You are on page 1of 8

Chemical Engineering and Processing 43 (2004) 823–830

Gas-liquid mass transfer coefficient in stirred tanks interpreted


through bubble contamination kinetics
S.S. Alves∗ , C.I. Maia, J.M.T. Vasconcelos
Department of Chemical Engineering, Instituto Superior Técnico, Centro de Eng. Biológica e Quı́mica, 1049-001 Lisboa, Portugal

Received 23 December 2002; received in revised form 18 May 2003; accepted 19 May 2003

Abstract

Experimental data on the average mass transfer liquid film coefficient (kL ) in an aerated stirred tank are presented. Liquid media used were
tap water, electrolyte solutions and water with controlled addition of tensioactive material. Values of kL range from those expected for bubbles
with a mobile surface to those expected for rigid bubbles. These data are quantitatively interpreted in terms of bubble contamination kinetics,
using a stagnant cap model, according to which bubbles suddenly change from a mobile interface to a rigid condition when surface tension
gradients, caused by surfactant accumulation, balance out shear stress.
© 2003 Elsevier B.V. All rights reserved.

Keywords: Tank; Bubble; Kinetics


1. Introduction vS
· D2/3 · ν−1/6
rigid
kL =c· (1)
d
Mass transfer effectiveness in gas–liquid contactors is
most often expressed by means of the volumetric mass trans- where d is the bubble diameter, D is the diffusivity, vS is
fer coefficient (kL a). This may be correlated, for example, the bubble-liquid relative velocity (slip velocity), ν is the
with power input per unit volume and gas superficial veloc- kinematic viscosity of the liquid and c is a constant value
ity, but the resulting correlations do not achieve any degree of ≈0.6. Experimental values of c have been found to vary
of generality. Too many phenomena contribute to the values from 0.42 to 0.95 [12,13].
of the film coefficient, kL and of the specific area a and their If the bubble has a mobile interface, then kL , which will be
combined effect cannot easily be predicted. Separation of named kL mobile , is given by the penetration model solution
kL and a in the volumetric mass transfer coefficient is thus [14]:
a first step for a better understanding of the underlying phe- 
vS
nomena. While separate determination of kL and a has been kL
mobile
= 1.13 · D1/2 (2)
carried out by a number of authors in bubble columns [1–9] d
and air-lifts [10], problems related to reliable measure- A considerable amount of literature data on bubble ab-
ment of kL make determination in stirred tanks particularly sorption in water [15–24] shows that experimental kL falls
difficult. between the limits defined by Eqs. (1) and (2), which may
The mass transfer film coefficient kL is a major function differ by a factor of >5 for small bubbles. The scatter of
of bubble rigidity. If a bubble is rigid, then kL , which will data is attributed to different methods of bubble release, to
rigid
be named kL in this case, is theoretically obtained by the different measurement techniques and to different system
equation proposed by Frössling [11] from laminar boundary purities.
layer theory: The dependence of bubble rigidity on bubble size and
on surface contamination has been widely recognized (e.g.
∗ Corresponding author. Tel.: +351-1-8417-188; Refs. [1,4,7,25–27]). There is experimental evidence sug-
fax: +351-1-8499-242. gesting that bubbles may be free of surface-active impurities
E-mail address: salves@alfa.ist.utl.pt (S.S. Alves). when they are formed, but that their behavior changes in

0255-2701/$ – see front matter © 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0255-2701(03)00100-4
824 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823–830

time as contaminants accumulate at the interface. This would 2. Model


explain why kL depends on time, both under uncontrol-
led conditions [12,16,17,28–32] or controlled conditions The model employs average values of bubble size, gas
[33,34]. holdup, specific interfacial area and bubble residence time
Contaminants with the greatest effect on the interface in the tank to calculate an average gas liquid film coefficient,
mobility, thus on mass transfer, are insoluble [21]. Surface kL .
contamination is particularly consistent with the observed For a population of n bubbles, kL is defined as
asymmetry of interface circulation that is implicit in Savic’s n 
 k Li ·dt
stagnant cap model [35,36]. According to this model, the in- tRi · di2
soluble monolayer of adsorbed surfactant is dragged by the i=1
kL = (3)
adjacent liquid towards the bottom of the bubble, where a N

