Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

  • Upload
    lili100

  • View
    217

  • Download
    0

Embed Size (px)

Citation preview

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    1/8

    Chemical Engineering and Processing 43 (2004) 823830

    Gas-liquid mass transfer coefficient in stirred tanks interpretedthrough bubble contamination kinetics

    S.S. Alves, C.I. Maia, J.M.T. Vasconcelos

    Department of Chemical Engineering, Instituto Superior Tcnico, Centro de Eng. Biolgica e Qumica, 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 ofkL 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

    Mass transfer effectiveness in gasliquid contactors is

    most often expressed by means of the volumetric mass trans-

    fer coefficient (kLa). This may be correlated, for example,

    with power input per unit volume and gas superficial veloc-

    ity, but the resulting correlations do not achieve any degree

    of generality. Too many phenomena contribute to the values

    of the film coefficient, kL and of the specific area a and their

    combined effect cannot easily be predicted. Separation of

    kL and a in the volumetric mass transfer coefficient is thus

    a first step for a better understanding of the underlying phe-

    nomena. While separate determination ofkL and a has been

    carried out by a number of authors in bubble columns [19]

    and air-lifts [10], problems related to reliable measure-

    ment of kL make determination in stirred tanks particularly

    difficult.

    The mass transfer film coefficient kL is a major function

    of bubble rigidity. If a bubble is rigid, then kL, which will

    be named krigidL in this case, is theoretically obtained by the

    equation proposed by Frssling [11] from laminar boundary

    layer theory:

    Corresponding author. Tel.: +351-1-8417-188;

    fax: +351-1-8499-242.

    E-mail address: [email protected] (S.S. Alves).

    krigid

    L= c vS

    dD2/3 1/6 (1)

    where d is the bubble diameter, D is the diffusivity, vS is

    the bubble-liquid relative velocity (slip velocity), is the

    kinematic viscosity of the liquid and c is a constant value

    of0.6. Experimental values of c have been found to vary

    from 0.42 to 0.95 [12,13].

    If the bubble has a mobile interface, then kL, which will be

    named kLmobile, is given by the penetration model solution

    [14]:

    kmobileL = 1.13

    vS

    dD1/2 (2)

    A considerable amount of literature data on bubble ab-

    sorption in water [1524] shows that experimental kL falls

    between the limits defined by Eqs. (1) and (2), which may

    differ by a factor of >5 for small bubbles. The scatter of

    data is attributed to different methods of bubble release, to

    different measurement techniques and to different system

    purities.

    The dependence of bubble rigidity on bubble size and

    on surface contamination has been widely recognized (e.g.

    Refs. [1,4,7,2527]). There is experimental evidence sug-

    gesting that bubbles may be free of surface-active impurities

    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

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    2/8

    824 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823830

    time as contaminants accumulate at the interface. This would

    explain why kL depends on time, both under uncontrol-

    led conditions [12,16,17,2832] or controlled conditions

    [33,34].

    Contaminants with the greatest effect on the interface

    mobility, thus on mass transfer, are insoluble [21]. Surface

    contamination is particularly consistent with the observedasymmetry of interface circulation that is implicit in Savics

    stagnant cap model [35,36]. According to this model, the in-

    soluble monolayer of adsorbed surfactant is dragged by the

    adjacent liquid towards the bottom of the bubble, where a

    stagnant cap region builds up. When the amount and type of

    contaminant are unknown, Savics model describes a lim-

    iting case where the surface tension varies from its value

    for a pure system, at the bubble front, to a minimum at

    the rear [21]. The gradient of surface tension so generated

    opposes the surface flow and increases the drag, up to the

    point where it balances the viscous stress at the surface,

    leading to immobilization. This hypothesis [37] is generally

    accepted to explain the retardation of terminal velocity bysurfactants [36]. It has also been demonstrated that the tran-

    sition bubble behaviour from that of a fluid sphere to that

    of a rigid particle is sharper for increased Reynolds number

    [38].

