scispace - formally typeset
Open AccessJournal ArticleDOI

Numerical simulation of the aerobreakup of a water droplet

Reads0
Chats0
TLDR
In this paper, the aerobreakup of a spherical water droplet in the flow behind a normal shock wave is simulated using the compressible multicomponent Euler equations in a finite-volume scheme with shock and interface capturing.
Abstract
We present a three-dimensional numerical simulation of the aerobreakup of a spherical water droplet in the flow behind a normal shock wave. The droplet and surrounding gas flow are simulated using the compressible multicomponent Euler equations in a finite-volume scheme with shock and interface capturing. The aerobreakup process is compared with available experimental visualizations. Features of the droplet deformation and breakup in the stripping breakup regime, as well as descriptions of the surrounding gas flow, are discussed. Analyses of observed surface instabilities and a Fourier decomposition of the flow field reveal asymmetrical azimuthal modulations and broadband instability growth that result in chaotic flow within the wake region.

read more

Content maybe subject to copyright    Report

J. Fluid Mech. (2018), vol. 835, pp. 1108–1135.
c
Cambridge University Press 2017
doi:10.1017/jfm.2017.804
1108
Numerical simulation of the aerobreakup of a
water droplet
Jomela C. Meng
1,
and Tim Colonius
1
1
California Institute of Technology, Pasadena, CA 91125, USA
(Received 29 October 2016; revised 28 October 2017; accepted 3 November 2017;
first published online 29 November 2017)
We present a three-dimensional numerical simulation of the aerobreakup of a spherical
water droplet in the flow behind a normal shock wave. The droplet and surrounding
gas flow are simulated using the compressible multicomponent Euler equations
in a finite-volume scheme with shock and interface capturing. The aerobreakup
process is compared with available experimental visualizations. Features of the droplet
deformation and breakup in the stripping breakup regime, as well as descriptions of
the surrounding gas flow, are discussed. Analyses of observed surface instabilities and
a Fourier decomposition of the flow field reveal asymmetrical azimuthal modulations
and broadband instability growth that result in chaotic flow within the wake region.
Key words: breakup/coalescence, drops, shock waves
1. Introduction
The study of droplet aerobreakup has historically been motivated by three
applications: bulk dissemination of liquid agents, raindrop damage during supersonic
flight, and secondary atomization of liquid jets in turbomachinery. Much of the
aerobreakup literature has focused research efforts on characterizing and mapping
various breakup regimes (e.g. Lane 1951; Engel 1958; Hanson, Domich & Adams
1963; Ranger & Nicholls 1968; Hsiang & Faeth 1995), calculating characteristic
breakup times (e.g. Ranger & Nicholls 1968; Hsiang & Faeth 1992), quantifying
dependence on parameters such as density and viscosity ratios (e.g. Hanson et al.
1963; Theofanous et al. 2012), predicting final drop size distributions (e.g. Ranger
& Nicholls 1968; Pilch & Erdman 1987), and quantifying unsteady drag properties
(e.g. Engel 1958; Simpkins & Bales 1972; Joseph, Belanger & Beavers 1999).
Unfortunately, these experimental and theoretical research efforts have resulted in
many, and often conflicting, phenomenological models describing the aerobreakup
process, and to date, a definitive understanding of aerobreakup remains elusive
(Khosla, Smith & Throckmorton 2006).
Beginning with the work of Hinze (1949), the Weber number, We = ρ
g
u
2
g
D
0
, has
been the principal parameter used to delineate the various regimes of aerobreakup.
Traditionally, there exist ve distinct regimes that are well established in the literature
(Pilch & Erdman 1987; Guildenbecher, López-Rivera & Sojka 2009). They are,
Email address for correspondence: jomela.meng@caltech.edu
Downloaded from https://www.cambridge.org/core. Caltech Library, on 20 Dec 2017 at 20:54:57, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2017.804