stagnant cap region builds up. When the amount and type of di2
contaminant are unknown, Savic’s model describes a lim- i=1
iting case where the surface tension varies from its value where tR, is the bubble residence time. This equation may
for a pure system, at the bubble front, to a minimum at be simplified if all bubbles are assumed of average size and
the rear [21]. The gradient of surface tension so generated only two possible values of kL are considered, depending on
opposes the surface flow and increases the drag, up to the surface mobility: kmobile for fully mobile bubble and kL
rigid
L
point where it balances the viscous stress at the surface, for a rigid bubble of the given average diameter. Eq. (3) then
leading to immobilization. This hypothesis [37] is generally becomes:
accepted to explain the retardation of terminal velocity by rigid
surfactants [36]. It has also been demonstrated that the tran- kLmobile · t mobile + kL · (tR − t mobile )
kL = (4)
sition bubble behaviour from that of a fluid sphere to that tR
of a rigid particle is sharper for increased Reynolds number
where tmobile stands for the time span where bubbles behave
[38].
like mobile fluid particles. kmobile
L is given by the penetration
A particularly clear picture of the phenomenon started rigid
to emerge from the experiments of Schulze and Schlünder model solution (Eq. (2)). kL is using Frössling’s equation
[39]. Mass transfer coefficient (kL ) was determined from the Eq. (1). The average bubble residence time, tR , in Eq. (4)
dissolution rate of free-floating bubbles held stationary in a is calculated dividing the gas volume by the gas volumetric
downward water flow. A period of large initial kL was sud- flowrate, Q:
denly followed by a much lower kL value, consistent with VL · ε
tR = (5)
Eq. (1). The time span of the initial regime was so short in Q · (1 − ε)
Schulze and Schlünder’s system that it could only be de-
where VL is the liquid volume and ε is the overall gas
tected with highly soluble gases. More recently, the use of
holdup. An expression for the calculation of tmobile has been
a water cleaning system in a similar apparatus [40] allowed
deduced [40] assuming the stagnant cap model of bubble
the duration of the regime of large kL to be expanded by or-
surface contamination [36]:
ders of magnitude, so that it could easily be observed with
air and other slowly dissolving gases. Moreover, the initial d 1/2 · ln (d / htrans )
t mobile = k · (6)
large kL value was consistent with Eq. (2), thus showing that Csurf
the abrupt change to the slower mass transfer regime was
connected with surface immobilization. A simple model was where d is the bubble average diameter, Csurf is the surfac-
developed [40] based on the stagnant cap concept [37,41] to tant concentration, k is a constant of unknown value related
theoretically interpret contamination times for various sizes with surfactant properties and htrans is the bubble clean
of bubbles. Larger bubbles remain mobile for a longer pe- segment height at the transition point from mobile to rigid.
riod, as they are slower to accumulate enough contaminant Eqs. (1)–(6) constitute a model for predicting kL . Besides
for transition to rigidity, explaining the common knowledge bubble diameter, two parameters are required: htrans , which
that larger bubbles tend to be mobile while smaller bubbles was found earlier [40] to have a value of 6×10−4 m and the
tend to be rigid [21]. The model was also used to success- ratio k/Csurf , which depends on experimental conditions.
fully interpret experimental mass transfer data in airlift and
bubble column contactors [42]. 3. Experimental
This paper is an attempt at extending the analysis to
stirred tanks. kL a data from a double turbine stirred tank The experimental set-up consisted of a 0.292 m diameter,
are combined with previous data on local bubble diame- flat-bottomed, fully baffled Perspex vessel. Agitation was
ter and on local gas holdup obtained in the same apparatus provided by two 0.096-m standard Rushton turbines set at
[43,44] to determine experimental values of the film coeffi- clearances of 0.146 and 0.438 m, respectively above the tank
cient kL and try to interpret these in terms of the proposed base. The tank dimensions are shown in Fig. 1, together with
theory. the location of sampling points for local gas hold-up and
S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823–830 825

Fig. 1. Tank dimensions and location of experimental sampling points


(䊉). Distances in mm.

