13
51-1 Efficient Numerical Calculation of Evaporating Sprays in Combustion Chamber Flows R. Schmehl, G. Klose, G. Maier and S. Wittig Lehrstuhl und Institut für Thermische Strömungsmaschinen Universität Karlsruhe (T.H.) 76128 Karlsruhe, Germany Summary Representing two different conceptual approaches, either Eule- rian continuum models or Lagrangian particle models are com- monly applied for the numerical description of dispersed two phase flows. Taking advantage of the positive features inherent to each model, a combination approach is presented in this study for the efficient computation of liquid fuel sprays in combustor flows. In the preconditioning stage, Eulerian transport equa- tions for gas phase and droplet phase are solved simultaneously in a block-iterative scheme based on a coarse discretization of spray boundary conditions at the nozzle. Due to the close cou- pling of both phases, the time expense of this approximate flow field computation is not much higher as for single phase flows. In the refinement stage, Lagrangian droplet tracking is applied with a detailed discretization of initial conditions. To account for complete interaction between gas phase and droplets, gas flow solution and droplet tracking are concatenated by an iter- ative procedure. In this stage, the numerical description of the spray is enhanced by additional modeling of droplet breakup. Results of numerical simulations are compared with measure- ments of the two phase flow in a premix duct of a LPP research combustor. 1 Notation cp specific heat capacity D droplet diameter D0.5 mass median diameter D32 Sauter mean diameter D0.632 characteristic diameter f body force h enthalpy ˙ H enthalpy flux H energy transfer rate I momentum transfer rate k turbulent kinetic energy ˙ m mass flux M mass transfer rate On Ohnesorge number P pressure Pr Prandtl number ˙ Q conductive heat flux Re Reynolds number S source term Sc Schmidt number T temperature Tu degree of turbulence U velocity component We Weber number Y mass fraction Greek Symbols α heat transfer coefficient α k liquid volume fraction β off axis angle ε dissipation rate of k Γ diffusion coefficient μ dynamic viscosity ν kinematic viscosity ρ density τ shear stress Subscripts 0 initial state g,d gas, droplet int interface t turbulent vap vapor 2 Introduction Improving modern gas turbine efficiencies by increasing pres- sure and temperature levels of the combustion process, essen- tially requires sophisticated combustion concepts in order to meet todays strict limitations on pollutant emissions. Funda- mental to these low emissions concepts is a characteristic strat- egy to inject and mix the liquid fuel with the compressed air flow, avoiding local stoichiometric combustion conditions as far as possible. Two promising approaches in this context are the concepts of Lean-Premix-Prevaporize (LPP) and Rich-Quench- Lean (RQL) combustion. In order to develop advanced com- bustor designs with the required flow characteristics, a better understanding of the two phase flow physics is necessary. Two phase flow effects typical for premix ducts of LPP combustors or prefilming air blast atomizers are summarized in Fig. 1. Evaporation Dispersion + Droplet Breakup Spray-wall Interaction Wall Film Flow Atomization Figure 1: Two phase flow effects in a LPP premix duct Due to the enormous increase in computing performance, Com- putational Fluid Dynamics (CFD) offers a promising potential for efficient combustor design and optimization. In particular when compared to experimental studies at elevated pressures, CFD analysis may be employed to reduce turn-around times and costs of combustor design significantly. On the other hand, complex flow phenomena such as turbulence, atomization or chemical reaction still represent some of the most challenging topics for CFD tools. Basically, two different conceptual approaches may be em- ployed for the numerical description of dispersed two phase flows [3]. In analogy to single phase gas flow, the Eulerian ap- proach is based on a continuum model of the spray, resulting in transport equations describing the propagation and evaporation of this droplet phase [28], [6]. In the Lagrangian approach, the spray is modeled by superposition of trajectories calculated for large numbers of representative droplets. Each of the two basic approaches is characterized by specific advantages and restric- tions. In the Eulerian method, the transport equations of the droplet phase are appended to the gas phase transport equations, result- ing in a compact description of the interacting two phase flow system. The essential advantage is a simultaneous solution of the interacting flow fields of gas phase and spray by a single numerical method. Applying a standard block-iterative solver for systems of linearized equations, the information exchange

Efficient Numerical Calculation of Evaporating Sprays in ...rschmehl.home.xs4all.nl/CV/publications/agard98_paper.pdf · spray is enhanced by additional modeling of droplet breakup

  • Upload
    others

  • View
    10

  • Download
    0

Embed Size (px)

Citation preview

  • 51-1

    Efficient Numerical Calculation of Evaporating Spraysin Combustion Chamber Flows

    R. Schmehl, G. Klose, G. Maier and S. WittigLehrstuhl und Institut für Thermische Strömungsmaschinen

    Universität Karlsruhe (T.H.)76128 Karlsruhe, Germany

    SummaryRepresenting two different conceptual approaches, either Eule-rian continuum models or Lagrangian particle models are com-monly applied for the numerical description of dispersed twophase flows. Taking advantage of the positive features inherentto each model, a combination approach is presented in this studyfor the efficient computation of liquid fuel sprays in combustorflows. In the preconditioning stage, Eulerian transport equa-tions for gas phase and droplet phase are solved simultaneouslyin a block-iterative scheme based on a coarse discretization ofspray boundary conditions at the nozzle. Due to the close cou-pling of both phases, the time expense of this approximate flowfield computation is not much higher as for single phase flows.In the refinement stage, Lagrangian droplet tracking is appliedwith a detailed discretization of initial conditions. To accountfor complete interaction between gas phase and droplets, gasflow solution and droplet tracking are concatenated by an iter-ative procedure. In this stage, the numerical description of thespray is enhanced by additional modeling of droplet breakup.Results of numerical simulations are compared with measure-ments of the two phase flow in a premix duct of a LPP researchcombustor.

    1 Notation

    cp specific heat capacityD droplet diameterD0.5 mass median diameterD32 Sauter mean diameterD0.632 characteristic diameterf body forceh enthalpyḢ enthalpy fluxH energy transfer rateI momentum transfer ratek turbulent kinetic energyṁ mass fluxM mass transfer rateOn Ohnesorge numberP pressurePr Prandtl numberQ̇ conductive heat fluxRe Reynolds numberS source termSc Schmidt numberT temperatureTu degree of turbulence

    U velocity componentWe Weber numberY mass fraction

    Greek Symbolsα heat transfer coefficientαk liquid volume fractionβ off axis angleε dissipation rate of kΓ diffusion coefficientµ dynamic viscosityν kinematic viscosityρ densityτ shear stress

    Subscripts0 initial stateg, d gas, dropletint interfacet turbulentvap vapor

    2 IntroductionImproving modern gas turbine efficiencies by increasing pres-sure and temperature levels of the combustion process, essen-

    tially requires sophisticated combustion concepts in order tomeet todays strict limitations on pollutant emissions. Funda-mental to these low emissions concepts is a characteristic strat-egy to inject and mix the liquid fuel with the compressed airflow, avoiding local stoichiometric combustion conditions as faras possible. Two promising approaches in this context are theconcepts of Lean-Premix-Prevaporize (LPP) and Rich-Quench-Lean (RQL) combustion. In order to develop advanced com-bustor designs with the required flow characteristics, a betterunderstanding of the two phase flow physics is necessary. Twophase flow effects typical for premix ducts of LPP combustorsor prefilming air blast atomizers are summarized in Fig. 1.

