2021 Singh PRFluids
2021 Singh PRFluids
2021 Singh PRFluids
The physical problem based on a shock-accelerated bubble has long been a fasci-
nating subject in the study of the Richtmyer-Meshkov (RM) instability. In this study
the behavior of a shock-accelerated heavy cylindrical bubble under the nonequilibrium
conditions of diatomic and polyatomic gases is investigated numerically. For this pur-
pose, a two-dimensional system of unsteady physical conservation laws derived from the
Boltzmann-Curtiss kinetic equations is solved by employing an explicit mixed-type modal
discontinuous Galerkin method with uniform meshes. For validation, the numerical results
are compared with available experimental and computational results, and are found to be in
good agreement. The results demonstrate that the effects of different physical properties,
including thermal nonequilibrium and bulk viscosity associated with the viscous excess
normal stress on diatomic and polyatomic gases, play a significant role in describing the
RM instability during the interaction between a planar shock wave and a heavy bubble.
The effects of diatomic and polyatomic gases result in a substantial change in the flow
morphology with complex wave patterns, vortex creation, vorticity generation, and bubble
deformation. In contrast to monatomic gas, the generation of larger rolled-up vortex chains,
a different kind of outward jet formation, and a large mixing zone with strong and large
expansion are observed in diatomic and polyatomic gases. A detailed study of the effects
of diatomic and polyatomic gases is investigated through the vorticity generation, degree
of nonequilibrium, the evolution of enstrophy, and dissipation rate. Furthermore, the time
variations of the shock trajectories and the interface scales are investigated from the view-
point of quantitative analysis. Finally, the effects of nonequilibrium parameters, including
bulk viscosity and index of inverse power law, are also investigated. The present work
can be seen as a supplement to the RM instability research to examine the nonequilibrium
effects of diatomic and polyatomic gases on the dynamics of a shock-accelerated heavy
cylindrical bubble.
DOI: 10.1103/PhysRevFluids.6.044001
I. INTRODUCTION
The interaction between shock wave and gas bubble interface has received considerable attention
in many natural and engineering applications. Studies include the inertial confinement fusion [1],
supernova explosions [2], medical shock wave lithotripsy [3], scramjet combustion process [4,5],
astrophysics [6], and so on. In such flows, when a shock wave interacts with a gas bubble inter-
face with different physical parameters, the coupling between them dominates the flow field and
produces a complex flow pattern. Because of this complexity, a disturbance is generated which
propagates along with the shock waves, resulting in a large deformation of bubble shape and
*
satyvir.singh@ntu.edu.sg
subsequent turbulent mixing which occurs along the gas interface owing to the mismatch of acoustic
impedances inside and outside of the bubble. This physical problem is known as shock-bubble
interaction or shock-accelerated bubble. During the interaction process, initially the gas bubble is
compressed and accelerated, then stretched by generating vortices at the interface between two fluids
having different densities, which leads to the Richtmyer-Meshkov (RM) instability [7,8]. The RM
instability plays a significant role in the creation of the bubble morphology due to the baroclinic
phenomena.
The physical problem of the shock-accelerated bubble has been regularly studied over the past
few decades to understand the physical structure of the RM instability. The flow configurations of
the shock-accelerated bubble essentially depend on the strength of the incident shock wave and the
Atwood number. If ρb and ρg are the densities of the bubble gas and its surrounding gas, respectively,
an Atwood number is then evaluated as
ρb − ρg
At = , −1 At 1. (1)
ρb + ρg
For At < 0, i.e., the bubble gas is lighter than the ambient gas, the transmitted shock wave within
the bubble propagates faster than the outside incident shock wave, and the bubble deforms into a
divergent configuration. Whereas, for At > 0, the transmitted shock wave propagates slower than
the incident shock wave, and the transmitted shock wave is the focus at the downstream pole of the
bubble and collapses into a single shock-focusing point with high pressure. Consequently, the bubble
deforms into a convergent configuration. The representative schematic view of flow morphology in
the shock-accelerated bubble with divergent pattern (At < 0) and convergent pattern (At > 0) is
illustrated in Fig. 1.
The study of a shock-accelerated bubble has been an important topic as it represents the closest
model to the RM instability. Throughout the past few decades, significant attempts have been made
to explain the physical phenomena of the shock-accelerated bubble experimentally, theoretically,
and numerically. The shock-accelerated light bubble problem was first investigated experimentally
in the 1960s by Markstein [9] and Rudinger and Somers [10] to describe a complex turbulent flow
field phenomenon with long-living vortex rings. Subsequently, for the spherical and cylindrical
with lighter and heavier gas cavities, Haas and Sturtevant [11] experimentally studied the shock-
accelerated bubble, and explored the wave pattern phenomenon for both divergent and convergent
situations. Later, Jacobs [12,13] experimentally studied the interaction between of a weak plane
shock in air and a helium column using a laser-induced fluorescence visualization technique.
Recently, Layes et al. [14,15] investigated experimentally the interaction between the plane shock
wave and spherical gas inhomogeneity soap bubbled using a high-speed camera shadowgraph
diagnostic. Balakumar et al. [16] studied for the first time the structure of the concentration and
velocity fields in an air–SF6 –air fluid layer subjected to an impulsive acceleration by a shock
wave Richtmyer-Meshkov instability using simultaneous particle-image velocimetry and planar
laser-induced fluorescence (PLIF). Later they studied experimentally and verified theoretically the
dependence of amplitude and wave number of initial perturbations on mixing in a shock-accelerated,
Richtmyer-Meshkov unstable fluid layer [17].
