You are on page 1of 44

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/228635807

Modeling Primary Atomization

Article  in  Annual Review of Fluid Mechanics · January 2008


DOI: 10.1146/annurev.fluid.40.111406.102200

CITATIONS READS
346 1,549

2 authors:

Mikhael Gorokhovski Marcus Herrmann


Ecole Centrale de Lyon Arizona State University
83 PUBLICATIONS   1,299 CITATIONS    102 PUBLICATIONS   2,006 CITATIONS   

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Mikhael Gorokhovski on 05 June 2014.

The user has requested enhancement of the downloaded file.


Modeling Primary Atomization 1

Modeling Primary Atomization

Mikhael Gorokhovski

LMFA UMR 5509 CNRS Ecole Centrale de Lyon, 36 avenue Guy de Collongue,

69131 Ecully Cedex, France ; Mikhael.Gorokhovski@ec-lyon.fr

Marcus Herrmann

Department of Mechanical Engineering, Stanford University, Bldg. 500,

ME/FPC, 488 Escondido Mall, Stanford, CA 94305-3030, USA;

Marcus.Herrmann@stanford.edu

Key Words spray, interface, DNS, breakup model, propulsion

Abstract This review concerns recent progress in primary atomization modeling. The numer-

ical approaches based on direct simulation are described first. While DNS offers the potential

to study the physical processes during primary atomization in detail, thereby supplementing

experimental diagnostics, it also introduces severe numerical challenges. These challenges and

the numerical methods to address them are outlined, highlighting some recent efforts in per-

forming detailed simulation of the primary atomization process. The second part is devoted to

phenomenological models of primary atomization. Since earlier conventional models of break-up

are well reported in available literature, only two recent developments have been highlighted:

(i) stochastic simulation of the liquid jet depletion in the framework of fragmentation under

scaling symmetry; (ii) primary atomization in terms of RANS mixing with a strong variation of

density.
.. 2000 1056-8700/97/0610-00

CONTENTS

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

Direct Numerical Simulation of Primary Atomization . . . . . . . . . . . . . . . . 6

Length and time scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

Describing the motion of the phase interface . . . . . . . . . . . . . . . . . . . . . . . . 8

Treating the discontinuity in material properties . . . . . . . . . . . . . . . . . . . . . . 10

Surface tension forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

Topology changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

Current state of the art . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

Phenomenological Approach in Primary Atomization Modeling . . . . . . . . . . . 18

Stochastic primary atomization under scaling symmetry . . . . . . . . . . . . . . . . . 18

Eulerian RANS approach and transport of the mean interface density . . . . . . . . . . 23

The future . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2
Modeling Primary Atomization 3

1 Introduction

Primary atomization represents an example of a complex gas-liquid flow: near

the nozzle, the liquid, initially introduced as a continuous jet, disintegrates into

filaments and drops by interaction with the gas. In addition to its academic inter-

est, understanding primary atomization is important in analyzing the combustion

process in propulsion and power related applications. Primary atomization is also

important in a wide variety of other applications (liquid metals, environmental

spraying, printing, food processing, etc.), as a significant parameter of the sub-

sequent droplet formation and dispersion. This has led researchers to focus on

theoretical and experimental analysis of this phenomenon.

When the liquid jet is injected into annular co-flowing gas at a high relative

velocity, the jet breaks up due to momentum transfer from the gas to the liquid.

This type of break-up is often referred to as air-blast atomization (19, 23, 27, 54).

A possible physical model was suggested in experimental studies by Hopfinger,

Villermaux, Cartellier and Lasheras, and by their PhD students, Rehab (74, 75),

Raynal (72), Marmottant (57, 58), Hong (42, 43) and Varga (97, 98). According

to these studies, the liquid jet break-up is caused by instabilities occurring at

the gas/liquid interface. After exiting the nozzle, the surface of the liquid jet

is subject to the Kelvin-Helmholtz instability. As this instability develops, the

most unstable wavelength, λKH , is determined by the thickness of the incoming

vorticity layer formed in the gas flow inside the nozzle (72), and by the liquid-to-

gas density ratio. These primary waves are exposed to the high-speed gas stream

and accelerated. This leads to the development of secondary interfacial instabil-

ities. It is shown in (42, 43, 97, 98) that the acceleration of primary waves gives

rise to a Rayleigh-Taylor instability with characteristic wavelength λRT , which


4 Modeling Primary Atomization

may be expressed in terms of inflow conditions. The Rayleigh-Taylor instability

manifests itself as a rapid transverse modulation which is followed by stripping

of drops.

When a high-speed liquid jet is discharged into stagnant dense air, less of

a consensus has been reached in identifying the dominant mechanism of initial

break-up: many interdependent phenomena may instigate this process. The ex-

perimental study showed that amongst the important phenomena are turbulence

(26, 87, 105, 106) and cavitation pockets (4, 5, 15, 88) within injection holes, drop

shedding (81), unsteadiness of the injection velocity (22).

It has been established that a large number of dimensionless scaling parame-

ters are involved in primary atomization. In air-blast atomizers, these include the
ρg (ug − ul )2
Weber number W e = ; the non-dimensional thickness of the incom-
σ
δg λRT
ing boundary layer from inside the nozzle ; the ratio characterizing
Dg − Dl λKH
the spectrum of drops produced near the injector; the gas-to-liquid density, mass
ρg ug (Dg2 − Dl2 ) ρg u2g
and dynamic pressure ratios ρl /ρg , m = and M = ; the
ρl ul Dl2 ρl u2l
µl
Ohnesorge number Oh = √ , measuring the relative importance of interfa-
ρl σDl
cial viscous stress and surface tension. For a high-speed liquid jet atomized in

still air, the list of relevant parameters also includes the level of turbulence and

ltur ktur 1/2ρl u2l
the cavitation number in the injection hole: , and CN = ,
Dl ul pg − pvap
where pvap is the saturation vapor pressure at the given temperature. Evaluating

the influence of all these parameters in different regimes of break-up, and in re-

alistic operating conditions (high velocities and pressures; small length-and-time

scales, strongly non-homogeneous 3D gas-liquid flow) is a difficult task for ex-

perimentalists. The numerical modeling of primary atomization may provide a

way of overcoming this problem and this review concerns recent progress in such
Modeling Primary Atomization 5

modeling.

The first category of numerical approaches concerns direct simulation of pri-

mary atomization, which is the subject of Section 2. These approaches are based

on integration of the Navier-Stokes equations, identifying the gas/liquid interface

at each time step. Using this technique, liquid fragments break apart when the

progressively stretched filament becomes smaller than the typical size of a numer-

ical cell. Such approaches provide accurate local estimates of the gas/liquid mix-

ture close to the injector, and may help in understanding the physics of primary

atomization. However, when the Weber number is very large, the computational

expense associated with the resolution of all length scales is very high. When the

relative velocity between liquid and gas is more than, say, a hundred meters per

second, droplets with diameters of a dozen or so microns may be produced, which

is usually smaller than the resolution of the numerical grid. As a result, alongside

direct numerical simulation, there is a need to develop phenomenological models,

with more modest CPU requirements. The objective of conventional models is

to represent statistically the essential features of the initial break-up. Although

these models contain the mechanisms of the initial break-up of the liquid jet (sur-

face instabilities in Reitz’s et. al. models (6, 36, 69, 76) and drop shedding (107);

”spontaneous” break-up in Tanner’s model (92), jet turbulence in Huh and Gos-

man’s model (44), cavitation in Arcoumanis and Gavaises’s model (2,3,51)), they

use round ”blobs” injected from the nozzle exit and hence do not really describe

the primary atomization process. To compensate for this to some extent, Sec-

tion 3 is devoted to two recent developments in phenomenological modelling of

primary atomization.
6 Modeling Primary Atomization

2 Direct Numerical Simulation of Primary Atomization

The detailed physical mechanisms by which primary atomization occurs are as

of this date not well understood. Thus direct numerical simulation (DNS) of the

primary atomization process can serve as a valuable tool to study the involved

mechanisms in detail. In DNS, the goal is to resolve all time and length scales

solving the governing equations directly. Since most primary atomization applica-

tions occur at low Mach numbers and the two fluids involved are immiscible, the

flow is governed by the unsteady Navier-Stokes equations in the variable density,

incompressible limit,

∂ρu   
+ u · ∇ρu = −∇p + ∇ · µ ∇u + ∇T u + ρg + T σ (1)
∂t
∇ · u = 0. (2)

Here, u is the velocity, ρ the density, p the pressure, µ the dynamic viscosity, g

the gravitational acceleration, and T σ the surface tension force which is non-zero

only at the location of the phase interface xf .