    EvaporationDispersion +Droplet

    BreakupSpray-wallInteraction

    Wall FilmFlowAtomization

    Figure 1: Two phase flow effects in a LPP premix duct

    Due to the enormous increase in computing performance, Com-putational Fluid Dynamics (CFD) offers a promising potentialfor efficient combustor design and optimization. In particularwhen compared to experimental studies at elevated pressures,CFD analysis may be employed to reduce turn-around timesand costs of combustor design significantly. On the other hand,complex flow phenomena such as turbulence, atomization orchemical reaction still represent some of the most challengingtopics for CFD tools.Basically, two different conceptual approaches may be em-ployed for the numerical description of dispersed two phaseflows [3]. In analogy to single phase gas flow, the Eulerian ap-proach is based on a continuum model of the spray, resulting intransport equations describing the propagation and evaporationof this droplet phase [28], [6]. In the Lagrangian approach, thespray is modeled by superposition of trajectories calculated forlarge numbers of representative droplets. Each of the two basicapproaches is characterized by specific advantages and restric-tions.In the Eulerian method, the transport equations of the dropletphase are appended to the gas phase transport equations, result-ing in a compact description of the interacting two phase flowsystem. The essential advantage is a simultaneous solution ofthe interacting flow fields of gas phase and spray by a singlenumerical method. Applying a standard block-iterative solverfor systems of linearized equations, the information exchange

  • 51-2

    between phases is realized on the level of the non-linear itera-tions. As a consequence, computation times are generally smallcompared to the Lagrangian approach. However, each dropletinitial condition to be simulated requires the solution of an in-dividual set of 6 transport equations. This conceptual feature isa severe limitation for the discretization of complex sprays withwide ranges of initial droplet size, injection angles and veloci-ties. Furthermore, Eulerian methods are generally not suited forthe modeling of complex two phase phenomena like secondaryatomization or spray-wall interaction.In the Lagrangian method, a large number of droplet trajecto-ries has to be tracked to achieve a continuous distribution ofthe liquid phase. Due to the stochastic simulation of turbulentspray dispersion by random sampling of gas velocity fluctua-tions, identical droplet initial conditions lead to different trajec-tories. As a consequence, the liquid phase flow field is a statisti-cal quantity. To limit the maximum field deviations, the numberof simulated droplet trajectories has to be increased to values upto 104 - 106 for typical combustor flows. Effects of the sprayback on the gas flow such as aerodynamic dragging or evapo-ration cooling are recorded in droplet source terms, describingthe local interfacial transfer of mass, momentum and energy.Including these source terms in the gas flow transport equa-tions, complete phase interaction is taken into account by aniterative concatenation of gas flow computation and Lagrangiandroplet tracking. Besides increased computation times, this it-erative procedure entails an artificial decoupling of gas flow andspray. In particular for flows with intense phase interaction, thismay cause severe convergence problems, requiring strong relax-ation of source term fields. Nevertheless, Lagrangian methodsare commonly preferred for practical CFD analysis due to sig-nificant advantages regarding complex spray discretization andmodeling of flow phenomena such as secondary droplet breakupor droplet-wall interaction.In this study, a new hybrid approach is presented combiningEulerian and Lagrangian methods to take advantage of the ca-pabilities inherent to both methods. In the first stage of thisprocedure, the Eulerian method is used for an efficient compu-tation of an approximate two phase flow field. A coarse dis-cretization of spray boundary conditions at the nozzle limitsthe size of the system of transport equations to a practical di-mension. Good convergence rates are achieved due to the closecoupling between gas flow and spray. In the refinement stage,iterative cycles of single phase gas flow computation and sub-sequent droplet tracking are employed to improve the qualityof the preconditioned two phase flow field. Taking advantageof the stochastical nature of the tracking approach, a fine dis-cretization of polydisperse sprays is achieved by random sam-pling of droplet initial conditions at the nozzle. The numeri-cal description of the spray is enhanced by optional modelingof secondary droplet breakup and spray-wall interaction. Sincemodeling of spray-wall interaction has been described in detailin Ref. [24], only secondary atomization of droplets is consid-ered in this study.To demonstrate the performance and accuracy of the numeri-cal methods discussed in this paper, the evaporating spray inthe premix duct of a LPP research combustor is simulated andassessed by measured droplet data. The experimental investiga-tion of this premix section has been a focus of various studies[15], [16], [13] and represents a valuable source of experimen-tal data.

    3 Eulerian approachThe Eulerian approach for the numerical description of dis-persed two phase flows is based on the assumption that the liq-

    uid phase represents an additional continuum penetrating thegas phase. In analogy to the continuum approach of singlephase flows, each phase is described by a set of transport equa-tions for mass, momentum and energy extended by interfacialexchange terms. This set of transport equations can be recastedinto a universal formulation which is discretized by a conser-vative Finite Volume method and solved by a block-iterativescheme for systems of linearized equations.

    3.1 Transport equations of the two phase flow

    3.1.1 Basic equations

    Except for the near region of the atomizer, the volume fractionof fuel in the flow field is low. In this dilute two phase flowregime, interactions between fuel fragments can be neglected.Starting from the basic Navier-Stokes equations, the instanta-neous transport equations for gas and droplet phase are derivedeither by spatial, temporal [9] or ensemble phase averaging.

    Gas phase:

    ∂tαgρg +

    ∂xjαgρgUg,j = Mint,g (1)

    ∂tαgρgUg,i +

    ∂xjαgρgUg,jUg,i =

    −αg ∂∂xi

    Pg +∂τi,j∂xj

    + αgρgfi + Iint,g,i (2)

    ∂tαgρghg +

    ∂xjαgρgUg,jhg =

    − ∂q̇j∂xj

    + Sh,g + Hint,g (3)

    Droplet phase:

    ∂tαdρd +

    ∂xjαdρdUd,j = Mint,d (4)

    ∂tαdρdUd,i +

    ∂xjαdρdUd,jUd,i =

    −αd∂

    ∂xiPg + αdρdfi + Iint,d,i (5)

    ∂tαdρdhd +

    ∂xjαdρdUd,jhd = Hint,d (6)

    The weighting factors αg and αd are a result of the averag-ing process and represent the local volume fractions of gas andliquid phases related by the following equation

    αg + αd = 1 (7)

    Dilute two phase flows are characterized by the conditions

    αd � 1 ; αg ≈ 1 (8)

    In this flow regime, the transport equations of the gas phaseapproach the standard single phase transport equations ex-tended by additional interfacial exchange terms Mint,g, Iint,gand Hint,g .

    3.1.2 Interfacial exchange terms

    The interfacial exchange terms describe the local rates of mass,momentum and energy transfer across the liquid-gas interface.Assuming a spherical shape of the droplets and a uniform inter-nal temperature distribution, the transfer rates may be estimatedfrom Lagrangian single droplet physics. The following modelexpressions are derived from Eq. (23) and the heat and mass

  • 51-3

    fluxes (28), (29) and (30) of an isolated droplet

    Mint,d = −Mint,g =6αdπD3

    ṁ∗vap (9)

    Iint,d,i = −Iint,g,i =6αdπD3

    �π

    8D2ρgCD |Ug − Ud|

    (Ug,i − Ud,i) + ṁ∗vap Ud,i � (10)Hint,d = −Hint,g =

    6αdπD3

    �Ḣ∗vap,s + Q̇

    cond,s � (11)3.1.3 Transport equation of the droplet diameter

    According to Ref. [6], a transport equation of the mean dropletdiameter can be established by combining transport equationsof droplet number and droplet mass (4), giving

    ∂tαdρdD +

    ∂xjαdρdUd,jD =

    8 αdπD2

    ṁvap (12)

    For non evaporating sprays this equation is identical to the con-tinuity equation.

    3.1.4 Turbulence modeling

    The transport equations derived so far are suited for the numer-ical description of sprays in laminar gas flows. Since combus-tors generally operate in the turbulent flow regime, the systemof transport equations (1) - (6), (12) is extended by introduc-ing turbulent fluctuations of the transport quantities followedby Reynolds averaging of the equations. With respect to the gasphase, the standard k-ε model is employed to model the trans-port terms resulting from correlations of fluctuating quantities.This procedure has been described in detail by several authors[12], [21]. The turbulence terms in the droplet phase trans-port equations are approximated by an algebraic model whichis based on a Bousinesq approach.

