2011 - T G Theofanous - Aerobreakup of Newtonian and Viscoelastic Liquids
2011 - T G Theofanous - Aerobreakup of Newtonian and Viscoelastic Liquids
ANNUAL
REVIEWS Further Aerobreakup of Newtonian
Click here for quick links to
Annual Reviews content online,
including:
• Other articles in this volume
and Viscoelastic Liquids
• Top cited articles
• Top downloaded articles
• Our comprehensive search
T.G. Theofanous
Departments of Chemical Engineering and Mechanical Engineering, Center for Risk Studies
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
661
FL43CH27-Theofanous ARI 19 November 2010 16:35
1. INTRODUCTION
Unless noted otherwise, these are taken as standard atmospheric pressure/temperature conditions.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
Cp
Video 1 Video
0.5
0 Cp
–0.5 1
–1 0.8
–1.5 0.6
–2 0.4
0.2
0
–0.2
–0.4
–0.6
–3 –2 –1 0 1 2 3 4 5 –3 –2 –1 0 1 2 3 4 5
Diameter Diameter
Figure 1
Pressure field following a shock interaction with a rigid, stationary sphere in subsonic (left, M S = 1.17, M = 0.35, and Re g =
1.57 × 104 ) and supersonic flow (right, M S = 2.65, M = 1.26, and Re g = 2.46 × 105 ). (Figures with the “Video” icon contain
embedded videos. To view these videos, go to http://arjournals.annualreviews.org/loi/fluid and download the PDF version of this
article.)
Scheme 1
Key dimensionless groups, p is the pressure on the surface of the drop, p∞ the free stream gas pressure, Pd
the flow dynamic pressure, d the drop diameter, Cs the shock speed, C g− /C g the speed of sound in gas
upstream/behind the shock, ug free stream gas velocity, μg /μl the gas/liquid dynamic viscosity, ρ g /ρ l the
gas/liquid density, σ the surface tension coefficient, θ the polymer chain (Rouse) relaxation time, and ε̇ the
elongational strain rate. The subscript 0 refers to initial conditions/values.
662 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
Scheme 2
Key scaling parameters and nomenclature, where a is the drop acceleration, Pd the flow dynamic pressure, d
the cross-stream drop diameter, h the thickness of flattened drop, Cs the shock speed, k the wave number, ug
the free stream gas velocity, t the time after shock impact, Ct a parameter accounting for aerodynamic
pressure gradients, ν l the liquid kinematic viscosity, ρ g /ρ l the gas/liquid density, σ the surface tension
coefficient. The subscript 0 refers to initial conditions/values, the subscript c to a criticality condition, and
the superscript ∗ refers to the maximum growth rate.
An important measure of the intensity of the interaction is given by the gas-flow dynamic
pressure (Pd = 1/2ρg u 2g ), and as a prompt consequence of the interfacial response to the forces just
mentioned we have a strongly coupled problem. At low dynamic pressures, the principal effect in
this coupling results from the overall drop-shape changes. At high dynamic pressures, we have
the additional effect of rapidly escalating interfacial morphologies. Eventually, at high-enough Deformation:
a measure of a drop’s
dynamic pressures, these finer-scale features dominate. We refer to the shape variations of the
shape departure from
(remaining) coherent-liquid mass as deformation. These shape changes result from drop-internal spherical; expressed in
motions that yield mass redistribution. Finer aspects of interfacial topologies initiate with the terms of the cross-
growth of unstable, infinitesimally small fluctuations that eventually yield detachment from the stream diameter
mother drop and entrainment into the gas flow. These we refer to as instabilities. normalized by that of
the original drop
Both deformation and instabilities are resisted by liquid inertia, viscous stresses, elastic stresses
(if present), and the surface tension force. The outcome defines the onset of breakup (first criti- Rayleigh-Taylor
piercing (RTP): a
cality), the pattern of bodily loss of coherence (regime of breakup), and the resulting particle-size
drop-breakup regime
distribution. For conditions that exceed those of the first criticality, the breakup regime may vary. governed by the
There are multiple variations of such regimes, but as shown below, they all can be categorized Rayleigh-Taylor
into two major classes: Rayleigh-Taylor piercing (RTP) and shear-induced entrainment (SIE). instability; a flattened
Besides the appropriate definition of these two regimes and the respective criticality measures, we drop penetrated by
one or more unstable
are interested in elucidating the key physics involved, as well as the nature of the processes leading
waves
to the final particle-size distributions. The developments presented herein are based on recent
The energetic dissemination of liquid agents concerns the sudden release of a liquid mass from a high-speed vehicle
disrupted either by collision with another fast-moving object or by detonation of a high-explosive charge found
within (Babarsky & Theofanous 2010). The outcomes of interest are the characteristics (number density and size
distributions) of the final, equilibrium cloud of particles falling at terminal velocities in the atmosphere. The physics
of the process couple across multiple length scales, from the nano- and mesoscales (the fracture of polymer-chain
bonds, rupture of elastic filaments), to the micro- and millimeter scales (aerobreakup in the sense discussed in this
article), to the macroscale (primary, gross disaggregation of the liquid mass). Collective effects of the dispersion
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
at the macroscale define the local aerodynamics (within the cloud) that affect aerobreakup at the length scale of
individual drops. Shock waves form ahead of the decelerating cloud, and gas compressibility is important throughout.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
Effective-field modeling is faced with significant theoretical and practical issues (e.g., inviscid phase interactions, high
relative velocities, particle fluctuations) (Lhuillier & Theofanous 2010), and the numerical solution of such models is
likewise challenging (e.g., nonhyperbolic, nonconservative system of equations, widely varying flow speeds, the need
to capture shocks and contact discontinuities) (Liou & Theofanous 2010). Moreover, one needs the constitutive
description of aerobreakup as discussed in this article. The problem differs from that of atomization in injection or
industrial-processing systems in that it is not amenable to detailed observation at the practice-prototype scales.
work by the author (Theofanous et al. 2007, Theofanous & Li 2008); generic remarks about past
work made here are to be understood as not including these two works.
664 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
2000, Theofanous et al. 2006, Villermaux 2007, Gorokhovski & Herrmann 2008, Babarsky &
Theofanous 2010).
