You are on page 1of 10

Chemical Engineering Science 57 (2002) 487496

www.elsevier.com/locate/ces
Experimental and modelling study of gas dispersion in a double
turbine stirred tank
S. S. Alves

, C. I. Maia, J. M. T. Vasconcelos
Centro de Eng. Biol ogica e Qumica, Department of Chemical Engineering, Instituto Superior T ecnico, 1049-001-Lisboa, Portugal
Received 7 February 2001; received in revised form 16 July 2001; accepted 7 August 2001
Abstract
Gas dispersion in a double turbine stirred tank is experimentally characterised by measuring local gas holdups and local bubble size
distributions throughout the tank, for three liquid media: tap water, aqueous sulphate solution and aqueous sulphate solution with PEG.
For all these media, bubble coalescence generally prevails over breakage. Where average bubble size decreases, this can be attributed to
the dierence in slip velocity between dierent sized bubbles. Most of the coalescence takes place in the turbine discharge stream.
A compartment model that takes into account the combined eect of bubble coalescence and breakage is used to simulate gas dispersion.
The model predicts spatial distribution of gas holdup and of average bubble size, with average bubble size at the turbines as an input.
Reasonable agreement between experiment and simulation is achieved with optimisation of two parameters, one aecting mainly the slip
velocity, the other related mainly to the bubble coalescence}breakage balance. Dierent sets of parameters are required for each of the
three liquid systems under study, but are independent of stirring}aeration conditions. The model only fails to simulate the smaller average
bubble diameters at the bottom of the tank. ? 2002 Elsevier Science Ltd. All rights reserved.
Keywords: Stirred tank; Gas dispersion; Bubble size
1. Introduction
Gas dispersion in stirred tanks is very important, since it
strongly inuences gasliquid mass transfer. It is an exceed-
ingly complex phenomenon, involving not only the com-
plexity of the continuous phase owitself and its interactions
with the dispersed phase, but also the behaviour and interac-
tions within the dispersed phase, including bubble breakage
and coalescence.
Modelling eorts have increased in recent years. Some
models predict the gas holdup distribution throughout the
tank, but make no attempt at predicting bubble sizes. These
include Eulerian}Eulerian CFD two-uid models which
assume a given constant bubble size to calculate the in-
teraction between phases (Gosman, Lekakou, Politis, Issa,
& Looney, 1992; Morud & Hjertager, 1996; Friberg, 1998;
Lane, Scwartz, & Evans, 2000), the Eulerian}Lagrangian
treatment of Patterson (1991) and the two-dimensional net-
work of zones proposed by Mann (1986) and Mann and
Hackett (1988).

Corresponding author. Tel.: +351-1-8417188; fax: +351-1-8499242.


E-mail address: salves@alfa.ist.utl.pt (S. S. Alves).
Prediction of average bubble sizes was included in mod-
els by Bakker and van den Akker (1994) and by Djeb-
bar, Roustan, and Line (1996), who performed population
balances which include a term to describe the combined
eects of bubble breakage and coalescence. A development
of Bakker and van den Akkers model is the prediction of
bubble size distributions (Venneker, 1999), allowed by in-
cluding population balances on classes of bubbles, together
with a more elaborate treatment of breakage}coalescence in-
spired in Tsouris and Tavlarides (1994) work.
CFD complexity implies that the nal results depend on
a considerable number of modelling options and assump-
tions and on a set of parameters over the value of which
there is but little experimental control. Very few data have
been published on the spatial distribution of bubble sizes
in stirred tanks to assess and guide the modelling eort. A
notable exception is the work by Greaves and co-workers
(e.g., Barigou, 1987; Barigou & Greaves, 1992), who used a
photoelectric capillary suction probe as the measuring tech-
nique in a single turbine stirred tank, with coalescing liquid
media. Takahashi and Nienow (1993) used a photographic
technique, requiring very lowgas holdups to measure bubble
sizes away from the tank wall. The same is true in the work
0009-2509/02/$ - see front matter ? 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0009- 2509( 01) 00400- 6
488 S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496
of Sch afer, W achter, and Durst (2000), who used Phase
Doppler Anemometry, with a mixture of silicone oils as the
liquid medium.
In this work, gas dispersion in a double turbine stirred
tank is experimentally characterised by measuring local gas
holdups and local bubble size distributions at dierent agi-
tation speeds in both coalescing and non-coalescing contin-
uous phases. Some modelling problems are then addressed
and emphasised using a simple compartment model that
takes into account bubble coalescence.
2. Experimental
The experimental set-up consisted of a 0.292 m diam-
eter, at-bottomed, fully baed Perspex vessel. Agitation
was provided by two 0.096 m diameter standard Rushton
turbines set at distances of 0.146 and 0.438 m, respectively
above the tank base. The experimental}modelling grid, con-
sisting of 30 cells and 28 sampling points, is shown in
Fig. 1.
All experiments were carried out in the vertical mid-plane
between two adjacent baes and a total liquid height of
0.584 m. The liquid media used were tap water, 0.3 M aque-
ous solution of sodium sulphate and 0.3 M aqueous solution
of sodium sulphate with 20 ppm PEG. Experimental condi-
tions used are shown in Table 1.
For local bubble size distributions, the technique devel-
oped by Greaves and co-workers (e.g., Barigou & Greaves,
1991) was used. It involves withdrawing, by means of a vac-
uum system, a continuous stream of gasliquid dispersion
through a short length calibrated capillary, 0.3 mm in diam-
eter. Gas bubbles are transformed into elongated slugs in-
side the capillary, which are then detected by a pair of LED
phototransistors. The bubble diameter detection limit is de-
termined by capillary diameter, thus approximately 0.3 mm.
Local gas hold-up was measured using the method sug-
gested by Tabera (1990) and Yang and Wang (1991). In
this method, the sampled gasliquid dispersion is withdrawn
through a glass tube of 3.0 mm internal diameter using a
peristaltic pump: both phases are separated and the volume
of each one is measured. A very important condition for
the accuracy of this method was found to be the orientation
of the probe against the liquid ow to approach isokinetic
sampling.
Total gas hold-up was evaluated by measurement of level
change after air sparging.
3. Model
3.1. Modelling grid
The radial coordinate is divided into 3 elements while
the axial coordinate is divided into 10 elements, resulting in
(50. 875) (53. 75) (41. 375)
(
7
3
.
2
5
)
(
6
2
.
7
5
)
(
2
0
)
(
2
0
)
(
6
2
.
7
5
)
(
7
3
.
2
5
)
(
7
3
.
2
5
)
(
6
2
.
7
5
)
(
6
2
.
7
5
)
(
7
3
.
2
5
)
A

