Вы находитесь на странице: 1из 60

CFD SIMULATION OF STIRRED TANKS: COMPARISON OF

TURBULENCE MODELS. PART I: RADIAL FLOW IMPELLERS


Jyeshtharaj B. Joshi,12 * Nandkishor K. Nere,13 Chinmay V. Rane,1 B. N. Murthy,1
Channamallikarjun S. Mathpati,1 Ashwin W. Patwardhan1 and Vivek V. Ranade4
1. Department of Chemical Engineering, Institute of Chemical Technology, Matunga, Mumbai 400019, Maharashtra, India
2. Homi-Bhabha National Institute, Trombay, Mumbai 400 094, Maharashtra, India
3. Global Pharmaceutical R and D, Abbott Laboratories, 1401 N Sheridan Rd, North Chicago, Illinois 60064
4. Chemical Engineering and Process Development Division, National Chemical Laboratory, Pune 411008, Maharashtra, India

A critical review of the published literature regarding the computational fluid dynamics (CFD) modelling of single-phase turbulent flow
in stirred tank reactors is presented. In this part of review, CFD simulations of radial flow impellers (mainly disc turbine (DT)) in a fully
baffled vessel operating in a turbulent regime have been presented. Simulated results obtained with different impeller modelling approaches
(impeller boundary condition, multiple reference frame, computational snap shot and the sliding mesh approaches) and different turbulence
models (standard k − ε model, RNG k − ε model, the Reynolds stress model (RSM) and large eddy simulation) have been compared with
the in-house laser Doppler anemometry (LDA) experimental data. In addition, recently proposed modifications to the standard k − ε models
were also evaluated. The model predictions (of all the mean velocities, turbulent kinetic energy and its dissipation rate) have been compared
with the experimental measurements at various locations in the tank. A discussion is presented to highlight strengths and weaknesses of
currently used CFD models. A preliminary analysis of sensitivity of modelling assumptions in the k − ε models and RSM has been carried out
using LES database. The quantitative comparison of exact and modelled turbulence production, transport and dissipation terms has high-
lighted the reasons behind the partial success of various modifications of standard k − ε model as well as RSM. The volume integral of predicted
energy dissipation rate is compared with the energy input rate. Based on these results, suggestions have been made for the future work in this area.
Nous présentons un examen critique de la littérature concernant la modélisation de la dynamique des fluides numérique (DFN) de l’écoulement
turbulent à une phase dans les réacteurs à cuve agitée. Dans cette partie de l’examen, nous présentons les simulations de DFN de turbines à
écoulement radial (principalement des turbines à disque (TD)) dans un réservoir entièrement cloisonné effectuées dans un régime turbulent. Les
résultats des simulations obtenus grâce à différentes approches de modélisation des turbines (couche limite turbulente, méthode des référentiels
multiples, snap-shot de modélisation numérique, maillage glissant) et à différents modèles de turbulence (modèle standard k-e, modèle RNG
k-e, modèle aux tensions de Reynolds et simulation des grandes échelles) ont été comparés aux données expérimentales internes d’allocation de
Dirichlet latente (ADL). De plus, les modifications des modèles standards k-e récemment proposées ont également été évaluées. Les prédictions
du modèle (de toutes les vitesses moyennes, de l’énergie cinétique turbulente et de son taux de dissipation) ont été comparées aux données
expérimentales relevées à différents endroits de la cuve. Une discussion présente les points forts et les points faibles des modèles de DFN actuellement
utilisés. Une analyse préliminaire de la sensibilité des hypothèses de modélisation liées aux modèles k-e et RSM a été menée en utilisant la base de
données LES. La comparaison quantitative des données exactes et modélisées de la production, du transport et de la dissipation de la turbulence
a mis en évidence les raisons qui expliquent la réussite partielle de plusieurs modifications apportées au modèle standard k-e ainsi qu’au modèle
aux tensions de Reynolds. L’intégrale de volume du taux de dissipation énergétique prévue a été comparée au taux d’intrant énergétique. Sur la
base de ces résultats, d’autres études à venir dans ce domaine ont été suggérées.

Keywords: CFD, stirred vessel, radial flow impellers, impeller models, k − ε models, RSM, LES

INTRODUCTION ∗ Author to whom correspondence may be addressed.

S
tirred tank reactors, in which one or more impellers are E-mail addresses: jb.joshi@ictmumbai.edu.in, jbjoshi@gmail.com
used to generate desired flow and mixing, are amongst the Can. J. Chem. Eng. 89:23–82, 2011
most widely used reactors in chemical and allied indus- © 2011 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.20446
tries. Stirred reactors offer unmatched flexibility and control over
Published online in Wiley Online Library
various transport processes occurring within the reactor. The (wileyonlinelibrary.com).
performance of a stirred reactor can be optimised by appro- This paper was meant to be part of the GLS 2010 Special Issue - PART A
priate adjustments of the reactor hardware and the operating (CJCE 20465 (Part B) shall be published soon).

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 23 |


parameters. Parameters like reactor and impeller shapes, aspect Therefore, in this part of review, an attempt has been made to
ratio of the reactor vessel, number, type, location and size of present a status report on the CFD as applied for the modelling of
impellers, degree of baffling, etc., provide effective handles to con- stirred tank reactors with radial flow impellers.
trol the performance of stirred reactors. However, the availability Capability of CFD to predict heat transfer coefficient in simple
of such a large number of parameters also makes the job of select- geometries such as pipeline reactor has been carefully analysed by
ing suitable configuration of the stirred reactor quite difficult. Thakre and Joshi (2002). They have used standard k − ε model,
The most important hurdle has been the lack of understanding of low Reynolds number k − ε model, and the Reynolds stress model
the underlying physics and hence the quantitative relationships (RSM) for the prediction of heat transfer coefficient in a single-
between the reactor parameters (hardware and operating) and the phase fully developed turbulent pipe flow. Even Muzzio and
performance objectives. Therefore, the available relationships are cowokers (Lamberto et al., 1996; Alvarez-Hernández et al., 2002;
empirical and cannot be considered reliable and/or over a wide Szalaia et al., 2004; Arratia et al., 2006) have worked on chaotic
range of applications encountered in practice. The combination mixing. In a similar way, it is wished that the predictive proce-
of such empirical relationships and the accumulated operating dures be developed for the heat transfer coefficient at the tank
experience has been the present state of the art. In this direc- wall, cooling coils, and other modern heat transfer geometries,
tion, Table 1 gives a partial list of previous attempts that have though the real flow is three-dimensional and not at all well devel-
provided empirical/semi-empirical relationships for the design oped because of internals. Further, the multiphase applications
of stirred tanks. For this transformation, vigorous attempts have are more commonly encountered in industry where there are
been made during the past 20 years by focusing on the physics major heat transfer needs. In addition, CFD simulations should
of fluid mechanics by using the tools, namely experimental fluid lead to such a hardware design where an optimum flow pat-
dynamics (EFD) and the computational fluid dynamics (CFD). tern (mean and turbulent) is also possible near the heat transfer
CFD is a body of knowledge and techniques to solve mathematical surface. The room available for the optimisation of heat transfer
models of fluid dynamics on digital computers. CFD was identi- is even bigger because of the distance between the locations of
fied as one of the critical enabling technologies for the chemical the heat transfer surface and the turbulence generating means
industry in Technology Vision 2020, The U.S. Chemical Industry. (impeller).
The application of CFD was expected to lead to shortened product- In the case of gas–liquid and liquid–liquid dispersions, the bub-
process development cycles, optimisation of existing processes to ble and drop size varies within the reactor because the size is
improve energy efficiency, and the efficient design of new products governed by the balance between the retaining interfacial/surface
and processes. Joshi and Ranade (2003) have provided a perspec- tension force and the breaking force due to turbulent (usually
tive of the current status of computational flow modelling with written in terms of turbulent energy dissipation rate, ε) and vis-
respect to above mentioned expectations. This manuscript in par- cous stresses. The coalescing nature of the dispersion is also
ticular makes an attempt to analyse the published literature and important. These flow parameters are the most difficult to predict
provide coherent picture of the current status on CFD as applied with the present status of knowledge. Till date, design engineers
to stirred tanks. In this review, emphasis is on assessing the state rely on empirical correlations and very large safety margins for
of the art and identifying possible ways of improving the state of prediction of dispersed phase hold-up, interfacial area, mass trans-
the art. fer coefficient in multiphase contacting. CFD simulation of these
In baffled stirred vessels, flow generated by the rotating systems available in literature could predict the average dispersed
impeller interacts with stationary baffles and generates complex, phase hold-up profile, but accurate estimation of other parameters
three-dimensional, re-circulating turbulent flow. The mean flow is a challenging area for CFD practitioners.
generated by the impeller causes bulk motion of the constituents, A CFD model is also useful for the prediction of residence
which is responsible for the convective transport of momentum, time distributions as a consequence of turbulence and mean flow
heat and mass. The generated turbulence determines the eddy field including, short-circuiting, by-pass, etc., in single and multi-
diffusion of momentum, heat and mass. Hence the overall perfor- phase systems. These models are also useful to understand issues
mance of the stirred vessel depends strongly on the flow field related to reactor dynamics and start-up/shutdown operations.
and the associated turbulence characteristics produced by the If the performance is found to be sensitive to fluid dynamics
impeller. The flow near the impeller region is unsteady in nature related issues, computational flow models can be used to obtain
due to the periodic passage of the blades, whereas, at a short dis- accurate information about the desired processes. Reaction engi-
tance away from the impeller, the mean flow behaves essentially neering models based on mixing cell framework with CFD models
at steady state (statistically stationary). The impeller details (size, to furnish the required information of convective and turbulent
number of blades, number of impellers and locations), vessel con- transport between different mixing cells are useful to interpret
figuration (number, location and size of baffles, vessel diameter, and extrapolate laboratory and pilot scale experiments. Detailed
liquid level, impeller clearance), and the physicochemical proper- simulations of reaction engineering models at different values of
ties of the process fluid have a profound impact on the flow field transport parameters (heat and mass transfer coefficient, mixing
generated in the stirred vessel. and so on) are carried out to identify operable windows and to
On the basis of flow produced, impellers are classified into two evolve quantitative demands on reactor hardware. Following are
broad categories (i) radial flow impellers (ii) axial flow impellers. some of potential applications of CFD modelling of stirred tank
The energy input is mainly used for producing turbulence in case reactors.
of radial flow impellers whereas for convection in case of axial
flow impellers. Radial flow impellers are more suited for appli- (1) Resolving conflicting process requirements: For most of the
cations such as gas–liquid, liquid–liquid, and other multiphase industrial processes, reactor has to carry out several functions
dispersions. Axial flow impellers are used for blending, heat trans- simultaneously. It is quite common to find that require-
fer, and solid suspension. Due to significant difference in the flow ments of these different expected functions of the reactor
pattern as well as non-homogeneous turbulence, the performance may be quite different or some times, even conflicting with
of turbulence models differs for both the classes of impellers. each other. For example, desired fluid dynamic characteristics

| 24 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


|
Table 1. Empirical/semi-empirical procedures for the design of stirred tanks: previous work

Sr. numberDesign parameters Refs.

1 Mixing time Peters and Smith (1969); Carreau et al. (1976); Brennan and Lehrer (1976); Khang and Levenspiel (1976); Pandit and Joshi (1983); Pandit et al.
(1984); Shiue and Wong (1984); Sano and Usui (1985); Raghava Rao and Joshi (1988); Raghava Rao et al. (1988); Takahashi et al. (1988); Saito
and Kamiwano (1989); Saito et al. (1990); Ranade et al. (1991); Rewatkar and Joshi (1991a); Rewatkar and Joshi (1991b); Carreau et al. (1992);

VOLUME 89, FEBRUARY 2011


Satoh et al. (1995); Nienow (1996); Lamberto et al. (1996); Patwardhan and Joshi (1999); Sahu et al. (1999); Patwardhan et al. (2003); Nere et

|
al. (2003); Patwardhan et al. (2004); Kumaresan and Joshi (2006); Kumaresan et al. (2006)
2 Power consumption Rushton et al. (1950a); Rushton et al. (1950b); Metzner and Otto (1957); Metzner et al. (1961); Godleski and Smith (1962); Bates et al. (1963);
Beckner and Smith (1966); Bourne and Butler (1969a); Novák and Rieger (1969); Nagata et al. (1970); Hall and Godfrey (1970); Fort et al.
(1971); Nienow and Miles (1971); Bertrand et al. (1980); Rewatkar and Joshi (1991a); Sarvanan et al. (1996); Karcz and Major (1998); Dohi et al.
(2004); Deshmukh and Joshi (2006)
3 Heat transfer Mizushina et al. (1966a); Mizushina et al. (1966b); Coyle et al. (1970); Edney and Edwards (1976); De Maerteleire (1978); Shamlou and Edwards
(1986); Wang and Yu (1989); Carreau et al. (1994)
4 Mass transfer Feranandes and Sharma (1967); Miller (1971); Levins and Glastonbury (1972); Kuboi et al. (1974); Boon-Long et al. (1978); Conti and Sicardi
(1982); Lai et al. (1988); Armenante and Kirwan (1989); Jadhav and Pangarkar (1991); Hiraoka et al. (1993); Barigou and Greaves (1996); Gezork
et al. (2001); Fishwick et al. (2003); Martı́n et al. (2010)
5 Scale-up procedures Rieger and Novák (1972); Khang and Levenspiel (1976); Bowen (1985); Costes and Couderc (1988); Obot (1993); Ogawa and Kuroda (1995);
Montante et al. (2003); Gimbun et al. (2009)
6 Multiphase flooding Miller (1974); Yung et al. (1979); Warmoeskerken and Smith (1985); Tanaka and Izumi (1987); Wong et al. (1987); Takahashi and Nienow (1992);
Birch and Ahmed (1997); Paglianti et al. (2000); Bombač and Žun (2006)
7 Gas–liquid dispersions Komasawa et al. (1970); Miller (1974); Joshi and Sharma (1976); Figueiredo and Calderbank (1979); Joshi and Kale (1979); Rewatkar and Joshi
(1991b); Rewatkar and Joshi (1991c); Rewatkar and Joshi (1993); Rewatkar et al. (1993); Mhetras et al. (1994); Tecante et al. (1996); Parsu Veera
et al. (2001); Joshi and Patwardhan (2002); Thatte et al. (2004); Scargiali et al. (2007a); Ford et al. (2008); Martı́n et al. (2010)
8 Gas–liquid–solid dispersion Raghava Rao and Joshi (1989); Rewatkar and Joshi (1992); Ebrahimi-Moshkabad and Winterbottom (1999); Micale et al. (2000); Nienow and
Bujalski (2002); Fishwick et al. (2003); Murthy et al. (2007b); Chen et al. (2009)
9 Liquid–liquid dispersion Sprow (1967); Chen and Middleman (1967); Brown and Pitt (1972); Weinstein and Treybal (1973); Coulaloglou and Tavlarides (1977); Brooks

|
(1979); Rounsley (1983); Davies (1985); Okufi et al. (1990); Kumar et al. (1991); Armenante et al. (1992); Ribeiro et al. (1996); Zhou and Kresta
(1998); Colenbrander (2000); Kamienski and Wójtowicz (2003); Derksen and van Den Akker (2007); Kadam et al. (2009)
10 Solid–liquid dispersion Patil et al. (1984); Raghava Rao et al. (1988); Rewatkar et al. (1989); Rewatkar and Joshi (1991a); Rewatkar and Joshi (1991b); Armenante et al.
(1992); Guiraud et al. (1997); Nocentini et al. (2000); Fishwick et al. (2003); Fajner et al. (2008); Unadkat et al. (2009)
11 Fractional dispersed phase Mlynek and Resnick (1972); McManamey (1979); Sembira et al. (1988); Pacek et al. (1994); Arjunwadkar et al. (1998); Kulkarni et al. (2001);
hold-up Brunazzi et al. (2004); Shewale and Pandit (2006); Sun et al. (2006); Boden et al. (2008); Kadam et al. (2009)
12 Size distribution and Magelli et al. (1990); Magelli et al. (1991); Nishikawa et al. (1994); Pacek et al. (1998); Sessiecq et al. (1999); Angeli and Hewitt (2000); Ochieng
concentration profile of and Lewis (2006); Montante et al. (2008); Fajner et al. (2008)
dispersed phase
13 Breakage and coalescence Madden and Damerell (1962); Komasawa et al. (1970); Mizoguchi et al. (1973); Brown and Pitt (1974); Molag et al. (1980); Narsimhan et al.
(1984); Das et al. (1987); Barigou and Greaves (2009); Shimizu et al. (1999); Zaccone et al. (2007)
14 Chemical reactions Komasawa et al. (1970); Robinson and Wilke (1973); Uchida et al. (1978); Leszek et al. (2004); Patwardhan et al. (2005); Milewska and Molga (2010)
15 Self-inducing impellers Joshi and Sharma (1977); Sawant and Joshi (1979); Joshi (1980); Sawant et al. (1981); Rielly et al. (1992); Forrester and Rielly (1994); Forrester et al.
(1998); Poncin et al. (2002); Murthy et al. (2007a); Scargiali et al. (2007b); Murthy et al. (2008a)
16 Flow visualisation Peters and Smith (1967); Bourne and Butler (1969b); Murakami et al. (1972); van’t Riet and Smith (1975); Günkel and Weber (1975); Hiraoka et al.
(1979); Kuriyama et al. (1982); Mochizuki and Takashima (1984); Kuboi and Nienow (1986); Yianneskis et al. (1987); Ranade and Joshi (1989);
Wu and Patterson (1989); Ranade and Joshi (1990a); Shervin et al. (1991); Mishra and Joshi (1991); Kumar et al. (1993); Mishra and Joshi (1993);
Chapple and Kresta (1994); Bakker and van den Akker (1996); Desai and Joshi (1996); Schäfer et al. (1997); Fentiman et al. (1998); Mishra et al.
(1998); Mavros et al. (2001); Alvarez-Hernández et al. (2002); Szalai et al. (2004); Nikiforaki et al. (2003); Kumaresan et al. (2005); Arratia et al.

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING


(2006); Kumaresan and Joshi (2006); Rielly et al. (2007); Murthy et al. (2007a); Hristov et al. (2008); Murthy et al. (2008b); Roy et al. (2010)

|
25
|
for blending and heat transfer are quite different (which impeller rotation speed. However, understanding of turbulent
require more bulk flow and less shear) from those desired for flow structures close to wall and their dependency on type
gas–liquid dispersion and mass transfer (which require more of impeller can help us to design energy efficient impellers.
shear). In case of aerobic fermentation, the oxygen trans- This understanding can be developed using advanced EFD
fer rate increases with hydrodynamic shear. However, the and CFD. These tools will give transient details of veloc-
validity, growth rate and the selectivity of microorganisms ity field inside the reactor. This database can be used to
(particularly animal and plant cell culture) and the activ- estimate various properties of flow structures such as age
ity of proteins and enzymes decreases with an increase in distribution, size and shape distribution, length scale distri-
shear. Such conflicting requirements make the task of evolv- bution, energy distribution, penetration depth distribution,
ing “wish list” for the desired fluid dynamics difficult. Reactor etc. This information can be used in combination with clas-
engineer has to achieve appropriate compromise between the sical theories of heat and mass transfer such as film theory,
conflicting process requirements to achieve the best results. penetration/surface newel theory, small eddy theory, large
Using a computation model, one can switch on and off vari- eddy theory, surface divergence theory. Most of these theo-
ous processes, which otherwise is not possible while carrying ries use assumed flow field and consider only one of the above
out experiments. Such parametric experimentation can give mentioned features. Accounting only one of these properties,
useful insight into interactions between different processes that too on an averaged scale may not be appropriate (Joshi
and can help to resolve the conflicting requirements. et al., 2009). To obtain accurate predictions, the velocity
(2) Translating batch data for continuous reactors: In most of the field data need to be analysed using eddy isolation algo-
cases, laboratory and bench scale experiments required to val- rithms based on zero crossing in fluctuating velocity signals,
idate the reactor concept are carried out in a batch mode. It wavelet based methods and proper orthogonal decomposition
is then necessary to translate (or to use) the data obtained to obtain structure age, size, energy, and shape distribution
in such experiments for designing continuous reactors. The and then obtain contribution of each and every flow structure
location of feed pipes, outlets and their influence on mixing and average it over space and time. Another way is to carry
and performance needs to be understood. Computational flow out direct numerical simulation (DNS) and/or LES of stirred
models can be of great help in this regard. tank with energy equation and obtain heat transfer coeffi-
(3) Scale-down/scale-up analysis: It is essential to analyse the pos- cient for various impeller designs. The objective is impeller
sible influence of scale of reactor on its fluid dynamics and should give appropriate combination of convection and tur-
eventually the performance. It should be noted that small- bulence close to impeller and tank surface can be designed
scale reactor would invariably have higher shear and more in such a way that the surface renewal frequency can be
rapid circulation than large-scale reactor. The interfacial area improved. Such kind of efforts has already been implemented
per unit volume of reactor normally reduces as the scale of in industry in heat exchangers, where inside surface is cor-
reactor increases. Scale-up/scale-down analysis is important rugated/microfins are provided. Microfins are found to be
to plan useful laboratory and pilot plant tests. It may be almost 50 times energy efficient compared to bare smooth
often necessary to use pilot reactor configuration, which is pipes for heat transfer applications. This is due to localised
not geometrically similar to the large-scale reactor in order to breaking of flow structures close to wall due to microfin struc-
maintain the similarity of the desired process. Conventionally ture.
such analysis is carried out based on certain empirical scal-
ing rules and prior experience. Computational flow modelling This brief review of the steps in engineering of stirred reactors
can make substantial contributions to this step by providing indicate that availability of large degrees of freedom regard-
quantitative information about the fluid dynamics. ing reactor configuration, impellers can be effectively exploited
(4) Testing of new reactor concepts: More often than not, develop- to evolve better reactor technologies. This, however, requires
ment of reactor technologies relies on prior experience. New detailed knowledge and understanding of the fluid dynamics in
reactor concepts are often sidelined due to lack of resources all the regions of the stirred reactors. In line with such a broad
(experimental facilities, time, funding, etc.) to test them. view, it was thought desirable to critically examine the predic-
Experimental studies have obvious limitations regarding the tive capabilities of CFD and to identify the limitations as regards
extent of parameter space that can be studied and their sub- to the expected accuracy. Second and third sections present the
sequent extrapolation beyond the studied parameter space. review of the published models for the turbulence (single phase)
A wide variety of impellers with different shapes are used and the impeller–baffle interaction, respectively. Fourth and fifth
in practice. Different practices of impeller clearance, etc., sections provide the comprehensive survey of CFD of stirred tank
are followed for different impellers and for different appli- with radial flow impellers using RANS and LES.
cations. Computational flow models, which allow “a priori” In order to appreciate the relative merits of these models,
predictions of the flow generated in a stirred reactor of any simulations of the flow field in the stirred tank using various
configuration (impellers of any shape), with just a knowledge impeller-modelling approaches along with the turbulence models
of geometry and operating parameters, can make valuable by the commercial code as well as in-house codes were carried out.
contributions in developing new reactor technologies. An attempt has been made to compare the model predictions of
(5) Development of theories for heat and mass transfer: In case all the three mean velocity components and the turbulent kinetic
of exo/endothermic reactions in stirred vessel, heat needs to energy with the experimental data with emphasis on bulk region.
supplied/removed from the reactor. External jackets or inter- Both the radial flow impeller (part I of this review) and axial flow
nal coils can be provided for this purpose. In majority of impellers (part II of this review) have been covered.
the cases, heat transfer coefficient inside the reactor dictates In part I, second section gives details of various turbulence
the overall heat transfer coefficient. This can be increased by modelling approaches such as DNS, large eddy simulation (LES),
increasing energy dissipated close to tank wall or outside sur- RSM, standard k − ε model, and modified two-equation models.
face of coils. This objective can be satisfied by increasing the The various impeller–baffle interaction models are presented in

| 26 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


third section. The present status of CFD simulation using Reynolds
stress Navier–Stokes (RANS) approach of radial flow impellers is
presented in fouth section. Fifth and sixth sections give compre-
hensive comparison of CFD predictions with experimental data in
bulk as well as impeller region. The CFD simulations of stirred
tanks with radial impellers using LES are reported in seventh
section. In RANS approach, assumptions are made to model the
terms containing fluctuating components and assessment of such
assumptions using LES database is presented in eighth section.
Ninth section covers the overall energy balance, knowledge of
which is expected to be useful in energy optimisation. It also
focuses a light on the comparison of Reynolds stresses (different
components) and the present status of modelling in brief. Finally,
the present status of CFD of stirred tanks has been summarised in
tenth section and possible suggestions have been given for future
work in the area CFD modelling of the stirred tank reactors for
radial flow impellers.

TURBULENCE MODELLING FOR SINGLE PHASE


SYSTEMS Figure 1. CFD simulation using RANS, LES, and DNS models: relative
scale resolutions in a typical energy spectrum.
Overview of Turbulence Modelling
CFD is widely used for understanding the turbulent transport are extremely accurate and non-dissipative tools for calculating
of mass, momentum, and heat in chemical process equipment. derivatives of discrete datasets. Later with the advent of computa-
Understanding of mixing and heat, mass transport rates in the tion power as well as robust and accurate numerical techniques,
reactors is essential for the reliable and efficient design. To achieve finite difference and finite element computations of DNS are possi-
this, information regarding the velocity, pressure, distribution of ble for complex geometries such as stirred vessels. However, DNS
kinetic energy and local dissipation is required in time as well as computation is very time consuming and has extensive storage
space domain. requirements and hence not possible for large-scale equipments
In CFD, the equations of continuity and motion should ideally at even moderate Reynolds numbers. To understand the large com-
be solved directly (DNS). However, because of huge computa- putational cost of DNS, consider that the ratio of the large eddies
tional demands for flows in a variety of equipments, the equations (energy containing) to the small (energy dissipating) scales is
3/2
of change are solved in combination with additional closure equa- proportional to Ret , where Ret is turbulent Reynolds number.
tions arising out of ensemble averaging (standard k − ε model Therefore to resolve all the scales in three-dimensions, the mesh
9/2
and RSM) or space filtering (LES). These various approaches (rel- size will be proportional to Ret . Thus, for high Reynolds number,
ative scale resolutions) can be best represented on an energy the grid size becomes prohibitive. Also added on is the fact that
spectrum as shown in Figure 1. The equations of change for a the simulations will be transient with very small time steps, since
three-dimensional incompressible system can be represented in the temporal resolution requirements are governed by the dissi-
the following form (Pope, 2000): pating scales, rather than the mean flow or the energy containing
eddies.
∂ ∂ui Until now, only DNS computations at moderate Reynolds num-
+ =0 (1)
∂t ∂xi bers are possible. An acknowledged limitation of DNS is its
restriction (by cost considerations) to low Reynolds number. For
 
∂ui ∂ui ∂p ∂ ∂ui channel flow, the approximate number of grid points needed, can
 + uj =− + ␮ (2) be estimated from the expression by Wilcox (1993):
∂t ∂xj ∂xi ∂xj ∂xj

NDNS = (0.088 Reh )9/4 (3)


CFD techniques can be classified as (i) DNS, (ii) LES, and (iii)
Reynolds averaged Navier–Stokes (RANS) approach. These tech-
where Reh is the Reynolds number based on the mean channel
niques are shown in Figure 2.
velocity and channel height. According to the above expression,
to compute a flow with a Reynolds number of 106 , which is
Direct Numerical Simulation (DNS) more realistic, we would require approximately 133 billion grid
In the DNS, it is theoretically possible to directly resolve the whole points, which is astronomical. To reduce this cost to some extent,
spectrum of turbulent scales. The main purpose of DNS is to solve Reynolds number scaling is used whenever possible depending
for the turbulent velocity field without the use of “turbulence mod- on the observed dependence of the flow on the Reynolds number
elling.” This condition means that the Navier–Stokes momentum without changing the essential physics. DNS can be used to
equation for fluid must be solved exactly. It is now possible to compute a specific fluid flow state. It can also be used to compute
examine fully developed turbulence flow fields at a microscale (up the transient evolution that occurs between one state and another.
to Kolmogorov scale) and perform extremely accurate calculations DNS using high-performance computers is an economical and
of flow. This requires accurate estimation of derivatives involved mathematically appealing tool for study of fluid flows with simple
in Navier–Stokes equation. The early DNS computations of turbu- boundaries, which become turbulent. DNS is used to compute
lence for simple geometries utilised the spectral methods, which fully nonlinear solutions of the Navier–Stokes equations, which

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 27 |


Figure 2. Overview of turbulence models.

capture important phenomena in the process of transition, as 0.15 m with Reynolds number of 7275. The authors used MRF
well as turbulence itself. DNS is mathematical, and therefore, can approach for simulation impeller–baffle interaction, though they
be used to create simplified situations that are not possible in an mentioned that for DNS simulations sliding mesh (SM) technique
experimental facility, and can be used to isolate specific phenom- is necessary. Their geometry was described by a block structured
ena in the transition process. As we are aiming at a nearly exact grid of 15 blocks. The grid size also was small (1 901 824) for DNS
solution (and not “the exact”) to specific turbulent flows utilising as compared to (2 082 816) for k − ε model for the same half simu-
limited computational resources, DNS is stressed as a research lation of the vessel. The average Kolmogorov scale was 6 mm and
tool and not as a brute-force solution to engineering problems. average grid size was 0.8 mm. As the impeller region has very
The objective of DNS is not necessarily to reproduce real-life high shear as well as energy dissipation, 35% of the grid points
flows (say the flow over an airplane), but to perform controlled were located in the impeller region (which consists only 7% of
studies that provide better insight, scaling laws, and turbulent reactor volume) to resolve the trailing vortices. Although both the
models to develop. In aerodynamics, DNS is associated with a simulations (DNS and k − ε model) were in good agreement with
large-scale computationally intensive solution procedure, which the experimental data the authors found that DNS captured few
may consume hundreds to thousands of Cray super-computing more details than k − ε model. The DNS simulations showed two
resources. The earliest use of DNS began in the 1970s and with small secondary vortices behind the blades which were missed by
the growth in the computational power today, it is getting more the k − ε model, even the flow structure predicted by k − ε model
and more popular day-by-day. DNS of stirred tanks has been car- was wavier than DNS. The shape of regions with increased tur-
ried out in literature by Bartels et al. (2000); Verzicco et al. (2004); bulent kinetic energy was closer to experimental results for DNS
Sbrizzai et al. (2006); and Gillissen and van den Akker (2009). as compared to k − ε model, but in the outer flow region k − ε
These simulations are reported at very low Reynolds number rang- model gave more uniform results of turbulent kinetic energy. The
ing from 1600 to 7300. The simulations carried out by Verzicco maximum energy of turbulent kinetic energy by DNS was much
et al. (2004); and Sbrizzai et al. (2006) were in unbaffled tank higher (k/U2tip = 0.118) and closer to the experiments, whereas
whereas Bartels et al. (2000); and Gillissen and van den Akker k − ε model (k/U2tip = 0.079) underpredicted.
(2009) carried out simulations in baffled tank. These two inves- Gillissen and van den Akker (2009) simulated six bladed Rush-
tigations have been briefly reported below. ton turbine and the impeller–baffle interaction was modelled with
Bartels et al. (2000) simulated stirred tank with Rushton turbine adaptive force field technique (AFT). Their grid size was of 1
using k − ε model and DNS and compared the results of both with billion and Reynolds number of 5000. They simulated 100 000
experimental results of Schäfer et al. (1998). The simulations were time steps for a month on 128 cores. The authors focused on
performed with silicon oil as working fluid in a tank diameter of comparison between LES and DNS. They also showed

| 28 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


 
visualisations of the vortical structures in the wake of a single ∂ui  ∂ui  ∂p ∂ ∂ui 
turbine blade.  + uj  =− + ␮ −ui uj  (8)
∂t ∂xj ∂xi ∂xj ∂xj

Large Eddy Simulation The Reynolds stresses (−∂ui uj /∂xj in Equation 8) arising out
In LES, large scales eddies are resolved and the small scales, which of ensemble-averaging procedure are modelled according to the
are isotropic in nature, are modelled using subgrid scale models. Boussinesq hypothesis and given by:
The major role of subgrid scale models is to provide proper dis-  
sipation for the energy transferred from the large scales to small 2 ∂ui  ∂uj 
scales. Because of these advantages, LES technique is gradually −ui uj  = kıij −␮t + (9)
3 ∂xj ∂xi
becoming a popular tool to investigate turbulent flows. LES stud-
ies have several valuable advantages. They are (i) much cheaper In order to close the system of equations (7)–(8), ␮t has been for-
(ii) easier to handle than field measurements. They provide (iii) mulated in many ways in the literature in terms of zero, one, and
three-dimensional fields of flow characteristics under (iv) well- two-equation models. Some of the popular models are discussed
controlled external condition. They are (v) more accessible in the below.
sense that LES can be run in the conditions where field measure-
ments would be difficult. Moreover, they are (vi) very helpful for Standard k − ε model
conceptual understanding of detail hydrodynamic of the chemi-
Standard k − ε model is essentially a high Reynolds number model
cal process system. The motivation behind LES is the recognition
and assumes the existence of isotropic turbulence and the spectral
that the large scales of the turbulence often dominate mixing,
equilibrium. In the case of standard k − ε model, ␮t is estimated
heat transfer, and other quantities of engineering interest, while
using turbulent kinetic energy (k) and rate of dissipation of tur-
the small scales are only of interest because of how they affect the
bulent kinetic energy (ε):
large ones. Furthermore, large scales are not universal and vary
according to the geometry and flows under consideration, while
k2
small scales exhibit a more or less universal behaviour, which ␮ t = C␮  (10)
considerably simplifies the task of modelling the SGS stresses. ε
In an LES, the flow field is decomposed into a large-scale or
The modelling of k and ε equations result into five turbulence
resolved component and a small-scale or subgrid-scale compo-
parameters C␮ , Cε1 , Cε2 , k and ε . These parameters have been
nent:
estimated from studies in simple flows. In the log-law region of
the boundary layer, experimentally it has been found that the
u = ū + u (4)
turbulence production and the dissipation terms are much larger
than the other terms. Turbulence in this region is considered to be
The conservation equations governing the filtered velocity
in local equilibrium. Neglecting transport and production in this
field are obtained by applying the filtering operation to the
flow, value of C␮ = 0.09 has been obtained. For the value of Cε2 ,
Navier–Stokes equations (Equations 1 and 2). We consider
experiments have been carried out in decaying grid turbulence.
spatially uniform filters, so that filtering and differentiation com-
Turbulence is generated when free flow goes through a grid which
mutes.
generates mean flow gradients and in turn via the production
When the Navier–Stokes equations for constant density (incom-
term generates turbulence. Sufficiently far downstream, the veloc-
pressible flow) are filtered, one obtains a set of equations very
ity gradients are zero and hence the production and dissipation
similar in form to the RANS equations:
terms are zero in contrast to log-law region in the boundary layer.
∂ū The turbulent diffusion term is also negligible. The balance of
=0 (5) turbulence convection and dissipation gives Cε2 = 1.92. The con-
∂xi
stant Cε1 is obtained by looking at the dissipation rate equation
∂ū ∂(ūū) 1 ∂p̄ ∂2 ū ∂ in the log-region of boundary layer. The turbulent diffusion is not
+ =− + 2− (6) negligible in this case unlike k equation. The simplifications of ε
∂t ∂xi  ∂xi ∂xi ∂xi
equation using velocity profiles in log law region and the numeri-
Various subgrid scale models are required for estimation of cal solution of these equations give Cε1 = 1.44, k and ε = 1.0 and
effect of subgrid scale stresses () such as Smagorinsky model, 1.3, respectively.
dynamic Smagorinsky model, scale similarity model, one equa- The standard k − ε model has performed satisfactorily in many
tion subgrid scale kinetic energy model, etc. These models are flows, but the applicability of this model is limited due to uncer-
discussed in Mathpati and Joshi (2007). tainties involved in the modelling of turbulence production,
turbulent transport and the assumptions in modelling dissipation
Reynolds Averaged Navier–Stokes (RANS) Approach rate equation. Further, the modelling of transport equations for
k and ε pose the difficulties to account for streamline curvature,
DNS and LES require huge computational power and are not pos- rotational strains, and the other body-force effects.
sible for very large-scale industrial contactors at high Reynolds
number. In most of the cases, it is necessary to understand the
Zonal model
gross flow patterns. This can be achieved by solving the ensemble-
averaged Navier–Stokes equation. In case of flows with very strong inhomogenity, the standard tur-
Equations (1) and (2) on ensemble-averaging reduce to the bulence model parameters fail to predict the flow field. Hence the
following form: entire domain is divided into various zones and a set of turbulence
model parameters (chosen on the basis of prior experience so as
∂ ∂ui  to improve upon the predictions) is assigned to each zone. Such
+ =0 (7) efforts in stirred tank are reported by Sahu et al. (1998). Thus
∂t ∂xi

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 29 |


Table 2. Source terms for variables in general transport equation

∂u1  ∂u2  ∂u3 


     
General transport equation: ∂x1 + ∂x2 + ∂x3 = ∂
∂x1 eff ∂
∂x + ∂
∂x2 eff ∂
∂x + ∂
∂x3 eff ∂
∂x + S
1 2 3
 eff S
1 0 0
∂p
u1  + t − ∂x
1
∂p
u2  + t − ∂x
2
∂p
u3  + t − ∂x
3
k  + t /k Pk −ε For all the variants of k−ε model
ε
ε  + t /k k (C1 Pk −C2 ε)  Standard k−ε model, zonal model
Pk 2
C1,CK kε Pk + C3,CK ε
ε −C2,CK k ε Optimised Chen Kim k−ε model
 
C1,RNG kε Pk −˛ kε ε−C2,RNG kε ε RNG k−ε model
  
∂ui  ∂uj  ∂ui  ∂uj 
Where Pk = t ∂xj + ∂xi ∂xj + ∂xi
∂u  u   ∂u  u  
 ku  u   ∂u  u  


RSM Transport equations for Reynolds stresses: ∂t
i j
+ uk i j
∂xk = ∂x1 Cs i j
ε ∂x1
i j
−C1 kε (ui uj −(2/3)ıij k)−C2 (Pij −(2/3)ıij P−Sij )−(2/3)ıij ε
 
   ∂ uj     ∂ui  1
Where Pij = − ui uk − uj uk and P = 2 Pii
∂xk ∂xk
∂ε ∂ui ε
 
Cε k
∂   ∂ε 2
Equation for the turbulent energy dissipation rate: ∂t + ∂xi = ∂xi ε ui ui  ∂x + Cε1 kε Gk −Cε2  εk
i

effectively, it identifies the different regions of distinct physics of the near wall flows. Table 2 gives the source terms along with
flow, modifies the values of the turbulence parameters accordingly the modified eddy viscosity formulation. It contains more num-
and effectively prescribes the zonal eddy viscosity distribution. ber of turbulence model parameters (Table 3) and is applicable
The selection and distribution of turbulence parameters in stirred to the swirling flows wherein strain gradients are large where the
tank is largely arbitrary. Nere et al. (2001) carried out CFD simula- simple eddy viscosity relation of Boussinesq leads to non-realistic
tion of stirred tanks using zonal modelling for C␮ . They have also values of the eddy viscosities.
developed a new constitute equation for eddy viscosity which gave
better agreement with experimental data for axial flow impellers. Multiscale models
Some investigators have used model parameters as a function of In contrast to the assumption in k − ε models as regards to the
geometrical and operating variables, rather than constant values. presence of equilibrium and hence the presence of a single time
For example in jet flows, model parameters are related with jet scale, these models assume the existence of two time scales Placek
half-width and centreline velocity (Launder et al., 1972; McGuirk et al. (1986). They have proposed a multiscale model for stirred
and Rodi, 1977; Rodi, 1993; Mathpati et al., 2009). tanks. It considers the existence of non-spectral equilibrium. The
energy spectrum is divided in three parts, namely production
RNG k − ε model region, a transfer region and a region in which dissipation takes
RNG k − ε model uses framework of two equations and these place. It assumes the existence of the spectral equilibrium only
equations are derived by application of renormalisation group between the transfer region and the region in which turbulence
methods to the original governing equations of momentum trans- is dissipated. Thus, it considers the presence of two time scales,
fer. This model gives the eddy viscosity relation based on the result namely one for the production range and the other for the dis-
of application of the renormalisation group theory where inertial sipation. It involves two equations: one for the transport of the
sub-range eddies are eliminated from the equations of motions turbulent kinetic energy of the large-scale vortices and the other
to yield averaged flow quantities at the integral scale of turbu- for transport of the turbulent kinetic energy of inertial sub-range
lence. This model is capable of modelling the distance between eddies in addition to the equation for the energy dissipation rate.
the production and dissipation of turbulent kinetic energy better Tables 2 and 3 give the source terms and the values of the con-
than the standard k − ε model. As opposed to the semi-empirical stants involved. Utility of this model needs to be extensively
models these are more fundamental and are capable of predicting proved for the prediction of flow pattern in stirred tanks.

