You are on page 1of 12

Chemical Engineering Science 55 (2000) 291}302

On the simulation of stirred tank reactors via computational


#uid dynamics
Alberto Brucato *, Michele Ciofalo, Franco Grisa" , Roberto Tocco 
Dipartimento di Ingegneria Chimica dei Processi e dei Materiali (DICPM), Universita% degli Studi di Palermo, Viale delle Scienze, 90128 Palermo, Italy
Dipartimento di Ingegneria Nucleare (DIN), Universita% degli Studi di Palermo, Viale delle Scienze, 90128 Palermo, Italy
Received 30 March 1999; accepted 1 April 1999

Abstract

Predictions of #ow "elds in a stirred tank reactor, obtained by computational #uid dynamics, were used for the simulation of
a mixing sensitive process consisting of two parallel reactions competing for a common reagent:
A#BPProd.1 A#CPProd.2.
Experimental data were obtained for A"OH\, B"Cu>> and C"ethyl-chloroacetate. For this reaction scheme the "nal

selectivity of the process, easily measured by a simple colorimetric analysis of the residual Cu>>, was found to depend on agitation
speed and therefore on the mixing history during the batch process. The #ow "eld-based three-dimensional simulations performed
here led to predictions that compared very well with the experimental data, though no adjustable parameters were used. Interestingly,
these encouraging results were obtained by modelling only the `macromixinga phenomenon, while `micromixinga phenomena were
neglected, i.e. the system was always considered as being locally perfectly micro-mixed. The good agreement found between simulation
predictions and experimental data retrospectively con"rms the negligibility of micromixing phenomena in the system investi-
gated.  1999 Elsevier Science Ltd. All rights reserved.

Keywords: Stirred tanks; Computational #uid dynamics; Macromixing; Micromixing; Parallel reactions

1. Introduction other hand, all these parameters may show considerable


gradients, which are strongly connected with the geo-
Due to their importance in the "eld of process indus- metry of the mixing tank and, in general, scale-up di!er-
tries, mixing operations have been for decades the subject ently for the various (often interacting) parameters of
of many investigations. In spite of these research e!orts, interest. All these inhomogeneities are wiped o! by the
many aspects of mixing processes are still to be properly lumping process, which is therefore bound to lead to
understood. As a consequence of this lack of knowledge, uncertainties on the "nal result.
considerable uncertainties, and associated losses of In order to avoid the limitations associated with the
money (Smith, 1990), are still involved in the set up of lumping process, distributed parameter models, based on
full-scale mixing processes. This lack of design reliability the actual hydrodynamics of mixing tanks, should be
is partly due to the fact that so far most investigations preferred. These require an accurate knowledge of local
have attempted to set up correlations among spatially properties of the #ow "eld in the mixer, i.e. a huge
averaged (`lumpeda) parameters, such as average tem- amount of information that in practice cannot be ob-
perature and reagents concentration (perfect mixing as- tained from experimentation. The only viable way for
sumption), average power dissipation per unit volume, gathering the required information is through numerical
average mass transfer coe$cients, and so on. On the simulation of the relevant #ow "elds.
The simulation of three-dimensional #ow "elds is
a complex, computation-intensive task. However, the
* Corresponding author. Tel.: #0039-091-6567216; fax: #0039-
091-6567280. never-ending decrease of computing costs and the conti-
E-mail address: abrucato@dicpm.unipa.it (A. Brucato) nuing development of commercial codes for computa-
 Present address: SULZER Italia S.p.A., Milano, Italy. tional #uid dynamics (CFD) have prompted several

0009-2509/00/$ - see front matter  1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 0 9 - 2 5 0 9 ( 9 9 ) 0 0 3 2 4 - 3
292 A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302

Nomenclature S mass source term, kg m\ s\


t time, s
Abs absorbance of solutions ¹ vessel diameter, m
A cell face surfaces, m u velocity components, m s\
G G
C concentration, mol m\ ; reference velocity (pDN), m s\
 
c impeller clearance, m < cell volume, m
CH reference initial ECA concentration, mol m\ X selectivity of the reactive process towards reaction
C initial CuSO concentration, mol m\ (4) de"ned
 
C initial ethyl-chloroacetate concentration, mol m\ > mass fraction
!
D impeller diameter, m
Greek symbols
D turbulent di!usivity, m s\
2
H liquid height into the vessel, m *x distance between centroids of adjacent cells, m
k kinetic constant of reaction (4), m mol\ s\ C molecular viscosity, kg m\ s\

mr element of the `relation matrixa, s\ e turbulent energy dissipation rate, W kg\
G
N agitation speed, s\ k turbulent viscosity, kg m\ s\
2
Q #ow rate, m s\ o liquid density, kg m\
r reaction rate, mol m\ s\ p turbulent Schmidt number
2