Numerical simulation of the aerobreakup of a water droplet 1109
in order of increasing We, the vibrational, bag, bag-and-stamen, stripping, and
catastrophic regimes.
The stripping regime, which is of primary interest in this paper, marks a transition
in breakup physics that fundamentally differs from that of the preceding breakup
regimes. Generally speaking, the stripping regime is characterized by an initial
deformation of the droplet into a disk-like shape. Following the shape change, droplet
fluid is observed to be stripped from the droplet’s periphery in a region near the
droplet equator (defined as a polar, or inclination, angle of ϕ = π/2). Ranger &
Nicholls (1968) postulated the ‘boundary layer stripping’ or ‘shear stripping’ model,
where boundary layers both inside and outside the drop become unstable at the
droplet equator, and are subsequently stripped off by the ambient gas flow. Liu &
Reitz (1997) later proposed an alternate mechanism known as ‘sheet thinning’, where
the droplet is initially flattened by the pressure gradient between the drop’s poles
(ϕ = 0, π) and equator. Once flattened, the strong inertial forces from the surrounding
flow draw a thin sheet of liquid off the periphery. This sheet is accelerated, stretched,
and bent in the direction of flow, and eventually breaks up into streamwise ligaments
that fragment into individual drops. Due to the flattened disk-like shape of the
deformed droplet, flow separation occurs for all practical values of the Reynolds
number, Re = ρ
g
u
g
D
0
g
, and the sheet thinning mechanism can be considered an
inviscid phenomenon with no dependence on Re (Guildenbecher et al. 2009).
The instability of the thin liquid sheet that is drawn from the droplet’s periphery in
the sheet thinning model is thought to be responsible for the generation of product
droplets. Liu & Reitz (1997) described a ‘stretched streamwise ligament breakup’
mechanism wherein, for low liquid flow rates of a planar liquid sheet sandwiched
between two shear air layers, streamwise vortical waves, alternating with thin liquid
membranes, would grow along the sheet. The membranes would burst first due to the
rotation of the streamwise vortices, leaving streamwise ligaments that subsequently
broke up. More recently, Jalaal & Mehravaran (2014) attributed the breakup to the
rise of Rayleigh–Taylor (RT) instability waves on the sheet. In this mechanism, the
Kelvin–Helmholtz (KH) instability generates axisymmetric waves at the liquid–gas
interface. The transient acceleration of these wave crests or rims (in the case of
the liquid jet) into the downstream air triggers a RT instability, which produces
‘transverse azimuthal modulations.
Recent work by Theofanous, Li & Dinh (2004) studying aerobreakup in rarefied
supersonic flows, and subsequent publications (Theofanous & Li 2008; Theofanous
2011; Theofanous et al. 2012), has substantially changed the overall understanding of
aerobreakup. Theofanous et al. (2004) proposed a reclassification into two principal
breakup regimes: Rayleigh–Taylor piercing (RTP) and shear-induced entrainment (SIE).
Theofanous & Li (2008) described SIE as a combination of shear-driven radial motion,
which results in the flattening, as well as instabilities on the stretched liquid sheet.
Perhaps most importantly, this reclassification argues that the catastrophic breakup
regime does not exist. Theofanous & Li (2008) contended that the wavy interface
on the upstream side of the droplet was an artefact created by a projected view of a
complex flow field, and that no RT waves exist or pierce the drop. Using laser-induced
fluorescence as their experimental visualization technique, SIE was proposed as the
terminal regime for We > 10
3
.
Recently, numerical simulation has emerged as a tool for studying aerobreakup.
Unfortunately, due to the high computational costs of fully three-dimensional (3D)
simulations, numerical aerobreakup studies have often invoked two-dimensional
(2D) (Zaleski, Li & Succi 1995; Igra & Takayama 2001a,b,c; Chen 2008) or
Downloaded from https://www.cambridge.org/core. Caltech Library, on 20 Dec 2017 at 20:54:57, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2017.804

