CFD de tratamiento de aguas.pdf

download CFD de tratamiento de aguas.pdf

of 12

Transcript of CFD de tratamiento de aguas.pdf

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    1/12

    Chemical Engineering Journal 172 (2011) 8495

    Contents lists available at ScienceDirect

    Chemical Engineering Journal

    journa l homepage: www.elsevier .com/locate /ce j

    CFD modelling ofhydrodynamics and degradation kinetics in an annular slurryphotocatalytic reactor for wastewater treatment

    Nana Qia,b, Hu Zhang b,, BoJin b,c,d, Kai Zhang a,e,

    a State Key Lab ofHeavyOil Processing, ChinaUniversity ofPetroleum, Beijing102249, Chinab School ofChemical Engineering, The University ofAdelaide, Adelaide, SA 5005,Australiac School ofEarth andEnvironmental Sciences, TheUniversity ofAdelaide, Adelaide, SA 5005,AustraliadAustralianWater Quality Centre, SAWater Corporation, Bolivar, SA 5110,Australiae National Engineering Lab for Biomass Power GenerationEquipment, NorthChina Electric PowerUniversity, Beijing 102206, China

    a r t i c l e i n f o

    Article history:

    Received 20 November 2010

    Received in revised form 17 May 2011

    Accepted 17 May 2011

    Keywords:

    Annular slurry photocatalysis reactor

    Water treatment

    Computational fluid dynamics

    Hydrodynamics

    Degradation kinetics

    a b s t r a c t

    Photocatalytic degradation process has been recognized as a low-cost, environmental friendly and sus-

    tainable technology for water and wastewater treatment. As a key carrier ofthe photocatalytic process,

    the semi-conduct photoreactor has been employed in many studies. Analysis and modelling ofhydrody-

    namics and degradation kinetics in the three phase system can provide useful information for process

    design, operation and optimization. We have modelled the hydrodynamics and degradation kinetics in

    an annular photocatalytic bubble column reactor using an Eulerian multi-fluid approach. All the results

    (gas holdup, fluid flow patterns and organic concentrations) were evaluated against experimental data

    from our lab or literature reports. The local information on gas holdup, solid particle concentration, light

    intensity as well as organic pollutant concentration has been obtained. The simulation results reveal that

    the degradation rate oforganic pollutants is controlled by the local hydrodynamics and light intensity.

    By combining CFD and degradation kinetics models, the mechanisms of governing the photocatalytic

    reaction can be determined. The computational models will also help optimize reactor design to improve

    the hydrodynamics and light intensity distribution and provide guided information for scale-up.

    2011 Elsevier B.V. All rights reserved.

    1. Introduction

    Around 4 billion people around the worldhaveno or little access

    to clean and sanitized water supply and millions ofpeople have

    lost their lives due to severe waterborne diseases [1]. Figures are

    expected to grow in the near future because of the increasing

    industrial practices and water contamination by harmful organic

    substances. Organicpollutants in wastewater from many manufac-

    turing industries, such as textile and paper industries, are toxic and

    biologically non-biodegradable [2], which cannot be removed by

    commonly used biological treatment process. These contaminants

    can be removed using conventional treatment processes, such as

    adsorption onto granular activated carbon, reverse osmosis and air

    Corresponding author at: School of Chemical Engineering, The University of

    Adelaide, SA 5005, Australian, Tel.: +61 8 83033810; fax: +61 8 83034373. Corresponding author at: National Engineering Lab for Biomass Power Genera-

    tion Equipment, North China Electric Power University, Beijing 102206, China.

    Tel.: +86 10 61772413.

    E-mail addresses: [email protected] (H. Zhang), [email protected]

    (K. Zhang).

    stripping [1,3,4]. Most ofthese processes are highly costly and none

    are environmental-friendly. Photocatalytic oxidation (PCO), how-

    ever, has been recognized as an effectiveprocess in the degradation

    and mineralization ofa wide variety ofpriority pollutants in water

    and wastewater [5,6], as well as a green sustainable process by

    producing environmentally harmless compounds (carbon dioxide,

    water and mineral acids) after degradation [7]. During the process,

    oxygen reactive species are generated on the surface oftitanium

    dioxide (TiO2) particles, and these oxidative species can degrade

    nearly all organic pollutants.

    The overall photocatalytic reactions taking place in a semi-

    conduct reactor can be divided into five independent steps: mass

    transfer of the organic contaminants in the liquid phase to the

    TiO2 surface; adsorption of the contaminants onto the photon-

    activated TiO2 surface; photocatalytic reaction for the adsorbed

    phase on the TiO2 surface; desorption ofthe intermediate(s) from

    the TiO2 surface; and mass transfer of the intermediate(s) from

    the interface region to the bulk fluid [8,9]. Great efforts have been

    made to improve the performance for above steps, including syn-

    thesis of new catalysts [10,11] and optimization of an annular

    slurry photoreactor system [12]; design ofphotocatalytic reactor

    [13]; and optimization ofoperational parameters [10]. It has been

    1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved.

    doi:10.1016/j.cej.2011.05.068

    http://dx.doi.org/10.1016/j.cej.2011.05.068http://dx.doi.org/10.1016/j.cej.2011.05.068http://dx.doi.org/10.1016/j.cej.2011.05.068http://www.sciencedirect.com/science/journal/13858947http://www.elsevier.com/locate/cejmailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.cej.2011.05.068http://dx.doi.org/10.1016/j.cej.2011.05.068mailto:[email protected]:[email protected]://www.elsevier.com/locate/cejhttp://www.sciencedirect.com/science/journal/13858947http://dx.doi.org/10.1016/j.cej.2011.05.068
  • 7/29/2019 CFD de tratamiento de aguas.pdf

    2/12

    N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495 85

    Nomenclature

    a specific area (m1)

    A constant

    CD drag force coefficient

    CL lift force coefficientCTD turbulent diffusion coeffitcient

    C1 parameters in the standard k model

    C2 parameters in the standard k modelC parameters in the standard k modeld diameter (m)

    D kinematic diffusivity in the liquid phase (m2 s1)

    Eo Etvs number

    F total interfacial force (N)g gravity acceleration (m s2)G spectral incident radiation (Wm2)

    Ib blackbody emission intensity (Wm2 sr1)

    Io radiation intensity leaving the boundary

    (W m2 sr1)