Although the shock-accelerated bubble is also the most numerically studied. Henderson et al. [18]
used a second-order Euler solver with an adaptive-mesh-refinement (AMR) scheme to investigate
the slow/fast refraction phenomena in the shock-accelerated bubble. Based on Haas and Sturtevant’s
experimental findings [11], Quirk and Karni [19] provided a detailed numerical investigation of
the shock-accelerated light/heavy bubble using a nonconservative shock-capturing scheme with the
AMR algorithm. Bagabir and Drikakis [20] investigated numerically the Mach number effects on
the shock-accelerated light bubble using high-resolution Godunov-type methods and an implicit
solver. Niederhaus et al. [21] provided an extensive numerical study of the shock-bubble interaction
for various combinations of heavy/light and light/heavy gases using a second-order piecewise linear
Eulerian Godunov code with the AMR algorithm. Hejazialhosseini et al. [22] presented a detailed
high resolution simulation for the visualization of both density and vorticity magnitude fields in the
044001-2
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 1. Representative schematic view of flow morphology in a shock-accelerated cylindrical bubble from
left to right direction for (a) divergent pattern with At < 0 case, and (b) convergent pattern with At > 0 case:
initial condition before interaction (left), during transmitted shock transit (middle), and immediately after
transmitted shock transit (right). IS incident shock; OI origin interface; RS reflected shock; TS1 first transmitted
shock; TS2 secondary transmitted shock; UZ unshocked zone; SF shock focusing; DS diffracted shock; RTS
reflected transmitted shock; MS Mach stem; TP triple point; SS slip surface.
shock-bubble interaction at different Mach numbers. Kundu and De [23] investigated numerically
the shock-accelerated Refrigerant-22 bubble problem using third-, fifth-, and ninth-order accurate
up-wind schemes and a fifth-order accurate weighted essentially nonoscillatory (WENO) scheme
by solving Navier-Stokes equations. Zhu et al. [24] simulated numerically the shock-heavy bubble
interaction using a high-order WENO scheme and studied the effects of different incident shock
Mach numbers on the shock-accelerated bubble dynamics. Wang et al. [25] studied numerically and
theoretically the vortex breakdown behavior of the scaling criterion in a shock-cylindrical bubble
interaction, and found that viscosity leads to the vortex breakdown. Recently, Singh [26] investigated
numerically the impacts of Atwood number on the flow morphology of a planar shock-accelerated
square bubble containing different gases.
Several existing theoretical and numerical studies on the shock-accelerated bubble have been
investigated assuming the negligible thermal nonequilibrium effects which are directly connected to
the internal modes of diatomic and polyatomic gases—the rotational and vibrational modes [27,28].
These studies are basically focused on the Euler or Navier-Stokes-Fourier (NSF) equations which
are presumed to be the de facto mathematical models for any possible flow problem, including
compressible flows. The NSF theory is developed upon a critical assumption proposed by Stokes
[29], so-called the Stokes’ hypothesis, in which the bulk viscosity is zero:
2 2
μb ≡ λ + μ = 0, equivalently, λ = − μ, (2)
3 3
where μ, λ, and μb represents the shear viscosity, the second coefficient of viscosity, and the bulk
viscosity of the fluid, respectively. In this hypothesis, the dilatational term (∇ · u) is assumed having
044001-3
SATYVIR SINGH AND MARCO BATTIATO
FIG. 2. Problem setup: schematic diagram of initial flow field and computational domain.
no important role in the viscous stress part although it may play a significant role in the compressible
gas flows. While the theory of the Stokes is definitely valid in the case of monatomic gases such as
argon, there is growing proof now that this is not the case with diatomic and polyatomic gases such
as nitrogen, methane, and carbon dioxide [27,28,30–36]. The viscous internal shock structure in
diatomic and polyatomic gases and the hypersonic penetration into the Mars atmosphere consisting
primarily of carbon dioxide provide examples of these circumstances.
Encouraged by these developments, in this study, numerical simulations have been conducted to
investigate the nonequilibrium behavior of a shock-accelerated heavy cylindrical bubble in diatomic
and polyatomic gases. To the best knowledge of the authors, no attempt has been made in the
past to examine the nonequilibrium behaviors of diatomic and polyatomic gases on shock-heavy
cylindrical bubble interaction. Furthermore, the current work can be seen as a supplement to the
RM instability research to examine the nonequilibrium effects of diatomic and polyatomic gases
on the dynamics of a shock-bubble interaction. The paper is structured as follows: In Sec. II the
problem description, computational model, and numerical method are presented. In Sec. III the grid
refinement analysis and the validation of numerical method are studied. In Sec. IV the computational
results for the shock-accelerated heavy cylindrical bubble in diatomic and polyatomic gases are
discussed in detail. First, the overall flow morphology visualization are presented. After that, the
effects of diatomic and polyatomic gases on the shock-accelerated heavy cylindrical bubble are
systematically analyzed to illustrate the differences in contrast to monatomic gas. Consequently,
the effects of nonequilibrium parameters, including bulk viscosity and index of inverse power law,
are conducted. Finally, concluding remarks with further development in line with the present study
are given in Sec. V.
044001-4
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
previous research works for the RM instability study. Therefore, we also consider here the SF6
gas with the density ρg = 5.97 kg/m3 filled inside the cylindrical bubble. For ambient gas, three
different gases—monatomic argon, diatomic nitrogen, and polyatomic methane—are considered in
this study. The left boundary is set to be inflow, while the upper, bottom, and right boundaries are
set to be outflow boundaries.
B. Governing equations
The shock-accelerated inhomogeneous flow problems are basically solved by simulating the
compressible multispecies flow model with a gas mixture [37,38]. Interestingly, it has been studied
in the previous research works of Picone and Boris [39], Samtaney and Zabusky [40], Quirk and
Karni [19], and Bagabir and Drikakis [20] that the different values of specific heat capacities γ
values for each gas does not affect the details of the vorticity generation qualitatively, particularly
the creation of large-scale structures. According to Quirk and Karni [19], for the problem of shock-
accelerated bubble “...the errors introduced by the single-gas model assumption are not catastrophic
and to some extent may be tolerated,” but such errors cannot be tolerated in other applications,
for example, air-fuel mixing in a supersonic combustion system, where temperature changes will
substantially affect the mixing. Recently, Latini and Schilling [41] investigated numerically the
growth dynamics of two- and three-dimensional single-mode reshocked air/SF6 RM instability
by considering a single specific heat ratio γ value. Based on the discussion mentioned above,
the present shock-accelerated SF6 cylindrical bubble problem is modeled through the unsteady
compressible laminar flow which assumes a single-component perfect gas with specific heat
ratio γ .
