddy viscosity ratio (P t/Pm) of 10 were specified at the inlet. geometrically towards the wall in the near-wall regions.
The expressions used to compute turbulent kinetic energy and Uniform grid spacing was used along the axial direction. Grid
dissipation rate from these values are given in the Appendix. A independence tests were performed using a coarse mesh of
relative average static pressure of zero across the outlet plane 96,150 nodes, a medium mesh of 252,250 nodes and a fine
was specified for the outlet boundary condition. The no-slip mesh of 403,600 nodes. These mesh sizes were chosen because
condition was assumed to apply at the pipe wall for the preliminary results indicated that the results, especially the
continuous (water) phase. The so-called scalable wall-function frictional pressure drop, are sensitive to the number of nodes in
version of k-H was used for the near-wall treatment. This the axial direction. Therefore, particular attention was paid to
required that y+ ≥ 11.5, which is lower than a typical standard the pressure drop obtained using the different mesh sizes.
wall-function model (y+ > 25). The SST model does not require Additionally, differences in the particle velocity- and
this criterion due to its automatic near-wall capability. Either concentration- distributions for the three different meshes were
no-slip or free-slip wall boundary conditions were used for the examined. The maximum difference occurred for frictional
dispersed phase depending on the test case (see Table 1). pressure gradient calculated using the medium and coarse grids;
however, the difference was less than 0.5%. The medium and
Solution Method: The numerical solution of the governing fine grids provided results that were almost identical. Based on
equations was obtained using ANSYS CFX, version 12.1. these tests, the medium mesh was used for all runs conducted
ANSYS CFX uses a finite volume method, but is based on a for comparison with the data of Hashemi et al [7]. For the
finite element approach to represent the geometry. Mass simulations of the Matousek [8] runs, a mesh of 560,500 was
conservation discretization on the non-staggered grid is an sufficient for grid independence. In all test cases using the

5 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


medium mesh, the y+ value is 54 for k-H and 56 for SST for profiles produced using the different boundary conditions are
Hashemi et al [7]. The corresponding value for Matousek [8] is nearly identical in shape, but the no-slip result provides superior
27. It is worth noting that in cases where two turbulence models agreement with experiments. Although it is not shown on the
were used, grid independence tests were performed using each figure, a similar observation was also made when the SST
model. model was used for the continuous phase.

4.0 DISCUSSION
In this section the results of the computed test cases are
evaluated and compared with the experimental values
determined by Hashemi et al [7] and Matousek [8]. As stated in
Section 1.0, the purpose of the present work is to assess
different dispersed phase turbulence models and non-drag
interphase forces. First, the local particle velocity and
concentration profiles obtained from numerical simulations are
compared with experimental data. Specifically, the Case 1 (see
Table 2) results reported by Hashemi et al [7] are compared
with the simulations. The numerical results from various
dispersed phase turbulence models and non-drag interphase a)
forces are compared to determine which model most accurately
predicts the experimental results. The model providing most
favourable comparison with the experimental data was then
used to compute the other flow conditions (Cases 2-6 in Table
2) and the results are discussed herein. Although neither of the
experimental studies provides information on turbulent intensity
or modulation, the accurate prediction of some integral
parameters such as frictional pressure drop and heat transfer
characteristics are highly dependent on the knowledge of the
effect that the dispersed phase has on the continuous phase.
Therefore, a discussion of the effect of the presence of the
particles on the fluid turbulence intensity obtained from the
numerical study conducted as part of this investigation is given. b)
This discussion provides qualitative insight into the fluid flow Figure 2: a) Normalized sand velocity and b) concentration
structure and also highlights the type of measurements needed distributions. See Table 1 for the meanings of the acronyms
for experimental verification. used in the legend.
It can be seen that treatment of the dispersed phase as
Dispersed Phase Turbulence Models: Figures 2a and 2b
a laminar flow does not provide good predictions of the
show the normalized mean particle velocity, Us (= us/uc) and
dispersed phase velocity distribution. In fact, the velocity
concentration (Is) profiles over the pipe cross section,
profile is more parabolic in shape which is typical of laminar
respectively. These figures show the results obtained from the
flow and, consequently different from other profiles, including,
simulations using various dispersed phase boundary conditions,
most critically, the experimental results. This is an interesting
and both dispersed and continuous phase turbulence models.
observation since this approach has been employed in the
The reader is referred to the legend of Figure 2 and the
previous work by Ekambara et al [4], although they did not
information of Table 1 for specifics. In addition, the
report the particle velocity distribution. On the other hand, Fig.
experimental data of Hashemi et al [7] are shown on the figure.
2b shows that in the near-wall region, the assumption of laminar
The results for k-e-DisNS1 and k-e-DisFS1 on Fig. 2a
flow for the disperse phase results in improved prediction of the
demonstrate the effect of using no slip and free slip as wall
concentration distribution; however, no difference was observed
boundary condition for the sand, with k-ε turbulence model for
among the results in the core region. The other results of Fig. 2
water. It appears that the free-slip boundary condition is better
show that increasing the Prandtl number (see Table 1) does not
able to predict the velocity profile. The runs implementing the
necessarily have a significant effect on the predicted results
no-slip boundary condition overestimated and underestimated
The results from SST-DisNS1 simulations were not
the velocity values in the core region and near-wall region,
shown in Fig. 2 because they are almost identical to SST-Homo
respectively. However, it can be seen from Fig. 2b, which shows
case (which is shown). It is clear from the results presented in
the local concentration profile, that the no-slip result is similar
Fig. 2 that the SST model (i.e. the SST-Homo case) provided
to the experimental result, especially in the near-wall region at
much better predictions of the velocity and concentration
the top of the pipe. It should be said that the concentration
distributions than the k-H model, particularly in the near-wall