    A particularly clear picture of the phenomenon started

    to emerge from the experiments of Schulze and Schlnder

    [39]. Mass transfer coefficient (kL) was determined from the

    dissolution rate of free-floating bubbles held stationary in a

    downward water flow. A period of large initial kL was sud-

    denly followed by a much lower kL value, consistent with

    Eq. (1). The time span of the initial regime was so short in

    Schulze and Schlnders system that it could only be de-tected with highly soluble gases. More recently, the use of

    a water cleaning system in a similar apparatus [40] allowed

    the duration of the regime of large kL to be expanded by or-

    ders of magnitude, so that it could easily be observed with

    air and other slowly dissolving gases. Moreover, the initial

    large kL value was consistent with Eq. (2), thus showing that

    the abrupt change to the slower mass transfer regime was

    connected with surface immobilization. A simple model was

    developed [40] based on the stagnant cap concept [37,41] to

    theoretically interpret contamination times for various sizes

    of bubbles. Larger bubbles remain mobile for a longer pe-

    riod, as they are slower to accumulate enough contaminant

    for transition to rigidity, explaining the common knowledge

    that larger bubbles tend to be mobile while smaller bubbles

    tend to be rigid [21]. The model was also used to success-

    fully interpret experimental mass transfer data in airlift and

    bubble column contactors [42].

    This paper is an attempt at extending the analysis to

    stirred tanks. kLa data from a double turbine stirred tank

    are combined with previous data on local bubble diame-

    ter and on local gas holdup obtained in the same apparatus

    [43,44] to determine experimental values of the film coeffi-

    cient kL and try to interpret these in terms of the proposed

    theory.

    2. Model

    The model employs average values of bubble size, gas

    holdup, specific interfacial area and bubble residence time

    in the tank to calculate an average gas liquid film coefficient,

    kL.

    For a population of n bubbles, kL is defined as

    kL =

    ni=1

    kLidt

    tRi d2i

    Ni=1

    d2i

    (3)

    where tR, is the bubble residence time. This equation may

    be simplified if all bubbles are assumed of average size and

    only two possible values ofkL are considered, depending on

    surface mobility: kmobileL for fully mobile bubble and krigidL

    for a rigid bubble of the given average diameter. Eq. (3) then

    becomes:

    kL =kmobileL t

    mobile + krigidL (tR t

    mobile)

    tR(4)

    where tmobile stands for the time span where bubbles behave

    like mobile fluid particles. kmobileL is given by the penetration

    model solution (Eq. (2)). krigidL is using Frsslings equation

    Eq. (1). The average bubble residence time, tR, in Eq. (4)

    is calculated dividing the gas volume by the gas volumetric

    flowrate, Q:

    tR =VL

    Q (1 )(5)

    where VL is the liquid volume and is the overall gas

    holdup. An expression for the calculation of tmobile has been

    deduced [40] assuming the stagnant cap model of bubble

    surface contamination [36]:

    tmobile = k d1/2 ln (d/htrans)

    Csurf(6)

    where d is the bubble average diameter, Csurf is the surfac-

    tant concentration, k is a constant of unknown value related

    with surfactant properties and htrans is the bubble clean

    segment height at the transition point from mobile to rigid.

    Eqs. (1)(6) constitute a model for predicting kL. Besides

    bubble diameter, two parameters are required: htrans, whichwas found earlier [40] to have a value of 6104 m and the

    ratio k/Csurf, which depends on experimental conditions.

    3. Experimental

    The experimental set-up consisted of a 0.292 m diameter,

    flat-bottomed, fully baffled Perspex vessel. Agitation was

    provided by two 0.096-m standard Rushton turbines set at

    clearances of 0.146 and 0.438 m, respectively above the tank

    base. The tank dimensions are shown in Fig. 1, together with

    the location of sampling points for local gas hold-up and

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    3/8

    S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823830 825

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

    (). 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 solutionof sodium sulphate with 20 ppm PEG (surface tension, 63

    mN/m). Operation conditions are presented in Table 1.

    The overall volumetric mass-transfer coefficient, kLa, 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 kLa value was calculated from

    kLa =Qperoxide

    2VLlogC(7)

    Table 1

    Experimental conditions

    Liquid phase Ref. N (s1) Q (m3 s1)

    Aqueous Na2SO4 0.3 M S-N1-Q1 4.2 0.000167

    S-N2-Q1 5.0 0.000167

    S-N3-Q1 7.9 0.000167

    S-N4-Q1 7.5 0.000167

    S-N5-Q1 10.0 0.000167

    S-N4-Q2 7.5 0.000333

    Aqueous Na2SO4 0.3 M

    with 20 ppm PEG

    PEG-N4-Q1 7.5 0.000167

    Water W-N4-Q1 7.5 0.000167

    Fig. 2. Local bubble size distributions for operating conditions S-N4-Q2

    (see Table 1).

