You are on page 1of 76

Prog. Energy Combust. Sci. 1983, Vol. 9, pp. 1-76. Printed in Great Britain. All rights reserved.

0360-1285/8350'00+ "50 Copyright 1983. Pergamon Press Ltd.

EVAPORATION AND COMBUSTION OF SPRAYS


G . M . FAETH Department of Mechanical Engineering, The Pennsylvania State University, U.S.A. NOMENCLATURE Symbol a A,A' A I, Ag Bo,BF C CB Co Cos C~ CI CL Cp d dp dpei D Da Da E acceleration of gravity empiricalconstants, Eq. (2.10) liquid and gas flow areas mass transfer driving potentials, Eqs. (5.10) and (5.11) particle concentration Bassett force coefficient, Eq. (5.26) drag coefficient Stokes drag coefficient turbulence model constants virtual mass drag coefficient, Eq. (5.26) lift coefficient specific heat at constant pressure nozzle or injector diameter drop diameter extinction diameters, Eqs. (5.21)-(5.23) disk diameter; 24 mass diffusivity Damk6hler number extinction Damk6hler numbers, Eqs. (5.18)(5.20) ignition Damk6hler number activation energy mixture fraction liquid fraction generalized spray distribution function particle drag force body force on particle forced convection extinction factor, Eq. (5.24) fugacity of species i particlelift force rate of change of drop velocity natural convection extinction factor, Eq. (5.24) Froude number, Table 10 square of the mixture fraction or concentration fluctuations rate of production of concentration fluctuations, Table 2 rate of production of turbulence kinetic energy, Table 2 heat transfer coefficient enthalpy of speciesi enthalpy of formation of species i total enthalpy internalenergy turbulence kinetic energy equilibrium constant; empirical constant, Eq. (5.32) empirical constants preexponentialconstant dissipationlength scale drop mass, number of elements mass transfer rate mass flux at drop surface power in power law expression; number of species; order of global reactions
1
JPECS 9-I/2-A

fik N Nu p P Pr Qg Qr r R RI Re Rp s Sc SMD Sp Srad S~ S~.j t tb te tp tt T u ~ Ucir ue Urej v v w We x y Y~ Z x,u ~ ~I,~9 ~, 6 Ah~ Ahp~ AN e ~ t/ 0 2 Af p Pt v

Dal E f fy fp Fo Fe FF F~ FL F~ FN Fr g
Gg

Gk
h h~ h H I k K K', K", K'" K, Le m th" n

number of drops in class k per unit time total number of drop classes Nusselt number pressure probability density function Prandtl number heat of combustion of species i heat of combustion per unit mass of oxidant, Eq. (5.16) radialdistance reactedness, Eq. (2.11); gas constant reaction rate, Eq. (2.9) Reynolds number rate of change of drop diameter stoichiometric oxidant to fuel ratio Schmidt number Sauter mean diameter source term in spray equation, Eq. (4.2) radiation sourceterm source term, Eqs. (2.1) and (2.5) exchange rate between phases, Eq. (4.1) time drop breakup time eddy lifetime time following aparticle drop transit time, Eq. (5.5) temperature axial velocity dimensionless drop velocity, Eq. (5.30) circulation velocity extinction velocity relative drop velocity radial velocity weighted radial velocity, Eq. (2.6) tangential velocity Weber number axialdistance coordinate distance mass fraction of species i coordinate distance vector average property factor, Eq. (3.1) liquid volume fraction, void fraction thermaldiffusivity Dirac delta function; thickness of film in film theory sensible enthalpy change of species i enthalpy deviation of species i number of particles in a generalized unit of volume dissipation rate of turbulence kinetic energy mass flux fraction of species i Kolmogoroffmicroscale generic property;spray angle thermalconductivity longitudinalintegral scale absolute viscosity turbulent viscosity kinematic viscosity

2 vq massfraction of elementj in species i


v~j, v~} stoichiometriccoefficients

G.M. FAETH dimensional spray models were routinely employed for performance estimations of liquid rocket engines over twenty years ago. The use of spray models during the development of furnaces, gas turbines, diesel engines, etc., has been less attractive, since complex models are needed for a realistic simulation and developmental tests can be carried out with relative ease. However, development problems have become more complex, often involving stringent pollution and performance objectives in addition to traditional design requirements. This has created new interest in models which can provide insight concerning system performance, since cut and dry developmental testing becomes less certain and more costly when faced with a multiplicity of goals. Evidence of increasing use of relatively sophisticated spray models for industrial development is provided by recent reports of their application for gas turbine systems, l 3 internal combustion engines,4 and fire safety systems. 5 At this time, however, while models can provide useful insight, their capabilities for a priori predictions are limited. Substantial development and experience with the use of spray models will be required before they become reliable design tools. The objective of this review is to describe recent spray evaporation and combustion models and their comparison with measurements. The discussion is limited to turbulent two- and three-dimensional spray processes found in furnaces, gas turbine combustors, internal combustion engines, etc. Models which attempt to describe the detailed structure of the flow will be emphasized. Individual drop processes are an important aspect of modeling sprays; therefore, recent work in this area will also be considered, concentrating on the behavior of drops in a spray environment. Aspects of spray modeling which are not considered here, particularly one-dimensional and lumped parameter models, are discussed in earlier reviews by Faeth, 6 Harrje and Reardon, 7 and Mellor. s The behavior of single-drops and sprays is also considered in Refs. 6-8, as well as in recent reviews by Williams,9 11 Law,12 Krier and Foo, 13 Chigier ~4 and Hedley et al. 1 s Within the class of spray models of interest here, two major categories can be distinguished: (1) Locally homogeneous flow (LHF) models, where the gas and liquid phases are assumed to be in dynamic and thermodynamic equilibrium, i.e. at each point in the flow both phases have the same velocity and temperature and are in phase equilibrium. This is a limiting case, which only accurately represents a spray consisting of infinitely small drops. (2) Separated flow (SF) models, where effects of finite rates of transport between the phases are considered. Models of this type are generally needed for quantitative predictions of practical sprays. Both approaches will be considered, even though SF models are of the greatest practical importance. LHF

variable of integration, Eq. (5.26) density surfacetension turbulent Prandtl/Schmidt number particle relaxation time, Eq. (5.6) characteristic chemical reaction time characteristic drop evaporation time, Eq. (3.41) Zdh characteristic drop heat-up time, Eq. (3.38) Zdm characteristic drop momentum exchange time, Eq. (3.37) dimensionlesstime, Eq. (5.30) Tp flow relaxation time, Eq. (3.36) Ts generic property p a~t aO z Zch zde

Subscripts
avg average value A air or ambient gases b completely burned state, bulk liquid centerline value crit critical value f liquid phase F fuel 9 gas phase P drop property pc end of potential core P combustion products t' reference condition drop surface S completely unburned state U 0 initial value oO ambient condition of spray

Superscripts
o

()' fluctuating quantity (-) time-averaged quantity


Other
() weighted average, Eq. (5.40) 1. INTRODUCTION Recent theoretical and experimental advances have created new interest in the development of models of spray processes. Theoretical progress includes the appearance of models of turbulent flow and combustion, the development of general-purpose computer codes capable of solving model equations, and enhanced computer capabilities. Experimental progress involves the development of improved instrumentation for evaluating model predictions, e.g. laser-Doppler anemometry, light scattering methods for measuring particle sizes, holography, etc. In this environment, the pace of model development has been unusually rapid. A major objective of modeling sprays is to reduce the time and cost of process development. Therefore, spray models have received greatest acceptance where modeling is relatively uncomplicated, testing is expensive, the penalties of test failure are severe, and time of development is limited. For example, one-

ideal gas value

Evaporation and combustion of sprays models have distinct advantages in some cases and also have been subjected to more complete evaluation than SF models, to date. Discussion of L H F models also provides a useful introduction to the treatment of gas phase flow processes for SF models. The review begins with a discussion of L H F models and their comparison with measurements. Conventional methods of treating drop transport characteristics are then considered, utilizing the results of L H F models to estimate the environment of drops in a spray. Recent SF models and their comparison with measurements are then discussed. The review concludes with a discussion of recem progress on modeling interactions between drops and the flow in both dilute and dense sprays, i.e. sprays having low and high liquid volume fractions, respectively.

accurate initial conditions at the injector exit. A second advantage of L H F models is that computations for sprays are essentially identical to computations for single-phase flows; therefore, existing computer codes can be employed which require shorter computation times than the more complex SF procedures. A third advantage is that L H F "models require far fewer empirical constants than SF models, which reduces the "art" involved in making predictions. Experience to date indicates that L H F models provide a reasonable first estimation of the extent and character of a spray process. Another attribute of L H F models is that they give the lower bound for the size of the spray process--all real sprays will require a larger volume due to finite interphase transport rates. Therefore, L H F model predictions can provide an indication of the potential process improvements by improving atomization, prior to ~esting. 2.1.2. Summary of LHF models Table 1 is a summary of L H F models that have been reported for sprays. The list is not exhaustive, but represents cases where well-defined experimental results were employed to test the predictions. Omitted from the list are several phenomenological models that embody the L H F approximation which have been developed to simulate spray combustion in diesel engines.Z9 33 Due to the transient three-dimensional nature of diesel combustion, convincing validation of the components of such models has not been achieved, although they have been useful in specific applications. The earlier models listed in Table 1, by Thring and Newby, 21 Newman and Brzustowski 22 and Shearer and Faeth, 23 employed integral models to represent the flow. These models were calibrated to only a limited degree with more basic and well-defined flows. Nevertheless, even these simplified models achieved broad agreement with overall phenomena such as flame lengths and spray boundaries. Later L H F models by Khalil and Whitelaw, 24 Khalil, 25 Shearer et al. 26 and Mao and coworkers 27'28 employed higher-order turbulence models which were calibrated using measurements in gaseous combusting and noncombusting flows. In these cases, predictions were compared with detailed measurements of spray structure. Under this level of scrutiny, the L H F models were found to invariably overestimate the rate of development of the flow, as expected due to their neglect of finite interphase transport rates. In the following, L H F models will be described in greater detail, with particular emphasis on the procedures employed in Refs. 24-28. Model predictions will then be compared with measurements. 2.2. Theory 2.2.1. Modeling the flow Under the LHF approximation, the flow model is essentially identical to a single phase flow. The only

2. LOCALLYHOMOGENEOUS FLOW APPROXIMATION 2.1. Introduction 2.1.1. Nature of the approximation The basic premise of L H F models is that rates of transport between the phases are fast in comparison to the rate of development of the flow field as a whole. This approximation requires that all phases have the same velocity and temperature, and that phase equilibrium is maintained at each point in the flow. Therefore, use of an L H F model implies that the process is mixing controlled. The dispersed phase must have infinitely small particle sizes for this model to be quantitatively correct. The L H F approximation is often still acceptable when particle sizes are finite. For example, cases where the dispersed phase is a gas or vapor and the continuous phase is a liquid are good candidates for L H F models due to the relatively low capacitance of bubbles. Therefore, L H F models have been extensively applied to bubbly flows and to processes of boiling and condensationJ 6 18 The application of the L H F approximation is more questionable for sprays. Nevertheless, it has long been recognized that spray and gas flames have similar appearance, suggesting that the spray acts similar to a single-phase flow and could be represented by the L H F approximation. Recent measurements by Onuma and Ogasawara 19 and Komiyama et al. 2 demonstrate this behavior more quantitatively. They find striking similarities between the structure of flames fueled with gases and with well-atomized sprays (maximum number densities for drop diameters of 10-20#m), when the two flows have the same stoichiometry and momentum. The major advantage of L H F models for sprays is that they require minimum information concerning injector characteristics, since initial drop size and velocity distributions play no role in the computations. Injector properties are difficult to obtain for spray nozzles and the performance of elegant SF models is discouragingly poor in the absence of

4~

TABLE 1. Summary of locally homogeneous flow (LHF) models of sprays Model b Integral, EQ, parabolic Integral, EQ, parabolic Integral, EQ, parabolic k-e-g, EQP, elliptic k-e-g, MEBU, elliptic
k-e-g, EQP, parabolic

Date Combustion of steam atomized spray in air, 0.1 MPa Evaporation of pressure atomized liquid near its critical point, 6-9 M P a Combustion of pressure atomized spray, no swirl, still air, SMD = 30/am (estimated),0.1 9 M P a Combustion of a pressure atomized spray, swirl, SMD = 45, 100/~m, 0.1 M P a Combustion of a pressure atomized spray, swirl, SMD = 45, 100#m, 0.1 M P a Evaporation of air atomized spray, no swirl, in still air, SMD = 29pm, 0.1 MPa Combustion of air atomized spray, no swirl, in still air, SMD = 35pm, 0.1 M P a Combustion of pressure atomized spray, no swirl, in still air, SMD = 30pm (estimated), 3-9 M P a

Reference

Flow configuration a

Exp erimentc

Assessment Qualitative agreement for flame length Good agreement for spray boundary

1953 Thring and Newby zl

Axisymmetric, boundary layer combustion

1971 Newman and Brzustowski z2

Axisymmetric, boundary layer evaporation

1977 Shearer and Faeth 23

Axisymmetric, boundary layer combustion

Spray and flame boundaries underestimated by 30-50 ~o, poor estimation of flow width Rate of development of process overestimated

1977 Khalil and Whitelaw z4

Axisymmetric, swirling flow with recirculation and combustion

1978 Khali125

Axisymmetric, swirling flow with recirculation and combustion

Improved prediction in some cases

1979 Shearer et al. 26

Axisymmetric, boundary layer evaporation

Rate of development of process overestimated. predicted mean velocities and mixture fraction 20-40 ~o below measurements near the injector Rate of development of process overestimated, flame length underestimated by 20 ~o Rate of development of process overestimated, spray lengths underestimated by 20 ~o

1980 Mao et al. z7


k-e.-9, EQP, parabolic k-e-9, EQP,

Axisymmetric, boundary layer combustion

1980 Mao et al. 28 parabolic

Axisymmetric, boundary layer combustion

"All cases shown here are steady, Refs. 29-33 describe transient phenomenological models of diesel spray combustion. b EQ implies local thermal equilibrium, EBU implies use of eddy-break-up model to estimate fuel concentration in conjunction with local thermal equilibrium, MEBU is the same as EBU except for an empirical modification for drop combustion, EQP implies use of local thermal equilibrium with probability density function for mean properties. Pressures are those of the spray environment.

Evaporation and combustion of sprays distinction between a two-phase L H F model and a single-phase flow model involves the representation of thermodynamic properties such as temperature, density, enthalpy, etc. (the state relationships). Therefore, relatively sophisticated flow models can be adopted for L H F models with little difficulty. Eliminating integral models from consideration, Reynolds 34 defines two major classes of turbulence models: full-field modeling (FFM) and large-eddy simulation (LES). The F F M method involves the use of partial differential equations describing suitably averaged quantities. The variables include a number of mean quantities (velocities, mixture fraction, etc.) and turbulence parameters (turbulence kinetic energy, turbulent stress components, etc.). Models must be constructed for various terms appearing in the governing equations for turbulence quantities, since they are not a closed set of equations. When these models are constructed, contributions from all scales of the turbulent motion must be considered. This presents difficulties since the large scale aspects of turbulent flows are frequently anisotropic while processes at small scales approach isotropy. Although some theoretical guidance is available for modeling these terms, models must be established by comparison with measurements and their generality is limited. The LES approach involves completing calculations of the time dependent, three-dimensional structure of the turbulent flow, with initial conditions constructed to reflect randomness. In this case, approximate modeling is only necessary for turbulence scales smaller than the computational grid spacing. The small scale of sub-grid scale turbulence is relatively universal in character and can be modeled more reliably than situations where the full range of turbulence scales must be considered. 34 While it appears that large-eddy simulations have the potential to remove many of the uncertainties of turbulence modeling, this approach is not as well developed as full-field models. At the present time, LES solutions generally require substantial computation times for practical problems, which make them less attractive for routine use as an aid to design. Reynolds 34 estimates that with projected computer development, routine use of LES models should be possible within a decade. However, spray models to date have generally employed the F F M approach and models of this type have the greatest potential for contributing to the design of spray processes and the development of spray models in the near future. Therefore, subsequent discussion will be limited to the F F M approach. F F M models may also be divided into two major classes: those based on time (Reynolds) averaged quantities, and those based on mass-weighted (Favre) averages. 35 Both methods are identical for constant density flows. Favre-averaging results in model governing equations for variable density flows which are almost identical to the equations found for constant density flows. This suggests that turbulence modeling

procedures developed for constant density flows might be applied without change for variable density flows. However, Lockwood and coworkers 36'37 point out that extension of modeling procedures in this manner can only be assessed empirically by comparison with measurements. When Reynolds averaging is used, the governing equations can only achieve the Favre-averaged form when a number of terms involving density fluctuations, which may not be small, are ignored. The solution of Reynolds-averaged equations then corresponds to a Favre-averaged solution, without computation of mass averages by solution of equations involving correlations of density and other variables. 36 Since this is only one of many approximations in turbulent flow models, Reynolds averaging has generally been adopted for sprays due to its computational convenience, even though either method could be applied. Therefore, the following disctission will be limited to the Reynolds-averaged model equations. Reynolds 34 and Bilger 3s describe Favreaveraging procedures in some detail. The bulk of Reynolds-averaged turbulence models follow procedures developed by workers at Imperial College. 3642 Since these methods have also been used for L H F spray models developed to date. 24-28 they will be described in some detail. The method involves a two-equation model of turbulence in which equations are solved for turbulence kinetic energy and its rate of dissipation. Local values of the effective turbulent viscosity are then determined from these quantities. The assumption of an isotropic turbulent viscosity is not satisfactory in all cases, e.g. in strongly swirling flows,43 however, the computational convenience and extensive validation of the k-e method have attracted most workers to this approach. 36 A second major assumption of most models is that the exchange coefficients of all species and heat are the same. This implies equality of both laminar and turbulent components of diffusivities. This is a reasonable approximation at high Reynolds numbers, where laminar effects are small, and for gases, where laminar diffusivities are nearly identical.44'4s (Bilger46 finds interesting exceptions for the low Reynolds number combustion of hydrogen in air.) Assuming equal exchange coefficients is less satisfactory for two-phase flows at low Reynolds numbers, under the L H F approximation, since the laminar diffusivity of even small drops is much smaller than that of gas molecules, aT The effect of laminar diffusion is generally not large in practical sprays, however, and the equal exchange coefficient approximation is reasonable in view of the other simplifications that must be made in modeling sprays. Comparison of spray models with detailed measurements primarily has been limited to steady axisymmetric flows in order to make the measurements tractable. Furthermore, buoyancy effects are usually small and the effect of buoyancy on turbulence properties has generally not been incorporated in spray models. With these limitations, the governing

G.M. FAEtH TABLE3. Source terms in Eqs. (2-5) q~ S,

equations can be placed in the following general form 36-4~

+!L(r",'Wl
r O r k a o ~?r}+So
(2.1) where q~ represents ~, ~, if, f , k, e, etc., cr~ is the turbulent Prandtl/Schmidt number of the variable ~b and S o is a source term. The designation q5 = 1 in Eq. (2.1) yields the equation of conservation of mass. The expressions for the source terms appear in Table 2. The variable H is the total enthalpy of the mixture including sensible, chemical and kinetic energy effects, since potential energy is generally negligible for typical spray processes, e.g.

+_a(po~-,6)

~ \Or/

f
/7

Srad (~D 2 Cg,p, C~Rf -gg -Co~p~

H = ~
and

Yihi+l(u2d-l)2q-w2)
h i = Ahi+hi.

e,

(2.2) Notes: 1. Positive sign is used for Sa for vertical upward flow. (2.3) 2. Turbulence model constants are assigned the following values: C~ = 0.09, C~, = 1.44, Cg, = 2.8, ak = 1.0, a, = 1.3, @ = 0.7, a 0 = 0.7, cr~, = 0.7, constant density flows C~, = Co~ = 1.89, variable density flows C,2 = Cg2 = 1.84. 3. Dilation and shear work terms have been ignored in the equation for ft.

Applying the k-e procedure, the effective turbulent viscosity is obtained from the following expression

p~ = C,fik2/~,.

(2.4)

TABLE2. Source terms in Eq. (2.1) q~


1

so
0

a f /p,
k Gk - )e

tga~ 1 8 ( t7 18#,'~

c~17~ 2#x,f ,6m2 a~


3r

y
/~ g
C~, G~- Co~,6~ g

0
Srad

C,Rf

Notes: 1.

=.,L2tt

F ff~aV

/0,~\ 2 /k2h

+W

/~mk 2

The empirical constants appearing in Eq. (2.4) and S o are assigned values based on matching predictions and measurements for well-defined flows. 3s Typical values of the constants are given as a note in Table 2. Some investigators employ slightly different values, original sources should be consulted for details. 36-4 Equation (2.1) provides the general elliptical form for steady, axisymmetric flow. This allows consideration of swirl and recirculation, which are often important for spray processes. The equations are also appropriate for premixed flames where axial diffusion is important even in non-recirculating flows. The governing equations can be simplified for nonswirling flows, particularly when the boundary layer approximations apply. The free jet is an important practical example, frequently used to evaluate spray models. The governing equations in this case can be written as follows 26 28,37.4-2

*)= (r#'
r dr \ a O ~ r /
(2.5) where*

J
2. Gg = IX, LkOx}
3. Turbulence model constants are assigned the following values: Cu = 0.09, C~, = 1.44, C,~ = 1.92, Co, = 2.8, Cg~ = 2.0, ak = 0.9, a, = 1.22, aI = tra = a 0 = a~, = 0.9. 4. Dilation and shear work has been ignored in the governing equation for/4.

~fo = fi~+ p" v'.

(2.6)

The parameters q~ and S~ appearing in Eq. (2.5) are summarized in Table 3, along with the appropriate

m * This involves retaining a density fluctuation term, p'v', which is normally omitted in solutions of Eq. (2.1).

Evaporation and combustion of sprays TABLE4. Summary of LHF model approximations Major assumptions l. 2. 3. 4. 5. LHF flow approximation Equal exchange coefficientsof all species and heat High Reynolds number, negligiblecontribution of laminar transport k-e turbulence model, no effect of buoyancy on turbulence properties Reynolds averaging neglecting density fluctuation terms or Favre averaging without computation of mass averages of properties Additional assumptions Adiabatic flow Negligibleradiation Low mach number Local chemical equilibrium Case 1 x/ x/ x/ ~/ Transport equations solved Mean quantities ti, f, if,f, a, f, if,f, ~. ti, f, N f,/4, radiation transport equation Case 2 x/ x/ ,/ ,/ Case 3

Turbulent quantities

k, e, 9

k, e, 9 or OYF

k, e, g

empirical constants. These constants were originally recommended by Lockwood and Naguib 42 based on measurements in single-phase flows. Recent measurements also support the choice of these values. 26'27 The fact that the constants are not the same as those in Table 2, and also differ for constant and variable density flows, represents an obvious lack of universality for k-e modeling. This difficulty has been recognized and tolerated for some time. 42 It is necessary to solve all the equations in either Table 2 or Table 3 for the most general cases. In many practical situations some phenomena can be ignored, which considerably simplifies the formulation. Several cases of interest are defined in Table 4, along with a summary of the major assumptions employed to derive Eqs. (2.1) and (2.5). Model properties in each of these cases will be described more fully in the following:

Case 1. The approximations of Case I (Table 4) are appropriate for many spray systems and have been used for several of the LHF spray models that have been evaluated to date. 24'26 28 Convective and radiative heat losses from evaporating sprays are generally small. These processes can frequently be ignored for high intensity combusting sprays as well. Allowable pressure drop limitations ordinarily imply low Mach numbers for most spray processes. The assumption of local chemical equilibrium yields a particularly simple formulation which is acceptable when production of pollutants is not considered. In the case of combusting flows, effects of premixed combustion must also be small, e.g. when the classical notions of diffusion flame modeling can be applied. With this formulation, the mean and turbulent quantities listed in Table 4 are determined directly by integration of the governing equations. Other scalar

properties, fi, T, Y~, etc. are found by employing a stochastic procedure which involves assuming a general form for the probability density function of the mixture fraction. Specification of the probability density function requires g, therefore, this approach is frequently denoted the k-~-g procedure. The method was initially suggested by Spalding,44'45 and was subsequently developed and applied to flames by Lockwood and Naguib 4z and Bilger. 35 When the Case 1 assumptions are employed, the instantaneous state of the mixture can be completely determined given the mixture fraction and pressure. The properties at each point in the flow correspond to the thermodynamic state attained when an amount f of injector fluid and (I-f) of ambient fluid, at their initial states, are adiabatically mixed and brought to thermodynamic equilibrium at the local pressure of the flow. Computation of this state only involves conventional adiabatic mixing or adiabatic flame temperature computations. In this manner, complex phenomena, such as dissociation and high pressure phase equilibrium, can be considered using routine thermodynamic state computations. The determination of the local state of the mixture as a function of mixture fraction (the state relationship) will be considered for both single- and twophase flows in the next section. Assuming for the moment that a state relationship O(f) has been found for all scalar properties (other t h a n f g, k and ~ which are known from solution of the governing equations), the mean value of 0 is determined from the following integral
ttl

0 = 1~ O ( f ) P ( f ) d f

(2.7)

where P(f) is the probability density function for f. The probability density function has been measured

G.M. FAETH method which has been applied to recent models of premixed combustion in furnaces.36 In this case R j- = {AfiYF, At~Yojs, A'fiYp/(l +s)}~/k (2.10)

for a number of flows. 3s'4s 50 Various functions have been suggested from these measurements, including clipped Gaussian, incomplete beta functions, rectangular waves, among others. 35'36'a2'45 Gosman et al. 36 note that the results are not particularly sensitive to the functional form used, although incomplete beta functions provide a computationally convenient formulation. Probability density distributions employed to date are most often characterized by two parameters, generally associated with the most probable value and the variance of the distribution. These parameters can be obtained by noting that f=

f0

fP(f)df;

g =

fi

(f-f)zP(f)df

(2.8)

where b o t h f a n d g are known from integration of the governing equations. Therefore, Eq. (2.8) provides two implicit equations which can be solved to determine any two-parameter P(f). Equation (2.7) can then be integrated, using the state relationship, to obtain f, T, Yi, etc. The use of more complex P ( f ) , requiring more than two parameters, has been suggested from time to time. However, this has not been pursued to a great degree, due to the need for modeling and computing additional moments of P ( f ) for use in equations analogous to Eq. (2.8). Case2. The approximation of local chemical equilibrium is relaxed in this formulation, which allows consideration of premixed combustion. This type of analysis for sprays is important for problems of flame stabilization; air atomizing injectors, where both air and fuel leave the injector; and diesel spray combustion, since air and fuel intermix prior to ignition even though pressure atomizing injectors are employed. When local chemical equilibrium is lost, the state of the mixture is no longer solely fixed by f, and additional scalar transport equations must be solved which require some means of specifying a rate of reaction in turbulent flow. This is an active area of research and only a few typical examples will be considered in the following. An approach that has been applied to both L H F and SF models of spray combustion involves application of Spalding's eddy-break-up model of turbulent reaction. 44'45's1'52 The method is used to describe the reaction rate of fuel (or fuel vapor for a SF model). The reaction rate of other major species is usually determined by stoichiometry under the assumption of a one-step reaction. The reaction rate is taken to be the smaller of either a global Arrhenius expression (based on mean properties) or the following eddy-break-up expression. 44'45'51,52 R: = -- C gfi,qly"F2~:/k (2.9)

where the s t a l e s t of the quantities in brackets is employed to compute the rate of reaction and the values of A and A' are roughly 4 and 2. Several other turbulent reaction rate formulas have been developed; 25'36'39'41'54'55 original sources should be consulted for details. Given the solution of the governing equations at a location, it is still necessary to determine the mean values of scalar quantities such as fi, T, Y~, etc. Lockwood, s6 describes a stochastic approach for this aspect of the solution which is closely related to the methods employed for Case 1 problems. In this case, the reaction process is represented by the reactedness

R_

Y~-Y~,.

Y,.b- Y,..

(2.11)

where i designates fuel, oxygen or product species and u and b designate completely unburnt or fully burnt conditions. Assuming a one-step reaction, R is independent of the major species chosen, although the approach could be extended to more complicated reaction schemes. 56 The method recognizes that local mean properties depend upon fluctuations of both f and R. In order to simplify the model, the f and R fluctuations are assumed to be uncorrelated and zones of reaction are taken to be thin so that the mixture is either completely burnt or completely unburnt. The time-averaged value of a fluid property O(f R) at any point is then given by

where P(R) is determined from/~ as follows P(R) = ( I - R t 6 ( R t + R 6 ( 1 - R ) (2.13)

where CR is a constant having a value on the order of unity and gYr is the square of the fuel mixture fraction fluctuation. The governing equation for gYr is identical to the g equation in Tables 2 and 3. Magnussen and Hjertager 53 propose a related

P ( f ) is found from f a n d g, using an assumed functional form for the probability density function, as described for Case 1. In order to apply the method, the state relationship must be determined for both R = 0 and 1, i.e. at the adiabatic mixing and full equilibrium limits. The second procedure involves computing mean scalar properties, ignoring the effect of fluctuations o f f and R, basing the calculations on the mean values determined from solution of the transport equations. This approach is computationally simpler than the stochastic method; however, this simplification is questionable when property variations are not linear i n f a n d R--which is generally the case. Case 3. In this formulation, effects of heat losses and kinetic energy are considered. The equation of radiation transfer must also be solved in order to yield S/~. Since radiative transport involves awkward integro-differential equations, a number of approximations of the equations have been developed to simplify their solution and specific references should be consulted for details. 35-42

Evaporation and combustion of sprays Knowing the local mixture fraction and enthalpy, the method described in Case 1 can be applied to obtain other scalar properties. The only change involves determining the state relationship for the local thermal energy content. Such an approach only approximates reality since the thermal energy content is not generally independent o f f For example, radiative heat losses are likely to be more significant for mixtures having a high temperature while low temperature mixtures are likely to absorb additional radiative energy. Since sprays generally involve moderate kinetic energies and radiative effects, errors due to this approximation are difficult to identify. 2.2.2. State relationships In this section, the state relationships required for solving the turbulence model equations will be considered, with particular emphasis on two-phase flows using the L H F approximation. In order to simplify the description, only Case 1 conditions, listed in Table 4, will be considered. The approach can be extended in a straight-forward manner to treat the other cases. In order to fix ideas, we shall consider experimental conditions which have been employed to evaluate both single- and two-phase flow models. This includes noncombusting and combusting gases, evaporating sprays and combusting sprays. 24-2s The spray results range from atmospheric pressure conditions where the ideal gas approximation can be used, to high pressure conditions when real-gas effects must be considered. Complete details concerning the construction of state relationships for these processes are provided by Shearer and Faeth 57 and Mao et al: s We assume that all properties of the injector and ambient flows, the mixture fraction, and the pressure are known. The general case of combustion in a high pressure environment is considered, where provision must be made for real-gas effects. The analysis is constructed for n species composed of m elements in two phases, gas and liquid. In practice, more phases can be present, e.g. due to the formation of soot and condensation of water into an immiscible liquid phase; however, the analysis can be readily extended to consider such instances. At high pressures, gases have appreciable solubility in liquids, therefore, all species are assumed to be present in both phases. The calculations must provide the mass fraction of each species as a whole and in each phase, the mass fraction of material which is liquid (liquid fraction,f:), the temperature, and the density; resulting in 3n+ 3 unknowns. The various mass fractions and the liquid fraction are related as follows species i, conservation of elements yields

i
i 1

vijYi = ~ vij(fYio+(1-f)Yi<);
i=1

j = 1,m.
(2.16)

In the gas phase, we have n-m-1 linearly independent chemical equilibrium reactions, since one equation is redundant due to Eq. (2.15), as follows v~j[-.//gj]~- ~ v~[oJgj];
j-1 j=l

i = 1, n-m-1.

This yields the following requirements for chemical equilibrium

~I F(~''~-~bl = K ~ " j=1

i = 1,n-m-1

(2.17)

where equilibrium is based on the fugacity of each species in order to conveniently treat non-ideal gases. Phase equilibrium requires that the fugacity of each species be the same in both phases Fg, = FT,, i = 1, n-1 (2.18)

since one of these equations is also redundant due to Eq. (2.15). Conservation of energy yields

(f : Yf, hf, + (1 - f :) Yghg,) = f h o + (1 - f )h~.


i=1

(2.19) The enthalpies in Eq. (2.19) are defined as follows

hp, = Ahp,+

dY~

f, COo' d T + h :

p =f

or

(2.20) where h , is the enthalpy of formation of species i as an ideal gas at T,, and Ahpi is the enthalpy deviation from the ideal gas state at T to the real state of the species in the flow. Finally, the density of the mixture is given by

p =
i=

r:;/pjl (1 -J))YofP,~

)1

(2.21)

Y~ = f s Y : , + ( 1 -f:)Yo,;
while by definition

i = 1,n

(2.14)

i=1

i=1

i=1

If we define v~j as the mass fraction of element j in

where P:i and Poi are the partial densities of species i in the two phases. Given an equation of state and thermodynamic property data, Eqs. (2.14 2.19) and (2.21) provide the 3n + 3 equations needed to solve for the composition, temperature and density of the mixture. The set of equations is nonlinear, and numerical solution is usually required. Examples of convenient package computer programs for their solution are CEC-7259 and TIGER. 6 The CEC-72 program is limited to the ideal gas equation of state, while TIGER allows the user to specify the equation of state and is more appropriate for use where real gas effects must be considered. Other solution methods which will be described as examples of state relationships are discussed in the following. Finding the state relationship for a nonreacting ideal gas mixture presents few difficulties, Shearer and Faeth 57 describe several examples of this type.

10

G.M. F E H AT
I I I

2000
v i,i

C M U TN P O A EJ _ T O B S I G R P N E_

~./

1.6

1600
II
,2o0

1.2

DENSITY

0.8 0.4
~

Q_

8 400 1.0' 0.8


Z

0.0

9
IO: h

0.6 0.4
^

_,,-co

-02

0.0

0.2

0.4 0.6 MIXTURE FRACTION

0.8

1.0

FIG. l. Scalar properties as a function of mixture fraction for a propane gas jet burningin air at atmospheric pressure. From Mao e t al. 27

Numerous results for combusting ideal gases are available in the literature. Lockwood and coworkers 36'37'42 complete the calculations while neglecting dissociation, while Bilger 35 and Mao e t al. 27 consider dissociation using CEC-7259 along with the JANAF thermochemical properties for the combustion gases. Figure 1 is an illustration of the state relationship for a propane jet burning in air at atmospheric pressure determined by Mao e t al. 27 The plot provides the temperature, density and mass fractions of major species as a function of mixture fraction. The nonlinear variation of properties with mixture fraction is evident. In this case, relating various properties by their mean values is questionable and the stochastic method is preferable. When a state relationship like Fig. 1 is constructed, some rather arbitrary decisions must be made concerning the extent of equilibration. Rates of equilibration are slow at low and high mixture fractions. For lack of other information, Fig. 1 was constructed so that equilibrium was maintained at all mixture fractions, but that soot does not form at temperatures

below 1000K. This approximation is not very critical at low mixture fractions, since results with and without dissociation are nearly identical. The approximation introduces errors in model predictions at high mixture fractions, however, which will be discussed later when measurements are compared with predictions for a combusting propane jet in air. The state relationship for a F r e o n - l l spray evaporating in air at atmospheric pressure is illustrated in Fig. 2, taken from Shearer e t al. 26 In this case, a twin-fluid injector was used, employing air as the atomizing gas, with both air and F r e o n - l l entering the injector at 300K. Since thermodynamic equilibrium is assumed at the exit of the injector, adiabatic saturation of air and Freon-11 in the injector causes the flow leaving the injector to have a temperature below 300K. The mass fraction of liquid Freon-ll decreases with decreasing f, reaching zero at f ~ 0.3. At this point, the mass fraction of F r e o n - l l vapor reaches a maximum while the temperature is a minimum. Density is a relatively nonlinear function of f d u e to the presence of liquid forf > 0.3.

