Sie sind auf Seite 1von 18

PARTICLE TECHNOLOGY AND FLUIDIZATION

Filtered Two-Fluid Models for Fluidized


Gas-Particle Suspensions
Yesim Igci, Arthur T. Andrews IV, and Sankaran Sundaresan
Dept. of Chemical Engineering, Princeton University, Princeton, NJ 08544

Sreekanth Pannala
Oak Ridge National Laboratory, Oak Ridge, TN 37831

Thomas OBrien
National Energy Technology Laboratory, Morgantown, WV 26507

DOI 10.1002/aic.11481
Published online March 28, 2008 in Wiley InterScience (www.interscience.wiley.com).

Starting from a kinetic theory based two-uid model for gas-particle ows, we rst
construct ltered two-uid model equations that average over small scale inhomogene-
ities that we do not wish to resolve in numerical simulations. We then outline a proce-
dure to extract constitutive models for these ltered two-uid models through highly
resolved simulations of the kinetic theory based model equations in periodic domains.
Two- and three-dimensional simulations show that the closure relations for the ltered
two-uid models manifest a denite and systematic dependence on the lter size. Lin-
ear stability analysis of the ltered two-uid model equations reveals that ltering
does indeed remove small scale structures that are afforded by the microscopic two-
uid model. 2008 American Institute of Chemical Engineers AIChE J, 54: 14311448, 2008
Keywords: circulating uidized beds, computational uid dynamics (CFD), uidization,
particle technology, uid mechanics

Introduction The number of particles present in most gas-particle ow


Chemical reactors that take the form of uidized beds and systems is large, rendering detailed description of the motion
of all the particles and uid elements impractical. Hence,
circulating uidized beds are widely used in energy-related
two-uid model equations24 are commonly employed to
and chemical process industries.1 Gas-particle ows in these
probe the ow characteristics, and species and energy trans-
devices are inherently unstable; they manifest uctuations
port. In this approach, the gas and particle phases are treated
over a wide range of length and time scales. Analysis of the
as interpenetrating continua, and locally averaged quantities
performance of large scale uidized bed processes through
such as the volume fractions, velocities, species concentra-
computational simulations of hydrodynamics and energy/spe-
tions, and temperatures of gas and particle phases appear as
cies transport is becoming increasingly common. In the pres-
dependent eld variables. The averaging process leading to
ent study, we are concerned with the development of hydro-
two-uid model equations erases the details of ow at the
dynamic models that are useful for simulation of gas-particle
level of individual particles; but their consequences appear in
ows in large scale uidized processes.
the averaged equations through terms for which one must de-
velop constitutive relations. For example, in the momentum
Correspondence concerning this article should be addressed to S. Sundaresan at balance equations, constitutive relations are needed for the
sundar@princeton.edu.
gas-particle interaction force and the effective stresses in the
2008 American Institute of Chemical Engineers gas and particle phases.

AIChE Journal June 2008 Vol. 54, No. 6 1431


Figure 1. Snapshots of the particle volume fraction eld in a large periodic domain of size 131.584 3 131.584
dimensionless units are displayed.
The physical conditions corresponding to these simulations are listed in Table 2. The domain-average particle volume fraction, h/si 5
0.05 Simulations were performed with different resolutions: (a) 64 3 64 grids; (b) 128 3 128 grids; (c) 256 3 256 grids; (d) 512 3 512
grids. The gray scale axis ranges from /s 5 0.00 (white) to /s 5 0.25 (black).

The general form of the two-uid model equations is fairly sought to capture these uctuations through numerical simu-
standard and this has permitted the development of numerical lation of microscopic two-uid model equations. Indeed,
algorithms for solving them. For example, open-source pack- two-uid models for such ows reveal unstable modes whose
ages such as MFIX4,5 and commercial software (e.g., Flu- length scale is as small as 10 particle diameters.30,31 This
ent1) can readily be applied to perform transient integration can readily be ascertained by simple simulations, as illus-
(of the discretized forms) of the balance equations governing trated in Figure 1. Transient simulations of a uidized sus-
reactive and non-reactive multiphase ows. The results gen- pension of ambient air and typical Fluid Catalytic Cracking
erated through such simulations are dependent on the postu- catalyst particles were performed (using MFIX4,5) in a Carte-
lated constitutive models, and a major focus of research over sian, two-dimensional (2D), periodic domain at different grid
the past few decades has been on the improvement of these resolutions; these simulations employed kinetic theory-based
constitutive models. (microscopic) two-uid model equations (summarized in
Through a combination of experiments and computer sim- Table 1 and briey discussed in the Microscopic Two-uid
ulations, constitutive relations have been developed in the lit- Model Equations section below). The relevant parameter val-
erature for the uidparticle interaction force and the effec- ues can be found in Table 2. The simulations revealed that
tive stresses in the uid and particle phases. In gas-particle an initially homogeneous suspension gave way to an inhomo-
systems, the interaction force is predominantly due to the geneous state with persistent uctuations. Snapshots of the
drag force. An empirical drag law that bridges the results of particle volume fraction elds obtained in simulations with
Wen and Yu6 for dilute systems and the Ergun7 approach for different spatial grid resolution are shown in Figure 1. It is
dense systems is widely used in simulation studies.2 In the readily apparent that ner and ner structures are resolved as
past decade, ab initio drag force models have also been the spatial grid is rened. Statistical quantities obtained by
developed via detailed simulations of uid ow around averaging over the whole domain were found to depend on
assemblies of particles.814 the grid resolution employed in the simulations and they
The Stokes number associated with the particles in many became nearly grid-size independent only when grid sizes of
gas-particle mixtures is sufciently large that particleparticle the order of 10 particle diameters were used (see Agrawal
and particlewall collisions do occur; furthermore, when the et al.30 for further discussion). Thus, if one sets out to solve
particle volume fraction is below 0.5, the particleparticle the microscopic two-uid model equations for gas-particle
interactions occur largely through binary collisions. The par- ows, grid sizes of the order of 10 particle diameters become
ticle phase stress in these systems is widely modeled through necessary. Moreover, such ne spatial resolution reduces the
the kinetic theory of granular materials.2,15,16 This kinetic time steps required, further increasing the computational
theory approach has also been extended to systems contain- effort. For most devices of practical (commercial) interest,
ing mixtures of different types of particles.2,1720 such extremely ne spatial grids and small time steps are
It is important to keep in mind that all these closures are unaffordable.32 Indeed, gas-particle ows in large uidized
derived from data or analysis of nearly homogeneous sys- beds and risers are often simulated by solving discretized
tems. Henceforth, we will refer to the two-uid model equa- versions of the two-uid model equations over a coarse spa-
tions coupled with constitutive relations deduced from nearly tial grid. Such coarse grid simulations do not resolve the
homogeneous systems as the microscopic two-uid model small-scale (i.e., subgrid scale) spatial structures which,
equations. For example, the kinetic theory based model equa- according to the microscopic two-uid equations and experi-
tions described and simulated in most of the literature refer- mental observation, do indeed exist. The effect of these unre-
ences fall in this category.2,1631 solved structures on the structures resolved in coarse-grid
A practical difculty comes about when one tries to solve simulations must be accounted for through appropriate modi-
these microscopic two-uid model equations for gas-particle cations to the closuresfor example, the effective drag
ows. Gas-particle ows in uidized beds and riser reactors coefcient in the coarse-grid simulations will be smaller than
are inherently unstable, and they manifest inhomogeneous that in the microscopic two-uid model to reect the tend-
structures over a wide range of length and time scales. There ency of the gas to ow more easily around the unresolved
is a substantial body of literature, where researchers have clusters30,31 than through a homogenous distribution of these

1432 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
Table 1. Model Equations for Gas-Particle Flows

@qs /s
r  qs /s v 0 (1)
@t
 
@ qg 1  /s  
r  qg 1  /s u 0 (2)
@t
 
@qg /s v
r  qg /s vv r  rs  /s r  rg f qs /s g (3)
@t
"   #
@ qg 1  /s u  
r  qg 1  /s uu 1  /s r  rg  f qg 1  /s g (4)
@t
 3   
@ 2 qs /s T 3
r qs /s Tv r  q  rs : rv Cslip  Jcoll  Jvis (5)
@t 2

Gas phase stress tensor


 
2
^g ru ruT  r  uI
rg pg I  l (6)
3

Gas-particle drag (Wen and Yu6)

3 qg 1  /s /s ju  vj
f bu  v; b CD 1  /s 2:65 (7)
4 d

8
>
< 24 1 0:15Reg 0:687 Reg < 1000 1  /s qg dju  vj
CD Reg ; Reg
>
: 0:44 Reg  1000 lg

