You are on page 1of 394

CONTRIBUTORS

Numbers in parentheses indicate the pages on which the authors contribution begin.
A. T. ANDREWS IV, Department of Chemical Engineering, Princeton University,
Princeton, NJ 08544, USA (65)
A. G. DIXON, Department of Chemical Engineering, Worcester Polytechnic
Institute, Worcester, MA 01609, USA (307)
L.-S. FAN, Department of Chemical and Biomolecular Engineering, The Ohio
State University, 140 West 19th Avenue, Columbus, OH 43210, USA (1)
R. O. Fox, Herbert L. Stiles Professor of Chemical Engineering, Iowa State
University, 3162 Sweeney Hall, Ames, IA 50011-2230, USA (231)
Currently on sabbatical at:
Swiss Federal Institute of Technology Zurich, ETHZ Institut fur Chemie-
und Bioingenieurwissenschaften ETH-Honggerberg/HCI H 109 (Gruppe
Morbidelli), CH-8093 Zurich, Switzerland
Y. GE, Department of Chemical and Biomolecular Engineering, The Ohio State
University, 140 West 19th Avenue, Columbus, OH 43210, USA (1)
J. A. M. KUIPERS, University of Twente, Faculty of Science & Technology, PO
Box 217, NL - 7500 AE Enschede, The Netherlands (65)
M. NIJEMEISLAND, Johnson Matthey Catalysts, Billingham, UK (307)
E. H. STITT, Johnson Matthey Catalysts, Billingham, UK (307)
S. SUNDARESAN, Department of Chemical Engineering, Princeton University,
Princeton, NJ 08544, USA (65)
H. E. A. VAN DEN AKKER, Delft University of Technology, Molenwindsingel 50,
NL 4105 HK Culemborg, The Netherlands (151)
M. A. VAN DER HOEF, Department of Science and Technology, University of
Twente, PO 217, NL - 7500 AE Enschede, The Netherlands (65)
M. VAN SINT ANNALAND, Department of Science and Technology, University of
Twente, PO 217, NL - 7500 AE Enschede, The Netherlands (65)
M. YE, Department of Science and Technology, University of Twente, PO 217,
NL - 7500 AE Enschede, The Netherlands (65)
ix
PREFACE
This issue attempts to give a feeling of the state-of-the-art of the application
of computational uid dynamics (CFD) in chemical engineering. It is, however,
not limited to a snap-shot but is aimed at providing a perspective: how did we
arrive at the present status and where do we go from here? To do so, contri-
butions from ve complementary contributions are brought together. From the
definition of CFD as the ensemble of all computational approaches that solve
for the spatial distribution of the velocity, concentration, and temperature
elds recalled by Fox, it is clear that a selection had to be made as to the topics
covered. In the wake of volume 30 on Multiscale Analysis the present volume
is organized from small to large: from bubbles and droplets in the rst
contribution, to a xed catalyst bed in the last one. The application of direct
numerical simulations (DNS) clearly is still limited to the small scale. Today
subgrid-scale (SGS) models are required to cover the full spectrum.
The reader will be confronted with some redundancy but this allows each
contribution to stand on its own. Also, a good balance is maintained between
the style of a tutorial and that of a research paper. Those who will read the
complete volume will realize that opinions can vary from looking at CFD as an
alternative for experimentation to emphasizing the need of experimental val-
idation. Some contributions are entirely limited to velocity and temperature
elds. Others, on the contrary, emphasize the difculties associated with the
combination of transport and reaction. The latter can introduce stiffness even
for laminar ow. Averaging (e.g. Reynolds-averaged NavierStokes, RANS) or
ltering (e.g. large eddy simulations, LES), performed to model velocity elds,
does not alleviate this difculty. Clearly, this is still quite a challenge.
The contribution from the Ohio State University by Ge and Fan is dealing
with the simulation of gasliquid bubble columns and gasliquidsolid uidized
beds. A scientist of a major engineering company told me a few years ago that
when he wanted to know how serious an academic group was about CFD, he
would ask whether they could simulate bubble columns. He would only engage
into further conversation if the answer was negative. The group from Columbus
is wise enough to focus on a single air bubble rising in water, and bubble
formation from a single nozzle. In a second part the hydrodynamics and heat
transfer phenomena of a liquid droplet in motion and during the impact process
with a hot at surface, as well as with a particle are studied. The applied
numerical techniques, such as the level set and immersed boundary method, are
outlined and important contributions are highlighted. Next, detailed imple-
mentations for particular problems are presented. Finally, numerous simulation
results are shown and compared with experimental data.
xi
The second contribution addresses the different levels of modeling that are
required in order to cover the full spectrum of length scales that are important
for industrial applications. It is a joint paper from Twente and Princeton Uni-
versity and claims to put Emphasis on technical details. The latter is a too
modest description of what is really offered to the reader. The recent devel-
opments in two leading research groups on the modeling of gas-uidized beds
are presented. The holy grail for those interested in the design of industrial units
being the closure of the model equations in general and SGS modeling in par-
ticular. The latest developments of both the ltering approach pursued at
Princeton University by Sundaresan and coworkers and the discrete bubble
model developed in Twente by the team of Kuipers are presented. The authors
realize fully that there is still a long way to go, as evidenced by their last
sentence: Finally, the adapted model should be augmented with a thermal
energy balance, and associated closures for the thermo-physical properties, to
study heat transport in large scale uidized beds, such as FCC-regenerators and
PE and PP gas-phase polymerization reactors. This is even more so because
inclusion of reaction kinetics remains beyond the scope of the contribution!
Chemical reactions come into the picture in the context of stirred turbulent
vessels in Chapter 3. Van den Akker from Delft strongly emphasizes the po-
tential of LES and DNS for reproducing not only the hydrodynamics of tur-
bulent stirred vessels but also for providing a basis for simulating a wide variety
of physical and chemical processes in this equipment. The author advocates the
use of the latticeBoltzmann (LB) technique to this purpose. Van den Akker
certainly belongs to those who believe that one can and should be much more
positive about the merits of CFD so far and about the term at which CFD will
replace and improve existing mixing correlations. To quote him: It may be
easier to measure the local and transient details of the turbulent ows in stirred
vessels and the spatial distributions in e.g. mixing rates and bubble, drop and
crystal sizes computationally than by means of experimental techniques! When
it comes to the design of chemical reactors the authors admit that CFD is
certainly not a panacea. Scale-up of many chemical reactors, in particular the
multi-phase types, is still surrounded by a fame of mystery indeed.
The importance of chemical-reaction kinetics and the interaction of the latter
with transport phenomena is the central theme of the contribution of Fox from
Iowa State University. The chapter combines the clarity of a tutorial with the
presentation of very recent results. Starting from simple chemistry and single-
phase ow the reader is lead towards complex chemistry and two-phase ow.
The issue of SGS modeling discussed already in Chapter 2 is now discussed with
respect to the concentration elds. A detailed presentation of the joint Prob-
ability Density Function (PDF) method is given. The latter allows to account
for the interaction between chemistry and physics. Results on impinging jet
reactors are shown. When dealing with particulate systems a particle size dis-
tribution (PSD) and corresponding population balance equations are intro-
PREFACE xii
duced. The author emphasizes that a balance between the degree of detail or
complexity of the chemistry and that of the physics should be maintained.
The last contribution comes from Dixon (Worcester Polytechnic Institute),
and Nijemeisland and Stitt (Johnson Matthey). The subject is another classic of
reactor engineering: the catalytic xed-bed reactor. Heat transfer issues on both
reactor scale and catalyst pellet scale are addressed. Steam reforming is used as a
typical example of a strongly endothermic reaction requiring high-heat uxes
through the reactor walls. The presence of the tube wall causes changes in bed
structure, ow patterns, transport rates and the amount of catalyst per unit
volume, and is usually the location of the limiting heat-transfer resistance.
Special attention is given to the modeling of the structure of a packed bed.
The importance of wall functions, to be applied not only at the reactor wall but
also at the external pellet surface, is stressed. The authors show ample results of
their own work without neglecting the contributions of others. At the end of this
chapter the reader will be convinced of the importance of the local nonuni-
formities in the temperature eld not only within a catalyst pellet but also from
one pellet to the other.
Let me conclude by thanking the authors for their willingness to contribute,
despite health problems for some of them, and for their exibility with respect to
timing.
Guy B. Marin
Ghent, Belgium
April 2006
PREFACE xiii
3-D DIRECT NUMERICAL SIMULATION OF GASLIQUID
AND GASLIQUIDSOLID FLOW SYSTEMS USING THE
LEVEL-SET AND IMMERSED-BOUNDARY METHODS
Yang Ge and Liang-Shih Fan

Department of Chemical and Biomolecular Engineering, The Ohio State University,


Columbus, OH 43210, USA
I. Introduction 2
II. Front-Capturing and Front-Tracking Methods 4
A. Level-Set Method 6
B. Immersed Boundary Method 9
III. System 1: Flow Dynamics of GasLiquidSolid Fluidized
Beds 11
A. Numerical Procedure for Solving the GasLiquid
Interface 12
B. Governing Equations for the GasLiquidSolid Flow 13
C. Modeling the Motion and Collision Dynamics of Solid
Particles in GasLiquidSolid Fluidization 14
D. Results and Discussions 16
IV. System 2: Deformation Dynamics of Liquid Droplet in
Collision with a Particle with Film-Boiling Evaporation 27
A. Simulation of Saturated Droplet Impact on Flat Surface
in the Leidenfrost Regime 29
B. Simulation of Subcooled Droplet Impact on Flat
Surface in Leidenfrost Regime 38
C. Simulation of DropletParticle Collision in the
Leidenfrost Regime 49
V. Concluding Remarks 58
References 61
Abstract
The recent advances in level-set and Immersed Boundary methods
(IBM) as applied to the simulation of complex multiphase ow systems
are described. Two systems are considered. For system 1, a computa-
tional scheme is conceived to describe the three-dimensional (3-D) bubble

Corresponding author. Tel.: +1-614-688-3262(o). E-mail: fan@chbmeng.ohio-state.edu


1
Advances in Chemical Engineering, vol. 31
ISSN 0065-2377
DOI 10.1016/S0065-2377(06)31001-0
Copyright r 2006 by Elsevier Inc.
All rights reserved
dynamics in gasliquid bubble columns and gasliquidsolid uidized
beds. This scheme is utilized to simulate the motion of the gas, liquid, and
solid phases, respectively, based on the level-set interface tracking
method, the locally averaged time-dependent NavierStokes equations
coupled with the Smagorinsky subgrid scale stress model, and the Lag-
rangian particle motion equations. For system 2, the hydrodynamics and
heat-transfer phenomena of a liquid droplet in motion and during the
impact process with a hot at surface, as well as with a particle, are
illustrated. The 3-D level-set method is used to portray the droplet surface
deformation whilst in motion and during the impact process. The IBM is
employed so that the particleuid boundary conditions are satised. The
governing equations for the droplet and the surrounding gas phase are
solved utilizing the nite volume method with the Arbitrary Lagrangian
Eulerian (ALE) technique. To account for the multiscale effect due to
lubrication-resistance induced by the vapor layer between the droplet and
solid surface or solid particle formed by the lm-boiling evaporation, a
vapor-ow model is developed to calculate the pressure and velocity dis-
tributions along the vapor layer. The temperature elds in all phases and
the local evaporation rate on the droplet surface are illustrated using a
full-eld heat-transfer model.
I. Introduction
Gasliquidsolid (three-phase) ow systems involve a variety of operating
modes of gas, liquid, and solid phases, including those with solid particles and/or
liquid droplets in suspended states. Commercial or large-scale operations using
three-phase ow systems are prevalent in physical, chemical, petrochemical,
electrochemical, and biological processes (Fan, 1989). In the gasliquidsolid
uidization systems with liquid as the continuous phase, the systems are char-
acterized by the presence of gas bubbles, which induce signicant liquid mixing
and mass transfer. The ow structure in the systems is complex due to intricate
coalescence and breakup phenomena of bubbles. The fundamental dynamics of
solids suspensions in the systems is closely associated with the particleparticle
collision and particlebubble interactive behavior. For three-phase ows that
occur in the feed nozzle area of a uid catalytic cracking (FCC) riser in gas oil
cracking, on the other hand, the gas phase is continuous where oil is injected
from the nozzle with the mist droplets formed from the spray in contact with
high-temperature catalyst particles (Fan et al., 2001). The droplets may splash,
rebound, or remain on the catalyst particle surface after the impact, and the oil is
evaporated and cracked into lighter hydrocarbons. Such contact phenomena
are also prevalent in the condensed mode operation of the Unipol process for
YANG GE AND LIANG-SHIH FAN 2
polypropylene or polyethylene production, where dropletparticle collisions in
the feed nozzle are also accompanied by intense liquid evaporation. In this
study, both systems involving three-phase uidization and evaporative droplet
and particle collisions are simulated using CFD based on the 3-D level-set and
immersed boundary method (IBM).
CFD is a viable means for describing the uid dynamic and transport
behavior of gasliquidsolid ow systems. There are three basic approaches
commonly employed in the CFD for study of multiphase ows (Feng and
Michaelides, 2005): the EulerianEulerian (E-E) method, the EulerianLag-
rangian (E-L) method, and direct numerical simulation (DNS) method. In
the E-E method (Anderson and Jackson, 1967; Joseph and Lundgren, 1990;
Sokolichin and Eigenberger, 1994, 1999; Zhang and Prosperetti, 1994, 2003;
Mudde and Simonin, 1999), both the continuous phase and the dispersed phase,
such as particles, bubbles, and droplets, are treated as interpenetrating contin-
uous media, occupying the same space as does the continuous phase with differ-
ent velocities and volume fractions for each phase. In this method, the closure
relationships such as the stress and viscosity of the particle phase need to be
formulated. In the E-L method, or discrete particle method (e.g., Tsuji et al.,
1993; Lapin and Lu bbert, 1994; Hoomans et al., 1996; Delnoij et al., 1997), the
continuous uid phase is formulated in the Eulerian mode, while the position
and the velocity of the dispersed phase, particles, or bubbles, is traced in the
Lagrangian mode by solving Lagrangian motion equations. The grid size used in
the computation for the continuous-phase equations is typically much larger
than the object size of the dispersed phase, and the object in the dispersed phase
is treated as a point source in the computational cell. With this method, the
coupling of the continuous phase and the dispersion phase can be made using the
Particle-Source-In-Cell method (Crowe et al., 1977). The closure relationship for
the interaction forces between phases requires to be provided in the E-L method.
In the DNS (Unverdi and Tryggvason, 1992a,b; Feng et al., 1994a,b; Sethian
and Smereka, 2003), the grid size is commonly much smaller than the object size
of the dispersed phase, and the moving interface can be represented by implicit
or explicit schemes in the computational domain. The velocity elds of the uid
phase are obtained by solving the NavierStokes equation considering the in-
terfacial forces, such as surface tension force or soliduid interaction force.
The motion of the object of the dispersed phase is represented in terms of a
time-dependent initial-value problem. With the rapid advances in the speed and
memory capacity of the computer, the DNS approach has became important in
characterizing details of the complex multiphase ow eld.
This paper is intended to describe recent progress on the development of
the level-set method and IBM in the context of the advanced front-capturing
and front-tracking methods. The paper is also intended to discuss the appli-
cation of them for the 3-D DNS of two complex three-phase ow systems as
described earlier.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 3
II. Front-Capturing and Front-Tracking Methods
In the DNS of multiphase ow problems, there are various methods available
for predicting interface position and movement, such as the moving-grid
method, the grid-free method (Scardovelli and Zeleski, 1999) and the xed-grid
front-tracking/front-capturing method. In the moving-grid method, which is
also known as the discontinuous-interface method, the interface is a boundary
between two subdomains of the grid (Dandy and Leal, 1989). The grid may be
structured or unstructured and even near-orthogonal, moving with the interface
(Hirt et al., 1974). It treats the system as two distinct ows separated by a
surface. When the interface moves or undergoes deformation, new, geometri-
cally adapted grids need to be generated or remeshed (McHyman, 1984). The
remeshing can be a very complicated, time-consuming process, especially when
it involves a signicant topology change, and/or a 3-D ow. Methods in which
grids are not required include the marker particle method (Harlow and Welch,
1965) and the smoothed particle hydrodynamics method (Monaghan, 1994).
The xed-grid method, which is also known as the continuous-interface
method, employs structured or unstructured grids with the interface cutting
across the xed grids. It treats the system as a single ow with the density
and viscosity varying smoothly across a nite-thickness of the interface. The
numerical techniques used to solve the moving interface problem with xed,
regular grids can be categorized by two basic approaches: the front-tracking
method (e.g., Harlow and Welch, 1965; Peskin, 1977; Unverdi and Tryggvason
1992a, b; Fukai et al., 1995) and the front-capturing method (e.g., Osher and
Sethian, 1988; Sussman et al., 1994; Kothe and Rider, 1995; Bussmann et al.,
1999). For a 3-D multiphase ow problem, the xed-grid method is the most
frequently used due to its efciency and relative ease in programming.
The front-tracking method explicitly tracks the location of the interface by the
advection of the Lagrangian markers on a xed, regular grid. The marker-and-
cell (MAC) method developed by Harlow and Welch (1965) was the rst front-
tracking technique applied in DNS, e.g., it was used by Harlow and Shannon
(1967) to simulate the droplet impact on a at surface without considering
the viscosity and the surface-tension forces in the momentum-conservation
equation. Fujimoto and Hatta (1996) simulated the impingement process of a
water droplet on a high-temperature surface by using a single-phase 2-D MAC
type solution method. The no-slip and free-slip boundary conditions are itera-
tively adopted on the liquidsolid interface for the spreading and recoiling
process, respectively. Fukai et al. (1995) developed the adaptive-grid, nite-
element method to track the droplet free surface in collision with a surface while
considering the wettability on the contact line. The front-tracking method
developed by Unverdi and Tryggvason (1992a, b) and Tryggvason et al. (2001)
leads to many applications in the simulation of droplet or bubble ow. In this
method, the location of the interface is expressed by discrete surface-marker
YANG GE AND LIANG-SHIH FAN 4
particles. High-order interpolation polynomials are employed to ensure a high
degree of accuracy in the representation of the interface. An unstructured sur-
face grid connecting the surface-marker particles is introduced within a volu-
metric grid to track the bubble front within the computational domain. Thus,
discretization of the eld equations is carried out on two sets of embedded
meshes: (a) the Eulerian uid grid, which is 3-D, cubical, staggered structured,
and nonadaptive; and (b) the Largrangian front grid, which is 2-D, triangular,
unstructured, and adaptive (Unverdi and Tryggvason 1992a, b). The innitely
thin boundary can be approximated by a smooth distribution function of
a nite thickness of about three to four grid spacing. The variable density
NavierStokes equations can then be solved by conventional Eulerian tech-
niques (Unverdi and Tryggvason 1992a, b). This method can be numerically
stiff as the density ratio of the two uids increases, and may pose difculties
when the appearance, the connection, the detachment, and the disappearance of
the gasliquid interface are encountered. Such interface behavior occurs in the
coalescence, breakup, or formation of bubbles and droplets in an unsteady ow.
The front-tracking method is therefore computationally intensive. Agresar et al.
(1998) extended the front-tracking method with adaptive rened grids near the
interface to simulate the deformable circulation cell. Sato and Richardson
(1994) developed a nite-element method to simulate the moving free surface of
a polymeric liquid. The IBM proposed by Peskin (1977) in studying the blood
ow through heart valves and the cardiac mechanics also belongs to the class of
front-tracking techniques. In the IBM method, the simulation of the uid ow
with complex geometry was carried out using a Cartesian grid, and a novel
procedure was formulated to impose the boundary condition at the interface.
Some variants and modications of this method were proposed in simulating
various multiphase ow problems (Mittal and Iaccarino, 2005). An introduction
to the IBM method is given in Section II.B.
The front-capturing method, on the other hand, is the Eulerian treatment of
the interface, in which the moving interface is implicitly represented by a scalar-
indicator function dened on a xed, regular mesh point. The movement of the
interface is captured by solving the advection equation of the scalar-indicator
function. At every time step, the interface is generated by piecewise segments
(2-D) or patches (3-D) reconstructed by this scalar function. In this method, the
interfacial force, such as the surface-tension force, is incorporated into the ow-
momentum equation as a source term using the continuum surface force (CSF)
method (Brackbill et al., 1992). This technique includes the volume of uid
(VOF) method (Hirt and Nichols, 1981; Kothe and Rider, 1995), the marker
density function (MDF) (Kanai and Mtyata, 1998), and the level-set method
(Osher and Sethian, 1988; Sussman et al., 1994).
In the VOF method, an indicator function is dened as: 0 for a cell with pure
gas, 1 for a cell with pure liquid, and 0 to 1 for a cell with a mixture of gas and
liquid. An interface exists in those cells that give a VOF value of neither 0 nor 1.
Since the indicator function is not explicitly associated with a particular front
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 5
grid, an algorithm is needed to reconstruct the interface. This is not an easy
task, especially for a complex dynamic interface requiring 3-D calculation.
Pasandideh-Ford et al. (1998) used a modied SOLA-VOF method to solve the
momentum and heat-transfer equations for droplet deposition on a steel sur-
face. Bussmann et al. (1999, 2000) developed a 3-D model to simulate the
droplet collision onto an incline surface and its splash on the surface, utilizing a
volume-tracking methodology. Mehdi-Nejad et al. (2003) also used the VOF
method to simulate the bubble-entrapment behavior in a droplet when it im-
pacts a solid surface. Karl et al. (1996) simulated small droplet (100200 mm)
impact onto the wall in the Leidenfrost regime using a VOF method. A free-slip
boundary condition and a 1801 contact angle were applied on the solid surface.
Harvie and Fletcher (2001a,b) developed an axisymmetric, 2-D VOF algorithm
to simulate the volatile liquid droplet impacting on a hot solid surface. The
vapor ow between the droplet and solid surface was solved by a 1-D, creeping
ow model, which neglects the inertial force of the ow. This model, despite
being accurate at a lower We, failed to reproduce the droplet dynamics at a
higher Weber number. Other front-capturing methods include the constrained
interpolation prole (CIP) method (Yabe, 1997), and the phase-eld method
(Jamet et al., 2001).
In the level-set method, the moving interface is implicitly represented by
a smooth level-set function (Sethian and Smereka, 2003). The level-set method
has proved capable of handling problems in which the interface moving speed
is sensitive to the front curvature and normal direction. A signicant advan-
tage of the level-set method is that it is effective in 3-D simulation of the
conditions with large topological changes, such as bubble breaking and merg-
ing, dropletsurface collisions with evaporation. In this study, the level-set
technique (Sussman et al., 1994) is employed to describe the motion of 3-D
gasliquid interfaces. In the following section a description of this technique
is given.
A. LEVEL-SET METHOD
The level-set method, which was rst derived by Osher and Sethian (1988), is
a versatile method for capturing the motion of a free surface in 2-D or 3-D on a
xed Eulerian grid. While similar to the VOF method, the level-set method also
uses an indicator function to track the gasliquid interface on the Eulerian
grid. Instead of using the marker particles or points to describe the interface, a
smooth level-set function is dened in the ow eld (Sussman et al., 1994).
Consider a nonbody conformal Cartesian grid which is used to simulate the
ow with a deformable interface G, as shown in Fig. 1. The whole computa-
tional domain is separated by the interface into two regions: O

and O
+
. The
value of the level-set function is negative in the O

region and positive in the


O
+
region, while the interface G is simply described as the zero level set of
YANG GE AND LIANG-SHIH FAN 6
the level-set function f, i.e.,
G xjfx; t 0

(1)
where x represents the position vector and t the time. Taking fo0 as being
inside the interface G (in O

) and f40 as being outside the interface G (in O


+
),
the level-set function has the form:
fx; t
o0; x 2 O

0; x 2 G
40; x 2 O

8
>
<
>
:
(2)
The evolution of f in a ow eld is given by the so-called weak-form equation:
@f
@t
V rf 0 (3)
where V is the velocity of uid, and is given by
V
V

; x 2 O

; x 2 G
V

; x 2 O

8
>
<
>
:
(4)
For gasliquid bubble ow, V

and V
+
are the gas and liquid velocities, respec-
tively, and the zero-level set of f marks the bubble interface, which moves with
time. For gas-droplets ows, on the other hand, V

and V
+
represent the
+
-

+
-
FIG. 1. The level sets of distance function for a smooth interface over a Cartesian grid.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 7
velocity of the liquid and gas phases, respectively, and the zero-level set of f
denes the droplet surface (Ge and Fan, 2005).
To compute the motion of two immiscible and incompressible uids such as a
gasliquid bubble column and gas-droplets ow, the uid-velocity distributions
outside and inside the interface can be obtained by solving the incompressible
NavierStokes equation using level-set methods as given by Sussman et al.
(1994):
@r
@t
r rV 0 (5)
@rV
@t
r rVV rp r s rg F
s
(6)
where F
s
is the surface tension force which is calculated by (Brackbill et al.,
1992):
F
s
skfdfrf (7)
k(f) is the curvature which can be estimated as r (rf/|rf|). A smooth d
function is dened as (Sussman et al., 1998; Sussman and Fatemi, 1999):
d
b
f
dH
b
f
df

1
2
1 cospf=b=b; f

ob
0; otherwise
(
(8)
where H
b
(f) follows the Heaviside formulation (Sussman et al., 1998; Sussman
and Fatemi, 1999) given by
H
b
f
1 f4b
0 fob
1
2
1
f
b

1
p
sinpf=b otherwise
8
>
<
>
:
(9)
The surface-tension force F
s
in Eq. (7) is smoothed and distributed into the
thickness of the interface. In order to circumvent numerical instability, the uid
properties such as density and viscosity in the interface region are determined
with a continuous transition:
rf r

H
b
f (10)
mf m

H
b
f (11)
Since the values for r(f), m(f), and the surface-tension force could be distorted
if the variation of rf along the interface is very large, the thickness of the
interface needs to be maintained uniformly, i.e. rf

1 (Sussman et al., 1998).


In the algorithm developed, the general level set function f(x,t) is replaced by a
YANG GE AND LIANG-SHIH FAN 8
distance function d(x,t), whose value represents the signed normal distance from
x to the interface. d(x,t) would satisfy rd j j 1 and d 0 for xAG (Sussman
et al., 1998).
Even if the initial value of the level-set function f(x,0) is set to be the distance
function, the level set function f may not remain as a distance function at t40
when the advection equation, Eq. (3), is solved for f. Thus, a redistance scheme
is needed to enforce the condition of rf

1. An iterative procedure was


designed (Sussman et al., 1998) to reinitialize the level-set function at each time
step so that the level-set function remains as a distance function while main-
taining the zero level set of the level-set function. This is achieved by solving for
the steady-state solution of the equation (Sussman et al., 1994, 1998; Sussman
and Fatemi, 1999):
@d
@t
sinf1 rf

(12)
dx; 0 fx (13)
until
rd j j 1 OD
2
(14)
where the sin function is dened as
sinf
1; fo0
0; f 0
1; f40
8
>
<
>
:
(15)
In Eq. (12), t is an articial time that has the unit of distance. The solutions
for Eq. (12) are signed distances and only those within a thickness of 35 grid
sizes from the interface are of interest (Sussman et al., 1994, 1998; Sussman and
Fatemi, 1999). Equation (12) needs to be integrated for 35 time steps using a
time step Dt 0.5D.
B. IMMERSED BOUNDARY METHOD
The IBM was originally proposed by Peskin (1977) to model the blood ow
through heart valves. Since then, this method has been extensively modied and
extended to simulate various uid ows in a complex geometrical conguration
using a xed Cartesian mesh (Unverdi and Tryggvason, 1992a,b; Udaykumar,
et al., 1997; Ye et al., 1999; Fadlun et al., 2000; Lai and Peskin, 2000; Kim et al.,
2001). In the IBM, the presence of the solid object in a uid eld is represented
by a virtual-body force eld, which is applied on the computational grid in the
vicinity of the solidow interface through a Dirac delta function (Lai and
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 9
Peskin, 2000). Various schemes have been proposed to calculate the virtual force
density in the literature. Goldstein et al. (1993) developed a virtual boundary
formulation to simulate the startup ow over a cylinder. In their formation, the
virtual force eld is calculated in a feedback manner in order to satisfy the
boundary condition at the solid surface. Mohd-Yusof (1997) developed an
alternative direct forcing scheme to evaluate the virtual force based on the N-S
equation at discrete times. Fadlum et al. (2000) extended the direct forcing
scheme of Mohd-Yusof (1997) to a 3-D nite-difference method. Instead of
evaluating and applying the virtual force, the velocity at the rst grid point
outside the solid boundary is estimated through a linear interpolation of the
moving velocity of the boundary and the velocity at the second external grid
point. Conceptually, this velocity interpolation scheme is equivalent to applying
the momentum force inside the ow eld (Kim et al., 2001). This scheme is more
efcient in 3-D because it has no adjustable constant and has no extra restric-
tion on the scale of the time step, which is required in the feedback-forcing
scheme. Kim et al. (2001) simulated the ow over complex geometry in a nite-
volume approach with staggered meshes. The momentum force and mass source
were applied on the immersed boundary to satisfy the no-slip boundary con-
dition and the ow continuity.
The basic idea of the IBM is that the presence of the solid boundary (xed or
moving) in a uid can be represented by a virtual body force eld
~
F
p
applied on
the computational grid at the vicinity of solidow interface. Thus, the Navier
Stokes equation for this ow system in the Eulerian frame can be given by
@rV
@t
r rVV rp r s rg
~
F
p
(16)
It is noted that the virtual body force
~
F
p
depends not only on the unsteady
uid velocity, but also on the velocity and location of the particle surface, which
is also a function of time. There are several ways to specify this boundary force,
such as the feedback forcing scheme (Goldstein et al., 1993) and direct forcing
scheme (Fadlun et al., 2000). In 3-D simulation, the direct forcing scheme
can give higher stability and efciency of calculation. In this scheme, the disc-
retized momentum equation for the computational volume on the boundary is
given as
V
t1
V
t
DtRHS
t
F
t
p
(17)
where RHS refers to all the terms in the right-hand side of Eq. (16) except the
virtual body force
~
F
p
. The virtual body force F
t
p
is used to maintain the uid
velocity to be equal to the particle velocity at the particle surface (i.e., no-slip
boundary condition), which is
V
t1
V
p
t (18)
YANG GE AND LIANG-SHIH FAN 10
where V
p
is the particle velocity. Thus, the discrete virtual force can be dened as
F
t
p
V
p
V
t
=Dt RHS
t
(19)
Since the computational grids are generally not coincident with the location of
the particle surface, a velocity interpolation procedure needs to be carried out in
order to calculate the boundary force and apply this force to the control volumes
close to the immersed particle surface (Fadlun et al., 2000).
Other than the virtual momentum force
~
F
p
, a virtual mass source/sink should
also be applied to the particle surface to satisfy the continuity for the control
volume containing the particle surface or the particle (Kim et al., 2001). The
mass source can be calculated by
q
t

1
DV
X
i
a
i
~
V
t
i
~n
i
Ds
i
(20)
where DV is the volume of the computation cell (control volume) and Ds
i
the
surface area of surface i of this cell. For a 3-D case, i 1, 2, y, 6. ~n
i
is
the normal vector of each face of the cell.
~
V
t
i
the uid velocity at each face of the
cell. a
i
the ag to indicate whether the virtual body force is applied to face i of
the cell or not. a
i
1 when the force is applied, otherwise it is zero. Therefore,
the continuity equation of the incompressible uid can be written as (Kim et al.,
2001):
r
~
V q (21)
III. System1: Flow Dynamics of GasLiquidSolid Fluidized Beds
The ows in a gasliquidsolid uidized bed or a gasliquid bubble column
are represented by two regimes, the homogeneous and the heterogeneous. In the
homogenous regime, the coalescence of bubbles does not occur and there is little
variation of bubble sizes. However, this is not the case in the heterogeneous
regime. The ow structure in the heterogeneous regime is complex due to sub-
stantial coalescence and breakup of bubbles. Both the E-E and the E-L methods
have proven to be more effective in modeling the homogenous regime than the
heterogeneous regime of gasliquid ow. In the simulation of the heterogeneous
regime of gasliquid ows using either the E-E or the E-L method, the challenge
lies in the establishment of an accurate closure relationship for the interphase
momentum exchange. The interphase momentum exchange is induced through
the drag force that liquid exerts on the bubble surface, the virtual mass force due
to the bubble and liquid inertial motion, and the lift force caused by the shear
ows around the bubbles. In gasliquid bubble columns and gasliquidsolid
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 11
uidized systems, the interstitial forces under the bubble coalescence and
breakup conditions are not well established. A computational model based on
the level-set methods given below provides some information on the much
needed closure relationship of the interphase momentum exchange noted above.
A. NUMERICAL PROCEDURE FOR SOLVING THE GAS LIQUID INTERFACE
The level-set technique described in Section II.A is employed to capture the
motion of 3-D gasliquid interfaces. The numerical procedures for solving the
gasliquid interface include nding the solution for the time-dependent Eqs. (3),
(5), and (6). Given f
n
and V
n
dened at cell centers at one time instant t
n
, f
n+1
,
and V
n+1
can be solved over a time increment at a new time instant
t
n+1
t
n
+Dt following the procedures given below:
Solve Eqs. (5) and (6) to obtain the velocity distribution in the ow eld V
n+1
using the Arbitrary-Lagrangian-Eulerian (ALE) scheme (Kashiwa et al., 1994).
Solve Eq. (3) to obtain f
n+1
using the second-order TVD-Runge-Kutta
method presented as follows:

f
n1
f
n
Dtf
tn
(22)
f
n1
f
n

Dt
2

f
tn1
f
tn
(23)
where f
tn
V
n
Df
n
and the time steps are the same as that used in calculating
V
n+1
, which is determined by restrictions due to the CourantFriedrichsLevy
(CFL) condition, gravity, viscosity, and surface tension.
Solve Eq. (12) to perform the redistancing.
Although, in principle, Eq. (12) would not alter the location of the zero-level
set of f, in practice, with numerical computation it may not be true. A redis-
tance operation is needed to maintain the volume conservation. Therefore, Eq.
(12) is modied to (Sussman et al., 1998):
@d
@t
sinf1 rf

l
ij
f f Lf; d l
ij
f f (24)
where
l
ij

R
O
ij
H
0
fLf; d
R
O
ij
H
0
ff f
(25)
and
f f H
0
f rf

(26)
YANG GE AND LIANG-SHIH FAN 12
B. GOVERNING EQUATIONS FOR THE GAS LIQUID SOLID FLOW
The gasliquidsolid ow is characterized by a wide range of physical length
scales, including small to large eddies in the bubble wake, and size in the milli-
meter range for solid particles and in the millimeter/centimeter range for gas
bubbles. The accurate description of the gas bubble surface and bubbling ow
requires the use of ne grids, while the tracking of the motion of solid particles
needs the grid size to be much larger than the particle sizes.
For simulation of a gasliquidsolid uidized bed, the locally averaged
NavierStokes equations (Anderson and Jackson, 1967) are used to describe the
liquid phase ow outside the gas bubble, and the gas phase ow inside the gas
bubble. Due to the large grid size used, the liquid phase turbulence needs to be
considered. In this study, a modied coefcient that illustrates the effect of the
bubble-induced turbulence for a subgrid scale (SGS) stress model is employed.
The level-set method and the numerical procedures described in Sections II.A
and III. are used to simulate the motion and the topological variation of the gas
bubble. The locally averaged governing equations of Eqs. (5) and (6) for liquid
ow outside the bubble and gas ow inside the bubble are given as:
@r
@t
r rV 0 (27)
@rV
@t
r rVV rp r s r s
sg
rg F
D
F
s
(28)
e represents the void fraction of liquid or gas and satises:

p
1 (29)
where e
p
is the void fraction of solid particles. t
sg
the SGS stress term. It is
modeled by the Smagorinsky (1963) model written as
s
sg
ij
n
T
@V
i
@x
j

@V
j
@x
i

(30)
where n
T
is dened as
n
T
C
s
l
2
S j j (31)
for bulk ow, and
n
T
C
s
f yl
2
S j j (32)
for walls with a wall function f(y). C
s
is the Smagorinsky coefcient, l D, and
S j j

2S
ij
S
ij
p
(33)
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 13
The volumetric uidparticle interaction force F
D
in Eq. (28) is calculated from
the forces acting on the individual particles in a cell:
F
D

P
f
k
d
DO
ij
, (34)
where f
d
is the uidparticle interaction force for a single particle and DO the
cell volume.
C. MODELING THE MOTION AND COLLISION DYNAMICS OF SOLID PARTICLES IN
GAS LIQUID SOLID FLUIDIZATION
The motion of a particle in the ow eld can be described in the Lagrangian
coordinate with the origin placed at the center of the moving particle. There are
two modes of particle motion, translation and rotation. Interparticle collisions
result in both the translational and the rotational movement, while the uid
hydrodynamic forces cause particle translation. Assuming that the force acting
on a particle can be determined exclusively from its interaction with the sur-
rounding liquid and gas, the motion of a single particle without collision with
another particle can be described by Newtons second law as
dx
p
dt
V
p
(35)
m
p
dV
p
dt
m
p
g
p
6
d
3
p
rp r s r s
sg
f
d
f
am
f
s
(36)
where x
p
and V
p
are the particle position and particle velocity, respectively, and
d
p
the diameter of the particle. The ve terms on the right-hand side of Eq. (36)
represent, respectively, the gravity force, the uid stress gradient force, the total
drag force, the added mass force and the bubble-surface-tension-induced force.
The Saffman, the Magnus, and the Basset forces are ignored.
Note that the lubrication effect due to particle collisions in liquid is signi-
cant. The liquid layer dynamics pertaining to the lubrication effect was exami-
ned by Zenit and Hunt (1999). Zhang et al. (1999) used a Lattice-Boltzmann
(LB) simulation to account for a close-range particle collision effect and
developed a correction factor for the drag force for close-range collisions, or the
lubrication effect. Such a term has been incorporated in a 2-D simulation based
on the VOF method (Li et al., 1999). Equation (36) does not consider the
lubrication effect. Clearly, this is a crude assumption. However, in the three-
phase ow simulation, this study is intended to simulate only the dilute solids
suspension condition (e
p
0.423.4%) with the bubble ow time of less than 1 s
starting when bubbles are introduced to the solids suspension at a prescribed e
p
.
YANG GE AND LIANG-SHIH FAN 14
The particle collision effect under this simulation condition, therefore, would
be small.
Note that depending on the manner in which the drag force and the buoyancy
force are accounted for in the decomposition of the total uidparticle inter-
active force, different forms of the particle motion equation may result (Jackson,
2000). In Eq. (36), the total uidparticle interaction force is considered to be
decomposed into two parts: a drag force (f
d
) and a uid stress gradient force (see
Eq. (2.29) in Jackson, 2000)). The drag force can be related to that expressed by
the WenYu equation, f
WenYu
, by
f
d
f
WenYu
(37)
The Wen and Yu (1966) equation is given by
f
WenYu

1
8
pd
2
p
C
D

2
r V V
p

V V
p
(38)
where the effective drag coefcient C
D
is calculated by
C
D
C
D0

4:7
(39)
In Eq. (39), C
D0
is a function of the particle Reynolds number, Re
p

rd
p
jV V
p
j=m. For rigid spherical particles, the drag coefcient C
D0
can be
estimated by the following equations (Rowe and Henwood, 1961):
C
D0

24
Re
p
1 0:15Re
p

0:687
; Re
p
o1000
0:44; Re
p
! 1000
(
(40)
The added mass force accounts for the resistance of the uid mass that is moving
at the same acceleration as the particle. Neglecting the effect of the particle
concentration on the virtual-mass coefcient, for a spherical particle, the volume
of the added mass is equal to one-half of the particle volume, so that
f
am

1
12
pd
3
p
r
DV
Dt

DV
p
Dt

(41)
When particles approach the gasliquid interface, the surface-tension force acts
on the particles through the liquid lm. The bubble-surface-tension induced
force can be described by
f
s

p
6
d
3
p
sKfdfrf (42)
When the particle inertia overcomes the surface-tension-induced force, the par-
ticle will penetrate the bubbles. Recognizing that particle penetration may not
lead to bubble breakage, details of bubble instability due to particle collision are
given in Chen and Fan (1989a, b).
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 15
To simulate the particleparticle collision, the hard-sphere model, which is
based on the conservation law for linear momentum and angular momentum, is
used. Two empirical parameters, a restitution coefcient of 0.9 and a friction
coefcient of 0.3, are utilized in the simulation. In this study, collisions between
spherical particles are assumed to be binary and quasi-instantaneous. The equa-
tions, which follow those of molecular dynamic simulation, are used to locate
the minimum ight time of particles before any collision. Compared with the
soft-sphere particleparticle collision model, the hard-sphere model accounts for
the rotational particle motion in the collision dynamics calculation; thus, only
the translational motion equation is required to describe the uid induced par-
ticle motion. In addition, the hard-sphere model also permits larger time steps in
the calculation; therefore, the simulation of a sequence of collisions can be more
computationally effective. The details of this approach can be found in the
literature (Hoomans et al., 1996; Crowe et al., 1998).
D. RESULTS AND DISCUSSIONS
The computation performed in this study is based on the model equations
developed in this study as presented in Sections II.A, III.A, III.B, and III.C
These equations are incorporated into a 3-D hydrodynamic solver, CFDLIB,
developed by the Los Alamos National Laboratory (Kashiwa et al., 1994). In
what follows, simple cases including a single air bubble rising in water, and
bubble formation from a single nozzle in bubble columns are rst simulated. To
verify the accuracy of the model, experiments are also conducted for these cases
and the experimental results are compared with the simulation results. Simu-
lations are performed to account for the bubble-rise phenomena in liquidsolid
suspensions with single nozzles. Finally, the interactive behavior between bub-
bles and solid particles is examined. The bubble formation and rise from multi-
ple nozzles is simulated, and the limitation of the applicability of the models is
discussed.
1. Single Air Bubble Rising in Water
The simulation for a single air bubble rising in water (density: 0.998 kg/cm
3
;
viscosity: 0.01 Pa s; surface tension: 0.0728 N/m) is performed in a 4 4
8 cm
3
3-D column. A uniform grid size of 0.05 cm is used for three dimensions
which generates 80 80 160 ( 1.024 10
6
) grid points in the computational
domain. Initially, a spherical air bubble is positioned at rest in this domain with
its center located 0.5 cm above the bottom and the liquid is quiescent. The free-
slip boundary conditions are imposed on all six walls. Note that the dimension
of the computational domain is selected based on numerical experiments. It is
found that, under both free-slip and no-slip wall boundary conditions, when the
distance of the bubble interface to the wall is more than twice as large as the
YANG GE AND LIANG-SHIH FAN 16
bubble diameter, there is practically little effect of the wall boundary conditions
on the simulation results. Thus, with an exception of the cases involving several
bubbles which migrate to the near-wall region through the zigzag motion that
will be discussed in a later section, the simulation results obtained in this study
are not affected by these wall boundary conditions. The time for an air bubble
of 0.8 cm in diameter rising from the initial position to the outlet of the column
is about 0.4 s, which corresponds to a bubble rise velocity of 18.75 cm/s.
The simulation was conducted using a Cray SV1 supercomputer at the Ohio
Supercomputer Center (OSC). The CPU time to compute the entire process of
bubble rising is 4 h.
The simulation results for the positions and the shape changes of the 0.8 cm
air bubble rising in water are shown in Fig. 2. The value for the Eotvos number
(E
0
gDrd
e
2
/s) and the Morton number (M gm
4
Dr/r
2
s
3
) are 8.5 and
2.5 10
7
, respectively. The time increment between two bubble images in
Fig. 2 is 0.05 s. As seen in the gure, the bubble shape undergoes continuous
changes from a sphere initially to an oblate ellipsoidal cap, and uctuates
between an oblate ellipsoidal cap and a spherical cap. The rectilinear motion
of a bubble in water exhibited in the gure, occurs for the rst several frac-
tions of a second of a single bubble rising (with symmetric wakes) in a quiescent
liquid even though the bubble Reynolds number (Re
b
) is in the wake shedding
regime (Re
b
4400). The computed results obtained in this study capture such
FIG. 2. Simulated positions and shape variations of a rising bubble in a water column. Initial
bubble diameter 0.8 cm and time increment 0.05 s.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 17
phenomena well. The liquid-eld disturbance would eventually induce an asym-
metric wake, yielding wake-shedding phenomena of the rising bubble.
The mesh-renement studies are conducted to examine the mesh effect on the
computation results. Simulations of a 0.8 cm air bubble rising in a 4 4 8 cm
3
water column (as shown in Fig. 2) are repeated at three different mesh reso-
lutions: from a lower resolution of 40 40 80 grid points with a grid size of
0.1 cm to a higher resolution of 100 100 200 grid points with a grid size of
0.04 cm. The simulated variations of bubble rise velocity and bubble aspect ratio
(height/width) with time are shown in Figs. 3a and b, respectively.
Fig. 3a indicates that the bubble-rise velocity measured based on the dis-
placement of the top surface of the bubble (U
bt
) quickly increases and ap-
proaches the terminal bubble rise velocity in 0.02 s. The small uctuation of U
bt
is caused by numerical instability. The bubble-rise velocity measured based on
the displacement of the bottom surface of the bubble (U
bb
) uctuates signi-
cantly with time initially and converges to U
bt
after 0.25 s. The overshooting of
U
bb
can reach 4550 cm/s in Fig. 3a. The uctuation of U
bb
reects the unsteady
oscillation of the bubble due to the wake ow and shedding at the base of
the bubble. Although the relative deviation between the simulation results of the
40 40 80 mesh and 100 100 200 mesh is notable, the deviation is insig-
nicant between the results of the 80 80 160 mesh and those of the
100 100 200 mesh. The agreement with experiments at all resolutions is
generally reasonable, although the simulated terminal bubble rise velocities
($20 cm/s) are slightly lower than the experimental results (21$25 cm/s). A
lower bubble-rise velocity obtained from the simulation is expected due to the
no-slip condition imposed at the gasliquid interface, and the nite thickness for
the gasliquid interface employed in the computational scheme.
The aspect ratio shown in Fig. 3b describes the change of the bubble shape
with time during the bubble rising. The simulation and the experimental results
generally agree well, as shown in the gure. It can also be seen that the simu-
lation results are only sensitive to the mesh size of 40 40 80 mesh and the
deviation between the results of 80 80 160 mesh and 100 100 200 mesh is
small. Thus, Figs. 3a and b indicate that a reasonable accuracy can be reached
in this bubble-rise simulation with a 80 80 160 mesh (grid size of 0.05 cm).
The simulation results on bubble velocities, bubble shapes, and their uctu-
ation shown in Fig. 3 are consistent with the existing correlations (Fan and
Tsuchiya, 1990) and experimental results obtained in this study. Bubble rise
experiments were conducted in a 4 cm 4 cm Plexiglas bubble column under
the same operating conditions as those of the simulations. Air and tap water
were used as the gas and liquid phases, respectively. Gas is introduced through a
6 mm nozzle. Note that water contamination would alter the bubble-rise prop-
erties in the surface tension dominated regime. In ambient conditions, this re-
gime covers the equivalent bubble diameters from 0.8 to 4 mm (Fan and
Tsuchiya, 1990). All the airwater experiments and simulations of this study are
carried out under the condition where most equivalent bubble diameters exceed
YANG GE AND LIANG-SHIH FAN 18
4 mm.These ow conditions correspond to the bubble inertial regime, and thus,
the extent of water contamination plays a negligible role in the determination of
the bubble-rise properties.
The thickness of the gasliquid interface is set as 3D based on the parameters
used in the case of Sussman (1998), with the same density-ratio on the interface
and similar Reynolds number. An interface thickness of 5D is also examined in
the simulation and no signicant improvement is observed. The accurate pre-
diction of the bubble shape (shown in Figs. 3 and 4) can be attributed, in part,
to the manner in which the surface-tension force is treated as a body force in
the computation scheme. Specically, since the surface-tension force acting on a
solid particle is considered only when a solid particle crosses the gasliquid
interface and the solid particle is considered as a point, the accuracy of the
calculation of this force can be expected if the surface tension is interpreted as a
body force acting on each grid node near the interface.
2. Bubble Formation from an Orice
The air-bubble formation from a single orice in water is simulated. The
computational domain is 2 2 4 cm
3
. A uniform grid size of 0.025 cm and
80 80 160 grid points are used to obtain convergent solutions for the bubble
formation process. This mesh-size effect is examined by comparing the simu-
lation results on the bubble-formation processes with experimental measure-
ments. As shown in Fig. 4, decreasing the mesh size or increasing the mesh
resolution from 40 40 80 to 80 80 160 improves the accuracy of the
prediction results on the bubble shape. Further increase in mesh resolution does
not practically change the simulation results.
Simulations are then performed for gas bubbles emerging from a single nozzle
with 0.4 cm I.D. at an average nozzle velocity of 10 cm/s. The experimental
measurements of inlet gas injection velocity in the nozzle using an FMA3306 gas
ow meter reveals an inlet velocity uctuation of 315% of the mean inlet
velocity. A uctuation of 10% is imposed on the gas velocity for the nozzle to
represent the uctuating nature of the inlet gas velocities. The initial velocity of
the liquid is set as zero. An inow condition and an outow condition are
assumed for the bottom wall and the top walls, respectively, with the free-slip
boundary condition for the side walls.
Fig. 5 shows the simulated air-bubble formation and rising behavior in water.
For the rst three bubbles, the formation process is characterized by three
distinct stages of expansion, detachment, and deformation. In comparison with
the bubble formation in the airhydrocarbon uid (Paratherm) system, the co-
alescence of the rst two bubbles occurs much earlier in the airwater system.
Note that the physical properties of the Paratherm are r
l
870 kg/m
3
,
m
l
0.032 Pa s, and s 0.029 N/m at 25 1C and 0.1 MPa. This is due to the
fact that, compared to that in the airParatherm system, the rst bubble in the
airwater system is much larger in size and hence higher in rise velocity leading
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 19
to a longer time for its coalescence with the second bubble. Beginning with the
third bubble, the formation and rising behavior of air bubbles in water shows
strongly asymmetric behavior. As is evident from Fig. 5, the bubble rises in a
spiral path or a zigzag path.
FIG. 3. (a) Comparison of the simulation results and experimental results of the bubble rise
velocity. (b) Comparison of the simulation and experimental results of the bubble aspect ratio.
YANG GE AND LIANG-SHIH FAN 20
In order to verify the simulation results, experiments on bubble behavior
in bubble columns are carried out under conditions similar to the simulations.
A 3-D rectangular bubble column with the dimension of 8 8 20 cm
3
is used
for the experiments. Four nozzles with 0.4 cm I.D. and a displacement of 2.4 cm
are designed in the experiments. For single-nozzle experiments, air is injected
into the liquid bed through one of the orices while the others are shut off. The
outlet air velocity from the nozzle is approximated using the measured bubbling
FIG. 4. Comparison of the experimental measurement and the simulation results with different
resolutions of air-bubble formation in water.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 21
FIG. 5. Simulation results of air-bubble formation from a single nozzle in water. Nozzle size 0.4 cm I.D. and nozzle gas velocity 10 cm/s.
Y
A
N
G
G
E
A
N
D
L
I
A
N
G
-
S
H
I
H
F
A
N
2
2
frequency and the initial bubble size. A high-speed video camera (240 frames/s)
is used to obtain the images of bubbles emerging from the orice in the liquid.
A common dimensionless number used to characterize the bubble formation
from orices through a gas chamber is the capacitance number dened as:
N
c
4V
c
gr
l
/pD
0
2
P
s
. For the bubble-formation system with inlet gas provided by
nozzle tubes connected to an air compressor, the volume of the gas chamber is
negligible, and thus, the dimensionless capacitance number is close to zero. The
gas-ow rate through the nozzle would be near constant. For bubble formation
under the constant ow rate condition, an increasing ow rate signicantly
increases the frequency of bubble formation. The initial bubble size also in-
creases with an increase in the ow rate. Experimental results are shown in
Fig. 6. Three different nozzle-inlet velocities are used in the airwater experi-
ments. It is clearly seen that at all velocities used for nozzle air injection, bubbles
rise in a zigzag path and a spiral motion of the bubbles prevails in airwater
experiments. The simulation results on bubble formation and rise behavior
conducted in this study closely resemble the experimental results.
FIG. 6. Experimental results on air-bubble formation and bubble rising in water. Nozzle size
0.4 cm I.D. and nozzle gas velocity (a) $6.0 cm/s; (b) $10.0 cm/s; (c) $14 cm/s.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 23
3. Gas Liquid Solid Fluidization
As noted earlier, to simulate the bubble motion in a gasliquid bubble column
accurately, ne grid sizes, 0.025 cm for airwater and 0.05 cm for airParatherm
system, should be used in the computation. This ne-grid computation yields
essentially the results of DNS. These grid sizes are smaller than the size of the
solid particle usually employed for the three-phase uidized bed operation. For
the particle size of 0.08 cm used in the present simulation of a three-phase
uidized bed, the computational grid size is required to be no less than 0.2 cm
in order to track both the bubble ow and the particle motion. Note that
the system simulated in this study is a dilute liquidsolid bed with a minimum of
particleparticle collisions and uniform particle distribution. Although a grid
size of D410 D
p
as generally used in the Lagrangian simulation of uidparticle
ows is preferable, the grid size used under the current simulation of three-phase
ows is acceptable. There were no numerical stability or convergence problems
encountered in the computation. For simulation of the bubble formation in a
gasliquid bubble column, a coarse grid size of 0.2 cm in a 4 4 8 cm
3
domain
with 21 21 41 grid points is used in this study. However, due to this large
grid size used, without any turbulence model, the simulation cannot accurately
track the discrete bubble-formation process. Specically, simulation without
consideration of the turbulent effects, the bubbles with distorted wake structure
are seen to be connected like a jet above a nozzle. An SGS stress model is thus
employed and incorporated into the code for subsequent simulation. The sim-
ulation of the gasliquid bubble column system indicates that experimental
results on a bubble formation in an airParatherm medium can be well de-
scribed when C
s
values are in a range of 1.0 to 1.2 with a grid size of 0.2 cm.Note
that the values for the Smagorinsky coefcient for single phase ow are 0.10.2.
The results are shown in Fig. 7(a).
Subsequently, simulations are performed for the airParathermsolid uidi-
zed bed system with solid particles of 0.08 cm in diameter and 0.896 g/cm
3
in
density. The solid particle density is very close to the liquid density (0.868 g/
cm
3
). The boundary condition for the gas phase is inow and outow for the
bottom and the top walls, respectively. Particles are initially distributed in the
liquid medium in which no ows for the liquid and particles are allowed through
the bottom and top walls. Free slip boundary conditions are imposed on the
four side walls. Specic simulation conditions for the particles are given as
follows: Case (b) 2,000 particles randomly placed in a 4 4 8 cm
3
column;
Case (c) 8,000 particles randomly placed in a 4 4 8 cm
3
column; and Case (d)
8,000 particles randomly placed in the lower half of the 4 4 8 cm
3
column.
The solids volume fractions are 0.42, 1.68, and 3.35%, respectively for Cases
(b), (c), and (d).
The bubble-formation process at different solids concentrations is shown in
Figs. 7(b)(d) and is compared with that without particles as shown in Fig. 7(a).
For the rst 0.3 s, little change is observed in the bubble-formation process
YANG GE AND LIANG-SHIH FAN 24
for the three solids concentrations used in this simulation. After 0.4 s, how-
ever, signicant changes can be found for the cases with high solids concen-
trations. This can be seen from the rst bubble in each case. When the solid
concentration is low or no solids are present, the rst bubble grows on the
FIG. 7. Simulation results of bubble formation and rising in Paratherm NF heat-transfer uid
with and without particles. Nozzle size 0.4 cm I.D., liquid velocity 0 cm/s, gas velocity 10 cm/s, and
particle density 0.896 g/cm
3
. (a) No particle; (b) 2000 particles; (c) 8000 particles; (d) 8000 particles.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 25
orice and connects to the second bubble. For the high solids concentra-
tion cases, the rst bubble is not well connected to the second bubble. This is
particularly true for Case (d) when the bubble rises into the solids-free region or
freeboard region of the bed. The solid particle entrainment is clearly observed in
Case (d).
FIG. 7 (Continued)
YANG GE AND LIANG-SHIH FAN 26
IV. System 2: Deformation Dynamics of Liquid Droplet in
Collision with a Particle with Film-Boiling Evaporation
The phenomena of evaporative liquid droplets impacting onto solid objects at
high temperatures are of relevance to many engineering problems, such as
sprinkler systems in the iron making or metal-casting processes, ink-jet spray-
painting, impingement of oil droplets on turbine engines, meteorology, and
spray coating of substrates. An evaporative liquid jet in gassolid ow systems
is also of interest to current technology applications in chemical, petroleum, and
materials processing industries, such as FCC, polyethylene synthesis (Kunii and
Levenspiel, 1991; Fan et al., 2001) and microelectronic materials manufactur-
ing. In FCC riser reactors, for example, gas oil at a low temperature is injected
into the riser from feed nozzles located at the bottom of the riser and the mist
droplets formed from the spray contact with high-temperature uidized catalyst
particles. The vaporized oil then carries the catalyst particle up through the
riser. In the feed nozzle region, the size of the droplet can be comparable or
signicantly smaller (or larger) than the size of particle. The droplet can always
have a different momentum, thus the collision between the catalytic particles
and oil droplet may have various modes. Fig. 8 shows some of the collision
modes existing in a feed-nozzle region (Zhu et al., 2000). Smaller droplets may
rebound from the surface of larger particles upon impact, and smaller particles
Large
Droplet
Small
Droplet
FIG. 8. Various modes of dropletparticle collisions.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 27
may penetrate through or penetrate but retain inside the larger droplets. Larger
droplets may break into smaller drops during the impact and/or remain at-
tached to the particle surface after the collision, which may intensify the particle
aggregation. Clearly, understanding the droplet and particle collision mechanics
are crucial to an accurate account of the momentum and heat transfer between
the droplet and solid object, which is important for prediction of hydrocarbon
product distributions in light of catalytic and the thermal-cracking reactions in
the riser. It is also relevant to the design of feed nozzles that provide desired
droplet properties for optimum droplet contact with catalyst particles in the
reactor.
In most of the applications, the solid objects (e.g., the catalyst particles in
FCC reactor) are always under high temperature, and the droplet impact proc-
esses are accompanied with intensive evaporation. The nature of the collision
of the droplet with the superheated objects exhibits a great diversity in hydro-
dynamic and thermodynamic properties, such as droplet splash and rebound,
wetting or nonwetting contact, nucleate boiling or lm boiling, and Marangoni
effect. Further, the droplet shape, the contact area and the cooling effectiveness
during the impact not only depend on such hydrodynamic forces as the inertia,
pressure, surface tension, and viscous forces but also on the degrees of the
surface superheating and the droplet subcooling (Inada et al., 1985). As
the solid temperature rises to superheated conditions, the characteristics of
liquidsolid contact signicantly change and the evaporation rate affects the
droplet hydrodynamics. Under this condition, the nonwetting contact may
develop during the collision, and the evaporation is under the lm-boiling re-
gime, or so called Leidenfrost regime (Gottfried et al., 1966). In the Leidenfrost
regime, the vapor pressure generated from the droplet evaporation prevents
the direct contact of the droplet with the solid objects. The heat transfer from
the hot objects to the droplet is also hindered due to the resistance of the vapor
layer existing between the droplet and the solid surface. In this work, a 3-D
numerical model is developed and the simulation is conducted to account for
the behavior of the dropletparticle collision in the Leidenfrost regime.
Experimental and numerical studies of droplets impacting onto a at surface
of varied temperatures have been extensively reported in the literature. The
effects of the initial droplet temperature on lm-boiling impact are signicant
(Inada et al., 1985; Harvie and Fletcher, 2001b). Depending on the initial
droplet temperature, there are two types of droplet impact: saturated impact
and subcooled impact. The saturated impact involves the initial temperature
at the boiling point of the liquid or saturation temperature of the liquid. The
subcooled impacts, on the other hand, involve the droplet initial impact
temperature below the liquid-saturation temperature. The experimental results
for these two types of impact are briey described below. The modeling and
numerical approaches used for the droplet impingement onto isothermal or
heated at wall are also given.
YANG GE AND LIANG-SHIH FAN 28
A. SIMULATION OF SATURATED DROPLET IMPACT ON FLAT SURFACE IN THE
LEIDENFROST REGIME
Wachters and Westerling (1966) rst presented a classication of the dynamic
regimes of the impact based on their experiments in which water drops with a
diameter of 2.3 mm impact on a polished gold surface at temperatures between
200 1C and 400 1C. They studied the saturated impact of water droplets and
found that for the impact with Weo30, where We 2r
l
V
2
R/s, the surface
tension of the droplet dominates the impact process, and the droplet recoils and
rebounds from the surface without disintegration. At 30oWeo80, the droplet
undergoes a similar spreading and recoiling process as that for Weo30. In the
rebounding process, the droplet may disintegrate into several smaller droplets
(secondary droplets) and the shape of the droplet may then become unstable.
For the impact with We480, the impact inertial force (or kinetic energy) is so
large that splashing occurs during the early stage of the impact, while the
droplet breaks up into a number of small droplets. Based on the measured heat
ux on the solid surface, Wachters and Westerling (1966) also estimated the
relative volume decrease of the droplet during the impact. It was found that,
when the solid temperature is higher than 200 1C, the averaged evaporation rate
of the droplet decreases with an increase in the surface temperature. At the
nonwetting condition when the surface temperature reaches 400 1C, the volume
(mass) change of the droplet due to the evaporation during the impact is slight
(0.20.3%) for a wide range of We.
Groendes and Mesler (1982) studied the saturated lm boiling impacts of a
4.7 mm water droplet on a quartz surface of 460 1C. The uctuation of the
surface temperature was detected using a fast-response thermometer. The maxi-
mal temperature drop of the solid surface during a droplet impact was reported
to be about 20 1C. Considering the lower thermal diffusivity of quartz, this
temperature drop implies a low heat-transfer rate on the surface. Biance et al.
(2003) studied the steady-state evaporation of the water droplet on a super-
heated surface and found that for the nonwetting contact condition, the droplet
size cannot exceed the capillary length.
Ge and Fan (2005) developed a 3-D numerical model based on the level-set
method and nite-volume technique to simulate the saturated droplet impact on
a superheated at surface. A 2-D vapor-ow model was coupled with the heat-
transfer model to account for the vapor-ow dynamics caused by the Leidenfrost
evaporation. The droplet is assumed to be spherical before the collision and the
liquid is assumed to be incompressible.
1. Hydrodynamic Model and Numerical Solution
In the level-set method, the free surface of the droplet is taken as the zero in
the level-set function f~x; t as given in Eq. (2). The motion of the interface is
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 29
traced by solving the HamiltonJacobi-type convection equation, given as Eq.
(3), in the computational domain. The mass loss of the droplet due to evapo-
ration during the impact process is neglected in surface-tracking based on the
experimental results of Wachters and Westerling (1966). With the level-set
method, the equation of motion of the uid follows the NavierStokes equation
as given by Eqs. (5) and (6). The density and viscosity are dened by Eqs. (10)
and (11).
The computational code used in solving the hydrodynamic equation is de-
veloped based on the CFDLIB, a nite-volume hydro-code using a common
data structure and a common numerical method (Kashiwa et al., 1994). An
explicit time-marching, cell-centered Implicit Continuous-uid Eulerian (ICE)
numerical technique is employed to solve the governing equations (Amsden
and Harlow, 1968). The computation cycle is split to two distinct phases: a
Lagrangian phase and a remapping phase, in which the Arbitrary Lagrangian
Eulerian (ALE) technique is applied to support the arbitrary mesh motion with
uid ow.
Let f
n
f~x; t
n
and
~
V
n

~
V~x; t
n
be the cell-centered level-set function and
velocity at time t
n
, respectively. The numerical procedures to solve the velocity
eld
~
V
n1
, and the level-set function f
n+1
at t
n+1
t
n
+Dt can be described
below:
(1) Compute the velocity eld
~
V
n1
by solving the governing equation, Eqs.
(56), using the cell-centered ICE technique and ALE technique (Kashiwa
et al., 1994).
(2) Solve the convection equation of f~x; t (Eq. (3)) to obtain the

f
n1
. The
high order (3rd order) essentially non-oscillatory (ENO) upwind scheme
(Sussman et al., 1994) is used to calculate the convective term
~
V
G
rf based
on the updated velocity eld
~
V
n1
. The time advancement is accomplished
using the second-order total variation diminishing (TVD) Runge-Kutta
method (Chen and Fan, 2004).
(3) Perform the redistance procedure to obtain the f
n+1
using

f
n1
as the initial
value. The detail of the redistance computation is given by Sussman et al.
(1998).
(4) Calculate the density and the viscosity of the eld using Eqs. (10)(11) with
the updated level-set function f
n+1
.
The time steps (Dt) for calculating the
~
V
n1
and the f
n+1
are the same, which
is determined by the CFL condition and under constraints of the viscous and
surface tension (Sussman et al., 1994).
Considering a surface temperature which is higher than the Leidenfrost tem-
perature of the liquid in this study, it is assumed that there exists a microscale
vapor layer which prevents a direct contact of the droplet and the surface.
Similar to Fujimoto and Hatta (1996), the no-slip boundary condition is adopted
at the solid surface during the droplet-spreading process and the free-slip
YANG GE AND LIANG-SHIH FAN 30
condition is applied for the recoiling and rebounding periods. The velocity at
the grid point inside the solid surface is solved together with whole domain but is
reset according to the relative boundary condition (Ge and Fan, 2005).
2. Vapor-Flow Model
As the thickness of the vapor layer (520 mm) is several orders of magnitude
smaller than the macroscale of the ow eld (i.e., the diameter of the droplet), it
would be impractical to use the same computation mesh for both macroow
and vapor-layer ow (Harvie and Fletcher, 2001a). Thus, a 2-D model is de-
veloped to simulate the dynamics of the vapor ow between the droplet and the
surface. For the lm-boiling impact problem, the vapor-layer model would
allow determination of the evaporation-induced pressure in the vapor layer
without neglecting the inertial force of the vapor ow. In the symmetrical co-
ordinates (x,l) shown in Fig. 9, assuming that the gas in the vapor layer is only
saturated vapor and neglecting the temporal term, the continuity and momen-
tum equations for incompressible vapor ows with gravitation terms neglected
are given by
@u
x
@x

u
x
x

@u
l
@l
0 (43)
u
x
@u
x
@x
u
l
@u
x
@l

@
@x
P
r

n
@
2
u
x
@l
2

@
2
u
l
@x@l

(44)
u
x
@u
l
@x
u
l
@u
l
@l

@
@l
P
r

n
@
2
u
l
@x
2

@
2
u
x
@x@l

1
x
@u
l
@x

1
x
@u
x
@l

(45)
where u
x
, u
l
are the vapor-ow velocities in x and l direction, respectively, n is
the kinematical viscosity of the vapor. To determine the relative signicance of

droplet
Vapor flow
O
Solid surface
FIG. 9. Coordinates for the vapor-layer model.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 31
each term in these motion equations, an order of magnitude analysis is made by
considering the following dimensionless groups:

x
x
R
; Z
l
d
; u
x

u
x
U
x
; u
l

u
l
U
l
; p
p
rU
2
x
;

t
tU
x
R
; Re
d

dxu
ld
n
(46)
where R is the droplet radius; d the vapor-layer thickness; Re
d
the local evapo-
ration Reynolds number; u
ld
(x) the local vapor velocity; and U
x
, U
l
the velocity
scalars in x, l directions, respectively. Two assumptions can be made in ac-
counting for the collision process: (a) The vapor-layer thickness is much smaller
than the radius of the droplet; (b) The velocity for the vapor ow is much larger
than the rates of variation of the vapor-layer thickness and breadth. Based on
these assumptions and the order of magnitude analysis, the x momentum equa-
tion can be simplied to:
u
x
@u
x
@x
u
l
@u
x
@l

@
@x
P
r

n
@
2
u
x
@l
2
(47)
The boundary conditions are:
l 0; u
x
x; 0 u
l
x; 0 0
l d; u
x
x; d u
l
x; u
l
x; d u
ld
x
x 0;
@
@x
0; x x
b
; p p
b
(48)
where p
b
is the pressure of the ambient gas at the outside edge of the vapor layer.
In the impact process that involves large temperature differences (DT)
between the surface and the droplet, such as the ones considered in this study
(e.g., DT300500 1C), the value for Re
d
is about 0.51.0. Thus, the inertial
force of the vapor ow would be of the same order of magnitude as the viscous
force, and cannot be neglected in Eq. (47) for the vapor-ow model.
To solve Eq. (47), a variable transformation is considered:
u
l
x; Z Zu
ld
x u
x
x; Z OxFZ (49)
u
ld
(x) can be calculated through the energy-balance equation at the
vapordroplet interface. O(x) and F(Z) are single-variable functions.
With this transformation, the solution for the x momentum can be converted
into that of an ordinary differential equation (ODE) of O(x):
F
00
Z Re
d
ZF
0
Z Re
d
FZ jx (50)
jx
d
2
nOx
@
@x
P=r (51)
YANG GE AND LIANG-SHIH FAN 32
The general solution of the Eq. (50) can be obtained in power series form.
Under the condition that Re
d
O1, F(Z) can be approximated by only in-
cluding the rst three terms in the power series with good accuracy:
FZ F
d
Z jxZ
1
2

Re
d
24

jx
Z
2
2

Re
d
24
Z
4

(52)
The averaged vapor-ow velocity is given by
u
x
x O
Z
1
0
F
d
dZ
1
2
u
ld
x
d
2
12g
1
3
20
Re
d

@
@x
p
r

(53)
The vapor-continuity equation can be expressed by
u
x
x
1
xdx
Z
x
0
x
0
u
ld
x
0
dx
0
(54)
The pressure distribution in the vapor layer can be obtained by solving Eqs.
(53) and (54) using a piecewise integration method (Ge and Fan, 2005). In this
procedure, the thickness of the vapor layer d(x) is obtained from the level-set
function. The u
ld
(x) is calculated by
u
ld
x _ m=r
v

@dx
@t
(55)
where the local evaporation rate m is dened by the heat-transfer model. The
vapor-pressure force simulated by this model is applied as an interfacial force to
the droplet bottom surface.
3. Heat-Transfer Model
Heat transfer occurs not only within the solid surface, droplet and vapor
phases, but also at the liquidsolid and solidvapor interface. Thus, the energy-
balance equations for all phases and interfaces are solved to determine the heat-
transfer rate and evaporation rate.
Inside the solid surface, the heat-conduction equation in 3-D coordinates is
@T
s
@t
a
s
@
@x
@T
s
@x

@
@y
@T
s
@y

@
@z
@T
s
@z

(56)
where T
s
(x,y,z) is the solid temperature and a
s
the thermal diffusivity of solid.
The heat transfer within the droplet is described by the following thermal-
energy transport equation with neglecting viscous dissipation:
@T
d
@t
u
@T
d
@x
v
@T
d
@y
w
@T
d
@z
a
d
@
2
T
d
@x
2

@
2
T
d
@y
2

@
2
T
d
@z
2

(57)
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 33
Using the same assumptions that were made in the vapor-layer model, the
energy-conservation equation for the incompressible 2-D vapor phase can be
simplied to a 1-D equation in boundary layer coordinates:
@
2
T
v
@Z
2
0 (58)
The radiative heat transfer across the vapor layer is neglected under the
condition that the solid temperature is lower than 700 1C (Harvie and Fletcher,
2001a,b). On the liquidvapor interface, the energy-balance equation is
k
v
T
ss
T
ds
d
_ mL
c
(59)
where k
v
is the thermal conductivity of the vapor; T
ss
and T
ds
are the temper-
atures of the solid surface and the droplet surface. The thermal boundary con-
dition at the solidvapor interface is
k
v
T
ss
T
ds
d
k
s
@T
s
@Z
(60)
where the heat ux in the Z direction is assumed to be much larger that that in
the x direction.
The numerical method used for solving the heat-transfer equation is similar to
that for solving the momentum equation, which is a nite-volume, ALE method
(Kashiwa et al., 1994).
4. Results and Discussion
To validate the model developed in the present study, the simulations are
rst conducted and compared with the experimental results of Wachters and
Westerling (1966). In their experiments, water droplets impact in the normal
direction onto a hot polished gold surface with an initial temperature of 400 1C.
Different impact velocities were applied in the experiment to test the effect of
the We number on the hydrodynamics of the impact. The simulation of this
study is conducted for cases with different Weber numbers, which represent
distinct dynamic regimes.
The simulation shown in Fig. 10 is an impact of a saturated water droplet of
2.3 mm in diameter onto a surface of 4001C with an impact velocity of 65 cm/s,
corresponding to a Weber number of 15. This simulation and all others pre-
sented in this study are conducted on uniform meshes (Dx Dy Dz D). The
mesh resolution of the simulation shown in Fig. 10 was 0.08 mm in grid size,
although different resolutions are also tested and the results are compared in
Figs. 11 and 12. The average time-step in this case is around 5 ms. It takes 4000
iterations to simulate a real time of 20 ms of the impact process. The simulation
YANG GE AND LIANG-SHIH FAN 34
FIG. 10. Water droplet impacts on a at surface. The initial droplet diameter is 2.3 mm and the
surface temperature is 400 1C. We 15.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 35
code is run on the cray-SV1 supercomputer at the OSC. The computing time of
this case is about 12 h.
Comparing the 3-D images simulated and the experimental photographs in
Fig. 10, it can be seen that the droplet shapes are well reproduced by the present
model. During the rst 3.5 ms of the impact (frames 13), a liquid lm with
attened disc shape is formed immediately after the impact. The inertial force
drives the liquid to continue spreading on the solid surface, while the surface
tension and the viscous forces resist the spreading of the liquid lm. As a result,
the droplet spreading speed decreases and the uid mass starts to accumulate at
FIG. 11. Simulated 3-D views of the impact for We 15 as a function of the mesh resolution.
YANG GE AND LIANG-SHIH FAN 36
the leading edge of the liquid lm (2.53.2 ms). After the droplet spreads to the
maximum extent, the liquid lm starts to shrink back to its center (frames 4
and 5) due to the surface-tension force at the edge of the lm. At 3.55ms (frame
4), the simulated droplet shows a concave structure with a void in the center,
which is also shown in the experimental photograph of 3.85 ms. This structure,
also called a ring structure, has also been widely reported in the literature. In the
cross-sectional images at 3.55 ms, the velocity eld shows that the inward ow
rst starts from the outer edge of the liquid lm, which conrms that recoiling
ow is driven by the surface tension. After 4.4 ms, the droplet continues to recoil
and forms an upward ow in the center of the droplet (frames 5 and 6), and this
leads to a bouncing of the droplet up from the surface (frame 7). The peanut-
shape droplet (also called dumb-bell shape by Harvie and Fletcher (2001b) shown
in the experimental photograph at 14.56 ms is reproduced in the simulation.
The impact process shown in Fig. 10 is also simulated at six different grid
resolutions, i.e., 0.150, 0.120, 0.100, 0.075, 0.060, and 0.050 mm in mesh sizes,
with the corresponding cells per droplet radius (CPR) of 7.6, 9.6, 11.5, 15.3,
19.0, and 23.0, respectively. The comparison of the 3-D images among three
resolutions is shown in Fig. 11. The corresponding CPRs of these resolutions
are 9.6, 11.5, and 15.3. It can be found that the simulated droplet shapes are
similar at all three resolutions during the spreading process and even the early
stage of the recoiling process (26 ms). The deviation appears in the late stage of
0 2 4 6 8 101 2 141 61 8
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
2.0
0.150mm grid size
0.120mm grid size
0.100mm grid size
0.075mm grid size
0.060mm grid size
0.050mm grid size
Experiment of Wachters and Westerling
6
R
/
R
0
Time after impact(ms)
FIG. 12. Spread factor of the droplet verse time for the impact condition given in Fig. 10 at
different mesh resolutions.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 37
the recoiling process (8 ms), while the droplet generated on coarser mesh
(0.12 mm) tends to be more uniform in structure and less elongated in the
vertical direction. The difference in the droplet shape for 0.1 mm mesh and
0.075 mm mesh is relatively small.
Fig. 12 shows the spread factors simulated on meshes with different reso-
lutions along with the measurement value of Wachters and Westerling (1966).
The spread factor is dened as the radius of the droplet on the solid surface
divided by the initial radius of the droplet. Although the convergence is not
perfect, the agreement between the experiment and the simulations is relatively
good for all resolutions. Consistent with the results of Fig. 11, the effect of
the mesh resolution on spread factor becomes notable after 8 ms since the
moment of impact, and the coarser resolution tends to yield a slower rebound-
ing process.
The simulations were also performed under same conditions as the case of
Fig. 10 but for higher impact velocities. The simulated-droplet dynamics and
heat-transfer rate at the solid surface at different impact velocities are given in
Ge and Fan (2005).
B. SIMULATION OF SUBCOOLED DROPLET IMPACT ON FLAT SURFACE IN
LEIDENFROST REGIME
Subcooled impacts, in which the initial temperature of the droplet is below
the liquid saturation temperature, are of primary interest in experiments since
the condition of the spray liquids is often of the ambient temperature in prac-
tical applications (Inada et al., 1985; Chandra and Avedisian, 1991; Chen and
Hsu, 1995). Chandra and Avedisian (1991) studied the collision dynamics of a
24 1C n-heptane droplet impacting on a metallic surface with a Weber number
equal to 43. The transition from the nucleate boiling to the lm boiling was
identied when the surface temperature rises from the boiling point (170 1C) to
above the Leidenfrost temperature (200 1C) of n-heptane. They found that un-
der the lm-boiling condition, the liquidsolid contact is hindered by the vapor
layer, as evidenced by the disappearance of the bubbles inside the liquid droplet
under the nucleate boiling condition. The contact angle of the liquid to the
surface was also reported to increase with an increase in the surface temper-
ature, and reached 1801 in the lm boiling condition. Qiao and Chandra (1996)
measured the temperature drop of a stainless steel surface during the impact of
the subcooled water and the n-heptane droplets in low gravity. They found that
when the surface temperature is above the superheat limit, the temperature drop
of the surface is relatively small for the impact of n-heptane droplet (less than
20 1C). But for the impact of the water droplet, the temperature drop of the
surface can reach 150 1C, which implies a high heat ux and the intermittent
contact of the liquid and the solid surface. Hatta et al. (1997) found that at low
impact We number, the dynamics of water droplet is almost independent of
YANG GE AND LIANG-SHIH FAN 38
the surface materials when the surface temperature is above the Leidenfrost
temperature.
The effects of the subcooling degree of the droplet on the lm-boiling impact
are studied by Inada et al. (1985). They found that the heat-transfer rate on
the solid surface during an impact of a 4-mm water droplet increases signi-
cantly with a decrease in the initial droplet temperature. The boiling regimes
were classied to represent different droplet dynamics and heat-transfer modes
at various droplet and surface temperatures. Inada et al. (1988) also measured
the thickness of the vapor lm between the impinging droplet and the surface
at various degrees of subcooling. Chen and Hsu (1995) measured the tran-
sient local heat ux at the surface of a presuperheated plate, which undergoes
the impingement of subcooled water droplets. A fast-response microthermo-
couple was designed to capture the instantaneous changes of the solid-surface
temperature. Although the droplet dynamics of the impact process was not
presented, they concluded that both the surface temperature and the degree
of the droplet subcooling are crucial to the intermittent contact mode at the
solid surface. At the lm-boiling regime with the surface temperature super-
heated at 400 1C, a subcooled droplet tends to disintegrate during the impact at
We 55.
In subcooled impact, the initial droplet temperature is lower than the sat-
urated temperature of the liquid of the droplet, thus the transient heat transfer
inside the droplet needs to be considered. Since the thickness of the vapor layer
may be comparable with the mean free path of the gas molecules in the sub-
cooled impact, the kinetic slip treatment of the boundary condition needs to be
applied at the liquidvapor and vaporsolid interface to modify the continuum
system.
1. Hydrodynamic Model
The ow eld of the impacting droplet and its surrounding gas is simulated
using a nite-volume solution of the governing equations in a 3-D Cartesian
coordinate system. The level-set method is employed to simulate the movement
and deformation of the free surface of the droplet during impact. The details of
the hydrodynamic model and the numerical scheme are described in Sections
II.A and 1 V.A.1.
During the subcooled droplet impact, the droplet temperature will undergo
signicant changes due to heat transfer from the hot surface. As the liquid
properties such as density r
l
(T), viscosity m
l
(T), and surface tension s(T) vary
with the local temperature T, the local liquid properties can be quantied once
the local temperature can be accounted for. The droplet temperature is simu-
lated by the following heat-transfer model and vapor-layer model. Since the
liquid temperature changes from its initial temperature (usually room temper-
ature) to the saturated temperature of the liquid during the impact, the linear
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 39
variation of liquid properties with the temperature is assumed, which is
gT g
0

T T
0
T
sa
T
0
g
sa
g
0
g r
l
; m
l
; s (61)
where T
0
and T
sa
are the initial and saturated temperatures of the liquid re-
spectively; g
0
and g
sa
are the liquid property at T
0
and T
sa
, respectively. The
boundary condition adopted at the solid surface is described in Section IV.A.1.
2. Heat Transfer Inside the Droplet and Across the Vapor Layer
For the subcooling impact, especially for the high subcooling degree case in
which the droplet initial temperature is much lower than the saturated temper-
ature, the heat transfer within the droplet is signicant and hence affects the
droplet evaporation rate. Neglecting the viscous dissipation, the equation of the
conservation of the thermal energy inside the droplet is given by
@T
@t

~
V rT a
l
r rT (62)
where a
1
is the thermal diffusivity of liquid. At other free surfaces of the droplet,
the adiabatic boundary condition is applied which is given by
~n
G
r
d
T 0 (63)
where ~n
G
is the normal vector of the droplet surface, which can be calculated
based on the level-set function:
~n
G

rf
rf

(64)
r
d
T is the temperature gradient which is evaluated only on the droplet side. The
heat-conduction equation inside the solid surface is given in Section IV.A.3.
The heat transfer across the vapor layer and the temperature distribution in
the solid, liquid, and vapor phases are shown in Fig. 13. In the subcooled
impact, especially for a droplet of water, which has a larger latent heat, it has
been reported that the thickness of the vapor layer can be very small and in
some cases, the transient direct contact of the liquid and the solid surface may
occur (Chen and Hsu, 1995). When the length scale of the vapor gap is com-
parable with the free path of the gas molecules, the kinetic slip treatment of the
boundary condition needs to be undertaken to modify the continuum system.
Consider the Knudsen number dened as the ratio of the average mean free
path of the vapor to the thickness of the vapor layer:
Kn
l
d
(65)
YANG GE AND LIANG-SHIH FAN 40
where l is the mean free path of molecule. Harvie and Fletcher (2001c) analyzed
the kinetic of the molecular behavior at the solid and evaporative surface for
0.01oKno0.1. Based on their simple kinetic theory, the effective temperature
discontinuity at liquidvapor surface and solidvapor surface can be given by
T
s2
T
s1
C
T;s
T
s1
T
d2
(66)
T
d1
T
d2
C
T;l
T
s1
T
d1
(67)
where T
s1
,T
s2
,T
d1
,T
d2
are the interface temperatures of solid surface and droplet
shown in Fig. 13. C
T
is dened by
C
T
Kn
9
4
g
5
4

2 s
t
s
t

(68)
where s
t
is the thermal accommodation coefcient dened by Harvie and
Fletcher (2001c). At the liquidvapor interface, the energy balance equation is
given by
k
v
T
s1
T
d1
d
q
l
_ mL
c
(69)
where k
v
is the thermal conductivity of the vapor; L
c
the latent heat of the
liquid; q
1
the heat ux at the droplet surface at the liquid side, which is given by
q
1
k
1
r
d
T.
Td

Solid surface
Droplet
Td1
Ts2 Ts1
Td2
Ts
qs
ql
Vapor layer
FIG. 13. Temperature distribution and heat ux across the vapor layer.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 41
3. Vapor-Layer Model with Kinetic Treatment at Boundary
The vapor-layer model developed in Section IV.A.2 is based on the contin-
uum assumption of the vapor ow. This assumption, however, needs to be
modied by considering the kinetic slip at the boundary when the Knudsen
number of the vapor is larger than 0.01 (Bird, 1976). With the assumption
that the thickness of the vapor layer is much smaller than the radius of the
droplet, the reduced continuity and momentum equations for incompressible
vapor ows in the symmetrical coordinates (x,l) are given as Eqs. (43) and (47).
When the Knudsen number of the vapor ow is between 0.01 and 0.1, the ow
is in the slip regime. In this regime, the ow can still be considered as a
continuum at several mean free paths distance from the boundary, but an
effective slip velocity needs to be used to describe the molecular interac-
tion between the gas molecules and the boundary. Based on the simple
kinetic analysis of vapor molecules near the interface (Harvie and Fletcher,
2001c), the boundary conditions of the vapor ow at the solid surface can be
given by
l 0; u
x
x; 0 F
s
@u
x
@l
x; 0; u
l
x; 0 0 (70)
and the boundary conditions at the droplet surface is
l d; u
x
x; d u
x;l
x F
l
@u
x
@l
x; d; u
l
x; d u
l;l
x (71)
where u
x,l
(x), u
l,l
(x) are the velocities of the droplet surface in x and l direction,
respectively; F
s
and F
l
are dened by
F
s
l
s
2 s
v;s
s
v;s
F
l
l
l
2 s
v;l
s
v;l
(72)
where l
s
, l
l
are the mean free path of the gas molecules near solid and droplet
surface.
Equations (43) and (47) can be solved by using the similar procedure as given
by Section IV.A.2 with the boundary condition given by Eqs. (70) and (71).
Thus the vapor pressure can be determined by
px p
0

Z
x
b
x
rjx dx (73)
where x
0
is the radius of the extent of the vapor layer andjris given by
jx
2nC
1
d
2
u
x

u
x
u
ll
d
B
1
B
2

u
2
x
x
B
1
(74)
YANG GE AND LIANG-SHIH FAN 42
where
B
1

C
2
1
5

C
1
C
2
2

2C
1
C
3
C
2
2
3
C
1
C
3
C
2
3
B
2

2C
2
1
15

C
1
C
2
3

C
2
2
6

2
3
C
1
C
3

1
2
C
2
C
3
C
1
3
2 2k
l
2k
s
y
l
2k
s
y
l
1 k
l
4k
s
6k
l
k
s
C
2

6 6k
l
2y
l
1 k
l
4k
s
6k
l
k
s
C
3
k
s
C
2
75
where k
l
F
l
=d; k
s
F
s
=d; y
l
u
x;l
= u
x
.
The averaged velocity of the vapor is expressed by Eq. (54). The pressure
distribution in the vapor layer can be obtained by solving Eqs. (54) and (73)
(75) by a piecewise integration method. Details of the solving procedure and how
to use the vapor pressure in ow eld calculation are given in Section IV.A.2.
4. Results and Discussion
Three different subcooled impact conditions under which experiments were
conducted and reported in the literature are simulated in this study. They are:
(1) n-heptane droplets (1.5 mm diameter) impacting on the stainless steel surface
with We 43 (Chandra and Avedisian, 1991), (2) 3.8 mm water droplets
impacting on the inconel surface at a velocity of 1 m/s (Chen and Hsu, 1995),
and (3) 4.0 mm water droplets impacting on the copper surface with We 25
(Inada et al., 1985). The simulations are conducted on uniform Cartesian
meshes (Dx Dy Dz D). The mesh size (resolution) is determined by
considering the mesh renement criterion in Section V.A. The mesh sizes in this
study are chosen to provide a resolution of CPR 15.
Fig. 14 shows the comparison of the photographs from Chandra and
Avedisian (1991) with simulated images of this study for a subcooled 1.5 mm
n-heptane droplet impact onto a stainless-steel surface of 200 1C. The impact
velocity is 93 cm/s, which gives a Weber number of 43 and a Reynolds number
of 2300. The initial temperature of the droplet is room temperature (20 1C). In
Fig. 14, it can be seen that the evolution of droplet shapes are well simulated by
the computation. In the rst 2.5 ms of the impact (frames 12), the droplet
spreads out right after the impact, and a disk-like shape liquid lm is formed on
the surface. After the droplet reaches the maximum diameter at about 2.1 ms,
the liquid lm starts to retreat back to its center (frame 2 and 3) due to the
surface-tension force induced from the periphery of the droplet. Beyond 6.0 ms,
the droplet continues to recoil and forms an upward ow in the center of the
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 43
FIG. 14. n-Heptane droplet collision with surface at 200 1C. Experimental images (right) are
presented by Chandra and Avedisian (1991). We 45. The size of the last frame is reduced.
YANG GE AND LIANG-SHIH FAN 44
droplet (frames 3 and 4), which leads to the bouncing of the droplet up from the
surface (frame 5). The photograph at 8.0 ms shows that the tip of the elongated
droplet separates from the main body of the droplet and the main body of the
droplet then breaks up into smaller secondary drops (frame 5). This
phenomenon was reproduced accurately in the simulation.
Fig. 15 shows the detailed structure of the droplet from a viewing angle of
601. Experimental images show that a hole is formed in the center of the droplet
for a short time period (3.44.8 ms) and the center of the liquid droplet is a dry
circular area. The simulation also shows this hole structure although a minor
variation exists over the experimental images. As the temperature of the surface
is above the Leidenfrost temperature of the liquid, the vapor layer between the
droplet and the surface diminishes the liquidsolid contact and thus yields a low
surface-friction effect on the outwardly spreading liquid ow. When the droplet
periphery starts to retreat due to the surface-tension effect, the liquid in the
droplet center still ows outward driven by the inertia, which leads to the
formation of the hole structure.
The impact process of a 3.8 mm water droplet under the conditions experi-
mentally studied by Chen and Hsu (1995) is simulated and the simulation results
are shown in Figs. 16 and 17. Their experiments involve water-droplet impact
on a heated Inconel plate with Ni coating. The surface temperature in this
simulation is set as 400 1C with the initial temperature of the droplet given
as 20 1C. The impact velocity is 100 cm/s, which gives a Weber number of 54.
Fig. 16 shows the calculated temperature distributions within the droplet and
within the solid surface. The isotherm corresponding to 21 1C is plotted inside
the droplet to represent the extent of the thermal boundary layer of the droplet
that is affected by the heating of the solid surface. It can be seen that, in the
droplet spreading process (07.0 ms), the bulk of the liquid droplet remains at its
initial temperature and the thermal boundary layer is very thin. As the liquid
lm spreads on the solid surface, the heat-transfer rate on the liquid side of the
dropletvapor interface can be evaluated by
q
drop
k
l
T
d2
T
d;bulk
d
T
(76)
where d
T
is the thickness of the thermal boundary layer; T
d2
and T
d,bulk
are the
droplet temperature at the surface and in the bulk, respectively. The thermal
boundary layer thickness can be estimated by (Pasandideh-Ford et al., 2001):
d
T

2d
0
Re
0:5
Pr
0:4
(77)
where Re is the Reynolds number of the impinging liquid ow and is dened
by Re rVd
0
=m; and V the liquid lm velocity. At the early stage of the
spreading, V is close to the initial impact velocity of the droplet, and thus, it
gives a thin thermal boundary layer as shown in frames 1 and 2 of Fig. 16. When
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 45
FIG. 15. Experimental photos (left) by Chandra and Avedisian (1991) and simulated images
(right) of the spreading droplet on surface at 200 1C. The formation of a hole in the center of the
liquid is captured.
YANG GE AND LIANG-SHIH FAN 46
the droplet spreads to the maximum extent and starts to recoil, the liquid
velocity diminishes to zero and the thermal boundary layer is disrupted. Until
then, the temperature rise inside the droplet becomes signicant (frames 3 and 4
of Fig. 16). The simulated temperature distribution inside the solid surface is
also shown in Fig. 16, in which the isotherm of 399 1C is chosen to represent the
area of temperature drop during the impact of the droplet.
The simulation of droplet impact shown in Fig. 16 is conducted under per-
fectly symmetrical conditions, which is not easy to achieve in the experiments.
FIG. 16. Simulated temperature eld in the liquid and solid phases.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 47
As the droplet is released from the nozzle and moves toward the superheated
surface, some uncontrollable factors such as the angle of dropping, obliquity of
the surface, and perturbation in the ambient conditions render it difcult to
maintain a perfectly normal collision between the droplet and solid surface. The
3-D simulation of this study is capable of reproducing the imperfect drop-
letsurface impact condition represented by asymmetrical collision. Fig. 17
shows the simulated 3-D images of the droplet under the same impact condition
as that shown in Fig. 16 but with a 5 cm/s tangential velocity. The simulated
solid surface temperature is shown in Fig. 18 with comparisons to the experi-
mental measurements of Chen and Hsu (1995). It can be seen that with a
small obliquity, there is a signicant effect on the movement behavior of the
droplet. Specically, in this case, the droplet moves away from the impact center
during the recoiling process, which leads to a faster recovery of the solid surface
temperature.
The simulations are further conducted under the experimental conditions of
Inada et al. (1985). In their experiments, 4.0 mm water droplets impact on a
heated platinum surface at a temperature up to 420 1C. The subcooling degree
FIG. 17. Droplet impacts on the at surface with a small tangential velocity. Other conditions are
the same as those in Fig. 16.
YANG GE AND LIANG-SHIH FAN 48
of the droplet (dT
sub
) can be varied from 2 1C to 88 1C. The droplet falls down
20 mm before it impacts on the surface, thus the velocity of the impact can be
estimated as 64 cm/s, which gives a Weber number of about 25. In this simu-
lation, the initial temperature of surface is xed on 420 1C, which ensure
the nonwetting contact between the droplet and the surface. Fig. 19 shows the
simulated solid temperature compared with the measured value at two loca-
tions. T
1
and T
2
were measured at 0.28 and 0.74 mm depth beneath the surface
(Inada et al., 1985). Simulated surface temperatures agreed well with the meas-
ured T
1
and T
2
; both decrease until 1314 ms after the impact. The simulated
temperature at the surface of the solid (T
w
) is also shown in Fig. 19. The
maximum temperature drop at the surface is about 60 1C, which occurs at about
13 ms after the impact.
C. SIMULATION OF DROPLET PARTICLE COLLISION IN THE LEIDENFROST REGIME
An efcient numerical model to describe the unsteady, 3-D uid ow during
the dropletparticle collision with evaporation is developed (Ge and Fan, 2005).
From the numerical point of view, the droplet and solid surface need to be
0 5 10 15 20 25 30 35 40 45 50
0
100
200
300
400
500
at z= 0.00mm
at z=-0.12mm
at z=-0.24mm
at z=-0.36mm
Experiment of Chen and Hsu(1995)
t
e
m
p
e
r
a
t
u
r
e

o
f

s
o
l
i
d

s
u
r
f
a
c
e

a
t

r
=
0

(

C
)
Time (ms)
FIG. 18. Simulated solid surface temperatures compared with the experiments of Chen and Hsu
(1995). The droplet impacts on the at surface with a small tangential velocity (5 cm/s) as shown in
Fig. 17.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 49
captured in the ow eld and the solid-ow boundary condition at the particle
surface needs to be satised. The numerical method adopted is a combination of
the level-set approach and the IBM. The evaporation effect is accounted for by
the vapor pressure force, which is calculated in a dynamics vapor-ow model.
Energy balance equations in each phase are solved with interface boundary
condition to give the temperature distribution and evaporation rate.
1. Hydrodynamic Model with Level-Set Method
In this model, two level-set functions (f
d
, f
p
) are dened to represent the
droplet interface (f
d
) and the moving particle surface (f
p
), respectively. The free
surface of the droplet is taken as the zero in the droplet level-set function
f
d
~x; t, and the advection equation (Eq. (3)) of the droplet level-set function
(fd) is solved to capture the motion of the droplet surface. The particle level-set
function (fp) is dened as the signed distance from any given point ~x in the
Eulerian system to the particle surface:
f
p
~x ~x
0
t

R
p
(78)
where ~x
0
t is the position vector of the center of the particle and R
p
the particle
radius. With this denition, f
p
40 when ~x is outside the particle, f
p
o0 when ~x
is inside the particle.
0 5 10 15 20
350
360
370
380
390
400
410
420
430
440
Simulated T
W
Simulated T
1
Measured T
1
Simulated T
2
Measured T
2
S
u
r
f
a
c
e

t
e
m
p
e
r
a
t
u
r
e

(

C
)
time after impact (ms)
FIG. 19. Simulated solid surface temperatures compared with the experiments of Inada et al.
(1985). T
1
and T
2
are the temperatures at locations inside the surface with Z
1
0.28 mm and
Z
2
0.74 mm.
YANG GE AND LIANG-SHIH FAN 50
The momentum equation for this 3-phase ow system in the Eulerian frame
can be given by
r
@
~
V
@t
r
~
V
~
V
!
rp r~g r 2m
~
D skf
d
df
d
rf
d

~
F
p
f
p

~
F
vapor
79
In this equation, the presence of the solid particle in the uid is represented by
a virtual boundary body force eld,
~
F
p
f
p
, dened by the IBM which will be
discussed in Section IV.C.2.
~
F
vapor
is vapor pressure force exerting on the
dropletparticle contact area due to the effect of the evaporation, which will be
discussed in vapor-layer model of Section IV.C.3.
2. Immersed-Boundary/Level-Set Method for Particle Flow Interaction
In the IBM, the presence of the solid boundary (xed or moving) in the uid
can be represented by a virtual body force eld
~
F
p
f
p
applied on the compu-
tational grid at the vicinity of solidow interface. Considering the stability and
efciency in a 3-D simulation, the direct forcing scheme is adopted in this model.
Details of this scheme are introduced in Section II.B. In this study, a new velocity
interpolation method is developed based on the particle level-set function (f
p
),
which is shown in Fig. 20. At each time step of the simulation, the uidparticle
boundary condition (no-slip or free-slip) is imposed on the computational cells
located in a small band across the particle surface. The thickness of this band can
be chosen to be equal to 3D, where D is the mesh size (assuming a uniform mesh
is used). If a grid point (like p and q in Fig. 20), where the velocity components of
the control volume are dened, falls into this band, that is
Dx4f
p
4Dx (80)
the velocity at these points will be redened using linear interpolation based on
the velocity and particle level-set function (f
p
) at the neighboring grid point. At
grid point p in Fig. 20, which is located in the band and outside the particle
surface, the uid velocity is determined by (for no-slip boundary condition):
U
p
V
p
t
f
p
n
X
n
i
U
p
0
;i
V
p
t
f
p
0
;i
(81)
where U
p
0 and f
p
0 are the uid velocity and particle level-set function at grid
point p
0
(in Fig. 20). It should be noted that only the neighboring grid points
located outside the band are chosen to interpolate the velocity at point p. The
velocities at these points (p
0
) are obtained by solving the N-S equation like other
points far from the interface. n is the total number of these neighboring points. It
can be seen that when f
p
0, meaning that the grid point p is right on the
particle surface, the velocity at p is equal to the particle velocity: U
p
V
p
(t).
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 51
The velocity at grid point q (in Fig. 20), which is located inside the small band
and on the particle side, is determined by interpolating the uid velocity at the
neighboring grid points outside the particle surface:
U
q
V
p
t
f
q
m
X
m
i
U
p;i
V
p
t
f
p;i
(82)
where m is the total number of neighboring grid points which are located out-
side the particle surface.
3. Vapor Layer and Heat-Transfer Model
As the vapor ows in the direction along the spherical surface of the particle,
a boundary layer coordinate (x, l, o) given in Fig. 21 is employed to describe
the vapor-layer equation. In this coordinate, the continuity and momen-
tum equations for incompressible vapor ows with gravitation terms neglected
P
p
q
q
P
P
Up
Up
x
x
Particle
interface
Fluid
FIG. 20. Velocity interpolation scheme based on the particle level-set function.
YANG GE AND LIANG-SHIH FAN 52
are given by
@u
x
@x

cotx=R
R
u
x

2
R
u
l

@u
l
@l
0 (83)
u
x
@u
x
@x
u
l
@u
x
@l

u
x
u
l
R

@
@x
P
r

n
@
2
u
l
@x@l

2
R
@u
x
@l

@
2
u
x
@l
2

(84)
u
x
@u
l
@x
u
l
@u
l
@l

u
2
x
R

@
@l
P
r

n
R
R
@
2
u
l
@x
2

@u
x
@x
R
@
2
u
x
@x@l

cot
x
R

@u
l
@x

u
x
R

@u
x
@l
!
85
where u
x
, u
l
are the vapor-ow velocities in x and l direction, respectively.
Based on the assumptions and the order of magnitude analysis described in
Section IV.A.2, and following the similar procedure of solution, the averaged
vapor-ow velocity is given by
u
x
x O
Z
1
0
F
d
dZ
1
2
u
ld
x
d
2
12g
1
3
20
Re
d

@
@x
p
r

(86)
The vapor continuity equation can be expressed by
u
x
x
1
xdx
Z
x
0
x
0
u
ld
x
0
dx
0
(87)
P
O
Particle surface

FIG. 21. Boundary layer coordinates.


SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 53
The u
ld
(x) is calculated using Eq. (55). The pressure distribution in the vapor
layer can be obtained by solving Eqs. (86) and (87). In this procedure, the
thickness of the vapor layer (d(x)) and the extent of the layer (x
b
) are obtained
through a vapor-layer identication scheme based on the particle and droplet
level-set function. Fig. 22 illustrates the details of the scheme.
As the particle is in motion, at every time step, a series of grid points near the
particle surface are rst identied to measure the vapor layer. As shown in Fig.
22, these grids points are in a small band around the surface and can be outside
the surface (y,i1, i, i+1, i+2,y) or inside the surface (y, i
0
1, i
0
, i
0
+1,
i
0
+2,y). If the droplet surface is represented by points (y, P
i1
, P
i
, P
i+1
,y)
in Fig. 22 and point P
i
is located on the mesh line between the mesh knots i and
i
0
, the vapor-layer thickness at P
i
can be calculated based on the values of the
level-set function at i and i dened as (f
d,i
,f
p,i
) and (f
d,i
0 ,f
p,i
0 ), respectively.
Since the level-set function is the signed distance from the computation knots to
the droplet and particle surface after the redistance process is performed, the
vapor-layer thickness (d
i
) at P
i
can be estimated by
d
i

f
d;i
f
p;i
f
d;i
0 f
p;i
0
2
(88)
The heat-transfer model described in Sections IV.A.3 and IV.B.2 can be ap-
plied to calculate the temperature distribution inside the droplet and energy

p,i
P
i
i
Particle Surface
-
d,i,
Droplet Surface

i
P
i+1
P
i-1
i-1
i+1
i-1 i
i+3
i+1
i+2
i+2

FIG. 22. Scheme for determining the thickness of the vapor layer.
YANG GE AND LIANG-SHIH FAN 54
balance at the interface. Inside the particle, the heat-conduction equation is
@T
@t
a
s
r rT (89)
where a
s
is the thermal diffusivity of particle. When the particle surface is in
contact with the droplet through the vapor layer, the boundary condition at the
particle surface is given by
k
v
T
ss
T
ds
d
k
s
~n
p
r
p
T
s
(90)
where ~n
p
is the normal vector of the particle surface, r
p
T the temperature
gradient which is evaluated only on the particle side. This boundary condition,
Eq. (90) is imposed on the computational cells, which are located in a small
band near the particle surface on the solid side.
4. Results and Discussion
Dropletparticle collision conditions under which experiments are conducted
are simulated in this study. In the experiment, the brass particle is heated on
a heating plate with adjustable temperature settings, and a high-speed camera
capable of capturing 500 frames per second will be used to record the drop-
letparticle collision process. A droplet is formed by the use of the syringe with
various needle sizes. In the experiments, acetone droplets with 1.62.2 mm di-
ameter impact onto the brass particles of sizes 5.5 mm or 3.6 mm.The particle
temperature is 200 1C300 1C, which is much higher than the boiling point of
acetone and ensures that the contact is nonwetting.
The simulation is conducted on uniform Cartesian meshes (Dx Dy
Dz D). The mesh sizes in this study are chosen to give a resolution of
CPR 15. Fig. 23 shows the comparison of the photographs with simulated
images for a 2.1 mm acetone droplet impact onto a 5.5 mm brass particle of
250 1C. The impact velocity is 45 cm/s. The initial temperature of the droplet is
close to the boiling point of acetone (56 1C). A comparison of the images shows
that the impact process is predicted well by this model. Similar to the impact on a
at surface, the droplet spreads on the particle surface at the rst stage
(05.5 ms), then recoils due to the vapor pressure force and surface tension force
(7.513.5 ms), and eventually rebounds away from the particle (17.527.5 ms).
The droplet lm reaches the maximum extent at about 7.0 ms, by which time the
radius of dropletparticle contact area is about 1.5 times that of the original
radius of the droplet. The total contact time is about 17 ms, which is relevant to
the rst-order vibration period of the oscillating drop.
Fig. 24 shows the collision process between a moving particle (1.5 mm in
diameter) and a large water droplet. A water droplet 2.5 mm in diameter has
an initial velocity of 25 cm/s and an initial temperature of 100 1C. The initial
velocity and temperature of the particle are 25 cm/s and 400 1C, respectively.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 55
The physical properties of the particle are the same as FCC particles. The
simulation is conducted using a 140 140 200 rectangular mesh covering a
7 mm 7 mm10 mm computational domain. Both the 3-D images and the
temperature eld are shown in this gure.
It can be seen in this gure that since the inertia (mass) of the particle is
smaller than that of the droplet, the particle velocity decreases rapidly as soon as
the particle collides with the droplet. After the collision, the surface tension of
-0.5ms
1.5ms
13.5ms
5.5ms
17.5ms
21.5ms
27.5ms 7.5ms
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.1 0.2 0.3 0.4
0.5 0.6 0
0.5
FIG. 23. Experimental photos (left) and simulated images (right) of the 2.1 mm acetone droplet
impact on 5.5-mm particle at 250 1C. Impact velocity V 45 cm/s.
YANG GE AND LIANG-SHIH FAN 56
0.8
0.7
t = 0.0ms
t = 10ms
t = 14ms
t = 17ms
t = 24ms
0.6
0.5
0.4
0.3
0.2
0.2
0.4 0.6 0.8
1 1.2
0
0.2
0.4
0.6
0.1
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.2
0.4 0.6 0.8
1 1.2
0
0.2
0.4
0.6
0.1
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.2
0.4 0.6 0.8
1 1.2
0
0.2
0.4
0.6
0.1
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.2
0.4 0.6 0.8
1 1.2
0
0.2
0.4
0.6
0.1
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.2
0.4 0.6 0.8
1 1.2
0
0.2
0.4
0.6
0.1
0.7
400
350
300
250
200
150
100
50
0
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
0.7
400
350
300
250
200
150
100
50
0
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
0.7
400
350
300
250
200
150
100
50
0
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
0.7
400
350
300
250
200
150
100
50
0
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
0.7
400
350
300
250
200
150
100
50
0
0.6
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
FIG. 24. Particle and droplet collision at the same initial velocity. The images on the right show
the temperature eld.
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 57
the droplet and the vapor-layer pressure induced by the evaporation drive the
particle to the inverse direction. Since the Weber number of the collision is
not very large (We 15), the droplet only undergoes small deformation without
splashing. After collision, the droplet still moves along its original path with
a decreased velocity, while the rebounding velocity of the particle is larger than
that of the droplet. It can also be found that the temperature drop of the particle
is not signicant in this condition.
V. Concluding Remarks
The most signicant advantage of the level-set method and the IBM in solv-
ing 3-D, three-phase ow problems is that the governing equations can con-
veniently be discretized and solved on xed, regular structured grids, rather than
resorting to the classic body-tted mesh approach. These methods are efcient
in utilizing computational resources while retaining the accuracy of the com-
putational results. The level-set method implicitly captures the motion of the
interface by solving the advection equation of the level-set function, and can be
easily implemented for 3-D interface tracking. The level-set methods are par-
ticularly effective in handling ow problems that involve changes in the topo-
logy of evolving interfaces, and in which the speed of the interface is sensitive to
local surface geometry, such as curvature and normal vector. Although, like
other front-capture techniques, the level-set method may suffer from some
problems associated with the preservation of mass conservation (Puckett et al.,
1997), the method is able to accurately compute interfacial ows with surface
tension force and other complex physical forces acting on the front. In the
conventional IBM, the moving front is represented by a nite number of dis-
crete Lagrangian markers along the interface, which move in the ow. In
general, this arrangement has the property of preservation of the clear interface
position and hence mass conservation; however, with this arrangement, it is
difcult to trace the interface with complicated shape and topological change in
three dimensions. The combination of the level-set method and the IBM as
described in this study can allow accuracy in interface movement, as well as
conservation in moving object mass when simulating the complex three-phase
interaction systems.
In system 1, the 3-D dynamic bubbling phenomena in a gasliquid bubble
column and a gasliquidsolid uidized bed are simulated using the level-set
method coupled with an SGS model for liquid turbulence. The computational
scheme in this study captures the complex topological changes related to the
bubble deformation, coalescence, and breakup in bubbling ows. In system 2,
the hydrodynamics and heat-transfer phenomena of liquid droplets impacting
upon a hot at surface and particle are analyzed based on 3-D level-set method
and IBM with consideration of the lm-boiling behavior. The heat transfers in
YANG GE AND LIANG-SHIH FAN 58
each phase are solved using a microscale vapor-ow model which is applied to
determine the vapor pressure force during the contact process between the
droplet and the superheated surface. The simulation model developed in this
study is capable of reproducing the droplet impaction process with signicant
heat transfer and phase change (evaporation). Both the saturated impact and
the subcooling impact are considered. The simulation results are found to be in
good agreement with the experimental measurements of the droplet deforma-
tion process and the surface-temperature variation.
Nomenclature
NOTATION
C
D
modied drag coefcient
C
D0
drag coefcient
C
s
Smagorinsky coefcient
~
D stress tensor
d distance function (or diameter)
d
p
diameter of particle
f
am
added mass force
f
d
uidparticle interaction force
f
dr
drag force
f
s
surface tension force
g gravitational acceleration
H
b
heaviside function
k heat conductivity
Kn Knudsen number
l length scale
L
c
latent heat
m
p
mass of particle
_ m mass evaporation rate
n normal vector
p pressure
Pr Prandtl number
q heat ux (or virtual mass source)
R radius of the droplet
Re particle Reynolds number
S shear strain rate
T temperature
t time
U velocity scale of the vapor ow
u velocity of the vapor ow
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 59
V velocity (or volume)
We Weber number
x position vector
GREEK LETTERS
a thermal diffusivity
D grid size
G gasliquid interface
F function dened in vapor-layer model (4.7)
O cell volume (or function dened in vapor-layer model (4.7))
b half of the thickness of the interface
d d function (or the thickness of the vapor-layer or thermal boundary
layer)
e void fraction
f level-set function
k curvature of the surface
l constant dened in Eq. (24) (or coordinate dened in vapor-layer
model)
m molecular viscosity
n
T
kinematic turbulent viscosity
r density
s surface tension (or thermal accommodation coefcient)
t viscous stress tensor, articial time
x coordinate dened in vapor-layer model
Z dimensionless variable dened in Eq. (46)
n kinematic viscosity of vapor ow
SUBSCRIPTS AND SUPERSCRIPTS
b boundary
d droplet
g gas phase
l liquid phase
o initial
p particle
r radial direction
s solid
v vapor phase
YANG GE AND LIANG-SHIH FAN 60
x vapor-layer coordinate
l vapor-layer coordinate
G gasdroplet interface
ij cell index
k particle index
sa saturated
sg sub-grid scale
REFERENCES
Agresar, G., Linderman, J. J., Triggvason, G., and Powell, K. G. J. Comp. Phys. 143, 346 (1998).
Amsden, A. A., and Harlow, F. H. J. Comp. Phys. 3, 80 (1968).
Anderson, T. B., and Jackson, R. Ind. Eng. Chem. Fundam. 6, 525539 (1967).
Biance, A. -L., Clanet, C., and Quere, D. Phys. Fluids 15(6), 16321637 (2003).
Bird, G. A., Molecular Gas Dynamics. Clarendon Press, Oxford, UK (1976).
Brackbill, J. U., Kothe, D. B., and Zemach, C. J. Comp. Phys. 100, 335 (1992).
Bussmann, M., Mostaghimi, J., and Chandra, S. Phys. Fluids 11(6), 14061417 (1999).
Bussmann, M., Mostaghimi, J., and Chandra, S. Phys. Fluids 12(12), 31213132 (2000).
Chandra, S., and Avedisian, C. T. Proc. R. Soc. Lond. A 432, 1341 (1991).
Chen, C., and Fan, L. -S. AIChE J 50, 288301 (2004).
Chen, J. C., and Hsu, K. K. J. Heat Transfer 117, 693697 (1995).
Chen, Y. -M., and Fan, L. -S. Chem. Eng. Sci. 44, 2762 (1989a).
Chen, Y. -M., and Fan, L. -S. Chem. Eng. Sci. 44, 117 (1989b).
Crowe, C. T., Sharma, M. P., and Stock, D. E. J. Fluids Eng. 99, 325 (1977).
Crowe, C., Sommerfeld, M., and Tsuji, Y., Multiphase Flows with Droplets and Particles. CRC
Press, NY (1998).
Dandy, D. S., and Leal, L. G. J. Fluid Mech. 208, 161 (1989).
Delnoij, E., Kuipers, J. A. M., and Van Swaaij, W. P. M. Chem. Eng. Sci. 52, 3623 (1997).
Fadlun, E. A., Verzicco, R., Orlandi, P., and Yusof, J. M. Comp. Phys. 161, 3560 (2000).
Fan, L. S., GasLiquidSolids Fluidization Engineering. Butterworths, Stoneham, MA (1989).
Fan, L. -S., and Tsuchiya, K., Bubble Wake Dynamics in Liquid and LiquidSolid Suspensions.
Butterworth-Heinemann, Stoneham, MA (1990).
Fan, L. -S., Lau, R., Zhu, C., Vuong, K., Warsito, W., Wang, X., and Liu, G. Chem. Eng. Sci. 56,
58715891 (2001).
Feng, J., Hu, H. H., and Joseph, D. D. J. Fluid Mech. 261, 95134 (1994a).
Feng, J., Hu, H. H., and Joseph, D. D. J. Fluid Mech. 277, 271301 (1994b).
Feng, Z. -G., and Michaelides, E. E. J. Comput. Phys. 202, 2051 (2005).
Fujimoto, H., and Hatta, N. J. Fluids Eng. 118, 142149 (1996).
Fukai, J., Shiiba, Y., Yamamoto, T., and Miyatake, O. Phys. Fluids 7(2), 236 (1995).
Ge, Y., and Fan, L. -S. Phys. Fluids 17, 027104 (2005).
Goldstein, D., Handler, R., and Sirovich, L. J. Comp. Phys. 105, 354 (1993).
Gottfried, B. S., Lee, C. J., and Bell, K. J. Int. J. Heat Mass Transfer 9, 11671187 (1966).
Groendes, V., and Mesler, R. Proceedings of the 7th International Heat Transfer Conference,
Munich, Federal Republic of Germany 1982, 1982, pp. 131136.
Harlow, F. H., and Welch, J. E. Phys. Fluid 8, 2182 (1965).
Harlow, F. H., and Shannon, J. P. J. Appl. Phys. 38, 3855 (1967).
Harvie, D. J. E., and Fletcher, D. F. Int. J. Heat Mass Transfer 44, 26332642 (2001a).
Harvie, D. J. E., and Fletcher, D. F. Int. J. Heat Mass Transfer 44, 26432659 (2001b).
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 61
Harvie, D. J. E., and Fletcher, D. F. J. Heat Transfer 123, 486491 (2001c).
Hatta, N., Fujimoto, H., Kinoshita, K., and Takuda, H. J. Fluids Eng. 119, 692699 (1997).
Hirt, C. W., Amsden, A. A., and Cook, J. L. J. Comput. Phys. 14, 227 (1974).
Hirt, C. W., and Nichols, B. D. J. Comput. Phys. 39, 201 (1981).
Hoomans, B. P. B., Kuipers, J. A. M., Briels, W. J., and van Swaaij, W. P. M. Chem. Eng. Sci. 51,
99118 (1996).
Inada, S., Miyasaka, Y., and Nishida, K. Bull. JSME 28, 26752681 (1985).
Inada, S., Miyasaka, Y., Sakamoto, K., and Hojo, K. J. Chem. Eng. Japan 21, 463 (1988).
Jackson, R., The Dynamics of Fluidized Particles. Cambridge University Press, NY (2000).
Jamet, D., Lebaigue, O., Coutris, N., and Delhaye, J. M. J. Comput. Phys. 169, 624 (2001).
Joseph, D. D., and Lundgren, T. S. Int. J. Multiphase ow 16, 3542 (1990).
Kanai, A., and Mtyata, H. Numerical simulation of bubbles in a boundary layer by Maker-
Density-Function. Proceedings of the 3rd International Conference on Multiphase Flow,
Lion, France (1998).
Karl, A., Anders, K., Rieber, M., and Frohn, A. Part. Part. Syst. Charact. 13, 186191 (1996).
Kashiwa, B. A., Padial, N. T., Rauenzahn, R. M., and Vanderheyden, W. B. Los Alamos National
Laboratory Research Report, LA-UR-93-3922 (1994).
Kim, J., Kim, D., and Choi, H. J. Comput. Phys. 171, 132150 (2001).
Kothe, D. B., and Rider, W. J. Los Alamos National Laboratory Research Report, LA-UR-94-3384
(1995).
Kunii, D., and Levenspiel, O., Fluidization Engineering. 2nd ed Butterworth-Heinemann, Boston
(1991).
Lai, M. -C., and Peskin, C. S. J. Comput. Phys. 160, 705719 (2000).
Lapin, A., and Lu bbert, A. Chem. Eng. Sci. 49, 3661 (1994).
Li, Y., Zhang, J., and Fan, L. -S. Chem. Eng. Sci. 54, 5101 (1999).
McHyman, J. Physica D 12, 396 (1984).
Mehdi-Nejad, V., Mostaghimi, J., and Chandra, S. Phys. Fluids 15(1), 173183 (2003).
Mittal, R., and Iaccarino, G. Annu. Rev. Fluid Meth. 37, 239261 (2005).
Monaghan, J. J. Comput. Phys. 110, 399 (1994).
Mudde, R. F., and Simonin, O. Chem. Eng. Sci. 54, 5061 (1999).
Osher, S., and Sethian, J. A. J. Comput. Phys. 79, 12 (1988).
Pasandideh-Ford, M., Bhola, R., Chandra, S., and Mostaghimi, J. Int. J. Heat Mass Transfer 41,
29292945 (1998).
Pasandideh-Ford, M., Aziz, S. D., Chandra, S., and Mostaghimi, J. Int. J. Heat Fluid Flow 22, 201
(2001).
Peskin, C. S. J. Computat. Phys. 25, 220 (1977).
Puckett, E. G., Almgren, A. S., Bell, J. B., Marcus, D. L., and Rider, W. J. J. Computat. Phys. 100,
269 (1997).
Qiao, Y. M., and Chandra, S. Int. J. Heat Mass Transfer 39(7), 13791393 (1996).
Rowe, P. N., and Henwood, G. A. Part. 1 Trans. Inst. Chem. Eng 39, 43 (1961).
Sato, T., and Richardson, S. M. Int. J. Numer. Meth. Fluids 19, 555 (1994).
Scardovelli, R., and Zeleski, S. Ann. Rev. Fluid Mechanics 31, 567 (1999).
Sethian, J. A., and Smereka, P. Annu. Rev. Fluid Meth. 35, 341372 (2003).
Smagorinsky, J. Mon. Weather Rev. 91, 99 (1963).
Sokolichin, A., and Eigenberger, G. Chem. Eng. Sci. 49, 5735 (1994).
Sokolichin, A., and Eigenberger, G. Chem. Eng. Sci. 54, 2273 (1999).
Sussman, M., Smereka, P., and Osher, S. J. Comput. Phys. 114, 146 (1994).
Sussman, M., Fatemi, E., Smereka, P., and Osher, S. Computers Fluids 114, 146 (1998).
Sussman, M., and Fatemi, E. SIAM J. Sci. Comput. 20, 1165 (1999).
Tryggvason, T., Bunner, B., Esmaeeli, A., Juric, D., Al-Rawahi, N., Tauber, W., Han, J., Nas, S.,
and Jan, Y. -J. J. Comput. Phys. 169, 708759 (2001).
Tsuji, Y., Kawaguchi, T., and Tanaka, T. Powder Technol 77, 7981 (1993).
YANG GE AND LIANG-SHIH FAN 62
Udaykumar, H. S., Kan, H. -C., Shyy, W., and Tran-Son-Try, R. J. Comput. Phys. 137, 366405
(1997).
Unverdi, S. O., and Tryggvason, G. J. Computat. Phys. 100, 25 (1992a).
Unverdi, S. O., and Tryggvason, G. Physica D 60, 70 (1992b).
Wachters, L. H. J., and Westerling, N. A. J. Chem. Eng. Sci. 21, 10471056 (1966).
Wen, C. Y., and Yu, Y. H. Chem. Eng. Prog. 62, 100 (1966).
Yabe, T. Interface Capturing and Universal Solution of Solid, Liquid and Gas by CIP Method.
Proceedings of the High-Performance Computing of Multi-Phase Flow, Tokyo, July 1819,
1997.
Ye, T., Mittal, R., Udaykumar, H. S., and Shyy, W. J. Computat. Phys. 156, 209240 (1999).
Yusof, J. M. Combined immersed boundaries/B-splines methods for simulations of ows in complex
geometries. CTR Annual Research Briefs, NASA Ames/Stanford University (1997).
Zenit, R., and Hunt, M. J. Fluids Eng. 121, 179 (1999).
Zhang, D. Z., and Prosperetti, A. J. Fluid Mech. 267, 185219 (1994).
Zhang, J., Fan, L. -S., Zhu, C., Pfeffer, R., and Qi, D. Powder Technol 106, 98 (1999).
Zhang, Z., and Prosperetti, A. J. Appl. Mech.-Trans. ASME 70, 6474 (2003).
Zhu, C., Wang, X., and Fan, L. -S. Powder Technol 111, 7982 (2000).
SIMULATION OF GAS LIQUID AND GAS LIQUID SOLID FLOW SYSTEMS 63
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS
M.A. van der Hoef
1
, M. Ye
1
, M. van Sint Annaland
1
, A.T. Andrews IV
2
,
S. Sundaresan
2
and J.A.M. Kuipers
1,
1
Department of Science & Technology, University of Twente, 7500 AE Enschede,
The Netherlands
2
Department of Chemical Engineering, Princeton University, Princeton, NJ 08544, USA
I. Introduction 66
A. Gas-Fluidized Beds 66
B. Numerical Models for Gas and Solid Flows 67
C. The Multi Level Modeling Approach for GasSolid
Flows 70
II. Lattice Boltzmann Model 74
A. From Lattice-Gas to Lattice-Boltzmann 74
B. The Lattice BhatnagarGrossKrook Model 78
C. Modeling Solid Particles 81
D. Results for the GasSolid Drag Force 83
III. Discrete Particle Model 86
A. Introduction 86
B. Particle Dynamics: The Soft-Sphere Model 89
C. Gas Dynamics 100
D. Interphase Coupling 102
E. Energy Budget 106
F. Results for the Excess Compressibility 107
IV. Two-Fluid Model 111
A. Introduction 111
B. GOVERNING Equations 113
C. General Kinetic Theory 115
D. Kinetic Theory of Granular Flow 119
E. Numerical Solution Method 120
F. Application to Geldart A Particles 127
V. Towards Industrial-Scale Models 131
A. The Limits of the Two-Fluid Model 132
B. State-of-the-Art on Dealing with Unresolved Structures 135
C. A Different Approach: The Discrete Bubble Model 141
VI. Outlook 143
References 146

Corresponding author. Tel: +31 53 489 3039; Fax: +31 53 489 2882 E-mail: j.a.m.kuipers@tnw.
utwente.nl
65
Advances in Chemical Engineering, vol. 31
ISSN 0065-2377
DOI 10.1016/S0065-2377(06)31002-2
Copyright r 2006 by Elsevier Inc.
All rights reserved
Abstract
Numerical models of gas-uidized beds have become an important
tool in the design and scale up of gas-solid chemical reactors. However, a
single numerical model which includes the solid-solid and solid-uid in-
teraction in full detail is not feasible for industrial-scale equipment, and
for this reason one has to resort to a multiscale approach. The idea is
that gas-solid ow is described by a hierarchy of models at different
length scales, where the particle-particle and uid-particle interactions
are taken into account with different levels of detail. The results and
insights obtained from the more fundamental models are used to develop
closure laws to feed continuum models which can be used to compute the
ow structures on a much larger (engineering) scale. Our multi-scale
approach involves the lattice Boltzmann model, the discrete particle
model, and the continuum model based on the kinetic theory of granular
ow. In this chapter we give a detailed account of each of these models as
they are employed at the University of Twente, accompanied by some
illustrative computational results. Finally, we discuss two promising ap-
proaches for modeling industrial-size gas-uidized beds, which are cur-
rently being explored independently at the Princeton University and the
University of Twente.
I. Introduction
A. GAS-FLUIDIZED BEDS
Gas-uidized beds consist of ne granular material (usually smaller than
5 mm) that are subject to a gas ow from below, large enough so that the gas
drag on the particles can overcome the gravity and the particles can uidize.
When in the uidized state, the moving particles work effectively as a mixer
resulting in a uniform temperature distribution and a high mass transfer rate,
which are benecial for the efciency of many physical and chemical processes.
For this reason, gas-uidized beds are widely applied in the chemical, petro-
chemical, metallurgical, environmental, and energy industries in large-scale op-
erations involving adhesion optimized coating, granulation, drying, and synthesis
of fuels and base chemicals (Kunii and Levenspiel, 1991). Lack of understanding
of the fundamentals of dense gasparticle ows in general has led to severe
difculties in the design and scale-up of these industrially important gassolid
contactors (van Swaaij, 1985). In most cases, the design and scale-up of uidized
bed reactors is a fully empirical process based on preliminary tests on pilot-scale
model reactors, which is a very time consuming and thus expensive activity.
Clearly, computer simulations can be a very useful tool to aid this design
and scale-up process. Basically, such simulations can be used for two different
M.A. VAN DER HOEF ET AL. 66
purposes. Firstly, to contribute to our understandingthat is, the simulations
are used to obtain a fundamental insight into the complex dynamic behavior of
dense uidized suspensions, which should lead to an understanding based on
elementary physical principles such as drag, friction, dissipation, etc.; this also
includes the testing of elementary assumptions in theoretical models, such as the
Maxwell velocity distribution of the particles. Secondly, the simulations can be
used as a design tool, where the ultimate goal is to have a numerical model with
predictive capabilities for the dense gasparticle ows encountered in engineer-
ing-scale equipment. Clearly, it will not be possible to have one single simulation
method that can achieve this, but one rather needs a hierarchy of methods for
modeling the ow phenomena on different length and time scales.
Obviously, these two items are not strictly separated; in contrast, the most
fruitful approach is when they are simultaneously followed, so that they can
mutually benet from each other. In this chapter, we want to focus on the use of
simulation methods as a design tool for gas-uidized bed reactors, for which we
consider gassolid ows at four distinctive levels of modeling. However, before
discussing the multilevel scheme, it is useful to rst briey consider the numer-
ical modeling of the gas and solid phase separately.
B. NUMERICAL MODELS FOR GAS AND SOLID FLOWS
1. Gas Phase
The description of a gas ow is well established from the micro- to macro-
scales. The length scale of a gas ow can be characterized by the local Knudsen
number Kn, which is dened as
Kn
l
L
where l is the mean free path of the molecules and L is the characteristic length
scale of the ow. The models to describe a gas ow are schematically shown in
Fig. 1. For large-scale systems with Kno0.01, the gas ow can be described by
ordinary uid dynamics where the macroscopic elds (such as density and ve-
locity) are formulated by NavierStokes equations in a three-dimensional (3D)
coordinate space, together with no-slip boundary conditions. A number of well-
developed numerical algorithms and meshing techniques in computational uid
dynamics (CFD) can be used to handle very complex geometries (Anderson,
1995). If the system becomes smaller, say 0.01oKno0.1, the NavierStokes
equations still hold, but caution must be exercised for the boundary conditions
because partial slip might exist between the gassolid interfaces. For a rareed
gas where Kn40.1, the continuum assumption breaks down, and the so-called
kinetic theory of (dense) molecular gases should be applied. Kinetic theory
differs from the ordinary uid dynamics as there is just one eld (the density
of molecules) in the phase space. The basic equation in kinetic theory, in the
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 67
simplest form, is the Boltzmann equation that describes the evolution of the
density function f in a six-dimensional (6D) space (three coordinates and three
velocity components) (Chapman and Cowling, 1970). At this scale, computa-
tional techniques such as molecular dynamics (MD) (Allen and Tildesley, 1990)
and direct simulation Monte Carlo (DSMC) approach (Bird, 1976) can be very
efcient. In these techniques, the motion of molecules is traced on an individual
basis. Gas pressure and other transport coefcients, including gas viscosity and
thermal conductivity, are obtained by methods from statistical mechanics
(Chapman and Cowling, 1970). Molecules can be treated either as hard spheres
or points with certain interaction potentials, depending on their physical prop-
erties. In the extreme case where the mean free path is very large compared to the
system sizes (i.e., Kn410), the molecules move freely and just collide with the
walls. This is the limit of free molecular ow, where the system behaves as an
ideal gas.
Clearly, there are two quite different types of models for a gas ow: the
continuum models and the molecular models. Although the molecular models
can, in principle, be used to any length scale, it has been almost exclusively
applied to the microscale because of the limitation of computing capacity at
present. The continuum models present the main stream of engineering appli-
cations and are more exible when applying to different macroscale gas ows;
however, they are not suited for microscale ows. The gap between the con-
tinuum and molecular models can be bridged by the kinetic theory that is based
on the Boltzmann equation.
2. Solid Phase
The methods used for modeling pure granular ow are essentially borrowed
from that of a molecular gas. Similarly, there are two main types of models: the
continuous (Eulerian) models (Dufty, 2000) and discrete particle (Lagrangian)
models (Herrmann and Luding, 1998; Luding, 1998; Walton, 2004). The con-
tinuum models are developed for large-scale simulations, where the controlling
equations resemble the NavierStokes equations for an ordinary gas ow. The
discrete particle models (DPMs) are typically used in small-scale simulations or
FIG. 1. The various levels of modeling gas ow.
M.A. VAN DER HOEF ET AL. 68
in the investigation of the detailed physics of granular ow. A kinetic theory of
granular ow (KTGF) has also been proposed to connect the microscale picture
of granular ow to the macroscale description (Jenkins and Savage, 1983; Lun
et al., 1984).
However, a granular ow differs signicantly from a molecular gas ow. The
collisions between molecules are elastic, and the kinetic energy is conserved in
isothermal systems. For the molecular gas, there is a well-dened equilibrium
state in the absence of external energy sources, and one can dene a thermal
temperature based on the internal kinetic energy. The interaction between
macroscopic particles, however, is far more complicated. The collision between
two macroscopic particles will come with surface friction and elasticplastic
deformation, which leads to the dissipation of kinetic energy. This inelasticity
forms the primary feature of granular ow that differentiates it from a molec-
ular gas (Campbell, 1990). Clearly, without any external energy sources, a
granular system will continuously cool down, and an equilibrium state can
never be reached.
To model granular systems, DPMs using the same techniques as MD meth-
ods can be used, where it is assumed that the particle motion can be well
described by the Newtonian equations. However, in order to establish a con-
tinuum description, a number of serious difculties are encountered when one
tries to describe the elds in phase space. First, an energy source term and a
dissipative term should be included in the Boltzmann equations, which com-
plicates the (approximate) solution. Also, particle sizes may show a certain
distribution even for the same type of materials. It is well known that a differ-
ence in particle sizes will result in the segregation of granular materials (e.g., the
Brazil nut effect). Furthermore, in most granular ows the effect of gravity
cannot be ignored, which introduces an anisotropy in the velocity uctuation of
particles. Clearly, the denitions of the particle-phase pressure and other trans-
port coefcients are not straightforward because normally a homogeneous
equilibrium state does not exist. For these reasons, the construction of a reliable
hydrodynamical model for granular ow offers a great challenge for both sci-
entists and engineers (Goldhirsch, 1999).
3. The Interphase Coupling
The prime difculty of modeling two-phase gassolid ow is the interphase
coupling, which deals with the effects of gas ow on the motion of solids and
vice versa. Elgobashi (1991) proposed a classication for gassolid suspensions
based on the solid volume fraction e
s
, which is shown in Fig. 2. When the solid
volume fraction is very low, say e
s
o10
6
, the presence of particles has a neg-
ligible effect on the gas ow, but their motion is inuenced by the gas ow for
sufciently small inertia. This is called one-way coupling. In this case, the gas
ow is treated as a pure uid and the motion of particle phase is mainly con-
trolled by the hydrodynamical forces (e.g., drag force, buoyancy force, and so
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 69
on), while the particleparticle interaction is assumed to be irrelevant. With
increasing solid volume fraction up to e
s
o10
3
, the effects of the particle phase
on the gas-phase ow pattern will become important. In this region, turbulent
structures encountered in gas ows can be modied by the presence of particles.
It is commonly accepted that particleparticle interactions still do not play a
dominant role in this regime, which we normally refer to as two-way cou-
pling. For even higher solid volume fractions (e
s
410
3
), the momentum of
particles will be transported not only by the free-ight mechanism but also by
the collisions between particles and particles with the conning walls. This
means that the particleparticle interaction will be very important and four-
way coupling should be taken into account. Note that it is precisely this dense-
particle regime that is important for the industrial applications of two-phase
ows. However, a numerical model that includes the solidsolid and soliduid
interactions in full detail is not feasible for industrial-scale equipment, and for
this reason one has to resort to a multilevel approach.
C. THE MULTI LEVEL MODELING APPROACH FOR GAS SOLID FLOWS
As mentioned previously, the construction of reliable models for large-scale
gassolid contactors is seriously hindered by the lack of understanding of the
fundamentals of dense gasparticle ows (van Swaaij, 1985). In particular, the
phenomena that can be related to the effective gasparticle interaction (drag
forces), particleparticle interactions (collision forces), and particlewall inter-
action are not well understood (Kuipers and van Swaaij, 1998; Kuipers et al.,
1998). The prime difculty here is the large separation of scales: the largest ow
structures can be of the order of meters, and yet, these structures are directly
inuenced by details of the particleparticle and particlegas interactions, which
take place on the scale of millimeters, or even micrometers. As shown above, for
both the gas and particle phase, continuum-(Eulerian) and discrete-(Lag-
rangian) type of models can be applied, depending on the length scales involved.
Thus, in order to model gassolid two-phase ows at different scales, one can
FIG. 2. Interphase coupling. Based on Elgobashi, Appl. Sci. Res. (1991).
M.A. VAN DER HOEF ET AL. 70
choose appropriate combinations of the gas- and solid-phase models, where in
all cases a four-way coupling is used either directly or effectively, depending on
the scale of the simulation. The basic idea is that the smaller scale models, which
take into account the various interactions (uidparticle, particleparticle) in
detail, are used to develop closure laws that can represent the effective coarse-
grained interactions in the larger scale models. Note that it is not guaranteed
that some subtle correlations between small- and large-scale processes exist,
which cannot be captured by effective interactions. However, experience has
shown that in many cases the main characteristics of gassolid ows can be well
described by the use of closure relations. In this chapter, we discuss three levels
of modeling: the lattice Boltzmann model (LBM), the DPM and the two-uid
model (TFM) based on the KTGF. In Fig. 3, we show a schematic represen-
tation of the three models, including the information that is abstracted from the
simulations, which is incorporated in higher scale models via closure relations,
with the aid of experimental data or theoretical results. We will next give a brief
description of each of these models.
1. Two-Fluid Model
At the largest scale, a continuum description is employed for both the solid
phase and the gas phase, and a CFD-type Eulerian code is used to describe the
time evolution of the local mass and momentum density of both phases (see
Refs. Kuipers et al., 1992 and Gidaspow, 1994 amongst others). In a more
sophisticated model, based on the KTGF, also the local granular temperature of
the solid phase is a dynamical variable, and thus included in the update. With
modernday computers, the TFM model can predict the ow behavior of
gassolid ows of systems with a linear dimension of the order of 1 m, denoted
as the engineering scale, corresponding typically to the size of pilot plants,
which is in between the laboratory scale (0.1 m) and the industrial scale (10 m).
The TFM relies heavily on closure relations for the effective solid pressure and
viscosity, and gassolid drag. The basic idea of the multiscale modeling is that
FIG. 3. Multilevel modeling scheme.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 71
these relations are obtained from kinetic theory and from numerical data col-
lected in the more detailed scale models.
2. Discrete Particle Model
At one level higher in detail (and thus smaller in scale), we have the DPM.
Here the continuum description for the solid phase is replaced by a description
with discrete particles, which are modeled by spheres (Hoomans et al., 1996,
2000). The ow eld is still continuous and updated by the same methods as in
the TFM, where the scale at which the gas ow eld is described is an order of
magnitude larger than the particles (a CFD-grid cell typically contains
O(10
2
)O(10
3
) particles). The motion of the particles is governed by Newtons
law, where the forces on the particles are integrated using standard schemes for
ordinary differential equations (ODEs). These forces follow from the interac-
tion with the uid phase and collisions with the other particles. Therefore, both
a drag-force closure and a collision model have to be specied for this level of
modeling. The advantage of this model is that it can account for the parti-
clewall and particleparticle interactions in a realistic manner. This model
allows one to validate (and modify) the viscosity and pressure closures derived
from the KTGF, which are used in the TFM simulations. Still, a closure law for
the effective momentum exchange between the two phases has to be specied for
this model. The system sizes that can be studied are of the order of O(10
5
)
particles, which corresponds (for millimeter-sized particles) to systems that have
a linear dimension of the order of 0.1 m (i.e., laboratory scale).
3. Lattice Boltzmann Model
At the most detailed level of description, the gas ow eld is modeled at scales
smaller than the size of the solid particles. The interaction of the gas phase with
the solid phase is incorporated by imposing stick boundary conditions at the
surface of the solid particles. This model thus allows one to measure the effec-
tive momentum exchange between the two phases, which is a key input in all the
higher scale models. A particularly efcient method to solve the ow eld be-
tween the spheres is the LBM (Ladd, 1994; Ladd and Verberg, 2001), although
in principle other direct numerical simulation (DNS) techniques can also be
used. The number of particles in such a simulation is typically around 500,
which is sufciently high to account for swarm effects. The goal of these sim-
ulations is to construct drag laws for dense gassolid systems. For low Reynolds
numbers (Re), the functional form of the drag law can be derived from theory
using the CarmanKozeny approximation, where the simulation data is then
used to determine the unknown parameters such as the Kozeny constant. For
higher Reynolds numbers, a theoretical evaluation of the functional form is not
possible and the drag law is simply constructed as the best possible t to the
simulation data, where the functional form is dictated by a compromise between
M.A. VAN DER HOEF ET AL. 72
simplicity and accuracy. For both low and high Reynolds numbers, the drag
laws are validated (and possibly adjusted) on the basis of pressure-drop data.
A graphical representation of the multilevel approach is shown in Fig. 4. All
three models are now commonly accepted and are widely used by a number of
research groups (both academic and industrial) around the world. In a recent
paper, we have given an overview of the three models as they are employed at
the University of Twente, together with some illustrative examples (Van der
Hoef et al., 2004). In this chapter, we will focus on the technical details of each
of the models, much of which has not been published elsewhere. The devel-
opment of detailed closure relations from the simulations, as indicated in Fig. 3,
is still ongoing. Some preliminary results for both the drag-force closures and
solid pressure will be presented in the Sections II and III. In this chapter, we will
FIG. 4. Graphical representation of the multilevel modeling scheme. The arrows represent a
change of model. On the left is a uidized bed on a life-size scale, a section of which is modeled by
the two-uid model (TFM) (see enlargement), where the shade of grey of a cell indicates the solid-
phase volume fraction. On the right, the same section is modeled using discrete particles. The gas
phase is solved on the same grid as in the TFM. The bottom graph shows the most detailed level,
where the gas phase is solved on a grid much smaller than the size of the particles. Note that in
reality, the separation in scales is much more extreme, and also that the section that can be modeled
by the TFM of the industrial-scale uidized bed is much smaller.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 73
only consider monodisperse systems, but nevertheless we will formulate the soft-
sphere model in Section III.B for general polydisperse systems where a particle a
has an individual radius R
a
. Apart from Section III.B, the size of the particles
are indicated by a single radius R or a diameter d.
Finally, we note that the TFM can simulate uidized beds only at engineering
scales corresponding typically to the size of pilot plants, and the industrial-scale
uidized bed reactors (diameter 15 m, height 320 m) are still far beyond its
capabilities. In Section V, we discuss two promising approaches for modeling
large-scale gassolid ow, which are currently being explored independently at
the Princeton University and the University of Twente. We stress, however, that
these approaches are still under development and that they should be recognized
as only preliminary.
II. Lattice Boltzmann Model
As mentioned in Section I, there are two fundamentally different types of
models to describe a gas ow: the continuum models and the molecular models.
In principle, the molecular models can be applied at any length scale; however, in
practice this is limited to microscopic scales only because of the limitation of
computer time. The continuum models present the main stream of engineering
applications and are more exible when applying to different macroscale gas
ows. The gap between the continuum and molecular models can be bridged,
however, by the lattice Boltzmann (LB) simulation model that applies at a me-
soscopic scale, which is in between fully microscopic and macroscopic scales.
The LB model that is currently the most widely usedthe lattice Bhatn-
agarGrossKrook (BGK) modelis nothing but a nite difference version of
the continuous, macroscopic BGK equation introduced in 1954 (Bhatnagar et
al., 1954). Historically, however, this LB model has evolved from the microscopic
lattice-gas simulation models for uids, and we will also follow this route here.
A. FROM LATTICE-GAS TO LATTICE-BOLTZMANN
1. Lattice-Gas Models
As mentioned earlier, in principle, one can model the dynamics of a simple
classical uid by means of MD simulations. This technique, although straight-
forward, is relatively time-consuming, and therefore not suitable for observa-
tion of large-scale macroscopic phenomena in the uid. However, one often
does not need such a detailed description of the microdynamics as provided by
MD. In such cases, it would be more efcient to strip the MD model down to its
barest essentials, where the only requirement is that the model behaves like a
uid macroscopically, but is still atomistic in characteri.e., the mechanism
underlying the uid motion is the movement of particles. From the derivation of
M.A. VAN DER HOEF ET AL. 74
the uid dynamics equations it is clear that a key ingredient in such a model
must be local conservation of mass and momentum. The simplest model to
think of would be one with a single species of particles moving on a lattice with
discrete velocities. In 1985, such a model was introduced for two-dimensional
(2D) uid ow by Frisch et al. (1986), where particles move on a triangular
lattice. Such a lattice has just enough symmetry to guarantee isotropic, mac-
roscopic equations of motion. The rules and basic idea of the FHP model are
illustrated in Fig. 5. Later, the model has been extended to three dimensions as
well. From the update rules it is clear that lattice-gas cellular automata (LGCA)
make an efcient simulation scheme, in particular on a parallel computer, since
the rules are completely local. Moreover, stability of the algorithm is guaran-
teed, since the update involves only bit manipulation, i.e., the update is exact
with no round-off errors. We will continue with a more formal description of
general LGCA models.
2. Denitions and Equation of Motion
In LGCA models, time and space are discrete; this means that the model
system is dened on a lattice and the state of the automaton is only dened at
regular points in time with separation dt. The distance between nearest-neighbor
sites in the lattice is denoted by dl. At discrete times, particles with mass m are
situated at the lattice sites with b possible velocities c
i
, where i A {1, 2, y, b}.
The set c
i
can be chosen in many different ways, although they are restricted by
the constraint that
r
0
r c
i
dt (1)
FIG. 5. Example of the time evolution in a small section of the FHP model. In the gure, each
arrow represents the velocity of a single particle. The particles are situated at the lattice sites. The
update of the lattice consists of two steps. First there are local collisions at all sites, simultaneously,
and such that locally the number of particles and momentum is conserved. Note that only some
cases lead to a new conguration. The next step is a propagation step: all the particles move
simultaneously according to their velocities to neighboring sites. Particles do not interact during this
step. Note that the gure only represents a small section of the lattice; therefore, we can only give the
complete nal conguration of the central site, as the state of the other sites after the propagation
depends on neighboring sites that are not shown. Next there will again be a collision step, etc.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 75
where r and r
0
must be lattice sites. Apart from Eq. (1), it has been proved
essential to have additional symmetry requirements on the velocity set in order
to get isotropic, macroscopic behavior from these models. These requirements
turn out to be that the even-rank tensors that can be constructed from the
velocity set are isotropic up to 4th rank, and the odd-rank tensors are zero. The
time evolution of the LGCA consists of two steps:
1. Propagation: All particles move in one time step dt from their initial lattice
position r to a new lattice position r
0
r+c
i
dt.
2. Collision: The particles at all lattice sites undergo a collision that conserves
the total number of particles and the total momentum at each site. The
collision rules may or may not be deterministic.
The state of the automaton at time t can be completely determined by the
boolean variable n
i
(r,t), which is equal to 1(0) if a particle is present (absent)
on site r with velocity c
i
. From this it follows that the local microscopic density ~ r
and ow velocity ~ u at site r are given by
~ pr; t m

i
n
i
r; t; ~ pr; t ~ ur; t m

i
n
i
r; tc
i
(2)
with m as the mass of the particles. The update of n
i
(r,t) (from propagation and
collision) can formally be written by the following equation of motion:
n
i
r c
i
dt; t dt n
i
r; t D
i
nr; t (3)
with
D
i
nr; t

s;s
0
s
0
i
s
i
xs; s
0
P
j
n
sj
j
1 n
j

1sj
With the sum is over all possible states s and s
0
of a single site, and x(s, s
0
) is a
collision function that is equal to one for states s, s
0
where s goes over into s
0
in a
collision, and zero for all other possible pairs of states. Note that expression
in Eq. (3) is the formal expression for the update. In a numerical code, the state
n
i
(r,t) of the systems is represented by a bbit word for every site r at time t. The
collision process can then be done by a very quick table lookup; whereas, for
the propagation step the bits from one word have to be put at the same bit
positions in the words describing the states of the neighboring sites. Despite its
extremely simplied microdynamical behavior, it turns out that the analogy of
these models with the real uid models is very close, such that the theoretical
framework of statistical mechanics of simple uids can be applied, to a great
extent, to these discrete uids. That is, starting from the formal expression in Eq.
(3), it can be proven that the macroscopic equations of motions are, in a well-
dened limit, equivalent to the NavierStokes equations (Frisch et al., 1987; Ernst
and Dufty, 1989). This solid theoretical basis makes the LGCA method not just a
toy model of computer scientists but also a numerical scheme that can be seriously
M.A. VAN DER HOEF ET AL. 76
considered for the study of hydrodynamic ow phenomena. However, such
an application is seriously hindered by two big drawbacks of LGCA. Firstly,
their inherent noisiness, which means that massive averaging is required to get
accurate numbers. And secondly, it turns out that the method is not suitable for
modeling uid ow at Reynolds numbers above Re 100, which is related to the
fact that the viscosity cannot be made lower than a certain value, since it is
dictated by the collision function x. It is for these reasons that the current class of
LGCA methods cannot compete with CFD methods for modeling large-scale
uid ow.
3. Averaged Equation of Motion
The two drawbacks mentioned above can be overcome, however, by consid-
ering the ensemble-averaged version of the microscopic equation of motion,
Eq. (3):
f
i
r c
i
dt; t dt f
i
r; t C
i
f r; t (4)
with f
i
(r,t) /n
i
(r,t)Sthe average occupation number of link i at site r and
time t, which is now a oating number between zero and one, and
C
i
f hD
i
ni

s;s
0
s
0
i
s
i
xs; s
0

j
f
s
j
j
1 f
j

1s
j
(5)
where in the second step the assumption is made that the particle occupation
numbers on a single site are not correlated, so that the average of the product
can be written as the product of the average. The ensemble averaged density r
and ow velocity u follow from Eq. (2):
r h ~ ri m

i
f
i
; ru h ~ r~ ui m

i
f
i
c
i
(6)
where we have omitted the space and time dependence. In its present form, the
collision operator in Eq. (5) is not very useful for simulations, since the update
of f
i
at each site requires the double sum over all possible states, where there are
over 16 million states (2
24
) for the 3D models. This problem can be circum-
vented by expanding the collision function about the equilibrium distribution
function f
i
eq
, for which it holds that C(f
i
eq
) 0:
f
i
r c
i
dt; t dt f
i
r; t

b
j1
L
ij
f
j
r; t f
eq
j
r; t (7)
where L is the linearized collision operator:
L
ij

@C
i
@f
j
_ _
f
j
f
eq
j
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 77
which can be evaluated directly from Eq. (5). Note that L, which has to be
calculated only once for a given set of collision rules, is now a small b b matrix,
compared to x that is a 2
b
2
b
matrix. Equation (7) can be directly converted
into an algorithm for simulation purpose. The advantage of an LB simulation is
that the system is essentially free of noise. Also, the linearized collision operator
need not necessarily be evaluated from an existing set of microscopic collision
rules x(s, s
0
). One is free to dene any operator L, which has the correct sym-
metry and conserves momentum and number of particles. As an example, for
the 2D hexagonal lattice, one can derive from the requirements of symmetry and
conservation that L should have the following form:
L
a b c d c b
b a b c d c
c b a b c d
d c b a b c
c d c b a b
b c d c b a
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
(8)
with
a
4
21

1
3
l; b
1
7

1
6
l; c
4
21

1
6
l; d
1
7

1
3
l
where l can take any value between 0 and 2 and is related to the kinematic
viscosity via
m
1
4l

1
8
_ _
r
dl
2
dt
This gives the possibility to make the viscosity arbitrarily small, so that sim-
ulations can be performed also at Reynolds numbers higher than 100.
B. THE LATTICE BHATNAGARGROSSKROOK MODEL
In the linearized LB equation Eq. (7), the ensemble averaged effect of the
particleparticle collision is now represented by a relaxation of the distribution
function f
i
to the equilibrium function f
i
eq
, where the matrix L
ij
does not nec-
essarily have to correspond to an existing set of collision rules. The question
now arises if L can be simplied even further to the form L
ij
ad
ij
, so that the
LB equation takes the form (with t dt/a)
f
i
r c
i
dt; t dt f
i
r; t
dt
t
f
i
r; t f
eq
i
r; t (9)
M.A. VAN DER HOEF ET AL. 78
At rst sight this does not seem possible because the specic form of L, for
instance Eq. (8), which followed from the requirement of symmetry. However, it
turns out that b, c, and d can be put to zero, if an equilibrium function f
i
eq
with
different weights is used for the different directions, so that the lack of symmetry
can be remedied (Qian et al., 1992; Ladd, 1994; Succi, 2001). Specically, the f
i
eq
should take the form
f
eq
i
a
ci
r 1
c
i
u
c
2
s

c
i
u
2
2c
4
s

u
2
2c
2
s
_ _
(10)
where the weight a
c
i
only depends on the magnitude c
i
of the velocity c
i
con-
nected to the link direction i, and c
s
is the speed of sound. For the popular
D3Q19 model (3D, 19 velocities), there are 6 velocities with c
i
1dl/dt, 12
velocities with c
i

2
p
dl=dt; and one zero velocity, c
i
0 (see Fig. 6). The
parameters that yield the proper equilibrium distribution in Eq. (10) are a
0

1/3, a
1
1/36, and a

2
p
1=18, and the speed of sound is usually set to c
s

1/3(dl/dt) (Ladd and Verberg, 2001; Succi, 2001).
To rst order, expression in Eq. (9) represents the nite difference form of the
well-known BGK equation:
@
@t
r c
i
_ _
f
i

1
t
f
i
f
eq
i
_
(11)
FIG. 6. The D3Q19 model.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 79
It can be shown that this equation will yield the familiar conservation of mass
and momentum equations
@
@t
r r ru 0;
@
@t
ru r ruu r p (12)
where the pressure tensor is equal to
p p

I m ru ru
T
_
l
2
3
m
_ _
r u
s


I
_
with

I the unit tensor, (ru)
ab
r
a
u
b
, ru
T
ab
r
b
u
a
, and the pressure p, shear
viscosity m, and bulk viscosity l given by
p c
2
s
r; m tc
2
s
r; l 0 (13)
As said, the lattice BGK in Eq. (9) is the nite difference version of Eq. (11) to
rst order in dt. To second order in dt, however, Eq. (9) represents the nite
difference version of a slightly different expression:
@
@t
r c
i
_ _
f
i

1
t
f
i
f
eq
i
_

dt
2t
@
@t
r c
i
_ _
f
i
f
eq
i
_
(14)
In the route to NavierStokes, it turns out that f
i
eq
in the second term on the
right-hand side (RHS) does not play a role, so that we can rewrite Eq. (14) as a
normal BKG equation with a different prefactor on the RHS:
@
@t
r c
i
_ _
f
i

dt
t dt=2
f
i
f
eq
i
_
We thus nd that the lattice BGK model describes, to second order in dt, the
uid according to the NavierStokes equation with a viscosity
m t
1
2
dt
_ _
c
2
s
r
where the extra term (1/2)dt is due to the nite difference scheme. As a dem-
onstration of how well the simple lattice BGK scheme in Eq. (9) can describe
uid ow, we show in Fig. 7 the velocity prole from a lattice BGK simulation
for forced Poiseuille ow and shear ow; it can be seen that excellent agreement
is found with the theoretical results that follow from the NavierStokes equa-
tion. Note that for the simulations shown in Fig. 7 we used stick-boundary
M.A. VAN DER HOEF ET AL. 80
conditions at the walls of the channel. The implementation of such conditions is
similar to the boundary conditions that are required to model large, solid ob-
jects in the LB model, which is described in Section II.C.
C. MODELING SOLID PARTICLES
In order to simulate large, moving particles in the LB model, we should dene
additional rules that describe the interaction of the LB gas with the surface of
the solid particles. One essential ingredient of the moving-boundary rules is that
these rules result, on average, in a dissipative force on the suspended particle.
An obvious choice of rules is those according to which the gas next to the solid
particle moves with the local velocity of the surface of the solid particle. In this
way, one models the hydrodynamic stick boundary condition; for a spherical
particle suspended in an innite 3D system, moving with velocity v, this will give
rise to a frictional force on the particle F 3pmdv, at least in the limit of low
Reynolds numbers Re rdv/m, where d is the hydrodynamic diameter of the
particle and m is the shear viscosity. A particular efcient and simple way to
enforce stick-boundary rules was introduced by Ladd (1994). First the bound-
ary nodes are identied, which are dened as the points halfway two lattice sites
that are inside and outside the particle and closest to the actual surface (see
Fig. 8, left graph, solid squares). For a static particle, the boundary rule is
simply that a distribution moving such that it would cross the boundary is
bounced back at the boundary node. Since this node is halfway the link, the
bounce-back rule has the effect so (see Fig. 8):
f
i
r; t dt f
i
0 r; t (15)
Forced Poiseuille flow
Lattice Boltzmann
Theory
Fluid velocity
Shear flow
Lattice Boltzmann
Theory
Fluid velocity
v
0
FIG. 7. Velocity prole for Poiseuille ow and shear ow. The points are the LBM data, and the
solid lines are the theoretical proles. For the simulations, we used the D3Q18 model with 25 lattice
sites across the channel.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 81
where i and i
0
are opposite links. This rule ensures that the uid velocity at the
boundary node indeed vanishes; the momentum at the boundary node at time
t+(1/2)dt is given by
r
b
u
b
f
i
0 r; t c
i
0
f
i
r; t dt c
i
(16)
Inserting Eq. (15) and using c
i
c
i
0
gives r
b
u
b
0. For nonstatic particles, the
local uid velocity must be set equal to the local boundary velocity v
b
. This can
be achieved by a simple modication of the bounce-back rule:
f
i
r; t dt f
i
0 r; t a c
i
(17)
where a is chosen such that u
b
v
b
. Note that only the component of v
b
in the
direction of the link can be set in this way. For more details we refer to the
papers by Ladd (1994) and Ladd and Verberg (2001). In Fig. 9 (right graph), we
show the LBM simulation results for the velocity of a single free-falling sphere
in an (effectively) unbounded uid. As can be seen from Fig. 9, the boundary
rule results in a terminal velocity according to the StokesEinstein friction force.
Note that the actual plateau value of the velocity is slightly smaller than the
theoretical prediction. This can be attributed to the fact that the radius of the
particle is not well dened because of the irregular shape of boundarynode
surface of the sphere. In fact, the free-falling sphere experiment (or a similar
experiment with periodic boundary conditions) is used for calibration purposes,
i
i
i
i
site
boundary
r
site
s
node
FIG. 8. Left graph: example of a boundary node map for a disc in a 2D hexagonal lattice. Right
graph: illustration of the bounce-back rule on an enlarged section of the boundary. The distribution
at site r that moves at time t into direction i
0
, instead of arriving at the (virtual) site s, is bounced at
the boundary node, and thus arrives back at site r at time t+dt, but is now headed in the opposite
direction i.
M.A. VAN DER HOEF ET AL. 82
i.e., the effective hydrodynamic radius of the spheres is obtained from the ter-
minal velocity, where it is assumed that the StokesEinstein relation holds.
D. RESULTS FOR THE GAS SOLID DRAG FORCE
The drag force from the gas phase on an assembly of spheres can, in principle,
be obtained from the terminal velocity in a sedimentation experiment. However,
the drag force can also be directly measured in the simulation from the change
in gas momentum due to the boundary rules. The change in gas momentum per
unit time, required to maintain stick-boundary conditions for particle a, is equal
to minus the total force F
g-s,a
that the gas phase exerts on particle a. This total
force has two contributions (see also Section III.D), namely the drag force F
d,a
due the uidsolid friction at the surface of the spheres and a force
F
p,a
V
a
rp due to the static pressure gradient rp, which drives the gas ow
past the spheres (V
a
pd
3
/6 is the volume of the sphere). From a balance of
forces it follows that Vrp

a
F
g-s,a
, with V the total volume of the system;
eliminating rp from the expressions gives F
d,a
eF
g-s,a
with e the volume
fraction of the gas phase. The procedure to obtain the drag force for mono-
disperse systems in an LB simulation is then as follows. N particles with di-
ameter d are distributed randomly in a box of n
x
n
y
n
z
lattice sites, so that
the gas fraction equals e 1Npd
3
/(6n
x
n
y
n
z
dl
3
). Some typical values are
N 54 and d 25dl, where periodic boundary conditions are used. All spheres
are forced to move with the same constant velocity v
sim
in some arbitrary di-
rection, so that the array of spheres moves as a static conguration through the
system. A uniform force is applied to the gas phase, to balance the total force
0 1000 2000 3000 4000 5000
time
0.000
0.002
0.004
0.006
P
a
r
t
i
c
l
e

v
e
l
o
c
i
t
y

Sedimentation of a single sphere
Simulation
v
term
= mg/3d
FIG. 9. Velocity of a single sphere in a 3D LB gas. The black line is the data from LBM, which has
the proper functional form v(t) v
N
(1 exp(-gt/v
N
)). The grey line is theoretical terminal velocity,
which is slightly higher than v
N
.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 83

N
a1
F
g!s;a
from the moving particles on the gas phase. From this it follows
that in a frame of reference where the particles are static, the supercial ow
velocity is equal to v
sim
, so that Re pd|v
sim
|/m, where m is the viscosity. Once
an equilibrium state is obtained, the average value

F
g!s
h

N
a1
F
g!s;a
=Ni
t
is
determined, with / S
t
a time average. The momentum exchange coefcient b,
as dened in Eq. (44), is then determined via
b

F
g!s
1
2
v
sim
V
a
since the relative velocity uv
a
in Eq. (44) corresponds to v
sim
/e in the LB
simulations and F
d,a
eF
g-s,a
. Note that the dimensionless quantity bd
2
/m will
only depend on the Reynolds number and the gas fraction e. Ergun (1952)
showed that when the experimental data for b for different e and Re is plotted in
a single {log(x), log(y)} graph with
x
Re
1
y
bd
2

m1
1
Re
all data fall onto a single curve y 150/x+1.75, which corresponds to
bd
2

m1
1501 1:75Re (18)
which is the famous Ergun equation. In Fig. 10, we show the data from ex-
tensive LB simulations (Van der Hoef et al., 2004; Beetstra et al., 2006) for a
range of gas fractions and Reynolds numbers, on the same {log(x), log(y)}
graph. We nd that our LBM data deviates substantially from the Ergun
equation: for low Re numbers the Ergun equation underestimates the drag
force, whereas for high Re numbers the Ergun equation overestimates the drag
force. A simple remedy would be to use different coefcients in the Ergun
equation, for instance 180 instead of 150 and 1.0 instead of 1.75. However, it
can be seen from Fig. 10 that not all data obey the functional form y A/x+B.
Note also that the Ergun equation was derived for packed beds, and is not
expected to be valid for high gas fractions; for that range, normally the Wen and
Yu Eq. (46) is used. However, we nd that this equation signicantly under-
predicts the drag force at higher Reynolds numbers (see Fig. 10). Based on our
data from the LB simulations, we suggest the following new drag-force cor-
relation that we write in the form of an Ergun-type equation (Beetstra et al.,
2006; Van der Hoef et al., 2005):
bd
2

m 1
A1 BRe (19)
M.A. VAN DER HOEF ET AL. 84
only with coefcients that depend on both e and Re:
A 180
18
4
1
1 1:5

1
p
_ _
and
B
0:31
1
31 8:4Re
0:343

1 10
3 1
Re
22:5
Expression in Eq. (19) is within 8% of all simulation data up to Re 1000.
Since this relation has been derived very recently (Beetstra et al., 2006), it has
not been applied yet in the higher scale models discussed in Sections III and IV.
However, the expression by Hill et al. in Eq. (47) derived from similar type of
LBM simulations is consistent with our data, in particular when compared to
the large deviations with the Ergun and Wen and Yu equations. So, we expect
that the simulation results presented in Section IV.F using the Hill et al.
correlation will not be very different from the results that would be obtained
with expression in Eq. (19). A more detailed account of the derivation of
expression in Eq. (19) and a comparison with other drag-force relations can be
found in Ref. Beetstra et al. (2006).
10
1
10
1
10
3
10
5
Re / (1 )
10
0
10
1
10
2
10
3
(

d
2
/

/
(
(
1

)
R
e
)
Ergun
WenYu (=0.8)
Hill et al (=0.5)
Equation (19) (=0.5)
LBM (=0.40.8)
FIG. 10. Normalized drag force at arbitrary Reynolds numbers and gas fractions. The symbols
represent the simulation data, the solid line the Ergun correlation Eq. (18), the dashed line the
WenYu correlation Eq. (46) for e 0.8, and the grey line the correlation by Hill et al. (2001a,b) Eq.
(47) and the long-dashed line Eq. (19), both for e 0.5.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 85
III. Discrete Particle Model
A. INTRODUCTION
DPMs offer a viable tool to study the macroscopic behavior of assemblies of
particles and originate from MD methods. Initiated in the 1950s by Alder and
Wainwright (1957), MD is by now a well-developed method with thousands of
papers published in the open literature on just the technical and numerical
aspects. A thorough discussion of MD techniques can be found in the book by
Allen and Tildesley (1990), where the details of both numerical algorithms and
computational tricks are presented. Also, Frenkel and Smit (1996) provide a
comprehensive introduction to the recipes of classical MD with emphasis on
the physics underlying these methods. Nearly all techniques developed for MD
can be directly applied to discrete particles models, except the formulation of
particleparticle interactions. Based on the mechanism of particleparticle in-
teraction, a granular system may be modeled either as hard-spheres or as
soft-spheres.
1. Hard-Sphere Model
In a hard-sphere system, the trajectories of particles are determined by mo-
mentum conserving binary collisions. The interactions between particles are
assumed to be pair-wise additive and instantaneous. In the simulation, the col-
lisions are processed one by one according to the order in which the events
occur. For not too dense systems, the hard-sphere models are considerably
faster than the soft-sphere models. Note that the occurrence of multiple col-
lisions at the same instant cannot be taken into account.
Campbell and Brennen (1985) reported the rst hard-sphere discrete particle
simulation used to study granular systems. Since then, the hard-sphere models
have been applied to study a wide range of complex granular systems. Hoo-
mans et al. (1996) used the hard-sphere model, in combination with a CFD
approach for the gas-phase conservation equations, to study gassolid two-
phase ows in gas-uidized beds. By using this model, they studied the effect of
particleparticle interaction on bubble formation (Hoomans et al., 1996) and
the segregation induced by particle-size differences and density differences
(Hoomans et al., 2000). This model has been further used in connection with
the kinetic theory of granular dynamics by Goldschmidt et al. (2001), high-
pressure uidization by Li and Kuipers (2002), and circulating uidized beds by
Hoomans (2000).
Similar simulations have been carried out by other research groups. Ouyang
and Li (1999) developed a slightly different version of this model. Helland et al.
(1999) recently developed a DPM in which hard-sphere collisions are assumed,
but where a time-driven scheme (typically found in the soft-sphere model) is
used to locate the collisional particle pair. Effect of the gas turbulence has also
M.A. VAN DER HOEF ET AL. 86
been taken into account in some hard-sphere models by Helland et al. (2000),
Lun (2000), and Zhou et al. (2004).
At high-particle number densities or low coefcients of normal restitution e,
the collisions will lead to a dramatical decrease in kinetic energy. This is the so-
called inelastic collapse (McNamara and Young, 1992), in which regime the
collision frequencies diverge as relative velocities vanish. Clearly in that case, the
hard-sphere method becomes useless.
2. Soft-Sphere Model
In more complex situations, the particles may interact via short- or long-
range forces, and the trajectories are determined by integrating Newtonian
equations of motion. The soft-sphere method originally developed by Cundall
and Strack (1979) was the rst granular dynamics simulation technique
published in the open literature. In soft-sphere models, the particles are
allowed to overlap slightly and the contact forces are subsequently calculated
from the deformation history of the contact using a contact-force scheme. The
soft-sphere models allow for multiple particle overlap, although the net contact
force is obtained from the addition of all pair-wise interactions. The soft-sphere
models are essentially time-driven, where the time step should be carefully cho-
sen in calculating the contact force. The soft-sphere models that can be found in
literature mainly differ from each other with respect to the contact-force scheme
that is used. A review of various popular schemes for repulsive interparticle
forces is presented by Scha fer et al. (1996). Walton and Braun (1986) developed
a model that uses two different spring constants to model the energy dissipation
in the normal and tangential directions. In the force scheme proposed by
Langston et al. (1994), a continuous potential of an exponential form is used,
which contains two unknown parameters: the stiffness of the interaction and an
interaction constant.
A 2D soft-sphere approach was rst applied to gas-uidized beds by Tsuji
et al. (1993), where the linear springdashpot modelsimilar to the one
presented by Cundall and Strack (1979)was employed. Xu and Yu (1997) in-
dependently developed a 2D model of a gas-uidized bed. However in their sim-
ulations, a collision detection algorithm that is normally found in hard-sphere
simulations was used to determine the rst instant of contact precisely. Based on
the model developed by Tsuji et al. (1993), Iwadate and Horio (1998) incorporated
van der Waals forces to simulate uidization of cohesive particles. Kafui et al.
(2002) developed a DPM based on the theory of contact mechanics, thereby
enabling the collision of the particles to be directly specied in terms of material
properties such as friction, elasticity, elasto-plasticity, and auto-adhesion.
It is also interesting to note that soft-sphere models have also been applied to
other applications such as gasparticle heat transfer by Li and Mason (2000)
and coal combustion by Zhou et al. (2003). Clearly, these methods open a new
way to study difcult problems in uidized bed reactors.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 87
3. Comparison between Hard- and Soft-Sphere Models
Although both hard- and soft-sphere models have been used in the simulation
of granular ow, each has its own characteristics that make them very efcient
in some cases, while inefcient in others. The two types of models are compared
in Table I. Hard-sphere models use an event-driven scheme because the inter-
action times are (assumed to be) small compared to the free-ight time of
particles, where the progression in physical time depends on the number of
collisions that occur. In contrast, in the soft-sphere models a time step that is
signicantly shorter than the contact time should be used. This directly implies
that the computational efciency of the soft-sphere model (compared to the
hard-sphere model) decreases when the ratio of the free-ight time to the con-
tact time increases, which is the case when the system becomes less dense. In the
soft-sphere models, a slight deformation of particles is allowed, so that multiple
contacts between several pairs of particles are possible, which should never
happen in the event-driven models. As mentioned above, a lower coefcient of
normal restitution may lead to the inelastic collapse in hard-sphere simulations.
Incorporation of cohesive forces, especially the pair-wise forces, is quite
straightforward in soft-sphere models. This is because the collisional process in
the soft-sphere model is described via the Newtonian equations of motion of
individual particles, that is, in terms of forces. In the hard-sphere system, the
update is not via forces (since they are, loosely speaking, either zero or innite),
but via a momentum exchange at contact. This means that for short-range
forces, such as the cohesive force, a kind of hybrid method for the interaction at
close encounters has to be devised, which is not straightforward. In contrast, for
systems with different size particles, it is the soft-sphere model that poses some
difculties. In a soft-sphere system using a linear springdashpot scheme, the
spring stiffness is dependent on the particle size. This means that in principle a
different spring stiffness should be used for calculating the contact forces be-
tween particles with different sizes, otherwise the computing efciency will drop
substantially.
TABLE I
COMPARISON BETWEEN HARD- AND SOFT-SPHERE MODELS. THE SYMBOLS INDICATE GOOD (++),
NORMAL (+), AND NOT SUITABLE ()
Hard-sphere Soft-sphere
Computing efciency ++ +
Multiple contacts ++
Dense systems ++
Incorporation of cohesive force + ++
Energy conservation during collisions ++ +
Multiple particle sizes ++ +
M.A. VAN DER HOEF ET AL. 88
In the following, we focus on the soft-sphere method since this really is the
workhorse of the DPMs. The reason is that it can in principle handle any
situation (dense regimes, multiple contacts), and also additional interaction
forcessuch as van der Waals or electrostatic forcesare easily incorporated.
The main drawback is that it can be less efcient than the hard-sphere model.
B. PARTICLE DYNAMICS: THE SOFT-SPHERE MODEL
1. The Equations of Motion
The linear motion of a single spherical particle a with mass m
a
and coordinate
r
a
can be described by Newtons equation:
m
a
d
2
r
a
dt
2
F
contact;a
F
pp;a
F
ext;a
(20)
where the RHS is the total force on the particle, which has three basic con-
tributions:
(i) The total contact force F
contact,a
is the sum of the individual contact forces
exerted by all other particles in contact with the particle a, which are nat-
urally divided into a normal and a tangential component:
F
contact;a

b2contactlist
F
ab;n
F
ab;t

(ii) The total external force F


ext
,
a
:
F
ext;a
F
g;a
F
d;a
F
p;a
which includes the gravitational force F
g,a
m
a
g, and forces exerted by the
surrounding gas phase: the drag force F
d,a
and a force F
p,a
from the pres-
sure gradient.
(iii) The sum of all other particleparticle forces F
pp,a
that can include short-
range cohesive forces F
coh,a
, which follow from the van der Waals inter-
action between the molecules that the particles are made up of, as well as
long-range electrostatic forces. In this chapter, we will only consider the
cohesive forces.
Note that for liquidsolid systems, Eq. (20) should also include the short-
range lubrication forces and the effects of other forces such as the virtual
mass force. But this is beyond the scope of this chapter.
Finally, the rotational motion of particle a is given by
I
a
do
a
dt
T
a
(21)
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 89
where I
a
is moment of inertia, o
a
the angular velocity, and T
a
the torque, which
depends only on the tangential component of the individual contact forces:
T
a

b2contactlist
R
a
n
ab
F
ab;t

with R
a
as the radius of particle a.
We will next give a more detailed description of the contact force, the co-
hesive force, and the integration of the equations of motionEqs. (20) and (21).
The description of the forces resulting from interaction with the gas phase is
given in Section III.D, whereas the dynamics of the gas phase itself is described
in Section III.C.
2. Contact Force
The calculation of the contact force between two particles is actually quite
involved. A detailed model for accurately computing contact forces involves
complicated contact mechanics (Johnson, 1985), the implementation of which is
extremely cumbersome. Many simplied models have therefore been proposed,
which use an approximate formulation of the interparticle contact force. The
simplest one was originally proposed by Cundall and Strack (1979), where a
linear-spring and dashpot model is employed to calculate the contact forces (see
Fig. 11 and 12). In this model, the normal component of the contact force
between two particles a and b can be calculated by
F
ab;n
k
n
d
n
n
ab
Z
n
v
ab;n
(22)
where k
n
is the normal spring stiffness, n
ab
the normal unit vector, Z
n
the normal
damping coefcient, and v
ab,n
the normal relative velocity. The overlap d
n
is
given by
d
n
R
b
R
a
r
b
r
a
j j
FIG. 11. Graphical representation of the linear springdashpot soft-sphere model. From Hoo-
mans, Ph.D. thesis, University of Twente (2000).
M.A. VAN DER HOEF ET AL. 90
with R
a
and R
b
denoting the radii of the particles. The normal unit vector is
dened as
n
ab

r
b
r
a
r
b
r
a
j j
(23)
The relative velocity of particles a and b is
v
ab
v
a
v
b
R
a
o
a
R
b
o
b
n
ab
(24)
where v
a
and v
b
are the particle velocities, and o
a
and o
b
the angular velocities.
The normal component of the relative velocity between particle a and b is
v
ab;n
v
ab
n
ab
n
ab
(25)
For the tangential component of the contact force, a Coulomb-type friction law
is used:
F
ab;t

k
t
d
t
Z
t
v
ab;t
for F
ab;t

m
f
F
ab;n

m
f
F
ab;n

t
ab
for F
ab;t

4m
f
F
ab;n

_
(26)
where k
t
, d
t
, Z
t
, and m
f
are the tangential spring stiffness, tangential displace-
ment, tangential damping coefcient, and friction coefcient, respectively. In
FIG. 12. The coordinate system used in the soft-sphere model.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 91
Eq. (26), the tangential relative velocity v
ab,t
, and tangential unit vector t
ab
are
dened as
v
ab;t
v
ab
v
ab;n
t
ab

v
ab;t
v
ab;t

The calculation of the tangential displacement d


t
requires some special attention
and will be addresses in Section III.B.3.
3. Tangential Displacement
Suppose that the tangential displacement at t
o
is equal to d
t
0
; then one would
expect that the displacement d
t
at time t follows by simply integrating the
tangential velocity (Hoomans, 2000):
d
t
d
t
0

_
t
t
0
v
ab;t
dt
This expression, however, is only justied for 2D systems, where the particles
are represented essentially by disks, which are conned in a single plane and the
particleparticle contact occurs along a line, as shown in Fig. 13. So, the tan-
gential component of the relative velocity is always in the same plane and no
coordinate transformation is required.
In a 3D system, however, it becomes more complicated. The particleparticle
contact now occurs in a plane. The tangential component of the relative velocity
is always in this plane and vertical to the normal unit vector according to the
denition. Since the normal unit vector is not necessarily situated in the same
plane at any time, it is desirable to transfer the old tangential displacement to
the new contact plane before we calculate the new tangential displacement. To
this end, a 3D rotation of the old tangential displacement should be applied. As
FIG. 13. The rotation of the contact plane during particleparticle collisions.
M.A. VAN DER HOEF ET AL. 92
the tangential velocity vector is always vertical to the normal unit vector, the 3D
rotation can be done around the vector determined by n
ab
n
0,ab
, as shown in
Fig. 14. So in a 3D situation, the tangential displacement is determined by
d
t
d
t
0
H
_
t
t
0
v
ab;t
dt (27)
where the rotation matrix is
H
qh
2
x
c qh
x
h
y
sh
z
qh
x
h
z
sh
y
qh
x
h
y
sh
z
qh
2
y
c qh
y
h
z
sh
x
qh
x
h
z
sh
y
qh
y
h
z
sh
x
qh
2
z
c
_
_
_
_
_
_
_
_
(28)
with h, c, s, and q are dened as
h
n
ab
n
0;ab
n
ab
n
0;ab

; c cos j; s sin j; q 1 c
and
j arcsin n
ab
n
0;ab

_ _
Are Eqs. (27) and (28) sufcient to describe the tangential displacement dur-
ing particleparticle contact? In the absence of friction, the answer is yes. When
we consider friction during particleparticle contactas pointed out by Brendel
FIG. 14. The transformation of tangential displacement vector.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 93
and Dippel (1998)the use of Eqs. (27) and (28) may give rise to unphysical
behavior for dense systems due to the allowance of an arbitrarily large tangen-
tial displacement (Eq. 28). In a dilute system, this will not be a problem since the
multiple-particle contacts do not happen frequently. In this case if the contacts
ends, the tangential displacements will be set to zero. In contrast for dense
systems, multiple-particle contacts are very common and the contact history for
a specic particle could be very long. The long contact history causes a relatively
large tangential displacement, which means that an extra friction force should
be taken into account. This problem can be overcome, however, by using the
method proposed by Brendel and Dippel (1998), where the tangential displace-
ment during the friction is calculated by d
t
m
f
|F
ab,n
|t
ab
/k
t
, so that
d
t

d
t
0
H
_
t
t
0
v
ab;t
dt for F
ab;t

m
f
F
ab;n

m
f
F
ab;n

t
ab
=k
t
for F
ab;t

4m
f
F
ab;n

_
_
_
(29)
4. Collision Parameters
To solve the Eqs. (20) and (21), we have to specify ve parameters: normal
and tangential spring stiffness k
n
and k
t
, normal and tangential damping
coefcient Z
n
and Z
t
, and the friction coefcient m
f
. In order to get a better
insight into how these parameters are related, it is useful to consider the equa-
tion of motion for the overlap in the normal direction dn:
m
eff
d
::
n
k
n
d
n
Z
n
d
:
n
(30)
which follows from Eq. (20) when only the normal contact force is taken into
account. In Eq. (30), m
eff
is the reduced mass of the two interacting particles a
and b:
1
m
eff

1
m
a

1
m
b
Equation (30) is the well-known differential equation of the damped harmonic
oscillator, the solution of which is
d
n
t u
0
=O expCt sinOt (31)
d
:
n
t u
0
=O expCtCsinOt OcosOt (32)
with u
0

_
d
n
0 as the initial relative velocity, and
O

O
2
0
C
2
_
O
0

k
n
=m
eff
_
C Z
n
=2m
eff

M.A. VAN DER HOEF ET AL. 94


The duration of a contact can be determined from d
n
(t
contact,n
) 0, which gives
t
contact,n
p/O, so that the relative velocity just after contact equals
d
:
n
t
contact;n
u
0
expCt
contact;n

According to the denition, the coefcient of normal restitution is given by


e
d
:
n
t
contact;n

d
:
n
0
exppC=O (33)
Thus, we can calculate the normal damping coefcient via
Z
n

2 ln e

m
eff
k
n
p

p
2
ln
2
e
_ ea0
Note that for e 0 we get O 0 according to Eq. (33), so that in that case
Z
n
2

k
n
m
eff
p
.
We can follow a similar procedure for the tangential springdashpot system.
So, the tangential damping coefcient is determined by
Z
t

2 ln e
t

m
0
eff
k
t
_

p
2
ln
2
e
t
_ e
t
a0
where m
0
eff
2 m
eff
/7 is the reduced mass of the two-particle system interaction
via a tangential linear spring. Note that m
0
eff
is different from m
eff
, since in a
tangential direction both the rotational and translational momentum must be
considered. In the case of particlewall contact, we shall simply treat particle b
as a big particle with an innite radius, so that we have
m
eff
m
a
m
0
eff

2
7
m
a
The contact force between two particles is now determined by only ve pa-
rameters: normal and tangential spring stiffness k
n
and k
t
, the coefcient of
normal and tangential restitution e and e
t
, and the friction coefcient m
f
. In
principle, k
n
and k
t
are related to the Young modulus and Poisson ratio of the
solid material; however, in practice their value must be chosen much smaller,
otherwise the time step of the integration needs to become impractically small.
The values for k
n
and k
t
are thus mainly determined by computational efciency
and not by the material properties. More on this point is given in the Section
III.B.7 on efciency issues. So, nally we are left with three collision parameters
e, e
t
, and m
f
, which are typical for the type of particle to be modeled.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 95
5. Cohesive Force
Cohesion between particles can arise from a variety of sources including van
der Waals forces, liquid bridging (i.e., capillary forces), sintering, and so on. Of
these forces, which become increasingly important as the particle size decreases,
the van der Waals force is generally accepted as the dominating cohesive force in
gas-uidized beds of ne particles (Geldart A and C particles), and will be
considered next. The van der Waals force is present between any two molecules
(polar or nonpolar) and follows from the interaction of the uctuating dipole
moments on the molecules. According to the London theory, the potential
energy of two molecules i and j at distance r
ij
, due to the van der Waals in-
teraction, is equal to jr C
6
r
6
ij
. The total energy between two macroscopic
bodies a and b, made of the same material, then equals:
V C
6

i on a f g

j on b f g
r
6
ij
C
6
r
2
_
v
a
_
v
b
dr
a
dr
b
r
a
r
b
j j
6
(34)
where in the last step we replaced S
i on a
by
_
V
a
dr
a
r(and similarly for b), where
r is the density of the material, which is justied since the number of molecules
present in the particles is very large. For two spheres with radii R
a
and R
b
,
where the centers are at position r
a
and r
b
, respectively, expression Eq. (34) can
be evaluated analytically (Hunter, 1986; Israelachvili, 1991):
V r
ab

A
6
2
r
2
ab
4

2
r
2
ab
ln
r
2
ab
4
r
2
ab
_ _ _ _
with
r
2
ab

r
2
ab
R
a
R
b

2
R
a
R
b
; r
ab
r
b
r
a
j j; A p
2
r
2
C
6
(35)
The parameter A is known as the Hamaker constant. The force on sphere a then
follows via
F
coh;a

@Vr
ab

@r
ab
n
ab

32 A
3 R
a
R
b
r
ab
n
ab
r
4
ab
r
2
ab
4
2
(36)
with n
ab
dened by Eq. (23). When the spheres are nearly touching (r
ab
-
R
a
+R
b
), and for R
a
R
b
R, the force in Eq. (36) can be simplied to
F
coh;a

AR
12s
2
ab
n
ab
s
ab
r
ab
2R
Note that Eq. (36) exhibits an apparent numerical singularity in that the van der
Waals interaction diverges if the surface distance between two particles
approaches zero. In reality, such a situation will never occur because of the
M.A. VAN DER HOEF ET AL. 96
short-range repulsion between particles. In the present model, we have not
included such a repulsion; however, we can avoid the numerical singularity by
dening a cut-off (maximal) value of the van der Waals force between two
spheres. In practice, it is more convenient to use the equivalent cut-off value for
the intersurface distance, s
ab
0
, instead of for the interparticle force.
The Hamaker constant A can, in principle, be determined from the C
6
co-
efcient characterizing the strength of the van der Waals interaction between
two molecules in vacuum. In practice, however, the value for A is also inu-
enced by the dielectric properties of the interstitial medium, as well as the
roughness of the surface of the spheres. Reliable estimates from theory are
therefore difcult to make, and unfortunately it also proves difcult to directly
determine A from experiment. So, establishing a value for A remains the main
difculty in the numerical studies of the effect of cohesive forces, where the
value for glass particles is assumed to be somewhere in the range of 10
21
joule.
6. Integrating the Equations of Motion
In the following section, we only consider the integration of the equation of
linear motion Eq. (20); the procedure for the equation of rotational motion, Eq.
(21), will be completely analogous. Mathematically, Eq. (20) represents an in-
itial-value ordinary differential equation. The evolution of particle positions
and velocities can be traced by using any kind of method for ordinary differ-
ential equations. The simplest method is the rst-order integrating scheme,
which calculates the values at a time t+dt from the initial values at time t (which
are indicated by the superscript 0) via:
v
a
v
0
a
a
0
a
dt; r
a
r
0
a
v
a
dt (37)
where a
a
is the acceleration:
a
a

F
contact;a
F
pp;a
F
ext;a
m
a
(38)
The rst-order integration scheme, however, will introduce a drift in the energy;
from Eq. (37), we have
v
a
a
0
a
dt
2
v
a

2
a
0
a
dt
2
2v
a
a
0
a
dt v
0
a

2
so
1
2
v
a

2
v
a
a
0
a
dt
1
2
a
2
a
dt
2

1
2
v
0
a

2
(39)
The rst term on the left of Eq. (39) is the reduced kinetic energy of the particle
at time t+dt, the second term is the work due to all kinds of external forces, and
the rst term on the right is the reduced kinetic energy at time t. The remaining
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 97
term a
2
a
dt
2
=2 is always positive, and this energy is introduced into the system
solely due to the numerical method, for each time step.
In the past decades, a large number of methods have been proposed to
achieve better energy conservation: for example, the Gear family of algorithms
and the family of Verlet algorithms (Frenkel and Smit, 1996). In our 3D code,
we have incorporated yet another type of method developed by Beeman, which
has a somewhat better energy conservation than the Verlet algorithm (Frenkel
and Smit, 1996). In the Beeman method, the position and velocity of particle a
are calculated via
r
a
r
0
a
v
0
a
dt
2
3
a
0
a

1
6
a
1
a
_ _
dt
2
v
a
v
0
a

1
3
a
a
dt
5
6
a
0
a

1
6
a
1
a
_ _
dt
where the superscript (1) denotes the values at time tdt. Note that the Bee-
manVerlet algorithm is not self starting, so it requires the storage of the old
value of the acceleration a
(1)
.
7. Efciency Issues: Spring Constants and Neighbor Lists
To perform simulations of relatively large systems for relatively long times, it
is essential to optimize the computational strategy of discrete particle simula-
tions. Obviously, the larger the time step dt, the more efcient the simulation
method. For the soft-sphere model, the maximum value for dt is dictated by the
duration of a contact. Since there are two different springdashpot systems in
our current model, it is essential to assume that t
contact,n
t
contact,t
, so that

p
2
ln e
2
k
n
=m
eff

p
2
ln e
t

2
k
t
=m
0
eff

If we further assume that e e


t
, then the relation between the normal and
tangential spring stiffness is
k
t
k
n

m
eff
m
0
eff

2
7
Based on the discussion in previous sections, we can calculate the time step by
dt
1
K
N
t
contact;n

1
K
N

p
2
ln e
2
k
n
=m
eff

M.A. VAN DER HOEF ET AL. 98


where K
N
is the minimum number of steps during one contact. Our experience is
that K
N
must not be less than 5, and is normally in the range 1550. It can thus
be seen that a smaller spring stiffness k
n
leads to a larger time step, and therefore
it is useful to rst perform a number of test simulations with different values for
k
n
. Another issue is the maximum overlap d
max
, which occurs at d
:
t 0. From
Eq. (31) it follows that
d
max
u
0
=O
0
exp C=O arcsinO=O
0

_
which must typically be less than 1% of the particle diameter.
A second way of speeding up the simulation is the use of neighbor lists and
cell list, which was originally developed for MD simulations (Allen and
Tildesley, 1990). The neighbor list contains a list of all particles within the
cut-off sphere of a particular particle, so that the separations do not need to
be calculated at each step, which is shown in Fig. 15. The neighbor list cut-off
s
cutoff
should be dened with care. A too small cut-off value may result in
some neighboring particles to be excluded from the list. In contrast, however,
a big cut-off value will greatly reduce the computational efciency. To speed
up the searching for neighbors, the particles in each uid cell in this research
are put into a corresponding list. All neighbors of a particle will then be
found either in the cell containing the particle or in an adjacent cell.
FIG. 15. The scheme of neighbor list and cell lists. The particle of interest is black; the grey
particles are within the neighbor list cut-off.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 99
C. GAS DYNAMICS
In the DPM the gas phase is treated as a continuum phase, the dynamics of
which can be described by a set of volume-averaged NavierStokes equations
(Kuipers et al., 1992). From mass conservation, we have
@r
@t
r ru 0 (40)
where r is the gas density, e the local porosity, and u the gas velocity. Mo-
mentum conservation gives that
@ru
@t
r ruu rp S
p
r t rg (41)
where p is the gas phase pressure, t the viscous stress tensor, g the gravitational
acceleration, and S
p
a source term that describes the momentum exchange with
the solid particles present in the control volume:
S
p

1
V
_

N
part
a1
bV
a
1
u v
a
dr r
a
dV (42)
Here V represents the local volume of a computational cell and V
a
the volume of
particle a. The d-function ensures that the drag force acts as a point force at the
(central) position of this particle. In Eq. (42), b is the momentum transfer
coefcient, which will be discussed in more detail in Section III.D. The gas
phase density r is calculated from the ideal gas law:
r
pM
RT
where R is the universal gas constant (8.314 J/(mol K)), T the temperature,
and M the molecular mass of the gas. The equation of state of the ideal gas
can be applied for most gases at ambient temperature and pressure. The
viscous stress tensor t is assumed to depend only on the gas motion. For gas-
uidized beds, the general form for a Newtonian uid (Bird et al., 1960) can
be used:
t l
2
3
m
_ _
r u

I mru ru
T
(43)
with l the gas phase bulk viscosity, m the gas phase shear viscosity, and

I the
unit tensor. Normally, the bulk viscosity of the gas phase can be set equal to
zero (Bird et al., 1960). Note that no turbulence modeling is taken into
M.A. VAN DER HOEF ET AL. 100
account. For dense gassolid uidization, this can be justied since the tur-
bulence is completely suppressed in the particle bed due to the high solids
volume fraction.
The numerical method for solving the set of Eqs. (40) and (41) is similar to the
method that is used in the TFM, which is discussed in detail in Section IV.E.
The time step by which the gas-phase is updated is typically one order of
magnitude larger than the time step dt that is used for updating the soft-sphere
system. The boundary conditions are taken into account by utilizing ctitious
cells at the boundaries and a ag-matrix concept, which allows different bound-
ary conditions to be applied for each single cell. A variety of boundary con-
ditions can be applied by specication of the value of the cell ag (i, j, k), which
denes the relevant boundary condition for the corresponding cell (i, j, k). A
typical set of boundary conditions used in a 2D simulation is shown in Fig. 16.
In Table II, we explain the meaning of each type of boundary condition.
Normally, the bottom distributor is dened as inux cells formulated by (i, j,
k) 4, where the void fraction is set to a constant value of 0.4.
FIG. 16. The typical set of boundary conditions used in 2D simulations.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 101
D. INTERPHASE COUPLING
For dense gassolid two-phase ows, a four-way coupling is required; how-
ever, the coupling between particles is managed in a natural way in DPMs. The
task is, therefore, only to nd a two-way coupling between the gas and the solid
phases, which satises Newtons third law. Basically, the gas phase exerts two
forces on particle a: a drag force F
d,a
due the uidsolid friction at the surface of
the spheres, and a force F
p,a
V
a
rp due to the pressure gradient rp in the gas
phase. We will next describe these forces in more detail, along with the pro-
cedure to calculate void fraction, which is an essential quantity in the equations
for the gassolid interaction.
1. Drag Force
The drag force that the gas phase exerts on a particle a, consistent with the
source term S
p
in expression Eq. (41), reads
F
d;a

V
a
b
1
u v
a
(44)
where b is the momentum exchange coefcient. The commonly used drag cor-
relations for b in the simulation of gas-uidized beds are the Ergun (1952)
equation for denser beds (eo0.8):
bd
2

m1
1501 1:75Re
a
(45)
and the Wen and Yu (1966) equation for dilute systems (e40.8):
bd
2

m1

3
4
C
d
Re
a

1:65
(46)
TABLE II
VALUES FOR THE CELL FLAG, WHICH DEFINE THE BOUNDARY CONDITIONS
(i, j, k) The type of cell
1 Interior cell, no boundary conditions have to be specied
2 Impermeable wall, free-slip boundaries
3 Impermeable wall, no-slip boundaries
4 Inux cell, velocities have to be specied
5 Prescribed pressure cell, free-slip boundaries
6 Continuous outow cell, free-slip boundaries
7 Corner cell, no boundary conditions have to be specied
M.A. VAN DER HOEF ET AL. 102
with Re
a
red
a
|uv
a
|/m the Reynolds number of particle a and C
d
the drag
coefcient, for which the expression by Schiller and Nauman (1935) is used:
C
d

241 0:15Re
0:687
a
=Re
a
Re
a
o10
3
0:44 Re
a
410
3
_
Note that the validity of both the Ergun and Wen and Yu equations has re-
cently been questioned on the basis of LB data, and alternative drag-force
correlations have been proposed. From LB simulations, Hill et al. (2001a, b)
suggest the following relation for Stokes ow (lim Re
a
-0):
bd
2

m1
A
o
1
with
A
o

180 o0:6
18
3
1
3

2
p 1
1=2

135
64
116:141
_ _
10:6811
2
8:481
3
8:161
4
40:6
_

_
whereas for Re
a
440, they found that the drag force increases linearly with Re
a
:
bd
2

m1
A
2
1 0:6057
3
1:908
3
1 0:209
2
_
Re
a
(47)
In the paper by Hill et al. (2001b), values for A
2
are only given for a nite
number of gas fractions
1
; however, A
2
is nearly the same as A
o
(Koch and Hill,
2001). Note that in Section II.D we suggest a different expression for b, also on
the basis of lattice Boltzmann simulations.
2. Force from the Pressure Gradient
The force on particle a due to the pressure gradient rp in the gas phase is
equal to
F
p;a
V
a
rp
Note that the reaction of this force (thus the two-way coupling) is incorporated
in the momentum conservation equation of the gas phase in the rst term on the
RHS of Eq. (41). The local value for rp at r
a
is obtained from a volume-
weighted averaging technique using the values of the pressure gradients at the
eight surrounding grid nodes. The volume-weighted averaging technique used to
1
In Ref. Hill et al. (2001b), values are listed for F
2
, which relates to A
2
via F
2
A
2
(1e)/(18e
3
).
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 103
obtain the local-averaged value

Q of a quantity Q
ijk
from the eight surrounding
computational nodes is shown in Fig. 17. The local-averaged value is calculated
as follows:

Q
Q
1
V
8
Q
2
V
7
Q
3
V
6
Q
4
V
5
Q
5
V
4
Q
6
V
3
Q
7
V
2
Q
8
V
1
DX DY DZ
where
V
1
d
x
d
y
d
z
V
2

~
d
x
d
y
d
z
V
3
d
x
~
d
y
d
z
V
4

~
d
x
~
d
y
d
z
V
5
d
x
d
y
~
d
z
V
6

~
d
x
d
y
~
d
z
V
7
d
x
~
d
y
~
d
z
V
8

~
d
x
~
d
y
~
d
z
with
~
d
x
DX dx;
~
d
y
DY dy;
~
d
z
DZ dz, and the distances d
x
, d
y
,
and d
z
necessary in this averaging techniqueare calculated from the position
of the particle in the staggered grid (see also Fig. 24). Note that the same
technique of volume weighting is also used to obtain local gas velocities and
local void fractions at the position of the center of the particle.
3. Void Fraction Calculation
From the position of each particle, we can calculate its contribution to the
local solid volume fraction e
s
in any specied uid cell. This local void fraction,
e 1 e
s
, is one of the key parameters that controls the momentum exchange
between the phases and should be determined with care.
FIG. 17. The scheme of volume-weighted averaging.
M.A. VAN DER HOEF ET AL. 104
For a 2D situation, the void fraction e(i, j) can be calculated on the basis
of the area occupied by the particles in the cell of interest. A particle can be
present in multiple cells, however, as shown in Fig. 18. Hoomans et al. (1996)
developed a method to account for the multiple cell overlap. The area of A
ii,jj
is
given by
A
ii;jj
R
2
a
d
1
d
2

1
2
d
1

1 d
2
1
_
d
2

1 d
2
2
_
arccos d
1
arcsin d
2
_ _
(48)
and area A
i,jj
by
A
i;jj
R
2
a

1
2
p d
1
d
2

1
2
d
1

1 d
2
1
_
d
2

1 d
2
2
_
arccosd
1
arccosd
2
_ _
(49)
with d
1
d
1
/R
a
and d
2
d
2
/R
a
(see Fig. 18). The area A
ii,j
can be calculated by
an equation similar to Eq. (49). However, the void fraction calculated in this
way is based on a 2D distribution of disks, whereas the empirical drag-force
correlations are derived for 3D systems. To correct for this inconsistency, the
void fraction calculated on the basis of area (e
2D
) is transformed into a 3D void
fraction (e
3D
) using the following equation:

3D
1
2

3
p
_
1
2D

3=2
FIG. 18. The multiple cell overlap of a single particle. From Hoomans, Ph.D. thesis, University of
Twente (2000).
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 105
In a true 3D situation, we can calculate the void fraction on the basis of actual
volume of the particles. However, no analytical expression is available for
volume V
ii,jj
. Hoomans et al. (1996) suggested the approximation
V
ii;jj
%
V
ii
V
a
V
jj
V
a
V
a
(50)
with V
ii
V
ii,jj
+V
ii,j
and V
jj
V
ii,jj
+V
i,jj
. The volume of the sphere caps V
ii
and V
jj
can be calculated exactly by
V
ii
V
a

1
4
1 d
1

2
2 d
1

V
jj
V
a

1
4
1 d
2

2
2 d
2

with d
i
d
i
/R
a
being the distance from the center of the particle to the cell
boundary relative to the radius of the particle and V
a
4pR
a
3
/3 the volume of
the particle. The error in the calculation of the porosity that is introduced by the
approximation in Eq. (50) is negligibly small when the particle radii are an order
of magnitude smaller than the size of the CFD-grid cell, which is required in any
case in order to have a grid-independent value of the porosity. In this context, it
is noteworthy that recently a new method has been developed that can generate
a grid-independent estimate of e, even when the size of the particles is of the
order of the size of the grid cells (Link et al., 2005).
E. ENERGY BUDGET
To relate the discrete particle simulations to the KTGFs, it is very useful to
analyze the detailed information of the energy evolution in the system. The total
energy balance of the system is obtained by calculating all relevant forms of
energy as well as the work performed due to the action of external forces.

Translational kinetic energy E


kin
trans
and rotational kinetic energy E
kin
rot
E
trans
kin

1
2

N
part
a1
m
a
v
a
v
a
E
rot
kin

1
2

N
part
a1
I
a
o
a
o
a
(51)

Potential energy from gravity


E
p

N
part
a1
m
a
g r
a

Potential energy of the normal spring and tangential spring


E
s

1
2

N
part
a1

b
k
n
d
2
ab;n
k
t
d
2
ab;t

where b4a and b A the contactlist of a, and d


ab,n
, d
ab,t
the overlap and relative
tangential displacement, respectively, of particle a and b.
M.A. VAN DER HOEF ET AL. 106

The work done by the external forces and the cohesive force in one time step
dt
W
ext
dt

N
part
a1
F
d;a
F
p;a
F
coh;a
v
a
Also, the energy dissipated during the particleparticle contact process has to
be considered and is determined by the following:

Energy dissipated by the normal and tangential spring in one time step dt
E
ds
dt

N
part
a1

b
Z
n
v
ab;n
v
ab;n
Z
t
v
ab;t
v
ab;t

Energy dissipated by the friction between particles in one time step dt


E
df
dt

N
part
a1

b
m
f
F
ab;n

t
ab;n
v
ab;t

where b4a and b A the contactlist of a.


