You are on page 1of 17

Comparative Analysis of CFD Models of Dense

Gas–Solid Systems
B. G. M. van Wachem, J. C. Schouten, and C. M. van den Bleek
DelftChemTech, Chemical Reactor Engineering Section, Delft University of Technology,
2628 BL Delft, The Netherlands
R. Krishna
Dept. of Chemical Engineering, University of Amsterdam, 1018 WV Amsterdam, The Netherlands
J. L. Sinclair
School of Chemical Engineering, Purdue University, West Lafayette, IN 47907

Many gas ] solid CFD models ha®e been put forth by academic researchers, go®ern-
ment laboratories, and commercial ®endors. These models often differ in terms of both
the form of the go®erning equations and the closure relations, resulting in much confu-
sion in the literature. These ®arious forms in the literature and in commercial codes are
re®iewed and the resulting hydrodynamics through CFD simulations of fluidized beds
compared. Experimental data on fluidized beds of Hilligardt and Werther (1986), Ke-
hoe and Da®idson (1971), Darton et al.(1977), and Kuipers (1990) are used to quanti-
tati®ely assess the ®arious treatments. Predictions based on the commonly used go®ern-
ing equations of Ishii (1975) do not differ from those of Anderson and Jackson (1967)
in terms of macroscopic flow beha®ior, but differ on a local scale. Flow predictions are
not sensiti®e to the use of different solid stress models or radial distribution functions, as
different approaches are ®ery similar in dense flow regimes. The application of a differ-
ent drag model, howe®er, significantly impacts the flow of the solids phase. A simplified
algebraic granular energy-balance equation is proposed for determining the granular
temperature, instead of sol®ing the full granular energy balance. This simplification does
not lead to significantly different results, but it does reduce the computational effort of
the simulations by about 20%.

Introduction
Gas]solid systems are found in many operations in the of CFD calculations. Sinclair Ž1997. gives an extensive intro-
chemical, petroleum, pharmaceutical, agricultural, biochemi- duction on applying CFD models for gasrsolid risers. Al-
cal, food, electronic, and power-generation industries. Com- though single-phase flow CFD tools are now widely and suc-
putational fluid dynamics ŽCFD. is an emerging technique cessfully applied, multiphase CFD is still not because of the
for predicting the flow behavior of these systems, as it is nec- difficulty in describing the variety of interactions in these sys-
essary for scale-up, design, or optimization. For example, tems. For example, to date there is no agreement on the ap-
Barthod et al. Ž1999. have successfully improved the perform- propriate closure models. Furthermore, there is still no
ance of a fluidized bed in the petroleum industries by means agreement on even the governing equations. In addition, pro-
posed constitutive models for the solid-phase stresses and the
interphase momentum transfer are partially empirical.
Correspondence concerning this article should be addressed to B. G. M. van
Wachem. CFD models of gas]solid systems can be divided into two
Current address for B. G. M. van Wachem and J. C. Schouten: Laboratory of groups, Lagrangian models and Eulerian models. Lagrangian
Chemical Reactor Engineering, Eindhoven University of Technology, P. O. Box 513,
5600 MB Eindhoven, The Netherlands. models, or discrete particle models, calculate the path and

AIChE Journal May 2001 Vol. 47, No. 5 1035


motion of each particle. The interactions between the parti- over the total space is normalized to unity:
cles are described by either a potential force Žsoft-particle
dynamics, Tsuji et al., 1993. or by collision dynamics Žhard- `
particle dynamics, Hoomans et al., 1995.. The drawbacks of 4p H0 g Ž r . r 2 dr s1. Ž1.
the Lagrangian approach are the large memory requirements
and the long calculation time and, unless the continuous
The ‘‘radius’’ l of function g is defined by
phase is described using direct numerical simulations ŽDNS.,
empirical data and correlations are required to describe the
l `
gas]solid interactions. Eulerian models treat the particle H0 g Ž r . r 2
dr s Hl g Ž r . r 2 dr , Ž2.
phase as a continuum and average out motion on the scale of
individual particles, thus enabling computations by this
method to treat dense-phase flows and systems of realistic If l is chosen to satisfy a< l < L, where a is the particle
size. As a result, CFD modeling based on this Eulerian radius and L is the shortest macroscopic length scale, aver-
framework is still the only feasible approach for performing ages defined should not depend significantly on the particu-
parametric investigation and scale-up and design studies. lar functional form of g or its radius.
This article focuses on the Eulerian approach and com- The gas-phase volume-fraction e Ž x . g and the particle
pares the two sets of governing equations, the different clo- number density nŽ x . at point x are directly related to the
sure models, and their associated parameters that are em- weighting function g:
ployed in the literature to predict the flow behavior of
gas]solid systems. Unfortunately, many researchers propose
governing equations without citing, or with incorrectly citing, e Ž x. gs HV g Ž N x y yN . dV y Ž3.
g
a reference for the basis for their equations. Both Anderson
and Jackson Ž1967. and Ishii Ž1975. have derived multiphase nŽ x . s Ý g Ž N x y x p N . , Ž4.
flow equations from first principles, but the inherent assump- p
tions in these two sets of governing equations constrain the
types of multiphase flows to which they can be applied. One where Vg is the fluid-phase volume, and x p is the position of
of the objectives of our current contribution is to show how the center of particle p. The local mean value of the fluid-
these two treatments differ; it is shown that Ishii’s Ž1975. phase point properties, - f )g , is defined by
treatment is appropriate for a dispersed phase consisting of
fluid droplets, and that Anderson and Jackson’s Ž1967. treat-
ment is appropriate for a dispersed phase consisting of solid e Ž x . g - f )g Ž x . s HV f Ž y . g Ž N x y yN . dV . y Ž5.
particles. In the case of a solid dispersed phase, many re- g

searchers and commercial CFD codes employ kinetic theory


concepts to describe the solid-phase stresses resulting from The solid-phase averages are not defined like the fluid-
particle]particle interactions. Various forms of the constitu- phase averages, since the motion of the solid phase is deter-
tive models based on these concepts have been applied in the mined with respect to the center of the particle, and average
literature. The qualitative and quantitative differences be- properties need only depend on the properties of the particle
tween these are shown in this article. The predictions of CFD as a whole. Hence, the local mean value of the solid-phase
simulations of bubbling fluidized beds, slugging fluidized point properties is defined by
beds, and bubble injection into fluidized beds incorporating
these various treatments are compared to the ‘‘benchmark’’ n Ž x . - f )s Ž x . s Ý f s Ž N x y x p N . . Ž6.
experimental data of Hilligardt and Werther Ž1986., Kehoe p
and Davidson Ž1971., Darton et al. Ž1977., and Kuipers Ž1990..
The average space and time derivatives for the fluid and
solid phases follow from the preceding definitions. The aver-
Governing Equations
aging rules are then applied to the point continuity and mo-
Most authors who refer to the origin of their governing mentum balances for the fluid. For the solid phase, the aver-
equations refer to the work of Anderson and Jackson Ž1967. aging rules are applied to the equation of motion of a single
or Ishii Ž1975.. Anderson and Jackson Ž1967. and Jackson particle p:
Ž1997. wwith correction in Jackson Ž1998.x use a formal math-
ematical definition of local mean variables to translate the ­ ©s
point Navier-Stokes equations for the fluid and the Newton’s r sVp HS s Ž y . n Ž y . ds q Ý
s f q p q r sVp g , Ž7.
equation of motion for a single particle directly into contin- ­t p
g y
q/ p
uum equations representing momentum balances for the fluid
and solid phases, as earlier suggested by Jackson Ž1963.. The where ©s is the particle velocity, r s is the particle density, Vp
point variables are averaged over regions that are large with is the volume of particle p, sg is the gas-phase stress tensor,
respect to the particle diameter, but small with respect to the S p denotes the surface of particle p, and f q p represents the
characteristic dimension of the complete system. A weighting resultant force exerted on the particle p from contacts with
function, g Ž N x y yN ., is introduced in forming the local aver- other particles.
ages of system point variables, where N x y yN denotes the The resulting momentum balances for the fluid and solid
separation of two arbitrary points in space. The integral of g phases, dropping the averaging brackets - ) on the vari-

