You are on page 1of 20

Chemical Engineering Science 61 (2006) 12171236

www.elsevier.com/locate/ces
CFDsimulations of two stirred tank reactors with stationary
catalytic basket
P. Magnico
a,
, P. Fongarland
b
a
Laboratoire de Gnie des Procds Catalytiques, CNRS-ESCPE Lyon, 43 bd du 11 Novembre 1918, B.P. 2077, 69616 Villeurbanne Cedex, France
b
Laboratoire de Catalyse de Lille, USTL, 59655 Villeneuve dAscq Cedex, France
Received 24 September 2004; received in revised form 29 April 2005; accepted 8 July 2005
Available online 3 October 2005
Abstract
Among the different systems used for laboratory kinetic investigation, stationary catalytic basket stirred tank reactors (SCBSTRs) allow
one to study triphasic reactions involving shaped catalyst with large size. The hydrodynamics of these complex reactors is not well known
and has been studied experimentally in only a few cases. Despite the difference in the design of two commercial SCBSTRs reported in
these works, the local measurements of the liquidsolid mass transfer coefcient inside the catalytic basket revealed the same velocity
prole. The aim of the present work is therefore to investigate more accurately the hydrodynamics of the two reactors by means of CFD
in order to compare the effect of the blade/bafe hydrodynamic interaction on the ow pattern. Owing to the geometrical complexity of
the reactors, the hydrodynamic investigation is based on the kc model and the BrinkmanForsheimer equations. The agreement at the
local level with the experimental data (PIV and mass transfer measurements) validates this preliminary work performed with the standard
values of the parameters present in the turbulent model and the BrinkmanForsheimer equations. The simulations reveal in both reactors
a ring-shaped vortex around the impeller in the agitation region. The high axial location of its centre induces a reverse ow at the tips of
the basket. Owing to the uid friction in the porous medium, the azimuthal ow in the core region is transformed into a radial ow in
the basket where the ow decreases abruptly. Vertical vortices are located at the blade tips and at the downstream face of the bafes or
they are located in the basket on both sides of the bafes, depending on the design and the location of the bafes. At the inner radius
interface of the basket, the vertical blade impeller induces a rather homogeneous velocity prole, but the pitched blade impeller imposes a
high velocity at the plane of symmetry. Therefore the simulations demonstrate that two different local velocity patterns and two different
porous media may induce the same mass transfer properties.
2005 Elsevier Ltd. All rights reserved.
Keywords: Computational uid dynamic; Particle image velocimetry; Stationary catalytic basket; Continuous stirred tank reactor; Mass transfer
1. Introduction
Several kinds of laboratory reactors are used to study the
kinetics of triphasic reactions catalysed by a solid phase:
triphasic xed bed reactors (trickle-bed or packed bed),
structured reactors like monolith or microreactors, and also
perfectly mixed reactors with catalyst in suspension or main-
tained in a basket (Perego and Peratello, 1999; Dudukovic

Corresponding author. Tel.: +33 04 72 43 17 63;


fax: +33 04 72 43 16 73.
E-mail addresses: pmo@lobivia.cpe.fr, magnico@arcadie.cpe.fr
(P. Magnico).
0009-2509/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2005.07.025
et al., 2002). In petroleum engineering, and especially in
the case of hydrotreatment reaction, trickle-bed reactors are
generally employed for kinetic investigations because they
are used in industrial processes. But these integral reactors
are not the most adapted for kinetic measurement. Continu-
ous stirred tank reactors may be preferred since they give the
apparent rate directly by the measurement of inlet and out-
let concentrations. Furthermore, detection of mass or heat
transfer limitations is easier in this kind of reactor, which
are a major concern to measure intrinsic kinetic.
Contrary to slurry perfectly mixed reactors where cata-
lyst is suspended, a catalytic basket stirred tank is set up
in order to use shaped catalytic particles with a size greater
than 1 mm. The solid particles, too heavy to be suspended,
1218 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
are xed in a static annular basket around an impeller
(Arroyo et al., 2000, 2001; Goto and Saito, 1984) or in a ro-
tating basket moved by a shaft (Goto and Saito, 1984; Turek
and Winter, 1990; Teshima and Ohashi, 1977; Warna et al.,
2002). For the case of the stationary catalytic basket stirred
tank reactor (SCBSTR), also called RobinsonMahoney
reactor (Mahoney et al., 1978), the hydrodynamics of the
liquid phase is a major concern even for these laboratory
reactors. Since the stationary basket is placed around the
impeller, a signicant effect of the basket on the ow pat-
tern induced by the impeller can be expected. This may
have an important impact on the liquidsolid mass transfer
all around the basket.
Few experimental hydrodynamic characterizations have
been carried out until now for these reactors (Pavko et al.,
1981; Mitrovic, 2001; Fongarland, 2003). The reason lies in
the opacity of the xed bed inside the catalytic basket, which
prevents one investigating close to the turbine. Two SCB-
STRs commercialized byAutoclave Engineers (AE) and Parr
Instruments have been studied. The velocity eld of the AE
reactor has been investigated in the region between the bas-
ket and the external tank wall by means of the particle im-
age velocimetry (PIV) technique (Mitrovic, 2001). In order
to investigate the velocity eld inside the basket, the local
liquidsolid mass transfer has also been studied by means of
dissolution of naphthol pellets. The second reactor has been
investigated as well by the mass transfer characterization
(Fongarland, 2003). The PIV data reveal a high heterogene-
ity of the velocity eld inside the basket. Despite the differ-
ence on the impeller and the bafe design, the experimental
results also show that the two reactors have similar local
properties of mass transfer. In fact the mass transfer measure-
ments are not accurate enough to estimate the supercial ve-
locity prole because (a) it is necessary to determine a Sher-
wood correlation to link the velocity to the liquidsolid mass
transfer coefcient (k
s
), (b) it allows one to investigate the
eld of the supercial velocity module only, and (c) the spa-
tial resolution of the experiments is the basket width which
is much larger than the scale of the velocity heterogeneity.
Therefore the interest of the present work is to use the
computational uid dynamic (CFD) approach to go deeper
in the investigation and to describe the hydrodynamics in
the catalytic basket and in the region round the impeller.
Moreover, until now no hydrodynamics study has been car-
ried out by means of CFD in xed basket stirred tank re-
actors. Only a brief study has been performed in the case
of a spinning basket reactor (Warna et al., 2002). We show
in the present paper that (a) the CFD approach allows one
to quantify the hydrodynamic interaction of the blades with
the bafes, (b) the same liquidsolid mass transfer proper-
ties can be obtained with two different ow patterns and two
different porous media, and (c) the velocity eld inside the
basket cannot be compared with the outside velocity even
close to the outer basket surface.
The simulations are carried out by means of the commer-
cial software package FLUENT for three rotation speeds.
The use of the numerical approach in order to characterize
such a complex reactor (presence of a heterogeneous porous
medium, complexity of the impeller design, close distance
between the blades and the bafes, multiphase transport)
represents a challenge, particularly at the interfaces of the
catalytic basket and in the region close to the tank wall where
the turbulence is not well established. Therefore the goal of
the present work is to sketch the mean ow pattern inside
the catalytic basket and around the impeller. In order to suc-
ceed, several simplications are carried out on the reactor
design and on the hydrodynamic model (assumption of ow
periodicity, stationary ow, use of the kc model) despite
the non-periodic and non-stationary ow and despite the
anisotropic turbulence close to the impeller. But, the agree-
ment at the local level with the experimental data (hydrody-
namics and mass transfer) validates this preliminary work
performed with the standard values of the parameters present
in the turbulence model and the BrinkmanForsheimer
equations.
The paper is organized as follows. In Section 2 we
describe briey the geometry of the two reactors and the
experimental methodologies. In Section 3, the computa-
tional domains are described. The modelling approach and
its limitations are also explained. In Section 4, the mesh
sensitivity is checked. The simulations are also validated
against the PIV data. The velocity eld in the basket and
close to the impeller is described in Section 5 for the two
SCBSTRs. In a nal section, the agreement of the computed
k
s
with experimental data conrms the validation of the
numerical approach. The supercial velocity proles inside
and outside the basket of the two reactors are compared. In
this section, the discussion sheds light on the inadequacy of
the solid dissolution rate measurements to characterize the
velocity eld in the basket.
2. Reactor geometry and experimental investigations
2.1. Reactor geometry
The AE reactor, which has a volume of 0.9 l, was inves-
tigated by Mitrovic (2001) (see Fig. 1). It has an inner di-
ameter of 8 cm and a height of 18 cm. The uid is stirred
by a six-blade radial-type turbine and a pair of four 45

pitched blade axial-type turbines (Fig. 1a). The diameter of


the blades is 3.2 cm. The radial turbine is 5.5 cm high and
0.2 cm thick; the axial turbine is 1 cm high and 0.2 cm thick
(Fig. 1b and c). The catalytic basket has an inner diameter
and an outer one of 4.25 and 5.75 cm, respectively. The bas-
ket height is 9 cm. Two horizontal rings (0.9 cm width) close
the top and the bottom of the basket. Two kinds of bafes
equip the reactor. Four outer bafes (9 cm0.4 cm) are lo-
cated along the catalytic basket at the outer interface. Four
inner bafes (9 cm 1.6 cm) are inside the basket and in
the stirred region. The basket and the turbines are located at
11.4 cm respectively from the top of the reactor. The gas is
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1219
(4)
(5)
(3)
(1)
(2)
(1)
(2)
(6)
15
(7)
(a) (b) (c)
Fig. 1. Sketch of the AE reactor: (1) six-blade radial-type turbine; (2) four pitched blade turbine; (3) catalytic basket; (4) internal bafe; (5) external
bafe; (6) ring closing the basket at the top and at the bottom.
injected from the bottom of the reactor, but at the top there
is no gas/liquid interface.
The Parr reactor investigated by Fongarland (2003) has
a volume of 300 ml (Fig. 2). It has a diameter of 6.4 cm
and a height of 10 cm. The interface gas/liquid is located at
8.2 cm from the reactor base. The uid is stirred by a gas-
inducing turbine which is composed of three 17

