You are on page 1of 14

5548

Ind. Eng. Chem. Res. 2004, 43, 5548-5561

Simulation of Heterogeneous Structure in a Circulating


Fluidized-Bed Riser by Combining the Two-Fluid Model with the
EMMS Approach
Ning Yang,* Wei Wang, Wei Ge, Linna Wang, and Jinghai Li
Multiphase Reaction Laboratory, Institute of Process Engineering, Chinese Academy of Sciences,
P.O. Box 353, Beijing 100080, P. R. China

To consider the critical effect of mesoscale structure on the drag coefficient, this paper presents
a drag model based on the energy-minimization multiscale (EMMS) approach. The proposed
structure parameters are obtained from the EMMS model, and then the average drag coefficient
can be calculated from the structure parameters and further incorporated into the two-fluid
model to simulate the gas-solid flow in a circulating fluidized-bed riser. Simulation results
indicate that the simulated flow structures are different for the EMMS-based drag model and
the hybrid model using the Wen and Yu correlation and the Ergun equation. The former shows
its improvement in predicting the solids entrainment rate, the mesoscale heterogeneous structure
involving clusters or strands, and the radial and axial voidage distributions. The simulation
results support the idea that the average drag coefficient is an important factor for the twofluid model and suggest that the EMMS approach could be used as a kind of multiscale closure
law for drag coefficient.
1. Overview of Modeling Approaches
In circulating fluidized beds (CFBs), gas and particles
exhibit a complex flow behavior that is quite different
from that in the regime of dilute transport. For instance,
a relatively dense suspension exists with a solids volume
fraction ranging from 0.1 to 0.3 if a sufficiently effective
cyclone is provided to return the solids that escape from
the top to the lower part of the system, and the effective
slip velocities between gas and particles are usually 2030 times terminal velocity. To explain these complex
phenomena, Yerushalmi et al.1 and Li and Kwauk2
supposed that most particles did not remain single, but
tend to aggregate in the form of so-called clusters or
strands. Later, these ideas were confirmed by experimental investigations such as measurements employing
high-speed photography,3 video cameras and optical
fibers ,4 and capacitance probes.5 It was found that the
clusters are generally irregular in shape with variable
size. This leads to significant local heterogeneity at the
scale of both time and space, posing great challenges
for the analysis and modeling of gas-solid flow behavior
in CFB systems. Conventional empirical models, however, cannot be employed to investigate the flow structure extensively. Recent decades have witnessed the
emergence of models and simulation based on computational fluid dynamics (CFD).
Generally speaking, these models fall into three
categories, namely, two-fluid models, discrete-particle
models, and pseudo-particle models. The two-fluid
models (TFMs) are based on a pure Eulerian treatment
of both gas and solid phases as continuous and fully
interpenetrating, that is, both the gas and solid phases
are described by the locally averaged conservation
equations with closure laws including the respective
constitutive correlations for stress-strain relationship
of gas and solids, and the transfer correlations for
describing the interactions between the two phases.6-10
* To whom correspondence should be addressed. E-mail:
nyang@home.ipe.ac.cn.

The second category is the so-called discrete particle


model (DPM) in which the gas phase is treated as a
continuum and the solid phase is treated as discrete
particles. The gas phase is still described by locally
averaged conservation equations with the gas-solid
interaction terms representing the coupling between the
two phases; while the particle motion is resolved into
the suspension process in which individual particle is
tracked by the application of Newtons second law of
motion, and the collision process in which the hard
sphere or soft sphere approach is used to describe the
particle-particle interactions11-15. The third category
is a kind of pure Lagrangian model for both the gas and
solid phases, treating the fluid phase as composed of
pseudo-particles16 and attributing all the interactions
to collision processes between pseudo-particles and real
particles. For the last two categories, enormous computational resources are needed to handle the motion
of individual particles, thereby hindering, or perhaps
even rendering impossible, their application in industrial systems, at least at present. On the other hand,
two-fluid models are employed in current commercial
CFD packages to simulate real industrial systems
because of their relatively lower computational costs.
One of the key points for two-fluid models is how to
establish the constitutive correlations for the relationship between the solid stress and the strain. Some
researchers17,18 have used empirical correlations to
calculate the so-called solids viscosity and solids
pressure, whereas the kinetic theory of granular flow
(KTGF), originating from the kinetic theory for nonuniform dense gases, has been employed by others.19,20 In
the KTGF, the solids pressure and viscosity are regarded as functions of the so-called granular temperature, a measure of the particles velocity fluctuations.
An additional partial differential equation for the balance of the fluctuating energy of the particles was
required. Although the KTGF seems to be a more
reasonable approach to calculating the solids stress than
empirical correlations, Polashenski and Chen21 and

10.1021/ie049773c CCC: $27.50 2004 American Chemical Society


Published on Web 07/27/2004

Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004 5549

Bussing and Reh22 found that their experimental data


for solids stress were quite different from the results
predicted with the KTGF. Bussing and Reh22 suggested
further that particle clusters should be considered as a
momentum transferring structure for the KTGF.
Another key point for TFMs is how to establish the
correlations for the interphase drag coefficient, which
represents the intensity of momentum transfer between
the gas phase and the solid phase. In the literature,
several synonyms for this parameter have been used,
such as interphase friction coefficient, interphase exchange coefficient, and interphase momentum transfer
coefficient. All of them, in essence, refer to the average
drag coefficient for particles in a unit volume. In most
literature reports, the drag coefficient was calculated
by modifying the standard drag coefficient for individual
particles23,24 or was derived from the pressure drop
equations for fixed beds such as the Ergun equation.25
In other works, the drag coefficient was calculated either
from some pressure drop correlations such as that of
Gibilaro et al.26 or some correction factors such as that
proposed by Di Felice.27 Recently, Mostoufi and Chaouki28 and Yang and Renken29 proposed some improved
drag correlations based on their respective experimental
results. However, almost all of the above correlations
were derived from experimental results for liquid-solid
fluidized-bed or fixed-bed systems. The question is
whether these correlations could represent the real
mechanisms of gas-solid interactions in CFB systems,
which are, in common cases, more heterogeneous than
liquid-solid fluidized-bed or fixed-bed systems.
By using the so-called energy-minimization multiscale
(EMMS) model, Li et al.30 compared the calculation
results of slip velocities and drag coefficients for dense
cluster phases, dilute broth phases, and the interphase
between clusters and dilute broths, showing their
dramatic difference. Li and Kwauk31 investigated the
average drag coefficients for different connections of
dense and dilute phases, as well as the uniform distribution of particles in a unit volume, showing the strong
dependence of the drag coefficient on simple structural
changes. Their further study32 indicated that both local
and global structural changes led to a decrease in the
average drag coefficient. However, the present drag
coefficient correlations used in TFMs relate the drag
coefficient only to the local average voidage and slip
velocity, losing sight of the structure of control volumes.
OBrien and Syamlal33 reported that the present drag
correlations must be corrected to account for cluster
effects for fine particles and further claimed that the
unphysical adjustment of the solids stress was not the
correct approach. On the basis of the experimental
pressure drop data, they presented an empirical correlation for two operating conditions. Qi et al.34 reported
that the particles fed into the riser would be elutriated
immediately, leading to a rather dilute flow if the Ergun
equation was used to calculate the drag force in simulation. They claimed that the present drag correlations
were suitable only for lower gas velocities and coarse
particles, in which case the terminal velocity was equal
or close to the superficial gas velocity. Agrawal et al.35
reported that the so-called coarse-grid simulation,
which completely ignored the subgrid microstructure,
would overestimate the drag force. Zhang and VanderHeyden36 proposed a two-average approach for conservation equations to resolve the added-mass force and
drag force due to mesoscale structure from those at the