6 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


region. The poor performance of the k-ε cases may be due to its dispersion coefficient in the Lopez correlation is increased. A
well-known inability to model the flow in the near-wall region comparison of the concentration profiles of SST-HomoL10 and
because of over-prediction of the eddy viscosity. The SST SST-HomoL50 indicates that, the optimal value is likely
model is less prone to this deficiency since it accounts for the between 10 and 50, but apparently closer to 50. This suggests
transport of the turbulent shear stress [19]. that the turbulence dispersion model does not have a significant
Using the SRC Two-Layer model [10], the frictional impact on the predicted results. It must be noted that the
pressure gradient was calculated to be 1.09 kPa/m for this flow effectiveness of the Lopez turbulence dispersion model will
condition. The numerical results obtained here underestimated vary from system to system, meaning that the optimal
the frictional pressure gradient irrespective of turbulence model, coefficient must be determined individually for each system,
with the most significant difference occurring when the free slip and can vary greatly from one to another. Another interesting
boundary condition was employed for the dispersed phase. For observation from Fig. 3 is that the effect of the turbulent
example, the frictional pressure gradient predictions for the k-e- dispersion force is clearly demonstrated by varying the Lopez
DisNS1, k-e-DisFS1 and SST-Homo test cases, respectively, coefficient values. In fact, this force has a greater effect on
are 0.91, 0.56 and 0.94 kPa/m, representing differences from concentration profile than on velocity distribution or frictional
the SRC model predictions of -16.4%, -48.7% and -14.3%, pressure gradient. Also, the improved predictions obtained
respectively. The relatively more accurate value produced using when larger dispersion coefficients are used indicate that this
SST-Homo is consistent with the observations made for sand force is dominant in the present flow condition. This is
velocity and concentration distributions (shown in Fig. 2) where consistent with the experimental observation made by Hashemi
the SST-Homo gave a better prediction. Also, with the poor et al [7].
performance of the cases using the free slip boundary condition
for dispersed phase, it is tempting to conclude that the zero
shear stress at the pipe wall assumption for the sand particle,
especially at the pipe invert, might not hold for the present
work. The velocity of the dispersed phase near the wall might in
fact be retarded by wall friction effects. In the absence of
experimental data in the near-wall region, one can estimate that
the thickness of viscous sublayer (y+ ≤ 5) is in fact little more
a) SST-HomoL1 SST-HomoL50 SST-Homo
than one-half the size of the sand particle diameter (assuming
the viscous sublayer thickness is based on the properties of the
continuous phase only) . Therefore, if any sand particle found
itself in this region of the flow, the incoming upstream flow
might not necessarily be able to lift or slide it along due to the
wall viscous friction.
Non-drag Interphase Forces: As described in the previous
paragraphs, solving the same transport equations for both
phases using the SST model gave better predictions and as a
result, it will be used as the reference in the subsequent
discussion. Here, a comparison is done for various non-drag
interphase forces: turbulence dispersion, lift and wall
lubrication. First, the turbulence dispersion force was activated
using the FAD model (SST-Homo) and the Lopez model with b)
dispersion coefficients of CTD = 0.15 (SST-HomoL015), CTD =
1 (SST-HomoL1), CTD = 10 (SST-HomoL10), and CTD = 50 Figure 3: Comparison of various non-drag forces for
(SST-HomoL50). Use of the FAD model for the turbulence concentration distribution
dispersion force in combination with both lift and wall Concentration and Multifluids: The results for various
lubrication forces (SST-HomoWL) was also tested (see Table bulk velocities and solids concentrations are compared in Fig.
1). Although not shown here, the velocity distribution and 4. Figure 4a compares velocity profiles from the present results
frictional pressure gradient were found to be insensitive to both and the measured data by Hashemi et al [7]. In the region y/D ≤
models used for the turbulence dispersion force and other 0.7, all the present results were higher than the measured data
interphase forces such as lift and wall lubrication. Similarly, as by Hashemi et al [7] by a maximum of 6%, irrespective of bulk
shown in Fig. 3b, the concentration distributions from both velocity and concentration. Note that if one estimates the bulk
SST-Homo and SST-HomoWL tests are indistinguishable. velocities from the Hashemi et al [7] data, the values are
Figure 3 shows that the accuracy of the concentration somewhat higher than those used in the present work or even
distribution predictions improves when the value of the reported in their work (a difference of 6%), but these

