You are on page 1of 15

Powder Technology 354 (2019) 615–629

Contents lists available at ScienceDirect

Powder Technology

journal homepage: www.elsevier.com/locate/powtec

CFD simulations of a full-loop CFB reactor using coarse-grained Eulerian–


Lagrangian dense discrete phase model: Effects of modeling parameters
Adnan Muhammad a,b, Nan Zhang b,⁎, Wei Wang a,b
a
State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, PR China
b
University of Chinese Academy of Sciences, Beijing 100049, PR China

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

Article history: The hydrodynamics of a 3D full-loop circulating fluidized bed (CFB) reactor were investigated using a coarse-
Received 28 December 2018 grained dense discrete phase model (DDPM). The effects of changes in key modeling parameters such as the
Received in revised form 16 April 2019 drag force, particle number per parcel, and particle–particle/particle–wall restitution coefficients were thor-
Accepted 13 June 2019
oughly investigated. Previous experimental data were used as a benchmark for validating the numerical results.
Available online 15 June 2019
In terms of the drag force, the simulation results show that the effect of mesoscale structures on drag force must
Keywords:
be considered in the DDPM approach. Regarding the number of particles per parcel, the predictions indicate
Mesoscale structures parcel-independent behavior at all of the coarse-graining ratios tested (dcl/dp = 55, 95, and 125), which is further
EMMS proof that parcel-independent results can be achieved when the parcel diameter is in the size range of clusters.
Coarse-graining The effect of energy dissipation from particle–particle (epp) collisions shows that better results are generated
DDPM when epp is in the non-ideal range (0.1–0.9) than for the ideal case (epp = 1.0). However, the particle–wall res-
CFB titution coefficient (epw) has little effect on hydrodynamic behavior in either the ideal or non-ideal collision
CFD range. In addition, the effect of the particle size distribution on the full-loop hydrodynamics was investigated.
The predicted results are similar to those from mean particle diameter simulations, but the distributions for dif-
ferent particles in the CFB are different. This would certainly affect the heat and mass transfer and reactions, and
will be further investigated in our ongoing simulations of combustion in CFBs.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction dimensionally means that the theoretical progress has been slow.
In contrast, by considering the effect of mesoscale structures
Circulating fluidized bed (CFB) reactors are common in many in- based on energy analysis, numerical simulations have given rea-
dustrial applications, such as fluid catalytic cracking (FCC), coal com- sonable results [11,12].
bustion, coal gasification, and Fischer–Tropsch synthesis. Compared Computational fluid dynamics (CFD) approaches for the numerical
to conventional coal fired boilers, CFB technology is increasingly simulation of gas–solid CFB reactors can be classified into two groups:
popular because of several unique and attractive features such as (1) the Eulerian–Eulerian approach, commonly known as the two-
fuel flexibility, better mixing efficiency, higher heat and mass trans- fluid model (TFM), and (2) the Eulerian–Lagrangian approach, i.e., the
fer rates and lower NOx emissions [1]. Despite all these applications discrete particle model (DPM). In the TFM approach, the gas and solid
and advantages, the design and scale-up of such reactors remain phases are treated as interpenetrating continua [21,22]. Mass and mo-
challenging because of their inherently spatio-temporal mesoscale mentum conservation equations are solved for each of the phases, and
structures [2–11]. These mesoscale structures feature nonlinear the interaction between particles is modeled using the kinetic theory
non-equilibrium behavior, which further impacts the overall perfor- of granular flow (KTGF) [23,24]. In the DPM approach, the gas phase
mance of CFB reactors in terms of their hydrodynamics, heat and motion is solved in a similar manner to the Eulerian TFM approach,
mass transfer, and reaction kinetics [5,12–14]. whereas the particle motion is solved by tracking the trajectories of
In the literature, there are several reports on the characteristics solid particles using Newton’s second law of motion [25–28]. To handle
of mesoscale structures elicited using experimental methods particle–particle collisions in the DPM approach, two different treat-
[15–20]. However, the spatio-temporal characteristics and diffi- ments are adapted: (1) an explicit particle–particle collision method,
culty of measuring these structures non-intrusively and three- e.g., the discrete element model (DEM) [26,28] or coarse-grained DEM
[29–33], or (2) an implicit particle–particle collision method, e.g., the
⁎ Corresponding author. multiphase particle in cell model (MP-PIC) [34,35] or dense discrete
E-mail address: nzhang@ipe.ac.cn (N. Zhang). phase model (DDPM) [36]. In the former method, the solver tracks the

https://doi.org/10.1016/j.powtec.2019.06.016
0032-5910/© 2019 Elsevier B.V. All rights reserved.
616 A. Muhammad et al. / Powder Technology 354 (2019) 615–629

trajectories of individual particles or groups of particles, called parcels Mass conservation equation for gas phase:
(each parcel consists of a number of real particles having the same
mass, velocity, density etc.), and collisions between particles are ∂   
ε g ρg þ ∇∙ εg ρg vg ¼ 0 ð1Þ
modelled using a soft sphere approach. In the latter method, the parti- ∂t
cles are tracked as parcels, but collisions between particles are modeled
using a simple stress–strain relation or KTGF. For large-scale industrial Momentum conservation equation for gas phase:
CFB reactors involving trillions of particles distributed across a wide
∂   
range of sizes, the DDPM/MP-PIC approaches are a good choice because ε g ρg vg þ ∇∙ εg ρg vg vg ¼ −εg ∇p þ ∇∙τg þ εg ρg g
∂t  
(1) the particle size distribution (PSD) is resolved naturally by Lagrang- þ K gp vp −vg ð2Þ
ian particle tracking, (2) the number of particles is reduced by tracking
parcels instead of individual particles, and (3) the computation speed is where εg is the gas phase volume fraction, ρg is the gas phase density, vg
accelerated because the collisions between particles are modeled using is the gas phase velocity, vp is the solid phase velocity, p is the gas pres-
KTGF. With all these advantages, the DDPM approach has recently been sure shared by both the gas and solid phases, τg represents the stress
used to analyze the flow behavior in bubbling fluidized beds [37,38], tensor of the gas phase, g is gravitational acceleration, and Kgp is the in-
CFBs [39–45], and cyclones [46]. However, the accuracy of the numerical terphase momentum exchange coefficient from the solid to the gas
results depends on the selection of appropriate model parameters (drag phase per unit volume of cells.
force, number of particles per parcel, and particle–particle restitution Particle equation of motion for solid phase:
coefficient) and boundary conditions (particle–wall tangential and nor-
 
mal coefficients) [33,40,44,47].
dvp   g ρp −ρg
In comparison to TFM and MP-PIC, fewer articles have been pub- ¼ F D vg −vp þ þ F interaction ð3Þ
dt ρp
lished on DDPM and there is insufficient information about selecting ap-
propriate model parameters and boundary conditions [37,40–43,45]. In
where FD(vg − vp) is the acceleration due to the drag force, g(ρp−ρg)/ρp
our previous study [44], an attempt was made to evaluate the effect of
represents acceleration due to gravity, ρp is the particle density, and
mesoscale structures on drag force models and the number of particles
Finteraction is the additional acceleration on the particles resulting from
per parcel using riser-only simulations with the DDPM approach. The
inter-particle interactions.
simulation results showed that the effect of mesoscale structures on
Integrating the particle equation of motion gives the particle posi-
drag force needs to be considered in DDPM and the parcel diameter
tion xp, which can be expressed as follows:
should fall into the size range of clusters to give reasonable parcel-
independent results and CPU cost. dxp
In this study, we extend the DDPM approach to full-loop simulations ¼ vp ð4Þ
dt
of a semi-industrial-scale CFB reactor. The effects of key model parame-
ters (drag force, number of particles per parcel, and particle–particle Acceleration due to the drag force FD(vg − vp) is defined as follows:
restitution coefficient) and boundary conditions (particle–wall tangen-
tial and normal coefficients) on the full-loop simulations are evaluated   3 μg  
F D vg −vp ¼ C d Rep 2
vg −vp ð5Þ
in detail. After successfully validating the model parameters against ex- 4 ρp d p
perimental measurements, the effect of PSD on the hydrodynamics of
this full-loop CFB is discussed. where μg is the gas viscosity, dp is the particle diameter, Cd is the drag
coefficient, and Rep is the particle Reynolds number, which can be
expressed as follows:
2. Numerical modeling
 