pitched
blades (Fig. 2a). The blade dimensions are 2.4 cm in diam-
eter, 4.3 cm high and 0.2 cm thick. The gas is transferred to
the liquid phase through the shaft by means of two holes
located above the gas/uid interface and at the horizontal
plane of symmetry of the blades. The basket has an inner
diameter of 3.4 cm, an outer diameter of 5 cm and a height
of 5.1 cm. As in the AE reactor, two horizontal rings (0.3 cm
width) close the basket. The reactor has two kinds of bafes.
Three inner bafes of cylindrical form are located inside the
basket and three outer ones are located along the tank wall
(5.1 cm height, 4 mm width and 3 mm large). The bottom of
the blades and of the basket are located at 3.9 and 1.3 cm,
respectively, from the bottom of the reactor.
2.2. Experimental methodologies
The velocity measurements were performed by means of
the PIV method in the region between the basket lled with
PMMA cylinders (0.61 cm long and 0.217 cm in radius)
and the vessel wall. Three vertical planes of measurement
(6 cm6 cm) were located at the azimuthal angle 0 of 35

,
0

and +35

from one of the inner bafes. To obtain a pro-


le all along the basket at each angle, two planes dened
by the axial location 0 cm<z <6 cm and 3 cm<z <9 cm
from the bottom of the basket were necessary. In the shared
zone (3 cm<z <6 cm), the data accuracy was evaluated and
each velocity component was averaged. In order to measure
the azimuthal components, a vertical plane tangent to the
basket grid is needed. The intersection between two perpen-
dicular planes forms a straight line along which the velocity
eld was analysed. The axial component (Z direction) mea-
sured in the two planes was also averaged. The measure-
ment frequency was 10 s
1
. The statistical average was per-
formed over 200 measurements which were not locked with
the impeller orientation. Therefore the velocity data are av-
eraged over all the impeller orientations. The agitation speed
c ranged from 500 to 1500 rpm.
In order to estimate the performance of the SCBSTR,
Mitrovic and Fongarland measured the liquidsolid mass
transfer coefcient with the dissolution of naphthol grains
in water and in n-Heptane at standard conditions of temper-
ature and pressure. Two approaches were used: the global
approach consisting in lling all the basket with naph-
thol pellets (see Tables 1 and 2) and the local approach
consisting in lling at an axial position only a 1 cm thick
slice, the remainder of the basket being lled with inert
grains of PMMA. The local experiments in the AE reactor
gave a characteristic time of dissolution in water ranging
from 9 10
3
to 1.7 10
4
s and a time of dissolution in
n-Heptane ranging from 1600 to 3100 s, depending on the
agitation speed. In the Parr reactor lled in n-Heptane, the
characteristic time ranged from 300 to 1200 s. These times
are much greater than the macromixing one (few seconds)
measured experimentally. To prove the coherence between
the hydrodynamic measurement outside the basket and the
measurements of the local dissolution rate, Mitrovic has
determined, for both uids, a Sherwood correlation which
links the local supercial velocity and the liquidsolid mass
transfer coefcient of naphthol. The parameters were de-
termined by means of a cylindrical xed bed reactor, the
length of which was equal to the basket width (Mitrovic,
2001). The Reynolds number Re
ks
ranged from 2 to 55. The
correlation used in the AE reactor, after a more accurate
analysis, seems to give good predictions of the liquidsolid
mass transfer coefcient of naphthol in n-Heptane, but a
large deviation is observed for k
s
values in water.
1220 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
(6)
(5)
(4)
(1)
(2)
(1)
(1c)
(1b)
(3)
(a) (b) (c)
Fig. 2. Sketch of the Parr reactor: (1) three pitched blade gas-inducing-type turbine; (1b) gas suction hole; (1c) gas ejection hole; (2) catalytic basket;
(3) ring closing the basket at the top and at the bottom; (4) external bafe; (5) internal bafe; (6) liquid/gas interface.
Table 1
Geometrical characteristics of the grains and values of the permeability of the catalytic basket inside the AE reactor
d
cyl
(m) L
cyl
(m) d
p
(m) d
ks
(m) : (1/m
2
) [ (1/m)
PMMA 0.00217 0.0061 0.002765 0.0033 0.47 5.3 10
7
3233
Naphthol 0.002 0.0058 0.00256 0.00259 0.47 6.193 10
7
3490
Table 2
Geometrical characteristics of the grains and values of the permeability of the catalytic basket inside the Parr reactor
d
cyl
(m) L
cyl
(m) d
p
(m) d
ks
(m) : (1/m
2
) [ (1/m)
PMMA 0.00217 0.0061 0.00276 0.0033 0.6 1.453 10
7
1172
Naphthol 0.002 0.0058 0.00256 0.00259 0.6 1.695 10
7
1266
3. Numerical methodology
The interaction of the bafes with the blades makes the
ow non-stationary. Macroinstabilities coming from the in-
teraction of the impinging jet with the tank wall, the swirling
ow at the tank bottom or the shedding of the trailing vortex
from the impeller, for example, also generate non-stationary
ows. The macroinstabilities have a frequency much lower
than the impeller rotation one, depending on the tank and
the impeller design. Therefore the ow pattern cannot be
periodic. Hydrodynamics is also characterized by complex
turbulent properties such as anisotropy, unclear frequential
separation in the power spectrum between the macroinsta-
bilities and the turbulent uctuations (Galletti et al., 2005).
The complete study of the ow pattern requires outstand-
ing numerical tools and large computational requirements
(Roussinova et al., 2003; Hartmann et al., 2004). To the best
of the authors knowledge, most of these studies are carried
out in simple reactors such as the standard mixing vessel
stirred by a Rushton impeller. Owing to the complexity of
the SCBSTR geometry and therefore of the ow, we decided
to simplify the design and to use an unsophisticated hydro-
dynamic approach in order to study the mean ow pattern
only.
3.1. Geometry of the computational domain
In order to reduce the computational requirements, geo-
metrical simplications are carried out in the following way.
In the AE reactor, we replace the conical-shaped base by a
horizontal one and we do not take into account the basket
support ((7) in Fig. 1c). In the Parr reactor, we do not take
into account the gas volume and the free surface gas/liquid
at the top of the reactor. We assume that the impeller, the
basket and the bafes are located at the centre of the tank.
Therefore, the domain in the two reactors can be divided into
two symmetrical parts by a central horizontal plane, and the
simulations are performed in the lower half of the reactor.
Owing to the geometrical periodicity in the azimuthal direc-
tion, the computational domain is reduced to a sector with
periodic boundary conditions set at the two opposite vertical
planes. The computational domain of the Parr reactor has
dimensions of 6.4 cm, 4.1 cm and 60

in radial, axial and


P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1221
azimuthal directions, respectively. The dimensions of the AE
reactor domain are 4 cm, 9 cm and 180