Evaporation and combustion of sprays


3 0 0 [ ~~ ~
,

11

r\
~-E

10

1.0

oI

I
, ,

I
,

~ 0.6 U~
.,c

~"

0.4

0.2

0,0

0.0

0.2

0.4

0.6

0.8

1.0

FIG. 2. Scalar properties as a function of mixture fraction for a Freon-ll spray evaporating in air at atmospheric pressure. From Shearer et al. 26

The state relationship for a pressure atomized npentane spray, taken from Mao e t al., 2v is illustrated in Fig. 3. Conditions involve combustion in air at atmospheric pressure with all reactants initially at 300K. This state relationship was constructed assuming ideal gases and neglecting the solubility of nonfuel gases in the liquid phase, which is justified at atmospheric pressure. For these approximations the fugacity expressions of Eq. (2.18) are replaced by the vapor pressure relationship of the fuel. Dissociation was considered at low mixture fractions, using CEC-72. Dissociation was ignored at high mixture fractions, where the presence of liquid must be considered and temperatures are relatively low. The low mixture fraction region for the combusting spray, illustrated in Fig. 3, is qualitatively similar to the combusting gas jet, illustrated in Fig. 1. However, the presence of liquid significantly alters the high mixture fraction region, yielding properties more similar to the evaporating spray pictured in Fig. 2. The mixture temperature varies relatively slowly where liquid is present, due to heat of vaporization requirements. The mass fraction of fuel vapor reaches a maximum at the mixture fraction where the liquid disappears, f ~ 0.89. The mixture density is a minim u m near the stoichiometric mixture ratio, f ~ 0.06, but increases rapidly in the region where liquid is present. When the pressure of a spray process is increased, real-gas effects and the presence of dissolved gases in the liquid phase must be considered. The latter effect is illustrated in Fig. 4, taken from C a n a d a and Faeth. 61
i I i i i

2600

I000

~1800 bJ
n,*

~
0

0.1MPo

IO0 A TEM

PERATURE

IO >i-Z hi e",

UJ a.

m
W I--

IOOO
I

200 1.0 .8

PENTANE LIQUID'-~/1 PENTANE VAPOR'--~I/~ '~/Z~

z
~

.6

fVt
~
.8 /\ -1 //I
.2

"Fx G N 4 oY E ~ | /C R O / FAB N
0~

.4 .6 MIXTURE FRACTION

FIG. 3. Scalar properties as a function of mixture fraction for n-pentane spray burning in air at atmospheric pressure. From Mao et al. 27

12
'

G.M. FAETH momentum and the state of the fluid, based on the LHF approximation, must be specified. The simplicity for defining injector exit conditions is one of the main advantages of LHF models. For calculations of parabolic flows, 26 28,42 modified versions of GENMIX 63 have generally been employed due to the widespread availability of this routine. A number of package computer programs are also available for elliptic flows; the TEACH 64 computer program has frequently been used in the past. 2.3. C o m p a r i s o n with M e a s u r e m e n t s LHF models have been compared with several combusting and non-combusting two-phase flows. Test conditions for these comparisons are summarized in Table 5. The models were also calibrated using a wide range of single-phase flows, to fix the model constants. 2.3.1. N o n c o m b u s t i n g f l o w s In this section we shall examine the comparison between predictions and measurements for the model employed by Shearer et al. 26 This model involved solution of Eq. (2.5) with source terms given in Table 3. The flow was modeled using the Case 1 approximations of Table 4, which were satisfactory for all flows considered, z6 The model of Shearer et al. z6 was compared with a variety of single- and two-phase flows, as follows: (1)

FUEL N2 .2 .I z .06 .04 .02

H20 -/

;~0!; .....

g,
I~ .Ol
.006 DO4 .002

PROPANOL-I
LIQUID PHASE

zo

- - - GAS PHASE
,:o zo
PRESSURE (arm)

40 6o ,oo

[ I

FIG.4. Phase equilibrium concentrations for propanol-1 and n-heptane/air combustion at high pressures. From Canada and Faeth.61 This is a plot of equilibrium gas and liquid phase compositions as a function of pressure, for propanol-1 and n-heptane drops evaporating in the presence of their stoichiometric combustion products when burned in air. The concentration of dissolved gas increases with increasing pressure, reaching levels of 5-10% (molal) for a pressure of 4 M P a (roughly 40atm). At sufficiently high pressures the thermodynamic critical point is reached, where all properties of both phases are the same. Dissolved gas levels reach values of 50-60% (molal) at the thermodynamic critical point for the conditions illustrated in Fig. 4. Real-gas properties and dissolved-gas effects must be considered at operating conditions corresponding to combusting sprays in gas turbines and diesel engines. This is easily accomplished when constructing the state relationship for an LHF calculation. For example, Fig. 5 is an illustration of the state relationship for a pressure atomized n-pentane spray, burning in air at 3 MPa, computed by Mao et al. 28 Both reactants were initially at 300K. Liquid is absent and real-gas effects are small at low mixture fractions, therefore, CEC-72 results were employed for this portion of the plot. Real-gas effects were considered at high mixture fractions, using the Redlich-Kwong equation of state with multicomponent mixing rules developed by Prausnitz and Chueh. 62 The results are qualitatively similar to the combusting spray at atmospheric pressure, illustrated in Fig. 3. However, for f > 0.8, where liquid is present, solubility effects are evident from the nonlinear behavior of the mass fractions of gases and the liquid fuel. 2.2.3. N u m e r i c a l solution Boundary conditions for LHF calculations are specified in the same manner as gas flows. When liquid injectors must be treated, only the mass flow rate,

240o

000 200D

~ 160C D < 127 120D :~ LtJ Eaoo 4oc I.C .8 z o


I-.6

O0

%
>..
CO z Ld a

rv u_ co co

.4
.2

O.

,2

.6

.8

IO

MIXTURE

FRACTION

FIG. 5. Scalar properties as a function of mixture fraction for an n-pentane spray burning in air at 3MPa. From Mao et alfl 8

TABLE 5. Summary of test conditions used to evaluate L H F models a Combusting Mao et al. 2v Mao et al. z8 using data Shearer and Faeth 23 B(0.2 m m dia.) n-pentane kerosene C(0.25 mm dia. Khalil and Whitelaw 24 Combusting Combusting I Combusting II Khalil and Whitelaw 24

Spray type

Evaporating

Reference

Shearer et al. 26

Injector b n-pentane and air 80 350 1517 2000

A(1.194 mm dia.)

A(1.194 m m dia.)

C(0.50 mm dia.) kerosene


o

Injector fluid 225 1548

Freon-11 and air

Injector flow rate, mg/s Gas Liquid

2000

3
0.211 0.138 5.7, 8.7, 11.7 105 69 28, 79 e 3, 6, 9 5.0 11.6c 35 d 0.097 0.121 0.138
o-

Injector pressure, M Pa Gas Liquid

1.52

1.52

Jet momentum, mN Initial velocity, m/s SMD, #m

132 74.5 c 29 d

45 f 0.101

1 O0f

Ambient pressure, M Pa

0.097

0.101

a Reactants at room temperature. b Injector types: A, Spraying systems air atomizing with no swirl; B, Spraying systems pressure atomizing with no swirl; C, Pressure atomizing with swirl. c Estimated assuming LHF flow at the injector exit. d Measured at x / d = 170 in cold flow. e Estimated SMD and maximum diameter. f Estimated.

14

G..M. FAETH

O.e ~ _

ISOTHERMAL SINGLE-PHASE JET

0.6 0.4
0.0i t 1.0~

~ ~
THE VARIABLE DENSFFY SING_L E-PHASEJET

0.6

0.4 0.2
0,0

% A THEORY~~ ~ ' x ~ '\~~-..Ii~@o z~

0.8 O. 0.4 0.2 0.0 0.00 THEORY~~q EVAPORATING SPRAY"


i I J

SY OL x/d SOURCE 35 HETSRONI a SOKOLOV 40 WYGNANSKI e~FIEDLER ~:


~
, ~ ~ I __

5,oJSTUDY

O.OB

O. I 6 r/x

0.24

0.32

FIG. 6. Radial profiles of mean axial velocity for various single- and two-phase noncombusting jets. From Shearer et al. 26

constant density single-phase jet (data from Shearer et al. 26 Wygnanski and Fiedler,65 Hetsroni and Sokolov, 66 and Becker et a/.67); (2) variable density single-phase jet (data from Shearer et al. 26 and Corrsin and Uberoi68); (3) air jet into water, which yields a bubbly flow that is representative of typical applications of L H F models 16.17 (data from Tross69); and (4) evaporating Freon-11 spray in air (data from Shearer, et a/.26). Predicted and measured radial profiles of mean axial velocity, mean mixture fraction and Reynolds stress are illustrated in Figs. 6-8. These figures include results for constant and variable density single-phase jets and for an evaporating spray. The dimensionless radial variable, r / x , is employed on the figures so that capabilities for predicting the flow width can be

evaluated. The model predicts similarity in these coordinates over the range of data, therefore, only a single theoretical curve is shown in each case. The various experiments and the theoretical results are all in reasonably good agreement. Even though a wide variety of jets are considered, they are remarkably similar when plotted in this manner. The flows are more distinguishable when the axial direction is considered. Mean axial velocities and mixture fractions, along the centerline, are illustrated in Figs. 9 and 10. The results tend to separate according to the initial density ratio of the flow P o / P ~ with development of the flow being more rapid for low values of this parameter (the range shown on the figure goes from 0.0012 for the air jet in water to 6.88 for the evaporating spray in air).

Evaporation and combustion of sprays 1.0 0.8 0.6 0.4 0.2 0.0
,.o
_ SITY SI NGL.E-PHASE a ET -

15

, ,
, 1

SYMBOL' x / j ' SOURCE ~" 28 8ECKER~al


oo 170

"~,

Ul ,',A

540 i. STUD Y 510J

]p RESENT

'S?~ ~;ET H##AMs


T H E O R Y ~ N

0.8 0.6 0.4 0.2

0.0

i.d
0.8 0.6 0.4 0.2 0.0 0.00
I

PORATING

SPRAY

THEORY-J
I I

NO~u
I I I I

0.08

O. 16 r/x

0.24

0.32

FIG. 7. Radial profiles of mean mixture fraction for various single- and two-phase noncombusting jets. From Shearer et al. 26

Results in Figs. 9 and 10 indicate that the model provides a reasonably good prediction of the axial variation of mean quantities for the single-phase flows; however, there are greater discrepancies between predictions and measurements for the twophase flows. The velocity predictions for the air jet into water are reasonably good; however, the predicted value~ of mean mixture fraction overestimate the measurements. This is disappointing since bubbles have low inertia in water for these test conditions which should favor the use of the L H F approximation. Difficulties with both theory and experiment could be contributing to the error. This flow has the greatest density variation of all the flows considered by a wide margin. Therefore, the variable density

approximations of the Reynolds averaged model are particularly questionable in this case. Another factor is that the probe used by Tross 69 tends to underestimate the void fraction, and thus the mixture fraction, due to surface tension effects. Additional theoretical and experimental consideration of this flow would be desirable. For the spray data illustrated in Figs. 9 and 10, predicted centerline velocities are 10-20 % lower than the measurements, while predicted centerline mixture fractions are 40 % below the measurements. This overestimation of the rate of development of the spray was also observed for other spray properties. For example, theory indicates that liquid should disappear on the centerline near x / d = 165, while the measurements

16
0 0 2 I 'Z~,

G.M. FAETH
I - I I [ I

~ o
~o''~ av ~

ISOTHERMAL SINGLE~E-.PHASEE JEt

,~

<] ~,~o~" ~THEORY

0.02 ~

1"

oJ o

a ~ao o o "~u
oo,
~

/%
x A

VARIABLE.DENSITY

SINGLETpHASE JET

RY

0.00 0.02

! \ /~/ ; o O ~
0.01 / I
/
,

SYMBOL x/d SOURCE ,: 501WYGNANSK, ~ 60[&FIEDLER ",o r,~


I

a~

o 170] PRESENT [] 340)' ST


EVAPORATING SPRAY
I I -I

THEORY/
I I I

0.00 0.00

0.08

0.16
r/x

0.24

0.32

FIG. 8. Radial profiles of Reynolds stress for various single- and two-phase noncombusting jets. From Shearer et al. 16 showed that some liquid was still present at x/d = 510. 26 The temperature measurements also approached ambient conditions more slowly than predicted. The density ratio of the spray was similar to the variable density jet, which was predicted quite well; therefore, variable density effects are not a prime source of error. Finite interphase transport rates are the major difficulty as will be shown later by means of drop-life-history predictions in this flow. The model of Mao et al. 27'28 is identical to that used by Shearer e t al. 26 including the same empirical constants. The only distinction between the computations involved changes in the state relationship in order to treat the various experimental conditions. The flows considered include: (1) n-propane gas jet burning in air at atmospheric pressure (data from Mao et al.ZV), (2) air-atomized n-pentane spray burning in air at atmospheric pressure (data from Mao et al.27), and (3) pressure atomized n-pentane spray burning in air at pressures of 3-9 MPa (data from Shearer and Faeth23). In all cases, the surroundings were stagnant and the flames were attached at the exit of the injector. The major properties of the two-phase flows are summarized in Table 5. Original references should be consulted for specific details concerning test conditions and instrumentation.

2.3.2. Combusting flows Evaluation of LHF predictions for combusting flows have been undertaken by Mao e t a / . 2v'28 and Khalil and Whitelaw. 24 Both these studies were preceded by extensive development of the turbulence model for single-phase combusting flows. 36 42

Evaporation and combustion of sprays


I I I I I l l l I I I I [ [ I I I I

17
I
I I

THEORY i ii ii iii iv v I.O

SYMBOL o ~ v P z~ D

TYPES OF FLOW HEATED AIR JET ISOTHERMAL AIR JET ISOTHERMAL AIR JET VARIABLE DENSITY JET EVAPORATING SPRAY AIR-WATER JET

SOURCE CORRSIN PRESENT STUDY WYGNANSKI ~ FIEDLER PRESENT STUDY PRESENT STUDY TROSS

,# ,#
A

o.,l

+\+

NNV 0 I
O,OI I I t I I i I II I I ]

V
I I I I ;I

,i ii
I I I I

I0 x/d

I00

FIG.9. Axialvariation of mean centerline axial velocity for various single- and two-phase noncombusting jets. From Shearer et al. 26

The results for the combusting gas jet were employed to examine the capabilities of the model with a combusting single-phase flow having the same general geometry as the sprays. The state relationship for these computations is illustrated in Fig. 1. Predicted and measured mean axial velocities and temperatures along the centerline are illustrated in Fig. 11. The comparison between predictions and measurements is excellent. The predictions of temperature are unusually good, Lockwood and coworkers a7,42 normally find maximum mean temperature errors of 100-200K with this model when dissociation is neglected, while Kent and Bilger+s observe similar errors with a Favre-averaged model which allows for dissociation. Such errors are probably more representative of models of this type. The radial variation of predicted and measured mean axial velocities and Reynolds stress for the n-propane flame are illustrated in Figs. 12 and 13. Due to large density variations in this flow, velocity and Reynolds stress profiles are not similar at the stations shown, in contrast to the results for noncombusting jets illustrated in Figs. 6 and 8. In general, predictions are in good agreement with measurements. Similar results for mean temperatures are illustrated
JPECS 9-I/2-B

in Fig. 14. Good agreement between theory and experiment is observed. At axial locations prior to x / d = 200, the flame zone is located off-axis and has a maximum temperature somewhat below the maximum value along the axis. All flame temperatures are lower than the adiabatic flame temperature for stoichiometric combustion in air of 2270K, c.f. Fig. 1, due to turbulent unmixedness. This effect is represented in the calculations by the probability density function, the mean square of the mixture fraction fluctuations, the mean mixture fraction and the temperature estimation of the state relationship (particularly near the stoichiometric condition). The results illustrated in Fig. 14 indicate that this aspect of the model is effective. Predicted and measured species concentrations in the propane flame are illustrated in Fig. 15. Predictions of fuel, oxygen and nitrogen concentrations are generally quite good. The least satisfactory aspect of the model involves predictions of product concentrations, particularly in the high mixture fraction portion of the flow. This region corresponds to mixture fractions where rather arbitrary decisions concerning the degree of equilibration had to be made during construction of the state relationships and this is probably

18
i i r i i i

G.M. FAETH
i i i i T I I [ I d I I I I I I

el
i ii iii
v

THEORY SYMBOL o <1 A []

TYPES OF FLOW HEATED AIR JET ISOTHERMAL AIR JET VARIABLE DENSITY JET EVAPORATING SPRAY AIR-WATER JET

SOURCE CORRSlN BECKER PRESENT STUDY PRESENT STUDY TROSS

iv

O.I ,o

O.OI
D 0

.~

OD

O.OOI

I I II

Iqll

VI

I0 x/d

I00

Fig. 10. Axial variation of mean centerline mixture fraction for various single- and two-phase noncombustingjets. From Shearer et al. 26

the major source of error. Kent and Bilger48 are more successful with predictions of product species concentrations using a similar model for combusting hydrogen jets where the high reactivity of hydrogen favors the assumption of complete chemical equilibrium when constructing the state relationships. The results of Fig. 15 suggest that hydrocarbons have more significant chemical kinetic effects, particularly in high mixture fraction regions. Since the model gave reasonably good results for gas flames the next step in its evaluation involved consideration of a combusting air atomized n-pentane spray at atmospheric pressure. 27 The presence of air in the primary injector flow implies a region of premixed combustion near the injector exit. However, the primary air flow was far below stoichiometric requirements, c.f. Table 3, and this was handled by assuming that primary air and fuel react immediately to maintain equilibrium. This approximation could be

avoided by employing a Case 2 model (Table 4), but such calculations have not been reported as yet. The state relationship for these computations corresponds to Fig. 3, except that the mixture fraction scale must be shifted to account for the presence of air in the flow leaving the injector, c.f. Ref. 27. Predicted and measured mean axial velocities and temperatures along the axis of the n-pentane spray are illustrated in Fig. 16. In contrast to the combusting gas jet, where predictions and measurements agree, the spray is developing more slowly than predicted. Measured velocities are higher than predicted near the end of the potential core. The maximum temperature position is estimated to be x/d = 215, while the measured position is x/d = 275, about 30% farther from the injector. Good agreement between predictions and measurements is only achieved somewhat downstream of the maximum temperature location, e.g. x/d > 300, which is beyond the region where

Evaporation and combustion of sprays


I0.0
COMBUSTING PROPANE JET DATA SYMBOL CENTERLINE DATA o 1 Ir.MPP..RA ~ o VELOCITY - THEORY

19

8OO

1.0 I . , ~

,4oo--.
Ld

S
io

uJ I0.I

iooo

0.01

I0 x/d

I00

I000

FIG. 11. Axialvariation of mean axial velocity and temperature for n-propane gas jet burning in air at atmospheric pressure. From Mao et al. 27

average position of the edge of the spray determined from shadowgraphs. The predicted spray boundary was taken as the point where the mean liquid fraction reaches zero. The state relationships used in these computations were similar to Fig. 4, but varied with pressure. Real-gas effects and the solubility of nonfuel gases in the liquid phase were considered. The comparison between predicted and measured spray boundaries for n-pentane sprays burning at high pressure is illustrated in Fig. 20. Both predictions and measurements indicate that the extent of the spray boundary is reduced as the ambient pressure is increased, however the theory overestimates the magnitude of the reduction. Predicted spray lengths are 10-20 ~o less than the measurements while greater errors are observed for the radial limits of the spray. The comparison between predictions and measurements illustrated in Fig. 20 is encouraging, but a more detailed evaluation of the model at high pressures is needed. It is unlikely that regions of low drop density were completely resolved on the shadowgraphs and at high pressures, large density gradients in gases can be misinterpreted as regions of spray. It seems reasonable to suppose, however, that drops were present outside the predicted spray boundary and that the LHF model overestimated the rate of development of high pressure sprays, similar to the findings observed for sprays at atmospheric pressure.

drops are present. This is expected, since the LHF model can treat gas flows in this geometry accurately and errors incurred in two-phase flow regions eventually decay away. Predicted and measured radial profiles of axial velocity and Reynolds stress are illustrated in Figs. 17 and 18. The variables on these figures are normalized in the same manner as the results for the combusting gas jet. When plotted in this manner, there is good agreement between predictions and measurements, however, it should be recalled that the centerline velocities used to normalize these results are not predicted very accurately for x/d < 300. Predicted and measured radial profiles of mean temperatures are illustrated in Fig. 19. Predictions of flow width are satisfactory and the complete profiles are predicted quite well for x/d > 300. However, there are large discrepancies between predictions and measurements near the injector, where overestimation of the rate of development of the flow is most pronounced. Measurements of species concentrations indicated similar problems with overestimation of the rate of development of the flow, in addition to the equilibrium difficulties discussed earlier. 27 Concentration predictions only became comparable to results for the gas flame for x/d > 300, which is beyond the two-phase flow region. The model was compared next with experimental results for a pressure atomized n-pentane spray burning in air at pressures of 3, 6 and 9 MPa. 28 The only measurements available to test these predictions were spray boundaries, measured by Shearer and Faeth. 2s The measured spray boundaries were the

o st

co.s,.G
x/d=540

JET 1

0.4 o0'2 0.8

"~o

~.o

"" I
o o o

O.Z_: ~
1.0 0.8 0.6

0.4 0.2 0.0


0.04

o
o o

x/d=74.5

0.08

0.12 r/x

0.16

0.20

FIG. 12. Radial variation of mean axial velocity for an n-propane gas jet burning in air at atmospheric pressure. From Mao et al. 27

20

G.M. FAETH

0.020

'

' COM;USTINGPROPANE dET


-

T______HEORY

0.010

x/d=510 o o

0.00 , 0.020

0.010 I#

40

I%o.oo.
0.020 x 0.010 / d = l ~

0.00' 0.020 QOIO / d= .5

/
0.00 ~
i

0.04

0.08

0.12
r/x

'

'

0.16

-0.20

FIG. 13. Radial variation of Reynolds stress for n-propane gas jet burning in air at atmospheric pressure. From Mao et al. 27 Mao e t al. 28 also report predictions of the structure of the combusting n-pentane sprays at high pressure. The predicted variation of mean quantities along the spray axis at a pressure of 3 MPa is illustrated in Fig. 21. Mean liquid fraction, mixture fraction and velocity decrease monotonically with increasing distance from the injector. The concentration of fuel vapor reaches a maximum at the position where the liquid disappears. The mean temperature reaches a maximum well downstream of the end of the two-phase region, where the fuel and oxygen concentrations overlap. This behavior generally corresponds to trends observable in the state relationship, c.f. Fig. 4, where liquid is confined to high mixture fraction regions and maximum temperatures occur at low mixture fractions which coincide with low concentrations of both fuel and oxygen. Within the flow, turbulent fluctuations modify the picture given by the state relationship somewhat, however, the general properties are similar. Figure 22 is an illustration of radial profiles of mean and turbulent quantities in a combusting spray at 3 MPa. Three axial locations are shown: before, near and beyond the position where the temperature reaches a maximum along the centerline. Liquid is confined to a cool region along the centerline where little oxygen is present. Fuel vapor extends beyond the

dET 1800 d 5

1400
w

/d
w a.

,,, I 0 0 0 I-

60(~

0.00

0.04

0.08

012

0.16

0.20

rlx FIG. 14. Radial variation of mean temperature for an npropane gasjet burning in air at atmospheric pressure.From
Mao et
al. 27

Evaporation and combustion of sprays


DATA SYMBOL SPECIES o o2
O

21

0,60

COMBUSTING PROPANE JET x/d = 74.5

, COMBUSTING

, , PROPANE JET

, '~AT ^~ ~.. SYMBOL SPECIES


o 0
c, v

0.50
z 0.40
O '~ [3 0 0

o
o -[3 [3 O "~

N2 x I/2 C02 CO

0.50

x/d=170

N2 11/2
co~ co

02

C3H s
H2 THEORY O O 1/2N 2 h03 O3

H20

z 0.40 o
Io <~ ~

0 D --

HEO C~H a THEORY

o
O

o
" ~ ' - - 1 / 2 Na

[3
/

/----------'~

0.50

0.30

02

m 0.20

/
0.1C

o,o~-

\,,

o-,~

fo2
0.0( , ~'-.-~o o

co---

"~5..f 9.,7~--.._ "o...~ /--H~O 0.08 0,12


r/x

0.006 e c " ~ o ~ ~ 0,00 0.04 0.08 0.12


r/x

0.00

0.04

0.16

0.20

0,16

0.20

COMBUSTING

PR'OPANE JET '

'

DATA'
SPECIES

' 02

DATA'
SPECIES

o5o~
z o 0 <
h

:":~

SYMBOL o 0

.cOMBUSTI N_.GG PRIOPANE JET ' 0,501 x/d=510

SYMBOL

o
D A O

o~
N2xl/2 C02 CO H20 THEORY O Q I/EN 2

O --

N2 x 112 CO 2 CO H20 THEORY n

0.40

~ ~ k ~ _ l / 2 N2

Z 0 )0 u.

0.40
D u

re

0.50

0.30

(,9 69

F
0.20

Oz

(n

0.20

~
o % o
" Z
/--CO2

0.10

O.lO o
O
0.08

)FH2
0
0.12

~
0.04

"
0.20

0.00' 0.00

0.04

0.08

0.12 r/w

0.16

0.20

0,0 0.00

0.16

r/x

FIG. 15. Radial variation of mean species concentrations for an n-propane gas jet burning in air at atmospheric pressure. From Mao et a l ? 7 liquid boundary and contacts oxygen in the high temperature regions of the flow. Thus the general nature of the process is similar to a gaseous diffusion flame, except that liquid evaporation provides an extended source of fuel. Concentration fluctuations are highest in the flame zone, near the injector. Larger concentration fluctuations tend to reduce maximum temperature levels in the flow, when the mean mixture fraction is near the stoichiometric value. The behavior predicted in Figs. 21 and 22 is qualitatively similar to existing measurements within combusting sprays which have n o s w i r l . 6'14'19'27 Khalil and Whitelaw 24 compared their predictions with measurements on an open combusting kerosene spray with swirl, e.f. Table 5, for test conditions. The liquid flow from the injector was concentric with a shroud flow of air. The injector tended to have a hollow cone pattern. The model involved solution of Eq. (2.1), with source terms from Table 2, under the Case 1 approximations of Table 4. Construction of the state relationship, however, was considerably simplified in comparison to the results discussed earlier for two-phase flows. Rather than determining the properties of the equilibrium gas-liquid mixture at high mixture fractions, the flow was simply replaced by a gaseous kerosene flow at the injector exit having the same mass velocity and energy content as the spray. This approximation underestimates the density of the flow leaving the injector by a wide margin. Examination of the effect of initial density ratio on initial flow development, c.f. Figs. 9 and 10, indicates that this approach would tend to overestimate the initial rate of development of the flow in comparison to the correct state relationship for the LHF approximation. Prior to examining the combusting sprays, Khalil and Whitelaw 24 tested their model for combusting gas flows, using the measurements of Bilger and Kent 7 for a hydrogen diffusion flame in a coflowing stream of air. The model yielded predictions of mean velocities

22
I0.0

G.M. F~TH
t

COMBUSTING N-PENTANEJET DATA SYMBOL CENTERLINEDATA o TEMPERATURE o VELOCITY


-THEORY

0.020

COMBUSTING N-PENTANE JET o DATA THEORY


x/d=510

1800

0.010
0.000 o o

1.0

o o

14oo

0'020I 0 . 0 1 0 ~ x/d=340

io

0.1

iooo

1%0.000~> 0.020 0.01

"".-o o o

x/d=170

OO 0.01
I0 100

0.000.~

"---o

I000

0"020I
x/d =74.5

I I

x/d FIG. 16. Axial variation of mean axial velocity and temperature for an n-pentane spray burning in air at atmospheric pressure. F r o m Mao e t al. 27

O . O l O ~
O,O 0 0 J ~
I I I ~ I ~ l

0.00

0,04

0.08

0.12 r/x

0.16

0.20

'T~i ' coM~sz,~& 1 PROPANE


dET

FIG. 18. Radial variation of Reynolds stress for an n-pentane spray burning in air at atmospheric pressure. From Mao e t
al. 27

o.6r " o z~
0.8-

. "~o
"~

T~E~
COMBUSTINGN-PENTANEJET DATA SYMBOL x/d o 165 o 170 v
1400 340

,o~.~.

\o

--~oo 1

1800

510

0.2

b1.0
0.8 0.6 0.4

I000' ~. THEORY x/d 74.5 I0

x/d=74.5

600

0.2 0.0 0.04 0.08 0.12 r/x 0.16 0.20

0.00

0.04

0.08

QI2 r/x

0.16

0.20

FIG. 17. Radial variation of mean axial velocity for an npentane spray burning in air at atmospheric pressure. From Mao e t al. 27

FIG. 19. Radial variation of mean temperature for a combusting n-pentane spray burning in air at atmospheric pressure. From Mao e t al. 27

Evaporation and combustion of sprays I


9MPo

23

......

40

--

MEASURED PREDICTED

-"
6MPa

~a

"~ 4D

40 t

3MPa

o L.__
0

40

80
x/d

120

160

FIG. 20. Predicted and measured spray boundaries for a pressure atomized n-pentane spray burning in high pressure air. From MaD et al. ~8

within 5~o and mean temperatures within 200K of measurements along the centerline of the flow. This level of accuracy is typical of turbulent combustion models of boundary layer flows. Notably, Kent and Bilger4s report predictions of similar accuracy for this flow, using Favre averaging. Predicted and measured mean axial velocities and temperatures along the centerline of Flame I (Table 5) are illustrated in Fig. 23. As expected, the LHF model predicts that the maximum temperature location along the centerline is nearer to the injector than measured. The effect of drop size on the comparison between predictions and" measurements can be
i i i

examined using the radial profiles of mean temperature illustrated in Fig. 24. The results for Flame I (having an SMD of 45#m) are much better approximated by the theory than the results for Flame II (having an SMD of 100#m). Therefore, while the measurements tend to approach the LHF predictions as the SMD of the spray becomes smaller, the agreement is still relatively poor for a spray having an SMD of 45 #m. Presumably, use of the correct state relationship for the high mixture fraction region would improve the predictions somewhat, since the higher liquid density would tend to retard the development of the flow, however, such predictions have not been reported for these flames. In a later study, Khali125 was able to achieve better predictions for a portion of the data illustrated in Figs. 23 and 24. This was accomplished by introducing an eddy-break-up model for the reaction of fuel, which included additional empirical parameters to account for heterogeneous combustion of drops. While such an approach tends to slow the rate of development of the flow, allowing a better match of predictions and measurements in some instances, it is difficult to generalize the method. Furthermore, this approach cannot correct for the effect of relative velocity between the drops and gas, which has an important influence on the dynamics of the flow. Therefore, the usefulness of the extension is limited and it has not been pursued in subsequent studies. 2.4. S u m m a r y The previous comparison between predictions and measurements indicates that the LHF models provide a useful qualitative description of both noncombusting and combusting sprays. LHF models are relat i i i |

1.0

~o/~o~~O
,/

N-PENTANE IN AIR L-~3MPo

.'~''~A--

_j

2000

"'~

.,

/ !j,'" ; ,"
-i I /! !
I I I I ! I /

1500

laJ I:E ::3 I'r,,.' hi

--

.~;

/
/

"1000

I 4

l/
', /
"0110 20
L I

5oo
I

4i0

60 80 I00 x/d

II

200

400 600

I000

FIG.21. Predictions of mean quantities along the centerline of a pressure atomized n-pentane spray burning in air at 3 MPa. From Mao et al. 2s

24

G.M. FAETH

1.2

x/d=50

5/5c--~
g/4g c

/-r
\

1600 1200 8OO


i 400

.4

I0 u-'T~qvli 2~ / / U c
, ,_ 1 I ~t~'//''~''~

>)<X'c
_ ~'p

1.2

x/d=350 U/~c

2000

1600 t.u r,,YN 2

.4

g/4 g I0 uivi/ue 2
t ~

\
? ~
i ~

1200 < 12:


t,i n 1-

1--

YOz
J

800 40O 2000 i600

,,,

8t

U/Uc ~

__

9N2

1200 8O0

'41

g / 4 g ~
1

.24

.16

08

0 r/x

.08

.16

.24

4OO

FIG. 22. Predicted radial variation of mean quantities for a pressure atomized n-pentane spray burning in air at 3 MPa. From Mao e t al. 28

200O

U(m/s}
=o

" Theory ...Measurement

s " -- "J~-~

/ /

lO \

/.,7,,,

///
D = 200 mm I 5 X/D I 10

1000

FIG. 23. Predicted and measured mean velocities and temperatures along the centerline of a kerosene spray burning in air at atmospheric pressure. From Khalil and Whitelaw. 24 tively easy to use since they only require general i n f o r m a t i o n concerning injector characteristics a n d they draw heavily on existing models of single-phase flow processes. F o r test conditions considered thus far, the L H F model invariably overestimates the rate of developm e n t of the flow ( S M D = 28-100/~m, pressures of 0 . 1 - 9 M P a ) . The results of Khalil a n d Whitelaw 24

Evaporation and combustion of sprays


2000
X/D = 3.64

25

:ooo! ' ~ f
/~ \

,o~.,~
EXPERIMENT

THEORY A 1500

~/
Ld

\ F 'oo

U.I I--

,?!2
5OO

I000

5OO

i
r/D

o
r/D

FIG. 24. Predicted and measured radial profiles of mean temperature for kerosene sprays burning in air at atmospheric pressure. From Khalil and Whitelaw.24 show that the L H F approximation is improved as the SMD of the spray is decreased. This behavior suggests that finite interphase transport rates are important for sprays having SMD as small as 28 #m. The appropriate size range for L H F computations can be established more quantitatively by means of drop-lifehistory computations. Such computations will be considered in Section 3. In situations where radiation, premixed reaction, kinetic energy, etc. can be ignored (corresponding to the Case I assumptions of Table 4) the k-e-g statistical approach provides a useful method of analysis. This method requires no parameters in addition to those needed for single-phase noncombusting flows. Determination of the state relationships requires conventional adiabatic mixing or adiabatic flame computations, which must be completed only once for each operating condition. Therefore, little prior experience is required to carry out computations, computation times are minimized, and complex phenomena--such as high pressure phase equilibrium and dissociation can be considered in a relatively routine manner. When premixed combustion must be considered, the joint probability density function method developed by Lockwood s6 provides an attractive extension of the Case 1 statistical model. Turbulent reaction rate parameters must be specified for this model, however, guidance is available from past studies of premixed single-phase reaction processes.36'45'5 3 This method requires construction of a second state relationship for unreacted conditions, however, only an adiabatic mixing computation is involved. Evaluation of this approach for premixed spray combustion processes has not received much attention and more work in this area is warranted. The general case where the L H F approximation must be applied to a spray process involving radiation, premixed combustion, kinetic energy, etc., naturally involves more empirical constants. Finding scalar properties using the statistical method becomes increasingly unwieldy in this case, since multidimensional joint probability density functions must be specified. Therefore, most investigators simply employ mean properties directly in state relationships, ignoring the effect of fluctuations. Models of this complexity are difficult to evaluate and the reliability of these procedures for LHF model computations is not yet established.
3. CURRENT MODELS OF SINGLE DROP BEHAVIOR