Whereas some references to employing elasticity in fuel injection systems can be found, as
Rayleigh-Taylor
well as a futile attempt (as a fuel antimisting agent) to reduce the extent of fire in aircraft crashes, (RT) instability: an
the principal practical context of thickening is in targeting bulk chemical weapons (and similarly interfacial instability
aircraft delivery of fire retardants and pesticides, albeit at much lower speeds). In such applications, found when a light
too fine of an atomization results in atmospheric dispersal far greater, and accordingly ground- phase is accelerated
into a heavy phase
deposition density levels much lower, than those sought. Liquid elasticity enters as the means
to suppress breakup (to shift the particle-size distributions to larger length scales) to an extent Kelvin-Helmholtz
instability: an
consistent with a particular application.
interfacial instability in
The elastic property is attained by dissolving high-molecular-weight polymers at concen-
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
6. The wide range of conditions and essential difficulties in empirical (field) testing (Babarsky streams
& Theofanous 2010) make basic understanding of the isolated drop problem indispensable. Shear-induced
Readiness for defensive risk management from chemical and biological weapons is receiving a entrainment with
ruptures (SIER): the
new urgency in recent years.
SIE breakup regime
found with elastic
liquids; rather than
1.3. A Canonical Problem in Interfacial Dynamics capillary breakup,
Elementary considerations suggest that the interface in the vicinity of the forward stagnation films and filaments are
subject to rupture
point is subject to Rayleigh-Taylor (RT) instability, whereas further away, toward the equator,
initiated by fractures
Kelvin-Helmholtz instability becomes increasingly important (Patel & Theofanous 1981). (Due in the underlying
to the extremely high acoustic impedance mismatch at the interface, Richtmeyer-Meshkov molecular-chain
instabilities are negligible.) The actual manifestation of these two classical instabilities is highly network
transient as they are strongly affected by the evolving overall drop deformation (and curvature
at the forward stagnation point) and the attendant changes in acceleration and relative velocity.
The interfacial curvature (Krechentnikov 2009) and the acceleration affect the propensity for
RT instability. The deformation alters the shear-force patterns at the interface, and thereby
the mean flow development in the outer liquid layers, which further couples to the interfacial
roughness induced by the Kelvin-Helmholtz instability. This mean flow induces straining around
the forward stagnation area, and this is known to suppress RT instability (Batchelor 1987,
Theofanous & Li 2008). Moreover, there is a third kind of coupling in that the aerodynamics
that drive these liquid flow processes are strongly coupled to the drop’s shape and interfacial
morphologies, and therefore to these processes themselves.
It is this series of complex and intertwined mechanics that determines whether the gas flow
can go through the drop (RTP) or not, in which case it will have to go around, in a continuing
shearing and entraining action (SIE) (Theofanous et al. 2004, 2007; Theofanous & Li 2008).
With increasing dynamic pressure (or Weber number), RTP and SIE are the first and second
major criticalities noted above. Both scale with viscosity as explained below, and so does the
effective intensity of the breakup process. Thus, rather than low/high dynamic pressure, or the
customary reference to the Weber number (which below we show to be irrelevant at Oh > 1),
it is principally important to refer to the relevant regime (RTP or SIE). If the rate of breakup is
significantly retarded by viscosity (as is the case with Oh 1), then with loss of relative velocity
(dynamic pressure) an SIE process may revert to RTP. With elastic liquids, breakup involves
ruptures of extending liquid threads, and the SIE becomes SIER.
Remarkably, and in contrast to all previous experimental work on the classical RT and Kelvin-
Helmholtz instabilities, in which the dominant waveforms are introduced as initial conditions
[starting with the pioneering work of Lewis (1949)], in the study of aerobreakup we begin with
a surface-tension-stabilized interface. Yet, as demonstrated below, the results are perfectly well
reproducible, even under the most apparently chaotic regimes. Thus our problem offers the next
Laser-induced
fluorescence (LIF): canonical step to the study of interfacial instabilities, with access to the multiscale complexity
a method of of shock-induced, multiphase mixing under competing physics. In the same vein, the problem
visualization of provides a basic foundation to test ab initio numerical simulations in this special class of interfacial
interfacial flows.
morphologies by light
emitted within the
liquid mass at
1.4. Experimental Methods
wavelengths distinct
from those of the laser Lane, Prewett, and Edwards, working in the Chemical Defense Experimental Establishment,
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Porton Down, United Kingdom, used a gas gun to produce atmospheric blasts (as discussed
in Taylor 1949), but most subsequent work, with single or string-out arrays of drops, employed
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
standard gas-driven shock tubes. Driver gases with a high speed of sound (i.e., helium) can enhance
the performance of a shock tube at the upper end of dynamic pressures. A special case is that of
Waldman et al. (1972), who used hydrogen explosions to achieve shock Mach numbers of up to
11. These are rather extreme conditions that likely introduced high temperatures and extraneous
liquid-flashing effects (Theofanous & Li 2008). Another special case is that of Theofanous et al.
(2004, 2007), who employed a pulse supersonic wind tunnel that produced Mach-3 flows at static
pressures in the range 2 × 104 to 10 Pa. Such high flow Mach numbers are not possible in a normal
shock tube, and the rarefied conditions open up new opportunities for observation and testing of
scaling laws. Obstructed flow geometries provide another way to tailor drop-flow interactions and
thereby provide a means to probe other regimes of deformation and breakup (Theofanous et al.
2007). In another testing setup, a drop is allowed to enter the flow field of a free gas jet, but this
introduces unwanted complications due to transit through the shear layer.
The principal task in measurement is that of visualization, and the apparently formidable chal-
lenge has been in achieving the necessary spatial/temporal resolutions. For low-viscosity liquids,
particle size and interfacial features with length scales of approximately tens of micrometers and
velocities of a few hundred ms−1 are common in the SIE regime; thus spatial resolutions of over 100
pixels per millimeter and exposure times of a few nanoseconds are basic requirements. Otherwise,
breakup at SIE is veiled in a dark smear well-known from shadowgraphs (with typical exposure
times in the microsecond range) taken from long ago (Engel 1958, Simpkins & Bales 1972) up
to recently ( Joseph et al. 2002). The RTP is easier to visualize, especially at the lower portion
(first RTP mode) of its range, as in the classical (photographic) images of Hanson et al. (1963)
depicting bag breakup. However, in shadowgraphy, as in photography, the significant interfa-
cial structures during the inception period remained unknown due to light reflection/refraction
processes that intensify with decreasing size and/or increasing multiplicity of the objects being
visualized. Moreover, these complications become prohibitive even at slightly higher dynamic
pressures.
