You are on page 1of 14

Numerical Simulation of a Semi-Industrial Scale CFB Riser using

Coarse-Grained DDPM-EMMS Modelling

Muhammad Adnan,1,2 Nan Zhang ,1,2* Fangfang Sun1,3 and Wei Wang1,2
1. State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing, 100190,
P. R. China
2. University of Chinese Academy of Sciences, Beijing, 100049, P. R. China
3. School of Resources and Safety Engineering, China University of Mining & Technology, Beijing, 100083, P. R. China

In this work, a coarse-grained Dense Discrete Phase Model (DDPM) coupled with the Energy-Minimization Multi-Scale (EMMS) drag, was used to
simulate the hydrodynamics of a 3D semi-industrial scale Circulating Fluidized Bed (CFB) with Geldart A particles. The effect of particle numbers in a
parcel, an important parameter in a coarse-grained approach, was investigated. It was found that when parcel diameter falls into the size range of
clusters, simulation results are insensitive to particle number in a parcel and generally agree with the experimental data. Further, a comparison
between DDPM and Eulerian Two-Fluid Modelling (TFM) approaches was investigated, both coupled with EMMS and conventional homogenous
drag. The simulation results suggested that the effect of meso-scale structures on drag force needs to be considered in both modelling approaches.

Keywords: circulating fluidized bed, CFD, dense discrete phase model, coarse-grained simulation

INTRODUCTION particle phase is solved by tracking individual particles.[18–21] This


approach can resolve the PSD naturally due to the presence of a

C
irculating Fluidized Bed (CFB) reactors are used in many Lagrangian discrete phase, but it is normally applicable only to
industrial applications such as coal gasification, coal small scale units due to the presence of a huge number of particles
combustion, Fluid Catalytic Cracking (FCC), and Fischer- in realistic systems.
Tropsch synthesis (FT). In these reactors, gas-solid flow features In recent years, development in Eulerian-Lagrangian discrete
the formation of complex meso-scale structures, e.g. clusters, over phase approaches includes the establishment of a family of coarse-
scales ranging from a single particle to global reactor.[1–4] These grained models, such as Dense Discrete Phase Model (DDPM),[22]
meso-scale structures have great impact on the performance of Multiphase Particle-in-Cell (MP-PIC) model,[23] and Coarse-
CFB reactors, in terms of hydrodynamics, heat and mass transfer, grained CFD-DEM model.[24,25] All these modelling approaches
as well as reaction kinetics,[5–7] and thus have attracted many track parcels instead of real particles, which reduces the number of
researchers to explore these effects by using Computational Fluid computational particles in simulations and accelerates the
Dynamics (CFD). computational speed.[26] The major difference between these
In CFD simulations, coarse grids are normally necessary for coarse-grained modelling approaches lies in their particle-particle
modelling large industrial systems. These coarse grid simulations interaction treatment.[25]
cannot resolve meso-scale structures accurately when coupled with In DDPM and MP-PIC modelling approaches, the interaction
conventional homogenous drag force models.[4,8] To resolve this between parcels is calculated implicitly by using the Kinetic
issue, various methods have been reported in the literature by Theory of Granular Flow (KTGF) or particle phase normal
considering the effects of unresolved, sub-grid, meso-scale struc- stresses.[27] The roots of DDPM and MP-PIC modelling approaches
tures.[7,9,10] One such example is the Energy-Minimization Multi- are similar to each other. A much bigger time step can be used in
Scale (EMMS) modelling approach proposed by Li and Kwauk[4] these approaches due to the implicit particle-particle interaction
and another is the filtered approach as developed by Sundaresan treatment. In particular, the time step for the particle phase could
and his colleagues.[11–13] Now it has been widely accepted that be the same as for the fluid phase, which accelerates the
such meso-scale modelling does give better predictions than computational speed and makes these approaches suitable for
using conventional homogenous drag models.[7,9,14,15] industrial applications.[28] While in the coarse-grained CFD-DEM
Generally modelling approaches for gas-solid flow in dense CFB modelling approach, the interaction between parcels is modelled
reactors can be classified into two categories: (1) Eulerian- explicitly by using the soft sphere model,[20] where the normal and
Eulerian Two Fluid Model (TFM), and (2) Eulerian-Lagrangian the tangential forces are calculated using the spring-dashpot
CFD-Discrete Element Model (CFD-DEM). In the TFM modelling
approach, gas and solid phases act as interpenetrating con-
tinua.[16,17] Mass and momentum conservation equations are * Author to whom correspondence may be addressed. E-mail address:
solved for each of the phases separately. The disadvantage of this nzhang@ipe.ac.cn
approach is the larger computational time required when there is Can. J. Chem. Eng. 96:1403–1416, 2018
particle size distribution (PSD). That restricts its applications to © 2017 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.23071
large industrial units. In the CFD-DEM modelling approach, gas Published online 13 November 2017 in Wiley Online Library
phase equations are solved similarly to as in TFM, while the (wileyonlinelibrary.com).