Kinetic theory model for particle phase stress

rs ps  glb r  vI  2ls S (8)

1
1
where rv rvT  r  vI
ps qs /s 1 4g/s go T; S
2 3
    
2a l 8 8 3
ls 1 /s ggo 1 g3g  2/s go glb
3 go g2  g 5 5 5
p
l 5qs d pT
l ; l ;
1 q /2bl
2 g
96
s s 0T

256l/2s go 1 ep 1
lb ; g ; go ; /s;max 0:65; a 1:6
5p 2 1  /s =/s;max 1=3

Kinetic theory model for pseudo-thermal energy ux

q ks rT (9)

  
k 12 12 64
where ks 1 g/s go 1 g2 4g  3/s go 41  33gg2 /2s g2o
go 5 5 25p
p
k 75qs d pT
k ; k
1 6bk 48g41  33g
5qs /s 2 go T

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1433
Table 1. (Continued)
Kinetic theory model for rate of dissipation of pseudo-thermal energy through collisions

48 q /2
Jcoll p g1  g s s go T3=2 (10)
p d

Effect of uid on particle phase uctuation energy (Koch and Sangani16)

54/s lg T
Jvis Rdiss ; where
d2

3/1=2 135
Rdiss 1 ps / ln /s 11:26/s 1  5:1/s 16:57/2s  21:77/3s  /s go ln0:01 (11)
2 64 s

81/s l2g ju  vj
Cslip p W; where (12)
go d3 qg pT

R2d
W ;
1 3:5/s1=2 5:9/s
8
>
> 1 3/s =21=2 135=64/s ln/s 17:14/s
>
< ; /s < 0:4
1 0:681/s  8:48/2s 8:16/3s
Rd
>
> 10/s
>
: 0:7; /s  0:4
1  /s 3

particles. Qualitatively, this is equivalent to an effectively ered in highly resolved simulations of a set of microscopic
larger apparent size for the particles. two-uid model equations) over the whole (periodic) domain,
One can readily pursue this line of thought and examine were very different from those used in the microscopic two-
the inuence of these unresolved structures on the effective uid model and that they depended on size of the periodic
interphase transfer and dispersion coefcients which should domain. They also demonstrated that all the effects seen in
be used in coarse-grid simulations. Inhomogeneous distribu- the 2D simulations persisted when simulations were repeated
tion of particles will promote bypassing of the gas around in three dimensions (3D) and that both 2D and 3D simula-
the particle-rich regions and this will necessarily decrease the tions revealed the same qualitative trends. Andrews et al.31
effective interphase mass and energy transfer rates. Con- performed many highly resolved simulations of uidized gas-
versely, uctuations associated with the small scale inhomo- particle mixtures in a 2D periodic domain, whose total size
geneities will contribute to the dispersion of the particles and coincided with that of the grid size in an anticipated large-
the gas, but this effect will be unaccounted for in the coarse- scale riser ow simulation. Using these numerical results,
grid simulations of the microscopic two-uid models. they constructed ad hoc subgrid models for the effects of the
Researchers have approached this problem of treating ne-scale ow structures on the drag force and the stresses,
unresolved structures through various approximate schemes. and examined the consequence of these subgrid models on
OBrien and Syamlal,33 Boemer et al.34 and Heynderickx the outcome of the coarse-grid simulations of gas-particle
et al.35 pointed out the need to correct the drag coefcient to ow in a large-scale vertical riser. They demonstrated that
account for the consequence of clustering, and proposed a these subgrid scale corrections affect the predicted large-
correction for the very dilute limit. Some authors have used scale ow patterns profoundly.31
an apparent cluster size in an effective drag coefcient clo- Thus, it is clear that one must carefully examine whether a
sure as a tuning parameter,36 others have deduced corrections microscopic two-uid model must be modied to introduce
to the drag coefcient using an energy minimization multi-
scale approach.37 The concept of particle phase turbulence
has also been explored to introduce the effect of the uctua- Table 2. Physical Properties of Gas and Solids
tions associated with clusters and streamers on the particle d Particle diameter 7.5 3 1026 m
phase stresses.38,39 However, a systematic approach that qs Particle density 1500 kg/m3
combines the inuence of the unresolved structures on the qg Gas density 1.3 kg/m3
lg Gas viscosity 1.8 3 1025 kg/m s
drag coefcient and the stresses has not yet emerged. The ep Coefcient of restitution 0.9
effects of these unresolved structures on interphase transfer vt Terminal settling velocity 0.2184 m/s
and dispersion coefcients remain unexplored. v2t /g Characteristic length 0.00487 m
Agrawal et al.30 showed that the effective drag law and vt/g Characteristic time 0.0223 s
qsvt2 Characteristic stress 71.55 kg/m s2
the effective stresses, obtained by averaging (the results gath-

1434 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
the effects of unresolved structures before embarking on (a.k.a. granular energy) associated with the particles, which
coarse-grid simulations of gas-particle ows. In the study by is used to compute the local granular temperature; the parti-
Andrews et al.,31 the ltering was done simply by choosing cle phase stress is then expressed in terms of the local
the lter size to be the grid size in the coarse-grid simulation particle volume fraction, granular temperature, rate of defor-
of the ltered equations. Furthermore, the corrections mation, and particle properties. There are several different
accounting for the effects of the structures that would not be closures for the terms appearing in the granular energy equa-
resolved in the coarse-grid simulations were extracted from tion as well. Thus, it must be emphasized that while the gen-
highly resolved simulations performed in a periodic domain eral forms of the continuity, momentum, and granular energy
whose size was chosen to be the same as the lter size; this equations are common among most of the microscopic two-
imposed periodicity necessarily limited the dynamics of the uid models discussed in the literature, there are variations
structures in the highly resolved simulations and so the accu- in the closure relations. Thus, the exact form of the closures
racy of using the subgrid models deduced from such restric- for the microscopic two-uid model is still evolving. Never-
tive simulations is debatable. theless, the microscopic two-uid models are robust in the
The rst objective of the present study is to develop a sys- sense that when they are augmented with physically reasona-
tematic ltering approach and construct closure relationships ble closures, they do yield all the known instabilities in gas-
for the drag coefcient and the effective stresses in the gas particle ows, which in turn lead to persistent uctuations
and particle phases that are appropriate for coarse-grid simu- that take the form of bubble-like voids in dense uidized
lations of gas-particle ows. Briey, we have performed beds and clusters and streamers in dilute systems.3,40,41 Thus,
highly resolved simulations of a kinetic theory based two- all sets of constitutive relations which capture these small
uid model in a large periodic domain, and analyzed the scale instabilities can be expected to lead to similar conclu-
results using different lter sizes. In this case, as the lter sions regarding the structure of the closures for the ltered
size is considerably smaller than the periodic domain size, equations. With this in mind, we have selected one set of
the microstructures sampled in the ltered region are not closures for the kinetic theory-based microscopic equations
constrained by the periodic boundary conditions. The present (see Table 1). Further discussion of these equations and an
approach also exposes nicely the lter size dependence of extensive review of the relevant literature can be found in
various quantities. Agrawal et al.30 As the closures for the microscopic two-
It should be emphasized that the present study neither uid models improve, one can easily repeat the analysis
challenges the validity of the microscopic two-uid model described here and rene the ltered closures.
equations such as the kinetic theory based equations nor
assumes that they are exactly correct. Instead, it uses these
microscopic equations as a starting point and seeks modica- Filtered Two-Fluid Model Equations
tions to make them suitable for coarse-grid simulations. (If a The two-uid model equations are coarse-grained through
ne grid can be used to resolve all the structures contained a ltering operation that amounts to spatial averaging over
in the microscopic two-uid model equations, the present some chosen lter length scale. In these ltered (a.k.a.
analysis is unnecessary; however, such high resolution is nei- coarse-grained) equations, the consequences of the ow
ther practical nor desirable for the analysis of the macroscale structures occurring on a scale smaller than a chosen lter
ow behavior.) As more accurate microscopic two-uid mod- size appear through residual correlations for which one
els emerge, one can readily use such models to rene the must derive or postulate constitutive models. If constructed
results presented here. properly, and if the several assumptions innate to the lter-
The second objective of the present study is to demon- ing methodology hold true, the ltered equations should
strate that ltering does indeed remove small scale structures produce a solution with the same macroscopic features as
that are afforded by the microscopic two-uid models. If l- the nely resolved kinetic theory model solution; however,
tering has been done in a meaningful manner, the ltered obtaining this solution should come at less computational
equations should yield coarser structures than the micro- cost.
scopic two-uid model (from which the ltered equations Let /s(y,t) denote the particle volume fraction at location
were derived). We will demonstrate that this is indeed the y and time t obtained by solving the microscopic two-uid
case through one-dimensional linear stability analysis of the model. We can dene a ltered particle volume fraction
ltered model equations. /s x; t as
Z
Microscopic Two-Fluid Model Equations /s x; t Gx; y/s y; tdy
The general form of the two-uid model equations for V1

gas-particle ows is fairly standard. However, several choices where G(x,y) is a weight function that depends on x2y
have been discussed and analyzed in the literature for the and V1 denotes the region over whichR the gas-particle ow
constitutive relations for the uid-particle drag force and the occurs. The weight function satises V1 Gx; ydy 1. By
effective stresses.2,3,12 We consider a system consisting of choosing how rapidly G(x,y) decays with distance measured
uniformly sized particles and focus on the situation, where from x, one can change the lter size. We dene the uctua-
the particles interact only through binary collisions. In the ki- tion in particle volume fraction as
netic theory approach, the continuity and momentum equa-
tions for the gas and particle phases are supplemented by an
equation describing the evolution of the uctuation energy /0s y; t /s y; t  /s y; t:

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1435
Filtered phase velocities are dened according to coming from the particle phase velocity uctuations, see Eq.
Z 17. As noted by Agrawal et al.30 and Andrews et al.,31 the
/s x; tvx; t Gx; y/s y; tvy; tdy contribution due to the velocity uctuations is much larger
V1
than the microscopic particle phase stress even for modestly
large lter sizes and this will be seen clearly in the results
and presented below. Thus, when realistically large lter sizes (of
Z the order of 100 particle diameters or more) are employed,
1  /s x; t
ux;t Gx; y1  /s y;tuy;tdy one can neglect the rs contribution for all practical purposes
for the particle volume fraction range analyzed in this study.
V1
Therefore, at least as a rst approximation, it is not necessary
Here, u and v denote local gas and particle phase veloc- to include a ltered granular energy equation in the analy-
ities appearing in the microscopic two-uid model. We then sis.31 This, however, does not imply that the granular energy
dene the uctuating velocities as: equation (see Eq. 5 in Table 1) is not important in gas-parti-
cle ows. The granular energy equation and the parameters
v0 y;t vy;t  
vy;t and u0 y;t uy;t  u
y;t: (such as the coefcient of restitution) contained in it will
inuence the details of the small scale structures, which in
Applying such a lter to the continuity Eqs. 1 and 2 in turn will affect the velocity uctuation term in the ltered
Table 1, we obtain particle phase stress.
The ltered gas-particle interaction force F  includes a l-
@qs /s tered gas-particle drag force f and a term representing corre-
r  qs /s 
v 0 (13)
@t lated uctuations in particle volume fraction and the (micro-
scopic two-uid model) gas phase stress gradient, see Eq. 18.
and Before one can analyze the ltered two-uid model equa-
@qg 1  /s
tions, constitutive relations are needed for the residual corre-
r  qg 1  /s 
u 0; (14)  Ss, and r
lations F, g in terms of ltered particle volume frac-
@t tion, velocities, and pressure. Furthermore, as these are l-
tered quantities, the constitutive relations capturing them will
as the ltered continuity equations, where it has been
necessarily depend on the details of the uctuations being
assumed that the gas density does not vary appreciably over
averaged, but these details will depend on the location in the
the representative region of the lter. These are identical in
process vessel. For example, one can anticipate that uctua-
form to the microscopic continuity equations in Table 1.
tions in the vicinity of solid boundaries will be different
Repeating this analysis with the two microscopic momentum
from those away from such boundaries. Accordingly, it is
balance Eqs. 3 and 4 in Table 1, we obtain the following l-
entirely reasonable to expect that the constitutive models for
tered momentum balances:
these residual correlations should include some dependence
  on distance from boundaries. (This is well known in single
@qs /s v
r  qs /s 
v  r
v r  Rs  /r g phase turbulent ows.) In the present study, we do not
@t address the boundary effect, but focus on constitutive models
F  qs /s g 15 that are applicable in regions away from boundaries as it is
an easier rst problem to address. It is assumed that the con-
" #
@qg 1  /s 
u stitutive relations for the residual correlations will depend on
r  qg 1  /s 
uu 1  /s r  r
g local ltered variables and their gradients.
@t
In rapid gas-particle ows with qs/s  qg (1 2 /s), it is
 qg 1  /s g
 r  qg 1  /s u0 u0  F 16 invariably the case that qg 1  /s u0 u0  qs /s v0 v0 , and we
simplify the ltered gas phase stress as:
Here
X g pg I
r (19)
rs qs /s vv  qs /s vv r
s qs /s v0 v0 (17)  as
We express F
s

 f  /0 r  r0 :  f  /0 r  r0 be 
F u
v (20)
F s g (18) s g

The ltered momentum balance equations are nearly iden- where be is a ltered drag coefcient to be found. The
tical (in form) to the microscopic momentum balances in Ta- /0s r  r0g term in Eq. 20 can also add a dynamic part, resem-
ble 1. One exception is that the ltered gas phase momentum bling an apparent added mass force4244; however, as Andrews45
balance now contains an additional term, found such a dynamic part to be much smaller than the drag
r  qg 1  /s u0 u0 . In the class of problems we consider force term in Eq. 20, we will limit ourselves to Eq. 20.
here the contribution of this term is much weaker than the We begin our analysis by postulating the following ltered
rst and third terms on the right hand side of the gas phase particle phase stress model:
momentum balance equation (as qs/s  qg (1 2 /S) in
most of the ow domain in the problem considered here). X 2
The effective particle phase stress, Ss, includes the ltered pse Ilbe r
vIlse r vT  rvI
v r (21)
s
3
microscopic stress rs and a Reynolds stress-like contribution

1436 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
where pse ps 13 qs /s v0x v0x qs /s v0y v0y qs /s v0z v0z is the l-
tered particle phase pressure;lse and lbe are the ltered parti-
cle phase shear and bulk viscosity, respectively. As the simu-
lations described below do not permit an evaluation of lbe ,
we do not consider this term in the present analysis and we
simplify Eq. 21 as
X 2
pse Ilse r vT  r
v r vI: (22)
s
3

The ltered particle phase shear viscosity is dened as


@vy
lse ls qs /s v0x v0y = @v
@y @x .
x

We now seek closure relations for be , pse , and lse by l-


tering computational data gathered from highly resolved sim-
ulations of the microscopic two-uid model equations.

Detailed Solution of Microscopic Two-Fluid


Model Equations
As already noted in the Filtered Two-uid Model Equa-
tions section, we restrict our attention to closures for be , pse
and lse in ow regions far away from solid boundaries. A Figure 2. Snapshot of the particle volume fraction eld
simple and effective manner by which solid boundaries can in a large periodic domain of size 131.584 3
be avoided is to consider ows in periodic domains. The l- 131.584 dimensionless units are displayed.
tering operation does not require a periodic domain; how- Simulations were performed with 512 3 512 grid points.
Overlaid is a pictorial representation of region averaging,
ever, as each location in a periodic domain is statistically where regions of varying size are isolated and treated as indi-
equivalent to any other location, statistical averages can be vidual realizations. Regions (lters) having dimensionless
gathered much faster when simulations are done in periodic lengths of 4.112, 8.224, 16.448, and 32.896 are shown as
shaded subsections.
domains. With this in mind, all the analyses described here
have been performed in periodic domains. Agrawal et al.30
have already shown that the results obtained from 2D and
3D periodic domains are qualitatively similar, but differ
somewhat quantitatively; therefore, we have focused rst on tered) values. (Such region averaging is equivalent to setting
2D simulations in the present study to bring forward the lter the weight function to an appropriate non-zero constant
size dependence of the closures for the residual correlations, everywhere inside the region and to zero outside.) Note that
as 2D simulations are computationally less expensive. We one can choose a large number of different regions of the
will present several 3D simulation results at the end to bring same size inside the overall domain and thus many region-
forth the differences between 2D and 3D closures. averaged values can be extracted from each instantaneous
snapshot. When the system is in a statistical steady state, one
can construct tens of thousands of such averages by repeating
Two-Dimensional Simulations
the analysis at various time instants.
We have performed many sets of highly resolved simula- Returning to Figure 2, note that the averages over different
tions (of the set of microscopic two-uid model equations regions at any given time are not equivalent; for example, at
for a uidized suspension of particles presented in Table 1) the given instant, different regions (even of the same size)
in large 2D periodic domains using the open-source software will correspond to different region-averaged particle volume
MFIX.5 These simulations are identical to those described by fractions, particle and uid velocities. Thus, one cannot sim-
Agrawal et al.,30 except that our simulations are now done ply lump the results obtained over all the regions; instead,
for much larger domain sizes. Agrawal et al.30 averaged the they must be grouped into bins based on various markers and
residual correlations over the entire domain (i.e., the lter statistical averages must be performed within each bin to
size is the same as their domain size), but as our simulation extract useful information. Our 2D simulations revealed that
domains are much larger, the computationally generated the single most important marker for a region is its average
data can now be averaged using a range of lter sizes that particle volume fraction. Therefore, we divided the permissi-
are smaller than the domain size. ble range of ltered particle volume fraction 0
/s <
After an initial transient period that depends on the initial /s;max 0:65 into 1300 bins (so that each bin represented a
conditions, persistent, time-dependent and spatially inhomo- volume fraction window of 0.0005) and classied the ltered
geneous structures develop. Figure 2 shows an instantaneous data in these bins. (Strictly speaking, one would expect to
snapshot of the particle volume fraction eld in one such 2D use two-dimensional bins, involving /s and a Reynolds num-
simulation. One can then select any region of desired size ber based on ltered slip velocity, to classify the ltered drag
(illustrated in the gure as gray squares of different sizes) coefcient; however, the Reynolds number dependence was
and average any quantity of interest over all the cells inside found to be rather weak for the cases investigated in this
that region; we refer to such results as region-average (or l- study.) For each snapshot of the ow eld in the statistical