    ∂tαdρd +

    ∂xjαdρdUd,j =

    ∂xj

    �µt,dSct,d

    αd∂xj � + Mint,d (13)

    ∂tαdρdUd,i +

    ∂xjαdρdUd,jUd,i =

    ∂xj � αdµt,d � ∂Ud,i∂xj + ∂Ud,j∂xi ��� +αdρdfi − αd

    ∂xiPg + Iint,d,i (14)

    ∂tαdρdhd +

    ∂xjαdρdUd,jhd =

    ∂xj

    �αd

    µt,dPrt,d

    ∂hd∂xj � + Hint,d (15)

    ∂tαdρdD +

    ∂xjαdρdUd,jD =

    8 αdπD2

    ṁvap +

    ∂xj

    �αd

    µt,dPrt,d

    ∂D

    ∂xj � (16)Double and triple correlations involving fluctuations of αd areneglected on the right hand side of Eq. (14). A value of 0.9for the turbulent Schmidt and Prandtl numbers, Sct,d, Prt,d ischosen for the present calculations. Using this value, Eq. (13)effectively is a transport equation of a scalar in the asymptoticalcase of a vanishing droplet diameter. A fundamental assumptionof this approach is the dependence of the turbulent viscosity ofthe droplet phase µt,d on local mean flow properties [17], [10]

    νt,dνt,g

    =µt,dµt,g

    ρgρd

    =1

    1 +

    �tdtg � 2 n (17)

    Thus, the ratio of the kinematic viscosities of droplet and gasphase is postulated to be a function of the characteristic timescales of both phases. In Ref. [10], a value of 0.25 for theempirical parameter n is suggested. The time scale td which isdenoted as droplet relaxation time characterizes the ability of adroplet to follow turbulent gas velocity fluctuations:

    td =4

    3

    ρdρg

    D2

    CD Red νg. (18)

    Originally, the characteristic time of the gas flow turbulence tgis taken to be the dissipation time scale given by Eq. 25. Inthis formulation, the droplet phase turbulence model fails todescribe the crossing trajectory effect which has a significantinfluence on turbulent droplet dispersion [27]. According tothe turbulence modeling of the Lagrangian approach, the ex-tended version of the model considers a second characteristictime scale. This crossing time tc is the time required by adroplet to cross the current coherent turbulence structure whichis estimated from Eq. 26. Combining both time scales, the gasphase time scale is now defined as

    tg = min[te, tc] (19)

    The validation of this enhanced turbulence model is based onthe basic experiment described in Ref. [27].

    3.2 Discretization of polydisperse spraysTo complete the numerical description of the spray, boundaryconditions of the droplet phase have to be specified at the atom-izer nozzle. However, most sprays of technical importance arecharacterized by a broad variety of initial droplet diameters andvelocities. Since each individual droplet phase boundary condi-tion theoretically requires the numerical solution of a separateset of transport equations (13) - (16), a polydisperse spray hasto be discretized by a limited number of representative dropletclasses. In practice, the computational effort increases at leastlinearly with the number of droplet classes employed. As a con-sequence, the maximum number of classes is restricted by theCPU time and memory capacity available.For dilute sprays, direct interaction between droplet classes canbe neglected although each class is coupled to the gas phase bythe closure equation

    Mint,g = −nc�

    k=1

    Mint,d,k (20)

    Iint,g = −nc�

    k=1

    Iint,d,k (21)

    Hint,g = −nc�

    k=1

    Hint,d,k (22)

    4 Lagrangian approach4.1 Spray dispersionThe Lagrangian simulation of dispersed two phase flow is basedon the tracking of statistically significant droplet parcels in thegas flow. Each parcel is represented by one droplet and is deter-mined by discretization of the continuous spectra of droplet ini-tial conditions in the near field of the atomizer. The tracking isbased on the integration of the droplets equation of motion com-bined with an empirical correlation for the aerodynamic dragcoefficient CD,

    d~uddt

    = −34

    ρgρd

    CDD

    |~ud − ~ug| (~ud − ~ug) (23)

    CD = 0.36 + 5.48Re−0.573d +

    24

    Red(24)

  • 51-4

    In order to simulate the effect of turbulent spray dispersion, theturbulence structure of the gas flow field is modeled by a ran-dom process along the droplet trajectories [5], [18]. In this con-cept, the local turbulence structure is characterized by the lengthscale le and dissipation time scale te of eddies representing thecoherent flow structures

    le = C1

    k3

    2

    ε, (25)

    In addition to the life time scale te, a crossing time scale tc iscalculated from ���� ���� � tc

    t0

    (~ug − ~ud)dt���� ���� = le, (26)

    taking into account the droplet dynamics. Each time the smallerone of both time scales is elapsed, the droplet enters a neweddy. Consequently, the random process generates a new ve-locity fluctuation ~u′g from a Gaussian distribution which is de-termined by

    µ = 0 , σ = � 23k. (27)

    This velocity fluctuation remains constant for the period ofdroplet-eddy interaction and is added to the local value of thegas flow velocity.

    4.2 Spray evaporationIn this study, droplet evaporation is simulated by means of theUniform Temperature model [4], [26], [2]. This computation-ally effective droplet model is based on the assumption of ahomogeneous internal temperature distribution in the dropletand phase equilibrium conditions at the surface. The analyti-cal derivation of this model does not consider contributions toheat and mass transport by forced convection by the gas flowaround the droplet. Since diffusive time scales in the surround-ing gas phase are much smaller than in the droplet fluid, a quasistationary description of the gas phase is applied. Using refer-ence values for variable fluid properties (1/3-rule), an integra-tion of the radially symmetric differential equations yields an-alytical expressions for the transport fluxes ṁvap, Q̇cond,s andḢvap,s. At this point, convective transport is taken into accountby two empirical factors [1] resulting in the corrected fluxesṁ∗vap, Q̇

    cond,s and Ḣ∗

    vap,s [23]

    ṁ∗vap = cfm ṁvap, (28)

    Q̇∗cond,s = πD2 α∗(Td − Tg), (29)

    Ḣ∗vap,s = ṁ∗

    vap cp,vap,ref (Td − Tg). (30)Vapor mass flux and heat transfer coefficient are calculated asfollows

    ṁvap = 2πD ρg,refΓim,ref ln1 − Yvap,g1 − Yvap,s

    , (31)

    α∗ = cfh

    ṁvap cp,vap,refπD2

    exp

    �ṁvap cp,vap,ref

    2πD λg,ref � − 1 , (32)cfm = 1 + 0.276 Re

    1

    2 Sc1

    3 , (33)

    cfh = 1 + 0.276 Re1

    2 Pr1

    3 . (34)

    The balance equations of the droplet reduce to ordinary differ-ential equations,

    d

    dtmd = −ṁ∗vap, (35)

    d

    dt(md hd) = −Q̇∗cond,s − Ḣ∗vap,s, (36)

    which can be appended to the differential equations describingthe droplet motion, Eq. 23 .

    4.3 Secondary droplet breakupAt low relative velocities, the spherical shape of the droplets ispreserved by the dominating effects of surface tension and vis-cous forces in the liquid. With increasing velocities, the desta-bilizing aerodynamic forces on the droplet surface are intensi-fying, resulting in deformation, oscillations and disintegrationof the droplets.