The total energy of the system is then equal to
E
tot
E
p
E
trans
kin
E
rot
kin
E
s
E
st
W
ext
E
ds
E
df
F. RESULTS FOR THE EXCESS COMPRESSIBILITY
In previous work, we have mainly used the DPM model to investigate the
effects of the coefcient of normal restitution and the drag force on the for-
mation of bubbles in uidized beds (Hoomans et al., 1996; Li and Kuipers,
2003, 2005; Bokkers et al., 2004; Van der Hoef et al., 2004), and not so much to
obtain information on the constitutive relations that are used in the TFMs. In
this section, however, we want to present some recent results from the DPM
model on the excess compressibility of the solids phase, which is a key quantity
in the constitutive equations as derived from the KTGF (see Section IV.D.). The
excess compressibility y can be obtained from the simulation by use of the virial
theorem (Allen and Tildesley, 1990).
y
1
6mN
part

b
F
ab
r
ab
(52)
where the sums are over all particles, with the restriction that a6b. In Eq. (52),
F
ab
is the interaction force between particles a and b. For the soft-sphere model
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 107
as presented in the previous sections, in the absence of cohesive forces, F
ab
is
equal to the sum of F
ab,n
and F
ab,t,
as given by Eqs. (26) and (21), and then only
when particles a and b are in contact. Furthermore, in Eq. (52),

y is the average
granular temperature, which can be dened as the average over the total volume
of the local granular temperature dened by Eq. (59). In the absence of any drift
velocity, m

y 2E
trans
kin
=3N, with E
trans
kin
as the total translational kinetic energy
given by Eq. (51). Note that for the hard-sphere model there are no forces, and a
different procedure is required. In that case, the solids pressure (and thus the
excess compressibility) can be obtained from the average number of collisions
per unit time (Allen and Tildesley, 1990).
We have performed simulations for 500 particles with periodic boundary
conditions and no gas phase present. Owing to the inelastic collisions, the par-
ticles will continuously dissipate energy, which would eventually cause the par-
ticles to come to a quiescent state. In this work, we therefore drive the system by
two different techniques: (1) rescaling the particle velocities every time step,
according to the desired granular temperature; (2) accelerating the particles
randomly. Method (2) is most robust but less efcient. The rescaling procedure,
however, does not attain an equilibrium state for high solid fractions. For this
reason, the random acceleration procedure is used to simulate the denser system
with a solid fraction higher than 0.45, while the rescaling procedure is used for
lower solid fractions. For more details on the procedures, we refer to a recent
paper (Ye et al., 2005). All the parameters are normalized by the particle radius,
particle density, and granular temperature.
First, we should check whether the soft-sphere model gives results compa-
rable to those from the hard-sphere model, since the approximate theories of
granular ow are based on the latter model. To this end, we carried out several
sets of simulations with particles starting from either random positions or face-
centered cubic (FCC) positions. The thermodynamic properties of the hard-
sphere system for these two congurations have been well documented by many
researchers (Alder and Wainwright, 1957; Carnahan and Starling, 1969; Hoover
and Ree, 1969; Erpenbeck and Wood, 1984). In Fig. 19, we show our simulation
results for smooth, elastic, and cohesiveless spheres in periodic boundary do-
mains, where at the start of the simulation the particles are placed in an FCC
grid. For such systems, Hoover and Ree (1969) observed a phase transition
from the uid state to the solid state at y 7.27. As can be seen, both the hard-
sphere and soft-sphere simulations clearly display this transition point. For the
uid state, our simulation data from both models is in very good agreement
with the CarnahanStarling equation of state (Carnahan and Starling, 1969).
y
ES

4
s
2
2
s
1
s

3
(53)
The conclusion is that the soft-sphere model can be used as an alternative for
the hard-sphere model, as far as the calculation of the excess compressibility is
concerned.
M.A. VAN DER HOEF ET AL. 108
Next, we consider a system of inelastic spheres (ISs). As can be seen from Eq.
(81), the KTGF predicts that the excess compressibility y
IS
of ISs is a linear
function of the coefcient of normal restitution e,
y
IS

1 e
2
y
ES
(54)
where y
ES
is the excess compressibility of elastic spheres (ESs). In Fig. 20, we
show our simulation results for the excess compressibility of ISs, both for the
soft-sphere and the hard-sphere model. The solid fraction in the initial cong-
uration is xed at 0.05. It is shown that for this dilute system, the simulation
results of both models are in very good agreement with the prediction in
Eq. (54) from the KTGF (solid line). Note that the Eq. (54) is derived under the
assumption that the particles are only slightly inelastic, i.e., e$1.0.
In Fig. 21, the excess compressibility is shown as a function of the solid
fraction for different coefcients of normal restitution e. These results are
compared with the Eq. (54), where the excess compressibility y
ES
is taken from
either the MaAhmadi correlation (Ma and Ahmadi, 1986) or the Car-
nahanStarling correlation. As can be seen, the excess compressibility agrees
well with both correlations for a solid fraction e
s
up to 0.55. For extremely dense
systems, i.e., e
s
40.55, the MaAhmadi correlation presents a much better
estimate of the excess compressibility, which is also the case for purely elastic
particles (see Fig. 23).
FIG. 19. Simulation results for both the soft-sphere model (squares) and the hard-sphere model
(the crosses), compared with the CarnahanStarling equation (solid-line). At the start of the sim-
ulation, the particles are arranged in a FCC conguration. Spring stiffness is K 70,000, granular
temperature is y 1.0, and coefcient of normal restitution is e 1.0. The system is driven by
rescaling.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 109
Up to this point, we have neglected the cohesive van der Waals forces between
the particles, which is only justied if particles are larger than say 1 mm. Pres-
ently, the van der Waals forces have not been included in the KTGF; a rst step
would be to consider the effect of such forces on the excess compressibility by
also including the interparticle force of Eq. (36) in F
ab
of Eq. (52). In Fig. 22, the
results for the excess compressibility for different Hamaker constants A are
shown. For simplicity, a coefcient of normal restitution e 1.0 is used. We
consider two different Hamaker constants: A 3.0 10
12
and A 3.0 10
10
(in units where r
s
1, R 1, and y 1). From Fig. 22, we see that for these
two Hamaker constants, the simulation results differ only slightly from the
prediction in Eq. (54), where y
ES
is calculated from the MaAhmadi correlation,
which suggests that cohesion has only a weak inuence on the excess com-
pressibilityat least for the values of Hamaker constant that we studied. In this
context, it should be noted that the quantication of the cohesive force is not
straightforward since there is no reference force (such as gravitational force) in
these systems. We consider these systems as slightly cohesive since the ratio of
the cohesive potential and the average kinetic energy per particle is small, i.e.,
j 6.25 10
8
$6.25 10
6
. At the same time, the ratio between the cohesive
force and contact force ranges from 1.11 10
5
to 1.11 10
3
. If a strong
cohesive force is present, particles in the system may form complicated struc-
tures, whereas a homogeneous state is one of the basic assumptions underlying
0.0 0.2 0.4 0.6 0.8 1.0
e
0.5
0.6
0.7
0.8
0.9
1.0
y
I
S

/

y
E
S
y
IS
= y
ES
(1+e)/2
Hardsphere
Softsphere
FIG. 20. Excess compressibility y
IS
for a system of inelastic hard spheres, as function of the
coefcient of normal restitution, for one solid fraction (e
s
0.05). The excess compressibility has
been normalized by the excess compressibility y
ES
of the elastic hard spheres system. Other sim-
ulation parameters are as in Fig. 19.
M.A. VAN DER HOEF ET AL. 110
the KTGF. A more detailed analysis of the effect of the cohesive force on the
excess compressibility can be found in Ref. Ye et al. (2005).
IV. Two-Fluid Model
A. INTRODUCTION
In the EulerEuler models, i.e., the TFMs, it is assumed that both the gas and
the solid phase are interpenetrating continua. This continuous approach is es-
pecially useful and computationally cost-effective when the volume fractions of
the phases are comparable, or when the interaction within and between the
phases plays a signicant role in determining the hydrodynamics of the system.
As discussed before, it is relatively straightforward to model the gas phase, for
instance by the use of well-established CFD techniques. The challenge is to
establish an accurate hydrodynamic description of the particulate phase.
0.0 0.2 0.4 0.6
Solid fraction
s
0
10
20
30
40
y
I
S
0.0 0.2 0.4 0.6
Solid fraction
s

0
10
20
30
40
y
I
S
0.0 0.2 0.4 0.6
Solid fraction
s
0
10
20
30
40
y
I
S
0.0 0.2 0.4 0.6
Solid fraction
s

0
10
20
30
40
y
I
S
e = 1.00
e = 0.95
e = 0.90
e = 0.80
FIG. 21. The excess compressibility from soft-sphere simulations, with random initial particle
positions, for different coefcients of normal restitution e: (a) e 1.0 (top-right); (b) e 0.95 (top-
left); (c) e 0.90 (bottom-right); (d) e 0.80 (bottom-left). The simulation results (symbols) are
compared with Eq. (54) based on the MaAhmadi correlation (solid line) or the CarnahanStarling
correlation (dashed line). The spring stiffness is set to k
n
70,000.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 111
Anderson and Jackson (1967, 1968, 1969) and Ishii (1975) have separately de-
rived the governing equations for TFMs from rst principles. Although the
details of constructing the averaged equations are different, the nal equations
are essentially the same. The TFMs differ signicantly from each other as
different closures for the solid stress tensor are used.
There are basically three types of approaches to dene the solid stress tensor,
or more specically the solid viscosity. In the early hydrodynamic models
developed by Jackson and his co-workers (Anderson and Jackson, 1967; And-
erson et al., 1995), Kuipers et al., (1992), and Tsuo and Gidaspow (1990)the
viscosity is dened as an empirical constant, and also the dependence of the
solid phase pressure on the solid volume fraction is determined from experi-
ments. The advantage of this model is its simplicity, the drawback is that it does
not take into account the underlying characteristics of the solid phase rheology.
In another class of models, pioneered by Elghobashi and Abou-Arab (1983)
and Chen (1985), a particle turbulent viscosity, derived by extending the concept
of turbulence from the gas phase to the solid phase, has been used. This is the
so-called k model, where the k corresponds to the granular temperature and
is a dissipation parameter for which another conservation law is required. By
coupling with the gas phase k turbulence model, Zhou and Huang (1990)
developed a k model for turbulent gasparticle ows. The k models do not
FIG. 22. The effect of the cohesive force on the excess compressibility. The coefcient of normal
restitution is e 1.0, and granular temperature is T 1.0. The Hamaker constant is A 3.0 10
12
(circles) and 3.0 10
10
(crosses).
M.A. VAN DER HOEF ET AL. 112
include the effect of particleparticle collisions, and so these models are re-
stricted to dilute gasparticle ows.
Signicant contributions to the modeling of gassolid ows have been made
by Gidaspow and co-workers (1994), who combined the kinetic theory for the
granular phase with continuum representations for the particle phase. There are
a number of other studies using this approach. Sinclair and Jackson (1989)
predicted the core-annular regime for steady developed ow in a riser. Ding et
al. (1990) simulated a bubbling uidized bed. Transient simulations and com-
parisons to data were done by Samuelsberg and Hjertager (1996). Nieuwland et
al. (1996) investigated a circulating uidized bed using the KTGF. Detamore et
al. (2001) have performed an analysis of scale-up of circulating uidized beds
using kinetic theory.
One of the strengths of the KTGF, although still under development, is that it
can offer a very clear physical picture with respect to the key parameters (e.g.,
particle pressure, particle viscosity, and other transport coefcients) that are
used in the TFMs. The TFMs based on KTGF requires less ad hoc adjustments
compared to the other two types of models. Therefore, it is the most promising
framework for modeling engineering-scale uidized beds.
B. GOVERNING EQUATIONS
In the TFM, both the gas phase and the solid phase are described as fully
interpenetrating continua using a generalized form of the NavierStokes equa-
tions for interacting uids. The continuity and momentum equations for the gas
phase are given by expressions identical to Eqs. (40) and (41), except for the
gassolid interaction term:
@r
@t
r ru 0 (55)
@ru
@t
r ruu rp bu u
s
r s rg (56)
with t as the viscous stress tensor of the gas phase given by Eq. (43). The
continuity and momentum equations for the particle phase are given by a sim-
ilar set of equations:
@
s
r
s

@t
r
s
r
s
u
s
0 (57)
@
s
r
s
u
s

@t
r
s
r
s
u
s
u
s

s
rp rp
s
bu u
s
r s
s

s
r
s
g (58)
where e
s
1e and u
s
is the velocity of the solid phase. Note that r
s
is the
material density of the solid phase, so that the local mass per unit volume is
equal to r
s
e
s
. Obviously, the numerical scheme for updating the solid phase is
now analogous to (and synchronous with) that of the gas phase, the details of
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 113
which are given in Section IV.E. Since the concept of particles has disappeared
completely in such a modeling, the effect of particleparticle interactions can
only be included indirectly, i.e., via the effective solid pressure p
s
and the effec-
tive solid stress tensor s
s
. A description that allows for a more detailed de-
scription of particleparticle interactions follows from the KTGF, which
expresses the pressure and the solid stress tensor as a function of the local
granular temperature y, which is dened from the uctuation in the velocity of
the individual solid particles. More precisely, the granular temperature at r is
dened as
y
1
3
1
N
r

N
r
a1
v
a
u
s

2
_ _
(59)
where /.S is an ensemble average, and the sum is over all N
r
particles in a small
control volume dV around r. Note that also the solid density and velocity as
they appear in Eqs. (57) and (58) can be dened from the positions and mom-
enta of the individual particles by similar type of averages
2
:

s
r
s

1
dV

N
r
a1
m
a
_ _

s
r
s
u
s

1
dV

N
r
a1
m
a
v
a
_ _
(60)
For particles of equal mass, we thus have e
s
r
s
mn with n the local number
density of particles. From the KTGF, the time evolution of the granular tem-
perature is given by
3
2
@
@t

s
r
s
y r
s
r
s
yu
s

_ _
p
s

I t
s
: ru
s
r q
s
3by g (61)
with q
s
the kinetic energy ux and g the dissipation of kinetic energy due to
inelastic particle collisions. In Eqs. (58) and (61), there are three unknown
quantities (pressure, stress tensor, and energy ux), which must be expressed in
terms of the three basic hydrodynamic variables (density, velocity, and tem-
perature), in order to get a closed set of equations. This is the subject of the
KTGF, and the resulting closures will be presented in Section IV.D. However,
before doing so, we will rst give a brief description of the general principles of
kinetic theory.
2
Note that for dV-0 the local density and momentum density can be written as
e
s
p
s

a
m
a
d(r r
i
) and e
s
p
s
u
s

a
m
a
v
a
d (r r
i
), which are the expressions that are usually found
in literature.
M.A. VAN DER HOEF ET AL. 114
C. GENERAL KINETIC THEORY
In this section, we will only discuss the basic principles of kinetic theory,
where for detailed derivations we refer to the classic textbook by Chapman and
Cowling (1970), and a more recent book by Liboff (1998). Of central impor-
tance in the kinetic theory is the single particle distribution function f
s
(r, v),
which can be dened as the number density of the solid particles in the 6D
coordinate and velocity space. That is, f
s
(r, v, t) dv dr is the average number of
particles to be found in a 6D volume dv dr around r, v. This means that the
local density and velocity of the solid phase in the continuous description are
given by
r
s
r; t
_
1
1
m f
s
r; v; tdv (62)
and
r
s
r; tu
s
r; t
_
1
1
mv f
s
r; v; tdv (63)
where the local density is dened as r
s
r
s

s
with r
s
as the material density of
the solid particles. The granular temperature, dened by Eq. (59), follows from
r
s
r; tyr; t
1
3
_
1
1
mv u
s

2
f
s
r; v; tdv (64)
The evolution of the one-particle distribution function f
s
can be described by the
Boltzmann equation
@
@t
f
s
r; v; t v rf
s
r; v; t C (65)
which is basically a continuity equation, where the second term on the left-hand
side (LHS) represents the change of f
s
in time due to streaming and the collision
function C on RHS represents the change of f
s
due to particleparticle inter-
actions. Conservation of mass, momentum, and energy in a collision gives that
C should satisfy
_
1
1
Cdv 0;
_
1
1
Cvdv 0;
_
1
1
Cu
2
dv 0
Taking the same integrals (
_
ydv,
_
yvdv, and
_
yu
2
dv) of the Boltzmann
equation Eq. (65), making use of Eqs. (62) and (63), yields
@
@t
r
s
r r
s
u
s
0 (66)
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 115
@
@t
r
s
u
s
r r
s
u
s
u
s
r p (67)
3
2
@
@t
r
s
y r r
s
yu
s

_ _
p : ru
s
r q
s
(68)
with
p
_
1
1
mVVf
s
r; v; tdv; q
_
1
1
mV
2
2
Vf
s
r; v; tdv; V v u
s
In principle, one should solve the Boltzmann equation Eq. (65) in order to arrive
at explicit expressions for the pressure tensor p and heat ux q, which proves not
possible, not even for the simple BGK equation Eq. (11). However, one can
arrive at an approximate expression via the ChapmanEnskog expansion, in
which the distribution function is expanded about the equilibrium distribution
function f
s
eq
, where the expansion parameter is a measure of the variation of the
hydrodynamic elds in time and space. To second order, one arrives at the
familiar expression for p and q
p p
s

I s
s
; q
s
k
s
ry (69)
with

I is the unit tensor, and
s
s
m
s
ru
s
ru
s

T
_
l
s

2
3
m
s
_ _
r u
s

I
_
(70)
where ru
ab
r
a
u
b
; ru
T
ab
r
b
u
a
. Inserting the above expression for p and q
into Eqs. (67) and (68) will give the NavierStokes equations, where the pa-
rameters k
s
, l
s
, m
s
, and ps can be calculated (at least in princeiple) when the
collision function C is known. For the simple BGK equation Eq. (11), this will
result in the relations of Eq. (13). For an accurate description of the solid phase,
however, one requires a much more detailed expression for C, which contains
the details of the particleparticle interactions. Although this is a laborious
route, it opens a possibility for making a link between the microscopic details
of particle collisions and the macroscopic transport coefcients. Apart from
the details of the particleparticle interactions, C does also depend on the joint
probability function f
2
s
(r
1
, v
1
, r
2
, v
2
, t), provided that the interactions between
the particles are pair-wise additive (generally for n-body interactions, C will
depend on f
s
(n)
). In order to get a closed equation, f
2
s
should be described in
term of f
s
. If the velocities v
1
and v
2
are not correlated, one can write
f
2
s
r
1
; v
1
; r
2
; v
2
; t gr
12
;
s
f
s
r
1
; v
1
; t f
s
r
2
; v
2
; t
M.A. VAN DER HOEF ET AL. 116
where g(r
12
, e
s
) is the pair distribution function, which depends only on the
distance r
12
jr
2
r
1
j and the solid fraction.
For sufciently low density, g 1, the collision function takes the form
C
1
m
2
_
dOsO
_
dv
0
v v
0
f
s
~ vf
s
~ v
0
f
s
vf
s
v
0

_
(71)
where we have omitted the r, t argument of f
s
. In Eq. (71), ~ v, ~ v
0
, are the velocities
of the two particles involved after the collision, which can be constructed from
the initial velocities v, v
0
from conservation of energy and momentum:
~ v v aa v v
0
~ v
0
v
0
aa v
0
v
with a as the unity vector along the line connecting the two centers of the
particle before the collision. Furthermore, in Eq. (71), s(O) represents the
cross-section and O is the solid angle in which the particle is scattered. More
details on these concepts can be found in the standard literature (Chapman and
Cowling, 1970; Liboff, 1998). Using this form of the collision function, it can be
shown that pressure p
s
, shear viscosity m
s
, and thermal conductivity k
s
in Eqs.
(69) and (70) are given by
m
id
s

5
96
pr
s
d

y
p
_
; k
id
s

75
384
pr
s
d

y
p
_
; p
id
s

s
r
s
y (72)
where d is the diameter of the particles, and the superscript id (ideal) indicates
that the expressions are for the limit of a dilute gas, for which the pressure is
given by the ideal gas law.
For high densities, g cannot be set equal to one, and the collision function
becomes much more complex and so is not given here. It turns out, however,
that instead of using the full radial distribution function, it is sufcient to use
the value at contact r R, so that we dene a new function:
w
s
gR;
s

In the standard Enskog theory (SET), the shear viscosity and thermal con-
ductivity of ESs are found to be equal to
3
m
ES
s
m
id
s
1
wbr
s

4
5
0:7614wbr
s
_ _
br
s
(73)
3
See Chapman and Cowling (1970). Note that the true expression for m
ES
s
reads
m
ES
s
c
1
m
id
s

1
xbr
s

4
5

4
25
1
12
pc
2
wbr
s
br
s
, with c
1
c
2
1.016. In most expressions in literature, c
1
is set equal to 1; in expression Eq. (82) of Gidaspow, both c
1
c
2
1, which is the cause of the
slightly different coefcient 0.771, compared to 0.7614 in Eq. (75). For practical purposes, the
difference is negligible. Similar remarks can be made about k.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 117
k
ES
s
k
id
s
1
wbr
s

6
5
0:7574wbr
s
_ _
br
s
(74)
with b 2e
s
pd
3
/3 m, so that bp
s
4e
s
. Note that the pressure of a dense system
is directly related to the radial distribution function at contact (Chapman and
Cowling, 1970; Hansen and McDonald, 1986):
p
ES
s
p
id
s
1 y
ES
y
ES
wbr
s
4w
s
with y
ES
the excess compressibility of the elastic hard-sphere system. Thus in the
Enskog theory, the transport coefcients are completely determined by the y
ES
:
m
ES
s
4m
id
s

s
1
y
ES

4
5
0:7614y
ES
_ _
(75)
k
ES
s
4k
id
s

s
1
y
ES

6
5
0:7574y
ES
_ _
(76)
Various expressions for y
ES
have been proposed in literature based on the virial
coefcients and simulation data. Most of these have the following general form:
y
ES

n0
c
n
4
s

n1
1
s
=
cp

b
(77)
with e
cp
the close-packed solid fraction, at which the pressure diverges. In
Table III, we summarize the parameters found by different authors. A com-
parison of expression in Eq. (77) with the MD simulation data from Alder and
Wainwright (1960) and Woodcock (1981) is shown in Fig. 23. In our current
TABLE III
VALUES FOR THE PARAMETERS IN EQ. (77)
CS MA SSM TC
e
cp
1 0.64356 0.6435 0.6875
a 1 3 1 1
b 3 0.67802 0.76 1
c
o
1 1 1 1
c
1
1/8 0.625 0.3298 0.2613
c
2
0 0.2869 0.08867 0.05968
c
3
0 0.070554 0.01472 0.005905
c
4
0 0 0.0005396 0.001191
c
5
0 0 0.0003574 0.0004455
c
6
0 0 0.0005705 0.0004818
c
7
0 0 0.0001212 0.00003636
c
8
0 0 0.0001151 0.00008182
Note: CS: Carnahan and Starling (J. Chem. Phys. 51, 635 (1969)); MA: Ma and Ahmadi (J. Chem.
Phys. 84, 3449 (1986)); SSM: Song, Stratt and Mason (J. Chem. Phys. 88, 1126 (1988)); TC: To-
bochnik and Chapin (J. Chem. Phys. 88, 5824 (1988)).
M.A. VAN DER HOEF ET AL. 118
version of the TFM, we use the expression by Ma and Ahmadi (1986) (see also
Fig. 21). Alder et al. (1970) have also measured the shear viscosity in MD
simulations of dense hard-sphere systems. It was found that the Enskog ap-
proximation in Eq. (75) is very accurate up to e
s
0.3; however, for higher solid
fractions the theory signicantly underestimates the shear viscosity up to a
factor of two for e
s
E0.5.
D. KINETIC THEORY OF GRANULAR FLOW
In the KTGF, the dissipation of energy in collisions is included in the Enskog
theory. Currently, only the effect of the coefcient of normal restitution has
been considered, although it is anticipated that friction also plays an important
role. The derivation of the constitutive equations for ISs can be found in the
book by Gidaspow (1994) and the papers by Jenkins and Savage (1983), Lun
et al. (1984), Ding and Gidaspow (1990), and Nieuwland et al. (1996). Here,
we will present the expressions for p
s
, m
s
, and K
s
from the book of Gidaspow
(1994) (Eqs. (T.9.1), (9.183), (9.250), (9.262), (9.268), and (9.272)):
p
IS
s
1 21 e
s
g
s
r
s
y (78)
m
IS
s

5
96
pr
s
d

y
p
_
2
1 eg
1
4
5
1 e
s
g
_ _
2

4
5

2
s
r
s
dg1 e

y
p
_
(79)
0.4 0.5 0.6 0.7
solid fraction
s
0
20
40
60
80

y
E
S
CarnahanStarling
MaAhmadi
SongStrattMason
MD (Woodcock)
MD (Alder & Wainwright)
FIG. 23. Comparison of the expressions from Eq. (77) and Table III with data from MD sim-
ulations.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 119
k
IS
s

75
384
pr
s
d

y
p
_
2
1 eg
1
6
5
1 e
s
g
_ _
2
2
2
s
r
s
dg1 e

y
p
_
(80)
where g is the value of the radial distribution function of a hard-sphere uid at
contact and e is the coefcient of normal restitution. Note that from Eq. (78) it
follows that the excess compressibility of the IS system is equal to
y
IS
21 e
s
g
1 e
2
y
ES
(81)
that is, the dissipation in the collisions reduces the excess compressibility by a
factor of (1+e)/2. Replacing 2(1+e)e
s
g in Eqs. (79) and (80) by y
IS
and using
expression in Eq. (72) for m
s
id
and k
s
id
gives
m m
id
s
4
s
1
y
IS
1
2
5
y
IS
_ _
2

48
25p
y
IS
_ _
m
id
s
4
s
1
y
IS

4
5
0:771y
IS
_ _
(82)
k k
id
s
4
s
1
y
IS
1
3
5
y
IS
_ _
2

32
25p
y
IS
_ _
k
id
s
4
s
1
y
IS

6
5
0:767y
IS
_ _
(83)
which are of the same form as the Enskog expressions in Eqs. (75) and (76), with
y
ES
replaced by y
IS
.
3
It thus turns out, like for the elastic hard spheres, that the
constitutive equations are completely determined by the excess compressibility,
and that the general form of the Enskog equations is not affected by the dis-
sipation of energy in the collisions.
Note that in the granular temperature equation Eq. (61), there is one extra
term that is absent in the SET, namely the dissipation of uctuating kinetic
energy g. From the KTGF follows that
g
3
2
1 ey
IS
r
s

s
y
4
d

y
p
_
r u
s
_ _
E. NUMERICAL SOLUTION METHOD
Owing to the tendency of inelastic particles to contract in high-density clus-
ters, and the strong nonlinearity of the particle pressure near the maximum
packing density, special attention has to be paid to the numerical implemen-
tation of the model equations. Most classic constant property TFMs are
solved using computational methods based on the implicit continuous Eulerian
(ICE) method pioneered by Harlow and Amsden (1975). The implementation is
based on a nite difference technique and the algorithms closely resemble the
SIMPLE algorithm (Patankar and Spalding, 1972), whereby a staggered grid is
employed to reduce numerical instability. A detailed discussion on the appli-
cation of this numerical technique to TFMs for gas-uidized beds is presented
by Kuipers et al. (1992).
M.A. VAN DER HOEF ET AL. 120
In principle, this numerical solution method can be straightforwardly applied
to modern TFMs with closure laws according to the KTGF. However, when
doing so, the numerical stability of the TFM is highly affected by the value of
the coefcient of normal restitution. Problems that can be handled with ac-
ceptable time steps of 10
4
s for ideal particles (e 1) require time steps of 10
5
s when the coefcient of normal restitution is taken to be 0.97, and unacceptably
small time steps of 10
6
s have to be taken when the coefcient of normal
restitution is reduced below 0.93. This extreme sensitivity to the value of the
coefcient of normal restitution is caused by the fact that particle volume frac-
tions at the next time level are estimated without taking into account the strong
nonlinear dependence of the particle pressure on the particle volume fraction. A
new numerical algorithm, which estimates the new particle volume fraction
taking the compressibility of the particulate phase more directly into account, is
presented in this section.
1. Discretization of the Governing Equations
The set of conservation equations, supplemented with the constitutive equa-
tions, boundary, and initial conditions cannot be solved analytically, and a
numerical method must be applied to obtain an approximate solution. There-
fore, the domain of interest is divided into a number of xed Eulerian cells
through which the gassolid dispersion moves. A standard nite difference
technique is applied to discretize the governing equations.
4
The cells are labeled
by indices i, j, and k located at their centers, and a staggered grid conguration
is applied. According to this conguration the scalar variables are dened at the
cell centers, whereas the velocities are dened at the cell faces, as indicated in
Fig. 24. Furthermore, different control volumes have to be applied for mass and
granular energy conservation on one hand and the momentum conservation
equations on the other. The control volumes for mass and granular energy
conservation coincide with the Eulerian cells, whereas the control volumes for
momentum conservation in all three directions are shifted half a cell with re-
spect to the Eulerian cells. Applying rst-order time differencing and fully im-
plicit treatment of the convective uxes, the discretized form of continuity
equation for the solid phase, Eq. (57), becomes

s
r
s

n1
i;j;k

s
r
s

n
i;j;k

dt
dx
h
s
r
s
u
s;x
i
n1
i
1
2
;j;k
h
s
r
s
u
s;x
i
n1
i
1
2
;j;k
_ _

dt
dy
h
s
r
s
u
s;y
i
n1
i;j
1
2
;k
h
s
r
s
u
s;y
i
n1
i;j
1
2
;k
_ _

dt
dz
h
s
r
s
u
s;z
i
n1
i;j;k
1
2
h
s
r
s
u
s;z
i
n1
i;j;k
1
2
_ _
0 84
4
This part is based upon Chapter 2 of the thesis of Goldschmidt (2001).
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 121
where the superscripts n and n+1 indicate that the quantities are at the old and
the new time, respectively. For the discretization of all convective mass,
momentum, and uctuating kinetic energy uxes the second-order accurate
Barton scheme (Centrella and Wilson, 1984; Hawley et al., 1984) is applied. A
schematic representation of this scheme for the convective transport of a
quantity D (e.g., er) by a velocity V
i+1/2
(e.g., u
x
) is given in Fig. 25. In the
discretization of the momentum Eq. (58), the terms associated with the gas and
solid pressure gradients are treated fully implicitly. The interphase momentum
transfer term is treated in a linear implicit fashion, and all other terms are
treated explicitly. The discretization of the solid phase momentum in Eq. (58)
for the x-direction is given by

s
r
s
u
s;x

n1
i
1
2
;j;k
A
n
i
1
2
;j;k

s

n1
i
1
2
;j;k
dt
dx
p
n1
i1;j;k
p
n1
i;j;k
_ _

dt
dx
p
s

n1
i1;j;k
p
s

n1
i;j;k
_ _
dtb
n
i
1
2
;j:k
u
x
u
s;x

n1
i
1
2
;j;k
85
where momentum convection, viscous interaction, and gravity have been
collected in the explicit term A
n
. The equation for the y-direction is obtained by
FIG. 24. Positions at which the key variables are evaluated for a typical computational cell in the
staggered-grid conguration.
M.A. VAN DER HOEF ET AL. 122
substituting y for x, B for A, and a change of subscripts:
. . .
i
1
2
;j;k
) . . .
i;j
1
2
;k
. . .
i1;j;k
) . . .
i;j1;k
and the equation for the z-direction is obtained by the substituting z for x, C for
A, and a change of subscripts
. . .
i
1
2
;j;k
) . . .
i;j;k
1
2
. . .
i1;j;k
) . . .
i;j;k1
Note that the mass and momentum equations for the gas phase can simply be
obtained by replacing e
s
-e, r
s
-r, u
s
-u in Eqs. (84) and (85), and dropping
the terms concerning the particle-pressure gradient.
The granular energy equation is solved in a fully implicit manner. The
solution of the equation however proceeds through a separate iterative
procedure that solves the granular temperature equations for the whole
computational domain when this is required by the main solution procedure
discussed in the next paragraph. In this separate iterative procedure, the terms
regarding convective transport and generation of uctuating kinetic energy by
viscous shear are explicitly expressed in terms of the most recently obtained
granular temperature y*. The granular energy dissipation term is treated in a
semi-implicit manner, whereas all other terms are treated fully implicitly. The
FIG. 25. Schematic representation of the Barton scheme for the convective ux of a quantity D by
velocity V
i+1/2
in the x-direction.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 123
applied discretization of the granular temperature equation is given by
3
2

s
r
s
y
n1
i;j;k

3
2

s
r
s
y
n
i;j
D

i;j;k
dt3b
n
i;j;k
y
n1
i;j;k
dt
g
y
_ _

i;j;k
y
n1
i;j;k
p
s

n1
i;j;k
dt
dx
u
s;x

i
1
2
;j;k
u
s;x

i
1
2
;j;k
_ _
p
s

n1
i;j;k
dt
dy
u
s;y

i;j
1
2
;k
u
k;y

i;j
1
2
;k
_ _
p
s

n1
i;j;k
dt
dx
u
s;z

i;j;k
1
2
u
k;y

i;j;k
1
2
_ _

dt
dx
k
s

i
1
2
;j;k
1
dx
y
n1
i1;j;k
y
n1
i;j;k
_ _
k
s

i
1
2
;j;k
1
dx
y
n1
i;j;k
y
n1
i1;j;k
_ _
_ _

dt
dx
k
s

i;j
1
2
;k
1
dy
y
n1
i;j1;k
y
n1
i;j;k
_ _
k
s

i;j
1
2
;k
1
dy
y
n1
i;j;k
y
n1
i;j1;k
_ _
_ _

dt
dz
k
s

i;j;k
1
2
1
dz
y
n1
i;j;k1
y
n1
i;j;k
_ _
k
s

i;j;k
1
2
1
dz
y
n1
i;j;k
y
n1
i;j;k1
_ _
_ _
86
In this equation, the superscript (*) indicates that a term is computed based
upon the most recent information, which complies with the (n+1)th time level
when all iterative loops have converged. Further, the convective transport and
viscous generation of uctuating kinetic energy have been collected in the
explicit term D*. The iterative solution procedure for the granular energy
equations continues until the convergence criteria
y
n1
i;j;k
y

i;j;k
oe
y
y
n1
i;j;k
(87)
are simultaneously satised for all cells within the computational domain. For a
typical value of e
y
10
6
, this takes only a couple of iterations per time step.
2. Solution Procedure of the Finite Difference Equations
The numerical solution of the discretized model equations evolves through a
sequence of computational cycles, or time steps, each of duration dt. For each
computational cycle, the advanced (n+1)-level values at time t+dt of all key
variables have to be calculated for the entire computational domain. This
calculation requires the old n-level values at time t, which are known from either
the previous computational cycle or the specied initial conditions. Then each
computational cycle consists of two distinct phases:

calculation of the explicit terms A


n
, B
n
, and C
n
in the momentum equations
for all interior cells and

implicit determination of the pressure, volume fraction, and granular


temperature distributions throughout the computational domain with an
iterative procedure. The implicit phase consists of several steps.
M.A. VAN DER HOEF ET AL. 124
The rst step involves the calculation of the mass residuals of the solid phase
(D
s
)
i,j,k
and the gas phase (D
g
)
i,j,k
from the continuity equations, for all interior
cells:
D
s

i;j;k

s
r
s

i;j;k

s
r
s

n
i;j;k

dt
dx
h
s
r
s
u
s;x
i

i
1
2
;j;k
h
s
r
s
u
s;x
i

i
1
2
;j;k
_ _

dt
dy
h
s
r
s
u
s;y
i

i;j
1
2
;k
h
s
r
s
u
s;y
i

i;j
1
2
;k
_ _

dt
dz
h
s
r
s
u
s;z
i

i;j;k
1
2
h
s
r
s
u
s;z
i

i;j;k
1
2
_ _
88
D
g

i;j;k
r

i;j;k
r
n
i;j;k

dt
dx
hru
x
i

i
1
2
;j;k
hru
x
i

i
1
2
;j;k
_ _

dt
dy
hru
y
i

i;j
1
2
;k
hru
y
i

i;j
1
2
;k
_ _

dt
dz
hru
z
i

i;j;k
1
2
hru
z
i

i;j;k
1
2
_ _
89
If the convergence criteria
D
g

i;j;k
oe
g
r

i;j;k
(90)
D
s

i;j;k
oe
s

s
r
s

i;j;k
(91)
are not satised for all computational cells (typically e
g
e
s
10
6
), a whole
eld pressure correction is calculated, satisfying
D
g

i;j;k
J
g

n
i;j;k
dp
i;j;k
J
g

n
i1;j;k
dp
i1;j;k
J
g

n
i1;j;k
dp
i1;j;k
J
g

n
i;j1;k
dp
i;j1;k
J
g

n
i;j1;k
dp
i;j1;k
J
g

n
i;j;k1
dp
i;j;k1
J
g

n
i;j;k1
dp
i;j;k1
92
where (J
g
)
n
represents the Jacobi matrix for the gas phase. This matrix contains
the derivatives of the defects D
g
with respect to the gas phase pressure, for which
explicit expressions can be obtained from the continuity equation for the gas
phase in combination with the momentum equations. To save computational
effort, the elements of the Jacobi matrix are evaluated at the old time level. The
banded matrix problem corresponding to Eq. (92) is solved using a standard
ICCG sparse matrix technique. Once new pressures have been obtained, the
corresponding new gas phase densities are calculated.
So far, the solution procedure has been exactly the same as the SIMPLE
procedure that is usually applied for the solution of the classic TFMs with
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 125
constant property closure equations. In the next step however, the standard
procedures continue with the computation of the new velocities from the
coupled momentum equations, after which the new volume fractions are
obtained from the solid phase mass balances, and only then the new solid
pressures are determined. This regularly leads to excessive compaction and
extremely high particle pressures in areas where the particle packing densities
are close to random close packing. Therefore, the new solution procedure
computes the particle volume fractions, taking the compressibility of the solid
phase more directly into account. Similar to the pressure correction for the gas
phase, a whole eld particle volume fraction correction is computed, satisfying
D
s

i;j;k
J
s

n
i;j;k
d
s

i;j;k
J
s

n
i1;j;k
d
s

i1;j;k
J
s

n
i1;j;k
d
s

i1;j;k
J
s

n
i;j1;k
d
s

i;j1;k
J
s

n
i;j1;k
d
s

i;j1;k
J
s

n
i;j;k1
d
s

i;j;k1
J
s

n
i;j;k1
d
s

i;j;k1
93
In this Eq. (J
s
)
n
is the Jacobi matrix for the solid phase, which contains the
derivatives of the mass residuals for the particulate phase to the solid volume
fraction. Explicit expressions for the elements of the Jacobi matrix can be
obtained from the continuity for the solid phase and the momentum equations.
For example for the central element, the following expression is obtained from
the solid phase continuity equation, in which the convective terms are evaluated
with central nite difference expressions:
J
s

n
i;j;k

@D
s

i;j;k
@
s

i;j;k
r
s

i;j;k

dt
dx
@
s
r
s
u
s;x

i
1
2
;j;k
@
s

i;j;k

@
s
r
s
u
s;x

i
1
2
;j;k
@
s

i;j;k
_ _

dt
dy
@
s
r
s
u
s;y

i;j
1
2
;k
@
s

i;j;k

@
s
r
s
u
s;y

i;j
1
2
;k
@
s

i;j;k
_ _

dt
dz
@
s
r
s
u
s;z

i;j;k
1
2
@
s

i;j;k

@
s
r
s
u
s;z

i;j;k
1
2
@
s

i;j;k
_ _
94
The derivatives of the mass uxes to the solid volume fractions can
subsequently be obtained from the solid phase momentum equations. From
Eq. (85), the discretized x-momentum equation, the derivatives of the mass
uxes in the x-direction can easily be obtained, e.g.,
@
s
r
s
u
s;x

i
1
2
;j;k
@
s

i;j;k

1
2
dt
dx
p

i1;j;k
p

i;j;k
_ _

dt
dx
@p
s
@
s
_ _

i;j;k
dtb
n
i
1
2
;j;k
@u
x
u
s;x

i
1
2
;j;k
@
s

i;j;k
95
M.A. VAN DER HOEF ET AL. 126
The second term on the RHS of this equation shows that the compressibility
of the solid phase is taken directly into account in the estimation of the new
particle volume fractions. Furthermore, the expression for the derivatives of the
velocities to the solid pressure can be obtained by combination with the x-
momentum equation for the gas phase that results in
@ru
x

i
1
2
;j;k
@
s

i;j;k

1
2
dt
dx
p

i1;j;k
p

i;j;k
_ _
dtb
n
i
1
2
;j;k
@u
x
u
s;x

i
1
2
;j;k
@
s

i;j;k
(96)
Together with Eq. (93), this equation forms a set of equations from which
explicit expressions for the derivatives of the velocities can readily be obtained.
Expressions for the y- and z-direction and for the other elements of the Jacobi
matrix are obtained in a similar manner.
After the new solid volume fractions have been obtained from Eq. (93), new
particle pressures are calculated, where after new velocities can be obtained
from the coupled momentum equations. Next, new granular temperatures are
calculated from the granular energy equations by an iterative procedure
described in Section IV.E.1. Finally, the new mass residuals (D
g
)
i,j,k
and (D
s
)
i,j,k
are computed and the convergence criteria are checked again.
Though this new algorithm still requires some time step renement for
computations with highly inelastic particles, it turns out that most computations
can be carried out with acceptable time steps of 10
5
s or larger. An alternative
numerical method that is also based on the compressibility of the dispersed
particulate phase is presented by Laux (1998). In this so-called compressible
disperse-phase method the shear stresses in the momentum equations are
implicitly taken into account, which further enhances the stability of the code in
the quasi-static state near minimum uidization, especially when frictional shear
is taken into account. In theory, the stability of the numerical solution method
can be further enhanced by fully implicit discretization and simultaneous solution
of all governing equations. This latter is however not expected to result in faster
solution of the TFM equations since the numerical efforts per time step increase.
F. APPLICATION TO GELDART A PARTICLES
A great challenge in CFD modeling of gassolid two-phase ows is to obtain
realistic predictions of the uidization behavior of small particles such as
Geldart A particles (Geldart, 1973), for which the standard TFM has so far
failed to predict even the bubbling uidization. Ferschneider and Mege (1996)
found a major overprediction of bed expansion in a bubbling bed of FCC
particles, and Bayle et al. (2001) obtained the same results in a turbulent bed of
FCC particles. Recently, Lettieri et al. (2003) used a particlebed model,
originally developed by Chen et al. (1999), to investigate the homogeneous
uidization of Geldart A particles. It has been demonstrated that a
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 127
homogeneous expansion can be obtained in this particlebed model. However in
this model, an articial particle-phase elasticity force is required. McKeen and
Pugsley (2003) used the two-uid CFD code MFIX to simulate a freely
bubbling bed of FCC catalyst for U
0
0.050.2 m/s and compared their
simulation results with ECT data. In accordance with ndings of Ferschneider
and Mege (1996), McKeen and Pugsley (2003) also found that the standard
CFD model greatly overpredicted bed expansion without any modications of
the drag closures. By using a scale factor of 0.25 for the commonly used
gassolid drag laws, they found that their simulation results are in accordance
with experimental observations. They argued that this is due to the formation of
clusters with a size smaller than the CFD grid size. Such small-scale clusters
have not been reported before, in particular for particles with a size of 75 mm.
Although the van der Waals force can play a role in the uidization of
Geldart A particles, it is not clear how this force affects the gassolid drag. The
inuence of the cohesion on the KTGF has not been carefully checked so far.
Recently, Kim and Arastoopour (2002) tried to extend the kinetic theory to
cohesive particles; however, their nal expression for the particular phase stress
is very complex. A simpler route would be to assume that the Enskog expres-
sions in Eqs. (75)(76) still hold for cohesive particles, only with a modied
excess compressibility. However at present, it proves difcult to give an accurate
estimate of the deviation of y due to the cohesive force (see Fig. 22). Moreover,
as discussed in Section III.F, also the magnitude of the cohesive force itself (i.e.,
the Hamaker constant) is not known. For this reason, we will only study the
effect of the gasparticle drag in this section, where we use two different models:
(i) the ab-initio drag model in Eq. (47) derived from detailed scale LB sim-
ulations and (ii) the empirical drag model in Eq. (46). Note that for the latter
model, the literature values for the exponent n are extremely scattered (Morgan
et al., 1971). In Table IV, we show the results for n from different experiments
for Geldart A particles, which are clearly much higher than the value n 4.65,
originally obtained by Richardson and Zaki (1954). In this section, we show
results using the Wen and Yu expression with the commonly used value
n 4.65, and with the highest reported value n 9.6, from the experiments by
Lettieri et al. (2002).
For the simulations we use a 2D TFM as described in the previous sections.
The simulation conditions are specied in Table V. The gas ow enters at the
bottom through a porous distributor. The initial gas volume fraction in each
uid cell is set to an average value of 0.4 and with a random variation of 75%.
Also for the boundary condition at the bottom, we use a uniform gas velocity
with a superimposed random component (10%), following Goldschmidt et al.
(2004).
The simulations show that for low gas velocities (U
0
0.009 m/s), the com-
monly used exponent n 4.65 does not yield a realistic bed expansion dynamics
for Geldart A particles. By using a large exponent (n 9.6), which was deter-
mined by gas uidization of Geldart A particles, we can get a bed expansion
M.A. VAN DER HOEF ET AL. 128
around 31% of the initial bed height, which is much closer to the experimental
results (Geldart, 1973). Basically, a larger exponent n in Eq. (46) will lead to a
higher drag at the same gas velocity. It can thus be argued that at low-gas
velocities the drag force is underestimated by the commonly used drag models.
The question arises what the physical origin is of such large exponents. One
possibility is that they are caused by microstructures that form from small-scale
instabilities and perhaps other mechanisms. Also, the experiments by Lettieri et
al. (2002) showed a much larger apparent terminal velocity, which is indicative
of a much larger effective size. If such microstructures cannot be captured by the
CFD grid, then the use of a modied drag function, such that the experimental
bed expansion is obtained, is a possible way to go about. It should be stressed,
however, that this type of approach is rather ad hoc and not in the spirit of the
multiscale modeling strategy.
It has been reported by several researchers (Ferschneider and Mege, 1996;
Bayle et al., 2001; McKeen and Pugsley, 2003) that an overestimated bed ex-
pansion was found at a high-gas velocity ($0.2 m/s). We also carried out several
simulations for a high gas velocity, U
0
0.2 m/s. We still use the drag model
given by Eq. (46) with an exponent n 4.65. The simulation domain, however,
TABLE V
SIMULATION CONDITIONS
Parameters Value Parameter Value
Gas shear viscosity 1.8 10
5
Pa s CFD cells 30 45
Gas temperature 293 K Size of the cell 5 5 mm
2
Gas pressure 1.01 10
5
Pa Particle diameter 75 mm
Gas constant 8.314 J/(mol K) Particle density 1,500 kg/m
3
CFD time step 1.0 10
4
s Coefcient of restitution 0.97
TABLE IV
EXPONENT N FOR GELDART A PARTICLES
d
p
(mm) n
Lettieri et al. (2002), Newton and Gates (2002): Gas-uidization
71 9.6
57 9.0
49 8.2
Massimilla et al. (1972): Gas-uidization ()
60 7.12
53 6.86
45 6.1
Lewis and Bowerman (1952): Liquid-uidization
8
6
8.3
Whitmore (1957): Liquid-sedimentation
65 9.5
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 129
is enlarged so that a high bed expansion can be accommodated. The compu-
tational domain is composed of 30 70 cells, and the size of each cell still
remains as 5 5 mm
2
. With such a high gas velocity the bed in fact is in the
turbulent uidization regime. In Fig. 26, we show the results obtained at differ-
ent points in time when the bed reaches a dynamical equilibrium. Clearly, the
particle phase displays a turbulence-like ow pattern. Also, an overestimation
of bed height is found in the simulations, which is around 100% of the initial
bed height.
We also carried out a set of simulations using Eq. (47) as a drag model, which
was based on the data of LB simulations. The results are shown in Fig. 27. As
can be seen, no big differences can be observed compared to the results from the
drag model given by Eq. (46) with an exponent n 4.65.
A similar simulation was also carried out by McKeen and Pugsley (2003).
They also found an overestimation of the bed height, compared to their ex-
perimental results. They argued that a factor should be used to scale down the
FIG. 26. The bed expansion dynamics of Geldart A particles from the TFM. The supercial gas
velocity U
0
is set to 0.2 m/s. The exponent n of the Wen and Yu equation is set to 4.65. No cohesion
is considered here. The results are, from the right to left, taken at t 9.6, 9.7, 9.8, 9.9, and 10.0 s.
M.A. VAN DER HOEF ET AL. 130
drag force in this regime in order to obtain a better agreement with the ex-
periments. In Fig. 28, we show the results of our simulations with a drag force
(n 4.65) scaled down by a factor 0.15. A signicant decrease of the bed height
is found, with a bed expansion that is around 16% of the initial bed height, close
to the experimental observations (McKeen and Pugsley, 2003).
V. Towards Industrial-Scale Models
In Section I, we mentioned that the TFM can simulate uidized beds at
engineering scales (height 12 m), and that the large-scale industrial uidized-
bed reactors (diameter 15 m, height 320 m) are still far beyond its capabilities.
Clearly, it would be highly desirable to predict the properties of gassolid ows
at the industrial scale; however at present, there is no fully evolved model
based on fundamental principleswhich is capable of this. In this section, we
outline some new ideas in this direction that have been developed both at the
FIG. 27. The same as in Fig. 26, but now using the LB drag model in Eq. (47) from Hill et al.
(2001b), with A
2
A
o
.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 131
Princeton University and at the University of Twente. Before doing so; how-
ever, it is rst important to understand why the current class of TFMs is not
suitable for describing large-scale gassolid ows.
A. THE LIMITS OF THE TWO-FLUID MODEL
Let us step back and examine the TFM and the closures we described thus far
in the chapter. Recall that the details of ow at the level of individual particles
are erased by the averaging process leading to the TFM equations, and that
their consequences appear in the averaged equations through terms which have
to be closed. The size of the averaging region was not explicitly considered
anywhere in the derivation of the TFM equations or the closures, and it was
implicitly assumed that a separation of scale existsnamely, the size of the
averaging region is much larger than the particle sizebut is much smaller than
the scale of the macroscopic ow structure that we wish to study by solving the
TFM equations. The assumption of such a separation of scales underlies the
very formulation of continuum models.
FIG. 28. The same as in Fig. 26, but with the drag force scaled by a factor of 0.15.
M.A. VAN DER HOEF ET AL. 132
Furthermore, the closures for the uidparticle drag and the particle-
phase stresses that we discussed were all derived from data or analysis of
nearly homogeneous systems. In what follows, we refer to the TFM equa-
tions with closures deduced from nearly homogeneous systems as the micro-
scopic TFM equations. The kinetic theory based model equations fall in this
category.
We illustrated how these equations are discretized over an appropriate nu-
merical grid and also showed some sample results. One can readily appreciate
that one must choose the grid sizes in the numerical solution of the TFM
equations to be smaller than the shortest length scale at which the TFM equa-
tions afford inhomogeneities. This requirement leads to a practical difculty
when one tries to solve these microscopic TFM equations for gasparticle ows,
as discussed below.
Gasparticle ows in uidized beds and riser reactors are inherently unstable
and they manifest inhomogeneous structures over a wide range of length and
time scales. There is a substantial body of literature where researchers have
sought to capture these uctuations through numerical simulation of micro-
scopic TFM equations, and it is now clear that TFMs for such ows do reveal
unstable modes whose length scale is as small as ten particle diameters (e.g., see
Agrawal et al., 2001; Andrews et al., 2005).
This is illustrated in Fig. 29. Transient simulations of a uidized suspension of
ambient air and typical uid catalytic cracking catalyst particles were performed
(using MFIX (Syamlal et al., 1993; Syamlal, 1998, which is an open-domain
code for solving multiphase ow problems) in a 2D periodic domain at different
grid resolutions. These simulations employed kinetic theory-based (microscopic)
TFM equations; see Agrawal et al. (2001) for a summary of the equations,
closures, and parameter values used in the simulations. Although there are some
slight differences between the closure expressions used by these authors and
those described (as illustrative examples) in this article, the differences are only
quantitative and not qualitative, so there is no need to present these closures
here. A pressure drop that is commensurate with the weight of the gasparticle
mixture in the periodic box was applied across the box in the vertical direction,
which provided the driving force for the upow of the uidizing gas. The sim-
ulations revealed that an initially homogeneous suspension gave way to an
inhomogeneous state with persistent uctuations. Snapshots of the particle
volume fraction elds obtained in simulations with different number of spatial
grids are shown in Fig. 29.
It is readily apparent that ner and ner structures get resolved as the
number of spatial grids is increased. Statistical quantities, such as average slip
velocity between the gas and particle phases, obtained by averaging over the
whole domain, were found to depend on the grid resolution employed in the
simulations and they became nearly grid-size independent only when grid sizes
of the order of a few (E10) particle diameters were used. Thus, if one sets out
to solve microscopic TFM equations, grid sizes of the order of few particle
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 133
diameters are required; such ne spatial grids (and the fact that inhomoge-
neous structures extend down to this ne scale) limit the time steps that can be
taken as well. For most devices of practical (commercial) interest, such ex-
tremely ne spatial grids and small time steps are unaffordable (e.g., see Sun-
daresan, 2000). Indeed, gasparticle ows in large uidized beds and risers are
often simulated by solving discretized versions of the TFM equations over
coarse spatial grids. Such coarse-grid simulations do not resolve the small-scale
(i.e., subgrid scale) spatial structures that, according to the microscopic TFM
equations, do indeed exist. The effect of these unresolved structures must be
brought to bear on the structures resolved in coarse-grid simulations through
appropriate modications to the closuresfor example, the effective drag
coefcient in the coarse-grid simulations will be smaller than that in the orig-
inal TFM to reect the tendency of the gas to ow around the unresolved
clusters. Qualitatively, this is equivalent to an effectively larger apparent size
for the particles.
One can readily pursue this line of thought and examine the inuence of these
unresolved structures on the effective interphase transfer and dispersion coef-
cients that should be used in coarse-grid simulations. Inhomogeneous distribu-
tion of particles will promote by passing of the gas around the particle-rich
regions and this will necessarily decrease the effective interphase mass and en-
ergy transfer rates. Similarly, uctuations associated with the small-scale in-
homogeneities will contribute to the rate of dispersion of the particles and the
gas, but they will be unaccounted for in the coarse-grid simulations of the
microscopic TFM equations.
FIG. 29. Snapshots of particle volume fraction elds obtained while solving a kinetic theory-based
TFM. 75 mm uid catalytic particles in ambient air. Simulations were done over a 16 32 cm pe-
riodic domain. The average particle volume fraction in the domain is 0.05. Dark (light) color
indicates regions of high (low) particle volume fractions. (See Refs. Agrawal et al., 2001; Andrews et
al., 2005) for other parameter values.) Source: Andrews and Sundaresan (2005).
M.A. VAN DER HOEF ET AL. 134
B. STATE-OF-THE-ART ON DEALING WITH UNRESOLVED STRUCTURES
Researchers have approached this problem of treating unresolved structures
through various approximate schemes. OBrien and Syamlal (1993) and Boemer
et al. (1994) pointed out the need to correct the drag coefcient to account for
the consequence of clustering and proposed a correction for the very dilute
limit. Some authors have used apparent cluster size in an effective drag-coef-
cient closure as a tuning parameter; for example see McKeen and Pugsley
(2003), who attribute the larger apparent size to interparticle attractive forces,
and others have deduced corrections to the drag coefcient using energy min-
imization multiscale approach (see Yang et al., 2004). The concept of particle-
phase turbulence has also been explored to introduce the effect of the uctu-
ations associated with clusters and streamers on the particle-phase stresses
(Dasgupta et al., 1994; Hrenya and Sinclair, 1997). However, a systematic ap-
proach that combines the inuence of the unresolved structures on the drag
coefcient and the stresses that can arise even when interparticle forces are not
important has not yet emerged.
One can summarize the multiscale character of TFM simulations using coarse
spatial grids as follows. When confronted with the task of performing simu-
lation of gasparticle ows in large process vessels, one faces constraints on
affordable grid resolution; this can lead to unresolved subgrid structures that
would have been obtained if only the TFM equations were solved on a ne
spatial grid. The consequence of these subgrid structures on the ow pattern
resolved by the coarse-grid simulations should be brought in through appro-
priate corrections to the closure relations. If one simply uses the closures in the
microscopic TFM without adding the corrections, then there is no guarantee
that the obtained solution is a true solution for the TFM equations that one sets
out to solve.
This is well known in other contexts, such as single-phase turbulence. Large
eddy simulations introduce corrections to the uid-phase stress through subgrid
models; for example, Smagorinsky, in his pioneering work (Smagorinsky, 1963),
introduced a model for subgrid viscosity correction.
Agrawal et al. (2001) pointed out that, in gasparticle ows such as those
encountered in uidized beds and riser ows, one should include subgrid cor-
rections for not only the effective particle and uid-phase stresses but also the
effective drag. They showed that the effective drag law and the effective stresses
obtained by averaging (the results gathered in highly resolved simulations of a
set of microscopic TFM equations, such as that corresponding to the most
resolved snapshot in Fig. 29 over the whole (periodic) domain were very differ-
ent from those used in the microscopic TFM and that they depended on size of
the domain over which simulations were carried out (Agrawal et al., 2001). They
also found that all the effects seen in the 2D simulations persisted when sim-
ulations were repeated in three dimensions (3D) and that both 2D and 3D
simulations revealed the same qualitative trends.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 135
Andrews et al. (2005) performed many highly resolved simulations of uid-
ized gasparticle mixtures in a 2D periodic domain, whose size coincided with
that of the grid size in an anticipated large-scale riser ow simulation. Through
such highly resolved simulations, they constructed ad hoc subgrid models for the
effects of the ne-scale ow structures on the drag force and the stresses, and
examined the consequence of these subgrid models on the outcome of the
coarse-grid simulations of gasparticle ow in a large-scale vertical riser. They
have demonstrated that these subgrid scale corrections can affect the predicted
ow patterns profoundly.
Thus, there is no doubt that one must carefully examine whether the micro-
scopic TFM equations must be modied to introduce the effects of unresolved
structures before embarking on coarse-grid simulations of gasparticle ows in
chemical reactors. At the same time, the ad hoc method employed by Andrews et
al. (2005), namely performing highly resolved simulations in periodic domains
whose linear dimensions are the same as those of the grids, is not a rigorous
approach to take either; for example, one can anticipate that the periodic
boundary conditions imposed in such highly resolved simulations would place
some restrictions (on the small-scale ow structure) that would be absent in the
real, large-scale ow. Thus, alternate approaches to constructing closures suit-
able for coarse-grid computations must be developed.
Adopting the approach pursued in large eddy simulations, one can start with
the TFM equations and perform a ltering operation, where the averaging is
done over a lter length scale that is somewhat larger than the grid size to be
used in the coarse-grid simulation of large-scale process vessels and over high
(temporal) frequencies. The mathematical steps involved in ltering any version
of the microscopic TFM are conceptually straightforward (e.g., see Zhang and
VanderHeyden, 2002) and will not be presented here. We simply note that the
dominant terms in the ltered equations can be recast in exactly the same form
as the original TFM equations; however, effective stresses, interphase interac-
tion force term, etc. will now involve additional contributions resulting from the
ltering process. (It is because of this similarity that one can use the same
platform such as MFIX to perform integration of the ltered equations as well.)
Insight into these closures for the additional contributions resulting from the
ltering process can be gained through analysis of computational data gathered
through highly resolved simulations in sufciently large domains, while ensuring
that the overall ow domain simulated is considerably larger than the region
over which the ltering operation is performed. This is illustrated below by
some results obtained by Andrews and Sundaresan (2005).
Consider a highly resolved simulation of a set of microscopic TFM equations
for a uidized suspension of particles in a large periodic domain. The ltering
operation does not require a periodic domain; however, as each location in a
periodic domain is statistically equivalent to any other location, statistical av-
erages can be gathered much faster when simulations are done in periodic do-
mains. After an initial transient period that depends on the initial conditions,
M.A. VAN DER HOEF ET AL. 136
persistent, time-dependent, and spatially inhomogeneous structures develop.
Fig. 30 shows an instantaneous snapshot of the particle volume fraction eld in
one such 2D simulation (performed using MFIX) and the cells (i.e., ne grids)
used in the simulations. One can then zoom in any region of desired size and
average any quantity of interest over all the cells inside that region, and obtain
region-averaged (ltered) values. Note that one can choose a large number of
regions inside the overall domain and thus several region-averaged values can be
constructed for any quantity of interest from each instantaneous snapshot.
When the system is in a statistical steady state, one can construct tens of thou-
sands of such averages by repeating the analysis at various time instants.
Returning to Fig. 30, note that the averages over different regions at any
given time are not equivalent; for example, at the given instant, different regions
(of the same size) will correspond to different region-averaged particle volume
fractions, particle and uid velocities, and so on. Thus, one cannot simply lump
the results obtained over all the regions; instead, the results must be grouped
FIG. 30. Snapshot of particle volume fraction elds obtained while solving a kinetic theory-based
TFM. Fluid catalytic particles in air. Simulations were done over a 16 16 cm periodic domain.
128 128 cells (shown in the gure). The average particle volume fraction in the domain is 0.05.
Dark (light) color indicates regions of high (low) particle volume fractions. Squares of different sizes
illustrate regions (i.e., lters) of different sizes over which averaging over the cells is performed.
Source: Andrews and Sundaresan (2005).
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 137
into bins based on various markers and perform statistical averages within each
bin to get useful information. The 2D simulations of Andrews and Sundaresan
(2005) revealed that the single most important marker for regions is the average
particle volume fraction in that region. Therefore, in order to expose the effects
of particle volume fraction on the ltered (i.e., region-averaged) quantities, they
classied the region-averaged data into bins of particle volume fraction and
evaluated the ltered slip velocity, uidparticle interaction force, etc., and
averaged each of these quantities within each bin. From such bin statistics, they
calculated the ltered drag coefcient, ltered particle-phase pressure, and
ltered particle-phase viscosity as functions of ltered particle volume fraction.
Fig. 31 shows the variation of the ltered drag coefcient as a function of the
ltered (i.e., region-average) particle volume fraction for various lter sizes.
5
Each point represents the average of many realizations in a bin. (Here, the
ltered drag coefcient is dened as the region-average drag force divided by the
region-average slip velocity.) It is clear from Fig. 31 that the ltered drag
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
0
10
20
30
40
50
60
70
80
90

s
bin
e
f
f
e
c
t
i
v
e

d
r
a
g

c
o
e
f
f
i
c
i
e
n
t
1x1 cells = 0.125cm
2x2 cells = 0.250cm
4x4 cells = 0.50cm
8x8 cells = 1.0cm
16x16 cells = 2.0cm
FIG. 31. Filtered drag coefcient (in CGS units) extracted from simulations over 16 16 cm
domain using 128 128 cells. Source: Andrews and Sundaresan (2005).
5
Strictly speaking, one should use 2D bins involving particle volume fraction and a Reynolds
number based on slip velocity to classify the ltered drag coefcient; however in these simulations,
the Reynolds number effect was found to be weak and hence the data were collapsed to just volume
fraction bins.
M.A. VAN DER HOEF ET AL. 138
coefcient depends on the size of the lter used in the analysis. This gure
includes results obtained from three different simulations corresponding to
three different average particle volume fractions in the domain (0.05, 0.15, and
0.40). The larger the lter size the smaller is the drag coefcient , the reason
being that the averaging (i.e., ltering) is being performed over larger and larger
clustersthe larger the clusters, the greater is the bypassing of the gas around
the clusters and hence lower is the apparent drag coefcient.
Fig. 32 shows the variation of ltered particle-phase pressure as a function of
the ltered particle volume fraction for various lter sizes. Here the ltered
particle-phase pressure includes the pressure arising from the streaming and
collisional parts captured by the kinetic theory and the sub-lter-scale Reyno-
lds-stress like velocity uctuations (see Agrawal et al., 2001 for further details).
Indeed, the contributions resulting from the sub-lter-scale velocity uctuations
swamp the kinetic theory pressure, indicating that at the coarse-grid scale one
can even ignore the kinetic theory contributions to the pressure! This gure
clearly shows that the ltered particle-phase pressure increases with lter size,
and this is a direct consequence of the fact that the energy associated with the
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
0
20
40
60
80
100
120
140
160
180
200

s
bin
p
s
1x1 cells = 0.125cm
2x2 cells = 0.250cm
4x4 cells = 0.50cm
8x8 cells = 1.0cm
16x16 cells = 2.0cm
FIG. 32. Filtered particle phase pressure (in CGS units) extracted from simulations over 16
16 cm domain using 128 128 cells. Source: Andrews and Sundaresan (2005). The ltered particle-
phase pressure includes the Reynolds stress-like uctuations and the kinetic theory pressure.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 139
velocity uctuations increases with lter length (analogous to what one has in
single-phase turbulence).
The trends presented in Figs. 31 and 32 qualitatively similar to those pre-
sented earlier by Agrawal et al. (2001) and Andrews et al. (2005) who, for the
sake of simplicity, did simulations on much smaller domains and let the lter
size be the same as the domain size. This shows clearly that the effects leading to
the type of results presented in Figs. 31 and 32 are robust.
Andrews and Sundaresan (2005) have also extracted the ltered particle-
phase viscosity from these simulations and found that at low particle volume
fractions (0.00.25), the ltered viscosity varies nearly linearly with particle
volume, and that it increases monotonically (and nearly linearly) with lter size.
A nal piece of the proof-of-concept calculations is to compare the predic-
tions obtained by solving the ltered TFM equations with highly resolved sim-
ulations of the microscopic TFM equations. For this purpose, Andrews and
Sundaresan (2005) performed simulations of the microscopic TFM equations in
a 16 32 cm periodic domain at various resolutions (e.g., see Fig. 29). From
these simulations, they extracted domain-average quantities in the statistical
steady state (see Agrawal et al., 2001 for a discussion of how these data are
gathered). Fig. 33 shows the domain-average slip velocity between the gas and
particle phases at various grid resolutions (shown by the squares connected by
0 10 20 30 40 50 60 70 80 90
0
20
40
60
80
100
grids in lateral direction
t
i
m
e

a
n
d

d
o
m
a
i
n

a
v
e
r
a
g
e
d

s
l
i
p

v
e
l
o
c
i
t
y
kinetic theory model
filter size = 2.0cm
FIG. 33. Comparison of the domain-average slip velocity (in cm/s) determined by solving a mi-
croscopic TFM and the corresponding ltered TFM. 16 32 cm periodic domain. Domain-average
particle volume fraction 0.05. Number of grids in the vertical direction is twice that in the lateral
direction. Source: Andrews and Sundaresan (2005).
M.A. VAN DER HOEF ET AL. 140
the bold solid line in this gure). After sufcient grid resolution, this quantity
clearly levels off, indicating convergence in a statistical sense. They also per-
formed computations with the ltered TFM, using the computationally gen-
erated closures (e.g., drag and particle-phase pressure closures shown in Figs. 31
and 32, and particle-phase viscosity closure, not shown) for a 2 2 cm lter.
The domain-average slip velocity obtained by solving the ltered equations at
different grid resolutions are shown in Fig. 33 as triangles (connected by the thin
solid line).
Fig. 33 reveals two important features. Firstly, at coarse resolutions, the
domain-average slip velocity obtained by solving the microscopic TFM changes
appreciably with grid resolution; in contrast, the grid-size dependence of the slip
velocity computed by solving the ltered TFM is much weaker. Secondly, at
sufciently high-grid resolution, both approaches yield comparable predictions,
and this is an important rst step in validating the ltered TFM approach.
Another result that is not evident in Fig. 33 concerns the computational times
required for gathering the statistical steady-state values of various quantities
(such as the slip velocity shown in Fig. 33); at comparable grid resolutions, the
computational time required to solve the ltered equations is much smaller than
that for the microscopic equations. This can be attributed to the fact that the
structures obtained in the solution of the ltered equations are comparatively
coarser than those for the microscopic TFM equations.
C. A DIFFERENT APPROACH: THE DISCRETE BUBBLE MODEL
An alternative scheme to tackle the problem of large-scale ow structures is
being pursued at Twente University. In this model the bubbles, as observed in
the DPM and TFM models of gas-uidized beds, are considered as discrete
entities. This is the so-called discrete bubble model, which has been successfully
applied in the eld of gasliquid bubble columns (Delnoij et al., 1997). The idea
to apply this model to describe the large-scale solids circulation that prevail in
gassolid reactors is new, however, and involves some slight modications of
the equivalent model for gasliquid systems (Bokkers et al., 2005a). To this end,
the emulsion phase is modeled as a continuumlike the liquid in a gasliquid
bubble columnand the larger bubbles are treated as discrete bubbles. Note
that granular systems have no surface tension, so in that respect there is a
pronounced difference with the bubbles present in gasliquid bubble columns.
For instance, the gas will be free to ow through a bubble in gassolid systems,
which is not the case for gasliquid systems. As far as the numerical part is
concerned, the DBM strongly resembles the DPM as outlined in Section III,
since it is also of the EulerLagrange type with the emulsion phase described by
the volume-average NavierStokes equations:
@
e
r
e

@t
r
e
r
e
u
e
0 (97)
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 141
with r
e
, e
e
, and u
e
the density, volume fraction, and ow velocity, respectively,
of the emulsion phase. Momentum conservation gives that
@
e
r
e
u
e

@t
r
e
r
e
u
e
u
e

e
rp S
E
r s
e

e
r
e
g (98)
where the symbols take their usual meaning, and the subscript e indicates
emulsion phase. The term S
E
accounts for the two-way coupling between the
dispersed phase and the continuous phase. The bubbles are considered as dis-
crete elements that are tracked individually according to Newtons second law
of motion:
m
b
dv
b
dt
F
tot
(99)
where F
tot
is the sum of different forces acting on a single bubble:
F
tot
F
g
F
p
F
d
F
W
F
VM
(100)
As in the DPM model, the total force on the bubble has contributions from
gravity (F
g
), pressure gradients (F
p
), and drag from the interaction with emulsion
phase (F
d
). The sum of F
g
and F
p
is equal to (p
e
p) V
b
g, with V
b
the volume of
the bubble. For the drag force on a single bubble (diameter d
b
), the correlations
for the drag force on a single sphere are used, only with a modied drag co-
efcient C
d
, such that it yields the relation v
br
0:711

gd
b
_
by Davies and Tay-
lor (1950) for the rise velocity of a single bubble. Note that in Eq. (100), there are
two forces present that are not included in the DPM, namely the wake force F
W
and the virtual mass force F
VM
. The wake force, accounting for the acceleration
of a bubble in the wake of a leading bubble, is neglected in this application;
whereas for the virtual mass force, the relation by Auton (1983) is used:
F
VM

DI
Dt
I ru
_ _
; I 0:5r
e
V
b
v
b
u
e

An advantage of this approach to model large-scale uidized bed reactors is


that the behavior of bubbles in uidized beds can be readily incorporated in the
force balance of the bubbles. In this respect, one can think of the rise velocity,
and the tendency of rising bubbles to be drawn towards the center of the bed,
from the mutual interaction of bubbles and from wall effects (Kobayashi et al.,
2000). In Fig. 34, two preliminary calculations are shown for an industrial-scale
gas-phase polymerization reactor, using the discrete bubble model. The geom-
etry of the uidized bed was 1.0 3.0 1.0 m (wh d). The emulsion phase
has a density of 400 kg/m
3
, and the apparent viscosity was set to 1.0 Pa s. The
density of the bubble phase was 25 g/m
3
. The bubbles were injected via 49
nozzles positioned equally distributed in a square in the middle of the column.
M.A. VAN DER HOEF ET AL. 142
In Figs. 34a and 34c, snapshots are shown of the bubbles that rise in the
uidized bed with a supercial gas velocity of 0.1 m/s (a) and 0.3 m/s (c). It is
clearly shown that the bubble holdup is much larger with a supercial gas
velocity of 0.3 m/s. However, the number of bubbles in this case is too large,
since bubble coalescence has not been accounted for in these simulations. In
Figs. 34b and 34d, time-averaged plots are shown of the emulsion velocity after
100 s of simulation. The large convection patterns, upow in the middle and
downow along the wall, are clearly demonstrated. Coalescence, which is a
highly prevalent phenomenon in uidized beds, is not included in the simu-
lations described above. However, since all the bubbles are tracked individu-
ally, it is relatively straightforward to include this in the DBM. In the latest
version of the DBM (Bokkers et al., 2005a), a simple coalescence model is
included, which was found to have a large effect on the macroscale circulation
pattern. In this model, all the bubbles that collide will coalesce till a max-
imum size, where the largest bubbles start to breakup.
VI. Outlook
In this chapter, we have discussed three levels of modeling for dense
gassolid ows, with the emphasis on the technical details of each of the
models, which have not been published elsewhere. Up till now, the models
have mainly been used to obtain qualitative information, that is, to acquire
insight into the mechanisms underlying the gassolid ow structures. How-
ever, the ultimate objective of the multiscale approach is to obtain quantitative
FIG. 34. Snapshots of the bubble hold-up in the DBM model without bubble coalescence, and the
time average vector plot of the emulsion phase after 100 s of simulation; (a)+(b) u
0
0.1.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 143
information that can be fed into the higher level models. Such a program has
just started at the University of Twente, and in Sections II.D and III.F we
have shown some rst results from this line of research. Much remains to be
done, however; specically for each level:
(i) For the LBM, we have presented a new drag-force correlation for ideal
systems: homogeneous (unbounded) static arrays of perfect monodisperse
spheres; in uidized beds, however, such systems are hardly ever encoun-
tered. Therefore, the next step(s) would be to consider the effects of het-
erogeneity, mobility, and polydispersity. With respect to the latter, our LB
studies for binary systems at very low Reynolds numbers showed that the
drag-force relations that are currently used for such systems are wrong by
up to a factor of 5, for relatively moderate diameter ratios of 1:4 (Van der
Hoef et al., 2005). For higher Reynolds numbers (100 and 500), the results
show a similar trend (Beetstra et al., 2006). Some preliminary results on size
segregation indicate that our new drag-force relations have a large effect on
the segregation phenomena in binary uidized systems (Beetstra et al.,
2006). With respect to heterogeneous structures, it will be obvious that the
drag force of clusters of spheres will be very different from that of a ho-
mogeneous suspension. We have recently performed an LBM study for
small clusters (close-to-sphere and H-shaped), and the conclusion was that
the effective drag of the cluster was equivalent to that of a large sphere that
has the same (projected) surface area as the cluster, perpendicular to the
direction of ow (Beetstra et al., 2006).
(ii) In the DPM, future work will be focused on measuring the particulate
pressure and viscosity, where it will be of particular interest to test how well
the general Enskog relation in Eq. (75) between the viscosity and excess
compressibility holds, which follows from kinetic theory, and if necessary
adjust the equations on the basis of the simulation data. The next step
would be to include the effect of the gas phase (drag), particle friction, van
der Waals interactions, and also polydispersity. Note that the determina-
tion of the viscosity is not straightforward, and has to our knowledge not
been measured previously in discrete particle simulations of uidized beds.
One option is to use methods from MD simulations (Allen and Tildesley,
1990)): statistical mechanics of nonequilibrium systems gives that the shear
viscosity is equal to the time integral (GreenKubo integral) of the
stressstress correlation function /s
xy
(0)s
xy
(t)S, where s
xy
is the mi-
croscopic stress tensor, which can be written in terms of the particle po-
sitions, velocities, and forces (Hansen and McDonald, 1986; Allen and
Tildesley, 1990). The second method is by measuring the velocity decay of
the impact of a large sphere (diameter D) in a uidized bed. When we
assume that the collisions of the large intruder with the small bed particles
take the effect of a StokesEinstein drag force $ 3pm
s
Du on the large
particle, with u the velocity of the intruder, then the effective solids viscosity
M.A. VAN DER HOEF ET AL. 144
can be obtained directly. Our discrete particle simulations of such an
impact in a bed of monodisperse particles are reported in Ref. Lohse et al.
(2004).
(iii) With respect to the TFM, the main challenges are the simulations of po-
lydisperse systems and ne powders. For the latter, we saw in Section IV.F
that without an ad hoc scaling of the drag force, the current class of TFMs
cannot predict the uidization properties of A-powders. For even smaller
particles, cluster-like structures might be formed, so that the application of
the current version of KTGF should be seriously questioned in any case. In
other words, it would be unlikely that a simple reduction of only gaspar-
ticle drag, as suggested by McKeen and Pugsley (2003), would sufce in
that case. For multidisperse uidized beds, the current class of TFMs still
lacks the capability of describing quantitatively particle mixing and segre-
gation rates. Recently, we have extended the KTGF to bidisperse systems
(Bokkers, 2005; Bokkers et al., 2006), and the next challenge would be to
extend this to general polydisperse systems. Also in the current KTGF, the
effect of particleparticle friction is not incorporated. A recent simulation
study using the DPM showed that particle friction has a large inuence on
the mixing behavior when a single bubble is injected into the system
(Bokkers et al., 2006). It was also found that the effects of lack of friction
could not be remedied by using a smaller coefcient of normal restitution,
which implies that friction should be taken into account explicitly in the
KTGF.
(iv) The models for describing industrial-scale gassolid ow are clearly still in
the preliminary stage. In this chapter, we have outlined two promising
approaches and noted that much remains to be done before their usefulness
as tools for the design and scale-up of chemical reactors can be ascertained.
In Section V.B, we have demonstrated the potential value of the ltering
approach. Many challenges still remain. It must be veried that results of
the type shown in Figs. 31 and 32 persist in 3D; simple predictive theories to
capture the lter-size dependence of the ltered drag coefcient and stresses
must be developed, and the viability of this approach should be validated
by comparison with experimental data.
For the discrete bubble model described in Section V.C, future work will be
focused on implementation of closure equations in the force balance, like
empirical relations for bubble-rise velocities and the interaction between bub-
bles. Clearly, a more rened model for the bubblebubble interaction, includ-
ing coalescence and breakup, is required along with a more realistic
description of the rheology of uidized suspensions. Finally, the adapted
model should be augmented with a thermal energy balance, and associated
closures for the thermophysical properties, to study heat transport in large-
scale uidized beds, such as FCC-regenerators and PE and PP gas-phase
polymerization reactors.
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 145
REFERENCES
Agrawal, K., Loezos, P. N., Syamlal, M., and Sundaresan, S. J. Fluid Mech 445, 151185 (2001).
Alder, B. J., and Wainwright, T. E. J. Chem. Phys. 27, 1208 (1957).
Alder, B. J., and Wainwright, T. E. J. Chem. Phys. 33, 1439 (1960).
Alder, B. J., Gass, D. M., and Wainwright, T. E. J. Chem. Phys. 53, 3813 (1970).
Allen, M. P., and Tildesley, D. J., Computer Simulations of Liquids. Oxford Science Publications,
Oxford, UK (1990).
Anderson, J. D., Computational Fluid Dynamics: The Basics with Applications. McGraw-Hill,
New York, USA (1995).
Anderson, T. B., and Jackson, R. Ind. Eng. Chem. Fundam. 6, 527 (1967).
Anderson, T. B., and Jackson, R. Ind. Eng. Chem. Fundam. 7, 12 (1968).
Anderson, T. B., and Jackson, R. Ind. Eng. Chem. Fundam. 8, 137 (1969).
Anderson, K., Sundaresan, S., and Jackson, R. J. Fluid Mech. 303, 327 (1995).
Andrews, A. T.IV, Loezos, P. N., and Sundaresan, S. Ind. Eng. Chem. Res. 44, 60226037 (2005).
Andrews, A. T. IV., and Sundaresan, S. Closures for ltered two-uid model equations of gaspar-
ticle ows, Manuscript in preparation (2006).
Auton, T. R., The dynamics of bubbles, drops and particles in motion in liquids, PhD thesis,
University of Cambridge (1983).
Bayle, J., Mege, P., and Gauthier, T., Dispersion of bubble ow properties in a turbulent FCC
uidized bed, in Fluidization X (M. Kwauk, J. Li, and W. -C. Yang Eds.), pp. 125.
Engineering Foundation, New York, USA (2001).
Beetstra, R., van der Hoef, M. A., and Kuipers, J. A. M. Drag force from lattice Boltzmann
simulations of intermediate Reynolds number ow past mono- and bidisperse arrays of
spheres, Manuscript submitted to AIChE J. (2006).
Beetstra, R., van der Hoef, M. A., and Kuipers, J. A. M. Numerical study of segregation using a new
drag force correlation for polydisperse systems derived from lattice Boltzman simulations,
Manuscript submitted the Chem. Eng. Sci. (2006, in press).
Beetstra, R., van der Hoef, M. A., and Kuipers, J. A. M., A lattice-Boltzmann simulation study of
the drag coefcient of clusters of spheres, Comput. Fluids 35, 966970 (2006).
Bhatnagar, P., Gross, E., and Krook, M. Phys. Rev. 94, 511 (1954).
Bird, G. A., Molecular Gas Dynamics and Direct Simulation of Gas Flows. Oxford University
Press, Oxford, UK (1976).
Bird, R. B., Stewart, W. E., and Lightfood, E. N., Transport Phenomena. John Wiley & Sons,
New York, USA (1960).
Boemer, A., Qi, H., Hannes, J., and Renz, U. Modelling of solids circulation in a uidised bed
with Eulerian approach. 29th IEAFBC Meeting in Paris, France, Nov. 2426, 1994
(1994).
Bokkers, G. A., Multi-level modelling of the hydrodynamics of gas phase polymerization reactors,
PhD thesis, University of Twente (2005).
Bokkers, G. A., Van Sint Annaland, M., and Kuipers, J. A. M., Comparison of continuum models
using kinetic theory of granular ow with discrete particle models and experiments: extent of
particle mixing induced by bubbles. Proceedings of Fluidization XI, May 914, 2004,
187194, Naples, Italy (2004).
Bokkers, G. A., Laverman, J. A., Van Sint Annaland, M., and Kuipers, J. A. M., Modelling of
large-scale dense gassolid bubbling uidised beds using a novel discrete bubble model,
Chem. Eng. Sci. 61, 55905602 (2006).
Brendel, L., and Dippel, S., Lasting contacts in MD simulations, in Physics of Dry Granular
Media (H. J. Herrmann, J. P. Hovi, and S. Luding Eds.), pp. 313318. Kluwer Academic
Publishers, Dordrecht (1998).
Campbell, C. S. Ann. Rev. Fluid Mech. 22, 57 (1990).
M.A. VAN DER HOEF ET AL. 146
Campbell, C. S., and Brennen, C. E. J. Fluid Mech. 151, 167 (1985).
Carnahan, N. F., and Starling, K. E. J. Chem. Phys. 51, 635 (1969).
Centrella, J., and Wilson, J. R. Astrophys. J. Suppl. Ser. 54, 229 (1984).
Chapman, S., and Cowling, T. G., The Mathematical Theory of Theory of Non-Uniform Gases.
Cambridge University Press, Cambridge, UK (1970).
Chen, C. P., Studies in two-phase turbulence closure modeling, Ph.D. Thesis, Michigan State Uni-
versity, USA (1985).
Chen, Z., Gibilaro, L. G., and Foscolo, P. U. Ind. Eng. Chem. Res. 38, 610 (1999).
Cundall, P. A., and Strack, O. D. L. Geotechnique 29, 47 (1979).
Dasgupta, S., Jackson, R., and Sundaresan, S. AIChE J 40, 215228 (1994).
Davies, R. M., and Taylor, G. I. Proc. R. Soc. Lond. textbfA200, 375390 (1950).
Delnoij, E., Kuipers, J. A. M., and van Swaaij, W. P. M. Chem. Eng. Sci. 52, 3623 (1997).
Detamore, M. S., Swanson, M. A., Frender, K. R., and Hrenya, C. M. Powder Technol 116, 190
(2001).
Ding, J., and Gidaspow, D. AIChE J 36, 523 (1990).
Dufty, J. W. J. Phys.: Condens. Mater. 12, 47 (2000).
Elghobashi, S. E., and Abou-Arab, T. W. Phys. Fluids 26, 931 (1983).
Elgobashi, S. Appl. Sci. Res. 48, 301304 (1991).
Ergun, S. Chem. Eng. Prog. 48, 89 (1952).
Ernst, M. H., and Dufty, J. W. J. Stat. Phys. 58, 57 (1989).
Erpenbeck, J. J., and Wood, W. W. J. Stat. Phys. 35, 321 (1984).
Ferschneider, G., and Mege, P. Revue de l Institut franc-ais du petrole 51, 301 (1996).
Frenkel, D., and Smit, B., Understanding Molecular Simulation: From Algorithms to Applica-
tions. Academic Press, San Diego, USA (1996).
Frisch, U., dHumie` res, D., Hasslacher, B., Lallemand, P., Pomeau, Y., and Rivet, J-P. Complex Sys
1, 649 (1987).
Frisch, U., Hasslacher, B., and Pomeau, Y. Phys. Rev. Lett. 56, 1505 (1986).
Geldart, D. Powder Technol 7, 285 (1973).
Gidaspow, D., Multiphase Flow and Fluidization: Continuum and Kinetic Theory Descriptions.
Academic Press, Boston, USA (1994).
Goldhirsch, I. Chaos 9, 659 (1999).
Goldschmidt, M. J. V., Beetstra, R., and Kuipers, J. A. M. Powder Technol. 142, 23 (2004).
Goldschmidt, M. J. V., Kuipers, J. A. M., and van Swaaij, W. P. M. Chem. Eng. Sci. 56, 571 (2001).
Hansen, J. -P., and McDonald, I. R., Theory of Simple Liquids. Academic Press, London, UK
(1986).
Harlow, F. H., and Amsden, A. A. J. Comput. Phys. 17, 19 (1975).
Hawley, J. F., Smarr, L. L., and Wilson, J. R. Astrophys. J. Suppl. Ser. 55, 211 (1984).
Helland, E., Occelli, R., and Tadrist, L. CR Acad. Sci. II B 327, 1397 (1999).
Helland, E., Occelli, R., and Tadrist, L. Powder Technol 110, 210 (2000).
Herrmann, H. J., and Luding, S. Contin. Mech. Thermodyn. 10, 189 (1998).
Hill, R. J., Koch, D. L., and Ladd, A. J. C. J. Fluid Mech. 448, 213 (2001a).
Hill, R. J., Koch, D. L., and Ladd, A. J. C. J. Fluid Mech. 448, 243 (2001b).
Hoomans, B. P. B., Granular dynamics in gassolids two-phase ows. Ph.D. dissertation, University
of Twente, Enschede, The Netherlands (2000).
Hoomans, B. P. B., Kuipers, J. A. M., and van Swaaij, W. P. M. Powder Technol 109, 41 (2000).
Hoomans, B. P. B., Kuipers, J. A. M., Briels, W. J., and van Swaaij, W. P. M. Chem. Eng. Sci. 47, 99
(1996).
Hoover, W. G., and Ree, F. H. J. Chem. Phys. 49, 3609 (1969).
Hrenya, C. M., and Sinclair, J. L. AIChE J 43, 853869 (1997).
Hunter, R. J. Foundations of Colloid Science. vol. IClarendon Press, Oxford, UK (1986).
Ishii, M., Thermo-Fluid Dynamic Theory of Two-Phase Flow. Eyrolles, Paris, France (1975).
Israelachvili, J., Intermolecular & Surface Forces. Academic Press, London, UK (1991).
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 147
Iwadate, M., and Horio, M., Agglomerating uidization of wet powders and group c powders: a
numerical analysis, in Fluidization IX (L. S. Fan, and T. Knowlton Eds.), pp. 293. En-
gineering Foundation, Durango, USA (1998).
Jenkins, J. T., and Savage, S. B. J. Fluid Mech. 130, 187 (1983).
Johnson, K. L., Contact Mechanics. Cambridge University Press, Cambridge, UK (1985).
Kafui, K. D., Thornton, C., and Adams, M. J. Chem. Eng. Sci. 57, 2395 (2002).
Kim, H., and Arastoopour, H. Powder Technol 122, 83 (2002).
Kobayashi, N., Yamazaki, R., and Mori, S. Powder Technol 113, 327 (2000).
Koch, D. L., and Hill, R. J. Annu. Rev. Fluid Mech. 33, 619 (2001).
Kuipers, J. A. M., and van Swaaij, W. P. M. Adv. Chem. Eng. 24, 227 (1998).
Kuipers, J. A. M., Hoomans, B. P. B, and van Swaaij, W. P. M. Hydrodynamic Modeling of Gas-
Fluidized Beds and their Role for Design and Operation of Fluidized Bed Chemical Reactors.
Proceedings of the Fluidization IX conference, 1530, Durango, USA (1998).
Kuipers, J. A. M., van Duin, K. J., van Beckum, F. H. P., and van Swaaij, W. P. M. Chem. Eng. Sci.
47, 1913 (1992).
Kunii, D., and Levenspiel, O., Fluid Engineering. Butterworth Heinemann series in Chemical
Engineering, London, UK (1991).
Ladd, A. J. C. J. Fluid Mech. 271, 285 (1994).
Ladd, A. J. C., and Verberg, R. J. Stat. Phys. 104, 1191 (2001).
Langston, P. A., Tu uzu n, U., and Heyes, D. M. Chem. Eng. Sci. 49, 1259 (1994).
Laux, H. Modeling of dilute and dense dispersed uid-particle ow, PhD Thesis, NTNU
Trondheim, Norway (1998).
Lettieri, P., Cammarata, L., Micale, G. D. M., and Yates, J. Int. J. Chem. Reactor Eng. 1, A5 (2003).
Lettieri, P., Newton, D., and Yates, J. G. Powder Technol 123, 221 (2002).
Lewis, E. W., and Bowerman, E. W. Chem. Eng. Prog. 48, 603 (1952).
Li, J., and Kuipers, J. A. M. Powder Technol 127, 173 (2002).
Li, J., and Kuipers, J. A. M. Chem. Eng. Sci. 58, 711 (2003).
Li, J., and Kuipers, J. A. M. Chem. Eng. Sci. 60, 1251 (2005).
Li, J., and Mason, D. J. Powder Technol 112, 273 (2000).
Liboff, R. L., Kinetic Theory. John Wiley & Sons, New York, USA (1998).
Link, J. M., Cuypers, L. A., Deen, N. G., and Kuipers, J. A. M. Chem. Eng. Sci. 60, 3425 (2005).
Lohse, D., Bergmann, R., Mikkelsen, R., Zeilstra, C., Van der Meer, D., Versluis, M., Van der
Weele, K., Van der Hoef, M. A., and Kuipers, J. A. M. Phys. Rev. Lett. 93, 198003 (2004).
Luding, S., Collisions & contacts between two particles, in Physics of Dry Granular Media (H. J.
Herrmann, J. P. Hovi, and S. Luding Eds.), p. 119. Kluwer Academic Publishers, Dordrecht
(1998) Chpt. 5.
Lun, C. C. K. Int. J. Multiphase ow 26, 1707 (2000).
Lun, C. K. K., Savage, S. B., Jeffrey, D. J., and Chepurniy, N. J. Fluid Mech. 140, 223 (1984).
Ma, D., and Ahmadi, G. J. Chem. Phys. 84, 3449 (1986).
Massimilla, L., Dons, G., and Zucchini, C. Chem. Eng. Sci. 27, 2005 (1972).
McKeen, T. R., and Pugsley, T. S. Powder Technol 129, 139 (2003).
McNamara, S., and Young, W. R. Phys. Fluids 4, 496 (1992).
Morgan, J. P., Taylor, R. W., and Booth, F. L. Powder Technol 4, 286 (1971).
Nieuwland, J. J., van Sint Annaland, M., Kuipers, J. A. M., and van Swaaij, W. P. M. AIChE J 42,
1569 (1996).
OBrien, T. J., and Syamlal, M Particle cluster effects in the numerical simulation of a circulating
uidized bed, in (A. Avidan, Ed.) Circulating Fluidized Bed Technology IV , Proceedings of
the Fourth International Conference on Circulating Fluidized Beds, Hidden Valley Confer-
ence Center and Mountain Resort, August 15, 1993, Somerset, PA, (1993).
Ouyang, J., and Li, J. Chem. Eng. Sci. 54, 2077 (1999).
Patankar, S. V., and Spalding, D. B. Int. J. Heat Mass Transfer 15, 1787 (1972).
Qian, Y., dHumie` res, D., and Lallemand, P. Europhys. Lett. 17, 479 (1992).
M.A. VAN DER HOEF ET AL. 148
Richardson, J. F., and Zaki, W. N. Trans. Instn. Chem. Eng. 32, 35 (1954).
Samuelsberg, A., and Hjertager, B. H. AIChE J 42, 1536 (1996).
Scha fer, J., Dippel, S., and Wolf, D. E. J. Phys. I Fr. 6, 5 (1996).
Schiller, L., and Nauman, A. V.D.I. Zeitung 77, 318 (1935).
Sinclair, J. L., and Jackson, R. AIChE J 35, 1473 (1989).
Smagorinsky, J. Monthly Weather Rev 91, 99164 (1963).
Succi, S., The Lattice Boltzmann Equation for Fluid Dynamics and Beyond. Oxford Science
Publications, Clarendon Press, Oxford (2001).
Sundaresan, S. AIChE J 46, 11021105 (2000).
Syamlal, M., MFIX Documentation: Numerical Techniques. DOE/MC-31346-5824. NTIS/
DE98002029 (1998).
Syamlal, M., Rogers, W., and OBrien, T. J., MFIX Documentation. U.S. Department of Energy,
Federal Energy Technology Center, Morgantown, WV (1993).
Tsuji, Y., Kawaguchi, T., and Tanaka, T. Powder Technol 77, 79 (1993).
Tsuo, Y. P., and Gidaspow, D. AIChE J 36, 885 (1990).
Van der Hoef, M. A., Beetstra, R., and Kuipers, J. A. M. J. Fluid Mech. 528, 233 (2005).
Van der Hoef, M. A., Van Sint Annaland, M., and Kuipers, J. A. M. Chem. Eng. Sci. 59, 5157
(2004).
van Swaaij, W. P. M., Chemical reactors, in Fluidization Academic Press, London, UK J. F.
Davidson, and R. Clift, Eds.), (1985).
Walton, O. R., Numerical simulation of inelastic frictional particleparticle interaction, in Par-
ticulate Two-Phase Flow (M. C. Roco Ed.), pp. 884. Butterworth-Heineman, Boston, USA
(2004) Chap. 25.
Walton, O. R., and Braun, R. L. J. Rheol. 30, 949 (1986).
Wen, C. Y., and Yu, Y. H. Chem. Eng. Prog. Symp. Ser. 62, 100 (1966).
Whitmore, R. L. J. Inst. Fuel 30, 328 (1957).
Woodcock, L. V. Ann. N.Y. Acad. Sci. 37, 274 (1981).
Xu, B. H., and Yu, A. B. Chem. Eng. Sci. 52, 2785 (1997).
Yang, N., Wang, W., Ge, W., Wang, L., and Li, J. Ind. Eng. Chem. Res. 43, 55485561 (2004).
Ye, M., Van der Hoef, M. A., and Kuipers, J. A. M. Chem. Eng. Res. Des. 83(A7), 833 (2005).
Zhang, D. Z., and VanderHeyden, W. B. Int. J. Multiphase Flow. 28, 805822 (2002).
Zhou, H., Flamant, G., Gauthier, D., and Flitris, Y. Chem. Eng. Res. Des. 81, 1144 (2003).
Zhou, H., Flamant, G., Gauthier, D., and Lu, J. Chem. Eng. Res. Des. 82, 918 (2004).
Zhou, L. X., and Huang, X. Q. Sci. China 33, 428 (1990).
MULTISCALE MODELING OF GAS-FLUIDIZED BEDS 149
THE DETAILS OF TURBULENT MIXING PROCESS AND
THEIR SIMULATION
Harry E.A. Van den Akker

Department of Multi-Scale Physics, Faculty of Applied Sciences, Delft University of


Technology, Delft, The Netherlands
I. Introduction 152
A. The Role of Turbulence 154
B. The Role of Computational Fluid Dynamics 155
C. The Scope of this Review 158
II. Various types of uid ow simulations 159
A. Direct Numerical Simulations 160
B. Large Eddy Simulations 161
C. Reynolds Averaged NavierStokes Simulations 163
D. The Simulation of Processes in a Turbulent
Single-Phase Flow 165
E. The Computational Fluid Dynamics of Two-Phase
Flows 167
III. Computational Aspects 171
A. Finite Volume Techniques 171
B. The Size of the Computations 173
C. Lattice-Boltzmann Techniques 175
D. A Mutual Comparison of Finite Volume and
Lattice-Boltzmann 176
IV. Boundary Conditions 178
A. Moving Boundaries 178
B. Curved Boundaries 180
C. The Domain and the Grid 181
V. Simulations of Turbulent Flows in Stirred Vessels 183
A. Turbulence Properties 183
B. Validation of Turbulent Flow Simulations 186
VI. Operations and Processes in Stirred Vessels 189
A. Mixing and Blending 190
B. Suspending Solids 192
C. Dissolving Solids 196
D. Precipitation and Crystallization 197
VII. Stirred GasLiquid and LiquidLiquid Dispersions 203
A. Bakkers GHOST! Code 204
B. Vennekers DAWN Code 205
C. Further Simulations 207
D. A Promising Prospect 209

Corresponding author E-mail: H.E.A.vandenAkker@TNW.TUDelft.nl


151
Advances in Chemical Engineering, vol. 31
ISSN 0065-2377
DOI 10.1016/S0065-2377(06)31003-4
Copyright r 2006 by Elsevier Inc.
All rights reserved.
VIII. Chemical Reactors 209
A. Mechanistic Micromixing Models 210
B. A Lagrangian Approach 211
C. A Eulerian Probabilistic Approach 213
D. A Promising Prospect 214
IX. Summary and Outlook 216
A. The Various Computational Fluid Dynamics Options 216
B. The Promises of Direct Numerical Simulations and
Large Eddy Simulations 217
C. An Outlook 218
Acknowledgements 222
References 223
Abstract
This chapter is devoted to turbulent mixing processes carried out
in - mainly - stirred vessels. It reviews rst a number of turbulent ow
characteristics as far as relevant to a wide variety of single-phase and
two-phase mixing processes and, secondly and most importantly, the
details of the advanced Computational Fluid Dynamics (CFD) tech-
niques required for simulating such processes with a large degree of
condence. The processes considered comprise blending, solids suspen-
sion, dissolution, precipitation, crystallization, chemical reactions, and
dispersing gases and immiscible liquids.
The emphasis in this chapter is on the fruitful application of Large
Eddy Simulations for reproducing the local and transient ow conditions
in which these processes are carried out and on which their performance
depends. In addition, examples are given of using Direct Numerical
Simulations of ow and transport phenomena in small periodic boxes
with the view to nd out about relevant details of the local processes.
Finally, substantial attention is paid throughout this chapter to the at-
tractiveness and success of exploiting lattice-Boltzmann techniques for
the more advanced CFD approaches.
I. Introduction
Mixing is an operation inherent to numerous processes encountered in the
chemical process industries. Mixing devices such as stirred tanks occur abun-
dantly in plants and processing facilities. This review focuses on stirred vessels
being operated under turbulent-ow conditions. Their design and scale-up, their
operation, and their often seemingly conicting performances under varying
HARRY E. A. VAN DEN AKKER 152
conditions have always troubled plant engineers and fascinated researchers. The
turbulent-ow phenomena and mixing rates encountered in stirred vessels as
well as the processes being dependent on proper and effective mixing have been
the subject of numerous research studies since the early 1950s (e.g., Rushton
et al., 1950; Kramers et al., 1953).
The early studies often documented Power NumberReynolds Number rela-
tionships and related to mixing and circulation times (Holmes et al., 1964;
Voncken et al., 1964). Many scale-up rules in terms of nondimensional numbers
were derived for a variety of operations such as blending, aeration (Westerterp
et al., 1963), and suspending solids (Zwietering, 1958), usually on the basis of
integral investigations for specic stirrer/vessel combinations. One was just
interested in the global ow eld characteristics and in the overall performance
of the stirred vessel as a whole. This already was a major step forward in
comparison with the ction of a continuous stirred tank (reactor) widely used in
the eld of chemical engineering.
Current industrial interests, however, are far beyond this basic level of under-
standing. Companies are continuously looking for process improvements
and for options of debottlenecking plants, in the context of improving per-
formance, protability, competitiveness, and sustainability. The result is an in-
creasing demand on local ow information since in real life many processes
exhibit substantial spatial variations in, e.g., bubble or drop size (Tsouris and
Tavlarides, 1994; Luo and Svendsen, 1996; Schulze et al., 2000; Venneker et al.
2002) or in crystal size (Ten Cate et al., 2000; Hollander et al., 2001a,b; Rielly
and Marquis, 2001). Further, the yield and selectivity of many chemical reactors
may depend on the rate the various chemical species involved are brought into
intimate contact and this rate may vary spatially as well (Bakker and Fasano,
1993; Bakker, 1996; Akiti and Armenante, 2004).
Such spatial variations in, e.g., mixing rate, bubble size, drop size, or crystal
size usually are the direct or indirect result of spatial variations in the turbulence
parameters across the ow domain. Stirred vessels are notorious indeed, due to
the wide spread in turbulence intensity as a result of the action of the revolving
impeller. Scale-up is still an important issue in the eld of mixing, for at least
two good reasons: rst, usually it is not just a single nondimensional number
that should be kept constant, and, secondly, average values for specic para-
meters such as the specic power input do not reect the wide spread in tur-
bulent conditions within the vessel and the nonlinear interactions between ow
and process. Colenbrander (2000) reported experimental data on the steady
drop size distributions of liquidliquid dispersions in stirred vessels of different
sizes and on the response of the drop size distribution to a sudden change in
stirred speed.
Knowledge of spatial variations in bubble, drop, and crystal sizes is often
desired or required, but extremely hard to obtain experimentally. Intrusive
measuring and sampling probes may disturb ow and process locally. Taking
samples may affect the sizes: in the sampling procedure, samples may experience
THE DETAILS OF TURBULENT MIXING PROCESS 153
ow conditions different from those at the sampling position as a result
of which particle size may change. Optical techniques may provide a solution
but under specic conditions (optical accessability and density) and in specic
geometries only. Knowledge of local ow variables and local mixing and trans-
fer rates is hard to obtain experimentally as well, since the turbulent ow in
most stirred vessels in industry is inhomogeneous and often time dependent, and
comprises a wide range of spatial scales and associated temporal scales.
CFD might provide a way of elucidating all these spatial variations in ow
conditions, in species concentrations, in bubble drop and particle sizes, and
in chemical reaction rates, provided that such computational simulations are
already capable of reliably reproducing the details of turbulent ows and their
dynamic effects on the processes of interest. This Chapter reviews the state of
the art in simulating the details of turbulent ows and turbulent mixing proc-
esses, mainly in stirred vessels. To this end, the topics of turbulence and CFD
both need a separate introduction.
A. THE ROLE OF TURBULENCE
The spatial scales in turbulent mixing range from the size of the impeller
(blade) down to the typical size of the smallest, so-called Kolmogorov eddies in
which the turbulent kinetic energy is dissipated into heat due to the action
of friction. In many applications, mixing at the small scales is of paramount
importance since many rate-limiting phenomena take place within these
Kolmogorov scale and are dominated by the dynamics of these Kolmogorov
eddies. Kresta and Brodkey (2004) present a valuable discussion on the role
of turbulence in mixing applications and on the role of time and length scales
and of scale-up rules. One could say that in turbulent mixing the details
really matter! Deriving local rates of energy dissipation within these Kolmogorov
eddies from experimental velocity data is at its best a tedious activity not viable
for chemical engineers in industry. Computational simulations may provide a
way out.
The turbulent-ow structures in stirred tanks are highly 3-D and complex
because of the complex geometry of the device. Vortical structures, high tur-
bulence levels, and high rates of energy dissipation particularly in the vicinity
of the impeller dominate the turbulent ow in stirred tanks. Under the action of
the revolving impeller, the uid is circulated through the tank. Bafes along
the tank wall prevent the liquid from carrying out a solid-body rotation about
the impeller axis and enhance mixing, partly via vessel-size macroinstabilities.
Turbulent ows in stirred vessels are very complex indeed.
As a result, the turbulent-ow eld in a stirred vessel may be far from
isotropic and homogeneous. Some of the cornerstones of turbulence theory,
however, start from the assumption that production and dissipation of tur-
bulent kinetic energy balance locally. In many chemical engineering ows, this
HARRY E. A. VAN DEN AKKER 154
assumption may not be satised everywhere. In the complex ows of most
process equipment, turbulence intensity and other turbulence properties vary
spatially. These spatial variations may induce ows, diffusion, and dispersion
across the ow domain (see, e.g., Ducci and Yianneskis, 2006). This implies that
turbulence is not necessarily dissipated where it has been produced: it rst may
be transported toward another part of the ow domain to get dissipated there.
As a result, strictly speaking, the common concepts of turbulence theory may
not be valid and applicable.
Turbulence research still focuses on canonical cases of high Reynolds Number
ows in order to deepen their understanding of the subject (the most important
unsolved problem of classical physics, according to the great physicist Richard
Feynman). Typical examples are grid/isotropic turbulence (in the absence of
walls), channel and pipe ow, and free shear layers and jets (see, e.g., Dimotakis,
2005) or rather simple hydraulic problems (Rodi, 1984). At the same time,
turbulence research usually refrains from considering practical chemical engi-
neering problems such as the complications and practicalities of stirred vessels.
Chemical engineers, however, have to nd practical ways for dealing with
turbulent ows in ow devices of complex geometry. It is their job to exploit
practical tools and nd practical solutions, as spatial variations in turbulence
properties usually are highly relevant to the operations carried out in their
process equipment. Very often, the effects of turbulent uctuations and their
spatial variations on these operations are even crucial. The classical toolbox
of chemical engineers falls short in dealing with these uctuations and its effects.
Computational Fluid Dynamics (CFD) techniques offer a promising alternative
approach.
B. THE ROLE OF COMPUTATIONAL FLUID DYNAMICS
Nowadays, the effects of, e.g., vessel design and operation conditions can be
assessed by means of computational simulations indeed. To mention just a few
examples: standard CFD simulations of stirred vessels are perfectly suited to
sort out the effect of varying impeller clearance in terms of eliminating dead
zones or reducing their size, or the effect of impeller speed on the degree of
solids suspension. Various commercial software vendors offer efcient CFD
packages, which in many cases can be seen as real workhorses for industrial
applications. CFD has tempestuously developed into a very versatile tool, not
only in the hands of uid ow experts, but also in those of chemical engineers.
Since computational resources have increased substantially at sharply de-
creasing costs, detailed computational information about the ows can now-
adays be obtained even at a fraction of the cost of the corresponding
experiments. Consequently, computational modeling of the ow has become an
attractive alternative route of describing ows and improving and scaling up
operations in stirred tanks. Papers such as Ditl and Rieger (2006) presenting
THE DETAILS OF TURBULENT MIXING PROCESS 155
design correlations drafted on the basis of experimental data collected in long
series of tests will soon be a thing of the past.
In the meantime, CFD can even do more. It may be easier to measure the
local and transient details of the turbulent ows in stirred vessels and the spatial
distributions in, e.g., mixing rates and bubble, drop, and crystal sizes com-
putationally than by means of experimental techniques! A good example is the
spatial distribution of the kinetic energy dissipation rate e (see Micheletti et al.,
2004). Such a questfrom commercial interests (market pull)in the details of
turbulent ows and associated processes then urges for really powerful CFD
techniques. On the reverse, the expectations as to the potential of pursuing a
computational route toward a better understanding of turbulent ows and
processes are also fed by the remarkable progress made recently in the eld of
CFD (technology push).
This paper deals with the advanced CFD of turbulent stirred vessels. The
advances attained in recent years in the eld of CFD really matter for the degree
of accuracy and condence at which the performance of stirred reactors and of
other operations carried out in stirred vessels can be simulated, as these pro-
cesses and operations may strongly depend on the details of the physics and
chemistry involved. The latter details require the more advanced CFD tech-
niques indeed.
Recently, Paul et al. (2004) compiled many engineering design principles de-
veloped in the eld of mixing over the last 30 years into the NAMF (North
American Mixing Forum) Handbook of Mixing. This Bible of Mixing pro-
vides within its 1,375 pages a wealth of information and design guidelines
for the practicing engineer who needs to both identify and solve mixing pro-
blems. It also contains a few chapters and sections reviewing achievements and
promises of CFD for the eld of mixing. In Chapter 5, Marshall and Bakker
(2004) present an overview on Computational Fluid Mixing; their review,
however, is rather limited in scope and in computational methods covered.
Patterson et al. (2004) devote some 30% of their Chapter 13 to computational
simulations of Mixing and Chemical Reactions. In the remaining Chapters of
the Handbook, CFD is largely presented as an immature technique which may
become relevant for the practicing engineer in due course only.
The present author thinks one can and should be much more positive about
the merits of CFD so far and about the term at which CFD will replace and
improve existing mixing correlations. This message was already passed on at
earlier occasions (Van den Akker, 1997, 2000). Substantial progress has been
made in exploiting CFD not just for simulating the turbulent-ow eld but also
with the view of representing the various processes carried out in stirred vessels.
Owing to the enormous growth in computer power and the proliferation of
models and numerical methods, CFD nowadays is very powerful and versatile.
This does not mean that every ow or process can be mimicked in all detail,
since Direct Numerical Simulations (DNS) resolving all turbulent uctuations
involved in ow and process will keep requiring insurmountable amounts of
HARRY E. A. VAN DEN AKKER 156
computer power and simulation times for the next decades. Current computer
power, however, already offers various ways out which are of inestimable value
to practicing engineers indeed.
First of all, the increased computer power makes it possible to switch to
transient simulations and to increase spatial resolution. One no longer has to be
content with steady ow simulations on relatively coarse grids comprising
10
4
10
5
nodes. Full-scale Large Eddy Simulations (LES) on ne grids of
10
6
10
7
nodes currently belong to the possibilities and deliver realistic repro-
ductions of transient ow and transport phenomena. Comparisons with quan-
titative experimental data have increased the condence in LES. The present
review stresses that this does not only apply to the hydrodynamics but relates
to the physical operations and chemical processes carried out in stirred vessels
as well. Examples of LES-based simulations of such operations and processes
are due to Hollander et al. (2001a,b, 2003), Venneker et al. (2002), Van Vliet
et al. (2005, 2006), and Hartmann et al. (2006).
A further option is to forget about simulating the ow and the processes in
the whole vessel and to zoom into local processes by carrying out a DNS for a
small box. The idea is to focus on the ow and transport phenomena within
such a small box, such as mass transport and chemical reactions in or around
a few eddies or bubbles, or the hydrodynamic interaction of a limited number
of bubbles, drops, and particles including their readiness to collisions and
coalescence. Examples of such detailed studies by means of DNS are due to Ten
Cate et al. (2004) and Derksen (2006b).
The effect of this small box being immersed in the dynamic ambiance of the
turbulent ow is mimicked by using periodic boundary conditions, which pro-
tect the inside of the box against restricting effects of the boundaries. By im-
posing typical (averaged) conditions of, e.g., ow, rate of energy dissipation,
temperature, species concentrations, and/or volume fractions, we then are ca-
pable of studying how the ow and the various processes of interest evolve as a
result of the governing models and equations. Where experimental techniques
often fail in elucidating the mechanisms behind certain phenomena and pro-
cesses, computational simulations are perfectly capable of doing this. This is a
most welcome aspect of computational simulations indeed.
As a matter of fact, one may think of a multiscale approach coupling
a macroscale simulation (preferably, a LES) of the whole vessel to meso or
microscale simulations (DNS) of local processes. A rather simple, off-line way
of doing this is to incorporate the effect of microscale phenomena into the
full-scale simulation of the vessel by means of phenomenological coefcients
derived from microscale simulations. Kandhai et al. (2003) demonstrated the
power of this approach by deriving the functional dependence of the single-
particle drag force in a swarm of particles on volume fraction by means of
DNS of the uid ow through disordered arrays of spheres in a periodic box;
this functional dependence now can be used in full-scale simulations of any
ow device.
THE DETAILS OF TURBULENT MIXING PROCESS 157
One may object that such computational simulations are too advanced and
too much time consuming for practicing engineers. Similar doubts as to the
usefulness of CFD for chemical engineers were raised in the early days of CFD.
Nowadays, however, CFD software is used throughout the chemical process
industries for a wide variety of applications indeed. This is at least partly due to
the development of very versatile and robust commercial CFD software that has
turned into a valuable tool for chemical engineers in industry. In addition, in the
last 15 years, the size of the computational simulations has increased substan-
tially, keeping more or less equal pace with Moores law saying the number of
transistors on integrated circuits doubles every 2 years. There is no reason why
computer power and CFDin a type of playing leapfrog processwould not
keep growing at the same pace. This means that what is advanced and time
consuming today will be a routine job in 3 or 5 years. Of course, academia may
have the lead in this development.
C. THE SCOPE OF THIS REVIEW
This review paper is restricted to stirred vessels operated in the turbulent-ow
regime and exploited for various physical operations and chemical processes.
The developments in the eld of computational simulations of stirred vessels,
however, are not separated from similar developments in the elds of, e.g.,
turbulent combustion, ames, jets and sprays, tubular reactors, and multiphase
reactors and separators. Fortunately, there is a strong degree of synergy and
mutual cross-fertilization between these various elds. This review paper focuses
on aspects specic to stirred vessels (such as the revolving impeller, the resulting
strong spatial variations in turbulence properties, and the macroinstabilities)
and on the processes carried out in them.
Because of the above interactions, it is impossible to distinctly mark the start
of the CFD of stirred vessels. The rst papers on CFD simulations of stirred
vessels were due to Harvey and Greaves (1982), Placek and Tavlarides (1985),
Placek et al. (1986), Middleton et al. (1986), and Ranade et al. (1989). Patterson
was certainly one of the pioneers exploring the options for computational simu-
lations of turbulent reactors (see, e.g., Patterson, 1985). From about 1990 on-
wards, the triennial European Conferences on Mixing, the bi-annual meetings
of the NAMF and the regular IChemE Fluid Mixing Events provided the oor
for numerous oral presentations on computational simulations in the mixing
eld as well as for benecial exchanges of ideas between academic researchers
and industrial users.
The results of some of the early simulations in the eld of stirred vessels are
still reported in this review paper. They may serve to illustrate the substantial
progress made since the early days. Several of the strong simplications of those
days are no longer required indeed. Most importantly, incorporating physical
HARRY E. A. VAN DEN AKKER 158
operations and chemical processes into the uid ow simulations has become
quite viable, as this review intends to demonstrate.
This review strongly focuses on the potential of LES and DNS for repro-
ducing not only the hydrodynamics of turbulent stirred vessels but also for
providing a basis for simulating a wide variety of physical and chemical pro-
cesses in this equipment. The rst journal paper presenting simulation results
obtained by means of LES for a stirred vessel was due to Eggels (1996), who was
also the rst to exploit a lattice-Boltzmann (LB) technique to this purpose. Of
course, LES and LB do not necessarily go along, LB being just an attractive
solution technique perfectly suited for parallel computing. Since, in the wake of
this pioneering Eggels paper, the topics of LES and LB have become leading
themes in the research group of the present author, this review contains many
references to work of Derksen et al. (from 1996 onwards) and to many PhD
theses and associated papers from this group. Gradually, however, LES (in a
Finite Volume, FV, context) is also receiving attention from other research
groups and from the commercial software vendors.
II. Various types of uid ow simulations
Computational uid dynamics techniques are not capable of fully resolving
the highly turbulent ow in most industrial applications within a reasonable
time span. In a DNS, all scales of the motion are simulated by means of
the classical Navier Stokes (NS) equations. For simulating turbulent ows, this
would require that the spacing of the computational grid be sufciently ne to
even resolve the smallest eddies the name of Kolmogorov is associated with. As
the Kolmogorov length scale, Z
K
, is as small as Re
3/4
times some macroscopic
dimension (see, e.g., Tennekes and Lumley, 1972), a 3-D DNS scales with Re
9/4
(e.g., Wilcox, 1993; Moin and Kim, 1997). This really limits the applicability of
DNS to rather low Reynolds numbers.
Even nowadays, a DNS of the turbulent ow in, e.g., a lab-scale stirred vessel
at a low Reynolds number (Re 8,000) still takes approximately 3 months on 8
processors and more than 17 GB of memory (Sommerfeld and Decker, 2004).
Hence, the turbulent ows in such applications are usually simulated with the
help of the Reynolds Averaged Navier Stokes (RANS) equations (see, e.g.,
Tennekes and Lumley, 1972) which deliver an averaged representation of the
ow only. This may lead, however, to poor results as to small-scale phenomena,
since many of the latter are nonlinearly dependent on the ow eld (Rielly and
Marquis, 2001).
The exponential increase in computational power along with a drastic drop in
price for computer hardware has led to the ability to solve industrial ows by
means of LES. The major advantage of LES is that closure is applied to just the
THE DETAILS OF TURBULENT MIXING PROCESS 159
uctuations that are smaller than the grid spacing, whereas the large turbulent
scales are solved explicitly. This means that grid spacing also acts as a low pass
lter: uctuations smaller than the grid spacing are ltered out, i.e., not resolved
in the simulation.
It is then assumed that due to this separation in scales, the so-called subgrid
scale (SGS) modeling is largely geometry independent because of the universal
behavior of turbulence at the small scales. The SGS eddies are therefore more
close to the ideal concept of isotropy (according to which the intensity of
the uctuations and their length scale are independent of direction) and, hence,
are more susceptible to the application of Boussinesqs concept of turbulent
viscosity (see page 163).
Large eddy simulations ask for a ne grid to realize the above separation in
scales. LES are therefore computationally more expensive than RANS-based
simulations. It is not an option to take refuge to coarse grids in order to reduce
simulation times, as LES on a coarse grid does not make sense physically. The
grid spacing should be such that at least (a substantial) part of the inertial
subrange can be resolved.
Large eddy simulations yield much more detailed simulation results indeed,
not only because the grids used easily comprise millions of grid cellsthese
numbers being larger than common in RANS simulations by at least one
order of magnitudebut also because the simulations are inherently transient
and reproduce the dynamics of a large proportion of the wide spectrum
of eddies. In fact, LES are positioned somewhere in between DNS and RANS.
Derksen (2003) carried out LES on a grid of 1.4 10
7
nodes while a DNS
would have required a grid of as many as 10
12
nodes, the latter number being
far beyond computational capabilities both currently and in the foreseeable
future.
We will now treat the various CFD options in some more detail.
A. DIRECT NUMERICAL SIMULATIONS
The term direct in Direct Numerical Simulations indicates that the ow is
fully resolved by solving, without any modeling, the classical NS equations
@v
@t
v rv
1
r
rp nr
2
v (1)
and that not any motion or eddy at whatever scale is ignored, provided that
the calculation grid is sufciently ne. The latter condition implies that rening
the grid would (hardly) change the ow eld resulting from the simulation. As
a matter of fact, laminar ows belong to the type of ows excellently viable
to DNS.
Other important targets for DNS are the turbulent ows at Reynolds num-
bers up to say 10,000 in simple geometries (such as straight channels or curved
HARRY E. A. VAN DEN AKKER 160
pipes), the local ow eld around a single particle, or the turbulent two-phase
ow in a periodic box of limited size, again under the proviso that the calcu-
lation grid is sufciently ne to capture all details of the ow at the scale of
particle or box.
B. LARGE EDDY SIMULATIONS
In LES, it is accepted that the ow is not fully resolved: turbulent eddies
smaller than the grid spacing D are not solved explicitly. These small eddies do
contribute, however, to the redistribution of momentum within the ow eld.
The resulting set of NS equations then runs as
@~ v
@t
~ v r~ v
1
r
r~ p nr
2
~ v r s (2)
where ~ v and ~ p now denote the variables to be resolved in the simulation. Note
that the latter term in this set of equations represents the effect of SGS stresses
that are not calculated explicitly. After having introduced the decomposition
s s
0