Table 3. Various turbulence models and the turbulence parameters

Turbulence model Turbulence parameters

Standard k − ε model (Launder and Spalding, 1974) C␮ = 0.09, Cε1 = 1.44, Cε2 = 1.92, k = 1.0, ε = 1.3
Modified k − ε model (Ranade and Joshi, 1989) C␮ = 0.125, Cε1 = 1.59, Cε2 = 1.62, k = 1.0, ε = 1.3
New k − ε model (Nere et al., 2001) C␮ = (1/RUtrip )(k 2 /ε), Cε1 = 1.44, Cε2 = 1.92, k = 1.0, ε = 1.3
High Re RNG k − ε model C␮ = 0.0845, Cε1 = 1.42, Cε2 = 1.68, k,RNG = 0.7194, ε = 1.3
Chen–Kim k − ε model C␮ = 0.09, Cε1 = 1.15, Cε2 = 1.90, Cε3 = 0.25, k = 0.75, ε = 1.15
Optimised Chen–Kim k − ε model C␮ = 0.09, Cε1 = 1.36, Cε2 = 1.90, Cε3 = 0.04, k = 0.75, ε = 1.15
RSM C␮ = 0.09, Cε1 = 1.44, Cε2 = 1.92, Cs = 0.22, C1 = 1.8, C2 =
 
0.60, C1 = 0.5, C2 = 0.18, k = 1.0, ε = 1
LES  = 0.42
Cs = 0.23, Smagorinsky constant (for homogeneous isotropic
turbulence in the inertial sub-range)

| 30 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Chen–Kim k − ε model sified into two categories, namely steady and unsteady state.
Chen–Kim k − ε model (Chen and Kim, 1987; Jenne and Reuss, These have been briefly described below:
1999) assumes the existence of non-equilibrium between the pro-
duction of turbulence and its dissipation and hence makes use Steady State Approaches
of two time scales: one for the production range and the other The model equations are solved in steady state mode and these
for the dissipation range. Thus it contains two constants (Cε1 and include the black box or impeller boundary condition (IBC)
Cε3 ) instead of Cε1 (in the modelled equation of energy dissipation approach, the source-sink approach (SS), multiple reference frame
rate) in k − ε model. Overall, it uses the concept of the three- (MRF) approach, inner–outer (IO) approach and the snap shot
equation model (consideration of two time scales) but effectively approach. Figure 3A–D gives a schematic representation of the
uses only two equations for the transport of the turbulent charac- impeller modelling approaches.
teristics (similar to two-equation models). For equilibrium flows
the Chen–Kim k − ε model reduces to the standard k − ε model. Impeller boundary condition approach (IBC)
The optimised version of Chen–Kim k − ε model (OPT-CK) sug- Here for the sake of simplicity, the impeller is treated as a
gests the optimised values of Cε1 and Cε3 based on the objective black box. The action of the impeller is simulated by provid-
of minimising the deviation between the measured and predicted ing the boundary conditions (experimentally measured fluid flow
length scales along the impeller centreline. Tables 2 and 3 depict characteristics) on a selected surface surrounding the impeller
the source terms and the values of constants used in Chen–Kim (Figure 3A). Various authors have used IBC approach (Harvey
k − ε model and its optimised version. This model has been found and Greaves, 1982a,b; Ranade and Joshi, 1990b; Kresta and Wood,
to give improved predictions of mean flow pattern. 1991). The boundary conditions consist of mean velocities (radial,
axial, and tangential) and the turbulent quantities such as the tur-
Reynolds Stress Model (RSM) bulent kinetic energy (k) and its dissipation rate (ε), that have to
It is clear that the standard k − ε model inherently fails to pre- be measured experimentally for the system under consideration.
dict properly the anisotropic flow situations (Reynolds, 1987; Thus, if the impeller design is changed or the tank geometry is
Launder, 1990; Hanjalic,́ 1994). Further, the modelling of trans- modified, it becomes necessary to again measure these boundary
port equations for k and ε clearly brings out the difficulties to conditions in order to simulate the new system. This was the start-
account for streamline curvature, rotational strains, and the other ing point of CFD simulations in stirred tanks. Once the boundary
body-force effects. RSM, in theory, can circumvent all the above conditions are specified, the transport equations are solved in rest
mentioned deficiencies and also it has an ability to predict more of the flow domain to predict the flow field prevailing in entire
accurately each individual stress. A RSM solves six equations for tank.
the Reynolds stress tensor and another equation for the dissipa- The applicability of black box approach is severely limited by
tion rate. The equation of turbulent stresses contains the terms, the availability of the experimental data. The impeller design, ves-
namely pressure strain rate (which contains fluctuating pressure sel geometry, operating conditions, physicochemical properties,
velocity gradients) and the flux of Reynolds stresses. In addition etc., have a profound effect on the flow generated by the impeller
to the modelling of the terms in the turbulent energy dissipa- and hence this approach cannot be used to predict the flow gener-
tion rate equation, these terms need to be modelled accurately. ated by a variety of impeller-vessel configurations. Also different
Therefore, though RSM takes into consideration the transport of choices of the IBCs lead to significantly different predicted flow
turbulent stresses, its parameters are not universal (i.e., model field and hence the selection of the impeller swept boundaries
needs to be calibrated for different types of flows). The pressure also poses a problem. Numerical simulations using this approach
strain term in RSM is responsible for making turbulence isotropic can at the most be extended to the geometries very much similar
and redistribution of energy between components u 21 , u 22 , and to that for which the experimental data are available. Also exten-
u 23 . Further, the extended system often has difficulties to produce sion of such an approach to multi-phase flows and to industrial
converged solutions and thus it is computationally expensive. The scale reactors is not feasible because it is virtually impossible to
modelled equations are given in Table 2 while the typical values obtain (from experiments) accurate boundary conditions for such
of the constants are given in Table 3. systems. More importantly, this approach cannot be used to make
Algebraic stress model (ASM) is the simplification of the RSM “a priori” simulations. It cannot, therefore, be used as a design
wherein the differential transport terms are approximated in terms tool.
of algebraic expressions. This model implicitly determines the
Reynolds stresses (locally) as a function the turbulent kinetic Source-sink approach (SS)
energy, its dissipation rate and the mean velocity gradients. In this approach the rotation of impeller is modelled as a source of
Thus computational expenses are reduced. These models are momentum, whereas baffles are represented as sinks of momen-
inherently less general and less accurate than Reynolds-stress tum. Pericleous and Patel (1987) first proposed this approach and
models. then it was modified and further used by Xu and McGrath (1996);
and Patwardhan (2001) to simulate the flow pattern produced by
MODELLING OF THE IMPELLER–BAFFLE the pitched blade turbines of different angles in turbulent regime.
In this approach, the impeller blades are replaced with a number
INTERACTION of blade sections by dividing the blade into a number of vertical
The overall development of the CFD models for the prediction of strips from the hub to tip. If any of such strip is curved (the angle
flow pattern in stirred tanks has been carried out on the two fronts, changes from the top to bottom of such a strip), the strip is fur-
namely modelling of the impeller (rotating)–baffle (stationary) ther divided into a number of sections so that every section (of
interaction and the modelling of turbulence. the blade) can be assumed to be practically flat. The blade section
Different approaches have been adopted for the modelling of inside each strip is approximated to an airfoil and airfoil aerody-
impeller–baffle interaction. These approaches can be broadly clas- namics is applied. Axial and tangential momentum sources are

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 31 |


Figure 3. Overview of impeller–baffle interaction modelling approaches.

similar to that given by Pericleous and Patel (1987) (these source the system attains a satisfactory numerical convergence. Each of
terms can easily be calculated in terms of the lift and drag coef- the inner and outer simulations is carried out under steady state
ficients by using the velocity–force diagram for the blade section assumptions in its own frame of reference. As the two frames
under consideration) and the radial momentum source is derived are different, the information that is iteratively exchanged is cor-
by assuming that the interaction between the impeller and fluid rected for the relative motion and averaged over the azimuthal
results in forces perpendicular to the blade surfaces. The bound- direction. The main feature of this approach is the existence of
ary conditions on the tank wall and baffles are specified as the the overlapping region, common to the inner and outer zones,
solid surfaces. Further, the impeller blades are also modelled as which provides the iterative matching of the two solutions. The
walls. In this approach, the full tank needs to be modelled and extent of this region and the exact location of its boundaries are
not a quarter of the tank as is practiced in some other cases. largely arbitrary.

Inner–outer approach (IO) Multiple reference frame technique (MRF)


In this approach the vessel is divided into two partly overlap- In this approach, the mesh block associated with the impeller is
ping zones. Daskopoulos and Harris (1996); and Brucato et al. rotated with it, while the outer block associated with the baffles is
(1998) have performed simulations using inner–outer approach. kept stationary (Figure 3B). In MRF approach, there is no overlap
The region containing the impeller is called as inner region and between inner and outer regions. In this approach, flow character-
the bulk of the vessel including the baffles is called as the outer istics of the inner region are solved using a rotating framework.
region. First, a simulation of the flow in the inner domain is These results are used to provide boundary conditions for the
carried out in a reference frame rotating with the impeller, with outer region (after averaging over azimuthal direction). The flow
arbitrary boundary conditions imposed on the boundary surface in the outer region is solved using stationary framework. The solu-
of inner zone. A first trial flow field is thus computed in the whole tion of the outer region is used to provide boundary conditions
impeller region, including the distributions of velocity, turbulent for the inner region. A few iterations over inner and outer region
kinetic energy and its dissipation on the inner surface of the outer lead to a converged solution. Further, the choice of this surface
zone. The latter distributions are used as the boundary conditions between inner and outer region is not arbitrary, since it has to be
for a first simulation of the flow in outer domain that is carried selected a priori as a surface where flow variables do not change
out as black box approach (IBC method). Now the information on appreciably either with angular location or with time. Full geom-
the flow field, including a first estimate of mean velocities, turbu- etry needs to be modelled and the impeller blades are modelled
lent kinetic energies and the energy dissipation rates on the outer as walls.
boundary surface of the inner domain are available. These values Brucato et al. (1994); and Harris et al. (1996) have applied
are in turn used for the second inner simulation and so on, until inner–outer method to simulate flow in stirred vessels whereas

| 32 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Marshall et al. (1996) have used a MRF approach. The MRF with respect to the laboratory reference frame, while the grid is
approach is computationally less intensive than the inner–outer assumed to rotate with the impeller speed. This is done by explic-
method. This approach is available with the various commercially itly adding acceleration terms to the momentum equation as the
available CFD codes. The MRF and IO techniques though are capa- rotation of the grid results in acceleration terms which are com-
ble of simulating the impeller action using steady simulations; pletely equivalent to the body forces arising in the non-inertial
they are computationally expensive as compared to the black box frames and one can assume that the impeller action is properly
approach. IO approach is in fact defines inner and outer zones modelled. The two regions are implicitly coupled at the interface
with finite overlap where as in MRF approach; there is no overlap separating the two blocks via a sliding-grid algorithm, which takes
between inner and outer regions. into account the relative motion between the two sub-domains
and performs the required interpolations. In order for the peri-
Snapshot approach (SNAP) odic boundary conditions to be imposed along the  direction, the
In order to model the impeller rotation, Ranade and Dommeti computational domain must include an integer number of both
(1996) have proposed a computational snapshot approach. In this blades and baffles; therefore, its minimum azimuthal extent (in
approach, the whole solution domain is divided into two regions radians) is 2 divided by the highest common factor of number
(Figure 3C). In the inner region surrounding the impeller, time of blades and the number of baffles. SM approach has a potential
derivative terms are approximated in terms of spatial derivatives. to generate “a priori” predictions without requiring any experi-
In the outer region, time derivative terms are usually quite small mental input. It can, therefore, be used as a design tool to screen
in magnitude in comparison with the other terms in the govern- different configurations.
ing equations and are, therefore, neglected. The time dependent The SM approach suffers from the large computational
term in impeller swept volume was approximated in terms of expenses. One is required to add explicitly the acceleration terms
impeller rotation speed and the spatial gradient. This spatial to the momentum equations (accounting for the moving grid sys-
derivative term is added to the source term in the steady state tem). Numerical problems at the grid interface are inherently
equation for the general transport variable. In order to repre- created particularly, when the blade shape is complex. Also the
sent the rotation of the impeller blades and hence the suction requirement of the interpolation procedure at the interface, which
and ejection of the fluid from the back and front side of the results into the reduced efficiency and potentially also the numer-
blade, respectively, they have used additional mass sink and ical accuracy of the scheme, as compared to the single (stationary
source terms for the cells adjacent to the impeller. This approach or fixed) grid approach. The presence of rotating grids requires
was described as snapshot approach as it gives a picture of proper care to be taken while handling the grids at the interface
(steady) flow prevailing in a stirred vessel for a certain position between the two regions, that is, mass flow continuity should be
of impeller blades. The principal components of the snapshot satisfied at the interface. This requires a lot of book keeping, as
approach consists of the approximation of the time dependent a result, the method of solution is complex, and computational
term in terms of the spatial derivative and the assumption of its requirements are enormous. With the advent of faster comput-
negligible magnitude in the bulk region and hence its removal ers together with larger storage capacity, this approach is slowly
from the momentum equations to be solved. Further, the assump- becoming feasible. The time required to get the converged solution
tion that the baffles have no effect on flow (subsequently sink or ranges between days to weeks on the presently available fastest
sources) in-between the impeller blades, which leads to the same computers. As it relies on solution of full time varying flow in a
impeller box irrespective of the impeller–baffle interaction seems stirred vessel, its computational requirements are greater by order
to inadequate/incorrect. In addition, the model predictions have of magnitude than those required by the steady state simulations.
not been validated over the entire flow domain for all the flow Because of the excessive computational requirements, there are
variables. restrictions on number of computational cells that can be used
for the simulations. Such a limitation may make “a priori” pre-
dictions of desired flow characteristics such as energy dissipation
Unsteady State Approaches
rates, shear rates near impeller blades, etc., difficult. The results
The time-dependent interaction of the impeller and the fluid obtained using this approach are not yet sufficiently validated
is modelled. This approach makes the use of the SM, moving- for turbulent regime and the extension to multiphase systems is
deforming grid technique (MDG) and impeller modelling within rather difficult. All these considerations make SM approach less
lattice-Boltzmann-LES framework (AFT). attractive as a reactor-engineering tool.

Sliding-mesh approach (SM)


Murthy et al. (1994) have made use of SM approach to model the Moving-deforming grid technique
impeller rotation in stirred tanks. In this approach, full transient This was developed by Perng and Murthy (1994), wherein, a sin-
simulations are carried out using the two grid zones (Figure 3D). gle grid spans both the stationary and moving parts. The grid
The flow domain in a stirred vessel is divided into two cylindrical, attached to the impeller moves with it and causes deformation of
non-overlapping sub-domains. Inner domain contains the rotat- grid throughout the vessel including the impeller and the baffles.
ing impeller while the outer contains the stationary baffles. Each In this approach the grid is time dependent and it moves and con-
domain is grided as a separate block. The outer grid is fixed in the sequently deforms during the calculations. The grid around the
laboratory reference frame, while the inner grid rotates with the impeller rotates and after certain deformation (predefined skew-
impeller. The detailed geometry of impeller needs to be modelled. ness of the grid for minimum desired mesh quality) is achieved,
Impeller blades are modelled as solid rotating walls. Flow within the grid is brought back to its original form and the properties are
the impeller blades is solved using the usual transport equations transferred to the restored grid in a conservative fashion, conse-
unlike the black box approach described earlier. Further, unlike quently taking care of the interfacial continuity. The grid motion
in the MRF and IO approaches, in SM, the flow equations in the is accounted for by transforming the time derivative in Eulerian
inner block (containing the rotating impeller) are formally written conservation equations into Lagrangian form.

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 33 |


Though this approach seems to closely simulate the transfer coefficient, etc. In this context, validated CFD model can
impeller–baffle interactions, it suffers from many noteworthy dif- be a promising tool. However, the CFD is still in the developmental
ficulties apart from the disadvantages offered by SM approach. stage because of the inadequate understanding of the physics of
The main difficulty is in controlling the grid quality and as a turbulence and the available computational speed. In both of these
consequence the numerical accuracy of the results. The computa- directions, a continuous development has been occurring and,
tional expenses are the highest amongst all the impeller modelling during the past 20 years, a fairly good progress has been achieved
approaches. in turbulence modelling as well as numerical techniques.
In view of the above, in the past, the flow generated by disc tur-
bine (DT) and pitched blade turbines (PBT) have been subjected
Impeller modelling within lattice-Boltzmann-LES to detailed experimental and computational studies. The initial
framework (adaptive force field technique, AFT) objectives were preliminary such as the estimation of gross flow
Lattice-Boltzmann technique is an alternative discretisation tech- parameters and the average flow field. However, during the past
nique which offers many advantages such as the computational 10 years, the objectives have become deeper such as the character-
efficiency, ease of parallelisation and its unique feature of capabil- isation of turbulent flow, flow instabilities, etc., and tailoring the
ity of forcing the boundary conditions at any spatiotemporal point. impeller design so as to get the desired flow field. Majority of the
In this approach fluid is assumed to be a system comprising of work has been devoted to the modelling of the mean and the tur-
many particles which obey the conservation laws. The conserva- bulence flow field and its comparison with the experimental flow
tion of mass and momentum are applied for each of the particles measurements. However, the comparison of predictive capabili-
and equations are obtained which are equivalent to the incom- ties of various turbulence models has been preliminary even for
pressible Navier–Stokes equations. Here each particle is assumed the flow generated by conventional impeller designs like DT and
to occupy a corner of the lattice and after each time step particles PBT. It is well known that the k − ε model assumes the isotropy
are supposed to move a lattice distance and collide with other for turbulence, and therefore, anisotropic models such as RSM
particles. The forces acting on the flow are calculated in such a and the LES model are being recommended for the simulation
way that the field has prescribed velocities at points within the of the complex three-dimensional flows. However, RSM has got
domain. This technique is called the AFT. The impeller and tank shortcomings like, non-universal model parameters, numerical
wall are defined as a set of control points on their surface. At the difficulties and is computationally expensive by an order of mag-
points on the impeller surface tangential velocities are in accor- nitude as compared to the k − ε model. Further, the RSM model
dance with the speed of impeller rotation and the distance of the does not capture the time-dependent nature of the flow. This limi-
surface point from the axis of rotation are prescribed. On the tank tation is overcome by the LES approach. During the last few years,
wall, the tangential velocity is set zero. To maintain the fixed val- the ability to resolve all but the smallest turbulence scales using
ues at the surface points, a control algorithm is used, wherein at LES has become more viable. LES can potentially produce more
each time step, mismatch between the actual flow velocity and accurate results by modelling only the smallest scales, which tend
the prescribed flow velocity at the control points is calculated and to be more isotropic, while fully resolving the turbulence at the
accordingly adapts the force field in such a way that it suppresses larger scales. From a practical point of view, the use of LES or DNS
mismatch. This AFT eliminates the necessity to grid the different as design tools is far from utility due to the high computational
impeller-vessel geometries. New designs can be simulated just by costs associated with them. For these reasons, it is envisioned
defining the surface points. Though this technique offers accurate that the RANS equations associated with turbulence modelling
modelling of impeller rotation and models the flow in detail, and will be the main CFD tool used by the practitioners and part of
has proven its accurateness of the prediction of flow pattern, it the research community, at least in the near future. As a result of
suffers from the limitation of the requirement of large demand on this trend, there is a need to improve the accuracy and reliability
both the time and storage front. Though it cannot be used widely of the solutions of turbulent flow fields obtained from the RANS
due to enormous computational requirements, it can be used as a equations.
research tool for better insight and detailed understanding of the In this sub-section, the studies pertaining to the CFD mod-
complex flow in stirred tanks. It can also be used as an alterna- elling of flow pattern in stirred tanks agitated by radial flow
tive to the experimental characterisation of the flow. Also it can impellers has been discussed. Table 4 provides the details of sim-
be of great help in understanding the flows near the walls and the ulations carried out by various investigators. Harvey and Greaves
regions in which measurements are difficult. (1982a,b) have simulated the flow in turbulence with IBC (black
box) approach using the standard k − ε turbulence model for the
first time. They have treated baffles as a source of a pressure-
RANS SIMULATION OF RADIAL FLOW IMPELLERS induced drag on swirling component of the flow. 3-D effects are
Stirred vessels are extensively used in the chemical process transferred mainly through pressure gradient term, which tends
industries over a wide range of applications. In majority of to deplete and reinforce the swirl momentum field before and after
cases, the flow field in baffled stirred vessels is highly tur- the baffle. They have given the comparison for the radial and axial
bulent hence it is three-dimensional, complex and chaotic in velocity that were in qualitative agreement. Deviations as much
nature. During the last 25 years, there have been continuous as 25% in radial velocity and 100% in tangential velocity were
efforts on understanding these flows using both sophisticated observed. Comparison for radial, axial, tangential velocities, and
experimental and computational fluid dynamic tools. Despite the the turbulent kinetic energy was presented mainly in the impeller
sophistication available in the flow visualisation (laser Doppler discharge stream. No attempt was made to validate the predictions
anemometry [LDA], ultrasound velocity profiler [UVP], particle of the energy dissipation rate.
image velocimetry [PIV], and hot film anemometry [HFA]) it is Middleton et al. (1986) have carried out the CFD simulation of
practically not possible/economical to obtain detailed local flow the turbulent flow using the black box approach with the standard
information throughout the vessel such as velocity profiles, tur- k − ε turbulence model. No quantitative comparison of the mean
bulence properties, energy spectra, mixing time, heat and mass flow and turbulent characteristics was presented in graphical

| 34 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


|
Table 4. RANS simulation of radial flow impellers

Refs. Impeller Geometric details Grid Impeller model RANS model

Harvey and Greaves (1982a, Rushton turbine D = 0.228 m, T = 0.456 m, flat enclosing 950 rpm Black box approach k − ε model
1982b) wall at the top, impeller blade
height = 0.022 m
Middleton et al. (1986) Rushton turbine T = 0.31 m, HL = 0.42, D = 0.15 m: N = 600, 100, 50 rpm Black box approach k − ε model

VOLUME 89, FEBRUARY 2011


T = 0.91 m, HL = 0.91 m, D = 0.30 m,

|
C/T = 0.33
Placek et al. (1986) Rushton turbine T = 0.294 m, D/T = 0.33, and 0.25 Re = 63 300 and 72 800 Black box approach kp − kT − ε multiple-scale
model
Pericleous and Patel (1987) Straight blade turbine, Source-sink approach
Rushton turbine
Ju et al. (1990) Rushton turbine Three tank geometries: T = 0.292 m, 26 × 30 × 25 and Black box approach Non-isotropic k − ε model
viscosity = 5 centiStokes, D = 76, N = 250: 23 × 30 × 25
I, D = 102, N = 150: II, D = 102, N = 200:
III
Ranade and Joshi (1990b) Rushton turbine T = 0.3 m, C/T = 1/2, D/T = 1/3, H = T Black box approach k − ε model
Kresta and Wood (1991) Rushton turbine T = 0.456 m, D/T = 0.5 and 0.33, H = T, 30 × 30 × 26 Black box approach k − ε model
C/T = 0.5, N = 950 rpm, FLUENT
Luo et al. (1993) Rushton turbine T = 0.294 m, C/T = 0.33, D/T = 0.33, 0.5, 28 × 45 × 120 (r, z, ,) MRF k − ε model
0.2
Perng and Murthy (1993) Rushton turbine 21 × 19 × 24 MRF
Dong et al. (1994) Eight straight bladed Unbaffled vessel, 100 rpm, T = 0.1 m, 39 × 11 × 23 MRF k − ε model
turbine D/T = 0.33, C/T = 0.33, H = T
Luo et al. (1994) Rushton turbine T = 0.292 m, C/T = 0.33, H = T, D/T = 0.33 Source-sink, MRF
Brucato et al. (1994) Rushton turbine 60 × 72 × 26 Inner outer approach k − ε model
Tabor et al. (1996) Rushton turbine T = 0.27 m, D = 0.093 m, C/T = 0.33 50 000, 120 000 Source-sink MRF and k − ε model
SM

|
Lee et al. (1996) Rushton Turbine T = 0.1 m, H = T, D/T = 0.33 46 016, 102 296, 138 632,
N = 2165 rpm
Re = 40 000 SM
Ciofalo et al. (1996) Rushton turbine, T = 0.6 m, D = 0.3 m, C = 0.3 m, and 26 × 46 × 16 and MRF k − ε model, second-order
straight blade turbine T = 0.19 m 20 × 30 × 16 differential stress model
Rigby et al. (1997) Rushton and flat bladed T = 0.294 m, w = 0.2D, D/T = 1/3, C/T = 1/3 SM
turbine
Ranade (1997) Rushton turbine 28 × 28 × 78, Source-sink approach k − ε model
Re = 32 000—5135 Snapshot approach
regime
Jenne and Reuss (1997) Rushton turbine T = 0.444 m, D/T = 0.33 = C/T N = 333 rpm Black box approach k − ε models, Chen–Kim
modified k − ε and RNG
k − ε models
Venneker and van den Akker Rushton turbine T = 0.286 m, D = 0.0953 30 × 28 × 50 Re = 5500 Black box approach k − ε model, Power law
(1997) (n = 0.77) model
Jaworski et al. (1997) Rushton turbine T = 0.15 m, D/T = 0.33, C/T = 0.33, H = T 20 × 28 × 40, Re = 17 000 SM k − ε and RNG k − ε model
Ng et al. (1998) Rushton turbine T = 0.1 m, H = T, D/T = 0.33, C/T = 0.33 46 016, 239 468 SM k − ε model
N = 2165 rpm

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING


Brucato et al. (1998) Rushton turbine C/T = 0.5 and 0.33 55 × 24 × 70, Re 32 000, Black box approach, k − ε model

|
67 000 IO, SM
(Continued )

35
|
|
36
|
Table 4. (Continued )

Refs. Impeller Geometric details Grid Impeller model RANS model

Jenne and Reuss (1999) Rushton turbine T = 0.444 m, H = T, C/T = 0.5, D/T = 0.33 Black box approach k − ε model, Chen–Kim
k − ε RNG-k − ε model
and optimised
Chen–Kim k − ε model
Derksen and van den Akker Rushton turbine D/T = 0.33, C/T = 0.33 6 00 000, Re = 29 000 Adaptive force-field
(1999) technique
Bartels et al. (2000) Rushton turbine T = H = 0.152 m, D/T = 0.33, C/T = 0.33, MRF DNS and k − ε model
Re = 7275
Ranade et al. (2001) Rushton turbine 80 × 83 × 95 Snap-shot approach k − ε model
Lane et al. (2000) Rushton turbine Geometry: same as that of Luo et al. (1994) SM and MRF

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING


48 × 39 × 60 Re = 48 000, k − ε model

|
N = 300 rpm
Montante et al. (2001) Rushton Turbine T = 0.29, H = T, D = T/3, C/T = 0:12; 0:15; 80 136, RPM = 250, SM and inner–outer Std k − ε model, RNG k − ε
0:2; 0:25; and 0.33. CFX software Re = 40 000 model, RSM
Jones et al. (2001) Paddle impeller H = T = 0.1 m, D = T/4, Case 1: C = T/2, 94 500–338 000 for 1/8th MRF k − ε model, modified k − ε
Case 2: C = T/3, FLUENT region, RPM = 100, model, RNG k − ε
Re = 3273 model, k − ω model, and
modified k − ω model
Ranade et al. (2002) Rushton turbine T = H = 0.15 m, C = T/3, D = T/3, FLUENT 269 667 for PBT (for 1/4th Computational k − ε model
region), 630 800 for snapshot
Rushton (for half region)
Kukuková et al. (2005) Rushton turbine H = T = 0.29 m, D = T/3, C = T/3, FLUENT 586 680 for Rushton MRF k − ε model
Deglon and Meyer (2006) Rushton turbine H = T = 0.15 m, D = T/3, C = T/3, FLUENT 33 000–1 900 000, MRF k − ε model
Re = 40 000
Guha et al. (2006) Rushton turbine H = T = 0.2 m, D = T/3, C = T/3, FLUENT 120–720 compartments, MRF k − ε model
150–350 RPM
Javed et al. (2006) Rushton turbine T = 0.3 and 0.15, H = T, D = T/3, C = T/3, 112 480, 138 348, SM k − ε model
FLUENT software Re = 24 000, 48 000
Ochieng et al. (2008) Rushton turbine T = 0.38, D = 0.33T, H = T, C = 0.33 T and 216 000, 436 000, and MRF and SM k − ε model, LES
0.15T 700 000, RPM: 300
Alopaeus et al. (2009) Total 17 cases. Majority T/H = 1, C = T/3, D = T/3, volume of 36 000–800 000, power MRF k − ε model, SST k − ω
with Rushton liquid = 14 dm3 , 194 dm3 and input = 1.49 W/kg to model

|
Turbine, last two 20 000 dm3 , CFX software 15.93 W/kg
with Combijet and
Phasejet

VOLUME 89, FEBRUARY 2011


|
form. The predicted flow field was used to study the effect of less importance as regards to their usefulness in improving the
mixing on the yield of a typical consecutive reaction. Good flow predictions as long as the proper boundary conditions are
comparison was sought (presented in tabular form) for the con- provided. The radial velocity and the turbulent kinetic energy
centrations. were well-predicted (maximum deviations < 20%) in the impeller
Placek et al. (1986) have used the kp − kT − ε multiple-scale discharge stream. Local velocity in the bulk was not compared.
model to model the turbulence and the black box approach Average energy dissipation in the bulk was found to be equal to the
for simulating the action of the impeller. The flow pattern was dissipation in the impeller stream (=50% each). They have pro-
obtained by solving the transport equations for vorticity, stream posed the need to accurately specify the turbulence characteristics
function and tangential momentum. The Reynolds stresses were near impeller.
modelled by means of effective viscosity based on the three- Luo et al. (1993) have solved the time-dependent equations in
equation model of turbulence. Contour plots for k, ω, ε and the two domains that were fixed with respect to their frame of ref-
steam function were presented. Detailed quantitative comparison erence (MRF) using the standard k − ε turbulence model. Finite
of the local velocity and turbulent characteristics was not given at volume mesh at the interface was allowed to slide. Better agree-
various locations in the tank. Comparison was presented near the ment as compared to the steady state and a body force impeller
tank wall only. Comparison of the mean velocity was presented model was obtained. Comparison for only tangential and axial
at three different angular locations. velocity was provided. Most of the region was above the impeller
Pericleous and Patel (1987) carried out the simulation of the and was excluded for the comparison purpose. Also tangential
flow using source-sink approach. For straight blade turbine, the velocity was underpredicted (by as much as 500%) for most of
predicted radial jet and the tangential velocity at the impeller the radial region while maximum deviation in axial velocity was
centre plane was found to agree well with the experimental mea- found to be around 20%. No comparison for the turbulent char-
surements. For radial flow impellers, the radial velocity agreed acteristics was presented.
within 10–20%, but the prediction of axial velocity was found to Perng and Murthy (1994) have illustrated the moving-
be off sometimes by as much as a factor of 2. Whereas, for axial deforming grid technique to simulate the action of the rotating
flow impellers, the axial velocity was found to be in good agree- impeller. They have used the same to perform 2- and 3-D unsteady
ment and radial velocity was found to be in poor agreement. The simulations of laminar and turbulent flows in baffled mixing tanks
streamline plots were given to see the flow field qualitatively. The with a Rushton turbine. Only vector plots of the velocity field were
turbulence characteristics were not compared. presented. No quantitative comparison for the mean velocities, the
Hutchings et al. (1989) have performed the simulation of the turbulent kinetic energy and its dissipation rate was presented.
flow field using MRF. They have treated the baffles as the localised Dong et al. (1994) have simulated stirred tank flow using MRF
sources of angular momentum and axisymmetric sink of circum- and k − ε model. Flow away from the impeller region was well
ferential momentum at the tank walls. It was, however, applicable predicted while the magnitudes of axial and radial velocities
to only simple tank geometry with negligible impeller–baffle were overestimated in near impeller region. The turbulent kinetic
interaction. energy was overpredicted in the impeller stream. It was fairly
Ju et al. (1990) have proposed non-isotropic k − ε model to predicted well in upper region of tank. Good comparison was
model the turbulence in the stirred tanks. Black box approach observed between the experimental and the predicted contour
was used to simulate the action of rotating impeller. Overall good plots for velocity vector. Predicted energy dissipation rate com-
comparison was observed with non-isotropic model in the region parison was not presented and attempt was not made to evaluate
of comparison. No comparison for turbulent characteristics was power number based on the predicted ε field.
presented. Richardson number was included for the inclusion Luo et al. (1994) have used MRF approach to model the impeller
of the effect of curvature. They have also given the comparison action and have illustrated the predictive ability of MRF simula-
of the radial velocity in angular direction that was closer with tion. They have observed that the interface conditions between
the non-isotropic model. Radial and tangential velocities were the inner and the outer domain are unsteady and there exist
found to be in better agreement (deviations less than 20% and the particular location where for all practical purposes steady
10%, respectively) in impeller plane as compared to that given flow conditions can be assumed. Interface coupling was per-
by isotropic turbulence model. Axial and radial velocities were formed by the velocity transformation. Good comparison of the
underpredicted by as much as 50% and 35%, respectively, in most radial (maximum deviation < 20%) and the axial velocity (maxi-
of the radial region (x1 /R > 0.5) while the tangential velocity was mum deviation < 15%) was obtained while poor comparison for
overpredicted by maximum of 10% by non-isotropic model. tangential velocity (with deviations as much as 700% and qual-
Ranade and Joshi (1990b) have simulated the flow produced by itatively opposite in most of the radial region) was sought. No
the radial flow impeller using black box approach in conjunction comparison was presented for the turbulence characteristics.
with k − ε turbulence model. They have systematically investi- Brucato et al. (1994) have used inner outer approach along with
gated the sensitivity of the IBCs and the turbulence parameters the standard k − ε turbulence model. Only vector plots showing
of k − ε model on the predictions. Axial velocity predictions were the flow field were presented. Comparison of local velocity with
found to be in good agreement with their in-house measurements. the experimental values was not given. Variation of power num-
Good comparison of radial, tangential velocities, and the turbulent ber obtained from the energy dissipation rate and pressure with
kinetic energy in the impeller centre plane was observed. While radial velocity and turbulent kinetic energy was shown. They have
poor comparison was sought in rest of the region. proposed that quantitative agreement can be further improved by
Kresta and Wood (1991) have shown the importance of 3-D the use of more advanced turbulence models. Fairly good predic-
simulations using the black box approach. They have extended tions were obtained for power (maximum deviation 20%) and
Kolar’s swirling radial jet model of impeller region by using two- flow number (maximum deviation 15%).
equation k − ε model to obtain direct estimate of the turbulent Tabor et al. (1998) have made use MRF approach and the
kinetic energy and the energy dissipation rate on impeller bound- SM approach. The standard k − ε turbulence model was used to
ary. They have concluded that the modified k − ε models are of model the turbulence. MRF technique was not able to capture