Motivated by the pioneer work of Curtiss [42], the two-dimensional conservation laws for un-
steady compressible viscous fluids derived rigorously from the Boltzmann-Curtiss kinetic equation
for diatomic and polyatomic gases can be written in dimensionless form as [27,28]
∂U
+ ∇ · Finv (U) + ∇ · Fvis (U, , Q, ) = 0, (3)
∂t
where the conservation vector U, the inviscid flux vector Finv , and the viscous flux vector Fvis are
defined as
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
ρ ρu 0
⎢ ⎥ 1 ⎣
U = ⎣ ρu ⎦, Finv = ⎣ ρuu + γ M 2 pI ⎦, Fvis =
1
+ I ⎦. (4)
Re
ρE ρE + 2 p u
1 ( + I) · u + 1
Q
γM E cPr
Here ρ is the mass density, u = (u, v) is the velocity vector, p is the pressure, γ is the specific heat
ratio, I is the unit second rank tensor, E is the total energy, stands for the stress tensor, is the
excess normal stress, and Q is the heat flux. The total energy density is defined as
p 1
ρE = + ρ|u|2 . (5)
γ (γ − 1)M 2 2
The dimensionless parameters such as Mach number (M ), Reynolds number (Re), Eckert number
(E c), and Prandtl number (Pr) can be defined as
ur ρr ur L Cp r μr
M≡√ , Re ≡ , E c ≡ (γ − 1)M 2 , Pr ≡ , (6)
γ Ru Tr μr kr
where the subscript r stands for the reference state, L denotes the reference length, Ru is the universal
gas constant, and Cp denotes the heat capacity per mass at constant pressure. In the present study the
sound speed of diatomic or polyatomic gas before the shock wave is chosen as the reference velocity
ur , resulting in M = 1. Nonetheless, we keep the reference Mach number M in the equations in
order to show its role in the formulation. Also, the bubble radius is used for the reference length,
i.e., L = Rb = 25 mm. The nondimensional equation of state is expressed as ρr = Pr /Tr , where
044001-5
SATYVIR SINGH AND MARCO BATTIATO
TABLE I. Physical properties of monatomic, diatomic, and polyatomic gases at standard conditions.
Pr = 101, 325 Pa and Tr = 273 K are used for the reference pressure and temperature, respectively.
The specific heat ratio γ is considered to be 1.667 for argon gas, 1.4 for nitrogen gas, and 1.31 for
methane gas. The Prandtl number (Pr) can be also determined by means of Eucken’s relationship
4γ
Pr = . (7)
9γ − 5
According to the classical Navier-Fourier model [27,28,31], the stress tensor (), the excess
normal vector (), and the heat flux vector (Q) are evaluated by the Newtonian law of shear and
bulk viscosity, and the Fourier law of heat conduction, respectively, as follows:
= −2μ[∇u]2 , = −μb ∇ · u, Q = −k∇T, (8)
where T is the absolute temperature, and the symbols μ, μb , and k represent the Chapman-Enskog
shear viscosity, the bulk viscosity, and the thermal conductivity, respectively. These expressions for
the Chapman-Enskog linear transport coefficients are employed as [31]
μ = T s, μb = fb μ, k = T s+1 , (9)
where s stands for the index of the inverse power laws of gas molecules, given as
1 2
s= + . (10)
2 υ −1
Here the parameter υ is the exponent of the inverse power laws for the gas particle interaction
potentials. The value of s is assumed to be 0.81 for argon gas, 0.78 for nitrogen, and 0.84 for
methane gas [43]. The factor fb = μb /μ is the ratio of the bulk viscosity to the shear viscosity. Its
value can be determined using a sound wave absorption measurement. The viscosity ratio fb values
for argon, nitrogen, and methane gases are considered to be 0.0, 0.8, and 1.33, respectively, based
on experiments [44]. Three type of gases—monatomic argon, diatomic nitrogen, and polyatomic
methane—are considered as the ambient gases in this work and the physical properties of these
gases at standard conditions are displayed in Table I.
In the above expressions, Ms denotes the shock Mach number. Here the value of shock Mach number
is taken as 1.22. Subscript 1 and 2 denote to the left- and right-hand sides of the shock wave,
respectively.
044001-6
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
044001-7
SATYVIR SINGH AND MARCO BATTIATO
FIG. 3. Grid refinement study: numerical schlieren images with different mesh sizes for shock-accelerated
SF6 cylindrical bubble surrounded by nitrogen gas at time t = 8.
mesh points 1600×800. Furthermore, the density distribution profiles along the centerline of the
computed bubble are also illustrated in Fig. 4 to demonstrate the grid sensitivity. The results show
that Mesh 6 with mesh points 1600×800 is very close to the asymptotic range. Based on the
above-mentioned finding, all the numerical simulations are performed on Mesh 6.
FIG. 4. Grid refinement study: density distribution profiles for different mesh sizes along the centerline of
the shock-accelerated SF6 cylindrical bubble surrounded by nitrogen gas at time t = 8.
044001-8
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 5. Validation of the numerical solver: comparison of numerical schlieren images between the experi-
mental results of Hass and Sturtevant [11], the computational results of Quick and Karni [19], and the present
numerical results for a shock-accelerated R22 cylindrical bubble surrounded by air at different time instants.
044001-9
SATYVIR SINGH AND MARCO BATTIATO
FIG. 6. Validation of the numerical solver: comparison of computed characteristic interface points (IS, UI,
DI, TS, and RS) between the experimental results of Hass and Sturtevant [11] and the present numerical results
for a shock-accelerated R22 cylindrical bubble surrounded by air. The circle (◦) symbol and the solid line (−)
represent the experimental and present computational results, respectively.
considered as the surrounding gas of the bubble. The physical properties of these gases are illustrated
in Table I, while, the evaluated Atwood numbers for ambient gas-bubble gas configurations are listed
in Table II.
This section is divided into three parts: In the first part, the overall flow morphology of a shock-
accelerated SF6 cylindrical bubble in nitrogen gas is discussed in depth. In the second part, the
effects of diatomic and polyatomic gases on the flow morphology of the shock-accelerated SF6
cylindrical bubble are investigated systematically to highlight the differences in comparison with
monatomic gas. This investigation includes the flow morphology visualization, vorticity generation,
degree of thermal nonequilibrium, enstrophy and dissipation rate, and the quantitative analysis based
on the shock trajectories and interface features for deeper understanding of the effects in diatomic
and polyatomic gases. Finally, the effects of nonequilibrium parameters, including bulk viscosity
and index of inverse power law on shock-accelerated SF6 cylindrical bubble are investigated in the
last part.
