ArticlePDF Available

Abstract and Figures

Shock-tube experiments on eight kinds of two-dimensional multi-mode air–SF $_6$ interface with controllable initial conditions are performed to examine the dependence of perturbation growth on initial spectra. We deduce and demonstrate experimentally that the amplitude development of each mode is influenced by the mode-competition effect from quasi-linear stages. It is confirmed that the mode-competition effect is closely related to initial spectra, including the wavenumber, the phase and the initial amplitude of constituent modes. By considering both the mode-competition effect and the high-order harmonics effect, a nonlinear model is established based on initial spectra to predict the amplitude growth of each individual mode. The nonlinear model is validated by the present experiments and data in the literature by considering diverse initial spectra, shock intensities and density ratios. Moreover, the nonlinear model is successfully extended based on the superposition principle to predict the growths of the total perturbation width and the bubble/spike width from quasi-linear to nonlinear stages.
Content may be subject to copyright.
J. Fluid Mech. (2021), vol.928, A37, doi:10.1017/jfm.2021.849
Richtmyer–Meshkov instability on
two-dimensional multi-mode interfaces
Yu Liang1,2, Lili Liu1, Zhigang Zhai1,,JuchunDing
1,TingSi
1and
Xisheng Luo1,
1Advanced Propulsion Laboratory, Department of Modern Mechanics, University of Science and
Technology of China, Hefei 230026, PR China
2NYUAD Research Institute, New York University Abu Dhabi, Abu Dhabi 129188, UAE
(Received 14 February 2021; revised 1 August 2021; accepted 23 September 2021)
Shock-tube experiments on eight kinds of two-dimensional multi-mode air–SF6interface
with controllable initial conditions are performed to examine the dependence of
perturbation growth on initial spectra. We deduce and demonstrate experimentally that
the amplitude development of each mode is influenced by the mode-competition effect
from quasi-linear stages. It is confirmed that the mode-competition effect is closely
related to initial spectra, including the wavenumber, the phase and the initial amplitude of
constituent modes. By considering both the mode-competition effect and the high-order
harmonics effect, a nonlinear model is established based on initial spectra to predict
the amplitude growth of each individual mode. The nonlinear model is validated by the
present experiments and data in the literature by considering diverse initial spectra, shock
intensities and density ratios. Moreover, the nonlinear model is successfully extended
based on the superposition principle to predict the growths of the total perturbation width
and the bubble/spike width from quasi-linear to nonlinear stages.
Key words: shock waves, turbulent mixing
1. Introduction
Richtmyer–Meshkov (RM) instability is initiated when a shock wave interacts with an
interface between two fluids of different densities (Richtmyer 1960; Meshkov 1969), and
further induces mushroom-shaped flow structures such as bubbles (light fluids penetrating
into heavy ones) and spikes (heavy fluids penetrating into light ones), which finally may
cause a flow transition to turbulent mixing (Zhou, Robey & Buckingham 2003; Zhou 2007;
Zhou et al. 2019). Over the past few decades, the RM instability has become a subject of
intensive research due to its crucial role in various industrial and scientific fields such
Email addresses for correspondence: sanjing@ustc.edu.cn,xluo@ustc.edu.cn
© The Author(s), 2021. Published by Cambridge University Press 928 A37-1
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
as inertial confinement fusion (ICF) (Lindl et al. 2014) and supernova explosion (Kuranz
et al. 2018). For example, the RM instability determines the seeds of Rayleigh–Taylor
(RT) instability (Rayleigh 1883; Taylor 1950) that develops during the implosion in ICF
(Goncharov 1999). The mixing of hot fuel inside with cooler shell material outside,
induced by RM and RT instabilities in the target of ICF, significantly reduces and
even eliminates the thermonuclear yield (Miles et al. 2004). The RM instability on a
single-mode interface has been extensively studied due to its fundamental significance
(Brouillette 2002; Zhou 2017a,b;Zhaiet al. 2018). However, the initial perturbation in
reality is essentially multi-mode with wavenumbers spanning many orders of magnitude,
and whether the perturbation growth of a multi-mode RM instability depends on the initial
spectrum or not is crucial to ICF (Miles et al. 2004) but remains unclear.
Theoretically, there are mainly six kinds of models describing the perturbation growth of
a multi-mode interface: the linear model, the modal model, the potential model, the vortex
model, the perturbation expansion model and the group theory approach. Based on the
principle that each individual mode develops independently in linear stages, Mikaelian
(2005) proposed the linear model to describe the multi-mode interface evolution by
summing the time-varying amplitude growth of each mode. Haan (1989) found that
the constituent modes with similar wavelengths of a multi-mode interface add up to
create an effective local large amplitude, and, therefore, the onset of the nonlinear stage
of a multi-mode perturbation is earlier than that of the classical single-mode case.
Subsequently, Haan (1991) proposed the modal model with second-order accuracy to
quantify the mode-competition effect on the perturbation growth of each mode in the
early nonlinear stage. The modal model and its extended types have achieved a wide range
of validation in RT instability issues (Remington et al. 1995;Oferet al. 1996; Elbaz &
Shvarts 2018), but their application to the RM instability is still lacking. Assuming that
mode competition is absent before a bubble reaches its asymptotic growth, the potential
model was proposed by Alon et al. (1994) and Layzer (1955) to predict the eventual
average bubble distribution and the growth rate. However, the potential model is invalid
when the Atwood number (defined as A=2ρ1)/(ρ2+ρ1),withρ1and ρ2being the
densities of light fluid and heavy fluid, respectively) is low. When the Atwood number
approaches zero, the vortex model (Jacobs & Sheeley 1996) was adopted by Rikanati,
Alon & Shvarts (1998) to make up the bubble asymptotic growth rate. Note that both the
potential model (Alon et al. 1994,1995;Oronet al. 2001) and the vortex model (Rikanati
et al. 1998)involve a self-similar growth of the bubble front which is independent of the
initial spectrum, and both models obtain a 1/tdecay for the late-time bubble growth rate
in a multi-mode RM instability. The perturbation expansion model developed by Zhang &
Sohn (1997) was extended by Vandenboomgaerde, Gauthier & Mügler (2002) to predict
the early nonlinear amplitude growth of the constituent modes of a multi-mode interface
by retaining only the terms with the highest power in time. The group theory approach
(Abarzhi 2008,2010; Pandian, Stellingwerf & Abarzhi 2017) identifies the connection
between the symmetry properties of the interface morphology and the relative phases of
waves constituting the interface perturbation.
