Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

download Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

of 13

Transcript of Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    1/13

    Journal of Biotechnology 97 (2002) 89101

    Mass transfer correlations for rotating drum bioreactors

    Matthew T. Hardin a,*, Tony Howes b, David A. Mitchell c

    a Chemical and Materials Engineering Department, Uni6ersity of Auckland, Auckland, New Zealandb Department of Chemical Engineering, Uni6ersity of Queensland, St Lucia, Qld 4072, Australia

    c Departamento de Bioqumica, Uni6ersidade Federal do Parana, Cx. P. 19046, 81531-990 Curitiba, Parana, Brazil

    Received 6 December 2001; received in revised form 18 March 2002; accepted 27 March 2002

    Abstract

    Evaporative cooling is extremely important for large-scale operation of rotating drum bioreactors (RDBs). Outlet

    water vapour concentrations were measured for a RDB containing wet wheat bran with the aim of determining the

    mass transfer coefficient for evaporation from the bran bed to the headspace. Mass transfer was expressed as the mass

    transfer coefficient times the area for transfer per unit volume of void space in the drum. Values of ka% were

    determined under combinations of aeration superficial velocities ranging from 0.006 to 0.017 ms1 and rotation rates

    ranging from 0 to 9 rpm. Mass transfer coefficients were evaluated using a variety of residence time distributions

    (RTDs) for flow in the gas phase including plug flow and well-mixed and a Central Jet RTD based on RTD studies.

    If plug flow is assumed, the degree of holdup at low effective Peclet ( Peeff) numbers gives an apparent under-estimate

    of ka% compared with empirical correlations. Values of ka% calculated using the Central Jet RTD agree well with values

    of ka% from literature correlations. There was a linear relationship between ka% and effective Peclet number:ka%=2.32103Peeff. 2002 Elsevier Science B.V. All rights reserved.

    Keywords: Mass transfer; Rotating drum; Solid state fermentation; Scale-up; Sherwood number

    Nomenclature

    area available for mass transfer per unit of void volume (m1)a

    area available for mass transfer per unit of headspace (m1)a %

    dimensionless water concentration calculated from Eq. (13)cC water concentration in dry air (kg kg1)

    dimensionless water concentration in headspacecairwater concentration in dry air of headspace (kg kg1)CAIR

    cbran dimensionless concentration of water at the bran surface

    www.elsevier.com/locate/jbiotec

    * Correponding author.

    E-mail address: [email protected] (M.T. Hardin).

    0168-1656/02/$ - see front matter 2002 Elsevier Science B.V. All rights reserved.

    PII: S 0 1 6 8 - 1 6 5 6 ( 0 2 ) 0 0 0 5 9 - 7

    mailto:[email protected]:[email protected]
  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    2/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 8910190

    cdead dimensionless concentration of water in the dead zone of the drum in the Central Jet RTD

    water concentration in dry air entering the drum (kg kg1)CINwater concentration in dry air exiting the drum (kg kg1)COUT

    cplug dimensionless concentration of water in the Central Jet of the Central Jet RTD

    saturation concentration of water in dry air at the bed temperature (kg kg1)CSATconstant of bed viscosity (dimensionless)CV

    D axial diffusion coefficient from the Central Jet RTD (dimensionless)

    particle diameter (m)ddrum diameter (m)Ddrum

    Dh hydraulic diameter of drum (m)

    acceleration due to gravity (ms2)g

    h maximum height of the bed (m)

    constant in Eq. (3) defined in Eq. (4)K

    mass transfer coefficient estimated using the Central Jet RTD (dimensionless)k

    area factor used by Stuart (1996) (dimensionless)n

    constant of bed porosity (dimensionless)N

    effective Peclet number across the surface of the bran (dimensionless)Peeffdegree of exchange between the different regions in Central Jet RTD (dimensionless)QmixReynolds number (dimensionless)Re

    s thickness of the mobile layer at the bed surface (m)

    Schmidt number (dimensionless)Sc

    Sherwood number (dimensionless)Sh

    t dimensionless time (number of residence times)

    velocity of the particles at the bottom of the moving layer (ms1)u0ueff effective velocity of gas over the surface of the bed (ms

    1)

    superficial gas velocity through drum (ms1)ugaverage particle velocity in the moving layer (ms1)upvolume of the stagnant region in the Central Jet RTD (dimensionless)Vdeadgas kinematic viscosity (m2 s1)6g

    x length along the drum (dimensionless)

    rotational speed of the drum (s1)z

    h % generic expression for mass transfer area available per unit volume (Eq. (1))

    mass transfer coefficient used by Blumberg and Schlunderi

    coefficient for fitted correlation of Blumberg and Schlunder

    diffusivity of water in air at the particle surface (m2 s1)lgdynamic angle of repose of bed ()k

    s generic expression for mass transfer coefficient (Eq. (1))

    generic expression for time term (Eq. (1))~

    V generic concentration term (Eq. (1))

    concentration term in Eq. (1)V1V2 concentration term in Eq. (1)

    factor in Eq. (35)x

    1. Introduction

    Rotating drum bioreactors (RDBs) have poten-

    tial to provide better heat and mass transfer

    characteristics than solid state fermentation (SSF)

    bioreactors with static beds, while providing gen-

    tler agitation than bioreactors with internal stir-

    rers. Furthermore, the absence of internal moving

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    3/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 89101 91

    parts for mixing makes design, construction and

    operation simpler, and the low pressure-drop

    across the bioreactor greatly reduces the operating

    costs associated with the aeration system. The

    gentle agitation associated with the tumbling mo-

    tion of the substrate bed minimises damage to the

    substrate particles, and makes RDBs ideal for

    those micro-organisms which can tolerate some

    shear damage, but which are affected deleteriously

    by more vigorous mixing.

