You are on page 1of 22

P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL

November 25, 1997 11:20 Annual Reviews AR049-04

Annu. Rev. Fluid Mech. 1998. 30:85–105


Copyright c 1998 by Annual Reviews Inc. All rights reserved

DROP AND SPRAY FORMATION


FROM A LIQUID JET
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

S. P. Lin
Mechanical and Aeronautical Engineering Department, Clarkson University, Potsdam,
by Princeton University Library on 10/24/05. For personal use only.

New York 13699; e-mail: gw02@ sun.soe.clarkson.edu

R. D. Reitz
Mechanical Engineering Department, University of Wisconsin, Madison, Wisconsin,
53706; e-mail: reitz@engr.wisc.edu

KEY WORDS: jet instability, breakup regimes, atomization, sprays, surface tension

ABSTRACT
A liquid jet emanating from a nozzle into an ambient gas is inherently unstable.
It may break up into drops of diameters comparable to the jet diameter or into
droplets of diameters several orders of magnitude smaller. The sizes of the drops
formed from a liquid jet without external control are in general not uniform. The
sizes as well as the size distribution depend on the range of flow parameters in
which the jet is produced. The jet breakup exhibits different characteristics in
different regimes of the relevant flow parameters because of the different physical
mechanisms involved. Some recent works based on linear stability theories aimed
at the delineation of the different regimes and elucidation of the associated phys-
ical mechanisms are reviewed, with the intention of presenting current scientific
knowledge on the subject. The unresolved scientific issues are pointed out.

1. INTRODUCTION
The breakup of a liquid jet emanating into another fluid has been quantitatively
studied for more than a century. Plateau (1873) observed that the surface energy
of a uniform circular cylindrical jet is not the minimum attainable for a given
jet volume. He argued that the jet tends to break into segments of equal length,
each of which is 2π times longer than the jet radius, such that the spherical
drops formed from these segments give the minimum surface energy if a drop
is formed from each segment. Rayleigh (1879a, b) showed that the jet breakup

85
0066-4189/98/0115-0085$08.00
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

86 LIN & REITZ

is the consequence of hydrodynamic instability. Neglecting the ambient fluid,


the viscosity of the jet liquid, and gravity, he demonstrated that a circular
cylindrical liquid jet is unstable with respect to disturbances of wavelengths
larger than the jet circumference. Among all unstable disturbances, the jet is
most susceptible to disturbances with wavelengths 143.7% of its circumference.
Rayleigh also considered the cases of a viscous jet in an inviscid gas (1892a)
and an inviscid gas jet in an inviscid liquid (1892b). He showed that if the mass
of the gas is neglected, the most amplified disturbance in the first case possesses
an infinitely long wave length, and that for the second case it is 206.5% of the
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

jet circumference.
Tomotika (1935) showed that an optimal ratio of viscosities of the jet and
the ambient fluid exists for which a disturbance of finite wavelength attains
by Princeton University Library on 10/24/05. For personal use only.

the maximum growth rate. Chandrasekhar (1961) took into account the liquid
viscosity and the liquid density, which was neglected by Rayleigh, and showed
mathematically that the viscosity tends to reduce the breakup rate and increase
the drop size. He also showed that the physical mechanism of the breakup of a
viscous liquid jet in a vacuum is capillary pinching. The theoretical results of
Rayleigh and Chandrasekhar appear to be in agreement with the experiments
of Donnelly & Glaberson (1966) and Goedde & Yuen (1970).
Weber (1931) considered the effects of the liquid viscosity as well as the
density of the ambient fluid. His theoretical prediction did not agree well with
experimental data, as pointed out by Sterling & Sleicher (1975), who improved
Weber’s theory with partial success. Taylor (1962) showed that the density of
the ambient gas has a profound effect on the form of the jet breakup. For a
sufficiently large gas inertia force (which is proportional to the gas density)
relative to the surface tension force per unit of interfacial area, the jet may
generate at the liquid-gas interface droplets with diameters much smaller than
its own diameter. This Taylor mode of jet breakup is the so-called “atomization”
that leads to fine spray formation.
The number of publications following the above pioneering works is indeed
very large owing to the increasingly wide applications of the jet breakup pro-
cesses. There have been several review articles in this area (e.g. Sirignano
1993). The latest ones are by Chigier & Reitz (1996) and Lin (1996). In this
review, we focus on the physical mechanisms that cause the onset of the jet
breakup at the liquid-gas interface. The nonlinear evolution after the onset of
jet breakup is not considered. The physical mechanism of breakup frequently
remains the same during the nonlinear evolution, although the nonlinear theory
may produce additional quantitative results. For example, the satellite droplets
formed from the ligament between two main drops are not predicted by lin-
ear theories, but the mechanism of the satellite formation remains capillary
pinching of the Rayleigh mode.
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 87

In addition, previous works on the breakup processes of a liquid jet in another


liquid (which are relevant to the emulsification process) are not reviewed. Nor
do we review works that focus on the application of the jet breakup process,
although they are interesting and are of considerable importance in technology
and science. Even within the area of the fluid dynamics of the onset of liquid
jet breakup in a gas, many important works are not commented on explicitly.
Most of these important works are cited in the papers discussed in this chapter.
In Section 2, works on the delineation of different regimes of jet breakup with
correlations in terms of relevant flow parameters are reviewed. These correla-
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

tions are based mainly on temporal linear stability theory. Section 2 provides a
framework for the discussion in Section 3 of the physical mechanisms at work
in the different regimes. The elucidations of the physical mechanisms are based
by Princeton University Library on 10/24/05. For personal use only.

on absolute and convective instability theories. A general critical discussion


of the recent work in the field is given in Section 4. Section 5 summarizes the
scientific issues that remain to be investigated.

2. BREAKUP REGIMES
The breakup of a liquid jet injected through a circular nozzle hole into a stagnant
gas has been studied most frequently. Previous studies have established that
the spray properties are influenced by an unusually large number of parameters,
including nozzle internal flow effects resulting from cavitation, the jet velocity
profile and turbulence at the nozzle exit, and the physical and thermodynamic
states of both liquid and gas (e.g. Wu et al 1992, Eroglu et al 1991, and Reitz
& Bracco 1979). The precise mechanisms of breakup are still being researched
(e.g. Lin 1996, Chigier & Reitz 1996). However, linear stability theory can
provide qualitative descriptions of breakup phenomena and predict the existence
of various breakup regimes. It is noteworthy that the influence of nozzle internal
flow effects is included only empirically in most jet breakup theories. These
effects are known to be important, particularly for high-speed jet breakup.
Jet breakup phenomena have been divided into regimes that reflect differ-
ences in the appearance of jets as the operating conditions are changed. The
regimes are due to the action of dominant forces on the jet, leading to its breakup,
and it is important that these forces be identified in order to explain the breakup
mechanism in each regime (Reitz & Bracco 1986). The case of a round liq-
uid jet injected into a stagnant gas is shown in Figure 1. Four main breakup
regimes have been identified that correspond to different combinations of liquid
inertia, surface tension, and aerodynamic forces acting on the jet. These have
been named the Rayleigh regime, the first wind-induced regime, the second
wind-induced regime, and the atomization regime (Figure 1) (Reitz & Bracco
1986).
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

88 LIN & REITZ


Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org
by Princeton University Library on 10/24/05. For personal use only.