Experimentally, shock-tube experiments were performed to investigate two-bubble
competition (Sadot et al. 1998), and the results showed that the growth of the larger (or
smaller) bubble is promoted (or suppressed). Dimonte & Schneider (2000) conducted a
series of three-dimensional linear electric motor experiments to investigate multi-mode
RT and RM instabilities, and found that the density ratio has a limited effect on the
self-similar growth factor for the bubble. When the density ratio is large, the self-similar
growth factor for the spike is clearly larger than that for the bubble counterpart.
928 A37-2
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
The multi-mode RM instability of two liquids was investigated by Niederhaus & Jacobs
(2003), and the development of the multi-mode perturbation was found to be strongly
dependent on the relative amplitudes of initial modes. The growth of the multi-mode
interface perturbation created by the gas curtain technique shows a weak dependence
on the initial conditions (Balasubramanian, Orlicz & Prestridge 2013). Experiments of
a dual-mode interface RM instability under high-Mach-number conditions have been
performed (Di Stefano et al. 2015a,b), and the results indicated that new modes are
generated from the mode-competition effect, and the perturbations of these modes grow
and saturate over time. The dual-mode RM instability under weak shock conditions
was also considered, from which the mode-competition effect on the RM instability
development cannot be ignored when the wavenumber of one constituent mode is twice
that of the other constituent mode (Luo et al. 2020). The mixing of a multi-mode interface
was investigated by Mohaghar et al. (2017) using density and velocity statistics, and the
flow shows a distinct memory of initial conditions, the long-wavelength perturbation
having a strong influence on the interface development. Recently, developments of
quasi-single-mode interfaces created by the soap film technique in the early nonlinear
stage have been studied, and the effect of high-order modes on the perturbation growth
was highlighted to distinguish from single-mode perturbation (Liang et al. 2019). A
near-sinusoidal interface dominated by one mode was generated by a novel membraneless
technique where cross-flowing air was separated from SF6by an oscillating splitter plate
(Mansoor et al. 2020), and the effects of the initial amplitude on the perturbation width
growth and mixing transition have been discussed, and earlier mixing transitions for higher
amplitude-to-wavelength ratio cases are noted from experiments.
Numerically, it is commonly realized that the phases of the constituent modes influence
multi-mode perturbation growth (Vandenboomgaerde et al. 2002; Miles et al. 2004;
Pandian et al. 2017). Besides, the self-similar growth factor of the late-time RM
instability has a dependence on the scale of the initial spectrum. Specifically, a broadband
perturbation leads to a larger bubble growth factor than a narrowband counterpart
(Thornber et al. 2010; Liu & Xiao 2016; Thornber 2016; Groom & Thornber 2020).
Although significant progress on the multi-mode RM instability has been made, the
quantitative relation between initial conditions and perturbation growth is still unclear
mainly because a general nonlinear theory for predicting the multi-mode perturbation
width growth is absent, and elaborate experiments on the multi-mode RM instability
with controllable initial conditions are very limited. In our previous work, the extended
soap-film technique was utilized to create a classical two-dimensional (2-D) single-mode
perturbation (Liu et al. 2018), a 2-D multi-mode interface dominated by only one mode
(Liang et al. 2019)anda2-D multi-mode interface dominated by two modes (Luo
et al. 2020). The initial perturbations of these interfaces were precisely designed and
the initial conditions were well controlled. In this work, a 2-D complex multi-mode
interface constituted of various modes is first formed, and shock-tube experiments on
the developments of eight kinds of air–SF6multi-mode interface are performed. Then,
new nonlinear theories based on the initial spectrum, shock intensity and density ratio are
proposed to predict each mode amplitude growth and the total perturbation width growth
of a 2-D multi-mode interface.
2. Experimental method
The extended soap-film technique, which has been widely used in our previous work
(Ding et al. 2017;Liuet al. 2018; Liang et al. 2019), is adopted to generate a periodic
928 A37-3
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
x
y
o
Shock
140 mm
Air SF6
100 mm 10 mm
7 mm
Soap film
Rectangular
frame
140 mm
Shock
Filaments
Interface
w0
x
y
z
z= –3.5 mm
z=0 mm
z= 3.5 mm
(a)(b)
Figure 1. Schematics of soap-film interface generation (a) and the initial configuration (b).
multi-mode interface with a controllable initial shape to separate SF6from air. Such a
technique can largely eliminate the short-wavelength perturbations, diffusion layer and
three-dimensionality of the formed interface (Liu et al. 2018; Liang et al. 2019). As shown
in figure 1(a), two transparent devices with an inner height of 7.0 mm and a width of
140.0 mm are first made using acrylic plates (3.0 mm in thickness). A groove (0.7 mm
in thickness and 0.5 mm in width) with a multi-mode shape is then manufactured on the
internal side of each plate by a high-precision engraving machine. Then, two thin filaments
(1.0mm in thickness and 0.5 mm in width) with the same multi-mode shape are mounted
into the grooves of the upper and lower plates, respectively, to produce desired constraints.