The decisive developments began with the introduction of laser-induced fluorescence (LIF;
Theofanous & Li 2008) and evolved toward the requisite space/time resolutions mentioned above
over the past few years. The method allows the visualization of light emitted within the liquid
masses, thus highlighting interfacial features, and it is employed with appropriate filters to admit
only the wavelength of fluorescence. The quantities of dye needed are very small, and testing
confirms that all relevant liquid properties remain unaltered. Initial implementation of LIF was
with a Phantom 7 (Vision Research) 1-megapixel digital video camera, synchronized to a copper-
vapor laser (Oxford Instruments, model LS20-50), allowing resolutions of up to 50 pixels mm−1
and time exposures of 20 ns at rates up to 50 kHz. The subsequent introduction of a Phantom 12
666 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
camera, and installation of quartz optical-quality windows, allowed the resolution to be improved
to 200 pixels mm−1 , but it was still somewhat limited by the 1-megapixel format of the camera
and the longer-than-desired exposure time achievable with the copper-vapor laser. The most
Tri-butyl phosphate
recent implementation involves multiple single-frame, 11-megapixel digital cameras (allowing the (TBP): a water-like
200 pixels mm−1 , but over a much larger field of view), and an Nd:YAG laser system (New Wave liquid; a good solvent;
Research, model Tempest 10 Hz) capable of delivering a single ∼100-MW pulse of 5-ns duration. because of its
Triggered with a time accuracy of ∼1 μs, these single-frame shots provide the equivalent of 1- extremely low vapor
pressure, it finds
MHz video. It turns out that the multiscale nature of the phenomena investigated requires the
common use as a
synthetic use of all this instrumentation, including even shadowgraphy, which at the upper end of simulant of persistent
the space/time resolutions afforded attains a significant new utility. nerve agents (e.g., VX)
Data presented herein were obtained in the ASOS shock-tube facility [e.g., viscous liquids (T.G.
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
PSBMA: solution of
Theofanous, V.V. Mitkin & C.L. Ng, manuscript in preparation), elastic liquids (V.V. Mitkin, C.L. polystyrene butyl
Ng & T.G. Theofanous, manuscript in preparation)]. The ASOS facility is a large-scale shock tube, methacrylate polymer
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
with a flow area of 200 × 200 mm, capable of delivering (with helium in the driver) shock Mach (molecular weight of
∼2 million) in TBP at
numbers of up to 3.5 (dynamic pressures of up to 2 MPa). Reference is also made to some earlier
3.8% by weight
data obtained in the ALPHA facility (Theofanous et al. 2004, 2007), which is a large-scale, pulse,
Breakup time: time
supersonic wind tunnel with a flow area of 200 × 200 mm, capable of delivering Mach-3 flows at
required for complete
(flow static) pressure levels down to 10 Pa. The consistency of results between these two facilities loss of a drop’s bodily
has been demonstrated in the overlap region of operation. Fluids utilized include water, tri-butyl coherence
phosphate (TBP), glycerol, and a wide array of silicone oils of varying viscosities. For all tests
reported here, elasticity was introduced by dissolving (3.8% of ) poly-styrene-butyl-methacrylate
(PSBMA) in TBP. This non-Newtonian liquid is strongly shear-thinning, exhibits very large elas-
ticity modulus at the high extensional rates (De > 1) found in the SIE regime, and is highly resistant
to breakup. Its zero-shear viscosity is about 300 times that of water, and the solvent (TBP) viscosity
is four times that of water. The surface tension of PSBMA is the same as that of TBP and is quite
close to all silicone oils utilized in this work, which is about one-third that of water, whereas that
of glycerol is similar to that of water. Unless otherwise noted, initial drop diameters are in the 2–
3-mm range. Reference to a particular experimental result will be made according to the following
format: FLUID NAME run, M S = xx, We = xx, Oh = xx, and Re g = xx. For elastic fluids, addi-
tional definition is necessary; however, the use of PSBMA is sufficient for our purposes, and so is the
above specification of an experimental run. The zero time is taken at the instant of shock impact.
in Section 5. We conclude with a summary of key points and an outlook to future developments
(Section 6). Major components in this outlook are the use of direct numerical simulations for
enhancing understanding at the local scale and the use of effective field methods for relating the
local behaviors to the global dynamics (primary atomization), and Sections 5 and 6 provide a
preview of current work on these areas as well.
It will become apparent that this vision, and the results that support it, differs significantly from
those found in past work. As a reasonably complete summary of past work has become available in
a very recent review article (Guildenbecher et al. 2009), rather than elaborating on disagreements,
we focus this presentation on what is correct.
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Video
TBP run, Ms = 1.46, We = 7 × 103, Oh = 1.8 × 10–2, Reg = 3.7 × 104, TI = 0.14
Figure 2
A sampling of early-time interfacial morphologies spanning the classes of liquids and flow conditions considered. On the left is an
oblique (200 ) view and on the right is a right-angle view of the same drop. The line in the left panel represents a length scale of 2 mm.
(The video includes a series of still frames, each showing for 5 s.)
668 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
conditions that can now be recognized as SIE (Theofanous & Li 2008). In a similar vein, past
work, starting with Ranger & Nicholls (1969), deduced drag coefficients, and a time scaling for
deformation and breakup (t + ), by enveloping such shadow images over time periods that are five
times longer than the time required for complete loss of coherence (Section 5.1). [The same scaling
is also implied in Engel’s (1958) analysis; however, comparison to experimental data suffered from
the same difficulty of being unable to recognize what portion of the drop was intact.] More-to-
the-point analysis of deformation under aerodynamically imposed normal stresses on the surface
of a drop (Theofanous et al. 2007), at Oh 1, yielded the t + scaling, but an attempt to account
for Mach number effects (the parameter Ct ) is perhaps much too rough for the intended purpose.
Accordingly we take Ct = 1 in the data treatment here. The role of liquid viscosity or elasticity
on deformation was not addressed in past work.
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
A measure of viscous effects is provided by the Ohnesorge number. We assume that when
Oh > 1, deformation is controlled solely by drop acceleration and liquid viscosity, and this leads
to a characteristic time tV+ , a scaling that emerges also in RT instability analysis. As noted above, this
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
controls the timing of RTP. A similarly motivated viscous length scale (l V+ ) controls the criticality
itself (Section 3.1).
