Sie sind auf Seite 1von 28

Chemical Engineering Science 63 (2008) 5468-- 5495

Contents lists available at ScienceDirect


Chemical Engineering Science
journal homepage: www. el sevi er . com/ l ocat e/ ces
Assessment of standard kc, RSMandLES turbulence models ina baffledstirredvessel
agitated by various impeller designs
B.N. Murthy, J.B. Joshi

Department of Chemical Engineering, Institute of Chemical Technology, University of Mumbai, Matunga, Mumbai 400 019, India
A R T I C L E I N F O A B S T R A C T
Article history:
Received 25 September 2007
Received in revised form 22 May 2008
Accepted 24 June 2008
Available online 2 July 2008
Keywords:
Stirred vessel
CFD
Disc turbine
Pitched blade turbine
Hydrofoil
RANS
LES
LDA
In the present work, laser-doppler anemometry measurements as well as CFD simulations have been
performed for the flow generated by various impellers, namely disc turbine (DT), a variety of pitched
blade down flow turbine impellers varying in blade angle (Standard PBTD60, 45 and 30) and hydrofoil
(HF) impeller. The tank was fully baffled, and the flow regime was turbulent. The objective of the present
work was to carry out a detailed investigation of the predictive capabilities of the various turbulence
models, i.e. the standard k c model, Reynolds-stress transport model (RSTM) and large eddy simulations
(LES). In case of LES, effect of subgrid scales on the resolved scales has been modeled by dynamic one
equation subgrid-scale model. The simulated values of the mean axial, radial and tangential velocities
along with the turbulent kinetic energy have been compared with the measured LDA data. It has been
identified that the present SGS LES model performs well for predicting all the flow variables. Whereas,
RSM and standard k c model underpredict the turbulent kinetic energy profiles significantly in the
impeller region. RSM can capture well all the mean flow characteristics and the standard k c model fails
to simulate the mean flow associated with the strong swirl. Energy content of the precessional vortex has
been quantified for all the five impeller designs. Intermediate frequencies inbetween the mean circulation
and the precession instability have been identified having a non-dimensional frequency of 0.04 to 0.07
for all the impeller designs under consideration.
2008 Elsevier Ltd. All rights reserved.
1. Introduction
Stirred vessels are extensively used in the chemical process in-
dustries over a wide range of applications. In majority of cases, the
flow field in baffled stirred vessels is highly turbulent hence it is
three-dimensional, complex and chaotic in nature. During the last
25 years, there have been continuous efforts on understanding these
flows using both sophisticated experimental and computational fluid
dynamics tools.
In view of the above, in the past, the flow generated by disc tur-
bine (DT) and pitched blade turbines (PBT) have been subjected to
detailed experimental and computational studies. The initial objec-
tives were preliminary such as the estimation of gross flow param-
eters and the average flow field. However, during the past 10 years,
the objectives have become deeper such as the characterization of
turbulent flow, flow instabilities, etc., and tailoring the impeller de-
sign so as to get the desired flow field. Majority of the work has
been devoted to the modeling of the mean and the turbulence flow
* Corresponding author. Tel.: +912224145616; fax: +912224145614.
E-mail address: jbj@udct.org (J.B. Joshi).
0009-2509/$ - see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2008.06.019
field and its comparison with the experimental flow measurements.
However, the comparison of predictive capabilities of various turbu-
lence models has been preliminary even for the flow generated by
conventional impeller designs like DT and PBT. It is well known that
the k c model assumes the isotropy for turbulence, and therefore,
anisotropic models such as Reynolds stress model (RSM) and the
large eddy simulation (LES) model are being recommended for the
simulation of complex three-dimensional flows. However, RSM has
got shortcomings like, non-universal model parameters, numerical
difficulties and is computationally expensive by an order of magni-
tude as compared to the k c model. Further, the RSM model does
not capture the time dependent nature of the flow. This limitation is
overcome by the LES approach. During the last few years, the abil-
ity to resolve all but the smallest turbulence scales using LES has
become more viable. LES can potentially produce more accurate re-
sults by modeling only the smallest scales, which tend to be more
isotropic, while fully resolving the turbulence at the larger scales.
From a practical point of view, the use of LES or DNS as design
tools is far from easy due to the high computational costs associated
with them. For these reasons, it is envisioned that the RANS equations
associated with turbulence modeling will be the main CFD tool used
by the practitioners and part of the research community, at least in
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5469
the near future. As a result of this trend, there is a need to improve
the accuracy and reliability of the solutions of turbulent flow fields
obtained from the RANS equations. In light of this discussion, the
RANS models can be improved by selecting the model parameters
on the basis of the understanding of mean and the turbulent flow
fields. For this purpose, it is desirable that, an extensive benchmark
database is made available with equally extensive exercise of the
comparative performance of these turbulence models. Therefore, one
of the main objectives of the present work is to present a comparative
study of k c, RSM and LES with inhouse LDA measurements.
2. Previous work
As regards to standard k c modeling, extensive work (Ranade
et al., 1992; Jenne and Reuss, 1999; Brucato et al., 1998; Jaworski and
Zakrzewska, 2001; Joshi and Patwardhan, 1999; Nere et al., 2001,
2003; Aubin et al., 2004) has been reported in the published litera-
ture. Hence, the present work reviews the past computational work
on the flow generated by DT and PBT using RSM and LES models.
2.1. Reynolds stress modeling
There have been two attempts partially focused on the RSM for
the flow generated by Rushton turbine in a baffled stirred vessel.
Table 1
Previous work on LES of the stirred tank
Authors, Impeller-SGS
model, Grid size
System investigated Flow
variable
Region of comparison Remarks
Eggels (1996), Adaptive force field
technique to model the action
of impeller, Standard Smagorinsky
model, 1.73 10
6
, 13.8 10
6
Rushtine turbine, T=0.48m, H=
T, C/T = 0.33, D/T = 0.33, N =
250rpm, Re =107, 000
u1 0.144< z/w< 0.176 Qualitative flow was presented. Comparison was given
only in the impeller region. Tangential velocity and turbu-
lent kinetic energy comparison was missing. Energy bal-
ance was missing.
u
1
0.37< r/R< 0.78
0< r/R< 1
z =0
u2 0< z/T < 0
u
2
r/R =0.34
Revstedt et al. (1998), Impeller
motion was modeled by specify-
ing momentum source term
Rushtine turbine, T=0.44m, H=
T, C/T = 0.5, D/T = 0.33, N =
165rpm, Re =60, 000
u1 z/T =0.33 Qualitative flow features were presented. Flow and pump-
ing numbers have been in good agreement with exper-
imentally measured values. Mean axial velocity was in
agreement with the experimental data all over the do-
main. The mean radial velocity and turbulent kinetic en-
ergy were compared only in the impeller region. Spectral
analysis was reported.
0< r/R< 1
u2 0.25< z/T < 0.75
0< r/R< 1
No sgs model, 32,768, 262,144 K z/T =0.33
0< r/R< 1
Derksen and Van den Akker
(1999), Adaptive force field
technique to model the action of
impeller, Standard Smagorinsky
sgs model, 6 10
6
Rushtine turbine, T=0.48m, H=
T, C/T = 0.33, D/T = 0.33, N =
68rpm, Re =29, 000
u1 0.144< z/w< 0.176 Detailed information was given for phase resolved veloc-
ities, kinetic energy and dissipation rate distribution all
over the domain. The vortex paths, both above and below
the impeller disk, were predicted correctly. Axial velocity
comparison was not given. However, comparison was re-
stricted only to the impeller region. Power number was
predicted well.
r/R =0.5, 0.65
u3 0.144< z/w< 0.176
r/R =0.5, 0.65
K 0.144< z/w< 0.176
r/R =0.5, 0.65
Bakker et al. (2000), Sliding mesh
technique, standard smagorinsky's
sgs model, 5,27,000 for PBTD
7,63,000 for DT
PBTD: T = 0.292m, D/T =
0.35 C/T =0.46, N =60rpm, DT:
T = 0.202m, C/T = D/T =
1
3
, N =
290rpm
Realistic qualitative flow patterns have been presented and
compared with flow field snapshots of PIV. Quantitative
comparison was missing.
Revstedt et al. (2000), Impeller
motion was modeled by specify-
ing momentum source term,
Scale similarity sgs model, Scalar
transport was studied
6RT, 6SRGT: multiple impeller
system, T = 0.8m, H = 1.5T,
inter impeller clearance=0.25T,
bottom clearance = 0.5T. Re =
100, 000
u3 0< r/R< 1.0 This was the first study which employed multiple impeller
system. This study gave more importance to prediction of
global parameters (Np and NQP). The mean axial velocity
and turbulent kinetic energy were predicted well. Spectral
analysis has been carried out for the radial velocity com-
ponent. LES scalar mixing studies have been performed.
All the flow variables were not compared with experi-
mental data.
z/T =0
k 0< r/R< 1.0
z/T =0
Derksen (2001), Adaptive force
field technique in Lattice Boltz-
mann frame work. Standard
Smagorinsky sgs and structured
function sgs models, 180
3
, 240
3
and 360
3
4 bladed PBTD, Impeller-vessel
geometry similar to the one in-
vestigated by Schafer et al.
(1998), Re =7300
u3 z/T =0.145, 0.276 Spatial resolution has significant impact on the overall
average flow field results. The impact of a slightly differ-
ent subgrid-scale (SGS) model (a structure function model
instead of a Smagorinsky model) was studied. No signifi-
cant effect change in the overall and phase-averaged flow
field results due to change in subgrid scale model. Only
the formation and strength of the tip vortex differed for
the two SGS approaches (vortex formed in the structure
function simulations was found to be weaker). Validation
was made only with in the impeller region.
0< r/R< 1.0
u2 z/T =0.145, 0.276
0< r/R< 1.0
k z/T =0.145, 0.276
0< r/R< 1.0
First one, Bakker and Van den Akker (1998) employed the simpli-
fied RSM, i.e. algebraic stress model (ASM) using the IBC method
to model the flows produced by a Rushton turbine. Their objective
was to improve the predictive capabilities of CFD modeling by ac-
counting anisotropy using less computational intensive ASM model
(simplified RSM). Their study concluded that the results predicted by
the ASM compare better with the experimental data than those pre-
dicted by the standard k c model. Oshinowo et al. (2000) performed
the CFD study using different turbulence models like, k c, RNG k c
and RSM for the prediction of tangential velocity distribution in a
baffled vessel using multiple reference frame (MRF) model. The tan-
gential velocity distribution above the impeller has been correctly
predicted. They attributed the occurrence of the counter-intuitive
reverse swirl in the simulation to poor convergence and coarse grid
density. It may be pointed out, however, that both the aforesaid in-
vestigations have shown comparisons in only a small region while
most of the vessel region remained unexplored. More importantly,
both the studies have not presented the predictive capability of CFD
models for the turbulent kinetic energy and the turbulent energy
dissipation rate.
2.2. Large eddy simulations
Table 1 summarizes the previous work on the LES of the flow
pattern generated in a stirred vessel. It gives details with regard to
5470 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
Table 1 (Continued.)
Authors, Impeller-SGS model, Grid
size
System investigated Flow
variable
Region of
comparison
Remarks
Revstedt et al. (2002), Impeller mo-
tion was modeled by specifying mo-
mentum source term, Scale
similarity sgs models
6RT, 6SRGT: multiple impeller
system, T = 0.8m, H = 1.5T,
inter impeller clearance=0.25T,
bottom clearance = 0.5T. Re =
100, 000
u1 0< r/R< 1 They presented results both in a fixed and moving frame
of reference. There was a slight over prediction of the
mean tangential velocity component due to insufficient
grid resolution. Comparison was made between Rushton
and Scaba turbines using their predictions. Finally, this
study explained the high levels of turbulence of flow gen-
erated by Rushton than Scaba turbine with the help of
spectral density of radial fluctuating component obtained
from LES.
z/T =0.33
u3 0< r/R< 1
z/T =0.33
k 0< r/R< 1
z/T =0.33
Roussinova et al. (2003), Sliding
mesh technique, Standard
Smagorinsky sgs model, 500,000
PBTD454, T = 0.24, D/T = 0.5,
W/D=0.2, dh/D=0.22, h/D=0.20,
C/T =0.5, N =200, Re =48, 000
u1 z/T =0.5 This study compared predicted values of the radial profile
of the mean axial velocity. It focused on the identification
of low frequency macro instabilities. Their predicted value
of frequency of macro instability was compared well with
LDA measurement.
0< r/R< 1
Yeoh et al. (2004), Sliding deform-
ing mesh technique, Standard
Smagorinsky sgs model, 490,000
DT6, T =0.10, D/T =0.33, C=T/3,
H=T, N=36.08rps, Re =40, 000
u1 0< r/R< 1 The predictions were compared with LDA data of mean
and rms velocities and energy dissipation. Both standard
k c and LES models predicted the mean flow field well.
LES out performed standard k c with regard to the k
predictions. The power number obtained from integration
of the energy dissipation was predicted to within 15% of
the measured value. This study also made the comparison
only in the impeller region.
z/T =0.33
u3 0< r/R< 1
z/T =0.33
k 0< r/R< 1
z/T =0.33
c 0< r/R< 1
z/T =0.33
Hartmann et al. (2004a), Standard
Smagorinsky and Voke sgs models,
Adaptive force field technique
T = 0.15, D = T/3, C = T/3, H =
T, baffle clearance=0.017T, N=
7rps, Re =7300
u1 0.52< z/R< 0.8 Predictive capabilities of LES and RANS were assessed. The
simulated phase-averaged flow fields were in good agree-
ment with the experimental results. LES overpredicted
the tangential velocity at the center of the impeller tip.
Their simulations revealed that Smagorinsky sgs model
performs better than Voke sgs model especially when it
comes to prediction of trailing vortex pair. It was shown
that LES better predicted turbulent kinetic energy in the
impeller discharge flow where RANS underperforms. They
found nearly isotropy in the circulation loops. But in the
impeller stream, the boundary layers, and at the separa-
tion points turbulence was found more anisotropic due to
high shear rate. They did not see relative merits of using
Voke sgs model.
0.366< r/R< 0.64
u3 0.52< z/R< 0.8
0.366< r/R< 0.64
k 0.52< z/R< 0.8
0.366< r/R< 0.64
Yeoh et al. (2005), Sliding deform-
ing mesh technique, Standard
Smagorinsky sgs model, 490,000
DT6, T =0.10, D/T =0.33, C=T/3,
H=T, N=36.08rps, Re =40, 000
This was the first attempt to study the mixing time. LES
provided a very detailed evolution pattern of the concen-
tration field in space and time. At the impeller midsection
a substantial amount of scalar is trapped in a region be-
tween two vertical baffles. In the impeller region it was
observed that instantaneous concentrations being as high
as three times the equilibrium value. They cited that LES
can be used to identify stagnant zones, mixing inhomo-
geneities due to the vessel geometry, and also for optimal
location of feed pipe.
Alcamo et al. (2005), standard
Smagorinsky sgs model, sliding
mesh technique, 761,760
DT6, unbaffled vessel, T = 0.19,
D/T = 0.5, C = T/3, H = T, N =
200rpm, Re =3 10
4
, unbaffled
vessel
u1 0.26< r/R< 0.44 The LES predictions were compared with experimental
data on an unbaffled tank. The simulated values compared
very well with the measured data as far as tangential ve-
locity are concerned. And a fairly good comparison ob-
tained for radial velocity profiles as well. The LES could
captured the existence of pair of trailing vortices in un-
baffled tanks.
0.033< z/T < 0.16
u3 0.27< z/R< 0.105
0.033< z/R< 0.16
u2 0.033< z/R< 0.16
Derksen et al. (2007), standard
Smagorinsky and mixed sgs models,
adaptive force field technique, 80
3
,
128
3
and 160
3
DT6, non-standard geometry,
T = 0.44, H = 1.86T, D =
2
3
T,
C=0.0375T, Re=14, 000, 82,000,
350,000
There was no significant difference between the perfor-
mance of the two sgs models. Further, the particle track-
ing technique has been extended for the study of mixing
performance. They reported that after some 510 impeller
revolutions uniform conditions have been obtained, except
for the bottom region that is characterized by the pres-
ence of fresh feed. However, LES predictions have been
discussed more qualitatively as non-availability of the ex-
perimental data for the industrial scale equipment.
the vessel and the impeller geometry and the subgrid scale model
along with the grid size. The table also gives remarks on the qual-
ity of comparison of predicted radial, axial, and tangential velocity
components, and turbulent kinetic energy (k) with the experimental
data.
The very first study of LES for stirred vessel was performed by
Eggels (1996) using a lattice-Boltzmann discretization scheme and
the adaptive force field technique for the impeller rotation. The sim-
ulations have been performed for coarse and fine mesh and the
corresponding uniform grid size of 4mm (1.73 10
6
) and 2mm
(13.8 10
6
), respectively. 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 lo-
cal flow field in the vicinity of the impeller which was not possible
with the RANS based turbulence models. Therefore, this study pre-
sented the instantaneous and the mean flowfield in both vertical and
horizontal planes which would help in initial understanding of the
given system. However, there were differences in the quantitative
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5471
Fig. 1. Impeller designs used for study. (A) DT; (B) PBTD60; (C) PBTD45; (D) PBTD30 and (E) hydrofoil impeller.
predictions of the mean radial and mean axial velocities especially
in the near impeller region. The discrepancies were due to insuffi-
cient statistical sampling time which was limited by the comput-
ing time with the fine grid and adaptive force field technique used
for the impeller modeling. Further, the quantitative comparison has
been restricted to a few flow variables (axial and radial velocity) and
also at limited locations (one radial and one axial line) and energy
balance was not reported.
Revstedt et al. (1998) investigated both the mean flow and the
spectral distribution of flow by power density function at various
locations in a baffled strirred tank. In this study, the effect of sgs
scale motion on the large scale was not taken into account and the
process of energy dissipation was numerically treated by using third
order upwind scheme. The LES could capture the trailing vortices
behind the blades and their detailed movement away from the im-
peller with increasing distance from the blade. These observations
were in close agreement with the expensive and time-consuming
experiment. They demonstrated that the LES could generate instan-
taneous flow field like sophisticated flow measuring techniques and
sometimes even better as far as the flow in the impeller region and
near wall regions are concerned. The mean axial velocities have been
better predicted. Comparison for the mean radial velocity at only one
axial location was given (impeller center line), which was very good
at all the locations except 0.5< x
1
/D< 0.6. They have also obtained a
fairly good prediction for the turbulent kinetic energy in the impeller
center plane. Acceleration of the fluid at the impeller tip because of
the fluid entrainment due to the trailing vortex pair was well simu-
lated. Prediction of radial and tangential velocity with the angle was
poor. They have suggested that the boundary conditions on the im-
peller region should be improved. Their instantaneous data from LES
supported the blade frequency and the existence of Kolmogorov's