ρg dp vg −vp
Rep ¼ ð6Þ
2.1. DDPM μg

A DDPM is used for the full-loop CFB simulations. This model is a After obtaining FD(vg − vp) from Eq. (5), Kgp(vp − vg) in Eq. (2) can
variant of the standard DPM approach in which the gas phase is be calculated as follows:
solved using the Navier–Stokes equations and the motion of the par-
ticle phase is solved using Newton’s second law of motion. One dis-   XN  
advantage of the standard DPM approach is that it does not take K gp vg −vp ¼ mi F D vg −vpi =V cell ð7Þ
i¼1
into account the volume fraction of the discrete particle phase,
thus restricting its applications to dilute multiphase flows. The
where N = Np ∙ Nparcel is the total number of particles in a cell, mi is the
DDPM approach overcomes this limitation by taking into account
mass of particle i, and vpi represents the velocity of particle i.
the volume fraction of the discrete particle phase and can be applied
The Finteraction force in DDPM is modeled using the solid phase stress
to both dilute as well as dense multiphase flow applications [48].
tensor (τp) given by KTGF:
Moreover, the particle equation of motion in DDPM is not solved
for the individual particles. The solver tracks groups of particles, usu- 1
ally called parcels. Each parcel consists of several numbers of real F interaction ¼ − ∇∙τp ð8Þ
ρp
particles which have the same mass, velocity, position etc. The parcel
properties such as velocity and position are first calculated from the The following expression, based on the work of Lun et al. [49], is used
particle equation of motion on Lagrangian coordinate and then inter- to model τp:
polated back to the Eulerian coordinate, which gives mean solid ve-
 
locity and volume fraction in each Eulerian cell. The particle-particle   2
τp ¼ −pp I þ εp μ p ∇vp þ ∇vp T þ εp λp − μ p ∇:vp I ð9Þ
and gas-particle interaction forces in DDPM are also modelled on the 3
Eulerian coordinate, which are further introduced back to the parti-
cle equation of motion as the source by using interpolation operators where pp represents the solid pressure, I is the unit stress tensor, μp rep-
[34]. The governing equations of the gas and solid phases in the resents the particle dynamic viscosity, λp is the bulk viscosity, and vp
DDPM approach are listed below. represents the average velocity vector of the solid phase.
A. Muhammad et al. / Powder Technology 354 (2019) 615–629 617

Table 1
Closure laws used for the solid stress tensor based on KTGF

Parameters Closure law

Solid pressure Lun et al. [49]


Shear viscosity Gidaspow [21]
Kinetic viscosity Gidaspow et al. [24]
Collisional viscosity Gidaspow et al. [24]
Frictional viscosity Schaeffer [50]
Bulk viscosity Lun et al. [49]
Diffusion coefficient Gidaspow et al. [24]
Collisional dissipation Lun et al. [49]
Energy exchange term Gidaspow et al. [24]
Granular temperature Algebraic equation
Radial distribution function Lun et al. [49]

The solid-phase pressure pp in Eq. (9) is also modeled with the help
of KTGF in DDPM and can be expressed as follows:

 
pp ¼ εp ρp ϴ p þ 2ρp 1 þ epp ε2p g o ϴ p ð10Þ

where epp is the particle–particle restitution coefficient, go is the radial


distribution function, and ϴp is the granular temperature.
The radial distribution function, go in Eq. (10) is a correction factor
that modifies the probability of collisions between grains when the
Fig. 1. Schematic of the semi-industrial-scale CFB reactor [55].
solid granular phase becomes dense. The following correlation of Lun
et al. [49] is used for the calculation of go:
The kinetic and collisional shear viscosities are modeled by using the
" #−1 expression of Gidaspow et al. [24]:
 1=3
εp
g o ¼ 1− ð11Þ pffiffiffiffiffiffiffiffiffi

ε p; max 10ρp dp ϴ p π 4   2
μ p;kin ¼   1 þ g o εp 1 þ epp ð13Þ
96εp 1 þ epp g o 5

In Eq. (9), solid shear and bulk viscosities are required for the calcu-  
4   ϴ p 1=2
lation of shear and normal stresses. To calculate these viscosities, the μ p;col ¼ ε p ρp dp g o 1 þ epp ð14Þ
KTGF only is not sufficient and plastic theory is also required for the cal- 5 π
culation of frictional stresses that prevail when the solids loading is
high. Thus, the solids shear viscosity for the dense granular flows is The optional frictional shear viscosity is modeled by using the ex-
the sum of kinetic, collisional, and frictional viscosities. pression of Schaeffer [50]:

pp sin∅
μ p ¼ μ p;kin þ μ p;col þ μ p;fr ð12Þ μ fir ¼ pffiffiffiffiffiffiffi ð15Þ
2 I2D

Table 2
Fitting formula of Hd calculated from EMMS-matrix model [9]

Range Fitting formula, Hd = a(Rep + b)c/Hd, max, where Hd, max = 3.04, εg, mf = 0.448, and εg, mf = 0.9997

εg, mf ≪ εg ≤ 0.5 ε g− 0:50778 2


a ¼ 0:7934 þ 1:1175 exp½−2ð 0:06956 Þ 
b=0
c=0
0.5 ≪ εg ≤ 0.6

50:25633 −1
εg
a ¼ 0:05204 þ 1:9743 1 þ ð Þ
0:53027

−1 (
−1 )
ε g −0:50084 −1 ε g− 0:60902
b ¼ 0:14294−0:15047 1 þ expð 0:0047 Þ 1− 1 þ exp−ð 0:02736 Þ
34:39893 −1

εg
c ¼ 0:56038−0:58092 1 þ ð Þ
0:54322
0.6 ≪ εg ≤ 0.98 a = (429.90532 − 434.69389εg)−0.5868
εg −1:02901

−1:58454 þ 7:56707ðε g −1:02901Þ−19:6441ðε g −1:02901Þ2
c = (0.05377 − 0.05394εg)0.163018
0.98 ≪ εg ≤ εg, max
(
−1 )
ε g− 0:995945 −1 ε g− 0:999915
a ¼ 0:36777 þ 5:13409 1 þ exp−ð 0:00447 Þ 1− 1 þ exp−ð 0:00044 Þ

ε g− 0:9986 2
b ¼ 0:0175 þ 0:71397 exp½−0:5ð 0:00186 Þ 

ε g− 0:9983 2
c ¼ 0:32769−0:25491 exp½−0:5ð 0:00969
Þ 

εg ≪ εg, mf or εg ≫ εg, max Hd = 1


618 A. Muhammad et al. / Powder Technology 354 (2019) 615–629

Fig 3. PSD curve of FCC particles.

used to model collisional dissipation energy.


 
12 1−e2pp g o
Fig. 2. Geometry and mesh of the full-loop CFB. 3=2
γϴ p ¼ pffiffiffi ρp ε2p ϴ p ð19Þ
dp π

The solid phase bulk viscosity, λp in Eq. (9) is accounted by using the The summary of all closure laws for the solid stress term and granu-
correlation of Lun et al. [49] which can be defined as follows:
lar temperature based on the KTGF are given in Table 1.
 