Figure 1 (a) Rayleigh breakup. Drop diameters larger than the jet diameter. Breakup occurs many
nozzle diameters downstream of nozzle. (b) First wind-induced regime. Drops with diameters of
the order of jet diameter. Breakup occurs many nozzle diameters downstream of nozzle. (c)
Second wind-induced regime. Drop sizes smaller than the jet diameter. Breakup starts some
distance downstream of nozzle. (d) Atomization regime. Drop sizes much smaller than the jet
diameter. Breakup starts at nozzle exit.

At low jet velocities, the growth of long-wavelength, small-amplitude distur-


bances on the liquid surface promoted by the interaction between the liquid and
ambient gas is believed to initiate the liquid breakup process. The existence
of these waves is clearly demonstrated in Figure 1a and b. For high-speed
liquid jets, the breakup is thought to result from the unstable growth of short-
wavelength waves (Figure 1c and d) (Reitz & Bracco 1982). The breakup drop
sizes are on the order of the jet diameter in the Rayleigh and first wind-induced
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 89

Λ
L
L
Λ
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org
by Princeton University Library on 10/24/05. For personal use only.

A
B
C
D

U
Figure 2 Schematic diagram of the jet breakup length curve.

breakup regimes. The drop sizes are very much less than the jet diameter in the
second wind-induced and atomization regimes.
A convenient method for categorizing jet breakup regimes is to consider
the length of the coherent portion of the liquid jet or its unbroken length, L,
as a function of the jet exit velocity, U (Figure 2) (e.g. Leroux et al 1996).
Beyond the dripping flow regime, the breakup length at first increases linearly
with increasing jet velocity, reaches a maximum, and then decreases (regions A
and B). Drops are pinched off from the end of the jet, with diameters comparable
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

90 LIN & REITZ

to that of the jet (Figure 1a and b). These first two breakup regimes, which are
reasonably well understood, correspond to the Rayleigh and first wind-induced
breakup regimes.
The form of the breakup curve in these two regimes is well predicted by linear
stability theories such as that of Sterling & Sleicher (1975). In this temporal
stability theory, it is assumed that the interface, r = a, of a circular jet of radius,
a, is perturbed by an axisymmetric wave with a Fourier component of the form
η = η0 exp(ωt + ikx), (1)
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

where η = η(x, t) is the displacement of the liquid surface. A cylindrical


coordinate system is used that moves in the axial direction, x, at the jet velocity,
U. The fluid is located at the origin, the nozzle exit x = 0, when t = 0. η0 is
by Princeton University Library on 10/24/05. For personal use only.

an initial disturbance level (the initial amplitude of the perturbation), k is the


wave number of the disturbance, and ω is the complex frequency, the real part
of which, ωr, is the growth rate. The stability of the liquid surface to linear
perturbations is examined and a dispersion equation is derived that relates the
complex frequency, ω, of an initial perturbation of infinitesimal amplitude, to
its wavelength λ (or wavenumber k = 2π/λ). The relationship also includes
the physical and dynamical parameters of the liquid jet and the surrounding gas
(e.g. Reitz & Bracco 1986), and there exists a maximum growth rate or most
unstable wave. Further discussion of the characteristics of the linear solution
is given in the next section.
Reitz (1987) generated curve-fits of numerical solutions to the dispersion
equation for the maximum growth rate (ωr = ) and for the corresponding
wavelength (λ = 3) of the form
3 (1 + 0.45 Z 0.5 )(1 + 0.4 T 0.7 )
= 9.02 (2a)
a (1 + 0.87 We1.67
2 )
0.6

ρ1 a 3 0.34 + 0.38 We1.5


 = 2
(2b)
σ 0.5 (1 + Z )(1 + 1.4 T 0.6 )
where Z = We0.5 1 /Re1, T = ZWe2 , We1 = ρ 1U a/σ , We2 = ρ 2U a/σ ,
0.5 2 2

and Re1 = Ua/ν 1. U is the relative velocity between the jet and the gas,
and the subscripts 1 and 2 identify properties based on the liquid and the gas,
respectively. As can be seen from Equations 2a and 2b, the maximum wave
growth rate increases and the corresponding wavelength decreases rapidly with
increasing Weber number, which is the ratio of the inertia force to surface
tension force acting on the jet. The effect of the liquid viscosity (which appears
in the Reynolds number, Re, and the Ohnesorge number, Z) is seen to reduce
the wave growth rate and to increase the wave length significantly as the liquid
viscosity increases.
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 91

At low jet velocities (small Weber numbers) it is reasonable to assume that


disruption of the jet occurs when the dominant wave’s amplitude is equal to the
jet radius. In this case, the jet breakup time, τ , is given by a = η0 exp (τ ),
and the breakup length is predicted to be
L = U τ = U/ ln(a/η0 ) (3)
For low-speed jets in the Rayleigh breakup regime, the parameter ln(a/η0)
has been determined from experiments to be roughly equal to 12. For an
inviscid liquid it is readily seen that Equation 3, when combined with Equation
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

2b, predicts the linear increase in jet breakup length with jet velocity at low
gas densities, since τ is independent of the jet velocity. A maximum in the
breakup length curve is predicted if the gas density is non-zero, i.e. the theory
by Princeton University Library on 10/24/05. For personal use only.

predicts that aerodynamic effects are responsible for the decrease in the breakup
length as the Weber number is increased beyond the maximum point. However,
discrepancies have been found between the predicted location of the maximum
point and experimental data.
The shape of the breakup curve has been reviewed by many researchers,
including Grant & Middleman (1966) and McCarthy & Malloy (1974) who
discussed the effects of the ambient gas, fluid properties, and nozzle design.
Leroux et al (1996) pointed out that the location of the maximum point de-
pends on nozzle parameters [presumably through the influence of the initial
disturbance term, ln(a/η0)], and also on the magnitude of the gas density itself.
Indeed, Leroux et al (1996) proposed empirical modifications to account for
these effects, which extend the theory of Sterling & Sleicher (1975).
Beyond the first wind-induced breakup regime (region B, Figure 2) there is
even more confusion about the breakup-length trends. For example, Haenlein
(1932) reported that the jet breakup length increases again with increasing jet
velocity (region C), and then abruptly reduces to zero (region D). McCarthy
& Malloy (1974) reported that the breakup length continually increases. More
recently, Hiroyasu et al (1991) discovered discontinuous elongations and short-
enings of the jet with changes in the jet velocity. These apparent anomalies are
associated with changes in the nozzle internal flow patterns caused by separation
and cavitation phenomena, which also exhibited hysteresis effects. Jets from
cavitating nozzles were found to have very short breakup lengths. Detached
flow jets have long breakup lengths. These phenomena may help explain the
previous discrepancies in measurements of breakup lengths in the spray liter-
ature, since only recently have investigators paid attention to nozzle flow and
geometry effects.
Equation 3 predicts that the breakup length decreases continuously as the jet
velocity is increased when the effect of the gas density is significant. How-
ever, the validity of the assumption that L = Uτ becomes questionable for
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