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 37 |


the transient details of the flow near the impeller such as 100%. Good power number and flow number comparison was
periodic unsteadiness that could be well predicted by the SM observed. Also the deviations in the energy dissipation rate were
technique. But the overall vortex pattern could be captured at the as much as 10–100%. Power dissipation was underpredicted by
less computational expenses No comparison was presented with 20%. Energy dissipation rate was given only for the centreline and
the experimental turbulence characteristics. High-resolution MRF was found to be poor. Near the bottom of the vessel, axial turbu-
gave better comparison for the velocity components near the blade lence intensity predictions showed poor comparison. Comparison
edges as compared to that obtained with SM approach (that were for the radial turbulence intensity was poor and got worst near
overpredicted by SM) Comparison was presented for the locations the bottom. Peak in the energy dissipation rate was not simulated
at x1 /R = 0.37 and 0.78. The radial and tangential components of well. Angular variation of dimensionless radial and axial veloc-
the velocity near the tank edges were showed deviation as much ity was given at an axial location (z = BW/4), but no comparison
as 30% while good agreement for other two velocity components with the experimental data was reported.
(deviation <10%) was sought. Radial and tangential velocities Jenne and Reuss (1997) have modelled the flow with the
were closely predicted by high-resolution MRF while axial veloc- k − ε turbulence model and used black box approach to describe
ity showed deviations as much as 100% near the impeller. They impeller rotation. They have proposed the 2-D model (that
have found that the MRF was better as compared to SM in pre- assumes the variation of the liquid distribution independent of
dicting the behaviour of impeller stream and enlisted the possible the angle), which can be used to calculate power, flow numbers
causes. They have recommended MRF for low D/T ratio and sug- and impeller efficiency. Model depends on the number of paddles.
gested its inapplicability for the small blade-baffle gap and the They have concluded that the energy required to sustain the liquid
cases where the time dependent features are of importance. motion in the mixer was strictly related to the velocity distribu-
Lee et al. (1996) have used SM technique to simulate the tion in the vessel itself. Their model reasonably represented the
impeller rotation. They have discussed the capabilities and primary and secondary circulation.
the limitations of the SM technique. Contour plots of both Kuncewicz et al. (1997) have proposed the two-dimensional
the experimental and the predicted turbulent kinetic energy model for the prediction of flow pattern in stirred vessel. Tangen-
profiles were presented. Severe underpredictions (maximum devi- tial velocity prediction was very poor (maximum deviation 100%)
ation = 200%) were observed in case of the turbulent kinetic while axial velocity agreed reasonably well (∼25). Liquid velocity
energy predictions. No comparison for mean axial and tangential was found to be proportional to power drawn.
velocities and the turbulent energy dissipation rate was presented. Venneker and van den Akker (1997) have attempted the simu-
No comparison in the upper part of the vessel was reported. lation of the flow using the black box approach along with the
Instantaneous radial velocity was overpredicted in most of the k − ε turbulence model. Radial velocities agreed well through-
region (x1 /R > 0.5) with maximum departure of 35% from exper- out the tank while axial velocity predictions were satisfactory in
imental data. most of the region (underpredicted by as much as 50% below
Ciofalo et al. (1996) have used rotating frame of reference along the impeller). The turbulent kinetic energy predictions were good
with the k − ε turbulence model and the second-order differential except near the shaft and the bottom of the impeller where they
stress model (DSM). Poor axial velocity predictions were observed were severely underpredicted. No attempt was made to validate
near the vessel centre. DSM gave good tangential velocity predic- the predicted energy dissipation rate. They have concluded that
tions in the upper region. Satisfactory comparison was obtained the difference between the experimental data and the predictions
for the radial velocity. The shape of the free surface and the due to inner–outer technique may be due to the overestimation
tangential velocity were in excellent agreement with the exper- of the eddy viscosity and anisotropy of the turbulent flow in the
imental results. The predicted power numbers were significantly stirred vessels. They also have suggested that the qualitatively cor-
smaller than the experimental values. They were the first one to rect solution can be obtained by standard k − ε model provided the
use the second-order turbulence closure wherein they have incor- experimental data at the impeller tip were used as the boundary
porated the effect of the Coriolis force. DSM gave good power conditions.
and flow number comparison. They have found that only DSM Jaworski et al. (1997) have carried out CFD simulation using
could capture the complex interrelation between the main and the SM approach along with the standard k − ε and RNG k − ε
the secondary flow field, turbulence stresses and the inertial forces turbulence models. They have found no effect of the turbulence
(rotation induced) in the baffled vessel. For Rushton turbine pre- model in the bulk region. Both the models used, underpredicted
dicted power number (15–30%) and the flow number were found the turbulence levels. Overall prediction of flow number was good.
to be in good agreement with the experimental data. Comparison of all the three velocity components and their fluc-
Rigby et al. (1997) have employed SM technique to simulate tuations near the impeller tip was presented as a function of
the action of rotating impeller. Comparison with the experimen- angular co-ordinate only. No spatial comparison was presented.
tal velocities was shown in tangential direction only. Reasonable Mean velocity predictions were in good comparison while both the
agreement with the experimental data for the location of impeller models failed to capture the trailing vortex structure. Turbulence
was observed. Periodicity (from stationary frame of reference) was was significantly underpredicted. Standard k − ε model yielded
well predicted. No comparison was sought for the validation of the smaller errors as compared to the RNG k − ε model for turbu-
turbulent flow characteristics. Comparison of all the three veloc- lence level. RMS components showed significant departure from
ity components showed slight underpredictions in the impeller the isotropy in the impeller discharge stream. CFD with any of the
centre plane and slightly above the impeller. models did not pick up trailing vortex region. Better prediction of
Ranade (1997) has used the snap shot approach to simulate turbulence with the standard k − ε model. Good comparison for
the action of rotating impeller and used the standard k − ε model the power (within 20%) and flow number was obtained.
of turbulence. The radial and the axial velocities were under- Ng et al. (1998) have made a use of sliding grid technique along
predicted. Also the radial velocity was predicted qualitatively with the standard k − ε model to predict the flow pattern in stirred
opposite in the vessel centre region. The turbulence intensities tank by radial flow impeller. They have discussed the effect of var-
in radial and axial direction were under predicted by as much as ious parameters such as the angle broadening, data rate and the

| 38 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


cyclic variation on the measured velocities. They have suggested Wechsler et al. (1999) have carried out simulation of the tur-
further improvement in turbulence modelling for the good pre- bulent flow induced by pitched blade turbine using steady and
diction of the turbulent kinetic energy in near impeller region. unsteady state approach. In steady state approach relative motion
Comparison of the radial velocity, kinetic energy, and the energy between the impeller blades and the fluid was neglected while
dissipation rate in the impeller stream at four different angular that was taken into consideration in case of unsteady state simu-
locations (at angular locations of 2 = 0◦ , 15◦ , 30◦ , and 45◦ ) were lations (modified form of the SM approach). They have presented
presented. Radial velocity was found to be overpredicted by devia- extensive comparison with experimental data for all the three
tion amounting to as much as 35% while severe underpredictions components of velocity and the turbulent kinetic energy at dif-
(as much as 200%) were observed for the turbulent kinetic energy ferent axial locations. The region near the liquid surface was not
and its dissipation rate. Poor predictions of the energy dissipation considered for comparison purpose. Radial and tangential velocity
rate (maximum deviation of 400%) were observed. predictions were good (∼10–15%) while turbulent kinetic energy
Brucato et al. (1998) have carried out CFD simulations of the was underpredicted by as much as It can be seen that axial velocity
flow in stirred tanks using various impeller models, namely black was off from the experimental data by as much as 50%.
box approach, inner–outer technique and sliding grid approach Oshinowo et al. (2000) performed the CFD study using different
along with the standard k − ε turbulence model. They have turbulence models like, k − ε, RNG k − ε and RSM for the predic-
presented a comparison with alternative modelling approaches. tion of tangential velocity distribution in a baffled vessel using
Comparison of the experimental and the predicted velocity vector MRF technique to model the impeller–baffle interaction. The tan-
plots for axial flow impeller was given which could be seen as gential velocity distribution above the impeller has been correctly
qualitatively good. They have shown that the predictions given predicted. They attributed the occurrence of counter-intuitive
by inner–outer approach were comparable to those given by the reverse swirl in the simulation results to poor convergence and
SM approach Predictions of the axial velocity were found to be the coarse grid density in their simulations.
in good agreement with SM while those were underpredicted by Bartels et al. (2000) have carried out the simulation of the flow
black box approach. pattern produced by Rushton turbine at Reynolds number of 7275
Jenne and Reuss (1999) have presented a critical assessment by both the standard k − ε model and the DNS using MRF. They
of the variants of the standard k − ε model and recommended have presented the velocity vector plots and the contour plots of
the optimised version of the Chen–Kim model for better predic- the turbulent kinetic energy. They have observed that the shape
tions. The comparison was presented only in the region below of the contour plots of the turbulent kinetic energy in the impeller
the impeller. Good agreement for all the velocity components plane as a result of DNS matches closely to the one experimen-
as a result of simulation with optimised k − ε model (deviation tally observed as compared to the standard k − ε model. Also the
less than 10%) was obtained. No comparison for the turbulence secondary vortices near the junction of the hub and the disc are
characteristics was presented. They have concluded that a more shown by DNS while standard k − ε model failed to show this flow
complex consideration of the different time scales of turbulence feature. Both the simulations were able to predict the upward incli-
(three equation or four equation eddy viscosity models) or the nation of the flow in impeller discharge region and presence of the
effects of anisotropy (RSM) are of minor importance. They also ring vortices. They have concluded that both the simulations show
concluded that the spectral local non-equilibrium is major while good comparison of the observed flow features as regards to the
anisotropy is of minor consideration. large-scale flow field while DNS performs better as compared to
Micale et al. (1999) have carried out CFD simulation of the flow the standard k − ε model when the close comparison is consid-
pattern using black box, SM, and inner–outer technique along ered. We can conclude that the standard k − ε model is capable
with the standard k − ε turbulence model. They have simulated of predicting most of the flow features of the flow prevailing in
the three cases, namely parallel flow, merging flow, and diverging the stirred vessel at very much lower computational expenses as
flow. Though the good power number comparison was observed compared to the DNS.
as a result of SM, 50% less by black box approach and 20% less Ranade et al. (2001) have carried out the simulation of the flow
by inner–outer technique (IO), it was not calculated by the inte- pattern produced by the Rushton turbine by using a snap shot
grating the energy dissipation rate throughout the tank. It was approach. They have used the standard k − ε model to model the
obtained as the product of the torque and the angular velocity. turbulence. They have presented the comparison of the flow char-
The torque was calculated by summing the contributions due to acteristics, namely in impeller region. They have observed very
the pressure drop between the upstream and downstream faces of good comparison of the flow features near the impeller. Snap shot
the baffles and the wall shear stress on the horizontal and lateral approach was found to predict a pair of trailing vortices. Closer
walls. Comparison was presented for the predicted flow number quantitative agreement (∼10% max) was obtained for the radial
NQ . Similar results with IO and SM and significantly different with velocity (∼150% max) while axial and tangential velocity (∼50%
black box approach. Closer comparison for the predicted radial max) predictions were found to be far away from the experimental
velocity was sought for parallel flow while comparison for merg- data while the turbulent kinetic energy was severely underpre-
ing and diverging flows was not good. They have observed the dicted (by as much as ∼200%) with the experimental data in the
presence of strong axial component (predicted) by SM as com- impeller region.
pared to IO. Good qualitative agreement between the presented Montante et al. (2001) carried out simulations to study the
experimental and the predicted contours of velocity vectors. Tur- effect of impeller clearance on the prediction of flow patterns.
bulent kinetic energy was severely underpredicted by both the They varied the clearance ratio (C/T) from 0.12 to 0.33. They
SM and IO. Also the applicability of black box approach was rec- used three different turbulence models (i) standard k − ε model,
ommended for if the flow is insensitive to the details of the flow (ii) RNG k − ε model, and (iii) RSM. Impeller–baffle interaction
at impeller boundaries and accurate flow data is available and was modelled using SM and inner–outer approach. They have
inapplicable to multiple impeller system. They have observed that reported the velocity vector plots to show the clearance effect
the black box approach fails to predict the power number and its on number of circulation loops. The velocity profiles from three
dependence on the impeller-vessel configuration. different turbulence models were almost similar and due to the

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 39 |


lack of experimental data in the plots, conclusion regarding the al. (2000) performed the CFD study using different turbulence
superiority of the model cannot be made. models like, k − ε, RNG k − ε and RSM for the prediction of
Kukuková et al. (2005) carried out mixing studies like other tangential velocity distribution in a baffled vessel using MRF
investigators. In addition to velocity profile, concentration profile model. The tangential velocity distribution above the impeller
and mixing time, they have also reported power and pumping has been correctly predicted. They attributed the occurrence of
number of the impeller. The simulation results of axial, radial, counter-intuitive reverse swirl in the simulation results to poor
and tangential velocities were in good agreement with experi- convergence and the coarse grid density in their simulations. It
ments. The simulated mixing time values were underpredicted may be pointed out, however, that both the aforesaid investiga-
by 20–45% as compared with experiments. tions have shown the comparisons in a small region and most of
Guha et al. (2006) carried out compartmental modelling to the vessel region remained unexplored. More importantly, both
study turbulent mixing. They have carried out simulation of reac- the studied have not presented the predictive capability of CFD
tive system (first- and second-order system) as well as effect of models for the turbulent kinetic energy and the turbulent energy
turbulent mixing on multiple reactions. They have also studied dissipation rate.
the effect of feed location on the desired product yield. From the foregoing literature review it follows that, in majority
Javed et al. (2006) carried out mixing studies using CFD as well of the studies, the comparison between the CFD predictions and
as experiments. They have solved scalar transport equation to esti- the experimental measurements has been made over a relatively
mate mixing time. Turbulent Schmidt number was chosen as 0.7. small region around the impeller or the impeller stream. Very few
They considered eight different locations in bulk and impeller predictions have been validated against the experimental data for
region as concentration measurement locations. The comparison all the flow variables and turbulence characteristics throughout
of mean axial, tangential, and radial velocities was reported and the vessel. Knowledge of the turbulent kinetic energy and its dis-
simulation results were in very good agreement with the experi- sipation rate is essential for the understanding of transport of the
mental data. However, simulation under predicted the turbulent flow variables. Also these are required in prediction of micromix-
kinetic energy. They have also reported the transient evolution of ing, turbulent length scales, shear stresses, and drop/bubble sizes.
tracer concentration. They observed that (i) the predicted tempo- Very few attempts have been made to establish the energy balance
ral and spatial distributions of tracer concentration reveal a direct from the predicted values of the energy dissipation rate ε through-
correspondence with the mean flow structure and levels of turbu- out the tank. The total turbulent energy dissipation in the vessel
lence in the tank. (ii) The time required for the first concentration has been rarely compared with the energy-input rate through the
peak to appear and its value, and the time required for concentra- impeller (estimated using the power number of the impeller).
tion profiles to approach the fully mixed condition depend on the Comparison of the predicted power number (which is the measure
position in the tank. the energy balance in the stirred tanks) is not reported in most
Ochieng et al. (2008) carried out mixing studies in stirred tank of the studies. Severe under/overprediction of turbulent kinetic
at low clearance. They have shown that at a low impeller clear- energy for all the simulation approaches and the turbulence mod-
ance, the Rushton turbine generates a flow field that evolves from els could be observed. Also the dissipation of turbulent kinetic
the typical two loops to a single loop flow pattern similar to that energy (hence the eddy viscosity) has received little attention
of an axial impeller. This single loop flow pattern resulted in an as regards to its prediction and validation. This may be partly
increase in axial flow and a decrease in mixing time at a constant because of the difficulty in experimental measurements of these
power number. They have also carried out experiments and simu- parameters and the consequent difficulty in reliable validation
lation by adding a draft tube in the tank to improve axial velocity with the experimental data. Further, potentially strong alterna-
and were able to achieve significant improvement in mixing per- tives (SM approach and MRF) suffer from the requirement of
formance. The simulation results of axial and radial velocity were expensive computation. Most of the simulations have been carried
in good agreement with experiments whereas turbulent kinetic out with two-equation standard k − ε model. This model contains
energy was underpredicted using simulations. three parameters (C␮ , Cε1 , Cε2 ). The values of these parameters
Alopaeus et al. (2009) carried out CFD simulation using zonal have been estimated on the basis of simple flows (boundary layer,
modelling. The model is based on tank division into two nested jets, flow downstream of a grid). The validity of using the values
regions in a systematic manner. By gradually increasing the of these parameters for the simulation of flow in stirred vessel is
inner zone volume, starting from the impeller swept volume and questionable. Practically no attempt has been made to understand
finally approaching the total tank volume, continuous curves are the significance of these parameters and their values in a typical
obtained for the most important factors characterising small- and stirred vessel configuration. In addition to the works of Sahu et al.
large-scale mixing. They have analysed two important variables (1998); and Nere et al. (2001), it is very important to understand
(i) pumping number and (ii) turbulent kinetic energy dissipation the validity of k − ε model for stirred tank reactors.
rate. They have simulated 17 cases varying vessel size, discretisa-
tion methods, grid size, turbulence model (namely standard k − ε
model and SST k − ω model) and type of impeller. COMPARISON OF IMPELLER–BAFFLE
There have been hardly two–three attempts partially focused
on the RSM for the flow generated by Rushton turbine in a baf- INTERACTION MODELS
fled stirred vessel. First one, Bakker and van den Akker (1994) Impeller–baffle interaction modelling plays a crucial role in pre-
employed the simplified RSM, that is, ASM using the IBC method diction of flow patterns in stirred vessels. The summary of
to model the flows produced by a Rushton turbine. Their objec- various models has already been provided in third section. In
tive was to improve the predictive capabilities of CFD modelling this section, comparison of these models has been carried out
by accounting anisotropy using less computational intensive ASM at four locations (x2 /R = 0.586, x2 /R = 0.24, x2 /R = −0.24, and
model (simplified RSM). Their study concluded that the results x2 /R = −0.933) (Figure 4). This analysis is presented using two
predicted by the ASM compare better with the experimental data different turbulence models, namely (i) standard k − ε model (ii)
than those predicted by the standard k − ε model. Oshinowo et RSM. An extensive comparison of the CFD predictions has been

| 40 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


ferent zones to improve the predictions. In case of RSM, all the
stress equations are solved to get more accurate turbulence pro-
duction. But in RSM, the modelled equation for turbulent kinetic
dissipation rate equation is same as that of standard k − ε model
(many assumptions to model terms containing fluctuating com-
ponents) and hence the deviations with experimental data can
be expected. If we have good grid resolution and all the shear
rates are accurately predicted in impeller as well as bulk zone,
then good agreement with RSM can be expected. The fifth and
sixth sections give the comparison between CFD predictions and
experimental data.

Comparison at x2 /R = 0.586
Figure 5A shows the comparison of radial velocity using standard
k − ε model and various impeller–baffle interaction models. It can
be seen that SNAP model (line 4, 28%) gave the best fit amongst
all the models. IBC (line 1, 44%), MRF (line 2, 52%), and SM (line
3, 88%) underpredicted the radial velocity. Even the axial velocity
predictions (Figure 5B) were better with SNAP model (51%), fol-
lowed by IBC (73%), MRF (79%), and SM (95%) models. None of
the models could predict tangential velocity (Figure 5C). The pre-
dictions of radial velocity with RSM as turbulence model and IBC
Figure 4. Schematic of stirred tank reactor with locations at which and MRF as impeller–baffle interaction model are given in (Figure
comparison between experiment and CFD data has been carried out.
6A). MRF (52%) gave almost similar results with RSM and stan-
dard k − ε model, while IBC (70%) gave more deviation with RSM.
illustrated with the experimental data obtained with the help of Even in case of axial velocity (Figure 6B), though both the models
in-house LDA. The values of error are reported in brackets in the under predicted, MRF (76%) with RSM gave better results than
analysis. Table 5 gives the quantitative error involved with all the standard k − ε model. IBC (81%) with RSM gave poor results as
turbulence models for mean and turbulent quantities. The error compared with standard k − ε model. Figure 6C shows the tangen-
is estimated as follows: tial velocity predictions of IBC and MRF with IBC. Both the models
(>300%) showed only qualitative agreement with experimental
Expt Value − Predicted Value results with RSM.
Error = × 100 (11)
Expt Value Figure 7A shows the prediction of turbulent kinetic energy.
Though IBC gave a peak at x1 /R = 0.1, which was not shown by
experiments, it gave satisfactory match (45%) in the rest of the
Simulation Details region. MRF (92%), SM (76%), and SNAP (65%) underpredicted.
Turbulent kinetic energy with RSM (Figure 7B) gave satisfactory
The geometrical details of the system investigated are already
results with IBC (45%), while underpredictions were observed
shown in Figure 4. The impeller speed was kept at 4.5 rps
with MRF (92%).
(Re = 45 000) to ensure the turbulent flow regime. The impeller
was placed at H/3 from the vessel bottom. All the simulations
have been carried out by using an in-house and the commercial Comparison at x2 /R = 0.24
code FLUENT. Pressure–velocity coupling has been taken care of Figure 8A shows the comparison of radial velocity using different
by using SIMPLE scheme. Grid independence studies have been impeller–baffle interaction models using standard k − ε turbu-
carried out using three different grid sizes, namely 45, 68, and lence model. It can be seen that IBC model (line 1, 835%) shows
85 K. Finally 85 K grid size has been chosen for all the simula- better agreement compared to other models (MRF: line 2, 246%,
tions. Non-uniform grids are used to maintain higher grid density SM: line 3, 1005%, SNAP: line 4). However, maximum value of
in high shear regions. velocity with experiment was observed close to x1 /R = 0.66 and
In some cases, visual observation of plots showed good agree- with IBC at x1 /R = 0.42. SM and SNAP approach do not predict
ment, whereas RMS error was high. This was because agreement negative radial velocity, whereas MRF approach shows very poor
was good over major portion; only at some locations deviations agreement with experimental data. Figure 8B shows comparison
were high making the RMS error values large. Other possible rea- of axial velocity. It can be seen that SNAP shows very good agree-
sons for relatively large deviations are (i) number of grid points ment with experimental data over the entire range. IBC predictions
for all the simulations have been decided after grid sensitivity (103%) were in good agreement up to x1 /R = 0.4, beyond which
study. In this grid sensitivity study, comparison of mean veloc- it was insignificant under prediction. SM (51%) and MRF (118%)
ity components with experimental data using IBC approach and showed under prediction over entire range and agreement was
standard k − ε model was carried out. Hence possibility of error very poor. Figure 8C shows comparison of tangential velocity. It
at this resolution for different impeller models and turbulence can be seen that no model is able to predict tangential velocity
models cannot be ruled out. (ii) The turbulence in stirred vessels neither in qualitative nor quantitative sense. The source of error
is strongly anisotropic and not homogeneous. The standard k − ε in prediction can be either from inappropriate impeller model or
model and its modifications cannot account for such anisotropy turbulence model. To check the sensitivity of turbulence model on
due to inherent assumptions involved in these models. At best, overall prediction, similar comparison of impeller models (MRF
one can use different combinations of model parameters in dif- and IBC) is presented using RSM. Figure 9A shows comparison of

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 41 |


Table 5. Quantitative error

x2 /R = 0.586 x2 /R = 0.24 x2 /R = −0.24 x2 /R = −0.933

Axial location Max error RMS error Max error RMS error Max error RMS error Max error RMS error

Impeller model: IBC


STD k − ε
Radial velocity 69 44 3164 835 13 283 2916 4037 1184
Axial velocity 158 73 301 103 990 238 150 84
Tangential velocity 4896 1178 331 129 2265 563
TKE 127 45 637 398 495 353 68 49
RNG k − ε
Radial velocity 71 47 2602 765 13 557 2968 1365 386
Axial velocity 80 66 275 93 1016 244 131 70
Tangential velocity 3883 896 462 159 2693 661
TKE 84 57 474 279 367 239 82 68
RSM
Radial velocity 87 70 10 496 2442 17 265 3780 444 153
Axial velocity 167 81 239 91 971 233 236 114
Tangential velocity 19 536 4718 361 120 1432 380
TKE 159 92 756 542 1258 743 84 58
MOD k − ε
Radial velocity 92 53 2563 574 623 226 179 105
Axial velocity 113 59 239 83 569 157 137 40
Tangential velocity 6849 2650 902 285 4542 1509
TKE 64 30 249 192 127 92 75 62
ZON k − ε
Radial velocity 80 56 1400 378 900 305 1468 391
Axial velocity 67 57 256 91 536 146 285 95
Tangential velocity 5342 2067 621 199 7921 2547
TKE 83 70 380 274 113 61 76 70
NEW k − ε
Radial velocity 69 51 2116 578 819 272 1326 419
Axial velocity 157 66 282 97 460 124 224 79
Tangential velocity 1751 682 554 179 10 769 3470
TKE 90 70 168 114 48 18 83 79
OPTCK k − ε
Radial velocity 69 51 1770 563 781 365 1326 419
Axial velocity 157 66 159 60 151 40 224 79
Tangential velocity 1628 536 575 192 22 706 7414
TKE 96 88 83 48 89 63 94 91
Impeller model: MRF
STD k − ε
Radial velocity 85 52 826 246 383 179 2867 974
Axial velocity 153 79 61 51 177 62 355 108
Tangential velocity 1246 312 175 81 1833 449 3649 1364
TKE 97 92 95 90 93 84 87 82
RSM
Radial velocity 81 52 2882 638 606 226 3463 1187
Axial velocity 162 76 75 53 58 45 393 114
Tangential velocity 3571 818 226 89 3153 772 1854 560
TKE 96 91 94 89 94 86 87 81
Impeller model: SM
STD k − ε
Radial velocity 93 88 3487 1005 1909 781 1134 358
Axial velocity 121 95 399 118 624 170 100 97
Tangential velocity 100 96 9171 3322 200 106
TKE 80 76 91 88 93 83 89 84
RNG k − ε
Radial velocity 97 91 3141 865 2286 780 724 238
Axial velocity 188 94 460 144 114 99
Tangential velocity 104 85 5033 1722 468 165
TKE 99 88 291 135 100 99

| 42 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 5. Prediction of mean velocity components using standard k − ε
model at x2 /R = 0.586. Impeller: disc turbine. (A) Radial velocity (B) axial
velocity (C) tangential velocity. Line 1: IBC, line 2: MRF, line 3: SM, line 4: Figure 6. Prediction of mean velocity components using RSM at
SNAP. x2 /R = 0.586. Impeller: disc turbine. (A) Radial velocity (B) axial velocity
(C) tangential velocity.  Experimental, line 1: IBC, line 2: MRF.
radial velocity and it can be seen that using IBC (line 1, 2442%)
and RSM together show significant overprediction than IBC and with the experimental data. MRF could predict the crossover of
standard k − ε model (Figure 5A). A significant improvement is velocity from positive to negative accurately. Figure 9C shows
observed in the wall region when MRF is used with RSM (638%), comparison of tangential velocity. The predictions have signifi-
whereas in the bulk effect of turbulence model was insignificant. cantly improved in central region using RSM than standard k − ε
Figure 9B shows comparison of axial velocity. It can be seen that model for both IBC (120%) and MRF (89%).
axial velocity profiles are relatively insensitive to turbulence mod- Figure 10A and B shows the comparison of turbulent kinetic
els. MRF model predictions (53%) were in qualitative agreement energy using standard k − ε model and RSM, respectively. It can

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 43 |


Figure 7. CFD prediction of turbulent kinetic energy for disc turbine at
x2 /R = 0.586. (A) Standard k − ε model:  experimental, line 1: IBC, line
2: MRF, line 3: SM, line 4: SNAP. (B) RSM:  experimental, line 1: IBC,
line 2: MRF.

be seen that predictions from all the impeller models for turbulent
kinetic energy were in poor agreement by both the turbulence
models. IBC model (with k − ε model: 398% and with RSM:
542%) overpredicted, whereas other models severely under pre-
dicted (factor of 10).

Comparison at x2 /R = −0.24
The comparison of mean and turbulence quantities in below
impeller plane has been presented in Figures 11–13. Figure 11A–C
shows comparison of mean flow using standard k − ε model. Simi-
lar to x2 /R = 0.24, IBC (2916%) shows only qualitative agreement
with the experimental data below impeller for radial velocity Figure 8. Prediction of mean velocity components using standard k − ε
(Figure 11A). SNAP and MRF have totally different trend than model at x2 /R = 0.24. Impeller: disc turbine. (A) Radial Velocity (B) axial
experimental data, whereas SM has shown significant under pre- velocity (C) tangential velocity.  Experimental, line 1: IBC, line 2: MRF,
diction (by a factor of 4). The reason behind good agreement using line 3: SM, line 4: SNAP.
IBC is it uses actual velocity field information (experimental data)
in the impeller zone, whereas in other methods, actual experi- comparison of impeller models (MRF and IBC) using RSM. From
mental data are not used for the simulation. Figure 11B shows Figure 12A, it can be seen that, a change in the turbulence model
the comparison of axial velocity. All the models show only qual- did not improve the results for MRF model, but the performance
itative match with the experimental data. SNAP approach was of IBC model with RSM (3780%) was deteriorated as compared
found to be in best agreement, followed by IBC (238%). MRF to standard k − ε model. Figure 11B shows that the effect of tur-
results were better than SM, but both the models under predict the bulence model on axial velocity (233%) is not significant and
experimental data. Figure 11C shows the comparison of tangential profiles are almost similar as in Figure 11B. The tangential velocity
velocity. IBC showed good agreement with the experimental data, (Figure 12C) predictions using IBC were in very good agree-
and other methods showed poor agreement. Figure 12 shows the ment with experimental data, whereas MRF results were in under

| 44 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 10. CFD prediction of turbulent kinetic energy at x2 /R = 0.24.
Impeller: disc turbine. (A) Standard k − ε model:  experimental, line 1:
IBC, line 2: MRF, line 3: SM, line 4: SNAP (B) RSM:  experimental, line 1:
IBC, line 2: MRF.

magnitude of the axial velocity. The tangential velocity (Figure


14C) predictions were also qualitative by all the models, SM
(106%) gave the best match amongst the models, while others
predicted (>400%). Radial velocity (Figure 15A) predictions with
RSM did not show either qualitative or quantitative match with
both the models. Figure 15B shows axial velocity comparison by
IBC and MRF models using RSM. Only qualitative agreement was
shown by MRF (114%) with RSM as turbulence model, while
IBC gave negative velocities. Neither IBC nor MRF could predict
tangential velocity (Figure 15C) with RSM (>300%).
The turbulent kinetic energy (Figure 16A) was predicted sat-
Figure 9. Prediction of mean velocity components using RSM at
x2 /R = 0.24. Impeller: disc turbine. (A) Radial velocity (B) axial velocity isfactorily by SNAP (49%), followed by IBC (49%), MRF (82%),
(C) tangential velocity.  Experimental, line 1: IBC, line 2: MRF. and SM (84%). IBC (58%) overpredicted turbulent kinetic energy,
while MRF (81%) underpredicted with RSM (Figure 16B).
prediction. Figure 13A and B shows the comparison of turbulent
kinetic energy using standard k − ε model and RSM, respectively COMPARISON OF RANS MODELS
(with k − ε model: 353% and with RSM: 743%). The observations In the previous section, detailed comparison of various
are almost similar to x2 /R = 0.24. impeller–baffle interaction models was presented. In this section,
comparison of various turbulence models (refer Table 2) used
Comparison at x2 /R = −0.933 for closure in Reynolds averaged Navier–Stokes equation is pre-
None of the models predicted radial velocity (Figure 14A) either sented. Various turbulence models, namely the standard k − ε,
quantitatively or qualitatively. Figure 14B shows the axial veloc- RNG k − ε, RSM have been used with impeller models such as
ity comparison, IBC, MRF, and SNAP models followed the trend IBC, MRF, and SM. Further, simulations with the k − ε model
qualitatively but underpredicted, while SM model gave negligible with the turbulence parameters suggested by Abujelala and Lilley

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 45 |


Figure 11. Prediction of mean velocity components using standard k − ε
model at x2 /R = −0.24. Impeller: disc turbine. (A) Radial velocity (B) axial Figure 12. Prediction of mean velocity components using RSM at
velocity (C) tangential velocity.  Experimental, line 1: IBC, line 2: MRF, x2 /R = −0.24. Impeller: disc turbine. (A) Radial velocity (B) axial velocity
line 3: SM, line 4: SNAP. (C) tangential velocity.  Experimental, line 1: IBC, line 2: MRF.

(1984) (MOD k − ε), zonal model (ZON k − ε by Sahu et al., 1998)


yielded closer comparison followed by RNG k − ε model and RSM
and optimised Chen–Kim k − ε model (OPT-CK, Jenne and Reuss,
(standard k − ε model: 44%, RNG: 47%, RSM: 70%). IBC under-
1999) model have been carried out using IBC approach.
predicted axial velocity (Figure 17A2) with all the turbulence
models and the predictions were found to be insensitive to the
Comparison Using IBC Impeller Model turbulence models employed (standard k − ε model: 73%, RNG:
66%, RSM: 81%). Relatively closer comparison was sought with
Comparison at x2 /R = 0.586 the standard and the RNG k − ε models as compared to the RSM
IBC (Figure 17A1) with all the turbulence models was able to give in the region with x1 /R < 0.1. IBC with all the models predicted
satisfactory comparison of radial velocity. Standard k − ε model tangential velocity (Figure 17A3) trend qualitatively (standard

| 46 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 13. CFD prediction of turbulent kinetic energy at x2 /R = −0.24.
Impeller: disc turbine. (A) standard k − ε model:  experimental, line 1:
IBC, line 2: MRF, line 3: SM, line 4: SNAP. (B) RSM:  Experimental, line
1: IBC, line 2: MRF.

k − ε model: 1178%, RNG: 896%, RSM: 4718%). Standard and


RNG k − ε models resulted into consistently severe underpredic-
tions. Closer comparison was sought by RSM for x1 /R < 0.3 while
overprediction was observed in rest of the radial region. The k-
predictions (Figure 17A4) by standard k − ε (45%) and RNG k − ε
(57%) models were similar quantitatively with minor underpre-
dictions up to x1 /R = 0.4. Closer comparison was sought in rest of
the radial region. RSM (92%) resulted into satisfactory compari-
son.