particle length scale, showing their essential roles in


the macroscopic momentum equations. Helland et al.37
studied the influence of porosity functions on correction
factors for the drag coefficient, indicating that homogeneous fluidization is unstable to small porosity perturbations because of nonlinear drag-inducing cluster
formation and that the collision parameters have a
significant influence on the shape and voidage of
clusters. The work of Li and Kuipers38 indicated also
that nonlinear drag plays an essential role in the
pattern formation of heterogeneous flow structures in
dense fluidized beds. Zhang and Reese39 reported that
the average drag force needs not only to incorporate the
influence of solids volume fraction but also to address
the effect of the random fluctuational motion of particles.
On the other hand, experimental results have suggested that the effective drag coefficient decreases upon
the formation of clusters and, therefore, that the Wen
and Yu correlation and Ergun equation were inadequate
for characterizing CFB systems. For example, Gunn and
Malik40 found that, if particles were grouped in aggregates for a given rate of fluid flow and overall voidage
of the system, the measured drag coefficient decreased
because of the decrease of the gas flow penetrating the
aggregates and the increase of the gas flow passing
around the aggregates. Muller and Reh41 found that the
formation of particle strands resulted in considerable
drag reduction, leading to a longer acceleration region
in risers. In view of the deficiency of the correlation of
Wen and Yu and the Ergun equation for CFB systems,
Li42 and Bai et al.43 proposed their respective empirical
correlations for the drag coefficient based on the average
values of experimental pressure drop data over the
whole cross-sectional area. Recently, Makkawi and
Wright44 also pointed out that it was far from accurate
for most numerical models to assume that the exponent
of the correction factor for the drag coefficient was a
constant, and they proposed several correlations of the
exponent of the correction factor for the wall and for
the center region of a fluidized bed on the basis of their
experimental measurements.
2. Motivation of the Present Study
Judging the survey of literature, it can be concluded
that the present drag correlations are too crude to reflect
the true mechanisms of gas-solid interactions for
heterogeneous systems, calling for a further investigation of the relationship between the drag coefficient and
the formation of mesoscale structure. In our opinion,
schemes to solve this problem can be classified into four
categories, as described below.
2.1. Tracking the Interface between Gas and
Particles. This category includes so-called direct numerical simulation (DNS), pseudo-particle modeling
(PPM), and the lattice-Boltzmann method (LBM). By
using these models, the interface between each particle
and the surrounding gas can be tracked and the
interactions between them need not to be described
explicitly. In DNS (Hu45), the moving body-fitted unstructured grids were frequently generated with the
motion of particles. The influence of particles on fluid
was reflected by the nonslip boundary condition of fluid
flow field which was then computed by solving the
Navier-Stokes equation, whereas the influence of fluid
on particles was reflected by the integration of the stress
on each element around the surface of particles. In PPM

5550 Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004

Figure 1. Multiscale resolution of structure and gas-solid interaction in the EMMS model.

(Ge and Li16) and LBM, the gas phase is resolved on a


length scale smaller than the particle size and the drag
force is calculated by the collision process between
pseudo-particles and real particles.
2.2. Refining the Computational Grids. The simulation work of some researchers35,46 showed that the
two-fluid model with high grid resolution, using the
conventional correlations for the average drag coefficient
or even using correlations for the standard drag coefficient for a single particle, is capable of simulating the
mesoscale structure. This might be due to the fact that
heterogeneity is weakened when the grid size is reduced.
2.3. Extracting the Closure Correlations from
Microscopic Simulations. With all of these modeling
approaches, it is generally agreed that a cascade description, i.e., extracting the closure correlations for the
drag coefficient from microscopic simulations, can suggest a practical way to explore the multiscale heterogeneity. For example, Agrawal et al.35 took the grid size
used in the coarse-grid simulation of a large unit as the
domain size, imposed periodic boundary conditions in
all directions, and carried out a two-fluid simulation to
obtain the average slip velocity, which could then be
used to extract an approximate drag law. Recently, an
attempt was made to compute the drag force with LBM
by Kandhai et al.47
2.4. Establishing a Structure Model for Control
Volumes. Although the above-mentioned schemes are
logical and fundamental, they are generally difficult to
implement at present because of the complexity of the
models and the enormous computational cost. A simple
but reasonable approach is to assume a structure for
each control volume. A structure model is established
to describe the multiscale heterogeneity through the
resolution of structure and the definition of structure
parameters. The average drag coefficient can then be
calculated from structure models.
The motivation of this study arises from the fourth
scheme because of its simplicity. The structure model
is established according to the so-called energy-minimization multiscale model (EMMS) proposed by Li and
Kwauk.48 The relationship between the average drag
coefficient for a control volume and the structure

parameters is deduced from the structure model. Also,


a simplified drag model is incorporated into the twofluid model to simulate the gas-solid flow behavior in
the riser section of a circulating fluidized bed. As an
extension of our previous work,49,50 this paper combines
the analysis of flow structure and the implementation
of combinations of the two models. Furthermore, this
approach is employed to investigate the effect of the
solids inventory on the axial voidage profile.
3. Dependence of the Drag Coefficient on the
Flow Structure
3.1. Reiteration of the Structure Model. Figure 1
shows the concept of structure resolution according to
the definition of the EMMS model. The heterogeneous
structure is resolved into a particle-rich dense cluster
phase and a gas-rich dilute phase. Eight parameters are
assigned to describe the phases, namely, c, f, dcl, Ugc,
and Upc for the dense phase and f, Ugf, and Upf for the
dilute phase. Correspondingly, the gas-solid interaction
is resolved into the interactions between gas and
particles within each of the phases and the interaction
between the dense cluster phase and the dilute broth
phase. In addition to the resolution of the structure and
of the gas-solid interactions, the EMMS model stipulates further that the total energy consumption per unit
mass of particles, NT, can be resolved into components
Nst that is needed for suspending and transporting
particles and Nd that is dissipated in other dissipative
processes. The particle-fluid flow is subject not only to
mass and momentum conservation but also to the
system stability condition, i.e., to the minimization of
the energy consumption, Nst, for suspending and transporting particles. For an extensive discussion of the
derivation and development of the EMMS model, the
interested reader is referred to the works of Li and
Kwauk,48 Xu and Li,51 Li et al.,52 and Ge and Li.53
Originally, this model was proposed to describe multiscale heterogeneity for global fluidized-bed systems.
In this study, we extend this model into the control
volume of the two-fluid model and take the average
acceleration, a, of a control volume into consideration.
The mathematical model is formulated by considering

Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004 5551

the mass and momentum balances, the correlations for


structure parameters, and the stability condition. The
mass balances for gas and solids require

Ug ) fUgc + (1 - f)Ugf

(1)

Up ) fUpc + (1 - f)Upf

(2)

For the dense phase, the resulting drag force comprises


that induced by gas-particle interactions inside the
dense phase and that induced by gas-cluster interactions between the cluster phase and the surrounding
dilute broth. For the dilute phase, the drag force refers
only to that induced by gas-particle interactions inside
the dilute phase. Then, the momentum equations for
particles in the dense and dilute phases per unit volume
can be expressed respectively as

f(1 - c)
f
3
3
CDc
FgUsc2 + CDi FgUsi2 )
4
dp
4 dcl
f(1 - c)(Fp - Fg)(g + a) (3)