92 LIN & REITZ

high-speed jets, because the breakup mechanism is no longer due to capillary


pinching, but is now due to the unstable growth of short-wavelength surface
waves (Figure 2). In fact, as the jet velocity is increased, it becomes difficult
to define a precise breakup length, and probability density functions are found
useful to quantify the breakup length (e.g. Leroux et al 1996).
The details of the unstable growth of short-wavelength waves on the surface of
the liquid jet near the nozzle exit are obscured by the dense spray that surrounds
the jet. However, it is generally believed that the jet consists of an unbroken
inner liquid core in the vicinity of the nozzle exit, and droplets are stripped
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

from the core by the action of aerodynamic forces at the liquid-gas interface
(Reitz & Bracco 1982). Attempts have been made to measure the length of
the core region by using intrusive techniques such as electrical conductivity
by Princeton University Library on 10/24/05. For personal use only.

measurements (e.g. Chehroudi et al 1985, Hiroyasu et al 1991), and laser sheet


visualization (e.g. Gulder et al 1994, Dan et al 1997). The core length depends
on the liquid/gas density ratio and only weakly on the fluid properties and the
jet velocity.
These trends can be demonstrated by using Taylor’s (1940) analysis of high-
speed liquid jet breakup. Taylor considered the rate of mass loss per unit length
of the jet caused by droplet erosion from the liquid surface resulting from the
unstable growth of short-wavelength surface waves and showed that the breakup
length of a high-speed jet is given by
L/a = B(ρ1 /ρ2 )1/2 / f (T ) (4)
where T is Taylor’s parameter, T = ρ 1/ρ 2(Re1/We1)2, and the function √ f (T )
has been approximated from Taylor’s numerical results as f (T ) = 3/6 [1 −
exp(−10T )] by Dan et al (1997).
The constant B in Equation 4 has a recommended value of 4.04 for typical
diesel spray nozzles (Chehroudi & Bracco 1985). However, nozzle internal
design effects are clearly important for high-speed jet breakup. It is known that
high-speed liquid jets in jet cutting applications remain intact for a distance
of many diameters away from the nozzle. On the other hand, modern diesel
injectors employ very similar injection pressures, but diesel spray breakup starts
at the nozzle exit (e.g. Figure 1d). The significant differences in the interior
nozzle design features of these two applications account for their different
performances. Diesel nozzles are typically short-length holes with sharp-edged
inlets, whereas jet cutting nozzles consist of contoured nozzles to minimize
initial disturbance levels to the liquid flow.
Criteria for predicting the onset of the breakup regimes have been reviewed
by Chigier & Reitz (1996). Consideration of the balance between the liquid
inertia force and the surface tension force of a free column of liquid led Ranz
(1956) to the criterion that dripping no longer occurs from the nozzle exit
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 93

(i.e. a jet is formed) if WeL > 8, where WeL = ρ 1U2(2a)/σ . The criterion
Weg ≡ ρ 2U2(2a)/σ < 0.4 corresponds to the point where the inertia force of
the surrounding gas reaches about 10% of the surface tension force. Ranz
(1956) suggested that this would mark the beginning of the first wind-induced
breakup regime, where the effects of the ambient gas are no longer negligible.
Numerical results of Sterling & Sleicher (1975) indicate that the maximum in
the jet breakup length (see Figure 2) occurs when WeL = 1.2 + 3.41 Z 10.9 , where
Z1 ≡ We0.5L /ReL, ReL ≡ U(2a)/ν 1. This could also indicate the importance of
aerodynamic effects, so that the criteria for Rayleigh breakup (see Figure 1a)
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

would be
We L > 8 and Weg < 0.4 or 1.2 + 3.41 Z 10.9 . (5)
by Princeton University Library on 10/24/05. For personal use only.

Note, however, that nozzle turbulence and other flow effects are not included
in Equation 5. Ranz (1956) argued that the gas inertia force is of the same order
as the surface tension force when Weg = 13. This could serve as a definition
of the end of the first wind-induced regime (see Figure 1b), which then occurs
when
1.2 + 3.41 Z 10.9 < Weg < 13 (6)
In this case, Weg > 13 marks the onset of the second wind-induced regime,
where the interaction with the surrounding gas starts to become dominant.
Miesse (1955) suggested the criterion Weg > 40.3 to predict the onset of the
atomization regime, the point at which breakup appears to start at the nozzle
exit (see Figure 1d). Thus, the criteria for breakup in the second wind-induced
regime are
13 < Weg < 40.3 (7)
In the second wind-induced regime, the breakup starts some distance down-
stream of the nozzle exit, and a smooth unbroken section of the jet is visible
downstream of the nozzle exit (Figure 1c).
As mentioned previously, no account is made of nozzle internal flow effects
in the above correlations. To address this shortcoming, Reitz (1978) assumed
that atomization corresponds to a critical value of the breakup length/nozzle
diameter ratio. With this assumption, the onset of atomization is predicted to
occur when
ρ2 /ρ1 > K f (T )−2 (8)
In this case, the parameter K was obtained from experiments on atomizing jets
and was found to be a function of the nozzle geometry, where
K = (0.53[3.0 + (`/2a)]1/2 − 1.15)/744 (9)
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

94 LIN & REITZ

and `/2a is the nozzle length-to-diameter ratio. K empirically accounts for the
effect of initial disturbances in the flow caused by nozzle internal flow phenom-
ena such as turbulence, cavitation, and flow separation. Equation 9 includes
the effect of liquid viscosity and nozzle internal flows, and it predicts that at-
omization is favored at high gas densities and for sharp inlet edge nozzles, with
small length-to-diameter ratio. These trends agree with experiments reported
in the literature (Hiroyasu et al 1991, Reitz 1978, Reitz & Bracco 1979).
When injection takes place into a coflowing gas, additional breakup regimes
are observed, as described by Chigier & Reitz (1996). This situation is of
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

much practical interest and it is frequently used in air-blast coaxial atomizers


to improve the quality of atomization and to maintain it over a wide range
of liquid flow rates. High gas velocities (up to sonic) are generated by high-
by Princeton University Library on 10/24/05. For personal use only.