1110 J. C. Meng and T. Colonius
axisymmetric (Han & Tryggvason 2001; Aalburg, Leer & Faeth 2003; Wadhwa, Magi
& Abraham 2007; Chang, Deng & Theofanous 2013) approximations. Additionally,
fluid density ratios are often assumed to be small, or the fluids are considered to
be incompressible (Han & Tryggvason 2001; Aalburg et al. 2003; Quan & Schmidt
2006; Jalaal & Mehravaran 2014; Xiao, Dianat & McGuirk 2014; Castrillon Escobar
et al. 2015; Jain et al. 2015). Previous work from the authors (Meng & Colonius
2015) simulated the early stages of two-dimensional aerobreakup, with comparison
to the experimental work of Igra & Takayama (2001c). Qualitative features of the
breakup process, such as the presence of a transitory equatorial recirculation region
and an upstream jet in the wake, were discussed. Additionally, a parametric study
varying incident shock strength was performed to study the effects of the transition
between subsonic and supersonic post-shock flow. A novel method of calculating
the cylinder’s centre-of-mass properties was also utilized to obtain accurate unsteady
acceleration and drag histories.
Using numerical simulation, the purpose of this paper is to elucidate the physical
breakup mechanisms responsible for the fully three-dimensional, compressible
aerobreakup of a single water droplet in air. While a direct numerical simulation
would be ideal for such an investigation, the computational grid required to fully
resolve all scales of the aerobreakup problem is intractable on currently available
computational resources; a compromise must be made. We thus consider the ‘inviscid’
case, which is in turn regularized by artificial viscosity and numerical diffusion
of the interface. The effects of these approximations are discussed in detail in
§ 3.2. Additionally, we will attempt to interpret the numerical results in a manner
consistent with the uncertainties introduced by the approximations. These results
fill a gap in the current state of droplet aerobreakup knowledge associated with
the underlying fundamental flow physics that dictates the experimentally observed
phenomena. Building upon the computational efforts of Coralic & Colonius (2014),
we utilize a compressible multicomponent flow solver to numerically investigate this
fundamental fluid dynamics problem. The rest of this paper is organized as follows.
In § 2, we describe the problem set-up and introduce the governing equations. The
numerical algorithm is subsequently described in § 3, along with the calculation of
various flow quantities used in the analysis. We describe the evolution of the liquid
droplet, show its centre-of-mass properties, and compare with available experimental
visualizations in §§ 4, 4.1 and 4.2. The flow field in the gas phase is investigated in
§ 4.3. The mechanisms of surface instabilities are discussed in § 5.1, and are followed
by a Fourier decomposition analysis in § 5.2. Finally, concluding remarks are made
in § 6.
2. Physical modelling
2.1. Problem description
In the laboratory, shock tubes are most often used to generate large relative velocities
between the liquid droplet and surrounding gas flow. Normal shock waves, in and
of themselves, have little effect on the droplet. Instead, they serve as a reliable and
repeatable way to create a high-speed flow around the droplet, which is responsible
for the droplet’s subsequent deformation and disintegration (Hanson et al. 1963;
Ranger & Nicholls 1968; Joseph et al. 1999). In figure 1, a schematic is shown of
the initial condition and computational grid. The shock wave of strength M
s
= 1.47
(modelled as a step discontinuity in fluid properties) is travelling in air towards
the water droplet, which is initialized as a spherical interface with diameter D
0
on
the cylindrical grid. The drop and the ambient air downstream of the shock are
Downloaded from https://www.cambridge.org/core. Caltech Library, on 20 Dec 2017 at 20:54:57, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2017.804