Comparison at x2 /R = 0.24
IBC with all the turbulence models though predicted radial veloc- Figure 14. Prediction of mean velocity components using standard k − ε
ity (Figure 17B1) trend qualitatively, the maximum was found to model at x2 /R = −0.933. Impeller: disc turbine. (A) Radial velocity (B)
axial velocity (C) tangential velocity.  Experimental, line 1: IBC, line 2:
be shifted away from the vessel wall (standard k − ε model: 835%, MRF, line 3: SM, line 4: SNAP.
RSM: 279%, RNG k − ε model: 765%). RSM overpredicted radial
velocity in the central region while satisfactory agreement can be
seen result in rest of the region. While other turbulence models see that the sensitivity of the predictions to the turbulence models
yielded relatively closer agreement of radial velocity throughout. is limited to the region with x1 /R < 0.4 where RNG k − ε and STD
Axial velocity (Figure 17B2) predictions of all the simulations k − ε models resulted into larger underprediction of tangential
were practically similar (i.e., insensitive to turbulence models). velocity. All the turbulence models largely overpredicted turbulent
All the turbulence models closely predicted the axial velocity x1 /R kinetic energy (Figure 17B4) with the highest deviation as a result
up to 0.4 while underpredictions were observed in rest of the of RSM (542%) followed by the STD (398%) and RNG (159%)
radial region (0.4 < x1 /R < 0.8). All the models resulted into poor k − ε model in most of the radial region (x1 /R > 0.4). Good agree-
predictions (120–398%) of tangential velocity (Figure 17B3). RSM ment was observed for x1 /R < 0.4 with RNG k − ε model while
gave relatively closer agreement for x1 /R up to 0.4. Thus one can STD k − ε model yielded closer agreement for x1 /R up to 0.15.

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 47 |


Figure 16. CFD prediction of turbulent kinetic energy at x2 /R = −0.933.
Impeller: disc turbine. (A) standard k − ε model:  experimental, line 1:
IBC, line 2: MRF, line 3: SM, line 4: SNAP. (B) RSM:  experimental, line
1: IBC, line 2: MRF.

k − ε (2968%) were relatively closer over most of the radial coor-


dinate (except near the vessel centre), while RSM yielded large
overprediction (3780%) in the region 0.1 < x1 /R < 0.6. Very good
agreement of the axial velocity (Figure 17C2) could be seen as a
result of simulations with IBC with all the turbulence models. All
the turbulence models yielded quantitatively similar axial velocity
predictions indicating insensitivity to the turbulence models. Very
good agreement of the tangential velocity (Figure 17C3) could be
found as a result of simulations with IBC with all the turbulence
models. Further the turbulence models resulted into quantita-
tively similar tangential velocity predictions (deviations within
120% and 159%) again leading to the insensitivity of w predic-
Figure 15. Prediction of mean velocity components using RSM at tions towards the turbulence models. Severe overpredictions in k
x2 /R = −0.933. Impeller: disc turbine. (A) Radial velocity (B) axial velocity
(C) tangential velocity.  Experimental, line 1: IBC, line 2: MRF.
(Figure 17C4) were observed as a result of all the simulations with
IBC. The magnitude in the deviation was the least with the RNG
k − ε (239%) followed by STD k − ε (353%) model. RSM resulted
Thus one can conclude that turbulence models result in different into the worst k-predictions (743%) throughout. RNG k − ε model
kinetic energies in bulk while the turbulence models do not sig- gave closer k predictions for x1 /R < 0.3.
nificantly affect the mean flow field (except tangential velocity)
in the bulk region. At x2 /R = −0.933
IBC with all the turbulence models employed resulted into the
Comparison at x2 /R = −0.24 poor radial velocity (153–1184%) predictions (Figure 17D1).
IBC resulted in good prediction of radial velocity trend (Figure No specific trend as a result of the change in the turbulence
17C1). The predictions as a result of STD k − ε (2916%) and RNG model was observed. Axial velocity trend (Figure 17D2) was well

| 48 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 17. Comparison of predictions with IBC. Impeller: disc turbine. A1–A4: at x2 /R = 0.586, B1–B4: at x2 /R = 0.24; C1–C4: at x2 /R = −0.24; D1–D4:
at x2 /R = −0.933.  Experimental (1) STD k − ε, (2) RNG k − ε, (3) RSM.

captured by all the turbulence models. IBC with STD k − ε (84%) throughout the radial region. The deviation was found to be as
and RNG k − ε (70%) resulted into relatively closer comparison much as 100%. All the three turbulence models resulted into sim-
as compared to that given by RSM (114%). With each of the simu- ilar axial velocity predictions. Its predictions were opposite for
lations severe underprediction was the result. Interestingly all the x1 /R < 0.45 and it was overpredicted for rest of the region. Tan-
models were able to predict the upflow near the vessel wall. None gential velocity was severely underpredicted by all the turbulence
of the turbulence models was able to satisfactorily predict the models. RSM gave relatively closer comparison. Further STD and
tangential velocity component (Figure 17D3) with IBC. Though RNG k − ε models resulted into quantitatively similar predictions.
the standard k − ε (563%) and RNG k − ε (661%) model resulted IBC using all the turbulence models yielded satisfactory compar-
into relatively closer comparison near the wall, they failed to cap- ison. Predictions of k as a result of all the simulations were more
ture the qualitative trend of the tangential velocity in the rest of or less similar.
the radial region. Though RSM (380%) was able to qualitatively
follow the tangential velocity trend, severe underprediction (neg- Discussion for the simulations with the variants of
ligible swirling) was the result throughout the radial region. IBC the k − ε model
resulted into overall closer comparison for k (Figure 17D4) with
Figure 18 gives the comparison between the experimental flow
all the turbulence models employed. RSM (58%) overpredicted k
characteristics produced by DT and the predictions of the various
in the central region while standard (49%) and RNG (68%) k − ε
simulations with the variants of k − ε models (MOD k − ε, ZON
models underpredicted throughout the radial region. RNG k − ε
k − ε, and OPTCK k − ε models) using IBC approach.
model yielded largest underpredictions.

Comparison at x2 /R = 0.586
Comparison at x2 /R = 0.0 (impeller centre plane) All the variants of k − ε models (Figure 18A1) could capture the
Simulations with IBC using all the three turbulence models qualitative behaviour of radial velocity throughout the region.
resulted into consistent overprediction of mean radial velocity All the simulations underpredicted (∼50%) u throughout the

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 49 |


Figure 18. Comparison of predictions with IBC using k − ε variants. Impeller: disc turbine. A1–A4: at x2 /R = 0.586, B1–B4: at x2 /R = 0.24; C1–C4: at
x2 /R = −0.24; D1–D4: at x2 /R = −0.933.  Experimental (1) MOD k − ε model, (2) ZON k − ε model, (3) NEW k − ε model, (4) OPTCK k − ε model.

radial region. Again u predictions were found to be insensi- for 0.33 < x1 /R < 0.5 by all the k − ε variants. All the k − ε mod-
tive (with minor difference in magnitude) as a result of all the els resulted into similar predictions for axial velocity (Figure
turbulence models. Axial velocity (Figure 18A2) trend was pre- 18B2). MOD k − ε gave qualitatively opposite tangential velocity
dicted by all the k − ε variants. Underpredictions (57–72%) were for x1 /R <0.2 (Figure 18B3). Thus overall poor comparison was
given by all the models in the region with 0.2 < x1 /R < 0.65 and sought by all the k − ε models. Even the trend was not reproduced
x1 /R > 0.8. Relatively closer comparison by ZON k − ε model by any of the k − ε variants. It can be seen from (Figure 18B4)
was observed for x1 /R < 0.2. All the k − ε variants were suc- that the k was overpredicted more or less by all the models. Only
cessful in predicting the radial reversal and all the k − ε models MOD k − ε model was able to closely predict the experimentally
severely overpredicted the tangential velocity as much as 500% observed k for x1 /R > 0.26 while it resulted in severe underpre-
(Figure 18A3) for 0.13 < x1 /R < 0.3. MOD and OPT-CK k − ε mod- dictions in rest of the region. ZON k − ε model overpredicted k
els resulted in qualitatively opposite comparison for x1 /R < 0.4 (274%) throughout the radial coordinate. Though the OPTCK
and 0.13 < x1 /R < 0.4, respectively. All the models yielded similar k − ε model gave the closer comparison for 0.3 < x1 /R < 0.8, it
comparison for x1 /R > 0.5 with minor differences. The turbu- overpredicted the k (as much as 200%) in rest of the radial region.
lent kinetic energy (Figure 18A4) was underpredicted by all the
k − ε models throughout the radial co-ordinate. The deviation was
between 30% and 88%. Comparison at x2 /R = −0.24
It can be observed from Figure 18C1 that the experimental radial
velocity trend was closely followed by all the k − ε models for
Comparison at x2 /R = 0.24 x1 /R > 0.2. All the k − ε variants employed, resulted into qual-
Almost all the models underpredicted radial velocity nominally itatively opposite radial velocities near the vessel centre while
for x1 /R > 0.6, the deviation being maximum (as much as 25%) severe overpredictions for 0.2< x1 /R < 0.4. Minor deviations were
for x1 /R > 0.35 (Figure 18B1). Qualitatively opposite predictions observed for x1 /R > 0.5. Maximum deviation was due to ZON k − ε
were observed for x1 /R < 0.33 while overprediction was the result for x1 /R < 0.4. Closer comparison was obtained with the OPTCK

| 50 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 19. Comparison of predictions with MRF. Impeller: disc turbine. A1–A4: at x2 /R = 0.586, B1–B4: at x2 /R = 0.24; C1–C4: at x2 /R = −0.24; D1–D4:
at x2 /R = −0.933.  Experimental (1) STD k − ε, (2) RSM.

k − ε for x1 /R between 0.45 and 0.65 and other k − ε variants for velocity, clearly depicts the success of all the k − ε models in pre-
x1 /R > 0.65. All the k − ε models yielded good axial velocity pre- dicting the upflow near the vessel wall. ZON k − ε resulted into
dictions throughout the radial coordinate. It was found to be more the closer axial velocity comparison (40%) followed by OPT-CK
or less insensitive to the type of k − ε model (Figure 18C2). Tan- k − ε model (79%). While deviations were higher (296%) with
gential velocity predictions were largely overpredicted for x1 /R up the MOD k − ε model. Qualitatively opposite tangential velocities
to 0.3 while underpredictions (as much as 200%) was the result (Figure 18D3) were obtained as a result of all the simulations
for x1 /R > 0.3 by all the k − ε models (Figure 18C3). Predictions using all the k − ε models (except ZON k − ε model). Again MOD
of the w were more or less similar as a result of all the simu- k − ε model gave highest deviation for x1 /R < 0.4. While more
lations with k − ε models (except MOD k − ε in the region with or less similar tangential velocity predictions were the result for
x1 /R < 0.3). All the k − ε variants resulted in quantitatively similar x1 /R > 0.5 by all the k − ε models. All the variants of k − ε mod-
k predictions for x1 /R < 0.3 (Figure 18C4). Excellent comparison els severely underpredicted k (Figure 18D4) throughout the radial
was observed for x1 /R > 0.4 by ZON and OPTCK k − ε models distance. Comparison was relatively closer with ZON k − ε model
while overpredictions (100%) and underpredictions (as much as (62%). The simulations using MOD k − ε model yielded k com-
150%) were seen by NEW and MOD k − ε models, respectively. parison with maximum underpredictions throughout the radial
coordinate. Thus no k − ε variant was able to predict k near the
Comparison at x2 /R = −0.933 liquid surface.
All the variants of k − ε model predicted the radial velocity for x1 /R
up to 0.45 with minor deviations while qualitatively and quanti- Comparison at x2 /R = 0.0 (impeller centre plane)
tatively poor comparison (100–774%) was sought in rest of the All the variants of k − ε model resulted into more or less similar
radial region. MOD k − ε model though gave closer comparison predictions of radial velocity. It was underpredicted by as much as
for x1 /R < 0.4 it resulted into the largest deviation for x1 /R > 0.4 10% near the tip of impeller while the differences between the pre-
(Figure 18D1). Figure 18D2 which shows the comparison for axial dicted and the experimental data were found to decrease. Though

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 51 |


all the turbulence models resulted into qualitative predictions of a good trend and it was severely underpredicted (449–772%) by
the axial velocity, severe underpredictions (as much as 400%) all the models employed with MRF in most of the radial region
were observed due to all the simulations. The tangential veloc- (0 < x1 /R < 0.6). Both the STD k − ε and RSM were able to give rel-
ity was severely underpredicted with consistency throughout the atively closer comparison of the tangential velocity near the vessel
radial region by all the simulations. Quantitatively no difference wall. Large underpredictions in k (Figure 19C4) were observed in
was observed. IBC with all the variants of the k − ε model resulted both the simulations with MRF, the deviations being severe (as
into underprediction of the turbulent kinetic energy. Thus it can much as 86%) for x1 /R > 0.5.
be seen that the mean velocity predictions are nearly independent
of the turbulence model used. Comparison at x2 /R = −0.933
MRF was not able to capture even the trend of the experimen-
Comparison Using MRF Impeller Model tal radial velocity (Figure 19D1) in conjunction with any of the
turbulence models. Overall radial velocity predictions by both the
Comparison at x2 /R = 0.586 simulations (>900%) using MRF were poor. It was able to better
MRF with standard k − ε and RSM models underpredicted the predict the axial velocity (Figure 19D2) using standard (108%)
radial velocity (Figure 19A1) magnitude of underprediction was and RSM (114%) models near the vessel centre and the near
higher as compared to that given by IBC. Predictions of both the wall region. While severe underpredictions were the result in
simulations (52%) were nearly the same indicating insensitiv- the rest of the region, with highest deviations with the standard
ity to the turbulence models. MRF with each of the turbulence k − ε. MRF was able to replicate the tangential velocity profile
models severely underpredicted the axial velocity (Figure 19A2) (Figure 19D3) though with higher deviations (>500%) from the
throughout the radial region. Underpredictions by standard k − ε experimental data. Near the wall satisfactory comparison was
(79%) and RSM (76%) were higher as compared to that given by obtained with all the turbulence models while severe underpre-
IBC. Predictions could be found to be insensitive to the turbulence dictions were seen in rest of the radial region. It resulted into
models. As regards to the tangential velocity component (Figure relatively better tangential velocity comparison as compared to
19A3) RSM (818%) though was successful in predicting the trend the IBC. MRF with both the turbulence models (81%) underpre-
a maximum was predicted at a location close to the vessel cen- dicted k (Figure 19D4) throughout the radial region. The standard
tre. Standard k − ε model (312%) severely underpredicted up to k − ε model and RSM which yielded quantitatively similar k
x1 /R = 0.4, above which it showed good comparison with exper- predictions.
imental data. MRF resulted into very low values of the turbulent
kinetic energies (Figure 19A4) with standard k − ε (92%) and Comparison at x2 /R = 0.0 (impeller centre plane)
RSM (91%) the turbulence models employed. Turbulence models As a result of simulation with MRF approach, radial velocity was
had no significant effect on the k-predictions, which can be again underpredicted by as much as 100% for x1 /R up to 0.66 while
seen as in agreement with the insensitivity of the mean velocities excellent comparison was sought for x1 /R > 0.66. Similarly both
towards the turbulence models. the axial and tangential velocities were underpredicted (maxi-
mum deviation being 60%) for x1 /R < 0.8 by both the STD k − ε
Comparison at x2 /R = 0.24 model and RSM. As regards to the turbulent kinetic energy, each
MRF with each of the turbulence models resulted into similar of the turbulence models resulted in severe underprediction with
radial velocity (Figure 19B1) with severe underprediction from relatively closer comparison with STD k − ε model. Thus predic-
the experimental values. RSM model yielded the highest devia- tions of all the three mean velocity components were not affected
tions (638%). It gave similar axial velocity (Figure 19B2) with as a result of change in turbulence models.
all the turbulence models with marginal underpredictions from
the experimental data indicating insensitivity of v towards the Comparison Using SM Impeller Model
turbulence models. STD k − ε resulted into relatively large under-
prediction of tangential velocity up to x1 /R = 0.4, while in rest of Comparison at x2 /R = 0.586
the region predictions by both the turbulence models were similar It can be seen from Figure 20A1 that the SM with standard k − ε
(Figure 19B3). The first maximum in w was found to be shifted (88%) and RNG k − ε (91%) models severely underpredicted the
towards vessel centre by RSM while severe underpredictions were radial velocity. Again the turbulence models had no significant
observed as a result of each of the simulations in the region with effect on the predictions. SM with all the turbulence models
x1 /R < 0.5. The turbulent kinetic energy predictions (Figure 19B4) yielded severe underpredictions (95%) of axial velocity (Figure
were similar as a result of both the turbulence models employed 20A2) predictions. SM with all the turbulence models resulted
in conjunction with MRF indicating the insensitivity towards the into the almost flatter tangential velocity profile (Figure 20A3)
turbulence models. Severe underpredictions of k (∼90%) were representing absence of the swirling velocity component clearly
observed throughout the radial region. showing the failure to predict the tangential velocity. SM with all
the models resulted into the severe underprediction (76%) of k
Comparison at x2 /R = −0.24 (Figure 20A4) which could be seen as the smaller compared to
MRF with both the turbulence models failed to predict the radial that given by MRF.
velocity (Figure 19C1) quantitatively. Predictions as a result of
RSM significantly differed (226%) from the predictions of the Comparison at x2 /R = 0.24
STD k − ε model (179%). MRF was able to closely predict (minor SM was able to capture the trend in radial velocity (Figure 20B1)
underpredictions, deviations within 45% and 62%) axial velocity only for x1 /R > 0.65 while it failed to do so in the remaining
(Figure 19C2) with both the turbulence models. The turbulence region. All the turbulence models resulted into similar predictions
models again showed insignificant effect on the axial velocity pre- for x1 /R < 0.45 while difference was larger in rest of the region.
dictions. The predicted tangential velocity (Figure 19C3) showed RNG k − ε model resulted into closer comparison for x1 /R > 0.65.

| 52 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 20. Comparison of predictions with SM. Impeller: disc turbine. A1–A4: at x2 /R = 0.586, B1–B4: at x2 /R = 0.24; C1–C4: at x2 /R = −0.24; D1–D4:
at x2 /R = −0.933.  Experimental (1) STD k − ε, (2) RNG k − ε.

STD k − ε model with SM failed to indicate axial velocity (Figure seen from (Figure 20C4) that all the turbulence models result
20B2) trend and hence the downflow in near the wall (118%) into overall poor prediction of the turbulent kinetic energy when
while RNG k − ε model resulted into the wider downflow region used with SM. A peak of k could be seen with RNG k − ε model,
(94%). As can be seen from Figure 20B3, SM again failed to predict which is similar to the peak, observed in tangential velocity as
the tangential velocity with all the turbulence models employed. a result of a SM simulation with RNG k − ε model. In rest of the
It resulted almost into insignificant values of tangential veloci- region severe underpredictions (83–135%) was observed with all
ties throughout the radial region. The k-predictions (Figure 20B4) the turbulence models.
were again found to be in poor agreement (severe underpredic-
tion, 88%) with the experimental data. It was closely predicted Comparison at x2 /R = −0.933
by only RNG k − ε in a narrow region nearby to x1 /R = 0.5. SM was not able to capture even the trend of the experimental
radial velocity (Figure 20D1) using any of the turbulence models.
Comparison at x2 /R = −0.24 Again SM with all the turbulence models failed to predict the axial
SM with both the turbulence models overpredicted the radial velocity component (Figure 20D2). It showed very negligible val-
velocity (Figure 20C1) for x1 /R up to 0.55 while qualitatively ues of the axial velocity. SM with all the turbulence models failed
opposite trend was observed in rest of the region. It with all to predict tangential velocity (Figure 20D3) both qualitatively and
the turbulence models employed resulted in good comparison quantitatively. It showed almost absence of the swirling velocity.
(144–170%) of the trend in axial velocity (Figure 20C2). Large It can be seen from Figure 20D4, that the SM failed again to predict
underpredictions (1722%) were observed from the measured the turbulent kinetic energy.
tangential velocities (Figure 20C3). RNG k − ε model was able
to predict the tangential velocity closely for x1 /R < 0.4 while Comparison at x2 /R = 0.0 (impeller centre plane)
STD k − ε model resulted into severe underpredictions (3322%) SM approach resulted in satisfactory comparison of the radial
throughout the radial region (except for x1 /R < 0.15). It can be velocity using both the STD and RNG k − ε models for x1 /R > 0.5

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 53 |


ble to enhance the gas induction rate and mass transfer coefficient
together.
Further, it has been demonstrated (using both the EFD and
CFD techniques) that protein deactivation was mainly caused by
turbulent kinetic energy (Ghadge et al., 2003) (normal stresses).
Extensive data have been reported on the deactivation rate with
respect to normal stress for different reactor designs. Hence by
performing detailed hydrodynamic analysis one can arrive at the
better agitator configuration for carrying out enzymatic reactions
and similar applications. Finally, the dynamics of small eddies
influence the relative motion and, consequently, the coalescence
and breakup of small particles, droplets, and bubbles. This discus-
sion clearly indicates that the design, scale-up and optimisation
of industrial agitated tanks requires a thorough understanding of
turbulence and transport phenomena over a wide range of length
and time scales. RANS based models due to their inherent assump-
tions (such as local isotropy, truncation of higher order fluctuating
terms, empirical modelling, etc.) cannot capture local transient
flow dynamics and in some cases even predictions regarding gross
flow patterns are inaccurate. In case of LES modelling approach,
large and energetic eddies are solved and small scales are mod-
elling, as they are isotropic in nature. In view of this, LES holds
promise for understanding the local flow dynamics even in large-
scale stirred tank reactors.
The accurate prediction of the following hydrodynamic vari-
ables using LES in stirred tank reactor will greatly contribute to
establish the reliable relationship between the design objectives
and process parameters. The literature efforts on LES simulation
of stirred vessels are reported in Table 6.
Simulations have been performed for the following hydrody-
namic:
Figure 20. Continued

(1) Bulk flow with respect to impeller design, internals, and ves-
while underpredictions of as much as 80% could be observed for sel configuration
x1 /R < 0.5. RNG k − ε model yielded relatively closer comparison. (a) Axial Flow
As regards to the axial velocity, STD k − ε model gave closer com- (b) Radial Flow
parison for x1 /R > 0.5 while it was underpredicted severely for (c) Tangential Flow
x1 /R < 0.45. The tangential velocity predictions were more or less (2) Gross flow parameters
similar as a result of each of the simulations. It was marginally (a) Primary flow number
underpredicted throughout. The turbulent kinetic energy was (b) Secondary flow number
severely underpredicted by STD k − ε model with the maximum (c) Power number
departure of around 400%. Though the RNG k − ε model was able (3) Quantification of bulk velocities in the near wall regions
to predict the trend of the turbulent kinetic energy deviations of (mainly for heat transfer)
as much as 80% were observed. (4) Local distribution of turbulent kinetic energy with respect to
geometric and operating variables
LARGE EDDY SIMULATION OF RADIAL FLOW (5) Quantification of local length scale distribution
(6) Trailing vortices
IMPELLERS (a) Shape with respect to impeller blade design
To meet the various process objectives, it is essential to main- (b) Quantification of associated energy
tain the desired local hydrodynamics by manipulating both the (c) Extent up to which their effect is significant
geometric variables and operating conditions. For instance, in (d) Symmetry/asymmetry with respect to disc (in case of
crystallisation the size distribution, shape, crystalline structure Rushton turbine) when impeller is located at (1/3)T
and purity of the product are affected by the dynamics of small and (1/2)T.
eddies which influence the mass transfer and degree of super- (7) Rate of turbulent kinetic energy dissipation
saturation in the immediate vicinity of precipitating and growing (a) Local distribution (crystalliser, drop and bubble size
crystals. Therefore, crystalliser design should ensure that homoge- estimation, instantaneous reaction)
neous distribution of fine length scales and local shear rate should (8) Local and global shear rate distribution with respect to geo-
be at the desired level. Similarly in the case of hollow gas inducing metric and operating conditions
impellers, which are highly recommended to handle hazardous, (9) Energy spectrum
expensive gases and when per pass conversion is low. Sufficient (a) Identification of dominant frequencies with respect to
low-pressure zone must be created for the induction of gas and geometric and operating parameters
at the same time with good dispersion characteristics to enhance (b) Quantification of instabilities (precessing, jet and cir-
mass transfer rate. Different design alternatives can be made possi- culation instabilities)

| 54 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


(10) Mixing and characterisation of near wall turbulence with

1.73 × 106 , 13.8 × 106


respect to design conditions.

803 , 1283 , and 1603


803 ,1283 and 1603

575 000–1 200 000


32 768, 262 144
The very first study of LES for stirred vessel was performed by
Eggels (1996) using a lattice-Boltzmann discretisation scheme.
Impeller–baffle interaction model Grid points

This numerical scheme was employed because of its compu-

1 275 567
49 0000

490 000
761 760

615 000

480 000
750 000

763 000
6 × 106
tational efficiency and the ease of implementation on parallel

2403

1203

2403
processors. The main objectives were to make the lattice-
Boltzmann scheme based solver more general to simulate flows
in real complex geometries and to understand the detailed local
Adaptive force field technique

Adaptive force field technique

Adaptive force field technique

Adaptive force field technique

Adaptive force field technique

Adaptive force field technique


flow field in the vicinity of the impeller which was not possible
Momentum source approach

with the RANS based turbulence models. Therefore, this study


Sliding deforming mesh

Sliding deforming mesh


Sliding deforming mesh

Sliding mesh technique


Sliding mesh technique
presented the instantaneous and the mean qualitative flow field

Inner–outer algorithm

MRF and sliding mess


in both vertical and horizontal planes which would help in initial
understanding of the given system. Eggels (1996) had reported
predictions for the both coarse mesh and fine mesh. Logically,

Sliding mesh
Sliding mesh

Sliding mesh
fine mesh was found to give better predictions than the coarse
mesh. However, there were differences in the quantitative predic-
tions of the mean radial and mean axial velocities especially in the
near impeller region. The discrepancies were due to insufficient
statistical sampling time which was limited by the computing
Reynolds number

14 000, 82 000,

14 000, 82 000,
time with the fine grid and adoptive force field technique used
for the impeller modelling. Further, the quantitative comparison
350 000

350 000

has been restricted to a few flow variables (axial and radial veloc-
107 000
60 000
29 000
40 000

40 000
30 000

45 000
46 600
29 000

56 250
24 000
22 000

45 000

ity) and also at limited locations (one radial and one axial line)
7300

9720

and energy balance was not established. The instantaneous flow


structures and corresponding turbulence characteristics were not
T = 0.48 m, H = T, C/T = 0.33, D/T = 0.33

T = 0.48 m, H = T, C/T = 0.33, D/T = 0.33

Non-standard geometry, T = , H = 1.86T,

given much consideration.


T = 0.44 m, H = T, C/T = 0.5, D/T = 0.33

T = 0.15, D = T/3, C = T/3, H = T, baffle

T = 0.292, H = T, D = 0.35T, C = 0.46T

Two year later Revstedt et al. (1998) investigated both the mean
T = 0.10, D/T = 0.33, C = T/3, H = T

T = 0.10, D/T = 0.33, C = T/3, H = T

flow and spectral distribution of flow by power density function


T = 0.147, H = T, D = T/2, T/3, T/4
T = 0.19, D/T = 0.5, C = T/3, H = T

H = 1.86T, D = 2/3T, C = 0.0375T


T = 0.29, H = T, D = T/3, C = T/3
T = 0.27, H = T, D = T/3, C = T/3
T = 0.27, H = T, D = T/2, C = T/2

at various locations in a baffled stirred tank. It has been the finite


T = 0.3, H = T, D = T/3, C = T/3

difference based solver where the effect of SGS scale motion on


the large scale was not taken into account and the process of
H = T = 0.45, C = D = T/3
D = 2/3T, C = 0.0375T

H = T = 0.3, C = D = T/3

energy dissipation was numerically treated by using second-order


clearance = 0.017T

upwind scheme. In this study authors have made an attempt to


Geometric details

reduce the computational demand of LES by using momentum


source approach for impeller modelling. The LES could capture
the trailing vortices behind the blades and their detailed move-
ment away from the impeller with increasing distance from the
blade. These observations were in close agreement with the expen-
sive and time-consuming experiment. They demonstrated that
Dynamic one equation model

the LES could generate instantaneous flow field like sophisti-


Standard smagorinsky model

Standard smagorinsky model


Standard smagorinsky model

Standard smagorinsky model


Standard smagorinsky model

Standard smagorinsky model


Standard smagorinsky model
Standard smagorinsky model
Standard smagorinsky model
Standard smagorinsky model

Standard smagorinsky model


Dynamic smagorinsky model
Standard Smagorinsky and
Standard smagorinsky and

Standard smagorinsky and

cated flow measuring techniques and sometimes even better as


far as the flow in the impeller region and near wall regions are
concerned. The LES predicted time averaged axial velocity was
mixed sgs models
voke SGS models

shown to agree with the experimental data of Costes and Coud-


mixed model
No SGS model

erc (1988) at various locations. Comparison for the mean radial


Table 6. Previous work on LES of the stirred tank

SGS model

velocity at only one axial location was given (impeller centre line),
which was very good at all the locations except 0.5 < x1 /D < 0.6.
They have also obtained a fairly good prediction for the turbulent
kinetic energy in the impeller centre plane and a good agreement
between the experimental and predicted value of pumping num-
Derksen and van den Akker (1999)

ber. Acceleration of the fluid at the impeller tip (because of the


Bakker and Oshinowo (2004)

fluid entrainment due to the trailing vortex) pair was well simu-
lated. Prediction of radial and tangential velocity with the angle
Murthy and Joshi (2008)

Murthy and Joshi (2008)


Hartmann et al. (2004a)

Hartmann et al. (2006)

was poor. They have suggested that the boundary conditions on


Delafosse et al. (2008)
Revstedt et al. (1998)

Derksen et al. (2007)

Derksen et al. (2007)


Alcamo et al. (2005)

Jahoda et al. (2007)


Zhang et al. (2006)

the impeller region should be improved. Their instantaneous data


Yeoh et al. (2004)

Yeoh et al. (2005)

Fan et al. (2007)

from LES supported the blade frequency and the existence of Kol-
Eggels (1996)

mogorov’s −5/3 region in the energy spectrum. The LES data


clearly represented periodicity in the impeller region and away
from the impeller the extent of periodicity was found to be negli-
Refs.

gible. However, the detailed comparison of all the flow variables


with the experimental data throughout the vessel has not been

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 55 |


reported. Authors should have explored the trailing vortex posi- prediction of the precessing vortices present in a stirred tank. The
tion by specifying vessel top as a wall. Further, comparison has detailed results gave insight in the structure and characteristic
not been reported for the turbulent kinetic energy dissipation rate. frequency of the vortices, and drew attention to the precessing
Derksen and van den Akker (1999) took forward Eggels (1996) vortex as a point for improvement in the existing modelling tech-
study with a more refined forcing algorithm especially to account niques. Furthermore, the results form a promising perspective for
for the impeller motion. The real potential of LES was partially the application of the LES methodology in future studies consider-
explored as the simulations could identify a detailed local flow in ing the significance of the precessing vortex in multiphase mixing
the wakes behind each blade which were found to be significantly processes. However, before the quantification of instabilities, LES
higher than the impeller tip speed (two times tip speed). Further, model should have been compared with experimental data of tan-
away from the impeller the fluctuations become less coherent. As gential velocity and turbulent kinetic energy profiles. Further, the
far as the quantitative comparison is concerned, good agreement study did not include the other types of instabilities such as jet and
has also been obtained between the experimental data and the intermediate instabilities. Also, an attention is needed to under-
simulation of the radial profiles of the radial velocity. However, the stand the instabilities associated with non-standard geometries.
tangential velocities have been overpredicted (by maximum 15%) Yeoh et al. (2004) have performed LES using sliding deform-
and it was attributed to a lack of spatial resolution and improper ing technique and finite volume based solver. The main objective
treatment of the SGS viscosity. The axial profiles of random (turbu- of this work was to assess the capability of the LES, coupled
lent), coherent (pseudo turbulence) and total kinetic energy have with the sliding-deforming methodology, of yielding an improved
been fairly predicted when compared with the LDA phase resolved simulation of the flow inside a stirred vessel. The results were
experimental data. Having obtained realistic kinetic energy dis- compared with detailed experimental data of mean and RMS
tribution, they extended their LES predictions to calculate the velocities and energy dissipation rates. Results from the stan-
turbulent kinetic energy dissipation rate. It was observed that the dard RANS–k − ε model were also presented and assessed in the
dissipation rate distribution throughout the tank is very inhomo- light of flow features and characteristics obtained from LES. In
geneous. Further, the impeller power number was calculated from general, there was very good qualitative and quantitative agree-
the estimation of torque on the impeller and was found to be ment between the LES predictions and the measured data. In this
slightly overpredicted. However, the comparison for axial veloc- paper, it was shown that, both from the local isotropic distribution
ity was not presented and all the predictions have been reported contour plots and the turbulent kinetic energy quantitative com-
only in the impeller region. Information about local shear rates as parisons at the impeller disc level, the equilibrium and isotropy
well as physics associated with the instabilities would have been hypothesis was indeed applicable in most parts of the tank except
a valuable addition to their work. for the impeller discharge stream. Comparisons of the LES and
Further, Hartmann et al. (2004a) extended their previous LES RANS predictions of the mean velocity and turbulent quantities
model to investigate the precessional vortex phenomenon. Macro with experimental data have indicated clearly that substantial
instability phenomena was included in the in this work. The improvement in the predictions can be achieved with LES that
sequences of snapshots of the flow field in a horizontal plane should facilitate the more reliable design of mixing processes.
below and above the impeller showed a vortical structure moving The standard Smagorinsky model, although relatively prelimi-
around the tank centreline in the same direction as the impeller. nary compared with more advance turbulence models such as
The vortex below the impeller was found to be more pronounced the dynamic model, was proven to be able to simulate accurately
in strength and size compared to the vortex above the impeller, the fluid phenomena both qualitatively and quantitatively. The
and both vortices move with a mutual phase difference. The pre- power number obtained from integration of the energy dissipa-
sented time series of the velocity components reveal, besides the tion rate was predicted to be within 15% of the measured value,
random turbulent fluctuations, two types of coherent fluctuations while the comparison of the local values of ε in the impeller stream
with time scales separated by more than two-orders of magnitude. was very encouraging, given the approximations involved in the
The fluctuations in the impeller out-flow region were domi- measurements. The data presented comprise the first quantitative
nated by the passage of the impeller blades, whereas fluctuation comparisons of the distribution of turbulent energy dissipation
levels close to the tank centreline were dominated by the low- in stirred vessel with direct measurements of ε and show much
frequency motion of a vortical structure. Authors have reported promise for improved prediction of both ε and k in mixing pro-
the frequency analysis and it was found that the characteris- cesses. Similar to most of the earlier investigations, this work has
tic frequencies of f = 0.0255 N and f = 0.0228 N at flow Reynolds also presented the comparison at only one axial location (impeller
numbers of 20 000 and 30 000, respectively. These values of the centre plane).
characteristic MI frequency have been in good agreement with Hartmann et al. (2004b) also made a detailed study to evalu-
reported (turbulent) frequencies found experimentally in similar ate the predictive capabilities of LES and RANS. They have used
flow geometry. At a Reynolds number of 12 500, a second fre- finite volume based solver, whereas Yeoh et al. (2004) employed
quency peak at f = 0.092 N was observed, which was consistent the lattice-Boltzmann solver. Further, the RANS based simulations
with the experimentally observed frequency, the so-called laminar were performed with the shear stress transport (SST) in place
frequency, in the literature. Further, the following observations of standard k − ε model. The RANS, LES predictions have been
have been reported. In the impeller outflow region, the flow was compared with the LDA measurements of radial, tangential veloc-
dominated by the passage of the impeller blades. In the bulk flow ities and turbulent kinetic energy in the impeller zone. The radial
region, the flow was dominated by the low-frequency precessing velocities of both the models were in good agreement with the
vortex, and at least 100 impeller revolutions were needed for an experimental data. As regards to the tangential velocity, the over-
accurate prediction of the kinetic energy contained in the velocity all performance of LES was found to be much superior to RANS
fluctuations. By means of a low-pass filtering procedure, it has simulations. Their simulations revealed that the Smagorinsky SGS
been observed that, in the top and bottom parts of the tank, a model performed better than Voke SGS model, especially, when
significant amount (up to 44%) of the kinetic energy is related the prediction of trailing vortex pair is concerned. It was further
to the precessing vortex. LES has proven its significance in the shown that the LES predictions of the turbulent kinetic energy in