3.1. Introduction The behavior of individual drops in a spray must be examined in order to assess the validity of L H F models and to undertake SF models. This involves drop-life-history computations to yield the size, velocity, temperature and composition of individual drops as a function of position in the flow. In this section, a basic model of drop processes is described which employs features most frequently adopted for SF models of sprays and have been subjected to experimental evaluation. Less commonly treated effects, e.g. dense spray phenomena, drop ignition, drop dispersion, etc., will be discussed later. The simplified drop model will be employed to evaluate the L H F approximation for sprays. This study will be conducted by computing drop-lifehistories, employing the results of L H F computations to estimate the local ambient conditions of the drops within the spray. Computations of this type provide a means of determining the range of drop sizes where interphase transport rates are sufficiently fast for the L H F approximation to apply. The results also yield insight concerning drop processes in sprays, providing a useful introduction to the more complex phenomena described in Sections 4 6. The discussion begins with a description of the simplified model and evaluation of its predictions by comparison with measurements made under well-controlled conditions. Drop-life-history computations in

26

G.M. FAETH relative velocities of 10-100m/s). In contrast, drop lifetimes in a spray are on the order of 1t0ms, two to three orders of magnitude greater, justifying the quasisteady flow assumption. For motionless drops, the second characteristic time is controlling, indicating flow development times comparable to drop lifetimes at high pressures (10-100atm.). However, motionless drops are rarely of interest for practical sprays, preserving the utility of the quasisteady flow assumption. The radial velocity of the liquid surface due to the evaporation of liquid is neglected. This assumption is related to assumption 3. The approximation is valid for a moving drop, as long as the liquid surface velocity is small in comparison to the relative velocity of the drop. This is generally the case, except when the liquid has uniform properties and is very near the thermodynamic critical point. Such conditions are rarely encountered and normally indicate the point where the drop is no longer treated as a separate entity in any event. Surface regression velocities are more important for motionless drops and the assumption should be reexamined when such cases are of interest. 72 Effects of drag and forced convection are represented by empirical correlations. This is necessary since accurate treatment of flow around spheres is impractical due to excessive computation requirements. Gas phase transport is based on mean ambient properties and the effect of turbulent fluctuations is ignored. This assumption is adequate as long as the intensity of fluctuations is small and all fluctuating parameters needed to compute transport rates vary linearly. Sprays only approximate these requirements. For example, mixture fraction is the primary fluctuating quantity in a combusting spray while temperature, which is a prime variable for drop heat transfer, varies nonlinearly with f near the stoichiometric mixture ratio, c.f. Figs. 3 and 5. Since the practice is widespread, however, mean properties will be employed for the present and the assumption will be reexamined in Section 5. The liquid surface is assumed to be in thermodynamic equilibrium with negligible temperature jump due to finite rates of evaporation. Furthermore, the effect of surface tension is neglected when determining phase equilibrium at the liquid surface. These assumptions are generally satisfactory for spray analysis, e.g. atmospheric pressure and above for drops having diameters greater than 1 ~tm.6 Bellan and Summerfield 73 present extensive results concerning surface equilibrium properties of small drops. The pressure is assumed to be constant and

sprays are then discussed for the test conditions examined in Section 2. These results involve both evaporating and combusting sprays, at pressures extending to the region where near-critical phenomena must be considered; therefore, these features will be incorporated in the drop-life-history model. 3.2. Theory 3.2.1. Description of the drop model In general, drop models must be approximate in order to control computation times, particularly for SF spray analysis. In the following, a simplified approach, which incorporates most features of interest, will be described. Earlier reviews also provide extensive discussion of single drop models, w~5 The analysis must consider effects of the relative motion of the drop with respect to the gas and the fact that the entire process is transient, i.e. liquid and ambient conditions vary through the lifetime of a drop in a spray. Most practical fuels are blends and dissolved gases are important at high pressures. Therefore, practical drop models must consider multicomponent liquid phases. These basic requirements can still be satisfied while making the following assumptions, typical of most drop models :~5 (1) The drop is assumed to be spherical. It is wellknown that moving drops deform; however, existing correlations for drag and convection incorporate this effect implicitly, treating the drop as an equivalent sphere. 7'71 (2) The spray is assumed to be dilute. Under this assumption drop collisions are ignored and the effect of adjacent drops on drop transport rates are neglected, i.e. drag and convection correlations for drops having infinite spacing are employed without correction. The general limitations of this assumption are not welldefined at present; however, moving monodisperse particles can be treated in this manner if the center-to-center distance between particles is greater than two particle diameters. 17 For equal spacing, this requirement implies a void fraction greater than roughly 90 ~o, depending upon the configuration of the drop array. The presence of dense sprays also influences the gas phase governing equations for T P F models. This aspect of the dense spray problem will be discussed further in Section 6. (3) The flow around the drop is assumed to be quasisteady, i.e., the flow immediately adjusts to the local boundary conditions and drop size at each instant of time. The appropriate characteristic times for development of the gas phase flow field are either dp/lup-~1 or d2/(ctt or D), whichever is smaller. The first parameter is usually controlling for moving drops, yielding characteristic times on the order of 0.1 10ps (drop diameters of 10-100#m,

(4)

(5)

(6)

(7)

(8)

Evaporation and combustion of sprays equal to the local mean ambient pressure. This approximation is also satisfactory, except for small drops at low pressure. 6 Only concentration diffusion is considered, neglecting thermal diffusion, and the Dufour effect is neglected in the heat flux equation. Errors incurred by these approximations are on the order of 10 %; therefore, complicating the model by including these phenomena is rarely warranted. 6 Radiation between the drop and its surroundings is neglected. Convective heat transfer rates of drops in sprays are high since the particles are small, which reduces the importance of radiation. Another factor helping to justify this approximation is that gaseous radiation bands are generally not coincident with absorption bands of most liquid fuels. 6'74 Radiation is more important when there is significant continuum radiation from hot surfaces and soot, or where there are absorbing particles present in the liquid, "e.g. coal slurries. When radiation must be considered, however, the extension of the model is straightforward. Oxidation and decomposition processes are neglected in the flow field surrounding the drop. A drop in an oxidizing atmosphere can be ignited with an oxidizing zone completely surrounding the drop (envelope flame); ignited, but with the oxidizing zone stabilized in the wake (wake flame); or not ignited, with reaction being completed in the bulk gas phase. Drop transport rates for the latter two cases are accurately represented by the present approximation, 75-77 however, the presence of an envelope flame enhances transport rates. A related phenomena involves decomposition of the fuel vapor, even in the absence of an oxidizing atmosphere. If the drop is a monopropellant liquid, e.g. a hydrazine or nitrate ester, transport rates are enhanced, while decomposition of nonmonopropellants results in reduced transport rates. 6'78 The importance of reaction effects can be characterized by the Damk6hler number, which is the ratio of a characteristic residence time to a characteristic chemical reaction time, e.g. Da = dp/(rch]u p fil) for a moving drop. The L H F predictions discussed in Section 2, indicated that liquid is generally confined to cool portions of the flow, tending to yield relatively large "Cch and low Damk6hler numbers. Therefore, assuming negligible reaction in the drop flow field is a reasonable first approximation. Further discussion of reaction near drops will be deferred until Section 5. The gas phase Lewis number is often assumed to be unity in drop models, 6 however, this approximation is not adopted here. Predictions of drop behavior at high pressures are

27

(9)

(10)

known to be influenced by Lewis number values. 61 Furthermore, the Lewis number of high molecular weight materials, often encountered in analysis of sprays, departs significantly from unity. Therefore, a conservative approach with respect to the unity Lewis number assumption appears to be warranted. (13) The properties of the gas flow field are assumed to be constant at each instant of time. An effective binary diffusivity, specific heat and molecular weight are used for all species. Properties are evaluated at an average condition, defined as follows:

~avg = ~q~sg'}- (1--~)~

(3.1)

(11)

where q5 represents both species mass fractions and temperature. 2w28 The empirical factor ~ is selected to obtain agreement between model predictions and existing measurements, wherever possible. Since property variations are large, particularly in combusting sprays, calibration of c~ in this manner reduces uncertainties of drop computations pending development of reliable methods for treating variable property effects for flow around drops. 6 Related to the selection of c~ are the property correlations for individual species and the mixing rules for the properties of mixtures which are used in the computations. Complete details for the results to be considered in this section are presented in Refs. 57 and 58. Original sources should be consulted for the SF models to be considered later since methods tend to vary. (14) Chemical reaction is neglected in the liquid phase. This approximation is analogous to the low Damk6hler number assumption of the gas phase, except that a more appropriate residence time is the drop lifetime. Drop lifetimes are longer than gas phase residence times; however, liquid temperatures are below boiling or critical temperatures and these low values generally yield low Damk6hler numbers. Exceptional cases can be found, e.g., liquid reactions were o0served in heavy oil drops by Masdin and Thring. 79 However, the approximation is acceptable for most cases where spray models have been evaluated to date. (15) Transport processes within the drop are treated by considering the three limiting cases, which bound actual behavior, summarized in Table 6. Transport processes within the liquid phase have characteristic times, d~,/(~j, or DI) and dp/ucirc, which are comparable to drop lifetimes, since liquid diffusivities and circulation velocities are low. Therefore, a quasisteady approximation analogous to the gas phase is not appropriate. Sirignano and co-

(12)

28

G. M. FAETH TABLE 6. Limiting cases for treating liquid phase transport in drop-life-history computations Case 1. Thin skin model
cqf D/

Notes Bulk liquid properties same as injected condition. Surface temperature and composition differ from bulk liquid. Bulk liquid composition same as injected liquid. Surface composition differs from bulk liquid. Temperature is spatially uniform, T b = T s. Both temperature and composition are spatially uniform in liquid. Tb = Ts,

2. Uniform temperature model

oo

3. Uniform state model

oo

oo

~bS = ~sS.

workers s'sl have shown that internal circulation within moving drops is a relatively ineffective agent for mixing and severely complicates estimations of liquid phase transport. Exact analysis is not very feasible even when surface active agents inhibit circulation and much less so when circulation must be considered. A reasonable means of circumventing this problem is by examining the three limiting cases in Table 6. The first case, or thin skin model, involves the assumption that only an infinitely thin surface layer is heated and has the composition changes required by phase equilibrium, while the bulk of the liquid remains at its initial state. For a quasisteady gas phase, the formulation of the thin skin model is identical to that of an evaporating sphere continuously supplied with liquid. The second case, or uniform temperature model, involves the assumption of infinite thermal diffusivity. This corresponds to the "well mixed" drop models often used in the past, 6 although the findings of Sirignano and coworkers, s'sl indicate that this terminology is
SURFACE LAYER T

not appropriate. In this case, the drop temperature is spatially uniform, but time varying, while a thin skin at the surface supports concentration changes required for equilibrium. The third limiting case, or uniform state model, extends the concept of temperature uniformity to species concentrations. In this case both temperature and composition are spatially uniform, but time varying. A fourth limiting case is possible, involving zero liquid thermal diffusivity and infinite mass diffusivities; however, this limit is not very useful since mass diffusivities in liquids are generally lower than thermal diffusivities. For the present assumptions, density is the only liquid property required by the analysis. The density is determined from the equation of state for the bulk liquid condition. 3.2.2. Gas phase transport The gas phase must be analyzed to determine expressions for heat and mass transfer rates. One method of completing this analysis is to solve the transport problem for a motionless drop in an infinite

,,fT
i I
I I

::fT

=,y
!
I i

:I

YAf~ |
"~ 0

N
I i

YAf
~YF r O

I I

I.,. i,r,.
i
r

\i

~y~
r

rio

II

rp

o| O

I rp I
UNIFORM STATE

THIN SKIN

UNIFORMTEMPERATURE
FIG. 25. Sketch of the drop vaporization models.

Evaporation and combustion of sprays stagnant medium and employ an empirical multiplicative correction to allow for convection. T M A second method is to employ the film theory technique widely used in the chemical engineering literature, sa This involves analyzing a spherically symmetric layer of thickness 6, where the value of 6 is determined from empirical expressions involving convection parameters. Both methods have been employed in recent drop-life-history computations and yield the same results when the Lewis number is unity. 2~28'84'85 For nonunity Lewis number, there is no evidence at this time that either method is preferred. The first procedure will be described in the following since it has the simplest notation. The basic gas phase model, prior to application of the convection corrections, is illustrated in Fig. 25. A binary system is illustrated, involving an injected fluid and an ambient gas, denoted F and A. The structure of both the drop and the gas phase is shown for the three limiting cases defined in Table 6, at a pressure sufficiently high so that dissolved gas effects are appreciable. The thin skin and uniform temperature models both involve surface layers whose properties are different than the bulk liquid, while the uniform state model involves uniform properties throughout the drop. In all cases, bulk liquid properties are known from the results of the drop-life-history computations to the instant under consideration. Properties of the surface layer, if present, and the gas adjacent to the surface are determined from phase equilibrium requirements and surface boundary conditions, as described later. For the present, we assume that gas phase properties at the surface are known. The governing equations for mass and heat transfer are discussed by Williamss2 and Bird et al. sa For the present spherically symmetric and quasisteady conditions, with a system consisting of n species, they have the following form" Conservation of mass" d By definition
i=1

29

L ~i = 1

(3.8)

however, the values of ei can be greater or less than zero, depending upon whether the species i is evaporating or condensing at the liquid surface. Specification of the e~ depends on the liquid transport model being used and will be discussed later. The heat transfer coefficient is defined as follows:

h=

or/sg/

and the boundary conditions are

r=rp;
r=oo;

T=T~, T=T,,

Y/=Y/sg, Y/= Yi,

i=l,n i=l,n

(3.10)

where the local flow conditions are represented by mean values. Integrating Eq. (3.3) and applying the boundary conditions of Eq. (3.10) yields the following expression for the net mass transfer rate

2xdppD

th

=ln( ~-~

\ Yi~g- 8i/'

i = 1, n,

Re = O.
(3.11)

Integrating Eq. (3.4) in the same manner gives the expression for the heat transfer coefficient as follows:

hdp

f rhCp'~ /r

f tfiCp "~ 11 '

O.

(3.12)
The quantity on the LHS of Eq. (3.12) is the Nusselt number. When rh ~ 0, Nu = hdp/2 ~ 2 from Eq. (3.12), which is the familiar value for a sphere in the absence of convection and mass transfer. Net outward mass transfer rates reduce the value of the Nusselt number, similar to the effect of blowing. The appropriate expressions for drag coefficient and the convection corrections for Eqs. (3.11) and (3.12) must now be determined. For the present, we adopt conventional practice and ignore virtual mass, accelerative, Magnus and Saffman lift forces in the expression for drag, as well as the counterparts of these terms in the expressions for mass and heat transfer. This approximation is satisfactory if the drop density is much greater than the ambient density and velocity gradients are small. The appropriate correlations after simplification in this manner have been described earlier. 6 The standard drag coefficient for solid spheres provides a good correlation for evaporating drops, s6 An expression proposed by Putnam 87 is convenient

dr (r2pv) = 0.

(3.2)

Conservation of species:

~dr [ r (2p v Y i - p D d Yd r J]_ l = 0 ; i~


Conservation of energy:

i=l,n.

(3.3)

zrL r [pvCp(rwhere & Y~= 1.


i=1

d F 2/

dT\]

(3.4) (3.5)

Integration of Eq. (3.2) yields

Co = 24(1+ReZ/3/6)/Re,
(3.6) where

Re < 1000 (3.13)


(3.14)

r2pv = rh/4x = const.

Re = dplup- ill/v.

where th is the net mass transfer rate of the drop, which includes both injected and ambient species. The mass transfer rate of each species can be conveniently represented by the mass flux fraction, defined as follows 82 rhi = ~rh, i = 1, n. (3.7)

The following expressions extend the correlation to higher Reynolds numbers, with similar accuracy, ss C D = 2 7 R e -84 , Re<80 80 ~< Re <~ 104 . (3.15)

= 0.271Re '217,

30

G.M. FAETH the energy reaching the surface of the drop is utilized to provide the energy required to gasify the evaporating material,

A useful formula for the multiplicative correction for heat and mass transfer, proposed by Faeth and Lazar, 89 is h
hRe = 0
or

nd~h(T- T~) = L e'ith(hisg-hibl)"


i-1

(3.22)

rh
rhRe=O

-1~

0.278Rel/2(pr or Sc) 1/3


(1

+l.232/(Re(Pr

o r Sc)4/3)) 1/2

(3.16)

The phase equilibrium requirements at the liquid surface are as follows L Y~f= L Y~0 = 1
i=1 i=1

where hR~=O and ~hRe=0 are found from and (3.12). Equation (3.16) has good properties at low Reynolds numbers and reasonable correlation of existing data
2 0 0 0 . 6 , 6 1 , 89

Eqs. (3.11) asymptotic provides a for Re <

(3.23) (3.24)

Fisf = Fiso, i = 1, n-1

3.2.3. Governin 9 Equations for the Drop In the following, a Lagrangian reference frame is adopted, following individual drops as a function of time as they pass through the flow field. The drop position is then given by xp = xp0+ jlu~dt (3.17)

where xp0 is the initial location of the drop. The instantaneous velocity is determined by solution of the equation for conservation of momentum m~-=duP --(rc/8)d~pCDlup--fil(up--~). (3.18)

where the fugacities of Eq. (3.24) are determined from the equation of state. Equations (3.8), (3.11), (3.12), (3.16) and (3.21)-(3.24) provide the 3n+ 3 equations needed to define surface conditions and transport rates. Given initial conditions and mean properties of the flow, the solution proceeds by numerical integration of Eqs. (3.17), (3.18) and (3.20). This analysis is simplified considerably if the bulk liquid is a pure material and solubility can be neglected. Denoting the bulk liquid species, F, Eq. (3.11) becomes

2~d.pO

rhRe=0 - - I n ( I - R .

~ / "

(3.25)

Equation (3.22) is also simplified, as follows,

ndEph(T - T~) = th(hvsg -- hFsf)


The three limiting cases for the liquid phase all imply uniform bulk liquid properties; therefore, the size and mass of the drop are related as follows:

(3.26)

while Yvso is determined, knowing the liquid surface temperature and the pressure, from the vapor pressure characteristics of the liquid

m = (x/6)pbyd3p

(3.19)

YFso = f(T~, p).

(3.27)

where Pbf is determined from the equation of state for the current bulk liquid temperature and composition. The remainder of the analysis varies, depending upon the treatment of the liquid phase, The formulation for the three limiting cases defined in Table 6 is described in the following. Case 1, thin skin model. In this case, the temperature and composition of the liquid phase remain at the injected condition,, while the liquid surface has different properties and composition, as dictated by equilibrium and transport requirements. Prediction of transport rates involves 3n+3 unknowns; Y~y, Y~sg and ei, for i = 1,n, as well as T~, h, and ~h. With these quantities known, conservation of mass of the drop yields dm
-dt =

The concentrations of the remaining species at the liquid surface are needed for the property evaluations of Eq. (3.1). These concentrations are found by noting that the argument of the logarithm of Eq. (3.11) must be the same for all species. Therefore, since ei = 0 for i @ F, we have Y~s0= ~(1 - YFso)/(1 -- YF), i (= F. (3.28)

- rh.

(3.20)

The composition of the bulk liquid is fixed, therefore, the mass flux fractions of the bulk liquid species must be equal to their bulk liquid mass fractions. Furthermore, the non-bulk liquid species only penetrate an infinitely thin surface layer and their mass flux fractions are zero. Therefore, ei = Yib~, i = 1, n-1. (3.21)

Case 2, uniform temperature model. In this case only the composition of the bulk liquid remains at the injected condition, while the surface composition varies as required by local conditions. The temperature of the surface and bulk liquid are the same and are known at each instant. Prediction of transport rates involves 3n+2 unknowns; Y~s, Y~s0and ei, for i = 1,n, as well as h and rh (or rh" = th/ndZ). The formulation of the uniform temperature model is very similar to the thin skin model. The main difference is that Eq. (3.22) is replaced by the equation of conservation of energy for the drop
(pfCpr

p/6)~t=h(T-Tp)-

dTp

L clth"(hiso-hibf)
i=1

(3.29) Equations (3.8), (3.11), (3.12), (3.16), (3.21), (3.23)and (3.24) remain unchanged and are sufficient to deter-

There is no bulk heating for this case; therefore, all

Evaporation and combustion of sprays mine the 3 n + 2 transport parameters. In this case, Eqs. (3.17), (3.18), (3.20) and (3.29) must be integrated numerically. The formulation of the uniform temperature model when only a single species is present in the liquid is also similar to the thin skin model, except that Eq. (3.22) is replaced by
(pyCpsdp/6) dt

31

drp

= h(T-

Tp) - rh"(hFso - hFsf).

(3.30) Case 3, uniform state model. In this case, the composition and temperature of the bulk liquid and the surface are the same and are known at each instant. Prediction of transport rates involves 2 n + 2 unknowns; Y~sgand e i for i - - 1,n, as well as h and hi. Equations (3.8), (3.11), (3.12), (3.16), (3.23)and (3.24) are sufficient to determine all these unknowns. Conservation of species in the liquid phase yields
dm i

dt

ei~h, i = 1, n.

(3.31)

Conservation of energy is correctly given by Eq. (3.29). The mass fraction of each species in the liquid phase is computed in the conventional manner, knowing the rni. In this case, Eqs. (3.17), (3.18), (3.29) and (3.31) must be integrated numerically. When only a single species is present in the liquid phase, the uniform state model is identical to the uniform temperature model. 3.3. M o d e l C a l i b r a t i o n There have been numerous studies of the evaporation and combustion of individual drops. ~a5 In general, models similar to the one described in Section 3.2 have been successful in predicting droplife-histories and correlating gasification rates at wetbulb conditions.* Good predictions are usually obtained since there are large property variations in the flow field around drops, which provide a great deal of latitude for adjusting "average" properties to match particular experimental results. Unfortunately, this latitude also results in substantial uncertainties when accurate a p r i o r i predictions are needed. Therefore, it is always desirable to calibrate drop models with test data for individual drops under carefully controlled conditions. This procedure involves adjusting ct in Eq. (3.1), for the present formulation, until a reasonable match is obtained between predictions and measurements. Appropriate calibration data is limited for combusting spray conditions; some measurements are cited in Ref. 6 and additional results will be discussed in the following. *When a drop is subjected to a fixed environment at moderate pressures, it approaches a wet-bulb state after a short transient heating or cooling period. At the wet-bulb state, all the heat reaching the drop is utilized for the heat of vaporization of the evaporating liquid and the drop temperature remains essentially constant (aside from slight changes in the wet-bulb temperature due to differences in the effect of convection on heat and mass transfer rates when the Lewis number is not unity).

Examples of the calibration of the present model, for conditions representative of the evaporating and combusting sprays examined using the L H F model, are summarized in Table 7. Predicted and measured gasification rates are compared, considering the wetbulb state for supported drops and the steady evaporation condition for porous spheres (which are kept wetted by a continuous supply of fuel). Since bulk heating is nearly complete at the wet-bulb state, the uniform temperature model was employed for these conditions. The thin skin model was employed for the porous sphere experiments since this model incorporates the correct boundary conditions for this case. The porous sphere experiments were conducted at sufficiently high pressures so that portions of the model involving real gas effects could be verified. In general, the comparison between predictions and measurements is satisfactory, using ~ = 0.9 for all the results. This choice of ct depends on the property correlations and mixing rules for mixture properties used in the computations. The same procedure was used throughout the computations summarized in Table 7; other methods would yield different optimum values of ct. Satisfactory predictions of complete drop-lifehistories were also obtained for the calibration tests conducted with supported drops. 26'27'57'90 A portion of the results summarized in Table 7 was obtained by supporting drops at various positions in a turbulent diffusion flame (actually the combusting propane gas jet considered in Figs. 11-15). This procedure provides a realistic simulation of conditions which are encountered in combusting sprays. A drop moving through a combusting spray encounters local species concentrations and temperatures which are largely prescribed by the adiabatic equilibrium state for the local mixture fraction, barring significant heat losses and nonequilibrium effects. Therefore, obtaining calibration data for a similar range of mixture fractions in a flame tests the drop-life-history model for the most appropriate range of conditions. The predictions shown in Table 7 for drops supported in flames ignored the presence of envelope flames even when oxygen was present at the test location. A few cases where envelope flames could be observed are identified, although luminosity from the flame itself makes observation of drop envelope flames difficult in some regions of the flow. When oxygen concentrations are low, there is little enhancement of transport rates due to the presence of envelope flames; therefore, neglecting their presence has little influence on the results. One condition for a JP-10 drop, however, involved a clearly visible envelope flame with a relatively high local oxygen concentration. In this case, the predicted gasification rate is significantly lower than the measured value, due to the presence of the envelope flame. Farther downstream for the same fuel, however, where gas temperatures are lower and oxygen concentrations are higher, the drop could not sustain a stable envelope flame. Therefore, the evaporative prediction is again close to the measurement. Envelope flames

TABLE 7. S u m m a r y of drop transport calibrations a Gas composition (mass fractions) N2 02 CO CO 2 HzO C3H 8 Comment b Gasification rate (mg/s) Measured Predicted

Pressure (MPa)

Relative velocity (m/s)

Gas temperature (K)

Shearer 0.77 0.77 0.23 0.23

e t al. 26

suspended Freon-11 drops, 150 p m in diameter, in a flowing air stream; predictions using the uniform temperature model, neglecting ambient gas solubility. 0.00211 0.00155 0.00234 0.00153

0.097 0.097

6.0 3.0

298 298

Mao 0.66 0.62 0.66 0.72 0.00 0.01 0.12 0.18 0.02 0.08 0.02 0.00 0.07 0.10 0.07 0.03 0.06 0.16 0.10 0.07 0.17 0.01 0.00 0.00

et al. 27

suspended n-pentane drop, 600 # m in diameter, in a propane diffusion flame; predictions using the uniform temperature model, neglecting ambient gas solubility. Combusting? 1.03 0.814 0.559 0.310 0.985 0.800 0.479 0.299

0.097 0.097 0.097 0.097

22.1 9.93 4.35 2.92

1008 1841 1271 779

Szekely and Faeth 9 suspended JP-10 drop, 420 4 5 0 # m in diameter, in a propane diffusion flame; predictions using the uniform temperature model, neglecting ambient gas solubility. 0.645 0.689 0.677 0.688 0.699 0.006 0.021 0.076 0.093 0.122 0.079 0.051 0.039 0.031 0.016 0.101 0.094 0.079 0.075 0.066 0.161 0.145 0.129 0.113 0.097 0.008 0.000 0.000 0.000 , 0.000
e t al. 28

0.097 0.097 0.097 0.097 0.097

9.93 8.14 6.15 4.82 4.35

1841 1903 1824 1541 1271

Combusting? Combusting? Combusting

0.572 0.517 0.479 0.501 0.298 using the thin skin model. 23.0 26.1 32.4 23.6 32.4

0.582 0.519 0.428 0.345 0.290

Canada and Faeth 61 porous sphere flowing n-heptane, 9500/~m in diameter, in exhaust from a c o m b u s t i o n chamber; predictions by M a o 0.54 0.54 0.54 0.54 0.54 0.46 0.46 0.46 0.46 0.46

1.45 2.07 2.35 2.79 3.45

0.317 0.316 0.332 0.335 0.285

1255 1255 1255 1255 1255

20.6 23.5 25.2 27.0 27.2

a ~ = 0.9. Correlations for species properties and mixing rules for mixture properties described in Refs. 57 and 58. b Drop evaporating with no envelope flame unless noted otherwise; combusting denotes envelope flame observed; combusting ? denotes possible envelope flame but not observable due to flame luminosity.

Evaporation and combustion of sprays will be considered further in Section 5, for the present we can conclude that their presence has little influence on drop transport rates in diffusion flames, as long as the drops do not penetrate beyond the maximum temperature position in the flame. Even though the calibration experiments listed in Table 7 are helpful, a number of deficiencies remain. It is difficult to measure drag for small supported drops; therefore, predictions of drag remain unassessed aside from the original tests by Yuen and Chen s6 used to establish the present drag correlation. The calibration conditions in Table 7 involve relatively large supported particles rather than the small moving particles of interest in sprays. This implies differences in the Reynolds number range, the relative intensity and scale of turbulent fluctuations, a: well as possible influences of turbulent diffusion of particles on transport rates. Removing these deficiencies to provide an accurate calibration of drop models will require the development of innovative experimental techniques. 3.4. Drop-Life-Histories in Sprays With the drop transport model calibrated, aside from the exceptions noted earlier, the results of drop300
I I I

33

life-history calculations can be employed to examine the behavior of drops in sprays. In the following, the evaporating F r e o n - l l spray and the combusting n-pentane sprays summarized in Table 5 will be considered. In each case, the results of the L H F model computations are employed to specify the local ambient conditions of drops passing along the centerline of the spray. A range of drop sizes will be considered, with initial conditions specified by the injector exit state determined for the L H F model. Predicted drop-life-histories along the centerline of the evaporating F r e o n - l l spray are illustrated in Fig. 26. The diameter, velocity and temperature of the drop, as well as the L H F prediction of the velocity and temperature of the gas, are shown as a function of distance from the injector. Drops having initial diameters of 10, 30 and 50#m are considered, employing the uniform temperature model for liquid phase transport. Dissolved gases are neglected since gas solubilities are low at atmospheric pressure. Comparison of drop and gas velocities in Fig. 26 indicates that drops as small as 10 ~tm initial diameter have appreciable relative velocities (slip) in the region just downstream of the potential core. Predicted fluid temperatures reach a minimum downstream of the
I i I I I I

260

\
LHF THEORY

t
O/Oo------

U/Uo~

0o:10 prn

1 I

~0

.Ol
lo lOO x/d looo

FIG. 26. Predicted drop-life-histories along the centerline of a Freon-ll spray evaporating in air at atmospheric pressure. From Shearer et al. 26
JPECS 9-I/2-C

34

G.M. FAETH
i i i i i i k

injector and then increase toward the ambient temperature of the spray at large x/d. The drop temperatures lag the fluid temperatures initially and then decrease monotonically toward a constant value below the gas temperature. In the region where drop temperatures are nearly constant, predicted vapor concentrations are relatively low; therefore, the drops approach the wet-bulb temperature appropriate for evaporation in pure air. The period where the drop temperature varies, however, extends over much of the lifetime of the particle justifying the need for considering transient effects. The drop diameter decreases relatively slowly near the end of the potential core, since vapor concentrations are high and gas temperatures are low in this region. Farther downstream, higher gas temperatures and lower vapor concentrations yield higher rates of evaporation. However, since drop mass is proportional to d 3, appreciable liquid is evaporated in the relatively flat portions of the diameter plots near the end of the potential core. In this region, since drop temperatures are decreasing, the evaporation process is enhanced by flashing. The test spray had maximum drop diameters on the order of 50#m. The calculations illustrated in Fig. 26 suggest that drops of this size should be present along the ccnterline for x/d >~ 510 (the measuring location farthest from the injector) which agrees with observations. In contrast, the L H F model predicts that all liquid should be evaporated at x/d = 165. The droplife-history computations indicate that finite interphase transport rates are important for this evaporating spray, probably accounting for much of the discrepancy between L H F predictions and the measurements. It appears that initial drop diameters less than 10pro would be required to obtain quantitative accuracy with the L H F model for these conditions. Predicted drop-life-histories along the centerline of the air atomized n-pentane spray burning in air at atmospheric pressure are illustrated in Fig. 27. These predictions employ the uniform temperature model for liquid phase transport, neglecting solubility. The mean velocity, mean temperature, and mean concentrations of fuel and oxygen in the flow, as well as the drop velocity and diameter, are shown for initial drop diameters of 20, 40 and 100 #m. Thermodynamic equilibrium is assumed for the L H F model; therefore, the temperature of the flow leaving the injector is 390K, due to reaction of the atomizing air and the fuel. This causes the drops to heat-up and evaporate as they pass through the potential core. In contrast, the drops are assumed to have the same velocity as the flow at the injector exit and slip only appears downstream of the end of the potential core. Assuming that equilibrium in the potential core is realistic, drops having an initial diameter less than 20#m disappear before the end of the potential core and do not contribute to slip. Drops larger than this, however, have appreciable slip in the region just downstream of the end of the potential core. The actual spray had a SMD of 35 #m (c.f. Table 5); therefore, the results illustrated in Fig.

.p/%,%(~)=ao ~ 40 N ioo~
i ~a. } ~ i i L i

,2

2000

dp/d

i,~'~
i
.6 o

I000
BOO g
600 \ ::3 400 ~

~: b. '/)
,2
[ I

~ FUELVAPOR ~" / OXYGEN/~


[ I i

MASS FRACTION

I.d o.. :~.1

200

'lo

ao

40

60 eoloo

zoo

400 600

x/d

FIG. 27. Predicted drop-life-histories along the centerline of an n-pentane spray burning in air at atmospheric pressure. From Mao et al.2~ 27 suggest that effects of finite interphase transport rates were appreciable and probably caused much of the discrepancy between the predictions of the L H F model and the measurements. It appears that a spray having initial drop diameters less than 20 #m would be required for the L H F model to be quantitatively
accurate.

The drop-life-history results illustrated in Fig. 27 indicate that drops having initial diameters less than 100 #m disappear before reaching the flame position (taken as the point where the mean mixture ratio is stoichiometric, which roughly corresponds to the maximum temperature position of the flow). Maximum drop diameters were less than 90pm for these tests and in agreement with the drop-life-history results, no drops were observed at x/d = 340. 27 Therefore, the drops in this spray were limited to an environment having low oxygen concentrations where the presence of envelope flames has little influence on transport rates. Only large drops having initial diameters greater than 100#m would penetrate the flame and reach conditions where it would be important to determine whether envelope flames were present in order to properly model the latter stages of the lifetime of the drop. Drop temperature changes are too small to show clearly on the scale of Fig. 27. However, the calculations indicated that the drops did not reach a fixed wet-bulb temperature during their lifetime, due to variations in local flow conditions. Therefore, it is necessary to treat transient effects of drop heating throughout the lifetime of the drop. The use of the thin skin model yields essentially the same results as the uniform temperature predictions illustrated in Fig. 27. This behavior is observed since sensible heating of the liquid is a relatively small fraction of the total energy of gasification for n-pentane at atmospheric pressure, due to the relatively high volatility of this fuel.

Evaporation and combustion of sprays


I I I

35
I I

1.0 .6

~0 ~ el. .2 ,r'-d n XP.-/ d P'~-_....

"~

"k
s

1.0
-

.6 .4
I>~

\ .._..~-~o ~

-XF-~oo~m -

A
-_

\.I

il\._
!I \.

il

Z
_

.2

.06 .04

J iil
-~-~''-7
_

i iI ",i
--"X/j~ ./" /~ , ~o ~m
~00 ~m

3./s
Z
I I
I000

2000

"-"

i000 600 400

--~- -~

200

I
20

Pl
40

I
60

]
i00

I
200

400 6OO

x/d
FIG. 28. Predicted drop-life-historiesalong the centerlineof a pressure atomized n-pentane spray burningin air at 3 MPa. Flow properties _ _ , critical temperature . Drop properties: thin skin model , uniform temperature model , vanishing point of drop by evaporation .