    A major hurdle that currently prevents wide-

    spread use of SSF is the difficulty of regulating

    the temperature in large-scale fermentations. In

    an RDB, heat exchange, and hence removal of

    metabolic heat, is limited to convective cooling

    (from the surface of the drum to the surroundings

    or from the substrate to the headspace air) and

    evaporative cooling. As the fermentations increasein scale evaporative cooling becomes more impor-

    tant because the ratio of heat produced to surface

    area for convection declines. One of the main

    advantages of the mixed bed in an RDB is that it

    is practical to add water to the substrate during

    the fermentation, for example, by spraying a fine

    mist onto the surface of the moving bed (Barstow

    et al., 1988). This means that evaporative cooling

    can be used as the major mechanism of heat

    removal without the danger of restricting growth

    due to drying out of the substrate.

    Since mass transfer is directly proportional to

    the surface area available for particle/gas contact,

    it follows that any activity that improves this

    surface area will improve mass transfer. Both

    increasing the rotation speed and adding lifters

    will improve particle/gas contact. Evaporative

    mass transfer in rotating drums has been studied

    from the point of view of drying (Friedman and

    Marshall, 1949) but attempts to develop gener-

    alised models are comparatively recent (Riquelmeand Navarro, 1986). There have been few at-

    tempts to quantify these effects for SSF. In the

    current work, a series of experiments was per-

    formed to quantify the mass transfer as a function

    of fill depth, rotation speed and air inlet.

    Mass transfer is typically characterised by an

    equation of the form:

    #V

    (~=sh %(V1V2) (1)

    where V is a concentration of the component in

    question (e.g. mass water per unit mass dry air),

    V1V2 is the driving force for mass transfer

    (both V1 and V2 are concentration terms), s is the

    mass transfer coefficient (length per unit time),which characterises the rate at which transportoccurs in response to the driving force, h % is the

    area available for mass transfer per unit volume

    being transferred into (area per unit volume) and

    ~ is an expression of time. All of these terms may

    be expressed in a variety of units or even in

    dimensionless terms (e.g. time expressed as num-

    ber of residence times).

    For simplicity it is commonly assumed in rotat-

    ing drums that the mass transfer coefficient is

    constant over the length of the drum (Blumbergand Schlunder, 1996; Jauhari et al., 1998) and

    therefore, the challenge lies in establishing the

    driving force as a function of axial position within

    the drum. In order to estimate the driving force as

    a function of position, it is necessary to employ a

    model that describes the flow pattern in theheadspace to estimate the concentration of the

    species being transported as a function of posi-

    tion. Once a model for the concentration profile

    down the length of the drum is established, the

    mass transfer coefficient can be found easily.For the purposes of this analysis, it assumed

    that the moisture content of the substrate is con-

    trolled by water sprays (Barstow et al., 1988).

    This means that the driving force for evaporation

    depends only on the gas phase water concentra-

    tion as the substrate surface is saturated with

    water. This implies that film resistance from the

    liquid to the gas phase limits mass transfer hence,

    the system acts as if it is in the constant-rate

    drying period. In practice, many SSFs are carried

    out in the falling rate period. This means that

    evaporative transfer is limited by internal diffu-

    sion of water within the substrate and hence, is

    greatly affected by substrate composition and

    temperature.

    This paper describes an investigation of the

    effect on the mass transfer coefficient, expressed

    as the lumped parameter ka%, of varying the drum

    rotation rate and the velocity of the gas, for a

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    4/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 8910192

    drum operating in the slumping and tumbling

    regimes during a constant-rate drying period. It

    compares three gas flow patterns; plug flow, per-

    fect mixing and the Central Jet residence time

    distribution (RTD), derived from earlier experi-

    ments in our laboratory (Hardin et al., 2001) with

    an empirical correlation from literature (Blum-

    berg and Schlunder, 1996).

    2. Materials and methods

    2.1. Experimental procedure

    The reactor was a stainless steel rotating drum

    (Fig. 1), with inside dimensions of a radius of 280

    mm and a length of 800 mm. A variable speed

    drive capable of between 0 and 9 rpm turned thedrum.

    For simplicity, no organisms were grown. Since

    the rate of metabolic heat and water production

    varies with phase of growth of the organism and

    substrate composition, the use of organisms

    would have added complicating factors, making it

    difficult to compare different runs solely on the

    basis of drying. Damp Defiance type 2032 wheat

    bran (47% water wet weight, 2 mm diameter) was

    used to imitate one type of substrate typically

    used in these processes (Stuart, 1996; Marsh et al.,2000). The bran and water were mixed in the

    drum at full speed (9 rpm) for approximately 1 h.