VOLUME 96, JUNE 2018 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 1403
model and the friction collisional laws. A smaller time step is NUMERICAL MODELLING
needed in the coarse-grained CFD-DEM modelling approach for
particle phase calculations due to the explicit particle-particle TFM Approach
interaction treatment, because higher values could result in
In the TFM approach, the solid phase is treated as a continuum and
serious overlapping of parcels and thus numerical instabilities
its properties are modelled similarly to the gas phase. The
during simulations.[25] Although the computational speed of the
governing equations for gas and solid phases are derived from the
coarse-grained CFD-DEM modelling approach is faster than the
concept of averaging.[35,36] Interaction forces between particles in
conventional CFD-DEM model due to tracking of coarse-grained
the TFM approach are modelled by using KTGF.[37] The mass and
parcels, it is still formidable to industrial applications due to the
momentum conservation equations for each phase can be defined
explicit particle-particle interaction treatment.
as follows.
Adamczyk et al.[27] simulated particle transport phenomena in
Mass conservation equation for gas phase:
an experimental scale 3D CFB riser using the DDPM modelling
approach with Geldart B particles. Simulation results showed that
@   
the DDPM can be used for modelling the particle transport process eg rg þ r  eg rg~
vg ¼ 0 ð1Þ
in a CFB. The effect of the geometrical simplifications on predicted @t
results was also discussed in their study. After validating the Momentum conservation equation for gas phase:
models from experimental scale, they further implemented them
to large-scale CFB boiler combustion modelling.[29] User defined @   
functions (UDFs) were used to control the boiler mass load, eg rg~
v g þ r  eg rg~ v g ¼ eg rp þ r  t g þ eg rg~
v g~ g
@t  
particle recirculation process, and interphase momentum ex- þKgp ~ vp  ~
vg ð2Þ
change coefficient. Predicted results showed the applicability of
the DDPM modelling approach for combustion simulations.
Mass conservation equation for particle phase:
Recently, Farid et al.[30] simulated a 340 MWe CFB boiler using
the DDPM modelling approach. The effect of coal feeder positions
@   
and coal feeding rates were investigated. Results were presented in ep rp þ r  ep rp~
vp ¼ 0 ð3Þ
@t
terms of distribution of gas temperature, solid volume fraction,
pressure, and mass fractions of combustion products, which were Momentum conservation equation for particle phase:
compared with operating data collected from a boiler. Li et al.[28]
simulated two different CFB risers with simplified 2D domains @   
using the MP-PIC modelling approach coupled with EMMS and ep rp~
v p þ r  ep rp~
v p~
v p ¼ ep rp  rpp þ r  t p
@t  
homogenous drag models for Geldart A particles. They concluded
þep rp~
g þ Kgp ~vg  ~
vp ð4Þ
that the MP-PIC approach correctly predicted the hydrodynamics
phenomena with the EMMS drag but failed with the homogenous
drag. Nikolopoulos et al.[25] simulated the hydrodynamics of a where the subscripts g and p represent the gas and particle phases,
pilot-scale 3D CFB carbonator using the coarse-grained CFD-DEM e represents the volume fraction, r is the density, ~v is the velocity
model (Geldart A particles). The results obtained with the EMMS vector, p is the gas pressure shared by both gas and solid phases, pp
drag showed better agreement than with homogenous drag. Lu is the solid pressure term, t represents the stress tensor, ~ g is the
et al.[31] developed a coarse-grained discrete particle method based gravitational acceleration, and K represents the interphase
on the EMMS model (DPM-EMMS). They first simulated fluidiza- momentum exchange coefficient.
tion phenomena in a fixed bed to validate their models with an The stress tensor for gas phase, t g , which represent the viscous
original DPM model. Then they simulated a lab-scale riser with effects can be calculated as follows:[38]
different geometries and operating conditions. Their results    
2
agreed well with the experiments. t g ¼ eg mg rv g þ rv g þ eg lg  mg r  ~
~ ~ T
vgI ð5Þ
From the above discussion of coarse-grained modelling 3
approaches and literature survey, it can be concluded that the
DDPM approach can be used for industrial applications due to its where mg represents the gas viscosity, lg is the bulk gas phase
acceptable computational cost and easy handling of PSD. viscosity, and I is the unit stress tensor.
Conventional homogenous drag models may be suitable for For solving the multiphase transport equations in the TFM
Geldart B particles, while for Geldart A particles, a drag model approach, closure laws for the solid stress tensor and interphase
with consideration of meso-scale structures is needed.[8,32] In momentum exchange coefficient are needed to be defined. Details
addition, previous studies of the DDPM approach have shown of all the closure laws used in the present study for the solid stress
that more particles in a parcel can speed up simulation,[33,34] but tensor and interphase momentum exchange coefficient are given
with poorer accuracy. Therefore, the selection of appropriate in the next section.
number of particles in a parcel is important. On the basis of the The disadvantage of using the TFM modelling approach is that,
above two concerns, the objectives of the present work are, firstly, when considering the PSD, additional mass and momentum
to integrate the EMMS drag model with the DDPM approach for conservation equations are required for each granular phase, which
the case of Geldart A particles, and secondly, to analyze the effect is numerically intensive for large-scale industrial applications. To
of particle number in a parcel on CFB hydrodynamics. Moreover, a circumvent this problem, the DDPM modelling approach in
general comparison between DDPM and TFM approaches, both commercial Fluent software has been recently introduced.[22,27,29,30]
coupled with EMMS and homogenous drag models, is also
DDPM Approach
presented in this work to explore the capabilities of the DDPM
modelling approach for CFB simulations. All the results are In the DDPM approach, gas phase equations are solved
validated with the experimental data. similarly to the Eulerian TFM approach, while the particle

1404 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 96, JUNE 2018
phase is modelled by using Newton’s equation of motion in a solid volume fraction is assigned to Eulerian coordinates which
Lagrangian reference frame. This approach can resolve the PSD can be calculated as follows:
naturally due to the presence of a Lagrangian discrete phase and is
applied for large industrial-scale reactor design applications.[27,29,30] eg ¼ 1  ep ð11Þ
In this approach, the particle equation of motion is not solved for  
individual particles. The solver tracks groups of particles, usually Acceleration due to drag force F D ~vg  ~
v p is defined as follows:
called parcels. Each parcel consists of several numbers of real
particles which have the same mass, velocity, position, etc. The   3 mg  
FD ~vg  ~
v p ¼ CD Rep vg  ~
~ vp ð12Þ
parcel properties such as velocity and position are calculated from the 4 2
rp dp
particle equation of motion on Lagrangian coordinates and then
interpolated back to Eulerian coordinates, which gives mean solid where mg is the gas viscosity, dp is the particle diameter, CD
velocity and volume fraction in each Eulerian cell. In the DDPM represents the drag coefficient, and Rep is the particle Reynolds
approach, particle-particle interaction force is also modelled with number which can be expressed as follows:
KTGF on Eulerian coordinates and then it is introduced back to the
particle equation of motion as the source by using interpolation  
rg dp ~ vp
vg  ~
operators.[23,39] The purpose of utilizing the interpolation operators Rep  ð13Þ
in the DDPM approach is to resolve the inter-particle stresses which mg
are difficult to calculate for each particle in the dense phase flow
   
simulations. Governing equations for DDPM approach are as follows. After obtaining FD ~vg  ~
v p from Equation (12), Kgp ~vg  ~
v p in
Mass conservation equation for gas phase: Equation (7) can be calculated as follows:

@      XN  
eg rg þ r  eg rg~
vg ¼ 0 ð6Þ Kgp ~vg  ~
vp ¼ mi F D ~vg  ~
v pi =V cell ð14Þ
@t
i¼1

Momentum conservation equation for gas phase:


where N ¼ N p  N parcel is total number of particles in a cell, mi is
mass of particle i, and ~v pi represents velocity of particle i.
@   
eg rg~
v g þ r  eg rg~ v g ¼ eg rp þ r  t g þ eg rg~
v g~ g The particle velocity and position calculated from Equation (8)
@t   and Equation (9) depend on the evaluated solid stress tensor (on
þKgp ~ vp  ~
vg ð7Þ Eulerian coordinate).
The solid phase stress tensor, t p , which accounts for particle-
where eg is the gas volume fraction, rg is the gas density, ~ v g is particle interactions can be calculated as follows:[38]
velocity of the gas phase, t g represents the stress tensor of the gas
phase, ~g is the gravitational acceleration, and Kgp represents the    
2
t p ¼ pp I þ ep mp r~ v p þ ep lp  mp r  ~
v p þ r~
T
interphase momentum exchange coefficient from solid to gas vpI ð15Þ
3
phase per unit volume of cell.
Particle equation of motion equates the particle inertia with the
where pp represents solid pressure, I is unit stress tensor, mp
forces acting on each particle as defined by the following:
represents particle dynamic viscosity, lp is the bulk viscosity, and
  ~
v p represents the average velocity vector of particle phase. The
d~
vp   ~g rp  rg rp tp closure law equations for granular pressure,[40] granular dynamic
¼ FD ~vg  ~
vp þ  r ð8Þ
dt rp rp rp viscosity,[17] and granular bulk viscosity,[40] based on the KTGF
  are given in Table 1. The dynamic viscosity of a granular phase is a
where ~v p is the particle
 velocity,
 FD ~vg  ~ v p is the acceleration further combination of three viscosity components, i.e. kinetic,[41]
due to drag force, ~ g rp  rg =rp represents acceleration due to collisional,[41] and frictional (see Table 1).[42]
gravity, rp is the particle density, and t p is the particle phase stress The granular temperature term, Qp , in constitutive equations
tensor which accounts for the interaction between particles on a (for granular pressure and granular dynamics viscosity) is
Eulerian grid. proportional to the kinetic energy of the particles random motion
Integrating the particle equation of motion gives the particle which can be defined as follows:
position ~x p which can be expressed as follows:
1
d~
xp Qp ¼ ~v p;i~
v p;i ð16Þ
¼~
vp ð9Þ 3
dt
where ~v p;i represents the ith component of the fluctuating solid
velocity. The transport equation derived from KTGF can be
After getting the particle position from Equation (9), we can expressed as follows:[43]
calculate the solid volume fraction in a given cell as follows:

3 @   
N p N parcel V p rp ep Qp þ r  rp ep~
v p Qp
ep ¼ ð10Þ 2 @t
V cell  
¼ ðpp I þ t p Þ : r~
v p þ r  kQp rQp  gQp þ fgp ð17Þ
where N p is the number of particles in a parcel, N parcel is the 
number of parcels in a cell, V p is volume of a particle, and V cell is where pp I þ t p Þ : r~ v p represents energy generation due to
volume of a cell. To obtain the void fraction of a gas, the calculated solid stress tensor, kQp rQp is diffusion energy term, g Qp is the

VOLUME 96, JUNE 2018 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 1405
Table 1. Closure laws used for solid stress tensor based on KTGF[37]

Parameters and closure models Closure law equations


[40]
Solids pressure by Lun et al. pp ¼ ep rp Qp þ 2rp ð1 þ ess Þep 2 go Qp

Solids viscosity by Gidaspow[17] mp ¼ mp;kin þ mp;col þ mp;fr


pffiffiffiffiffiffiffi
10r d p Qp p 2
[41]
Kinetic viscosity by Gidaspow et al.
mp;kin ¼ 96eppð1þepp Þg 1 þ 45go ep ð1þepp Þ
o

[41]  = 1
Collisional viscosity by Gidaspow et al. Q
mp;col ¼ 45 ep rp d p go ð1 þ epp Þ pp
2

Frictional viscosity by Schaeffer[42] mp;fr ¼


pp sin f
pffiffiffiffiffi
2 I2D
 =
Solids bulk viscosity by Lun et al.[40]
1
Q 2

lp ¼ 43 ep rp d p go ð1 þ epp Þ pp
pffiffiffiffiffiffiffiffi qffiffiffiffiffi
Diffusion coefficient by Gidaspow et al.[41] 150rp d p ðQpÞ 2 Q
kQp ¼ 384ð1þepp Þgo 1 þ 65ep go ð1þepp Þ þ 2rp ep 2 d p ð1 þ epp Þgo pp

Collisional dissipation by Lun et al.[40] g Qp ¼


12ð1e2pp Þgo
pffiffiffi rp e2p Qp=
3
2

dp p

Energy exchange term by Gidaspow et al.[41] fgp ¼ 3K gp Qp

Radial distribution function by Lun et al. [40]  1 =


1
ep
go ¼ 1  ep;max
3

 
collisional dissipation of energy, and fgp is the energy exchange 3 ep eg rg ~ v g  2:65
vp  ~
between fluid and solid phases. The formulation of closure laws Kgp ¼ CD eg HD ð20Þ
4 dp
for diffusion energy coefficient, collisional dissipation energy, and
energy exchange between fluid and solid phases are defined in
Table 1. If the granular temperature equation is solved as a partial where CD is drag coefficient and HD is fitting function. The drag
differential equation then the convective term has to be calculated coefficient CD can be defined as follows:
from the velocity field passed down from Lagrangian particle
8 h i
< 24 1 þ 0:15eg Rep 0:687
tracks. This cannot run stably for the present simulation cases. To
overcome this problem Fluent provides alternative algebraic Rep < 1000
CD ¼ eg Rep ð21Þ
treatment for Equation (17), where the convective and diffusive :
0:44 Rep  1000
terms are neglected. The simplified form of granular temperature
can be represented as follows:

ðpp I þ t p Þ : r~
v p  g Qp þ fgp ¼ 0 ð18Þ
Drag Force Correction
Drag force correction based on the EMMS/matrix model[9] is
Interphase exchange coefficient, Kgp , which corresponds to
implemented in the present work. The EMMS model resolved the
interaction between gas and solid phases, is one of the most
heterogeneous gas-solid flow into a particle-rich, dense phase and
important parameters for gas-solid flow applications. In the
a gas-rich, dilute phase. Eight parameters were defined for the
present work two different correlations for the interphase
dense and dilute phases (Ugc, Usc, egc, Ugf, Usf, egf, f, dcl) in the
momentum exchange coefficient were tested and compared.
original version of the EMMS model,[4] which were closed by a set
The first one is the interphase momentum exchange coefficient
of conservation equations and a specific stability condition, which
based on the work of Gidaspow et al.,[41] which assumes
has been proven to be beyond the single criterion of non-
homogenous flow conditions in each computational cell and is
equilibrium thermodynamics.[44] Extension to this model for
defined as follows:
transient flow applications was done by Yang et al.[7] and Wang
8   and Li,[9] where three additional acceleration terms were
>
> 3 ep eg rg ~ v g  2:65
vp  ~
>
> eg for eg > 0:80 introduced (ac, af, ai), and were further solved through a two-
< 4 CD dp
  step scheme EMMS/matrix model to obtain structure-dependent
Kgp ¼  
>
>
> ep 1  eg mg rg ep ~ vg
vp  ~ drag. In the EMMS/matrix model, ten variables were defined for
> 150
: þ 1:75 for eg  0:80 the dense and dilute phases (Ugc, Usc, egc, f, dcl, ac, Ugf, Usf, egf, af),
eg d2p dp
while only seven nonlinear equations based on the hydrodynam-
ð19Þ ics were established for these variables. Therefore a stability
condition is needed to close the variables, that is, the minimization
The second one is the interphase momentum exchange of energy consumption needed for suspension and transportation
coefficient based on the work of the EMMS model (explained in of particles per unit mass of particles, Nst (Nst ! min).[4]
the next section),[9] which considers the effects of heterogeneity in The momentum balance equations for dense phase, inter-phase,
each computational cell and is defined as follows: and dilute phase in EMMS/matrix model can be defined as follows:

1406 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 96, JUNE 2018
where N st is energy needed for suspension and transportation of
 
3 1  egc    particles while N T is the total energy. The correlations for N st and
mc F c ¼ Cdc rg U 2slip;c ¼ 1  eg rp  rg ðac  gÞ ð22Þ N T can be defined as follows:
4 dp
"   #
3 f    rp  rg egf  eg
mi Fi ¼ Cdi r U 2 ¼ f eg  egc rp  rg ðai  gÞ ð23Þ N st ¼ Ug    f ð1  f ÞU gf g ð31Þ
4 dcl g slip;i rp 1  eg