5
3
region in the energy spectrum. However, the detailed comparison
of all the flow variables with the experimental data throughout the
vessel has not been reported.
Derksen and Van den Akker (1999) took forward the Eggels (1996)
study with a more refined forcing algorithm. The real potential of
LES was explored as the simulations could identify a detailed lo-
cal flow in the wakes behind each blade which were found to be
5472 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
D/T = 1/3
H = T
C
T/10
z
r
T = 0.30m
Z = 0.010m
Z = 0.044m
Z = 0.082m
Z = 0.10m
Z = 0.118m
Z = 0.154m
Z = 0.244m
Z = 0.190m
D
Fig. 2. Geometrical details of the stirred vessel.
-0.2
-0.1
0
0.1
-0.4
-0.2
0
0.2
0.4
-0.4
-0.2
0
0.2
0.4
-0.2
-0.1
0
0.1
0.2
0.3
-0.4
-0.2
0
0.2
0.4
-0.4
-0.2
0
0.2
0.4
-0.4
-0.2
0
0.2
0.4
-0.2
-0.1
0
0.1
0.2
0
NORMALISED RADIAL COORDINATE, r/R (-)
1
2
3
1
2
3
1
2
3
2
3
1
3
1
2
1
2
3
1
2 3 1
2
3
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

A
X
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
2
>
/
U
t
i
p
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.5 0.7 0.9
Fig. 3. Comparison between the simulated and experimental profiles of the dimensionless mean axial velocity for disc turbine at various axial levels. (A) H = 0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
significantly higher than the impeller tip speed (two times tip speed).
As far as the quantitative comparison is concerned, good agreement
has been obtained between the experimental data and the simula-
tion of the radial profiles of the radial velocity. However, the tan-
gential velocities were overpredicted (by maximum 15%) and it was
attributed to a lack of spatial resolution and improper treatment of
the SGS viscosity. The axial profiles of random (turbulent), coher-
ent (pseudoturbulence) and total kinetic energy have been shown
to agree with the LDA phase resolved experimental data. Further, it
was observed that the dissipation rate distribution throughout the
tank is very inhomogeneous. However, the comparison for axial ve-
locity was not presented and all the predictions have been reported
only in the impeller region.
Revstedt et al. (2000, 2002) have made the first attempt to study
the flow generated in multiple impeller system and the spatial dis-
tribution of inert tracer using the LES. The objective of this study
was to investigate the influence of the impeller type (Rushton and
Scaba) on the flow structure and scalar transportation. This study
gave relatively more emphasis to the prediction of global param-
eters (N
P
and N
QP
) rather than the detailed estimation of mean
and turbulent flow field. Further, only axial velocity and turbulent
kinetic energy were compared with the experimental data (by con-
stant temperature anemometry in the impeller center plane). In
addition, there were significant discrepancies in the prediction of
turbulent kinetic energy in the impeller region. They attributed it to
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5473
-0.4
-0.3
-0.2
-0.1
0
-0.16
-0.12
-0.08
-0.04
0
-0.05
0
0.05
0.1
0.15
0
0.2
0.4
0.6
0.8
1
-0.1
0
0.1
0.2
0.3
-0.02
-0.01
0
0.01
0.02
-0.03
-0.02
-0.01
0
0.01
0.02
0.03
0
-0.1
-0.05
0
0.05
0.1
NORMALISED RADIAL COORDINATE, r/R (-)
2 3
1
1 2
3
2
1
3
1
2 3
1
2
3
1
2
3
3
1 2
1
3
2
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

R
A
D
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
1
>
/
U
t
i
p
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
Fig. 4. Comparison between the simulated and experimental profiles of the dimensionless mean radial velocity for disc turbine at various axial levels. (A) H = 0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
the improper treatment of impeller rotation and inadequate mesh
resolution. However, they have clearly demonstrated the advantages
of LES over RANS using the data generated from LES by carrying
out the spectral analysis of both the velocity and scalar flow field. It
revealed the large spatial inhomogeneities in the concentration fields
and low-frequency variations seemed to dominate the concentration
fluctuations in the impeller stream. They concluded that, at equal
power input, the center plane velocities and the volumetric flow
in the impeller stream did not differ with respect to the impeller
type (DT and Scaba). Also, for the Scaba type impeller, the periodic
fluctuations associated with the blade passing, were found to be less
pronounced due to the absence of trailing vortices.
Derksen (2001) employed the previously developed lattice-
Boltzmann based solver for the flow generated by 4-PBTD45. This
study focused on the influence of the spatial resolution and sgs
modeling on the flow predictions. Furthermore details can be seen
in Table 1. It has been observed that the spatial resolution has signif-
icant impact on the overall average flow field results. No significant
effect was observed in the overall and phase-averaged flow field re-
sults due to a change in the subgrid scale model. Only the formation
and strength of the tip vortex differed for the two SGS approaches
(vortex formed in the structure function simulations was found to
be weaker). Validation was made only within the impeller region.
For the PBTD45 impeller, Roussinova et al. (2003) performed LES
simulations using sliding mesh technique. The mean axial veloc-
ity in the impeller center plane has been in agreement with the
experimentally measured values. The study was extended for the
identification of low frequency macroinstabilities. Their predicted
value of frequency of macroinstability agreed very well with those
obtained from LDA measurement. However, the comparison of the
LES predicted mean radial, and tangential velocities and the tur-
bulent kinetic energy with the experimental data have not been
reported.
Hartmann et al. (2004a) employed LES to investigate the preces-
sional vortex phenomenon with the help of LES. This study opted the
same numerical techniques as those of Derksen and Van den Akker
(1999). Their LES model could capture the vortical structure moving
around the tank centerline in the same direction as the impeller. Us-
ing LES, they also observed that the strength of the vortex below the
impeller to be much stronger and pronounced in size as compared
to that prevailing above the impeller and both vortices move with a
mutual phase difference which was qualitatively in good agreement
with the experimental data. Their LES model could accurately predict
the characteristic MI (macroinstability) frequency. It could even pre-
dict a second frequency peak at f =0.092N at a Reynolds number of
12,500. Using LES data, they observed that the flow was dominated
5474 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
-0.05
0
0.05
0.1
0
-0.06
-0.01
0.04
0.09
0.14
-0.05
0
0.05
0.1
0.15
0
0.1
0.2
0.3
0
0.05
0.1
0.15
0.2
0
0.02
0.04
0.06
0.08
0.1
0
0.03
0.06
0.09
0
0.02
0.04
0.06
0.08
0.1
NORMALISED RADIAL COORDINATE, r/R (-)
1
2
3
1
2
3
1
2
3
1
2
3
1 2 3
3
1
2
2 3
1
3
1 2
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