There is also the shock-passage timescale (tP+ ) that defines the period of highly enhanced drag
due to the asymmetry of the pressure loading. Deformations are scaled by d 0 , and when the tV+
scaling is relevant we also have the scaled initial diameter D0,V as an additional parameter.
It can be readily shown that the characteristic time for acceleration is much longer than tI+ , so
that deformation and RTP at Oh 1 occur always at virtually undiminished relative velocity.
This is far from the case when Oh > 1.
a Video b Video
Figure 3
Samples of deformation histories leading to Rayleigh-Taylor piercing. (a) Water run, M S = 1.01, We = 12, Oh = 2 × 10−3 , and
Re g = 3.3 × 103 . (b) Glycerol run, M S = 1.24, We = 980, Oh = 1.8, and Re g = 2.27 × 104 .
2.0
1.6 a b
1.8
1.4
1.6
D D
1.4
1.2
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
1.2
1.0
0 0.4 0.8 1.2 0 2 4 6
TI TV
Figure 4
Scaled deformation transients near Rayleigh-Taylor piercing criticality at (a) Oh 1 and (b) 1 < Oh < 560. In the latter case, tV+ is
based on instantaneous acceleration (deduced from displacement histories).
(Section 6). Deformation transients measured experimentally over a wide range of viscosities and
flow conditions (Figure 4, with deformations obtained in the manner illustrated in Figure 5a)
confirm the scaling approach introduced above. Furthermore, the data show that the role of the
DV 0 parameter is secondary. The timing of RTP is on the order of T I ∼ 1 and T V ∼ 5 for
Oh 1 and Oh > 1, respectively. In the latter case, allowance for rebound may increase the
scaled RTP timing to as much as seven.
4
a 6
b
Position in cross-stream (mm)
5
3
Velocity (m s–1)
2
3
2
1
Front
Center of mass
1
0
0
0 1 2 3 4 0 0.1 0.2 0.3 0.4 0.5
Position in flow direction (mm) Time (ms)
Figure 5
(a) Drop shapes and displacements, shown in 50-μs intervals, for the glycerol case illustrated in Figure 3, and (b) deduced
center-of-mass and drop-front velocities. The drop-front acceleration is a constant at 104 ms−2 , and the center-of-mass accelerations
range from 2 × 103 to 2.1 × 104 ms−2 . The characteristic times are tI+ = 500 μs and tV+ = 250 μs.
670 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
3. RAYLEIGH-TAYLOR PIERCING
a Cp b
0.9
0.7 Cp
0.5 1.0
0.3 0.7
0.1 0.4
–0.1 0.1
–0.3 –0.2
–0.5 –0.5
–0.7 –0.8
–0.9 –1.1
–1.1 –1.4
–1.3 –1.7
–1.5 –2.0
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
–2 –1 0 1 2 3 –2 –1 0 1 2 3
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
Diameter Diameter
c
Cp
1.0
0.6
0.2
–0.2
–0.6
–1.0
–1.4
–1.8
–2.2
–2.6
–3.0
–2 –1 0 1 2 3
Diameter
Figure 6
Steady-state pressure distributions from numerical simulations with some of the drop (axisymmetric) shapes shown in Figure 5.
Notable are the low-pressure regions that develop behind the tip (doughnut-shaped) region of the highly deformed drop.
In Figure 9 we can see that RTP is also obtained with elastic liquids, but with bags that are
elastic and retract radially to form long filaments. The same situation occurs at the next higher
mode (Figures 10 and 11), a sort of bag-and-stamen configuration, but again with elastic recovery
to nearly the drop’s original form. As discussed in Section 4, aerobreakup of such liquids must
involve fractures in the entangled polymer-chain network that lead to ruptures of the stretching
liquid filaments. The placement of these two data points in Figure 8 suggests that elasticity effects
on this special kind of RTP are not significant. Consistent with the low interfacial shear found at
such conditions, the Ohnesorge number for the PSBMA data points in Figure 8 is based on the
zero-shear viscosity.
672 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
8 2.0
a Sphere b Center of mass
Shape 8 MuSiC-SIM simulations
6 1.6
Drag coefficient (CD)
2 0.8
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
0 0.4
0 5 10 15 20 25 0 0.2 0.4 0.6 0.8 1.0
Tp TI
Figure 7
(a) Drag coefficient transients for a sphere, and shape 8 of Figure 5, computed for the flow conditions of the experiment, and
(b) comparison of computed steady-state values, obtained for the shapes applicable to selected time instances, with the experiment
values (deduced from the center-of-mass displacement data). The dashed portion of the experimental result in panel b indicates
uncertainty due to insufficient frequency of data points in this early portion.
flattened drop. The difference in the present case is that we must account for viscous dissipation
and, as a consequence, for the finite thickness of liquid involved. The basis is provided by the
dispersion relation (Mikaelian 1996), which adapted to our problem reads
1/2
K V2
NV = −K V2 + K V4 + KV 1− tanh(H V K V ) . (1)
Oh 20 DV 0
105 Video
a b
104
103
Wec
102
Hinze (1955)
101 Hsiang & Faeth (1995)
Present, Newtonian
Present, PSBMA
100 –2
10 10–1 100 101 102 103
Oh
Figure 8
(a) The first criticality [Rayleigh-Taylor piercing (RTP)] in the classical coordinates used to reflect the effect of viscosity. Open symbols
denote the absence of breakup (drops were followed until they reached terminal velocities, far downstream from the position of initial
interaction). Half-full symbols denote the occurrence of the elementary RTP-mode (bag breakup), as shown in panel b (glycerol run,
M S = 1.17, We = 390, Oh = 1.9, and Re g = 1.26 × 104 ). The Oh for the PSBMA data on this plot is based on the zero-shear viscosity
of the polymer solution.