3 1  egf    !
mf F f ¼ Cdf rg U 2slip;f ¼ 1  egf rp  rg ðaf  gÞ ð24Þ rp  rg
4 dp NT ¼ Ug  g ð32Þ
rp
where the inertial term for the interphase (ai) based on equal pressure
drop assumption (for dense and dilute phases) can be defined as All the related definitions and model parameters used in
follows: Equations (22–24) and Equations (26–29) are summarized in
    Table 2. The EMMS/matrix model follows the two-step scheme for
ð1  f Þ½ 1  eg ðac  gÞ  1  egf ðaf  gÞ the calculation of structure dependent drag force. In the first step,
ai ¼   þg ð25Þ
f eg  egc meso-scale structure parameters (egc, dcl) are calculated on the
global reactor scale by solving the set of conservation Equations
(22–24 & 26–29) and stability condition Equation (30) under the
The mass balance equations for gas and solid phases can be
specified operating conditions (Ug0,Gs). Then ten local variables
defined as follows:
are reduced to eight variables (Ugc, Usc, Ugf, Usf, egf, f, ac, af). In the
second step, the remaining eight variables are calculated by
U g ¼ fU gc þ ð1  f ÞU gf ð26Þ
solving a set of conservation equations locally on each grid scale
with the minimization of Nst, which can be further simplified by
U s ¼ fU sc þ ð1  f ÞU sf ð27Þ considering egf ! emax and af ! g. Finally, the deterministic
solution of the remaining six unknown variables (Ugc, Usc, Ugf,
The overall voidage based on dense and dilute phase can be Usf, f, ac) is obtained by six conservation equations. For a detailed
defined as follows: description of the two-step EMMS/matrix scheme and its
algorithm, the reader is referred to Wang and Li.[9]
eg ¼ fegc þ ð1  f Þegf ð28Þ The structure dependent drag coefficient obtained from the
EMMS/matrix model approach is defined as follows:
The cluster diameter used for Geldart A particles in the EMMS/
matrix model follows its original definition and is inversely e2g
Kgp-EMMS ¼ ðfMc F c þ Mi F i þ ð1  f ÞMf Ff Þ ð33Þ
proportional to Nst, i.e. the energy consumption needed for U slip
suspension and transportation of particles and is defined as follows:
And heterogeneity index (HD ) can be calculated by the following
dcl ½U s =ð1  emax Þ  ðU mf þ emf U s =ð1  emf ÞÞg expression:
¼   ð29Þ
dp N st rp = rp  rg  gðU mf þ emf U s =ð1  emf ÞÞ
Kgp-EMMS
HD ¼ ð34Þ
Kgp-Wen&Yu
To satisfy the above nonlinear Equations (22–24 & 26–29) a
stability condition Nst ! min is assumed to specify which can be
defined as follows: where Kgp-EMMS is the interphase momentum exchange coefficient
obtained from EMMS/matrix drag,[9] and Kgp-Wen&Yu is the
    interphase momentum exchange coefficient obtained from the
N st U g 1  eg  fU gf egf  eg ð1  f Þ
¼   ¼ min ð30Þ Wen-Yu equation.[45]
NT U g 1  eg

Table 2. Summary of parameters and definitions used in the EMMS/matrix model[9]

Parameters Dense phase Dilute phase Inter-phase

Characteristic diameter dp dp d cl
Voidage egc egf 1f
Usc egc Usf egf
 
Superficial slip velocity U e
Uslip;c ¼ Ugc  esc Uslip;f ¼ Ugf  esf Uslip;i ¼ Ugf  scesc gf ð1  f Þ
Characteristic Reynolds number rg d p Uslip;c rg d p Uslip;f rg d cl Uslip;i
Rec ¼ mg Ref ¼ mg Rei ¼ mg
Standard drag coefficient C dc0 ¼ Re
24
c
þ Re3:6
c
0:313 C df0 ¼ Re
24
f
þ Re3:6
f
0:313 C di0 ¼ Re
24
i
þ Re3:6
i
0:313

4:65 4:65
Effective drag coefficient C dc ¼ C dc0 egc C df ¼ C df0 egf C di ¼ C di0 ð1  f Þ4:65
Number density 1egc
mc ¼ pd 3 =6
1egf
mf ¼ pd 3 =6 mi ¼ pd 3f =6
p p cl

Drag force on each particle or cluster pd 2 r pd 2 r pd 2cl rg


F c ¼ C dc 4 p 2g U2slip;c F f ¼ C df 4 p 2g U2slip;f F i ¼ C di 4 2 U2slip;i
Drag force in unit volume F dc ¼ mc F c F df ¼ mf F f F di ¼ mi F i

VOLUME 96, JUNE 2018 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 1407
Table 3. Fitted formula of heterogeneity index calculated from EMMS/matrix model with Ug ¼ 3.5 m/s, Iinventory ¼ 198 kg, Hinitial ¼ 1.93 m,
dp ¼ 6  105 m, and Gs ¼ 86.1 kg/(m2  s)

Fitting function HD ¼ AðRe þ BÞC =HD;max , where HD;max ¼ 3:04,


Range eg;mf ¼ 0:448, and eg;max ¼ 0:9997
h  i
eg;mf < eg  0:5 eg 0:507782
2 0:06956
A ¼ 0:7934 þ 1:1175e
B¼0
C¼0
0:5 < eg  0:6  50:25633
1
eg
A ¼ 0:05204 þ 1:9743 1 þ 0:53027
" e 0:50084 #1( eg 0:60902
1 )
g

B ¼ 0:14294  0:15047 1 þ e 0:0047


1  1 þ e 0:02736

 34:39893
1
eg
C ¼ 0:56038  0:58092 1 þ 0:54322

0:6 < eg  0:98  0:5868


A ¼ 429:90532  434:69389eg
e
g 1:02901
B ¼ 1:58454þ7:56707ðe e 2
g 1:02901Þ19:64414ð g 1:02901Þ

 0:163018
C ¼ 0:05377  0:05394eg
" eg 0:995945 #1( eg 0:999915
1 )
0:98 < eg  eg;max
A ¼ 0:36777 þ 5:13409 1 þ e 0:00447 1  1 þ e 0:00044

h eg 0:99862 i
0:5
B ¼ 0:0175 þ 0:71397e
0:00186

h eg 0:998632 i
0:5
C ¼ 0:32769  0:25491e
0:00969

eg < eg;mf or eg > eg;max HD ¼ 1

The Wen-Yu drag model is determined by the following Zhang et al.,[46] nearly 540 000 cell elements with average cell size
expression: of 0.1 m are used. Special attention is needed for mesh generation
  in DDPM simulations in that the mass capacity of particles in a cell
3 ep eg rg ~ v g  2:65
vp  ~ should not exceed its maximum packing limit (0.63).
Kgp-Wen&Yu ¼ CD eg ð35Þ The maximum cell mass capacity can be calculated by using the
4 dp
following correlation:
where drag coefficient (CD ) can be calculated from Equation (21).
For the present simulation cases, the fitting functions of HD
calculated from the EMMS/matrix model are taken from the work
of Zhang et al.[46] (see Table 3). Drag correction based on the
EMMS/matrix model is implemented through UDF. It is important
to note that the slip velocity calculation in Equation (35) for TFM
simulations is based on cell-averaged values (for both gas and
solid phases), while in DDPM simulations the slip velocity is
calculated by taking the velocity of each particle surrounded by
the gas phase instead of its cell-averaged value.