    I spectral radiation intensity (Wm2 sr1)

    k turbulence kinetic energy per unit mass (m2 s2)

    ko,ls liquid solidorganicmass transfer coefficient (m s1)

    K apparent reaction rate constant (s1)Ka absorption coefficient (m

    1)

    Ks scattering coefficient (m1)

    M Morton numberP pressure (Pa)

    qr radiative flux (W)

    r position vector (m)

    Re Reynolds numbers director vector (m)

    S radiation intensity source term (W m2 sr1)

    ScT turbulence Schmidt number

    So,s reaction source oforganic in solid phase (s1)t time (s)

    T local absolute temperature (K)

    u velocity vector ofgas or liquid (ms1

    )ug superficial gas velocity (m s1)Y mass fraction

    Greek letters

    volume fraction turbulence dissipation rate (m2 s3) viscosity (Pas) light frequency (Hz) density (kgm3) surface tension (Nm1)t turbulent Prandtl numbertl liquid turbulent Schmidt numberk turbulence model constant for the k equation

    k turbulence model constant interphase masstransfersource termoforganicpol-

    lutants per unit volume (kgm3) in-scattering phase function

    s(x,y,z,t) solid particle concentration solid angle (sr)

    Superscripts

    D drag

    L liftT turbulent dispersion

    Subscripts

    b bubble

    c catalyst

    g gas-phase

    i gas, liquid or solid phase

    j other phase

    l liquid-phase

    o organicp particle

    ref reference

    s solid-phase

    t turbulence

    demonstrated that there are a number of systemic factors which

    could affect the oxidation rates and efficiency ofthe photocatalytic

    system, including TiO2 loading [1,10,11] pH [10,11,14,15], tem-

    perature [1], dissolved oxygen [1,10,11], contaminants and their

    loading [1,10,11], light wavelength [1] and light intensity [1,16].

    Kinetics studies are often conducted to assemble the above opera-tional parameters into a simple model for scalingup andoptimizing

    photoreactor system. However, Emeline et al. [17] argued that

    the kinetic data alone does not establish the actual mechanism of

    photocatalytic reaction, which must be supported by other evi-

    dences outside the field ofkinetics. These evidences include an

    understanding of the hydrodynamics and the radiation intensity

    distributions in the multiphase photoreactors. This requirement

    becomes essential for further scale-up and optimization.

    Chong et al. [10] have reported a promising photocatalytic pro-

    cess by using a newly synthesized titania impregnated kaolinite

    nano-photocatalyst in an annular slurry photoreactor. Congo Red,

    which is an organic substance often used as a surrogate for organic

    pollutants was completely degraded after 4 h irradiation. Chemical

    oxygen demand (COD) was reduced to 20% in the system in 2 h. Itwas noticed that the degradation rate was relatively fast and then

    became gradually slow. The significance ofhydrodynamics inside

    the slurry reactor has also been experimentally demonstrated by

    providing agitation and controlled aeration [10]. The role ofhydro-

    dynamics in the process, however, remains unclear as complicated

    interactions exist between ranges ofoperational parameters.

    Computational fluid dynamics (CFD) techniques have been

    widely used in stirred reactors [18], bubble columns [1921], flu-

    idized beds [22] and loop reactor [23,24]. CFD models have been

    developed to address the hydrodynamics in the photocatalytic

    bubble column reactors [19,25]. However, most studies are only

    focused on hydrodynamics while kinetics models have not been

    included. In this study, the radiationintensitydistributionand reac-

    tion kinetics models were combined with CFD models to elucidate

    the degradation mechanism in the photocatalytic system.

    2. Mathematicmodels

    2.1. Hydrodynamicmodels

    Three-dimensional transient CFD models were developed to

    compute the local hydrodynamics of the gasliquidsolid three-

    phase annular photocatalytic bubble column reactor. An Eulerian

    multi-fluid approach was adopted to describe the flow behaviours

    of each phase. Fluid is considered to be the continuous phase,

    whereas gas bubbles and solid particles are assigned to the dis-

    persed phases. Based on the principles of conservation of mass

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    3/12

    86 N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495

    and momentum, the continuity and momentumequations for each

    phase in a photocatalytic bubble column reactor are given below:

    (ii)

    t+ (ii ui) = 0 (1)

    where, and u stand for the volume fraction, density and velocityvector, respectively. The subscript i represents gas, liquid or solid

    phase.

    (ii ui)t

    + (ii ui ui) = iPi + (ii( ui + ( ui)T

    ))

    + Fi + ii g (2)

    where P, , and gare the pressure, viscosity and gravity accelera-tion, respectively. Fi is the interfacial forces acting on phase i dueto the presence ofthe other phase,j.

    2.1.1. Interface models

    Virtual mass force is ignored in comparison with drag, buoyancy

    and turbulent dispersion forces [2628]. Accordingly, the interfa-

    cial forces ofgreat significance are assumed to be drag force, lift

    force and interphase turbulent dispersion force in this study:

    Fi = FDi + FLi + FTi (3)

    where FDi

    , FLi

    and FTi

    are the interfacial forces due to drag force, lift

    force, turbulent dispersion, respectively.

    The drag component of the gasliquid interfacial force term is

    given as:

    FDlg =3

    4

    CD,lgdg

    l gug ul (ug ul) (4)

    where CD is drag force coefficient. The Grace relation is chosen

    for the drag force coefficient because bubbles are experimentally

    observed to be spherical and dispersed. The drag force coefficient

    [29,30] is:

    CD =

    4

    3

    gdbu2T

    l (5)

    wheredb stands for the mean bubble diameter, is the differencein density between the liquid and gas phases, and uT is the bubble

    terminal rise velocity. It can calculated as follows:

    uT =lldb

    M0.149(J 0.857) (6)

    In the above expression, M, the Morton number related to the

    fluid property, is defined by:

    M =4lg

    2l3

    (7)

    where is surface tension andJis given by:

    J= 0.94H0.751 2 < H 59.3 (8a)

    J= 3.42H0.441 H > 59.3 (8b)

    In Eqs. (8a) and (8b), H is expressed as:

    H =4

    3EoM0.149

    lref

    0.14(9)

    where ref is the molecular viscosity oftap water under a refer-ence temperature and pressure [29] and Eo, the Etvs number, is

    defined as:

    Eo =gd2

    b

    (10)

    The lift force, FLi

    , exerting on the liquid phase from the gas phase

    is given by:

    FLlg = CLl g(ug ul) ( ul) (11a)

    The turbulent dispersion force, FTi

    , is calcuated by the model of

    Lopez de Bertodano [31]:

    FTlg = CTDCDtltl

    gg

    ll (12a)

    where, CTD is the momentum transfer coefficient for the inter-

    phase drag force, CD is the drag coefficient as described in Eq. (5),

    tl stands for turbulent viscosity, and tl is the liquid turbulentSchmidt number. g and l are the gas and liquid phase volume

    fractions, respectively.

