12.07.2015 Views

Modeling of Bubble Column Reactors: Progress and Limitations

Modeling of Bubble Column Reactors: Progress and Limitations

Modeling of Bubble Column Reactors: Progress and Limitations

SHOW MORE
SHOW LESS
  • No tags were found...

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Ind. Eng. Chem. Res. 2005, 44, 5107-51515107<strong>Modeling</strong> <strong>of</strong> <strong>Bubble</strong> <strong>Column</strong> <strong>Reactors</strong>: <strong>Progress</strong> <strong>and</strong> <strong>Limitations</strong>Hugo A. Jakobsen,* Håvard Lindborg, <strong>and</strong> Carlos A. DoraoDepartment <strong>of</strong> Chemical Engineering, Norwegian University <strong>of</strong> Science <strong>and</strong> Technology,NTNU, Sem Sæl<strong>and</strong>s vei 4, NO-7491 Trondheim, Norway<strong>Bubble</strong> columns are widely used for carrying out gas-liquid <strong>and</strong> gas-liquid-solid reactions ina variety <strong>of</strong> industrial applications. The dispersion <strong>and</strong> interfacial heat- <strong>and</strong> mass-transfer fluxes,which <strong>of</strong>ten limit the overall chemical reaction rates, are closely related to the fluid dynamics<strong>of</strong> the system through the liquid-gas contact area <strong>and</strong> the turbulence properties <strong>of</strong> the flow.There is thus considerable interest, within both academia <strong>and</strong> industry, to improve the limitedunderst<strong>and</strong>ing <strong>of</strong> the complex multiphase flow phenomena involved, which is preventing optimalscale-up <strong>and</strong> design <strong>of</strong> these reactors. In this paper, the progress reported in the literature duringthe past decade regarding the use <strong>of</strong> averaged Eulerian multifluid models <strong>and</strong> computationalfluid dynamics (CFD) to model vertical bubble-driven flows is reviewed. The limiting steps inthe model derivation are the formulation <strong>of</strong> proper boundary conditions, closure laws determiningturbulent effects, interfacial transfer fluxes, <strong>and</strong> the bubble coalescence <strong>and</strong> breakage processes.Examples <strong>of</strong> both classical <strong>and</strong> more recent modeling approaches are described, evaluated, <strong>and</strong>discussed. Physical mechanisms <strong>and</strong> numerical modes creating bubble movement in the radialdirection are outlined. Special emphasis is placed on the population balance modeling <strong>of</strong> thebubble coalescence <strong>and</strong> breakage processes in two-phase bubble column reactors. The constitutiverelations used to describe the bubble-bubble <strong>and</strong> bubble-turbulence interactions, the bubblecoalescence <strong>and</strong> breakage criteria, <strong>and</strong> the daughter size distribution models are discussed witha focus on model limitations. The dem<strong>and</strong> for amplified modeling, more accurate <strong>and</strong> stablenumerical algorithms, <strong>and</strong> experimental analysis providing data for proper model validation isstressed.1. IntroductionChemical reaction engineering (CRE) has emerged asa methodology that quantifies the interplay betweentransport phenomena <strong>and</strong> kinetics on a variety <strong>of</strong> scales<strong>and</strong> allows the formulation <strong>of</strong> quantitative models forvarious measures <strong>of</strong> reactor performance. The abilityto establish such quantitative links between measures<strong>of</strong> reactor performance <strong>and</strong> input <strong>and</strong> operating variablesis essential in optimizing the operating conditionsin manufacturing, in determining proper reactor design<strong>and</strong> scale-up, <strong>and</strong> in correctly interpreting data inresearch <strong>and</strong> pilot-plant work.A starting point for reaction engineers is the formulation<strong>of</strong> a reactor model for which the basis is themicroscale species mass <strong>and</strong> enthalpy balances. Forpractical applications, the direct solution <strong>of</strong> these equationsis excessively costly, <strong>and</strong> simplifications or averagerepresentations are usually introduced.The choice <strong>of</strong> averages (e.g., global reactor volume,cross-sectional area, or length) over which the balanceequations are integrated (averaged) determines the level<strong>of</strong> sophistication <strong>of</strong> the reactor model. It is very commonin tubular reactors to have flow predominantly in onespatial direction, say, z. The major gradients then occurin that direction. For many cases, then, the crosssectionalaverage values <strong>of</strong> concentration <strong>and</strong> temperatureare used instead <strong>of</strong> the local values. In this way,a one-dimensional dispersion model is obtained. If theconvective transport is completely dominant over the* To whom correspondence should be addressed. E-mail:jakobsen@chemeng.ntnu.no. Tel.: + 47 73594132. Fax: + 4773594080.diffusive transport, the diffusive term can be neglected.The resulting equations are denoted the ideal plug-flowreactor (PFR) model. When the entire reactor can beconsidered to be uniform in both concentration <strong>and</strong>temperature (i.e., because <strong>of</strong> very large dispersioncoefficients), one can neglect gradients in all spatialdirections <strong>and</strong> integrate the equations globally over allspatial dimensions (assuming convective flows at theboundaries), leading to the ideal reactor model <strong>of</strong> thecontinuous stirred tank reactor (CSTR). For morecomplex flow patterns, more elaborated <strong>and</strong> completemodels are required where the flow fields are describedvia the solution <strong>of</strong> the Navier-Stokes equations. Theunderst<strong>and</strong>ing <strong>of</strong> the complex flow phenomena involvedas well as the solution <strong>of</strong> these vector equations makethe problem much more difficult to analyze within theconstraint <strong>of</strong> reasonable costs <strong>and</strong> efforts. In these cases,the full set <strong>of</strong> governing equations can be averaged overlocal but finite spatial <strong>and</strong> temporal scales to obtainformulations that are solved with feasible space <strong>and</strong>time resolutions. The latter type <strong>of</strong> models can also beobtained using other local averaging procedures aswell.Two-phase <strong>and</strong> slurry bubble columns are widely usedin the chemical <strong>and</strong> biochemical industries for carryingout gas-liquid <strong>and</strong> gas-liquid-solid (catalytic) reactionprocesses. 1-6 The historical development <strong>of</strong> bubblecolumn modeling was discussed in a recent paper byDudukovic. 4 In the past, the axial dispersion model wascommonly applied for both the gas <strong>and</strong> liquid (slurry)phases in bubble columns. 2,7 The balance equationsdetermining the liquid- (slurry-) phase axial dispersionmodel can be written as10.1021/ie049447x CCC: $30.25 © 2005 American Chemical SocietyPublished on Web 01/20/2005


5108 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005d(F L v S z,L ω i,L )<strong>and</strong>dz) ɛ L F L D z,Ld 2 ω i,Ldz 2/+ k L a(F L,i-F L,i ) +ɛ L∑γ i,r R L,r (1)rdT L d 2 T LSF L C P,L v z,Ldz ) ɛ L λ z,L + h dz 2 W a W (T W - T L ) +ɛ L∑(-∆H r )R r,L (2)rIn these balance equations, all terms should bedescribed at the same level <strong>of</strong> accuracy. It certainly doesnot pay to have the finest description <strong>of</strong> one term in thebalance equations if the others can only be very crudelydescribed. Current dem<strong>and</strong>s for increased selectivity<strong>and</strong> volumetric productivity require more precise reactormodels <strong>and</strong> also force reactor operation to churn turbulentflow, which, to a great extent, is unchartedterritory. An improvement in accuracy <strong>and</strong> a moredetailed description <strong>of</strong> the molecular-scale events describingthe rate <strong>of</strong> generation term in the heat- <strong>and</strong>mass-balance equations has, in turn, pushed forward aneed for a more detailed description <strong>of</strong> the transportterms (i.e., in the convection/advection <strong>and</strong> dispersion/conduction terms in the basic mass <strong>and</strong> heat balances).Experimental evidence shows that the liquid axialvelocity is far from being flat <strong>and</strong> independent <strong>of</strong> theradial space coordinate, 2 <strong>and</strong> the use <strong>of</strong> a cross-sectionalaverage velocity variable seems not to be sufficient. Theback-mixing induced by the global liquid flow patternhas commonly been taken into account by adjusting theaxial dispersion coefficient accordingly. However, eventhough (slurry) bubble column performance <strong>of</strong>ten canbe fitted with an axial dispersion model, decades <strong>of</strong>research have failed to produce a predictive equationfor the axial dispersion coefficient.Hence, research is in progress to quantify theseparameters based on first principles (e.g., see Jakobsenet al., 8 Joshi, 9 Rafique et al., 10 <strong>and</strong> references therein).However, despite their simple construction, the fluiddynamics observed in these columns is very complex.Even though CFD modeling concepts have been extendedover the past two decades in accordance withthe rapid progress in computer performance, the modelcomplexity required to resolve all <strong>of</strong> the importantphenomena in these systems is still not feasible withinreasonable time limits. The multifluid model is foundto represent a trade<strong>of</strong>f between accuracy <strong>and</strong> computationalefforts for practical applications.Unfortunately, the present models are still on a levelaiming at reasonable solutions with several modelparameters tuned to known flow fields. For predictivepurposes, these models are hardly able to predictunknown flow fields with a reasonable degree <strong>of</strong> accuracy.It appears that CFD evaluations <strong>of</strong> bubblecolumns by use <strong>of</strong> multidimensional multifluid modelsstill have very limited inherent capabilities to fullyreplace the empirical-based analyses (i.e., in the framework<strong>of</strong> axial dispersion models) in use today. 11,12 Aftertwo decades <strong>of</strong> performing fluid dynamic modeling <strong>of</strong>bubble columns, it has been realized that there iscertainly a limit to how accurate one will be able informulating closure laws using the Eulerian framework.A severe problem is that, although the Eulerian modelingprospects are not outst<strong>and</strong>ing, no other conceptsavailable are favorable for the purpose <strong>of</strong> predictingbubble column flows. A more relevant question is thus:What can be achieved by adopting the Eulerian multifluidmodeling framework for the purpose <strong>of</strong> describingchemical processes operated in bubble column reactors?The authors intend the following summary to providecertain guidelines for the discussion but, <strong>of</strong> course, nocomplete answer to this question.Fluid Dynamic <strong>Modeling</strong>. Considering modeling infurther detail, the general picture from the literatureis that the forces acting on the dispersed phase are 8inertia, gravity, buoyancy, viscous, pressure, lift, wall,turbulent stress, turbulent dispersion, steady-drag, <strong>and</strong>added-mass forces. In the latest papers (i.e., publishedafter the review by Jakobsen et al. 8 ) performing 3Dsimulations, the force balances in vertical bubbly flowswere found to be determined by only a few <strong>of</strong> theseforces. The axial component <strong>of</strong> the momentum equationfor the gas phase is dominated by the pressure <strong>and</strong>steady-drag forces only, indicating that algebraic slipmodels might be sufficient, 13-15 whereas most multifluidmodels also retain the inertia <strong>and</strong> gravity (buoyancy)terms. The axial momentum balance for the liquid phaseconsiders the inertia, turbulent stress, pressure, steadydrag,<strong>and</strong> gravity forces. Only pressure <strong>and</strong> buoyancyforces are acting on a motionless bubble in a liquid atrest. In the radial <strong>and</strong> azimuthal directions, the forcebalances generally includes the steady-drag force, i.e.,a force that opposes motion, whereas the pertinentforces causing motion are more difficult to define. In avery short inlet zone, the wall friction is likely to inducea radial pressure gradient that pushes the gas bubblesaway from the wall, whereas a few column diametersabove the inlet, the radial pressure gradient vanishes.It is still an open question whether this pressuregradient is sufficient to determine the phase distributionsobserved in these systems. It is expected that thepresence <strong>of</strong> the wall induce forces that act on thedispersed particles farther from the inlet, but there isno general acceptance on the physical mechanisms <strong>and</strong>formulations <strong>of</strong> these forces. It is also a matter <strong>of</strong>discussion whether this wall effect should be taken intoaccount indirectly through the liquid wall friction or,in view <strong>of</strong> the model averaging performed, directly as aforce in the gas-phase equations. The bulk lift forces donot induce any radial bubble movement without aninitial velocity gradient, so other forces are importantin the initial phase <strong>of</strong> the flow pattern development. Inthe generalized drag formulation, the interfacial couplingis expressed as a linear sum <strong>of</strong> independent forces;this point <strong>of</strong> view is probably not valid when the voidfraction exceeds a few percent. Moreover, the parametrizationsused for the coefficients occurring in theinterfacial closures vary significantly, especially athigher void fractions. Single-particle drag, added-mass,<strong>and</strong> lift coefficients are most frequently used, whereasswarm corrections have been included in some codes.For higher void fractions, other rather empirical correctionshave been introduced as well. 16The flow <strong>of</strong> the continuous phase is considered to beinitiated by a balance between the interfacial particlefluidcoupling <strong>and</strong> wall friction forces, whereas the fluidphaseturbulence damps the macroscale dynamics <strong>of</strong> thebubble swarms, thereby smoothing the velocity <strong>and</strong>volume fraction fields. Temporal instabilities induced


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5109Figure 1. Classification <strong>of</strong> regions accounting for the macroscopic flow structures: left, 2D bubble column; 21 right, 3D bubble column. 23by the fluid inertia terms create inhomogeneities in theforce balances. Unfortunately, proper modeling <strong>of</strong> turbulenceis still one <strong>of</strong> the main open issues in gas-liquidbubbly flows, <strong>and</strong> this flow property can significantlyaffect both the stresses <strong>and</strong> the bubble dispersion. 15It was shown by Svendsen et al., 17 among others, thatthe time-averaged experimental data on the flow patternin cylindrical bubble columns apparently becomesclose to axisymmetric. Fair agreement between experimental<strong>and</strong> simulated results are generally obtained forthe steady velocity fields in both phases, whereas thesteady phase distribution is still a problem. Therefore,it was anticipated that 2D axisymmetric simulationswould capture the pertinent time-averaged flow patternneeded for the analysis <strong>of</strong> many (not all) mechanisms<strong>of</strong> interest for chemical engineers. Sanyal et al. 18 <strong>and</strong>Krishna <strong>and</strong> van Baten, 19 for example, stated that 2Dmodels provide good engineering descriptions, althoughthey are not able to capture the high-frequency unsteadybehavior <strong>of</strong> the flow, <strong>and</strong> can be used for approximatelypredicting the low-frequency time-averaged flow <strong>and</strong>void patterns in bubble columns.The early 2D steady-state model proposed by Jakobsen20 was able to capture the global flow pattern fairlywell, provided that a large negative lift force wasincluded. However, after the first elaborated experimentalstudies on 2D rectangular bubble columns werepublished by Tzeng et al. 21 <strong>and</strong> Lin et al., 22 it wascommonly accepted that time-average computationscannot provide a rational explanation <strong>of</strong> the transportprocesses <strong>of</strong> mass, momentum, <strong>and</strong> energy between thebubbles <strong>and</strong> liquid. The experimental data obtainedwere analyzed, <strong>and</strong> sketches <strong>of</strong> their interpretations <strong>of</strong>the dynamic flow patterns in both 2D <strong>and</strong> 3D columnswere given, as shown in Figure 1. It was concluded thata proper bubble column model should consider thetransient or instantaneous flow behavior. A few yearslater, Sokolichin <strong>and</strong> Eigenberger, 24 Sokolichin et al., 25<strong>and</strong> Sokolichin <strong>and</strong> Eigenberger 26 claimed that dynamic3D models were needed to provide sufficient representations<strong>of</strong> the high-frequency unsteady behavior <strong>of</strong> theseflows. Very different dynamic flow patterns can resultin quantitatively similar long-time-averaged flow pr<strong>of</strong>iles.This limits the use <strong>of</strong> long-time-averaged flowpr<strong>of</strong>iles for validation <strong>of</strong> bubbly flow models. van denFigure 2. Lateral movement <strong>of</strong> the bubble hose in a flat bubblecolumn. 28 Reprinted with permission from Elsevier.Akker, 12 among others, questioned the early turbulencemodeling performed (or rather the lack <strong>of</strong> any) in thesestudies <strong>and</strong> argued that the apparently realistic simulations<strong>of</strong> the transient flow characteristics could benumerical modes rather than physical ones (see alsoSokolichin et al. 15 ). Insufficiently fine grids might havebeen used in the simulations, resulting in numericalinstabilities that could be erroneously interpreted asphysical ones. However, after the observation <strong>of</strong> Sokolichin<strong>and</strong> Eigenberger was reported, intensive focus onthese instability issues were made, mainly by researchersfrom the fluid dynamics community. The experimentaldata provided by Becker et al. 27,28 (see Figure2) still serve as a benchmark test <strong>and</strong> are <strong>of</strong>ten usedfor validation <strong>of</strong> dynamic flow models. The numericalinvestigations were restricted to bubbly flow hydrodynamics(i.e., no reactive systems were analyzed),where additional simplifications were made: isothermalconditions, no interfacial mass transfer, constant liquiddensity, gas density constant or depending on localpressure as described by the ideal gas law, <strong>and</strong> nobubble coalescence <strong>and</strong> breakage.However, after the dynamic flow structures wereobserved, bubble columns have generally been simu-


5110 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005lated using either 2D or 3D dynamic models for bothcylindrical 18,29-36 <strong>and</strong> rectangular 2D <strong>and</strong> 3D columngeometries. 14,15,34,35,37-43 The gas is introduced adoptingboth uniform <strong>and</strong> localized feedings at the bottom <strong>of</strong> thecolumn. The modeling <strong>of</strong> systems uniformly gassed atthe bottom is more difficult than the modeling <strong>of</strong> partlyaerated ones. Simulating systems with continuous liquidflow is also more difficult than keeping the liquid inbatch mode. Finally, it is also noted that the differentresearch groups applied both commercial codes(e.g., CFX, FLUENT, PHOENICS, CFDLIB, ASTRID,NPHASE) <strong>and</strong> several in-house codes in which theinherent choices <strong>of</strong> numerical methods, discretizations,grid arrangements, <strong>and</strong> boundary implementations varyquite widely. These numerical differences alter thesolutions to some extent, so it should not be expectedthat the corresponding simulations will provide identicalresults. An open <strong>and</strong> unified research code available forall research groups could assist by eliminating anymisinterpretations <strong>of</strong> numerical modes as physicalmechanisms <strong>and</strong> visa versa.Considering the interfacial <strong>and</strong> turbulent closures forvertical bubble-driven flows, no extensive progress hasbeen observed in the later publications. However, twodiverging modeling trends seem to emerge because <strong>of</strong>the lack <strong>of</strong> underst<strong>and</strong>ing <strong>of</strong> the phenomena involved<strong>and</strong> how to deal with these phenomena within anaverage modeling framework. One group <strong>of</strong> papersconsiders only phenomena that can be validated withthe existing experimental techniques <strong>and</strong> thus containsa minimum number <strong>of</strong> terms <strong>and</strong> effects. Papers in theother group include a large number <strong>of</strong> weakly foundedtheoretical hypothesis <strong>and</strong> relations intended to resolvethe missing mechanisms.Steady-State or Dynamic Simulations, Closures,<strong>and</strong> Numerical Grid Arrangements. Not only dynamicmodels have been adopted investigating thesephenomena. Lopez de Bertodano, 44 for example, used a3D steady finite-volume method (FVM) (PHOENICS)with a staggered grid arrangement to simulate turbulentbubbly two-phase flow in a triangular duct. In thisstudy, the lift, turbulent dispersion, <strong>and</strong> steady-dragforces were assumed to be dominant. St<strong>and</strong>ard literatureexpressions were adopted for the drag <strong>and</strong> liftforces, whereas a crude model for the turbulent dispersionforce was developed. An extended k-ɛ model,considering bubble-induced turbulence, was also developedfor the liquid-phase turbulence. It was proposedthat the shear-induced turbulence <strong>and</strong> the bubbleinducedturbulence could be superimposed. The lift forcewas found to be essential to reproducing the experimentallyobserved wall void peaks satisfactorily. Anglartet al. 45 adopted many <strong>of</strong> the same closures withina 2D steady version <strong>of</strong> the same code (PHOENICS),predicting low-void bubbly flow between two parallelplates. They found satisfactory agreement with experimentaldata when applying drag, added-mass, lift, wall(lubrication), <strong>and</strong> turbulent diffusion forces in theirstudy. The extended k-ɛ model <strong>of</strong> Lopez de Bertodano 44was applied for the liquid-phase turbulence. In a morerecent paper, Antal et al. 46 adopted very similar closuresfor 3D steady-state bubble column simulations using theNPHASE code. The NPHASE CMFD code employs anFVM on a collocated grid. A three-field multifluid modelformulation was used to simulate two-phase flow in abubble column operating in the churn-turbulent flowregime. The gas phase was subdivided into two fields(i.e., small <strong>and</strong> large bubbles) to more accuratelydescribe the interfacial momentum-transfer fluxes. Thethird field was used for the liquid phase. The modelresults were validated against a few time-averaged datasets for the liquid axial velocity <strong>and</strong> the gas volumefraction. Global flow patterns for all three fields <strong>and</strong> theoverall gas volume fractions were shown. The simulationswere in fair agreement with the experimentalobservations. Using a 2D in-house FVM code with astaggered grid arrangement, Dhotre <strong>and</strong> Joshi 47 predictedthe flow pattern, pressure drop, <strong>and</strong> heat-transfercoefficient in bubble column reactors. The model usedcontained steady-drag, added-mass, <strong>and</strong> lift forces, aswell as a reduced pressure gradient formulated as anapparent form drag. Turbulent dispersion was takeninto account by use <strong>of</strong> mass-diffusion terms in thecontinuity equations. A low-Reynolds-number k-ɛ modelwas incorporated, merely constituting a st<strong>and</strong>ard k-ɛmodel with modified treatment <strong>of</strong> the near-wall region.The turbulence model used contained an additionalproduction term accounting for the large-scale turbulenceproduced within the liquid flow field because <strong>of</strong>the movement <strong>of</strong> the bubbles. A semiempirical mechanicalenergy balance for the gas-liquid system wasimposed. The simulated results were in surprisinglygood agreement with experimental literature data onthe axial liquid velocity, gas volume fraction, frictionmultiplier, <strong>and</strong> heat-transfer coefficient.Deen et al. 40 used the lift force in addition to thesteady-drag <strong>and</strong> added-mass forces in their dynamic 3Dmodel to obtain the transverse spreading <strong>of</strong> the bubbleplume that is observed in experiments. A prescribedzero-void wall boundary was used to force the gas tomigrate away from the wall. The continuous-phaseturbulence was incorporated in two different ways,using either a st<strong>and</strong>ard k-ɛ or a VLES model. Theeffective viscosity <strong>of</strong> the liquid phase was composed <strong>of</strong>three contributions: the molecular, shear-induced turbulent,<strong>and</strong> bubble-induced turbulent viscosities. Thecalculation <strong>of</strong> the turbulent gas viscosity was based onthe turbulent liquid viscosity as proposed by Jakobsenet al. 8 These simulations were performed using thecommercial code CFX, so an FVM on a collocated gridwas employed. Sample results simulating a square 3Dcolumn at low void fractions using the 3D VLES model<strong>of</strong> Deen et al. 40 are shown in Figure 3. Krishna <strong>and</strong> vanBaten 29,35 used a steady-drag force as the only interfacialmomentum coupling in their transient 2D <strong>and</strong> 3Dthree-fluid models with an inherent prescribed <strong>and</strong> fixedbimodal distribution <strong>of</strong> the gas bubble sizes. The twobubble classes were denoted small <strong>and</strong> large bubbles.The small bubbles were in the range <strong>of</strong> 1-6 mm,whereas the large bubbles were typically in the range<strong>of</strong> 20-80 mm. An unfortunate simplification made inthe model is that there are no interactions or exchangesbetween the small <strong>and</strong> large bubble populations, as willbe discussed later. An FVM on an unstaggered grid wasused to discretize the equations (i.e., in CFX). No-slipconditions at the wall were used for both phases. A k-ɛmodel was applied for the liquid-phase turbulence,whereas no turbulence model was used for the dispersedphases. To prevent a circulation pattern in which theliquid flowed up near the wall <strong>and</strong> came down in thecore, the large bubble gas was injected on the inner 75%<strong>of</strong> the radius. The time-averaged volume fraction <strong>and</strong>velocity pr<strong>of</strong>iles calculated from the predicted 3D flowfield were in reasonable agreement with experimental


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5111Figure 3. Snapshots <strong>of</strong> the instantaneous isosurfaces <strong>of</strong> R d ) 0.04<strong>and</strong> liquid velocity fields after 30 <strong>and</strong> 35 s for the LES model. 40Reprinted with permission from Elsevier.data. Lehr et al. 32 used a similar three-fluid model(implemented in CFX), combined with a simplifiedpopulation balance model for the bubble size distribution.The simplified population balance relation usedcontained semiempirical parametrizations for the bubblecoalescence <strong>and</strong> breakage phenomena. It was concludedthat the calculated long-time-average volume fractions,velocities, <strong>and</strong> interfacial area density were in goodagreement with the experimental data. Pfleger et al. 39applied a two-fluid model using the same code (CFX) toa 2D rectangular column with localized spargers. It wasconcluded that a 3D model including the steady-dragforce <strong>and</strong> a st<strong>and</strong>ard k-ɛ model is sufficient to correctlycapture the unsteady behavior <strong>of</strong> bubbly flow with verylow gas void fractions. The dispersed phase was consideredlaminar. Sanyal et al. 18 <strong>and</strong> Bertola et al. 34 usedsimilar two-fluid models <strong>and</strong> FVMs on collocated gridarrangements (FLUENT <strong>and</strong> CFDLIB) to simulate bothcylindrical <strong>and</strong> rectangular columns, confirming theresults found by Pfleger et al. 39 It should be noted,however, that Bertola et al. 34 solved the k-ɛ turbulencemodel for both phases, whereas Sanyal et al. 18 adoptedan approach based on Tchen’s theory 48-51 to predictturbulence in the dispersed phase. Mudde <strong>and</strong> Simonin38 performed both 2D <strong>and</strong> 3D simulations <strong>of</strong> a 2Drectangular column using a similar FVM on a collocatedgrid arrangement (ASTRID). Their two-fluid modelcontained an extended k-ɛ turbulence model formulationfor the liquid-phase turbulence, as well as drag <strong>and</strong>added-mass forces. The dispersed-phase turbulence wasassumed to be steady <strong>and</strong> homogeneous <strong>and</strong> wasdescribed by the extended Tchen’s theory approachmentioned above. This code predicted reasonable highfrequencyoscillating flows only when the added-massforce was included; without this force low-frequency,almost steady flows were obtained. Using an FD scheme,Tomiyama et al. 52 reported that the transient transversemigration <strong>of</strong> bubble plumes in vessels can be wellpredicted including steady-drag, added-mass, <strong>and</strong> liftforces in their two-fluid model describing laminar flowscontaining very low void fractions (i.e., below 0.5%).Later, Tomiyama 53 <strong>and</strong> Tomiyama <strong>and</strong> Shimada 54 usedthe bubble-induced turbulence model <strong>of</strong> Sato <strong>and</strong> Sekoguchi55 <strong>and</strong> Sato et al. 56 <strong>and</strong> an extended k-ɛ model intheir work on turbulent flows. In a recent paper,Sokolichin et al. 15 concluded that the model by Sato <strong>and</strong>Sekoguchi 55 for bubble-induced turbulence stronglyunderestimates the turbulence level in a number <strong>of</strong> testcases. Another disadvantage <strong>of</strong> this approach consists<strong>of</strong> their local properties, because it considers the increase<strong>of</strong> the turbulence intensity only locally in thereactor where the gas phase is actually present. Inreality, the turbulence induced by the bubbles at somegiven point can spread <strong>and</strong> affect regions farther fromthe turbulence source. Oey et al. 14 applied a 3D tw<strong>of</strong>luidmodel containing the steady-drag, added-mass, <strong>and</strong>turbulent dispersion forces together with an extrasource term in the k-ɛ turbulence model to account forthe effects <strong>of</strong> the interface in their in-house staggeredFVM code (ESTEEM). Because <strong>of</strong> the controversy in theliterature regarding dispersed-phase turbulence, bothlaminar <strong>and</strong> turbulent gas simulations were performed.In the turbulence case, the extended Tchen’s theoryapproach was adopted. The liquid tangential velocitycomponents close to the wall were found using wallfunctions, whereas no wall friction was taken intoaccount for the dispersed phase. They found that, in 3D,the steady-drag force was sufficient to capture the globaldynamics <strong>of</strong> the bubble plume whereas the other forcesmoreover had secondary effects only. Furthermore, Oeyet al. 14 also concluded that the discretization schemesadopted for the convection terms in the fluid momentum<strong>and</strong> dispersed-phase continuity equations had a severeimpact on the solutions. This is an important aspect <strong>of</strong>any multiphase flow modeling: the numerical <strong>and</strong>modeling issues cannot be investigated (completely)separately, as their interplay is <strong>of</strong> considerable importance.8,11,57Most <strong>of</strong> the studies mentioned above adopted a kind<strong>of</strong> k-ɛ model to describe the liquid turbulence in thesystem, whereas there is less consensus regardingwhether the dispersed phase should be consideredturbulent or laminar, or even whether any deviatingstress terms should remain in the dispersed-phaseequations at all. However, even the k-ɛ model predictionsare questioned by Deen et al. 40 <strong>and</strong> Bove et al. 43This group showed that only low-frequency unsteadyflow is obtained using the k-ɛ model because <strong>of</strong> overestimation<strong>of</strong> the turbulent viscosity. These modelpredictions were found not to be in satisfactory agreementwith the more high-frequency experimental results.On the other h<strong>and</strong>, when a 3D Smagorinsky LES(large-eddy simulation) model was used instead, thestrong transient movements <strong>of</strong> the bubble plume observedin experiments were captured.For completeness, it should also be mentioned that,in a few recent papers, 58,59 it has been claimed that 2Dmixture model formulations can be used to reproducethe time-dependent flow behavior <strong>of</strong> 2D bubble columns.It has been found that the crucial physics can becaptured by 2D models if suitable turbulence models areused. The predictions reported by Bech 58 rely on the


5112 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005inclusion <strong>of</strong> a mass-diffusion term in the dispersedphasecontinuity equation, whereas the predictions <strong>of</strong>Cartl<strong>and</strong> Glover <strong>and</strong> Generalis 59 rely on the inclusion<strong>of</strong> a Reynolds stress model.However, the great majority <strong>of</strong> the investigationsreported conclude that, for both rectangular <strong>and</strong> cylindricalcolumns, the high-frequency instabilities are 3D<strong>and</strong> have to be resolved through the use <strong>of</strong> 3D models.An apparent conclusion drawn from these papers (althoughnot explicitly stated), that the fluid dynamicshad to be essentially perfect before the chemistryrelatedtopics could be considered, might have been asevere hindrance in the development <strong>of</strong> integratedmodels <strong>of</strong> interest for chemical engineers. The mainlimitation is said to be the tremendous CPU dem<strong>and</strong>sneeded for the high-frequency instabilities (assumingthat these are important for the chemical conversion),making such 3D simulations infeasible for most researchgroups. Unfortunately, there are several indicationsthat these local high-frequency dynamics <strong>and</strong>coherent structures <strong>of</strong> the flow are important in determiningthe conversion <strong>of</strong> the system. 15,37,42,60,61 Mixingtimes predicted by steady flow codes, for example, arefound not to be in accordance with experimental data,whereas mixing times predicted by high-frequencytransient flow codes are in fair agreement with thecorresponding measurements.The interfacial <strong>and</strong> turbulence closures suggested inthe literature also differ in considering the anticipatedimportance <strong>of</strong> the bubble size distributions. It thusseemed obvious for many researchers that furtherprogress on the flow pattern description was difficultto obtain without a proper description <strong>of</strong> the interfacialcoupling terms, <strong>and</strong> especially <strong>of</strong> the contact area orprojected area for the drag forces. The bubble columnresearch thus turned toward the development <strong>of</strong> adynamic multifluid model that is extended with apopulation balance module for the bubble size distribution.However, the existing models are still restrictedin some way or another because <strong>of</strong> the large CPUdem<strong>and</strong>s required by 3D multifluid simulations.Multifluid Models <strong>and</strong> <strong>Bubble</strong> Size Distributions.To gain insight into the capability <strong>of</strong> the presentmodels to capture physical responses to changes in thebubble size distributions, a few preliminary analyseshave been performed adopting the multifluid modelingframework.Carrica et al. 13 developed the most comprehensivemultifluid/population balance model known to date (tothe knowledge <strong>of</strong> the authors) for the description <strong>of</strong>bubbly two-phase flow around a surface ship. In themultifluid model part, the inertia <strong>and</strong> shear stresstensors were assumed to be negligible for the gas bubblephases or groups. The interfacial momentum transferterms included for the different bubble groups aresteady-drag, added-mass, lift, <strong>and</strong> turbulent dispersionforces. Algebraic turbulence models were used both forthe liquid-phase contributions <strong>and</strong> for the bubbleinducedturbulence. The two-fluid model was solvedusing an FVM on a staggered grid. In the populationbalance part <strong>of</strong> the model the intergroup transfermechanisms included were bubble breakage <strong>and</strong> coalescence<strong>and</strong> the dissolution <strong>of</strong> air into the ocean. Fifteensize groups were used with bubble radii at normalpressure between 10 <strong>and</strong> 1000 µm. It was found thatintergroup transfer is very important in these flows fordetermining both a reasonable two-phase flow field <strong>and</strong>the bubble size distribution. The population balance wasdiscretized using a multigroup approach. It was pointedout that the lack <strong>of</strong> validated kernels for bubble coalescence<strong>and</strong> breakage limited the accuracy <strong>of</strong> the modelpredictions. Politano et al. 62 adapted the populationbalance model <strong>of</strong> Carrica et al. 13 for the purpose <strong>of</strong> 3Dsteady-state simulation <strong>of</strong> bubble column flows. However,no details regarding the necessary model modificationswere provided. In a later study by Politano et al., 63a 2D steady-state version <strong>of</strong> this model was applied forthe simulation <strong>of</strong> polydisperse two-phase flow in verticalchannels. The two-fluid model was modified using anextended k-ɛ model for the description <strong>of</strong> liquid-phaseturbulence. A two-phase logarithmic wall law wasdeveloped to improve on the boundary treatment <strong>of</strong> thek-ɛ model. The interfacial momentum-transfer termsincluded for the different bubble groups were the steadydrag,added-mass, lift, turbulent dispersion, <strong>and</strong> wallforces. The two-fluid model equations were discretizedusing an FD method. The bubble mass was discretizedin three groups. The effect <strong>of</strong> bubble size on the radialphase distribution in vertical upward channels wasinvestigated. Comparing the model predictions withexperimental data, it was concluded that the model isable to predict the transition from the near-wall gasvolume fraction peaking to the core peaking beyond acritical bubble size.Tomiyama 53 <strong>and</strong> Tomiyama <strong>and</strong> Shimada 54 adoptedan (N + 1)-fluid model for the prediction <strong>of</strong> 3D unsteadyturbulent bubbly flows with nonuniform bubble sizes.Among the (N + 1) fluids, one fluid corresponds to theliquid phase, <strong>and</strong> the N fluids correspond to gas bubbles.To demonstrate the potential <strong>of</strong> the proposed method,unsteady bubble plumes in a water-filled vessel weresimulated using both (3 + 1)-fluid <strong>and</strong> two-fluid models.The gas bubbles were classified <strong>and</strong> fixed in threegroups only, so a (3 + 1)- or four-fluid model was used.The dispersions investigated were very dilute so thebubble coalescence <strong>and</strong> breakage phenomena wereneglected, whereas the inertia terms were retained inthe three bubble-phase momentum equations. No populationbalance model was then needed, <strong>and</strong> the phasecontinuity equations were solved for all phases. It wasconfirmed that the (3 + 1)-fluid model gave betterpredictions than the two-fluid model for bubble plumeswith nonuniform bubble sizes.As mentioned earlier, three-fluid models have alsobeen used by a few groups. 32,35 Krishna <strong>and</strong> van Baten 35solved the Eulerian volume-averaged mass <strong>and</strong> momentumequations for all three phases. However, no interchangebetween the small- <strong>and</strong> large-bubble phases wasincluded so each <strong>of</strong> the dispersed bubble phases exchangesmomentum only with the liquid phase. Nopopulation balance model was used as the bubblecoalescence <strong>and</strong> breakage phenomena were neglected.Lehr et al. 32 extended the basic three-fluid model byincluding a simplified population balance model for thebubble size distribution.In most studies reported so far, two-fluid models havebeen used, 13,31,33,41,62,64,65 assuming that all <strong>of</strong> the particleshave the same average velocities. In other words,possible particle collisions due to buoyancy effects areneglected even though these contributions have not beenproven insignificant. This means that the two momentumequations for the two phases are solved togetherwith the continuity equation <strong>of</strong> the liquid phase <strong>and</strong> theN population balance equations for the dispersed