studies aimed at its application to the case of mixing only the `macro-mixinga phenomena in the tank, though
tanks (see for example Brucato, Ciofalo, Grisa" the reaction pair employed practically coincided with
& Micale, 1998 and references therein). It can be stated in that used in a number of works aimed at characterising
general that the predicted #ow "elds show a reasonable `micro-mixinga phenomena (e.g. Rice & Baud, 1990;
agreement with the experimental data, but quantitative Thorma, Ranade & Bourne, 1991; Baldyga & Bourne,
discrepancies can usually be observed (Daskopoulos 1992; Baldyga, Bourne & Hearn, 1997).
& Harris, 1996). These depend on the inadequacy of In the present work, CFD-based simulations of a dif-
the spatial resolution as well as of turbulence models, e.g. ferent mixing sensitive reaction scheme are carried out
on the hypothesis of isotropic turbulence implicitly as- and compared with original experimental data. Both
sumed by the commonly adopted `k}ea model (Ciofalo, simulations and experiment regarded the concentration
Brucato, Grisa" & Torraca, 1996). dynamics in a batch stirred reactor in which a pair of
Improvements in the accuracy of results are expected parallel-competitive reactions takes place. This is a mix-
from the "ner and "ner griddings of the computational ing-sensitive reaction scheme which can be represented in
domain allowed by nowadays computers (Lee, Ng the following way:
& Yianneskis, 1996) and especially from the application
and tuning of more advanced turbulence models such as A#BPProd. 1, (1)
RNG k}e, ASM, RSM and large eddy simulation (LES). A#CPProd. 2. (2)
In the meantime, though the currently available #ow "eld
predictions are not entirely accurate yet, it is worth If equimolar quantities of A, B and C are introduced in
starting to explore the potential of CFD-based simula- a batch system, the "nal products distribution depends
tions of mixing sensitive processes. Among these, the on the kinetics of the two reactions and, if these are not
simulation of multiphase systems is still an area that too slow with respect to mixing times, also on the mixing
requires further developments of fundamental models history of the components.
(Kuipers & Swaaij, 1997), while the simulation of homo- Details on the experiments carried out are given "rst;
geneous complex reaction schemes is probably the most the #ow "eld prediction is then described, and the way in
viable one at present. which the information from such simulations was utilised
In this last "eld relatively few works have appeared so for carrying out process simulations is subsequently de-
far. In the pioneering work by Middleton, Pierce and tailed. Simulations results are "nally compared with the
Lynch (1986) not only CFD was used for the "rst time to experimental data obtained.
simulate the three-dimensional #ow "eld in a stirred
vessel, but the results were also used to simulate the
selectivity of a well-known parallel-consecutive reaction 2. Experimental work
scheme. The comparison of simulation results with ex-
perimental data was highly encouraging. Interestingly, Experimental data were obtained in the 5.5 l vessel
these results were obtained by taking into consideration (¹"0.19 m) illustrated in Fig. 1. It is a fully ba%ed tank,
A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302 293

Fig. 1. Stirred vessel employed for the experimentation.

stirred by a Rushton turbine impeller (D/¹"0.5) driven


by a 500 W, speed controlled, DC motor.
This vessel was utilised in batch mode to run a mix-
ing-sensitive pair of reactions. The reaction pair chosen is Fig. 2. Validation of the kinetic model for reaction (4).
a modi"cation of that proposed by Baldyga and Bourne
(1990), and consists of the following two parallel-com-
may be stoichiometrically complex due to the stability of
petitive reactions:
over-hydroxylated ionic species (Cotton & Wilkinson,
NaOH#CuSO PCu(OH) #Na SO , (3) 1988). The reaction with Cu>> is not only stoichiometri-
      
cally simpler, but also the strong coloration of cupric ions
NaOH#ClCH COOCH CH PClCH COONa
    in water makes it possible to measure their concentration
#CH CH OH, (4) by simple colorimetry. The wavelength utilised to this
 
purpose was 800 nm; at this wavelength the absorbance
i.e. the precipitation of cupric hydroxide and the alkaline
of cupric ions shows a maximum, while the absorbance of
hydrolysis of ethyl-chloroacetate (ECA), both competing
all other species is zero, so that interference with the
for the NaOH added to the system.
cupric ions detection is avoided.
Both reactions are second order but the "rst one is
Preliminary calibration tests showed that a linear de-
much faster, having a kinetic constant which is about
pendence of absorbance on cupric ions concentration
nine orders of magnitude larger than the second one; this
exists in the concentration range of interest. Other pre-
implies that, to all practical purposes, the "rst reaction
liminary tests were performed in order to check the
can be considered as instantaneous with respect to the
negligibility of cupric over-hydroxylated species in the
second one.
solutions of interest, as well as the reliability of the whole
The value of the kinetic constant of reaction (4), at the
measurement technique. To this end the absorbance of
temperature of 203C, is 0.023 m mol\ s\ (Kirby,
the solutions obtained by subsequent additions of so-
1972). A preliminary run was conducted by carrying out
dium hydroxide to a cupric sulphate solution was mea-
only reaction (4). In this case the pH dynamics of an ECA
sured, after separation of the precipitated cupric
solution to which an equimolar amount of NaOH had
hydroxide. As a di!erence with similar attempts made
been suddenly added was recorded. The run was conduc-
with Fe> ions, these measurements resulted in a linear
ted in a small vigorously agitated beaker so that a perfect
plot of absorbance vs added sodium hydroxide, which
mixing assumption could be made with negligible errors.
agreed very well with that predicted on the base of the
In Fig. 2, the pH dynamics predicted by a second-order
stoichiometry of reaction (4) assuming a zero re#ectance
kinetics with the above value of the kinetic constant is
for all other species. The symmetric test consisting in
compared with that experimentally observed, and a good
subsequent additions of cupric sulphate to a solution of
agreement can be observed. The slight discrepancies ob-
sodium hydroxide gave rise to similar results.
servable at long times are in fact meaningless due to the
A typical experimental run consisted of the following
very small residual reagent concentrations. These results
steps:
validate both the second-order kinetics and the above
stated value for the kinetic constant, which were then (a) the reactor was half-"lled with distilled water;
retained in the following simulations. weighted amounts of copper sulphate and ethyl-chloro-
The reaction scheme adopted here di!ers slightly from acetate were then added and the vessel content was
that proposed by Baldyga and Bourne (1990), who used mixed until both reagents were completely dissolved. The
Fe> ions instead of Cu>>. The same authors recog- amounts of reagents were such that, when the system was
nised, however, that the reaction between Fe> and OH\ brought to the "nal volume, concentrations C "10.5