    In the liquidsolid system, the drag component of the

    liquidsolid interfacial force is similar to that ofgasliquid:

    FDls =3

    4

    CD,lsds

    lsus ul (us ul) (13)

    The drag coefficient exerted by the solid phase on the liquid

    phase, CD,ls is obtained by the Schiller Naumann drag model [30]:

    CD,ls= max24Re (1 + 0.15Re

    0.687

    ),0.44

    (14)

    where

    Re =lds

    us ull

    (15)

    The lift force, FLi

    , is determined by:

    FLls = CLls(us ul) ( ul) (11b)

    where CL is a non-dimensional lift coefficient, and it is assumed to

    a constant, the value is listed in Table 1 [32].

    The turbulent dispersion force, FTi

    , for liquidsolid interaction is

    similar to that ofgasliquid:

    FT

    ls= C

    TDCD

    tl

    tls

    s

    l

    l (12b)

    All items are explained in Eq. (12a), except that s is the solidvolume fraction.

    2.1.2. Turblence model

    The standard k model for single phase flows has beenextended for the three phase flows for simulating the turbulence

    in the present study, which can be described as:

    t(llkl) + (ll ulkl) =

    l

    l +

    tlk

    kl

    + lPl

    lll (16)

    t(lll) + (llull) = ll + tl

    l

    +l

    lkl

    (C1pl C2ll) (17)

    where C1, C2,k, and are parametersin the standardkmodeland the values are listed in Table 1. The liquid phase turbulent vis-

    cosity is calculated based on the Sato enhanced turbulence model

    [33]:

    tl = tl,s + tl,b (18)

    where tl,s is the conventional shear-induced turbulent viscosity,which is obtained by the standard k model as:

    tl,s = Clk2

    (19)

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    4/12

    N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495 87

    Table 1

    Constantsused in the CFD models.

    CL C1 C2 C k D (m2 s1) ScT ko,ls (m s

    1) ds (m) db (m)

    0.5 1.45 1.9 0.05 1.0 1.3 4.4 1010 0.9 2.0 104 3.5 106 2 103

    where C is a constant, which is listed in Table 1. tl,b is a bubble-

    induced component ofturbulent viscosity given by:

    tl,b = C,bl gdbug ul (20)

    andthe particle-induced componentofturbulentviscositygiven

    by:

    tl,p = C,plsdsus ul (21)

    The gas and solid phase turbulence is modelled using a zero

    equation model, in which gas turbulent viscosity is proportional to

    liquid phase turbulent viscosity [27,30]:

    tg =gl

    tlt

    (22a)

    ts =sl

    tst

    (22b)

    wheretis a turbulentPrandtl number relating the dispersed phasekinematic eddy viscositytgandtl to the continuous phase kine-

    matic eddy viscositytl.

    2.2. Radiation transport equation

    The radiation transport equation (RTE) is employed to charac-

    terize the light intensity distribution inside the photoreactor. The

    robust form of the RTE was first published by Spadoni et al. [34],

    while it was more recently applied to the photocatalytic system

    [35].

    A radiation balance is made along a given direction ofprop-

    agation together with the appropriate boundary conditions [34].

    Assumptions are made to simplify the simulation: shadowing,

    reflection, and refraction effects have little influence on the finaldistribution [25]. The spectral radiative transfer equation (RTE) can

    be written as:

    dI(r, s)

    ds= (Ka + Ks)I(r, s) + KaIb(v, T)

    +Ks4

    4

    I(r, s)(s s

    ) d + S (23)

    where rand s are the position and director vector, respectively;Kaand Ks represent wavelength-averaged absorption and scattering

    coefficient, which were calculated from the equations proposed by

    Pareek [13] as shown in Eq. (24a); v is the light frequency; s is the

    photon path length; Ib is blackbody emission intensity (which is

    negligible whenoperated at ambienttemperature)and I is spectral

    radiation intensity which depends on position (r) and direction (s);Tis local absolute temperature; is solid angle; is in-scatteringphase function; and S is the radiation intensity source term.

    Ka = 0.2758 Wcat (24a)

    Ks = 3.598 Wcat (24b)

    where Wcat is the catalyst loading of6 g/dm3.

    The RTE is a first order integro-differential equation for I in a

    fixed direction, s. To solve this equation within a domain, a bound-

    ary conditionfor I is required. The diffusely emitting and reflecting

    opaque boundary is defined as:

    I(rw, s) = (rw)Ib(v, T) +w(rw)

    ns

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    5/12

    88 N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495

    Fig. 1. Schematic diagram ofthe photoreactor used in the simulation.

    Base on the assumptions,the transportequationsfor the organic

    pollutants are described as:

    t(llYo,l) + (llYo,l ul) = [l(lDo,l +

    T,lScT,l

    )(Yo,l)]

    So,l o,ls (29a)

    t(ssYo,s) = So,s + o,ls (29b)

    where Yis the mass fraction oforganic pollutants in liquid or solid

    phase; D is the kinematic diffusivity ofCongo Red in water, which

    is listedin Table 1 [20];t is the turbulence viscosityofeach phase;ScT is the turbulence Schmidt number, the value is listed in Table 1

    [30]; S is the user specified masssource termoforganic species; and

    is the interphase mass transfer source term oforganic pollutantsper unit volume; The subscript l, s means the liquid and solid phase,

    respectively; o denotes organic species.

    The balance for species in each phase is given blow:

    Yo,l + Yw,l = 1 (30a)

    Yo,s + Yc,s = 1 (30b)

    where the subscriptw andcmeans water and catalyst, respectively.