7 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


discrepancies are not significant based on the relative the dispersed phase is not monosized and thus it should be
insensitivity of particle concentration distribution on bulk treated as a bimodal mixture in the simulations. Here, we
velocity. The normalized velocity distribution is also compare results of our simulations with the experimental results
independent of both bulk velocity and concentration for the of Matousek [8]. Matousek [8] presented data only on
measured data in the region y/D ≤ 0.7. A similar observation concentration for both monosize and bimodal, and frictional
was made for the present results throughout the entire pipe. pressure gradient for the monosized mixture. Generally, the
Irrespective of the bulk concentration, in the region y/D > 0.7 shape of velocity profiles was not altered in any of the fluids, as
and for uc = 3 m/s, the shape of calculated velocity profile is shown in Fig. 5a. It is evident from Fig. 5b that a comparison
different from that of the measured data. Since the trend in this between the present and experimental data for the monosized
region is only for uc = 3 m/s and is not consistent, it is not clear dispersed phase showed a very good agreement, except in the
if it is representative of the flow behavior or is simply an bottom wall region. The overestimation of the concentration at
artifact. Very good agreement was observed between the the bottom wall suggests that this region may need special
computational and experimental results for the concentration treatment, which is consistent with earlier observations made
profile, as shown in Fig. 4b. This good agreement is not for Hashemi et al [7] in their smaller pipe. It should be noted
surprising since the shape of local concentration distribution that a computation with free slip condition at the wall for the
tends to be insensitive to small changes in the bulk velocity, dispersed phase was also performed. The results were identical
again suggesting that the 6% discrepancy reported by Hashemi to those presented in Fig. 5 (where the no-slip condition was
et al [7] is not important. Also, the normalized concentration applied).
distribution depends on both bulk velocity and concentration in
Table 4: Frictional pressure gradient, Px
both results (not shown on the figure). The frictional pressure
gradients obtained from these test runs are similar to those uc (m/s) C Px (kPa/m) Difference (%)
estimated using SRC model, as shown in Table 4. Present SRC
2 0.2 0.852 0.874 -2.60
3 0.2 1.760 1.666 5.60
0.3 1.940 1.932 0.40
Figure 5b shows that a discrepancy exists between the
simulation results and the experimental data for the bimodal test
case. In fact, even the shape of the profile from the
computational result is a little different. Although, the effect of
the third fluid is seen on the plot, the extent is much less than
that of the experimental measurement. The present result is no
different from those reported by Chen et al [3] for coal-water
slurry in a much smaller pipe. One probable reason for this
discrepancy is the coupling of the phases. The interaction
a) between the two dispersed phases was not accounted for. In
other words, the only interphase interactions were between the
dispersed and continuous phases. In a dense slurry, the
collisions among fine and medium sands should be expected to
be important, and ignoring such phenomenon might be
erroneous. Nonetheless, Fig. 5 shows that the bimodal (fine and
medium sand) mixture is distributed more uniformly than
monosized mixture, especially near the top of the pipe. The
relatively uniform distribution of the bimodal mixture is due to
the stronger particle-particle interactions in that mixture. It is
likely that the fine particles which could be thought of as
occupying the ‘void’ spaces between the medium particles
reduced the average distance between particles, thereby
b) contributing to particle-particle interactions. The iso-contour of
Figure 4: a) Normalized sand velocity and b) concentration volume fraction for individual dispersed phases, shown in Figs.
distributions; experimental data by Hashemi et al (Cases 2-4 in 6(a, b), indicate that the fine is more uniformly distributed than
Table 2) the medium sand. Consequently the bimodal mixture is more
In most computations the dispersed phase is treated as uniformly distributed than the monosized mixture, as shown in
a monosize mixture, i.e. a single particle size is used to Figs. 6(c, d).
represent the entire size distribution. For many slurry flow tests, For the bimodal test case, the frictional pressure
gradient (2.12 kPa/m) is 17.5% lower than the prediction