294 A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302

and C "21 mol m\ were obtained for CuSO and speeds were not experimentally investigated because at
! 
ECA, respectively. such speeds the surface aeration phenomenon becomes
(b) the vessel content was then brought to the desired important, and the consequent changes of the system
"nal volume by adding cold and/or warm distilled water, hydrodynamics cannot presently be properly accounted
in order to obtain the desired temperature of 203C. No for in CFD simulations. The initial concentrations of
thermostatting was needed in practice, as each experi- NaOH and CuSO were always kept at their reference

mental run was completed in a few tenths of seconds and values (21 and 10.5 mol/m\, respectively), while the in-
the "nal temperature was always found to be within itial ECA concentration was either the reference one
0.53C of the starting value; (21 mol m\) or halved or doubled with respect to it. The
(c) 20 ml of NaOH solution were suddenly introduced experimental data obtained in terms of X are reported in
by gently, yet rapidly, pouring the content of a small Table 1 along with the relevant experimental conditions.
beaker directly on the surface of the liquid, close to the Observation of Table 1 shows that the selectivity of the
agitator shaft, in an angular position midway between process towards reaction (4) increases while increasing
ba%es. The whole operation was completed in less than the initial ECA concentration, as it could be expected by
a second; the amount of soda in the solution was to give an considering that in this case reaction (b) is increasingly
average concentration of 21 mol m\ in the whole vessel; favoured. In all cases selectivity is of the order of 10\,
(d) reactions were allowed to complete and a sample of in agreement with literature results (Baldyga & Bourne,
the reacted mixture was withdrawn. This was "rst centri- 1990; Baldyga, Bourne & Yang Yang, 1993; Bourne,
fuged at 4000 rpm for 20 min in order to remove the cupric Gholap & Rewatkar, 1995; Baldyga et al., 1997). These
hydroxide precipitate; the absorbance of the solution was results can be compared with the values in the range of
then measured in order to assess the amount of unreacted 10\ that would be predicted if the system was modelled
cuprous ions. This last quantity actually coincides, apart as being perfectly mixed. In other words, in the present
from a stoichiometric coe$cient of 2, with the amount of case the assumption of perfect mixing inside the tank
NaOH reacted with the ethyl-chloroacetate; leads to results which are in error by several orders of
(e) process selectivity towards reaction (4) was then magnitude with respect to reality.
computed as the amount of NaOH reacted via reaction In order to understand this quite surprising result, one
(4), divided by the total amount of NaOH reacted. This should consider that when the concentrated A reagent
last quantity could easily be estimated on the basis of the (NaOH) reaches any volume portion of the agitated
amount of unreacted Cu>>, as detailed in the following. vessel, it reacts almost instantly with the relatively small
amount of reactant B (Cu>>) present there. As a conse-
It is worth noting that contingent small overfeedings of quence reactant B suddenly disappears from that volume,
NaOH (reagent A) might give rise to signi"cant overes- leaving a high concentration of A in contact with reac-
timations of the selectivity X, especially in the range of tant C (ethyl-chloroacetate) and therefore reaction (4) is
small X values. In order to reduce the sensitivity of the allowed to occur. The evolution of the process concerns
technique to such errors, a slight excess (2%) of CuSO
 the involvement, by turbulent dispersion, of further por-
with respect to NaOH was actually used, so that even if tions of the unreacted mixture of B and C formerly
the selectivity towards reaction (4) had been zero, a cer- introduced in the vessel, while the reacting plume is
tain amount of Cu>> would have been left anyway in the carried across the vessel by convective #ows. During all
solution. This amount was checked by means of `blanka this time reagent B engulfed in the plume suddenly disap-
reaction runs, which were performed exactly as the usual pears by reaction with A, but always leaving signi"cant
runs, apart from the fact that no ethyl-chloroacetate was concentrations of A which is therefore allowed to further
added to the system. Indicating with Abs , Abs and
M @ react with C.
Abs , respectively, the absorbances of the solution before
D It is clear that the "nal result of the process, namely its
the introduction of NaOH, at the end of the blank "nal selectivity, will depend on the whole history of the
experiment and at the end of a normal experiment, the mixing process and in turn on the characteristics of the
selectivity X of the process towards reaction (4) was turbulent #ow "eld existing in the vessel. Observation of
simply computed as
Prod. 2 Abs !Abs
X" " D @. (5) Table 1
Prod. 1#Prod. 2 Abs !Abs
M @ Experimental values of selectivity (X)

Agitation speed (rpm)

3. Experimental results Initial ECA conc. (mol m\) 60 90 180 240


10.5 0.130
21 0.243 0.202 0.142 0.124
Experimental runs were performed at various agita- 42 0.390 0.336 0.226 0.194
tion speeds ranging from 60 to 240 rpm; higher agitation
A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302 295