The left-hand sides of eqs 6-8 represent the three


kinds of gas-solid interactions proposed by the EMMS
model, i.e., the interactions between gas and particles
in the dense phase (eq 6) and in the dilute phase (eq 8)
and the interaction between the dense cluster phase and
the dilute broth phase (eq 7). The right-hand sides,
correspondingly, represent the effective part of the
particle weight plus the inertial force balanced by the
three respective drag forces. In other words, these
equations show that only part of the gravitational and
inertial force for the dense phase (kc) is balanced by the
drag force induced by the gas-particle interactions
inside the dense phase, and the remaining part (ki) is
balanced by the gas-cluster interaction between the
dense cluster and the dilute broth, whereas all of the
gravitational and inertial force for the dilute phase is
balanced by the gas-particle interactions inside this
phase. Furthermore, the resultant drag force for all of
the particles per unit volume FD should be balanced by
the weight of particles plus the inertial force

FD )

Us
) (Fp - Fg)(1 - )(g + a)
 

(12)

and

3 (1 - f)(1 - f)
C
FgUsf2 )
4 Df
dp
(1 - f)(1 - f)(Fp - Fg)(g + a) (4)
where a denotes the average acceleration term. According to the EMMS model, the gas pressure drop in the
dense phase should be balanced by that in the dilute
phase plus that induced by the interaction between
dense clusters and the dilute broth

CDc

1 - c
1 - f
f
1
FgUsc2 ) CDf
FU 2+
FU 2
C
dp
dp g sf
1 - f Didcl g si
(5)

where represents the average drag coefficient for a


control volume (type A, see Gidaspow20) and Us is the
superficial slip velocity.
In addition to the mass and momentum equations,
an equation should be specified to relate the overall
voidage to those of the dense phase and the dilute phase,
i.e.

 ) cf + f(1 - f)

The hydrodynamic equivalent cluster diameter can be


correlated to the energy consumption for suspending
and transporting Nst, as proposed by Li and Kwauk48
and Xu and Kato54

dp

Rearrangement of eqs 3-5 leads to the following three


equivalent equations

dcl )

f(1 - c)

CDc FgUsc2 dp2 ) f(Fp - Fg)(g + a)(1 - c)kc


2
4
dp3/6
(6)
f

CDi FgUsi2 dcl2 ) f(Fp - Fg)(g + a)(1 - c)ki


2
4
dcl3/6
(7)
(1 - f)(1 - f)
3

dp /6

CDf FgUsf2 dp2 )


2
4
(Fp - Fg)(g + a)(1 - f)(1 - f)kf (8)

where

1-
kc )
1 - c

(9)

 - c
1 - c

(10)

ki )

kf ) 1

(11)

(13)

Up

(1 - max)
NstFp

(Fp - Fg)

)]

Umf + Upmf
(1 - mf)

(Umf + Upmf)
(1 - mf)

g
(14)

where

Nst ) Ug -

f - 
Fp - Fg
f(1 - f)Uf (g + a)
(15)
1-
Fp

Here, the parameter max denotes the maximum voidage


for the aggregates of particles (0.9997, see Matsen55).
Finally, the stability condition is proposed to close the
above equations. That is, the energy consumption for
suspending and transporting, Nst, with respect to the
total energy, NT, tends to be minimal at the stable
condition so that

Nst [Ug(1 - ) - fUf(f - )(1 - f)]


)
) min
NT
Ug(1 - )

(16)

where

NT )

Fp - Fg
Ug(g + a)
Fp

(17)

5552 Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004
Table 1. Summary of Related Expressions for the EMMS Model
dense phase

dilute phase

-4.7

effective drag coefficient

CDc ) CD0cc

standard drag coefficient

CD0c )

-4.7

CDi ) CD0i(1 - f)-4.7

CDf ) CD0ff

24
3.6
+
Rec Re 0.313

CD0f )

FgdpUsc
g

characteristic Reynolds number

Rec )

superficial slip velocity

Usc ) Ugc -

drag force acting on a


single particle or cluster

dp2 Fg 2
U
Fdense ) CDc
4 2 sc

number of particles or
clusters per unit volume

Mc )

f(1 - c)

CD0i )

FgdpUsf
g

Rei )

24
3.6
+
Rei Re 0.313
i

FgdclUsi
g

Usf ) Ugf -

fUpf
1 - f

Usi ) Ugf -

Fdilute ) CDf

dp2 Fg 2
U
4 2 sf

Fcluster ) CDi

Mf )

dp3/6

24
3.6
+
Ref Re 0.313
f

Ref )

cUpc
1 - c

interphase

(1 - f)(1 - f)
dp3/6

Mi )

fUpc
(1 - f)
1 - c

dcl2 Fg 2
U
4 2 si

f
dcl3/6

FD ) McFdense + MiFcluster + MfFdilute

resulting drag force

FD

average acceleration

a)

average drag coefficient:

FD
)
Us/

(Fp - Fg)(1 - )

-g

The relevant parameters in eqs 1-17 are listed in


Table 1. With the value of Ug, Gs, and  specified, the
structure parameters can be calculated by solving the
above nonlinear equations, which constitute a nonlinear
optimization problem.
3.2. More Insight into the Structural Heterogeneity. Computation was carried out with the specified
superficial gas and solids velocities for the fluid catalytic
cracking (FCC)/air system (Fp ) 930 kg/m3, dp ) 54 m,
Ug ) 1.52 m/s, Gs ) 14.3 kg/m2s). Our previous work49
showed that, with decreasing average voidage , the
solids volume fraction in the dense phase f(1 - c)
increased and that in the dilute phase (1 - f)(1 - f)
decreased. Meanwhile, the superficial gas velocity in the
dilute phase, Ugf, increased, and that in the dense phase,
Ugc, decreased. The calculation results indicate that the
particles tend to enter into clusters instead of suspending in dilute broth, whereas the gas tends to pass
around, instead of penetrating through, the dense
cluster phase. These conclusions are conceptually consistent with the observed phenomena in particle-fluid
systems, and, in turn, they show the reasonableness of
the EMMS approach.
As defined in eqs 9 and 10, kc and ki indicate the
effective part of the weight plus inertial force in the
dense phase balanced by different gas-solid drag forces,
and their variation is closely related to  and c. Our
previous work49 showed that c f mf when  f mf,
whereas c f  when  f 1. Therefore, we can deduce
from eqs 9 and 10 that kc f 1 and ki f 0 when  f mf,
whereas kc f 0 and ki f 1 when  f 1. Figure 2 shows
that kc increases whereas ki decreases with decreasing
average voidage . Because of the difficulties inherent
in solving the nonlinear equations when  f max, the
right sides of the curves are plotted with dashed lines.
The left ( f mf) and right ( f 1) sides of the curves
can be considered to represent the two extreme cases
of fluidization respectively, i.e., particulate fluidization
and dilute transport. Thus, we found that the absolute
value of the difference, k, between kc and ki, as
represented by vertical lines in Figure 2, can be used
to estimate the extent of heterogeneity of the system.

Figure 2. Variation of structure parameters with average voidage.

