Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Chemical Engineering and Processing 43 (2004) 1511–1517

Gas–liquid mass transfer coefficient in stirred tanks interpreted through


models of idealized eddy structure of turbulence in the bubble vicinity
V. Linek∗ , M. Kordač, M. Fujasová, T. Moucha
Department of Chemical Engineering, Prague Institute of Chemical Technology, Technická 5, CZ-166 28 Prague 6, Czech Republic

Received 9 September 2003; received in revised form 26 September 2003; accepted 10 February 2004
Available online 24 April 2004

Abstract

Experimental data on the average mass transfer liquid film coefficient (kL ) in an aerated tank stirred by two Rushton turbines on common
shaft are presented. Liquid media used were distilled water and 0.5 M sodium sulphate solution. Volumetric mass transfer coefficient (kL a)
was measured by the dynamic pressure method with pure oxygen absorption. Specific interfacial area a was taken from Alves et al. [Chem.
Eng. Proc., in press] who measured data on local gas hold-up and local average bubble diameter in the same apparatus and batches. Values
of kL are quantitatively interpreted in terms of correlations based on idealized eddy structures of turbulence in the bubble vicinity, namely by
“eddy” model by Lamont and Scott [AIChE J. 16 (1970) 513] in the form of kL = 0.523(eν/ρ)0.25 (D/ν)1/2 , which fits the data with the mean
deviation of 4.7%. It is shown that the decisive quantity to correlate kL in the stirred tank is power dissipated in the liquid phase rather than
the bubble diameter and the slip velocity as assumed by Alves et al.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Mass transfer coefficient; Bubble; Dynamic pressure method

1. Introduction surface due to the contamination), kL is (according to Alves


et al. [3]) given by the equation proposed by Frössling [1]
Mass transfer from swarm of bubbles into turbulent liq- 
vsl 2/3 −1/6
uid controls the rate of many chemical and biochemical pro- kL = c1 D ν (1)
cesses; hence many papers were devoted to the problem. In d
spite of this, the mechanism of the transfer remains uncer- or (according to Calderbank and Moo-Young [2]) by the
tain. Various widely different, but yet partially successful, relation
approaches exist. They predict different, often contradictory,    
influence of operational conditions on mass transfer coeffi- ρνg 1/3 D 2/3
kL = 0.31 (2)
cient kL . Two of the most frequently used concepts assume ρ ν
the mechanism of mass transport in the liquid phase is due
to a renewal of the liquid at the bubble surface. For the bubbles that have completely mobile surface (large
The first group of the models (“slip velocity” models) as- bubbles in pure liquid or fresh bubble), Alves et al. [3] have
sumes a gross mean flow of fluid relative to the bubble (slip proposed for calculation of kL the equation, which follows
velocity) and a bubble surface rigidity control of this re- from the well-known penetration model with exposition time
newal. The surface rigidity depends on bubble size (accord- equal to d/vsl
ing to Calderbank and Moo-Young [2]) or on a degree of the 
vsl 1/2
bubble surface contamination by surface-active impurities kL = 1.13 D (3)
(according to Alves et al. [3]). For the rigid bubbles (small d
bubbles d < 1 mm or bubbles with completely immobile and Calderbank and Moo-Young [2] have proposed the re-
lation
   
∗ Corresponding author. Tel.: +42-2-24353298; fax: +42-2-3117335. ρνg 1/3 D 1/2
kL = 0.42 (4)
E-mail address: linekv@vscht.cz (V. Linek). ρ ν

0255-2701/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.cep.2004.02.009
1512 V. Linek et al. / Chemical Engineering and Processing 43 (2004) 1511–1517