in radial, axial and


azimuthal directions, respectively.
Versicco et al. (2004) showed that the periodic assump-
tion cannot be applied to the uid ow and gives inaccurate
results. In particular, the authors found that the radial veloc-
ity is overestimated if the simulations are performed on a
domain lower than 1/4th of the tank in the azimuthal direc-
tion. They used an immersed boundary method coupled with
the direct numerical simulation (DNS) and a Reynolds av-
erage NavierStokes (RANS) approach to compute the ow
eld at a Reynolds number (Re
tip
) of 1636. In the past, sev-
eral authors used the periodic boundary conditions in order
to reduce the computational cost (Ranade, 1997; Brucato
et al., 1998; Sheng et al., 1998). The numerical results at
Reynolds numbers higher than 9000 do not show any over-
estimation of the radial component of the mean velocity. In
the present study, half and 1/3rd of the vessel volume in the
azimuthal direction are used for the AE and the Parr reactor,
respectively. Therefore the recommendation made by Vers-
icco et al. is respected. If the periodic boundary conditions
are not used, the computational requirement would be too
large: 2 250 000 and 3 300 000 cells are needed for the
AE and the Parr reactor respectively, in this case (see below).
From now, we dene the inner zone as the cylindrical
region delimited by the inner radius face of the catalytic
basket. Its bottom is located at the base of the ring below
the basket. In the AE reactor the bottom is at z = 4.5 cm,
and in the Parr reactor it is at z = 1.5 cm. The outer zone
denes the volume of the reactor minus the inner zone and
the porous medium. The basket faces dened by the inner
radius and the outer one are named inner boundary and outer
boundary, respectively. The bottom of the basket denes the
origin of axial coordinate.
3.2. Turbulence model
In the experimental works described in Section 2, the im-
peller rotation speed ranged from 500 to 1500 rpm, giving a
Reynolds number Re
tip
running from 13 000 to 40 000 and
from 9000 to 25 000 for the AE and the Parr reactor, respec-
tively. Therefore the turbulence can be assumed well estab-
lished in most of the reactor volume if the reactor is suitably
bafed and if the free gas/liquid surface is not too deformed
(Nagata, 1975). But even if the Reynolds number range is
high enough around the impeller, we will see that the small
permeability of the catalytic basket prevents the uid from
travelling through with a sufciently high velocity in accor-
dance with the experimental observations. The turbulence
cannot be well established in the outer region and it should
be relevant to use a turbulent model available at moderate
Reynolds number.
In the zone around the impeller the anisotropic turbulence
would be suitably modelled by large eddy simulation (LES)
(Hartmann et al., 2004; Revsteldt et al., 1998; Sheng et al.,
2000). The Reynolds stress model (RSM) has the advan-
tage of being simpler and taking into account the turbulence
anisotropy. But, as in the kc model, the turbulence scales
are not resolved properly. Moreover, all the RANS models
give the same mean ow eld accurately (Campolo et al.,
2003; Sheng et al., 1998), but underestimate the turbulent
kinetic energy process and do not model accurately the mix-
ing (Yeoh et al., 2005). The interest of LES is to solve by
means of DNS the large scales of the anisotropic motions.
The small scales are supposed to be universal and are mod-
elled by a sub-grid model. The cut-off length scale lies in
the inertial sub-range of turbulence and the two scales are
clearly separated by means of a lter, the width of which
is of the order of the grid size (Germano, 1992). But in the
presence of a porous medium, in which the ow is laminar,
only the kc and kc RNG models, used outside the basket,
are implemented in the software FLUENT.
3.3. Hydrodynamic model in the catalytic basket
The catalytic basket is lled with cylinders. The annular
xed bed is assumed to be an effective mediumcharacterized
by its porosity and permeability. The laminar ow eld is
computed with the BrinkmanForsheimer equations which
couple the NavierStokes equations with the Ergun ones in
the following way (Ergun, 1952; Macdonald et al., 1979):
j
jt

U +

U


U =

P
j

v
K

U v

U, (1)
where
1
K
=: +
1
v
[|

U| (2)
with
: =150
(1 )
2

3
d
2
p
and [ =1.75
(1 )

3
d
p
. (3)
In expression (3), d
p
is the equivalent diameter of a spher-
ical particle. For a cylinder the diameter has the following
expression:
d
p
=
6L
cyl
2 +4L
cyl
/d
cyl
. (4)
Tables 1and 2 show the characteristic parameters of the
porous medium inside the catalytic basket for each SCB-
STR. The empirical relation (2) is available in the range of
Reynolds number Re
p
less than 1000. In the simulations,
the Reynolds number reaches a maximum value of 340 at
1500 rpm. Therefore the use of the BrinkmanForsheimer
equations (1) is justied.
Expressions (2) and (3) imply the assumption of isotropy
and homogeneity of the medium. The hypothesis of effective
medium is justied if the size of the xed bed is much
larger than the grain size (Magnico, 2003). In the SCBSTR,
the basket thickness to the cylinder length ratio is around
1222 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
4. Therefore the xed bed should be considered as a two-
phase structure in which the ow eld is computed at the
pore scale. Considering the pellet size and their random
spatial arrangement, the basket boundaries cannot be plane
and must present an important roughness. Moreover the pore
mouths constitute the holes at the interfaces. Therefore, at the
outer boundary, the ow would be constituted of jets coming
from the pores and in the inner boundary where the ow
is essentially azimuthal, the roughness would inuence the
ux through the boundary. These considerations are all the
more important because the anisotropic shape of cylinders
increases the heterogeneity of the packing close to the basket
boundaries. The above assumptions (effective medium with
homogeneous boundaries and isotropy) must be considered
as very rough. But separating the two phases in the catalytic
basket would require too large resources. The anisotropic
assumption may be introduced by dening the parameters :
and [ as second-order tensors. The problem is therefore to
guess the value of each term of the tensors.
3.4. Numerical method
Two numerical methods are commonly used in stirred re-
actors in order to reproduce the effect of the impeller rota-
tion: the multiple reference frame (MRF) approach (Luo et
al., 1994) and the sliding mesh approach. In the two meth-
ods, the reactor is divided into two regions: (1) a cylindrical
one containing the impeller and (2) the remainder of the re-
actor. The momentum equations are solved in the rotating
frame reference in region (1) and in the laboratory frame
reference in region (2). The MRF method solves the steady
state momentumequations, considering the impeller position
as xed. In the case of a strong impeller bafe interaction
or a mixing structure analysing, the sliding mesh approach
is the most appropriate because informations are exchanged
through the sliding interface at each time step. In the reac-
tors studied here, the inner bafes and the blades are close
to each other. Therefore, the sliding mesh approach is nec-
essary to compute the velocity eld inside the inner region.
But with the software version 6.1, the sliding mesh fails in
this context producing overows at the rst iterations what-
ever the value of the time step. When periodic surfaces and
non-conformal interfaces are present in the computational
domain, the computation of the new position of the nodes
during the sliding mesh simulation may fail (private com-
munication with FLUENT France). In the region between
the catalytic basket and the tank wall, we will see that the
velocity eld is not very sensitive towards the blades ori-
entation, allowing one to assume that in this zone the MRF
method is accurate enough. The interface position between
the two regions (1) and (2) must be far from the blade tips
in order to minimize its effect on the numerical solution
(Oshinowo et al., 2000). No static solid must be present in
region (1) and the interface must be of cylindrical symme-
try. Therefore, in order to move the interface away from the
blade tips, the interface is located along the inner radius of
the basket in the Parr reactor and at the inner bafe edges in
the AE reactor. But the interface position remains too close
to the blade tips. For the two reactors, the base of region (1)
is located at the bottom of the tanks.
The mesh adopted in the AE reactor is prismatic around
the radial turbine, tetrahedral around the pitched blade tur-
bine and hexahedral elsewhere. The cell number is 247 362.
Along the three vertical straight lines where the PIV data
are analysed, the mesh is rened in order to have a better
numerical accuracy locally. The rened volume is a cylin-
der 0.6 cm in radius and 4.5 cm high. In the Parr reactor,
the mesh is tetrahedral around the blades and the inner baf-
es, and hexahedral elsewhere. The cell number is 300 000.
No-slip conditions are dened at the blades wall and the
standard wall treatment used in turbulent ows is imposed
at the tank wall. The momentum equations are solved with
the coupling SIMPLE algorithm and the second upwind dis-
cretization scheme. The pressure is computed by means of
the PRESTO scheme.
4. Hydrodynamic validation
In this section, the numerical methodology is validated
with the AE reactor by checking the mesh sensitivity and
by comparing the numerical results against the PIV data.
Therefore the basket and the reactor are assumed to be lled
with PMMA cylindrical particles and water, respectively.
The values of the parameters dening the permeability of
the porous medium are reported in Tables 1 and 2.
4.1. Mesh sensitivity
In order to check the mesh independence, the renement
is performed in three ways: (a) in regions where the velocity
gradients are large, (b) around the surface separating the two
MRF zones, and (c) along the straight lines where PIV mea-
surements were performed. In the last case the renement
region is a cylinder 6 mm in radius and 4.5 cm high. Fig. 4
displays the radial proles of the three components of the
normalized velocity at different axial locations in a vertical
plane dened by 0=60

, midway between two blades (see


Fig. 3a). In the domain of computation, the inner bafes are
located at 80

and +10

, and the outer bafes are located


at 35

and +55

. The impeller rotation speed is 1500 rpm


(U
tip
= 2.51 m/s). The axial locations of the radial proles
are shown in Fig. 3b which displays the velocity vector eld
in the vertical plane (2a). Owing to the high velocity in the
inner region, the vector length is set constant all over the
plane. Every other vector is skipped. The proles obtained
with renement (a) and (b) are compared with the proles
computed without any renement.
The renement (a) takes place along the blades, the inner
bafes and the wall at the bottom of the tank. The nal cell
number is 397 000. This renement has an effect on the radial
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1223
Fig. 3. (a) Sketch of the blades/bafes conguration in the computational
domain of the AE reactor: (1a) plane of periodicity; (2a) vertical plane
located at 60

. (b) Vector velocity eld in the plane (2a). The rotation


speed is 1500 rpm. The vector length is set constant. Every other vector
is skipped. (1) Radial location of the blade tips, (1

) axial location of the


blade bottom, (2) inner boundary, (3) outer boundary, (4) reactor wall, (5)
ring closing the basket tips. (a) z =0.65 cm; (b) z =1.5 cm; (c) z =2 cm;
(d) z =2.5 cm; (e) z =4.5 cm (symmetry plane location).
component near the symmetry plane only in the inner region.
At the surface separating the two MRF zones, the azimuthal
component displays a small discontinuity in the vicinity of
the plane of symmetry. As in (a), the renement (b) has no
major effect on the velocity. It decreases the discontinuity of
the azimuthal component a little. In the outer region and in
the porous medium, the velocity proles remain unchanged.
Along the PIV measurement lines, the velocity is sensitive
to the local renement (c) if the lines are located in front of
the inner bafes. The sensitivity is less than 10% depending
on the blade orientation. In the kc model, the production of
turbulent energy and the turbulent viscosity are expressed in
terms of the velocity gradient and are more sensitive to the
mesh renement. Therefore, renement (b) is chosen to de-
scribe the spatial properties of the turbulence (see Section 5).
-0.2
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
r (m)
U
'
r