Fig. 1. Modelling grid and location of experimental sampling points ().
Distances in mm. A aeration compartment.
30 compartments, as shown in Fig. 1, to coincide with the
experimental sampling grid. Axial symmetry is assumed.
3.2. Liquid ow
The general unaerated liquid ow pattern assumed for
the tank is shown in Fig. 2, where ow rates between
S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496 489
Table 1
Experimental conditions used in individual tests
Liquid phase N (rpm) Q
q
(l}min)
Aqueous Na
2
SO
4
0.3 M 300 10
Aqueous Na
2
SO
4
0.3 M 450 10
Aqueous Na
2
SO
4
0.3 M 450 20
Na
2
SO
4
0.3 M with 20 ppm de PEG 450 10
Water 450 10
Q
p
/2 2 Q
p
Q
p
4 Q
p
2 Q
p
Q
p
/2
Q
p
'/2
2 Q
p
'
4 Q
p
'
Q
p
'
2 Q
p
' Q
p
'/2
Q
p
/2 3/2 Q
p
2 Q
p
Q
p
/2 3/2 Q
p
2 Q
p
Q
p
/2 3/2 Q
p
2 Q
p
Q
p
/2 2 Q
p
Q
p
'/2 3/2 Q '
p
2 Q
p
'
Q
p
'/2 3/2 Q '
p
2 Q
p
'
Q
p
'/2 3/2 Q '
p
2 Q
p
'
Q
p
'/2 2 Q
p
'
3/2 Q
p
3/2 Q
p
'
Fig. 2. Unaerated liquid ow pattern assumed for the tank, with ow rates
between compartments given as a function of the pumping ow rate.
compartments are given as a function the pumping ow rate,
Q
P
. This results in a circulation rate, Q
C
, above and below
the impeller which doubles Q
P
, in agreement with several
authors for single Rushton turbine placed at mid-height in
a tank (Costes & Couderc, 1988; Magni, Costes, Bertrand,
& Couderc, 1988; Ranade & Joshi, 1990; Kusters, Wigers,
& Thoenes, 1991). For these authors, the ratio Q
C
}Q
P
ranges from 1.85 to 2.3. The pumping ow rate is given by
Q
P
= N
q
ND
3
, (1)
where N is the turbine speed, D is the turbine diameter.
The pumping number N
q
is equal to 0.75 015 for single
impellers with a turbine to tank diameter ratio D}1 =
1
3
,
according to a critical review of the literature by Revill
(1982). Mishra and Joshi (1994) found that N
q
decreases
to 0.600.65 for double turbine, even when the distance
between the two turbines is large enough (1.51 to 2.251) to
produce independent circulation patterns and for the power
number of each turbine to be the same as would be in a single
turbine arrangement. A value of N
q
= 0.62 will be used.
Liquid ow splits are based on ow patterns such as given
in Rutherford, Lee, Mahmoudi, and Yanneskis (1996) and
Friberg (1998); for such a course grid as the one adopted in
this work there is little uncertainty about these ow splits.
Under aerated conditions, for moderate aeration rates such
that the gas holdups are relatively small, the liquid ow
circulation still follows the pattern shown in Fig. 2, but the
pumping rate must be corrected for gassed conditions. For
low gas holdups, the ratio between aerated and unaerated
pumping ow rates is equal to the ratio between aerated and
unaerated power (Bakker & van den Akker, 1994):
(Q
P
)
q
(Q
P
)
u
=
P
q
P
. (2)
The ratio P
q
}P is known for the geometry at play (Vas-
concelos, Alves, & Barata, 1995). Liquid velocities, t
L
, at
boundaries between compartments are calculated from the
liquid exchange owrates shown in Fig. 2 and the known
boundary surface areas.
3.3. Gas phase ow
At steady state, the continuity equation for the gas phase
is