Gas–liquid slip velocities vsl are assumed to be close to The overall volumetric mass transfer coefficients kL a were
single bubble terminal velocities, vt , in still water on which measured by the peroxide decomposition steady-state tech-
a correction for turbulence is introduced [3]. nique with air as a gas phase. It was shown [11], however,
The second group of models (“eddy” models) assumes that the steady-state methods with air absorption/desorption
that this renewal is due to the small scale eddies of the gives substantially lower kL a values at high mixing intensi-
turbulent field rather than to any gross mean flow of fluid ties as compared to pure oxygen absorption.
relative to the bubble. The models are based on idealized Aim of this paper is to show that if the overall volu-
eddy structures of turbulence in the bubble vicinity. Lamont metric mass transfer coefficients kL a are measured by the
and Scott [4] assumed that the mass transfer in the liquid non-problematic pressure dynamic method [12] with pure
is affected preponderantly by the small scales of turbulent oxygen absorption in the same apparatus as Alves, the rel-
motion (the scales extend from the smallest viscous motions evant kL data are well interpreted in terms of the “eddy”
into the inertial motions). In any case, these motions are model by Lamont and Scott [4]. The paper should support
much smaller in scale than the gas bubbles and they “attack” opinion that the decisive quantity to correlate kL in stirred
the surface of bubbles in the same way from any direction. tanks is power dissipated in the liquid phase rather than the
As a result, the size of the gas bubble is not a very critical bubble diameter and their slip velocity.
parameter for the estimation of kL . They deduced [4] the
following relation for kL
 1/4  1/2 2. Experimental
eν D
kL = c5 (5)
ρ ν The experiments were performed in a fully baffled (four
Different authors predict different values of the constant c5 : baffles, 1/10 of the vessel diameter width) cylindrical ves-
0.301 [5], 0.4 [4], 0.592 [6], 1.13 [7]. sel of i.d. 0.29 m equipped with two standard Rushton tur-
These two groups of models are contradictory. The “eddy” bines. The impeller diameter was equal to 1/3 of the tank
models predict an increase while the “slip velocity” mod- diameter as well as the bottom impeller clearance and the
els a decrease of kL with increasing turbulence intensity, inter-impeller distance was equal to the tank diameter. An
i.e. with increasing power dissipated in the liquid. This is aeration gas was introduced 2 cm under the bottom im-
due to the fact, that kL from rigid bubbles is lower than peller through a ring distributor with diameter 4 cm and with
from the ones with a mobile surface by a factor around 5. nine orifices (1 mm in diameter). The membrane-covered
The small bubbles, which either have lower kL (according polarographic type oxygen probes were located in the liq-
to Calderbank’ variant of “slip velocity” models) or become uid batch at the impeller’s planes. A sketch of the vessel
rigid more quickly (according to Alves’ variant) are gener- indicating locations of oxygen probes is in Fig. 1. The ex-
ated in larger amount just at higher turbulence intensities. As periments were performed in distilled water and in 0.5 M
a result this leads to the decrease of average mass transfer Na2 SO4 aqueous solution, which represents non-viscous co-
coefficient in dispersion with increasing power dissipated in alescent and non-coalescent batch, respectively. The vessel
the liquid. Literature data on kL in stirred tanks supporting was filled with the liquid to the height of two times the ves-
both these models can be found. Contradictory results on the sel diameter. Superficial velocities of aeration gas (air in the
mass transfer coefficients kL determined by various authors measurements of power and gas hold-up, pure oxygen in
can be partially explained by errors in kL a values used in the measurements of volumetric mass transfer coefficient)
their evaluation from the relation kL = kL a/a. It was shown were 2.12, 4.24 or 8.48 mm s−1 . The experiments were per-
[8], that the kL a values measured by dynamic methods based formed in the impeller speed range from 4.17 to 14.17 s−1
on the inlet gas exchange at high mixing intensities are un- at 20 ◦ C.
derestimated even by multiples due to an incorrect flow pat- The impeller power was measured by a strain gauge on the
tern of gas phase used in the model for the kL a evaluation. It impeller shaft. The gas hold-up was determined by visual in-
was demonstrated [8] on the data of different authors [9,10] spection from the difference between the batch level height
that the reported decrease of kL with increasing power input with and without aeration. The volumetric mass transfer
dissipation in the liquid does not reflect real phenomena but coefficient was measured by the dynamic pressure method
is only the result of error in the measuring methods of kL a (DPM) described earlier [12]. Physical correctness of the
at high mixing intensities. method has been verified by the agreement between kL a
Alves et al. [3] presented data on kL in a double Rushton measured by absorption of air and pure oxygen: physically
turbine stirred tank. Their experimental kL falls between the consistent kL a data should not depend on the driving force
limits defined by Eqs. (1) and (3), which differ by a factor of the absorption, i.e., the kL a values should be the same no
around 5. The data are quantitatively interpreted in terms of matter what oxygen concentration in aeration gas is used.
bubble contamination kinetics, using a stagnant cap model, The principle of the DPM lies in the fast change of the to-
according to which bubbles suddenly change from a mobile tal pressure in the vessel and the consequent recording the
interface to a rigid condition when surface tension gradients, oxygen probes response in liquid phase in both stages. In
caused by surfactant accumulation, balance out shear stress. evaluation of the kL a the dynamic of the oxygen probes was
V. Linek et al. / Chemical Engineering and Processing 43 (2004) 1511–1517 1513