    Many granules of bran formed during this time,

    and were broken up by hand before the bran was

    used for experimental purposes. The bran was

    removed from the drum and, after overnight stor-

    age at 4 C, was autoclaved at 121 C for 1 h to

    simulate the treatment the substrate normally un-

    dergoes. The bran was cooled to approximately

    ambient temperature before recording of results

    began.

    The humidity of the inlet air was measured at

    the inlet point of the drum while the bran was

    cooling. The inlet humidity was checked prior toeach series of trials. While there was variation in

    inlet humidity between series of trials, within each

    series the inlet air humidity was constant. The

    outlet humidity was measured on-line using an

    Electro-tech Model 5612B humidity probe and the

    relative humidity logged. The relative humidities

    were used to find water concentrations in the gas

    streams by use of the Antoine equation.

    2.2. Analysis of experimental results

    The actual surface area available for the mass

    transfer is a function of particle size and exposure

    of individual particles to the gas stream. Since this

    is extremely difficult to calculate, in this analysis

    the surface area per unit volume of headspace, is

    combined with the mass transfer coefficient k to

    give a ka% term. This is very similar to the way

    that oxygen transfer is handled in submerged

    liquid culture.

    The effect of rotation rate on mass transfer

    between the bed and headspace will be related to

    the flow regimes that correspond to the various

    rotation rates (Fig. 2). As the speed of rotation of

    the drum is increased, the regime of flow of the

    particles changes. At very low speeds the particles

    exhibit slumping flow where the solid bed periodi-

    cally slips back down the wall. At higher speeds

    the material on the top of the bed tumbles back

    down the face of the bed in a continuous cascade.

    At still higher speeds material is thrown into the

    air before landing again (cataracting). Beyond a

    certain critical speed there is sufficient centrifugal

    force to hold the bed against the wall. The exact

    transition between these regimes is a function of

    rotation speed particle size distribution and parti-

    cle cohesiveness and is difficult to predict. Flow

    regimes beyond tumbling were not observed in the

    rotational speed range of these experiments.

    Fig. 1. Experimental apparatus. The drum is 800 mm long

    560 mm diameter. The humidity probes measure the moisture

    content of the air as close as possible (within 50 mm) of the

    drum inlet and outlet.

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    5/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 89101 93

    Fig. 2. Flow regimes in rotating drums showing direction of

    rotation (anti-clockwise) and the direction of motion for the

    substrate bed.

    tion of radial gradients. The thickness of the

    mobile layer depends on the material, fill depth

    and the rotation speed.

    Blumberg and Schlunder (1996) studied mass

    transfer in rotating drums and developed an im-

    plicit equation for the thickness (s) of this mobile

    layer:

    pz [Ddrum(hs)(hs)2]Ks5/2

    +2pzs(Ddrum/2h+s)=0 (3)

    where z is the rotation rate in revolutions per s,

    Ddrum is the drum diameter, h is maximum bed

    height (Fig. 3) and K is a constant that depends

    on the material. This equation can be solved fairly

    simply by numerical methods. K is a function of

    the porosity of the bed and the dynamic angle of

    repose of the material (k). K can be calculated as

    follows:

    K=2

    3gsink

    CVd21/2

    3

    5

    12

    11N+

    18

    17N2

    12

    23N3+

    3

    29N4

    (4)

    where N is a dimensionless constant factor of

    porosity which is equal to 0.8 for most materials,

    CV is a dimensionless constant factor of bed vis-

    cosity (0.6 for most materials), g is the accelera-

    tion due to gravity (ms

    2

    ), k is the dynamic angleof repose of the bed and d is the particle diameter

    in metres (Savage, 1979). Blumberg and Schlunder

    (1996) give the full derivation of these equations.

    Note that Eq. (4) is different from the equation

    given in Blumberg and Schlunder due to a typo-

    graphical error in that paper. In the current work

    the dynamic angle of repose was measured as

    3792 through the Perspex end of the drum.

    Blumberg and Schlunder (1996) go on to show

    that the average velocity of the particles in the

    moving layer can then be calculated from:up=Ks

    3/2+u0 (5)

    where up is the average flow velocity of the parti-

    cles in the moving layer and u0 is the velocity of

    the particles in the moving layer in contact with

    the non-moving bed and is given by:

    u0=2pz(Ddrum/2h+s) (6)

    The maximum rotation rate used in the experi-

    ments was 9 rpm. This did not allow material to

    go beyond the cascading flow regime (Fig. 2).

    When material is in the cascading regime there is

    a mobile top layer that slides down the face of the

    bed (Fig. 3). It is assumed that mass transfer

    between the headspace and bed occurs only in this

    layer as individual particles in this layer are in

    rapid motion with respect to the gas phase. Con-

    versely, particles deeper in the bed have little

    contact with the gas. Movement of particleswithin the bed ensures that, ultimately, the entire

    bed is exposed to the air and minimises the forma-

    Fig. 3. Detail of drum in the tumbling regime showing the

    direction of travel for the mobile layer as well as the nomen-

    clature for equations for calculating the mobile layer thickness.