Although a DNS approach reduces epistemic uncertainty, the numerical re-

quirements are extremely challenging:

• length and time scales vary over several orders of magnitude,

• the phase interface constitutes a discontinuity in material properties,

• surface tension forces are singular forces active only at the phase interface,

• turbulence and the turbulent primary atomization process are inherently

three dimensional,

• topology changes of the phase interface occur frequently.

While the second and third properties are typical for most flows involving liq-
Modeling Primary Atomization 7

uid/gas interfaces, the others are specifically significant in primary atomization

DNS.

The following sub-sections will address the numerical requirements outlined

above and discuss state of the art numerical solution strategies for to address

them.

2.1 Length and time scales

In a DNS, the goal is to reduce epsitemic uncertainty by resolving all necessary

time and length scales inherent in the flow and thus eliminating the need to model

the effect of any unresolved scales. In a single phase turbulent flow, the smallest

length scale that must be resolved is the Kolmogorov length scale η. Multiphase

flows add an additional smallest length scale that requires resolution, the size

of the smallest liquid structure ζ. Unfortunately, when topology change of the

phase interface is involved, that length scale approaches zero, since the process

of pinching involves at the very instance of breakup, a zero sized connecting

ligament. Such resolution requirements together with the break down of the

continuum assumptions inherent in the Navier-Stokes equations makes simulation

of the details of the topology change process unfeasible in the context of DNS of

the turbulent primary breakup process. Thus, modeling has to be introduced to

address topology change. This will be discussed in a later section.

Using adequate topology change models, the smallest liquid structure ζ that

then should be resolved is of the order of the smallest drop generated by the

initial breakup. Depending on the interface tracking/capturing method used,

this implies a grid resolution of about 2-5 grid cells per ζ.

As an example, consider the primary breakup of a Diesel jet. There, smallest


8 Modeling Primary Atomization

drop diameters near the nozzle of the order of 1 µm have been observed (104)

with the Kolmogorov length scale being of the same order of magnitude. This

implies an about one order of magnitude finer resolution requirement for the

primary breakup DNS as compared to a single phase DNS. Fortunately, such

high resolution is required only at and near the phase interface. Away from it,

the usual single phase DNS resolution requirements apply.

2.2 Describing the motion of the phase interface

To perform a DNS of the primary atomization process, the location and motion

of the phase interface has to be described with high accuracy. Neglecting phase

transition effects, the phase interface constitutes a material surface whose motion

is described by
dxf
= v(xf ) (3)
dt

where the subscript f refers to a point on the phase interface. The straightforward

way of describing the motion of the phase interface is to solve Eq. (3) for a

collection of marker particles placed onto the phase interface. This is the so-called

interface tracking approach. While flow solver grid nodes can be used as marker

particles, thereby ensuring phase interface conforming flow solver grids, such an

approach would require almost constant regridding or rezoning (84) due to the

complex flow field at the phase interface during primary atomization. Instead, in

the surface marker method, the marker points are connected to form a separate

surface grid that is advected according to Eq. (3) in a Lagrangian way inside

a fixed flow solver grid (93, 94). Although not inherently mass conserving, this

approach nonetheless shows excellent mass conservation properties. It’s drawback

is that topology changes are not handled automatically, see Sec. 2.5. It has thus
Modeling Primary Atomization 9

not been applied to the DNS of primary breakup, yet.

An alternative to interface tracking methods are interface capturing methods.

Here, the phase interface is imbedded into a fixed grid via the use of an additional

scalar. Two principle methods have emerged in this class: the Volume of Fluid

method (40,62,84) and the level set method (65,66,85) . In the former method, a

marker function ψ is defined as the liquid volume fraction in each computational

grid cell,
1
Z
ψ= H(x − xf )dx (4)
V V

with H the Heaviside function, resulting in a simple advection equation describing

the motion of the phase interface,

∂ψ
+ v · ∇ψ = 0 . (5)
∂t

One of the major advantages of the Volume of Fluid method making it a popular

interface tracking method is that it can be constructed to be inherently liquid

volume conserving. However, due to the discontinuous nature of ψ at the phase

interface, see Eq. (4), special geometric algorithms have to be employed to solve

Eq. (5) to avoid unacceptable numerical diffusion of ψ (35, 77).

In level set methods, a scalar φ is defined to be equal to some constant value

φ0 at the location of the phase interface, and φ > φ0 in fluid 1 and φ < φ0 in

fluid 2. From this definition and Eq. (3) the level set equation follows,

∂φ
+ v · ∇φ = 0 . (6)
∂t

Since the above definition of the level set scalar defines φ only at the location of

the phase interface, there is significant latitude in defining φ away from it. For

numerical reasons, one would like φ to be a smooth function, thus one of the

most popular definition is that of a signed distance function, i.e. |∇φ| = 1, see
10 Modeling Primary Atomization

for example (89). The major drawback of the level set method is that it does not

inherently preserve liquid volume. Since volume errors are proportional in size to

the employed grid resolution, grid refinement strategies, like structured adaptive

mesh refinement (63) or the Refined Level Set Grid Method (RLSG) (39) can be

employed to limit their impact. An alternative to grid refinement is to augment

and correct the level set scalar by an additional numerical scheme, for example

the CLSVOF or MCLS methods coupling level set and volume of fluid (71,90) or

the particle level set method, coupling marker particles and level sets (25).

Defining φ as the signed distance, is, as indicated above, not a requirement.

Instead, in the so-called conservative level set method (64), φ0 = 0.5 and φ away

from the interface is defined to be a smeared out version of the liquid volume

fraction ψ, i.e.
1 d
   
φ= tanh +1 , (7)
2 2

where d is the minimum distance to the phase interface . Solving Eq. (6) in con-

servative form in conjunction with re-initializing φ according to Eq. (7) does show

good volume conservation properties (64) even in complex interface geometries

(20).

So far, in DNS of the primary atomization process, the VOF (10,100), CLSVOF

(60), conservative level set (20), and RLSG method (46) have all been successfully

employed.

2.3 Treating the discontinuity in material properties

In most technical applications involving primary atomization, the smallest scales

that need to be resolved in a DNS, i.e. the Kolmogorov scale η and the size

of the smallest drops ζ are still orders of magnitude larger than the thickness
Modeling Primary Atomization 11

of the phase interface itself. Thus, on the resolved scales, the phase interface

constitutes a discontinuity in material properties. Under the common assumption

that material properties α, like density and viscosity, are constant in each of the

fluids, their respective values at any point x in the computational domain are

defined as

α(x) = α2 + H(x − xf )(α1 − α2 ) , (8)

where indices 1 and 2 denote values in fluid 1, respectively 2, and H is the

Heaviside function.

Two different numerical approaches are commonly used to deal with the jump

in material properties. In the first, geared towards finite volume schemes, the

discrete control volume value αcv of any quantity α(x) is defined as the control

volume average,

1
Z
αcv = α(x)dx = α2 + ψ(α1 − α2 ) , (9)
Vcv Vcv

where the last part follows from Eqs. (8) and (4). This approach is thus ideally

suited for Volume of Fluid methods and the conservative level set method, since

the volume fraction ψ is readily available. Note, that Eq. (9) does treat the phase

interface as a discontinuity in the finite volume sense.

For finite difference methods, Eq. (8) directly holds at the node location xi ,

however, some amount of artificial smearing is typically introduced, resulting in

αi = α2 + ψ(α1 − α2 ) (10)

for VoF methods and

αi = α2 + H (φ − φ0 ) (α1 − α2 ) (11)

for level set methods, where H is a smeared out version of the Heaviside function.

