You are on page 1of 10

Minerals Engineering 19 (2006) 10591068

This article is also available online at:


www.elsevier.com/locate/mineng

CFD modelling of stirred tanks: Numerical considerations


D.A. Deglon

a,*

, C.J. Meyer

Mineral Processing Research Unit, Department of Chemical Engineering, University of Cape Town, Private Bag Rondebosch,
Cape Town 7700, South Africa
Center for Research in Computational and Applied Mechanics, University of Cape Town, Private Bag Rondebosch, Cape Town 7700, South Africa
Received 8 December 2005; accepted 2 April 2006
Available online 8 June 2006

Abstract
The aim of this paper is to demonstrate that the Multiple Reference Frames (MRF) impeller rotation model and the standard ke
turbulence model, as commonly used in engineering CFD simulations of stirred tanks, can accurately model turbulent uid ow provided
very ne grids coupled with higher-order discretization schemes are used. The MRF model has been found to give adequate results for
the steady-state simulation of stirred tanks but the ke turbulence model generally under or over-predicts turbulence. In this study the
CFD software Fluent 6 is used to simulate ow in a small baed tank of standard geometry agitated by a Rushton turbine impeller.
Simulations are conducted on four grids of signicantly dierent resolution using the upwind, central and QUICK discretization
schemes. CFD model results are evaluated in terms of the predicted ow eld, power number, mean velocity components and turbulent
kinetic energy using published experimental data. The general ow eld and mean uid velocity predictions are not strongly inuenced
by either the grid resolution or discretization scheme. However, turbulent kinetic energy predictions are strongly inuenced by both the
grid resolution and discretization scheme. In this study a grid consisting of nearly 2 million control volumes in one half of a 15 cm diameter stirred tank, combined with a high-order discretization scheme, was required to accurately predict the turbulent kinetic energy. These
represent conditions which are considerably more numerically intensive than used in most similar studies and suggests that the poor predictions of turbulence obtained using the ke turbulence model, often noted in the literature, may be due to numerical errors rather than
inadequacies in the turbulence model.
 2006 Elsevier Ltd. All rights reserved.
Keywords: Computational uid dynamics; Modelling; Agitation

1. Introduction
Stirred/agitated tanks are widely used in the mineral and
metallurgical industries e.g. leach tanks, mechanical otation cells, crystallizers, mixer-settlers, etc. Computational
Fluid Dynamics (CFD) provides a tool for determining
detailed information on uid ow (hydrodynamics) which
is necessary for modelling subprocesses in stirred tanks.
A CFD model of a stirred tank requires one to select,
amongst others, an appropriate grid resolution, discretization scheme, impeller rotation model and turbulence
model. The selection of these, largely numerical consider*

Corresponding author. Tel.: +27 216502980; fax: +27 216505501.


E-mail address: dad@chemeng.uct.ac.za (D.A. Deglon).

0892-6875/$ - see front matter  2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2006.04.001

ations, can have a dramatic inuence on both the accuracy


of the CFD simulation and the associated computational
expense. This is particularly important when performing
CFD simulations of large industrial vessels (engineering
studies) as the number of control volumes in the grid
increases enormously with increasing tank size. As a consequence, until there are no longer computational limitations, CFD models will always maintain a compromise
between reasonable accuracy and reasonable computational expense. In particular, engineering CFD models
often reduce computational expense by using the pseudosteady state MRF impeller rotation model and the engineering standard ke turbulence model. The MRF model
has been found to give similar results to the more sophisticated transient Sliding Mesh impeller rotation model for

1060

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068

Nomenclature
a
impeller blade width (m)
b
impeller hub diameter (m)
C
o bottom impeller clearance (m)
Cl, C1e, C2e constants of the ke turbulence model
d
impeller disc diameter (m)
D
impeller diameter (m)
h
impeller blade height (m)
H
height of uid in the tank (m)
k
turbulent kinetic energy (m2/s2)
N
impeller speed (s 1)
Np
impeller power number, P/(qN3D5) (dimensionless)
NRe
impeller Reynolds number (qND2)/l (dimensionless)
P
power drawn by impeller (W)
r
radial distance from impeller centre (m)
t
time (s)
T
tank diameter (m)
Utip
impeller tip velocity (m/s)