Table 1
Parameters C1 , C2 and C3 in correlations (6), (8) and (9) obtained by the
experimental data regression
Batch C1 C2 C3 S.D. (%)

Specific power dissipated by agitators, correlation (6)


Water 0.155 3.141 −0.365 4.6
0.5 M Na2 SO4 0.208 2.932 −0.393 5.2
Gas hold-up, correlation (8)
Water 0.2346 0.3091 0.7347 5.6
0.5 M Na2 SO4 0.0505 0.4903 0.5788 9.3
Volumetric mass transfer coefficient, correlation (9)
Water 0.0108 0.699 0.581 9.6
0.5 M Na2 SO4 0.978E−3 1.233 0.55 10

3.2. Gas hold-up

The gas hold-up as a function of the specific power dissi-


pated by agitators in the liquid and superficial gas velocity
was fitted by empirical relation
ε = C1 (eagit )C2 vC
s
3 (8)

The parameters C1 , C2 and C3 obtained by the experimental


data regression are given for both batches used in Table 1.

3.3. Volumetric mass transfer coefficient

The overall volumetric mass transfer coefficients were


Fig. 1. Scheme of the apparatus. expressed as a function of the total specific power dissipated
in the liquid and gas superficial velocity as follows (the usage
of e gave closer correlation than the usage of an agitator
taken into account [12] and the kL a value for each stage power eagit )
of the vessel was evaluated. Fast response oxygen probes
with time constants above 0.8 s−1 were used (polypropylene kL a = C1 eC2 vC
s
3 (9)
membrane, 10 ␮m thick). An average kL a value was calcu-
lated as an arithmetic mean of both stage values. The arith- C1 , C2 and C3 values are given for both batches in Table 1.
metic mean was used as it was shown [13] that its value is Experimental data are fitted with the correlations very well
not affected by axial dispersion of both phases. as is demonstrated in Fig. 2. Standard deviation of the data
is 10%.
The overall volumetric mass transfer coefficients were
measured by Alves et al. [3] using the peroxide decomposi-
3. Results
tion steady-state technique with manganese dioxide as the
catalyst in the same apparatus as it is used in this work.
3.1. Power dissipated
The only difference is higher bottom impeller clearance
(0.146 m instead of 0.097 m). Such small difference has
The specific power dissipated by agitators in the liquid
inexpressive effect on mass transfer characteristics of dis-
was well fitted by the following relations
persion. Operational conditions used in this work (impeller
eagit = C1 N C2 vC
s
3 (6) speed 4.17–14.17 s−1 , gas superficial velocity 0.00212,
0.00424 and 0.00848 m s−1 ) fully overlap the conditions
The parameters C1 , C2 and C3 obtained by the experi- under which Alves et al. [3] performed their experiments
mental data regression are given for both batches used in (impeller speed 4.2–10 s−1 , gas superficial velocity 0.00249
Table 1. These empirical correlations fit the experimental and 0.00497 m s−1 ). The liquid media used by Alves were
data with high accuracy (standard deviations lower than 5%, water and 0.3 M aqueous solution of Na2 SO4 . In this work
see Table 1). The total specific power dissipated in the liquid 0.5 M solution of Na2 SO4 is used. It is well known that
by agitator and by rising bubbles was calculated from the addition of an electrolyte to water reduces the rate of
bubble coalescence resulting in an increase in kL a up to a
e = eagit + ρgvs (7) limiting concentration beyond which a further increase in
1514 V. Linek et al. / Chemical Engineering and Processing 43 (2004) 1511–1517

evaluation of the characteristics at operational conditions at


which Alves et al. [3] performed their measurement.
Table 2 brings together average tank data on operational
conditions, gas hold-up, bubble diameter, specific area and
volumetric mass transfer coefficients measured by Alves and
those calculated from correlations (6) to (9). There is very
good agreement in gas hold-up values but there is strong
disagreement in volumetric mass transfer coefficients. kL a
values calculated from correlation (9) are up to six times
higher than Alves’ kL a data. The reliability of our data (9)
on kL a measured in two-impeller configuration supports the
agreement with those determined [11] in one-impeller con-
figuration by dynamic pressure method for air and pure oxy-
gen absorption and by steady-state sulphite feeding method
for pure oxygen absorption. The one-impeller data were cor-
related [11] by the equation
kL a = 3.11 × 10−4 e1.24 v0.4
s (10)

which gives up to 20% higher data compared to Eq. (9).