    The reaction source is defined as:

    S = K Y (31a)

    Although the Langmuir-Hinshelwood (L-H) model is used for

    most kinetics studies in the photomineralization, the pseudo-

    first-order model could be more applicable when the organics

    Fig. 2. Gas phase velocity vectors in a vertical plane ofthe photocatalytic reactor.

    concentration is low. The above equation is employed to describe

    the loss oforganic species in the solid phase due to chemical reac-

    tion. K is the apparent reaction rate constant, which is a function

    of local incident radiation G(x,y,z) [39] and solid particle concen-

    tration s(x,y,z,t). As there is no suitable expression for the rateconstant Kin the Chongs photocatalytic system [10], Kis assumed

    to be proportional to the incident radiation and solid particle con-

    centration:

    K=AG(x, y, z)s(x, y, z, t) (31b)

    where A is a constant; G(x,y,z) is gained by integrating Iwhich is

    computed from Eq. (28) and s(x,y,z,t) is updated from the CFDcalculation from Eq. (1).

    The interphase mass transfer between the liquid and solidphase

    is defined as:

    o,ls = ko,ls as (s Y s Yo,s) (31c)

    where ko,ls is the liquid solid organic mass transfer coefficient, the

    value is listed in Table 1 (see details in [40]); Yo,s is the saturatedmass fraction in solid phase, which can be calculated from the

    adsorption isotherm; and as is photocatalyst specific area, equal

    to 6s/ds, s is the volume fraction ofsolid particle and ds is theparticle diameter.

    Because the organic pollutant concentration is quite low, its

    degradation reaction and interphase mass transfer are neglected

    for the continuity equations (Eq.(1)) andthe momentum equations

    Eq. (2) [30].

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    6/12

    N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495 89

    3. Numerical procedure

    3.1. Geometry and grid generation

    As can be seen in Fig.1, a conical bottomed and internally irradi-

    ated photocatalytic reactor withUV light being placed in the central

    core axis ofthe reactor is used [10]. The reactor has a diameter of

    0.149 m and a height of0.05 m.The initial slurry height is 0.0322 m,

    which will be used for all simulations in this work. The details

    for the experimental equipment characteristics can be found in

    Ref. [10]. An unstructured numerical grid has been implemented

    with a total number of278,235 elements, after a grid independent

    investigation ofthree levels ofgrids.

    3.2. Initial and boundary conditions

    In this work, the bubble size is treated to be uniformly dis-

    tributed in the multiphase reactor. As (d1.50 uglg0.5/) 1 in our

    study (whered0 is the diameterofthe orificeandugis the superficial

    gas velocity at the distribution plate, the size ofbubbles produced

    at the orifice db is calculated by Chen et al. [41]:

    db = 2.9d0gl

    1/3(32)

    The calculated result ofthe bubble diameter db at the orifice is

    listed in Table 1.

    Thegas inlet is providedat a uniform inlet gas velocity according

    to the experiment. In order to prevent a significant liquid loss by

    introducing air into the reactor, an inlet air velocity from zero to its

    final value over the first 2.5 real seconds is used [42]. At the gas out-

    let, an outlet condition at the top ofthe reactor is defined based on

    the proposal from Padial et al. [42], which was successfully adopted

    by Michele and Hempel [43]: the periphery area is defined as the

    opening boundary where only liquid can leave the computation

    domain; the inner area is defined as the degassingboundarywhere

    only gas can leavethe computation domain; the solid is kept within

    the reactor [30,44]. Along the reactor walls and the UV light tube,gas and solid phase are treated as free-slip, while for liquid phase

    no-slip boundary condition is applied [45]. Initially, the reactor is

    filledwith liquidand the solidparticles distribute uniformly within

    the reactor, and the liquid and particles are stationary. The den-

    sity and diameter ofphotocatalysts are 4507 kg m3 and 3.5m,

    respectively. The initial organic concentration is 40 mg/dm3.

    In order to solve the RTE described above sufficient boundary

    conditions need to be defined, which are source boundary condi-

    tion and wall properties. A line source model is used to determine

    the source boundary condition [13]. The inner quartz wall through

    which the UV radiation passes into the reacting media is assumed

    has an emissivity of1. While the outer wall made ofstainless steel

    was assumed has an emissivity of0.1 [30].

    3.3. Numerical procedure

    The simulations are carried out in a platform ofANSYS CFX

    12.0 software [30]. The high resolution scheme were selected as

    the iteration scheme to numerically solve the mass and momen-

    tum equations (Eqs. (1), (2)), while the implicit first order upwind

    scheme were selected to numerically solve the turbulence equa-

    tions (Eqs. (16), (17)) and mass transfer equations (Eq. (29a)),

    respectively [46,47]. Two different time steps are applied for vari-

    ables changingwith time.When calculating the distribution for gas

    bubbles and solids, a time step of0.01 s was used, while for mass

    transfer between liquid and solid phase and chemical reaction of

    theorganicspecies,alargetime step,0.1 s,was applied.A maximum

    residual convergent target of1 104 was set for all simulations.

    Fig. 3. liquidphase velocity vectors in a vertical plane ofthe photocatalytic reactor.

    4. Results and discussion

    4.1. Flow patterns

    Hydrodynamics, transport, and mixing properties depend

    strongly on the prevailing flow pattern in bubble columns [48].

    Three flow patterns, i.e., bubbly flow, transitional flow, and tur-

    bulent flow patterns, have been observed according to the upward

    movement ofthe bubble swarm in the column [21]. The gas- and

    liquid-phase velocity vectors within the column are presented in

    Fig. 2 when the simulations reach steady state. It can be seen from

    Fig. 2, at a superficial gas velocity of0.036 m s1, corresponding to

    10 dm3/min ofair flow rate, the gas bubble swarm starts to swing

    towards the column wallin the middle partofthe reactor and move

    towards the central tube at the gas outlet. There are no vorticesformed in the column while the movement pathways ofgas bub-

    bles are skewed. The gas bubbles move relatively slower in the

    central part in the middle ofthe reactor and in the region close

    to the reactor wall in the upper part.

    The liquid phase is carried upward by the bubble swarm and

    flows downward in the spiral manner between the central bubble

    swarm and column wall. The liquid downward produces several

    vortices as can be seen from the liquid vectors in a vertical plane

    (Fig. 3). It can be noticed that the vortices in the left and right are

    not symmetric, which is due to the gas bubbles move upward in a

    spiral manner. And these asymmetric vortices can be captured by

    the three-dimensional simulation. In the current geometric config-

    uration at a gas flow rate of 10 dm3/min, three vortices are found

    in the left and right, respectively. Two large vortices are located in

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    7/12

    90 N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495

    the upper and lower part of the reactor, while the middle one is

    relatively small and away from the centre.

