You are on page 1of 9

Research Paper

Journal of Chemical Engineering of Japan, Vol. 47, No. 5, pp. 373381, 2014

CFD Prediction of Effects of Impeller Design on Floating Solids


Mixing in Stirred Tanks with Pitched Blade Turbines
Shengchao Qiao, Rijie Wang, Xiaoxia Yang and Yuefei Yan
School of Chemical Engineering and Technology, Tianjin University, 92 Weijin Road, Nankai District,
Tianjin 300072, China
Keywords: Floating Solids, CFD, Flow Pattern, Stirred Tank, PBT, Two-Phase Flow
Computational fluid dynamics (CFD) simulations were performed to study effects of the flow pattern on the mixing quality of floating solids in stirred tanks with up- and down-pumping pitched blade turbine (PBTU and PBTD). The variables
about impeller design have influences in the flow pattern and four key ones, impeller diameter, clearance from tank
bottom, blade angle and number, were investigated in this work. EulerEuler multiphase model along with mixture k
turbulence model was adopted. The relative motion between rotating impeller and stationary baffles was modeled with
the multiple reference frame (MRF) approach. The predicted results were compared and explained with the experimental
or theoretical results in published literature. Qualitative and reasonable agreement was achieved.

Introduction
O-surface suspension of floating solids, which float at
the liquid surface without agitation, is commonly encountered in the processes involving mechanical stirring, such
as fermentation, mineral flotation, sewage treatment and
polymerization reactions. In a stirred tank, floating solids
are firstly drawn down from the liquid surface and then
distributed by the impeller throughout the tank. Hence, the
most critical place in a stirred tank is the liquid surface and
the diculty of solids drawdown depends much on the flow
pattern (Khazam and Kresta, 2008), which is strongly aected by the impeller design (Kumaresan and Joshi, 2006). Two
typical flow patterns produced by dierent PBT design are
single-eight and double-eight (Sharma and Shaikh, 2003).
For the single-eight flow pattern, a single circulation loop
fills the entire tank. While for the double-eight pattern, a
secondary circulation loop is formed in addition to a primary loop. An important performance criteria, the standard
deviation () of local solid concentrations, is widely used to
quantify the mixing quality in solidliquid systems and evaluate the competing impeller design (Khopkar et al., 2006).
Current methods used for studying the impeller design
are computational fluid dynamics (CFD), experimental fluid
dynamics (EFD) and empirical correlation method. CFD
method permits the numerical modeling of varied geometries and can provide qualitative or quantitative flow hydrodynamics underlying the system, especially the velocity field
and turbulent flow field. Such information can be obtained
using EFD method, but only by some expensive and complex equipment like particle image velocimetry. Three emReceived on November 18, 2013; accepted on January 6, 2014
DOI: 10.1252/jcej.13we313
Correspondence concerning this article should be addressed to X. Yang
(E-mail address: xxytju@163.com).
Vol. 47 No.52014
Copyright
2014The Society of Chemical Engineers, Japan

pirical correlations for complete drawdown of floating solids


in stirred tanks are found in published literature (Joosten
et al., 1977; Takahashi and Sasaki, 1999; Tagawa et al.,
2006), but they are more applicable to their preferred configurations. The complex interactions between the variables
about impeller design, additionally, make it unlikely to use
a simple empirical correlation to reliably predict the mixing
performance in even a small range. Hence, CFD method has
been promoted as a powerful tool to study the impeller design and multiphase flow.
Complete drawdown of floating solids is usually required
for practical applications, because the maximum surface
area of particles is available for chemical reaction, mass or
heat transfer. The corresponding impeller speed (N) for
complete drawdown is defined as the critical impeller speed
(NCS). If N is increased beyond NCS, the rate of mass transfer
process increases very slowly, and it is cost ineective to further increase the power input (P).
Previous works (Bakker and Frijlink, 1989; Kuzmani and
Ljubii, 2001; zcan-Takin and McGrath, 2001) indicated
that the mixed flow impeller, PBT, was more energy ecient
to achieve complete drawdown than the radial flow impellers. Kuzmani and aneti (1999), Kuzmani and Ljubii
(2001) proposed a new EFD method for determining complete drawdown of floating solids and associated NCS based
on a special behavior of mixing time with respect to N, and
then studied eects of pitched brade turbine (PBT) design in
a stirred tank.
However, CFD studies on the floating solids mixing in
stirred tanks with PBT and especially its dependence on the
impeller design can be hardly found in published literature.
Hence, the main objective of this work is to perform CFD
simulations to illustrate the flow pattern produced by different impeller design and compare its eect on the mixing
quality, especially complete drawdown of floating solids in a
stirred tank similar to that of Kuzmani and aneti (1999),
373

Kuzmani and Ljubii (2001).