1
3
tr s I (3)
the rst term at the right-hand side of Eq. (3) is modeled with the help of an
effective SGS viscosity coefcient n
e
:
s
0
n
e
r~ v r~ v
T

(4)
while the second term at the right-hand side of Eq. (3) is incorporated into the
pressure term of Eq. (2).
This implies that all eddies larger than the grid size are explicitly resolved in a
LES. A ow eld obtained by means of LES therefore is inherently transient
and 3-D. To take advantage of the concept of LES, which is particularly aimed
at resolving a great deal of the time and length scales of a turbulent-ow eld,
ne grids should be used. This implies that LES applied to a ow domain of
some size are computationally quite demanding. Restricting LES to 2-D in view
of saving computational time does not make sense physically, as the dynamics
of turbulent ows in process equipment is inherently 3-D.
In dealing with the SGS terms, Revstedt et al. (1998, 2000) and Revstedt and
Fuchs (2002) did not use any model; rather, they assumed these terms were just
as small as the truncation errors in the numerical computations. This heuristic
approach lacks physics and does not deserve copying. A most welcome aspect of
LES is that the SGS stresses may be conceived as being isotropic, i.e., insensitive
to effects of the larger scales, to the way the turbulence is induced and to the
complex and varying boundary conditions of the ow domain. Exactly this
THE DETAILS OF TURBULENT MIXING PROCESS 161
feature renders modeling attractive, in contrast with modeling all turbulent
stresses as done in RANS-based simulations (see the next section).
The most widely used model for the SGS stresses is due to Smagorinsky
(1963) and involves a SGS eddy viscosity, n
e
, which is related to the local
resolved deformation rate
~
S:
n
e
c
2
s
D
2

~
S
2
q
(5)
with
~
S
2

1
2
r~ v r~ v
T

: r~ v r~ v
T

(6)
While the theoretical value (based on homogeneous, isotropic turbulence) of the
Smagorinsky coefcient c
s
amounts to 0.165 (Mason and Callen, 1986), in many
simulation studies lower values for c
s
proved to result in a better reproduction
of experimental data. This may have to do with the abundant presence of shear
ows in process equipment. Derksen (2003) reported that varying c
s
values in
the range 0.080.14 does not have a large impact on the simulation results. A
value of 0.12 is recommended.
At the basis of the Smagorinsky SGS model is the assumption of equilibrium
between production and dissipation of turbulent kinetic energy in the inertial
subrange of eddy sizes. In stirred tanks, however, there is hardly any position
where this equilibrium prevails. Furthermore, there is not a good reason why
the Smagorinsky coefcient should be constant across the ow domain. Other
more specic SGS models have therefore been proposed and also investigated
as to their impact on the resulting ow elds and turbulence characteristics.
Hartmann et al. (2004a) assessed the usability of an SGS model due to Voke
(more geared to low-Reynolds number turbulence), while Derksen (2001) in-
vestigated a so-called structure function SGS model for a turbulent viscosity that
depends on eddy size. Derksen (2006a) supplemented the standard Smagorinsky
SGS model with wall-damping functions to bring the eddy viscosity explicitly to
zero at solid walls, since in physical reality velocity uctuations and subgrid
stresses are zero at walls. Recently, FLUENT 6.2 came with several new SGS
models. Further renements in SGS modeling may be expected to improve the
accuracy of LES.
An inherent property of the LES approach is that the simulated ow eld is
no longer steady, but exhibits a transient character due to the presence and
motion of large-scale eddies. The LES methodology has proven to be a powerful
tool for studying and visualizing stirred tank ows (Eggels, 1996; Derksen et al.
1999; Bakker et al., 2000; Derksen, 2001; Bakker and Oshinowo, 2004), as it
inherently takes the unsteady and periodic behavior of the ow (around impeller
and bafes) into account.
HARRY E. A. VAN DEN AKKER 162
C. REYNOLDS AVERAGED NAVIER STOKES SIMULATIONS
The focus of RANS simulations is on the time-averaged ow behavior
of turbulent ows. Yet, all turbulent eddies do contribute to redistributing
momentum within the ow domain and by doing so make up the inherently
transient character of a turbulent-ow eld. In RANS, these effects of the full
range of eddies are made visible via the so-called Reynolds decomposition of
the NS equations (see, e.g., Tennekes and Lumley, 1972, or Rodi, 1984) of the
ow variables into mean and uctuating components. To this end, a clear dis-
tinction is required between the temporal and spatial scales of the mean ow
on the one hand and those associated with the turbulent uctuations on the
other hand.
Via this Reynolds decomposition and after subsequent averaging all terms of
the NS equations, the so-called turbulent or Reynolds stresses u
i
u
j
emerge in
the transport equations, where these stresses represent the additional averaged
momentum transport due to the eddies. These stresses may be resolved explicitly
from separate transport equations which in sufx notation (usual in the eld of
turbulence) look as follows:
@u
i
u
j
@t
u
k
@u
i
u
j
@x
k
P
ij
D
ij

ij
P
ij
(7)
in which the rst three terms of the right-term side denote the production,
diffusion, and dissipation of the turbulent stresses, respectively, while the last
term is the so-called pressure-strain term that represents the redistribution of the
turbulent kinetic energy among the three coordinate directions that makes the
turbulence more isotropic (Rodi, 1984). Several of these terms need modeling
for which a gamut of choices is available. In principle, this approach implies
the need of solving nine more partial differential equations per grid cell. As a
result, CPU times required for computational simulations on the basis of some
Reynolds Stress Model (RSM) are relatively high.
A cure against these longer CPU times is the Algebraic Stress Model (ASM)
described by, e.g., Rodi (1984) and used and recommended by, e.g., Bakker
(1992) and Bakker (1996). Most commercial codes do no longer support an ASM.
Usually, however, the stresses are modeled with the help of a single turbulent
viscosity coefcient that presumes isotropic turbulent transport. In the RANS-
approach, a turbulent or eddy viscosity coefcient, n
t
, covers the momentum
transport by the full spectrum of turbulent scales (eddies). Frisch (1995) recol-
lects that as early as 1870 Boussinesq stressed turbulence greatly increases vis-
cosity and proposed an expression for the eddy viscosity. The eventual set of
equations runs as
@V
@t
V rV
1
r
rP n n
t
r
2
V (8)
THE DETAILS OF TURBULENT MIXING PROCESS 163
In its turn, the turbulent viscosity may be position dependent and generally
may be modeled in terms of a model, very usually a ke model:
n
t
C
m
k
2

(9)
where k is the concentration of turbulent kinetic energy in J/kg (or m
2
/s
2
) and e
is the rate of dissipation (in W/kg, or m
2
/s
2
/s) of this turbulent kinetic energy.
These two concentrations k and e are generally conceived as the most important
parameters describing a turbulent-ow eld. In their turn, their spatial distri-
butions within the turbulent-ow domain may be calculated from the following
transport equations for k and e, respectively:
@k
@t
u
i
@k
@x
i
P
k
D
k
P
k
(10)
the right-hand side of which contains similar terms as the above transport
equations for the turbulent stresses, and
@
@t
u
i
@
@x
i
P

O (11)
in which the last term denotes the destruction of e. The interested reader is
referred to, e.g., Rodi (1984) for the meaning of all these right-hand terms and
their modeling. The assumption often used in classical turbulence theory that
production and dissipation of turbulent kinetic energy balance locally, is found
by putting in Eq. (10) all terms but the rst and third at the right-hand side
equal to zero.
These two transport equations for k and e form an inherent part of any ke
model of RANS-simulations. As the result of closing the turbulence modeling
such that no further unknown variables and equations are introduced, the
e-equation does contain some terms that are still the result of modeling, albeit at
the very small scales (e.g., Rodi, 1984).
The (isotropic) eddy viscosity concept and the use of a ke model are known
to be inappropriate in rotating and/or strongly 3-D ows (see, e.g., Wilcox,
1993). This issue will be addressed in more detail in Section IV. Some researchers
prefer different models for the eddy viscosity, such as the ko model (where o
denotes vorticity) that performs better in regions closer to walls. For this latter
reason, the ke model and the ko model are often blended into the so-called
Shear-Stress-Transport (SST) model (Menter, 1994) with the view of using these
two models in those regions of the ow domain where they perform best. In spite
of these objections, however, RANS simulations mostly exploit the eddy vis-
cosity concept rather than the more delicate and time-consuming RSM turbu-
lence model. They deliver simulation results ofin many casesreasonable or
sufcient accuracy in a cost-effective way.
HARRY E. A. VAN DEN AKKER 164
RANS-based simulations exploiting the eddy viscosity concept just reproduce
the average ow eld and the spatial distribution of turbulence properties such
as k and e. As such, RANS-based simulations are excellently suited for iden-
tifying dead zones, recirculatory ows, short-circuiting between entrance
and exit, and further undesired ow features. Even in transient RANS-based
simulations, however, it is not a priori clear which part of the uctuations
is temporally resolved and which part is taken care of by the turbulence
model. This inherent property of RANS-based simulations especially raises
concerns in the case of ows exhibiting no clear spectral separation between
low-frequency coherent motions (such as macroinstabilities, precessing vortices,
and trailing vortices) and turbulent uctuations (making part of the cascade of
eddies or whirls typical of turbulence); and: A mechanistic picture of turbu-
lence cannot be treated on the average since such ows are dynamic. (Praturi
and Brodkey, 1978).
Yet (steady) RANS-based simulations are attractive as they relatively cheaply
deliver a quick impression of the overall ow eld in the vessel. Effects on the
overall ow eld of varying the position of impeller, feed pipe, withdrawal pipe,
and/or heat coil can easily be explored.
Note that the Eqs. (1), (2), and (8) are really and essentially different due to
the absence or presence of different turbulent transport terms. Only by incor-
porating dedicated formulations for the SGS eddy viscosity can one attain that
LES yield the same ow eld as DNS. RANS-based simulations with their
turbulent viscosity coefcient, however, essentially deliver steady ow elds
and as such are never capable of delivering the same velocity elds as the
inherently transient LES or DNS, irrespectively of the renement of the com-
putational grid!
D. THE SIMULATION OF PROCESSES IN A TURBULENT SINGLE-PHASE FLOW
For simulating computationally the spatial and temporal evolution of both
physical and chemical processes in mixing devices operated in a turbulent single-
phase mode, two essentially different approaches are available: the Lagrangian
approach and the Eulerian technique. These will be explained briey.
In the Lagrangian approach, individual parcels or blobs of (miscible) uid
added via some feed pipe or otherwise are tracked, while they may exhibit
properties (density, viscosity, concentrations, color, temperature, but also vorti-
city) that distinguish them from the ambient uid. Their path through the
turbulent-ow eld in response to the local advection and further local forces (if
applicable) is calculated by means of Newtons law, usually under the assump-
tion of one-way coupling that these parcels do not affect the ow eld. On
their way through the tank, these parcels or blobs may mix or exchange mass
and/or temperature with the ambient uid or may adapt shape or internal
velocity distributions in response to events in the surrounding uid.
THE DETAILS OF TURBULENT MIXING PROCESS 165
In real life, the parcels or blobs are also subjected to the turbulent uctuations
not resolved in the simulation. Depending on the type of simulation (DNS,
LES, or RANS), the wide range of eddies of the turbulent-uid-ow eld is not
necessarily calculated completely. Parcels released in a LES ow eld feel both
the resolved part of the uid motion and the unresolved SGS part that, at
best, is known in statistical terms only. It is desirable that the forces exerted by
the uid ow on the particles are dominated by the known, resolved part of
the ow eld. This issue is discussed in greater detail in the next section in the
context of tracking real particles. With a RANS simulation, the turbulent
velocity uctuations remaining unresolved completely, the effect of the turbu-
lence on the tracks is to be mimicked by some stochastic model. As a result,
particle tracking in a RANS context produces less realistic results than in an
LES-based ow eld.
An early example of tracking uid parcels in a stirred tank can be found in
Bouwmans (1992) and Bouwmans et al. (1997). Bakker (1996) used a tracking
routine with the view to provide a Lagrangian description of micromixing in a
stirred tank chemical reactor. Lapin et al. (2004) recently described a compu-
tational strategy for travelling along the lifelines of single cells (i.e., tracking
them) in stirred bioreactors in order to characterize the dynamics of the hetero-
geneous cell population and to study the impact of spatial and dynamic varia-
tions in concentrations of substrates and products across the reactor.
The motions of the individual uid parcels may be overlooked in favor of a
more global, or Eulerian, description. In the case of single-phase systems, con-
vective transport equations for scalar quantities are widely used for calculating
the spatial distributions in species concentrations and/or temperature. Chemical
reactions may be taken into account in these scalar transport equations by
means of source or sink terms comprising chemical rate expressions. The per-
tinent transport equations run as
@T
@t
v rT kr
2
T q (12)
and
@c
@t
v rc Dr
2
c r (13)
In Eq. (13), r stands for the production (or consumption) of the species of
interest due to a chemical reaction, while in Eq. (12) q represents the heat
production, e.g., due to one of more chemical reactions. Equation (13) is often
referred to as the Convection-Diffusion-Reaction (CDR) equation.
Since turbulent uctuations not only occur in the velocity (and pressure) eld
but also in species concentrations and temperature, the convection diffusion
equations for heat and species transport under turbulent-ow conditions also
comprise cross-correlation terms, obtained by properly averaging products of
HARRY E. A. VAN DEN AKKER 166
velocityconcentration, velocitytemperature, concentrationtemperature, and
concentrationconcentration uctuations, on the analogy of the Reynolds
stresses in the NS equations (e.g., Patterson, 1985; Ranade, 2002). The challenge
is still to nd appropriate closure relations for these cross-correlation terms:
these may be either phenomenological or mechanistic (micromixing models) or
probabilistic (exploiting probability density functions, PDFs).
In stirred chemical reactors, unlike in combustion and with other gas-phase
reactions, these closure terms should take into account that for liquids the
Schmidt number (Sc n=D) is in the order 1001,000, and, hence, the role of
species diffusion at scales within the Kolmogorov eddies should explicitly be
taken into account (Kresta and Brodkey, 2004). Essential is that diffusion of
chemical species is governed by the Batchelor length scale Z
B
which obeys to
Z
B
Z
K
Sc
1=2
(14)
which for large Sc numbers is much smaller than the Kolmogorov length
scale Z
K
indeed. Such so-called micromixing processes have to be described by
means of micromixing models which will be dealt with in some greater detail in
Section VIII.
These convective transport equations for heat and species have a similar
structure as the NS equations and therefore can easily be solved by the same
solver simultaneously with the velocity eld. As a matter of fact, they are much
simpler to solve than the NS equations since they are linear and do not involve
the solution of a pressure term via the continuity equation. In addition, the
usual assumption is that spatial or temporal variations in species concentration
and temperature do not affect the turbulent-ow eld (another example of one-
way coupling).
E. THE COMPUTATIONAL FLUID DYNAMICS OF TWO-PHASE FLOWS
On the analogy of simulating the process of adding blobs of a miscible
liquid, two-phase ow in stirred tanks in a RANS context may be treated in
two ways: EulerLagrange or EulerEuler, with the second, dispersed phase
treated according to a Lagrangian approach and from a Eulerian point of view,
respectively.
1. Euler Lagrangian Approach
The EulerLagrangian approach is very common in the eld of dilute dis-
persed two-phase ow. Already in the mid 1980s, a particle tracking routine
was available in the commercial CFD-code FLUENT. In the EulerLagrangian
approach, the dispersed phase is conceived as a collection of individual particles
(solid particles, droplets, bubbles) for which the equations of motion can be
solved individually. The particles are conceived as point particles which move
THE DETAILS OF TURBULENT MIXING PROCESS 167
across the ow domain in response to the turbulent-ow eld of the carrier
phase. The consequence of treating particles as point particles is that the detailed
ow between the particles in response to the presence and motion of the par-
ticles is not resolved.
For the hydrodynamics forces acting on the particles, mostly single-particle
expressions are used; this implies that hydrodynamic interactions between par-
ticles are ignored completely. In many cases, direct interactions of particles
owing to collisions are ignored as well. All thisalong with the computational
burden that increases linearly with the number of particles being trackedmay
limit the practical applicability of the method to dilute systems with relatively
low volume fractions of dispersed phase and/or to ow domains of small size.
By feeding back the reactive forces exerted by the particles on the continuous
carrier phase, two-way coupling between the two phases is obtained.
Without much discussion, one may anticipate that particle inertia, gravity,
and drag force need to be part of the equations of motion describing the motion
and paths of the particles. Since a stirred tank is very inhomogeneous and
exhibits strong gradients in velocity, pressure, and stress elds, it is difcult to
estimate a priori if more exotic uid-particle forces such as the Saffman lift
force, the Magnus force, and the history force may play a role of signicance
either globally or locally. For a concise summary about these forces and for
expressions for these forces, the reader is referred to, e.g., Derksen (2003).
Added mass may be important for bubbly ows. It is obvious that in such a
Lagrangian approach distributions in particle size may easily be taken into
account.
As discussed earlier in the context of tracking miscible parcels or blobs,
particles travel through the resolved average or uctuating velocity eld as well
as feel the unresolved velocity uctuations. Since the major uid-particle force
may be the drag force, the uids velocity eld is of primary importance, the
turbulent velocity uctuations inclusively, whether or not they are resolved in
the simulation; uctuations in pressure and stresses may be secondary. Supply-
ing stochastic variations on top of the resolved velocity eld mimics the
unresolved uctuations and brings the expected seemingly erratic paths of the
particles about.
Of course, the role of the articially introduced stochastics for mimicking the
effect of all eddies in a RANS-based particle tracking is much more pronounced
than that for mimicking the effect of just the SGS eddies in a LES-based
tracking procedure. In addition, the random variations may suffer from lacking
the spatial or temporal correlations the turbulent uctuations exhibit in real life.
In RANS-based simulations, these correlations are not contained in the steady
spatial distributions of k and e and (if applicable) the Reynolds stresses from
which a typical turbulent time scale such as k/e may be derived. One may
try and cure the problem of missing the temporal coherence in the velocity
uctuations by picking a new random value for the uids velocity only after a
certain period of time has lapsed.
HARRY E. A. VAN DEN AKKER 168
In LES-based simulations, just the SGS part of the turbulence spectrum needs
to be mimicked by stochastics. The idea is that the resolved eddies have the
biggest impact on the paths of the particles indeed. This requires not only that
the resolved velocity uctuations should be stronger than the (estimated) SGS
uctuations, but also that the particle relaxation time should be larger than
the time-step applied in the LES. Meeting the latter criterion implies that the
time step of the LES is capable of keeping up with the time scale the particle
needs to respond to changes in the ow eld of the surrounding uid. In the
context of LES, picking a new random velocity only after (part of) a time k
sgs
/e
has lapsed may cure the problem of missing the coherence in the SGS eddies
when mimicking the effect of the SGS eddies on the particle tracks. Here, k
sgs
stands for the kinetic energy associated with the SGS eddies only and has to be
estimated. All these issues have been extensively discussed by, e.g., Derksen
(2003, 2006a).
2. Euler Euler Approach
In the complete Eulerian description of multiphase ows, the dispersed phase
may well be conceived as a second continuous phase that interpenetrates the
real continuous phase, the carrier phase; this approach is often referred to as
two-uid formulation. The resulting simultaneous presence of two continua is
taken into account by their respective volume fractions. All other variables such
as velocities need to be averaged, in some way, in proportion to their presence;
various techniques have been proposed to that purpose leading, however, to
different formulations of the continuum equations. The method of ensemble
averaging (based on a statistical average of individual realizations) is now gene-
rally accepted as most appropriate.
In the two-uid formulation, the motion or velocity eld of each of the two
continuous phases is described by its own momentum balances or NS equations
(see, e.g., Rietema and Van den Akker, 1983 or Van den Akker, 1986). In both
momentum balances, a phase interaction force between the two continuous
phases occurs predominantly, of course with opposite sign. Two-uid models
therefore belong to the class of two-way coupling approaches. The continuum
formulation of the phase interaction force should reect the same effects as
experienced by the individual particles and discussed above in the context of the
Lagrangian description of dispersed two-phase ow.
One therefore has to decide here which components of the phase inter-
action force (drag, virtual mass, Saffman lift, Magnus, history, stress gradients)
are relevant and should be incorporated in the two sets of NS equations.
The reader is referred to more specic literature, such as Oey et al. (2003), for
reports on the effects of ignoring certain components of the interaction force
in the two-uid approach. The question how to model in the two-uid formu-
lation (lateral) dispersion of bubbles, drops, and particles in swarms is relevant
THE DETAILS OF TURBULENT MIXING PROCESS 169
as well: various models are available. See also the discussion on page 204 as
to (Eq. (19)).
Another important issue in two-uid models is about modeling the turbulent
stresses under two-phase conditions (e.g., Van den Akker, 1998). At this mo-
ment, there is still no consensus on a universal two-phase turbulence model.
Generally, turbulence in the continuous phase may be generated by shear due to
large-scale velocity gradients felt by the continuous phase itself (just like in
single-phase ows) as well as by the presence and relative motion of the dis-
persed phase particles. The ratio at which these two mechanisms contribute to
the generation of turbulence may be an important factor in drafting a universal
model. In addition, the dispersed phase may exhibit a turbulent-ow behavior in
response to the turbulent motions of the continuous phase in which it is em-
bedded; this response may depend on several time scales and their interaction
(Oey et al., 2003). Lance et al. (1991) suggested that the motion of bubbles
promotes a return to isotropy (see also Van den Akker, 1998). A universal
model, however, is not available right now.
In dense systems such as encountered in solids suspension, particleparticle
interaction may be important as well. Then, the closure of solid-phase stresses is
an important issue for which kinetic theory models and solids phase viscosity
may be instrumental (see, e.g., Curtis and Van Wachem, 2004).
As a matter of fact, in comparison with the EulerLagrangian approach, the
complete Eulerian (or EulerEuler) approach may better comply with denser
two-phase ows, i.e., with higher volume fractions of the dispersed phase, when
tracking individual particles is no longer doable in view of the computational
times involved and the computer memory required, and when the physical in-
teractions become too dominating to be ignored. Under these circumstances, the
motion of individual particles may be overlooked and it is wiser to opt for a more
supercial strategy that, however, still has to take the proper physics into account.
Precisely owing to the continuum description of the dispersed phase, in
EulerEuler models, particle size is not an issue in relation to selecting grid cell
size. Particle size only occurs in the constitutive relations used for modeling the
phase interaction force and the dispersed-phase turbulent stresses.
In the case of droplets and bubbles, particle size and number density may
respond to variations in shear or energy dissipation rate. Such variations are
abundantly present in turbulent-stirred vessels. In fact, the explicit role of the
revolving impeller is to produce small bubbles or drops, while in substantial
parts of the vessel bubble or drop size may increase again due to locally lower
turbulence levels. Particle size distributions and their spatial variations are
therefore commonplace and unavoidable in industrial mixing equipment. This
seriously limits the applicability of common EulerEuler models exploiting just
a single value for particle size. A way out is to adopt a multiuid or multiphase
approach in which various particle size classes are distinguished, with mutual
transition paths due to particle break-up and coalescence. Such models will be
discussed further on.
HARRY E. A. VAN DEN AKKER 170
III. Computational Aspects
As the continuity equation, the NS equations, and the transport equations for
the turbulent variables are highly nonlinear, any CFD-calculation is essentially
iterative. Generally, the convergence rate of simulations depends on the number
of grid points and on the number of equations to be solved.
The number of grid points is associated with the desired or required degree of
detail and accuracy of the simulations given the type of simulation selected. In
running a DNS (provided it is doable) one is interested in a fully resolved ow
eld and the grid should be sufciently ne to catch all motions. In running a
LES, the grid should be sufciently ne for the subgrid scales to become in-
dependent of the macroow.
The number of equations to be solved is, among other things, related to the
turbulence model chosen (in comparison with the k e model, the RSM involves
ve more differential equations). The number of equations further depends on
the character of the simulation: whether it is 3-D, 2
1/2
-D, or just 2-D (see below,
under The domain and the grid). In the case of two-phase ow simulations, the
use of two-uid models implies doubling the number of NS equations required
for single-phase ow. All this may urge the development of more efcient so-
lution algorithms. Recent developments in computer hardware (faster proces-
sors, parallel platforms) make this possible indeed.
Various numerical techniques are available for discretizing the set of partial
differential equations to be solved. Discretizing essentially is the method of con-
verting the (partial) differential into algebraic equations by transforming (partial)
derivatives into nite-difference formulations. First of all, most (commercial) ow
simulation codes exploit the nite-volume (or nite-difference) method that has
been discussed extensively by Shyy (1994). An introduction to the concept can be
found with, e.g., Abbott and Basco (1989) and Shaw (1992). Many more tech-
niques are available for discretization, such as nite-element, spectral, arc-length,
and lter-scheme methods, which are beyond the scope of the present review. The
result is a (large) set of algebraic equations anyway: one algebraic equation per grid
point for each ow variable that connects the value of a particular ow variable at
a particular grid point with those at a number of neighboring grid points.
A. FINITE VOLUME TECHNIQUES
Most commercial CFD-codes are based upon the ideas and numerical
methods developed back in the 1970s at Imperial College London by Spalding,
Patankar, Gosman, and others:

the FV formulation (the natural balance formulation of the NS equations and


the continuity equation),
THE DETAILS OF TURBULENT MIXING PROCESS 171

the staggered grid concept (velocity components and scalar quantities such as
pressure are not dened at the same mesh points),

the common discretization schemes (central, upwind, quadratic upwind),

the pressurevelocity coupling according to the Semi-Implicit Method for


Pressure Linked Equations (SIMPLE) or related algorithms, and

the matrix solvers for the resulting sets of algebraic equations.


Useful reviews of these basic elements of CFD can be found with Patankar
(1980), Abbott and Basco (1989), Shaw (1992), and Ranade (2002). In the mean-
time, substantial progress has been realized in developing more effective and
powerful numerical techniques. Several of them have made it into the common
commercial CFD packages. Just as an example, several of the commercial ven-
dors have incorporated the option of collocated grids. A few more important
issues should be highlighted here.
First of all, the current iterative solution procedures solve the various mo-
mentum equations successively; although this highly segregated solution tech-
nique does not require large computer memory capacity, it does result in slow
convergence and, hence, in long CPU times. Reducing the degree of separation,
i.e., coupling several momentum equations and solving them simultaneously,
substantially speeds up the convergence rate of the calculation and reduces the
number of iterations required (Van Santen et al., 1996); the larger matrices do
make greater demands on computer memory.
Second, the various discretization schemes common to all commercial soft-
ware may suffer from the appearance of spurious oscillations (wiggles) in
regions of strong gradients and from numerical or false diffusion, or smearing
(Shyy, 1994). Finding the optimum scheme to avoid numerical diffusion is
often not very easy, especially in convection-dominated ows, when direction of
ow and grid orientation do not match everywhere. Wiggles may pose a serious
problem in solving turbulent ows, multiphase ows, and species transport as
the pertinent equations contain variables that are inherently positive, such as
k, e, phase fractions and concentrations. For these variables, wiggles may not
be tolerated as they could give rise to negative values which may result in
divergence of the algorithm. Although adaptive grid techniques, e.g., local grid
renement, may cure the problem of oscillations, the most promising among
the modern schemes seems to be the antidiffusion concept of Total Variation
Diminishing (TVD) that has proliferated in quite a few variants. By adding
subgrid points, TVD schemes do increase the calculational burden. In FLUENT
6.2, a bounded central differencing scheme is switched on for LES, replacing the
second-order upwind scheme which by default discretizes the convective terms
in RANS-based simulations.
Third, writing the discretized equations in matrix form results in sparse matri-
ces, often of a tri-diagonal form, which traditionally are solved by successive
under- or over-relaxation methods using the tri-diagonal matrix algorithm
HARRY E. A. VAN DEN AKKER 172
(TDMA). In the 1990s, however, two novel classes of methods have entered the
scene, viz.

Krylov subspace methods (such as Conjugate Gradient CG, the improved


BiCGSTAB, and GMRES) in combination with preconditioners for matrix
manipulations aimed at enhanced convergence, and

Multigrid methods in which convergence is improved by solving the equations


in, e.g., a nested iteration on multiple grids, starting, e.g., on a coarse grid and
then moving to a ner grid, et cetera.
For a more detailed description of these more sophisticated solution methods
the reader is referred to, e.g., Vuik (1993) and Shyy (1994). Adopting these more
sophisticated solution techniques becomes more important with increasing
number of partial differential equations to be solved, such as in two-phase ow
CFD (see, e.g., Lathouwers and Van den Akker, 1996; Van Santen et al., 1996;
Lathouwers, 1999).
In the last decade, most new algorithms, schemes, solvers, and precondi-
tioners have found their way into most commercial software packages. Multi-
grid solvers are also available. Furthermore, all CFD vendors have developed
powerful pre- and post processing routines.
B. THE SIZE OF THE COMPUTATIONS
The number of grid points used for CFD-studies of stirred tanks strongly
varies with the type of CFD and generally increases in time. While for Bakker
and Van den Akker (1994a, b) the maximum number of grid points amounted
to some 25,000, these days hundreds of thousands grid points are not uncom-
mon for RANS-simulations. Usually, DNS and LES are carried out with the
view to arrive at very ne spatial and temporal resolutions; then, substantially
more grid points are needed than with RANS-based simulations. For his LES,
Eggels (1996) used two uniform grids comprising 1.73 10
6
and 13.8 10
6
nodes, respectively. Derksen and Van den Akker (1999) used up to 6 10
6
grid
points, Hollander et al. (2000) some 2 10
6
, Derksen (2001), Lu et al. (2002),
and Hartmann et al. (2006) up to some 13.8 10
6
, while Ten Cate et al. (2000)
went as high as 35 10
6
grid points for an industrial crystallizer.
In addition, adding more transport equations for simulating physical or
chemical processes and running CFD for multiphase ows increases the size of
the computational jobs, both in the number of equations to be solved and in
terms of the difculty of getting the solution converged.
Advanced CFD simulations (both in terms of numbers of grid points and
partial differential equations) therefore require increasing amounts of computer
memory and CPU-time. Chemical engineers increasingly get familiar with the
idea of exploiting CFD, though still mostly of the RANS-type. Gradually, the
THE DETAILS OF TURBULENT MIXING PROCESS 173
advantages of LES become more obvious. These days, it becomes attractive to
speed up CFD simulations by running them on parallel computer platforms: a
cluster of pcs operating under, e.g., LINUX, or a massive parallel machine
(e.g., a CRAY). In this way, larger CFD simulations can be run overnight or
over the weekend. In addition, computer hardware (processors, memory) keep
becoming faster and cheaper. Just to illustrate what currently is becoming fea-
sible: Derksen (2006a) carried out his LES job on a platform of 12 CPUs, and
needed 7 h of wall-clock time for a single impeller revolution; as a mater of fact,
such a simulation yields an incredible degree of detail.
Parallel computing requires software made suitable for operating on parallel
processors. Decomposition of the ow domain and attributing each domain to a
separate processor is the common procedure. A fast communication between
the various processors is crucial, not to partly spoil the gain obtained by going
parallel.
Although commercial CFD vendors make versions of their software suited
for parallel computing, it is precisely the promises of parallel computing
that turn the Lattice Boltzmann techniques into an attractive option. The
locality of the collision operation in this technique (see below) along with a
high computational efciency allows for simulating complex ow systems
with high spatial and temporal resolution. Furthermore, the efciency of the
scheme hardly depends on the complexity of the ow domain. Compared to
the conventional FV solvers of the current commercial CFD codes, which
should be robust and suited for a wide variety of ow problems and ow
conditions, a LB solver is faster by at least one order of magnitude (see, e.g.,
Table I). All these properties make LB very attractive for LES in complex
geometries and even very competitive in comparison with the conventional FV
techniques.
TABLE I
NUMERICAL SETTINGS OF TWO SIMULATIONS ON A GAS CYCLONE WITH DIFFERENT NUMERICAL
SCHEMES: A RANS-BASED SIMULATION (HOEKSTRA, 2000) AND A LES DUE TO DERKSEN AND VAN DEN
AKKER (2000)
RANS LES
Numerical scheme Finite volume (FLUENT) Lattice-Boltzmann
Spatial and temporal properties 3-D, steady state 3-D, transient
Number of grid nodes 1.2 10
5
4.9 10
6
Total wall-clock time per simulation
a
117 h
b
864 h
c
Wall-clock time per grid node per step
d
1.7 10
4
s 6.1 10
6
s
Both simulations were run on an HP Convex S-Class computer.
a
Based on a simulation on one processor.
b
Converged after 20,000 iterations.
c
Required 95,500 time steps for obtaining reliable ow eld statistics.
d
One step denotes one time step or one iteration step.
HARRY E. A. VAN DEN AKKER 174
C. LATTICE-BOLTZMANN TECHNIQUES
Very promising with a view to simulating turbulent ows is the LB scheme.
The method originates from lattice-gas automata, which for uid ow appli-
cations were introduced by Frisch et al. (1986) and by McNamara and Zanetti
(1988) and has been studied, rened, and applied ever since. In spite of LB being
relatively new and still under construction, the method receives an increasing
amount of attention among scientists and engineers. Rather than taking the
discretized NS equations as a starting point for the numerical analysis, a discrete
system that mimics uid ow is designed directly.
Although LB therefore nowadays may be considered as a solver for the NS
equations, there is denitely more behind it. The method originally stems from
the lattice gas automaton (LGA), which is a cellular automaton. In a LGA, a
uid can be considered as a collection of discrete particles having interaction
with each other via a set of simple collision rules, thereby taking into account
that the number of particles and momentum is conserved.
In the LB technique, the uid to be simulated consists of a large set of
ctitious particles. Essentially, the LB technique boils down to tracking a col-
lection of these ctitious particles residing on a regular lattice. A typical lattice
that is commonly used for the effective simulation of the NS equations (Somers,
1993) is a 3-D projection of a 4-D face-centred hypercube. This projected lattice
has 18 velocity directions. Every time step, the particles move synchronously
along these directions to neighboring lattice sites where they collide. The col-
lisions at the lattice sites have to conserve mass and momentum and obey the so-
called collision operator comprising a set of collision rules. The characteristic
features of the LB technique are the distribution of particle densities over the
various directions, the lattice velocities, and the collision rules.
Such an approach is conceptually different from the continuum description of
momentum transport in a uid in terms of the NS equations. It can be dem-
onstrated, however, that, with a proper choice of the lattice (viz. its symmetry
properties), with the collision rules, and with the proper redistribution of par-
ticle mass over the (discrete) velocity directions, the NS equations are obeyed at
least in the incompressible limit. It is all about translating the above charac-
teristic LB features into the physical concepts momentum, density, and vis-
cosity. The collision rules can be translated into the common variable viscosity,
since colliding particles lead to viscous behavior indeed. The reader interested in
more details is referred to Succi (2001).
Lattice-Boltzmann is an inherently time-dependent approach. Using LB for
steady ows, however, and letting the ow develop in time from some starting
condition toward a steady-state is not a very good idea, since the LB time
steps need to be small (compared to, e.g., FV time steps) in order to meet the
incompressibility constraint.
The LB method is especially attractive if complexly shaped boundaries are
involved; see, e.g., Chen and Doolen (1998). Eggels (1996) was the rst to study
THE DETAILS OF TURBULENT MIXING PROCESS 175
the ow near a Rushton turbine in a bafed stirred tank reactor by means of a
LES, thereby using LB on a uniform grid. Derksen and Van den Akker (1998,
1999) performed a similar study for a tank reactor where the ow was driven by
either a pitched blade turbine or a Rushton turbine. These simulations revealed
a range of ow eld characteristics, such as the turbulence statistics and the
trailing vortex structure near the impeller. The computer code they developed is
based on a LB scheme proposed by Somers (1993). Boundary conditions can be
imposed through locally forcing the uid to a prescribed velocity, or (in case of
no-slip walls) by simple bounce-back rules for the ctitious LB particles. In an
agitated vessel, the action of the revolving impeller is described by means of an
adaptive force-eld procedure (Derksen and Van den Akker, 1998, 1999).
Lattice-Boltzmann can simply be conceived as an alternative method of nd-
ing a solution for the NS equations and is now being used for a growing number
of applications, ranging from laminar blood ow through irregularly shaped
veins (Artoli et al., 2003), the swirling ow in a vessel with a revolving bottom
(Derksen et al., 1996) and the vortex street behind a cylinder in cross-ow
(Derksen et al., 1997) to turbulent ow in industrial devices such as stirred
vessels, swirl tubes and reverse ow cyclones (Derksen and Van den Akker,
1998, 1999, 2000; Derksen, 2001, 2002a, b). There is also a commercial LB code
available (Power FLOW, EXA Corp., USA) that is used for ow problems
related to, e.g., the automotive and aerospace industries.
A disadvantage of LB techniques is that they require a (locally) uniform and
cubic computational grid. This raises problems with curved boundaries (see
Section III). Furthermore, the ow is calculated in great detail everywhere,
irrespective of the degree of turbulence; in some parts of the ow domain,
particularly in the more quiescent parts of the vessel, the LES may turn into
some type of DNS, the SGS contribution to the transport equation becoming
quite obsolete. Rohde (2004) developed a technique for local grid renement
which, however, has not yet been applied to the cylindrical geometry of a stirred
vessel. Lu et al. (2002) successfully attempted the use of a nonuniform grid in
their LB LES of the turbulent ow in a stirred tank.
Eggels and Somers (1995) used an LB scheme for simulating species transport
in a cavity ow. Such an LB scheme, however, is more memory intensive than a
FV formulation of the convective-diffusion equation, as in the LB discretization
typically 18 single-precision concentrations (associated with the 18 velocity
directions in the usual lattice) need to be stored, while in the FV just 2 or 3
(double-precision) variables are needed. Scalar species transport therefore can
better be simulated with an FV solver.
D. A MUTUAL COMPARISON OF FINITE VOLUME AND LATTICE-BOLTZMANN
Lattice-Boltzmann exhibits some inherent properties that favor the speed
of the simulations, and therefore make the method a serious candidate and
HARRY E. A. VAN DEN AKKER 176
alternative for FV techniques for simulating ows in process industry equip-
ment. These properties are:

The hydrodynamic quantities such as the velocities and velocity gradients are
determined locally in a LB code. As a result, the local character of the LB
technique is strongly in favor of (massive) parallellization of the computa-
tional job via domain decomposition: the communication between CPUs
relates to grid cells very near to the subdomain boundaries only. On the
contrary, in solving the NS equations iteratively with a FV solver the eld
properties result in long-range variations propagating across the boundaries
of subdomains and, hence, requiring intense communication between CPUs
dealing with neighboring subdomains.

Since LB describes the NS equations in the incompressible limit, the local


pressures can directly be obtained from the local densities and the speed of
sound. Hence, a distinct step for calculating the pressures via a Poisson
equation (derived from the continuity equation) as required in incompressible
FV schemes, is absent in LB.

Implementing complex boundaries in LB simulations is relatively easy com-


pared to doing so for FV techniques (Chen and Doolen, 1998). In view of the
usually complex boundaries of process equipment, particularly in the case of
stirred vessels with a revolving impeller, this is a distinctive advantage.

More basically, LB with its collision rules is intrinsically simpler than most
FV schemes, since the LB equation is a fully explicit rst-order discretized
scheme (though second-order accurate in space and time), while temporal
discretization in FV often exploits the CrankNicolson or some other mixed
(i.e., implicit) scheme (see, e.g., Patankar, 1980) and the numerical accuracy in
FV provided by rst-order approximations is usually insufcient (Abbott and
Basco, 1989). Note that fully explicit means that the value of any variable at
a particular moment in time is calculated from the values of variables at the
previous moment in time only; this calculation is much simpler than that with
any other implicit scheme.
A comprehensive and more extensive overview of the pros and cons of LB
with respect to applications can be found in Succi (2001). By the way, LB
methods are continuously improved to increase speed and accuracy, particularly
by introducing grid renement techniques and advanced techniques for arbi-
trarily shaped boundaries (e.g., Rohde et al., 2002, 2003, 2006; Rohde, 2004).
It makes sense to compare the implications (in terms of simulation times) of
using FV vs. LB in simulating turbulent-ow elds in process devices. Hoekstra
(2000) demonstrated the numerical implications of applying different numerical
schemes in an industrial application. He compared the outcome of his RANS
simulation for a gas cyclone with that of a LES carried out by Derksen and Van
den Akker (2000). Table I presents a number of numerical features of the two
types of simulations.
THE DETAILS OF TURBULENT MIXING PROCESS 177
Not only did the results of the LES agree much better with the experimental
data than those of the RANS simulation, the LB code was also roughly 30 times
faster than the FV code in terms of simulation time per grid node per step. To
put it in a more practical sense: performing this LES on, e.g., 18 processors with
the FV code would take 2 months, whereas a simulation as such with LB would
take 2 days only.
Another comparison is due to Van Wageningen et al. (2004) who performed a
similar study (in terms of the numerical scheme used) on unsteady laminar ow
in a Kenics
s
static mixer. They found that the LB code was 500600 times
faster than FLUENT in terms of simulation time per grid node per time step and
that FLUENT used about 5 times more memory than LB.
The difference in speed between a LB code and a FV code in the above studies
partly originates from their different character: the FV code in a general-
purpose commercial CFD code should be robust and suited in many applica-
tions, while the LB codes used are of a research type and usually strongly
dedicated and geared to a specic job.
IV. Boundary Conditions
Solving sets of (partial) differential equations inherently requires the spec-
ication of boundary conditions and, in case of transient simulations, also in-
itial conditions. This is not as simple as it looks like, especially for turbulent
ows in complex process equipment.
Whenever a free surface is present at some (mean) xed position, most CFD
codes assume it to be strictly at, while in the direction normal to the free
surface velocities and gradients of most variables are taken zero. Usually, this is
accomplished by dening mirror cells at the free surface. It is not clear what
the effect is of the use of such mirror cells on the ow eld in the upper part of
the vessel in comparison with real life where the surface is not necessarily at.
A. MOVING BOUNDARIES
An aspect of CFD in stirred vessels that needs separate discussion is the issue
of the revolving impeller and the way its motion is dealt with in the simulations.
In the early days, see, e.g., Bakker and Van den Akker (1994a), a black box
representing the impeller swept volume was often used in RANS simulations,
with boundary conditions in the outow of the impeller which were derived
from experimental data. The idea behind this approach was that such near-
impeller data are hardly affected by the rest of the vessel and therefore could be
used throughout. Generally, this is not the case of course. Furthermore, this
approach necessitates the availability of accurate experimental data, not only
HARRY E. A. VAN DEN AKKER 178
with respect to the average velocity components, but also for k and e. The latter
variable in particular can hardly be measured directly. Nowadays, this approach
is no longer used, also due to the steep increase in computer power which no
longer urges for such drastic simplications.
In those early days, when computer power was limited, often use was made of
a symmetry assumption: each quarter of the vessel containing one of the four
bafes at the vessel wall was supposed to behave identically; hence, a steady
ow in the RANS approach was simulated in just a quarter vessel. Such strong
simplications are no longer in use. Precessing vortices moving around the
vessel centerline contribute to ow unsteadiness and, therefore, exclude models
that just assume ow steadiness or allow for domain reductions through geo-
metrical symmetries. The most correct response to this ow unsteadiness is the
concept of LES.
Later on, in 1994, novel options were introduced such as Sliding Mesh (SM)
and Multiple Frames of Reference (MFR) in which the ow domain is divided
into two parts, each with their own meshing; one mesh is connected to the
stationary vessel wall and the other one to the revolving impeller. The inter-
action of the ow elds at either side of the interface between these two meshes
requires delicate bookkeeping of uxes and forces among cells moving with
respect to one another at the interface. Of course, these methods are a drastic
improvement over the black box description of the impeller swept area applied
in the early days of stirred vessel CFD. The mesh associated with the impeller in
SM is perfectly capable of simulating the unsteady ow around the impeller
blades including the trailing vortices.
The MFR technique introduced by Luo et al. (1993, 1994) starts from a
steady-state description of the ow eld and therefore ts in RANS-based simu-
lations only. The SM approach (e.g., Murthy et al., 1994; Bakker et al., 1997) is
a fully transient method, also applicable in LES (see, e.g., Bakker et al., 2000;
Jahoda et al., 2006; Gao and Min, 2006; Gao et al., 2006). Yeoh et al. (2004a, b)
adopted a Sliding and Deforming Mesh (SDM) technique in which the two
grids do not only slide with respect to each other but also feel shear resulting
in deforming interface cells. Keeping mass and momentum conserved in this
technique requires bookkeeping somewhat different from that with the SM
technique.
Generally, however, it is unclear what (with SM) the effect is of the position
of the interface between stationary and moving meshes on the simulated un-
steadiness of the overall ow in the vessel. In addition, simulations making use
of SM and MFR may suffer from slower or poor convergence. Of course, the
transient SM technique is more accurate, though at the cost of larger computer
time consumption. Montante et al. (2006) even reported that MFR yielded
unphysical ndings, viz. a region of opposite swirl.
Harvey et al. (1995) and Harvey and Rogers (1996) proposed a multiblock
impeller-tted grid structure for dealing with the exact geometry of the impeller.
The rst of these two papers introduces an approximate steady-state method
THE DETAILS OF TURBULENT MIXING PROCESS 179
that solves the viscous ow with the impeller at one position with respect to the
bafes ignoring its relative motion, while the second paper is about an unsteady
moving grid approach that now takes the relative motion of the impeller with
respect to the bafes into account. This approach, however, never made it into
to the turbulent-ow regime.
Ranade and Van den Akker (1994) and Ranade and Dommeti (1996) intro-
duced a snapshot approach in which the impeller was put in a standstill and the
revolving ow in the simulation was obtained by imposing velocity jets at one or
more xed positions of the impeller blades. At a rst glance, realistic ow elds
were obtained, energy dissipation levels and the total amount of energy dis-
sipated being in the right order of magnitude. A more detailed assessment of this
approach, however, reveals that the equations used are not invariant to the type
coordinate transformation used. Furthermore, elementary turbulence theory
(e.g., Tennekes and Lumley, 1972) indicates that, with a view to the ow eld in
the impeller swept volume, a jet issuing from the face of an impeller blade may
not be equivalent to a wake behind an impeller blade, as jets and wakes obey
different laws for their expansion in downstream direction. As a result, the
shape of the impeller connected zones of high k and e values may not be pre-
dicted condently. Ranades snapshot approach (still in use, see, e.g., Khopkar
et al., 2006) discussed further on page 207 should therefore be abandoned, in
spite of all explanations devoted to it (Ranade, 2002).
Applying Immersed or Embedded Boundary Methods (Mittal and Iaccarino,
2005) circumvents the whole issue of the friction between the more or less steady
overall ow in the bulk of the vessel and the strongly transient character of
the ow in the zone of the impeller. These methods are introduced below. In the
context of a LES, Derksen and Van den Akker (1999) introduced a forcing
technique for both the stationary vessel wall and the revolving impeller. They
imposed no-slip boundary conditions at the revolving impeller and at the sta-
tionary tank wall (including bafes). To this purpose, they developed a specic
control algorithm.
B. CURVED BOUNDARIES
In stirred vessels and static mixers the ow domain is bounded by complex
boundaries due to the curvature of containing walls, the revolving impeller axis
and/or static mixing elements.
While in the early days a staircase representation of a curved boundary in
a cubic grid was quite common, commercial CFD software nowadays ex-
ploits boundary tted meshing. The staircase representation usually was not a
problemin terms of mass conservation or of introducing artefacts such as
additional small eddiesas long as the steps did not involve more than a single
grid cell. The currently widely adopted boundary tted meshing, i.e., generating
body-conformal either structured or unstructured grids providing adequate
HARRY E. A. VAN DEN AKKER 180
local resolution with a minimum but large number of grid cells, requires either
commercial grid generating software or extensive coding (usually far beyond the
reach of academic groups). The impact of a poor grid on accuracy and on
convergence properties of the solver may not be underestimated.
Whenever a cubic grid is mandatoryeither due to coding limitations from
the part of academic groups or due to the inherent properties of, e.g., LB
techniquesand a staircase approach is to be avoided (e.g., for a revolving
impeller axis) one can take refuge to some immersed boundary method (see,
e.g., Mittal and Iaccarino, 2005). One may distinguish between

embedded boundary methods using cut cells: these methods adjust cell volume
and face areas to the geometry of the body in the ow domain; while in a FV
approach this method guarantees global and local conservation of mass and
momentum, it creates problems in an explicit formulation of the FV scheme;

immersed boundary methods exploiting boundary forcing methods: a bound-


ary is simply treated as a force exerted or felt at the position of the boundary;
this position may even be time dependent, e.g., in the case of a revolving
impeller; and

immersed boundary methods using ghost cells: ghost cells are boundary cells
the centers of which are lying outside the ow domain; in this approach,
values of variables in these ghost cells are required to satisfy, e.g., a Neumann
boundary condition (velocity gradients normal to a wall zero, velocity com-
ponent along a wall zero at the wall).
Which of the various immersed or embedded boundary methods is best
generally or for a particular caseis still an open question. Thornock and Smith
(2005) introduced a Cell Adjusted Boundary Force Method for a stirred vessel.
All methods proposed so far have their own pros and cons. Immersed boundary
methods are also exploited in LB techniques (e.g., Derksen and Van den Akker,
1999). Rohde (2004) investigated the use of triangular facets for representing a
spherical particle.
C. THE DOMAIN AND THE GRID
With a view to any simulation, a few important items have to be addressed.
First of all, it has to be decided whether the ow to be simulated is 2-D, 2
1/2
-D,
or 3-D. When the ow is, e.g., axis-symmetrical and steady, a 2-D simulation
may sufce. For a ow eld in which all variables, including the azimuthal
velocity component, may not depend on the azimuthal coordinate, a 2
1/2
-D
simulation may be most appropriate. Most other cases may require a full 3-D
simulation. It is tempting to reduce the computational job by casting the
3-D ow eld into a 2-D mode. The experience, however, is that in 2-D simu-
lations the turbulent viscosity tends to be overestimated; in this way, the ow
THE DETAILS OF TURBULENT MIXING PROCESS 181
may become more or less diffusion dominated and less capable of sustaining
a transient behavior. Furthermore, between 2-D and 3-D, the wall to volume
ratio is different, turbulence intensity may go down, and hydrodynamic stability
may be affected. Of course, the above remarks apply to RANS-based simu-
lations only, as LES are inherently transient and 3-D.
A second choice to be made relates to the size of the ow domain. It may
be worthwhile to limit the calculational job by reducing the size of the ow
domain, e.g., by identifying an axis or plane of symmetry, or, in a cylindrical
vessel with bafes mounted on the wall, due to periodicity in the azimuthal
direction. Commercial software accomplishes these choices by means of sym-
metry cells and cyclic cells, respectively; although such choices reduce the size
of the simulation, they may eliminate the possibility of nding the real (asym-
metric, unstable, or transient) 3-D ow eld. The presence of feed pipes or drain
or withdrawal pipes may also make the use of symmetry or cyclic cells impos-
sible. Again, this issue only plays a role in RANS-type simulations.
A third issue is how many grid cells should be used for the domain size
selected. In general, using more computational cells implies more detailed in-
sight in the ow eld as a result of the simulation, at the cost of longer CPU
times (although sometimes convergence rate may increase as a result of in-
creasing the number of grid cells.) This is a trade-off that is to be decided upon,
each time a new ow eld is to be simulated. In some cases, a stepwise approach
may be pursued to zoom into a particular zone within a ow device (see, e.g.,
Stekelenburg et al., 1994).
A fourth issue is the use of local grid renement: many commercial codes
offer this modality, often with unstructured grids. The rationale behind the idea
of introducing local grid renement techniques is that the grid is only rened in
those parts of the calculation domain in which the ow exhibits strong spatial or
temporal gradients. By doing so, one does not waste grid cells in parts of the
domain where the ow does not vary signicantly. Rohde et al. (2006) explored
the use of a generic, mass-conservative local grid renement technique within a
LB technique. He successfully attempted this technique for various rather sim-
ple cases; so far, it has not been exploited in a cylindrical stirred vessel in which
the ow would require a ne grid not only in the impeller swept region but also
near to the bafes.
Anyhow, the result of any ow simulation should be grid independent, i.e.,
the ow eld should not be different if a ner grid is used for the simulation.
This test is to be carried out always and everywhere, and may really be described
as a conditio sine qua non.
Finally, it is good to quote Shaw (1992, p. 227): the most important asset in a
CFD analysis process is the analyst, who actually translates the engineering
problem into a computational simulation, runs the CFD solver and analyzes the
results. It is the skill of this person, or set of persons, that will determine whether
all the hardware and software will be utilized in the best possible way and
produce good quality results. CFD is certainly not a panacea that may solve all
HARRY E. A. VAN DEN AKKER 182
possible design and optimization problems. It rather is a tool that in the hands
of a well-trained professional may provide valuable insights in the local phe-
nomena and processes, which take place in (bio)reactors and all sorts of process
equipment including mixing devices. CFD-simulations may provide a good
starting point for a mechanistic description of operations such as blending,
suspending solids, dispersing gas, and carrying out (bio)chemical reactions.
V. Simulations of Turbulent Flows in Stirred Vessels
When exploiting computational techniques for studying the mixing perform-
ance of stirred tanks and static mixers, the rst question is how well the
turbulent-ow eld and the characteristic turbulence properties are predicted
by the various forms of CFD discussed above. The different forms of CFD
(RANS, LES, DNS) and the various strategies, turbulence models, and sub-
models used by them may suffer from a different degree of validity in the
various mixing devices considered. The impact of a limited validity of the
models used may vary from case to case. That is why validation of simulation
results is still an urgent issue. Sometimes, the results of computational simu-
lations just look qualitatively correct. It is important to check code performance
by means of quantitative experimental data as to average quantities as well as
local and transient uctuations.
The next question then is whether the processes taking place inside this tur-
bulent-ow eld can be modeled with condence. We will now rst consider the
rst question.
A. TURBULENCE PROPERTIES
It is worthwhile to verify whether or not some basic assumptions of turbu-
lence theory (local equilibrium between production and dissipation of turbu-
lence, isotropy, homogeneous turbulence) which are used in modeling certain
aspects of momentum transport via turbulent eddies are met with in stirred
tanks. To be honest, at hardly any position in a stirred tank the production rate
of turbulent kinetic energy balances the rate of its dissipation. Production of
turbulence mainly takes place in the impeller swept volume, while a much larger
part of the vessel takes part in dissipating the turbulence. In addition, the action
of a revolving impeller and its interaction with nearby bafes turns the ow
intrinsically unsteady. As a result, the turbulent ow in a stirred vessel is cer-
tainly not an equilibrium ow, as presumed in using, e.g., the Smagorinsky SGS
model in LES.
Most RANS-based simulations make use of the ke model for taking into
account the momentum transport due to the turbulent eddies. This model is an
THE DETAILS OF TURBULENT MIXING PROCESS 183
eddy-viscosity model and as such it assumes isotropic turbulent transport. The
question is whether everywhere in a stirred tank the turbulent ow is locally
isotropic. This issue might have to be explained a bit further as local isotropy
should be distinguished from just isotropy. Local isotropy means that the
uctuations can be modeled with an (isotropic) eddy viscosity, while isotropy
has been dened as having the same uctuation levels in the three coordinate
directions. Local isotropy does not imply isotropy of the uctuations: a ke
model can predict nonisotropic uctuations.
The question whether or not stirred tank ow is locally isotropic, may be
investigated with the help of a LES which resolves a great deal of the Reynolds
stresses. To this end, the Reynolds stress data are best presented in terms of the
so-called anisotropy tensor a
ij
and its invariants A
1
, A
2,
and A
3
.
The anisotropy tensor is related to the turbulent stresses, of course, and is
dened as
a
ij

u
i
u
j
k

2
3
d
ij
(15)
Its rst invariant A
1
is equal to zero by denition. The second and third in-
variants of this tensor are A
2
a
ij
a
ji
and A
3
a
ij
a
jk
a
ki
, respectively. The range
of physically allowed values of A
2
and A
3
is bounded and represented by the so-
called Lumley triangle in the (A
3
, A
2
) plane (Lumley, 1978). The dis-
tance|A| O(A
2
2
+A
3
2
) from the isotropic state, i.e., from the origin (A
2
0,
A
3
0), is a measure of the degree of anisotropy. See also Escudie and Line
(2006) for a more extensive discussion as to how to quantify and visualize how
different from isotropic turbulence a stirred vessel is.
Phase-averaged values of|A|in a plane midway between two bafes of a
stirred tank have been plotted in Fig. 1 (from Hartmann et al., 2004a) for two
different SGS models (Smagorinsky and Voke, respectively) in LES carried out
in a LB approach. The highest values, i.e., the strongest deviations from isot-
ropy, occur in the impeller zone, in the boundary layers along wall and bottom
of the tank, and at the separation points at the vessel wall from which the
anisotropy is advected into the bulk ow. In the recirculation loops, the tur-
bulent ow is more or less isotropic.
Fig. 2 (also from Hartmann et al., 2004a) shows how the values of the in-
variants A
2
and A
3
found in a LES using a Smagorinsky SGS model are dis-
tributed within the Lumley triangle. Most but not all of the points are clustered
in the lower part of this triangle. This implies that in a large part of a stirred
tank the turbulent ow is more or less isotropic and the use of a turbulent
viscosity and a ke model are permitted. On the whole, however, turbulent ow
is not really isotropic. This may also explain why in RANS-based simulations
turbulent kinetic energy levels generally are underestimated.
As far as e is concerned: it is extremely difcult to measure its spatial dis-
tribution and its wide range of values experimentally. LES may provide an
HARRY E. A. VAN DEN AKKER 184
FIG. 1. Phase-averaged plots of the anisotropy distance|A|in a plane midway between two bafes
in a stirred vessel provided with a Rushton turbine, as obtained by means of LES, with two different
SGS models: (a) the Smagorinsky model; (b) the Voke model. Reproduced with permission from
Hartmann et al. (2004a).
FIG. 2. This plot shows to which degree, according to a LES, the turbulence in a plane midway
between two bafes in a stirred vessel provided with a Rushton turbine can be typied. For clarity,
not all grid points in such a plane have been used for this plot. According to Lumley (1978), the
borders represent different types of turbulent ows: 3-D isotropic turbulence, 2-d axis-symmetric
turbulence, 2-D turbulence, and 1-D turbulence. Most but not all points are concentrated in the
lower part of this Lumley triangle. Reproduced with permission from Hartmann et al. (2004a).
THE DETAILS OF TURBULENT MIXING PROCESS 185
attractive alternative. Micheletti et al. (2004) present a valuable discussion on
this issue.
B. VALIDATION OF TURBULENT FLOW SIMULATIONS
In view of the different requirements as to computer power, it is very worth-
while to compare the outcome of RANS and LES simulations mutually and/
or with quantitative experimental data. Several authors have done this, e.g.,
Derksen and Van den Akker (1998, 1999), Derksen (2001), Lu et al. (2002),
Ranade (2002), Derksen (2002b), and Yeoh et al. (2004a,b). In most cases, the
experimental data have been obtained by means of Laser-Doppler Anemometry
(LDA, or LDV): see, e.g., Yianneskis et al. (1987), Wu and Patterson (1989),
Scha fer et al. (1997, 1998), and Derksen et al. (1999). In this review, we will
mainly refer to the validation study due to Hartmann et al. (2004a).
In comparing RANS results and LES results with LDA data, the focus is rst
on the global, phase-averaged ow eld: see Fig. 3 (from Hartmann et al.,
2004a) that relates to a plane midway between two bafes. The two types of
simulations capture the dominant ow feature, viz. the two large circulation
loops, more or less to the same extent. The RANS simulation better predicts the
position of the point where the upper loop separates from the vessel wall. As far
as turbulent kinetic energy (k) levels are concerned, k-values obtained in a LES
relate to the resolved turbulent eddies. Hartmann et al. (2004a) argued that the
SGS uctuations hardly contribute to the k-levels. It is evident from Fig. 4 that
the RANS simulation underestimates the kinetic energy levels created by the
blades of the Rushton impeller. The same conclusion applies to phase-resolved
kinetic energy data (see Fig. 5). The RANS simulation is hardly capable of
catching the remainders of the trailing vortices created by the preceding impeller
blade, while LES nicely reproduces this succession of vortices. Whenever one is
interested in details of the turbulent-ow eld because they may affect the
performance of the mixing device, one should really consider carrying out a
LES. The ndings due to Yeoh et al. (2004a) are completely in line with this.
As long as the interest is in elds of averaged velocity components and in
overall mixing patterns, RANS-based simulations may sufce. Examples of
such satisfactory simulation results are plentiful, e.g., Marshall and Bakker
(2004) and Montante et al. (2006). When, however, the interest is in the details
of the turbulent-ow eld and in processes affected by these details, LES is
the option to be preferred. From the ndings reproduced above and from the
validation studies of Derksen and Van den Akker (1999) and Derksen (2001) the
general conclusion is that, as long as the spatial resolution is sufcient, LB LES
deliver results in excellent agreement with experimental turbulence data.
Large eddy simulations explicitly resolves the inherently unsteady character
of the turbulent ow in a stirred tank into account, including the periodic
phenomena associated with the motion of the impeller and their interaction with
HARRY E. A. VAN DEN AKKER 186
FIG. 3. Velocity vector elds and levels of turbulent kinetic energy in a plane midway between two
bafes in a stirred vessel, according to LDA data (a); a RANS-based simulation (b); and two LES (c)
and (d). Reproduced with permission from Hartmann et al. (2004a).
THE DETAILS OF TURBULENT MIXING PROCESS 187
the bafes. Global as well as subtle ow features are in quantitative agreement
with experimental data. Typical examples of such a quantitative agreement
arefor six-bladed Rushton turbinesthe path along which the trailing vor-
tices developing at the impeller blades are swept into the bulk of the tank as well
as the turbulent kinetic energy levels in the wakes of the impeller blades; in a
vessel equipped with a pitched-blade turbine, we could mention the primary and
secondary recirculations in the phase-averaged ow eld.
Derksen (2001) and Hartmann et al. (2004a) demonstrated that various
choices in SGS modeling did not have a big impact on the quality of the ow
eld predictions. In addition, all these studies did not reveal a signicant effect
of the value of the Smagorinsky coefcient within the range 0.080.14.
One of the complications in stirred tank ows is the presence of macroin-
stabilities (i.e., low-frequency mean ow variations) that may affect the mixing
performance. Various authors have distinguished between various types and in-
vestigated their occurrence and their frequencies under varying operating con-
ditions and with several types of vessels and impellers (Yianneskis et al., 1987;
Haam et al., 1992; Myers et al., 1997; Hasal et al., 2000; Nikiforaki et al., 2002).
FIG. 4. Phase-averaged plots of turbulent kinetic energy in the vicinity of the Rushton impeller as
found in different types of simulations as indicated. Reproduced with permission from Hartmann
et al. (2004a).
HARRY E. A. VAN DEN AKKER 188
Roussinova et al. (2003) and Hartmann et al. (2004b) found that the LES
methodology is excellently capable of reproducing various types of macroin-
stabilities. While Nikiforaki et al. (2002) experimentally found precessing fre-
quencies in the range 0.0130.018 N, the LES of Hartmann et al. (2004b) arrived
at 0.0228 and 0.0255 N as the dominant frequencies, where N stands for the
impeller speed (in number of revolutions/s). The discrepancy between experi-
mental and numerical frequencies is a challenge for further improving parti-
cularly the SGS-model and some numerical settings.
VI. Operations and Processes in Stirred Vessels
In simulating physical operations carried out in stirred vessels, generally one
has the choice between a Lagrangian approach and a Eulerian description.
While the former approach is based on tracking the paths of many individual
uid elements or dispersed-phase particles, the latter exploits the continuum
concept. The two approaches offer different vistas on the operations and require
different computational capabilities. Which of the two approaches is most
FIG. 5. Phase resolved plots of velocity vector eld and turbulent kinetic energy in a plane 15
o
behind an impeller blade (obtained by sampling data only when the measuring point is at the
specied position with respect to the impeller blades). Not all vectors have been plotted for clarity.
Reproduced with permission from Hartmann et al. (2004a).
THE DETAILS OF TURBULENT MIXING PROCESS 189
suited is hard to say: it depends on the details of the issue of interest and on the
computational power available.
In whichever approach, the common denominator of most operations in
stirred vessels is the common notion that the rate e of dissipation of turbulent
kinetic energy is a reliable measure for the effect of the turbulent-ow charac-
teristics on the operations of interest such as carrying out chemical reactions,
suspending solids, or dispersing bubbles. As this e may be conceived as a con-
centration of a passive tracer, i.e., in terms of W/kg rather than of m
2
/s
3
, the
spatial variations in e may be calculated by means of a usual transport equation.
In the context of the RANS-methodology, this e is also required for solving
the spatial distributions of the velocity components, while in LES e just serves
the purpose of providing a basis for modeling the operation(s) of interest.
Even when the number of grid cells in a LB LES simulation of a stirred vessel
1.1 m
3
in size amounts to some 36 10
6
grid cells, this implies a cell size, or grid
spacing, of 5 mm only. Even a cell size of just a few millimeters makes clear that
substantial parts of the transport of heat and species as well as all chemical
reactions take place at scales smaller than those resolved by the ow simula-
tion. In other words: concentrations of species and temperature still vary and
uctuate within a cell size. The description of chemical reactions and the trans-
port of heat and species therefore ask for subtle approaches to these SGS
uctuations.
A. MIXING AND BLENDING
One of the most crucial (design) parameters in blending two miscible liquids
or distributing a particular miscible addition over a heel of liquid is the so-called
mixing time, i.e., the time needed to achieve complete homogenization or a
predetermined degree of homogeneity (see, e.g., Grenville and Nienow, 2004).
In this review, the focus is on blending operations carried out with low viscosity
uids in the turbulent regime.
Kramers et al. (1953) were among the rst to study mixing times as a function
of bafe position and impeller rotational speed. Results of several experimental
studies have been combined into empirical correlations relevant to industrial
applications (Procha zka and Landau, 1961; Hoogendoorn and Den Hartog,
1967; Sano and Usui, 1985; Grenville, 1992; Ruszkowski, 1994; Nienow, 1997).
Grenville and Nienow (2004) present a concise review as to such correlations.
Bouwmans (1992; see also Bouwmans et al., 1997) used a particle tracking
technique in a RANS ow eld to estimate trajectories of neutral and buoyant
additions, to construct Poincare sections of additions crossing specic horizon-
tal cross-sectional planes, to predict probabilities of surfacing for buoyant
additions, and to mimic the temporal response of conductivity probes.
Exploiting the Eulerian point of view, Ranade et al. (1991) was one of the rst
to simulate mixing times for a vessel provided with a pitched blade turbine; in
HARRY E. A. VAN DEN AKKER 190
this early work, the impeller swept area was still modeled as a black box just
delivering inlet conditions for the turbulent ow in the remainder of the vessel.
Various authors (Osman and Varley, 1999; Jaworski et al., 2000; Bujalski et al.,
2002) used SM or MFR techniques in RANS-based simulations for Rushton
turbines driven ow combined with species transport equations to predict mix-
ing times, but arrived at values 23 times higher than the experimental values.
This is in line with the ndings for Rushton turbines that in such RANS-based
simulations turbulence levels are underpredicted and that mixing across the
central plane of the discharge plane is poorly reproduced. Montante and
Magelli (2004) studied a homogenization process in a bafed vessel stirred with
various sets of Rushton turbines; while effects of varying impeller number and
spacing were correctly forecasted, a very low turbulent Schmidt number had
to be adopted for obtaining good quantitative agreement in terms of tracer
response curves. RANS-based simulations of mixing in tanks provided with
pitched blade turbines prove to underpredict mixing times, probably owing to
mesoscale concentration uctuations not really reproduced by the simulation.
All these ndings of disappointing quantitative agreement with experimental
data stem from the inherent drawback of the RANS-approach that there is no
clear distinction between the turbulent uctuations modeled by the Reynolds
stresses and (mesoscale) uctuations. In LES, however, the distinction between
resolved and unresolved turbulence is clear and relates to the cell size of the
computational grid chosen.
The LES methodology has recently been applied in a number of studies
simulating in a stirred tank the mixing in response to the addition of a tracer.
Yeoh et al. (2004a) carried out an FV-simulation that matched the experimental
set-up of Lee (1995); the focus of their study is on mixing patterns, on traces of
concentrations at certain monitoring points, and on a comparison of predicted
mixing time with correlations from literature. Hartmann (2005) and Hartmann
et al. (2006) coupled an FV solver for the species transport to an LB ow solver.
In his study, different from Yeoh et al. (2005), the passive scalar is injected at
zero speed to avoid an effect of jet mixing on mixing times. Hartmann aimed at
reproducing the experiments of Distelhoff et al. (1997) and focused on the effect
of impeller size and injection position on mixing time. A sample of his results is
presented in Fig. 6. For the time being, the results due to Hartmann still suffer
from an unphysical mass increase caused by the current implementation of his
novel immersed boundary method using ghost cells.
Comparative studies on simulating mixing times by means of the traditional
RANS approach and the more sophisticated LES are due to Gao and Min
(2006), Gao et al. (2006), and Jahoda et al. (2006). They all show that
RANS-based simulations fail in reproducing the transient responses of probes
monitoring the local tracer concentrations, while LES is able to mimic
the experimental traces quite accurately (see Fig. 7, from Jahoda et al., 2006).
The latter traces strongly resemble those presented by Hartmann (2005) and
Hartmann et al. (2006).
THE DETAILS OF TURBULENT MIXING PROCESS 191
B. SUSPENDING SOLIDS
An important class of stirred tank applications involves suspending discrete
solid particles in a continuous liquid phase where the turbulent-liquid-ow
FIG. 6. A sample of Hartmanns results (2005). The tracer is injected (with zero speed) at the top
of the tank in a plane midway between two bafes (black dot in the plots). The concentration elds
shown are also in the midway bafe plane. T/D 3, c
N
denotes the nal uniform concentration.
FIG. 7. Time traces of normalized concentration as seen by a probe in the lower part of a vessel
in simulations of a mixing time experiment. The vessel is provided with two Pitched Blade Turbines.
Three different types of simulations are shown, where ske stands for a standard ke simulation and
sm, mrf, and les have the usual meaning. Reproduced with permission from Jahoda et al. (2006).
HARRY E. A. VAN DEN AKKER 192
generated by the impeller induces particle motion and should prevent sedimen-
tation. Such suspensions are for instance encountered in industrial crystalliza-
tion and in catalytic slurry reactors. The usual procedure in designing such
processes is making use of the Zwietering correlation (1956). Note that in the
NAMF Handbook of Mixing, Chapter 10 on Solid-Liquid Mixing does not
refer to any CFD results.
Although the Zwietering correlation provides valuable guidance to chemical
engineers in, e.g., solving practical engineering problems and dealing with scal-
ing-up issues, at a more fundamental level there are many unresolved issues
such as:

What is actually going on at the particle scale (in terms of, e.g., heat and mass
transfer, or mechanical load on particles as a result of particleparticle and
particleimpeller collisions) and how are these microscale events affected by
the larger-scale phenomena?

On the reverse, how does the presence of particles affect local and global ow
features in the vessel such as the vortex structure in the vicinity of the impeller,
power consumption, circulation and mixing times, and the spatial distribution
of turbulence quantities; more specically: colliding particles have an impact
on the liquids turbulence (Ten Cate et al., 2004) while local particle con-
centrations affect the effective (slurry) viscosity which may be useful in the
macroow simulations?
Most of these issues may best be studied by DNS, while other can better be
tackled by LES. Anyhow, RANS-based simulations are not very suited to this
purpose as the turbulence in the RANS-approach is not resolved at all but just
modeled. Below, typical DNS-, LES-, and RANS-based simulations of solids
suspensions will be reviewed in succession.
1. A Direct Numerical Simulations Approach
In view of secondary nucleation in crystallizers, Ten Cate et al. (2004) were
interested in nding out locally about the frequencies of particle collisions in a
suspension under the action of the turbulence of the liquid. To this end, they
performed a DNS of a particle suspension in a periodic box subject to forced
turbulent-ow conditions. In their DNS, the ow eld around and between
the interacting and colliding particles is fully resolved, while the particles are
allowed to rotate in response to the surrounding turbulent-ow eld.
Ten Cate et al. (2004) were able to learn from their DNS about the mutual
effect of microscale (particle scale) events and phenomena at the macroscale: the
particle collisions are brought about by the turbulence, and the particles affect
the turbulence. Energy spectra conrmed that the particles generate uid
motion at length scales of the order of the particle size. This results in a strong
increase in the rate of energy dissipation at these length scales and in a decrease
THE DETAILS OF TURBULENT MIXING PROCESS 193
of the turbulence at larger length scales. All these details were obtained just by
restricting the DNS to a small periodic box. Details of their approach will be
dealt with further on; a typical result is Fig. 12.
2. Results of Large Eddy Simulations
Derksen (2003), on the contrary, was interested in simulating the process of
solids suspension in a stirred tank; to this end, he tracked particles in the whole
tank in a Lagrangian sense, considering the particles as point particles and not
resolving the detailed ow eld between the particles. In other words, Derksen
applied a more supercial view on the particle suspension by dropping details,
and was rewarded with a picture of the full tank: see Fig. 8. Yet, Derksen was
able to track just over 6.7 million particles, to include effects such as particle
rotation, particleparticle collisions, particleimpeller collisions and even
two-way coupling, and to include uidparticle interaction forces such as the
Saffman force, the Magnus force, and forces due to stress gradients. Tracking all
these particles was done in a turbulent-ow eld obtained via an LB LES.
Derksen (2006a) continued along this line of approach andby means of a
clever strategymimicked the long-time behavior of solids suspension in an
unbafed tall stirred tank equipped with four hydrofoil impellers (Lightnin
s
A310). The time span covered by his LES amounted to some 20,000 impeller
revolutions (some 20 min). Running a LES for a Reynolds number of 1.6 10
5
over the entire time span is not an option, and for that reason a particular ow
FIG. 8. This is a snapshot of a spatial particle distribution. The plane shown is the horizontal
cross-section just below the disc of a Rushton turbine in a at-bottomed stirred tank. The impeller
revolves in the counter clockwise direction. Particle size is some 0.468 mm; Re 1.5? 10
5
;
volume fraction amounts to 3.6%; number of particles tracked in the simulation just over 6.7
million. Reproduced with permission from Derksen (2003).
HARRY E. A. VAN DEN AKKER 194
time series of sufcient length is stored and repeatedly played for tracking the
paths of 20,000 particles 0.33 mm in diameter and with a density of 2,450 kg/m
3
.
The advantage of this approach is that use is made of the highly resolved ow
information of the LES. The simulations of the particle response to the typical
ow eld created by the four impellers showed the counterintuitive behavior
observed by Pinelli et al. (2001): almost all solid particles rise to the top of
the tank.
3. Results of Reynolds Averaged Navier Stokes Simulations
A very different type of CFD-simulation results are those due to, e.g., Montante
and Magelli (2005) who studied solids suspension in stirred tanks by means of
two-uid CFD simulations, i.e., a RANS-type of EulerEuler simulations. These
authors tested two commercial CFD codes, viz. FLUENT 6.0 and CFX4, for
various formulations of the two-uid model, of the uidparticle interaction force,
and of the ke turbulence model for multiphase ow. Montante and Magelli just
focused on predicting axial proles of the solids concentration for three bafed
stirred tanks agitated with single and multiple impellers; they evaluated the effect
of the various model formulations on these proles and compared their various
predicted curves with experimental data. This type of simulation delivers data
relevant for engineering purposes of limited scope and depth only.
In this context, one component of the phase interaction force may need sepa-
rate discussion: the drag force. While most authors use the SchillerNauman
equation
C
D

24
Re
p
1:0 0:15Re
0:687
p

(16)
for the relation between particle drag coefcient and the particle Reynolds
number (see, e.g., Ranade, 2002, or Derksen, 2003). Brucato et al. (1998),
however, reported experimental data showing that free stream turbulence may
signicantly increase particle drag coefcients. They proposed a novel corre-
lation for predicting the effect of free stream turbulence on the particle drag
coefcient:
C
Dt
C
D
C
D
8:76 10
4
d
p
Z
K

3
(17)
in which C
D t
denotes the particle drag coefcient in a turbulent ow and C
D
stands for the usual particle drag coefcient as given by Eq. (16). Equation (17)
implies that the drag coefcient for particles the size of which is smaller than
or comparable with that of the smallest turbulent eddies (represented by the
local Kolmogorov length scale Z
K
) is hardly affected by free-stream turbulence,
while the uid-mechanical interaction of larger particles with turbulent eddies
becomes signicant and leads to an increase of particle drag.
THE DETAILS OF TURBULENT MIXING PROCESS 195
Montante et al. (2000), Micale et al. (2000), and Montante and Magelli (2005)
incorporated this novel correlation into their two-uid simulations of solids
suspension in stirred vessels and arrived at pretty good agreement with experi-
mental data. These authors claim that this correlation due to Brucato et al.
(1998) plays an essential role in arriving at this good agreement: introducing
the turbulence effect reduces the tendency of the larger particles to collect in the
bottom part of the vessel. One may argue, however, that the introduction of
a higher particle drag coefcient is rather articial and just compensates for
shortcomings in the two-uid formulation used, such as ignoring all other
components of the phase interaction force. By the way, Micale et al. (2004) in
simulating particle suspension height ignored this effect of free stream turbu-
lence and identied it as a second-order effect only.
In some way, introducing an increased particle drag by means of Eq. (17)
resembles the earlier proposal raised by Bakker and Van den Akker (1994b) to
increase viscosity in the particle Reynolds number due to turbulence (in agree-
ment with the very old conclusion due to Boussinesq, see Frisch, 1995) with the
view of increasing the particle drag coefcient and eventually the bubble hold-
up in the vessel. Lane et al. (2000) compared the two approaches for an aerated
stirred vessel and found neither proposal to yield a correct spatial gas distri-
bution.
In addition, it is dubious whether this new correlation due to Brucato et al.
(1998) should be used in any EulerLagrangian approach and in LES which
take at least part of the effect of the turbulence on the particle motion into
account in a different way. So far, the LES due to Derksen (2003, 2006a) did not
need a modied particle drag coefcient to attain agreement with experimental
data. Anyhow, the need of modifying particle drag coefcient in some way
illustrates the shortcomings of the current RANS-based two-uid approach of
two-phase ow in stirred vessels.
The present author wonders whether, due to its very nature (particularly, the
various assumptions as to averaging and the various modeling uncertainties), a
RANS-based two-uid approach is suited for reproducing the details of solids
suspension in a stirred vessel. May be we should be satised with the gross
predictions of the current RANS methods and turn to LES for the details
of those processes which are dominated by the spatially distributed turbulence.
It is really a valid question how long we should keep trying and improving the
RANS two-uid approach as now the increased computer power brings the
much more sophisticated LES within reach.
C. DISSOLVING SOLIDS
Properly simulating a dissolution process of solid particles in a stirred vessel
operated in the turbulent-ow regime urges for a very detailed simulation of the
turbulent-ow eld itself. Just reproducing the overall ow pattern by means of
HARRY E. A. VAN DEN AKKER 196
a RANS-type of simulation is not sufcient! Our estimate is that a two-uid
simulation (on the basis of RANS) yields too rough an approximation of the
process. The dissolution process may too strongly depend on the heavily uc-
tuating ow and concentration elds around particles whichas the result of
the presence and action of eddies of various scalesmay move chaotically with
respect to one another while being advected through the vessel.
One really may need an inherently transient LES to capture all these details.
The ner the grid for such a LES, the more reliably the local transient con-
ditions may be taken into account in reproducing this turbulent mass transfer
process (while ignoring the issue of supplying the heat for the dissolution which
may also depend on a proper representation of the turbulent-ow eld).
An additional important issue is how many particles have to be tracked for a
proper representation of the transient spatial distribution of the particles over
the vessel.
Hartmann et al. (2006) reported very detailed simulation results (see also
Hartmann, 2005) (Fig. 9). Their LB simulation was restricted to a lab-scale
vessel 10 L in size only for which 240
3
lattice cells a bit smaller than 1 mm
2
were
used; the temporal resolution was 25 ms only. A set of 7 million mono-disperse
spherical particles 0.3 mm in size was released in the upper 10% part of the
vessel. At the moment of release, the local volume fraction amounted to 10%.
The particle properties were those of calcium chloride. The simulation was
carried out on 30 parallel processors of an SGI Altrix 3700 system and required
for 6 weeks for 100 impeller revolutions.
Figure 9 presents some typical results for the spatial distribution of the dis-
solving particles, their sizes, and their size distribution. Initially, the particles
respond to the centrifugal forces in the vortices and collect at the outer regions
of the vortices, giving rise to streaky patterns. The dissolution process results in
decreasing particle sizes and, hence, in decreasing inertia to the effect that
gradually they start behaving like uid tracers. Overall, the solids and scalar
concentrations become more homogeneous in the course of the dissolution
process. The simulation even predicts that a plot of the Sauter mean diameter
d
32
vs. time may exhibit a minimum (not shown here).
D. PRECIPITATION AND CRYSTALLIZATION
The counterparts of dissolving particles are the processes of precipitation and
crystallization the description and simulation of which involve several addi-
tional aspects however. First of all, the interest in commercial operations often
relates to the average particle size and the particle size distribution at the com-
pletion of the (batch) operation. In precipitation reactors, particle sizes strongly
depend on the (variations in the) local concentrations of the reactants, this
dependence being quite complicated because of the nonlinear interactions of
uctuations in velocities, reactant concentrations, and temperature.
THE DETAILS OF TURBULENT MIXING PROCESS 197
The same applies to crystallizers, in which particle sizes and particle number
concentrations not only depend on nucleation and growth from supersaturated
mother liquid, but are also affected by shear-dominated agglomeration and
by secondary nucleation as a result of particleparticle and particleimpeller
collisions. Some of the subprocesses involved may be limited to specic and
different parts of the vessel: e.g., nucleation may be restricted to a ame-like
region around the outlet of a feed pipe (Van Leeuwen, 1998). In addition, in
FIG. 9. Snapshots of particle sizes and their spatial distribution in a vertical midway bafe plane
at two moments in time, along with the pertinent respective overall particle size distributions. The
diameter of the particles is enlarged by a factor of 10 for reasons of clarity. Grey colors represent
particle size with respect to the original particle size. Reproduced with permission from Hartmann
(2005).
HARRY E. A. VAN DEN AKKER 198
many cases, many more product properties are relevant, such as color, texture,
and particle morphology.
Gradually, the insight is growing that all these properties are not only
affected by the averaged process conditions in reactor or crystallizer, but also
by their spatial and temporal variations felt by the particles during their stay in
and on their pathways through the vessel (see, e.g., Rielly and Marquis, 2001).
The rst response of modelers was to introduce compartmental modeling
(Van Leeuwen, 1998; Bermingham et al., 1998; Ten Cate et al., 2000): the vessel
is divided into several compartments, each compartment being considered as
(more or less) ideally mixed and described in terms of averaged values of the
process parameters such as temperature, concentrations, and rate of turbulent
kinetic energy dissipation; these averaged values may vary strongly from com-
partment to compartment. Although this approach is an improvement over
the traditional method of lumping all variations into a single average value, an
important problem still is how to decide on the number and size of the com-
partments. Carrying out RANS-type simulations would be an option for
selecting proper compartmentalization on a case-to-case basis.
Seckler et al. (1995), Van Leeuwen et al. (1996), and Wei and Garside (1997)
were among the rst to exploit commercial CFD codes (FLUENT and
PHOENICS) for simulating precipitation reactors (of a particular simplied
design) by adding to their codes some simple precipitation kinetics, i.e., relations
for nucleation and particle growth. Van Leeuwen et al. (1996) included the rst
four moments of the crystal size distribution in their simulations. Van Leeuwen
(1998) was the rst to study precipitation in a stirred vessel; among other things,
he explored the option for extending his RANS-type simulations with a routine
involving the use of PDFs to account for variations in the species concentrations
(see one of the next sections on chemical reactors). He did not arrive at a
satisfactory agreement with experimental data.
As discussed in several of the above sections, LES is much better suited to
represent the spatial and temporal variations in a turbulent-ow eld. Of
course, this is very relevant in precipitation reactors and crystallizers where the
particle formation is a highly nonlinear process. Progress is being made in
developing LES including detailed models (e.g., PDFs, or methods of moments)
for specic reactive systems (e.g., combustion, polymerization) under specic
conditions in simple geometries (e.g., tubular reactors): see also one of the next
sections on chemical reactors. In turbulent agitated precipitation reactors and
crystallizers, however, mixing is so intense and complex and so heavily dom-
inated by the revolving impeller that so far no one has succeeded in simulating
the full process of nucleation, particle growth, agglomeration, and particle
break-up and in arriving at a reasonable prediction of the eventual particle size
distribution.
Certain isolated aspects of precipitation and crystallization which have suc-
cessfully been studied by means of LES, are discussed below to illustrate the
progress being made in the eld of precipitation and crystallization. The LB
THE DETAILS OF TURBULENT MIXING PROCESS 199
technique exploited in these studies is very helpful, since LB easily allows par-
allel computing and the related computational acceleration. These detailed
studies may make a similar simulation possible as presented above for a dis-
solution process.
1. Agglomeration
Hollander (2002) and Hollander et al. (2001a,b, 2003) studied agglomeration
in a stirred vessel by adding a single transport equation for the particle number
concentration m
0
(actually, the rst moment of the particle size distribution)
@m
0
@t
r ~ vm
0

1
2
b
0
m
2
0
(18)
Note that the particle diffusion term is ignored, just like particle dispersion
due to SGS motions (this was found justied in a separate simulation). The
shape of the sink term in the right-hand term of this equation is due to Von
Smoluchowski (1917) while the local value of the agglomeration kernel b
0
is
assumed to depend on the local 3-D shear rate according to a proposition due to
Mumtaz et al. (1997).
This additional Eq. (18) was discretized at the same resolution as the ow
equations, typical grids comprising 120
3
and 180
3
nodes. At every time step,
the local particle concentration is transported within the resolved ow eld.
Furthermore, the local ow conditions yield an effective 3-D shear rate that can
be used for estimating the local agglomeration rate constant b
0
. Fig. 10 (from
Hollander et al., 2003) presents both instantaneous and time-averaged spatial
distributions of b
0
in vessels agitated by two different impellers; color versions
of these plots can be found in Hollander (2002) and in Hollander et al. (2003).
The results of these simulations conrm the suspicion of a large spread in
b
0
-values across the vessel. Furthermore, the simulations show that agglomer-
ation does not occur in the impeller region, as the hydrodynamic conditions
(shear!) are too severe for agglomerates to survive. A bulk region can be dened
in which agglomeration conditions are benecial. Due to the typical structure of
the ow (i.e., trailing vortices, large-scale turbulent motion), the vessel contains
large gradients in agglomeration rates. It is this effect that causes large differ-
ences in the local particle consumption rate and leaves the vessel with a very
uneven distribution in particle concentration.
Fig. 11 presents the results of some 30 simulations for various conditions and
two impeller types in terms of the mean agglomeration rate constant observed in
the various simulations vs. the vessel-averaged shear rate found in the simu-
lations. The simulations all started from the dotted curve for relating local
agglomeration rate constant to local shear rate. A clear decrease in the maxi-
mum of b
0
as well as a shift toward higher average shear rates was found which
are caused by the local nature of the nonlinear ow interactions only. These
HARRY E. A. VAN DEN AKKER 200
FIG. 11. The discrepancy between the original kinetic relation due to Mumtaz et al. (1997) and the
observed relation between mean agglomeration rate constant b
0
and volume-averaged shear rate.
Symbols refer to individual numerical simulations (LES). RT stands for Rushton Turbine, PBT for
Pitched Blade Turbine. Reproduced with permission from Hollander et al. (2001b).
FIG. 10. Results of LES-based simulations of an agglomeration process in two vessels: one ag-
itated by a Rushton turbine (left) and one agitated by a Pitched Blade Turbine (right). The two plots
show the agglomeration rate constant b
0
normalized by the maximum value, in a vertical cross-
sectional plane midway between two bafes and through the center of the vessel. Each of the two
plots consists of two parts: the right-hand parts present instantaneous snapshots; the left-hand parts
present spatial distributions of time-averaged values after 50 impeller revolutions. Reproduced with
permission from Hollander et al. (2003).
THE DETAILS OF TURBULENT MIXING PROCESS 201
nonlinear effects may play an important role in our mediocre understanding of
the scale-up behavior of agglomeration processes in stirred vessels.
The computational demands for Hollanders simulations (in 2000) were typ-
ically of the order of 1 CPU week on a PentiumIII/500 MHz at 400 MB of
memory per run. The computer code was fully parallelized and was run on a 12
CPU Beowulf cluster.
2. Colliding Particles
Finally, Ten Cate et al. (2001) studied secondary nucleation as a result
of crystalcrystal collisions by means of a two-step approach. The rst step
involved a LES of the complete crystallizer in which the liquid phase (typically
containing 1020%v of solids) is treated as a single phase with a homogeneous
density and viscosity. (In principle, density and viscosity may be allowed to vary
locally in the simulation, by invoking the help of a particle dispersion routine
(e.g., Liu, 1999) and coupling the resulting particle concentrations back to a
SGS model to modify the local values of density and viscosity.) The spatial grid
resolution (some 5.0 mm) used by Ten Cate typically was at least 10 times the
crystal size; the grid comprised some 36 million cells.
The second step in Ten Cates two-step approach was to focus on crystal
crystal interaction by means of an explicit two-phase DNS of the turbulent
suspension that completely resolves the translational and rotational motions
and collisions of the spherical particles plus the turbulence of the liquid between
the particles. The particle motions are driven by the turbulent ow and the
particles affect the turbulent ow of the liquid in between. When particles
approach one another down to a distance smaller than the grid spacing, lubri-
cation theory is exploited to bridge the gap between them.
At the start of a simulation, the particles (up to 3,900) are placed randomly,
without mutual contact and with zero velocity, into a fully developed turbulent
single-phase ow eld in a periodic box. This starting situation is subjected to a
precalculated force eld; this forcing involves a divergence-free white-noise sig-
nal, distributed over the wave number domain as a Gaussian distribution about
a desired wave number, with a characteristic root mean square velocity and a
characteristic length scale. In this way, turbulent conditions are generated which
accurately recover a priori set values such as the Kolmogorov length scale, the
integral length scale, and the integral time scale, derived from the LES at some
position somewhere in the crystallizer. For details, the reader is referred to Ten
Cate (2002) and Ten Cate et al. (2004).
The evolution of the two-phase turbulence depends on the initial random
position of the particles, the motion of which modies the turbulent-ow eld
directly. These DNS are therefore a nice example of two-way coupling between
the two phases: see Fig. 12. From these DNS, detailed knowledge can be derived
as to the frequency of the particleparticle collisions and the forces involved
HARRY E. A. VAN DEN AKKER 202
(as a function of local conditions in a crystallizer) which may yield quantitative
data on secondary nucleation (fragmentation of crystals due to collisions).
VII. Stirred GasLiquid and LiquidLiquid Dispersions
Similar problems as encountered in simulating agitated suspensions also
play a role in agitated gasliquid and liquidliquid systems. Again, there is
the dilemma whether to use a Lagrangian approach or to apply a two-uid,
EulerEuler, model. Then, there is again the question of one-way coupling vs.
two-way coupling. Thirdly, the issue of taking the particle size distribution into
account (yes or no) should be addressed. Further, particleparticle interaction
may not be ignored in many cases, and then the additional problem is coa-
lescence of bubbles or droplets, and their break-up. All these issues have not
been tackled simultaneously in the past, not in the least because of computa-
tional limitations.
Furthermore, the physics of the interaction between turbulence and bubbles
in the complex ow of a stirred vessel, with its implications for coalescence and
break-up of bubbles and drops, is still far from being understood. Up to now,
simple correlations are available for scale-up of industrial processes; generally,
these correlations have been derived in experimental investigations focusing on
the eventual mean drop diameter and the drop size distributions as brought
1.210
1
1.210
-2
0.4
FIG. 12. Snapshot from a two-phase DNS of colliding particles in an originally fully developed
turbulent ow of liquid in a periodic 3-D box with spectral forcing of the turbulence. The particles
(in blue) have been plotted at their position and are intersected by the plane of view. The arrows
denote the instantaneous ow eld, the colors relate to the logarithmic value of the nondimensional
rate of energy dissipation.
THE DETAILS OF TURBULENT MIXING PROCESS 203
about by the power input (mainly) via the impeller given the uid properties
(e.g., Colenbrander, 2000).
In the 1980s, Issa and Gosman (1981), Pericleous and Patel (1987), and
Tra ga rdh (1988) made the rst attempts to simulate aerated stirred vessels
computationally. Their results were rather approximate indeed, and were not
validated by means of experimental data.
A. BAKKERS GHOST! CODE
In the early 1990s, Bakker and Van den Akker (1991, 1994) introduced an
approximate but effective EulerEuler approach (see also A. Bakkers PhD
Thesis, 1992): on the basis of a single-phase RANS ow eld calculated by
FLUENT, a code named GHOST! calculated local and averaged values of
bubble size d
b
, gas hold-up a, and specic mass transfer rate k
l
a.
Their GHOST!-code essentially consisted of a mass balance for the gas, a
transport equation for the bubble number density n
b
, and a force balance for a
single bubble, respectively, which run as
r a u r r D a

S
g
(19)
@n
b
@t
r n
b
u o n
b1
n
b

S
g
V
b;in
(20)
r
l
V
b
w
2
r
^ r r
l
gV
b
^ z C
D
1
2
r
l
u
s
j ju
s
p
4
d
2
b
(21)
respectively. Of course, the variables n
b
, d
b,
and a are interrelated. The second
term in the mass balance, Eq. (19) stands for the extra transport of the bubbles
due to the larger turbulent eddies; usually, this turbulent transport is taken into
account by a separate term in the momentum balance for the dispersed phase
which, however, in Bakkers approach is replaced by the simple force balance,
Eq. (21), for a single bubble. The coefcient o in the transport equation for n
b
is
just a relaxation parameter that stands for the rate at which n
b
responds to the
local turbulence, i.e., adaptseither by coalescence (in the bulk of the stirred
vessel) or by bubble break-up (in the impeller swept domain of the vessel)to
the local equilibrium number density n
bN
The latter variable corresponds to the
local maximum stable bubble size d
bN
whichaccording to Hinze (1955)
depends on the local value of the specic rate of energy dissipation e:
d
b1
C
b1
12
s
r
l

3=5

2=5
(22)
Bakker increased the local values of e as obtained with FLUENT by a con-
tribution related to the slip velocity of the bubbles. Working with a single
HARRY E. A. VAN DEN AKKER 204
relaxation parameter o is a phenomenological description avoiding detailed
relations for coalescence and bubble break-up.
The Eulerian gas velocity eld required in both the mass balance and the
above transport equation for n
b
is found by an approximate method: rst, the
complete eld of liquid velocities obtained with FLUENT is adapted downward
because the power draw is smaller under gassed conditions; next, in a very
simple way of one-way coupling, the bubble velocity calculated from the above
force balance is just added to this adapted liquid velocity eld. This procedure
makes a momentum balance for the bubble phase redundant; this saves a lot of
computational effort.
Finally, Bakker and Van den Akker calculated local values for the specic
mass transfer rate k
l
a, by estimating local k
l
-values from local values of the
Kolmogorov time scale O(n/e) and by deriving local values of the specic in-
terfacial area a from local values for bubble size and bubble hold-up.
In spite of all the simplications Bakker and Van den Akker applied and
given the black box approach for the impeller swept domain, their simulations
resulted in values for the bubble size just below the liquid surface, overall hold-
up, and average k
l
a values which are in good agreement with their experimental
data (see Table II). The major step forward they made was the acquisition of the
different spatial distributions of average bubble size (see Fig. 13), bubble hold-
up and k
l
a as effected by three common impeller types. As a matter of fact, their
approach may be restricted to low values of the gas hold-up.
B. VENNEKERS DAWN CODE
Venneker et al. (2002) extended the GHOST! approach due to Bakker and
Van den Akker (1994b) by replacing the single transport equation for the
TABLE II
COMPUTATIONAL AND EXPERIMENTAL RESULTS FOR OVERALL HOLD-UP A, SIZE /D
B,OUT
S OF THE BUBBLES
LEAVING THE LIQUID HEEL, AND OVERALL SPECIFIC MASS TRANSFER RATE K
L
A (FROM: BAKKER, 1992)
S. no. a (%) /d
b,out
S (mm) k
l
a (l/s)
Exp. Sim Exp Sim Exp Sim
1 DT 4.770.2 4.9 3.25 2.91 0.038 0.038
2 A315 4.670.2 4.2 3.76 3.59 0.035 0.036
3 A315 4.870.2 4.3 3.82 0.038 0.036
4 PBT 4.170.3 4.1 3.44 3.39 0.036 0.037
5 PBT 1.170.3 1.0 2.00 0.011 0.013
Note: The respective impellers used are a classical Rushton turbine (DT), a hydrofoil impeller
(A315) manufactured by Lightnin, and a Pitched Blade impeller (PBT). The cases 1 through 4 all
relate to a supercial gas rate of 3.6 mm/s only, with impeller speeds varying between 5 and 10/s (gas
ow numbers between 0.01 and 0.02); cases 2 and 3 differ in sparger size and position.
THE DETAILS OF TURBULENT MIXING PROCESS 205
bubble number density including the effective relaxation parameter o by a
number of population balance equations. Actually, these population balance
equations are convective-diffusive transport equations for the bubble number
density in a specic bubble size class and include separate birth and death terms
which take coalescence and break-up into account. In addition, the procedure
for adapting the liquid ow eld to the lower power draw under gassed con-
ditions has been improved on the basis of experimental ndings due to Rous ar
and Van den Akker (1994). For the rest, the same approximations are made as
in GHOST! with respect to gas velocities, rate of energy dissipation, and specic
mass transfer rate.
Venneker et al. (2002) used as many as 20 bubble size classes in the bubble
size range from 0.25 to some 20 mm. Just like GHOST!, their in-house code
named DAWN builds upon a liquid-only velocity eld obtained with FLUENT,
now with an anisotropic Reynolds Stress Model (RSM) for the turbulent
momentum transport. To allow for the drastic increase in computational bur-
den associated with using 20 population balance equations, the 3-D FLUENT
ow eld is averaged azimuthally into a 2-D ow eld (Venneker, 1999, used a
less elegant simplication!)
The agreement between simulation results and experimental data is encour-
aging (see Fig. 14), although the simulation gives higher hold-up values in the
upper part of the vessel while the overall hold-up is lower in the simulation than
FIG. 13. Spatial distributions of bubble size in three vessels agitated by different impellers: a
classical Rushton turbine (DT), a hydrofoil impeller (A315) manufactured by Lightnin, and a
Pitched Blade Impeller (PBT). The gas ow numbers in these simulations are in the range 0.010.02.
These simulation results have been obtained by using GHOST! Reproduced with permission from
Bakker (1992).
HARRY E. A. VAN DEN AKKER 206
in the experiment. The latter discrepancy may be due to the use of optical probes
overlooking bubbles smaller than, say, 1 mm.Venneker et al. (2002) present
some more comparisons between computational and experimental results. One
should realize, however, that for validation purposes hardly any detailed ex-
perimental data as to bubble size distributions in stirred vessels are available.
This same shortage of experimental data hampers the assessment of the so-called
MUltiple-SIze Group approach MUSIG due to Lo (2000) as incorporated in the
commercial CFD code CFX.
C. FURTHER SIMULATIONS
In comparison with Bakker and Van den Akker (1994b) and Venneker et al.
(2002), Khopkar et al. (2005) applied a more sophisticated two-uid approach
including a standard ke turbulence model. Using the incorrect snapshot ap-
proach due to Ranade (2002), their simulation results (for gas ow numbers
being 4 times higher than those of Bakker and Van den Akker, 1994b) still
exhibit major discrepancies with respect to experimental data. One of the
alpha (%)
10.0
9.0
8.0
7.0
6.0
5.0
4.0
3.0
2.0
1.0
0.0
FIG. 14. Spatial distribution of the gas hold-up in a turbulent stirred vessel lled with a 0.075%
Keltrol solution in water and agitated by a Rushton turbine: (a) experimental data obtained by
means of an optical probe; (b) computational result from DAWN. Overall hold-up amounts to some
3.1% in the simulation and 3.7% in the experiment. Reproduced with permission from Venneker
(1999), improved.
THE DETAILS OF TURBULENT MIXING PROCESS 207
striking features is that their liquid velocities even in the outow of the Rushton
impeller are pretty much overestimated in the simulation, may be due to the use
of the snapshot approach (cf. the dicussion on page 180). The spatial distri-
butions of gas hold-up found in the simulations are compared with experimental
data obtained by means of computed tomography (CT); there is substantial
room for improvement (see Fig. 15) in spite of the much more sophisticated
type of simulation. The paper due to Gentric et al. (2005) exploits several fea-
tures and options of the two-uid mode of the commercial code STAR-CD and
illustrates its capabilities in comparing the mixing performance of two industrial
mixing vessels, but does not present validation by means of experimental data.
The combination of two-phase ow, turbulence, and a revolving impeller
poses tremendous simulation problems and still requires excessive computer
time or power. While in bubble columns and gas lift loops population balances
are used with some success (see, e.g., Wang et al., 2006), the ow eld in a
stirred tank is so much more complicated and turbulent and dominated by the
revolving impeller that implementing them in stirred vessel simulations still
causes serious convergence problems. Laakkonen et al. (2006) reduced such
problems by restricting his simulation to a multiblock approach subdividing the
stirred vessel into just 23 ideally mixed subregions: this approach actually is a
kind of network of zone approach (see also, e.g., Hristov et al., 2004, who even
used 36,000 zones) extended with population balances and may not be named
CFD indeed. The use of the so-called Quadrature Method of Moments
(QMOM) has been suggested as a proper tool for combining population bal-
ances with CFD (Marchisio et al., 2003), but so far has not been used for
simulating gassed stirred tanks.
FIG. 15. Comparison of simulated and experimental gas hold-up distribution in a horizontal plane
10 cm above the bottom of the vessel. The gas ow number amounts to 0.084. Reproduced with
permission from Khopkar et al. (2005).
HARRY E. A. VAN DEN AKKER 208
D. A PROMISING PROSPECT
Just like in the context of simulating solids suspension, one may wonder
whether much may be expected from just sticking to the two-uid approach
combined with population balances. A better way ahead might rather be to
combine population balances with LES, while proper relations for the various
kernels used for describing coalescence and break-up processes could be deter-
mined from DNS of periodic boxes comprising a certain number of bubbles (or
drops). The latter simulations would serve to study the detailed response of
bubbles or drops to the ambient turbulent ow.
An attractive framework for investigating these phenomena is provided by
Derksen (2006b) who carried out DNS of liquidliquid dispersions in a 3-D
periodic box. Derksen investigated the response of a turbulent dispersion of
droplets to a history comprising rst a rapidly increasing turbulent activity, then
a quasi-steady situation of high turbulence intensity and nally a rapid decay in
turbulence intensity; this history may be equivalent to what a uid package
experiences during its passage through the impeller stream. Again, Derksen
applied a particular LB method for mimicking the two phases. The drop
size distribution and the Sauter mean diameter were tracked in time. Further-
more, the presence of the droplets affected the turbulence spectrum because of
the small-scale uid motions induced by the droplets, on the analogy of the
interaction between solid particles and turbulence in the work of Portela and
Oliemans (2003) and Ten Cate et al. (2004).
VIII. Chemical Reactors
So far, most (stirred) chemical reactors have been designed and scaled up by
traditional methods exploiting simple conceptssuch as continuous stirred
tank reactor and residence time distributionand scale-up rules involving
usually a single dimensionless number such as the (mixing) Damko hler number
being the ratio of the turbulent macrotime scale to the characteristic reaction
time scale. In addition, various types of local mixing times and their ratios
are used to characterize or categorize the interaction of mixing and chemical
reactions at scale-up (Patterson et al., 2004).
These methods hardly take spatial distributions of velocity eld and chemical
species or transient phenomena into account, although most chemical reactors
are operated in the turbulent regime and/or a multiphase ow mode. As a result,
yield and selectivity of commercial chemical reactors often deviate from the
values at their laboratory or pilot-scale prototypes. Scale-up of many chemical
reactors, in particular the multiphase types, is still surrounded by a fame of
mystery indeed. Another problem relates to the occurrence of thermal runaways
due to hot spots as a result of poor local mixing effects.
THE DETAILS OF TURBULENT MIXING PROCESS 209
Patterson (1985) presented a concise review of the early developments in
computational modeling of second-order chemical reactions and of more com-
plicated and multiple reaction sets which are affected by an intermediate rate of
local turbulent mixing. At that moment in time, closing the cross-correlation
terms stemming from the turbulent uctuations by means of micromixing
models was still in its infancy. He also just hinted on the use of PDFs. Fur-
thermore, the limited computer power of those days kept detailed simulations
and their assessment impossible. Stirred vessels in particular were too difcult a
type of ow devices to allow for application of rigorous CFD techniques, al-
though some attempts were made with a very small number of zones or mixing
segments only.
A. MECHANISTIC MICROMIXING MODELS
In the 1980s, Bourne along with a long series of co-workers at ETH Zurich
developed a mechanistic micromixing approach in which lamellar structures
were central. His lamellar structures represent the small ow structures of the
size of the Kolmogorov length scale within which molecular diffusion is the
mechanism bringing the chemical species into the intimate contact required for a
chemical reaction. The best reference might be two papers due to Baldyga and
Bourne (1984a, b).
Such lamellar structures have also been described and modeled by Ranz
(1979) and Ottino (1980) in the context of chemical reactions in laminar ows.
In Bournes micromixing models for chemical reactors operated in the turbulent-
ow regime, various assumptions are raised as to the engulfment, the defor-
mation, and the lifetime of these lamellar structures which, along with the
diffusion of the reacting species, all affect the yield of the chemical reactions
taking place within these structures. Actually, a CDR equationsee Eq. (13)
is solved explicitly for a chemical species within a single Kolmogorov eddy. The
most appealing models Baldyga and Bourne proposed for the evolution of
such eddies are the Strain (St) Model, the Shear (Sh) Model, the Engulfment-
Deformation-Diffusion (EDD) model, and then the simpler Engulfment (E)
model. Bourne applied his technique to various sets of competing parallel or
consecutive model reactions each carried out in a fed batch reactor.
In the 1990s, Bakker and Van den Akker (1994, 1996)see also R.A.
Bakkers PhD thesis (1996)continued this mechanistic modeling approach by
attempting a completely deterministic description of the 3-D small-scale ow
eld in which the chemical reactions take place at the pace the various species
meet. Starting point is a lamellar structure of layers intermittently containing
the species involved in the reaction. These authors conceived such small-scale
structures as Cylindrical Stretched Vortex (CSV) tubes being strained in the
direction of their axis andas a resultshrinking in size in a plane normal to
HARRY E. A. VAN DEN AKKER 210
their axis. Such CSV tubes showed up around 1990 in several studies exploiting
DNS of turbulent ows in a periodic box. The evolution of a CSV tube during
its lifetime can be found by means of an analytical solution of the vorticity
equation. For the parameters typical of the turbulent-ow eld in a stirred
reactor, however, the vorticity distribution does not result in substantial wind-
ing of material lines during the pertinent short time scales. As a result, the main
effect of the stretching of the CSV is just the exponential shrinking rate of more
or less at, only slightly curling material layers.
Consequently, Bakker (1996) described the concentration evolution of a
single layer subjected to the vorticity eld of a single CSV by means of a one-
dimensional differential equation where both the nondimensional time and the
nondimensional spatial coordinate contain the exponential shrinking rate. In
this respect, the CSV approach differs from the various Bourne models in which
the successive generation of several multiple-layer stacks is required and vortex
age is a crucial element.
B. A LAGRANGIAN APPROACH
In addition, Bakker and Van den Akker (1994, 1996) were the rst to track
the path such structures follow in the turbulent-ow eld of a fed batch reactor
computationally. This is extremely relevant as both vortex age (in Bournes
multiple-layer models) and Kolmogorov length scale strongly depend on the
spatially strongly varying e. Precisely this latter variable exhibits a very inhomo-
geneous spatial distribution that only can be estimated by means of CFD. The
idea is that during the microscale process of mixing and reaction the macroow
eld advects the reaction zone throughout the reactor, thereby exposing the
zone to regions of varying e. The ow eld and the spatial e-distribution were
obtained via a RANS-type of simulation (FLUENT), while the tracking was
done by means of a Discrete Random Walk approach. (It should be kept in
mind that at the time of their simulations LES was not really an option yet!) In
addition to their own CSV model, Bakker and Van den Akker also validated
some of Bournes micromixing models.
Some typical results from their simulations are presented in Fig. 16 in which
the yield X
Q
of the product Q from the slow reaction of a set of two competitive
reactions in a fed batch reactor has been plotted vs. impeller speed for two
micromixing models, viz. their own CSV model and Bournes EDD model; their
simulation results are compared with experimental data from Bourne and Yu
(1991). For the cases shown, the CSV model may perform better than Bournes
EDD model, in particular when A is fed near to the impeller where mixing is
most intense.
An alternative but similar approach (Akiti and Armenante, 2004) is to dene
the reaction zones (or blobs) as a separate phase distinct for the ambient uid
THE DETAILS OF TURBULENT MIXING PROCESS 211
and to track these reaction zones by means of a Volume-of-Fluid (VOF) tech-
nique, which may be conceived being a pseudo-multiphase model, originally
designed by Hirt and Nichols (1981). Rather than a ke model, Akiti and
Armenante used a RSM model to reproduce the turbulence characteristics of
the ow eld needed for the tracking procedure.
The most important drawback of using a Lagrangian approach for simu-
lating (micro) mixing in chemical reactors is that some model is required
for describing the formation of the Kolmogorov-scale ow structures at
some (which?) distance from the mouth of the feed tube. This so-called feed
discretization, aimed at dening the starting conditions (size, number, concen-
trations) for the lamellar structures, may have an unknown impact on the
eventual yield.
FIG. 16. The yield X
Q
of the product Q of the slower reaction of a set of two competitive parallel
reactions in a fed batch reactor plotted vs. impeller speed (in /s). The experimental data are due to
Bourne and Yu (1991); the crosses refer to feeding reactant A at the top of the vessel, while the
diamonds refer to feeding more closely to the impeller. The various types of lines refer to simulations
as specied in the legend. Reproduced with permission from R. A. Bakker (1996).
HARRY E. A. VAN DEN AKKER 212
C. A EULERIAN PROBABILISTIC APPROACH
An alternative approach (e.g., Patterson, 1985; Ranade, 2002) is the Eulerian
type of simulation that makes use of a CDR equationsee Eq. (13)for each
of the chemical species involved. While resolution of the turbulent ow down to
the Kolmogorov length scale already is far beyond computational capabilities,
one certainly has to revert to modeling the species transport in liquid systems in
which the Batchelor length scale is smaller than the Kolmogorov length scale by
at least one order of magnitude: see Eq. (14). Hence, both in RANS simulations
and in LES, species concentrations and temperature still uctuate within a
computational cell. Consequently, the description of chemical reactions and the
transport of heat and species in a chemical reactor ask for subtle approaches as
to the SGS uctuations.
In order to obtain a realistic estimate for the reaction rate, the joint distri-
bution of the reactants at the smallest turbulent scales is required. Any model
disregarding this joint distribution may lead to an erroneous estimate of the
reaction rate. For instance, ltering the reaction term, on the analogy of the
LES lter for the uid ow, would result in an over-prediction of the reaction
rate due to the segregation at the subgrid scales. It may be just due to peculiar
operating or mixing conditions when, such as in the FLUENT simulations re-
ported on page 845 of Patterson et al. (2004), CFD simulations ignoring SGS
uctuations result in yield predictions close to experimental data. The value of
the Damkohler number, denoting the ratio of the turbulent macrotime scale to
the characteristic reaction time scale, plays an important role as well.
During the years, quite some proposals have been raised as to closure equa-
tions in the CDR equations for the spatial species distributions. These closure
equations relate to the correlation terms in general and to the SGS uctuations
in species concentrations in particular. These proposals substantially differ in
degree of sophistication. Patterson et al. (2004) present several examples of
closure models which were reasonably successful in reproducing a particular set
of experimental data. This, however, does not necessarily say something about
their universal applicability. Bakker and Fasano (1993) applied the so-called
Magnussen model and arrived at reasonable yield predictions for a competitive-
consecutive reaction system in a stirred reactor (see also Marshall and Bakker,
2004). This Magnussen model, originally derived for combustion, locally cal-
culates several reaction rates as a function of both mean concentrations and
turbulence levels and then selects the lower rate for the source term in the CDR
equation.
Particular attention is to be paid to closure models exploiting various types
of PDFs such as beta, presumed, or full PDFs (e.g., Baldyga, 1994; Fox, 1996,
2003; Ranade, 2002). While PDFs have successfully been exploited for descri-
bing chemical reactions in turbulent ames, tubular reactors (Baldyga and
Henczka, 1997), and a Taylor-Couette reactor (Marchisio and Barresi, 2003),
they have never been used successfully in stirred reactors so far.
THE DETAILS OF TURBULENT MIXING PROCESS 213
D. A PROMISING PROSPECT
At the end of this review on chemical reactors, special room is reserved for
a very promising approach, although this approach, too, has not yet been
applied for simulating a stirred chemical reactor. Van Vliet et al. (2001,
2005) exploited an elegant probabilistic approach (see also Van Vliet, 2003),
where the PDF methodology incorporates the joint scalar information by solv-
ing the transport equation for the full joint scalar PDF (Pope, 1985). In this
way, the second-order and thus nonlinear reaction terms in the CDR equation
are kept in closed form, making further modeling of the chemical reaction term
redundant.
In order to implement the PDF equations into a LES context, a ltered
version of the PDF equation is required, usually denoted as ltered density
function (FDF). Although the LES ltering operation implies that SGS mode-
ling has to be taken into account in order to capture micromixing effects, the
reaction term remains closed in the FDF formulation. Van Vliet et al. (2001)
showed that the sensitivity to the Damkohler number of the yield of competitive
parallel reactions in isotropic homogeneous turbulence is qualitatively well
predicted by FDF/LES. They applied the method for calculating the selectivity
for a set of competing reactions in a tubular reactor at Re 4,000.
Although this LES/FDF methodology is a promising technique, the (current)
drawback is the high computational costs involved to obtain a numerical so-
lution of the FDF transport equation. In the above study due to Van Vliet
(2003), the LES uid transport was computed with the help of an LB solver on a
5 10
6
computational grid. Solving a transport equation for the joint PDF of
the chemical species is most effectively done in a Lagrangian Monte-Carlo (MC)
manner: the chemical composition of the ow as a function of time and space is
represented by a collection of ctitious particles that are randomly released in
the ow domain and that carry with them the full chemical composition. The
assembly of MC particles is tracked through physical and chemical space by a
set of stochastic ordinary differential equations, where the random term rep-
resents diffusion. These equations need closure as to the way the particles in-
teract with their direct chemical environment, more specically for the scalar
energy dissipation rate. The model used is the rather common Interaction by
Exchange with the Mean (IEM) model in which a mixing frequency describes
the mixing at SGS.
Van Vliet et al. (2005) tracked 1 10
8
computational nodes to obtain a stoc-
hastic solution of their FDF equations. In order to deal with the high com-
putational costs, the code was run in parallel on a Linux cluster of 11 dual
AMD Athlon (TM) MP 1800+processors. In this way, about one turbulent
macro time scale (or 8,000 computational steps) per 2 days was computed.
Van Vliet et al. (2004, 2006) investigated the formation of hot spots and
reactor efciency in various geometrical congurations of a tubular reactor
for manufacturing Low-Density Polyethylene (LDPE) by means of the above
HARRY E. A. VAN DEN AKKER 214
LES/FDF-approach. An In situ Adaptive Tabulation (ISAT) technique (due
to Pope) was used to greatly reduce (by a factor of 5) the CPU time needed to
solve the set of stiff differential equations describing the fast LDPE kinetics.
Fig. 17 shows some of the results of interest: the occurrence of hot spots in the
tubular LDPE reactor provided with some feed pipe through which the initiator
(peroxide) is supplied. The 2004-simulations were carried out on 34 CPUs
(3 GHz) with 34 GB shared memory, but still required 34 h per macroow time
scale; they served as a demo of the method. The 2006-simulations then dem-
onstrated the impact of installing mixing promoters and of varying the inlet
temperature of the initiator added.
The above simulations as to the occurrence of hot spots once more illustrate
the power and promises of LES over RANS-type simulations. The hot spots can
never be found by means of a RANS-type of simulation. The same technique
was used by Van Vliet et al. (2006) to study the inuence of the injector geo-
metry and inlet temperature on product quality and process efciency in
the LDPE reactor. On the contrary, the RANS-based simulations due to
R. A. Bakker and Van den Akker (1994, 1996) were pretty much suited to arrive
at yield predictions for a fed batch reactor as a whole.
So far, to the best of our knowledge, the above LES/FDF-approach has not
been applied to stirred chemical reactors in which the turbulent-ow eld is far
more complex than in a tubular reactor. This LES/FDF-approach, however,
may be the way to go, as it provides highly detailed information on turbulent
reactive ows with the usage of a minimum of modeling assumptions. Although
the high computational demands make LES/FDF simulations currently acces-
sible to academic research groups only, the continued exponential growth of
FIG. 17. Three different representations (using increasing temperature thresholds) of hot spots in
a tubular LDPE reactor as found by the LES/FDF-methodology due to Van Vliet et al. (2004).
THE DETAILS OF TURBULENT MIXING PROCESS 215
computer resources will make them a versatile tool for process and geometry
optimization of turbulent reactive ows in the process industries.
IX. Summary and Outlook
The above review has shown that 20 years of developing CFD-techniques has
yielded us substantial simulation capabilities for studying and predicting mixing
under turbulent conditions.
The start in the 1980s was slow and with much trial and error. In the early
1990s, we had to be content with RANS-based simulations onspeaking af-
terwardscoarse grids of limited size only. With increasing computer power
and memory becoming available at lower cost, ner and larger grids offered the
potential of getting more detailed pictures, among other things via DNS. LES
entered the mixing scene, although their proliferation suffered from slow con-
vergence of the FV solvers. In the late 1990s, however, LB solvers entered the
mixing eld and, owing to being faster and better geared to parallellization,
made LES much more attractive and viable.
A. THE VARIOUS COMPUTATIONAL FLUID DYNAMICS OPTIONS
Nowadays, it is therefore essential to distinguish between the various main
CFD options for dealing with turbulent mixing issues, viz.