(: t
G
) = Q
q
, (3)
where : is the local gas holdup and Q
q
is the local aera-
tion rate per unit volume, which, in the present case, is zero
throughout the vessel, except at the aeration point (Fig. 1).
The gas velocity, t
G
, is given by the sum of the liquid ve-
locity, t
L
, and the slip velocity, t
S
:
t
G
=t
L
+t
S
. (4)
The axial slip velocity is obtained through a force balance
on the bubbles, assumed spherical:
j
L
qJ
b
= C
D
1
2
j
L
t
2
s

4
d
2
b
, (5)
490 S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496
Table 2
Normalised values of the rms turbulent velocity in a standard unaerated tank stirred by a Rushton turbine
a
Reference u

}(ND)
Exit turbine Near the wall Average for the bulk of the tank
Cutter (1966) 0.33 (r = 0.27 T) 0.21 (r = 0.45 T)
Ranade and Joshi (1990) 0.23 (r = 0.45 T) 0.13
Kusters et al. (1990) 0.30 (r = 0.27 T) 0.18 (r = 0.37 T) 0.10
Dyster et al. (1993) 0.34 (r = 0.27 T) 0.25 (r = 0.45 T)
Lee and Yianneskis (1998) 0.34 (r = 0.27 T) 0.19 (r = 0.38 T)
Deglon et al. (1998) 0.34 (r = 0.20 T) 0.24 (r = 0.34 T) 0.14
Values used in this work 0.33 (r = 0.27 T) 0.22 (r = 0.45 T) 0.12
a
Range of geometries and conditions: 0, 1 61(m) 60, 4, H = 1, D = 1}3, 1}3 6C 61}2, 0, 8 6ND(m}s) 63, 8 (r= radial coordinate).
where d
b
is the bubble diameter, J
b
is the bubble volume,
j
L
is the liquid density and q is the acceleration of gravity.
Centripetal force on bubbles is neglected. The drag coe-
cient, C
D
, depends on bubble Reynolds number and can be
calculated using appropriate correlations (Morsi & Alexan-
der, 1972). Since these correlations are valid only in a stag-
nant uid, turbulence is accounted for using a modied bub-
ble Reynolds number, Re

b
, as suggested by Bakker and van
den Akker (1994):
Re

b
=
j
L
t
s
d
b
p

, (6)
where p

is a modied liquid viscosity, dened as the sum


of the liquid viscosity, p
L
, and a term proportional to the
eddy viscosity, p
E
p

= p
L
+ C

p
E
= p
L
+ C

j
L
u

(7)
assuming that the eddy viscosity is proportional to turbulent
velocity, u

, by analogy with gas kinetic theory (McComb,


1990). C

is a model parameter. It was preferred to calculate


p
E
as a function of u

, rather than of the turbulent kinetic


energy and its dissipation rate, as expressed by Bakker and
van den Akker (1994), because there is little agreement in
the literature regarding the value the latter variable (Sahu,
Kumar, Patwardhan, & Joshi, 1999).
3.4. Description of the dispersion process
In this model the dispersed phase is characterized by its
local holdup, :, and average size, d
b
. Bubble average vol-
ume, J
b
, is related to the bubble number density, n
b
, and :
by
n
b
=
:
J
b
. (8)
At steady state, the conservation equation of bubble num-
ber density is

(n
b
t
G
) = n
b
+
Q
q
J
b0
, (9)
where Q
q
is the local gassing rate per unit volume and J
b0
is the average bubble volume at the gas source. The change
in bubble number density due to breakage}coalescence pro-
cesses, n
b
, is calculated in the discretized version of Eq. (9)
as a fraction, F, of the density of bubbles entering the vol-
ume element:
n
b
= F(n
b
t
G
A)
in
}J, (10)
where J is the compartment volume. F is given by
F = C

:d
1
b
u

(11)
an expression inspired in the gas kinetic theory. Allowance
for bubble breakage in the tank could be introduced by a
term in Eq. (11), as in Bakker and van den Akker (1994).
However, since coalescence dominates over breakage ev-
erywhere except in the turbine swept volume, this would un-
necessarily complicate the model, adding an extra parameter
with likely identiability problems. It is deemed preferable
to keep a single lumped parameter (C

in Eq. (11)) which


reects the combined eects of coalescence eciency and
of competing breakage.
3.5. Turbulence data
Local turbulence data, namely the turbulent velocity u