(1) (4) (3) (2)
(e)
(d)
(c)
(b)
(a)
adaptation inside the inner region
adaptation vs velocity gradients
-0.95
-0.85
-0.75
-0.65
-0.55
-0.45
-0.35
-0.25
-0.15
-0.05
0.05
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
r (m)
U
'
t

(1) (4) (3) (2)
(e)
(d)
(c)
(b)
(a)
-0.3
-0.1
0.1
0.3
0.5
0.7
0.9
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
r (m)
U
'
z

(1) (4) (3) (2)
(e)
(d)
(c)
(b)
(a)
(a)
(b)
(c)
Fig. 4. Radial prole of the normalized velocity in the vertical plane
located at 60

at several axial locations displayed in Fig. 3b (AE


reactor). The rotation speed is 1500 rpm. (a) Radial component; (b) axial
component; (c) azimuthal component. (1), (2), (3), (4), (a), (b), (c), (d),
(e): see Fig. 3.
4.2. Validation using PIV data
The stationary hypothesis of the ow imposed by the
MRF approach can be circumvent by computing the ow
eld for different angular orientations of the impeller. The
ow eld computed at each orientation is considered as in-
stantaneous. In the AE reactor, an accurate computation is
necessary to compare the results of the simulations against
experimental data. The ow eld is computed for four im-
peller orientations 0
(blade/bafe)
: 0

, 7.5

, 15

, 22.5

where 0
(blade/bafe)
is the azimuthal angle between a vertical
blade and an inner bafe arbitrarily chosen together. Each
orientation corresponds to an instantaneous conguration.
We do not take into account the azimuthal orientation of the
pitched blades because we assume that they induce a global
effect of uid pumping and do not inuence the ow locally
in the porous medium and in the outer region. Therefore the
1224 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
-0.02
0
0.02
0.04
0.06
0.08
0.1
0.12
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
U
r

(
m
s
-
1
)
Z (m)
1000 rpm 1500 rpm
500 rpm
-0.2
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 U
z

(
m
s
-
1
)
Z (m)
1000 rpm
1500 rpm
500 rpm
(a)
(b)
Fig. 5. Axial prole of the velocity averaged over four blade/bafe ori-
entations in the AE reactor at 4 mm from the catalytic basket in front of
the inner bafes. (a) Radial component; (b) axial component. Lines with
symbols: experiments; full lines: numerical simulations.
impeller has an azimuthal periodicity of 30

. The experi-
mental measurements were located at 35

, 0

and +35

from an inner bafe. The two inner bafes present inside the
computational domain are geometrically equivalent. There-
fore, the stationary velocity proles along the catalytic bas-
ket are averaged over height instantaneous ones.
In Fig. 5, the computed stationary velocity proles, along
the basket in front of an inner bafe and at 4 mm from the
outer boundary, are compared to the experimental ones. The
radial and the axial components of the velocity are repre-
sented for the three agitation speeds. The origin of the axial
coordinate is the bottom of the basket. The vertical bars rep-
resent the variation amplitude of the instantaneous velocity
at z = 0, 2.25, 4.5, 6.75 and 9 cm. Fig. 5b also reports the
difference between the axial component measured in the ra-
dial plane and in the azimuthal one by means of vertical bars
at z =1.5, 3.5, 5.5 and 7.5 cm.
In Fig. 5a, the experimental curves display local maxima,
the axial positions of which are mostly independent of the
impeller rotation speed. This observation means certainly
that the ow at the outer boundary is not homogeneous
owing to the presence of pore mouths, which would induce
ow jets. Flow disequilibrium between the lower and the
upper part of the basket can be observed in this gure and in
Fig. 5b. The ux is more important in the upper part. The
difference of the radial velocity between the two parts com-
pared to the mean value is 74% (500 rpm), 4% (1000 rpm)
and 38% (1500 rpm). In Fig. 5b, the ux disequilibrium
also induces a smaller axial velocity at the top of the basket
-0.05
-0.025
0
0.025
0.05
0.075
0.1
0.125
0.15
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
U
r

(
m
s
-
1
)
Z (m)
+35
-35
Fig. 6. Axial prole of the radial velocity in the AE reactor at 4 mm
from the catalytic basket and at 35

and +35

from the inner bafes.


Rotation speed: 1500 rpm. Lines with symbols: experiments; full lines:
numerical simulations.
compared to the velocity at the bottom and an offset of
the experimental curves of 1 cm to the bottom. This ux
inhomogeneity can be explained by the heterogeneity of the
packing which would be more permeable in its upper part.
But this cannot explain the erratic variation of the disequi-
librium with the agitation speed because the same packing
was used. The other explanation is the discrepancy of the
measurements between the upper and the lower vertical
plane of PIV measurement as observed in the dissertation
(Mitrovic, 2001).
In the case of the radial velocity, a good accordance be-
tween the simulations and the experimental data is displayed
at the centre of the porous medium. Close to the tips of the
basket, the computed radial velocity curves decrease more
rapidly than the experimental ones. On the contrary, the axial
component displays a discrepancy in the region z =4.5 cm
owing to the dissymmetry of the experimental curves and a
good accordance at the tips of the basket. But at 1500 rpm,
a discrepancy appears also in these regions. The numerical
simulations have shown that the ow eld in the outer re-
gion is sensitive to the blade orientation despite the porous
medium. The amplitude of variation of the radial compo-
nent ranges from 13% to 46% at z = 4.5 and 2.25 cm and
from 160% to 450% at z =0 cm in the range of the impeller
rotation speed investigated in this work.
Fig. 6 displays the proles of the radial component of the
velocity at 0 = 35

and +35

from an inner bafe. The


agitation speed is 1500 rpm. As in Fig. 5, the vertical bars
represent the variation amplitude of the instantaneous veloc-
ity. The proles show that the amplitude of the maximum
is slightly dependent on the azimuthal location. But, at the
tips of the basket, the negative values of the radial velocity
are greater than the ones near the inner bafe. This means
that far from the inner bafe, the uid enters the basket from
its outer boundary at its tips. The radial velocity is posi-
tive only close to the inner bafe. Therefore the simulations
agree with the experimental observations but with a higher
amplitude especially in the case of 0 =35

. Two reasons
can explain this observation:
(a) The numerical simulations locate the main vortex at
2.5 cm above the bottom of the basket (and 2.5 cm below
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1225
the top of the basket in the symmetrical part of the reactor)
in the inner zone. The axial location may be too high. It
depends on the permeability of the packing (see Section 5.1
for the discussion).
(b) The PIV measurements were performed in coinci-
dence with an angular orientation of the blades. At 1500 rpm,
the impeller performs two and a half rotations during 0.1 s.
Therefore, if we neglect azimuthal orientation of the pitched
blades, the velocity measurements were carried out with
the same impeller position. This observation is also valid
for the two other rotation speeds. The comparison with the
simulations at 1500 rpm gives an optimal agreement for the
0
(blade/bafe)
value of 15

. But this angle cannot be the same


at each impeller rotation speed. Unfortunately, this cannot
be checked because the experimental data at +35

and 35

are not available at the other agitation speeds.


5. Description of the ow eld
5.1. Autoclave engineers reactor
Inside the inner region two kinds of vortex characterize
the uid ow: (a) a main ring-shaped vortex around the im-
peller (Fig. 3b), with a radial location around 1.61.7 cm
and a mean axial one around 2 cm (i.e., 1.35 cm above the
bottom of the vertical blades), and (b) secondary vortices
following the blade tips. No trailing vortices take place be-
hind the blades as in the case of the Rushton impeller (Lee
and Yianneskis, 1998).
The radial position of the main vortex is independent of
the azimuthal angle 0. In Fig. 4a (c = 1500 rpm, U
tip
=
2.5 m/s), at z =2 cm the sign of the radial velocity changes
at r =1.6 cm. Above and below this axial position the curves
have opposite sign, conrming the main vortex position. At
the inner boundary, the radial velocity decreases abruptly in
the axial region 2 cm<z <2.5 cm. Above the main vortex,
the radial velocity ranges from 0.08 U
tip
to 0.055 U
tip
. Be-
low, the velocity is constant and equal to 0.02 U
tip
and
remains negative inside the catalytic basket. This means that
the uid goes through the basket from the outer boundary.
The vortex size is important enough to induce a rotational
movement to the uid inside the porous medium so that at
the outer boundary the radial component velocity is negative
for z <2 cm.
The size and form of the vortex depend on its azimuthal
position towards an inner bafe. Fig. 7 displays the colour
map of the radial velocity at the outer boundary in the
case of the instantaneous blade/bafe orientation shown in
Fig. 3a. Upstream the inner bafe located at 0 =+10

, the
vortex size decreases as the azimuthal position is closer to
the bafe so that at the bottom of the basket the radial veloc-
ity is less and less negative. Just upstream the inner bafe
(+10

<0 <+20

), the radial velocity is positive all along


the axial position because the vortex is so small that all
over the porous media thickness (2.13 cm<r <2.88 cm) the
Fig. 7. Colour map of the radial component of the velocity normalized
by U
tip
at the outer interface (r = 2.88 cm) in the AE reactor in the
case of the blades/bafes conguration sketched in Fig. 3a. The rotation
speed is 1500 rpm. (1) Bottom of the interface; (2) top of the interface;
(3) angular position of the blade at 0 = 30