| 56 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


the impeller discharge flow were substantially better than RANS compared with in-house experimental data obtained by means
which invariably underperformed. Using LES, they found near of PIV, as well as, when possible, with the literature results. An
isotropy in the circulation loops. However, in the impeller stream, excellent agreement was found between simulation results and
the boundary layers, and at the separation points, the turbulence experimental measurements, especially regarding mean tangen-
was found to be more anisotropic due to high shear rates. Overall, tial velocities. This is a significant contribution. Good agreement
they did not see the relative merits of using Voke SGS model. Fur- was also observed for other flow quantities, such as radial aver-
ther, they calculated the energy dissipation rate by assuming local age velocities. Also the presence and trajectory of a pair of trailing
equilibrium between production and dissipation at and below vortices, whose existence has long been known in baffled tanks
subgrid scale level and compared with the RANS predictions in and as recently confirmed also in unbaffled tanks, was correctly
the baffle mid-plane. It was observed that the RANS based model simulated. The good agreement observed between the simulation
yielded higher values of the energy dissipation rate than the LES. results and the experiment is particularly meaningful, due to fact
However, qualitatively both the models were found to predict very that the flow field in unbaffled vessels is a particularly severe
inhomogeneous distribution of the energy dissipation throughout benchmark for turbulence models.
the tank. Again, this study also has presented comparisons in a Jahoda et al. (2007) carried out LES simulation using FLUENT
limited region. Further, axial velocity comparison has not been to predict mixing time. They have carried out RANS simula-
given. tion using two impeller modelling approaches (i) MRF (ii) SM
Bakker and Oshinowo (2004) carried out LES to predict large- approach. They have shown the comparison for axial and radial
scale vortical structures using various impellers. Their work velocity component. Scalar transport equation was solved to get
reported vector plots and isovorticity surfaces close to impeller the transient concentration profiles and then mixing time. Unlike
in stirred vessel. They did not provide any quantitative compari- other investigators, they have estimated the power number and
son. Fan et al. (2007) investigated flow instabilities in stirred tank flow number using RANS and LES simulation. Power number
using SM approach with CFX5 software. They have carried our predicted by RANS simulation was 20% lower compared to LES.
simulations using standard k − ε model and LES and compared The estimated flow number was almost same for all simulations
the results with PIV data. They have shown the comparison of whereas mixing time was underpredicted using MRF approach
axial and radial velocity along the axial distance. They have also and overpredicted using SM with standard k − ε model. Mixing
reported vector and contour plots to show transient circulation time prediction using LES were in good agreement with the exper-
patterns. They have reported time series of axial and radial veloc- imental values.
ity. It would have been very useful if energy spectra at various Zhang et al. (2006) carried out LES simulation using improved
locations were analysed. inner–outer approach. They have shown the comparison of mean
Yeoh et al. (2005) extended their previous work to study the axial, radial, and tangential profiles as well as turbulent kinetic
mixing of inert tracer using LES. This was the first attempt to energy. The comparison was carried out with the experimental
study the detailed mixing phenomenon using LES. The work data of Wu and Patterson (1989). They have also presented the
presented here has clearly demonstrated that the LES method vector plots to show transient vortical structures in the impeller
coupled with the sliding-mesh methodology can be used to char- region.
acterise the transient mixing state in a stirred vessel and to Derksen et al. (2007) extended their previous lattice-Boltzmann
provide a very detailed evolution pattern of the concentration scheme based LES solver for industrial scale crystalliser. Their
field in space and time during the process. The transient analysis work was focused on the assessing the feasibility of using a
revealed that, at the impeller midsection, a substantial amount of computationally efficient LES to quantify the fine scale turbulent
scalar was trapped in a region between two vertical baffles. Very structures in an industrial size crystalliser. The effects of spatial
large temporal fluctuations were detected close to the impeller, resolution and the sub-grid scale model (Smagorinsky and the
with instantaneous concentrations being as high as three times mixed scale models) were also investigated. No significant differ-
the equilibrium value. The time history of scalar concentration ence was observed in the predictions of the two sub-grid scale
revealed that mixing can vary considerably across the vessel models. Further, the particle tracking technique was employed
and the use of experimental techniques that provide informa- for the study of mixing performance. Practically uniform condi-
tion at one or more locations with intrusive probes is likely to tions were obtained after 5–10 impeller revolutions with respect
lead to a less realistic mixing time. The work presented was of to concentration of fluid elements, except for the bottom region
importance for industrial process prediction as it offers not only which gets characterised by the presence of fresh feed. However,
the final mixing time but also the localised mixing character- LES predictions have been discussed more qualitatively because
istics during the intermediate mixing stages at various points of the non-availability of the experimental data from the industrial
across the vessel. In addition effects associated with stagnant scale equipment.
zones, mixing inhomogeneities due to the vessel geometry, opti- Murthy and Joshi (2008) have carried out critical analysis stan-
mal location of feed pipes, etc., can be easily identified. They have dard k − ε model, RSM and LES. The comparison was carried out
concluded that the good agreement of the predicted mixing time with in-house laser Doppler velocimetry (LDV) experiments. The
with available correlations indicates the capability of this mod- simulations were carried out using one equation dynamic sub-
elling approach to accurately simulate mixing characteristics in grid scale model. LES results were relatively superior compared
a stirred vessel and therefore should be employed for the opti- to other turbulence models. They observed that RSM and standard
misation of industrial stirred vessel design. This work, however, k − ε model underpredicted the turbulent kinetic energy profiles
did not include the effect of various instabilities on local mix- significantly in the impeller region. RSM can capture well all
ing time. Authors should have extended the model to see the the mean flow characteristics and the standard k − ε model fails
effect of various internals on the local mixing time and stagnation to simulate the mean flow associated with the strong swirl. In
points. addition to mean and turbulent quantity comparison, they have
Alcamo et al. (2005) presented LES simulation of unbaffled tank reported energy associated with precessional vortex and jet insta-
stirred by a Rushton turbine. The numerical predictions were bility. The frequency of the jet instability was found to be linearly

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 57 |


related to the rotational speed. They have also established the
overall energy balance using both integral ε and torque-based
approaches.
Zadghaffari et al. (2010) simulated a six bladed Rushton tur-
bine and compared the mean velocities, turbulent kinetic energy,
energy dissipation rate with experimental data of Wu and Pat-
terson (1989). Though the mean velocities were in very good
agreement with the experimental data, the turbulent kinetic
energy and energy dissipation rate showed only qualitative agree-
ment. They compared, power number results obtained from
simulation with experimental data of Walas (1990); and Zadghaf-
fari et al. (2009) and mixing time with that of Murthy et al. (1994).
They found good agreement between simulation and experimen-
tal results.
CFD simulations have been performed with LES turbulence
model for the five different impellers. The simulation results of
the dimensionless mean axial velocity, mean radial velocity, mean
tangential velocity, and turbulent kinetic energy have been plot-
ted against the LDA experimental data. The comparison of radial
profiles of these parameters has been made at eight different axial
levels, x2 /R = 0.01 m (A), 0.044 m (B), 0.082 m (C), 0.1 m (D),
0.118 m (E), 0.154 m (F), 0.190 m (G), and 0.244 m (H). It may be
pointed out that the exercise of comparison covers practically the
entire region in the vessel both near and away from the impeller.
In case of DT, the high-speed impeller stream impinges on
the tank wall and changes the direction three times to return to
impeller again. This recirculatory flow exists in the bulk region
of the tank. Near the eye of circulation, very small mean veloc-
ities exist. The radial profiles of axial velocities at various axial
locations are shown in Figure 21A–H. It can be seen that maxi-
mum axial velocities exist near the wall and are of the order of
0.20–0.30 times the tip speed. However, the axial velocity changes
more sharply in the near wall region compared to that in the near
axis region of the vessel. As one moves away vertically from the
impeller swept region, axial velocity initially increases, attains a
maximum and then decreases. It is evident that the predictions of
all the axial velocity profiles by the LES in good agreement with
Figure 21. Comparison between LES simulation and experimental data
the experimental data. for dimensionless mean axial velocity. Impeller: disc turbine. (A)
As regards to the radial component, impeller rotation generates H = 0.01 m; (B) H = 0.044 m; (C) H = 0.082 m; (D) H = 0.1 m; (E)
radially outward flow through the vertical surface of swept vol- H = 0.118 m; (F) H = 0.154 m; (G) H = 0.190 m; and (H) H = 0.244 m: 
ume. This high-speed radial jet entrains surrounding fluid and experimental LES.
slows down as they approach the tank wall. It can be noted from
Figure 22A–H LES predictions agree well with the experimental
mean radial velocity profiles at all the axial levels. One million LES provides macro and the reliable predictions of the turbulent
grids were used for this simulation. flow, the predictions can further be useful for better predictions
Radial profiles of the mean tangential velocity are depicted in of macro and micromixing, heat and mass transfer.
Figure 23A–H. Again, LES simulations capture the experimental
mean tangential velocity profiles quite well. ASSESSMENT OF ASSUMPTIONS IN RANS
Figure 24A–H illustrates the comparison for turbulent kinetic
energy throughout the tank. In the impeller central plane (Figure MODELS
24D) the profile of turbulent kinetic energy is practically simi- It has already been emphasised that the RANS approach is popular
lar to those of radial velocity and tangential velocity. Below the than LES and DNS for stirred tank simulation due to lesser compu-
impeller (Figure 24C), turbulent kinetic energy also exhibits two tational power requirement. However, RANS predictions require
maxima as in the case of tangential velocity. The first maxima experimental validation. Tuning the model parameters has only
is at x1 /R = 0.27 and the second one occurs closer to the wall at partially succeeded in improving the predictions for turbulent
x1 /R = 0.92. The former is generated by the rotating impeller and kinetic energy, dissipation rate. The Reynolds averaging process
latter is because of the shear flow. The kinetic energies below the introduces extra terms containing correlations of fluctuating com-
impeller are much smaller than those in the impeller centre plane ponents. These terms are modelled making some assumptions
and above the impeller (Figure 24E), it is mainly because of the about the turbulent transport processes. Due to the modelling
upward inclination of trailing vortex pair. It can be observed that assumptions and truncation of higher order terms, a model devel-
good predictive capabilities of LES can be clearly seen for the tur- oped for one purpose deviates significantly in other types of flows.
bulent kinetic energy predictions. Here one can justify the utility In the literature abundant information is available on compar-
of LES even at the expense of computational resources. As the ison of various turbulence models for different types of flows

| 58 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 23. Comparison between LES simulation and experimental data
Figure 22. Comparison between LES simulation and experimental data for dimensionless mean tangential velocity. Impeller: disc turbine. (A)
for dimensionless mean radial velocity. Impeller: disc turbine. (A) H = 0.01 m; (B) H = 0.044 m; (C) H = 0.082 m; (D) H = 0.1 m; (E)
H = 0.01 m; (B) H = 0.044 m; (C) H = 0.082 m; (D) H = 0.1 m; (E) H = 0.118 m; (F) H = 0.154 m; (G) H = 0.190 m; and (H) H = 0.244 m: 
H = 0.118 m; (F) H = 0.154 m; (G) H = 0.190 m; and (H) H = 0.244 m:  experimental LES.
experimental LES.

and with relatively fine mesh in the impeller region in order to


with the objective of understanding the failure of the models. The better resolve the strong velocity flow field. In the present LES
next question arises, how does parameter tuning and additional simulations, the length scales were resolved between 200 ␮m and
constants in a model proposed for one type of flow improve the 4.2 mm in contrast to Kolmogorov’s length scale of 20 ␮m. All the
predictability even in some other type of flow. One possible rea- simulations were carried out using Fluent software. The modelled
son for such observation can be an error made by one assumption terms as obtained from standard k − ε model and the RSM have
can be nullified by making one more error which will compen- been compared with that obtained from LES (which is considered
sate for the first one (Sciberras and Coleman, 2007; Mathpati et to be exact and close to DNS).
al., 2009). But such an exercise will not improve the true reli- The modelled and exact terms in TKE have been compared.
ability of the model and deviations can be observed in the same The variations throughout the reactor are almost 3–4 orders of
geometry at different flow conditions (which can be characterised magnitude. Hence the contours are zoomed in at four locations.
by the level of power consumption per unit mass). Hence it really These four locations are shown in Figure 25. Region A is in the
becomes necessary to quantify the effect of each modification in impeller plane, region B is below the impeller whereas region C
the model. shows the above impeller plane. Region D is close to wall in the
In this section, analysis of the important terms in the trans- impeller zone, where liquid jet discharging from the impeller hits
port equation for turbulent kinetic energy and Reynolds stress the sidewall. This is similar to impinging jet (stagnation flow) type
equation has been carried out. The major emphasis is given to tur- situation. Figure 26 shows the turbulence production term in TKE
bulent production, turbulent transport and pressure strain terms. equation. Analysis over entire plane shows that modelled term
The geometric and operating conditions are maintained same as is 4–5 times lower than the exact term. The exact and modelled
discussed in previous sections. The grid size was increased to terms are comparable in below impeller and above impeller region
1 million hexahedral cells for the standard k − ε model and the (Figure 26BE, CE, BM, and CM). The differences can be observed
RSM. In the case of LES, 4.2 million hexahedral cells were used in the impeller zone (Figure 26AE and AM). Figure 27 shows the

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 59 |


Figure 25. Comparison of various turbulence models at different
locations in stirred tank reactor. (A) Impeller discharge stream (B) below
impeller region (C) above impeller region (D) impinging jet in wall.

off the bottom. In particular, it should be used in creating enough


flow to suspend the particles in the reactor. In dispersion appli-
cations (gas–liquid and liquid–liquid) systems, drop size or the
bubble size is markedly influenced by local values of the turbu-
lent energy dissipation rate (ε). In biochemical operations, the
value of maximum turbulent energy dissipation rate (εmax ) may
also govern the viability and the growth rate of microorganisms,
Figure 24. Comparison between LES simulation and experimental data particularly plant and animal cells. In such cases, the total power
for dimensionless turbulent kinetic energy profiles. Impeller: disc turbine. input or the average rate of energy dissipation (ε̄) is not suffi-
(A) H = 0.01 m; (B) H = 0.044 m; (C) H = 0.082 m; (D) H = 0.1 m; (E) cient to completely describe the system. The same power input per
H = 0.118 m; (F) H = 0.154 m; (G) H = 0.190 m; and (H) H = 0.244 m: 
experimental LES. unit mass can result in widely different distributions of the turbu-
lence energy dissipation when different impellers are used with
the same tank geometry. As a result of this, different impellers
turbulence transport term in TKE equation. Quantitative differ- show different performance characteristics even when compared
ence can be seen in region A (i.e., impeller zone). Region D, which at the same total power input. Thus, the measurement of ε requires
is similar to the impinging jet type flow, shows that the region of the measurement of fluctuating velocity gradients in all the three
positive and negative transport term is accurately predicted by directions. This would mean that the fluctuating velocity in all
exact terms using LES, whereas modelling does not capture this the three directions would have to be measured over very short
physics (Figure 27DE and DM). Figure 28 shows the pressure distances (the distance between the measurement points would
strain term in impeller region. This particular graph shows the have to be of the same order of magnitude as the smallest eddy
importance of pressure strain term. In the impeller zone, radial size).
flow is more significant than the axial flow. It can be seen that Turbulent energy dissipation rate is defined as:
except axial normal stress, all other terms are negative in radial
direction. This means that the energy is extracted out from other ε = 2Sij Sij (12)
components and fed to the axial normal stress which further helps
in restoring the isotropy and helps in the liquid recirculation. where the fluctuating strain rate Sij is defined as:
 
1 ∂ui ∂uj
ENERGY BALANCE AND REYNOLDS STRESSES Sij = + (13)
2 ∂xj ∂xi
Any process or equipment has to be energy efficient to be eco-
nomically viable. In agitated reactors, energy efficiency means Clearly, this is not feasible by the currently available measure-
the effective utilisation of the input energy to achieve the process ment techniques. Therefore, the turbulent energy dissipation rate
objective. For example, in the case of solid suspension, energy can only be estimated. Over the years, a number of different meth-
has to be spent near the vessel bottom to lift the solid particles ods have been proposed for the estimation of ε. These have been

| 60 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Figure 26. Turbulence production term in TKE equation in stirred tank with disc turbine.

Figure 27. Turbulence transport term in TKE equation in stirred tank with disc turbine.

reviewed by Kresta and Wood (1993); and Sahu et al. (1999). One where u is a characteristic velocity scale, L is a characteristic
of the simplest ways of estimating the turbulent energy dissipa- length scale, and A is a constant of proportionality. Different meth-
tion rates is to assume that the eddy transfers all its energy within ods reported in the previous literature use different methods for
one lifetime. This results in the following estimate of ε: estimation of u and L.
In the published literature, the characteristic velocity scale has
u 3 been taken as the
√ RMS velocity or the square root of the turbulent
ε=A (14) kinetic energy ( k). The characteristic length scale has also been
L

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 61 |


Figure 28. Production term in dissipation rate equation in stirred tank with disc turbine.

defined in a number of ways: (i) as a simplest approximation, the methods yield different values of L and hence different estimates
characteristic length scale has been taken to be proportional to of ε.
the impeller diameter. The proportionality constant has been cho- The energy dissipation rate can also be estimated from Equa-
sen arbitrarily to be between 0.1 and 1 (Kresta and Wood, 1993). tions (11) and (12) by assuming Taylor’s frozen turbulence
Moreover, these authors have not considered the possible varia- hypothesis (Kresta and Wood, 1993). Further, if the turbulence
tion of the characteristic length scale throughout the vessel. The is isotropic, then, ε can be estimated using the following relation-
assumption of constant L is totally unrealistic; (ii) the length scale ship:
has been estimated using the auto-correlation function (Wu and
Patterson, 1989). The length scale can be estimated by first calcu- uRMS
2
lating the integral time scale () using auto-correlation function ε = 15 (17)

2
as:
∞ Here, the
is the Taylor microscale, which can be estimated
r= r()d (15) from the curvature of the auto-correlation function.
The energy associated with the fluctuating velocity component
0
represents the energy associated with the turbulent motion. The
Here the integration may be carried out for a long time or up Fourier transform of the velocity fluctuations give the energy con-
to the first zero crossing. Commonly the integration is carried out tributions of the fluctuations at different frequencies. This can
up to the first zero crossing. then be represented as a plot of energy associated with a par-
Anandha Rao and Brodkey (1972) have reported that the oscil- ticular frequency versus the frequency. This plot is commonly
lations in the auto-correlation function were due to the periodicity called the power spectrum or the energy spectrum (Mujumdar et
of the flow especially near the impeller blades. Thus, if the inte- al., 1970). The spectra are commonly represented in terms of the
gration is carried out for a long time, the negative values of the wave number. The wave number can be related to the frequency
auto-correlation function may actually reduce the value of the as:
integral time scale.
The Taylor microscale can be obtained from the curvature of 2f
k1 = (18)
the auto-correlation function as: V

d2 r()

−2
=0= 2 (16) If the turbulence is assumed to be isotropic, the energy dis-
d 2

t sipation rate can be estimated from the one-dimensional energy
spectrum using the following relationship:
The above time scales can be converted to the length scale by
multiplication with the mean velocity or the RMS velocity. Thus
the two ways of obtaining the time scale and the two ways of ∞
getting the length scale from the time scale result into four possible ε = 15 k12 E(k1 )dk1 (19)
ways of obtaining an estimate of L. It is quite likely that these 0

| 62 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


According to the Kolmogorovs hypothesis, the Energy spectrum It was observed that the angular momentum was conserved in
in the inertial sub-range can also be used to obtain the energy the impeller discharge stream, except in the region close to wall
dissipation rate (Wernersson and Tragardh, 1998) using the fol- where the flow diverged. The total energy flowing through the
lowing equation: impeller discharge stream as a function of the radial distance was
also computed. It was found that of the total energy given by the
E(k1 ) = Aε2/3 k −5/3 (20) impeller about 60–80% was dissipated in the impeller discharge
stream and the remainder was dissipated in the rest of the tank.
The local rate of energy dissipation was estimated using Equation
Previous Work (13). The characteristic velocity scale was taken equal to the RMS
Ranade and Joshi (1990b); Kresta and Wood (1993); and Sahu et velocity and the characteristic length scale was calculated from the
al. (1999) have estimated the energy dissipation rates by different integration of the correlation function. The value of A was taken
methods and have clearly shown that the different methods yield as 1.1. The local values of ε were integrated to obtain estimated
dramatically different values of ε. One check on the values of ε is of the energy dissipated in different regions of the tank. It was
that the integration of the local ε values must equal the total power observed that about 20% of the energy was dissipated within the
given by the impeller, that is the power number calculated from impeller swept volume, 50% was dissipated in the impeller dis-
the integration of the local ε values must match the experimentally charge stream and 30% was dissipated in the remaining portion
observed (with the help of shaft torque or the torque table) power of the tank. However, in order to match the total power dissipated
number. Patwardhan (2001) has studied this aspect for a wide by integration of the local ε values with the power dissipated mea-
variety of axial downflow impellers. It was thought desirable to sured with the torque table, the value of A had to be reduced to
review in detail the various methods for the estimation of ε and 0.34. The ratio of the local energy dissipation rate to the average
the results arising out of these. energy dissipation rate was also computed. It was observed that
Kim and Manning (1964) have investigated the turbulence the ratio, ε/ε̄, was very high near the impeller tip (about 70) and
characteristics of stirred vessels. Their experimental set-up con- reduced rapidly to about 3.5 near the tank wall, and was about
sisted of 0.6 m diameter vessel fitted with standard Rushton 0.26 outside the impeller discharge stream.
turbines having diameters 0.12–0.17 m. That is, the D/T ratio Mujumdar et al. (1970) have measured the turbulence charac-
was varied only between 0.2 and 0.28. The impeller speed was teristics in 0.38 m diameter stirred reactor using 0.12 and 0.15 m
varied over the range 100–800 rpm, and the liquid viscosity was diameter Rushton turbines with the help of a hot wire anemome-
varied from 0.8 to 100 cp. The intensity of turbulence was mea- ter. The turbulent intensity (the ratio of the RMS velocity to the
sured with the help of a pressure transducer. The RMS velocity impeller tip speed) was found to be independent of the impeller
was related to the output of the pressure transducer, which was speed and the diameter. Based on the velocity–time data, the
used to calculate the intensity of turbulence. This was then auto-correlation function was computed. Using this, the energy
used to compute the energy spectra. They have observed that spectrum and the integral length scales were calculated with the
the energy spectra were practically unaffected by impeller size, help of Equation (14) and the mean velocity. The shape of the
impeller speed, and probe position within the impeller discharge energy spectrum function was found to be independent of the
stream. location in the impeller discharge stream.
Cutter (1966) has investigated the turbulence characteristics Anandha Rao and Brodkey (1972) have investigated the turbu-
in a 0.3 m stirred tank fitted with a 0.1 m diameter Rushton tur- lence characteristics in the impeller discharge stream of a straight
bine. The local fluid velocity was measured by photographing blade turbine. The experiments were carried out in 1 0.3 m diam-
lycopodium particles with a short exposure time so that the par- eter tank and the D/T ratio was 0.33. The measurements were
ticles showed-up as streaks. Photographs were taken through the made with the help of a 3-D pitot tube. They have observed
bottom and side of the vessel. The photographs taken through the that the intensity of turbulence (ratio of the RMS velocity to the
bottom were used to analyse the radial and tangential velocity local mean velocity) was practically independent of the radial dis-
and the correlation coefficient at different distances. The pho- tance in the impeller discharge stream. The velocity–time data
tographs taken through the side of the vessel were used to obtain was also used to compute the auto-correlation function. The auto-
mean velocities. From the data, it was found that the normalised correlation function was used to calculate the integral length scale
fluctuating velocity components (normalised with respect to the using Equation (14) and for computing the energy spectrum. The
impeller tip velocity) were independent of the impeller speed. The energy spectrum showed peak corresponding to the impeller rota-
decay of turbulence as a function of the radial and axial distance tional speed followed by a region exhibiting −5/3 slope. The
away from the impeller was also found. The photographic mea- Taylor microscale was calculated in two ways: (i) using the RMS
surements allow estimation of the local velocity as a function of velocity and the velocity gradients, and (ii) the curvature of the
the distance. This enabled computation of the correlation coeffi- auto-correlation function. It was observed that the microscale cal-
cient as a function of the distance. It was found that the correlation culated using either of the two definitions were practically the
coefficient as a function of the distance was independent of the same. The microscales calculated were used to estimate the local
impeller speed. From the correlation plots, the scale of turbulence energy dissipation rate using Equation (16).
was calculated in the radial and tangential direction by integra- Komasawa et al. (1974) have measured the turbulence charac-
tion. It was observed that the radial scale increased approximately teristics with the help of particle trajectories obtained with the
as square root distance from the impeller tip. The tangential scale help of high-speed camera. The experiments were conducted in
was always lower than the radial scale and it increased up to a cer- 0.10 and 0.15 m diameter stirred vessels equipped with Rushton
tain distance from the tip, beyond which, it again reduced. It was turbines. Using the velocity–time data the cross correlation as well
also observed that the turbulence was close to isotropic near the as auto-correlation were obtained. Energy spectra were obtained
impeller blades, and the anisotropy increased with an increase in by taking Fourier transform of the auto-correlation function and
distance from the impeller tip. A balance of the angular momen- were found to be in agreement with the previous literature.
tum was also established and was used to calculate the torque. Using the auto-correlation function, the Taylor microscale was

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 63 |


calculated. It was observed that the Taylor microscale increased velocity was measured with the help of a hot wire anemometer.
with an increase in the radial distance in the impeller discharge The energy dissipation rate was calculated from the energy spec-
stream. They have also compared the energy spectra in the trum function using Equation (18). It was observed that, in the
impeller discharge stream in stirred vessels having different diam- impeller discharge stream, the local energy dissipation rate was
eters with that for a pipe flow for the same value of the local energy substantially higher (3–10 times) than the average energy dis-
dissipation rate. It was observed that the three spectra were prac- sipation rate. In the region above the impeller, the local values
tically identical. This would indicate that the turbulence structure of energy dissipation rate were about 2–5 times smaller than the
at a point is nearly independent of the details of the turbulence average energy dissipation rate. The distribution of dimensionless
generating geometry. In other words, the small scales of turbu- ε was found to be independent of the speed of stirring.
lence were independent of the geometry and depended only on Laufhutte and Mersmaan (1985) have measured the energy dis-
the local energy dissipation rates. sipation rates with the help of LDV. The experiments were carried
Günkel and Weber (1975) have investigated the turbulence out in a 0.19 m diameter stirred reactor equipped with a Rush-
characteristics of Rushton turbines in a 0.45 and 0.91 m diameter ton turbine. The rate of energy dissipation was calculated using
stirred vessels. The working fluid was air, and the measurements Equation (13), using the RMS velocity and the impeller diameter
were made with the help of hot wire anemometry. They have as the velocity and the length scale, respectively. The value of the
observed that the turbulence intensity in the impeller discharge constant A was varied so that the integration of the local ε val-
stream was independent of the impeller speed. The energy spec- ues matches the power input calculated from the impeller power
trum exhibited two slopes with the magnitudes −5/3 and −7, number. In this manner the value of the constant C was found to
respectively, in the impeller discharge stream as well as in the be 6. As seen from Equation (14), the ε values are inversely pro-
bulk of the vessel. This indicates that eddies of all scales are portional to the length scale. Since the length scales were taken to
present in the impeller discharge stream as well as the bulk of the be same as impeller diameter, the value of the constant C also had
vessel. The rate of flow of energy was calculated through all the to be increased substantially, in order to obtain the overall energy
faces of the impeller. From these calculations, it was observed that balance. The contour plots of the ratio of local energy dissipation
very little energy dissipation occurred in the impeller region. The rate to the average energy dissipation rate were also given. It was
integral scale of turbulence was calculated using Equation (14). observed that the value of this ratio was very high in the vicinity
This was used to compute the local turbulent energy dissipation of the impeller (close to 8), and it decayed gradually to about 0.1
rates. The local energy dissipation rates were integrated to find near the liquid surface and the bottom of the stirred vessel. This
the total energy dissipated in the stirred vessel. The power num- means that the local energy dissipation rate varies by a factor of
ber was predicted using the total energy dissipation rate. It was about 80 in the stirred vessel.
observed that the power number so predicted was about 20–30% Placek et al. (1986) have carried out CFD simulations of flow
less than that reported in the literature. Nishikawa et al. (1979) generated by Rushton turbines in stirred tanks. Their CFD results
have computed the energy spectrum for stirred reactors with diam- indicate even a wider range in the predicted values of ε. The ratio
eters in the range of 0.15–0.60 m and equipped with Rushton of the ε near the impeller tip to that in the bulk was found to about
turbines. The local velocity measurements were made with the 300. However, they have not compared the total energy dissipated
help of HFA. Using the velocity measurements, the energy spec- by integration of ε with the power drawn by the agitator.
trum was calculated. It was observed that, at a particular point in Costes and Couderc (1988) have investigated the turbulence
the impeller discharge stream, the dimensionless energy spectrum characteristics of stirred reactors having 70 and 200 L as working
was found to be independent of the impeller discharge stream. The volume. The impeller used was a standard Rushton turbine. The
dimensionless energy spectrum was found to be strongly depen- local velocity measurements were made with the help of laser
dent upon the distance away from the impeller, and hence it was Doppler velocimeter. Using these measurements the turbulence
concluded that the turbulence was anisotropic in the stirred ves- intensity was calculated in the entire vessel. The instantaneous
sel. They have also calculated the Taylor microscale of turbulence velocity measurements were used to compute the auto-correlation
using the energy spectrum. It was observed that the length scale function and the energy spectrum. The integral time scale was
was much higher in the impeller discharge stream and reduced calculated using Equation (15). The integral length scale was
dramatically by about five times away from the impeller region. calculated using the mean velocity. The Taylor microscale was
This was because the length scale was itself obtained from the calculated from curvature of the auto-correlation function. The
energy spectrum. As a result the regions of low values of turbu- dimensionless power spectrum at the same dimensionless dis-
lence intensity also showed low values of the length scale. van tance away from the impeller were found to be independent of
der Molen and van Mannen (1978) have used LDV to measure the speed of rotation and the size of the vessel. The power spec-
the turbulence characteristics of stirred vessels having diameters trum showed peak corresponding to the blade passage frequency,
in the range 0.12–0.90 m. They have observed that the dimension- indicating that the velocity–time data contained a periodic com-
less mean radial velocity in the impeller discharge stream decayed ponent. The local energy dissipation rate was calculated using
with an increase in the radial distance away from the impeller as Equation (14) with the help of RMS velocity and the impeller
x−7/6
1 . The power spectra was calculated by taking Fourier trans- diameter. It was observed that the ratio of the local energy
form of the cross correlation of the two velocity components. The dissipation rate to the average dissipation rate (ε/εavg ) varied sub-
power spectra for three different vessels were compared at the stantially in the vessel. Near the impeller tip, the ratio was about
same dimensionless distance in the impeller discharge stream and 5–10, whereas, away from the impeller in the vessel it reduced
at the same level of power consumption per unit volume. It was to about 0.05. This means that the energy dissipation rate varies
observed that in the high wave number region, all the three spec- by a factor of 100–200. The value of the constant A in Equation
tra were identical. This is similar to the observations made by (14) had to be made to about 4.42 in order to match the integra-
Komasawa et al. (1974). tion of the local energy dissipation rate with the impeller power
Okamoto et al. (1981) have measured the energy dissipation number. This means that the C values range from 2.0 to 6.8. This
rates for stirred vessels having diameters 0.15–0.60 m. The local represents nearly a three times variation in the value of C. Since

| 64 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


they have also used the impeller diameter to calculate the length fluid (pseudoplastic) was used, the profile of energy dissipa-
scales, the value of C also deviated greatly from unit like Laufhutte tion rate were found to be dramatically altered for all the three
and Mersmaan (1985). impellers.
Wu and Patterson (1989) have investigated the turbulence Schäfer et al. (1997) have investigated the flow generated by a
characteristics of 0.27 m diameter stirred vessel equipped with Rushton turbine operating in 0.15 m diameter tank with√ the help
a Rushton turbine. The measurements were made with the help of LDV. The characteristic velocity scale was taken as k, and the
of a LDV. The instantaneous velocity measurements were used volume averaged velocity was computed as:
to calculate the auto-correlation function. The turbulence inten-
sity was found to be independent of the stirrer speed. The energy k 3/2 dV
u =
3
spectrum function was also computed and it showed two peaks (21)
VL
corresponding to the blade passage frequency and two times the
blade passage frequency. The integral microscale was computed The average length scale in the reactor was computed using
from Equation (15), and the length scale was calculated by multi- Equation (14) with the measured value of average energy dissi-
plying with mean velocity. The Taylor microscale was calculated pation rate. The average length scale calculated in this manner
from the curvature of the auto-correlation function. The local was found to be about D/8.7, more specifically; it was found to
length scales were found to be in the range of one-half of the be about half the impeller blade width. This is in good agreement
blade width. The local energy dissipation rates were calculated with the results of Kresta and Wood (1993).
using Equation (13) with the help of turbulent kinetic energy and Wernersson and Tragardh (1998) have investigated the turbu-
Lres is the resultant microscale given as (L21 + L22 + L23 )0.5 . The local lence characteristics in a 0.8 m diameter tank equipped with a
energy dissipation rate was found to vary by a factor of 20–30 Rushton turbine. The turbulence characteristics were measured
within the stirred vessel. Since the estimated length scales were with the help of a constant temperature anemometry (CTA). The
of much smaller as compared to impeller diameter, the value of energy spectrum was observed to be independent of the speed
the constant C had to be kept equal to 0.85 (unlike Laufhutte and of rotation. The spectrum also showed peaks corresponding to
Mersmaan, 1985) in order to get good match between the total the blade passage frequencies. The local energy dissipation rate
energy dissipated by integration of the local ε values and that cal- was calculated from Equation (19), and the value of A was taken
culated from the impeller power number. From the local ε values as 0.47. It was observed that the energy dissipation rates in the
they have also concluded that about 60% of the total energy was impeller discharge stream were about 200–400 times those in the
dissipated in the impeller discharge stream and the remaining was bulk.
dissipated in the rest of the vessel. Wernersson and Tragardh (1999) have extended their mea-
Kresta and Wood (1991) have carried out CFD simulations of surements for 1.88 and 2.1 m diameter stirred vessels. The 0.8,
flow generated by a Rushton turbine. The CFD predictions of ε in 1.88, and 2.1 m diameter vessels were equipped with two three
the impeller discharge stream agreed well with the experimental and four Rushton turbines, respectively. Based on these mea-
measurements of previous workers. Based on their CFD simula- surements they have concluded that the values of the energy
tions, it was concluded that about 50% of the energy is dissipated dissipation rate in the impeller discharge stream could be scaled
in the impeller region and the rest is dissipated in the bulk. with the help of impeller diameter and the cube of the impeller
Kresta and Wood (1993) have briefly reviewed the various tip velocity. Using these scaling parameters, the values of εD/U3tip
methods used to estimate the energy dissipation rate. They have were found to be proportional to (1/x1 ) in the impeller discharge
also experimentally investigated the turbulence characteristics stream.
produced by a pitched blade turbine operating in a 0.15 m diame- Jenne and Reuss (1999) have employed k − ε model and its vari-
ter stirred vessel. The instantaneous velocity was measured with ations for the prediction of flow produced by a Rushton turbine.
the help of a LDV. They have observed that the Reynolds shear They have observed that the prediction of ε profiles were very
stress measured with the help of LDV are much smaller in mag- sensitive to the type of k − ε model employed. The ratio εmax /ε̄
nitude as compared with the normal stresses (turbulent kinetic was found to vary between 43 and 15 depending on the type of
energy). The turbulent energy dissipation rate was calculated k − ε model employed. The average length scale in the vessel was
from various methods reported in the previous literature. It was computed as:
observed that the methods based on the turbulent kinetic energy