An alternative approach that avoids smearing of the phase interface discontinuity


12 Modeling Primary Atomization

is the ghost fluid approach (41, 45) applicable to finite difference methods. Here,

although formally still in a single fluid formulation, the two phases are effectively

solved separately and coupled by imposing the appropriate jump conditions at

the phase interface. The method relies on knowing these jump conditions and

being able to evaluate xf in Eq. (8) within each computational cell with good

accuracy. It is thus ideally suited for level set methods.

2.4 Surface tension forces

Even though Weber numbers of technically relevant turbulent primary atomiza-

tion processes calculated using integral scales are typically very large, seemingly

indicating that surface tension forces are unimportant, primary atomization can

occur on scales significantly smaller than the integral scale, resulting in relatively

small local Weber numbers. In fact, there is experimental evidence (58) that the

actual primary breakup process is governed by capillary forces determining the

resulting drop size distribution. Hence the accurate treatment of surface tension

forces is crucial for DNS. On scales relevant to DNS, surface tension forces T σ

are singular forces active only at the location of the phase interface,

T σ = σκnf δ(x − xf ) = −∇ · ([1 − n ⊗ n] σδ(x − xf )) , (12)

where κ is the mean curvature of the phase interface, nf is its normal vector,

and δ is the delta function concentrated on the phase interface. The first part

of Eq. (12) forms the basis of the so called Continuum Surface Force (CSF)

method (14), whereas the last part is the basis for the Continuum Surface Stress

(CSS) method (55, 84). Depending on the chosen interface capturing method,

two different approaches are common to discretize the delta function in Eq. (12).
Modeling Primary Atomization 13

For VoF methods, interpreting ψ as a color function (14),

nf δ(x − xf ) ≈ ∇ψ , (13)

whereas for level set methods, a regularized, spread out version of the delta

function is commonly used (24, 70). Both effectively spread the surface tension

force into a small neighborhood normal to the phase interface. The alternative

approach is to use the ghost fluid method (41,45). Here, the surface tension force

is acting through the pressure jump condition at the phase interface,

[[p]] = σκ + 2[[µ]]nT · ∇u · n , (14)

and thus does not require a discretization of the delta function. As is obvious

from Eq. (14), surface tension forces result in a change of pressure across the

phase interface. In the CSF and CSS method it is therefore crucial, that on a

discrete level, the pressure gradient across the phase interface, Eq. (1) and the

surface tension force balance exactly. While this is guaranteed in VoF schemes

on staggered grids, this is not the case in collocated formulations or level set

formulations using spread out delta functions, resulting in significant so-called

spurious currents. These erroneous velocities limit the range of capillary/Laplace

numbers that can be simulated in an accurate and stable manner (84). For

collocated schemes, Francois et al. (29) recently proposed a so-called balanced

force algorithm for VoF methods on Cartesian grids that ensures the required

discrete balance, while Herrmann (39) presented a corresponding scheme for level

set methods on structured and unstructured grids.

Using either the ghost fluid method or the CSF/CSS method in the discrete

balance form is however not sufficient. The phase interface curvature has to

be calculated with good accuracy as well. This is especially problematic for VoF
14 Modeling Primary Atomization

method, since ψ is discontinuous at the phase interface, thus requiring convolution

operations (103), a height function approach (37, 91), or higher order interface

reconstruction methods like PROST (77). For level set methods, discretization

of the curvature at node locations κ = ∇ · (∇φ/|∇φ|) (66) is at best first order

accurate requiring surface based evaluation of curvature (39) for higher accuracy.

2.5 Topology changes

As discussed in Sec. 2.1, resolution of the actual instance of topology change

cannot be resolved in a DNS of turbulent primary atomization. Pinching models

are thus required. The choice of model used, depends on the interface track-

ing/capturing method. In interface capturing methods, topology change is ini-

tiated automatically, once two interface segments enter into the same computa-

tional grid cell. These methods thus contain an inherent breakup length scale of

the order of the local grid size, making the topology change process grid depen-

dent, a clearly undesired feature. However, it can be argued that while in most

primary atomization situations the exact moment of topology change might not

be predicted accurately, nevertheless the error in volume of the broken off liquid

structure is only of the order of the volume of the local grid cell, thus not signifi-

cantly affecting drop sizes larger than several local grid cell volumes. Still, due to

the inherent grid dependency, it is incumbent on an interface capturing DNS to

demonstrate grid independence. It should be noted that interface tracking meth-

ods based on marker particles never result in topology changes unless they are

”manually” introduced. While this intervention is a cumbersome and at times

costly procedure (93), these methods nonetheless offer the potential to introduce

a physics based breakup length scale, independent of the local grid size. Some
Modeling Primary Atomization 15

similar efforts have been presented in (38) for the RLSG method.

2.6 Current state of the art

Direct numerical simulation of the turbulent primary atomization process is still

in its infancy due to the associated high computational cost and numerical chal-

lenges. Nonetheless a number of studies have shown the potential for analyzing

the atomization process using numerical methods. Most studies are not DNS

in the sense outline above. They employ a single phase LES formulation in the

single phase regions of the flow and extend that formalism to regions containing

the phase interface. They thus neglect all LES subgrid terms that arise from the

presence of the phase interface and the associated jump in material properties and

surface tension forces. As such, they can be called LES in name only and should

rather be viewed as quasi-DNS, or under-resolved DNS of the phase interface

combined with LES of the single phase regions of the flow (9, 10). Nonetheless,

quasi-DNS/LES simulations can yield valuable insight into the primary atomiza-

tion process and the resulting drop size distribution, if it can be shown that the

small scale non-resolved phase interface dynamics will not influence the larger

scale, resolved phase interface dynamics. A measure for this pre-requisite is the

grid-convergence of the resolved scale drop size distribution: refining the grid

should not alter the resulting resolved scale drop size distribution.

Using the quasi-DNS/LES approach, de Villiers et al. (100) performed rela-

tively coarse simulations of the Diesel jet breakup under typical operating condi-

tions. They performed a one level grid convergence study and found that although

their resulting drop size distributions are of χ2 type, for different grid sizes, these

distributions are significantly different, even for larger drop sizes. This indicates
16 Modeling Primary Atomization

that for their employed grid resolution, the quasi-DNS/LES assumption is ques-

tionable.

Bianchi et al. (9, 10) reported on similar Diesel jet atomization simulations

using a slightly higher grid resolution. They studied the influence of the turbu-

lence in the injector pipe and found that this has a strong impact on both intact

core length and atomized mass, but only a weak impact on the resulting drop

size distribution. However, grid dependency results were not reported leaving the

question of applicability of the quasi-DNS/LES approach unanswered.

While the former two studies use VoF interface capturing and are in the atom-

ization regime, Menard et al. (60) employ the CLSVOF method in conjunction

with a ghost fluid method to study the Diesel jet atomization in the second wind

induced regime. Instead of simulating part of the injection system, they prescribe

turbulent fluctuations at the nozzle exit with a procedure due to Klein et al. (47).

They do not apply any LES subgrid models in the single phase regions and thus

aim to perform a DNS. However, their grid resolution is too coarse for a true DNS

with ∆x/η ≈ 3 and more importantly ζ limited to drops with Weber numbers

larger than 10 resolved by a single grid point. Still, their grid sizes are a factor 2

finer than that of Bianchi et al. (10). Unfortunately no grid convergence results

or resulting drop size distributions are reported.

Finally, Desjardins et al. (20) report promising results for Diesel jet atom-

ization using a conservative level set approach in conjunction with a ghost fluid

method. Figure 1 shows two snapshots of the phase interface at different times

during the injection. In the upper frame, most of the droplets have been gener-

ated by shedding of the mushroom shape at the tip of the jet. However, in the

lower frame, generation and break-up of ligaments occur in many different places
Modeling Primary Atomization 17

throughout the whole length of the jet. A zoom on the middle part of the jet is

presented in Fig. 2. The surface topology appears extremely complex, composed

of stretched out ligaments and droplets. A clear liquid core can be seen behind

these structures. Many different droplet sizes and shapes are obtained, ranging

from the smallest possible resolvable droplets to large liquid blobs more than

ten computational cells wide. However, as with most other DNS simulations, no

quantitative comparison to experimental data and no grid convergence study was

performed.

