A model for the prediction of the size and velocity distribution of daughter droplets created by the breakup of an unstable parent droplet is proposed. The basis of the model is the maximum entropy formalism, which states that the most probable joint probability density function (JPDF) for the daughter droplet population is the one that maximizes the Bayesian entropy conditional on the enforcement of a set of constraints, which are the conservation laws for the problem. The result is a closed-form expression for the JPDF. Validation against experimental and Direct Numerical Simulations data over the bag, multimode, and sheet-thinning breakup regimes is included. Predictions from the model show that the daughter droplet velocity distribution widens as the droplet size decreases. This result is due to a heightened sensitivity to drag force with lower droplet inertia and coincides with spray behavior. The velocity distribution is found to be near Gaussian. The model does not treat size and velocity as independently distributed, as generally assumed in the literature. In fact, marginal conditional densities derived from JPDF show that the distribution of size and velocity are interrelated.

The conventional picture of spray atomization consists of a primary breakup process, where the liquid core undergoes fragmentation, and a secondary breakup process, where the unstable liquid ligaments or droplets (referred to as parent droplets) are exposed to further breakup.1 The scope of the present work is on the development of a new model for predicting the size and velocity distribution of the population of daughter or secondary droplets generated by the breakup of an unstable parent droplet. Existing experimental investigations on droplet breakup (e.g., see Hinze,2 Simpkins and Bales,3 Krzeczkowski,4 Pilch and Erdman,5 Wierzba,6 Hsiang and Faeth,7,8 and Gelfand9) have shown that the droplet breakup process is primarily characterized by the dimensionless gas-based Weber number,

WeG=ρGup,o2Dp,oσ,
(1)

where ρG is the density of the surrounding gas medium, up,o is the initial slip velocity magnitude between the parent droplet and the surrounding gas flows, Dp,o is the initial size of the parent droplet in diameter, and σ is the droplet surface tension coefficient. The WeG gauges the ratio of the destabilizing drag force to the restoring surface tension force. The fact that this number is a gas-based Weber number rather than liquid-based is a result of dynamic pressure from the gas side, which is responsible for the deformation and breakup of the droplet.10 If this ratio is above a critical value (WeG,c), the drag force outweighs surface tension, inciting the disturbances on the droplet surface, which, in turn, causes the droplet to break up. A value of WeG,c equal to 12 has been experimentally confirmed by Pilch and Erdman5 and Lefebvre and McDonell.10 Depending on the magnitude of the liquid-based Ohnesorge number, OhL, this non-dimensional quantity also affects the breakup characteristics and breakup threshold (Ref. 7, Fig. 1). However, once OhL0.1, the influence of OhL is essentially negligible, and WeG controls the breakup process. For all cases considered in the present work, OhL is at most O(102), and thus, we limit our considerations to the WeG alone.

FIG. 1.

A schematic description of three breakup events, leading to different size distributions of daughter droplet populations pertaining to the same parent droplet condition.

FIG. 1.

A schematic description of three breakup events, leading to different size distributions of daughter droplet populations pertaining to the same parent droplet condition.

Close modal

Regarding the detailed studies of the droplet breakup dynamics, readers can refer to the review papers by Gelfand,9 Guildenbecher et al.,11 and Theofanous.12 A greater value of WeG is associated with a more violent droplet breakup mode. A classification of the breakup modes includes bag (WeG 20), multimode (WeG 60), and sheet-thinning breakups (WeG 100), as documented in a series of experimental studies of droplet breakup by Faeth and his group.1,7,8,13,14 These works were mainly focused on a single droplet breakup condition. In the recent paper by Wang et al.,15 more complex breakup cases where parent droplets are in tandem formation are considered and are experimentally investigated.

Classical approaches for modeling secondary droplet population in droplet breakup generally treat droplet size and velocity as independently distributed.16–19 Besides its use in droplet breakup, the size-velocity independence assumption is also adopted in spray impingement on a flat surface.20 To date, numerous models17,19,21–40 aimed at describing the droplet size statistics based on various approaches have been proposed. The first set of models is completely empirical, where an optimal fitting function is sought. Prominent examples include Nukiyama-Tanasawa,21 Rosin-Rammler,22 log-normal,23 upper-limit,24 chi-squared,25 and root-normal distribution functions.26 The main problem with this approach is that the model performance can significantly deteriorate if applied outside the experimental regime considered during the curve-fitting process.

The second set of models has been developed within the context of Lagrangian–Eulerian spray simulations. A popular model in this category is the Taylor Analogy Breakup (TAB) model advanced by O'Rourke and Amsden.17 Under this model, O'Rourke and Amsden treated the competing mechanisms of destabilizing and stabilizing forces acting on the droplet as a damped/forced harmonic motion. The resulting equation is a linearized ordinary differential equation governing the temporal evolution of droplet deformation. A drawback to the TAB model is that it under-estimates droplet size in diesel sprays.27 To remedy this problem, Tanner27 proposed the enhanced TAB model (ETAB) by assuming that the ratio of droplet generation rate to the droplet number is constant, and different constants are proposed corresponding to different droplet breakup regimes. Another noticeable drawback of TAB is that it cannot capture the non-linearity in large droplet deformation.28 As such, Ibrahim et al.28 proposed the droplet deformation and breakup (DDB) model, where they assumed that the time rate of change of the droplet internal energy (sum of the kinetic and potential energies) was due purely to work done on the droplet due to pressure and viscous force. Ibrahim et al.28 showed that for the medium Weber number case, such as Weber number = 50, DDB outperforms TAB. Rimbert et al.29 later improved DDB by modifying the way of pressure effect treatment. Recent advances in the TAB method can be found in Obenauf and Sojka,40 where an improved breakup criterion is introduced along with the inclusion of turbulence effects within the droplet.

Another route in estimating the secondary droplet size statistics is the Rayleigh–Taylor (RT) model based on Rayleigh–Taylor instability. One early use of the RT model was reported by Bellman and Pennington,41 where the surface tension and viscosity effects on the liquid-gas interface were analyzed in detail. Bower et al.42 later applied the RT instability to study droplet breakup in sprays. A systematic modification of the RT model was subsequently presented by Reitz and his group.19,30,31 Among these RT adaptations, the one proposed by Beale and Reitz19 is generally regarded as the most polished version.

A modeling strategy that employs the stochastic nature of the droplet breakup process has also been attempted. For instance, Gorokhovski and Saveliev32 proposed a stochastic model using Kolmogorov's theory. In this model, the mean number of daughter droplets is assumed to be independent of the parent droplet size, and as such, the droplet size distribution can be shown to take the form of a power function. Apte et al.33 proposed a hybrid Large-Eddy Simulations (LES)-stochastic model and apply it to spray simulations. The model of Gorokhovski and Saveliev is later revised by the same authors,34 with particular attention to the asymptotic behavior of the droplet size evolution.

Models based on the joint use of the probability theories and molecular dynamics are also reported in the literature. These models aim at calculating droplet breakup in turbulent flows. One early model was proposed by Luo and Svendsen.35 The model assumes that a breakup event occurs only when turbulent eddies hit a droplet. By analyzing the eddies' collision frequency, a droplet breakup probability function can be obtained. Recent progress in this methodology has extended the models' capabilities to a full spectrum of turbulence, where the energy-containing and the dissipation subranges are also present.37–39 

While the models mentioned above provide reasonable estimates of droplet size, they have some fundamental shortcomings. The crucial issue is that they assume a highly idealized physical process in their description of the droplet breakup physics. Specifically, as revealed by Direct Numerical Simulations (DNS),43–47 the breakup process begins with severe droplet deformation, leading into a bag/thin sheet formation, interfacial instabilities, ligament destabilization, and fragment dispersion. Moreover, the dynamics are highly dependent on the Weber number. Hence, the actual breakup process is far more complex than what the idealized views suggest. For example, the TAB model17 assumes a simplified oscillation treatment, significantly overlooking the underlying complexity.

Another issue is that the models generally treat droplet size and velocity as independently distributed borrowing from ideas used in spray modeling.16–20 Again, the reason for this assumption is the reduction of modeling complexity. However, this assumption does not agree with experimental results.8,48 For instance, Hsiang and Faeth8 show that there is a non-linear correlation between daughter droplet velocity and daughter droplet size. Likewise, the work by Guildenbecher et al.48 demonstrates that the droplet velocity distribution shrinks as droplet size increases. A detailed analysis of this assumption will be presented in Sec. IV E.

Due to the significant complexity of the breakup process and the need to model it without the overburdening computational expense of DNS, an attractive alternative is to employ the maximum entropy formalism (MEF).49 This principle is based on an underlying JPDF that describes the secondary droplet size and velocity distributions. The principle then states that the JPDF that maximizes the statistical entropy while simultaneously obeying a set of physical constraints is the most likely JPDF and the one chosen to characterize a given breakup process. Hence, the application of MEF targets the results of hydrodynamic breakup rather than all of the intermediate dynamics. Since the results are of interest in many applications, the use of MEF is well suited for this purpose, and it is the one implemented in the present work.

Thus far, most attempts at employing MEF in breakup problems have targeted the full liquid jet atomization and resulting spray formation problem.50–64 In those models, the maximization of statistical entropy is coupled to the constraints imposed by physical conservation laws. For the interested reader, informative reviews are provided by Babinsky and Sojka65 and Dumouchel.60 

In the case of droplet breakup, the MEF has scarcely been applied. To the authors' best knowledge, only the paper by Bodaghkhani et al.64 employs this method, in which their treatment inherits the same formulation as in primary atomization of sprays. Hence, the present work represents one of the first studies applying MEF in secondary droplet breakup problems. The new features of our model include the following:

  1. A new statistical description of the breakup process is presented, forming the basis for the ensemble-averaged equations, which are subsequently solved via the MEF method.