Therefore, the bulging of the filament into the flow is less than 0.3 mm, and has a negligible
effect on the flow field. A small rectangular frame wetted by soap solution (78 % distilled
water, 2 % sodium oleate and 20 % glycerine by mass) is pulled along the filaments,
and a quasi-2-D soap-film interface is immediately generated, as shown in figure 1(a).
Subsequently, the auxiliary framework is gently inserted until it is completely connected
to the corresponding device. After that, the framework with a soap film on its surface is
slowly inserted into the test section of the shock tube. To form an air–SF6interface, the
air on the right-hand side of the interface is replaced by SF6. To minimize the effect of
the shock-tube walls on interface evolution, a short flat part with 10mm on each side of
the perturbed interface is adopted, as sketched in figure 1(b), and its effect on the interface
evolution is limited (Luo et al. 2019).
In the Cartesian coordinate system, as sketched in figure 1(b), the multi-mode interface
investigated can be described as a sum of three cosine modes:
y=
3
1
a0
kncos(knx+φkn), x[60,60] mm,(2.1)
where a0
kn,knand φknrespectively denote the initial amplitude, wavenumber and phase
of the nth constituent mode with n=1, 2 and 3. To illustrate the influences of the initial
amplitude, wavenumber and relative phase on the 2-D multi-mode RM instability, eight
different kinds of multi-mode interface are designed in this work. The initial spectrum and
the initial perturbation width (w0, sketched in figure 1b) of the multi-mode interface in
928 A37-4
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
Case a0
k1a0
k2a0
k3k1k2k3φk1φk2φk3a0
k1k1a0
k2k2a0
k3k3w0w0bw0s
IP-s 1.0 1.0 1.0 105 209 314 0 0 0 0.1 0.2 0.3 4.3 3.0 1.3
k3AP-s 1.0 1.0 1.0 105 209 314 0 0 π0.1 0.2 0.3 3.5 1.5 2.0
k2AP-s 1.0 1.0 1.0 105 209 314 0 π0 0.1 0.2 0.3 4.3 1.3 3.0
AP-s 1.0 1.0 1.0 105 209 314 0 ππ 0.1 0.2 0.3 3.5 2.0 1.5
IP-h 3.0 2.0 1.0 105 209 314 0 0 0 0.3 0.4 0.3 8.0 6.0 2.0
k3AP-h 3.0 2.0 1.0 105 209 314 0 0 π0.3 0.4 0.3 7.5 4.0 3.5
k2AP-h 3.0 2.0 1.0 105 209 314 0 π0 0.3 0.4 0.3 8.0 2.0 6.0
AP-h 3.0 2.0 1.0 105 209 314 0 ππ 0.3 0.4 0.3 7.5 3.5 4.0
Table 1. Initial spectrum and initial perturbation width of multi-mode interfaces formed in the present work.
Here a0
kn,knand φkndenote the initial amplitude, wavenumber and phase of the nth constituent mode,
respectively; w0,w0band w0sdenote the initial total perturbation width, bubble width and spike width of
the multi-mode interface, respectively. The unit for the amplitude and width is mm and for the wavenumber is
m1.
different cases are listed in table 1. In this work, φk1is kept as 0 and φk2and φk3are varied.
For convenience, the following notation is adopted: φk2=φk3=0 (in-phase (IP) case);
φk2=φk3=π(anti-phase (AP) case); φk2=0, φk3=π(k3AP case); φk2=π,φk3=0
(k2AP case). The influence of the initial amplitude of mode knon the RM instability can
be studied by varying a0
knkn(Mansoor et al. 2020;Sewellet al. 2021). In this work, the
cases of IP-s, k3AP-s, k2AP-s and AP-s are classified as small-w0cases, whereas the cases
of IP-h, k3AP-h, k2AP-h and AP-h are classified as large-w0cases.
The experiments are performed in a horizontal shock tube with a 140 mm ×13 mm
cross-sectional area. This type of tube has been widely used in shock–interface interaction
studies (Luo, Wang & Si 2013;Zhaiet al. 2014;Luoet al. 2015; Ding et al. 2017).
The ambient pressure and temperature are 101.3kPa and 299.5±1.0 K, respectively. In
experiments, the incident shock wave with velocity (vs)of409±1ms
1(the incident
shock Mach number (M)is1.18)movesfromairtoSF
6. The ambient air is considered
as pure and the test gas is a mixture of air and SF6,the mass fraction of SF6being
0.97 ±0.01 calculated according to one-dimensional gas dynamics theory. Meanwhile, the
transmitted shock velocity (vt) and the speed jump of the interface (Δv) can be calculated
as 182 ±1ms
1and 65.5±0.5ms
1, respectively. The post-shock Atwood number A+
(defined as A+=+
2ρ+
1)/(ρ+
2+ρ+
1),withρ+
2and ρ+
1being the densities of shocked
test gas and air, respectively) is 0.66 ±0.01. The flow field is monitored using high-speed
schlieren photography. The frame rate of the high-speed video camera (FASTCAM SA5,
Photron Limited) is 62 500 f.p.s. with a shutter time of 1 μs. The spatial resolution of the
schlieren images is 0.4mm pixel1. The visualization window of the flow field is within
the range x[50,50] mm, as shown with the grey zone in figure 1(b). For each case, at
least three experimental runs are performed, and the experiments have a good repeatability.
The relative differences of the data among diverse experimental runs are within 3 %.
The three-dimensional feature of the initial soap-film interface is discussed. Because the
gases on both sides of the interface are at ambient pressure, the soap-film interface formed
has a zero mean curvature, and the geometry can be characterized as minimum surface
(Luo et al. 2013; Liang et al. 2021). For instance, half of the perturbation width of the
soap-film interface in case IP-h or k3AP-h is 4 mm (see table 1), indicating that the total
amplitude of the soap-film interface on the boundary slice (i.e. z3.5 mm) is 4 mm.