the steady-state simulation of stirred tanks, suggesting that


the additional computational expense is unjustied (Koh
et al., 2003; Koh and Schwarz, 2005). However, the ke turbulence model has been found to poorly predict turbulence
in stirred tanks which is important as models for subprocesses occurring in these vessels often require an accurate
description of turbulence (e.g. particle-bubble collision in
otation). This paper demonstrates that the MRF impeller
rotation model and the standard ke turbulence model may
be used to accurately model turbulent uid ow in a stirred
tank, provided very ne grids coupled with higher-order
discretization schemes are used. The paper begins with an
overview of the inuence of grid resolution, discretization
scheme, impeller rotation model and turbulence model on
the CFD modelling of stirred tanks. The paper focuses
on comparison of ndings which are of relevance to the
small laboratory vessel used in this study.
1.1. Grid resolution
The resolution of the computational grid is a key factor
in any CFD simulation as this is directly related to the
computational cost of the solution. Grids reported in the
various studies of stirred tanks range from an order of
about 104 to about 106 control volumes for the simulation
of stirred tanks of similar dimensions. For example, Brucato et al. (1998) used about 44,000 control volumes on
one quarter of a 19 cm diameter stirred tank, while Bartels
et al. (2002) used approximately 2 million control volumes
on one half of a 15 cm diameter stirred tank. To determine
the grid resolution required to get a solution that is both
physically and numerically accurate, it is necessary to perform simulations on successively rened grids until no

w
z

bae width (m)


vertical distance from tank bottom (m)

Greek symbols
e
rate of dissipation of turbulence energy (m2/s3)
q
density (kg/m3)
rk, re constants of the ke turbulence model
h
tangential distance from impeller center plane
(rad)
l
dynamic viscosity (kg/ms)
Abbreviations
CFD computational uid dynamics
LDV laser Doppler velocimetry
MRF multiple reference frames
PIV
particle image velocimetry
RMS root-mean-square
SM
sliding mesh

notable dierence in the predicted values of the important


solution variables are observed. Ng et al. (1998) performed
grid sensitivity tests on successively rened grids from
about 100,000 to 240,000 control volumes. It was shown
that the successively ner grids yielded equivalent predictions of mean velocity proles which were in good agreement with experimental LDV data but that the turbulent
kinetic energy was severely under-predicted on all grids,
with the ner grids showing only marginal improvement.
Montante et al. (2001) reported similar ndings using grids
from about 23,000 to 145,000 control volumes. The relatively small improvement in the turbulent kinetic energy
observed on the ner grids led these authors to conclude
that further grid renement was unlikely to result in better
predictions. However, it should be noted that the level of
successive grid renements mentioned in these, and other
studies, was relatively small and may not display accurate
grid dependency of the turbulent kinetic energy. For example, the work of Wechsler et al. (1999), which involved grid
sensitivity studies of up to over a million control volumes,
showed that extremely ne grids are needed to isolate
numerical and model errors in the prediction of turbulence
quantities. Therefore it is possible that some of the unsatisfactory solutions for the turbulent kinetic energy observed
in the literature are a result of inadequate grid resolution
i.e. too coarse a grid.
1.2. Discretization scheme
Some authors have investigated the eect of the discretization scheme on the accuracy of the predicted ow and in
most cases have shown that the choice of discretization
scheme has little or no eect on the solution. Brucato