  2. The model is derived in a secondary droplet breakup framework, which differs from the existing MEF models derived based on the spray atomization framework.

  3. New sub-models for calculating the source terms included in momentum and energy conservation equations are developed.

In what follows, we begin with a statistical description of a droplet breakup event in Sec. II which forms the foundation for the development of the MEF model, presented in Sec. III. Results consisting of a qualitative assessment of the model performance, experimental and numerical validation, impact of the uncertainty in the sub-model for momentum and energy losses, and an examination of the independence of the size-velocity distribution are all presented in Sec. IV. Finally, a summary of the work together with the concluding thoughts is given in Sec. V.

Consider the breakup of an unstable parent droplet into a population of secondary or daughter droplets. Due to the stochastic nature of the breakup process, the discrete size and velocity distribution of the daughter droplet population, and the total number of daughter droplets, are not expected to remain the same from realization to realization (breakup event to breakup event). That is, if we were to consider, for example, three different instances of an unstable parent droplet characterized by some velocity, size, fluid, and ambient conditions, the resulting daughter droplet populations would differ. This result is illustrated in Fig. 1, indicating the randomness of the breakup phenomenon.

Despite the randomness of the breakup event, it must satisfy the following conversation laws for mass, momentum, and energy:

i=1MqπρL6Di3=πρL6Dp,o3,
(2)
i=1MqπρL6Di3ui=πρL6Dp,o3up,oΔmom,
(3)
i=1Mq12πρL6Di3ui2+i=1MqσπDi2=12πρL6Dp,o3up,o2+σπDp,o2ΔE,
(4)

where subscript q denotes an arbitrary breakup event, Mq is the number of secondary droplets created, Dp,o is the initial parent droplet size in diameter, Di is the daughter droplet diameter, up,o is the initial velocity of the parent droplet, and ui is the daughter droplet velocity. Also, σ is the surface tension coefficient, ρ is the mass density, and subscripts (L, G) denote liquid and gas. Furthermore, Δmom and ΔE represent the momentum and energy losses during the breakup process due to the aerodynamic drag force exerted by the surroundings.

To arrive at Eqs. (2)–(4), we make the following assumptions:

  1. The lateral velocity of daughter droplet populations is assumed to be negligible. Since the breakup model is targeted for use in spray simulations, the velocity of the liquid in the near-field of the spray, where the secondary droplet process mostly occurs, is heavily weighted toward the initial parent velocity direction with a minor contribution from the radial or lateral directions.

  2. The ambient velocity is treated as zero, i.e., we are considering the general case of spray formation in a quiescent environment (this can be generalized to a non-zero condition without much trouble).

  3. All daughter droplets are assumed to be spherical. As suggested by Helenbrook and Edwards,66 the assumption of sphericity holds for WeG<O(10). In our calculations, we have confirmed that the great majority of daughter droplet populations satisfy this criterion. However, there are secondary droplets with WeG>12, and thus further breakup is applied to these unstable droplets.

For the energy conservation [Eq. (4)], both kinetic energy and surface energy are considered as a single constraint. In some MEF papers, such as Sellens and Brzustowski,50 Sellens,52 and Chin et al.,53 it is argued that using a single energy constraint cannot provide information on how the total energy is distributed between these two energy modes. Hence, they introduced two separate energy conservation constraints, one for kinetic energy and the other for surface energy. However, employing separate conservation equations for surface energy and kinetic energy suggests a violation of the conservation laws established in continuum mechanics; thus, this approach is not employed in the present work.

We begin the statistical description by introducing the JPDF, which is denoted by fDu(D,u). Following convention, it obeys the normalization requirement,

ΩfDu(D,u)dDdu=1.
(5)

The expected value or ensemble average of any quantity Drus (r and s) can be obtained from

Drus=Ω(Drus)fDu(D,u)dDdu,
(6)

where the sample space, Ω2, is spanned by D[Dmin,Dmax] and u[umin,umax]. Here, Dmin, Dmax, umin, and umax are, respectively, the lower and upper bounds of D and u, both of which are discussed below in Sec. III A. For secondary droplet statistics, an equivalent manner of expressing an ensemble average is

Drus=limq{1M1+M2++Mq(i=1M1Diruis+i=1M2Diruis++i=1MqDiruis)},
(7)

where M1, M2, and Mq are the total number of secondary droplets in realization 1, 2, or generally in realization q. The expected value of M can be similarly obtained from

M=limq{M1+M2++Mqq}.
(8)

To obtain an ensemble-averaged mass conservation equation, we begin with Eq. (7) evaluated at r =3 and s =0, namely, the expected value of D3,

D3=limq{1M1+M2++Mq(i=1M1Di3+i=1M2Di3++i=1MqDi3)}.
(9)

Since the mass conservation expression in Eq. (2) holds for any breakup event, i.e.,

i=1M1Di3=i=1M2Di3==i=1MqDi3=Dp,o3,
(10)

this allows to rephrase Eq. (9) as

D3=limq{1M1+M2++Mq(Dp,o3+Dp,o3++Dp,o3)}=limq{qDp,o3M1+M2++Mq}.
(11)

Noting that the parent droplet size Dp,o does not change from realization to realization yields

D3=Dp,o3limq{qM1+M2++Mq}.
(12)

Introducing Eq. (8) in this expression gives the ensemble-averaged form of mass conservation as

MD3=Dp,o3.
(13)

For momentum, we start with the expected value of D3u given by

D3u=limq{1M1+M2++Mq(i=1M1Di3ui+i=1M2Di3ui++i=1MqDi3ui)}.
(14)

From Eq. (3), the conservation of momentum holds for all realizations, i.e.,

i=1M1Di3ui=i=1M2Di3ui==i=1MqDi3ui=[Dp,o3up,oΔmom(π/6)ρL],
(15)

where Δmom is a deterministic quantity that is only dependent on parent droplet conditions, and thus it remains unchanged in all breakup events. Substituting Eq. (15) into Eq. (14) gives

D3u=limq{(Dp,o3up,oΔmom(π/6)ρL)+(Dp,o3up,oΔmom(π/6)ρL)++(Dp,o3up,oΔmom(π/6)ρL)M1+M2++Mq}=limq{q(Dp,o3up,oΔmom(π/6)ρL)M1+M2++Mq}.
(16)

Using the expression for M in Eq. (8) yields the ensemble-averaged form for momentum conservation,

MD3u=Dp,o3up,oΔmom(π/6)ρL.
(17)

Last, for energy conservation, the goal is to obtain an expected value expression for the kinetic and surface energy term. Beginning with

πρL12D3u2+σπD2=limq{i=1M1(πρL12Di3ui2+σπDi2)+i=1M2(πρL12Di3ui2+σπDi2)++i=1Mq(πρL12Di3ui2+σπDi2)M1+M2++Mq},
(18)

we note that energy is conserved for each breakup event, namely,

i=1M1πρL12Di3ui2+σπDi2=i=1M2πρL12Di3ui2+σπDi2==i=1MqπρL12Di3ui2+σπDi2=πρL12Dp,o3up,o2+σπDp,o2ΔE.
(19)

Substituting Eq. (19) into Eq. (18) yields

πρL12D3u2+σπD2=limq{q[πρL12Dp,o3up,o2+σπDp,o2ΔE]M1+M2++Mq}.
(20)

Again, since the parent droplet conditions do not change from realization to realization, this yields the ensemble-averaged form of energy conservation,

M(πρL12D3u2+σπD2)=πρL12Dp,o3up,o2+σπDp,o2ΔE,
(21)

where Eq. (8) has been used.

In the MEF model, explicit use of fDu(D,u) is made. Hence, the expected values appearing in Eqs. (13), (17), and (21) are expressed in terms of this JPDF, namely,

MΩ(D3)fDu(D,u)dDdu=Dp,o3,
(22)
MΩ(D3u)fDu(D,u)dDdu=Dp,o3up,oΔmom(π/6)ρL,
(23)
πρL12MΩ(D3u2)fDu(D,u)dDdu+σπMΩ(D2)fDu(D,u)dDdu=πρL12Dp,o3up,o2+σπDp,o2ΔE.
(24)

Substitution for M from Eq. (13) transforms Eqs. (22)–(24) into

1D3Ω(D3)fDu(D,u)dDdu=1,
(25)
1D3up,oΩ(D3u)fDu(D,u)dDdu=1Δ¯mom,
(26)
1D3up,o2Ω(D3u2)fDu(D,u)dDdu+12WeLDp,oD3Ω(D2)fDu(D,u)dDdu=1+12WeLΔ¯E,
(27)

where Δ¯mom=Δmom/[(π/6)ρLDp,o3up,o] and Δ¯E=ΔE/[(π/12)ρLDp,o3up,o2] are momentum loss and energy loss over the breakup process in dimensionless form. These terms are discussed in Sec. II A, Also, WeL=(ρLup,o2Dp,o)/σ is the liquid-based Weber number. Equations (25)–(27), along with the normalization constraint of fDu(D,u) defined in Eq. (5), constitute a set of constraints that fDu(D,u) must satisfy.