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1437
FCC particles. Thus, one can readily appreciate that this lter
size is quite small compared to the macroscopic dimensions
of typical process vessels. The various symbols in this gure
refer to computational data obtained by solving the micro-
scopic two-uid model equations at different resolution lev-
els. Simulations were performed using different domain-aver-
age particle volume fractions so that every (ltered) volume
fraction shown here would have many realizations. This g-
ure indicates that at a sufciently high resolution the results
did become nearly independent of the grid size used in the
simulations to generate the computational data. Typically,
when the grid size was smaller than the lter size by a factor
of 16 or more (so that there were at least 256 grids inside
the ltering region in 2D simulations), the ltered
results were found to be essentially independent of the grid
Figure 3. The variation of the dimensionless ltered resolution.
drag coefcient with particle volume frac- The effect of (periodic) domain size on the ltered drag
tion, determined by ltering the computa- coefcient was explored by performing simulations with two
tional data gathered from simulations in a different domain sizes. Figure 4 presents the dimensionless
large periodic domain of size 131.584 3 ltered drag coefcient for two different lter sizes and two
131.584 dimensionless units, is presented. different domain sizes. It is clear that for both lter sizes, the
The dimensionless lter length 5 8.224. The ltered drag results are essentially independent of domain size. In general,
coefcient includes contributions from the drag force and the ltered results were found to be independent of the do-
the pressure uctuation force. Data used for ltering were main size as long as the lter size was smaller than one-
generated by running simulations for domain-average parti-
cle volume fractions of 0.05, 0.15, 0.25, and 0.35. The g- fourth of the domain size. (The lter size dependence seen in
ure shows results obtained by ltering data generated at this gure is discussed below.)
different grid resolutions as marked in the legend. The top The results presented in Figures 3 and 4, and in many of
curve corresponds to result obtained with 256 3 256 grids.
the gures below, were generated by combining results
obtained from simulations with many different specied do-
steady state, we considered a ltering region around each main-average particle volume fractions. Figure 5 shows the
grid point in the domain and determined the ltered particle variation of the ltered drag coefcient with ltered particle
volume fraction /s , ltered slip velocity  u v, ltered volume fraction, with results obtained from simulations with
uid-particle interaction force, and so forth. This combina- different domain-average particle volume fractions indicated
tion of ltered quantities represents one realization and it with different symbols. Although the domain-average particle
was placed in the appropriate ltered particle volume fraction volume fraction affects the ltered drag coefcient slightly
bin, determined by its volume fraction value. In this manner,
a large number of realizations were generated from each
snapshot. This procedure was then repeated for many snap-
shots. The many realizations within each bin were then aver-
aged to determine ensemble-averaged values for each ltered
quantity. From such bin statistics, the ltered drag coef-
cient, the ltered particle phase normal stresses and ltered
particle phase viscosity were calculated as functions of l-
tered particle volume fraction. For example, the ltered drag
coefcient is taken to be the ratio of the ltered drag force
and the ltered slip velocity, each of which has been deter-
mined in terms of the volume fraction. All the results are
presented as dimensionless variables, with qs,  v, and g repre-
senting characteristic density, velocity, and acceleration.
Figure 3 shows the variation of the dimensionless ltered
drag coefcient, be;d be vt =qs g as a function of /s for
one particular lter size. Even though all the results are pre-
sented in terms of dimensionless units, it is instructive to Figure 4. The effect of domain size on the dimension-
consider some dimensional quantities to help visualize the less ltered drag coefcient is displayed.
physical problem better. Most of the 2D ltered results pre- Data used for ltering were generated by running simula-
sented in this manuscript are based on computational data tions at domain-average particle volume fractions of 0.02,
gathered in a 131.584 3 131.584 (dimensionless units) 0.05, 0.10, 0.15, 0.20, 0.25, and 0.35 for two different
square periodic domains of sizes: 131.584 3 131.584
square periodic domain; this domain size translates to 0.64 m dimensionless units (512 3 512 grids) and 32.896 3
3 0.64 m for the FCC particles (whose physical properties 32.896 dimensionless units (128 3 128 grids). The top two
curves correspond to a dimensionless lter length of 2.056,
are given in Table 2). The dimensionless lter size of 8.224 while the bottom two are for a dimensionless lter length
used in Figure 3 corresponds to a lter size of 0.04 m for the of 4.112.

1438 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
of the lter length only. It can be given in a simple algebraic
form as

32Frf2 63:02Frf1 129


be b
Frf3 133:6Frf2 66:61Frf1 129

Intuitively, one can expect that the clusters will not grow
beyond some critical size and that at sufciently large lter
sizes the ltered drag coefcient will become essentially in-
dependent of the lter size. It is clear from Figure 6b that
this critical lter size is denitely larger than the largest lter
size shown there. Simulations using much larger domains are
needed to identify this critical size, but we have not pursued
this issue in the present study; instead, we have focused on a
qualitative understanding as to how the ltered quantities
Figure 5. The effect of domain-average particle volume depend on lter size for modest lter sizes.
fraction on the dimensionless ltered drag
coefcient is displayed.
Simulations were performed in a square domain of size
131.584 3 131.584 dimensionless units and 512 3 512 grid
points and domain-average particle volume fractions of
0.05, 0.10, 0.15, 0.20, 0.25, and 0.35 (shown by different
symbols in each curve). Results are presented for dimen-
sionless lter lengths of 2.056 (top curve), 4.112 (middle
curve) and 8.224 (bottom curve).

(particularly for volume fractions above 0.20), this effect is


clearly much smaller than that of the lter size. Physically,
this implies that the lter size dependence manifested in this
and other gures largely stems from the inhomogeneous
microstructure inside the ltering region and the ltered drag
coefcients are either independent of or only weakly depend-
ent on the conditions prevailing outside the ltered region (at
least over the range of particle volume fractions over which
the data were collected).
Figure 6a shows the variation of the (dimensionless) ltered
drag coefcient, as a function of ltered particle volume frac-
tion for various lter sizes. The results for the four smallest l-
ter sizes are likely to decrease somewhat if simulations could
be done at higher resolutions, but as noted earlier in the con-
text of Figure 3 the results for all larger lter sizes are essen-
tially independent of grid size. It is clear that the ltered drag
coefcient decreases substantially with increasing lter size,
and this can readily be rationalized. As the lter size is
increased, the averaging is being performed over larger and
larger clusterslarger clusters allow greater bypassing of the
gas resulting in lower apparent drag coefcient. The uppermost
curve in Figure 6a is the intrinsic drag law; the lter size here
is simply the grid size used in the simulations of the micro- Figure 6. (a) The variation of the dimensionless ltered
scopic two-uid model equations (which is equivalent to no l- drag coefcient with particle volume fraction
tering at all). For typical FCC particles (see Table 2), a dimen- for various lter sizes (listed in the legend in
sionless lter size of 2.056 is equivalent to 0.01 m, and so dimensionless units) is shown.
even at small lter sizes (from an engineering viewpoint) an Simulations were performed in a square periodic domain of
size 131.584 3 131.584 dimensionless units and using 512
appreciable reduction occurs in the effective drag coefcient. 3 512 grid points. Data used for ltering were generated
Figure 6b shows how the ratio of the ltered drag coef- by running simulations for domain-average particle volume
cient to the microscopic drag coefcient changes with parti- fractions of 0.01, 0.02, 0.03, 0.04, 0.05, 0.10, 0.15, 0.20,
0.25, 0.30, and 0.35. The dimensionless lter lengths are
cle volume fraction for several lter lengths. It is clear from shown in the legend. (b) An alternative representation of
this gure is that this ratio is only weakly dependent on the the ltered drag coefcient. The variation of the dimension-
less ltered drag coefcient with particle volume fraction
particle volume fraction for the range of 0.03 \ /s \ 0.30. for various lter sizes (listed in the legend in dimensionless
The ratio within this range can be represented as a function units) is shown. All conditions are as in Figure 6a.