1036 May 2001 Vol. 47, No. 5 AIChE Journal


ables, are as follows: Table 1. Governing Equations Applied to Gas–Solid Flow
Continuity equations
­ ­e g
rg e g ©g q ©g ?=©g s= Ž e g sg . q=? Ž e g ©g . s 0
­t ­t
­e s
q=? Ž e s ©s . s 0
­t
yÝ HS sg ? n Ž y . g N x y yN ds y q r g e g g Ž8.
p p Momentum equations of Jackson (1997)
­ ©g b
­ rg q ©g ?=©g s=? t g y=P y Ž ©g y ©s .q r g g
­t eg
rg es ©s q ©s ?=©s s Ý g N x y x p N HS s n Ž y . ds
­t p
g y
­ ©s ­ ©g b
p
rs e s q ©s ?=©s y r g e s q ©g ?=©g s Ž ©g y ©s .
­t ­t eg
q=? ss q r s e s g . Ž9. q e s Ž r s y r g . g q=? ts y=Ps
in alternative form:
­ ©g
The first term on the righthand side of the gas-phase equa- rg e g q ©g ?=©g s e g=? t g y e g=P y b Ž ©g y ©s .q e g r g g
tion of motion represents the effect of stresses in the gas ­t
phase, the second term on the righthand side represents the ­ ©s
rs e s q ©s ?=©s s e s=? t g y e s=P q=? ts y=Ps q b Ž ©g y ©s .
traction exerted on the gas phase by the particle surfaces, ­t
and the third term represents the gravity force on the fluid. q e s rs g
The first term on the righthand side of the solid-phase equa- Momentum equations of Ishii (1975)
tion of motion represents the forces exerted on the particles ­ ©g
rg e g q ©g ?=©g sy e g=P q=? e g t g q e g r g g y b Ž ©g y ©s .
by the fluid, the second term on the righthand side repre- ­t
sents the force due to solid]solid contacts, which can be de- ­ ©s
rs e s q ©s ?=©s sy e s=P q=? e sts q e s r s g q b Ž ©g y ©s .
scribed using concepts from kinetic theory, and the third term ­t
represents the gravity force on the particles. The averaged applied to gas-solid flow ŽEnwald et al., 1996.:
shear tensor of the gas phase can be rewritten with the New- ­ ©g
rg e g q ©g ?=©g sy e g=P q=? e g t g q e g r g g y b Ž ©g y ©s .
tonian definition as ­t
­ ©s
rs e s q ©s ?=©s sy e s=P q=? ts y=Ps q e s r s g q b Ž ©g y ©s .
mg T ­t
sg sy Pg I q =©g q Ž =©g . , Ž 10.
eg Definitions
2
t i s 2 m i Di q l i y m i tr Ž Di . I
ž /
3
where the gas-phase volume-fraction is introduced in the vol- 1 T
ume process. Di s w =©i q Ž =©i . x
2
Note that the forces due to fluid traction are treated dif-
ferently in the fluid-phase and solid-phase momentum bal- Note: The explanation of the symbols can be found in the Notation.
ances. In the particle phase, only the resultant force acting
on the center of the particle is relevant; the distribution of
stress within each particle is not needed to determine its mo- difference in the manner in which the resultant forces due to
tion. Hence, in the solid-phase momentum balance, the re- fluid tractions act on the surfaces of the particles is a key
sultant forces due to fluid traction acting everywhere on the distinction between the Jackson Ž1997. and Ishii Ž1975. for-
surface of the particles are calculated first, after which these mulations. In the Ishii Ž1975. formulation, applicable to fluid
are averaged to the particle centers. In the fluid-phase mo- droplets, the fluid-droplet traction term is the same in the
mentum balance, the traction forces at all elements of gas phase and the dispersed-phase governing equations.
fluid]solid interaction are calculated, and then are averaged The integrals involving the traction on a particle surface
to the location of the surface elements. Hence, the fluid-phase have been derived by Nadim and Stone Ž1991. and are given
traction term is given as in Jackson Ž1997. as

b
Ý HS sg ? n Ž y . g N x y yN ds y s Ý g N x y x p NHS sg ? n Ž y . ds y Ý g N x y x p NHS sg ? n Ž y . ds y s e Ž ©g y ©s . q r g e s g
p p p p p p g

Df © g
q rg es Ž 12.
½
y=? a Ý g N x y x p N
p
HS
p
5
sg ? n Ž y . n Ž y . ds y q O Ž = 2 . , Dt

Ž 11. =? a Ý g N x y x p N HS sg ? n Ž y . n Ž y . ds y sy= Ž e s Pg . ,
p p

which is a result of a Taylor series expansion in g N x y yN Ž 13.


about the center of the particle with radius a. Here terms of
O Ž = 2 . and higher have been neglected. Note that the first where b is the interphase momentum transfer coefficient.
term on the righthand side of Eq. 11 is the same as the fluid The final equations of motion for both phases according to
traction term in the particle-phase momentum balance. The Jackson Ž1997. are shown in Table 1, both in the form as

AIChE Journal May 2001 Vol. 47, No. 5 1037


originally presented in his article, and in an equivalent alter- ligible, and
native form, which is merely a linear combination of the orig-
inal equations. Mk sdragq - Pk )=e k . Ž 18.
In Ishii’s Ž1975. formulation, the fluid and dispersed phases
are averaged over a fixed volume. This volume is relatively The momentum equations for the gas phase and the dis-
large compared to the size of individual molecules or parti- persed phase following the original work of Ishii Ž1975. are
cles. A phase indicator function is introduced, X Ž r ., which is shown in Table 1. Many researchers and commercial codes
unity when the point r is occupied by the dispersed phase, modify Ishii’s Ž1975. equations to describe gas]solid flows
and zero if it is not. Averaging over this function leads to the Žsuch as Enwald et al., 1996.. These modified equations are
volume fraction of both phases: also shown in Table 1. When Ishii’s Ž1975. equations are ap-
plied to gas]solid flows, the solid-phase stress tensor is not
1 multiplied by the solid volume fraction, since the volume-
es s HV X Ž r . dV ,
r Ž 14.
V fraction functionality is already accounted for in the kinetic
theory description.
where V is the averaging volume. Since both the continuous Comparing the Ishii Ž1975. and Jackson Ž1997. momentum
and dispersed phases are liquids, they are treated the same balances, the differences are twofold. First, Jackson Ž1997.
in the averaging process. Hence, the momentum balances for includes the solid volume fraction multiplied by the gradient
both phases are the same, of the total gas-phase stress tensor in the solid-phase mo-
mentum balance, whereas Ishii Ž1975. only includes the solid
­e k r k - ©k ) volume-fraction multiplied by the gradient of the pressure.
q=? Ž e k r k - ©k ) - ©k ) . Second, in the Ishii Ž1975. approach in the gas-phase mo-
­t mentum balance, the pressure carries the gas volume fraction
sy= Ž e k - Pk ) . q=? Ž e k -t k ) . q e k r k g q Mk , Ž 15. outside the gradient operator; the shear stress carries the gas
volume fraction inside the gradient operator. In Jackson
Ž1997. both stresses are treated equally with respect to the
where k is the phase number and Mk is the interphase mo-
gas volume fraction and the gradient operators. When the
mentum exchange between the phases, with M g q M s s 0. In
gas-phase shear stress plays an important role, these differ-
the Ishii Ž1975. formulation, the distribution of stress within
ences may be significant near large gradients of volume frac-
both phases is important since the dispersed phase is consid-
tion, that is, near bubbles or surfaces.
ered as fluid droplets. Hence, ‘‘jump’’ conditions are used to
determine Mk . The interphase momentum transfer is defined
as Closure Relations
Kinetic theory
1 Closure of the solid-phase momentum equation requires a
Mk sy Ý
Lj
Ž Pk n k y n k ? t k . description for the solid-phase stress. When the particle mo-
j
tion is dominated by collisional interactions, concepts from
1 gas kinetic theory ŽChapman and Cowling, 1970. can be used
sÝ Ž Ž - Pki )y Pk . n k y - Pki ) n k y n k to describe the effective stresses in the solid phase resulting
j Lj
from particle streaming Žkinetic contribution. and direct col-
? Ž -t ki )yt k . q n k ? -t ki ) . , Ž 16. lisions Žcollisional contribution.. Constitutive relations for the
solid-phase stress based on kinetic theory concepts have been
derived by Lun et al. Ž1984., allowing for the inelastic nature
where L j is the interfacial area per unit volume, Pk is the
of particle collisions.
pressure in the bulk of phase k, - Pki ) is the average pres-
Analogous to the thermodynamic temperature for gases,
sure of phase k at the interface, t k denotes the shear stress
the granular temperature can be introduced as a measure of
in the bulk, and -t ki ) represents the average shear stress
the particle velocity fluctuations.
at the interface. The terms Ž - Pki )y Pk . n k and n k ? Ž -t ki
)yt k . are identified by Ishii Ž1975. as the form drag and
1
the skin drag, respectively, making up the total drag force. Us - ©Xs2 ). Ž 19 .
The other terms can be written out as 3

Mk sdragq - Pk )=e k q Ž - Pki )y - Pk ) . =e k Since the solid-phase stress depends on the magnitude of
these particle-velocity fluctuations, a balance of the granular
y Ž =e k . ? -t ki ). Ž 17. energy Ž 32 Q . associated with these particle-velocity fluctua-
tions is required to supplement the continuity and momen-
According to Ishii and Mishima Ž1984., the last term on tum balance for both phases. This balance is given as
the righthand side is an interfacial shear term and is impor-
tant in a separated flow. According to Ishii Ž1975., the term 3 ­
Ž - Pki )y - Pk ) . only plays a role when the pressure at Ž e s r sQ . q=? Ž e s r sQ©s . s y=Ps I qts :=©s
ž /
2 ­t
the bulk is significantly different from that at the interface, as
in stratified flows. For many applications both terms are neg- y=? Ž k s=Q . ygs y Js , Ž 20.

1038 May 2001 Vol. 47, No. 5 AIChE Journal


where the first term on the righthand side represents the cre- mean free path to tend toward infinity, and the solids viscosi-
ation of fluctuating energy due to shear in the particle phase, ties tends toward a finite value as the solid volume fraction
the second term represents the diffusion of fluctuating en- tends to zero. Hence, by constraining the mean free path, the
ergy along gradients in Q, gs represents the dissipation due limit of the Hrenya and Sinclair Ž1997. shear viscosity expres-
to inelastic particle ]particle collisions, and Js represents the sion correctly tends to zero as the solid volume fraction ap-
dissipation or creation of granular energy resulting from the proaches zero. In dense solid systems Ž e s ) 0.05., there is no
working of the fluctuating force exerted by the gas through difference in the predicted solids viscosity of Lun et al. Ž1984.
the fluctuating velocity of the particles. Rather than solving and Hrenya and Sinclair Ž1997.. The Syamlal et al. Ž1993.
the complete granular energy balance given in Eq. 20, some solids shear viscosity also tends to zero as the solid volume
researchers ŽSyamlal et al., 1993; Boemer et al., 1995; Van fraction tends to zero. In this case, however, this solids shear
Wachem et al., 1998, 1999. assume the granular energy is in a
steady state and dissipated locally, and neglect convection and
Table 2. Solids Shear Viscosity
diffusion. Retaining only the generation and the dissipation
terms, Eq. 20 simplifies to an algebraic expression for the Lun et al. (1984)
granular temperature: 8
5't Q 1 8 es 1q h Ž 3h y2. e s g 0
5

0 s y=Ps I qts :=©s ygs .