Table 1 shows that the process is indeed mixing sensitive, model (Launder & Spalding, 1974) with standard values
as the selectivity decreases with the increase of agitation of model coe$cients.
speed. This is the expected trend if one considers that Due to periodicity considerations only one quarter of
increasing agitation speed means increasing mixing the tank was simulated. The grid employed is illustrated
rates and therefore getting closer to the perfect mixing in Fig. 3 and includes 49*30*30 (axial*radial*circum-
hypothesis. ferential) cells, a discretisation su$cient to obtain prac-
The amazingly high values of selectivity towards reac- tically grid-independent results (Brucato et al., 1998). As
tion (4) are therefore to be explained on the basis of the
concentration inhomogeneities existing at any instant
during the evolution of the process. Such concentration
inhomogeneities may exist over several length scales,
from vessel size down to molecular size. The related
unmixedness phenomena take di!erent names: `macro-
mixinga is the term employed when concentration di!er-
ences at the vessel scale are considered (i.e. from
centimetres to tens of centimetres for the vessel here
employed) while the term `micromixinga is usually em-
ployed to characterise concentration variations at the
scale of Kolmogoro! eddies (usually in the range of tens
of microns) down to molecular scale (Baldyga et al.,
1997). The term `mesomixinga has also been introduced
(Baldyga & Bourne, 1992) to characterise concentration
variations at intermediate scales (here millimetres to cen-
timetres). In the present case concentration gradients at
all of these scales might in principle be signi"cant and
therefore a!ect the "nal results. Interestingly, the same
reaction pair was used in the past to validate micromix-
ing models as well as mesomixing model (Baldyga &
Bourne, 1990; Baldyga et al., 1993; Bourne et al., 1995;
Baldyga et al., 1997). Macromixing was not taken into
consideration in these works, probably due to the di$-
culties involved in properly accounting for such e!ects.
However, concentration gradients certainly exist at the
vessel scale during the whole process evolution. For
instance, when X"0.1, 90% of the initially uniformly
distributed B reactant has to come into physical contact
with A, implying that at least 90% of the vessel content
has to mix with the concentrated A-solution before the
reactive process ends. This also clearly involves concen-
tration histories on the whole vessel scale, and not only in
the proximity of the injection point.
In the present work an attempt was made to explicitly
take into account the complete mixing history in the
vessel in order to verify how much of the "nal result obs-
erved is to be related to the macroscale inhomogeneities
and how much is to be attributed to unmixedness at
smaller scales.

4. Flow 5eld simulation

The CFDS-FLOW3D code (release 2.3) was used to


simulate the #ow "eld in the tank. This computer code
makes use of a "nite volume method to solve the time-
averaged Navier}Stokes equations. The turbulent stres-
ses were modelled using the well-known k}e turbulence Fig. 3. Computational grid employed for the simulations.
296 A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302

regards the simulation of the impeller, it was obtained by rates on each of the six faces of a cell, which are inter-
imposing on the cylindrical surface swept by impeller tips polated from the velocities at cell centroids, and the
experimental values of velocity, turbulent energy and turbulent, or eddy, viscosity (k "C ok/e), from which
R I
dissipation rate. The values utilised to this purpose were turbulent di!usivities can be deduced by means of turbu-
extrapolated from published experimental LDA data lent Schmidt numbers. In fact, convective and di!usive
(Wu & Patterson, 1989). #uxes for a generic passive scalar (e.g. chemical species)
The `hybrid di!erencing schemea (HDS) was em- can be computed from the above quantities.
ployed for the convective terms and the #ow equations Typical results of these CFD simulations are shown in
were coupled using the SIMPLEC algorithm (Van Door- Fig. 4, where velocity vector plots on signi"cant planes
mal & Raithby, 1984), with under-relaxation factors set have been reported (the dots indicate the centroids of
at 0.3 and 0.4 for velocities and turbulence quantities, cells and are the starting points of each vector). It can be
respectively. 2000 SIMPLEC iterations were performed observed that the general #ow patterns agree quite well
in order to ensure a high degree of convergence. This with published experimental information and the posi-
required a CPU time of about 6 h on the HP712 work- tion of the circulation loci on the plane located midway
station employed. between consecutive ba%es (Fig. 4a) is realistic. More-
For each cell, the output of the CFD simulation in- over, the simulation predicts a displacement of circula-
cludes the values of mean velocity (;), pressure (p), turbu- tion loci towards the impeller (Fig. 4b) as one moves to
lent kinetic energy (k) and dissipation rate (e) at cell vertical planes closer to the ba%e planes; a feature that
centroids. For the subsequent process simulations, how- has been experimentally observed (Costes & Couderc,
ever, the only relevant quantities are the convective #ow 1989). On the other hand, it is worth mentioning that in

Fig. 4. Predicted vector plots (N"90 rpm): (a) plane midway between ba%es; (b) plane immediately in front of ba%es.
A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302 297