T
A
N
G
E
N
T
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
3
>
/
U
t
i
p
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
Fig. 5. Comparison between the simulated and experimental profiles of the dimensionless mean tangential velocity for disc turbine at various axial levels. (A) H=0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
by the low-frequency precessional vortex in the bulk flowregion and
it was associated with a significant amount of the kinetic energy.
Yeoh et al. (2004) have performed LES using sliding deforming
technique. The main objective was to assess the capability of the LES
modeling technique, coupled with the sliding deformation method-
ology (SDM) and relative performance with RANS approach. RANS
have been performed using the standard k c model. The predicted
mean radial and tangential velocities and turbulent kinetic energy
were compared with LDA data only in the impeller center plane. For
all the three variables good qualitative and quantitative agreement
was obtained using both the RANS and the LES models. However,
LES made superior predictions of the correct radial profiles than the
standard k c model. With regard to the prediction of the turbulent
kinetic energy, LES clearly outperformed the standard k c model.
The power number obtained from integration of the energy dissi-
pation was predicted within 15% of the experimentally measured
value. Similar to earlier investigations, this work has also presented
the comparison at only one axial location (impeller center plane).
Hartmann et al. (2004b) made a detailed study to evaluate the
predictive capabilities of LES and RANS. The main difference between
this study and Yeoh et al. (2004) was, this study employed the lattice-
Boltzmann solver as against the finite volume based solver. Further,
the RANS based simulations were performed with the shear-stress-
transport (SST model) in place of standard k c model. The RANS, LES
predictions have been compared with the LDA measured radial, tan-
gential velocities and turbulent kinetic energy in the impeller zone.
The radial velocities of both the models were in good agreement
with the experimental data. With regard to the tangential velocity,
the overall performance of LES was found to be much superior to
RANS simulations. Their simulations revealed that the Smagorinsky
sgs model performed better than Voke sgs model, especially, as far
as the prediction of trailing vortex pair is concerned. It was further
shown that the LES predictions of the turbulent kinetic energy in
the impeller discharge flow were substantially better than the RANS
which invariably underperformed. Overall they did not see the rel-
ative merits of using Voke sgs model. Further, they calculated the
energy dissipation rate by assuming local equilibrium between pro-
duction and dissipation at and below subgrid scale level and com-
pared with the RANS predictions in the baffle midplane. This study
also presented comparisons in a limited region. Further, axial veloc-
ity comparison has not been given.
Yeoh et al. (2005) extended their previous work to study the
mixing of inert tracer using LES. This was the first attempt to study
the detailed mixing phenomenon using LES. It was clear from the
study that the LES provides a very detailed evolution pattern of the
concentration field in space and time. It was at the impeller mid-
section that a substantial amount of scalar gets trapped in a region
between two vertical baffles. In the impeller region they observed
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5475
0
0.005
0.01
0.015
0
0
0.005
0.01
0.015
0
0.01
0.02
0.03
0.04
0.05
0
0.05
0.1
0.15
0.2
0
0.02
0.04
0.06
0.08
0.1
0
0.005
0.01
0.015
0.02
0
0.004
0.008
0.012
0
0.003
0.006
0.009
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

T
U
R
B
U
L
E
N
T

K
I
N
E
T
I
C

E
N
E
R
G
Y
,

k
/
U
t
i
p

(
-
)

1
2
3
1
2
3
3
1
2
3
1
2
1
2
3 2
1
3
2 1
3
2
1
3
2
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Fig. 6. Comparison between the simulated and experimental profiles of the dimensionless turbulent kinetic energy profiles for disc turbine at various axial levels. (A)
H = 0.01m; (B) H = 0.044m; (C) H = 0.082m; (D) H = 0.1m; (E) H = 0.118m; (F) H = 0.154m; (G) H = 0.190m and (H) H = 0.244m: Experimental; (1) LES model; (2)
RSM and (3) k c model.
instantaneous concentrations being as high as three times the steady
state value. They have clearly brought out the utility of LES in identi-
fying the stagnant zones, mixing inhomogeneities due to the vessel
geometry and optimal location for the feed pipe.
Alcamo et al. (2005) presented the predictive capabilities of LES
for flow field generated in an unbaffled cylindrical vessel. The LES
predictions were better compared with the experimental data ob-
tained by both inhouse PIV technique and with the available litera-
ture data. The simulated values agreed very well with the measured
tangential velocities and fairly well with the radial velocities. The LES
could capture the existence of pair of trailing vortices in unbaffled
tanks.
Recently Derksen et al. (2007) extended their previous lattice-
Boltzmann scheme based LES solver for industrial scale crystallizer.
It has been focused on the assessing the feasibility of using a
computationally efficient LES to quantify the fine scale turbulent
structures in an industrial size crystallizer. The effects of spatial
resolution and the subgrid scale model (Smagorinsky and the mixed
scale models) were also investigated. No significant difference was
observed in the predictions of the two subgrid scale models. Further,
the particle tracking technique was employed for the study of mix-
ing performance. Practically uniform conditions were obtained after
510 impeller revolutions with respect to concentration of fluid
elements, except for the bottom region which gets characterized
by the presence of fresh feed. However, LES predictions have been
discussed more qualitatively because of the non-availability of the
experimental data from the industrial scale equipment.
The foregoing reviewbrings out the superior potential of LES over
the RANS models. However, the LES predictions have been compared
with the experimental measurements in a limited region (less than
10% of the tank volume) close to the impeller. Further, the effect of
impeller design has been investigated over a limited range of power
number. Further, it cannot be taken for granted that, if any turbulence
model predicts the flow in one region (even most complex), it would
have the similar ability over the entire computational domain. If this
was true, the model parameters of k c, RSM which were selected
to get better prediction in the near impeller zone, the same parame-
ters should have been able to capture the flow quantitatively in the
bulk region. In order to overcome such a limitation, attempts have
been made for the standard k c model to identify the set of model
parameter values in different zones, i.e. zonal modeling. The impor-
tant message is to realize the existence of non-similar featured flows
in different zones. For instance, the literature reports clearly show
that various flow structures with certain characteristics (existence,
shape and energy content) appear in different regions in the tank.
For example, some features include processional vortex motion near
shaft, trailing vortex behind the blade and jet instabilities (belowim-
peller region). In addition, these characteristic features strongly de-
pend upon the impeller design. Therefore, it was thought desirable
to investigate the flow patterns generated by a wide spectrum of
5476 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
-0.2
-0.1
0
0.1
0.2
-0.5
-0.3
-0.1
0.1
0.3
0.5
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0
-0.4
-0.2
0
0.2
0.4
0.6
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.1
-0.15
-0.1
-0.05
0
0.05
0.1
3
1
2
2
1
3
2
1
3
2
1
3
2
1
3
2
1
3
2
1
3
2
1
3
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

A
X
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
2
>
/
U
t
i
p
NORMALISED RADIAL COORDINATE, r/R (-)
0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Fig. 7. Comparison between the simulated and experimental profiles of the dimensionless mean axial velocity for PBTD60 at various axial levels. (A) H=0.01m; (B) H=0.044m;
(C) H =0.082m; (D) H =0.1m; (E) H =0.118m; (F) H =0.154m; (G) H =0.190m and (H) H =0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
open impellers. For instance, DT produces strong radial and tangen-
tial flow with complex trailing vortices, whereas the hydrofoil (HF)
impeller generates flow with practically weak trailing vortices and
less complexity. In view of such a wide possible variation, flows gen-
erated by five impeller designs namely DT, PBTD60, PBTD45, PBTD30
and an HF impeller have been investigated in the present work.
Further, in the present work, for the first time, comprehensive
RSM has been employed to simulate both radial and axial flow im-
pellers. In the literature, it is quite evident that the predictive ca-
pabilities of full fledged RS model has yet to receive due attention.
However, there have been quite a few studies, which have used al-
gebraic stress model (simplified RSM) which were excellent starting
points. Further, one of the significant terms in the Reynolds stress
transport equations is pressure strain term, which has been mod-
eled using quadratic pressure strain model, which has been demon-
strated to give superior performance in a range of basic shear flows,
including plane strain, rotating plane shear and axisymmetric ex-
pansion/contraction. On LES front, the standard Smagorinsky model
has been widely used in the literature for stirred tank simulations.
However, it is well known that the Smagorinsky model is essentially
an algebraic model in which subgrid-scale stresses are parameter-
ized using the resolved velocity scales. The underlying assumption
is the local equilibrium between the transferred energy through the
resolved scale and the dissipation of kinetic energy at unresolved
subgrid scales. In order to overcome this limitation, the dynamic ki-
netic energy subgrid-scale model has been used in this study. In this
model, transport equation for subgrid-scale turbulence kinetic en-
ergy is explicitly solved.
It is well known that the precessing vortex is an intermittent,
non-stationary, pseudo periodic and large scale structure. This en-
tire phenomena is named precessing instability. This is mainly due
to a precessional motion of a vortex around the shaft. These are
low-frequency in nature with high amplitude of oscillatory motions
in the low turbulence regions of a vessel which have the capabil-
ity of transporting substances over relatively long distances. These
low-frequency motions can severely affect the transport phenom-
ena (local mean flow pattern heat and mass transfer rates, gas/solid
hold-up, etc.). Apart from the transport phenomena, these frequen-
cies at higher amplitude can affect the solid components inside the
reactors causing permanent damage to baffles, internals, sensors, etc.
Therefore, it was thought desirable to quantify the energy content of
the precessing vortex with respect to all the impeller designs under
consideration. In addition, scanty information is available in the lit-
erature on the possible interaction between the mean circulation and
the precessing instability which subsequently leads to intermediate
instabilities (between circulation and precessional vortex). Hence, in
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5477
-0.2
-0.1
0
0.1
0.2
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
-0.4
-0.2
0
0.2
0.4
0.6
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0
-0.4
-0.2
0
0.2
0.4
-0.2
-0.1
0
0.1
0.2
0.1
-0.1
-0.05
0
0.05
NORMALISED RADIAL COORDINATE, r/R (-)
1
3
2
2
1
3
2
1
3
2
1
3
2
1
3
2
1
3
2
1
3
1
2
3
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

A
X
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
2
>
/
U
t
i
p
0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
Fig. 8. Comparison between the simulated and experimental profiles of the dimensionless mean axial velocity for PBTD45 at various axial levels. (A) H=0.01m; (B) H=0.044m;
(C) H =0.082m; (D) H =0.1m; (E) H =0.118m; (F) H =0.154m; (G) H =0.190m and (H) H =0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
this study an attempt has been made to identify the intermediate
instabilities with respect to various impeller designs.
In this work, all the simulations have been performed using un-
steady sliding mesh approach, second-order discretization schemes
and fine mesh throughout the computational domain. All the compu-
tations have been performed using FLUENT 6.2. (2005), whose suc-
cessful application has been documented in the literature (Bakker
et al., 2001; Rocchi and Zasso, 2002; Roussinova et al., 2003; Hu and
Kazimi, 2006; Wang et al., 2006; Murthy et al., 2007). The details of
the turbulence modeling are discussed in Section 4.
3. LDA measurements
LDA measurements were carried out in a 0.29m i.d. transparent
acrylic vessel equipped with various impellers located at a clearance
(C) equal to H/3 from the tank bottom. The vessel was fitted with
four baffles having width 1/10 of the vessel diameter (fully baffled
conditions). To minimize the effect of curvature on the intersect-
ing beams, the vessel was placed inside a square vessel. A standard
DT, PBTD60, PBTD45, PBTD30, and HF (D = 0.1m) were used as the
impellers. The impeller was centrally located and driven by a vari-
able speed DC motor. The flow measurements were performed in
the midbaffle plane for all the configurations. LDA measurements
were carried with a single point, two-component LDAsystem(Ar-ion,
5W laser system) following the work of Ranade and Joshi (1989) and
Ranade et al. (1992). A large number of samples (300,000500,000)
were recorded for each run in which the data rate varied from 300
to 1500Hz.
LDA data collected for all the five impellers contain periodic-
ity due to the rotary motion of impeller. The periodic component
contributes significantly to the flows, especially in the near im-
peller region. Hence direct processing of the data may lead to the
overestimates of the turbulent kinetic energy. Thus it is essential to
remove the signal corresponding to the impeller blade passage fre-
quency and its simple low-frequency harmonics. For the validation
of the turbulence models, which essentially model the turbulence,
data free of regular (i.e. periodic) component are required. In view
of this, the periodic component arising out of the impeller rotation
from the velocitytime series has been removed. The following are
the details.
The data obtained due to LDA are random in nature with re-
spect to time. For any kind of transformation, it is desirable that the
time series should be expressed in terms of the variable at equal
time intervals. This conversion of a random time series to a time
series in terms of discrete data at equal time intervals is termed
here as equispacing of the data. For this purpose, linear interpolation
5478 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
-0.1
-0.05
0
0.05
0.1
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0
-0.2
-0.1
0
0.1
0.2
-0.3
-0.2
-0.1
0
0.1
0.2
-0.2
-0.1
0
0.1
0.2
-0.08
-0.06
-0.04
-0.02
0
0.02
0.04
0.1
-0.06
-0.04
-0.02
0
0.02
NORMALISED RADIAL COORDINATE, r/R (-)
1
2
3
1
2
3
1
2
3
3
2
1
1
2
3
2
1
3
1
2
3
1
2
3
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