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    4/8

    826 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823830

    where Qperoxide is the peroxide molar addition to the liq-

    uid volume V and logC is the logarithmic mean between

    the oxygen concentration in the liquid bulk, CL and the one

    in equilibrium with the gas. The outlet oxygen concentra-

    tion in the gas phase was calculated assuming a constant

    volumetric gas flow across the vessel, which is accurate

    within 5%. kLa was determined at least twice under thesame experimental conditions with a reproducibility within

    20%.

    Fig. 3. Local specific areas, a (m1), 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.

    4. Results and discussion

    Data on local gas hold-up and local average bubble di-

    ameter [43,44], obtained from bubble size distributions, as

    shown in Fig. 2, allow local specific areas to be determined

    for the tank. From local data, the average interfacial specific

    area may easily be calculated using

    a =

    tank aiVi

    V(8)

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    5/8

    S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823830 827

    Table 2

    Experimental gas holdup, bubble size and volumetric mass transfer coefficient; calculated specific area and film coefficient for various tank conditions

    Conditions d32 (m) a (m1) kLa (s

    1) kL (ms1)

    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 0.000141S-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 Vthe 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, kLa and film coefficient given by the ratio kLa/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 Higbies Eq. (2) for bub-

    bles with mobile surface and from Frosslings Eq. (1) for

    rigid bubbles are also presented in Fig. 4. These depend

    upon gasliquid slip velocity, which, apart from the turbines

    discharge jet, may be assumed to be a rise velocity. For therelatively low gas holdups at play, rise velocities are close

    to single bubble terminal velocities, which, both for rigid

    and for mobile bubbles rising in still water, may be esti-

    mated using correlations of experimental data, given in Clift

    et al. [21]. It is however known that turbulence consider-

    ably reduces bubble mean rise velocity, up to 50% [46,47].

    A correction for turbulence was introduced by a factor as-

    sumed equal for bubbles with both rigid and mobile surface.

    This factor was adjusted by noticing that bubbles in 20 ppm

    PEG solution are mostly rigid due to the relatively high con-

    centration of contaminant, as observed in [42]. Their value

    of kL should therefore fall on the Frssling line. This was

    achieved by a 35% reduction on rise velocities, as calculated

    from terminal velocities.

    Simulated kL, calculated by applying the simple model

    described in Section 2, is also superimposed on Fig. 4. While

    the previously determined value of 6104 m was used for

    parameter htrans [40], parameter k/Csurf is expected to vary

    with the liquid medium, since both the contaminant and its

    concentration are probably different, thus affecting k and

    Csurf.

    Bubbles in tap water are the closest to Higbies line, since

    the water is relatively clean and the bubbles are relatively

    large. Bubbles in salt solution (non-coalescing medium), but

    Fig. 4. Liquid film mass transfer coefficient versus average bubble dia-

    meter. , Salt solution; , water; , PEG solution; - - -, simulated kL.

    without PEG addition, behave in a manner that is interme-

    diate between that of bubbles in tap water and in PEG solu-

    tion. This is because they are of intermediate size, but also

    because the liquid medium has intermediate contaminant

    concentration. Preparation of the salt solution with techni-

    cal sodium sulphate is likely to have introduced a level of

    contaminant higher than what existed in the tap water. This

    explains why parameter k/Csurf that fits the data for salt so-

    lution is approximately half of that which fits tap water. If

    the contaminant were similar, this would mean contaminant

    concentration in the salt solution about double of that in

    tap water.

    PEG (20 ppm), on the other hand, is certainly a higher

    concentration than the trace levels of surfactant present either

    in tap water or in salt solution. It causes bubbles to rigidify

    very quickly after formation. In previous work carried out

    both in airlifts and in a bubble column, with considerably

    larger bubbles and lower overall residence times, antifoam

    concentrations >10 ppm invariably led to rigid bubble values

    of kL [42].

    The effect of bubble size on kL which is apparent in

    the simulation curves is also clear from the salt solution

    experimental points. It is due to the fact that larger bub-

    bles take longer to rigidify. Thus, they tend to move away

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    6/8

    828 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823830

    Fig. 5. Fraction of bubbles with mobile interface versus average bubble

    diameter.

    from Frsslings line, while smaller bubbles spend a greater

    fraction of their residence time in the rigid regime, thusapproaching it. The estimated fraction of bubbles in the

    tank that are mobile, x = tmobile/tR, is presented in Fig. 5.