We begin with the equation of motion for a droplet,18 

16πDp,o3ρLdup(t)dt=πDp2(t)8ρGCD(t)|uEul(xd(t))up(t)|(uEul(xd(t))up(t))+πDp3(t)6ρLg,
(28)

where only the drag and gravitational terms are accounted for since ρLρG. In this expression, up(t) denotes the parent droplet velocity, Dp(t) the parent droplet size, CD(t) the drag coefficient, uEul(xd(t)) the gas velocity at the droplet position xd(t), and g the gravity. In the current model, droplet size and drag coefficient are treated as functions of time. Considering the breakup of a droplet into a quiescent environment, such that uEul(xd(t)) is set to zero (non-zero values for gas velocity can also be considered), Eq. (28) can be simplified to

16πDp,o3ρLdup(t)dt=πDp2(t)8ρGCD(t)up(t)(up(t)),
(29)

where Dp(t),up(t), and CD(t) are unknowns. In this expression, it is being acknowledged that drag forces dominate, and gravitational effects are almost negligible. This material is illustrated in greater detail in  Appendix A.

Address first Dp(t). This parameter is obtained by matching the experimental findings of a droplet breakup reported by Hsiang and Faeth,7 where this quantity is described by

Dp(t)Dp,oDmaxDp,o=0.6214ttchar,
(30)

over the range of (10<WeG<105). Note that our interest is for WeG>12, and thus the above relationship applies. Here, Dmax is the maximum droplet size at the onset of the breakup, which can be correlated with WeG7 as

DmaxDp,o=1+0.19WeG1/2.
(31)

The characteristic time for breakup has been reported by Nicholls and Ranger67 as

tchar=Dp,o(ρL/ρG)1/2up,o.
(32)

Combining the two empirical correlations in Eqs. (30) and (31) yields the final form of Dp(t),

Dp(t)=Dp,o(1+0.1181WeG1/2ttchar).
(33)

The change in drag coefficient due to a deforming droplet, CD(t), is given empirically by Hsiang and Faeth7 as

CD(t)=0.9283Dp(t)Dp,o0.4883.
(34)

The relation in Eq. (34) holds for ReG in the range of 1000–2500, where ReG=(ρGDp,oup,o)/μG, and μG is the dynamic viscosity of the gas. In the limits of no deformation, i.e., Dp(t)=Dp,o, Eq. (34) converts back to the drag coefficient for a solid sphere.68 In our calculations, the test ReG range is 773–2446, corresponding to a range for WeG from 12 to 120. Hence, Eq. (34) is mostly applicable.

Returning to the parent droplet's equation of motion, substitutions for Dp(t) [Eq. (33)] and CD(t) [Eq. (34)] are incorporated into Eq. (29), giving

16πDp,o3ρLdup(t)dt=π{Dp,o(1+0.1181WeG1/2ttchar)}28ρG{0.9283(1+0.1181WeG1/2ttchar)0.4883}up2(t).
(35)

This expression can be manipulated to yield

dup(t)up2(t)=34ρGρL1Dp,o{0.9283(1+0.1181WeG1/2ttchar)30.4883(1+0.1181WeG1/2ttchar)2}dt.
(36)

Integrating from t =0 to an arbitrary value of t gives

1up(t)1up,o=0t34ρGρL1Dp,o{0.9283(1+0.1181WeG1/2ttchar)30.4883(1+0.1181WeG1/2ttchar)2}dt.
(37)

Solving this equation for up(t) gives

up(t)={1up,o+34ρGρLtDp,o(0.44+0.1068WeG1/2tchart+0.01068WeGtchar2t2+0.0003823WeG3/2tchar3t3)}1.
(38)

Let tbreak denote the time required to reach the droplet breaking point. Then the corresponding velocity at this time is

up(tbreak)={1up,o+34ρGρLtbreakDp,o(0.44+0.1068WeG1/2tchartbreak+0.01068WeGtchar2tbreak2+0.0003823WeG3/2tchar3tbreak3)}1.
(39)

Pilch and Erdman's experimental correlation of tbreak5 is

tbreak=1.9(WeG12)0.25tchar,
(40)

where tchar was previously defined in Eq. (32). Substituting this empirical correlation into the equation for up(tbreak) [Eq. (39)] results in

up(tbreak)={1up,o+1.425up,oρGρLΦ(0.44+0.2029WeG1/2Φ+0.03855WeGΦ2+0.002622WeG3/2Φ3)}1,
(41)

where Φ=(WeG12)0.25.

With expressions for the final droplet velocity at the breakup point, the momentum loss (Δmom) is given by

Δmom=π6Dp,o3ρL(up,oup(t)),
(42)

or in dimensionless form as

Δ¯mom=Δmom(π/6)ρLDp,o3up,o=1up(tbreak)up,o.
(43)

Substituting Eq. (39) into Eq. (43) yields the final form of Δ¯mom as

Δ¯mom=1{1+1.425ρGρLΦ(0.44+0.2029WeG1/2Φ+0.03855WeGΦ2+0.002622WeG3/2Φ3)}1.
(44)

Since Eq. (44) shows that Δ¯mom is a function of WeG and the liquid-gas density ratio (ρL/ρG), an instructive exercise is to plot a 2D iso-surface contour of Δ¯mom in terms of WeG and ρL/ρG. The result is shown in Fig. 2. The parameter range considered for WeG is 20–100, covering the bag, multimode, and sheet-thinning breakup regimes. For ρL/ρG, the range is between 20 and 1000, which is the expected range employed in engineering sprays. The results from Fig. 2 demonstrate that Δ¯mom is not sensitive to changes in WeG, remaining nearly unchanged as this parameter is varied. However, it is highly sensitive to ρL/ρG, particularly as this quantity decreases from approximately 265 to 1000. This behavior is attributed to the functional dependence of momentum loss to WeG and ρL/ρG, which from Eq. (44) can be estimated to be Δ¯mom(ρL/ρG)1/2 and Δ¯momWeG1/4.

FIG. 2.

Distributions of Δ¯mom as a function of WeG and ρL/ρG.

FIG. 2.

Distributions of Δ¯mom as a function of WeG and ρL/ρG.

Close modal

For ΔE, we consider a kinetic energy form, which is related to Δmom via

ΔE=Δmom22(π/6)ρLDp,o3
(45)

or in dimensionless form as

Δ¯E=Δ¯mom2.
(46)

Thus,

Δ¯E={1{1+1.425ρGρLΦ(0.44+0.2029WeG1/2Φ+0.03855WeGΦ2+0.002622WeG3/2Φ3)}1}2.
(47)

Based on the same parameter ranges considered in Fig. 2, the global maximum of Δ¯E is only around 0.04.

An unlikely, but potential issue, exists with the proposed equations for Δ¯mom and Δ¯E at the precise stability limit, i.e., at WeG = 12. At this boundary, Δ¯mom=1 and Δ¯E=1. This result indicates that 100% of the parent droplet momentum is consumed by drag force during the breakup process. In this circumstance, the momentum equation defined in Eq. (26) indicates that the expected value of daughter droplet momentum is zero, which is unphysical. This problem is caused by the singularity of the breakup time, tbreak=1.9(WeG12)0.25tchar, which diverges to infinity at WeG = 12. For WeG sufficiently close to 12, it will take an impractically long time for the breakup to occur, therefore depleting all the parent droplet momentum. To avoid this unphysical situation, another empirical time suggested by Guildenbecher et al.11 in initiating a bag breakup event, 2tchar, is also considered. The revised expression for tbreak is

tbreak=min[1.9(WeG12)0.25tchar,2tchar].
(48)

The maximum entropy principle states that potentially infinite numbers of distributions can satisfy a set of physical constraints. Among these, the most likely probability distribution is the one that maximizes the measure of entropy as explained by Kapur.49 The most classical measure is the Shannon entropy,56,69,70

SShannon=ΩfDu(D,u)ln[fDu(D,u)]dDdu.
(49)

Unfortunately, the Shannon entropy does not take any underlying physics-based or empirical information into account. For hydrodynamic breakup problems, exercising the maximum entropy principle does not generally yield a globally proper JPDF. A noticeable issue has been reported by Babinsky and Sojka65 in the limit as D0, where fDu(D,u) is expected to be zero, but an implementation based on the maximization of the Shannon entropy produces an fDu(D,u) that does not approach zero in this limit. The expectation of this zero limit comes from the fact that below some minuscule droplet size, there are no more droplets. For circumventing this issue, the Bayesian entropy is used instead of the Shannon entropy, which is defined57 as

SBayesian=ΩfDu(D,u)ln[fDu(D,u)fo(D)]dDdu,
(50)

where fo(D) is a priori droplet size distribution known empirically for small droplet populations characterized by DD30. Here, D30 is the mass mean diameter and is defined as

D30=(Ω(D3)fDu(D,u)dDduΩ(D0)fDu(D,u)dDdu)1/3=(Ω(D3)fDu(D,u)dDdu)1/3=D31/3.
(51)

Comparing Shannon entropy with Bayesian entropy shows that Shannon entropy is a particular case of Bayesian entropy with fo(D)=1, implying that all small droplets have the same size, which is generally not valid in breakup problems.65 The details of fo(D), D30, and their relation to the present MEF work will be discussed later in Sec. III A.