Geometry and Mesh


A semi-industrial scale CFB is selected for the present work, whose
details can be found in Herbert et al.[47] A schematic diagram of the
complete CFB unit is shown in Figure 1. This CFB unit consists of a
riser, cyclone, down-comer, and L-valve. Due to the limited
computation time, here only the riser part is simulated. The riser is
8.5 m high and its inner diameter is 0.411 m. This riser geometry
has been used in previous 3D, full-loop simulations with
simplified secondary air inlets configuration.[46] In the present
work, we kept the secondary air inlets configuration the same as
the experimental setup.[47] Figure 2 shows the 3D domain (left)
and mesh of the CFB riser (right). Based on the previous work of Figure 1. Schematic diagram of ETH CFB-riser.[45]

1408 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 96, JUNE 2018
Table 4. Boundary conditions and model parameters

Operation parameters Value

Riser diameter (m) 0.411


Riser height (m) 8.5
Gas density (kg/m3) 1.225
Gas viscosity (kg/(m  s)) 1.7894  105
Particle density (kg/m3) 1400
Particle diameter (mm) 60
Primary air inlet (m/s) 1.2
Secondary air inlet (m/s) 12.4
Outlet (Pa) 101 325
Wall specularity coefficient 0.6
Particle wall restitution coefficient 0.9
Wall reflection coefficient for 0.7 for both normal and
DDPM tangential
Initial bed height (m) 1.93
Bed inventory (kg) 198
Packing limit 0.63
Transition factor for DDPM 1

particles at an initial bed height of 1.93 m with volume fraction of


solid phase being 0.552, i.e. 198 kg. While in DDPM simulations
the patching of particles for 3D simulations is usually done by
using either surface injection or file injection options. In the
present investigation cases, we have patched the particles in the
CFB domain (at targeted mass of 198 kg) by using the surface
injection option available in the discrete phase injection panel.
Figure 2. 3D geometry of CFB riser (left) and mesh (right).
The parcels are injected from the inlet bottom surface of the riser.
Solid particles after patching in the CFB riser are fluidized due to
primary and secondary air flows, and re-circulated through the
mcell;max ¼ ep rp V cell ð36Þ side inlet to keep constant solids inventory. Such a scheme by
fixing solids inventory was called the CFBC mode of computa-
where ep is particle phase volume fraction, rp is density of particle, tion,[48] which was found to be the same as the scheme by fixing
and V cell is volume of cell. In DDPM simulations the mass capacity solids flux in terms of axial profile of voidage but different in terms
of a cell can be exceeded during the Lagrangian particle tracing.[29] of the time to reach steady state.
In this situation, the excess mass should be redistributed Pressure outlet boundary condition is specified at the top outlet
iteratively to its neighbouring cells. If the number of continuous with atmospheric pressure prescribed. At the wall boundary, no-
phase iterations used in the calculation procedure is too small then slip condition is prescribed for the gas phase, while for the solid
the masses evaluated on Eulerian and Lagrangian coordinates can phase the partial slip condition is used,[49] where the specularity
differ. So during the calculation procedure it is important to use a coefficient is taken as 0.6. For DDPM simulations, the wall
sufficient number of continuous phase iterations to keep the mass reflection boundary condition is also specified at the wall, where
difference between Euler and Lagrangian coordinates close to each its value is set to 0.7 for both normal and tangential directions.
other. To judge this situation during each time step a UDF function Detailed settings for simulations are given in Table 4.
can be used.[27,29,30] In the present study, we did not use the UDF
function for the evaluation of mass difference between Euler and Solution Procedure
Lagrangian coordinates due to considering the computational cost. The ANSYS Fluent1 solver is used for the DDPM and TFM
Based on the previous researchers’ experience and investiga- approaches. The phase coupled simple method is chosen for
tions,[27,29,30] we have used a sufficient number of continuous coupling between pressure and velocity. The second-order upwind
phase iterations (above 50) which keeps the mass difference scheme is used for the momentum conservation equation, while for
between Euler and Lagrangian coordinates in an acceptable range. the volume fraction the Quadratic Upwind Interpolation for
In our simulations the mass difference between Euler and Convection Kinetics (QUICK) scheme is used. The first order
Lagrangian coordinates is checked after completing the simu- implicit scheme is chosen for the transient formulation. The node-
lations by using Equation (36) in all simulation cases, and it is based averaging option is enabled in the numeric tab of the discrete
found to be less than 15 %. phase model. The purpose of enabling the node-based averaging was
to distribute the effects of discrete phase parcels to its neighbouring
Boundary Conditions and Simulation Setup mesh nodes, which in its original treatment are applied only to a
The CFB riser consists of inlets, an outlet, and wall boundaries. specific fluid cell that contains the parcel. The advantage of using
Primary air enters at the bottom part of the CFB riser while node-based averaging was to reduce the grid dependency of discrete
secondary air enters from the side inlets. The particles with phase simulations, since each parcel’s effect on the flow solver is
minimum fluidization voidage are initially patched up to a certain distributed more smoothly across its neighbouring cells.[38]
height according to the solid inventory in experiment in both TFM The distribution of any variable averaged on mesh nodes can be
and DDPM approaches. In TFM, a region is adapted for patching of defined as follows:

VOLUME 96, JUNE 2018 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 1409
of Kaufmann et al.[51] The numeric scheme for integration of
X  
wnode ¼ N in; parcel  w ~
xp  ~
k
x node wp ð37Þ particle equation of motion is chosen as trapezoidal and implicit.
k More details about the particle treatment options and its numeric
can be found in Fluent help.[50] The time step used for transient
where wnode represents accumulation of the particle variables on calculations in both modelling is 0.0005 s. Solid flux in the riser is
the node for all k parcels, wp is the particle variable, N in;parcel is the monitored at an elevation of 7 m. When the simulations reached
number in the parcel, and w is weighting function or kernel. quasi-steady state (Figure 3), we started the time-averaged
Several kernels are available in Fluent software,[50] such as nodes analysis. In the present work, all the cases were simulated for a
per cell, shortest distance, inverse distance, and Gaussian. In the total time of 22 s, where the last 10 s were considered for data
present study Gaussian kernel is chosen, which can be defined as analysis.
follows:
0  k 2 1