For co-axial atomizing configurations, Kim et al. (46) used a RLSG method in

conjunction with a balanced force algorithm to perform near DNS simulations of

a low density ratio liquid jet. They report qualitative agreement with the postu-

lated subsequent physical processes involved in atomization, determined exper-

imentally by Marmottant and Villermaux (58), see Fig. 3, i.e. initial Rayleigh

limit Kelvin-Helmholtz instability, followed by a Rayleigh-Taylor instability re-

sulting in ligament formation that are subsequently stretched by the surrounding

fast moving gas stream resulting in pinch off followed by a capillary breakup of

the detached ligament into drops of varying sizes. In order to limit the computa-

tional cost, broken-off small near spherical drops are removed from the interface

capturing representation and added as drops into a Lagrangian spray model pre-

serving their volume, center of mass, and momentum. They performed a limited

grid convergence study observing indications for grid converged drop size distri-

butions for drops that are resolved by at least two grid points (46).

In addition to the discussed studies, a number of two-dimensional studies of

primary atomization have been performed (13, 68, 108, 109). It must be stressed,

however, that these cannot be considered DNS of turbulent primary atomization,


18 Modeling Primary Atomization

due to the inherent three dimensional nature of the break up process. They

do, however, represent a class of models for the atomization process, with their

inherent, albeit questionable, modeling assumption that initial breakup occurs

in the form of torus detachment from the liquid core. However, none of the

quasi-DNS simulations discussed above, seem to support this hypothesis.

Finally, the numerical studies of Klein (48) and Pan and Suga (67) should be

mentioned. The former performs three dimensional DNS of turbulent film phase

interfaces (48), whereas the latter simulates laminar liquid jets in the second wind

induced regime (67). Neither study, however, involves breakup and thus avoids

a number of critical issues associated with atomization.

3 Phenomenological Approach in Primary Atomization Model-

ing

3.1 Stochastic primary atomization under scaling symmetry

This approach, referred to as the stochastic approach, was initially applied to the

simulation of secondary atomization (1, 30, 31, 78, 80, 101) and has been further

developed for primary air-blast atomization in (32, 33). It combines LES in the

turbulent gas flow with stochastic simulation of the liquid jet depletion: the 3D

structure of the liquid core with surrounding drops that are pinched off from

the core, are simulated stochastically. The frequency of break-up is assumed to

be high enough in order to consider the liquid jet depletion as a fragmentation

cascade with scaling symmetry. The λRT -to-λKH ratio, and the kinetic energy

transfer from the gas to the liquid are parameters of the stochastic process.

Statistical universalities. Kolmogorov (49) considered the break-up of solid

carbon particles as a random discrete process. He assumed the probability of


Modeling Primary Atomization 19

breaking each parent particle into a given number of parts is independent of the

size of the parent particle. In the context of the central limit theorem he predicted

that, after a large number of break-up events, such a discrete process should

yield a log-normal distribution of particle sizes. Using a scaling formulation,

Kolmogorov’s scenario states that each break-up event reduces the typical size of

a fragment, r ⇒ αr (rM +1 = rM eln α ), by an independent random multiplier α,


R1
which is governed by the fragmentation spectrum q(α), q(α)dα = 1.
0

An alternative to Kolmogorov’s discrete stochastic process is provided by the

mathematical models described in (17, 18, 28, 59, 73, 111, 112), which are based

on the analytical solution of the fragmentation equation, without appealing to

the central limit theorem. In these models, the particle-size distribution evolves

continuously with increasing time, instead of the growing number of breakage

events in discrete models.

In the case of constant fragmentation frequency ν, the kinetic fragmentation

equation for the normalized distribution function f (r, t) has the following form:

Z1
1 ∂f (r) 1 r
 
= f q(α)dα − f (r) (15)
ν ∂t α α
0

Z1
1 r
 
The steady-state solution of this equation, (f (r) = f q(α)dα), is the
α α
0
delta function

f (r) = δ(r) (16)

and the following question can be asked: How does the distribution f (r, t) evolve

towards the ultimate steady-state solution (16)? Since the solution of the frag-

mentation equation (15) requires knowledge of q(α), which is, in principal, un-

known, this question cannot be completely answered. However, due to scaling

symmetry, the evolution of f (r, t) goes through at least two intermediate asymp-
20 Modeling Primary Atomization

totic distributions. Evaluating these distributions does not require the knowledge

of the entire spectrum q(α) - only its first two logarithmic moments, and, at large

times, only the ratio of these moments determine the behavior of the solution of

Eq.(15).

These two universal forms are given in (31) as follows. The asymptotic solution

of Eq.(15) can be derived as:


 
1 hln αi2
rf (r, t) = r D E exp − D 2 E νt ×
 
2
2π ln α νt 2 ln α
 2  hln αi
ln Rr r
 
hln2 αi
× exp − D E  (17)
2 ln2 α νt R

D E R1
where R = exphln ri0 denotes the initial length scale and lnl α = lnl αq(α)dα.
0

This expression confirms the main result of Kolmogorov 1941 concerning the log-

normal asymptotic shape of the size distribution resulting from fragmentation

under scaling symmetry. Increasing time still further, it is seen that the ”log-

normal” multiplier in Eq.(17) tends to unity and hence the long-time particle-

size distribution is determined by a power law with just one universal parameter
D E
ln2 α / hln αi:
hln αi
r
 
hln2 αi
rf (r, t → ∞) ∝ (18)
R

The power law (18) implies that size of particle has a fractal distribution. The

fractal geometry of an atomizing liquid jet has been observed photographically

in many experiments (see, for example, (16, 57, 86, 110)). The identification of
D E
ln2 α / hln αi may be seen from the solutions for the logarithmic moments of
D E D E
the continuous model (1) (34): hln ri ≈ νt hln αi; (ln r − hln ri)2 ≈ νt ln2 α ,
D E
where the mean quantity is denoted by angled brackets. Hence the ratio (ln r − hln ri)2 / hln
D E
has the time-independent constant value ln2 α / hln αi, regardless of dispersion
Modeling Primary Atomization 21
D E
(ln r − hln ri)2 :

D E D E
(ln r − hln ri)2 / hln ri = ln2 α / hln αi (19)

The statistical universality of fragmentation also implies that at large times

(high fragmentation frequency ν), the fragmentation equation reduces to the

Fokker-Planck equation. Indeed, the kernel in the fragmentation equation (15)


∞ n
1 r 1 ∂
  X 
may be expanded as f = (−1)n r f (r) lnn α (Saveliev et. al.
α α n=0
n! ∂r
D E
(82,83)), leading to an infinite number of terms containing the moments lnk α in

the fragmentation equation. However the asymptotic distribution (17), depends


D E
solely on the first two logarithmic moments, hln αi and ln2 α . This means that,
D E
although each term with lnk α , k > 2, may be significant, setting them to zero

will not change the long time asymptotics of the solution. The consequence is

that at high fragmentation frequency ν, Eq.(15) takes exactly the following form:

1 ∂f (r) ∂rf (r) 1 D 2 E ∂ ∂rf (r)


= − hln αi + ln α r (20)
ν ∂t ∂r 2 ∂r ∂r

The Fokker-Planck equation Eq.(20) describes the log-Brownian stochastic pro-

cess. The corresponding stochastic equation is


r D E
ṙ = ν hln αi r + ν ln2 α /2rΓ(t) (21)

where Γ(t) is the Langevin process: hΓ(t)i = 0; hΓ(t)Γ(t0 )i = 2δ(t − t0 ).

Stochastic model of liquid jet depletion. In the technique (32, 33, 49),

the random interface of the liquid core is simulated using the spatial trajectories

of constituent stochastic particles. The one-point distribution, f (x, t; r), defined

such that f (x, t; r)dr is the probability that the radial location of the liquid core

interface, r, at axial position x and time t lies in the element dr about r, is given

by f (x, t; r) = hδ(r(x, t) − rSP )i, where rSP is the position of a stochastic particle
22 Modeling Primary Atomization

and δ(r(x, t) − rSP ) denotes the Dirac function. The main steps in the modeling

are:

(i) stochastic particles are injected one after another, thus yielding the random

geometry of the liquid core; each particle trajectory ends after a length of time

determined by kinetic energy transfer from the gas flow to the liquid.