RANS simulations: usually exploiting some ke turbulence model, intended


for global information on the average ow eld and the global transport
phenomena in full-scale process equipment, with additional output (of limited
condence level) on spatial distributions of k and e;

DNS simulations: delivering fully resolved transient elds of velocities and


other variables in either a ow domain of limited size under laminar or very
moderately turbulent ow conditions or in a periodic box with some pre-
scribed turbulence level;

LES: with some model of the SGS ow and transport phenomena, suited for
reproducingat the level of the grid cell sizerather detailed transient elds
of velocities and other transport variables in full-scale process equipment
operated under turbulent-ow conditions.
Commercial CFD software has become a reliable tool for carrying out simu-
lations for laminar ows andbased on RANSfor turbulent ows. Practising
engineers gradually have become convinced about the usefulness of RANS-
based simulations. This review, however, emphasizes that CFD now has much
more to offer. For practicing engineers confronted with mixing problems, it is
HARRY E. A. VAN DEN AKKER 216
important to realize that CFD is not inherently restricted to just the average
single-phase ow eld, but gradually is becoming more and more capable of
dealing with the details of turbulent eddies and two-phase ows. The perform-
ance of many physical operations and the yield and selectivity of many chemi-
cally reacting systems strongly depend on nonlinear interactions at the small
scales of turbulent ows.
An example: Hollander et al. (2001a) nicely demonstrated how the strong
inhomogeneities in stirred-tank ow result in unpredictable scale-up behav-
iour and that the impact of the detailed hydrodynamics and of the non-
uniform spatial particle distribution on agglomeration rate is larger and
more complex than usually assumed; their study once more illustrated the
risks of scale-up on the basis of keeping a single non-dimensional number.
Sophisticated CFD, especially on the basis of LES, offers an attractive
alternative indeed.
Compared to RANS simulations, DNS and LES are much better geared to
reproducing these small-scale processes. RANS simulations focus on the aver-
age ow only and by their nature just model the small scales rather rudiment-
arily. On the contrary, a DNS resolves all uid motions and a LES resolves
most part of the turbulence spectrum, i.e., all eddies larger than the grid cell size.
While DNS nowadays can be used for turbulent ows at Reynolds numbers up
to say 10,000 in simple geometries (channels, curved tubes) only, LES are quite
feasible for complex geometries, certainly when LB techniques are adopted.
B. THE PROMISES OF DIRECT NUMERICAL SIMULATIONS AND LARGE EDDY
SIMULATIONS
At the moment, DNS and LES for turbulent ows are still the playground
of academic research groups. These groups are making substantial progress,
however, in developing dedicated software forand building up competence
insimulating multiphase ows, transport phenomena, many types of physical
operations, and chemical reactions. Such dedicated software makes it possible
to dig into the details of the mechanisms of a variety of ow and transport
phenomenaoften beyond the current capabilities of experimental techniques.
That is why this review paper is anadmittedly provocativeplea for starting
the exploitation of the advantages of DNS and LES.
An example: rather than linking average bubble size to just or essentially the
(overall) power input of a particular vessel-impeller combination, dedicated
CFD (preferably DNS and LES) allows for studying (tracking) the response
of bubble size to local and spatial variations in the turbulence levels in a
stirred vessel. In this way, the validity of certain modeling assumptions may
be afrmed or disproved. Particularly, effects of spatial variations in e which
THE DETAILS OF TURBULENT MIXING PROCESS 217
remain hidden in the traditional engineering techniques, may surface as a
result of such dedicated CFD approaches. This type of dedicated CFD sim-
ulations offer a better and closer look into the details of ow and transport
phenomena than experimental techniques which, e.g., still are not capable of
delivering reliable high-resolution e-values.
The advantage of LES over RANS-based simulations is that in the former
approach modeling the effect of the unresolved scales of the ow is easier and
more straightforward, just because the SGS eddies are distinctly separated from
the vessel boundary conditions andas a resulttheir behavior is closer to the
ideal of isotropy rendering universal turbulence modeling feasible. This makes the
outcome of simulations less sensitive to deciencies in turbulence modeling.
Thisalong with the inherently transient character and the degree of detail of the
simulationsturns LES highly suited as a base for simulating physical operations
and chemical reactions carried out in stirred vessels. Whenever the performance
of these processes is strongly dependent on turbulent mixing, the degree to which
CFD simulations can be trusted depends on the ability to reproduce the com-
plicated nonlinear interactions of ow and transport phenomena across the vari-
ous turbulence scales. LES is then the CFD option to be recommended.
The present author even wonders whether we should not be satised with the
gross predictions of the current RANS methods and turn to LES for the details of
those single-phase and multiphase mixing processes which are dominated by the
spatially distributed turbulence. It is really a valid question how long we should
keep trying and improving the various RANS methods now the increased com-
puter power brings the much more sophisticated LES within reach. The very
nature of the RANS approach itselfparticularly the basic assumptions as to
averaging and the various modeling uncertainties as to turbulence and multiphase
owmay really set limits to its exploitation. The modest demands on computer
resources RANS-based simulations require these days are no excuse in this respect.
In addition, DNS of turbulent ow in a periodic box offer interesting oppor-
tunities for studying in a fully resolved mode the intimate details of the ow eld,
its interaction with particles and the mutual interaction between particles (in-
cluding particleparticle collisions and coalescence). Such simulations may yield
new insights; see, e.g., Ten Cate et al. (2004) and Derksen (2006b). The same can
be said about our understanding of particleturbulence interactions in wall-
bounded ows: this has increased due to Portela and Oliemans (2003) exploiting
both DNS and LES and due to Ten Cate et al. (2004).
C. AN OUTLOOK
Nowadays, CFD research at academia is heavily engaged in attempts to in-
clude microscale transport phenomena and microscale processes in the dedicated
codes under development with a view to reproduce such divergent processes as
HARRY E. A. VAN DEN AKKER 218
blending, dissolution, crystallization, precipitation, coalescence and redispersion
of bubbles and droplets, suspending solids, and chemical reactions. Essential
physical challenges are in nding proper models for the details of the ow. In
single-phase ows, we need better models for the unresolved contribution of
microscale transport phenomena such as micromixing, while multiphase ow
CFD looks for better models for the mutual interaction of turbulence and dis-
persed phase particles and for the interaction force(s) between the dispersed
phase particles and the ambient continuous phase. The devil is in the detail here
fully applies.
In developing multiphase ow CFD and in combining CFD with population
balances and various types of PDF approaches, one needs to keep the size of the
computational job under controlin spite of the overwhelming growth in com-
putational power (processor speed, memory, communication tools). This requires
on the one hand efcient and effective numerical tools and on the other hand
clever strategies for handling the enormous amounts of data. Local grid rene-
ment techniques may be of great help in avoiding an unnecessary degree of detail.
The development of the above more dedicated LES and DNS is promoted by
the introduction of LB techniques into the world of turbulent mixing simula-
tions. LB techniques provide a viable alternative for the more classical FV
solvers of the commercial CFD software, in particular in the context of parallel
simulations on multiple processors. LB techniques are also inherently faster
than FV techniques due to the locality of their operations. In addition, complex
boundaries are easier to implement in the LB approach than with FV solvers.
Substantial improvements in LB techniques have been effectedin terms of
immersed or embedded boundary methods for dealing with moving and curved
boundaries (impeller blades, solid particles) and of grid renement techniques
which have had a positive impact on the fast proliferation of dedicated CFD
tools. Here, too, the details of the computational techniques do matter.
Finally, the large number of processors used in many of the parallel simu-
lations cited is striking. It illustrates the enormous progress made in the size of
the simulations academic groups have realized. The falling prices of such pro-
cessors and the ease at which these can be combined into platforms for parallel
simulations may have the effect thatjust like in the past decade with RANS-
based simulationspretty soon industrial users can afford such dedicated and
detailed simulations, both LES and DNS, and can benet from their outcome in
dealing with their commercial targets.
NOTATION
a specic interface area
a
ij
anisotropy tensor, comprising, essentially, the turbulent stresses
made nondimensional with the turbulent kinetic energy k
THE DETAILS OF TURBULENT MIXING PROCESS 219
A distance to origin in (A
3
, A
2
) plane
A
1
, A
2
, invariants of anisotropy tensor a
ij
A
3
c concentration
c
s
Smagorinsky constant (in SGS modeling)
C
b1
coefcient in relation for local maximum bubble size d
bN
C
D
drag coefcient
C
Dt
drag coefcient in a free stream turbulence
C
m
coefcient in model equation for n
t
(in RANS models)
d
b
bubble size
d
bN
local maximum stable bubble size
D diffusion (or dispersion) coefcient
D
ij
specic rate of production of turbulent stresses
D
k
specic rate of production of turbulent kinetic energy
D
e
specic rate of production of e (Kolmogorov eddies)
g gravitational acceleration constant
I unity tensor
k concentration of turbulent kinetic energy
k
l
mass transfer coefcient
k
sgs
turbulent kinetic energy contained in the SGS eddies
m
0
particle number concentration
n
b
particle number density
n
b1
local equilibrium number density
N impeller speed (number of revolutions per unit of time)
p pressure
~ p pressure as resolved in LES
P average pressure (in RANS context)
P
ij
specic rate of production of turbulent stresses
P
k
specic rate of production of turbulent kinetic energy
Pe specic rate of production of e (Kolmogorov eddies)
q specic heat production rate
r specic rate of chemical reaction producing or consuming a
particular species
^ r radial vector component
~
S local resolved deformation rate (in LES)
S
g
source term in mass balance for gas phase (due to gas supply)
t time
T temperature
U
i
U
j
components of velocity vector v (in sufx notation)
U
k
u
i
u
j
average turbulent, or Reynolds, stresses
u gas velocity vector
U
s
slip (or: relative) velocity vector
v uid velocity vector
HARRY E. A. VAN DEN AKKER 220
~ v uid velocity vector as resolved in LES
V average uid velocity vector (in RANS context)
V
b
bubble volume
V
b,in
bubble volume at position of gas supply
w azimuthal component of velocity vector
x
i
, x
k
spatial coordinate
X
Q
yield of product Q
^ z vertical vector component
GREEK SYMBOLS
a volume fraction of gas
b
0
agglomeration coefcient
d
ij
Kronecker delta
D grid spacing
e specic rate at which turbulent kinetic energy is dissipated (in the
Kolmogorov eddies)

ij
specic rate at which turbulent stresses are dissipated
Z
B
Batchelor length scale (proportional to penetration depth for
diffusion)
Z
K
Kolmogorov length scale (smallest scale in turbulent ow)
k thermal conductivity coefcient
n kinematic viscosity coefcient
n
e
effective SGS viscosity coefcient (in LES)
n
t
turbulent viscosity coefcient (in RANS)
P
ij
specic rate of production of turbulent stresses
P
k
specic rate of production of turbulent kinetic energy
r uid density
r
l
liquid density
s interfacial tension
s shear stress tensor as resolved in LES
s
0
part of s, see Eq. (3)
o effective break-up/agglomeration coefcient (a kind of relaxation
parameter)
o vorticity
O specic rate of destruction of e
DIMENSIONLESS NUMBERS
Re Reynolds number
Sc Schmidt number
THE DETAILS OF TURBULENT MIXING PROCESS 221
ABBREVIATIONS
ASM Algebraic Stress Model
CDR Convection-Diffusion-Reaction
CFD Computational Fluid Dynamics
CSV Cylindrical Stretched Vortex
CT Computed Tomography
DNS Direct Numerical Simulation
E Engulfment
EDD Engulfment-Deformation-Diffusion
FDF Filtered Density Function
FV Finite Volume
IEM Interaction by Exchange with the Mean
ISAT In situ Adaptive Tabulation
LB Lattice-Boltzmann
LDA Laser Doppler Anemometry
LDV Laser Doppler Velocimetry
LDPE Low Density Poly Ethylene
LES Large Eddy Simulation
LGA Lattice Gas Automaton
MC Monte Carlo
MFR Multiple Frames of Reference
NS NavierStokes
PDF Probability Density Function
QMOM Quadrature Method of Moments
RANS Reynolds Averaged Navier Stokes
RSM Reynolds Stress Model
SGS Sub Grid Scale
SDM Sliding and Deforming Mesh
SM Sliding Mesh
Sh Shear
St Strain
TVD Total Variation Diminishing
VOF Volume of Fluid
ACKNOWLEDGEMENTS
First of all, Dr. Jos J. Derksen of the Department of Multi-Scale Physics at
Delft University of Technology is gratefully acknowledged for a fruitful long-
time collaboration and for his critical review of the draft paper. The author is
also indebted to all former PhD students of the Kramers Laboratorium voor
HARRY E. A. VAN DEN AKKER 222
Fysische Technologie of Delft University of Technology for contributing
through their PhD projects and theses to the development of the views and
capabilities described in this chapter.
REFERENCES
Abbott, M. B., and Basco, D. R., Computational Fluid Dynamics: An Introduction for
Engineers. Longman Scientic & Technical, Harlow (UK) (1989).
Akiti, O., and Armenante, P. M. AIChE J 50, 566577 (2004).
Artoli, A. M., Hoekstra, A. G., and Sloot, P. M. A. J. Mod. Phys. B 17(12), 9598 (2003).
Bakker, A., Hydrodynamics of stirred gasliquid dispersions, Ph.D. Thesis, Delft University of
Technology, Delft, Netherlands (1992).
Bakker, R. A., Micromixing in chemical reactors: models, experiments and simulations, Ph.D.
Thesis, Delft University of Technology, Delft, Netherlands (1996).
Bakker, A., and Fasano, J. B., Time-Dependent, Turbulent Mixing and Chemical Reaction in
Stirred Tanks, AIChE Symposium. Series No 299 90 7178 (1993).
Bakker, A., LaRoche, R. D., Wang, M. H., and Calabrese, R. V. Chem. Eng. Res. Des. 75A, 4244
(1997).
Bakker, A., Oshinowo, L. M., and Marshall, E. M., The Use of Large Eddy Simulation to Study
Stirred Vessel Hydrodynamics. Proceedings of the 10th European Conference on Mixing,
Delft, Netherlands, 247254 (2000).
Bakker, A., and Oshinowo, L. M. Chem. Eng. Res. Des. 82(A9), 11691178 (2004).
Bakker, A., and Van den Akker, H. E. A., A Computational Study on Dispersing Gas in a Stirred
Reactor. Proceedings of the 7th European Conference on Mixing, Brugue, Belgium
199207. Also in: Fluid mechanics of mixing: modelling, operations and experimental tech-
niques, (R. King, Ed.) Fluid Mechanics and its Applications, 10, 3746. Kluwer
Academic Publishers (1991).
Bakker, R. A., and Van den Akker, H. E. A. Chem. Eng. Des. 72A, 733738 (1994).
Bakker, A., and Van den Akker, H. E. A. Chem. Eng. Res. Des. 72A, 583593 (1994a).
Bakker, A., and Van den Akker, H. E. A. Chem. Eng. Res. Des. 72A, 594605 (1994b).
Bakker, R. A., and Van den Akker, H. E. A. Chem. Eng. Sci. 51, 26432648 (1996).
Baldyga, J. Chem. Eng. Sci. 49, 19852003 (1994).
Baldyga, J., and Bourne, J. R. Chem. Eng. Commun. 28, 243258 (1984a).
Baldyga, J., and Bourne, J. R. Chem. Eng. Commun. 28, 259281 (1984b).
Baldyga, J., and Henczka, M., Turbulent mixing and parallel chemical reactions in a pipe: appli-
cation of a closure model, Recents Progre`s en Genie des Procedes 11, 341348 (1997).
Bermingham, S. K., Kramer, H. J. M., and Van Rosmalen, G. M. Comp. Chem. Eng. 22, 355362
(1998).
Bourne, J. R., and Yu, S., An Experimental Study of Micromixing Using Two Parallel Reactions.
Proceedings of the 7th European Conference on Mixing, Brugues, Belgium, 6775 (1991).
Bouwmans, I., The blending of liquids in stirred vessels, Ph.D. Thesis, Delft University of Tech-
nology, Delft, Netherlands (1992).
Bouwmans, I., Bakker, A., and Van den Akker, H. E. A. Chem. Eng. Res. Des. 75A, 777783 (1997).
Brucato, A., Grisa, F., and Montante, G. Chem. Eng. Sci. 53, 32953314 (1998).
Bujalski, W., Jaworski, Z., and Nienow, A. W. Chem. Eng. Res. Des. 80, 97104 (2002).
Chen, S., and Doolen, G. D. Ann. Rev. Fluid Mech 30, 329364 (1998).
Colenbrander, G. W., Experimental Findings on the Scale-Up Behaviour of the Drop Size Dis-
tribution of LiquidLiquid Dispersions in Stirred Vessels. Proceedings of the 10th European
Conference on Mixing, Delft, Netherlands, 173180 (2000).
THE DETAILS OF TURBULENT MIXING PROCESS 223
Curtis, J. S., and Van Wachem, B. AIChE J 50, 26382645 (2004).
Derksen, J. J. Chem. Eng. Res. Des. 79A, 824830 (2001).
Derksen, J. J. Lecture Notes Comput. Sci. 2329, 713722 (2002a).
Derksen, J. J. Flow Turbulence Combustion 69, 333 (2002b).
Derksen, J. J. AIChE J 49, 27002714 (2003).
Derksen, J. J. Chem. Eng. Res. Des. 84(A1), 3846 (2006a).
Derksen, J. J., Multi-scale simulations of stirred liquidliquid dispersions. 12th European Con-
ference on Mixing, Bologna, Italy, pp. 447454 (2006b).
Derksen, J. J., Doelman, M. S., and Van den Akker, H. E. A. Exp. Fluids 27, 522532 (1999).
Derksen, J. J., Kooman, J. L., and Van den Akker, H. E. A., A parallel DNS implementation for
conned swirling ow, in HPCN Challenges in Telecomp and Telecom: Parallel Simulation
of Complex Systems and Large-Scale Applications (L. Dekker, et al., Eds.), pp. 237244.
Elsevier, Amsterdam (1996).
Derksen, J. J., Kooman, J. L., and Van den Akker, H. E. A., Parallel ow simulations by means of
a latticeBoltzmann scheme, In: B. Hertzberger, P. Sloot (Eds.), High-Performance Com-
puting and Networking, Lecture Notes in Computer Science 1225, 524530 (1997).
Derksen, J. J., and Van den Akker, H. E. A., Large eddy simulation of stirred tank ow by means
of a lattice-Boltzmann scheme, In: C. R. Kleijn, S. Kawano (Eds.), ASME Proceedings
Volume PVP-377-2, ASME, 1116 (1998).
Derksen, J. J., and Van den Akker, H. E. A. AIChE J 45, 209221 (1999).
Derksen, J. J., and Van den Akker, H. E. A. AIChE J 46, 13171331 (2000).
Dimotakis, P. E. Annu. Rev. Fluid Mech. 37, 329356 (2005).
Distelhoff, M. F. W., Marquis, A. J., Nouri, J. M., and Whitelaw, J. H. Can. J. Chem. Eng. 75,
641652 (1997).
Ditl, P., and Rieger, F. Chem. Eng. Progr. 102(1), 2230 (2006).
Ducci, A., and Yianneskis, M. Chem. Eng. Sci. 61, 27802790 (2006).
Eggels, J. G. M. Int. J. Heat Fluid Flow 17(3), 307323 (1996).
Eggels, J. G. M., and Somers, J. A. Int. J. Heat Fluid Flow 16(5), 357364 (1995).
Escudie , R., and Line , A. Chem. Eng. Sci. 61, 27712779 (2006).
Fox, R. O. Rev. Inst. Franc- . du Petrole 51(2), 215243 (1996).
Fox, R. O., Computational Models for Turbulent reacting Flows. Cambridge University Press,
Cambridge, UK (2003).
Frisch, U., Turbulence, the Legacy of A. N. Kolmogorov. Cambridge University Press,
Cambridge, UK (1995).
Frisch, U., Hasslacher, B., and Pomeau, Y. Phys. Rev. Lett. 56, 15051508 (1986).
Gao, Z., and Min, J. Chinese J. Chem. Eng. 14, 17 (2006).
Gao, Z., Min, J., Smith J. M., and Thorpe, R. B., Large Eddy Simulation of Mixing Time in a
Stirred Tank with Duals Rushton Turbines. 12th European Conference on Mixing,
Bologna, Italy, pp. 431438 (2006).
Gentric, C., Mignon, D., Bousquet, J., and Tanguy, P. A. Chem. Eng. Sci. 60, 22532272 (2005).
Grenville, R. K., Blending of viscous and pseudo-plastic uids, Ph.D. Thesis, Craneld Institute
of Technology, Craneld (UK) (1992).
Grenville, R. K., and Nienow, A. W., Blending of miscible liquids, In: NAMF Handbook of
Industrial Mixing (E. L. Paul, V. A. Atiemo-Obeng, and S. M. Kresta, Eds.), Wiley,
Hoboken (NJ, USA) (2004).
Haam, S., Brodkey, R. S., and Fasano, J. B. Ind. Eng. Chem. Res. 31, 13841391 (1992).
Hartmann, H., Detailed simulations of liquid and liquid-solid mixingturbulent agitated ow and
mass transfer, Ph.D. Thesis, Delft University of Technology, Delft, Netherlands (2005).
Hartmann, H., Derksen, J. J., Montavon, C., Pearson, J., Hamill, I. S., and Van den Akker, H. E. A.
Chem. Eng. Sci. 59, 24192432 (2004a).
Hartmann, H., Derksen, J. J., and Van den Akker, H. E. A. AIChE J 50, 23832393 (2004b).
Hartmann, H., Derksen, J. J., and Van den Akker, H. E. A. Chem. Eng. Sci. 61, 30253032 (2006).
HARRY E. A. VAN DEN AKKER 224
Harvey, P. S., and Greaves, M. Trans. IChemE 60, 201210 (1982).
Harvey, A. D., Lee, C. K., and Rogers, S. E. AIChE J 41, 21772186 (1995).
Harvey, A. D., and Rogers, S. E. AIChE J 42, 27012712 (1996).
Hasal, P., Montes, J. L., Boisson, H. C., and Fort, I. Chem. Eng. Sci. 55, 391401 (2000).
Hinze, J. O. AIChE J 1, 289295 (1955).
Hirt, C. W., and Nichols, B. D. J. Comput. Phys. 39, 201225 (1981).
Hoekstra, A. J., Gas ow eld and collection efciency of cyclone separators, Ph.D. Thesis, Delft
University of Technology, Delft, Netherlands (2000).
Hollander, E. D., Shear induced agglomeration and mixing, Ph.D. Thesis, Delft University of
Technology, Delft, Netherlands (2002).
Hollander, E. D, Derksen, J. J., Bruinsma, O. S. L., Van Rosmalen, G. M., and Van den Akker, H.
E. A., A Numerical Investigation into the Inuence of Mixing on Orthokinetic Agglom-
eration. Proceedings of the 10th European Conference on Mixing, Delft, Netherlands,
221230 (2000).
Hollander, E. D., Derksen, J. J., Portela, L. M., and Van den Akker, H. E. A. AIChE J 47,
24252440 (2001a).
Hollander, E. D., Derksen, J. J., Bruinsma, O. S. L., Van den Akker, H. E. A., and Van Rosmalen,
G. M. Chem. Eng. Sci. 56, 25312541 (2001b).
Hollander, E. D., Derksen, J. J., Kramer, H. M. J., Van Rosmalen, G. M., and Van den Akker, H.
E. A. Powder Technol. 130, 169173 (2003).
Holmes, D. B., Voncken, R. M., and Dekker, J. A. Chem. Eng. Sci. 19, 201208 (1964).
Hoogendoorn, C. J., and Den Hartog, A. P. Chem. Eng. Sci. 22, 16891699 (1967).
Hristov, H. V., Mann, R., Lossev, V., and Vlaev, S. D. Trans. IChemE, Food Bioproducts Process
82(C1), 2134 (2004).
Issa, R., and Gosman, A. D., The Computation of Three-Dimensional Turbulent Two-Phase Flow
in Mixer Vessels. Proc. 2nd Int. Conf. Num. Meth. Lam. Turb. Flows, Venice, Italy (1981).
Jahoda, M., Mos te k, M., Kukukova , A., and Machon , V., CFD Modelling of Liquid Homoge-
nisation in Stirred Tanks With One and Two Impellers Using Large Eddy Simulation. 12th
European Conference on Mixing, Bologna, Italy, pp. 455462 (2006).
Jaworski, Z., Bujalski, W., Otomo, N., and Nienow, A. W. Chem. Eng. Res. Des. 78, 327333 (2000).
Kandhai, D., Derksen, J. J., and Van den Akker, H. E. A. AIChE J 49, 10601065 (2003).
Khopkar, A. R., Rammohan, A. R., Ranade, V. V., and Dudukovic, M. P. Chem. Eng. Sci. 60,
22152229 (2005).
Khopkar, A. R., Kasat, G. R., Pandit, A. B., and Ranade, V. V. Chem. Eng. Sci. 61, 29212929
(2006).
Kramers, H., Baars, G. M., and Knoll, W. H. Chem. Eng. Sci. 2, 3542 (1953).
Kresta, S. M., and Brodkey, R. S., Turbulence in mixing applications, Ch.2, In: Paul, E. L., Atiemo-
Obeng, V. A., Kresta, S. M. (Eds.),NAMF Handbook of Industrial Mixing, Wiley,
Hobken (NJ, USA) (2004).
Laakkonen, M., Alopeus, V., and Aittamaa, J. Chem. Eng. Sci. 61, 218228 (2006).
Lance, M., Marie , J. L., and Bataille, J. J. Fluids Eng. 113, 295300 (1991).
Lane, G. L., Schwarz, M. P., and Evans, G. M., Modelling of the Interaction Between Gas and
Liquid in Stirred Vessels. Proceedings of the 10th European Conference on Mixing, Delft,
Netherlands, 197204 (2000).
Lapin, A., Mu ller, D., and Reuss, M. Ind. Eng. Chem. Res. 43, 46474656 (2004).
Lathouwers, D., Modelling and simulation of turbulent bubbly ow, Ph.D. Thesis, Delft
University of Technology, Delft, Netherlands (1999).
Lathouwers, D., and Van den Akker, H. E. A., A numerical method for the solution of two-uid
model equations. Proceedings of the Fluids Eng. Div. 1996 Summer Meeting, San Diego
(CA, USA), ASME, New York (USA), Vol. 1, 121126 (1996).
Lee, K. C., An experimental investigation of the trailing vortex structure and mixing characteristics
of mixing vessels, Ph.D. Thesis, Kings College, London, (UK) (1995).
THE DETAILS OF TURBULENT MIXING PROCESS 225
Liu, S. Chem. Eng. Sci. 54, 873891 (1999).
Lo, S., Application of population balance to CFD modelling of gasliquid reactors. Conference
on Trends in Numerical and Physical Modelling for Industrial Multiphase Flows, Carge` se,
Corse 2729 September (2000).
Lu, Z., Liao, Y., Qian, D., McLaughlin, J. B., Derksen, J. J., and Kontomaris, K. J. Comput.
Physics 181, 675704 (2002).
Lumley, J. Adv. Appl. Mech. 24, 123176 (1978).
Luo, J. Y., Gosman, A. D., Issa, R. I., Middleton, J. C., and Fitzgerald, M. K. Trans. IChemE. 71A,
342344 (1993).
Luo, J. Y., Issa, R. I., and Gosman, A. D. IChemE Symp. Ser. 136, 549556 (1994).
Luo, H., and Svendsen, H. F. AIChE J 42, 12251233 (1996).
Marchisio, D. L., and Barresi, A. A. Chem. Eng. Sci. 58, 35793587 (2003).
Marchisio, D. L., Pikturna, J. T., Fox, R. O., Vigil, R. D., and Barresi, A. A. AIChE J 49, 12661276
(2003).
Marshall, E., and Bakker, A., Computational Fluid Mixing. Fluent Inc. Lebanon, NH; also as
Ch. 5 In: Paul, E. L., Atiemo-Obeng, V. A., Kresta, S. M. (Eds.), NAMF Handbook of
Industrial Mixing, Wiley, Hoboken (NJ, USA) (2004).
Mason, P. J., and Callen, N. S. J. Fluid Mech. 162, 439462 (1986).
McNamara, G., and Zanetti, G. Phys. Rev. Lett. 61, 23322335 (1988).
Menter, F. R. AIAA J 32(8), 269289 (1994).
Micale, G., Montante, G., Grisa, F., Brucato, A., and Godfrey, J. Chem. Eng. Res. Des. 78,
435444 (2000).
Micale, G., Grisa, F., Rizzuti, L., and Brucato, A. Chem. Eng. Res. Des. 82, 12041213 (2004).
Micheletti, M., Baldi, S., Yeoh, S. L., Ducci, A., Papadakis, G., Lee, K. C., and Yianneskis, M.
Chem. Eng. Res. Des. 82, 11881198 (2004).
Middleton, J. C., Pierce, F., and Lynch, P. M. Chem. Eng. Res. Des. 64, 1822 (1986).
Mittal, R., and Iaccarino, G. Annu. Rev. Fluid Mech. 37, 239261 (2005).
Moin, P., and Kim, J., Sci. Am., January 4652 (1997).
Montante, G., Micale, G., Brucato, A., and Magelli, F., CFD Simulation of Particle Distribution
in a Multiple-Impeller High-Aspect-Ratio Stirred Vessel. Proceedings of the 10th European
Conference on Mixing, Delft, Netherlands, 125132 (2000).
Montante, G., and Magelli, F. Chem. Eng. Res. Des. 82, 11791187 (2004).
Montante, G., and Magelli, F. Int. J. Comp. Fluid Dynam. 19, 253262 (2005).
Montante, G., Bakker, A., Paglianti, A., and Magelli, F. Chem. Eng. Sci. 61, 28072814 (2006).
Mumtaz, H. S., Hounslow, M. J., Seaton, M. J., and Paterson, W. R. Chem. Eng. Res. Des. 75,
152159 (1997).
Murthy, J. Y., Mathur, S. R., and Choudhury, D. IChemE Symp.Ser. 136, 341345 (1994).
Myers, K. J., Ward, R. W., and Bakker, A. ASME J. Fluids Eng. 119, 623632 (1997).
Nienow, A. W. Chem. Eng. Sci. 52, 25572565 (1997).
Nikiforaki, L., Montante, G., Lee, K. C., and Yianneskis, M. Chem. Eng. Sci. 58, 29372949 (2002).
Oey, R. S., Mudde, R. F., and Van den Akker, H. E. A. AIChE J 49, 16211636 (2003).
Osman, J. J., and Varley, J. IChemE Symp.Ser. 146, 1522 (1999).
Ottino, J. M. Chem. Eng. Sci. 35, 13771391 (1980).
Patankar, S. V., Numerical Heat Transfer and Fluid Flow. Hemisphere Publishing Corporation,
New York (USA) (1980).
Patterson, G. K., Modelling of turbulent reactors, Ch. 3, In: Mixing of Liquids by Mechanical
Agitation. (J. J. Ulbrecht and G. K. Patterson, Eds.), Gordon and Breach Science Pub-
lishers, New York (USA) (1985).
Patterson, G. K., Paul, E. L., Kresta, S. M., and Etchells, A. W. III., Mixing and chemical
reactions, Ch. 13, In: NAMF Handbook of Industrial Mixing (E. L. Paul, V.A. Atiemo-
Obeng, and S. M. Kresta, Eds.), Wiley, Hoboken (USA) (2004).
HARRY E. A. VAN DEN AKKER 226
Paul, E. L., Atiemo-Obeng, V. A., and Kresta, S. M., NAMF Hand book of Industrial Mixing.
Wiley, Hoboken (USA) (2004).
Pericleous, K. A., and Patel, M. K. Physico Chem. Hydrodynam 8, 105123 (1987).
Pinelli, D., Nocentini, M., and Magelli, F. Chem. Eng. Commun. 188, 91107 (2001).
Placek, J., and Tavlarides, L. AIChE J 31, 11131120 (1985).
Placek, J., Tavlarides, L., Smith, G. W., and Fort, I. AIChE J 32, 17711785 (1986).
Pope, S. B. Prog. Energy Combust. Sci. 11, 119192 (1985).
Portela, L. M., and Oliemans, R. V. A. Int. J. Numer. Meth. Fluids 43, 10451065 (2003).
Praturi, A. K., and Brodkey, R. S. J. Fluid Mech. 89, 251272 (1978).
Procha zka, J., and Landau, J. Coll. Czech. Chem. Commun 26, 29612973 (1961).
Ranade, V. V., Computational Flow Modeling for Chemical Reactor Engineering, Volume 5 of
Process Systems Engineering (G. Stephanopoulos and J. Perkins, Eds.), Academic Press, San
Diego (CA, USA) (2002).
Ranade, V. V., Bourne, J. R., and Joshi, J. B. Chem. Eng. Sci. 46, 18831893 (1991).
Ranade, V. V., and Dommeti, S. M. S. Chem. Eng. Res. Des. 74A, 476484 (1996).
Ranade, V. V., Joshi, J. B., and Marathe, A. G. Chem. Eng. Commun. 81, 225248 (1989).
Ranade, V. V., and Van den Akker, H. E. A. Chem. Eng. Sci. 49, 51755192 (1994).
Ranz, W. E. AIChE J 25, 4147 (1979).
Revstedt, J., Fuchs, L., and Tra ga rdh, Ch. Chem. Eng. Sci. 53, 40414053 (1998).
Revstedt, J., Fuchs, L., Kova cs, T., and Tra ga rdh, Ch. AIChE J 46, 23732382 (2000).
Revstedt, J., and Fuchs, L. Chem. Eng. Technol. 25, 443446 (2002).
Rielly, C. D., and Marquis, A. J. Chem. Eng. Sci. 56, 24752493 (2001).
Rietema, K., and Van den Akker, H. E. A. Int. J. Multiphase Flow 9, 2136 (1983).
Rodi, W.,Turbulence models and their application in hydraulicsa state of the art review,
International Association for Hydraulic Research, Delft (NL), reprinted in 1984 (1984).
Rohde, M., Extending the Lattice-Boltzmann methodnovel techniques for local grid renement
and boundary conditions, Ph.D. Thesis, Delft University of Technology, Delft, Netherlands
(2004).
Rohde, M., Derksen, J. J., and Van den Akker, H. E. A., Phys. Rev. E 65, Paper No. 056701 (2002).
Rohde, M., Kandhai, D., Derksen, J. J., and Van den Akker, H. E. A., Phys. Rev. E 67, Paper No.
066703 (2003).
Rohde, M., Kandhai, D., Derksen, J. J., and Van den Akker, H. E. A. Int. J. Numer. Meth. Fluids
51(7), 439468 (2006).
Rous ar, I., and Van den Akker, H. E. A., LDA Measurements of Liquid Velocities in Sparged
Agitated Tanks with Single and Multiple Rushton Turbines. 8th European Conference on
Mixing, Cambridge, UK IChemE Symp. Ser., 136, 8996 (1994).
Roussinova, V., Kresta, S. M., and Weetman, R. Chem. Eng. Sci. 58, 22972311 (2003).
Rushton, J. H., Costich, E. W., and Everett, H. J. Chem. Eng. Progr 46, 395404 467476 (1950).
Ruszkowski, S., A Rational Method for Measuring Blending Performance, and Comparison of
Different Impeller Types. Proceedings of the 8th European Conference on Mixing,
Cambridge, UK, pp. 283291 (1994).
Sano, Y., and Usui, H. J. Chem. Eng Japan 18, 4752 (1985).
Scha fer, M., Ho fken, M., and Durst, F. Chem. Eng. Res. Des. 75A, 729736 (1997).
Scha fer, M., Yianneskis, M., Wa chter, P., and Durst, F. AIChE J 44, 12331246 (1998).
Schulze, K., Ritter, J., and Kraume, M., Investigations of Local Drop Size Distributions and Scale-
Up in Stirred LiquidLiquid Dispersions. Proceedings of the 10th European Conference on
Mixing, Delft, Netherlands, 181188 (2000).
Seckler, M. M., Bruinsma, O. S. L., and Van Rosmalen, G. M. Chem. Eng. Commun. 135, 113131
(1995).
Shaw, C. T., Using Computational Fluid Dynamics. Prentice Hall International Ltd, Hemel
Hempstead (UK) (1992).
THE DETAILS OF TURBULENT MIXING PROCESS 227
Shyy, W. Adv. Heat Transfer 24, 191275 (1994).
Smagorinsky, J. Monthly Weather Rev. 91, 99164 (1963).
Somers, J. A. Appl. Sci. Res. 51, 127133 (1993).
Sommerfeld, M., and Decker, S. Chem. Eng. Technol. 27(3), 215224 (2004).
Stekelenburg, A. J. C., Van der Hagen, T. H. J. J., and Van den Akker, H. E. A. Int. J. Num. Meth.
Heat Fluid Flow 4, 143158 (1994).
Succi, S., The Lattice Boltzmann Equation for Fluid Dynamics and Beyond. Oxford University
Press, New York (USA) (2001).
Ten Cate, A., Turbulence and particle dynamics in dense crystal slurriesa numerical study by
means of lattice-Boltzmann simulations, Ph.D. Thesis, Delft University of Technology,
Delft, Netherlands (2002).
Ten Cate, A., Bermingham, S. K., Derksen, J. J., and Kramer, H. M. J., Compartmental Modeling
of a 1,100 L Crystallizer Based on Large Eddy Flow Simulation. Proceedings of the 10th
European Conference on Mixing, Delft, Netherlands, 255264 (2000).
Ten Cate, A., Derksen, J. J., Kramer, H. J. M., Van Rosmalen, G. M., and Van den Akker, H. E. A.
Chem. Eng. Sci. 56, 24952509 (2001).
Ten Cate, A., Derksen, J. J., Portela, L. M., and Van den Akker, H. E. A. J. Fluid Mech. 519,
233271 (2004).
Tennekes, H., and Lumley, J. L., A First Course in Turbulence. MIT Press, Cambridge (MA,
USA) (1972).
Thornock, J. N., and Smith, P. J. WIT Trans. Built Environ 84, 573583 (2005).
Tra ga rdh, Ch., A Hydrodynamic Model for the Simulation of an Aerated Agitated Fed-Batch
Fermentor. Proceedings of the 2nd International Conference on Bioreactor Fluid Dynamics,
Cambridge, UK 117134 (1988).
Tsouris, C., and Tavlarides, L. L. AIChE J 40, 395406 (1994).
Van den Akker, H. E. A.,On Status and Merits of Computational Fluid Dynamics. In: Nienow,
A. W. (Ed.), Proceedings of the 4th International Conference on Bioreactor and Bioprocess
Fluid Dynamics, BHR, Edinburgh, UK 407432 (1997).
Van den Akker, H. E. A. ERCOFTAC Bull 36, 3033 (1998).
Van den Akker, H.E.A., Momentum Equations in Dispersed Two-Phase Flows. In:
Cheremisinoff, N.P. (Ed.), Encyclopedia of Fluid Mechanics, Gulf Publishing Company,
Houston (TX, USA), Vol. 3, Chapter 15, 371400 (1986).
Van den Akker, H. E. A., Computational uid dynamics: more than a promise to chemical reaction
engineering. Plenary paper presented at CHISA, Prague, CZ Paper #1270 (2000).
Van Leeuwen, M. L. J., Precipitation and mixing, Ph.D. Thesis, Delft University of Technology,
Delft, Netherlands (1998).
Van Leeuwen, M. L. J., Bruinsma, O. S. L., and Van Rosmalen, G. M. Chem. Eng. Sci. 51,
25952600 (1996).
Van Santen, H., Lathouwers, D., Kleijn, C. R., and Van den Akker, H. E. A., Inuence of
segregation on the efciency of nite volume methods for the incompressible NavierStokes
equations. Proceeding of the Fluids Eng. Div. 1996 Summer Meeting, San Diego (CA,
USA), ASME, New York (USA), Vol. 3, 151158 (1996).
Van Vliet, E., Turbulent reactive mixing in process equipment, Ph.D. Thesis, Delft University of
Technology, Delft, Netherlands (2003).
Van Vliet, E., Derksen, J. J., and Van den Akker, H. E. A., Modelling of Parallel Competitive
Reactions in Isotropic Homogeneous Turbulence Using a Filtered Density Function Ap-
proach for Large Eddy Simulations. Proc. PVP01 3rd Int. Symp. on Comput. Techn. for
Fluid/Thermal/Chemical Systems with Industrial Appl., Atlanta, GE, USA (2001).
Van Vliet, E., Derksen, J. J., and Van den Akker, H. E. A., A numerical study of a low-density
polyethylene tubular reactor using a 3-D FDF/LES approach, AIChE 2004 Annual Mtg.,
Austin, TX, USA (2004).
Van Vliet, E., Derksen, J. J., and Van den Akker, H. E. A. AIChE J 51, 725739 (2005).
HARRY E. A. VAN DEN AKKER 228
Van Vliet, E., Derksen, J. J., and Van den Akker, H. E. A., Numerical Study on the Turbulent
Reacting Flow in the Injector Region of an LDPE Tubular Reactor. Proceedings of the 12th
European Conference on Mixing, Bologna, Italy, pp. 719726 (2006).
Van Wageningen, W. F. C., Kandhai, D., Mudde, R. F., and Van den Akker, H. E. A. AIChE J 50,
16841696 (2004).
Venneker, B. C. H., Turbulent ow and gas dispersion in stirred vessels with pseudo plastic uids,
Ph.D. Thesis, Delft University of Technology, Delft, Netherlands (1999).
Venneker, B. C. H., Derksen, J. J., and Van den Akker, H. E. A. AIChE J 48, 673685 (2002).
Voncken, R. M., Holmes, D. B., and Den Hartog, H. W. Chem. Eng. Sci. 19, 209213 (1964).
Von Smoluchowski, M. Phys. Chem. 92, 129156 (1917).
Vuik, C., Fast iterative solvers for the discretized incompressible NavierStokes equations, Delft
University of Technology, TMI TR93-98 (1993).
Wang, T., Wang, J., and Jin, Y. AIChE J 52, 125140 (2006).
Wei, H., and Garside, J. Chem. Eng. Res. Des. 75, 219227 (1997).
Westerterp, K. R., Van Dierendonck, L. L., and De Kraa, J. A. Chem. Eng. Sci. 18, 157176 (1963).
Wilcox, D. C., Turbulence Modelling for CFD. DCW Industries Inc., La Canada (CA) (1993).
Wu, H., and Patterson, G. K. Chem. Eng. Sci. 44, 22072221 (1989).
Yeoh, S. L., Papadakis, G., and Yianneskis, M. Chem. Eng. Res. Des. 82(A7), 834848 (2004a).
Yeoh, S. L., Papadakis, G., Lee, K. C., and Yianneskis, M. Chem. Eng. Technol. 27, 257263
(2004b).
Yeoh, S. L., Papadakis, G., and Yianneskis, M. Chem. Eng. Sci. 60, 22932302 (2005).
Yianneskis, M., Popiolek, Z., and Whitelaw, J. H. J. Fluid Mech. 175, 537555 (1987).
Zwietering, Th. N. Chem. Eng. Sci. 8, 244253 (1958).
THE DETAILS OF TURBULENT MIXING PROCESS 229
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL
REACTORS
Rodney O. Fox

Department of Chemical Engineering, Iowa State University, Ames, IA 50011-2230, USA


I. Introduction 231
II. Computational Fluid Dynamics for Reacting Systems 234
A. Basic Formulation of CFD Models 234
B. Length and Time Scales in Turbulent Flows 238
C. Models for SubGrid Scale Phenomena 244
D. Reactor Systems Amenable to CFD 250
III. Mixing-Dependent, Single-Phase Reactions 252
A. AcidBase and Equilibrium Chemistry 254
B. Consecutive-Competitive and Parallel Reactions 257
C. Detailed Chemistry 267
IV. Production of Fine Particles 273
A. Mixing-Dependent Nucleation and Growth 275
B. Brownian and Shear-Induced Aggregation and
Breakage 278
C. Multivariate Population Balances 282
V. Multiphase Reacting Systems 287
A. Multiuid CFD Models 288
B. Interphase Mass/Heat-Transfer Models 296
C. Coupling with Chemistry 299
VI. Conclusions and Future Perspectives 300
Acknowledgments 302
References 302
I. Introduction
The eld of chemical reaction engineering (CRE) is intimately and uniquely
connected with the design and scale-up of chemical reacting systems. To achieve
the latter, two essential elements must be combined. First, a detailed knowledge
of the possible chemical transformations that can occur in the system is required.
This information is represented in the form of chemical kinetic schemes, kinetic
rate parameters, and thermodynamic databases. In recent years, considerable
progress has been made in this area using computational chemistry and carefully

Corresponding author. Tel: +1 515 294 9104; Fax: +1 515 294 2689. E-mail: rofox@iastate.edu
231
Advances in Chemical Engineering, vol. 31
ISSN 0065-2377
DOI 10.1016/S0065-2377(06)31004-6
Copyright r 2006 by Elsevier Inc.
All rights reserved.
controlled experiments to isolate the chemical kinetics in the absence of transport
effects. The second essential element is detailed knowledge of the transport phe-
nomena in chemical reacting systems. Indeed, because commercial chemical
reactors are almost always operated in a regime where mass and energy transport
affect or even control the product distribution from the system, understanding
the coupling between transport processes and chemical reactions is an essential
step in the design and scale-up of chemical reacting systems.
From its beginning, the holy grail of CRE has been a computational model
that is capable of predicting reactor performance based on the fundamental
molecular-scale parameters describing the chemical kinetics and the transport
coefcients. In principle, the latter can be measured experimentally in small-
scale laboratory experiments (or estimated using computational chemistry). The
chemical reaction engineer then incorporates this information into a compu-
tational model to predict the behavior of the plant-scale reactor. By avoiding
the need for pilot-scale experiments, this experiment-free scale-up approach
should result in more rapid process development at much lower cost. Admit-
tedly, while CRE has made considerable progress toward this goal, much work
remains to be accomplished. Nevertheless, due to the tremendous expansion in
computing power over the last 30 years, computational models used in CRE can
now account for much more detail than was previously thought possible. This
trend is unlikely to abate, and thus, to remain relevant to industry, chemical
reaction engineers of the future must become adept at employing detailed ow
models for chemical reacting systems.
For example, over the past 15 years computational uid dynamics (CFD) has
become an important tool in CRE for understanding the coupling between
transport processes and chemical reactions. The denition of CFD has in the
process expanded from its original emphasis on uid dynamics (i.e., momentum
transport) to include mass and energy transport, as well as detailed chemical
reactions. One might argue that it would, therefore, be more accurate to refer to
the eld as computational transport phenomena. On the other hand, because
CFD models rarely resolve all of the relevant scales (as described below), one
might also argue that computational chemical reaction engineering is a more
accurate description. However, both of these names are perhaps too broad and
lose the essential focus on the fact that CFD always includes a description of
momentum transport in the uid phase(s). Thus, the original name continues to
be used to designate all computational approaches that solve for the spatial
distribution of the velocity, concentration, and temperature elds. CFD models
have also been developed for multiphase systems, and commercial CFD codes
now offer a wide range of options for modeling chemical reactors.
Despite the many advances, the users of CFD codes must keep in mind that
the underlying transport equations are based on models, which may or may not
be valid for a particular application. This fact often escapes the minds of
newcomers to the eld who are typically overwhelmed by the numerical issues
associated with convergence, grid-independence, and post-processing. Even
RODNEY O. FOX 232
many CFD experts tend to avoid the issue by working exclusively in an
application area where acceptably accurate models are already available. Un-
fortunately (at least for industrial users!), the application of CFD to chemical
reactor analysis introduces new modeling challenges with each new reactor type.
In the simplest case of laminar ow with fully resolved concentration and tem-
perature elds, the models have a strong fundamental basis that typically
involves molecular-scale transport coefcients. These CFD models are based on
the so-called microscopic balance equations that are taught to undergraduate
students in chemical engineering courses on transport phenomena. The exten-
sion of the microscopic balance equations to multiphase ow systems is also
well understood, but brings with it a wide range of new ow phenomena in even
the simplest cases.
In reality, most of the applications in CRE for which CFD is used cannot be
treated using the microscopic balance equations alone. Instead, CFD models are
introduced to describe the effects of unresolved phenomena on the resolved
quantities. These models introduce phenomenological transport coefcients,
much like the ones developed in CRE models for chemical reactors. In fact, in
many cases, spatial transport is dominated by convection and the molecular-
scale spatial transport can be neglected. Nevertheless, just as in classical CRE
models for interphase mass/energy transport, the molecular-scale transport
coefcients appear in the dimensionless numbers used to formulate the phen-
omenological coefcients. Indeed, because the accuracy of CFD predictions are
strongly dependent on the accuracy of these so-called subgrid-scale (SGS)
models, the modeling skills developed in CRE over its long history are a crucial
component in the development of CFD for chemical reactor design and anal-
ysis. In fact, it would not be pretentious to claim that, due to their considerable
abilities to deal with the coupling between chemical reactions and transport
phenomena, chemical reaction engineers are uniquely qualied to develop the
SGS models needed for CFD modeling of chemical reacting systems.
In the remainder of this work we review the current status of CFD models for
chemical reacting systems. In some cases (e.g., single-phase reacting ow) the
current models are fully predictive in the limits of high and low Reynolds
numbers, and quite accurate in the transition region between these limits. In
other cases (e.g., multiphase reacting ow), the predictive abilities of current
CFD models are, in general, limited. Nevertheless, for particular multiphase
reactors (e.g., gassolid reacting ow), powerful models exist and are making
their way into commercial CFD codes. The goal of the presentation will not be
to describe every model in detail, but rather to indicate the current status of
models for treating reacting ows and to point out areas where further research
is needed. The reader interested in a deeper understanding of the particular
aspects a model will be pointed to the appropriate research literature for further
reading. Moreover, consistent with our desire to use CFD for chemical reactor
analysis and design, we will not discuss models whose primary purpose is to
describe nonreacting ows. For single-phase reactors, excellent descriptions of
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 233
such models are available in the literature. Likewise, CFD models for multi-
phase ows are described by other authors in this issue.
The organization of the material is as follows. In Section II we provide a
general introduction to CFD models for chemical reacting systems, and to the
critical modeling issues that arise in their development and application. In
Section III we describe the current state of the art in CFD models for single-
phase reacting ows. In Section IV we extend the discussion of single-phase
reacting ows to include systems that produce ne particles that follow the
continuous-phase ow. In these systems, the principal novelty is the inclusion of
a population balance model for the particulate phase. In Section V, we describe
the current state of the art in CFD models for multiphase reacting ows. Because
this last area is the least developed, but most rapidly advancing, we will limit our
discussion to CFD models that can potentially be used to describe plant-scale
reactors (i.e., multiuid models and related mean-eld descriptions). Even for
these models, we will not cover the details on how momentum exchange is
treated between phases. Rather, we will focus our attention on factors that affect
directly the chemical reactions. Finally, in Section VI conclusions are drawn
concerning the current status of CFD models for chemical reactor analysis, and
an attempt is made to point out the research directions where progress can be
expected in the near future.
II. Computational Fluid Dynamics for Reacting Systems
In this Section we give an overview of the formulation of CFD models for
reacting systems, with particular emphasis on systems requiring SGS models.
For the reader to understand the procedure followed to create a CFD model for
a chemical reactor, we cover rst the basic formulation. Then, because the SGS
models are often needed due to the ow being turbulent, we next review the
principal length and time scales present in turbulent transport. We then give
examples of SGS phenomena and their corresponding models in turbulent
reacting ows. Finally, we end the section with a brief discussion of the types of
reactor systems that can currently be treated using CFD.
A. BASIC FORMULATION OF CFD MODELS
When applying CFD to model a chemical reactor, we are interested in
knowing how the basic quantities (density, velocity, concentrations, etc.) vary
with the spatial location in the reactor at a given time instant. The starting point
for developing a CFD model is the microscopic balance equation, described in
detail in standard textbooks on transport phenomena (Bird et al., 2002). Letting
F denote a quantity of interest, the general form of its microscopic balance
RODNEY O. FOX 234
equation is
@F
@t
= UF = J
f
_ _
S
f
(1)
where U is the convective velocity, J
f
is a molecular-scale model for the diffu-
sive ux, and S
f
is a molecular-scale source term. Typical examples of quantities
of interest are uid density r, species mass fractions rY
a
, and the uid
momentum rU. Likewise, for multiphase systems similar quantities are of in-
terest, but for each individual phase present in the reactor. The generalized
source term S
f
will then include mass/momentum/heat-transfer between phases.
For complex uids (e.g., non-Newtonian ows), molecular-scale models for J
f
and S
f
can be quite complicated and can lead to numerical difculties, requiring
specially developed CFD solvers.
As mentioned in Section I.A, CFD codes were originally developed to solve
for the uid momentum for Newtonian uids, for which the right-hand side of
Eq. (1) is well understood (Bird et al., 2002). However, even for such uids, it is
not possible to accurately solve the microscopic balance equation for Reynolds
numbers commonly observed in chemical reactors. It is thus necessary to dis-
tinguish between direct-numerical simulations (DNS) and CFD models using
phenomenological descriptions of the turbulence. The two most widely used
CFD approaches for describing turbulent ow are large-eddy simulations (LES)
and Reynolds-averaged NavierStokes (RANS) models (Pope, 2000). In both
approaches, it is no longer possible to solve Eq. (1) directly for F due to the
enormous computational cost. Instead, Eq. (1) is ltered (LES) or ensemble-
averaged (RANS), yielding
@
~
F
@t
=
~
U
~
F =

uf = J
f
_ _

~
S
f
(2)
This transport equation cannot be solved directly because it involves several
unclosed terms. The SGS ux

uf represents the spatial transport of F by the
unresolved velocity uctuations. Models for this term can generally be written
in the form of a generalized transport equation:
@
~
F
@t
=
~
U
~
F =
~
J
f

~
J
Tf
_ _

~
S
f

~
S
Tf
(3)
where the SGS diffusive ux is denoted by
~
J
Tf
and SGS source term by
~
S
Tf
. To
distinguish this expression from Eq. (1), we will refer to Eq. (3) as the CFD
transport equation. Thus, only in the (rare) case of DNS will the CFD transport
equation correspond to the microscopic balance equation.
In chemical reacting systems, the Reynolds number of the ow is not the only
source of computational challenges. Indeed, even for laminar reacting ows the
chemical source term can be extremely stiff and tightly coupled to the diffusive
transport terms. Averaging, as done above to treat turbulent ows, does not
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 235
alleviate this difculty. Thus, turbulent reacting ows offer many difcult
challenges and require specialized models to describe the coupling between
molecular diffusion and chemical reactions (Fox, 2003). We will look at some of
the more widely applicable models in Section III.
Keeping in mind the discussion leading to Eq. (3), the formulation of a CFD
model for a chemical reactor can be broken down into the following broad
steps:
(i) First we must identify the set of state variables needed to completely de-
scribe the reactor. Typical examples are
~
F 2 ~ r; ~ r
~
U; ~ r
~
Y
a
; ~ r
~
T; ~ r
~
f
b
_
where, in addition to the quantities introduced earlier,
~
T is the uid
temperature and

f
b
is a set of scalar quantities. The latter are introduced to
dene, for example, the closure for the chemical source term (Fox, 2003)
and the turbulence model (Pope, 2000). Note that the identication of the
state variables is analogous to what is done in classical CRE models.
Thus, chemical reaction engineers are generally well acquainted with the
methodology needed to complete this step. The only new quantity that does
not appear in lumped CRE models is the uid velocity. However, chem-
ical engineers are typically introduced to momentum balances in courses on
transport phenomena, and thus understand its signicance.
(ii) The next and arguably the most difcult step is to nd closures for the CFD
transport equation, expressed in terms of the state variables. For example,
in turbulent ows the diffusive-ux terms can often be modeled successfully
as gradient-diffusion terms:
~
J
f

~
J
Tf
D
f
D
Tf
_ _
=
~
F (4)
where D
f
is a molecular-diffusion coefcient and D
Tf
is a turbulent-diffu-
sion coefcient. In high-Reynolds-number ows, D
f
is negligible compared to
D
Tf
. Note that in general D
Tf
will depend on the scalar quantities

f
b
ap-
pearing in the turbulence model. Closure of the source terms in Eq. (3) is
much more difcult, and requires fundamental knowledge about how the
local ow eld interacts with the quantity of interest (e.g., how the local
turbulence level affects the rates of diffusive mixing and chemical reactions at
the subgrid scale). Nevertheless, the nal closures must be expressed as fol-
lows in terms of the state variables:
~
S
f
~ r;
~
U;
~
Y
a
;
~
T;
~
f
b
_ _
and
~
S
Tf
~ r;
~
U;
~
Y
a
;
~
T;
~
f
b
_ _
Note that these closures describe SGS phenomena and hence are essentially
local in space (i.e., interior to a computational grid cell). For this reason, it is
RODNEY O. FOX 236
often possible to use DNS of statistically homogeneous systems (i.e., for which
the [ltered] state variables
~
F do not depend on x) to develop closure models
for
~
S
f
and
~
S
Tf
. This procedure has been widely used in single-phase tur-
bulence modeling (Pope, 2000; Fox, 2003), and more recently in multiphase
ow systems (e.g., Bunner and Tryggvason, 2003; Nguyen and Ladd, 2005).
For the latter, the generalized source terms include mass/momentum/heat-
transfer between phases, and as discussed in Section V the closure models
involve dimensionless parameters such as the particle Reynolds number.
(iii) The coupled system of CFD transport equations now appears as follows in
closed form:
@
~
F
@t
=
~
U
~
F = D
f
D
Tf
_ _
=
~
F
~
S
f

~
S
Tf
(5)
and the remaining task is to nd a suitable numerical algorithm to solve
them. Fortunately, CFD experts have developed powerful and robust al-
gorithms for solving equations in the form of Eq. (5), and these are now
available in commercial CFD codes. Thus, from the perspective of the
chemical reaction engineer working in industry, the efcient application of
CFD to chemical reactor analysis and design will inevitably involve the use
of a commercial CFD code. The next step in the CFD model formulation
will thus be to introduce the closure models developed in the previous step
into the CFD code. This is facilitated in most commercial CFD codes by
the availability of so-called user-dened scalars. In many cases, the basic
turbulence and multiphase models will already be available in a commercial
code. The chemical reaction engineer will thus only need to add the
specialized closure models (in terms of
~
f
b
) needed to describe the state
variables in a particular application.
(iv) Once the CFD model equations have been implemented in the code, the
next step is to create a computational grid to capture the specic geometry
of the chemical reactor. The qualities of the grid strongly affect the accu-
racy and the speed of convergence of the numerical algorithm. Thus, for
complex reactor geometries, it may make sense to hire a specialist in grid
generator to carry out this step.
(v) The remaining steps involve solving the CFD model and postprocessing of
the results. The latter is greatly facilitated by the built-in functions available
in most commercial CFD codes. It is at this point that reactor analysis and
design actually come into full play. By experimenting with variations in the
operating conditions and reactor geometry, the CFD model can be used to
enhance product selectivity and reactor performance.
When applying the steps outlines above, the prudent engineer will start by
modeling an existing reactor for which plant-scale data are available for val-
idation of the CFD results. If the agreement is poor, usually this will be due to
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 237
inadequate choices for the state variables and/or closure models. Nevertheless,
one should also examine the computational results to see if there are numerical
errors leading, for example, to inconsistencies in the mass, species, or energy
balances. Getting acceptable agreement may take several iterations of changes
in the closures. For many cases, this process can be facilitated by breaking it
down into independent steps (e.g., ow-eld predictions can be validated before
adding the chemistry). After reasonable agreement between the model and data
is obtained, the CFD model can be safely used to explore alternative design
scenarios.
B. LENGTH AND TIME SCALES IN TURBULENT FLOWS
As mentioned before in Eq. (3), the most common source of SGS phenomena
is turbulence due to the Reynolds number of the ow. It is thus important to
understand what the principal length and time scales in turbulent ow are, and
how they depend on Reynolds number. In a CFD code, a turbulence model will
provide the local values of the turbulent kinetic energy k and the turbulent
dissipation rate e. These quantities, combined with the kinematic viscosity of the
uid n, dene the length and time scales given in Table I. Moreover, they dene
the local turbulent Reynolds number Re
L
also given in the table.
The integral scale of a turbulent ow characterizes the largest and most
energetic ow structures. In a CFD simulation, the local grid size will be pro-
portional to the integral length scale L
u
. Likewise, the characteristic lifetime of
the largest eddies is proportional to the integral time scale t
u
. The Kolmogorov
scale characterizes the smallest ow structures and is resolved by neither LES
nor RANS simulations (only in DNS). Note that the ratios of the length and
time scales are as follows:
L
u
Z
Re
3=4
L
and
t
u
t
Z
Re
1=2
L
(6)
Thus, as the local turbulent Reynolds number increases, the separation between
the scales will increase. As a general rule, Re
L
will be proportional to the
TABLE I
THE PRINCIPAL LENGTH AND TIME SCALES, AND REYNOLDS NUMBERS CHARACTERIZING A TURBULENT FLOW
DEFINED IN TERMS OF THE TURBULENT KINETIC ENERGY k, AND TURBULENT DISSIPATION RATE , AND THE
KINEMATIC VISCOSITY m
Quantity Integral scale Kolmogorov scale
Length L
u
k
3/2
/e Z (n
3
/e)
1/4
Time t
u
k/e t
Z
(n/e)
1/2
Reynolds number Re
L
k
2
/en Re
Z
1
RODNEY O. FOX 238
macroscopic Reynolds number for the ow (i.e., Re dened in terms of a char-
acteristic ow velocity and length scale.)
In general, for a xed-ow geometry, the integral length scale will remain
approximately constant (e.g., in a turbulent jet the integral length scale is
proportional to the jet diameter). Likewise, the integral scale velocity, dened
by L
u
/t
u
, will be proportional to the characteristic velocity of the ow (e.g., the
jet velocity). Thus, as the Reynolds number increases (e.g., to enhance turbulent
mixing), Z, t
u
, and t
Z
will all decrease. In the CFD simulation, the grid will
remain approximately the same and the time step must decrease to follow t
u
.
This implies that at high Reynolds numbers less and less of the small-scale ow
structures are captured by the CFD simulation.
To estimate the amount of turbulent kinetic energy lost when ltering at a
given grid size, it is useful to introduce a normalized model energy spectrum
(Pope, 2000) as follows:
E
u
k Ck
5=3
f
L
k f
Z
k (7)
where k is the dimensionless wavenumber (inverse length), and the Kolmogorov
constant is C 1.61 (based on the most recent DNS (Watanabe and Gotoh,
2004)). The nondimensional cut-off functions are dened by
f
L
k
k
k
2
c
L

1=2
_ _
5=3p
0
(8)
and
f
Z
k exp b k
4
=Re
3
L
c
4
Z

1=4
c
Z
_ _ _ _
(9)
wherein p
0
2 and b 5.2. The parameters c
L
and c
Z
are found by applying
two integral constraints as follows:
1
_
1
0
E
u
k dk (10)
and
Re
L

_
1
0
2k
2
E
u
k dk (11)
Note that the nal form of the energy spectrum depends only on the local
turbulent Reynolds number. As an example, spectra found with different Re
L
are shown in Fig. 1.
In the normalized energy spectrum, k 1 corresponds to the inverse of the
local integral length scale and k Re
3=4
L
to the inverse of the local Kolmogorov
length scale. The range of wavenumbers in Fig. 1 over which the slope is 5/3 is
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 239
called the inertial range. Thus, for the ow to be considered turbulent (as op-
posed to transitional ow), Re
L
must be larger than approximately 20. In con-
trast, high-Reynolds-number turbulence (i.e., with a signicant inertial range)
does not begin until Re
L
is larger than 450. In RANS turbulence models de-
signed for low-Reynolds-number turbulent ows, the model parameters are
functions of Re
L
, and as the local turbulent Reynolds number approaches zero,
the microscopic balance equation (Eq. 1) is recovered. In contrast, in LES
turbulence models the lter size is typically xed at some Reynolds-number-
independent wavenumber k
c
410. Thus, the fraction of turbulent kinetic energy
captured by LES can be found from
f
c

_
k
c
0
E
u
k dk (12)
and varies from f
c
1 for small Re
L
up to a constant value less than one for
very large Re
L
(Pope, 2000).
In the discussion above, we have considered only the velocity eld in a tur-
bulent ow. What about the length and time scales for turbulent mixing of a
scalar eld? The general answer to this question is discussed in detail in Fox
(2003). Here, we will only consider the simplest case where the scalar eld f is
inert and initially nonpremixed with a scalar integral length scale L
f
that is
approximately equal to L
u
. If we denote the molecular diffusivity of the scalar
by G, we can use the kinematic viscosity to dene a dimensionless number in the
following way:
Sc
n
G
(13)
called the Schmidt number. In gases, typical values of the Schmidt number are
near unity, while in liquids values near 1,000 are quite common. The Schmidt
10
7
10
6
10
5
10
4
10
3
10
2
10
1
10
0
10
-1
10
0
10
-2
10
-4
10
-6
10
-8
10
-10
10
-12
E
u
Re
L
= 10
0
Re
L
= 10
2
Re
L
= 10
4
Re
L
= 10
6
Re
L
= 10
8

FIG. 1. The normalized model turbulent energy spectrum for a range of Reynolds numbers.
RODNEY O. FOX 240
number and the Kolmogorov length scale can be used to dene the Batchelor
length scale as follows:
l
B
Sc
1=2
Z Sc
1=2
Re
3=4
L
L
u
(14)
which is the length scale where molecular diffusion occurs. In a nonpremixed
turbulent ow seen under magnication (e.g., using planar laser-induced u-
orescence), the smallest observable structures over which concentration gradi-
ents are seen have characteristic size l
B
. Note that for large Sc, the Batchelor
scale can be very small even at low Reynolds numbers.
The degree of local mixing in a RANS simulation is measured by the scalar
variance hf
02
i, which ranges from zero for complete mixing (i.e., f hfi is
uniform at the SGS) up to f
max
hfihfi f
min
where hfi is the mean
concentration and f
max
and f
min
are the maximum and minimum values, re-
spectively. The rate of local mixing is controlled by the scalar dissipation rate e
f
(Fox, 2003). The scalar time scale analogous to the turbulence integral time
scale is (Fox, 2003) as follows:
t
f

2hf
02
i

f
(15)
In a RANS simulation of scalar mixing, a model for e
f
must be supplied to
compute hf
02
i. In fully developed turbulence, t
f
can be related to t
u
by con-
sidering the scalar energy spectrum, as rst done by Corrsin (1964).
To determine how the scalar time scale dened in Eq. (15) is related to the
turbulence integral time scale given in Table I, we can introduce a normalized
model scalar energy spectrum (Fox, 2003) as follows:
E
f
k C
OC
Re
3bk5=4
L
k
bk
f
L
kf
B
k (16)
where the scaling exponent is dened by
bk 1
2
3
7 6f
D
k
_
f
Z
k (17)
and ObukhovCorrsin constant is C
OC
0.670.68 (Sreenivasan, 1996; Watanabe,
and Gotoh, 2004; Yeung et al., 2005). In the model spectrum, the diffusive-scale
cut-off function is dened by
f
D
k 1 c
D
Sc
dk=2
k=Re
3=4
L
_ _
exp c
D
Sc
dk=2
k=Re
3=4
L
_ _
(18)
with c
D
2.59, and the diffusive exponent is
dk
1
2

1
4
f
Z
k (19)
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 241
The Batchelor-scale cut-off function is dened by
f
B
k 1 c
d
Sc
dk
k=Re
3=4
L
_ _
exp c
d
Sc
dk
k=Re
3=4
L
_ _
(20)
wherein the scalar-dissipation constant c
d
is found by applying an integral con-
straint as follows:
Re
L
Sc
_
1
0
2k
2
E
f
kdk (21)
Note that the scalar-dissipation constant computed from Eq. (21) depends only on
Re
L
and Sc.
In Fig. 2, the normalized model scalar energy spectrum is plotted for a xed
Reynolds number (Re
L
10
4
) and a range of Schmidt numbers. In Fig. 3, it is
shown for Sc 1000 and a range of Reynolds numbers. The reader interested in
the meaning of the different slopes observed in the scalar spectrum can consult
Fox (2003). By denition, the ratio of the time scales is equal to the area under
the normalized scalar energy spectrum as follows:
t
f
2t
u

_
1
0
E
f
kdk (22)
Thus, from Figs. 2 and 3 we can conclude that the time-scale ratio will depend
on Re
L
and Sc.
In the literature on turbulent mixing, the mechanical-to-scalar time-scale ratio
is dened by
R
2t
u
t
f
(23)
10
6
10
5
10
4
10
3
10
2
10
1
10
0
10
-1
10
-1
10
-3
10
-5
10
-7
10
-9
E

Sc = 0.05
Sc = 0.5
Sc = 5
Sc = 50
Sc = 500
Sc = 5000
FIG. 2. Normalized model scalar energy spectra for a range of Schmidt numbers and Re
L
10
4
.
RODNEY O. FOX 242
and is plotted based on Eq. (16) for a range of Schmidt numbers as a function of
Reynolds number in Fig. 4. Note that for very large Reynolds numbers, R is
independent of Sc and approaches a constant value of R C/C
OC
2.37, i.e.,
the ratio of the Kolmogorov and the ObukhovCorrsin constants. In contrast,
for Re
L
less than 10
6
, R is strongly dependent on both the Reynolds and
Schmidt numbers. The dependence on Reynolds number is especially signicant
for Schmidt numbers far from unity. For example, liquid-phase reactors used
for material processing (Mahajan and Kirwan, 1993, 1996; Johnson and
Prudhomme, 2003a,b) have high Schmidt numbers and operate at low to mode-
rate Reynolds numbers (Liu and Fox, 2006). In a CFD simulation, t
u
can be
found from the turbulence model and t
f
from the data in Fig. 4. Thus, the
curves in Fig. 4 dene a SGS model for t
f
parameterized in terms of Re
L
and Sc.
10
7
10
6
10
5
10
4
10
3
10
2
10
1
10
0
Sc = 0.05
Sc = 0.5
Sc = 5
Sc = 50
Sc = 500
Sc = 5000
Re
L
4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
R
FIG. 4. Mechanical-to-scalar time-scale ratio found from the normalized model scalar energy
spectra.
10
6
10
8
10
4
10
2
10
0
10
0
10
-2
10
-4
10
-6
10
-8
10
-10
10
-14
10
-12
Re
L
= 10
0
Re
L
= 10
2
Re
L
= 10
4
Re
L
= 10
6
Re
L
= 10
8

FIG. 3. Normalized model scalar energy spectra for a range of Reynolds numbers and Sc 10
3
.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 243
We will revisit this topic in Section III when discussing CFD models for mixing-
sensitive reactions. Note that while the discussion above applies to RANS tur-
bulence models, the method can be extended to LES by integrating over the SGS
wavenumbers (i.e., starting at k
c
).
In summary, we have seen that the principal length and time scales in a single-
phase turbulent ow depend on the local turbulent Reynolds number Re
L
. In a
CFD code, standard turbulence models will provide the local values of k and e.
Given the uid viscosity, it will thus be possible to compute the local turbulent
Reynolds number and related integral-scale quantities such as L
u
and t
u
. Using
the model energy spectra, we have also shown how the scalar mixing time t
f
depends on the Schmidt number and local turbulent Reynolds number. In
principle, a similar analysis could be carried out for multiphase turbulent ows
to understand the scaling laws for the length and time scales and their depend-
ence on the relevant dimensionless numbers. Unfortunately, DNS of multiphase
ow is still in its infancy and experimental measurements of energy spectra are
difcult to obtain. Nevertheless, we can expect signicant progress in our un-
derstanding of turbulent multiphase ows using DNS for particular systems
(e.g., gassolid ows) in the coming years.
C. MODELS FOR SUBGRID SCALE PHENOMENA
As noted earlier, in most applications of CFD to chemical reactor design and
analysis the CFD transport equation (Eq. 3) will require SGS closures. (Here we
use subgrid scale to refer to LES and RANS models for terms not fully resolved
by the computational grid.) Thus, one of the principal tasks of a chemical reaction
engineer is to develop the SGS models that accurately describe the chemistry and
physics occuring at the unresolved scales of the ow. As discussed above, for a
single-phase turbulent ow, the unresolved scales are those smaller than the in-
tegral length and time scales (L
u
and t
u
, respectively). Obviously, the SGS models
will be highly dependent on the type of ow under consideration (e.g., single-phase
vs. multiphase, nonreacting vs. reacting, etc.), and a complete listing of all such
models would be lengthy and uninformative to the general reader. Thus, instead of
giving general examples, in this section we will demonstrate how the SGS model is
developed for a particular example (single-phase turbulent mixing). In the sub-
sequent sections we will extend this model to reacting scalars of various types.
Readers interested in more details on the models can consult Fox (2003).
One of the rst questions that arises when considering a chemical reactor is
Can the reactor be considered perfectly mixed?. In CRE, this question implies
at least two physical situations:
1. The reactor is perfectly macromixed if the mean concentration at every point
in the reactor is equal to the volume average.
2. The reactor is perfectly micromixed if the instantaneous, local concentration
at every point in the reactor is equal to the local mean concentration.
RODNEY O. FOX 244
The classical CRE model for a perfectly macromixed reactor is the contin-
uous stirred tank reactor (CSTR). Thus, to x our ideas, let us consider a stirred
tank with two inlet streams and one outlet stream. The CFD model for this
system would compute the ow eld inside of the stirred tank given the inlet
ow velocities and concentrations, the geometry of the reactor (including bafes
and impellers), and the angular velocity of the stirrer. For liquid-phase ow with
uniform density, the CFD model for the ow eld can be developed independ-
ently from the mixing model. For simplicity, we will consider this case. Nev-
ertheless, the SGS models are easily extendable to ows with variable density.
Following the steps for formulation of a CFD model introduced earlier, we
begin by determining the set of state variables needed to describe the ow.
Because the density is constant and we are only interested in the mixing prop-
erties of the ow, we can replace the chemical species and temperature by a
single inert scalar eld x(x, t), known as the mixture fraction (Fox, 2003). If we
take x 0 everywhere in the reactor at time t 0 and set x 1 in the rst inlet
stream, then the value of x(x, t) tells us what fraction of the uid located at
point x at time t originated at the rst inlet stream. If we denote the inlet
volumetric ow rates by q
1
and q
2
, respectively, for the two inlets, at steady state
the volume-average mixture fraction in the reactor will be

x
q
1
q
1
q
2
(24)
Thus, the reactor will be perfectly mixed if and only if x

x at every spatial
location in the reactor. As noted earlier, unless we conduct a DNS, we will not
compute the instantaneous mixture fraction in the CFD simulation. Instead, if
we use a RANS model, we will compute the ensemble- or Reynolds-average
mixture fraction, denoted by hxi. Thus, the rst state variable needed to describe
macromixing in this system is x h i. If the system is perfectly macromixed, x h i

x
at every point in the reactor. The second state variable will be used to describe
the degree of local micromixing, and is the mixture-fraction variance hx
02
i.
When the variance is zero, the uid is perfectly micromixed so that x x h i. The
maximum value of the variance at any point in the reactor is x h i1 x h i, and
varies from zero in the feed streams to a maximum of 1/4 when x h i 1=2.
At this point, we should clarify an all-to-common misconception in the
chemical-engineering literature concerning the meaning of Reynolds average.
Unfortunately, many textbooks and journal articles still dene it as a time
average or a volume average over an interval that is not too long, but not too
short. This denition confuses methods for estimating the expected value from
experimental data for a single realization (i.e., time and volume averages), which
are statistics, with the underlying expected values with respect to all possible
realizations. In general, a statistic will be different every time it is computed,
while an expected value is constant at a given point in space and time. Thus,
when deriving closures for expected values such as x h i and hx
02
i, we start with a
general transport theory based on the joint probability density function (PDF)
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 245
as described in Pope (2000) and Fox (2003). Space or time averages only come
into the picture when we must validate the predictions of the CFD model
against experimental data. For example, if the ow is statistically stationary,
then the time average of x(x, t) can be used to estimate x h i(x). (Note that by
denition of statistically stationary the expected values will not depend on
time.) Likewise, if the ow is statistically homogeneous, then the volume average
of x(x, t) can be used to estimate x h i(t). In chemical reactors, the ow is almost
never homogeneous (if it were, CFD would not be needed). Nevertheless, one
still nds authors who confuse micromixing, rigorously dened in terms of the
variance hx
02
i, with deviations of the mean x h i from its volume-average.
In reality, such uctuations correspond to poor macromixing and are a
mathematical artifact caused by lumping inhomogeneous ow into a homo-
geneous model (e.g., by modeling laminar ow in a tubular reactor using a plug-
ow model). Finally, we should note that identical statistical concepts can be
used to derive CFD models for scalar mixing in low-Reynolds-number chaotic
ows encountered in high-viscosity mixing. The principal difculty in these
ows is nding general state variables to describe the length and time scales of
the ow.
The remaining state variables in our CFD model for turbulent mixing are
needed to describe the ow eld in the reactor. In a RANS model for turbulent
ow, the mean velocity U h i appears in the Reynolds-average momentum bal-
ances. The latter is closed by providing a turbulence model for the Reynolds
stresses (Pope, 2000). If a turbulent-viscosity-based model is used, two state
variables are introduced to describe the local turbulent integral time scale and
length scale (see Table I). Common choices are the turbulent kinetic energy k,
and the turbulent dissipation rate, e. The set of state variables used to described
turbulent mixing in the reactor are thus
~
F 2 U h i; k; ; x h i; hx
02
i
_
Note that when solving the CFD transport equations, the mean velocity and
turbulence state variables can be found independently from the mixture-fraction
state variables. Likewise, when validating the CFD model predictions, the ve-
locity and turbulence predictions can be measured in separate experiments (e.g.,
using particle-image velocimetry [PIV]) from the scalar eld (e.g., using planar
laser-induced uorescence [PLIF]).
Now that the state variables have been determined, we can go to steps (ii) and
(iii), which involve nding closed CFD transport equations. The derivation of
the RANS equations is described in detail in Fox (2003), and will not be re-
peated here. Instead, we will simply give the CFD transport equations and
discuss the closures appearing in the equations. The ve transport equations are
@r U h i
@t
= r U h i U h i = m m
T
_ _
= U h i =p rg (25)
RODNEY O. FOX 246
@rk
@t
= r U h ik = m m
T
=s
k
_ _
=k P
k
r (26)
@r
@t
= r U h i = m m
T
=s

_ _
=

k
C
1
P
k
C
2
r (27)
@r x h i
@t
= r U h i x h i = DD
T
= x h i (28)
and
@rhx
02
i
@t
= r U h ihx
02
i
_ _
= DD
T
=s
x
=hx
02
i
2D
T
= x h i j j
2
r
x
29
Note that although the density is constant, we have included it in the transport
equations to be consistent with the formulation used in commercial CFD codes.
The left-hand sides of Eqs. (25)(29) have the same form as Eq. (5) and
represent accumulation and convection. The terms on the right-hand side can be
divided into spatial transport due to diffusion and source terms. The diffusion
terms have a molecular component (i.e., m and D), and turbulent components.
We should note here that the turbulence models used in Eqs. (26) and (27) do not
contain corrections for low Reynolds numbers and, hence, the molecular-diffu-
sion components will be negligible when the model is applied to high-Reynolds-
number ows. The turbulent viscosity is dened using a closure such as
m
T
rC
m
k
2
= (30)
The turbulent diffusivity is dened by introducing a so-called turbulent Schmidt
number Sc
T
:
D
T
m
T
=Sc
T
(31)
which should not be confused with the molecular Schmidt number Sc. Like the
other diffusion-model constants (i.e., s
k
, s
e
, and s
x
), Sc
T
has been determined
using canonical turbulent mixing experiments (see Pope (2000) and Fox (2003)
for details). We should note, however, that these constants must sometimes
be adjusted for noncanonical ows.
The source terms on the right-hand sides of Eqs. (25)(29) are dened as
follows. In the momentum balance, g represents gravity and p is the modied
pressure. The latter is found by forcing the mean velocity eld to be solenoidal
= U h i 0. In the turbulent-kinetic-energy equation (Eq. 26), P
k
is the source
term due to mean shear and the nal term is dissipation. In the dissipation
equation (Eq. 27), the source terms are closures developed on the basis of the
form of the turbulent energy spectrum (Pope, 2000). Finally, the source terms
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 247
for the mixture-fraction variance (Eq. 29) are due to production by mean mix-
ture-fraction gradients and dissipation by micromixing. As written, Eq. (29) is
not yet closed: we need to add a model for the mixture-fraction dissipation rate
e
x
. Using Eq. (23), the latter can be modeled by

x
R

k
hx
02
i (32)
where R depends on Re
L
and Sc as shown in Fig. 4. The CFD model for
turbulent mixing is now complete, and can be solved to investigate the degree of
macro- and micromixing in a chemical reactor.
The next step in the CFD model formulation involves adding Eqs. (25)(29)
to a CFD code. For this particular example, this step is facilitated in some
commercial CFD codes that have the model already included in the standard
release of the code. The nal step is to solve the model and to compare with
experimental data when available. In this step, it may be useful to dene new
variables during postprocessing to quantify the degree of mixing, mixing zones
or the characteristic times for macro- and micromixing. See Liu and Fox (2006)
for examples of how this can be done using output from a CFD mixing model.
The CFD model developed above is an example of a moment closure.
Unfortunately, when applied to reacting scalars such as those considered in
Section III, moment closures for the chemical source term are not usually
accurate (Fox, 2003). An alternative approach that yields the same moments
can be formulated in terms of a presumed PDF method (Fox, 1998). Here we
will consider only the simplest version of a multi-environment micromixing
model. Readers interested in further details on other versions of the model can
consult Wang and Fox (2004).
The basis idea behind multi-environment models is that the mixture fraction
at any location in the reactor can be approximated by a distribution function in
the form of a sum of delta functions as follows:
f
x
z; x; t

N
n1
p
n
x; t d z x
n
x; t (33)
where p
n
is the mass fraction of environment n, and x
n
the mixture fraction in
environment n. Using the denition of mixture-fraction moments, we have
x h i

N
n1
p
n
x
n
(34)
and
hx
02
i

N
n1
p
n
x
2
n
x h i
2
(35)
RODNEY O. FOX 248
In other words, if we know p
n
and x
n
at every point in the reactor, then we can
compute up to 2N1 independent mixture-fraction moments.
The simplest model of this type is the two-environment model (N 2) for
which the independent state variables in the CFD model are
~
F 2 U h i; k; ; p
1
; x
1
; x
2
_
In theory, this model can be used to x up to three moments of the mixture
fraction e:g:; hxi; hx
2
i; and hx
3
i. In practice, we want to choose the CFD
transport equations such that the moments computed from Eqs. (34) and (35)
are exactly the same as those found by solving Eqs. (28) and (29). An elegant
mathematical procedure for forcing the moments to agree is the direct quad-
rature method of moments (DQMOM), and is described in detail in Fox (2003).
For the two-environment model, the transport equations are
@rp
1
@t
= r U h ip
1
_ _
= DD
T
=p
1
(36)
@rp
1
x
1
@t
= r U h ip
1
x
1
_ _
= DD
T
=p
1
x
1
rgp
1
p
2
x
2
x
1

D
T
x
1
x
2
p
1
=x
1

2
p
2
=x
2

2
_ _
37
and
@rp
2
x
2
@t
= r U h ip
2
x
2
_ _
= DD
T
=p
2
x
2
rgp
1
p
2
x
1
x
2

D
T
x
2
x
1
p
1
=x
1

2
p
2
=x
2

2
_ _
38
where p
2
1p
1
. Summing together Eqs. (37) and (38) and using Eq. (34), the
reader can easily show that Eq. (28) is recovered. With a little algebra (Fox,
2003), one can also show using all three equations and Eq. (34) that Eq. (29) will
be recovered if we let the micromixing parameter be
g
R
2k
(39)
Although we will not do so here, with a little more work one can use Eqs.
(36)(38) to nd the transport equation for hx
3
i. The two-environment model
thus provides an extra piece of information that can be compared to exper-
imental data.
The next step would be to implement the CFD transport equation for the state
variables in a CFD code. This is a little more difcult for the two-environment
model (due to the gradient terms on the right-hand sides of Eqs. 37 and 38)
than for the moment closure. Nevertheless, if done correctly both models will
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 249
predict exactly the same values for the mean and variance of the mixture fraction.
(See Wang and Fox (2004) and Liu and Fox (2006) for specic examples.) The
real advantage of the two-environment model comes when dealing with reacting
scalars. Unlike the moment method, multi-environment models can easily be
extended to multiple reacting scalars with virtually no changes in the model
formulation and, more importantly and surprisingly, are often nearly as accurate
as more sophisticated closures (Wang and Fox, 2004). We will look at examples
of how this is done in Sections III and IV.
In summary, we have presented two different SGS models for single-phase
turbulent mixing of an inert scalar. The goal of this presentation was not to
show the reader the specic details of how models are derived and tested, but
rather to show how a rather complicated physical process can be modeled by
adding additional scalars to a CFD model in the form of Eq. (5). Once the
equations are in this form, they can be solved in a commercial CFD code for
arbitrarily complex reactor geometries. The primary task faced when developing
a CFD model for a new reacting system is to develop closures in terms of an
appropriate set of state variables. For chemical reaction engineers, the usual
starting point will be an existing CFD model for the uid phase(s), which has
been developed and (hopefully!) validated by experts in uid dynamics. Given
such a model for momentum transport, the chemical reaction engineer can focus
on the signicant task of describing local mass/heat-transfer and chemical re-
actions. Thus, the availability of accurate models for momentum transport is
the baseline requirement when faced with a new reactor system, and essentially
determines which systems are amenable to CFD.
D. REACTOR SYSTEMS AMENABLE TO CFD
It would be difcult to construct an exhaustive list of reactor systems that can
be treated to some degree using CFD. However, we can give a partial list with a
few examples to illustrate the technical issues. First, the simplest systems to treat
with CFD are laminar-ow reactors for which the microscopic transport equation
can be solved directly (i.e., no SGS modeling is required.) For such reactors, it is
possible to use detailed chemical kinetics in complex ow geometries involving
heat and mass-transfer and catalytic surfaces. Nevertheless, even for laminar
systems computational difculties can arise, for example, when the working with
liquid systems wherein the Schmidt (Prandtl) number is much larger than unity.
For such cases the scalar eld will require a much ner grid than the velocity eld
to fully resolve all of the chemistry and physics (i.e., reaction-diffusion layers).
One might therefore consider using a micromixing model for the scalar elds to
describe the molecular mixing below the grid resolution of the velocity eld. In
principle, such a model would have the same form as those used for turbulent
reacting ows (see Section III), but with a micromixing rate (or local scalar
dissipation rate) found from the local strain-rate tensor of the velocity eld.
RODNEY O. FOX 250
Finally, we can also mention that laminar-ow systems with non-Newtonian
uids often require special numerical algorithms that are usually not available in
CFD codes designed mainly for turbulent ows.
Turbulent single-phase ow reactors can also be treated quite accurately with
current CFD technology. The key issues in this case are the SGS models and the
modeling of heat/mass transfer and reactions at ow boundaries. These issues
arise in the CFD transport equation due to the inability to resolve the smallest
scales of the ow or boundary layers. For turbulent reacting ows, it is now
possible to handle relatively complex chemistry. Nevertheless, due to the com-
putational cost, the total number of chemical species that can be transported by a
CFD code for a large computational grid is on the order of 10100. Furthermore,
due to numerical stiffness of many kinetic schemes, simply adding a large number
of scalar equations coupled through the chemical source terms leads to unreal-
istically long computing times. It is thus still very much of interest to nd smart
methods for reducing the number of transported scalars needed to describe
complex chemistry. Several useful methods have been proposed to do this (e.g.,
using the quasi steady state for free radicals (Kolhapure et al., 2005), but methods
based on tabulation in terms of a set of progress variables (e.g., Fiorina et al.,
2005) appear to be promising for complex gas-phase reaction systems. Difcult
complications arise, however, if the chemical reactor has multiple inlets or re-
cycled product streams. Such reactors cannot be classied as either nonpremixed
or premixed (which are the types that can be most easily handled using tabu-
lation), and the number of degrees of freedom in scalar phase space is large
enough that it is very difcult to determine a priori an appropriate set of progress
variables to describe the ow. It, thus, may be necessary to carry a large number
of scalar elds to describe such reactors, but one can still use tabulation schemes
(Raman et al., 2004) to handle the numerical stiffness of the chemical source term.
Multiphase reactor systems offer many challenges to CFD modelers. In terms
of complexity, uidsolid systems are more amenable to CFD modeling than
gasliquid systems. Nevertheless, progress has been made in both elds. For
uidsolid systems, we can distinguish (see Section IV) between reactor systems
with ne particles that follow exactly the uid and systems with solid-particle
velocities different than the uids velocity. In the rst case, the solids can be
treated as a scalar eld advected by the (single-phase) uid. In the second case, the
solid phase must have its own momentum equation that is coupled to the mo-
mentum equation for the uid. The momentum equation for the solid phase
requires many modeling assumptions to describe all possible ow regimes. From
the point of view of the uid dynamics, in general, dilute uidsolid systems (e.g.,
circulating uidized beds), dominated by uid-phase turbulence, are easier to deal
with than dense systems (e.g., bubbling uidized beds). However, the addition of
chemical species and reactions is challenging in both cases. In theory, a general
CFD code for uidsolid ows must account for homogeneous reactions occur-
ing in the uid (or solid) phase and heterogeneous reactions occuring at the
interface between phases. Given that the solid phase is very often a complex
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 251
porous material with microchannels that cannot be resolved by the CFD code, it
will be necessary to develop subgrid-scale models to describe mass/heat transfer
to the particle surface and through the pores. Indeed, in many cases, the solid-
phase and surface reactions will be mass-transfer limited and the overall con-
version predicted by the CFD code will be determined by the model used to
describe mass/heat transfer between the uid and solid phases. Additional com-
plications arise for uidsolid systems when the solid particles change in size due,
for example, to surface growth, breakage, or agglomeration. It is then necessary
to include a description of the particle size distribution (Fan et al., 2004). Finally,
we should note that although turbulence models have made considerable progress
for dilute gassolid systems (Minier and Peirano, 2001), the same cannot be said
for dense systems. As a result, CFD simulations of bubbling uidized beds are
usually done without turbulence models (see Section V) and require relatively ne
computational grids to capture integral-scale properties of the ow (e.g., total
pressure drop). The high computational cost of such models makes them intrac-
table for analyzing plant-scale uidized-bed reactors.
From the perspective of CFD, the most difcult reactor systems are gasliquid
and liquidliquid ows. Using the denser phase as the reference phase, such
systems range from dilute (e.g., liquid sprays) to dense (e.g., bubbly ow). From
the point of view of the uid dynamics, these systems are challenging because the
interface between the phases is deformed by (and deforms) the ow. A completely
general CFD model would need to keep track of the uid velocity in each phase
and the location and velocity of the interface. Although it is possible to use this
approach for specic model problems, it is intractable for actual reactor systems
where a less-detailed approach must be applied. For example, continuum model
can be used that describes gas and liquid as interpenetrating uids. However, it is
then necessary to introduce models for momentum, mass, and energy exchange
between the phases that describe the unresolved processes occuring at the phase
interphase. Unlike in uidsolid ows where the interfaces are rigid, in gasliquid
ows the interface can change due to the uid dynamics and chemical/physical
processes occuring at the interface. Moreover, under industrial conditions where
the volume fraction of the gas phase is often very high, turbulence and bubble
coalescence and breakage must be accounted for in the CFD model (Sanyal et al.,
2005). Unfortunately, it is very difcult to validate (and thus to improve) mul-
tiphase turbulence models using modern laser-based measurement techniques.
CFD models for gasliquid chemical reactors remain, therefore, the least devel-
oped and should be applied with caution for reactor design and analysis.
III. Mixing-Dependent, Single-Phase Reactions
CFD models for single-phase chemical reactors are by far the most advanced
and widely used in industry. The number of different chemical reactors that can be
RODNEY O. FOX 252
modeled by CFD is very broad and ranges for laminar ow reactors with detailed
gas-phase chemistry coupled with catalytic-surface chemistry to complex turbulent
hydrocarbon ames. For many of the more complicated ow congurations, spe-
cialized CFD codes have been developed to take advantage of the particular
characteristics of the ow. Thus, we will not attempt to describe the entire range of
ow phenomena that can be predicted using CFD in any detail. For turbulent
reacting ows, a recent monograph (Fox, 2003) covering specic SGS models with
abundant references to the literature is recommended as a starting point for an-
yone wanting to know more about the subject. Here, we will conne our attention
to a few specic examples of single-phase turbulent reacting ows to give the
reader a general idea of the key modeling issues in the context of CFD.
The topics in this section are arranged in the order of the computational dif-
culty faced when treating the chemical source term. For completeness, we should
note that the simplest case, which simply neglects SGS uctuations of the scalar
elds, will not be discussed for two reasons. First, its implementation in a CFD
code is trivial (at least for cases where it is accurate!) and it is the usual default
model in a commercial CFD code. Second, since it is accurate in the limit where the
reaction rates are slow compared to the ow time scales, CFD is often not required
to understand how to scale up chemical reactors with slow chemistry. The dis-
cussion here will thus proceed in the opposite direction: starting with very fast
chemistry and progressing toward so-called nite-rate chemistry. As discussed in
Fox (2003), the speed of the reactions is taken with respect to the resolved scales of
the uid ow. Thus, in a DNS, the smallest characteristic time scale is the
Kolmogorov time scale t
Z
(see Table I), and a fast reaction occurs on time scales
shorter than t
Z
. In contrast, in RANS simulations the ow time scale is given by the
local integral time scale t
u
. In comparison, for a classical CRE model the ow
time scale is the residence time, which is typically much larger than t
u
. Therefore,
we can conclude that CFD models will start to have utility for reactor analysis
whenever the reaction time scales are smaller than the residence time of the reactor.
Before looking at specic SGS models, we should highlight highly exothermic
chemical reactions (e.g., combustion). The CFD models for these systems are
complicated by the fact that the reaction rates change dramatically across the
ow domain depending on the local temperature. Thus, these systems can be-
have as not only nonreacting ows under ambient conditions but also innitely
fast reactions once ignited. For this reason, combustion models for premixed
and nonpremixed systems are usually formulated very differently (Peters, 2000;
Poinsot and Veynante, 2001; Veynante and Vervisch, 2002). In contrast, if we
consider fast, nearly isothermal reactions in the liquid phase the range of
behaviors is more limited in terms of the observed reaction rates. For example,
it does not make sense to discuss CFD models for a premixed acidbase re-
action, because once mixed the reaction occurs instantaneously. For this reason,
liquid-phase reactions that are sensitive to mixing are almost always operated
under nonpremixed conditions. We will thus limit our attention to these cases in
the following discussion.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 253
A. ACID BASE AND EQUILIBRIUM CHEMISTRY
Acidbase reactions are the archetypical instantaneous reactions. If we let A
denote the acid concentration and B the base concentration, the chemical source
term for both the acid and base can be expressed as
SA; B kAB (40)
where the rate constant k is extremely large. In essence, acid and base cannot
coexist at the same spatial location so that either A 0 when B40, or B 0
when A40. These zones with excess acid or base are separated by stoichiometric
surfaces whereon A B 0. In a CFD simulation of an acidbase reaction it
makes no sense to try to solve the problem directly using the chemical source
term. Indeed, even if k were only moderately large, including the source term
will lead (at best) to very slow convergence.
To overcome this difculty, we can introduce a new variable x dened in
terms of a linear combination of A and B such that the chemical source term for
x is null. Consider an acidbase reaction of the form
ArB !P (41)
The microscopic transport equations for A and B are, respectively,
@A
@t
U =A = G
A
=AkAB (42)
and
@B
@t
U =B = G
B
=B rkAB (43)
where G denotes the molecular-diffusion coefcient. Note that if we multiply
Eq. (42) by r and then subtract Eq. (43), we can eliminate the chemical source
term as follows:
@ rAB
@t
U =rAB r
2
G
A
rAG
B
B (44)
However, the diffusion term is now more complicated and cannot be closed
unless A and B are known separately.
As discussed for the general case in Fox (2003), to proceed further we must
assume that G
A
EG
B
so that Eq. (44) can be written as
@ rAB
@t
U = rAB Gr
2
rAB (45)
RODNEY O. FOX 254
The applicability of this approximation depends on the relative importance of
the convection and the diffusion terms, and it becomes more accurate for cases
dominated by convection (i.e., at large Reynolds numbers).
As discussed earlier, acidbase reactions are always nonpremixed. For exam-
ple, a semi-batch reactor could initially be lled with base at concentration B
0
and
acid is added with concentration A
0
. Likewise, a continuous reactor could be run
with two feed streams: one for acid and one for base. For both of these examples,
the degree of mixing between the acid stream and the rest of the reactor contents
can be quantied by introducing the mixture fraction x, which obeys
@x
@t
U =x Gr
2
x (46)
with boundary conditions x 1 in the acid inlet stream and x 0 in the base inlet
stream (and inside the reactor at t 0). For reactors with more than two inlet
streams, it is possible to dene a mixture-fraction vector n (Fox, 2003), which
obeys Eq. (46) and has N components with the properties 0 x
n
1 and
S
N
n1
x
n
1. The modeling approaches discussed below for a single mixture frac-
tion component x can thus be extended to n to treat more complex ow con-
gurations (Fox, 2003).
The mixture fraction as dened above describes turbulent mixing in the re-
actor and does not depend on the chemistry. However, by comparing Eqs. (45)
and (46), we can note that they have exactly the same form. Thus, for the
acidbase reaction, the mixture fraction is related to rAB by
x
rAB B
0
rA
0
B
0
(47)
and the stoichiometric mixture fraction is given by
x
st