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    6/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 8910194

    The effect of superficial air velocity on mass

    transfer was also investigated. The air flow rate

    is expressed in terms of the superficial velocity

    (volumetric gas flow rate divided by unfilled

    cross-sectional area). The superficial velocity has

    an effect on the RTD of the air in the drum.

    The Central Jet RTD of Hardin et al. (2001),

    which is summarised below, is used in this paperto determine an estimate for the mass transfer

    coefficient based on realistic RTDs. Blumberg

    and Schlunder (1996) developed other correla-

    tions that are based on the Reynolds number of

    the gas flow through the drum and over the

    particles. Their correlations are described in Ap-

    pendix A.

    The overall mass transfer depends on the ef-

    fective velocity of the gas over the particles,

    which is simply the vector sum of the gas veloc-

    ity and the particle velocity (Eq. (5)). If plugflow is assumed, the two are at right angles to

    one another for a horizontal drum so the effec-

    tive velocity is as given in Eq. (7):

    ueff=ug2+up2 (7)

    In Blumberg and Schlunder (1996), an effec-

    tive Peclet number, Peeff, was calculated on the

    basis of the plug flow assumption, regardless of

    the real flow pattern. This simplification is help-

    ful for scale-up and allows comparison with lit-erature results. The effective velocity is

    combined with the diffusion coefficient (lg in

    m2 s1) and the particle diameter (d, in m) into

    an effective Peclet number, Peeff (Eq. (8)):

    Peeff=ueffd

    lg(8)

    The diffusivities used in this work are derived

    from correlations in McCabe et al. (1985).

    The Peclet number is a dimensionless group

    giving the ratio of convective flow to diffusive

    flow. The two extremes for Pe are (represent-

    ing a case where diffusion is insignificant) and 0

    (representing a case where convective transfer is

    insignificant). Peeff is used as a characteristic for

    our correlations, as it is a simple description of

    the gas flow over the particles. It also allows

    simple comparison between the different meth-

    ods of estimating ka%. Note, however, that the

    gas RTD is far from plug flow and the Peeff is

    not a physical quantity within the context of the

    Central Jet RTD.

    The driving force for the mass transfer is the

    most difficult part to estimate accurately because

    it depends on the RTD. Two extreme idealised

    flow patterns are the plug flow and well-mixedflow patterns. Plug flow assumes a front of gas

    moving through the drum with an axial concen-

    tration gradient along the drum with no axial

    mixing of the gas. The well-mixed distribution

    assumes that the gas is uniformly well-mixed

    within the drum and the exit concentration of a

    substance is equal to the internal concentration

    which is the same at every point in the drum.

    The flow pattern of the gas affects the concen-

    tration of water in the space above the bran.

    For the plug flow analysis the concentration ofwater in the air above the bran (CAIR) was esti-

    mated as the log mean of the inlet and outlet

    driving forces. For the well-mixed case, the out-

    let humidity was used to calculate the outlet

    moisture content of the air and this value was

    used as CAIR. Neither of these RTDs is strictly

    true. RTD studies of the system (Hardin et al.,

    2001) led to a three-parameter RTD involving a

    Central Jet region with axial dispersion, a sur-

    rounding stagnant region and a rate of exchange

    between the two regions. This Central Jet RTD

    accounts for back mixing within the drum, and

    is also used to analyse the mass transfer results

    obtained in the current work.

    The Central Jet RTD can be described by two

    partial differential equations:

    (cplug

    (t=D

    (2cplug

    (x2

    1

    (1Vdead)

    (cplug

    (x

    +Qmix

    (1Vdead)

    (cdeadcplug) (9)

    (cdead

    (t=

    Qmix

    Vdead(cdeadcplug) (10)

    where c refers to the dimensionless water con-

    centration (Eq. (12)), D refers to the axial dis-

    persion in the Central Jet, Vdead is the relative

    volume of the stagnant region and Qmix is the

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    7/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 89101 95

    relative rate of exchange between the stagnant

    (dead) and jet (plug) regions. The parameters D,

    Vdead and Qmix are strictly empirical and depend of

    the superficial gas velocity and possibly other

    factors including drum geometry. A fuller descrip-

    tion of the parameters is found in Hardin et al.

    (2001).

    If it is assumed that the bran lies entirely withinthe dead region then Eq. (10) can be modified to

    describe the transfer of water from the bran (Eq.

    (11)). Also the effective Peclet number should be

    expressed only in terms of the surface particle

    velocity, as this is the only motion of particles

    relative to the gas in the dead region. Hence, the

    effective Peclet number used in this paper is as given

    in Eq. (12).