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068

et al. (1998) compared a hybrid scheme (upwind-central


dierencing) and the high-order QUICK scheme and found
that predicted mean velocities did not dier appreciably.
These authors concluded that for their model of 97,000
control volumes, numerical diusion eects associated with
the rst-order upwind discretization scheme were not signicant and that turbulent diusion was largely dominant.
However, no comparison of the eect of discretization on
turbulence quantities was presented and it is not clear
whether the results were grid independent. Aubin et al.
(2004) investigated three discretization methods (rst-order
upwind, upwind-central hybrid, QUICK) on a grid of
155,000 control volumes and found that the choice of the
discretization scheme had no eect on the mean velocities,
except that the upwind scheme was found to under-predict
the swirling region below the impeller. However, all three
discretization schemes were found to under-predict the turbulent kinetic energy, this being most severe for the rstorder upwind scheme in the impeller discharge stream.
1.3. Impeller rotation modelling
Modelling the impeller rotation is complex as the relative motion between the rotating impeller blades and the
stationary baes causes a cyclic variation of the solution
domain. Approaches to modelling the impeller rotation
include using experimentally determined impeller boundary
conditions, specifying momentum source/sink terms on the
impeller blades or incorporating rotating and stationary
reference frames. Two commonly used models are the Multiple Reference Frames and Sliding Mesh models. In these
models, the solution domain is divided into an inner region
containing the rotating impeller and an outer region containing the stationary baes. For the MRF model,
steady-state calculations are performed with a rotating reference frame in the impeller region and a stationary reference frame in the outer region. In this way, the eects of the
blade rotation are accounted for by virtue of the frame of
reference, allowing for explicit modelling of the impeller
geometry. For the SM model, the impeller region is allowed
to slide relative to the outer region in discrete time steps
and time-dependent calculations are performed using
implicit or explicit interpolation of data at successive
time-steps. Being time dependent, the SM method is the
more accurate representation of the actual phenomenon
of the impeller rotation but is computationally demanding.
1.4. Turbulence modelling
The ke model is the most established turbulence model
for engineering ows and has been widely used for modelling turbulent ow in stirred tanks. In most of these studies,
poor prediction of turbulence quantities has been attributed to deciencies in this model, especially the inherent
assumption of isotropic turbulence and limitations in predicting swirling or recirculating ow (Abujelala and Lilley,
1984; Armentante et al., 1997; Jenne and Reuss, 1999). Sev-

1061

eral authors have compared ow predictions in stirred


tanks using variations of the ke model such as the ChenKim and Renormalized Group (RNG) ke models
(Ranade et al., 1989; Jenne and Reuss, 1999; Jaworski
and Zakrzewska, 2002; Aubin et al., 2004). Generally, the
dierent variations have resulted in only slight changes in
turbulence predictions, and in some cases the standard
model gave superior results. It has been suggested that a
turbulence model that is not based on the assumption of
isotropic turbulence should give better results. Armentante
et al. (1997) found that the Algebraic Stress Model indeed
gave superior results to the standard ke turbulence model.
However, in some published studies the Reynolds Stress
Model based on non-isotropic turbulence was found to
yield turbulent kinetic energy proles showing a larger
deviation from experimental values than those obtained
using the standard ke model (Montante et al., 2001;
Jaworski and Zakrzewska, 2002). The use of large eddy
simulation for modelling ow in stirred tanks has been
reported to result in good agreement with experimental
data for both mean velocities and turbulence quantities
(Eggels, 1996; Derksen and Van Den Akker, 1999; Hartmann et al., 2004). However, due to limitations on grid
requirements and the associated computational expense,
LES techniques are impractical for many research
problems.
The aim of this paper is to investigate the eects of grid
resolution and discretization scheme on the CFD simulation of uid ow in a stirred tank using the MRF impeller
rotation model and the standard ke turbulence model. In
particular, the paper aims to determine whether the poor
predictions of turbulence obtained using the ke turbulence
model, often noted in the literature, may be due to numerical errors associated with the choice of grid resolution and
discretization scheme. The CFD model predictions are
compared to the experimental data of Rushton et al.
(1950) and Wu and Patterson (1989).
2. System description and computational method
2.1. Stirred tank conguration
The system investigated in this study consists of a 15 cm
diameter cylindrical tank with four equally spaced baes
and agitated by a standard six-bladed Rushton turbine
impeller (cf. Fig. 1). This system was chosen as it is more
or less a research standard conguration for stirred tanks
and is suciently small to investigate very ne grids.
Water at 25 C (density = 997 kg/m3, dynamic viscosity =
0.00089 kg/ms) was used as the test uid. In this study,
results are compared with the experimental data of Wu
and Patterson (1989). These authors published an extensive
phase-averaged experimental data set (LDV) in the form of
axial proles of mean velocities and RMS turbulent velocities at various radial distances from the impeller, measured
in a plane mid-way between the baes. It is well known
that measurements obtained in the impeller discharge