3   ϴ p 1=2 2.2. Drag force modeling
λp ¼ εp ρp dp g o 1 þ epp ð16Þ
4 π
In gas–solid CFB systems, the motion of particles is driven by
To model the solid phase stress term, the granular temperature is re- the drag force, and so this is one of the most important parameters.
quired which is also derived from the KTGF. The transport equation for The complex mesoscale structures of CFB reactors mean that the
the granular temperature based on KTGF can be expressed as follows: gas–solid flow is a nonlinear non-equilibrium system. These meso-
scale structures cannot be resolved accurately using homogenous


3 ∂    drag or the least energy dissipation criterion [12]. To resolve this
ρp εp ϴ p þ ∇: ρp εp vp ϴ p ¼ −pp I þ τp
2 ∂t   issue, a number of researches have employed the EMMS scheme,
: ∇vp þ ∇: kϴ p ∇ϴ p −γ ϴ p þ ∅gp ð17Þ which considers the flow heterogeneity in each computational
cell when calculating the drag force [8,9,51,52]. Recently, research
where (−ppI + τp) : ∇ vp represents energy generation due to the solid has shown that this kind of drag model is also important in
stress tensor, kϴ p ∇ ϴp is the diffusion energy term, γϴ pis the collisional Eulerian–Lagrangian approaches, but there has been less focus on
dissipation of energy, and ∅gp is the energy exchange between the fluid this aspect [32,33,44,53]. In the present work, we implemented
and solid phases. In the present work, an algebraic treatment is applied an EMMS-matrix drag model [9] for the full-loop CFB reactor and
to the granular temperature.
The diffusion coefficient term, kϴ p is modeled by using the expres- Table 4
sion of Gidaspow et al. [24]: Parameters required for Rosin–Rammler PSD

pffiffiffiffiffiffiffiffiffiffi
Parameters Value
150ρp dp ðϴπÞ 6   2
kϴp ¼   1 þ εp g o 1 þ epp Mean diameter (μm) 72
384 1 þ epp g o 5
rffiffiffiffiffiffi Minimum diameter (μm) 25
  ϴp Maximum diameter (μm) 130
þ 2ρp ε2p dp 1 þ epp g o ð18Þ
π Spread parameter 3.3
Number of diameters 9

The term γϴ p in Eq. (17) represents energy dissipation rate in the pth
solid phase due to particles collisions. The expression of Lun et al. [49] is
Table 5
Boundary and initial conditions.
Table 3
Properties of gas and solid phases Boundary and initial conditions Gas phase Solid phase

Primary air inlet (m/s) 1.2 –


Physical property Value
Secondary air inlet (m/s) 2.3 –
3
Gas density (kg/m ) 1.225 Fluidized air inlet (m/s) 0.003 –
Gas viscosity (kg/(m.s)) 1.7894×10−5 Outlet (Pa) 101325 –
Solid density (kg/m3) 1400 Wall No slip Partial slip
Particle diameter (m) 6×10−5 Solids inventory – 0445 kg
A. Muhammad et al. / Powder Technology 354 (2019) 615–629 619

Fig. 4. Solid flux monitored at a riser height of 7 m.

compared the results with those from the conventional Gidaspow


drag model [24].
The interphase momentum exchange coefficient (Kgp) based on the
EMMS-matrix drag model [9] can be defined as follows:

3 εp εg ρg vp −vg −2:65
K gp−EMMS ¼ Cd εg  Hd ð20Þ
4 dp

where Cd is the drag coefficient and Hd is the heterogeneity index. The


drag coefficient Cd can be defined as follows:
8 h
< 24  0:687 i
1 þ 0:15 εg Rep ; Rep ≪1000
C d ¼ εg Rep ð21Þ
:
0:44 ; Rep ≥1000

Table 2 presents the fitting function formula for Hd calculated from Fig. 5. Instantaneous snapshots of solid volume fraction distribution in full-loop: (a)
the EMMS-matrix drag model for the CFB reactor considered in this EMMS-matrix drag; (b) Gidaspow drag.
study. This fitting function formula is implemented through a user-
defined function (UDF) provided by the software. Note that the slip ve-
locity calculated in Eq. (20) is based on the velocity of each particle
surrounded by the gas phase, instead of the cell-averaged value.
The interphase momentum exchange coefficient (Kgp) based on the
conventional Gidaspow drag model [24] can be defined as follows:

3 εp εg ρg vp −vg −2:65
Kgp−WenYu ¼ Cd εg ; when εg ≫0:80 ð22Þ
4 dp
 
εp 1−εg μ g ρg vp −vg
Kgp−Ergun ¼ 150 2
þ 175 ; when εg ≤0:80 ð23Þ
εg dp dp

2.3. Parcel concept

In parcel-based approaches, the number of particles per parcel is a


key parameter in terms of the simulation speed and accuracy of results
[31,33,40]. The highest possible accuracy may be achieved when the
ratio of parcels to particle size is equal to unity. As this ratio increases
from unity, the accuracy of the simulation results tends to deteriorate.
In large industrial applications, where trillions of particles are involved,
it is almost impossible to achieve a ratio of unity because of the tremen-
dous computational cost. Therefore, selecting a proper parcel to particle
ratio that ensures acceptable computational cost and sufficient numer- Fig. 6. Comparison of cross-sectionally averaged axial solid volume fraction profiles
ical accuracy is important. No general and intrinsic criterion for choos- between experimental data and simulation results predicted by the EMMS-matrix and
ing a suitable parcel to particle ratio has yet been developed. Gidaspow drag models.
620 A. Muhammad et al. / Powder Technology 354 (2019) 615–629

However, some researchers have reported that, for gas–solid fluidized 2.4. Geometry and mesh
bed systems involving distinct mesoscale structures, the parcel size
should be sufficiently small to resolve the mesoscale structures accu- A 3D semi-industrial-scale full-loop CFB reactor is selected for the nu-
rately [31,54]. In the present work, we used a parcel to particle size merical simulations. A schematic diagram of the CFB is shown in Fig. 1.
ratio for full-loop CFB simulations based on our previous work [44]. The CFB consists of four major parts, i.e., riser, cyclone separator, down-
The correlation in the parcel to particle size ratio can be defined as fol- comer, and L-valve. The inner diameter of the CFB riser is 0.411 m and
lows: its height is 8.5 m. Details of the experimental setup have been reported
elsewhere [55], and are not discussed in the present work.
1 3 Fig. 2 shows the geometry and mesh of the full-loop CFB reactor. The
mparcel ρp πdparcel :ε cl d 3 geometry shown in Fig. 2(a) is similar to that in the original experimen-
Np ¼ ¼ 6 ¼ cl
:εcl ð24Þ
mparticle 1 3 dp tal setup, except for a small simplification to the secondary air inlet sec-
ρp πdp
6 tion, where dozens of secondary air inlet ports are replaced with four
equispaced secondary air inlet ports. This kind of simplification was
where Np represents the number of particles per parcel, mparcel is the also made in the work of Zhang et al. [56] for the same riser geometry.
mass of each parcel, mparticle is the mass of a single particle, dcl/dp is The mesh of the CFB, shown in Fig. 2(b), is created using the Gambit ®
the ratio of the cluster diameter to the particle diameter, and εcl is the software. Hexahedral cell elements are selected in most parts of the
volume fraction of particles in a cluster; this value is preliminary set to mesh, except for the L-valve and the connection between the riser and
0.5. The size of the parcel in terms of coarse-graining ratio(dcl/dp), cho- cyclone separator, where tetrahedral elements are selected. The total
sen in the present study (55–125) was recommended from the value of number of cell elements used in the present work is about 220,000,
cluster diameter in Wang and Li’s work [9], where the size of mesoscale with an average cell size of 0.1 m, as in the work of Zhang et al. [56]. Ac-
structures is calculated from the original definition of cluster diameter cording to our experience [56], this number of cells ensures mesh-
for Geldart A particles used in the EMMS model [3]. independent results.