B
0
rA
0
B
0
(48)
Using the fact that A and B cannot coexist at the same spatial location, we then
nd
A
x x
st
A
0
B
0
=r
_ _
if x4x
st
0 otherwise
_
(49)
and
B
0 if x4x
st
x
st
x rA
0
B
0
otherwise
_
(50)
Thus, the CFD simulation need to only treat the turbulent mixing problem for
the mixture fraction. Once x (or its statistics) are known, the acid and base
concentrations can be found from Eqs. (49) and (50), respectively.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 255
The acidbase reaction is a simple example of using the mixture fraction to
express the reactant concentrations in the limit where the chemistry is much
faster than the mixing time scales. This idea can be easily generalized to the case
of multiple fast reactions, which is known as the equilibrium-chemistry limit. If
we denote the vector of reactant concentrations by / and assume that it obeys a
transport equation of the form
@/
@t
U =/ Gr
2
/ S / (51)
then in the equilibrium limit we need to only consider the solution to a simpler
equation that includes only the chemical source term as follows:
d/
dt
S / (52)
In fact, we are only interested in the value of / for t N subject to initial
conditions that depend on the mixture fraction as follows:
/
0
x x/
1
1 x /
2
(53)
where /
1
is the reactant concentration vector in the rst inlet stream (dened
by x 1) and /
2
is the reactant concentration vector in the second inlet stream.
When formulating the equilibrium-chemistry approximation we implicitly
assume that the solution to Eq. (52) for large t depends only on the mixture
fraction, and not on the mixing history of the uid element. For some mixing-
sensitive reactions (see Section III.B below), this assumption does not hold and
the equilibrium-chemistry approximation is not applicable. These reactions are
typically irreversible and the nal product distribution depends on the mixing
path in concentration phase space traversed by the uid particle. In general, the
equilibrium-chemistry approximation should only be used for systems of fast
reversible reactions. For this case, /
N
(x) found from solving Eq. (52) will de-
pend only on x. Note that adding a reverse reaction to the acidbase reaction
(Eq. 41) discussed above will not change the basic conclusion that A and B can
be determined from x. Only the nal formulae (Eqs. 49 and 50) will change, and
these can be found using the methods described in Fox (2003).
In a turbulent ow, the local value (i.e., at a point in space) of the mixture
fraction will behave as a random variable. If we denote the probability density
function (PDF) of x by f
x
(z) where 0rzr1, the integer moments of the mixture
fraction can be found by integration:
hx
n
i
_
1
0
z
n
f
x
z dz (54)
In most applications, the moments of principal interest are the mean hxi and
variance hx
02
i hx
2
i hxi
2
. The most widely used approach for approximating
RODNEY O. FOX 256
f
x
is the presumed PDF method wherein the PDF depends only on a small set of
moments. For example, the beta PDF can be used and has the functional form
f
x
z
a b 1 !
a 1 ! b 1 !
z
a1
1 z
b1
(55)
where a and b depend on the mean and variance as follows:
a x h i
x h i 1 x h i
hx
02
i
1
_ _
and b
1 x h i
x h i
a (56)
The spatial dependencies of x h i and hx
02
i are found by solving Eqs. (28) and (29),
respectively.
A typical CFD model for acidbase and equilibrium chemistry solves Eqs.
(25)(29), and then uses Eq. (55) to approximate f
x
. Once f
x
is known, the
expected values of the reactant concentrations are computed by numerical
quadrature from the formula
h/i
_
1
0
/
1
z f
x
z dz (57)
For example, A h i and B h i can be computed using Eqs. (49) and (50), respectively.
Note that instead of Eq. (55), we could use the simpler expression for f
x
given by
Eq. (33), which avoids the need for numerical quadrature. In both cases, the mean
and variance of the mixture fraction are identical (and thus both models account
for nite-rate mixing effects.) In practical applications, the differences in the
predicted values of h/i can often be small (Wang and Fox, 2004).
B. CONSECUTIVE-COMPETITIVE AND PARALLEL REACTIONS
To go beyond the equilibrium-chemistry limit to consider cases where some of
the reaction rates are nite compared to the ow time scales, we need an efcient
method to solve for chemical species with chemical-source terms. The straight-
forward approach for doing this is to simply solve a transport equation for each
chemical species with its corresponding chemical-source term. However, it is often
the case that one or more of the reaction steps is very fast compared to the ow
time scales, leading to numerical difculties or poor convergence. An elegant
method for avoiding this problem is to rewrite the transport equations in terms of
the mixture fraction and a set of reaction-progress variables (Fox, 2003).
Some typical examples of the reactions that can be treated in this manner are
consecutivecompetitive reactions:
AB!
k
1
R
AR!
k
2
S 58
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 257
and parallel reactions as follows:
AB!
k
1
P
1
AC!
k
2
P
2
59
In most applications, the rst reaction in each set is an acidbase reaction so
that k
1
is very large. For Eq. (59), B and C are premixed and added to A under
conditions such that B is in stoichiometric excess to A. Likewise, for Eq. (58), B
is reacted in stoichiometric excess with A to produce the desired product R.
Under these conditions, the rst reaction in each set is favored. However, if
mixing occurs with the same time scale as the second reaction, the undesired by-
product (S in Eq. (58) and P
2
in Eq. (59)) will be produced. Thus, the amount of
by-product produced is a sensitive measure of the quality of mixing in the
chemical reactor.
The description of Eqs. (58) and (59) in terms of the mixture fraction and
reaction-progress variables is described in detail by Fox (2003). Here we will
consider a variation of Eq. (59) wherein the acid acts as a catalyst in the second
reaction (Baldyga et al., 1998):
AB!
k
1
P
1
AC!
k
2
AP
2
60
This parallel reaction set was used, for example, by Johnson and Prudhomme
(2003a) to investigate the quality of mixing in a conned impinging-jets reactor.
Following the steps outline in Fox (2003), the reactant concentrations in Eq.
(60) can be written in terms of the mixture fraction x and two reaction-progress
variables Y
1
and Y
2
as
c
A
A
0
1 x 1 x
s1
Y
1
(61)
c
B
B
0
x x
s1
Y
1
(62)
and
c
C
C
0
x x
s2
Y
2
(63)
where the two stoichiometric mixture fractions are
x
s1

A
0
A
0
B
0
and x
s2

A
0
A
0
C
0
(64)
and A
0
, B
0
, and C
0
are the inlet concentrations of reactants A, B, and C,
respectively. Note that in the absence of chemical reactions, the reaction-
progress variables are dened such that Y
1
Y
2
0.
RODNEY O. FOX 258
The microscopic transport equations for the reaction-progress variables can
be found from the chemical species transport equations by generalizing the
procedure used above for the acidbase reactions (Fox, 2003). If we assume that
G
A
EG
B
EG
C
, then the transport equations are given by
@Y
a
@t
U =Y
a
Gr
2
Y
a
S
a
x; Y
1
; Y
2
for a 1; 2 (65)
where the chemical-source terms are
S
1
x; Y
1
; Y
2

k
1
B
0
x
s1
c
A
c
B
B
0
x
s1
k
1
1 x
1 x
s1
Y
1
_ _
x
x
s1
Y
1
_ _
(66)
and
S
2
x; Y
1
; Y
2

k
2
C
0
x
s2
c
A
c
C
B
0
x
s1
k
2
1 x
1 x
s1
Y
1
_ _
x
x
s2
Y
2
_ _
(67)
Note that since the reaction rates must always be nonnegative, the chemically
accessible values of the reaction-progress variables will depend on the value of
the mixture fraction. We will discuss this point further by looking next at the
limiting case where the rate constant k
1
is very large and k
2
is nite.
In many applications, due to the large value of k
1
, the rst reaction is es-
sentially instantaneous compared to the characteristic ow time scales. Thus, if
the transport equation is used to solve for Y
1
, the chemical-source term S
1
will
make the CFD code converge slowly. To avoid this problem, Y
1
can be written
in terms of x by setting the corresponding reaction-rate expression (S
1
) equal to
zero as follows:
Y
11
min
x
x
s1
;
1 x
1 x
s1
_ _
(68)
It is then no longer necessary to solve a transport equation for Y
1
and the
numerical difculties associated with treating the rst reaction with a nite-rate
chemistry solver are thereby avoided.
As with the acidbase reaction, Eq. (68) implies that A and B cannot coexist
at any point in the ow. Using this innite-rate approximation, we need only
solve transport equations for x and Y
2
, where the source term for Y
2
is now
S
21
x; Y
2
B
0
x
s1
k
2
1 x
1 x
s1
Y
11
_ _
x
x
s2
Y
2
_ _
(69)
Note that S
2N
must be nonnegative, and thus the expression above only holds
for x and Y
2
values that satisfy this condition. For all other values, S
2N
is null.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 259
Applying Eq. (68), we nd that when S
2N
is nonzero, it equals
S
21
x; Y
2
A
0
k
2
1
x
x
s1
_ _
x
x
s2
Y
2
_ _
if 0 x x
s1
and 0 Y
2
x=x
s2
70
The region in xY
2
composition space where this chemical source term is non-
zero is shown in Fig. 5. Note that the maximum conversion of C occurs when
x x
s1
and corresponds to Y
2max
x
sl
/x
s2
or (using Eq. 63) to c
C
0 (i.e.,
complete conversion).
As mentioned earlier, in applications of Eq. (60) the reactor is operated with
excess B so that x
s1
o1/2. If the mixing in the reactor is good, the mixture
fraction in all uid particles at the exit of the reactor will be equal to the mean
x x h i41=2. Thus, if mixing were much faster than the characteristic reaction
time of Eq. (70) ((A
0
k
2
)
1
), then the chemical-source term in Eq. (70) would be
zero and no reaction would occur so that Y
2
0 at the exit. In contrast, any
local deviations from perfect mixing can lead to zones in the reactor where
xrx
s1
and hence to the production of the by-product Y
2
. In the opposite limit
where k
2
is large compared to the mixing rates, the maximum attainable value
for Y
2
when x
s1
rxr1 is the mixing line (Fox, 2003), dened by
Y
2mix
x Y
2 max
1 x
1 x
s1
_ _
for x
s1
x 1 (71)

0 1
0
non-zero
source term
'mixing' line

s1
Inaccessible region
Y
2max
Y
2
FIG. 5. Region in xY
2
phase space with non-zero chemical source term and the mixing line.
RODNEY O. FOX 260
and shown as a dashed line in Fig. 5. Using this expression, we nd that the
maximum attainable conversion is
X
max

x
s1
1

x

x1 x
s1

for x
s1


x 1 (72)
The accessible region in xY
2
phase space for the reaction given in Eq. (60) is
represented by the triangular region in Fig. 5 found by connecting the feed
points and the maximum conversion point. Phase-space trajectories begin at the
two feed streams [stream 1: (0; 0) and stream 2: (1; 0)]. If x
s1
is less than the
outlet value of the mixture fraction, then the amount of by-product formed is
determined by the amount of time spent in the region with nonzero source term
(t
mix
) and the characteristic time of the second reaction (t
r
). If t
r
is large com-
pared to t
mix
, then the by-product concentration will be near zero. If the inverse
is true, then the by-product concentration will be near X
max
.
The CFD model for the reaction given in Eq. (60) in the limit where the rst
reaction is very fast must account for uctuations in x and Y
2
due to turbulent
mixing. In general, this is done by solving for their joint PDF (Fox, 2003),
denoted here by f (z, y). There are several ways this can be accomplished:
1. Solve the joint PDF transport equation.
2. Assume a functional form for the joint PDF.
3. Assume a functional form for the conditional PDF of Y
2
given x and use a
presumed PDF for f
x
.
Method 1 will be discussed in Section III.C. Method 3 can be implemented in
several different forms (Baldyga, 1994; Klimenko and Bilger, 1999; Fox and
Raman, 2004), but the lowest order approximation requires a model for the
conditional expected value of Y
2
given that x z (denoted by hY
2
jzi) where
hY
2
j0i hY
2
j1i 0. By denition, hY
2
jzi will be a single-valued function of
mixture fraction and will lie in the triangular region in Fig. 5. The simplest such
model is the one proposed by Baldyga (1994), which uses a linear-interpolation
procedure to nd the conditional moment from the unconditional moment hY
2
i
(Fox, 2003). Here we will look at a multi-environment model that is based on
method 2.
The multi-environment model for the joint PDF generalizes Eq. (33) by
writing
f z; y; x; t

N
n1
p
n
x; t dz x
n
x; t dy Y
2n
x; t (73)
where Y
2n
is the value of Y
2
corresponding to environment n. Here we will
consider only the two-environment model (N 2) where the CFD models for
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 261
p
1
, x
1
and x
2
are given by Eqs. (36), (37), and (38), respectively. Similarly, the
CFD models for the reaction-progress variable in the two environments are
@rp
1
Y
21
@t
= r U h ip
1
Y
21
_ _
= DD
T
=p
1
Y
21
rp
1
S
21
x
1
; Y
21

rgp
1
p
2
Y
22
Y
21

D
T
Y
21
Y
22
p
1
j=Y
21
j
2
p
2
j=Y
22
j
2
_ _
74
and
@rp
2
Y
22
@t
= r U h ip
2
Y
22
_ _
= DD
T
=p
2
Y
22
rp
2
S
21
x
2
; Y
22

rgp
1
p
2
Y
21
Y
22

D
T
Y
22
Y
21
p
1
j=Y
21
j
2
p
2
j=Y
22
j
2
_ _
75
Except for the chemical source term, these equations have the same form as
those used for the mixture fraction. Note that the chemical source term (S
2N
) is
evaluated using the mixture fraction and reaction-progress variable in the par-
ticular environment. The average chemical source term hS
21
x; Y
2
i will thus
not be equal to S
21
hxi; hY
2
i unless micromixing occurs much faster than the
second reaction.
The CFD model for the reaction given in Eq. (60) with k
1
N has state
variables
~
F 2 U h i; k; ; p
1
; x
1
; x
2
; Y
21
; Y
22
_
By denition of the reaction-progress variables, Y
21
and Y
22
are zero for the
inlet streams, and nonnegative inside the reactor due to the chemical source
term. Once the CFD model has been solved, the reactant concentrations in each
environment n are found from
c
An
A
0
1 x
n
1 x
s1
Y
1n
(76)
c
Bn
B
0
x
n
x
s1
Y
1n
(77)
and
c
Cn
C
0
x
n
x
s2
Y
2n
(78)
where
Y
1n
min
x
n
x
s1
;
1 x
n
1 x
s1
_ _
(79)
The mean reactant concentrations are then dened by (p
2
1p
1
)
c
A
h i p
1
c
A1
p
2
c
A2
(80)
c
B
h i p
1
c
B1
p
2
c
B2
(81)
RODNEY O. FOX 262
and
c
C
h i p
1
c
C1
p
2
c
C2
(82)
The overall conversion of C, denoted by X, is computed using
X 1
c
C
h i
C
0
x h i
(83)
The value of X at the reactor outlet is a sensitive measure of the degree of mixing
in the reactor. If X51, then mixing in the reactor is rapid compared to the
second reaction in Eq. (60). In contrast, if XE1, then mixing is slow.
The CFD model described above has been used by Liu and Fox (2006) to
simulate the experiments of Johnson and Prudhomme (2003a) in a conned
impinging-jets reactor. In these experiments, two coaxial impinging jets with
equal ow rates are used to introduce the two reactant-streams. The jet Re-
ynolds number Re
j
determines the uid dynamics in the reactor. Typical CFD
results are shown in Fig. 69 for a jet Reynolds number of Re
j
400 and a
reaction time of t
r
4.8 msec. The latter is controlled by xing the inlet con-
centrations of the reactants. Further, details on the reactor geometry and the
CFD model can be found in Liu and Fox (2006).
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
1.0
0.9
0.8
0.7
0.6
0.5
0.5
0.4
0.3
0.2
0.1
0.0
p
1
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
1.0
0.9
0.8
0.7
0.6
0.5
0.5
0.4
0.3
0.2
0.1
0.0
p
2
FIG. 6. Volume fractions p
1
and p
2
in the cross-section of the conned impinging-jets reactor.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 263
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
1.00
0.91
0.82
0.73
0.64
0.55
0.45
0.36
0.27
0.18
0.09
0.00

1
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
1.00
0.91
0.82
0.73
0.64
0.55
0.45
0.36
0.27
0.18
0.09
0.00

2
FIG. 7. Mixture fractions x
1
and x
2
in the cross-section of the conned impinging-jets reactor.
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
0.67
0.61
0.55
0.49
0.43
0.37
0.30
0.24
0.18
0.12
0.06
0.00
Y
21
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
0.67
0.61
0.55
0.49
0.43
0.37
0.30
0.24
0.18
0.12
0.06
0.00
Y
22
FIG. 8. Reaction-progress variables Y
21
and Y
22
in the cross-section of the conned impinging-jets
reactor.
RODNEY O. FOX 264
In Fig. 6, the volume fractions for each environment p
1
and p
2
are shown in a
cross-section of the reactor, which includes the inlet jets and the outlet tube.
Note that because p
2
1p
1
and the inlet ow rates are equal, the contour plots
are symmetric about the vertical axis. For the same reason, p
1
p
2
1/2 on the
vertical axis. In the left-hand inlet stream p
1
1, corresponding to reactant A.
In the right-hand inlet stream p
2
1, corresponding to reactants B and C. On
the outlet cross-section mixing is nearly complete so that p
1
Ep
2
. Finally, note
that since Eq. (36) does not contain a term for micromixing, the distribution of
p
1
and its deviation from 1/2 measures the degree of macromixing at any point
in the reactor. We can therefore conclude the reactor is fairly well macromixed
everywhere except in the region near the inlet jets.
In Fig. 7, the mixture fractions in each environment x
1
and x
2
are shown. By
denition of the inlet conditions, in the inlet tubes x
1
0 and x
2
1. The var-
iations away from the inlet values represent the effect of micromixing. For ex-
ample, if we set g 0 in Eqs. (36) and (37) to eliminate micromixing, then x
1
and
x
2
would remain at their inlet values at all points in the reactor. Note that the
spatial distributions of x
1
and x
2
are antisymmetric with respect to the vertical
axis (as would be expected from the initial conditions.) In the outlet tube, x
1
and
x
2
are very near the perfectly micromixed value of 1/2. Finally, by comparing
Fig. 6 and Fig. 7, we can observe that macromixing occurs slightly faster than
micromixing in this reactor (i.e., p
n
are closer to their outlet values than are x
n
.)
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
619
563
506
450
394
338
281
225
169
113
56
0
c
C1
X(mm)
Z
(
m
m
)
-2 -1 0 1 2
-2
-1
0
1
2
3
4
5
619
563
506
450
394
338
281
225
169
113
56
0
c
C2
FIG. 9. Reactant concentrations c
C1
and c
C2
in the cross-section of the conned impinging-jets
reactor.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 265
The results shown in Figs. 6 and 7 can be combined to compute the mean
mixture fraction hxi and its variance hx
02
i from Eqs. (34) and (35), respectively.
Example plots are shown in Liu and Fox (2006) and, as expected, they agree
with the solution found by solving the moment transport equations directly
(Eqs. 28 and 29).
In Fig. 8, the reaction-progress variables in each environment Y
21
and Y
22
are
shown. By denition of the inlet conditions, in the inlet tubes Y
21
Y
22
0.
Recall that Y
2
is produced by the second (nite-rate) reaction in Eq. (60). Thus,
as observed in the plots, it is largest in zones in the reactor where A is in excess.
The largest values are found near the left wall of the reactor where the con-
vective velocity is low and A is in slight stoichiometric excess. The residence time
for uid particles in this region is relatively long compared to the reaction time.
In general, Y
21
is larger than Y
22
, which is easily explained by the fact that A
enters the reactor in environment one. Finally, note that at the reactor outlet Y
2
is not uniformly mixed across the tube. Thus, despite the high energy dissipation
in the reactor (as measured by the pressure drop), the macromixing at the outlet
is not complete (Liu and Fox, 2006).
In Fig. 9, the distribution of reactant C is shown in each environment. As c
C
is a linear combination of x and Y
2
(Eq. 78), we can distinguish features of both
Fig. 7 and Fig. 8 in the plots in Fig. 9. In particular, because C is injected in the
right-hand inlet stream, c
C2
and x
2
appear to be quite similar. Finally, as shown
in Liu and Fox (2006), the CFD predictions for the outlet conversion X are in
excellent agreement with the experimental data of Johnson and Prudhomme
(2003a). For this reactor, the local turbulent Reynolds number Re
L
is relatively
small. The good agreement with experiment is thus only possible if the effects of
the Reynolds and Schmidt numbers are accounted for using the correlation for
R shown in Fig. 4. Further details on the simulations and analysis of the CFD
results can be found in Liu and Fox (2006).
The example reactions considered in this section all have the property that
the number of reactions is less than or equal to the number of chemical species.
Thus, they are examples of so-called simple chemistry (Fox, 2003) for which
it is always possible to rewrite the transport equations in terms of the mix-
ture fraction and a set of reaction-progress variables where each reaction-
progress variablereaction-progress variable 4 depends on only one reaction.
For chemical mechanisms where the number of reactions is larger than the
number of species, it is still possible to decompose the concentration vector
into three subspaces: (i) conserved-constant scalars (whose values are null
everywhere), (ii) a mixture-fraction vector, and (iii) a reaction-progress vector.
Nevertheless, most commercial CFD codes do not use such decomposi-
tions and, instead, solve directly for the mass fractions of the chemical spe-
cies. We will thus look next at methods for treating detailed chemistry
expressed in terms of a set of elementary reaction steps, a thermodynamic
database for the species, and chemical rate expressions for each reaction step
(Fox, 2003).
RODNEY O. FOX 266
C. DETAILED CHEMISTRY
In a CFD model with detailed chemistry, the user must provide a chemical
mechanism involving K chemical species A
b
of the form (Fox, 2003)

K
b1
u
f
bi
A
b

k
f
i
k
r
i

K
b1
u
r
bi
A
b
for i 2 1; . . . ; I reactions (84)
The rate constants (k
f
i
and k
r
i
) and the stoichiometric coefcients (u
f
bi
and u
r
bi
) are
all assumed to be known. Likewise, the reaction rate functions R
i
for each
reaction step, the equation of state for the density r, the specic enthalpies for
the chemical species H
k
, as well as the expression for the specic heat of the uid
c
p
must be provided. In most commercial CFD codes, user interfaces are avail-
able to simplify the input of these data. For example, for a combusting system
with gas-phase chemistry, chemical databases such as Chemkin-II greatly sim-
plify the process of supplying the detailed chemistry to a CFD code.
The reaction rates R
i
will be functions of the state variables dening the
chemical system. While several choices are available, the most common choice
of state variables is the set of species mass fractions Y
b
and the temperature T.
In the literature on reacting ows, the set of state variables is referred to as the
composition vector /:
/ Y
1
; . . . Y
K
; T
T
where the mass fractions sum to unity: Y
1
+?+Y
K
1. The microscopic bal-
ance equation for the composition vector has the form of Eq. (1) (Bird et al.,
2002). For a turbulent reacting ow, the CFD transport equation will thus have
the form of Eq. (3) after averaging.
With detailed chemistry, the most difcult term to close in the CFD transport
equation will be the averaged chemical source term

S/: As described in detail
in Fox (2003), the chemical source term depends on the reaction rates R
i
, which
can be highly nonlinear in the composition vector /. For this reason, simple
closures that neglect correlations between components of the composition vec-
tor are usually inaccurate. Nevertheless, the default closure for detailed chem-
istry in most commercial CFD is the so-called laminar-chemistry
approximation:

S/ S
~
/. In words, this closure approximates the average
chemical source term by its value evaluated at the average composition vector.
In general, the laminar-chemistry approximation overpredicts the reaction rate
of the principal reactants, which in reality will be reduced by nite-rate tur-
bulent mixing (Fox, 2003). The simplest example is the reaction rate for the
acidbase reaction

S k

AB 0, which is null because acid and base do not


coexist. The laminar-chemistry approximation for this case is

S k
~
A
~
B which
forces either

A or

B to be null for large k. In reality, this is usually not the case in
a turbulent ow unless hx
02
i 0.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 267
To represent correlations between the components of /, we can introduce the
joint composition PDF denoted by f
/
(w), where w is the sample-space compo-
sition vector (Fox, 2003). Starting from the microscopic transport equation for /,
it is possible to derive a CFD transport equation for f
/
(w; x, t), in which the
chemical source term appears in closed form. This CFD transport equation is the
starting point for transported PDF methods. In transported PDF methods, clo-
sures are required for turbulent dispersion and molecular mixing. However, once
such closures are introduced into the CFD transport equation, we are still faced
with the computational challenge of dealing with a large number of independent
variables (w; x, t). One method for overcoming this difculty is to use
MonteCarlo simulation techniques wherein f
/
is represented by a nite number
of so-called notional particles (Fox, 2003). Typically, 50100 notional particles
are needed for each CFD grid cell to accurately capture the correlations between
the components of / and to control statistical errors. For comparison, the
laminar-chemistry approximation uses in effect one notional particle per grid cell.
The transported PDF method will therefore be more computationally demand-
ing, but represents the state of the art for treating detailed chemistry coupled to
turbulent mixing. Reader interested in further details on PDF methods and the
corresponding simulation codes are advised to consult Fox (2003). There are
many reported applications of transported PDF methods to combusting systems
(see, for example, results for a turbulent diffusion ame in Raman et al. (2004),
and Fox (2003) for other references). Applications to chemical process systems
are rarer, but some recent examples are Raman et al. (2003) and Liu et al. (2004).
The relatively high cost of transported PDF methods has led us to explore
lower-cost methods for approximating the CFD transport equation for f
/
(Wang
and Fox, 2004). In principle, transported PDF simulations can be made cheaper
by using only a small number of notional particles N. However, the random
uctuations introduced by the simulation method lead to statistical errors that
scale like 1=

N
p
: Thus, when N is small the statistical error is quite signicant.
Ideally, we would like to have an acceptably accurate method that works for N
as small as one, and whose accuracy increases for larger N in a well-dened
deterministic manner. Our method for accomplishing this task is called the direct
quadrature method of moments (DQMOM) (Fox, 2003; Wang and Fox, 2004;
Marchisio and Fox, 2005). We briey describe the resulting CFD model below.
Readers interested in more details should consult the references given above.
The application of DQMOM to the closed composition PDF transport equa-
tion is described in detail by Fox (2003). If the IEM model is used to describe
micromixing and a gradient-diffusivity model is used to describe the turbulent
uxes, the CFD model will have the form
@rp
n
@t
= r U h ip
n
_ _
= DD
T
=p
n
for n 1; . . . ; N 85
RODNEY O. FOX 268
and
@rp
n
f
an
@t
= r U h ip
n
f
an
_ _
= DD
T
=p
n
f
an
rp
n
S
a
/
n
_ _
rp
n
g hf
a
i f
an
_ _
rb

an
86
where the subscript n denotes the environment and the subscript a denotes
the component of the composition vector. Thus, p
n
is the mass fraction of
environment n, and /
n
is the composition vector in environment n. The reader
will recognize these equations as an N-environment generalization of the
two-environment model introduced earlier. Note that, as in transported
PDF simulations, the chemical source term S
a
(/
n
) appears in closed form in
Eq. (86).
By denition, the sum of the mass fractions is unity: p
1
+?+p
N
1. Thus,
one of the equations for p
n
in Eq. (85) is redundant. In Eq. (86), f
an
is one of the
K+1 components of the composition vector in environment n. The mean com-
position hf
a
i appearing in the micromixing model is dened by
hf
a
i

N
n1
p
n
f
an
(87)
Note that if the mass fractions are used to dene the composition vector, then
by denition

K
a1
f
an
1 (88)
where the sum is over all the K chemical species. This implies that the sum of
Eq. (86) over all species must yield Eq. (85), and thus that one of the chemical
species equations is redundant. Redundancy can be avoided in the CFD model
by solving Eq. (85) only for n 1,y, N1, and Eq. (86) for n 1,y, N but
with only K1 mass fractions. To avoid numerical errors, the mass fraction of
the species not solved for should be relatively large and, preferably, correspond
to a chemically inert species. The composition vector will then have K (including
temperature) components, and a total of N(K+1)1 transport equations must
be solved to represent the model in the CFD code. Alternatively, since the sum
of the chemical source term in Eq. (86) over all chemical species is null, the CFD
model can solve only Eq. (86) for a 1,y, K+1 and not use Eq. (85). This
leads to N(K+1) equations (i.e., one more than required
1
), but has the benet
that all of the transport equations have the same form.
The nal term in Eq. (86) is the correction term b

an
; which comes from
applying DQMOM to the transport equation for the composition PDF (Fox,
1
The extra equation is the mass continuity equation for r.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 269
2003). This term is computed by solving a linear system of equations dened for
each a 1,y, K+1 by

N
n1
f
m1
an
b

an

N
n1
m1p
n
f
m2
an
D
T
j=f
an
j
2
for m 1; . . . ; N 89
This expression results from forcing the transport equations for the rst N
moments of f
a
, denoted by f
m
a
_
; to agree with the composition PDF transport
equation (Fox, 2003). For example, with N 2 the linear system in Eq. (89) can
be written in matrix form as
1 1
f
a1
f
a2
_ _
b

a1
b

a2
_ _

0
p
1
D
T
=f
a1

2
p
2
D
T
=f
a2

2
_ _
(90)
Solving for b

an
yields
b

a1

D
T
f
a2
f
a1
p
1
=f
a1

2
p
2
=f
a2

2
_ _
(91)
and
b

a2

D
T
f
a2
f
a1
p
1
=f
a1

2
p
2
=f
a2

2
_ _
(92)
The reader will recognize these terms as having of the same form as the cor-
rection terms in the two-environment model discussed earlier. With N 1,
b

a1
0 and the model reduces to the laminar-chemistry approximation. With
N 2, additional information is obtained concerning the second-order mo-
ments of the composition vector. Likewise, by using a larger N, the Nth-order
moments are controlled by the DQMOM correction terms found from Eq. (89).
As noted earlier, the sum of the mass fractions is unity and thus Eq. (86) will
be consistent with Eq. (85) only if the sum of the correction term b

an
over all
chemical species a 1,y, K is null. In general, this will not be the case if Eq.
(89) is used. Another difculty that can arise is that the mass fractions in two
environments may be equal, e.g., f
a1
f
a2
, and thus the coefcient matrix in
Eq. (89) will be singular. This can occur, for example, in the equilibrium-chem-
istry limit where the compositions depend only on the mixture fraction, i.e.,
/ /
N
(x). For chemical species that are not present in the feed streams, the
equilibrium values for x 0 and x 1 are zero, but for intermediate values of
the mixture fraction, the equilibrium values are positive. This implies that the
equilibrium values will be the same for at least two values of the mixture frac-
tion in the range 0oxo1. Thus, in the equilibrium limit it is inevitable that two
environments will have equal mass fractions for certain species at some point in
the ow eld. Since singularity implies an underlying correlation between
RODNEY O. FOX 270
components of the composition vector, in most cases these difculties can be
overcome by computing the correction terms using a set of moments that is
different than hf
m
a
i (i.e., one can include cross moments hf
mp
a
f
p
b
i. For example,
one possible choice of moments that also ensures that the sum of the correction
terms is null is to use the cumulative mass fractions. We will describe this choice
next, but the reader should keep in mind that other choices may be required if
the coefcient matrix in Eq. (89) becomes singular.
Let Y
a
denote the mass fractions of the K chemical species describing the
reacting ow. By denition,

K
a1
K Y
a
1: Assuming that the chemical species
are numbered such that the major species (e.g., reactants) appear rst,
2
followed
by the minor species (e.g., products), we can dene a linear transformation by
X
b

b
a1
Y
a
for b 1; . . . ; K (93)
Note that by denition X
K
1 and thus X
b
is the cumulative mass fraction of
the rst b species. The inverse transformation corresponding to Eq. (93) is
Y
1
X
1
and Y
a
X
a
X
a1
for a ! 2 (94)
The correction terms can be computed using either Y
a
or X
a
. However, it is clear
that the correction terms will depend on which choice is used since the moments
controlled by DQMOM are different (i.e., hY
m
a
i vs. hX
m
a
i). Thus, for example,
with xed N one set of moments may lead to singular correction terms, but not
the other. In general, we can continue to solve for the mass fractions in the CFD
model, but with the correction terms computed using the cumulative mass
fractions as follows.
1. Given the mass fractions in each environment Y
an
, use Eq. (93) to compute
the cumulative mass fractions X
an
.
2. Use the integer moments of X
a
to compute the correction terms as follows:

N
n1
X
m1
an
b

an

N
n1
m1p
n
X
m2
an
D
T
rX
an
j j
2
for m 1; . . . ; N 95
3. Compute the correction terms for the mass fractions b

an
using Eq. (94) as
follows:
b

1n
b

1n
and b

an
b

an
b

a1n
for a ! 2 (96)
2
In the computer code, a sorting algorithm can be used to put the mean mass fractions hY
a
i in
descending order before dening X
b
. By keeping track of the order of the indices, one can easily
dene the inverse transformation needed to compute Y
a
from X
b
.
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 271
Note that since X
Kn
1, Eq. (95) leads to the degenerate case where
b

Kn
0, and thus the sum of the correction terms over all species is null.
Also, if one of the mass fractions is null
3
(say Y
g
0) then X
g
X
g1
and
thus the correction term for such mass fractions will be null due to Eq. (96).
As discussed in detail in Wang and Fox (2004), on the one hand, for non-
reacting systems the DQMOM approach with N environments will exactly re-
produce the moment equations for each chemical species up to order N.
However, N cannot be chosen to be too large because the coefcient matrix in
Eq. (89) can become poorly conditioned. In contrast, the moments estimated
from the transported PDF simulations will have statistical uctuations that can
only be reduced by time/ensemble averaging. In this sense, DQMOM is pref-
erable. On the other hand, for reacting systems the moments of chemical species
obey transport equations that contain unclosed averages involving the nonlinear
chemical source term. In both approaches, these averages are approximated by
f
m1
a
S
a
/
_

_
c
m1
a
S
a
w f
/
dw %

N
n1
p
n
f
m1
an
S
a
/
n
_ _
(97)
where in a transported PDF code p
n
is the particle weight and /
n
is the particle
composition vector (Fox, 2003). The integral in Eq. (97) is just the denition of
the expected value. It can be seen that the integral is replaced by a nite sum
over a set of N quadrature points. Thus, the accuracy of the moments will
depend on the degree of nonlinearity of S(/). For a weakly nonlinear source
term, a low-order quadrature method may be adequate and DQMOM would be
preferred. Such cases have been successfully investigated by Wang and Fox
(2004). In contrast, for strong nonlinearities (e.g., typical of combusting sys-
tems), a large value of N may be needed to approximate the integral in Eq. (97)
and transported PDF simulations would be preferred.
In practice, it may be difcult to determine in advance which method is best
to use for a particular application. For example, the CFD results may be more
sensitive to large-scale inhomogeneities in the ow eld than to the chemical
source term closure. A rational approach to determine whether a more detailed
SGS model is needed might be to start with N 1 (laminar-chemistry approx-
imation) and compare the predicted mean chemical species elds to the two-
environment model (N 2). If the differences are small, then the simpler model
is adequate. However, if the differences are large, then the CFD simulation can
be repeated with N 3 and the results compared to N 2. Naturally, once this
procedure has converged, it will still be necessary to validate the CFD results
with experimental data whenever possible. Indeed, it may be necessary to
3
If the mass fractions are sorted in descending order, then all of the null mass fractions will be
grouped together at the end of the array. The procedure can thus be simplied by using only the non-
zero mass fractions to dene X
b
. In practice, one can dene a lower limit for Y
a
and set the
correction term to zero for mass fractions below this limit.
RODNEY O. FOX 272
improve the turbulent transport and micromixing models before satisfactory
agreement is obtained even with the state-of-the-art transported PDF method.
IV. Production of Fine Particles
The CFD models discussed in the previous section considered mass balances
for a nite number of chemical species. In this section, we will extend these
models to include systems wherein a second phase is produced. Such systems
include aerosols (Friedlander, 2000; Wright et al., 2001), reactive precipita-
tion (Pohorecki and Baldyga, 1983; Garside and Tavare, 1985; Pohorecki
and Baldyga, 1988; Villermaux and David, 1988; Marcant and David, 1991;
Mahajan and Kirwan, 1993; David and Marcant, 1994; Seckler et al., 1995;
Mahajan and Kirwan, 1996; Aoun et al., 1999; Johnson and Prudhomme,
2003b,c), colloids (Oles, 1992; Sandku hler et al., 2003, 2005; Waldner et al.,
2005), and ame synthesis of nanoparticles (Kodas et al., 1987; Akhtar et al.,
1991; Xiong and Pratsinis, 1991; Kruis et al., 1993; Pratsinis et al., 1996; Zhu
and Pratsinis, 1997; Briesen et al., 1998; Pratsinis, 1998; Kammler and Pratsinis,
1999, 2000; Kammler et al., 2001; Mueller et al., 2004a,b; Tani et al., 2004a,b)
(to name just a few examples). We will limit the discussion to systems wherein
the particles are not too large, in other words, to particles that on average
follow the local uid velocity (Davies, 1949). Systems with larger particles will
be considered in Section V. In terms of the CFD model, the primary difference
between ne and large particles is that the former can be treated as pseudo
species by extending the mass balances and chemical kinetic expressions that we
have already considered (Piton et al., 2000; Johannessen et al., 2000, 2001). In
contrast, for larger particles, a separate momentum balance is required for each
phase (Fan et al., 2004), which signicantly complicates the solution procedure
used to solve the CFD model.
The denition of a ne particle can be made more quantitative by intro-
ducing the article Stokes number St, which measures the particle-to-uid re-
sponse-time ratio to changes in the velocity (Fuchs, 1964):
St
gr
s
d
2
s
12r
f
n
f
(98)
In this denition, r
s
and r
f
are the solid and uid densities, respectively. The
characteristic diameter of the particles is d
s
(which is used in calculating the
projected cross-sectional area of particle in the direction of the ow in the drag
law). The kinematic viscosity of the uid is n
f
and g is a characteristic strain rate
for the ow. In a turbulent ow, g can be approximated by 1/t
Z
when d
s
is
smaller than the Kolmogorov length scale Z. (Unless the turbulence is extremely
intense, this will usually be the case for ne particles.) Based on the Stokes
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 273
number, we can neglect the momentum equation for the solid phase when
Sto0.14 (Dring, 1982). Note that for systems with growing or aggregating
particles, the characteristic diameter of the largest particles should be used to
compute the Stokes number. Thus, as a general rule of thumb, in a liquid-phase
turbulent ow solid particles (with r
s
Er
f
) will follow the uid if d
s
rZ. Since
typical minimum values for the Kolmogorov length scale in practical systems
are in the range 10100 mm, nucleation and growth of nanoparticles and crys-
tals, and colloidal aggregation can all be treated as systems involving ne par-
ticles.
A fundamental modeling challenge that arises when dealing with particulate
systems is the need to describe the particle size distribution (PSD). Depending
on the application (Randolph and Larson, 1988; Ramkrishna, 2000), this is
done by dening a population balance equation (PBE) governing a number
density function (NDF) n(l; x, t). The latter is dened by an internal coor-
dinate l, which may correspond to mass, volume, or (as done here) length.
4
In
words, n(l; x; t, is the number concentration of particles with lengths in the
range [l, l+dl]) and thus it has units of (number)/(volume length).
The NDF is very similar to the PDFs introduced in the previous section to
describe turbulent reacting ows. However, the reader should not confuse them
and must keep in mind that they are introduced for very different reasons. The
NDF is in fact an extension of the nite-dimensional composition vector / to
the case of an innite number of scalars (parameterized by 0ploN). Thus,
even in the case of laminar ow where the PDFs are not needed, the NDF still
introduces an extra dimension (l) to the problem description. The choice of the
state variables in the CFD model used to solve the PBE will depend on how the
internal coordinate is discretized. Roughly speaking (see Ramkrishna (2000)
for a more complete discussion), there are two approaches that can be em-
ployed:
1. Sectional methods that represent the NDF by a histogram (Kumar and
Ramkrischna, 1996).
2. Quadrature methods that approximate integral constraints (e.g., moments) of
the NDF (McGraw, 1997).
For cases with only one internal coordinate, either approach can be imple-
mented in a CFD code (but the computational cost for the same accuracy can be
very different). However, for cases with more than one internal coordinate, only
the quadrature methods are computationally tractable on current computers.
Thus, in the examples below, we will describe only CFD models based on the
4
The preferred choice of internal coordinates is discipline dependent. Nevertheless, the conser-
vation of solid mass will imply constraints on particular moments of the PSD. In general, given the
relationship between the various choices of coordinates, it is possible (although not always practical)
to rewrite the PBE in terms of any choice of internal coordinate.
RODNEY O. FOX 274
quadrature approach. Details on particular applications of this approach can be
found in our recent publications (Marchisio et al., 2001a,b, 2003a,b; Wang and
Fox, 2003; Wang et al., 2005a).
A. MIXING-DEPENDENT NUCLEATION AND GROWTH
As a rst example of a CFD model for ne-particle production, we will
consider a turbulent reacting ow that can be described by a species concen-
tration vector c. The microscopic transport equation for the concentrations is
assumed to have the standard form as follows:
@c
@t
= Uc = G=c Sc (99)
All of the chemical species, except one, will be assumed to be completely soluble.
The one partially insoluble species will nucleate and grow a solid phase. A
typical example is A+B-P where P is a sparingly soluble compound. The rates
of nucleation J and molecular surface growth G can be functions of the local
concentration vector c, the particle size l, and the local turbulence properties.
Neglecting aggregation and breakage processes, a microscopic PBE for this
system can be written as follows:
@n
@t
= Un = G
n
=n Jl; c
@
@l
Gl; cn (100)
Note that we have used the uid velocity U to describe convection of particles,
which is valid for small Stokes number. In most practical applications, J is a
highly nonlinear function of c. Thus, in a turbulent ow the average nucleation
rate will depend strongly on the local micromixing conditions. In contrast, the
growth rate G is often weakly nonlinear and therefore less inuenced by tur-
bulent mixing.
The quadrature approach for treating Eq. (100) introduces a moment trans-
formation dened by
m
k

_
1
0
l
k
nldl for k 0; 1; . . . ; 1 (101)
Applying this transformation to Eq. (100) yields
@m
k
@t
= Um
k
= G=m
k
J
k
c kG
k1
c (102)
where the moments of the nucleation function are dened by
J
k
c
_
1
0
l
k
Jl; c dl (103)
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 275
and the growth rates for the moments are dened by
G
k
c
_
1
0
l
k
Gl; cnl dl (104)
Note that in the special case of size-independent growth, this term can be ex-
pressed as a closed function of the moments, i.e., G
k
c Gcm
k
. Note also
that when deriving Eq. (102) we have neglected the size-dependence of G
n
. This
is justied in turbulent ows and, in any case, to do otherwise would require a
micromixing model that accounts for differential diffusion (Fox, 2003).
In principle, any functional form could be used for the nucleation rate.
However, to simplify the discussion, we will assume that only particles of zero
size are formed by nucleation so that Eq. (103) becomes
J
k
c d
0;k
Jc (105)
where d
j
,
k
is the Kronecker delta and J(c) contains the dependence of the nu-
cleation rate on the local composition vector. Note that under this assumption
nucleation only appears in the equation for the moment of order k 0.
Due to the presence of the unknown NDF inside the integral in Eq. (104), the
growth term is not closed (i.e., it cannot be computed exactly in terms of the
moments unless G is independent of l). In the quadrature method of moments
(QMOM), the integral is approximated by a sum over a set of M weights (w
m
)
and M abscissas (l
m
):
G
k
c

M
m1
w
m
l
k
m
Gl
m
; c (106)
The weights and abscissas are determined in QMOM by forcing them to agree
with the quadrature approximation of the rst 2 M moments:
m
k

M
m1
w
m
l
k
m
for k 0; 1; . . . ; 2M 1 (107)
This system of 2 M nonlinear equations is ill-conditioned for large M, but can be
efciently solved using the product-difference (PD) algorithm introduced by
McGraw (1997). Thus, given the set of 2 M moments on the left-hand side of
Eq. (107), the PD algorithm returns w
m
and l
m
for m 1,y, M. The closed
microscopic transport equation for the moments can then be written for
k 0,y, 2 M1 as
@m
k
@t
= Um
k
= G=m
k
d
0;k
Jc k

M
m1
w
m
l
k1
m
Gl
m
; c (108)
RODNEY O. FOX 276
To proceed further, this expression and Eq. (99) must be averaged to nd the
CFD transport equations for the species concentrations and the moments.
By dening the composition vector to include the species concentrations and
the moments as follows:
/ c
1
; . . . ; c
K
; m
0
; . . . ; m
2M1

T
we can observe that the microscopic transport equations have the same form as
those used for detailed chemistry in Section III.C. Thus, any turbulent reacting
ow model that can be used for detailed chemistry can also be used as a CFD
model for ne-particle production. For example, using the DQMOM approach
to treat the composition PDF leads to Eqs. (85) and (86) with source terms for
the moments given by the last two terms in Eq. (108) as follows:
S
k
/ d
0;k
Jc k

M
m1
w
m
l
k1
m
Gl
m
; c (109)
Note that each environment in the micromixing model will have its own set of
concentrations c
an
and moments m
kn
, reecting the fact that the PSD is coupled
to the chemistry and will thus be different at every SGS point in the ow. The
PD algorithm is applied separately in each environment to compute the weights
(w
mn
) and abscissa (l
mn
) from the quadrature formula as follows:
m
kn

M
m1
w
mn
l
k
mn
for k 0; 1; . . . ; 2M 1 (110)
Thus, the source terms for each environment S(c) and S
k
(/) will be closed. Of
particular interest are the local nucleation rates J(c
n
). As discussed in Wang and
Fox (2004), due to poor micromixing the local nucleation rates can be much
larger than those predicted by the average concentrations J(hci). This results in a
rapid increase in the local particle number density m
0n
due to the creation of a
very large number of nuclei. As discussed below, this will have signicant con-
sequences on the local rate of aggregation.
The CFD model for nucleation and growth can now be solved to determine
the average species concentrations hci and the average moments of the NDF
hm
k
i. However, to properly interpret the computational results, care must be
taken in dening the averages for terms involving the moments. Starting from
the denition of the moments from the microscopic NDF (Eq. 101), the average
moments are dened by
m
k
h i
_
1
0
l
k
hnli dl (111)
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 277
where hnli is the Reynolds-average NDF. For the multi-environment model,
the Reynolds-average quantities are dened by
m
k
h i

N
n1
p
n
m
kn
and hnli

N
n1
p
n
n
n
l (112)
where n
n
(l) is the NDF in environment n. Using the denition of the weights
and abscissas (Eq. 110), the average moments can be expressed as follows:
m
k
h i

N
n1
p
n

M
m1
w
mn
l
k
mn

M
m1
hw
m
l
k
m
ia

M
m1
w
m
h i l
m
h i
k
(113)
Thus the weights and abscissas for the average NDF cannot be found to be
averaging those for the NDF in each environment. Due to the nonlinear re-
lationship between the moments and weights and abscissas, this result is not
surprising.
5
However, it does illustrate that hw
m
i and hl
m
i are not the relevant
quantities needed to reconstruct hm
k
i.
B. BROWNIAN AND SHEAR-INDUCED AGGREGATION AND BREAKAGE
As mentioned above, when local nucleation rates are high, the local number
density of particles will be large and aggregation will be favored. For very small
particles (e.g., submicron), the dominant aggregation mechanism is Brownian
motion (Einstein, 1905). Physically, the particles move by random walks driven
by momentum exchange with solvent molecules. When two particles collide,
they will stick together with a probability p
B
to form a doublet. By this mech-
anism, larger clusters are eventually formed with a fractal dimension d
f
(Meakin, 1988; Sorensen, 2001; Lattuada et al., 2003a,b). If the volume fraction
of particles is high enough, the clusters can reach innite size in a nite time
5
If we interpret the weights and abscissas as a delta-function representation of the NDF:
n
n
l

M
m1
w
mn
dl l
mn

then the average NDF is represented by N+M delta functions:


hn l i

N
n1

M
m1
p
n
w
mn
dl l
mn

By using the PD algorithm, this set of NM delta functions is reduced to a set of only M delta
functions, but with the same values for the rst 2 M moments hm
k
i as the original set. It should be
obvious that this cannot be accomplished by simply averaging the weights and abscissas.
RODNEY O. FOX 278
span through a process called gelation (Sandku hler et al., 2005). However, if the
clusters are growing in a turbulent ow eld, once their characteristic size is
greater than approximately one micron other shear-driven processes become
dominant (Oles, 1992). These are typically classied as shear-induced growth,
breakage and restructuring. As in the case of chemical kinetics, aggregation and
breakage models are required to describe these phenomena. From the perspec-
tive of developing CFD models, we can assume that such kinetic expressions
are available. Thus, our focus here will be on how to implement the PBE in a
CFD simulation to predict the moments of the PSD.
The simplest aggregation and breakage models can be formulated in terms of
the NDF n(u), which uses volume as the independent variable.
6
The microscopic
transport equation for the NDF has the form (Wang et al., 2005a,b)
@n
@t
= Un = G
n
=n Ju; c
@
@u
Gu; cn
Au Bu 114
where A(u) and B(u) are the aggregation and breakage terms, respectively.
Although we do not do so here, these terms can be assumed to be dependent on
the species concentrations c without changing the form of the CFD model. For
binary aggregation and breakage, the aggregation term can be expressed as
follows (Ramkrishna, 2000):
Au
1
2
_
u
0
bu s; snu sns ds
nu
_
1
0
bu; snsds 115
where b is the aggregation kernel. A typical breakage term has the form
Bu
_
1
u
bujsasnsds aunu (116)
where b is the daughter-size distribution and a is the breakage kernel.
For Brownian aggregation, the aggregation kernel can be written as follows
(Elimelech et al., 1995; Sandku hler et al., 2003):
bu; s
2p
B
k
B
T
3m
u
1=d
f
s
1=d
f
_ _
u
1=d
f
s
1=d
f
_ _
(117)
6
It is also possible to use length as the independent variable as described in Wang et al. (2005b).
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 279
while for shear-induced aggregation it has the form (Oles, 1992; Elimelech et al.,
1995) as follows:
bu; s gau; s u
1=d
f
s
1=d
f
_ _
3
(118)
A general expression can be found by combining these two cases (Melis et al.,
1999). In these expressions, k
B
is the Boltzmann constant, T is the uid tem-
perature (Kelvin), m is the uid viscosity, g is the local shear rate, and a is an
efciency factor. For shear-induced breakage, the kernel is usually t to exper-
imental data (Wang et al., 2005a,b). A typical form is (Pandya and Spielman,
1983) as follows:
au c
1
g
c
2
u
c
3
=d
f
(119)
where c
1
c
3
are empirical tting parameters. A typical daughter-size distribu-
tion is
bujs du fs du 1 f s (120)
where f 1/2 corresponds to equal-size daughters and f51 corresponds to
erosion (i.e., a very small and a very large daughter).
At this point, we should step back and make a few comments concerning
Eq. (114). First, it should be obvious to the reader that many modeling assump-
tions have already been invoked to arrive at forms for the aggregation and
breakage terms and the rate functions needed to dene them. Therefore, we should
keep in mind that, just as when working with chemical kinetics, the CFD pre-
dictions can be no better than the basic physical/chemical models used to describe
the source terms. By their nature, aggregation and breakage terms are much more
difcult to formulate accurately than gas-phase chemical mechanisms. Thus, we
should expect a greater degree of mismatch with experimental data for CFD
solutions for aggregation-breakage systems than we are accustomed to with, for
example, combustion systems (Raman et al., 2004). Due to the uncertainties in the
kinetics for aggregation and (especially) breakage, the use of a highly sophis-
ticated SGS model for turbulent mixing (e.g., transported PDF or CMC models)
is likely unwarranted for most systems of interest. In fact, at present the greatest
need in this domain is carefully designed experiments to accurately measure the
rate constants appearing in the aggregation and breakage expressions.
A second general observation can be made by comparing the aggregation
terms (Eq. 115) to the breakage terms (Eq. 116): The former is second order in
n(u) and the latter is rst order. This implies that when n(u) is very large (e.g.,
due to high local nucleation), aggregation will be favored. In fact, in shear-
dominated systems gelation is prevented only by virtue of the fact that the
aggregation efciency a drops off rapidly for large clusters. Thus, in systems
RODNEY O. FOX 280
with nucleation, growth, aggregation, and breakage (e.g., aggregating nano-
particles), the PSD can be nonzero over a very wide range of cluster volumes
(i.e., 34 orders of magnitude is not uncommon). If sectional methods are used
to approximate n(u), such systems typically require a relatively large number of
sections (e.g., 25100) for reasonable accuracy (Marchisio et al., 2003b). For
this reason, when used in a CFD model, sectional methods require an unfavor-
able trade-off between reasonable computational cost and accuracy. Quad-
rature methods, which have lower cost for equivalent accuracy (Marchisio et al.,
2003b), are thus a better choice for combining with CFD to describe these
complex systems.
A nal observation concerns the shear rate g appearing in both the aggre-
gation and breakage kernels. Since from the outset we have assumed that the
clusters are small compared to the Kolmogorov length scale, the local shear rate
seen by a cluster is the instantaneous shear (e/u)
1/2
, where e is the uctuating
dissipation rate (Fox, 2003). The uctuating dissipation is very different than the
average dissipation e computed by the turbulence model (Fox and Yeung, 2003).
On average hi , but e(t) uctuates strongly on a characteristic time scale
proportional to the Kolmogorov time scale t
Z
. Whether these uctuations must
be accounted for in the CFD model depends on the characteristic aggregation
time. Marchisio et al. (2006) estimate that when the local solid volume fraction
exceeds 10
3
, the uctuations must be included in the CFD model. Otherwise, g
can be set equal to (e/u)
1/2
. In any case, we see from this discussion that g scales
with the local turbulent Reynolds number like g$Re
L
1/2
. Thus, shear-induced
aggregation and breakage will be important phenomena in turbulent ows and
their importance will increase with increasing Reynolds number.
We now turn to the question of developing a CFD model for ne-particle
production that includes nucleation, growth, aggregation, and breakage.
Applying QMOM to Eq. (114) leads to a closed set of moment equations as
follows:
@m
k
@t
= Um
k
= G=m
k
S
k
/ (121)
where the moment source term is given by (Marchisio et al., 2003a)
S
k
/ d
0;k
Jc k

M
m1
w
m
u
k1
m
Gu
m
; c

M
m1
w
m
au
m
b
k
m
u
k
m
_ _

1
2

M
m1

M
p1
w
m
w
p
u
m
u
p

k
u
k
m
u
k
p
_ _
bu
m
; u
p
122
and
b
k
m

_
1
0
u
k
buju
m
du (123)
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 281
Thus, just as we saw with Eq. (109), the moment source term has the form found
with detailed chemistry (i.e., the right-hand side of Eq. (122) depends only on
/).
7
The CFD transport equation can therefore be developed along the same
lines that we discussed earlier for nucleation and growth. In other words, the
DQMOM model can be used to describe micromixing of the moments of the
PSD at the sub-grid scale, along with turbulent transport models to describe
macromixing. One new factor that can arise when dealing with aggregation is
that the moment source terms can be stiff. To handle this problem in CFD
simulations, Wang and Fox (2003) successfully implemented a tabulation
method originally designed for combustion chemistry.
C. MULTIVARIATE POPULATION BALANCES
The CFD model described above is adequate for particle clusters with a
constant fractal dimension. In most systems with uid ow, clusters exposed to
shear will restructure without changing their mass (or volume). Typically re-
structuring will reduce the surface area of the cluster and the fractal dimension
will grow toward d
f
3, corresponding to a sphere. To describe restructuring,
the NDF must be extended to (at least) two internal coordinates (Selomulya
et al., 2003; Zucca et al., 2006). For example, the joint surface, volume NDF can
be denoted by n(s, u; x, t) and obeys a bivariate PBE.
With two (or more) internal coordinates, numerical approaches for the PBE
using sectional methods become intractable in the context of CFD. A practical
alternative is to use a nite number of samples to approximate the NDF in
terms of delta functions as follows:
ns; u

M
m1
w
m
ds s
m
du u
m
(124)
The weights w
m
and abscissas s
m
and u
m
are related to the bivariate moments by
m
k
1
k
2

__
s
k
1
u
k
2
ns; udsdu

M
m1
w
m
s
k
1
m
u
k
2
m
(125)
Thus, it would be natural to attempt to extend the QMOM approach to handle a
bivariate NDF. Unfortunately, the PD algorithm needed to solve the weights and
abscissas given the moments cannot be extended to more than one variable. Other
methods for inverting Eq. (125) such as nonlinear equation solvers can be used
(Wright et al., 2001; Rosner and Pykkonen, 2002), but in practice are compu-
tationally expensive and can suffer from problems due to ill-conditioning.
7
There are, however, important differences. For example, in detailed chemistry the source terms
do not depend on the ow quantities. In contrast, all of the rate functions for particulate systems can
potentially depend on the local ow quantities such as the instantaneous shear rate.
RODNEY O. FOX 282
To overcome the difculty of inverting the moment equations, Marchisio and
Fox (2005) introduced the direct quadrature method of moments (DQMOM).
With this approach, transport equations are derived for the weights and ab-
scissas directly, thereby avoiding the need to invert the moment equations dur-
ing the course of the CFD simulation. As shown in Marchisio and Fox (2005),
the NDF for one variable with moment equations given by Eq. (121) yields two
microscopic transport equations of the form
@w
m
@t
= Uw
m
= G=w
m
a
m
a

m
(126)
and
@w
m
u
m
@t
= Uw
m
u
m
= G=w
m
u
m
b
m
b

m
(127)
where the source terms a
m
and b
m
are found from the moment source terms S
k
,
and a

m
and b

m
are correction terms that depend on G9ru
m
9
2
. The role of the
correction terms is to ensure that Eqs. (126) and (127) are mathematically
equivalent to Eq. (121), and they result from the diffusion terms during the
nonlinear change of variables (Eq. 107). Thus, they will be null in the absence of
molecular diffusion and their forms do not depend on the moment source terms.
The extension of DQMOM to bivariate systems is straightforward and, for
the surface, volume NDF, simply adds another microscopic transport equation
as follows:
@w
m
s
m
@t
= Uw
m
s
m
= G=w
m
s
m
c
m
c

m
(128)
Example calculations for a bivariate system can be found in Marchisio and Fox
(2006) and Zucca et al. (2006). We should note that for multivariate systems the
choice of the moments used to compute the source terms is more problematic
than in the univariate case. For example, in the bivariate case a total of 3 M
moments must be chosen to determine a
m
, b
m
and c
m
. In most applications,
acceptable accuracy can be obtained with 3rMr5. Thus, the maximum
number of moments that will be required is 15, and one might decide to use
bivariate moments up to order four as shown below:
m
0,0
,
m
1,0
, m
0,1
,
m
2,0
, m
1,1
, m
0,2
,
m
3,0
, m
2,1
, m
1,2
, m
0,3
,
m
4,0
, m
3,1
, m
2,2
, m
1,3
, m
0,4
.
However, there is no guarantee that this set of moments will be linearly inde-
pendent. Even if they do work, we can note that compared to the univariate case
where the ten moments m
0
m
9
would be employed, the accuracy (as measured
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 283
by the highest-order moment used) decreases as the number of internal coor-
dinates increases. Thus, the only way to obtain equivalent accuracy would be to
increase M.
Leaving aside the question of determining which moments to use, we also
need a consistent method for deriving a CFD transport equation for turbulent
reacting ows from the microscopic transport equations for the weights and
abscissas. In particular, since the moments have a nonlinear dependence on the
weights and abscissas, the denition of the micromixing model in terms of
the weights and abscissas must be consistent with that used for the moments.
8
The development of a consistent model is most easily done by proceeding in
three independent steps as follows:
1. Find the multi-environment model for the case where the moment source terms
are null and there is no micromixing: a
m
a

m
b
m
b

m
c
m
c

m
0:
2. Extend the model to include micromixing, but no moment source terms:
a
m
b
m
c
m
0.
3. Extend the model to include the moment source terms due to nucleation,
growth, etc.
We will now briey illustrate how these steps are carried out.
The multi-environment model for turbulent transport of the bi-variate mo-
ments in the absence of moment source terms has the form
@p
n
@t
= Up
n
= G
T
=p
n
(129)
and, for arbitrary values of k
1
and k
2
@p
n
m
k
1
k
2
n
@t
= Up
n
m
k
1
k
2
n
= G
T
=p
n
m
k
1
k
2
n
b

k
1
k
2
n
(130)
where the bi-variate moments in environment n are given in terms of the
weights and abscissas by
m
k
1
k
2
n

M
m1
w
mn
v
k
1
mn
s
k
2
mn
(131)
As discussed earlier with other transported scalars, the correction term appearing
on the right-hand side of Eqs. (130) is found by solving a linear system dened by

N
n1
m
k1
k
1
k
2
n
b
n
k
1
k
2
n

N
n1
k 1p
n
m
k2
k
1
k
2
n
G
T
=m
k
1
k
2
n

2
for k 1; . . . ; N (132)
8
We develop the CFD equations using the DQMOM model for micromixing. Nevertheless, care
must also be taken when using other micromixing models, including transported PDF methods.
RODNEY O. FOX 284
Note that in this expression k is unrelated to k
1
and k
2
, which are determined by
the choice of moment m
k
1
k
2
n
. Equation (132) denes the correction term b
n
k
1
k
2
n
for a
xed set of indices (k
1
, k
2
, n). In other words, given the weights and abscissas we
can use Eq. (131) to compute the bi-variate moments, which can then be used in
Eq. (132) to solve for b
n
k
1
k
2
n
.
The multi-environment model for the weights and abscissas has the form
@p
n
w
mn
@t
= Up
n
w
mn
= G
T
=p
n
w
mn
A
mn
A

mn
(133)
@p
n
w
mn
u
mn
@t
= Up
n
w
mn
u
mn
= G
T
=p
n
w
mn
u
mn

B
mn
B

mn
134
and
@p
n
w
mn
s
mn
@t
= Up
n
w
mn
s
mn
= G
T
=p
n
w
mn
s
mn

C
mn
C

mn
135
where the nal two terms on the right-hand sides are null in the absence of
micromixing and moment source terms. The correction terms
A
n
mn
; B
n
mn
and C
n
mn
are determined such that they are equivalent to the cor-
rection term b
n
k
1
k
2
n
in Eq. (130). In other words, if we use Eq. (131) to replace
m
k
1
k
2
n
in Eq. (130), then Eqs. (133)(135) will be equivalent to Eq. (130) when
the correction terms satisfy the following linear system:
1 k
1
k
2

M
m1
u
k
1
mn
s
k
2
mn
A
n
mn
k
1

M
m1
u
k
1
1
mn
s
k
2
mn
B
n
mn
k
2

M
m1
u
k
1
mn
s
k
2
1
mn
C
n
mn
b
n
k
1
k
2
n
k
1
k
1
1p
n

M
m1
w
mn
v
k
1
2
mn
s
k
2
mn
G
T
=v
mn
j j
2
k
1
k
2
p
n

M
m1
w
mn
v
k
1
1
mn
s
k
2
1
mn
G
T
=v
mn
=s
mn

2
k
2
k
2
1p
n

M
m1
w
mn
v
k
1
mn
s
k
2
2
mn
G
T
=s
mn
j j
2
136
Note that the correction terms are proportional to G
T
and result from
turbulent velocity uctuations (represented by a gradient-diffusion model). For
the multi-environment model the composition vector is dened by
/ w
1
; . . . ; w
M
; w
1
v
1
; . . . ; w
M
v
M
; w
1
s
1
; . . . ; w
M
s
M

T
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 285
Thus, a total of 3 M variables are needed to describe the evolution of the
bi-variate NDF in a turbulent ow. The rst step in the construction of the
CFD model is now complete.
In the second step we must add the micromixing terms from the DQMOM
model to Eqs. (133)(135). However, as we discussed earlier, we need to keep in
mind that micromixing conserves the moments of the NDF, and not the weights
and abscissas (see Eq. 113). The micromixing model in environment n for the bi-
variate moments has the form
M
k
1
k
2
n
g hm
k
1
k
2
i m
k
1
k
2
n
_ _
(137)
where the moments in environment n are given by Eq. (131) and the average
moments are found by averaging over all N environments:
hm
k
1
k
2
i

N
n1
p
n
m
k
1
k
2
n
(138)
Thus, given the weights and abscissas, the micromixing term for the moments is
closed. Applying DQMOM, the micromixing source terms (which are added to
the right-hand sides of Eqs. (133)(135)) can be shown to obey for each
n 1,y, N the linear system dened by
1 k
1
k
2

M
m1
u
k
1
mn
s
k
2
mn
A
mn
k
1

M
m1
u
k
1
1
mn
s
k
2
mn
B
mn
k
2

M
m1
u
k
1
mn
s
k
2
1
mn
C
mn
p
n
M
k
1
k
2
n
139
Note that the right-hand side of this expression contains the closed micromixing
term for the moments (Eq. 137). To nd the 3 M micromixing source terms
(A
mn
, B
mn
, C
mn
) from this expression, we must choose a set of 3 M bi-variate
moments. Note that because the moment equations are closed when only
micromixing is considered, the chosen moments will be reproduced exactly. A
convenient choice is to use the uncoupled moments m
k0
and m
0k
. (Note that this
same choice should then be used in Eq. (136).) This yields the linear system
1 k

M
m1
u
k
mn
A
mn
k

M
m1
u
k1
mn
B
mn
p
n
M
k0n
for k 0; . . . ; 2M 1 140
which can be solved to nd A
mn
and B
mn
for each n 1,y, N, and the linear
system
k

M
m1
s
k1
mn
C
mn
p
n
M
0kn
k 1

M
m1
s
k
mn
A
mn
for k 1; . . . ; M 141
RODNEY O. FOX 286
which can be solved to nd C
mn
. The second step in the construction of the CFD
model is now complete.
The third step is to add the moment source terms due to nucleation, growth,
aggregation, and restructuring. The exact form of these terms will depend on the
models used to describe these processes (see for example Eq. 122). However, if
we denote the source term for the bi-variate moments as S
k
1
k
2
/, where / is the
composition vector including all variables needed to describe the kinetics, then
the source terms (including micromixing) in the CFD model (A
mn
, B
mn
, C
mn
) can
be found by simply adding p
n
S
k
1
k
2
/
n
to the right-hand side of Eq. (139). The
extension of DQMOM to multi-variate systems with more than two internal
coordinates is discussed in Marchisio and Fox (2005). As far as the CFD model
equations are concerned, no additional complications arise for higher-
dimensional systems. Nevertheless, from a practical point of view the reader
should keep in mind that going to higher dimensions will necessarily diminish
the accuracy of the method as compared to the uni-variate case.
We end here with our discussion of CFD models for ne-particle production.
The reader hopefully has a good feel for the issues involved and at least a
cursory understanding of the available models. The current status of the eld is
such that the CFD tools available for the analysis of ne-particle systems are
adequate for most applications. Currently, the weakest link in the modeling
process is description of physical processes such as aggregation and breakage,
and their coupling to the ow eld. As mentioned earlier, there is a need for
carefully designed experiments with local measurements of the PSD and
turbulence elds that can be used for CFD validation. As in the eld of
turbulent mixing, the ideal experiment would be spatially homogeneous, or a
most one-dimensional, to focus on the source terms for the chemical reactions
and the moments. It would also be useful to conduct a few idealized
experiments using direct numerical simulations (DNS) for homogeneous ne-
particle systems undergoing nucleation, growth, aggregation, and breakage.
Although such experiments cannot be directly compared to real systems, they
would still be useful for calibrating CFD models used for micromixing and
shear-induced aggregation and breakage.
V. Multiphase Reacting Systems
The CFD models considered up to this point are, as far as the momentum
equation is concerned, designed for single-phase ows. In practice, many of the
chemical reactors used in industry are truly multiphase, and must be described
in the context of CFD by multiple momentum equations. There are, in fact,
several levels of description that might be attempted. At the most detailed level,
direct numerical simulation of the transport equations for all phases with fully
resolved interfaces between phases is possible for only the simplest systems. For
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 287
example, for uid-particle systems the NavierStokes equation must be solved
for the uid phase between the particles with enough detail to capture
momentum transport at the particle surfaces (Nguyen and Ladd, 2005). At the
same time, Newtonian equations for the particle positions and momenta must be
solved simultaneously to account for uid surface forces and particleparticle
collisions. Obviously, such a detailed model could not be used to describe a large
chemical reactor such as a uidized bed. Less costly methods have thus
been developed and are described in other chapters of this issue. In general, the
methods available in commercial CFD codes are based on the so-called
multiuid model (Drew and Passman, 1999) that makes no attempt to capture
the details at the interfaces between phases. Instead, the uid at the subgrid scale
is described by the volume or mass fractions for each phase much in the same
way that environments are used to describe micromixing in single-phase ows).
We will thus look briey at the structure of multiuid models and describe some
of the modeling assumptions that are required for multiphase reacting ows.
A. MULTIFLUID CFD MODELS
In this section, we will look briey at multiuid CFD models. Our primary
objective is to understand the modeling issues that arise and how they are dealt
with in CFD codes. To x ideas, we will look at a gassolid system (e.g., a
uidized bed) wherein the solid particles undergo growth, aggregation, and
breakage. Unlike in Section IV, we will assume here that the particle Stokes
number (St) can be larger than 0.14. Thus, it will be necessary to account for
particle momentum separately from the gas phase (or at least to account for the
effect of the second phase in the momentum balance). Nevertheless, the particle
size distribution (PSD) can be accounted for using DQMOM (Fan et al., 2004).
For simplicity, we will assume that the particle density r
s
is constant and
independent of particle size, and that the gas density r
g
is constant. At the
subgrid scale, the two phases are described by the volume fractions a
g
and a
s
for
the gas and solid phases, respectively. By denition, a
g
+a
s
1. In addition, the
DQMOM representation of the solid phase will introduce a volume fraction for
each abscissa a
sm
for m 1,y, M. By denition, a
s1
a
sM
a
s
. (See Fan
et al. (2004) for details.)
The multiuid CFD model at its most basic level consists of mass and mo-
mentum balances for each phase. For the present example, the mass balance
for the gas phase can be written as follows:
@r
g
a
g
@t
= r
g
a
g
U
g

M
m1
M
gm
(142)
where M
gm
is the mass-transfer rate from the gas to the solid phases. As we will
discuss later (Section V.B), a model must be provided to close the mass-transfer
RODNEY O. FOX 288
term. In words, r
g
a
g
is the mass of gas per unit volume of the multiphase
mixture (i.e., gas and solid phases), whereas r
g
is the mass of gas per unit
volume of gas. The gas velocity U
g
appears in the convection term for the gas
phase, and is normally not equal to the solids velocities.
The mass balance for the solid phases (m+1,y, M) can be written as follows:
@r
s
a
sm
@t
= r
s
a
sm
U
sm
M
gm
3k
v
r
s
l
2
m
b
m
2k
v
r
s
a
m
(143)
The solid velocities U
sm
(one for each abscissa l
m
) appear in the convection term.
The source terms a
m
and b
m
are found from the DQMOM representation of
aggregation and breakage of the solid particles (Fan et al., 2004). Each particle
phase is represented by its volume fraction (instead of its weight w
m
) and its
characteristic length l
m
. Note that growth of solid particles (with constant den-
sity r
s
) requires mass transfer from the gas phase, represented by M
gm
. In the
absence of growth the total solids volume fraction a
s
does not change. Thus, the
aggregation and breakage terms will cancel as follows:

M
m1
3l
2
m
b
m
2a
m
_ _
0 (144)
This property will result from applying DQMOM to the aggregation and
breakage terms in the PBE (Fan et al., 2004).
While the mass balances given above are relatively straightforward (assuming
that a suitable closure can be derived for the mass-transfer terms), the mo-
mentum balances are signicantly more complicated. In their simplest forms,
they can be written as follows:
@r
g
a
g
U
g
@t
= r
g
a
g
U
g
U
g
a
g
=p = r
g

M
m1
f
gm
r
g
a
g
g 145
and
@r
s
a
sm
U
sm
@t
= r
s
a
sm
U
sm
U
sm
a
sm
=p = r
sm
f
gm

M
n1
f
smn
r
s
a
sm
g 146
for the gas and solid phases, respectively. The two terms of the left-hand sides of
the momentum balances correspond to accumulation and convection. Note that
the conserved quantities (for example r
g
a
g
U
g
) are the momentum of a given
phase per unit volume of the mixture (hence the appearance of a
g
, etc.), and that
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 289
each phase has its own characteristic convection velocity (e.g., U
g
). The terms
on the left-hand side of Eqs. (145) and (146) account for changes in momentum
of each phase. Obviously, since momentum is exchanged between phases at the
interfaces and we are not resolving the interfaces, such phase-interaction terms
will require models. In contrast, body forces (i.e., gravity g in this example)
appear in closed form.
The rst term on the right-hand side of Eqs. (145) and (146) is a pressure term
shared by both phases. The purpose of this term (when r
s
and r
g
are constant) is
to ensure that the volume-average velocity, dened by
U
vol
a
g
U
g

M
m1
a
sm
U
sm
(147)
is solenoidal in the absence of mass transfer. Indeed, dividing the mass balances
(Eqs. 142 and 143) by the densities and neglecting the mass-transfer terms lead to
= U
vol
0 (148)
Thus, just as for incompressible single-phase ow, the pressure p constrains
the velocity elds to ensure (in the case of multiphase ows) that the sum of the
phase volume fractions equals unity. In the presence of mass transfer, the right-
hand side of Eq. (148) is nonzero; nevertheless, the role of the pressure is still the
same. Finally, we should note that in gassolid ows the maximum volume fraction
of the solid phase is less than unity due to physical constraints (i.e., when particles
are close packed there is still room for the gas phase so that 0oa
g
). To accom-
modate this constraint, it is common to introduce a solid-pressure term p
s
that
becomes extremely large when a
g
approaches its minimum value (e.g., a

g
0.4).
The second term on the right-hand side of Eqs. (145) and (146) contains the
viscous-stress models s
g
and s
sm
. Even for laminar ow, suitable forms for these
models are difcult to determine a priori. Typical models used in CFD introduce an
effective viscosity m
eff
for each phase, and describe the viscous stresses as follows.
r
x
m
eff;x
rU
x
rU
x

T
_
x g; sm (149)
Leaving aside the difcult question of whether this model holds for multiphase
ows, we still have the problem of determining m
eff,x
in terms of the computed
properties of the ow. The reader should appreciate that choosing an effective
viscosity for a multiphase ow is much more complicated than just adding a
turbulence model as done in single-phase turbulent ows. Indeed, even for a case
involving two uids (e.g., two immiscible liquids) for which the molecular viscosi-
ties are constant, the choice of the effective viscosities is not obvious. For example,
even if the mass-average velocity dened by
U
mass

r
g
a
g
U
g

M
m1
r
s
a
sm
U
sm
r
g
a
g
r
s
a
s
(150)
RODNEY O. FOX 290
were laminar, the ow around individual particles could be turbulent (as measured
by the particle Reynolds number dened below) and the effective viscosity should
reect this fact. The simplest models account for particle-scale turbulence using
an expression of the form
m
eff;g
m
g
1 C
s
a
s
Re
s
(151)
where m
g
is the molecular viscosity of the gas phase and Re
s
is a particle Reynolds
number dened by
Re
s

r
g
d
s
jU
s
U
g
j
m
g
(152)
and d
s
is the characteristic diameter of the particles. The model constant C
s
is order
unity, but must be t to experimental data. The effective viscosity in Eq. (151) has
the desired behavior in the limits where a
s
or Re
s
are very small; however, it is
unlikely to be accurate when the product of these terms is large.
The situation for cases where U
mass
is also turbulent is even more compli-
cated. First, such large-scale turbulence can be due to a variety of physical
phenomena, and thus have different characteristics. For example, large-scale
turbulence can (as in single-phase ows) be introduced through the boundary
conditions (e.g., turbulent jets) or by using mixing devices (e.g., stirred tanks).
For such cases, it may be possible to make suitable modications to single-phase
turbulence models to arrive at useful expressions for the effective viscosity. In
contrast, large-scale turbulence that arises due to internal properties of a mul-
tiphase ow (e.g., density differences between phases) is more difcult to de-
scribe by simple modications of standard turbulence models.
Second, due to the difculty of accessing multiphase ows with laser-based ow
diagnostics, there is very little experimental data available for validating multi-
phase turbulence models to the same degree as done in single-phase turbulent
ows. For example, thanks to detailed experimental measurements of turbulence
statistics, there are many cases for which the single-phase k-e model is known to
yield poor predictions. Nevertheless, in many CFD codes a multiphase k-e model
is used to supply multiphase turbulence statistics that cannot be measured exper-
imentally. Thus, even if a particular multiphase turbulent ow could be adequately
described using an effective viscosity, in most cases it is impossible to know
whether the multiphase turbulence model predicts reasonable values for m
eff
.
Third, many of the multiphase ows of interest to chemical engineers are in
regimes where both particle-scale and large-scale turbulence are signicant. For
example, in gasliquid bubble columns the particle Reynolds number (based on
the bubble rise velocity) is typically large. Thus, even for low gas-ow rates,
particle-scale turbulence will be signicant. However, at low gas-ow rates and
with uniform sparging, a bubble column will have no large-scale turbulence (i.e.,
the ow regime will be homogeneous) (Garnier et al., 2002; Harteveld et al.,
2003), and thus only the effective viscosity of individual particles should be
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 291
included in the model. As the gas ow rate is increased to a critical value of the
gas holdup (which can be as high as a
g
0.55 (Mudde, 2005)), the ow will
become unstable and large-scale turbulence will be generated. Although it has
been attempted in the literature (Thorat and Joshi, 2004), it is unlikely that a two-
uid k-e turbulence model has the necessary mathematical structure to correctly
predict ow transitions or even the turbulence levels observed in homogeneous
bubbly ows. In contrast, a two-uid model using only a particle-scale effective
viscosity and appropriate force models can predict ow transitions (Monahan
et al., 2005) in reasonable agreement with experiments (Harteveld et al., 2003).
Nevertheless, it is likely that different CFD models will be required for different
ow regimes (e.g., homogeneous vs. churn turbulent) and the user must be careful
not to extend a particular model beyond its range of applicability.
Returning to the momentum equations, the third term on the right-hand side
of Eqs. (145) and (146) contains the gassolid momentum-exchange models
f
gm
. Likewise, the fourth term on the right-hand side of Eqs. (146) contains the
solidsolid momentum-exchange model f
smn
. Note that because solidsolid in-
teractions conserve momentum, the latter must be dened such that

M
m1

M
n1
f
smn
0 (153)
Determination of accurate models for f
gm
and f
smn
is nontrivial (Drew and
Passman, 1999), and no consensus exists on the exact forms needed to describe
particular ows. Nevertheless, it is generally acknowledged that the momentum-
exchange model must include drag terms with forms similar to
f
gm
a
g
a
sm
C
D
Re
s

3m
g
Re
s
4d
2
s
U
sm
U
g
(154)
where C
D
(Re
s
) is a drag coefcient, and
f
smn
a
sm
a
sn
C
mn
U
sn
U
sm
(155)
where C
mn
depends on the properties of the solid phases (Gao et al., 2006). Note
that the drag models depend on the velocity difference between two phases, and
thus can be nonzero even for cases where the velocity elds are uniform in time
and space. Other forces that can be included depend on gradients (temporal or
spatial) of the velocities or volume fractions (Drew and Passman, 1999), and
thus are only signicant for inhomogeneous ows. However, as can be shown
using linear stability analysis (Batchelor, 1988; Lammers and Biesheuvel, 1996;
Minev et al., 1999; Jackson, 2000; Johri and Glasser, 2002; Sankaranarayanan
and Sundaresan, 2002), spatially uniform solutions to the multiuid model are
usually unstable, implying that the stationary, homogeneous solution to the
multiuid model is not representative of the ow. Thus, even when simulating
homogeneous ows, it is important to include all relevant forces when
RODNEY O. FOX 292
comparing numerical simulations with experiments (Monahan et al., 2005).
Moreover, the computational requirements in terms of gird size needed to attain
grid-independent solutions are relatively high for the laminar two-uid model
(Monahan et al., 2005). In the context of modeling chemical reactors, it will be
necessary to develop CFD models for the unresolved scales when applying
multiuid models to real reactors. Although procedures for developing such
models are still being actively investigated and no clear consensus has yet to
emerge (Sundaresan, 2000), here we will limit ourselves to a brief discussion of
the relevant issues.
To simplify the presentation, let us consider the transport equation for U
mass
found by summing together Eqs. (145) and (146):
@^ rU
mass
@t
= ^ rU
mass
U
mass
= ^ r u
d
u
d
=p = ^ r ^ rg (156)
where the phase-average density, dened by ^ r r
g
a
g
r
s
a
s
; obeys
@^ r
@t
= ^ rU
mass
0 (157)
Note that these expressions (Eqs. 156 and 157) appear deceptively simple (i.e., as
if the problem can be reduced to modeling a variable-density, single-phase ow)
because we have hidden the difcult terms in the denition of some new
symbols! First, the phase-average stress r is dened by
^ r r
g

M
m1
r
sm
(158)
and (based on the model in Eq. (149)) it is not a simple function of U
mass
. Second,
a new multiphase stress term u
d
u
d
has be introduced and is dened by
^ r u
d
u
d
r
g
a
g
u
g
u
g
r
sm
a
sm

M
m1
u
sm
u
sm
(159)
where u
g
U
g
U
mass
and u
sm
U
sm
U
mass
are the differences between
the phase velocities and the mass-average velocity. We can note that for a
constant-density, two-phase system with r
g
6 r
s
, U
g
and U
s
are related to U
mass
and U
vol
by
U
g

r
s
a
g
r
s
r
g

U
vol

^ r
a
g
r
s
r
g

U
mass
U
s

r
g
a
s
r
s
r
g

U
vol

^ r
a
s
r
s
r
g

U
mass
160
Thus the two-uid model can be formulated in terms of any two velocities chosen
from the set U
g
, U
s
, U
mass
and U
vol
, which might be useful, for example, to
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 293
examine the limiting case where r
g
(r
s
. The closures for shear stresses ^ r and
convection due to differences in the phase velocities u
d
u
d
in Eq. (156) are
necessarily ow dependent. Nevertheless, the simplest closures might use Eq.
(149) with U
mass
and an effective viscosity depending on ^ r and assume that the
slip velocity between phases u
d
is a known constant (e.g., depending on density
difference and bubble size). We should also note that it is possible to write the
model in terms of the volume-average velocity U
vol
(Eq. 147). However,
the resulting expressions are more complicated than Eqs. (156) and (157).
Up to this point we have not introduced any modeling concepts to deal with
large-scale turbulence. However, if the Reynolds number corresponding to
Eq. (156) is large enough, the velocity eld U
mass
will become turbulent. In this
case, the computational resources needed to resolve all of the relevant ow
scales will increase drastically, and the multiuid CFD model will no longer be
tractable for analyzing chemical reactors. To deal with this difculty, we can
introduce a multiphase turbulence model based on Reynolds averaging Eq.
(156). Because ^ r is not constant, we will in fact use the Favre average. For
example, if the Reynolds-average velocity is denoted by hU
mass
i, then the Favre-
average velocity is dened by

U
mass

h ^ rU
mass
i
h ^ ri
(161)
Note that we have also introduced the Reynolds-average, phase-average density
h^ ri. Applying the Favre average to Eq. (157) yields a closed expression for the
mass balance as follows:
@h^ ri
@t
= h ^ ri

U
mass
0 (162)
Applying the same process to Eq. (156) yields
@h ^ ri

U
mass
@t
= h ^ ri

U
mass

U
mass
= h ^ ri u
mass
u
mass

= h ^ ri

u
d
u
d
=hpi = ^ r h i h ^ rig 163
where the velocity uctuations due to large-scale turbulence are denoted by
u
mass
U
mass

~
U
mass
. The most important unclosed terms in Eq. (163) are the
turbulence stresses u
mass
u
mass
and the Favre-average multiphase stresses

u
d
u
d
.
The rst of these is usually closed by introducing a multiphase turbulence model
with appropriate modications to include the effect of interfacial momentum
exchange on production (dissipation) of large-scale turbulence. The second term

u
d
u
d
is not directly related to turbulent velocity uctuations. Instead, it will
depend on correlations between the phase-average density ^ r and the velocity
difference u
d
. For example, if the bubble rise velocity U
r
were constant and
independent of the gas volume fraction, then

u
d
u
d
% U
2
r
e
v
e
v
where e
v
g/|g|is
RODNEY O. FOX 294
the unit vector in the vertical direction. More generally, the turbulent two-uid
model for gasliquid ow should have properties similar to turbulence gener-
ated by buoyancy in single-phase ows (Riley and DeBruynKops, 2003). Like-
wise, in gassolid ows u
d
will depend on the Favre-average drag term, and thus
on the particle Stokes number through correlations between gas- and solid-
phase velocity uctuations (Fan and Zhu, 1998). Although computationally
expensive using present-day computers (Agrawal et al., 2001; Zhang and
VanderHeyden, 2002), it might be instructive to use direct simulations of the
laminar two-uid model (i.e., before Favre averaging) to parameterize multi-
phase turbulence models as has been done for stably stratied ows (Shih et al.,
2005). This possibility is especially attractive because it offers access to ow
statistics that cannot be measured experimentally. At the very least, it might
allow us to distinguish between the adequacy of the various multiphase tur-
bulence models available in the literature (Mudde, 2005).
The goal of the discussion above was obviously not to describe multiphase
turbulence models, but rather to point out the difculties encountered when
trying to derive a consistent set of transport equations. Although it is usually
not done, it is worthwhile to think of the averaging process used to arrive at
Eqs. (162) and (163) in two distinct steps: (1) ensemble averages over different
phase congurations to derive the laminar multiuid model (i.e., Eq. 146)
that can be used to describe multiphase ows without large-scale turbulence,
and (2) Reynolds or Favre averages (or even LES) to describe turbulent
multiphase ows. Ideally, direct numerical simulations (DNS) of two-phase
ows with resolved interfaces could be used to develop two-uid models for
laminar multiphase ows. For example, the recent work of Nguyen and Ladd
(2005) uses DNS to understand the sedimentation of mono- and poly-disperse
hard-sphere suspensions when the large-scale ow is laminar, and the work of
Bunner and Tryggvason (2003) uses DNS to investigate bubbly ows. If a
two-uid model (which does not resolve the interfaces) could be derived that
adequately reproduces these DNS data, then it could be used to investigate
the effects of large-scale turbulence that arises, for example, when the system is
subjected to shear (Lakehal et al., 2002). The results from direct simulations of
the two-uid model in the turbulent regime could then be used to develop
and validate multiphase turbulence models along the lines suggested by
Sundaresan (2000). Fortunately, with the continuing advances in computer
power, steady advances in DNS of two-phase systems can be expected. There is
thus reason to be optimistic that more powerful multiphase turbulence models
will eventually be available for modeling practical systems such as chemical
reactors.
In the current state of the art, almost all multiphase CFD models available
in commercial codes use some type of turbulence model based on extending
models originally developed for single-phase ows. Such CFD models are
thus meant to describe fully turbulent ows (as opposed to laminar or tran-
sitional ows). Nevertheless, many of these models have not been validated
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 295
experimentally in the context of dense multiphase ows for the reasons dis-
cussed earlier, and thus should be used with caution even for turbulent mul-
tiphase ows. In any case, there is still considerable room for improvement of
multiphase CFD models through comparison with carefully designed experi-
ments for canonical ows. Even more so than for single-phase turbulence, it can
be expected that particular models will have limited ranges of applicability and
will have to be tuned for multiphase ows with different physics. For ex-
ample, a two-uid model for a solidliquid slurry in an agitated reactor will
require different physical models than a churn-turbulent bubble column. In the
rst case, large-scale turbulence is generated by the agitation system, while in
the second case, it is generated by buoyancy and interfacial dynamics. However,
for specic ows of interest to the chemical industry, it should be possible to
develop reliable CFD models for multiphase ow dynamics that can be used to
investigate scalar transport and chemical reactions needed to model chemical
reactors. In the next section, we will thus look briey at the additional models
needed to describe the transport and production of chemical species and
thermal energy.
B. INTERPHASE MASS/HEAT-TRANSFER MODELS
Our discussion of multiphase CFD models has thus far focused on describing
the mass and momentum balances for each phase. In applications to chemical
reactors, we will frequently need to include chemical species and enthalpy
balances. As mentioned previously, the multiuid models do not resolve the
interfaces between phases and models based on correlations will be needed to
close the interphase mass- and heat-transfer terms. To keep the notation simple,
we will consider only a two-phase gassolid system with a
g
+a
s
1. If we
denote the mass fractions of N
sp
chemical species in each phase by Y
ga
and Y
sa
,
respectively, we can write the species balance equations as
@r
g
a
g
Y
ga
@t
= r
g
a
g
Y
ga
U
g
= J
ga
M
a
R
ga
, (164)
and
@r
s
a
s
Y
sa
@t
= r
s
a
s
Y
sa
U
s
= J
sa
M
a
R
sa
. (165)
The terms J
ga
and J
sa
are the diffusive uxes of species a in the gas and solid
phases, respectively. Note that in addition to molecular-scale diffusion, these
terms include dispersion due to particle-scale turbulence. The latter is usually
modeled by introducing a gradient-diffusion model with an effective diffusivity
along the lines of Eqs. (149) and (151). Thus, for large particle Reynolds num-
bers the molecular-scale contribution will be negligible. The term M
a
is the
RODNEY O. FOX 296
mass-transfer rate from the gas to the solid phase for species a. By denition, the
mass fractions sum to unity so that S
N
sp
a1
J
ga
S
N
sp
a1
J
sa
0 and M
g
S
N
sp
a1
M
a
(see Eq. 142). The terms R
ga
and R
sa
are the reaction rates for species a in each
phase. By denition, mass is conserved so that S
N
sp
a1
R
ga
0 and S
N
sp
a1
R
sa
0.
Although we do not write them explicitly here, the reader can appreciate that
the enthalpy balances for each phase will have a form similar to Eqs. (164) and
(165), and can be used to determine the temperatures T
g
and T
s
of each phase.
Likewise, mass transfer will lead to corresponding terms in the momentum
balances (Eqs. 145 and 146) (Bird et al., 2002).
CFD models for turbulent multiphase reacting ows do not solve the
laminar two-uid balances (Eqs. 164 and 165) directly. First, Reynolds
averaging is applied to eliminate the large-scale turbulent uctuations. Using
Eq. (164) as an example, we can apply Reynolds averaging to nd (with r
g
constant)
@r
g
ha
g
i

Y
ga
@t
= r
g
ha
g
i

Y
ga

U
g
_ _
= r
g
ha
g
i

Y
00
ga
u
00
g
_ _
= hJ
ga
i hM
a
i hR
ga
i 166
The Reynolds-average gas volume fraction ha
g
i is found from
@r
g
ha
g
i
@t
= r
g
ha
g
i

U
g
_ _
hM
g
i (167)
wherein the Reynolds-average mass-transfer term hM
g
i is unclosed. The Favre-
average gas velocity

U
g
is dened by

U
g

ha
g
U
g
i
ha
g
i
(168)
and its transport equation is found by Reynolds averaging Eq. (145). Although
we do not write it out explicitly here, the reader should appreciate that the
Reynolds-average gas-phase momentum equation has a number of unclosed
terms that require models.
Returning to Eq. (166), the third term on the left-hand side involves the
turbulent scalar uxes, dened by

Y
00
ga
u
00
g

ha
g
Y
00
ga
u
00
g
i
ha
g
i
(169)
where the scalar and velocity uctuations are dened by Y
00
ga
Y
ga


Y
ga
and
u
00
g
U
ga


U
g
; and respectively. The usual model for the scalar uxes is gra-
dient diffusion as follows:

Y
00
ga
u
00
g
G
T
=

Y
ga
(170)
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 297
where G
T
is a turbulent diffusivity that is computed from the multiphase tur-
bulence model. The CFD model can then be written as
@r
g
ha
g
i

Y
ga
@t
= r
g
ha
g
i

Y
ga

U
g
_ _
= G
eff
=
~
Y
ga
_ _
hM
a
i hR
ga
i 171
where we have combined the (usually negligible) particle-scale diffusive ux and
the turbulent uxes into an effective-diffusion coefcient (G
eff
).
As mentioned earlier, since the interfaces between phases are not resolved in
the CFD model, the Reynolds-average mass-transfer terms (hM
a
i), and the
Reynolds-average reaction rates (hR
ga
i) in Eq. (171) must be modeled in terms
of known quantities. This situation is very much like classical reaction engi-
neering models for multiphase reactors with the important difference that all
quantities are known locally. Such quantities include
a
g
_
;

Y
ga
;

Y
sa
;

T
g
;

T
s
;

U
g
;

U
s
and local multiphase turbulence statistics. Note that these variables, although
local, tell us nothing about the internal structure of the phases (i.e., subgrid-
scale information). For example,

T
s
represents the Favre-average temperature
of a solid particle consistent with the Favre-average enthalpy of a single
particle.
9
If, as is often the case, the temperature varies strongly between the
center and outer surface of a particle, a SGS model will be required to account
for this effect on, for example, the chemical reactions. The principal advantage
of using the CFD model over a classical CRE model is thus the ability to
account for the effect of local uid dynamics (e.g., j