    (cdead

    (t=

    Qmix

    Vdead(cdeadcplug)+ka(cbrancdead)

    (11)

    Peeff=uPd

    lg(12)

    The concentrations in the above differential

    equations are dimensionless. The dimensional con-

    centrations (C) are normalised with the saturation

    water concentration at the temperature of the bran

    (CSAT) (i.e. assuming that the water activity of the

    bran is 1) to give the variable c, where:

    c=CSATCAIR

    CSATCIN(13)

    This implies cbran=0 at the surface of the bran

    and cplug=1 at the entry to the drum. The actual

    water concentrations, CAIR and CIN, are calculated

    from the measured relative humidity. The concen-

    tration of water at the surface of the bran particle

    is assumed to be equal to CSAT. This must be true

    if the steady state analysis is to hold, that is, the

    data are collected during the constant-rate drying

    period. Tests on the water activity of the bran after

    the trials showed this to be so. The saturation

    humidity for the temperature is found from the

    Antoine equation and by assuming that the air

    behaves as an ideal gas.

    During the constant-rate drying period, the drum

    is at steady state therefore, the left hand sides of Eq.

    (9) and Eq. (11) are equal to zero i.e.

    0=D(2cplug

    (x2

    1

    (1Vdead)

    (cplug

    (x

    +Qmix

    (1Vdead)(cdeadcplug) (14)

    0=Qmix

    Vdead(cdeadcplug)+ka(cbrancdead) (15)

    Combining Eq. (14) and Eq. (15) to eliminate

    cdead gives:

    0=D(2cplug

    (x2

    1

    (1Vdead)

    (cplug

    (x

    +Qmix

    (1Vdead)

    Qmixcplug

    Vdead+Kcbran

    Qmix

    Vdead+K

    cplug

    (16)

    which is a second order differential equation of the

    form:

    0=Ay+By %+f(y) (17)

    with the boundary conditions:

    x=0, cplug=1 (18)

    x=1, cplug=CSATCOUT

    CSATCIN(19)

    Eq. (17) has the solution:

    cplug=Aexp(R1x)+Bexp(R2x) (20)

    where A and Bare constants and R1 and R2 are the

    roots of the equation

    Dy2y

    (1Vdead)+

    Qmix

    (1Vdead)

    Qmixy

    Vdead+kacbran

    Qmix

    Vdead

    +ka

    y

    =0 (21)

    The second boundary condition is used to elim-

    inate either A or B, based on the magnitude of the

    roots, with the first boundary condition used to set

    the retained constant equal to 1. Arranging the

    roots so that the magnitude of R1\0 and R2B0,

    the final solution is:

    cplug=exp(R2x) (22)

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    8/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 8910196

    Fig. 4. Conceptual model of water transport from bran bed to stagnant region to central jet region and thence to the exit of the

    drum. Note that the bran bed is assumed to be in contact only with the stagnant region.

    The outlet concentration is required and is rep-

    resented as:

    cplugx=1=exp(R2) (23)Rearranging R2 allows the lumped mass trans-

    fer coefficient ka% to be calculated from:

    NTU=Qmixf

    1f(24)

    where f is given by the expression:

    f=1Vdead

    4QmixD

    11Vdead

    2DlnC*2

    1

    1Vdead

    2n(25)

    and NTU represents the number of transfer units.

    ka=NTU

    Qmix(26)

    where ka is in terms of transfer area per unit of

    volume of the dead space in m3. Thence

    ka%=kaVdead (27)

    Conceptually, the movement of water vapour

    from the bran to the exit air can be imagined as

    shown in Fig. 4.

    It can be shown that the NTU will be ln(c) for

    plug flow and (1/c1) for the well-mixed case

    (Levenspiel, 1999), where c is the water concentra-

    tion as defined in Eqs. (5)(13). To calculate ka%

    using the Blumberg and Schlunder (1996) correla-

    tion, the mass transfer coefficients derived from

    the Blumberg and Schlunder (1996) correlations

    were multiplied by the flat area of the bed surface

    and divided by the void space of the drum.

    3. Results

    Outlet humidities were measured for the drum

    operated under different conditions of aeration

    rate, rotation rate and fractional filling. The re-

    sults were analysed using four different methods

    of estimating the apparent ka% number for each

    operating condition. The results of this analysis

    are in Table 1. The variation in fill depth changes

    the moving layer thickness and the average veloc-

    ity of the moving layer. This change is captured in

    the effective Peclet number (Peeff), however,

    changes in gas flow rate do not substantially

    change the Peeff for a given rotation speed. The

    calculated Peeff is more sensitive to changes in the

    rotation rate, which affect the average velocity of

    the particles in the mobile layer.

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    9/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 89101 97

    Table1

    Comparisonofka%derived

    fromdifferentmethods

    Peeff

    Estimatedka%

    %Filling

    Rotationspeed

    Airr

    ate

    (lmin

    1)

    (rpm)

    Plugflow

    Ce

    ntralJet

    BlumbergandSchlunder(1996)

    Well-mixed

    correlation

    (H

    ardinetal.,

    2001)