    4.3.1 Classification of breakup mechanisms

    A common practice to classify secondary droplet atomizationprocesses is based on two characteristic groups of parameters,

    We =ρg u

    2rel D

    σd, On =

    µd√ρd D σd

    . (37)

    The Weber number is a measure of the strength of aerodynamicforces relative to surface tension forces, whereas the Ohnesorgenumber assesses the damping effect of viscous friction in thedroplet against surface tension. In the Weber number rangefrom We = 1 up to a critical value We = Wec, non-destructivedroplet deformation and oscillation is observed. As illustratedin Fig. 2, three different mechanisms govern the breakup ofdroplets for increased Weber numbers typical for flows in gasturbine combustors. From these visualizations it is obvious that

    We=70

    ShearBreakup

    We=20

    MultimodeBreakup

    We=10

    BagBreakup

    Figure 2: Breakup mechanisms of water droplets (Ref. [22])

    a common feature of all three mechanisms is an initial deforma-tion of the droplet into a disc shape. After this deformation pe-riod, various complex flow phenomena lead to the final dropletbreakup depending on the intensity of the aerodynamic forces.Exceeding the critical Weber number, the first mechanism ob-served is bag breakup. This process is characterized by the for-mation of a thin hollow bag of droplet fluid stretching from atoroidal rim. The thin film of this bag is eventually burstinginto a cloud of tiny droplets, followed by a disintegration of thetoroidal rim into significantly larger fragments. With increasingaerodynamic forces, a transition to more complex bag struc-tures is observed. In this multimode or stamen breakup regimethe aerodynamic flow interaction is forming an additional fluidfilament in the center of the bag structures which is alignedwith the relative flow velocity. For even higher Weber numbers,shear breakup is observed. This mechanism is fundamentallydifferent to the preceding mechanisms and is characterized by arapidly disintegrating film of fluid continuously stripped off therim of the disc shaped droplet by shear forces.Fig. 3 summarizes the results of various experimental stud-ies [11], [19] in a breakup regime map, indicating the relevantmechanism corresponding to a specific combination of Ohne-sorge and Weber number. For On > 0.1 a significant influenceof viscosity is observed. The transitions between the differentmechanisms are given by the following functions

  • 51-5

    10-3 10-2 10-1 100 101

    On

    10

    20

    30

    40

    50

    We

    Deformation

    BagBreakup

    MultimodeBreakup

    ShearBreakup

    Figure 3: Breakup regime map (◦ : Present flow simulation)

    • Critical Weber number (Transition to bag breakup):

    Wec = 12 (1 + 1.077 On1.6) (38)

    • Transition to multimode breakup:

    We = 20 (1 + 1.200 On1.5) (39)

    • Transition to shear breakup:

    We = 32 (1 + 1.500 On1.4) (40)

    4.3.2 Deformation and breakup times

    Basically, the breakup process can be subdivided in two stages:Initial deformation and further deformation with disintegration.It is convenient to express the relevant times of the breakup pro-cess in terms of the characteristic time of shear breakup

    t∗ =D0urel

    � ρdρg . (41)As stated in Ref. [8], the time of initial droplet deformationtdef has a constant value independent of the specific breakupmechanism

    tdeft∗

    = 1.6. (42)

    However, the breakup time tb measured from begin of defor-mation until final destruction strongly depends on the specificmechanism. A fit to a large number of experimental data isgiven in Ref. [19]

    tbt∗

    =

    ������ �����6 (We − 12)−0.25 12 < We < 18

    2.45 (We − 12)0.25 18 < We < 4514.1 (We − 12)−0.25 45 < We < 351

    0.766 (We − 12)0.25 351 < We < 26705.5 2670 < We.

    (43)

    For On > 1, liquid viscosity is the dominating parameter of thebreakup process resulting in the following correlation

    tbt∗

    = 4.5 (1 + 1.2 On0.74). (44)

    4.3.3 Droplet drag

    Deformation of the droplet prior to breakup leads to a significantincrease of aerodynamic drag. Due to the resulting acceleration,the droplet generally experiences substantial displacement frominitial deformation until final breakup. With respect to the de-formation period, several authors [8] report a linear increase

    of the droplet size from D0 up to a maximum value given by(On < 0.1, We < 100)

    DmaxD0

    = 1 + 0.19√

    We. (45)

    The effect of higher Ohnesorge numbers is taken into accountby using a corrected Weber number in the above and followingequations.

    Wecorr =We

    1 + 1.077 On1.6. (46)

    As suggested in Ref. [8], a linear transition of the drag coeffi-cient from the sphere shape value to the disc shape value is usedin the present study to model the aerodynamic properties of theflattening droplet.In the following period of disintegration, droplet drag dependson the specific breakup mechanism. As illustrated in Fig. 3, thebag and filament structures observed prior to breakup are verycomplex. According to Refs. [11], [22] the toroidal rim evolv-ing in bag breakup is expanding to seven times the initial dropletdiameter, whereas in multimode breakup a maximum diameterof six times the initial diameter is reached (see Fig. 6). Atthis time however, a major part of the droplet cross section con-sists of a thin fluid film accelerated in direction of the relativevelocity thus decreasing the aerodynamic drag. To bypass thedifficulties of describing these opposing effects, the drag of thedisintegrating droplet is calculated from the constant disc statereached at the end of the deformation period. In shear breakup,the size of the disc shaped droplet is continuously decreasing toits final value at tb.

    4.3.4 Secondary droplet sizes

    Based on an extended experimental study covering the com-plete range of breakup mechanisms, a single correlation for theSauter mean diameter D32 has been derived in Ref. [8] for allthree mechanisms (On < 0.1, We < 1000)

    D32D0

    = 6.2 On0.5 We−0.25 (47)

    The exponents in this correlation have been determined by theauthors from an approximate analysis of the droplet internalflow during shear breakup, leaving only a constant factor as aparameter for the fitting to experimental data. Using the Webernumber given by Eq. 46 to account for viscosity effects, addi-tional fitting of the exponents leads to an improved correlation

    D32D0

    = 1.5 On0.2 We−0.25corr . (48)

    This correlation which is used for the flow simulations in thepresent study and the experimental data is shown in Fig. 4.Although the above correlation is valid for the complete range

    0.1 0.2 0.3 0.4On0.2Wecorr

    -0.25

    0.1

    0.2

    0.3

    0.4

    0.5

    D32

    /D0

    WaterGlycerol 42%Glycerol 63%n-HeptaneEthyl Alcohol1.5 On0.2 Wecorr

    -0.25

    Figure 4: Improved correlation for D32 (Data from Ref. [8])

    of breakup conditions, the distribution function of the droplet

  • 51-6

    diameter is substantially different for the various mechanismsof secondary breakup.Bag and multimode breakupConsidering bag or multimode breakup, the volume distributionof the droplet fragments is approximated by a root normal dis-tribution [25], given by the following density function

    f(D) =x

    2√

    2π σ Dexp � −1

    2 � x − µσ � 2 � , (49)and the parameters

    x = � DD0.5

    , µ = 1, σ = 0.238 (50)

    For a volume distribution with this distribution function, themass median diameter D0.5 is related to the Sauter mean di-ameter D32 by

    D0.5D32

    = 1.2 (51)

    Shear breakupAccording to Ref. [8], the volume distribution resulting fromshear breakup is characterized by a bimodal density functionwith a maximum at small diameters and a second maximum atlarge diameters. From the experimental data illustrated in Fig.5, it is concluded that the fine fraction of the droplet fragmentscorresponds to approximately 80% of the total cumulated vol-ume. These droplets result from film fragments stripped off the

    0 1 2D/D0.5

    0

    20

    40

    60

    80

    100

    Cum

    ula

    tive

    Vo

    lum

    e%

    distributiondistribution

    distribution

    ExperimentRoot normal

    Figure 5: Cumulative droplet volume (Data from Ref. [8])

    disc-shaped droplet by shear forces. The 20% in the large di-ameter range not specified in Fig. 5 represent the contributionof the core droplet fragment left by the stripping process. Thediameter Dc of this core droplet is estimated from the criticalWeber number, evaluating Eq. (38) at flow conditions at theinstant of breakup. As illustrated by the curve in Fig. 5, thevolume distribution of the fine fraction of the droplet spectrumafter shear breakup can be approximated by a root normal dis-tribution based on a reduced Sauter mean diameter

    D32,red =4 D32 Dc

    5 Dc − D32. (52)

    In this equation, the Sauter mean diameter of the completedroplet spectrum is evaluated from Eq. 48.