(as characteristic velocity scale) and mean velocity (to calculate A1 k 3/2
the length scale) gave very large values of ε. Whereas, the meth- L= √ dV (22)
3 ε
ods based on the auto-correlation function gave very low values VL
of ε. The other methods which give values of ε between these two
extremes seem to work well. More specifically, the estimation of ε The ratio of the average length scale to the blade width was
based on the turbulent kinetic energy and the length scale equal found to vary between 0.62 and 0.26 for the various k − ε models
to one-tenth of the impeller diameter was recommended. employed. No attempt was made to compare the total energy dis-
Mavros and Baudou (1997) have investigated the turbulence sipated from integration of the local ε values to the power input
characteristics produced by propeller, Rushton turbine and A310 to the stirrer.
impeller. The measurements were carried out with the help of Sheng et al. (2000) have employed PIV technique to estimate the
LDV. The turbulent energy dissipation rate was estimated from turbulent energy dissipation rates. The PIV window was selected
Equation (14). The characteristic velocity scale was taken equal so as to identify eddies within the inertial sub-range. Since the
to RMS velocity and the length scale was obtained from the inte- eddies in this range are in a state of equilibrium (they receive
gration of the auto-correlation function (Eulerian time scale). energy from the higher scale and transfer it to the smaller scale
The ratio of ε to the average energy dissipation rate was found at about the same rate), the turbulent energy dissipation rate
to vary from 0.2 to 40 for Rushton turbines, 0.1 to 20 for pro- was estimated as the rate of turbulent kinetic energy transfer to
peller and 0.1 to 100 for A310 impeller. When a shear thinning these scales. The contour plots of turbulent energy dissipation

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 65 |


rates were also given. The estimated values of ε were then inte- tain smaller eddies. However every eddy has its own characteristic
grated to get the total energy dissipation rate in different regions velocity, life, and size and is not affected by its location in the large
of the reactor. It was observed that out of the total energy sup- size eddies. In other words, after isolation of an individual eddy,
plied by the impeller, 30% is dissipated within the impeller swept the large eddy motion has no net effect on the local characteristics
region, and about 50% is dissipated in the impeller discharge of the small eddies. Eddies of same size need not have the same life
stream. (time taken by it to cross the measurement volume) as it is decided
Ng and Yianneskis (2000) have employed the CFD model using by the surrounding net liquid motion and the eddy shapes. Indi-
SM technique for the prediction of the turbulent energy dissi- vidual eddy has no effect on the characteristic parameters (viz.
pation rate for a Rushton Turbine. They have reported that, out size, life, dissipation nature, etc.) of the neighbouring eddies in
of the total energy given by the impeller, 12% is dissipated in the time scale. For these eddies, the characteristics/dynamics can
the impeller swept region, 31% is dissipated in the impeller dis- be understood by analysing the instantaneous velocity–time data
charge stream, and remaining 57% is dissipated in the bulk properly. Whenever an eddy passes through the measurement
region. point, because of the local turbulence the velocity fluctuates about
Patwardhan (2001) has developed a new model to predict the a mean value. Subtracting the mean value, we get the fluctuating
flows generated by impellers based on the blade geometry and the part of the velocity. Fluctuating velocity has positive and neg-
speed of rotation. He has evaluated the total power imparted to ative values and it crosses the time axis at many locations. As
the liquid in three ways: (i) by integration of the local ε values mentioned earlier, after subtracting the mean value, the fluctu-
predicted from CFD simulations, (ii) based on angular momentum ating part changes from positive to negative or vice versa and it
going out of the impeller zone, and (iii) based on the forces acting indicates that all the fluid elements that are accounted during the
on the impeller blades. This has been done for a wide variety of measurement have the same averaged kinetic properties. This can
impeller designs varying in blade angle, blade twist, blade width, be considered as the existence of an eddy. Therefore, in the plot of
impeller diameter, etc. In all the cases, the power delivered to the fluctuating velocity (ui  ) versus time as the velocity vector crosses
liquid computed by the above three methods was compared with the time axis, one eddy ends and the other eddy enters the vol-
the impeller power number, and good agreement was observed in ume of measurement. It has been proposed that the time gap ( t)
all the cases. between the two successive crossings represents one eddy. The
Baldi et al. (2004) had used 2-D, 3-D PIV and four channel LDV characteristic velocity scale is estimated as square root of turbu-
to measure the dissipation rate using disc turbine. The measure- lent kinetic energy of that eddy and length scale is estimated as
ment were carried out in vessel of T = 100 and 294 mm diameter the ratio of velocity and time scale. The turbulent kinetic energy
and impeller diameter D = T/3. Dissipation rate was measured dissipation rate (ε) is estimated as the ratio of cube of velocity
at Re = 40 000 directly from the measurements of Reynolds stress scale to length scale.
gradients by analysing PIV images over interrogation areas down
to 0.1 mm. Similar LDV data was obtained with a resolution of
around 50 ␮m. Comparison of the energy dissipation rate predictions
as a result of various impeller models
Comparison of Energy Dissipation Rate At x2 /R = 0.586, IBC resulted into severe underpredictions of the
energy dissipation rate with either of the turbulence models. Only
In view of the above discussion, the present work is directed
RSM could yield closer comparison with MRF for x1 /R > 0.5 while
towards addressing the following issues: (i) comparison of the ε
it gave severe underprediction in rest of the radial region. All the
values predicted by CFD simulations with the ε values estimated
turbulence models failed to predict the energy dissipation rate
from in-house LDV data and processing using the eddy isola-
using SM approach. Further, the turbulent energy dissipation rates
tion strategy; (ii) distribution of energy delivered by the impeller
were severely underpredicted (maximum deviations ∼300%) by
in different regions of the vessel as predicted by CFD simula-
snap shot approach throughout the radial region. At x2 /R = 0.24,
tions; (iii) comparison of the impeller power number predicted
it can be seen from the poor prediction (severe underprediction)
by integration of the ε values, with the experimentally observed
of the energy dissipation rate by all the turbulence models using
power number. All the above three issues will be addressed for
IBC approach. Both the turbulence models (STD k − ε model and
different impeller modelling approaches and different turbulence
RSM) with MRF resulted into severe underprediction throughout
models.
the radial region. While SM was found to give the poor compar-
ison using both the standard and RNG k − ε turbulence models.
Estimation of energy dissipation rate using LDV data Further snap shot approach severely overpredicted (x1 /R < 0.35,
For the estimation of energy dissipation rate using LDV data, eddy by as much as 250%) the turbulent energy dissipation rate check
isolation methodology has been used. In this strategy, it is hypoth- plot. At x2 /R = −0.24, IBC with the RSM yielded closer predictions
esised that eddies are a group of fluid elements, which have the throughout the radial region. While standard and RNG k − ε mod-
same hydrodynamic properties restricted to its spatial location at els resulted in overpredictions in the region with x1 /R > 0.5. MRF
a particular instant. For the identification of eddies, Luk and Lee could not predict ε and gave negligible values of using any of the
(1986) approach is adopted. turbulence models used. The energy dissipation rate predictions
Instantaneous velocity–time data obtained using LDA at vari- were closer for 0.45 < x1 /R < 0.55 with RNG k − ε model while
ous spatial locations is subjected to the analysis and following underpredictions was the result with standard k − ε turbulence
assumptions are made while processing the data. Whenever a sin- models in rest of the radial region. Further, the predicted energy
gle eddy is passing through the measurement volume, it maintains dissipation rates as result of snap shot approach showed lower
its existence till it passes completely. In other words, no breakage values (less by ∼100%) for x1 /R < 0.6 while more (by ∼100%)
of an eddy takes place while it is present in the measurement vol- for x1 /R > 0.7 as compared to the experimentally estimated val-
ume formed at the intersection of the beams. Very large eddies ues. At x2 /R = −0.933, the simulations with IBC using any of the
(generally of the order of the inner cylinder diameter) may con- turbulence models resulted into prediction of negligible values

| 66 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


dissipation rate throughout the radial co-ordinate while all ZON
k − ε model resulted into overpredictions up to x1 /R = 0.2. At
x2 /R = −0.24, it can be seen that NEW k − ε predicted good ε
comparison for x1 /R > 0.5 followed by ZON and OPT-CK k − ε
models while it underpredicted in rest of the radial region. MOD
k − ε model yielded severe underpredictions of ε throughout. At
x2 /R = −0.933, all the k − ε models resulted in negligible values
of the energy dissipation rate. At x2 /R = 0.0, the turbulent energy
dissipation rate predictions were satisfactory as a result of all
the simulations with the variants of k − ε models. Magnitudes of
predicted ε were more or less for all the simulations.

Comparison of Power Numbers


Based on the overall distribution of the energy dissipation rate,
power number has been estimated and compared with exper-
imental data. To establish the energy balance, the total power
dissipation rate was calculated by volume integration of the pre-
dicted turbulent energy dissipation rate (ε) as:

2 H R
Utip
3
P= εx1 dx1 dx2 dx3 (23)
R
0 0 0

Experimentally, the total power consumption was measured by


a torque table (the accuracy of experimental measurements was
validated by measuring power consumption for a standard DT for
which the power number of five has been reported). From the pre-
dicted energy dissipation rate (Equation 23) and the experimental
power consumption (torque table), the values of power number
(NP ) were estimated using the following equation:

Figure 29. Comparison of energy dissipation rate at x2 /R = 0.0. Impeller:


disc turbine. (A) IBC:  experimental, line 1: STD k − ε model, line 2: P = NP N 3 D5 (24)
RNG k − ε model, line 3: RSM. (B) MRF:  experimental, line 1: STD k − ε
model, line 2: RSM. (C) SM:  experimental, line 1: STD k − ε model, line
2: RNG k − ε model. (D) SNAP:  experimental, line 1: STD k − ε model. Power numbers obtained by using Equation (24) as a result
of all the simulations of the flow generated in stirred tank are
reported in Table 7. The underpredictions were by about 40% for
of the turbulent energy dissipation rates. The energy dissipation variants of k − ε models when MRF was used. RNG k − ε model
rate comparison could be seen as poor as a result of MRF simu- coupled with MRF yielded only 5% underprediction, while RSM
lation using the standard k − ε model. All the turbulence models gave an overprediction of 45%. Power number predicted by vari-
employed resulted into insignificant energy dissipation rates when ants of k − ε models with SM were overpredicting by 48%, RNG
used with SM approach. The predicted energy dissipation values when coupled with SM gave 40% overprediction and RSM gave
due to snap shot approach were higher by two to three times than only 1% underprediction.
the experimentally measured one. At x2 /R = 0.0, IBC with both the
standard and RNG k − ε models yielded severe overpredictions ε
(Figure 29A) while it gave relatively closer comparison with RSM Reynolds Stresses
(flatter profile). Closer comparison was sought with MRF using Reynolds stresses are amongst the foremost importance charac-
RSM while standard k − ε model severely overpredicted the dissi- teristics, which govern the transport processes. Their distribution
pation rate (Figure 29B). SM was found to fail in predicting the ε in the stirred reactors is useful in the study of de-activation of
by each of the turbulence models (Figure 29C). Similar compar- enzymes/microorganisms in biochemical operations. Enzymes
ison was the result of computational snap shot approach (Figure activity is significantly depends on the shear stress distribution
29D). in the reactor. Some enzymes get deactivated in high shear flow
field. In case of solid suspension applications, to keep particles in
Discussion for the simulations with the variants of the state of suspension, turbulent dispersion forces are important
the k − ε model which in turn depend in Reynolds stresses. In case of mass trans-
At x2 /R = 0.586, the turbulent kinetic energy dissipation rates fer at gas–liquid and liquid–liquid systems, the true mass transfer
were closely predicted by OPT-CK k − ε model for x1 /R > 0.13 coefficient is dictated by the interface normal Reynolds stresses.
and x1 /R < 0.87 while it severely overpredicted ε in near ves- Higher Reynolds stresses lead to enhanced rate of surface renewal
sel centre and wall region. All other models were found to yield and increase in mass transfer rate. This section reviews the litera-
underprediction throughout the radial coordinate. At x2 /R = 0.24, ture as regards to the measurement of Reynolds stresses followed
all the k − ε models gave satisfactory comparison of the energy by assessment of the predictive capability of the CFD models.

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 67 |


Table 7. Effect of grid size on power number as a function of the turbulence model and the impeller modelling approach

Impeller Impeller model Turbulence model Grid size inimpeller swept region Grid size Np ( based) Np (ε based)

DT IBC k−ε — 85 000 — 2.71


DT IBC RNG k − ε — 85 000 — 2.24
DT IBC RSM k − ε — 85 000 — 1.83
DT MRF k−ε 57 888 461 496 4.43 1.35
DT MRF k−ε 42 240 311 328 4.38 2.71
DT MRF k−ε 38 720 293 040 4.34 5.38
DT MRF k−ε 28 800 194 432 4.27 0.98
DT MRF RNG k − ε 28 800 194 432 4.27 4.51
DT MRF RSM 42 368 311 328 4.32 2.34
DT SM k−ε 38 720 194 432 4.36 1.19
DT SM k−ε 38 720 293 040 4.30 1.48
DT SM k−ε 42 368 311 328 4.92 2.91
DT SM k−ε 57 888 461 496 1.89 1.59
DT SM RNG k − ε 42 368 311 328 4.92 2.91
DT SM RNG k − ε 57 888 461 496 5.25 5.35
DT IBC Zonal k − ε — 85 000 — 1.84
DT IBC MOD k − ε — 85 000 — 1.34
DT IBC NEW k − ε — 85 000 — 1.54
DT IBC OPT-CK — 85 000 — 2.63

Previous work of LDV. The Reynolds stresses were also estimated from the
Drbohlav et al. (1978) have investigated the turbulence charac- relationship:
teristics produced by a Rushton turbine in a 1.0 m diameter tank.  
2 ∂ui  ∂uj 
The Reynolds normal stresses and shear stresses were measured u i u j  = kıij + T + (25)
with the help of a pitot tube. The measured values of the Reynolds 3 ∂xj ∂xi
stresses and the mean velocity gradients were used for the calcu-
lation of the eddy viscosity. It was observed that the eddy viscosity The values of eddy diffusivity were estimated from the relation:
values reduce with an increase in the radial distance away from
the impeller tip. No attempts were made to obtain the distribution k2
 T = C␮ (26)
of the eddy viscosity in the rest of the vessel. ε
Mahoust et al. (1988) have measured the turbulence charac-
teristics of stirred vessels equipped with a Rushton turbine. The The values of k were measured and that of ε were taken as 15
vessel diameter was 0.2 m and the measurements were made with times the average energy dissipation rate. It was observed that
the help of LDV. They have made simultaneous measurements of the predicted Reynolds stresses showed the same trend as that
the radial and tangential velocity components, which enabled the observed experimentally. However, their magnitude was as much
calculation of the Reynolds shear stresses. It was observed that, as 50% higher or lower in the impeller discharge stream.
in the impeller discharge stream and close to the impeller tip, the Derksen and van den Akker (1999) performed three-
Reynolds shear stress u i u j  was about three times smaller than dimensional, angle-resolved LDV measurements of the turbulent
the normal stresses. Whereas, near the wall, it was found to be flow field (Re = 2.9 × 104 ) in the vicinity of a Rushton turbine in
about 10 times smaller than the normal stresses. a baffled mixing tank. They have presented results on the average
Mahoust et al. (1989) have established energy balance by flow field, as well as on the complete set of Reynolds stresses. The
computing the power transmitted from the torque calculations. It anisotropy of the turbulence has been characterised by the invari-
was observed that the power number based on torque was about ants of the anisotropy tensor. The trailing vortex structure, which
4.5, whereas, the experimentally observed power number was is characteristic for the flow induced by a Rushton turbine, is
close to 5. demonstrated to be associated with strong, anisotropic turbulent
Kemoun et al. (1994) have investigated the turbulent stresses activity.
produced by standard Rushton turbines. The experiments were Galletti et al. (2004) used 3-D LDV to measure Reynolds stresses
carried out in a 0.2 m diameter stirred vessel and the measure- and triple products in stirred tank with DT. They presented the pro-
ments were made with the help of a LDV. Their measurements files of ensemble-averaged normal and shear stresses throughout
have shown that the three components of the turbulent stresses the entire vessel. They observed that near the tank bottom and
as well as the shear stresses are not similar in shape and are not vessel axis, the three normal stresses differed significantly. In this
equal in magnitude. This indicates that the turbulence is highly region, radial-tangential shear stress was highest and the other
anisotropic. Also a comparison of the turbulent shear stresses and two being negligible.
the normal stresses shows that the former are smaller by as much
as five times than the latter. Comparison of Reynolds stresses
Derksen et al. (1997) have measured the Reynolds stresses in Extensive comparison of predicted Reynolds stresses by all the
the stirred vessel equipped with a Rushton turbine. The mea- simulations with the experimental data of Kumar (1997) was
surements were made in a 0.3 m diameter tank with the help carried out. Figure 30 shows the comparison for the predicted

| 68 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


trially (even for the single-phase system). In industrial
practice one often comes across multiple impeller systems
consisting of impellers of different types, different D/T
ratios, and placed at various locations within the reactor.
Industrial stirred reactors often use internals like draft tubes,
cooling coils, nozzles, etc. Further, the physicochemical
properties of process stream can vary significantly. It is
important to predict the flow field under such a wide variety
of situations mentioned above and compare the CFD model
predictions with the experimental measurements of mean
velocity field and turbulence properties. Only after such
comparisons have been made, conclusions can be drawn
about the predictive capability of the various turbulence
models and the impeller modelling approaches.
(3) It can be concluded from the Table 4 that the CFD models
have focussed on a limited region in stirred tank, which is
about 30% of the tank volume. Very little attention has been
paid to the accuracy of predictions near the reactor walls,
baffles, liquid surface and in-between the impeller blades.
Accurate prediction of flow characteristics in these regions
is important to predict the mixing characteristics, heat
transfer coefficients, bubble/drop breakage by impellers,
onset and rates of surface aeration and gas induction,
etc. A comprehensive program needs to be undertaken
to experimentally measure the mean velocity field and
turbulence characteristics in the regions mentioned above
and compare them with the CFD model predictions. Fur-
ther, the validation of the predictions of the CFD models
should be carried out by rigorous quantitative comparison.
Qualitative comparisons in the form of contour plots may
not lead to any conclusion as regards to the accuracy of
predictions. In addition, one should not compare the results
of improvement in impeller/turbulence models by mere
comparison with the predictions of the other models but
with the accurate experimental measurements.
(4) As mentioned earlier, the working fluid has been water. It
Figure 30. Comparison of Reynolds stresses. Impeller: disc turbine. (A)
is important to cover a very wide range of Re and for all the
Comparison for u  1 u  2  (B) comparison for u  2 u  3 .  Experimental (1): impeller designs. Urgent attention is needed towards flow
at x2 /R = 0.586; (2): at x2 /R = 0.24; (3): at x2 /R = −0.24; (4): at visualisation and fluids over a wide range of characteristics.
x2 /R = −0.933. Future studies need to include the investigation of close
clearance impellers.
(5) Though some attempts (Lee and Yianneskis, 1997; Schäfer
Reynolds stresses, namely u 1 u 2  (Figure 30A1–A4) and u 2 u 3  et al., 1997, 1998, etc.) have been made to measure the flow
(Figure 30B1–B4), as a result of a simulation by IBC approach characteristics in the impeller region, their scope has been
using standard k − ε model at all the four axial locations. It can limited to the standard impellers and geometries. There is
be seen that the comparison was poor both qualitatively and a need to carry out reliable and extensive measurements of
quantitatively. Similar observations were noted for the predic- angle resolved velocity components and quantification of
tions of Reynolds stresses due to other simulations (comparison the random and the periodic components of the turbulent
is not shown here). The discrepancy in the comparison may be kinetic energy. Further, accurate measurements with specific
attributed to the errors in experimental measurement of Reynolds emphasis on the impeller swept region and close to the wall
stresses and/or turbulence modelling. are required. The reliable measurements of the flow charac-
teristics should be carried out near wall/baffle region, near
CONCLUSIONS AND SUGGESTIONS FOR FUTURE liquid surface and in the geometries with internals (such as
draft tubes, cooling coils, etc.). Further, accurate measure-
WORK ment techniques for the energy dissipation rates need to be
(1) From the extensive literature presented, it can be concluded developed. All these need to be carried out for a variety of
that, most of the CFD studies have focussed on the mod- impeller types with different geometric configurations. All
elling of flows generated by two types of impellers, namely the experimental data should be processed to quantify error
Rushton turbine and the pitched blade turbines in standard bars and confidence limits. The reliable sets of experimental
geometries. Further, most of these studies have been limited data are expected to be useful for the further validation and
to water as the working fluid. subsequent development of CFD models. For the good exper-
(2) There have been practically no attempts to simulate wide imental practices the reader is referred to papers by Ng et al.
variety of stirred reactor configurations encountered indus- (1998); Baudou et al., (1997); Schäfer et al. (1997, 1998).

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 69 |


(6) As regards to the grid size, from the comparison presented may become transitional and even laminar with increasing
in this work, it can be concluded that though complete distance from the blades. Hence the MI frequency must
grid independent solution is difficult, one can reasonably be investigated extensively to find its effect on mixing.
succeed in getting grid size, which can yield the solution Finally, the CFD models need to focus in future on the
with a fair accuracy. Refined grid near the impeller blade, in design objectives, namely, mixing time, heat transfer
the region near the impeller tip, near vessel walls and baffles coefficients, critical impeller speed for the gas dispersion,
should be used and the effect should be quantified. Further, flooding/loading transitions, critical impeller speed for solid
it is desirable to use higher order differencing scheme to suspension, gas hold-up, bubble/drop size, mass transfer
reduce numerical errors to the extent possible. Once these coefficients, etc.
issues are resolved, one can comfortably proceed for the (9) Murthy and Joshi (2008) measured three-dimensional mean
validation of CFD models. The present impeller and turbu- flow field and the turbulent kinetic energy of the baffled
lence models are capable of predicting the overall flow both stirred vessel agitated by five impellers (DT, PBT (60◦ , 45◦ ,
near and away from the impeller for standard geometries and 30◦ ) and a hydrofoil) have been measured using LDA.
only qualitatively. They are capable of predicting the gross They performed three-dimensional CFD simulations by the
flow features produced by the single and multiple impellers standard k − ε, the Reynolds stress transport and LES tur-
(parallel, merging and parallel flows as a result of change in bulence models for all the above cases. The pressure strain
the inter-impeller clearance between the Rushton turbines terms in RSM was modelled based on the proposal in the lit-
(in the presence and absence of cooling coils); transition erature that is suitable for the present flow configuration. In
of double to single loop due to decrease in the clearance case of LES, for the first time, one equation dynamic subgrid
between the impeller and the vessel bottom and the trailing scale model has been successfully employed for the stirred
vortices behind the rotating blades). However, they lack vessel geometry. The predicted flow fields by all the three
significantly in quantitative predictions, particularly the turbulence models were comprehensively compared with
tangential velocity and the turbulent kinetic energy. the experimental data. The following conclusions can be
(7) From the extensive comparison presented in fifth section, drawn from their work. (a) As for as mean flow predictions
it can seen that the quantitative comparison of axial and are concerned, the RSM performed better than the standard
radial velocities in the bulk region is good with the IBC k − ε model when comparisons were made with DT. The
approach while other impeller modelling approaches result results reaffirm that the RSM can out-perform the k − ε
into very poor comparison for the mean and turbulence model when recirculations are present. This is due to the
quantities. Due to the requirement of the experimental data, overestimation of the eddy viscosity which is a general char-
IBC approach cannot be used for the screening of a variety of acteristic of the k − ε model. In general, it can be concluded
impeller designs in different geometric configurations. This that the standard k − ε model performs well when the flow
can be avoided if impeller rotation is modelled properly. In is unidirectional that is with less swirl and weak recircula-
this regard SM approach offers a promising alternative for tion. (b) Both the standard k − ε model and anisotropy RSM
the actual rotation of the impeller. One needs to critically fail to predict the turbulent kinetic energy profiles in the
examine the effects of impeller design and other internal impeller region when the flow is dominated by the unsteady
mountings for the flows in stirred vessel involving strong coherent flow structures. (c) Using both the LDA and LES
interaction between the impeller, vessel wall and the baffles tools, the strength of the precessing vortex instability has
where the vessel boundary has a significant effect on the been quantified for all the five impeller designs. It has been
flow generated by the impeller. observed that the impeller which generates strong swirl flow
(8) The flow patterns encountered in the stirred vessels are of generate equally strong was the main reason for such insta-
complex type that depends on various operating conditions bilities. Therefore, among the impeller designs considered
and geometry. An efficient design of stirred vessel requires in the present study, the DT produces strongest instabilities
knowledge of the hydrodynamics of the flow pattern, and the HF generates the weakest instabilities. (d) The fre-
power consumption, and mixing time. Generally, overall quency of the jet instability was found to be linearly related
design parameters can give only the global picture but the to the rotational speed (∼0.13–0.2 N, Hz) of all the impeller
local information on hydrodynamics is mandatory for the designs under consideration. The energies associated with
reliable design of reactors. Such local parameters are very the jet instability for all the four impellers, that is, DT,
much affected by the temporal intermittent flow called PBTD60, PBTD45, PBTD30, and HF, have been found to
macroinstabilities (MIs). MIs are complex spatiotemporal be 42 m2 /s2 , 38 m2 /s2 , 0.64 m2 /s2 , 8 m2 /s2 , and 12 m2 /s2 ,
phenomenon spanning a large portion of the reactor. respectively. (e) Further, consistent occurrence of inter-
These are low frequency in nature with high amplitude mediate instabilities have been observed having frequency
of oscillatory motions in the low turbulence regions of a 0.04–0.07 N which lies in between the precessional and the
vessel which have the capability of transporting substances jet instability. The origin of these kinds of instabilities could
fed to a mixing process over relatively long distances. The be due to the interaction of precessing vortex instability
magnitude of the MIs varies with the location in the vessel with either the mean flow or jet/circulation instabilities.
and these low frequency motions can severely affect the (f) Having demonstrated that the LES model provides pre-
transport phenomena (local heat and mass transfer rates, dictions that agree well with the measurements, and these
gas/solid holdup, etc.). Apart from the mixing operation, simulations capture relatively many flow features, these
these frequencies at higher amplitude can affect on the solid simulations can be further be utilised to explore a variety
surfaces inside the reactors causing permanent damage for of issues with a reasonable confidence. (g) Energy balance
baffles, internals, sensors, etc. At low Reynolds number, has been established using both integral ε and torque-based
the MI peaks are more pronounced near the impeller. Even approaches. It can be concluded that the torque-based
if the flow near to the impeller is turbulent (high NRe ), it approach seems to be more promising for the estimation

| 70 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


of power number for a new impeller design and using by more sophisticated techniques (such as three-dimensional
computationally economical RANS based simulations. electrical resistivity technique), such kind of measurements
(10) As discussed in ninth section, there are a number of have now become possible.
methods of estimation of energy dissipation rate (ε). These (14) Various studies have been reported as regards to the effect
methods stem from the fact that there are a large number of impeller design on the mixing time and the power
of ways of defining the characteristic velocity scale and consumption. Some of the conclusions drawn suggest that
the characteristic length scale. However, any method has the mixing time is function of impeller design while some
to satisfy the overall energy balance. That is, the volume- lead to the same mixing time irrespective of impeller design
integrated rate of energy dissipation must be equal to the at the same power consumption. A critical analysis has
energy supplied by the impeller. Any method of estimation been presented which shows that the liquid phase mixing
of ε that does not satisfy the overall energy balance cannot in turbulent flow regime is the flow controlled process and
be considered to be realistic. Studies need to be undertaken hence depends on the impeller design.
to check, which methods of estimation of ε satisfy the over- (15) Most of the studies indicate that the axial flow impellers
all energy balance and which methods do not. The energy in general and marine propellers in particular are superior
supplied by the impeller can also be calculated from the in mixing performance as compared to the radial flow
outflow of angular momentum and the torque acting on the impellers as they require smaller amount of energy to
impeller blades. The energy supplied using these methods produce vigorous fluid circulation throughout the tank.
also needs to be compared with the volume integrated ε. (16) Experimental data on the mixing time have been analysed on
(11) A quite a few good measurement techniques have been the basis of equal power consumption. It has been observed
reported in the literature. Simple measurement techniques that the mixing time varies inversely with the cube root of
(visual and conductometric) though give quick estimates of the power consumption per unit volume of the liquid. The
mixing times; they suffer from the accuracy of measurement constants in the proportionality relationship are found to
while the most sophisticated techniques (e.g., 3-D ERT) are be impeller design dependent. Also they are found to be the
expensive and require a certain kind of expertise. In this function of the range of operating parameters investigated.
regards radioactive liquid tracer technique presents a good (17) The second most important parameter to have the major
alternative because of its non-intrusiveness, applicability impact on the mixing performance has been found to be
in case of non-transparent vessels and its stability at severe the impeller diameter. The momentum flux of the fluid
operating conditions. discharging out of the impeller increases with an increase in
(12) Most of the extensive studies reported in literature make impeller diameter leading to lowering of circulation times.
use of the conductivity measurements in a vessel after the Further, it also affects secondary flow number. All these
addition of tracer technique due to its simplicity. Extensive ultimately result into the lower mixing times as a result of
studies have been reported to quantify the effect of various an increase in impeller diameters. The mixing time varies
parameters, namely the location of tracer addition, volume inversely with the square of the impeller diameter.
of tracer, time of tracer addition, location and size of the (18) In the work of Patwardhan and Joshi (1999), it was proved
detection probe. Further, in case of visual assessment of that blending application prefers all the energy in the form
mixing, it has been suggested that the indicators in an of mean and minimal turbulent kinetic energy (even <5%).
acidic pH range should be used as the indicators in alkaline In contrast, colloidal mills, homogenisers and emulsifiers
pH range suffer from the time lag. need highly turbulent flows. The ultrasound reactors are the
(13) The foremost important parameter in the evaluation of the types which still puzzle, even though it has important appli-
mixing performance is the definition of the mixing. Most cations at specific chemical/biochemical process industries.
of the investigations define the mixing time as the time Hence, the acoustic streaming and associated turbulence
required to achieve a concentration/temperature of 95% characteristics of the flow governing the rate of the transport
of the final average concentration/temperature. It has been processes, the detailed understanding of the fluid mechanics
reported that the time required for attainment of certain (velocity and stress fields, turbulence structure, distribution
percentage approach to the final average concentration of energy dissipation rates, etc.) must be investigated and
depends on the impeller design. In view of this it is cross check with the postulate of Patwardhan and Joshi
suggested to outline a unique mixing criterion which can (1999) on the mixing front. All the previous work reported
be a correct measure of uniformity and invariant of the have used either intrusive measurement techniques or
impeller-vessel configuration. In this regard the criterion visual observations techniques like streak photography to
suggested by Harnby et al. (1991) can be seen as the estimate the velocity magnitude prevailing in the ultrasonic
promising one. This criterion is based on the variance of reactors. However, no previous work has yet reported about
the concentration measurements at number locations: the turbulence levels prevailing in such reactors.
 (19) For a given type of impeller, the parameters such as blade
1
n

2 = (Ci −C0 )2 number, blade width, blade height, disc thickness (in case
n−1 of DT), have been found to have a profound impact on
i=0
the overall flow pattern, power number of impeller and the
The value of the variance is representative of the extent mixing time. It has been reported that an increase in the
of the overall mixing and should be decided as per the blade width can be more energy efficient as compared to
desired extent of mixing. This criterion has scarcely been the increase in impeller diameter. Further, the number and
used in the past investigations. This may be due to the dimensions of the impeller blades (thickness, in particular)
fact that it requires a number of measurements of the can be of great help in the optimisation of the mixing pro-
concentration at number of points in order to eliminate cesses. Further, in view of this a great deal of care should be
the effect on variance. With the advent/development of taken while comparing mixing performance of same and/or

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 71 |


different impellers types from the published literature as the impeller-vessel configurations. For such studies the work of
difference in mixing time may also arise due to the difference Patwardhan and Joshi (1999) can be treated as a guideline.
in blade and disc thicknesses, hub and shaft diameters, etc. (24) An extensive CFD studies in conjunction with the LDA
(20) The internals like the baffles and draft tube have been measurements of the flow pattern suggest that the con-
found to affect the flow pattern and hence the mixing vective motion controls the mixing process to a large
performance. The studies have been reported which guide extent in the turbulent regime. Further, the relationship,
the selection of the number, width, and the positioning of  mix ∝ N1/3
P T
2/3
/NQS which relates the mixing time to both
the baffles. Further, few studies indicate that the use of draft the power number and secondary flow number is suggested
tube for the better mixing operation. for the evaluation of the mixing performance of alternative
(21) Position of the impeller with respect to the vessel bottom impeller designs on the basis of equal power consumption
(off bottom clearance) and the location of the impeller shaft per unit mass.
relative to the central axis (impeller eccentricity) have got a (25) The results of CFD studies indicate that there is enormous
significant impact on the flow pattern and the mixing perfor- scope to modify the local values of mean flow and eddy
mance thereof. It has been suggested to use higher impeller viscosity to achieve substantially higher levels of energy
off-bottom clearance in case of the axial flow impellers while efficiency in the mixing process which promises substantial
low clearance for radial flow impellers in order to achieve savings in terms of operating cost.
energy efficient mixing operation. The eccentricity should
be chosen so as to maintain the equality of the distances
between the wall baffle and the impeller tip on one side and NOMENCLATURE
the impeller tip and vessel on the other side where single A empirical constant in Equation (14)
loop formation occurs with maximum flow passage through A1 empirical constant in Equation (22)
the impeller body. Lower or higher eccentricities have been C impeller clearance from bottom, m
found to have detrimental impact on the overall mixing C␮ , Cε1 , Cε2, Cε3 turbulence parameters in the k − ε model
performance. In addition, the eccentric impellers are more D impeller diameter, m
beneficial in case of unbaffled vessel. Further, the impellers dV differential volume in the reactor, m3
mounted on the angled shaft are found to perform better as E energy associated with wave number k1 ,
compared to the impellers mounted on vertical shaft. f frequency, s−1
(22) Many times, process engineers are required to evaluate the Gk turbulence production term in two-equation k − ε
performance of existing equipment, with a view to modify or model
improve its performance. In that case, a very quick estimate H liquid height in the vessel, m
of the mixing time may be required, by the time CFD mod- k dimensionless turbulent kinetic energy (with
elling takes a shape of a very common utility and becomes a U2tip )
short time activity. In such cases the correlations like the one k1 wave number as in Equation (18), (1)/m
proposed by Rewatkar and Joshi (1991c) (which was arrived kp kinetic energy of large vortices in multiple scale
after the analysis of mixing time data due to extensive exper- model
imentations in large sized vessels and over a wide range of kT kinetic energy of transfer eddies in multiple scale
parameters) will be useful. Further, such kind of correlations model
should be developed by using corrected data for the differ- L characteristic length scale, m
ence in operating range and impeller dimensions in addition Lres resultant microscale, m
to the errors in mixing time measurements. In addition it Li miscroscale in ith direction, m
is suggested to prepare and make use of comprehensive N impeller speed, s−1
worksheet like that proposed by Zlokarnik (2001) for the NQS impeller flow number
handy selection of the impeller-vessel configuration for a NP impeller power number
desired mixing time at a given size of operation based on the NDNS number of grids required for DNS in Equation (3)
reliable data due to the recent experimental and theoretical Pk turbulence production term, kg m−1 s−3
investigations with an emphasis on the high efficiency Pij production term, m2 s−3
impellers. p instantaneous pressure, N/m2
(23) In view of the complex interdependency of a wide paramet- p time-averaged pressure, N/m2
ric space (i.e., impeller design, impeller and tank diameters, p̄ filtered pressure, N/m2
impeller location, baffle configuration, draft tube, etc.) R radius of the vessel, m
screening for an optimum configuration using experimen- Re Reynolds number, ND2 /␮
tation becomes a mammoth and almost impossible task. Reh Reynolds number based on channel height and
In this regards CFD can be of great help in screening out mean velocity
various configurations (incorporating a variety of miscella- Sij strain rate tensor, 1/s
neous impeller designs and internals which are difficult to S source term of
fabricate and time consuming to study experimentally) for T vessel diameter, m
mixing process qualitatively and can significantly reduce t time, s
the experimentation. Further, CFD can be of great help in ui instantaneous velocity vector
order to quantify the contribution of the flow or the power Utip impeller tip velocity, m s−1
consumption and the role of impeller design. In view of this ū mean velocity, m s−1
it is suggested that the thorough analysis of the flow pattern uRMS r.m.s. velocity, m/s
and hence the mixing time should be carried out using the ui  time-averaged velocity in ith direction, m/s
detailed simulation of the mixing process with a variety of u i u j  Reynolds shear stress, m2 /s2

| 72 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


u fluctuating velocity component, m/s REFERENCES
VL volume of liquid in the reactor, m3
w baffle width, m Abujelala, M. T. and D. G. Lilley, “Limitations and Empirical
x 1 , x2 , x3 space coordinates in radial, axial, and tangential Extensions of the k − ε Model as Applied to the Turbulent
direction respectively, m Confined Swirling Flow,” Chem. Eng. Commun. 31, 223–236
(1984).
Alcamo, R., G. Micale, F. Grisafi, A. Brucato and M. Ciofalo,
Greek Symbols
“Large-Eddy Simulation of Turbulent Flow in an Unbaffled
ıij Kronecker delta Stirred Tank Driven by a Rushton Turbine,” Chem. Eng. Sci.
ε dimensionless turbulent energy dissipation rate (by 60, 2303–2316 (2005).
U3tip /R) Alopaeus, V., P. Moilanen and M. Laakkonen, “Analysis of
ε̄ average rate of turbulent kinetic energy dissipation, Stirred Tanks with Two-Zone Models,” AIChE J. 55,
m2 /s3 2545–2552 (2009).
εmax maximum rate of turbulent kinetic energy dissipation, Alvarez-Hernández, M. M., T. Shinbrot, J. Zalc and F. J. Muzzio,
m2 /s3 “Practical Chaotic Mixing,” Chem. Eng. Sci. 57, 3749–3753
eff effective diffusivity, m2 s−1 (2002).
transport property Anandha Rao, M. and R. S. Brodkey, “Continuous Flow Stirred
 von Karman constant Tank Turbulence Parameters in the Impeller Stream,” Chem.

t Taylor microscale, m Eng. Sci. 27, 137–156 (1972).