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1439
Figure 7. The contribution of the (dimensionless) pres- Figure 9. The variation of ltered dimensionless slip
sure uctuation term to the dimensionless l- velocity with ltered particle volume fraction
tered drag coefcient shown earlier in Figure is shown for various dimensionless lter
6a is presented. lengths shown in the legend.
All conditions are as in Figure 6a. The dimensionless lter These results were generated from the same set of simula-
lengths are shown in the legend. tion data that led to Figure 6a.

It was seen earlier, Eq. 18, that the ltered drag force curve in this gure corresponds to the intrinsic drag expres-
includes contributions from two terms. The second term is sion extracted simply using a lter size equal to the grid size
essentially equal to /0s rp0g as the deviatoric stress in the of the simulations. The two obvious regions manifested by
gas phase is quite small. The contribution from this term to this uppermost curve can be traced to a Reynolds number
the ltered drag coefcient is presented in Figure 7, while (Reg) effect present in the Wen and Yu6 drag expression
the total contribution due to both terms was shown earlier in used in the simulations. The ltered slip velocity in the verti-
Figure 6a. Although the overall ltered drag coefcient cal direction, as a function of /s , is shown in Figure 9 for
decreases with increasing lter size (Figure 6a), the contribu- various lter sizes. Here, the bottommost curve is for
tion from /0s rp0g rst increases with the lter size and then the case where the lter size is the same as the grid size; the
decreases (Figure 7). However, /0s rp0g contributes no more inverse relationship between the local slip velocity and the
than 25% of the overall ltered drag coefcient. So, over this range particle volume fraction is clear. It can be seen from Eq. 7
 comes from f.
of lter sizes, the primary contribution to F that b increases with /s and Reg; in the uppermost curve in
The results presented in Figure 6a are plotted in Figure 8 Figure 8, the effect of /s dominates at high /s values, while
on a logarithmic scale which shows: (a) the typical Richard- the Reg effect leads to a reversal of trend at very small /s
sonZaki46 form for /s not too close to zero, and (b) at small values. To establish this point, we carried out simulations
/s values, a clear departure from this trend. The uppermost where the intrinsic drag coefcient expression (see Eq. 7)
was modied by setting CD 5 24/Reg (so that only the
Stokes drag remained). Figure 10 shows the results obtained
from these simulations, cf. Figure 8. The uppermost curve in
Figure 10 does not show the reversal of trend at very small
/s values, establishing Reynolds number effect as the reason
for the difference between the shapes of the uppermost
curves in Figures 8 and 10.
Let us now consider the other curves in both Figures 8 and
10, which are for lter sizes larger than grid size. All of these
curves exhibit a RichardsonZaki like behavior at high volume
fractions and a reversal of trend at very low particle volume frac-
tions. This behavior is not due to an Reg effect in the intrinsic
drag law, as Figure 10 does not have any such dependence, and
so one has to seek an alternate explanation. The results presented
in Figure 9 indicate that one cannot capture this effect through a
Reynolds number term involving the ltered slip velocity. Note
that for large lter sizes, the slip velocity manifests a peak at
some intermediate /s ; for /s values to the left of this peak, the
Figure 8. The results shown earlier in Figure 6a are ltered slip velocity decreases as /s is decreased, while the
plotted on a natural logarithmic scale. quantity plotted in Figures 8 and 10 increase with decreasing
Here Q /b/d , where bd is the dimensionless ltered drag /s . Thus, if we seek to capture the data in Figures 8 and 10 in
coefcient, s/gs is particle volume fraction, and /g 5 1 2
/s is the gas volume fraction. The dimensionless lter the low /s region through a Reynolds number dependence
lengths are shown in the legend. (based on the ltered slip velocity), it will involve a negative

1440 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
Figure 10. These results are analogous to those shown
earlier in Figure 8, with the only difference
being that the intrinsic drag force model used
in the simulations that led to the present g-
ure did not include a Reynolds number de-
pendence.
The dimensionless lter lengths are shown in the legend.

order dependence, which makes no physical sense. Therefore


we attribute the trend reversal seen in Figures 8 and 10 at small
/s values to just the inhomogeneous microstructure inside the
lter region. At low /s values, an increase in /s increases both
the cluster size and particle volume fraction in the clusters; the
gas ows around these clusters and the resistance offered by
these clusters decreases with increasing cluster size. Large lter
sizes average over larger clusters and so the extent of drag
reduction observed increases with lter size. At sufciently
large /s values, the clusters begin to interact and hindered drag
sets in. This behavior is clearly reected in the vertical slip ve-
locity corresponding to large lter sizes, see Figure 9. The slip
velocity increases with /s at small /s values, consistent with
larger and/or denser clusters; it then decreases with increasing
/s when the clusters begin interact with each other.
It is interesting to note in Figure 9 that the dimensionless
slip velocity, in the limit of zero particle volume fraction
(/s ! 0), differs from unity. In our simulations with various
domain-average particle volume fractions, regions with /s ! 0
appeared in the dilute phase surrounding the clusters; here the
slip velocity was almost always larger than the terminal veloc-
ity. This implies that the gas in the dilute phase was constantly Figure 11. (a) Dimensionless apparent terminal velocity
engaged in accelerating the particles upward. This can happen for different dimensionless lter lengths,
only if the clusters are dynamic in nature with active, continual extracted from results in Figure 8 (2D) for
exchange of particles between the clusters and the dilute phase. the range 0.075 /s 0.30 and those
Linear ts of the data in Figure 8 over the particle volume extracted from results in Figure 19 (3D) for
fraction range 0:075
/s
0:30 were used to estimate the range 0.075 /s 0.25.
dimensionless apparent terminal velocity Vt, app and an appa- The solid lines in Figure 8 are based on the apparent ter-
rent RichardsonZaki exponent, NRZ,app. minal velocity shown here and the apparent Richardson
  ! Zaki exponent in Figure 11b. (b) Apparent Richardson
be vt
Zaki exponent for different dimensionless lter lengths,
be;d
ln ln  extracted from results in Figure 8 (2D) for the range 0.075
qs g/s 1  /s /s 1  /s
/s
0.30. The solid line in Figure 8 for a lter length
of 2.056 is based on this apparent terminal velocity in Fig-
NRZ;app  1 ln1  /g  lnVt;app : ure 11a and the apparent RichardsonZaki exponent shown
here. (c) Apparent RichardsonZaki exponent for different
dimensionless lter lengths, extracted from results in Figure
The variation of Vt,app and NRZ,app with dimensionless lter 19 (3D) for the range 0.075
/s
0.25.
size, Fr1
f 5 gDf/v2t , are shown in Figures 11a (diamonds), b,