To obtain an fDu(D,u) that maximizes the Bayesian entropy while satisfying the physical conservation laws, we employ the Lagrange multiplier method described in Kapur.49 This JPDF is further refined by enforcing that the small size droplet distribution (DD30) must follow a prescribed probability density function (PDF) given by fo(D). First, the constraints imposed on fDu(D,u)[see Eqs. (5) and (25)–(27)] are re-cast in the form of gi(fDu(D,u))=0,i= 0, 1, 2, and 3, namely,

g0(fDu(D,u))=DminDmaxuminumaxfDu(D,u)dudD1=0,
(52)
g1(fDu(D,u))=1D3DminDmaxuminumax(D3)fDu(D,u)dudD1=0,
(53)
g2(fDu(D,u))=1D3up,oDminDmaxuminumax(D3u)fDu(D,u)dudD(1Δ¯mom)=0,
(54)
g3(fDu(D,u))=1D3up,o2DminDmaxuminumax(D3u2)fDu(D,u)dudD+12WeLDp,oD3DminDmaxuminumax(D2)fDu(D,u)dudD(1+12WeLΔ¯E)=0,
(55)

corresponding to the normalization requirement, mass, momentum, and energy conservation.

Next, four Lagrange multipliers, λi, i= 0, 1, 2, and 3, are introduced into the following function:

L(fDu(D,u),λ)=SBayesiani=03λigi(fDu(D,u)).
(56)

Under the Lagrange multiplier method, the optimum fDu(D,u) is obtained from

L(fDu(D,u),λ)fDu(D,u)=0.
(57)

Substituting all relevant equations [see Eqs. (50) and (52)–(55)] into Eq. (57) yields

fDu(D,u){DminDmaxuminumaxfDu(D,u)ln(fDu(D,u)fo(D))dudDλ0(DminDmaxuminumaxfDu(D,u)dudD)λ1(1D3DminDmaxuminumax(D3)fDu(D,u)dudD)λ2(1D3up,oDminDmaxuminumax(D3u)fDu(D,u)dudD)λ3(1D3up,o2DminDmaxuminumax(D3u2)fDu(D,u)dudD+12WeLDp,oD3DminDmaxuminumax(D2)fDu(D,u)dudD)+(λ0+λ1+λ2(1Δ¯mom)+λ3(1+12WeLΔ¯E))}=0.
(58)

After some manipulations the above equation becomes

ln(fDu(D,u)fo(D))=1λ0λ1D3D3λ2D3uD3up,oλ3(D3u2D3up,o2+12WeLDp,oD2D3).
(59)

Letting λ0*=1+λ0, the optimum JPDF that maximizes Bayesian entropy in the presence of the constraints is found to be

fDu(D,u)=fo(D)exp{λ0*λ1D3D3λ2D3uD3up,oλ3(D3u2D3up,o2+12WeLDp,oD2D3)}.
(60)

Overall, the MEF system consists of four equations, Eqs. (52)–(55) with six unknowns, namely, [λ0*,λ1,λ2,λ3,fo(D), and D3]. To have a unique solution to this system requires introducing two more independent equations that fo(D) and D3 have to satisfy.

A mathematical form for fo(D) that matches the empirical size distributions for DD30 is sought. Dumouchel59 indicates that fo(D) should be related to D, due partially to the role of surface tension forces in characterizing liquid-gas interfaces. Continuing this idea, Dumouchel59 suggests that fo(D) should follow a power-law distribution with respect to D for small droplets. Following this suggestion, we assign fo(D) the well-established Nukiyama–Tanasawa distribution,10 given by

fo(D)=BDpexp[(D/C)q],
(61)

where B is a constant, C is the size parameter, and p and q are distribution parameters.

We will treat fo(D) as a PDF, which automatically implies obeying the normalization requirement. González-Tello et al.71 suggest that in spray problems, the size of small droplets characterized by DD30 is much smaller than the size parameter C, i.e., DC. Expanding out the exponential term in Eq. (61) yields

fo(D)=BDp1+(D/C)q+(D/C)2q2!+(D/C)3q3!+.
(62)

Also, Li and Tankin51 point out that the constant q is typically greater than one. Hence, the entire denominator in Eq. (62) can be approximated with good accuracy to be equal to one. Furthermore, the parameter p is determined by matching the empirical observations of Li and Tankin51 and Lefebvre and McDonell10 in liquid jet breakup problems, which gives p= 2. With all of this information, fo(D)=BD2, and from the normalization requirement

DminDmaxuminumaxfo(D)dudD=1,
(63)

gives

B=3(Dmax3Dmin3)(umaxumin).
(64)

Following the procedure documented by Movahednejad et al.61 for spray problems, we assign the following respective bound values:

Dmin=0,Dmax=3D31/3,umin=0,andumax=3up,o.
(65)

Hence, the final form of fo(D) is

fo(D)=BD2=127D2D3up,o.
(66)

Regarding D3, it is constructed by performing a non-linear regression to the DNS data from Jain et al..46 The correlation is then calibrated in line with spray measurements of D30, i.e., D31/3, from Kim et al.,57 yielding

D3=Dp,o3(5.5392WeG2.2207)1.
(67)

Equation (67) shows that D31/3 is correlated with up,o via D31/3up,o1.48. The droplet breakup experiments from Wolfe and Andersen72 also report a similar empirical correlation of D31/3up,o1.33, which provides some degree of reassurance. Combining fo(D) together with the general JPDF in Eq. (60) gives the final expression for fDu(D,u) as

fDu(D,u)=127D2D3up,oexp{λ0*λ1D3D3λ2D3uD3up,oλ3(D3u2D3up,o2+12WeLDp,oD2D3)}.
(68)

The results are divided into five sections. In Sec. IV A, a qualitative assessment of fDu(D,u) is completed to confirm whether the key features of this distribution resemble those predicted in previous MEF spray breakup studies, namely, in the works of Sellens,52 Li et al.,73 Ayres et al.,56 Movahednejad et al.,61 Yan et al.,62 and Fu et al.63 In Sec. IV B, the MEF model is compared to experiments from Hsiang and Faeth.7,8 A similar comparison is reported in Sec. IV C, but this time the source is from a DNS by Jalaal and Mehravaran.43 Since the MEF model includes source terms to address the effects of drag force on droplet momentum and energy, the sensitivities of uncertainties in these source terms to fDu(D,u) are studied and presented in Sec. IV D. Finally, the independence of size and velocity distributions is examined.

The characteristics of fDu(D,u) are examined for a case consisting of the breakup of a droplet within the multimode breakup regime. This regime is chosen because of its relatively complex breakup mechanism compared to the bag and sheet-thinning regimes.74 As such, it can serve as a good benchmark for assessing the model performance. The droplet, in this case, has a diameter (Dp,o) of 200 μm and a velocity (up,o) of 135.07 m/s, with a respective Weber number (WeG) of 60. The physical properties are listed in Table I. The predicted fDu(D,u) using the MEF model is shown as a 2D surface plot in Fig. 3.

TABLE I.

Physical properties for the cases in analyzing fDu(D,u).

 Liquid density (water) ρL = 997 kg/m3  
 Gas density (air) ρG = 1.1839 kg/m3  
 Coefficient of surface tension σ = 0.072 kg/s2  
 Liquid density (water) ρL = 997 kg/m3  
 Gas density (air) ρG = 1.1839 kg/m3  
 Coefficient of surface tension σ = 0.072 kg/s2  
FIG. 3.

Illustrations of the predicted distribution, fDu(D,u), of secondary droplets resulting from the breakup of a parent droplet at WeG = 60.

FIG. 3.

Illustrations of the predicted distribution, fDu(D,u), of secondary droplets resulting from the breakup of a parent droplet at WeG = 60.

Close modal

The first noticeable characteristic from Fig. 3 is that the velocity distribution becomes wider with decreasing drop size. This behavior is attributed to an increased sensitivity to drag force with small droplet inertia. Similar findings have also been reported for sprays by Sellens.52 A second characteristic is that fDu(D,u) decays significantly for droplets below approximately 2 μm, reaching zero at D 2 μm. This result is attributed to the fo(D), i.e., small droplet distribution. Another feature is that fDu(D,u) exhibits the near-symmetric velocity distribution about up,o. With regard to spray problems, similar behavior was also documented in Sellens and Brzustowski,50 where the authors state that the symmetric behavior is very similar to a Gaussian profile. Sellens and Brzustowski50 explain this observation by arguing that the liquid momentum is a linear function in the velocity space, and the liquid kinetic energy is a square function of velocity. These two constraints taken in combination should allow for the estimation of the mean and variance of velocity. With the information of mean and variance, a Gaussian distribution is the most likely distribution that can be deduced. As a means of confirming this near-Gaussian behavior for the velocity distribution, the respective marginal velocity density function,

fu(u)=DminDmaxfDu(D,u)dD,
(69)

is compared against a Gaussian distribution in Fig. 4. For this Gaussian distribution, the mean value is equal to the parent droplet velocity up,o, and the standard deviation is 0.26 up,o. Overall, good qualitative matching is observed, confirming expectations from Sellens and Brzustowski.50 

FIG. 4.

The comparison between the predicted marginal velocity distribution, fu(u), and the Gaussian distribution with the mean value up,o and a standard deviation of 0.26 up,o.

FIG. 4.

The comparison between the predicted marginal velocity distribution, fu(u), and the Gaussian distribution with the mean value up,o and a standard deviation of 0.26 up,o.