Possible reasons for too low kL a values measured by
Fig. 2. Comparison of kL a data with the correlation (9).
Alves et al. [3] in comparison with our data are 2:

electrolyte concentration no longer has an effect on kL a. (i) It was shown [11] that steady-state sulphite feeding
The concentrations 0.5 and 0.3 M are higher than the limit- method for air absorption gives kL a values substan-
ing value for Na2 SO4 . This follows from Fig. 3, where kL a tially lower (two times) than those for pure oxygen ab-
data measured [14] by dynamic start-up method 0 → O2 sorption in the region of kL a > 0.1 s−1 . This is caused
(the batch is degassed at the beginning of experiment) in by the fact that the residence times of individual bub-
one-impeller configuration in sodium sulphate solutions of bles differ considerably. The small bubbles, with lower
different concentrations are compared: kL a data produced tendency to coalesce, stay in the reactor significantly
in stirred vessel in 0.5 and 0.2 M Na2 SO4 do not differ. longer than the mean residence time and have higher
All these facts support conclusion that the correlations surface to volume ratio. During absorption of a diluted
presented in this paper for e, ε and kL a can be used for gas, a part of bubbles is therefore formed in the ves-
sel, the content of which is in equilibrium with the sur-
rounding liquid and which is ineffective with respect
to absorption. As a result, kL a data measured by di-
luted gas are lower because they do not include these
bubbles. In other words, the interfacial area a in kL a
values measured by diluted gas (air) is not a geometric
surface area of all the bubbles in dispersion but only
its part that is effective in the mass transfer. The geo-
metric surface area of all bubbles is active only in the
kL a data measurement by some steady-state methods
with pure oxygen absorption or by the pressure dy-
namic method regardless whether air or pure oxygen
absorption is used.
(ii) Presence of magnesium dioxide in dispersion (in a form
of fine powder of approximately 20 ␮m Sauter mean
diameter at the concentration of 0.8 g l−1 [15]) can en-
hance the rate of bubble coalescence and as a result may
diminish interfacial area. It was shown [15] that MnO2
does not influence kL a values in coalescent batch of
water. In non-coalescent batches, however, MnO2 may
enhance the bubble coalescence significantly. The kL a
Fig. 3. Comparison of kL a data measured by Linek et al. [14] in values measured by Alves et al. [3] in sulphate solution
one-impeller configuration by start-up dynamic method in sodium sul- are as low as they correspond rather to water (coales-
phate solution of various concentrations. cent batch) than to non-coalescent batches. Compare
V. Linek et al. / Chemical Engineering and Processing 43 (2004) 1511–1517 1515

Table 2
Average tank data on operational conditions, gas hold-up, bubble diameter, specific area and volumetric mass transfer coefficients measured by Alves
et al. [3] and those calculated from correlations (6) to (9)
Batch N (s−1 ) vs e (W m−3 ) d23 (mm) a (m−1 ) ε (×102 ) kL a (×102 s−1 ) kL (×104 m s−1 )
(mm s−1 ) (this work) (Ref. [3]) (Ref. [3])
Ref. [3] This work Ref. [3] This work Ref. [3] This work

0.3 M Na2 SO4 4.2 2.49 173 2.3 47 1.8 1.8 0.80 2.08 1.69 4.42
0.3 M Na2 SO4 5.0 2.49 271 1.67 78 2.2 2.3 1.3 3.61 1.66 4.63
0.3 M Na2 SO4 6.3 2.49 510 1.52 135 3.3 3.3 2.8 7.88 2.07 5.84
0.3 M Na2 SO4 7.5 2.49 832 1.24 213 4.4 4.2 3.0 14.4 1.41 6.77
0.3 M Na2 SO4 10.0 2.49 1901 0.9 512 7.2 6.3 6.2 39.9 1.21 7.80
0.3 M Na2 SO4 7.5 4.97 666 1.34 233 5.0 5.5 3.0 16.0 1.29 6.88
Water 7.5 2.49 800 2.89 53 2.5 2.3 1.7 3.54 3.19 6.89