    The above theory may be used to estimate average values

    of kL for each of the two halves of the tank, assuming that

    there is no bubble recirculation from the top to the bottom

    half of the tank and that the air/liquid interface lose no con-

    taminant in the top turbine. kL calculation for each half of

    the tank follows the same steps as for the whole tank, only

    using local average values of bubble size and gas holdup

    for the two halves of the tank, obtained from data in pre-

    vious work [43,44]. Results are presented in Table 3. What

    they indicate is rather surprising, namely, that the volumet-

    ric mass transfer coefficient for the salt solution is signifi-cantly higher in the bottom half of the tank, in spite of the

    lower specific transfer area in that region. This is because

    the film coefficient is much higher in the bottom half, since

    the fraction of clean bubbles is much larger there. There are

    very few experimental data to assess these simulated results.

    While they appear to agree with the experimental results by

    Moucha et al. [48], they disagree with results by Alves and

    Vasconcelos [49] and Linek et al. [50].

    Table 3

    Experimental gas holdup and bubble size; calculated specific area; simulated fraction of clean bubbles, film coefficient and volumetric mass transfercoefficient for top and bottom halves of the tank for various tank conditions

    Conditions Location d32 (m) a (m1) x KL (ms

    1) kLa (s1)

    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.

    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, kmobileL , to those

    expected for rigid bubbles, krigidL ,which are much lower.

    Bubbles in PEG solution behave as rigid bubbles, whilebubbles in tap water behave closer to having a mobile inter-

    face. Bubbles in salt solution have intermediate values ofkL.

    For the same liquid medium (salt solution) smaller bub-

    bles result in lower values of kL, closer tokrigidL .

    These data can be 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 gradi-

    ents caused by surfactant accumulation balance out shear

    stress.

    Appendix A. Nomenclature

    a specific interfacial area based on the liquid

    volume (m1)

    c constant in Eq. (1)

    C concentration (mol m3)

    d, d32 bubble diameter, Sauter mean diameter (m)

    D gas diffusivity in the liquid (m2 s1)

    htrans height of the clean segment at the bubble

    front (m)

    k constant in Eq. (6) (mole m7/2 s)

    kL liquid-side mass transfer coefficient (m s1)

    kLa volumetric mass transfer coefficient referredto the liquid volume (s1)

    n number of bubbles

    N agitation rate (s1)

    Q gassing rate (m3 s1)

    Qperoxide peroxide solution addition flow rate (mol s1)

    t time (s)

    tR residence time (s)

    V volume (m3)

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    7/8

    S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823830 829

    vS slip velocity (m s1)

    x fraction of bubbles with mobile interface

    Greek symbol

    overall fractional gas holdup

    Superscripts and subscripts

    i refers to zone i in the tank or to an

    individual bubble

    L liquid

    mobile mobile interface

    rigid rigid interface

    surf surfactant

    References

    [1] W.W. Eckenfelder, E.L. Barnhart, The effect of organic substances

    on the transfer of oxygen from air bubbles to water, AIChE J. 7

    (1961) 631634.

    [2] K. Akita, F. Yoshida, Bubble size, interfacial area, and liquid-phase

    mass transfer coefficient in bubble columns, Ind. Eng. Chem. Proc.

    Des. Dev. 13 (1974) 8491.

    [3] D.N. Miller, Interfacial area, bubble coalescence and mass transfer

    in bubble column reactors, AIChE J. 29 (1983) 312319.

    [4] J.J. Jeng, J.R. Maa, Y.M. Yang, Surface effects and mass transfer in

    bubble column, Ind. Eng. Chem. Proc. Des. Dev. 25 (1986) 974978.

    [5] M.H.I. Baird, N.V.R. Rao, Characteristics of a countercurrent recip-

    rocating plate bubble column. II. Axial mixing and mass transfer,

    Can. J. Chem. Eng. 66 (1988) 222231.

    [6] T. Miyahara, M. Kurihara, M. Asoda, T. Takahashi, Gas-liquid inter-

    facial area and liquid-phase mass transfer coefficient in sieve platecolumns without downcomer operating at high gas velocities, J.

    Chem. Eng. Jpn. 23 (1990) 280285.

    [7] F. Kawase, M. Moo-Young, The effect of antifoam agents on mass

    transfer in bioreactors, Bioproc. Eng. 5 (1990) 169173.