ž / Ž 21.
ms s
96
rs d s
ž h g0
q
5 /ž 2yh /
768 2
he g x
q
Because the generation and dissipation terms dominate in 25p s 0
4 Q 1 r s d s g 0 Ž 1q e .Ž 3r2 ey1r2. e s2
dense-phase flows, it is anticipated that this simplification is
a reasonable one in dense regions of flow.
s e s2 r s d s g 0 Ž1q e .
5
(p
q 'Qp
15 Ž 3r2y1r2 e .
1 r s d s e s Ž 3r4 eq1r4 . 10 rs d s
q 'Qp q 'Qp
6 Ž 3r2y er2 . 96 Ž 1q e .Ž 3r2y1r2 e . g 0
Solid-phase stress tensor
Syamlal et al. (1993)
The solids pressure represents the normal solid-phase
4 e s d s r s'p Q
Q
forces due to particle]particle interactions. In the literature
there is general agreement on the form of the solids pres-
m s s e s2 r s d s g 0 Ž1q e .
5
2
p
q (6 Ž 3y e .
sure, given by Lun et al. Ž1984. as 1q Ž 1q e .Ž 3ey1. e s g 0
5
4 2 Q 1 Ž 1q e .Ž 3r2 ey1r2 .
Ps s r s e sQ w 1q2 Ž 1q e . g 0 e s x s e s r s d s g 0 Ž1q e .
5 p
(
q 'Qp r s d s g 0
15 Ž 3r2y er2 .
e s2

s r s e sQq2 g 0 r s e s2 Q Ž 1q e . . Ž 22. 1 e s d s r s'p Q


q
12 Ž 3r2y er2 .
The first part of the solids pressure represents the kinetic Gidaspow (1994)
contribution, and the second part represents the collisional 5'p
contribution. The kinetic part of the stress tensor physically 4 Q 2 r d 'Q 4
2
96 s s
represents the momentum transferred through the system by
particles moving across imaginary shear layers in the flow;
ms s
5
(
e s2 r s d s g 0 Ž1q e .
p
q
Ž 1q e . g 0
? 1q g 0 e s Ž 1q e .
5
4 Q 1
the collisional part of the stress tensor denotes the momen-
tum transferred by direct collisions.
s e s2 r s d s g 0 Ž1q e .
5
( p
q
15
1
'Qp rs d s g 0 Ž1q e. e s2
10 rs d s
The solids bulk viscosity describes the resistance of the q 'Qp rs d s e s q 'Qp
6 96 Ž 1q e . g 0
particle suspension against compression. In the literature,
there also is general agreement on the form of the solids bulk Hrenya and Sinclair (1997)
viscosity, given by Lun et al. Ž1984. as
5'p Q 1 1 8 es 1q8r5h Ž 3h y2. e s g 0

ls s
4
e s2r s d s g 0 Ž 1q e . ( Q
. Ž 23.
ms s
96
rs d s
ž
1q
lm f p h g 0
R
q
5 /ž 2yh /
3 p 768 2
q he g
25p s 0
4 Q 1 r s d s g 0 Ž 1q e . Ž 3r2 ey1r2 . e s2
However, the kinetic theory description for the solids shear
viscosity often differs between the various two-fluid models.
s e s2 r s g 0 Ž1q e .
5
(p
q 'Qp
15
lm f p
Ž 3r2y er2 .
Gidaspow Ž1994. does not account for the inelastic nature of
r s d s e s 1r2 1q
particles in the kinetic contribution of the total stress, as Lun
et al. Ž1984. do, claiming this correction is negligible. The
1
q 'Qp
ž ž R
q3r4 ey1r4

lm f p
/ /
6
solids shear viscosity of Syamlal et al. Ž1993. neglects the ki-
netic or streaming contribution, which dominates in dilute- 10
Ž 3r2y1r2 e . 1q
R ž rs d s
/
phase flow. Hrenya and Sinclair Ž1997. follow Lun et al. q 'Qp
96 lm f p
Ž1984., but constrain the mean free path of the particle by a
dimension characteristic of the actual physical system. This is
Ž 1q e .Ž 3r2y1r2 e . g 0 1q
R ž /
opposed to the Lun et al. Ž1984. theory, which allows the Note: The symbols can be found in the Notation.

AIChE Journal May 2001 Vol. 47, No. 5 1039


Table 3. Solids Thermal Conductivities
Lun et al. (1984)
25'p Q 8 96 e s 1q12r5h 2 Ž 4h y3. e s g 0
ks rs d s
128 ž h g0
q
5
512
/ž 41y33h /
q he s2 g 0
25p
2
Q r s d s g 0 Ž 1r2q er2 . Ž 2 ey1. e s2
9
s 2 e s2 r s d s g 0 Ž1q e .
p
( 8
q 'Qp
Ž 49r16y33r16 e .
15 e s r s d s Ž e 2r2q1r4 eq1r4 .
q 'Qp
16 Ž 49r16y33r16 e .
25 rs d s
q 'Qp
64 Ž 1q e .Ž 49r16y33r16 e . g 0
Syamlal et al. (1993)
15d s r s e s'Qp 12 16
ks 1q h 2 Ž 4h y3. e s g 0 q Ž 41y33h . he s g 0
4 Ž 41y33h . 5 15p
2
Q r s d s g 0 Ž 1r2q er2 . Ž 2 ey1. e s2
9
Figure 1. Comparison of solids shear viscosities from
different kinetic theory models: es 0.9, e max
s 2 e s2 r s d s g 0 Ž1q e . ( p
15
8
q 'Qp
Ž 49r16y33r16 e .
e s rs d s
s 0.65. q 'Qp
32 Ž 49r16y33r16 e .
Gidaspow (1994)
75
viscosity limit is reached because the kinetic contribution to k dil s r d 'p Q
384 2 s s
the solids viscosity is neglected. 2 6 Q
Table 2 presents the forms for the solids shear viscosity as
presented in the original articles as well as in an equivalent
ks
Ž 1q e . g 0
1q
5
Ž 1q e . g 0 e s k dil q2 e s2 r s d s g 0 Ž1q e .
2
p
(
Q 9 r s d s g 0 1r2q er2 2 ey1. e s2
form so that all of the models can be easily compared. Figure
1 shows a comparison of the constitutive models for the solids
s 2 e s2 r s d s g 0 Ž1q e . ( p
q
8
'Qp
lm f p
Ž . Ž
Ž 49r16y33r16 e .
shear viscosity as a function of the solid volume fraction. All e s r s d s e r2q1r4 eq1r4q
2

models yield the same solids shear viscosity at high solids vol- q
15
'Qp
ž R /
16 lm f p
ume fractions. Syamlal et al. Ž1993. deviate from the others
for solid volume fractions less than 0.3. Hrenya and Sinclair
25
Ž 49r16y33r16 e . 1q
ž R
rs d s
/
Ž1997. show a rapid decrease in solids shear viscosity at ex- 'Up
q
tremely small particle concentrations. 64 lm f p
Ž 1q e .Ž 49r16y33r16 e . 1q
ž R / g0

Conducti©ity of granular energy Note: The symbols can be found in the Notation.
Similar to the solids shear viscosity, the solids thermal con-
ductivity, k , consists of a kinetic contribution and a colli-
sional contribution. Gidaspow Ž1994. differs from Lun et al.
Ž1984. only in the dependency of the solids thermal conduc-
tivity on the coefficient of restitution. Syamlal et al. Ž1993.
neglect the kinetic contribution to the thermal conductivity.
Hrenya and Sinclair Ž1997. follow Lun et al. Ž1984., but con-
strain the mean free path of the particle by a dimension char-
acteristic of the actual system. Hence, the limit of their con-
ductivity expression, as with the shear viscosity, correctly tends
to zero when approaching zero solid volume fraction. Syamlal
et al. Ž1993. also correctly predict zero for the conductivity at
zero solid volume fraction by neglecting the kinetic contribu-
tion.
Table 3 presents the forms for the solids thermal conduc-
tivity as presented in the original articles, as well as in an
equivalent form so that all of the closure models can be eas-
ily compared. Figure 2 shows a quantitative comparison of
the constitutive models for the solids thermal conductivity as
a function of the solid volume fraction. All models yield the
same thermal conductivity at high solid volume fraction. Figure 2. Comparison of solids thermal conductivity
Syamlal et al. Ž1993. deviate from the others for solids vol- from different kinetic theory models: es 0.9,
ume fraction less than 0.3. Hrenya and Sinclair Ž1997. show a e max s 0.65.