Fig. 7. Comparison of cross-sectionally averaged radial solid volume fraction profiles between experimental data and simulation results predicted by the EMMS-matrix and Gidaspow drag
models at four different CFB riser heights: (a) H = 1.54 m; (b) H = 2.59 m; (c) H = 4.21 m; (d) H = 7.08 m.
A. Muhammad et al. / Powder Technology 354 (2019) 615–629 621

2.5. Material properties Table 6


Parameters for number of particles per parcel in simulations.

The present study considers air and solid materials in the full-loop Coarse-graining ratio Number of particles in Total number of parcels
simulations. The properties of both materials are summarized in (dcl/dp) parcel (Np) (Nparcels)
Table 3. The FCC particles used in the present simulations belong to 65 1.4×105 7.7×106
Geldart group A, with a particle density of 1400 kg/m3 and size range 95 4.3×105 2.6×106
of 25–130 μm. The particle size distribution obtained from experimental 125 9.8×105 1.5×106
data is shown in Fig. 3. The PSD parameters used in the present work are
presented in Table (See Table 4).
Table 7
2.6. Boundary and operating conditions Time-averaged solid flux predicted from different dcl/dp ratios.

Coarse-graining ratio (dcl/dp) Solid flux, Gs (kg/(m2.s))


The full-loop CFB geometry used in the present study consists of a Experiment Simulation % Error
primary air inlet, secondary air inlet, fluidized air inlet, and outlet
65 86.1 90 4.5
boundaries, as shown in Fig. 2(b). The primary air inlet is specified ac-
95 86.1 90 4.5
cording to the experimental air flow rate, whereas the secondary air 125 86.1 89 3.4
inlet is determined by the difference between the total gas flow rate
and primary air contribution. A fluidized air inlet is specified at the bot-
tom of the L-valve, and its value is set to the minimum fluidization ve- a no-slip condition is specified for the gas phase, whereas a partial-slip
locity. A pressure outlet boundary condition is specified at the top condition is applied for the solid phase [49], where the specularity coef-
outlet, where atmospheric pressure is prescribed. At the wall boundary, ficient is 0.6. A wall reflection boundary condition is also specified at the

Fig. 8. Comparison of cross-sectionally averaged radial solid velocity profiles between experimental data and simulation results given by the EMMS-matrix and Gidaspow drag models at
four different CFB riser heights: (a) H = 1.54 m; (b) H = 2.59 m; (c) H = 4.21 m; (d) H = 7.08 m.
622 A. Muhammad et al. / Powder Technology 354 (2019) 615–629

Fig. 11. Effect of dcl/dp on CPU cost.

is used to inject the parcels into the domain). This value of the solid in-
ventory is calculated from the corresponding experimental data for the
pressure drop in the full-loop domain. After being injected into the do-
main, the parcels are fluidized by the primary and secondary air. The
solids leaving from the top of the riser enter the cyclone separator,
where they are separated from the gas. The separation efficiency in
Fig. 9. Snapshots of solid volume fraction distribution simulated from different dcl/dp
ratios: (a) dcl/dp = 65; (b) dcl/dp = 95; (c) dcl/dp = 125. the cyclone separator is reported to be close to 100%, with only a
small portion of the solids (less than 1%) exiting from the outlet of the
cyclone separator during the simulations, reflecting the optimum de-
wall. Details of all the boundary conditions and their operating values sign of the cyclone. Similar cyclone separation efficiency is reported in
are summarized in Table 5. the work of Zhang et al. [56], who conducted TFM simulations using
the same full-loop geometry. The separated solids from the cyclone sep-
2.7. Simulation setup arator flow through the down-comer and then return to the riser
through the L-valve to complete the recirculation cycle.
The simulations are started by injecting parcels from the inlet sur-
face of the riser with a target mass of 445 kg (surface injection option

Fig. 10. Comparison of cross-sectionally averaged axial solid volume fraction profiles Fig. 12. Comparison of cross-sectionally averaged axial solid volume fraction profiles
between simulations and experiments at different dcl/dp ratios. between simulations and experiment predicted for different values of epp.
A. Muhammad et al. / Powder Technology 354 (2019) 615–629 623

2.8. Solution method EMMS and Gidaspow drag models are 90.1 and 379.5 kg/(m2.s), respec-
tively. The EMMS drag model predicted a solid circulation rate much
The ANSYS Fluent ® solver is used to complete the simulation. The closer to the experimental measurements, whereas the Gidaspow drag
phase-coupled SIMPLE method is chosen for the discretization of the model overestimated the solid circulation rate.
pressure and velocity terms. A second-order upwind scheme is chosen Fig. 5 shows instantaneous snapshots of the solid volume fraction
for the momentum conservation equation, and all remaining equations calculated from the EMMS and Gidaspow drag models. In Fig. 5(a), the
are discretized using the QUICK scheme. The simulation time step size solid volume fraction in CFB is distributed heterogeneously, showing
for all transient calculations is set to 0.0005 s. The solid flux is continu- dense aggregates at the bottom and dilute suspension at the top. The
ously monitored in the riser. All cases were simulated for a total physical highest solid volume fraction is predicted at the bottom of the down-
time of 22 s, with the last 10 s used for data analysis. comer and loop seal regions. The flow state at the bottom portion of
the down-comer is a typical bubbling fluidized bed, which further acts
3. Results and discussion as a solid seal to prevent riser gas leakage. These predictions are consis-
tent with empirical knowledge and experimental evidence. In contrast
3.1. Effect of drag force to Fig. 5(a), (b) shows a fairly uniform solid volume fraction distribution
over the whole CFB loop, which is the opposite of experimental
To verify the simulation results against experimental measurements, observations.
the two drag models mentioned in the drag force modeling section In gas–solid simulations, homogeneous drag models tend to overes-
were tested. Fig. 4 shows the results of the time-averaged solid circula- timate the drag force, lifting a large portion of particles upward, which
tion rate given by these two different drag models. The time-averaged results in a uniform solid volume fraction distribution. By considering
solid circulation rate data for both simulation cases were monitored at the effect of mesoscale structures, the EMMS drag model generates a
a riser height of 7 m. The experimental solid circulation rate for the pres- lower drag force, which predicts a non-uniform solid volume fraction
ent CFB riser is 86.1 kg/(m2.s), whereas the values predicted by the distribution in the CFB, as shown in Fig. 5(a).