␮t turbulent viscosity, kg/m s Angeli, P. and G. Hewitt, “Drop Size Distributions in Horizontal
 kinematic viscosity, m2 /s Oil–Water Dispersed Flows,” Chem. Eng. Sci. 55, 3133–3143
t turbulent diffusivity, m2 /s (2000).
 density of liquid, kg m−3 Arjunwadkar, S. J., K. Saravanan, A. B. Pandit and P. R.
ε constant in two-equation k − ε model Kulkarni, “Optimizing the Impeller Combination for
k turbulent Prandtl number Maximum Hold-Up with Minimum Power Consumption,”
 integral time scale, s Biochem. Eng. J. 1, 25–30 (1998).
ω specific dissipation rate, s−1 Armenante, E. M. and D. J. Kirwan, “Mass Transfer to
Microparticles in Agitated Systems,” Chem. Eng. Sci. 44,
Superscripts and Subscripts 2781–2796 (1989).
1, 2, 3 axis indexes of space coordinates Armenante, P. M., Y. Huang and T. Li, “Determination of the
t turbulent Minimum Agitation Speed to Attain the Just Dispersed State
in Solid–Liquid And Liquid–Liquid Reactors Provided with
Multiple Impellers,” Chem. Eng. Sci. 47, 2865–2870 (1992).
Abbreviations Arratia, P. E., J. Kukura, J. Lacombe and F. J. Muzzio, “Mixing
AFT adaptive force field technique of Shear-Thinning Fluids with Yield Stress in Stirred Tanks,”
ASM algebraic stress model AIChE J. 52, 2310–2322 (2006).
CFD computational fluid dynamics Bakker, A. and L. M. Oshinowo, “Modelling of Turbulence in
CTA constant temperature anemometry Stirred Vessels Using Large Eddy Simulation,” Chem. Eng.
DNS direct numerical simulation Res. Des. 82, 1169–1178 (2004).
DSM differential stress model Bakker, A. and H. E. A. van den Akker, “Single-phase flow in
DT disc turbine stirred reactors,” Chem. Eng. Sci. 72, 583–593 (1994).
EFD experimental fluid dynamics Bakker, R. A. and H. E. A. van den Akker, “A Lagrangian
IBC impeller boundary condition approach Description of Micromixing in a Stirred Tank Reactor Using
IO inner outer 1D-Micromixing Model in a CFD Flow Field,” Chem. Eng. Sci.
LDV laser Doppler velocimetry 51, 2643–2648 (1996).
LES large eddy simulation Baldi, S., A. Ducci and M. Yianneskis, “Determination of
MDG moving-deforming grid technique Dissipation Rate in Stirred Vessels Through Direct
MOD modified Measurement of Fluctuating Velocity Gradients,” Chem. Eng.
MRF multiple reference frame technique Technol. 27 (3), 275–281 (2004).
MI macro-instability Barigou, M. and M. Greaves, “Gas Holdup and Interfacial Area
OPT-CK optimised Chen–Kim Distributions in a Mechanically Agitated Gas–Liquid
PBTD pitched blade downflow turbine Contactor,” Trans. Inst. Chem. Engrs. - A: Chem. Eng. Res.
PIV particle image velocimetry and Des. 74 (Part A), 397–405 (1996).
RANS Reynolds averaged Navier–Stokes Barigou, M. and M. Greaves, “Bubble-Size Distributions in a
RMS root mean square Mechanically Agitated Gas–Liquid Contactor,” Chem. Eng.
RNG renormalised group Sci. 47, 2009–2025 (1992).
RSM Reynolds stress model Bartels, C., M. Breuer and F. Durst, “Comparison between Direct
SGS subgrid scale Numerical Simulation and k−ε Prediction of the Flow in a
SM sliding mesh technique Vessel Stirred by a Rushton Turbine,” Proc. 10th Euro. Conf.
SNAP snapshot approach Mix. 239–246 (2000).
SS source-sink approach Bates, R. L., E. L. Fondy and R. R. Corpstein, “An Examination
SST shear stress transport of Some Geometric Parameters of Impeller Power,” Ind. Eng.
ZON zonal Chem. Process Des. Dev. 2, 310–314 (1963).

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 73 |


Baudou, C., C. Xuereb, J. Costes, G. Ranchin and J. Bertrand, Carreau, P. J., J. Paris and P. Guérin, “Mixing of Newtonian and
“Laser Doppler Measurements of Flow Field and Turbulent Non-Newtonian Liquids: Screw Agitator and Draft Coil
Flow Parameters in a Stirred Tank Equipped with Two System,” Can. J. Chem. Eng. 70 (6), 1071–1082 (1992).
Industrial Propellers,” Proc. 9th Euro. Conf. Mix. 11–18 Carreau, R. J., J. Paris and E. Guérin, “Heat Transfer to
(1997). Newtonian and Non-Newtonian Liquids in a Screw Agitator
Beckner, J. L. and J. M. Smith, “Anchor-Agitated Systems: Power and Draft Coil System,” Can. J. Chem. Eng. 72 (6), 966–974
Input with Newtonian and Pseudo-Plastic Fluids,” Trans. Inst. (1994).
Chem. Eng. 44, T224 (1966). Chapple, D. and S. Kresta, “The Effect of Geometry on the
Bertrand, J., J. P. Couderc and H. Angelino, “Power Stability of Flow Patterns in Stirred Tanks,” Chem. Eng. Sci.
Consumption, Pumping Capacity and Turbulence Intensity in 49, 3651–3660 (1994).
Baffled Stirred Tanks: Comparison Between Several Turbines,” Chen, Y. S. and S. W. Kim, “Computation of Turbulent Flows
Chem. Eng. Sci. 35, 2157–2163 (1980). Using an Extended k − ε Turbulence Closure Model,” NASA
Birch, D. and N. Ahmed, “The Influence of Sparger Design and CR-179204 (1987).
Location of Gas Dispersion in Stirred Vessels,” Trans. Inst. Chen, H. T. and S. Middleman, “Drop Size Distribution in
Chem. Eng. 75 (Part A), 487–496 (1997). Agitated Liquid–Liquid Systems,” AIChE J. 13, 989–995
Boden, S., M. Bieberle and U. Hampel, “Quantitative (1967).
Measurement of Gas Hold-Up Distribution in a Stirred Chen, L., Y. Bao and Z. Gao, “Void Fraction Distributions in
Chemical Reactor Using X-Ray Cone-Beam Computed Cold-Gassed and Hot-Parged Three Phase Stirred Tanks with
Tomography,” Chem. Eng. J. 139, 351–362 (2008). Multi-Impeller,” Chin. J. Chem. Eng. 17, 887–895
Bombač, A. and I. Žun, “Individual Impeller Flooding in Aerated (2009).
Vessel Stirred by multiple-Rushton Impellers,” Chem. Eng. J. Ciofalo, M., A. Brucato, F. Grisafi and N. Torraca, “Turbulent
116, 85–95 (2006). Flows in Closed and Free Surface Unbaffled Tanks Stirred by
Boon-Long, S., C. Laguerie and J. E. Couderc, “Mass Transfer Radial Impellers,” Chem. Eng. Sci. 51, 3557–3573 (1996).
from Suspended Solids to a Liquid in Agitated Vessels,” Colenbrander, G. W., “Experimental Findings on the Scale-Up
Chem. Eng. Sci. 33, 813–819 (1978). Behaviour of the Drop Size Distribution of Liquid/Liquid
Bourne, J. R. and H. Butler, “Power Consumption of Helical Dispersions in Stirred Vessels,” 10th Euro. Conf. Mix.
Ribbon Impellers in Viscous Liquids,” Trans. Inst. Chem. Eng. 173–180 (2000).
47, T263–270 (1969a). Conti, R. and S. Sicardi, “Mass Transfer from Freely-Suspended
Bourne, J. R. and H. Butler, “An Analysis of the Flow Produced Particles in Stirred Tanks,” Chem. Eng. Commun. 14, 91–98
by Helical Ribbon Impellers,” Trans. Inst. Chem. Eng. 47, (1982).
159–168 (1969b). Costes, J. and J. R. Couderc, “Study by Laser Doppler
Bowen, R. L., “Agitation Intensity: Key to Scaling Up Anemometry of the Turbulent Flow Induced by a Rushton
Flow-Sensitive Liquid Systems,” Chem. Eng. 92 (6), 159–168 Turbine in a Stirred Tank: Influence of the Size of the
(1985). Units—I. Mean Flow and Turbulence,” Chem. Eng. Sci. 43,
Brennan, D. J. and I. H. Lehrer, “Impeller Mixing in Vessels 2751–2764 (1988).
Experimental Studies on the Influence of Some Parameters Coulaloglou, C. A. and L. L. Tavlarides, “Description of
and Formulation of a General Mixing Time Equation,” Trans. Interaction Processes in Agitated Liquid–Liquid Dispersions,”
Inst. Chem. Eng. 54, 139–152 (1976). Chem. Eng. Sci. 32, 1289–1297 (1977).
Brooks, B. W., “Drop Size Distributions in an Agitated Coyle, C. K., H. E. Hirschland, B. J. Michel and J. Y. Oldshue,
Liquid/Liquid Dispersion,” Trans. Inst. Chem. Eng. 57, “Heat Transfer to Jackets with Close Clearance Impellers in
210–212 (1979). Viscous Materials,” Can. J. Chem. Eng. 48 (3), 275–278
Brown, D. E. and K. Pitt, “Drop Size Distribution of Stirred (1970).
Non-Coalescing Liquid–Liquid System,” Chem. Eng. Sci. 27, Cutter, L. A., “Flow and Turbulence in a Stirred Tank,” AIChE J.
577–583 (1972). 12, 35–44 (1966).
Brown, D. E. and K. Pitt, “Effect of Impeller Geometry on Drop Das, P. K., R. Kumar and D. Ramkrishna, “Coalescence of Drops
Break-Up in a Stirred Liquid–Liquid Contactor,” Chem. Eng. in Stirred Dispersion. A White Noise Model for Coalescence,”
Sci. 29, 345–348 (1974). Chem. Eng. Sci. 42, 213–220 (1987).
Brucato, A., M. Ciofalo, F. Grisafi and G. Micale, “Complete Daskopoulos, P. and C. K. Harris, “Three Dimensional CFD
Numerical Solution of Flow Fields in Baffled Stirred Vessels: Simulations of Turbulent Flow in Baffled Stirred Tanks: An
The Inner–Outer Approach,” Proc. 8th Euro. Conf. Mix. Assessment of the Current Position,” Inst. Chem. Eng. Symp.
155–162 (1994). Ser. 140, 1–113 (1996).
Brucato, A., M. Ciofalo, F. Grisafi and G. Micale, “Numerical Davies, J. T., “Drop Sizes of Emulsions Related to Turbulent
Prediction of Flow Fields in Baffled Stirred Vessels: A Energy Dissipation Rates,” Chem. Eng. Sci. 40, 839–842
Comparison of Alternative Modelling Approaches,” Chem. (1985).
Eng. Sci. 53, 3653–3684 (1998). De Maerteleire, E., “Heat Transfer to a Helical Cooling Coil in
Brunazzi, E., C. Galletti, A. Paglianti and S. Pintus, “An Mechanically Agitated Gas–Liquid Dispersions,” Chem. Eng.
Impedance Probe for the Measurements of Flow Sci. 33, 1107–1113 (1978).
Characteristics and Mixing Properties in Stirred Slurry Deglon, D. A. and C. Meyer, “CFD Modelling of Stirred Tanks:
Reactors,” Chem. Eng. Res. Des. 82, 1250–1257 (2004). Numerical Considerations,” Miner. Eng. 19, 1059–1068
Carreau, P. J., I. Patterson and C. Y. Yap, “Mixing of Viscoelastic (2006).
Fluids with Helical-Ribbon Agitators. I—Mixing Time and Delafosse, A., A. Liné, J. Morchain and P. Guiraud, “LES and
Flow Patterns,” 54, 135–142 (1976). URANS Simulations of Hydrodynamics in Mixing Tank:

| 74 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Comparison to PIV Experiments,” Chem. Eng. Res. Des. 86, Ford, J. J., T. J. Heindel, T. C. Jensen and J. B. Drake, “X-Ray
1322–1330 (2008). Computed Tomography of a Gas-Sparged Stirred-Tank
Derksen, J. J. and H. E. A. van den Akker, “Large Eddy Reactor,” Chem. Eng. Sci. 63, 2075–2085 (2008).
Simulations on the Flow Driven by a Rushton Turbine,” Forrester, S. E. and C. D. Rielly, “Modelling the Increased Gas
AIChE J. 45, 209–2221 (1999). Capacity of Self-Inducing Impellers,” Chem. Eng. Sci. 49,
Derksen, J. J. and H. E. A. van Den Akker, “Multi-Scale 5709–5718 (1994).
Simulations of Stirred Liquid–Liquid Dispersions,” Chem. Forrester, S. E., C. D. Rielly and K. J. Carpenter, “Gas-Inducing
Eng. Res. Des. 85, 697–702 (2007). Impeller Design and Performance Characteristics,” Chem.
Derksen, J. J., J. H. Stockmann and H. E. A. Akker, Eng. Sci. 53, 603–615 (1998).
“Three-Dimensional Laser Doppler Anemometry in a Stirred Fort, L., H. Valešová and V. Kudraa, “Studies on Mixing. XXVII.
Tank,” Proc. 9th Euro. Conf. Mix. 11, 81–87 (1997). Liquid Circulation in a System with Axial Mixer and Radial
Derksen, J. J., K. Kontomaris, J. B. McLaughlin and H. E. A. van Baffles,” Collect. Czech. Chem. Commun. 36, 164–185
den Akker, “Large-Eddy Simulation of Single-Phase Flow (1971).
Dynamics and Mixing in an Industrial Crystallizer,” Chem. Galletti, C., E. Brunazzi, S. Pintus, A. Paglianti and M.
Eng. Res. Des. 85, 169–179 (2007). Yianneskis, “A Study of Reynolds Stresses, Triple Products
Desai, R. B. and J. B. Joshi, “Flow Pattern in the Vicinity of and Turbulence States in a Radially Stirred Tank with 3-D
Gas–Liquid Interface in a Stirred Cell,” J. Chem. Eng. Jpn. 29, Laser Anemometry,” Chem. Eng. Res. Des. 82, 1214–1228
568–576 (1996). (2004).
Deshmukh, N. A. and J. B. Joshi, “Surface Aerators: Power Gezork, K. M., W. Bujalski, M. Cooke and A. W. Nienow, “Mass
Number, Mass Transfer Coefficient, Gas Hold Up Profile and Transfer and Hold-up Characteristics in a Gassed, Stirred
Flow Patterns,” Trans. Instn. Chem. Engrs - A: Chem. Eng. Vessel at Intensified Operating Conditions,” Chem. Eng. Res.
Res. and Des., 84, 977–992 (2006). Des. 79, 965–972 (2001).
Dohi, N., T. Takahashi, K. Minekawa and Y. Kawase, “Power Ghadge, R. S., S. B. Sawant and J. B. Joshi, “Enzyme
Consumption and Solid Suspension Performance of Deactivation in a Bubble Column, a Stirred Vessel and an
Large-Scale Impellers in Gas–Liquid–Solid Three-Phase Inclined Plane,” Chem. Eng. Sci. 58, 5125–5134 (2003).
Stirred Tank Reactors,” Chem. Eng. J. 97, 103–114 (2004). Gillissen, J. J. J. and H. E. A. van den Akker, “Direct Numerical
Dong, L., S. T. Johansen and T. A. Engh, “Flow Induced by an Simulation of a Turbulently Stirred Tank,” 13th European
Impeller in an Unbaffled Tank—II. Numerical Modelling,” Conference on Mixing London, 14–17 (2009).
Chem. Eng. Sci. 49, 3511–3518 (1994). Gimbun, J., C. D. Rielly and Z. K. Nagy, “Modelling of Mass
Drbohlav, J., I. Fort, K. Maca and J. Placek, “Turbulent Transfer in Gas–Liquid Stirred Tanks Agitated by Rushton
Characteristics of Discharge Flow from the Turbine Impeller,” Turbine and CD-6 Impeller: A Scale-Up Study,” Chem. Eng.
Collect. Czech. Chem. Commun. 43, 3148–3163 (1978). Res. Des. 87, 437–451 (2009).
Ebrahimi-Moshkabad, M. and J. M. Winterbottom, “The Godleski, E. S. and J. C. Smith, “Power Requirements and Blend
Behaviour of an Intermeshing Twin Screw Extruder with Times in the Agitation of Pseudoplastic Fluids,” AIChE J. 8,
Catalyst Immobilised Screws as a Three-Phase Reactor,” Catal. 617–620 (1962).
Today 48, 347–355 (1999). Guha, D., M. P. Dudukovic and P. A. Ramachandran,
Edney, H. G. S. and M. E. Edwards, “Heat Transfer to “CFD-Based Compartmental Modelling of Single Phase
Non-Newtonian and Aerated Fluids in Stirred Tanks,” Trans. Stirred-Tank Reactors,” AIChE J. 52, 1836–1846 (2006).
Inst. Chem. Eng. 54, 160–166 (1976). Guiraud, P., J. Costes and J. Bertrand, “Local Measurements of
Eggels, J. G. M., “Direct and Large-Eddy Simulation of Turbulent Fluid and Particle Velocities in a Stirred Suspension,” Chem.
Flow using the Lattice-Boltzmann Scheme,” Int. J. Heat Fluid Eng. J. 68, 75–86 (1997).
Flow 17, 307–323 (1996). Günkel, A. A. and M. E. Weber, “Flow Phenomena in Stirred
Fajner, D., D. Pinelli, R. S. Ghadge, G. Montante, A. Paglianti Tanks,” AIChE J. 21, 931–948 (1975).
and F. Magelli, “Solids Distribution and Rising Velocity of Hall, K. R. and J. C. Godfrey, “Power Consumption by Helical
Buoyant Solid Particles in a Vessel Stirred with Multiple Ribbon Impellers,” Trans. Inst. Chem. Eng. 48, T201–208
Impellers,” Chem. Eng. Sci. 63, 5876–5882 (2008). (1970).
Fan, J., Y. Wang and W. Fei, “Large Eddy Simulations of Flow Hanjalic,́ K., “Advanced Turbulence Closure Models: A View of
Instabilities in a Stirred Tank Generated by a Rushton Current Status and Future Prospects,” International J. of Heat
Turbine,” Chin. J. Chem. Eng. 15, 200–208 (2007). and Fluid Flow 15 (3), 178–203 (1994).
Fentiman, N. J., N. St Hill, K. C. Lee, G. R. Paul and M. Harnby, N., M. F. Edwards and A. W. Nienow, “Mixing in
Yianneskis, “A Novel Profiled Blade Impeller for Process Chemical Industries,” Butterworth-Henmman
Homogenisation of Miscible Liquids in Stirred Vessels,” Publications, New Delhi (1991).
Chem. Eng. Res. Des 76, 835–842 (1998). Harris, C. K., D. Roekaerts and F. J. J. Rosendal, “Computational
Feranandes, J. B. and M. M. Sharma, “Effective Interfacial Area Fluid Dynamics for Chemical Reactor Engineering,” Chem.
in Agitated Liquid–Liquid Contactors,” Chem. Eng. Sci. 22, Eng. Sci. 51, 1569–1594 (1996).
1267–1282 (1967). Hartmann, H., J. J. Derksen and H. E. A. van den Akker,
Figueiredo, M. M. L. and R. H. Calderbank, “The Scale-Up of “Macro-Instabilities Uncovered in a Rushton Turbine Stirred
Aerated Mixing Vessels for Specified Oxygen Dissolution Tank by Means of LES,” AIChE J. 60, 2383–2393
Rates,” Chem. Eng. Sci. 34, 1333–1338 (1979). (2004a).
Fishwick, R. P., J. M. Winterbottom and E. H. Stitt, “Effect of Hartmann, H., J. J. Derksen, C. Montavo, J. Pearson, I. S. Hamill
Gassing Rate on Solid–Liquid Mass Transfer Coefficients and and H. E. A. van den Akkar, “Assessment of Large Eddy and
Particle Slip Velocities in Stirred Tank Reactors,” Chem. Eng. RANS Stirred Tank Simulations by Means of LDA,” Chem.
Sci. 58, 1087–1093 (2003). Eng. Sci. 59, 2419–2432 (2004b).

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 75 |


Hartmann, H., J. J. Derksen and H. E. A. Van Den Akker, Joshi, J. B. and M. M. Sharma, “Mass Transfer Characteristics of
“Mixing Times in a Turbulent Stirred Tank by Means of LES,” Horizontal Agitated Contactors,” Can. J. Chem. Eng. 54 (6)
AIChE J. 52, 3696–3706 (2006). 560–565 (1976).
Harvey, P. S. and M. G. Greaves, “Turbulent Flow in an Agitated Joshi, J. B. and M. M. Sharma, “Mass Transfer and
Vessel Part-I: A Predictive Model,” Chem. Eng. Res. Des. 60, Hydrodynamic Characteristics of Gas Inducing Type of
195–200 (1982a). Agitated Contactors,” Can. J. Chem. Eng. 55 (6) 683–695
Harvey, P. S. and M. G. Greaves, “Turbulent Flow in an Agitated (1977).
Vessel Part-II: Numerical Solution and Model Predictions,” Joshi, J. B., M. V. Tabib, S. S. Deshpande and C. S. Mathpati,
Chem. Eng. Res. Des. 60, 201–210 (1982b). “Dynamics of Flow Structures and Transport Phenomena—1:
Hiraoka, S., I. Yamada and K. J. Mizoguchi, “Two Dimensional Experimental and Numerical Techniques for Identification
Model Analysis of Flow Behaviour of Highly Viscous and Energy Content of Flow Structures,” Ind. Eng. Chem.
Non-Newtonian Fluid in Agitated Vessel with Paddle Res. 48, 8244–8284 (2009).
Impeller,” Chem. Eng. Jpn. 12, 56–62 (1979). Ju, S. Y., T. M. Mulvahill and R. W. Pike, “Three-Dimensional
Hiraoka, S., N. Kamei, Y. Kato, Y. Tada, K. Asai, S. Hibino and T. Turbulent Flow in Agitated Vessels with a Nonisotropic
Yamaguchi, “Mass Transfer Volumetric Coefficient and Viscosity Turbulence Model,” Can. J. Chem. Eng. 68 (1),
Droplet Diameter in Liquid–Liquid Dispersions Stirred with a 3–16 (1990).
Paddle Impeller with Wire Gauze,” J. Chem. Eng. Jpn. 26, Kadam, B., J. B. Joshi, S. B. Koganti and R. N. Patil, “Dispersed
227–229 (1993). Phase Hold-Up, Effective Interfacial Area and Sauter Mean
Hristov, H., B. Stephan, H. Uwe, K. Holger, H. Günther and S. Drop Diameter in Annular Centrifugal Extractors,” Chem.
Wilfried, “A Study on the Two-Phase Flow in a Stirred Tank Eng. Res. Des. 87, 1379–1389 (2009).
Reactor Agitated by a Gas-Inducing Turbine,” Chem. Eng. Kamienski, J. and R. Wójtowicz, “Dispersion of Liquid–Liquid
Res. Des. 86, 75–81 (2008). Systems in a Mixer with a Reciprocating Agitator,” Chem.
Hutchings, B. J., R. J. Weetmann and B. R. Patel, “Computation Eng. Process. 42, 1007–1017 (2003).
of flow fields in mixing tanks with experimental verification,” Karcz, J. and M. Major, “An Effect of a Baffle Length on the
ASME, Annual meeting, San Francisco, 12, 10–15 (1989). Power Consumption in an Agitated Vessel,” Chem. Eng.
Jadhav, S. V. and V. G. Pangarkar, “Particle–Liquid Mass Transfer Process. 37, 249–256 (1998).
in Mechanically Agitated Contactors,” Ind. Eng. Chem. Res. Kemoun, A., F. Lusseyran, M. Mahouast and J. Mallet,
30, 2496–2503 (1991). “Experimental Determination of the Complete Reynolds
Jahoda, M., M. Moštěk, A. Kukuková and V. Machoň, “CFD Stress Tensor in Fluid Agitated by a Rushton Turbine,” Inst.
Modelling of Liquid Homogenisation in Stirred Tanks with Chem. Eng. Symp. Ser. 136, 399–406 (1994).
One and Two Impellers Using Large Eddy Simulation,” Chem. Khang, S. J. and O. Levenspiel, “New Scale-Up and Design
Eng. Res. Des. 85, 616–6625 (2007). Method for Stirrer Agitated Batch Mixing Vessels,” Chem.
Javed, K. H., T. Mahmud and J. M. Zhu, “Numerical Simulation Eng. Sci. 31, 569–577 (1976).
of Turbulent Batch Mixing in a Vessel Agitated by a Rushton Kim, W. J. and F. S. Manning, “Turbulence Energy and Intensity
Turbine,” Chem. Eng. Process. 45, 99–112 (2006). Spectra in a Baffled, Stirred Vessel,” AIChE J. 10, 747–751
Jaworski, Z., K. N. Dyster, I. P. T. Moore, A. W. Nienow and M. (1964).
S. Wyszynski, “The Use of Angle Resolved LDA Data to Komasawa, I., S. Morioka and T. Ohtake, “Studies of the
Compare Two Differential Turbulence Models Applied to Interaction Rate of Dispersed Phase and Resulting Chemical
Sliding Mesh CFD Flow Simulations in a Stirred Tank,” Proc. Reaction in a Stirred Tank,” Kagaku Kogaku 34, 538 (1970).
9th Euro. Conf. Mix. 11, 187–194 (1997). Komasawa, I., R. Kuboi and T. Otake, “Fluid and Particle Motion
Jenne, M. and M. Reuss, “Fluid Dynamic Modelling and in Turbulent Dispersion—I. Measurement of Turbulence of
Simulation of Gas–Liquid flow in Baffled Stirred Tank Liquid by Continual Pursuit of Tracer Particle Motion,” Chem.
Reactors,” Proc. 9th Euro. Conf. Mix. 11 (52), 201–208 (1997). Eng. Sci. 29 (3) 659–668 (1974).
Jenne, M. and M. Reuss, “A Critical Assessment on the Use of Kresta, S. M. and P. E. Wood, “Prediction of the
k − ε Turbulence Models for Simulation of the Turbulent Three-Dimensional Turbulent Flow in Stirred Tanks,” AIChE
Liquid Flow Induced by a Rushton Turbine in Baffled Stirred J. 37, 448–460 (1991).
Tank Reactors,” Chem. Eng. Sci. 54, 3921–3941 (1999). Kresta, S. M. and P. E. Wood, “The Flow Field Produced by a
Jones, R. M., A. D. Harvey III and S. Acharya, “Two-Equation Pitched Blade Turbine: Characterisation of Turbulence and
Turbulence Modelling for Impeller Stirred Tanks,” ASME J. Estimation of the Dissipation Rate,” Chem. Eng. Sci. 48,
Fluids Eng. 123, 640–648 (2001). 1761–1774 (1993).
Joshi, J. B., “Modifications in the Design of Gas Inducing Kuboi, R. and A. W. Nienow, “Intervortex Mixing Rates in
Impellers,” Chem. Eng. Commun. 5, 213–218 (1980). High-Viscosity Liquids Agitated by High-Speed Dual
Joshi, J. B. and D. D. Kale, “Effect of Drag Reducing additives on Impellers,” Chem. Eng. Sci. 41, 123–134 (1986).
Mass Transfer Characteristics in Gas–Liquid Contactors-Gas Kuboi, R., I. Komasawa, T. Otake and M. Iwasa, “Fluid and
Inducing Contactors,” Chem. Eng. Commun. 3, 15–19 (1979). Particle Motion in Turbulent Dispersion—III Particle–Liquid
Joshi, J. B. and A. W. Patwardhan, “Measurement of Gas Hydrodynamics and Mass-Transfer in Turbulent Dispersion,”
Hold-Up Profiles in Stirred Tank Reactors by Gamma Ray Chem. Eng. Sci. 29, 659–668 (1974).
Attenuation Technique,” Trans. Instn. Chem. Engrs. 80A, Kukuková, A., M. Moštěk, M. Jahoda and V. Machoň, “CFD
411–412 (2002). Prediction of Flow and Homogeneisation in Stirred Vessel:
Joshi, J. B. and V. V. Ranade, “Computational Fluid Dynamics Part I Vessel with One and Two Impeller,” Chem. Eng.
for Designing Process Equipment: Expectations, Current Technol. 28, 1125–1132 (2005).
Status and Path Forward,” Ind. Eng. Chem. Res. 42, Kulkarni, A. A., J. B. Joshi, V. Ravi Kumar, and B. D. Kulkarni,
1115–1128 (2003). “Application of Multiresolution Analysis for Simultaneous

| 76 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Measurement of Gas and Liquid Velocities and Fractional Gas Luo, J. Y., A. D. Gosman, R. I. Issa, J. C. Middleton and M. K.
Hold-Up in Bubble Column Using LDA,” Chem. Eng. Sci. 56, Fitzgerald, “Full Flow Field Computation of Mixing in Baffled
5037–5048 (2001). Stirred Vessels,” Chem. Eng. Res. Des. 71, 342–344
Kumar, P., “Design of Bioreactors. Ph.D. Thesis,” University of (1993).
Mumbai, Mumbai (1997). Luo, J. Y., A. D. Gosman and I. R. Issa, “Prediction of Impeller
Kumar, S., R. Kumar and K. S. Gandhi, “Alternative Mechanisms Induced Flows in Mixing Vessels Using Multiple Frames of
of Drop Breakage in Stirred Vessels,” Chem. Eng. Sci. 46, Reference,” Proc. 8th Euro. Conf. Mix. 155–162 (1994).
2483–2489 (1991). Madden, A. J. and G. L. Damerell, “Coalescence Frequencies in
Kumar, S., R. Kumar and K. S. Gandhi, “A Simplified Procedure Agitated Liquid–Liquid Systems,” AIChE J. 8, 233–239
for Predicting dmax in Stirred Vessels,” Chem. Eng. Sci. 48, (1962).
3092–3096 (1993). Magelli, F., D. Fajner, M. Nocentini and G. Pasquali, “Solid
Kumaresan, T. and J. B. Joshi, “Effect of Impeller Design on the Distribution in Vessels Stirred with Multiple Impellers,”
Flow Pattern and Mixing in Stirred Tanks,” Chem. Eng. J. Chem. Eng. Sci. 45, 615–625 (1990).
115, 173–193 (2006). Magelli, F., D. Fajner, M. Nocentini, G. Pasquali, V. Marisko and
Kumaresan, T., N. K. Nere and J. B. Joshi, “Effect of Internals on P. Ditl, “Solids Concentration Distribution in Slurry Reactors
Flow Pattern and Mixing in Stirred Tanks,” Ind. Eng. Chem. Stirred with Multiple Axial Impellers,” Chem. Eng. Process.
Res. 44, 9951–9961 (2005). 29, 27–32 (1991).
Kumaresan, T., N. K. Nere and J. B. Joshi, “Effect of Internals on Mahoust, M., R. David and G. Cognet, “Periodic Phenomena
the Flow Pattern and Mixing in Stirred Tanks,” Ind. Eng. Generated in the Discharge Flow by a Rushton Turbine in a
Chem. Res. 45, 4849–4850 (2006). CSTR and Investigation of the Reynolds Stresses,” Proc. 6th
Kuncewicz, C., A. Heim and M. Pietrzykowski, “A Model of Euro. Conf. Mix. 23–28 (1988).
Laminar Liquid Flow for Flat Bladed Paddles,” Proc. 10th Mahoust, M., G. Cognet and R. David, “Two Component LDV
Euro. Conf. Mix. 113–120 (1997). Measurements in a Stirred Tank,” AIChE J. 35, 1770–1777
Kuriyama, M., H. Inomata, K. Aral and S. Saito, “Numerical (1989).
Solution for the Flow of Highly Viscous Fluid in Agitated Marshall, E., A. Haidari and S. Subbiah, AIChE Meeting held at
Vessel with Anchor Impeller,” AIChE J. 28, 385–391 (1982). Chicago, November (1996).
Lai, R., S. Kumar, S. N. Upadhyay and Y. D. Upadhya, Martı́n, M., F. J. Montes and M. A. Galán, “Mass Transfer Rates
“Solid–Liquid Mass Transfer in Agitated Newtonian and from Bubbles in Stirred Tanks Operating with Viscous
Non-Newtonian Fluids,” Ind. Eng. Chem. Res. 27, 1246–1259 Fluids,” Chem. Eng. Sci. 65, 3814–3824 (2010).
(1988). Mathpati, C. S. and J. B. Joshi, “Insight into Theories of Heat
Lamberto, D. J., F. J. Muzzio, P. D. Swanson and A. L. and Mass Transfer at the Solid–Fluid Interface Using Direct
Tonkovich, “Using Time-Dependent RPM to Enhance Mixing Numerical Simulation and Large Eddy Simulation,” Ind. Eng.
in Stirred Vessels,” Chem. Eng. Sci. 51, 733–741 (1996). Chem. Res. 46, 8525–8557 (2007).
Lane, G. L., M. P. Schwarz and G. M. Evans, “Modelling of the Mathpati, C. S., S. S. Deshpande and J. B. Joshi, “Computational
Interaction between Gas and Liquid in Stirred Vessels,” Proc. and Experimental Fluid Dynamics of Jet Loop Reactor,”
10th Euro. Conf. Mix. 197–204 (2000). AIChE J. 55, 2526–2544 (2009).
Laufhutte, H. D. and A. B. Mersmaan, “Dissipation of Power in Mavros, P. and C. Baudou, “Quantification of the Performance of
Stirred Vessel,” Proc. 5th Euro. Conf. Mix. 331–340 (1985). Agitators in Stirred Vessels: Definition and Use of an Agitation
Launder, B. E., “Current Capabilities for Modelling Turbulence in Index,” Chem. Eng. Res. Des. 75 (8) 737–745 (1997).
Industrial Flows,” Appl. Sci. Res. 48 (3–4), 247–269 (1990). Mavros, P., R. Mann, S. D. Vlaev and J. Bertrand, “Experimental
Launder, B. E. and D. B. Spalding, “The Numerical Computation Visualisation and CFD Simulation of Flow Patterns Induced
of Turbulent Flows,” Comput. Methods. Appl. Mech. Eng. 3, by a Novel Energy-Saving Dual-Configuration Impeller in
269–289 (1974). Stirred Vessels,” Chem. Eng. Res. Des. 79, 857–866 (2001).
Launder, B. E., A. P. Morse, W. Rodi and D. B. Spalding, “The McGuirk, J. and W. Rodi, “A Mathematical Model for a Vertical
Prediction of Free Shear Flows—A Comparison of Six Jet Discharging Into a Shallow Lake,” Proc. 17th Congress of
Turbulence Models,” NASA Rep. SP-311 (1972). the International Association For Hydraulic Research,
Lee, K. C. and M. Yianneskis, “Measurement of Temperature and 579–586 (1977).
Mixing Time in Stirred Vessels with Liquid Crystal McManamey, W. J., “Sauter Mean and Maximum Drop
Thermography,” Proc. 9th Euro. Conf. Mix. 121–128 Diameters of Liquid–Liquid Dispersions in Turbulent Agitated
(1997). Vessels at Low Dispersed Phase Hold-Up,” Chem. Eng. Sci.
Lee, K. C., K. Ng and M. Yianneskis, “Sliding Mesh Predictions 34, 432–434 (1979).
of the Flows around Rushton Impellers,” IChemE. Symp. Ser. Metzner, A. B. and R. E. Otto, “Agitation of Non-Newtonian
140, 1–12 (1996). Fluid,” AIChE J. 3, 3–10 (1957).
Leszek, R., P. Machniewski, A. Milewska and E. Molga, “CFD Metzner, A. B., R. H. Feehs, H. L. Ramos, R. E. Otto and J. D.
Modelling of Stirred Tank Chemical Reactors: Homogeneous Tuthill, “Agitation of Viscous Newtonian and Non-Newtonian
and Heterogeneous Reaction Systems,” Chem. Eng. Sci. 59, Fluids,” AIChE J. 7, 3–9 (1961).
5233–5239 (2004). Mhetras, M. B., A. B. Pandit and J. B. Joshi, “Effect of Agitator
Levins, D. M. and J. R. Glastonbury, “Particle-Liquid Design on Hydrodynamics and Power Consumption in
Hydrodynamics and Mass Transfer in a Stirred Vessel Part. Mechanically Agitated Gas–Liquid Reactors,” Proc. 8th Euro.
II—Mass Transfer,” Trans. Inst. Chem. Eng. 50, 132–146 Conf. Mix. Cambridge, U.K., 375–382 (1994).
(1972). Micale, G., A. Brucato, F. Grisafi and M. Ciofalo, “Prediction of
Luk, S. and Y. H. Lee, “Mass Transfer in Eddies Close to Flow Fields in a Dual Impeller Stirred Vessel,” AIChE J. 45,
Air–Water Interface,” AIChE J. 32, 1546–1555 (1986). 445–464 (1999).