1062

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068

Fig. 1. Conguration of the stirred tank and Rushton turbine impeller.

stream overestimate turbulence since such measurements


include the non-random variation in the mean velocity
due to the passage of the impeller blades. In their results,
Wu and Patterson (1989) separated the total uctuating
velocities into purely turbulent and periodic components.
In this study, a range of impeller speeds corresponding
to Reynolds numbers covering the laminar and turbulent
ow regimes (0.5 6 NRe 6 60,000) were simulated. It
should be noted that the experimental studies of Wu and
Patterson (1989) were conducted in a larger unit of equivalent geometry and results are reported for NRe = 29,000.
However, experimental studies have shown that for turbulent ow in geometrically similar vessels, velocity values
normalized by the impeller tip speed are independent of
the rotational speed of the impeller or the size of the stirred
tank (Costes and Courdec, 1988; Wu and Patterson, 1989;
Dyster et al., 1993). Therefore the quantitative comparison
with the experimental data of Wu and Patterson (1989) are
considered to be valid for the turbulent Reynolds number
simulations investigated in this study.
2.2. CFD methodology
The CFD software Fluent 6.1 was used in this study.
For the turbulent Reynolds number range (NRe P 7000),
the standard ke turbulence model and model constants
were used (cf. Table 1). To account for the viscous ow

Table 1
Standard ke model constants
Cl

rk

re

C1e

C2e

0.09

1.0

1.3

1.44

1.92

region near the solid surfaces, standard wall functions as


proposed by Launder and Spalding (1974) were applied.
Since the stirred tank and expected ow show rotational
symmetry, only half of the tank geometry was used as
the calculation domain (i.e. 180 geometry, 2 baes, 3
impeller blades) by applying cyclic boundary conditions.
The impeller rotation was modelled using the Multiple Reference Frame (MRF) impeller model. Initial simulations
performed using the transient Sliding Mesh model gave
only marginally improved results, as noted by Koh et al.
(2003) and Koh and Schwarz (2005), suggesting an unnecessary computational expense. For example, using equivalent grids and discretization schemes, power numbers
predicted by the two impeller rotation models diered by
under 2%. The MRF impeller model provides a frozen
eld or snapshot solution of the ow domain. The Wu
and Patterson (1989) experimental data used in this study
has the periodic component of turbulence removed and is
compared to the MRF baseline turbulence, corresponding
to a minimum in the angle-resolved turbulent kinetic
energy. The impeller blades, disc, baes and tank walls

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068


Table 2
Distribution of control volumes for Grids 1 to 4

Grid
Grid
Grid
Grid

1
2
3
4

Blade, r z

Total

22
41
63
84

42
84
126
168

36
66
102
136

68
12 16
18 24
24 32

33,000
230,000
800,000
1,900,000

were modelled as innitely thin surfaces with non-slip


boundary conditions. Fully structured, non-uniform grids
with hexahedral body-tted control volumes were used in
this study. In the region encompassing the impeller discharge stream, contained within 1.5 blade heights above
and below the impeller and extending horizontally across
the tank, the number of control volumes was increased to
resolve the large ow gradients expected in this region.
The grid was also rened near the impeller blade and bae
surfaces.
Grid sensitivity studies were conducted on four grids of
signicantly dierent resolution using upwind dierencing
(cf. Table 2). To investigate the eect of other discretization
schemes, simulations were run on the nest grid using the
upwind, central and QUICK discretization schemes.
Experiments have shown that the region where ow is
strongly inuenced by the periodic passage of the blades
extends to a radius of D/2 away from the impeller tip
and 1.5 blade heights above and below the impeller disc
(Lee and Yianneskis, 1994). Therefore, the inner volume
where the ow equations were computed in a rotating reference frame was dened using this boundary. The discretized equations were solved iteratively using the SIMPLEC
algorithm for pressure-velocity coupling and the solution
was considered converged when the total residuals for the
continuity equation, scaled by its largest absolute value
over the rst 5 iterations, dropped to below 10 4.
3. Results and discussion
3.1. Flow eld prediction
An accurate CFD model should be able to predict characteristic features of a ow eld in a stirred tank. The
model was able to predict the typical ow patterns
observed for a Rushton turbine in a baed tank using even
the coarsest grid (e.g. radial discharge stream, upper and
lower circulation zones, bae eddies). However, more subtle ow phenomena, such as the formation of the trailing
vortex pair generated at the tips of the impeller blade, were
not resolved under these conditions. Fig. 2 shows mean
velocity vectors in the vicinity of the impeller blade for
Grids 2, 3 and 4. As can be seen, the ability of the CFD
model to predict the trailing vortex pair depends on the resolution of the grid. On the vectors of the grids consisting of
830,000 and 1,900,000 control volumes, the trailing vortex
pair is clearly evident. However, this phenomenon is not
resolved on the second coarsest grid containing 230,000