,
is required for calculations of the slip velocity, t
S
, and the
bubble breakage}coalescence parameter, n
b
.
A survey of the literature for unaerated conditions reveals
some measure of general agreement in the normalised val-
ues of the rms turbulent velocity, u

}ND, for the various


locations in a standard tank, whether measured or calcu-
lated using CFD (Table 2). The values of u

}ND used in
the simulation, also shown in the table, were average values
taken from the authors quoted. A single average value was
used for the bulk of the tank, since u

}ND is found not to


vary a great deal for each author, except along the turbine
discharge stream.
The unaerated values of u

were corrected for aeration


assuming that (Bakker & van den Akker, 1994):
c
q
= c
u
P
q
P
+ c
b
,
c
u
P
q
P
, (12)
S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496 491
hence
u

q
u

P
q
P

1}3
(13)
where c
q
, c
u
and c
b
are the turbulent energy dissipation val-
ues, respectively, for aerated liquid, unaerated liquid and the
contribution due to bubbles, the latter assumed negligible.
u

q
and u

u
are the aerated and unaerated turbulent velocities.
4. Results and discussion
4.1. Experimental results
Fig. 3 shows both experimental and simulated average
bubble sizes at several locations in the tank, for the condi-
tions presented in Table 1. Diameters are represented as a
function of a spatial coordinate which follows the longest
liquid ow circulation path from each turbine back to the
same turbine, as illustrated in the gure. Diameters are vol-
ume averaged, d
43
, which was found to be the simplest def-
inition compatible with the model equations. As the surface
mean diameter is more useful for mass transfer calculations,
the relationship between d
43
and d
32
was examined and ap-
pears in Fig. 4, which shows a simple relationship between
the two diameters for all of the operating conditions studied.
Fig. 5 attempts a comparison of d
32
bubble size diam-
eters obtained by dierent authors using dierent meth-
ods, including the suction probe (Barigou, 1987) and
video}photographic methods (Martin, 1995; Machon, Pacek,
& Nienow, 1997; Bouai & Roustan, 1998). Since the op-
tical methods only measure bubble diameters near the tank
wall, the sampling location considered in the suction probe
experiments was also chosen to be near the wall. From
this gure we may draw the following conclusions: (i) all
experimental diameters lie within 35% maximum error
from the diameter predicted by the regression; (ii) scatter
within each method is of the same order of magnitude as
the overall scatter; (iii) results of both methods overlap,
although Barigou (1987) tended to obtain larger diameters.
Fig. 6 shows both experimental and simulated local gas
holdup at several locations in the tank, for the same sets of
conditions. Integration of the experimental local gas holdup
for the tank was consistent with the measured overall gas
holdup (Table 3).
From Fig. 3, it can be seen that the general trend is for
bubble size to increase along the circulation path, which
indicates that bubble coalescence prevails over breakage.
Coalescence is usually observed to be most intense in the
turbines discharge streams, which may be explained by a
larger collision frequency due to higher turbulence. Another
point of intense coalescence is in circulation loop 3, below
the upper turbine, near the wall, where upcoming bubbles
from the lower part of the tank meet the downcoming half of
the upper turbine discharge stream. This is common to both
so-called non-coalescing media (sulphate solution, sulphate
(f) (e)
(c) (a) (b)
(d)
0
1
2
3
4
0.0 0.1 0.2 0.3 0.4
x (m)
d
b

(
m
m
)
0
1
2
3
4
d
b

(
m
m
)
0
1
2
3
4
d
b

(
m
m
)
0
1
2
3
4
0.0 0.1 0.2 0.3 0.4
x (m)
0
1
2
3
4
0.0 0.1 0.2 0.3 0.4
x (m)
0
1
2
3
4
0
1
2
3
4
0
1
2
3
4
0
1
2
3
4
0.0 0.1 0.2 0.3 0.4
x (m)
d
b

(
m
m
)
0
1
2
3
4
d
b

(
m
m
)
0
1
2
3
4
d
b

(
m
m
)
0
1
2
3
4
d
b

(
m
m
)
Loop 4 Loop 4
Loop 3 Loop 3
Loop 2 Loop 2
Loop 1 Loop 1
Loop 4
Loop 3
Loop 2
Loop 1 Loop 1
Loop 4
Loop 3
Loop 2
0
1
2
3
4
0
1
2
3
4
0
1
2
3
4
Loop 1
Loop 4
Loop 3
Loop 2
0
1
2
3
4
d
b