Fig. 13. Comparison of cross-sectionally averaged radial solid volume fraction profiles between experimental data and simulations predicted using different values of epp at riser heights of:
(a) H = 1.54 m; (b) H = 2.59 m; (c) H = 4.21 m; (d) H = 7.08 m.
624 A. Muhammad et al. / Powder Technology 354 (2019) 615–629

Fig. 6 shows the cross-sectionally averaged axial solid volume profile Fig. 7 compares the predicted cross-sectionally averaged radial solid
along the height of the CFB riser alongside the experimental measure- volume fraction profiles with the experimental data. The experimental
ments. The experimental data in Fig. 6 show a high solid volume fraction data were collected at four different CFB riser heights (H = 1.54, 2.59,
at the bottom of the riser and low solid volume fraction at the top. The 4.21, 7.08 m). As can be seen in Fig. 7, the experimental data indicate
simulation results from the EMMS drag model captured the dense bot- that the solid concentration is dense near the walls and dilute in the
tom part reasonably well, whereas the Gidaspow drag model failed to middle. At the bottom portion, at heights of H = 1.54 and 2.59 m, the
capture the solid volume fraction in this dense bottom part of the simulation results from the EMMS drag model are in good agreement
riser. The predicted profile given by the EMMS drag model shows a with the experimental measurements, whereas the Gidaspow drag
clear heterogeneous distribution in the CFB riser, whereas the Gidaspow model fails to capture the solid volume fraction near the walls. At
drag model predicted an almost uniform profile. However, the improve- greater heights, the difference in the solid volume fraction profiles pre-
ment in the cross-sectionally averaged axial solid volume fraction pro- dicted by both the EMMS and Gidaspow drags decreased, leading to
file results given by this DDPM-EMMS modeling approach is not as big comparable results. Moreover, the radial solid volume fraction profiles
as in previous TFM-EMMS simulations [56] for the same CFB loop. predicted by the EMMS drag model near the walls at heights of H =
Note that the development of the EMMS model and its respective theory 1.54 m, 4.21 m, and 7.08 m are lower than the experimental data. This
were based on a TFM modeling approach using the Eulerian framework kind of discrepancy in the radial solid volume fraction profiles has
for both the gas and solid phases. A new EMMS drag model based on a been reported in previous 2D [17,57,58], 3D [59], and 3D full-loop
Lagrangian framework will be more promising for the DDPM approach simulations [56].
in the future, as this modeling framework can handle the PSD more eas- Fig. 8 shows the cross-sectionally averaged radial solid velocity pro-
ily than the continuum-based approach and can easily be applied to files measured at four different CFB riser heights. Experimental data for
large-scale industrial CFB reactors. the radial solid velocity profiles in Fig. 8 show a core-annulus structure.

Fig. 14. Comparison of cross-sectionally averaged radial solid velocity profiles between experimental data and simulations predicted using different values of epp at riser heights of:
(a) H = 1.54 m; (b) H = 2.59 m; (c) H = 4.21 m; (d) H = 7.08 m.
A. Muhammad et al. / Powder Technology 354 (2019) 615–629 625

Fig. 15. Comparison of cross-sectionally averaged axial solid volume fraction profiles
between experimental data and simulations predicted using different particle–wall
restitution coefficients: (a) epw-n; (b) epw-t.

The solid velocity near the walls is lower and particles move downward, Fig. 16. Comparison of cross-sectionally averaged axial solid volume fraction profiles
while in the core region, the velocity of solid particles is higher and the measured from PSD and mean particle diameter simulations.
particles move upward. This phenomenon is predicted well by the
EMMS drag model at all measured heights, except for a small discrep- Fig. 10 compares the cross-sectionally averaged axial solid volume
ancy in the core region, where the predicted solid velocity is higher fraction profiles with the experimental data for various numbers of par-
than the experimental measurements. This kind of discrepancy in the ticles per parcel. All the simulated cases exhibit a similar S-shaped axial
solid velocity profile results from the EMMS drag model simulations solid volume fraction profile and are generally in reasonable agreement
also occurred in our previous 3D riser-only simulations [44], and re- with the experimental observations.
quires further investigation. However, the radial solid velocity profiles The cross-sectionally averaged radial solid velocity and solid
simulated by the Gidaspow model failed to capture the core-annulus volume fraction profiles are almost the same, and are therefore
structure at any of the measured heights (H = 1.54 m, 2.59 m, 4.21 m, not shown here. From the above results, it can be concluded that
and 7.08 m). the hydrodynamic behavior of the full-loop CFB reactor is not sen-
sitive to the number of particles per parcel if the parcel diameter
falls into the size range of clusters. However, the computational
3.2. Effect of particle number per parcel cost is rather different. Fig. 11 shows the effect of dcl /d p on CPU
cost. It can be seen that the number of particles per parcel has a
Three different numbers of particles per parcel were simulated. The
details of the simulation cases in terms of coarse-graining ratio (dcl/dp),
particle number per parcel (Np), and the total number of parcels
(Nparcels) are listed in Table 6. The EMMS drag model was used in all
the simulation cases. Table 7 summarizes the calculated solid flux.
From Table 7, it can be seen that the difference in solid flux results is
fairly small. The time-averaged results from all of the particle number
per parcel simulations are close to the experimental measurements,
and the relative error between the experimental data and the simula-
tions is in the range 3.4–4.5 %, which is acceptable.
Fig. 9 shows instantaneous snapshots of the solid volume fraction
distribution simulated using different numbers of particles per parcel.
From these results, it can be seen that all the simulations predict the
macroscopic dense bottom and dilute top regions of the CFB. However,
finer flow structures are better captured by decreasing the number of
particles per parcel.

Table 8
Solid flux predicted using PSD and mean particle diameter.

Simulation cases Total number of Solid flux, Gs(kg/(m2. s))


parcels,
Experiment Simulation %
(Nparcels)
Error

PSD 4.6×106 86.1 81 −5.9


Mean particle 1.5×106 86.1 88 2.2
Fig. 17. Instantaneous profiles of solid volume fraction distribution in riser obtained with
diameter
different particle diameters.
626 A. Muhammad et al. / Powder Technology 354 (2019) 615–629

significant effect on the computational speed. Although the hydro- Figs. 13 and 14 depict the cross-sectionally averaged radial solid
dynamic results are similar for all cases, when dcl/dp = 65 and Np volume fraction and velocity profiles predicted from different epp
= 1.4×10 5 , the time required to complete 22-s simulations is at four different measured CFB riser heights (H = 1.54 m, 2.59 m,
four times that when d cl /d p = 125. Therefore, d cl /d p = 125 was 4.21 m, and 7.08 m). From these results, it is apparent that the
used in subsequent simulations. cross-sectionally averaged radial solid volume fraction and velocity
profiles predicted are insensitive to epp in the inelastic particle colli-
sion range (epp = 0.1, 0.3, 0.6, and 0.9) and are in reasonable agree-
3.3. Effect of particle–particle restitution coefficient ment with the experimental measurements. However, the radial
solid volume fraction and radial solid velocity profiles predicted
The particle–particle restitution coefficient (epp), which defines the from perfectly elastic particle–particle collisions (epp = 1.0) deviate
energy dissipation resulting from collisions between particles, has a sig- from the experimental measurements (see Fig. 13(b)–(c) and Fig. 14
nificant effect on several solid phase parameters (used in closure law (a)–(b)). The predicted solid volume fraction under perfectly elastic
terms for calculating the solid stress equation based on KTGF), such as particle collision (epp = 1.0) in Fig. 13(b) is overestimated near the
the solid pressure, solid shear viscosity, and solid bulk viscosity. The walls, whereas in Fig. 13(c), its value is underestimated near
value of epp normally varies from 0 to 1. The lower value implies inelas- the walls. Similar to the radial solid volume fraction profile results,
tic particle collisions, whereas the upper value suggests perfectly elastic the radial solid velocity predicted from epp = 1.0 is overestimated
particle collisions. Because of the difficulty of measuring this parameter at riser heights of H = 1.54 m and 2.59 m. These results are consis-
in experimental studies, the exact value of epp is hard to determine. tent with those from previous studies by Goldschmidt et al. [60],
Therefore, values of epp = 0.1, 0.3, 0.6, 0.9, and 1.0 were simulated Yin et al. [47], and Upadhyay and Park [61]. Thus, epp can predict
and their effect on the simulation results was examined. the hydrodynamic behavior of fluidized bed reactors better under
Fig. 12 shows the effect of epp on the cross-sectionally averaged the condition of non-ideal particle–particle collisions than with
axial solid volume fraction profiles. In the inelastic particle collision ideal particle–particle collisions (epp = 1.0), and this should be con-
range (epp = 0.1, 0.3, 0.6, and 0.9), the effect is almost negligible. sidered during the simulations.
All the simulation results for the cross-sectionally averaged axial
solid volume fraction profiles compare well with the experimental 3.4. Effect of particle–wall restitution coefficient
measurements. By comparison, the cross-sectionally averaged axial
solid volume fraction profiles predicted from ideal particle collisions The particle–wall restitution coefficient (epw) accounts for the
(epp = 1.0) vary with the riser height in the range 3–5 m. At these energy loss resulting from inelastic particle–wall collisions. The par-
heights, the predicted distribution of cross-sectionally averaged ticle rebounds off the boundary with a change in momentum defined
axial solid volume fraction is lower than in the inelastic particle col- by the restitution coefficient. The particle–wall restitution coeffi-
lision range (epp = 0.1, 0.3, 0.6, and 0.9). However, the predicted re- cient in discrete numerical simulations includes normal and tangen-
sults are still within an acceptable range of the experimental tial coefficients. The normal coefficient of restitution (epw−n) defines
measurements. the momentum in the direction normal to the wall that is retained