1063

control volumes. These results suggest that it is possible


to predict the general ow eld in a stirred tank using relatively coarse grids but predicting more subtle phenomena
in the ow eld requires considerably ner grids.
3.2. Power number prediction
An accurate CFD model should be able to predict
important parameters such as the overall power input to
a stirred tank. The impeller power number Np has been
commonly used to check the validity of CFD simulations
of ow in stirred tanks (Dyster et al., 1993; Brucato
et al., 1998; Bartels et al., 2001). Consequently, power numbers were calculated from the predicted torque for a range
of Reynolds numbers covering laminar and turbulent ow
regimes and compared to the experimental data of Rushton
et al. (1950). Table 3 shows the inuence of dierent modelling approaches on the predicted power number for
NRe = 40,000. The power number predicted by the CFD
model improves signicantly as the grid resolution is
increased, the change being over 20% between the coarsest
and nest grids. The power number prediction also
improves as the order of the discretization scheme is
increased, the change being about 6.5% between the
upwind and QUICK schemes. The best prediction is about
10% lower than the experimental value. A possible reason
for this discrepancy is that the impeller blades and disc
were modelled as innitely thin walls. Studies have shown
that the impeller blade thickness inuences the power number (Rutherford et al., 1996).
Fig. 3 shows a loglog plot of the predicted power numbers for the full range of Reynolds numbers using the nest
grid and the QUICK discretization scheme. The power
numbers predicted by the CFD model over the full range
of impeller Reynolds numbers shows the typical power
curve relationship. The power numbers are in good agreement with the experimental data of Rushton et al. (1950),
with a slight under-prediction at higher Reynolds numbers
as noted previously. From these results one may conclude
that the grid resolution and discretization scheme both
inuence the accuracy of the predicted power number
and that it is possible to accurately predict power number
using very ne grids and high-order discretization schemes.
3.3. Mean velocity prediction
An accurate CFD model should be able to predict the
mean uid velocity components in a stirred tank. Fig. 4
shows axial proles of the normalized mean radial and tangential velocities near the impeller for dierent grids. The
velocity proles predicted with the coarsest grid (Grid 1)
deviate substantially from the results of the ner grids.
However, the results of the three ner grids are comparable
with each other and the experimental data of Wu and Patterson (1989). The results of the two nest grids are almost
identical suggesting that further grid renement would
have no eect on the solution. The CFD model predictions

1064

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068

Fig. 2. Mean velocity vectors near the impeller for dierent grid resolutions.

Table 3
Modelled power numbers for dierent grid resolutions and discretization
schemes (NRe = 40,000)
Model

Np

Rushton et al. (1950)


Grid 1, upwind dierencing
Grid 2, upwind dierencing
Grid 3, upwind dierencing
Grid 4, upwind dierencing
Grid 4, central dierencing
Grid 4, QUICK

6.07
4.13
4.69
4.95
5.07
5.31
5.40

do deviate from the experimental data on the lower side of


the impeller (i.e. for z/H < 0.33). Here, the CFD results are
largely symmetrical whereas the experimental data is
slightly skewed towards the upper side of the impeller. In
addition, although the experimental data is shown as a continuous prole for illustrative purposes, the data actually
consists of 10 data points through which a smooth curve
has been interpolated. Consequently, the comparisons
between CFD model predictions and experimental data
in both this and subsequent sections should be interpreted

with the nature of the experimental data set in mind i.e.