    4.3.5 Secondary droplet velocities

    Due to the dominating influence of aerodynamic forces on smalldroplets, tiny breakup products are immediately dragged withthe gas flow. So, an accurate modeling of initial velocities is notnecessary in general. These considerations apply in particularto the tiny droplet fragments generated by bursting of film bagsor by shear induced film stripping. The motion of large droplets

    in turn is dominated by inertia forces. As a consequence, themodeling of initial velocities of large droplet fragments has asignificant influence on the dispersion behavior of the spray.As a first approximation, the fragments generated by dropletbreakup inherit the velocity of the parent droplet due to momen-tum conservation. In bag or multimode breakup, a transversevelocity component of droplet fragments is observed inducedby the transverse spreading motion of droplet fluid during theexpansion of the toroidal rim. This transverse velocity compo-nent is responsible for increased dispersion of sprays with sec-ondary atomization. For an approximate estimation, the growthvelocity of the rim is determined from time-resolved visualiza-tions of breakup processes. Fig. 6 indicates that the ring has

    0 0.5 1 1.5t/tb

    0

    2

    4

    6

    8

    10

    Dr,

    max

    /D0

    We=20

    We=10We=15

    D

    r,maxD

    r,max

    Figure 6: Growth of the toroidal rim (Water droplet, [22])

    an extension of about seven times the original droplet diameterin bag breakup against six times in multimode breakup. Theseobservations agree with the values reported in Ref. [11]. Basedon these results, the transverse velocity component is estimatedas

    vt =Dr,max − D02(tb − tdef )

    . (53)

    With respect to multimode breakup, a fraction of the dropletfluid is concentrating on the axis of the disintegrating droplet(see Fig. 2, We = 20). Due to the alignment of this prolatefilament with the flow, the fragments of this structure have notransverse velocity component. The volume fraction of the fil-ament is estimated from a Weber number based interpolationbetween the limiting values of 0% for bag breakup and 20%(core droplet) for shear breakup.

    4.3.6 Stochastical simulation of droplet breakup

    The period of droplet disintegration is specified by the char-acteristic times tdef , and tb of the breakup process. In shearbreakup mode, the secondary droplets are continuously gener-ated in the time from tdef to tb, whereas in bag or multimodebreakup mode significant generation of fragments is observedduring the second half of this time period [22]. Instead of focus-ing on a realistic simulation of each individual breakup event,the computational implementation makes use of the Lagrangiantrajectory superposition approach involving large numbers ofdroplet parcels.During droplet deformation, the cross sectional area of thedroplet and the drag coefficient are continuously increased upto their maximum values at tdef . A certain time later, theparent droplet trajectory is terminated and a fixed number ofchild droplets is generated by random sampling of initial con-ditions. Each secondary trajectory is assigned an equal fractionof the volume flux. To limit the number of secondary dropletsto be tracked, only 3 child droplets have been modeled perbreakup event in the present flow simulation in which 10000parent droplets are injected per Lagrangian step. In analogy

  • 51-7

    to the stochastical modeling of fuel atomization and gas flowturbulence effects, the superposition of large numbers of trajec-tories leads to continuous and thus realistic representation ofsecondary atomization.Modeling a single shear breakup event, the time of droplet dis-integration is determined as a random number with uniformdistribution between tdef and tb. To model bag and multi-mode breakup events, the time of disintegration is sampled inthe second half of this interval. The initial size of the childdroplets is determined as a random number with a root nor-mal distribution or from stability criteria (core droplet in shearbreakup). Droplets generated by rim fragmentation are pro-vided with an additional transverse velocity component. In bagbreakup mode, virtually all larger fragments possess a trans-verse velocity component whereas in multimode breakup a vol-ume fraction of up to 20% of the largest fragments is startingwithout dispersive transverse momentum. The limiting value of20% represents the transition to shear breakup and correspondsto the volume fraction of the core droplet. In this breakupregime all secondary droplets inherit the velocity of the parentdroplet since no significant spreading motion is observed.

    1mm

    Formation of Bag

    Breakup

    Injection

    Beginning Deformation

    Figure 7: Breakup of a 60 µm droplet

    Fig. 7 illustrates the computation of a deforming and disin-tegrating droplet in the near field of the nozzle. Compared tothe trajectory of a rigid spherical droplet (dashed line), a sig-nificant deflection of the deforming droplet due to increasedaerodynamic forces is observed. This single droplet computa-tion clearly demonstrates that advanced modeling of secondarybreakup has to take into account the time scales of the breakupprocess since the droplets experience considerable displace-ments before their final disintegration.

    4.4 Iterative solution procedureTracking of a statistically significant number of droplet parcelsand superposition of their trajectories yields a flow field ap-proximation of the dispersed liquid phase. However, a simpleLagrangian two step calculation consisting of a gas flow com-putation and subsequent droplet tracking does not take accountof spray effects on the gas flow such as aerodynamic draggingor evaporation cooling. In particular for evaporating sprays incombustor flows, these effects have a significant influence onthe overall two phase flow field. To establish mutual informa-tion exchange between both phases, Lagrangian two step cyclesare concatenated in an iterative procedure with droplet sourceterms being updated during each tracking step. Representinglocal transfer rates of mass, momentum and energy from sprayto gas flow, droplet source terms are included in the gas flowcomputation of the following iteration cycle. This iterative ap-proach is illustrated in the lower part of Fig. 8.The separated computation of gas and liquid phase flow fieldsand the iterative exchange of interfacial transfer data entails anartificial decoupling of both phases. In particular for two phaseflows with intense phase interaction, this effect leads to a crit-ical overestimation of droplet source terms in the first iteration

    cycles. To achieve convergence of the iterative procedure, arelaxation of the droplet source term fields is employed. Recur-sive damping of the source terms on the level of the iterationcycles is realized by the following equation

    Si+1

    d,φ = αφSid,φ + (1 − αφ)S

    i

    d,φ . (54)

    According to Eq. 54, only a fraction of the source termsrecorded during the previous droplet tracking, Sid,φ, is con-

    tributing to the source term Si+1

    d,φ included in the current gas gasflow solution. The second contribution is calculated from thesource term included in the previous gas flow computation. Forgas flows which are substantially influenced by the fuel spray,strong relaxation of droplet source terms may be necessary toenforce convergence of the iterative procedure. In these cases anincreased number of two stage iterations may be necessary forcomplete consideration of phase interaction [24]. In the presentflow simulation, only weak relaxation (αφ > 0.5) is required toachieve a convergent solution for the two phase flow field within10 two stage iterations.

    5 The Hybrid procedureAs indicated in the preceding sections, the continuum descrip-tion of the Eulerian method has the advantage of close cou-pling of gas and liquid phase in a single numerical scheme.Thus, computation times are rather short as long as the num-ber of droplet classes used for the discretization of the spray issmall. Consequently, Eulerian methods are particularly suitedfor an approximate but efficient computation of polydispersetwo phase flows in combustors. Lagrangian methods in turnachieve a high resolution discretization of complex spray struc-tures by tracking large numbers of droplet parcels of variousinitial sizes and velocities. However, the price to be payed for

    SourcesDroplet

    Gas Phase

    Trajectories

    FieldFlow

    Droplet SourcesFlow Field,

    Droplet PhaseGas Phase

    Lagrangian Method

    Eulerian Method

    Figure 8: Structure of the Hybrid procedure

    a realistic spray representation is high. Due to the artificial de-

  • 51-8

    coupling of the two phase flow computation by separate solu-tion schemes for each phase and iterative realization of phaseinteraction, total computation times are rather large [24]. Forflow cases where strong relaxation of droplet source terms isrequired, time expenses can grow to practically unmanagableextents.For such two phase flows, a reduction of computational effort isachieved by preconditioning the two phase flow field by meansof an Eulerian method based on a coarse discretization of thespray. This is in fact the basic idea of the Hybrid procedure: Atwo stage combination of both methods in order to reduce totalcomputation times by Eulerian preconditioning yet maintainingthe detailed modeling of spray physics in a Lagrangian refine-ment stage. This approach is schematically illustrated in Fig. 8.Following an approximate computation of the two phase flow,the flow field and the droplet source terms are passed to the re-finement stage. Here, Lagrangian iteration cycles are based ona fine discretization of droplet injection conditions and an ad-vanced modeling of secondary droplet breakup. Since the twophase flow field calculated by the Eulerian method already ac-counts for spray effects on the gas flow, the droplet source termsrecorded during subsequent tracking steps are rather close to thefinal flow result. Consequently, the number of iterations is sig-nificantly reduced compared to a standard Lagrangian simula-tion.