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5113Figure 4. Simulated steady-state results obtained with (s) <strong>and</strong> without (- --) the population balance at the axial level 2 m above thecolumn inlet. Points (‚) are experimental data. 76 v d s ) 8 cm/s.phases. 13 An alternative, <strong>of</strong>ten used when commercialCFD codes are used because <strong>of</strong> limited access to thesolver routines, is to solve the full two-fluid model inthe common way using the dispersed-phase continuityequation together with the two momentum equations<strong>and</strong> the liquid-phase continuity. Within the IPSA-likecalculation loop, the N - 1 population balance equationsare solved in another step considering additional transportequations for scalar variables. Unfortunately, usingthis approach, it can be difficult to ensure mass conservationfor the dispersed phase, <strong>and</strong>/or negativeconcentrations can be predicted for the last class. Fromthe solution <strong>of</strong> the size distribution <strong>of</strong> the dispersedphase, the Sauter mean diameter is calculated. Thisdiameter is then used to compute the contact area, sothat the two-way interaction between the flow <strong>and</strong> thebubble size distribution is established. When dilutedispersions are considered, the interfacial momentumtransferfluxes due to particle collisions, coalescence,<strong>and</strong> breakage phenomena are normally neglected. Politanoet al. 62 are probably the only ones to have consideredthe interfacial momentum-transfer fluxes betweenthe bubble groups induced by bubble coalescence <strong>and</strong>breakage, but still, they did not include the bubblecollisions resulting in rebound.Several recent studies indicate that algebraic slipmodels are sufficient for modeling the flow pattern inbubble columns, 13-15 as only the pressure <strong>and</strong> steadydragforces dominate the axial component <strong>of</strong> the gasmomentum balance. Therefore, the population balancemodel can be merged with an algebraic slip model toreduce the computational cost required for preliminaryanalysis. 66 Furthermore, by adopting this concept, therestriction used in the two-fluid model <strong>of</strong> assuming thatall <strong>of</strong> the particles have the same velocity can beavoided. This means that only one set <strong>of</strong> momentum <strong>and</strong>continuity equations for the mixture is solved, togetherwith the N population balance equations for the dispersedphases. The individual phase <strong>and</strong> bubble-classvelocities are calculated from the mixture velocitiesusing algebraic relations.Even simpler approaches are used in solving a singletransport equation for one moment <strong>of</strong> the populationbalance only, determining a locally varying mean particlesize or the interfacial area density. 36,67-69Luo 70 initiated the population balance investigationswithin our group. The main contribution was a closuremodel for binary breakage <strong>of</strong> fluid particles in fullydeveloped turbulence flows based on isotropic turbulence<strong>and</strong> probability theories. 71 The authors also claimedthat this model contains no adjustable parameters,although a better phrase might be “no additionaladjustable parameters”, as both the isotropic turbulence<strong>and</strong> probability theories involved contain adjustableparameters <strong>and</strong> distribution functions.Hagesaether et al. 72-75 continued the populationbalance model development within the framework <strong>of</strong> anidealized plug-flow model, whereas Bertola et al. 66combined the extended population balance module witha 2D algebraic slip mixture model for the flow pattern.Bertola et al. 66 studied the effect <strong>of</strong> the bubble sizedistribution on the flow fields in bubble columns. Theyused an extended k-ɛ model to describe the turbulence<strong>of</strong> the mixture flow. Two sets <strong>of</strong> simulations wereperformed, i.e., both with <strong>and</strong> without the populationbalance. Four different superficial gas velocities, i.e., 2,4, 6, <strong>and</strong> 8 cm/s were used, <strong>and</strong> the superficial liquidvelocity was set to 1 cm/s in all cases. The populationbalance contained six prescribed bubble classes withdiameters set to d 1 ) 0.0038 m, d 2 ) 0.0048 m, d 3 )0.0060 m, d 4 ) 0.0076 m, d 5 ) 0.0095 m, <strong>and</strong> d 6 )0.0120 m.Figures 4 <strong>and</strong> 5 show simulated <strong>and</strong> experimentalresults for the superficial gas velocity v sg ) 8 cm/s.Figure 4 shows the axial <strong>and</strong> radial liquid velocity


5114 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005Figure 5. Calculated steady-state bubble number density at the axial level 2 m above the column inlet. Points (‚) are experimentaldata. 76 v d s ) 8 cm/s.components, the axial gas velocity component, <strong>and</strong> thegas fraction 2.0 m above the column inlet. Figure 5shows the number density in each class 2.0 m above theinlet.The results from the two simulations (i.e., with <strong>and</strong>without the population balance) are nearly identical. Inboth cases, the simulated results are in fair agreementwith the experimental data, but in the center <strong>of</strong> thereactor, the deviations between the simulated <strong>and</strong>experimental velocity <strong>and</strong> void fraction pr<strong>of</strong>iles arerather large. The number density 2.0 m above the inletis shown in Figure 5. In general, it was found that theinitial bubble size was not stable <strong>and</strong> was furtherdetermined by break-up <strong>and</strong> coalescence mechanisms.The simulation provides results in fair agreement withthe experimental data for classes 3-6, for which thebubble number densities are <strong>of</strong> the same order <strong>of</strong>magnitude as the experimental data. In bubble classes1 <strong>and</strong> 2, the experimental bubble number densities areconsiderably underestimated in the simulations.However, in other cases, the model predictions deviatemuch more from each other <strong>and</strong> were in poor agreementthe experimental data on measurable quantities suchas phase velocities, gas volume fractions, <strong>and</strong> bubblesize distributions. An obvious reason for this discrepancyis that the breakage <strong>and</strong> coalescence kernels relyon ad hoc empiricism determining the particle-particle<strong>and</strong> particle-turbulence interaction phenomena.The existing parametrizations developed for turbulentflows are high-order functions <strong>of</strong> the local turbulentenergy dissipation rate that is <strong>of</strong>ten determined by thek-ɛ turbulence model. This approach is not accurateenough, 32,33,77 as this model variable (i.e., ɛ) merelyrepresents a closure for the turbulence integral lengthscale with model parameters fitted to experimental datafor idealized single-phase flows. The population balancekernels are also difficult to validate on the mesoscalelevel, as the physical mechanisms involved (e.g., consideringeddies <strong>and</strong> eddy-particle interactions) arevague <strong>and</strong> not clearly defined. If possible, the coalescence<strong>and</strong> breakage closures should be reparametrizedin terms <strong>of</strong> measurable quantities. Well-planed experimentalanalysis <strong>of</strong> the mesoscale phenomena are thenrequired to provide data for proper model validation.Initial <strong>and</strong> Boundary Conditions. Initial <strong>and</strong>boundary conditions are also very important parts <strong>of</strong>any model formulation. However, there is still verylimited knowledge regarding the formulation <strong>of</strong> properboundary conditions even for simple bubble columns. 11,12Inlet boundary conditions for the local gas volumefraction, bubble size distribution, <strong>and</strong> local gas velocitycomponents are difficult to determine, although anumber <strong>of</strong> more or less well founded suggestions havebeen reported in the literature. 43,66,78 The complex flowsin this section have been studied experimentally inseveral papers, 61,79,80 <strong>and</strong> there is clearly a need forimprovements as Harteveld et al. 61 found significantdiscrepancies between their experimental observations<strong>and</strong> the present model predictions regarding the onset<strong>of</strong> the significant flow instabilities. Harteveld et al. 61claimed that uniform aeration gives only a very smallentrance region <strong>and</strong> no large-scale circulation or coherentstructures. This visual observation clearly contradictsmost <strong>of</strong> the model predictions cited above.The specification <strong>of</strong> proper outlet boundary conditionsis also a problem for these flows, as the recirculatingmotion <strong>of</strong> the liquid phase continues as long as gasbubbles are present at sufficiently high fluxes. Consistentformulations <strong>of</strong> such boundaries have not yet beenreported. 81 This limitation restricts the use <strong>of</strong> explicitdiscretization schemes to situations with low gas fractionsor cases where the local recirculating flux containsvery little gas.For implicit solution methods, approximate <strong>and</strong> rathercrude outlet conditions are used for the flow variables.The boundary is usually idealized considering the liquid


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5115phase in batch mode only. 78 Following the experimentallaboratory practice <strong>of</strong> keeping the height <strong>of</strong> the gasliquiddispersion lower than the actual column height,the top surface <strong>of</strong> the column can be modeled as anoutlet for both the gas <strong>and</strong> liquid phases. It is anticipatedthat the solution <strong>of</strong> the model equations willdetermine the actual height <strong>of</strong> the gas-liquid dispersion<strong>and</strong> only gas will exit from the column outlet. Inadopting this approach, one has to ensure that themodeling closures are well posed <strong>and</strong> that the discretizationscheme applied to the governing equations aswell as the iterative solver used is capable <strong>of</strong> h<strong>and</strong>lingthe steep gradients in the volume fraction <strong>and</strong> densitypr<strong>of</strong>iles (discontinuities) that occur when the continuousphase change from being liquid below the gas-liquidinterface to gas above it. Because <strong>of</strong> the large densitydifference between these phases, such an attempt very<strong>of</strong>ten leads to nonphysical pressure <strong>and</strong> velocity valuesclose to the interface <strong>and</strong> encounters severe convergencedifficulties. To enhance the convergence behavior in theinterface region, empirically adjusted smooth pr<strong>of</strong>iles<strong>of</strong> the continuous-phase density pr<strong>of</strong>ile <strong>and</strong>/or the voidfraction pr<strong>of</strong>ile might help to maintain a fairly stablesolution. 82 For incompressible flows, this numericalapproach could enforce mass conservation, whereas forreactive <strong>and</strong> other density-varying systems, mass conservationcan be a severe problem so this boundarytreatment should be avoided.As numerical instabilities <strong>and</strong> inaccurate mass conservationare frequently encountered in solving problemscontaining sharp interfaces within the calculationdomain, the solution domain is <strong>of</strong>ten restricted to theheight <strong>of</strong> the gas-liquid dispersion. In this case, thelocal liquid velocity components normal to the outletplane are fixed at zero, as there is no net flux throughthe column outlet cross section. The top surface <strong>of</strong> thesolution domain is assumed to coincide with the freesurface <strong>of</strong> the dispersion, which might or might not beassumed flat. This assumption is rather crude, as anexact value <strong>of</strong> the gas-liquid dispersion height is notknown a priori. To induce apparently physical flowcharacteristics at the outlet, approximate boundariesare required for the other flow variables. The tangentialshear stress <strong>and</strong> the normal fluxes <strong>of</strong> all scalar variablesare set to zero at the free surface. The gas bubbles arefree to escape from the top surface. In commercial codes,this implementation might not be possible, <strong>and</strong> furtherapproximations have been proposed. In some cases, thetop surface <strong>of</strong> the dispersion is defined as an “inlet”where the normal liquid velocity is set to zero <strong>and</strong> thenormal gas velocity is set to an approximate terminalrise velocity. In other cases, the top surface <strong>of</strong> thedispersion is modeled as a “no shear wall”. This boundarysets both the gas <strong>and</strong> liquid velocities to zero. Torepresent escaping gas bubbles, an artificial gas sinkis defined for all <strong>of</strong> the grid cells attached to the topsurface. A similar approach was used by Lehr et al. 32The free surface assumed to be located at the top <strong>of</strong> thecolumn was replaced by an apparent semipermeablewall. In this way, the gas could leave the system,whereas the liquid surface acted as a frictionless wallfor the liquid. The liquid was considered to leave thesystem through an outflow at the periphery <strong>of</strong> thecolumn.Bove et al. 43 specified an outlet pressure boundaryinstead <strong>and</strong> determined the axial liquid velocity componentsin accordance with a global mass balance. Thisapproach is strictly valid only when the variations inliquid density due to interfacial mass transfer or temperaturechanges are negligible, as the local changesare not known a priori.Furthermore, in many industrial systems, the liquidphase is not operated in batch mode; rather, a continuousflow <strong>of</strong> the liquid phase has to be allowed. However,because <strong>of</strong> numerical problems, most reports on bubblecolumn modeling introduce the simplifying assumptionthat the continuous phase is operated in batch mode.Further work is needed on the continuous-mode boundaryconditions.The assumption <strong>of</strong> cylindrical axisymmetry in thecomputations prevents lateral motion <strong>of</strong> the dispersedgas phase <strong>and</strong> leads to an unrealistic radial phasedistribution that has also been reported by otherauthors. 35 Krishna <strong>and</strong> van Baten 35 obtained betteragreement with experiments when a 3D model wasapplied. However, experience shows that it is verydifficult to obtain reasonable time-averaged radial voidpr<strong>of</strong>iles even in 3D simulations, even though the predictedradial velocity pr<strong>of</strong>iles seem reasonable.At the wall boundary, both the 2D <strong>and</strong> 3D turbulentviscosity-basedmodels rely on the assumption that thesingle-phase logarithmic law <strong>of</strong> the wall is still valid forbubble-driven upward flows in bubble columns. Troshko<strong>and</strong> Hassan 83,84 claimed that this assumption is reasonablefor dilute <strong>and</strong> downward bubbly flows, but notrecommended for upward bubble-driven flows as foundin bubble columns. A corresponding two-phase logarithmicwall law was derived with the intention that itwould be in better agreement with experimental dataon homogeneous bubbly flows.Numerical Schemes <strong>and</strong> Algorithms. In terms <strong>of</strong>progress on the numerical side, several novel schemes<strong>and</strong> algorithms for solving the fluid dynamic part <strong>of</strong> themodel have been published (the solution methods intendedfor the population balance equation will bediscussed later). This work has concentrated on severalitems. Most important, as mentioned earlier, one avoidsusing the very diffusive first-order upwind schemeswhile discretizing the convective terms in the multifluidtransport equations. Instead, higher-order schemes, e.g.,second- or third-order TVD (total variation diminishing)schemes, which are much more accurate, have beenimplemented in the codes. 8,14,25,26,57 The numericaltruncation errors induced by the discretization schemeadopted for the convective terms can severely alter thenumerical solution, <strong>and</strong> this can destroy the physicsreflected by the model equations. It is also well-knownthat the strong coupling between the phasic equationsprevents efficient <strong>and</strong> robust convergence for the implicititeration process. 85 Exchanging the well-knownpartial elimination algorithm (PEA) <strong>of</strong> Spalding 86,87 <strong>and</strong>reducing the interaction between the phasic velocitiesin the drag terms <strong>of</strong> the momentum equations with acoupled solver 46,88-93 that simultaneously iterates on onevelocity component <strong>of</strong> all phases seems to improve thenumerical stability <strong>and</strong> the overall convergence rate. 94In addition, the discretization schemes together with thesolution algorithms lead to large sparse linear systems<strong>of</strong> algebraic equations that need to be solved. Previously,the TDMA algorithm was applied to multidimensionalproblems to determine a linewise Gauss-Seidel approach.During the past decade, there have been developmentsin full-field solvers along the lines <strong>of</strong> Krylovsubspace methods <strong>and</strong> in the field <strong>of</strong> multigrid


5116 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005schemes. 90,91,94-96 These methods make effective use <strong>of</strong>sparsity <strong>and</strong> are efficient methods for the solution <strong>of</strong>large linear systems. 96 Furthermore, the original interphase-slipalgorithm (IPSA) <strong>of</strong> Spalding 86 was developedto introduce an implicit coupling between pressure <strong>and</strong>volume fractions. The algorithm contains an attempt toapproximate the simultaneous change <strong>of</strong> volume fractions<strong>and</strong> velocity with pressure. However, most versions<strong>of</strong> the IPSA algorithm were merely extensions <strong>of</strong>the single-phase SIMPLE approach; thus, the pressurewas computed by assuming that all <strong>of</strong> the velocitycomponents, but not the volume fraction variable,depend on pressure changes. The pressure-volumefraction relationship was not considered in a satisfactoryimplicit manner. Therefore, the extension <strong>of</strong> the singlephasepressure-velocity coupling to multifluid modelsled to low convergence rates <strong>and</strong> pure robustness <strong>of</strong> theiterating procedure. 20 Lately, numerical methods originallyintended for multiphase models, rather thanbeing extended single-phase approaches, have beeninvestigated. 46,89,91-93 A fully implicit coupling <strong>of</strong> thephasic continuity <strong>and</strong> compatibility equations within theframework <strong>of</strong> pressure-volume fraction-velocity correctionschemes seems to have potential if the resultingset <strong>of</strong> algebraic equations can be solved by an efficient<strong>and</strong> robust parallel solver. 93 However, so far, severestability problems have been identified within theiterative solution process. The numerical properties <strong>of</strong>the resulting set <strong>of</strong> algebraic equations are not optimizedfor robust solutions. In summary, significantimprovements <strong>of</strong> the numerics have been obtainedduring the past decade, but the present algorithms arestill far from being sufficiently robust <strong>and</strong> efficient.Further work on the numerical solution methods in theframework <strong>of</strong> FVMs should proceed along the pathssketched above. Very few attempts exploring the capabilities<strong>of</strong> alternative methods such as FEMs 97 <strong>and</strong> fullyspectral methods have been reported.Chemical Reaction Engineering. As a consequence<strong>of</strong> the large number <strong>of</strong> modeling limitationsdiscussed above, caution concerning the predictivepower <strong>of</strong> the multifluid model applied to bubbly flowsis certainly justified. However, even though there areobviously many open questions <strong>and</strong> shortcomings relatedto the fluid dynamic modeling <strong>of</strong> bubble columns,preliminary attempts have been performed in predictingchemical reactive systems. For example, the early CFDmodels that were developed in our group to describeglobal steady flow pattern were tested aiming at predictingchemical conversion <strong>of</strong> a reactive system operatedin a bubble column. 20,98 The system investigatedwas CO 2 absorption in a methyldiethanolamine (MDEA)solution. The starting point for the numerical investigationwas a steady two-fluid flow model tuned to the air/water system. This air/water model was then appliedto the reactive system without retuning <strong>of</strong> any modelparameters but updating the physical properties inaccordance with the reactive system. It was found that,with this model, the global flow pattern, the interfacialmass-transfer fluxes, <strong>and</strong> the conversion were all stilldifficult to predict because <strong>of</strong> the limited accuracyreflected by the interfacial coupling models <strong>and</strong> especiallythe relations used for the contact area (<strong>and</strong> theprojected area). Several semiempirical models for thelocally varying mean bubble size <strong>and</strong> contact areas havebeen suggested 20,99 with limited success. The accuracy<strong>of</strong> the experimental data used for model validation wasalso questioned.According to Dudukovic et al., 5 still no fundamentalmodels for the interfacial heat- <strong>and</strong> mass-transfer fluxeshave been coupled successfully to the flow models, <strong>and</strong>reliable reactor performance predictions based on thesemodels are not imminent. The mechanisms <strong>of</strong> coalescence<strong>and</strong> breakage are far from being sufficientlyunderstood yet.The physicochemical hydrodynamics determining thebubble coalescence <strong>and</strong> breakage phenomena cannot becaptured by any continuum model formulation. Therefore,the fluid dynamic models described above do notsignificantly improve on the prediction <strong>of</strong> the interfacialtransfer fluxes <strong>and</strong> thus the chemical conversion, as thepertinent physics on bubble, interface, <strong>and</strong> molecularscales still have to be considered using empirical parametrizations.A multiscale reactor model with aninherent two-way coupling between mechanisms onscales ranging from discrete molecular ones (moleculardynamics) to continuum macroscale reactor scales shouldbe elucidated. 100The weakest link in modeling reactive systems withinbubble columns is thus still the fluid dynamic part,considering multiphase turbulence modeling, interfacialclosures, <strong>and</strong> especially the impact <strong>and</strong> descriptions <strong>of</strong>bubble size <strong>and</strong> shape distributions. For reactive systems,the estimates <strong>of</strong> the contact areas <strong>and</strong> thus theinterfacial mass-transfer rates are likely to contain largeuncertainties. The preliminary simulations performedto date clearly indicate that future work should alsoconsider the possibility <strong>of</strong> developing more efficient <strong>and</strong>stable numerical models for the integrated multiphase/population balance models. Bertola et al. 66 found thatthe CPU dem<strong>and</strong> was increased by about 15% for everybubble class added to a simulation.Outline <strong>of</strong> the Paper. In the following sections, afairly rigorous modeling framework is outlined with afocus on population balance modeling. Emphasis isplaced on the closures for the coalescence <strong>and</strong> breakageterms. Bottlenecks that need further considerations areidentified. The numerical methods considered goodc<strong>and</strong>idates in solving the population balance model forbubbly flows are the moment, volume (i.e., the class <strong>and</strong>multigroup), <strong>and</strong> spectral methods. The optimal choice<strong>of</strong> numerical solution methods is discussed. Finally,concluding remarks are given.2. <strong>Modeling</strong> FrameworkAverage multifluid models with an integrated populationbalance module have been found to represent atrade<strong>of</strong>f between accuracy <strong>and</strong> computational efforts forpractical applications. If bubbles <strong>of</strong> different mass haveto be considered, separate continuity <strong>and</strong> momentumbalances are required for each bubble size <strong>and</strong> for thecontinuous liquid phase in a rigorous model formulation.The multifluid model equations are listed in thefollowing sections, together with the interfacial closuresthat are adopted in most gas-liquid (two-fluid) analyses.Multifluid Formulation. The multifluid model representsa direct extension <strong>of</strong> the well-known two-fluidmodel <strong>and</strong> is described in detail in Carrica et al., 13Pfleger et al., 30 Pfleger et al., 39 Tomiyama <strong>and</strong> Shimada,54 <strong>and</strong> Fan et al. 101 The governing set <strong>of</strong> equationsconsists <strong>of</strong> the continuity <strong>and</strong> momentum equations for


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5117N + 1 phases; one phase corresponds to the liquid phase,<strong>and</strong> the remaining N phases are gas bubble phases.The fundamental form <strong>of</strong> the multifluid continuityequation for phase k readswhere F k is the density <strong>and</strong> R k is the volume fraction <strong>of</strong>phase k. The first term on the left-h<strong>and</strong> side denotesthe transient change <strong>of</strong> mass within a control volume,the second term denotes the convective flux <strong>of</strong> massthrough the surfaces <strong>of</strong> the control volume, <strong>and</strong> the termon the right-h<strong>and</strong> side describes the net mass-transferflux to phase k from all other phases l.The phasic volume fractions also satisfy the compatibilityconditionIn a consistent manner, the momentum balance forphase k yields∂∂∂t (R k F k ) + ∇‚(R k F k v k ) ) ∑Γ k,l (3)l)1The terms on the left-h<strong>and</strong> side <strong>of</strong> eq 5 denote the inertiaforce, whereas the terms on the right-h<strong>and</strong> side denoteall <strong>of</strong> the additional forces acting on phase k. These arethe pressure force, the deviating normal <strong>and</strong> shearstresses, the gravitational force, <strong>and</strong> the interfacialmomentum-transfer terms accounting for all momentumtransferfluxes between phase k <strong>and</strong> the other N phases.When sufficiently dilute dispersions are considered,only particle-fluid interactions are significant, <strong>and</strong> thetwo-fluid closures can be adopted. For the gas bubblephases (d), the interaction with the continuous liquidphase (c) <strong>and</strong> the wall (w) through the last term on theright-h<strong>and</strong> side <strong>of</strong> eq 5 is expressed asi.e., the sum <strong>of</strong> steady-drag, added-mass, lift, turbulentdiffusion, <strong>and</strong> wall forces, respectively. All <strong>of</strong> the forceterms are multiplied by the liquid fraction because <strong>of</strong>the reduced liquid volume available at considerable gasloadings.The net interfacial momentum-transfer term for theliquid phase (i.e., excluding the wall forces) can bewritten asThe drag force is given byNN+1∑ R k ) 1 (4)k)1∂t (R k F k v k ) + ∇‚(R k F k v k v k ) )-R k ∇p - ∇‚(R k σj k ) +NR k F k g +∑M k,l (5)l)1N+1∑ M d,l ≈ M d,c + M d,w )R c F d )R c F D,d +R c F V,d +l)1R c F L,d +R c F TD,d +R c F d,w (6)N∑d)1M c,d )-∑M d,c (7)d)1F D,d )-F D,d (v d - v c ) )- 34d s,dR d F c C D,d |v d - v c |(v d -Nv c ) (8)The drag coefficient can be estimated using the relationsuggested by Tomiyama et al. 102C D,d ) max{ min [ ARe B,d(1 +0.15Re B,d 0.687 ),For pure systems, the parameter A ) 16, whereas forcontaminated systems A ) 24. The Eötvös number, E 0,d ,is given as<strong>and</strong> Re B,d is the particle Reynolds number3ARe B,d] , 3E 0,d8(E 0,d + 4)} (9)E 0,d ) g z |F c -F d |d 2s,d(10)σ IRe B,d ) F c d s,d |v d - v c |µ c(11)The acceleration <strong>of</strong> the liquid in the wake <strong>of</strong> the bubblecan be taken into account through the added-mass forcegiven by Ishii <strong>and</strong> Mishima 103F V,d )-R d F c C V,d( Dv dDt - Dv cDt) (12)where C V,d ) 0.5 is derived for potential flow.The lift force on the dispersed phase due to shear inthe liquid phase is expressed as 104F L,d )-R d F c C L,d (v d - v c ) × (∇ ×v c ) (13)where C L,d ) 0.5 is derived for potential flow.The dispersion <strong>of</strong> bubbles in turbulent liquid flow canbe modeled as suggested by Carrica et al. 13F TD,d )-ν c,tR d Sc t,dF D,d ∇R d (14)where the Schmidt number is defined as Sc t,d ) ν c,t /ν d,t .An additional lift force that pushes the dispersedphase away from the wall was suggested by Antal etal. 105 to be given byF W,d ) max(0, C w1 + C w2d s,dy)R d F c|v d - v c | 2d s,dn w (15)where C w1 )-0.1 <strong>and</strong> C w1 ) 0.35. This force, eq 15,represents an extension <strong>of</strong> the original model <strong>of</strong> Antalet al. 105 A defect in the original model, namely, that abubble located far from the wall is attracted to the wall,has been removed. 63Turbulence Closures. It is expected that futuremultifluid models will adopt VLES turbulence closures.40 However, as noted in the Introduction, so far,the st<strong>and</strong>ard (single-phase) k-ɛ model is usually adoptedas the basis for describing liquid-phase turbulence inEulerian multiphase simulations. The structure <strong>of</strong> theturbulence model equations thus corresponds to thewell-known generalized scalar transport equation formultiphase flow. 39 A few extensions <strong>of</strong> the st<strong>and</strong>ard k-ɛmodel have also been used, accounting for bubbleinducedturbulence.The turbulence models adopted in the sample simulationspresented in this paper represent three versions


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5119- 1 F dn∫∇‚(R d / F d n v d /// )dv - 1 F cn∫∇‚(R c / F c n v c /// )dv)- ∆tF dn∫ V∇‚(R d / ∇φ n+1 )dv - ∆tF cn∫ V∇‚(R c / ∇φ n+1 )dv(26)5. The void fractions are updated by solving thecontinuity equation for the dispersed phase using thenewest estimate <strong>of</strong> the R k <strong>and</strong> v k variables∫ V( R n+1 d F n d -R n nd F d∆t) dv )-∫ V∇‚(R / d F n d v n+1 d )dv (27)For variable-density flows, a suitable equation <strong>of</strong> state(EOS) is used to calculate the variations in the densities.The fractional-step concept applied consists <strong>of</strong> successiveapplications <strong>of</strong> the predefined operators determiningparts <strong>of</strong> the transport equation. The convective<strong>and</strong> diffusive terms are further split into their componentsin the various coordinate directions. The timetruncationerror in the splitting scheme is <strong>of</strong> first order.The convective terms are solved using a second-orderTVD scheme in space <strong>and</strong> a first-order explicit Eulerscheme in time. The TVD scheme applied was constructedby combining the central difference scheme <strong>and</strong>the classical upwind scheme by adopting the “smoothnessmonitor” <strong>of</strong> van Leer 108 <strong>and</strong> the monotonic centeredlimiter. 109 In the turbulence model, the diffusive termswere discretized with a second-order central differencescheme in space <strong>and</strong> a first-order semiimplicit Eulerscheme in time.4. Sample Results <strong>and</strong> DiscussionIn this section, the capabilities <strong>and</strong> limitations <strong>of</strong> 2Daxisymmetric two-fluid models for simulating cylindricalbubble column reactor flows are discussed. Modellimitations are explained, <strong>and</strong> sample results are givento illustrate the properties <strong>of</strong> the interfacial, turbulence,<strong>and</strong> wall interaction closures in use today. The aim isto evaluate whether a 2D model can be sufficient inproviding reasonable predictions when descriptions <strong>of</strong>variable bubble size distributions, chemical reactivesystems, continuous flow <strong>of</strong> the liquid phase, compressibleeffects, etc., are to be included in our fluid dynamicanalysis.In this work, simulations were performed with threedifferent turbulence closures for the liquid phase. Ast<strong>and</strong>ard k-ɛ model was used to examine the effect <strong>of</strong>shear-induced turbulence (case a). In the first alternative,case b, both shear- <strong>and</strong> bubble-induced turbulenceare taken into account by linearly superposing theturbulent viscosities obtained from the k-ɛ model <strong>and</strong>the model <strong>of</strong> Sato <strong>and</strong> Sekoguchi. 55 A third approach,case c, is similar to case b in that both shear- <strong>and</strong>bubble-induced turbulence contributions are considered.However, in this model formulation, case c, the bubbleinducedturbulence contribution is included through anextra source term in the turbulence model equations. 8As mentioned earlier, it has been shown by Svendsenet al., 17 among others, that the time-averaged experimentaldata on the flow pattern in cylindrical bubblecolumns apparently becomes close to axisymmetric.Therefore, it was anticipated that steady-state 2Daxisymmetric simulations could capture the pertinenttime-averaged flow pattern in these columns <strong>and</strong> thata time-averaged flow description was sufficient for theanalysis <strong>of</strong> interest for chemical reaction engineers.Because 2D computations are much faster to perform,it is also important to learn as much as possible fromsuch simulations before one performs dynamic 3Dcomputations that actually might not provide anyfurther information from a chemical engineering point<strong>of</strong> view. Therefore, in line with the early CFD analyses<strong>of</strong> bubble column flows, in this work, we are merelyinterested in the time-averaged behavior <strong>of</strong> the flowsystem, comparing the model results with time-averagedexperimental data. As we proceed in discussingour results, we intend to reveal the pertinent reasonswhy this type <strong>of</strong> bubble column simulation has beenperformed for more than two decades with a ratherlimited degree <strong>of</strong> success.First, the lack <strong>of</strong> proper boundary conditions, asdiscussed in the Introduction, is further elucidated. Atthe column outlet, numerical problems arise for highersuperficial gas velocities, because <strong>of</strong> the intense recirculatingflow pattern produced by the free surfaceboundary that redistributes the liquid flow. Furthernumerical problems are observed for the case <strong>of</strong> operatingthe bubble column in a continuous liquid mode.The assumption <strong>of</strong> cylindrical axisymmetry in thecomputations prevents lateral motion <strong>of</strong> the dispersedgas phase <strong>and</strong> leads to an unrealistic radial phasedistribution wherein the maximum void is away fromthe centerline (Figures 6 <strong>and</strong> 7), as also reported byother authors. 35 Krishna <strong>and</strong> van Baten 35 obtainedbetter agreement with experiments by applying a 3Dmodel. Even though the time-dependent terms in our2D model were retained merely to ease the numericalsolution <strong>of</strong> the 2D pseudo-steady-state problem, so thatthe resulting transients predicted by our 2D model areconsidered numerical modes, it is interesting to notethat the flow development during the first 3-5 s<strong>of</strong>thesimulations determines whether the fully developedsteady-state flow pattern will result in wall, intermediate,or center void peaking (apparently multiple steadystates). If the gas phase moves toward the wall or thecenter, there are no 3D secondary flows available toallow for any redistribution <strong>of</strong> the phases. Comparingour 2D axisymmetric model results at steady state withexperimental data on the axial velocities <strong>and</strong> voidfraction pr<strong>of</strong>iles at the axial level 0.3 m above the inlet,as in Figures 7 <strong>and</strong> 8, it can be observed that themagnitude <strong>of</strong> the simulated liquid axial velocity componentis overestimated at the centerline in cases a <strong>and</strong>c, whereas the higher effective viscosity level in case byields an estimation <strong>of</strong> the liquid axial velocity closerto the experimental results. The simulated axial velocitiesin the two phases are more tightly coupled than inthe experiment, resulting in overestimation <strong>of</strong> the axialgas velocity component at the centerline <strong>and</strong> underestimationtoward the wall. The corresponding void pr<strong>of</strong>ileis in fair agreement with experiment, but with anunphysical local minimum at the center that has alsobeen observed by other investigators. The cross-sectionalaveragedvoid, however, is overestimated by about 25-35%, so the corresponding cross-sectional averaged gasvelocity is underestimated. This is an indication that3D effects are important at least in the inlet section.There is only one way to improve on this axisymmetricmodel limitation: a full 3D model is needed. However,studies reported in the literature show that it is verydifficult to obtain reasonable time-averaged radial voidpr<strong>of</strong>iles even in 3D simulations, even though the pre-


5120 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005Figure 6. Axial liquid velocity, axial gas velocity, gas voidage, <strong>and</strong> viscosity pr<strong>of</strong>iles as a function <strong>of</strong> column radius at z ) 2.0 m after 80s (steady-state) using only drag <strong>and</strong> added mass as interfacial forces: ×, experimental data; 76 s, case a; ‚‚‚, case b; - - -, case c. Gridresolution, 20 × 72; time resolution, 2 × 10 -4 s.Figure 7. Axial liquid velocity, axial gas velocity, gas voidage, <strong>and</strong> viscosity pr<strong>of</strong>iles as a function <strong>of</strong> column radius at z ) 0.3 m after 80s (steady-state) using only drag <strong>and</strong> added mass as interfacial forces: ×, experimental data; 76 s, case a; ‚‚‚, case b; - - -, case c. Gridresolution, 20 × 72; time resolution, 2 × 10 -4 s.dicted radial velocity pr<strong>of</strong>iles seem reasonable. Alternatively,or rather in addition, the deviation betweenthe experimental data <strong>and</strong> the simulated results mightalso indicate that the inlet boundary conditions, phasedistribution, <strong>and</strong> interfacial coupling are inaccurate <strong>and</strong>not in agreement with the real system, as will bediscussed later.The present simulations indicate that the most importantmechanism for radial movement <strong>of</strong> the gas, asthe gas starts entering the bottom <strong>of</strong> the column, is theliquid wall friction (Figure 8). In our 2D pseudotransientsimulations, the wall friction creates a significantpressure gradient at the inlet that pushes thegas toward the center line. An effect <strong>of</strong> the axisymmetricboundary condition is that it reflects most <strong>of</strong> themomentum in the fluid streams. Thus, if the effectiveviscosity in the bulk is low, only a small amount <strong>of</strong> thekinetic energy <strong>of</strong> the liquid will dissipate into heat, <strong>and</strong>the reflected fluxes <strong>of</strong> liquid <strong>and</strong> gas might havesignificant radial velocity components. In contrast, atthe wall, the kinetic energy seems to be dissipated to alarger extent because <strong>of</strong> the wall friction. The radial