Predicted drop-life-histories along the centerline of the pressure atomized n-pentane spray burning in air at 3 MPa are illustrated in Fig. 28. Both the thin skin and uniform temperature liquid phase transport models are illustrated since there are significant differences in the predictions of these two limits in this case. Due to the high pressure, real-gas effects and the solubility of noninjected species in the liquid were considered during the computations. Drop velocity diameter and bulk (uniform temperature model) or surface (thin skin model) temperatures are illustrated for drops having initial diameters of 10 and 100~m. The larger size is representative of the largest drops found in the test spray for cold flow conditions. Flow properties obtained from the LHF model are also illustrated in the figure, including: mean velocity, tem-

perature, fuel liquid fraction, fuel vapor mass fraction and oxygen mass fraction. Both predictions illustrated in Fig. 28 indicate that only drops having initial diameters less than 10pm have velocities and temperatures nearly equal to the flow and disappear near the spray boundary estimated by the LHF model, c.f. Fig. 20. Since the test spray contained drops larger than 10 #m in diameter, these findings indicate that finite drop transport rates are significant for the test conditions and the LHF model would be expected to overestimate the rate of development of the flow. However, the departure of the test conditions from the requirements of the LHF model is less than for the atmospheric pressure sprays illustrated in Figs. 26 and 27. The drop velocity predictions for the two liquid

36

G.M. FAETH critical pressure of the liquid. Past work on critical conditions for individual drops, either evaporating in heated gases or burning in an oxidizing environment, also indicates that pressures on the order of twice the critical pressure of the pure liquid are required to reach the critical point. 61'9~94 The predictions showed that some drops reached the thermodynamic critical point for the pressure atomized n-pentane spray burning in air at 9 MPa. The drop-life-history predictions for this condition are illustrated in Fig. 29. The variables shown and many of the features of the computations are the same as Fig. 28. The differences in velocities and temperatures between the drops and the flow are roughly the same as the results for 3 MPa, since the interphase transport rates and the L H F prediction of the rate of flow development both increase with increasing pressure. Therefore, it appears that high gas phase densities alone are no guarantee of improved accuracy for LHF predictions, since the rate of flow development plays a major role in the validity of the L H F approximation. The L H F predictions for the conditions of Fig. 29 indicated that spray boundaries were established by conventional evaporation of the liquid, rather than by the liquid reaching its critical point, due to the relatively low temperature levels within the spray boundaries. In contrast, the drop-life-history calculations yield instances where drops reach their thermodynamic critical point at this pressure, particularly large drops which penetrate to high temperature regions of the flow. For example, using the thin skin model, a drop having an initial diameter of 100#m evaporates in a conventional manner, while 200pm drop reaches its critical point near the maximum temperature position in the flow. The uniform temperature model gives similar trends, but indicates that smaller drops reach the critical point, since this model predicts higher liquid temperatures than the thin skin model toward the end of the drop lifetime when bulk heating is nearly complete. Further increases in pressure reduces the initial drop diameter required for drops to reach their critical point for both drop models. The results illustrated in Fig. 29 indicate that there are appreciable velocity differences between the drops and the flow when the critical point is reached. Surface tension approaches zero as the critical point is neared, suggesting that drop deformation and breakup would become important as this condition is approached. Some aspects of drop breakup, helpful in examining this phenomena, will be considered in Section 3.5. The environment of the sprays considered thus far was motionless. It is of interest to examine drop-lifehistories in moving ambient flows which are more typical of practical applications. Boyson and Swithenbank 95 report recent results of this type for a spray injected into a heated air stream within a cylindrical chamber at atmospheric pressure. A cylindrical baffle, having a diameter half that of the chamber, yields a recirculating zone at the upstream end of the chamber, roughly simulating the flow patterns encountered in

models illustrated in Fig. 28 are similar. The relatively small injector diameter results in a large rate of deceleration of the flow; therefore, there is significant slip for the 100pm diameter drop in the region just downstream of the potential core. In contrast to velocity, drop temperature and diameter predictions are significantly different for the two liquid diffusivity limits. Drop surface temperatures are relatively independent of size for the thin skin model, therefore, only a single surface temperature line is shown for this case. On the other hand, the uniform temperature model indicates that the rate of heating decreases as the drop size increases. Predictions for both liquid phase models indicate that drop temperatures tend to follow the temperature of the flow near the injector while farther downstream drop temperatures approach a constant value even though flow temperatures are increasing. Steady evaporation at a constant wet-bulb temperature is never achieved, however, since flow properties are changing continuously. The uniform temperature model predicts relatively low liquid surface temperatures in the region just downstream of the potential core. Fuel vapor concentrations are relatively high in this region, causing condensation of fuel vapor on the drops. The combined effect of condensation and reduced liquid densities due to heating cause predicted drop diameters for the uniform temperature model to increase for a time. In contrast, the thin skin model provides for no bulk heating and yields sufficiently high surface temperatures so that condensation does not occur; resulting in a monotonically decreasing drop diameter as distance from the injector increases. The critical temperature of the liquid varies slightly with distance from the injector due to changes in the composition of the flow. Predicted drop temperatures remain well below the local critical temperature for both liquid models. These differences in the two models indicate that better methods of treating liquid phase transport are needed in order to accurately treat high pressure sprays. The results illustrated in Fig. 28 indicate that only large droplets having initial diameters greater than 100pm would penetrate the flame zone and reach the region where oxygen concentrations are significant. Maximum drop diameters in the test spray for cold flow conditions are estimated to be smaller than 80 #m; therefore, envelope flames probably play only a small role in drop transport processes in this spray. The drops simply evaporate in the cool core of the spray, yielding fuel vapor which oxidizes farther downstream in the flame zone. Computations conducted for the pressure atomized n-pentane spray burning in air at 6 MPa gave results similar to those illustrated in Fig. 28. Even though this pressure is almost twice the critical pressure of pure n-pentane (3.37 MPa91), the drops did not reach the thermodynamic critical point. This behavior is observed since the mole fraction of fuel vapor at the drop surface was significantly less than unity, even when bulk heating was nearly complete. Furthermore, the presence of dissolved gases tends to raise the

Evaporation and combustion of sprays


I
I I

37
I I I

I ~

1.0
O

.6 .4

[u , 10pmJ~'~.~x ~_ ~ , o N~#, "~.N -~ 2 o o -

~m

~O

.2 ~dp/~_o~lO 1.0 .6 .4
,H

Lt m

N~.

YFf - "

"\

)!

...,~_,~

100 um J ~

~)

o .2
"=1

L_

\
\

.1 .06 .04

gi

I
|

\ %7//

z
\

2000

~-~ i000 600 400


E-~
T ,

/' CRIT. TEMP. /

pm 200 gm pm
_

200

. pl 20

i0 ~m -x i00 ~rn-~ V00 pm--~ I I I [ I I t 40 60 i00 200 400 600 i000 .x/d

Fla. 29. Predicted drop-life-histories along the centerline of a pressure atomized n-pentane spray burning in air at 9 MPa. Flow properties___ , critical temperature .... Drop properties: thin skin model__, uniform temperature model ____, vanishing point of drop by evaporation , drop reaches thermodynamic critical point O.

gas turbine combustors. The spray is injected from the center of the baffle with a cone angle of 45 and an initial velocity roughly twice the inlet velocity of the heated air stream. Boyson and Swithenbank 95 employ a simplified model of the flow. The gas flow was predicted using a k-e turbulence model, but the presence of the spray was ignored entirely. This procedure was justified by the argument that the momentum of the inlet flow dominates the process, except for the region close to the injector. The drop-life-history model was also highly simplified. The drop lifetime was divided into two parts, a heat-up period where no evaporation occurs (terminated when the drop reaches the wetbulb temperature which can be approximated by the

boiling temperature at atmospheric pressure 82), and a steady evaporation period A convection correction, similar to Eq. (3.16), was applied to the transport rates in both periods. Particle drag was represented by the standard correlation for solid spheres. The calculations were not calibrated for any particular fluid. A portion of the predictions obtained by Boyson and Swithenbank 9s are illustrated in Fig. 30. The upper part of this figure shows streamlines in the recirculation zone and the development of the axial velocity distribution The lower part of the figure shows the trajectories of individual drops in the flow. Two predictions are shown for each drop size, one allowing for the heat-up period and one assuming that steady evaporation was initiated at the injector exit.

38

G.M. FAETH

I .0

2,0

3.0

4.0

5.0

NOMIN~ CONE

1.0

y/r 0.5

~'B~FLE

'~

50~__~

_ _

WITHOUT H~ATUP
~ m ,

l'

- - I -

1.0

2.0

3,0

4.0

~'.o
x/r

FIG. 30. Predicted mean velocities and drop trajectories for drops injected into a heated and recirculating flow. From Boyson and Swithenbank. 95

Due to the relatively low pressure and the fact that the interior structure of the spray was not considered, the effect of considering heat-up was not large for these conditions. An interesting feature of the computations is that the drop trajectories deviate to a small degree from their initial direction in the recirculation region, and only respond to deflection by the flow near the end of their lifetime, when the drop is small. This behavior is somewhat similar to the results employing the LHF model to represent the structure of the spray, illustrated in Figs. 26-29, where the larger particles also respond relatively slowly to mean flow variations. Weak response to mean flow variations also implies a weak response to turbulent fluctuations which cause turbulent drop diffusion. The possibility of ignoring turbulent diffusion, except for the smallest drops, can provide an important simplification for SF spray models. Similar to the earlier evaluations of L H F models, the results of Fig. 30 indicate that drops having initial diameters greater than 10/~m have appreciable slip. Nevertheless, an LHF prediction would provide a better estimation of the flow structure, with little additional effort, than simply ignoring the presence of the spray.

3.5. DropBreakup Thus far, we have considered the behavior of drops having initial diameters up to maximum sizes found when calibrating the injector (normally in cold flow) in order to evaluate L H F predictions of sprays. This approach is deficient, since use of the injector for a particular application can involve different flow conditions and drop temperatures, which changes the drop size distribution from the calibration condition. An obvious example involves high pressure combusting sprays where drops approach their thermodynamic critical point, leading to shattering effects which are not present in cold flow tests. In the following, we consider aspects of drop shattering, with particular reference to the combusting high pressure sprays illustrated in Figs. 28 and 29. Past studies of drop breakup have generally subjected drops to abrupt changes of relative velocity, observing the occurrence, the mode, and the time required for breakup. Measurements conducted using shock wave disturbances of various strengths past the drop yield the following correlation for the critical Weber number required for drop breakup, v
Wecrit = 1/2 Recrit/2

(3.32)

Evaporation and combustion of sprays where


We = plup-filZdp/(2G,).

39

(3.33)

Two modes of drop breakup are generally observed. For values of the Weber number near the limit given by Eq. (3.32), "bag" breakup occurs. In this case, liquid at the forward stagnation point is deflected through the center of the drop, forming a bag-shaped liquid shell with the opening at the upstream end of the drop. The bag expands and eventually shatters into a number of small drops. The second or stripping mode of drop breakup is observed when the Weber number is much greater than the limit of Eq. (3.32). In this case, breakup occurs by liquid shearing off the periphery of the drop. In both cases, breakup is not instantaneous. Wolfe and Anderson 96 provide expressions to estimate the time of breakup, for shock wave disturbances, and the mean drop size after breakup. Ranger and Nicholls 97 suggest the following expression for breakup time in the stripping mode
tblup-fil(p/py)l/2/dp = 3.5-5.5

(3.34)

where the range results from different test conditions and a value of 4.0 is recommended in the absence of other information. 7 Applying the results of Ref. 96, or Eqs. (3.32) and (3.34), to sprays involves considerable uncertainties. Property selection rules, particularly for combustion at high pressures, have not been established for breakup phenomena. Furthermore, the Weber number varies more gradually in a spray than in the shock wave disturbances used to establish the empirical expressions. In spite of these limitations, it still is of interest to examine the breakup properties of the high pressure sprays considered using the L H F model. Weber number and drop breakup times were computed for a 100pm drop at 3 and 9 MPa, employing the drop-lifehistory results of Figs. 28 and 29 to estimate drop properties. These results indicated that Weber numbers for a drop having an initial diameter of 100#m exceeded the critical Weber number for drop breakup a short distance downstream of the potential core (x/d ~ 40-50) for both pressures. Predicted drop breakup times were roughly two orders of magnitude smaller than the drop lifetime, suggesting rather abrupt shattering of the drops. Drops having an initial diameter of 10#m only exceeded drop breakup conditions as the critical point was approached, for the uniform temperature model at 9 MPa. These findings suggest that the region where liquid is present in these flames is probably overestimated by the trajectories for the 100/~m drops in Figs. 28 and 29. Furthermore, drop shattering in high pressure combusting sprays, where most drops approach their thermodynamic critical point, undoubtedly limits maximum drop sizes to values substantially below cold flow calibrations, tending to increase the effectiveness of L H F predictions for these conditions. 3.6. Characteristic Times for Sprays Many investigators have considered characteristic

times for flow and drop processes in order to gain a better understanding of sprays, c.f. Refs. 7-15 and references cited therein. It has been difficult to obtain quantitative conclusions from these considerations due to property uncertainties and the fact that characteristic times are only order of magnitude representations of drop transport processes. In the present case, however, the drop-life-history computations can be used to support results using characteristic times, providing a convenient method of generalizing some of the results of the drop-life-history calculations. This involves comparing the characteristic relaxation time of the flow with various characteristic times for drop transport processes. Those circumstances where the flow relaxation time is long in comparison to interphase transport times represent situations where the L H F approximation is likely to yield satisfactory predictions. We confine our attention to the flow induced by the spray itself. The variation of mean quantities along the centerline of sprays, c.f. Figs. 9, 10, 16 and 21, indicates that at least over a narrow range of x, downstream of the end of the potential core,

Uo--Uoo

Uc-U ~ fc ~
--

fo

[(x/d),,d(x/d)]",

x / d >1 (x/d),, o

(3.35) where n is a constant of order unity. Choosing velocity defect to represent the rate of flow decay the following flow relaxation time is obtained, based on conditions at the end of the potential core.
d ( u c - uo~) - t d(x/d),c

(3.36)

~s = ~

n(Uo--U~)"

Since (x/d)pc does not vary greatly for different sprays, Eq. (3.36) indicates that sprays tend to have small relaxation times since the diameters of most injectors are small and good atomization requires a reasonably large value of u 0. This is one of the main reasons that slip, etc. was significant for drops as small as 10pm, in the drop-life-history results illustrated in Figs. 26-29. Characteristic times for the interphase transport of momentum, heat and mass must now be considered. Combining Eqs. (3.17)-(3.19), the following expression for the characteristic time of momentum exchange is obtained
dup - 1 4pydpuo rn,~ = dx = 3pC~(~o ZUoo) z

(3.37)

based on conditions at the end of the potential core with the local flow velocity represented by the ambient velocity. In general, Reynolds numbers for drops at this state are higher than the Stokes flow regime and Co must be evaluated from Eqs. (3.13)(3.15). Heat transfer to the drop influences the characteristic times of heat-up and evaporation. A characteristic time for heat-up can be obtained from Eq. (3.29), neglecting the effect of mass transfer in this

40 period, as follows

G.M. FAETH influenced by surface temperature estimate since this quantity appears in the argument of a logarithm. The definition of the characteristic times of interphase transport is completed by selecting a reference state for properties and an appropriate drop size. A convenient reference state for properties is the local ambient condition of the spray since only order of magnitudes are required for characteristic time considerations. The maximum drop size is perhaps the best measure of effects of slip, etc., however, SMD was considered in the following since this parameter can be more readily estimated for sprays. These characteristic times are obviously only crude measures of spray processes, therefore, we shall limit consideration to the cases where drop-life-histories were computed--illustrated in Figs. 26-29. In this manner, the present approximate discussion can be supported by more accurate computations. Table 8 is a summary of the results, giving the characteristic times of the flow and drop interphase transport processes, as well as the ratio of these quantities. There is rough agreement between the trends obtained from the ratios of interphase and flow relaxation times in Table 8, and the drop-life-history results illustrated in Figs. 26-29. For example, the ratios of interphase transport to flow relaxation times are largest for the evaporating F r e o n - l l spray while the drop-lifehistory computations of Fig. 26 indicated the greatest slip, etc. for this flow. Conversely, these ratios are smallest for the combusting air atomized n-pentane spray, in agreement with the fact that Fig. 27 indicates the smallest effect of slip, etc. for this flow. Therefore, although drop-life-history calculations provide a more reliable means of evaluating interphase transport effects in sprays, the approximate characteristic times are a useful indication of the potential performance of a L H F model while requiring much less computational effort. 3.7. Summary A conventional drop transport model has been described which can provide an estimation of droplife-histories in a spray. Major features of the model include assuming a quasisteady gas phase, treating

"rdh= ( , f m a x - - T o ) / ~ - =

,,,./dTp]

p f f p f d 2 ( T f m a x - To)

122(h/hRe=O)(T- TI)max
(3.38)

employing NURe=O= 2. The rate of temperature increase has been normalized by the total temperature rise during heat-up. In the following, Eq. (3.38) will be evaluated based on conditions at the end of the potential core, with the local flow having the ambient velocity and a temperature yielding the maximum temperature difference between the drop and the flow, e.g. the flame temperature for a combusting spray. The drop evaporation time will be based on the thin skin approximation for the liquid phase, assuming a single liquid component. Solving Eq. (3.22) for rhRe=0, we have thRe=0 = r~d~hRe=0(T- T~)/(h~g-hbs). (3.39)

The effect of mass transfer on heat transfer must be considered for an evaporating drop. This is accomplished by substituting Eq. (3.12) into Eq. (3.39), yielding
rhRe = oCp _

In 1 -t- c/T~ J

(3.40)

2nd.)o

Equation (3.40) corresponds to the usual expression obtained for steady evaporation at the wet-bulb state, 6 except that the energy required to heat the fluid from Tb to T~is included in the total enthalpy required for gasification. The characteristic drop evaporation time can now be obtained from Eqs. (3.12), (3.19), (3.20) and (3.40) as follows

p fC pd2e
Zde

~'122(~__m / in [1 + C,(T- T~)max]~ "

\mge=o/

L "

h~o-hbf

]J
(3.41)

Equation (3.41) is evaluated based on the same conditions as Zah. The surface temperature is taken to be either the boiling point of the liquid (for low pressures) or the critical point (for high pressures). The characteristic evaporation time is not strongly

TABLE8. Summary of characteristic times for sprays considered in Figs. 26-29 Air atomized evaporating freon-ll spray, 0.097 MPa 26 74.5 1.194 29 18 290 660 -3000 2.3 Air atomized combusting n-pentane spray, 0.097 MPa 27 11.6 1.194 35 40 4100 880 20 1600 0.2 0.004 0.4 Pressure atomized combusting n-pentane spray, 3 MPa 28 69 0.2 28 30 90 30 30 150 0.3 0.4 1.7 Pressure atomized combusting n-pentane spray, 9 MPa 29 69 0.2 28 25 70 10 20 90 0.1 0.3 1.2

Flow Figure uo (m/s) d (mm) SMD (/~m)

(x/d)pc
zs (#s) Zdm(#s) "r~h(/.is) Zde(/~s)
"[dm/'~s

Zdh/'q
zde/Zs

--

10

Evaporation and combustion of sprays convection and drag by empirical expressions, neglecting envelope flames, and employing limiting cases for liquid phase transport (thin skin, uniform temperature and uniform state models). Most of these features have been in use for models of spray combustion in liquid rocket engines for some time. v Satisfactory predictions of drop transport were obtained with a reasonable selection of physical properties (ct = 0.9 in Eq. (3.1)). Past evaluations of models of this type also indicate good performance. 6'7'9-15 Due to uncertainties of physical properties, however, it is recommended that drop transport models be calibrated with measurements obtained from well-controlled experiments, whenever possible. Drop-life-history calculations provide a systematic method for evaluating the potential accuracy of L H F predictions of sprays. Differences in drop and flow velocities and temperatures provide a direct indication of the effect of finite interphase transport rates and the maximum spray drop sizes where the L H F model can be expected to give accurate results. Cold flow injector measurements provide an indication of maximum expected drop diameters for assessment of L H F models, however, secondary drop breakup should be considered to ensure that such sizes can be sustained in the injector application under consideration. Methods of estimating drop shattering are not as highly developed as other drop processes and should receive further attention in order to obtain realistic spray models. The present drop-life-history computations suggest that finite interphase transport rates were important for the evaporating and combusting sprays considered in Section 2.3, probably accounting for much of the discrepancy between L H F predictions and the measurements. It was generally found that drops having initial diameters less than 10#m would be required to satisfy the L H F approximation. Consideration of characteristic times indicated that small flow relaxation times, due to small injector dimensions, were the primary cause of this behavior. Since most practical injectors are small, this suggests that SF modeling will be generally required in order to accurately represent sprays. L H F predictions should be at least subject to error at high pressures, where approach to the critical point results in the shattering of large drops, and for internally mixed air atomizing injectors which have relatively large exit diameters.

41

4. SEPARATED FLOW MODELS

4.1. Introduction Separated flow models of multiphase flow specifically treat the finite rate exchange of mass, momentum and energy between the phases. Current models generally average over processes which occur on scales comparable to the drop size, i.e. no attempt is made to accurately model the details of the flow field around or within individual drops due to

practical limitations of computer storage and computation costs. Therefore, the exchange processes between phases must be modeled independently, usually by employing empirical expressions for drop drag and heat and mass transfer. Processes within the liquid phase are also simplified by employing one of the limiting cases described in Table 6 (usually Case 2). Most separated flow analyses of evaporating and combusting sprays employ an approach variously identified as the "particle-source-in-cell" (PSIC) 98-11 or "discrete droplet model" (DDM), 17 e.g. Refs. 98108. This involves dividing the spray into representative samples of discrete drops whose motion and transport are tracked through the flow field, using a Lagrangian formulation similar to the drop-lifehistory computations discussed in Section 3. This procedure corresponds to a statistical (Monte Carlo) computation for liquid properties since a finite number of particles are used to represent the entire spray. A Eulerian formulation is employed to solve the governing equations for the gas phase, similar to the solution of the flow equations for L H F models. The effect of drops on the gas phase is considered by introducing appropriate source terms in the gas phase equations of motion. D D M eliminate errors due to numerical diffusion in the liquid flow solution and it is generally found that the number of particle classes required to achieve a satisfactory representation of the spray is not excessive. The formulation is also convenient for considering a relatively complete representation of drop transport processes. A second approach for separated flow analysis of sprays involves a statistical or "continuous droplet model" (CDM), 17 described by Williams.82 In this case, drop properties are represented by a statistical distribution function defined in a multidimensional space of drop diameter, position, time, velocity, temperature, concentration, etc. Conservation principles yield a transport equation for the distribution function which is solved along with the gas conservation equations to provide all properties of the spray. Similar to DDM, the governing equations for the gas phase include appropriate source terms for effects due to the presence of drops. The CDM approach is best applied when only a few phenomena must be considered, e.g. the motion of monodisperse nonevaporating particles. When variations in drop size, concentration and temperature must be considered, the dimensions of the distribution function must be increased. Since drops tend to have a limited range of positions, diameters, velocities, temperatures, etc. the distribution function is relatively sparse, leading to burdensome computer storage requirements. The third main method for modeling sprays employs a continuum formulation of the conservation equations for both phases (CFM). The motion of both drops and gas are treated as though they were interpenetrating continua. With this formulation, the

TABLE9. Summary of properties of representative discrete droplet models of sprays

Reference(date) Spray from sprinkler evaporating in air Steady 2-D Dilute, completely interacting phases Steady 2-D Dilute, completely interacting phases Steady 2-D Dilute, completely interacting phases Steady 3-D Dilute, completely interacting phases Transient 2-D Dilute but considering void fraction, cornpletely interacting phases Combusting spray in a cylindrical furnace Combusting spray in a cylindrical furnace Combusting spray in gas turbine combustors Combusting spray in DISC engine

Crowe and coworkers (197478)98 i01 Alpert and Mathews (1979) 5 E1 Banhawy and Whitelaw (1980) 12 Mongia and coworkers (1979) z'a Butler et al. (1980)`*

Anderson et al. (1981) 1

Gosman and coworkers (1981) ~3,t'*

Swithenbank and coworkers (198081)95'105'106

Gosman and Johns (1980) l7 Evaporating spray in a DISC engine Transient 2-D Dilute but considering void fraction, completely interacting phases

~lication

Evaporating sprays

Dimensionsa Spray densityb

Steady 2-D Dilute, completely interacting phases

Spray evaporation in premixing prevaporizing passages Steady 2-D Dilute, gas flow calculation neglects liquid, otherwise interacting

Combusting spray in can and annular gas turbine combustors Steady 3-D Dilute, completely interacting phases

Gas phase modelc

Variables ~b d

1,a,f, fl, Yv, k,e

1,a,e, eF, k,~,

1,a,~,T, YF, k,~

1, a,e,B,f,k,~,g

1,a,v,n, rr, Yox,

1,a,~,~,B,Y F,
Yco, YFOX, k, e 1, li, iT,if, H, Y f k-e ~ox, k, e l,u,v, ff,/~, Y r k-e

1,a,e,w, Er,~F,

1,a,~,L~
1, ti, f, L Ft~ Scalar eddy diffusivity.

1,~,e,n, eF, k,~


1, ti, ~5,/7, ~v k-e Single-step reaction with global Arrhenius rate constant None

Entropy 1,if, ~ , / t , f k-e-g 1,ff, v, I4, Yr k-e Single-step reaction with rate determined by eddy dissipation control expression of Magnussen and Hjertager 53 None Stochastic model based on doubledelta pdf with chemical equilibrium

YPR'k, e

Variables ~bwith drop source terms Turbulence model 1, if, ~, T,, Fv k-~ including effect of buoyant production on k None

1,if, O,//, Y F k-e

None Eddy viscosity or k-e

Reaction model

None

None

Single step reaction with rates found from smallest of eddy breakup and global Arrhenius expressions None

Pollutant model

None

None

None

None

Two-step reaction involving fuel and CO as reactants, Rates determined from eddy breakup model and global Arrhenius expressions None

None

Radiation model Algorithm No. spatial nodes (typ) 210-440 400

None TEACH 64

None ADD lOS

None TEACH6,*

None TEACH0,*.111

4 flux 11z TEACH.T64

6 flux 117 SIMPLE1 is 7500

6 flux 112 CORA31 ~9 3400

NO prediction by extended Zeldovich mechanism using Arrhenius rate constants None ICED_ALEI20-122 835

None RPM~2`*.125 400-450

Liquid phase modele xv, uv, Tv, d p Transient heat-up with mass transfer controlled vaporization xp, uv, Tp, mv Transient heat-up and evaporation with mass transfer controlled vaporization. No envelope flames xp, u~,,Hp, mp Transient heat-up and evaporation with mass transfer controlled vaporization. No envelope flames xv, up, Tv, mv Transient heat-up and evaporation with mass transfee controlled vaporization

Dependent variable xp, up, Tt,,d p

xl,, up, Tp, d v

Drop life history model f

Transient heat-up with mass transfer controlled vaporization

Transient heat-up with mass transfer controlled vaporization. Approximate treatment of vapor pressure of blends

xv, uv, Tp (heatup), d v Two-stage; heat-up to boiling temperature with no evaporation followed by heat transfer controlled vaporization

xv, uv, T~ (heatup), dr, Two-stage: heat-up to boiling temperature neglecting evaporation followed by evaporation at rates assuming the presence of envelope flames

xv, uv, Tv (heatup), dp Two-stage: heat-up to boiling temperature neglecting evaporation followed by heat transfer controlled vaporization

Drop dispersion model None None None None None None

Gradient diffusion using empirical turbulent Prandtl/ Schmidt number

None

Drop breakup model None None None None None None None

Stochastic drop diffusion model using k and e to estimate eddy sizes and lifetimes None None None

Drop collision model None None None None None

None

Determined employing results of Wolfe and Anderson. 96 Developed following Swinbank. 19 Also considers drops striking wall and evaporating or rebounding 5 Compared with measurements of air velocity and liquid mass flux by Heskestad et 360 None
al.1 ~ o et al. 2

No. drop classes (typ) Model evaluation 5-20 Compared with mean velocity and temperature measurements for a burning kerosene spray in a furnace by K halil et al. 126 5-100 Compared with cold and hot flow (Jet A and natural gas fuel) combustor measurements of Bruce

15 None

1250 Compared with measurements where drop evaporation was small 1

Compared with nonevaporating spray penetration measurements of Hiroyasu and Kadota 123 for ambient pressures of 0.1-5 M P a

1200 None

Comments Predictions and measurements in good agreement, but better spray characterization needed for definitive evaluation Fair agreement between predictions and measurements aside from region near injector. Strong effect of initial spray conditions on predictions.

Levels of drop evaporation were too small for definitive assessment of model

Compared with kerosene fuel furnace experiments of Founti et al.l ~3 yielding relatively poor agreement. Particle dispersion model evaluated using results of Refs. 114-116 with fair agreement Particle dispersion was found to have a relatively small effect on evaporating and combusting spray predictions Fair predictions of cold flow velocity and hot flow composition and radiation measurements

Penetration predictions were good, but only involve a test of a portion of the model

a The dimensionality indicated is that of the gas flow solution. Eulerian formulation, allowing for elliptic flow, of the gas phase conservation equations is employed for cases where the boundaries of the flow field are motionless. For the cases shown, moving flow field boundaries are treated by a Lagrangian Eulerian computational grid (a deformable grid). b Dilute two-phase flow implies that while the particles interact with the gas phase, they do not interact with each other, i.e. drop collisions are ignored, and results for single drops in an infinite media are employed to describe interphase transport rates. In some cases 4'17 the volume fraction occupied by the liquid is considered when writing the gas phase conservation equations, however, the remainder of the models assume that the volume fraction of liquid is negligibly small. c The models in this table all employ a Lagrangian solution for particle motion, tracking each drop or group of drops from the time of injection to the time of disappearance by gasification or exit from the flow field. d Major dependent variables of the gas phase conservation equations, c.f. Tables 1 and 2. Notation q5 = 1 designates the equation of conservation of mass. e Injector exit conditions are specified by dividing the liquid flow into drop classes based on ranges of initial diameter, velocity, direction, and initial location in the plane of the injector exit. Transient versions also distinguish drop classes by time of injection. 4' l or rThe drop-life-history models ignore virtual mass, Magnus, Basset, etc., effects. Gas flow around the drop is taken to be quasisteady, with negligible radius regression rate and empirical expressions for drag and convection beat and mass transfer, for mean ambient conditions. The drop temperature is assumed to be uniform, with thermodynamic equilibrium at the surface and negligible radiation. Only concentration diffusion is considered with constant gas phase properties. Aside from Mongia et al. 2'3 it is assumed that no envelope flames are present.

44

G.M. FAETH fraction for the liquid phase is assumed to be negligibly small in most models. The spray models of Butler et al. 4 and Gosman and Johns lv are an exception to this practice, where the volume occupied by liquid is considered in the gas phase conservation equations. Allowing for void fraction variations is a step toward a proper treatment of the dense spray; however, drop collisions, etc., will also have to be treated in order to obtain a complete model of this region. Therefore, the models listed in Table 9 are primary applicable to the dilute regions of the spray. The manner in which injector conditions are specified in a separated flow model has a strong bearing on the need to treat the dense spray region. Past measurements of drop size and velocity distributions have generally been made at some distance from the injector, in regions approaching a dilute spray, due to experimental limitations. In a sense this has limited the practical application of models to the dilute region of the spray in any event. 4.2.2. Gas phase model All the models in Table 9 employ a Eulerian formulation for gas motion and a statistical Lagrangian formulation for particle motion. This involves dividing the liquid flow at the injector exit (or some other location where drop properties are assumed known) into a finite number of drop classes. Each drop class is assigned an initial diameter, velocity, direction, temperature, concentration, position and a time of injection (for transient sprays). The life-history of each class is then computed throughout the flow field, generally following procedures similar to those described in Section 3. The computation for any class is terminated when the drops either disappear due to gasification or pass out of the flow field. Current gas phase models are generally restricted to dilute sprays and employ k-~ or k-e- 9 models of the turbulence; therefore, we shall restrict consideration to this case. In order to simplify the formulation we shall also only consider steady flow (Butler et al. 4 and Gosman and Johns 1~ present the gas phase formulation for transient sprays while also considering nonunity void fraction). Under these assumptions, the governing equations for the gas phase are identical to those presented in Section 2 (Eqs. (2.1) (2.6) and Tables 2 and 3), except that new source terms must be added to represent interactions with the liquid phase. The solution of the gas phase governing equations employs a finite-difference grid of computational cells. The source terms due to drops for the Eulerian gas phase solution are determined by computing changes in drop properties as they traverse each computational cell. A typical computational cell for the gas phase and several drop group trajectories are illustrated in Fig. 31. For each drop class, k, the number of drops proceeding along the class trajectory per unit time, ilk, is known from the boundary condition at the injector exit. Then the exchange rate between the phases, for any generic property ~b, is given by the

governing equations for both phases are similar, however, the formulation is inconvenient when a range of particle sizes, and effects of drop heat-up, etc. must be considered. There are also difficulties in establishing the proper representation of turbulent stresses and transport when considering a dispersed drop phase. In this section, the main features of existing DDM, CDM and C F M will be considered along with their comparison with measurements. Since spray modeling efforts to date have largely employed DDM, this approach will be considered first and in the greatest detail. 4.2. Discrete Droplet Models 4.2.1. Summary of D D M Table 9 is a summary of the properties of recent DDM. The list is by no means complete, a number of earlier models are described in previous reviews of sprays ~15 and new models are appearing with great frequency. The models in Table 9 have been selected to represent typical features of current models, a range of applications, and the extent and limitations of efforts to evaluate model predictions. The number of entries on even this limited list indicates that there is no dearth of separated flow models for evaporating and combusting sprays. It will become evident that the main limitation to the practical use of separated flow models involves gaining experience with them and properly calibrating both their numerous internal features and injector properties so that reliable predictions can be obtained. The applications considered by the models listed in Table 9 involve both evaporating and combusting sprays. Cases of evaporating sprays include drop vaporization in ducts and in the prevaporizer passages of aircraft gas turbine combustors, 1'98-11 commercial sprinkler systems for the control of fires, 5 and the evaporation period of sprays injected into reciprocating direct-injection-stratified-charge (DISC) engines. 17 Models of combusting sprays have been constructed for furnaces, lz gas turbine combustors 2'3 and DISC enginesff The dimensions of the gas phase flow solutions are indicated in the table. Models reported thus far include two- and three-dimensional steady flows as well as two-dimensional transient flows. There is no fundamental difficulty involved with extending these models to three-dimensional transient flow. Such calculations have not been reported as yet due to the important practical limitations of computer costs and storage required to obtain adequate numerical closure. It is generally assumed that the spray is dilute in these models. This implies that although particles interact with the gas phase, they do not interact with each other. Therefore, drop collisions are ignored (except for Anderson et al.a) and transport expressions determined for single drops in an infinite medium are used to estimate interphase transport rates, i.e. drop spacing is taken to be infinitely large. Consistent with these assumptions, the volume

Evaporation and combustion of sprays following :98 101


N

45

S,.~ = ~
k=l

tik[(mkC~k)i,--(mkdPk)out]i, ~

(4.1)

where i,j designates the computational cell, N is the total number of drop classes, and m k is the mass of the drops in drop class k. The variables q~ for which drop source terms are considered are summarized in Table 9 for the various models. Generally drop source terms are provided in the equations for conservation of mass (~b = 1), momentum (u), energy (/~ or [) and injected species (Y~ or f ) . Most of the D D M in Table 9 employ a k-e model of turbulence. When a higher-order turbulence model is used, drop source terms should also appear in the governing equations for k, e, 9, etc. as well. Present models neglect these effects, largely since models for predicting turbulence production, dissipation, etc. by particle motion are not highly developed. This procedure is acceptable at the limit of a dilute spray where such effects should be small due to low particle densities. Methods for treating the influence of drops on turbulence properties will be discussed in Section 6. Combusting sprays are generally modeled by assuming that drops simply evaporate with no envelope flames present. The fuel vapor is assumed to subsequently react in the gas phase with rates determined in the same manner as for gas flames, c.f. Section 2. Therefore, the drops within the spray are assumed to simply act as distributed sources of fuel vapor. E1 Banhawy and Whitelaw a2 employ a statistical formulation for turbulent reaction using a k-e-9 turbulence model with an assumed pdf for mixture fraction. The other models in Table 9 employ various phenomenological expressions for turbulent reaction rates. Mongia and coworkers; z'3 Swithenbank and coworkers, 95'15'16 and Gosman and Johns 17 employ a rate taken as the smallest of either a global Arrhenius rate or an eddy breakup turbulent reaction rate. Gosman and Johns 17 employ the Magnussen and Hjertager 53 expression given in Eq. (2.10) for the eddy breakup reaction rate, while the other models employ the Spalding 44'45 expression of
,i,j~'l GAS FLOW COMPUTATIONAL

CELL k+i~
r" - m ~ I

I
I

I I

~:t,j

I
I I I

i:l, i

k L--m TYPICAL CLAS ' ~ DROP ~ TRAJECTORY i,j -I

FIG. 31. Typical computational cell and drop class trajectories for DDM.