1040 May 2001 Vol. 47, No. 5 AIChE Journal


rapid decrease in thermal conductivity at extremely small Table 4. Radial Distribution Function
particle concentration. Carnahan and Starling (1969)
1 3e s e s2
Dissipation and generation of granular energy g0 s q q
1y e s 2 Ž 1y e s . 2 2 Ž 1y e s . 3
Jenkins and Savage Ž1983. represent the dissipation of Lun and Sa®age (1986) y2 .5 e s,max
granular energy due to inelastic particle]particle collisions as es

4 Q
ž
g 0 s 1y
e s, max /
gs s 3 Ž 1y e 2 . e s2r s g 0 Q
ž (
ds p
y=? ©s .
/ Ž 24. Sinclair and Jackson (1989)
es 1r3
y1

For small mean-field gradients associated with a slight parti-


g 0 s 1y
ž e s, max /
cle inelasticity, the term =? ©s is typically omitted, as in Lun Gidaspow (1994) y1
et al. Ž1984.: 3 es 1r3

e s2r s g 0
g0 s
5
1y
ž e s, max /
gs s12 Ž 1y e 2 . Q 3r2 . Ž 25.
d s'p

The rate of energy dissipation per unit volume resulting from Radial distribution function
the action of the fluctuating force exerted by the gas through The solid-phase stress is dependent on the radial distribu-
the fluctuating velocity of the particles is given by Js s tion function at contact. Lun et al. Ž1984. employed the Car-
b Ž©Xs ? ©Xsy©Xg ? ©Xs.. According to Gidaspow Ž1994., the term ©Xs ? ©Xs nahan and Starling Ž1969. expression for the radial distribu-
is equal to 3U. The second term, ©Xg ? ©Xs, is neglected by Gi- tion function. The Carnahan and Starling Ž1969. expression,
daspow Ž1994.. However, Louge et al. Ž1991. have proposed a however, does not tend toward the correct limit at closest
closure for this second term based on the work of Koch Ž1990. solids packing. Because particles are in constant contact at
for the dilute flow regime, which we apply here: the maximum solid volume fraction, the radial distribution
function at contact tends to infinity. Therefore, alternative
b d s Ž © g y ©s .
2 expressions to the Carnahan and Starling Ž1969. expression
Js s b 3Qy . Ž 26 . have been proposed by Gidaspow Ž1994., Lun and Savage
4e s r s'p Q Ž1986., and Sinclair and Jackson Ž1989., which tend to the
correct limit at closest packing. These various forms of the
Using the closure of Louge et al. Ž1991. for ©Xg ? ©Xs, we have radial distribution function are given in Table 4 and are plot-
found that this term is of the same order of magnitude as ted in Figure 3 as a function of the solid volume fraction,
©Xs ? ©Xs. It should be noted, however, that the term as proposed along with the data from molecular simulations of Alder and
by Louge et al. Ž1991. is originally meant for the dilute flow Wainright Ž1960. and the data from experiments of Gidaspow
regime and does not tend to zero at closest solids packing. and Huilin Ž1998.. The expression of Gidaspow Ž1994. most
Therefore, Sundaresan Žprivate communication, 1999. has closely coincides with the data over the widest range of solid
proposed dividing this term by the radial distribution func- volume fractions. The expression of Gidaspow Ž1994., how-
tion to correct the closure in this limit of closest solids pack-
ing.
Recently, Sangani et al. Ž1996. have derived an equation
for ©Xs ? ©Xs, and Koch and Sangani Ž1999. have derived an equa-
tion for ©Xg ? ©Xs, especially for dense solid flows. With these
correlations, the expression for the rate of energy dissipation
resulting from fluctuations is

2
m s e sQ b 2 d s Ž © g y ©s .
Js s 3 R diss y SU , Ž 27 .
d s2 4e s r s'p Q

where R diss can be interpreted as the effective drag coeffi-


cient, which is determined as a result of a fit of numerical
simulations ŽSangani et al., 1996., and SU is an energy source:

1
SU s Rs b 2 , Ž 28.
2'p

where R s represents the energy source due to a specified


mean force acting on the particles and is obtained by a fit of Figure 3. Radial distribution functions: computational
numerical simulations. When the solids volume fraction ap- data of Alder and Wainright (1960) vs. experi-
proaches the maximum packing limit, R s tends to zero. mental data of Gidaspow and Huilin (1998).

AIChE Journal May 2001 Vol. 47, No. 5 1041


Johnson and Jackson Ž1987. propose a semiempirical equa-
tion for the normal frictional stress, Pf :

n
Ž e s y e s, min .
Pf s Fr p , Ž 32 .
Ž e s, max y e s .

where Fr, n, and p are empirical material constants, and


e s ) e s, min , e s, min are the solid-volume fraction when fric-
tional stresses become important; Fr, n, and p are material-
dependent constants. The frictional shear viscosity is then re-
lated to the frictional normal stress by the linear law pro-
posed by Coulomb Ž1776.

sx y s Pf sin f , Ž 33.

where f is the angle of internal friction of the particle. Rep-


Figure 4. Solids shear viscosity from different radial resentative values for the empirical constants employed in
distribution functions. Eqs. 32 and 33 are given in Table 5.
Solids shear viscosity follows Hrenya and Sinclair Ž1997.; e
s 0.9, R s 0.01525 m, and e max s 0.65.
Another approach, originally from Schaeffer Ž1987., was
employed by Syamlal et al. Ž1993. to describe the frictional
stress in very dense gas]solid systems:
ever, does not approach the correct limit of one as the solid n
volume-fraction approaches zero. Figure 4 presents the effect Pf s A Ž e s y e s, min . Ž 34.
of these different expressions for the radial distribution func-
tions on the solids shear viscosity. A difference of up to a Pf ? sin f
m fs .
factor of 2 in viscosity can result. 2 2 2
­ us ­ ®s ­ ®s ­ us ­ us ­ ®s

Frictional stress
es) 1
6 ž ­x
y
­y / ž / ž / ž
q
­y
q
­x
2
q
1
4 ­y
q
­x /
At high solid volume fraction, sustained contacts between Ž 35.
particles occur. The resulting frictional stresses must be ac-
counted for in the description of the solid-phase stress. Zhang
Values of As10 25, ns10, e s,min s 0.59, and f s 258 are typ-
and Rauenzahn Ž1997. conclude that particle collisions are
ically employed.
no longer instantaneous at very high solid volume fractions,
The approaches of Johnson and Jackson Ž1987. and Syam-
as is assumed in kinetic theory. Several approaches most of
lal et al. Ž1993. are compared in Figure 5. It can be seen that
which originated from geological research groups, have been
resulting normal frictional stress can differ by orders of mag-
presented in the literature to model the frictional stress. The
nitude.
models for frictional stress are very empirical and should be
used with caution. Typically, the frictional stress, sf , is writ-
ten in a Newtonian form: Interphase transfer coefficient
Generally, the form drag and skin drag are combined in
sf s Pf I q m f =©q Ž =© .
T
. Ž 29. one empirical parameter, the interphase drag constant b , in
the modeling of the momentum transfer between the two
phases. The drag coefficient b is typically obtained experi-
The frictional stress is added to the stress predicted by ki- mentally from pressure drop measurements in fixed, flu-
netic theory for e s ) e s, min : idized, or settling beds. Ergun Ž1952. performed measure-
ments in fixed liquid]solid beds at packed conditions to de-
Ps s P kinetic q Pf Ž 30. termine the pressure drop. Wen and Yu Ž1966. have per-
formed settling experiments of solid particles in a liquid over
m s s m kinetic q m f . Ž 31 . a wide range of solid volume fractions, and have correlated

Table 5. Empirical Parameters of Eqs. 32 and 33 by Various Researchers


Fr wNrm2 x n p esmi n f d s w m mx r s wkgrm3 x Material Reference
0.05 2 3 0.5 28 8 150 2500 Not specified Ocone et al. Ž1993.
3.65=10y3 2 0 40 } 25.08 1800 2980 Glass Johnson and Jackson Ž1987.
4.0=10y3 2 0 40 } 25.08 1000 1095 Polystyrene Johnson and Jackson Ž1987.
0.05 2 5 0.5 28.58 1000 2900 Glass Johnson et al. Ž1990.

1042 May 2001 Vol. 47, No. 5 AIChE Journal


Ž1952. for solid volume fractions larger than 0.2. The motiva-
tion for this hybrid drag description of Gidaspow Ž1994. is
unclear because the Wen and Yu Ž1966. expression includes
experimental drag data for solid volume fractions larger than
0.2. Moreover, a step change in the interphase drag constant
is obtained at the ‘‘crossover’’ solid volume fraction of 0.2,
which can possibly lead to difficulties in numerical conver-
gence. The magnitude of this discontinuity in b increases with
increasing particle Reynolds number. The drag coefficients
are summarized in Table 6 and are compared quantitatively
in Figure 6 for a range of solid volume fractions at a fixed
particle Reynolds number.

Simulations
The impact on the predicted flow patterns of the differ-
ences in the governing equations and constitutive models are
compared for the test cases of a freely bubbling fluidized-bed,
Figure 5. Different expressions for the frictional normal
a slugging fluidized bed, and a single bubble injection into a
stress.
fluidized bed. The particles in a fluidized bed move accord-
ing to the action of the fluid through the drag force, and
their data and those of others for solids concentrations, 0.01 bubbles and complex solid mixing patterns result. Typically,
F e s F 0.63. Syamlal et al. Ž1993. use the empirical correla- the average solid volume fraction in the bed is fairly large,
tions of Richardson and Zaki Ž1954. and Garside and Al-Bi- averaging about 40%, whereas in the the freeboard of the
bouni Ž1977. to determine the terminal velocity in fluidized fluidized bed Žthe top. there are almost no particles Ž e s f
and settling beds expressed as a function of the solid volume 10y6 ..
fraction and the particle Reynolds number. From the termi- The simulations in this work were carried out with the
nal velocity, the drag force can be readily computed. commercial CFD code CFX 4.2 from AEA Technology, Har-
The drag model of Gidaspow Ž1994. follows Wen and Yu well, UK, employing the Rhie-Chow ŽRhie and Chow, 1983.
Ž1966. for solid volume fractions lower than 0.2 and Ergun algorithm for discretization. For solving the difference equa-

Table 6. Drag Coefficients


Wen and Yu (1966)
3 Ž 1y e s . e s r g N ©g y ©s N
b s CD Ž1y e s .y2.65
4 ds
Rowe Ž1961.
24 0.687
1q0.15 ŽŽ 1y e s . Re p . if Ž 1y e s . Re p -1,000
Re p Ž 1y e s .
CD s
¼ 0.44 Re p s
d s r g N © g y ©s N
mg
if Ž 1y e s . Re p G1,000