    0.0

    472

    0.0

    545

    30

    21.9

    0.0

    649

    155

    0.9

    0.0

    369

    0.0

    940

    0.0

    460

    0.0

    504

    0.0

    822

    30

    1.8

    32.9

    155

    3.6

    0.1

    323

    155

    0.1

    241

    30

    49.1

    0.1

    041

    0.0

    487

    0.1

    373

    0.1

    584

    0.0

    507

    61.7

    30

    0.1

    124

    5.4

    155

    0.0

    520

    155

    0.1

    252

    0.1

    885

    30

    72.5

    0.1

    180

    7.2

    0.1

    248

    0.0

    536

    0.2

    024

    0.2

    156

    9

    155

    30

    82.3

    0.0

    505

    155

    0.0

    538

    0.0

    714

    22.5

    19.8

    0.0

    731

    0.9

    0.0

    880

    0.0

    861

    0.0

    616

    1.8

    155

    0.1

    016

    29.5

    22.5

    0.0

    694

    155

    0.1

    337

    0.1

    125

    22.5

    43.7

    0.1

    264

    3.6

    0.1

    400

    5.4

    0.1

    351

    155

    22.5

    54.9

    0.1

    314

    0.0

    708

    0.1

    742

    0.1

    556

    0.0

    736

    0.1

    419

    63.9

    7.2

    155

    22.5

    0.0

    777

    155

    0.1

    141

    0.1

    741

    22.5

    72.2

    0.1

    569

    9

    0.0

    468

    155

    0.0

    449

    0.0

    570

    15

    17.3

    0.0

    630

    0.9

    0.0

    564

    0.0

    666

    0.0

    511

    0.0

    722

    1.8

    25.6

    15

    155

    0.0

    864

    0.0

    570

    0.0

    746

    0.0

    838

    3.6

    155

    15

    37.5

    0.0

    923

    0.0

    985

    0.0

    625

    5.4

    155

    0.1

    011

    46.5

    15

    0.0

    656

    155

    0.1

    164

    0.1

    114

    15

    54.1

    0.1

    095

    7.2

    0.1

    225

    0.0

    698

    0.1

    179

    0.1

    232

    155

    9

    15

    60.6

    0.1

    423

    0.1

    587

    0.0

    492

    61.3

    30

    0.1

    203

    5.4

    92.6

    0.0

    635

    118

    0.1

    286

    0.1

    587

    30

    61.8

    0.1

    562

    5.4

    0.1

    617

    0.0

    694

    0.1

    746

    0.1

    583

    5.4

    137

    30

    61.8

    0.1

    707

    0.0

    760

    0.1

    799

    0.1

    584

    155

    5.4

    30

    61.8

    0.1

    692

    0.1

    583

    30

    0.0

    837

    5.4

    174

    0.1

    852

    61.9

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    10/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 8910198

    Estimating a ka% using either the simple well-

    mixed RTD or the plug flow RTD yields values of

    ka% quite different from those estimated using the

    Central Jet RTD and also quite different from the

    values of ka% derived from the correlations of

    Blumberg and Schlunder (1996). Blumberg and

    Schlunder (1996) validated their correlation with a

    large number of different literature values. The

    values of ka% estimated using the more complex

    flow patterns of Hardin et al. (2001) seem to be

    more realistic than those estimated using the plug

    flow and well-mixed assumptions, based on their

    similarity to the numbers obtained using the ap-

    proach of Blumberg and Schlunder (1996). Blum-

    berg and Schlunder validated their approach

    against a wide range of experimental data, and

    therefore, the numbers obtained using their ap-

    proach should be reasonably close to the realvalues.

    Interestingly, the results of Blumberg and

    Schlunder (1996) were for smaller scales (maxi-

    mum diameter was 0.3 m) and, in general, higher

    length to diameter ratios than the experiments

    performed here. The effect of the difference in

    geometry on the air RTD probably caused the

    slight discrepancies between the estimates using

    the Central Jet RTD and the estimates found

    using the correlation of Blumberg and Schlunder

    (1996). These deviations from plug flow tend toincrease as drums increase in diameter to length

    ratio. Scaling up on the basis of a plug flow RTD

    is ill-advised as the RTD deviates from plug flow

    as scale increases with geometric similarity being

    maintained (Mecklenburgh and Hartland, 1975).

    The values of ka% estimated from the Central Jet

    assumption are probably the most reliable of the

    four estimations. They are based on a gas flow

    RTD which was experimentally determined for

    the drum (Hardin et al., 2001), and not on a plug

    flow RTD, which was a key assumption of theapproach of Blumberg and Schlunder (1996). The

    method of Blumberg and Schlunder for estimating

    the ka% differs from the simple plug flow analysis.

    The difference lies in the use of the correlations

    described in Appendix A. These modify the plug

    flow ka% based on correlations derived from dry-

    ing studies.

    The similarity, at small drum scales, between

    the values of the Central Jet ka% and ka% estimated

    using the Blumberg and Schlunder (1996) correla-

    tion shows that the convective mass transfer is

    independent of the composition of the particle

    itself. This is probably to be expected in the

    constant-rate drying period where the rate-limit-

    ing step is diffusion of water from the saturatedsurface of the particle. The materials used by

    Blumberg and Schlunder were mainly uniformly

    sized, spherical, inorganic particles. The values of

    ka% evaluated in the present trials were performed

    using damp wheat bran, which is non-uniform in

    size, non-spherical and organic. This similarity in

    ka% suggests workers in SSF can access data from

    the drying literature, even when it was obtained

    with very different particle types, provided they

    are operating in the constant-rate drying regime.