    The simulation results are qualitatively in agreement with the

    three-dimensional flow structuresexaminedby Chenet al. [49] and

    Qi et al. [21] in bubble column reactors.The flow pattern falls within

    the regime between the homogeneous bubbly flow and the transi-

    tional flow [21], where liquid starts to form small local circulations

    while there are no vortices in the gas phase.

    4.2. Distribution ofgas bubbles and solid particles

    The gas flow pattern has an impact on the gas holdup distri-

    bution in the reactor, while the gas velocity plays a significant

    role as well. Gas bubbles along the movement pathways rise faster

    than those in other regions, which means less bubbles are trapped

    inside the reactor. An asymmetric gas holdup distribution profile is

    obtained from the three dimensional simulation (Fig. 4a). It can be

    seen that the highest gas holdup regions on the right side ofFig. 4a

    are close to the light tube, however these regions as shown in the

    black regions in Fig.4a are discontinuously presented in the vertical

    plane. The discontinuous regions are due to the spiral movement

    ofgas bubble swarm [5052]. In the right side, a relative high gas

    holdup region is located along the movement pathways ofgas bub-

    ble swarm. Interestingly, there is quitelow gas holdup in the regionclose to the central tube in the upper part of the reactor and this

    region may potentially decrease the degrading efficiency ofthe col-

    umn. In the left side of Fig. 4a, a completely different gas holdup

    distribution is presented. In the middle ofthe reactor, few gas bub-

    bles are trapped inside this region. However, gas holdup is much

    higher in the upper part ofthe reactor.

    The gas holdup at an axial position of0.2 m as referred in Fig. 4a

    is plotted along the radial direction in Fig. 4b. It can be seen that

    at higher gas velocity (0.054 m s1), a sharp drop in gas holdup is

    seen in the central region; while the gas holdup values increase in

    the middle ofthe reactor between the central light tube and the

    reactor wall. The increase can be explained by bubbles trapped in

    the liquid vortex illustrated in Fig. 3. By decreasing the gas bubble

    velocity,the gas holdup becomes morehomogeneous while the fewgas bubbles are trapped inside the reactor.

    In contrast to the heterogeneous distribution of gas bubbles

    inside the reactor, the solid catalysts are distributed uniformly in

    themajorityofthe reactor, except in the lower andupper part ofthe

    reactor (Fig. 5a), which is corresponding to the higher gas holdup

    region (Fig. 4a). The lower concentration of solid catalyst in this

    region may decrease the degradation efficiency. By reducing the

    aeration rate to one tenth ofthe original value, the solids distribu-

    tion is homogeneously in the reactor (Fig. 5b). Because at low gas

    flow velocities, the gas bubbles move up in a homogeneous regime,

    and gas hold up in the reactor is ignorable. In this context, the solid

    distribution is not affected by the gas bubble movement. The com-

    parison between thehigher andlower flow rates can be further seen

    from Fig. 5c, where the solids concentration is plotted along theaxial direction at the radial point ofr/R= 0.27. It can be found that

    smaller gas inlet velocity, but greater than particle terminal set-

    tling velocity,can makemore uniformlydistributed photocatalysts,

    which will be ofbenefit to the organic degradation process. How-

    ever,lower gas flow rates can leadto agreat reductionin gas holdup

    or dissolved oxygen concentration. Furthermore, Chong et al. [10]

    estimated the minimum aeration rate was 5 dm3/min (0.018 m s1)

    for the critical solids loading of6 g/dm3. Given the photocatalysts

    solution is quite dilute, a higher flow rate can be used to keep all

    particles in homogeneous suspension and then a lower flow rate

    is used to maintain the homogeneous suspension. However as the

    dissolved oxygen plays an important role in TiO2 photocatalytic

    reaction to assure sufficient electron scavengers present to trap the

    excited conduction-band electron from recombination[10] an opti-

    Fig. 4. Volume fraction profiles ofgas phase. (a) Simulated gas holdup. (b) Radial

    profile ofgas holdup under different gas inlet velocities (z=0.2 m).

    mized gas flow rate should be used to counteract negative effects

    ofheterogeneous solid distribution.

    4.3. Radiation intensity

    The gas holdup and catalyst distribution affect the degrada-

    tion oforganic pollutants, while the light intensity distribution

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    8/12

    N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495 91

    Fig. 5. Volume fraction profiles ofsolid phase. (a) Normalized solid concentration (ug=0.036 m s1). (b) Normalized solid concentration (ug=0.0036 m s

    1). (c) Normalized

    solid concentration under different air inlet velocity (r/R= 0.27).

    inside the reactor plays a significant role in the degradation as

    well. The light incident radiation profile in the reactor (Fig. 6) can

    be obtained by solving the radiation transport equation (Eq. (28))

    in CFX 12.0. As the solid particles are nearly uniformly distributed

    insidethe reactor (Fig.5b) andthe concentration ofparticlesis quite

    low, the adsorption and scattering coefficients are assumed to be

    proportionalto the catalyst loading [13]. The incident radiationdis-

    tribution is characterized by a decrease radially from a maximum

    near the UV light tube to a minimum at the outer wall ofthe reac-

    tor, which is in good agreement with reports by Pareek et al. [13].

    Because the distribution is nearly identical along the axial and cir-

    cumferential directions, a simple expression ofG related to rcan be

    derived as

    G(r) =11

    (r/0.0745)2

    (33)

    The sharp drop in incident radiation away from the central light

    tube leads to a rapid decrease in degradation efficiency in the pho-

    toreactor. A better design ofphotoreactor is required to achieve

    uniform distribution ofthe light incident radiation, such as intro-

    ducing multiple lamps in the reactor [13]. This problem can be also

    alleviated by increasing the total irradiated surface area ofcatalyst

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    9/12

    92 N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495

    1.00.80.60.40.20.0

    0

    100

    200

    300

    400

    500

    RadiationIn

    tensity(Wm

    -2)

    r/R (-)

    light tube

    Fig. 6. Radial profile ofradiation intensity.

    per unit volume, for example, utilization ofslurry photocatalytic

    reactor [11].