Larger values of k represent weaker heterogeneity of


the system. The intersection of the two curves where
k is equal to zero indicates that the gravitational and
inertial forces of the dense cluster phase is one-half
balanced by the drag force resulting from gas-particle
interactions inside the dense phase, whereas the other
half is balanced by the drag force due to the gas-cluster
interactions. At the right side of the intersection, k
increases with increasing voidage, indicating that the
system heterogeneity decreases quickly and approaches
the state of dilute transport in the extreme case. At the
left side, k first increases gradually with decreasing
voidage, indicating essentially the decrease in heterogeneity; its sudden increase at a turning point indicates
the tendency that the system becomes homogeneous
more quickly and approaches another extreme case,
namely, particulate fluidization. The turning point here
can be considered to reflect a kind of regime transition
and also appeared in the variation of other structure
parameters such as f, dcl, and c with voidage, as
presented in our previous work.49
3.3. Structure-Dependent Drag Coefficient. The
overall momentum equation (eq 12) shows that the
average acceleration term a is determined by only the
average drag coefficient if the superficial slip velocity
and the average voidage are specified, indicating that,

Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004 5553

Figure 3. Variation of drag coefficients with average voidage.

if cannot be accurately calculated for the two-fluid


model, a and its related flow field cannot be obtained
correctly. In fact, it is the average acceleration term a
that determines the simulation accuracy of the flow field
and the corresponding concentration field when twofluid models are used.
Summing the left and right sides of eqs 6-8 leads to

FD ) McFdense + MiFcluster + MfFdilute

(18)

Combining eq 12 and 18, one obtains

2
(M F
+ MiFcluster + MfFdilute)
Us c dense

3 (1 - )
Fg|ug - up|CD0-2.7
4 dp

(22)

Comparing eqs 19 and 22, one can see that the structure
information is considered in the former, whereas it is
neglected in the latter.
Figure 3a compares the results of correction factors
for effective drag coefficients computed using the EMMS
approach and the Wen and Yu/Ergun correlations. The
correction factors are calculated from

) /0

(23)

where

(1 - ) 1

F (u - up)|ug - up|2 dp2CD0-4.7


3 2 g g
4
d
6 p
(20)

The resulting drag force per unit volume is also defined


in the two-fluid model as

FD ) (ug - up)


(19)

where the relevant parameters, as given in Table 1, can


be calculated from the structure parameters.
For comparison, we now give the derivation of the
drag correlation used in the common two-fluid models.
Assuming that the effective drag force for each particle
in a control volume is the same, i.e., ignoring the effect
of the heterogeneous structure, the resultant drag force
for all of the particles per unit volume, FD, should be
equal to the effective drag force for a single particle
multiplied by the number of particles in the unit volume

FD )

tion employed in the present two-fluid models

(21)

Combining eqs 20 and 21, one obtains the drag correla-

0 )

3 (1 - )
Fg|ug - up|CD0
4 dp

(24)

It can be seen that the correction factor computed with


the EMMS model is much smaller than that computed
with the Wen and Yu/Ergun correlations, which is in
reasonable agreement with the conclusions from experimental observations that the drag coefficient decreases
as a result of cluster formation. At the left side of the
curve calculated with the EMMS model, the correction
factor increases abruptly with decreasing average voidage at the previously discussed turning point and
approaches the curve calculated with the Ergun equation. The extended dashed lines at the right side of
curves indicate that the correction factors from both the
Wen and Yu correlation and the EMMS model approach
unity. This extreme value corresponds to the standard
drag coefficient.
Figure 3b compares the average drag coefficient
computed with the EMMS model and the Wen and Yu/
Ergun correlations. It also shows that the value computed with the EMMS model is much less than that
computed with the Wen and Yu/Ergun correlations.

5554 Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004

Furthermore, the two curves tend to coincide with each


other at the left and right sides.
Figure 3c shows the average acceleration a at the
cross section of a CFB as computed with the EMMS
model and the Wen and Yu/Ergun correlations. The data
for the local superficial gas velocity, solids flow rate, and
average voidage are taken from the experimental results
of Bader et al.56 It can be seen that using the Wen and
Yu correlation leads to much higher values of the
average acceleration than using the EMMS model,
especially in the wall region. The higher value of
acceleration seems to be unreasonable unless a large
part of it could be balanced by that due to particleparticle interactions. Muller and Reh41 reported that,
for a particle strand of 1-mm diameter falling in
stagnant air, the particle velocity is much higher than
the terminal velocity for a single suspended particle, and
the acceleration of particles inside the strand is equal
to the acceleration of free fall. The latter result implies
that no drag force acts on the particles in the strand.
In some literature reports, the effect of the drag force
is expressed as a function of the average particle
relaxation time , which indicates the speed of response
for particles to entrainment by the surrounding gas and
can be generally formulated as

Us
F U 2 d 2C ) Fp dp3
2 g s4 p D
6


(25)

Thus, the average relaxation time can also be given


by

Us
(g + a)

(26)

Figure 3d shows the computed particle average relaxation time . The results indicate that the value
computed with the Wen and Yu correlation is much less
than that computed with the EMMS model. This implies
that, for the Wen and Yu correlation, the response of
particles to the entrainment of surrounding gas is very
fast, and the simulated solids entrainment rate for gas
would be higher than that from the EMMS model.
4. Combination with the Two-Fluid Model
4.1. Model Description. Table 2 summarizes the
mass and momentum conservation equations and constitutive equations of the two-fluid model. For simplicity, empirical models were employed to calculate the
solids pressure and viscosity. The gradient term of solids
pressure was employed to prevent the solid phase from
being too densely packed. The correlation for solids
viscosity was taken from the work of Miller and Gidaspow.57 To compare the effect of the average drag
coefficients, we adopted two kinds of drag correlations.
Drag model A refers to the hybrid of the Wen and Yu
correlation and the Ergun equation. For drag model B,
we employ a simplified version of the EMMS model to
derive the correction factors for different cases of
operating conditions, as shown in Table 2. Because the
computation procedure for solving the nonlinear equations involves two layers of iteration and the optimization process for the EMMS model, it would be difficult
to directly apply this model to each control volume with
the local superficial gas velocity and solids flow rate as
the input parameters. To circumvent this problem, we
first simplified the model by assuming that the voidage

of the dense phase is a constant. Then, with the overall


superficial gas velocity and solids flow rate as the input
parameters, we can calculate the average drag coefficient for a specified voidage and thereby obtain a
correlation for the relationship () between the correction factor for the drag coefficient and the voidage.
Extending this correlation into each control volume and
employing the local slip velocity and average voidage,
the average drag coefficient for each control volume
can thus be calculated as

3 (1 - )
Fg|ug - up|CD0()
4 dp

(27)

The simulation was carried out for the riser section


of a circulating fluidized bed, as illustrated in Figure 4.
The gas entered the riser from the bottom with a
specified superficial velocity and left the bed at the top
from two side outlets. Solids were first stacked up to a
certain height in the bottom of the riser. When fluidized,
the solids were allowed to leave from the top outlets and
in turn be fed back into the riser from the bottom inlets
by specifying that the solids flow rate at the two side
inlets was equal to that at the top outlets. The initial
bed height was evaluated from the experimental data
of the axial voidage profiles in the riser. For the
operating conditions with the same superficial gas
velocities and solids circulation rates but different solids
inventories, the evaluated initial bed heights were
different because of the difference in experimental axial
voidage profiles. In this way, we could consider the effect
of solids inventory on the simulated axial voidage
profiles. The initial bed voidage was set to be incipient
fluidization voidage. For simplicity, the grids adopted
in the simulation were uniformly distributed in both the
axial and radial directions. For the gas phase, the
nonslip boundary condition was used at the wall, so that

ug,w ) vg,w ) 0

(28)

For the solid phase, the free-slip boundary condition was


chosen (Ding and Gidaspow58)

up,w ) 0,

vp,w ) -

dp
1/3

(1 - )

vp,w
x

(29)