044001-10
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 7. Overall flow morphology visualization of the shock-accelerated SF6 cylindrical bubble in nitrogen
gas: time evolution of density contour (top), and numerical schlieren images (bottom).
cylindrical SF6 bubble surrounded by nitrogen gas ( fb = 0.8) accelerated by a planar incident shock
wave with Ms = 1.22 at different time instants. The propagation speed of the incident shock wave
inside the SF6 cylindrical bubble is smaller than that of the outside medium due to the larger acoustic
impedance (i.e., Z = ρ c). As the incident shock (IS) wave passes through the cylindrical bubble,
a curved transmitted shock (TS1) wave inside the bubble is observed which takes the convergent
shape owing to the difference in acoustic impedance. Before interaction with the IS wave, the initial
state of the SF6 cylindrical bubble interface can be clearly observed (t = 0).
When the IS wave reaches the bubble interface, the bubble begins to compress inwards and the
TS1 wave propagating downstream inside the bubble is generated and a reflected shock (RS) wave
travels upstream simultaneously (t = 2). During this period, an unshocked zone (UZ) is also gener-
ated inside the bubble. At the beginning, the TS1 wave is slightly converging and can be clearly seen
inside the volume and the compressed volume forms a crescent shape (t = 3–5). When the IS wave
passes across the top point of the bubble interface, the curvature of the TS1 wave becomes larger.
A curved diffracted shock (DS) wave outside the bubble is also observed to be connected with the
top-and-bottom ends of the two straight sections of the IS wave. The curved TS1 continuously stands
back the IS wave that moves through outside bubble. Later, the ends of the DS wave that linked with
the IS wave are strongly distorted outside the bubble. It is observed that when the TS1 wave crosses
the bubble, the part of the bubble-ambient gas interface behind it is thicker than the undisturbed
interface in front of it. This is because of the roll-up of small scale vortices that are produced by the
baroclinic vorticity phenomena, i.e., the local pressure and density gradient misalignment.
044001-11
SATYVIR SINGH AND MARCO BATTIATO
FIG. 8. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble: density
contours (top), and numerical schlieren images (bottom) in (a) argon gas ( fb = 0), (b) nitrogen gas ( fb = 0.8),
and (c) methane gas ( fb = 1.33) at different time instants.
As time passes, a shock-focusing (SF) zone is created towards the downstream end point of the
bubble on its axis due to the converging the DS and TS1 waves (t = 5.3). Later, the TS1 wave
converges to a very small zone at the interior downstream interface (t = 5.5–6.0). Meanwhile, a
reversed transmitted shock (RTS) wave is observed in the downstream interior bubble, which then
moves upstream simultaneously. As the IS wave moves ahead, a Mach configuration [Mach stem
(MS), and triple point (TP)] is generated in the downstream interior bubble surface (t = 6–6.5).
After focusing, the diffracted shock emerges from the bubble downstream end and a secondary trans-
mitted shock (TS2) wave is generated which expands outward downstream (t = 6.7). Subsequently,
a bulge is generated at the gas interface (t = 7). Shortly afterward, the bulge moves downstream,
and consequently, an outward jet is formed which is driven by the peak pressure caused by the
complex interaction among shocks inside the bubble near the centerline of the downstream (t = 8).
This physical phenomena of an outward jet formation is also observed in the past research studied
[19,23,24]. Also, the size of rolled-up vortices increases and appear conspicuously at the interface
between the bubble and surrounding gas. The outward jet grows along the generated rolled-up
vortices with time (t = 8–20).
044001-12
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 9. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble: pressure
contours in (a) argon gas ( fb = 0), (b) nitrogen gas ( fb = 0.8), and (c) methane gas ( fb = 1.33) at different
time instants.
observed as sluggish and weakly due to the high density of the ambient gas as shown in Fig. 8(a).
As the IS wave passes, a reversed transmitted shock (RTS) wave is observed inside bubble, which
moves upstream simultaneously. Subsequently, a small size of sharp outward jet is found in the
centerline near the right interface of the bubble. In case of nitrogen gas ( fb = 0.8), the interaction
process become stronger, and the outward jet becomes larger in comparison with the argon gas, as
shown in Fig. 8(b). In addition, the size of rolled-up vortices increases with intense strength and
appear conspicuously at the interface between the bubble and surrounding gas due to the baroclinic
vorticity deposition. While, in methane gas ( fb = 1.33), the certain areas near the bubble interface
becomes larger at a long time, as shown in Fig. 8(c). Interestingly, a sharp outward SF6 jet in
argon gas is observed, while a different kind of SF6 outward jet having a small structure on the jet
interface due to the second instabilities is formed in nitrogen and methane gases. Later on, the SF6
jet is elongated with its head weakening because of the diffusion and the mixing of two gases and
the interface develops strong vortex rings (t = 10–15). Then, the interface breaks up and lots of
small scale structures are generated because of the Kelvin-Helmholtz instability.
Figure 9 shows the effects of diatomic and polyatomic gases on the pressure flow field of the
shock-accelerated SF6 cylindrical bubble at four different time instants t = 4, 5, 6, and 7. It can
be easily seen that there are significant differences in the pressure flow field for the different
gases after the interaction. In order to explore the effects of diatomic and polyatomic gases on the
shock-accelerated SF6 cylindrical bubble, the pressure distribution is displayed along the centerline
of the computed bubble after the IS wave impacts as shown in Fig. 10. As seen from the pressure
profile, the fastest shock propagation inside the bubble is observed in methane gas, while the
slowest shock propagation is noticed in argon gas. After the interaction of the DS wave, a high
044001-13
SATYVIR SINGH AND MARCO BATTIATO
FIG. 10. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble: pres-
sure distribution obtained along the centerline of the computed bubble in (a) argon gas ( fb = 0), (b) nitrogen
gas ( fb = 0.8), and (c) methane gas ( fb = 1.33) at different time instants.
pressure region is created in downstream interface which can be visualized in Fig. 9. This effect
only continues few moments and is disappeared because of the TS1 wave interaction and the
pressure diffusion outside of the bubble. The downstream interface travels upwards, guided by the
comparatively higher pressures. The highest pressure is produced when the TS1 and DS waves meet
each other within the bubble and the unshocked zone (UZ) disappears as seen in Fig. 7. The largest
pressure jump is witnessed in methane gas, while the smallest pressure jump is observed in argon
gas. The internal pressures of the bubble decrease constantly but are much higher than external
pressures with the divergence of the generated shocks as shown in Fig. 10. The produced outward
jet in shock-accelerated bubble is closely related to the pressure perturbation.