8 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


obtained from the SRC model (2.57 kPa/m). For the monosize The results of the bimodal flow test case seem to
test case where Matousek [8] reported experimentally suggest that turbulent kinetic energy is suppressed over the
determined pressure gradient, the results are almost identical: entire cross-section of the pipe; however, the magnitude is, on
2.183 was measured and the simulation provided a value of average, smaller than for the other test cases. This is consistent
2.198 kPa/m. with the relative uniformity observed for the concentration
distributions (see Fig. 5b).

a) Fine Sand b) Medium Sand

a)

c) Bimodal (C = 0.43) d) Monosize (C = 0.43)


Figure 6: Iso-contours of sand volume concentration

b)
Figure 5: a) Normalized sand velocity and b) concentration
distributions for multifluid with data from Matousek [8] (Cases
5-6 in Table 2)
Turbulence Kinetic Energy: In this section the turbulent
kinetic energy of slurry flows are compared to that expected for
single phase flow under identical conditions. The comparison is
shown as so-called turbulent modulation, TM ( k s / kl -1),
which is expressed as a per cent of the value expected for
single-phase flows. In the regions y/D > 0.5 and y/D < 0.12, the
presence of the dispersed phase suppresses the turbulent kinetic
energy irrespective of the bulk velocity, concentration, and pipe
Figure 7: Turbulence Modulation
diameter, as shown in Fig. 7. Outside these regions, there is
turbulence enhancement for C = 0.2 for both uc = 2 and 3 m/s,
CONCLUSIONS
and C = 0.43 for uc = 6 m/s. For the other flow conditions, there
For this study, several test cases exploring the effects
is suppression in the entire pipe. For the dense slurry flows that
of turbulence models, wall boundary conditions, interstitial
are considered here, particle-particle collisions are known to be
forces and bimodal mixture were performed to determine their
a source of particle kinetic energy. It should be noted that at
effect on numerical predictions for sand velocity and
higher bulk concentrations, the flows are relatively
concentration distributions, and frictional pressure drop. The
homogeneous as a result of the competing forces, including
predicted results were compared with the experimental
particle-particle collisions, turbulent dispersion, drag and
measurements of Hashemi et al [7] and Matousek [8].
viscous forces (as mentioned earlier).