A
X
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
2
>
/
U
t
i
p
0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.3 0.5 0.7 0.9
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Fig. 9. Comparison between the simulated and experimental profiles of the dimensionless mean axial velocity for PBTD30 at various axial levels. (A) H=0.01m; (B) H=0.044m;
(C) H =0.082m; (D) H =0.1m; (E) H =0.118m; (F) H =0.154m; (G) H =0.190m and (H) H =0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
technique was used. Further, the signal obtained after equispacing
was again processed using fast Fourier transform (FFT). The power
spectrum was obtained and the frequencies corresponding to the
impeller blade passage frequency (6 4.5Hz) and its harmonics
were identified. The data corresponding to these frequencies were
removed and the time series was reconstructed using the inverse
Fourier transformation. The data free of periodicity was subsequently
subjected to the noise removal using the multiresolution analysis
(Kostelich and Yorker, 1988).
Fig. 2 shows eight axial locations at which LDA data were col-
lected. The data were processed for the estimation of all the three
mean components of the velocity and the turbulent kinetic energy.
It may be emphasized again that, in this work, only random part of
the turbulent kinetic energy has been presented which was made
free from periodic component.
4. Turbulence modeling and flow governing equations
4.1. Standard k c model
In a turbulent flow, if we assume that the turbulence responds
rather quickly to changes in the mean flow we would expect the
Reynolds stresses themselves to be related to the mean rate of strain
as given by
u

i
u

j
=v
t

ju
i

jx
j
+
ju
j

jx
i

2
3
ok (1)
It involves six unknown eddy viscosities. Unfortunately the viscosi-
ties are related to each other in a manner that would be difficult to
describe in general. Hence, the most commonly used assumption is
to treat the eddy viscosity as a scalar quantity.
The standard k c model (Jones and Launder, 1972; Sahu et al.,
1999) is essentially a high Reynolds number model and assumes
the existence of isotropic turbulence and the spectral equilibrium. It
proposes the relation for the eddy viscosity in terms of k and c with
the help of the turbulence parameter C
j
. In addition, the triple ve-
locity correlations in the transport equation for the energy dissipa-
tion rate are modeled with the help of two more constants (C
c1
and
C
c2
) whose values have been derived from the measurement of flow
characteristics of simple two-dimensional equilibrium flows. In the
k c model, the length and the time scales are built up from the tur-
bulent kinetic energy and the dissipation rate using the dimensional
arguments.
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5479
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
0
0
0.1
0.2
0.3
0.4
0.5
-0.2
-0.1
0
0.1
0.2
0.3
0.4
-0.15
-0.11
-0.07
-0.03
-0.2
-0.1
0
0.1
-0.15
-0.11
-0.07
-0.03
-0.1
-0.08
-0.06
-0.04
-0.02
0
0.1
-0.02
-0.01
0
0.01
0.02
0.03
2
1
3
3
1
2
1
2
3
1
2
3
1
2
3
1
2
3
3
1
2
3
2
1
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

R
A
D
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
1
>
/
U
t
i
p
0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.3 0.5 0.7 0.9
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
Fig. 10. Comparison between the simulated and experimental profiles of the dimensionless mean radial velocity for PBTD60 at various axial levels. (A) H = 0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
The k c turbulence models suffer from the necessity of model-
ing a number of quantities for which reliable experimental data are
desirable under a large number of flow conditions. While this ne-
cessity is a fundamental weakness of the k c approach, a further
uncertainty lies in the assumption that the turbulent kinetic energy
and its dissipation rate are necessary and sufficient turbulence vari-
able for the simulation of turbulent flows. Nonetheless, the model is
widely used and has been attributed to some significant simulation
successes.
4.2. Reynolds stress model
It is clear that the standard k c is not equipped to pre-
dict anisotropic turbulent flows (Reynolds, 1987; Launder, 1990;
Hanjalic, 1994). Further, the modeling of transport equations for k
and c clearly brings out the difficulties to account for streamline
curvature, rotational strains, and the other body-force effects. RSM,
in theory, circumvents all the above mentioned deficiencies and
also it has an ability to predict more accurately each individual
stress.
The transport equation for the Reynolds stresses is given as
jt
ij
jt
+u
k

jt
ij
jx
k
=

t
ik
ju
j

jx
k
+t
jk
ju
i

jx
k

j
jx
k
C
ijk
+H
ij
c
ij
+v
2
t
ij
. (2)
The terms C
ijk
, H
ij
and c
ij
in Eq. (2) are given as
H
ij
=

ju

j
jx
i
+
ju

i
jx
j

=Pressure strain correlation (3)


c
ij
=2v

ju

i
jx
k
ju

j
jx
k

=Dissipation rate correlation (4)


C
ijk
=u

i
u

j
u

k
+
1
j
(P

i
o
jk
+P

j
o
ik
)
= Third-order diffusion correlation (5)
The third-order diffusion correlation, C
ijk
, pressure strain, H
ij
and
dissipation rate tensor c
ij
need to be modeled to close the set of the
5480 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
0
0.2
0.4
0.6
0
0
0.2
0.4
-0.05
0
0.05
0.1
-0.15
-0.11
-0.07
-0.03
0.3
-0.2
-0.1
0
0.1
-0.2
-0.15
-0.1
-0.05
0
-0.12
-0.08
-0.04
0
-0.06
-0.04
-0.02
0
0.02
0.04
NORMALISED RADIAL COORDINATE, r/R (-)
2
2
1
3
1
3
2
1
3
2
2
1
3
2
1
3
1
3
1
2
3
1
3 2
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

R
A
D
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
1
>
/
U
t
i
p
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.5 0.7 0.9
Fig. 11. Comparison between the simulated and experimental profiles of the dimensionless mean radial velocity for PBTD45 at various axial levels. (A) H = 0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
governing equations. The turbulent diffusion is modeled as per Lien
and Leschziner (1994) in FLUENT 6.2: j/jx
k
((v
t
/o
k
) jt
ij
/jx
k
), where
o
k
= 0.82 and the turbulent viscosity, j
t
, estimated from the k and
c with C
j
= 0.09. It is obvious that the highly anisotropic flow due
to impeller rotation and subsequent interaction between the strong
impeller jet and the bulk flow suggests that the production term and
the pressure strain correlation play a significant role in the prediction
of turbulent stresses. In the present work, the quadratic pressure
strain model proposed by Speziale et al. (1991), which is known to
improve the accuracy of flow field with streamline curvature, has
been used to model the pressurestrain term of the RSM.
To obtain the boundary conditions for the Reynolds stresses at
the wall, the equation for the turbulent kinetic energy (k) was solved.
Further, the equation for the dissipation rate (c) of turbulent kinetic
energy was solved to obtain the dissipation rate (c
ij
) term in the
Reynolds stress transport equation.
4.3. Large eddy simulations
LES equations of turbulent flows are formally derived by ap-
plying a filtering operation to the NavierStokes equation and
assuming that filtering and differentiation operators' commute
(Leonard, 1974). The filtering process effectively filters out the ed-
dies whose scales are smaller than the filter width or grid spacing
used in the meshing. The resulting equations have the structure
as the original equation plus additional terms, called subgrid scale
stresses (SGS). The filtered equations are used to compute the dy-
namics of the large-scale structures, while the effect of the small
scale turbulence is modeled using a SGS model. Following are the
flow governing equations for LES:
ju
i
jx
i
=0 (6)
j
jt
(u
i
) +u
j
ju
i
jx
j
=
1
j
jP
jx
i
+
j
jx
j

[v +v
t
]
ju
i
jx
j

(7)
Where we have used the incompressibility constraint to simplify the
equation and the pressure is now modified to include the trace term
t
kk
o
ij
/3. The SGS resulting from the filtering operation are unknown
and require modeling. Most of the past work on LES of stirred ves-
sels have employed algebraic models in which subgrid-scale stresses
are parameterized using the resolved velocity scales. The underlying
assumption is the local equilibrium between the transferred energy
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5481
-0.1
0
0.1
0.2
0.3
0
0.05
0.1
0.15
0.2
0
-0.06
-0.03
0
0.03
0.06
0.09
0.12
0.15
-0.06
-0.04
-0.02
0
-0.2
-0.1
0
0.1
-0.04
-0.03
-0.02
-0.01
0
0.1
-0.05
-0.04
-0.03
-0.02
-0.01
0
0
0.01
0.02
0.03
0.1
NORMALISED RADIAL COORDINATE, r/R (-)
1
2
3
3
1
2
3
1
2
2
1
3
2 1
3
1
2
3
1
2
3
1
2
3
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

R
A
D
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
1
>
/
U
t
i
p
0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.3 0.5 0.7 0.9
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.6
Fig. 12. Comparison between the simulated and experimental profiles of the dimensionless mean radial velocity for PBTD30 at various axial levels. (A) H = 0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
through the grid-filter scale and the dissipation of kinetic energy
at small subgrid scales. The subgrid-scale turbulence can be better
modeled by accounting for the transport of the subgrid-scale tur-
bulence kinetic energy. The dynamic kinetic energy SGS model in
FLUENT replicates the model proposed by Kim and Menon (1997),
which has been used in the present work. The subgrid-scale kinetic
energy is defined as (Kim and Menon, 1997)
k
sgs
=
1
2
(u
2
k
(u
k
)
2
) (8)
The subgrid-scale eddy viscosity, v
t
, is computed using k
sgs
as
(Kim and Menon, 1997)
v
t
=C
k
k
1/2
sgs
A (9)
The unkown, k
sgs
is obtained by solving its transport equation
given by
jk
sgs
jt
+u
j
jk
sgs
jx
j
= t
ij
ju
j
jx
j
C
c
k
3/2
sgs
A
+
j
jx
j

v
t
o
k
jk
sgs
jx
j

(10)
In the above equations, the model constants, C
k
and C
c
, are deter-
mined dynamically.
5. Numerical details
5.1. Geometrical details and grid generation
The geometrical details of the problem investigated are shown
in Figs. 1 and 2. The cylindrical vessel of diameter T = 0.30m and
fitted with four baffles of width, T/10 was used as the stirred vessel
to study the flow characteristics of DT, PBTD and HF. In all the cases,
the diameter of the impeller was D =0.1m. The clearance between
the impeller and the vessel bottom was kept at T/3 and the impeller
was centrally located. The height of liquid level (H) was equal to the
tank diameter (T). The impeller rotational speed was 4.5rps and wa-
ter was used as the working fluid. Hexahedral elements were used
for meshing the geometry and a good quality of mesh was ensured
throughout the computational domain using GAMBIT mesh gener-
ation tool. In order to ensure better quality mesh, impeller blades
and baffles were considered with zero thickness. The computational
mesh consisted of 575,000 ( 90 60 106, z r 0) hexahedral
cells for the standard k c model and the RSM. In the case of LES,
5482 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
0
0.05
0.1
0.15
0.2
0.25
0
0
0.1
0.2
0.3
0.4
0.5
0
0.2
0.4
0.6
0.8
-0.2
-0.1
0
0.1
0.2
0.3
0
0.03
0.06
0.09
0.12
0
0.02
0.04
0.06
0.1
-0.06
-0.04
-0.02
0
0.02
0.04
0.06
0
0.01
0.02
0.03
1
3
2
3
1
2
3
1
2
3
1
2
1
2
3
1
2
3
3
1
2
NORMALISED RADIAL COORDINATE, r/R (-)
3 1
2
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