(
m
m
)
0
1
2
3
4
5
0
1
2
3
4
5
0
1
2
3
4
5
0
1
2
3
4
5
0.0 0.1 0.2 0.3 0.4
x (m)
Loop 1
Loop 2
Loop 3
Loop 4
x
x
x
x
Fig. 3. Experimental () and simulated () average bubble diam-
eters as a function of position along liquid circulation loops: (a)
Identication of liquid circulation loops. (b) Aqueous Na
2
SO
4
0.3 M
solution, N = 300 rpm, Q
q
= 10 l}min. (c) Aqueous Na
2
SO
4
0.3 M
solution, N = 450 rpm, Q
q
= 10 l}min. (d) Aqueous Na
2
SO
4
0.3 M
solution, N = 450 rpm, Q
q
= 20 l}min. (e) Aqueous Na
2
SO
4
0.3 M
with 20 ppm de PEG, N = 450 rpm, Q
q
= 10 l}min. (f) Tap water,
N = 450 rpm, Q
q
= 10 l}min.
solution+PEG) and to the coalescing medium (tap water).
A decrease in bubble size along the liquid circulation paths
only occurs at the bottom of the tank and above the two
turbines, particularly at low stirrer speeds. When this occurs,
the gas holdup also decreases. The explanation is that, at
those locations only small bubbles, with small slip velocities,
tend to follow the downwards liquid ow, particularly at
low circulation velocities.
The eect of stirrer speed on gas dispersion can be as-
sessed by comparing Figs. 3(a) with (b) (bubble sizes)
and 6(a) with (b) (gas holdups). At higher stirring speed,
492 S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496
0.00
0.50
1.00
1.50
2.00
2.50
3.00
3.50
4.00
4.50
0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00 4.50 5.00
d43 (mm)
d
3
2

(
m
m
)
Fig. 4. Relationship between d
43
and d
32
for all operating conditions under study.
0.40
0.60
0.80
1.00
1.20
1.40
1.60
1.80
2.00
2 4 6 8 9
ln(Pg/V)
l
n
(
d
3
2
)
Barigou and
Greaves, 1992
Martin, 1995
Machon et al,
1997
Bouaifi and
Roustan, 1998
This work
Correlation
7 5 3
Fig. 5. Sauter mean bubble diameter near tank wall as a function of power input per unit tank volume, with water as the liquid medium.
smaller bubble sizes are found in the turbines discharge
stream, due to higher energy dissipation. In the bulk of the
tank, after the eect of coalescence, the dierence in bubble
diameter between the two conditions is not very signicant,
while at the bottom of the tank and immediately above the
turbines bubble sizes are actually larger for higher turbine
speeds, when larger liquid circulation velocities manage to
re-circulate larger bubbles.
The eect of aeration on gas dispersion can be assessed by
comparing Figs. 3(b) with (c) and 6(b) with (c). Larger aer-
ation leads to slightly larger bubbles, due to the part played
by a larger gas holdup on coalescence.
The dierence between a so-called non-coalescing
medium, sulphate solution, and a coalescing medium, tap
water, can be examined by comparing Figs. 3(a) with (e)
(bubble sizes) and 6(a) with (e) (gas holdups). As expected
bubble sizes are larger with tap water, since coalescence
is more ecient, and, as a consequence, gas holdups also
tend to be smaller.
The eect of adding a surfactant (PEG) to an already
non-coalescing solution (Fig. 3(d) vs. (b)) leads to further
repression of coalescence and to slightly lower bubble size
diameters.
4.2. Simulation vs. experiment
The proposed model calculates local average bubble sizes
and local gas holdups throughout the tank, with experi-
mental values of average bubble size produced by each of
the turbines as an input. Two parameters are adjusted: C

(Eq. (7)), which inuences the gasliquid slip velocity, and


hence the gas holdup, and C

(Eq. (11)), related mainly


to coalescence eciency and bubble size. The eect of the
two parameters is not independent, however, since bubble
size aects gas holdup and vice-versa.
Table 4 presents optimised parameters for the three liq-
uid media under study. C

, which expresses coalescence


eciency, is greatest for water and smallest for the sul-
phate+PEG aqueous solution, as expected. The value of
C

reaches an upper limit when the calculation of the frac-


tion F of bubbles that disappear by coalescence (Eq. (10))
S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496 493
Fig. 6. Experimental (bold) and simulated (within brackets) local gas holdups: (a) Aqueous Na
2
SO
4
0.3 M solution, N = 300 rpm, Q
q
= 10 l}min.
(b) Aqueous Na
2
SO
4
0.3 M solution, N = 450 rpm, Q
q
= 10 l}min. (c) Aqueous Na
2
SO
4
0.3 M solution, N = 450 rpm, Q
q
= 20 l}min. (d) Aqueous
Na
2
SO
4
0.3 M with 20 ppm de PEG, N = 450 rpm, Q
q
= 10 l}min. (e) Tap water, N = 450 rpm, Q
q
= 10 l}min.
becomes grater than one, which makes no physical sense.
This limit was reached for water in the upper turbine dis-
charge stream.
C

is the only parameter which acts on the gasliquid slip


velocity. It was introduced as a correction due to turbulence,
but its eect is the same as a decrease in drag coecient,
494 S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496
Table 3
Experimental and simulated global gas holdup
Liquid medium Operating conditions Experimental Total gas holdup Total gas holdup
N, Q
q
(rpm, l}min) total gas holdup integrated from integrated from
experimental simulated local
local holdups holdups
Aqueous NaSO
4
solution 300, 10 2.2 2.2 2.8
Aqueous NaSO
4
solution 450, 10 4.4 4.4 3.8
Aqueous NaSO
4
solution 450, 20 5.0 5.4 5.0
Aqueous NaSO
4
sol.+PEG 450, 10 5.2 5.2 5.3
Tap water 450, 10 2.5 n.a. 3.4
Table 4
Adjusted model parameters
Liquid medium
Aqueous 0.3 M Na
2
SO
4
Tap water 0.3 M Na
2
SO
4
with 20 ppm de PEG
C