    [8] A. Cockx, M. Roustan, A. Line, G. Hebrard, Modelling of mass

    transfer coefficient KL in bubble columns, Trans. Inst. Chem. Eng.

    73A (1995) 627631.

    [9] M. Bouaifi, G. Hebrard, D. Bastoul, M. Roustan, A comparative

    study of gas hold-up, bubble size, interfacial area and mass transfer

    coefficients in gas-liquid reactors and bubble columns, Chem. Eng.

    Proc. 40 (2001) 97111.

    [10] M.Y. Chisti, M. Moo-Young, Airlift reactors: characteristics, appli-

    cations and design considerations, Chem. Eng. Commun. 60 (1987)

    195242.

    [11] N. Frssling, bber die verdnstung fallenden tropfen (Evaporation of

    falling drops), Gerlands Beitage Geophys. 52 (1938) 170216.

    [12] R.M. Griffith, Mass transfer from drops and bubbles, Chem. Eng.

    Sci. 12 (1960) 198213.

    [13] A.C. Lochiel, P.H. Calderbank, Mass transfer in the continuous phase

    around axisymmetric bodies of revolution, Chem. Eng. Sci. 19 (1964)

    471484.

    [14] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena,

    Wiley, New York, 1960, pp. 537542.

    [15] P.H. Calderbank, M.B. Moo-Young, The continuous phase heat and

    mass-transfer properties of dispersions, Chem. Eng. Sci. 16 (1961)

    3954.

    [16] F.H. Deindoerfer, A.E. Humphrey, Mass transfer from individual gas

    bubbles, Ind. Eng. Chem. 53 (1961) 755759.

    [17] P.H. Calderbank, A.C. Lochiel, Mass transfer coefficients, velocities

    and shapes of carbon dioxide bubbles in free rise through distilled

    water, Chem. Eng. Sci. 19 (1964) 485503.

    [18] P.H. Calderbank, Gas absorption from bubbles, The Chemical Engi-

    neer (1967) October, CE209CE233.

    [19] S.A. Zieminski, D.R. Raymond, Experimental study of the behavior

    of single bubbles, Chem. Eng. Sci. 23 (1968) 1728.

    [20] D.R. Raymond, S.A. Zieminski, Mass transfer and drag coefficients

    of bubbles rising in dilute aqueous solutions, AIChE J. 17 (1971)

    5765.

    [21] R. Clift, J.R. Grace, M.E. Weber, Bubbles, Drops, and Particles, Aca-

    demic Press, London, 1978, pp. 3541 (pp. 125137 and 169199).

    [22] M. Motarjemi, G.J. Jameson, Mass transfer from very small

    bubblesthe optimum bubble size for aeration, Chem. Eng. Sci. 33

    (1978) 14151423.

    [23] J.H. Hills, C.J. Abbott, L.J. Westall, A simple apparatus for the

    measurement of mass transfer from gas bubbles to liquids, Trans.

    Inst. Chem. Eng. 60 (1982) 369372.

    [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) 4854.

    [25] K. Koide, S. Yamazoe, S. Harada, Effects of surface active substances

    on gas holdup and gas liquid mass transfer in bubble column, J.Chem. Eng. Jpn. 18 (1985) 287292.

    [26] F. Kudrewizki, P. Rabe, Hydrodynamics and gas absorption in gassed

    stirred tanks in presence of tensids, Chem. Eng. Sci. 42 (1987)

    19391944.

    [27] A. Moro, C.I. Maia, M.M.R. Fonseca, J.M.T. Vasconcelos, S.S.

    Alves, Effect of antifoam addition on gasliquid mass transfer in

    stirred fermenters, Bioproc. Eng. 20 (1999) 165172.

    [28] M.H.I. Baird, J.F. Davidson, Gas absorption by large rising bubbles,

    Chem. Eng. Sci. 17 (1962) 8793.

    [29] J.H. Leonard, G. Houghton, Mass transfer and velocity of rise phe-

    nomena for single bubbles, Chem. Eng. Sci. 18 (1963) 133142.

    [30] J.H. Hills, C.J. Abbott, L.J. Westall, A simple apparatus for the

    measurement of mass transfer from gas bubbles to liquids, Trans.

    Inst. Chem. Eng. 60 (1982) 369372.