T
A
N
G
E
N
T
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
3
>
/
U
t
i
p
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.5 0.7 0.9
0.3 0.5 0.7 0.9 0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
Fig. 13. Comparison between the simulated and experimental profiles of the dimensionless mean tangential velocity for PBTD60 at various axial levels. (A) H =0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
1,275,567 ( 10069184, zr0) hexahedral cells were used and
with relatively fine mesh in the impeller region in order to better
resolve the strong velocity flow field. In the present LES, the length
scales were resolved between 200m and 4.2mm, which are of the
order of the Taylor microscale (z), i.e. 6.5, 6.7, 7.1, 7.3 and 7.5mm
for DT, PBTD60, PBTD45, PBTD30 and HF impeller, respectively.
Therefore, one can expect the realistic LES predictions from this grid
resolution. However, the present grid resolution is much higher as
compared with the Kolmogorov's length scale (p) of 46m
((v
3
/c)
1/4
), for the flow generated by DT, i.e. smallest scale in the
flow generated among the five impeller designs in the present
study. Further, the Kolmogorov length scale is 53m for PBTD60,
57m for PBTD54, 64m for PBTD45 and95m for HF impeller, re-
spectively. It may be noted that these estimations have been made
by considering c as the average energy dissipation rate.
5.2. Method of solution
In this work, all the computational work has been carried out us-
ing the commercially available software FLUENT 6.2. The discretized
form of the governing equations for each cell was obtained such
that the conservation principles are obeyed on each cell. The second-
order implicit scheme was used for time discretization in all the
turbulence models. Further, the second-order central bounded dif-
ference scheme was used for spatial discretization in case of LES and
the second-order upwind scheme for RANS based models. All the
discretized equations were solved in a segregated manner with the
PISO (pressure implicit with splitting of operators) algorithm. PISO
involves one predictor step and two corrector steps and may be seen
as an extension of SIMPLE (semi-implicit method for pressure-linked
equations), with a further corrector step to enhance it. PISO is a
pressurevelocity calculation procedure developed originally for the
non-iterative computation of unsteady compressible flows. There-
fore, PISO has better performance in unsteady simulation than SIM-
PLE series algorithm. To improve the efficiency of this calculation,
the PISO algorithm adopts two additional corrections: neighbor cor-
rection and skewness correction. As far as LES run is concerned,
to ensure smooth and better convergence initially k c simulations
have been performed until the complete steady state flow field is
obtained and then k c results have been used as the initial guess
values for the LES. In the present work, all the solutions were con-
sidered to be fully converged when repeated iterations do not de-
crease the sum of residuals below 1 10
5
. Here, the residual R
calculated as the imbalance in algebraic equation summed over all
the computational cells. FLUENT scales the residual using a scal-
ing factor representative of the flow rate of variable through out
the domain. Standard model constants have been used for all the
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5483
0
0.1
0.2
0.3
0
0.1
0.2
0.3
0.4
0.5
0
0.2
0.4
0.6
0
0.1
0.2
0
0.1
0.2
0
0.04
0.08
0.12
0
0.02
0.04
0.06
0.08
0
-0.06
-0.04
-0.02
0
0.02
0.04
NORMALISED RADIAL COORDINATE, r/R (-)
3
1
2
1
2
3
3
1
2
2 1
3
2
1
3
2
1
3
2
1
3
1
2 3
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

T
A
N
G
E
N
T
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
3
>
/
U
t
i
p
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
Fig. 14. Comparison between the simulated and experimental profiles of the dimensionless mean tangential velocity for PBTD45 at various axial levels. (A) H =0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
turbulence models. Grid independent study has been carried out only
for the case of DT using two grid resolutions for the standard k c,
Reynolds-stress transport models (RSTMs). The total number of cells
in the three directions for the two cases, 475,000 ( 95 50 100,
z r 0) and 575,000 ( 9060106, z r 0), respectively. It has
been found that both the grids gave very similar profiles of turbulent
kinetic energy and the values of power number were found to be
practically the same. However, all the RANS based simulations have
been performed for the grid resolution of 575,000 ( 90 60 106,
z r 0). In the case of LES, the presented turbulent kinetic energy
is a sum of resolved and unresolved (obtained fromsolving transport
equation for the sgs kinetic energy equation) turbulence kinetic en-
ergy. The periodic component of all the variables was removed in all
the simulations. In the present work, simulations have been initially
performed with the time step size of 0.0001s and the correspond-
ing CFL number is 0.7. As the solution progressed the time step size
has gradually been increased to 0.001s for LES, and 0.01s for RANS
in order to save computational time. For the same reason implicit
time stepping scheme has been employed in the present work. The
major advantage of this kind of methods is that the time step is not
limited by stability reasons, i.e. CFL condition. This means that, con-
trary to explicit schemes, stability is ensured for any value of the
time step. In this way a smaller number of iterations are required to
complete the simulation, leading to an important gain in CPU time.
However, the level of accuracy of the solution at high CFL numbers is
low. Therefore, when a high level of accuracy is needed, for instance
the resolution of unsteady and transient phenomena, the time step
must be kept small. In this situation, explicit codes will perform bet-
ter than implicit ones because of their smaller computational cost
per iteration. The simulation was performed for a time span of 90s
in the case of LES. In the case of RANS the simulations were per-
formed for the flow time of 9s, which corresponds to 44 impeller
revolutions. All the computations have parallely been performed on
an AMD64, 32 (16 nodes) processors cluster with a total 32GB RAM,
2.4GHz processor speed.
6. Results and discussions
CFD simulations have been performed with three turbulence
models, namely standard k c, RSM and LES for the five different
impellers. The simulation results of the dimensionless mean axial
velocity, mean radial velocity, mean tangential velocity and turbu-
lent kinetic energy have been plotted against the LDA experimental
data. The comparison of radial profiles of these parameters has
been made at eight different axial levels, z = 0.01m (A), 0.044m
(B), 0.082m (C), 0.1m (D), 0.118m (E), 0.154m (F), 0.190m (G) and
0.244m (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.
5484 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
0
0.05
0.1
0.15
0.2
0.25
0
0
0.1
0.2
0.3
0
0.1
0.2
0.3
0.4
0
0.05
0.1
0.15
0.3
0
0.05
0.1
0
0.01
0.02
0.03
0.04
0.1
0
0.02
0.04
-0.02
-0.01
0
0.01
0.02
NORMALISED RADIAL COORDINATE, r/R (-)
3
1
2
3
1
2
3
1
2
2
1
3
1
3
2
1
3
2
3
1
2
3
1
2
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

T
A
N
G
E
N
T
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
3
>
/
U
t
i
p
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0.5 0.7 0.9
0.3 0.5 0.7 0.9 0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9
Fig. 15. Comparison between the simulated and experimental profiles of the dimensionless mean tangential velocity for PBTD30 at various axial levels. (A) H =0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
The following section discusses the performance of various tur-
bulence models for all the five impellers. However, in order to make
the explanation clear and simple, the discussion on the predictive
capabilities of all the three models is divided into the DT, all the
axial flow impellers together, i.e. PBTD (60, 45 and 30) and HF. It
is well known that a DT produces strong radial discharge with in-
tense swirling. In the case of axial flow impellers, PBTD60 develops
a strong downward flow with strong tangential velocity component
in the impeller region. On the other hand, PBTD45, PBTD30 and HF
produce tangential component in the decreasing order of magni-
tude. Therefore, it would be easy to assess the inadequacy of the
model formulation and underlying assumptions involved in different
models.
6.1. Disc turbine
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 velocities exist. The
radial profiles of axial velocities at various axial locations are shown
in Fig. 3(AH). It can be seen that maximum axial velocities exist
near the wall and are of the order of 0.200.30 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 all the three turbulent models are in good agreement with the
experimental data.
With regard to the radial component, impeller rotation generates
radially outward flow through the vertical surface of swept volume.
This high speed radial jet entrain surrounding fluid and slows down
as they approach the tank wall. It can be noted from Fig. 4(AH)
that both RSM and LES predictions agree well with the experimental
mean radial velocity profiles at all the axial levels. Whereas, standard
k c model exhibits some disparity at all the levels particularly above
the impeller.
Radial profiles of the mean tangential velocity are depicted in
Fig. 5(AH). Again, LES and RSM simulations capture the experimen-
tal mean tangential velocity profiles quite well. In case of standard
k c model, predictions mainly deviate in the near impeller region.
Fig. 6(AH) illustrates the comparison for turbulent kinetic en-
ergy throughout the tank. In the impeller central plane (Fig. 6D) the
profile of turbulent kinetic energy is practically similar to those of
radial velocity and tangential velocity. Below the impeller (Fig. 6C),
turbulent kinetic energy also exhibits two maxima as in the case of
tangential velocity. The first maximum is at r =0.27 and the second
one occurs closer to the wall at r =0.92. The former is generated by
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5485
0
0.025
0.05
0.075
0.1
0
0
0.04
0.08
0.12
0.16
0.2
0
0.05
0.1
0.15
0.2
0.25
0
0.02
0.04
0.06
0
0.02
0.04
0.06
0.08
0.1
0
0.01
0.02
0.03
0.04
0
0.02
0.04
0.1
0
0.005
0.01
0.015
0.02
0.025
3
1
2
3
1
2
3
1
2 3
1
2
3
1
2
1
3
2
3
1
2
1
2
3
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

T
U
R
B
U
L
E
N
T

K
I
N
E
T
I
C

E
N
E
R
G
Y
,

k
/
U
t
i
p

(
-
)

2
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.3 0.5 0.7 0.9 0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9 0.1 0.3 0.5 0.7 0.9
0.3 0.5 0.7 0.9
Fig. 16. Comparison between the simulated and experimental profiles of the dimensionless turbulent kinetic energy for PBTD60 at various axial levels. (A) H =0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
the rotating impeller and latter is because of the shear flow. The ki-
netic energies below the impeller are much smaller than those in the
impeller center plane and above the impeller (Fig. 6E), it is mainly
because of the upward inclination of trailing vortex pair. It can be
observed that at many axial levels, significantly in the impeller re-
gion, the standard k c and RSM models consistently underpredict
the turbulent kinetic energy. The good predictive capabilities of LES
can be clearly seen for the turbulent kinetic energy predictions. Here
one can justify the utility of LES even at the expense of computational
resources. As the LES provides macro and the reliable predictions of
the turbulent flow, the predictions can further be useful for better
predictions of macro- and micromixing, heat and mass transfer.
6.2. Axial flow impellers
Pitched blade down flow turbine (PBTD) develops a downward
jet below the impeller. The jet entrains fluid and the velocity gra-
dients become less sharp as they approach the tank bottom. The
axial jet impinges on the base and moves radially along the base
and after approaching the wall, turns upward in a well defined wall
jet. The following section discusses the mean axial, radial and tan-
gential velocity profiles and the turbulent kinetic energy profiles for
all the three axial flow impellers. Figs. 79 show the comparison
for the radial profiles of the mean axial velocity, whereas the mean
radial velocity comparison is depicted in Figs. 1012. Figs. 1315
showthe mean tangential velocity and the turbulent kinetic PBTD30,
respectively.
Figs. 7C9C clearly show that the maximum axial velocity in-
creases with an increase in the blade angle upto 60

. The width of
high speed jet issuing from the impeller also increases with an in-
crease in the blade angle. The magnitude of the maximum velocity
decreases with a decrease in the blade angle (0.25U
tip
for 30

to
0.55U
tip
for 60

). In the bulk (Figs. 79A, B, FH), the mean axial ve-


locity is relatively small near the axis at all the axial levels. Further,
the mean axial velocity increases steadily till a maximum is reached.
The maximum shifts towards the wall as the flow approaches the
base and the magnitude of the maximum decreases as the flow be-
comes predominantly radial. After the peak value, axial velocity de-
creases steadily and the flow turns upwards near the wall. Regarding
CFD modeling, in all the three cases, axial velocity predictions by all
the three turbulence models are in good agreement with the exper-
imentally measured values over the entire computational domain.
The comparison of the radial profiles of the mean radial velocity
is shown in Figs. 1012. In the impeller center plane (Figs. 10D12D)
radial flow is towards the center at all the radial locations and it
increases gradually from the wall and a maximum occurs near the
impeller blade tip. Just below the impeller (Figs. 10C12C), the mean
radial flow near the axis is feeble where the axial flow is dominant.
It increases gradually and a maximum occurs at r/R = 0.3 and just
5486 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
0
0.025
0.05
0.075
0
0.02
0.04
0.06
0.08
0.1
0
0.04
0.08
0.12
0.16
0
0.01
0.02
0.3
0
0.02
0.04
0
0.01
0.02
0.03
0
0.02
0.04
0
0
0.005
0.01
0.015
3
1
2
1
2
3
2
3
1
1
2
3
1
2
3
1
2
3
3
2
1
1
2
3
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

T
U
R
B
U
L
E
N
T

K
I
N
E
T
I
C

E
N
E
R
G
Y
,

k
/
U
t
i
p

(
-
)