Eq. (2.9). Mongia and coworkers 2'3 use the most complex reaction mechanism of the cases shown, involving a two-step reaction which co,asiders the formation of carbon monoxide as an intermediate. The global kinetic parameters for this model were obtained by fitting predictions to measurements in a gas turbine combustor. Only Butler et al. 4 attempt pollutant predictions within a DDM. This involves predictions of NO, using an extended Zeldovich mechanism with Arrhenius rate constants, ignoring the effects of turbulence. Methods for predicting pollutants for turbulent gas flames are more highly developed, and these procedures could be adopted for D D M or LHF models with little difficulty, c.f. Caretto. x27 Several of the models listed in Table 9 treat radiation, following procedures developed for turbulent gas flows. The multiflux methods (specifically 4 and 6 flux) developed by workers at Imperial College have generally been employed. The effect of absorption and scattering by drops is ignored, an approximation which is only appropriate for dilute sprays. Similar to the use of Arrhenius expressions in the reaction models, mean properties are employed in the transport equations for radiation, ignoring the effect of turbulence. Naturally, this practice is questionable due to the nonlinear characteristics of the terms. The computer algorithms used to solve the gas phase equations of motion in separated flow models are generally adapted from single-phase flow models. The steady-flow elliptic codes TEACH and SIMPLE, developed at Imperial College,64'111,~ 1s are widely available. The ICED-ALE code developed at Los Alamos, 12-122 and RPM developed by Gosman and coworkers' 24.125 provide a Eulerian-Lagrangian grid scheme which is useful for the transient recirculating flows with moving boundaries that are encountered for sprays within reciprocating engines. The number of spatial nodes employed for the gas phase computations are summarized in Table 9. These grids are relatively coarse in comparison to injector dimensions and it is unlikely that numerical closure of the gas phase flow in the near injector region has been achieved in any of the cases shown. For example, the average rectilinear distance between grid nodes for the computations of Mongia and coworkers, 2'3 and El Banhawy and Whitelaw, 12 is 4 and 20mm, respectively. The other computations listed in Table 9 have grid dimensions which are similar. These distances are 1-2 orders of magnitude larger than the diameters of typical injector passages and such grids clearly cannot resolve the small scale features of the gas flow near the injector (the L H F calculations discussed in Section 2, Refs. 26-28, required 30 or more grid nodes across the local width of the spray to obtain adequate numerical closure). Therefore, averaging gas phase properties over spatial grids that are larger than a small fraction of the local width of the spray fails to provide an adequate representation of the actual environment experienced by drops within the spray.

46

G.M. FAETH realistic approximation in cases where premixed combustion does not predominate the process. Drop transport rates are strongly influenced by the selection rules used to determine average properties. This should be established by calibration of the calculations using measurements for individual drops. Unfortunately, this was not done in any of the computations listed in Table 9. Most of the models listed in Table 9 assume that the drops follow fixed trajectories, as they might in a laminar flow, and ignore turbulent particle diffusion. Exceptions to this practice are models by Jurewicz et al? ~ where a gradient drop diffusion model with an empirical turbulent Prandtl/Schmidt number is employed and Gosman and Ioannides, ~3 who employ a stochastic drop diffusion model. Turbulent drop diffusion will be considered in greater detail in the next section. We note for the present that Gosman and Ioannides ~3 find that effects of drop diffusion on spray predictions are small in comparison to uncertainties concerning the proper initial conditions of the spray for the cases they examined. Anderson et al. 1 consider other features important to the dynamics of drops in sprays including drop breakup, employing the procedure described in Section 3.5, and drop collisions. They also propose methods for treating drop collisions with surfaces. Their methods, however, have not been subjected to evaluation by comparison with experimental results. Drop trajectory calculations are undertaken for several drop classes when using the D D M approach. Each class is identified by its initial size, velocity, direction and position at the exit of the injector, and time of injection, as well, if the process is transient..For the cases listed in Table 9, known spray properties were limited to SMD and spray angle, which is not sufficient to properly specify the injector exit condition since the joint distributions of size, velocity, direction and position are required. Given the SMD, Simmons 128 provides a useful relationship for determining the drop size distribution for various injector types. However, adequately estimating initial velocities and directions is less satisfactorily resolved, requiring the use of relatively ad hoc assumptions. The problem of adequately specifying injector exit conditions constitutes the major impediment to the proper use and evaluation of separated flow models of sprays. 4.2.4. R e s u l t s In the following, we will consider a representative sample of the results of the models summarized in Table 9. This will include the steady evaporating and combusting sprays considered by Alpert and Mathews, 5 E1 Banhawy and Whitelaw, 12 and Boyson et al. 16 as well as the transient sprays examined by Butler et al. 4 Alpert and Mathews 5 considered a water spray resulting from a fire sprinkler mounted near the ceiling and injecting downward in still air. Their computations were compared with measurements of air

4.2.3. Liquid phase model The drop models employed for the computations summarized in Table 9 fall into two categories: transient heat-up models and two-stage models. The transient heat-up models are similar to the formulation of the drop-life-history computations discussed in Section 3. The liquid phase is treated by assuming a uniform liquid temperature at each instant of time (Case 2 of Table 6). The rate of mass transfer is then determined using the difference in fuel vapor concentration between the surface of the drop and its local surroundings as the driving potential. This procedure correctly represents the fact that a drop within a spray has no fixed wet-bulb temperature due to the continuous change in the properties of its surroundings. The presence of envelope flames is generally ignored and empirical expressions, found for single drops, are used to represent drop drag and the effect of forced convection on drop heat and mass transfer rates. Aside from the work of Mongia and coworkers, 2'3 the two-stage drop models involve computation of drop heat-up to a fixed wet-bulb temperature, ignoring evaporation, followed by evaporation employing the temperature difference between the drop and its surroundings as the driving potential for mass transfer. The wet-bulb temperature is taken to be the boiling temperature of fuel at the pressure of the spray. Effects of drag and forced convection are estimated using empirical relations for single drops, similar to the more complete transient heat-up models. Two-stage drop models simplify computations, but their capability for accurately representing drop processes within a spray is questionable. In high temperature surroundings, containing low concentrations of fuel vapor, drop wet-bulb temperatures approach the boiling temperature at pressures near atmospheric pressure. 6'7z However, drops spend much of their lifetime in the cool core region of the spray where gas temperatures are relatively low and fuel vapor concentrations are high. In this region, it is relatively inaccurate to estimate mass transfer rates by a heat transfer controlled driving potential. This approach also requires ad hoc assumptions concerning wet-bulb temperatures when the pressure of the spray exceeds the critical pressure of the fuel. Mongia and coworkers 2'3 employ a two-stage drop model, but use a driving potential appropriate for steady combustion in the presence of an envelope flame to estimate gasification rates during the second stage. This approach is inconsistent with their gas phase reaction model, where fuel vapor is assumed to react in the bulk gas phase (when an envelope flame is present, only combustion products reach the gas phase). Furthermore, past observations within spray flames tend to support the bulk gas phase reaction mechanism. ~9"27 The potential importance of envelope flames during spray combustion, and current capabilities for estimating their stability, will be discussed in Section 5. Based on existing information, ignoring envelope flames, similar to most of the studies listed in Table 9, appears to be the most

Evaporation and combustion of sprays velocity and liquid mass flux from a sprinkler, completed by Heskestad e t al. H The initial drop size distribution was estimated using the measurements of Dundas~ 29 for sprinkler systems. Lack of information on initial velocity and direction distributions precluded complete assessment of the model, although reasonable selections of these parameters yielded encouraging agreement with the measurements. The authors also point out that their computational grid was too coarse near the injector. The conditions of the flow considered by Alpert and Mathews s did not involve significant drop evaporation, therefore, hydrodynamic aspects of the problem were emphasized. Some typical results of the calculations are illustrated in Fig. 32. This involves calculated streamlines of the airflow induced by the spray as well as the outermost trajectories of two drop size classes and the outer envelope of the entire spray. The injector condition was a relatively coarse spray, mass median drop size of 850#m, with a hollow cone pattern. The small drops rapidly dissipate their initial momentum which causes them to collect near the centerline of the spray as they fall under the action of gravity and are deflected by the induced air motion. In contrast, larger drops penetrate greater radial distances. Since spray properties are generally measured in cold flow at some distance from the injector, effects of this type can erroneously be interpreted as stratification of the initial drop size distribution with respect to angle. This indicates the importance of analyzing even cold flow injector property measurements with a separated flow model in order to obtain a proper specification of injector exit conditions for other operating conditions. E1 Banhawy and Whitelaw ~2 completed predictions for the steady kerosene fueled combustor measurements of Khalil e t al.~ 26 This involved a water cooled cylindrical combustor with a swirling spray flame along the axis. The fuel nozzle was a swirlatomizing injector. The sensitivity of the calculations to injector exit conditions was examined by considering five cases, involving different initial distributions of drop size, velocity and direction. Figure 33 is an illustration of the comparison between the predictions (for one distribution) of mean gas temperature and axial velocity and the measurements. The discrepancies between predictions and measurements are relatively large near the injector, possibly due to the excessively large grid spacing for the gas phase solution in this region. Uncertainties in the drop evaporation rate computations, lack of adequate information on injector exit conditions, and the known limitations of k-e turbulence models for highly swirling recirculating flows were also advanced as reasons for the discrepancies. In particular, the results were found to be very sensitive to changes in estimated spray properties at the injector exit. For a given distribution, however, the solution closed rapidly as the number of drop size ranges was increased, e.g. results for 5 and 20 classes were essentially the same. Boyson e t al. 16 do not compare their predictions
Symbol N d(mm)-Outer Trajectory 0.33 1.50 Radial Axis, r (m)

47

E E

FIG. 32. Calculated streamlines of induced air flow and outer droplet trajectories for a ceilingmounted sprinkler in still air (hollow cone injection).Alpert and Mathews.5

1600

...

I~X

--

Pred.,case

1200 800

%~.. ~ ~6oo"'~-2''''I~E
4200 ~ - - - - - - ~ o 8oo
"c

,5 Io
5

o ~

.-~ o xla ~n

1600

15

o
0

0 X
II

1200
o

I0 II

400

I
I X

1200 E ,~ 800 400

,
. . . . . . .

4oo

1600 1200

, *

(~ H

4001

I0

05

0 '

FIG. 33. Radial profiles of temperature and axial velocityfor a combusting kerosene spray in a cylindricalfurnace. Predictions from E1 Banhawy and Whitelaw,~2 measurements from Khalil et al. 126

48

G.M. FAETH

q, - -

FIG. 34. Predicted velocities, temperatures and drop trajectories for combusting hollow cone spray in a cylindrical gas turbine combustion chamber. Boyson et al.16 with measurements, however, the results are of measurements of Hiroyasu and Kadota 123 with an interest as an example of computations for a practical encouraging degree of agreement. However, these steady gas turbine combustor configuration. A can- computations also involved significant uncertainties type combustor is considered with a swirling primary with respect to the specification of injector exit conair flow, concentric to a fuel injector which produces a ditions and a numerical grid that is too coarse to hollow cone pattern. Primary and secondary dilution properly resolve the near injector region. Computed results by Butler et al. 4 for a spray air streams enter the combustor via three sets of within a DISC engine are illustrated in Fig. 36. orifices, each consisting of six equally spaced holes. Therefore, the flow is three-dimensional with a Velocity vectors, particle distribution, gas temperarepeated flow pattern allowing computations to be ture and local evaporated fuel-air mass ratio are shown at a time just before ignition by a spark. Due to confined to a 60 sector of the combustor. Some typical results of the computations of Boyson the small size of the combustion chamber, more features of the spray appear to be resolved than was et al. 16 appear in Fig. 34. Predicted mean velocities, the case for the gas turbine combustor results temperatures and drop trajectories (20 initial drop appearing in Fig. 34. However, the grid is still relasize ranges and 18 angular injection locations) are shown. The flow results in a strong recirculation zone tively coarse in comparison to the flow width near the near the inlet. The grid is able to resolve the main injector exit. features of the dilution jets, however, it is too coarse to 4.3. C o n t i n u o u s D r o p l e t M o d e l s indicate the cooling effects of the spray near the Rather than represent the drops in a spray by a injector exit. For this computation condition, a large number of drops strike the combustor wall and some finite sample, as in DDM, the CDM approach describes drop properties by a continuous statistical undergo deflection by the dilution jets. Butler et al. 4 report predictions of the transient distribution function defined at all positions in the spray processes encountered in reciprocating internal flow field. Development of this method for sprays is combustion engines. Figure 35 is an illustration of due to Williams.82 CDM to date have been limited to predicted spray tip position as a function of time cases where the drops of pure liquids are evaporating following the initiation of injection from a single at a steady wet-bulb condition. The following orifice injector. The predictions are compared with the discussion will be limited to this case to simplify the

Evaporation and combustion of sprays


I I /
/ 80 Experimental 70 P = 3 MPa P= 1 MPa

49

I p = 0.1 MPa

Computed
60

Y P = 5 MPa

"-0

E c" .9

E5f////
40~ ' 30 20 10

c~ o~ a.

Scaled Cross-Section of the Spray att=3ms P = 3 MPa

. <.~'.

:%: ..c.

or
0

I
1

I
2

[
3

Time, ms

FIG. 35. Spray tip penetration as a function of time for various gas pressures. Predictions from Butler et al., 4 measurements from Hiroyasu and Kadota.~23

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ,,a ,it ,rl iilllllll, pl


. . . . . . . . . . . . . . . . . .

i, ~ ,i I

. . . . . . . . . . . . .

i r . . . . . . . . . . . ii t . . . . . . . . . . . . i I, ii, ii, lIT ii I . . . . . . . . . .


11~, iiIq' rl
. . . . . . . . . . . . . . . . . . . . .

.,=,.

,::.4 -'E=:,

. . . . . . . . . .
. . . . . . . . . .

.~.ll.c

~.~

. . . . . . . . . . . . .

(a) Velocity v e c t o r s

(b) Particles

=1.6

4= 14.0
L----

(c) Temperature

contours
et al. 4

(d) Fuel-air ratio contours

FIG. 36. Velocity, particle, temperature and fuel vapor distributions predicted for a DISC engine. Butler

formulation, although these restrictions are removable. F o r C D M , the spray is described by a generalized distribution function fp, where f p ( r r, xp, up,
JPECS 9-I/2-D

t)drpdxpdup is the probable number of drops with radii in the range drp about rp, located in the volume dxp about xp, with velocities in the range dup about up, at time t. An equation governing the time rate of

50

G.M. FAETH 4.4. Continuum Formulation A third method for separated flow analysis of sprays is the continuum formulation model (CFM). CFM treat drops as a continuous fluid or droplet phase, so that the gas and drop phases are described as interpenetrating continua which interact with each other. There has been substantial recent progress in modeling multidimensional two-phase flows using CFM. 135-137 The method is also widely employed for analysis of safety margins in nuclear reactors/39 The approach yields a particularly concise formulation when a single fluid phase is adequate to describe the process. For sprays, however, multiple fluid phases must be considered based on ranges of drop sizes. Past experience indicates the CFM become unwieldy and costly in terms of computer time and storage requirements when multiple phases are needed. 4'13 In particular, CFM require small computational grids for all phases near the injector exit in order to reduce numerical diffusion errors. This only compounds difficulties already present for the gas phase solution. Application of D D M at least eliminates numerical diffusion problems for the liquid phase portion of the calculation. Travis eta/. 138 discuss the relationship between CFM and CDM. They show that the multifluid equations correspond to moments of the spray equation. 4.5. Summary Existing separated flow models for evaporating and combusting sprays have been described. The models can be grouped into three classes, discrete droplet models (DDM), continuous droplet models (CDM) and continuous fluid models (CFM), based on the method used to describe the liquid phase. DDM have been favored by recent workers since this approach provides a convenient formulation for considering complex drop transport problems, minimizes computer storage and time requirements, and eliminates problems of numerical diffusion in the near-injector region for the liquid phase solution. Separated flow models for sprays are appearing with great frequency. Most of these formulations have been constructed for the limit of a dilute spray using the DDM procedure. This involves computing droplife-histories assuming a uniform liquid temperature and allowing for heat-up and mass transfer controlled vaporization. Generally, the presence of envelope flames and transport phenomena associated with high pressures and turbulent fluctuations have been ignored for drop transport computations. The transport of mass, momentum and energy between the phases has been considered in most recent models. The gas phase is usually represented by higher-order turbulent transport and reaction models, with the governing equations solved using an elliptical flow algorithm. Computations with these models and limited evaluation by comparison with experimental results have revealed several major problem areas. The foremost difficulty involves the strong sensitivity of the

change of fp can be derived following procedures analogous to kinetic theory of gases, as follows :82 ~fp

Ot ~-~rr(Rpfp)+ V~. (upfp)+ V~- (Fpfp) = Sp


(4.2)

where

Rp

drp dup = d ~ - ' F P - dt

(4.3)

and Sp is a source term giving the net increase offp due to drop breakup, nucleation and collisions. At the limit of a dilute spray, the expressions for Rp and F r can be obtained as described in Section 3.2.3. Phenomenological expressions can also be developed for Sp, 1's2 although this term is often ignored. The formulation is then completed by defining boundary conditions at the injector exit and the boundary surfaces of the flow. The solution for spray properties involves integration of Eq. (4.2) along with the governing equations for the gas phase. This presents serious computer storage problems due to the large number of dimensions required to specify fp. For example, fp involves eight dimensions for a three-dimensional transient problem, even under the simplifying assumptions employed to find Eq. (4.2). This problem is compounded when additional factors, such as drop temperature and composition variations are considered, which increases the dimensions of fp. Furthermore, drop densities, velocities, etc. vary rapidly in the near injector region imposing requirements for closely spaced grids for all the dimensions offp if numerical diffusion is to be avoided--difficulties that are circumvented by the Lagrangian formulation for the liquid phase in DDM. These factors have attracted most workers to DDM in recent years. DDM and CDM naturally only represent different ways of describing a spray and both methods yield identical results for comparable transport assumptions. In fact, fp = AN/dr dxp dup, where AN is the number of particles contained in the volume drdxpdup for a statistically significant sample using the DDM approach. 13 This relationship highlights the computational problems with CDM--since it clearly requires finer grids and more computations to define the gradient of the drop density in a multidimensional space than it does to determine the drop density itself as a function of position. Past applications of the CDM have employed relatively simplified representations of spray phenomena, generally neglecting drop heat-up and blended fuels. The earliest CDM also ignored the influence of drops on the gas motion. TM 133 More recently, Haselman and Westbrook 134 have performed fully coupled computations for a two-dimensional representation of a constant volume cylindrical combustion chamber with injection at various locations. In general, these models have not been compared with experiments. Gany et al. ~35 also report computations using this method.

Evaporation and combustion of sprays computations to drop size, velocity and direction distributions at the exit of the injector. No evaluation of separated flow models to date has involved adequate information concerning these spray properties. Thus the computations become largely an exercise to determine whether reasonable assumptions concerning spray properties can yield flow properties which correspond to the measurements. In a sense, this is hardly more than a sophisticated correlation procedure. Real progress toward the development of practical spray modeling procedures, useful for design, depends on the development of convenient methods of completely characterizing spray properties. Fortunately, recent progress in the application of optical techniques for drop size and velocity measurements suggests that this requirement may be satisfied in the near future? 4 A second difficulty with past computations using separated flow models is that adequate numerical closure has not been achieved in the near injector region, due to limitations of computer storage and costs. The use of D D M efficiently controls this problem for the liquid phase solution, however, adequate numerical closure and acceptable levels of numerical diffusion still require a fine grid for the gas phase solution near the injector. This becomes particularly critical for frequently encountered cases where the spray is confined to a relatively narrow boundary layer region, which cannot be efficiently handled by elliptical flow formulations. Modeling such regions using boundary layer formulations, which require only the solution of parabolic equations, is an obvious and well-established solution to this problem. This implies a need for greater application of the modular modeling approach suggested by Harsha and Edelman, 141-142 for separated flow computations for sprays. Another deficiency of existing separated flow computations is the absence of independent calibration of drop transport models. Reasonable changes in selection rules for average properties can result in an order of magnitude variation in predictions of drop transport rates. 6 Calibration of such predictions, using measurements in environments representative of conditions experienced by drops within a spray, is an obvious requirement of a reliable computation procedure. There are also uncertainties in other aspects of separated flow models for sprays. Major problems include dense spray phenomena, interactions between drops and turbulence structure, turbulent diffusion of drops, effects of envelope flames, collisions of drops with each other and with surfaces, and adequate modeling of swirling and recirculating flows that are frequently encountered in practical spray processes. It can be concluded that while capabilities for completing separated flow calculations for transient multidimensional spray evaporation and combustion processes have clearly been demonstrated, much remains to be done before such models become reliable design tools.
5. DILUTE SPRAY PROCESSES

51

5.1. Introduction The major assumptions of past separated flow models of sprays tend to be similar, c.f. Table 9. In most cases, the spray is assumed to be dilute, turbulent dispersion of drops is ignored, drops are assumed to evaporate with no envelope flames present, and the influence of dynamic effects and turbulence on drop transport rates is neglected. Progress toward determining the adequacy of these approximations, and eliminating them where appropriate, is discussed in the following. Consideration is given first to the dilute spray approximation, emphasizing methods of identifying the region of a spray where the approximation is realistic. Approximations which are frequently employed for analysis of dilute sprays are then considered, including turbulent dispersion of drops; the impact of envelope flames on drop transport rates and methods for estimating their appearance (ignition) and stability; and the effect of dynamic processes (drop and flow acceleration, etc.) and turbulent fluctuations on drop transport rates. Consideration of dense spray processes, e.g. drop interactions, effect of drops on turbulence properties, drop collisions, etc. will be deferred until Section 6. 5.2. Dilute Spray Region The dilute spray approximation generally implies that while particles interact with the gas phase, direct interactions between drops can be neglected. Therefore, single-drop formulas are employed to describe drop transport rates, the presence of drops is neglected in the constitutive equations describing the transport characteristics of the gas phase, and drop collisions are ignored. At present, the liquid fraction requirements for these assumptions are not well defined. Of the several phenomena that must be considered, the effect of inter-drop spacing on drop transport rates is best understood. For monodisperse particles in a convective environment, earlier work has shown that singleparticle transport formulas apply reasonably well if the center-to-center distance between particles is greater than two particle diameters. ~7 Assuming spherical, close-packed particles with equal particle spacing, this requirement implies a liquid fraction less than 0.08. The dilute spray region should be examined on a case-by-case basis, since real sprays generate nonuniform liquid flux distributions and quality of atomization varies with operating conditions. A general indication of the dilute spray region can be obtained, however, if simplifying assumptions are accepted. Results of this type are developed in the following for pressure atomized sprays. Pressure atomized sprays are of particular interest since the liquid fraction is unity at the injector exit; therefore, they tend to have larger dense spray regions than air atomized sprays. We take a simple one-dimensional viewpoint,

52

G.M. FAETn sided injector passages having large length-todiameter ratios and no swirl. In this case, spray angles in the range 12-25 have been observed for injection into stagnant gases (the latter angle roughly corresponds to the natural angle of spread of a fully developed constant density jet). 57'143 For this range of spray angles, values of~ I < 0.08 are reached for x/d of 6-12. Larger spray angles are obtained using diverging injector passages or swirl, yielding a corresponding reduction in x/d required to reach liquid fraction values appropriate for the dilute spray region. For combusting solid cone sprays in a stagnant environment, the region where drops are present typically extends to values of x/d on the order of 100, c.f. Figs. 20, 28-29. If the results of Fig. 38 are reasonable, this suggests that the dilute spray assumption is adequate for the bulk of the two-phase portion of the flow. This conclusion is reached with some reservations. Spray angles are generally based on measurements completed much farther from the injector than x/d of 6-12 and similar to single-phase jets the flow tends to be narrower near the injector. Furthermore, assuming uniform liquid flux and constant liquid velocity tends to underestimate the distance required to reach a given liquid fraction. 5.3. Turbulent Drop Dispersion Turbulent dispersion of drops has generally been neglected in past separated flow models of sprays (Refs. 101, 103, 130 are exceptions). Neglecting dispersion implies that drops follow deterministic trajectories prescribed by their initial condition at the injector exit and mean gas properties throughout the flow field. In contrast, drops are assumed to disperse due to turbulence at the same rate as the gas phase in locally homogeneous flow models. Actual behavior lies between these two limits, depending on the inertial properties of the drops and the turbulence properties of the gas phase. Techniques for treating turbulent dispersion are rapidly developing and some representative methods are discussed in the following. The effect of various approximations for treating turbulent particle dispersion is illustrated in Fig. 39, utilizing the measurements of Yuu et al. 145 Experimental conditions consisted of a dust laden round air jet, having an initial diameter of 8 mm, in stagnant air. The dust consisted of nearly monodisperse slag particles having a mass mean diameter of 20 pm with a standard deviation of 2.5 #m. The flow was highly dilute, since solid fractions in the injected flow were in the range 0.4-2 10 -6. Measurements of the radial variation of particle concentration are compared with predictions for a L H F model, a deterministic separated flow model where turbulent particle dispersion is ignored, and a separated flow model which incorporates a stochastic turbulent dispersion model to be described later. All three models assume that both phases were directed axially at the exit of the nozzle. The fact that the particles had a lower velocity than the gas at the nozzle exit was considered for the

~x

SPRAYBOUNDARY ~ "
FIG. 37. Injector geometry for liquid fraction estimations. defining the total area of the two-phase flow by the spray angle. The liquid flow is assumed to be distributed uniformly over the cross-section of the spray, evaporation is neglected, and the velocity of the liquid is assumed to be equal to the injector exit velocity throughout the high liquid fraction region. The assumption of constant liquid velocity is partly justified by the drop-life-history computations illustrated in Figs. 26-30. These results show that drops tend to maintain their original injected velocity in the near-injector region. Liquid fraction is defined as the ratio of the liquid and total flow areas for a one-dimensional flow. 16

O~f Af/(Af -{-Ag).


=

(5.1)

We examine the variation of this parameter for the solid cone spray configuration illustrated in Fig. 37. Since the liquid velocity is assumed to be constant, and evaporation is neglected, the liquid phase flow area remains equal to the injector exit area. In contrast, the total flow area increases with distance from the injector exit. Simple geometrical considerations, based on the notation illustrated in Fig. 37, yield the following expression for liquid fraction ~I = [1 + 2(x/d) tan (0/2)] - 2 (5.2)

where 0 is the spray angle. The estimated variation of liquid fraction with axial distance from the injector is illustrated in Fig. 38 for solid cone sprays. Results for spray angles from 5-80 are shown. Spray angles are narrowest for straight| , 0

.
"o

tO x/d

100

FIG. 38. Estimated liquid fraction variation in the axial direction for a solid cone spray.

Evaporation and combustion of sprays


I r | ! I

53

Theory; x/d = - - 50 ---20

El Data; Yuu,et al.,x/d=20-50

.4

I
y
t X

no
~
~ ~'%

\E)
x\
% ~ r stochast ic ~v x /)'diffusion O~ Up=30 m/s

\'.
Uo, 50 m/s 0 0 I I .04 I r/x

"ab..
I .08 I .12

FIG. 39. Turbulent particle diffusion in an axisymmetric jet. Predictions from Shuen et al., 144 measurements from Yuu et al. ~4~ separated flow models. A k - e gas phase model was employed, similar to the model used by Shearer e t al. 26 The flow was sufficiently dilute so that the presence of the solid phase had a negligible influence on gas flow properties. Of the three predictions illustrated in Fig. 39, the separated flow model which incorporates a stochastic turbulent diffusion model agrees best with the measurements. The deterministic separated flow model underestimates the radial spread of the particles and also indicates that the particles are limited to a progressively narrower range of r / x with increasing distance from the nozzle exit--which is contrary to observations. The difficulty with deterministic models is that they treat particle motion as equivalent to motion in a laminar flow. Therefore, particles only move in the radial direction due to the drag forces exerted by the radial component of velocity and they tend to accumulate in regions where the radial velocity component is zero. Soo ~7 presents a number of examples of the tendency for laminar motions to concentrate particles in regions where the transverse velocity component is zero. For the conditions illustrated in Fig. 39, the locally homogeneous flow approximation overestimates the radial spread of the particles. This will not always be the case since the effective turbulent diffusivity of particles can be greater or less than the gas, depending on the inertial properties of the particles and the structure of the flOW. 114 These findings support the need for reliable methods to predict particle dispersion during separated flow computations. Two main methods have been employed to compute particle dispersion in sprays. The earliest investigations modeled turbulent dispersion using a gradient diffusion expression which employs a turbulent gas/particle exchange coefficient (or equivalently, a gas/particle turbulent Prandtl/ Schmidt number). ~~'~48 Some method of estimating the gas/particle turbulent exchange coefficient is

required. Hinze, 114 Yuu e t al. x~5 Goldschmidt e t al. t46 and Lilly 147 review theoretical and experimental studies of gas/particle exchange coefficients. Existing continuum analysis for gas/particle exchange coefficients has a limited range of application, since the process must be substantially simplified in order to obtain closed form solutions.114 The main difficulty is that dispersion is influenced by both particle and turbulence properties, adding a whole new dimension to the problem of estimating a turbulent exchange coefficient. Lack of theoretical guidance on values of gas/ particle exchange coefficients is compounded by difficulties with existing measurements. Yuu e t al. 145 point out that measurements are often suspect, since the experiments were analyzed assuming that mean gas and particle velocities were the same and that turbulent gas/particle exchange coefficients were constant over the cross-section of the flow. These requirements are rarely satisfied in practice. Even in cases where the experiments were properly analyzed, it is unlikely that the results could be directly applied to a particular spray since both particle and turbulence properties must be the same, because both sets of characteristics influence dispersion properties. Gradient diffusion models of particle dispersion are not particularly convenient for incorporation in the Lagrangian particle motion calculations which are employed for the popular D D M approach. Jurewicz, Stock and Crowe 11,148 propose circumventing this problem by representing the effect of particle diffusion as an effective drift velocity which is found from the gradient diffusion expression. Unfortunately, this introduces errors due to numerical diffusion which compromises one of the advantages of the Lagrangian formulation. Hotchkiss and Hirt, 149 cited by Dukowicz, 13 propose another method for employing estimated turbulent gas/particle exchange coefficients to compute particle dispersion. In this case, the turbulent exchange coefficient is related to the variance of the probability density function of particle position (assumed to be Gaussian) at the end of each computational time step. Given the variance, the distribution is randomly sampled to obtain the diffusional increment of particle position. This procedure does not provide any real advantages over the stochastic methods to be discussed next, unless the exchange coefficient can be prescribed very simply. The method has not been applied to separated flow computations for sprays. Recent analyses of particle dispersion in sprays, e.g. Gosman and Ioannides 13 and Butler e t al. ~ have employed stochastic methods which involve modeling particle dispersion directly. A turbulence model provides the instantaneous properties of the environment of a particle during a Lagrangian computation of particle motion. This involves random sampling to determine gas flow properties in addition to considering a representative number of particle trajectories. Past computations with stochastic particle dis-

54

G.M. FAETH assumed to interact with an eddy for a time taken as the smaller of either the eddy lifetime, te, or the transit time, tt, required for the particle to traverse the eddy. The eddy and transit times are determined by assuming that the characteristic size of the randomly sampled eddy is the dissipation length scale, Le, given by
L e = C~/4k3/2/~.

persion models have used a variety of turbulence models. Yuu et al. ~45 employ empirical correlations of turbulence intensities and length scales for their computations of particle dispersion in jets. Gosman and Ioannides, 13 Shuen et al. 144 and Buckingham and Siekhaus ~5'151 employ a k-e turbulence model. Peskin and Kau 152 use a more sophisticated approach, where a large eddy simulation is utilized to represent the gas phase during a theoretical study of particle dispersion in a channel. The main features of the stochastic particle dispersion model of Gosman and Ioannides 13 are discussed in the following. This method employs a k-e turbulence model and is easily adapted for use in most current D D M of sprays. Furthermore, this model has been employed to estimate the effect of dispersion in evaporating and combusting sprays, yielding results of some importance to spray modeling. Gosman and Ioannides ~3 compute the motion of a particle using the Lagrangian formulation described in Section 3. The density ratio p f / p is assumed to be large; therefore, virtual mass, Bassett forces, etc. are neglected. For particle dispersion calculations, the instantaneous gas velocity replaces the mean velocity in the governing equations. As particles pass through the flow, they are assumed to interact with individual eddies as indicated in Fig. 40. Flow properties within each eddy are assumed to be constant. Each interaction deflects the particle as dictated by the instantaneous eddy velocity. Therefore, the particle trajectory is determined similar to a random walk computation until the particle passes out of the region under consideration. Overall particle behavior is obtained by averaging over a statistically significant number of particles. The instantaneous gas velocity within an eddy is obtained from the mean gas velocity and k, which are known from the solution of the governing equations of the gas phase. The turbulence is assumed to be isotropic with fluctuating components having a Gaussian distribution for this computation. The standard deviation of the distribution is taken to be (2k/3) ~/2. The distribution is randomly sampled when a particle enters an eddy to obtain the instantaneous velocity as u = fi + u'. Following Brown and Hutchinson, ~53 a particle is

(5.3)

The eddy lifetime is then estimated as


te = L J l u ' J .

(5.4)

The transit time was determined from the following equation


tt = -zln(1-Le/(Zlu-up]))

(5.5)

where z is the particle relaxation time


z = ~p dp/(pCD ]u -

up[).