    6 Simulation of a LPP premix duct flowThe performance and accuracy of the numerical methods pre-sented is demonstrated by a simulation of the two phase flowin the premix duct of a LPP combustor. The experimental in-vestigation of this combustor has been part of an extended re-search project on low emission combustion concepts [15], [16],[13]. A detailed description of the test rig and the measurementtechniques is presented in a parallel study [20]. The combustorsection of interest for the present flow simulation is illustratedin Fig. 9. Compressed air is supplied to the cylindrical duct

    Figure 9: Premix zone of the LPP research combustor

    (l = 124mm, di = 44.6mm) by two coaxial annular ducts.The fuel is injected into the gas flow by a pressure swirl atom-izer aligned with the duct axis. The nozzle diameter is about1 mm. In order to perform PDPA measurements, optical ac-cess to the flow is given by circumferencial slits in the ductliner at various axial positions. For spray visualizations, themetallic liner is substituted by a quartz glass cylinder. Premixand reaction zones are separated by an arrangement of swirlervanes acting as a flame stabilizer. The complete configurationillustrated in Fig. 9 is enclosed in a pressure casing with wa-ter cooled window ports. The outer coaxial annular air flow(bypass air) is shielding the duct liner from droplet impact andfilm formation. The highly accelerated inner flow (atomizationair) is focused directly onto the conical fuel sheet generated bythe pressure swirl atomizer. The fundamental idea of this atom-ization concept is a further reduction of droplet sizes by high

    velocity aerodynamic interaction between spray and gas flow.The operating point of the premix duct investigated in this studyis specified by the flow parameters summarized in Table 1. The

    Gas Flow (Air) Fuel (Diesel)

    ṁg 213 g/s ṁfuel 6 g/s

    Tg 753 K Tfuel 350 K

    pg 4 bar vsheet 30 m/s

    Tug 15 % βsheet 40◦

    Table 1: Parameters at the inlet of the premix duct (z = 0mm)computational domain of the flow simulation is illustrated inFig. 10 including (from left to right) intake section, coaxial an-nular ducts, premix zone, swirler vanes, reaction zone, dilutionholes and burnout zone. The axial coordinate is measured from

    Premix

    zone

    Figure 10: Computational domain (30◦-Segment)

    the atomizer nozzle. To model the evaporation behavior of thediesel spray, tetradecane is used as a single component dieselsubstitute in the present flow simulation. A detailed descrip-tion of the calculated non-reacting single phase gas flow in theintake and premix zone is given in Ref. [20].

    6.1 Discretization of the sprayIn order to derive droplet phase boundary conditions for the Eu-lerian method and droplet initial conditions for the Lagrangianmethod, the spray is visualized in the near field of the atomizer(0 < z < 10 mm). A side view on the three dimensional spraycone is given by the flashlight shadowgraph in Fig. 11(a). Thepicture was taken under atmospheric, cold conditions withoutbypass air flow, using water as a fuel substitute. It is evidentthat the outer region of the spray is dominated by larger fuelfragments. To get an impression of the spray structure inside

    Figure 11: (a) Flashlight shadowgraphy, (b) Laser light sheet

    the cone, a laser light sheet photograph is shown in Fig. 11(b),which was taken under real operation conditions of the combus-tor. In contrast to the coarse structure of the outer spray region,this cross view reveals a very fine droplet distribution in thespray cone.

  • 51-9

    The spray visualizations indicate two basic processes that gov-ern the atomization of the liquid fuel [15]. Due to its high swirl,the fuel is leaving the nozzle as a conical sheet. According toFig. 11(a), this sheet is completely disintegrating along a dis-tance of 1 to 2 mm. In this immediate near zone of the noz-zle, prompt atomization is the governing process. This mecha-nism is controlled mainly by the internal sheet dynamics [14].Further disintegration of sheet fragments is induced by aerody-namic interaction with the high velocity gas flow which pene-trates the spray. The resulting small secondary fragments aredragged by the gas flow into the core flow as indicated in Fig.11(b).With respect to the numerical simulation, it is obvious that arepresentation of the complex two phase flow in the nozzle nearfield by non-interacting droplets is a rather crude approximationof the physical reality. But despite of this simplified modeling offuel atomization, the simulation of spray dispersion and evapo-ration in the premix duct agrees well with the experimental data.In the following sections, two strategies are presented to derivedroplet injection conditions at the nozzle. The Eulerian calcu-lation is essentially based on droplet data measured at a down-stream position of the atomizer. It is important, that secondaryatomization has ceased completely and droplets are spherical atthis position. Basically, the procedure starts from an assumeddiscretization of the injection conditions of the droplet phase. Insubsequent optimization iterations, the computed droplet datais fitted to the corresponding measured data. In contrast to thisapproach, the Lagrangian calculation is based on a rather crudeapproximation of the fuel sheet disintegration in the nozzle nearfield. Here, a far more physical description of the spray in thesecondary atomization region is achieved by modeling dropletdeformation and breakup. A similar strategy is described in Ref.[7] where pressure swirl fuel injection into a diesel engine com-bustion chamber is discussed.

    6.1.1 Eulerian method, droplet phase boundary conditions

    Since secondary droplet breakup is observed in a spray regionup to 50 mm downstream of the atomizer, measured dropletdata at z = 55 mm is used for an optimization of droplet phaseboundary conditions. The droplet size distribution of the sprayis discretized by means of 3 diameter classes. The volume fluxfraction (or normalized volume flux) of each class is describedby a Rosin-Rammler distribution evaluated at the representativeclass diameter Di

    V̇i

    V̇= 1 − exp � − � DiD0.632 � n � , D0.632 = D0.50.693 1n . (55)

    Values of D0.5 = 46µm for the mass median diameter andn = 4.8 for the spreading parameter are determined as a goodfit to the reference droplet data at z = 55 mm. The mean injec-tion velocity of the droplet phase of 30 m/s is derived from themean velocity of the liquid sheet. The remaining parameterswith significant influence on the spray structure are the injec-tion angles of the droplet phases. Introducing 3 angle classesper diameter class leads to a final discretization of the spray bymeans of 9 droplet phases with different boundary conditionsat the nozzle. Table 2 summarizes the correlations betweendroplet phase diameter, normalized volume flux and injectionangle employed for the Eulerian two phase flow simulation inthe present study.

    6.1.2 Lagrangian method, droplet initial conditions

    In the Eulerian calculation, the boundary conditions of thedroplet phase at the nozzle have to account implicitly for sec-ondary atomization in an extended downstream flow region.

    ���[ ��� ] V̇i/V̇ [1] βi [◦]

    0 - 41.2 4/15 341/30 71/30 21

    41.2-50.7 4/15 301/30 101/30 20

    50.7-∞ 4/15 271/30 13.51/30 20.5

    v f3 0 m / s

    Table 2: Droplet phase boundary conditions

    It is evident that small secondary droplets originating frombreakup of larger sheet fragments in outer flow regions may notbe reproduced by a spray representation as described in the pre-vious section. The approximation by rigid spherical particlesof different size injected into the contracting, high velocity gasflow leads to a separation of droplet sizes. As a consequence,tiny and small droplets are captured in the axis region of thecore flow unless they are injected with unphysically large radialvelocity components.In the Lagrangian calculation, secondary breakup of dropletsis taken into account during trajectory integration. Thus, onlyprompt atomization of the conical sheet has to be considered forthe formulation of droplet initial conditions. The basic idea is toinject most of the fuel in form of large droplets with sizes simi-lar to the characteristic sheet thickness of about 100 to 200 µm.Due to this coarse primary spray structure and the high rela-tive velocities in the nozzle near zone, the critical conditionsof droplet breakup are significantly exceeded as indicated bythe data points mapped in Fig. 3. Modeling of delayed dropletdeformation, drag increase and breakup results in a fine sec-ondary spray contribution to the core flow region in the premixduct. Very good agreement to the experimental droplet data is

    Droplet diameter:

    • 10 size classes equally spaced from 0 to 200 µm• Rosin-Rammler distribution of V̇i/V̇• D0.5 = 88.5 µm, n = 3

    Droplet velocities:

    • Sampled, Gaussian distribution• v = 30 m/s, σv = 5 m/s

    Injection angle:

    • Sampled, Gaussian distribution• β = 40◦, σβ,i = 30◦, . . . , 5◦

    • βmax = 45◦ (clipping value)

    ������������������������������

    ������������������������������

    βsheet

    ������

    vsheet

    ������������������������������������

    Table 3: Droplet initial conditions

    achieved by using the spray discretization summarized in Table3. Droplet velocity and injection angle are sampled as randomnumbers with Gaussian distributions. From Fig. 11(b) it is ob-vious that the disintegration of the conical sheet is responsiblefor a fine primary contribution to the spray in the core flow. Tomodel this effect approximately, the variance σβ,i of the injec-tion angle β is correlated with droplet size resulting in valuesof 30◦ for the smallest droplets (D from 0 to 20µm) up to 5◦

    for the largest droplets (D from 180µm to 200µm)

  • 51-10

    6.2 Results

    To illustrate the calculated two phase flow in the premix duct,contour plots of the Eulerian flow simulation are discussed first.Comparing the axial gas velocities of the single phase and thetwo phase flow calculations from Fig. 12 indicates that spray-induced deceleration of the gas flow is limited to the core flowof the duct. In particular in the nozzle near zone, the axial gasvelocity is decreased by up to 40 m/s due to the aerodynamicacceleration of the droplet phase.The influence of fuel evaporation is indicated by the gas flowtemperature and fuel vapor concentration in Figs. 13 and 14.Although the gas phase experiences a substantial temperaturedrop across the whole core flow region, significant concentra-tions of fuel vapor are not calculated in the first half of theduct. This delay in vapor generation is a consequence of lowevaporation rates in the transient heating phase of the droplets.This conclusion is confirmed by an analysis of Lagrangian sin-gle droplet computations, which indicate that heatup and prop-agation time scales of typical droplets are of the same order.In total, 28% of the injected fuel is evaporated in the Euleriansimulation, in contrast to a value of 42% in the Lagrangian sim-ulation. The difference is caused by the secondary atomizationmodeling in the tracking algorithm, resulting in considerablenumbers of small, rapidly evaporating droplet fragments. Atthis point it should be noted that the total fraction of evaporatedfuel substantially depends on the D32-correlation used for thesecondary breakup modeling. Summarized over all calculatedbreakup events, Eq. 47, which is actually not used, leads to afragment mass median diameter of D0.5 = 53µm, whereas Eq.48 results in a value of D0.5 = 38µm.To compare the calculated axial volume flux density αd Ud withPDPA measurements, it is weighted by the annular area andnormalized by the total axial volume flux

    V̇r

    V̇=

    2πr+0.5mm�

    r−0.5mm

    αd Ud r dr

    2π24mm�

    0

    αd Ud r dr

    . (56)

    This normalized volume flux is illustrated in Fig. 15 and repre-sents the fraction of the total liquid volume flux which passes anannular fraction of the duct cross section. The contour plot in-cludes mean trajectories determined by integration of the meanaxial velocity of the droplet phase. The trajectories are evalu-ated for the two limiting size classes and clearly demonstratethe influence of initial droplet momentum on the propagation ofthe droplet phase.The second point of discussion is concerning the comparisonof experimental droplet data with Eulerian and Lagrangian twophase flow simulations. In this context, the Hybrid procedure isused as a computational tool to accelerate the Lagrangian cal-culation. The overall reduction of computation time achievedis about 30% of the time required for a standard Lagrangianflow simulation without Eulerian preconditioning. Radial pro-files of number averaged two phase flow variables are presentedat axial positions z = 20, 55 and 90 mm. With respect to themean axial velocities of the droplets shown in Figs. 16 and 17,both numerical methods predict a maximum in the core flow(r < 6 mm) which is not observed in experiment. In partic-ular at z = 55 and 90 mm the axial droplet velocities in thecore flow region are overestimated by the Lagrangian calcula-tion. Basically, this effect is a result of an insufficient spray-induced flow deceleration. This conclusion is supported by theunderestimated liquid volume flux in the core flow region at thecorresponding axial positions as shown in Fig. 19. With re-spect to the Eulerian result shown in Fig. 18, it is evident that

    the unavailability of secondary breakup models requires smallinjection angles of the droplet phase to meet the radial volumeflux profile in the second half of the duct flow. Consequently,substantial deviations are observed in the upstream flow regionwhere secondary atomization occurs. As illustrated by Figs. 20and 21, Eulerian and Lagrangian flow simulations predict rathersimilar radial profiles of the Sauter mean diameter. Althoughthe calculated values are deviating from experimental data byup to 20 µm, both flow simulations reproduce the trend in theevolution of the droplet size spectrum.

    7 ConclusionsEvaporating fuel sprays in combustor flows are characterizedby high rates of mass, momentum and enthalpy transfer be-tween spray and gas flow. Fuel atomization, spray dispersionand evaporation are often complicated by additional physicaleffects such as droplet-wall interaction, shear driven evaporat-ing wall films or secondary breakup of droplets.The primary objective in this study is the design of a compu-tational tool for an efficient numerical simulation of combus-tor two phase flows including advanced modeling of secondaryatomization physics. Compared to state of the art Lagrangianspray simulations, a significant reduction of computation timeis achieved by the presented Hybrid procedure. Basically, thiscomputational strategy is a two stage combination of an Eule-rian and a Lagrangian method. The Eulerian method is usedas an efficient preconditioner for the interacting two phase flowfield, in order to reduce the number of subsequent Lagrangiangas flow computation - droplet tracking iterations. In this re-finement stage, advanced modeling of secondary breakup ofdroplets including bag, multimode and shear mechanisms isused to improve the physical description of the spray.To assess the accuracy of the two fundamental numerical ap-proaches, Eulerian and Lagrangian flow simulations are com-pared with droplet data measured in the two phase flow of a LPPcombustor premix duct. The atomization concept employed inthis premix duct achieves a substantial improvement of fuel at-omization by aerodynamic breakup of droplets in an extendedflow region downstream of the nozzle. With respect to the Eule-rian flow simulation, extrapolation of measured droplet data tothe point of injection with implicit consideration of secondaryatomization effects results in a rather crude approximation ofthe spray structure. Since secondary breakup of droplets is mod-eled in detail by the tracking algorithm, the numerical descrip-tion of the spray structure in the Lagrangian flow simulation issignificantly improved.It is evident that the computational acceleration achieved by theHybrid procedure significantly depends on the structure of thetwo phase flow. With respect to the rather uncritical premix ductflow presented in this study, it is certainly questionable whetherthe moderate time savings are worth the additional complexityof the CFD program. However, technical practice offers a suf-ficient number of critical two phase flow applications, wherestrong relaxation of droplet source terms requires an excessivenumber of Lagrangian coupling iterations [24]. In these flowcases, the Hybrid procedure has an increased potential for sub-stantial speedup of flow simulations.

    AcknowledgementThis work has been founded by the Department of Education,Science, Research and Technology of Germany under contractNo. 50-807642. The support is gratefully acknowledged.The authors would like to thank Mr. Michael Willmann, BMWRolls-Royce, Germany, for his support concerning the develop-

  • 51-11

    ment of droplet breakup models.

    References[1] B. Abramzon and W.A. Sirignano. Droplet Vaporisation

    Models for Spray Combustion Calculations. InternationalJournal of Heat and Mass Transfer, 32:1605–1618, 1989.

    [2] S. K. Aggarwal and F. Peng. A Review of Droplet Dynam-ics and Vaporization Modeling for Engineering Calcula-tions. ASME–Journal of Engineering for Gas Turbinesand Power, 117:453–461, 1995.

    [3] C. T. Crowe. Review - Numerical Models for Dilute Gas-Particle Flows. ASME–Journal of Fluids Engineering,104:297–303, 1982.

    [4] G. M. Faeth. Evaporation and Combustion of Sprays.Progress in Energy and Combustion Science, PergamonPress, 9:1–76, 1983.

    [5] A. D. Gosman and E. Ioannides. Aspects of ComputerSimulation of Liquid-Fueled Combustors. Journal of En-ergy, 7(6):482–490, 1983.