; (4) angular position of


the blade at 0 = +30

; (5) outer bafe at 0 = 35

; (6) outer bafe at


0 =+55

; (7) inner bafe at 0 =+10

.
radial velocity remains positive. This is visualized in Fig. 8,
which displays the velocity vector eld in three horizontal
planes axially located on both sides of the main vortex. Only
the vectors located at r >1.7 cm are reported. In front of the
bafe (rung with dotted lines), the vortex disappears. In the
region z <2 cm close to the bafe (Fig. 8a and b), the uid,
pushed azimuthally to the upstream face of the inner bafes,
is deviated to the porous medium. Therefore the ow runs
through the basket radially along these bafes. But the pres-
sure downstream the inner bafes is so low that the uid,
going outside the basket, moves round these bafes and en-
ters the basket through the outer boundary. The low pressure
induced by the bafes amplies the uid ux at the tips of
the basket and increases the vortex intensity (see Fig. 7). It
also induces an important azimuthal ow close to the inner
bafes in the outer region.
The position of the vortex depends on the permeability
of the packing as displayed in Fig. 9. Simulations with dif-
ferent values of grain size reveal that the axial and the ra-
dial locations increase with the packing permeability, i.e., it
displaces away from the blade tips to the symmetry plane.
Fig. 9a displays the velocity vector eld in the case of a
grain size d
p
equal to 8 times the size of the PMMA parti-
cles. The position of the vortex is at the inner boundary and
at 3.5 cm above the bottom of the basket. Therefore, the neg-
ative magnitude of the radial velocity at the outer boundary
increases with the permeability. If the grain size is 2 times
lower than the PMMA size (Fig. 9b), the vortex is displaced
down to the vertical blade bottom. The low axial position
1226 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
Fig. 8. Vector velocity eld on three horizontal planes (with r >1.7 cm) inside the AE reactor with the blade orientation displayed in Fig. 3a. The rotation
speed is 1500 rpm. The vector length is set constant. Every other vector is skipped. (a) z =1 cm; (b) z =2 cm; (c) z =4.5 cm.
of the vortex induces a positive radial velocity at the tips of
the basket. But the vortex interacts with the ow induced by
the pitched blades so that its form and its position depend
on the azimuthal position.
In order to decrease the ow from the outer boundary at
the tips of the basket, we have to slide the vortex down, i.e.,
we must decrease the permeability without decreasing the
ux through the basket. Two trials were carried out. First,
the tips of the basket (0 cm<z <1.5 cm) were assumed
less permeable (:

= 2:, [

= 2[). The numerical results


showed that the velocity eld remains identical. Second, the
porous medium lled with cylindrical particles should be
anisotropic especially close to the tips of the basket. Simu-
lations were performed assuming that the second-order ten-
sors :
ij
and [
ij
are diagonal in the cylindrical reference.
Using :
rr
= :, [
rr
= [, :
00
= :
zz
= 2:, [
00
= [
zz
= 2[ in
the entire porous medium, the computed velocity prole re-
mained also unchanged all along the basket. Therefore the
numerical simulations are sensitive to :
rr
and [
rr
only. But
changing their value also changes the ux through the porous
medium.
Secondary vertical vortices are located in front of the
blades close to the inner boundary (rung with full lines in
Fig. 8). But their axial location depends on the blade az-
imuthal position towards an inner bafe. These vortices are
produced at the downstream face of the inner bafes at the
horizontal plane of symmetry (Fig. 8c). When a blade passes
in front of one of these bafes, it catches the vortex from
this bafe. During the blade displacement, the vortex fol-
lows the blade tip and moves down in the axial direction
until it reaches the blade bottom. When the blade reaches
the following inner bafe, the vortex disappears. Owing to
the small distance between the blade tips and the basket, the
vortex induces an azimuthal ow inside the porous medium.
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1227
Fig. 9. Velocity eld at 1500 rpm for two particle diameters. The vector
length is set constant. Every other vector is skipped. (a) d
p
=2.2 cm; (b)
d
p
= 0.138 cm; (a) axial location of the main vortex centre in the case
of PMMA particles; (b) symmetry plane; (1), (1

), (2), (3), (4), (5): see


Fig. 3.
The production of the turbulent kinetic energy and the
turbulent dissipation is induced by the local shear stress.
Fig. 10 shows the radial variation of the normalized tur-
bulent kinetic at several axial positions at 1500 rpm in the
same vertical plane as in Fig. 3 and with the same blades
and bafes locations. The curves display two sources of tur-
bulent kinetic energy: the main vortex centre and the inner
boundary. In the range of agitation speed studied, the max-
imum value of k

is around 0.2 and around 0.1 at the inner


boundary and in the main vortex, respectively. At the inner
boundary, the maximum is located at z 22.5 cm which
corresponds to the axial position of the main vortex. At this
position the velocity eld is azimuthal and axial inside the
inner zone and becomes suddenly radial and axial inside the
porous medium. This sudden change of direction creates a
large velocity gradient, which induces a transfer of the ki-
netic energy from the mean ow to the turbulent one. On
both sides of this axial location, the normalized kinetic en-
ergy decreases because the ow eld is more and more ra-
dial in the inner region as the axial position deviates from
the maximum location. In the catalytic basket, the turbu-
lent kinetic energy vanishes because the ow is assumed to
be laminar. In the outer zone, k

remains null meaning that


the porous medium makes the ow laminar in this zone,
i.e., all the turbulent energy is dissipated inside the turbulent
zone.
5.2. Parr reactor
The main vortex is located at z=1.3 cm (i.e., 0.8 cm above
the blade bottom) and at r = 1.4 cm. In Fig. 11 the radial
proles of the radial component velocity are displayed at
several axial positions on both sides of the vortex location.
In the domain of computation, the pitched blade is oriented
at 0 = +30

, the bafes are located at 0

and the verti-


cal plane is midway between two blades, i.e., at 0 = 30

(Fig. 11a). The curves at z = 2.6 and 0.45 cm represent


the prole at the symmetry plane and at the blade bot-
tom, respectively. The impeller rotation speed is 1500 rpm
(U
tip
=1.88 m/s). Fig. 11b displays the velocity vector eld
in the plane (2a) and shows the axial position of the curves
in Fig. 11c with respect to the main vortex. Fig. 11c dis-
plays the vortex position by the change of the sign of the
radial velocity at z = 1.5 cm and r = 0.15 cm and by the
opposite sign of the proles at z = 2 and 1 cm, i.e., on
both sides of the vortex centre. The curves (a) and (b) show
that below the main vortex (0 cm<z <1 cm), the uid is
pumped through the porous medium to the impeller. Con-
trary to the AE reactor, above the main vortex (curves (d)
and (e)), the radial velocity increases rapidly with the ax-
ial position and reaches a maximum in the axial direction
at the symmetry plane. This maximum at the radial position
(1) (U
r
(r =1.2 cm) =0.43 U
tip
) is much greater than in the
case of the AE reactor. The computation conrms that the
uid is radially ejected along the symmetry plane owing to
1228 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
r (m)
k
'
(1) (4) (3) (2)
(e)
(d)
(c)
(b)
(a)
Fig. 10. Radial proles of normalized turbulent kinetic energy at several axial positions in the plane (60

), midway between two vertical blades inside


the AE reactor (see Fig. 3a). The rotation speed is 1500 rpm. (1), (2), (3), (4), (a), (b), (c), (d): see Fig. 3.
Fig. 11. Flow pattern in the plane (0 =30

), midway between two pitched blades inside the Parr reactor. The rotation speed is 1500 rpm. (a) Sketch of
the blades/bafes orientation. (1a) Planes of periodicity, (2a) vertical plane location (0 = 30

). (b) Vector velocity eld in the plane (2a). The vector


length is set constant. Every other vector is skipped. (1) Radial location of the blade tips, (1

) axial location of the blade bottom, (2) inner boundary,


(3) outer boundary, (4) reactor wall, (5) ring closing the basket tips, (a) z =0.5 cm; (b) z =1 cm; (c) z =1.5 cm; (d) z =2 cm; (e) z =2.5 cm. (c) Radial
prole of the normalized velocity at different axial locations displayed in (b).
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1229
Fig. 12. Vector velocity eld on three horizontal planes (with r >1.6 cm) inside the Parr reactor with the blade orientation displayed in Fig. 11a. The
rotation speed is 1500 rpm. (a) z =0.2 cm; (b) z =0.5 cm; (c) z =1.5 cm.
the design of the blades and proves the efciency of the im-
peller to introduce gas bubbles in a stirred reactor. The sud-
den decrease is also observed along the inner boundary. At
the symmetry plane the radial velocity is equal to 0.3 U
tip
and decreases to 0.05 U
tip
at z =2 cm. The velocity proles
induced by the two impellers are very different at the inner
boundary. But we will see that the liquidsolid mass transfer
is the same in the two reactors.
Fig. 12 displays the velocity vector eld in three planes
below the main vortex and for a radial distance higher than
1.5 cm. In the inner zone, the ow eld is mostly azimuthal.
But close to the bottom of the basket the ow is radial
and azimuthal (Fig. 12a). As we move closer to the main
vortex centre along the axial direction, the ow becomes
azimuthal and axial. Inside the porous medium, the ow is
mostly radial and axial. Below the main vortex, two vortices
take place on both sides of each inner bafe in the region
0.5 cm<z <1.25 cm (Fig. 12b). They induce an azimuthal
ow over the width of the porous medium. But the ow
remains mostly radial at the inner boundary. Above the main
vortex (z >1.25 cm), the change of the ow direction occurs
inside the porous medium over a depth much smaller than
the grain size (Fig. 12c). Around the plane of symmetry, the
impeller imposes a radial ow in the inner region and in the
porous medium (see Fig. 11c).
The important change of owdirection at the inner bound-
ary induces an intense shear stress and therefore an impor-
tant turbulent production as shown in Fig. 13. In this gure
the vertical plane is the same one as in Fig. 11a. All the tur-
bulence is dissipated inside the inner region. But contrary
to the AE reactor, there is no other maximum of turbulent
energy at the main vortex location. In the range of agitation
speed investigated, the power number N
p
is computed as-
suming that the dissipation takes place mainly in the inner
zone. N
p
is dened as
N
p
=N