1.CFD Modeling
1.1Model equation
A commercial software Fluent 6.3 (ANSYS, Inc.) was
used for CFD modeling. The multiphase flow was simulated
with the Euler granular multiphase (EGM) model based
on the EulerEuler calculation, in which dierent phases
are treated mathematically as the interpenetrating continua.
Only a brief account of CFD model was given as following
and further details could be found in the literature (Ranade,
2002). The motion of each phase is governed by respective
mass and momentum conservation equations without considering mass transfer:

( )+ (i i ui ) = 0
t i i

(1)

( u ) + (i i ui ui ) = ip + ( i )
t i i i
+ i i g + Mi

(2)

Here i=l and s denote the continuous phase, liquid, and


dispersed phase, solid; is phase volume fraction; u is mean
velocity vector; p is the pressure shared by all phases; is
stressstrain tensor; g is gravitational acceleration; Mi is the
inter-phase force term.
1.2Inter-phase force term
Only the contribution of drag force was considered in
this work, since it had been reported that no-drag forces,
that is, lift force, virtual mass force and turbulent dispersion
force, had little eect on the solidfluid hydrodynamics in
stirred tanks (Ljungqvist and Rasmuson, 2001; Montante et
al., 2001). Drag force exerted by the dispersed phase on the
continuous phase is calculated as
FD =

3 CD
u ul (us ul )
4 dp l s s

(3)

Where dp is the particle diameter and CD is the drag coefficient obtained from the Brucato model (Brucato et al.,
1998), which could accounts for the influence of the free
stream turbulence on solid distribution and had been successfully employed in solidliquid systems, especially for
low solid loadings (Kasat et al., 2008; Ochieng and Onyango,
2008). Thus CD is calculated by the following equations:
3
CD / CD0
= 8.76 10/4 ( dp / )
CD0

CD0 =

24
1+ 0.15Re p0.687
Rep

(4)
(5)

Here CD0 refers to the value in the quiescent liquid and is


the Kolmogorov length scale.
1.3Turbulence model
Three extensions of the standard k- turbulence model
374

(mixture, dispersed and per-phase) are developed to simulate the turbulent motion in multiphase systems. For the
mixture k model, only a couple of k and equations with
the physical properties of the mixture are solved and two
phases are assumed to share the same values of k and .
Montante and Magelli (2005) compared the axial profiles of
solid concentration predicted by three models, and found
the mixture k model was the most proper turbulence
model due to its less computational demand and qualitatively fair representation of the solid distributions. Moreover, this model is very applicable to that the density ratio
between the dispersed phase and continuous phase is close
to 1 (ANSYS, 2009). The solid and liquid densities simulated
in this work were 840 and 996 kg/m3, respectively. Hence,
the mixture k model was adopted.
1.4Solution domain and system
The investigated system consisted of a flat-bottomed cylindrical tank with the diameter T=0.32 m and liquid height
H=T. Four full baes of width 1/10 T were symmetrically
mounted perpendicular to the wall. For an up- and downpumping PBT with a fixed blade height h=1/5 D and blade
thickness t=2 mm, the studied impeller design variables
were: impeller diameter, D=0.250.41 T; o-bottom clearance, C=0.330.66 H; blade inclined angle to the horizontal,
= 3060; blade number, n=4 and 6. To reduce the computational requirement and numerical eort, one quarter
and one half of the symmetric domain was modeled, respectively, for 4- and 6-bladed PBT. The simulated continuous
phase and dispersed phase were tap water (l =996 kg/m3;
l =1103 Pa s) and polyethylene particles (dp =205 m;
s =840 kg/m3). For all the simulations, the dilute particles
with a mean solid loading of 5 kg/m3 (0.6% v/v) were initially uniformly distributed in the liquid.
1.5Solution method
The relative motion between rotating impeller and stationary baes was modeled by the MRF approach. The divided solution domain consisted of a central cylinder region
associated with the impeller discharge stream in a rotating
frame of reference and the rest region in a stationary frame
of reference. The diameter and height of the central cylinder region were 0.6 T and 4 h, respectively. Hexahedral and
tetrahedral elements were used to mesh the computational
domain, and finer grid was used in the impeller region to
ensure an adequate prediction of turbulent characteristic
produced by the impeller.
The periodic boundary condition was used for the left
and right halves of the symmetry plane. The symmetry
boundary condition (zero normal velocity gratitude) was assumed on the free liquid surface without considering the air
entrainment from the headspace and large surface vortex,
which is suppressed by the full baes. The no-slip boundary condition with standard wall function was assumed on
the tank wall, the surfaces of the baes, impeller and shaft.
All terms of governing equations were discretized by
a second-order upwind scheme and the phase coupled
Journal of Chemical Engineering of Japan