The simulation was carried out by employing the


commercial CFD package CFX4.4 (AEA Technology) in
which the above-mentioned two drag models were
incorporated through user-defined routines. The highly
coupled partial differential equations of the two-fluid
model were solved by the interphase slip algorithm
(IPSA) of Spalding.59 The simulation parameters are
listed in Table 3.
4.2. Results and Discussion
4.2.1. Solids Entrainment Rate. The solids mass
flux at the top outlets can be obtained by dynamically
checking the outlet solids velocity and concentration in
this location during the simulation. Figure 5 illustrates
the variation of the outlet solids mass flux with time
for two solids inventories. At initial times (t < 5 s), the
outlet solids mass flux is almost equal to zero. Because
the bed expands with the incipient fluidization voidage
in this period, the Ergun equation is always used by
both drag models. After this period, the curves of outlet
solids flux oscillate in different ways. For drag model

Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004 5555
Table 2. Mathematical Equations in the Two-Fluid Model
Conservation Equations
(1) Mass Conservation Equations
(a) gas phase

(F ) + (Fgug) ) 0
t g

(b) solid phase

[(1 - )Fp] + [(1 - )Fpup] ) 0


t
(2) Momentum Conservation Equations

(a) gas phase

(F u ) + (Fgugug) ) -pg + Tg - (ug - up) + gFgg


t g g

(b) solid phase

[(1 - )Fpup] + [(1 - )Fpupup] )


t
-(1 - )pg + Tp - (up - ug) + (1 - )Fpg
Constitutive Equations
(1) Stress-Strain Correlations

(a) gas phase

2
Tg ) g [ug + (ug)T] - (ug)I
3

(b) solid phase

2
Tp ) -ppI + (1 - )p [up + (up)T] - (up)I
3

(a) solids viscosity


(b) solids pressure

(2) Model for Solids Viscosity and Pressure


p ) 0.5(1 - )
pp ) 10-8.686+6.385(1 - )
(3) Model for Gas-Solid Interaction

(a) drag model A

(b) drag model B

{
{

3 (1 - )
Fg|ug - up|CD0-2.7
4 dp
2

(1 - ) g

150

dp

(1 - )2g
2

dp

( > 0.8)

(1 - )Fg|ug - up|
( e 0.8)
+ 1.75
dp

3 (1 - )
Fg|ug - up|CD0()
4 dp
150

( > *)

(1 - )Fg|ug - up|
( e *)
+ 1.75
dp

* ) 0.74 for Ug ) 1.52 m/s, Gs ) 14.3 kg/m2s

0.0214
(* <  e 0.82)
4( - 0.7463)2 + 0.0044
0.0038
() ) -0.0101 +
(0.82 <  e 0.97)
4( - 0.7789)2 + 0.0040
( > 0.97)
- 31.8295 + 32.8295
-0.5760 +

A, the curve rises abruptly and falls back because of


the resistance of the outlets, and finally, it oscillates
steadily at an average of about 39.03 kg/m2s for I ) 15
kg. Considering drag model B, the outlet solids mass
flux always oscillates steadily around an average of
about 12.17 kg/m2s at the same solids inventory.
According to the experiments of Li and Kwauk,48 the
outlet solids flux (14.3 kg/m2s) was calculated by
measuring the pressure drop and accumulated bed
height above the closed butterfly valve in the standpipe
within a certain time. It can be seen that drag model A
overpredicted whereas drag model B well predicted the
solids mass flux. The higher solids entrainment rate
predicted by drag model A obviously results from the
overprediction of the average drag coefficient and
therefore the average acceleration a for each control
volume. The same phenomenon can also be found for I
) 20 kg in Figure 5b that drag model A greatly
overpredicted the average value of the solids outlet mass
flux (77.8 kg/m2s), whereas the simulation result from
drag model B (12.1 kg/m2s) was comparable to the
experimental result.

4.2.2. Dynamic Flow Structure. Figures 6 and 7


illustrate the gas and solids velocity vectors from drag
models A and B, respectively, at t ) 20 s. Figure 8 shows
snapshots of the contours for the voidage distribution
at t ) 20 s. It can be seen that the simulated flow
structures for the two drag models are different. For
drag model A, the directions of the gas and solids
velocity vectors are always upright, and the distribution
of solids concentration is rather homogeneous across the
whole bed except for the denser region near the two side
inlets at the bottom. The phenomena predicted by drag
model A are much more like the flow structure of dilute
transport than that of fast fluidization. For drag model
B, the solids velocity vectors are upright at the top and
chaotic at the bottom, showing that the lateral motion
of the solids is strong at the bottom, and the solids
themselves are entrained upward in a snakelike wriggling way. The corresponding snapshot of the solids
concentration distribution, as shown in Figure 8, shows
that the top section of the bed is dilute and the bottom
section is dense and in a turbulent state. Clearly
observed is the formation of clusters protruding from

5556 Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004

Figure 4. Schematic drawing of the simulated riser.


Table 3. Parameter Settings for the Simulation
parameter

value

particle diameter
particle density
time interval
grid size (x)
grid size (y)
maximum voidage for particle aggregation
voidage of the dense phase
superficial gas velocity
solids inventory I ) 15 kg
solids inventory I ) 20 kg

54 m
930 kg/m3
5.0 10-4 s
2.25 10-3 m
3.5 10-2 m
0.9997
0.69
1.52 m/s
H ) 1.225 m
H ) 1.855 m

the wall and of strands above the bottom dense region.


Particles fall along the wall and then stack together to
form radial protrusions. Some particles are dynamically
extracted from the protrusions by the up-flowing gas
and reaggregate into strands above the bottom dense
region. The cluster protrusions resist the motion of upflowing gas and solids, leading consequently to the
wriggling movement. These phenomena predicted by
drag model B are in reasonable agreement with the
experimental observations of Rhodes et al.,60 who found
that the predominant motion is radially inward at the
interface between the upper dilute region and the lower
dense region. Davidson61 insisted that clusters of particles detached from the film falling against the wall
and further decomposed in the dilute phase to contribute particles to the dilute phase. Because coarse grids
are used in this study, the simulated structure might
represent a kind of group behavior of microstructures.
4.2.3. Time-Average Flow Structure. The timeaverage voidage can be obtained by averaging the
calculated transient values over 20-30 s, and the timeaverage velocity for a control volume can be obtained
as follows

ui,j )

t i,jui,j
t i,j

(30)

where i,j and ui,j denote the transient voidage and


velocity for a control volume, respectively.

Figure 5. Comparison of outlet solids mass flux.

Figure 6. Simulated velocity vector with drag model A at t ) 20


s: (a) gas, I ) 15 kg; (b) solids, I ) 15 kg; (c) gas, I ) 20 kg; (d)
solids, I ) 20 kg.

Figure 9 shows the radial profiles of the resulting


time-average voidage based on drag models A and B for
I ) 15 kg. Figure 10 displays the radial voidage profiles
for I ) 20 kg. With experimentally measured voidages
obtained using optical fibers, Tung et al.62 and Zhang
et al.63 correlated the radial voidage profile with the
cross-sectional average voidage, thus allowing easy
comparisons of our simulated profiles with calculations
from their correlation. It can be seen that, for the

Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004 5557

Figure 7. Simulated velocity vector with drag model B at t ) 20


s: (a) gas, I ) 15 kg; (b) solids, I ) 15 kg; (c) gas, I ) 20 kg; (d)
solids, I ) 20 kg.