(5.6)

Equation (5.5) is found by solving the linearized equation of motion of a particle in a uniform flow. In both Eqs. (5.5) and (5.6), u - u e refers to conditions at the start of the interaction of the particle with an eddy. When L~ > rlu-up[, Eq. (5.5) has no solution. This can be interpreted as the eddy having captured a particle so that the interaction time becomes te.103 The stochastic dispersion model predictions were checked by comparison with both exact analysis for simplified conditions and measurements. Comparison with analysis involved the dispersion of "marked" fluid particles introduced at a constant rate from a point source into a homogeneous and isotropic turbulent flow, with long diffusion times. Hinze 114 provides an analytical dispersion expression for this situation, assuming a constant turbulent diffusivity. Comparable results for the stochastic model were obtained by fixing values of ~, k and e and computing the turbulent diffusivity of the analytical solution as C~pk2/e, which is consistent with the turbulence model. The comparison between the analytical and stochastic predictions is illustrated in Fig. 41. The stochastic predictions of Shuen et al. 44 are shown along with those of Gosman and Ioannides 13 on the figure. The two models are identical, except that the former evaluates t e a s
t e = L J ( 2 k / 3 ) 1/2.

(5.7)

ED O,

/3

-X

FIG. 40. Sketch of the particle trajectory model.

In both cases, roughly 1000 particle trajectories were computed in order to obtain overall particle behavior. The particles were assumed to be small, to correspond with fluid particles; therefore, all interaction times are given by te. The stochastic predictions of Shuen et al. ~44 are in reasonably good agreement with the analysis of Hinze, a14 while the original proposal of Gosman and Ioannides 103 tends to underestimate the rate of dispersion. The stochastic model was also compared with the measurements of Snyder and Lumley. ~~s In this case, the dispersion of solid particles in a uniform turbulent flow downstream of a grid was measured. The comparison between predictions and measurements is

Evaporation and combustion of sprays


I I I I I I I

55

1.5 ~ \ / ,"/

STOCHASTIC MODEL
te FROM:

GOSMAN IOANNIDES ~ PRESENT STUDY ANALYTICAL SOLUTION

1.0

0.5

I ~'''--~'r''-'--

indicates that most recent studies have used some form of stochastic dispersion model. This approach is readily adapted for D D M of sprays and provides a convenient means of handling nonlinear interaction effects, e.g. high Reynolds number drag, etc. The most elegant computations employ large eddy simulations. 152 However, existing computations with k-~ full-field models have been encouraging, 13'144 which should motivate further development of these methods. Results thus far suggest that drop dispersion does not have a large effect on the overall properties of evaporating and combusting sprays, particularly in comparison to uncertainties in the specification of injector exit conditions. 13 Further assessment of this conclusion is needed. Presumably, as the accuracy of injector specifications improves, accurate calculation of particle dispersion properties will become more important. 5.4. D r o p Ignition and C o m b u s t i o n 5.4.1. I m p o r t a n c e o f envelope f l a m e s Whether envelope flames are present around individual drops in a combusting spray has generated much discussion in the literature. When an envelope flame appears, drop heat and mass transfer rates are increased from the values found for evaporation without combustion in the flow field around the drop. The presence of envelope flames also modifies the mechanism of fuel oxidation. An envelope flame has a diffusion flame structure where the fuel vapor reacts to yield combustion products. Therefore, drops act as a source for combustion products for the bulk gas phase when envelope flames are present. In contrast, in the absence of envelope flames, drops act as a source of fuel vapor for the bulk gas phase. In this case, the vapor eventually reacts in the bulk gas phase similar to a gas fueled flame. The presence of a diffusion flame around individual drops also modifies the mechanism of pollutant production in spray flames. In particular, envelope flames have been proposed as the cause for increased NO and particulate production in spray flames.154-~56 Most separated flow models of combusting sprays reported thus far have ignored envelope flames, c.f. Table 9. There is justification for this approach in some circumstances. Onuma et al. ~9 and Mao et al. 2v
I GLASS =COPPER I I I HLOLLOW~ .,~" I

i0

20 r (mml

30

40

FIG. 41. Analytical and numerical solutions for the transverse distribution of infinitely small particles from a point source. From Shuen et al. 144 illustrated in Fig. 42. The agreement is encouraging. Both models yield the same results for heavier particles, since interaction times are largely controlled by particle transit times which are computed in the same manner. Shuen et al. 144 also obtained good predictions of the particle laden jet experiments, discussed earlier in reference to Fig. 39. Gosman and Ioannides ~3 applied their stochastic dispersion model to the evaporating spray measure-ments of Tishkoff et al. ~ 6 and the combusting spray measurements of Founti et al. ~13 While turbulent fluctuations and dispersion were considered in the equations of motion, mean properties or a fixed evaporation rate constant were employed for other transport quantities. Uncertainties in injector exit conditions limited quantitative assessment of the entire separated flow model for the sprays. The main conclusion for the two cases considered was that the effect of dispersion is small in comparison to the effect of uncertainties in the distributions of drop size, velocity and direction at the injector exit. This brief consideration of particle dispersion
l I l

MEASUREMENTS,SNYDER8~ LUMLEY
0 HOLLOW GLASS 400 ACORN POLLEN

r~
-

PRED,CT,O.S
~ A N N I D E S __ __

/
f ~ . F l

o j

-I
.~./~.......-

1~

zOo

. v ~ / ~

-'~ /

~"

"'CORN

POLLEN"

~i"~lg~-~

I00

200 t(ms)

300

400

FIG. 42. Particle dispersion in a grid generated turbulent duct flow. From Shuen et al. 1.4

56

G.M. FAETH 5.4.2. Drop Combustion Theories Drop transport in the presence of envelope flames has been the subject of numerous investigations, In the following, a brief account of envelope flame theory will be presented and the results will be employed to indicate the effect of envelope flames on drop transport rates within a spray flame environment. Further discussion of the properties of envelope flames can be found in recent review articles. 6 15 The analysis of envelope flames will employ the same assumptions as the drop evaporation analysis of Section 3, except as noted in the following. We assume that fuel and oxidizer react in stoichiometric proportions within an infinitely thin flame sheet which completely surrounds the drop. Reaction rates within the flame sheet are assumed to be infinitely large, so that the concentrations of fuel and oxidant are zero at the flame. In order to reduce the complexity of the transport equations, the same average properties are employed on both sides of the flame sheet and the Lewis number is taken to be unity. These approximations are typical of most models of envelope flames. 6-15 The property assumptions are easily removed if required. The structure of the gas phase near evaporating and combusting drops, under present assumptions, is sketched in Fig. 43. The same surface and ambient conditions are shown for the two cases. It is evident that the presence of an envelope flame increases temperature and concentration gradients at the drop surface, enhancing drop transport rates. The objective of the analysis is to provide drop heat and mass transfer rates given conditions at the drop surface and in the surrounding gas. These results can then be incorporated in drop-life-history computations. A multicomponent liquid phase is considered. Following Law et al. s l the overall stoichiometry is taken to he
ei[F,] + ~ (~i/v)[0] ~ Products
n n

observe appreciable quantities of unburned gaseous hydrocarbons within the core of well-attached spray diffusion flames, suggesting that most drops are simply evaporating with no envelope flames present at least for these test conditions. The drop-life-history computations for attached spray flames discussed in Section 3 also support the view that most drops evaporate before reaching regions of the flame where appreciable concentrations of oxygen are present. Komiyama et al. 2 considered the stability of envelope flames for their tests of ducted spray flames using both air and pressure atomizing injectors. Based on existing correlations for envelope flame stability, they conclude that envelope flames were not present for their test conditions. Chigier and McCreath 157 also concluded that envelope flames were not present for their open spray flame tests, based on photographic evidence. While the absence of oxidant precludes the existence of envelope flames, there are many instances where drops evaporate within a spray environment which contains oxygen and envelope flames could be present. For example, large drops can enter the flame zone with air in their immediate environment in premixed/prevaporizing spray combustors. Prior to ignition in diesel combustion chambers, air and drops are mixed, resulting in an initial period of premixed combustion. Spray flames in steady combustors are not always attached, yielding regions of premixed combustion of air and drops. Finally, even if the flame is attached and the mean concentration of oxidant is small, drops passing through a spray can encounter oxygen rich eddies. Whether envelope flames are present during such interactions has an important bearing on the selection of proper mean ambient conditions for estimation of drop transport rates, the mechanism of fuel oxidation, and the production of pollutants. Clearly, envelope flames can be an important aspect of spray combustion processes. In the following, some of the main properties of envelope flames are considered, including their effect on drop transport rates, their stability and their initiation (ignition).

(5.8)

where ~i is the mass flux fraction of i, defined in Eq. (3.7). Effects of forced convection are treated using the

EVAPORATION

~T
ENVELOPE ,~FLAME r rp r

rp

FIG. 43. Sketch of gas phase structure for evaporating and combusting drops.

Evaporation and combustion of sprays empirical expressions of Eq. (3.16). The quasisteady transport equations can be solved for Re = 0, where they have spherical symmetry, to yield
a | |

57

rhCp -ln[(l+By)(l+Bo)], 2rcdp.~


where

Re=O

(5.9)

~2 E

~ ~ ~( l o p e t flame observed

By
Bo=

= (1

YJ/Y~

(5.10) (5.11) (5.12)


t~

Yo~/(~ ~,lv)

heory,dp=435~m

ei = Y~(I+By)/By.

The quantity Y,~ is the mass fraction of nonfuel species at the drop surface. This quantity is obtained from the equation of state for liquid/vapor equilibrium, knowing the liquid temperature and composition and the pressure. Basing the heat transfer coefficient on the flame temperature,

O
0 data, dp=420-450pn

e~

0
solution of the transport equations yields

I I 1 equivalence ratio

hd~ = 2 1 n 2 By

[(1 +By)(1 +Bo)],

Re

= 0. (5.14)

FIG. 44. Steady evaporation rates of JP-10 drops within an open turbulent propane diffusion flame. Data from Szekely andFaeth.9 suitable reference state for computing average properties. The present formulation has been tested for pure fuels in combusting gas environments having particular relevance to analysis of drop-life-histories in spray flames. This includes measurements for supported drops in laminar and turbulent flames for pressures in the range 0.1-6.8MPa. 61'89'9 Other comparisons are discussed in recent review articles on the.combustion of drops and sprays. 6 15 As a first example we consider the measurements of Szekely and Faeth 9 for JP-10 drops evaporating at various positions along the centerline of an open turbulent propane flame in still air. A portion of this data was discussed in Section 3 and is summarized in Table 7. More complete results are reported in Fig. 44, where the burning or evaporation rate constant is illustrated as a function of the fuel equivalence ratio of the bulk gas phase. Predictions for noncombusting and combusting drops are also shown on the figure. For equivalence ratios greater than unity, the ambient oxygen concentration is small and both theories yield the same burning rate constant, which is in good agreement with the data. As the equivalence ratio decreases below unity, the theories begin to diverge. In the region where envelope flames were observed, the combusting drop model yields best agreement with the data. As the equivalence ratio continues to decrease, however, envelope flames are no longer stable and the measurements shift to the predictions for evaporating drops. The results indicate that the models give reasonable predictions for both evaporating and combusting drops in sprays, as long as the appropriate branch of the solution can be selected for equivalence ratios less than unity. Gasification rate predictions for both evaporating

Neglecting dissociation, the flame temperature is ~-~ = LC,(To _ T~) t-1

I+~r(By+Bo)
(5.15)

where Qr is the heat released per unit mass of oxidizer consumed at the flame front.

Qr=~eiQ~/(~i/v).

(5.16,

If an envelope flame is present, Eqs. (5.9)-(5.15) replace Eqs. (3.11) and (3.12) when determining drop heat and mass transfer rates. Otherwise the computation of drop-life-histories is the same as for evaporating drops. In cases where the surroundings of a drop contain oxygen, but a stable envelope flame is not present, drop transport rates can be approximated by setting Yo~ = 0 in Eqs. (5.9)-(5.15) which retrieves the formulation for an evaporating drop. At moderate pressures, single component drops in a fixed surroundings reach a steady wet-bulb temperature (aside from minor effects of non-unity Lewis number which have been ignored in this analysis). In this circumstance, the rate of evaporation is given by

2rcdv2

rhCp _ In [1 +

(C,(T~ - T~)+ QrYo~)/heo], Re = 0


(5.17)

where T~can be approximated by the boiling temperature of the fuel with little error. 7z Numerous comparisons between predictions of envelope flame models and measurements have been reported. 6-~2 In general, models similar to Eqs. (5.9)(5.17) provide a satisfactory method for correlating data once the results are calibrated by selecting a

58

G.M. FAE~

Re:O 2.4

EVAP.

,~"

O 0

~ .04 ,08 .12 .16 .20

FIG. 45. Influence of the presence or absence of an envelope flame on the steady evaporation of drop within a combusting n-pentane spray. and combusting drops are illustrated for a wider range of mixture fractions in Fig. 45. Drops are considered in a combusting n-pentane spray, taking gas properties as a function of mixture fraction from Fig. 3. Steady evaporation rates were computed for a drop having a negligible velocity relative to the gas phase. In this case, the gasification rates of both evaporating and combusting drops agree within 10~o for fuel equivalence ratios greater than 0.5. As the mixture fraction decreases below this value, however, it becomes increasingly important to determine whether an envelope flame is present. 5.4.3. Stability of envelope flames The stability of envelope flames, which provides a necessary condition for their existence, will be considered in this section. Drop ignition, which provides a sufficient condition for envelope flames, will be considered in the next section. Attention will be limited to diffusion flames, e.g. a fuel droplet in an oxidizing atmosphere, since this case is of greatest interest for current spray combustion models. We have seen that the quasisteady flow approximation is usually acceptable for the analysis of drop processes. Recognizing this, most theoretical studies of envelope flame stability have adopted the quasisteady flow approximation in order to simplify the analysis. Under this assumption, a drop continuously fed with fuel is being considered, which is a configuration that has often been employed for measurements of envelope flame stability. Analysis has also been generally limited to a one-step reaction using Arrhenius kinetics. Extinction occurs when the rate of reaction is not sufficient to consume the reactants transported to the flame zone and also provide sufficient chemical energy release to maintain high flame temperatures in spite of energy transport from the flame zone. Damk6hler 158 pointed out that' the balance between transport and reaction rates, which is crucial to flame stability, can be conveniently represented as the ratio of a characteristic diffusive or convective time to a characteristic chemical reaction time (the Damk6hler number). Therefore, most subsequent analyses have sought critical Damk6hler numbers for extinction. The flow configuration has a strong influence on the characteristic diffusive or convective time. Several cases have been considered for envelope flames including: the spherically symmetric flow around a motionless drop in the absence of body forces; the forward stagnation point of a moving drop, which represents the region of envelope flame attachment for forced convection conditions; and the lower stagnation point of a motionless drop in a gravitational field, which represents the region of envelope flame attachment for natural convection conditions. Forced convection conditions are of greatest interest for modeling spray combustion processes; however, spherically symmetric flow provides a useful limit for low Reynolds numbers. Furthermore, an understanding of extinction in the presence of natural convection is valuable for interpreting experimental results. Therefore, all three cases will be considered in the following. Analysis of extinction for spherically symmetric flow has been undertaken by Lorell e t al., 159 Agafanova et al. 16 and Polymeropoulis and Peskin, 161 using numerical methods. A problem with

Evaporation and combustion of sprays numerical solutions, however, is that it is difficult to distinguish between true extinction and numerical instabilities which are often present near limits. Numerical methods also do not provide much guidance concerning simplified methods for correlating stability results for use in spray combustion models. Finally, computations are costly and these studies only examined a limited range of parameters. Some investigators have employed approximate analytical methods in order to reduce the cost of computations and avoid numerical stability problems, e.g. Tarifa et al. 162 and Peskin and coworkers. 163 164 In this case, reaction is confined to a limited region, whose thickness and reaction properties are prescribed by a closure approximation. The objection to this approach is that the closure approximation introduces uncertainties in the results which are difficult to evaluate. The most extensive results concerning envelope flame stability have been obtained using the method of matched asymptotic expansions at the limit of large activation energy for the global reaction. In particular, Law x65 has shown that the analysis of Lifian 166 can be adapted to diffusion flames around drops. This approach yields relatively simple expressions for the extinction criterion, useful for data correlation and parametric studies, with mathematical rigor--at least at the large activation energy limit. Analysis of extinction for forced convection conditions has been limited to large Reynolds numbers where boundary layer approximations apply at the forward stagnation point. Sami and Ogasawara, 167 describe a semi-empirical analysis of this problem using film theory ideas. Most work in this area, however, has employed the method of matched asymptotic expansions at large activation energies. This includes studies by Fendell, ~6s Krishnamurthy eta/. 1 6 9 and Wu et alJ 7 The method of matched asymptotic expansions has also been applied to envelope flame stability under natural convection conditions by Wu et al. ~7 The general nature of the extinction process is similar for various convective and diffusive conditions when examined in terms of the Damk6hler number. The typical variation of burning or evaporation (gasification) rate of a drop with Damk6hler number is sketched in Fig. 46. Fixed kinetic constants are considered, treating E/RT~ as a parameter. Low Damk6hler numbers correspond to small diffusion or convection times, i.e. fast transport rates. This condition is encountered with small drops under spherically symmetric or natural convection conditions and with small drops and high ambient velocities for forced convection conditions. Low values of E/RT~ are encountered when the ambient temperature is near the adiabatic flame temperature, e.g. for equivalence ratios near unity in Figs. 44 and 45. The sketch in Fig. 46 indicates that the drop gasification rate is single-valued for such conditions; therefore, reaction effects gradually increase with increasing Damk6hler number and no well-defined

59

increasing

_ ~ " ~ " ~

"

e0

I'-.....
ign

I
Da E Damkohler Number

I
Da I

FIG. 46. Variation of drop gasification rate with Damk6hler number.

extinction condition is observed. At low values of Damk6hler number, the rate of reaction influences the gasification process which is reflected by a significant variation of the gasification rate with varying Damk6hler number. At high Damk6hler numbers, however, the gasification rate approaches a limiting value where high reaction rates yield a thin BurkeSchuman flame. In this region, rates of diffusion control the process and the actual rate of reaction has little influence on the rate of gasification. As E/RTo~ increases, the curves in Fig. 46 tend toward a characteristic S shape. The lowest branch of the curve involves only small levels of reaction and is characteristic of an evaporating drop. The uppermost branch represents a combusting drop, approaching the Burke-Schuman limit at high Damk6hler numbers. The middle branch is unstable and drop gasification along this line is not observed. Extinction is identified by the minimum Damk6hler number on the upper curve, Da~. Quasisteady ignition is identified by the maximum Damk6hler number on the lower curve, Dal, although this would only be true in the event that kinetic parameters appropriate when a flame is present are also representative of preignition kinetics. Wu et alJ 7 find expressions for the extinction Damk6hler number for spherically symmetric and forced and natural convective conditions, making the following assumptions: global bimolecular reaction with preexponential factor K, at the limit of large values of E/RT~ ; only concentration diffusion with negligible radiation, dissociation in the flame, and viscous dissipation; unity Lewis number; and an ideal gas mixture with Cp, Krp, p2D (forced and natural convection) or pD (spherically symmetric) constant. The analysis yields the following expressions for the range of Damk6hler numbers where a stable envelope flame is possible:
DaF = Krpdp/uo~ > DaeF, forced convection

(5.18)
Das = K,pd2 /D s >=DaEs, spherically symmetric

(5.19)

60

G.M. FAETH DaN = Krp(dp/a) 1/2 > DaeN, natural convection. (5.20) were used, with the drop simulated by a porous sphere continuously fed with fuel. The test method involved igniting the sphere and then gradually increasing the velocity until the envelope flame either extinguished or was transformed into a wake flame. Experimental conditions and the nature of the results for these investigations are summarized in Table 10. These experiments were conducted either in air or in oxygenenriched air. The agreement between the trends of extinction velocity obtained from asymptotic analysis and the measurements summarized in Table 10 can be examined by recalling that Eq. (5.21) yields the following expression for extinction velocity u~vE ~ dpp n- 1 (5.25)

The extinction Damk6hler numbers, Da~i, appearing on the RHS of these equations are functions of E/R, Yo~, v'y/V'o, Cp, Q, hso, To~ and T~. Wu et al. 17 should be consulted for complete formulas for the DaEi. In general, the DaE~ tend to decrease with increasing flame temperature and are relatively weak functions of pressure (except near the thermodynamic critical point) and drop surface temperature. The extinction criteria appearing in Eqs. (5.18)(5.20) all yield a minimum drop size for a stable envelope flame. The DaEi are nearly constant for a fixed ambient temperature and oxygen concentration, therefore, Eqs. (5.18)-(5.20) yield the following proportionalities between extinction diameter and the pressure, the velocity and the gravitational field: dpEF ~ u ~ p - 1 forced convection dpE~ ~ p - 1 to (5.21)

p - 3/2, spherically symmetric (5.22) (5.23)

dp~s ~ ap- a, natural convection

where the range of the pressure exponent for the spherically symmetric case results from the selection of either pDr or p2Df as constants? 7 Equations (5.18)-(5.23) are appropriate for a global secondorder reaction. For an nth order reaction, the pressure exponents in Eqs. (5.21)-(5.23) should be increased by the value (2-n). iv Equations (5.21)-(5.23) indicate that the extinction diameter should decrease as the pressure increases for all three transport configurations. The effect of pressure is greatest for natural convection conditions. For forced convection conditions, the extinction diameter is directly proportional to velocity, at least for Reynolds numbers sufficiently high for the boundary layer approximations to apply. Wu et al. 17 also find an interesting relationship between extinction diameters for natural and forced convection conditions. For the lowest order of approximation, at the same ambient conditions and pressure, dpEF - - 9 ( U ~ ] ( F F I 4 dpt~N \ adp~v / \ F N / " (5.24)

The extinction factors, Fv and Fu, have nearly the same value for the combustion of typical hydrocarbons when Y0~ > 0.1. The other term on the RHS of Eq. (5.24) can be recognized as the Froude number for the process, based on the extinction diameter for forced convection conditions. At large Froude numbers, dpEv >> dpEn, which implies that natural convection has only a minor influence on extinction. The criterion of Eq. (5.24) is useful for the design and evaluation of extinction experiments--at least for cases where the ambient velocity is in upflow. There have been several experimental studies of envelope flame extinction under forced convection conditions. 77,x67,xTx-174 Quasisteady configurations

for a global nth-order reaction. 17 Only the earliest measurements of Spalding 77'171 correspond to the diameter trend indicated by Eq. (5.25). Wu et al. 17 argue that the lower power of diameter observed by Agoston et al. 172 could be due to natural convection effects for their downflow configuration, however, the Froude number of these measurements is reasonably high. The upflow experiments of Sami and Ogasawara 167 yield even poorer agreement with Eq. (5.25) even though they involve relatively high Froude numbers where natural convective effects should be small. The average pressure dependence observed by Sami and Ogasawara 167 and Gollahalli and Brzustowski 174 suggests global reaction orders in the range 1.3-1.5, which is low in comparison to other hydrocarbon flame processes. Furthermore, Sami and Ogasawara 167 observed an unusual peak in the extinction velocity of methanol and n-butanol near 3atm, which is difficult to explain using current theories (similar peaks were sought but not observed for hydrocarbons in Ref. 174). Quantitative evaluation of predicted trends for extinction velocity with changes in ambient temperature and concentration has not been reported. Wu et al. 17 also attempted comparison of the spherically symmetric and natural convection extinction diameter predictions with measurements reported by Tarifa et al. 162 While comparison of trends was encouraging, unfortunately the data was obtained in an intermediate Grashoff number range where neither limiting case was appropriate. None of the data obtained in the investigations summarized in Table 10 is directly applicable for evaluation of drop ignition during spray combustion processes. Since the measurements do not agree with trends expected from existing theory, theoretical extrapolation of the data is questionable. It appears that additional study is necessary to resolve conditions for envelope flame stability of drops within combusting sprays. It would be particularly desirable if extinction experiments were conducted using ambient compositions representative of spray flames, i.e. a range of mixture ratios rather than heated or oxygen enriched air flows. The extinction of wake flames has received rela-

Evaporation and combustion of sprays TABLE10. Summary of envelope flame extinction experiments for forced convection conditions Reference" (date) Fuels Spalding ~7'17~ (1953) kerosene gasoline benzine ethyl alcohol air (~30C
1

61

Agostonetal.

172

(1957) n-butyl alcohol

Sami and Ogasawara 16v (1970) methanol n-butanol kerosene air 20-600C 0.4~16 0.8-6.5 2.1-10 upflow 30-5000 17-1700
dip '3 pi/3

Gollahalli and Brzustowski lv3'~v4 (1973,1975) n-pentane n-heptane

Environment: Composition Temperature (K) Pressure (atm) Velocity (m/s) Sphere diameter (mm) Flow orientation Reynolds number b Froude number c Dependence of extinction velocity on: Diameter Pressure Ambient temperature Ambient composition Other factors

oxygen enriched air Y% = 0.23 0.33 29-38C


1

air room temperature 1-24 0.5 2 6 upflow 140-18000 4 70

0.7-2.3 7 26 crossflow 310 3800 7 22


de

0.9-2.3 3.15 13.2 downflow 180-2100 25 200


-v1/2 d
n.a.

n,a,

n.a. increase with increasing T~ n.a.


--

pi/2 n.a. n.a.

n.a. rapid increase with increasing Y% Increased turbulence intensity reduces u~E' turbulence scale also has an effect

increase with increasing To~ n.a.

a All quasisteady experiments using a porous sphere continuously fed with fuel.
b R e = d.uooE/v ~ . e r r = uSE/ade.~/

tively little attention, even though their presence influences the mechanism of fuel oxidation in spray flames. Spalding 77'171 presents some wake flame extinction data for volatile fuels evaporating in pure air. 5.4.4. D r o p i g n i t i o n General reviews of drop and spray ignition are presented in Refs. 6-15. The present discussion is limited to aspects of spontaneous drop ignition pertinent to modeling drop processes in dilute cornbusting sprays. The main objective of drop ignition studies has

~-~

Ignition Locus

..

been to determine the time required for a stable envelope flame to appear (the ignition delay) when a drop is subjected to a given environment. Theoretical models of drop ignition are generally quite similar to the extinction models discussed in the previous section, i61-i65'i75'i76 Major assumptions of these models are: one-step global chemical reaction; quasisteady gas phase, i.e. the gas phase structure adapts instantly to imposed boundary conditions at the drop surface and the surroundings; and spherical symmetry with no forced or natural convection effects. Initial studies employed either numerical integration of the governing equations or approximated the properties of the reaction zone in order to obtain a closed form solution, i61-i64 More recent studies by Law and coworkers, 165j75'176 employ asymptotic analysis at the limit of large activation energy, adapting the extensive numerical results of Lifian.166 When the quasisteady assumption is made, the gas phase analysis can be separated from the analysis of drop heat-up. The gas phase analysis yields the drop gasification rate as a function of Damk6hler number at fixed ambient conditions, for various fuel vapor mass fractions at the drop surface.* A typical gas phase solution is sketched in Fig. 47. The lower * The drop surface temperature is frequently employed as a parameter, ~65'175'176 however, the temperature is directly related to vapor mass fraction at a given pressure through the vapor pressure relationship.

Damkohler Number

FIG. 47. Variation of drop gasification rate with Damk6hler number for the pre-ignition region.

62

G.M. FAETH diffusional and convection transport of the crucial intermediate in the model--effects which become more important when forced convection effects are significant. Clearly, the understanding of drop ignition processes is not very highly developed at present. Perhaps this is not surprising in view of current difficulties in modeling envelope flame stability. Further theoretical and experimental studies are needed to resolve drop ignition processes, particularly with regard to forced convection conditions in environments representative of drops in combusting sprays. Until this is done, no reliable assessment can be made concerning the presence of envelope flames around drops in combusting sprays. 5.5. Drop Transport in Sprays The conventional analysis of drop transport properties, described in Section 3, is only appropriate for steady nonturbulent flow past a sphere; therefore, fundamental uncertainties appear when the analysis is applied to the nonsteady turbulent processes encountered by drops within a spray. In this section we examine the validity of the conventional analysis in more detail, beginning with effects of drop motion in a nonturbulent fluid and then considering effects of turbulence. 5.5.1. Effects of drop motion There have been numerous studies of the effects of the motion of spheres on their drag and transport properties. Soo, 17 Gauvin and coworkers, 182'183 and Putnam et al.XS4present extensive reviews concerning the drag and transport properties of moving spheres. Aspects of sphere dynamics are also more recently reviewed by Durst et al. 185 with reference to the response of seeding particles used for Laser-Doppler anemometry, and by Hinze, 186 with respect to turbulent dispersion of particles. We begin by considering the motion of a rigid spherical particle in a nonturbulent fluid. The effect of lift forces, which appear when there are significant velocity gradients normal to the trajectory of the particle, are also ignored for the present. In these circumstances, the general equation of particle motion (the B-B-O equation which includes effects studied by Bassett, Boussinesq and Oseen) can be written as follows :17
/z d 3

portion of the S-shaped curve is shown since only this region is relevant for pre-ignition conditions. Similar to Fig. 46, the locus of ignition is determined by conditions where the tangent to the gasification rate curve is vertical. Employing results similar to Fig. 47, Law and coworkers 165'175'176 determine the ignition delay by computing the drop heat-up process and finding the time required to reach the locus of ignition conditions. The trajectory of the heat-up process can also be located on the plot of gasification rate as a function of Damk6hler number--two examples are illustrated in Fig. 47, assuming that the initial drop temperature is quite low, yielding a small initial fuel vapor concentration at the drop surface. As the drop heats-up, the surface concentration of fuel increases while the Damk6hler number decreases since the drop diameter becomes smaller due to evaporation. For a sufficiently high Damk6hler number, e.g. curve a-b in Fig. 47, the drop reaches the ignition locus before heat-up is complete and a stable envelope flame is formed. However, for low Damk6hler numbers, e.g. curve c-d in Fig. 47, the drop reaches its wet bulb state having a fuel vapor concentration below the ignition locus and continues to evaporate throughout its lifetime without igniting. Law 175 has completed computations using the quasisteady ignition model and compared the results with a portion of the drop ignition data reported in Refs. 177 179. These test conditions involved drops in heated air or oxygen enriched air, with relatively small drop Reynolds numbers. The data from the various sources were scattered; however, kinetic parameters could be found which provided a fair correlation of the measurements. This finding is encouraging; however, it will be necessary to extend the model so that forced convection conditions can be treated before the approach can be applied to estimate drop ignition properties in practical sprays. Other troublesome aspects of this approach involve the quasi-steady assumption and the use of a global one-step reaction. Since ignition properties must often be found for low temperature environments, reaction rates can be slow leading to failure of the quasisteady approximation. Chain-branching effects are frequently important for ignition processes, in which case a one-step reaction mechanism is not adequate. Faeth and Olson 178 develop a different model of drop ignition which allows for non-quasisteady reactions, by postulating that ignition occurs when the concentration of a crucial chain-branching intermediate reaches a critical value in the gas phase. This approach provided a correlation of their drop ignition data in stagnant heated air environments. Sangiovani and Kestin 18'181 extended the crucial intermediate model to treat forced convection effects using film theory. The predictions were tested using data on drop ignition in the post-flame region of a fiatflame burner. The extended model was able to correlate a portion of the data; however, trends with respect to velocity variations were not predicted correctly. This behavior was attributed to neglect of

pp~~

dup

--

- ~dppCDlu-upl(u-up)-~dp Or + ~ d . p C l ~ ( u - up) + ~ dp(~pp)


7I" 3

7~

7C 3 ~P

1/2

CB J,po ~

[", (d/d~)(u-u.) d" +

Fe

(5.26)

where the time derivative is taken following the motion of the particle d ~ ~3 (5.27)

dtp -- c3t ~-Up~r"

Evaporation and combustion of sprays The term on the left-hand side of Eq. (5.26) represents the inertial force of the sphere. Taken in order, the terms on the right-hand side of Eq. (5.26) represent: the drag force on the sphere, which conventionally includes both skin friction and form drag; the force on the sphere due to static pressure gradients in the flow; the force on the sphere due to the inertia of fluid displaced by its motion, which is often called the virtual mass term; the Bassett term, which allows for effects of the deviation of the flow from a steady flow pattern around the sphere; and the external or body force term, e.g. the force due to gravity. Equation (5.26) was originally developed for low Reynolds number flow around an accelerating particle where C o = 2 4 / R e and C I = CB= 1.17 Therefore, accelerating particles at low Reynolds numbers have a drag coefficient which is independent of the rate of acceleration while the coefficient of the virtual mass term is identical to the value found for inviscid flow around a sphere, j7 At higher Reynolds numbers, most investigators have adopted the practice that the form of the equation remains the same; however, the coefficients are assumed to be functions of Reynolds number and acceleration number, e.g. 187 190

63

also neglected for the present. In order to simplify examination of the order of magnitude of the terms in Eq. (5.26), we introduce the dimensionless variables

fi = U/Upo, zp = 18#tp/ppd~

(5.30)

yielding the following equation of particle motion.


dl~p C D (ff_ffp)+l p 2

+ ( 9p

~1/2f,,

d a ( f f - a p ) de (5.31)

c-ci(-?