pressure gas flows passing through annular orifices surrounding the liquid jet.
The high coflowing gas velocity transmits momentum to the liquid interface.
Large-scale eddy structures in the gas flow impact upon the liquid jet, causing
stretching, destabilization, and flapping of the liquid jet. Eroglu et al (1991)
measured breakup lengths of round liquid jets in annular coaxial air streams
and found that the breakup length decreases with increasing Weber number and
increases with increasing liquid jet Reynolds number according to the relation
L/2a = 0.5We−0.4
L Re0.6
g (10)
where L is the liquid intact length, a is the central tube inner radius, and the
Weber and Reynolds numbers are based on the relative velocity between the
gas and the liquid.
Jet breakup in coaxial flows is highly unsteady, and unstable liquid structures
are observed to disintegrate in a time-varying, bursting manner. Farago &
Chigier (1992) refer to these as pulsating and super-pulsating breakup processes.
At high air-flow rates, the unstable liquid cylindrical jet undergoes a flapping
motion and can be transformed into a curling liquid sheet. The sheet becomes
stretched into a membrane bounded by thicker rims, which finally burst into
ligaments and drops of various sizes.
Farago & Chigier (1992) classified coaxial jet disintegration into three main
categories: (a) Rayleigh-type breakup where the mean drop diameter is of the
order of the jet diameter—both axisymmetric breakup (for Weg < 15) and non-
axisymmetric breakup patterns (for 15 < Weg < 25) were observed; (b) jet
disintegration via the stretched-sheet mechanism, which produces membrane-
type ligaments (25 < Weg < 70)—in this case, the diameter of the drops formed
is considerably smaller than the diameter of the jet; and (c) jet disintegration
via fiber-type ligaments (100 < Weg < 500)—at even higher air-flow rates,
fibers are formed that peel off the liquid-gas interface. This breakup mecha-
nism resembles the short-wavelength breakup mechanism of jets in the second
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 95

wind-induced and atomization regimes mentioned above. The atomization be-


gins with the unstable growth of short-wavelength waves, the formation of
fibers, and their peeling off from the main liquid core. The fibers break into
droplets by the nonaxisymmetric Rayleigh-type jet disintegration mode. Again,
the drop diameter is much smaller than the jet diameter.

3. BREAKUP MECHANISMS
Jet Instability
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

The regimes of jet breakup have been delineated above with correlations in
terms of relevant parameters. The results of recent works based on the the-
ory of absolute and convective instability of liquid jets enable us to elucidate
by Princeton University Library on 10/24/05. For personal use only.

the different physical mechanisms responsible for the jet breakup in the various
regimes. In the aerodynamic theory of spontaneous jet breakup without external
excitation, it is assumed that the onset of breakup is caused by the amplification
of natural disturbances in a jet. Any arbitrary form of disturbance can be con-
structed by superposition of all Fourier components. Each Fourier component
has the form A exp[ikx + ωt], where A is the wave amplitude, k = kr + iki is
the complex wave number whose real and imaginary parts give, respectively,
the number of waves over a distance 2π and the exponential spatial growth rate
per unit distance in the axial x-direction, and ω = ωr + iωi is the complex
wave frequency, the real and imaginary part of which give, respectively, the
exponential temporal growth rate and the frequency of the Fourier wave.
Not all the Fourier components are capable of extracting energy from the jet
system and amplifying, however. The Fourier components must have special
values of k and ω, which depend on specific characteristics of the jet system,
in order to grow from initially infinitesimal amplitudes. Mathematically, (k, ω)
is determined as the eigenvalue of a linear system containing relevant flow
parameters. The eigenvalues or the characteristic values are so determined that
the condition of the existence of a nontrivial solution of the system is satisfied.
This condition is the so-called characteristic equation or dispersion relation. In
the linear aerodynamic theory of jet instability, the finite amplitude disturbances
introduced outside or inside the nozzle by the various means mentioned above
are excluded from consideration.
Absolute Instability and Formability of a Jet
In the pioneering works cited above, the liquid jet is considered to be infinitely
long and k is assumed to be real. Thus the disturbance must grow or decay
everywhere in space at the same time rate ωr. However, Keller et al (1972)
noted that the disturbance initiating from the nozzle tip actually grows in space
as it is swept downstream to break up the jet into drops, leaving a section of jet
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

96 LIN & REITZ

intact near the nozzle tip. They set k to be complex and allow the disturbance
to grow in space as well as in time in a semi-infinite weightless inviscid jet in
a vacuum. They found that Rayleigh’s results are relevant only in the case of
large Weber number, WeL (WeL = ρ LU 22a/σ is based on the liquid density).
They also showed that in the limit of WeL → ∞, the spatial growth rate kI can
be inferred from the temporal growth rate ωr by the relation ki = ± ωr + O
(1/WeL), while the disturbance travels at the jet velocity. For Weber number less
than the order of one, they found a new mode of faster-growing disturbances
whose wavelengths are so long that they may not be actually observable.
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

Using the theory of absolute and convective instability (Briggs 1964, Bers
1983), Leib & Goldstein (1986b) showed that the new mode actually corre-
sponds to absolute instability arising from a saddle-point singularity in the
by Princeton University Library on 10/24/05. For personal use only.

characteristic equation. The unstable disturbances in an absolutely unstable jet


must propagate in both upstream and downstream directions. Thus, the unsta-
ble disturbances expand in space over the course of time. This contrasts with
what is observed in a Rayleigh jet, wherein unstable disturbances grow over
time as they are convected in a wave packet in the downstream direction with
the group velocity dωi/dkr (Lighthill 1987, Mei 1989). For a brief introduction
to the theory of absolute and convective instability in the context of jet breakup,
see the recent work of Lin (1996).
The critical Weber number WeLc, below which an inviscid jet under weight-
less condition in vacuum is absolutely unstable, and above which the jet is
convectively unstable, was found by Leib & Goldstein (1986b) to be π . When
the viscosity of the jet is taken into account, the critical Weber number depends
on the Reynolds number ReL = U(2a)/ν 1, where ν 1 is the liquid viscosity
as shown in Figure 3 with the curve Q = 0, which was obtained by Leib &
Goldstein (1986a). The other two curves for nonvanishing values of Q ≡ ρ 2/ρ 1
were obtained by Lin & Lian (1989), who were motivated to find out whether
the absolute instability discovered by Leib & Goldstein is physical or mathe-
matical, arising from the neglected ambient gas effect. It turns out that the effect
of the gas density is to increase WeLc (Figure 1). Thus, the gas density promotes
absolute instability in the sense that the given jet that is convectively unstable
may be made absolutely unstable by increasing the ambient gas density.
The absolute instability cannot be suppressed by either the gas compress-
ibility (Zhou & Lin 1992a,b, Li & Kelly 1992) or by the gas viscosity (Lin
& Lian 1993). Absolute instability is a real physical phenomenon, at least in
the absence of gravity, which is neglected in the theories. However, jet ab-
solute instability under weightless conditions has not yet been reported in the
literature. The delineation of transition from convective to absolute instabil-
ity in the absence of gravity is yet to be completed (Honohan 1995, Vihinen
1996). In such a delineation, the critical Weber number is a function of ReL,
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org
by Princeton University Library on 10/24/05. For personal use only.
P1: ARS/kja
November 25, 1997
P2: HCS/plb
11:20
QC: MBL/agr
T1: MBL
Annual Reviews
AR049-04