bubble size, data which were determined in previous work


[43,44]. The liquid media used were tap water 0.3 M aqueous
solution of sodium sulphate and 0.3 M aqueous solution
of sodium sulphate with 20 ppm PEG (surface tension, 63
mN/m). Operation conditions are presented in Table 1.
The overall volumetric mass-transfer coefficient, kL a, was
measured at 25 ± 0.5 ◦ C, using the peroxide decomposition
steady-state technique with manganese dioxide as the cat-
alyst [45]. Measurements of the dissolved oxygen concen-
tration CL were performed using two oxygen meters, WTW
Oxi340, equipped with galvanic probes WTW Cell Ox 325.
The kL a value was calculated from
Qperoxide
kL a = (7)
2VL log C
Table 1
Experimental conditions
Liquid phase Ref. N (s−1 ) Q (m3 s−1 )
Aqueous Na2 SO4 0.3 M S-N1-Q1 4.2 0.000167
S-N2-Q1 5.0 0.000167
S-N3-Q1 7.9 0.000167 Fig. 2. Local bubble size distributions for operating conditions S-N4-Q2
S-N4-Q1 7.5 0.000167 (see Table 1).
S-N5-Q1 10.0 0.000167
S-N4-Q2 7.5 0.000333
Aqueous Na2 SO4 0.3 M PEG-N4-Q1 7.5 0.000167
with 20 ppm PEG
Water W-N4-Q1 7.5 0.000167
826 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823–830

where Qperoxide is the peroxide molar addition to the liq- 4. Results and discussion
uid volume V and log C is the logarithmic mean between
the oxygen concentration in the liquid bulk, CL and the one Data on local gas hold-up and local average bubble di-
in equilibrium with the gas. The outlet oxygen concentra- ameter [43,44], obtained from bubble size distributions, as
tion in the gas phase was calculated assuming a constant shown in Fig. 2, allow local specific areas to be determined
volumetric gas flow across the vessel, which is accurate for the tank. From local data, the average interfacial specific
within ± 5%. kL a was determined at least twice under the area may easily be calculated using
same experimental conditions with a reproducibility within 
a i Vi
± 20%. a = tank (8)
V

Fig. 3. Local specific areas, a (m−1 ), for various operating conditions (see Table 1): (a) S-N2-Q1; (b, c) two runs of S-N4-Q1; (d) S-N4-Q2; (e)
PEG-N4-Q1; (f) W-N4-Q1.
S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823–830 827

Table 2
Experimental gas holdup, bubble size and volumetric mass transfer coefficient; calculated specific area and film coefficient for various tank conditions
Conditions ε d32 (m) a (m−1 ) kL a (s−1 ) kL (ms−1 )

S-N1-Q1 0.018 0.00230 47 0.008 0.000169


S-N2-Q1 0.022 0.00167 78 0.013 0.000166
S-N3-Q1 0.033 0.00152 135 0.028 0.000207
S-N4-Q1 0.044 0.00124 213 0.030 0.000141
S-N5-Q1 0.072 0.00090 512 0.062 0.000121
S-N4-Q2 0.050 0.00134 233 0.030 0.000129
PEG-N4-Q1 0.052 0.00121 253 0.022 0.000087
W-N4-Q1 0.025 0.00289 53 0.017 0.000319
Conditions as explained in Table 1.