general the agreement of simulations with experimental ematic viscosity k /o and the turbulent Schmidt number
2
#ow "elds is not completely quantitative yet, probably p , for which a value of 0.8 was adopted for all species
2
due to the simpli"cations introduced by the turbulence (Yakhot, Orszag & Yakhot, 1987). The molecular di!us-
model (Ciofalo et al., 1996). ivity can be neglected in comparison with the eddy di!us-
The physical properties of the #uid were always as- ivity for highly turbulent #ows, as usually are those
sumed to coincide with those of pure water. It was involving stirred low-viscosity liquids.
considered, in fact, that the minor e!ects on the #ow "eld Facilities to solve a number of such equations are
due to the small density and viscosity changes caused by included in most commercial CFD codes, including
the presence of reactants and products (including the CFDS-FLOW3D. In the present work, however, the
solid precipitate) were negligible in comparison with the choice was made of solving the set of Eq. (6) by means of
other approximations. a purposely written post processing code. In this way the
#uid dynamic simulation could be limited to one quarter
of the tank volume, as previously detailed, while, due to
5. Process simulation the point injection method used in the experiments and the
consequent loss of periodicity of concentration "elds, Eq.
In principle, the #uid dynamic information obtained (6) had to be solved in the whole vessel volume. More-
from CFD computations might allow the modelling of over, by using a post processing program it was possible
both macro and micromixing phenomena. In practice, to preserve full control of the computation, an especially
macromixing phenomena are dealt with more easily as worthy feature for model development purposes.
the information made available by such computations In order to discretise Eq. (6), one can consider the
(#uid velocities and turbulence intensity) is su$cient for vessel volume as split into many lumped cells, which are
the solution of local mass balances and almost no addi- interconnected by known convective and turbulent-dif-
tional modelling is required. On the contrary, the simula- fusive #uxes. This information allows the solution of
tion of micromixing phenomena requires subgrid models, the time-dependent mass and energy balances in each
which are still a research area (Kuipers & Swaaij, 1997). cell. In this way, it is virtually possible to simulate any
In this work only macromixing phenomena are ac- physicochemical process, provided that all relevant ki-
counted for, while micromixing phenomena have been netics are known and that its evolution does not signi"-
neglected. In other words, the agitated system was con- cantly a!ect the #uid-dynamic behaviour of the #uid
sidered as being, at any instant, well micromixed and phase. The hypothesis implicitly made is, in fact, that the
poorly macromixed. This was considered as a "rst ap- process evolution does not a!ect the #ow "eld in the tank
proach to the problem and was prompted by the pre- (passive scalar assumption).
viously mentioned consideration that, if a reacting The way in which Eq. (6) were discretised is now
system is prone to be a!ected by micromixing phe- detailed. Considering a generic cell (i, j, k) in the vessel
nomena, macroscale inhomogeneities are bound to occur volume, its faces can be numbered according to Fig. 5.
as well, so that macromixing phenomena will always A discretised mass balance equation for each chemical
have to be simultaneously dealt with. A separate treat- species can therefore be written as
ment for `mesomixinga modelling is not required with dC 
the present approach, as the #uid dynamic discretisation < " (Q ) C ) (Q '0)#Q ) C ) (Q (0))
dt G G G G G
can be made "ne enough fully to account for the concen- G
tration gradients at mesomixing scales. In the following  (C !C)
# AD G #r ) <, (7)
the term `macromixinga will be meant as inclusive of G 2 *x
`mesomixinga phenomena. G G
As far as macromixing is concerned, once the informa- where the "rst sum on the RHS represents the convective
tion on local values of mean velocity components and exchange, the second sum is the turbulent-di!usion term
turbulence intensity has been generated, the following and the last term is the generation rate due to chemical
advection-dispersion equation has to be solved for each reaction; < is the volume of the cell on which the mass
of the species participating in the reaction processes: balance is being carried out; C and r are, respectively, the
concentration and the reaction rate of the species under
*o> consideration in the same cell; C is the concentration of
G
G#
) +oU> !(C #D )
> ,"S , (6)
*t G G 2 G G the same species in the adjacent cell identi"ed by the
subscript i according to Fig. 5; Q is the time-averaged
G
where U is the velocity vector, while indexed quantities convective #ow through the ith cell face, with a positive
refer to the generic species `ia: > is the mass fraction, S is sign for convective #ows entering into the cell and
G G
the source term due to the chemical reaction, C and a negative sign in the opposite case; A is the area of face i;
G G
D are, respectively, the molecular and eddy di!usivities. the logical operations which appear in the "rst term
2
D is computed as the ratio between the turbulent kin- result into one (1) if the test clause is satis"ed, while the
2
298 A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302

mr "A /<[u (u '0)#D /*x ],


    R 
mr "A /<[u (u '0)#D /*x ],
    R 


 
mr "1/< ) !u A (u '0)# u A (u (0)
 G G G G G G
G G
!D A /*x !D A /*x
R   R  
!D A /*x !D A /*x !D A /*x
R   R   R  
!D A /*x
R   