EMMS-based model (drag model B), the simulated


results are in good agreement with that of experimental
correlation in the dense annulus, but underpredict the
voidage at the dilute core. Because most particles are
aggregated in the annulus for fast fluidization, the
underprediction of the voidage in the dilute core would
have little effect on the accuracy of hydrodynamic
simulations. It should be noted that the simulated
voidage near the wall is generally larger than that from
Tung et al.s correlation. According to the recent work

of Xu et al.,64 this might indicate that the simulation is


more representative of experiment because the Tung et
al. correlation actually underestimates the voidages on
the wall. For the hybrid model based on the Wen and
Yu correlation and the Ergun equation (drag model A),
the simulated radial voidage profile is rather flat, which
actually shows the homogeneous structure discussed in
the previous sections. It can be seen that the simulation
results obtained with drag model A deviate significantly
from the experimental correlation, overpredicting the
voidage in the annulus and underpredicting the voidage
in the core. Figure 10b shows that the simulated voidage
obtained with drag model A is much lower than those
measured in experiments and obtained from the simulation with drag model B. Because the solids entrained
to the outlet are all fed back into the inlet, the average
solids concentration for the whole riser is always held
as a constant. For drag model A, the overprediction of
the voidage in the bottom annular region would lead to
an underpredicted voidage for the top region. If a solidsfeeding loop were not allowed and a constant solids flow
rate were specified at the bottom inlets, the solids
concentration predicted by drag model A would be
rather small. In this case, using drag model A would
overpredict the voidages throughout the bed, and the
system then appears to transport particles quickly, as
reported by Qi et al.34
Figure 11 shows the time-average axial voidage
profiles from calculations based on the two drag models.
The plotted experimental axial voidage profiles are from
Li and Kwauk.48 For drag model A, the resulting axial
voidage profile is almost vertical except near the inlet
and outlet. Using drag model B improves the prediction,
leading to the S-shaped voidage profile, although deviations still exist probably because of the inaccurate
evaluation of the solids inventory in the riser at the
initial time and the unrealistic setup of the inlet and
outlet boundary conditions in the 2D simulation.
To illustrate the effect of the solids inventory on the

Figure 8. Simulated voidage distribution at t ) 20 s: (a) I ) 15 kg, drag model A; (b) I ) 15 kg, drag model B. (c) I ) 20 kg, drag model
A; (d) I ) 20 kg, drag model B.

5558 Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004

Figure 11. Axial voidage profile for (a) I ) 15 kg, (b) I ) 20 kg:
- - -, drag model A; s, drag model B; 9, experimental.

Figure 9. Radial voidage profile for I ) 15 kg.

Figure 12. Effect of solids inventory on axial voidage profile:


- - -, I ) 15 kg; s, I ) 20 kg.

Figure 10. Radial voidage profile for I ) 20 kg.

axial voidage profile, results for the axial voidage profile


simulated with drag model B are replotted in Figure
12. As discussed previously, the setting of the solids
inventory in the simulation is achieved by setting
different initial bed heights evaluated from experimental results of axial voidage profiles corresponding to
different solids inventories. It can be seen that the only
difference between the two curves is the inflection point.
The cross-sectional voidages in both the dilute top region
and the dense bottom region are invariable. Moreover,
the simulated values of the outlet solids mass flux for

the two solids inventories, as shown in Figure 5, are


almost the same, indicating that the entrainment
reaches the state of saturation. This reasonably agrees
with the conclusions from experimental findings that,
when the solids inventory I is increased within a certain
range at a given superficial gas velocity Ug, the solids
flow rate Gs remains essentially constant, and the only
variation for the S-shaped voidage profile is the movement of the inflection point. Therefore, it is insufficient
to determine the axial voidage profile by specifying only
Ug and Gs, as reported by Li et al.31 and Xu and Gao.65
Figure 13 shows the time-average axial velocity
profiles for gas and solids in radial direction for I ) 15
kg obtained with the two drag models. It can be seen
from Figure 13a and c that, for drag model A, the shape
of the solids velocity curve looks too similar to that of
the gas velocity, indicating that the solids track the gas
flow very quickly and the particle relaxation time should
be very short. This phenomenon can also be found in
Figure 6, where the gas and solids velocity fields in the
whole bed simulated using drag model A are almost the
same, showing the quick response of the particles to gas
entrainment as well. The results from the simulation
based on drag model B are obviously different, especially
in the bottom region, as shown in Figure 7. Moreover,
the velocity fields of the gas and solids simulated with
drag model A are higher than those obtained with drag

Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004 5559

Figure 13. Simulated velocity profile for gas and solids for I ) 15 kg.

model B. Thus, the choice of a suitable drag model is


necessarily needed for CFD simulation.
5. Conclusions and Prospects
Complexity in gas-solid flow in circulating fluidized
beds lies in the formation of heterogeneous structures
existing at different time and space scales. Although the
emergence of CFD-based models provides a powerful
tool for simulating flow structure in CFB systems,
models that can reflect the heterogeneous nature arising
from the mutual compromise among different kinds of
interactions are lacking for such complex systems. The
objective of this study was to understand the relationship between the mesoscale heterogeneity and the gassolid drag coefficient. Compared to the hybrid drag
model of the Wen and Yu correlation and the Ergun
equation, the drag model presented in this study was
established on the basis of a structure model that takes
the effect of mesoscale structure into consideration.
Consequently, a combination of the drag model with the
two-fluid model shows the reasonableness and improvement in simulations of the dynamic formation of local
heterogeneous structure, such as the formation of
clusters or strands, and in predictions of the timeaverage macroscale heterogeneity. Although this EMMSbased model needs to be further validated for other
systems, it would be expected to be used as a closure
law for drag coefficients.
Recently, Ge and Li53 proposed that the two-fluid
model could be extended to the so-called two-phase twofluid model. This means that the particles and fluid in
the dense and dilute phases could all be regarded as
continua, with similar but double average conservation
equations of the present TFM. The interactions between
each pair of them could be determined in a further
generalized version of the EMMS-based drag model.
Therefore, future work might be needed to improve the
proposed model and incorporate it into a two-phase twofluid model. It is also worthwhile to establish a kind of

multiscale approach to evaluate the relative dominance


and mutual compromise among different interactions
occurring between gas and solids, solids and solids, and
mixture and wall. This might allow for a proper evaluation of the magnitude of the acceleration terms induced
by these interactions in the average conservation equations of the two-fluid models.
Acknowledgment
This work was financially supported by the National
Natural Science Foundation of China (Nos. 20336040,
20221603).
Notation
a ) average acceleration of particles for a control volume,
m/s2
CD ) effective drag coefficient for a particle
CDc, CDf ) effective drag coefficients for a particle in the
dense phase and the dilute phase, respectively
CDi ) effective drag coefficient for a cluster
CD0 ) standard drag coefficient for a particle
dcl ) cluster diameter, m
dp ) particle diameter, m
f ) volume fraction of dense phase
g ) gravitational acceleration, m/s2
Gs ) solids flow rate, kg/m2s
H ) initial bed height, m
I ) solids inventory, kg
Nst ) mass-specific energy consumption for suspending and
transporting particles, W/kg
NT ) mass-specific total energy consumption for particles,
W/kg
Ug ) superficial fluid velocity, m/s
Ugc, Ugf ) superficial fluid velocities in the dense phase
and in the dilute phase, respectively, m/s
Umf ) superficial fluid velocity at minimum fluidization,
m/s
Up ) superficial particle velocity, m/s