9 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


The present work demonstrates that, in absence of any [7]. Hashemi, S. A., Sadighian, I., Shah, I. S., and Sanders,
specific information on the particle-wall interaction, using no R. S., 2011, “Solid Concentration Fluctuations in Highly
slip for the dispersed phase will provide (relatively) more Concentrated Liquid-Solids Pipe Flows,” Journal of
accurate predictions of the flow. The turbulence model used for Fluid Mechanics, submitted.
the continuous phase has an impact on the dispersed phase [8]. Matousek, V., 2002, “Pressure Drops and Flow Patterns
velocity and dispersed phase concentration distribution, with in Sand-Mixture Pipes”, Experimental Thermal and
the SST model being superior, especially in the near-wall Fluid Sciences, 26, 693-702.
region. Choice of turbulence model does not, however, have a [9]. Sanders, R. S., Schaan, J., Hughes, R., and Shook, C. A.,
significant impact on the prediction of the pressure gradient. In 2004, “Performance of Sand Slurry Pipelines in the Oil
flow conditions where there is possibility of stratification or Sands Industry,” The Canadian J. of Chem. Engineering,
insignificant slip among the phases, solving the same transport 82 (4), 850-857.
equation for the turbulence field will give better prediction. [10]. Gillies, R. G., Shook, C. A., and Xu, J., 2004,
The use of the Lopez model to compute the turbulence “Modelling Heterogeneous Slurry Flows at High
dispersion force was shown to provide accurate results for the Velocities,” The Canadian J. of Chemical Engineering,
dispersed phase velocity and concentration distributions, as well 82, 1060-1065.
as pressure gradient predictions, provided the ‘optimal’ [11]. Wilson, K. C., Sanders, R. S., Gillies, R. G., and Shook,
coefficient value is used – which is likely to vary from system to C. A., 2010, “Verification of the Near-wall Model for
system. For the horizontal slurry flows described here, one can Slurry Flow,” Powder Technology, 197, 247-253.
ignore the effects of wall lubrication and lift forces; however, [12]. Ling, J., Skudarnov, P. V., Lin, C. X., and Ebadian, M.
this is unlikely to be the case for all situations [11]. Better A., 2003, “Numerical Investigations of Liquid–Solid
predictions can be obtained for bimodal flows if particle- Slurry Flows in a Fully Developed Turbulent Flow
particle interactions between the dispersed phases can be Region”, Inter. J. of Heat and Fluid Flow, 24, 389–398.
accounted for. In general, a very good agreement was observed [13]. Gidaspow, D., 1994, “Multiphase Flow and Fluidization:
between the predicted and the compared experimental results. Continuum and Kinetic Theory Descriptions”, Academic
ACKNOWLEDGMENTS Press, New York.
The support of the Natural Sciences and Engineering [14]. Antal, S., Lahey, R., and Flaherty, J., 1991, “Analysis of
Research Council of Canada (NSERC), Canadian Natural Phase Distribution in Fully-developed Laminar Bubbly
Resources Limited, Fort Hills LLP, Nexen, Inc., Saskatchewan Two Phase Flow”, Int. J Multiphase Flow, 635.
Research Council, Shell Canada Energy, Syncrude Canada Ltd [15]. Burns, A. D. B., Frank, T., Hamill, I., and Shi, J-M,
and Total E&P Canada through the NSERC Industrial Chair in 2004, “Drag Model for Turbulent Dispersion in Eulerian
Pipeline Transport Processes (RSS) is gratefully acknowledged. Multiphase Flows”, 5th International Conference on
Multiphase Flow, ICMF, Yokohama, Japan.
REFERENCES [16]. Lopez de Bertodano, M., 1991, “Turbulent Bubbly Flow
[1]. Krampa-Morlu, F. N., Bergstrom, D. J., Bugg, J. D., in a Triangular Duct”, Ph.D. Thesis, Rensselaer
Sanders, R. S., and Schaan, J., 2004, “Numerical Polytechnic Institute, Troy, New York.
Simulation of Dense Coarse Particle Slurry Flows in a [17]. Rhie, C. M. and Chow, W. L., 1983, “Numerical Study of
Vertical Pipe”, 5th International Conference on the Turbulent Flow Past an Airfoil with Trailing Edge
Multiphase Flow, ICMF’04, Yokohama, Japan. Separation”, AIAA Journal, 21 (11), 1525-1535.
[2]. Lin, C. X. and Ebadian, M. A, 2008, “A Numerical [18]. Barth, T. J. and Jesperson, D. C, 1989, “The Design and
Study of Developing Slurry Flow in the Entrance Region Application of Upwind Schemes on Unstructured
of a Horizontal Pipe”, Comput. & Fluids, 37, 965–974. Meshes”, AIAA Paper 89-0366.
[3]. Chen, L., Duan, Y., Pu, W., and Zhao, C., 2009, “CFD [19]. Menter, F. R., 1994, “Two-equation Eddy-viscosity
Simulation of Coal-Water Slurry Flowing in Horizontal Turbulence Models for Engineering Applications”,
Pipelines”, Korean J. Chem. Eng., 26(4), 1144-1154. AIAA-Journal, 32 (8), 1598 - 1605.
[4]. Ekambara, K., Sanders, R. S., Nandakumar, K., and [20]. Lun, C. K. K., Savage, S. B., Jeffrey, D. J., and
Masliyah, J. H., 2009, “Hydrodynamic Simulation of Chepurnity, N., 1984, “Kinetic Theory for Granular
Horizontal Slurry Pipeline Flow Using ANSYS-CFX”, Flow: Inelastic Particles in Couette Flow and Slightly
Ind. Eng. Chem. Res., 48, 8159–8171. Inelastic Particles in a General Flow Field”, Journal of
[5]. Bossio, B. M., Blanco, A. J., and Hernandez, F. H., Fluid Mech., 140, 223-235.
2009, “Eulerian-Eulerian Modeling of Non-Newtonian
Slurries Flow in Horizontal Pipes”, Proc. of the Fluids
Eng. Div. Sum. Mtng., FEDSM2009, Vail, CO, USA.
[6]. Lahiri, S. K. and Ghanta, K. C., 2010, “Slurry Flow
Modelling by CFD”, Chemical Industry & Chemical
Engineering Quarterly, 16 (4), 295-308.