    Fig. 5 shows ka% estimated on the basis of theCentral Jet flow RTD (solid circles) and also

    using the Blumberg and Schlunder correlation

    (hollow circles) as a function of effective Peclet

    number. The line in the figure is the regression

    line for estimates using the Central Jet RTD,

    which was forced through the origin. The overall

    correlation for estimates of ka% using the Central

    Jet RTD is ka%=02.32103Peeff. As can be

    seen from Fig. 5 it applies for a wide range of

    Fig. 5. ka% vs. effective Peclet number. The estimates using the

    Blumberg and Schlunder (1996) method are shown as hollow

    circles and the estimates using the Central Jet RTD as solid

    circles. The line is the line of best fit for the ka% estimated using

    the Central Jet RTD (forced through the origin).

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    11/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 89101 99

    Peclet numbers with a reasonable correlation

    (R2=0.75).

    4. Discussion

    4.1. Comparison of methods of estimation

    The Blumberg and Schlunder (1996) method of

    estimating ka% differs from the technique using the

    Central Jet RTD developed in this chapter in

    three main ways. The two major differing assump-

    tions are, firstly, that the area for mass transfer is

    simply the flat area of the bed surface, and sec-

    ondly, that the gas flow regime through the drum

    is plug flow. The third is the use of correlations

    from the drying literature (Appendix A).

    Calculation of the ka% using the Central Jet

    RTD produces a term that includes the wholesurface area available for mass transfer. The cor-

    relations from the Central Jet RTD correlation

    therefore, include the effect on mass transfer of

    increasing mobile layer thickness with increasing

    rotational speed. The mobile layer means that the

    headspace/bed interface is not something sharp,

    but actually has a finite thickness, of the order of

    3 cm. This contrasts with the assumption in the

    work of Stuart (1996) of a very thin static air

    layer of 3 mm thickness where mass transfer was

    taking place. Stuarts thin layer was convoluted

    by increasing the rotation speed thus, increasing

    its surface area, with the fold increase in area

    being described by a factor n which varied be-

    tween 1 and 5. Note that this assumption (a very

    thin layer) was explicit in the work of Blumberg

    and Schlunder (1996). The values of ka% estimated

    in the current work incorporate the area of mass

    transfer so, do not make any assumption as to the

    thickness of the layer in which mass transfer

    occurs.

    The Central Jet RTD is also, by its very nature,

    a more realistic assessment of the gas flow regime

    and hence, driving force down a given drum axis.

    One disadvantage of predicting ka% from the Cen-

    tral Jet RTD is there is currently a lack of infor-

    mation on how to predict Central Jet parameters

    for different drum geometries and sizes. This im-

    plies that each different reactor must be character-

    ised in terms of its gas flow RTD before any

    attempt is made to estimate ka%. Further work is

    required to develop correlations for geometry and

    scale for Central Jet parameters. However, the use

    of an empirical RTD better characterises the gas

    flow and accommodates a wide variety of condi-

    tions. Characterisation of the RTD using the Cen-

    tral Jet RTD allows a more accurate estimate ofka% when back mixing becomes more prevalent.

    Also, only one calculation is needed in our ap-

    proach whereas, Blumberg and Schlunder require

    the worker to calculate four Sherwood numbers

    and combine these with a Reynolds-number-

    derived factor to produce a predicted Sherwood

    number. In our approach, after the RTD in a

    drum has been characterised the calculation of

    mass transfer coefficients for different conditions

    is trivial.

    4.2. Implications for design and operation of

    RDBs

    To date the most complete model of SSF pro-

    cesses in RDBs was developed by Stuart (1996).

    This model dealt with variations in rotation speed

    by a factor n, which varied the surface area avail-

    able for mass transfer from the flat area of the top

    of the bed (n=1) to a factor of five times this

    area for cataracting flow. However, the values

    assumed for this factor n were simply educated

    guesses and where not based on any theoretical or

    experimental evidence. Note that, in the present

    work, in which a 56-cm diameter drum was oper-

    ated at 9 rpm, cataracting flow of the particles

    was not observed. Nine rpm represents 16% of the

    critical speed, and as expected, at this speed a

    tumbling motion was observed. The correlations

    predicted from the Central Jet RTD demonstrate

    a much greater effect of rotation speed on mass

    transfer than that assumed by Stuart (1996). Sher-

    wood numbers increased by a factor of approxi-

    mately 4 over the range of rotation speeds from

    zero to 9 rpm.

    From a design perspective the correlations de-

    veloped in this paper (and shown in Fig. 5) give a

    mass transfer coefficient for a given Peclet num-

    ber. The effective Peclet number takes into ac-

    count the fill depth, the gas flow rate and the

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    12/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 89101100

    rotational speed. This figure allows the worker to

    estimate the ka% that corresponds to their calcu-

    lated Peeff, and therefore, to calculate the mass

    transfer coefficient in their system. This greatly

    simplifies design and calculation of mass transfer.