Alves’ kL a value 0.062 s−1 for sulphate solution with from the pertinent kL a-values. The resulting kL values are
kL a = 0.066 s−1 for water calculated from correlations plotted against total power dissipated in the liquid in Fig. 4.
(6) through (9) at N = 10 s−1 and vs = 0.00249 m s−1 . Our kL data for all batches are well fitted by equation
 0.244  1/2
eν D
kL = 0.48 (11)
3.4. Specific interfacial area ρ ν
with the mean deviation of 4.6%. If the theoretical value 1/4
The specific interfacial area was measured by Alves et al.
of the power exponent is used the equation changes to
[3]. They measured local gas hold-up and local bubble size
 1/4  1/2
in 33 points in the vessel. From these local data they cal- eν D
culated the average gas hold-ups ε and interfacial areas a, kL = 0.523 (12)
ρ ν
which are shown in Table 2. The batches in which the mea-
surements were performed did not contain catalyst MnO2 . with the mean deviation of 4.7%. Similar correlation of
Area measured by the method represents geometric surface Alves’ kL -data yields
of all bubbles in dispersion.  −0.0497  1/2
−3 eν D
kL = 2.16 × 10 (13)
ρ ν
3.5. Mass transfer coefficient
with the mean deviation of 24.4%. In contradiction to our
Mass transfer coefficients calculated by the ratio kL a/a are data Alves’ data show a weak fall with increasing power
given in Table 2. kL a values calculated from correlation (9) dissipated in the liquid.
were up to six times higher than Alves’ kL a data. Accord- In Fig. 5, our and Alves’ kL data are plotted against bub-
ingly, the same disagreement exists between kL calculated ble diameter together with the lines representing limiting

Fig. 4. Liquid film mass transfer coefficient versus total power dissipated in the liquid. Filled symbols: water; empty symbols: salt solution (vs = 2.5 mm s−1 );
empty symbols with +: salt solution (vs = 4.97 mm s−1 ).
1516 V. Linek et al. / Chemical Engineering and Processing 43 (2004) 1511–1517

Fig. 5. Liquid film mass transfer coefficient versus average bubble diameter. Filled symbols: water; empty symbols: salt solution (vs = 2.5 mm s−1 );
empty symbols with +: salt solution (vs = 4.97 mm s−1 ).

values for the rigid, Eqs. (1) and (2), and the fully mobile ing power dissipated in the liquid. It seems that the reported
bubbles, Eqs. (3) and (4). The lines which follow from Alves decrease of kL with increasing power input dissipation in the
et al. [3] model of bubble contamination kinetics in water liquid is a result of mistake in kL a data. On the contrary, the
and sulphate solutions are also included in the figure. To fit data presented in this work increase with the power dissi-
experimental kL onto the model they [3] used three empir- pated in the liquid phase according to Eq. (12). The “eddy”
ical constants: k/Csurf = 27.5 for water, k/Csurf = 15 for model [4] based on turbulence theory gives successful fit-
sulphate solutions and a 35% reduction on rise velocities, as ting of all our kL results for water (coalescent batch) and
calculated from terminal velocities vsl = 0.65vt . The model sulphate solution (non-coalescent batch), contrary to “slip
fits Alves’ kL data with the mean deviation of 9.7%. Our kL velocity” model [3], which is based on bubble diameter and
data for sulphate solutions are well fitted by Eq. (3) for fully bubble-liquid relative velocity.
mobile bubble surface with vsl = 0.65 vt , but kL value for
water is higher by 64%. If the empirical reduction on rise
Acknowledgements
velocities is not used, i.e. vsl = vt , the model (3) does not
fit the data neither for sulphate solutions (being up to 35%
This work was supported by the Grant agency of Czech
higher) nor for water (being by 32% lower), see Fig. 5.
Republic through the project No. 104/98/1126 and by Czech
We conclude that Eq. (12) with one empirical parameter is
Ministry of Education through the scientific project No.
much better for fitting kL data measured in all batches used.
223400007.