Fig. 18. Instantaneous snapshots of PSD in full-loop CFB.


A. Muhammad et al. / Powder Technology 354 (2019) 615–629 627

from the particles following a collision with the boundary. Similarly, fraction of the particles gradually decreases). The lowest value of the
the tangential coefficient of restitution (epw−t) defines the amount solid volume fraction, close to zero, occurs when dp = 130 μm, for
of momentum in the direction tangential to the wall that is retained. which the distribution is completely homogeneous in the riser.
In the literature, there is little information about selecting an appro- Fig. 18 shows instantaneous snapshots of the PSD predicted from the
priate value for the particle–wall restitution coefficient under the DDPM approach. From Fig. 18, it can be seen that the segregation be-
DDPM approach. To optimize this parameter for numerical simula- tween dense and dilute zones occurs after 11 s. The larger-diameter par-
tions of full-loop CFB reactor, several different values of epw were ticles can be seen in the bottom section of the riser, and as the riser
simulated. height increases, the diameter of the particles decreases. Hence the
From the simulated results for the cross-sectionally averaged smaller-diameter particles from the top of the riser move toward the cy-
axial solid volume fraction profiles (Fig. 15), it can be seen that the clone where, after separation from the gas phase, they fall downward
effect of particle–wall normal (epw−n) and tangential (epw−t) coeffi- and then return back to the riser section through the loop seal. In both
cients on the CFB riser hydrodynamics is minor, which is consistent the cyclone and down-comer regions, our results produce a comparable
with the results of Neri and Gidaspow [62], McKeen and Pugsley PSD. Such mixing behavior in full-loop CFB reactors (as expected from
[63], Chen et al. [64], and Yin et al. [47]. All cases predict full-loop the simulations) can be predicted easily from the DDPM approach with-
CFB reactor hydrodynamics that are in reasonable agreement out any additional computational resources to obtain time-averaged re-
with the experimental measurements. Note that, in the simulations sults, as required in the TFM approach.
to determine the effect of the particle–wall normal coefficient
4. Conclusions
(epw−n), the value of epw−t was fixed to 0.9; in determining the effect
of epw−t, epw−n was set to 0.9.
The hydrodynamics of a full-loop semi-industrial-scale CFB were in-
vestigated using the DDPM approach. The effects of the drag force, par-
3.5. Effect of particle size distribution ticle number per parcel, particle–particle restitution coefficient, and
particle–wall restitution coefficient were examined, and the PSD in
Particles in gas–solid CFB reactors usually have a wide range of sizes, the full-loop CFB was resolved. The results presented in this paper sug-
and the size distribution significantly influences the performance of the gest the following important conclusions:
reactor in terms of its hydrodynamics (mixing and conversion) and 1. The EMMS drag model coupled with DDPM can resolve the sub-grid
chemical reaction kinetics. Therefore, it is important to consider the ef- scale structures in a full-loop CFB reactor better than the Gidaspow
fect of PSD in CFB modeling applications. However, in the Eulerian– drag. The predicted hydrodynamics results using EMMS drag showed
Eulerian modeling approach, a wider PSD means more phases, and heterogeneous flow in the whole CFB loop, whereas the Gidaspow
thus more computational resources. In the DDPM approach, which drag model predicted a homogeneous distribution, which is the op-
treats the particles as parcels, the PSD is resolved naturally using accept- posite of the experimental data.
able computational resources. In this section, various PSDs are applied 2. The effect of the number of particles per parcel for full-loop simula-
to the full-loop CFB reactor application using the DDPM-EMMS model- tions was investigated. Similar to our previous investigation [44]
ing approach. The results obtained from each PSD case are compared (for 3D riser-only simulations), the present study showed that
with those from mean particle diameter simulations and the experi- when the parcel diameter is in the size range of clusters (dcl/dp =
mental measurements. 65–125), the simulation results are parcel-independent and produce
Table 8 compares the solid flux results predicted from the PSD and hydrodynamics that are in reasonable agreement with experimental
mean particle diameter simulations. From Table 8, it can be seen that measurements.
the difference in solid flux results predicted by the two approaches is 3. The effect of particle–particle (epp) and particle–wall (epw) resti-
not large. Both cases predicted similar solid flux results, and these are tution coefficients on the full-loop hydrodynamics was investi-
generally in reasonable agreement with the experimental measure- gated. The simulation results in the non-ideal particle–particle
ments, although the relative error from the mean particle diameter sim- restitution coefficient range (epp = 0.1–0.9) were insensitive to
ulations is less than that from the PSD simulations. this parameter and gave similar results. However, in the ideal
Fig. 16 compares the cross-sectionally averaged axial solid volume range (epp = 1.0), some deviations between the simulations
fraction profiles predicted from PSD and mean diameter simulations and experimental measurements were apparent. By comparison,
against the experimental data. The results predicted by both numerical the particle–wall restitution coefficient (epw) had no significant
simulations overlap and are generally in reasonable agreement with the effect on the hydrodynamic behavior of the full-loop CFB reactor
experimental measurements. in the present study.
Fig. 17 shows the instantaneous profiles of solid volume fraction dis- 4. PSD in the full-loop CFB was resolved using the DDPM-EMMS model-
tribution in the riser for nine different particle diameters (calculated ing approach. The predicted results were almost the same as those
from particle position and parcel mass history data). The solid volume from mean particle diameter simulations. However, the distributions
fraction distribution in the riser varies with the particle diameter. For of different particles in the CFB are different, and this would affect the
smaller (dp = 25) and larger particle diameters (dp = 117, 130 μm), heat and mass transfer and reactions. These effects will be investi-
the solid volume fraction in the riser is distributed uniformly. Generally, gated in our ongoing simulations of combustion in CFBs.
smaller diameter particles (dp = 25) are easier to fluidize than larger
ones. Hence, the axial profile exhibits a uniform trend, whereas for the Nomenclature
larger diameter particles (dp = 117,130 μm), the uniform axial solid vol- Cd Drag coefficient
ume fraction distribution in the riser is the result of the lower mass frac- dcl Cluster diameter (m)
tion of these particles. However, for all other particle diameters (dp = dp Particle diameter (m)
38, 51, 64, 76, 92, and 104 μm), the predicted profiles have a heteroge- dparcel Parcel diameter (m)
neous trend (dense volume fraction at bottom and dilute at top). More- epp Particle–particle restitution coefficient
over, the solid volume fraction distribution in the bottom of the riser epw−n Particle–wall normal coefficient
increases gradually with the diameter of the particles at first (greater epw−t Particle–wall tangential coefficient
mass fraction of these particles present in CFB). After attaining a maxi- g Gravitational acceleration (m/s2)
mum value at dp = 76 μm, the solid volume fraction at the bottom of Gs Solid flux (kg/(m2∙s))
the riser decreases with further increases in particle diameter (mass
628 A. Muhammad et al. / Powder Technology 354 (2019) 615–629