2
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0.5 0.7 0.9
Fig. 17. Comparison between the simulated and experimental profiles of the dimensionless turbulent kinetic energy for PBTD45 at various axial levels. (A) H =0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
beyond the tip of the impeller (at r/R=0.4) radial velocity is very small
like the velocities near the axis. Radial profiles of the mean radial
velocity in the bulk region below the impeller (Figs. 1012A and B)
show that, the radial flow is towards the wall and velocity increases
right from the axis and attains a maximum. The radial location of
this maximum shifts towards the wall as one moves towards the
vessel bottom. The magnitude of the maximum increases near the
base. After attaining the maximum value, radial velocity decreases
and becomes small near the wall where the axial flow is dominant.
Above the impeller in the bulk (Figs. 1012FH), the flow is towards
the center at all the radial locations. In all the cases, all the turbulent
models are able to capture the mean radial velocity profiles at all the
axial levels. However, in the impeller region, standard k c model
underpredict the mean radial velocity (though it is weak) generated
by the PBTD60.
Figs. 1315 show the mean tangential velocity comparison. It
can be noticed that the tangential flow is in the direction of the
impeller rotation at all the axial levels. Just below the impeller
(Figs. 13C15C), it increases and decreases steeply between r =0.15
and 0.3. Beyond this point, it is almost constant from r = 0.4 on-
wards. The maximum velocity is comparable with the maximum ax-
ial velocity at this level. Impeller with blade angle of 60

generates
significant tangential velocity components along the vertical surface
of the impeller region which do not exist for the impellers of blade
angles 45

and 30

. These tangential velocities result into a local


maximum in the resultant velocity along the vertical periphery. The
position of this maximum appears slightly below the impeller cen-
ter plane. This shift may be because of the strong axial downward
flow. The entering and leaving angles of the flow also decrease with
a decrease in the blade angles. It can be seen that the LES and RSM
predictions are in good agreement with the strong tangential flow
generated by the PBTD60. However, due to strong swirling motion,
standard k c model predictions show disparity with the experimen-
tal data mainly in the impeller region. In case of PBTD45, standard
k c model is unable to capture the tangential flow field. However, in
the case of PBTD30, standard k c model is quite successful in simu-
lating the relatively weak tangential velocities. Therefore, this study
clearly brings out quantitatively the limitations of the standard k c
model.
The profiles of turbulent kinetic energy are shown in Figs. 1618.
In the impeller center plane (Figs. 16D18D) turbulent kinetic energy
is more or less constant. Just below the impeller (Figs. 16C18C), the
value of k increases sharply from the axis upto r =0.3 and decreases
drastically beyond that point. From r =0.4 onwards turbulent kinetic
energy is very small and constant. It can be seen that, the values of
k behave similar to the mean velocity with respect to blade angle.
Impeller with blade angle of 60

generates intense turbulent flow.


The maximum value of resultant intensity is 0.45U
tip
, whereas it is
0.15U
tip
for the impeller with blade angle of 30

. There is a good
agreement between the experimentally measured turbulent kinetic
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5487
0
0.01
0.02
0.03
0
0
0.02
0.04
0.06
0
0.04
0.08
0
0.01
0.02
0.03
0.04
0.05
0
0.02
0.04
0.06
0.08
0.1
0
0.005
0.01
0.015
0.02
0
0.001
0.002
0.003
0.004
0.005
0
0.001
0.002
0.003
0.004
1
3
2
3
1
2
3
2
1
3
1
2
1
3
2
1 2 3
1
2
3
3
1 2
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

T
U
R
B
U
L
E
N
T

K
I
N
E
T
I
C

E
N
E
R
G
Y
,

k
/
U
t
i
p

(
-
)

2
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
0.3 0.5 0.7 0.9 0.1 0.3 0.5 0.7 0.9
0.1 0.3 0.5 0.7 0.9 0.1 0.3 0.5 0.7 0.9
Fig. 18. Comparison between the simulated and experimental profiles of the dimensionless turbulent kinetic energy for PBD30 at different axial levels. (A) H =0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
energy and LES predictions for all the axial flow impellers. However,
there are discrepancies as shown in Fig. 16F, and it may be attributed
for not long enough statistics. On the other hand, both the RANS
based models fail to simulate the turbulent kinetic energy associated
with the unsteady large scale motion generated by PBTD60 and 45.
However, the RSM, standard k c models can be seen to be successful
along with LES in predicting the relatively weak turbulent kinetic
energy generated by the PBTD30. It can further be noticed that, the
turbulent kinetic energy has been well captured in the bulk region
of the tank (both below and above impeller) by all the three models.
Since, the validation of blade angle from 60

to 30

shows a
trend in the variation of all the flow parameters, it was thought
desirable to investigate the flow generated by a HF impeller. It may
be pointed out that the power number of HF, PBTD30, PBTD45 and
PBTD60 are 0.27, 1.3, 2.24 and 3.1, respectively. Further, the rate of
mean to turbulent kinetic energy (at all the locations) is the highest
for a HF and decreases with an increase in the blade angle. Further,
though all these axial flowimpellers generate a mixed flow, the radial
component generated by HF is the weakest and increases with an
increase in the blade angle. In view of the strong convective motion
generated by an HF, these impellers are widely used for the flow
controlled operation.
Fig. 19(AH) depict the comparison of experimental axial velocity
with the predictions of the three turbulence models. It can be seen
that, at all the axial locations, all the three models give excellent pre-
dictions. A comparison of radial profiles of the mean radial velocity
are shown in Fig. 20(AH). In the impeller center plane (Fig. 20D)
radial flow is towards the impeller center at all the radial locations
and it increases gradually from the wall and a maximum occurs near
the impeller blade tip. Just below the impeller (Fig. 20), the mean
radial flow near the axis is feeble where the axial flow is dominant.
It increases gradually and a maximum occurs at r/R = 0.2 and just
beyond the tip of the impeller (r/R=0.4), the radial velocity remains
more or less constant. Radial profiles of the mean radial velocity in
the bulk region below the impeller (Fig. 20A and B) show that the
radial flow is towards the wall and the velocity increases right from
the axis and attains a maximum. The radial location of this maxi-
mum shifts towards the wall (at z =0.046m it was r/R =0.3 and at
z =0.01m it was r/R=0.6) as one moves towards the vessel bottom.
The magnitude of the maximum increases near the base. After at-
taining the maximum value, radial velocity decreases and becomes
small near the wall where the axial flow is dominant. Above the im-
peller in the bulk (Fig. 20FH), the flow is towards the center at all
the radial locations. In all the cases, all the three turbulence models
are able to capture the mean radial velocity profiles.
Fig. 21(AH) showthe mean tangential velocity comparison. It can
be noticed that the tangential flow is in the direction of the impeller
rotation at all the axial levels. Just below the impeller (Fig. 20C), it
increases and decreases steeply between r = 0.15 and 0.3. Beyond
this point, it remains practically constant fromr=0.5 onwards. These
tangential velocities result into a local maximum in the resultant
velocity along the vertical periphery. However, the tangential flow
5488 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
-0.25
-0.15
-0.05
0.05
0.15
0.25
-0.2
-0.1
0
0.1
0.2
-0.04
-0.02
0
0.02
-0.1
-0.05
0
0.05
0.1
0
-0.3
-0.2
-0.1
0
0.1
0.2
0
-0.2
-0.1
0
0.1
-0.08
-0.06
-0.04
-0.02
0
0.02
0.04
0.06
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

A
X
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
2
>
/
U
t
i
p

2
1 3
2
1
3
2
1
3
-0.15
-0.1
-0.05
0
0.05
0.1
3
2
1
2
1
3
2
1
3
2
1
3
2
1
3
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
0.5 1
Fig. 19. Comparison between the simulated and experimental profiles of the dimensionless mean axial velocity for HF at various axial levels. (A) H=0.01m; (B) H=0.044m;
(C) H =0.082m; (D) H =0.1m; (E) H =0.118m; (F) H =0.154m; (G) H =0.190m and (H) H =0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
field generated by HF impeller is relatively weaker than the PBTD30
impeller. Further, the measured mean tangential velocities at all the
axial locations except at a few locations (Fig. 21D), have been found
to be in good agreement with the experimental data.
The profiles of turbulent kinetic energy are shown in Fig. 22. In
the impeller center plane (Fig. 22D) turbulent kinetic energy is more
or less constant. Just below the impeller (Fig 22C), the value of k
increases sharply from the axis up to r=0.2 and decreases drastically
beyond that point. From r =0.4 onwards turbulent kinetic energy is
very small and constant. It could be noticed that the turbulence levels
generated by HF impeller are relatively low than that generated by
the PBTD30. And the turbulent kinetic energy has been well captured
in the bulk region of the tank (both below and above impeller) by
all the three models.
6.3. Discussion
In case of DT, the flow pattern generated by the impeller is a
mixed radialtangential flow. The values of radial and tangential
velocities are maximum at the impeller tip. In the impeller center
plane, with increasing radial distance, the tangential velocity
decreases more rapidly than the radial component. As far as the
mean radial and tangential velocity predictions are concerned the
RSM outperforms the standard k c model in the impeller region
(Figs. 4 and 5). It can be attributed to the estimation of all the
components Reynolds stress which takes care of anisotropy, an im-
provement captured by RSM models and not captured by the scalar
eddy viscosity approach employing a Boussinesq constitutive rela-
tion. In addition, RSM accounts for the effects of streamline curva-
ture, and the rapid changes in the strain rate more rigorously than
the standard k c model. In the bulk region where isotropy exists,
both the velocity components get satisfactorily predicted well by
both the standard k c and the RSM model. In case of LES, since it
directly solves for the instantaneous scales, the entire mean flow
has been captured well.
PBT impellers generate a mixed axialradial flow in the form of
a jet which spreads radially as it progresses towards the base of the
vessel entraining fluid adjacent to the impeller. After hitting the base,
part of the axial momentum gets converted to radial component
along the base towards the wall of the vessel. As it can be seen
from the results, the axial mean velocity flow field of all the three
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5489
0
0.02
0.04
0.06
0.08
-0.03
-0.02
-0.01
0
0.01
0.02
0.03
-0.08
-0.06
-0.04
-0.02
0
-0.06
-0.04
-0.02
0
0.02
-0.01
-0.005
0
0.005
-0.05
0
0.05
0.1
0.15
0.2
0.25
0
-0.06
-0.04
-0.02
0
0.02
-0.04
-0.02
0
0.02
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

R
A
D
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
1
>
/
U
t
i
p
2
1
3
1
2
3
1
2
3
1
2
3
1
2
3
1
2
3
2
3
1
1
2
3
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
Fig. 20. Comparison between the simulated and experimental profiles of the dimensionless mean radial velocity for HF at various axial levels. (A) H=0.01m; (B) H=0.044m;
(C) H =0.082m; (D) H =0.1m; (E) H =0.118m; (F) H =0.154m; (G) H =0.190m and (H) H =0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
impellers has been predicted well by all the three turbulence models.
However, the mean radial and tangential velocities in the impeller
region have not been well predicted by the standard k c due to the
strong anisotropic nature of the flow as discussed above. Both the
RSM and LES faithfully captured the flow associated with the swirl
component like that generated by a DT.
The total kinetic energy is high in the jet, in the vicinity of the
impeller. Indeed, in the region near the impeller blade, the tangen-
tial velocities of the trailing vortices are high and the associated pe-
riodic energy is significant. The trailing vortices vanish far from the
impeller region. Just away from the impeller blade, kinetic energy
associated with the periodic motion decreases, however, the total
turbulent kinetic energy remains practically constant. Which indi-
cates a continuous conversion of kinetic energy (periodic compo-
nent) into turbulent kinetic energy. For this situation of flow, both
the standard k c and RSM fail to capture such a transfer process
due to unsteady and complex nature of flow structures in the im-
peller region. Therefore, in the near impeller regions, these models
consistently underpredicted the turbulent kinetic energy.
In the bulk region, all the three models were found to predict
close to the experimental data. These deviations can be attributed
to the energy transfer mechanism in each of the model and its abil-
ity to handle intermittency in the flow. The best agreement with
the data comes from the LES predictions which are able to cap-
ture the dynamic behavior of the coherent structures. Therefore, LES
predictions for all the five impellers, have been in good agreement
with the experimentally measured turbulent kinetic energy. How-
ever, the results reaffirm that the RSM outperforms the k c model
when swirling flows and recirculations are present, i.e. flow in the
vicinity of the DT and PBTD60 impellers. In the cases of PBTD30 and
HF, it has been shown that, due to the relatively much lower inten-
sity of swirl motion, the results show that the standard k c model
can even predict the flow fields equally well along with the RSM
and LES.
6.4. Instantaneous snap shots of flow field
The instantaneous vector field obtained from LES in a vertical
midbaffle plane of the stirred vessel is depicted in Fig. 23AD for
DT, PBTD60, PBTD45 and PBTD30 impellers, respectively. These re-
ports instantaneous velocity vector plots at the time instant of t =5s
(corresponding to 23.5 impeller revolutions). As can be seen from
Fig. 23 that the instantaneous flow field is highly complex and ran-
dom in nature. First of all the flow is never symmetry. Further,
it should be noted that the flow field is unsteady, and that these
5490 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
0
0.05
0.1
0.15
0.2
0
0.02
0.04
0.06
0.08
0.1
0.12
0
0.02
0.04
0.06
-0.01
-0.005
0
0.005
0.01
0.015
0
0
0.01
0.02
0.03
0.04
0
0
0.05
0.1
0.15
0.2
0
0.02
0.04
0.06
0.08
0
0.01
0.02
0.03
0.04
NORMALISED RADIAL COORDINATE, r/R (-)
D
I
M
E
N
S
I
O
N
L
E
S
S