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 77 |


Micale, G., V. Carrara, F. Grisafi and A. Brucato, “Solids Agitated by Various Impeller Designs,” Chem. Eng. Sci. 63,
Suspension in Three-Phase Stirred Tanks,” Chem. Eng. Res. 5468–5495 (2008).
Des. 78, 319–326 (2000). Murthy, J. Y., S. R. Mathur and D. Choudhary, “CFD Simulation
Middleton, J. C., F. Peirce and P. M. Lynch, “Computation of of Flows in Stirred Tank Rectors using a Sliding Mesh
Flow Fields and Complex Reaction Yield in Turbulent Stirred Technique,” Proc. 8th Euro. Conf. Mix. 155–162 (1994).
Reactors and Comparison with Experimental Data,” Chem. Murthy, B. N., N. A. Deshmukh, A. W. Patwardhan and J. B.
Eng. Res. Des. 64, 18–22 (1986). Joshi, “Hollow Self-Inducing Impellers: Flow Visualisation
Milewska, A. and E. Molga, “Safety Aspects in Modelling and and CFD Simulation,” Chem. Eng. Sci. 62, 3839–3848
Operating of Batch and Semibatch Stirred Tank Chemical (2007a).
Reactors,” Chem. Eng. Res. Des. 88, 304–319 (2010). Murthy, B. N., R. S. Ghadge and J. B. Joshi, “CFD Simulations of
Miller, D. N., “Scale-Up of Agitated Vessels. Mass Transfer from Gas–Liquid-Solid Stirred Reactor: Prediction of Critical
Suspended Solute Particles,” Ind. Eng. Chem. Proc. Des. Dev. Impeller Speed for Solid Suspension,” Chem. Eng. Sci. 62,
10, 365–453 (1971). 7184–7195 (2007b).
Miller, D. K., “Scale-Up of Agitated Vessels Gas–Liquid Mass Murthy, B. N., R. B. Kasundra and J. B. Joshi, “Hollow
Transfer,” AIChE J. 20, 445–453 (1974). Self-Inducing Impellers for Gas–Liquid–Solid Dispersion:
Mishra, V. P. and J. B. Joshi, “LDA Measurements of Gas–Liquid Experimental and Computational Study,” Chem. Eng. J. 141,
Flows in Mechanically Agitated Reactors,” Proc. 7th Euro. 332–345 (2008a).
Conf. Mix. 247–255 (1991). Murthy, B. N., N. Deshmukh, A. W. Patwardhan and J. B. Joshi,
Mishra, V. P. and J. B. Joshi, “Flow Generated by a Disc “Hollow Self-Inducing Impellers: Flow Visualisation and CFD
Turbine—III: Effect of Impeller Diameter, Impeller Location Simulation-II,” Chem. Eng. Sci. 63, 4031–4032 (2008b).
and Comparison with Other Radial Flow Impellers,” Trans. Nagata, S., M. Nishikawa, H. Tada, H. Hirabayashi and S. J.
Instn. Chem. Engrs. (U.K.).-A: Chem. Eng. Res. Des. 71, Gotoh, “Power Consumption of Mixing Impellers in Bingham
563–573 (1993). Plastic Liquids,” Chem. Eng. Jpn. 3, 237–243 (1970).
Mishra, V. P., P. Kumar and J. B. Joshi, “Flow Generated by a Narsimhan, G., G. Nejfelt and D. Ramkrishna, “Breakage
Disc Turbine in Aqueous Solution of Polyacrylamide,” Chem. Functions for Droplets in Agitated Liquid–Liquid
Eng. J. 71, 11–21 (1998). Dispersions,” AIChE J. 30, 457–467 (1984).
Mizoguchi, K., E. O’Shima, H. Inoue and I. Inoue, “Break-Up Nere, N. K., A. W. Patwardhan and J. B. Joshi, “Prediction of
and Coalescence Rate of Dispersed Liquid Droplets in an Flow Pattern in Stirred Tanks: New Constitutive Equation for
Agitated Vessel,” Kagaku Kogaku 37, 521 (1973). Eddy Viscosity,” Ind. Eng. Chem. Res. 40, 1755–1772 (2001).
Mizushina, T., R. Ito, Y. Murakami and M. Kiri, “Experimental Nere, N. K., A. W. Patwardhan and J. B. Joshi, “Liquid Phase
Study of the Heat Transfer of Newtonian Fluid to the Wall of Mixing in Stirred Vessels: Turbulent Flow Regime,” Ind. Eng.
Agitated Vessel,” Kagaku Kogaku 30, 719 (1966a). Chem. Res. 42, 2661–2698 (2003).
Mizushina, T., R. Ito, Y. Murakami and S. Tanaka, Ng, K. and M. Yianneskis, “Observations on the Distribution of
“Experimental Study of the Heat Transfer of Non-Newtonian Energy Dissipation in Stirred Vessels,” Trans. Inst. Chem.
Fluids to the Wall of Agitated Vessel,” Kagaku Kogaku 30, 819 Eng. 78A, 334–347 (2000).
(1966b). Ng, K., N. J. Fentiman, K. C. Lee and M. Yianneskis,
Mlynek, Y. and W. Resnick, “On Local Hold-Up and Average “Assessment of Sliding Mesh CFD Predictions and LDA
Drop Size in a Liquid–Liquid Stirred System,” Can. J. Chem. Measurements of the Flow in a Tank Stirred by a Rushton
Eng. 50 (1), 134–135 (1972). Impeller,” Chem. Eng. Res. Des. 76A, 737–747 (1998).
Mochizuki, M. and L. Takashima, “Distribution of Pressure on Nienow, A. W., “Gas–Liquid Mixing Studies: A Comparison of
the Surface of Blade of Turbine Impeller,” Kagaku Kougaku Rushton Turbines with Some Modern Impellers,” Trans. Inst.
Ronbunshu 10, 399 (1984). Chem. Eng. 74 (Part A), 417–423 (1996).
Molag, M., G. E. H. Joosten and A. A. H. Drinkenburg, “Droplet Nienow, A. W. and W. Bujalski, “Recent Studies on Agitated
Breakup and Distribution in Stirred Immiscible Two-Liquid Three-Phase (Gas–Solid–Liquid) Systems in the Turbulent
Systems,” Ind. Eng. Chem. Fundam. 19, 275–281 (1980). Regime,” Chem. Eng. Res. Des. 80, 832–838 (2002).
Montante, G., K. C. Lee, A. Brucato and M. Yianneskis, Nienow, A. W. and D. Miles, “Impeller Power Numbers in Closed
“Numerical Simulations of the Dependency of Flow Pattern Vessels,” Ind. Eng. Chem. Proc. Des. Dev. 10, 41–43 (1971).
on Impeller Clearance in Stirred Vessels,” Chem. Eng. Sci. 56, Nikiforaki, L., G. Montante, K. C. Lee and M. Yianneskis, “On
3751–3770 (2001). the Origin, Frequency and Magnitude of Macro-Instabilities of
Montante, G., D. Pinelli and F. Magelli, “Scale-Up Criteria for the the Flows in Stirred Vessels,” Chem. Eng. Sci. 58, 2937–2949
Solids Distribution in Slurry Reactors Stirred with Multiple (2003).
Impellers,” Chem. Eng. Sci. 58, 5363–5372 (2003). Nishikawa, M., K. Ashiwake, N. Hashimoto and S. Nagata,
Montante, G., D. Horn and A. Paglianti, “Gas–Liquid Flow and “Agitation Power and Mixing Time in Off-Centering Mixing,”
Bubble Size Distribution in Stirred Tanks,” Chem. Eng. Sci. Int. Chem. Eng. 19, 153–159 (1979).
63, 2107–2118 (2008). Nishikawa, M., T. Kayama, S. Nishioka and S. Nishikawa, “Drop
Mujumdar, A. S., B. Huang, D. Wolf, M. E. Weber and W. J. M. Size Distribution in Mixing Vessel with Aeration,” Chem.
Douglas, “Turbulence Parameters in a Stirred Tank,” Can. J. Eng. Sci. 49, 2379–2384 (1994).
Chem. Eng. 48 (4), 475–483 (1970). Nocentini, M., D. Pinelli and F. Magelli, “Dispersion Coefficient
Murakami, Y., K. Fujimoto, T. Shimada, A. Yamada and K. J. and Settling Velocity of the Solids in Agitated Slurry Reactors
Asano, “Evaluation of Performance of Mixing Apparatus for Stirred with Multiple Rushton Turbines,” Chem. Eng. Sci. 57,
High Viscosity Fluids,” Chem. Eng. Jpn. 5, 297–303 (1972). 1877–1884 (2000).
Murthy, B. N. and J. B. Joshi, “Assessment of Standard k − ε, Novák, V. and E. Rieger, “Homogenisation with Helical Screw
RSM and LES Turbulence Models in a Baffled Stirred Vessel Agitators,” Trans. Inst. Chem. Eng. 47, T335–340 (1969).

| 78 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Obot, N. T., “Design Mixing Processes Using the Frictional Law Patwardhan, A. W., J. B. Joshi, S. Fotedar and T. Mathew,
of Corresponding States,” Chem Eng. Progress 47–54 (1993). “Optimisation of Gas–Liquid Reactor Using Computational
Ochieng, A. and A. E. Lewis, “Nickel Solids Concentration Fluid Dynamics,” Chem. Eng. Sci. 60, 3081–3089 (2005).
Distribution in a Stirred Tank,” Miner. Eng. 19, 180–189 Pericleous, K. A. and M. K. Patel, “The Modelling of Tangential
(2006). and Axial Agitators in Chemical Reactors,” Physico. Chem.
Ochieng, A., A. M. S. Onyango, A. Kumar, K. Kiriamiti and P. Hydrodyn. 8, 105–123 (1987).
Musonge, “Mixing in a Tank Stirred by a Rushton Turbine at Perng, C. Y. and J. Y. Murthy, “A moving-deforming-mesh
a Low Clearance,” Chem. Eng. Process. 47, 842–851 technique for simulation of flow in mixing tanks,” In: Process
(2008). Mixing: Chemical and Biochemical Applications pp. 37–41.
Ogawa, K. and C. Kuroda, “A New Scale-Up Rule and Evaluation (1993).
of Traditional Rules from a Viewpoint of Energy Spectrum Perng, C. Y. and J. Y. Murthy, “A Moving Deforming Mesh
Function,” in: G.B. Tatterson (Ed.), AIChE Symposium Series, Technique for Simulation of Flow in Mixing Tanks,” Proc. 8th
Industrial Mixing Fundamentals with Application, no. 305, Euro. Conf. Mix. 37–39 (1994).
1995, 95–101. Peters, D. C. and J. M. Smith, “Fluid Flow in the Region of
Okamoto, Y., M. Nishikawa and K. Hashimoto, “Energy Anchor Agitator Blades,” Trans. Inst. Chem. Eng. 45,
Dissipation Rate Distribution in Mixing Vessels and Its Effect T360–366 (1967).
on Liquid–Liquid Dispersion and Solid–Liquid Mass Peters, D. C. and J. M. Smith, “Mixing in Anchor Agitated
Transfer,” Int. Chem. Eng. 21, 88–94 (1981). Vessels,” Can. J. Chem. Eng. 47 (3), 268–271 (1969).
Okufi, S., E. S. Perez de Ortiz, and H. Sawistowski, “Scale-Up of Placek, J., L. L. Talvarides, G. W. Smith and I. Fort, “Turbulent
Liquid–Liquid Dispersions in Stirred Tanks,” Can. J. Chem. Flow in Stirred Tanks, Part II: A Two Scale Model of
Eng. 68 (3), 400 (1990). Turbulence,” AIChE J. 32, 1771–1786 (1986).
Oshinowo, L., Z. Jaworski, K. N. Dyster, E. Marshall and A. W. Poncin, S., C. Nguyen, N. Midoux and J. Breysse,
Nienow, “Predicting the Tangential Velocity Field in Stirred “Hydrodynamics and Volumetric Gas–Liquid Mass Transfer
Tanks Using Multiple Reference Frames (MRF) Model with Coefficient of a Stirred Vessel Equipped with a Gas-Inducing
Validation by LDA Measurements,” Proc. 10th Euro. Conf. Impeller,” Chem. Eng. Sci. 57, 3299–3306 (2002).
Mix. 281–289 (2000). Pope, S. B., “Turbulent Flows,” Cambridge University Press,
Pacek, A. W., A. W. Nienow, and I. P. T. Moore, “On the structure Cambridge, UK (2000).
of turbulent liquid–liquid dispersed flows in an agitated Raghava Rao, K. S. M. S. and J. B. Joshi, “Liquid Phase Mixing
vessel,” Chem. Eng. Sci. 49, 3485–3498 (1994). in Mechanically Agitated Three Phase Reactors,” 6th Euro.
Pacek, A., C. Man, and A. Nienow, “On the Sauter mean Conf. Mix, Pavia, Italy, May 24–26, BHRA 6, 427–433 (1988).
diameter and size distributions in turbulent liquid/liquid Raghava Rao, K. S. M. S. and J. B. Joshi, “Gas Phase Dispersion
dispersions in a stirred vessel,” Chem. Eng. Sci. 53, and Power Consumption in Mechanically Agitated Three
2005–2011 (1998). Phase Reactors,” Int. J. Eng. Fluid Mech. 2, 305–320 (1989).
Paglianti, A., S. Pintus and M. Giona, “Time-Series Analysis Raghava Rao, K. S. M. S., V. B. Rewatkar and J. B. Joshi, “Critical
Approach for the Identification of Flooding/Loading Impeller Speed for Solid Suspension in Mechanically Agitated
Transition in Gas–Liquid Stirred Tank Reactors,” Chem. Eng. Solid Liquid Contactors,” AIChE J. 34, 1332–1340 (1988).
Sci. 55, 5793–5802 (2000). Ranade, V. V., “An Efficient Computational Model for Simulating
Pandit, A. B. and J. B. Joshi, “Mixing in Mechanically Agitated Flow in Stirred Vessels: A Case of Rushton Turbine,” Chem.
Contactors, Bubble Columns and Modified Bubble Columns,” Eng. Sci. 52, 4473–4484 (1997).
Chem. Eng. Sci. 38, 1189–1215 (1983). Ranade, V. V. and S. M. S. Dommeti, “Computational Snapshot
Pandit, A. B., N. Tenefrancia, J. B. Joshi and K. K. Tiwari, of Flow Generated by Axial Impellers in Baffled Stirred
“Power Consumption and Homogenisation Time for Vessels,” Chem. Eng. Res. Des. 74, 476–484 (1996).
Oleoresin Suspensions,” Chem. Eng. J. 28, 25–38 (1984). Ranade, V. V. and J. B. Joshi, “Flow Generated by Pitched Bladed
Parsu Veera, U., A. W. Patwardhan and J. B. Joshi, Turbine Part I: Experimental,” Chem. Eng. Commun. 81,
“Measurement of Gas Hold-Up Profiles in Stirred Tank 197–224 (1989).
Reactors by Gamma Ray Attenuation Technique,” Trans. Inst. Ranade, V. V. and J. B. Joshi, “Flow Generated by a Disc Turbine
Chem. Eng. 79A, 684–688 (2001). I: Experimental,” Trans. Instn. Chem. Engrs. (U.K.).-A: Chem.
Patil, V. K., J. B. Joshi and M. M. Sharma, “Solid–Liquid Mass Eng. Res. Des. 68, 19–33 (1990a).
Transfer Coefficient in Mechanically Agitated Contactors,” Ranade, V. V. and J. B. Joshi, “Flow Generated by a Disc Turbine:
Trans. Instn. Chem. Engrs. (U.K.). -A: Chem. Eng. Res. Des. Part II Mathematical Modelling and Comparison with
62, 247–254 (1984). Experimental Data,” Chem. Eng. Res. Des. 68A, 34–43
Patwardhan, A. W., “Prediction of Flow Characteristics and (1990b).
Energy Balance for a Variety of Axial Flow Impellers,” Ind. Ranade, V. V., J. R. Bourne, J. B. Joshi, “Fluid Mechanics and
Eng. Chem. Res. 40, 3806–3816 (2001). Blending in Agitated Tanks” Chem. Eng. Sci. 46, 1883–1893
Patwardhan, A. W. and J. B. Joshi, “Relation Between Flow (1991).
Pattern and Blending in Stirred Tanks,” Ind. Eng. Chem. Res. Ranade, V. V., M. Perrard, C. Xuereb, N. Le Sauze and J.
38, 3131–3143 (1999). Bertrand, “Trailing Vortices of Rushton Turbine: PIV
Patwardhan, A. W., A. B. Pandit and J. B. Joshi, “The Role of Measurements and CFD Simulations with Snapshot
Convection and Turbulent Dispersion in Blending-I,” Chem. Approach,” Chem. Eng. Res. Des. 79 (1), 3–12 (2001).
Eng. Sci. 58, 2951–2962 (2003). Ranade, V. V., Y. Tayalia and H. Krishnan, “CFD Predictions of
Patwardhan, A. W., A. B. Pandit and J. B. Joshi, “The Role of Flow Near Impeller Blades in Baffled Stirred Vessels:
Convection and Turbulent Dispersion in Blending-II,” Chem. Assessment of Computational Snapshot Approach,” Chem.
Eng. Sci. 59, 729 (2004). Eng. Commun. 189, 895–922 (2002).

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 79 |


Revstedt, J., L. Fuchs and C. Tragardh, “Large Eddy Simulations Sahu, A. K., P. Kumar and J. B. Joshi, “Simulation of Flow in
of the Turbulent Flow in a Stirred Reactor,” Chem. Eng. Sci. Stirred Vessels with Axial Flow Impeller: Zonal Modelling
53, 4041–4043 (1998). and Optimisation of Parameters,” Ind. Eng. Chem. Res. 37,
Rewatkar, V. B. and J. B. Joshi, “Role of Sparger Design in 2116–2130 (1998).
Mechanically Agitated Gas–Liquid Reactor-I: Power Sahu, A. K., P. Kumar, A. W. Patwardhan and J. B. Joshi, “CFD
Consumption,” Chem. Eng. Technol. 14, 333–347 (1991a). Modelling and Mixing in Stirred Tanks,” Chem. Eng. Sci. 54,
Rewatkar, V. B. and J. B. Joshi, “Role of Sparger Design in 2285–2293 (1999).
Mechanically Agitated Gas–Liquid Reactors-II: Liquid Phase Saito, E. and M. J. Kamiwano, “An Extended Technique for
Mixing,” Chem. Eng. Technol. 14, 386–393 (1991b). Predicting the Mixing Times of High- Viscosity Liquid in a
Rewatkar, V. B. and J. B. Joshi, “Effect of Impeller Design on Mixer-Mixing Systems with Molecular Diffusion of Solute,” J.
Liquid Phase Mixing in Mechanically Agitated Reactors,” Chem. Eng. Jpn. 22, 491–496 (1989).
Chem. Eng. Commun. 102, 1–34 (1991c). Saito, E., K. Aral and M. J. Kamiwano, “An Extended Technique
Rewatkar, V. B. and J. B. Joshi, “Effect of Addition of Alcohol on for Predicting the Mixing Times of High-Viscosity Liquid in a
Design Parameters in Mechanically Agitated Three Phase Mixer-Mixing Systems with Molecular Diffusion and Reaction
Reactors,” Chem. Eng. J. 49, 107–117 (1992). of Solute,” J. Chem. Eng. Jpn. 23, 222–227 (1990).
Rewatkar, V. B. and J. B. Joshi, “Role of Sparger Design on Gas Sano, Y. and H. J. Usui, “Interrelations among Mixing Time,
Dispersion in Mechanically Agitated Gas/Liquid Reactors,” Power Number and Discharge Flow Rate Number in Baffled
Can. J. Chem. Eng. 71 (2) 278–291 (1993). Mixing Vessels,” Chem. Eng. Jpn. 18, 47–52 (1985).
Rewatkar, V. B., K. S. M. S. Raghava Rao and J. B. Joshi, “Some Sarvanan, K., V. D. Mundale, A. W. Patwardhan and J. B. Joshi,
Aspects of Solid Suspension in Mechanically Agitated “Power Consumption in Gas Inducing Type Mechanically
Reactors,” AIChE J. 35, 1575–1582 (1989). Agitated Reactors,” Ind. Eng. Chem. Res. 35, 1583–1602
Rewatkar, V. B., A. J. Deshpande, A. B. Pandit and J. B. Joshi, (1996).
“Gas Hold-Up Behaviour of Mechanically Agitated Satoh, K., T. Menju, M. Mochizuki and A. Shono, “Mixing Times
Gas–Liquid Reactors Using Pitched Blade Downflow Turbine,” of Liquids in Gas–Liquid Contactors with Mechanical
Can. J. Chem. Eng. 71 (2), 226–237 (1993). Agitation,” Kagaku Kogaku Ronbunshu 21, 137 (1995).
Reynolds, W. C., “Fundamentals of Turbulence for Turbulence Sawant, S. B. and J. B. Joshi, “Critical Impeller Speed for the
Modelling and Simulation,” Lecture Notes for Von Karman Onset of Gas Induction in Gas Inducing Types of Agitated
Institute, AGARD CP-93 (1987). Contactors,” Chem. Eng. J. 18, 87–91 (1979).
Ribeiro, L. M., P. F. R. Regueiras, M. M. L. Guimarães, C. M. N. Sawant, S. B., J. B. Joshi, V. G. Pangarkar and R. D. Mhaskar,
Madureira and J. J. C. Cruz-Pinto, “The Dynamic Behaviour “Mass Transfer and Hydrodynamics Characteristics of Denver
of Liquid–Liquid Agitated Dispersions—II. Coupled Type of Floatation Cells” Chem. Eng. J. 21, 11–119
Hydrodynamics and Mass Transfer,” Comp. Chem. Eng. 21, (1981).
543–558 (1996). Sbrizzai, F., V. Lavezzo, R. Verzicco, M. Campolo and A.
Rieger, E. and V. Novák, “Scale-Up Method for Power Soldatia, “Direct Numerical Simulation of Turbulent Particle
Consumption of Agitators in the Creeping Flow Regime,” Dispersion in an Unbaffled Stirred-Tank Reactor,” Chem. Eng.
Chem. Eng. Sci. 27, 39–44 (1972). Sci. 61, 2843–2851 (2006).
Rielly, C. D., G. M. Evans, J. F. Davidson and K. J. Carpenter, Scargiali, F., A. D’Orazio, F. Grisafi and A. Brucato, “Modelling
“Effect of Vessel Scaleup on the Hydrodynamics of a and Simulation of Gas–Liquid Hydrodynamics in
Self-Aerating Concave Blade Impeller,” Chem. Eng. Sci. 47, Mechanically Stirred Tanks,” Chem. Eng. Res. Des. 85,
3395–3402 (1992). 637–646 (2007a).
Rielly, C. D., M. Habib and J. P. Sherlock, “Flow and Mixing Scargiali, F., R. Russo, F. Grisafi and A. Brucato, “Mass Transfer
Characteristics of a Retreat Curve Impeller in a Conical-Based and Hydrodynamic Characteristics of a High Aspect Ratio
Vessel,” Chem. Eng. Res. Des. 85, 953–962 (2007). Self-Ingesting Reactor for Gas–Liquid Operations,” Chem.
Rigby, G. D., G. M. Evans and G. J. Jameson, “Bubble Breakup Eng. Sci. 62, 1376–1387 (2007b).
from Ventilated Cavities in Multiphase Reactor,” Chem. Eng. Schäfer, M., M. Höfken and F. Durst, “Detailed LDV
Sci. 52, 3677–3684 (1997). Measurements for Visualisation of the Flow Field Within a
Robinson, C. W. and C. R. Wilke, “Oxygen Absorption in Stirred Stirred-Tank Reactor Equipped with a Rushton Turbine,”
Tanks: A Correlation for Ionic Strength Effects,” Biotechnol. Chem. Eng. Res. Des. 75, 729–736 (1997).
Bioeng. 15, 755–782 (1973). Schäfer, M. M., M. Yianneskis and F. Durst, “Trailing Vortices
Rodi, W., “Turbulence Models and Their Application in Behind 45◦ Pitched Blade Impeller,” AIChE J. 44, 1233–1246
Hydraulics,” IAHR/AIRH Monograph, A. A. Balkema, Ed. (1998).
Rotterdam (1993). Sciberras, M. A. and G. N. Coleman, “Testing of
Rounsley, R. R., “Oil Dispersion with a Turbine Mixer,” AIChE J. Reynolds-Stress-Transport Closures by Comparison with DNS
29, 597–603 (1983). of an Idealized Adverse-Pressure-Gradient Boundary Layer,”
Roy, S., S. Acharya and M. Cloeter, “Flow Structure and the Eur. J. Mech. B Fluids 26, 551–582 (2007).
Effect of Macro-Instabilities in a Pitched-Blade Stirred Tank,” Sembira, A. N., J. C. Merchuk and D. Wolf, “Characteristics
Chem. Eng. Sci. 65, 3009–3024 (2010). of a Motionless Mixer for Dispersion of Immiscible
Rushton, J. H., E. W. Costich and H. J. Everett, “Power Fluids—III. Dynamic behaviour of the average drop size and
Characteristics of Mixing Impellers Part 1,” Chem. Eng. dispersed phase hold-up,” Chem. Eng. Sci. 43, 373–377
Progress 46, 395–404 (1950a). (1988).
Rushton, J. H., E. W. Costich and H. J. Everett, “Power Sessiecq, P., P. Mier, F. Gruy and M. Cournil, “Solid Particles
Characteristics of Mixing Impellers Part 2,” Chem. Eng. Concentration Profiles in an Agitated Vessel,” Chem. Eng.
Progress 46, 467–476 (1950b). Res. Des. 77, 741–746 (1999).

| 80 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |


Shamlou, R. A. and M. E. Edwards, “Heat Transfer to Viscous Establish Scaling Rules,” Chem. Eng. Sci. 33, 1161–1168
Newtonian and Non-Newtonian Fluids for Helical Ribbon (1978).
Mixers,” Chem. Eng. Sci. 41, 1957–1967 (1986). van’t Riet, K. and J. M. Smith, “The Trailing Vortex System
Sheng, J., H. Meng and R. O. Fox, “A Large Eddy PIV Method for Produced by Rushton Turbine Agitators,” Chem. Eng. Sci. 30,
Turbulence Dissipation Rate Estimation,” Chem. Eng. Sci. 55, 1093–1105 (1975).
4423–4434 (2000). Venneker, B. C. H. and H. E. A. van den Akker, “CFD
Shervin, C. R., D. A. Raughley and R. A. Romaszewski, “Flow Calculations of the Turbulent Flow of Shear-Thinning Fluids
Visualisation Scaleup Studies for the Mixing of Viscoelastic in Agitated Tanks,” Proc. 9th Euro. Conf. Mix. 11, 179–186
Fluids,” Chem. Eng. Sci. 46, 2867–2873 (1991). (1997).
Shewale, S. D. and A. B. Pandit, “Studies in Multiple Impeller Verzicco, R., M. Fatica, G. Iaccarino and P. Orlandi, “Flow in an
Agitated Gas–Liquid Contactors,” Chem. Eng. Sci. 61, Impeller-Stirred Tank Using an Immersed-Boundary Method,”
489–504 (2006). AIChE J. 50, 1109–1118 (2004).
Shimizu, K., K. Minekawa, T. Hirose and Y. Kawase, “Drop Walas, S. M., “Chemical Process Equipment,”
Breakage in Stirred Tanks with Newtonian and Butterworth-Heinemann Co, USA (1990).
Non-Newtonian Fluid Systems,” Chem. Eng. J. 72, 117–124 Wang, K. and S. Yu, “Heat Transfer and Power Consumption of
(1999). Non-Newtonian Fluids in Agitated Vessels,” Chem. Eng. Sci.
Shiue, S. J. and C. W. Wong, “Studies on Homogenisation 44, 33–40 (1989).
Efficiency of Various Agitators in Liquid Blending,” Can. J. Warmoeskerken, M. M. C. G. and J. M. Smith, “Flooding of Disc
Chem. Eng. 62 (5) 602–609 (1984). Turbines in Gas–Liquid Dispersions: A New Description of
Sprow, E. B., “Distribution of Drop Sizes Produced in Turbulent the Phenomenon,” Chem. Eng. Sci. 40, 2063–2071
Liquid–Liquid Dispersion,” Chem. Eng. Sci. 22, 435–444 (1985).
(1967). Wechsler, K., M. Breuer and F. Durst, “Steady and Unsteady
Sun, H., Z. Mao and G. Yu, “Experimental and Numerical Study Computations of Turbulent Flows Induced,” Trans. Am. Soc.
of Gas Hold-Up in Surface Aerated Stirred Tanks,” Chem. Eng. Mech. Eng. J. Fluids Eng. 121, 318–329 (1999).
Sci. 61, 4098–4110 (2006). Weinstein, B. and R. E. Treybal, “Liquid–Liquid Contacting in
Szalaia, E. S., P. Arratia, K. Johnson and F. J. Muzzio, “Mixing Unbaffled, Agitated Vessels,” AIChE J. 19, 304–312
Analysis in a Tank Stirred with Ekato IntermigJ Impellers,” (1973).
Chem. Eng. Sci. 59, 3793–3805 (2004). Wernersson, E. S. and C. Tragardh, “Scaling of Turbulence
Tabor, G., A. D. Gosman and R. I. Issa, “Numerical simulation of Characteristics in a Turbine-Agitated Tank in Relation to
the flow in a mixing vessel stirred by a Rushton Turbine,” Agitation Rate,” Chem. Eng. J. 70, 37–45 (1998).
IChemE Symposium Series 140, 25–34 (1996). Wernersson, E. S. and C. Tragardh, “Scale-Up of Rushton
Tabor, A., A. D. Gosman and R. A. Issa, “Numerical Simulation Turbine-Agitated Tanks,” Chem. Eng. Sci. 54, 4245–4256
of the Flow in a Mixing Vessel Stirred by a Rushton Turbine,” (1999).
IChemE Symp. Ser 140, 25–34 (1998). Wilcox, D. C., “Turbulence Modelling for CFD,” DCW Industries,
Takahashi, K. and A. W. Nienow, “Effect of Gas Density on Inc., Canada (1993).
Power Consumption in Aerated Vessel Agitated by a Rushton Wong, C. W., J. R. Wang and S. T. Huang, “Investigations of
Turbine,” J. Chem. Eng. Jpn. 25, 432–434 (1992). Fluid Dynamics in Mechanically Stirred Aerated Slurry
Takahashi, K., T. Yokota and H. Konno, “Mixing of Pseudoplastic Reactors,” Can. J. Chem. Eng. 65 (3), 412–419 (1987).
Liquid in a Vessel Equipped with a Variety of Helical Ribbon Wu, H. and G. K. Patterson, “Laser-Doppler Measurements of
Impellers,” J. Chem. Eng. Jpn. 21, 63–68 (1988). Turbulent-Flow Parameters in a Stirred Mixer,” Chem. Eng.
Tanaka, M. and T. Izumi, “Gas Entrainment in Stirred-Tank Sci. 44, 2207–2221 (1989).
Reactors,” Chem. Eng. Res. Des. 65, 195–198 (1987). Xu, Y. and G. McGrath, “CFD Predictions in Stirred Tank Flows,”
Tecante, A., E. Britode la Fuente, L. Choplin and P. A. Tanguy, Chem. Eng. Res. Des. 74, 471–475 (1996).
“Advances in Engineering Fluid Mechanics: Multiphase Yeoh, S. L., G. Papadakis and M. Yianneskis, “Numerical
Reactor and Polymerization System Hydrodynamics,” Pages Simulation of Turbulent Flow Characteristics in a Stirred
431-453, 1996. Vessel Using the LES and RANS Approaches with the
Thakre, S. S. and J. B. Joshi, “Momentum, Mass and Heat Sliding/Deforming Mesh Methodology,” Chem. Eng. Res. Des.
Transfer in Single Phase Turbulent Flow,” Rev. Chem. Eng. 82, 834–848 (2004).
18, 283–293 (2002). Yeoh, S. L., G. Papadakis and M. Yianneskis, “Determination of
Thatte, A. R., R. S. Ghadge, A. W. Patwardhan, J. B. Joshi and G. Mixing Time and Degree of Homogeneity in Stirred Vessels
Singh, “Local Gas Holdup Measurement in Sparged and with Large Eddy Simulation,” Chem. Eng. Sci. 60, 2293–2302
Aerated Tanks by Gamma-Ray Attenuation Technique,” Ind. (2005).
Eng. Chem. Res. 43, 5389–5399 (2004). Yianneskis, M., Z. Popiolek and J. E. Whitelaw, “An
Uchida, S., H. Moriguchi, H. Maejima, K. Koide and S. Experimental Study of the Steady and Unsteady Flow
Kageyama, “Absorption of Sulfur Dioxide into Limestone Characteristics of Stirred Reactors,” J. Fluid Mech. 175,
Slurry in a Stirred Tank Reactor,” Can. J. Chem. Eng. 56 (6), 537–555 (1987).
690–697 (1978). Yung, C. N., C. W. Wong and C. L. Chang, “Gas Holdup and
Unadkat, H., C. D. Rielly, G. K. Hargrave and Z. K. Nagy, Aerated Power Consumption in Mechanically Stirred Tanks,”
“Application of Fluorescent PIV and Digital Image Analysis to Can. J. Chem. Eng. 57 (6), 672–676 (1979).
Measure Turbulence Properties of Solid–Liquid Stirred Zaccone, A., A. Gäbler, S. Maaß, D. Marchisio and M. Kraume,
Suspensions,” Chem. Eng. Res. Des. 87, 573–586 (2009). “Drop Breakage in Liquid–Liquid Stirred Dispersions:
van der Molen, K. and H. R. E. van Mannen, “Laser Doppler Modelling of Single Drop Breakage,” Chem. Eng. Sci. 62,
Measurements of the Turbulent Flow in Stirred Vessels to 6297–6307 (2007).

| VOLUME 89, FEBRUARY 2011 | | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | 81 |


Zadghaffari, R., J. S. Moghaddas and J. Revstedt, “A Mixing
Study in a Double-Rushton Stirred Tank,” Comp. Chem. Eng.
33, 1240–1246 (2009).
Zadghaffari, R., J. S. Moghaddas and J. Revstedt, “Large-Eddy
Simulation of Turbulent Flow in a Stirred Tank Driven by a
Rushton Turbine,” Comp. Fluids 39, 1183–1190 (2010).
Zhang, Y., C. Yang and Z. Mao, “Large Eddy Simulation of Liquid
Flow in a Stirred Tank with Improved Inner–Outer Iterative
Algorithm,” Chin. J. Chem. Eng. 14, 321–329 (2006).
Zhou, G. and S. M. Kresta, “Evolution of Drop Size Distribution
in Liquid–Liquid Dispersions for Various Impellers,” Chem.
Eng. Sci. 53, 2099–2113 (1998).
Zlokarnik, M., “Stirring: Theory and Practice,” Wiley-VCH,
Weinheim, Germany (2001).

Manuscript received November 30, 2009; revised manuscript


received August 4, 2010; accepted for publication August 6, 2010.

| 82 | THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING | | VOLUME 89, FEBRUARY 2011 |

Вам также может понравиться