    [6] M. Hallmann, M. Scheurlen, and S. Wittig. Computa-tion of Turbulent Evaporating Sprays: Eulerian Versus La-grangian Approach. Transactions of the ASME, 117:112–119, 1995.

    [7] Z. Han, S. Parrish, P. V. Farrell, and R. D. Reitz. Mod-eling Atomization Processes of Pressure-Swirl Hollow-Cone Fuel Sprays. Atomization and Sprays, 7:663–684,1997.

    [8] L.-P. Hsiang and G. M. Faeth. Near-Limit Drop Defor-mation and Secondary Breakup. International Journal ofMultiphase Flow, 18(5):635–652, 1992.

    [9] M. Ishii. Thermo-Fluid Dynamic Theory of Two PhaseFlow. Eyrolles, 1975.

    [10] M. Krämer. Untersuchungen zum Bewegungsverhaltenvon Tropfen in turbulenter Strömung in Hinblick auf Ver-brennungsvorgänge. Dissertation, Universität Karlsruhe,1988.

    [11] S. A. Krzeczkowski. Measurements of Liquid DropletDisintegration Mechanisms. International Journal of Mul-tiphase Flow, 6:227–239, 1980.

    [12] B.E. Launder and D.B. Spalding. The Numerical Compu-tation of Turbulent Flows. Computer Methods in AppliedMechanics and Engineering, 3:269–289, 1974.

    [13] W. Layher, G. Maier, and Wittig. S. Schadstoffemissionenund Betriebsverhalten einer LPP-Brennkammer. In DGLRJahrestagung in Bremen, Germany, 1998.

    [14] A. H. Lefebvre. Atomization and Sprays. HemispherePublications, New York, 1998.

    [15] G. Maier, M. Willmann, and S. Wittig. Development andOptimization of Advanced Atomizers for Application inPremix Ducts. In 97-GT-56. ASME, 1997.

    [16] G. Maier and S. Wittig. Effects of Liquid Properties onthe Operating Performance of Air-Assisted Pressure SwirlAtomizers. In ILASS Europe, pages 193–199, 1998.

    [17] W. K. Melville and K. N. C. Bray. A Model of the Two-Phase Turbulent Jet. International Journal of Heat andMass Transfer, 22:647–656, 1979.

    [18] D. Milojević. Lagrangian Stochastic-Deterministic (LSD)Predictions of Particle Dispersion in Turbulence. Particleand Particle Systems Characterization, 7:181–190, 1990.

    [19] M. Pilch and C. A. Erdman. Use of Breakup Time Dataand Velocity History Data to Predict the Maximum Sizeof Stable Fragments for Acceleration-Induced Breakup ofa Liquid Drop. International Journal of Multiphase Flow,13(6):741–757, 1987.

    [20] K. Prommersberger, G. Maier, and S. Wittig. Validationand Application of a Droplet Evaporation Model for RealAviation Fuel. In RTO-MP-14, pages 16.1–16.12, 1998.

    [21] W. Rodi. Turbulence Models and their Application in Hy-draulics - A State of the Art Review. IAHR, 1984.

    [22] W. Samenfink. Sekundärzerfall von Tropfen. In Atom-ization and Sprays, Short Course. Institut für ThermischeStrömungsmaschinen, Universität Karlsruhe (TH), 1995.

    [23] R. Schmehl. Theory and Application of Single Compo-nent Evaporation Models. Technical report, Institut fürThermische Strömungsmaschinen, Universität Karlsruhe,http://itsnova.mach.uni-karlsruhe.de/∼schmehl, 1998.

    [24] R. Schmehl, H. Rosskamp, M. Willmann, and S. Wittig.CFD Analysis of Spray Propagation and Evaporation In-cluding Wall Film Formation and Spray/Film Interactions.In ILASS ’98 Europe, pages 546–555, 1998.

    [25] H. C. Simmons. The Correlation of Drop-Size Dis-tributions in Fuel Nozzle Sprays; Part I: The Drop-Size/Volume-Fraction Distribution. ASME–Journal of En-gineering for Power, 99:309–314, 1977.

    [26] W. A. Sirignano. Fuel Droplet Vaporization and SprayCombustion Theory. Progress in Energy and CombustionScience, Pergamon Press, 9:291–322, 1984.

    [27] W. Snyder and J. L. Lumley. Some Measurements ofParticle Velocity Autocorrelation Functions in a TurbulentFlow. Journal of Fluid Mechanics, 48:41–71, 1971.

    [28] S. Wittig, M. Hallmann, M. Scheurlen, and R. Schmehl.A New Eulerian Model for Turbulent Evaporating Spraysin Recirculating Flows. In AGARD-CP-536, pages 37.1–37.11, 1993.

  • 51-12

    0.0 15.0 30.0 45.0 60.0 75.0 90.0 105.0 120.0Ugas [m/s] :

    Figure 12: Axial gas velocity: Single phase calculation (top half) and two phase calculation (bottom half)

    460.00 490.00 520.00 550.00 580.00 610.00 640.00 670.00 700.00 730.00 760.00Tgas [K] :

    Figure 13: Calculated gas temperature

    0.000 0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.040 0.045 0.050Cfuel [-] :

    Figure 14: Calculated vapor concentration

    0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50VolFlux [-] :

    33 mµ57 mµ

    Figure 15: Normalized volume flux and mean trajectories

  • 51-13

    40

    60

    80

    100

    ud

    [m/s

    ]

    z=90 mm

    40

    60

    80

    100

    ud

    [m/s

    ]z=20 mm

    z=90 mm

    40

    60

    80

    100

    u d[m

    /s]

    z=55 mm

    Gas Phase, CalculationDroplets, ExperimentDroplets, Calculation

    0 5 10 15 20r [mm]

    Figure 16: Mean axial velocities (Euler)

    0 5 10 15 2040

    60

    80

    100

    u d[m

    /s]

    z=55 mm

    0 5 10 15 2040

    60

    80

    100

    ud

    [m/s

    ]

    z=20 mm

    Gas Phase, Calculation

    Droplets, CalculationDroplets, Experiment

    0 5 10 15 20r [mm]

    40

    60

    80

    100

    ud

    [m/s

    ]

    z=90 mm

    Figure 17: Mean axial velocities (Lagrange)

    0

    0.1

    0.2

    0.3

    0.4

    Vo

    lFlu

    x

    z=90 mm

    0

    0.1

    0.2

    0.3

    0.4

    Vol

    Flu

    x

    z=55 mm

    0

    0.1

    0.2

    0.3

    0.4

    Vol

    Flu

    x

    z=20 mm

    0 5 10 15 20

    0 5 10 15 20

    0 5 10 15 20r [mm]

    Figure 18: Normalized volume flux (Euler)

    0 5 10 15 200

    0.1

    0.2

    0.3

    0.4

    Vo

    lFlu

    x

    z=55 mm

    0 5 10 15 200

    0.1

    0.2

    0.3

    0.4

    Vol

    Flu

    x

    z=20 mm

    0 5 10 15 20r [mm]

    0

    0.1

    0.2

    0.3

    0.4

    Vo

    lFlu

    x

    z=90 mm

    Figure 19: Normalized volume flux (Lagrange)

    0 5 10 15 20r [mm]

    0 5 10 15 200

    25

    50

    75

    100

    D32

    [µm

    ]

    z=55 mm

    0

    25

    50

    75

    100

    D32

    [µm

    ]

    z=20 mm

    0

    25

    50

    75

    100

    D32

    [µm

    ]

    z=90 mm

    z=55 mm

    Figure 20: Sauter mean diameter (Euler)

    0 5 10 15 200

    25

    50

    75

    100

    D32

    [µm

    ]

    z=20 mm

    0 5 10 15 200

    25

    50

    75

    100

    D32

    [µm

    ]

    z=55 mm

    0 5 10 15 20r [mm]

    0

    25

    50

    75

    100

    D32

    [µm

    ]

    z=90 mm

    Figure 21: Sauter mean diameter (Lagrange)