"! mr
G
G
The mr 's terms depend only on #uid dynamics in-
G
formation; therefore they do not change in time during
process simulation, provided that also the #ow "eld does
Fig. 5. Connection scheme for the generic cell: each face bears the index not change. Therefore, they can be evaluated once for all
of the adjacent cell.
at the beginning of the process simulation and stored in
a four-dimension `relation's matrixa mr(i, j, k, l) where
the "rst three indexes identify the cell on which the mass
result is zero (0) in the other case. This expresses the fact balance is being computed while the fourth, ranging from
that, if the convective #ow rate on the i-face is entering 0 to 6, identi"es the six neighbouring cell, with the zero
into the cell (Q '0), then the contribution of this term to index indicating the considered cell itself. Once the rela-
G
the accumulation rate of the species under consideration tion's matrix has been computed and stored, the time
is given by C *Q , where C is the concentration in the cell derivatives of all concentrations can be computed
G G G
adjacent to the i-face. On the other hand, if the convective straightforwardly by means of Eq. (8).
#ow rate is leaving the cell (Q (0) then the above A simple explicit technique, namely the Heun method
G
contribution is negative and given by C*Q ; *x is the (Chapra & Canale, 1989), was used for the numerical
G G
distance between the centroids of the considered cell and solution of the mass balance di!erential (8). This method
of the ith adjacent one; D is the eddy di!usivity at each allows an (over)estimation of the error made at each in-
2
face; in the present case, for all species, a value of 0.8 was tegration step, so that the adoption of an adaptive time
assumed for the turbulent Schmidt number. As the CFD step is possible, a feature that was utilised here.
code employed computes turbulent viscosity at cell cen- In the present case the "rst reaction is extremely fast
troids rather than at cell faces, for each face of the control with respect to the second, so that resolving the actual
volume the relevant value was obtained by linear inter- time dynamics for both reactions would result in a sti!
polation from values at cell centroids; r is a generation set of equations. This problem was overcome simply by
term due to chemical reactions and it is a function of assuming the "rst reaction to be instantaneous with
species concentration in the cell under consideration. respect to the second one, a reasonable assumption in
Such an equation can be written for each chemical view of the large di!erence between the two kinetic con-
component and for each cell of the agitated volume. To stants. In practice, this means that after the convective
compute the concentration dynamics of all components and turbulent contributions to the concentration levels in
in each cell, it is then necessary to solve this set of linear each cell had been computed, concentrations in each cell
ordinary di!erential equations (ODE). were updated by letting the "rst reaction occur up to
In order to speed up the numerical solution of the complete consumption of the limiting reactant before
ODEs set, Eq. (1) may be rewritten as subtracting the sink terms due to the second reaction.
This `instantaneous treatmenta was obviously performed
dC 
" mr ) C #mr ) C#r (8) at each of the two steps involved in the Heun method for
dt G G 
G every time increment.
where: As regards the injection modelling, it was assumed that
the exact amount of NaOH suddenly appeared in the cell
mr "A /<[!u (u (0)#D /*x ],
    R  corresponding to the actual injection point. Though the
mr "A /<[!u (u (0)#D /*x ], real injection was not instantaneous, it took less than
    R 
a second and it was checked that simulating a continuous
mr "A /<[!u (u (0)#D /*x ],
    R  injection over a few seconds did not signi"cantly a!ect
mr "A /<[u (u '0)#D /*x ], the results.
    R 
A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302 299

As described above, the process simulations made use


of the #ow "eld (velocities and eddy viscosities) computed
by CFDS-FLOW3D. Each process simulation was stop-
ped when the entire amount of reactant A (NaOH) had
been consumed by the reaction. The reaction time varied
from 2.5 to 12 s, depending on agitation speed and initial
ECA concentration. In all runs the overall mole balance
was successfully checked. The CPU time required for
each run was of the order of 10 h on the HP-712 worksta-
tion utilised.

6. Simulation results and discussion

Several simulations of the reactive process under in-


vestigation were carried out, covering the experimental
range of values of reactant concentrations and agitation Fig. 6. Selectivity towards reaction (4). Experimental data: (䊏)
speeds. C "21; (䉱) C "42 (mol m\) Simulations results: ***
! !
The simulation results obtained are shown as lines in C "21; - - - - - C "42 (mol m).
! !
Fig. 6, where the experimental data have been reported as
symbols for comparison purposes. A satisfactory agree-
ment between simulation results and experimental data
can be observed. Not only the order of magnitude of the
selectivity is well predicted, but even the actual values in
many cases do practically coincide with experimental
data. Also the trends with both agitation speed and ECA
concentration appear to be well predicted.
The agreement tends to worsen at the highest agitation
speeds and ECA concentrations, where selectivities are
slightly overpredicted, which implies some underestima-
tion of mixing intensity. This behaviour cannot be ex-
plained by the fact that micromixing phenomena were
neglected, as their inclusion in any form would have
resulted in a poorer mixing and therefore in a more or
less pronounced increase of predicted selectivities. Nor
can it be explained by the fact that an instantaneous in-
jection was simulated though the real injection extended
over a "nite length of time, since simulations performed Fig. 7. Selectivity towards reaction (4) vs initial ECA concentration.
by spreading the injection over a few seconds (a feature
allowed by the process simulation procedure) did not
result in signi"cantly di!erent "nal selectivity predic-
tions. Probably the observed discrepancies, which are not data were used, so that the simulations performed can be
particularly important in view of the several orders of regarded as being fully predictive.
magnitude already recovered over the perfect mixing It is interesting to compare the above results with
assumption, are somehow due to shortcomings in the previous work (Baldyga & Bourne, 1990) in which prac-
simulated #ow "elds. tically the same reaction scheme and tank geometry was
In Fig. 7 the predicted trend of selectivity vs initial adopted in order to validate micromixing models, on the
ECA concentration for runs conducted at an agitation basis of the assumption that only microscopic concentra-
speed of 90 rpm is compared with the experimentally tion gradients a!ected the observed results. The agree-
observed trend. As can be seen, the predictions obtained ment of model predictions with experimental data was
by computer simulations are again in very good agree- optimised by tuning model parameters, namely the speci-
ment with experimental data. "c energy dissipation in the reaction volume. In sub-
The noticeable agreement between predicted and ex- sequent papers (Baldyga et al., 1993; Bourne et al., 1995;
perimental data is an especially interesting result no Baldyga et al., 1997) it was recognised that in the case of
adjustable parameters were employed for the simula- instantaneous common reagent injections, mesomixing
tions: only published literature kinetic and anemometric may play a signi"cant, though not exclusive, role.
300 A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302