Numerical simulation of the aerobreakup of a water droplet 1111
Post-shock
air
Pre-shock
air
NRBC
NRBC
NRBC
NRBC
r
z
FIGURE 1. (Colour online) Schematic of the 3D initial condition and non-uniform
computational domain at 1:10 of the actual resolution. The grid extends radially outward
for 6D
0
, and the axial extents are 7D
0
6 z 6 15.5D
0
.
initialized at rest. The entire 3D flow field is simulated with no enforced symmetries.
Non-reflective boundary conditions (NRBC), applied at all computational domain
boundaries, approximately extend the surrounding air to infinity. The simulation is
performed on the computational domain = [−7D
0
, 15.5D
0
] × [0, 6D
0
] × [0, 2π]
with a spatial resolution of (N
z
, N
r
, N
θ
) = (800, 600, 320) grid cells. The grid is
stretched towards the axial boundaries using a hyperbolic tangent function. The most
refined portion of the grid is located near the initial location of the droplet and
in the region of the near-field wake. In this region, the nominal axial and radial
grid resolution corresponds to 100 cells per original droplet diameter. The azimuthal
resolution is chosen such that the cells near the spherical droplet interface are close
to regular. Previous testing of grid resolution sensitivities in 2D (Meng & Colonius
2015; Meng 2016) suggests the present spatial resolution captures the salient flow
features without being computationally cumbersome. The simulation is performed with
a constant Courant–Friedrichs–Lewy (CFL) number of 0.2. Since the initial condition
is axisymmetric, the gas is initially seeded with small random radial and azimuthal
velocity perturbations, the largest of which have approximate magnitudes of O(10
4
u
s
),
where u
s
is the post-shock gas velocity. These white-noise-type perturbations are
generated via the Fortran compiler’s intrinsic random number generator, and are
applied to both the pre-shock and post-shock gas in the initial condition.
2.2. Governing equations
We model the flow with the compressible multicomponent Euler equations. In addition
to being compressible, each fluid is considered immiscible and does not undergo phase
change. In the absence of mass transfer and surface tension, material interfaces are
Downloaded from https://www.cambridge.org/core. Caltech Library, on 20 Dec 2017 at 20:54:57, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2017.804

1112 J. C. Meng and T. Colonius
simply advected by the local flow velocity. For a system of two fluids, this gives
rise to a five-equation model, introduced in its inviscid form by Allaire, Clerc &
Kokh (2002). Implications of the choice to neglect viscous and capillary effects are
addressed in detail in § 3.2. The ve-equation model, (2.1), consists of individual
continuity equations for each of the fluids, (2.1a) and (2.1b), mixture momentum,
(2.1c), and energy, (2.1d), equations, and a transport equation for the gas volume
fraction, (2.1e):
g
ρ
g
)
t
+ ·
g
ρ
g
u) = 0, (2.1a)
l
ρ
l
)
t
+ ·
l
ρ
l
u) = 0, (2.1b)
u)
t
+ · u u + pI) = 0, (2.1c)
E
t
+ · ((E + p)u) = 0, (2.1d)
α
g
t
+ u · α
g
= 0, (2.1e)
where ρ is the density, α is the volume fraction, u is the velocity vector, p is the
pressure, E is the total energy defined as E = ρε + (1/2kuk
2
, ε is the specific
internal energy, and the subscripted g, l represent, respectively, the gas and liquid
fluids. Each equation in the ve-equation model, (2.1), is evolved independently of
all others. While alternate models for multicomponent flows exist within the literature
(e.g. Kapila et al. 2001; Murrone & Guillard 2005; Saurel, Petitpas & Berry 2009;
Pelanti & Shyue 2014), the present model is chosen for its simplicity and desirable
conservation properties. The simplified interface model theoretically ensures volume
fraction positivity, and does not require regularization to fully specify shock jump
conditions. Though it is not capable of physically handling mixture regions, the model
is sufficient for the immiscible fluids assumed in the multicomponent flows of interest.
The simulation of material interfaces is made possible by the volume-of-fluid
method, which belongs to the broader class of interface-capturing schemes. One
common characteristic of these schemes is the relaxation of the natural sharpness
of material discontinuities. That is, the interfaces are allowed to numerically diffuse,
resulting in an interface region of small, but finite, thickness. Within this interface
region, a non-physical fluid mixture exists, which is appropriately treated using
mixture rules that are defined in § 2.2.1. For numerical stability purposes, material
interfaces are not initialized as sharp discontinuities, but are smeared over a few grid
cells. Based on previous results (Johnsen 2007; Coralic 2015) and sensitivity testing
by the authors for the specific problem of aerobreakup, this initial artificial smearing
over a few cells is known to have negligible impact on the computational results.
2.2.1. Equation of state
The five-equation model, (2.1), is closed with the specification of an appropriate
equation of state (EOS) that relates the fluid densities, pressures, and internal energies.
The stiffened gas EOS (Harlow & Amsden 1971), is used in our flow solver to model
both gases and liquids:
p = 1ε γ π
, (2.2)
Downloaded from https://www.cambridge.org/core. Caltech Library, on 20 Dec 2017 at 20:54:57, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1017/jfm.2017.804