10 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


APPENDIX A

MATHEMATICAL MODEL

A.1 Governing Equations


The mathematical governing equations solved by ANSYS CFX are:
Continuity Equation:
w
U i D i  ’ x U i D i u&i 0 (1)
wt
Momentum Equation:
&
§ wul &
D l Ul ¨
& ·
& &

 ul x ’ul ¸ D l ’p  ’ x D l Pl ’ul  (’ul )T  M (2)
© wt ¹
&
§ wu & & ·
&
&
&
D s U s ¨ s  u s x ’u s ¸ D s ’p  ’ x D s P s ’u s  (’u s )T  ’ x W s  M  D s U s  Ul g
w
(3)
© t ¹
Volume Fraction Equation:
l ,s o l ,s

¦U ¦D
1
’ ˜ (D i U i u i ) 0 and i 1 (4)
i i i
&
where α is the concentration or volume fraction, index i is for either solid (s) or liquid (l), u is the velocity vector for respective phase,
ρ is the density, g is acceleration of gravity, p is the thermodynamic pressure which is the sum of dynamic, static and turbulent
fluctuation components, M is the sum of interphase forces, P is effective viscosity which is the sum of the dynamic viscosity and eddy
&
or turbulent viscosity, and W s is the solid stress tensor. It is worth commenting that, in two-fluid approach, it is assumed that both
fluids/phases share the same pressure field. In the subsequent sections, model equations for eddy viscosity, solid stress tensor, and
interphase forces are discussed.
A.2 Closure Models
A.2.1 Turbulence Models
For the present study, the continuous phase is assumed to be turbulent and therefore, effective viscosity is computed using
two-equation turbulence models, the standard linear-eddy-viscosity k-H model and shear stress transport which hereafter referred to as
k-H and SST, respectively. The general expression for effective viscosity, P is given as:
Pl Plm  Plt (5)
where m and t denote, respectively, dynamic and eddy components of the viscosity. The k-H assumes that, the turbulence viscosity is
linked to the turbulence kinetic energy, k and dissipation, H via the relation:
k2
Plt CP U (6)
H l
where CP is a constant of value 0.09 and all other variables have their usual meanings. The values of k and H are computed directly
from differential transport equations for the turbulence kinetic energy and turbulence dissipation rate which are given in Eqns. (7) and
(8).
&
§ wk w ul k · w §¨ §¨ P t · wk ·¸
D l U l ¨¨  ¸ Dl
¸ P  l ¸  D l Pk  U l H (7)
© wt wxi ¹ wxi ¨© ¨© V k ¸¹ wxi ¸¹
&
§ wH w ul H · w §¨ §¨ P lt ·¸ wH ·¸ H
¨
D l Ul ¨  ¸ Dl
¸ P   D l cH 1 Pk  cH 2 U l H (8)
¨ ¨ ¸
V H ¹ wxi ¹ ¸
© wt wxi ¹ wxi © © k

where cH1, cH2, Vk and VH are constants with values 1.44, 1.92, 1.0 and 1.3, respectively, Pk is the production term which is responsible
for the generation of the turbulent kinetic energy and is a function of the mean velocity gradient.

11 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


The equations solved for SST model are given as:
&
§ wk w ul k · w §¨ §¨ Plt ·¸ wk ·¸
¨
D l Ul ¨  ¸ Dl P   D l Pk  EU l kZ (9)
© wt wxi ¸¹ wxi ¨© ¨© V k1 ¸¹ wxi ¸¹
&
§ wZ w ul Z · w §¨ §¨ Plt ·¸ wZ ·¸ Z V wk wZ
D l U l ¨¨  ¸ Dl
¸ P   D l O Pk  U l E1Z 2  2(1  F ) U l Z 2
¨ ¨ ¸
V Z1 ¹ wxi ¹ ¸ Z wxi wxi