results should be comparable rather than identical.
Fig. 5 shows axial proles of the normalized mean radial
and tangential velocities near the impeller for dierent discretization schemes. The velocity proles obtained with the
upwind, central and QUICK discretization schemes are
almost identical on this nest grid and are comparable with
the experimental data. From these results one may conclude that mean uid velocity predictions are not strongly
inuenced by the grid resolution or discretization scheme
provided a grid which is not too coarse is used. This is consistent with the ndings of researchers such as Ng et al.
(1998), Montante et al. (2001) and Aubin et al. (2004) as
discussed in the introductory section.
3.4. Turbulent kinetic energy prediction
An accurate CFD model should be able to predict turbulence parameters such as RMS uctuating velocities or
turbulent kinetic energy in a stirred tank. As noted previously, CFD models often predict mean uid velocities in
the tank with reasonably accuracy but tend to under or

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068

1065

1000

Grid 4
QUICK

Rushton et al. (1950)


Computation

NP

100

10

1
0.1

10

100

1000

10000

100000

NRe
Fig. 3. Comparison of experimental and modelled power numbers.

0.45

0.45

NRe = 40 000
Upwind Differencing
r/T = 0.185

NRe = 40 000
Upwind Differencing
r/T = 0.185
0.40

z/H

z/H

0.40

0.35

0.30

0.25
0.00

0.35

0.30

0.20

0.40

0.60

0.80

0.25
0.00

Uradial/Utip
Grid 1

Grid 2

0.20

0.40

0.60

0.80

Utangential/Utip
Grid 3

Grid 4

Wu and Patterson

Fig. 4. Normalised mean radial and tangential velocity proles near the impeller for dierent grid resolutions.

over-predict turbulence quantities i.e. turbulence is the


problem area. This is generally ascribed to deciencies in
the ke turbulence model rather than to numerical errors
associated with the grid resolution or discretization
scheme. Fig. 6 shows axial proles of the normalized turbulent kinetic energy near the impeller and in the impeller discharge stream for dierent grids. The proles for all the
grids show qualitative agreement with the experimental
data but there is a dramatic improvement in the predicted
turbulent kinetic energy as the grid resolution increases.
Here, grid independence is not yet achieved on even the n-

est grid of nearly 2 million control volumes. However,


although the results of the nest grid are comparable with
the experimental data near the impeller, the CFD model
still signicantly under-predicts the turbulent kinetic
energy in the impeller discharge stream. Nonetheless, it
should be noted that the rst-order upwind scheme was
used in the above simulations and this has been shown to
signicantly under-predict turbulence in the impeller discharge stream (Aubin et al., 2004).
Fig. 7 shows axial proles of the normalized turbulent
kinetic energy near the impeller and in the impeller dis-

1066

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068


0.45

0.45

NRe = 40 000
Grid 4
r/T = 0.185

NRe = 40 000
Grid 4
r/T = 0.185
0.40

z/H

z/H

0.40

0.35

0.30

0.35

0.30

0.25
0.00

0.20

0.40

0.60

0.25
0.00

0.80

0.20

Uradial/Utip
Upwind Differencing

0.40

0.60

0.80

Utangential/Utip
Central Differencing

QUICK

Wu and Patterson

Fig. 5. Normalised mean radial and tangential velocity proles near the impeller for dierent discretization schemes.

0.45

0.45

NRe = 40 000
Upwind Differencing
r/T = 0.185

NRe = 40 000
Upwind Differencing
r/T = 0.285
0.40

z/H

z/H

0.40

0.35

0.30

0.25
0.00

0.35

0.30

0.02

0.04

0.06

0.08

0.25
0.00

2
k/Utip

Grid 1

Grid 2

0.02

0.04

0.06

0.08

0.10

2
k/Utip

Grid 3

Grid 4

Wu and Patterson

Fig. 6. Normalised turbulent kinetic energy proles near the impeller and in the impeller discharge stream for dierent grid resolutions.