U
g


U
s
j) on the mass/heat-
transfer rate between phases. In CFD codes, this is typically done by using
correlations for hM
a
i written in terms of the local Sherwood (or Nusselt)
number and particle Reynolds number (modied perhaps by a function of ha
g
i
to account for particleparticle interactions). In the case where large-scale
mixing is innitely rapid, these correlations will reduce to the classical CRE
models for homogeneous multiphase reactors. However, such cases are rare
(and need not be modeled using CFD), and it is more likely that large-scale
mixing will be rate limiting at certain locations within the reactor. Indeed, it is
exactly in such cases that CFD modeling will be of most benet for reactor
design and analysis.
9
This discussion also applies to the original variable T
s
, which represents the ensemble-average
temperature of particles located at a particular point at a given time. Basically, we know the total
enthalpy of each particle, but we do not know how it is distributed inside any given particle. Since
the reaction rate can be very sensitive to the local temperature, we will need a SGS model to describe
the coupling between intraparticle transport processes and chemical reactions.
RODNEY O. FOX 298
C. COUPLING WITH CHEMISTRY
The species balances given above (Eqs. 164 and 165) include the reaction
source terms on the right-hand sides. However, for these expressions to be
useful in CFD modeling, the user must supply the reaction rate functions and
the kinetic parameters. In addition, just as in single-phase turbulent reacting
ows (Fox, 2003), it may be necessary to account for micromixing effects on the
chemical kinetics by using SGS models in the Reynolds-average transport
equations (e.g., for hR
ga
i in Eq. (171)). For example, consider the parallel
reactions
AB !R
AC !S
with A in the gas phase and excess B and C in the liquid phase. If the rst
reaction is very rapid, then in the absence of micromixing effects in the liquid
phase only R would be produced. However, if A cannot be rapidly mixed into
the liquid phase (after mass transfer from the gas phase), then some S will be
produced as an unwanted by-product. In many ways, this situation is analogous
to the reaction systems discussed in Section III.B. If we dene a mixture fraction
x that is unity in the gas phase (i.e., for pure A) and initially zero in the liquid
phase, then the degree of conversion of the rst reaction (Y
1N
) can be para-
meterized by the value of x in the liquid phase (see Eq. 68). However, the rate of
the second reaction will depend on the local mixture fraction in a nontrivial
manner (see Eq. 70). Thus, if we simply ignore SGS uctuations in the liquid
mixture fraction we will likely severely underestimate the extent of the second
reaction. In theory, it is possible to write a presumed PDF transport model for
the mixture fraction in the liquid phase. However, unlike in single-phase tur-
bulence, the source term for the mixture fraction in the liquid phase is the mass-
transfer term and the sink term for mixture-fraction uctuations will depend on
the rate of molecular mixing in different regions around the interphase (e.g.,
boundary layer, wake, far eld). A similar situation is encountered in spray
combustion (Reve illon et al., 2004) where evaporating liquid droplets act as
source terms for reactants in the gas phase. It thus may be useful to adapt SGS
models for spray combustion (at least the parts modeling the mixture fraction
mean and variance) to describe SGS mixing in more general settings such as
gasliquid ows. One can also use direct simulations of bubbly reacting ows
(Khinast, 2001; Khinast et al., 2003; Koynov and Khinast, 2004; Raffensberger
et al., 2005) to explore the validity of SGS models developed for two-phase
reacting ows.
From the discussion above, we should keep in mind that even if no SGS
micromixing model is used to describe the multiphase ow, it may often be the
case that chemical reactions (and indeed micromixing) will be limited by mass/
heat transfer between the phases. Because the multiuid model (see Eqs. 164 and
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 299
165) includes mass-transfer terms, the reaction rates will usually be mass trans-
fer limited in cases where the chemical kinetics are fast. Thus, since we are
already relying on correlations to calculate the mass/heat-transfer rates, it may
not be fruitful to try to include a detailed description of micromixing inside each
phase. Indeed, it will more likely be the case that improving the correlations
(e.g., to include the effect of chemical reactions on local mass transfer) will have
a greater impact on the accuracy of CFD model predictions.
Finally, to conclude our discussion on coupling with chemistry, we should
note that in principle fairly complex reaction schemes can be used to dene the
reaction source terms. However, as in single-phase ows, adding many fast
chemical reactions can lead to slow convergence in CFD simulations, and the
user is advised to attempt to eliminate instantaneous reaction steps whenever
possible. The question of determining the rate constants (and their dependence
on temperature) is also an important consideration. Ideally, this should be done
under laboratory conditions for which the mass/heat-transfer rates are all faster
than those likely to occur in the production-scale reactor. Note that it is not
necessary to completely eliminate mass/heat-transfer limitations to determine
usable rate parameters. Indeed, as long as the rate parameters found in the lab
are reliable under well-mixed (vs. perfect-mixed) conditions, the actual mass/
heat-transfer rates in the reactor will be lower, leading to accurate predictions of
chemical species under mass/heat-transfer-limited conditions.
VI. Conclusions and Future Perspectives
From the brief overview of CFD models presented in this work, the reader will
hopefully have gained an initial appreciation of the utility and power of CFD tools
for chemical reactor analysis and design. In Section II we have discussed the basic
formulation and specic steps needed to set up a CFD model. We have also
introduced the key concept of subgrid-scale (SGS) modeling and its importance in
describing unresolved phenomena in reactor-scale CFD models. However, it is
worth repeating here that the development of SGS models for chemical reactions
and molecular transport is a natural extension of traditional chemical reaction
engineering modeling activities, and thus one of the key areas where chemical
engineers can have a large impact on the eld. In Section III we gave an example
of an SGS model developed for mixing-sensitive chemical reactions. This simple
model for mass-transfer-limited chemical reactions can be easily extended to other
applications such as high-Schmidt-number laminar ows. In Section IV we dis-
cussed efcient methods for adding a population balance equation to a CFD code
to model the production of ne particles. More generally, the ability to represent
the evolution of a population of entities (e.g., particles, bubbles, drops) in the
context of CFD results in a very powerful tool for describing complex reacting
ows. At present, methods based on the direct quadrature method of moments
RODNEY O. FOX 300
(DQMOM) are still in their initial stages of development. Nevertheless, they have
already been applied with great success by a growing number of researchers to a
wide variety of problems. The versatility of DQMOM was demonstrated in Sec-
tion IV by applying it twice to the one modeling problem: rst to model micro-
mixing between uid elements containing a bivariate number density function
(NDF), and then to represent the bivariate NDF itself. Finally, in Section V we
discussed the challenges associated with CFD models for multiphase reacting
ows. Although there are still a number of open problems to be solved in mul-
tiphase ow CFD, when used with caution existing CFD models can be used for at
least qualitative analysis of chemical reactors.
In my opinion, the perspective for future developments in the eld of CFD
modeling of chemical reactors is quite strong. On the one hand, the continued
growth of computational power both through faster computers and better
algorithms will make it possible to solve more and more complex problems. On
the other hand, we are fortunate in this eld that the basic microscopic balance
equations are known, even though they lead to complex multiscale, multiphysics
phenomena. The accessibility of large-scale computing facilities will allow us to
explore this complexity using direct simulations for specic academic pro-
blems that can be used to test the SGS models needed for CFD simulations of
industrial reactors. In general, advances in the development of SGS models will
require collaboration between computational physicist/chemist working on di-
rect simulation of academic problems and chemical reaction engineers develo-
ping multiphysics models. Indeed, the reader should appreciate that it is almost
never the case that a reliable SGS model can be developed by simply analyzing
the results from a large-scale direct simulation. Inversely, SGS model developed
without validation against detailed experimental or direct simulation data are
usually of limited value. Instead, the more fruitful approach is to rst develop a
tentative SGS model based on a preliminary understanding of the physics/
chemistry of the problem, and then to design a large-scale direct simulation to
test key assumptions/predictions of the model. This dialogue between model
development and model validation is continued until a suitably reliable SGS
model is found. SGS models developed in this manner have a strong funda-
mental underpinning and have a much greater chance of being applicable to a
wide range of operating conditions.
In summary, my recommendation for future progress in the eld is not to
follow the deceptively simple path of rushing toward the application of large-
scale CFD simulations to complex industrial reactor systems if the basic SGS
models have not rst been shown to be reliable on academic problems.
10
Rather, I would recommend that we proceed more cautiously with adequate
attention given to the development of the fundamental physical understanding
required to develop reliable CFD models. While this path will obviously require
10
Despite this word of caution, one should not lose site of the fact that there are many industrial
reactor systems that can be accurately simulated with existing SGS models!
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 301
patience and perseverance, in the long run it will undoubtedly be the surest way
to attain the chemical reaction engineers long-sought goal of experiment-free
scale-up of chemical reactors.
ACKNOWLEDGMENTS
The authors work in the area of CFD analysis of chemical reactors has been
supported nearly continuously for the last 15 years by the U.S. National Science
Foundation. The work on gassolid multiphase ows and population balances
was funded by the U.S. Department of Energy. The author would also like to
acknowledge support from several companies, including Air Products and
Chemicals, BASF, BASELL, BP Chemicals, Dow Chemical, DuPont Engi-
neering, and Univation Technologies. Last, but not least, the author wishes to
acknowledge his many collaborators over the years who are many in number to
name them individually.
REFERENCES
Agrawal, K., Loezos, P. N., Syamlal, M., and Sundaresan, S. J. Fluid Mech. 445, 151185 (2001).
Akhtar, M. K., Xiong, Y., and Pratsinis, S. E. AIChE J. 37, 15611570 (1991).
Aoun, M., Plasari, E., David, R., and Villermaux, J. Chem. Eng. Sci. 54, 11611180 (1999).
Baldyga, J. Chem. Eng. Sci. 49, 19852003 (1994).
Baldyga, J., Bourne, J. R., and Walker, B. Can. J. Chem. Eng. 76, 641649 (1998).
Batchelor, G. K. J. Fluid Mech. 193, 75110 (1988).
Bird, R. B., Stewart, W. E., and Lightfoot, E. N., Transport Phenomena. 2nd edn. John Wiley &
Sons, New York, USA (2002).
Briesen, H., Fuhrmann, A., and Pratsinis, S. E. Chem. Eng. Sci. 53, 41054112 (1998).
Bunner, B., and Tryggvason, G. J. Fluid Mech. 495, 77118 (2003).
Corrsin, S AIChE J. 10, 870877 (1964).
David, R., and Marcant, B. AIChE J. 40, 424432 (1994).
Davies, C. N. The sedimentation and diffusion of small particles. Proceedings of the Royal
Society of London. Series A, Mathematical and Physical Sciences. 200, 110113 (1949).
Drew, D. A., and Passman, S. L., Theory of Multicomponent Fluids. Springer-Verlag, Inc., New
York, USA (1999).
Dring, R. P. ASME J. Fluid Eng 104, 15 (1982).
Einstein, A. Annalen der Physik 17, 549560 (1905).
Elimelech, M., Gregory, J., Jia, X., and Williams, R. A., Particle Deposition and Aggregation,
Measurement, Modelling and Simulation. Butterworth-Heinemann, Woburn (1995).
Fan, L. S., and Zhu, C., Principles of Gas-Solid Flows. Cambridge University Press, New York
(1998).
Fan, R., Marchisio, D. L., and Fox, R. O. Powder Technol. 139, 720 (2004).
Fiorina, B., Gicquel, O., Vervisch, L., Carpentier, S., and Darabiha, N. Combust. Flame 140,
147160 (2005).
Fox, R. O. Chem. Eng. Process. 37, 521535 (1998).
RODNEY O. FOX 302
Fox, R. O., Computational Models for Turbulent Reacting Flows. Cambridge University Press,
Cambridge, UK (2003).
Fox, R. O., and Raman, V. Phys. Fluids 16, 45514564 (2004).
Fox, R. O., and Yeung, P. K. Phys. Fluids 15, 961985 (2003).
Friedlander, S. K., Smoke, Dust, and Haze. 2nd edn Oxford University Press, New York, USA
(2000).
Fuchs, N. A., The Mechanics of Aerosols. Pergamon Press, New York, USA (1964).
Gao, D., Fan, R., Subramaniam, S., Fox, R. O., and Hoffman, D. K., J, Fluid Eng. 128, 6268
(2005).
Garnier, C., Lance, M., and Marie , J. L. Exp. Therm. Fluid Sci. 26, 811815 (2002).
Garside, J., and Tavare, N. S. Chem. Eng. Sci. 40, 14851493 (1985).
Harteveld, W. K., Mudde, R. F., and Van Den Akker, H. E. A. Can. J. Chem. Eng. 81, 389394
(2003).
Jackson, R., The Dynamics of Fluidized Particles. Cambridge University Press, Cambridge, UK
(2000).
Johannessen, T., Pratsinis, S. E., and Livbjerg, H. Chem. Eng. Sci. 55, 177191 (2000).
Johannessen, T., Pratsinis, S. E., and Livbjerg, H. Powder Technol. 118, 242250 (2001).
Johnson, B. K., and Prudhomme, R. K. AIChE J. 49, 22642282 (2003a).
Johnson, B. K., and Prudhomme, R. K. Aust. J. Chem. 56, 10211024 (2003b).
Johnson, B. K., and Prudhomme, R. K. Phys. Rev. Lett. 91, 91 (2003c).
Johri, J., and Glasser, B. J. AIChE J. 48, 16451664 (2002).
Kammler, H. K., Mueller, R., Senn, O., and Pratsinis, S. E. AIChE J. 47, 15331543 (2001).
Kammler, H. K., and Pratsinis, S. E. J. Nanopart. Res. 1, 467477 (1999).
Kammler, H. K., and Pratsinis, S. E. Chem. Eng. Process. 39, 219227 (2000).
Khinast, J. G. AIChE J. 47, 23042319 (2001).
Khinast, J. G., Koynov, A., and Leib, T. M. Chem. Eng. Sci. 58, 39613971 (2003).
Klimenko, A. Y., and Bilger, R. W. Prog. Energ. Combust. Sci. 25, 595687 (1999).
Kodas, T. T., Friedlander, S. K., and Pratsinis, S. E. Ind. Eng. Chem. Res. 26, 19992007 (1987).
Kolhapure, N. H., Fox, R. O., Daiss, A., and Ma hling, F. -O. AIChE J. 51, 585 (2005).
Koynov, A., and Khinast, J. G. Chem. Eng. Sci. 59, 39073927 (2004).
Kruis, F. E., Kusters, K. A., Pratsinis, S. E., and Scarlett, B. Aerosol Sci. Tech. 19, 514526 (1993).
Kumar, S., and Ramkrischna, D. AIChE J. 51, 13111332 (1996).
Lakehal, D., Smith, B. L., and Milelli, M. J. Turbul. 3, 121 (2002).
Lammers, J. H., and Biesheuvel, A. J. Fluid Mech. 328, 6793 (1996).
Lattuada, M., Wu, H., and Morbidelli, M. J. Colloid Interf. Sci. 268, 96105 (2003a).
Lattuada, M., Wu, H., and Morbidelli, M. J. Colloid Interf. Sci. 268, 106120 (2003b).
Liu, Y., and Fox, R. O. AIChE J. 52, 731734 (2006).
Liu, Y., Raman, V., Fox, R. O., and Harvey, A. D. Chem. Eng. Sci. 59, 51675176 (2004).
Mahajan, A. J., and Kirwan, D. J. J. Phys. D Appl. Phys. 26, B176B180 (1993).
Mahajan, A. J., and Kirwan, D. J. AIChE J. 42, 18011814 (1996).
Marcant, B., and David, R. AIChE J. 37, 16981710 (1991).
Marchisio, D. L., and Fox, R. O. J. Aerosol Sci. 36, 4373 (2005).
Marchisio, D. L., Fox, R. O., and Barresi, A. A. AIChE J. 47, 664676 (2001a).
Marchisio, D. L., Fox, R. O., Barresi, A. A., and Baldi, G. Ind. Eng. Chem. Res. 40, 51325139
(2001b).
Marchisio, D. L., Pikturna, J. T., Fox, R. O., Vigil, R. D., and Barresi, A. A. AIChE J. 49,
12661276 (2003a).
Marchisio, D. L., Soos, M., Sefcik, J., and Morbidelli, M. AIChE J. 52, 158173 (2006).
Marchisio, D. L., Vigil, R. D., and Fox, R. O. J. Colloid Interf. Sci. 258, 322334 (2003b).
McGraw, R. Aerosol Sci. Tech. 27, 255265 (1997).
Meakin, P. Adv. Colloid Interfac. Sci. 28, 249331 (1988).
Melis, S., Verduyn, M., Storti, G., Morbidelli, M., and Baldyga, J. AIChE J. 45, 13831393 (1999).
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 303
Minev, P. D., Lange, U., and Nandakumar, K. J. Fluid Mech. 394, 7396 (1999).
Minier, J. -P., and Peirano, E. Phys. Rep. 352, 1214 (2001).
Monahan, S. M., Vitankar, V. S., and Fox, R. O. AIChE J. 51, 18971923 (2005).
Mudde, R. F. Toward modeling and simulations of industrial bubbly ows, in 11th Workshop
on Two-Phase Flow Predictions, Institut fu r Verfahrenstechnik, Martin-Luther-Universita t,
Halle-Wittenberg, Germany (2005).
Mueller, R., Jossen, R., Kammler, H. K., Pratsinis, S. E., and Akhtar, M. K. AIChE J. 50,
30853094 (2004a).
Mueller, R., Jossen, R., Pratsinis, S. E., Watson, M., and Akhtar, M. K. J. Am. Ceram. Soc. 87,
197202 (2004b).
Nguyen, N. -Q., and Ladd, A. J. C. J. Fluid Mech. 525, 73104 (2005).
Oles, V. J. Colloid Interf. Sci. 154, 351358 (1992).
Pandya, J. D., and Spielman, L. A. Chem. Eng. Sci. 38, 19831992 (1983).
Peters, N., Turbulent Combustion. Cambridge University Press, Cambridge, UK (2000).
Piton, D., Fox, R. O., and Marcant, B. Can. J. Chem. Eng. 78, 983993 (2000).
Pohorecki, R., and Baldyga, J. Chem. Eng. Sci. 38, 7983 (1983).
Pohorecki, R., and Baldyga, J. Chem. Eng. Sci. 43, 19491954 (1988).
Poinsot, T., and Veynante, D., Theoretical and Numerical Combustion. R. T. Edwards,
Philadelphia, USA (2001).
Pope, S. B., Turbulent Flows. Cambridge University Press, Cambridge, UK (2000).
Pratsinis, S. E. Prog. Energ. Combust. Sci. 24, 197219 (1998).
Pratsinis, S. E., Zhu, W., and Vemury, S. Powder Technol. 86, 8793 (1996).
Raffensberger, J. A., Glasser, B. J., and Khinast, J. G. AIChE J. 51, 14821496 (2005).
Raman, V., Fox, R. O., and Harvey, A. D. Combust. Flame 136, 327350 (2004).
Raman, V., Fox, R. O., Harvey, A. D., and West, D. H. Ind. Eng. Chem. Res. 42, 25442557 (2003).
Ramkrishna, D., Population Balances. Academic Press, San Diego, USA (2000).
Randolph, A. D., and Larson, M. A., Theory of Particulate Processes. 2nd edn. Academic Press,
San Diego, USA (1988).
Reve illon, J., Pera, C., Massot, M., and Knikker, R. J. Turb. 5, 127 (2004).
Riley, J. J., and DeBruynKops, S. M. Phys. Fluids 15, 20472059 (2003).
Rosner, D. E., and Pykkonen, J. J. AIChE J. 48, 476491 (2002).
Sandku hler, P., Sefcik, J., Lattuada, M., Wu, H., and Morbidelli, M. AIChE J. 49, 15421555
(2003).
Sandku hler, P., Sefcik, J., and Morbidelli, M. J. Phys. Chem. B 108, 2010520121 (2005).
Sankaranarayanan, K., and Sundaresan, S. Chem. Eng. Sci. 57, 35213542 (2002).
Sanyal, J., Marchisio, D. L., Fox, R. O., and Dhanasekharan, K. Ind. Eng. Chem. Res. 44,
50635072 (2005).
Seckler, M. M., Bruinsma, O. S. L., and Van Rosmalen, G. M. Chem. Eng. Commun. 135, 113131
(1995).
Selomulya, C., Bushell, G., Amal, R., and Waite, T. D. Chem. Eng. Sci. 58, 327338 (2003).
Shih, L. H., Koseff, J. R., Ivey, G. N., and Ferziger, J. H. J. Fluid Mech. 525, 193214 (2005).
Sorensen, C. M. Aerosol Sci. Technol. 35, 648687 (2001).
Sreenivasan, K. R. Phys. Fluids 8, 189196 (1996).
Sundaresan, S. AIChE J. 46, 11021105 (2000).
Tani, T., Takatori, K., and Pratsinis, S. E. J. Am. Ceram. Soc. 87, 523525 (2004a).
Tani, T., Takatori, K., and Pratsinis, S. E. J. Am. Ceram. Soc. 87, 365370 (2004b).
Thorat, B. N., and Joshi, J. B. Exp.l Therm. Fluid Sci. 28, 423430 (2004).
Veynante, D., and Vervisch, L. Prog. Energ. Combust. Sci. 28, 193266 (2002).
Villermaux, J., and David, R. Journal de la Chimie Physique 85, 273 (1988).
Waldner, M. H., Sefcik, J., Soos, M., and Morbidelli, M. Powder Technol. 156, 226234 (2005).
Wang, L., and Fox, R. O. Chem. Eng. Sci. 58, 43874401 (2003).
Wang, L., and Fox, R. O. AIChE J. 50, 22172232 (2004).
RODNEY O. FOX 304
Wang, L., Marchisio, D. L., Vigil, R. D., and Fox, R. O. J. Colloid Interf. Sci. 282, 380396 (2005a).
Wang, L., Vigil, R. D., and Fox, R. O. J. Colloid Interf. Sci. 285, 167178 (2005b).
Watanabe, T., and Gotoh, T. New J. Phys. 6, 4075 (2004).
Wright, D. L., Mcgraw, R., and Rosner, D. E. J. Colloid Interf. Sci. 236, 242251 (2001).
Xiong, Y., and Pratsinis, S. E. J. Aerosol Sci. 22, 637655 (1991).
Yeung, P. K., Donzis, D. A., and Sreenivasan, K. R. Phys. Fluids 17, 081703 (2005).
Zhang, D. Z., and VanderHeyden, W. B. Int. J. Multiphas. Flow 28, 805822 (2002).
Zhu, W., and Pratsinis, S. E. AIChE J. 43, 26572664 (1997).
Zucca, A., Marchisio, D. L., Barresi, A. A., and Fox, R. O. Chem. Eng. Sci. 61, 8795 (2006).
CFD MODELS FOR ANALYSIS AND DESIGN OF CHEMICAL REACTORS 305
PACKED TUBULAR REACTOR MODELING AND CATALYST
DESIGN USING COMPUTATIONAL FLUID DYNAMICS
Anthony G. Dixon
1
, Michiel Nijemeisland
2
and E. Hugh Stitt
2
1
Worcester Polytechnic Institute, Department of Chemical Engineering, Worcester,
MA 01609, USA
2
Johnson Matthey Catalysts, Billingham, Cleveland, UK
I. Introduction 308
A. CFD and Packed Reactor Tube Modeling 308
B. CFD Approaches to Interstitial Flow in Fixed Beds 312
II. Principles of CFD for Packed-Tube Flow Simulation 315
A. CFD Basics and Turbulence Modeling 315
B. Packed Bed CFD Model Development 325
C. Packed Bed CFD Simulation Issues 334
D. Validation of CFD Simulations for Packed Beds 342
III. Low-N Packed Tube Transport and Reaction Using CFD 348
A. Hydrodynamics and Pressure Drop 348
B. Mass Transfer, Dispersion, and Reaction 352
C. Heat Transfer 356
IV. Catalyst Design for Steam Reforming Using CFD 363
A. Steam Reforming and Principles of Catalyst Design 363
B. CFD Simulation of Reformer Tube Heat Transfer with
Different Catalyst Particles 367
C. Reaction Thermal Effects in Spheres Using CFD 372
D. Reaction Thermal Effects in Cylinders Using CFD 378
V. Future Prospects 381
References 386
Abstract
Computational uid dynamics (CFD) is rapidly becoming a standard
tool for the analysis of chemically reacting ows. For single-phase
reactors, such as stirred tanks and empty tubes, it is already well-
established. For multiphase reactors such as xed beds, bubble columns,
trickle beds and uidized beds, its use is relatively new, and methods are
still under development. The aim of this chapter is to present the ap-
plication of CFD to the simulation of three-dimensional interstitial ow
in packed tubes, with and without catalytic reaction. Although the use of
307
Advances in Chemical Engineering, vol. 31
ISSN 0065-2377
DOI: 10.1016/S0065-2377(06)31005-8
Copyright r 2006 by Elsevier Inc.
All rights reserved
CFD to simulate such geometrically complex ows is too expensive and
impractical currently for routine design and control of xed-bed reac-
tors, the real contribution of CFD in this area is to provide a more
fundamental understanding of the transport and reaction phenomena in
such reactors. CFD can supply the detailed three-dimensional velocity,
species and temperature elds that are needed to improve engineering
approaches. In particular, this chapter considers the development of
CFD methods for packed tube simulation by nite element or nite
volume solution of the governing partial differential equations. It dis-
cusses specic implementation problems of special relevance to packed
tubes, presents the validation by experiment of CFD results, and reviews
recent advances in the eld in transport and reaction. Extended discus-
sion is given of two topics: heat transfer in packed tubes and the design
of catalyst particles for steam reforming.
I. Introduction
A. CFD AND PACKED REACTOR TUBE MODELING
The design of chemical reactors to make useful chemical products must be
able to accommodate new catalysts, new feedstocks, and new product speci-
cations, while facing ever-tighter economic, environmental, safety, and social
acceptability constraints. The reactor designer must strive for higher yields, less
energy use, smaller reactor capital costs, and other aspects of sustainable
processing. To accomplish these objectives, the reactor designer must be able
to understand, quantify, and control the individual chemical and physical phe-
nomena present in reactors. An essential part of such a task is the development
of predictive models of reactor behavior, based on the true representation of the
physical and chemical processes that occur, on different length scales. Capturing
the real physics and chemistry is especially important for heterogeneous reac-
tors, such as gassolid or liquidsolid xed beds, and gasliquidsolid trickle
beds. It is necessary to know the spatial distribution of reactants, catalysts,
inerts, and products in detail (Lerou and Ng, 1996).
A catalytic xed bed reactor is a (usually) cylindrical tube that is randomly
lled with porous catalyst particles. These are frequently spheres or cylindrical
pellets, but other shapes are also possible. The use of rings or other forms of
particles with internal voids or external shaping is on the increase. During
single-phase operation, a gas or liquid ows through the tube and over the
catalyst particles, and reactions take place on the surfaces, both interior and
exterior, of the particles.
Single-phase catalytic xed bed reactors are the main reactor type used for large-
scale heterogeneously catalyzed gas-phase reactions. Frequently, multitubular
ANTHONY G. DIXON ET AL. 308
xed beds with low tube-to-particle diameter ratio (N) are used for strongly
exothermic reactions such as partial oxidations and selective hydrogenations
as well as strongly endothermic reactions such as steam reforming of methane.
In these processes heat must be rapidly transferred into or out of a narrow
reactor tube, while the need to reduce compressor costs dictates a low pressure
drop along the tube, and so the particle size cannot be too small. These con-
straints combine to give tubes with low values of N. The presence of the tube
wall has a strong inuence on heat transfer, reaction rates, and selectivity for
these reactor tubes, which have proved to be exceptionally difcult to model.
Current xed bed reactor models have been based on fairly strong simplifying
assumptions, such as pseudo homogeneity, unidirectional plug ow, effective
transport parameters, and uniform catalyst pellet surroundings. These simpli-
cations have been motivated in the past by the need for computational savings,
which continues to become less of an issue. The complex structure of random-
packed tubes has also prompted a simplied approach. These idealized models
have led, however, to problems. Even the most advanced models today cannot
quantitatively predict reactor behavior if independently determined kinetics and
transport parameters are used (Schouten et al., 1994; Landon et al., 1996). The
effects of tube and catalyst pellet design changes are masked by the use of
effective parameters and simplied models. Reaction engineers are in agreement
that the entire eld has neglected the role of uid ow in reactor modeling. For
xed beds, a better understanding of uid ow through arrays of realistic cat-
alyst particle shapes would be of great help, with special attention to the prob-
lematical wall region. The presence of the tube wall causes changes in bed
structure, ow patterns, transport rates, and the amount of catalyst per unit
volume, and is usually the location of the limiting heat transfer resistance.
New techniques in experimentation and computation that allow us to
understand and model xed-bed phenomena at the particle or subparticle level
are needed. The desire to measure uid ow inside the bed has led several
researchers to use noninvasive experimental methods. McGreavy et al. (1984,
1986) used laser Doppler velocimetry (LDV) in low-N packed beds, for both
liquid and gas experiments, although only results using liquids were presented.
Particle tracking methods were used by Rashidi et al. (1996) and also by
Stephenson and Stewart (1986). The latter authors used marker bubbles as a
noninvasive method to measure the radial distribution of ow of a matched-
refractive index uid in transparent packed beds of equilateral cylinders
with N 10.7. They found that the local supercial velocity attained its global
maximum at 0.2d
p
from the wall and its global minimum at 0.5d
p
from
the wall. Both studies found an oscillatory radial velocity prole. This type of
prole was conrmed recently, again using LDV (Giese et al., 1998), for a
column with a tube-to-particle diameter ratio of approximately 9. Comparisons
were made with the extended Brinkman model, and good agreement was
obtained when an adjustable effective viscosity was introduced into the term for
wall effects.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 309
A noninvasive experimental method that has been used to obtain local ow
patterns in xed beds is magnetic resonance imaging (MRI). This method can
show ow patterns in complicated geometries. The method has so far been
restricted to relatively low ow rates, and to uids that can produce a suitable
signal for measurement, such as water. Gas ow has rarely been investigated by
MRI techniques. Generally the packed beds used for MRI have had a consid-
erably higher tube-to-particle diameter ratio, which will result in less pro-
nounced wall effects. Qualitatively the MRI results show generally accepted
ow concepts such as ow increase in bed voids, as well as inhomogeneous
velocity distribution in different pores (Sederman et al., 1997, 1998). The larger
tube-to-particle diameter ratio also allows for a statistical view of the velocity
distribution over the column cross section. When averaged over a long evo-
lution time, the data approached Gaussian behavior (Park and Gibbs, 1999).
With a tube-to-particle diameter ratio of 6.7 and relatively low ow rates, the
velocity prole was roughly parabolic with the maximum being near the center
of the tube. Also, negative velocities or reversed ow within the bed were shown
(Kutsovsky et al., 1996). Recent applications of the noninvasive MRI technique
have been made by Suekane et al. (2003) to a cubic-packing xed bed unit cell
and by Yuen et al. (2002) to isothermal reacting ows. Ren et al. (2005) used
NMR methods to obtain proles of velocity in a narrow tube.
All these methods allow observations of ow inside the xed bed without
disturbing the bed structure and can thus further our understanding, but each is
subject to severe limitations. LDV requires windows for optical access and is
restricted to beds of very low N where such voids occur naturally. It also re-
quires the uid to be refractive-index-matched with the transparent material of
the column. MR methods are mainly used for liquids, and techniques to allow
imaging of fast ows are starting to appear (Gladden et al., 2005). Particle
tracking methods require observation and counting of the markers, and prob-
lems with choice of uid similar to those with LDV are found. Experimental
methods do not yet let us get to conditions of interest for xed bed reactor
design, i.e., gas-phase high ow rates at elevated temperatures with conduction
and reaction in catalyst particles. While experimental techniques continue to
improve, a complementary approach is to take advantage of the recent advances
in scientic computing to simulate the ow elds. Computational uid dynamics
(CFD) has become a standard tool in the eld of chemical engineering. The
general setup of most CFD programs allows for a wide range of applications,
and several commercial packages have introduced chemical reactions into the
CFD code allowing rapid progress in the use of CFD within the eld of chemi-
cal reaction engineering (Bode, 1994; Harris et al., 1996; Kuipers and van
Swaaij, 1998; Ranade, 2002).
The application of CFD to packed bed reactor modeling has usually involved
the replacement of the actual packing structure with an effective continuum
(Kvamsdal et al., 1999; Pedernera et al., 2003). Transport processes are then
represented by lumped parameters for dispersion and heat transfer (Jakobsen
ANTHONY G. DIXON ET AL. 310
et al., 2002). The reactions that take place in the porous catalyst particles are
represented by source or sink terms in the conservation equations (Ranade,
2002) and corrected for volume fraction and particle transport limitations. The
velocity eld can be obtained from a modied momentum balance (Bey and
Eigenberger, 1997) or a form of the Brinkman-Forcheimer-extended Darcy
(BFD) equation (Giese et al., 1998). These approaches provide an averaged
supercial velocity eld, usually in the form of a radially varying axial com-
ponent of velocity, which is an improvement over the classical assumption of
plug ow (constant unidirectional ow). These velocity elds have been used in
improved models of xed bed transport and reaction (Winterberg et al., 2000).
The disadvantages of the BFD approach have been the continued lumping of
transport processes, thus obscuring the physical basis of the model, and the
necessity to introduce an effective viscosity for the bed to bring computed and
experimental velocity proles into agreement (Bey and Eigenberger, 1997; Giese
et al., 1998). This form of CFD in xed beds is an extension of the classical
pseudo-continuum approach, in which both uid and solid phases are modeled
as inter-penetrating continua, i.e., as if they coexisted at every point in the tube
(Fig. 1a).
An alternative and complementary use of CFD in xed bed simulation has
been to solve the actual ow eld between the particles (Fig. 1b). This approach
does not simplify the geometrical complexities of the packing, or replace them
by the pseudo-continuum that is used in the rst approach. The governing
equations for the interstitial uid ow itself are solved directly. The contrast is
thus between the interstitial ow eld type of simulation and the supercial ow
(a) pseudocontinuum CFD
(b) interstitial CFD
FIG. 1. Comparison of (a) pseudo-continuum and (b) interstitial CFD approaches to packed-tube
simulation.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 311
eld models of the BFD approach (Tobis , 2000). The equations of the inter-
stitial approach are well established and relatively straightforward; however,
the geometric modeling and grid generation become complicated and the com-
putational demands rise signicantly (Ranade, 2002). Owing to the computa-
tional requirements, the approach can so far be applied only to small, periodic
regions of the reactor. It is therefore useful mainly as a learning tool, from
which we can develop detailed insight into xed bed ow structures and un-
derstand how they inuence transport and reaction. The understanding from
this more rigorous approach to CFD can then be used to inform the simplifying
decisions made in the development of the more computationally tractable
pseudo-continuum xed bed models.
B. CFD APPROACHES TO INTERSTITIAL FLOW IN FIXED BEDS
The development of CFD calculations of interstitial ow in xed beds has
increased and become more realistic, as greater computational power has be-
come more available over the recent years. Calculations of velocity and pressure
proles for creeping ow between spheres were done by Snyder and Stewart
(1966) using Galerkins method, followed by Srensen and Stewart (1974)
using specially designed collocation methods. They were able to obtain the
velocity and temperature proles in cubic arrays of spheres, a highly symmetric
arrangement. Their calculations yielded insight into the behavior of the heat
transfer coefcient for particle-to-uid heat and mass transfer, over a wide
range of values of the Peclet number Pe RePr, where Re is the particle
Reynolds number, based on supercial velocity (rv
0
d
p
/m) and Pr is the Prandtl
number (mc
p
/k
f
). Flow through cubic arrays of particles was also studied by
Lahbabi and Chang (1985) by analytical methods, with focus on ow transitions.
Dalman et al. (1986) investigated ow around two spheres near a wall using
two-dimensional (2D) nite element models in an axisymmetric radial plane.
This study showed that eddies formed between the spheres, which led to regions
of poor heat transfer. Lloyd and Boehm (1994) also did a 2D study, with eight
spheres in line, to determine the inuence of the sphere spacing on the drag
coefcients and the particle-uid heat transfer coefcient. A 3D nite element
method was used by Mansoorzadeh et al. (1998) to simulate ow past a heated/
cooled sphere at moderate Re. They found good agreement between calculated
drag coefcients and a literature correlation, and that the axisymmetry of the
wake ow broke down at Re 400. Heat transfer from or to the sphere
increased the drag coefcient at higher Re. It was found that heat transfer from
the spheres decreased with decreased sphere spacing. McKenna et al. (1999)
used discrete particle CFD to obtain valuable insight into the effect of particle
size on particle-uid heat transfer during olen polymerization in a uidized
bed. They used a commercial code, Fluent, to conduct a 2D CFD study of small
clusters of catalyst particles and a 3D study of a single catalyst sphere close to a
ANTHONY G. DIXON ET AL. 312
wall. They explored two ways to include the local energy effects of reaction. The
rst was by specifying a surface heat ux for systems of two touching particles
or three particles with one small hot particle touching two larger, colder ones.
Their second approach was to utilize a constant volumetric heat source in par-
ticles with solid conduction included, to allow for dilution of active sites and
heat generation as the particle grows. Their main nding was that signicant
heat is removed from these particles by conduction, as well as the usually
assumed path of convection.
Three-dimensional (3D) models have been developed more recently. A simple
three-sphere model (Derkx and Dixon, 1996) focused on obtaining wall
heat transfer coefcients. An eight-sphere model followed (Logtenberg and
Dixon, 1998a, b) in which the packing was modeled as two layers of four
spheres, perpendicular to the ow in a tube with a tube-to-particle diameter
ratio of N 2.43. This study was limited by the absence of contact points
between the spheres and the wall and between the spheres themselves. Sub-
sequently, a 3D 10-sphere model was developed, with N 2.68, incorporating
contact points between the particles and between the particles and the wall
(Logtenberg et al., 1999), which used spherical dead volumes with estimated
diameters, around the contact points. These studies focused on using CFD to
obtain the traditional radial heat transfer modeling parameters such as the wall
heat-transfer coefcient (h
w
) and the effective radial thermal conductivity (k
r
),
and gave reasonable qualitative agreement with experimental estimates. Other
heat transfer work in xed beds has explored the use of CFD to simulate
ow and transport in structured packings (Von Scala et al., 1999) and to in-
vestigate the effects of roughness gaps in particleparticle heat conduction
(Lund et al., 1999).
So far, there have only been a few modeling studies to try to link local uid
ow to bed structure. Chu and Ng (1989) and later Bryant et al. (1993) and
Thompson and Fogler (1997) used network models for ow in packed beds.
Different beds were established using a computer simulation method for cre-
ating a random bed. The model beds were then reduced to a network of pores,
and either ow/pressure drop relations or Stokes law was used to obtain a ow
distribution.
Several groups have studied the connection between uid ow and bed
structure in complete particle beds. Esterl et al. (1998) and Debus et al. (1998)
applied a computational code by Nirschl et al. (1995) to nd ow proles in a
square channel, using an adapted chimera grid. This grid consisted of a struc-
tured grid, based on the owing medium, which was overlaid by a separate
structured grid, based on the packing particles. Calculated pressure drops were
compared against predicted pressure drops using, amongst others, Erguns rela-
tion for a bed with the same porosity; the simulation data gave the same order
of magnitude. Simulations were performed in beds with up to 300 spheres,
although the bed for which results were discussed consisted of 120 spheres. One
of the aspects that may have affected the accuracy of the simulations is that the
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 313
bed was only approximately ve layers deep, resulting in a ow that mainly
consisted of inlet and outlet effects.
Still other research groups (Georgiadis et al., 1996; Maier et al., 1998; Manz
et al., 1999; Zeiser et al., 2001) have used the lattice Boltzmann method (LBM)
for simulation of ow in a xed bed of spheres. A dense packing of spheres in a
cylindrical column was created from experimental observations, such as MRI,
or by computer simulation, for example by using a raining and compression
algorithm. The created packing geometry was then divided into an equidistant
Cartesian grid, where individual elements were labeled as solid or uid regions
as in a marker-and-cell approach. A high resolution of the grid made it possible
to obtain accurate ow proles. Recently (Zeiser et al., 2002; Freund et al.,
2003; Yuen et al., 2003) simple reactions have been added to the simulation,
showing species and conversion proles inside the bed. Limitations of the LBM
are that simulations of turbulence are expensive, as the method corresponds to a
direct numerical simulation (DNS) approach to the Navier-Stokes equations,
and that it is difcult to include heat transfer. Incorporating heat transfer into
xed bed simulations is extremely challenging, due to the need to mesh and
solve both convection in the voids and conduction in the particles. For LBM
thermal models in the uid, a multispeed approach must be used, usually from
two to four speeds is feasible. Due to the small number of speeds, the variation
in temperature is restricted. In addition, LBM methods are inherently transient
and thus more computationally expensive than steady-state differential equa-
tion-based approaches, and they suffer from instability, which is worse for
multispeed methods (Chen and Doolen, 1998).
A growing number of studies are appearing in which CFD methods are being
used to simulate multi-physics ow and heat transfer at higher ow rates in xed
beds. Calis et al. (2001) applied the commercial code CFX-5.3 to a structured
packing of spheres. They simulated ow in a number of channels of square
cross-section lled with spherical particles. Several different types of structured
packing were investigated, all based on structured packing of spheres. The re-
petitive sections had varying N, from 1 to 4. Values for pressure drop obtained
from the simulations were validated against experimental values. The turbulence
models used (ke and RSM) showed similar results, with an average error from
the experimental values of about 10%. Pressure drop in a structured packing
has also been studied by Larachi et al. (2003) and Petre et al. (2003), who used
the Fluent CFD code to obtain ow elds that allowed them to construct
submodels of different contributions to the overall pressure drop.
Pressure drop and dispersion were the focus of work by Magnico (2003) who
simulated ow at lower Re by direct numerical simulation (DNS) in beds of
spheres with an in-house code. Tobis (2000) simulated a small cluster of four
spheres with inserts between them to compare to his experimental measurements
of pressure drop. Gunjal et al. (2005) also focused on ow and pressure drop
through a small cell of spheres, in order to validate the CFD approach by
comparison to the MRI measurements in the same geometry made by Suekane
ANTHONY G. DIXON ET AL. 314
et al. (2003). Romkes et al. (2003) extended the study of Calis et al. to include
calculations for the heat transfer coefcient for a single sphere in an innite
medium, the wall heat transfer coefcient for laminar and turbulent ow in an
empty tube, and the particle-to-uid heat transfer coefcient in a packed tube.
Heat transfer in full beds and bed segments has been the main focus of our own
work (Dixon and Nijemeisland, 2001; Nijemeisland and Dixon, 2001, 2004)
with special emphasis on ow patterns and heat transfer near the tube wall.
Guardo et al. (2004, 2005) have performed similar calculations for pressure drop
and wall heat transfer rates. Our most recent work has extended ow and heat
transfer simulations to include source or sink terms in the solid particles, to
emulate the energetic effects of reaction (Dixon et al., 2003; Nijemeisland et al.,
2004). Several of these contributions will be discussed in more detail below.
The focus of the remainder of this chapter is on interstitial ow simulation by
nite volume or nite element methods. These allow simulations at higher ow
rates through turbulence models, and the inclusion of chemical reactions and
heat transfer. In particular, the conjugate heat transfer problem of conduction
inside the catalyst particles can be addressed with this method.
II. Principles of CFD for Packed-Tube Flow Simulation
A. CFD BASICS AND TURBULENCE MODELING
CFD may be loosely thought of as computational methods applied to the
study of quantities that ow. This would include both methods that solve
differential equations and nite automata methods that simulate the motion of
uid particles. We shall include both of these in our discussions of the appli-
cations of CFD to packed-tube simulation in Sections III and IV. For our
purposes in the present section, we consider CFD to imply the numerical
solution of the NavierStokes momentum equations and the energy and species
balances. The differential forms of these balances are solved over a large
number of control volumes. These small control volumes when properly com-
bined form the entire ow geometry. The size and number of control volumes
(mesh density) are user determined and together with the chosen discretization
will inuence the accuracy of the solutions. After boundary conditions have
been implemented, the ow and energy balances are solved numerically; an
iteration process decreases the error in the solution until a satisfactory result has
been reached.
Commercially available CFD codes use one of the three basic spatial disc-
retization methods: nite differences (FD), nite volumes (FV), or nite ele-
ments (FE). Earlier CFD codes used FD or FV methods and have been used in
stress and ow problems. The major disadvantage of the FD method is that it is
limited to structured grids, which are hard to apply to complex geometries and
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 315
are mostly used for stress calculations in beams, etc. A 3D structured grid
results in a grid with all rectangular elements. The rectangular elements can
undergo limited deformation to t the geometry, but the adaptability of the grid
is limited.
The FV and FE methods support both structured and unstructured grids, and
therefore can be applied to a more complex geometry. An unstructured
grid is a 2D structure of triangular cells or a 3D structure of tetrahedral
cells, which is interpolated from user-dened node distributions on the surface
edges or a triangular surface mesh, respectively. The interpolation part of the
creation process is less directly inuenced by the user in an unstructured mesh
than in a structured mesh because of the random nature of the unstructured
interpolation process. This aspect does however allow the mesh to adapt
more easily to a complex geometry. The FE method is in general more accurate
than the FV method, but the FV method uses a continuity balance per con-
trol volume, resulting in a more accurate mass balance. FV methods are more
appropriate for ow situation, whereas FE methods are used more in stress
and conduction calculations, where satisfying the local continuity is of less
importance.
The equations for both laminar and turbulent ows, and the nite volume
methods used to solve them, have been presented extensively in the literature
(Patankar, 1980; Mathur and Murthy, 1997; Ranade, 2002; Fluent, 2003). The
following summary focuses on aspects of particular concern for simulation of
packed tubes and also those options chosen for our own work.
1. Navier Stokes Equations
The general equation used for conservation of mass (the continuity equation)
may be written as follows:
@r
@t

@ ru
i

@x
i
S
m
(1)
The source term S
m
contains the mass added through phase changes or user-
dened sources. In general, and in the simulations described here, the source
term was equal to zero.
The equation for conservation of momentum in direction i and in a non-
accelerating reference frame is given by
@ru
i

@t

@ru
i
u
j

@x
j

@p
@x
i

@t
ij
@x
j
rg
i
F
i
(2)
In this balance p is the static pressure, t
ij
is the stress tensor, and rg
i
is the
gravitational body force. F
i
is an external body forces component; it can include
forces from interaction between phases, centrifugal forces, Coriolis forces, and
ANTHONY G. DIXON ET AL. 316
user-dened sources. For single-phase ow through packed tubes it is usually
zero. The stress tensor t
ij
for a Newtonian uid is dened by
t
ij
m
@u
i
@x
j

@u
j
@x
i
!

2
3
m
@u
l
@x
l
d
ij
(3)
Here m is the molecular viscosity; the second term on the right-hand side of
the equation is the effect of volume dilation.
2. RANS and the Standard k e Turbulence Model
A time-dependent solution of the NavierStokes equations for high-Reynolds-
number turbulent ows in complex geometries is currently beyond our com-
putational capabilities. Two methods have been developed to transform the
NavierStokes equations so that the small-scale turbulent uctuations do not
have to be directly simulated. These are Reynolds averaging (RANS) and l-
tering or Large-Eddy simulation (LES). Both methods introduce additional
terms in the governing equations that need to be modeled in order to achieve
closure. LES has not yet been applied to packed-tube modeling to any signi-
cant extent.
With RANS the solution variables in the NavierStokes equations are
decomposed into mean, u
i
, and uctuating u
0
i
components, and integrated over
an interval of time large compared to the small-scale uctuations. When this is
applied to the standard NavierStokes equations (Eqs. (1)(3)), the result is
@ru
i

@t

@ru
i
u
j

@x
j

@p
@x
i

@
@x
j
m
@u
i
@x
j

@u
j
@x
i

2
3
m
@u
l
@x
l
!

@ru
0
i
u
0
j

@x
j
4
The velocities and other solution variables are now represented by Reynolds-
averaged values, and the effects of turbulence are represented by the Reynolds
stresses, ru
0
i
u
0
j
that are modeled by the Boussinesq hypothesis:
ru
0
i
u
0
j
m
t
@u
i
@x
j

@u
j
@x
i

2
3
rk m
t
@u
i
@x
i

d
ij
(5)
The k e turbulence model was developed and described by Launder and
Spalding (1972). The turbulent viscosity, m
t
, is dened in terms of the turbulent
kinetic energy, k, and its rate of dissipation, e.
m
t
rC
m
k
2

(6)
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 317
The turbulent kinetic energy and its dissipation rate are obtained from the
adapted transport equations.
@rk
@t

@ru
i
k
@x
i

@
@x
i
m
m
t
s
k

@k
@x
i
!
G
k
G
b
r (7)
@r
@t

@ru
i

@x
i

@
@x
i
m
m
t
s


@
@x
i
!
C
1

k
G
k
1 C
3
G
b

C
2
r

2
k
(8)
In these equations, G
k
is the generation of turbulent kinetic energy, k, due to
turbulent stress, and is dened by
G
k
ru
0
i
u
0
j
@u
j
@x
i
(9)
G
b
is the generation of turbulent kinetic energy, k, due to buoyancy,
G
b
bg
i
m
t
Pr
t
@T
@x
i
(10)
Here, Pr
t
is the turbulent Prandtl number for temperature or enthalpy and b
is the thermal expansion coefcient,
b
1
r
@r
@T

p
(11)
The default values of the constants C
1e
1.44, C
2e
1.92, C
m
0.09,
s
k
1.0, s
e
1.3, and Pr
t
0.85 have been established from experimental
work with air and water, and have been found to work well for a wide range of
wall-bounded and free shear ows (Launder and Spalding, 1972).
In a system with both heat and mass transfer, an extra turbulent factor, k
t
, is
included which is derived from an adapted energy equation, as were e and k. The
turbulent heat transfer is dictated by turbulent viscosity, m
t
, and the turbulent
Prandtl number, Pr
t
. Other effects that can be included in the turbulent model
are buoyancy and compressibility.
The energy equation is solved in the form of a transport equation for static
temperature. The temperature equation is obtained from the enthalpy equation,
by taking the temperature as a dependent variable. The enthalpy equation is
dened as,
@rh
@t

@ru
i
h
@x
i

@
@x
i
l l
t

@T
@x
i

@
P
j
h
j
J
j
@x
i

Dp
Dt
t
ik

eff
@u
i
@x
k
S
h
(12)
ANTHONY G. DIXON ET AL. 318
In this equation S
h
includes heat of chemical reaction, any interphase
exchange of heat, and any other user-dened volumetric heat sources. l
t
is
dened as the thermal conductivity due to turbulent transport, and is obtained
from the turbulent Prandtl number
l
t

c
p
m
t
Pr
t
(13)
The enthalpy h is dened as
h
X
j
Y
j
h
j
(14)
where Y
j
is the mass fraction of species j and,
h
j

Z
T
T
ref
c
p;j
dT (15)
For problems involving gradients in chemical species, the convection-diffusion
equations for the species are also solved, usually for N1 species with the Nth
species obtained by forcing the mass fractions to sum to unity. Turbulence can
be described by a turbulent diffusivity and a turbulent Schmidt number, Sc
t
,
analogous to the heat transfer case.
3. Alternative Turbulence Models
The Reynolds-averaged approach is widely used for engineering calculations,
and typically includes models such as SpalartAllmaras, k e and its variants,
ko, and the Reynolds stress model (RSM). The Boussinesq hypothesis, which
assumes m
t
to be an isotropic scalar quantity, is used in the SpalartAllmaras
model, the k e models, and the k o models. The advantage of this approach is
the relatively low computational cost associated with the computation of the
turbulent viscosity, m
t
. For the SpalartAllmaras model, one additional trans-
port equation representing turbulent viscosity is solved. In the case of the k e
and ko models, two additional transport equations for the turbulence kinetic
energy, k, and either the turbulence dissipation rate, e, or the specic dissipation
rate, o, are solved, and m
t
is computed as a function of k and either e or o.
Alternatively, in the RSM approach, transport equations can be solved for each
of the terms in the Reynolds stress tensor. An additional scale-determining
equation (usually for e) is also required. This means that seven additional
transport equations must be solved in 3D ows.
Our group has made extensive use of the RNG k e model (Nijemeisland and
Dixon, 2004), which is derived from the instantaneous NavierStokes equations
using the Renormalization Group method (Yakhot and Orszag, 1986) as
opposed to the standard k e model, which is based on Reynolds averaging. The
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 319
major differences, in application, from the standard k e model are different
empirical constants in the k and e balances and extra terms in the turbulent
dissipation balance (e). The Renormalization group methods are a general
methodology of model building based on the stepwise coarsening of a problem.
The main idea is that the RNG theory is applicable to scale-invariant pheno-
mena that do not have externally imposed characteristic length and time scales.
In the case of turbulence, the RNG theory is applicable to the small-scale
eddies, which are independent of the larger scale phenomena that create them.
The RNG theory as applied to turbulence reduces the Reynolds number to an
effective Reynolds number (Re
eff
) by increasing an effective viscosity (m
eff
).
Through this process the small-scale eddies are eliminated, which reduces com-
putational demand considerably. The new equation for the variation of the
effective viscosity is as follows:
m
eff
m
mol
1
3A
4m
3
mol

4
d

!
1=3
!
d
(16)
where A is a constant derived by the RNG theory, is the eddy length scale, and

d
is the Kolmogorov dissipation scale. So in this case when the eddy length
scale is the Kolmogorov scale, the effective viscosity is the molecular viscosity.
This equation then gives the interpolation formula for m
eff
() between the mole-
cular viscosity m
mol
valid at dissipation scales and the high Reynolds number
limit L )
d
.
Using the denition for the turbulent viscosity (m
t
m
eff
m
mol
), which gives
a result similar to the standard k e model with only a small difference in the
modeling constant, the effective viscosity is now dened as a function of k and e
in Eq. (16) in algebraic form.
m
eff
m
mol
1

C
m
m
mol
s
k

p
" #
2
(17)
The differential form of this equation is used in calculating the effective
viscosity in the RNG k e model. This method allows varying the effective
viscosity with the effective Reynolds number to accurately extend the model to
low-Reynolds-number and near-wall ows.
The transport equations for the turbulent kinetic energy, k, and the turbu-
lence dissipation, e, in the RNG k e model are again dened similar to the
standard k e model, now utilizing the effective viscosity dened through the
RNG theory. The major difference in the RNG k e model from the standard
k e model can be found in the e balance where a new source term appears,
which is a function of both k and e. The new term in the RNG k e model makes
the turbulence in this model sensitive to the mean rate of strain. The result is a
model that responds to the effect of strain and the effect of streamline curvature,
ANTHONY G. DIXON ET AL. 320
a feature that is nonexistent in the standard k e model. The inclusion of this
effect makes the RNG k e model more suitable for complex ows.
The RNG model provides its own energy balance, which is based on the
energy balance of the standard k e model with similar changes as for the k and e
balances. The RNG k e model energy balance is dened as a transport equation
for enthalpy. There are four contributions to the total change in enthalpy: the
temperature gradient, the total pressure differential, the internal stress, and the
source term, including contributions from reaction, etc. In the traditional tur-
bulent heat transfer model, the Prandtl number is xed and user-dened; the
RNG model treats it as a variable dependent on the turbulent viscosity. It was
found experimentally that the turbulent Prandtl number is indeed a function of
the molecular Prandtl number and the viscosity (Kays, 1994).
Guardo et al. (2004, 2005) have compared various turbulence models for a
four-layer packed tube with N 3.92 containing 44 spheres. In their rst report
(Guardo et al., 2004) they compared laminar, standard k e, and Spalart
Allmaras models, and in the follow-up work (Guardo et al., 2005) they extended
this to cover the SpalartAllmaras, standard k e, RNG k e, and realizable k e
and k o models. Based on comparisons of computed pressure drop to the
predictions of the Ergun equation, they conclude that the SpalartAllmaras
equation is preferable. This is explained by the fact that it is formulated to be
valid all the way to solid surfaces, thus avoiding problems with wall functions
(discussed below). Our comparisons of the standard k e, RNG k e, and RSM
models (Nijemeisland and Dixon, 2001) showed that there were no signicant
differences in the results. Similar results for their geometry were found by Calis
et al. (2001). The RNG k e model was chosen in our work because it deals
better with ow with high streamline curvature and high strain rates, such as
would be expected in packed tubes.
4. Wall Functions
Turbulent ows in packed tubes are strongly inuenced by the solid surfaces,
both the tube wall and the surfaces of the packing. Collectively, solid surfaces
are referred to as walls in the CFD literature, and in this section we will
continue that tradition. Besides the no-slip boundary condition on the velocity
components that has to be satised, the turbulence is also changed by the
presence of the wall. Very close to the wall, the tangential velocity uctuations
are reduced by viscous damping and the normal uctuations are reduced by
kinematic blocking. In the outer part of the near-wall region, in contrast, tur-
bulence is increased by the production of turbulence kinetic energy due to the
large gradients in mean velocity.
The near-wall region is conceptually subdivided into three layers, based on
experimental evidence. The innermost layer is the viscous sublayer in which the
ow is almost laminar, and the molecular viscosity plays a dominant role.
The outer layer is considered to be fully turbulent. The buffer layer lies between
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 321
the viscous sublayer and the fully turbulent layer, and the effects of molecular
viscosity and turbulence are equally important. To numerically resolve a so-
lution in the sublayer requires a very ne mesh, since the sublayer is thin and
gradients there are large. Models that are modied to enable the viscosity-
affected region to be resolved with a mesh all the way to the wall, including the
viscous sublayer such as the SpalartAllmaras and k o models, are termed low-
Reynolds number models.
To save computational effort, high-Reynolds number models, such as k e and
its variants, are coupled with an approach in which the viscosity-affected inner
region (viscous sublayer and buffer layer) are not resolved. Instead, semiempiri-
cal formulas called wall functions are used to bridge the viscosity-affected
region between the wall and the fully turbulent region. The two approaches to
the sublayer problem are depicted schematically in Fig. 2 (Fluent, 2003).
The standard wall function (Launder and Spalding, 1974) has been widely
used for industrial ows. The wall function is based on the assumption that the
velocity obeys the log law-of-the-wall
U


1
k
lnEy

(18)
where
U


U
P
C
1=4
m
k
1=2
P
t
w
=r
(19)
Wall Function Approach Near-Wall Model Approach
buffer &
sublayer
t
u
r
b
u
l
e
n
t

c
o
r
e
wall functions.
The viscosity-affected region is not The near-wall region is resolved
used.
High-Re turbulence models can be
?
resolved, instead is bridged by the all the way down to the wall.
throughout the near-wall region.
The turbulence models ought to be valid
FIG. 2. Near-wall treatments (reproduced from Fluent Inc., Version 6.1 Manual, 2003, by per-
mission).
ANTHONY G. DIXON ET AL. 322
y


C
1=4
m
k
1=2
P
y
P
n
(20)
and k and E are universal constants, and U
P
is the mean velocity at P, the
centroid of the cell next to the wall, and y
P
is the distance of point P from the
wall. We shall follow the original reference and present the wall functions in
terms of y

and U

, although the usual notation in the turbulence eld is to use


y

t
w
=r
p
y
P
n
and U


U
P

t
w
=r
p
:
It is important to place the rst near-wall grid node far enough away from the
wall at y
P
to be in the fully turbulent inner region, where the log law-of-the-wall
is valid. This usually means that we need y

43060 for the wall-adjacent cells,


for the use of wall functions to be valid. If the rst mesh point is unavoidably
located in the viscous sublayer, then one simple approach (Fluent, 2003) is to
extend the log-law region down to y

11.225 and to apply the laminar


stressstrain relationship: U

for y

o11.225. Results from near-wall


meshes that are very ne using wall functions are not reliable.
The heat ux to the wall and the wall temperature are related through a wall
function
T
w
T
P
rc
p
C
0:25
m
k
0:5
P
_ q
00
w
Pr
t
1
k
lnEy

1
2
rPr
C
0:25
m
k
0:5
P
_ q
00
w
Pr
t
U
2
P
Pr Pr
t
U
2
c

(21)
where P can be computed using (Launder and Spalding, 1974)
P
p=4
sinp=4
A
k

0:5
Pr
Pr
t
1

Pr
t
Pr

0:24
(22)
where T
P
is the temperature at the cell adjacent to the wall, T
w
is the temper-
ature at the wall, Pr
t
is the turbulent Prandtl number, U
c
is the mean velocity
magnitude at the edge of the thermal conduction layer, and A, k and E are
universal constants. An analogous approach is used for species transport.
The standard wall function is of limited applicability, being restricted to cases
of near-wall turbulence in local equilibrium. Especially the constant shear stress
and the local equilibrium assumptions restrict the universality of the standard
wall functions. The local equilibrium assumption states that the turbulence
kinetic energy production and dissipation are equal in the wall-bounded con-
trol volumes. In cases where there is a strong pressure gradient near the wall
(increased shear stress) or the ow does not satisfy the local equilibrium con-
dition an alternate model, the nonequilibrium model, is recommended (Kim and
Choudhury, 1995). In the nonequilibrium wall function the heat transfer pro-
cedure remains exactly the same, but the mean velocity is made more sensitive to
pressure gradient effects.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 323
Apart from renements to the standard wall function approach, such as the
nonequilibrium model, there has been little progress in the 30 years since its rst
development. Craft et al. (2002) have suggested that this is because those
who work on fundamental phenomena in the near-wall region regard any wall
functions as a basically inadequate approach, whereas for industrial applica-
tions for which resolution of the boundary layer is usually impractical, the more
advanced wall functions have not given clearly better results than the simpler
original formulation. More recently, there has been a renewed interest in develo-
ping improved wall functions, based on sub-grid models of the near-wall cell
(Craft et al., 2002) or on analytical integration of simplied boundary-layer
equations to get proles to be used in the near-wall cell (Utyuzhnikov, 2005).
The development of improved wall function treatments, especially for situations
where the near-wall node must be located in the buffer region, will be of great
importance to the application of CFD to packed tubes, as will be further dis-
cussed below in the context of mesh generation. An enhanced wall function
option that uses blending functions to obtain a single equation valid for all three
near-wall layers has been developed in Fluent, but has yet to be extensively
tested for packed-tube ows.
5. Finite Volume Methods and Codes
There is considerable difference of opinion in the CFD research community
regarding the use of commercial CFD codes versus the development of in-house
code. The main advantage of the latter is that the user has complete control of
the code, and can modify it as he/she sees t. This approach is therefore es-
sential for those whose research is directed toward the development of improved
turbulence models, wall functions, or numerical algorithms. The downside of
the in-house code is that it is usually restricted to specialized, if not simple,
computational domains, and that post-processing and visualization facilities are
usually primitive, in the absence of a third-party post-processing package. When
the research is directed more toward the use of CFD methods to obtain insight
or detailed information about a complex geometry such as a packed tube, the
mesh generation and post-processing facilities available in commercial codes
can be invaluable. Some of the commercial codes that have been used for
packed-tube simulations are those by Fluent Inc. (FLUENT, FIDAP), ANSYS,
Inc. (CFX), and Computational dynamics, Ltd. (Star-CD). Other codes that are
often used originate in the governmental or academic sectors, e.g., CFDLIB
from Sandia National Labs and PHOENICS, which is available as both share-
ware and commercially from Simuserve Ltd. CFD codes, can be modied to
include extra terms, such as the Ergun terms for porous regions (Tierney et al.,
1998). Some have interfaces for user-dened code (e.g., FLUENT) while others
make the original source code available (e.g., CFDLIB).
The two main approaches to CFD for packed-tube simulations have been
LBM and FV methods. The LBM codes are in-house codes; the methods behind
ANTHONY G. DIXON ET AL. 324
them are well described in Chen and Doolen (1998) and are beyond the scope of
this chapter. FV methods that directly discretized the NavierStokes equations
have been extensively described by Patankar (1980), Mathur and Murthy
(1997), and Ranade (2002). Briey, the governing equations are integrated over
control volumes to obtain discrete algebraic equations that conserve the quan-
tity on the control volume. Variables are stored on a collocated grid with a local
pressure gradient correction. The pressurevelocity coupling is achieved by the
SIMPLE or the SIMPLEC scheme, and the equations are stabilized by under-
relaxation. Second-order centered differences are used for spatial discretization
of the viscous terms, while interpolation of face values from cell center values is
done by rst- or second-order upwind differences for the convective terms. The
equations are solved iteratively, and either the energy equation can be a part of
the main iteration in a coupled calculation or the ow eld can be calculated
rst and then substituted into the energy equation in a segregated calculation.
The solution of the large sets of algebraic equations is carried out by multigrid
schemes, to accelerate convergence by computing corrections on a series of
coarse grid levels. This is particularly important for unstructured meshes, as the
line iterative methods used for structured meshes are unavailable. Multigrid
methods use the property that on coarse meshes the global error can be reduced
quickly, while on ne meshes the local error can be reduced by ne-grid
relaxation schemes. Further details of the computational methods may be found
in standard references, and are not repeated here as our focus is on the use and
specialization of CFD methods for packed tubes, and the issues arising from
this application area.
B. PACKED BED CFD MODEL DEVELOPMENT
To create a useful CFD simulation the model geometry needs to be dened
and the proper boundary conditions applied. Dening the geometry for a CFD
simulation of a packed tube implies being able to specify the exact position and,
for nonspherical particles, orientation of every particle in the bed. This is not
an easy task. Our experience with different types of experimental approaches
has convinced us that they are all too inaccurate for use with CFD models.
This leads to the conclusion that the tube packing must either be computer-
generated or be highly structured so that the particle positions can be calculated
analytically.
Some groups have worked with models of unit cells of periodic repeating
packing, such as cubical or rhombohedral, which are representative of the bed
far from the containing walls (Tobis , 2000; Gunjal et al., 2005). Others have
developed models of complete tube cross-sections in relatively short-axial seg-
ments of tubes with low enough N to be completely structured, for example
N 2 (Nijemeisland and Dixon, 2001) and 1 rN r2 in a square cross-section
channel (Calis et al., 2001). Several groups used computer-generated packings,
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 325
at N 7 (Esterl et al., 1998) and at N 2.42 and N 4 in square channels
(Calis et al., 2001) and at N 5.96 and N 7.8 in circular tubes (Magnico,
2003). Others used various ad hoc methods based on a combination of structure
and experimental observation to guide model development for N 4 (Dixon
and Nijemeisland, 2001) and N 3.92 (Guardo et al., 2004, 2005). In this
section we briey review and illustrate some of these approaches, including our
wall-segment (WS) approach that aims to retain the essential near-wall phe-
nomena in a reduced-size model.
1. Computer-generated Packed Tubes
The geometric complexity of random-xed bed structures has usually been
handled by a statistical description of the geometry, in terms of void fraction
proles and pore size distributions. This approach gives a practical way to
implement bed structure into the ow model by retaining the statistical char-
acteristics of the void space without having to introduce the real void structure,
since that is complicated and its 3D structure will vary with repacking (Jiang
et al., 2002). For CFD simulation, however, the exact structure of a particular
instance of the bed must be known. For truly random large-N beds, the CFD
results would then have value only if averaged over an ensemble of realizations
of the possible bed structures.
For low-N packed tubes, a different point of view can be taken. The packing
of such tubes is quite far from random, as the ordering of the particles imposed
by the tube walls can extend from two- to ve-particle diameters into the
packing. In addition, the packing of the particles against the tube wall results in
a small number of repeatable structures. Thus, if the focus of the research is the
phenomena in the near-wall region, then CFD simulations in quite-structured
arrays of particles can be regarded as a good representation of the full-size tube.
The advantage of computer generation of such arrays is that we get complete
information on the position and orientation of the particles, with high accuracy.
The drawbacks of the routine use of computer-generated packs are as follows:
several algorithms do not allow for the inuence of containing walls or require
such walls to be planar (rectangular ducts); few algorithms exist for nonspher-
ical particles; it is difcult to generate dense or loose packing as required, and a
particular algorithm usually gives one or the other case; and validation of the
computer results still requires some development. Despite these problems,
packing algorithms are being used to provide geometric models for CFD simu-
lations. Some of these instances are reviewed in this section, along with some of
the more recent advances in computer generation algorithms.
The approaches to computer-generated packings have been classied into
two main classes (Liu and Thompson, 2000): sequential deposition (SD) and
collective rearrangement (CR). Typical of SD are the drop-and-roll algorithms,
in which particles are added one by one to the packing and allowed to move
until they nd a gravitationally stable position. These algorithms usually result
ANTHONY G. DIXON ET AL. 326
in loose packings of porosity e 0.42 or higher. The CR algorithms (sometimes
called Monte Carlo methods) typically begin with a prescribed number of
particles that are moved randomly to either eliminate overlaps if the initial
arrangement allowed them, or decrease voidage if the initial arrangement was
nonoverlapping. A prescribed voidage can often be achieved, but it may be at
some considerable computational cost.
It is also possible to classify the algorithms by whether the container is rec-
tangular or cylindrical, and by whether the particles are spherical or nonspheri-
cal. One of the earliest algorithms that has been used for CFD geometry
generation is an SD method for spheres in rectangular ducts (Chan and Ng,
1986), which was used by Esterl et al. (1998) and Debus et al. (1998). Soppe
(1990) used a hybrid method, combining the raining technique and Monte Carlo
rearrangement, again for spheres in a rectangular duct. This method was
adapted by Freund et al. (2003) on a 3D equidistant orthogonal lattice to give
packings of spheres with 1.1 oN o20.3. Reyes and Iglesia (1991) also used an
SD algorithm for slow settling of hard spheres, but extended the methodology
to cylindrical container walls and arbitrary-size distributions of spheres. Nolan
and Kavanagh (1992) developed a CR algorithm for spheres in a cylinder,
and then extended it to nonspherical particles (Nolan and Kavanagh, 1995) by
representing arbitrary shapes as assemblies of component spheres with varying
size distributions. Spedding and Spencer (1995) used SD for spherical packings
in cylinders with no edge effects, since N was equal to 50 in their simulations of
liquid rivulet ow over spheres.
Periodic boundary conditions for the walls were used by Yang et al. (1996)
who wanted an efcient method for large numbers of random-size spheres
packed into a hexahedral domain. They used a hybrid algorithm with random
placement to generate the initial distribution of spheres and rearrangement to
stable positions via a drop-and-roll procedure to eliminate overlaps. Liu and
Thompson (2000) studied the extent to which the choice of boundary conditions
affected internal-packing structure for spheres in rectangular ducts in CR
algorithms. They found that even with periodic boundaries, intended to remove
structure from the packings, CR algorithms could induce a self-assembly proc-
ess that resulted in packings with ordered structures.
Mueller (1997) paid particular attention to wall effects in his use of an SD
method for spheres in cylindrical tubes. His algorithm starts with a base layer of
spheres, and then adds each new sphere to one of the two different types of
positions, wall sphere positions and inner sphere positions, maintaining stability
under gravity. He compared different procedures for determining the next site
for addition of a new sphere. Best results were obtained when spheres were
added to the lowest vertical positions of the two types of positions according to
a set percentage. A more recent renement of the method (Mueller, 2005) uses a
packing parameter for the layers above the base layer, and gave good agreement
with experimental voidage proles for 2rNr20. Muellers approach seems to
be the most useful for generating packed tubes of spheres at low N for CFD
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 327
simulations, and it was adopted by Magnico (2003) for his simulations at
N 5.96 and N 7.8. Maier et al. (1998) used a CR technique to generate
spherical bead packs in cylinders with nonpermeable walls and rectilinear
domains with periodic boundaries, which they then used in their own, lattice
Boltzmann simulations.
Nonspherical particle packings present more difculties than spherical ones
and have thus received less attention. Notable efforts have been made by
Nandakumar et al. (1999) and Jia and Williams (2001). Nandakumar et al.
(1999) represented arbitrary packing objects by a 3D polygonal model with a set
of triangular surfaces. Their method did not simulate the dynamics of the
packing process, but rather sought stable equilibrium positions for the objects
using a collision detection algorithm. They obtained good agreement with ex-
perimental data for packings of spheres, Raschig rings and Pall rings. Jia and
Williams (2001) took a completely different, digital approach, in which both
particle shapes and packing space were digitized. Shapes could be represented
by a collection of pixels (in 2D) and particles moved in random directions on
a square lattice, with higher probability of moving down to simulate the
gravity-induced packing process. Extensions to 3D and mixed particle sizes
have been reported by Caulkin et al. (2006). Finally, Theuerkauf et al. (2006)
reported on the use of the discrete element method (DEM) to generate packing
structures in packed beds of spheres for use in CFD modeling, for low-N beds.
DEM is a well-established method that uses particle mechanical properties
to inuence a nal packing structure during a settling/deposition type process.
The particles can be of arbitrary shape, and a soft contact approach is used,
which allows particles to overlap during the process. The authors used a com-
mercial DEM code, and have presented results on packings of cylinders by
DEM in conferences. This appears to be a very promising method, which was
validated by comparison to experimental radial voidage proles.
2. Complete Wall Models of Packed Tubes
In packed beds with regular particles, such as spheres, several packing struc-
tures can be identied. First, a wall-induced structure is regularly found, in
which the spheres arrange themselves along the wall in staggered rings (Mueller,
1997). A dense sphere packing can be identied in the center of the bed, when
this is located far enough from the wall. The third structure is a more random
transition between the very regular wall and center structures. In the literature,
the majority of research is directed toward determining where the transition
from the wall-induced structure to the central structure takes place. This is
important when a radial porosity prole is used in modeling, in order to de-
termine when to use the bed average porosity value. Generally, it is concluded
that the effect of the wall has dissipated at about four particle diameters from
the wall (Benenati and Brosilow, 1962) with only minimal contribution at about
two particle diameters from the wall (Schuster and Vortmeyer, 1981).
ANTHONY G. DIXON ET AL. 328
In beds with very high N, the central structure will dominate throughout the
bed. In low-N beds the wall-induced structure will dominate as the inuence of
the wall on the structure penetrates relatively deep into the bed. As an example,
consider a tube with a diameter of 100, in arbitrary units, and two different
spherical packing materials with diameters of 10 (N 10) and 1 (N 100);
further assume that the wall-induced structure is recognizable as such for four
layers of spheres from the wall. In the N 10 bed the rst four layers at the wall
occupy 96 volume % of the bed. In the N 100 bed, the wall-inuenced region
only makes up 15.6 volume % of the bed.
Once the structure of the bed is understood, it must be implemented as
a computational geometry. For this the exact positions of the spheres must be
established so that they can be placed in the CFD simulation geometry. For the
N 4 bed that we studied (Dixon and Nijemeisland, 2001) there was a regular
structure in the bed, which made it possible to obtain the sphere locations using
geometric relations. The geometry layout was divided into a nine-sphere wall
induced structure and a three-sphere central structure. The nine-sphere wall
layer was redistributed regularly along the wall. The specic tube-to-particle
diameter ratio allowed for an almost exact t of nine spheres along the tube
wall, as shown in Fig. 3.
All the spheres in a layer were supported by two spheres of the layer below
and the column wall, creating a stable packing structure. As the tube-to-particle
diameter ratio of the bed was only four, the entire packing structure was con-
trolled by the inuence of the wall. Nevertheless, the packing was divided into
an immediate wall layer and a central section, but this should not be taken
to imply that the central structure was not wall inuenced. Although a three-
sphere planar structure would almost t within the nine-sphere wall layer, there
was just not enough room at the same axial coordinate. When, however, the
1-1
1-2
1-3
1-4
1-5 1-6
1-7
1-8
1-9
(a)
(b)
FIG. 3. Complete wall N 4 geometries: (a) top view of a layer of spheres, with nine wall spheres
marked, showing, three spheres in the center region; (b) perspective view of the two-layer full bed of
cylinders.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 329
axial coordinate of this three-sphere structure was located between two nine-
sphere wall layers, there was enough room. It was eventually found that the
three-sphere central stacking was supported by the nine-sphere wall layers. The
additional spacing in the central structure results in identical layer spacing of
both the nine-sphere wall layers and the three-sphere center layers, creating a
stable overall structure. The three-sphere central structure was a spiral repetitive
structure in which spheres were supported by only one sphere from the central
structure and two spheres from the wall structure. The spiral nature of the
structure necessitated a larger overall bed, to be able to accommodate periodic
boundaries. The spiral needed six layers for the central layer spheres to return to
their original positions; therefore, the full-bed model had to be made over six
layers, or 72 spheres in total.
More recently, we have created full-bed periodic two-layer models for pack-
ings of cylindrical particles (Taskin et al., 2006). The geometry shown in Fig. 3b
was created specically for comparison of the WS model with cylindrical par-
ticles described in the following section, with the structures identical within a
1201 segment of the bed.
3. Wall Segments
Experience in CFD modeling in packed tubes has taught us that the number
of control volumes increases rapidly with the increasing size of the geometry. To
be able to solve for certain details in the model, such as areas where particles in
the packing touch each other or the tube wall, a high level of detail is necessary
for the required accuracy of the simulation. More particles in a model simu-
lation will lead to more high-detail areas, increasing the computational size of a
complete-wall, full-bed simulation very quickly.
This led us to the conclusion that we needed to formulate a model that
focused on a small number of catalyst particles and their direct neighbors, for an
accurate description of the heat transfer and ow processes taking place on the
local scale. The near-wall region is the most interesting area since this is where
the largest heat transfer gradient occurs. A smaller model was needed with
representative particle and periodic boundaries so that it could be compared
with any near-wall position in the full-bed geometry. The cylindrical shape of
the geometry pointed to the use of a wedge-shaped segment. A WS consisting of
a third of the tube circumference (1201 segment) and two axial layers of particles
would allow for appropriate detail throughout the simulation geometry. The
relationship of the WS to the complete wall geometry is shown in Fig. 3. Within
the WS geometry an appropriately large buffer zone around the area of interest
was used to limit the effects of the boundary conditions on the ow properties in
that area. The smaller WS geometry also gave us the freedom to adjust the mesh
density to the requirements of the different packing situations.
The major added difculty of this approach was the addition of several new
boundaries on which appropriate boundary conditions had to be implemented.
ANTHONY G. DIXON ET AL. 330
The most important parts of creating a segment model are the application of the
physical boundary conditions and the positioning of the internals to allow for
the symmetry and periodic boundary conditions. Without properly applying
boundary conditions the simulation results cannot be compared to full-bed
results, both as a concept and as a validation, since the segment now is not really
a part of a continuous geometry. Our approach was to apply symmetry bound-
aries on the side planes parallel to the main ow direction, thereby mimicking
the circumferential continuation of the bed, and translational periodic bound-
aries on the axial planes, as was done in the full-bed model.
For a WS model, periodic boundary conditions on the top and bottom
boundaries are necessary, and a layout that allows for these conditions as well
as the symmetry conditions is needed. The N 4 full-bed model had nine-
sphere wall layers and three-sphere center layers, as described in the pre-
vious section. The 1201 section would contain three wall-layer spheres and
one center-layer sphere, for a fully symmetrical layout. In the full-bed model
six layers were necessary to create periodic conditions, because of the spiral
structure in the three-sphere center layers. Using six layers in the segment
model, however, would defeat the purpose of creating a segment model (re-
duction of size), so it was chosen to slightly adjust the central layer positions
to allow for periodic boundary conditions over two layers of spheres, as shown
in Fig. 4. There are several sphere segments from the three-sphere central layer
that can be seen in the top view. One of the spheres from the central layer
structure is completely enclosed in the segment model. The two other spheres,
toward the symmetry walls of the segment, are truncated both by the symmetry
wall and the top and bottom of the model. The smaller dashed semicircles
FIG. 4. The 1201 wall-segment (WS) model: top (spheres) and front (cylinders).
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 331
within the sphere-outlines depict the intersection of the spheres with the top and
bottom planes.
To simulate industrial packed tubes, we wanted to extend our WS approach
to cylindrical particles, based on the model created for the N 4 spheres WS. A
1201 segment with two axial layers of particles and translational periodic
boundaries on the column inlet and outlet was used for the cylindrical particle
beds. An experimental study was performed by packing cylindrical particles
with a 1:1 ratio in a bed with a tube-to-cylinder diameter ratio of 4. Dense
particle packing structures were created using the Unidense method. From a
large selection of packing structures, it could be seen that there were distinct
common particle situations near the tube wall, such as the 451 angled particle, or
axially aligned particles. The most common arrangement of cylindrical particles
near the wall of the column proved to be the one with the round surface of the
cylinder against the wall of the column with the particle axis at an angle with
the axis of the column (see Fig. 4). To represent this most common orientation
the base geometry was created with the main particle axis at a 451 angle with the
column axis, as shown. The peripheral particles were then placed in the geo-
metry, based on situations encountered in the physical examples, allowing for
the periodic boundary conditions.
4. Contact Points
In a packed tube there are a large number of contact points where the par-
ticles touch each other and the tube wall. This results in very narrow regions in
the computational domain, which creates problems in meshing the packed tube
geometry. Automatic mesh generation can result in volumes that are extremely
skewed around the contact points, meaning that some of their surfaces can be
much larger than others within the one tetrahedron. Manual mesh generation at
contact points would involve an unrealistic amount of work as the number of
contact points increases rapidly as the number of particles in a tube increase to
more realistic levels.
Magnico (2003) included contact points with a ne structured mesh and
successfully carried out DNS simulations of laminar ow at low to moderate
Re. However, Esterl et al. (1998) and Calis et al. (2001) both reported that gaps
were introduced between the spheres in their geometries. Tobis (2000) was able
to avoid the problem, as his experimental set-up had particles glued together
which he modelled in his CFD simulation. Gunjal et al. (2005) did not report
their treatment of the problem, but from their gures it seems that they, too,
introduced gaps between the spheres. A different approach seems to have been
used by Guardo et al. (2004), who actually increased the sizes of their spheres by
1% so that they overlapped slightly. They encountered no convergence dif-
culties, but the impact of this change on heat transfer and velocity distributions
has not yet been reported. For our simulations, it was not possible to obtain
ANTHONY G. DIXON ET AL. 332
convergence under turbulent ow conditions when actual contact points of the
spheres with each other or the wall were incorporated.
Our approach (Nijemeisland and Dixon, 2001) was to create a gap between
the different entities in the geometry, and then verify that the ow eld was not
disturbed by the gap. To choose the correct gap size a number of different
models were created with a small number of spheres, and differing diameters, to
allow for different gap sizes. The largest sphere size that would permit a tur-
bulent model to be solved was 99.5% of the original sphere size. Other models
were created with sphere sizes of 99%, 97% and 95%. These models were
compared using velocity distribution histograms of uid elements near the
contact points. The uid elements for comparison were selected by limiting the
uid zone to a small area approximately 0.5 cm square around the contact point
(Fig. 5a). In the ve different geometries air was owed through the bed at a
Reynolds number of about 20. Velocity magnitude data were taken from the
different geometries and compared.
It was shown that when the gaps were larger (the 95% and 97% sphere sizes)
the velocity distribution tended to move to higher velocities. Both the 99.5%
and 99% sphere size models showed negligible difference from the touching
models velocity distribution (see Fig. 5b). For our heat transfer studies at
moderate Re it was decided to create models with 99% spheres. This was chosen
because this gap size allowed for easier construction and faster convergence
than the 99.5% spheres model, without any loss in accuracy. Calis et al. (2001)
also chose to reduce their spheres to 99% of the full size. For our later work
under more extreme steam reforming simulation conditions, higher pressure and
higher ow rates, it was necessary to re-evaluate the particle size reduction. It
was found that at these high ow rates a 99% reduction of the spheres would
0
0.2
0.4
0.6
0.8
1
0 0.002 0.004 0.006 0.00
Flow velocity (m/s)
8 0.01 0.012 0.014 0.016 0.018 0.02
f
r
a
c
t
i
o
n
touching spheres(100%)
99.5% sphere diameter
99% sphere diameter
97% sphere diameter
95% sphere diameter
FIG. 5. (a) Selected elements around the sphere contact points for mesh comparison; (b) Velocity
histograms for comparison of the different gap sizes, v
in
0.01 m/s. Reprinted from Chemical
Engineering Journal, Vol. 82, Nijemeisland and Dixon, Comparison of CFD simulations to exper-
iment for convective heat transfer in a gas-solid xed bed, pp. 231246, Copyright (2001), with
permission from Elsevier.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 333
result in considerable amounts of ow in the gaps, creating an unrealistic
ow eld. The gap size was reduced until the no-ow area around the contact
point was re-established. The resulting particle size reduction was 99.5%. These
results demonstrate the necessity to evaluate the approach to contact points on
small reduced models before implementing it on the simulation model.
C. PACKED BED CFD SIMULATION ISSUES
1. Packed Bed Flow Regimes
When running a CFD simulation, a decision must be made as to whether to
use a laminar-ow or a turbulent-ow model. For many ow situations, the
transition from laminar to turbulent ow with increasing ow rate is quite
sharp, for example, at Re 2100 for ow in an empty tube. For ow in a xed
bed, the situation is more complicated, with the laminar to turbulent transition
taking place over a range of Re, which is dependent on the type of packing and
on the position within the bed.
Several studies have looked at questions such as the transition to turbulence,
the level of turbulence intensity in the void space, and the delineation of ow
regimes in xed bed ow. Some of the earliest works investigated turbulence
intensity for gas ow in beds of spheres. Mickley et al. (1965) used hot-wire
anemometry for air ow in a rhombohedral packing at Re 4,780 and 7,010.
They concluded that the dispersion coefcients for the bed were determined by
side-stepping of the uid stream as it passes between particles, as eddy diffu-
sivity in the voids was much smaller than the bed dispersion. They found that
eddy shedding did not occur in the packing voids, and that high local heat
transfer coefcients in spherical packings must be due to turbulence intensity in
the voids, which was as high as 50%, measured behind the seventh layer. Van
der Merwe and Gauvin (1971) also found no eddy shedding except on the rst
bank of spheres in their regular packings over the range 2,500oReo27,000, and
turbulence intensity values around 25%. Turbulence measurements at much
lower Re (501470) were made by Kingston and Nunge (1973) in a rho-
mbohedral array, who found that particle geometry and packing conguration
had a large effect on maximum intensity and transition Re.
Early studies of the transition to turbulence relied on ow visualization
techniques for liquid ow through arrays of spheres. Jolls and Hanratty (1966)
found a transition from steady to unsteady ow in the range 110oReo150 for
ow in a dumped bed of spheres at N 12, and they observed a vigorous
eddying motion that they took to indicate turbulence at Re 300. In regular
beds of spheres, Wegner et al. (1971) found completely steady ow with nine
regions of reverse ow on the surface of the sphere for Re 82, and similar ow
elements but with different sizes in an unsteady ow at Re 200. Dybbs and
Edwards (1984) used laser anemometry and ow visualization to study ow
ANTHONY G. DIXON ET AL. 334
regimes of liquids in hexagonal packings of spheres and rods. They determined
that there are four ow regimes for different ranges of Reynolds number, based
on interstitial or pore velocity Re
i
Re/e:
(1) Re
i
o1: Viscous or creeping ow in which pressure drop is linearly propor-
tional to interstitial velocity and ow is dominated by viscous forces;
(2) 10 r Re
i
r 150: Steady laminar inertial ow in which pressure drop de-
pends nonlinearly on interstitial velocity and and boundary layers in the
pores become pronounced with an inertial core appearing in the pores;
(3) 150 r Re
i
r 300: Unsteady laminar inertial ow in which laminar wake
oscillations appear in the pores and vortices form at around Re
i
250;
(4) Re
i
4300: Highly unsteady ow, chaotic and qualitatively resembling tur-
bulent ow.
These results are in reasonable agreement with the earlier work in terms of
Re, and also the authors make the point that some workers used the term
turbulent to cover the unsteady range.
This general classication of ow regimes appears to have been generally
accepted, with some reservations about the transitions. Lati et al. (1989)
sought to remove the inuence of the observer by using microelectrodes as
electrochemical sensors, to get more precise regime transitions. They later cor-
rected their results (Rode et al., 1994) to include the transfer function of the
electrochemical probe and gave the transition to time-dependent chaotic ow as
110oReo150, but noted that this was not necessarily fully developed turbu-
lence. They made measurements on the tube wall and found that ow was
extremely nonhomogeneous at different spatial locations in a packed tube. This
result was conrmed by Seguin et al. (1998a) in their study of the end of the
stable, laminar regime, which they found to occur at Re 113 inside the bed,
but at Re 135 at the wall. A similar study from the same group (Seguin et al.,
1998b) to determine the transition to the turbulent regime found that the tran-
sition was gradual and not at the same Re at all locations. They obtained
stabilization of the uctuation rate, corresponding to local turbulence, at 90%
of the electrodes for Re4600.
The effects of inertia on ows in both ordered and random arrays of spheres
have been approached by lattice-Boltzmann simulations for small Re (Hill et al.,
2001a) and for moderate Re (Hill et al., 2001b). These studies illustrate the
scaling of drag force with Re, and especially extend the experimental results to
beds of more dilute solid fraction. An interesting point of view was provided by
Niven (2002), who claimed that the transition to nonlinear behavior was not due
to turbulence, but to expansion, contraction, and changes in direction of the
ow, i.e., an increase in inertial forces relative to viscous ones. This viewpoint
has been challenged by Stevenson (2003), who suggests that the transition from
laminar ow to turbulence may occur at much lower Re in a packed tube than
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 335
an empty one, due to the reduced viscous damping of radial velocity compo-
nents caused by ow instabilities.
The detail of the uid mechanical studies of ow in particle arrays stands in
sharp contrast to the highly simplied pictures used for reaction engineering.
From the above survey, it appears that it would be safe to set up CFD sim-
ulations with a steady laminar ow model for Reo100, and to use RANS
turbulence models for Re4600, and simply to avoid the unsteady intermediate
range. A similar point of view was taken by Gunjal et al. (2005), who used a
laminar model up to Re
i
204.74 and turbulent models for Re
i
1,000 and
2,000. They also argued that the onset of unsteady ow is delayed by particles
packing closely together, and that the uid is conned and stabilized by neigh-
boring spheres. Magnico (2003) suggests that the stationary hypothesis can
be used up to larger Re, and regarded as a coarse approximation. At present, the
use of CFD in the transition range can be informed by the intended use of
the simulations. If details of the ow environment of the particles are essential,
then a very ne mesh should be used and unsteady laminar calculations must
be performed. So far this approach has not been taken, to our knowledge.
Alternatively, if improved bed-scale modeling is sought, which incorporates
more realistic uid ow on the scale of particles or larger, then an approximate
steady ow can be obtained by either turbulent or laminar models, and often
both are used and compared to see if the bed scale ow features are the same.
2. Mesh Generation
An important part of CFD modeling is the construction of the mesh, espe-
cially in complex geometries such as xed beds. The mesh strongly affects the
accuracy of the simulation. It has to be chosen with enough detail to describe
the processes accurately and with a degree of coarseness that enables solution
within an acceptable amount of time. When an optimal density has been found,
rening the mesh will increase the model size without displaying more ow
detail. When it is coarsened, the mesh may obscure possibly essential parts of
the ow detail. The mesh determines a large part of creating an acceptable
simulation.
The two main types of mesh are structured and unstructured. In a structured
mesh there are families of grid lines, and grid lines of the same family do
not cross each other and cross each member of the other families only once
(Ranade, 2002). Block-structured grids allow local renement of structured
grids. Unstructured grids are typically made up of tetrahedral cells in 3D, and
can be locally rened anywhere. They are very suitable for complex geometries
such as those found in packed tubes. The Chimera grid in Fig. 6a was used by
Agarwal (1999) to illustrate types of grids for aircraft wings. Similar Chimera
grids were used by Nirschl et al. in simulations of a single sphere between
moving walls, and by Esterl et al. (1998) and Debus et al. (1998) for packed tube
simulations. Each sphere had its own prismatic tted mesh to resolve the
ANTHONY G. DIXON ET AL. 336
boundary layers, which was overlaid on a coarser main mesh for the tube. They
had to work with a very ne mesh to capture the gradients in the laminar ow in
the narrow gaps between the particles.
The grid shown in Fig. 6b was developed by Calis et al. (2001) and consisted
of ve layers of prismatic cells on the walls of the spheres and tube, and un-
structured tetrahedral cells in between. To obtain grid-independent pressure
drops under laminar ow they had to restrict the rst layer of prismatic cells to
be 0.052 mm thick. The thickness then increased for the following four layers.
The tetrahedral cells were 0.4 mm in size. In their later work (Romkes et al.,
2003), which included heat transfer, they had to reduce the size of the rst layer
of prismatic cells by a factor of three under laminar ow.
Other workers have also commented on the need for ne meshes on the wall
surface in laminar ow. Magnico (2003) claimed that the spatial resolution or
cell size must be less than d
p
/40, while for LBM simulations, Zeiser et al. (2002)
recommended cell size between d
p
/30 and d
p
/20 and Freund et al. (2003) used a
grid of size d
p
/30. These cell sizes are in line with the experiences of Tobis (2000)
who used mesh sizes of 1 mm and 2 mm with particles of diameter 38 mm, and
our own experience in which we used average cell sizes in the range 0.51 mm
depending on the simulation conditions, for particles of 25.4 mm diameter.
The preferred range for the thickness of the near-wall cell layer is y
+
430.
However, this is difcult to achieve in packed tubes. The cells sizes are con-
strained by the need to t in between the gaps and/or narrow spaces between
particles, so they cannot be too large. This can result in the y
+
values being too
small for proper application of wall functions. The alternative to use small
enough cells to resolve the boundary layer (y
+
o1) increases the computational
FIG. 6. Examples of types of meshes developed to resolve laminar ow around particles: (a)
Chimera grid. Reprinted, with permission, from the Annual Review of Fluid Mechanics, Volume 31
r1999 by Annual Reviews www.annualreviews.org; (b) Unstructured grid with layers of prismatic
cells on particle surfaces. Reprinted from Chemical Engineering Science, Vol. 56, Calis et al., CFD
Modeling and Experimental Validation of Pressure Drop and Flow Prole in a Novel Structured
Catalytic Reactor Packing, pp. 17131720, Copyright (2001), with permission from Elsevier.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 337
cost unacceptably, at present. For turbulent ow, Calis et al. (2001) employed
the k e model with wall functions on a tetrahedral mesh of size 1.0 mm.They
noted that the velocity varies widely near particle surfaces, so that y
+
can also
vary by a factor of over 40. Their average y
+
was in the range 4.4440 for their
range of Re. Guardo et al. (2004) report 4oy
+
o12, and our simulations result
in similar values for y
+
. Romkes et al. (2003) said that the y
+
criterion for the
use of wall functions was met in their work only at the highest Re, and for their
lowest Re they obtained y
+
o1.
The effects on the computed ow of applying the wall functions outside
their preferred range are not yet clear. Guardo et al. (2005) claim that the
SpalartAllmaras model performs well under such conditions, due to the cou-
pling of wall functions with damping functions. They found good predictions of
pressure drop despite the range of y
+
, and that this quantity was not sensitive to
near-wall cell size. The heat transfer coefcient, in contrast, was very sensitive.
Romkes et al. (2003) found that heat transfer under turbulent ow was only
weakly dependent on y
+
. Clearly this area requires further investigation.
We determined an appropriate mesh density for our simulations by comparing
results from several different mesh sizes (Derkx and Dixon, 1996; Logtenberg
and Dixon, 1998a, 1998b; Logtenberg et al., 1999). An optimal mesh density was
chosen from these previous studies on small particle clusters and additional
studies for the WS simulation geometries used later (Nijemeisland and Dixon,
2004). Typical examples are shown in Fig. 7 for the unstructured grids.
It was shown that there were no differences between the ow solutions
whether a completely ne mesh was used or a locally rened mesh. This mesh
had a node spacing equal to the size of the gap at the sphere contact points at
these locations, gradually grading toward a four times coarser node spacing
near the voids in the geometry. The node distributions on the sphere surface
were also graded from ne near the contact point to coarser away from the
FIG. 7. Typical examples of unstructured meshes: (a) the surface mesh on a number of spheres and
a section of the cylinder and (b) a section of the interior mesh in a plane indicated in part (a).
ANTHONY G. DIXON ET AL. 338
contact points, resulting in a graded mesh both in the uid region as well as in
the solid region, the inside of the particles. The results of the graded meshing
can be seen in Fig. 7b. Mesh gradation was dened using a rstlast-ratio
principle, in the pre-processing GAMBIT package. In this method a changing
node density on an edge in the geometry is created by two user-specied
parameters, an average node spacing, and the size ratio between the rst and the
last node on the edge. The average node spacing in the graded mesh was set to
0.03 inch, and the rstlast-ratio was set to 0.25, resulting in a factor 4 differ-
ence in node spacing between the ne and coarse regions. The indicated settings
create a mesh where the node spacing is 0.015 inch near the contact points (the
nely meshed regions) and 0.06 inch at the coarsely meshed regions.
3. Conjugate Heat Transfer
When conjugate heat transfer through solid particles in the tube is to be
included, the energy balance must be solved in the solid particles, in addition to
the uid ow regions. The energy balance for a solid region is dened by:
@ rh
@t

@
@x
i
l
@T
@x
i

S
h
(23)
The last term S
h
is the volumetric heat source, which may include user-dened
energy source terms. The sensible enthalpy, h, is dened as
h
Z
T
T
ref
c
p
dT (24)
which is consistent with Eq. (14) since c
p

P
j
Y
j
c
p;j
.
The inclusion of heat transfer in the particles means that the solid regions must
also be meshed, in addition to the uid regions. This can considerably increase
the mesh size. If the solid particles are only subjected to simple conduction,
steady-state gradients within them are not likely to be large, and it may be
expected that a coarse grid would sufce. This may not be true if heat effects of
reaction are to be included via the source terms. The use of a ne mesh in the
uid region, especially in the vicinity of the contact points, requires the use of a
matching ne mesh in the solid particles. Even with mesh coarsening toward the
center of the particles, the increase in mesh size may be substantial. One way to
deal with this may be to use nonconformal grids, in which the grid points in the
solid do not have to coincide with the grid points in the uid at the uidsolid
interface. This may be a limited remedy, as curvature in the surfaces of the
particles will limit the extent to which there can be a mismatch between the grids.
4. Boundary Conditions
A no-slip boundary condition is used on all impermeable solid surfaces, but
the choice of boundary conditions for the inlet and outlet of the model is not so
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 339
straightforward. It is generally accepted that when ow enters a structured or
randomly packed xed bed, an area near the entrance of the bed has an un-
developed ow prole as the uid is adjusting from one ow environment to
another. A similar situation is found near the exit of the bed, where a sudden
change in pressure drop is experienced and the ow relaxes before it actually
exits the bed. In a full model, in which an actual bed inlet and outlet were
modeled, a large portion of the bed would need to be modeled just to eliminate
entrance and exit effects from the central portion of the bed that was the main
focus of the study. In a real industrial packed tube, in contrast, the entrance and
exit of the bed are usually small compared to the length, and an active bed may
be preceded or followed by inert packed sections that condition the ow.
In our N 4 models we use translational periodic ow boundaries at both
the ow inlet and ow outlet of the column, in which all variables except
pressure are identical at the periodic planes. The total mass ow is supplied
to the model at the inlet boundary, to give the desired Reynolds number. By
imposing the translational periodic boundaries a generic, developed ow solu-
tion is obtained. Since the translational periodic boundary denes the column
inlet to be identical to the column outlet, there is no ow development in the
bed; a steady-state ow situation is obtained. The periodic boundaries remove
the effects of an entrance or exit effect in the bed. The overall size of the model
can now be greatly reduced. Other groups have used periodic conditions at the
inlet and outlet of a packed tube (Magnico, 2003) or on the ow surfaces of unit
cells (Tobis , 2000; Gunjal et al., 2005), whereas Guardo et al. (2004) set a
constant velocity at their tube inlet and a constant pressure at the outlet. Calis
et al. (2001) and Nijemeisland and Dixon (2001) simulated lengths of empty
tube before and after their packed regions.
Clearly, the pressure itself cannot be periodic even if the geometry of the
model is periodic. Instead, the pressure drop is periodic. The local pressure
gradient is decomposed into two parts: the gradient of a periodic component,
superimposed on the gradient of a linearly varying component. The linearly
varying component of the pressure corresponds to the familiar packed-tube
pressure drop. Its value is not known before the simulation; it must be iterated
on until the mass ow rate that was specied is achieved in the computational
model.
When heat transfer is included in the CFD model for a packed tube, the tube
wall is heated (or cooled for some applications). Either temperature or heat ux
is given at the wall. The gas heats up as it passes through the tube, and the
temperature eld cannot be treated as periodic. Since we want to investigate
the energy penetration into the bed from the wall, it is necessary to reinstate the
generic bed section as a section in a larger bed, dening different inlet and outlet
conditions. For the uid, the inlet temperature should be specied. This could
also be done for the solid particles; however, an alternative is to choose the inlet
heat ux to be zero for the solid regions. Outlet conditions are required for the
solid regions only, and again a zero heat ux condition can be used. For full-bed
ANTHONY G. DIXON ET AL. 340
models this has the disadvantage that a considerable part of the simulated tube
may be atypical of the bed as a whole, due to the uniform inlet temperature
prole. In addition, for short-bed segments, the temperature prole may not
develop enough for the effects of ow throughout the entire model to be felt. To
calculate the development of the temperature prole into the bed, a series of
simulations have to be performed.
To overcome thermal entry effects, the segments may be virtually stacked
with the outlet conditions from one segment that becomes the inlet conditions
for the next downstream section. In this approach, axial conduction cannot
be included, as there is no mechanism for energy to transport from a down-
stream section back to an upstream section. Thus, this method is limited to
reasonably high ow rates for which axial conduction is negligible compared to
the convective ow of enthalpy. At the industrial ow rates simulated, it is a
common practice to neglect axial conduction entirely. The objective, however,
is not to simulate a longer section of bed, but to provide a developed inlet
temperature prole to the test section.
5. Convergence
The discretized equations of the nite volume method are solved through an
iterative process. This can sometimes have difculty converging, especially when
the nonlinear terms play a strong role or when turbulence-related quantities
such as k and e are changing rapidly, such as near a solid surface. To assist in
convergence a relaxation factor can be introduced:
f
new
p
a
f
f
new
p
1 a
f
f
old
p
(25)
The relaxation factor (a
f
) is a multiplier for the change in the solution var-
iable. When this factor is less than unity, the process is called under-relaxed.
When under-relaxed, the iteration process is slower, since the step change is
small, but less likely to diverge. Commercial codes, such as Fluent, will typically
recommend values for the under-relaxation factors that work well with a wide
range of ows. For packed-tube simulations, we have usually needed to reduce
the default values by a small amount, usually 0.2. Some simulations, however,
did not converge until very small values, of the order of 0.1 or lower, were used
and values in this range for pressure and velocity under-relaxation factors have
also been suggested by Gunjal et al. (2005).
To determine when a solution is converged usually involves examining
the residual values. The residual value is a measure of the imbalance in the
discretized equation, summed over all the computational cells in the domain.
Residuals can be obtained for continuity, velocity components, and turbulence
variables. Again, it is common practice to set a cut-off value for the nor-
malized residual values. When the set value is reached, the iteration process
is stopped. Our experience with packed-tube simulations, especially if low
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 341
under-relaxation factors are used, is that the cut-off values should be set very
low, and the iteration continued until the residuals have leveled out. This will
normally be a good deal higher for turbulent ows than for laminar ones, and
ow residuals are higher than those for energy. In addition, it is a good idea to
monitor other measures of convergence besides the residuals, such as pressure
drop and/or an averaged wall shear stress or exit temperature (Guardo et al.,
2004; Gunjal et al., 2005). We have seen apparently level residuals, while the
pressure drop slowly changed, and a substantially different nal ow eld was
eventually obtained, often after several thousand iterations. Following apparent
convergence, it is essential to check both mass and energy balances, as well as
performing grid independence studies and comparing to experimental results, to
have condence in the solution. Just because a simulation has converged, does
not mean that it is necessarily reliable.
Convergence of the discretized system can sometimes be problematic, espe-
cially for some of the turbulence models. In such cases, using a laminar ow
solution as an initial guess for the turbulent ow eld can be helpful. Similarly,
when using options such as nonequilibrium wall functions or enhanced wall
treatment, convergence can be facilitated by rst obtaining a solution using the
k e model with standard wall functions, and then switching to the desired
model.
D. VALIDATION OF CFD SIMULATIONS FOR PACKED BEDS
1. Flow Field Validation
Since the CFD methodology is not specically designed for application in
constrained geometries, such as particle packed beds, it is necessary to verify if
the simulated results are valid. Although the CFD code is based on fundamental
principles of ow and heat transfer, some of the boundary issues are modeled
using empirical data not necessarily appropriate for xed bed applications.
Validation of CFD ow calculations has generally taken one of the two forms.
In the rst, noninvasive velocity measurements inside the packed bed have been
made, and compared to velocities computed from a model of either the entire
experimental bed or a representative part of it. In the second form, computed
pressure drops have been compared to either measured values or established
correlations for pressure drop in xed beds, such as the Ergun equation (Ergun,
1952).
The direct experimental verication of the computed ow eld requires
noninvasive measurements of velocity components inside the packed tube. One
very promising technique for this is magnetic resonance (MR) as described in
a recent review by Gladden (2003). She showed pictures of 3D MR visualization
of axial velocity for ow of water in packings of spheres, and her group has used
MR to connect the 3D structure of a packed bed to the transport phenomena in
ANTHONY G. DIXON ET AL. 342
it (Sederman et al., 1998) and to validate their lattice Boltzmann code for low-
Re simulations of water owing through arrays of spheres (Manz et al., 1999).
They have also combined MR spectroscopy with imaging to get noninvasive
measurements of chemical conversion inside the bed (Yuen et al., 2002).
Maier et al. (1998) also used LBMs to simulate ow through a column
of beads for N 10, which they then compared to the NMR data of Lebon
et al. (1996). They obtained encouraging agreement for the qualitative features
such as negative velocities; however, they saw differences in peak values of
normalized velocity, some of which they attributed to longitudinal dispersion.
Most recently, Gunjal et al. (2005) conducted CFD simulations of laminar ow
through a simple cubic unit cell corresponding to the set-up used by Suekane
et al. (2003) in their MRI studies. Comparisons were made over the range 12.17
rRe
i
r204.74 and showed good agreement for axial components of velocity.
At the highest ow rate inertial ow was dominant, with jet-like ow being
observed in both experiments and simulations. Some discrepancies were seen at
the lowest ow rate, which may have been attributable to experimental dif-
culties in maintaining a constant low ow rate. Overall, the CFD simulations
correctly captured the inertial ow structures, including vortices. Comparisons
were also made between the nite-volume CFD calculations and the LBM
simulations of Hill et al. (2001a), with good results.
Freund et al. (2003) chose to validate their LBM simulations using laser
Doppler anemometry (LDA) measurements. Experiments for two packings
of spheres at N 4 and N 6.15 were used, at a ow rate corresponding to
Re 50. Good agreement was found for the radial proles of axial velocity,
with the simulations reproducing the typical oscillations in velocity in the wall
region. Some discrepancies were seen near the bed center, but the authors noted
that while packing structure of experiment and model were identical near the
wall due to the ordering there, some differences existed in the bed center.
No direct validations have yet been reported for the unsteady laminar or
turbulent ow regimes, due to lack of experimental data at the higher ow rates.
Measurements of pressure drop, however, can provide an indirect means of
checking on the computations at higher ows, although most comparisons have,
in fact, been made at relatively low ow rates. Many groups compare their
results to the original Ergun (1952) equation for pressure drop (Gunjal et al.,
2005, Guardo et al., 2004; 2005), or to later modications or different corre-
lations that take wall effects into account (Esterl et al., 1998; Freund et al., 2003;
Magnico, 2003). In general, good agreement is claimed, especially for the lower
Re range; however, many comparisons are presented on log-log graphs where it
is harder to see differences.
There are few reported comparisons to experimental pressure drop data taken
by the same workers. An exception is Calis et al. (2001) who compared CFD,
the Ergun correlation and experimental data for N 12. They found 10%
error between CFD and experimental friction factors, but the Ergun equation
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 343
over-predicted by 80%, which they attributed to a poor prediction of the tur-
bulent contribution. Their results extended to Re 40,000.
The validation of CFD codes using pressure drop is most reliable when actual
experimental data are taken in equipment identical to the situation that is being
simulated. Existing literature correlations such as the Ergun equation are
known to have shortcomings with respect to wall effects, particle shape effects,
application to ordered beds and validity at high Re. The applicability of liter-
ature correlations to typical CFD simulation geometries needs to be examined
critically before fruitful comparisons can be made.
2. Heat Transfer Validation
In this section a short description of a comparison between experimental and
simulation results for heat transfer is illustrated (Nijemeisland and Dixon,
2001). The experimental set-up used was a single packed tube with a heated wall
as shown in Fig. 8. The packed bed consisted of 44 one-inch diameter spheres.
The column (single tube) in which they were packed had an inner diameter of
two inches. The column consisted of two main parts. The bottom part was an
unheated 6-inch packed nylon tube as a calming section, and the top part of the
column was an 18-inch steam-heated section maintained at a constant wall
temperature. The 44-sphere packed bed lls the entire calming section and part
of the heated section leaving room above the packing for the thermocouple
cross (Fig. 8) for measuring gas temperatures above the bed.
A radial temperature measurement consisted of establishing and recording
a steady-state temperature prole for a combination of a specic bed length,
Reynolds number, and angle of thermocouple cross. A total of four
thermocouple cross
insulation
heated wall
packing
steam in
rotameter
dryer
air in
steam out
TC
P
P
r/r
t
= 0.30
r/r
t
= 0.56
r/r
t
= 0.80
r/r
t
= 0.91
r/r
t
= 0.70
r/r
t
= 0.46
FIG. 8. Experimental setup and detail of the thermocouple cross.
ANTHONY G. DIXON ET AL. 344
thermocouple-cross positions were used for a measurement by rotating the cross
151, 301, and 451 from the initial orientation. By rotating the thermocouple
cross a good spread of data points was ensured, giving a full picture of the
angular spread of the radial temperature prole. Particle Reynolds numbers
were varied from 373 to 1922; bed lengths were varied from 0.132 through
0.42 m.
Fig. 9 shows comparisons of CFD results with experimental data at a
Reynolds number of 986 at three of the different bed depths at which experi-
ments were conducted. The proles are plotted as dimensionless temperature
versus dimensionless radial position. The open symbols represent points from
CFD simulation; the closed symbols represent the points obtained from experi-
ment. It can be seen that the CFD simulation reproduces the magnitude and
trend of the experimental data very well. There is some under-prediction in the
center of the bed; however, the shapes of the proles and the temperature drops
in the vicinity of the wall are very similar to the experimental case. More ex-
tensive comparisons at different Reynolds numbers may be found in the original
reference. This comparison gives condence in interstitial CFD as a tool for
studying heat transfer in packed tubes.
3. Wall-Segment vs. Full-Bed Validation
When simulations are done in a WS model, the results need to be validated
against a full-bed model. The main reason for this is not only to see if the WS
model results are representative for a full bed but also to check that the sym-
metry boundaries, which are relatively close to all parts of the segment model,
do not inuence the solution.
Re = 986
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
r/R

z=0.276m CFD z=0.276m Exp. z=0.132m CFD z=0.132m Exp.