where ai are the local experimental specific areas, Vi the


corresponding compartment volumes and V the tank volume.
Results are shown in Fig. 3. While large specific areas near
the turbines discharge are due to lower bubble diameter (see
Fig. 2), in other points of the tank they result from high local
gas hold-up.
Table 2 brings together average tank data on gas hold-up
and bubble diameter [43,44], specific area (as calculated
through Eq. (8)), experimental volumetric mass transfer co-
efficient, kL a and film coefficient given by the ratio kL a/a.
The resulting kL values are plotted against bubble diameter
in Fig. 4. These results have an estimated random error of
≈30%.
Theoretical values obtained from Higbie’s Eq. (2) for bub-
bles with mobile surface and from Frossling’s Eq. (1) for Fig. 4. Liquid film mass transfer coefficient versus average bubble dia-
rigid bubbles are also presented in Fig. 4. These depend meter. 䊏, Salt solution; 䊉, water; 䉱, PEG solution; - - -, simulated kL .
upon gas–liquid slip velocity, which, apart from the turbines
discharge jet, may be assumed to be a rise velocity. For the
relatively low gas holdups at play, rise velocities are close without PEG addition, behave in a manner that is interme-
to single bubble terminal velocities, which, both for rigid diate between that of bubbles in tap water and in PEG solu-
and for mobile bubbles rising in still water, may be esti- tion. This is because they are of intermediate size, but also
mated using correlations of experimental data, given in Clift because the liquid medium has intermediate contaminant
et al. [21]. It is however known that turbulence consider- concentration. Preparation of the salt solution with techni-
ably reduces bubble mean rise velocity, up to 50% [46,47]. cal sodium sulphate is likely to have introduced a level of
A correction for turbulence was introduced by a factor as- contaminant higher than what existed in the tap water. This
sumed equal for bubbles with both rigid and mobile surface. explains why parameter k/Csurf that fits the data for salt so-
This factor was adjusted by noticing that bubbles in 20 ppm lution is approximately half of that which fits tap water. If
PEG solution are mostly rigid due to the relatively high con- the contaminant were similar, this would mean contaminant
centration of contaminant, as observed in [42]. Their value concentration in the salt solution about double of that in
of kL should therefore fall on the Frössling line. This was tap water.
achieved by a 35% reduction on rise velocities, as calculated PEG (20 ppm), on the other hand, is certainly a higher
from terminal velocities. concentration than the trace levels of surfactant present either
Simulated kL , calculated by applying the simple model in tap water or in salt solution. It causes bubbles to rigidify
described in Section 2, is also superimposed on Fig. 4. While very quickly after formation. In previous work carried out
the previously determined value of 6×10−4 m was used for both in airlifts and in a bubble column, with considerably
parameter htrans [40], parameter k/Csurf is expected to vary larger bubbles and lower overall residence times, antifoam
with the liquid medium, since both the contaminant and its concentrations >10 ppm invariably led to rigid bubble values
concentration are probably different, thus affecting k and of kL [42].
Csurf . The effect of bubble size on kL which is apparent in
Bubbles in tap water are the closest to Higbie’s line, since the simulation curves is also clear from the salt solution
the water is relatively clean and the bubbles are relatively experimental points. It is due to the fact that larger bub-
large. Bubbles in salt solution (non-coalescing medium), but bles take longer to rigidify. Thus, they tend to move away
828 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823–830

5. Conclusions

Experimental data on the average mass transfer liquid film


coefficient, kL , in an aerated stirred tank range from those
expected for bubbles with a mobile surface, kLmobile , to those
rigid
expected for rigid bubbles, kL ,which are much lower.
Bubbles in PEG solution behave as rigid bubbles, while
bubbles in tap water behave closer to having a mobile inter-
face. Bubbles in salt solution have intermediate values of kL .
For the same liquid medium (salt solution) smaller bub-
rigid
bles result in lower values of kL , closer tokL .
These data can be quantitatively interpreted in terms of
bubble contamination kinetics, using a stagnant cap model,
Fig. 5. Fraction of bubbles with mobile interface versus average bubble according to which bubbles suddenly change from a mobile
diameter. interface to a rigid condition when surface tension gradi-
ents caused by surfactant accumulation balance out shear
stress.
from Frössling’s line, while smaller bubbles spend a greater
fraction of their residence time in the rigid regime, thus
approaching it. The estimated fraction of bubbles in the Appendix A. Nomenclature
tank that are mobile, x = tmobile /tR , is presented in Fig. 5.
The above theory may be used to estimate average values a specific interfacial area based on the liquid
of kL for each of the two halves of the tank, assuming that volume (m−1 )
there is no bubble recirculation from the top to the bottom c constant in Eq. (1)
half of the tank and that the air/liquid interface lose no con- C concentration (mol m−3 )
taminant in the top turbine. kL calculation for each half of d, d32 bubble diameter, Sauter mean diameter (m)
the tank follows the same steps as for the whole tank, only D gas diffusivity in the liquid (m2 s−1 )
using local average values of bubble size and gas holdup htrans height of the clean segment at the bubble
for the two halves of the tank, obtained from data in pre- front (m)
vious work [43,44]. Results are presented in Table 3. What k constant in Eq. (6) (mole m−7/2 s)
they indicate is rather surprising, namely, that the volumet- kL liquid-side mass transfer coefficient (m s−1 )
ric mass transfer coefficient for the salt solution is signifi- kL a volumetric mass transfer coefficient referred
cantly higher in the bottom half of the tank, in spite of the to the liquid volume (s−1 )
lower specific transfer area in that region. This is because n number of bubbles
the film coefficient is much higher in the bottom half, since N agitation rate (s−1 )
the fraction of clean bubbles is much larger there. There are Q gassing rate (m3 s−1 )
very few experimental data to assess these simulated results. Qperoxide peroxide solution addition flow rate (mol s−1 )
While they appear to agree with the experimental results by t time (s)
Moucha et al. [48], they disagree with results by Alves and tR residence time (s)
Vasconcelos [49] and Linek et al. [50]. V volume (m3 )