2 lUrell lUroll dtp

i=I,B.
(5.28)

Drop-life-history computations indicate that drop Reynolds numbers are generally less than 104 in sprays. Typical values of acceleration numbers for drops can be estimated if we consider a drop injected into a motionless gas and neglect all terms in Eq. (5.26) except the drag term, yielding dp dlurel] = 0.75Co(p/pp). Ur21 dtp (5.29)

For values of the Reynolds number greater than unity, Eq. (5.29) suggests that the acceleration number is less than 10, with the largest values encountered at low Reynolds numbers and high pressures. For this range of Reynolds and acceleration numbers, it has been found that Co can be represented by the standard drag curve for steady flow around spheres; while Cl and Cn are relatively independent of the acceleration number and can be approximated by their values at low Reynolds numbers, i.e. CI : CB = 1.187 190 These determinations have been made for both rectilinear and oscillatory sphere motions, with spatially uniform ambient pressures. With the equation of motion established, it is now of interest to examine conditions where the usual simplified form, e.g. Eq. (3.18), is valid. For this consideration, the pressure gradient term appearing in Eq. (5.26) will be neglected since mean static pressure gradients are generally small for spray processes. Turbulent fluctuations can also contribute to the pressure gradient term, 17 but this will be neglected under the assumptions stated at the beginning of this section. Body force terms are easily evaluated and are

where Cos = 24/Re, is the Stokes drag, while Co is the standard drag expression from Eqs. (3.13) or (3.15). Since Co~Cos/> 1, Eq. (5.31) indicates that the virtual mass and Bassett terms only provide a small contribution to the drag at atmospheric pressure where P/Pn is on the order of 10-3 for most sprays. At higher pressures the coefficients of the virtual mass and Bassett terms approach unity suggesting a greater need to include them. A mitigating factor, however, is that Reynolds numbers in sprays tend to increase with pressure as well, causing the coefficient of the drag term to become larger throughout much of the lifetime of a drop in a spray. Therefore, recomputation of drop motion for the high pressure conditions of Figs. 28 and 29, including the virtual mass and Bassett terms, did not greatly change the results for these conditions. Each case, however, should be evaluated independently. Calculations of particle motion are frequently encountered where only the virtual mass term is considered, while the Bassett term is neglected. Equation (5.31) reveals that these two terms are of similar order of magnitude and must generally be considered together. Computations of Hjelmfelt and Mockros T M and A1-Taweel and Carley 192 yield the same conclusion. Another phenomenon that can influence the motion of a particle is the Magnus effect which appears when the particle is rotating. Drop rotation in sprays can occur by virtue of velocity gradients within injectors since the fluid vorticity still takes some time to decay after the drops are formed. Drops also tend to acquire the local vorticity of the gas phase causing them to rotate when they are passing through a gas flow with a velocity gradient normal to their trajectory. For the Reynolds number range encountered in sprays, these effects cause the drop to be deflected toward the side where the peripheral motion of the particle is in the same direction as relative gas flow. Saffman, 193 investigated the lift force on a sphere at low Reynolds numbers in a shear flow. Assuming that the sphere is rotating freely, so that its angular velocity is one-half the local velocity gradient, the ratio of the lift and drag forces is IFDI -- 12n (5.32)

64

G.M. FAETH effect could still be comparable to processes such as turbulent drop dispersion and further examination of this phenomenon would be desirable. Effects of virtual mass, Bassett, Magnus, Saffman lift and static pressure gradient forces influence drop heat and mass transfer rates as well as drag. However, since the present discussion has indicated that these phenomena only moderately influence drop motion in typical spray environments, this aspect will not be considered here. 5.5.2. EJfects of turbulence A turbulent environment modifies drop transport rates from the values estimated using single drop correlations which were generally obtained with flows having small ambient turbulence levels. The effect has invariably been ignored in past discrete drop models of sprays, e.g. the models summarized in Table 9. It has occasionally been suggested that effects of turbulence on transport rates can be ignored when the drop diameter is smaller than the turbulent microscales. 197'198 This assessment is not correct, since scale has been found to have only a secondary effect on the modification of transport rates by turbulence, asz'a83'199-28 Even if this view were correct, however, it will be shown that significant portions of practical sprays have turbulent microscales that are comparable to drop diameters, justifying consideration of effects of turbulence on drop transport. In order to relate the spray environment to experimental conditions used in studies of effects of turbulence on particle transport it is useful to have a general indication of the turbulence intensities and scales encountered by drops in sprays. In typical sprays, turbulence intensities of the gas phase are on the order of 10-40 Vo, aside from regions of recirculation or near the edge of sprays in a stagnant environment, c.f. Refs. 14,24-28 and 198 for examples. This is not the relevant measure of the turbulence intensity experienced by a drop moving through a spray, however, since the relative velocity between the drop and the gas can be much smaller than the mean velocity of the gas. Therefore, the turbulence intensity determined by an observer on a drop is generally higher than the value for the gas flow and values in excess of 100~o are not unusual, even in regions of maximum gas velocities within the flow field. Turbulence macro- and microscales experienced by drops can vary widely depending upon the geometry of the system; however, some generalizations can be obtained in the near-injector region, where the flow has properties similar to a jet. In the following, we limit consideration to a round injector with no swirl and estimate scales employing the L H F approximation. The results discussed in Section 2 indicated that the turbulence properties of evaporating and combusting sprays are similar to constant density jets near the L H F limit. Since macroscales are primarily dependent on the geometry of the flow, the longitudinal integral

where K = 81.2 and r is the coordinate normal to the trajectory. This ratio can be examined for spray conditions using the L H F results to approximate the velocity gradient within a spray, e.g. Figs. 9 and 12. This yields the maximum velocity gradient as follows: #u

~ = K'duo/x 2, x/d > (x/d)pc

(5.33)

where K' is a constant on the order of 100. Combining Eqs. (5.32) and (5.33) yields

F~ = K,,Re~/2(dp/d)l/2(d/x) ' x/d > (x/d)p c


(5.34) where K" is a constant on the order of ten. Since d/dp is normally 10-100 for the largest drops, Eq. (5.33) indicates that the lift force is at least an order of magnitude smaller than the drag force, beyond the potential core of the spray, x/d > 10, for the low Reynolds number regime where this analysis is valid. Saffman 193 reaches the same general conclusion for freely rotating particles in shear flows. An alternative estimate of lift and drag forces at low Reynolds numbers was obtained by Robinow and Keller. 17'194 Use of their expressions for a freely rotating particle also suggests that lift forces are small in comparison to drag forces for low Reynolds number flows with freely rotating particles. Lift forces for freely rotating particles have been studied by Eichhorn and Small 195 for Reynolds numbers in the range 85-230. For present purposes, their lift coefficient measurements can be roughly correlated as follows:

~?rJ "

(5.35)

Noting that drag coefficients for spheres are in the range 0.44-1, for this range of Reynolds numbers, combining Eqs. (5.33) and (5.35) yields

F~ = K,,(dp/d)Z(d/x)4 ' x/d > (x/d)p c (5.36)


where K " is a constant on the order of 1000. Equation (5.36) also yields the conclusion that lift forces are at least an order of magnitude smaller than drag forces in sprays, when dp/d < 1 and x/d > 10. The use of the lift coefficient results of Maccoll, 196 for rotating spheres at high Reynolds number, but neglecting the region of negative drag, yields the same conclusion. The preceding results suggest that effects of Saffman lift forces on freely rotating drops are generally small, throughout the dilute spray region, x/d > 10. The effect of residual rotation of the particles due to processes within the injector, or freely rotating particles in the near injector region where velocity gradients are greater than the estimate of Eq. (5.33), however, could still result in significant Magnus effects; therefore, the phenomenon deserves further attention in spray models. Even in the dilute spray region, where Magnus forces are small, their

Evaporation and combustion of sprays scale can be estimated from the constant density jet measurements of Wygnanski and Fiedler,65 yielding A f j x = 0.04 in the self-preserving region of the flow, with somewhat higher values off-axis. Therefore, in this region d~- >~ 0.04 . (5.37)

65

critical Reynolds number for transition and the relative turbulent intensity
Recr(U'/Urel) 2
=

45

(5.39)

For well atomized sprays, d/dp is smallest for constant area injectors with no swirl, falling in the range 10 100 for the largest drops in a spray. Under these conditions, Ay/d v >~ 4 40 in the dilute spray region, x/d > 10, although smaller values could be encountered near the injector. This suggests that the drops are a relatively small feature in comparison to the scale of the larger turbulent eddies tLroughout much of the dilute spray region, similar to the view taken in the turbulent drop dispersion calculations discussed earlier. Turbulent microscales are dependent on the Reynolds number of the flow. In the following, the Kolmogoroff microscale, ~/4 = (v3/e), will be used to represent microscales. 21 For present purposes we assume that production and dissipation of turbulent kinetic energy are equal and employ the k-~ model for an axisymmetric jet to yield the following estimate

La2c&/xJ

\dv/\d/"
(5.38)

The Reynolds number of a spray Re:, = ucx/v does not vary to a great degree with position in the self-preserving region and is usually greater than 104 for practical sprays. Then if k/a~ and the velocity gradient in Eq. (5.38) are estimated from LHF results for the self-preserving region of evaporating and combusting sprays,26-2s,57 58 we find q/dp~lO-3(d/dp)(x/d). Employing did v in the range 10 100 which represents the larger drops in a spray as before, we find that ~/ d v for 10 < x/d < 100, which corresponds to the bulk of the dilute spray region. Higher Reynolds numbers would cause this range to extend to higher values of x/d. Therefore, we conclude that turbulent microscales are comparable or smaller than the larger drops in many practical sprays, particularly in the near-injector region. The effect of turbulence on drag properties has been studied and reviewed by Gauvin and coworkers, 182'199'2 and Soo. 17 One of the most remarkable effects that is encountered is the reduction of the Reynolds number for transition to the supercritical flow regime (defined by convention as the Reynolds number where the steeply sloped portion of the standard drag curve reaches the value 0.3). 182'199'200 At low turbulence intensities, this transition occurs at Recr ~ 105, which is well above the range usually of interest for drops in sprays; however, at high turbulence intensities, transition has been observed for Recr as low as 400.182 Torobin and Gauvin 199 find the following relationship between the
JPECS 9-I/2-E

for nonevaporating spheres having relative turbulence intensities greater than 10 ~. The effect of turbulence scale on transition is weak in this range (macroscale to diameter ratio in the range 2-6). The effect of evaporation from the surface has not been found to influence Recr to a great degree, although the evaporation rates examined are low in comparison to drop vaporization rates in sprays. At high turbulence levels, drag coefficients are higher than the standard drag curve for Re < Re,. r. After the abrupt drop at transition to values below the standard drag curve, the drag coefficient increases for a time, before resuming its general downward trend with increasing Reynolds number. Working formulas for drag coefficient in turbulent flows for both the subcritical and supercritical regimes are summarized by Clift and Gauvin. 183 The unusually low values of drag coefficient, in comparison to the standard drag coefficient, reported by Ingebo 22 and others in highly turbulent flows for limited Reynolds number ranges, have been attributed to this early transition phenomenon.199 Effects of ambient turbulence intensities on heat and mass transfer rates of spheres are discussed in Refs. 203-208, for Reynolds numbers of interest to spray processes. The experiments upon which these results are based were conducted in flows where only velocity fluctuations were significant; therefore, the effects of simultaneous velocity, composition and temperature fluctuations, which must be considered in a spray environment, remain largely unresolved. It is generally agreed that the relative turbulence intensity is the most important factor in representing the effect of turbulence on heat and mass transport rates, while the scale of the turbulence assumes a secondary role, similar to the behavior observed for drag. 23-28 An interesting effect of scale, however, is that when the ratio of the Eulerian scale to diameter is 1.6 for spheres, the influence of turbulence on transport rates is maximized.27 The magnitude of the effect and the optimum value of scale ratio are virtually identical to that observed by van der Hegge Zijnen29 for cylinders in crossflow, where he suggests that the optimum value corresponds to a resonance between free-stream fluctuations and eddy shedding. Raithby and Eckert 2s also observe some effect of scale on transport rates for Reynolds numbers on the order of 104; however, due to a relatively coarse selection of scale ratios an optimum was not observed. Galloway and Sage zv provide correlations for the effect of turbulence on heat and mass transfer to spheres, covering the subcritical regime with Reynolds numbers in the range 1.2-3 x 105, turbulence intensities in the range 0.5 23 ~o with Pr and Sc > 0.6. This data base involves only low rates of mass transfer, although Boulos and Pei z5 find similar behavior for high rates of blowing at the surface of a sphere. High

66

G.M. FAETH adopted as the most appropriate approximation for the response of the liquid phase to rapid variations in the environment of a drop; the presence of envelope flames is neglected; and the standard drag curve and convection correction expressions are employed. Under these assumptions, the instantaneous drag and mass transfer rate of a drop are functions of the instantaneous velocity and mixture fraction of the flow, given the diameter, velocity, initial temperature and initial composition of the drop. The mean transport rate can be obtained by integration over the joint pdf of velocity and mixture fraction. Applying the assumption that u and f are uncorrelated, the mean rate of evaporation becomes

turbulence intensities influence transport rates to the greatest degree for large Reynolds number, e.g. a 23 ~o turbulence intensity increases the Nusselt number by less than 1 ~o at Re = 1, while the increase is 23 ~o at Re = 1000. z6 A limitation of these results is the relatively low turbulence intensity levels of the data in comparison to the range of interest for drops in sprays, particularly when it is noted that extrapolation of the correlations to high intensities is very uncertain. For example, available correlations suggest that Nusselt numbers become proportional to (u'/ Urel)" at large turbulence intensities; however, existing measurements find n in the range 0.035-2, 2o4 zos which results in wide differences in extrapolated values. Drops in sprays experience temperature and concentration fluctuations as well as velocity fluctuations. Unfortunately, there is little information available in the literature concerning simultaneous variations of all these ambient properties on transport rates, particularly in environments representative of spray combustion processes. In the following the potential magnitude of the effect will be considered, using a simplified description of both turbulence and drop transport properties. Consider the effect of turbulent fluctuations on drop transport within a combusting spray. As a reasonable approximation, we employ the results of L H F computations, using the k-c.-9 turbulence model, to specify mean and turbulent properties within the spray. Mixture fraction and velocity fluctuations are assumed to be statistically independent; therefore, turbulence statistics are specified by ti a n d f a n d their variances 2k/3 and 9, assuming isotropic turbulence for the velocity field. Similar to the particle dispersion calculations described earlier, velocity fluctuations are described by a Gaussian distribution. Consistent with the LHF (k-e-9) model described earlier for sprays, a clipped Gaussian pdf will be employed for concentration fluctuations. Properties of the surroundings of the drop are prescribed by the state relationship. To fix ideas we consider an n-pentane spray burning in room temperature air at atmospheric pressure with the state relationship illustrated in Fig. 3. For present purposes, drop transport properties will be computed assuming a quasisteady gas phase. This assumption is more questionable for evaluating effects of turbulence than was the case for computing drop-life-histories, since characteristic times for turbulent fluctuations are small--particularly at high spray Reynolds numbers--in comparison to characteristic times for changes in mean flow properties. Furthermore, turbulence frequencies perceived by an observer on the drops will be further increased if there is a large relative velocity between the drop and the gas. Since transient development increases transport rates, it is expected that the present analysis will tend to underestimate effects of turbulent fluctuations. The remaining assumptions concerning drop transport are similar to those described in Section 3.2.1. In particular, the thin skin model in Table 6 (case 1) is

(rh) =

df ff~(dp,uv, u,f)P(f)P(u)
(5.40)

which becomes a function of k and 9 through the influence of these parameters on P ( f ) and P(u). A similar expression can be written for Fo. In contrast to Eq. (5.40), current spray models employ mean values of all properties (velocity, temperature and composition) to compute transport rates. In this case, k and 9 influence transport rates only indirectly, due to their effect on mean values alone. In the following, an average transport rate computed in the conventional manner will be designated by an overbar, e.g. n~. A comparison between these two methods of averaging to obtain drop transport rates is illustrated in Fig. 48, for drops within a combusting n-pentane spray at atmospheric pressure. Average properties used in these computations were computed following Mao et al. 27 The ratios Qh)/rfi and (Fo)/FD are plotted as a function of mean mixture fraction for R e = 0 . 1 and 1000, with gl/2/f and U'/are 1 a s parameters, both having values of 0.1 and 1.0. These parameter ranges are reasonably representative of conditions in sprays. The effect of f on the ratios is small, which is surprising in view of the rapid variation of scalar properties near the stoichiometric mixture fraction, c.f. Fig. 3. The largest discrepancies between the two methods of averaging occur at high Reynolds numbers and relative turbulence intensities. The effect of fluctuations is generally much greater for drag than for the evaporation rate, e.g. (rh)/n~ reaches values of 1.2-1.4 for Re = 1000 and U'/Urel = 1, while (Fo)/Fo is in the range 3-4. The present analysis is relatively crude, however, these high ratios for drag are comparable to values summarized by Clift and Gauvin as3 at high turbulence intensities for similar values of Reynolds number. It would be premature to consider the results illustrated in Fig. 48 to be anything more than a qualitative indication of significant potential effects of turbulent fluctuations. Data are needed to verify the predictions and effects of transient development of the flow, the liquid phase model, turbulence scale, etc., need to be evaluated. The effect of envelope flames, and their stability, should also be considered when the

Evaporation and combustion of sprays


0.4

67

g"~/~
0.1 0.1

u'/U
0.1 I 0.1 I
-I 2

0.2

3
I-E

I I

0
s

R e = 0.1 Re=O.I -O2 0.01 I 0:1 ~


I O.Or 0,1

0,4

-........ 0.2

,....,--

7
I.E
0
Re = 1000

7
A

Re

I000

-0"20.01

I 0.1
"T

0
0.01

I
O. I

FIG.48. Effect of turbulence on evaporation rates and drag of drops in an n-pentane spray burning in room temperature air at atmospheric pressure. drop interacts with eddies containing oxygen. The results suggest that current practice tends to underestimate drop transport rates in sprays--particularly drag forces. The effects appear to be sufficiently large to justify further study. 5.6. assessment are that other processes important to dense sprays, drop collisions, effect of drops on turbulence properties, etc. are not necessarily small at the same condition. Furthermore, small Reynolds numbers cause drops to influence their neighbors to a greater degree than is the case for the conditions considered in Ref. 17. These effects will be considered in greater detail in the next section, which examines dense spray phenomena. Stochastic analysis provides a more satisfactory approach for treating drop dispersion by turbulence than earlier attempts to correlate turbulent diffusivities of particles. Recent use of the stochastic method to predict particle dispersion in jets has yielded encouraging agreement with measurements. Gosman and Ioannides la have employed this theory to examine effects of turbulent dispersion in evaporating and combusting sprays, showing that the effect was small for cases they considered. If this is true for a broader class of sprays it will be very helpful, since stochastic dispersion models require extensive computations. However, more study will be required to complete this assessment. It has been shown that drop transport rates are significantly affected by the presence of envelope flames at low fuel equivalence ratios. Current informa-

Summary

Several aspects of the dilute region of evaporating and combusting sprays were examined, including: the extent of the dilute spray region, drop dispersion by turbulence, effects of envelope flames and their ignition and stability, and effects of drop motion and turbulence on drop transport properties in sprays. These phenomena all have a bearing on assumptions commonly adopted for models of spray processes. The extent of the dilute spray region was estimated assuming that the region corresponded to the zone where inter-drop spacing was greater than two drop diameters. This criterion was chosen since existing measurements indicate that transport rates in the presence of convection for this spacing approach the values found for single drops. 17 This assumption in conjunction with a simplified model of the nearinjector region suggested that the dilute spray approximation is reasonable for much of the twophase region of sprays, x/d > 10. Limitations on this

68

G.M. FAETH The dense spray region is near the injector for wellatomized sprays although its boundaries are not welldefined. In the downstream direction, the dense spray gradually evolves into a dilute spray. This limit was discussed in Section 5, based only on effects of drop spacing on interphase transport rates; however, factors such as drop collisions or modifications of turbulence by particles may persist to a greater degree, extending the dense spray region in a manner that is not well understood at this time. Upstream of the dense spray region, the liquid phase is less dispersed, yielding flows more typical of low void fraction, twophase flows in tubes, e.g. Reitz 21 observed what he termed a churn-flow region very near the injector for well-atomized pressure atomized sprays. Flow regimes of this type might be handled using conventional two-phase flow analysis, 16 but their consideration is beyond the scope of the present review. Progress in understanding processes within the dense spray region has been limited, largely due to experimental difficulties. The region is difficult to probe, since its small size for practical injectors limits spatial resolution. Two-phase flow effects at intermediate void fractions also create difficulties in probing the flow or even penetrating it optically in order to make observations. In spite of these problems, there has been progress in developing models of dense spray processes which will be discussed in the following. Overall models of dense sprays will be considered first of all, followed by a description of studies of specific dense spray processes, e.g. interphase transport rates, drop collisions, and the effect of drops on turbulence properties.
6.1. Dense Spray Models

tion on the ignition and stability of envelope flames is not adequate to provide predictions of the presence or absence of envelope flames in practical spray combustion processes. More experiments are particularly needed to evaluate existing models of envelope flame stability for conditions representative of sprays. Current drop ignition models are not very highly developed, particularly for convection conditions, and improved analysis and experiments are also needed in this area. Drop-motion models conventionally consider particle inertia and quasisteady drag, ignoring virtual mass, pressure gradients, Bassett, Saffman lift and Magnus forces. The present discussion has indicated that this approximation is reasonable at moderate pressures. Magnus forces are a possible exception in the near-injector region, if the injection process results in high levels of initial particle rotation. Saffman lift forces, due to mean velocity gradients, appear to be an order of magnitude smaller than drag forces for freely rotating particles and are generally a secondary effect. Virtual mass and Bassett forces, however, become important for low Reynolds number particles in high pressure sprays (P/Pt, = 0(1)) and both should be considered under these conditions. Effects of turbulence on drop transport properties appear to be quite significant although this phenomenon has generally been ignored in spray models appearing to date. Turbulence intensities perceived by particles in sprays can be quite high, since relative velocities are generally lower than mean gas velocities. Existing information indicates that effects of turbulence scale are secondary, so that small drop sizes do not provide a basis for ignoring effects of turbulence. Furthermore, turbulence microscales are comparable to or smaller than the diameters of the largest drops in sprays over much of the dilute spray region. Preliminary considerations indicate that turbulence has a most pronounced effect on droplet drag and a lesser influence on evaporation rates, particularly for high Reynolds numbers and relative turbulence intensities. Turbulence generally increases drag and evaporation rates, suggesting that current separated flow spray models tend to underestimate interphase transport rates. More experimental results are needed to establish these trends, particularly at the high relative turbulence intensities encountered by drops in sprays.
6. DENSE SPRAY PROCESSES

The dense spray region is characterized by discrete drops in a more-or-less continuous gas phase. Unlike the dilute spray region, however, void fractions are low and interactions between drops are important within the dense spray region. This implies that effects of drop spacing on transport rates, drop collisions, modifications of turbulence properties due to the presence of drops and the volume occupied by the liquid phase become significant within the dense spray region.

Several of the models summarized earlier, c.f. Table 9, consider some dense spray phenomena. For example, Anderson et aU treat drop collisions while Butler et al. 4 and Gosman and Johns 17 allow for the volume occupied by the liquid in the governing equations for the gas phase. A more comprehensive analysis of dense spray phenomena has been reported by O'Rourke and Bracco. 211 This model was evaluated by comparison with experiments of Hiroyasu and Kodota lz3 which involved measurements of spray tip penetration and drop sizes during transient axisymmetric injection into a pressurized chamber. Drop vaporization was negligible for the experimental conditions; therefore, evaluation of the model was limited to the dynamics of an isothermal spray within a constant pressure chamber. However, the formulation inclttdes effects of drop heat-up and vaporization, as well as gas phase conservation of energy, for a fully interacting twophase flow. O'Rourke and Bracco 2H use a discrete droplet model (DDM) with a Lagrangian treatment of the liquid phase. Turbulence was represented by a constant eddy diffusivity, based on a correlation for constant density jets, along with a turbulent Prandtl/Schmidt number of unity. The effect of

Evaporation and combustion of sprays turbulent fluctuations on the drag and heat and mass transfer properties of drops was considered using stochastic model based on a constant turbulence intensity; however, turbulent drop dispersion was neglected. Dense spray phenomena that were considered included: the effect of nonunity void fraction on interphase transport rates; the effect of drop collisions; and allowance for nonunity void fraction in the governing equations for the gas phase. Computations were initiated at a void fraction of 0.9, which was estimated to correspond to x/d = 10 for the injector used in the experiments. This void fraction was suggested as a limit for the dilute spray region in Section 5.2, based on relatively small effects of void fraction on interphase transport rates for higher values of void fraction. In agreement with this assessment, O'Rourke and Bracco 2x 1 found that their computations were not strongly influenced by effects of void fraction on interphase transport rates. In contrast, O'Rourke and Bracco 21a found significant effects of drop collisions in the region considered in their computations. For example, choosing an initial SMD of 6 #m, which is consistent with an aerodynamic breakup correlation due to Reitz 21 for the injector conditions, the computed downstream SMD increased to 48/~m, which compares reasonably well with the value 42 #m measured by Hiroyasu and Kodota. 123 While predicted drop growth was smaller when initial drop sizes were assumed to be larger, e.g. an initial SMD of 42 pm yielded a downstream SMD of 70#m, effects of collisions appear to be very significant in this flow. It should be noted that transient sprays in a stagnant environment tend to maximize collision effects since drops injected later in the injection period move through gases having a higher velocity and tend to overtake and collide with drops injected earlier in the process. Nevertheless, the results provide ample motivation for developing a reliable model of drop collision processes in sprays. O'Rourke and Bracco 21x were able to achieve reasonably good predictions of the spray tip penetration measurements of Hiroyasu and Kodota, ~23 while at the same time rationalizing the small initial SMD from the correlation of Reitz 21 with the larger SMD measured downstream as a result of drop collisions. The authors point out, however, that many features of their model require more complete assessment. 6.2. Drop Transport in Dense Sprays There is little direct information available on drop transport processes in dense sprays. One method of treating dense spray transport, suggested by Putnam et al. 184 and used by O'Rourke and Bracco, 2~ is to adapt results measured in fluidized and packed beds. A second procedure, involves studying transport processes in specified drop arrays. Both these approaches will be considered in the following. The literature on transport in packed and fluidized beds is extensive and it is beyond the scope of the present review to discuss these results in any detail.

69

The method of adapting the results to dense sprays will be considered, however, following the approach suggested by O'Rourke and Bracco. 211 In this case the particle drag coefficient was synthesized to yield the fluidized bed results of Richardson and Zaki, 212 at low void fractions, while approximating Bachelor's z 13 theoretical perturbation analysis at high void fractions to yield 21 Co = 24(o;o265 + Re2/3o~1Ts/6)Re. (6.1)

Equation (6.1) is identical to Eq. (3.13) at c~ = 1 and is 0 in fair agreement with several other expressions developed for dispersed flows at moderate void fractions. 2a4'215 Since the effect of % was not large in the computations of Ref. 211, however, the accuracy of Eq. (6.1) for dense sprays has not been adequately assessed. Analogous expressions were developed for effects of nonunity void fraction on convective heat and mass transfer rates, however, these portions of the model were not evaluated. 211 There have been a number of theoretical studies of flow and transport properties of arrays of spheres. Happel and Brenner 216 summarize early work in this field. Tal and Sirignano 217 discuss more recent computations and also present new results covering Reynolds numbers of interest for sprays. In order to control computer costs, array calculations are generally simplified by considering a single sphere surrounded by a fluid cell with boundary conditions selected to approximate the array. Results are influenced by the geometry chosen for the cell, which introduces uncertainties, although the computations have provided fair agreement with measurements.216,2 a7 If successfully exploited, this method has the potential for examining effects of circulation, evaporation and combustion in dense sprays that would be difficult to investigate by experiment. There have been numerous theoretical studies of interactions between drops in arrays in the absence of forced convection. Labowsky 21s and Labowsky and Rosner 2~9 review much of this work. Recent calculations of this type consider arrays of evaporating and combusting drops allowing for Stefan flows induced by evaporation, although the results are limited to quasisteady gas-phase conditions. When forced convection is absent, drops interact even for very large spacings. For example, the lifetime of combusting drops in an equally spaced array of four drops was found to increase by 20 ~o over the lifetime of a single drop, for interdrop spacings of 10 drop diameters. 2a8 In contrast, spacings of this magnitude would result in negligible effects of adjacent drops on transport rates when forced convection is present, even for drop Reynolds numbers as low as unity. 17'216'217 Since most practical sprays have appreciable drop Reynolds numbers and involve turbulent flow, consideration of multidrop effects along the lines of Refs. 218 and 219 appears to have limited utility in comparison to the spray models described in other sections of this review.

70 6.3. D r o p Collisions

G.M. FAETH particles is reduced since less momentum is exchanged between the phases during a turbulent fluctuation. Abramovich221 and Owen 22: propose models of these processes, where turbulent diffusivities are adjusted for the presence of particles while considering effects of slip. Melville and Bray 223 describe a model of particulate flow, with small interphase slip, employing constant eddy diffusivities for momentum and particle transport. The influence of particles on turbulent transport was considered by adopting expressions due to Abramovich221 and Owen. 222 The predictions were evaluated using the measurements of Laats and Fishman 225'226 for a round jet containing powders of various sizes. Best results were obtained for these conditions using the approach suggested by Owen 222 for negligible slip between the phases. Danon et al. 224 describe a k-l model for two-phase jets. The length scale was not modified from the value appropriate for a constant density jet; however, a term representing the added dissipation due to the presence of particles was included in the governing equation for k, in a manner which partially compensated for particle inertia. The model was evaluated using the data of Hetsroni and Sokolov 66 for a round jet containing oil droplets. The basic model was not in good agreement with these measurements. The comparison between predictions and measurements was improved by multiplying the rates of production and dissipation of k by a coefficient which was a very strong function of void fraction. The authors conclude that there is a substantial and unexplained influence of particles on turbulence properties of jets, even at low particle concentrations. The conclusions of Danon et al. 224 and the empirical corrections of their basic model are probably incorrect, due to errors in the original experimental results of Hetsroni and Sokolov. 66 During the early stages of developing the LHF model described in Section 2,227 an attempt was made to evaluate the model using the data for air jets containing oil drops from Ref. 66. Similar to Danon et al. 224 it was found that effects of the drops were unexplainably high in view of their small momentum and estimated slip. However, further study indicated that Sokolov and Hetsroni 66 significantly underestimated the effect of drop impacts on the hot wire used to make velocity measurements (even though a clipper circuit was employed to remove the highest peaks in the hot wire signal due to drop impacts). The effect biased their velocities upward whenever drops were present in the flow; therefore, gas velocities were overestimated-particularly at the centerline where drop concentrations were highest tending to make velocity profiles appear narrower as particle concentrations increased, when plotted as ~/ffc. Based on these observations it is recommended that the experiments be repeated using a procedure less influenced by the presence of drops for velocity measurements, prior to application of the results for evaluation of effects of particles on turbulence,z27

The results of O'Rourke and Bracco 2xi suggest that drop collisions can have an important effect on spray processes--particularly in the dense spray region. Modeling collisions involves numerous uncertainties, however, and only a few studies have considered the process for sprays. Putnam et al. Is4 review early work in the field. Anderson et a l J develop a model for collisions between drops, as well as between drops and surfaces, for steady conditions. O'Rourke and Bracco 2~ include effects of drop collisions in their model of transient spray development. The methods used in these studies are only briefly described in the following--original sources should be consulted for details. Both Anderson et al. 1 and O'Rourke and Bracco 211 employ DDM. Therefore, collisions occur between drop groups as they pass through the flow field. Both occurrence and outcome of drop collisions must be modeled. In both studies, drop groups are assumed to collide when they are in the same computational cell (drops within a group do not collide since they are moving with the same velocity). Rates of collision are estimated using results developed in cloud physics. Upon collision, several outcomes are possible: coalescence, elastic rebound, or shattering into two or more drops of different size. Criteria for coalescence during collisions with both drops and surfaces are developed by Anderson et aL 1 based on an approach suggested by Swinbank.~90'Rourke and Bracco 2a1 adopt results from Brazier-Smith et al. 22 concerning coalescence of free drops. The results of collision can lead to formation of many new drop groups, which rapidly exhausts computer storage capabilities.2~1 Therefore, proliferation of classes is reduced by modifying the properties of colliding groups, while satisfying conservation principles. The collision models of Refs. 1 and 211 represent interesting attempts to treat this important process in a comprehensive spray model. It is clear, however, that these models involve numerous approximations which must still be justified. Additional research on drop collisions in sprays is obviously needed, which should be facilitated by recent advances in laser-based methods of sizing drops and measuring their velocities. 6.4. Effects o f D r o p s on T u r b u l e n c e In the near-injector region, the turbulent gas-phase flow is largely initiated by the transfer of momentum from the liquid phase. While the particles create the motion, however, their presence also influences turbulence production, dissipation and scale. Several theoretical investigations of the turbulence properties of two-phase jets have been reported, for the limit where mean gas and particle velocities are nearly identical.2za-22'~ When there is negligible slip between the phases, the particles reduce turbulence fluctuations since the gas flow must accelerate the additional mass of the particles. If there is appreciable slip between the phases, however, the effect of the

Evaporation and combustion of sprays Buckingham and Siekhaus 151 describe a k-e turbulence model which allows for effects of particles on turbulence properties. The model is applied to flows containing small solid particles, considering added dissipation of particle interactions in the governing equations for k and e. This analysis also includes a stochastic particle dispersion model as discussed in Section 5.3. The computations are not compared directly with experiments, however, the results suggest that particles tend to damp turbulent motions primarily due to inertial effects, which reduces turbulent exchange coefficients. Complimentary experimental results for a particle laden flow also indicate that the presence of particles tends to reduce eddy scale sizes, in qualitative agreement with the predictions. T M More extensive evaluation will be required, however, in order to adequately establish empirical coefficients, appearing in terms representing particle/turbulence interactions, which are employed in the model. 6.5. S u m m a r y The dense spray region involves low void fractions, effects of nearby drops on drop transport rates, appreciable rates of drop collisions and significant effects of drops on the turbulence properties of the flow. Spray models developed by Anderson et al., ~ Butler et al. 4 and Gosman and Johns t7 treat some aspects of the dilute spray region; however, the model reported by O'Rourke and Bracco 211 represents one of the first comprehensive treatments of dense sprays. The model of Buckingham and Siekhaus T M was developed for dense solid particle flows but could be modified to treat dense spray processes. The predictions obtained by O'Rourke and Bracco 211 indicate that effects of adjacent drops on drop transport rates are relatively small for void fractions greater than 0.9; however, drop collision effects are still significant at relatively high void fractions. The predictions of Buckingham and Siekhaus T M indicate that particles tend to reduce turbulence levels and turbulent exchange coefficients primarily due to inertial effects, although additional production of turbulence in particle wakes was ignored in these computations. The results of these computations provide incentive for further study of dense spray problems, but are not definitive since many aspects of the models have not yet been evaluated by comparison with measurements. There is a pressing need to overcome the experimental difficulties of the dense spray region so that further progress can be made toward developing reliable models for this region. 7. CONCLUSIONS 1. The L H F approximations for sprays implies infinitely fast interphase transport rates so that the results are only strictly applicable to a spray containing infinitely small drops. The advantages of the method are that necessary injector specifications are

71

minimal; the computations are no more difficult than computations for single-phase processes, allowing adoption of existing computer algorithms; and the results are generally qualitatively correct, at least providing a useful bound for the gains that could be realized by improving atomization properties for practical injectors. The LHF model has been evaluated in both evaporating and combusting sprays. These results indicate that finite interphase transport rates are important for most practical sprays and that L H F models generally overestimate the rate of development of the flow, even for spray SMD as small as 10#m. This effect is due to the small dimensions of injector passages which create high rates of flow deceleration near the injector. Droplife-history computations, using the L H F predictions to estimate conditions within the spray, provide a means of estimating the potential accuracy of L H F predictions for a given application. 2. Several separated flow spray models, which allow for finite interphase transport rates, have been reported. Most recent models have adopted the discrete droplet formulation (DDM), which involves a Eulerian description of the gas phase (including source terms to represent transport from drops), while drop properties are determined by Lagrangian computations for the life-histories of groups of drops using a statistically significant sample. These models are largely limited to the dilute spray region where the void fraction is assumed to be unity, effects of adjacent drops on drop transport rates and drop collisions are ignored, and drops are assumed to have a negligible influence on turbulence properties. Evaluation of separated flow models for dilute sprays has been limited. The computations exhibit a strong sensitivity to the drop size, velocity and direction distributions specified at the injector exit. Existing measurements in sprays have not documented these properties sufficiently to obtain a convincing evaluation of predictions. Computations reported thus far have also probably not achieved adequate numerical closure, particularly in the near injector region where dense grids are needed due to small injector passage dimensions. The capabilities of a given model to adequately estimate drop-lifehistories should also be calibrated if reliable results are to be obtained, since property variations result in considerable uncertainties in these predictions. In spite of these uncertainties and limitations, several recent instances of the application of spray models to assist practical development efforts have been reported.1 5 3. Several approximations concerning dilute spray processes were considered in this review, including: drop breakup, turbulent drop dispersion, effects of envelope flames, and effects of drop dynamics and turbulence on drop transport rates. Effects of drop breakup become important near the injector, particularly at high pressures where low surface tension reduces the stability of the drop. More information should be developed on conditions for

72

G.M. FAETH fractions greater than 0.9. Additional evaluation of this approach is needed considering effects of drop gasification, circulation and possible combustion. The effects of polydisperse drops on correlations of transport rates at low void fractions also need to be examined. A recent dense spray model developed by O'Rourke and Bracco 211 indicates that drop collisions are very important near the injector. Current models of drop collision processes are not highly developed and have not received experimental confirmation, which should motivate more study of this problem. Past studies indicate that drops affect turbulence production, dissipation and scale. Results obtained thus far suggest that drops tend to damp fluctuations, reducing turbulent exchange coefficients; however, existing theories involve numerous simplifications which still must be evaluated. The author's research on spray combustion has been supported by NASA Grants NGR 39-009077, NSG-3306 and NAG 3-190, under the technical management of R. J. Priem, C. J. Marek and R. Tacina, respectively, of the Lewis Research Center. The author also acknowledges the assistance of L-D. Chen, J-S. Shuen, A. S. P. Solomon and G. A. Szekely,Jr., in preparing this review.
Acknowledgement

drop breakup in spray environments as well as the outcome and rate of breakup. Stochastic analysis has provided a more satisfactory treatment of turbulent drop dispersion than earlier attempts to correlate turbulent diffusivities of particles. Recent analysis of effects of turbulent dispersion on the overall properties of evaporating and combusting sprays suggest that dispersion effects are small in comparison to other uncertainties in spray modeling, at least for the cases considered. This assessment should be extended to a broader range of spray conditions. Envelope flames were shown to have an important influence on drop transport rates when the fuel equivalence ratio of the gas around the drop was small. Existing information on ignition and stability of envelope flames is incomplete and not consistent; therefore, additional experimental and theoretical effort is required before the importance of envelope flames on the properties of combusting sprays can be determined. Conventional models of drop transport generally ignore dynamic effects which cause virtual mass and Bassett forces. This approach is adequate for most practical sprays near atmospheric pressure; however, both virtual mass and Bassett effects become more important at high pressures. Saffman lift forces, which appear for freely rotating particles in a velocity gradient, are generally a second-order effect for most sprays. Residual rotation of drops, due to vorticity generated within the injector passage, yields Magnus forces which deserve further attention in the nearinjector region. Drops experience higher turbulence intensities than the flow as a whole, since their relative velocities can be small in comparison to mean gas velocities. The larger drops in a spray have diameters comparable to turbulence microscales over much of the dilute spray region. Furthermore, studies have shown that scale has only a secondary effect on the enhancement of particle transport rates by turbulence. It appears that the effect of velocity fluctuations on particle drag is very significant for sprays. Unfortunately, existing data do not extend to the turbulence intensity levels encountered by drops in sprays suggesting the need for further measurements. Velocity fluctuations have a smaller influence on heat and mass transfer rates. Aside from possible effects of envelope flames, which were not evaluated, concentration fluctuations appear to be reasonably represented using mean properties, which is current practice. 4. Important dense spray phenomena include effects of adjacent drops on drop transport rates, effects of drop collisions, and effects of drops on turbulence properties. In this region, the fact that the void fraction is not unity must also be considered in the governing equations for the gas phase. Effects of neighboring drops on drop transport rates, for the convective environment of sprays, were examined using data from fluidized and packed beds. The results indicate that these effects are small for void

REFERENCES

1. ANDERSON,0. L., CHIAPPETTA, L. M., EDWARDS, D. E. and McVEY, J. B. Analytical Modeling of Operating

2.
3.