Furthermore, the effects of diatomic and polyatomic gases can be extended to visualize the evo-
lution of the bubble shape, in particular, generation of rolled-up vortices and jets during interaction
with the IS wave, as shown in Fig. 11. The bubble appears to be compressed by the IS wave along
the x direction, and its top and bottom edges have been pushed forward near the horizontal axis
of symmetry as compared to its middle section. At the beginning of interaction, this compression
starts as soon as the IS wave hits the upstream end of the bubble. For all three gases, the downstream
interface hardly moves, while the upstream side presses forward under the influence of the IS wave
at the early instants, as shown in Figs. 11(a)–11(c). As time goes, the upper and lower interfaces of
the bubble also fold toward the x-axis, and the bubble takes the convergent shape. Some small scale
rolled-up vortices are generated on the upper and lower interfaces due to the baroclinic vorticity
044001-14
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 11. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble:
evolution of the bubble shape showing early compression in (a) argon gas ( fb = 0), (b) nitrogen gas ( fb = 0.8),
and (c) methane gas ( fb = 1.33).
generation. In case of argon gas, the size of rolled-up vortices are smaller in comparison to nitrogen
and methane gases. These rolled-up vortices are continuously increasing over time. At later times,
the flow field is completely controlled by the rolled-up vortices. A different type of outward jets
is observed near the downstream boundary in different ambient gases which gradually grows with
time, as shown in Figs. 11(a)–11(c).
2. Vorticity generation
In the study of a shock-accelerated cylindrical bubble, the vorticity occurring due to the mis-
alignment of the pressure and the density gradients plays a vital role in understanding the flow
morphology phenomena during the interaction process. When the IS wave crosses the bubble,
vorticity is generated in the flow field. It can be expressed as the curl of the velocity
ω = ∇ × u. (14)
The essential physics occurred during interaction between the IS wave and the gas bubble can be
explained by the equation of vorticity transport. This equation includes several physically distinctive
044001-15
SATYVIR SINGH AND MARCO BATTIATO
FIG. 12. Schematic diagram of the vorticity generation on the interface of the heavy cylindrical bubble
during and after initial shock wave transit.
044001-16
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 13. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble:
contours of vorticity distribution in (a) argon gas ( fb = 0), (b) nitrogen gas ( fb = 0.8), and (c) methane gas
( fb = 1.33) at different time instants.
In a shock-accelerated bubble, the vorticity is basically generated at the bubble interface due
to the misalignment of the density and pressure gradients. Figure 13 illustrates the effects of the
diatomic and polyatomic gases on the vorticity distribution of the shock-accelerated SF6 cylindrical
bubble at different time instants. Initially, the vorticity is equal to zero everywhere. When the IS
wave passes across the bubble, the baroclinic vorticity is deposited locally on the bubble interface,
where the discontinuity between SF6 and the ambient gas exists. At the top and bottom locations of
the bubble where the density and pressure gradients are orthogonal, the magnitude of the vorticity
is maximum and it is zero at the interface along the axis of the bubble where the density and
pressure gradients are collinear. According to the direction of the IS wave (from left to right), a
significant quantity of negative vorticity is generated at the upper-half of the bubble interface, while
a significant quantity of positive vorticity at the lower-half of the bubble interface. After shock
focusing, the deformed shock wave produces a small quantity of positive (negative) vorticity in
the filaments connecting the rolled-up vortices. Also, some quantity of positive (negative) vorticity
is gathered on the upper-half (lower-half) plane of the generated outward jet head. The potential
explanation for the gathering of dominant positive (negative) vorticity in the outward jet heads on the
upper-half (lower-half) plane is the higher amplitude attenuation of the deformed shock wave when
it hits the downstream pole of the bubble. This quantity of positive (negative) vorticity on the upper
(lower) jet head promotes the outward jet movement, which implies the relationship between the jet
formation and the vorticity deposition. It can be observed that there are substantial gaps in vorticity
distribution for the various gases after the interaction. For argon gas, a small quantity of vorticity
occurs at the generated rolled-up vortices on the bubble interface, as shown in Fig. 13(a). These
rolled-up vortices are more pronounced for nitrogen and methane gases, as seen in Figs. 13(b)–13(c).
044001-17
SATYVIR SINGH AND MARCO BATTIATO
When the IS wave impact with bubble at about t = 10–15, the vorticity along the jet is diffused more
and the formed rolled-up vortices accumulate a significant quantity of vorticity. In conclusion, the
generation and distribution of vorticity plays a dominant role in the diatomic and polyatomic gases
when the rolled-up vortices are formed.
The vorticity on the bubble interface plays a crucial role in gas mixing inside and outside the
bubble. Therefore, in order to obtain a better understanding on the physics of the vorticity generation
in diatomic and polyatomic gases, a detailed investigation can be further presented based on the
four important spatial integrated fields; (i) average vorticity (ωav ), (ii) absolute dilatation vorticity
(|ωdil |), (iii) absolute baroclinic vorticity (|ωbar |), and (iv) absolute viscous vorticity (|ωvis |). The
spatial integrated field of average vorticity is defined as
|ω|dxdy
ωav (t ) = D , (16)
D dxdy
where D represents the entire computational domain. The spatial integrated field of absolute
dilatation vorticity is computed as follows:
|ωdil |(t ) = − |ω(∇ · u)|dxdy. (17)
D
The spatial integrated field of absolute baroclinic vorticity is given as
1
|ωbar |(t ) = (∇ρ × ∇ p)dxdy. (18)
ρ2
D
Figure 14 illustrates the effects of diatomic and polyatomic gases on the spatial integrated fields
of average vorticity, absolute dilatational vorticity, absolute baroclinic vorticity, and absolute vis-
cous vorticity in the shock-accelerated SF6 cylindrical bubble. It can be observed that the values of
these spatial integrated fields in argon gas are the smallest among the three gases when the incident
and reflected shock waves collide with the bubble. It is obvious that the spatial integrated fields
are substantially enhanced in case of methane gas. Besides, these spatial integrated fields increase
with time in all three cases, which implies more and more ambient gas has been entrained into
the distorted SF6 cylindrical bubble. The vortices produced by the shock wave-bubble interaction
encourage the mixing of ambient gas and SF6 cylindrical bubble. When the RS wave impinges
on the distorted SF6 bubble again, the growth rate of the spatial integrated fields is the largest,
which indicates that the vorticities are greatly raised during this period as shown in Fig. 14. Then,
the growth rate slows down due to the influence of highly raised viscosity in the flow field. Thus,
it may be concluded that the evolution of the spatial integrated fields for vorticity and its related
components are not a simply monotonic relationship with the ambient gas.