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5121Figure 8. Gas <strong>and</strong> liquid velocity components <strong>and</strong> gas voidagepr<strong>of</strong>iles as a function <strong>of</strong> column radius showing center peaking inthe bubble column after 4 s (left) <strong>and</strong> 8 0s (right). Grid resolution,20 × 72; time resolution, 2 × 10 -4 s.boundary conditions thus make the radial flow unphysicallysensitive to the turbulence level in the column. Ifthe resulting radial velocity away from the center is highenough, a large fraction <strong>of</strong> the fluids seems to reach thewall, <strong>and</strong> the buoyancy force will ensure that the gasflows upward along the wall. This inlet flow patternresults in wall peaking, as shown in Figure 9. Alternatively,if the resulting radial velocity away from thecenter become lower because <strong>of</strong> a higher effectiveviscosity in the fluid phase, only negligible fluid streamswill reach the wall. The liquid will then entrain the gasdownward along the wall as the liquid continuityenforces a circulation cell, <strong>and</strong> a void center peakingpr<strong>of</strong>ile occurs as shown in Figure 8. Apparently, similarnumerical modes can be produced in dynamic 3Dsimulations with insufficient interfacial <strong>and</strong> turbulenceclosures.The limited accuracy reflected by the inlet boundarycondition <strong>and</strong> the prediction <strong>of</strong> bubble size distributionphenomena is considered next. The specifications <strong>of</strong>proper inlet conditions for the phase distribution, localvelocities, bubble size distributions, turbulence quantitiesare not trivial. The approach used here relies onthe possibility <strong>of</strong> estimating an average bubble size orbubble size distribution at the inlet zone <strong>and</strong> a netterminal velocity <strong>and</strong> neglects the effects <strong>of</strong> the localFigure 9. Gas <strong>and</strong> liquid velocity components <strong>and</strong> gas voidagepr<strong>of</strong>iles as a function <strong>of</strong> column radius showing wall peaking inthe bubble column after 4 s (left) <strong>and</strong> 8 0s (right). Grid resolution,20 × 72; time resolution, 2 × 10 -4 s.bubble jets found at the locations <strong>of</strong> the holes in the gasdistributor. This approach thus induces a diffusive effectdistributing the momentum uniformly over the columncross section. A better approach could be to resolve theflow through the gas distributor, as an integrated part<strong>of</strong> the model.Bertola et al. 66 studied the effect <strong>of</strong> the bubble sizedistribution on the flow fields in bubble columns usinga 2D axisymmetric mixture model with an inherentpopulation balance. The initial bubble size was foundnot to be stable <strong>and</strong> was further determined by breakup<strong>and</strong> coalescence mechanisms that were to a largeextent influenced by the local turbulence level. Even forthe pure flow calculations, the common simplifyingassumption <strong>of</strong> a uniform phase distribution at the inletwas found questionable, as the breakage <strong>and</strong> coalescencekernels rely on ad hoc empiricism determining theparticle-particle <strong>and</strong> particle-turbulence interactionphenomena. For reactive systems, the estimates <strong>of</strong> thecontact areas <strong>and</strong> thus the interfacial mass-transfer rateat the inlet are likely to contain large uncertainties.Turbulence model limitations are difficult to avoideven for single-phase flows. To enable multiphasesimulations with the k-ɛ model, a few model adjustmentswere required to avoid divergence <strong>of</strong> the numer-


5122 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005ical algorithms. Accounting for shear-induced turbulenceonly, case a, a minimum turbulent energy level,k min ) 5 × 10 -3 m 2 /s 2 , was used in the first few seconds<strong>of</strong> the pseudo-dynamic simulation to ensure a sufficientlyhigh viscosity level at the inlet to avoid havingthe turbulence level die out before the gas entered intothe column. For the same reason, the contribution <strong>of</strong>the bubble-induced turbulence in case b was enforcedby a factor <strong>of</strong> 10 during the first 3s<strong>of</strong>thesimulationuntil the k-ɛ-model started to produce turbulent energy.In the case <strong>of</strong> introducing the bubble-induced turbulencethrough a source term in the turbulence model, case c,the start-up <strong>of</strong> the simulation was performed in thesame way as for case a.Comparisons between the steady-state pr<strong>of</strong>iles ataxial levels z ) 2.0 m <strong>and</strong> z ) 0.3 m for cases a-c aregiven in Figures 6 <strong>and</strong> 7, respectively. For z ) 2.0 m,all <strong>of</strong> the models give reasonable velocities comparedto the experimental data. The turbulent viscosity <strong>of</strong> caseb is approximately 20% higher than that <strong>of</strong> case a, whichresults in lower velocities that are closer to the experimentaldata. The turbulent viscosity pr<strong>of</strong>ile in case clies between the two other pr<strong>of</strong>iles. However, thevelocities in case c are about equal to the velocities incase b.For z ) 0.3 m, there are larger discrepancies betweenthe simulated results <strong>and</strong> the experimental data, asdiscussed earlier. The voidage pr<strong>of</strong>iles observed in theexperiment are lower than at z ) 2.0 m, whereas thepr<strong>of</strong>iles from the simulations are approximately thesame. The discrepancy indicates that the physics at thebottom <strong>of</strong> the column is not sufficiently captured by the2D model.The interfacial coupling <strong>and</strong> wall forces are difficultto parametrize with sufficient accuracy. However, thesteady-drag force is found to be the most importantinterfacial force in the system, as expected from literatureobservations. Reasonable agreement with experimentaldata is achieved by using this interfacial forceonly. A variety <strong>of</strong> different formulations <strong>of</strong> this force,<strong>and</strong> especially <strong>of</strong> the drag coefficient, can be found inthe literature, valid for deformable <strong>and</strong> spherical particles<strong>and</strong> extended for different flow regimes, swarmeffects, <strong>and</strong> pure <strong>and</strong> contaminated fluid systems. Kurul<strong>and</strong> Podowski 111 compared several expressions for thedrag coefficient in a boiling channel <strong>and</strong> showed thatthe different correlations for bubbly flows appeared tohave minor effects on the predicted void fraction pr<strong>of</strong>iles.The different drag coefficients applied by Deen et al., 40Krishna <strong>and</strong> van Baten, 35 <strong>and</strong> Tomiyama 53 deviate byless than 15% in our simulations, indicating the accuracy<strong>of</strong> this parameter.However, in several papers on bubble column modeling,14,34 drag coefficient correlations valid for nondeformablespheres, 112,113 which give rise to drag forcecoefficients being about one-fourth the values <strong>of</strong> theother coefficients mentioned above for high Reynoldsnumbers, were adopted. When these spherical particlecoefficient correlations were applied in our simulations,all <strong>of</strong> the pr<strong>of</strong>iles became flat. The gas is thus not ableto entrain the liquid to the center in the bottom <strong>of</strong> thecolumn to create circulation cells. The literature pr<strong>of</strong>iles,on the other h<strong>and</strong>, seem very similar to the experimentaldata. It is speculated that this deviation can beexplained by the different grid arrangements used inthe different codes. It is believed, however, that the zerovoid fraction occurring at the wall should be considereda net effect <strong>of</strong> the forces acting on the bubbles, thusprescribing the wall void fraction results in an overconstrainedset <strong>of</strong> equations, <strong>and</strong> should be avoided. 105Further examination <strong>of</strong> these issues is required to avoidany misinterpretations <strong>of</strong> the physical phenomenainvolved.Implementing the added-mass force has barely anyinfluence on the steady-state solution; this conclusionis in agreement with the observations reported by Deenet al., 40 among others. Deen et al. 40 explained thisnegligible effect by the fact that the simulations yield aquasi-stationary state in which there is only a littleacceleration. The bubble jets observed close to thedistributor plate are then not considered. However, theconvergence rate <strong>and</strong> thus the computational costs aresignificantly improved when this force is implemented.The net effect <strong>of</strong> the transverse lift force is to pushthe gas radially inward or outward in the columndepending on the sign <strong>of</strong> the lift coefficient. The force isinduced by a velocity gradient <strong>and</strong> can therefore onlyreinforce or smooth out gradients in the flow fields. Inour 2D simulations, this force alone cannot change acenter peak flow pattern to a wall peak pattern or viceversa. This observation is quite surprising, as the flowpattern produced by the early steady-state model <strong>of</strong>Jakobsen 20 was totally dominated by this force. Areasonable explanation could be that the simulationsby Jakobsen 20 were strongly influenced by numericaldiffusion, as the resolution in the axial direction waslimited.Effects similar to those <strong>of</strong> the lift force are observedwhen a turbulent dispersion force that is proportionalto the gradient <strong>of</strong> the gas voidage is implemented. Theonly effect seen is that it smoothes out sharp gradientsin the domain until the pr<strong>of</strong>iles become completely flat,as the models tested seemed to overestimate the diffusioneffect. This trend is probably related to the lowaccuracy reflected by the effective viscosity variabledetermined by the turbulence model. The modeling <strong>of</strong>this force should also be seen in connection with theinherent numerical diffusion caused by the numericalapproximation <strong>of</strong> the convective terms.The radial wall lift force proposed by Antal et al. 105requires a certain minimum resolution <strong>of</strong> the grid closeto the wall. The choice <strong>of</strong> model coefficients is crucialfor the development <strong>of</strong> the pr<strong>of</strong>iles as these parametersdetermine the magnitude <strong>of</strong> the force as well as theoperative distance from the wall. The magnitude <strong>of</strong> theforce is highest for larger bubbles, <strong>and</strong> it seems to bevery sensitive to the bubble size. However, in our 2Dmodel simulations, the force could not be used to alterthe flow development significantly except very close tothe wall. There was no mechanism available for transportingthe gas further toward the center.Lopez de Bertodano 44 correctly suggested that severalunresolved phenomena close to the wall influence thephase distribution. A net axial friction force acting onthe gas in the vicinity <strong>of</strong> the wall was derived. In thepresent work, the net effect <strong>of</strong> this force is to reducethe time scale for the flow development from the gasstarts entering the column. However, this force has asignificant effect on the solution when the drag coefficientfor nondeformable spheres is used, as it causescenter peaking, whereas the pr<strong>of</strong>iles become flat withoutthis force as mentioned above. This indicates that a noslipcondition specified for the gas phase as well as afixed zero void fraction on the wall, which is used


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5123as boundary conditions in some numerical codes,will give reasonable results even with this drag coefficient.Another limiting model simplification adopted in mostreports on bubble column modeling is the assumption<strong>of</strong> a constant gas density. Simulating columns that areseveral meters high, adopting an ambient pressureboundary at the outlet, cannot be consistent with thissimplifying assumption, as density gradient effectsbecome important. Similar inconsistency problems arisewhen reactive systems are simulated, as the gas densityhas to be allowed to vary in accordance with thechemical process behavior. In addition, in our experience,interfacial mass-transfer fluxes <strong>and</strong> chemicalreactions <strong>of</strong>ten induce numerical stability problems <strong>and</strong>the lack <strong>of</strong> proper convergence.Finally, multiphase flow simulations are numericallyunstable considering both 2D <strong>and</strong> 3D models because<strong>of</strong> the applications <strong>of</strong> low-accuracy discretizations, largedensity differences between the phases, ill-conditionedimplementations <strong>of</strong> interfacial closures in the limit <strong>of</strong>zero void or zero liquid fractions, <strong>and</strong> large gradientsor discontinuities in the flow fields that can occur inthese flow systems. Therefore, further studies on theseissue are highly recommended.Preliminary results using our 2D model show thataccurate discretization schemes for the dispersed-phasecontinuity equation <strong>and</strong> for the convection terms withinthe fluid momentum equations are crucial for accuratepredictions <strong>of</strong> the flow development. Large discretizationerrors can change the flow development from center towall peaking or visa versa. Similar observations werereported by Oey et al. 14 regarding the simulation <strong>of</strong> 2Drectangular columns in 3D.The physical interpretation <strong>and</strong> relevance <strong>of</strong> dynamic2D analysis <strong>of</strong> bubble column flows have been matters<strong>of</strong> debate for many years. Judging from the experimentalanalyses discussed in the Introduction, it is evidentthat bubble column flows should be considered dynamic<strong>and</strong> 3D in any rigorous fluid dynamic analysis. Turbulenceis a dynamic 3D phenomenon. However, directnumerical simulations (DNS) <strong>of</strong> these systems is stillnot feasible, so some kind <strong>of</strong> filtering or averaging isneeded. Dynamic 3D simulations by use <strong>of</strong> filteredVLES models are computationally intensive <strong>and</strong> performaccurately only when the cut<strong>of</strong>f turbulence lengthscale (eddy size) is placed at a spectral gap in theturbulence energy spectrum, thereby ensuring that theenergy flux between the resolved <strong>and</strong> subgrid scales isvery small. Accurate modeling <strong>of</strong> these fluxes itself isthus necessary only when the subgrid-scale fluxes getlarge in comparison with the resolved fluxes (as in mostVLES simulations 114 ). Unfortunately, in engineeringpractice, the cut<strong>of</strong>f length scale is <strong>of</strong>ten apparentlychosen arbitrarily as no energy gap is found in theenergy spectrum, requiring more sophisticated subgridclosures. This might be one <strong>of</strong> the reasons why mostdynamic 3D simulations <strong>of</strong> bubble column flows havebeen performed using Reynolds-averaged models. Amongthe Reynolds-averaged turbulence models available, thesimple k-ɛ model is adopted in almost all <strong>of</strong> theinvestigations reported in the literature. There is experimentalevidence, however, that the flow in bubblecolumns is highly anisotropic. 115 It is thus an openquestion whether a 3D simulation using a k-ɛ model(predicting isotropic turbulence) provides the necessary3D effects or whether perhaps Reynolds stress or VLESmodels are required to represents the dynamics <strong>of</strong> theseflows. However, because <strong>of</strong> the impact <strong>of</strong> phenomenathat cannot be resolved by continuum models, it is notexpected that even the phase distributions predicted bythe use <strong>of</strong> VLES models will always be physicallyrealistic.Furthermore, when unsteady turbulent flow simulationsusing Reynolds-averaged models are performed,providing physical interpretations <strong>of</strong> the model resultsis not trivial. Schlichting <strong>and</strong> Gersten 116 suggested thatthese flow variables could be interpreted as timeaveragequantities whereby the time interval used inthe averaging operator had been chosen large enoughto include all turbulent fluctuations but still smallenough to contain no effect from the transient, orperiodic, parts. This concept is based on the assumptionthat the transient process occurs very slowly or that thefrequency <strong>of</strong> the oscillation is very small <strong>and</strong> lies outsidethe turbulent spectrum. Telionis 117 discussed an extension<strong>of</strong> the st<strong>and</strong>ard Reynolds averaging approach: atriple decomposition, where the instantaneous quantitiesare decomposed into three parts. Recall that, in thest<strong>and</strong>ard Reynolds decomposition, we split the instantaneousquantities into a time-independent time average(mean) <strong>and</strong> a fluctuation, whereas in the tripledecomposition approach, the mean motion is also dependenton time. The mean is thus generally made up<strong>of</strong> a time-independent part <strong>and</strong> a time-dependent part.In other cases, the application <strong>of</strong> the st<strong>and</strong>ard Reynoldsaveraging concept has been further extended to simulations<strong>of</strong> transients that are within the turbulent spectrum.To compare the simulated results obtained withthis type <strong>of</strong> model with experimental data that areaveraged over a long time period to give steady-statedata (representing the whole spectrum <strong>of</strong> turbulence),both the modeled <strong>and</strong> the resolved scales have to beconsidered in much the same manner as for VLES. 118For bubbly flows, this model interpretation inducesanother severe problem related to the bubble-inducedturbulence phenomena. For steady-state simulationsusing the st<strong>and</strong>ard Reynolds-averaged models, theturbulent kinetic energy variable, k, denotes a measure<strong>of</strong> the mean energy considering all time scales withinthe flow (i.e., for the whole turbulence energy spectrum).In a typical energy spectrum, most <strong>of</strong> the energy isaccumulated on the larger scales <strong>of</strong> turbulence <strong>and</strong> verylittle on the smaller scales. This k quantity is thussometimes used as a measure <strong>of</strong> the energy level on theintegral scales (i.e., the larger energy containing scales<strong>of</strong> turbulence). Considering the energy spectrum interms <strong>of</strong> turbulence length scales, the energy-containingscales represented by the k-ɛ model are thus normallymuch larger than the bubble size. The inclusion <strong>of</strong>turbulence production due to the bubbles’ relativemotion is therefore based on the assumption <strong>of</strong> aninverse cascade <strong>of</strong> turbulence. 8 For dynamic simulations,the k quantity can represent scales less than orat the same order <strong>of</strong> magnitude as the particle size. Inthe cases where the time averaging operator is chosenin such a way that the dispersed phase is resolved, nobubble-induced turbulence mechanisms should be includedin the turbulence model.In summary, one can certainly conclude that dynamicsimulations <strong>of</strong> bubble column flows are not a trivial task.Physical interpretations <strong>of</strong> the simulated 3D dynamicsshould be performed with caution when any kind <strong>of</strong>turbulence model is adopted.


5124 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005Considering the much simpler approach used in oursimulations, it was found that the 2D steady-statebubble column model has very limited inherent capabilities<strong>of</strong> predicting bubble column flows with sufficientaccuracy. However, after tedious adjustments <strong>of</strong> modelparameters, boundary conditions, time <strong>and</strong> space resolutions,numerical implementations, <strong>and</strong> solution algorithms,it is <strong>of</strong>ten possible to reproduce known flowfields to a certain extent. The global liquid flow patternis usually captured fairly well, whereas the correspondinggas fields are more questionable <strong>and</strong> difficult toevaluate because <strong>of</strong> the lack <strong>of</strong> reliable experimentaldata. Unfortunately, the possibility <strong>of</strong> achieving acompletely erroneous steady-state flow pattern wherethe liquid flows up along the wall <strong>and</strong> down in the core<strong>of</strong> the column seems ever-present. The almost definitiverejection <strong>of</strong> this reactor model formulation is related tothe fact that it is <strong>of</strong>ten difficult to reproduce the gasvolume fraction fields with sufficient accuracy. Thelatter limitation would severely affect the contact areas,the interfacial heat- <strong>and</strong> mass-transfer fluxes, <strong>and</strong> theprojected areas, as well as the interfacial momentumtransferfluxes, making reliable predictions <strong>of</strong> theperformance <strong>of</strong> chemical processes impossible.It is speculated that the more specific model limitationsthat have to be eliminated to enable reliable modelpredictions can be found among the topics outlined inthe following. On the physical side, the model has to bemade 3D <strong>and</strong> dynamic, proper initial <strong>and</strong> boundaryconditions have to be formulated, the high-frequencyturbulence mechanisms <strong>and</strong> lower-frequency coherentstructures have to be resolved (requiring sufficientlyaccurate turbulence closures), the fluid particleturbulenceinteraction mechanisms as well as interparticleinteractions are probably important, the description<strong>of</strong> the interactions between the dispersedphase <strong>and</strong> the turbulent wall boundary layer structureshould be extended, the pertinent phase distributionmechanisms have to be identified <strong>and</strong> described withsufficient accuracy, bubble size <strong>and</strong> shape distributionshave to be considered (i.e., the microscopic bubblecoalescence <strong>and</strong> breakage processes have to be described),the interfacial momentum closures have to bedescribed with higher accuracy (especially the transverseforces), the free surface at the top <strong>of</strong> the columnmight have to be resolved, the mechanical <strong>and</strong> turbulenceenergy balances in the system have to be fulfilled(for validation <strong>of</strong> the interfacial transfer fluxes <strong>and</strong> theenergy dissipation rates), phasic density variationsshould be allowed, <strong>and</strong> the impact <strong>of</strong> heat- <strong>and</strong> masstransfermechanisms <strong>and</strong> chemical conversion mightneed to be investigated in further detail. On the numericalside, any algorithms, grid arrangements, unphysicalsource implementations (e.g., unphysical modelextensions implemented to avoid ill-conditioned implementations<strong>of</strong> interfacial <strong>and</strong> turbulence closures in thelimit <strong>of</strong> zero void fractions), or discretization truncationerrors producing solutions merely determined by numericalmodes rather than physical mechanisms shouldbe removed (if possible). Finally, proper convergencecriteria should be formulated <strong>and</strong> fulfilled for every timestep in the simulation.In view <strong>of</strong> the comprehensive list <strong>of</strong> limitationsprovided above, it is concluded that a truly predictivemodel is not imminent. However, the more pressingimprovements required soon are those that will enablea sufficiently accurate description <strong>of</strong> the time-averagedphase distribution in the system. These improvementsshould also ensure that one would not achieve, on atime-averaged basis, downward flow <strong>of</strong> liquid in the core<strong>of</strong> the column <strong>and</strong> upward flow <strong>of</strong> liquid close to the wallthat completely disagree with the available experimentalobservations. The authors believe that, in thiscontext, one <strong>of</strong> the most important model extensionsrequired is the inclusion <strong>of</strong> a proper description <strong>of</strong> thebubble size <strong>and</strong> shape distributions. Further work inour group continues to elucidate these issues along thelines described in the remaining sections <strong>of</strong> this paper.5. Population Balance FrameworkThe chemical engineering community began the firstefforts <strong>of</strong> association with the concepts <strong>of</strong> the populationbalance in the 1960s.Two fundamental modeling frameworks emerge forformulating the early population balances, in quite thesame way as the kinetic theory <strong>of</strong> gases <strong>and</strong> thecontinuum theory were proposed deriving the governingconservation equations in fluid mechanics (e.g., appendixesin Williams 119 ). A third less rigorous approachis also used in formulating the population balancedirectly on the volume-averaging scale considering theinternal coordinates (i.e., particle size), an analogue tomultiphase mixture models.Considering dispersed two-phase flows, a few researchgroups have thus preferred to describe the dispersedphase on the basis <strong>of</strong> a statistical Boltzmann-typeequation determining the temporal <strong>and</strong> spatial rates <strong>of</strong>change <strong>of</strong> a suitably defined distribution function.Performing “Maxwellian” averaging, integrating allterms over the whole velocity space to eliminate thevelocity dependence, one obtains the generic populationbalance equation. Reliable closures are required for theunknown terms resulting from the averaging process.However, this theory provides rational means <strong>of</strong> underst<strong>and</strong>ingthe one-way coupling between discrete particlephysics <strong>and</strong> the average continuum properties. Theprocedure sketched above has much in common with thegranular theory (i.e., that could be defined as the flow<strong>of</strong> powder in a vacuum) <strong>of</strong> solid particles. 120 However,the great majority <strong>of</strong> research groups within the chemicalengineering community adopted an approach basedon an extension <strong>of</strong> the classical continuum theoryinstead developing the population balance equation. Themain reason for this choice was probably that the basicconcept is familiar to most chemical engineers frombasic courses in fluid mechanics, whereas the principles<strong>of</strong> kinetic theory might be less accessible from a pedagogicpoint <strong>of</strong> view. On the other h<strong>and</strong>, the continuumtheory gives only an average representation <strong>of</strong> thedispersed phase (i.e., on a macroscopic level relative tothe particle scale) <strong>and</strong> does not provide any informationon the unresolved mechanisms formulating the populationbalance closures. In most cases, the balance principleis applied by formulating the particle numberbalance on the integral form. Extended versions <strong>of</strong> theLeibnitz <strong>and</strong> Gauss theorems are then required totransfer the integral balances to the differential form.For turbulent flows, the governing microscopic continuumequations on the differential form are then time-(Reynolds-) averaged after being volume averaged toobtain trackable volume <strong>and</strong> time resolutions, givingrise to additional closure requirements. 121 The thirdgroup <strong>of</strong> population balances are thus formulated directlyon the averaging scales.The Leibnitz <strong>and</strong> Gauss


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5125theorems are used to cast all <strong>of</strong> the terms in theequation into a volume integral. The governing integralequation is thereafter converted to the differential form.In this formulation, the closures are purely empiricalparametrizations based on intuitive relationships ratherthan sound scientific principles.Even though the kinetic theory approach is consideredmore general, the basic balance formulation might needto be reformulated for every novel system consideredto ensure that the pertinent physics on the discreteparticle level is consistent with the problem in question.An optimized choice <strong>of</strong> distribution function definitionmight be necessary, <strong>and</strong> novel closures might be required,making the approach rather dem<strong>and</strong>ing theoretically.The population balance concept was first presentedby Hulburt <strong>and</strong> Katz. 122 Rather than adopting thest<strong>and</strong>ard continuum mechanical framework, 123 the modelderivation was based on the alternative continuumapproach based on particle statistics familiar fromclassical statistical mechanics. The main problemsinvestigated stem from solid particle nucleation, growth,<strong>and</strong> agglomeration.R<strong>and</strong>olph 124 <strong>and</strong> R<strong>and</strong>olph <strong>and</strong> Larson, 125 on theother h<strong>and</strong>, formulated a generic population balancemodel based on the extended continuum mechanicalframework. Their main concerns were solid particlecrystallization, nucleation, growth, agglomeration/aggregation, <strong>and</strong> breakage.Ramkrishna 126,127 adopted the concepts <strong>of</strong> R<strong>and</strong>olph<strong>and</strong> Larson to explore biological populations. An outline<strong>of</strong> the population balance model derivation from thecontinuum mechanics point <strong>of</strong> view was discussed.Similar approaches are also frequently used in thetheory <strong>of</strong> aerosols in which the gas is the continuousphase 121,128 <strong>and</strong> in chemical, mechanical, <strong>and</strong> nuclearengineering describing multiphase droplet flow dynamics;129,130 such an approach was also used for incompressiblebubbly two-phase flows by Guido Lavalle etal. 131Carrica et al. 13 developed the population balance fromkinetic theory using the particle mass as the internalvariable while investigating compressible bubbly twophaseflow around a surface ship, whereas most earlierwork on solid particle crystallization used particlevolume (or diameter). The two formulations are completelyequivalent in the case <strong>of</strong> incompressible gases.However, in flows where compressibility effects in thegas are important (as in the case <strong>of</strong> experimental bubblecolumns operated at atmospheric conditions), the use<strong>of</strong> mass as an internal variable was found to beadvantageous because it is conserved under pressurechanges. Lehr et al., 32 Millies <strong>and</strong> Mewes, 68 Lehr <strong>and</strong>Mewes, 36 <strong>and</strong> Pilon et al. 132 sketched a possible alternativeformulation using particle volume (diameter) as theinternal coordinate. In their approach, several growthterms have to be considered to express the effects <strong>of</strong> gasexpansion due to changes in gas density in accordancewith a suitable EOS. The use <strong>of</strong> a mass density form <strong>of</strong>the population balance derived according to the kinetictheory approach (i.e., instead <strong>of</strong> the more commonnumber density) has also been discussed, having severaladvantages in reactor technology. 133,134In chemical engineering, Coulaloglou <strong>and</strong> Tavlarides135 were among the first to introduce the simplermacroscopic formulation, describing the interactionprocesses in agitated liquid-liquid dispersions. Driftingfrom the fundamental equations, the closures becamean integrated part <strong>of</strong> the discrete numerical discretizationscheme adopted (i.e., there is no clear split betweenthe numerical scheme <strong>and</strong> the closure laws).Lee et al. 136 <strong>and</strong> Prince <strong>and</strong> Blanch 77 adopted thebasic ideas <strong>of</strong> Coulaloglou <strong>and</strong> Tavlarides 135 to formulatethe population balance source terms directly on theaveraging scales to perform an analysis <strong>of</strong> bubblebreakage <strong>and</strong> coalescence in turbulent gas-liquid dispersions.The source term closures were completelyintegrated parts <strong>of</strong> the discrete numerical schemeadopted. The number densities <strong>of</strong> the bubbles were thusdefined as the number <strong>of</strong> bubbles per unit mixturevolume <strong>and</strong> not as a probability density in accordancewith the kinetic theory <strong>of</strong> gases.Luo <strong>and</strong> Svendsen 71 extended the work <strong>of</strong> Coulaloglou<strong>and</strong> Tavlarides, 135 Lee et al., 136 <strong>and</strong> Prince <strong>and</strong> Blanch, 77formulating the population balance directly on themacroscopic scales where the closure laws for the sourceterms were integrated parts <strong>of</strong> the discrete numericalscheme used to solve the model equations. The works<strong>of</strong> Millies <strong>and</strong> Mewes, 68 Lehr <strong>and</strong> Mewes, 36 Lehr et al., 32Hagesaether et al., 73-75 <strong>and</strong> Wang et al. 137 all adoptedthe governing population balance <strong>of</strong> Luo <strong>and</strong> Svendsen.71The extension <strong>of</strong> these modeling concepts to bubbledrivengas-liquid flows is not as straightforward as onemight think from either a physical or a computationalpoint <strong>of</strong> view. Many physical or physicochemical mechanismsare still not understood or possibly even notdiscovered yet, <strong>and</strong> as the pertinent interfacial phenomenainvolved are expected to interfere at the molecularscale, tremendous computational efforts arerequired. Further challenges are foreseen consideringreactive systems in industrial chemical reactors, as thechanges in species compositions, heats <strong>of</strong> reaction, <strong>and</strong>surface processes are strongly linked.However, preliminary attempts have already beenmade in the modeling <strong>of</strong> bubble coalescence <strong>and</strong> breakagein gas-liquid systems, adopting the populationbalance framework <strong>and</strong> mathematical modeling toolsproposed by the above-mentioned pioneers. Integratedmultiphase flow/population balance models have beenapplied to bubble column reactors <strong>and</strong> two-phase stirredvessels with limited success. 31,32,41,65,66,138,139 The simulationsreveal that these models are not able to predictnonequilibrium bubble size distributions sufficientlyaccurately as a function <strong>of</strong> space <strong>and</strong> time because <strong>of</strong>the very restricted predictive capabilities reflected bythe present population balance kernels for both bubblecoalescence <strong>and</strong> breakage.In the following subsections, two forms <strong>of</strong> the populationbalance are formulated in accordance with thecontinuum theory. First, a macroscopic balance isformulated directly on the averaging scales in terms <strong>of</strong>number density functions. A corresponding set <strong>of</strong> sourceterm closures is presented as well. Second, a morefundamental microscopic population balance as well asfairly general expressions for the source terms areformulated in terms <strong>of</strong> number probability densities.This balance is then averaged using the popular timeafter-volumeaveraging procedure. The fairly generalform <strong>of</strong> the constitutive relations adopted for the bubblecoalescence <strong>and</strong> breakage phenomena are formulatedpurely on the basis <strong>of</strong> physical reasoning <strong>and</strong> intuitiveinterpretations <strong>of</strong> the mechanisms involved. Furtherlinks to the discrete particle-scale phenomena areusually expressed on the averaging scales by extrapo-


5126 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005lating relationships <strong>and</strong> concepts from the kinetic theory<strong>of</strong> gases. However, the basic kinetic theory <strong>of</strong> gases doesnot provide any coupling between the particles <strong>and</strong> theambient fluid. In a more generalized framework, however,the kinetic theory concept allows for the introduction<strong>of</strong> external forces determining the interactionbetween the fluid <strong>and</strong> the dispersed particles. Theapproach described by Ramkrishna 127 is outlined.Macroscopic Population Balance Equation. Inthis subsection, the macroscopic population balanceformulation <strong>of</strong> Luo 70 is outlined. In the work <strong>of</strong> Luo, 70no growth terms were considered; the balance equationthus contained a transient term, a convection term, <strong>and</strong>four source terms due to binary bubble coalescence <strong>and</strong>breakage. The population equation was expressed as∂n i∂t + ∇‚(v i n i ) ) B B,i - D B,i + B C,i - D C,i ( 1sm 3)(28)where n i is the number density with units <strong>of</strong> 1/m 3 <strong>and</strong>v i is the mass-average velocity vector, v i ) (∑ N i)1 -n i v i,p F g V i )/(∑ Ni)1 n i F g V i ). [Note that, alternatively, theparticle number-average, v i,ni ) (∑ N i)1 n i v i,p )/(∑ N i)1 n i );size-average, v i,di ) (∑ N i)1 n i v i,p d i )/(∑ Ni)1 n i d i ); surface-average,v i,ai ) (∑ N i)1 n i v i,p a i )/(∑ Ni)1 n i a i ); or volume-average,v i,Vi ) (∑ N i)1 n i v i,p V i )/(∑ Ni)1 n i V i ) velocity vector could beused as well. 140 ] The source terms express the bubblenumber birth <strong>and</strong> death rates per unit dispersionvolume for bubbles <strong>of</strong> size d i at time t due to coalescence<strong>and</strong> breakage. The source terms are assumed to befunctions <strong>of</strong> bubble size d i , bubble number n i , <strong>and</strong> timet.The birth <strong>of</strong> bubbles <strong>of</strong> size d i due to coalescence stemsfrom the coalescence between all bubbles <strong>of</strong> size smallerthan d i . Hence, the birth rate for bubbles <strong>of</strong> size d i , B C,i ,can be obtained by summing all coalescence events thatform a bubble <strong>of</strong> size d i . This givesB C,i )d i /2∑ Ω C (d j :d i - d j )d j )d min( 13) (29)smwhere d min is the minimum bubble size <strong>and</strong> depends onthe minimum eddy size in the system. The source termdefinition implies that bubbles <strong>of</strong> size d j coalesce withbubbles <strong>of</strong> size (d i - d j ) to form bubbles <strong>of</strong> size d i . Theupper limit <strong>of</strong> the sum stems from symmetry considerationsor avoidance <strong>of</strong> counting the coalescence betweenthe same pair <strong>of</strong> bubble sizes twice.Similarly, the death <strong>of</strong> bubbles <strong>of</strong> size d i due tocoalescence stems from coalescence between two bubblesin class d i or between one bubble in class d i <strong>and</strong> otherbubbles. Hence, the bubble death rate for bubbles <strong>of</strong> sized i , D C,i , can be calculated asd max -d iD C,i ) ∑ Ω C (d j :d i )d j )d min( 13) (30)smwhere d max is the maximum bubble size in the system.The upper limit indicates that the bubble size formedby coalescence must not exceed d max .The birth <strong>of</strong> bubbles <strong>of</strong> size d i due to breakage stemsfrom the breakage <strong>of</strong> all bubbles larger than d i . Thebreakage birth rate, B B,i , can be obtained by summingall breakage events that form bubbles <strong>of</strong> size d id maxB B,i ) ∑ Ω B (d j :d i )d j )d i( 13) (31)smThe death <strong>of</strong> bubbles <strong>of</strong> size d i due to breakage stemsfrom breakage <strong>of</strong> bubbles within this class; thusD B,i ) Ω B (d i )( 1sm 3) (32)The local gas volume fraction in any grid cell can becalculated as R g ) ∑ N i)1 n i (π/6)d 3 i .The macroscopic model formulation has numericalproperties similar to those <strong>of</strong> the discrete discretizationscheme (discussed later) <strong>and</strong> can be solved directlywithout further considerations.To ensure number <strong>and</strong> mass conservation, Hagesaetheret al. 73-75 extended this model by adopting anumerical procedure to redistribute the bubbles on pivotpoints in accordance with the discrete solution method.This modification <strong>of</strong> the solution strategy was introducedbecause <strong>of</strong> an inherent model limitation, i.e., thebreakage closure <strong>of</strong> Luo <strong>and</strong> Svendsen 71 was not necessarilyconservative, as will be explained later.Macroscopic Source Term Closures. In accordancewith the work <strong>of</strong> Coulaloglou <strong>and</strong> Tavlarides 135<strong>and</strong> Prince <strong>and</strong> Blanch, 77 Luo 70 assumed that allmacroscopic source terms determining the death <strong>and</strong>birth rates could be defined as the product <strong>of</strong> a collisiondensity <strong>and</strong> a probability. Thus, modeling <strong>of</strong> bubblecoalescence means modeling <strong>of</strong> a bubble-bubble collisiondensity <strong>and</strong> a coalescence probability, whereasmodeling <strong>of</strong> bubble breakage means modeling <strong>of</strong> aneddy-bubble collision density <strong>and</strong> a breakage probability.Models for the collision densities were derived assumingthat the mechanisms <strong>of</strong> the bubble-bubble <strong>and</strong>eddy-bubble collisions are analogous to those <strong>of</strong> collisionsbetween molecules as in the kinetic theory <strong>of</strong>gases. 135Models for the Binary <strong>Bubble</strong> Coalescence Rate,Ω C (d i :d j ). For coalescence between bubbles <strong>of</strong> class, d i ,<strong>and</strong> bubbles <strong>of</strong> class, d j , the coalescence rate is expressedasΩ C (d i :d j ) ) ω C (d i :d j ) p C (d i :d j )( 1sm 3) (33)Models for the <strong>Bubble</strong>-<strong>Bubble</strong> Collision Density,ω C (d i :d j ). Although not necessarily valid for bubblecolumns, the dispersions are considered sufficientlydilute that only binary collisions need to be considered.A collision <strong>of</strong> two bubbles can occur when the bubblesare brought together by the surrounding liquid flow orby body forces such as gravity. At least three sources <strong>of</strong>relative motion can be distinguished: motion inducedby turbulence in the continuous phase; motion inducedby mean velocity gradients; <strong>and</strong> motion induced bybuoyancy (or, more generally, body forces) arising fromdifferent bubble slip velocities, wake interactions, orhelical/zigzag trajectories. 140 However, most studies onbubble columns have been restricted to considering onlymodels for the contribution <strong>of</strong> turbulence to coalescence;the contributions from mean velocity gradients <strong>and</strong> bodyforces are generally neglected without proper validation.These collision mechanisms are generally both signifi-