0.00007 0.00022 0.00010


C

0.030 0.0680 0.021


C
D
. If slip velocities are calculated without this correction
(C

= 0), considerably low modelling estimates of the gas


holdups are obtained. C

increases when PEG is added to an


aqueous sulphate solution and this is likely to be a slowing
down eect due to accumulation of tensioactive material at
the interphase rather than a turbulence eect, which would
be more appropriately described by an increase in the drag
coecient C
D
, but that would require an additional parame-
ter. C

also increases for coalescing conditions (tap water)


which might be explained since the eect of turbulence on
bubble rise velocity increases with bubble diameter (Bakker
& van den Akker, 1994).
Comparison between simulated and experimental results
are presented in Fig. 3 (average bubble sizes) and in Fig. 6
(local gas hold-ups). Disagreement between simulation and
experiment is most signicant at points where bubble size
decreases along the liquid circulation paths, i.e., most clearly
at the bottom of the tank (second half of loop 1). The expla-
nation is that, at those locations, only small bubbles, with
small slip velocities, tend to follow the downwards liquid
ow, particularly at low circulation velocities. Since the
model uses a single average bubble size, it cannot discrimi-
nate between large and small bubbles and hence cannot sim-
ulate this eect. To be able to simulate the dierence in be-
haviour between bubbles of dierent diameters, the model
would have to include population balances on classes of
bubbles.
The other location where considerable disagreement is
found between simulation and experiment is where upcom-
ing bubbles from the lower half of the tank meet the down-
coming stream from the upper turbine ( 0.20 m along loop
3 in Fig. 3). An intuitive explanation is that bubbles travel-
ling in opposite directions have a larger chance of coalesc-
ing, something which is cannot be described in a model as
simple as the one proposed. A similar, but less signicant
disagreement also happens in loop 1.
Fig. 7 presents a simulation of the gas ow in a typical
situation in the tank. This qualitatively agrees with what is
expected (e.g., Manikowski, Bodemeier, L ubbert, Bujalski,
& Nienow, 1994).
5. Conclusions
(1) For all liquid media under study, bubble coalescence
prevails over breakage throughout the tank. Where av-
erage bubble size decreases this can be attributed to the
dierence in slip velocity between dierent sized bub-
bles.
(2) It is found that bubble size increases most quickly in
the turbine discharge stream, and immediately below
the point where this meets the tank wall.
(3) Modelling of gas dispersion throughout the tank may be
achieved with a simple compartment model that takes
into account bubble coalescence, but neglects breakage.
The model predicts spatial distribution of gas holdup
and of average bubble size, with average bubble size at
turbine as an input.
(4) Reasonable agreement between experiment and simu-
lation is achieved with optimisation of two parameters,
one aecting mainly the slip velocity, the other related
mainly to the bubble coalescence}breakage balance.
(5) Dierent sets of parameters are required for each of the
three liquid systems under study. These parameters are
independent of stirring}aeration conditions.
(6) The model fails to simulate the smaller average bub-
ble diameters at the bottom of the tank, which are due
to dierent behaviour of bubbles of dierent diameters.
S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496 495
0.56 2.23
1.15 4.67
2.44 0.59
0.38 1.50
3.15 1.43
1.20 0.19
0. 42 1. 26
2.68
0. 42 1. 26
0.73 2.26
1. 99
0.73
1. 99
0. 24 0. 71
1.95
0. 24 0. 71
1.95
1.19 1.01
1. 20
0.19
1. 20
2.26
1.01
0.14 0.41 0.45
0.14
2.68
0.45 0.41
1.00
Fig. 7. Simulation of gas ow pattern in the tank in a typical situation
(450 rpm, 20 l}min aeration rate, non-coalescing medium). Numbers rep-
resent the gas ow rate across each boundary, normalised by the aeration
ow rate.
To simulate this, the model would have to include pop-
ulation balances on classes of bubbles.
Notation
C turbine clearance from tank bottom, m
C

model parameter
C

model parameter
d
b
bubble diameter, m
d
32
surface based mean bubble diameter, d
32
=