(ii) the radial motion of a stochastic particle is governed by Eq.(21), where the
λRT
D E  
fragmentation parameter (19) is determined from ln2 α / hln αi = ln .
λKH
The axial particle velocity is taken equal to the convection velocity of liquid
√ √
ρg u g + ρ l u l
surface waves (21), ul = √ √ . The path of each stochastic particle
ρg + ρl
lies in a plane with random azimuth.

(iii) the statistics of the core surface are used to obtain the size and position

characteristics of the drops surrounding the core, sampled with the total mass

rate equal to inflowing liquid mass rate.

(iv) LES is used to simulate the gas flow and is conditioned by the presence

of the liquid core. The blobs formed around the core are dragged along by such

a conditioned flow of gas and subject to secondary atomization and coalescence,

both modeled by an inter-droplet collision mechanism (101).

Figure 4 shows comparisons with measurements of Werquin (102) for two liquid

injection velocities of the near injector distribution of liquid fraction in a round

liquid jet atomized by parallel co-flow of air. It is seen that the simulation re-

produces qualitatively the experimental observation. Figures 5 shows simulation

results for an atomizing liquid jet injected at different velocities into high-speed

(140 m/s) co-flowing gas. Here, the inlet conditions correspond to those of the

experimental study by Lasheras et al. (53). Evidently, this approach cannot

reproduce the real filamentary structure of an atomizing spray. However, the


Modeling Primary Atomization 23

results of such statistical simulation agree with physical intuition: increasing the
ρg u2g
parameter M = (99) causes the liquid core to shorten and the diameter of
ρl u2l
drops around the core to decrease. What is more, the predicted far-field of the

spray has length scales in reasonable agreement with measurements.

3.2 Eulerian RANS approach and transport of the mean inter-

face density

This approach was proposed by Vallet, Burluka and Borghi in (95,96) and applied

in simulation of primary atomization of a Diesel spray (7,8,12) and air-blast atom-

izers (52, 61). The approach is based on homogeneous formulation of two-phase

medium: in each unit volume, two random motions may exist simultaneously;

the liquid and the gas move through each other independently. Each motion

is connected with its own ”effective” mass: motion of the liquid is connected

with ρ∗l = αρl , where α is the instantaneous liquid volume fraction and ρl is

the constant density of the liquide, whereas motion of the gas is connected with

ρ∗g = (1 − α)ρg , where ρg is the constant gas density. The total true density of the

gas-liquid mixture is the sum ρ = αρl + (1 − α)ρg . Introducing the local liquid

mass fraction Yl , one writes

1 Yl 1 − Yl
= + (22)
ρ ρl ρg

For such a gas-liquid mixture, the bimodal mass-weighted pdf (11) of Yl -distribution

may be defined: Pe (Yl ) = γδ(Yl − 1) + (1 − γ)δ(Yl ), where γ = Yel is the Favre-

averaged liquid mass fraction. With this pdf and Eq.(22), the time-averaged mass

fraction, Y l , is equal to the time-averaged liquid volume fraction: Y l = α, and,


ρg
also, ρYel = ρl α, ρ = . The sum of two motions is governed
1 − Yel (1 − ρg /ρl )
by the mixture conservation law of the density ρ and of the mass flow density
24 Modeling Primary Atomization

ρui = ρu
ei (u
ei is the local mass-averaged velocity). In mass weighting of the last

one, the contribution of liquid is significant; thereby the introduced here ”two-
1
phase turbulent kinetic energy” ρke = ρ(ui − u
ei )(ui − u
ei ) contains the velocity
2
slip between liquid and gaseous phases.

This approach considers the primary atomization in the framework of turbulent

RANS (11) mixing of a jet with high density (assigned to the liquid) spreading

in the gaseous environment. Similar to classical assumption in turbulence, it is

assumed that the turbulent Reynolds number, along with the turbulent Weber

number, are high; thereby the turbulent dispersion of the jet is independent

of those numbers. At small scales, following Kolmogorov (50), the formation

of droplets of maximum stable size is linked to the surface tension σ, the gas

viscosity µg and also, to the turbulent stretching.

Large-scale mixing. The turbulent dispersion of liquid jet is computed from

standard transport equation for ρYel . However the diffusion term in this equa-

00 Y 00 , is coupled with the local slip velocity


tion, containing the turbulent flux ρug
i l

00 Y 00 = ρY
ρug el (ul, i − u
ei ) (ul, is the time-mean velocity of the liquid fluid parti-
i l i

cle). Hence its gradient closure is not completely appropriated. Additionally to


∂p
this, due to high density variation, the pressure-gradient drift fluxes u00j and
∂xi
∂p
Yl00 00 Y 00 , ρug
contribute significantly to the turbulent transport ρug
i l
00 u00 and ρk.
i j
e
∂xi
This motivated authors to use the technique of algebraic stress modelling (79).

With the bimodal shape of PDF, the pressure driven fluxes may be obtained in

the following form:


! !
  1 1 1 1
Yl00 = ρYel 1 − Yel − 00 Y 00
; u00 j = ρug
i l − (23)
ρ l ρg ρl ρg
 
(last expression in Eqs.(23) is probably based on expansion ui (Yl ) ≈ a+b Yl − Yel .)

Transport equation of the mean interface density. The new Eulerian


Modeling Primary Atomization 25

variable Σ, a mean amount of interfacial surface per unit volume, is introduced

in this approach. The knowledge of local Σ allows to estimate the mean radius of

round droplets r = 3ρYel /ρl Σ, which are available for advection in a Lagrangian

way.

Similar to the Marble and Broadwell’s flame surface density in turbulent com-

bustion (56), the transport equation of Σ is presumed with terms of convection,

diffusion, production (surface stretching) and destruction (coalescence):


!
∂Σ ∂ u
ej Σ ∂ ∂Σ 1 1 2
+ = DΣ + Σ− Σ (24)
∂t ∂xj ∂xj ∂xj τprod τdestr

In this equation, the production of interface density is addressed to the turbulent

stretching; expressions of turbulent time scales are used. In equilibrium, when

turbulent stretching is balanced by capillary forces, i.e. the critical Weber number
1 1
is of order of unity, it follows from Eq.(24) that = which may be
τdestr τprod Σeq
1 ρl req
expressed in terms of maximum stable radius req : = . Different
τdestr 3τprod ρYel
expressions of req are available from (30) and (96). In RANS computation of

the rocket combustion (52), Jay, Lacas and Candel modify Eq.(24) taking into

account for the Kelvin-Helmholtz instability effect.

The Eulerian mixing model of primary atomization naturally lends itself to the

RANS codes which are in widespread use. In the case of a Diesel spray atomiza-

tion, Figure 6 (8) shows that this model matches well the DNS presented in (60).

At the same time, the model is affected by the drawbacks of the RANS approach:

the presumed form of Eq.(24) is supposed to be universal, independently of the

specific geometry of the flow and its boundary conditions; the transport of Σ,

which is modelled in (24) by diffusion-like hypothesis, and thereby, neglects the

spatial grouping of liquid elements, is of concern.


26 Modeling Primary Atomization

4 The future

The advent of ever more powerful computers will enable the simulation of the

primary atomization process with ever higher resolution, enabling simulations in

the near future employing at least adequate resolution. This puts simulating the

atomization process as a whole into reach, if the primary atomization simulation

is coupled to a secondary atomization model. Developments along these lines

have been reported in (38, 46) transferring broken off, near spherical small liq-

uid structures into a Lagrangian particle/parcel model that describes secondary

atomization.