JET BREAKUP
97

Figure 3 Critical Weber number as a function of the Reynolds number. The jet is absolutely unstable below each curve of constant Q.
The jet is convectively unstable in the rest of the parameter space.
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

98 LIN & REITZ

Q, and N ≡ ν 2/ν 1, where ν 2 is the viscosity of the surrounding gas. Exist-


ing theoretical results show that for a given set of ReL, Q, and N, a liquid jet
may be made absolutely unstable by reducing the Weber number to be below
WeLc. The theoretical prediction that the unstable disturbance must propagate
in both downstream and upstream directions when the jet velocity is smaller
than that corresponding to WeLc signifies that absolute instability occurs when
the jet inertia is not sufficiently large to carry downstream all of the unstable
disturbances that derive their energy from the surface tension. Thus, surface
tension remains the source of instability. Part of the unstable disturbances
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

will propagate back to the nozzle tip to interrupt the formation of a jet of any
length. It is likely that the phenomenon of absolute instability also exists in a
jet in the presence of gravity, because the physical mechanism is unlikely to
by Princeton University Library on 10/24/05. For personal use only.

be altered significantly by the gravity-induced variation in the thickness and


the velocity of the jet along its axis. Thus, the transition from absolute to
convective instability signifies the beginning of the formability of a liquid jet.
The parameter range in which a liquid jet cannot be formed may be termed the
absolutely unstable regime. The origin of nonformability of a jet is the surface
tension.

Capillary Pinching with Wind Assistance


The jet breakup in the Rayleigh, the first wind-induced, the second wind-
induced, and the atomization regimes defined in the previous section are all
the manifestations of convective instability. As mentioned earlier, the unstable
disturbances amplify in time as they are convected in a group in the down-
stream direction with the group velocity. However, the physical mechanism of
breakup in the second wind-induced and atomization regimes is fundamentally
different from that in the other regimes. Neglecting the viscosity of gas, Lin &
Creighton (1990) calculated from the Navier-Stokes equations the mechanical
energy budget of a liquid jet in a parameter range of convective instability. The
stability analysis in this range was completed earlier by Lin & Lian (1990).
They expressed the total time rate of changes of the disturbance kinetic en-
ergy in a controlled volume of the jet, over a wavelength of the most amplified
disturbance, as the sum of the rate of work done by various relevant forces, i.e.

E = Pg + P` + S + V + D, (11)

where Pg is the rate of work done by the gas pressure fluctuation at the liquid-gas
interface, P` is the rate of work done by the liquid pressure fluctuation at the
inlet and outlet of the control volume, S is the rate of work done by the surface
tension, V is the rate of work done by the liquid viscous stress, and D is the rate
of viscous dissipation of mechanical energy.
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 99

Some typical values of ( pg, p` , s, v, d) ≡ (Pg, P` , S, V, D)/E for various


breakup regimes are shown in Table 1. The first four rows in this table belong
to the Rayleigh and the first wind-induced breakup regimes. The last four rows
belong to the second wind-induced and the atomization regimes. The values
kr = 2πa/λm in the fourth column are the wave numbers corresponding to
the wavelength, λm, of the most amplified disturbance predicted by the linear
theory for the flow parameters specified in the first three columns. Q = 0.0013
corresponds to the case of a water jet in air under one atmosphere. The high
Reynolds number jet at low gas density depicted in the fourth row is closest
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

to the idealized Rayleigh jet. The presence of the low density gas increases
only slightly the most amplified wave number, 0.696, predicted by Rayleigh.
Moreover, s dominates all other work terms. This is a classical case of Rayleigh
by Princeton University Library on 10/24/05. For personal use only.

breakup by capillary pinching, which produces drops of a diameter comparable


to the jet diameter.
Capillary pinching remains the mechanism of breakup for the low Reynolds
number jet depicted in the first row (Table 1). Note that the value of kr is
almost four times smaller than that predicted by Rayleigh. As the drop size
is inversely proportional to kr, the liquid viscosity tends to increase the drop
size considerably. This regime, which is not shown in Figure 1, may be termed
the Weber-Chandrasekhar regime to emphasize the important role of liquid
viscosity.
With Q kept at 0.0013 in rows 2 and 3, pg increases with increasing ReL
almost to the same order of magnitude as s. All other work terms remain
insignificant. Thus, as the relative speed of gas-to-liquid (wind speed) increases,
the gas pressure fluctuation assists significantly the capillary force to break up
the liquid jet. Nevertheless the capillary force remains dominant over the
wind force. Comparing the values of kr in these two rows with that of the
Rayleigh jet, it is seen that the drop size in this first wind-induced breakup
regime may be slightly larger or smaller than that in the Rayleigh regime,

Table 1 Energy budget in jet breakup

ReL 103/WeL Q × 103 kr s pg p` v d

2 1.25 1.3 0.1669 117.5 3.5 −0.2 0.2 −21.6


102 1.25 1.3 0.5725 97.7 26.2 −4.9 1.6 −20.6
4 × 102 2.50 1.3 0.7701 65.5 41.8 0.0 0.1 −7.4
4 × 104 1.25 0.1 0.7088 96.2 3.5 0.4 0.0 −0.1
11016 0.882(−2) 1.3 35.417 −296.0 547.7 −2.1 −0.1 −149.5
36720 0.882(−2) 1.3 40.051 −215.1 351.7 −2.0 0.1 −34.7
67411 0.882(−2) 1.3 41.580 −210.4 331.6 −1.9 0.0 −19.3
116122 0.802(−2) 1.3 42.368 −214.7 332.1 −1.0 −0.2 −16.2
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

100 LIN & REITZ

depending on the flow parameters (see Figure 1b). Since the surface tension is
mainly responsible for the breakup in this regime, and the gas inertia force only
assists in the breakup, rather than inducing the breakup, it is probably more
appropriate to call this regime the wind assisted breakup regime instead of the
first wind-induced breakup regime. The second wind-induced breakup regime
is genuinely wind-induced, as is explained below.