Hd Heterogeneity index [16] U. Lackermeier, C. Rudnick, J. Werther, A. Bredebusch, H. Burkhardt, Visualization of


flow structures inside a circulating fluidized bed by means of laser sheet and image
I Unit tensor processing, Powder Technol. 114 (2001) 71–83.
Kgp Interphase momentum exchange coefficient (kg/(m3∙s)) [17] V. Mathiesen, T. Solberg, B. Hjertager, An experimental and computational study of
kϴ p Diffusion coefficient (kg/(m∙s)) multiphase flow behavior in a circulating fluidized bed, Int. J. Multiph. Flow 26
(2000) 387–419.
mparcel Mass of parcel (kg) [18] C.R. Muller, D.J. Holland, A.J. Sederman, M.D. Mantle, L.F. Gladden, J. Davidson,
mparticle Mass of particle (kg) Magnetic resonance imaging of fluidized beds, Powder Technol. 183 (2008)
Np Number of particles per parcel 53–62.
[19] L. Reh, J. Li, Measurement of voidage in fluidized beds by optical probes, Circulating
Nparcel Number of parcels Fluidized Bed Technol. 3 (1991) 163–170.
p Gas pressure shared by both the gas and solid phases (Pa) [20] J. Werther, Measurement techniques in fluidized beds, Powder Technol. 102 (1999)
Rep Particle Reynolds number 15–36.
[21] D. Gidaspow, Multiphase Flow and Fluidization: Continuum and Kinetic Theory De-
vp Particle velocity (m/s)
scriptions, Academic Press, 1994.
xp Particle position (m) [22] H. Enwald, E. Peirano, A.E. Almstedt, Eulerian two-phase flow theory applied to flu-
idization, Int. J. Multiph. Flow 22 (1996) 21–66.
[23] S. Chapman, T.G. Cowling, The Mathematical Theory of Non-Uniform Gases: an Ac-
Greek letters count of the Kinetic Theory of Viscosity, Thermal Conduction and Diffusion in Gases,
ε Voidage Cambridge University Press, 1970.
[24] D. Gidaspow, R. Bezburuah, J. Ding, Hydrodynamics of circulating fluidized beds: ki-
εcl Cluster voidage netic theory approach, in fluidization VII, Proceedings of 7th Engineering Founda-
ρ Density (kg/m3) tion Conference on Fluidization, 1992.
τ Stress tensor (Pa) [25] P.A. Cundall, O.D. Strack, A discrete numerical model for granular assemblies,
Geotechnique 29 (1979) 47–65.
μ Shear viscosity (kg/(m∙s)) [26] Y. Tsuji, T. Kawaguchi, T. Tanaka, Discrete particle simulation of two-dimensional
λp Bulk viscosity (kg/(m·s)) fluidized bed, Powder Technol. 77 (1993) 79–87.
ϴp Granular temperature (m2/s2) [27] B. Xu, A. Yu, Numerical simulation of the gas–solid flow in a fluidized bed by com-
bining discrete particle method with computational fluid dynamics, Chem. Eng.
γϴ p Collisional dissipation of energy (kg/(m∙s3)) Sci. 52 (1997) 2785–2809.
[28] K. Chu, A. Yu, Numerical simulation of complex particle–fluid flows, Powder
Technol. 179 (2008) 104–114.
Subscripts [29] M. Sakai, S. Koshizuka, Large-scale discrete element modeling in pneumatic convey-
g Gas phase ing, Chem. Eng. Sci. 64 (2009) 533–539.
[30] M. Sakai, Y. Yamada, Y. Shigeto, K. Shibata, V.M. Kawasaki, S. Koshizuka, Large–scale
p Solid phase
discrete element modeling in a fluidized bed, Int. J. Numer. Methods Fluids 64
(2010) 1319–1335.
[31] L. Lu, J. Xu, W. Ge, Y. Yue, X. Liu, J. Li, EMMS-based discrete particle method
Acknowledgment (EMMS–DPM) for simulation of gas–solid flows, Chem. Eng. Sci. 120 (2014)
67–87.
This work was financially supported by the Joint Funds of the [32] L. Lu, J. Xu, W. Ge, G. Gao, Y. Jiang, M. Zhao, X. Liu, J. Li, Computer virtual experiment
on fluidized beds using a coarse-grained discrete particle method—EMMS-DPM,
National Natural Science Foundation of China under Grant Nos. Chem. Eng. Sci. 155 (2016) 314–337.
U1710251, 21625605, and 91834302. [33] A. Nikolopoulos, A. Stroh, M. Zeneli, F. Alobaid, N. Nikolopoulos, J. Strohle, S. Karellas,
B. Epple, P. Grammelis, Numerical investigation and comparison of coarse grain
CFD–DEM and TFM in the case of a 1 MWth fluidized bed carbonator simulation,
References Chem. Eng. Sci. 163 (2017) 189–205.
[34] M. Andrews, P. O'rourke, The multiphase particle-in-cell (MP-PIC) method for dense
[1] P. Basu, Circulating Fluidized Bed Boilers: Design, Operation and Maintenance, particulate flows, Int. J. Multiph. Flow 22 (1996) 379–402.
Springer, 2015. [35] D. Snider, An incompressible three-dimensional multiphase particle-in-cell model
[2] H. Li, Y. Xia, Y. Tung, M. Kwauk, Micro-visualization of clusters in a fast fluidized bed, for dense particle flows, J. Comput. Phys. 170 (2001) 523–549.
Powder Technol. 66 (1991) 231–235. [36] B. Popoff, M. Braun, A Lagrangian approach to dense particulate flows, 6th Int. Conf.
[3] J. Li, M. Kwauk, Particle-Fluid Two-Phase Flow: The Energy-Minimization Multi- on Multiph. Flow, Leipzig, Germany, 2007.
Scale Method, Metallurgical Industry Press, 1994. [37] S. Cloete, S. Amini, S.T. Johansen, A fine resolution parametric study on the numer-
[4] M. Horio, H. Kuroki, Three-dimensional flow visualization of dilutely dispersed ical simulation of gas–solid flows in a periodic riser section, Powder Technol. 205
solids in bubbling and circulating fluidized beds, Chem. Eng. Sci. 49 (1994) (2011) 103–111.
2413–2421. [38] S. Cloete, S. Amini, The dense discrete phase model for simulation of bubbling fluid-
[5] K. Agrawal, P.N. Loezos, M. Syamlal, S. Sundaresan, The role of meso-scale structures ized beds: validation and verification, 9th Int. Conf. on Multiph. Flow, Firenze, Italy,
in rapid gas–solid flows, J. Fluid Mech. 445 (2001) 151–185. 2016.
[6] D.Z. Zhang, W.B. VanderHeyden, The effects of mesoscale structures on the macro- [39] S. Cloete, S. Johansen, M. Braun, B. Popoff, S. Amini, Evaluation of a Lagrangian dis-
scopic momentum equations for two-phase flows, Int. J. Multiph. Flow 28 (2002) crete phase modeling approach for resolving cluster formation in CFB risers, 7th
805–822. Int. Conf. on Multiph. Flow, Tampa, Florida, 2010.
[7] A. Harris, J. Davidson, R. Thorpe, The prediction of particle cluster properties [40] E.M. Ryan, D. DeCroix, R. Breault, W. Xu, E.D. Huckaby, K. Saha, S. Dartevelle, X. Sun,
in the near wall region of a vertical riser, Powder Technol. 127 (2002) Multi-phase CFD modeling of solid sorbent carbon capture system, Powder Technol.
128–143. 242 (2013) 117–134.
[8] N. Yang, W. Wang, W. Ge, J. Li, Choosing structure-dependent drag coeffi- [41] W.P. Adamczyk, P. Kozołub, A. Klimanek, R.A. Białecki, M. Andrzejczyk, M.
cient in modeling gas–solid two-phase flow, China Particuology 1 (2003) Klajny, Numerical simulations of the industrial circulating fluidized bed
38–41. boiler under air-and oxy-fuel combustion, Appl. Therm. Eng. 87 (2015)
[9] W. Wang, J. Li, Simulation of gas–solid two-phase flow by a multi-scale CFD 127–136.
approach–extension of the EMMS model to the sub-grid level, Chem. Eng. Sci. 62 [42] W.P. Adamczyk, P. Kozołub, G. Kruczek, M. Pilorz, A. Klimanek, T. Czakiert, G. Węcel,
(2007) 208–231. Numerical approach for modeling particle transport phenomena in a closed loop of
[10] J. Gao, J. Chang, C. Xu, X. Lan, Y. Yang, CFD simulation of gas solid flow in FCC strip- a circulating fluidized bed, Particuology 29 (2016) 69–79.
pers, Chem. Eng. Sci. 63 (2008) 1827–1841. [43] W.P. Adamczyk, A. Klimanek, R.A. Białecki, G. Węcel, P. Kozołub, T. Czakiert, Compar-
[11] W. Wang, B. Lu, N. Zhang, Z. Shi, J. Li, A review of multiscale CFD for gas–solid CFB ison of the standard Euler–Euler and hybrid Euler–Lagrange approaches for model-
modeling, Int. J. Multiph. Flow 36 (2010) 109–118. ing particle transport in a pilot-scale circulating fluidized bed, Particuology 15
[12] Y. Tian, J. Geng, W. Wang, Structure-dependent analysis of energy dissipation in (2014) 129–137.
gas–solid flows: beyond nonequilibrium thermodynamics, Chem. Eng. Sci. 171 [44] M. Adnan, N. Zhang, F. Sun, W. Wang, Numerical simulation of a semi-industrial
(2017) 271–281. scale CFB riser using coarse–grained DDPM–EMMS modelling, Can. J. Chem. Eng.
[13] W. Dong, W. Wang, J. Li, A multiscale mass transfer model for gas–solid riser 96 (2017) 1403–1416.
flows: Part 1–Sub-grid model and simple tests, Chem. Eng. Sci. 63 (2008) [45] W.P. Adamczyk, K. Myohanen, E.-U. Hartge, J. Ritvanen, A. Klimanek, T. Hyppanen,
2798–2810. R.A. Białecki, Generation of data sets for semi-empirical models of circulated fluid-
[14] W. Dong, W. Wang, J. Li, A multiscale mass transfer model for gas–solid riser flows: ized bed boilers using hybrid Euler-Lagrange technique, Energy 143 (2018)
part II–sub-grid simulation of ozone decomposition, Chem. Eng. Sci. 63 (2008) 219–240.
2811–2823. [46] P. Kozołub, A. Klimanek, R.A. Białecki, W.P. Adamczyk, Numerical simulation of a
[15] A. Carlos Varas, E. Peters, N. Deen, J. Kuipers, Solids volume fraction measurements dense solid particle flow inside a cyclone separator using the hybrid Euler–
on riser flow using a temporal–histogram based DIA method, AICHE J. 62 (2016) Lagrange approach, Particuology 31 (2017) 170–180.
2681–2698.
A. Muhammad et al. / Powder Technology 354 (2019) 615–629 629