928 A37-5
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
The maximum wavelength of the soap-film interface is 60 mm, and the interface height is
7 mm. Based on our previous work (Luo et al. 2013; Liang et al. 2021), the amplitude of
the soap film on the symmetry slice (i.e. z=0 mm) is 3.75 mm. Therefore, the amplitude
ratio of the symmetry slice over the boundary slice is 93.75 %. The absolute difference
between the amplitudes on the symmetry slice and the boundary slice is about 0.25 mm,
and is smaller than the size of a schlieren image’s pixel. As a result, it is believed that the
interface height of 7 mm can reduce the three-dimensional effect in this work.
The boundary layer effect on the interface evolution is also considered. After the
incident shock with Mof 1.18 impacts the interface, the flow behind the transmitted shock
can be regarded as laminar and incompressible. As a result, the displacement thickness of
the boundary layer (δ) can be approximately calculated using the following expression:
δ=1.72μymax
ρΔv,(2.2)
where ymax (100mm measured from experiment) is the maximum distance that the
interface moves when image recording ends. In this study, μ=1.83 ×105Pa s (=
1.60 ×105Pa s) is the viscosity coefficient of the ambient (test) gas, ρ=1.2kgm
3
(=5.3kgm
3) is the density of the ambient (test) gas and Δv65.5ms
1. According
to (2.2), the displacement thickness of the boundary layer is calculated to be about 0.26 mm
for ambient gas and 0.12mm for test gas, which is much smaller than the inner height of
the acrylic plates (7.0 mm). Therefore, the effect of the boundary layer on the interface
evolution is negligible.
After a shock wave impacts the soap film, the soap solution is atomized into tiny droplets
(Cohen 1991; Hosseini & Takayama 2005; Ranjan et al. 2005). Our previous work (Luo
et al. 2013;Siet al. 2015;Leiet al. 2017) revealed that the dimension of the atomized
droplets is within 1–10 μm, and a portion of tiny soap droplets follows the evolving
interface nicely and can be utilized for light scattering illuminated by a laser. Besides,
it is recommended to mix atomized olive oil droplets with a diameter of around 1 μmwith
SF6when injecting the test gas into the test section of a shock tube. Using the atomized
soap droplets and oil droplets as tracer particles, it is worth looking forward to adopting
aparticle image velocimetry system to capture the velocity and vorticity contours of an
evolving interface initially generated with soap-film technology.
3. Results and discussion
3.1. Experimental observation and quantitative results
The schlieren images of the shocked multi-mode interface for small-w0cases are shown
in figure 2. It is evident that the phases of the constituent modes greatly affect the initial
interface shape and the later interface evolution. Taking the IP-s case as an example, after
the transmitted shock just leaves the interface, the shocked interface retains its initial
shape (69 μs). Then, the perturbation on the interface grows gradually but the interface
remains single-valued, which indicates that the interface evolves in early nonlinear
stages. Subsequently, vortices appear on the spikes, and the interface morphology acts
as multi-valued (261 μs). Due to the bubble-merging process (Sadot et al. 1998), the large
spike between the large bubble and small bubble skews towards the large bubble, whereas
the small spike between two large spikes remains symmetric (581 μs), which qualitatively
agrees with the group theory analysis (Abarzhi 2008,2010; Pandian et al. 2017). Finally,
the scales of vortices are comparable to the total perturbation width, and the interfacial
morphology shows a strong nonlinearity (1061μs).
928 A37-6
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
69
68
66
261
260
258
581
580
578
574
IP-s
k
3
AP-s
k
2
AP-s
AP-s
Interface
TS
Vortices LS
SS
LB
SB
w
UP DP
x
y
1061
1060
1058
105462 254
(a)
(b)
(c)
(d)
Figure 2. Schlieren images of multi-mode interface evolution for small w0cases. TS, transmitted shock; w,
interface perturbation width; LS, large spike; LB, large bubble;SS, small spike; SB, small bubble; UP, upstream
point of interface; DP, downstream point of interface. Numbers denote time in μs, and similarly hereinafter.
The Schlieren images of the shocked multi-mode interface for large-w0cases are shown
in figure 3. The interface morphologies in large-w0cases are qualitatively similar to
the corresponding small-w0ones. However, for the large-w0cases, there is a greater
misalignment between the pressure gradient of the shock wave and the density gradient of
the interface, resulting in more baroclinic vorticity production and the earlier appearance
of vortices. Besides, at late time in the k2AP case (1062 μs), the spike structures
on the multi-valued interface break, and the whole interface becomes chaotic, which
indicates that the transition may occur earlier when the initial interface amplitude is larger
(Mohaghar et al. 2019; Mansoor et al. 2020).
The captured interface morphology is distinct such that the interface contours in all
cases can be extracted by an image processing program, as indicated by the insets in
figures 4 and 5. The mean ycoordinate in each image is taken as the average position of
the local interface. Spectrum analysis is then performed on the averaged interface contour
before the interface becomes multi-valued, and the amplitudes of three constituent modes
are acquired, as shown in figures 4 and 5. Time is normalized as τn=k1|vR
kn|t,wherevR
kn
is the Richtmyer growth rate of mode kncalculated by the impulsive theory (Richtmyer
1960):
vR
kn=ZckA+Δva0
kncoskn), (3.1)
928 A37-7
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
261
259
262
IP-h
k
3
AP-h
k
2
AP-h
AP-h
581
579
582
580
1061
1059
1062
1060
69
67
70
69 260
x
y
(a)
(b)
(c)
(d)
Figure 3. Schlieren images of the multi-mode interface evolution for large-w0cases.
in which Zc(=1Δv/vs) is the shock compression factor and is equalto0.84inall
cases. In the present coordinate system, vR
knis positive if φkn= 0, but becomes negative
if φkn=π. The amplitude is scaled as ηn=k1|akn(t)Zca0
kncoskn)|,withakn(t)the
time-varying amplitude of mode kn.