Interfacial Stress Against Surface Tension


For the parameter range specified in the last four rows (Table 1), kr is found
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

to be more than one order of magnitude larger than in the previous four cases.
Thus, the drop radii produced in these parameter ranges are much smaller than
the jet radius (see Figure 1c and d ). In contrast to the previous four cases, the
by Princeton University Library on 10/24/05. For personal use only.

pressure work-term dominates the surface tension work-term in the last four
rows. In fact the surface tension term is negative. That is to say, the surface
tension acts against the formation of small droplets generated by the interfacial
pressure fluctuation in the second wind-induced and atomization regimes. Part
of the kinetic energy in a jet is converted through the pressure work-term to the
surface energy in the droplets. Thus the second wind-induced and atomization
regimes are genuinely wind-induced.
It is seen (Table 1) that while the atomization and second wind-induced
regimes exist in the parameter range WeL  Q, the rest of the regimes exist for
WeL ≤ 1. It has been shown (Kang & Lin 1987) that the unstable disturbances
in the atomization regime scale with the gas capillary length c = σ /ρ 2U2. The
condition WeL  Q implies that c is much smaller than the jet diameter (Lin &
Lian 1990). The interfacial shear stress fluctuation will augment the pressure
fluctuation in the breakup process, if the gas viscosity is considered.
It should be remembered that the linear stability theory is not capable of
differentiating between the second wind-induced and the atomization regimes.
The linear theory can only predict the onset of instability, which produces in-
terfacial waves of different length scales depending on the parameters. The
nonlinear processes of pinching off a small droplet from the interface, subse-
quent to the onset, and the continuous generation of droplets from the receding
interface toward the core of a jet are involved in reaching the atomization state
(see Figure 1d). However, the physical mechanism involved in the initial stage
of the second wind-induced and the atomization regimes may be the same.

4. DISCUSSION
Interfacial Shear Layer
A serious defect common to all of the above reviewed works is the lack of
a rigorous treatment of the effect of the gas viscosity. Sterling & Sleicher
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 101

(1975) assumed with Benjamin (1959) that the Kelvin-Helmholtz model can
be applied locally along the interfacial wave with an arbitrary correction factor
that—because of the viscous effect—is used to reduce the pressure distribution
predicted by the Kelvin-Helmholtz model. Thus the possibility of generation
of droplets by shear waves is missed. Lin & Lian (1990) modeled the liquid-gas
interfacial boundary layer with the boundary layer over a wavy solid surface in
order to estimate the interfacial shear effect. This is not satisfactory because a
fluid-fluid interfacial shear layer is fundamentally different from that of a solid-
fluid boundary layer. The effect of gas viscosity was rigorously analyzed by
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

Lin & Ibrahim (1990) with temporal theory, and by Lin & Lian (1993) with the
theory of absolute and convective instability for a viscous liquid jet in a viscous
gas in a vertical circular pipe. The basic flow is an exact solution of the Navier-
by Princeton University Library on 10/24/05. For personal use only.

Stokes equation. Unfortunately, the numerical results for the case of very strong
interfacial shear were not sufficiently accurate to allow the drawing of definitive
conclusions on the effect of the gas viscosity on the atomization process. The
relative importance of the shear stress fluctuation to the pressure fluctuation in
the atomization process in various parameter ranges remains unknown.

Nonaxisymmetric Disturbances
The above description is based on works that assume that axisymmetric distur-
bances are more unstable than asymmetric ones. Rayleigh was able to prove
that asymmetric temporal disturbances in an inviscid jet are all stable. Tem-
porally stable waves are also convectively stable evanescent waves (Huerre &
Monkewitz 1990). For ReL = O (10), Lin & Webb (1994) showed that the
asymmetric disturbances are evanescent waves in the parameter range 10−4 ≤
Q ≤ 10−2, 10 ≤ We ≤ 103. Yang (1992) demonstrated that temporally grow-
ing asymmetric long-wavelength disturbances may become dominant when
the Weber number of an inviscid jet in an inviscid gas is in the atomization
regime. However, the viscosity of the liquid tends to bring down the temporal
growth rate of nonaxisymmetric disturbances (Avital 1995), and the maximum
temporal growth rates of axisymmetric disturbances remain higher than those
of asymmetric ones except when the jet is almost inviscid, for example when
ReL = 105 and WeL = 104 (Li 1995).
A similar conclusion was reached by EA Ibrahim (personal communication),
who investigated convectively unstable asymmetric disturbances in a viscous
jet emanating into an inviscid gas, when WeL and ReL are of order 103 and
higher. There is also experimental evidence of asymmetric disturbances in
the atomization regime mentioned above (Meister & Scheele 1969, Taylor &
Hoyt 1983, Eroglu et al 1991). However, the appearance of non-axisymmetric
disturbances may also be caused by the secondary instability after the onset of
instability from axisymmetric disturbances. In the Rayleigh, the wind-induced,
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

102 LIN & REITZ

and the wind-assisted regimes, the asymmetric disturbances may be brought out
prominently by the swirl in the liquid jet (Ponstein 1959, Kang & Lin 1989) or
less prominently by the swirl in the gas (Lin & Lian 1990).

Success and Shortcoming of Linear Theory


Despite these shortcomings, the linear stability analysis started by Rayleigh pro-
vides a qualitative description of the physical mechanisms involved in various
regimes of jet breakup. The linear theories have even enjoyed some reason-
able semiquantitative comparisons with experiments in the atomization regime
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

(Kang & Lin 1987, Reitz & Bracco 1982). These comparisons include the
intact length, spray angle, and droplet size, which scales with the gas capillary
length σ /ρ 2U2. The quantities that have just been mentioned are the products
by Princeton University Library on 10/24/05. For personal use only.

of highly nonlinear processes. The reason such a reasonable comparison based


on linear theory is possible is probably due to the fact that the physical mech-
anisms at work in various breakup regimes are already basically determined at
the onset of instability. The nonlinear evolution subsequent to the onset only
modifies quantitatively the physical mechanisms.
On the other hand, the excellent quantitative comparison between the theory
of Rayleigh and the experiments of Goedde & Yuen (1970) and Donnelly &
Glaberson (1966) is probably fortuitous. While Rayleigh neglected the exis-
tence of the ambient gas, the liquid viscosity, and gravity, none of these are
neglected in experiments. Moreover, the Rayleigh jet breakup is due to capil-
lary pinching, and yet the experiments do not seem to be sensitive to the Weber
number. A complete delineation of where each regime should start and end
in the parameter space (Q, N, Re, We, Fr) may be made within the framework
of linear theory only if the various effects of nozzle flows on jet instability
are known or assumed known. Here, the Froude number, Fr = U 2/g(2a),
represents the ratio of inertial to gravitational force.