4.

5.

6. 7.

8. 9. 10. 11. 12. 13. 14.


15. 16.

Characteristics of Premixing-Prevaporizing Fuel-Air Mixing Passages, Report No. UTRC 80-102, Vol. I and II, United Technologies Research Center, East Hartford (1980). BRUCE,T. W., MONGIA,H. C. and REYNOLDS,R. S. Combustor Design Criteria Validation, Vol. I-IIl, USARTL-TR-78-55(A,B,C),Fort Eustis, VA (1979). MONGIA,H. C. and SMITH,K. An Empirical/Analytical Design Methodology for Gas Turbine Combustors, AIAA Paper No. 78 998 (1978). BUTLER,T. D., CLOUTMAN, L. D., DUKOWICZ,J. K., RAMSHAW, J. D. and KRIEGER, R. B. Combustion Modeling in Reciprocating Engines (J. N. MATTAVI and C. A. AMANN,eds.), pp. 231-264, Plenum Press, New York (1980). ALPERT, R. L. and MATHEWS, M. K. Calculation of Large-Scale Flow Fields Induced by Droplet Sprays, Technical Report No. FMRC J.1.OEOJ4.BU, Factory Mutual Research Corp., Norwood, MA (1979). FAETH,G. M. Prog. Energy Combust. Sci. 3, 191 (1977). HARRJE, D. T. and REARDON, F. H. (ed.), Liquid Propellant Rocket Combustion Instability, NASA SP194, pp. 37 102,Washington (1972). MELLOR,A. M. Seventeenth Symposium (International) on Combustion, pp. 377-387, The Combustion Institute, Pittsburgh (1979). WILLIAMS, Combustion of Sprays of Liquid Fuels, Elek A. Science, London (1976). WILLIAMS, Prog. Energy Combust. Sci. 2, 167 (1976). A. WILLIAMS, Combust. Flame 21, 1 (1973). A. LAW,C. K. Prog. Energy Combust. Sci. (in press). KRIER,H. and Foo, C. L. Oxidation Combust. Rev. 6, 111 (1973). CHIGIER, A. Pro 9. Energy Combust. Sci. 2, 97 (1977). N. HEDLEY,A. B., NURUZZAMIN, S. M. and MARTIN,G. A. F. J. Inst. Fuel 44, 38 (1971). COLLIER, J. G. Convective Boiling and Condensation, McGraw-Hill, London (1973).

Evaporation and combustion of sprays 17. SOD, S. L. Fluid Dynamics of Multiphase Systems, Blaisdell, Waltham, MA (1967). 18. WALLIS~ G. One-Dimensional Two-Phase Flow, McGraw-Hill, New York (1969). 19. ONUMA,Y. and OGASAWARA,M. Fifteenth Symposium (International) on Combustion, pp. 453-465, The Combustion Institute, Pittsburgh (1975). 20. KOMIYAMA, K., FLAGAN, R. C. and HEYWOOD, J. B. Sixteenth Symposium (International) on Combustion, pp. 549-560, The Combustion Institute, Pittsburgh (1977). 21. THRING, M. W. and NEWBY, M. P. Fourth Symposium (International) on Combustion, pp. 789-796, Williams and Wilkins, Baltimore (1953). 22. NEWMAN,J. A. and BRZUSTOWSKI,T. A. AIAA J 9, 1595 (1971). 23. SHEARER,A. J. and FAETH,G. M. Combustion of Liquid Sprays at High Pressures, NASA CR-135210 (1977). 24. KHALIL, E. E. and WHITELAW, J. H. Sixteenth Symposium (International) on Combustion, pp. 569-576, The Combustion Institute, Pittsburgh (1977). 25. KHALIL, E. E. A Simplified Approach for the Calculation of Free and Confined Spray Flames, AIAA Paper No. 78-029 (1978). 26. SHEARER, A. J., TAMURA, H. and FAETH, G. M. J. of Energy 3, 271 (1979). 27. MAO, C-P., SZEKELY,G, A., JR. and FAETH, G. M. J. of Energy 4, 78 (1980). 28. MAD, C-P., WAKAMATSU,Y. and FAETH, G. M. Eighteenth Symposium (International) on Combustion, pp. 337-347, The Combustion Institute, Pittsburgh (1981). 29. SHAHED,S. M., FLYNN, P. F. and LYN, W. T. Combustion Modeling in Reciprocating Engines (J. N. MATTAVIand C. A. AMANN,eds.), pp. 345-368, Plenum Press, New York (1980). 30. HIROYASU, H. and KODOTA, T. SAE Trans. 85, 513 (1976). 31. SHAHED,S. M., Cmu, W. S. and YUMLU, V. S. SAE Trans. 82, 338 (1973). 32. CHIU, W. S., SHAHED, S. M. and LYN, W. T. SAE Trans. 85, 502 (1976). 33. KHAN, I. M., GREEVES, G. and PROSERT, D. M. Proc. I. Mech. Engr. C142, 295 (1971). 34. REYNOLDS, W. C. Combustion Modeling in Reciprocating Engines (J. N. MATTAVI and C. A. AMANN, eds.), pp. 41-68, Plenum Press, New York (1980). 35. BILGER,R. W. Prog. Energy Combust. Sci. 1, 87 (1976). 36. GOSMAN,A. D., LOCKWOOD, F. C. and SALOOJA, A. P. Seventeenth Symposium (International) on Combustion, pp. 747-760, The Combustion Institute, Pittsburgh (1979). 37. GOSMAN, A. D., LOCKWOOD, F. C. and SYED, S. A. Sixteenth Symposium (International) on Combustion, pp. 1543-1555, The Combustion Institute, Pittsburgh (1977). 38. LAUNDER,B. A. and SPALDING, D. B. Mathematical Models of Turbulence, Academic Press, London (1972). 39. HUTCHINSON, P., KHALIL, E. E., WHITELAW, J. H. and WIGLEY,G. J. Heat Trans. 98, 276 (1976). 40. ELGHOBASHI,S. E. and PUN, W. M. Fifteenth Symposium (International) on Combustion, pp. 1353-1365, The Combustion Institute, Pittsburgh (1975). 4l. HUTCHINSON,P., KHALIL, E. E. and WH1TELAW,J. N. J. Energy 1,212 (1977). 42. LOCKWOOD, F. C. and NAGUIB, A. S. Combust. Flame 24, 109 (1975). 43. KOOSINLIN, M. L. and LOCKWOOD, F. C. AIAA J 12, 547 (1974). 44. SPALDING, B. Chem. Engr. Sci. 26, 95 (1971). O. 45. SPALDING, B. Combust. Sci. Tech. 13, 3 (1976). D. 46. BILGER, R. W. Molecular Transport Effects in Turbulent Diffusion Flames at Moderate Reynolds Number, AIAA Paper No. 81-0104 (1981).

73

47. DAVIES,C. N. Aerosol Science, pp. 393 468, Academic Press, New York (1976). 48. KENT, J. W. and BILGER, R. W. Sixteenth Symposium (International) on Combustion, pp. 1643 1656, The Combustion Institute, Pittsburgh (1977). 49. KENNEDY,I. M. and KENT, J. H. Seventeenth Symposium (International) on Combustion, pp. 279-287, The Combustion Institute, Pittsburgh (1979). 50. MDREAU, P. Eighteenth Symposium (International) on Combustion, pp. 993-1000, The Combustion Institute, Pittsburgh (1981). 51. SPALDING,D. B. Thirteenth Symposium (International) on Combustion, pp. 649 657, The Combustion Institute, Pittsburgh (1971 ). 52. MASON,H. B. and SPALDING,D. B. Combustion Institute European Symposium, pp. 601-606, Academic Press, New York (1973). 53. MAGNUSSEN, B. F. and HJERTAGER, n. W. Sixteenth Symposium (International) on Combustion, pp. 719-729, The Combustion Institute, Pittsburgh (1977). 54. BORGIaI,R. Adv. Geophys. 18B, 349 (1974). 55. BRAY,K. N. C. and Moss, J. B. Acta Astronautica 4, 291 (1977). 56. LOCKWOOD,F. C. Combust. Flame 29, 111 (1977). 57. SHEARER, A. J. and FAETR, G. M. Evaluation of a Locally Homogeneous Model of Spray Evaporation, NASA CR-3198 (1979). 58. MAO, C-P., SZEKELY, G. A. JR. and FAETH, G. M. Evaluation of a Locally Homogeneous Flow Model of Spray Combustion, NASA CR-3202 (1980). 59. GORDON,S. and MCBRIDE, B. J. Computer Program for Calculation of Complex Chemical Equilibrium Compositions, Rocket Performance, Incident and Reflected Shocks, and Chapman-Jouguet Detonations, NASA SP-273, Washington (1971). 60. COWPERTHWAITE, M. and ZWISLER, W. H. TIGER Computer Program Documentation, Stanford Research Institute, NTIS Report No. AD-A-002791, Washington, DC (1974). 61. CANADA, S. and FAETH, G. M. Fifteenth Symposium G. (International) on Combustion, pp. 1345-1354, The Combustion Institute, Pittsburgh (1975). 62. PRAUSNITZ, J. M. a n d CHUEH, P. t.: Computer Calculations for High Pressure Vapor-Liquid Equilibria, Prentice-Hall, Englewood Cliffs, NJ (1968). 63. SPALDING,D. B. GENMIX: A General Computer Programfor Two-Dimensional Series, Parabolic Phenomena, Pergamon Press, Oxford (1978). 64. GOSMAN, A. D. and PUN, W. M. Lecture notes for a course entitled Calculation of Recirculating Flows, Mechanical Engineering Department, Imperial College (1973). 65. WYGNANSKI,I. and FIEDLER. H. E. J. Fluid Mech. 38, 577 (1969). 66. HETSRONI,G. and SOKOLOV, M. J. Appl. Mech. 38, 314 (1971). 67. BECKER,H. A., HO'I'TEL, W. C. and WILLIAMS,G. C. J. Fluid Mech. 30, 285 (1967). 68. CORRSIN,S. and UBEROI, M. S. Further Experiments on Flow and Heat Transfer in a Heated Turbulent Air Jet, NACA Report No. 988, Washington (1950). 69. TROSS,S. R. Characteristics of a Submerged Two-Phase Free Jet, M.S. Thesis, The Pennsylvania State University (1974). 70. BILGER, R. W. and KENT, J. H. Measurements in Turbulent Jet Diffusion Flames, Charles Kolling Research Laboratory Technical Note F-41, The University of Sydney (1972). 71. YUEN, M. C. and CHEN,L. W. Combust. Sci. Tech. 14, 147 (1976). 72. WILLIAMS, A. J. Chem. Phys. 33, 133 (1960). F. 73. BELLAN,J. and SUMMEREIELD,M. Combust. Flame 33, 107 (1978). 74. BEREAD,A. and HIBBARD,R. Effect of Radiant Energy

74

G.M. FAETH
Flows, pp. 12.27-12.33, University Park, PA (April 1977). EL BANHAWY, Y. and WHITELAW, J. H. AIAA J. 18, 1503 (1980). GOSMAN,A. D. and IOANNIDES,E. Aspects of Computer Simulation of Liquid-Fueled Combustors, AIAA Paper No. 81-0323 (1981). GOSMAN,A. D., IOANNIDES,E., LEVER,D. A. and CL1FFE, K. A. A Comparison of Continuum and Discrete Droplet Finite-Difference Models Used in the Calculation of Spray Combustion in Swirling Turbulent Flows, AERE Harwell Report TP865 (1980). SWITHENBANK,J., TURAN, A. and FELTON, P. G. Gas Turbine Combustor Design Problems (A. H. LEFEBVRE, ed.), pp. 249-314, Hemisphere Publishing, Washington (1980). BOYSON,F., AYERS, W. H., SWITHENBANK,J. and PAN, Z. Three-Dimensional Model of Spray Combustion in Gas Turbine Combustors, AIAA Paper No. 81-0324 (1981). GOSMAN,A. D. and JOHNS, R. J. R. Computer Analysis of Fuel-Air Mixing in Direct-Injection Engines, SAE Paper No. 800091 (1980). ANDERSON,O. L. Comput. Fluids 8, 391 (1980). SWINBANK, C. Nature 4051, 849 (1947). W. HESKESTAD,G., KUNG, H. C. and TODTENKOPF, N. F. Air Entrainment in Water Spray Curtains, ASME Paper No. 76-WA/FE-40 (1976). PUN, W. M. and SPALDING,D. B. A General Computer Program for Two-Dimensional Elliptic Flows, Report HTS 176/2, Mechanical Engineering Department, Imperial College, London (1977). GOSMAN,A. D. and LOCKWOOD, F. C. Fourteenth Symposium (International) on Combustion, pp. 661-671, The Combustion Institute, Pittsburgh (1973). FOUNTI, M., HUTCHINSON, P. and WHITELAW, J. H. Measurements and Calculations of a Kerosene Fueled Flow in a Model Furnace, Report FS/79/19, Mechanical Engineering Department, Imperial College, London (1979). HINZE, J. O. Turbulence, pp. 427 428, McGraw-Hill, New York (1975). SNYDER,W. H. and LUMLEY,J. L. J. Fluid Mech. 48, 41 47 (1971). TISHKOFE,J. M., HAMMOND,D. C. Jr. and CHRAPLYVY, A. R. Diagnostic Measurements of Fuel Spray Dispersion, ASME Paper No. 80-WA/HT-35 (1980). PATANKAR,S. V. and SPALDING,D. B. Fourteenth Symposium (International) on Combustion, pp. 605-614, The Combustion Institute, Pittsburgh (1973). PATANKAR, S. V. Numerical Prediction of ThreeDimensional Flows, Studies in Convection: Theory Measurements and Applications (B. E. LAUNDER, ed.), Vol. 1, Academic Press, New York (1975). CHAM Ltd., London (1980). HIRT, C. W., AMSDEN, A. A. and COOK, J. L., J. Comp. Phys. 4, 227-253 (1974), AMSDEN,A. A. and HIRT, C. W. YAQUI: An Arbitrary Lagrangian-Eulerian Computer Program for Fluid Flow at all Speeds, Los Alamos Scientific Laboratory Report LA-5100 (1973). NORTON, J. L. and RUPPEL, H. M. YAQUI User's Manual for Fireball Calculations, Los Alamos Scientific Laboratory Report LA-6261-M (1976). HIROYASU, H. and KODOTA, T. Fuel Droplet Size Distribution in Diesel Combustion Chamber, SAE Trans. 83, 2615-2624, Paper No. 740715 (1974). GOSMAN,A. D. and JOHNS, R. J. R. Development of a Prediction Tool for In-Cylinder Gas Motion in Engines, SAE Paper No. 780315 (1978). GOSMAN,A. D., JOHNS,R. J. R. and WATK1NS,A. D. Combustion Modelin 9 in Reciprocatin9 Enyines (J. N. MATTAVI and C. A. AMANN, ed.), pp. 69-129, Plenum Press, New York (1980).

on Vaporization and Combustion of Liquid Fuels, NACA RME-52109 (1952). 75. KODOTA, T., HIROYASU, H. and OVA, H. Japan Soe. Mech. Engrs. 41, 2475 (1975). 76. GOLLAHALLI, S. R. and BRZUSTOWSKI, T. A. Fourteenth Symposium (International) on Combustion, pp. 1333-1344, The Combustion Institute, Pittsburgh (1973). 77. SPALDING,D. B. Fourth Symposium (International) on Combustion, pp. 847 864, The Combustion Institute, Pittsburgh (1953). 78. ALDRED,J. W., PRATT,J. C. and WILLIAMS,A. Combust. Flame 17, 139 (1971). 79. MASDIN, E. G. and TrtRING, M. W. J. lnst, Fuel, 351 (1962). 80. PRAKASH, S. and SIRIGNANO, W. A. Int. J. Heat Mass Transfer 21,885 (1978); Ibid. 23, 253 (1980). 81. LAW, C. K., PRAKASH, S. and SIRIGNANO, W. A. Sixteenth Symposium (International) on Combustion, pp. 605-617, The Combustion Institute, Pittsburgh (1977). 82. WILLIAMS,F. A. Combustion Theory, Addison-Wesley, Reading, MA (1965). 83. BIRD, R. B., STEWART, W. E. and LIGHTEOOT,E. M. Transport Phenomena, John Wiley, New York (1960). 84. KODOTA,T. and HIROYASU,H. Trans. Japan Soc. Mech. Engrs. 42, 1216 (1976). 85. SZEKELY,G. A. JR. and FAETH,G. M. Combustion of Carbon Slurry Drops in a Turbulent Diffusion Flame, AIAA Paper No, 81-0322 (1981). 86. YUEN, M. C. and CHEN, L. W. Combust. Sci. Tech. 14, 147 (1976). 87. PUTNAM,A. ARS J. 31, 1467 (1961). 88. DICKERSON, R. A. and SCHUMAN, M. D. J. Spacecraft, 99 (1960). 89. FAETH,G. M. and LAZAR,R. S. AIAA J. 9, 2165 (1971). 90. SZEKELY,G. A. JR. and FAETr~, G. M. AIAA J. 20, 422 (1982); and ibid., Effects of Envelope Flames on Drop Gasification Rates in Turbulent Diffusion Flames, Combustion and Flame (in press). 91. REID,R. C., PRAUSNITZ,J. M. and SHERWOOD,T. K. The Properties of Gases and Liquids, 3rd Ed., McGraw-Hill, New York (1977). 92. FAETH, G. M., DOMINIOS, D. P., TULPINSKY, J. F. and OLSON, D. R. Twelfth Symposium (International) on Combustion, pp. 9-18, The Combustion Institute, Pittsburgh (1969). 93. LAZAR, R. S. and FAETrt, G. M. Thirteenth Symposium (International) on Combustion, pp. 801-811, The Combustion Institute, Pittsburgh (1971). 94. CANADA,G. S. and FAETH,G. M. Fourteenth Symposium (International) on Combustion, pp. 1345 1354, The Combustion Institute, Pittsburgh (1973). 95. BOYSON, F. and SWITrlENBANK, J. Seventeenth Symposium (International) on Combustion, pp. 443-453, The Combustion Institute, Pittsburgh (1979). 96. WOLVE, H. E. and ANDERSON, W. H. Kinetics, Mechanism and Resulting Droplet Sizes of the Aerodynamic Breakup of Liquid Drops, Aerojet General Rept. No. 0395-04(18)SP (1964). 97. RANGER,A. A. and N1CHOLLS, J. A. AIAA J. 7, 285 (1969). 98. CROWE,C. T., SHARMA,M. P. and STOCK,D. E. J. of Fluids Engr. 99, 325 (1977). 99. CROWE, C. T. A Computational Model for the GasDroplet Flow in the Vicinity of an Atomizer, Paper No. 74-25, Western States Section, The Combustion Institute (1974). 100. CROWE, C. T. A Numerical Model for the Gas Droplet Flow Field Near an Atomizer, First International Conference on Liquid Atomization and Spray Systems, Tokyo, August (1978). 101. JUREWICZ,J. T., STOCK, D. T. and CROWE, C. T. The Effect of Turbulent Diffusion on Gas Particle Flow in an Electric Field, First Symposium on Turbulent Shear

102. 103. 104.

105.

106.

107. 108. 109. 110. 111.

112. 113.

114. 115. 116. 117. 118.

119. 120. 121.

122. 123. 124.

125.

Evaporation and combustion of sprays 126. KHALIL,K. H., EL MAHALLAWY,F. M. and MONEIB,H. A. Sixteenth Symposium (International) on Combustion, pp. 135-143, The Combustion Institute, Pittsburgh (1977). 127. CARETTO,L. S. Prog. Energy Combuat. Sci. l, 47 71 (1976). 128. SIMMONS,H. C. Engr.for Power99, 309-319 (1977). 129. DUNDAS, P. H. The Scaling of Sprinkler Discharge: Prediction of Drop Size, Progress Report No. 10 in Optimization of Sprinkler Fire Protection, Technical Report RC 73-T-40, Factory Mutual Research Corporation, Norwood, MA (June 1974). 130. DUKOWICZ,J. K. J. Comp. Phys. 35, 229 (1980). 131. BRACCO,F. V. Combust. Sci. Tech. 8, 69 (1973). 132. BRACCO, F. V., GUPTA, H. L., KRISHNAMURTFIY, L., SANTAVICCA, D. A., STEINBERGER,R. L. and WARSAW, V. SAE Paper No. 760114 (1976). 133. WESTBROOK,C. K. Sixteenth Symposium (International) on Combustion, pp. 1517-1526, The Combustion Institute, Pittsburgh (1977). 134. HASELMAN,L. C. and WESTBROOK, C. K. SAE Paper No. 780138 (1978). 135. GANY, A., MANNHEIMER TIMNAT,T. and WOLFSHTEIN, M. Acta Astronautica 3, 241 (1976). 136. HARLOW,F. H. and AMSDEN,A. A. J. Comp. Phys. 17, 19 (1975). 137. HARLOW, F. H. and AMSDEN, A. A. J. Comp. Phys. 18, 440 (1975). 138. TRAVlS,J. R., HARLOW, F. H. and AMSDEN,A. A. Nucl. Sci. Engng. 61, 1 (1976). 139. DEHAYE, J. M., GIOT, M. and RIETHMULLER, M. L. Thermohydraulics of Two-Phase Systems for Industrial Design and Nuclear Engineering, pp. 95 179, Hemisphere Publishing Corp., Washington (1981 ). 140. JONES,A. R. Prog. Energy Combust. Sci. 3, 225 (1977). 141. HARSHA, P. T. and EDELMAN, R. B. Application of Modular Modeling to Ramjet Performance Prediction, AIAA Paper No. 78 944 (1978). 142. EDELMAN, R. B. and HARSHA, P. T. Modeling Techniques for the Analysis of Ramjet Combustion Processes, AIAA Paper No. 80-1190 (1980). 143. VARAPRASAD,C. M. and KAR, S. J. Engr.for Power 99, 225 (1977). 144. SHUEN,J-S., CHEN, L-D. and FAETH, G. M. Evaluation of a Stochastic Model of Particle Dispersion in a Turbulent Round Jet, AIChE J. (in press). 145. Yuu, S., YASUKOUCHI,N., H1ROSAWA,Y. and JOTAKI, T. AIChE J. 24, 509 (1978). 146. GOLDSCHMIDT, V., HOUSEHOLDER, M. K., AHMAD, G. and CHUANG, S. C. Progress in Heat and Mass Transfer, Vol. 6 (G. HETSRONI, ed.), pp. 487 508, Pergamon Press, Oxford (1972). 147. LILLY,G. P. Ind. Eng. Chem. Fundam. 12, 268 (1973). 148. JUREWICZ,J. T. and STOCK, D. E. A Numerical Model for Turbulent Diffusion in Gas-Particle Flows, ASME Paper No. 76-WA/FE-33 (1976). 149. HOTCHKISS, R. S. and HIRT, C. W. Particulate Transport in Highly Distorted Three-Dimensional Flow Fields, Los Alamos Scientific Laboratory Preprint LADC-72-364 (1972). 150. BUCKINGHAM,A. C. Turbulent Dusty Gas Motion with Weak Statistical Coupling, AIAA Paper No. 79 1484 (1979). 151. BUCKINGHAM, C. and SIEKHAUS, J. Interaction of A. W. Moderately Dense Particle Concentrations in Turbulent Flow, AIAA Paper No. 81 0346 (1981). 152. PESKIN, R. L. and KAU, C. J. J. Fluids Engr. 101, 319 (1979). 153. BROWN,D. J. and HUTCI-nNSON,P. J. Fluids Engr. 101, 265 (1979). 154. BRACCO, F. Fourteenth Symposium (International) on Combustion, pp. 831-842, The Combustion Institute, Pittsburgh (1973).

75

155. ALTENKIRCH,R. A., SHAHED, S. M. and SAWYER,R. F. Combust. Sci. Tech. 5, 147 (1972). 156. KESTIN,A. Combust. Sci. Tech. 6, 115 (1972). 157. CHIGIER,N. and MCCREATH, C. G. Acta Astronautica 1, 687 (1974). 158. DAMKOHLER,G. Z. Elektrochem 42, 846 (1936). 159. LORELL,J., WISE, H. and CARR, R. E. J. Chem. Phys. 25, 325 (1956). 160. AGAFANOVA,F. A., GUREVlCH, M. Q. and PALIEV, I. I. Soviet Phys.-Tech. Phys. 2, 169 (1958). 161. POLYMEROPOULIS, C. E. and PESKIN, R. L. Combust. Flame 13, 166 (1969). 162. TARIFA,C. S., DEL NOTARIO, P. P. and MORENO, F. G. Eighth Symposium (International) on Combustion, pp. 1035-1056, Williams and Wilkins, Baltimore (1962). 163. PESKIN,R. L. and WISE, H. AIAA J. 4, 1646 (1966). 164. PESKIN, R. L, POLYMEROPOULIS,C. E. and YEH, P. S. AIAA J. 5, 2173 (1967). 165. LAW, C. K. Combust. Flame 24, 89 (1975). 166. LIlqAN,A. Acta Astronautica l, 1007 (1974). 167. SAMI, H. and OGASWARA,M. JSME Bull. 13, 395 (1970). 168. FENDELL,F. E. J. Fluid Mech. 21,291 (1965). 169. KRlSHNAMUR~Y, L., WILLIAMS,F. A. and Seshadri, K. Combust. Flame 26, 363 (1976). 170. Wu, S., LAW, C. K. and FERNANDEZ-PELLO, A. C. A Unified Criterion for the Convective Extinction of Fuel Particles, to be published. 171. SPALDING,D. B. Fuel 32, 169 (1953). 172. AGOSTON, G. A., WISE, H. and ROSSER, W. A. Sixth Symposium (International) on Combustion, pp. 708-717, Reinhold, New York (1957). 173. GOLLAHALLI,S. R. and BRZUSTOWSKI,T. A. Fourteenth Symposium (International) on Combustion, pp. 13331344, The Combustion Institute, Pittsburgh (1973). 174. GOLLAHALLI, S. R. and BRZUSTOWSKI, T. A. Fifteenth Symposium (International) on Combustion, pp. 409-417, The Combustion Institute, Pittsburgh (1975). 175. LAW, C. K. Combust. Flame 31,285 (1978). 176. LAW, C. K. and CHUNG, S. H. Combust. Sci. Tech. 22, 17 (1980). 177. EL-WAKIL,M. M. and ABDOU,M. I. Fuel45, 177 (1963). 178. FAETH, G. M. and OLSON, D. R. SAE Trans. 77, 1793 (1968). 179. WOOD,B. J. and ROSSER,W. A. AIAA J. 7, 2288 (1969). 180. SANGtOVANNI, J. J. and KESTIN, A. S. Sixteenth Symposium (International) on Combustion, pp. 577-592, The Combustion Institute, Pittsburgh (1977). 181. SANGIOVANNL J. and Kestin, A. S. Combust. Sci. Tech. J. 16, 59 (1977). 182. TOROmN, L. B. and GAUVlN, W. H. Can. J. Chem. Eng. 37, 129 (1959); ibid. 37, 224 (1959); ibid. 38, 142 (1960); ibid. 38, 189 (1960). 183. CLIFT, R. and GAUVIN,W. H. Can. J. Chem. Eng. 49, 439 (1971). 184. PUTNAM,A. A. et al. Injection and Combustion of Liquid Fuels, WADC Technical Report 56-344, Chapt. 5 (1957). 185. DURST,F., MELLING,A. and WHITELAW,J. H. Principles and Practice of Laser-Doppler Anemometry, pp. 275294, Academic Press, London (1976). 186. HINZE, J. O. Turbulence, pp. 460-482, McGraw-Hill, New York (1975). 187. ODAR,F. and HAMILTON,W. E. J. Fluid Mech. 18, 302 (1964). 188. KARANEILIAN,S. K. and KOTAS, J. J. J. Fluid Mech. 87, 85 (1978). 189. ODAR, F. J. Fluid Mech. 25, 591 (1966). 190. HAMILTON, W. S. and LINDEL, J. A. Proc. A.S.C.E. 97, 805 (1971). 191. HJELMFELT, T. JR. and MOEKROS, L. F. Appl. Sci. Res. A. 16, 149 (1966). 192. AL-TAWEEL,A. M. and CARLEY,J. F. AIChE Symposium Series 67, No. 116, 114 (1971).

76

G.M. FAETH Experiments, Stratified Charge Automotive Engines Conference, The Institution of Mechanical Engineers, London (1980). RICHARDSON,J. F. and ZAKI, W. N. Trans. Inst. Chem. Engrs. 32, 35 (1954). BATCHELOR,G. K. J. Fluid Mech. 52, 245 (1972). ROWE,P. N. Trans. Inst. Chem. Engrs. 39, 175 (1961). ISHn, M. and ZUBER,N. AIChE. 3. 25, 843 (1979). HAPPEL, J. and BRENNER, H. Low Reynolds Number Hydrodynamics, Noordhoff International Publishing Corp., Leyden, Holland (1973). TAL, R. and SIRIGNANO, A. Heat Transfer in Sphere W. Assemblages at Intermediate Reynolds Numbers: A Cylindrical Cell Model, ASME Paper 81-WA/HT44 (1981). LABOWSKY,M. Combust. Sci. Tech. 18, 145 (1978), also ibid. 22, 217 (1980). LABOWSKY, M. and ROSNER, D. E. Evaporation and Combustion of Fuels, Advances in Chemistry Series 166, p. 63, American Chemical Society, Washington, DC (1978). BRAZIER-SMITH,P. R., JENNINGS, S. G. and LATHAM,J. Proc. R. Soc. London 326A, 393 (1972). ABRAMOVlCH,G. N. Int. J. Heat Mass Transfer 14, 1039 (1971). OWEN, P. R. J. Fluid Mech. 39, 407 (1969). MELVXLLE,W. K. and BRAY, K. N. C. Int. J. Heat Mass Transfer 22, 647 (1979), ibid. 279 (1979). DANON, H., WOLFSHTEIN, M. and HETSRONI, G. Int. J. Multiphase Flow 3, 223 (1976). LAATS, M. K. and FISHMAN, F. A. Heat Transfer --Soviet Res. 2, No. 6, 7 (1970). Ibid., Fluid Dynamics 5, 333 (1970). SREARER, A. J. and FAERY, G. M. Evaluation of a Locally Homogeneous Model of Spray Evaporation and Combustion, AIAA Paper No. 78-1042 (1978).

193. SAra~MAN,P. G. J. Fluid Mech. 22, 385 (1965). 194. ROBINOW,S. I. and KELLER, J. B. J. Fluid Mech. 11,447 (1961). 195. EICHHORN, R. and SMALL, S. J. Fluid Mech. 20, 513 (1964). 196. MACCOLL,J. W. J. Roy. Aero. Soc. 32, 777 (1928). 197. FAETH, G. M. Spray Combustion Models A Review, AIAA Paper No. 79-0293 (1979). 198. DRlSCOLL,J. F. and PELACCIO,D. C. Combust. Sci. Tech. 21,205 (1980). 199. TOROBIN, L. B. and GAUVIN, W. H. AIChE J. 7, 615 (1961). 200. CLAMEN, A. and GAUVIN, W. H. AIChE J. 15, 184 (1969). 201. HINZE, J. O. Turbulence, p. 121, McGraw-Hill, New York (1975). 202. INGEBO,R. D. NACA TN 3762 (1956). 203. CLAMEN,A. and GAUVIN, W. H. Can. J. Chem. Eng. 46, 73 (1968); ibid. 46, 223 (1968). 204. LAVENDER,W. J. and PEt, D. C. T. lnt. J. Heat Mass Transfer 10, 529 (1967). 205. BOULOS,M. I. and PEI, D. C. Can. J. Chem. En9. 47, 30 (1969). 206. GALLOWAY,T. R. and SAGE,B. H. Int. J. Heat Mass Transfer 7, 283 (1964). 207. GALLOWAY,T. R. and SAGE, B. H. Int. J. Heat Mass Transfer 10, 1195 (1967). 208. RAITHBY,G. D. and ECKERT, E. R. G. Int. J. Heat Mass Transfer 11, 1233 (1968). 209. VANDER HEGGEZIJNEN, B. C. Appl. Scient. Res. 7A, 205 (1957-8). 210. REITZ, R. D. Atomization and Other Breakup Regimes of a Liquid Jet, Ph.D. Thesis 1375-T, Princeton University (1978). 211. O'ROURKE, P. J. and BRACCO,F. V. Modeling of Drop Interactions in Thick Sprays and a Comparison with

212. 213. 214. 215. 216. 217.

218. 219.

220. 22l. 222. 223. 224. 225. 226. 227.

You might also like