a Video b Video
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Figure 9
Illustration of the Rayleigh-Taylor piercing regime with viscoelastic liquids. (a) The first mode (bag) (PSBMA run, M = 3, We = 76,
Oh = 1.35, and Re = 115) obtained in the ALPHA facility and (b) the second mode (PSBMA run, M S = 1.07, We = 150, Oh = 1.35,
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
This elegant result was obtained for planar waves by extending Layzer’s (1955) potential flow
theory. A similar derivation using cylindrical waves for the first mode [the J0 (kr) eigenfunction]
leads to essentially the same dispersion relation. For the case of a semi-infinite viscous fluid, this
solution was shown to be a rather poor approximation (Mikaelian 1996, figure 2) to (exact) results
derived by numerical means. It is easy to see that this resulted from a trivial error in the derivation
(a factor of 2 is missing from Mikaelian’s equation 24); in fact the agreement is quite good. Another
comparison with numerical results was for the case of finite thickness imposed by rigid, nonslip
walls. The agreement was fair, not surprisingly, as such a boundary condition undercuts the basis of
the theoretical construct employed (the slip condition engendered in the potential eigenfunctions
employed for the analytical result). In our case the bounding interface is free (the back side of the
flattened drop), and we can expect much better performance. Thus we conclude that Equation 1
is appropriate for our purposes. As expected, it exhibits the same (capillary) cutoff wave number
irrespective of viscosity.
The quantitative trends in this two-parameter solution are illustrated in Figure 12a. We can
see that the value of Oh 2 DV 0 ∼ 10 provides a demarcation above which viscous effects become
a Video b
Figure 10
Illustration of a first approach to shear-induced entrainment with fallback to Rayleigh-Taylor piercing (a), and elastic reconstitution of
the highly disrupted drop as observed at a position 64 cm downstream (b). PSBMA run, M S = 1.09, We = 250, Oh = 1.35, and
Re g = 5.9 × 103 .
674 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
Video
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
Figure 11
Oblique and right-angle laser-induced fluorescence images from a run similar to that shown in Figure 10. (PSBMA run, M S = 1.11,
We = 345, Oh = 1.26, and Re g = 7.8 × 103 .) The line represents a length scale of 2 mm. Images taken at 1.34 and 1.50 ms following
shock impact, which correspond to T I = 1.45 and 1.63. (The video includes a series of still frames, each showing for 5 s.)
a 0.5 b
HV Oh2DV M Oh
0.1 106 0.4 3.0 10–2
0.1 103 3.0 10–2
0.4
0.1 102 0.25 1.9
0.1 1 0.25 5.2
2.0 106 0.3 0.3 23
0.3 2.0 103 0.8 1.6 × 102
2.0 102 1.2 3.1 × 102
NV
NV
0.1
0.1
0 –3 0 –3
10 10–2 10–1 100 101 102 10 10–2 10–1 100 101
KV KV
Figure 12
(a) Dispersion relation of Rayleigh-Taylor instability in viscous-scaled coordinates. (b) The dots show the wave numbers of
experimentally observed onsets of Rayleigh-Taylor piercing (RTP). The Mach-3 cases were carried out in ALPHA (rarefied flow) at
Weber numbers of 9 (dashed line) and 28 (solid line). The off-peak arrow in the latter suggests a higher-mode RTP as illustrated in the
inset. The viscous scaling was made with actual (measured) acceleration at the time just prior to RTP.
dominant (and surface tension becomes irrelevant, of course away from the cutoff limits, which
move further away from the peak-growth regions with increasing Oh 2 DV 0 ). This diminished
role of surface tension at high values of Oh is also possible to discern from the criticality data
trends in Figure 8, as noted above. Direct incorporation of this data in the present framework
is shown in Figure 12b, the remarkable feature of which is the proximity of RTP-active wave
numbers at criticality to the peak growth regions (around K V∗ ) of the dispersion curves. The
same is true for the low-viscosity branch of Figure 8a, but here the peak is quite close to the
cutoff (surface tension is important). Moreover, here we can see the positioning of a higher RTP
mode obtained in rarefied Mach-3 flow (ALPHA). The data plotted in Figure 12b include all
experimental runs for which instantaneous accelerations could be deduced, as needed for the
scaling employed. The positioning of this data on the wave-number axis is based on the deformed
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
diameter, under the assumption K V = π/DV , and the acceleration at the moment just prior to
instability.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
In conclusion, we now need to develop a more appropriate way of expressing RTP criticality
for the viscous branch of Figure 8a. For this, we first note (Figure 13a) that the peak-growth
−1/4
region can be expressed, to a very good approximation for H V < 2, simply by K V∗ = 0.9 H V .
It is then trivial to derive the first-mode RTP criticality (one-half wavelength fitting exactly within
the frontal area of a flattened drop) as
DV 0 D2 = 4.6 (2)
or equivalently as
PV ,c D8 = 130. (3)
As shown in Figure 13 these theoretical interpretations are fully supported by the experimental
data. Moreover, because all relevant dynamics have been accounted for, it is now clear that there
are no viscous limits to RTP, which is a conclusion supported by the experimental data extending
out to a viscosity of 100,000 times that of water. Notably, as suggested by Equation 3, the first RTP
criticality can be expressed simply by a critical gas-dynamic pressure that scales with a characteristic
liquid-dynamic pressure (based on a liquid density and the viscous velocity u + V ).
4. SHEAR-INDUCED ENTRAINMENT
676 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
10 8
a Exact dependence b
Analytical approximation
90% envelope
95% envelope 6
Experimental data
DV0, c D 2
K *V
4
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
2
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
0.1 0
1 2 0 100 200 300
HV Oh
180
c
120
PV, c D 8
60
Figure 13
A theoretical interpretation of Rayleigh-Taylor piercing (RTP) criticality. (a) Variation of the maximum-growth wave number with
viscosity-scaled, flattened-drop thickness. (b) Comparison of the theoretical prediction of RTP criticality (DV 0 D2 = 4.6) (line) with
experimental data ( points). (c) Comparison of theoretical prediction of RTP criticality (PV ,c D8 ∼ 130) (line) with the experimental data
( points). Viscous scaling is as in Figure 12.
near-stagnation region remains relatively flat all the way until the complete destruction of the drop
(Section 5). The area of the smooth windward-surface region seems to decrease with increasing
flow-dynamic pressure and perhaps somewhat with liquid viscosity, and elasticity, but in no case is
the frontal area penetrated by unstable RT waves. It is thus suggested that the smoothness results
from straining liquid motions in the forward stagnation region induced by the shearing action in
the periphery. Direct numerical simulations of the type discussed in Section 6 hopefully will aid
in our understanding of these complex mechanics.