Table 3
Experimental gas holdup and bubble size; calculated specific area; simulated fraction of clean bubbles, film coefficient and volumetric mass transfer
coefficient for top and bottom halves of the tank for various tank conditions
Conditions Location ε d32 (m) a (m−1 ) x KL (ms−1 ) kL a (s−1 )

S-N2-Q1 Top 0.031 0.00167 110 0 0.000078 0.0085


Bottom 0.013 0.00164 46 0.62 0.000401 0.0184
S-N4-Q1 Top 0.053 0.00122 269 0 0.000085 0.0229
Bottom 0.033 0.00128 157 0.73 0.000198 0.0310
S-N4-Q2 Top 0.069 0.00144 285 0 0.000082 0.0233
Bottom 0.035 0.00120 175 0.67 0.000291 0.0509
PEG-N4-Q1 Top 0.060 0.00125 307 0 0.000084 0.0259
Bottom 0.039 0.00116 198 ∼0 0.000086 0.0170
Conditions as explained in Table 1.
S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823–830 829

vS slip velocity (m s−1 ) [17] P.H. Calderbank, A.C. Lochiel, Mass transfer coefficients, velocities
and shapes of carbon dioxide bubbles in free rise through distilled
x fraction of bubbles with mobile interface water, Chem. Eng. Sci. 19 (1964) 485–503.
[18] P.H. Calderbank, Gas absorption from bubbles, The Chemical Engi-
neer (1967) October, CE209–CE233.
Greek symbol [19] S.A. Zieminski, D.R. Raymond, Experimental study of the behavior
ε overall fractional gas holdup of single bubbles, Chem. Eng. Sci. 23 (1968) 17–28.
[20] D.R. Raymond, S.A. Zieminski, Mass transfer and drag coefficients
of bubbles rising in dilute aqueous solutions, AIChE J. 17 (1971)
Superscripts and subscripts 57–65.
i refers to zone i in the tank or to an [21] R. Clift, J.R. Grace, M.E. Weber, Bubbles, Drops, and Particles, Aca-
individual bubble demic Press, London, 1978, pp. 35–41 (pp. 125–137 and 169–199).
L liquid [22] M. Motarjemi, G.J. Jameson, Mass transfer from very small
bubbles—the optimum bubble size for aeration, Chem. Eng. Sci. 33
mobile mobile interface
(1978) 1415–1423.
rigid rigid interface [23] J.H. Hills, C.J. Abbott, L.J. Westall, A simple apparatus for the
surf surfactant measurement of mass transfer from gas bubbles to liquids, Trans.
Inst. Chem. Eng. 60 (1982) 369–372.
[24] F. Kawase, M. Moo-Young, Correlations for liquid-phase mass trans-
fer coefficients in bubble column reactors with Newtonian and
non-Newtonian fluids, Can. J. Chem. Eng. 70 (1992) 48–54.
References [25] K. Koide, S. Yamazoe, S. Harada, Effects of surface active substances
on gas holdup and gas liquid mass transfer in bubble column, J.
[1] W.W. Eckenfelder, E.L. Barnhart, The effect of organic substances Chem. Eng. Jpn. 18 (1985) 287–292.
on the transfer of oxygen from air bubbles to water, AIChE J. 7 [26] F. Kudrewizki, P. Rabe, Hydrodynamics and gas absorption in gassed
(1961) 631–634. stirred tanks in presence of tensids, Chem. Eng. Sci. 42 (1987)
[2] K. Akita, F. Yoshida, Bubble size, interfacial area, and liquid-phase 1939–1944.
mass transfer coefficient in bubble columns, Ind. Eng. Chem. Proc. [27] A. Morão, C.I. Maia, M.M.R. Fonseca, J.M.T. Vasconcelos, S.S.
Des. Dev. 13 (1974) 84–91. Alves, Effect of antifoam addition on gas–liquid mass transfer in
[3] D.N. Miller, Interfacial area, bubble coalescence and mass transfer stirred fermenters, Bioproc. Eng. 20 (1999) 165–172.
in bubble column reactors, AIChE J. 29 (1983) 312–319. [28] M.H.I. Baird, J.F. Davidson, Gas absorption by large rising bubbles,
[4] J.J. Jeng, J.R. Maa, Y.M. Yang, Surface effects and mass transfer in Chem. Eng. Sci. 17 (1962) 87–93.
bubble column, Ind. Eng. Chem. Proc. Des. Dev. 25 (1986) 974–978. [29] J.H. Leonard, G. Houghton, Mass transfer and velocity of rise phe-
[5] M.H.I. Baird, N.V.R. Rao, Characteristics of a countercurrent recip- nomena for single bubbles, Chem. Eng. Sci. 18 (1963) 133–142.