044001-18
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 14. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble: spatial
integrated fields of (a) average vorticity (ωav ), (b) absolute dilatational vorticity (|ωdil |), (c) absolute baroclinic
vorticity (|ωbar |), and (d) absolute viscous vorticity (|ωvis |).
044001-19
SATYVIR SINGH AND MARCO BATTIATO
FIG. 15. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble:
contours of (a) degree of thermal nonequilibrium (R), (b) magnitude of viscous stress (||), (c) excess normal
stress (||), and (d) magnitude of heat flux (|Q|) in argon gas (left), nitrogen gas (middle), and methane gas
(right) at time t = 10.
on the spatial integrated fields during the interaction. These spatial integrated fields are computed
as follows. Figure 16 shows the spatial integrated fields of the degree of thermal nonequilibrium
044001-20
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 16. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble: spatial
integrated fields of (a) degree of thermal nonequilibrium (R), (b) magnitude of viscous stress (||), (c) excess
normal stress (||), and (d) magnitude of heat flux (|Q|).
and its associated components. The results reveal that during the interaction, the degree of thermal
nonequilibrium and its associated components attain significant values in diatomic and polyatomic
gases after passing the bubble. It can also be seen that viscous stress and excess normal stress are the
most prevailing contributors of the degree of thermal nonequilibrium during the shock-accelerated
SF6 cylindrical bubble, accompanied by heat flux.
Moreover, the viscous effects of diatomic and polyatomic gases can be investigated by introducing
the area-weighted dissipation rate of kinetic energy,
ε(t ) = E (x, y, t )dxdy, (22)
D
044001-21
SATYVIR SINGH AND MARCO BATTIATO
FIG. 17. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble:
contours of (a) enstrophy ( ) and (b) dissipation rate (ε) in argon gas (left), nitrogen gas (middle), and methane
gas (right) at time t = 10.
where E (x, y, t ) represents the dissipation rate per unit volume which is defined as
Here ii is the viscous shear stress, is the excess normal stress, and Si j is the strain rate defined
as Si j = ∂ui /∂x j .
Figure 17 illustrates the effects of diatomic and polyatomic gases on enstrophy and dissipation
rate in the shock-accelerated SF6 cylindrical bubble at time t = 10. It can be easily observed
that there are significant differences in enstrophy and dissipation rate for the different gases after
the interaction. The higher values of enstrophy and dissipation rate are found inside the created
rolled-up vortices of the deformed bubble interface. In contrast to monatomic gas, the enstrophy and
dissipation rate during the interaction are substantially enhanced in diatomic and polyatomic gases.
In order to further explore the effects of diatomic and polyatomic gases on the shock-accelerated
SF6 cylindrical bubble, the spatial integrated fields of enstrophy and dissipation rate are computed as
shown in Fig. 18. The enstrophy and dissipation rate during the interaction can be seen substantially
improved for diatomic and polyatomic gases in contrast to monatomic gas. For diatomic and
polyatomic gases, trends of these evolution profiles are basically the same. To be specific, the
enstrophy increases when the IS and RS waves impinge on the bubbles. Subsequently, the enhanced
vorticities promote the mixing of gases inside and outside the gas bubble and thus accelerate the
transfer and consumption of vorticity energy, which can weaken the enstrophy intensity in the bubble
region gradually, as shown in Fig. 18(a). This physical phenomena can also be confirmed in the time
evolution of the dissipation rate as shown in Fig. 18(b). Moreover, nonmonotonic trends are found
in enstrophy and dissipation rate throughout the interaction process.
044001-22
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 18. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble: spatial
integrated fields of (a) enstrophy ( ) and (b) dissipation rate (ε).
044001-23
SATYVIR SINGH AND MARCO BATTIATO
FIG. 19. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble: shock
trajectories (IS, ISB, and TS1) for the computed bubble in (a) argon gas ( fb = 0), (b) nitrogen gas ( fb = 0.8),
and (c) methane gas ( fb = 1.33). The definitions of the IS, ISB, and TS1 waves are inserted.
FIG. 20. Effects of diatomic and polyatomic gases on the shock-accelerated SF6 cylindrical bubble:
temporal variations of the interfacial characteristic scales—(a) the height and (b) the length of the evolving
interface for the computed bubble. The definitions of the height and the length of the evolving interface are
inserted.
044001-24
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
TABLE III. Selected nonequilibrium parameters in monatomic, diatomic, and polyatomic gases.
specific values of fb and s in argon, nitrogen, and methane gases are considered, as shown in
Table III. For visualizing the bulk viscosity effect in diatomic and polyatomic gases, the contours
of density, numerical schlieren images, enstrophy, and dissipation rate at time t = 10 are illustrated
in Fig. 21. The numerical results are presented for fb = 0, 0.8 with same s = 0.74 in nitrogen gas,
while the values fb = 0, 1.33 with same s = 0.84 are chosen for methane gas. It can be visualized as
a non-negligible difference on the flow morphology of a shock-accelerated SF6 cylindrical bubble, in
particular, creation of the outward jet and the rolled-up vortices based on the absence and presence
of the excess normal stress, i.e., = 0 and = 0, as shown in Fig. 21(a). Note that the excess
normal stress is proportional to fb as = − fb μ∇ · u. Interestingly, the deformed bubble interface is
found more elongated vertically at fb = 0 in comparison to fb = 0. Moreover, a significant increase
in enstrophy and dissipation rate is also observed with increasing fb value in both nitrogen and
methane gases, as seen in Figs. 21(b)–21(c).