The high-viscosity branch in Figure 14a suggests that at SIE criticality the Weber number
is proportional to the Ohnesorge number. Straightforward reduction of this observation leads
to the perhaps more revealing scaling relation shown in Figure 16. It involves a characteristic
106
a Present, Newtonian b Video
Present, PSBMA
105
104
Wec
103
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
102
101
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
Figure 14
(a) The second criticality (shear-induced entrainment) in the classical coordinates used to reflect the effect of viscosity. Open symbols
denote the occurrence of entrainment as illustrated in panel b (water run, M S = 1.23, We = 750, Oh = 2.25 10−3 , and
Re g = 2.2 × 104 ), even though, depending on the proximity to the critical condition, this may revert to Rayleigh-Taylor piercing as
illustrated in the video in Figure 3. Half-full symbols indicate higher modes of RTP; they develop increasingly between the first and
the second criticalities. The Oh for the PSBMA data is based on the solvent viscosity.
Video
TBP run, Ms = 1.21, We = 1.6 × 103, Oh = 1.8 × 10–2, Reg = 1.6 × 104, TI = 0.26
Figure 15
Detailed morphologies of the shear-induced entrainment regime spanning the classes of liquids and flow conditions considered. Scales
and views are the same as in Figure 2. (The video includes a series of still frames, each showing for 5 s.)
678 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
105
104
PVC,c
103
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
102
10–1 100 101 102
Oh
Figure 16
An interpretation of shear-induced entrainment criticality physics based on the characteristic viscous and
capillary velocities of the liquid (see Scheme 2).
liquid-dynamic pressure based on the product of two different velocities: the viscous velocity (u + V)
seen already in the scaling of RTP and a capillary velocity (u C+ ) that may be thought of as providing
a limiting balance between the thrust of a liquid eddy (or a wave) within the drop and the surface
tension force that restrains its detachment. Again we look forward to numerical simulations that
will provide the means to pursue these suggestions.
Behaviors similar to SIE are seen with elastic fluids, but with two exceptions: (a) Rather than
the particulation of entrained mass seen with Newtonian liquids, here we see the creation of
sheets, which rebound elastically (unless the dynamic pressure is high enough to cause fractures;
see below) toward reconstituting the original drop (Figure 10). (b) Because of the shear-thinning
property of this fluid, the onset of this phenomenon is characterized by an Oh based on the
viscosity of the pure solvent.
a Video
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
Figure 17
(a) Illustration of the shear-induced entrainment with rupture regime at flow conditions near its onset
(PSBMA run, M S = 1.45, We = 8.5 × 103 , Oh = 1.6 × 10−2 , and Re g = 4.47 × 104 ). (b) Laser-induced
fluorescence image detail at 241 μs (T I = 1.2) obtained in a repeat run. On the left is an oblique (200 ) view
and on the right is a right-angle view of the same drop. The line represents a length scale of 2 mm.
680 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
Screen
Vacuum chamber
Target
Nozzle Test liquid
2r0
Bullet
Guide rod
Rubber
Bullet
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
Piston
Gus gun
Figure 18
Schematic of the IRIS rheometer. It operates with microliter liquid quantities that are shot against an
inverted cone at velocities of up to 300 ms−1 .
We have been able to reproduce such rupture phenomena in the apparatus illustrated in
Figure 18 (Mitkin et al. 2008). Sample results are shown in Figure 19. In a preliminary, simplified
analysis we used the rebounds seen in the data to estimate the effective elasticity moduli at strain
rates of up to a few thousand inverse seconds, and on this basis evaluated the magnitude of the
elastic tension at the inception of ruptures. The magnitudes of these limit tensions appear to be
the property of the particular polymer-solvent pair, to scale with concentration, and to be con-
sistent with a molecular-bond fracture mechanism. Accordingly, a topological change associated
with rupture is suggested to be the consequence of an avalanche of polymer-chain fractures that
suddenly release elastic tension at levels unsupportable by the solvent itself.
The application of such results to aerobreakup would lead to a means of computing a filament’s
elongation rates, under the action of aerodynamic forces and evolving elastic tensions, yielding
Figure 19
Illustration of (a) elastic retraction, (b) propagation, and (c) rupture, obtained with increasing impact velocity in the IRIS experiment.
These are PSBMA runs at impact velocities of 19, 33, and 65 ms−1 , respectively.
eventually ruptures if the magnitude of applicable limit tension was reached. We can expect that
such work, aided by mesoscale simulations of entangled polymer solutions (Kivotides et al. 2010),
should open the way to the pertinent scaling laws and ultimately a priori prediction of all key
features of interest to aerobreakup.
final, equilibrium particle-size distribution, at the time that all breakup ceases and particles attain
terminal velocities in the atmosphere. Such a source cloud is then delivered to an atmospheric
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
transport code to predict ground depositions (and thereby ground effects). To assess the damage
of raindrop impact in supersonic flight, one is interested in the detailed transformation of the drop
(the relative amounts of mass in the core drop and its satellite cloud) as it transits the distance
between the bow shock and the point of impact. Two-phase combustion and certain chemical
processing applications require even greater detail on two-phase mixing, and this includes the
cloud dynamics along with corresponding particle-size-distribution transients. Whereas past work
is not useful in any of these areas, it is instructive to mention the main issues briefly here.
In the dissemination area, the task was approached principally by field tests. These tests have
been carried out at scales from tens to hundreds of kilograms, statically or dynamically, over a
time period of approximately 60 years, continuing to this day. They used witness cards of various
kinds to capture drops on the ground, and various kinds of optical and chemical methods to
determine deposition densities and particle-size distributions at the grid positions of the cards.
The most recent tests have also employed high-speed video to record the overall, gross features
of the cloud dynamics. Among the many problematic aspects of these tests are difficulties in
targeting the test grid, including the adverse effect of winds, scarcity of sampling stations, and use
of poorly characterized and calibrated witness cards. The latter difficulties are further aggravated
by overlapping stains as the absorbed liquid drops diffused within the substrate material. In no
case could complete material recovery be assured. A recent critical review concluded that none
of the results of these tests could withstand normal scientific scrutiny (Babarsky & Theofanous
2010).