charge stream for dierent discretization schemes. It is


clear from this gure that, when using the higher-order central dierencing and QUICK discretization schemes on this
very ne grid, the prediction of turbulent kinetic energy in
the impeller discharge stream is substantially improved.
The CFD model predictions for the nest grid and the
two higher-order discretization schemes are comparable

with the experimental data. From these results one may


conclude that turbulent kinetic energy predictions are
strongly inuenced by the grid resolution and the discretization scheme with very ne grids and high-order discretization schemes required for accurate CFD predictions.
These represent conditions which are numerically intensive
and suggest that the poor predictions of turbulence

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068


0.45

0.45

NRe = 40 000
Grid 4
r/T = 0.185

NRe = 40 000
Grid 4
r/T = 0.285
0.40

z/H

z/H

0.40

0.35

0.30

0.25
0.00

1067

0.35

0.30

0.02

0.04

0.06

0.08

0.25
0.00

0.02

2
k/Utip

Upwind Differencing

0.04

0.06

0.08

0.10

2
k/Utip

Central Differencing

QUICK

Wu and Patterson

Fig. 7. Normalised turbulent kinetic energy proles near the impeller and in the impeller discharge stream for dierent discretization schemes.

obtained using the ke turbulence model, often noted in the


literature, may be due to numerical errors rather than inadequacies in the turbulence model.
4. Conclusions
The eects of grid resolution and discretization scheme
on the CFD simulation of uid ow in an impeller stirred
tank using the MRF impeller rotation model and the standard ke turbulence model have been investigated. The
CFD model predictions have been compared with the
experimental data of Rushton et al. (1950) and Wu and
Patterson (1989). The accuracy of the CFD model has been
evaluated in terms of the predicted ow eld, power number, mean velocity components and the turbulent kinetic
energy. From these results one may conclude that:
Flow eld: The general ow eld can be predicted using
relatively coarse grids but predicting more subtle phenomena in the ow eld, such as the formation of trailing vortices, requires very ne grids.
Power number: The accuracy of the power number prediction is strongly inuenced by the grid resolution and
the discretization scheme. It is possible to accurately
predict the power number using very ne grids and
high-order discretization schemes.
Mean velocity: The accuracy of the mean uid velocity
predictions are not strongly inuenced by the grid resolution or discretization scheme, provided a grid which is
not too coarse is used.
Turbulent kinetic energy: The accuracy of the turbulent
kinetic energy predictions are strongly inuenced by
the grid resolution and the discretization scheme with

very ne grids and high-order discretization schemes


required for accurate CFD predictions.
This suggests that computationally light conditions
may be used if only the mean uid ow is required but that
computationally intensive conditions are required for
accurate prediction of turbulence. In this study a grid consisting of nearly 2 million control volumes in one half of a
15 cm diameter stirred tank, combined with a high-order
discretization scheme, was required to accurately predict
the turbulent kinetic energy. These represent conditions
which are considerably more numerically intensive than
used in most similar studies and suggests that the poor predictions of turbulence obtained using the ke turbulence
model, often noted in the literature, may be due to numerical errors rather than inadequacies in the turbulence model.
This supports the ndings of researchers such as Wechsler
et al. (1999) and Aubin et al. (2004) who noted that very ne
grids and higher-order discretization schemes are necessary
to limit numerical errors in modelling turbulence.
Acknowledgements
Financial support was provided by the Mineral Processing Research Unit (MPRU) and Center for Research and
Applied Mechanics (CERECAM) at the University of
Cape Town.
References
Abujelala, M.T., Lilley, D.G., 1984. Limitations and empirical extensions
of the ke model as applied to turbulent conned swirling ows.
Chemical Engineering Communications 31, 223236.