Close modal

Since the droplet breakup severity is heavily influenced by WeG, it is expected that fDu(D,u) would reflect this sensitivity. To examine this aspect, we consider a parent water droplet having a size of 200 μm and impose different initial velocities of 77.98, 135.07, and 174.38 m/s, corresponding to WeG of 20, 60, and 100, respectively. The choice of these Weber numbers ensures that the three regimes of the breakup, namely, bag, multimode, and sheet-thinning breakup, are considered. The surrounding medium is air, and physical properties are the same as those listed in Table I. Surface plots of fDu(D,u) pertaining to the three different cases are shown Fig. 5. Similar to the previous case, the density function is symmetric about the parent droplet velocity. The maximum value of the JPDF, fDu(D,u)|max, is located at (D, u) = (12.86 μm, 74.71 m/s), (5.75 μm, 131.01 m/s), and (3.95 μm, 169.74 m/s) corresponding, respectively, to WeG = 20, 60, and 100. Hence, there is a natural shift to higher velocities and smaller drop sizes with increasing WeG.

FIG. 5.

Distributions of fDu(D,u) in (a) bag (WeG = 20), (b) multimode (WeG = 60), and (c) sheeting-thinning breakup regimes (WeG = 100).

FIG. 5.

Distributions of fDu(D,u) in (a) bag (WeG = 20), (b) multimode (WeG = 60), and (c) sheeting-thinning breakup regimes (WeG = 100).

Close modal

A related trend displayed in Fig. 5 is the concentration of progressive smaller daughter droplets with increasing breakup intensity, i.e., larger values for WeG. This pattern is quantified by calculating the size variance of fDu(D,u), i.e., σD2 = DminDmaxuminumax(DD)2fDu(D,u)dudD. This variance, σD2, is found to be equal to 22.43, 4.56, and 2.14 μm2, corresponding again to the respective WeG = 20, 60, and 100 cases. This characteristic is one of the most distinct features of the breakup process with increasing WeG.

Breakup predictions from the MEF model in the bag, multimode, and sheet-thinning breakup regimes are compared to the experiments from Hsiang and Faeth.7,8 Since in Hsiang and Faeth's work,7 a variable drop size is used for different test fluids, which include water, n-heptane, ethyl alcohol, mercury, and glycerol solutions, in the model predictions, we chose to let the initial parent droplet size match the one for water drop conditions, which is 1000 μm. The results are quantified using a normalized cumulative volume distribution denoted by CDF(D3) as a function of droplet size for different WeG. The parent droplet is formed using water and its initial velocity is systematically varied to let the parent droplet have a respective WeG equal to 20, 60, and 100. The surrounding medium is air, and physical properties are the same as those listed in Table I.

The comparisons between model solutions and the experimental measurements are shown in Fig. 6. In the figure, the x-axis is the droplet diameter normalized by the mass medium diameter MMD, which is implicitly defined by

DminMMDD3fD(D)dDDminDmaxD3fD(D)dD=0.5.
(70)
FIG. 6.

Comparisons between the model estimates and Hsiang and Faeth's experimental measurements:7,8 (a) bag (WeG = 20), (b) multimode (WeG = 60), and (c) sheeting-thinning breakup regimes (WeG = 100).

FIG. 6.

Comparisons between the model estimates and Hsiang and Faeth's experimental measurements:7,8 (a) bag (WeG = 20), (b) multimode (WeG = 60), and (c) sheeting-thinning breakup regimes (WeG = 100).

Close modal

The y-axis of Fig. 6 shows the values of CDF(D3) calculated by

CDF(D3)=DminDD3fD(D)dDDminDmaxD3fD(D)dD,
(71)

where fD(D) is the marginal size density function given by

fD(D)=uminumaxfDu(D,u)du.
(72)

Considering the results shown in Fig. 6, at least two observations can be made. The first is that for values D/MMD below one, the agreement with experimental data is appreciably better than for the range D/MMD greater than one. The exception to this trend is for the sheet-thinning regime employing WeG = 100, where the level of agreement between larger and smaller droplets relative to MMD is approximately the same. A second observation concerns the noticeable deviation between the MEF predicted curves and the experiments for the larger secondary droplet populations. MEF does not predict as many larger droplets as experiments would indicate.

1. MMD/D32

Another metric that can be used in validating the droplet size characteristics is the ratio of MMD to D32, where D32 is the Sauter mean diameter defined as

D32=DminDmaxD3fD(D)dDDminDmaxD2fD(D)dD.
(73)

This metric was originally advanced by Simmons,26 where the author showed that MMD/D32 was universally equal to 1.2 under various spray operating conditions. Hsiang and Faeth7,8 and Guildenbecher et al.48 later show that this conclusion still holds in droplet breakup experiments in the bag, the multimode, and the sheet-thinning breakup regimes. Our calculations show that the value of MMD/D32 is 1.1 in each WeG case reported, agreeing with the experimental values within 10%.

We compare our predicted marginal droplet size distributions, fD(D), to results from DNS reported by Jalaal and Mehravaran.43 In these simulations, the authors43 employed the open-source solver Gerris, a geometric Volume-of-Fluid (gVoF) algorithm with adaptive mesh refinement (AMR) capabilities developed by Popinet,75 to capture the liquid–gas interface. A total of three cases are considered having Weber numbers and parent droplet sizes given, respectively, by (WeG, D) = (37.6, 3.2×104μm), (WeG, D) = (47.8, 3.9×104μm), and (WeG, D) = (59, 4.5×104μm). The physical properties are the same as listed in Table I. In the work of Jalaal and Mehravaran,43 the breakup is initiated by the falling action of the droplet due to gravity.

Results and comparisons are shown in Fig. 7 according to each of the three conditions considered. The overall agreement is reasonable for the larger secondary droplet, i.e., D103μm. For the smaller droplet distribution, the deviation between the model and the DNS data is substantial. Besides recognizing the possibility of modeling error in this part of the droplet size spectrum, other sources of discrepancy should be considered. For instance, in all Weber numbers provided, the fD(D) from the DNS goes to zero at roughly half Δxmin, where Δxmin is the minimum mesh size. Hence, there is not enough spatial resolution to capture these small-scale features. Previous work in our group performing similar DNS76 indicates that at least O(10) cells should be employed across the diameter of a droplet to capture its large scale dynamics. Thus, it can be argued that insufficient grid resolution is being employed by the authors43 for the smallest part of the daughter droplet population.

FIG. 7.

Comparison against DNS data from Jalaal and Mehravaran43 at (a) WeG = 37.6, (b) WeG = 47.8, and (c) WeG = 59. The vertical dashed line in each plot indicates the minimum grid size used in the DNS.

FIG. 7.

Comparison against DNS data from Jalaal and Mehravaran43 at (a) WeG = 37.6, (b) WeG = 47.8, and (c) WeG = 59. The vertical dashed line in each plot indicates the minimum grid size used in the DNS.

Close modal

Consider the momentum and energy conservation equations. Losses in momentum and energy are accounted for through Δ¯mom and Δ¯E, respectively. Since these terms are obtained from models, they are guaranteed not to be perfect. For the sake of argument, let us assume that we have a ±30% uncertainty in Δ¯mom, which will transcend to ±69% uncertainty in Δ¯E since both terms are related by Δ¯E=Δ¯mom2.

A plot of Δ¯mom shown in Fig. 2 (see Sec. II A) indicates that its greatest sensitivity is due to changes in liquid-gas density ratio instead of the Weber number. Hence, in the following analysis, we consider three different scenarios corresponding to ρL/ρG = 1000, 100, and 50. All of which correspond to WeG = 60, i.e., the same multimode regime studied previously in Sec. IV A. In each case, the parent droplet has an initial size of 200 μm, a density of 1000 kg/m3, and a surface tension coefficient of 0.072 kg/s2.

At each density ratio, the nominal value for Δ¯mom is varied ± 30%, and its effects on the resulting secondary droplet population are examined. Here, we are interested in the effects of the uncertainty of Δ¯mom on fDu(D,u) as a function of ρL/ρG, not on the changes of fDu(D,u) as a result of variations of density ratio. This point is subtle but important. The results, expressed in terms of mean values with respect to size and velocity and the respective standard deviations, are included in Table II. The mean values of these statistics are normalized by the respective nominal quantities. For instance, at ρL/ρG=1000, the mean diameter corresponding to Δ¯mom±(0.3)Δ¯mom is divided by the mean diameter evaluated at Δ¯mom. For standard deviations, they are normalized by the mean values. This operation gives a direct indication of the sensitivity of each statistic to the uncertainty associated with Δ¯mom.

TABLE II.

Effects of the uncertainty of Δ¯mom on mean values for secondary droplet size and velocity along with their respective standard deviations.

ρL/ρG = 1000ρL/ρG = 100ρL/ρG = 50
Δ¯mom30%Nominal−30%30%Nominal−30%30%Nominal−30%
D/Dnom 1.006 0.991 1.014 0.967 1.032 0.881 
σD/D 0.458 0.467 0.473 0.439 0.463 0.529 0.452 0.504 0.790 
u/unom 1.003 0.996 1.003 0.997 0.999 1.018 
σu/u 0.452 0.423 0.386 0.550 0.505 0.446 0.565 0.513 0.479 
ρL/ρG = 1000ρL/ρG = 100ρL/ρG = 50
Δ¯mom30%Nominal−30%30%Nominal−30%30%Nominal−30%
D/Dnom 1.006 0.991 1.014 0.967 1.032 0.881 
σD/D 0.458 0.467 0.473 0.439 0.463 0.529 0.452 0.504 0.790 
u/unom 1.003 0.996 1.003 0.997 0.999 1.018 
σu/u 0.452 0.423 0.386 0.550 0.505 0.446 0.565 0.513 0.479 

The general trend observed from the results in Table II is one of increasing uncertainty impacts with decreasing density ratio. Only the mean velocity shows negligible variation with different values of ρL/ρG, but, even in this case, there is a slighter larger variation at ρL/ρG=50 in comparison to the other cases. This heightened sensitivity to uncertainties in Δ¯mom at the low end of the density ratio spectrum is a direct result of increased momentum exchange between the droplet and its surroundings due to large gas density. Specifically, the low-density ratio is governed by a higher gas density since ρL remains the same. Hence, any variations in Δ¯mom at higher ρG will have greater impact on the characteristics of fDu(D,u).