V
rz
jc dV
j(2c/60)
3
d
5
imp
,
1230 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
0
0.05
0.1
0.15
0.2
0.25
0.3
0 0.004 0.008 0.012 0.016 0.02 0.024 0.028 0.032
r (m)
k
'
(1)
(4) (3)
(2)
(e)
(d)
(c)
(b)
(a)
Fig. 13. Radial proles of normalized turbulent kinetic energy at several axial positions in the plane (30

), midway between two vertical blades inside


the Parr reactor. The rotation speed is 1500 rpm. (1), (2), (3), (4), (a), (b), (c), (d): see Fig. 11.
where N is the ratio of the reactor volume to the computa-
tional domain one and V
rz
is the volume of the inner zone.
N = 4 for the AE reactor and N = 6 for the Parr reactor.
For the two reactors N
p
is found to be constant and equal to
9.0 5% for the three agitation speeds. Observing the high
intensity of the turbulent kinetic energy and its high gradi-
ent close to the inner boundary, a renement of the meshing
is carried out as mentioned in Section 4.1. The kc RNG
model is also used at 1500 rpm. This turbulent model com-
putes the same ow eld as the kc model. On the contrary,
the magnitude of the turbulent quantities k and c, averaged
over the inner zone, decreases by a factor of 2.2 and 3.2,
respectively. But these quantities cannot be measured inside
the inner zone. For a more detailed analysis, a robust model
of turbulence and the sliding mesh approach must be used.
6. Liquidsolid mass transfer properties
6.1. Proles computed by CFD and comparison with
experimental data
The local measurement of k
s
allows one to estimate a
velocity prole inside the basket and to compare it to the
prole computed by CFD. However, the velocity deduced
from the Sherwood correlation is very sensitive to the Sher-
wood number (U Sh
2
) and to experimental uncertainties.
Therefore, it is relevant to compare the measured and the
computed k
s
proles.
In order to compute the dissolution time in each slice and
to compare this time to the experimental data, the dissolved
species transport over the entire reactors must be simulated
with the liquidsolid mass transfer. But the MRF approach
does not conserve the mass of dissolved species and the local
velocity used by the software inside the MFRregion (1) is the
relative velocity and not the absolute one. Therefore the k
s
Table 3
Physical properties of water of n-Heptane at standard condition
j (Pa s) j (kg/m
3
) D (m
2
/s)
Water 0.001 1000 7.1 10
10
n-Heptane 0.0004 684 2.56 10
9
proles are computed by means of the local velocity module
inside the porous medium and the Sherwood correlation.
Then the k
s
proles are compared to the experimental k
s
estimated from the time to reach 95% of the equilibrium
concentration (Fongarland, 2003) or from the time evolution
of the dissolved naphthol concentration (Mitrovic, 2001).
In the numerical simulations, the permeability difference
between the slice of naphthol and the remaining packing
lled with PMMA is not taken into account.
The liquidsolid mass transfer coefcient is determined
by means of the Sherwood number, which has the general
expression
Sh =
k
s
d
ks
D
=a +bRe
c
ks
Sc
d
, (5)
where Sc is the Schmidt number, Re
ks
is the Reynolds num-
ber dened with d
ks
, the diameter of a sphere which has the
same surface as a naphthol cylinder. The values of the four
parameters a, b, c and d determined by Mitrovic are 0, 1.1,
0.48 and 0.33, respectively, for Reynolds numbers Re
ks
rang-
ing from 2 to 55. These values were computed with diffusion
coefcient values of 1.17 10
9
and 3.44 10
10
m
2
s
1
in water and n-Heptane, respectively. But these are not the
values computed by means of the Wilke and Chang correla-
tion under standard conditions of temperature as mentioned
in the dissertation (see Table 3). Therefore another expres-
sion of the Sherwood number was found (Fig. 14):
Sh =1.1Re
0.41
ks
Sc
0.4
. (6)
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1231
0.0E+00
1.0E-05
2.0E-05
3.0E-05
4.0E-05
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
n-heptane
water
U (ms
-1
)
K
s

(
m
s
-
1
)
Mitrovic's correlation (2001)
Present work
Fig. 14. Variation of the mass transfer coefcient with the supercial velocity. Comparison of Mitrovics correlation and of the corrected one against
experimental data (Mitrovic, 2001).
0.0E+00
2.0E-05
4.0E-05
6.0E-05
8.0E-05
1.0E-04
1.2E-04
1.4E-04
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
n-heptane
water
k
s
(
m
s
-
1
)
Z (m)
1500 rpm 1000 rpm
500 rpm
1500 rpm
1000 rpm
500 rpm
0.0E+00
2.5E-05
5.0E-05
7.5E-05
1.0E-04
1.3E-04
1.5E-04
0 0.01 0.02 0.03 0.04 0.05
k
s
(
m
s
-
1
)
Z (m)
1500 rpm 1000 rpm
500 rpm
(a)
(b)
Fig. 15. Axial proles of the liquidsolid mass transfer coefcient k
s
of naphthol for three rotation speeds: (a) in n-Heptane and water in the AE reactor;
(b) in n-Heptane in the Parr reactor. Symbols: experiments; lines: simulations.
The k
s
proles inside the catalytic basket in the AE re-
actor are displayed in Fig. 15a. for water and n-Heptane.
The experimental accuracy of k
s
was estimated at 12%.
The simulations overestimate the mass transfer coefcient
in water but a good agreement can be observed in the case
of n-Heptane. Mitrovic compared the velocity measured by
PIV in the vicinity of the symmetry plane to the interstitial
velocity deduced from the Sherwood correlation and the
porosity. The interstitial velocity was used because the ve-
locity oscillations along the basket were assumed to be gen-
erated by ow jets, i.e., the PIV velocity is the velocity in-
side the pores. So only the maxima have been taken into
account. But Mitrovic did not succeed in matching the PIV
velocity to the velocity deduced from k
s
. The PIV veloc-
ity was always greater by a factor ranging from 1.7 to 3.2,
depending on the agitation speed. In fact the velocity mea-
sured outside the catalytic basket is not representative of the
inside velocity owing to the high heterogeneity of the ow
eld in the basket and especially at the inner boundary as
explained below in Section 6.2.
1232 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05
Blade axial position
inner boundary
outer boundary
S(r = 0.02733 m)
porous medium
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0.18
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
U