Gidaspow (1994) applies the Ergun (1952) equation for higher ®olume fractions:
e s2m g 7 e s r g N © g y ©s N
150 q if e g ) 0.2
Ž 1y e s . d s2 4
bs
¼
3
C
4 D
Ž .
ds
ds
1y e s e s r g N ©g y ©s N
Ž 1y e s . y2 .65 if e s F 0.2

Syamlal et al. (1993)


3 e s Ž 1y e s . r g
b s CD N © g y ©s N
4 Vr 2 d s
Dalla Valle Ž1948.
2
Vr
ž
CD s 0.63q4.8 ( / Re
Garside and Al-Dibouni Ž1977.
1
Vr s
2
' 2
ay0.06 Req Ž 0.06 Re . q0.12 Re Ž 2 by a. q a2
as Ž1y e s . 4.14
1.28
0.8 Ž 1y e s . if e s G 0.15
bs
½ Ž 1y e s . 2.65 if e s - 0.15

AIChE Journal May 2001 Vol. 47, No. 5 1043


tions, the higher-order total variation diminishing ŽTVD.
scheme, Superbee is used. This TVD scheme incorporates a
modification to the higher-order upwind scheme Žsecond or-
der.. The time discretization is done with the second-order
backward-difference scheme. The solution of the pressure
from the momentum equations requires a pressure correction
equation, correcting the pressure and the velocities after each
iteration; for this, the SIMPLE ŽPatankar, 1980. algorithm is
employed. The calculated pressure is used to determine the
density of the fluid phase; the simulations are performed al-
lowing for compressibility of the gas phase. The grid spacing
was determined by refining the grid until average properties
changed by less than 4%. Due to the deterministic chaotic
nature of the system, the dynamic behavior always changes
with the grid. The simulations of the slugging fluidized bed
and the freely bubbling fluidized bed were carried out for 25
s of real time. After about 5 s of real time, the simulation has
reached a state in which averaged properties stay unchanged. Figure 6. Different expressions for the interphase drag
Averaged properties, such as bubble size and bed expansion, coefficient as a function of solid volume-frac-
were determined by averaging over the last 15 s of real time tion: Rep s 45.
in each simulation. A bubble is defined as a void in the solid
phase with a solid volume fraction less than 15%. The bubble
at this boundary is fixed to a reference value, 1.013=10 5 Pa.
diameter is defined as the diameter of a circle having the
Neumann boundary conditions are applied to the gas flow,
same surface as the void in the solid phase; this is called the
requiring a fully developed gas flow. For this, the freeboard
equivalent bubble diameter.
of the fluidized bed needs to be of sufficient height; this is
validated through the simulations. In the freeboard, the solid
Boundary conditions volume fraction is very close to zero, and this can lead to
All the simulations are carried out in a two-dimensional unrealistic values for the particle-velocity field and poor con-
rectangular space in which front and back wall effects are vergence. For this reason, a solid volume fraction of 10y6 is
neglected. The left and right walls of the fluidized bed are set at the top of the freeboard. This way the whole freeboard
treated as no-slip velocity boundary conditions for the fluid is filled with a very small number of particles, which gives
phase, and the free-slip velocity boundary conditions are em- more realistic results for the particle phase velocity in the
ployed for the particle phase. A possible boundary condition freeboard, but does not influence the behavior of the flu-
for the granular temperature follows Johnson and Jackson idized bed itself.
Ž1987.: The bottom of the fluidized bed is made impenetrable for
the solid phase by setting the solid phase axial velocity to
zero. For the freely bubbling fluidized bed and the slugging
n ? Ž k =Q . s
fluidized bed, Dirichlet boundary conditions are employed at
pr s e s'3Q 3Q the bottom with a uniform gas inlet velocity. To break the
w X N ©slip N 2 y Ž 1y ew2 . , Ž 36 . symmetry in the case of the bubbling and slugging beds, ini-
es 1r3
2 tially a small jet of gas is specified at the bottom lefthand
6 e s, max 1y
ž e s, max / side of the geometry. In the case of the bubble injection, a
Dirichlet boundary condition is employed at the bottom of
the fluidized bed. The gas inlet velocity is kept at the mini-
where the lefthand side represents the conduction of granu- mum fluidization velocity, except for a small orifice in the
lar energy to the wall, the first term on the righthand side center of the bed, at which a very large inlet velocity is speci-
represents the generation of granular energy due to particle fied. Finally, the solid-phase stress, as well as the granular
slip at the wall, and the second term on the righthand side temperature at the top of the fluidized bed, are set to zero.
represents dissipation of granular energy due to inelastic col-
lisions. Another possibility for the boundary condition for the Initial conditions
granular temperature is proposed by Jenkins Ž1992.:
Initially, the bottom part of the fluidized bed is filled with
particles at rest with a uniform solid volume fraction. The gas
n ? Ž k =Q . sy ©slip ? M y D, Ž 37 . flow in the bed is set to its minimum fluidization velocity. In
the freeboard a solid volume fraction of 10y6 is set, as ex-
where the exact formulations of M and D depend upon the plained earlier. The granular temperature is initially set to
amount of friction and sliding occurring at the wall region. 10y10 m2 ? sy2 .
Simulations we have done with an adiabatic boundary condi-
tion at the wall Ž =Qs 0. show very similar results. Test Case
The boundary condition at the top of the freeboard Žfluid- With increasing gas velocity above the minimum fluidiza-
phase outlet. is a so-called pressure boundary. The pressure tion velocity, Um f , bubbles are formed as a result of the in-

1044 May 2001 Vol. 47, No. 5 AIChE Journal


putational meshes are also shown in Figure 7. The test cases
are discussed in greater detail in the following sections.

Freely bubbling fluidized beds


In the freely fluidized-bed case, the gas flow is distributed
uniformly across the inlet of the bed. Small bubbles form at
the bottom of the fluidized bed that rise, coalesce, and erupt
as large bubbles at the fluidized-bed surface. In order to
evaluate model predictions, we use the Darton et al. Ž1977.
bubble model for bubble growth in freely bubbling fluidized
beds. This model is based upon preferred paths of bubbles
where the distance traveled by two neighboring bubbles be-
fore coalescence is proportional to their lateral separation.
Darton et al. Ž1977. have validated their model with mea-
surements of many researchers. Their proposed bubble-
growth equation for Geldart type B particles is

0.4 0.8
D b s 0.54 Ž UyUm f . Ž hq4'A .
0 gy0 .2 , Ž 38 .

where D b is the bubble diameter, h is the height of the bub-


ble above the inlet of the fluidized bed, U is the actual super-
ficial gas inlet velocity, and A 0 is the ‘‘catchment area’’ that
characterizes the distributor. For a porous-plate gas distribu-
'
tor, Darton et al. Ž1977. propose 4 A 0 s 0.03 m.
Werther and Molerus Ž1973. have developed a small capac-
Figure 7. Computational grid of simulated fluidized itance probe and the statistical theory to measure the bubble
beds with the gas inlet boundary condition. diameter and the bubble rise velocity in fluidized beds using
Ž a . Freely bubbling fluidized bed; Žb. slugging fluidized bed, this probe. This capacitance probe can be placed in the flu-
Žc . bubble injection into a fluidized bed.
idized bed at different heights and radial positions in the bed.
The bubble rise velocity is determined by placing two verti-
cally spaced probes and correlating the obtained data. The
herent instability of the gas]solid system. The behavior of capacitance probe measures the bubbles passing it, that is,
the bubbles significantly affects the flow phenomena in the the bubble is pierced by the capacitance probe. The duration
fluidized bed, that is, solid mixing, entrainment, and heat and of this piercing is dependent upon the size of the bubble, the
mass transfer. The test cases in this comparative study are rise velocity of the bubble, and the vertical position of the
used to investigate the effect of different closure models and bubble relative to the probe.
governing equations on the bubble behavior and bed expan- Hilligardt and Werther Ž1986. have done many measure-
sion. Simulation results of each test case are compared to ments of bubble size and bubble velocity under various con-
generally accepted experimental data and Žsemi.empirical ditions using the probe developed by Werther and Molerus
models. The system properties and computational parame- Ž1973. and have correlated their data in the form of the
ters for each of the test cases are given in Table 7; the com- Davidson and Harrison Ž1963. bubble model. Hilligardt and

Table 7. System Properties and Computational Parameters


Freely Bubbling Slugging Bubble Injection into
Parameter Description Fluidized Bed Fluidized Bed Fluidized Bed ŽKuipers, 1990.
r s wkgrm3 x Solid density 2,640 2,640 2,660
r g wkgrm3 x Gas density 1.28 1.28 1.28
m g wPa ? sx Gas viscosity 1.7=10y5 1.7=10y5 1.7=10y5
d s w m mx Particle diameter 480 480 500
e Coefficient of restitution 0.9 0.9 0.9
e ma x Max. solid volume fraction 0.65 0.65 0.65
Um f wmrsx Minimum fluidization velocity 0.21 0.21 0.25
D T wmx Inner column diameter 0.5 0.1 0.57
Ht wmx Column height 1.3 1.3 0.75
Hm f wmx Height at minimum fluidization 0.97 0.97 0.5
e s, m f Solids volume fraction 0.42 0.42 0.402
at minimum fluidization
D x wmx x-mesh spacing 7.14=10y3 6.67=10y3 7.50=10y3
D y w mx y-mesh spacing 7.56=10y3 7.43=10y3 1.25=10y2