As discussed, previous spray models often assume that the velocity and size distributions of daughter droplets are independent of each other16–20,53 without actually examining the validity of this assumption. To inspect this treatment in the present work, we simply consider the respective conditional densities. For reference, if D and u are independently distributed, then the corresponding JPDF is simply a product of the marginal densities, i.e., hDu(D,u)=hD(D)hu(u). Consequently, the distribution of velocity conditional on D = Da is given by

hu|D=Da(u)=hDu(D=Da,u)uminumaxhDu(D=Da,u)du=hD(D=Da)hu(u)uminumaxhD(D=Da)hu(u)du=hD(D=Da)hu(u)hD(D=Da)uminumaxhu(u)du=hu(u),
(74)

which simply gives the same hu(u) for any value of Da due to independence. Similar dependencies can also be shown for the marginal size distribution conditional on a specified velocity of daughter droplet, i.e., hD|u=ua(D).

Regarding the present JPDF corresponding to the MEF, its conditional marginal density is

fu|D=Da(u)=fDu(D=Da,u)uminumaxfDu(D=Da,u)du=fDu(D=Da,u)fD(Da),
(75)

where fD(D) has a closed-form expression given by (see  Appendix B for the derivation)

fD(D)=erf(ξmax)erf(ξmin)54πDλ3D3exp{λ0*λ1D3D3+λ3(D3D3(λ22λ3)212WeLDp,oD2D3)}.
(76)

Here, ξ(u)=(λ3D3)/D3[u/up,o+λ2/(2λ3)],ξmax=ξ(umax),ξmin=ξ(umin), and erf(ξ(u))=(2/π)0ξ(u)exp(x2)dx is the error function of ξ(u).

Substituting Eqs. (76) and (68) into Eq. (75) gives

fu|D=Da(u)=2Da2D3up,oexp{λ0*λ1Da3D3λ2Da3uD3up,oλ3(Da3u2D3up,o2+12WeLDp,oDa2D3)}(erf(ξmax)erf(ξmin))πDaλ3D3exp{λ0*λ1Da3D3+λ3(Da3D3(λ22λ3)212WeLDp,oDa2D3)},
(77)

which after some manipulations can be simplified as (see  Appendix C)

fu|D=Da(u)=2DaDaup,o(erf(ξmax)erf(ξmin))λ3πD3exp{Da3D3(λ2uup,oλ3(u2up,o2+λ224λ32))}.
(78)

This expression clearly shows that the marginal density conditional on D = Da is explicitly dependent on Da, i.e., the independently distributed condition does not apply. In particular, fu|D=Da(u) is exponentially correlated with Da3. A numerical illustration is given in Fig. 8. Similar conclusions also hold for the marginal density of size conditional on a prescribed velocity.

FIG. 8.

Illustrations of marginal velocity distributions conditional on D = Da. The parent droplet parameters are Dp,o = 200 μm and up,o=135.07m/s. The fluid properties employed are listed in Table I.

FIG. 8.

Illustrations of marginal velocity distributions conditional on D = Da. The parent droplet parameters are Dp,o = 200 μm and up,o=135.07m/s. The fluid properties employed are listed in Table I.

Close modal

Motivated by the complexity of the droplet breakup phenomena and the associated burdensome costs of performing DNS, the present effort provides an alternative treatment offered by the maximum entropy formalism. Hence, the model is not targeted at all of the complicated intermediate dynamics of the breakup process; instead, it is focused on the final outcome in terms of secondary droplet distribution with respect to velocity and size. In various practical applications, it is precisely the results of the breakup process that are of interest, and thus the present model suits this purpose. This model would also be useful in situations where many droplets exist under hydrodynamically unstable conditions and are too expensive to be resolved by a DNS. The model accounts for all the conservation laws and includes submodels for momentum and energy losses during the breakup. It provides in closed form an expression for the most probable joint probability density function in size and velocity space based on the maximization of the Bayesian entropy. While we would prefer an empiricism-free solution, we acknowledge that due to the complexity of the phenomena, which is breakup regime dependent and which is characterized by pronounced changes in droplet shape and fragmentation, some empirical relations have been introduced. This was primarily done in the closure of fo(D) and D3.

Predictions from the model show that the droplet velocity distribution widens as the droplet size decreases. This result is due to a heightened sensitivity to drag force with lower droplet inertia and coincides with spray behavior documented by Sellens.52 In addition, the velocity distribution is found to be near the Gaussian distribution. For the size distribution, it is similar to a root-normal distribution, agreeing with Guildenbecher et al.'s experimental observations on droplet breakup.48 As the Weber number increases, the model shows that the size distribution skews toward more concentrated smaller droplet sizes in agreement with experiments. Further analysis shows that the size and velocity are dependently distributed based on an inspection of the conditional marginal densities. This last feature contrasts with common assumptions made in the literature claiming that size and velocity are independently distributed.

Support from Caterpillar Inc. and the Army Research Laboratory is gratefully acknowledged. A portion of this research was sponsored by the Army Research Laboratory and was accomplished under Cooperative Agreement No. W911NF-20-2-0181. The views and conclusions contained in this document are those of the authors and should not be interpreted as representing the official policies, either expressed or implied, of the Army Research Laboratory or the U.S. Government. The U.S. Government is authorized to reproduce and distribute reprints for Government purposes notwithstanding any copyright notation herein. CEI is acknowledged for granting use of their post-processing software, EnSight.

The authors declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Let us return to the droplet momentum equation defined in Eq. (28). The drag force term is

πDp2(t)8ρGCD(t)|uEul(xd(t))up(t)|(uEul(xd(t))up(t)),
(A1)

while the gravitational term is

πDp3(t)6ρLg.
(A2)

Recalling uEul(xd(t)) = 0 in the present work, the difference in the magnitude between these two terms is

πDp2(t)8ρGCD(t)|up(t)|2πDp3(t)6ρL|g|=πDp2(t)6ρL(34ρGρLCD(t)|up(t)|2Dp(t)|g|).
(A3)

Since

O(34ρGρLCD(t)|up(t)|2)O(103101)O(1011)O(104)O(1103),
(A4)

and

O(Dp(t)|g|)O(105103)O(10)O(104102),
(A5)

it then follows from Eq. (A3) that the gravitational term is nearly negligible in the droplet momentum equation.

The other metric for evaluating the influence of the gravitational effects is the Bond number (Bo),

Bo=ρL|g|Dp,o2σ,
(A6)

which quantifies the ratio of gravitational force to surface tension force. A 2D surface plot of Bo is shown in Fig. 9 as a function of ρL and Dp,o. In this calculation, the surface tension coefficient is given as 0.072 kg/m3; the maximum values of Dp,o and ρL are given as 1000 μm and 1000 kg/m3. The results show that Bond number is O(102)O(101), which confirms the almost negligible effect of gravity.

FIG. 9.

Bond number range in the present model, which shows the negligible effect of graviational contributions.

FIG. 9.

Bond number range in the present model, which shows the negligible effect of graviational contributions.

Close modal

The procedure begins with substituting the final form of fDu(D,u) [Eq. (68)] into the definition of fD(D) [Eq. (72)], yielding

fD(D)=uminumaxfDu(D,u)du=uminumax127D2D3up,oexp{λ0*λ1D3D3λ2D3uD3up,oλ3(D3u2D3up,o2+12WeLDp,oD2D3)}du.
(B1)

Equation (B1) can be carried out analytically as follows. First, we re-write Eq. (B1) as

fD(D)=127D2D3up,oexp(λ0*λ1D3D3λ312WeLDp,oD2D3)uminumaxexp(λ2D3uD3up,oλ3D3u2D3up,o2)du.
(B2)

The integrand of Eq. (B2) after some manipulations can be re-cast as

uminumaxexp(λ2D3uD3up,oλ3D3u2D3up,o2)du=exp{λ3D3D3(λ22λ3)2}uminumaxexp{λ3D3D3(uup,o+λ22λ3)2}du.
(B3)

Perform change of variable by defining

ξ(u)=λ3D3D3(uup,o+λ22λ3),
(B4)

whose derivative is

dξ(u)du=λ3D3D3(1up,o).
(B5)

With Eqs. (B4) and (B5), the RHS of Eq. (B3) can be re-written as

exp{λ3D3D3(λ22λ3)2}uminumaxexp{λ3D3D3(uup,o+λ22λ3)2}du=exp{λ3D3D3(λ22λ3)2}up,oD3λ3D3ξminξmaxexp(ξ2)dξ=exp{λ3D3D3(λ22λ3)2}up,oD3λ3D3π2(erf(ξmax)erf(ξmin)),
(B6)

where ξmax=ξ(umax),ξmin=ξ(umin), and erf(ξ(u))=(2/π)0ξ(u)exp(x2)dx is the error function of ξ(u).