Citations
More filters
Journal ArticleDOI

Rayleigh–Taylor and Richtmyer-Meshkov instability induced flow, turbulence, and mixing. II

TL;DR: In this article, Zhou et al. presented the initial condition dependence of Rayleigh-Taylor (RT) and Richtmyer-Meshkov (RM) mixing layers, and introduced parameters that are used to evaluate the level of mixedness and mixed mass within the layers.
Journal ArticleDOI

Supersonic spray combustion subject to scramjets: Progress and challenges

TL;DR: A review of the research on supersonic spray combustion that has been conducted in the past few decades and focuses on the key physiochemical processes and associated fluid physical mechanisms is provided in this article.
Journal ArticleDOI

An assessment of multicomponent flow models and interface capturing schemes for spherical bubble dynamics

TL;DR: The inadequacy of the traditional 5-equation model for spherical bubble collapse problems is demonstrated and the corresponding advantages of the augmented model of Kapila et al.
Journal ArticleDOI

Numerical symmetry-preserving techniques for low-dissipation shock-capturing schemes

TL;DR: It is shown that symmetry-breaking relates to vanishing numerical viscosity and is driven systematically by algorithmic floating-point effects which are no longer hidden by numerical dissipation, and a systematic procedure is proposed to deal with such errors by numerical and algorithmic formulations which respect floating- point arithmetic.
Journal ArticleDOI

Near-surface dynamics of a gas bubble collapsing above a crevice

TL;DR: In this article, the impact of a collapsing gas bubble above rigid, notched walls is considered, and a generic configuration is investigated numerically using a second-order accurate compressible multi-component flow solver in a two-dimensional axisymmetric coordinate system.
References
More filters
Journal ArticleDOI

Use of breakup time data and velocity history data to predict the maximum size of stable fragments for acceleration-induced breakup of a liquid drop

TL;DR: In this article, a triangular relationship based on the concept of a critical Weber number, breakup time data and velocity history data is presented which permits prediction of the maximum size of stable fragments.
Journal ArticleDOI

Near-limit drop deformation and secondary breakup

TL;DR: In this paper, the properties of drop deformation and secondary breakup were observed for shock wave initiated disturbances in air at normal temperature and pressure Test liquids included water, glycerol solutions, n-heptane, ethyl alcohol and mercury to yield Weber numbers (We) of 05-1000, Ohnesorge numbers (Oh) of 00006-4, liquid/gas density ratios of 580-12,000 and Reynolds numbers (Re) of 300-16,000.
Journal ArticleDOI

Two-phase modeling of deflagration-to-detonation transition in granular materials: Reduced equations

TL;DR: In this paper, the Baer-Nunziato model is reduced to a two-phase mixture model with unequal phase velocities and phase pressures, and the reduced models are hyperbolic and thermodynamically consistent with the parent model, but they cannot be expressed in conservation form and hence require a regularization in order to specify the jump conditions across shock waves.
Journal ArticleDOI