AIChE Journal May 2001 Vol. 47, No. 5 1045


Werther propose a variant of the Davidson and Harrison their theoretical analysis, which led to the result that
Ž1963. model for predicting the bubble rise velocity as a func-
tion of the bubble diameter, Hmax y Hm f UyUm f
s , Ž 42.
Hm f u bub
u b s c Ž UyUm f . q wn gd b ,
' Ž 39 .
where u bub is the rise velocity of a slug without influence of
where w is the analytically determined square root of the the gas phase,
Froude number of a single rising bubble in an infinitely large
homogeneous area. Pyle and Harrison Ž1967. have deter- w
mined that w s 0.48 for a two-dimensional geometry, whereas
u bub s
2
'gD T Ž 43.
in three dimensions the Davies-Taylor relationship gives w s
0.71. The symbols c and n , added by Hilligardt and Werther or
Ž1986., are empirical coefficients based on their data, which
w
are dependent upon the type of particles and the width and u bub s '2 gD T , Ž 44.
height of the fluidized bed. For the particles and geometry 2
employed in this study, Hilligardt and Werther Ž1986. pro-
pose c f 0.3 and n f 0.8. Proposals of values for c and n corresponding to Eqs. 40 and 41. Hence, they also propose
under various fluidization conditions, determined by simula- upper and lower bounds on the maximum bed expansion.
tions, are given by Van Wachem et al. Ž1998..
Hilligardt and Werther Ž1986. also measured bed expan- Bubble injection in fluidized beds
sion under various conditions. Predictions of the bed expan-
Single jets entering a minimum fluidized bed through a
sion from the simulations are compared to these data.
narrow single orifice provide details of bubble formation and
growth. Such experiments were carried out by Kuipers Ž1990..
Slugging fluidized beds Kuipers Ž1990. reported the shape of the injected bubble as
In the case of the slugging fluidized beds, coalescing bub- well as the quantitative size and growth of the bubble with
bles eventually reach a diameter of 70% or more of the col- time using high-speed photography. The superficial gas-inlet
umn diameter, resulting from either a large inlet gas velocity velocity from the orifice was Us10 mrs, and the orifice was
or a narrow bed. The operating conditions employed in the ds1.5=10y2 m wide.
simulations correspond to the slugging conditions reported
by Kehoe and Davidson Ž1971., who present a detailed study Results and Discussion
of slug flow in fluidized beds. The experiments of Kehoe and
Predictions based on simulations of these three test cases
Davidson Ž1971. were performed in slugging fluidized beds of
are used to compare the different governing and closure
2.5-, 5-, and 10-cm diameter columns using Geldart B parti-
models. For this comparative study, only one particular clo-
cles from 50-m m to 300-m m diameter and with superficial gas
sure model is varied at a time, to determine the sensitivity of
inlet velocities of up to 0.5 mrs. X-Ray photography was used
the model predictions to that particular closure. No coupling
to determine the rise velocity of slugs and to determine the
effects were investigated. The default governing equations are
bed expansion. Kehoe and Davidson Ž1971. use their data to
those given by Jackson Ž1997., and the default closure models
validate two different equations for the slug rise velocity, both
are the solid-phase stress of Hrenya and Sinclair Ž1997., the
based on two-phase theory:
radial distribution function of Lun and Savage Ž1986., the
frictional model of Johnson and Jackson Ž1987. with empiri-
w
u slug sUyUm f q 'gD Ž 40. cal values given by Johnson et al. Ž1990., the complete
T
2 granular energy balance neglecting Js , and the drag coeffi-
w cient model of Wen and Yu Ž1966.. For animations of some
u slug sUyUm f q
2
'2 gD T , Ž 41. of the sim ulations, please refer to our W ebsite
http:rrwww.tcp.chem.tue.nlr;scrrwachemrcompare.html.

where w is the analytically determined square root of the


Froude number of a single rising bubble. Equation 40 is the Go©erning equations
exact two-phase theory solution, and Eq. 41 is a modification Simulations of the slugging bed case were performed with
of Eq. 40, based on the following observations: both the Ishii Ž1975. and the Jackson Ž1997. governing equa-
1. For fine particles Ž - 70 m m. the slugs travel symmetri- tions. Figure 8 shows the predicted maximum bed expansion
cally up in the fluidized bed, so the slug rise velocity is in- with increasing gas velocity during the slug flow and the two
creased by coalescence. correlations of Kehoe and Davidson Ž1971.. Figure 9 shows
2. For coarser particles Ž ) 70 m m. the slugs tend to move the increasing slug rise velocity with increasing gas velocity.
up the walls, which also increases their velocity. Clearly, the exact formulation of the governing equation does
According to Kehoe and Davidson Ž1971., Eqs. 40 and 41 not have any significant influence on the prediction of these
give upper and lower bounds on the slug rise velocity. Fur- macroscopic engineering quantities, and both CFD models
thermore, Kehoe and Davidson Ž1971. measured the maxi- do a good job at predicting these quantities. Microscopically,
mum bed expansion Ž Hmax . during slug flow. They validated however, there does seem to be a difference in the predic-

1046 May 2001 Vol. 47, No. 5 AIChE Journal


Figure 8. Predicted maximum expansion of a slugging
fluidized bed with increasing gas velocity with
the governing equations of Jackson (1997)
and Ishii (1975), and the additional term J s in
the granular energy equation.
The predictions are compared with the two-phase theory as
proposed and validated by Kehoe and Davidson Ž1971 ..

Figure 10. Rising bubble in a slugging fluidized bed


tions, as indicated in Figure 10. The flow of the gas phase in predicted by (a) employing the governing
areas of large solid volume-fraction gradient is slightly differ- equations of Jackson (1997), and by (b) em-
ent, leading to a different solids distribution. Specifically, ploying the governing equations of Ishii
Figure 10 shows that the Jackson Ž1997. governing equations (1975) at the same real time; the lines are
produce a more round-nosed bubble shape than the Ishii contours of equal solid volume fraction.
Ž1975. equations, because the path of the gas phase is differ-
ent.
Solids stress models
The exact solid-phase stress description does not influence
either the freely bubbling or the slugging fluidized-bed pre-
dictions, as is expected from Figure 1; this figure shows that
between 0.4 and 0.6 solids volume fraction, which is domi-
nant in the cases studied, all solids-phase stress predictions
are equal. Moreover, the influence of the radial distribution
upon the stress does not give rise to any variation in the pre-
dictions of the engineering quantities associated with these
simulations; the variation of the solids phase stress as a func-
tion of radial distribution function, shown in Figure 4, is small
between 0.4 and 0.6 solids volume fraction, as long as the
Carnahan and Starling Ž1969. equation is not employed. From
the magnitude of the terms on the solid-phase momentum
balance during simulations of fluidized beds, it can be con-
cluded that gravity and drag are the dominating terms and
that solids-phase stress predicted by kinetic theory plays a
minor role.

Figure 9. Predicted slug rise velocity with increasing Drag models


gas velocity with the governing equations of
Coordinating with results of the comparison of the drag
Jackson (1997) and Ishii (1975), and the addi-
models shown in Figure 6, the Syamlal et al. Ž1993. drag leads
tional term J s in the granular energy equation.
to a lower predicted pressure drop and lower predicted bed
The predictions are compared with the two-phase theory as
proposed and validated by Kehoe and Davidson Ž1971 .. The expansion than the other two drag models. Figure 11 shows
constant w s 0.48. the average simulated bed expansion employing different drag

AIChE Journal May 2001 Vol. 47, No. 5 1047


Figure 11. Predicted bed expansion as a function of gas Figure 13. Predicted bubble rise velocity as a function
velocity based on different drag models and of the bubble diameter at U s 0.54 m r s
with and without frictional stress. based on different drag models and com-
The predictions are compared to the experimental data of
pared to the experimental correlation of Hilli-
Hilligardt and Werther Ž1986.. The spread in the simula- gardt and Werther (1986).
tion data with the drag model of Gidaspow Ž1994. is indi- The vertical lines indicate the spread of the simulated
cated by the line. bubble rise velocity.

models in the freely bubbling fluidized-bed case, compared spread in the simulations is fairly large, all of the investigated
to measurements of Hilligardt and Werther Ž1986.. The drag drag models are in agreement with the equation put forth by
model of Syamlal et al. Ž1993. underpredicts the bed expan- Darton et al. Ž1977.. Figure 13 shows the predicted bubble
sion compared to the findings of Hilligardt and Werther rise velocity employing different drag models in a freely bub-
Ž1986., and therefore also underpredicts the gas holdup in bling fluidized bed, compared to the empirical correlation of
the fluidized bed. Hilligardt and Werther Ž1986.. All of the investigated drag
Figure 12 shows the simulated bubble size as a function of models are in fairly good agreement with the empirical corre-
the bed height when employing different drag models, com- lation.
pared with the Darton et al., Ž1977. equation. Although the Because the bubble sizes predicted by the different drag
models are all close, while the predicted bed expansion dif-
fers between the models, variations in the predicted solid-
volume fraction of the dense phase exist between the models,
with the Syamlal et al. Ž1993. drag model predicting the high-
est solid volume fraction in the dense phase.
Figure 14 shows the quantitative bubble-size prediction for
a single jet entering a minimum fluidized bed based on the
drag models of Wen and Yu Ž1966. and Syamlal et al. Ž1993.,
which are compared to the experimental data of Kuipers
Ž1990.. Moreover, in Figure 15 we show the resulting qualita-
tive predictions of the bubble growth and shape, and also
compare these with photographs by Kuipers Ž1990.. The Wen
and Yu Ž1966. drag model yields better agreement with
Kuipers’ Ž1990. findings for both the bubble shape and size
than the Syamlal et al. Ž1993. drag model. The Syamlal et al.
Ž1993. drag model underpredicts the bubble size and pro-
duces a bubble that is more circular in shape than in the
experiments of Kuipers Ž1990. and in the simulations with
the Wen and Yu Ž1966. drag model.
Figure 12. Predicted bubble size as a function of bed
height at U s 0.54 mr rs based on different Frictional stress
drag models and compared to the correla-
tion of Darton et al. (1977). Frictional stresses can increase the total solid-phase stress
The vertical lines indicate the spread of the simulated by orders of magnitude, and is an important contributing force
bubble size. in dense gas]solid modeling. The simulation of the single jet