In small-w0cases, as shown in figure 4, it is clear that the dimensionless amplitudes
of modes k1and k2are larger than those of mode k3in IP-s and AP-s cases, but smaller
than those of mode k3in k3AP and k2AP cases. In other words, the low-order modes
(modes k1and k2) in IP-s and AP-s cases dominate the flow, whereas the high-order
mode (mode k3)ink3AP and k2AP cases dominates the flow, which agree with the
observations in figure 2. For example, in the IP-s case in figure 2(a), the late-time interface
is dominated by long-wavelength structures (a large bubble and two groups of large and
small spikes), but in the k3AP case in figure 2(b), the late-time interface is dominated
by four short-wavelength structures (four pairs of spikes and bubbles). Therefore, the
mode-competition effect plays a role in the amplitude development of the constituent
modes in early stages. Since the small perturbation hypothesis is satisfied for each
constituent mode, the impulsive theory should be valid to predict the amplitude growth
of each constituent mode if the mode-competition effect is ignored. However, compared
with the predictions of the impulsive theory, in the IP-s case, mode k1development is
promoted but mode k3development is suppressed, while mode k2development is not
obviously influenced by the mode-competition effect. Differently, in the k3AP-s case,
the developments of both modes k1and k2are suppressed, while the development of
mode k3is not obviously influenced by the mode-competition effect. Therefore, the
928 A37-8
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
0.1 0.2 0.3 0.40
0.1
0.2
0.3
0.4
69 261213165117
k1
Run2
mode k3
Run1
k3
k2
mode k1mode k2
Present
Haan-RM
Ofer-RM Impulsive
theory
IP-s
0.1 0.2 0.3 0.4
0
0.1
0.2
0.3
0.4
68 260212164116
k1
k3
k2
k3AP-s
0.1 0.2 0.3 0.40
0.1
0.2
0.3
0.4
66 258210162114
k1
k3
k2
k2AP-s
0.1 0.2 0.3 0.40
0.1
0.2
0.3
0.4
62 254206158110
k1
k3
k2
AP-s
ηn
ηn
τnτn
(a)(b)
(c)(d)
Figure 4. The dimensionless amplitudes of three constituent modes obtained from small-w0cases. The insets
show the interface contours for the spectrum analysis in which numbers indicate time in μs. Run1 and run2
represent typical experimental runs. The black dashed line represents the impulsive theory (Richtmyer 1960).
The coloured solid lines, dashed lines and dash-dotted lines represent the amplitudes of three constituent modes
calculated by the Haan-RM model (3.4), the Ofer-RM model (3.7)and the present model (3.10), respectively,
and similarly hereinafter.
mode-competition effect is greatly influenced by the initial wavenumber and the initial
phase of the constituent mode.
In large-w0cases, as shown in figure 5, similar to the corresponding small-w0cases,
the amplitudes of both modes k1and k2are larger than those of mode k3in IP-h and
AP-h cases, but smaller than those of mode k3in k3AP-h and k2AP-h cases. However,
compared with the impulsive theory, mode k2development in the IP-h case is suppressed
whereas mode k3development in the k3AP-h case is promoted by the mode-competition
effect, which is different from the results of corresponding small-w0cases. Therefore,
the initial amplitude of the constituent mode also affects the mode-competition effect. In
summary, the effect of mode competition is closely related to the initial spectra, including
the wavenumber, the phase and the initial amplitude of the constituent modes.
Generally, the linear stage is defined by the fact that Fourier modes evolve separately
(Drazin & Reid 2004; Chandrasekhar 2013). Because it is difficult to obtain the starting
point of the linear stage in reality, usually used within the framework of RM instability
is that the linear stage occurs as long as mode ksatisfies ak(t)k. This constant α
928 A37-9
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
0.1 0.2 0.3 0.4
0
0.1
0.2
ηn
ηn
τnτn
0.3
0.4
52 18014811684
k1
k3
k2
Run2
mode k3
Run1
mode k1mode k2
Present
Haan-RM
Ofer-RM Impulsive
theory
IP-h
0.1 0.2 0.3 0.40
0.1
0.2
0.3
0.4
51 17914711583
k1
k3
k2
k3AP-h
0.1 0.2 0.3 0.40
0.1
0.2
0.3
0.4
54 18215011886
k1
k3
k2
k2AP-h
0.1 0.2 0.3 0.40
0.1
0.2
0.3
0.4
0.5
36 22818013284
k1
k3
k2
AP-h
(a)(b)
(c)(d)
Figure 5. The dimensionless amplitudes of three modes obtained from large-w0cases.
depends essentially on the accuracy required. As a result, it is generally accepted that the
necessary condition of the linear RM instability is a0
kk1 (Mügler & Gauthier 1998;
Collins & Jacobs 2002; Mikaelian 2003; Niederhaus & Jacobs 2003; Mariani et al. 2008;
Vandenboomgaerde et al. 2014; Mansoor et al. 2020). However, the linear stage still exists
for a very large initial a0
kkat the expense of a large reduction in the duration of the linear
stage (Rikanati et al. 2003; Dell, Stellingwerf & Abarzhi 2015;Zhaiet al. 2016;Dellet al.
2017). It should be noted that there are harmonics growing with time in these large initial
αcases but they are (almost) negligible. Therefore, this linear stage within large initial α
is actually a quasi-linear stage. As listed in table 2,thecriterion of dimensionless time,
i.e. τ
n(=kn|vR
kn|t), of mode knbetween the quasi-linear stage and the nonlinear stage is
evaluated from experiments. It is found that τ
nof the three initial modes are less than
the generally accepted criterion of dimensionless time of 0.7 for the single-mode RM
instability (Niederhaus & Jacobs 2003), indicating that the large initial a0
knknresults in a
very short quasi-linear stage and the mode competition advances the nonlinearity of the
multi-mode RM instability.