SIMPLE algorithm was used for pressurevelocity coupling.


The time step was set to 0.001 s and the transient calculations were considered to be converged when the total residuals for all the equation dropped below 104. Further
checks for the convergence were made by observing whether
the volume average of and mass integral of for the whole
domain keep constant.

2.Results and Discussion


2.1Preliminary simulation
Preliminary simulations were performed to simulate the
flow produced by the standard up- and down-pumping
pitched blade turbine (PBTU) and (PBTD) of D=0.33 T,
C=0.33 H, = 45, n=4 at N=4001,000 rpm. The number
of grid cells varied from 300 k to 900 k was used to check
the sensitivity of simulated results with respect to the grid
size. The comparison of every two consecutive cases at an
interval of 100 k showed a reduction in the dierence between the predicted velocity field and solid distributions, but
the corresponding computation time increased as the grid
was fined. Around 600 k grid cells were found sucient to
describe the key flow characteristics, and the mass integral
of for the whole domain showed no significant changes,
less than 3%, with further grid refinement. Considering its
good compromise between accuracy and computational effort, the simulations were performed with 600 k grid cells
for one quarter of the analyzed region for 4-bladed PBT. For
6-bladed PBT, around 1.2 million grid cells of same size are
used to mesh one half of the analyzed region.
The velocity fields produced by PBTD in a stirred tank
of Kresta and Wood (1993) were simulated to indicate the
validity of CFD simulations. The axial circulation loop was
mainly responsible for solid suspension. Thus the results of
four cases for dierent C were shown in the form of axial
liquid velocity (Vz) over a radial traverse at the same distance from the lower edge of blades (Figure 1). Good agreement was achieved between EFD and CFD results, especially
for the radial positions of the peaks. Moreover, predicted
pumping numbers (NQ) increased from 0.54 to 0.78 as C
was increased from 0.155 D to 0.93 D, which also agreed well

Fig. 1 Axial velocity profiles from experiment and simulation for


cases of C/D=0.1551
Vol. 47 No. 5 2014

with experiment data.


Therefore, this work performed CFD simulations to study
eects of impeller design on the flow pattern and floating
solids mixing.
2.2Effect of impeller diameter
The eect of impeller diameter was firstly investigated. CFD studies were carried out with PBTU and PBTD
of diameters D1 =0.08 m (D1/T=0.25), D2 =0.106 m
(D2/T=0.33) and D3 =0.132 m (D3/T=0.41) positioned at
C=0.106 m (C/H=0.33).
This work extended a way to predict NCS based on the
variation of with respect to N. This way had been successfully employed by Khopkar et al., (2006) and Murthy et
al., (2007) for sinking solids, which stay at the tank bottom
without agitation. Bohnet and Niesmak (1980) firstly quantified the degree of homogeneity for solidliquid mixing
using , which is defined as:
=

1
m

ci

i=1

avg

( c

1)2

(6)

Where m denotes the number of sampling locations for


measuring local solid concentration (c). The precise calculation of requires the data of c in the whole tank, available
only from CFD results. The degree of homogeneity improves
as decreases, and complete drawdown is achieved when
lies between 0.2 and 0.8.
The distributions of local solid concentration at 600 rpm
for PBTD and PBTU are displayed, respectively, in Figures
2(a) and (b). For PBTD of diameter D1, the particles accumulate around the shaft. As D is increased, many particles
are drawn down and caught in the circulation flow produced