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1441
respectively. Here, Df denotes the lter size. Both increase
with lter size.
Figure 12a shows the variation with /s of the dimension-
less ltered kinetic theory pressure, ps;d ps =qs v2t , for the
simulations discussed earlier in connection with Figures 4
and 8. At very low /s values, the ltered kinetic theory pres-
sure is essentially independent of lter size, but at larger /s
values distinct lter size dependence becomes clear. Figure
12b shows the dimensionless total particle phase pressure
pse;d pse =qs v2t as a function of /s for various lter sizes.
Here, the ltered particle phase pressure includes the pres-
sure arising from the streaming and collisional parts captured
by the kinetic theory and the sublter-scale Reynolds-stress
like velocity uctuations (see text below Eq. 21). Comparing
Figures 12a, b, it is seen that the contribution resulting from
the sublter-scale velocity uctuations is much larger than
the kinetic theory pressure indicating that, at the coarse-grid
scale, one can ignore the kinetic theory contributions to the
pressure. It is also clear from Figure 12b that the ltered
pressure increases with lter size, a direct consequence of
the fact that the energy associated with the velocity uctua-
tions increases with lter length (as in single phase turbu-
lence). Once again, results obtained from simulations with
different domain-average particle volume fractions collapse
on to the same curves (as earlier in Figures 5 and 6 for the
ltered drag coefcient), conrming that the ltered quanti-
ties largely depend on quantities inside the ltering region.
The data presented in Figure 12b could be captured by an
expression of the form a/s(1 2 b/s) with b  1.80. The para-
meter a increases with lter size, see Figure 12c.
Figures 13a, b show the variation with /s of dimensionless
ltered kinetic theory viscosity, ls;d ls g=qs v2t , and the
(dimensionless) ltered particle phase shear viscosity,
lse;d lse g=qs v3t . The latter includes the streaming and colli-
sional parts captured by the kinetic theory (shown in Figure
13a) and that associated with the sublter-scale velocity uc-
tuations. It is readily seen that for large lter sizes, the con-
tribution from the sublter scale velocity uctuations domi-
nate, and the ltered particle phase viscosity increases appre-
ciably with lter size. Once again, results from simulations
with different domain-average particle volume fractions col-
lapse on the same curves, adding further support to the via-
bility of the ltering approach. The data presented in Figure
13b could be captured by an expression of the form c/s(1 2
d/s) with d  0.86. The parameter c increases with lter size;
see Figure 13c.
It is mentioned in passing that we have studied the robust-
ness of the ltered statistics against small changes in the sec- Figure 12. (a) The variation of the dimensionless l-
ondary model parameters (namely, the coefcient of restitu- tered kinetic theory pressure with particle
tion, density ratio, etc.) and found that they are much less volume fraction is presented for different
important than the dimensionless lter size; so, capturing the dimensionless lter lengths.
effect of the dimensionless lter size on the dimensionless The results were extracted from simulations mentioned in
ltered drag coefcient is indeed the most important the caption for Figure 6a. The dimensionless lter lengths
are shown in the legend. (b) The variation of the dimen-
challenge. sionless ltered particle phase pressure with particle vol-
Agrawal et al.30 and Andrews et al.31 determined domain- ume fraction is presented for different dimensionless lter
averaged drag coefcient, particle phase pressure and viscos- lengths. The results were extracted from simulations men-
tioned in the caption for Figure 6a. The dimensionless l-
ity by averaging their kinetic theory simulation results over ter lengths are shown in the legend. (c) The coefcient
the entire periodic domain. In contrast, we have performed a of the dimensionless ltered particle phase pressure
in Figure 12b represented as a/s(1 2 b/s) (for /s

the averaging over regions that are much smaller than the 0.30) is plotted against the dimensionless lter length.
periodic domain, so that the ltered statistics are not affected b  1.80 for all lters.
by the periodic boundary conditions. It is interesting to

1442 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
observe that the lter size dependences of all these ltered
quantities obtained in our study are qualitatively identical to
those reported in the studies of Agrawal et al.30 and Andrews
et al.31 This further conrms that the robustness of the role
played by lter size.

Three-Dimensional Simulations
Figure 14 shows a snapshot of the particle volume fraction
eld in a 3D periodic domain, and the presence of particle-
rich strands is readily visible. Figure 15 shows the effect of
grid resolution on the ltered drag coefcient. As seen earlier
in Figure 3 for 2D simulations, the dependence of the ltered
drag coefcient on grid resolution becomes weaker as the l-
ter size increases. At the lower grid resolution, the lter size
of 1.028 is only twice the grid size and when the grid resolu-
tion is increased, the ltered drag coefcient changes appre-
ciably. For a lter size of 4.112, there are 512 and 4096
grids inside the lter volume in the two simulations; these
are quite large and so the ltered drag coefcient manifests
only a weak dependence on resolution.
Figure 16 displays the variation of ltered drag coefcient
with particle volume fraction for different lter sizes. As the
grid size used in these simulations is 0.257 dimensionless
units, the uppermost curve corresponds to using no lter at
all. The next curve corresponding to the lter size of 0.514
has only eight grids inside the ltering volume and so is
likely to change if simulations with greater resolutions are

Figure 13. (a) The variation of the dimensionless l-


tered kinetic theory viscosity with particle
volume fraction is presented for different
dimensionless lter lengths.
The results were extracted from simulations mentioned in
the caption for Figure 6a. The dimensionless lter lengths
from the top curve to the bottom curve are shown in the
legend. (b) The variation of the dimensionless ltered
particle phase viscosity with particle volume fraction is Figure 14. A snapshot of the particle volume fraction
presented for different dimensionless lter lengths. The
results were extracted from simulations mentioned in eld in a large periodic domain of size 16.448
the caption for Figure 6a. The dimensionless lter lengths 3 16.448 3 16.448 dimensionless units is
are shown in the legend. (c) The coefcient c of the shown.
dimensionless ltered particle phase viscosity in Figure
13b represented as c/s(1 2 d/s) (for /s
0.30) is plot- Simulation was performed using 64 3 64 3 64 grid
ted against the dimensionless lter length. d  0.86 for points. The domain-average particle volume fraction, h/si
all lters. 5 0.05.

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1443
Figure 15. The effect of grid resolution on the dimen-
sionless ltered drag coefcient is pre- Figure 17. The effect of the domain-average particle
sented. volume fraction on the dimensionless l-
Simulations were performed in a cubic periodic domain of
tered drag coefcient is presented.
size 16.448 3 16.448 3 16.448 dimensionless units and Simulations were performed in a cubic domain of size
at two different grid resolutions (323 and 643). The ltered 16.448 3 16.448 3 16.448 dimensionless units using 64
drag coefcients were calculated for dimensionless lter 3 64 3 64 grid points. The ltered drag coefcients
lengths of 1.028 and 4.112. Data used for ltering were were calculated for dimensionless lter lengths of 1.028
generated by running simulations for domain-average par- (top curve) and 2.056 (bottom curve). Data used for lter-
ticle volume fractions of 0.05, 0.10, 0.15, 0.20, and 0.35. ing were generated by running simulations for domain-av-
erage particle volume fractions of 0.05, 0.10, 0.15, 0.20,
and 0.25 (shown by different symbols in each curve).
performed. The results for lter sizes larger than 2.056 are
expected to be nearly independent of grid resolution. It is ltered drag coefcient is largely determined by the inhomoge-
clear from Figure 16 that the lter size dependence of the l- neous microstructure inside the ltering volume. The results
ter drag coefcient seen earlier in the 2D simulations persist presented in Figure 16 are plotted on a logarithmic scale in
in 3D as well. Figure 19. RichardsonZaki like behavior at high particle vol-
As in the case of 2D simulations, the ltered drag coef- ume fractions and a reversal of the trend at lower volume frac-
cient obtained from 3D simulations at different domain-average tions, seen earlier in 2D simulations (see Figure 8), persist in
particle volume fractions collapse onto the same curve (over 3D as well. Filter size dependence of the apparent terminal ve-
the range of volume fractions displayed), see Figure 17. Fur- locity and the exponent in the RichardsonZaki regime, are
thermore, Figure 18 illustrates the ltered drag coefcient is shown in Figures 11a, c. The apparent terminal velocity
indeed independent of the domain size. These suggest that the increases with lter size, just as it did for 2D simulations; how-
ever, the RichardsonZaki exponent shows a slight decline
with increasing lter size, in marked contrast to 2D simulations

Figure 16. The variation of the dimensionless ltered


drag coefcient with particle volume frac-
tion for various lter sizes (listed in the
Figure 18. The effect of domain size on the dimension-
legend in dimensionless units) is shown.
less ltered drag coefcient is presented for
Simulations were in a square domain of size 16.448 3 a dimensionless lter length of 2.056.
16.448 3 16.448 dimensionless units, using 64 3 64 3
64 grid points. Data used for ltering were generated by Simulations were performed at domain-average particle vol-
running simulations for domain-average particle volume ume fractions of 0.05, 0.15, 0.25, and 0.35 in two different
fractions of 0.01, 0.02, 0.05, 0.10, 0.15, 0.20, 0.25, and cubic periodic domains of sizes: 16.448 3 16.448 3 16.448
0.35. The dimensionless lter lengths from the top curve dimensionless units (64 3 64 3 64 grids) and 8.224 3
to the bottom curve are shown in the legend. 8.224 3 8.224 dimensionless units (32 3 32 3 32).