Combining Eqs. (B2), (B3), and (B6) yields the final form of fD(D),

fD(D)=erf(ξmax)erf(ξmin)54πDλ3D3exp{λ0*λ1D3D3+λ3(D3D3(λ22λ3)212WeLDp,oD2D3)}.
(B7)

The expression for fu|D=Da(u) [Eq. (77)] can be simplified as follows:

fu|D=Da(u)=2Da2D3up,oexp{λ0*λ1Da3D3λ2Da3uD3up,oλ3(Da3u2D3up,o2+12WeLDp,oDa2D3)}(erf(ξmax)erf(ξmin))πDaλ3D3exp{λ0*λ1Da3D3+λ3(Da3D3(λ22λ3)212WeLDp,oDa2D3)}=2Da2D3up,oexp{λ2Da3uD3up,oλ3(Da3u2D3up,o2+12WeLDp,oDa2D3)}(erf(ξmax)erf(ξmin))πDaλ3D3exp{λ3(Da3D3(λ22λ3)212WeLDp,oDa2D3)}
=2Da2D3up,oexp{λ2Da3uD3up,oλ3(Da3D3u2up,o2+Da3D3(λ22λ3)2)}(erf(ξmax)erf(ξmin))πDaλ3D3=2DaDaup,o(erf(ξmax)erf(ξmin))λ3πD3exp{Da3D3(λ2uup,oλ3(u2up,o2+λ224λ32))}.
(C1)
1.
G. M.
Faeth
,
L.-P.
Hsiang
, and
P.-K.
Wu
, “
Structure and breakup properties of sprays
,”
Int. J. Multiphase Flow
21
,
99
127
(
1995
).
2.
J. O.
Hinze
, “
Fundamentals of the hydrodynamic mechanism of splitting in dispersion processes
,”
AIChE J.
1
(
3
),
289
295
(
1955
).
3.
P. G.
Simpkins
and
E. L.
Bales
, “
Water-drop response to sudden accelerations
,”
J. Fluid Mech.
55
(
4
),
629
639
(
1972
).
4.
S. A.
Krzeczkowski
, “
Measurement of liquid droplet disintegration mechanisms
,”
Int. J. Multiphase Flow
6
(
3
),
227
239
(
1980
).
5.
M.
Pilch
and
C. A.
Erdman
, “
Use of breakup time data and velocity history data to predict the maximum size of stable fragments for acceleration-induced breakup of a liquid drop
,”
Int. J. Multiphase Flow
13
(
6
),
741
757
(
1987
).
6.
A.
Wierzba
, “
Deformation and breakup of liquid drops in a gas stream at nearly critical weber numbers
,”
Exp. Fluids
9
(
1
),
59
64
(
1990
).
7.
L.-P.
Hsiang
and
G. M.
Faeth
, “
Near-limit drop deformation and secondary breakup
,”
Int. J. Multiphase Flow
18
(
5
),
635
652
(
1992
).
8.
L.-P.
Hsiang
and
G. M.
Faeth
, “
Drop properties after secondary breakup
,”
Int. J. Multiphase Flow
19
(
5
),
721
735
(
1993
).
9.
B. E.
Gelfand
, “
Droplet breakup phenomena in flows with velocity lag
,”
Prog. Energy Combust. Sci.
22
(
3
),
201
265
(
1996
).
10.
A. H.
Lefebvre
and
V. G.
McDonell
,
Atomization and Sprays
(
CRC Press
,
2017
).
11.
D. R.
Guildenbecher
,
C.
López-Rivera
, and
P. E.
Sojka
, “
Secondary atomization
,”
Exp. Fluids
46
(
3
),
371
402
(
2009
).
12.
T. G.
Theofanous
, “
Aerobreakup of newtonian and viscoelastic liquids
,”
Annu. Rev. Fluid Mech.
43
,
661
690
(
2011
).
13.
W.-H.
Chou
,
L.-P.
Hsiang
, and
G. M.
Faeth
, “
Temporal properties of drop breakup in the shear breakup regime
,”
Int. J. Multiphase Flow
23
(
4
),
651
669
(
1997
).
14.
W.-H.
Chou
and
G. M.
Faeth
, “
Temporal properties of secondary drop breakup in the bag breakup regime
,”
Int. J. Multiphase Flow
24
(
6
),
889
912
(
1998
).
15.
Z.
Wang
,
T.
Hopfes
,
M.
Giglmaier
, and
N. A.
Adams
, “
Experimental investigation of shock-induced tandem droplet breakup
,”
Phys. Fluids
33
(
1
),
012113
(
2021
).
16.
R. D.
Reitz
and
R.
Diwakar
, “
Effect of drop breakup on fuel sprays
,”
Report No.
860469 (
SAE Technical Paper
,
1986
).
17.
P. J.
O'Rourke
and
A. A.
Amsden
, “
The TAB method for numerical calculation of spray droplet breakup
,”
Report No.
872089 (
SAE Technical Paper
,
1987
).
18.
A. A.
Amsden
,
P. J.
Orourke
, and
T. D.
Butler
, “
KIVA-II: A computer program for chemically reactive flows with sprays
,”
Report No. LA-11560-MS
(
Los Alamos National Laboratory
,
1989
).
19.
J. C.
Beale
and
R. D.
Reitz
, “
Modeling spray atomization with the Kelvin-Helmholtz/Rayleigh-Taylor hybrid model
,”
Atomization Sprays
9
(
6
),
623
(
1999
).
20.
M. F.
Trujillo
,
W. S.
Mathews
,
C. F.
Lee
, and
J. E.
Peters
, “
Modelling and experiment of impingement and atomization of a liquid spray on a wall
,”
Int. J. Engine Res.
1
(
1
),
87
105
(
2000
).
21.
S.
Nukiyama
and
Y.
Tanasawa
, “
Study of liquid atomization
,”
Trans. Soc. Mech. Engrs. (Jpn.)
5
,
136
136
(
1939
).
22.
P.
Rosin
, “
Laws governing the fineness of powdered coal
,”
J. Inst. Fuel
7
,
29
36
(
1933
).
23.
R. S.
Bevans
, “
Mathematical expressions for drop size distributions in sprays
,” in
Conference on Fuel Sprays
(
University of Michigan
,
1949
).
24.
R. A.
Mugele
and
H. D.
Evans
, “
Droplet size distribution in sprays
,”
Ind. Eng. Chem.
43
(
6
),
1317
1324
(
1951
).
25.
H.
Hiroyasu
, “
Mathematical expression for drop size distribution in sprays
,” Topical Report No. NASA-CR-72272 (
University of Wisconsin-Madison
,
1967
).
26.
H. C.
Simmons
, “
The correlation of drop-size distributions in fuel nozzle sprays
,”
J. Eng. Power
99
(
3
),
309
319
(
1977
).
27.
F. X.
Tanner
, “
Liquid jet atomization and droplet breakup modeling of non-evaporating diesel fuel sprays
,”
Report No. 970050
(
SAE Technical Paper
,
1997
).
28.
E. A.
Ibrahim
,
H. Q.
Yang
, and
A. J.
Przekwas
, “
Modeling of spray droplets deformation and breakup
,”
J. Propul. Power
9
(
4
),
651
654
(
1993
).
29.
N.
Rimbert
,
A.
Hajjar
,
S.
Castrillon-Escobar
,
R.
Meignen
, and
M.
Gradeck
, “
A new look at the droplet deformation and breakup model
,” in
26th Annual Conference on Liquid Atomization and Spray Systems
(
ILASS—Europe
) (
2014
).
30.
T. F.
Su
,
M. A.
Patterson
,
R. D.
Reitz
, and
P. V.
Farrell
, “
Experimental and numerical studies of high pressure multiple injection sprays
,”
Technical Report No. 960861
(
SAE Technical Paper
,
1996
).
31.
M. A.
Patterson
and
R. D.
Reitz
, “
Modeling the effects of fuel spray characteristics on diesel engine combustion and emission
,”
Technical Report No. 980131
(
SAE Technical Paper
,
1998
).
32.
M. A.
Gorokhovski
and
V. L.
Saveliev
, “
Analyses of Kolmogorov's model of breakup and its application into Lagrangian computation of liquid sprays under air-blast atomization
,”
Phys. Fluids
15
(
1
),
184
192
(
2003
).
33.
S. V.
Apte
,
M.
Gorokhovski
, and
P.
Moin
, “
LES of atomizing spray with stochastic modeling of secondary breakup
,”
Int. J. Multiphase Flow
29
(
9
),
1503
1522
(
2003
).
34.
M. A.
Gorokhovski
and
V. L.
Saveliev
, “
Statistical universalities in fragmentation under scaling symmetry with a constant frequency of fragmentation
,”
J. Phys. D: Appl. Phys.
41
(
8
),
085405
(
2008
).
35.
H.
Luo
and
H. F.
Svendsen
, “
Theoretical model for drop and bubble breakup in turbulent dispersions
,”
AIChE J.
42
(
5
),
1225
1233
(
1996
).
36.
F.
Lehr
,
M.
Millies
, and
D.
Mewes
, “
Bubble-size distributions and flow fields in bubble columns
,”
AIChE J.
48
(
11
),
2426
2443
(
2002
).
37.
L.
Han
,
S.
Gong
,
Y.
Li
,
N.
Gao
,
J.
Fu
,
Z.
Liu
 et al, “
Influence of energy spectrum distribution on drop breakage in turbulent flows
,”
Chem. Eng. Sci.
117
,
55
70
(
2014
).
38.
L.
Han
,
S.
Gong
,
Y.
Ding
,
J.
Fu
,
N.
Gao
, and
H.
Luo
, “
Consideration of low viscous droplet breakage in the framework of the wide energy spectrum and the multiple fragments
,”
AIChE J.
61
(
7
),
2147
2168
(
2015
).
39.
J.
Solsvik
,
V. T.
Skjervold
,
L.
Han
,
H.
Luo
, and
H. A.
Jakobsen
, “
A theoretical study on drop breakup modeling in turbulent flows: The inertial subrange versus the entire spectrum of isotropic turbulence