M
E
A
N

T
A
N
G
E
N
T
I
A
L

V
E
L
O
C
I
T
Y
,

<
u
3
>
/
U
t
i
p
3
1
2
3
1
2
3
1
2 3
1
2
3
1
2
3
1
2
1
2 3
3
1
2
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.5 1
Fig. 21. Comparison between the simulated and experimental profiles of the dimensionless mean tangential velocity for HF at various axial levels. (A) H = 0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
images show snapshots of the flow field only. In the case of DT, a
strong radial discharge flow can be observed, which is most intense
in the plane containing the impeller. Large and irregular secondary
recirculation structures are present throughout the vessel. For the
axial flow impellers, PBTD60 tend to generate much complex and
irregular instantaneous flow than PBTD45 and PBTD30. It is mainly
due to strong interaction between the axial and tangential flow com-
ponents. With a decrease in the blade angle, the flow can be seen to
be less and less complex. This particular feature provides a justifica-
tion for the good predictive capabilities of the standard k c model
for the PBTD30 and HF impellers.
The mean flow and turbulence fields in a fully baffled vessel
stirred by various impeller designs (DT, PBTD60, 45 and 30, and HF) at
one clearance have been investigated with laser-Doppler anemom-
etry (LDA) and LES to characterize the instabilities present in such
flows. Time-resolved velocity measurements were made and the fre-
quency content of the velocity recordings was analyzed with FFT
techniques. The study aims to identify the flowinstabilities and asso-
ciated energy with them as well as frequency with respect to the im-
peller design. The frequency of the precessing vortex instability was
found to be linearly related to the rotational speed ( 0.015 0.02N,
Hz) of the impeller and to be essentially independent of the im-
peller design. The LDA data and LES predictions indicated clearly that
the precessing vortex instability stems from a precessional motion
about the vessel axis, similar to the precession encountered in most
swirling flows.
In the present study, the Reynolds number of the flowwas 45,000
for all the impeller designs under consideration. To estimate the
energy, time-resolved velocity measurements were made and the
data obtained due to LDA are random in nature with respect to time.
For any kind of transformation, it is desirable that the time series
should be expressed in terms of the variable at equal time intervals.
This conversion of a random time series to a time series in terms of
discrete data at equal time intervals is termed here as equispacing of
the data. For this purpose, linear interpolation technique was used.
Further, the signal obtained after equispacing was again processed
using FFT. The power spectrum was obtained and the frequencies of
instabilities were identified and corresponding energy on y-axis is
energy associated with that instability.
The energy of precessing vortex instability for both DT and
PBTD60 impellers near the vessel surface was found to be highest
and about 9m
2
s
2
.
And for PBTD45 and PBTD30 impellers the en-
ergy was found to be 1.8 and 1.35m
2
s
2
which is one-fifth of DT and
PBTD60. Whereas for HF impeller the energy content is 0.8m
2
s
2
,
which is relatively less among the impeller designs considered in
this study. It can be mainly attributed to the intensity of swirl
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5491
0
0.002
0.004
0.006
0.008
0
0.005
0.01
0.015
0.02
0.025
0
0.002
0.004
0.006
0
0.001
0.002
0
0.002
0.004
0.006
0.008
0
0
0.002
0.004
0.006
0.008
0.01
0
0.002
0.004
0.006
0.008
0
0.001
0.002
0.003
0.004
D
I
M
E
N
S
I
O
N
L
E
S
S

T
U
R
B
U
L
E
N
T

K
I
N
E
T
I
C

E
N
E
R
G
Y
,

k
/
U
t
i
p

(
-
)

NORMALISED RADIAL COORDINATE, r/R (-)
3
1
2
3
1
2
3
1
2
3
1
2
3
1
2
3
1
2
3
1
2
3 1
2
2
0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0.2 0.4 0.6 0.8 1
Fig. 22. Comparison between the simulated and experimental profiles of the dimensionless mean turbulent kinetic energy for HF at various axial levels. (A) H =0.01m; (B)
H=0.044m; (C) H=0.082m; (D) H=0.1m; (E) H=0.118m; (F) H=0.154m; (G) H=0.190m and (H) H=0.244m: Experimental; (1) LES model; (2) RSM and (3) k c model.
generated by DT and PBTD60 than the other impeller designs. It can
be realized that the flow generated by HF is mainly a convective
flow.
Also, the macroinstability is triggered by the complex interac-
tion of the impinging jet from the impeller discharge stream with
either the tank wall or bottom. The resulting oscillation in the cir-
culation pattern is called as jet instability. The frequency of the jet
instability was found to be linearly related to the rotational speed
( 0.13 0.2N, Hz) of the impeller and to be essentially independent
of the impeller design for a given impeller clearance. The energies
associated with the jet instability for all the five impellers, i.e. DT,
PBTD60, PBTD45, PBTD30 and HF, have been found to be 42, 38, 0.64,
8 and 12m
2
s
2
, respectively.
Further, there has been an apparent discrepancy in the literature
on the possible interaction between the jet instability and the pre-
cessing instability which subsequently leads to intermediate insta-
bilities (between jet/circulation and precessional vortex). Therefore,
in this study, an attempt has been made to identify the intermediate
instabilities with respect to various impeller designs. It can be seen
in Fig. 24(AE) that the occurrence of some intermediate instabili-
ties in the frequency range of 0.20.3 (i.e. f/N=0.04 0.07) for all the
impeller designs. Energy content of these intermediate instability
have been found to be 8m
2
s
2
(DT), 8m
2
s
2
(PBTD60), 3m
2
s
2
(PBTD45), 2m
2
s
2
(PBTD30) and 1m
2
s
2
(HF). These kind of
instabilities might be part of the cascading of the large scale insta-
bilities. Further a separate detailed study of identification, quantifi-
cation and relating these instabilities to design objectives is under
progress.
6.5. Energy balance
It was thought desirable to assess the performance of the differ-
ent turbulence models by establishing the energy balance. The total
power dissipation rate was calculated by volume integration of the
predicted turbulent kinetic energy dissipation rate (c) as
P =

2
0

H
0

R
0
cr dr dz d0

2
0

H
0

R
0
r dr dz d0
(11)
In LES model, c has been estimated based on the SGS viscosity ob-
tained as
c =t
ij
S
ij
=2v
SGS
S
ij
S
ij
=(C
S
A)
2
|S
ij
|
3
(12)
By assuming local equilibrium between production and dissipation
at and below subgrid-scale level, the energy dissipation rate in the
LES can be coupled to the deformation rate.
From the predicted energy dissipation rate (Eq. (13)), the values
of power number (N
P
) for all the five impellers were obtained using
5492 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
Fig. 23. The instantaneous vector field in a vertical midbaffle plane: (A) DT; (B) PBTD60; (C) PBTD45 and (D) PBTD30.
the following equation:
N
P
=
P
jN
3
D
5
(13)
Table 2 shows the comparison of power numbers predicted by the
standard k c, RSM and LES models for all the five impellers with the
experimentally measured values. The power consumption was mea-
sured by measuring torque on table, below which the tank remains
stationary. For this purpose, the torque table was restrained from ro-
tating by a string and the force on the string was then measured by
connecting it to a cantilever type load cell. The pre-calibrated load
indicator displays the load on the torque table. Ten readings were
taken for each set and average was used for power measurement.
The predicted N
P
by LES can be seen to be in close agreement with
the experimental values which implies good overall energy balance.
Whereas, both the standard k c and RSM models consistently un-
derpredicted the power numbers for all the impellers.
The energy balance for all the impellers was also established using
an alternative procedure. The power consumption P is calculated as
the product of torque on the impeller blades and the angular velocity.
This is then used for the estimation of power number and it can be
expressed as follows:
N
P
=
2NM
jN
3
D
5
(14)
where torque (M) exerted in all blades was computed by summing
the cross product of the pressure and viscous forces vectors with
corresponding distance vector for each computational cell on the
impeller surface. The predictions of the power numbers for all the
five impellers are shown in Table 3. The power number (N
P
) has
been well predicted by all the three turbulent models. It can be
noted that, the power number calculated fromthe integration of local
epsilon value obtained from RANS based models tends to be under-
predicted in the range of 2025% for DT and PBTD60 impellers, a sim-
ilar behaviour has already been reported in the literature (Brucato
et al., 1998; Patwardhan, 2001). For the other two pitched blade im-
pellers (PBTD45 and PBTD30) and HF impeller a close agreement
between predictions and the experimental values can be seen. On
the other hand, LES found to closely predict the power number
with the maximum of 7% deviation in the case of DT. However,
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5493
Fig. 24. Power spectra for the different impeller designs (Z/T =0.27, r/R =0.6): (A) DT; (B) PBTD60; (C) PBTD45; (D) PBTD30 and (E) HF.
for PBTD60, PBTD45, PBTD30 and HF impellers close agreement has
been obtained between the experimentally measured values and LES
predictions.
7. Conclusions
Three-dimensional mean flow field and the turbulent kinetic en-
ergy of the baffled stirred vessel agitated by five impellers (DT, PBT
(60

, 45

and 30

) and an HF) have been measured using LDA. For


all the above cases three-dimensional CFD simulations have been
performed by the standard k c, the Reynolds stress transport and
large eddy simulation turbulence models. The pressure strain terms
in RSM was modeled based on the proposal in the literature that is
suitable for the present flow configuration. In case of LES, for the
first time, one equation dynamic subgrid scale model has been suc-
cessfully employed for the stirred vessel geometry. The predicted
5494 B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495
Table 2
Comparison of the power number predicted (integral c based approach) by standard k c, RSM and LES turbulence models
Turbulence model Disc turbine (Exp NP =5.1) PBTD60 (Exp NP =3.1) PBTD45 (Exp NP =2.24) PBTD30 (Exp NP =1.3) HF (Exp NP =0.27)
Standard k c 3.9 2.6 1.9 1.1 0.25
RSM 4.1 2.75 1.9 1.1 0.25
LES 4.7 2.85 2.1 1.2 0.26
Table 3
Comparison of the power number predicted (torque based approach) by standard k c, RSM and LES turbulence models
Turbulence model Disc turbine (Exp NP =5.1) PBTD60 (Exp NP =3.1) PBTD45 (Exp NP =2.24) PBTD30 (Exp NP =1.3) HF (Exp NP =0.27)
Standard k c 4.9 2.95 2.3 1.35 0.28
RSM 5.0 3.2 2.2 1.25 0.28
LES 5.2 3.0 2.1 1.38 0.28
flow fields by all the three turbulence models were comprehensively
compared with the experimental data. The following conclusions can
be drawn from the present work.
1. As for as mean flow predictions are concerned, the RSM per-
formed better than the standard k c model when comparisons
were made with DT. The results reaffirm that the RSM can out-
perform the k c model when recirculations are present. This is
due to the overestimation of the eddy viscosity which is a general
characteristic of the k c model. In general, it can be concluded
that the standard k c model performs well when the flow is uni-
directional that is with less swirl and weak recirculation.
2. The standard k c model tends to satisfactorily represent all the
flow parameters generated by a hydrofoil (HF) impeller. Since the
HF impeller produces such a weak swirl flow, these results show
that all the three turbulence models can predict the flow field
equally well.
3. Both the standard k c model and anisotropy RSM fail to predict
the turbulent kinetic energy profiles in the impeller region when
the flow is dominated by the unsteady coherent flow structures.
4. Using both the LDA and LES tools, the strength of the precess-
ing vortex instability has been quantified for all the five impeller
designs. It has been observed that the impeller which generates
strong swirl flow generates equally strong was the main reason
for such instabilities. Therefore, among the impeller designs con-
sidered in the present study, the DT produces strongest instabil-
ities and the HF generates the weakest instabilities.
5. The frequency of the jet instability was found to be linearly related
to the rotational speed ( 0.13 0.2N, Hz) of all the impeller de-
signs under consideration. The energies associated with the jet in-
stability for all the four impellers, i.e. DT, PBTD60, PBTD45, PBTD30
and HF, have been found to be 42, 38, 0.64, 8 and12m
2
s
2
, re-
spectively.
6. Further, consistent occurrence of intermediate instabilities have
been observed having frequency of 0.040.07N which lies inbe-
tween the precessional and the jet instability. The origin of these
kind of instabilities could be due to the interaction of precess-
ing vortex instability with either the mean flow or jet/circulation
instabilities.
7. Having demonstrated that the LES model provides predictions
that agree well with the measurements, and these simulations
capture relatively many flow features, these simulations can be
further utilized to explore a variety of issues with a reasonable
confidence.
8. Energy balance has been established using both integral c and
torque based approaches. It can be concluded that the torque
based approach seems to be more promising for the estimation
of power number for a new impeller design and using by compu-
tationally economical RANS based simulations.
Notation
C impeller clearance from the tank bottom, m
C
j
, C
c1
, C
c2
turbulence model parameters in the k c model
C
k
, C
c
SGS model constants
d
w
distance of a grid point closest wall, m
D impeller diameter, m
DT disc turbine
D
ij
diffusion of Reynolds stresses, m
2
s
3
D
c
turbulent diffusion of dissipation, m
2
s
4
f frequency, s
1
H liquid height, m
HF hydrofoil impeller
k turbulent kinetic energy, m
2
s
2
l
o
characteristic length scale, m
M torque, Nm
N impeller rotation speed, s
1
N
P
power number of the impeller
N
QP
primary flow number of the impeller
P power dissipation, m
2
s
3
PBTD30 pitched blade downflow turbine with a blade angle
of 30