© wt wxi ¹ wxi © © k
(10)
Unlike Eqn. (6), the eddy viscosity is computed from:
OU k
Plt (11)
max(OZ , JF1 ) l
where specific dissipation rate, Z = H/k, respectively, Vk1, VZ, VZ, E, E1 and O are constants with values 2.0, 2.0, 0.856, 0.09, 0.0828
and 5/9, J is the invariant measure of the strain rate, and the blending factors (F and F 1) are given as:
§ § k 500Q · 4 UkV Z 2 · § 2 UV Z 2 wk wZ ·
F tanh(I 4 ); I min ¨ max¨ , ¸, ¸; C
DkZ max¨¨ ,1.0 u 10 10 ¸¸
¨ ¨ EZ y y 2Z ¸ C 2 ¸
Z w w
© © ¹ DkZ y ¹ © xi xi ¹
(12)
§ 2 k 500Q ¸ ·
F1 tanh(I12 ); I1 max ¨ , (13)
¨ EZ y y 2Z ¸
© ¹
For boundary conditions, the turbulent kinetic energy, k is computed from k = 1.5Iuc2 and dissipation rate, H from H = CP U k2/(PPo);
where I is turbulent intensity in decimal value and Po is eddy viscosity ratio (Pt /P) for the liquid. For homogeneous computations, the
mixture properties will be used.
A.2.2 Kinetic Theory of Granular Flow
In Eqn. (3), the stresses due to both collision and kinetic momentum transfer or inter-particle collisions is accounted for via
&
the solid stress tensor, W s and is given below:
& & &
W s ’ps  *s\  K s\ (14)
where ps is solid pressure which is momentum transport due to particle velocity fluctuation and particle interaction, Γs is shear viscosity
as a result of translational motion and collision of the particles, Ks is bulk viscosity accounting for the resistance of the solid body to
&
dilation, and \ is strain rate vector. Based on the idea of kinetic theory of gasses, the inelastic collision of particles can be taken into
&
account. Thus, the solid stress is assumed to depend on the solid phase granular temperature, T u cs / 3 , coefficient of restitution, e
taken here as 0.9, and particle diameter, d. Based on an isotropic assumption, local generation and dissipation of fluctuation energy is
equal, the relation between the solid stress tensor and granular temperature is:
&
& wu s
Ws
wx j
J s; J s §4§ T
3 1  e 2 U sI s2 g oT ¨ ¨
¨d ¨ S
& ··
 divu s ¸ ¸
¸¸
(15)
© © ¹¹
where γs is the dissipation of fluctuating energy in the solid phase, g0 is the contact radial distribution function, and all other variables
have their usual meanings. There are several correlations available in ANSYS CFX for computing, g0, however, the present results are
based on the correlation by Gidaspow [13]:
g0
0.6 1  I s / I m 1 / 3
1
(16)
where Im is maximum packing volume fraction.
The expressions for solid pressure and bulk viscosity are given in Eqns. (17) and (18) (Lun et al [19]), respectively.
p s U sI sT  2U sI s2T 1  e g o (17)
T
U sI s2 g o d 1  e
4
Ks (18)
3 S
Meanwhile, the solids shear viscosity is given as the sum of collisional and kinetic contributions. There are several correlations for
each contribution. However, ANSYS CFX implements only the collisional contribution since the correct form of the expression is
widely agreed on, and consequently, in the present work the kinetic contribution is omitted. Therefore, the collisional solid shear
viscosity is given as:

12 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


T
U sI s2 g o d 1  e
4
*s,col (19)
5 S
It is important to note that, due to singularity of some of the above mentioned equations, the numerical implementations might be
different from the expression given here, especially using limiting values in some situations.
A.2.3 Interphase Forces
The interphase momentum fluxes, M become:
M = MD + ML + MWL + MTD + Mex (20)
where subscripts D, L, WL, TD and ex, respectively, correspond to drag, lift, near-wall lubrication, turbulence dispersion, and et cetera
forces including virtual mass transfer which is ignored when the dispersed phase is solid. Another force that is ignored is the particle
induced turbulence on the continuous phase as a result of wake forming behind the particles despite it can be significant for large
particles. The drag force, MD is given as:
1 & & & &
C D U l I s u l  u s u l  u s
3
MD (21)
4 d
The correlations given by Gidaspow [13] were used to compute the drag coefficient, CD. The lift force, ML which is due to the liquid
shear is expressed as a function of the slip velocity and the curl of the liquid phase velocity:
& & &
M L CL Ul Is ul  u s u ’ u ul (22)
where CL is a lift coefficient and a value of 0.5 is used in the present study. The correlation given by Antal et al [14] is used for wall
lubrication force, MWL:
ªC C º & & 2
M WL  max « 1  2 ,0» U l I s u l  u s n w (23)
¬ d y w ¼
where yw is the distance to the nearest wall, nw is the unit normal pointing away from the wall, C1 = -0.01 and C2 = 0.05.
Two expressions were used for turbulent dispersion forces: Favre average drag model (FAD) and Lopez de Bertodano model.
The FAD expression (Burns et al [15]) is given as:
C Pt & & § ’I s ’Il ·
CTD D l I s ul  u s ¨¨
5
M TD  ¸ (24)
¸
6 d © Is Il ¹
where Eqn. (24) include 0.9 as the turbulent Schmidt number and a value of 1.0 is used for constant CTD. In Lopez de Bertodano [16]
model, the dispersion force is a function of density, continuous phase turbulent kinetic energy, and gradient of volume fraction, and is
given as:
M TD CTDkl Ul ’Il (25)
Here, the values of CTD have several ranges and more empirical. This constant varies from 0.1 to 0.5 for medium or large size (>3 mm)
bubbles in the ellipsoidal particle regime. However, very large value up to 500 is required for flow regimes involving very small
bubbles or very small solid particles. In the present work, 0.15, 1, 10, and 50 are used.

13 Copyright © 2012 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 01/29/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like