,”
Chem. Eng. Sci.
149
,
249
265
(
2016
).
40.
D. G.
Obenauf
and
P. E.
Sojka
, “
Theoretical deformation modeling and drop size prediction in the multimode breakup regime
,”
Phys. Fluids
33
(
9
),
092113
(
2021
).
41.
R.
Bellman
and
R. H.
Pennington
, “
Effects of surface tension and viscosity on Taylor instability
,”
Q. Appl. Math.
12
(
2
),
151
162
(
1954
).
42.
G.
Bower
,
S. K.
Chang
,
M. L.
Corradini
,
M.
El-Beshbeeshy
,
J. K.
Martin
, and
J.
Krueger
, “
Physical mechanisms for atomization of a jet spray: A comparison of models and experiments
,”
Technical Report No. 881318
(
SAE Technical Paper
,
1988
).
43.
M.
Jalaal
and
K.
Mehravaran
, “
Fragmentation of falling liquid droplets in bag breakup mode
,”
Int. J. Multiphase Flow
47
,
115
132
(
2012
).
44.
D.
Stefanitsis
,
I.
Malgarinos
,
G.
Strotos
,
N.
Nikolopoulos
,
E.
Kakaras
, and
M.
Gavaises
, “
Numerical investigation of the aerodynamic breakup of droplets in tandem
,”
Int. J. Multiphase Flow
113
,
289
303
(
2019
).
45.
D.
Stefanitsis
,
G.
Strotos
,
N.
Nikolopoulos
, and
M.
Gavaises
, “
Numerical investigation of the aerodynamic breakup of a parallel moving droplet cluster
,”
Int. J. Multiphase Flow
121
,
103123
(
2019
).
46.
M.
Jain
,
R. S.
Prakash
,
G.
Tomar
, and
R. V.
Ravikrishna
, “
Secondary breakup of a drop at moderate Weber numbers
,”
Proc. R. Soc. A
471
(
2177
),
20140930
(
2015
).
47.
S. S.
Jain
,
N.
Tyagi
,
R. S.
Prakash
,
R. V.
Ravikrishna
, and
G.
Tomar
, “
Secondary breakup of drops at moderate Weber numbers: Effect of density ratio and Reynolds number
,”
Int. J. Multiphase Flow
117
,
25
41
(
2019
).
48.
D. R.
Guildenbecher
,
J.
Gao
,
J.
Chen
, and
P. E.
Sojka
, “
Characterization of drop aerodynamic fragmentation in the bag and sheet-thinning regimes by crossed-beam, two-view, digital in-line holography
,”
Int. J. Multiphase Flow
94
,
107
122
(
2017
).
49.
J. N.
Kapur
,
Maximum-Entropy Models in Science and Engineering
(
John Wiley & Sons
,
1989
).
50.
R. W.
Sellens
and
T. A.
Brzustowski
, “
A simplified prediction of droplet velocity distributions in a spray
,”
Combust. Flame
65
(
3
),
273
279
(
1986
).
51.
X.
Li
and
R. S.
Tankin
, “
Droplet size distribution: A derivation of a Nukiyama-Tanasawa type distribution function
,”
Combust. Sci. Technol.
56
(
1–3
),
65
76
(
1987
).
52.
R. W.
Sellens
, “
Prediction of the drop size and velocity distribution in a spray, based on the maximum entropy formalism
,”
Part. Part. Syst. Charact.
6
(
1–4
),
17
27
(
1989
).
53.
L. P.
Chin
,
P. G.
LaRose
,
R. S.
Tankin
,
T.
Jackson
,
J.
Stutrud
, and
G.
Switzer
, “
Droplet distributions from the breakup of a cylindrical liquid jet
,”
Phys. Fluids A
3
(
8
),
1897
1906
(
1991
).
54.
J.
Cousin
and
C.
Dumouchel
, “
Coupling of classical linear theory and maximum entropy formalism for prediction of drop size distribution in sprays: Application to pressure-swirl atomizers
,”
Atomization Sprays
6
(
5
),
601
(
1996
).
55.
V.
Semiao
,
P.
Andrade
, and
M.
da Graca Carvalho
, “
Spray characterization: Numerical prediction of Sauter mean diameter and droplet size distribution
,”
Fuel
75
(
15
),
1707
1714
(
1996
).
56.
D.
Ayres
,
M.
Caldas
,
V.
Semiao
, and
M.
da Graca Carvalho
, “
Prediction of the droplet size and velocity joint distribution for sprays
,”
Fuel
80
(
3
),
383
394
(
2001
).
57.
W. T.
Kim
,
S. K.
Mitra
,
X.
Li
,
L. A.
Prociw
, and
T. C. J.
Hu
, “
A predictive model for the initial droplet size and velocity distributions in sprays and comparison with experiments
,”
Part. Part. Syst. Charact.
20
(
2
),
135
149
(
2003
).
58.
X.
Li
and
M.
Li
, “
Droplet size distribution in sprays based on maximization of entropy generation
,”
Entropy
5
(
5
),
417
431
(
2003
).
59.
C.
Dumouchel
, “
A new formulation of the maximum entropy formalism to model liquid spray drop-size distribution
,”
Part. Part. Syst. Charact.
23
(
6
),
468
479
(
2006
).
60.
C.
Dumouchel
, “
The maximum entropy formalism and the prediction of liquid spray drop-size distribution
,”
Entropy
11
(
4
),
713
747
(
2009
).
61.
E.
Movahednejad
,
F.
Ommi
, and
S. M.
Hosseinalipour
, “
Prediction of droplet size and velocity distribution in droplet formation region of liquid spray
,”
Entropy
12
(
6
),
1484
1498
(
2010
).
62.
K.
Yan
,
Z.
Ning
,
M.
, and
C.
Sun
, “
Study on droplet size and velocity distributions of a pressure swirl atomizer based on the maximum entropy formalism
,”
Entropy
17
(
2
),
580
593
(
2015
).
63.
Q.-F.
Fu
,
J.-J.
Wang
, and
L.-J.
Yang
, “
Application of maximum entropy principle to predict droplet size distribution for swirl injectors
,”
Iranian J. Sci. Technol., Trans. Mech. Eng.
41
(
4
),
305
313
(
2017
).
64.
A.
Bodaghkhani
,
B.
Colbourne
, and
Y. S.
Muzychka
, “
Prediction of droplet size and velocity distribution for spray formation due to wave-body interactions
,”
Ocean Eng.
155
,
106
114
(
2018
).
65.
E.
Babinsky
and
P. E.
Sojka
, “
Modeling drop size distributions
,”
Prog. Energy Combust. Sci.
28
(
4
),
303
329
(
2002
).
66.
B. T.
Helenbrook
and
C. F.
Edwards
, “
Quasi-steady deformation and drag of uncontaminated liquid drops
,”
Int. J. Multiphase Flow
28
(
10
),
1631
1657
(
2002
).
67.
J. A.
Nicholls
and
A. A.
Ranger
, “
Aerodynamic shattering of liquid drops
,”
AIAA J.
7
(
2
),
285
290
(
1969
).
68.
W. R. A.
Goossens
, “
Review of the empirical correlations for the drag coefficient of rigid spheres
,”
Powder Technol.
352
,
350
359
(
2019
).
69.
C. E.
Shannon
, “
A mathematical theory of communication
,”
Bell Syst. Tech. J.
27
(
3
),
379
423
(
1948
).
70.
X.
Li
,
R. S.
Tankin
,
B.
Sztal
, and
J.-M.
Most
, “
Derivation of droplet size distribution in sprays by using information theory
,”
Combust. Sci. Technol.
60
(
4–6
),
345
357
(
1988
).
71.
P.
González-Tello
,
F.
Camacho
,
J. M.
Vicaria
, and
P. A.
González
, “
A modified Nukiyama–Tanasawa distribution function and a Rosin–Rammler model for the particle-size-distribution analysis
,”
Powder Technol.
186
(
3
),
278
281
(
2008
).
72.
H. E.
Wolfe
and
W. H.
Andersen
,
Kinetics, Mechanism, and Resultant Droplet Sizes of the Aerodynamic Breakup of Liquid Drops
(
Aerojet-General Corporation
,
1964
).
73.
X.
Li
,
L. P.
Chin
,
R. S.
Tankin
,
T.
Jackson
,
J.
Stutrud
, and
G.
Switzer
, “
Comparison between experiments and predictions based on maximum entropy for sprays from a pressure atomizer
,”
Combust. Flame
86
(
1–2
),
73
89
(
1991
).
74.
Z.
Dai
and
G. M.
Faeth
, “
Temporal properties of secondary drop breakup in the multimode breakup regime
,”
Int. J. Multiphase Flow
27
(
2
),
217
236
(
2001
).
75.
S.
Popinet
, “
Gerris: A tree-based adaptive solver for the incompressible Euler equations in complex geometries
,”
J. Comput. Phys.
190
(
2
),
572
600
(
2003
).
76.
R.
Bardia
,
Z.
Liang
,
P.
Keblinski
, and
M. F.
Trujillo
, “
Continuum and molecular-dynamics simulation of nanodroplet collisions
,”
Phys. Rev. E
93
(
5
),
053104
(
2016
).