3.2. Linear and nonlinear theories
To quantitatively describe the 2-D multi-mode RM instability development, linear and
nonlinear theories based on the impulsive theory, the modal model and the interpolation
model have been established.
928 A37-10
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
Case IP-s k3AP-s k2AP-s AP-s IP-h k3AP-h k2AP-h AP-h
τ
10.06 0.07 0.06 0.08 0.13 0.18 0.17 0.22
τ
20.64 0.32 0.32 0.34 0.36 0.40 0.36 0.52
τ
30.66 0.72 0.54 0.72 0.30 0.48 0.42 0.66
Table 2. The criterion of dimensionless time (τ
n) of mode knbetween the quasi-linear stage and the
nonlinear stage.
I. Impulsive theory. If each mode of a multi-mode interface satisfies |ak(t)k|1, the
whole interface evolves linearly. Previous studies (Sadot et al. 1998; Mikaelian 2005;Di
Stefano et al. 2015a,b; Liang et al. 2019)considered that the mode-competition effect is
negligible in quasi-linear stages. Therefore, for an initial light–heavy interface, the linear
amplitude growth rate (vl
k) of mode kcan be described by the impulsive theory (3.1), i.e.
vl
k=vR
k. For an initial heavy–light perturbed interface, the Richtmyer growth rate should
be modified as vR
k=(Zc+1)kA+Δva0
kcosk)/2 (Meyer & Blewett 1972).
When a0
kis comparable to its wavelength or/and the shock intensity is large, the
high-amplitude effect or/and the high-Mach-number effect will inhibit vR
k(Rikanati et al.
2003;Dellet al. 2015,2017;Guoet al. 2020). Here, the high-amplitude effect and the
high-Mach-number effect are considered independently.Then the modified Richtmyer
growth rate (vMR
k)is
vMR
k=βΔvtanh(RkvR
kΔv), (3.2)
where Rk(=1/[1 +(ka0
k/3)(4/3)]) is the reduction factor proposed by Dimonte &
Ramaprabhu (2010) to quantify the high-amplitude effect on mode k.Forsmall-w0cases,
Rk1=0.97, Rk2=0.93 and Rk3=0.91; for large-w0cases, Rk1=0.91, Rk2=0.88 and
Rk3=0.91. Parameter β(=1Δv/vt) is the reduction factor proposed by Hurricane
et al. (2000) to quantify the high-Mach-number effect on all modes, and β=0.64 in all
cases.
II. Modal model. When the mode competition starts to play a role, the interface evolution
enters the early nonlinear stage. Haan (1991) first proposed a modal model applicable to
the 2-D RT instability, and then deduced the modal model for the 2-D RM instability with
zero acceleration (g=0) and assuming |ak(t)||a0
k|, i.e. ignoring the initial amplitude
terms, as
ak(t)=al
k(t)+1
2Ak
k
al
k(t)al
k (t)×1
2ˆ
k·
k1
2
k·
k,(3.3)
where k =kk;k,kand k R2;and ˆ
kis the unit vector k/k.Amplitude al
k(t)is the
linear amplitude of mode k. Taking the first derivative of (3.3) and using the post-shock
physical parameters, the ‘Haan-RM’ model can be obtained to calculate the early nonlinear
amplitude growth rate ven
k(t)of mode kas
ven
k(t)=vMR
k+A+k2
k
vMR
kvMR
k+k
k<k
vMR
kvMR
kkt,(3.4)
where k,kand k R+. The first term on the right-hand of the Haan-RM model indicates
that the development of each mode of a multi-mode RM unstable interface is still strongly
928 A37-11
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
influenced by its independent perturbation growth. At t0, the Haan-RM model reduces
to ven
k(t)vl
k=vMR
k, which indicates that the mode-competition effect is very weak and
can be ignored. The second term on the right-hand side is the mode-competition term, and
it is evident that as tincreases, the mode-competition effect plays a more important role in
influencing the RM instability. The first sum term in the mode-competition term represents
the generation of mode kfrom high-order modes and indicates the bubble-merging
process. The second sum term in the mode-competition term represents the generation
of mode kfrom the interaction of low-order modes, which is related to the bubble-spike
asymmetry and the total mixing rate decrement.
However, the initial amplitude terms cannot be ignored in deducing the modal model for
the whole process of the RM instability, especially when the initial amplitude of mode kis
large. Now, the modal model for the RM instability including the initial amplitude terms
is re-derived. Based on the modal model solved by Ofer et al. (1996)to second-order
accuracy with g=constant for the RT instability,
ak(t)=al
k(t)+1
2Ak
k
al
k(t)al
k+k(t)1
2
k<k
al
k(t)al
kk(t),(3.5)
we take the second derivative of (3.5) with time, and get
d2ak(t)
dt2=Agka0
k+1
2A2gk
k
[ka0
kal
k+k(t)+2k(k+k)a0
ka0
k+k
+(k+k)a0
k+kal
k(t)]1
2
k<k
[ka0
kal
kk(t)
+2k(kk)a0
ka0
k+k+(kk)a0
kkal
k(t)].(3.6)
Similar to Richtmyer (1960), the constant gis replaced by an impulsive acceleration δtΔv
(δt=0whent=0andδt=1whent>0) and the post-shock physical parameters are
adopted. Through integrating (3.6) with time, ven
kfor the RM instability can be expressed
as a superposition of the linear amplitude growth rate vl
kand the weakly nonlinear
modification vwn
k(t):
ven
k(t)=vl
k+vwn
k(t), (3.7)
with
vl
k=vMR
k+1
2A+k
kvMR
kZca0
k+k+vMR
k+kZca0
k1+2k
k+1
1
2
k<kvMR
kZca0
kk+vMR
kkZca0
k1+2k
k1 (3.8)
and
vwn
k(t)=A+k
k
vMR
kvMR
k+k1
2
k<k
vMR
kvMR
kkt.(3.9)
Here, (3.7)–(3.9)arecalled the ‘Ofer-RM’ model. Different from vl
k=vMR
kindicated
by the Haan-RM model, vl
kin the Ofer-RM model is a superposition of vMR
kwith the
928 A37-12
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
Case vMR
k1vl
k1vMR
k2vl
k2vMR
k3vl
k3vMR
k4vl
k4vMR
k5vl
k5vMR
k6vl
k6
IP-s 3.75.27.17.910.18.803.102.403.4
k3AP-s 3.73.37.25.110.311.300.10 2.603.7
k2AP-s 3.72.17.26.110.311.403.30 2.603.7
AP-s 3.74.17.19.210.29.500.102.603.6
IP-h 10.115.812.914.210.15.2010.704.903.6
k3AP-h 10.412.213.25.210.415.901.80 5.103.8
k2AP-h 10.14.912.912.510.115.5010.70 4.903.6
AP-h 10.38.813.421.510.35.201.704.903.6
Table 3. Comparison of the modified Richtmyer growth rate (vMR
kn) calculated by (3.2) with the linear
amplitude growth rate (vl
kn) calculated by (3.7). The unit for the growth rate is m s1.