Fig. 2 Solid distributions for (a) PBTD and (b) PBTU at C=0.33 H
(from left to right, D1 =0.25 T, D2 =0.33 T, D3 =0.41 T)
375

Fig. 3 vs. N for three diameters of PBTD

Fig. 5 Velocity fields for (a) PBTD and (b) PBTU at C=0.33 H (from
left to right, D1 =0.25 T, D2 =0.33 T, D3 =0.41 T)

Fig. 4 vs. N for three diameters of PBTU

by the impeller. For PBTU, this phenomenon seems to be


more remarkable. Additionally, the eye of the primary circulation loop in which the particles are caught moves downward as D is increased.
Both curves of versus N for PBTD and PBTU show that
for all impeller diameters decrease sharply with increasing
N and finally tend to flat o (Figures 3 and 4). For PBTD,
has an increasing-decreasing behavior as D is increased
at same N, opposite to for PBTU. Following the description of Murthy et al., (2007) who found a sharp reduction
in as N approached to NCS, NCS for D1 and D2 of PBTD
lie between 700 and 800 rpm, while NCS for D3 lies around
600 rpm. These NCS results are in reasonable agreement with
EFD results of Kuzmani and aneti (1999), thus it can
infer that complete drawdown is achieved at a lower N for
larger impeller (D3).
This result can be attributed to the increased liquid velocity and decreased decay in the turbulence during the flow
path, both resulted from an increase in D. The single-eight
flow patterns produced by three diameters PBTD are displayed with the same scale at 700 rpm in Figure 5(a). As D
is increased, the circulation flow near the bottom becomes
more vigorous and its eye moves outward. However, this
also leads to an accumulation of the particles caught in the
circulation loop at high N, and larger impeller makes this effect more significant. This is also the cause that the ultimate
376

for larger impeller (D2 and D3) is higher than smaller one
(D1).
For PBTU of D3 =0.41 T, complete drawdown can be
achieved at a much lower N around 500 rpm. But the ultimate is almost twice higher than the other two. The mixing quality of floating solids seems sensitive to D for PBTU.
This is attributed to the double-eight flow pattern produced
by the up-pumping mode of PBT. Seeing from the velocity
fields produced by PBTU at 700 rpm shown with the same
scale in Figure 5(b), the impeller discharge stream hits the
tank wall to generate two loops and an increase in D makes
the primary loop in the lower part of the tank become stronger and its eye moves down and outward. The particles are
more easily trapped in the downward flow of the primary
loop once drawn down from the liquid surface, and such an
eect becomes more noticeable as D is increased.
Comparisons of for three D of PBTU and PBTD indicate that PBTU is overall less ecient for o-surface suspension of floating solids on the condition of C=0.33 H.
For PBTU, the diculty to achieve complete drawdown
increases due to the double-eight flow pattern. Moreover,
this flow pattern requires more P to achieve complete drawdown. In previous EFD studies, some authors (Gray, 1987;
Sharma and Shaikh, 2003; Khazam and Kresta, 2009) found
the similar phenomena when the double-eight flow pattern
was responsible for o-surface suspension of floating solids
or o-bottom suspension of sinking solids.
2.3Effect of impeller clearance
The location of the impeller in the stirred tank, that
is, o-bottom clearance may also aect the solid distribution. So the eects were studied with PBTU and PBTD
of D=0.106 m (D/T=0.33) positioned at three clearances:
C1 =0.106 m (C1/H=0.33), C2 =0.157 m (C2/H=0.49) and
C3 =0.212 m (C3/H=0.66).
Journal of Chemical Engineering of Japan

Fig. 6 vs. N for three clearances of PBTU

Fig. 8 Solid distributions for PBTU of D=0.33 T: (a) C1 =0.33 H; (b)