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5127Figure 10. Sketch <strong>of</strong> a collision tube <strong>of</strong> an entering bubble movingthrough the tube with a velocity v p. The bubbles within the tubeare assumed to be “frozen” or stationary.cant, which greatly complicates the construction <strong>and</strong>validation <strong>of</strong> coalescence models.The parametrization <strong>of</strong> the collision densities wasobviously influenced by the more fundamental studieson formulating the source terms at the microscopic levelfollowed by some kind <strong>of</strong> averaging <strong>and</strong> a discretenumerical discretization scheme, as will be furtheroutlined in a later section. At this point, it is simplyobserved that the units <strong>of</strong> the volume-average coalescencefrequency correspond to the units <strong>of</strong> the effectiveswept volume rate in the kinetic theory <strong>of</strong> gases.Therefore, rather than the actual collision density, thevolume swept by a moving particle (i.e., bubble or eddy)multiplied by the particle density squared is used todefine the collision densities. 77,135,138To derive an expression for the effective swept volumerate, one usually considers a particle as it travels in astraight path from one collision to the next in amonodisperse dispersion, as sketched in Figure 10. Theparticle’s speed <strong>and</strong> direction <strong>of</strong> motion changes witheach collision. Further imagine that, at a given instant,all particles except for the one in question are frozen inposition <strong>and</strong> this particle moves with an average speed,v p . At the instant <strong>of</strong> collision, the center-to-centerdistance <strong>of</strong> the two particles is d. The collision crosssection <strong>of</strong> the target area <strong>of</strong> the particle is σ A ) 1 / 4 πd 2 .It should be noted that the collision cross-sectional areais defined in several ways. In one approach, inspiredby kinetic theory, the moving particle is approximatedas a point-like particle. Thus, the cross section <strong>of</strong> themoving particle is not considered in calculating theeffective cross-sectional area; hence, σ A ) 1 / 4 πd 2 .Inthecase <strong>of</strong> small particles, this approximation might besufficient. However, an improved estimate <strong>of</strong> the effectivecross-sectional area is achieved by taking intoaccount the size <strong>of</strong> the moving particle as well; thus σ A) πd 2 , as sketched in Figure 11. In time ∆t, the movingparticle sweeps out a cylindrical volume <strong>of</strong> length v p ∆t<strong>and</strong> cross section 1 / 4 πd 2 . Any particle whose center isin this cylinder will be struck by the moving particle.The number <strong>of</strong> collisions in the time ∆t is f 1 ( 1 / 4 πd 2 v p ∆t),where f 1 is assumed to be locally uniformly distributedin space. The effective swept volume rate h C (d) is givenbyh C (d) ) σ A v p ) 1 4 πd2 v p( m3s) (34)where σ A ) 1 / 4 πd 2 is the cross-sectional area <strong>of</strong> the socalled“collision tube”.Figure 11. Sketch <strong>of</strong> a collision tube <strong>of</strong> a bubble, defining thecollision diameter <strong>and</strong> the effective cross-sectional area.The collision density <strong>of</strong> a single particle is defined asthe number <strong>of</strong> collisions per unit time <strong>and</strong> length(diameter)ω f1(d) ) f 1 (d) h C (d) ) f 1 (d,t) 1 14 πd2 v p [(35)s (m)]This relation represents a very crude collision model fora dispersion containing only one type <strong>of</strong> particle.Sometimes, the collision density representing thenumber <strong>of</strong> collisions per unit mixture volume <strong>and</strong> perunit time <strong>and</strong> length (diameter) is a more convenientquantity. Multiplying the collision density <strong>of</strong> a singleparticle by the particle number density, one obtains themodified collision densityω f1 f 1(d) ) h C (d) f 1 f 1 ) 1 2 f 1 f 1114 πd2 v p[ m 3 s (m) (m)](36)The factor <strong>of</strong> 1 / 2 appears because h(d) represents twicethe number <strong>of</strong> collisions.In accordance with the pioneering work <strong>of</strong> Smoluchowski,141 similar considerations can be repeated fordispersions containing two types <strong>of</strong> particles havingdiameters d, d′ <strong>and</strong> particle densities f 1 , f′ 1 . The resultingcollision density is given byω C (d,d′) ) h C (d,d′) f 1 f′ 1 )πf 1 f′ 14( d + d′1vrel2 )2[ m 3 s (m) (m)] (37)where the collision frequency is calculated using themean collision diameter d ≡ (d + d′)/2, as sketched inFigure 12. Note that, in this case, the 1 / 2 factor is notincluded.Kolev 130 discussed the validity <strong>of</strong> these relations forfluid particle collisions considering the obvious discrepanciesresulting from the different nature <strong>of</strong> the fluidparticle collisions compared with the r<strong>and</strong>om molecularcollisions. The basic assumptions in kinetic theory thatthe molecules are hard spheres <strong>and</strong> that the collisionsare perfectly elastic <strong>and</strong> obey the classical conservationlaws do not hold for real fluid particles because theseparticles are deformable <strong>and</strong> elastic <strong>and</strong> can agglomerateor even coalesce after r<strong>and</strong>om collisions. The collisiondensity is thus not truly an independent function<strong>of</strong> the coalescence probability.


5128 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005subrange <strong>of</strong> turbulence, then δv 2 (d) ) C(ɛd) 2/3 . A relationshipbetween the constant C <strong>and</strong> the Kolmogorovconstant C k (i.e., the parameter in the energy spectrumfunction for the inertial subrange) can be deduced fromthe power-law spectrum <strong>and</strong> the second-order structurefunction, C ) (27/55)Γ(1/3)C k . 144,145 Setting C k ) 1.5 inaccordance with experimental data yields C ≈ 2.0.Interpreting the second-order structure function (i.e.,because it is a two-point correlation function) as arelative particle velocity (i.e., actually a mean velocityobtained using one-point statistics seems more appropriate),the collision density suggested by Prince <strong>and</strong>Blanch 77 becomesFigure 12. Sketch <strong>of</strong> the mean collision diameter <strong>and</strong> the effectivecross-sectional area.As mentioned above, bubble-bubble collisions canoccur as a result <strong>of</strong> a variety <strong>of</strong> mechanisms. Prince <strong>and</strong>Blanch 77 modeled bubble coalescence in bubble columnsconsidering bubble collisions due to turbulence, buoyancy,<strong>and</strong> laminar shear through analysis <strong>of</strong> the coalescenceprobability (efficiency) <strong>of</strong> collisions. It wasassumed that the collisions from the various mechanismswere cumulative.The collision density resulting from turbulent motionwas expressed as a function <strong>of</strong> bubble size, concentration,<strong>and</strong> velocity in accordance with the work <strong>of</strong>Smoluchowski 141ω T π(d,d′) ≈ f 1 f′ 14( d + d′ 22 )2 (vjt,d + vj 2 t,d′ ) 1/21m 3 s (m) (m)] (38)Estimates <strong>of</strong> the turbulent velocities have been obtainedusing certain relations developed in the classicaltheory on isotropic turbulence due to Kolmogorov. 142 Atthis point, the population balance closures seem tobecome rather vague, as the application <strong>of</strong> the classicalturbulence relations developed for fluid velocity fluctuationsor vortices to describe flows <strong>of</strong> discrete fluidparticles are not always in accordance with the basicturbulence theory restrictions. The theory <strong>of</strong> Kolmogorov142 states that, if the distance λ between two pointsin the flow field is much smaller than the turbulencemacroscale L but much larger than the Kolmogorovmicroscale λ d , then the second-order velocity structurefunction is a function <strong>of</strong> only the turbulent energydissipation rate ɛ <strong>and</strong> the magnitude <strong>of</strong> the length λ,δv 2 (λ) ) C(ɛλ) 2/3 . This two-point velocity structurefunction should not be confused with the normal component<strong>of</strong> the Reynolds stresses, which is a one-pointvelocity correlation function.A summary <strong>of</strong> the theory involved is given by Pope 143(Chapter 6 <strong>and</strong> Appendix G). By definition, the secondordervelocity structure function is the covariance <strong>of</strong> thedifference in velocity between two points in physicalspace. Considering the second-order structure functiondefined by Kolmogorov 142 for the continuous fluid flowturbulence, it can be expressed at a separation equal tothe bubble diameter d as δv 2 (d) ) [v z (z+d)-v z (z+d)] 2 .If the magnitude <strong>of</strong> diameter d lies within the inertial[ω T πC (d i :d j ) ≈ n i n j4( d i + d j2) 2 (vj t,di + vj t,dj ) 1/2 ≈0.089πn i n j (d i + d j ) 2 ɛ 1/3 (d 2/3 i + d 2/3 j ) ( 1/2 1 ) (39)sm 3(Note that the collision density was originally expressedin accordance with the macroscopic population balance.70 If otherwise acceptable, it seems that the coalescenceclosures formulated directly within the macroscopicframework can be transformed <strong>and</strong> expressedin terms <strong>of</strong> probability densities <strong>and</strong> used within theaverage microscopic balance framework without furtherconsiderations if one adopts a discrete numerical discretizationscheme.)Note that the parameter value used by Luo 70 <strong>and</strong> Luo<strong>and</strong> Svendsen 71 is a factor <strong>of</strong> 4 larger than this, one asthey used a different definition <strong>of</strong> the effective crosssectionalarea.In this interpretation, two bubbles being dispersed ina large-scale fluid eddy (i.e., larger than the bubblescale, d) are assumed to be advected together with thefluid without approaching each other.However, the validity <strong>of</strong> the Kolmogorov hypothesisdescribing bubbly flows is not yet strictly verified. Forconvenience, the relationship for turbulent fluid velocityfluctuations has thus been interpreted <strong>and</strong> extrapolatedin many ways. There seems to be some confusion in theliterature regarding different interpretations <strong>of</strong> theKolmogorov velocity scale <strong>and</strong> the second-order structurefunction. 143 The second-order structure function,δv 2 (d), has thus (apparently erroneously) been used asan estimate for an imaginary absolute turbulent bubblevelocity in the inertial subrange <strong>of</strong> isotropic turbulence,in other cases considered a relative bubble velocity in aturbulent flows, 77 <strong>and</strong> as a third possibility been useda measure <strong>of</strong> intrabubble oscillations.The buoyancy collision density ω B C (d i :d j ) is expressedas (Prince <strong>and</strong> Blanch; 77 Williams <strong>and</strong> Loyalka, 128 p 164)ω B πC (d i :d j ) ≈ n i n j4( d i + d j2) 2 |vj r,di - vj r,dj | ( 1 )sm 3 (40)where vj r,di is the rise velocity <strong>of</strong> the particle. (Again, notethat the collision density was originally expressed inaccordance with the macroscopic population balanceformulation. 70 )The functional form <strong>of</strong> the collision rate due tolaminar shear is given by (Friedl<strong>and</strong>er, 121 p 200; Williams<strong>and</strong> Loyalka, 128 p 170)


)3( ω LS 4C (d i :d j ) ≈ n i n j3 (d i + d dv cjdR) ( 3) 1 (41)smwhere v c is the continuous-phase circulation velocity <strong>and</strong>R is the radial coordinate <strong>of</strong> the column. The term(dv c /dR) is the average shear rate. [Again, the collisiondensity was originally expressed in accordance with themacroscopic population balance formulation (e.g, Luo 70 ).]The net coalescence frequency <strong>of</strong> bubbles <strong>of</strong> diametersd i <strong>and</strong> d j was then calculated by superposing thedifferent bubble collisions mechanisms as a linear sum<strong>of</strong> contributions multiplied by a common efficiency (asimilar closure can be obtained for the average microscopicformulation if one adopts a discrete numericaldiscretization scheme, as will be described later) 77p C ∝ f(∆t coal /∆t col ) (44)The coalescence efficiency thus represents the fraction<strong>of</strong> particles that coalesce out <strong>of</strong> the total number <strong>of</strong>particles that have been colliding. The functional relationshipgiven by Coulaloglou <strong>and</strong> Tavlarides, 135 i.e.p C ≈ exp(-∆t coal /∆t col ) (45)is <strong>of</strong>ten used for the modeling <strong>of</strong> bubble coalescenceprobabilities for bubble column flows.Ω C (d i :d j ) ) [ω T C (d i :d j ) + ω B C (d i :d j ) +The duration <strong>of</strong> such interactions is limited, <strong>and</strong>coalescence will occur only if the intervening film canω LS C (d i :d j )]p C (d i :d j )( 1drain to a sufficiently small thickness to rupture in thesm3)(42)time available.Coalescence is obviously possible if there is enoughHagesaether 146 adopted this approach when modeling contact time for completing the coalescencebubble column dispersions <strong>and</strong> found that, with thechoice <strong>of</strong> parameter values used in his analysis, the∆t col > ∆t coal (46)turbulent contribution dominated the collision rate forthe bubbles in the system. However, the most important Luo, 70 Luo <strong>and</strong> Svendsen, 71 <strong>and</strong> Hagesaether et al. 73-75model parameter as the turbulent energy dissipation adopted this approach.rate is difficult to compute, <strong>and</strong> only crude estimates The complexity <strong>of</strong> the film draining phenomenawere obtained. 32,70,77 In addition, Kolev 130 argued that involved is a severe problem <strong>and</strong> might be best illustratedby briefly discussing an early modeling at-the frequency <strong>of</strong> coalescence <strong>of</strong> a single bubble shouldrather be determined for individual efficienciestempt. Oolman <strong>and</strong> Blanch 147 derived an expression forthe coalescence times in stagnant fluids by examiningΩ C (d i :d j ) ) ω T C (d i :d j ) p T C (d i :d j ) + ω B C (d i :d j ) p B C (d i :d j ) + the time required for the liquid film between bubblesto thin from an initial thickness to a critical value whereω LS C (d i :d j ) p LS C (d i :d j ),( 13) sm (43)Kolev 130 further assumed that, for the coalescenceprocesses induced by buoyancy <strong>and</strong> nonuniform velocityfields, the forces leading to collisions inevitably acttoward coalescence for contracting particle free paths.Therefore, the probabilities must have the followingranges: 0 E p C T (d i :d j ) E 1 <strong>and</strong> p C B (d i :d j ) ) p C LS (d i :d j ) ≈ 1.However, no firm conclusions on the relative importance<strong>of</strong> the various contributions have yet been drawn,as the turbulent particle velocity closures are at bestinaccurate.Models for the Probability <strong>of</strong> Coalescence, p C -(d i :d j ). The probability or efficiency <strong>of</strong> oscillatory bubblecoalescence, e.g., induced by turbulent fluctuations, isexpected to be determined by physical mechanisms onvarious scales. The coalescence process is assumed tooccur in three consecutive stages. First, bubbles collide,trapping a small amount <strong>of</strong> liquid between them underthe action <strong>of</strong> the continuous phase. Second, this liquidthen drains over a period <strong>of</strong> time from its initialthickness until the liquid film separating the bubblesreaches a critical thickness, under the action <strong>of</strong> the filmhydrodynamics. The hydrodynamics <strong>of</strong> the film dependson whether the film surface is mobile or immobile, <strong>and</strong>the mobility, in turn, depends on whether the continuousphase is pure or a solution. Third, at this point, filmrupture occurs because <strong>of</strong> film instability, resulting ininstantaneous coalescence (i.e., usually not modeled infurther detail). The probability <strong>of</strong> coalescence is thusdefined as a function <strong>of</strong> the ratio <strong>of</strong> the time intervalwithin which the bubbles are touching each other, calledInd. Eng. Chem. Res., Vol. 44, No. 14, 2005 5129the collision time interval, ∆t col , <strong>and</strong> the time intervalrequired to push out the surrounding liquid <strong>and</strong> toovercome the strength <strong>of</strong> the capillary microlayer betweenthe two bubbles, called the coalescence timeinterval, ∆t coal130rupture occurs. The original model considers the flowrate <strong>of</strong> fluid from the liquid film by capillary pressure,augmented by the Hamaker contribution (reflecting themutual attraction <strong>of</strong> fluid molecules on opposite sides<strong>of</strong> the liquid film) at very low film thicknesses, bubbledeformation, <strong>and</strong> the changes in the concentration <strong>of</strong>surfactant species- dhdt { ) 8R 2 d F c[ -4cRT( dσ Idc) 2( 2 + h 2σ 3)]}I+ A 1/2r b 6πh(47)where h is the film thickness, R d is the radius <strong>of</strong> theliquid disk between the coalescing bubbles, R is the gasconstant, T is the temperature, A is the Hamakerconstant, σ I is the surface tension, <strong>and</strong> c is the concentration<strong>of</strong> a surfactant species.To solve this equation, sufficient initial conditions arerequired. The rather arbitrary initial thickness used forthe film in air-water systems was set to h 0 ) 1 × 10 -4m, whereas the final thickness was set at h f ) 1 × 10 -8m. However, in practice, several <strong>of</strong> the effects includedin the model are very cumbersome to describe, <strong>and</strong> noaccurate solution <strong>of</strong> this equation have yet been found.Several simplifying assumptions are usually adopted(e.g., no Hamaker contributions, no surface impurities,simplified interface geometry, etc.), making analyticalsolutions possible. A typical coalescence time was foundby integration, ∆t ij ) (r ij 3 F c /16σ I ) 1/2 ln(h 0 /h f ). r ij denotesthe equivalent bubble radius.A similar collision time, ∆t col , was suggested by Luo 70for the flowing bubble column system, assuming deformable<strong>and</strong> fully mobile interfaces


5130 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005∆t col ) 2t max ≈ (1 + ξ ij) (F d /F c + C V ) 3F c d i3(1+ ξ 2 ij )(1 + ξ 3 ij ) σ I(48)where the bubble ratio is ξ ij ) d i /d j , C V is the addedmasscoefficient, <strong>and</strong> t max is the time between the firstcontact <strong>and</strong> when the film area between the twocolliding bubbles reaches its maximum value.The corresponding coalescence time, ∆t coal , first suggestedby Chester 148 <strong>and</strong> further extended by Luo, 70yields∆t coal ≈ 0.5F c vj ij d 2i(49)(1 + ξ ij ) 2 σ Iwhere vj ij ) (vj i2 + vj j2 ) 1/2 ) vj i (1 + ξ-2/3 ij ) 1/2 <strong>and</strong> vj i is themean turbulent bubble velocity.To determine the mean approach velocity <strong>of</strong> bubblesin turbulent collisions, a series <strong>of</strong> unfortunate assumptionswas made by Luo <strong>and</strong> Svendsen. 71 First, inaccordance with earlier work on fluid particle coalescence,77,135 the colliding bubbles were assumed to takethe velocity <strong>of</strong> the turbulent fluid eddies having thesame size as the bubbles. Luo <strong>and</strong> Svendsen 71 furtherassumed that the turbulent eddies in liquid flows wouldhave approximately the same velocity as neutrallybuoyant droplets in the same flow. Utilizing the experimentalresults obtained in an investigation on turbulentmotion <strong>of</strong> neutrally buoyant droplets in stirred tanksreported by Kuboi et al., 149,150 the mean square dropletvelocity was expressed as v 2 rms (d) ) v 2 (d) ) 2.0(ɛd) 2/3 .The experimental data also indicated that the turbulentvelocity component distributions <strong>of</strong> droplets follow aMaxwellian distribution function (a brief description <strong>of</strong>the properties <strong>of</strong> the Maxwellian state is given byGidaspow 120 ); thus, vj drops ) (8v 2 rms (d)/3π) 1/2 ) 1.70(ɛd) 1/3 . It was further noticed that the value <strong>of</strong> thecoefficient was sensitive to the density ratio betweenthe continuous <strong>and</strong> dispersed phases, F d /F c . This approachwill be discussed in further detail in the breakagesection. Note, however, that Brenn et al. 151 investigatedunsteady bubbly flow with very low void fractions<strong>and</strong> concluded that the velocity probability densityfunctions <strong>of</strong> bubbles <strong>and</strong> liquid are better describedusing two superimposed Gaussian functions.Hagesaether et al. 72 derived a model for film drainagein turbulent flows <strong>and</strong> studied the predictive capabilities.In line with all other attempts to estimate thesetime scales, it was concluded that the present modelsare not sufficiently accurate <strong>and</strong> that sufficiently detaileddata on bubbly flows are not available for modelvalidation. For droplet flows, it was found that the puredrainage process (without interfacial mass-transferfluxes) was predicted with fair accuracy, whereas noreliable coalescence criteria was found (see also Klaseboeret al. 152,153 ). Furthermore, it was concluded that ahead-on collision is not representative for all possibleimpact parameters. Orme 154 <strong>and</strong> Havelka et al., 155among others, noticed that the impact parameter is <strong>of</strong>great importance for the droplet-droplet collision outcomein gas flows. However, no such collision outcomemaps have yet been published for bubble-bubble collisions.Saboni et al. 156,157 developed drainage modelsfor partially mobile plane films to describe film drainage<strong>and</strong> rupture during coalescence in liquid-liquid dispersions,taking into account the interfacial tension gradientsgenerated by interfacial mass transfer. Theresulting Marangoni forces accelerated the film drainage,which, in general, corresponds to dispersed-tocontinuousphase transfer <strong>and</strong> diminishes film drainagein the negative case. Similar effects are expected tooccur for the gas-liquid systems operated in bubblecolumns, but no detailed experimental analysis on gasliquiddispersions has yet been reported.Models for the Macroscopic Breakage Rate, Ω B -(d i ,d j ). During the past decade, considerable attentionhas been focused on the macroscale modeling <strong>of</strong> bubblebreakage in gas-liquid dispersions. 67,77,137,145,158 A briefoutline <strong>of</strong> the important milestones is given next.Fluid particle breakage controls the maximum bubblesize <strong>and</strong> can be greatly influenced by the continuousphasehydrodynamics <strong>and</strong> interfacial interactions. Therefore,a generalized breakage mechanism can be expressedas a balance between external stresses, σ ij +σ t,ij , that attempt to disrupt the bubble <strong>and</strong> the surfacestress, σ I /d, that resists the particle deformation. Thus,at the point <strong>of</strong> breakage, these forces must balance, σ≈ σ I /(d/2). This balance leads to the prediction <strong>of</strong> acritical Weber number, above which the fluid particleis no longer stable. It is defined by 159We cr ) (σ ij + σ t,ij )d max2σ I(50)where d max is the maximum stable fluid particle size<strong>and</strong> σ reflects the hydrodynamic conditions responsiblefor particle deformation <strong>and</strong> eventual breakage.In the case <strong>of</strong> turbulent flow, particle breakage iscaused by velocity fluctuations resulting in pressurevariation along the particle surface (We cr ) F cv′ 2 c d max /2σ I ). In laminar flow, viscous shear in thecontinuous phase will elongate the particle <strong>and</strong> causebreakage [We cr ) (µ c (∂v z /∂r)d max )/2σ I ]. In the absence <strong>of</strong>net flow <strong>of</strong> the continuous phase such as rising bubblesin a liquid, the fluid particle breakage is caused byinterfacial instabilities due to Raleigh-Taylor <strong>and</strong>Kelvin-Helmholtz instabilities. 160In most bubble column analyses, the flow is consideredturbulent, <strong>and</strong> the effects <strong>of</strong> both viscous <strong>and</strong>interfacial instability are neglected basically withoutany further validation.The fluid particle fragmentation phenomena in ahighly turbulent flow are related to the fact that thevelocity in a turbulent stream varies from one point toanother (i.e., validated by two-point measurements 145 ).Therefore, different dynamic pressures will be exertedat different points on the surface <strong>of</strong> the fluid particle.Under certain conditions, this will inevitably lead todeformation <strong>and</strong> breakage <strong>of</strong> the fluid particle.According to Kocamustafaogullari <strong>and</strong> Ishii, 67 theforce due to the dynamic pressure can develop eitherthrough the local relative velocity around the particle,which appears because <strong>of</strong> inertial effects, or throughchanges in the eddy velocities over the length <strong>of</strong> theparticle.Most <strong>of</strong> the published literature on bubble breakageis derived from the theories outlined by Kolmogorov 161<strong>and</strong> Hinze. 159 In one phenomenological interpretation,bubble breakage occurs through bubble interactionswith turbulent eddies bombarding the bubble surface.If the energy <strong>of</strong> the incoming eddy is sufficiently highto overcome the surface energy, deformation <strong>of</strong> the


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5131surface results, which can finally lead to the formation<strong>of</strong> two or more daughter bubbles. For bubble breakageto occur, the bombarding eddies must be smaller thanor equal in size to the bubble, as larger eddies onlytransport the bubble. To model the breakup process, thefollowing simplifications are generally made: 71 (1) Theturbulence is isotropic. (2) Only binary breakage <strong>of</strong> abubble is considered. (3) The breakage volume ratio isa stochastic variable. (4) The occurrence <strong>of</strong> breakage isdetermined by the energy level <strong>of</strong> the arriving eddy. (5)Only eddies <strong>of</strong> a size smaller than or equal to the bubblediameter can cause bubble breakage.The basis <strong>of</strong> the macroscopic theories is the engineeringinterpretation considering the velocity fluctuationsas imaginary discrete entities denoted fluid slabs oreddies. Following this concept, a large number <strong>of</strong> eddiesexist in the flow having a size (diameter) distributionranging from the Kolmogorov microscale to the vesseldimensions.The basic ideas for the model development <strong>of</strong> Luo <strong>and</strong>Svendsen 71 were adopted from an earlier paper byCoulaloglou <strong>and</strong> Travlarides 135 considering droplet breakagein turbulent flows. In the model <strong>of</strong> Coulaloglou <strong>and</strong>Travlarides, 135 the breakage density was expressed asa product <strong>of</strong> an integral or average breakage frequency(∝ 1/t b ) <strong>and</strong> an integral (average) breakage efficiencydetermining the fraction <strong>of</strong> particles breaking. Thebreakage time (t b ) was determined from isotropic turbulencetheory, <strong>and</strong> the breakage efficiency was determinedfrom a probable fraction <strong>of</strong> turbulent eddiescolliding with droplets that have kinetic energy greaterthan the droplet surface energy.Prince <strong>and</strong> Blanch 77 further postulated that bubblebreakage is a result <strong>of</strong> collisions between particles <strong>and</strong>turbulent eddies <strong>and</strong> that the collision density can becalculated following arguments from the kinetic theory<strong>of</strong> gases. An integral (average) breakage densityΩ B (d i :d j ), with units 1/(s m 3 ), was thus expressed as aproduct <strong>of</strong> an average eddy-bubble collision densityω B (d i :d j ) <strong>and</strong> an average breakage efficiency p B (d i :d j );thus, Ω B (d i :d j ) ) ω B (d i :d j ) p B (d i :d j ). No explicit breakagefrequency function was given.Luo <strong>and</strong> Svendsen, 71 on the other h<strong>and</strong>, consideredthe individual eddy-particle collisions <strong>and</strong> argued that,for a bubble to break, each <strong>of</strong> the colliding eddies musthave sufficient energy to overcome the increase inbubble surface energy, as illustrated in Figure 13, <strong>and</strong>have a size on the order <strong>of</strong> the bubble diameter.The differential breakage density was then expressedas a product <strong>of</strong> an eddy-bubble collision probabilitydensity, ω B,λ (d i ,λ), <strong>and</strong> a breakage efficiency, p B (d i :d j ,λ),both <strong>of</strong> which depended on the eddy size (λ). The totalbreakage rate for a bubble <strong>of</strong> size d i is thenΩ B (d i ) )d max∑ Ω B (d i :d j )d j )d min( 13) (51)smwhereas the individual rate <strong>of</strong> breaking a parent bubble<strong>of</strong> size d i into the daughter size classes d j is expressedasdΩ B (d i :d j ) ) ∫ T ωB λmin(d i ,λ) p B (d i :d j ,λ) dλ( 13) (52)sm<strong>and</strong> the simple eddy-bubble collision probability density,ω B T (d i ,λ), has units 1/[s m 3 (m)]. The upper integrationlimit for the eddy size is based on the modelFigure 13. Sketch <strong>of</strong> the bubble breakage surface energy balance.The mean kinetic energy <strong>of</strong> an eddy <strong>of</strong> size λ breaking a bubble <strong>of</strong>size d i, ej(d i,λ), is assumed to be larger than the increase <strong>of</strong> thebubble surface energy required breaking the parent bubble d i intoa daughter bubble d j <strong>and</strong> a second corresponding daughter bubble,e s(d i,d j).Figure 14. Sketch <strong>of</strong> a collision tube <strong>of</strong> an entering eddy movingthrough the tube with a velocity vj λ. The bubbles within the tubeare assumed to be frozen or stationary.assumption that only eddies <strong>of</strong> size smaller than orequal to the bubble diameter can cause bubble breakage.A bubble being dispersed in a large-scale fluid eddy(larger than the bubble scale, d) is assumed to beadvected together with the eddies in the fluid. Unfortunately,the breakage frequency is found to be highlysensitive to the choice <strong>of</strong> integration limits. 146,158 Notethat no explicit breakage frequency was given.Luo <strong>and</strong> Svendsen 71 considered the collision density<strong>of</strong> eddies with a given velocity, vj λ , bombarding anumber, n i , <strong>of</strong> locally frozen bubbles, as illustrated inFigure 14ω B T (d i ,λ) ) n i f λ h C (d i ,λ) ) n i f λπ4 (d i + λ)2 vj λwhere f λ is the number <strong>of</strong> eddies <strong>of</strong> size between λ <strong>and</strong>λ + dλ with units <strong>of</strong> 1/[m 3 s (m)] <strong>and</strong> vj λ is the turbulentvelocity <strong>of</strong> eddies <strong>of</strong> size λ. Politano et al. 162 defined theeffective collision cross-sectional area as σ A ) (π/4)[(d i+ λ)/2] 2 , in accordance with the st<strong>and</strong>ard definition inkinetic theory (see also Kolev 130 ).The mean turbulent velocity <strong>of</strong> eddies with size λ inthe inertial subrange <strong>of</strong> isotropic turbulence was assumedto be equal to the velocity <strong>of</strong> the neutrally[1sm (m)] (53)3