FIG. 21. Effects of bulk viscosity ratio ( fb ) on the shock-accelerated SF6 cylindrical bubble: contours of
(a) density (top) and numerical schlieren images (bottom), (b) enstrophy ( ), and (c) dissipation rate (ε) in
nitrogen (s = 0.74) and methane (s = 0.84) gases at time t = 10.
044001-25
SATYVIR SINGH AND MARCO BATTIATO
FIG. 22. Effects of bulk viscosity ( fb ) on the shock-accelerated SF6 cylindrical bubble: spatial integrated
fields of (a) enstrophy ( ), (b) dissipation rate (ε), (c) average vorticity (ωav ), and (d) absolute baroclinic
vorticity (|ωbar |) in nitrogen (s = 0.74) and methane (s = 0.81) gases.
The nonequilibrium effects of bulk viscosity ratio can be further seen in the spatial integrated
fields of physical quantities—enstrophy, dissipation rate, average vorticity, and absolute baroclinic
vorticity, as shown in Fig. 22. From the evolution profiles, it is easy to find that a higher value of fb
value has a greater influence on the flow morphology which are quite responsible for nonequilibrium
phenomena in diatomic and polyatomic gases. Finally, the nonequilibrium effects of index of inverse
power law (s) on the shock-accelerated SF6 cylindrical bubble in diatomic and polyatomic gases are
investigated. Figure 23 shows the spatial integrated fields of degree of thermal nonequilibrium in
argon ( fb = 0), nitrogen ( fb = 0.8), and methane ( fb = 1.33) gases during the interaction process.
The results show that the values of degree of thermal nonequilibrium are observed in a decreasing
trend when the s value increases in all three gases.
V. CONCLUDING REMARKS
The physical problem based on a shock-accelerated bubble has long been a fascinating sub-
ject in the study of the Richtmyer-Meshkov (RM) instability. In this study the behavior of a
shock-accelerated heavy cylindrical bubble under the nonequilibrium conditions of diatomic and
polyatomic gases is investigated numerically. The emphasis is placed on the flow morphology visu-
alization, wave patterns, vorticity generation, interface movement and deformation, and quantitative
044001-26
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
FIG. 23. Effects of index of inverse power law (s) on the shock-accelerated SF6 cylindrical bubble: spatial
integrated fields of degree of nonequilibrium (R) in (a) argon gas ( fb = 0), (b) nitrogen gas ( fb = 0.8), and
methane gas ( fb = 1.33).
analysis. For this purpose, a two-dimensional system of unsteady physical conservation laws derived
from the Boltzmann-Curtiss kinetic equations is solved by employing an explicit mixed-type modal
discontinuous Galerkin method with uniform meshes. The computational model is validated by the
previous experimental and computational results for the shock-accelerated heavy cylindrical bubble
problem and are found to be in good agreement.
The numerical results demonstrate that the effects of different physical properties, including
thermal nonequilibrium and bulk viscosity associated with the viscous excess normal stress on
diatomic and polyatomic gases play a significant role in describing the RM instability during the
interaction between a planar shock wave and a heavy bubble. It is a possibility that other physical
properties, such as density, specific heat ratio, and so on, also have a contribution to the present
observations. The effects of diatomic and polyatomic gases result in a substantial change in the
flow morphology with complex wave patterns, vortex creation, vorticity generation, and bubble
deformation. In contrast to monatomic gas, the generation of larger rolled-up vortex chains, different
kinds of outward jet formation, and a large mixing zone with strong and large expansion are
observed in diatomic and polyatomic gases.
A detailed study of the effects of diatomic and polyatomic gases is investigated through vor-
ticity generation, degree of nonequilibrium, the evolution of enstrophy, and dissipation rate. It is
044001-27
SATYVIR SINGH AND MARCO BATTIATO
interesting to observe that both vorticity and degree of nonequilibrium play a significant role to
describe essential features in the study of the shock-accelerated bubble. Furthermore, the time
variations of the shock trajectories, and the interface scales, are investigated for the viewpoint of
quantitative analysis. The effects of nonequilibrium parameters, including bulk viscosity, which is
directly related to the rotational mode, and index of inverse power law are also investigated. The
numerical results demonstrated that both nonequilibrium parameters have a pivotal role in diatomic
and polyatomic gases away from equilibrium, and change substantially the flow fields of bubble
after interaction. A significant increase in enstrophy and the dissipation rate is found with increasing
the bulk viscosity ratio, whereas these physical quantities are observed in a diminishing trend with
increasing the index value.
The main goal of the present study is to investigate the nonequilibrium effects with different
physical properties of diatomic and polyatomic gases on the shock-accelerated heavy bubble with a
single-component gas model. Interestingly, the chemical reaction kinetics, including bulk viscosity
beyond the Stokes’ hypothesis, play an essential role in determining nonequilibrium effects of
diatomic and polyatomic gases in compressible multicomponent gas flows. It is also expected
that the nonequilibrium behavior will significantly affect the development of the RM instability
in compressible multicomponent gas flows. Therefore, in this context, the current work based
on the high-order discontinuous Galerkin method will be extended to investigate the impacts of
chemical reaction kinetics, including bulk viscosity, in a more controlled experiments on the flow
morphology in shock-accelerated polygonal bubbles with different types of interface shapes in
future investigations.
ACKNOWLEDGMENTS
The authors would like to acknowledge the financial support of the NAP-SUG grant program
funded by the Nanyang Technological University, Singapore. The authors thank the referees for
providing their very constructive comments and suggestions on the present paper.
[1] J. D. Lindl, R. L. McCrory, and E. M. Campbell, Progress toward ignition and burn propagation in inertial
confinement fusion, Phys. Today 45, 32 (1992).
[2] W. D. Arnett, J. N. Bahcall, R. P. Kirshner et al., Supernova 1987A, Annu. Rev. Astron. Astrophys. 27,
629 (1989).
[3] A. R. Jamaluddin, G. L. Ball, and T. J. Leighton, Shock/bubble interaction near a rigid boundary in shock
wave lithotripsy, in Shock Waves (Springer, Berlin, 2005).
[4] F. E. Marble, G. J. Hendricks, and E. E. Zukoski, Progress toward shock enhancement of supersonic
combustion processes, in Turbulent Reactive Flows (Springer, Berlin, 1989).