C2 =0.49 H; (c) C3 =0.66 H

Fig. 7 vs. N for three clearances of PBTD

For both PBTU and PBTD, decreases dramatically as C


is increased from C1 to C3 at same N (Figures 6 and 7). Thus
the predicted NCS is much lower at C2 and C3. Combined
with the conclusion obtained in previous section that PBTU
is overall less ecient at C1 =0.33 H, it infers that PBTU
seems particularly more sensitive to o-bottom clearance,
performing slightly better than PBTD at C2 and C3, but
worse than PBTD at C1. This agrees reasonably with that of
Khazam and Kresta (2009). The solid distributions at three
C for PBTU at 600 rpm are displayed in Figure 8. As C is
increased, more and more particles are pulled into the liquid
and follow the path of circulation loop, which rises to the
liquid surface at C2 and C3.
Two main drawdown mechanisms proposed by Khazam
and Kresta (2008) based on the turbulent fluctuations and
mean drag at the liquid surface are tried to explain the eect
of C on and NCS. The velocity fields predicted for PBTU at
three C at 600 rpm are displayed within the same range in
Figure 9(a). Although the maximum value is almost same
near the impeller tip, the double-eight flow pattern is generated at C1 and its secondary loop is characterized by low
velocity, low turbulence and radially-inward flow near the
liquid surface, which was reported not applicable to osurface suspension of floating solids (Khazam and Kresta,
2009). While at C2 and C3, the impeller discharge stream
directly acts on the particles near the liquid surface, so the
Vol. 47 No. 5 2014

Fig. 9 Velocity fields for (a) PBTU and (b) PBTD of D=0.33 T (from
left to right, C1 =0.33 H, C2 =0.49 H, C3 =0.66 H)

mean circulation velocity are relatively higher to be able to


draw down the particles at lower N. However, many particles may accumulate by the force of this strong single-eight
loop as N is further increased, which eventually decreases
the degree of homogeneity and leads to a slight increase in
(Figure 6). In other words, once the maximum homogeneity is achieved, further increase in N was not beneficial and
even be detrimental. Hosseini (2008) observed the similar
phenomenon in their researches about solidliquid mixing.
For PBTD, the discharge stream moves downward and
then hit the wall or the bottom to generate a diverging or
reverse flow in Figure 9(b). The contours of turbulent kinetic energy (k) at the liquid surface indicate that k is overall
highest at C3 and lowest at C1 in Figure 10. The decay of
377

Fig. 10 Contours of k at the liquid surface for PBTD of D=0.33 T: (a)


C1 =0.33 H; (b) C2 =0.49 H; (c) C3 =0.66 H

Fig. 11 vs. N for 30, 45 and 60 PBTU

turbulence during the flow path may decrease as PBTD


rises from C1 to C3. Thus the drawdown of floating solids
becomes easier at larger clearances.
To sum up, the higher the impeller is located, the more
vigorous is the action of the flow and turbulent fluctuations
to which floating solids are exposed, therefore complete
drawdown will commence at lower N.
To find an optimum C, the impeller further rises to position C4 (C4/H=3/4). Predicted for PBTU from 500 to
700 rpm show an increase from 0.147 to 0.175, which is
slightly lower than those at C3. But the corresponding P increases around 0.4%. For PBTD, P also increases more than
2.1%, whereas the homogeneity has no significant improvement. This means further increase in C cannot significantly
improve the energy eciency.
2.4Effect of blade inclined angle
Blade inclined angle () was found to have very strong effects on the flow pattern (Ranade, 1989). Therefore, the flow
patterns produced by 30 and 60 PBTU and PBTD were
studied in addition to 45 PBT used in previous sections.
The impeller diameter D/T=0.33 and impeller clearance
C/H=0.33.
decreases as is increased at same N, and an eect of
on floating solids mixing is more significant for PBTD
in Figures 11 and 12. N required for complete drawdown
decreases with an increase in . The highest homogeneity is
achieved for 60 PBTD with approaching 0.1 at 600 rpm in
Figure 12. The results exhibit a qualitative agreement with
the literature data (Driss et al., 2010). For PBTD, the solid
distributions for three at 600 rpm are displayed in Figure
13. For = 30, most particles accumulate near the shaft and
liquid surface, leading to a low homogeneity. For = 45,
many particles are pulled into the circulation loop near the
bottom. While for = 60, the formation of the regions with
high solid concentration is not observed.
It was reported that the pumping capacity was closely
related to the solid dispersion and it depended much on
(Mao et al., 1997). Therefore, the axial pumping number
NQax and radial pumping number NQrad are calculated, respectively, by the integration of axial volume flow rate (Qax)
378