n
i=1
d
3
b
i
}

n
i=1
d
2
b
i
, m
d
43
volume based mean bubble diameter, d
43
=

n
i=1
d
4
b
i
}

n
i=1
d
3
b
i
, m
D turbine diameter, m
F fraction of bubble number density entering the vol-
ume element
q acceleration of gravity, m s
2
H liquid height, m
n
b
bubble number density, m
3
n
b
change in bubble number density due to
breakage}coalescence, m
3
s
1
N turbine speed, s
1
or min
1
N
q
pumping number
P turbine power, W
P
q
aerated turbine power, W
Q
C
circulation rate, m
3
s
1
Q
q
local gassing rate per unit volume, s
1
Q
P
turbine pumping rate, m
3
s
1
r radial distance from centre of symmetry, m
Re

b
modied bubble Reynolds number
1 tank diameter, m
u

turbulent velocity, m s
1
u

q
turbulent velocity under aerated conditions, m s
1
u

u
turbulent velocity under unaerated conditions,
m s
1
J compartment volume, m
3
J
b
average bubble volume, m
3
J
b0
average bubble volume at the gas source, m
3
t
G
gas velocity, m s
1
t
L
liquid velocity, m s
1
t
S
gasliquid slip velocity, m s
1
Greek letters
: gas holdup
c
q
total turbulent energy dissipation rate under aera-
tion conditions, m
2
s
3
c
b
turbulent energy dissipation rate due to bubbles,
m
2
s
3
c
u
unaerated turbulent energy dissipation rate, m
2
s
3
p