rocating plate bubble column. II. Axial mixing and mass transfer, [30] J.H. Hills, C.J. Abbott, L.J. Westall, A simple apparatus for the
Can. J. Chem. Eng. 66 (1988) 222–231. measurement of mass transfer from gas bubbles to liquids, Trans.
[6] T. Miyahara, M. Kurihara, M. Asoda, T. Takahashi, Gas-liquid inter- Inst. Chem. Eng. 60 (1982) 369–372.
facial area and liquid-phase mass transfer coefficient in sieve plate [31] A. Brankovic, I.G. Curie, W.W. Martin, Laser-Doppler measurements
columns without downcomer operating at high gas velocities, J. of bubble dynamics, Phys. Fluids 27 (1984) 348–355.
Chem. Eng. Jpn. 23 (1990) 280–285. [32] F. Bischof, M. Sommerfeld, F. Durst, The determination of mass
[7] F. Kawase, M. Moo-Young, The effect of antifoam agents on mass transfer rates from individual small bubbles, Chem. Eng. Sci. 46
transfer in bioreactors, Bioproc. Eng. 5 (1990) 169–173. (1991) 3115–3121.
[8] A. Cockx, M. Roustan, A. Line, G. Hebrard, Modelling of mass [33] K. Koide, Y. Orito, Y. Hara, Mass transfer from single bubbles in
transfer coefficient KL in bubble columns, Trans. Inst. Chem. Eng. Newtonian liquids, Chem. Eng. Sci. 29 (1974) 417–425.
73A (1995) 627–631. [34] K. Koide, T. Hayashi, K. Sumino, S. Iwamoto, Mass transfer from
[9] M. Bouaifi, G. Hebrard, D. Bastoul, M. Roustan, A comparative single bubbles in aqueous solutions of surfactants, Chem. Eng. Sci.
study of gas hold-up, bubble size, interfacial area and mass transfer 31 (1976) 963–967.
coefficients in gas-liquid reactors and bubble columns, Chem. Eng. [35] P. Savic, Circulation and distortion of liquid drops falling through
Proc. 40 (2001) 97–111. a viscous medium, National Research Council of Canada, Rep. No.
[10] M.Y. Chisti, M. Moo-Young, Airlift reactors: characteristics, appli- MT-22, 1953, (cited in Ref. 36).
cations and design considerations, Chem. Eng. Commun. 60 (1987) [36] R.M. Griffith, The effect of surfactants on the terminal velocity of
195–242. drops and bubbles, Chem. Eng. Sci. 17 (1962) 1057–1070.
[11] N. Frössling, bber die verdünstung fallenden tropfen (Evaporation of [37] A. Frumkin, V.G. Levich, On surfactants and interfacial motion
falling drops), Gerlands Beitage Geophys. 52 (1938) 170–216. (in Russian), Zh. Fizichesk. Khimii 21 (1947) 1183–1204.
[12] R.M. Griffith, Mass transfer from drops and bubbles, Chem. Eng. [38] F. Takemura, A. Yabe, Rising speed and dissolution rate of a carbon
Sci. 12 (1960) 198–213. dioxide bubble in slightly contaminated water, J. Fluid Mech. 378
[13] A.C. Lochiel, P.H. Calderbank, Mass transfer in the continuous phase (1999) 319–334.
around axisymmetric bodies of revolution, Chem. Eng. Sci. 19 (1964) [39] G. Schulze, E.U. Schlünder, Physical absorption of single gas bubbles
471–484. in degassed and preloaded water, Chem. Eng. Proc. 19 (1985) 27–37.
[14] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, [40] J.M.T. Vasconcelos, S.C.P. Orvalho, S.S. Alves, Gas-liquid mass
Wiley, New York, 1960, pp. 537–542. transfer to single bubbles: effect of surface contamination, AIChE J.
[15] P.H. Calderbank, M.B. Moo-Young, The continuous phase heat and 48 (2002) 1145–1154.
mass-transfer properties of dispersions, Chem. Eng. Sci. 16 (1961) [41] I.E. Scriven, Dynamics of a fluid interface. Equation of motion for
39–54. Newtonian surface fluids, Chem. Eng. Sci. 12 (1960) 98–108.
[16] F.H. Deindoerfer, A.E. Humphrey, Mass transfer from individual gas [42] J.M.T. Vasconcelos, J.M.L. Rodrigues, S.C.P. Orvalho, S.S. Alves,
bubbles, Ind. Eng. Chem. 53 (1961) 755–759. R.I. Mendes, A. Reis, Effect of contaminants on mass transfer co-
830 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823–830