(
m
s
-
1
)
U

(
m
s
-
1
)
Z (m)
Z (m)
inner boundary
outer boundary
S(r = 0.0335 m)
porous medium
Blade axial position
(a)
(b)
Fig. 16. Axial proles of the velocity module in the Parr reactor. Rotation speed: 1000 rpm. (a) AE reactor; (b) Parr reactor.
In the simulations the velocity outside the basket has to
be compared to the supercial one inside the basket because
the packing is an effective medium. Using relation (6) with
the global experimental values of k
s
, the mean supercial
velocity inside the basket is 0.0115, 0.0145 and 0.05 m/s for
agitation speeds of 500, 1000 and 1500 rpm, respectively.
At 1000 rpm, the simulations give a mean value of 0.07 m/s.
Such a difference (a factor of 5) should induce a large dis-
agreement of the numerical results with the PIV data.
Therefore the k
s
measured in water is far too small than
what is expected by the correlation used in the simula-
tions. Several reasons can be mentioned to explain this dis-
crepancy: (a) The correlation was determined in a range of
Reynolds number smaller than the range investigated in the
AE reactor. The correlation must be extrapolated in order
to be used in the SCBSTR. (b) Mitrovic observed a drift of
concentration during the experiments of dissolution in water
in the AE reactor. This drift came from the increase of the
solvent temperature. Each of these experiments took several
hours and a mean value of the equilibrium concentration
was used to estimate the dissolution rate. No drift was ob-
served in the case of dissolution in n-Heptane, and the sim-
ulations predict a mass transfer coefcient in n-Heptane in
accordance with the experiments.
It must be mentioned that the computed velocity in n-
Heptane is only 5% upper than in water, even if n-Heptane
has a dynamic viscosity 41% greater than water. Hydrody-
namics depends on the physical properties of the uid if
the Reynolds number is small. Therefore in the outer re-
gion and in the porous medium, the ow should depend on
the solvent. But in the rst region where the ow is weakly
turbulent, the kc model gives a turbulent viscosity value
10 times the laminar one even at 500 rpm. In the porous
medium, the physical properties of the uid are introduced
by Darcys term : and by the local pressure gradient term.
The inertial term has a greater contribution than Darcys one
if the velocity module is greater than 0.016 m/s in water and
0.01 m/s in n-Heptane. Therefore Darcys term should have
effects only at the tips of the basket, as we will see below
and in Fig. 16. It was also observed that the global pressure
drop through the basket divided by the density depends on
the impeller rotation speed only. Therefore the computation
is not globally sensitive to the nature of the solvent.
Fig. 15b displays the k
s
proles in the Parr reactor. Each
dot represents a value averaged over three experiments
and the standard deviation, displayed by vertical bars at
1000 rpm, is globally less than 25% of the mean value. An
important disequilibrium between the upper and the lower
part of the basket characterizes the experimental proles
for the three impeller rotation speeds, so that half of the
experimental data at 1500 rpm have a smaller value than at
1000 rpm. This disequilibrium cannot be simulated because
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1233
Table 4
Computed and measured mass transfer coefcient k
s
Water
a
Water
c
n-Heptane
a
n-Heptane
c
n-Heptane
b
n-Heptane
d
500 rpm 1.15 10
5
2.96 10
5
5.73 10
5
6.3 10
5
6.55 10
5
6.52 10
5
1000 rpm 1.84 10
5
3.85 10
5
8.3 10
5
8.5 10
5
9.28 10
5
8.85 10
5
1500 rpm 2.53 10
5
4.55 10
5
10 10
5
9.98 10
5
9.37 10
5
10.3 10
5
a
Mitrovic (2001).
b
Fongarland (2003).
c
This work in the AE reactor.
d
This work in the Parr reactor.
the computational domain is assumed symmetric. If we
make the experimental proles symmetric by averaging the
k
s
values of two slices at equal distance from the symmetry
plane, the computed proles t much better the modied
experimental ones. These two observations explain the dis-
crepancy between the predictions of the simulations and the
experimental data. Table 4 reports the values of k
s
averaged
over the basket (predicted and measured). The simulations
agree with the experiments in the case of dissolution in
n-Heptane. In the two reactors, the predicted k
s
in the two
solvents varies as c
0.41
. This means that the mean velocity
inside the basket is linearly dependent on c. This is also
observed experimentally in the two reactors if the solvent is
n-Heptane. But in the AE reactor, the k
s
measured in water
varies as c
0.48
, and is systematically less than the simu-
lated one. Despite this discrepancy, the simulations conrm
the experimental observations which show that the two
SCBSTRs have the same global solidliquid mass transfer
properties.
The question is now: why the mass transfer properties of
the two reactors are similar even though the ow pattern is
completely different?
6.2. Comparison of the velocity proles inside and outside
the basket
Fig. 16 displays the axial proles of the velocity module
in the porous medium of the two reactors. The velocity is
averaged over 5 mm height slices and all over the azimuthal
angle 0. In Fig. 16a, the curve S(r =0.0335 m) is the prole
at 4 mm from the outer boundary. In Fig. 16b, the curve
S(r = 0.0273 m) is the prole at 2.33 mm from the outer
boundary, i.e., at 1/3rd the distance between the basket and
the tank wall as in the AE reactor. The curves named porous
medium represent the velocity proles averaged also over
the basket width. The agitation speed is 1000 rpm. At the
inner boundary, the velocity values are multiplied by 1/4th
owing to their large magnitude.
At this boundary of the AE reactor, the velocity remains
constant along most of the blade height and has a value
equal to 0.64 m/s(=0.4 U
tip
). On the contrary, in the Parr
reactor the velocity prole has a maximum at the symme-
try plane equal to 0.32 m/s (=0.25 U
tip
). This maximum
means that the uid is ejected from the pitched blades
centre in the radial direction (see Fig. 11c). But averaged
over the inner boundary, the velocity represents 15% and
17% of U
tip
for the Parr and the AE reactor, respectively
(i.e., 0.192 and 0.27 m/s). As explained in Section 5, from
the inner boundary to the outer one, the azimuthal ow is
transformed into a radial one. The azimuthal velocity and
the velocity module drop abruptly with r over a distance
much smaller than the grain size. But the radial velocity de-
creases slowly as 1/r owing to the conservation of the uid
ux through the porous medium. At the outer boundary, the
Parr reactor has a constant velocity of 0.073 m/s (0.06 U
tip
).
In the case of the AE reactor, the prole has a max-
imum located at the symmetry plane with a value of
0.09 m/s(0.05 U
tip
) and the velocity, averaged over the outer
boundary, is equal to 0.06 m/s (0.037 U
tip
). The permeabil-
ity of the AE reactor porous medium is smaller than the
Parr reactor one owing to the smaller value of the porosity
(see Tables 1 and 2). The consequence is a higher decrease
of the velocity from the inner to the outer boundary in the
case of the AE reactor despite the smaller width of the cat-
alytic basket. The other consequence is the smaller value
of the velocity averaged over the porous medium width in
the AE reactor (curves porous medium). It can also be
noticed that the mean velocity over the basket is equal to
0.07 and 0.09 m/s in the AE reactor and in the Parr reac-
tor, respectively. The Sherwood correlation and the mass
transfer coefcient k
s
are weakly sensitive to the velocity
(Sh U
0.41
). So the difference of the Sherwood coef-
cient between the two reactors is of the order of magnitude
of the experimental accuracy, i.e., 12%. Outside, the two
reactors have the same prole: a minimum at the symme-
try plane where the velocity is radial and a maximum at
the tips of the basket where the ow is axial. These pro-
les do not look like the proles of the velocity averaged
over the basket width even at the plane of symmetry. In
the case of the AE reactor the gap between the velocity
outside the basket and the velocity average over the bas-
ket width is 43% near the symmetry plane whatever the
agitation speed. The gap is around 50% in the case of the
Parr reactor.
Therefore the velocity outside the basket is not represen-
tative of the inside velocity owing to the great change of
the ow direction at the inner boundary even close to the
1234 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
symmetry plane. The permeability is much smaller in the AE
reactor. Even if the velocity at the inner boundary is greater
in this reactor than in the Parr reactor, it becomes smaller
when it is averaged over the basket width. But the velocity
difference between the two SCBSTRs cannot be measured
owing to the lack of accuracy of the experimental method.
Moreover, the abrupt drop of the velocity through the bas-
ket width involves a spatial resolution of the order of the
grain size which cannot be reached by the liquidsolid mass
transfer experiments. Therefore the mass transfer approach
is not adapted to local velocity investigations.
Finally, in both reactors, the average of the ow veloc-
ity prole inside the basket gives close mean velocity and
consequently the liquidsolid mass transfer coefcients ex-
perimentally measured are quite the same.
7. Conclusion
The ow eld is studied in two xed basket reactors by
means of the kc model and the BrinkmanForsheimer equa-
tions, assuming that the packing inside the basket is a ho-
mogeneous porous medium. In the two sets of equations,
standard values of the parameters are used. The geometry of
the simulated reactors is simplied: no gas/liquid interface
at the top of the Parr reactor, at bottom of the AE tank, and
assumption of a horizontal plane of symmetry midway from
the top and the bottom of the tank. Despite the use of the
MRF approach and the azimuthal periodicity assumption of
the ow pattern, the numerical results in the AE reactor are
validated by PIV measurements carried out along the cat-
alytic basket for three impeller rotation speeds (500, 1000
and 1500 rpm). Extending the hydrodynamic study to the
liquidsolid mass transfer characterization inside the basket,
the validation of the numerical approach is conrmed one
more time by the agreement with the measurements of the
local k
s
in the two SCBSTRs.
The simulations show that the ow structure has similar
characteristics in the two reactors. The structure does not
depend on the impeller rotation speed but its intensity is
linearly dependent on c. A ring-shaped vortex around the
impeller is located in the inner zone. Its position depends on
the permeability of the packing inside the basket. Using the
experimental conditions, the simulations localize the vortex
at 2 cm above the bottom of the basket in the AE reactor and
at 1.5 cm in the Parr reactor. The axial position of the vortex
induces a ow from the outer boundary to the inner one at
the tips of the basket. This was also observed by the PIV
measurements but with a lower amplitude. This difference
may be due to the high axial position of the main vortex
computed by CFD. Owing to the friction of the uid with
the packing, the porous medium induces a great change of
ow between the inner zone and the outer one. In the inner
zone the ow is mostly azimuthal and the turbulent energy
is dissipated at the inner boundary. In the outer zone, the
laminar ow is radial and axial. The porous medium, in
which the ow is also laminar, represents a transition region.
Close to the inner boundary the azimuthal component of the
velocity decreases abruptly over a distance much smaller
than the particle size and becomes negligible at the outer
boundary.
But the design of the blades and of the inner bafes
induces differences in the inner zone and in the porous
medium. In the AE reactor, vertical vortices are located at
the blade tips and at the downstream face of the bafes. The
vertical vortices induce locally an azimuthal ow over the
packing width. At the inner boundary, the vertical blade im-
peller induces a rather homogeneous velocity prole. At the
outer boundary, the velocity prole displays a maximum at
the plane of symmetry, i.e., the ow is much smaller at the
basket tips. In the Parr reactor, the vertical vortices are lo-
cated in the basket on both sides of the bafes. The pitched
blade impeller imposes a high velocity at the plane of sym-
metry. But compared to the AE reactor the ow seems to be
more homogeneous at the outer boundary and the reverse
ow is located in a thinner region at the basket tips.
In order to compare the mass transfer efciency of the
two reactors, the liquidsolid mass transfer was also inves-
tigated by Mitrovic. Coupling the k
s
measurements with a
Sherwood correlation, the velocity eld inside the basket
was also studied. But the experimental results display simi-
lar proles despite the difference of the two reactor designs.
In fact the geometry acts on the ow eld but also on the
compacity of the packing. At the inner boundary, the veloc-
ity in the AE reactor is greater than in the Parr reactor. But
owing to the smaller permeability of the AE reactor porous
medium, the velocity decreases more rapidly from the inner
to the outer boundary. The simulations show that the prole
of the velocity averaged over the basket width is identical in
the two reactors, demonstrating that two different local ve-
locity patterns with two different porous media may lead to
the same liquidsolid mass transfer properties. The simula-
tions also show that the high decrease of the velocity eld at
the inlet boundary imposes a spatial resolution smaller than
the particle size and that the k
s
measurement is inappropri-
ate to velocity investigations.
The geometrical complexity of the SCBSTR induces spe-
cic problem for the numerical model: low Reynolds tur-
bulence in the outer zone, high hydrodynamic interaction
between the inner bafes and the blades, and high velocity
gradient in the porous zone especially at the inner bound-
ary. Despite several numerical approximations, the simula-
tions describe the essential part of the hydrodynamics and
allow one to consider an optimization of the SCBSTR ge-
ometry in order to increase the mass transfer coefcient and
to decrease its radial and axial distribution. In order to go
deeper into hydrodynamic study, it will also be necessary to
use the sliding mesh approach, a more realistic turbulence
model and to suggest a better boundary condition at the in-
ner and the outer boundary. This will allow one to study the
dynamics of the hydrodynamics and of the chemical species
mixing.
P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236 1235
Notation
d
cyl
cylindrical diameter, m
d
imp
impeller diameter, m
d
ks
diameter of a sphere which has the same
surface as a cylinder, m
d
p
diameter of a sphere which has the same
specic surface as a cylinder, m
D naphthol diffusion coefcient, m
2
/s
k turbulent kinetic energy, m
2
/s
2
k
s
liquidsolid mass transfer coefcient, m/s
k