    Further work is required to understand the effect

    of scale on the Central Jet flow pattern, and

    therefore, on the values of the Central Jet RTDparameters. At the moment the correlations for

    Vdead, Qmix and D are valid only for the drum as

    described in Hardin et al. (2001). Different corre-

    lations based on scale and geometry, involving

    more RTD trials using different drums, are

    needed to improve the usefulness of the method

    and to test the ka%-effective Peclet number correla-

    tion. If these are found, scale-up on the basis of

    mass transfer phenomena within RDBs will be

    simplified.

    5. Conclusion

    The Central Jet RTD correlation offers a sim-

    ple yet accurate estimate of Sherwood numbers in

    RDBs for Peclet numbers up to 85. This will

    allow easy design of RDBs by improving esti-

    mates of evaporative cooling. Implicit in this cor-

    relation are the assumptions that the observed

    effects are independent of temperature and there

    are no changes in volume due to mass transfer.

    Also, this correlation is only tested for the con-

    stant-rate drying period.

    Acknowledgements

    David Mitchell thanks the Brazilian National

    Council of Research (Conselho Nacional de

    Pesquisa, CNPq) for a research scholarship and

    auxiliary funding.

    Appendix A. Brief description of Blumberg and

    Schlunder correlation

    The effective mass transfer coefficient, ig,eff esti-

    mated using the Blumberg and Schlunder correla-

    tion is calculated by combining the mass transfer

    coefficient around a single sphere, ig,p, and

    through a tube in the entrance region, ig,tube,

    using the hydraulic diameter of the free cross-sec-

    tion of the drum. These values were combined

    using:

    ig,eff=1.8ig,P+(1)ig,tube (28)

    where is evaluated from the fitted equation:

    =0.0015ueffd

    6g(29)

    where 6g is the gas kinematic viscosity. The two

    mass transfer coefficients from Eq. (28) are

    derived from the following correlations.

    Mass transfer around a single sphere:

    Shg,P=ig,Pd

    lg=2+

    Shg,lam

    2 +Shg,turb2 (30)

    where

    Shg,lam=0.664Red(Scg)1/3 (31)

    Shg,turb=0.037Red

    0.8Scg

    1+2.443Red0.1(Scg

    2/31)(32)

    Red=ugd

    6g(33)

    Mass transfer for forced convection in tubes:

    Shg,tube,lam=ig,tubeDh

    lg

    =3.663+0.73

    +(1.615(ReDScgDh/L)1/30.7)3

    + 2

    1+22Scg

    1/6(ReDScgDh/L)3

    n1/3(34)

    Shg,tube,turb=ig,tubeDh

    lg=

    (x/8)(ReD1000)Scg

    1+12.7x/8(Scg2/31)1+DhL2/3n (35)

    where:

    x= (1.82logReD1.64)2 (36)

    and

  • 7/27/2019 Correlaciones de Transferencia de Masa Para Hacer Girar Biorreactores de Tambor

    13/13

    M.T. Hardin et al. /Journal of Biotechnology 97 (2002) 89101 101

    ReD=ueffDh

    6g(37)

    References

    Barstow, L.M., Dale, B.E., Tengerdy, R.P., 1988. Evaporative

    temperature and moisture control in solid substrate fer-

    mentation. Biotechnol. Tech. 2, 237242.

    Blumberg, W., Schlunder, E.-U., 1996. Transversale

    Schuettgutbewegung und konvektiver Stoffubergang in

    Drehrohen. Teil 1: Ohne Hubschaufeln. Chem. Eng. Proc.

    35, 395404.

    Friedman, S.J., Marshall, W.R., 1949. Studies in rotary dry-

    ing, part IIheat and mass transfer. Chem. Eng. Prog. 45,

    573588.

    Hardin, M.T., Howes, T., Mitchell, D.A., 2001. Residence

    time distributions of gas flowing through rotating drum

    bioreactors. Biotechnol. Bioeng. 74, 145153.

    Jauhari, R., Gray, M.R., Masliyah, J.H., 1998. Gassolid

    mass transfer in a rotating drum. Can. J. Chem. Eng. 76,

    224232.

    Levenspiel, O., 1999. Chemical Reaction Engineering. Wiley,

    New York.

    McCabe, W.L., Smith, J.C., Harriott, P., 1985. Unit Opera-

    tions of Chemical Engineering. McGraw-Hill, Sydney.

    Marsh, A.J., Stuart, D.M., Mitchell, D.A., Howes, T., 2000.

    Characterizing mixing in a rotating drum bioreactor forsolid-state fermentation. Biotechnol. Lett. 22, 473477.

    Mecklenburgh, J.C., Hartland, S., 1975. The Theory of Back-

    mixing. Wiley.

    Riquelme, G.D., Navarro, G., 1986. Analysis and modelling

    of rotary dryerdrying of copper concentrate in drying of

    Solidsrecent international developments. In: Mujumdar,

    A.S. (Ed.), Wiley Eastern, New Delhi.

    Savage, S.B., 1979. Gravity flow of cohesionless granular

    materials in chutes and channels. J. Fluid Mech. 92, 53

    96.

    Stuart, D.M., 1996. Solid state fermentation in rotating drum

    bioreactors. PhD Thesis, University of Queensland.