    [31] A. Brankovic, I.G. Curie, W.W. Martin, Laser-Doppler measurementsof bubble dynamics, Phys. Fluids 27 (1984) 348355.

    [32] F. Bischof, M. Sommerfeld, F. Durst, The determination of mass

    transfer rates from individual small bubbles, Chem. Eng. Sci. 46

    (1991) 31153121.

    [33] K. Koide, Y. Orito, Y. Hara, Mass transfer from single bubbles in

    Newtonian liquids, Chem. Eng. Sci. 29 (1974) 417425.

    [34] K. Koide, T. Hayashi, K. Sumino, S. Iwamoto, Mass transfer from

    single bubbles in aqueous solutions of surfactants, Chem. Eng. Sci.

    31 (1976) 963967.

    [35] P. Savic, Circulation and distortion of liquid drops falling through

    a viscous medium, National Research Council of Canada, Rep. No.

    MT-22, 1953, (cited in Ref. 36).

    [36] R.M. Griffith, The effect of surfactants on the terminal velocity of

    drops and bubbles, Chem. Eng. Sci. 17 (1962) 10571070.

    [37] A. Frumkin, V.G. Levich, On surfactants and interfacial motion

    (in Russian), Zh. Fizichesk. Khimii 21 (1947) 11831204.

    [38] F. Takemura, A. Yabe, Rising speed and dissolution rate of a carbon

    dioxide bubble in slightly contaminated water, J. Fluid Mech. 378

    (1999) 319334.

    [39] G. Schulze, E.U. Schlnder, Physical absorption of single gas bubbles

    in degassed and preloaded water, Chem. Eng. Proc. 19 (1985) 2737.

    [40] J.M.T. Vasconcelos, S.C.P. Orvalho, S.S. Alves, Gas-liquid mass

    transfer to single bubbles: effect of surface contamination, AIChE J.

    48 (2002) 11451154.

    [41] I.E. Scriven, Dynamics of a fluid interface. Equation of motion for

    Newtonian surface fluids, Chem. Eng. Sci. 12 (1960) 98108.

    [42] J.M.T. Vasconcelos, J.M.L. Rodrigues, S.C.P. Orvalho, S.S. Alves,

    R.I. Mendes, A. Reis, Effect of contaminants on mass transfer co-

  • 7/28/2019 Gas-Liquid Mass Transfer Coefficient in Stirred Tanks Interpreted

    8/8

    830 S.S. Alves et al. / Chemical Engineering and Processing 43 (2004) 823830

    efficients in bubble column and airlift contactors, Chem. Eng. Sci.

    58 (2003) 14311440.

    [43] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, Experimental and mod-

    elling study of gas dispersion in a double turbine stirred tank, Chem.

    Eng. Sci. 57 (2002) 487496.

    [44] S.S. Alves, C.I. Maia, J.M.T. Vasconcelos, A.J. Serralheiro, Bubble

    size in aerated stirred tanks, Chem. Eng. J. 89 (2002) 109117.

    [45] J.M.T. Vasconcelos, A.W. Nienow, T. Martin, S.S. Alves, C.M.

    McFarlane, Alternative ways of applying the hydrogen peroxide

    steady-state method of kL a measurement, Chem. Eng. Res. Des. Part

    A 75 (1997) 467472.

    [46] P.D.M. Spelt, A. Biesheuvel, On the motion of bubbles in homoge-

    neous isotropic turbulence, J. Fluid Mech. 336 (1997) 221244.

    [47] R.E.G. Poorte, A. Biesheuvel, Experiments on the motion of gas

    bubbles in turbulence generated by an active grid, J. Fluid Mech.

    461 (2002) 127154.

    [48] T. Moucha, V. Linek, J. Sinkule, Measurement of kLa in

    multiple-impeller vessels with significant axial dispersion in both

    phases, Trans. Inst. Chem. Eng. Part A 73 (1995) 286290.

    [49] S.S. Alves, J.M.T. Vasconcelos, Mixing and oxygen transfer in

    aerated tanks agitated by multiple impellers, in: Bioreactor and

    Bioprocess Fluid Dynamics, Mechanical Engineering Publications,

    London, 1993, pp. 314 (Third International Conference).

    [50] V. Linek, T. Moucha, J. Sinkule, Gas-liquid mass transfer in vessels

    stirred with multiple-impellersI. Gasliquid mass transfer charac-

    teristics in individual stages, Chem. Eng. Sci. 51 (1996) 32033212.