However, due to fractal nature of primary atomization in a high Weber and

Reynolds number flow, the problem of full resolution of liquid scales will persist

even with very powerful computers. Introducing similarities of break-up at small

scales and discovering the renormalized formulation of Navier-Stokes equations,

which will describe interacting gas-liquid flow ”effectively”, similar to LES equa-

tions of turbulent flow, is the future of modelling primary atomization. Another

problem which motivates the future work stems from recent developments in a

high-pressure propulsion systems, when primary atomization takes place in super-

critical conditions. In these conditions, the surface tension and laminar viscosity

are negligible and the concept of interfacial surface has no more sense. Describing

primary atomization in such conditions is an open problem for researchers.

Literature Cited

1. S. Apte, M. Gorokhovski, and P. Moin. LES of Atomizing Spray With

Stochastic Modeling of Secondary Break-up. Int. J. of Multiphase Flow,

29:1503–1522, 2003.
Modeling Primary Atomization 27

2. C. Arcoumanis, M. Gavaises, and B. French. Effect of fuel injection processes

on the structure of diesel sprays. SAE paper 970799, 1997.

3. C. Arcoumanis, and M. Gavaises Linking nozzle flow with spray characteris-

tics in a Diesel fuel injection systems. Atom. Sprays, 8:307–347, 1998.

4. C. Arcoumanis, M. Badami, and M. Gavaises. Cavitation in Real-Size Multi-

Hole Diesel Injector Nozzles. Technical Report 2000-01-1249, SAE, 2000.

5. C. Badock, R. Wirth, A. Fath, and A. Leipertz. Investigation of Cavitation in

Real Size Diesel Injection Nozzles. Int. J. Heat and Fluid Flow, 20:538–544,

1999.

6. J. C. Beale, and R. D. Reitz. Modeling Spray Atomization With the Kelvin-

Helmholtz/ Rayleigh-Taylor Hybrid Model. Atomization Sprays, 9:623–650,

1999.

7. P.A. Beau, M. Funk, R. Lebas, and F.X. Demoulin. CavitationApplying

quasi-multiphase model to simulate atomization in diesel engines. SAE Paper

2005-01-0220, 2005.

8. P.A. Beau, T. Menard, R. Lebas, A. Berlement, S.Tanguy, and F.X. De-

moulin. Numerical jet atomization. Part II : modeling information and com-

parison with DNS. FEDSM2006, 98166, 2006.

9. G. M. Bianchi, P. Pelloni, S. Toninel, R. Scardovelli, A. Leboissetier, and

S. Zaleski. Improving the knowledge of high-speed liquid jets atomization by

using quasi-direct 3d simulation. Technical Report 2005-24-089, SAE, 2005.

10. G. M. Bianchi, F. Minelli, R. Scardovelli, and S. Zaleski. 3d large scale

simulation of the high-speed liquid jet atomization. Technical Report 2007-

01-0244, SAE, 2007.

11. R.W. Bilger. Turbulent flows with nonpremixed reactants. In P.A.Libby and
28 Modeling Primary Atomization

F.A.Williams, editors, Turbulent reacting flows, pages 65–114, Berlin, 1980.

Springer.

12. G. Blokkeel, F.X. Demoulin, and R. Borghi. Modeling of two-phase flows: An

Eulerian model for Diesel injection. In J.H. Whitelaw et al., editors, Thermo-

and-Fluid Dynamics process in Diesel Engines 2, pages 87–105, Berlin, 2004.

Springer.

13. T. Boeck, J. Li, E. Lopez-Pages, P. Yecko, and S. Zaleski. Droplet formation

in sheared liquid-gas layers. Theor. Comput. Fluid Dynamics, 21:59–76, 2007.

submitted to Theor. Comput. Fluid Dyn.

14. J. U. Brackbill, D. B. Kothe, and C. Zemach. A continuum method for

modeling surface tension. J. Comput. Phys., 100:335–354, 1992.

15. H. Chaves, and C. Ludwig. Characterization of Cavitation in Transparent

Nozzles Depending on the Nozzle Geometry. In ILASS 2005, Orleans, France,

2005.

16. B. Chehroudi, and D. Talley. The fractal geometry of round turbulent

cryogenic nitrogen jets at subcritical pressures. J. Atomization and Sprays,

14(1):81–91, 2004.

17. Z. Cheng, and S. Redner. Scaling Theory of Fragmentation. Phys. Rev.

Letters, 60(24), 1988.

18. Z. Cheng, and S. Redner. Kinetics of Fragmentation. J. Phys. A, 23:1233–

1258, 1990.

19. N.A. Chigier. The Physics of Atomization. Plenary Lecture. In ICLASS 91,

Gaithersburg, MD, USA, July, 1991.

20. O. Desjardins, V. Moureau, E. Knudsen, M. Herrmann, and H. Pitsch. Con-

servative level set/ghost fluid method for simulating primary atomization. In


Modeling Primary Atomization 29

ILASS Americas 20th Annual Conference on Liquid Atomization and Spray

Systems. ILASS Americas, 2007.

21. P.E. Dimotakis. Two-dimensional shear layer entrainment. AIAA,

24(11):1791–1796, 1986.

22. R. Domann, and Y. Hardalupas. Breakup model for accelerating liquid

jets. In Proc. 42nd AIAA Aerospace Science Meeting and Exhibition. Reno,

Nevada, USA, 2004.

23. C. Engelbert, Y. Hardalupas, and J.H. Whitelaw. Breakup phenomena in

coaxial airblast atomizers. Mathematical and Physical Sciences, Osborn

Reynolds Centenary Volume, 1941:189–229, 1995.

24. B. Engquist, A.-K. Tornberg, and R. Tsai. Discretization of dirac delta

functions in level set methods. J. Comput. Phys., 207(207):28–51, 2005.

25. D. Enright, R. Fedkiw, J. Ferziger, and I. Mitchell. A hybrid particle level

set method for improved interface capturing. J. Comput. Phys., 183:83–116,

2002.

26. G. M. Faeth, L. P. Hsiang, and P. K. Wu. Structure and Breakup Properties

of Sprays. Int. J. Multiphase Flow, 21:99–127, 1995.

27. Z. Farago, and N. Chigier. Morfological classification of disintegration of

round liquid jets in a coaxial air stream. Atom. Sprays, 2:137–153, 1992.

28. A.F. Fillipov. On the distribution of the sizes of particles which undergo

splitting. Theor. Prob. Appl, 4:275, 1961.

29. M. M. Francois, S. J. Cummins, E. D. Dendy, D. B. Kothe, J. M. Sicilian,

and M. W. Williams. A balanced-force algorithm for continuous and sharp

interfacial surface tension models within a volume tracking framework. J.

Comput. Phys., 213:141–173, 2006.


30 Modeling Primary Atomization

30. M. Gorokhovski. The Stochastic Sub-Grid-Scale Approach for Spray Atom-

ization. Atom. Sprays, 11:505–519, 2001.

31. M. Gorokhovski, and V. Saveliev. Further Analyses of Kolmogorov’s Model

of Breakup. Phys. Fluids, 15(1):184–192, 2003.

32. M. Gorokhovski, A. Chtab, and J. Jouanguy. Simulation of air-blast atomiza-

tion: ”floating guard” statistic particle method for conditioning of LES com-

putation; stochastic models of break-up and coalescence. In Proc. ICLASS-

2006, Kyoto, Japan, 2006.

33. M. Gorokhovski, J. Jouanguy, and A. Chtab. Stochastic simulation of the

liquid jet break-up in a high-speed air co-flow. ICMF 2007, Leipzig, Germany,

July 9 - 13, 2007.

34. M. Gorokhovski, and V. Saveliev. Statistical universalities in fragmentation

under scaling symmetry with the constant frequency of fragmentation. sub-

mitted to J. Stat. Phys., 2007.

35. D. Gueyffier, J. Li, A. Nadim, S. Scardovelli, and S. Zaleski. Volume of Fluid

interface tracking with smoothed surface stress methods for three-dimensional

flows. J. Comput. Phys., 152:423–456, 1999.

36. C. Habchi, and T. Baritaud. Modeling atomization and break-up in high

pressure Diesel Sprays. Technical Report 9700881, SAE Paper, 1997.

37. J. Helmsen, P. Colella, and E. G. Puckett. Non-convex profile evolution in

two dimensions using volume of fluids. Technical Report LBNL Technical

Report LBNL-40693, Lawrence Berkeley National Laboratory, 1997.