[47] S. Yin, W. Zhong, B. Jin, J. Fan, Modeling on the hydrodynamics of pressurized high- [57] A. Almuttahar, F. Taghipour, Computational fluid dynamics of a circulating flu-
flux circulating fluidized beds (PHFCFBs) by Eulerian–Lagrangian approach, Powder idized bed under various fluidization conditions, Chem. Eng. Sci. 63 (2008)
Technol. 259 (2014) 52–64. 1696–1709.
[48] A. Fluent, Release 15.0, Theory Guide, ANSYS Inc, Pittsburgh, 2013. [58] A. Almuttahar, F. Taghipour, Computational fluid dynamics of high density circulat-
[49] C. Lun, S.B. Savage, D. Jeffrey, N. Chepurniy, Kinetic theories for granular flow: inelas- ing fluidized bed riser: study of modeling parameters, Powder Technol. 185 (2008)
tic particles in Couette flow and slightly inelastic particles in a general flowfield, J. 11–23.
Fluid Mech. 140 (1984) 223–256. [59] Q. Wang, T. Niemi, J. Peltola, S. Kallio, H. Yang, J. Lu, L. Wei, Particle size
[50] D.G. Schaeffer, Instability in the evolution equations describing incompressible distribution in CPFD modeling of gas–solid flows in a CFB riser, Particuology 21
granular flow, J. Differ. Equ. 66 (1987) 19–50. (2015) 107–117.
[51] M. Zeneli, A. Nikolopoulos, N. Nikolopoulos, P. Grammelis, E. Kakaras, Application of [60] M. Goldschmidt, R. Beetstra, J. Kuipers, Hydrodynamic modelling of dense gas-
an advanced coupled EMMS-TFM model to a pilot scale CFB carbonator, Chem. Eng. fluidised beds: comparison of the kinetic theory of granular flow with
Sci. 138 (2015) 482–498. 3D hard-sphere discrete particle simulations, Chem. Eng. Sci. 57 (2002)
[52] B. Lu, W. Wang, J. Li, Searching for a mesh-independent sub-grid model for CFD sim- 2059–2075.
ulation of gas–solid riser flows, Chem. Eng. Sci. 64 (2009) 3437–3447. [61] M. Upadhyay, J.-H. Park, CFD simulation via conventional two-fluid model of a circu-
[53] F. Li, F. Song, S. Benyahia, W. Wang, J. Li, MP-PIC simulation of CFB riser with EMMS- lating fluidized bed riser: influence of models and model parameters on hydrody-
based drag model, Chem. Eng. Sci. 82 (2012) 104–113. namic behavior, Powder Technol. 272 (2015) 260–268.
[54] L. Lu, A. Konan, S. Benyahia, Influence of grid resolution, parcel size and drag models [62] A. Neri, D. Gidaspow, Riser hydrodynamics: simulation using kinetic theory, AICHE J.
on bubbling fluidized bed simulation, Chem. Eng. J. 326 (2017) 627–639. 46 (2000) 52–67.
[55] P. Herbert, R. Nicolai, L. Reh, The ETH experience: experimental database and [63] T. Mckeen, T. Pugsley, Simulation and experimental validation of a freely bubbling
results from the past 8 years, AIChE Symp. Ser. 95 (1999) 61–66. bed of FCC catalyst, Powder Technol. 129 (2003) 139–152.
[56] N. Zhang, B. Lu, W. Wang, J. Li, Virtual experimentation through 3D full-loop simu- [64] C. Chen, J. Werther, S. Heinrich, H.-Y. Qi, E.-U. Hartge, CPFD simulation of circulating
lation of a circulating fluidized bed, Particuology 6 (2008) 529–539. fluidized bed risers, Powder Technol. 235 (2013) 238–247.

You might also like