z=0.420m CFD z=0.420m Exp.
FIG. 9. Simulated and experimental radial temperature proles at Re 986.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 345
a. Comparing segment results to full-bed results. A ne mesh was created for
the WS and the full-bed models. For the full bed, the mesh was too large for
the 32-bit computers so it was reduced in size by excluding the solid parts (the
particle internals) from the mesh. This mesh is referred to here as the no-
sphere mesh. The nest possible mesh for the full bed that was created in-
cluding all bed internals will be referred to as the re-mesh. The no-sphere
full-bed ne mesh was 2.6 million control volumes (cv), while the WS reduced
this to only 756,700 cv. The full bed with the ne mesh was over 6 million cv,
which was not implemented at the time (but has since become possible with a
64-bit machine), while the full bed with the coarse mesh was 1.97 million cv. In
the initial comparisons velocity proles were compared for both the full-bed
meshes mentioned above (the no-sphere mesh and the re-mesh) as well as the
WS model. Only velocity prole comparisons were eligible, since the no-sphere
mesh could not give comparable energy solutions.
A section in the full-bed models was isolated that was comparable to the WS
model. The layout of these different sections was identical, except that the WS
model had a two-layer periodicity and the full-bed models had a six-layer
periodicity. To be able to make direct comparisons of velocity proles, several
sample-points needed to be dened. In the three different models seven tan-
gential planes were dened and on each plane three axial positions were dened.
This reduced the data to single radial velocity proles at corresponding posi-
tions in all three models, as shown in Fig. 10, for the WS model. Identical planes
were dened in the full-bed models. Some spheres and sample planes 4 and 5 are
not displayed to improve the visibility of the sample planes and lines. In the
right-hand part of the gure, plane 4 is shown with the axial positions at which
data were taken and compared.
Simulation data were collected and plotted from the three models. Flow
magnitudes were plotted separately for the three different components of the
ow, the axial velocity, v
z
, the radial velocity, v
r
, and the tangential velocity, v
y
for seven planes with three data-lines each. Selected results are shown in Fig. 11
FIG. 10. Comparison section with seven tangential planes and axial prole lines indicated.
ANTHONY G. DIXON ET AL. 346
for the axial and radial velocity components. All simulations were performed
at a particle Reynolds number of 420, under atmospheric conditions with
no temperature gradients. The supercial velocity in the simulations with a
Reynolds number of 420 is 0.58 m/s, and this was used to normalize the different
velocity components.
The velocity prole plots show interruptions in the velocity prole, where the
solid packing was located. In general, the data of the three different cases agreed
very well qualitatively; velocity highs and lows are shown at the same points in
the bed. Quantitatively, the data of the two full-bed models are practically
identical, indicating that the solutions were completely mesh independent.
The data from the WS model in some cases deviated slightly from the full-bed
models. This could be explained by the slightly different layout of the WS
model. Some spheres had to be relocated in the WS model to create a two-layer
periodicity from the six-layer periodicity in the full-bed models. The differences
in velocity magnitudes were mainly found in the transition area between the wall
layers and the center layers. The effect of slightly larger gaps between spheres
from the nine-sphere wall layers and the three-sphere central layers, due to the
sphere relocations, had a noticeable effect on the velocity prole. Differences
were also found in the central layer area where the sphere positions were not
identical.
b. Wall-segment mesh independence. In the previous section, two meshes were
compared for the full-bed models. Similarly for the WS model, two separate
meshes were created: a ne mesh, as reported above, and a coarse mesh with the
same mesh density as the full-bed coarse mesh. In Fig. 12, ow proles of the
ne mesh (labeled WS) are compared to the results from the coarse mesh. Also
included in this comparison are the results from a simulation performed in a
scaled-down version of the ne mesh model. A simulation geometry was created
at 1/8th the size of the original model, to see if the absolute size of the model had
a signicant inuence on the solution of the physical models, including the wall
-1
-0.5
0
0.5
1
1.5
2
2.5
0 0.2 0.4 0.6 0.8 1
r/r
t
v
r
/
v
i
n
wall segment
full bed, fine mesh, no spheres
full bed, coarse mesh
-1
-0.5
0
0.5
1
1.5
2
2.5
0 0.2 0.4 0.6 0.8 1
r/r
t
v
z
/
v
i
n
wall segment
full bed, fine mesh, no spheres
full bed, coarse mesh
FIG. 11. Full-bed and WS mesh comparisons of axial and radial velocity components (at Z3).
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 347
functions, which resolve the wall boundary layers. The results in Fig. 12 show
complete mesh independence of the WS geometry and no effect of the scaling.
III. Low-N Packed Tube Transport and Reaction Using CFD
The broad objective of using CFD to simulate interstitial ow in packed
tubes is to obtain a more fundamental understanding of the phenomena, taking
place within them. Some more specic uses of this approach are (i) to obtain
information (data) for conditions under which experiments are difcult or im-
practical, such as high temperatures or pressures; (ii) to obtain simulated meas-
urements where sensors cannot be located without disturbing the bed packing,
which would invalidate the measurements being sought; and (iii) to examine the
individual contributions to transport and reaction in isolation, or to break down
complex transport processes into their contributing mechanisms, in order to
establish more fundamental correlations for them. In this section we want to
examine the progress so far toward (iii) above, as CFD offers a very promising
means of making further progress on problems that have resisted traditional
approaches.
A. HYDRODYNAMICS AND PRESSURE DROP
As we have seen earlier, it is quite common for CFD pressure drop results to
be compared to well-established empirical correlations of experimental data as
part of the validation of the simulations. Some studies, in contrast, have used
CFD to obtain the contributing mechanisms to pressure drop. One area in
particular where this has been fruitful is structured packings. Corrugated-sheet
structured packings have been studied by Petre et al. (2003) and Larachi et al.
(2003). A layer of this type of structured packing is made of an ensemble of a
large number of Toblerone-like triangular ow channels having identical
-1
-0.5
0
0.5
1
1.5
2
2.5
0 0.2 0.4 0.6 0.8 1
r/r
t
v
z
/
v
i
n
wall segment
wall segment, coarse mesh
1/8th size wall segment
-1
-0.5
0
0.5
1
1.5
2
2.5
0 0.2 0.4 0.6 0.8 1
r/r
t
v
r
/
v
i
n
wall segment, fine mesh
wall segment, coarse mesh
1/8th size wall segment
FIG. 12. Wall-segment mesh comparisons for axial and radial velocity components.
ANTHONY G. DIXON ET AL. 348
cross-sections. These channels are the result of the voids between two adjacent
corrugated sheets that are placed together at opposite orientations.
Structured packing is expensive, so good geometric design of the elements is
important, to optimize performance. Design is still empirical, however, and the
limits of the correlations and design data are uncertain, since the constants for
describing hydrodynamics and mass transfer have been obtained by calibrating
models using laboratory experiments. CFD allows the rigorous uid transport
equations to be solved locally, to better understand the phenomena.
Petre et al. (2003) have presented an approach that combines a mesoscale and
a microscale predictive approach. They identify recurrent mesoscale patterns in
the packing, termed representative elementary units (REU), such as criss-cross-
ing junctions, two-layer transitions, entrance regions, etc. They choose these
REUs by identifying important dissipative phenomena at different parts of the
packing. They then use CFD to simulate each REU over a wide range of
Reynolds number from creeping to turbulent regimes. The next step is the
conjecture that total resistance to ow can be obtained from adding the resis-
tances of individual REUs. The simulations for each REU gave pressure drop at
the microscale, from which they extracted a correlation for its loss coefcient.
These coefcients were then used to predict the additive contribution to the total
bed pressure drop. The results were checked to see that they were not dependent
on the sizes of the REUs, and the bed pressure drops were compared to lit-
erature experimental data with good agreement, to within 1020%. A follow-up
paper (Larachi et al., 2003) used this approach together with CFD to evaluate
proposed geometric design changes of structured packings in terms of energy
loss and pressure drop.
Fig. 13 shows typical results in a plot of pressure drop per unit length against
the gas ow factor F
s
( vr
0.5
). Four types of REUs were identied and
simulated, and correlations were developed. The pressure drop arising from
criss-crossing elements was the largest contribution to pressure drop of the four.
The total pressure drop predicted from the individual contributions was in
excellent agreement with literature data.
Returning to more conventional tube packings, several investigators have
looked at the contributions to pressure drop in arrays of spheres. Esterl et al.
(1998) suggested that CFD could be used to numerically test the assumptions
made in deriving the analytical result that dissipation of energy is 75% due to
elongational effects and 25% due to shear effects. Dhole et al. (2004) inves-
tigated the contributions to drag of power-law uids in distended beds of
spheres at intermediate Re. Since their packings had high void fractions, they
could use an idealized cell model where each sphere was surrounded by an
envelope of uid. Their main focus was to write the drag coefcient as the sum
of a pressure component and a frictional component, and determine how these
behaved as functions of the power law uid index. Gunjal et al. (2005) analyzed
the relative contributions of shear drag and form drag for their unit cell pack-
ings of cubic and rhombohedral spheres. They found that the viscous
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 349
contribution was between 21% and 27% of the total, and was relatively con-
stant in the laminar regime of 12oReo200.
The inuence of bed geometry on frictional resistance has been studied sys-
tematically (Tobis , 2000, 2001, 2002) using turbulence promoters placed between
spheres in a cubic model packing. The experimental and CFD study began with
turbulent air ow through six model packings (Tobis , 2000), which consisted of a
base cubic arrangement of spheres with different turbulence promoters
rectangular bands, triangular prisms, or cylindrical tubesinserted between
them. Wall effects were partially reduced by experimental design, and the re-
maining effects were eliminated by analysis and did not play a part in this work.
The CFD calculations were checked against experiment and different mesh sizes,
and turbulence models were used and gave good agreement with each other. The
author tested the Ergun equation and concluded that the friction factor could
not be correlated using bed porosity e and hydraulic diameter d
h
as the only
measures of bed structure, which goes to explain why the constants in the Ergun
equation scatter between the various experimental investigations in the literature.
A second study (Tobis , 2001) used only the thin bands as inserts, but covered
more arrangements, and different numbers of bands were used in the packing.
The tested packings were identied as being made up of three arrangements of
representative elementary units (REUs). CFD was used to predict the friction
factors for the elementary REUs, and then these micro results were combined to
get the macro pressure drop. It was concluded that local bed anisotropy and/or
clear passages through the bed affect the pressure drop. Complex bed structures
can be handled by the macro-correlation/micro-CFD approach, where local bed
anisotropy is taken into account by micro-CFD, and channeling by formulas to
FIG. 13. Breakdown into the contributing dissipation mechanisms of dry pressure drops in vessels
containing Montz B1-250.45 structured packings. Reprinted from Chemical Engineering Science,
Vol. 58, Petre et al., Pressure Drop Through Structured Packings: Breakdown Into the Contributing
Mechanisms by CFD Modeling, pp. 163177, Copyright (2003), with permission from Elsevier.
ANTHONY G. DIXON ET AL. 350
combine the friction factors for the REUs. Typical results are shown in Fig. 14,
for a horizontal band REU, conguration B2.
The rst part of this gure shows the triangular surface mesh with periodic
conditions on the sides. The middle picture shows a view slightly tilted forward
to show path lines of ow accelerating around the sides of the insert, and
relatively slow ow in the wake region. Corresponding to these the third picture
presents the high amounts of turbulence kinetic energy in the ow trailing from
the sides of the insert. The author noted that the turbulence kinetic energy was
not homogeneous in these complex packings. There was ow circulation in
regions much smaller than the inter-particle space, and ow channeling through
clear passages in the packing that provided evidence that there was not complete
mixing in the packing interstitial space.
In the most recent paper in the series (Tobis , 2002), the method was extended
to chessboard periodic structures to look at different spatial arrangements of
the same packing elements. Experiments showed that pressure drop depended
not only on the resistance of the individual REUs but also on their spatial
arrangement. The experiments were compared to interstitial ow simulations by
CFD, supercial ow modeling by the modied Forchheimer equation, and a
simplied structural method involving macrocorrelations. Although limiting
cases of the macrocorrelations were useful, only the CFD simulations could
reproduce observed experimental results, and further development of the struc-
tural approach to provide less expensive models is needed.
Contributions to pressure drop have also been studied by lattice Boltzmann
simulations. Zeiser et al. (2002) postulated that dissipation of energy was due to
shear forces and deformational strain. The latter mechanism is usually missed
by capillary-based models of pressure drop, such as the Ergun equation, but
may be signicant in packed beds at low Re. For a bed of spheres with N 3,
they found that the dissipation caused by deformation was about 50% of that
FIG. 14. CFD predictions of turbulent ow (meshed REU geometry, air pathlines with spheres
hidden for clarity, contours of turbulent kinetic energy k) for conguration B2. Copyright 2001
From Turbulent Resistance of Complex Bed Structures by J. Tobis . Reproduced by permission of
Taylor and Francis, Inc., http://www.taylorandfrancis.com.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 351
caused by shear, over the range 0.1oReo20. The deformation contribution
decreased as Re increased. Later work by the same group (Freund et al., 2003)
over a higher range of Re concluded that correlations for pressure drop cannot
account for the inuences of local structure on global pressure drop. The au-
thors saw small changes in bed void fraction causing much larger changes in
pressure drop than could be explained by the nonlinear terms, which pointed to
local effects. An extremely thorough investigation of drag force on spheres, and
its dependence on Re, has been performed by Hill et al. (2001a, b), also using
LBMs. These studies examined the effects of uid inertia at small to moderate
Re, on ow in ordered and random arrays of spheres over a wide range of void
fractions.
Although most CFD work has focused on pressure drop, several studies
have also reported radial proles of axial velocity, which are also of interest
for simplied 1D models of xed bed uid ow. Zeiser et al. (2002) found an
oscillating prole, with two peaks at d
p
/8 and d
p
from the wall, for a tube
with N 3. Magnico (2003) found a qualitatively similar result for N 5.96
and N 7.8, but in those results the near-wall peak was lower, which differs
from all other workers. Zeiser et al. (2001) gave fairly qualitative pictures for
N 5 and N 6, while Freund et al. (2003) also found a maximum close to d
p
/
8 for N 4 and N 6.15. Overall, most studies show the near-wall peak at d
p
/8
with a magnitude of about 2.53 times the supercial average velocity, which is
lower than BFD-type models unless an effective viscosity is introduced. Good
agreement has been demonstrated in a few studies with LDA experiments, when
they have been available for the same geometry. The similarity of the velocity
proles near the wall for beds of spheres is not surprising, given the analogous
result for porosity. Toward the bed center, where the wall ordering decreases,
the agreement is not so good. Also, beds of cylinders and other shapes of
particles might be expected to give interesting results that may not be the same
as for spheres.
B. MASS TRANSFER, DISPERSION, AND REACTION
A number of studies have used CFD results to obtain information on dis-
persion or mass transfer in packed tubes, in addition to the usual hydrodynamic
results. Some relatively recent work has also included reaction. This area is so
far not as well developed as the pressure drop and ow elds results of the
previous section; however, some promising rst steps have been taken.
Structured packings were studied using CFD by Van Baten et al. (2001), who
wished to determine the radial liquid residence time distribution of a
KATAPAK-S-like structure for comparison to conventional xed beds for
use in heterogeneously catalyzed reactive distillation processes. Such structured
packings consist of catalyst particles sandwiched between corrugated sheets of
wire gauze that are sealed around the edge. A group of these sandwiches is then
ANTHONY G. DIXON ET AL. 352
bound together. An illustration of the criss-cross structure is given in Fig. 15
below. The Toblerone structure consisting of intersecting and connecting
triangular tubes has many similarities with the packings investigated by Petre et
al. (2003) and Larachi et al. (2003) that are discussed above. The computational
grid used for the simulations is also shown.
In these simulations, the authors used the pseudo-continuum approach for
the catalyst packing inside the criss-cross structure. The body force describing
the resistance to ow offered by the packing was modeled by the Ergun equa-
tion. The equation of continuity for a tracer was included with the equations
of conservation for total mass and momentum, so their simulation tracked
unsteady liquid ow through the illustrated assemblage of triangular tubes.
Liquid entered the system through the top eight tubes and exited through the
bottom eight tubes. At steady state, a pulse tracer was injected and tracer outlet
concentrations were monitored to obtain the RTD curves at different horizontal
positions. From these the axial and radial dispersion coefcients were estimated.
The authors found good agreement between the coefcients estimated from
CFD simulations and those estimated from experiments, and determined that
the radial dispersion coefcient for the KATAPAK-S was about an order of
FIG. 15. The KATAPAK-S structure and its computational grid. Reprinted from Chemical En-
gineering Science, Vol. 56, van Baten et al., Radial and Axial Dispersion of the Liquid Phase within
a KATAPAK-S
s
Structure: Experiments vs. CFD Simulations, pp. 813821, Copyright (2001), with
permission from Elsevier.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 353
magnitude higher than that in co-current down-ow trickle beds, which rep-
resented a desirable feature.
Dispersion in conventional packed beds of spheres has also been simulated by
CFD. Manz et al. (1999) compared lattice Boltzmann simulations to NMR
velocimetry and propagator measurements. Their ows were at 0.4oReo0.77
with 182oPeo350. Over these ranges, they evaluated contributions to the dis-
persion tensor: mechanical dispersion, i.e., dispersion arising from the stochastic
variations of the velocity eld due to the structure of the packing; Taylor dis-
persion, i.e., diffusion of uid molecules across streamlines; and holdup dis-
persion, caused by blocked regions of ow in the packing. At larger length
scales, dispersion was dominated by mechanical dispersion, while at smaller
length scales dispersion was determined by Taylor dispersion. At the largest
length scales, they suggested that holdup could play a signicant role.
Zeiser et al. (2001) also used LBM to examine dispersion. They simulated a
simple cubic packing only one sphere wide, with periodic boundary conditions
perpendicular to the main ow direction to emulate an innite array, and 40
spheres in the main ow direction to allow dispersion to develop. A pulse was
simulated across the transverse plane, which eventually became Gaussian and
allowed extraction of the longitudinal dispersion coefcient. They obtained a
correlation in terms of Pe
1.83
, whereas Taylor dispersion would predict Pe
2
. This
may mean that similarly to Manz et al. (1999), mechanical dispersion begins to
play a role even at quite low Re. In a more recent paper (Freund et al., 2005) the
same group used LBM to obtain ow elds in a simple cubic packing of spheres,
and in a random packing of N 5. They then used a particle tracking algo-
rithm, together with the ow elds, to obtain 3D concentration elds from
which they obtained axial and radial dispersion coefcients. They found good
agreement with experiment for dispersion in the cubic array, for which ow was
relatively simple, but for the more complex ows in the random bed the agree-
ment was less good, although still reasonable compared to the scatter of the
data. Freund et al. (2005) also performed simulations that gave residence time
behaviour, and warned that the residence time and dispersion coefcients ob-
tained from the popular tracer injection technique could depend on the tracer
injection location.
Dispersion in packed tubes with wall effects was part of the CFD study by
Magnico (2003), for N 5.96 and N 7.8, so the author was able to focus on
mass transfer mechanisms near the tube wall. After establishing a steady-state
ow, a Lagrangian approach was used in which particles were followed along
the trajectories, with molecular diffusion suppressed, to single out the connec-
tion between ow and radial mass transport. The results showed the ratio of
longitudinal to transverse dispersion coefcients to be smaller than in the lit-
erature, which may have been connected to the wall effects. The ow structure
near the wall was probed by the tracer technique, and it was observed that there
was a boundary layer near the wall of width about d
p
/4 (at Re
i
7) in which
there was no radial velocity component, so that mass transfer across the layer
ANTHONY G. DIXON ET AL. 354
would be by molecular transport only. At higher ow rates the layer was too
thin to be resolved by the authors structured mesh, but even at Re
i
200 it was
noted that the tracer particles were not easily transported across the region one
sphere diameter thick next to the wall. Maier et al. (2003) also investigated
wall effects on hydrodynamic dispersion using LBM for tubes at higher N, in
the range 1050, but at relatively low Re. They also found that the presence
of the tube wall enhances dispersion, due to the effects on bed structure and
velocity prole.
The inclusion of chemical reaction into CFD packed-tube simulations is a
relatively new development. Thus far, it has been reported only by groups using
LBM approaches; however, there is no reason not to expect similar advances
from groups using nite volume or nite element CFD methods. The study by
Zeiser et al. (2001) also included a simplied geometry for reaction. They simu-
lated the reaction A+B -C on the outer surface of a single square particle on
the axis of a 2D channel (Fig. 16).
The chemical species were treated as passive scalar tracers in the unsteady
LBM equations. The reaction was simulated as being mass-transfer limited at
low Re 166, with diffusivities in the ratios D
A
: D
B
: D
C
1 : 3 : 2. The
concentration elds shown in Fig. 16 are different for each species due to the
different diffusivities. The slow-diffusing species A is transported mainly by
convection and regions of high or low concentration correspond to features of
the ow eld. A more uniform eld is seen for the concentration of faster
FIG. 16. Snapshot of the 2D concentration eld for reactive uids. Reprinted from Chemical
Engineering Science, Vol. 56, Zeiser et al., CFD-Calculation of Flow, Dispersion and Reaction in a
Catalyst Filled Tube by the Lattice Boltzmann Method, pp. 16971704, Copyright (2001), with
permission from Elsevier.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 355
diffusing species B, while species C with intermediate diffusivity has a concen-
tration eld between the other two. The main conclusion was that the inho-
mogeneous ow eld led to a nonuniform concentration distribution around
the particle, in contrast to the usual assumption of uniform surroundings for
catalyst particles in standard reactor models.
A sequel to this study was presented later by the same group (Freund et al.,
2003), this time for the simple rst-order reaction A-B in a cylindrical bed of
spheres with N 5. The reaction was again taken to be mass-transfer limited
and to occur on the surfaces of the catalyst particles, but at a very low ow rate
at Re 6.5. It was found that concentration peaks occurred near the wall at
values close to the inlet value of species A, indicating that channeling was taking
place. There were also local peaks of product concentration that indicated areas
of high reactivity that could give rise to hotspots in practice.
Yuen et al. (2003) used MR visualization techniques to obtain information on
local chemical conversion for the liquid-phase esterication of methanol and
acetic acid in an optically opaque 3D xed bed of catalyst spheres. They used
LBM simulations to obtain a ow eld within the xed bed of catalyst, which
was validated by comparison to ow data also obtained from MR visualization.
Their experimental data showed signicant fractional variations in conversion
in the tube, at the same axial planes, which they were able to relate to hetero-
geneities in the ow eld. They compared local conversion and locally averaged
uid velocity over various length scales. When correlating over the length scale
of the entire bed, the effects of local uid mixing were lost in comparison to
heterogeneities in the macroscopic ow eld. Correlations at the smallest length
scale, the size of the packing (about 600850 mm) showed only a weak link
between ow and conversion, as local mixing caused by the bed structure was
incomplete at this scale. An intermediate scale of 1.5 mm1.5 mm500 mm
gave a correlation between the conversion and velocity data that was well tted
by a model based on the assumption of a kinetically controlled esterication
reaction. A suggested explanation for this result was that over these volumes
effective mixing between uid streamlines had occurred.
C. HEAT TRANSFER
There are only a few research groups that investigate heat transfer in packed
tubes using CFD, as it involves larger models, and the LBMs cannot, at present,
accommodate the energy balance. The added complication of meshing the solid
particles for the conjugate heat transfer problem makes heat transfer CFD
studies more computationally expensive. Most heat transfer CFD work is done
in small sections of packing, representing REUs or periodic elements of the bed.
Lund et al. (1999) reported nite element simulations of conduction between
two contacting spheres without uid ow. They modeled a small separation
between the spheres to allow for interparticle micro-asperity gaps, and then the
ANTHONY G. DIXON ET AL. 356
thermal conductivity of some of the center-most gap gas elements was increased
to that of the solid particles, to include the effect of deformation contact area.
From the CFD simulations they derived a correlation for effective bed thermal
conductivity that included the effects of the micro-asperity uid gap, the
deformation contact diameter, and the thermal conductivities of solid and uid.
They also considered different packing arrangements to obtain the dependence
on bed voidage.
One of the earliest heat transfer problems addressed by CFD is the particle-
to-uid heat transfer coefcient. Many experimental studies had disagreed on
the limiting value of the particle to uid Nusselt number (Nu
p
h
p
d
p
/k
f
) as Re
decreased to zero. The work of Srensen and Stewart (1974) provided denitive
calculations for creeping ow that gave the limiting value for spheres in a cubic
packing, and showed that it should be nonzero. CFD simulations at higher Re
for simple cubic and face-centered cubic arrangements were performed by
Gunjal et al. (2005). Their results for the FCC packing were in reasonable
agreement with literature correlations as far as the magnitude of Nu
p
, but the
trend with Re was somewhat different. The simulations for the SC packing gave
Nu
p
values well below the predicted ones. The authors explained these differ-
ences by the different ow structures in the interstitial space for the different
packing arrangements.
Some investigators have considered the particle-uid heat transfer for a
single particle in an innite medium or in a channel. This is a useful case for the
testing of CFD models and grids, as analytical solutions are available for the
limiting case Re -0. Romkes et al. (2003) simulated particleuid heat transfer
for a single particle in tubes of various sizes intended to approach an innite
domain, and also for composite structured packings (CSP) made up of square
channels packed with spheres to give 1 oN o2.05. They also carried out mass
transfer experiments with CSP using sublimation of naphthalene from test
particles. The CSP structures are designed to reduce pressure drop for packed
bed reactor operations, but this also has the effect of reducing heat and mass
transfer rates between particle and uid. Therefore, it is important to test
existing correlations for these transport coefcients, and develop new ones if
needed.
For the single sphere, Romkes et al. (2003) found that over the range
1oReo10
5
the average heat transfer rate computed by CFD was within 10% of
that predicted by well-established literature correlations. They also found that
the local heat transfer rate varied along the sphere surface and depended on the
angle relative to the stagnation point, in qualitative agreement with prior work.
They presented this dependence for high and low Re and showed ow and heat
transfer maps for various Re. For the CSP simulations, test particles with
higher surface temperatures were simulated, analogously to the mass transfer
experiments. The authors found that consistent results were obtained for differ-
ent locations and numbers of active particles, as long as they were not in the
rst two layers, for which the ow eld had not established periodic behavior, or
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 357
the last layer in the tube, for which exit effects had an inuence. Good agree-
ment with experiment was obtained when simulated Nusselt numbers were
compared to experimental Sherwood numbers, to about 15%. Correlations
were obtained for each packing (N value) separately, and an overall correlation
for N o2.00. The main lesson learned was that low-N beds are special cases, the
usual packed-tube particle-uid transport correlations deviated considerably
from both experiment and CFD simulation.
Our own groups work has focused on wall-bed heat transfer in low-N xed
bed tubes. At the high ow rates of industrial practice in such applications
as partial oxidation and steam reforming, the bulk of the resistance to radial
heat transfer lies in the vicinity of the tube wall; hence, this is a very important
problem. The standard approach to modeling radial heat transfer has been to
conduct experiments in heated or cooled tubes, measure radial temperature
proles at various axial positions in the tube, and t a 2D pseudo-homogeneous
model for temperature, under the assumption of plug ow. The tting para-
meters have been the effective radial thermal conductivity, k
r
, and the apparent
wall heat transfer coefcient, h
w
. The model equations and a discussion of the
limitations of this effective or lumped approach have been presented recently
(Dixon and Nijemeisland, 2001), although there are literally hundreds of papers
presenting variations on the experimental techniques, the types of model tted,
and the theoretical justications. Despite this effort, it is generally accepted that
approaches to wall heat transfer are empirical and in strong disagreement with
each other.
Our initial hopes for the use of CFD for packed-tube simulation were to
generate more accurate and consistent radial temperature proles, free of ex-
perimental artifacts, from which better estimates of k
r
and h
w
could be obtained.
In our earliest work we estimated these parameters in dimensionless form as k
r
/
k
f
and Nu
w
( h
w
d
p
/k
f
), from temperature elds for ow around small clusters
of three particles (Derkx and Dixon, 1996) and eight particles (Logtenberg
and Dixon, 1998a). This work was later extended to include the effects of
temperature-dependent uid properties (Logtenberg and Dixon, 1998b) and to a
cluster of ten contacting particles (Logtenberg et al., 1999). The results from
these studies showed reasonable agreement in magnitude between the para-
meters estimated from the CFD simulations, and values predicted from com-
monly used correlations. Trends with Re were followed reasonably well for k
r
/k
f
but were off for Nu
w
, which could be attributed to the use of small clusters of
particles that could not fully represent the ow patterns of full beds. We would
now regard these efforts as early attempts to check if CFD gave reasonable
results for the bed-scale temperature eld, by comparing the derived heat
transfer parameters to literature values.
With increased computer power, our next step was to construct full beds of
particles: N 2 for validation studies as described above in Section II.D.2, and
N 4 for further investigation of the temperature elds and near-wall transport
processes as described above in Section II.B.2. Some early ow maps and path
ANTHONY G. DIXON ET AL. 358
lines were given in Dixon and Nijemeisland (2001) while temperature elds were
reported in Nijemeisland and Dixon (2004). The intention was to extract radial
temperature proles and t the pseudo-homogeneous model to estimate the heat
transfer parameters as previously. Similar work has been published recently by
Guardo et al. (2004, 2005). However, we soon faced a dilemmaHow should
the CFD-generated radial temperature prole be sampled? The steep increase in
temperature near the wall, which had been far less apparent in cluster geome-
tries, could not be tted by the standard model for the case of a complete N 4
bed. This problem was well illustrated by Von Scala et al. (1999) for the struc-
tured packing KATAPAK-M
s
. Fig. 17 shows an example of this difculty.
In this gure, dimensionless temperature is plotted against dimensionless
radial coordinate. Twelve experimental data points are shown, and curve (a) is
tted by the usual methods to these points. It clearly does not t as well as
we would like either in the bed center or near the wall, due to the limitations of
the underlying model. The intercept at r/R 1 provides the temperature differ-
ence used to dene the wall coefcient h
w
. If an extra point were available closer
to the wall (marked as speculative in Fig. 17), such as might be provided by
CFD simulations, then the t to all points results in curve (b). This clearly
ts the data set even worse than curve (a) and gives a signicantly different
intercept and resulting h
w
. So, the parameter values that are estimated depend
on the location of the measured values. A more consistent approach might be to
drop the measurements closest to the wall, as for curve (c) shown, in which this
has been done for the three experimental points closest to r/R 1. This would
have the merit of at least tting the bed center temperatures well, and the
parameter estimates would have a greatly reduced dependence on the measure-
ment locations. The tted prole would be quite inaccurate near the wall,
FIG. 17. Least squares ts of the radial temperature prole in KATAPAK-M. Reprinted from
Chemical Engineering Science, Vol. 54, von Scala et al., Heat Transfer Measurements and Sim-
ulation of KATAPAK-M Catalyst Supports, pp. 13751381, Copyright (1999), with permission
from Elsevier.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 359
however. A decision would also have to be made as to which measurements to
accept for tting.
From this illustration we can see that the added detail of the radial temper-
ature prole near the wall that could be provided by CFD simulations does
not help in obtaining better estimates for the standard heat transfer parameters.
It also implies that experimental efforts to measure temperatures closer to the
wall are, in fact, counter-productive. Finally, it is clear that the standard model
with plug ow and constant effective transport parameters does not t satis-
factorily to temperature proles in low-N beds. These considerations have led us
to look for improved approaches to near-wall heat transfer.
At this stage it seemed clear that to improve near-wall heat transfer modeling
would require better representation of the near-wall ow eld, and how it was
connected to bed structure and wall heat transfer rates. Our early models of full
beds of spheres at N 4 were too large for our computational capacity when
meshed at the renement that we anticipated to be necessary for the detailed
ow elds that we wanted. We therefore developed the WS approach described
above in Section II.B.3.
In Fig. 18, ow path lines are shown in a perspective view of the 3D WS. By
displaying the path lines in a perspective view, the 3D structure of the eld, and
of the path lines, becomes more apparent. To create a better view of the ow
eld, some particles were removed. For Fig. 18, the particles were released in
the bottom plane of the geometry, and the ow paths are calculated from the
release point. From the path line plot, we see that the diverging ow around the
particle-wall contact points is part of a larger undulating ow through the pores
in the near-wall bed structure. Another ow feature is the wake ow behind the
middle particle in the bottom near-wall layer. It can also be seen that the uid is
transported radially toward the wall in this wake ow.
The second picture in Fig. 18 shows a temperature map for a vertical plane in
the middle of the WS. The tube wall is to the right of the picture, and the scale
has been chosen to emphasize the temperature gradients in the near-wall region.
FIG. 18. Path lines and a temperature map in the WS geometry.
ANTHONY G. DIXON ET AL. 360
Outlines of some of the particles are shown for reference. The boundary layer
next to the wall is clearly visible (lightest shade), while the progression of the
temperature eld through the solid particles can also be seen. The temperature
eld is observed to be 3D, with features that do not depend on the radial
coordinate alone.
When we want to look at the connection between the ow behavior and the
amount of heat that is transferred into the xed bed, the 3D temperature eld
is not the ideal tool. We can look at a contour map of the heat ux through the
wall of the reactor tube. Fig. 19 actually displays a contour map of the global
wall heat transfer coefcient, h
0
, which is dened by q
w
h
0
(T
w
T
0
) where T
0
is
a global reference temperature. So, for constant wall temperature, q
w
and h
0
are
proportional, and their contour maps are similar. The map in Fig. 19 shows the
local heat transfer coefcient at the tube wall and displays a level of detail that
would be hard to obtain from experiment. The features found in the map are the
result of the ow features in the bed and the packing structure of the particles.
From Fig. 19, it is clear that the structured packing near the wall causes a
pattern in the wall heat ux. To indicate the repeating sections lines were added
in Fig. 19. The dotted lines connect the particlewall contact points, the hor-
izontal line connects these contact points of spheres in the same layer, and the
vertical dotted lines connect particlewall contact points from spheres of alter-
nating layers. The section indicated by the solid box is the repeating section
selected for further investigation. The right-hand side of the box is centered at a
sphere-wall contact point and the left-hand side of the box is in between contact
points. The height of the box is identical to one packing-layer height.
For relating the wall heat ux and the near-wall ow patterns quantitatively
the separate pieces of information had to be linked. Detailed information on the
FIG. 19. Map of wall heat transfer coefcient for N 4 bed of spheres.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 361
near-wall ow eld was available in conjunction with an equally detailed map of
the wall heat ux, similar to Fig. 19. Each surface cell in a repetitive section of
the wall was linked directly to a number of uid volumes with similar tangential
and axial coordinates, up to about 5 mm into the uid. In this way we could
probe on a small scale whether the wall heat ux from each surface mesh
element (about 1.3 mm in size) could be linked to the corresponding local ow
features. Unfortunately, it was found that the quantitative comparison of near-
wall ow features with the local wall heat ux by itself did not show enough
distinct features to identify any trends. A number of features of the ow eld
were tried, such as velocity components, vorticity components, helicity, and
velocity component derivatives. No correlations were found for any of these
quantities, suggesting that we cannot relate wall heat ux to local ow eld on
this length scale.
By reducing the ow features to simple component magnitudes, larger scale
patterns such as the ow path were lost. To test whether wall heat ux could
be related to ow patterns on a larger length scale, a conceptual comparison was
made using the near-wall repetitive section with the wall heat ux map, as
shown in Fig. 19. To capture the 3D volume of the near-wall ow features, the
ow eld was simplied to a cartoon and the foremost features of the ow were
identied. When the ow features were compared to the wall heat ux, it could
be seen that the areas of low wall heat ux were located in the parts where the
main through-ow and the wake ow met. The high wall heat ux areas were
mostly located near the sphere-wall contact point and just upstream from that
area. The area just upstream from the contact point had a diverging ow,
consisting of strong axial and tangential components; the wake ow, however,
was also characterized by strong tangential ow, but combined with radial ow.
When the ow features were separated by component, as was done in the
quantitative analysis, these connections were lost. So, analysis at the length scale
of a particle was able to reveal connections between ow and wall heat ux
(Nijemeisland and Dixon, 2004).
The results of the analysis described above have suggested that to lump all
the near-wall heat transfer mechanisms into a wall heat transfer coefcient
idealized at the wall is too simple an approach for low-N beds. One of our
approaches will be based on separating the three main contributions to the extra
resistance in the near-wall region: the viscous boundary layer, the decrease in
stagnant conductivity of the bed due to decreasing solid fraction, and the re-
duction in radial convective transport due to increased channeling and changed
ow patterns near the wall. We intend to use full-bed simulations to generate
temperature elds such as those shown in Fig. 20, for N 4.
The temperature maps shown in Fig. 20 illustrate the development of the
temperature eld as the ow enters a tube heated at the wall. The rst (left-
hand) map shows the initial heating of the gas at the tube entrance. The de-
velopment of the boundary layer near the walls is clear and represents one
contribution to the heat transfer resistance in the wall region. The more rapid
ANTHONY G. DIXON ET AL. 362
penetration of higher temperatures into the bed through the solid particles
due to their higher thermal conductivity is also well shown. In the second (right-
hand) temperature map the arrangement of the spherical particles around
the tube wall is illustrated, and the more developed temperature eld further
down the tube. The more uniform temperatures in the bed center, due to con-
vective heat transfer, are evident, as are the temperature gradients across the
rst layer of particles next to the wall. The proper choice of simulation con-
ditions will allow us to separate the individual contributions due to the separate
heat transfer mechanisms, and our early results in this direction are promising
(Leising, 2005).
IV. Catalyst Design for Steam Reforming Using CFD
A. STEAM REFORMING AND PRINCIPLES OF CATALYST DESIGN
The steam reforming of methane to produce synthesis gas is becoming
a particularly important reaction recently, due to the increased availability of
natural gas as a feedstock. The use of reformed gas has historically been domi-
nated by hydrogen manufacture for commercial use, or for ammonia and
methanol synthesis. Of increasing importance, however, is the manufacture of
liquid fuels from remote or stranded natural gas using Fischer Tropsch chem-
istry, and the generation of hydrogen from natural gas liquid fuels to power
mobile and stationary fuel cells. The syngas generation section of such plants
comprises over 50% of the capital cost (Abbott et al., 1989). There is thus
a strong economic incentive to develop more efcient steam reforming tech-
nology, and a major step in this effort is the optimal design of catalyst particles
(Stitt, 2005).
FIG. 20. Temperature contour maps of respectively the x 0 plane of the rst stage in the N 4
geometry stacking, and the y 0 plane of the fourth stage. Main axial ow direction moves from left
to right in these pictures.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 363
Steam reforming is traditionally carried out in large red furnaces containing
many catalyst-containing tubes. There are several requirements in reforming
that might normally be considered mutually incompatible:

Low pressure drop and thus high voidage and large particles

High surface area for high activity which normally requires small particles

Good radial mixing for heat transfer that requires a large particle

High strength to avoid breakage during handling and charging (Note that no
catalyst is strong enough to resist the stresses of tube contraction during
cooling but that the fracture patterns are important to preventing pressure
drop build up).
The general outcome of this is that the reforming industry uniformly uses
pellets of a length to diameter ratio in the range 0.81.2 with internal holes.
Catalysts ranging from Raschig rings to pellets with 4 to 10 axial holes are now
commonly used. Tube-to-particle diameter ratios also vary but at the entry
region where good heat transfer is essential they will normally be 510. Smaller
catalyst particles tend to be used in the lower portion of the tube where the
reaction activity becomes a factor next to the heat transfer. This is especially
true of top red reformers.
A randomly packed tube, see Fig. 21 for example, is geometrically extremely
complex and thus hard to represent. The randomness of the packing makes it
hard to construct mappings with periodicity.
The impact of the tube-side heat transfer coefcients on the tube wall tem-
perature is shown in Fig. 22: a two-fold improvement in the coefcients facil-
itates approximately a 401C lowering of the tube wall temperature.
Tube wall temperature is an important parameter in the design and operation
of steam reformers. The tubes are exposed to an extreme thermal environment.
Creep of the tube material is inevitable, leading to failure of the tubes, which is
exacerbated if the tube temperature is not adequately controlled. The effects of
tube temperature on the strength of a tube are considered by use of the Larson-
Miller parameter, P (Ridler and Twigg, 1996):
P T
logt K
1000
(26)
where T material temperature [K], t time [h] and K is a material-dependent
constant. The value of this parameter is plotted against the rupture stress of the
material. Fig. 23 shows the Larson-Miller diagram for a typical cast high tem-
perature alloy (K 15), validated by the results of rupture tests at standard
conditions.
A high tube wall temperature also affects the performance of the catalyst.
Higher temperatures lead to increased carbon lay-down on the catalyst and a
resultant loss of catalytic activity, as well as potential catalyst breakage. Both
ANTHONY G. DIXON ET AL. 364
lead to a decrease in the local reaction rate. The effective reduction in the
reaction-originated heat sink may cause the tube to overheat locally or globally.
The effects of excessive temperatures on reformer tubes are in fact quite
dramatic. Fig. 24 shows photographs of reformer tube banks with poor per-
formance and tube over-heating. In Fig. 24a, the ame from the burner is visible
in the top of the photograph. On several of the tubes clear evidence of hot bands
700
740
780
820
860
0 1 2 3 4 5 6 7 8 9 10 11 12
Distance Down Tube [m]
T
u
b
e

W
a
l
l

T
e
m
p

[

C
]
Base case
Base case with
2x heat transfer
Top Fired Reformer
FIG. 22. Impact of tube-side heat transfer coefcients.
FIG. 21. Randomly packed tube packed with reforming catalyst.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 365
and hot patches can be seen. These may be the result of local deactivation, or
catalyst voids (settling) which lead to loss of the local endotherm and tube
failure. In Fig. 24b, a tube can be seen to have overheated over its entire length.
This emphasizes the thermally aggressive environment to which the reformer
tube is exposed.
From the Larson-Miller analysis, it is possible to derive more easily inter-
preted information relating to the effects of sustained high temperature on the
life of a tube. A common rule of thumb is that a tube wall temperature increase
of 201C will shorten a tube life by over 50%: from its design period of 10 years
to less than 5 years. The cost of a typical reformer tube is USD 6 0007 000.
With typical reformer sizes in the order of 300400 tubes and taking on-site
expenditure into account, this puts the cost of a complete re-tube in the range
P (Larson-Miller Parameter)
R
u
p
t
u
r
e

S
t
r
e
s
s

[
M
P
a
]

500
100
50
10
16 18 20 22 24 26
30%
FIG. 23. Larson-Miller diagram summarizing the results of 170 rupture tests on cast 25/20 chrome
nickel alloy (adapted from Ridler and Twigg, 1996).
FIG. 24. Photographs of primary steam reformer tube banks showing high tube wall temperature
features, (a) showing bands and hot patches and (b) showing an entire tube that has overheated.
ANTHONY G. DIXON ET AL. 366
USD 58 million. Avoidance of high tube temperatures, both globally and
locally is therefore at a premium.
The overall effect of catalyst pellet geometry on heat transfer and reformer
performance is shown in the simulation results presented in Table 1. The per-
formance of the traditional Raschig ring (now infrequently used) and a modern
4-hole geometry is compared. The benets of improved catalyst design in terms
of tube wall temperature, methane conversion and pressure drop are self-evident.
B. CFD SIMULATION OF REFORMER TUBE HEAT TRANSFER WITH DIFFERENT
CATALYST PARTICLES
For the steam reforming reaction, catalyst particles have been developed with
internal voids or holes, so as to increase both bed porosity and particle geo-
metric surface area. This results in a lower bed pressure drop and in increased
activity for the reforming reaction, respectively. It is not well established, how-
ever, what the effect of these features of the catalyst particle would be on wall
heat transfer in the tube. Standard heat transfer models characterize the actual
particle by using either the diameter of a sphere of equivalent volume to surface
area ratio or the diameter of a sphere of equivalent volume, depending on
the quantity being correlated. This is a fairly crude approach that can miss the
sometimes subtle effects of the particle design on near-wall uid ow, and thus
on convective heat transfer. Our objective in this work was to use CFD to
perform a more detailed assessment of the particle design for heat transfer.
Simulations were run in the WS conguration described earlier. Cylindrical
particles with length diameter 0.0254 m were placed in the segment in an
arrangement that approximated the most common situation seen from a series
of experimental packings in a transparent tube, with N 4. The arrangement
was constrained by the necessity for it to be axially periodic, so that a periodic
ow eld could be used. Different cases of the same arrangement were studied,
in which the particles had different void structures formed by using various
numbers and sizes of holes running parallel to the axis of the cylinder and
TABLE 1
BENEFITS OF MODERN CATALYST PELLET DESIGN ON REFORMER PERFORMANCE
17 mm Raschig Rings 17 mm Raschig Rings LxD 19 14 mm 4-Hole
Plant rate (relative) 100 112 112
Maximum TWT (1C) 921 940 921
CH
4
slip (mol% dry) 4.4 4.8 4.3
Approach to
Equilibrium. (1C) 3 6 2
Pressure Drop (kg/cm
2
) 2.3 3.1 2.8
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 367
placed symmetrically about that axis. A standard hole had diameter slightly
larger than a quarter of the particle diameter, a big hole had diameter
twice the standard diameter, while a small hole had diameter reduced by a
factor O2 of the standard diameter. The particles studied were designated as
full, 1-hole, 1-bighole, 3-hole, 4-hole, and 4-small-holes, with the obvious in-
terpretation. Symmetry conditions were retained on the sidewalls of the segment
geometry, in default of a better alternative. The focus of the simulations was
intended to be the cylinders in the center (horizontally) of the segment, which
were only marginally affected by the conditions on the side walls (Leising, 2005).
The ow and energy simulations were decoupled, so that the ow eld was
established under isothermal conditions and the temperature eld was then
established with a xed ow eld. The dependence of the ow on temperature
was judged to be small at industrial conditions. The simulation conditions (T, P,
x
i
) were for the reaction gas at the inlet conditions of a typical industrial meth-
anol plant steam reformer. Details and parameter values are available in the
original reference (Nijemeisland et al., 2004).
The ow simulations were run under isothermal conditions at the inlet tem-
perature of the reactor tube T
in
, with periodic conditions on top and bottom
surfaces. The RNG k e model was used for turbulence, with nonequilibrium
wall functions. Fig. 25a shows typical ow path lines for the 1-hole particles,
obtained by simulating the release of marker particles from the planes
r 0.045 m and r 0.05 m. Releasing the particles from these planes empha-
sizes the ow patterns near the tube wall. From the path lines it may be seen that
the ow is mostly axial, as expected; however, there are regions of ow with
a strong radial component, and also regions of backow (i.e., ow with a
negative axial component). The strongly axial ow is found between particles,
such as to the right of the central particle. Strong radial components are found
when the ow is displaced by a particle, such as above the central particle.
Backow often occurs in the wake of an obstruction, or where two particles
approach closely.
FIG. 25. Wall-segment geometry for 1-hole particles: orthographic projections showing (a) ow
path lines for particles released from vertical planes close to the tube wall; (b) ow path lines for
particles released from the bottom horizontal plane.
ANTHONY G. DIXON ET AL. 368
In Fig. 25b, the simulated marker particles were released from the bottom
surface, which generates path lines that show more detail of the ow inside the
WS, at lower radial coordinate values. The path lines reinforce the trends seen in
Fig. 25a, and it is also possible to see some evidence of ow through the center
voids of the particles. Most evident is the mix of spiraling and axial ow be-
tween the center front and center right particles.
It is of interest to determine the extent to which there is ow through the
interior holes of the particles, as the reaction activity is proportional to geo-
metric surface area under these conditions. So, it is important to know whether
the extra surface area provided by the holes is accessible to the ow. It is not
easy to see this internal ow from the path lines in Fig. 25, although there
appears to be ow through the center particle. To determine this more clearly
we constructed a surface that passed through the midpoint of the center particle,
perpendicular to its axis, for each of the particle geometries. This is shown as the
dark square in Fig. 26, which illustrates the results for the 4-hole particle.
The point of view for Fig. 26 is aligned with the axis of the particle, and the
perspective causes the sides of the particle to appear as the outer darker ring of
cells and the sides of the holes to appear as skewed rings of darker cells around
each hole. Contours of velocity magnitude are shown for the midway point of
each hole and demonstrate that there is clearly substantial ow through the
voids, but it is not symmetric inside the holes. By summing mass ow rates for
FIG. 26. View of velocity contours down internal voids or holes of the center 4-hole particle.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 369
the four internal cross-sections, we found that slightly over 10% of the total
mass ow for the segment actually passed through the interior voids of the
center particle. Within each void, the ow rate was higher in that part of the
hole that was toward the top of the particle, i.e., at the higher Z-coordinate. It
should be noted that the particles simulated here were quite large, and so were
the internal voids, as is typical for steam reforming. This result should not be
extrapolated to other applications with smaller particles and voids, which may
present much higher resistance to internal ow.
The simulations for temperature did not use periodic axial boundary con-
ditions, as the long-term objective was to incorporate heat sinks into the par-
ticles (see next subsection). Instead, the technique of virtual stacking was used,
as has been described above. All results were from the third segment in the stack
and used a tube wall boundary condition of constant wall heat ux, with the
inlet uid temperature for the rst stage set to the uniform T
in
used for the ow
simulation, and the top and bottom solid surface heat uxes set to zero. The
outlet uid required no thermal boundary condition. Each subsequent stage
used the outlet temperature eld from the previous stage as the inlet uid tem-
perature eld, and the rst two stages were regarded as being present to over-
come thermal entry effects and to give a more developed radial temperature
prole for the third stage comparisons of the particles. The uids thermal
properties were those of the reforming mixture, the solid thermal properties
were those of alumina, and radiation was neglected, having been previously
determined to be small.
Fig. 27a shows the temperature eld in the uid adjacent to the tube wall,
by means of a temperature contour map. The axes of the map are the axial
coordinate Z and the arc length along the curved tube wall, S. The contour map
shows one hot region and several colder regions in an overall temperature
distribution that was quite moderate. The hotter region in the center of the map
is associated with the strong axial ow component found there. The cold region
to the left of the center of the map (S-coordinate 0.0350.045, Z-coordinate
0.020.04) corresponds to the position of the curved section of the center par-
ticle in Fig. 25. In this area the ow is of average velocity, but has a uniform
direction and a reasonable radial component, creating the cooler spot.
The weaker hot region at the left side of the segment (S-coordinate 0.010.015,
Z-coordinate 0.010.04) corresponds again to a part of the WS where a strong
uniform axial ow was found, but the radial component of the ow was mini-
mal. The energy being put into the system could not be easily transported into
the bed, as the ow was more parallel to the tube wall, which resulted in a higher
near-wall temperature and lower energy uptake. Similar features could be found
near the right-hand side of the temperature plot (S-coordinate 0.08). The cool
spot near the right-hand side (S-coordinate 0.090.1, Z-coordinate 00.03) is a
result of the radial ow into the bed around the particles there.
Fig. 27b gives a more quantitative comparison between the radial temper-
ature proles for the full, 1-hole, and 1-bighole particles. There is a temperature
ANTHONY G. DIXON ET AL. 370
jump over the viscous boundary layer at the tube wall that is not shown here;
the proles are all for uid temperatures in the fully turbulent region. The
temperature proles are very similar, with the full cylinders being slightly
higher, which implied more effective thermal transport for the simulations
shown here. There was little difference between the 1-hole and 1-bighole par-
ticles, except toward the tube wall, which was ascribed to the difference in
porosity there. It is interesting that there was quite a large difference in ow
through the two different 1-hole geometries, which did not translate into a
strong difference in heat transfer effectiveness. There is a complex interaction
between changes in the amount of solid conduction and radial displacement of
ow, which facilitates convective heat transfer. Also, comparisons, such as the
above, do not take into consideration the effect of particle geometry on reac-
tion, which will further affect the heat transfer picture by providing heat sinks.
0 0.2 0.4 0.6 0.8 1
r / R
810
830
850
870
890
910
Solid cylinder
1-hole cylinder
1-big-hole cylinder
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
S [m]
0.01
0.02
0.03
0.04
0.05
Z

[
m
]
830
840
850
860
870
880
890
900
910
920
930
940
950
960
970
980
990
(a)
(b)
T

(
K
)
FIG. 27. (a) Near-wall temperature map for the 1-hole particles; (b) radial temperature proles for
solid cylinders and cylinders with two different sizes of internal void.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 371
The wall temperature maps shown in Fig. 28 are intended to show the quali-
tative trends and patterns of wall temperature when conduction is or is not
included in the tube wall. The temperatures on the tube wall could be calculated
using the wall functions, since the wall heat ux was specied as a boundary
condition and the accuracy of the values obtained will depend on their validity,
which is related to the y
+
values for the various solid surfaces. For the range of
conditions in these simulations, we get y
+
E 1314. This is somewhat low
for the k e model. The values of T
w
are in line with industrially observed
temperatures, but should not be taken as precise.
Fig. 28a shows the wall temperature map for the 1-hole particles that results if
the conduction in the tube wall is omitted from the model. It is quite similar to
the temperature map for the near-wall uid shown in Fig. 27a, and is relatively
rich in features such as regions of hot and cold temperatures, which correspond
to the local ow eld and positions of the particles. The temperatures are con-
siderably higher than in the uid, reecting the temperature difference across
the viscous boundary layer, which is substantial at the industrial ow rate of this
study. There is, correspondingly, a fairly wide range of wall temperatures. Wall
conduction was then included in the model, using thermal properties and tube
thickness of a typical high-temperature reforming tube made out of a high-alloy
steel. The results are presented in Fig. 28b, as a wall temperature map on the
same temperature scale as in Fig. 28a. It appears virtually featureless, and all the
temperature variations seem to have been smoothed out, due to the thermal
conduction in the wall that is now operating. This is misleading, however, as
may be seen in Fig. 28c, in which the same temperature eld is presented on a
scale chosen to bring out the temperature variations. We can see that the fea-
tures in this map are similar to those of Fig. 28a. The average wall temperatures
are the same with or without wall conduction, but the extremes in temperature
of Fig. 28a have been strongly mitigated by the wall conduction, as shown in
Fig. 28c. The distribution of temperatures still reects the distribution of tem-
peratures in the neighboring uid, which in turn reects the local variation in
heat transfer resistance near the wall. Temperature elds within the bed have
been shown to be unaffected by the inclusion of conduction in the tube wall
(Dixon et al., 2005).
C. REACTION THERMAL EFFECTS IN SPHERES USING CFD
Steam reforming is a heterogeneously catalyzed process, with nickel catalyst
deposited throughout a preformed porous support. It is empirically observed in
the industry, that conversion is proportional to the geometric surface area of
the catalyst particles, rather than the internal pore area. This suggests that
the particle behaves as an egg-shell type, as if all the catalytic activity were
conned to a thin layer at the external surface. It has been demonstrated by
conventional reaction-diffusion particle modelling that this behaviour is due to
ANTHONY G. DIXON ET AL. 372
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
S (m)
0.01
0.02
0.03
0.04
0.05
Z

(
m
)
980
990
1000
1010
1020
1030
1040
1050
1060
1070
1080
1090
1100
1110
1120
1130
1140
1150
1160
1170
1180
1190
1200
1-hole Tube wall Wall conduction
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
S (m)
0.01
0.02
0.03
0.04
0.05
Z

(
m
)
1050
1055
1060
1065
1070
1075
1080
1085
1090
1095
1100
1105
1110
1-hole Tube wall Wall conduction
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
S (m)
0.01
0.02
0.03
0.04
0.05
Z

(
m
)
980
990
1000
1010
1020
1030
1040
1050
1060
1070
1080
1090
1100
1110
1120
1130
1140
1150
1160
1170
1180
1190
1200
1-hole Tube wall No wall conduction
(a)
(b)
(c)
FIG. 28. Wall temperature maps for 1-hole particles (a) without wall conduction; (b) with wall
conduction, on the same temperature scale as (a); (c) with wall conduction, on a temperature scale
chosen to show nonuniform temperature features.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 373
the extremely strong diffusion limitations under steam reforming conditions
(Pedernera et al., 2003).
This study was carried out to simulate the 3D temperature eld in and around
the large steam reforming catalyst particles at the wall of a reformer tube, under
various conditions (Dixon et al., 2003). We wanted to use this study with
spherical catalyst particles to nd an approach to incorporate thermal effects
into the pellets, within reasonable constraints of computational effort and re-
alism. This was our rst look at the problem of bringing together CFD and
heterogeneously catalyzed reactions. To have included species transport in the
particles would have required a 3D diffusion-reaction model for each particle to
be included in the ow simulation. The computational burden of this approach
would have been very large. For the purposes of this rst study, therefore,
species transport was not incorporated in the model, and diffusion and mass
transfer limitations were not directly represented.
The approach that was used was to represent the energetic effects of reac-
tion through user-dened volumetric heat generation terms. For this approach,
the partial pressures of the species in the gas mixture were held constant at
the conditions corresponding to the position of interest in the reactor tube, and
reaction energy effects were allowed in the catalyst particles in an outer shell
only. Thus, the WS was regarded as representing a differential slice of the
reactor, with uniform species partial pressures. For each solid phase cell, if the
local position of the volume centroid was within an outer shell of the catalyst
particle, the reaction rates were calculated at the local solid temperature, and an
energy source term was included. The volumetric heat generation rate, Q
p
, in
units of W/m
3
(cat.) was given by the sum of the products of the reaction rates
and the heats of reaction. A cut-off value was used so that the reaction heat
effects were conned to the outer shell, and a parametric study was carried out
on the effects of the choice of the value. This was equivalent to setting the
effectiveness factor of the particle.
The simulation was run by rst determining the ow solution in the periodic
segment, and subsequently determining the energy solution. The solution of
ow and energy were decoupled, as the temperature-dependence of the gas
properties was not expected to inuence ow at the extremely high industrial
ow rates simulated here. This assumption allowed the ow to be treated as
periodic, independently of the temperature eld. The gas heated up slightly as
it passed through the segment, and as the reaction rates gave temperature-
dependent sinks/sources, the temperature eld could not be treated as periodic.
The ux through the tube wall of the segment was kept constant. Symmetry
conditions were applied to the sides of the WS.
The CFD calculations of the present work used conditions and compositions
from a Johnson Matthey detailed reformer model of a methanol plant steam
reformer with upwards ow, at typical operating conditions. Conditions were
chosen corresponding to three different axial positions along the tube, to reect
reaction rates typical of those close to the inlet, midway down the tube and close
ANTHONY G. DIXON ET AL. 374
to the outlet. Details of parameter values used in these simulations can be found
in the original reference (Dixon et al., 2003).
For our simulations with heat sinks in spheres, we chose not to stack runs, in
order to minimize the change in temperature across the WS, and instead carried
out a study of the effects of simplied boundary conditions. For most runs,
simple boundary conditions were used in which the uid entering at the base of
the segment was set uniformly to T
in
, and uxes through the solid areas on the
top and bottom planes were set to zero. No thermal boundary condition was
required for the ow outlet boundary.
From knowledge of the coordinates of the center points of the spherical
particles, we developed a simple criterion to select those control volumes whose
centroid lay within the cut-off from the particle surface. We then veried this
using a user-dened marker technique as shown in Fig. 29a. In the gure, uid
cells were tagged 0 and are intermediate in shade, solid particle inactive cells
were tagged 1 and are the lightest, while the selected solid active cells were
tagged 2 and are the darkest. As can be seen, the algorithm correctly selected the
cells at the particle edges. The algorithm did not select any interior cells. The
appearance of somewhat ragged edges is due to a 2D representation of a 3D
situation. A closer look is shown in Fig. 29b. Some tetrahedra appear to be
selected that are not at the surface of a particle. In these cases, the centroid of
the tetrahedron lies close enough to the surface for the cell to be selected, even
though the part of the cell that is intersected by the plane of Fig. 29b appears to
lie further from the surface. A similar explanation holds for cells that appear
to be next to a surface, but which are not selected, due to the position of
the centroid. Examination of the geometry in 3D conrms this interpretation of
the picture.
The simulation of the thermal effects of the steam reforming reaction was
based on a published reaction model (Hou and Hughes, 2001) for methane
FIG. 29. (a) Midplane cross-section of the WS packed with spheres, showing control volumes
found by selection algorithm, marked as darkest cells.; (b) close-up of gap between two spheres,
showing cell selection in detail.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 375
reforming over a Ni/Al
2
O catalyst. The model is based on the performance of
the steam reforming catalyst produced by ICI-Katalco (now Johnson Matthey
Catalysts), and consisted of the following three main reactions:
1: CH
4
+H
2
O CO+3H
2
2: CO+H
2
O CO
2
+H
2
3: CH
4
+2H
2
O CO
2
+4H
2
The nal rate expressions, which were used in the present work, were given by
Hou and Hughes (2001). In these rate expressions all reaction rate and equi-
librium constants were dened to be temperature-dependent through the
Arrhenius and vant Hoff equations. The particular values for the activation
energies, heats of adsorption, and pre-exponential constants are available in the
original reference and were used in our work without alteration.
The active shell thickness was determined by the cut-off value r
cut
, which
represented how much of the spherical catalyst particle was active and provided
an energy sink/source, depending on the local reaction rates. A position in the
particle was regarded as inert if its distance from the particle center was less
than r
cut
. The actual catalyst particles themselves were uniformly impregnated
with active metal catalyst. The use of r
cut
was a device to represent the egg-
shell nature of the reaction and diffusion in the catalyst particles, in the present
study of the energetic effects of steam reforming. It is, therefore, of interest to
investigate the effect of different values of r
cut
, or the equivalent values of the
effectiveness factor, Z.
The amount of heat actually taken up by the particles was an important
quantity, as tubes operate under heat transfer limited conditions near the tube
inlet. Fig. 30 shows a plot of Q against Z, where Q was the total energy ow into
the solid particles, for the entire segment. For inlet conditions, Q varied strongly
at lower Z, but was almost constant at higher values. As r
cut
/r
p
decreased from
0.95 to 0.0 and the effectiveness factor increased from nearly zero to one, the
active solid volume increased by a factor of 7. If the solid temperature had
remained the same, the heat sink would also have had to increase sevenfold.
This could not be sustained by the heat transfer rate to the particles, so the
particle temperature had to decrease. This reduced the heat sink and increased
the driving force for heat transfer until a balance was found, which is repre-
sented by the curve for the inlet in Fig. 30.
Subsequent to the simulations reported in Dixon et al. (2003), some runs were
carried out in which the virtual stacking was used, to provide a similar length of
heated tube as in the runs without heat sinks, for comparison purposes. Some
results for the third segment are given in Fig. 31, which shows a horizontal
plane at the vertical midpoint of the WS and a vertical plane at the horizontal
midpoint of the WS, under inlet tube conditions. The level of particle activity
was r
cut
/r
p
0.95, which was closest to the egg-shell picture.
ANTHONY G. DIXON ET AL. 376
We can see that for these conditions, the temperature elds inside the wall
particles are far from symmetric. Signicant temperature incursions appear
inside the spheres, and the inuence of the wall is strong. The spheres are hotter
close to the tube wall than on the side facing the center of the segment. The
interior particles appear to be more symmetrical in temperature. It is noticeable
that the particles are considerably lower in temperature than the surrounding
FIG. 31. Horizontal and vertical planes through the third-stacked WS, showing the temperature
elds in the uid and through the spherical catalyst particles, with 5% activity.
0 0.2 0.4 0.6 0.8 1
-100
0
100
200
300
400
500
Q

(
W
)
Mid-tube
Inlet
Outlet
FIG. 30. Heat ow into particles, as a function of effectiveness factor, for the three tube positions
studied in Dixon et al. (2003).
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 377
gas, except possibly near a contact point with the wall. This is an expected result
of the endothermic reactions.
At the inlet of the reactor tube, the gas mixture is rich in methane and steam
with some carbon dioxide. It is also at a relatively low temperature. The
watergas shift reaction (reaction 2) is energetically favored because of its low
activation energy, but the reactant carbon monoxide is not present to any sig-
nicant extent. Of the other two reactions, methane conversion to carbon di-
oxide (reaction 3) is favored over methane conversion to carbon monoxide
(reaction 1) at lower temperatures, as its activation energy E
3
is approximately
half the value of that of reaction 1, E
1
. Since this is not the desired product
mix for synthesis gas, the strategy would be to increase the tube temperatures
as quickly as possible to a range where reaction 1 is favored. Reaction 3 is
strongly endothermic, and reduces the heat-up rate of the catalyst by providing
a heat sink. Thus, efcient heat transfer in the early stages of the tube is es-
sential.
D. REACTION THERMAL EFFECTS IN CYLINDERS USING CFD
Our initial work on reaction thermal effects involved CFD simulations of
uid ow and heat transfer with temperature-dependent heat sinks inside
spherical particles. These mimicked the heat effects caused by the endothermic
steam reforming reaction. The steep activity proles in the catalyst particles
were approximated by a step change from full to zero activity at a point 5% of
the sphere radius into the pellet.
To extend these calculations to cylinders is more complicated, as both the
position and orientation of a cylinder must be obtained. To do that we followed
the sequence of operations used to position each cylinder, as shown in Fig. 32a,
for particle 1, the lower front particle in the wall segment (note that wire frames
of the previous positions are retained in each sketch for comparison). Similar
sequences were available for each of the other particles in the wall segment
model.
It was possible to follow the centre point and top centre point under the
transformations, and we developed criteria to select those control volumes
whose centroid lay within the cut-off from either the curved surface or the at
ends. We then veried this using the same user-dened marker technique as
for the spheres, and the results are shown in Fig. 32b. Again, uid cells were
intermediate in shade; solid particle inactive cells were the lightest, while the
selected solid active cells were the darkest. As can be seen, the algorithm cor-
rectly selected the cells at the particle edges, whether at or curved. The al-
gorithm did not select any interior cells. The appearance of somewhat ragged
edges is again due to a 2D representation of a 3D situation.
Following the verication step, we applied the heat sink methodology to a
WS with full cylinders, comprising a ow eld solution followed by three virtual
ANTHONY G. DIXON ET AL. 378
stacking thermal simulations. As mentioned previously, the main reason for
stacking is to mitigate the effects of a at inlet temperature prole, and to
develop the thermal penetration of the bed. For our simulations with heat sinks
in spheres, where we studied the effects of using simplied boundary conditions
at the inlet and outlet instead of virtual stacking, we found that the main
qualitative features of the solution were not changed, but there was denitely
some inuence at the detailed level, and on quantitative results. For the work
described here, we wished to compare simulations with and without heat sinks,
so we chose to use virtual stacking as in our earlier work without heat sinks.
This meant that the gas and solids would gradually heat up and depart from the
set reactor tube conditions, while the partial pressures stayed constant. We
regarded this as one of the simplications necessary for the heat sinks meth-
odology, which will be eliminated by the development of an improved approach
with proper modeling of species diffusion and reaction in the solid particles. We
are currently working on such an approach.
Comparisons for the full solid cylinders, with and without heat sinks, were
made in Nijemeisland et al. (2004) for conditions near the inlet of the reactor
tube. These showed that temperature proles changed drastically when heat
sinks were included. In Figs. 33 and 34, we show a similar comparison for
conditions typical of the middle of the reactor tube. We compare the planes at
the midpoint of the third stage, for the cases where the outer 5% of the particle
was active and where the entire particle was inactive. These are shown in Fig. 33.
Both temperature elds shown are on the same temperature scale. As ex-
pected, the simulation that included heat sinks resulted in much colder tem-
peratures. With no reaction heat sinks, the temperature eld is barely affected
by the presence of the particles, despite their higher thermal conductivity, as
FIG. 32. (a) Sequence of transformations 1-2-3-4 to place bottom front cylindrical particle;
(b) Midplane cross-section of the WS packed with cylinders, showing control volumes found by
selection algorithm, marked as darkest cells.
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 379
shown in the picture on the right. When active pellets are simulated, there is a
clear effect resulting in the particles being seen owing to their lower temper-
atures. It is, however, remarkable that at the level of reaction included in these
models, there is still considerable lack of symmetry in the particle temperatures
for the cylinders nearest the tube wall. This observation conrms that the as-
sumption of a symmetric temperature eld surrounding the catalyst particles
could lead to serious errors in estimating reaction rates and modeling the
reactor tube.
FIG. 33. Comparison of temperature elds in WS midplane for particles with active outer shell
(95% inactive, left diagram) and particles with no reaction heat effects (100% inactive, right di-
agram).
900
920
940
960
980
1000
1020
1040
1060
1080
1100
0 0.2 0.4 0.6 0.8 1
r/r
No sinks
Sinks
No sinks - fluid
Sinks - fluid
T

[
K
]
FIG. 34. Comparison of radial temperature proles for WS packed with full cylinders, with and
without heat sinks; solid symbols are for temperatures averaged over uid and solid, open symbols
for temperatures averaged over uid alone.
ANTHONY G. DIXON ET AL. 380
These qualitative observations can be expressed more quantitatively by
examining average radial temperature proles for the WS, as shown in Fig. 34.
The introduction of heat sinks into the catalyst particles decreases the radial
temperatures, whether they are averaged over both solid and uid cells or over
the uid cells alone. If we look at the average uid temperatures, the heat sinks
cause a decrease of about 4050 K, and the uid proles are fairly at in both
cases. For the situation in which there are no heat sinks, the radial temperature
prole is almost the same for the uid and for the uid and solid combined. This
corresponds to the observation in Fig. 32b for the 100% inactive case, in which
no differences between uid and solid could be seen. When heat sinks are
included, the solid is much colder, and maxima and minima in the overall
averaged prole appear, together with much larger differences in temperature
compared to the uid only prole. It is expected that the shape of the overall
temperature prole may be related to the distribution of the solid and its active
region; future work will investigate this further.
The extension of this work to include catalyst particles with internal voids is
more complex, as there are regions of catalytic activity adjacent to the internal
holes, complicating the testing procedure. A comparison of several different
catalyst congurations of internal voids has recently been completed, and a
description of the method, its verication, and the results obtained will be the
subject of a future publication.
V. Future Prospects
This article has attempted to review the issues in applying CFD to simulate
interstitial ow in packed tubes, with an emphasis on low-N tubes. The rapid
changes in computational capability of todays computers mean that the prob-
lems and limitations discussed here will also change rapidly. All we can do is to
try to extrapolate where the needs and areas of interest are likely to be in the
near future.
There continues to be a need for improved automatic generation of 3D
packings, for arbitrary shapes of particles, in cylindrical tubes, and with wall
effects present. Since a CFD simulation is based on a single instance of a
packing, it is necessary to have the details of that packing readily available. To
extend the simulations to consider ensembles of packings, for statistical analysis,
would require an easy, inexpensive method of generating the packings.
The LBM has been used to obtain some very interesting results and good
insight into packed tube ow, dispersion, and simple reactions. This method
however appears to us to have limitations for reactor applications due to its
complications in addressing high-Re, nonisothermal situations. Future develop-
ments in simulations of catalysis in reactor tubes are likely to be based on the
FV-CFD methods. The geometry creation, meshing, and post-processing tools
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 381
of the commercial versions of these methods are powerful, and a growing array
of turbulence model options promises adequate capability for simulating high-
Re ows. Abilities to simulate porous regions and species ows are available or
are rapidly being added in several packages, in addition to the traditional base
of ow and heat transfer.
One area that will be critical for xed bed modeling is that of wall functions.
Owing to the packing, a packed tube ow is dominated by interactions with
solid surfaces. As discussed above, resolving turbulent boundary layers will be
too expensive for a long time to come, while the demands of the geometry and
accurate calculation often make it difcult or impossible to obtain a mesh to
satisfy the conditions for wall functions to be strictly applicable. The develop-
ment of improved wall functions that can work with a range of mesh sizes would
be very helpful. Alternatively, several groups have indicated reasonable results
and agreement with experimental data from simulations that used meshes and
wall functions that were not strictly appropriate. More research is needed on
this question.
The research on the ow regimes in packed tubes suggests that laminar ow
CFD simulations should be reasonable for Re o100 approximately, and tur-
bulent simulations for Re 4600, also approximately. Just as RANS models
provide steady solutions that are regarded as time averages of the real time-
dependent turbulent ow, it may be suggested that CFD simulations in the
unsteady laminar inertial range 100 oRe o600 may provide a time-averaged
picture of the ow eld. As with wall functions, comparisons with experimental
data and an improved assessment of what information is really needed from the
simulations will inform us as to how to proceed in these areas.
Simulations with representative segments and unit cells employing periodic or
symmetry boundary conditions are likely to be necessary for the foreseeable
future. Although simulations of complete tube cross-sections would be pre-
ferred, these are anticipated to remain too costly for some time to come. This
will be especially true for turbulent ows and geometries that require ne
meshes or boundary-layer resolution.
The validation of CFD codes by comparison to reliable experiments is of the
highest importance. Especially promising is the use of MRI methods to non-
invasively provide ow elds and dispersion data. Major challenges will be to
extend MRI and similar methods such as LDV and particle tracking to a wider
range of conditions, and to develop noninvasive measurements of temperature
to improve verication of heat transfer simulations.
Applications of packed tube CFD to improve transport and reaction under-
standing and models are still in early stages. A number of studies have focused
on addressing the technical issues of using CFD for packed tubes, and have
presented qualitative results that have yielded insight into the phenomena.
Other studies have tried to use CFD simulations to extend experimental data,
and provide estimates of familiar transport parameters.
ANTHONY G. DIXON ET AL. 382
Some of the recent work on structured packings may provide a pointer
to future directions for developing models of transport in packed tubes of
particles. The identication of REUs in structured packings has provided a
fruitful approach to correlating pressure drop. Since low-N packings of parti-
cles are also strongly structured due to the wall ordering, can we simi-
larly identify REUs for these tubes, and develop contributions to dispersion
and heat transfer? Some of our work on near-wall heat transfer suggests that
this may be a way forward, to move on from the current approaches based on
neglect of bed structure and empirical correlation of bed effective transport
parameters.
The simulation of reacting ows in packed tubes by CFD is still in its earliest
stages. So far, only isothermal surface reactions for simplied geometries and
elementary reactions have been attempted. Heterogeneous catalysis with diffu-
sion, reaction, and heat transfer in solid particles coupled to the ow, species,
and temperature elds external to the particles remains a challenge for the
future.
A promising start has been made in packed tube CFD simulation, especially
at lower Re and for reduced geometries such as unit cells and bed segments.
Applications to transport and catalyst particle assessment are active areas of
research. We look forward to the insights that these simulations promise, to
more streamlined and easier application of the CFD methods, and to wider
applications such as two-phase ow in trickle beds.
Nomenclature
c
p
uid heat capacity (J/kg?K)
C
m
constant in Eqn. (6) (-)
C
1e
, C
2e
,
C
3e
constants in Eqn. (8) (-)
d
t
tube diameter (m)
d
p
particle diameter (m)
D
i
molecular diffusivity of species I (m
2
/s)
E empirical constant in law of the wall (-)
E
i
activation energy of species i (kJ/mol)
F
i
external body forces component in direction i (N/m
3
)
g
i
acceleration due to gravity in coordinate direction i (m/s
2
)
G
k
generation of turbulent kinetic energy due to stress (J/m
3
?s)
G
b
generation of turbulent kinetic energy due to buoyancy (J/m
3
?s)
h
i
uid enthalpy of species i (kJ/kg)
H uid enthalpy (kJ/kg)
h
o
particle to uid heat transfer coefcient (W/m
2
?K)
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 383
h
p
particle to uid heat transfer coefcient (W/m
2
?K)
h
w
wall-heat transfer coefcient (W/m
2
?K)
J
j
diffusion ux of species i (kg/m
2
?s)
k turbulent kinetic energy (J/kg)
k
P
turbulent kinetic energy at point P (J/kg)
k
f
uid thermal conductivity (W/m?K)
k
r
effective radial thermal conductivity (W/m?K)
k
t
turbulent thermal conductivity (W/m?K)
K material-dependent constant (Eqn. 26) (-)
eddy length scale (m)

d
Kolmogorov dissipation scale (m)
N tube-to-particle diameter ratio (d
t
/d
p
) (-)
P static pressure (Pa)
P Larson-Miller parameter (K)
q heat ux (W/m
2
)
Q heat uptake into catalyst particles (W/m
3
)
Q
p
heat generation in catalyst particle (W/m
3
(cat.))
r radial coordinate (m)
r
p
particle radius (m)
R tube radius (m)
S arc length (m)
S
h
volumetric heat source (J/m
3
?s)
S
m
volumetric mass source (kg/m
3
?s)
t time (s)
T temperature (K)
T
in
tube inlet temperature (K)
T
0
global reference temperature (K)
T
w
tube wall temperature (K)
u uctuating velocity component (m/s)
% u mean velocity component (m/s)
u
+
dimensionless mean velocity (-)
u
*
dimensionless mean velocity (Eqn. 19) (-)
u
P
mean velocity of uid at point P (m/s)
v
in
inlet velocity (m/s)
v
i
interstitial gas velocity (m/s)
v
0
supercial gas velocity (m/s)
X coordinate (m)
y coordinate normal to the wall (m)
y
P
distance from wall to point P (m)
y
+
dimensionless distance from wall (-)
y
*
dimensionless distance from wall (Eqn. 20) (-)
Y
i
mass fraction of species i (-)
Z axial coordinate (m)
ANTHONY G. DIXON ET AL. 384
GREEK LETTERS
a
f
relaxation factor for variable f (-)
b thermal expansion coefcient (K
1
)
d
ij
Kronecker delta function ( 1 if i j, 0 otherwise) (-)
e turbulence dissipation rate (J/kg?s)
e bed porosity (-)
y dimensionless temperature (TT
in
)/(T
wall
T
in
) (-)
k von Ka rma n constant (-)
l uid thermal conductivity (W/m?K)
l
t
turbulent thermal conductivity (W/m?K)
m uid dynamic viscosity (kg/m?s)
m
eff
effective viscosity (kg/m?s)
m
mol
molecular viscosity (kg/m?s)
m
t
turbulent viscosity (kg/m?s)
r uid density (kg/m
3
)
s
k
turbulent Prandtl number for k (-)
s
e
turbulent Prandtl number for e (-)
t deviatoric stress tensor (N/m
2
)
t
w
wall shear stress (N/m
2
)
u uid kinematic viscosity (m
2
/s)
O specic dissipation rate (s
1
)
DIMENSIONLESS NUMBERS
Particle Nusselt number Nu
h
p
d
p
k
f
Pe clet number Pe
rv
0
c
p
d
p
k
f
Prandtl number Pr
c
p
m
k
f
Turbulent Prandtl number Pr
t

c
p
m
t
k
f
Reynolds number (supercial) Re
rv
0
d
p
m
Reynolds number (interstitial) Re
i

rv
i
d
p
m
Reynolds number (effective) Re
eff

rv
0
d
p
m
eff
ABBREVIATIONS
BFD Brinkman-Forcheimer-Extended Darcy
CFD Computational Fluid Dynamics
CR Collective Rearrangement
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 385
CSP Composite Structured Packings
cv control volume
DEM Discrete Element Method
DNS Direct Numerical Simulation
FD Finite Difference
FE Finite Element
FV Finite Volume
LBM Lattice Boltzmann Method
LDA Laser Doppler Anemometry
LES Large Eddy Simulation
MRI Magnetic Resonance Imaging
RANS Reynolds-Averaged NavierStokes
REU Representative Elementary Unit
RNG Renormalized Group
RSM Reynolds Stress Model
RTD Residence Time Distribution
SD Sequential Deposition
TWT Tube Wall Temperature
WS Wall Segment
REFERENCES
Abbott, P. E. J., Conduit, M. R., and Manseld, K. Proceedings of World Methanol Conference,
paper 14-1, Crocco & Associates Inc, Houston, TX (1989).
Agarwal, R. Annual Review of Fluid Mechanics 31, 125 (1999).
Benenati, R. F., and Brosilow, C. B. AIChE J. 8, 359 (1962).
Bey, O., and Eigenberger, G. Chem. Eng. Sci. 52, 1365 (1997).
Bode, J. Comp. & Chem. Eng., 18 SUPPL, S247 (1994).
Bryant, S. L., Mellor, D. W., and Cade, C. A. AIChE J. 39, 387 (1993).
Calis, H. P. A., Nijenhuis, J., Paikert, B. C., Dautzenberg, F. M., and van den Bleek, C. M. Chem.
Eng. Sci. 56, 1713 (2001).
Caulkin, R., Fairweather, M., Jia, X., Gopinathan, N., and Williams, R. A. Comput. Chem. Eng 30,
1178 (2006).
Chan, S. K., and Ng, K. M. Chem. Eng. Comm. 48, 215 (1986).
Chen, S., and Doolen, G. D. Annual Review of Fluid Mechanics 30, 329 (1998).
Chu, C. F., and Ng, K. M. AIChE J. 35, 148 (1989).
Craft, T. J., Gerasimov, A. V., Iacovides, H., and Launder, B. E. Int. J. Heat Fluid Flow 23, 148
(2002).
Dalman, M. T., Merkin, J. H., and McGreavy, C. Computers and Fluids 14, 267 (1986).
Debus, C., Nirschl, H., Delgado, A., and Denk, V. Chem.-Ing.-Tech. 70, 415 (1998).
Derkx, O. R., and Dixon, A. G. Numerical Heat Transfer A 29, 777 (1996).
Dhole, S. D., Chhabra, R. P., and Eswaran, V. Chem. Eng. Res. Des. 82(A5), 642 (2004).
Dixon, A. G., and Nijemeisland, M. Ind. Eng. Chem. Res. 40, 5246 (2001).
Dixon, A. G., Nijemeisland, M., and Stitt, E. H. Int. J. Chemical Reactor Eng. 1, A11 (2003).
Dixon, A. G., Nijemeisland, M., and Stitt, E. H. Ind. Eng. Chem. Res. 44, 6342 (2005).
ANTHONY G. DIXON ET AL. 386
Dybbs, A., and Edwards, R. V., Fundamentals of Transport Phenomena in Porous Media
(J. Bear, and M. Corapcioglu, Eds.) p. 201. Martinus Nijhoff, Dordrecht (1984).
Ergun, S. Chem. Eng. Prog. 48, 89 (1952).
Esterl, S., Debus, C., Nirschl, H., and Delgado, A., ECCOMAS 984th Eur. Comput. Fluid Dyn.
Conf. (K. D. Papailiou, Ed.), p. 692. John Wiley & Sons Ltd, New York (1998).
Fluent, Users Guide version 6.1.2, Fluent Inc., Lebanon, NH (2003).
Freund, H., Bauer, J., Zeiser, T., and Emig, G. Ind. Eng. Chem. Res. 44, 6423 (2005).
Freund, H., Zeiser, T., Huber, F., Klemm, E., Brenner, G., Durst, F., and Emig, G. Chem. Eng. Sci.
58, 903 (2003).
Georgiadis, J., Noble, D. R., Uchanski, M. R., and Buckius, R. O. ASME J. Fluid Eng. 118, 434
(1996).
Giese, M., Rottscha fer, K., and Vortmeyer, D. AIChE J. 44, 484 (1998).
Gladden, L. F., Mantle, M. D., and Sederman, A. J. Adv. Chem. Eng. 30, 63 (2005).
Gladden, L. F. AIChE J. 49, 2 (2003).
Guardo, A., Coussirat, M., Larrayoz, M. A., Recasens, F., and Egusquiza, E. Ind. Eng. Chem. Res.
43, 7049 (2004).
Guardo, A., Coussirat, M., Larrayoz, M. A., Recasens, F., and Egusquiza, E. Chem. Eng. Sci. 60,
1733 (2005).
Gunjal, P. R., Ranade, V. V., and Chaudhari, R. V. AIChE J. 51, 365 (2005).
Harris, C. K., Roekaerts, D., Rosendal, F. J. J., Buitendijk, F. G. J., Daskopoulos, Ph., Vreenegoor,
A. J. N., and Wang, H. Chem. Eng. Sci. 51, 1569 (1996).
Hill, R. J., Koch, D. L., and Ladd, A. J. C. J. Fluid Mech. 448, 213 (2001a).
Hill, R. J., Koch, D. L., and Ladd, A. J. C. J. Fluid Mech. 448, 243 (2001b).
Hou, K., and Hughes, R. Chem. Eng. Journal 82, 311 (2001).
Jakobsen, H. A., Lindborg, H., and Handeland, V. Comput. Chem. Eng. 26, 333 (2002).
Jia, X., and Williams, R. A. Powder Technology 120, 175 (2001).
Jiang, Y., Khadilkar, M. R., Al-Dahhan, M. H., and Dudukovic, M. P. AIChE J. 48, 701 (2002).
Jolls, K. R., and Hanratty, T. J. Chem. Eng. Sci. 21, 1185 (1966).
Kays, W. M. J. Heat Transfer 116, 284 (1994).
Kim, S.-E., and Choudhury, D., in Separated and Complex Flows, ASME FED vol. 217, p. 273.
ASME, New York (1995).
Kingston, G., and Nunge, R. Can. J. Chem. Eng. 51, 246 (1973).
Kuipers, J. A. M., and van Swaaij, W. P. M. Adv. Chem. Eng. 24, 227 (1998).
Kutsovsky, Y. E., Scriven, L. E., Davis, H. T., and Hammer, B. E. Phys. Fluids 8, 863 (1996).
Kvamsdal, H. M., Svendsen, H. F., and Hertzberg, T. O. Olsvik. Chem. Eng. Sci. 54, 2697 (1999).
Lahbabi, A., and Chang, H. -C. Chem. Eng. Sci. 40, 435 (1985).
Landon, V. G., Hebert, L. A., and Adams, C. B., in Heat TransferHouston 1996 Proc. 31st
National Heat Transfer Conference, p. 134. AIChE Symp. Ser., 1996.
Launder, B. E., and Spalding, D. B., Lectures in Mathematical Models of Turbulence. Academic
Press, London p. 1 (1972).
Launder, B. E., and Spalding, D. B. Comp. Meth. Appl. Mech. Eng. 3, 269 (1974).
Larachi, F., Petre, C. F., Iliuta, I., and Grandjean, B. P. A. Chem. Eng. & Process. 42, 535 (2003).
Lati, M. A., Midoux, N., Storck, A., and Gence, J. N. Chem. Eng. Sci. 44, 2501 (1989).
Lebon, L., Oger, L., Leblond, J., Hulin, J. P., Martys, N. S., and Schwartz, L. M. Phys. Fluids 8, 293
(1996).
Leising, G., M.S. Thesis, Worcester Polytechnic Institute, Worcester, MA (2005).
Lerou, J. J., and Ng, K. M. Chem. Eng. Sci. 51, 1595 (1996).
Liu, G., and Thompson, K. E. Powder Technology 113, 185 (2000).
Lloyd, B., and Boehm, R. Numerical Heat Transfer 26, 237 (1994).
Logtenberg, S. A., and Dixon, A. G. Chem. Eng. & Process. 37, 7 (1998a).
Logtenberg, S. A., and Dixon, A. G. Ind. Eng. Chem. Res. 37, 739 (1998b).
Logtenberg, S. A., Nijemeisland, M., and Dixon, A. G. Chem. Eng. Sci. 54, 2433 (1999).
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 387
Lund, K. O., Nguyen, H., Lord, S. M., and Thompson, C. Can. J. Chem. Eng. 77, 769 (1999).
Magnico, P. Chem. Eng. Sci. 58, 5005 (2003).
Maier, R. S., Kroll, D. M., Bernard, R. S., Howington, S. E., Peters, J. F., and Davis, H. T. Phys.
Fluids 15, 3795 (2003).
Maier, R. S., Kroll, D. M., Kutsovsky, Y. E., Davis, H. T., and Bernard, R. S. Phys. Fluids 10, 60
(1998).
Mansoorzadeh, S., Pain, C. C., De Oliveira, C. R. E., and Goddard, A. J. H. Int. J. Numer. Meth.
Fluids 28, 903 (1998).
Manz, B., Gladden, L. F., and Warren, P. B. AIChE J. 45, 1845 (1999).
Mathur, S. R., and Murthy, J. Y. Numerical Heat Transfer 31, 195 (1997).
McGreavy, C., Kam, E., Foumeny, E. A., Guidoum, A., and Ikponmwosa, A. N., in 2nd In-
ternational Symposium on Application of Laser Anemometry to Fluid Mechanics, Lisbon,
(1984).
McGreavy, C., Foumeny, E. A., and Javed, K. H. Chem. Eng. Sci. 41, 787 (1986).
McKenna, T. F., Spitz, R., and Cokljat, D. AIChE J. 45, 2392 (1999).
Mickley, H. S., Smith, K. A., and Korchak, E. I. Chem. Eng. Sci. 20, 237 (1965).
Mueller, G. E. Powder Technology 92, 179 (1997).
Mueller, G. E. Powder Technology 159, 105 (2005).
Nandakumar, K., Shu, Y., and Chuang, K. T. AIChE J. 45, 2286 (1999).
Nijemeisland, M., and Dixon, A. G. Chem. Eng. J. 82, 231 (2001).
Nijemeisland, M., and Dixon, A. G. AIChE J. 50, 906 (2004).
Nijemeisland, M., Dixon, A. G., and Stitt, E. H. Chem. Eng. Sci. 59, 5185 (2004).
Nirschl, H., Dwyer, H. A., and Denk, V. J. Fluid Mech. 283, 273 (1995).
Niven, R. K. Chem. Eng. Sci. 57, 527 (2002).
Nolan, G. T., and Kavanagh, P. E. Powder Technology 72, 149 (1992).
Nolan, G. T., and Kavanagh, P. E. Powder Technology 84, 199 (1995).
Park, J., and Gibbs, S. J. AIChE J. 45, 655 (1999).
Patankar, S. V., Numerical Heat Transfer and Fluid Flow, p. 1 Hemisphere, Washington, D.C.
(1980).
Pedernera, M. N., Pin a, J., Borio, D. O., and Bucala , V. Chem. Eng. J. 94, 29 (2003).
Petre, C. F., Larachi, F., Iliuta, I., and Grandjean, B. P. A. Chem. Eng. Sci. 58, 163 (2003).
Ranade, V., Computational Flow Modeling for Chemical Reactor Engineering, p. 403 Academic
Press, New York (2002).
Rashidi, M., Tompson, A., Kulp, T., and Peurrung, L. J. Fluids Eng. 118, 470 (1996).
Ren, X., Stapf, S., and Blumich, B. AIChE J. 51, 392 (2005).
Reyes, S. C., and Iglesia, E. Chem. Eng. Sci. 46, 1089 (1991).
Ridler, D. E., and Twigg, M. V. Steam Reforming. pp. 225282 in: Twigg, M. V. (Ed.), Catalyst
Handbook (2nd Ed.) Manson Publishing, London, 1996.
Rode, S., Midoux, N., Lati, M. A., Storck, A., and Saatdjian, E. Chem. Eng. Sci. 49, 889 (1994).
Romkes, S. J. P., Dautzenberg, F. M., van den Bleek, C. M., and Calis, H. P. A. Chem. Eng. J. 96, 3
(2003).
Schouten, E. P. S., Borman, P. C., and Westerterp, K. R. Chem. Eng. Sci. 49, 4725 (1994).
Schuster, J., and Vortmeyer, D. Chemie Ingenieur Technik 53, 806 (1981).
Sederman, A. J., Johns, M. L., Bramley, A. S., Alexander, P., and Gladden, L. F. Chem. Eng. Sci.
52, 2239 (1997).
Sederman, A. J., Johns, M. L., Alexander, P., and Gladden, L. F. Chem. Eng. Sci. 53, 2117 (1998).
Seguin, D., Montillet, A., and Comiti, J. Chem. Eng. Sci. 53, 3751 (1998a).
Seguin, D., Montillet, A., Comiti, J., and Huet, F. Chem. Eng. Sci. 53, 3897 (1998b).
Snyder, L. J., and Stewart, W. E. AIChE J. 12, 167 (1966).
Soppe, W. Powder Technology 62, 189 (1990).
Srensen, J. P., and Stewart, W. E. Chem. Eng. Sci. 29, 819 (1974).
Spedding, P. L., and Spencer, R. M. Comput. Chem. Eng. 19, 43 (1995).
ANTHONY G. DIXON ET AL. 388
Stephenson, J. L., and Stewart, W. E. Chem. Eng. Sci. 41, 2161 (1986).
Stevenson, P. Chem. Eng. Sci. 58, 5379 (2003).
Stitt, E. H., pp. 185216 in Sustainable strategies for the upgrading of natural gas: fundamentals,
challenges and opportunities (Derouane, E., Parmon, V., Lemos, F. and Ramoa-Ribiero F.,
Eds.), NATO Science Series, Vol. 191, Springer, Dordrecht, Netherlands (2005).
Suekane, T., Yokouchi, Y., and Hirai, S. AIChE J. 49, 10 (2003).
Taskin, E., Dixon, A. G., and Stitt, E. H., submitted to Numerical Heat Transfer A (2006).
Theuerkauf, J., Witt, P., and Schwesig, D. Powder Technology 165, 92 (2006).
Thompson, K. E., and Fogler, H. S. AIChE J. 43, 1377 (1997).
Tierney, M., Nasr, A., and Quarini, G. Sep. Purif. Technol. 13, 97 (1998).
Tobis , J. Chem. Eng. Sci. 55, 5359 (2000).
Tobis , J. Chem. Eng. Comm. 184, 71 (2001).
Tobis , J. Ind. Eng. Chem. Res. 41, 2552 (2002).
Utyuzhnikov, S. V. Computers & Fluids 34, 771 (2005).
Van Baten, J. M., Ellenberger, J., and Krishna, R. Chem. Eng. Sci. 56, 813 (2001).
Van der Merwe, D. F., and Gauvin, W. H. AIChE J. 17, 519 (1971).
Von Scala, C., Wehrli, M., and Gaiser, G. Chem. Eng. Sci. 54, 1375 (1999).
Wegner, T. H., Karabelas, A. J., and Hanratty, T. J. Chem. Eng. Sci. 26, 59 (1971).
Winterberg, M., Tsotsas, E., Krischke, A., and Vortmeyer, D. Chem. Eng. Sci. 55, 967 (2000).
Yakhot, V., and Orszag, S. A. J. of Sci. Comput. 1, 1 (1986).
Yang, A., Miller, C. T., and Turcoliver, L. D. Phys. Rev. E. 53, 1516 (1996).
Yuen, E. H. L., Sederman, A. J., and F Gladden, L. Appl. Catal. A. 232, 29 (2002).
Yuen, E. H. L., Sederman, A. J., Sani, F., Alexander, P., and Gladden, L. F. Chem. Eng. Sci. 58, 603
(2003).
Zeiser, T., Lammers, P., Klemm, E., Li, Y. W., Bernsdorf, J., and Brenner, G. Chem. Eng. Sci. 56,
1697 (2001).
Zeiser, T., Steven, M., Freund, H., Lammers, P., Brenner, G., Durst, F., and Bernsdorf, J. Phil.
Trans. R. Soc. Lond. A 360, 507 (2002).
PACKED TUBULAR REACTOR MODELING AND CATALYST DESIGN 389
INDEX
A
Agglomeration, 198
Aggregation, 274, 275, 277282, 287289
Algebraic stress model, 163
Anisotropy, 184185
B
Batchelor length scale, 167, 213
Batchelor scale, 241, 242
BeemanVerlet algorithm, 98
Blending, 151, 153, 190, 219
Blending, suspending, 183
Boundary methods, 180, 181, 191, 219
Boussinesq, 160, 163, 196
Breakage, 252, 275, 278281, 287289
Break-up, 170, 199, 203206, 209
Bubble column, 2, 8, 11, 16, 18, 21, 24, 58
Bubble dynamics, 2
Bubble number density, 204
C
CarnahanStarling Equation, 108
Catalyst particles, 308, 310312, 315, 330,
352, 356, 363364, 367, 372, 374,
376378, 380381, 384
Chemical reaction engineering (CRE),
231233, 236, 244, 245, 253, 298, 310
Chemical reactions, 156157, 166, 183,
190, 209210, 217219, 310, 315, 319,
355
Coalescence, 157, 170, 203206, 209,
218219
Coarse-grid simulation, 136
Coefcient of normal restitution, 95
Cohesive force, 96, 110
Collision, 157, 168, 174175, 177, 193194,
198, 202203, 218
Collision operator, 77
Compartmentalization, 199
Composition vector, 267272, 274, 276,
277, 285, 287
Computational uid dynamics, 307, 310,
385
Computational uid dynamics (CFD)
code, 232, 233, 235, 237, 238, 244,
247253, 259, 266, 267, 269, 274, 288,
291, 298, 300
model, 232238, 244246, 248250, 252,
253, 257, 261263, 267269, 271,
273275, 277, 279, 281, 282, 286288,
292301
simulation, 238, 239, 243, 24, 252, 24,
255, 272, 279, 282, 283, 300, 301
Computer-generated packing, 325326
Contact force, 90
Continuum, 169170, 175, 189
Cross-correlation terms, 166, 167, 210
Crystallization, 151, 193, 197, 219
Crystallizer, 173, 193, 198, 199, 202203
D
Damko hler number, 209, 213214
Direct numerical simulation (DNS), 3,
151152, 160, 193, 217, 235, 237, 238,
239, 244, 245, 253, 287, 295
Direct quadrature method of moments
(DQMOM), 249, 268272, 277,
282284, 286289, 301
Discrete bubble model, 141
Discrete particle model, 72, 86
Dispersed, 167170, 189, 204, 219
Dispersed two-phase, 167, 169
Dispersion, 310, 314, 334, 343, 352355,
381383
Dissipated, 180
391
Dissipation, 154, 162164, 170, 183, 190,
193, 203204, 206, 214
Dissolution, 196197, 200, 219
DNS, 34, 24, 156157, 159160, 193194,
202, 209, 216219
Drag, 157, 168, 195196
Drag, virtual mass, 169
Droplet deformation, 59
Droplet evaporation, 28, 40
Dropletparticle collision, 3, 2728, 49, 55
D3Q19 model, 79
E
Eddy viscosity, 162165, 184
Energy spectrum
scalar, 241, 242
turbulent, 240, 247
Enskog theory, 117
Ergun equation, 84
EulerEuler, 167, 169, 170, 195, 204
EulerLagrangian, 167, 196
Eulerian, 152, 165167, 169170, 189190,
205, 213
Excess compressibility, 107, 109
F
Favre average, 294, 295, 298,
FHP model, 75
Film boiling, 2829, 31, 3839, 58
Film-boiling evaporation, 2, 27
Filtered drag coefcient, 138
Filtered particle-phase pressure, 138, 139
Finite element, 308, 312, 315, 355356, 386
Finite volume, 151, 159, 171, 174, 176, 308,
315316, 324, 341, 355, 386
Fixed bed reactor, 308310
Flow transitions, 312
Fluidized bed, 2, 11, 13, 24, 58
Front-capturing method, 4
Front-tracking method, 4
FV, 159, 171, 174178, 181, 191, 216, 219
G
Gas liquid solid ow, 3, 5, 7, 9, 11, 13, 15,
17, 19, 21, 23, 25, 27, 29, 31, 33, 35,
37, 39, 41, 43, 45, 47, 49, 51, 53, 55,
57, 59, 61, 63
Gasliquid, 151, 203
Gasliquid ow, 11
Gasliquidsolid ow, 1, 3, 13
Gasliquidsolid uidization, 2, 14, 24
Gassolid drag force, 83
Gas-uidized beds, 66
Geldart A particles, 127
Granular temperature, 115
Growth, 252, 274277, 279, 281, 282, 284,
287289
H
Hamaker constant, 96, 110
Hard-sphere model, 86
Heat transfer, 28, 3334, 3940, 5859,
308310, 312315, 318319, 321, 323,
330, 332334, 337340, 342, 344345,
356365, 367, 371372, 376, 378,
382384
History, 168
Hold-up, 204208
Hydrodynamics, 2, 4, 28, 34, 58
I
Immersed boundary method, 1, 3, 9
Impeller swept, 179180, 182183,
191, 204
Interaction force, 169170, 194, 195196,
219
Interstitial ow, 307, 311312, 315, 348,
351, 381
Isotropic, 155, 161, 184
Isotropy, 160, 170, 183184, 218
J
Jacobi matrix, 125
K
ke model, 164, 171, 183184, 212
ke turbulence model, 195, 207, 216
ko model, 164
Kinetic theory of granular ow, 69,
119
Knudsen number, 40, 42,
59, 67
Kolmogorov, 154, 159, 167, 195, 202, 205,
210213
INDEX 392
Kolmogorov
length scale, 238, 239, 241, 273, 274, 281
time scale, 238, 253, 281
L
Lagrangian, 152, 165170, 189, 194, 203,
211212, 214
Large eddy simulations (LES), 151152,
157, 160161, 186, 194, 217, 235,
238, 240, 244, 295
Lattice, 151, 175176, 197
Lattice Boltzmann (LB), 151, 159,
174175, 314, 328, 343, 351, 354355,
386
Lattice Boltzmann Model, 72, 74
Lattice-gas models, 74
LB, 159, 174175, 181182, 184, 186,
190191, 197, 199200, 209, 214,
216217, 219
LB LES, 194
LES, 157, 159162, 194197, 214, 216219
Level-set method, 23, 56, 8, 1213,
2930, 39, 5051, 58
Liquidliquid, 151, 203, 209
Local grid renement, 176, 182, 219
M
Macroinstabilities, 154, 158, 165, 188, 189
Magnus, 168, 194
Magnus, history, 169
Mass transfer, 193, 197, 204206
Mesh generation, 324, 332, 336
MFR, 179, 191
Micromixing, 152, 166167, 210211, 214,
219, 245, 246, 248250, 262, 265, 268,
269, 273, 275277, 282, 284288,
299301
Microscopic
balance equation, 233235, 240, 301
model, 233, 234, 240, 275
transport equation, 250, 254, 259, 267,
268, 275277, 279, 283, 284
Mixing time, 190192, 209
Mixture fraction
mean, 250, 257, 266, 299
transport equation, 257
variance, 245, 248, 250, 257, 266, 299
Moments, 199, 208
Momentum exchange coefcient, 84,
102
Multi-environment model, 248, 250, 261,
278, 284, 285
Multiuid model, 234, 288, 292, 293, 295,
296, 299
Multiphase ow
model, 234, 237, 244, 287, 291, 294296,
299
reacting, 233, 234, 244, 288, 297, 301
Multiple frames of reference (MFR), 179
N
NavierStokes, 159
NavierStokes equations, 100
Neighbor listsm, 98
Nucleation, 193, 198199, 202203,
274278, 280282, 284, 287
Number density function (NDF)
bivariate, 282, 286, 301
univariate, 274, 279, 283
Numerical diffusion, 172
Nusselt number, 298
O
ObukhovCorrsin constant, 241, 243
One-way coupling, 165, 167, 203
P
Packed bed, 309310, 313, 325, 328, 334,
342, 344, 351, 354, 357
Packed tube, 307308, 315317, 321,
324328, 330, 332, 335337, 340, 342,
344345, 348, 352, 354, 356, 364365,
381383
Parallel, 159, 171, 174, 197, 200, 210, 212,
214, 219
Parallellization, 216
Particle-image velocimetry (PIV), 246
Particle size distribution (PSD), 252, 274,
277, 279, 281, 282, 287, 288
PDF, 167, 199, 210, 213, 214, 219
Periodic box, 157, 161, 193194, 202, 209,
211, 216, 218
Planar laser-induced uorescence (PLIF),
241, 246
INDEX 393
Poincare , 190
Point particles, 167168, 194
Population balance, 206, 208209, 219
Population balance equation (PBE), 274,
275, 279, 282, 289, 301
Prandtl number, 250
Precipitation, 151, 197, 219
Pressure drop, 309, 313315, 321, 335,
337338, 340, 342344, 348352, 357,
364, 367, 383
Probability density function (PDF)
method, 248, 257, 268, 284
model, 281, 299
presumed, 248, 257, 261, 299
transported, 261, 268, 272, 273, 280
Product-difference (PD) algorithm,
276278, 282
Q
Quadrature method of moments
(QMOM), 276, 281, 282
R
Random, 168169, 202, 211
RANS, 160, 163, 195197, 215217
Reaction engineering, 336
Reaction-progress variable, 257259, 262,
266
Reaction rate, 213
Reaction thermal effects, 372, 378
Reactor, 199, 209
Relaxation, 169, 172, 204206
Residence time distribution, 209
Reynolds average, 245, 246, 278, 294,
297299
Reynolds averaged NavierStokes
(RANS), 151, 159, 163, 195
Reynolds-averaged NavierStokes
(RANS) models, 235, 238, 240, 241,
244246, 253
Reynolds number
particle, 237, 291, 296, 298
turbulent, 238240, 244, 266, 281
Reynolds stress model (RSM), 163,
206
RSM, 171, 212
RSM turbulence model, 164
S
Saffman force, 194
Saffman lift, 169
Saffman lift force, 168
Saturated droplet impact, 29
Scalar dissipation, 241, 242, 250
Schmidt number, 240244, 247, 250, 266,
300
Selectivity, 153, 209, 214, 217
SGS, 160162, 168169, 176, 184186,
188190, 200, 202, 213214, 216, 218
SGS eddy viscosity, 162, 165
SGS stresses, 161162
Sherwood number, 298
SIMPLE algorithm, 120
Single-phase ows
model, 244, 21, 287, 291, 295
nonreacting, 233, 244, 253
reacting, 233, 234, 253, 299
Sliding, 179
Slip, 176, 180, 204
SM, 179, 191
Smagorinsky, 162, 184185, 188
Smagorinsky SGS, 162, 183
Snapshot, 180, 207
Soft-sphere model, 87
Solids suspension, 155, 170, 193, 209
Species transport, 166, 172, 176, 191, 213
Springdashpot soft-sphere model, 90
Steam reforming, 308309, 333, 358,
363364, 367, 370, 372, 374376, 378
Stick-boundary rules, 81
Stokes number, 273275, 288, 295
Stresses, 162164, 167168, 170, 184, 191
Subgrid, 160, 171172, 213
Subgrid-scale
model, 233, 234, 243245, 250, 251, 23,
272, 280, 298301
phenomena, 234, 236, 238
Subgrid stresses, 162
Suspending, 151, 153, 192, 219
T
The Lattice BhatnagarGrossKrook
model, 78
Tracking, 166168, 170, 175, 189190,
194195, 211212, 217
INDEX 394
Turbulence, 203, 314315, 317, 319324,
334336, 341342, 350351, 368, 382,
385
Turbulence model, 165, 171, 183,
218
Turbulence spectrum, 169, 209, 217
Turbulent kinetic energy, 154, 162164,
183184, 186190
Turbulent kinetic energy dissipation, 199
Two-uid, 169170, 196, 207
Two-uid model, 71, 111
Two-Phase, 151, 161, 167, 170171, 173,
196, 202, 208, 217
Two-phase DNS, 203
Two-phase turbulence model, 170
Two-way coupling, 168169, 194,
202203
U
Unresolved, 166, 168, 191, 193,
218219
Unresolved SGS, 166
V
Variance
mixture-fraction, 245, 248, 250, 299
scalar, 241
Virtual mass force, 142
Voke, 162, 184185
Volume-of-uid, 212
Vorticity, 164165, 211
W
Wall effects, 309310, 327, 343344,
350, 354355, 381
Wall function, 321324, 337338, 342, 348,
368, 372, 382
Wall segment, 330, 378, 386
Wen and Yu eq., 84
Y
Yield, 153, 209215, 217
Z
Zwietering, 153, 193
INDEX 395

You might also like