efficients in bubble column and airlift contactors, Chem. Eng. Sci. [47] R.E.G. Poorte, A. Biesheuvel, Experiments on the motion of gas
58 (2003) 1431–1440. bubbles in turbulence generated by an active grid, J. Fluid Mech.
[43] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, Experimental and mod- 461 (2002) 127–154.
elling study of gas dispersion in a double turbine stirred tank, Chem. [48] T. Moucha, V. Linek, J. Sinkule, Measurement of kL a in
Eng. Sci. 57 (2002) 487–496. multiple-impeller vessels with significant axial dispersion in both
[44] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, A.J. Serralheiro, Bubble phases, Trans. Inst. Chem. Eng. Part A 73 (1995) 286–290.
size in aerated stirred tanks, Chem. Eng. J. 89 (2002) 109–117. [49] S.S. Alves, J.M.T. Vasconcelos, Mixing and oxygen transfer in
[45] J.M.T. Vasconcelos, A.W. Nienow, T. Martin, S.S. Alves, C.M. aerated tanks agitated by multiple impellers, in: Bioreactor and
McFarlane, Alternative ways of applying the hydrogen peroxide Bioprocess Fluid Dynamics, Mechanical Engineering Publications,
steady-state method of kL a measurement, Chem. Eng. Res. Des. Part London, 1993, pp. 3–14 (Third International Conference).
A 75 (1997) 467–472. [50] V. Linek, T. Moucha, J. Sinkule, Gas-liquid mass transfer in vessels
[46] P.D.M. Spelt, A. Biesheuvel, On the motion of bubbles in homoge- stirred with multiple-impellers—I. Gas–liquid mass transfer charac-
neous isotropic turbulence, J. Fluid Mech. 336 (1997) 221–244. teristics in individual stages, Chem. Eng. Sci. 51 (1996) 3203–3212.

You might also like