1048 May 2001 Vol. 47, No. 5 AIChE Journal


achieved solids packing is higher Žmaximum achieved solids
volume fraction increased from 0.630 to 0.649., and the bed
expansion is less. Moreover, the solid-phase stress in the
dense regions are significantly decreased because the pre-
dicted granular temperature in the dense region of flow is
very low Ž Qf10y5 m2 ? sy2 . due to the magnitude of the dis-
sipation term. When frictional stress is neglected in the simu-
lations, convergence difficulty arises because the maximum
solid volume fraction specified in the radial distribution func-
tion is approached and the derivative of the radial distribu-
tion function near maximum solid volume fraction is ex-
tremely steep. In order to still obtain convergence, we have
written the radial distribution function as a Taylor series ap-
proximation at very high solid volume fraction. Adding fric-
tional stress in the simulations prevents this problem, be-
cause then the solid volume fraction does not approach the
Figure 14. Bubble diameter as a function of time for a maximum packing value.
bubble formed at a single jet of U s10 mr
rs.
A comparison is made between the experiments of Kuipers Granular energy balance
Ž1990 ., model simulations using the drag coefficient of Wen
and Yu Ž1966 . with and without frictional stress, and model The influence of the additional generation and dissipation
simulations using the interphase drag constant of Syamlal
et al. Ž1993 .. term Js in the granular energy balance is determined in the
case of the slugging fluidized bed. Figure 8 shows the predic-
tions of the maximum bed expansion as a function of increas-
ing gas velocity for simulations with and without this addi-
entering a fluidized bed reveals that the size of the bubble is
tional term. Figure 9 also shows the predicted rise velocity of
not significantly influenced by the frictional stress, as shown
the slugs with and without this additional term Js . Although
in Figure 14. However, Figure 11 shows that the predicted
this additional term Js results in as much as 20% higher
bed expansion in the freely bubbling fluidized bed is signifi-
granular temperature values Žgranular temperature increased
cantly less without frictional stress. Moreover, the number of
from 0.138 m2 ? sy2 to 0.165 m2 ? sy2 ., this does not seem to
iterations for obtaining a converged solution is almost dou-
influence the predicted bed expansion or the slug rise veloc-
bled when frictional stress is omitted. Without frictional
ity. The exact formulation of Js ŽEq. 26 or 27. does not play a
stress, there is less air in the dense phase, the maximum
role in the predicted granular temperature.
Simulations of slugging fluidized beds were also performed
using the simplified algebraic granular energy equation, Eq.
21. There were no differences in predicted bed expansion,
bubble size, or bubble rise velocity due to this simplification
vs. using the full granular energy balance. This simplified
equation gives rise to deviations from full granular energy-
balance predictions of as much as 10% in the granular tem-
perature Žgranular temperature decreased from 0.138 m2 ? sy2
to 0.0127 m2 ? sy2 .. The computational effort for solving the
complete granular energy equation is about 20% higher than
calculating the granular temperature from the algebraic
equation. More simulation results of the freely bubbling flu-
idized bed case with the algebraic equation are given in van
Wachem et al. Ž1998..

Conclusions
In this article we have compared different formulations that
are employed in CFD models for gas]solid flow in the Eule-
rianrEulerian framework. We discussed the basis for the for-
mulation of the two different sets of governing equations
common to the two-fluid literature with respect to the nature
Figure 15. Experimental and simulated bubble shape of the dispersed phase. It is shown in detail that the model-
associated with a single jet at U s10 mr
rs ing of gas]solid flows requires different governing equations
and at t s 0.10 s and t s 0.20 s. than the modeling of gas]liquid flows. We also have com-
Comparison is made between the Ža . experiment of Kuipers pared various closure models both quantitatively and qualita-
Ž1990 .; Žb . model simulation using the interphase drag
constant of Wen and Yu Ž1966 .; Žc . model simulation using tively. For example, we have shown how the hybrid drag model
the interphase drag constant of Syamlal et al. Ž1993 .. proposed by Gidaspow Ž1994. produces a discontinuity in the

AIChE Journal May 2001 Vol. 47, No. 5 1049


drag coefficient, how an order-of-magnitude difference in the psempirical constant in frictional stress
normal stress is predicted by the various frictional stress P spressure, N ? my2
r spoint in space, m
models, and how the Syamlal et al. Ž1993. model predicts a Rscharacteristic length scale, m
lower bed expansion than with the other drag models. ResReynolds number
Finally, we have studied the impact of the two governing Sssurface, m2
equations and the various closure models on simulation pre- t stime, s
Usinlet Žsuperficial . gas velocity, m ? sy1
dictions in three fluidized-bed test cases. It is shown that the
Um f sminimum fluidization velocity, m ? sy1
resulting predictions based on the two sets of governing ©svelocity vector, m ? sy1
equations are similar on an engineering scale, but are differ- V svolume, m3
ent in terms of microscopic features associated with individ- Vr sratio of terminal velocity of a group of particles to that of an
ual bubbles or localized solids distributions. It is also shown isolated particle
x sposition vector, m
that the model predictions are not sensitive to the use of dif- X sphase indicator
ferent solids stress models or radial distribution functions. In D x sx-mesh spacing, m
dense-phase gas]solid flow, the different approaches in the D y sy-mesh spacing, m
kinetic theory modeling predict similar values for the solid-
phase properties. From an analysis of the individual terms on Greek letters
the momentum balance of the solid-phase momentum bal-
b sinterphase drag constant, kg ? my3 ? sy1
ance during the simulations, it can be concluded that gravity e svolume fraction
and drag are the most dominating terms; this is why the two h s1r2 Ž1q e .
different sets of governing equations predict similar results, f sangle of internal friction
and why the exact solid-phase stress prediction is of minor w ssquare root of the Froude number
w X sspecularity coefficient
importance. At a very high volume fraction, frictional stress g sdissipation of granular energy, kg ? my3 ? sy1
can influence the hydrodynamic prediction due to its large k ssolids thermal conductivity, kg ? my1 ? sy1
magnitude. Simplifying the granular energy balance by re- l ssolids bulk viscosity, Pa ? s
taining only the generation and dissipation terms is a reason- l m f p smean free path, m
m ssolids shear viscosity, Pa ? s
able assumption in the case of fluidized-bed modeling and
n sempirical coefficient
reduces the computational effort by about 20%. Finally, the c sempirical coefficient
manner in which the drag force is modeled has a significant r sdensity, kg ? my3
impact on the simulation results, influencing the predicted s stotal stress tensor, N ? my2
bed expansion and the solids concentration in the dense- t sviscous stress tensor, N ? my2
Qsgranular temperature, m2 ? sy2
phase regions of the bed.

Acknowledgments Subscripts
The investigations were supported Žin part. by the Netherlands bsbubble
Foundation for Chemical Research ŽSON., with financial aid from bubssingle bubble
the Netherlands Organization for Scientific Research ŽNWO.. This dil sdilute
support is largely acknowledged. B.G.M. van Wachem gratefully ac- f sfrictional
knowledges the financial support of the Netherlands Organization g sgas phase
for Scientific Research ŽNWO., the Stimulation fund for Internation- isinterface
alization ŽSIR., DelftChemTech, the Delft University Fund, and the k seither phase
Reactor Research Foundation ŽRR., for the expenses for visiting mf sminimum fluidization
Purdue University. min sminimum; kick-in value
max smaximum
Notation psparticle
sssolids phase
Asempirical constant slip sslip
A 0 scatchment area of distributor, m2 slug sslug
CD sdrag coefficient w swall
d s sparticle diameter, m
Ds sstrain rate tensor, sy1
Dsdiameter, m
D T sinner column diameter, m Literature Cited
escoefficient of restitution Alder, B. J., and T. E. Wainwright, ‘‘Studies in Molecular Dynamics:
f sfluid-phase point property II. Behaviour of a Small Number of Elastic Spheres,’’ J. Chem.
Fr sempirical material constant, N ? my2 Phys., 33, 1439 Ž1960..
g r . sweighting function
Ž Anderson, T. B., and R. Jackson, ‘‘A Fluid Mechanical Description
g sgravitational constant, m ? sy2 of Fluidized Beds,’’ Ind. Eng. Chem. Fundam., 6, 527 Ž1967..
g 0 sradial distribution function Barthod, D., M. Del Pozo, and C. Mirgain, ‘‘CFD-Aided Design Im-
hsheight of bubble in fluidized bed, m proves FCC Performance,’’ Oil Gas J., 66 Ž1999..
Hm f sminimum fluidization bed height, m Boemer, A., H. Qi, U. Renz, S. Vasquez, and F. Boysan, ‘‘Eulerian
Ht scolumn height, m Computation of Fluidized Bed Hydrodynamics}A Comparison of
J sfluctuating velocityrforce correlation, kg ? my3 ? sy1 Physical Models,’’ Proc. of the Int. Conf. on FBC, Orlando, FL, p.
Lsinterfacial area per unit volume, my1 775 Ž1995..
M sinterphase momentum exchange, N ? sy1 Carnahan, N. F., and K. E. Starling, ‘‘Equations of State for Non-At-
nsempirical constant in frictional stress tracting Rigid Spheres,’’ J. Chem. Phys., 51, 635 Ž1969..
nsnumber density Chapman, S., and T. G. Cowling, The Mathematical Theory of Non-
nsnormal vector, m Uniform Gases, Cambridge Univ. Press, 3rd Ed., Cambridge Ž1970..