5560 Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004
Upc, Upf ) superficial particle velocities in the dense phase
and in the dilute phase, respectively, m/s
Us ) superficial slip velocity, m/s
Usc, Usf, Usi ) superficial slip velocities in the dense phase,
the dilute phase, and the interphase, respectively, m/s
Greek Symbols
) drag coefficient in a control volume, kg/m3s
 ) voidage
c, f ) voidages of the dense phase and the dilute phase,
respectively
max ) maximum voidage for particle aggregation (0.9997)
mf ) voidage at minimum fluidization (0.5)
g, p ) gas or solids viscosities, respectively, Pas
Fg, Fp ) fluid and solids densities, respectively, kg/m3
) average particle relaxation time, s
) correction factor

Literature Cited
(1) Yerushalmi, J.; Turner, D. H.; Squires, A. M. The Fast
Fluidized Bed. Ind. Eng. Chem. Process Des. Dev. 1976, 15, 4753.
(2) Li, Y.; Kwauk, M. The Dynamics of Fast Fluidization. In
Fluidization; Grace, J. R., Matsen, J. M., Eds.; Plenum Press: New
York, 1980; pp 537-544.
(3) Yerushalmi, J.; Squires, A. M. The Phenomenon of Fast
Fluidization. AIChE Symp. Ser. 1977, 72, 44-47.
(4) Li, H.; Xia, Y.; Tung, Y.; Kwauk, M. Micro-visualization of
Clusters in a Fast Fluidized Bed. Powder Technol. 1991, 66, 231235.
(5) Brereton, C. M. H.; Grace, J. R. Microstructural Aspects of
the Behavior of Circulating Fluidized Beds. Chem. Eng. Sci. 1993,
48, 2565-2568.
(6) Gidaspow, D. Hydrodynamics of Fluidization and Heat
Transfer: Supercomputer Modeling. Appl. Mech. Rev. 1986, 39,
1-23.
(7) Kuipers, J. A. M.; Van Duin, K. J.; Van Beckum, F. P. H.;
Van Swaaij, W. P. M. A Numerical Model of Gas-Fluidized Beds.
Chem. Eng. Sci. 1992, 47, 1913-1924.
(8) Enwald, H.; Peirano, E.; Almstedt, A. E. Eulerian TwoPhase Flow Theory Applied to Fluidization. Int. J. Multiphase
Flow 1996, 22, 21-66.
(9) Samuelsberg, A.; Hjertager, B. H. Computational Modeling
of Gas/Particle Flow in a Riser. AIChE J. 1996, 42, 1536-1546.
(10) Nieuwland, J. J.; Van Sint Annaland, M.; Kuipers, J. A.
M.; Van Swaaij, W. P. M. Hydrodynamic Modeling of Gas/Particle
Flows in Riser Reactors. AIChE J. 1996, 42, 1569-1582.
(11) Tsuji, Y.; Kawaguchi, T.; Tanaka, T. Discrete Particle
Simulation of Two-Dimensional Fluidized Bed. Powder Technol.
1993, 77, 79-87.
(12) Hoomans, B. P. B.; Kuipers, J. A. M.; Briels, W. J.; Van
Swaaij, W. P. M. Discrete Particle Simulation of Bubble and Slug
Formation in a Two-Dimensional Gas-Fluidised Bed: A HardSphere Approach. Chem. Eng. Sci. 1996, 51, 99-118.
(13) Xu, B. H.; Yu, A. B. Numerical Simulation of the GasSolid Flow in a Fluidized Bed by Combining Discrete Particle
Method with Computational Fluid Dynamics. Chem. Eng. Sci.
1997, 52, 2785-2809.
(14) Ouyang, J.; Li, J. H. Discrete Simulations of Heterogeneous
Structure and Dynamic Behavior in Gas-Solid Fluidization.
Chem. Eng. Sci. 1999, 54, 5427-5440.
(15) Kafui, K. D.; Thornton, C.; Adams, M. J. Discrete ParticleContinuum Fluid Modelling of Gas-Solid Fluidised Beds. Chem.
Eng. Sci. 2002, 57, 2395-2410.
(16) Ge, W.; Li, J. H. Macroscale Phenomena Reproduced in
Microscopic SystemssPseudo-Particle Modeling of Fluidization.
Chem. Eng. Sci. 2003, 58, 1565-1585.
(17) Benyahia, S.; Arastoopour, H.; Knowlton, T. M. Prediction
of Solid and Gas Flow Behavior in a Riser Using a Computational
Multiphase Flow Approach. In Fluidization IX; Fan, L. S., Knowlton, T. M., Eds.; Engineering Foundation: New York, 1998; pp
493-500.
(18) Sun, B.; Gidaspow, D. Computation of Circulating Fluidized-Bed Riser Flow for the Fluidization VIII Benchmark Test.
Ind. Eng. Chem. Res. 1998, 38, 787-792.