    4.4. Degradation oforganicpollutants

    The organic degradation process can be described in Fig. 7. The

    organic molecules are first transferred to the boundary ofthe cat-

    alyst particles surface by convective flow, and then are adsorbed

    onto the surface of the catalyst particles. Those molecules are

    finally degraded through the photocatalysis. The degraded prod-

    ucts are released from the catalyst surface and transferred back

    to the bulk liquid stream or into the gas stream. Overall degrada-

    tion rate depends on the transfer rate ofthe pollutant molecules by

    mass-transfer mechanisms, and the degradation rate. The rate of

    mass transfer has been characterized by CFD simulations and the

    detailed information on the gas holdup or oxygen concentration,

    catalyst concentration, and incident radiation has been obtained.

    The rate ofdegradation,however, requires thedegradationkinetics.

    Currently there is no kinetics model available for this photocat-

    alytic reactor. Two simplified models have been used in this work

    as described below.

    The first case is to simulate the organic degradation under the

    same experimental conditions [10]. In this case,the organic species

    are adsorbed onto the catalyst surface for half an hour before the

    light source is introduced.The photocatalytic reaction occurs on the

    surface ofTiO2 nanoparticles where enough photons from the light

    source are received (So,l = 0 in Eq. (29a)). New organic molecules

    are constantly adsorbed onto the catalyst surface as soon as the

    molecules are consumed during degradation process. After 4 h, the

    distribution of organic concentration in the liquid phase is pre-

    sented in Fig.8a. It can be seen that most oforganic molecules have

    beendepletedin the regions close to the light tube.Bycloselyexam-

    ining the change in organic concentration at the axial position of

    0.2 m, theorganic molecules in the area close to the central tube are

    degraded faster than those in the area close to the reactor wall. This

    can be due to much higher incident radiation in the central region.

    However,the concentration oforganic pollutants increasessteadily

    along the radial direction and reaches a plateau when approach-

    ing the reactor wall region (Fig. 8b), resulting from a combination

    of incident radiation and particle concentration. Although Fig. 6

    shows that the incident radiation has a gradual decline away from

    the central tube, andthe absolute reductionrate seems to be minor.

    While the solidconcentration at a gas flow rate of10 dm3/min has asteady increase away from the central tube (Fig. 5c) and more solid

    particlescan result in increasing surface area ofoxidation platform,

    contributing to the degradation oforganic pollutants in the region

    awayfrom the central tube.The results demonstrate thatthe hydro-

    dynamics andthe incidentradiationdistributions in the multiphase

    photoreactors should be integrated with photocatalytical kinetics.

    Thus, we can have a better understanding ofthe degradation pro-

    cess. Herein, we compared the simulated averaged global organic

    concentration values with experimental data, as shown in Fig. 8c.

    At a high solid loading, the degradation trend is well predicted. A

    minor deviation between the simulation results and experimen-

    tal data was found in the initial degradation process at a low solid

    loading. The deviation is reduced and becomes negligible after 4 h.

    In the early stage of photocatalytic reaction when the organicspecies is adsorbed onto the catalyst surface, our simulation is

    based on the assumption that the reaction rate is much faster than

    (a)

    (c)

    (b)

    is the pollutant molecules, is the degraded products, is the catalysts and the big circle represent the

    solid particles.

    (d) (e) (f)

    Fig. 7. The process ofpollutant degradation. (a) Pollutant molecule in the bulk flow; (b) pollutant molecule is transferred to the catalyst surface; (c) pollutant molecule is

    absorbed onto the catalyst surface; (d) molecule is reacted in the presence ofcatalyst; (e) pollutant molecule is fullyconvertedto products; (f) products are released into the

    main stream.

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    10/12

    N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495 93

    Fig. 8. Normalized concentration oforganic in liquid phase. (a) Effect ofTiO2/K photocatalyst loading on organic degradation rate. (b) Simulated organic normalized con-

    centration. (c) Radial distribution oforganic normalized concentration at axial position of0.2 m. (d) Simulated organic degradation rate at solid loading of6 g dm3 (case

    2).

    the transfer process. In such a case, there is no accumulation of

    organic pollutants in the solid phase, i.e. (/t)(ssYo,s) = 0. There-fore, So,s =o,ls, or the amount of pollutants is transferred from

    liquid to solid ( in the Eq. (29a)) should be equal to that which

    is degraded in Eq. (31a). The degradation curve presented in Fig. 8d

    shows that the degradation process is much faster than the first

    case. 70% of organic pollutants have been mineralized in 2 h. By

    doubling the incident radiation ofthe original light source, 90% of

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    11/12

    94 N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495

    organic pollutants have been degraded within less than one hour,

    (data not shown). Our results revealed that the fluid dynamics and

    incident radiation distribution greatly affect the degradation pro-

    cess. We can see the mass transfer from liquid to solid phase is akey

    controlling mechanism in this photocatalyst reaction. Therefore,

    further design on improving the fluid dynamics, incident radiation

    distribution andmass transfer will be required basedon the current

    model.

    The photocatalytic degradation process involves a range of

    factors: incident radiation, contaminants and their concentra-

    tion, dissolved oxygen, temperature, pH and catalysts loading.

    Traditionally, a lump kinetic model has been used to interpret

    experimental data without consideration of dynamic changes in

    liquid holdup, incident radiation, catalyst distribution and organic

    pollutant concentration. This model may not reveal the actual

    governing mechanism for photocatalytic degradation. The above

    two cases are referred to two possible degradation mechanisms

    in the photo-assisted degradation process. The combined CFD

    and kinetic models can be used for investigating more possi-

    ble mechanisms. By comparing the predictions and experimental

    results, the actual mechanism ofthe photocatalytic reactionmay be

    determined.

    5. Conclusion

    Decontamination of polluted waters has become increasingly

    important and photocatalysis by titania impregnated kaolin-

    ite nano-photocatalyst in an annular slurry photoreactor is

    attractive as an environmentally friendly and effective process.

    Hydrodynamic characteristics, incident radiation distribution and

    degradation kinetics were simulated by employment of3D tran-

    sient CFD models in an annular photocatalytic bubble column

    reactor using an Eulerian multi-fluid approach. Simulation results

    of gas holdup, liquid flow patterns as well as organic concentra-

    tions in the liquid phase are in good agreement with corresponding

    experimental measurements from our lab or literature reports. The

    results revealed that organic degradation rate is highly dependenton the local fluid hydrodynamics and incident radiation distribu-

    tion.