PBTD45 pitched blade downflow turbine with a blade angle


of 45

PBTD60 pitched blade downflow turbine with a blade angle


of 60

P filtered pressure term for LES model, Nm


2
P
c
production of dissipation, m
2
s
3
P
ij
production in Reynolds stress transportequation,
m
2
s
3
r radial coordinate, m
R radius of the vessel, m
R
ij
Reynolds stress tensor, Nm
2
S
ij
strain rate of the resolved scales, s
1
t time, s
T tank diameter, m
u
i
time average of velocity, ms
1
u
1
time averaged radial velocity, ms
1
u
2
time averaged axial velocity, ms
1
u
3
time averaged tangential velocity, ms
1
u
i
filtered velocity, ms
1
U
tip
impeller tip velocity, ms
1
V volume of the computational cell, m
3
z axial coordinate, m
Greek letters
o Kronecker delta
A filter width, m
B.N. Murthy, J.B. Joshi / Chemical Engineering Science 63 (2008) 5468-- 5495 5495
c turbulent energy dissipation rate, m
2
s
3
p Kolmogorov length scale, m
0 tangential coordinate,rad
z Taylor microscale, m
v kinematics viscosity, m
2
s
1
H
ij
pressure strain term in RSM equations, m
2
s
3
j density of the fluid, kgm
3
o
k
turbulent Prandtl number for the turbulent kinetic
energy
o
c
turbulent Prandtl number for the dissipation rate
t
ij
shear stress in i-direction, Nm
2
t
o
characteristic time scale, s
t
sgs
subgrid-scale stress, Nm
2
t
sgs
subgrid scale eddy viscosity, m
2
s
1
[
c
turbulent destruction of dissipation, m
2
s
4
Subscripts and Superscript
i, j, k axis indexes of space coordinates
l molecular
sgs subgrid scale
t turbulence

fluctuating quantity
Acknowledgment
One of the authors, Mr B.N. Murthy gratefully acknowledge the
financial support during this work by the Department of Atomic
Energy (DAE), Government of India.
References
Alcamo, R., Micale, G., Grisafi, F., Brucato, A., Ciofalo, M., 2005. Large-eddy simulation
of turbulent flow in an unbaffled stirred tank driven by a Rushton turbine.
Chemical Engineering Science 60, 23032316.
Aubin, J., Flethcer, D.F., Xuereb, C., 2004. Modeling turbulent flow in stirred tanks with
CFD: the influence of the modeling approach, turbulence model and numerical
scheme. Experimental Thermal and Fluid Science 28, 431445.
Bakker, A., Van den Akker, H.E.A., 1998. Single phase flow in stirred reactors. Chemical
Engineering Research & Design 72, 583593.
Bakker, A., Oshinowo, L.M., Marshall, E.M., 2000. The use of large eddy simulation to
study stirred vessel hydrodynamics. In: Proceedings of 10th European Conference
on Mixing, Delft, Netherlands, pp. 247253.
Bakker, A., Oshinowo, L.M., Marshall, E.M., 2001. Realize great benefits from CFD.
Chemical Engineering Progress, 4553.
Brucato, A., Ciofalo, M., Grisafi, F., Michale, G., 1998. Numerical prediction of flow
fields in baffled stirred vessels: a comparison of alternative modeling approaches.
Chemical Engineering Science 55, 291302.
Derksen, J., 2001. Assessment of large eddy simulations for agitated flows. Chemical
Engineering Research & Design 79, 824830.
Derksen, J.J., Van den Akker, H.E.A., 1999. Large-eddy simulations on the flow driven
by a Rushton turbine. A.I.Ch.E. Journal 45, 209219.
Derksen, J.J., Kontomaris, K., McLaughlin, J.B., Van den Akker, H.E.A., 2007. Large-
eddy simulation of single-phase flow dynamics and mixing in an industrial
crystallizer. Chemical Engineering Research & Design 85, 169179.
Eggels, J.G.M., 1996. Direct and large-eddy simulation of turbulent fluid flow using
the lattice-Boltzmann scheme. International Journal of Heat and Fluid Flow 17,
307323.
FLUENT 6.2, 2005. User's Manual to FLUENT 6.2. Fluent Inc. Central Resource Park,
10 Cavendish Court, Lebanon, USA.
Hanjalic, K., 1994. Advanced turbulence closure models: a view of current status
and future prospects. Journal of Heat and Fluid Flow 15, 178203.
Hartmann, H., Derksen, J.J., Van den Akkar, H.E.A., 2004a. Macro-instabilities
uncovered in a Rushton turbine stirred tank by means of LES. A.I.Ch.E. Journal
60, 23832393.
Hartmann, H., Derksen, J.J., Montavo, C., Pearson, J., Hamill, I.S., Van den Akkar,
H.E.A., 2004b. Assessment of large eddy and RANS stirred tank simulations by
means of LDA. Chemical Engineering Science 59, 24192432.
Hu, L.W., Kazimi, M.S., 2006. LES benchmark study of high cycle temperature
fluctuations caused by thermal striping in a mixing tee. International Journal of
Heat and Fluid Flow 27, 5464.
Jaworski, Z., Zakrzewska, B., 2001. Modeling of the turbulence wall jet generated by
a pitched blade turbine impeller. The effect of turbulence model. In: Proceedings
of 10th European Conference on Mixing, pp. 187194.
Jenne, M., Reuss, M.A., 1999. Critical assessment on the use of k c turbulence
models for simulation of the turbulent liquid flow induced by a Rushton-turbine
in baffled stirred-tank reactor. Chemical Engineering Science 54, 39213941.
Jones, W.P., Launder, B.E., 1972. The prediction of laminarization with a two-
equation model of turbulence. International Journal of Heat and Mass Transfer 15,
301314.
Joshi, J.B., Patwardhan, A.W., 1999. Relation between flow pattern and blending in
stirred tanks. Industrial Engineering Chemical Research 38, 31313143.
Kim, W.W., Menon, S., 1997. Application of the localized dynamic subgrid-
scale model to turbulent wall-bounded flows. Technical Report AIAA-97-0210,
American Institute of Aeronautics and Astronautics, 35th Aerospace Sciences
Meeting, Reno.
Kostelich, E.J., Yorker, J.A., 1988. Noise reduction in dynamical systems. Physics
Reviews A 38, 16491656.
Launder, B.E., 1990. Phenomenological modeling: present and future?. In: Lumley,
J.L. (Ed.), Whither Turbulence? Turbulence at the Crossroads. Springer, Berlin,
pp. 439485.
Leonard, A., 1974. Energy cascade in large-eddy simulations of turbulent fluid flow.
Advances in Geophysics 18A, 237248.
Lien, F.S., Leschziner, M.A., 1994. Assessment of turbulence-transport models
including non-linear RNG eddy-viscosity formulation and second-moment
closure for flow over a backward facing step. Computers & Fluids 23, 9831004.
Murthy, B.N., Deshmukh, N.A., Patwardhan, A.W., Joshi, J.B., 2007. Hollow
self-inducing impellers: flow visualization and CFD simulations. Chemical
Engineering Science 62, 38393848.
Nere, N.K., 2001. Transport phenomena in multiphase reactors, Ph.D. Thesis.
University of Mumbai, Mumbai, India.
Nere, N.K., Patwardhan, A.W., Joshi, J.B., 2003. Liquid-phase mixing in stirred
vessels: turbulent flow regime. Industrial & Engineering Chemistry Research 42,
26612698.
Oshinowo, L., Jaworski, Z., Dyster, K.N., Marshall, E., Nienow, A.W., 2000. Predicting
the tangential velocity field in stirred tanks using the multiple reference frames
(MRF) model with validation by LDA measurements. In: Proceedings of 10th
European Conference on Mixing, Delft, Netherlands, pp. 247253.
Patwardhan, A.W., 2001. Prediction of flow characteristics and energy balance for a
variety of down flow impellers. Industrial & Engineering Chemical Research 40,
38063816.
Ranade, V.V., Joshi, J.B., 1989. Flow generated by pitched bladed turbine part i:
experimental. Chemical Engineering Communications 81, 197224.
Ranade, V.V., Mishra, V.P., Saraph, V.S., Deshpande, G.B., Joshi, J.B., 1992. Comparison
of axial flow impellers using LDA. Industrial & Engineering Chemistry Research
31, 23702379.
Revstedt, J., Fuchs, L., Tragardh, C., 1998. Large eddy simulation of the turbulent
flow in a stirredtank. Chemical Engineering Science 53, 40414053.
Revstedt, J., Fuchs, L., Kovac, T., Tragardh, C., 2000. Influence of impeller type on
the flow structure in a stirred reactor. A.I.Ch.E. Journal 46, 23732382.
Revstedt, J., Fuchs, L., Tragardh, C., 2002. Large eddy simulation of the turbulent
flow in stirred vessels. Chemical Engineering Technology 25, 443446.
Reynolds, W.C., 1987. Fundamentals of turbulence for turbulence modeling and
simulation. In: Lecture Notes for Von Karman Institute Agard. Report No. 755.
Rocchi, D., Zasso, A., 2002. Vortex shedding from a circular cylinder in a smooth and
wired configuration: comparison between 3D LES simulation and experimental
analysis. Journal of Wind Engineering and Industrial Aerodynamics 90, 475489.
Roussinova, V., Kresta, S.M., Weetman, R., 2003a. Low frequency macroinstabilities
in a stirred tank: scale-up and prediction based on large eddy simulations.
Chemical Engineering Science 53, 22972311.
Roussinova, V., Kresta, S.M., Weetman, R., 2003b. Low frequency macroinstabilities
in a stirred tank: scale-up and prediction based on large eddy simulations.
Chemical Engineering Science 58, 22972311.
Sahu, A.K., Kumar, P., Patwardhan, A.W., Joshi, J.B., 1999. CFD modeling and mixing
in stirred tanks. Chemical Engineering Science 54, 22852294.
Schafer, M., Yianneskis, M., Wachter, P., Durst, F., 1998. Trailing vortices around a
45

pitched-blade impeller. A.I.Ch.E. Journal 44, 12331246.


Speziale, C.G., Sarkar, S., Gatski, T.B., 1991. Modeling the pressure-strain correlation of
turbulence: an invariant dynamical systems approach. Journal of Fluid Mechanics
227, 245272.
Wang, Y., Yuan, G., Yoon, Y.K., Allen, M.G., Bidstrup, S.A., 2006. Large eddy simulation
(LES) for synthetic jet thermal management. International Journal of Heat and
Mass Transfer 49, 21732179.
Yeoh, S.L., Papadakis, G., Yianneskis, M., 2004. Numerical simulation of turbulent
flow characteristics in a stirred vessel using the LES and RANS approaches
with the sliding/deforming mesh methodology. Chemical Engineering Research
& Design 82, 834848.
Yeoh, S.L., Papadakis, G., Yianneskis, M., 2005. Determination of mixing time and
degree of homogeneity in stirred vessels with large eddy simulation. Chemical
Engineering Science 60, 22932302.

Das könnte Ihnen auch gefallen