mode-competition term that is related to the initial amplitudes of constituent modes. In
other words, the mode-competition effect influences each mode perturbation growth in the
quasi-linear stage of the multi-mode RM instability, which is different from the previous
view that the mode-competition effect can be ignored in the quasi-linear stage (Sadot
et al. 1998; Mikaelian 2005; Di Stefano et al. 2015a,b; Liang et al. 2019). Besides, when
ven
k(t)=0, mode kis fully saturated. Here, additional rules introduced by Ofer et al. (1996)
for the enforced post-saturation treatment in calculating ven
k(t)are adopted: (i) no weakly
nonlinear modifications of a saturated mode to low-order modes and (ii) the phases of the
harmonics generated by the saturated modes are opposite.
The predictions from the Haan-RM model and the Ofer-RM model are calculated as
shown in figures 4 and 5.Forsmall-w0cases, because the initial amplitudes of three
modes are small, both models give reasonable predictions of the experimental results.
Differently, for large-w0cases, the Ofer-RM model provides a better prediction of the
amplitude growth than the Haan-RM model for some constituent modes, such as modes
k1and k3in the IP-h case, mode k1in the k2AP-h case and modes k1and k2in the AP-h
case. Note that the amplitude developments of mode k1in the IP-h case and mode k3in the
AP-h case deviate from predictions of the impulsive theory from the very beginning, which
verifies that the mode-competition effect plays an important role in the early evolution of
a multi-mode interface. In summary, the Ofer-RM model is more applicable to describing
the multi-mode RM instability behaviour in the early nonlinear stage, especially when the
initial amplitudes of constituent modes are large. Meanwhile, the present experimental
results prove that the linear growth of each mode amplitude is also influenced by the
mode-competition effect.
Basedon(3.8), vl
kconsidering the mode-competition effect is calculated and compared
with vMR
kwithout the mode-competition effect, as listed in table 3 for all cases. The
difference between vl
kand vMR
kis larger in large-w0cases than in small-w0cases.
According to the modal analysis (Haan 1991;Oferet al. 1996; Miles et al. 2004), the
coupling of constituent modes will generate new harmonics. In our experiments, k2and
k3are integral multiples of k1, and thus no modes with wavenumber lower than k1will
be generated (Miles et al. 2004). However, three new harmonics with higher order, i.e.
harmonics k4(=k1+k3), k5(=k2+k3)andk6(=k3+k3), are generated if only the first
generation of new harmonics is considered (Ofer et al. 1996). The values of vl
kfor the three
generated harmonics are listed in table 3. It is evident that the new generated harmonics
cannot be ignored in the multi-mode RM instability development in early stages.
928 A37-13
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
III. Interpolation model. Although the Ofer-RM model generally gives a better
prediction than the Haan-RM model in large-w0cases, it still underestimates mode k2
development in the k3AP-s case and overestimates mode k2development in the AP-h case
when tis large. According to the Ofer-RM model, when t→∞,vl
kis neglected and ven
kis
proportional to tas long as vwn
kis non-zero. Actually, in the consideration of classical
single-mode RM instability, the late nonlinear amplitude growth rate (vln
k) of mode k
should be t1decay (Hecht, Alon & Shvarts 1994; Alon et al. 1995; Mikaelian 1998)
because of the suppression of high-order (three orders or greater) harmonics generated by
mode kitself (Velikovich & Dimonte 1996; Zhang & Sohn 1997; Nishihara et al. 2010;
Velikovich, Herrmann & Abarzhi 2014). In the multi-mode RM instability counterparts,
the perturbation width growth in the fully turbulent stage is proportional to tθas a result
of bubble merging (Alon et al. 1994,1995). Although the value of θhas not been
unified, it should be much lower than 1.0, as reviewed by Zhou (2017a,b). Therefore,
the perturbation width growth rate of a multi-mode interface should be proportional to
tθ1with 1<(θ1)<0 in the fully turbulent stage. As a result, the Ofer-RM model
with second-order accuracy is not applicable to describing the late-time 2-D multi-mode
RM instability, and its scope should be extended by considering the suppression from
high-order harmonics.