5. SUMMARY AND UNRESOLVED


SCIENTIFIC ISSUES
This study discusses various regimes of breakup of liquid jets injected into
both stagnant and coflowing gases. Available criteria for the transition between
the regimes are reviewed. The physical mechanisms at work in the different
breakup regimes are described. The influence of nozzle internal flow effects is
shown to be important, but these effects are only included empirically in current
wave-stability theories.
A useful area for future research would be the development of fundamentally
based models that account for the effect of nozzle internal flows on the liquid
breakup process. In addition, current breakup models need to be extended
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 103

to account for the nonlinear effects of liquid distortion, membrane formation,


and stretching on the atomization process. These phenomena are especially
important in liquid injections in a high-momentum coflowing gas.
The effect of liquid-gas viscous shear layers on the onset of instabilities,
which leads to various regimes of the jet breakup, remains to be rigorously
analyzed and tested. This knowledge is not only important for applications
involving jet breakup, it will also advance our scientific understanding of var-
ious processes in nature as well as in industries. The impact of the stick-slip
condition experienced by the jet liquid exiting the nozzle tip and the spatial
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

development of the liquid-gas interfacial shear layer (before the jet flow is fully
developed) on the receptivity of the jet to instability need to be investigated. The
effect of gravity in the absolute instability regime and the Weber-Chandrasekhar
by Princeton University Library on 10/24/05. For personal use only.

regimes, where the Froude number is small, remains to be elucidated. In the


atomization regime, Fr is so large that the gravitational effect may not be signif-
icant. A complete delineation of the jet breakup regimes in the entire parameter
space (Q, N, Re, We, Fr), even in the framework of linear theory, has not yet
appeared. The nonlinear studies, which take into account the finite amplitude
disturbances originated in the nozzle, and which elucidate the nonlinear process
subsequent to the onset of unstable waves of various-length scales, will con-
tribute to the understanding of the formation of sprays and drops from a liquid
jet.
ACKNOWLEDGMENTS
Support for SP Lin was provided in part by Army Research Office Grant
DAAL03-89-K-0179 and NASA Grant NAG3-1891. Support for R Reitz was
provided by the Army Research Office Contract DAAL03-86-K-0174.

Visit the Annual Reviews home page at


http://www.AnnualReviews.org.

Literature Cited

Avital E. 1995. Asymmetric instability of a vis- ity of a liquid jet. In Hydrodynamic and Hy-
cid capillary jet in an inviscid media. Phys. dromagnetic Stability, pp. 537–42. Oxford:
Fluids. 7:1162–64 Oxford Univ. Press. 652 pp.
Benjamin TB. 1959. Shearing flow over a wavy Chehroudi B, Bracco FV. 1985. On the intact
boundary. J. Fluid Mech. 6:161–205 core of full cone sprays. Soc. Automot. Eng.
Bers A. 1983. Space-time evolution of plasma Tech. Pap. 850126
instabilities—absolute and convective. In Chigier N, Reitz RD. 1996. Regimes of jet
Handbook of Plasma Physics, ed. M. Rosen- breakup and breakup mechanisms (physical
bluth, 1:452–516 Amsterdam: North Hol- aspects). In Recent Advances in Spray Com-
land bustion: Spray Atomization and Drop Burn-
Briggs RJ. 1964. Electron Stream Interaction ing Phenomena, ed. KK Kuo, 1:109–35. Re-
with Plasmas. Cambridge: MIT ston: AIAA
Chandrasekhar S. 1961. The capillary instabil- Dan T, Yamamoto T, Senda J, Fujimoto H. 1997.
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

104 LIN & REITZ

Effect of nozzle configurations for character- tion: Spray Atomization and Drop Burning
istics of non-reacting diesel fuel sprays. Soc. Phenomena, ed. KK Kuo, 1:137–60. Reston,
Automot. Eng. Tech. Pap. 970355 VA: AIAA
Donnelly RJ, Glaberson W. 1996. Experiments Lin SP, Creighton B. 1990. Energy budget in
on the capillary instability of a liquid jet. atomization. J. Aerosol. Sci. and Technol.
Proc. R. Soc. London Ser. A. 290:547–56 12:630–36
Eroglu H, Chigier N, Farago Z. 1991. Coaxial Lin SP, Ibrahim EA. 1990. Instability of a vis-
atomizer liquid intact lengths. Phys. Fluids cous liquid jet surrounded by a viscous gas in
A. 3:303–8 a pipe. J. Fluid Mech. 218:641–58
Farago Z, Chigier N. 1992. Morphological clas- Lin SP, Kang DJ. 1987. Atomization of a liquid
sification of disintegration of round liquid jets jet. Phys. Fluids. 30:2000–6
in a coaxial air stream. At. Sprays. 2:137–54 Lin SP, Lian ZW. 1989. Absolute instability in
Goedde EF, Yuen MC. 1970. Experiments on a gas. Phys. Fluids A. 1:490–93
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

liquid jet instability. J. Fluid Mech. 40:495– Lin SP, Lian ZW. 1990. Mechanism of atomiza-
512 tion. AIAA J. 28:120–26
Grant RP, Middleman S. 1966. Newtonian jet Lin SP, Lian ZW. 1993. Absolute and convective
stability. A.I.Ch.E. J. 12:699–78 instability of a viscous liquid jet surrounded
by Princeton University Library on 10/24/05. For personal use only.

Gulder OL, Smallwood GJ, Snelling DR. 1994. by a viscous gas in a vertical pipe. Phys. Flu-
Internal structure of transient full-one dense ids A. 5:771–73
diesel sprays. Int. Symp. COMODIA, pp. Lin SP, Webb RD. 1994. Nonaxisymmetric
355–60 evanescent waves in a viscous liquid jet. Phys.
Haenlein A. 1932. Uber den Zerfall eines Flus- Fluids. 6:2545–47
sigkeitsstrahls (On the disruption of a liquid McCarthy MJ, Malloy NA. 1974. Review of sta-
jet). NACA TM Report. 659 bility of liquid jets and the influence of nozzle
Hiroyasu H, Arai M, Shimizu M. 1991. Break- design Chem. Eng. J. 7:1–20
up length of a liquid jet and internal flow in a Mei CC. 1994. Applied Dynamics of Ocean Sur-
nozzle. Proc. ICLASS-91, Pap. 26 face Waves. Hong Kong: World Sci. 740 pp.
Honohan A. 1995. Experimental measurement Meister BJ, Scheele GF. 1969a. Drop formation
of the spatial instability of a viscous liquid jet from cylindrical jets in immiscible liquid sys-
at microgravity. MS thesis. Clarkson Univ., tems. AIChE J. 15:700–6
Potsdam, NY. 95 pp. Meister BJ, Scheele GF. 1969b. Prediction of jet
Huerre P, Monkewitz PA. 1990. Local and length in immiscible liquid system. AIChE J.
global instabilities in spatially developing 15:689–99
flows. Annu. Rev. Fluid Mech. 22:473–537 Miesse CC. 1955. Correlation of experimental
Kang DJ, Lin SP. 1989. Breakup of swirling liq- data on the disintegration of liquid jets. Ind.
uid jets. Int. J. Eng. Fluid Mech. 2:47–62 Eng. Chem. 47:1690–95
Keller JB, Rubinow SI, Tu YO. 1972. Spatial Plateau J. 1873. Statique Experimentale et The-
instability of a jet. Phys. Fluids. 16:2052–55 orique des Liquids Soumie aux Seules Forces
Leib SJ, Goldstein ME. 1986a. Convective and Moleculaire, vols. 1, 2. Paris: Cauthier Vil-
absolute instability of a viscous liquid jet. lars. 450 pp. 495 pp.
Phys. Fluids. 29:952–54 Ponstein P. 1959. Instability of rotating cylin-
Leib SJ, Goldstein ME. 1986b. The generation drical jets. Appl. Sci. Res. 8:425–57
of capillary instability on a liquid jet. J. Fluid Ranz WE. 1956. On sprays and spraying. Dep.
Mech. 168:479–500 Eng. Res., Penn State Univ. Bull. 65. 53 pp.
Leroux S, Dumouchel C, Ledoux M. 1996. The Rayleigh L. 1879a. On the capillary phe-
stability curve of Newtonian liquid jets. At. nomenon of jets. Proc. R. Soc. London.
Sprays 6:623–47 29:71–97
Li HS, Kelly RE. 1992. The instability of a liq- Rayleigh L. 1879b. On the instability of jets.
uid jet in a compressible air stream. Phys. Proc. London Math. Soc. 10:4–13
Fluids A. 4:2162–68 Rayleigh L. 1892a. On the instability of a cylin-
Li HS, Kelly RE. 1993. On the transfer of energy der of viscous liquid under capillary force.
to an unstable liquid jet in a coflowing com- Phil. Mag. 34:145–54
pressible airstream. Phys. Fluids A. 5:1273– Rayleigh L. 1892b. On the instability of cylin-
74 drical fluid surfaces. Phil. Mag. 34:177–
Li X. 1995. Mechanism of atomization of a liq- 80
uid jet. At. Sprays 5:89–105 Reitz RD. 1978. Atomization and other breakup
Lighthill J. 1987. Waves in Fluids, pp. 239–43 regimes of a liquid jet. PhD thesis. Princeton
Cambridge: Cambridge Univ. Press. 504 pp. Univ., Princeton, NJ. 231 pp.
Lin SP. 1996. Regimes of jet breakup and Reitz RD. 1987. Modeling atomization pro-
breakup mechanisms (Mathematical as- cesses in high-pressure vaporizing sprays.
pects). In Recent Advances in Spray Combus- Atom. Spray Technol. 3:309–37
P1: ARS/kja P2: HCS/plb QC: MBL/agr T1: MBL
November 25, 1997 11:20 Annual Reviews AR049-04