RESULTS AND DISCUSSION
  a3=2 ~xp  ~
x node 
w ~
xp  ~ exp@a A
k
x node ¼ ð38Þ Correlation for Particles per Parcel Based on Clusters
p Dx2
Particle number in a parcel, N p , which affects the accuracy of
results, is an important parameter for coarse-grained modelling
where Dx is the characteristic length scale of a cell containing
approaches.[28,34] Usually in the DDPM modelling approach, the
parcel and a is an additional parameter used to control the width of
standard treatment for particle number in a parcel is calculated by
the Gaussian distribution. Its value is taken as 6 based on the work
the ratio of parcel mass to the particle mass:[27,29,30,33,52–54]

 
mparcel N p  rp  16 pd3p rp  16 pd3parcel dparcel 3
Np ¼ ¼ ¼ ¼ ð39Þ
mparticle rp  16 pd3p rp  16 pd3p dp

In all previous studies, the parcels are only points of information


travelling through the domain. It is rather difficult to choose an
appropriate value, as a larger N p can get the simulations done
more quickly, while a lower value may give more reasonable
results. Considering that meso-scale structures are inherent in
CFB, we therefore assume that parcels are a kind of cluster in our
simulations, thus the particle number in a parcel can be
determined as follows:
 3
mparcel rp  16 pd3cl  ecl dcl
Np ¼ ¼ ¼  ecl ð40Þ
mparticle rp  16 pd3p dp

In Equation (40), dcl =dp is the ratio of cluster diameter to particle


diameter and ecl is the volume fraction of particles in a cluster, its
value is taken as 0.5 for the preliminary test. The schematic
diagram for the concept of coarse-grained particles (which reduces
the number of original particles in simulations) in two dimensions
based on Equation (40) is shown in Figure 4.

Figure 4. Representation of coarse-grained particle/cluster in which one


coarse-grained particle consists of several numbers of real particles where
the mass of the coarse-grained particle is mparcel ¼ Np  mp , mass of the
Figure 3. Solid flux monitored at 7 m height of CFB riser: (a) DDPM, (b) original particles is mp , and number of real particles in coarse-grained
TFM. particle is Np .

1410 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 96, JUNE 2018
Table 5. Solid flux results at different d cl =d p values

Gs (kg/(m2  s))

Number of particles in a Error


d cl =d p parcel, Np Experiment Simulation (%)

55 8.3  104 54 37.28


65 1.4  105 86.1 58 32.64
95 4.3  105 89 3.37
200 4.0  106 73 15.21

Parametric Analysis for DDPM Modelling Approach


The effect of dcl =dp was investigated, where the range of ratio, 55–
95, was recommended from the value of cluster diameter in Wang
and Li.[9] All cases were simulated with the EMMS drag model and
results obtained from these simulations were presented in terms of
solid flux and axial solids concentration distribution. Additionally,
the effect of dcl =dp on computational cost was also examined in
this work.

Effect of dcl =dp on solid flux


Solid flux is an important parameter to characterize the flow
pattern in CFB risers.[55] Table 5 shows the effect of dcl =dp on time-
averaged solid flux results. It is observed that all dcl =dp can give
reasonable time-averaged solid flux results with the EMMS drag. It
Figure 5. Effect of d cl =d p on cross-sectional time-averaged axial solid
is also observed that, when dcl =dp increases from 55–95, the time-
concentration distribution.
averaged solid flux gradually increases. If dcl =dp is further
increased to 200, a small decrease in solid flux results is observed.
The closest prediction of time-averaged solid flux with experi-
mental data is reported from dcl =dp ¼ 95. In summary, varying cost changes little, while when dcl =dp < 95, the CPU cost increases
dcl =dp does have an influence on time-averaged solid flux results, exponentially with decrease of dcl =dp . By considering the parcel-
but in the range we consider here, all the results are in an independent solid concentration profiles in Figure 5 and CPU cost
acceptable range. in Figure 6, we selected dcl =dp ¼ 95 for further simulations.

Effect of dcl =dp on hydrodynamics Comparison of DDPM and TFM Approaches


Figure 5 shows the effect of dcl =dp on time-averaged axial solid In this section, a comparison of DDPM and TFM approaches is
concentration profile. From Figure 5, it can be seen that increasing presented. In TFM, particle size distribution needs additional
dcl =dp does not show a significant influence on time-averaged axial computational resources to solve the equations of each particle
solid concentration profile results. Below a height of 8 m, the solid
concentration distribution from dcl =dp ¼ 200 is observed to be
different from the other three cases, while for dcl =dp ¼ 5595, the
simulated results show similar solid concentration distribution.
Generally, the results of solid concentration distribution obtained
from all dcl =dp values below 8 m height are in reasonable
agreement with experimental data. Above 8 m height, deviation
is found between cases, which may be attributed to the complex
interaction between configuration of outlet and settings of coarse-
grained parcels, and deserves more efforts in future.
Moreover, the effect of dcl =dp on time-averaged radial solid
volume fraction and velocity profile results is also investigated in
the present work and it is observed that at all dcl =dp values, radial
solid volume fraction and velocity profile results are insensitive to
dcl =dp ratio and are in good agreement with the experiment.
Therefore, results of radial solid concentration and radial solid
velocity profiles are not shown in the present work.

Effect of dcl =dp on CPU cost


Figure 6 shows the effect of dcl =dp on CPU costs. All the
simulations were carried out with parallel computing with the
same number of processors. It can be seen that a larger value of
dcl =dp results in decreased CPU cost. When dcl =dp > 95, the CPU Figure 6. Effect of d cl =d p on computational cost.

VOLUME 96, JUNE 2018 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 1411
concentration profile accurately. Snapshot results of time-aver-
Table 6. Solid flux results obtained from DDPM and TFM simulations
with two different drag models
aged axial solid concentration distribution in Figure 7 (right) from
DDPM-EMMS and TFM-EMMS simulations also show a reason-
Gs (kg/(m2  s)) able solid concentration distribution trend similar to experimental
observations than the DDPM-Gidaspow and TFM-Gidaspow
Drag model Experiment DDPM Error (%) TFM Error (%) simulations. Furthermore, it is also observed that solid concentra-
EMMS 86.1 89 3.4 142 65 tion distribution in the TFM approach from both EMMS and
Gidaspow 86.1 380 341 422 390 conventional homogenous drag at the solid inlet section is denser
than the DDPM modelling approach.
Radial solid concentration
size. In this preliminary work, we use the DDPM approach by Figure 8 shows time-averaged radial solids concentration
considering mean particle size and comparing it with TFM. To distribution at four different heights (1.54, 2.59, 4.21, and
consider the effect of meso-scale structures, the EMMS drag model 7.08 m). From Figure 8, experimental data shows that the radial
is coupled and compared with the homogenous drag model. solids concentration near the wall is dense and the centre is
dilute, confirming a core-annulus structure. Our simulations
Solid flux
with the EMMS drag model predicted similar results with the
Table 6 shows a summary of the simulated time-averaged solid experiments at all riser heights in both modelling approaches,
flux results against experimental data. It is clear that the while the conventional homogenous drag model failed to
homogenous drag model in both DDPM and TFM approaches capture the dense wall and dilute core region at the bottom
over-predicts the solid flux, while the results from the EMMS drag part of the riser (1.54 and 2.59 m heights). It is also observed
model are close to the experimental results. The predicted value of that radial solids concentration profile results from DDPM-
time-averaged solid flux from DDPM-EMMS simulations is much EMMS and TFM-EMMS are almost similar and are in good
closer to the experiment as compared to TFM-EMMS. In the TFM- agreement with the experiment.
EMMS approach, slightly higher solid flux is observed which is
also reported in the literature,[46,56,57] while the studies for solid Radial solid velocity
flux results with the DDPM-EMMS modelling approach are not Figure 9 shows time-averaged radial solid velocity distribution
reported in the literature and will be further discussed in future at four different heights (1.54, 2.59, 4.21, and 7.08 m).
work. Experimental results in Figure 9 show negative velocity near
the wall and positive velocity in the centre. The EMMS drag
Axial solid concentration model in the present simulations captures the right phenomena
Figure 7 shows the comparison of simulated time-averaged axial similar to experimental observations. In contrast, the homoge-
solid concentration profile results with the experiment (left) and nous drag model failed to capture radial solid velocity
snapshots of time-averaged axial solid concentration distributions distribution near the wall at all measured heights. Further-
(right). Experimental data in Figure 7 (left) shows high solids more, it is also observed that results of radial solid velocity
concentration at the bottom and low solids concentration at the from DDPM-EMMS and TFM-EMMS are almost similar except a
top, making an S-shaped axial solids concentration profile. The slightly higher value of radial velocity is observed in the core
EMMS drag model in our simulations from the DDPM and TFM region from the DDPM-EMMS approach.
approaches predicted the time-averaged axial solid concentration In summary, all results from the DDPM-EMMS approach in
profile in reasonable agreement with experimental observations, comparison to TFM-EMMS showed the capability of resolving the
while the homogenous drag model failed to capture the axial solid hydrodynamics phenomena in the CFB riser, which can be further