5132 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005buoyant droplets measured by Kuboi et al. 149,150 Kuboiet al. 149,150 found that the turbulent velocity <strong>of</strong> dropletscould be expressed by the Maxwell distribution function,giving the mean eddy velocity <strong>of</strong>1/2vj λ ) [ 8v2 (λ)3π](54)where v 2 (λ) ) (8β˜/2π)(ɛλ) 2/3 is in accordance with theclassic turbulence theory. The coefficient value is coincidentallythe same as that obtained by Kuboi etal., 149,150 (8β˜/3π) ≈ C ≈ 2.0. However, it was commentedby Kuboi et al. 149,150 that the comparison <strong>of</strong> theserelations cannot be very decisive, in view <strong>of</strong> the fact thatthere is a large difference between the processes <strong>and</strong>types <strong>of</strong> fluids used to obtain these relations. Kuboi etal. 149,150 also investigated the effect <strong>of</strong> a difference inparticle fluid density producing nonneutrally buoyantparticle flows <strong>and</strong> concluded that the parameter valuediscussed above is very sensitive to the density ratio.Therefore, the application <strong>of</strong> the above relation as anapproximation for the bubble velocity is highly questionable.To employ eq 53, the number probability density <strong>of</strong>eddies <strong>of</strong> a particular size must be determined. Luo <strong>and</strong>Svendsen 71 assumed that the turbulence is isotropic <strong>and</strong>that the eddy size <strong>of</strong> interest lies in the inertialsubrange. An expression for the number probabilitydensity <strong>of</strong> eddies as a function <strong>of</strong> wavelength for theseconditions was formulated using conceptual ideas fromAzbel 163 (p 85) <strong>and</strong> Azbel <strong>and</strong> Athanasios. 164 Theturbulent energy spectrum function, E(k), can be interpretedas the kinetic energy contained within eddies <strong>of</strong>wavenumbers between k <strong>and</strong> k + dk, or equivalently,<strong>of</strong> size between λ <strong>and</strong> λ + dλ, per unit mass. Arelationship between f λ <strong>and</strong> E(k) can thus be obtainedby formulating an energy balance for eddies beinginterpreted both as discrete entities <strong>and</strong> as a wavefunctionJE eddies (λ) ) E spectra (λ) (55)[ m (m)] 3f λ[ 2( 1 F πL6 ) λ3 vj λ2] ) E(k)F L (1 -R g( ) - dkThusJdλ) [ m (m)] 3(56)The functional form <strong>of</strong> the energy spectrum in theinertial subrange <strong>of</strong> turbulence is defined as 143E(k) ) C k ɛ 2/3 k -5/3[ m2s 2 (m) ](57)<strong>and</strong> the relationship between the wavenumber <strong>and</strong> thesize <strong>of</strong> the eddy (wavelength) is k ) 2π/λ (dk/dλ )-2πλ -2 ), vj λ 2 ) β(ɛλ) 2/3 , β ) 8β˜/3π, <strong>and</strong> β˜ ) 3 / 5 Γ( 1 / 3 )C k )2.41 [Γ( 1 / 3 ) ≈ 2.6789.] [The value <strong>of</strong> the latter parameteris difficult to determine as different authors give differentvalues: Luo 70 used β˜ ) 3 / 5 Γ( 1 / 3 )C k ) 2.41; Batchelor144 (p 123) defined β˜ ) 9 / 5 Γ( 1 / 3 )C k ) 7.23; Pope 143 (p232) obtained another value, β˜ ≈ 2.0; Martínez-Bazánet al. 165 referred to Batchelor 166 <strong>and</strong> claimed that β˜ ≈8.2; Risso <strong>and</strong> Fabre 145 referred to Batchelor 167 (p 120)<strong>and</strong> defined β˜ ) 27 / 55 Γ( 1 / 3 )C k ) 1.97, which is about thesame as the value given by Pope; 143 <strong>and</strong> Lasheras etal. 158 reviewed several breakage models <strong>and</strong> pointed outthat the value <strong>of</strong> this parameter ranges from about β˜ )2.045 to β˜ )8.2. Because <strong>of</strong> the disagreement on thisvalue in the literature, we have chosen to proceed withthe value <strong>of</strong> Luo. 70 ) The number density <strong>of</strong> eddies, f λ ,was thus defined as9C k 1f λ ) (1 -R g )β˜2 5/3 π 2/3 λ ) 0.822(1 -R g ) 14 λ [4 m (m)] 3(58)where R g is the volume fraction <strong>of</strong> the dispersed phase.However, it is not clear whether Luo <strong>and</strong> Svendsen 71distinguish between the theoretical parameter value forthe turbulent eddy velocities <strong>and</strong> the empirical parameterfor the bubble velocity in the model implementation.In this report, the “theoretical value” is adopted for theturbulent eddy properties.The collision density <strong>of</strong> eddies with sizes between λ<strong>and</strong> dλ on particles <strong>of</strong> size d i can be expressed asω B T (d i ,λ) ) π 4 (0.822)2.045(1 -R g )n i (ɛλ)1/3(d i + λ)2In dimensionless variables, this isω B T (ξ) ) 0.923(1 -R g )(ɛd i ) 1/3 n i(1 + ξ) 2where ξ ) λ/d i .To close the collision density model, a breakageprobability (efficiency) is needed. For each particulareddy hitting a particle, the probability for particlebreakage was assumed to depend not only on the energycontained in the arriving eddy but also on the minimumenergy required to overcome the surface area increasedue to particle fragmentation. The latter quantity wasdetermined by the number <strong>and</strong> sizes <strong>of</strong> the daughterparticles formed in the breakage processes. To determinethe energy contained in eddies <strong>of</strong> different scales,a distribution function <strong>of</strong> the kinetic energy for eddiesin turbulent flows is required. A Maxwellian distributionfunction might be a natural <strong>and</strong> consistent choice, 136as the eddy velocity is assumed to follow this distribution,but Luo <strong>and</strong> Svendsen 71 preferred an empiricalenergy distribution density function for fluid particlesin liquid developed by Angelidou et al. 168 The breakageprobability function used yields(Notice that this efficiency function is not necessaryvolume-or mass-conserving, as Luo <strong>and</strong> Svendsen 71considered the breakage efficiency function to be equalto the kinetic energy distribution function. It wouldprobably be better to consider the breakage distributionfunction as purely proportional to the empirical kineticenergy distribution function <strong>and</strong> determine the probabilityconstant by requiring bubble volume or mass[λ 41sm (m)] (59)3d i 2 ξ 11/3 (60)p B (d i :d j ,λ) ) 1 - ∫ 0e s,c 1ej(d i ,λ) exp [ - e s (d i ,d j )ej(d i ,λ) ] de′ s )1 - ∫ 0χ cexp(-χ) dχ′ ) exp(-χc ) (61)


flows in an agitated tank. The model <strong>of</strong> Luo <strong>and</strong>where C f (f Vij ) is defined as the increase coefficient <strong>of</strong>Svendsensurface area, which depends only on the breakagewas in very good agreement with thevolume fraction, f Vij ) d 3 j /d 3 experimental data, whereas the model <strong>of</strong> Prince <strong>and</strong>i . The e s (d i ,d j ) function isBlanchsymmetrical about the breakage volume fraction, f Vij )appears to overpredict the bubble size.0.5.In summary, most <strong>of</strong> the models discussed so far areCombining eqs 60, 61, <strong>and</strong> 52, the breakage density based on eddy collision arguments that rely on the<strong>of</strong> one particle <strong>of</strong> size d i that breaks into particles <strong>of</strong> interpretation that turbulent flows consist <strong>of</strong> a collectionsizes d j <strong>and</strong> (d 3 i - d 3 j ) 1/3 is given by<strong>of</strong> fluid slabs or eddies that are treated using relationshipsdeveloped in the kinetic theory <strong>of</strong> gases. Thisd engineering eddy concept is impossible to validateΩ B (d i :d j ) ) ∫ ωB λmin(d i ,λ)p B (d i :d j ,λ)dλ( 1regarding the eddy shape, the number densities, <strong>and</strong>smthe breakage mechanisms discussed above. Further, the) 0.923(1 -R resulting breakage models require the specification <strong>of</strong>g )n i( ɛ 1/3(64)d the minimum <strong>and</strong> maximum eddy sizes that are capablei2)<strong>of</strong> causing particle breakage, as well as a criterion for1 (1 + ξ) the minimum bubble size d min . Finally, the macroscopic∫ ξminformulation is considered less fundamental <strong>and</strong> lessξexp( - 12C f σ I11/3 βF c ɛ 2/3 d 5/3 i ξ dξ general than the average microscopic one (discussed inInd. Eng. Chem. Res., Vol. 44, No. 14, 2005 5133conservation within the breakage process.) In eq 61, χ) e s (d i ,d j )/ej(d i ,λ) is a dimensionless energy ratio, whereej(d i ,λ) is the mean kinetic energy <strong>of</strong> an eddy <strong>of</strong> size λ<strong>and</strong> e s (d i ,d j ) is the increase in surface energy when aparticle <strong>of</strong> diameter d i is broken into two particles <strong>of</strong>size d j <strong>and</strong> (d 3 i - d 3 j ) 1/3 . A critical dimensionless energyfor breakage to occur was thus defined as χ c ) [e s (d i ,d j )/<strong>and</strong> the initial force <strong>of</strong> the colliding eddy balance eachother. In contrast, Hagesaether et al. 74 <strong>and</strong> Wang etal. 137 assumed that, during breakage, the inertial force<strong>of</strong> the colliding eddy is <strong>of</strong>ten larger than the interfacialforce <strong>and</strong> bubble deformation is strengthened untilbreakage occurs. Therefore, the force balance <strong>of</strong> Lehret al. 32 might not be satisfied during bubble breakage.ej(d i ,λ)]| c ) [-(12C f σ I )/βF c ɛ 2/3 d 5/3 i ξ 11/3 ]| c .Politano et al. 162 studied the equilibrium betweenThe mean kinetic energy <strong>of</strong> an eddy with size λ, ej(d i ,λ), coalescence <strong>and</strong> breakage in homogeneous flows withwas expressed asisotropic turbulence using a population balance model.The population balance was solved using a multigroup2πapproach. The daughter size distribution function wasej(d i ,λ) )F c6 λ3vj λ2 )F 2βc3 λ11/3 ɛ 2/3 2β)F c3 ξ11/3 d 11/3 i ɛ 2/3approximated by a uniform function, by a delta function,(62) <strong>and</strong> by the model proposed by Luo <strong>and</strong> Svendsen. 71 Thewhereas the increase in surface energy was given bye s (d i ,d j ) ) πσ I [d 2 j + (d 3 i - d 3 j ) 1/3 - d 2 i ] )breakage rate was calculated using either the modelproposed by Luo <strong>and</strong> Svensen 71 or the model proposedby Prince <strong>and</strong> Blance. 77 Significant differences in theresulting bubble breakage rate, <strong>and</strong> therefore in theπσ bubble size distribution, were observed comparing theI d 2 i [f 2/3 Vij+ (1 - f Vij) 2/3 - 1] ) C f (f Vij)πσ I d 2 i (63)model performance with experimental data on bubblythe next section). An unfortunate consequence <strong>of</strong> thiswhere ξ min ) λ min /d, λ min /λ d ≈ 11.4 - 31.4, <strong>and</strong> λ d is themodeling framework is that a discrete numerical schemeKolmogorov microscale. Recall that C f is a function <strong>of</strong>d j , i.e., C f ) C f (f Vij ) ) C f (d 3 j /d 3 for the particle size is embedded within the source termi ). No explicit breakageclosures; thus, optimizing the numerical solution procedureis not a trivial task.frequency was determined in this model.A severe limitation <strong>of</strong> the original breakage densityclosure <strong>of</strong> Luo <strong>and</strong> Svendsen 71 is that the modelMicroscopic Population Balance Formulation.produces too many very small particles regardless <strong>of</strong> theThe dispersed-phase system is considered as a population<strong>of</strong> particles <strong>of</strong> the dispersed phase distributed notnumerical size resolution used; the model is thus notgrid-independent (i.e., in the particle size grid). Severalonly in physical space (i.e., in the ambient continuouscontributions have focused on the modification <strong>of</strong> thephase) but also in an abstract property space. 122,124 Inbasic model <strong>of</strong> Luo <strong>and</strong> Svendsen 71 intending to avoidthe terminology <strong>of</strong> Hulburt <strong>and</strong> Katz, 122 one refers tothis limitation. 32,74,137the spatial coordinates as external coordinates <strong>and</strong> theHagesaether et al. 74 <strong>and</strong> Wang et al. 137 extended theproperty coordinates as internal coordinates. The jointbasic breakage model <strong>of</strong> Luo <strong>and</strong> Svendsen 71 by assumingthat, when an eddy <strong>of</strong> size λ has kinetic energy e(λ),space <strong>of</strong> internal <strong>and</strong> external coordinates is referredto as the particle state space.the daughter bubble size is limited by two constraints: The quantity <strong>of</strong> basic interest is the average numberOne is that the dynamic pressure <strong>of</strong> the turbulent eddy <strong>of</strong> particles per unit volume <strong>of</strong> the particle state space.1 / 2 F c vj λ2 must be larger than the capillary pressure The population balance is thus an equation for theσ I /d′′, resulting in a minimum breakage fraction f Vmin (or number density <strong>and</strong> can be regarded as representing abubble size d j,min ). d′′ denotes the diameter <strong>of</strong> the smaller number balance on particles <strong>of</strong> a particular state.daughter particle (or two times the minimum radius <strong>of</strong>curvature). The second constraint is in accordance withthe work <strong>of</strong> Luo <strong>and</strong> Svendsen 71 that the eddy kineticenergy must be larger than the increase <strong>of</strong> the surfaceenergy during the breakage. These constraints resultin a maximum breakage fraction f Vmax (or bubble sized j,max ). Lehr et al. 32 derived a similar breakage densityfunction using only the capillary constraint <strong>and</strong> assumedthat the interfacial force on the bubble surfaceThe balance equation basically accounts for the variousways in which particles <strong>of</strong> a specific state can eitherform or disappear from the system. When particle statesare continuous (i.e., in the internal coordinates as well),then processes, which cause their smooth variation withtime, must contribute to the rates <strong>of</strong> formation <strong>and</strong>disappearance <strong>of</strong> specific particle types. Such processescan be viewed as convective processes because theyresult from convective motion in particle state space.


5134 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005The number <strong>of</strong> particles <strong>of</strong> a particular type can changeby processes that create new particles (“birth” processes)<strong>and</strong> destroy existing particles (“death” processes). Inbubble columns, birth <strong>of</strong> new particles can occur throughboth breakage <strong>and</strong> coalescence processes. The bubblebreakage <strong>and</strong> coalescence processes also contribute todeath processes, as a particle type that either breaks(into other particles) or coalesces with another particleno longer exists as such following the event. Thephenomenological utility <strong>of</strong> population balance modelslies in the convective processes as well as the birth <strong>and</strong>death processes.Let f 1 (x,r,t) be the average number <strong>of</strong> particles perunit volume <strong>of</strong> the particle state space at time t, atlocation x ≡ (x 1 , x 2 , ..., x d ) in property space (d representingthe number <strong>of</strong> different quantities associatedwith the particle) <strong>and</strong> r ≡ (r 1 , r 2 , r 3 ) in physical space.This average number density function is considered asmooth function <strong>of</strong> its arguments x, r, <strong>and</strong> t. Thus,f 1 (x,r,t) can be differentiated as many times as desiredwith respect to any <strong>of</strong> its arguments.The continuous-phase variables are represented by avector Y(r,t) ≡ (v c , v d , R c , R d , p, F c , F d , ...) that iscalculated from the governing continuum transportequations <strong>and</strong> the problem-dependent boundary conditions.If dV x <strong>and</strong> dV r denote infinitesimal volumes in propertyspace <strong>and</strong> physical space, respectively, located at(x, r), then the average number <strong>of</strong> particles in dV x dV ris given by f 1 (x,r,t)dV x dV r . The local (average) numberdensity in physical space, that is, the total number <strong>of</strong>particles per unit volume <strong>of</strong> physical space, denotedN(r,t), is given byN(r,t) ) ∫ Vxf 1 (x,r,t) dV x (65)Whereas the rate <strong>of</strong> change <strong>of</strong> external coordinatesrefers to motion through physical space, that <strong>of</strong> internalcoordinates refers to motion through an abstract propertyspace. Separate velocities can then be defined forthe internal coordinates, X4 (x,r,Y,t), <strong>and</strong> for the externalcoordinates, R4 (x,r,Y,t). It is thus possible to identifyparticle (number) fluxes. f 1 (x,r,t) R4 (x,r,Y,t) representsthe particle flux through physical space, <strong>and</strong> f 1 (x,r,t)X4 (x,r,Y,t) is the particle flux through internal coordinatespace.Let ψ(x,r) be an extensive property (i.e., the propertychanges with the quantity <strong>of</strong> matter present) associatedwith a single particle <strong>of</strong> state (x, r). The amount <strong>of</strong> thisextensive property contained in all <strong>of</strong> the particles inthe material population above, denoted ψ(t), is given byψ(t) ) ∫ V(t)ψ(x,r) f 1 (x,r,t) dV′ (66)where V(t) is considered a combined abstract materialvolume (i.e., containing V r <strong>and</strong> V x ) <strong>of</strong> a hypotheticalmedium containing the embedded particles.Using the preceding nomenclature, a balance <strong>of</strong> theproperty (ψ) for the total population <strong>of</strong> entities simultaneouslyfound in the volume, V, can be formulated ina suitable frame <strong>of</strong> reference. R<strong>and</strong>olph 124 <strong>and</strong> R<strong>and</strong>olph<strong>and</strong> Larson, 125 as well as Ramkrishna, 126,127 adopted theLagrangian point <strong>of</strong> view. In accordance with thest<strong>and</strong>ard approach for deriving the multifluid model, 100the Eulerian point <strong>of</strong> view is used in this contribution.In this framework, the control volume (CV) is fixed withno local translational velocity (v CV ) 0).The balance principle applied to an Eulerian controlvolume (CV) is, by definition, expressed as a balance <strong>of</strong>accumulation (the term on the left-h<strong>and</strong> side <strong>of</strong> theequation), net inflow by convection (the first term onthe right-h<strong>and</strong> side <strong>of</strong> the equation), <strong>and</strong> volumetricproduction (the second term on the right-h<strong>and</strong> side <strong>of</strong>the equation).The Eulerian transport equations can thus be cast inthe generalized form(rate <strong>of</strong> ψ accumulation in the CV) )(net inflow rate <strong>of</strong> ψ by convectionacross the CV surface) +(net source <strong>of</strong> ψ within the CV)This approach thus consider the balance principle <strong>of</strong>the quantity ψ within a fixed abstract volume (V)bounded by a fixed abstract surface area Addt ∫ V (ψf 1 )dV′ )- ∫ A(vψf 1 )‚n dA′ + ∫ Vφ dV′ (67)In this notation, n is an abstract outward-directed unitvector normal to the surface <strong>of</strong> an abstract controlvolume, d/dt is an extended total time derivative operator,v is the phase-space velocity, ψ is the balancedquantity, <strong>and</strong> φ is the net source term.The term on the left-h<strong>and</strong> side <strong>of</strong> eq 67 can betransformed using an extended Leibnitz theorem intoan integral over only volume because the contributionsfrom the fixed abstract surfaces (A) vanish as thesurface phase-space velocities become zero. In terms <strong>of</strong>the previously defined nomenclature, R<strong>and</strong>olph <strong>and</strong>Larson 125 expressed a generalized theorem in the formd∫ dt (ψf V(t) 1 )dV′ ) ∫ V(t)[ ∂(ψf 1 )+ ∇‚(v∂tCV ψf 1 )]where v CV denotes a surface phase-space velocity vector(≡ X4 + R4 ). Adopting an Eulerian frame instead, thetheorem simplifies considerablyThe convective transport terms in eq 67 can berewritten as a volume integral using an extendedversion <strong>of</strong> Gauss’ theoremEquation 67 can then be rewritten as a volume integralEquation 71 must be satisfied for any V, so the expressionsinside the volume integral must be equal to zero.The local instantaneous equation for a general balancedquantity ψ then yields (eq 72)This relation can be referred to as a general transportequation.dV′ (68)d∫ dt (ψf V 1 )dV′ ) ∂(ψf∫1 )VdV′ (69)∂t∫ A(vψf 1 )‚n dA′ ) ∫ V∇‚(vψf 1 )dV′ (70)∫ V ( ∂(ψf 1 )∂t+ ∇‚(vψf 1 ) - φ) dV′ ) 0 (71)∂(ψf 1 )+ ∇‚(vψf∂t1 ) - φ ) 0 (72)


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5135The corresponding source term φ contains the birth<strong>and</strong> death terms representing the net rate <strong>of</strong> production<strong>of</strong> particles <strong>of</strong> state (x, r) at time t in an environment <strong>of</strong>state Y, thus φ(x,r,Y,t). Considering the bubble coalescence<strong>and</strong> breakage phenomena only, the function φ-(x,r,Y,t) can be expressed as φ(x,r,Y,t) ) B C (x,r,Y,t) -D C (x,r,Y,t) + B B (x,r,Y,t) - D B (x,r,Y,t). With ψ(x,r) )1, the generalized population balance framework becomes∂f(x,r,t)+ ∇∂tr ‚(f(x,r,t)v r ) + ∇ x ‚(f(x,r,t)X4 )) B B (x,r,Y,t) - D B (x,r,Y,t) + B C (x,r,Y,t) - (73)D C (x,r,Y,t)where f(x,r,t) is the bubble number probability density[1/m 3 (x)], which varies with internal coordinates (x),spatial position (r), <strong>and</strong> time (t).The first term on the LHS <strong>of</strong> eq 73 is the change inthe bubble number probability density with time. Thesecond term is the change in the bubble numberprobability density due to convection. The third term isthe change in the bubble number probability density dueto convection in the internal coordinates denotingseveral particle growth phenomena. The RHS is thesource terms. The source terms represent the changein the number probability density, f 1 , due to particlebreakup <strong>and</strong> coalescence. B B <strong>and</strong> D B represent the birth<strong>and</strong> death, respectively, <strong>of</strong> bubbles due to breakage, <strong>and</strong>B C <strong>and</strong> D C , respectively, are the birth <strong>and</strong> death <strong>of</strong>bubbles due to coalescence.To close this model formulation, constitutive relationsare needed for both the growth <strong>and</strong> source term functions.This process is <strong>of</strong> extreme importance as itrepresents the weakest part <strong>of</strong> population balancemodeling. Possibly, one could expect that these termsdepend on the number density function (kinetic theory)at the instant in question, but the detailed nature <strong>of</strong>these relationships are generally unknown <strong>and</strong> requirefurther attention. To some degree, the complexity <strong>of</strong> thegrowth terms is also dependent on the choice <strong>of</strong> internalcoordinates or particle properties used to characterizethe dispersed phase, as mentioned earlier. In mostreports, the particle volume (diameter) is used, so thegas expansion due to pressure, temperature, <strong>and</strong> compositionchanges as well as interfacial mass-transferfluxes has to be incorporated through this term. Incontrast, using the particle mass as the internal coordinate,only the interfacial mass-transfer fluxes haveto be incorporated through this term.However, it is still not feasible to solve the microscopicpopulation balance equations derived above by directnumerical simulations, so some kind <strong>of</strong> averaging isrequired to enable solution with reasonable time <strong>and</strong>physical space resolutions. In accordance with previouswork, 8 the microscopic transport equation is averagedover a physical averaging volume <strong>and</strong> thereafter overtime. The result is∂〈f(x,t)〉+ ∇∂t r ‚(〈f(x,t)〉〈v r 〉) + ∇ x ‚(〈f(x,t)〉〈X4 〉)) 〈B B (x,Y,t)〉 - 〈D B (x,Y,t)〉 + 〈B C (x,Y,t)〉 - (74)〈D C (x,Y,t)〉One can now introduce number-density-weighted averagevelocities, i.e.<strong>and</strong>This makes the velocities in the population balancedifferent from the mass- or phase-weighted averagevelocities obtained solving the multifluid model. Accordingto Dorao <strong>and</strong> Jakobsen, 134 this discrepancy isan argument for the formulation <strong>of</strong> a mass densitypopulation balance instead <strong>of</strong> a number balance toachieve a consistently integrated population balancewithin the multifluid modeling framework.One can also perform st<strong>and</strong>ard Reynolds decomposition<strong>of</strong> the variables <strong>and</strong> then time average the equation.The typical turbulence modeling <strong>of</strong> the resulting covarianceterms gives rise to diffusive terms in the balanceequation. 121 The physics involved in these terms is notunderstood, however, making this modeling issue achallenging task<strong>and</strong>〈f(x,t)〉〈v r 〉 ) 〈f(x,t)〉 〈v r 〉 f (75)〈f(x,t)〉〈X4 〉 ) 〈f(x,t)〉 〈X4 〉 f (76)〈f(x,t)〉〈v r 〉 ≈ 〈f(x,t)〉 〈v r 〉 + 〈f(x,t)〉′〈v r 〉′ (77)〈f(x,t)〉〈X4 〉 ≈ 〈f(x,t)〉 〈X4 〉 + 〈f(x,t)〉′〈X4 〉′ (78)Provided that sufficient relations for the contact areaare available, the bubble growth term due to interfacialmass transfer can be modeled in accordance with thewell-known two-film theory <strong>and</strong> the ideal gas law. 32 Themodeling <strong>of</strong> the source <strong>and</strong> sink terms due to bubblebreakage <strong>and</strong> coalescence are more difficult from aphysical point <strong>of</strong> view, <strong>and</strong> the existing theory is rathercomplex <strong>and</strong> not easily available. In the followingsubsections, these source terms are thus elucidated infurther detail.Microscopic Source Term Closures. In any twophaseflow field, the initial bubble size is determinedin terms <strong>of</strong> the mechanisms <strong>of</strong> fluid particle generationsuch as formation <strong>of</strong> bubbles at an orifice, perforatedor sintered distributor plates, or other sparger devices.However, in most dispersions, the initial fluid particlesize might be too large or too small to be stable. In thesecases, the fluid particle size is further determined byeither the breakage or coalescence mechanism, respectively.When a fluid particle exceeds a critical value, theparticle interface becomes unstable, <strong>and</strong> breakage islikely to occur. Similarly, when fluid particles aresmaller than a certain critical dimension, coalescenceis likely to occur through a series <strong>of</strong> collision events.It has thus been common to relate particle breakageto a maximum attainable size <strong>of</strong> the particle <strong>and</strong> particlecoalescence to a minimum size. For determining themaximum <strong>and</strong> minimum sizes <strong>of</strong> fluid particles, severalcriteria for breakage <strong>and</strong> coalescence have been reportedin the literature. However, these criteria do not treat adistribution <strong>of</strong> fluid particle sizes, but describe theparticle size limits <strong>of</strong> breakage <strong>and</strong> coalescence.In the following discussion, a brief description <strong>of</strong> thephysical mechanisms that have been considered forformulating closure laws for the more elaborated source


5136 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005<strong>and</strong> sink terms within the population balance equationis provided. Emphasis is placed on a selected number<strong>of</strong> the novel closures suitable for bubble column modeling.A fairly general framework has also been formulatedfor the source terms considering bubble breakage <strong>and</strong>coalescence. 67,121,122,127,169 However, no st<strong>and</strong>ard notationhas yet emerged in the literature, so to avoid unnecessaryconfusion, the following subsections are presentedin accordance with the textbook <strong>of</strong> Ramkrishna. 127Considering the breakage terms, the breakage behaviors<strong>of</strong> different particles are assumed to be independent<strong>of</strong> each other. The average loss rate <strong>of</strong> particles<strong>of</strong> state (x, r) per unit time by breakage yieldsD B (x,r,Y,t) ) b B (x,r,Y,t) f 1 (x,r,t) (79)where b B (x,r,Y,t) has the dimension <strong>of</strong> reciprocal time<strong>and</strong> is <strong>of</strong>ten called the breakage frequency. It representsthe fraction <strong>of</strong> particles <strong>of</strong> state (x, r) breaking per unittime. The breakage processes are <strong>of</strong>ten considered tobe r<strong>and</strong>om in nature, so the modeling work usuallyadopts probabilistic theory, as will be outlined later.The average production rate for particles <strong>of</strong> state(x, r) originating from breakage <strong>of</strong> particles <strong>of</strong> all otherparticle states, considering both internal <strong>and</strong> externalcoordinates is given byB B (x,r,Y,t) ) ∫ Vr∫ Vxν(x′,r′,Y,t) b B (x′,r′,Y,t)P B (x,r|x′,r′,Y,t) f 1 (x′,r′,t) dV′ r dV′ x (80)where ν(x′,r′,Y,t) denotes the average number <strong>of</strong> particlesformed from the breakage <strong>of</strong> a single particle <strong>of</strong>state (x′, r′) in an environment <strong>of</strong> Y at time t. P B -(x,r|x′,r′,Y,t) denotes the probability density function forparticles from the breakage <strong>of</strong> a particle <strong>of</strong> state (x′, r′)in an environment <strong>of</strong> state Y at time t that have state(x, r). In the engineering literature, P B is commonlyreferred to as the daughter size distribution function{i.e., a probability density with units <strong>of</strong> 1/[m 3 (x)] <strong>and</strong>not a probability that is dimensionless} denoting thesize distribution <strong>of</strong> daughter particles produced uponbreakage <strong>of</strong> a parent particle.The integr<strong>and</strong> on the RHS represents the rate <strong>of</strong>formation <strong>of</strong> particles <strong>of</strong> state (x, r) formed by breakage<strong>of</strong> particles <strong>of</strong> state (x′, r′). That is, the number <strong>of</strong>particles <strong>of</strong> state (x′, r′) that breaks per unit time isb B (x′,r′,Y,t) f 1 (x′,r′,t) dV′ r dV′ x . Multiplying the aboveexpression by ν(x′,r′,Y,t) yields the number <strong>of</strong> newparticles resulting from the breakage processes,ν(x′,r′,Y,t) b B (x′,r′,Y,t) f 1 (x′,r′,t) dV′ r dV′ x , <strong>of</strong> which afraction P B (x,r|x′,r′,Y,t) dV′ r dV′ x represents particles <strong>of</strong>state (x, r).Usually, binary breakage is assumed, for whichν(x′,r′,Y,t) ) ν ) 2. However, experimental determination<strong>of</strong> this variable is highly recommended. The functionP B (x,r|x′,r′,Y,t) should also be determined fromexperimental observations or, if future underst<strong>and</strong>ingallows, by detailed modeling <strong>of</strong> the breakage process.P B (x,r|x′,r′,Y,t) is a probability density function <strong>and</strong>should satisfy the normalization condition (in theinternal coordinates)∫ VxP B (x,r|x′,r′,Y,t) dV′ x ) 1 (81)This function also needs to fulfill other obvious conservationlaw properties that constrain the breakageprocess. 127The coalescence functions are considered next. Thepopulation balance source term due to coalescence isusually defined as 127B C (x,r,Y,t) ) ∫ Vr∫ Vx1δ a C (x˜,r˜;x′,r′,Y)f 2 (x˜,r˜;x′,r′,t) ∂(x˜,r˜)∂(x,r)|(x′,r′) )dV′ x dV′ r (82)where δ represents the number <strong>of</strong> times identical pairshave been considered in the interval <strong>of</strong> integration, so1/δ corrects for the redundancy.a C (x˜,r˜;x′,r′,Y,t) denotes the coalescence frequency orthe fraction <strong>of</strong> particle pairs <strong>of</strong> states (x˜, r˜) <strong>and</strong> (x′, r′)that coalesce per unit time. The coalescence frequencyis defined for an ordered pair <strong>of</strong> particles, although froma physical viewpoint, the ordering <strong>of</strong> particle pairsshould not alter the value <strong>of</strong> the frequency. In otherwords, a C (x˜,r˜;x′,r′,Y,t) satisfies a symmetry property:a C (x˜,r˜;x′,r′,Y,t) ) a C (x′,r′;x˜,r˜,Y,t). It is thus essential toconsider only one <strong>of</strong> the above orders for a given pair <strong>of</strong>particles.To proceed, it is necessary to assume that it is possibleto solve for the particle state <strong>of</strong> one <strong>of</strong> the coalescingpair given those <strong>of</strong> the other coalescing particle <strong>and</strong> thenew particle (as the three variables are not independent).Thus, given the state (x, r) <strong>of</strong> the new particle<strong>and</strong> the state (x′, r′) <strong>of</strong> one <strong>of</strong> the two coalescingparticles, the states <strong>of</strong> the other coalescing particle isknown <strong>and</strong> denoted by [x˜(x,r|x′,r′),r˜(x,r|x′,r′)].As in the classical kinetic theory <strong>of</strong> gases, the pairdensity function f 2 is impossible to determine analytically,<strong>and</strong> some closure approximation has to be made.In the population balance derivation, one follows a quitecommon procedure in kinetic theory <strong>and</strong> makes thecoarse approximation f 2 (x˜,r˜;x′,r′,t) ≈ f 1 (x˜,r˜,t) f 1 (x′,r′,t).This assumption implies that there is no statisticalcorrelation between particles <strong>of</strong> states (x′, r′) <strong>and</strong> (x˜, r˜)at any instant t.The sink term due to coalescence is usually definedasD C (x,r,Y,t) ) ∫ Vr∫ Vxa C (x′,r′;x,r,Y)1f 2 (x′,r′;x,r,t) dV′ x dV′ r[ sm (x)] (83)3Although Carrica et al. 13 <strong>and</strong> Hagesaether 146 suggestedthat the most convenient choice <strong>of</strong> internalcoordinate for the bubble column case is the particlemass, m, for pedagogic <strong>and</strong> practical reasons, we usethe bubble diameter, d, as the only internal coordinatein this review.The expressions for the continuous source term functionsthus yieldB B (d,r,Y,t) ) ∫ Vr∫ d∞ν(d′,r′,Y) bB (d′,r′,Y)P B (d,r|d′,r′, Y) f 1 (d′,r′,t) dd′ dV′ rD B (d,r,Y,t) ) b B (d,r,Y) f 1 (d,r,t)


B C (d,r,Y,t) ) 1 ∫ d 12 V r∫ aC 0(d˜,r˜;d′,r′,Y) f 1 (d˜,r˜,t)V r∫ Vrb B (d,r,Y) f 1 (d,r,t) dV r ≈f 1 (d′,r′,t) dd′ dV′ r11V∞ r∫ Vrb B (d,r,Y) dV rV r∫ Vrf 1 (d,r,t) dV rD C (d,r,Y,t) ) f 1 (d,r,t)∫ V′r∫ aC 0(d,r,d′,r′,Y)1f 1 (d′,r′,t) dd′ dV′ rV r∫ Vr∫ Vrf 1 (d,r,t)a C (d,r,d′,r′,Y) f 1 (d′,r′,t) dV′ r dV rwhere d˜ )(d 3 - d′ 3 ) 1/3 . All <strong>of</strong> the source terms have thecommon units 1/[s m 3 ≈ 1 (m)].V r∫ Vrf 1 (d,r,t)The local source term formulation given above is veryseldom used in practical simulations; some kind <strong>of</strong> ∫ Vr∫ Vra C (d,r;d′,r′,Y) dV′ r dV r1averaging is required. In this work, the time-aftervolumeaveraging procedure is employed.dV rV r V r∫ Vrf 1 (d′,r′,t) dV′ r .Integrating the source terms over an averaging wherevolume, V r , yields〈B B (d,Y,t)〉 ) 1 〈ν(d′,Y)〉 )V r∫ 1 VrB B (d,r,Y,t) dV rV r∫ Vrν(d′,r′,Y) dV′ r (84)) ∫ d∞∫ Vrν(d′,r′,Y)〈b B (d′,Y)〉 ) 1 V r∫ Vrb B (d′,r′,Y) dV′ r ( 1 (85)s)∫ VrP B (d,r|d′,r′,Y) dV rb B (d′,r′,Y)fV 1 (d′,r′,t) dV′ r dd′ 〈〈P B (d|d′,Y)〉〉 )r1〈D B (d,Y,t)〉 ) 1 V r∫ VrD B (d,r,Y,t) dV r ) 1 V r∫ Vr∫ VrP B (d,r|d′,r′,Y) dV r dV′ r [(m)] 1 (86)V r∫ Vrb B(d,r,Y) f 1 (d,r,t) dV〈〈a C (d;d′,Y)〉〉 )r1〈B C (d,Y,t)〉 ) 1 V r∫ Vr∫ Vra C (d,r;d′,r′,Y) dV r dV′ r ( m3s) (87)V r∫ VrB C (d,r,Y,t) dV r) 1 ∫ 〈 fd 11 (d′,t)〉 ) 1 1V2 0 V r∫ Vr∫ Vra C (d˜,r˜;d′,r′,Y) f 1 (d˜,r˜,t)r∫ Vrf 1 (d′,r′,t) dV′ r[ m (m)] (88)3f The approximate volume-average source terms can thus1 (d′,r′,t) dV r dV′ r dd′be expressed as〈D C (d,Y,t)〉 ) 1 V r∫ VrD C (d,r,Y,t) dV r〈B B (d,Y,t)〉 )∞∫ 〈ν(d′,Y)〉〈bB∞ 1d(d′,Y)〉〈〈P B (d|d′,Y)〉〉〈f 1 (d′,t)〉 dd′) ∫ 0 V r∫ Vr∫ Vrf 1 (d,r,t) a C (d,r,d′,r′,Y)〈D B (d,Y,t)〉 ) 〈b B (d,Y)〉〈f 1 (d,t)〉f 1 (d′,r′,t) dV′ r dV r dd′To proceed Ramkrishna, 127 (pp 59 <strong>and</strong> 74) introduced 〈B C (d,Y,t)〉 ) 1 ∫ 〈〈aC (d˜;d′,Y)〉〉〈f2 0d1 (d˜,t)〉〈f 1 (d′,t)〉 dd′the usual assumptions in volume averaging, 170 neglectingall correlation terms occurring within the source∞term expressions, meaning that the averages <strong>of</strong> products 〈D C (d,Y,t)〉 ) ∫ 〈f1 0(d,t)〉〈〈a C (d,d′,Y)〉〉〈f 1 (d′,t)〉 dd′are approximated as products <strong>of</strong> averages.The following approximations are introducedPerforming turbulence modeling on these terms is avery difficult task. One usually assumes that thephysical time scales involved are much longer than the∫ VrP B (d,r|d′,r′,Y) dV rtime scales <strong>of</strong> the turbulent fluctuations so that no∫ Vrν(d′,r′,Y) b B (d′,r′,Y)×V correlation functions need to be considered.rThe time-after-volume-averaged source terms aref 1 (d′,r′,t) dV′ r listed below, dropping the averaging symbols for simplicityin the proceeding model derivation∫ Vrν(d′,r′,Y) dV′ r ∫ Vrb B (d′,r′,Y) dV′ r≈∞V rV rB B (d,Y,t) ) ∫ d ν(d′,Y) bB (d′,Y) P B (d|d′,Y) f 1 (d′,t) dd′(89)∫ Vr∫ VrP B (d,r|d′,r′,Y) dV r dV′ r ∫ Vrf 1 (d′,r′,t) dV′ rV r D B (d,Y,t) ) b B (d,Y) f 1 (d,t) (90)V rInd. Eng. Chem. Res., Vol. 44, No. 14, 2005 5137