The problem of raindrop impact was addressed in the laboratory by shock-tube tests. For this
application the relevant regime is SIE with Oh 1, and as shown above, available instrumentation
was not up to the task. A conservative estimate of breakup time was obtained from measured times
to produce a widely diffuse mist (deemed at this point to be structurally harmless). This subjective
result, T I ∼ 5, became pervasive for aerobreakup in general, and there is nothing to be found
on particle-cloud dynamics. Combustion-motivated studies are of similar scope and stature. [The
most ambitious work in this area is that of Faeth’s, as reported in several papers after Hsiang &
Faeth (1995), but these papers, based on visualization, provide no direct data nor do they provide
evidence necessary to judge the quality of the data reductions.]
682 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
Video
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
TBP run, Ms = 1.21, We = 1.6 × 103, Oh = 1.8 × 10–2, Reg = 1.6 × 104, TI = 0.53
Figure 20
Detailed morphologies of the advanced shear-induced entrainment regime, near complete loss of the drop’s bodily coherence, spanning
the classes of liquids and flow conditions considered. Scales and views are the same as in Figure 2. (The video includes a series of still
frames, each showing for 5 s.)
be well represented by the estimate T I ∼ 1 for Oh 1 and T V ∼ 7–9 for Oh > 1. In SIE,
especially at Oh 1, the task of determining tLC is quite more difficult. For this purpose, we need
the volume of the core drop as a function of time, and this requires the synergistic use of video
LIF and maximum-resolution LIF (as in Figure 20), along with shadow video (as in Figure 14b)
in mutually refining and refocusing the succession of repeat experimental runs. A sample result
is shown in Figure 21. The present state of our database suggests that the T LC ∼ 1 estimate is
appropriate for all Newtonian liquids of any viscosity.
1.0
0.8
0.6
V/V0
0.4
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
0.2
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
0.0
0.00 0.25 0.50 0.75 1.00
TI
Figure 21
Fractional volume of liquid remaining as a coherent entity, in the shear-induced entrainment regime, as a
function of time. (Water run, M S = 1.23, We = 780, Oh = 2.4 × 10−3 , and Re g = 2.2 × 104 .)
Figure 22
A digitally composed cloud at a position 10 cm downstream of the interaction (PSBMA run, M S = 1.95,
We = 4.4 × 104 , Oh = 1.6 × 10−2 , and Re g = 105 ). The black dot at the top right corner indicates the size
of the original drop. The other dots shown represent sizes from 1 mm to 0 in 100-μm increments.
684 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
task is feasible because the global cloud dynamics are known, as described above, and because the
experiments are highly reproducible both in the space and time domains.
In the case of the atmospheric dissemination problem, the use of such results is twofold: first,
in providing the local constitutive description of breakup needed in system-level computations
that address the disaggregation problem (Theofanous et al. 2006, Lhuillier & Theofanous 2010,
Liou & Theofanous 2010) and, second, in providing ideas for a direct scale-up methodology that
would allow one to relate laboratory tests to field conditions. Clearly these two efforts are mutually
reinforcing and therefore beneficial to both of the respective goals.
6. OUTLOOK
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
The problem of aerobreakup is complex in that multiple physics participate in a competitive man-
ner, and it is fascinating given the wide variety of outcomes obtained as a result of this competition
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
at various ranges of the parameters involved. Still, with the help of advanced visualization methods
that allow probing of the internal structures over the complete transient of dispersal, applied to
experiments that span wide ranges in viscosity and elasticity, it has been possible to unify the
behaviors and understand underlying physics in terms of rather simple concepts and scaling laws.
What remains to be done is to extend this understanding by means of direct numerical simulations
at the local level and in turn by effective field methods at the system scale (e.g., for the dissem-
ination problem, as discussed above). The experimental data provide important benchmarks for
these challenging numerical tasks.
For direct numerical simulations, we use the sharp interface method (Nourgaliev et al. 2008),
extended to compressible flows by the use of the exact Riemann problem to compute interface
velocities (C.H. Chang, X. Deng & T.G. Theofanous, manuscript in preparation). Initial results
from this MuSiC-SIM code, already verified by comparison to a comprehensive set of drag data
on spheres (Bailey & Hiatt 1972), applied to the evolving shapes of a glycerol drop, suggest that
the basic aerodynamics of the computation are fit for the purpose (Figure 7). Similarly promising
is an initial step toward fully coupled simulations. As shown in Figures 23 and 24, even though
a b
Figure 23
Laser-induced-fluorescence visualization of internal structure of clouds during early stages of dispersal. (a) PSBMA run, M S = 2.06,
We = 4.7 × 104 , Oh = 1.8 × 10−2 , and Re g = 105 . (b) TBP run M S = 1.09, We = 280, Oh = 1.6 × 10−2 , and Re g = 7.5 × 103 . The
right edges of these frames are 3 cm downstream of the original position of the drops. The line shown in panel a represents a length
scale of 2 mm, and its thickness is 50 μm. The two black dots at the upper-left corner in panel b represent 50 μm and 100 μm.
Figure 24
Laser-induced-fluorescence visualization of a glycerol drop at conditions (similar to Figure 8b) just past Rayleigh-Taylor piercing
criticality (M S = 1.2, We = 520, Oh = 1.82, and Re g = 1.57 × 104 ) in comparison with numerical simulations. The time is 400 μs
following shock impact. The flow is from the right. In the experiment, the bag shown above turns inside out at 700 μs, before it turns
around again to fully develop and burst. This oscillation probably results from the low-pressure region seen in the numerical simulation
(Figure 25) to develop behind the outer edge of the drop and/or a highly concentrated shear stress at the same outer edge of the drop
(it is an order of magnitude greater than that found at the drop’s equator when of spherical shape).
at a much coarser grid than that utilized in the quasistatic computations, the basic features of
the deformation transient are captured. A parallel version of the code is nearly ready, which will
allow fully coupled computations as well as lower viscosities and higher shock Mach numbers and
eventually will address elastic liquids.
Video
Time = 3
3.9e–006
9e– s
Figure 25
Numerical simulation of the experiment shown in Figure 24. The early portion of this video is shown at a
lower speed to allow a better visualization of the shock-wave dynamics.
686 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
SUMMARY POINTS
1. Across fluid properties and flow conditions, the first criticality results from RTP, and
the second criticality results from SIE—this is also the terminal criticality (there is no
catastrophic regime).