4. Conclusions Appendix A. Nomenclature

Average volumetric mass transfer coefficients measured a specific interfacial area based on the liquid
by Alves et al. [3] using peroxide decomposition steady-state volume (m−1 )
method with air as gas phase are six times lower compared to c1,5 constants in Eqs. (1) and (5)
the data measured by dynamic pressure method using pure C1,2,3 constants in Eqs. (6)–(9)
oxygen. The abnormally low kL a results presented in [3] are d, d32 bubble diameter, Sauter mean diameter (m)
discussed and shown to be probably wrong, which implies D gas diffusivity in the liquid (m2 s−1 )
that kL data from the same source are unreliable. This kL e total specific power dissipated in the liquid
values are up to six times lower and decrease with increas- volume (W m−3 )
V. Linek et al. / Chemical Engineering and Processing 43 (2004) 1511–1517 1517

eagit specific power dissipated by agitators in the [4] J.C. Lamont, D.S. Scott, An eddy cell model of mass transfer into
liquid volume (W m−3 )
surface of a turbulent liquid, AIChE J. 16 (1970) 513–519.
[5] Y. Kawase, B. Halard, M. Moo-Young, Liquid-phase mass transfer
g gravitational constant (m s−2 ) coefficients in bioreactors, Biotechnol. Bioeng. 39 (1992) 1133–
k/Csurf empirical constant of Alves et al. [3] model 1140.
of bubble contamination kinetics [6] B.D. Prasher, G.B. Wills, Mass transfer in an agitated vessel, Ind.
kL liquid-side mass transfer coefficient (m s−1 ) Eng. Chem. Process. Des. Dev. 12 (1973) 351–354.
[7] Y. Kawase, B. Halard, M. Moo-Young, Theoretical prediction of vol-
kL a volumetric mass transfer coefficient referred to umetric mass transfer coefficients in bubble columns for Newtonian
the liquid volume (s−1 ) and non-Newtonian fluids, Chem. Eng. Sci. 42 (1987) 1609–1617.
N impeller speed (s−1 ) [8] V. Linek, P. Beneš, V. Vacek, F. Hovorka, Analysis of differences
vs superficial velocity (m s−1 ) in kL a values determined by steady state and dynamic methods in
stirred tanks, Chem. Eng. J. 25 (1982) 77–88.
vsl slip velocity (m s−1 ) [9] C.W. Robinson, C.R. Wilke, Simultaneous measurement of interfacial
vt single bubble terminal velocity in still water area and mass transfer coefficients for a well-mixed gas dispersion
(m s−1 ) in aqueous electrolyte solutions, AIChE J. 20 (1974) 285–294.
[10] I.T.M. Hassan, C.W. Robinson, Mass transfer coefficients in mechan-
Greek letters ically agitated gas-aqueous electrolyte dispersions, Can. J. Chem.
Eng. 58 (1980) 198–205.
ε overall fractional gas hold-up [11] V. Linek, P. Beneš, J. Sinkule, Critical assessment of the steady-state
ν kinematic viscosity of the liquid phase (m2 s−1 ) Na2 SO3 feeding method for kL a measurement in fermentors, Biotech-
ρ density of the liquid phase (kg m−3 ) nol. Bioeng. 35 (1990) 766–770.
ρ =(ρliquid − ρgas ) (kg m−3 ) [12] V. Linek, P. Beneš, J. Sinkule, T. Moucha, Non-ideal pressure step
method for kL a measurement, Chem. Eng. Sci. 48 (1993) 1593–
1599.
[13] T. Moucha, V. Linek, J. Sinkule, Effect of liquid axial mixing on
References local kL a values in individual stages of multiple-impeller vessel,
Collect. Czech. Chem. Commun. 63 (1998) 2103–2113.
[1] N. Frössling, Über die Verdünstung fallenden Tropfen, Gerlands [14] V. Linek, P. Beneš, O. Holecek, Correlation for volumetric mass
Beitage Geophys. 52 (1938) 170–216. transfer coefficient in mechanically agitated aerated vessel for oxygen
[2] P.H. Calderbank, M.B. Moo-Young, The continuous phase heat and absorption in aqueous electrolyte solutions, Biotechnol. Bioeng. 32
mass transfer properties of dispersions, Chem. Eng. Sci. 16 (1961) (1988) 482–490.
39–61. [15] J.M.T. Vasconcelos, A.W. Nienow, T. Martin, S.S. Alves, C.M.
[3] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, Gas–liquid mass transfer McFarlane, Alternative ways of applying the hydrogen peroxide
coefficient in stirred tank interpreted through bubble contamination steady-state method of kL a measurement, Trans. Ind. Chem. E 75A
kinetics, Chem. Eng. Proc., in press. (1997) 467–472.

You might also like