It is clear that in the simulations performed here the Fig. 8 the time evolution of an `iso-surfacea at constant
mesomixing phenomena are probably resolved together Cu>> concentration of 0.021 mol m\ has been re-
with macromixing phenomena, but no account whatso- ported. The concentration level chosen corresponds to
ever is given of micromixing phenomena. The good  of the initial, uniformly spread, Cu>> concentration

agreement obtained between predicted and experimental and therefore the surfaces reported in Fig. 8 can be
results can therefore be regarded as a proof that only regarded as subdividing the vessel volume in a NaOH
meso- and macro-scale phenomena are actually impor- rich `reaction zonea (inside the iso-surfaces) where Cu>>
tant in the present case and that CFD techniques are an has already been completely depleted and only reaction
e!ective tool for properly accounting for such phe- 2 is taking place, and a Cu>> rich region where no
nomena. Interestingly, practically the same result was NaOH can be present and no reaction is taking place. At
obtained by Middleton et al. (1986) with a well-known any instant the fast reaction 1 is taking place in a narrow
parallel-consecutive reaction scheme, which has been layer just about the iso-surfaces shown.
widely adopted to validate micromixing models (Rice It can be observed that after 4 s from the start a fairly
& Baud, 1990; Thorma et al., 1991; Baldyga & Bourne, large portion of the vessel region above the impeller
1992; Baldyga et al., 1997). plane is involved in the reaction, while the vessel re-
In order to illustrate the spatial features of concentra- gion below the impeller is practically never interested.
tion dynamics for the system under investigation, in Clearly the process involves macro-scale inhomogeneities

Fig. 8. Concentration dynamics of Cu>> for the case N"90 rpm, C "21 mol m\. The iso-surfaces shown embed the reactor volume where the

Cu>> has been completely depleted and only the slower reaction (4) is taking place.
A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302 301

connected with the whole mixing history of the Finally, the fact that some mixing-sensitive reaction
reactants. processes appear to be properly dealt with by CFD-
By looking at Fig. 8, one may wonder whether a simple based simulations despite the still enduring inaccuracies
model based on the subdivision of the vessel volume in in the simulated #ow "elds, may be explained by the
just two well-mixed zones, one above and the other consideration that such inaccuracies might have opposite
below the impeller plane, may be su$cient for the simula- e!ects in the various reactor regions involved in the
tion of the present system. With such an approach the full reaction process. In such a case, the predicted "nal selec-
predictivity of the CFD-based approach would be lost, tivity would not be strongly a!ected by local #ow "eld
due to the need to estimate a proper inter-exchange rate inaccuracies. It is certainly worth, however, pursuing
parameter. It should be noted, however that, though after further improvements in the accuracy of the #ow "eld
4 s the reaction zone is fairly wide, by no means it in- simulations, in order to prevent the applicability range
cludes the whole upper region of the vessel. Also, at this of CFD-based simulations from being a!ected by such
point almost 95% of the NaOH initially added has uncertainties.
already reacted (it was 55% after 2 s) and from now on
the reaction zone tends to shrink until it completely
disappears shortly after 7 s. If one also considers that the 7. Conclusions
instantaneous selectivity X tends to decrease as time
elapses (in the present case it is about 0.24 after 0.5 s and Fully predictive CFD-based simulations of a parallel-
slowly decays towards its "nal value of 0.20) it is clear competitive, mixing-sensitive reaction scheme were car-
that the simpler model based on the segregation between ried out. Simulation results were compared with original
the upper and lower vessel regions would miss most of experimental data obtained in a standard stirred reactor,
the phenomena actually occurring in the tank. and a good agreement between the two was found, lead-
The present results indicate that the signi"cance of ing to the conclusion that CFD is a powerful tool for
micromixing phenomena is probably less widespread than modelling mixing sensitive processes.
is often assumed, but by no means they rule away the These encouraging results were obtained without giv-
possibility that micromixing phenomena may be impor- ing any consideration to micromixing phenomena,
tant in many practical situations. They only prove that for though a very similar reaction scheme has been used in
the reaction process investigated here, in the case of the past to validate micromixing models. This observa-
instantaneous injection of the common reactant, micro- tion implies that micromixing phenomena were actually
mixing does not seem to play any role. Had reaction 2 negligible in the investigated system. This is likely to be
been signi"cantly faster, the conclusions might well have the case also for another reaction scheme commonly
been di!erent. Even with the same reaction process, in employed for micromixing models development and vali-
the case of su$ciently slow common reactant injections, dation, when instantaneous injections of the concen-
micromixing may well start playing a signi"cant role. trated reactant are involved.
In any case, since it is unlikely to deal with systems that As macroscale inhomogeneities can now be e!ectively
are well macromixed and poorly micromixed, macromix- dealt with by means of CFD tools, they should never be
ing phenomena should always be properly accounted neglected. For cases where results cannot be explained in
for in simulations, and CFD tools appear to be well terms of macroscopic inhomogeneities, micromixing
suited for such a purpose. Micromixing phenomena models should be coupled, as sub-grid models, to the
should be dealt with by means of sub-grid models CFD-derived macromixing frame. In this case the para-
superimposed on macromixing CFD-based simulations, meters involved in micromixing modelling could prob-
as done for instance by Pipino and Fox (1994) for a ably be related to the local values of turbulence
tubular jet reactor. This would be needed only for those parameters, such as the speci"c energy dissipation rate,
cases in which the simple macromixing simulation which are provided by the same CFD simulation.
failed to yield a satisfactory agreement with experimental
observations.
Notably, the e!ect of superimposing subgrid models Acknowledgements
on a CFD-based macromixing simulation was also inves-
tigated by Harris, Roekaerts and Rosendal (1996), for The authors wish to thank AEA Technology for pro-
a more complex reaction scheme and a cylindrical jet- viding CFDS-Flow3D (now CFX4) under a special li-
stirred reactor. Both a probability density function cence arrangement.
(PDF) model and an extended eddy break-up (EBU#)
model were tested. The authors found, however, that References
results provided by simple macromixing modelling were
satisfactory on their own, and practically una!ected by Baldyga, J., & Bourne, J. R. (1990). The e!ect of micromixing on parallel
the inclusion of subgrid modelling. reactions. Chemical Engineering Science, 45, 907}916.
302 A. Brucato et al. / Chemical Engineering Science 55 (2000) 291}302