Fig. 12 vs. N for 30, 45 and 60 PBTD

Fig. 13 Solid distributions for PBTD of D=0.33 T at C=0.33 H: (a)


30; (b) 45; (c) 60

and radial volume flow rate (Qrad) of the impeller discharge


stream.
N Qax = Qax /ND 3

(7)

Qax = Vax dAax

(8)

N Qrad = Qrad /ND 3

(9)

Journal of Chemical Engineering of Japan

Table 1NQ for three of both PBT at 600 rpm


NQrad

NQax

[o]
30
45
60

PBTU

PBTD

PBTU

PBTD

0.569
0.917
1.307

0.531
0.965
1.180

0.081
0.177
0.207

0.064
0.150
0.181

Fig. 15 Solid distributions for 6-bladed PBTU of D=0.33 T at C=


0.33 H: (a) 600 rpm; (b) 800 rpm

Fig. 14 vs. N for 4- and 6-bladed PBTU and PBTD

Qrad = Vrad dArad

(10)

Where Vax and Vrad are axial and radial liquid velocity; the
axial integral region (Aax) is a circular surface with the same
diameter as the impeller just above and below the blade
edge, respectively, for PBTU and PBTD; the radial integral
region (Arad) is a cylindrical surface with the same height as
the blade at the impeller tip for both PBTU and PBTD.
In Table 1, NQ for PBTU is higher than that for PBTD and
NQax is much higher than NQrad for both PBTU and PBTD
at 600 rpm. Moreover, both pumping numbers, especially
NQax, increase as is increased, which agrees qualitatively
with EFD results of Ranade and Joshi (1989). zcan-Takin
and Wei (2003) reported that it was the axial flow to be
mostly responsible for o-surface suspension of floating
solids. Thus the diculty of floating solids mixing decreases
as is increased. However, the associated energy dissipation
behind the blades increases.
2.5Effect of blade number
4- and 6-bladed PBT are most widely used in stirred vessels. Thus the eect of blade number on floating solids mixing in stirred tanks was investigated by simulating the flow
produced by 4- and 6-bladed PBT. The D/T and C/H ratios
are both 0.33.
The curves of versus N show that 6-bladed PBT perform
better to achieve higher degree of homogeneity in Figure
14. For PBTU, even decrease from 0.93 to 0.23 at 400 rpm.
The maximum homogeneity is achieved at 500600 rpm,
beyond which the homogeneity has no improvement and
even begins to decrease. This can be attributed to the vigorous circulation flow discharged from 6-bladed PBTU. Seeing
from the solid distributions in Figure 15, both the primary
Vol. 47 No. 5 2014

Fig. 16Velocity fields for 4- and 6-bladed (a) PBTU (b) PBTD

Fig. 17 vs. P for 4- and 6-bladed PBTU and PBTD

and secondary circulation loops are very strong to make the


particles trapped in them. At 800 rpm, this phenomenon
becomes more remarkable. This also confirms that it is not
beneficial to further increase N once the maximum homogeneity is achieved.
The predicted velocity fields produced by 4- and 6-bladed
PBTU and PBTD at 700 rpm are displayed within the same
range in Figure 16. 6-bladed PBT generate a stronger loop
than 4-bladed PBT, though the positions of the loop center
have no significant dierence. Hence, 6-bladed PBT can pull
379

the particles down from the liquid surface at lower N. However, an increase in blade number makes the energy dissipation behind the blades also increase.
Thus although 6-bladed PBT can achieve better performance, it requires more P. So were plotted against P to
further compare the mixing ability between 4- and 6-bladed
PBT in Figure 17. The results point to that 6-bladed PBT are
more energy ecient than 4-bladed PBT.