(19) Sinclair, J. L.; Jackson, R. Gas-Particle Flow in a Vertical


Pipe with Particle-particle Interaction. AIChE J. 1989, 35, 14731486.
(20) Gidaspow, D. Multiphase Flow and Fluidization: Continuum and Kinetic Theory Description; Academic Press: Boston,
1994.
(21) Polashenski, W.; Chen, J. C. Measurement of Particle
Phase Stresses in Fast Fluidized Beds. Ind. Eng. Chem. Res. 1999,
38, 705-713.
(22) Bussing, W.; Reh, L. On Viscous Momentum Transfer by
Solids in Gas-Solids Flow through Risers. Chem. Eng. Sci. 2001,
56, 3803-3813.
(23) Richardson, J. F.; Zaki, W. N. Sedimentation and Fluidization. Trans. Inst. Chem. Eng. 1954, 32, 35-53.
(24) Wen, C. Y.; Yu, Y. H. Mechanics of Fluidization. Chem.
Eng. Prog. Symp. Ser. 1966, 62, 100-111.
(25) Ergun, S. Fluid Flow through Packed Columns. Chem.
Eng. Prog. 1952, 48, 89-54.
(26) Gibilaro, L. G.; Di Felice, R.; Waldram, S. P.; Foscolo, P.
U. Generalized Friction Factor and Drag Coefficient Correlations
for Fluid-Particle Interactions. Chem. Eng. Sci. 1985, 40, 18171823.
(27) Di Felice, R. The Void Function for Fluid-Particle Interaction Systems. Int. J. Multiphase Flow 1994, 20, 153-159.
(28) Mostoufi, N.; Chaouki, J. Prediction of Effective Drag
Coefficient in Fluidized Beds. Chem. Eng. Sci. 1999, 54, 851-858.
(29) Yang, J. Z.; Renken, A. A. Generalized Correlation for
Equilibrium of Forces in Liquid-Solid Fluidized Beds. Chem. Eng.
J. 2003, 92, 7-14.
(30) Li, J. H.; Chen, A.; Yan, Z.; Xu, G.; Zhang, X. ParticleFluid Contacting in Circulating Fluidized Beds. In Preprint
Volume for CFB-IV; Avidan, A. A., Ed.; AIChE: New York, 1993;
pp 49-54.
(31) Li, J. H.; Wen, L.; Ge, W.; Cui, H.; Ren, J. Dissipative
Structure in Concurrent-up Gas-Solid Flow. Chem. Eng. Sci.
1998, 53, 3367-3379.
(32) Li, J. H.; Kwauk, M. Exploring Complex Systems in
Chemical EngineeringsThe Multi-scale Methodology. Chem. Eng.
Sci. 2003, 58, 521-535.
(33) Obrien, T. J.; Syamlal, M. Particle Cluster Effects in the
Numerical Simulation of a Circulating Fluidized Bed. In Preprint
Volume for CFB-IV; Avidan, A. A., Ed.; AIChE: New York, 1993;
pp 430-435.
(34) Qi, H.; You, C.; Boemer, A.; Renz, U. Eulerian Simulation
of Gas-Solid Two-Phase Flow in a CFB-Riser under Consideration
of Cluster Effects. In Fluidization 2000: Science and Technology;
Xu, D., Mori, S., Eds.; Xian Publishing House: Xian, China, 2000;
pp 231-237.
(35) Agrawal, K.; Loezos, P. N.; Syamlal, M.; Sundaresan, S.
The Role of Mesoscale Structures in Rapid Gas-Solid Flows. J.
Fluid Mech. 2001, 445, 151-185.
(36) Zhang, D. Z.; VanderHeyden, W. B. The Effects of Mesoscale Structures on the Macroscopic Momentum Equations for TwoPhase Flows. Int. J. Multiphase Flow 2002, 28, 805-822.
(37) Helland, E.; Occelli, R.; Tadrist, L. Numerical Study of
Cluster Formation in a Gas-Particle Circulating Fluidized Bed.
Powder Technol. 2000, 110, 210-221.
(38) Li, J.; Kuipers, J. A. M. Gas-particle Interactions in Dense
Gas-Fluidized Beds. Chem. Eng. Sci. 2003, 58, 711-718.
(39) Zhang, Y.; Reese, J. M. The Drag Force in Two-Fluid
Models of Gas-Solid Flows. Chem. Eng. Sci. 2003, 58, 1641-1644.
(40) Gunn, D. J.; Malik, A. A. The Structure of Fluidized Beds
in Particulate Fluidization. In Proceedings of the International
Symposium on Fluidization; Dringkenbrug, A. A., Ed.; Netherlands University Press: Eindhoven, The Netherlands, 1967; pp
52-65.
(41) Mueller, P.; Reh, L. Particle Drag and Pressure Drop in
Accelerated Gas-Solid Flow. In Preprint Volume for CFB-IV;
Avidan, A. A., Ed.; AIChE: New York, 1993; pp 193-198.
(42) Li, Y. Hydrodynamics. In Advances in Chemical Engineering; Kwauk, M., Ed.; Academic Press: San Diego, 1994; pp 85146.
(43) Bai, D. R.; Jin, Y.; Yu, Z. Acceleration of Particles and
Momentum Exchanges between Gas and Solids in Fast Fluidized
Beds. In Fluidization: Science and Technology; Kwauk, M., Kunni,
D., Eds.; Science Press: Beijing, 1991; pp 46-55.
(44) Makkawi, Y. T.; Wright, P. C. The Voidage Function and
Effective Drag Force for Fluidized Beds. Chem. Eng. Sci. 2003,
58, 2035-2051.

Ind. Eng. Chem. Res., Vol. 43, No. 18, 2004 5561
(45) Hu, H. H. Direct Simulation of Flows of Solid-Liquid
Mixtures. Int. J. Multiphase Flow 1996, 22, 335-352.
(46) Zhang, D. Z.; VanderHeyden, W. B. High-Resolution Threedimensional Numerical Simulation of a Circulating Fluidized Bed.
Powder Technol. 2001, 116, 133-141.
(47) Kandhai, D.; Derksen, J. J.; Van den Akker, H. E. A.
Interphase Drag Coefficients in Gas-Solid Flows. AIChE J. 2003,
49, 1060-1065.
(48) Li, J. H.; Kwauk, M. Particle-Fluid Two-Phase Flow: The
Energy-Minimization Multi-scale Method; Metallurgical Industry
Press: Beijing, 1994.
(49) Yang, N.; Wang, W.; Ge, W.; Li, J. H. Analysis of Flow
Structure and Calculation of Drag Coefficient for Concurrent-up
Gas-Solid Flow. Chin. J. Chem. Eng. 2003, 11, 79-84.
(50) Yang, N.; Wang, W.; Ge, W.; Li, J. H. CFD Simulation of
Concurrent-Up Gas-Solid Flow in Circulating Fluidized Beds with
Structure-Dependent Drag Coefficient. Chem. Eng. J. 2003, 96,
71-80.
(51) Xu, G.; Li, J. H. Analytical Solution of the EnergyMinimization Multi-scale Model for Gas-Solid Two-Phase Flow.
Chem. Eng. Sci. 1998, 53, 1349-1366.
(52) Li, J. H.; Cheng, C.; Zhang, Z.; Yuan, J.; Nemet, A.; Fett,
F. N. The EMMS ModelsIts Application, Development and
Updated Concepts. Chem. Eng. Sci. 1999, 54, 5409-5425.
(53) Ge, W.; Li, J. H. Physical Mapping of Fluidization Regimess
The EMMS Approach. Chem. Eng. Sci. 2002, 57, 3993-4004.
(54) Xu, G.; Kato, K. Hydrodynamic Equivalent Diameter for
Clusters in Heterogeneous Gas-Solid Flow. Chem. Eng. Sci. 1999,
54, 1837-1847.
(55) Matsen, J. M. Mechanisms of Choking and Entrainment.
Powder Technol. 1982, 32, 21-33.
(56) Bader, R.; Findlay, J.; Knowlton, T. M. Gas/Solids Flow
Patterns in a 30.5-cm-Diameter Circulating Fluidized Bed. In
Circulating Fluidized Bed Technology II; Basu, P., Large, J. F.,
Eds.; Pergamon Press: Oxford, U.K., 1988; pp 123-137.

(57) Miller, A.; Gidaspow, D. Dense, Vertical Gas-Solid Flow


in a Pipe. AIChE J. 1992, 38, 1801-1805.
(58) Ding, J.; Gidaspow, D. A Bubbling Fluidization Model
Using Kinetic Theory of Granular Flow. AIChE J. 1990, 36,
523-538.
(59) Spalding D B. Numerical Computation of Multi-Phase
Fluid Flow and Heat Transfer. In Recent Advances in Numerical
Methods in Fluids; Taylor, C., Morgan, K., Eds.; Pineridge:
Swansea, U.K., 1980; pp 139-167.
(60) Rhodes, M. J.; Sollaart, M.; Wang, X. S. Flow Structure in
a Fast Fluid Bed. Powder Technol. 1998, 99, 194-200.
(61) Davidson, J. F. Circulating Fluidized Bed Hydrodynamics.
Powder Technol. 2000, 113, 249-260.
(62) Tung, Y.; Li, J.; Kwauk, M. Radial Voidage Profiles in a
Fast Fluidized Bed. In Fluidization88: Science and Technology;
Kwauk, M., Kunii, D., Eds.; Science Press: Beijing, 1988; pp 139145.
(63) Zhang, W.; Tung, Y.; Johnsson, F. Radial Voidage Profiles
in Fast Fluidized Beds of Different Diameters. Chem. Eng. Sci.
1991, 46, 3045-3052.
(64) Xu, G.; Sun, G.; Gao, S. Estimating Radial Voidage Profiles
for All Fluidization Regimes in Circulating Fluidized Bed Risers.
Powder Technol. 2004, 139, 186-192.
(65) Xu, G.; Gao, S. Necessary Parameters for Specifying the
Hydrodynamics of Circulating Fluidized Bed RiserssA Review and
Reiteration. Powder Technol. 2003, 137, 63-76.

Received for review March 22, 2004


Revised manuscript received May 31, 2004
Accepted June 9, 2004
IE049773C

You might also like