Baldyga, J., & Bourne, J. R. (1992). Interaction between mixing on Kuipers, J. A. M., & Swaaij, W. P. M. (1997). Application of computa-
various scales in stirred tank reactors. Chemical Engineering Science, tional #uid dynamics to chemical reaction engineering. Reviews in
47, 1839}1848. Chemical Engineering, 13, 1}118.
Baldyga, J., Bourne, J. R., & Yang Yang (1993). In#uence of feed pipe Launder, B. E., & Spalding, D. B. (1974). The numerical computation of
diameter on mesomixing in stirred tank reactors. Chemical Engin- turbulent #ows. Computer Methods in Applied Mechanics and Engin-
eering Science, 48, 3383}3390. eering, 3, 269}289.
Baldyga, J., Bourne, J. R., & Hearn, S. J. (1997). Interaction between Lee, K. C., Ng, K., & Yianneskis, M. (1996). Sliding mesh predic-
chemical reactions and mixing on various scales. Chemical Engin- tions of the #ows around Rushton impellers. Flud Mixing xfth
eering Science, 52, 457}466. International Chemical Engineering Symposium Series, 140,
Bourne, J. R., Gholap, R. V., & Rewatkar, V. B. (1995). The in#uence of 47}58.
viscosity on the product distribution of parallel reactions. Chemical Middleton, J. C., Pierce, F., & Lynch, P. M. (1986). Computation of
Engineering Journal, 58, 15}20. #ow "elds and complex reaction yield in turbulent stirred reactors
Brucato, A., Ciofalo, M., Grisa", F., & Micale, G. (1998). Numerical and comparison with experimental data. Chemical Engineering Re-
prediction of #ow "elds in ba%ed stirred vessels: a comparison of search Design, 64, 18}22.
alternative modelling approaches. Chemical Engineering Science, 53, Pipino, M., & Fox, R. O. (1994). Reactive mixing in a tubular jet
3653}3684. reactor: a comparison of PDF simulation with experimental data.
Chapra, S. C., & Canale, R. P. (1989). Numerical methods for engineers. Chemical Engineering Science, 49, 5229}5241.
New York: McGraw Hill. Rice, R. W., & Baud, R. E. (1990). The role of micromixing in the
Ciofalo, M., Brucato, A., Grisa", F., & Torraca, N. (1996). Turbulent scale-up of geometrically similar batch reactors. A.I.Ch.E. Journal,
#ow in closed and free-surface unba%ed tanks stirred by radial 36(2), 293}298.
impeller. Chemical Engineering Science, 51, 3557}3573. Smith, J. M. (1990). Industrial needs for mixing research. Transactions of
Costes, I., & Couderc, J. P. (1989). Study by laser doppler anemometry the Institution of Chemical Engineers, 68, 3}6.
of the turbulent #ow induced by a rushton turbine. Chemical Engin- Thorma, S., Ranade, V. V., & Bourne, J. R. (1991). Interaction between
eering Science, 43, 2751}2764. micro- and macro-mixing during reactions in agitated tanks. Cana-
Cotton, F. A., & Wilkinson, G. (1988). Advanced inorganic chemistry (53 dian Journal of Chemical Engineering, 69, 1135}1141.
Ed.). New York: Wiley. Van Doormal, J. P., & Raithby, G. D. (1984). Enhancements of the
Daskopoulos, Ph., & Harris, C. K. (1996). Three-dimensional CFD simple method for predicting incompressible #uid #ows. Numerical
simulations of turbulent #ow in ba%ed stirred tanks: An assessment Heat Transfer, 7, 147}163.
of the current position. I. Chem. E. Symposium Series, 140, 1}13. Wu, H., & Patterson, G. K. (1989). Laser-Doppler measurements of
Harris, C. K., Roekaerts, D., & Rosendal, F. J. J. (1996). Computational turbulent-#ow parameters in stirred tanks. Chemical Engineering
#uid dynamics for chemical reactor engineering. Chemical Engineer- Science, 44, 2207}2221.
ing Science, 51, 1569}1594. Yakhot, V., Orszag, S. A., & Yakhot, A. (1987). Heat transfer in
Kirby, A. I. (1972). Hydrolysis and formation of esters of organic acids. turbulent #uids * I. pipe #ow. International Journal of Heat and
In C. H. Bamford, & C. F. H. Tipper, Comprehensive chemical Mass Transfer, 30(1), 15}22.
kinetics, Vol. 10. Amsterdam: Elsevier.

You might also like