Conclusions
CFD simulations based on the EulerEuler approach
along with mixture k turbulence model were performed
to qualitatively investigate eects of impeller design on the
mixing quality of floating solids in stirred tanks with PBTU
and PBTD. The mean solid loading is 5 kg/m3 (0.6% v/v).
The main conclusions are as follows:
With an increase in D, as well as n, both and NCS decrease, but the associated P increases due to more energy
dissipation behind the blades.
The double-eight flow pattern requires more P for floating solids mixing.
For the impellers of dierent diameters positioned at
C=0.33 H, PBTU is overall less ecient than PBTD, especially for large impeller D3 =0.41 T.
Once the maximum homogeneity is achieved, further
increase in N is not beneficial and even be detrimental.
The drawdown performance is sensitive to C for both
PBTU and PBTD, but much more for PBTU. Further
increase in C above 0.66 H cannot further improve the
energy eciency for both PBT impellers.
has a more significant eect for PBTD and the impeller
pumping capacity can explain the eect of .
6-Bladed PBT are more energy ecient than 4-bladed
PBT.
CFD simulations show many promising results, which are
useful for extending the application of CFD to two-phase
flow involving floating solids and providing references for
impeller design for floating solids mixing in stirred vessels.
Nomenclature
A
C
CD
CD0
c
cavg
D
dp
FD
g
H
h
k
M
N
NCS

380

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

integral area
clearance from bottom to center of blades
drag coecient in turbulent liquid
drag coecient in quiescent liquid
local solid concentration
average solid concentration
impeller diameter
particle diameter
drag force
gravitational constant
liquid height
blade height
turbulent kinetic energy
inter-phase force
impeller speed
critical impeller speed

[m2]
[m]
[]
[]
[kg/m3]
[kg/m3]
[m]
[m]
[N/m3]
[m/s2]
[m]
[m]
[m2/s2]
[N/m3]
[rpm]
[rpm]

NQ
n
P
Q
Rep
t
T
V
v/v

=
=
=
=
=
=
=
=
=

pumping number
blade number
power input
volume flow rate
particle Reynolds number
blade thickness
tank diameter
velocity vector
solid volume/total volume

=
=
=
=
=
=
=
=

solid volume fraction


blade angle to the horizontal axis
turbulence dissipation rate
Kolmogorov length
density
viscosity
shear stress
standard deviation

[]
[]
[W]
[m3/s]
[]
[m]
[m]
[m/s]
[]
[]
[o]
[m2/s3]
[m]
[kg/m3]
[Pa s]
[N/m]
[]

Subscripts
l
= liquid phase
s
= solid phase
ax
= axial direction
rad
= radial direction
tip
= blade tip

Literature Cited
ANSYS, Inc.; Ansys Fluent Theory Guide, Canonsburg, U.S.A. (2009)
Bakker, A. and J. J. Frijlink; The Drawdown and Dispersion of Floating
Solids in Aerated and Unaerated Stirred Vessels, Chem. Eng. Res.
Des., 67, 208210 (1989)
Bohnet, M. and G. Niesmak; Distribution of Solids in Stirred Suspensions, Ger. Chem. Eng., 3, 5765 (1980)
Brucato, A., F. Grisafi and G. Montante; Particles Drag Coecient in
Turbulence Fluids, Chem. Eng. Sci., 53, 32953314 (1998)
Driss, Z., G. Bouzgarrou, W. Chtourou, H. Kchaou and M. S. Abid;
Computational Studies of the Pitched Blade Turbines Design
Eect on the Stirred Tank Flow Characteristics, Eur. J. Mech. B/
Fluids, 29, 236245 (2010)
Gray, D. J.; Impeller Clearance Eect on o Bottom Particle Suspension in Agitated Vessels, Chem. Eng. Commun., 61, 151158
(1987)
Hosseini, S.; SolidLiquid Mixing in Agitated Tanks: Experimental and
CFD Analysis, M.A.Sc., Thesis, Ryerson University, Canada (2008)
Joosten, G. E. H., J. G. M. Schilder and A. M. Broere; The Suspension
of Floating Solids in Stirred Vessels, Trans. Inst. Chem. Eng., 55,
220223 (1977)
Kasat, G. R., A. R. Khopkar, V. V. Ranade and A. B. Pandit; CFD Simulation of LiquidPhase Mixing in SolidLiquid Stirred Reactor,
Chem. Eng. Sci., 63, 38773885 (2008)
Khazam, O. and S. M. Kresta; Mechanisms of Solids Drawdown in
Stirred Tanks, Can. J. Chem. Eng., 86, 622634 (2008)
Khazam, O. and S. M. Kresta; A Novel Geometry for Solids Drawdown
in Stirred Tanks, Chem. Eng. Res. Des., 87, 280290 (2009)
Khopkar, A. R., G. R. Kasat, A. B. Pandit and V. V. Ranade; Computational Fluid Dynamics Simulation of the Solid Suspension in a
Stirred Slurry Reactor, Ind. Eng. Chem. Res., 45, 44164428 (2006)
Kresta, S. M. and P. E. Wood; The Mean Flow Field Produced by a 45
Pitched Blade Turbine: Changes in the Circulation Pattern due to
O Bottom Clearance, Can. J. Chem. Eng., 71, 4253 (1993)
Kumaresan, T. and J. B. Joshi; Eect of Impeller Design on the Flow