[5] Y. Yang, T. Kubota, and E. E. Zukoski, Applications of shock-induced mixing to supersonic combustion,
AIAA J. 31, 854 (1993).
[6] D. Arnett, The role of mixing in astrophysics, Astrophys. J. Suppl. 127, 213 (2000).
[7] R. D. Richtmyer, Taylor instability in shock acceleration of compressible fluids, Commun. Pure. Appl.
Math. 13, 297 (1960).
[8] E. E. Meshkov, Instability of the interface of two gases accelerated by a shock wave, Fluid Dyn. 4, 101
(1969).
[9] G. H. Markstein, A shock-tube study of flame front-pressure wave interaction, in 6th International
Symposium on Combustion (Elsevier, Amsterdam, 1957), Vol. 6, p. 387.
[10] G. Rudinger and L. M. Somers, Behavior of small regions of different gases carried in accelerated gas
flows, J. Fluid Mech. 7, 161 (1960).
[11] J. F. Haas and B. Sturtevant, Interaction of weak shock waves with cylindrical and spherical gas inhomo-
geneities, J. Fluid Mech. 181, 41 (1987).
[12] J. W. Jacobs, Shock-induced mixing of a light-gas cylinder, J. Fluid Mech. 234, 629 (1992).
044001-28
BEHAVIOR OF A SHOCK-ACCELERATED HEAVY …
[13] J. W. Jacobs, The dynamics of shock accelerated light and heavy gas cylinders, Phys. Fluids 5, 2239
(1993).
[14] G. Layes, G. Jourdan, and L. Houas, Experimental investigation of the shock wave interaction with a
spherical gas inhomogeneity, Phys. Fluids 17, 028103 (2005).
[15] G. Layes, G. Jourdan, and L. Houas, Experimental study on a plane shock wave accelerating a gas bubble,
Phys. Fluids 21, 074102 (2009).
[16] B. J. Balakumar, G. C. Orlicz, C. D. Tomkins, and K. P. Prestridge, Simultaneous particle-image
velocimetry-planar laser-induced fluorescence measurements of Richtmyer-Meshkov instability growth
in a gas curtain with and without reshock, Phys. Fluids 20, 124103 (2008).
[17] S. Balasubramanian, G. C. Orlicz, K. P. Prestridge, and B. J. Balakumar, Experimental study of initial
condition dependence on Richtmyer-Meshkov instability in the presence of reshock, Phys. Fluids 24,
034103 (2012).
[18] L. F. Henderson, P. Colella, and E. G. Puckett, On the refraction of shock waves at a slow-fast gas interface,
J. Fluid Mech. 224, 1 (1991).
[19] J. J. Quirk and S. Karni, On the dynamics of a shock-bubble interaction, J. Fluid Mech. 318, 129 (1996).
[20] A. Bagabir and D. Drikakis, Mach number effects on shock-bubble interaction, Shock Waves 11, 209
(2001).
[21] J. H. J. Niederhaus, J. A. Greenough, J. G. Oakley, D. Ranjan, M. H. Anderson, and R. A. Bonazza, A
computational parameter study for the three-dimensional shock-bubble interaction, J. Fluid Mech. 594,
85 (2008).
[22] B. Hejazialhosseini, D. Rossinelli, M. Bergdorf, and P. Koumoutsakos, High order finite volume methods
on wavelet-adapted grids with local time-stepping on multicore architectures for the simulation of shock-
bubble interactions, J. Comput. Phys. 229, 8364 (2010).
[23] A. Kundu and S. De, High resolution numerical simulation of a shock-accelerated refrigerant-22 bubble,
Comput. Fluids 193, 104289 (2019).
[24] Y. Zhu, Z. Yang, Z. Pan, P. Zhang, and J. Pan, Numerical investigation of shock-SF6 bubble interaction
with different mach numbers, Comput. Fluids 177, 78 (2018).
[25] Z. Wang, B. Yu, H. Chen, B. Zhang, and H. Liu, Scaling vortex breakdown mechanism based on viscous
effect in shock cylindrical bubble interaction, Phys. Fluids 30, 126103 (2018).
[26] S. Singh, Role of Atwood number on flow morphology of a planar shock-accelerated square bubble: A
numerical study, Phys. Fluids 32, 126112 (2020).
[27] S. Singh, A. Karchani, and R. S. Myong, Non-equilibrium effects of diatomic and polyatomic gases on the
shock-vortex interaction based on the second-order constitutive model of the Boltzmann-Curtiss equation,
Phys. Fluids 30, 016109 (2018).
[28] S. Singh, A. Karchani, K. Sharma, and R. S. Myong, Topology of the second-order constitutive model
based on the Boltzmann–Curtiss kinetic equation for diatomic and polyatomic gases, Phys. Fluids 32,
026104 (2020).
[29] G. Stokes, On the theories of the internal friction of fluids in motion, and of the equilibrium and motion
of elastic fluids, Trans. Cambridge Philos. Soc. 8, 287 (1845).
[30] G. Emanuel, Bulk viscosity of a dilute polyatomic gas, Phys. Fluids A 2, 2252 (1990).
[31] R. S. Myong, A generalized hydrodynamic computational model for rarefied and microscale diatomic gas
flow, J. Comput. Phys. 195, 655 (2004).
[32] N. Carlevaro and G. Montani, Bulk viscosity effects on the early universe stability, Mod. Phys. Lett. A
20, 1729 (2005).
[33] G. Billet, V. Giovangigli, and G. D. Gassowski, Impact of volume viscosity on a shock–hydrogen-bubble
interaction, Combust. Theory Model. 12, 221 (2008).
[34] Y. Zhu, C. Zhang, X. Chen, H. Yuan, J. Wu, S. Chen, C. Lee, and M. Gad-El-Hak, Transition in hypersonic
boundary layers: Role of dilatational waves, AIAA J. 54, 3039 (2016).
[35] F. Bahmani and M. S. Cramer, Suppression of shock-induced separation in fluids having large bulk
viscosities, J. Fluid Mech. 756, 1 (2014).
[36] S. Pan and E. Johnsen, The role of bulk viscosity on the decay of compressible, homogeneous, isotropic
turbulence, J. Fluid Mech. 833, 717 (2017).
044001-29
SATYVIR SINGH AND MARCO BATTIATO
044001-30