Figure 7. Comparison of predicted time-averaged axial solid concentration profile with experiment.

1412 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 96, JUNE 2018
Figure 8. Comparison of radial solid concentration profiles measured at four different heights (1.54, 2.59, 4.21, and 7.08 m) with the experiment: (a)
DDPM, (b) TFM.

extended to full loop simulations with PSD. The comparison of two modelling approaches. The following conclusions are drawn
different drag models showed that the EMMS drag model is needed from the present investigations.
for the DDPM approach as is the case in TFM.
1. The DDPM modelling approach, when used for CFB
simulations with Geldart A particles, should also consider
CONCLUSIONS the effect of meso-scale structures on drag force as in the TFM
Hydrodynamics behaviour of a semi-industrial CFB is investigated modelling approach, and the simulations with EMMS drag
in the present work with coarse-grained DDPM and TFM can give reasonable results.

VOLUME 96, JUNE 2018 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 1413
Figure 9. Comparison of radial solid velocity profiles measured at four different heights (1.54, 2.59, 4.21, and 7.08 m) with the experiment: (a) DDPM, (b) TFM.

2. Effect of particle number in a parcel was investigated and it homogenous drag. It has been found that coupling EMMS
has been found that when the parcel diameter is in the size drag is needed for both modelling approaches to correctly
range of clusters (dcl =dp ¼ 5595), the simulated results capture the hydrodynamics in a CFB.
show similar solids concentration distribution. Considering
the parcel-independent profiles and CPU cost, dcl =dp ¼ 95 was
ACKNOWLEDGEMENT
selected as optimum to reproduce the closest results to
experimental measurements. This work was financially supported by the National Basic
3. Comparison between DDPM and TFM approaches was Research Program of China (973 Program, Grant No.
investigated, both coupled with EMMS and conventional 2014CB744304) and National Natural Sciences Foundation of

1414 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 96, JUNE 2018
China (NSFC) under Grant Nos. 21625605 and 91334204. The f angle of internal friction (8)
authors would especially like to express their gratitude to g Qp collisional dissipation of energy (kg/(m  s3))
Wojciech Adamczyk at the Institute of Thermal Technology,
Silesian University of Technology for fruitful discussions about
Subscripts
the DDPM modelling approach and providing assistance with
the UDF for solids recycling stream. g gas phase
p particle phase

NOMENCLATURE
ac acceleration of particle in dense phase (m/s2) REFERENCES
af acceleration of particle in dilute phase (m/s2) [1] A. Harris, J. Davidson, R. Thorpe, Powder Technol. 2002,
ai acceleration of particle at interface (m/s2) 127, 128.
CD drag coefficient [2] M. Horio, H. Kuroki, Chem. Eng. Sci. 1994, 49, 2413.
dp diameter of particle (m)
[3] H. Li, Y. Xia, Y. Tung, M. Kwauk, Powder Technol. 1991, 66,
dparcel diameter of parcel (m)
231.
dcl diameter of cluster (m)
epp restitution coefficient for particle collisions [4] J. Li, M. Kwauk, Particle-Fluid Two-Phase Flow: The Energy-
f volume fraction of dense phase Minimization Multi-Scale Method, 1st edition, Metallurgical
~
g acceleration due to gravity (m/s2) Industry Press, Beijing 1994.
go radial distribution function [5] K. Agrawal, P. N. Loezos, M. Syamlal, S. Sundaresan, J. Fluid
Gs solid flux (kg/(m2  s)) Mech. 2001, 445, 151.
H riser height (m) [6] W. Dong, W. Wang, J. Li, Chem. Eng. Sci. 2008, 63, 2798.
HD heterogeneity index [7] N. Yang, W. Wang, W. Ge, J. Li, Chem. Eng. J. 2003,
I unit tensor 96, 71.
I 2D second invariant of the deviatoric stress tensor
[8] S. Benyahia, S. Sundaresan, Powder Technol. 2012, 220, 2.
Kgp-EMMS interphase exchange coefficient calculated from the
EMMS/matrix model (kg/(m3  s)) [9] W. Wang, J. Li, Chem. Eng. Sci. 2007, 62, 208.
Kgp-Wen&Yu interphase exchange coefficient calculated from [10] J. Wang, W. Ge, J. Li, Chem. Eng. Sci. 2008, 63, 1553.
Wen-Yu model (kg/(m3  s)) [11] A. T. Andrews IV, P. N. Loezos, S. Sundaresan, Ind. Eng.
k Qp diffusion coefficient (kg/(m  s)) Chem. Res. 2005, 44, 6022.
mparcel mass of a parcel (kg) [12] Y. Igci, A. T. Andrews, S. Sundaresan, S. Pannala, T. O’Brien,
mp mass of a particle (kg) AIChE J. 2008, 54, 1431.
mi mass of particle i (kg) [13] A. Ozel, J. Kolehmainen, S. Radl, S. Sundaresan, Chem. Eng.
Np number of particles in a parcel Sci. 2016, 155, 258.
N parcel number of parcels
[14] A. Nikolopoulos, K. Atsonios, N. Nikolopoulos, P. Gramme-
p gas pressure shared by both gas and solid phases
lis, E. Kakaras, Chem. Eng. Sci. 2010, 65, 4089.
(Pa)
Rep particle Reynolds number [15] Q. Zhou, J. Wang, Chem. Eng. Sci. 2015, 122, 637.
U gc superficial gas velocity for dense phase (m/s) [16] H. Enwald, E. Peirano, A.-E. Almstedt, Int. J. Multiphas. Flow
U gf superficial gas velocity for dilute phase (m/s) 1996, 22, 21.
U sc superficial particle velocity for dense phase (m/s) [17] D. Gidaspow, Multiphase Flow and Fluidization: Continuum
U sf superficial particle velocity for dilute phase (m/s) and Kinetic Theory Descriptions, 1st edition, Academic press,
~
vp velocity of particle (m/s) New York 1994.
Vp volume of a particle (m3) [18] K. Chu, A. Yu, Powder Technol. 2008, 179, 104.
V cell volume of cell (m3)
[19] P. A. Cundall, O. D. Strack, Geotechnique 1979, 29, 47.
V sz axial solid velocity (m/s)
~
xp position of a particle (m) [20] Y. Tsuji, T. Kawaguchi, T. Tanaka, Powder Technol. 1993,
77, 79.
[21] B. Xu, A. Yu, Chem. Eng. Sci. 1997, 52, 2785.
Greek Letters [22] B. Popoff, M. Braun, “A Lagrangian Approach to Dense
e voidage Particulate Flows,” 6th International Conference on Multi-
egc dense phase voidage phase Flow, ICMF, Leipzig, 9–13 July 2007.
egf dilute phase voidage [23] M. Andrews, P. O’Rourke, Int. J. Multiphas. Flow 1996,
ecl volume fraction of cluster 22, 379.
ep;max maximum packing limit [24] K. Chu, J. Chen, A. Yu, Miner. Eng. 2016, 90, 43.
r density (kg/m3)
[25] A. Nikolopoulos, A. Stroh, M. Zeneli, F. Alobaid, N.
t stress tensor (Pa) €hle, S. Karellas, B. Epple, P. Grammelis,
Nikolopoulos, J. Stro
m shear viscosity (kg/(m  s))
Chem. Eng. Sci. 2017, 163, 189.
mp;kin kinetic viscosity (kg/(m  s))
mp;col collisional viscosity (kg/(m  s)) [26] L. Lu, J. Xu, W. Ge, G. Gao, Y. Jiang, M. Zhao, X. Liu, J. Li,
mp;fr frictional viscosity (kg/(m  s)) Chem. Eng. Sci. 2016, 155, 314.
l bulk viscosity (kg/(m  s)) [27] W. P. Adamczyk, A. Klimanek, R. A. Białecki, G. Węcel, P.
Qp granular temperature (m2/s2) Kozołub, T. Czakiert, Particuology 2014, 15, 129.