modied liquid viscosity, Pa s


p
E
eddy viscosity, Pa s
p
L
liquid viscosity, Pa s
j
L
liquid density, kg m
3
Acknowledgements
The authors wish to thank Dr. A.J. Serralheiro for help
with the suction probe. Financial support by PRAXIS XXI,
Project No. 2}2.1}BIO}1061}95 and the research grant
awarded to C.I. Maia (PRAXIS XXI 4}4.1}BD}2935}96)
are also gratefully acknowledged.
496 S. S. Alves et al. / Chemical Engineering Science 57 (2002) 487496
References
Bakker, A., & van den Akker, H. E. A. (1994). A computational model for
the gasliquid ow in stirred reactors. Transactions of the Institution
of Chemical Engineers, Part A, 72, 594606.
Barigou, M. (1987). Bubble size, gas holdup and interfacial area
distributions in mechanically agitated gasliquid reactors. PhD thesis,
The University of Bath, United Kingdom.
Barigou, M., & Greaves, M. (1991). A capillary suction probe for bubble
size measurement. Measurements in Science and Technology, 2, 318
326.
Barigou, M., & Greaves, M. (1992). Bubble-size distributions in a
mechanically agitated gasliquid contactor. Chemical Engineering
Science, 47(8), 20092025.
Bouai, M., & Roustan, M. (1998). Bubble size and mass transfer
coecients in dual-impeller agitated reactors. Canadian Journal of
Chemical Engineering, 76, 390397.
Costes, J., & Couderc, J. P. (1988). Study by LDA of the turbulent ow
induced by a Rushton turbine in a stirred tank I. Mean ow and
turbulence. Chemical Engineering Science, 43(10), 27512764.
Cutter, L. A. (1966). Flow and turbulence in a stirred tank. AIChE
Journal, 12(1), 3545.
Deglon, D. A., OConnor, C. T., & Pandit, A. B. (1998). Ecacy of
a spinning disk as a bubble break-up device. Chemical Engineering
Science, 53(1), 5970.
Djebbar, R., Roustan, M., & Line, A. (1996). Numerical computation
of turbulent gasliquid dispersion in mechanically agitated vessels.
Transactions of the Institution of Chemical Engineers, Part A, 74,
492498.
Dyster, K. N., Koutsakos, E., Jaworsky, Z., & Nienow, A. W. (1993). An
LDA study of the radial discharge velocities generated by a Rushton
turbine. Transactions of the Institution of Chemical Engineers, Part
A, 71, 1123.
Friberg, P. C. (1998). Three-dimensional modelling and simulation of
gas}liquid ow processes in bioreactors. Dr.Ing. thesis, Telemark
Institute of Technology, Norway.
Gosman, A. D., Lekakou, C., Politis, S., Issa, R., & Looney, M. K.
(1992). Multidimensional modeling of turbulent two-phase ows in
stirred vessels. AIChE Journal, 38(12), 19461956.
Kusters, K. A., Wigers, J. G., Baaten, J. P. W. M. M., & Thoenes, D.
(1990). Laser Doppler measurements of the local energy dissipation
rate in stirred tanks to establish scaling rules. Proceedings of the
International Symposium on Applied Laser Techniques and Fluid
Mechanics (pp. 17). Lisbon, Portugal, 23.3.
Kusters, K. A., Wigers, J. G., & Thoenes, D. (1991). Numerical particle
tracking in a turbine agitated vessel. Seventh European Conference
on Mixing (pp. 429441). Brugges, Belgique.
Lane, G. L., Scwartz, M. P., & Evans, G. M. (2000). Modelling of the
interaction between gas and liquid in stirred vessels. In H.E.A. van
den Akker & J.J. Derksen (Eds.), Proceedings of the Tenth European
Conference on Mixing (pp. 197204). Amsterdam: Elsevier.
Lee, K. C., & Yianneskis, M. (1998). Turbulence properties of the impeller
stream of a Rushton turbine. AIChE Journal, 44(1), 13.
Machon, V., Pacek, A. W., & Nienow, A. W. (1997). Bubble sizes
in electrolyte and alcohol solutions in a turbulent stirred vessel.
Transactions of the Institution of Chemical Engineers, Part A, 75,
339348.
Magni, F., Costes, J., Bertrand, J., & Couderc, J. P. (1988). Study by LDA
of the ow induced by a Rushton turbine in a stirred tank: Inuences
of the geometry of the tank bottom and the position of the turbine.
Sixth European Conference on Mixing (pp. 714). Pavia, Italia.
Manikowski, M., Bodemeier, S., L ubbert, A., Bujalski, W., & Nienow, A.
W. (1994). Measurement of gas and liquid ows in stirred tank reactors
with multiple agitators. Canadian Journal of Chemical Engineers, 94,
769781.
Mann, R. (1986). Gasliquid stirred vessel mixers: Towards a unied
theory based on networks of zones. Chemical Engineering Research
Design, 64, 2334.
Mann, R., & Hackett, L. A. (1988). Fundamentals of gasliquid mixing
in a stirred vessel: An analysis using networks of backmixed zones.
Proceedings of the Sixth European Conference on Mixing (pp. 321
328). BHRA, Pavia.
Martin, T. (1995). Gas dispersion with radial and hydrofoil impellers
in uids with dierent coalescence characteristics. Ph.D. thesis,
Birmingham University, United Kingdom.
McComb, W. D. (1990). The physics of turbulence. Oxford: Clarendon
Press.
Mishra, V. P., & Joshi, J. B. (1994). Flow generated by a disc turbine: Part
IV: Multiple impellers. Transactions of the Institution of Chemical
Engineers, Part A, 72, 657668.
Morsi, S. A., & Alexander, A. J. (1972). An investigation of particle
trajectories in two-phase ow systems. Journal of Fluid Mechanics,
Part 2, 5, 193208.
Morud, K. E., & Hjertager, B. H. (1996). LDA measurements and CFD
modelling of gasliquid ow in a stirred vessel. Chemica Engineering
Science, 51, 233249.
Patterson, G. K. (1991). Measurements and modelling of ow in
gas sparged agitated vessels. Proceedings of the Seventh European
Conference on Mixing (pp. 209215). Brugges, Belgique.
Ranade, V. V., & Joshi, J. B. (1990). Flow generated by a disc turbine:
Part I. Experimental. Transactions of the Institution of Chemical
Engineers, Part A, 68, 1933.
Revill, B. K. (1982). Pumping capacity of disk turbine agitators
a literature review. Fourth European Conference on Mixing (pp. 11
24). Noordwijkerhout, Holland.
Rutherford, K., Lee, K. C., Mahmoudi, S. M. S., & Yanneskis, M. (1996).
Hydrodynamic characteristics of dual Rushton turbine stirred vessels.
AIChE Journal, 42(2), 332334.
Sahu, A. K., Kumar, P., Patwardhan, A. W., & Joshi, J. B. (1999). CFD
modelling and mixing in stirred tanks. Chemical Engineering Science,
54, 22852293.
Sch afer, M., W achter, P., & Durst, F. (2000). Experimental investigation
of local bubble size distribution in stirred vessels using Phase Doppler
Anemometry. In H.E.A. van den Akker & J.J. Derksen (Eds.),
Proceedings of the Tenth European Conference on Mixing (pp. 205
212). Amsterdam: Elsevier.
Tabera, J. (1990). Local gas hold-up measurement in stirred fermenters.
I. Description of measuring apparatus and screening of variables.
Biotechnology and Techniques, 4(5), 299304.
Takahashi, K., & Nienow, A. W. (1993). Bubble sizes and coalescence
rates in an aerated vessel agitated by a Rushton turbine. Journal of
Chemical Engineering Japan, 26(5), 536542.
Tsouris, C., & Tavlarides, L. L. (1994). Breakage and coalescence models
for drops in turbulent dispersions. AIChE Journal, 40, 396406.
Vasconcelos, J. M. T., Alves, S. S., & Barata, J. M. (1995). Mixing in gas
liquid contactors agitated by multiple turbines. Chemical Engineering
Science, 50(14), 23432354.
Venneker, B. C. H. (1999). Turbulent ow and gas dispersion in stirred
vessels with pseudoplastic liquids. Ph.D. thesis, Technische Universiteit
Delft.
Yang, J., & Wang, N. S. (1991). Local gas holdup measurement in aerated
agitated bioreactors. Biotechnology and Techniques, 5(5), 349354.

You might also like