5138 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005All <strong>of</strong> the source terms still have the common units, 1/[sm 3 (m)].Comparing the source term expressions, eqs 89-92,with eqs 29-32, it is clearly seen that the two formulationsgive rise to identical expressions for the sourceterms only under certain conditions. The macroscopicformulation is explicitly expressed in terms <strong>of</strong> a discretediscretization scheme <strong>and</strong> is very difficult to convert toother schemes.Coalescence Frequency Closures, a C (d; d′, Y).The literature contains several models <strong>and</strong> experimentalanalyses on bubble coalescence. 67,77,130,140,147,171,172Intuitively, bubble coalescence is related to bubblecollisions. The collisions are caused by the existence <strong>of</strong>spatial velocity differences among the particles themselves.However, not all collisions necessarily lead tocoalescence. Thus, the modeling <strong>of</strong> bubble coalescenceon these scales means the modeling <strong>of</strong> bubble collision<strong>and</strong> coalescence probability or efficiency mechanisms.The pioneering work on coalescence <strong>of</strong> particles to formsuccessively larger particles was carried out by Smoluchowski.173,174As for the collision density in the macroscopic modelformulation, the average collision frequency <strong>of</strong> fluidparticles is usually described assuming that the mechanisms<strong>of</strong> collision are analogous to those <strong>of</strong> collisionsbetween molecules as in the kinetic theory <strong>of</strong> gases. 140Furthermore, as pointed out before, the units <strong>of</strong> thevolume-average coalescence efficiency, a C (d;d′,Y), correspondto the units <strong>of</strong> the effective swept volume rate,h C (d;d′,Y), in the kinetic theory <strong>of</strong> gases. 138 Therefore,rather than the actual collision frequency, the volumeswept by a moving bubble <strong>of</strong> diameter d is calculated,from which the number <strong>of</strong> other bubbles with diameterd′ that are hit by this moving bubble <strong>of</strong> diameter d canbe indirectly determined.The volume-average coalescence frequency, a C (d;d′,Y),can thus be defined as the product <strong>of</strong> an effective sweptvolume rate, h C (d;d′,Y), <strong>and</strong> the coalescence probability,p C (d;d′,Y) (Prince <strong>and</strong> Blanch; 77 Kolev, 130 p 169; Coulaloglou<strong>and</strong> Tavlarides; 135 Venneker et al.; 138 Tsouris<strong>and</strong> Tavlarides 169 )which expresses the fact that not all collisions lead tocoalescence. <strong>Modeling</strong> coalescence thus means findingadequate physical expressions for h C (d;d′,Y) <strong>and</strong> p C -(d;d′,Y). Kamp et al., 140 among others, suggested thatmicroscopic closures can be formulated in line with themacroscopic population balance approach; thus, we c<strong>and</strong>efine<strong>and</strong>B C (d,Y,t) ) 1 2 ∫ 0daC (d˜;d′,Y) f 1 (d˜,t) f 1 (d′,t) dd′ (91)D C (d,Y,t) ) ∫ 0∞f1 (d,t) a C (d,d′,Y) f 1 (d′,t) dd′ (92)a C (d;d′,Y) ) h C (d;d′,Y) p C (d;d′,Y)h C (d;d′,Y) ) π 4 (d + d′)2 (vj t,d 2 + vj t,d′ 2 )( m3s) (93)( m3s) (94)p C (d;d′,Y) ) exp( - ∆t coal∆t col)(95)where suitable continuous closures are needed for themodel variables such as the mean turbulent bubbleapproach velocities <strong>and</strong> the time scales. Multifluidmodels can provide improved estimates for the meanturbulent bubble approach velocities as individualvelocity fields for each <strong>of</strong> the bubble phases are calculated.However, as discussed earlier, the existing closuresfor the time scales are very crude <strong>and</strong> need furtherconsideration.The coalescence terms in the average microscopicpopulation balance can then be expressed in terms <strong>of</strong>the local effective swept volume rate <strong>and</strong> the coalescenceprobability variables asB C (d,Y,t) )1∫ dhC (d˜;d′,Y) p2 0C (d˜;d′,Y) f 1 (d˜,t) f 1 (d′,t) dd′ (96)D C (d,Y,t) ) f 1 (d,t)∫ 0∞hC (d;d′,Y) p C (d;d′,Y) f 1 (d˜,t) dd′(97)All <strong>of</strong> the terms in the population balance equation thushave common units, 1/[m 3 s (m)], noting that thecollision densities for the average microscopic <strong>and</strong> themacroscopic frameworks have different units. However,by use <strong>of</strong> a discrete numerical scheme for the solution<strong>of</strong> the average microscopic model, the two formulationsbecome very similar.Models for the Breakage Frequency, b B (d). Formulatingaverage microscopic source term closure lawsfor the breakage processes usually means derivingrelations for the time-after-volume-average breakagefrequency, b B (d); the average number <strong>of</strong> daughterbubbles produced, ν; <strong>and</strong> the average daughter particlesize distribution, P B (d|d′).During the past several decades, a few models havebeen developed for the bubble breakage frequency,b B (d), under turbulent conditions. 137,158 Several differentcategories <strong>of</strong> breakage frequency models are distinguishedin the literature; models based on reactionkinetic concepts, 175 phenomenological models based onthe turbulent nature <strong>of</strong> the system, 169 <strong>and</strong> models basedon purely kinematic ideas. 165,176In an interesting attempt to overcome the limitationsfound in the turbulent breakage models described above,Martínez-Bazán et al. 165 (see also Lasheras et al. 158 )proposed an alternative model in the kinetic theory(microscopic) framework based on purely kinematicideas to avoid the use <strong>of</strong> the incomplete turbulent eddyconcept <strong>and</strong> the macroscopic model formulation.The basic premise <strong>of</strong> the model <strong>of</strong> Martínez-Bazán etal. 165 is that, for a particle to break, its surface has tobe deformed <strong>and</strong>, further, this deformation energy mustbe provided by the turbulent stresses produced by thesurrounding fluid. The minimum energy needed todeform a particle <strong>of</strong> size d is its surface energyE s (d) ) πσ I d 2 (J) (98)The surface restoring pressure is E s per unit volumeσ s (d) ) 6E sπd ) 6σ I3 d(Pa) (99)The size <strong>of</strong> these particles is assumed to be within theinertial subrange <strong>of</strong> turbulence, so the average deformationstress, which results from velocity fluctuations


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5139existing in the liquid between two points separated bya distance d, was estimated aswhere ∆v 2 (d) is the mean value <strong>of</strong> the velocity fluctuationsbetween two points separated by a characteristicdistance d <strong>and</strong> F c is the density <strong>of</strong> the continuous phase.When the turbulent stresses are equal to the confiningstresses, σ t (d) ) σ s (d), a critical diameter, d c , is definedsuch that particles with d < d c are stable <strong>and</strong> will notbreak. A particle with d > d c has a surface energysmaller than the deformation energy, σ s (d) < σ t (d), <strong>and</strong>thus, such a particle deforms <strong>and</strong> might eventuallybreak up in a time t b . Martínez-Bazán et al. 165 appliedKolmogorov’s universal theory valid for homogeneous<strong>and</strong> isotropic turbulent flow to estimate the mean value<strong>of</strong> the velocity fluctuations, ∆v 2 (d) ) β(ɛd) 2/3 .The critical diameter, d c ) (12σ I /βF) 3/5 ɛ -2/5 , is definedby the crossing point <strong>of</strong> the two curves determined byσ s (d) <strong>and</strong> σ t (d). Martínez-Bazán et al. 165 postulated, inaccordance with Newton’s law, that the acceleration <strong>of</strong>the particle interface during deformation is proportionalto the difference between the deformation <strong>and</strong> confinementforces acting on it. In other words, the probability<strong>of</strong> breaking a particle <strong>of</strong> size d in time t b increases asthe difference between the pressure produced by theturbulent fluctuations on the surface <strong>of</strong> the particle,1 /2 F c ∆v 2 (d), <strong>and</strong> the restoring pressure caused by surfacetension, 6σ I /d, increase. On the other h<strong>and</strong>, the breakagefrequency should decrease to a zero limiting value asthis difference <strong>of</strong> pressures vanishes. Thus, the particlebreakage time can be estimated ast b ∝σ t (d) ) 1 2 F c ∆v2 (d) (Pa) (100)dv breakage)where v breakage is the characteristic velocity <strong>of</strong> theparticle breakage process. The breakage frequency b B -(d,ɛ) is given bywhere the values β ) 8.2 <strong>and</strong> K g ) 0.25 were foundexperimentally for bubbly flows.The breakage frequency is zero for particles <strong>of</strong> size de d c , <strong>and</strong> it increases rapidly for particles larger thanthe critical diameter, d > d c . After reaching a maximumat d bB,max ) ( 9 / 4 ) 3/5 d c ≈ 1.63d c , the breakage frequencydecreases monotonically with the particle size. Themaximum breakage frequency, achieved at d bB,max, isgiven by b B,max (ɛ) ) 0.538K g β 1/2 ɛ 3/5 {12[σ I /(F c β)]} -2/5 .Although this approach avoids the eddy concept, it isstill restricted to homogeneous <strong>and</strong> isotropic turbulentflows, it contains several hypotheses that are not yetverified for bubble column flows, <strong>and</strong> it contains a fewadditional adjustable parameters that need to be fittedto many sets <strong>of</strong> experimental data. Furthermore, thisd∆v 2 (d) - 12 σ IF c db B (d,ɛ) ) 1 ∆v 2 (d) - 12 σ IF c d) Kt g )b dK g β(ɛd)2/3 - 12 σ IF c dd(101)(102)model concept has only been applied to systems (i.e.,turbulent water jets 165,176-178 ) where the turbulentdissipation rates are 2-3 orders <strong>of</strong> magnitude largerthan what is observed in bubble columns. The bubblesizes considered were also very small, up to about 1 mmonly. On the other h<strong>and</strong>, after careful investigations <strong>of</strong>the turbulent breakage processes in bubble columns, itmight be possible to extend the application <strong>of</strong> thisapproach to bubble column systems. The experimentaltechniques developed for investigating the breakage <strong>of</strong>bubbles on the centerline within the fully developedregion <strong>of</strong> turbulent water jets might initiate ideas onhow to study these phenomena in bubble columns. 177,178The digital image-processing technique 177,178 used inthese studies gives satisfactory results in very dilutetwo-phase flows, but the method is still questionablewhen the bubble concentration increases considerablyas for bubble columns operated in the heterogeneousflow regime.A severe drawback for integrated multifluid/populationbalance models is that all <strong>of</strong> the kernels suggestedin the literature are very sensitive to the turbulentenergy dissipation rate (ɛ). As mentioned earlier, thelocal ɛ variable is difficult to calculate from the k-ɛturbulence model because the equation for the dissipationrate merely represents a fit <strong>of</strong> a turbulent lengthscale to single-phase pipe flow data. Therefore, furtherwork is highly needed to elucidate the mechanisms <strong>of</strong>bubble breakage in turbulent flows. However, an alternativeto the eddy concept has been reported that mightmake the work on model validation easier. Perhaps thegreatest advantages lies in the fact that this closure isformulated within the average microscopic modelingframework, thereby avoiding the limitations <strong>of</strong> themacroscale formulation.At this point, care should be taken as the discretemacroscopic breakage rate models [Ω B (d i )] are <strong>of</strong>tenerroneously assumed to be equal to the average microscopicbreakage frequency (b B (b)).Number <strong>of</strong> Daughter <strong>Bubble</strong>s Produced, ν. In theaverage microscopic formulations, this parameter determinesthe average number <strong>of</strong> daughter particlesproduced by breakage <strong>of</strong> a parent particle <strong>of</strong> size d. Thebreakage <strong>of</strong> parent bubbles into two daughter bubblesis assumed in most investigations reported (ν ) 2).In a recent paper, Risso <strong>and</strong> Fabre 145 observed thatthe number <strong>of</strong> fragments depended on the specific shape<strong>of</strong> the parent bubble during the deformation process <strong>and</strong>varied in a wide range. In their experiments, a videoprocessingtechnique was used for studies <strong>of</strong> two groups<strong>of</strong> bubbles, one group being <strong>of</strong> sizes in the range between2 <strong>and</strong> 6 mm <strong>and</strong> the other <strong>of</strong> sizes in the range 12.4-21.4 mm. Two fragments were obtained in 48% <strong>of</strong> thecases, between 3 <strong>and</strong> 10 in 37% <strong>of</strong> the cases, <strong>and</strong> morethan 10 fragments in 15% <strong>of</strong> the cases. The breakageprocess is certainly not purely binary. Implementingthis effect instead <strong>of</strong> binary breakage is expected tosignificantly alter the number <strong>of</strong> smaller bubbles <strong>and</strong>the interfacial area concentration predicted by thepopulation balance equation.The experimental data <strong>of</strong> Risso <strong>and</strong> Fabre 145 alsoindicate that an equal-size daughter distribution is morecommon for bubble breakage than an unequal one. Incontrast, Hesketh et al. 179,180 investigated bubble breakagein turbulent flows in horizontal pipes <strong>and</strong> concludedthat an unequal-size daughter distribution is moreprobable than an equal-size one. The daughter particle


5140 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005size distribution model <strong>of</strong> Luo <strong>and</strong> Svendsen 71 relies onthe assumption that unequal-size breakage is moreprobable than equal-size breakage in accordancewith the observations <strong>of</strong> Hesketh et al., as will bediscussed in the next subsection. The deviating experimentalobservations indicate that further experimentalinvestigations are needed to determine reliable correlationsfor the physical systems <strong>and</strong> flow conditionsfor which the equal <strong>and</strong> unequal breakage outcomesoccur.Daughter Particle Distribution, P B (d|d′). Amongthe papers adopting the average microscopic formulation,two <strong>of</strong> the earliest models for the daughter particlesize distribution were proposed by Valentas et al. 181,182The daughter particle probability distribution functionssuggested by Valentas et al. were purely statisticalrelations. The first was a discrete model, in which aparent particle <strong>of</strong> diameter d was assumed to split intoequally sized daughter particles <strong>of</strong> diameter d/ν, whereν is the number <strong>of</strong> daughter particles formed. Thesecond model proposed by Valentas et al. was a logical,continuous analogue <strong>of</strong> the discrete daughter particleparticle distribution function (PDF). In this case, it wasassumed that the daughter particle sizes were normallydistributed about a mean value.Ross <strong>and</strong> Curl, 175 Coulaloglou <strong>and</strong> Travlarides, 135 <strong>and</strong>Prince <strong>and</strong> Blanch 77 argued that the particle breakagefrequency should be a function <strong>of</strong> the difference insurface energy between a parent particle <strong>and</strong> daughterparticles produced <strong>and</strong> the kinetic energy <strong>of</strong> a collidingeddy. Although Ross <strong>and</strong> Curl, 175 Coulaloglou <strong>and</strong>Travlarides, 135 <strong>and</strong> Prince <strong>and</strong> Blanch 77 developed morephysical models for the particle breakage frequency, thedaughter particle size distributions still relied uponstatistical relations.Among the most widely used phenomenological modelsbased on surface energy considerations is the oneproposed by Tsouris <strong>and</strong> Tavlarides. 169 Their expressionfor the daughter particle PDF yieldsP B (d|d′) )∫ dmine min + [e max - e(d′)]demin + [e max - e(d′)] dd′( 1 m) (103)Considering binary breakage only, Tsouris <strong>and</strong> Tavlarides169 postulated that the probability <strong>of</strong> formation<strong>of</strong> a daughter particle <strong>of</strong> size d′ is inversely proportionalto the energy required to split a parent particle <strong>of</strong> sized into a particle <strong>of</strong> size d′ <strong>and</strong> its complementary particle<strong>of</strong> size d˜ ) (d 3 - d′ 3 ) 1/3 . This energy requirement isproportional to the excess surface area generated bysplitting the parent particlee(d′) ) πσ I d′ 2 + πσ I d˜ 2 - πσ I d 2 ) πσ I d 2 [( d′d ) 2 +[ 1 - ( d′d ) 3 ] 2/3 - 1 ] (104)To avoid the singularity present in their daughter sizedistribution model for d′ ) 0 (i.e., the parent particledoes not break), a minimum particle size was defined,d min . The excess surface area relation reaches a minimumvalue when a particle <strong>of</strong> minimum diameter, d min ,<strong>and</strong> a complementary one <strong>of</strong> maximum size, d max ) (d 3- d min3 ) 1/3 , are formed. The minimum energy, e min ,isgiven bye(d min ) ) πσ I d 2 [( d mind) 2 + [ 1 - ( d mind) 3 ] 2/3 - 1] (105)This expression gives a minimum probability for theformation <strong>of</strong> two particles <strong>of</strong> the same size <strong>and</strong> amaximum probability for the formation <strong>of</strong> a pair madeup <strong>of</strong> a very large particle <strong>and</strong> a complementary verysmall one. The maximum energy corresponding to theformation <strong>of</strong> two particles <strong>of</strong> equal volumes or <strong>of</strong>diameters d′ ) (d 3 - d′ 3 ) 1/3 ) d /2 1/3 is e max ) πσ I d 2 [2 1/3- 1].Luo <strong>and</strong> Svendsen 71 derived a discrete expression forthe breakage density <strong>of</strong> a particle <strong>of</strong> diameter d i intotwo daughter particles <strong>of</strong> size d j <strong>and</strong> (d i3 - d j 3 ) 1/3 usingenergy arguments similar to those employed by Tsouris<strong>and</strong> Tavlarides. 169The significant difference between the Luo <strong>and</strong>Svendsen model <strong>and</strong> its predecessors is that the formergives both a partial breakage rate, that is, the breakagerate for a particle <strong>of</strong> size d i splitting into a particle <strong>of</strong>size d j <strong>and</strong> its complementary bubble, <strong>and</strong> an overallbreakage rate. The previous surface energy modelsprovided only an overall breakage rate. An expressionfor the daughter particle size distribution function canthus be calculated by normalizing the partial breakagerate by the overall breakage rate. As mentioned earlier,this variable was not required for the population balanceclosure but is rather a spin-<strong>of</strong>f from the primaryclosures.The Luo <strong>and</strong> Svendsen daughter particle size distributionis thus determined from the expressionP B (d i |d j ) ) Ω B (d i :d j )Ω B (d i )[-] (106)Note that the units <strong>of</strong> the discrete daughter bubble sizedistribution variable are different from the units obtainedby deriving the continuous daughter size distributionfunction from the microscopic formulations(1/m). It is thus not trivial how the model <strong>of</strong> Luo 70 shouldbe adopted within a more fundamental modeling approach.However, similarly to the model <strong>of</strong> Tsouris <strong>and</strong>Tavlarides, 169 the model <strong>of</strong> Luo <strong>and</strong> Svendsen predictedthat the probability <strong>of</strong> breaking a parent particle into avery small particle <strong>and</strong> a complementary large particleis larger than the probability <strong>of</strong> equal-size breakage(Figure 15). The distribution has a U-shape, with aminimum probability for the formation <strong>of</strong> two equallysized daughter particles <strong>and</strong> a maximum probability forthe formation <strong>of</strong> a very large daughter particle <strong>and</strong> itscomplement d min .Although Hinze 159 <strong>and</strong> Risso <strong>and</strong> Fabre, 145 amongothers, showed <strong>and</strong> discussed the diversity <strong>of</strong> shapes <strong>of</strong>the bubbles that can be found in turbulent flows, atpresent, sufficient experimental data do not exist in theliterature to assist in developing adequate daughterbubble probability density functions valid for bubblecolumns.In an attempt to overcome the limitations found inthe models for the daughter particle size distributionfunction described above based on the eddy collisionconcept, Martínez-Bazán et al. 176 proposed an alternativestatistical model based on energy balance principles.The model was originally intended for theprediction <strong>of</strong> air bubble breakage at the centerline <strong>of</strong> ahigh-Reynolds-number, turbulent water jet. It was


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5141P B (d/d′) ∝ [12 F c β(ɛd′)2/3 - 6σ I /d ]{12 F c β[ɛ(d3 -d′ 3 ) 1/3 ] 2/3 - 6σ I /d } (107)Relating d′ <strong>and</strong> (d 3 - d′ 3 ) 1/3 through the mass balancegiven above yieldsP B (d/d′) ∝ [12 F c β(ɛd)2/3 ] 2 [(d′/d) 2/3 - (d c /d) 5/3 ]{[1 -(d′/d ) 3 ] 2/9 - (d c /d) 5/3 } (108)Figure 15. Sketch <strong>of</strong> the breakage kernel <strong>of</strong> Luo <strong>and</strong> Svendsen. 71An unequal daughter size distribution is predicted by this model.In this case, the parent diameter size is, d i ) 0.006 m, <strong>and</strong> theturbulent energy dissipation rate is set at ɛ ) 1m 2 /s 3 .assumed that, when an air bubble is injected into theturbulent water jet, the velocity fluctuations <strong>of</strong> theunderlying turbulence result in pressure deformationforces acting on the bubble’s surface that, when greaterthan the confinement forces due to surface tension, willcause breakage.The model assumes that, when a parent particlebreaks, two daughter particles are formed (ν ) 2) withdiameters d′ <strong>and</strong> (d 3 - d′ 3 ) 1/3 . The validity <strong>of</strong> thisassumption is supported by the high-speed video images,presented by Martínez-Bazán et al. 176 The diametersare related through the conservation <strong>of</strong> mass, (d 3- d′ 3 ) 1/3 ) d[1 - (d′/d) 3 ] 1/3 . The particle-splitting processwas not considered purely r<strong>and</strong>om, as the pressurefluctuations <strong>and</strong> thus the deformation stress (σ t ) inhomogeneous <strong>and</strong> isotropic turbulence is not uniformlydistributed over all scales. It was further assumed thatthere is a minimum distance, d min , over which theturbulent stresses acting between two points separatedby this distance, 1 / 2 F c β(ɛd min ) 2/3 , are just equal to theconfinement pressure due to surface tension: σ t (d min )) σ s (d). In other words, at this distance, the turbulentpressure fluctuations are exactly equal to the confinementforces for a parent particle <strong>of</strong> size d. The probability<strong>of</strong> breaking <strong>of</strong>f a daughter particle with size d′< d′ min ) (12σ I /βF c d) 3/2 ɛ -1 should therefore be zero.The fundamental postulate <strong>of</strong> the Martínez-Bazán etal. 176 model is that the probability <strong>of</strong> splitting <strong>of</strong>f adaughter particle <strong>of</strong> any size such that d′ min < d′ < d isproportional to the difference between the turbulentstresses over a length d′ <strong>and</strong> the confinement forcesholding the parent particle <strong>of</strong> size d together. For theformation <strong>of</strong> a daughter particle <strong>of</strong> size d′, the differencein stresses is given by ∆σ t,d′ ) 1 / 2 F c β(ɛd′) 2/3 - 6σ I /d. Foreach daughter particle <strong>of</strong> size d′, a complementarydaughter particle <strong>of</strong> size (d 3 - d′ 3 ) 1/3 is formed with adifference <strong>of</strong> stresses given by ∆σ t,(d 3 -d′ 3 ) 1/3 ) 1 / 2 F c β[ɛ(d 3- d′ 3 ) 1/3 ] 2/3 - 6σ I /d.The model states that the probability <strong>of</strong> forming a pair<strong>of</strong> complementary daughter particles <strong>of</strong> sizes d′ <strong>and</strong>(d 3 - d′ 3 ) 1/3 from the splitting <strong>of</strong> a parent particle <strong>of</strong> sized is related to the product <strong>of</strong> the excess stressesassociated with the length scales corresponding to eachdaughter particle size. That isorP B (d*) ∝ [12 F c β(ɛd)2/3 ] 2 (d 2/3 - Λ 5/3 )[(1 - d 3 ) 2/9 - Λ 5/3 ](109)where Λ ) d c /d <strong>and</strong> d c is the critical diameter definedas d c ) (12σ I /βF c ) 3/5 ɛ -2/5 .The critical parent particle diameter defines theminimum particle size for a given dissipation rate <strong>of</strong>turbulent kinetic energy for which breakage can occur.The minimum daughter diameter defines the distanceover which the turbulent normal stresses just balancethe confinement forces <strong>of</strong> a parent particle <strong>of</strong> size d. Theminimum diameter, therefore, gives the minimumlength over which the underlying turbulence can pinch<strong>of</strong>f a piece <strong>of</strong> the parent particle. This length is notarbitrarily selected; rather, its determination is basedon kinematics.This model further assumes that the size <strong>of</strong> the parentparticle is in the inertial subrange <strong>of</strong> turbulence.Therefore, it implies that d min e d′ e d max provided thatd min > λ d , where λ d is the Kolmogorov length scale <strong>of</strong>the underlying turbulence. Otherwise, d min is taken tobe equal to λ d . However, no assumption needs to bemade about the minimum <strong>and</strong> maximum eddy size thatcan cause particle breakage. All eddies with sizesbetween the Kolmogorov scale <strong>and</strong> the integral scale aretaken into account.The daughter particle probability density function canbe obtained from the probability expression given aboved*∫ maxdmin *by utilizing the normalization condition: P(d*)d(d*) ) 1. The PDF <strong>of</strong> the ratio <strong>of</strong> diameters d* ) d′/d,P / B (d*), can then be written asP B / (d*) )d/∫maxdmin/(d* 2/3 - Λ 5/3 )[(1 - d* 3 ) 2/9 - Λ* 5/3 ](d* 2/3 - Λ 5/3 )[(1 - d* 3 ) 2/9 - Λ* 5/3 ]d(d*)(110)Note that P B (d′|d) ) P B / (d*)/d, <strong>and</strong> Λ ) d c /d ) (d min /d) 2/5 .In contrast to the collision-based phenomenologicalmodels for the daughter particle size distribution, thismodel predicts a symmetric distribution with the highestprobability for equal-size particle breakage in accordancewith the pure statistical model <strong>of</strong> Konno etal. 183 (Figure 16).To summarize, purely statistical models for thedaughter particle size distribution lack physical support.As for the breakage frequency models, the existingdaughter particle size models based on eddy collisionarguments rely on the assumption that turbulenceconsists <strong>of</strong> a collection <strong>of</strong> eddies that can be treatedusing relationships from the kinetic theory <strong>of</strong> gases.Martínez-Bazán et al., 176 on the other h<strong>and</strong>, made a first


5142 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005Figure 16. Sketch <strong>of</strong> the breakage kernel <strong>of</strong> Konno et al. 183 Anequal daughter size distribution is predicted by this model. Thekernel by Martínez-Bazán 176 is quite similar to this one, but theshape <strong>of</strong> the pr<strong>of</strong>ile is slightly different close to the minimum <strong>and</strong>maximum values <strong>of</strong> d′ 3 /d 3 .approach toward developing an average microscopickinematic/statistical model, where the physics wereincluded through a daughter size distribution functionexpressed on the basis <strong>of</strong> kinematic arguments. Alternatively,one can develop kinematic/statistical modelsin which an effective daughter size distribution functionrepresents an average <strong>of</strong> a large number <strong>of</strong> simulations<strong>of</strong> the physical breakage phenomena for a series <strong>of</strong>operating conditions.The deviating experimental observations reported inthe literature indicate that further experimental investigationsare needed to determine reliable correlationsfor which physical systems <strong>and</strong> under which flowconditions the equal <strong>and</strong> unequal breakage outcomesoccur.6. Numerical Discretization <strong>of</strong> the PopulationBalance EquationFor the average microscopic formulations, the closures<strong>and</strong> the numerical discretizations are split, so that anoptimal numerical solution method has to be found afterthe closure laws are derived. The numerical solutionmethods are, in general, problem-dependent <strong>and</strong> optimizedfor particular applications.In mathematical terms, the population balance equation(PBE) is classified as a nonlinear partial integrodifferentialequation (PIDE). Because analytical solutions<strong>of</strong> this equation are not available for most cases<strong>of</strong> practical interest, several numerical solution methodshave been proposed during the past two decades, asdiscussed by Williams <strong>and</strong> Loyalka 128 <strong>and</strong> Ramkrishna.127In general, the numerical solution <strong>of</strong> PIDEs consists<strong>of</strong> a three-step procedure. First, the basic set <strong>of</strong> PIDEsis discretized in the internal coordinates <strong>and</strong> expressedas a set <strong>of</strong> partial differential equations (PDEs). ThePDEs are then discretized in time <strong>and</strong> in the physicalspace coordinates using st<strong>and</strong>ard PDE discretizationtechniques in a second step. Finally, the resulting set<strong>of</strong> algebraic equations is solved by use <strong>of</strong> a suitablesolver.For most engineering calculations in the past, fluxes<strong>and</strong> integral results were considered sufficient interpretinggrowth, agglomeration, nucleation, <strong>and</strong> breakagedata <strong>and</strong> for fitting the mechanistic kernels for solidparticle suspensions. Therefore, only the first moments<strong>of</strong> the size density distribution function <strong>of</strong> the population(i.e., the mean, the st<strong>and</strong>ard deviation, etc.) are required.For the more complex fluid particle applicationsinvestigated in recent studies, local or semilocal sizedistributions are required, enabling better underst<strong>and</strong>ing<strong>of</strong> the physical phenomena involved <strong>and</strong> propermodel validation. However, many <strong>of</strong> the present CFDcodes still resort to integral formulations to minimizethe computational time <strong>and</strong> memory requirements. Itthus seems convenient to divide the numerical solutionmethods into two groups; methods solving the modelsfor the moments <strong>of</strong> the size distribution <strong>and</strong> thosesolving the models for the size distribution itself.Methods for the Moments <strong>of</strong> the Distribution.Basically, the method <strong>of</strong> moments converts the set <strong>of</strong>PIDEs into a set <strong>of</strong> PDEs in which each PDE representsa given moment <strong>of</strong> the population size density function.Defining the jth moment <strong>of</strong> the population densityfunction asthe PBE can be reduced to a set <strong>of</strong> PDEs by integratingthe basic transport equation over the whole particle sizedistributionor in terms <strong>of</strong> the momentsµ j ) ∫ 0∞d j f(r,d,t) dd (111)∫ 0∞[ ∂f∂t + ∇ r ‚(Ṙf ) + ∇ x ‚(Ẋf ) - D C + B C - D B +B B] dj dd for j ) 0, 1, ... (112)∂µ j∂t + ∫ 0∞[∇r ‚(Ṙf)]d j dd + ∫ 0∞[∇x ‚(Ẋf )]d j dd )∫ 0∞(DC + B C - D B + B B )d j ddfor j ) 0, 1, ... (113)Assuming that the convective velocities are independent<strong>of</strong> the property coordinate (i.e., the particle diameter),one can integrate by parts to obtain∂µ j∂t + dj fṘ| 0 ∞ - ∇ r ‚Ṙ∫ 0∞jd j-1 f dd + d j Ẋf| 0 ∞ -∇ x ‚Ẋ∫ 0∞jd j-1 f dd (114)) ∫ 0∞(DC + B C - D B + B B )d j dd for j ) 0, 1, ...Because the f(r,d,t) function goes to zero at the maximum<strong>and</strong> minimum boundaries, the equation can besimplified <strong>and</strong> rewritten as∂µ j∂t - j∇ r ‚(Ṙµ j-1 ) - j∇ x ‚(Ẋµ j-1 ) ) ∞∫ (DC 0+ B C - D B +B B )d j dd for j ) 0, 1, ... (115)with initial conditions µ j (r,d,t ) 0) ) µˆ j(r,d) for j ) 0, 1,...