=k/U
2
tip
normalized turbulent kinetic energy
K permeability, m
2
L
cyl
cylindrical particle length, m
P pressure, Pa
Re
p
=
Ud
p
v
Reynolds number
Re
tip
=
U
tip
d
imp
v
Reynolds number
Sc =
v
D
Schmidt number
Sh =
k
s
d
ks
D
Sherwood number
U
tip
tip velocity, m/s
U

=U/U
tip
normalized velocity
Greek letters
: Darcy permeability, 1/m
2
[ inertial term of the Ergun correlation, 1/m
c turbulence dissipation rate, 1/m
2
s
3
j dynamic viscosity, Pa s
v kinematic viscosity, m
2
/s
j uid density, kg/m
3
porosity
c impeller rotation speed, rpm
Subscripts
r radial coordinate
t azimuthal coordinate
z axial coordinate
References
Arroyo, J.A.M., Martens, G.G., Froment, G.F., Marin, G.B., Jacobs, P.A.,
Martens, J.A., 2000. Hydrocracking and isomerization of n-parafn
mixtures and a hydrotreated gasoil on Pt/ZMS-22: conrmation of
pore mouth and key-lock catalytic in liquid phase. Applied Catalysis
A: General 192, 922.
Arroyo, J.A.M., Thibaut, J.W., Marin, G.B., Jacobs, P.A., Martens,
J.A., Baron, G.V., 2001. Reaction pathways of 1-Cyclohexyloctane
in admixture with Dodecane on Pt/H-ZMS-22 Zeolite in three phase
hydroconversion. Journal of Catalysis 198, 2940.
Brucato, A., Ciofalo, M., Grisa, F., Micale, G., 1998. Numerical
prediction of ow elds in bafed stirred vessels: a comparison of
alternative modelling approaches. Chemical Engineering Science 53,
36533684.
Campolo, M., Sbrizzai, F., Soldati, A., 2003. Time-dependent ow
structures and Lagrangian mixing in Rushton-impeller bafed-tank
reactor. Chemical Engineering Science 58, 16151629.
Dudukovic, M.P., Larachi, F., Mills, P.L., 2002. Multiphase catalytic
reactors: a perspective on current knowledge and future trends.
Catalysis Reviews 44, 123246.
Ergun, S., 1952. Fluid ow through packed columns. Chemical
Engineering Progress 48, 8994.
Fongarland, P., 2003. Etude cintique de lhydrodsulfuration de composs
soufrs de gazoles en racteur continu parfaitement agit. Ph.D.
Dissertation, Claude Bernard University (Lyon).
Galletti, C., Pagliani, A., Yianneskis, M., 2005. Observations on the
signicance of instabilities turbulence and intermittent motions on uid
mixing processes in stirred reactors. Chemical Engineering Science
60, 23172331.
Germano, M., 1992. Turbulence: the ltering approach. Journal of Fluid
Mechanics 238, 325336.
Goto, S., Saito, T., 1984. Liquidsolid mass transfer in basket type three
phase reactor. Journal of Chemical Engineering of Japan 14, 324327.
Hartmann, H., Deksen, J.J., van den Akker, H., 2004. Macroinstability
uncovered in a Rushton turbine stirred tank by means of LES. A.I.Ch.E.
Journal 50, 23832393.
Lee, K.C., Yianneskis, M., 1998. Turbulence properties of the impeller
stream of a Rushton turbine. A.I.Ch.E. Journal 44, 1324.
Luo, J.Y., Issa, R.I., Gosman, A.D., 1994. Prediction of impeller-induced
ows in mixing vessels using multiple frames of reference. Transaction
of Institute of Chemical Engineers 136, 549556.
Macdonald, I.F., El-Sayed, M.S., Mow, K., Dullien, F.A.L., 1979.
Flow through porous mediathe Ergun equation revisited. Industrial
Engineering Chemical Fundamentals 18, 199208.
Magnico, P., 2003. Hydrodynamic and transport properties of packed beds
in small tube to sphere diameter ratio: pore scale simulation using an
Eulerian and a Lagrangian approach. Chemical Engineering Science
58, 50055024.
Mahoney, J.A., Robinson, K.K., Myers, E.C., 1978. Catalyst evaluation
with the gradientless reactor. Chemtech 8 (N12), 758763.
Mitrovic, M., 2001. Etude des transferts de matires dans un racteur
triphasique gazliquidesolide dinvestigation cintique (racteur
Robinson Mahoney). Ph.D. Dissertation, Claude Bernard University,
Lyon.
Nagata, S., 1975. Mixing: Principles and Applications. Wiley, New York.
Oshinowo, L., Jaworsky, Z., Dyster, K.N., Marshall, E., Nieno, A.W.,
2000. Predicting of tangential velocity eld in stirred tanks using
the multiple reference frames (MRF) model with validation by LDA
measurements. Technical Notes 108 Presented at the 10th European
Conference on Mixing, Delft (The Netherlands).
Pavko, A., Misic, D.M., Levec, J., 1981. Kinetics in three-phases reactors.
Chemical Engineering Journal 21, 149154.
Perego, C., Peratello, S., 1999. Kinetics and modelling of catalytic
reactions: from laboratory to industrial reactor. Catalysis Today 52,
133.
Ranade, V.V., 1997. An efcient computational model for simulating ow
in stirred vessels: a case of Rushton-turbine. Chemical Engineering
Science 52, 44734484.
Revsteldt, J., Fuchs, L., Trgardh, C., 1998. Large eddy simulations of
the turbulent ow in a stirred reactor. Chemical Engineering Science
53, 40414053.
Roussinova, V., Kresta, S., Weetman, R., 2003. Low frequency macro
instabilities in a stirred tank: scale up and prediction based on large
eddy simulations. Chemical Engineering Science 58, 22972311.
Sheng, J., Meng, H., Fox, R.O., 1998. Validation of CFD simulations of a
stirred tank using particle image velocity data. The Canadian Journal
of Chemical Engineering 76, 611625.
Sheng, J., Meng, H., Fox, R.O., 2000. A large eddy PIV method for
turbulence dissipation rate estimation. Chemical Engineering Science
55, 44234434.
Teshima, H., Ohashi, Y., 1977. Particle to liquid mass transfer in a rotating
catalytic basket reactor. Chemical Engineering of Japan 10, 7072.
Turek, F., Winter, H., 1990. Effectiveness factor in a three phase spinning
basket reactor: hydrogenation of butynediol. Industrial and Engineering
Chemical Research 29, 15461549.
1236 P. Magnico, P. Fongarland / Chemical Engineering Science 61 (2006) 12171236
Versicco, R., Fatica, M., Iaccarino, G., Orlandi, P., 2004. Flow in an
impeller-speller tank using an immersed-boundary method. A.I.Ch.E.
Journal 50, 11091118.
Warna, J., Ronnholm, M., Salmi, T., Keikko, K., 2002. Application of
CFD on a catalytic rotating basket reactor. Computer-Aided Chemical
Engineering 10, 10091014.
Yeoh, S.L., Papadakis, G., Yianneskis, M., 2005. Determination of mixing
time and degree of homogeneity in stirred vessels with large eddy
simulation. Chemical Engineering Science 60, 22922302.
Further reading
Sharp, K.V., Adrian, R.J., 2001. PIV study of small-scale ow structure
around a Rushton turbine. A.I.Ch.E. Journal 47, 766778.
Wilke, C.R., Chang, P., 1955. Correlation of diffusion coefcients in dilute
solutions. A.I.Ch.E. Journal V1 (N2), 262.

You might also like