1444 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
Figure 21. Dimensionless ltered particle phase vis-
Figure 19. The results shown earlier in Figure 16 are
cosity for different dimensionless lter
plotted on a natural logarithmic scale.
lengths, extracted from simulations men-
Here Q /b/d , where bd is the dimensionless ltered drag
coefcient, s/gs is particle volume fraction, and /g 5 1 2 tioned in the caption for Figure 16.
/s is the gas volume fraction. The dimensionless lter The dimensionless lter lengths are shown in the legend.
lengths are shown in the legend.

coarser structures than the microscopic two-uid models


(see Figure 11b). Thus, there are denite quantitative differen- would. We now demonstrate that this is indeed the case.
ces between 2D and 3D results; however, it is clear from Consider the fate of a small disturbance imposed on a uni-
Figures 8 and 19 that both 2D and 3D results are strikingly formly uidized bed of innite extent, as predicted by the l-
similar. tered model equations. It can readily be shown that the state
Figures 20 and 21 present ltered particle phase pressure of uniform uidization is most unstable to disturbances that
and viscosity extracted from 3D simulations and can be com- take the form of one-dimensional traveling waves having no
pared to Figures 12b and 13b, respectively. The strong lter horizontal structure.3 The growth rates of such one-dimen-
size dependence of these quantities is clearly present in both sional disturbances of various wavenumbers as predicted by
two- and three-dimensions. the ltered two-uid model equations, with closures deter-
mined in this study for different lter sizes, are presented in
Figures 22a, b for uniformly uidized beds at two different
Linear Stability Analysis of the Filtered
Two-Fluid Model Equations volume fractions. As seen in these gures, the state of uni-
form uidization is unstable over a range of wavenumbers
If the ltering process and the general form of the ltered (k), 0 \ k \ kHB, where kHB denotes the wavenumber at
constitutive models are meaningful, one would expect that which a Hopf bifurcation occurs.3 Disturbances whose wave-
the ltered model equations should afford considerably number exceed kHB will decay, while those in the range 0 \
k \ kHB will grow Thus, one can expect that the size of the
region 0 \ k \ kHB is a measure of the range of wavenum-
bers that one is most likely to see in transient simulations. It
is clear from these gures that kHB decreases monotonically
as the lter size is increased, revealing that the larger the l-
ter size the coarser the structures resolved in the ltered two-
uid model are. Thus, the ltering operation has indeed aver-
aged over the ne structures and generated equations and
constitutive models that are suitable for integration over
coarser grids.

Summary
We have presented a methodology where computational
results obtained through highly resolved simulations (in a
large periodic domain) of a given microscopic two-uid
model are ltered to deduce closures for the corresponding
Figure 20. Dimensionless ltered particle phase pres- ltered two-uid model equations. These ltered closures
sure for different dimensionless lter depend on the lter size and can readily be constructed for a
lengths, extracted from simulations men- range of lter sizes. To a good approximation, the dimen-
tioned in the caption for Figure 16. sionless ltered drag coefcient, particle phase pressure and
The dimensionless lter lengths are shown in the legend. particle phase viscosity can be treated as functions of only

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1445
the kinetic theory strives to model, are not important; these
uctuations inuence the inhomogeneous microstructure and
their velocity uctuations, and hence the closures for the l-
tered equations.
Linear stability analysis of the ltered two-uid model
equations, with closures corresponding to several different
lter lengths, showed that ltering is indeed erasing the ne
structure and only presenting coarser structures.
It is clear from our simulation results that there is a strik-
ing similarity between the 2D and 3D results. Although there
are quantitative differences between 2D and 3D, the follow-
ing characteristics were found to be common between them:
(a) The ltered drag coefcient decreased with increasing
lter size, and (b) The ltered particle phase pressure and
viscosity increased with lter size. It seems reasonable to
expect that the clusters will not grow beyond some critical
size; if this is indeed the case, the ltered drag coefcient,
and particle phase pressure and viscosity will become nearly
independent of the lter size beyond some critical value. It is
important to understand if such saturation occurs and, if so,
at what lter size. It is also important to incorporate the
effects of bounding walls on the ltered closures as compari-
son of the ltered model predictions with experimental data
cannot be pursued until this issue is addressed. These are fer-
tile problems for further research.
In the present study, the /0s rp0g term has been absorbed
into the ltered the drag force. Zhang and VanderHeyden42
and de Wilde43,44 argue that /0s rp0g should also include a
dynamic part (namely, an added mass force). Andrews45
found in his simulation study that the principal contribution
of /0s rp0g was to a ltered drag force term, which has been
included in our study. A more thorough investigation of the
dynamic contribution would also be of interest.
Figure 22. 1D Linear stability analysis (LSA) of the l-
tered equations extracted from the 2D sim-
ulations for various dimensionless lter Acknowledgments
lengths shown in the legend. (a) h/si 5 0.15
(b) h/si 5 0.25. This work was supported by the US Department of Energy (grants:
CDE-FC26-00NT40971 and DE-PS26-05NT42472-11) and the Exxon-
Mobil Research and Engineering Company. Andrews and Igci acknowl-
edge summer training on MFIX at the National Energy Technology Lab-
particle volume fraction and dimensionless lter size. The oratory, Morgantown, WV. The authors thank Madhava Syamlal, Chris
effective drag coefcient to describe the interphase interac- Guenther, Ronald Breault, and Soane Benyahia for their assistance
throughout the course of this study.
tion force in the ltered equations shows two distinct
regimes. At particle volume fractions greater than about
0.075, it follows an effective RichardsonZaki relationship,
Notation
and the effective R-Z exponent and apparent terminal veloc-
ity have an understandable physical interpretation in terms of CD 5 single particle drag coefcient
d 5 particle diameter, m
interactions between particle clusters instead of the individual ep 5 coefcient of restitution for particleparticle collisions
particles. At low particle volume fractions, the drag coef- f 5 interphase interaction force per unit volume in the micro-
cient shows an anomalous behavior that is consistent with scopic two-uid model, kg/m2 s2
the formation of larger and denser clusters with increasing f 5 ltered value of f, kg/m2 s2
particle volume fraction.  5 interphase interaction force per unit volume in the ltered
F
two-uid model, kg/m2 s2
The velocity uctuations associated with the very compli- Frf 5 Froude number based on lter size 5 v2t /gDf
cated inhomogeneous structures shown by the microscopic g, g 5 acceleration due to gravity, m/s2
two-uid simulations dictate the magnitudes of the ltered gO 5 value of radial distribution function at contact (see expres-
particle phase pressure, and viscosity. The contributions of sion in Table 1)
the kinetic theory pressure and viscosity to these ltered G(x,y) 5 weight function, m23
Jcoll 5 rate of dissipation of granular energy per unit volume by col-
quantities are negligibly small and so, for practically relevant lisions between particles, kg/m s3
lter sizes, one need not include the ltered granular energy Jvis 5 rate of dissipation of granular energy per unit volume by the
equation in the analysis. This, however, does not mean that relative motion between gas and particles, kg/m s3
the uctuations at the level of the individual particles, which NRZ,app 5 apparent RichardsonZaki exponent