38. M. Herrmann. Refined level set grid method for tracking interfaces. In Annual

Research Briefs-2005, pages 3–18. Center for Turbulence Research, 2005.

39. M. Herrmann. A balanced force refined level set grid method for two-phase
Modeling Primary Atomization 31

flows on unstructured flow solver grids. In Annual Research Briefs-2006.

Center for Turbulence Research, 2007. in print.

40. C. W. Hirt and B. D. Nichols. Volume of fluid (vof) method for the dynamics

of free boundaries. J. Comput. Phys., 39:201–225, 1981.

41. J.-M. Hong, T. Shinar, M. Kang, and R. Fedkiw. On boundary condition

capturing for multiphase interfaces. J. Sci. Comput., 31(1-2):99–125, 2007.

42. M. Hong, A. Cartellier, and E. Hopfinger. Atomisation and mixing in coax-

ial injection. In Proc. 4th Int. Conference on Launcher Technology ”Space

Launcher Liquid Propulsion”, Liege, Belgique, 3-6 December 2002.

43. M. Hong. Atomisation et mélange dans les jets coaxiaux liquide-gaz. PhD

Thesis, University of Grenoble, France, 2003.

44. K. Y. Huh, and A. D. Gosman. A phenomenological model of Diesel spray

atomization. In Proc. ICMF, Tsukuba, Japan, 1991.

45. M. Kang, R. Fedkiw, and X.-D. Liu. A boundary condition capturing method

for multiphase incompressible flow. J. Sci. Comput., 15(3):323–360, 2000.

46. D. Kim, O. Desjardins, M. Herrmann, and P. Moin. The primary breakup

of a round liquid jet by a coaxial flow of gas. In ILASS Americas 20th An-

nual Conference on Liquid Atomization and Spray Systems. ILASS Americas,

2007.

47. M. Klein, A. Sadiki, and J. Janicka. A digital filter based generation of inflow

data for spatially developing direct numerical or large eddy simulations. J.

Comput. Phys., 186:652–665, 2003.

48. M. Klein. Direct numerical simulation of a spatially developing water sheet

at moderate reynolds number. Int. J. Heat Fluid Fl., 26(5):722–731, October

2005.
32 Modeling Primary Atomization

49. A.N. Kolmogorov. On the Log-Normal Distribution of Particles Sizes During

Breakup Process. Dokl. Akad. Nauk. SSSR, XXXI(2):99–101, 1941.

50. A.N. Kolmogorov. On the Drops Breakup in the Turbulent Flow.

Gidromekanika, DAN, LXVI, NS:825–828, 1949.

51. S. C. Kong, P. K. Senecal, and R. D. Reitz. Developments in spray mod-

elling in diesel and direct-injection gasoline engines. Oil & Gas Science and

Technology - Rev. IFP, 54(2):197–204, 1999.

52. S. Jay, F. Lacas, and S. Candel. Combined surface density for dense spray

combustion. Combustion and Flame, 144:558–577, 2005.

53. J.C. Lasheras, E. Villermaux, and E.J. Hopfinger. Break-up and atomization

of a round water jet by a high-speed annular air jet. J.Fluid Mech., 357:351–

379, 1998.

54. J.C. Lasheras, and E.J. Hopfinger. Liquid jet instability and atomization in

a coaxial gas stream. Annu. Rev. Fluid Mech., 32:275–308, 2000.

55. B. Lafaurie, C. Nardonne, R. Scardovelli, S. Zaleski, and G. Zanetti. Modeling

merging and fragmentation in multiphase flows with SURFER. J. Comput.

Phys., 113:134–147, 1994.

56. F.E. Marble, and J.E. Broadwell. The coherent flame model for turbulent

chemical reactions. Tech. Rep. TRW-9-PV, Project Squid, Chaffee Hall,

Purdue University, West Lafayette, 1977.

57. P. Marmottant. Atomisation d’un liquide par un courant gazeux. PhD Thesis,

UJF, Grenoble, 2001.

58. P. Marmottant and E. Villermaux. On spray formation. J. Fluid Mech.,

498:73–111, 2004.

59. E.D. McGrady, and R.M. Ziff. ”Shattering” Transition in Fragmentation.


Modeling Primary Atomization 33

Phys. Rev. Letters, 58(9), 1987.

60. T. Menard, P. A. Beau, S. Tanguy, F. X. Demoulin, and A. Berlemont.

Primary break up modeling, part a: Dns, a tool to explore primary break up.

In ICLASS-2006, number ICLASS06-034, 2006.

61. N. Meyers, F. Dupoirieux, and M. Habiballah. An Eulerian model for

LOX/GH2 shear-coaxial jet atomization and combustion. In First European

Conference for Aerospace Sciences (EUCASS), Moscou, Russie, 2005.

62. W. Noh and P. Woodward. Slic (simple line interface calculation). In

A. van de Vooren and P. Zandbergen, editors, Proc. 5th Int. Conf. Fluid

Dyn., volume 59 of Lect. Notes Phys., pages 330–340, Berlin, 1976. Springer-

Verlag.

63. R. R. Nourgaliev and T. G. Theofanous. High-fidelity interface tracking in

compressible flows: Unlimited anchored adaptive level set. J. Comput. Phys.,

doi:10.1016/j.jcp.2006.10.031, 2007.

64. E. Olsson and G. Kreiss. A conservative level set method for two phase flow.

J. Comput. Phys., 210:225–246, 2005.

65. S. Osher and J. A. Sethian. Fronts propagating with curvature-dependent

speed: Algorithms based on Hamilton-Jacobi formulations. J. Comput.

Phys., 79:12–49, 1988.

66. S. Osher and R. Fedkiw. Level Set Methods and Dynamic Implicit Surfaces,

volume 153. Springer, New York, 2002.

67. Y. Pan and K. Suga. A numerical study on the breakup process of laminar

liquid jets into a gas. Phys. Fluids, 18(052101):1–11, 2006.

68. H. Park, S. S. Yoon, and S. D. Heister. A nonlinear atomization model for

computation of drop size distributions and spray simulations. Int. J. Numer.


34 Modeling Primary Atomization

Meth. Fluids, 48(11):1219–1240, August 2005.

69. M. A. Patterson, and R. D. Reitz. Modeling the effects of fuel spray character-

istics on Diesel engine combustion and emissions. Technical Report 980131,

SAE Paper, 1998.

70. C. S. Peskin. Numerical analysis of blood flow in the heart. J. Comput.

Phys., 25:220–252, 1977.

71. S. P. van der Pijl, A. Segal, and C. Vuik. A mass-conserving level-set method

for modelling of multi-phase flows. Int. J. Numer. Meth. Fluids, 47:339–361,

2005.

72. L. Raynal. Instabilite et entrainement a l’interface d’une couche de mélange

liquide-gaz. PhD Thesis, UJF, Grenoble, 1997.

73. S. Redner. Fragmentation. In H.J. Herrmann and S.Roux, editors, Statistical

models for the fracture of disordered media. Elsevier, Amsterdam, 1990.

74. H. Rehab. Structure de l’écoulement et mélange dans le champs proche des

jets coaxiaux. PhD Thesis, UJF, Grenoble, 1997.

75. H. Rehab, E. Villermaux, and E. J. Hopfinger. Flow regimes of large velocity

ratio coaxial jets. J.Fluid Mech., 345:357–381, 1997.

76. R. D. Reitz. Modeling atomization processes in high-pressure vaporizing

sprays. Atom. Spray Tech., 3:309–337, 1987.

77. Y. Renardy, and M. Renardy. Prost: A parabolic reconstruction of surface

tension for the volume-of-fluid method. J. Comput. Phys., 183:400–421, 2002.

78. N. Rimbert, and O. Sero-Guillaume. Log-stable laws as asymptotic solutions

to a fragmentation equation: Application to the distribution of droplets in a

high Weber-number spray. Phys. Rev. E, 69:056316, 2004.

79. W. Rodi. A new algebraic relation for calculating the Reynolds stresses.
Modeling Primary Atomization 35

ZAMM, 56:219–221, 1976.