VOLUME 96, JUNE 2018 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING 1415
[28] F. Li, F. Song, S. Benyahia, W. Wang, J. Li, Chem. Eng. Sci. €h€anen, S. Kallio, T. Hypp€anen, Particuology
[57] S. Shah, K. Myo
2012, 82, 104. 2015, 18, 66.
[29] W. P. Adamczyk, G. Węcel, M. Klajny, P. Kozołub, A.
Klimanek, R. A. Białecki, Particuology 2014, 16, 29.
[30] M. M. Farid, H. J. Jeong, K. H. Kim, J. Lee, D. Kim, J. Hwang, Manuscript received July 10, 2017; revised manuscript received
Fuel 2017, 192, 187. September 3, 2017; accepted for publication September 25, 2017.
[31] L. Lu, J. Xu, W. Ge, Y. Yue, X. Liu, J. Li, Chem. Eng. Sci. 2014,
120, 67.
[32] M. Xu, W. Ge, J. Li, Chem. Eng. Sci. 2007, 62, 2302.
[33] S. Cloete, S. Johansen, M. Braun, B. Popoff, S. Amini,
“Evaluation of a Lagrangian Discrete Phase Modeling
Approach for Resolving Cluster Formation in CFB Risers,”
7th International Conference on Multiphase Flow, ICMF,
Tampa, 30 May–4 June 2010.
[34] E. M. Ryan, D. DeCroix, R. Breault, W. Xu, E. D. Huckaby, K.
Saha, S. Dartevelle, X. Sun, Powder Technol. 2013, 242, 117.
[35] T. B. Anderson, R. Jackson, Ind. Eng. Chem. Fund. 1967, 6,
527.
[36] M. Ishii, Thermo-Fluid Theory of Two-Phase Flows, Eyrolles,
Paris 1975.
[37] S. Chapman, T. G. Cowling, The Mathematical Theory of
Non-Uniform Gases: An Account of The Kinetic Theory of
Viscosity, Thermal Conduction and Diffusion in gases, 3rd
edition, Cambridge University Press, New York 1970.
[38] ANSYS Fluent, 14.0, Theory guide, ANSYS, Canonsburg
2011.
[39] D. M. Snider, J. Comput. Phys. 2001, 170, 523.
[40] C. Lun, S. B. Savage, D. Jeffrey, N. Chepurniy, J. Fluid Mech.
1984, 140, 223.
[41] D. Gidaspow, R. Bezburuah, J. Ding, “Hydrodynamics of
Circulating Fluidized Beds: Kinetic Theory Approach,” 7th
Engineering Foundation Conference on Fluidization, Engi-
neering Foundation, Brisbane 3–8 May 1992.
[42] D. G. Schaeffer, J. Differ. Equations 1987, 66, 19.
[43] J. Ding, D. Gidaspow, AIChE J. 1990, 36, 523.
[44] Y. Tian, J. Geng, W. Wang, Chem. Eng. Sci. 2017, 171, 271.
[45] C. Y. Wen, Y. H. Yu, Chem. Eng. Prog. S. Ser. 1966, 162, 100.
[46] N. Zhang, B. Lu, W. Wang, J. Li, Particuology 2008, 6, 529.
[47] P. Herbert, R. Nicolai, L. Reh, “The ETH Experience:
Experimental Database and Results from the Past 8 Years,”
AIChE Annual Meeting, Miami, 15–21 November 1998.
[48] Y. Mei, M. Zhao, B. Lu, S. Chen, W. Wang, Particuology 2017,
31, 42.
[49] P. C. Johnson, R. Jackson, J. Fluid Mech. 1987, 176, 67.
[50] ANSYS Fluent, 14.0, User’s guide, ANSYS, Canonsburg 2011.
[51] A. Kaufmann, M. Moreau, O. Simonin, J. Helie, J. Comput.
Phys. 2008, 227, 6448.
[52] W. P. Adamczyk, P. Kozołub, A. Klimanek, R. A.
Białecki, M. Andrzejczyk, M. Klajny, Appl. Therm. Eng.
2015, 87, 127.
[53] W. P. Adamczyk, P. Kozołub, G. Kruczek, M. Pilorz, A.
Klimanek, T. Czakiert, G. Węc el, Particuology 2016,
29, 69.
[54] A. Klimanek, W. Adamczyk, A. Katelbach-Wo zniak, G.
Węcel, A. Szlęk, Fuel 2015, 152, 131.
[55] B. Lu, W. Wang, J. Li, Chem. Eng. Sci. 2011, 66, 4624.
[56] B. Lu, W. Wang, J. Li, Chem. Eng. Sci. 2009, 64, 3437.

1416 THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING VOLUME 96, JUNE 2018

You might also like