    The 3D transient CFD models coupled with the kinetics model

    are able to capture the global fluid dynamics and light distri-

    bution in the three-phase photocatalytic bubble column reactor,

    and provide more detailed local information on the gas holdup,

    solids concentration as well as organic pollutants concentra-

    tion. Two possible controlling mechanisms for photocatalytic

    degradation of organic pollutants are investigated. The results

    demonstrate that CFD combined with kinetics models can be

    applied to elucidate the governing mechanisms in the process.

    These models can be used for optimizing the design to accel-

    erate the degradation process, as well as facilitating scale-up

    strategies.

    Acknowledgements

    Financial support from the National Natural Science Founda-

    tion ofChina (No 51076043) and the Major State Basic Research

    Development Program ofChina (973 Program, 2009CB219801) is

    gratefully acknowledged. Ms. Qi would like to acknowledge finan-

    cial support from the China Scholarship Council (CSC) and the

    University ofAdelaide, Australia. We thank for Dr Meng Nan Chong

    and the co-workers research progress generatedin Water Environ-

    ment Biotechnology Laboratory at the University ofAdelaide. We

    would like to acknowledge Dr Kenneth Davey for his proof-reading

    the manuscript.

    References

    [1] S. Malato, P. Fernndez-Ibnez, M.I. Maldonado, J. Blanco, W. Gernjak, Decon-

    tamination and disinfection ofwater by solar photocatalysis: recent overviewand trends, Catal. Today 147 (2009) 159.

    [2] W. Xing, H.H. Ngo, S.H. Kim, W.S. Guo, P. Hagare, Physico-chemical processesfor landfill leachate treatment: experiments and mathematical models, Sep.Sci. Technol. 43 (2008) 347361.

    [3] A.S. Erses, T.T. Onay, O. Yenigun, Comparison ofaerobic and anaerobic degra-dation ofmunicipal solid waste in bioreactorlandfills,Bioresource Technol. 99

    (2008) 54185426.[4] K. Pirkanniemi, M. Sillanp, Heterogeneous water phase catalysis as an envi-

    ronmental application: a review, Chemosphere 48 (2002) 10471060.[5] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Environmental appli-

    cations ofsemiconductor photocatalysis, Chem. Rev. 95 (1995) 6996.[6] Y. Ohko, K.I. Iuchi, C. Niwa, T. Tatsuma, T. Nakashima, T. Iguchi, Y. Kubota, A.

    Fujishima, 17 beta-estrodial degradation by TiO2 photocatalysis as means ofreducing estrogenic activity, Environ. Sci. Technol. 36 (2002) 41754181.

    [7] D.F. Ollis, E. Pelizzetti, N. Serpone, Photocatalyzed destruction ofwater con-taminants, Environ. Sci. Technol. 25 (1991) 15221529.

    [8] J.M. Herrmann,Heterogeneousphotocatalysis: fundamentals and applicationsto the removal ofvarious typesofaqueous pollutants,CatalysisToday 53 (1999)115129.

    [9] H.S. Fogler, ElementsofChemical Reaction Engineering, thirded., Prentice HallPTR, NewJersey, 1998.

    [10] M.N. Chong, S. Lei, B. Jin, C.P. Saint, C.W.K. Chow, Optimisation ofan annularphotoreactor process for degradation of Congo Red using a newly synthe-sizedtitania impregnated kaolinite nano-photocatalyst,Sep. Purif. Technol. 67(2009) 355363.

    [11] M.N. Chong, B. Jin, H.Y. Zhu, C.W.K. Chow, C.P. Saint, Application ofH-titanatenanofibers for degradation of Congo Red in an annular slurry photoreactor,Chem. Eng. J. 150 (2009) 4954.

    [12] M.N. Chong, B. Jin, C.W.K. Chow, C.P. Saint, A new approach to optimise anannular slurryphotoreactorsystemfor the degradation ofCongoRed:statisticalanalysis and modelling, Chem. Eng. J. 152 (2009) 158166.

    [13] V. Pareek, S. Chong, M. Tad, A.A. Adesina, Light intensity distribution in het-erogenous photocatalytic reactors, Asia-Pac. J. Chem. Eng. 3 (2008) 171201.

    [14] J.J. Pignatello,Dark and photoassistedFe3+-catalyzeddegradationofchlorophe-noxy herbicides by hydrogen peroxide, Environ. Sci. Technol. 26 (1992)944951.

    [15] A.Safarzadeh-Amiri,J.R. Bolton,S.R. Cater, Theuse ofiron in advanced oxidationprocesses, J. Adv.Oxid. Technol. 1 (1996) 1826.

    [16] T. Wei, C. Wan, Heterogeneous photocatalytic oxidation ofphenol with tita-nium dioxide powders, Ind. Eng. Chem. Res. 30 (1991) 12931300.

    [17] A.V. Emeline, V.K. Ruabchuk, N. Serpone, Dogmas and misconceptions in het-ergeneous photocatalysis. Some elightened reflections, J. Phys. Chem. B 109(2005) 1851518521.

    [18] H. Zhang, P. Ayazi-Shamlou, Characterisation of the engineering flow in aminiaturised bioreactor by computational fluid dynamics, Chem. Res. ChineseUniversities 20 (2004) 489493.

    [19] J.E. Duran, M. Mohseni, F. Taghipour, Computational fluid dynamics modelingof immobilized photocatalytic reactors for water treatment, AIChE J. 12399(2010) 113, doi:10.1002/aic.

    [20] M.K. Inglesby, S.H. Zeronian, Diffusion coefficients for direct dyes in aqueousand polar aprotic solvents by the NMR pulsed-field gradient technique, DyesPigments 50 (2001) 311.

    [21] N.N. Qi, K. Zhang, Q.T. Song, X.G. Gao, H. Zhang, J.J . Lu, 2D and 3D simulationoffluid dynamics in the bubble reactor for liquid fuel synthesis: comparisonsagainst experiments, Energy Source Part A 31 (2009) 15981611.

    [22] K. Zhang,H. Zhang,J. Lovick, J.Y. Zhang,B.J.Zhang, Numerical computationandexperimental verification ofthejet region in a fluidized bed, Ind. Eng. Chem.Res. 41 (2002) 36963704.