80. V. Sabel’nikov, M. Gorokhovski, and N. Baricault. The extended IEM mixing

model in framework of the composition PDF approach; applications to Diesel

spray combustion. J. Comb. Theory and Modeling, 10(1):155–169, 2006.

81. K. A. Sallam, Z. Dai, and G. M. Faeth. Liquid break-up at the surface of

turbulent round liquid jets in still gases. Int. J. Multiphase Flows, 28:427–

449, 2002.

82. V. Saveliev, and K. Nanbu. Collision group and renormalization of the Boltz-

man collision integral. Phys. Rev. E, 65:051205, 2002.

83. V. Saveliev, and M. Gorokhovski. Group-theoretical model of developed

turbulence and renormalization of Navier-Stokes equations. Phys. Rev. E,

72:016302, 2005.

84. R. Scardovelli and S. Zaleski. Direct numerical simulation of free-surface and

interfacial flow. Annu. Rev. Fluid Mech., 31:567–603, 1999.

85. J. A. Sethian. A fast marching level set method for monotonically advancing

fronts. Proc. Natl. Acad. Sci. USA, 93:1591–1595, 1996.

86. U. Shavit, and N. Chigier. Fractal Dimensions of Liquid Jet Interface Under

Breakup. J. Atomization and Sprays, 5:525–543, 1995.

87. M. Stahl, M. Gnir, N. Damaschke, and C. Tropea. Laser Doppler mea-

surements of nozzle flow and optical characterisation of the generated spray.

ILASS-2005, Orleans, France, 2005.

88. M. Stahl, N. Damaschke, and C. Tropea. Experimental investigation of tur-

bulence and cavitation inside a pressure atomizer and optical characterization

of the generated spray. ILASS-2006, Kyoto, Japan, 2006.

89. M. Sussman, P. Smereka, and S. Osher. A level set method for computing
36 Modeling Primary Atomization

solutions to incompressible two-phase flow. J. Comput. Phys., 114:146, 1994.

90. M. Sussman and E. G. Puckett. A coupled level set and volume of fluid

method for computing 3d and axisymmetric incompressible two-phase flows.

J. Comput. Phys., 162(2):301–337, 2000.

91. M. Sussman. A second order coupled level set and volume-of-fluid method

for computing growth and collapse of vapor bubbles. J. Comput. Phys.,

187(1):110–136, 2003.

92. F. X. Tanner Development and validation of a cascade atomization and

drop breakup model for high-velocity dense sprays. Atomisation and Sprays,

164:211–242, 2004.

93. G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber,

J. Han, S. Nas, and Y.-J. Jan. A front-tracking method for the computations

of multiphase flow. J. Comput. Phys., 169:708–759, 2001.

94. S. O. Unverdi and G. Tryggvason. A front-tracking method for viscous,

incompressible, multi-fluid flows. J. Comput. Phys., 100:25–37, 1992.

95. A. Vallet, aand R. Borghi. An Eulerian model of atomization of a liquid jet.

ICMF’98, Lyon, France, Jule 8-12, 1998.

96. A. Vallet, A. Burluka, and R. Borghi. Development of an Eulerian model for

the atomization of a liquid jet. Atomization and Sprays, 11:619–642, 2001.

97. C.M. Varga. Initial break-up of a small diameter liquid jet by high-speed gas

stream. PHD Thesis of UCSD, San Diego, 2002.

98. C.M. Varga, J.C. Lasheras, and E.J. Hopfinger. Initial breakup of a small-

diameter liquid jet by a high-speed gas stream. J. Fluid. Mech, 497:405–434,

2003.

99. E. Villermaux. Mixing and spray formation in coaxial jets. Journal of Propul-
Modeling Primary Atomization 37

sion and Power, 14(5):807–817, 1998.

100. E. de Villiers, A. D. Gosman, and H. G. Weller. Large eddy simulation

of primary diesel spray atomization. Technical Report 2004-01-0100, SAE,

2004.

101. I. Vinkovic, C. Agguire, S. Simoens, and M. Gorokhovski. Large eddy sim-

ulation of droplet dispersion for inhomogeneous turbulent wall flow. Int. J.

Multiphase Flow, 32(3):344–364, 2006.

102. O. Werquin. Diagnostics des scalaires par plan laser dans les jets diphasiques

denses. PhD thesis, University of Rouen, France, 2001.

103. M. W. Williams, D. B. Kothe, and E. G. Puckett. Convergence and accuracy

of continuum surface tension models. In W. Shyy and R. Narayanan, editors,

Fluid Dynamics at Interfaces, pages 294–305. Cambridge University Press,

Cambridge, 1999.

104. K.-J. Wu, R. D. Reitz, and F. V. Bracco. Measurements of drop size at the

spray edge near the nozzle in atomizing liquid jets. Phys. Fluids, 29(4):941–

951, 1986.

105. P. K. Wu, L. K. Tseng, and G.M. Faeth. Primary Breakup in Gas/Liquid

Mixing Layers for Turbulent Liquids. Atomization and Sprays, 2:295–317,

1992.

106. P. K. Wu, and G.M. Faeth. Aerodynamic Effects on Primary Breakup of

Turbulent Liquids. Atomization and Sprays, 3:265–289, 1993.

107. Y. Yi and R. D. Reitz. Modeling the primary break-up of high-speed jets.

Atom. Sprays, 14(1):53–80, 2004.

108. S. S. Yoon and S. D. Heister. A nonlinear atomization model based on a

boundary layer instability mechanism. Phys. Fluids, 16(1):47–61, January


38 Modeling Primary Atomization

2004.

109. S. S. Yoon. Droplet distributions at the liquid core of a turbulent spray.

Phys. Fluids, 17(3):035103, March 2005.

110. W.X. Zhou, and Z.H. Yu. Multifractality of drop breakup in the air-blast

nozzle atomization process. Phys. Rev. E, 63:016302, 2000.

111. R.M. Ziff, and E.D. McGrady. The kinetics of cluster fragmentation and

depolymerisation. J. Phys. A, 18:3027–3037, 1985.

112. R.M. Ziff. New solutions to the fragmentation equation. J. Phys. A Math.

Gen., 24:2821–2828, 1991.


Modeling Primary Atomization 39

Figure 1: Two successive snapshots of the turbulent atomization of a liquid Diesel

jet (20).

Figure 2: Zoom into the chaotic structure of the interface of the atomizing liquid

Diesel jet (20).


40 Modeling Primary Atomization

Figure 3: Co-axial atomization mechanisms. Initial Rayleigh limit Kelvin-

Helmholtz instability (top left), Rayleigh-Taylor instability (top right), ligament

stretching (bottom left), and ligament detachment and drop generation (bottom

right) (46). RLSG captured phase interface is shown in blue, and Lagrangian

spray drops are depicted in yellow.


Modeling Primary Atomization 41

Figure 4: Modeling liquid core geometry and averaged liquid fraction field (32)

compared to measurements from (102). Here the round liquid jet issued from

the central tube (Dl = 1.8 mm) at ul = 0.68 m/s (on the upper part) and

ul = 1.36 m/s (on the bottom part) is atomized by a parallel flow of air issued

at ug = 60 m/sfrom the annular duct (Dg = 3.4 mm).


42 Modeling Primary Atomization

Figure 5: LES of the gas flow (velocity of injection ug = 140 m/s) combined with

stochastic simulation of the liquid jet depletion (33) in experimental conditions

(53); on the left-hand side, ul = 0.13 m/s; on the right hand-side, ul = 0.55 m/s;

coordinates are dimensioned in the inner nozzle orifice diameter. Dl = 3.8 mm;

the nozzle diameter of the annular air jet is Dg = 5.6 mm.


Modeling Primary Atomization 43

Figure 6: Application of Eulerian RANS model (8) (upper part) to the Diesel

spray atomization (uinj = 100 m/s, dinj = 100 µm, ρg = 25 km/m3 , ρl =

670 km/m3 ); on the bottom: comparison with mean distributions from quasi-

DNS of (60); from the left to the right : the liquid mass fraction, the interface

00 Y 00 , the kinetic turbulent energy.


density, the turbulent flux ρug
i l

View publication stats

You might also like