2. The first criticality selects the fastest-growing wave of RT instability (in the linear
regime); when Oh > 1, the role of the finite drop length scale in this selection is signif-
icant. Beyond moderately high viscosities, surface tension becomes irrelevant, and the
Weber number is not an appropriate criterion for this criticality. Rather, the process
scales with a characteristic liquid-dynamic pressure based on a viscous velocity.
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
FUTURE ISSUES
1. A deeper theoretical understanding needs to be developed of the SIE and SIER regimes,
including the high-strain-rate rheology of polymer solutions and molecular-chain-
fracture phenomena in relation to various polymer-chain relaxation mechanisms.
2. The key mechanics in all three aerobreakup regimes (RTP, SIE, SIER) and the transitions
among them need to be addressed by means of direct numerical simulations.
3. The particle-cloud dynamics and particle-size distributions for Newtonian and viscoelas-
tic liquids in millimeter-scale drops should be quantified in detail.
4. There is a need for the development of a means for effective-field simulation of cloud
dynamics and particle-size distributions and for benchmarking against experimental data
obtained with millimeter-scale drops. The use of so-developed numerical tools would
aid our understanding toward the scaling and simulation of large-scale (atmospheric
dissemination field) tests.
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
DISCLOSURE STATEMENT
The author is not aware of any affiliations, memberships, funding, or financial holdings that might
be perceived as affecting the objectivity of this review.
ACKNOWLEDGMENTS
This work was supported by the Joint Science and Technology Office, Defense Threat Reduction
Agency ( JSTO/DTRA), and the National Ground Intelligence Center (NGIC). We are grateful
to Dr. Richard Babarsky (NGIC) for his encouragement, cooperation, and support all through
this work from the very beginning. The lead experimentalists in the ALPHA and ASOS programs
were (successively) Dr. G.J. Li, Dr. C.L. Ng, and Dr. V.V. Mitkin. The lead code developer
for MuSiC-SIM is Dr. C.H. Chang and in the computations presented here participated also
Dr. X. Deng.
LITERATURE CITED
Includes comments of
20 commissioned Babarsky R, Theofanous TG. 2010. An assessment of the state-of-the-art on aerodynamic/explosive
reviewers and responses dissemination of chemical agents with perspectives for future work. Natl. Ground Intell. Center
by the authors; it is a
Rep. WF-67774
useful point of reference
Bailey AB, Hiatt J. 1972. Sphere drag coefficients for a broad range of Mach and Reynolds numbers.
for future field work.
AIAA J. 10:1436–40
Batchelor GK. 1987. The stability of a large gas bubble rising through liquid. J. Fluid Mech. 184:399–422
Standard reference for
drag on spheres in
Brodkey RS. 1967. The Phenomena of Fluid Motions. Reading, MA: Addison Wesley
transonic and Dodd KN. 1960. On the disintegration of water drops in an air stream. J. Fluid Mech. 9:175–82
supersonic flows Engel OG. 1958. Fragmentation of waterdrops in the zone behind an air shock. J. Res. Natl. Bur.
(200 < Re g < 105 ; Stand. 60:245–80
0.1 < M < 6). Gelfand BE. 1996. Droplet breakup phenomena in flows with velocity lag. Prog. Energy Combust. Sci. 22:201–
65
An early, careful, Gorokhovski M, Herrmann M. 2008. Modeling primary atomization. Annu. Rev. Fluid Mech. 40:343–66
voluminous paper that Guildenbecher DG, Lopez-Rivera C, Sojka PE. 2009. Secondary atomization. Exp. Fluids 46:371–402
asked most of the right Hanson AR, Domich EG, Adams HS. 1963. Shock tube investigation of the breakup of drops by air blasts.
questions; an important
Phys. Fluids 6:1070–80
point of reference in
Harper EY, Grube GW, Chang I. 1972. On the breakup of accelerating liquid drops. J. Fluid Mech. 52:565–
aerobreakup.
91
688 Theofanous
FL43CH27-Theofanous ARI 19 November 2010 16:35
Hinze JO. 1955. Fundamentals of the hydrodynamic mechanism of splitting in dispersion processes.
First theoretical and
AIChE J. 1:289–95
experimental
Hsiang LP, Faeth GM. 1992. Near-limit drop deformation and secondary breakup. Int. J. Multiphase Flow
consideration of
18:635–52
viscosity effects;
Hsiang LP, Faeth GM. 1995. Drop deformation and breakup due to shock-wave and steady disturbances. Int.
provides interesting
J. Multiphase Flow 21:545–60 historical remarks on
Joseph DD, Beavers GS, Funada T. 2002. Rayleigh-Taylor instability of viscoelastic drops at high Weber early developments,
numbers. J. Fluid Mech. 453:109–32 including some
Joseph DD, Belanger J, Beavers GS. 1999. Breakup of a liquid drop suddenly exposed to a high-speed airstream. little-known literature.
Int. J. Multiphase Flow 25:1263–303
Kivotides D, Wilkin SL, Theofanous TG. 2010. Stochastic entangled chain dynamics of dense polymer
solutions. J. Chem. Phys. 133:144903
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Krechetnikov R. 2009. Rayleigh-Taylor and Richtmyer-Meshkov instabilities of flat and curved interfaces.
J. Fluid Mech. 625:387–410 Presents a clever and
Lasheras JC, Hopfinger EJ. 2000. Liquid jet instability and atomization in a coaxial gas stream. Annu. Rev. powerful method that
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
RELATED RESOURCES
Thoroddsen ST, Etoh TG, Takehara K. 2008. High-speed imaging of drops and bubbles. Annu.
Rev. Fluid Mech. 40:257–85
Yarin AL. 2006. Drop impact dynamics: spashing, spreading, receding, bouncing. . . . Annu. Rev.
Fluid Mech. 38:159–92
Access provided by Indian Institute of Technology - Mumbai/ Bombay on 11/17/22. For personal use only.
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
690 Theofanous
FL43-FrontMater ARI 15 November 2010 11:55
Annual Review of
T. Mullin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Fish Swimming and Bird/Insect Flight
Annu. Rev. Fluid Mech. 2011.43:661-690. Downloaded from www.annualreviews.org
v
FL43-FrontMater ARI 15 November 2010 11:55
Indexes
Errata
An online log of corrections to Annual Review of Fluid Mechanics articles may be found
at http://fluid.annualreviews.org/errata.shtml
vi Contents