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5143The mathematical model characteristics <strong>of</strong> the sourceterms, i.e., D C , B C , D B , <strong>and</strong> B B , determine whether thesystem <strong>of</strong> equations is closed or not.To illustrate this problem, a pure breakage processis considered, i.e., D C ) B C ) 0, but D B * 0 <strong>and</strong> B B * 0.In addition, as an example only, a generalized formulation<strong>of</strong> the breakage death term (D B ) is analyzed in thefollowing discussion. Assuming that the breakage frequency,b B (d), can be represented by a polynomialrelation such as b B (d) ) ∑ N i)0 b Bi P Bi (d), where P Bi (d) )d j , integration <strong>of</strong> the source term givesN∞∫ bBi 0f(r,d,t)d j ∞dd ) ∑b Bi ∫ 0 d j f(r,d,t)d j dd )i)0N∑b Biµ i+j (116)i)0This implies that higher-order moments are introduced,so the system <strong>of</strong> PDEs cannot be closed analytically. Itis possible to show that similar effects will occur for theother source terms as well. This problem limits theapplication <strong>of</strong> the exact method <strong>of</strong> moments to theparticular case where one has constant kernels only. Inother cases, one has to introduce approximate closuresin order to eliminate the higher-order moments, therebyensuring that the transport equations for the moments<strong>of</strong> the PSD can be expressed in terms <strong>of</strong> the lower-ordermoments only (i.e., a modeling process very similar toturbulence modeling).For bubbly flows, most <strong>of</strong> the early papers adoptedeither a macroscopic population balance approach withan inherent discrete discretization scheme as describedearlier or rather semiempirical transport equations forthe contact area <strong>and</strong>/or the particle diameter. Actually,very few consistent source term closures exist for themicroscopic population balance formulation. The existingmodels are usually solved using discrete semiintegraltechniques, as will be outlined in the nextsubsection.Therefore, the approximate integral method is notwidely used to solve the population balance model forbubbly flows, as the kernels for these systems are rathercomplex, so it is very difficult to eliminate the higherordermoments developing closures with sufficient accuracy.Hounslow et al. 184 suggested that the difficulty incurredby the inclusion <strong>of</strong> complex closures for thesource terms investigating nucleation, growth, <strong>and</strong>aggregation <strong>of</strong> particulate suspensions can be relaxedby discretizing the population balance equation. Themoments <strong>of</strong> the size distribution were calculated fromthe discrete data resulting from the simulations. However,the direct solution <strong>of</strong> the equations containingphysical source term parametrizations by use <strong>of</strong> conventionalfinite-difference techniques was complicatedby the presence <strong>of</strong> the integral terms <strong>and</strong> resulted incomputational loads far greater than desirable. Hounslowet al. thus placed substantial emphasis on correctlydetermining both the moments <strong>and</strong> the particle sizedistribution. The novel numerical technique developed,later referred to as the class method, reinforces theconservative properties <strong>of</strong> a few lower-order momentsby transforming the equation through the introduction<strong>of</strong> a constant-volume correction factor that was foundvalid for several cases. Further details will be given laterin discussing discrete semi-integral <strong>and</strong> differentialmethods in general. The class method containing volumecorrection factors has not been widely used forbubbly flows mainly because <strong>of</strong> the complexity <strong>of</strong> thebubble breakage closures.When the population balance is written in terms <strong>of</strong>one internal coordinate (e.g., particle diameter or particlevolume), the closure problem mentioned above forthe moment equation can be successfully relaxed forsolid particle systems by the use <strong>of</strong> a quadratureapproximation.In the quadrature method <strong>of</strong> moments (QMOM)developed by McGraw 185 for the description <strong>of</strong> sulfuricacid-water aerosol dynamics (growth), a certain type<strong>of</strong> quadrature function approximation is introduced toapproximate the evolution <strong>of</strong> the integrals determiningthe moments. Marchisio et al. 186,187 extended the QMOMfor the application to aggregation-breakage processes.For the solution <strong>of</strong> crystallization <strong>and</strong> precipitationkernels, the size distribution function is expressed usingan expansion in delta functions 186,187Pf(d,t) ≈ ∑ω R δ(d - d R ) (117)R)0Defining the ith moment <strong>of</strong> the population densityfunction asµ i ) ∫ 0∞d i f(d,t) dd (118)<strong>and</strong> using a quadrature rule to approximate the integralyieldsP∞µ i ) ∫ 0 d i if(d,t) dd ≈ ∑ω R (t)d RR)1(119)Inserting the same quadrature approximation into thesource term integrals provides an approximate numericaltype <strong>of</strong> closure avoiding the closure problem <strong>of</strong>higher-order moments at the cost <strong>of</strong> model accuracy. 185This numerical approximation actually neglects thephysical effects <strong>of</strong> the higher-order moments. No reportsapplying this procedure to bubbly flows have been foundso far.Methods for the Distribution. This group <strong>of</strong> methodscontains a large variety <strong>of</strong> solution strategies, butonly a few popular techniques such as the finite-volume(FVM) methods <strong>and</strong> the direct quadrature method <strong>of</strong>moments (DQMOM) are discussed here.The main premise <strong>of</strong> these methods is that, inpractice, one is not necessarily interested in the numberdensity probability f 1 , but rather in the number densityN i (i.e., the number <strong>of</strong> bubbles <strong>of</strong> a particular size orsize interval per unit volume). The methods within thiscategory either divide the size domain into finite regionson which the number density is integrated to providebalances between the number <strong>of</strong> particles in each “bin”or solve the balance equations accurately for a limitednumber <strong>of</strong> discretization points on the size domain.A group <strong>of</strong> discrete techniques is based on theconventional finite-volume <strong>and</strong> finite-difference concepts<strong>and</strong> can be classified as finite-volume methods (FVMs).The wide use <strong>of</strong> the FVMs to solve PBEs is due to theirsimple construction <strong>and</strong> conservative characteristics, ina similar manner as the FVM is <strong>of</strong>ten preferred forformulating the CFD discretization algorithms. That is,the flux <strong>of</strong> a given property leaving through one face <strong>of</strong>


5144 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005the grid volume must equal the flux entering theneighboring ones. Adopting this method thus enforcesthe conservation <strong>of</strong> the conserved quantities. The growthterms are simply treated as flux processes, whereas thecoalescence <strong>and</strong> breakage processes are more difficultto h<strong>and</strong>le because the particles can disappear <strong>and</strong>appear in different regions <strong>of</strong> the domain. To fulfill theconservation properties considering these processes, itis necessary to employ some kind <strong>of</strong> ad hoc numerical“tricks” analogous to the linear source term manipulationsin the CFD algorithms. The different tricksemployed in the literature on PBEs have been characterizedas the so-called discrete, class, <strong>and</strong> multigroupmethods.In principle, the discrete <strong>and</strong> multigroup methods arevery similar. A slight difference is observed in the waythe number density is approximated. The discretemethod approximates the functional values at discretepivotal points within the size interval using deltafunctions, whereas the multigroup method divides thesize spectrum into a number <strong>of</strong> continuous subintegralsor groups. In addition, the mean value theorem issometimes applied in different ways. The multigroupmethod substitutes the mean value <strong>of</strong> the populationdensity in each interval in each integr<strong>and</strong> by f i ) N i /(d i+1 - d i ) <strong>and</strong> withdraws this factor from the integral,whereas the discrete method can be based on themultigroup approach or alternatively derived by estimatingthe kernels at the pivots <strong>and</strong> assigning the sizeinterval definition to the remaining factors <strong>of</strong> theintegrals. The two names used to refer to these methodsoriginate from the parallel historical developmentswithin two different fields <strong>of</strong> science. The discretemethod was developed in chemical engineering applicationsinvestigating biopopulations, whereas the multigroupapproach was developed in neutron <strong>and</strong> nuclearreactor physics. A well-established approach in nuclearengineering to solving the neutron density transportequations resorts to the multiple energy group equationsusing transfer cross sections. Similar ideas were laterused to determine a detailed description <strong>of</strong> the dynamicinteractions among groups <strong>of</strong> particles.Because some versions <strong>of</strong> the discrete or multigroupFVMs are adopted in nearly all multifluid CFD simulations,the main steps in the PDE discrete discretizationprocedures are outlined here. 127First, the dispersed phase is divided into size intervalsaccording to some criterion; in this case, we have chosenthe bubble diameter d i . The interval [d i , d i+1 ] is denotedI i , <strong>and</strong> the total number <strong>of</strong> particles in interval I i is givenbyN i (t) ≡ ∫ did i+1fi(d,t) dd (120)The population balance is integrated in particle size overthe subintervald∫i+1 ∂f(d,t)ddidd + ∫ i+1∇‚(f(d,t)vp∂tdi)ddd) ∫i+1(BBdi(d,t) - D B (d,t) + B C (d,t) - D C (d,t)) dd(121)The class boundaries are independent <strong>of</strong> time <strong>and</strong>spatial position, so ∂/∂t <strong>and</strong> ∇ can be moved outside theintegral∂∂t ∫ d id i+1f(d,t) dd + ∇‚ ∫ did i+1(f(d,t)v) dd(122)) ∫ did i+1[BB(d,t) - D B (d,t) + B C (d,t) - D C (d,t)] ddThe result <strong>of</strong> the integration is a balance equation forthe number densities, N i , in terms <strong>of</strong> the unknownnumber density probability, f 1 (d,t), <strong>and</strong> is, hence, stillunresolvable.A number-average particle velocity can be introducedto simplify the expression for the convective terms∫ did i+1[f(d,t)v] dd ) ∫ did i+1f(d,t) dd 〈v〉Ni ) N i 〈v〉 Ni(123)Inserting eqs 120 <strong>and</strong> 123 into eq 122 yields∂N i∂t + ∇‚(N i 〈v〉 N i)d) ∫i+1[BBdi(d,t) - D B (d,t) + B C (d,t) - D C (d,t)] dd(124)The integrals (with respect to d′) in the source terms(see eqs 89-92) are expressed as the sums <strong>of</strong> integralsover subintervals. Replacing the integral in each sourceterms with a sum yieldsMdB B (d,t) ) ∑ j+1v(d′)∫dj bB (d′)P B (d|d′) f 1 (d′,t) dd′ (125)j)iD B (d,t) ) b B (d) f 1 (d,t) (126)B C (d,t) ) 1 2 ∑ j)0i-1∫djd j+1aC[(d 3 - d′ 3 ) 1/3 ,d′] f 1 [(d 3 -d′ 3 ) 1/3 ,t] f 1 (d′,t) dd′ (127)MdD C (d,t) ) f 1 (d,t) ∑j+1aC∫dj (d,d′) f 1 (d′,t) dd′ (128)j)0The above expressions contain the continuous bubblenumber probability density, f(d,t). The source termsmust be expressed entirely in terms <strong>of</strong> the dependentvariable N i . This can be achieved by using the meanvalue theorem. 127 Note, as mentioned before, at thispoint, the discrete method <strong>of</strong> Ramkrishna 127 mightdeviate slightly from the multigroup method.The mean value theorem is used to cast the eqs 125-129 entirely in terms <strong>of</strong> N i <strong>and</strong> N j . In each interval, thevariables are replaced by a mean value.For example, using the discrete method, the meanvalue theorem can be used to expressMd∫i+1DC ddi(d,t) dd ) ∫i+1f1 ddi(d,t)∑j+1aC∫dj (d,d′)j)0f 1 (d′,t) dd′ ddMd≈ a C (x i ,x j )∫i+1f1 ddi(d,t) dd∑j+1f1∫dj (d′,t) dd′ )j)0MN i∑a C (x i ,x j )N j (129)j)0where x i <strong>and</strong> x j are pivot points in I i <strong>and</strong> I j , respectively.


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5145The pivot concentrates the particles in the interval ata single representative point. Thus, one can write thenumber probability density f 1 (d,t) as being given byM1f 1 (d,t) ) ∑N i δ(d - x i ) (130)i)0[ m (m)] 3The RHS <strong>of</strong> eq 131 is zero for d * x i <strong>and</strong> equals N i whend ) x i .Applying the mean value theorem to the source termsyieldsMd∫i+1BB ddi(d,t) dd ) ∑N j v(x j ) b B (x j )∫i+1PBdi(d|x j )ddj)i(131)d∫i+1DBdi(d,t) dd ) b B (x i )N i (132)i-1N j∫ did i+1BC(d,t) dd ) 1 2 ∑ j)0<strong>and</strong> inserting into eq 124 gives+ 1 2 ∑ j)0∑(x j +x k )∈I iN k a C (x k ,x j ) (133)Md∫i+1DCdi(d,t) dd ) N i∑N j a C (x i ,x j ) (134)j)0dN idt + ∇‚(N i 〈v〉 N i)Md) ∑N j v(x j ) b B (x j )∫i+1PBdi(d|x j )dd - b B (x i )N ij)ii-1N j∑ N k a C (x k ,x j ) - N i∑(x j +x k )∈I i j)0Equation 135 is not necessarily conservative because<strong>of</strong> the finite (i.e., in practice, rather coarse) size gridresolution, <strong>and</strong> some sort <strong>of</strong> numerical trick must beused to enforce the conservative properties. It is mainlyat this point in the formulation <strong>of</strong> the numericalalgorithm that the class method <strong>of</strong> Hounslow et al., 184the discrete method <strong>of</strong> Ramkrishna, 127 <strong>and</strong> the multigroupapproach used by Carrica et al., 13 among others,differ to some extent, as discussed earlier.The problem in question is related to the birth termsonly. The formation <strong>of</strong> a bubble <strong>of</strong> size d′ in size range(x i , x i+1 ) due to breakage or coalescence is representedby assigning fractions γ 1 (d,x i ) <strong>and</strong> γ 2 (d,x i+1 ) to bubblepopulation at pivots x i <strong>and</strong> x i+1 , respectively. This isnecessary because not all coalescence <strong>and</strong> breakageresult in a bubble that has a legitimate size. For thenonvalid daughter particles, one has to distribute thenew bubble in fractions γ 1 (d,x i ) <strong>and</strong> γ 2 (d,x i+1 ) <strong>of</strong> the twoneighboring size pivots. To determine these two fractions,one needs two balance equations. The first balanceequation relates to the total volume or mass <strong>of</strong> thebubbles, to ensure mass conservation. The second balanceequation is the number balance for the bubblesinvolved in the breakage <strong>and</strong> coalescence processesMN j a C (x i ,x j )i ) 0, 1, ..., M (135)γ 1 (d,x i ) V b,i (d) + γ 2 (d,x i+1 ) V b,i+1 (d) ) V b (d) (136)γ 1 (d,x i ) + γ 2 (d,x i+1 ) ) 1 (137)Alternatively, the second balance equation is sometimessubstituted by a bubble surface balance as one wantsto ensure good estimates <strong>of</strong> the contact areaγ 1 (d,x i ) a b,i + γ 2 (d,x i+1 ) a b,i+1 ) a b (d) (138)The fractions γ 1 (d,x i ) <strong>and</strong> γ 2 (d,x i+1 ) need to be embeddedin the source terms as part <strong>of</strong> the discretizationprocedure. Thus, the source terms are modified todN idt + ∇‚(N i 〈v〉 N i)Mx) ∑γ 1 (d,x i ) N j v(x j ) b B (x j )∫iPBxi(d|x-1 j )ddj)iMx+ ∑γ 2 (d,x i ) N j v(x j ) b B (x j )∫i+1PBxi(d|x j )dd - b B (x i )N ij)ii-1N j+ 1 2 ∑ j)0i-1N j+ 1 2 ∑ j)0∑(x j +x k )∈I iN k γ 1 (d,x i )a(x k ,x j )∑ N k γ 2 (d,x i )a(x k ,x j ) - N i∑(x j +x k )∈I i-1 j)0The breakage function, P B (d|d′) dd, is the number <strong>of</strong>particles formed between size d′ <strong>and</strong> d′ + dd′ dividedby the total number <strong>of</strong> size d particles broken. At thepivots, the corresponding breakage function is definedas P B (x i |x k )P B (x i |x k ) ) ∫ xi-1In this method, the smallest bubbles do not break up,<strong>and</strong> the largest bubbles are not involved in the coalescenceprocess. To cover a broad range in bubble volume(diameter), a large number <strong>of</strong> classes is needed, makingthis algorithm rather time-consuming. However, themethod provides information on the bubble size distributionthat is needed in the multifluid framework <strong>and</strong>can also be used for proper validation <strong>of</strong> the source termclosures. The simpler method <strong>of</strong> moments containingtransport equations for only a few moments such as themean bubble size, the variance, etc., cannot be used ina multifluid framework because <strong>of</strong> the lack <strong>of</strong> any bubblesize resolution <strong>and</strong> can thus be validated only in anaverage sense even if experimental data on the sizedistribution are provided. For very simple distributionfunctions, it might be possible to reconstruct the physicalsize distributions with sufficient accuracy utilizingthe information provided by a few moments only (i.e.,as provided by the moment models). However, as theexisting integral models provide two to five momentsonly, it is believed that the complex bubble size distributionscannot be reproduced with sufficient accuracy.It is then an open question as to whether the computationalcosts <strong>of</strong> solving transport equations for about10-15 moments, ensuring sufficient accuracy, are lessthan the corresponding costs <strong>of</strong> solving for the wholesize distributions using spectral methods. Further workin our group continues to elucidate this issue.Another method representing an extension <strong>of</strong> theQMOM method has obtained increasing attention forparticulate systems during the past several years.According to Fan et al., 101 one <strong>of</strong> the main limitationsMN j a C (x i ,x j )(139)x iPB (d|x k )dd + ∫ xix i+1PB(d|x k )dd (140)


5146 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005<strong>of</strong> the QMOM is that the solid phase is representedthrough the moments <strong>of</strong> the distribution, so the phaseaveragevelocity <strong>of</strong> the different solid phases must beused to solve the transport equations for the moments.Thus, to use this method in the context <strong>of</strong> multiphaseflows, it is necessary to extend QMOM to h<strong>and</strong>le caseswhere each particle size is convected by its own velocity.To address these issues, a direct quadrature method <strong>of</strong>moments (DQMOM) has been formulated by Fan etal. 101 DQMOM is based on the direct solution <strong>of</strong> thetransport equations for weights <strong>and</strong> abscissas <strong>of</strong> thequadrature approximation (i.e., quadrature approximationsfor the moments). Moreover, each node <strong>of</strong> thequadrature approximation is treated as a distinct solidphase. DQMOM is thus used as a module in multifluidmodels describing polydisperse solid phases undergoingsegregation, growth, aggregation, <strong>and</strong> breakage processesin the context <strong>of</strong> CFD simulations.However, in the literature, there seems to be someinconsistency regarding whether the QMOM is reallywell posed or not. 188To date, no reports have been published evaluatingthe behavior <strong>of</strong> this procedure for bubbly flows.7. Concluding RemarksThe past two decades have seen substantial developmentsin both experimental description <strong>and</strong> modeling<strong>of</strong> bubbly flows. However, the physical underst<strong>and</strong>ing<strong>of</strong> the local spatial <strong>and</strong> temporal scales is still verylimited, <strong>and</strong> thus, the modeling tools are restricted tovery low gas fractions.Average multifluid models determining pressure <strong>and</strong>phase velocities combined with an integrated populationbalance module predicting the phase <strong>and</strong> bubble sizedistributions have been found to represent a trade<strong>of</strong>fbetween accuracy <strong>and</strong> computational efforts for practicalapplications. During the past decade, it has become clearthat, to resolve the physics <strong>of</strong> bubble column flows,dynamic <strong>and</strong> three-dimensional fluid dynamic modelsare required.Considering the fluid dynamic part <strong>of</strong> the model, thegeneral picture from the literature is that time-averagedvelocity fields are reasonably well-predicted with models<strong>of</strong> this type. The prediction <strong>of</strong> phase distribution phenomena,on the other h<strong>and</strong>, is still a problem, particularlyat high gas flow rates.The limiting steps in the model development are theformulation <strong>of</strong> closure relations or closure laws determiningturbulence effects, interfacial transfer fluxes,<strong>and</strong> bubble coalescence <strong>and</strong> breakage processes. Improvedformulations <strong>of</strong> the boundary conditions are alsoneeded for flows with higher gas volume fractions,especially at the inlet because <strong>of</strong> the unresolved interactionbetween the fluid dynamics <strong>and</strong> the bubble size <strong>and</strong>phase distributions <strong>and</strong> at the outlet because <strong>of</strong> thehighly recirculating flow phenomena observed at the top<strong>of</strong> the column. In a similar manner, the wall boundaryneeds further development because <strong>of</strong> the complexinteraction between multiphase fluid dynamics, colloidchemistry, bubble coalescence <strong>and</strong> breakage, <strong>and</strong> theturbulence boundary layer structure. Finally, the tremendouscomputational requirements in terms <strong>of</strong> bothCPU time <strong>and</strong> memory have to be reduced by developingmore efficient parallel algorithms. The numerical accuracy<strong>and</strong> stability <strong>of</strong> these schemes are also <strong>of</strong> criticalimportance, as the numerical solution algorithm determinesan integrated part <strong>of</strong> the bubble column modeldevelopment.Considering the population balance modeling <strong>of</strong> bubblyflows, the general picture from the literature is thatthese models are at a very early stage <strong>of</strong> development<strong>and</strong> that the physicochemical fluid dynamics involvedare not yet sufficiently understood. For bubbly flows,there is even some confusion regarding the formulation<strong>of</strong> the governing population balance equation. Most <strong>of</strong>the reports on bubbly flows resort to a rather empiricalmacroscopic formulation, whereas only a few papersadopt microscopic <strong>and</strong> more generalized concepts. Most<strong>of</strong> the existing coalescence <strong>and</strong> breakage closures arethus preferably formulated directly within the macroscopicframework. The breakage closures are <strong>of</strong>ten basedon an incomplete discrete (eddy) interpretation <strong>of</strong> theturbulence structure. These formulations are very difficultor impossible to validate directly on the bubblescalelevel <strong>and</strong> are considered semiempirical in nature.In addition, the commonly used exponential relationsfor both the coalescence <strong>and</strong> breakage probabilities havenot been validated in a sufficient manner for bubblecolumn flows. The coalescence closures are also limitedbecause <strong>of</strong> the lack <strong>of</strong> underst<strong>and</strong>ing in formulatingaccurate coalescence criteria. Proper validation <strong>of</strong> theexisting breakage criteria for bubble column flows is alsorecommended. Both the turbulent breakage <strong>and</strong> coalescenceclosures are very sensitive to the turbulenceenergy dissipation rate, which is difficult to estimatewith sufficient accuracy using the st<strong>and</strong>ard turbulencemodels.Further, by the use <strong>of</strong> number probability densityfunctions within the population balance framework, anunfortunate inconsistency occurs regarding the interpretation<strong>of</strong> the velocity variable involved in the integratedmultifluid/population balance model. In themultifluid model formulation, the velocity variable isusually a mass-average quantity, whereas the correspondingvariable within the population balance is anumber-density-average variable. It would probably beadvantageous to formulate a population balance for theparticle mass rather than the number density to ensurebetter consistency between the two model parts havingsomewhat different origins.However, it is expected that the next generation <strong>of</strong>population balances will be formulated adopting themore fundamental statistical concepts on the microscopiclevel <strong>and</strong> that the coalescence <strong>and</strong> breakageclosures will be reformulated in a consistent manner.The formulation <strong>of</strong> optimal numerical solution methodssolving the integrated multifluid/population balancemodel is not a trivial task. In most cases, the populationbalance part <strong>of</strong> the model is solved by adopting thesimplest methods originally developed for solid particleanalysis. Because the mathematical properties <strong>of</strong> thephysical processes <strong>and</strong> closures involved can deviateconsiderably, the choice <strong>of</strong> solution strategy needsfurther consideration.In parallel to the modeling work, intensive effortshave to be focused on well-planned experimental programsto enable better underst<strong>and</strong>ing <strong>of</strong> the phenomenato be described by the microscopic modeling process <strong>and</strong>for model validation. In this respect, extended integrations<strong>and</strong> knowledge transfer between the modeling <strong>and</strong>experimental groups will be beneficial.


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5147AcknowledgmentThis work received support from the Research Council<strong>of</strong> Norway (Program <strong>of</strong> Supercomputing) through agrant <strong>of</strong> computer-time. The Ph.D. fellowships (H.L. <strong>and</strong>C.A.D.) financed by the Research Council <strong>of</strong> Norwaythrough a Strategic University Program (CARPET) aregratefully appreciated.NomenclatureLatin Lettersa C (x,r;x′,r′,Y,t) ) coalescence frequency (general), i.e.,fraction <strong>of</strong> particles <strong>of</strong> sizes d <strong>and</strong> d′ coalescing per unittime (l/s)a C (d,d′,Y) ) average coalescence frequency (m 3 /s)A ) parameter in the drag coefficient relationb B (d) ) breakage frequency (l/s)B B (x,r,Y,t), B C (x,r,Y,t) ) birth rates due to breakage <strong>and</strong>coalescence (general), respectively {1/[s m 3 (x)]}B B (d), B C (d) ) average birth rates due to breakage <strong>and</strong>coalescence, respectively, with d as the internal coordinate{1/[s m 3 (m)]}c ) concentration <strong>of</strong> surfactant species (kmol/m 3 )C D ) drag coefficientC f ) wall friction coefficientC k ) Kolmogorov energy spectrum parameter ()1.5)C L ) lift coefficientC V ) added-mass force coefficientC wb ) wall boundary coefficientC w1 ) wall lift coefficientC w2 ) wall lift coefficientd ) bubble diameter (m)d c ) critical diameter (m)d i ) diameter <strong>of</strong> particle in interval i (m)d(i) ) diameter <strong>of</strong> particle in class i (m)d′ ) diameter <strong>of</strong> daughter bubble (m)d′′ ) diameter <strong>of</strong> the smallest daughter bubble (m)d˜ )diameter <strong>of</strong> complementary daughter particle, d˜ )(d 3- d′ 3 ) 1/3 (m)d max , d min ) boundary diameters <strong>of</strong> class to be split (m)d s ) Sauter mean diameter (m)dV x ) infinitesimal volume in property spacedV r ) infinitesimal volume in physical space (m 3 )D B (x,r,Y,t), D C (x,r,Y,t) ) death rates due to breakage <strong>and</strong>coalescence (general), respectively {1/[s m 3 (x)]}D B (d), D C (d) ) average death rates due to breakup <strong>and</strong>coalescence, respectively, with d as the internal coordinate{1/[s m 3 (m)]}ej(d i ,λ) ) mean kinetic energy <strong>of</strong> an eddy (J)e s (d i ,d j ) ) increase in surface energy due to binary breakage(J)E ) constant <strong>of</strong> integrationE eddies (λ), E spectra (λ) ) kinetic energy contained within eddies<strong>of</strong> size between λ <strong>and</strong> λ + dλ {J/[m 3 (m)]}E(k) ) kinetic energy contained within eddies <strong>of</strong> wavenumbersbetween k <strong>and</strong> k + dk [m 2 (m)/s 2 ]E s ) minimum energy needed to deform a particle (J)E 0 ) Eötvös numberf(d,t) ) particle number density probability {1/[m 3 (m)]}f(x,r,t) ) particle number density (general) {1/[m 3 (m)]}f vij ) breakage volume fraction ()d j3 /d i3 )f λ ) number <strong>of</strong> eddies <strong>of</strong> size between λ <strong>and</strong> λ + dλ {1[m 3s (m)]h ) film thickness (m)h C ) effective swept volume rate (m 3 /s)h f ) final film thickness (m)h 0 ) initial film thickness (m)k ) turbulent kinetic energy (m 2 /s 2 ), or wavenumber (1/m)n i ) number density in the interval i (1/m 3 )N ) total number <strong>of</strong> particles per unit volume <strong>of</strong> physicalspace (1/m 3 )N i ) number density <strong>of</strong> particles in size interval i (1/m 3 )p ) pressure (N/m 2 )p B (d i :d j ) ) breakage probability (or breakage efficiency)p B (d i :d j ,λ) ) breakage probability (or breakage efficiency),dependent on eddy sizep C ) coalescence probability (or coalescence efficiency)P B (d|d′,Y) ) average probability density function, or daughtersize distribution function [1/(m)]P B (x,r|x′,r′,Y,t) ) probability density function (general), ordaughter size distribution function (general) {1/[m 3 (X)]}r b ) bubble radius (m)R ) molar gas constant [J/(mol K)]R d ) radius <strong>of</strong> liquid disk between coalescing bubbles (m)Re B ) particle Reynolds numberSc ) Schmidt numbert ) time coordinate (s)t b ) particle breakage time (s)T ) temperature (K)vj λ ) mean eddy velocity expressed by the Maxwell distributionfunction (m/s)v 2 (λ) ) second-order structure function (m 2 /s 2 )V ) volume, physical or abstractV i ) volume <strong>of</strong> a bubble in class iWe ) Weber numberx i , x j ) pivotal points in I i <strong>and</strong> I jy ) distance from the wall (m)y + ) inner boundary layer length scale (m)z ) axial coordinate (m)VectorsF i,k ) interfacial force (i ) D, V, L, TD, W) [kg/(m 2 s 2 )]g ) acceleration <strong>of</strong> gravity, or gravity force per unit mass(m/s 2 )M k,l ) interfacial momentum transfer to phase k fromphase l [kg/(m 2 s 2 )]n w ) unit vector normal to the walln ) outward-directed normal unit vectorr ) position vector in physical space (m)R4 ) external coordinates (m)v ) velocity (m/s)v d ) velocity <strong>of</strong> dispersed phase (m/s)v k ) velocity <strong>of</strong> phase k (m/s)v p ) velocity <strong>of</strong> particles (m/s)v rel ) relative velocity between the phases (m/s)x ) (internal) property vectorX4 ) internal coordinates (m)Y ) vector representing the continuous phase variablesGreek lettersR k ) volume fraction <strong>of</strong> phase kβ ) empirical model parameterβ˜ )empirical model parameter in turbulence theoryγ ) bubble fractionΓ ) incomplete gamma functionΓ k,l ) net mass-transfer flux to phase k from all other lphasesδ ) correction <strong>of</strong> redundancy factor, or delta functionɛ ) turbulent energy dissipation rateλ ) eddy size, or turbulent length scale (m)Λ ) bubble diameter ratioµ j ) jth moment <strong>of</strong> the population density functionµ k ) viscosity <strong>of</strong> phase k [kg/(m s)]ν(d,t) ) average number <strong>of</strong> particles formed from thebreakupξ ) eddy/bubble size ratio ()λ/d i )ξ ij ) bubble/bubble size ratio () d j /d i )F k ) density <strong>of</strong> phase k (kg/m 3 )σ A ) effective collision cross-sectional area (m 2 )


5148 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005σ I ) surface tension (N/m)σ k , σ ɛ ) Pr<strong>and</strong>tl numbersσj k ) viscous stress tensor (Pa)σ s ) surface restoring pressure (Pa)σ t ) average deformation stress (Pa)φ ) net source termχ ) energy ratio [e s (d i ,d j )/ej(d i ,λ)]χ c ) critical energy ratioψ ) extensive propertyω B (d i :d j ) ) eddy-bubble collision density [1/(s m 3 )]ω B I (d i :λ) ) eddy-bubble collision probability density {1/[sm 3 (m)]}ω C ) bubble-bubble collision density {1/[m 3 s (m) (m)]ω C B ) buoyancy collision density [1/(s m 3 )]ω C LS ) collision density due to laminar shear [1/(s m 3 )]ω f1 ) single-particle collision density [1/s (m)]ω C T ) collision density resulting from turbulent motion (sm 3 )Ω B ) breakage rate [1/(s m 3 )]Ω C ) coalescence rate [1/(s m 3 )]Literature Cited(1) Deckwer, W.-D.; Schumpe, A. <strong>Bubble</strong> columnsThe state <strong>of</strong>the art <strong>and</strong> current trends. Int. Chem. Eng. 1987, 27, 405.(2) Deckwer, W.-D. <strong>Bubble</strong> <strong>Column</strong> <strong>Reactors</strong>; John Wiley &Sons: Chichester, U.K., 1992.(3) Deckwer, W.-D.; Schumpe, A. Improved Tools for <strong>Bubble</strong><strong>Column</strong> Reactor Design <strong>and</strong> Scale-up. Chem. Eng. Sci. 1993, 48,889.(4) Dudukovic, M. P. Trends in Catalytic Reaction Engineering.Catal. Today 1999, 48, 5.(5) Dudukovic, M. P.; Larachi, F.; Mills, P. L. MultiphasereactorssRevisited. Chem. Eng. Sci. 1999, 54, 1975.(6) Dudukovic, M. P.; Larachi, F.; Mills, P. L. MultiphaseCatalytic <strong>Reactors</strong>: A Perspective on Current Knowledge <strong>and</strong>Future Trends. Catal. Rev. 2002, 44, 123.(7) Shah, Y. T.; Sharma, M. M. Gas-Liquid-Solid <strong>Reactors</strong>.In Chemical Reaction <strong>and</strong> Reaction Engineering; A Series <strong>of</strong>Reference Books <strong>and</strong> Textbooks; Carberry, J. J., Varma, A., Eds.;Marcel Dekker: New York, 1987; p 667.(8) Jakobsen, H. A.; Sannæs, B. H.; Grevskott, S.; Svendsen,H. F. <strong>Modeling</strong> <strong>of</strong> Vertical <strong>Bubble</strong>-Driven Flows. Ind. Eng. Chem.Res. 1997, 36, 4052.(9) Joshi, J. B. Computational flow modeling <strong>and</strong> design <strong>of</strong>bubble column reactors. Chem. Eng. Sci. 2001, 56, 5893.(10) Rafique, M.; Chen, P.; Dudukovic, M. P. Computationalmodeling <strong>of</strong> gas-liquid flow in bubble columns. Rev. Chem. Eng.2004, 20, 225.(11) Jakobsen, H. A. Phase distribution phenomena in twophasebubble column reactors. Chem. Eng. Sci. 2001, 56, 1049.(12) van den Akker, H. E. A. Computational Fluid Dynamics:More Than a Promise to Chemical Reaction Engineering, PaperPresented at the CHISA 2000 Conference, Praha, Czech Republic,Aug 27-31, 2000.(13) Carrica, P. M.; Drew, D.; Bonetto, F.; Lahey, R. T., Jr. APolydisperse Model for Bubbly Two-Phase Flow Around a SurfaceShip. Int. J. Multiphase Flow 1999, 25, 257.(14) Oey, R. S.; Mudde, R. F.; van den Akker, H. E. A.Sensitivity Study on Interfacial Closure Laws in Two-Fluid BubblyFlow Simulations. AIChE 2003, 49, 1621.(15) Sokolichin, A.; Eigenberger, G.; Lapin, A. Simulation <strong>of</strong>Buoyancy Driven Bubbly Flow: Established Simplifications <strong>and</strong>Open Questions. AIChE J. 2004, 50, 24.(16) Behzadi, A.; Issa, R. I.; Rusche, H. Modelling <strong>of</strong> dispersedbubble <strong>and</strong> droplet flow at high phase fractions. Chem. Eng. Sci.2004, 59, 759.(17) Svendsen, H. F.; Jakobsen, H. A.; Torvik, R. Local FlowStructures in Internal Loop <strong>and</strong> <strong>Bubble</strong> <strong>Column</strong> <strong>Reactors</strong>. Chem.Eng. Sci. 1992, 47, 3297.(18) Sanyal, J.; Vasquez, S.; Roy, S.; Dudukovic, M. P. Numericalsimulation <strong>of</strong> gas-liquid dynamics in cylindrical bubblecolumn reactors. Chem. Eng. Sci. 1999, 54, 5071.(19) Krishna, R.; van Baten, J. M. Rise Characteristics <strong>of</strong> Gas<strong>Bubble</strong>s in a 2D Rectangular <strong>Column</strong>: VOF Simulations vsExperiments. Int. Commun. Heat Mass Transfer 1999, 26, 965.(20) Jakobsen, H. A. On the <strong>Modeling</strong> <strong>and</strong> Simulation <strong>of</strong> <strong>Bubble</strong><strong>Column</strong> <strong>Reactors</strong> Using a Two-Fluid Model. Dr. Ing. Thesis,Norwegian Institute <strong>of</strong> Technology, Trondheim, Norway, 1993.(21) Tzeng, J.-W.; Chen, R. C.; Fan, L.-S. Visualization <strong>of</strong> FlowCharacteristics in a 2D <strong>Bubble</strong> <strong>Column</strong> <strong>and</strong> Three-Phase FluidizedBed. AIChE J. 1993, 39, 733.(22) Lin, T.-J.; Reese, J.; Hong, T.; Fan, L.-S. QuantitativeAnalysis <strong>and</strong> Computation <strong>of</strong> Two-Dimensional <strong>Bubble</strong> <strong>Column</strong>s.AIChE J. 1996, 42, 301.(23) Chen, R. C.; Reese, J.; Fan, L.-S. Flow Structure in Three-Dimensional <strong>Bubble</strong> <strong>Column</strong> <strong>and</strong> Three-Phase Fluidized Bed.AIChE J. 1994, 40, 1093.(24) Sokolichin, A.; Eigenberger, G. Gas-liquid flow in bubblecolumns <strong>and</strong> loop reactors: Part I. Detailed modeling <strong>and</strong> numericalsimulation. Chem. Eng. Sci. 1994, 49, 5735.(25) Sokolichin, A.; Eigenberger, G.; Lapin, A.; Lübbert, A.Dynamical numerical simulation <strong>of</strong> gas-liquid two-phase flows.Euler/Euler versus Euler/Lagrange. Chem. Eng. Sci. 1997, 52, 611.(26) Sokolichin, A.; Eigenberger, G. Applicability <strong>of</strong> the k-ɛturbulence model to the dynamic simulation <strong>of</strong> bubble columns:Part I. Detailed numerical simulations. Chem. Eng. Sci. 1999, 54,2273.(27) Becker, S.; Sokolichin, A.; Eigenberger, G. Gas-LiquidFlow in <strong>Bubble</strong> <strong>Column</strong>s <strong>and</strong> Loop <strong>Reactors</strong>: Part II. Comparison<strong>of</strong> Detailed Experiments <strong>and</strong> Flow Simulations. Chem. Eng. Sci.1994, 49, 5747.(28) Becker, S.; De Bie, H.; Sweeney, J. Dynamics flow behaviourin bubble columns. Chem. Eng. Sci. 1999, 54, 4929.(29) Krishna, R.; van Baten, J. M. Scaling up <strong>Bubble</strong> <strong>Column</strong><strong>Reactors</strong> with Highly Viscous Liquid Phase. Chem. Eng. Technol.2002, 25, 1015.(30) Pfleger, D.; Becker, S. <strong>Modeling</strong> <strong>and</strong> simulation <strong>of</strong> thedynamic flow behaviour in a bubble column. Chem. Eng. Sci. 2001,56, 1737.(31) Olmos, E.; Gentric, C.; Vial, Ch.; Wild, G.; Midoux, N.Numerical simulation <strong>of</strong> multiphase flow in bubble column reactors.Influence <strong>of</strong> bubble coalescence <strong>and</strong> breakup. Chem. Eng. Sci.2001, 56, 6359.(32) Lehr, F.; Millies, M.; Mewes, D. <strong>Bubble</strong>-Size Distributions<strong>and</strong> Flow Fields in <strong>Bubble</strong> <strong>Column</strong>s. AIChE J. 2002, 48, 2426.(33) Olmos, E.; Gentric, C.; Midoux, N. Numerical description<strong>of</strong> flow regime transitions in bubble column reactors by multiplegas-phase model. Chem. Eng. Sci. 2003, 58, 2113.(34) Bertola, F.; Vanni, M.; Baldi, G. Application <strong>of</strong> ComputationalFluid Dynamics to Multiphase Flow in <strong>Bubble</strong> <strong>Column</strong>s.Int. J. Chem. React. Eng. 2003, 1, 1.(35) Krishna, R.; van Baten, J. M. Scaling up <strong>Bubble</strong> columnreactors with the aid <strong>of</strong> CFD. Trans. Inst. Chem. Eng. 2001, 79,283.(36) Lehr, F.; Mewes, D. A transport equation for the interfacialarea density applied to bubble columns. Chem. Eng. Sci. 2001,56, 1159.(37) Mudde, R. F.; Lee, D. J.; Reese, J.; Fan, L.-S. Role <strong>of</strong>Coherent Structures on Reynolds Stresses in a 2-D <strong>Bubble</strong><strong>Column</strong>. AIChE J. 1997, 43, 913.(38) Mudde, R. F.; Simonin, O. Two- <strong>and</strong> three-dimensionalsimulations <strong>of</strong> bubble plume using a two-fluid model. Chem. Eng.Sci. 1999, 54, 5061-5069.(39) Pfleger, D.; Gomes, S.; Gilbert, N.; Wagner, H.-G. Hydrodynamicsimulations <strong>of</strong> laboratory scale bubble columns fundamentalstudies <strong>of</strong> the Eulerian-Eulerian modelling approach.Chem. Eng. Sci. 1999, 54, 5091.(40) Deen, N. G.; Solberg, T.; Hjertager, B. H. Large EddySimulation <strong>of</strong> the Gas-Liquid Flow in a Square Cross-Sectioned<strong>Bubble</strong> <strong>Column</strong>. Chem. Eng. Sci. 2001, 56, 6341.(41) Buwa, V. V.; Ranade, V. V. Dynamics <strong>of</strong> gas-liquid flowin a rectangular bubble column: Experimental <strong>and</strong> single/multigroupCFD simulations. Chem. Eng. Sci 2002, 57, 4715.(42) Buwa, V. V.; Ranade, V. V. Mixing in <strong>Bubble</strong> <strong>Column</strong><strong>Reactors</strong>: Role <strong>of</strong> Unsteady Flow Structures. Can. J. Chem. Eng.2003, 81, 402.(43) Bove, S.; Solberg, T.; Hjertager, B. H., Numerical Aspects<strong>of</strong> <strong>Bubble</strong> <strong>Column</strong> Simulations. Int. J. Chem. React. Eng. 2004, 2.(44) Lopez de Bertodano, M. Turbulent Bubbly Two-Phase Flowin a Triangular Duct. Ph.D. Thesis, Rensselaer PolytechnicInsitute, Troy, New York, 1992.(45) Anglart, H.; Andersson, S.; Podowski, M. Z.; Kurul, N. AnAnalysis <strong>of</strong> Multidimensional Void Distribution in Two-PhaseFlows. In Proceedings <strong>of</strong> the Sixth International Topic Meeting on