1068

D.A. Deglon, C.J. Meyer / Minerals Engineering 19 (2006) 10591068

Armentante, P.M., Luo, C., Chou, C., Fort, I., Medek, J., 1997. Velocity
proles in a closed vessel: comparison between experimental LDV data
and numerical CFD predictions. Chemical Engineering Science 52,
34833492.
Aubin, J., Fletcher, D.F., Xuereb, C., 2004. Modelling turbulent ow in
stirred tanks with CFD: the inuence of the modelling approach,
turbulence model and numerical scheme. Experimental Thermal and
Fluid Science 28, 431445.
Bartels, C., Breuer, M., Wechsler, K., Durst, F., 2002. Computational
uid dynamics applications on parallel-vector computers: computations of stirred vessel ows. Computers and Fluids 31, 6997.
Brucato, A., Ciofalo, M., Grisa, F., Micale, G., 1998. Numerical
prediction of ows in baed stirred vessels: a comparison of
alternative modelling approaches. Chemical Engineering Science 53,
36533684.
Costes, J., Courdec, J.P., 1988. Study by laser Doppler anemometry of the
turbulent ow induced by a Rushton turbine in a stirred tank:
Inuence of size of units, mean ow and turbulence. Chemical
Engineering Science 43, 27512764.
Derksen, J.J., Van Den Akker, H.E.A., 1999. Large eddy simulations on
the ow driven by a Rushton turbine. AIChE Journal 45, 209221.
Dyster, K.N., Kousakos, E., Jaworski, Z., Nienow, A.W., 1993. An LDA
study of the radial discharge velocities generated by a Rushton turbine:
Newtonian uids, Re P 5. Transactions of IChemE 71, 1123.
Eggels, J.M.G., 1996. Direct and large eddy simulations of the turbulent
uid ow using the lattice-Boltzman scheme. International Journal of
Heat and Fluid Flow 17, 307323.
Hartmann, H., Derksen, J.J., Montavon, C., Pearson, J., Hamill, I.S., Van
den Akker, H.E.A., 2004. Assessment of large eddy and RANS stirred
tank simulations by means of LDA. Chemical Engineering Science 59,
24192432.
Jaworski, Z., Zakrzewska, B., 2002. Modelling of the turbulent wall jet
generated by a pitched blade turbine impeller: the eect of turbulence
model. Transactions of IChemE 80, 846854.
Jenne, M., Reuss, M., 1999. A critical assessment on the use of ke
turbulence models for simulation of the turbulent liquid ow induced

by a Rushton-turbine in a baed stirred-tank reactor. Chemical


Engineering Science 54, 39213942.
Koh, P.T.L., Schwarz, M.P., 2005. CFD Modelling of bubble-particle
attachments in a otation cell. In: Proc. Centenary of Flotation
Symposium, Brisbane, pp. 201207.
Koh, P.T.L., Schwarz, M.P., Zhu, Y., Bourke, P., Peaker, R., Franzidis,
J.P., 2003. Development of CFD models of mineral otation cells. In:
Proc. Third International Conference on CFD in the Minerals and
Process Industries, Melbourne, pp. 171175.
Launder, B.E., Spalding, D.B., 1974. The numerical computation of
turbulent ows. Computer Methods in Applied Mechanics and
Engineering 3, 269289.
Lee, K.C., Yianneskis, M., 1994. The extent of periodicity of the ow in
vessels stirred by Rushton impellers. AIChE Symposium Series 90, 5
18.
Montante, G., Lee, K.C., Brucato, A., Yianneskis, M., 2001. Numerical
simulations of the dependency of ow pattern on impeller clearance in
stirred vessels. Chemical Engineering Science 56, 37513770.
Ng, K., Fentiman, N.J., Lee, K.C., Yianneskis, M., 1998. Assessment of
sliding mesh CFD predictions and LDA measurements of the ow in a
tank stirred by a Rushton turbine. Transactions of IChemE 76, 737
747.
Ranade, V.V., Joshi, J.B., Marathe, A.G., 1989. Flow generated by
pitched blade turbines II: simulation using ke model. Chemical
Engineering Communications 81, 225248.
Rushton, J.H., Costich, E.W., Everett, H.J., 1950. Power characteristics of
mixing impellersPart II. Chemical Engineering Progress 46, 467476.
Rutherford, K., Mahmoudi, S.M., Lee, K.S., Yianneskis, M., 1996. The
Inuence of Rushton impeller blade and disc thickness on the mixing
characteristics of stirred vessels. Transactions of IChemE 74, 369378.
Wechsler, K., Breuer, M., Durst, F., 1999. Steady and unsteady
computations of turbulent ows induced by a 4/45 pitched-blade
impeller. ASME Journal of Fluids Engineering 121, 318328.
Wu, H., Patterson, G.K., 1989. Laser-Doppler measurements of turbulent-ow parameters in a stirred mixer. Chemical Engineering Science
44, 22072221.

You might also like