1050 May 2001 Vol. 47, No. 5 AIChE Journal


Coulomb, C. A., ‘‘Essai sur Une Application des Regles´ de Maximis Kuipers, J. A. M., A Two-Fluid Micro Balance Model of Fluidized Beds,
et Minimis ` `
a Quelques Problemes de Statique, Relatifs ´
a l’Archi- PhD Thesis, Univ. of Twente, Twente, The Netherlands Ž1990..
tecture,’’ Acad. R. Sci. Mem.` Math, Phys. Di®ers. Sa®ants, 7, 343 Louge, M. Y., E. Mastorakos, and J. T. Jenkins, ‘‘The Role of Parti-
Ž1776.. cle Collisions in Pneumatic Transport,’’ J. Fluid Mech., 231, 345
Dalla Valle, J. M., Micromeritics, Pitman, London Ž1948.. Ž1991..
Darton, R. C., R. D. LaNauze, J. F. Davidson, and D. Harrison, Lun, C. K. K., and S. B. Savage, ‘‘The Effects of an Impact Velocity
‘‘Bubble Growth Due to Coalescence in Fluidized Beds,’’ Trans. Dependent Coefficient of Restitution on Stresses Developed by
Inst. Chem. Eng., 55, 274 Ž1977.. Sheared Granular Materials,’’ Acta Mech., 63, 15 Ž1986..
Davidson, J. F., and D. Harrison, Fluidized Particles, Cambridge Univ. Lun, C. K. K., S. B. Savage, D. J. Jefferey, and N. Chepurniy, ‘‘Kinetic
Press, Cambridge Ž1963.. Theories for Granular Flow: Inelastic Particles in Couette Flow
Enwald, H., E. Peirano, and A. E. Almstedt, ‘‘Eulerian Two-Phase and Slightly Inelastic Particles in a General Flowfield,’’ J. Fluid
Flow Theory Applied to Fluidization,’’ Int. J. Multiphase Flow, 22, Mech., 140, 223 Ž1984..
21 Ž1996.. Nadim, A., and H. A. Stone, ‘‘The Motion of Small Particles and
Ergun, S., ‘‘Fluid Flow through Packed Columns,’’ Chem. Eng. Prog., Droplets in Quadratic Flows,’’ Stud. Appl. Mech., 85, 53 Ž1991..
48, 89 Ž1952.. Ocone, R., S. Sundaresan, and R. Jackson, ‘‘Gas-Particle Flow in a
Garside, J., and M. R. Al-Dibouni, ‘‘Velocity-Voidage Relationship Duct of Arbitrary Inclination with Particle-Particle Interactions,’’
for Fluidization and Sedimentation,’’ Ind. Eng. Chem. Proc. Des. AIChE J., 39, 1261 Ž1993..
De®., 16, 206 Ž1977.. Patankar, S. V., Numerical Heat Transfer and Fluid Flow, Hemi-
Gidaspow, D., Multiphase Flow and Fluidization, Academic Press, San sphere, New York Ž1980..
Diego Ž1994.. Pyle, D. L., and D. Harrison, Chem. Eng. Sci., 22, 531 Ž1967..
Gidaspow, D., and L. Huilin, ‘‘Equation of State and Radial Distri- Rhie, C. M., and W. L. Chow, ‘‘Numerical Study of the Turbulent
bution Functions of FCC Particles in a CFB,’’ AIChE J., 44, 279 Flow Past an Airfoil with Trailing Edge Separation,’’ AIAA J1, 21,
Ž1998.. 1527 Ž1983..
Hilligardt, K., and J. Werther, ‘‘Local Bubble Gas Hold-Up and Ex- Richardson, J. F., and W. N. Zaki, ‘‘Sedimentation and Fluidisation:
pansion of GasrSolid Fluidized Beds,’’ Ger. Chem. Eng., 9, 215 I,’’ Trans. Inst. Chem. Eng., 32, 35 Ž1954..
Ž1986.. Rowe, P. N. ‘‘Drag Forces in a Hydraulic Model of a Fluidized Bed:
Hoomans, B. P. B., J. A. M. Kuipers, W. J. Briels, and W. P. M. van II,’’ Trans. Inst. Chem. Eng., 39, 175 Ž1961..
Swaaij, ‘‘Discrete Particle Simulation of Bubble and Slug Forma- Sangani, A. S., G. Mo, H.-K. Tsao, and D. L. Koch, ‘‘Simple Shear
tion in a Two-Dimensional Gas-Fluidised Bed: A Hard-Sphere Flows of Dense Gas-Solid Suspensions at Finite Stokes Numbers,’’
Approach,’’ Chem. Eng. Sci., 51, 99 Ž1996.. J. Fluid Mech., 313, 309 Ž1996..
Hrenya, C. M., and J. L. Sinclair, ‘‘Effects of Particle-Phase Turbu- Schaeffer, D. G., ‘‘Instability in the Evolution Equations Describing
lence in Gas-Solid Flows,’’ AIChE J., 43, 853 Ž1997.. Incompressible Granular Flow,’’ J. Differ. Eqs., 66, 19 Ž1987..
Ishii, M., Thermo-Fluid Dynamic Theory of Two-Phase Flow, Direc- Sinclair, J. L., ‘‘Hydrodynamic Modelling,’’ Circulating Fluidized Beds,
tion des Etudes et Recherches d’Electricite ´ de France, Eyrolles, J. R. Grace, A. A. Evidan, and T. M. Knowlton, eds., Blackie,
Paris Ž1975.. London, p. 149 Ž1997..
Ishii, M., and K. Mishima, ‘‘Two-Fluid Model and Hydrodynamic Sinclair, J. L., and R. Jackson, ‘‘Gas-Particle Flow in a Vertical Pipe
Constitutive Relations,’’ Nucl. Eng. Des., 82, 107 Ž1984.. with Particle-Particle Interactions,’’ AIChE J., 35, 1473 Ž1989..
Jackson, R., ‘‘The Mechanics of Fluidized Beds: Part I: The Stability Syamlal, M., W. Rogers, and T. J. O’Brien, ‘‘Mfix Documentation
of the State of Uniform Fluidization,’’ Trans. Inst. Chem. Eng., 41, Theory Guide,’’ U.S. Dept. of Energy, Office of Fossil Energy,
13 Ž1963.. Tech. Note Ž1993..
Jackson, R., ‘‘Locally Averaged Equations of Motion for a Mixture Tsuji, Y., T. Kawaguchi, and T. Tanaka, ‘‘Discrete Particle Simula-
of Identical Spherical Particles and a Newtonian Fluid,’’ Chem. tion of Two-Dimensional Fluidized Bed,’’ Powder Technol., 77, 79
Eng. Sci., 52, 2457 Ž1997.. Ž1993..
Jackson, R., ‘‘Erratum,’’ Chem. Eng. Sci., 53, 1955 Ž1998.. Van Wachem, B. G. M., J. C. Schouten, R. Krishna, and C. M. van
Jenkins, J. T., ‘‘Boundary Conditions for Rapid Granular Flow: Flat, den Bleek, ‘‘Eulerian Simulations of Bubbling Behaviour in Gas-
Frictional Walls,’’ J. Appl. Mech., 59, 120 Ž1992.. Solid Fluidised Beds,’’ Comput. Chem. Eng., 22, s299 Ž1998..
Jenkins, J. T., and S. B. Savage, ‘‘A Theory for the Rapid Flow of Van Wachem, B. G. M., J. C. Schouten, R. Krishna, and C. M. van
Identical Smooth, Nearly Elastic, Spherical Particles,’’ J. Fluid den Bleek, ‘‘Validation of the Eulerian Simulated Dynamic Be-
Mech., 130, 187 Ž1983.. haviour of Gas-Solid Fluidised Beds,’’ Chem. Eng. Sci., 54, 2141
Johnson, P. C., and R. Jackson, ‘‘Frictional-Collisional Constitutive Ž1999..
Relations for Granular Materials, with Application to Plane Shear- Wen, C. Y., and Y. H. Yu, ‘‘Mechanics of Fluidization,’’ Chem. Eng.
ing,’’ J. Fluid Mech., 176, 67 Ž1987.. Prog. Symp. Ser., 62, 100 Ž1966..
Johnson, P. C., P. Nott, and R. Jackson, ‘‘Frictional-Collisional Werther, J., and O. Molerus, ‘‘The Local Structure of Gas Fluidized
Equations of Motion for Particulate Flows and Their Application Beds I. A Statistically Based Measured System,’’ Int. J. Multiphase
to Chutes,’’ J. Fluid Mech., 210, 501 Ž1990.. Flow, 1, 103 Ž1973..
Kehoe, P. W. K., and J. F. Davidson, ‘‘Continuously Slugging Flu- Zhang, D. Z., and R. M. Rauenzahn, ‘‘A Viscoelastic Model for
idised Beds,’’ Inst. Chem. Eng. Symp. Ser., 33, 97 Ž1971.. Dense Granular Flows,’’ J. Rheol., 41, 1275 Ž1997..
Koch, D. L., ‘‘Kinetic Theory for a Monodisperse Gas-Solid Suspen-
sion,’’ Phys. Fluids A, 2, 1711 Ž1990..
Koch, D. L., and A. S. Sangani, ‘‘Particle Pressure and Marginal
Stability Limits for a Homogeneous Monodisperse Gas-Fluidized
Bed: Kinetic Theory and Numerical Simulation,’’ J. Fluid Mech.,
400, 229 Ž1999.. Manuscript recei®ed Dec. 7, 1999, and re®ision recei®ed Oct. 9, 2000.

AIChE Journal May 2001 Vol. 47, No. 5 1051

You might also like