A five-equation model for the simulation of interfaces between compressible fluids

TL;DR: In this article, a diffuse-interface method is proposed for the simulation of interfaces between compressible fluids with general equations of state, including tabulated laws, and the interface is allowed to diffuse on a small number of computational cells.
Journal ArticleDOI

On the dynamics of a shock-bubble interaction

TL;DR: In this article, a detailed numerical study of the interaction of a weak shock wave with an isolated cylindrical gas inhomogeneity is presented, focusing on the early phases of interaction process which are dominated by repeated refractions and reflections of acoustic fronts at the bubble interface.
Related Papers (5)
Frequently Asked Questions (14)
Q1. What contributions have the authors mentioned in the paper "Numerical simulation of the aerobreakup of a water droplet" ?

The authors present a three-dimensional numerical simulation of the aerobreakup of a spherical water droplet in the flow behind a normal shock wave. Features of the droplet deformation and breakup in the stripping breakup regime, as well as descriptions of the surrounding gas flow, are discussed. 

The droplet morphology is first described and compared to experimental visualizations of the SIE process. Analyses of the instabilities arising on the liquid sheet reveal discrepancies with the proposed ‘ stretched streamwise ligament breakup ’ mechanism, while some qualitative evidence for the rise of RT instability along the accelerated sheet can be found. 

Due to the flattened disk-like shape of the deformed droplet, flow separation occurs for all practical values of the Reynolds number, Re = ρgugD0/µg, and the sheet thinning mechanism can be considered an inviscid phenomenon with no dependence on Re (Guildenbecher et al. 2009). 

due to the unsteady and nonlinear nature of the aerobreakup problem, this analysis is limited in its efficacy, and instead shows broadband instability growth of all modes, as would be expected from impulsive forcing of the system. 

Entrainment of shed vortices is also associated with a temporary increase in upstream jet velocity that results in a cyclic pumping of fluid onto the back side of the droplet. 

As the liquid sheet flaps, generating longitudinal ripples, the shear layers, which are subject to KH instability, periodically shed vortices that are either entrained by the wake recirculation region, or are convected downstream. 

In the absence of surface tension modelling, there does not exist a numerical mechanism that approximates capillary effects, i.e. there is no capillary counterpart to numerical viscosity. 

The simulation of material interfaces is made possible by the volume-of-fluid method, which belongs to the broader class of interface-capturing schemes. 

The study of droplet aerobreakup has historically been motivated by three applications: bulk dissemination of liquid agents, raindrop damage during supersonic flight, and secondary atomization of liquid jets in turbomachinery. 

The droplet’s unsteady drag coefficient, when normalized using the deformed droplet diameter, briefly recovers that for a rigid sphere during the very early stages of aerobreakup. 

due to the high computational costs of fully three-dimensional (3D) simulations, numerical aerobreakup studies have often invoked two-dimensional (2D) (Zaleski, Li & Succi 1995; Igra & Takayama 2001a,b,c; Chen 2008) orD ownl oade dfr omh ttps ://w ww .cam brid ge.o rg/c ore. 

Taking advantage of the type of quantitative analysis allowed by simulations, integral expressions have been derived (Meng & Colonius 2015) for the droplet’s centre-of-mass velocity and acceleration that minimize unnecessary noise that would be introduced by differentiating position data:xc =∫ Ωαlρlx dV∫ 

More recently, an analysis of the viscous KH instability by Theofanous et al. (2012) found that ‘[w]ave numbers and growth factors of [KH] instability are consistently greater than those of [RT] instability by more than an order of magnitude. 

Based on previous results (Johnsen 2007; Coralic 2015) and sensitivity testing by the authors for the specific problem of aerobreakup, this initial artificial smearing over a few cells is known to have negligible impact on the computational results.