    [23] K. Zhang, N.N. Qi, J .Q. Jin, C.X. Lu, H. Zhang, Gas holdup and bubble dynamicsin a three-phase internal loop reactor with external slurry circulation, Fuel 89(2010) 13611369.

    [24] C.X. Lu, N.N. Qi, K. Zhang, J.Q. Jin, H. Zhang, Experiment and CFD simulation

    on gas holdup characteristics in an internal loop reactor with external liquidcirculation, Int. J. Chem. React. Eng. 7 (2009) A3.

    [25] D.A. Sozzi, F. Taghipour, UV reactor performance modeling by Eulerian andLagrangian methods, Environ. Sci. Technol. 40 (2006) 16091615.

    [26] D.Wiemann,D. Mewes,Calculationofflow fieldsin two andthree-phasebubblecolumns considering mass transfer,Chem. Eng. Sci. 60 (2005) 60856093.

    [27] T.A. Bartrand,B. Farouk, C.N. Haas,Countercurrent gas/liquid flow and mixing:implications for water disinfection, Int. J. Multiple Flow 35 (2009) 171184.

    [28] R. Pannee rs elvam, S. S avit hr iand, G .D. S ur ender, CFD m odeling of gas-liquid-solid mechanically agitated contactor, Chem. Eng. Res. Des. 86 (2008)13311344.

    [29] R. Clift, J.R. Grace, M.E. Weber, Bubbles, Drops and Particles, Academic Press,New York, 1978.

    [30] ANSYS Incorporated, ANSYS CFX-Solver Release 12.0, Canonsburg PA, USA,2009.

    [31] M.Lopez de Bertodano,Turbulent bubblyflow in a triangularduct, Ph.D.Thesis,Rensselaer Polytechnic Institute, Troy New York, 1991.

    [32] J.B. Joshi,Computational flow modelling and design ofbubblecolumn reactors,Chem. Eng. Sci. 56 (2001) 58935933.

  • 7/29/2019 CFD de tratamiento de aguas.pdf

    12/12

    N.Qi et al. / Chemical Engineering Journal 172 (2011) 8495 95

    [33] Y.Sato,M. Sadatomi,Momentum andheat transfer in two-phasebubble flow-I,Int. J. Multiple Flow 7 (1981) 167177.

    [34] G. Spadoni, E. Bandini, F. Santarelli, Scattering effects in photosensitized reac-tions,Chem. Eng. Sci. 33 (1978) 517524.

    [35] A.E. Cassano, C.A. Martin, R.J. Brandi, O.M. Alfano, Photoreactor analysisand design-fundamentals and applications, Ind. Eng. Chem. Res. 34 (1995)21552201.

    [36] A.E. Cassano, O.M. Alfano, Reaction engineering ofsuspended solid heteroge-neous photocatalytic reactors, Catal. Today 58 (2000) 167197.

    [37] R.L.Romero, O.M.Alfano,A.E.Cassano,Cylindricalphotocatalyticreactors. Radi-ation absorption and scattering effects produced by suspended fine particles

    in an annular space,Ind. Eng. Chem. Res. 36 (1997) 30943109.[38] M.Pasquali,F. Santarelli, J.F.Porter, P.L.Yue, Radiativetransferin photocatalytic

    systems, AIChE J. 42 (1996) 532537.[39] F. Denny,J. Scott,V. Pareek, G.D. Peng, R. Amal, CFDmodelingfor a TiO2-coated

    glass-bead photoreactor irradiated by optical fibres: photocatalytic degrada-tion ofoxalic acid, Chem. Eng. Sci. 4 (2009) 16951706.

    [40] D.C. Apters, L.S. Fan, Solid-liquid mass transfer in a gas-liquid-solid fluidizedbed, Chem. Eng. Sci. 41 (1986) 107115.

    [41] P.C. Chen, M.P. Dudukovic, J. Sanyal, Three-dimensional simulation of bub-ble column flows with bubble coalescence and breakup, AIChE J. 51 (2005)696712.

    [42] N.T. Padial , W .B . Vander He yden , R.M. Rauenzah n, S .L. Yarbro , Thr ee -dimensional simulationofa three-phase draft-tubebubble column, Chem. Eng.Sci. 55 (2000) 32613273.

    [43] V. Michele, D.C. Hempel, Liquid flow and phaseholdup-measurementand CFDmodelling for two- andthree-phase bubblecolumns, Chem. Eng. Sci. 57 (2002)18991908.

    [44] X.Q. Jia, J.P.Wen,H.L. Zhou, W.Feng, Q. Yuan, Local hydrodynamics modellingofa gas-liquid-solidthree-phasebubblecolumn,AIChE J. 53 (2007) 22212231.

    [45] D.A. Sozzi, F. Taghipour, UV reactor performance modelling by Eulerian andLagrangian methods, Environ. Sci. Technol. 40 (2006) 16091615.

    [46] A. Sokolichin, G. Eigenberger, Applicability of the standard k- turbulencemodel to the dynamic simulationofbubblecolumns: Part I. Detailed numericalsimulations, Chem. Eng. Sci. 54 (1999) 22732284.

    [47] M.E. Daz, A. Iranzo, D. Cuadra, R. Barbero, F.J. Montes, M.A. Galn, Numerical

    simulation of the gasliquid flow in a laboratory scale bubble column Influ-ence ofbubble size distribution and non-drag forces, Chem. Eng. J. 139 (2008)363379.

    [48] L.S.Fan, Gas-Liquid-Solid FluidizationEngineering, Butterworths,Boston, 1989.[49] R.C. Chen, J. Reeseand, L.S. Fan, Flow structure in a three-dimensional column

    and three-phase fluidized bubble bed, AIChE J. 40 (1994) 10931104.[50] G.Q. Yang, B. Du, L.S. Fan, Bubble formation and dynamics in gasliquidsolid

    fluidizationa review, Chem. Eng. Sci. 62 (2007) 227.[51] G.M.C. Glover, S.C. Generalis, Gasliquidsolid flow modelling in a bubble col-

    umn, Chem. Eng. Process. 43 (2004) 117126.[52] W. Chen, T. Hasegawa, A. Tsutsumi, K. Otawara, Scale-up effects on the time-

    averaged anddynamic behavior in bubble column reactors,Chem. Eng. Sci. 56(2001) 61496155.