1446 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal
pg,ps 5 gas and particle phase pressures in the microscopic two-uid 8. Hill RJ, Koch DL, Ladd AJC. The rst effects of uid inertia on
model, respectively, kg/m s2 ow in ordered and random arrays of spheres. J Fluid Mech. 2001;
pg , ps 5 ltered value of pg and ps, respectively, kg/m s2 448:213241.
ps;d 5 ps made dimensionless; ps;d ps =qs vt
2 9. Hill RJ, Koch DL, Ladd AJC. Moderate-Reynolds-number ows in
ordered and random arrays of spheres. J Fluid Mech. 2001;448:243
pse 5 ltered particle phase pressure, kg/m s
2
278.
pse;d 5 pse made dimensionless; pse;d pse =qs vt
2
10. Wylie JJ, Koch DL, Ladd AJC. Rheology of suspensions with high
q 5 ux of granular energy, kg/s3 particle inertia and moderate uid inertia. J Fluid Mech. 2003; 480:
Reg 5 single particle Reynolds number 95118.
t 5 time, s 11. Khandai D, Derksen JJ, van den Akker HEA. Interphase drag coef-
T 5 granular temperature, m2/s2 cients in gas-solid ows. AIChE J. 2003;49:10601063.
u, v 5 gas and particle phase velocities in the microscopic two-uid 12. Li J, Kuipers JAM. Gas-particle interactions in dense gas-uidized
model, respectively, m/s beds. Chem Eng Sci. 2003;58:711718.
, v 5 ltered gas and particle phase velocities, respectively, m/s
u 13. van der Hoef MA, Boetstra R, Kuipers JAM. Lattice Boltzmann
u0 , v0 5 uctuations in gas and particle phase velocities, respectively, simulations of low Reynolds number ow past mono- and bi-dis-
m/s perse arrays of spheres: results for the permeability and drag forces.
vt 5 terminal settling velocity, m/s J Fluid Mech. 2005;528:233254.
Vt,app 5 dimensionless apparent terminal velocity 14. Benyahia S, Syamlal M, OBrien TJ. Extension of Hill-Koch-Ladd
x, y 5 position vectors, m drag correlation over all ranges of Reynolds number and solids vol-
ume fraction. Powder Technol. 2006;162:166174.
15. Gidaspow D, Jung J, Singh RJ. Hydrodynamics of uidization using
Greek letters kinetic theory: an emerging paradigm. 2002 Fluor-Daniel Lecture.
Powder Technol. 2004;148:123141.
b 5 drag coefcient in the microscopic two-uid model, kg/m3 s 16. Koch DL, Sangani AS. Particle pressure and marginal stability limits
be 5 ltered drag coefcient, kg/m s for a homogeneous monodisperse gas uidized bed: kinetic theory
3
and numerical simulations. J Fluid Mech. 1999;400:229263.
be;d 5 dimensionless ltered drag coefcient 5 be vt =qs g 17. Huilin L, Yurong H, Gidaspow D, Lidan Y. Yukun Q. Size segrega-
/s, /g 5 particle and gas phase volume fractions, respectively tion of binary mixture of solids in bubbling uidized beds. Powder
/s , /g 5 ltered particle and gas phase volume fractions, respectively Technol. 2003;134:8697.
/0s 5 uctuation in particle phase volume fraction 18. Iddir H, Arastoopour H, Hrenya CM. Analysis of binary and ternary
h/si 5 domain-average particle volume fraction granular mixtures behavior using the kinetic theory approach. Pow-
/s,max 5 maximum particle volume fraction der Technol. 2005;151:117125.
qs, qg 5 particle and gas densities, respectively, kg/m3 19. Arnarson BO, Jenkins JT. Binary mixtures of inelastic spheres: sim-
Df 5 lter size, m plied constitutive theory. Phys Fluids. 2004;16:45434550.
rs, rg 5 particle and gas phase stress tensors in the microscopic two- 20. Jenkins J, Mancini F. Kinetic theory for binary mixtures of smooth,
uid model, respectively, kg/m s2 nearly elastic spheres. Phys Fluids A. 1989;1:20502057.
rs , rg 5 ltered values of rs and rg, respectively, kg/m s
2
21. Benyahia S, Arastoopour H, Knowlton TM. Simulation of particles
Ss 5 ltered total particle phase stress, kg/m s2 and gas ow behavior in the riser section of a circulating uidized
Gslip 5 rate of generation of granular energy per unit volume by bed using the kinetic theory approach for the particulate phase. Pow-
gas-particle slip, kg/m s3 der Technol. 2000;112:2433.
g, a, k, l 5 quantities dened in Table 1 22. Ding J, Gidaspow D. A Bubbling uidization model using kinetic-
k*, l* 5 quantities dened in Table 1 theory of granular ow. AIChE J. 1990;36:523538.
ks 5 granular thermal conductivity, kg/m s 23. Goldschmidt MJV, Kuipers JAM. van Swaaij WPM. Hydrodynamic
lg 5 gas phase viscosity, kg/m s modeling of dense gas-uidized beds using the kinetic theory of
l^g 5 effective gas phase viscosity appearing in the microscopic granular ow: effect of restitution coefcient on bed dynamics.
two-uid model (taken to be equal to lg itself in our simula- Chem Eng Sci. 2001;56:571578.
tions) kg/m s 24. Neri A, Gidaspow D. Riser hydrodynamics: simulation using kinetic
lb, lg 5 bulk and shear viscosities of the particle phase appearing in theory. AIChE J. 2000;46:5267.
the kinetic theory model, kg/m s 25. Lun CKK, Savage SB, Jeffrey DJ, Chepurniy N. Kinetic theories of
ls;d 5 ls made dimensionless; ls;d ls g=qs v3t granular ows: inelastic particles in Couette ow and slightly inelas-
tic particles in a general ow eld. J Fluid Mech. 1984;140:223
lbe , lse 5 bulk and shear viscosities of the particle phase appearing in 256.
the ltered two-uid model, kg/m s
26. Sinclair JL, Jackson R. Gas-particle ow in a vertical pipe with
lse;d 5 lse made dimensionless; lse;d lse g=qs v3t particle-particle interaction. AIChE J. 1989;35:14731486.
27. Pita JA, Sundaresan S. Gas-solid ow in vertical tubes. AIChE J.
1991;37:10091018.
28. Pita JA, Sundaresan S. Developing ow of a gas-particle mixture in
Literature Cited a vertical riser. AIChE J. 1993;39:541552.
29. Louge M, Mastorakos E, Jenkins JT. The role of particle collisions
1. Grace JR, Bi H. In: Grace JR, Avidan AA, Knowlton TM, editors. in pneumatic transport. J Fluid Mech. 1991;231:345359.
Circulating Fluidized Beds, 1st ed. New York: Blackie Academic & 30. Agrawal K, Loezos PN, Syamlal M, Sundaresan S. The role of
Professional, 1997; Chapter 1; pp 120. meso-scale structures in rapid gas-solid ows. J Fluid Mech. 2001;
2. Gidaspow D. Multiphase Flow and Fluidization. CA: Academic 445:151185.
Press, 1994; Chapter 4; pp 99152. 31. Andrews AT IV, Loezos PN, Sundaresan S. Coarse-grid simulation
3. Jackson R. The Dynamics of Fluidized Particles. Cambridge Univer- of gas-particle ows in vertical risers. Ind Eng Chem Res. 2005;44:
sity Press, 2000. 60226037.
4. Syamlal M, Rogers W, OBrien TJ. MFIX Documentation. Morgan- 32. Sundaresan S. Perspective: modeling the hydrodynamics of multi-
town, WV: U.S. Department of Energy, Federal Energy Technology phase ow reactors: current status and challenges. AIChE J. 2000;
Center, 1993. 46:11021105.
5. Syamlal M. MFIX Documentation: Numerical Techniques. DOE/ 33. OBrien TJ, Syamlal M. Particle cluster effects in the numerical sim-
MC-31346-5824. NTIS/DE98002029; 1998. Also see www.mx.org. ulation of a circulating uidized bed. In: Avidan A, editor. Circulat-
6. Wen CY, Yu YH. Mechanics of uidization. Chem Eng Prog Symp ing Fluidized Bed Technology. IV. Proceedings of the Fourth Inter-
Ser. 1966;62:100111. national Conference on Circulating Fluidized Beds, Hidden Valley
7. Ergun S. Fluid ow through packed columns. Chem Eng Prog. Conference Center and Mountain Resort, Somerset, PA, August 15,
1952;48:8994. 1993.

AIChE Journal June 2008 Vol. 54, No. 6 Published on behalf of the AIChE DOI 10.1002/aic 1447
34. Boemer A, Qi H, Hannes J, Renz U. Modelling of solids circulation 40. Glasser BJ, Sundaresan S, Kevrekidis IG. From bubbles to clusters
in a uidised bed with Eulerian approach. 29th IEA-FBC Meeting in in uidized beds. Phys Rev Lett. 1998;81:18491852.
Paris, France, Nov. 2426, 1994. 41. Glasser BJ, Kevrekidis IG, Sundaresan S. One- and two-dimensional
35. Heynderickx GJ, Das AK, de Wilde J, Marin GB. Effect of cluster- travelling wave solutions in gas-uidized beds. J Fluid Mech. 1996;
ing on gas-solid drag in dilute two-phase ow. Ind Eng Chem Res. 306:183221.
2004;43:46354646. 42. Zhang DZ, VanderHeyden WB. The effects of mesoscale structures
36. McKeen T, Pugsley T. Simulation and experimental validation of a on the macroscopic momentum equations for two-phase ows. Int J
freely bubbling bed of FCC catalyst. Powder Technol. 2003;129: Multiphase Flow. 2002;28:805822.
139152. 43. De Wilde J. Reformulating and quantifying the generalized added
37. Yang N, Wang W, Ge W, Wang L. Li, J. Simulation of heterogene- mass in ltered gas-solid ow models. Phys Fluids. 2005;17:114.
ous structure in a circulating uidized bed riser by combining the 44. De Wilde J. The generalized added mass revised. Phys Fluids. 2007;19:14.
two-uid model with EMMS approach. Ind Eng Chem Res. 2004; 45. Andrews AT IV. Filtered models for gas-particle hydrodynamics.
43:55485561. PhD Dissertation, Princeton University, Princeton, NJ, 2007.
38. Dasgupta S, Jackson R, Sundaresan S. Turbulent gas-particle ow in 46. Richardson JF, Zaki WN. Sedimentation and uidization. I. Trans
vertical risers. AIChE J. 1994;40:215228. Inst Chem Eng. 1954;32:3553.
39. Hrenya CM, Sinclair JL. Effects of particle-phase turbulence in gas-
solid ows. AIChE J. 1997;43:853869. Manuscript received Sept. 26, 2007, and revision received Feb. 14, 2008.

1448 DOI 10.1002/aic Published on behalf of the AIChE June 2008 Vol. 54, No. 6 AIChE Journal

Das könnte Ihnen auch gefallen