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5149Nuclear Reactor Thermal HydraulicssNURETH 6; Courtaud, M.,Delhaye, J. M., Eds.; 1993; Vol. 1, pp 139-153.(46) Antal, S. P.; Lahey, R. T., Jr.; Al-Dahhan, M. H. SimulatingChurn-Turbulent Flows in a <strong>Bubble</strong> <strong>Column</strong> using a Three Field,Two-Fluid Model. Paper Presented at the 5th InternationalConference on Multiphase Flow, ICMF′04, Yokohama, Japan, May30-Jun 4, 2004; Paper No. 182.(47) Dhotre, M. T.; Joshi, J. B. Two-Dimensional CFD Modelfor the Prediction <strong>of</strong> Flow Pattern, Pressure Drop <strong>and</strong> HeatTransfer Coefficient in <strong>Bubble</strong> <strong>Column</strong> <strong>Reactors</strong>. Trans. Inst.Chem.Eng., Chem. Eng. Res. Des. 2004, 82, 689.(48) Tchen, C. M. Mean value <strong>and</strong> correlation problems connectedwith the motion <strong>of</strong> small particles suspended in a turbulentfluid. Ph.D. Thesis, TU Delft, Delft, The Netherl<strong>and</strong>s, 1947.(49) Simonin, O. Eulerian formulation for particle dispersionin turbulent two-phase flows. In Fifth Workshop on Two-PhaseFlow Predictions; Sommerfeld, M., Wennerberg, P., Eds.; FRG:Erlangen, Germany, 1990; pp 156-166.(50) Simonin, O.; Viollet, P. L. Prediction <strong>of</strong> an oxygen dropletpulverization in a compressible subsonic c<strong>of</strong>lowing hydrogen flow.Numer. Methods Multiphase Flows 1990, 91, 73.(51) Viollet, P. L.; Simonin, O. Modelling dispersed two-phaseflows: Closure, validation <strong>and</strong> s<strong>of</strong>tware development. Appl. Mech.Rev. 1994, 47, S80.(52) Tomiyama, A.; Matsouka, T.; Fukuda, T.; Sakaguchi, T. ASimple Numerical Method for Solving an Incompressible Two-Fluid Model in a General Curvelinear Coordinate System. Adv.Multiphase Flow 1995, Elsevier: 241.(53) Tomiyama, A. Struggle with computational bubble dynamics.Presented at the Third International Conference on MultiphaseFlow, ICMF′98; Lyon, France, Jun 8-12, 1998.(54) Tomiyama, A.; Shimada, N. A Numerical Method forBubbly Flow Simulation Based on a Multi-Fluid Model. J. PressureVessel Technol. 2001, 23, 510.(55) Sato, Y.; Sekoguchi, K. Liquid Velocity Distribution in Two-Phase Bubbly Flow. Int. J. Multiphase Flow 1975, 2, 79.(56) Sato, Y.; Sadotomi, M.; Sekoguchi, K. Momentum <strong>and</strong> HeatTransfer in Two-Phase <strong>Bubble</strong> Flow. Int. J. Multiphase Flow 1981,7, 167.(57) Jakobsen, H. A. Numerical Convection Algorithms <strong>and</strong>Their Role in Eulerian CFD Reactor Simulations. Int. J. Chem.React. Eng. 2003, 1, 1.(58) Bech, K. <strong>Modeling</strong> <strong>of</strong> bubble-driven flow applied to electrolysiscells. Paper Presented at the 5th International Conferenceon Multiphase Flow, ICMF′04, Yokohama, Japan, May 30-Jun4, 2004; Paper 404.(59) Cartl<strong>and</strong> Glover, G. M.; Generalis, S. C. The modeling <strong>of</strong>buoyancy driven flow in bubble columns. Chem. Eng. Process.2004, 43, 101.(60) van den Akker, H. E. A. Coherent structures in multiphaseflows. Powder Technol. 1998, 100, 123.(61) Harteveld, W. K.; Mudde, R. F.; van den Akker, H. E. A.Dynamics <strong>of</strong> a <strong>Bubble</strong> <strong>Column</strong>: Influence <strong>of</strong> Gas Distribution <strong>of</strong>Coherent Structures. Can. J. Chem. Eng. 2003, 81, 389.(62) Politano, M. S.; Carrica, P. M.; Larreteguy, A. E. Gas/Liquid Polydisperse Two-Phase Flow <strong>Modeling</strong> for <strong>Bubble</strong> <strong>Column</strong><strong>Reactors</strong>. Paper Presented at the 3rd Mercosur Congress onProcess Systems Engineering 1st Mercosur Congress on ChemicalEngineering, Sante Fe, Argentina, Sep 16-20, 2001.(63) Politano, M. S.; Carrica, P. M.; Converti, J. A model forturbulent polydisperse two-phase flow in vertical channels. Int.J. Multiphase Flow 2003, 29, 1153.(64) Lo, S. Application <strong>of</strong> population balance to CFD modelling<strong>of</strong> bubbly flow via the MUSIG model. AEA Technol. 1996, AEAT-1096.(65) Lo, S. Some recent developments <strong>and</strong> applications <strong>of</strong> CFDto multiphase flows in stirred reactors. Presented at the AMIF-ESF Workshop: Computing Methods for Two-Phase Flow, Aussois,France, Jan 12-14, 2000.(66) Bertola, F.; Grundseth, J.; Hagesaether, L.; Dorao, C.; Luo,H.; Hjarbo, K. W.; Svendsen, H. F.; Vanni, M.; Baldi, G.; Jakobsen,H. A. Numerical Analysis <strong>and</strong> Experimental Validation <strong>of</strong> <strong>Bubble</strong>Size Distribution in Two-Phase <strong>Bubble</strong> <strong>Column</strong> <strong>Reactors</strong>.Multiphase Sci. Technol. 2005, 17 (1-4), 1-23.(67) Kocamustafaogullari, G.; Ishii, M. Foundation <strong>of</strong> theinterfacial area transport equation <strong>and</strong> its closure relations. Int.J. Heat Mass Transfer 1995, 38, 481.(68) Millies, M.; Mewes, D. Interfacial area density in bubblyflow. Chem. Eng. Sci. 1999, 38, 307.(69) White, A. J.; Hounslow, M. J. Modelling droplet sizedistributions in polydispersed wet-stream flows. Int. J. Heat MassTransfer 2000, 43, 1873.(70) Luo, H. Coalescence, break-up <strong>and</strong> liquid circulation inbubble column reactors. Dr. Ing. Thesis, Department <strong>of</strong> ChemicalEngineering, The Norwegian Institute <strong>of</strong> Technology, Trondheim,Norway, 1993.(71) Luo, H.; Svendsen, H. F. Theoretical Model for Drop <strong>and</strong><strong>Bubble</strong> Breakup in Turbulent Dispersions. AIChE J. 1996, 42,1225.(72) Hagesaether, L.; Jakobsen, H. A.; Svendsen, H. F. TheoreticalAnalysis <strong>of</strong> Fluid Particle Collisions in Turbulent Flow.Chem. Eng. Sci. 1999, 54, 4749.(73) Hagesaether, L.; Jakobsen, H. A.; Hjarbo, K.; Svendsen,H. F. A Coalescence <strong>and</strong> Breakup Module for Implementation inCFD-codes. Comput.-Aided Chem. Eng. 2000, 8, 367.(74) Hagesaether, L.; Jakobsen, H. A.; Svendsen, H. F. A Modelfor Turbulent Binary Breakup <strong>of</strong> Dispersed Fluid Particles. Chem.Eng. Sci. 2002, 57, 3251.(75) Hagesaether, L.; Jakobsen, H. A.; Svendsen, H. F. <strong>Modeling</strong><strong>of</strong> the Dispersed-Phase Size Distribution in <strong>Bubble</strong> <strong>Column</strong>s.Ind. Eng. Chem. Res. 2002, 41, 2560.(76) Jakobsen, H. A.; Hjarbo, K. W.; Svendsen, H. F. <strong>Bubble</strong>size distributions, bubble velocities <strong>and</strong> local gas fractions fromconductivity measurements; Report STF21 F93103; SINTEF AppliedChemistry: Trondheim, Norway, 1994.(77) Prince, M. J.; Blanch, H. W. <strong>Bubble</strong> Coalescence <strong>and</strong>Break-Up in Air-Sparged <strong>Bubble</strong> <strong>Column</strong>s. AIChE J. 1990, 36,1485.(78) Ranade, V. V. Computational Flow <strong>Modeling</strong> for ChemicalReactor Engineering; Academic Press: San Diego, 2002.(79) Ramakrishnan, S.; Kumar, R.; Kuloor, N. R. Studies inbubble formation-I: <strong>Bubble</strong> formation under constant flow conditions.Chem. Eng. Sci. 1969, 24, 731.(80) Geary, N. W.; Rice, R. G. <strong>Bubble</strong> Size Prediction for Rigid<strong>and</strong> Flexible Spargers. AIChE J. 1991, 37, 161.(81) Gresho, P. M.; Sani, R. L. On pressure boundary conditionsfor the incompressible Navier-Stokes equations. Int. J. Numer.Methods Fluids 1987, 7, 1111.(82) Padial, N. T.; V<strong>and</strong>erHeyden, W. B.; Rauenzahn, R. M.;Yarbro, S. L. Three-dimensional simulation <strong>of</strong> a three-phase drafttubebubble column. Chem. Eng. Sci. 2000, 55, 3261.(83) Troshko, A. A.; Hassan, Y. A. Law <strong>of</strong> the wall for two-phaseturbulent boundary layers. Int. J. Heat Mass Transfer 2001, 44,871.(84) Troshko, A. A.; Hassan, Y. A. A two-equation turbulencemodel <strong>of</strong> turbulent bubbly flows. Int. J. Multiphase Flow 2001,27, 1965.(85) Oliveira, P. J.; Issa, R. I. On the numerical treatment <strong>of</strong>interfaphase forces in two-phase flow. In Numerical Methods inMultiphase Flows; Crowe, C. T., Ed.; ASME Press: New York,1994; Vol. 185, p 131.(86) Spalding, D. B. The calculation <strong>of</strong> free-convection phenomenain gas-liquid mixtures. In Turbulent Buoyant Convection;Hemisphere: Washington, DC, 1977; p 569.(87) Spalding, D. B. Numerical computation <strong>of</strong> multiphase fluidflow <strong>and</strong> heat transfer. In Recent Advances in Numerical Methodsin Fluids; Taylor, C., Morgan, K., Eds.; Pineridge Press: Swansea,U.K., 1980; p 139.(88) Kelly, J. M.; Stewart, C. W.; <strong>and</strong> Cuta, J. M. VIPRE-02sA two-fluid thermal-hydraulics code for reactor core <strong>and</strong> vesselanalysis: Mathematical modeling <strong>and</strong> solution methods. NucearTechnol. 1992, 100, 246.(89) Kunz, R. F.; Siebert, B. W.; Cope, W. K.; Foster, N. F.;Antal, S. P.; Ettorre, S. M. A coupled phasic exchange algorithmfor three-dimensional multi-field analysis <strong>of</strong> heated flows withmass transfer. Comput. Fluids 1998, 27, 741.(90) Lathouwers, D.; van den Akker, H. E. A. A numericalmethod for the solution <strong>of</strong> two-fluid model equations. In Numer.Methods Multiphase Flows 1996, 236, 121.(91) Lathouwers, D. <strong>Modeling</strong> <strong>and</strong> simulation <strong>of</strong> turbulentbubbly flow. Ph.D. Thesis, The Technical University <strong>of</strong> Delft, Delft,The Netherl<strong>and</strong>s, 1999.(92) Antal, S. P.; Ettorre, S. M.; Kunz, R. F.; Podowski, M. Z.Development <strong>of</strong> a Next Generation Computer Code for the Prediction<strong>of</strong> Multicomponent Multiphase Flows. Presented at theInternational Meeting on Trends in Numerical <strong>and</strong> Physical


5150 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005<strong>Modeling</strong> for Industrial Multiphase Flow 2000, Cargese, France,Sep 27-29, 2000.(93) Kunz, R. F.; Yu, W.-S.; Antal, S. P.; Ettorre, S. M. AnUnstructured Two-Fluid Method Based on the Coupled PhasicExchange Algorithm; Report AIAA-2001-2672; American Institute<strong>of</strong> Aeronautics <strong>and</strong> Astronautics: Herndon, VA, 2001.(94) Vasquez, S. A.; Ivanov, V. A. A phase coupled method forsolving multiphase problems on unstructured meshes. In Proceedings<strong>of</strong> ASME FEDSM′00; ASME Press: New York, 2000; p 1.(95) van Santen, H.; Lathouwers, D.; Kleijn, C. R.; van denAkker, H. E. A. Influence <strong>of</strong> segregation on the efficiency <strong>of</strong> finitevolume methods for the incompressible Navier-Stokes equations.In Proceedings <strong>of</strong> the ASME Fluid Engineering Division Conference;ASME Press: New York, 1996; Vol. 238, N. 3, p 151.(96) Lindborg, H.; Eide, V.; Unger, S.; Henriksen, S. T.;Jakobsen, H. A. Parallelization <strong>and</strong> performance optimization <strong>of</strong>a dynamic PDE ractor model for practical applications. Comput.Chem. Eng. 2004, 28, 1585.(97) Caia, C.; Minev, P. A finite element method for an averagedmultiphase flow model. Int. J. Comput. Fluid Dyn. 2004, 18, 111.(98) Jakobsen, H. A.; Bourg, I.; Hjarbo, K. W.; Svendsen, H. F.Interaction Between Reaction Kinetics <strong>and</strong> Flow Structure in<strong>Bubble</strong> <strong>Column</strong> <strong>Reactors</strong>. In Parallel Computational FluidDynamicssTrends <strong>and</strong> Applications; Periaux, J., Jenssen, C. B.,Pettersen, B., Ecer, A., Eds.; Elsevier Science B. V.: Amsterdam,2001; pp 543-550.(99) Jakobsen, H. A.; Svendsen, H. F.; Hjarbo, K. W. On theprediction <strong>of</strong> local flow structures in internal loop <strong>and</strong> bubblecolumn reactors using a two-fluid model. Comput. Chem. Eng.1993, 17S, S531.(100) Jakobsen, H. A. General REActor Technology FUNdamentals.In Lecture Notes in Subject KP8103 Advanced Reactor<strong>Modeling</strong>; NTNU: Trondheim, Norway, 2004; p 890.(101) Fan, R.; Marchisio, D. L.; Fox, R. O. Application <strong>of</strong> thedirect quadrature method <strong>of</strong> moments to polydisperse gas-solidfluidized beds. Powder Technol. 2004, 139, 7.(102) Tomiyama, A.; Miyoshi, K.; Tamai, H.; Zun, I.; Sakaguchi,T. A <strong>Bubble</strong> Tracking Method for the Prediction <strong>of</strong> Spatial-Evolution <strong>of</strong> <strong>Bubble</strong> Flow in a Vertical Pipe. Presented at the ThirdInternational Conference on Multiphase Flow, ICMF′98, Lyon,France, Jun 8-12, 1998.(103) Ishii, M.; Mashima, K. Two Fluid Model <strong>and</strong> HydrodynamicConstitutive Relations. Nucl. Eng. Des. 1984, 82, 107.(104) Drew, D. A.; Lahey, R. T. Application <strong>of</strong> General ConstitutivePrinciples to the Derivation <strong>of</strong> Multidimensional Two-PhaseFlow Equations. Int. J. Multiphase Flow 1979, 7, 243.(105) Antal, S. P.; Lahey, R. T.; Flaherty, J. E. Analysis <strong>of</strong>Phase Distribution in Fully Developed Laminar Bubbly Two-PhaseFlow. Int. J. Multiphase Flow 1991, 17, 635.(106) Marié, J. L.; Moussali, E.; Lance, M. A First Investigation<strong>of</strong> a Bubbly Stress Measurements. Presented at the InternationalSymposium on Turbulence Modification in Multiphase Flows,ASME/JSME Annual Meeting, Portl<strong>and</strong>, OR, Jun 23-27, 1991.(107) Jakobsen, H. A.; Lindborg, H.; H<strong>and</strong>el<strong>and</strong>, V. A numericalstudy <strong>of</strong> the interactions between viscous flow, transport <strong>and</strong>kinetics in fixed bed reactors. Comput. Chem. Eng. 2002, 26, 333.(108) van Leer, B. Towards the Ultimate Conservation DifferenceScheme 2. Monotonicity <strong>and</strong> Conservation Combined in aSecond-Order Scheme. J. Comput. Phys. 1974, 14, 361.(109) van Leer, B. Towards the Ultimate Conservation DifferenceScheme 4. A New Approach to Numerical Convection. J.Comput. Phys. 1977, 23, 276.(110) Grienberger, J. Untersuchung und Modellierung vonBlasensäulen. Dr. Ing. Thesis, Der Technischen Fakultät derUniversität Erlangen-Nürnberg, Erlangen, Germany, 1992.(111) Kurul, N.; Podowski, M. Z. On the <strong>Modeling</strong> <strong>of</strong> MultidimensionalEffects in Bioling Channels. In Proceedings <strong>of</strong> the 1991National Heat Transfer Conference; ASME Press: New York, 1991;Vol. 5, pp 30-40.(112) Morsi, S. A.; Alex<strong>and</strong>er, A. J. An Investigation <strong>of</strong> ParticleTrajectories in Two-Phase Flow Systems. J. Fluid. Mech. 1972,55, 193.(113) Schiller, L.; Naumann, Z. Über die grundlegenden Berechnungenbei der Schwerkraftaufbereitung. Z. Ver. Deutsch. Inc.1935, 77, 318.(114) Schumann, U. Simulations <strong>and</strong> parametrizations <strong>of</strong> largeeddies in convective atmospheric boundary layers. In Proceedings<strong>of</strong> a workshop held at the European Centre for medium-rangeweather forecasts (ECMWF) on fine scale modelling <strong>and</strong> thedevelopment <strong>of</strong> parametrization schemes; ECMWF: London, 1992;pp 21-51.(115) Menzel, T. Die Reynolds-Schubspannung als wesentlicherParameter zur Modellierung der Strömungssstruktur in Blasensäulen<strong>and</strong> Airlift-Schlaufenreaktoren; VDI Verlag: Düsseldorf,Germany, 1990; VDI Reihe 3, Nr. 198.(116) Schlichting, H.; Gersten, K. Boundary-Layer Theory;Springer-Verlag: Berlin, 2000.(117) Telionis, D. P. Unsteady Viscous Flows; Springer-Verlag: New York, 1981.(118) Johanson, S. H.; Davidson, L.; Olsson, E. Numericalsimulation <strong>of</strong> vortex shedding past triangular cylinders at highReynolds number using a k-ɛ turbulence model. Int. J. Numer.Methods Fluids 1993, 16, 859.(119) Williams, F. A. Combustion Theory, 2nd ed.; Addison-Wesley Publishing Company: Redwood City, CA, 1985.(120) Gidaspow, D. Multiphase Flow <strong>and</strong> Fluidization-Continuum<strong>and</strong> Kinetic Theory Descriptions; Academic Press, HarcourtBrace & Company, Publishers: Boston, MA, 1994.(121) Friedl<strong>and</strong>er, S. K. Smoke, Dust, <strong>and</strong> Haze: Fundamentals<strong>of</strong> Aerosol Dynamics, 2nd ed.; Oxford University Press: New York,2000.(122) Hulburt, H. M.; Katz, S. Some problems in particletechnology: A statistical mechanical formulation. Chem. Eng. Sci.1964, 19, 555.(123) Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. TransportPhenomena; John Wiley & Sons: New York, 1960.(124) R<strong>and</strong>olph, A. D. A Population Balance for CountableEntities. Can. J. Chem. Eng. 1964, 42, 280.(125) R<strong>and</strong>olph, A. D.; Larson, M. A. Theory <strong>of</strong> ParticulateProcesses: Analysis <strong>and</strong> Techniques <strong>of</strong> Continuous Crystallization,2nd ed.; Academic Press, Harcourt Brace Jovanovich, Publishers:San Diego, CA, 1988.(126) Ramkrishna, D. The Status <strong>of</strong> Population Balances. Rev.Chem. Eng. 1985, 3, 49.(127) Ramkrishna, D. Population Balances: Theory <strong>and</strong> Applicationsto Particulate Systems in Engineering; AcademicPress: San Diego, CA, 2000.(128) Williams, M. M. R.; Loyalka, S. K. Aerosol Science Theory<strong>and</strong> Practice: With Special Applications to the Nuclear Industry;Pergamon Press: Oxford, U.K., 1991.(129) Kolev, N. I. Fragmentation <strong>and</strong> Coalescence Dynamicsin Multiphase Flows. Exp. Therm. Fluid Sci. 1993, 6, 211.(130) Kolev, N. I. Multiphase Flow Dynamics 2: Mechanical<strong>and</strong> Thermal Interactions; Springer: Berlin, Germany, 2002.(131) Guido, Lavalle G.; Carrica, P. M.; Clausse, A.; Qazi, M.K. A bubble number density constitutive equation. Nucl. Eng. Des.1994, 152, 213.(132) Pilon, L.; Fedorov, A. G.; Ramkrishna, D.; Viskanta, R.<strong>Bubble</strong> transport in three-dimensional laminar gravity-drivenflowsmathematical formulation. J. Non-Cryst. Solids 2004, 336,71.(133) Prasher, C. L. Crushing <strong>and</strong> Grinding Process H<strong>and</strong>book;John Wiley & Sons Limited; Chichester, U.K., 1987.(134) Dorao, C.; Jakobsen, H. A. The Kernel Effect in thePopulation Balance Equation. Poster Presented at CHISA 2004,Prague, Czech Republic, Aug 22-26, 2004.(135) Coulaloglou, C. A.; Tavlarides, L. L. Description <strong>of</strong>Interaction Processes in Agitated Liquid-Liquid Dispersions.Chem. Eng. Sci. 1977, 32, 1289.(136) Lee, C.-K.; Erickson, L. E.; Glasgow, L. A. <strong>Bubble</strong>Breakup <strong>and</strong> Coalescence in Turbulent Gas-Liquid Dispersions.Chem. Eng. Commun. 1987, 59, 65.(137) Wang, T.; Wang, J.; Jin, Y. A novel theoretical breakupkernel function for bubbles/droplets in a turbulent flow. Chem.Eng. Sci. 2003, 58, 4629.(138) Venneker, B. C. H.; Derksen, J. J.; van den Akker, H. E.A. Population Balance <strong>Modeling</strong> <strong>of</strong> Aerated Stirred Vessels Basedon CFD. AIChE J. 2002, 48, 673.(139) Fleischer, C.; Bierdel, M.; Eigenberger, G. Prediction <strong>of</strong><strong>Bubble</strong> Size Distributions in G/L-Contactors with PopulationBalances. Poster Presented at the Third German/Japanese Symposiumon <strong>Bubble</strong> <strong>Column</strong>s, Schwerte, Germany, Jun 13-15, 1994.(140) Kamp, A. M.; Chesters, A. K.; Colin, C.; Fabre, J. <strong>Bubble</strong>coalescence in turbulent flows: A mechanistic model for turbulenceinducedcoalescence applied to microgravity bubbly pipe flow. Int.J. Multiphase Flow 2001, 27, 1363.


Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5151(141) Smoluchowski, M. Versuch einer mathematischen theorieder koagulationskinetik kolloider lösungen. Z. Physik Chem. 1918,XCII, 129.(142) Kolmogorov, A. N. Local Structure <strong>of</strong> Turbulence inIncompressible Viscous Fluid for Very Large Reynolds Number.Dokl. Akad. Nauk. SSSR 1941, 30, 229.(143) Pope, S. B. Turbulent Flows; Cambridge UniversityPress: Cambridge, U.K., 2001.(144) Batchelor, G. K. The Theory <strong>of</strong> Homogeneous Turbulence;Cambridge University Press: Cambridge, U.K., 1982.(145) Risso, F.; Fabre, J. Oscillations <strong>and</strong> breakup <strong>of</strong> a bubbleimmersed in a turbulent field. J. Fluid. Mech. 1998, 372, 323.(146) Hagesaether, L. Coalescence <strong>and</strong> Break-up <strong>of</strong> Drops <strong>and</strong><strong>Bubble</strong>s; Dr. Ing. Thesis, Department <strong>of</strong> Chemical Engineering,The Norwegian University <strong>of</strong> Science <strong>and</strong> Technology, Trondheim,Norway, 2002.(147) Oolman, T. O.; Blanch, H. W. <strong>Bubble</strong> Coalescence inStagnant Liquid. Chem. Eng. Commun. 1986, 43, 237.(148) Chester, A. K. The modelling <strong>of</strong> coalescence processes influid-fluid dispersions: A review <strong>of</strong> current underst<strong>and</strong>ing. Trans.Inst. Chem. Eng. 1991, 69, 259.(149) Kuboi, R.; Komasawa, I.; Otake, T. Behavior <strong>of</strong> dispersedparticles in turbulent liquid flow. J. Chem. Eng. Jpn. 1972, 5, 349.(150) Kuboi, R.; Komasawa, I.; Otake, T. Collision <strong>and</strong> coalescence<strong>of</strong> dispersed drops in turbulent liquid flow. J. Chem. Eng.Jpn. 1972, 5, 423.(151) Brenn, G.; Braeske, H.; Durst, F. Investigation <strong>of</strong> theunsteady two-phase flow with small bubbles in a model bubblecolumn using phase-Doppler anemometry. Chem. Eng. Sci. 2002,57, 5143.(152) Klaseboer, E.; Chevallier, J. Ph.; Masbernat, O.; Gourdon,C. Drainage <strong>of</strong> the liquid film between drops colliding at constantapproach velocity. Paper Presented at the Third InternationalConference on Multiphase Flow, ICMF98, Lyon, France, Jun 8-12,1998.(153) Klaseboer, E.; Chevallier, J. Ph.; Gourdon, C.; Masbernat,O. Film drainage between colliding drops at constant approachvelocity: Experimental <strong>and</strong> modeling. J. Colloid Interface Sci.2000, 229, 274.(154) Orme, M. Experiments on Droplet Collisions, Bounce,Coalescence <strong>and</strong> Disruption. Prog. Energy Combust. Sci. 1997, 23,65.(155) Havelka, P.; Gotaas, C.; Jakobsen, H. A.; Svendsen, H.F. Droplet Formation <strong>and</strong> Interactions under Normal <strong>and</strong> Highpressures. Paper Presented at the 5th International Conferenceon Multiphase Flow, ICMF′04, Yokohama, Japan, May 30-Jun4, 2004.(156) Saboni, A.; Gourdon, C.; Chesters, A. K. The influence <strong>of</strong>inter-phase mass transfer on the drainage <strong>of</strong> partially mobile liquidfilms between drops undergoing a constant interaction force.Chem. Eng. Sci. 1999, 54, 461.(157) Saboni, A.; Alex<strong>and</strong>rova, S.; Gourdon, C.; Chesters, A. K.Inter-drop coalescence with mass transfer: Comparison <strong>of</strong> theapproximate drainage models with numerical results. Chem. Eng.J. 2002, 88, 127.(158) Lasheras, J. C.; Estwood, C.; Martínez-Bazán, C.; Montañés,J. L. A review <strong>of</strong> statistical models for the break-up <strong>of</strong> animmiscible fluid immersed into a fully developed turbulent flow.Int. J. Multiphase Flow 2002, 28, 247.(159) Hinze, J. O. Fundamentals <strong>of</strong> the Hydrodynamic Mechanism<strong>of</strong> Splitting in Dispersion Processes. AIChE J. 1955, 1, 289.(160) Drazin, P. G.; Reid, W. H. Hydrodynamic Stability;Cambridge Univerity Press: Cambridge, U.K., 1981.(161) Kolmogorov, A. N. On the Breakage <strong>of</strong> Drops in aTurbulent Flow. Dokl. Akad. Navk. SSSR 1949, 66, 825.(162) Politano, M. S.; Carrica, P. M.; Baliño, J. L. About bubblebreakup models to predict bubble size distributions in homogeneousflows. Chem. Eng. Commun. 2003, 190, 299.(163) Azbel, D. Two-Phase Flows in Chemical Engineering;Cambridge Univerity Press: Cambridge, U.K., 1981.(164) Azbel, D.; Athanasios, I. L. A Mechanism <strong>of</strong> LiquidEntrainment. In H<strong>and</strong>book <strong>of</strong> Fluids in Motion; Cheremisin<strong>of</strong>f,N., Ed.; Ann Arbor Science Publishers: Ann Arbor, MI, 1983; p473.(165) Martínez-Bazán, C.; Montañés, J. L.; Lasheras, J. C. Onthe breakup <strong>of</strong> an air bubble injected into a fully developedturbulent flow. Part 1. Breakup frequency. J. Fluid. Mech. 1999,401, 157.(166) Batchelor, G. K. The Theory <strong>of</strong> Homogeneous Turbulence;Cambridge University Press: Cambridge, U.K., 1956.(167) Batchelor, G. K. The Theory <strong>of</strong> Homogeneous Turbulence;Cambridge University Press: Cambridge, U.K., 1953.(168) Angelidou, C.; Psimopoulos, M.; Jameson, G. J. SizeDistribution Functions <strong>of</strong> Dispersions. Chem. Eng. Sci. 1979, 34,671.(169) Tsouris, C.; Tavlarides, L. L. Breakage <strong>and</strong> CoalescenceModels for Drops in Turbulent Dispersions. AIChE J. 1994, 40,395.(170) Soo, S. L. Particulates <strong>and</strong> Continuum; HemispherePublishing Corporation: New York, 1989.(171) Marrucci, G. A Theory <strong>of</strong> Coalescence. Chem. Eng. Sci.1969, 24, 975.(172) Thomas, R. M. <strong>Bubble</strong> Coalescence in Turbulent Flows.Int. J. Multiphase Flow 1981, 7, 709.(173) Smoluchowski, M. Drei vortrage uber diffusion, Brownschemolekularbewegung und koagulation von kolloidteilchen.Phys. Z. 1916, 17, 557.(174) Smoluchowski, M. Versuch einer mathematischen theorieder koagulationskinetik kolloider lösungen. Z. Phys. Chem. 1917,92, 129.(175) Ross, S. L.; Curl, R. L. Measurement <strong>and</strong> models <strong>of</strong> thedispersed phase mixing process. In Joint Chemical EngineeringConference; Symposium Series 139; AIChE Press: New York, 1973;Paper 29b.(176) Martínez-Bazán, C.; Montañés, J. L.; Lasheras, J. C. Onthe breakup <strong>of</strong> an air bubble injected into a fully developedturbulent flow. Part 2. Size PDF <strong>of</strong> the resulting daughter bubbles.J. Fluid. Mech. 1999, 401, 183.(177) Rodríguez-Rodríguez, J.; Martínez-Bazán, C.; Montañes,J. L. A novel particle tracking <strong>and</strong> break-up detection algorithm:application to the turbulent break-up <strong>of</strong> bubbles. Meas. Sci.Technol. 2003, 14, 1328.(178) Eastwood, C. D.; Armi, L.; Lasheras, J. C. The breakup<strong>of</strong> immersible fluids in turbulent flows. J. Fluid Mech. 2004, 502,309.(179) Hesketh, R. P.; Etchells, A. W.; Russell, T. W. F. Experimentalobservations <strong>of</strong> bubble breakage in turbulent flows. Ind.Eng. Chem. Res. 1991, 30, 845.(180) Hesketh, R. P.; Etchells, A. W.; Russell, T. W. F. <strong>Bubble</strong>breakage in pipeline flows. Chem. Eng. Sci. 1991, 46, 1.(181) Valentas, K. J.; Amundson, N. R. Breakage <strong>and</strong> Coalescencein Dispersed Phase Systems. Ind. Eng. Chem. Fundam.1966, 5, 271.(182) Valentas, K. J.; Amundson, N. R. Breakage <strong>and</strong> Coalescencein Dispersed Phase Systems. Ind. Eng. Chem. Fundam.1966, 5, 533.(183) Konno, M.; Aoki, M.; Saito, S. Simulations model forbreak-up process in an agitated tank. J. Chem. Eng. Jpn. 1980,13, 67.(184) Hounslow, M. J.; Ryall, R. L.; Marshall, V. R. A DiscretizedPopulation Balance for Nucleation, Growth, <strong>and</strong> Aggregation.AIChE J. 1988, 34, 1821.(185) McGraw, R. Description <strong>of</strong> Aerosol Dynamics by theQuadrature Method <strong>of</strong> Moments. Aerosol Sci. Technol. 1997, 27,255.(186) Marchisio, D. L.; Vigil, R. D.; Fox, R. O. Quadraturemethod <strong>of</strong> moments for aggregation-breakage processes. J. ColloidInterface Sci. 2003, 258, 322.(187) Marchisio, D. L.; Vigil, R. D.; Fox, R. O. Implementation<strong>of</strong> the quadrature method <strong>of</strong> moments in CFD codes for aggregation-breakageproblems. Chem. Eng. Sci. 2003, 58, 3337.(188) Press: W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery,B. P. Numerical Recipes in Fortran 77. The Art <strong>of</strong> ScientificComputing, 2nd ed.; Cambridge University Press: Cambridge,U.K., 1992; Vol. 1.Received for review June 24, 2004Revised manuscript received September 16, 2004Accepted September 30, 2004IE049447X

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!