JET BREAKUP 105

Reitz RD, Bracco FV. 1979. On the dependence graphy—techniques and methods. Exp. Flu-
of the spray angle and other spray parameters ids. 1:113–20
on nozzle design and operating conditions. Tomotika S. 1935. On the instability of a cylin-
Soc. Automot. Eng. Tech. Pap. 790494 drical thread of a viscous liquid surrounded
Reitz RD, Bracco FV. 1982. Mechanism of by another viscous fluid. Proc. R. Soc. Lon-
atomization of a liquid jet. Phys. Fluids. don Ser. A. 150:322–37
25:1730–42 Vihinen I. 1996. Absolute and convective in-
Reitz RD, Bracco FV. 1986. Mechanisms of stability of a liquid jet in microgravity. MS
breakup of round liquid jets. The Encyclope- thesis. Clarkson Univ., Potsdam, NY. 104 pp.
dia of Fluid Mechanics, ed. N Cheremisnoff, Weber CZ. 1931. Zum Zerfall eines Flus-
3:233–49. Houston: Gulf sigkeitsstrahles. Math. Mech. 11:136–54
Sirignano WA. 1993. Fluid dynamics of sprays. Wu PK, Tseng LK, Faeth GM. 1992. Primary
J. Fluid Eng. 115:345–78 breakup in gas/liquid mixing layers for tur-
Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

Sterling AM, Sleicher CA. 1975. The instabil- bulent liquids. At. Sprays 2:295–318
ity of capillary jets. J. Fluid Mech. 68:477– Yang HQ. 1992. Asymmetric instability of a liq-
95 uid jet. Phys. Fluids A. 4:681–89
Taylor GI. 1962. Generation of ripples by wind Zhou ZW, Lin SP. 1992a. Absolute and convec-
by Princeton University Library on 10/24/05. For personal use only.

blowing over viscous fluids. In The Scien- tive instability of a compressible jet. Phys.
tific Papers of G.I. Taylor, ed. GK Batche- Fluids A. 4:277–82
lor, 3:244–54. Cambridge: Cambridge Univ. Zhou ZW, Lin SP. 1992b. Effects of compress-
Press ibility on the atomization of liquid jets. J.
Taylor JJ, Hoyt JW. 1983. Water jet photo- Propuls. Power. 8:736–40
Annual Review of Fluid Mechanics
Volume 30, 1998

CONTENTS
Lewis Fry Richardson and His Contributions to Mathematics,
0
Meteorology, and Models of Conflict, J.C.R. Hunt
Aircraft Laminar Flow Control , Ronald D. Joslin 1

Vortex Dynamics in Turbulence, D. I. Pullin, P. G. Saffman 31

Interaction Between Porous Media and Wave Motion, A. T. Chwang, A.


53
T. Chan

Drop and Spray Formation from a Liquid Jet, S. P. Lin, R. D. Reitz 85

Airplane Trailing Vortices, Philippe R. Spalart 107


Annu. Rev. Fluid. Mech. 1998.30:85-105. Downloaded from arjournals.annualreviews.org

Diffuse-Interface Methods in Fluid Mechanics, D. M. Anderson, G. B.


139
McFadden, A. A. Wheeler
by Princeton University Library on 10/24/05. For personal use only.

Turbulence in Astrophysics: Stars, V. M. Canuto, J. Christensen-


167
Dalsgaard
Vortex-Body Interactions, Donald Rockwell 199

Nonintrusive Measurements for High-Speed, Supersonic, and Hypersonic


231
Flows, J. P. Bonnet, D. Grésillon, J. P. Taran

Renormalization-Group Analysis of Turbulence, Leslie M. Smith, Stephen


275
L. Woodruff

Control of Turbulence, John Lumley, Peter Blossey 311

Lattice Boltzmann Method for Fluid Flows, Shiyi Chen, Gary D. Doolen 329

Boiling Heat Transfer, V. K. Dhir 365

Direct Simulation Monte Carlo--Recent Advances and Applications, E.S.


403
Oran, C.K. Oh, B.Z. Cybyk
Air-Water Gas Exchange, B. Jähne, H. Haußecker 443
Computational Hypersonic Rarefied Flows, M. S. Ivanov, S. F.
469
Gimelshein

Turbulent Flow Over Hills and Waves, S. E. Belcher, J. C. R. Hunt 507

Direct Numerical Simulation: A Tool in Turbulence Research, Parviz


539
Moin, Krishnan Mahesh

Micro-Electro-Mechanical-Systems (MEMS) and Fluid Flows, Chih-


579
Ming Ho, Yu-Chong Tai

Fluid Mechanics for Sailing Vessel Design, Jerome H. Milgram 613

Direct Numerical Simulation of Non-Premixed Turbulent Flames, Luc


655
Vervisch, Thierry Poinsot

You might also like