Journal of Chemical Engineering of Japan

Pattern and Mixing in Stirred Tanks, Chem. Eng. J., 115, 173193
(2006)
Kuzmani, N. and B. Ljubii; Suspension of Floating Solids with UpPumping Pitched Blade Impellers; Mixing Time and Power Characteristics, Chem. Eng. J., 84, 325333 (2001)
Kuzmani, N. and R. aneti; Influence of Floating Suspended Solids
on the Homogenization of the Liquid Phase in a Mixing Vessel,
Chem. Eng. Technol., 22, 943950 (1999)
Ljungqvist, M. and A. Rasmuson; Numerical Simulation of the TwoPhase Flow in an Axially Stirred Vessel, Chem. Eng. Res. Des., 79,
533546 (2001)
Mao, D., L. Feng, K. Wang and Y. Li; The Mean Flow Field Generated
by a Pitched Blade Turbine: Changes in the Circulation Pattern
due to Impeller Geometry, Can. J. Chem. Eng., 75, 307316 (1997)
Montante, G. and F. Magelli; Modelling of Solids Distribution in
Stirred Tanks: Analysis of Simulation Strategies and Comparison
with Experimental Data, Int. J. Comput. Fluid Dyn., 19, 253262
(2005)
Montante, G., G. Micale, F. Magelli and A. Brucato; Experiments and
CFD Predictions of Solid Particle Distribution in a Vessel Agitated with Four Pitched Blade Turbines, Chem. Eng. Res. Des., 79,
10051010 (2001)
Murthy, B. N., R. S. Ghadge and J. B. Joshi; CFD Simulations of Gas
LiquidSolid Stirred Reactor: Prediction of Critical Impeller Speed
for Solid Suspension, Chem. Eng. Sci., 62, 71847195 (2007)

Vol. 47 No. 5 2014

Ochieng, A. and M. S. Onyango; Drag Models, Solids Concentration


and Velocity Distribution in a Stirred Tank, Powder Technol., 181,
18 (2008)
zcanTakin, G. and C. McGrath; Draw Down of Light Particles in
Stirred Tanks, Chem. Eng. Res. Des., 79, 789794 (2001)
zcanTakin, G. and H. Wei; The Eect of Impeller-to-Tank Diameter Ratio on Draw Down of Solids, Chem. Eng. Sci., 58, 20112022
(2003)
Ranade, V. V.; Computational Flow Modelling for Chemical Reactor
Engineering, Academic Press, pp. 3683, New York, U.S.A. (2002)
Ranade, V. V. and J. B. Joshi; Flow Generated by Pitched Blade Turbines I: Measurements Using Laser Doppler Anemometer, Chem.
Eng. Commun., 81, 197224 (1989)
Sharma, R. N. and A. A. Shaikh; Solid Suspension in Stirred Tanks
with Pitched Blade Turbines, Chem. Eng. Sci., 58, 21232140
(2003)
Tagawa, A., N. Dohi and Y. Kawase; Dispersion of Floating Solid
Particles in Aerated Stirred Tank Reactors: Minimum Impeller
Speeds for OSurface and Ultimately Homogeneous Solid Suspension and Solids Concentration Profiles, Ind. Eng. Chem. Res.,
45, 818829 (2006)
Takahashi, K. and S. I. Sasaki; Complete Drawdown and Dispersion of
Floating Solids in Agitated Vessel Equipped with Ordinary Impellers, J. Chem. Eng. Japan, 32, 4044 (1999)

381

You might also like