An interpolation model proposed by Dimonte & Ramaprabhu (2010) (DR model) for
predicting 2-D single-mode amplitude growth covers the entire time domain from the
early to late nonlinear stages of the RM instability, and it has been well verified by several
independent experiments (Dimonte et al. 1996;Sadotet al. 1998; Niederhaus & Jacobs
2003; Jacobs & Krivets 2005) through considering diverse amplitude-to-wavelength ratios,
shock intensities and density ratios. In this work, the scope of the Ofer-RM model is
extended in combination with the DR model, and then vln
kcan be expressed as an average
of the bubble amplitude growth rate vln
kb(t)with the spike amplitude growth rate vln
ks(t)of
mode k:
vln
k(t)=1
2[vln
kb(t)+vln
ks(t)],(3.10)
with
vln
kb/ks(t)=ven
k(t)1+(1∓|A+|)|kven
k(t)t|
1+Cb/s|kven
k(t)t|+(1∓|A+|)Fb/s|kven
k(t)t|2,
Cb/s=4.5±|A+|+(2∓|A+|)(kak
0)
4,Fb/s=1±|A+|.
(3.11)
Finally, (3.2), (3.7)and(3.10) constitute the ‘present’ model for describing the 2-D
multi-mode RM instability in this work. Notably, the self-similar law shows that the
perturbation width growth rate of a multi-mode interface in the fully turbulent stage has
atθ1dependence and therefore approaches zero when t→∞. According to (3.10)and
(3.11), when t→∞,vln
k(t)also approaches zero but shows a t1dependence, i.e. θ=0,
which violates the self-similar law. However, the t1asymptotic dependence agrees with
the potential model (Alon et al. 1994,1995;Oronet al. 2001), the vortex model (Rikanati
et al. 1998)and recent experimental findings (Mansoor et al. 2020).
The predictions of the present model for our experiments are shown in figures 4
and 5, and a good agreement between them is achieved. Besides, data from literature
are extracted to validate the present model. First, the numerical results from figures 9
and 10 of Vandenboomgaerde et al. (2002)withbothM(=1.0962) and A+(=0.764)
928 A37-14
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
t (μs)
an (mm)
an (μm)
0 100 200 300 400
–6
–4
–2
0
2
4
6
k1
k3
k3
k2
k1
k2
Figure 9
Vandenboomgaerde et al. (2002)
Figure 10
t (ns)
20 22 24 26 28 30
0
5
10
15
20
k3
k2
k1
Presentexp num Di Stefano et al. (2015)
Figure 5
(a)(b)
Figure 6. The amplitudes of modes obtained from (a) the numerical results in Vandenboomgaerde et al.
(2002)and(b) the experimental and numerical results in Di Stefano et al. (2015b). Coloured lines represent the
predictions calculated with the present model.
similar to those in our experiments are extracted, as shown in figure 6(a). The initial
interface consists of three modes: a0
k1=0.35 ×103,a0
k2=1.9055 ×103and a0
k3=
1.072 ×103m; k1=274.855 m1,k2=3k1/7andk3=4k1/7. For the data in figure 9
of Vandenboomgaerde et al. (2002), φ1=φ2=φ3=0, while for the data in figure 10,
φ1=φ2=φ3=π. For these two cases, Rk1=0.97, Rk2=0.93 and Rk3=0.95; β=
0.80. It is found that the predictions of the present model agree well with all the numerical
results (Vandenboomgaerde et al. 2002). Further, the experimental and numerical results
(Di Stefano et al. 2015b) with a much larger M(=8) and a smaller A+(=1/3) compared
with our experimental conditions are predicted by the present model. The initial interface
in the literature (Di Stefano et al. 2015b) consists of two modes: a0
k1=5×106m,
a0
k2=0.5a0
k1;k1=6.28 ×104m1,k2=2k1;φ1=0, φ2=π;Rk1=0.91, Rk2=0.91;
β=0.35. The amplitude growths of the two constituent modes and the harmonic k3
generated by the coupling of modes k1and k2are extracted from figure 5 of Di Stefano
et al. (2015b), as shown in figure 6(b). It is found that not only the constituent mode
amplitudes, but also the generated new harmonic amplitude are well predicted by the
present model. Overall, the present model established in this work by considering both the
mode-competition effect and the high-order harmonics effect is applicable to the nonlinear
RM instability before the transition.
3.3. The perturbation width growth
The memory of the perturbation width growth of a multi-mode interface on its initial
spectrum is crucial to RM instability research. The perturbation width w(t)of a
multi-mode interface is defined as the streamwise distance from the upstream point
(UP) to the downstream point (DP) of an interface, as shown in figure 2(b). The
variations of the perturbation width measured from the schlieren images for small-and
large-w0cases are shown in figures 7(a)and8(a), respectively. The time is normalized
as τw=0.5k1vMR
wt,wherevMR
w=3
1vMR
kn[cos(knxDP)cos(knxUP )],withxUP and xDP
being the xcoordinates of UP and DP, respectively. The half of the perturbation width
is scaled as ηw=0.5k1[w(t)Zcw0]. To compare the perturbation width between the
928 A37-15
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
0.5 1.0 1.5 2.00
0.5
1.0
1.5
Impulsive
theory
DR model
0.5 1.0 1.5 2.0
0
0.5
1.0
1.5
k3AP-s
k2AP-s
IP-s
exp Sum
Bubble Bubble
Spike
Spike
AP-s
0.5 1.0 1.5 2.0
0
1.5
2.0
2.5
3.0
0.5
1.0
0.5
0
1.5
1.0
0.5 1.0 1.5 2.0
0
1.5
2.0
0.5
1.0
0.5
0
1.5
1.0
ηwζw
ηb/sζb/s
τwτw
(a)(b)
(c)(d)
Figure 7. Comparison of the perturbation width (a), growth rate of the perturbation width (b), widths of
bubble and spike (c)and width growth rates of bubble and spike (d)insmall-w0cases. Black dashed lines
and dash-dotted lines represent the single-mode linear and nonlinear amplitude growth rates calculated by the
impulsive theory (Richtmyer 1960) and DR model (Dimonte & Ramaprabhu 2010), respectively. Coloured solid
lines represent the predictions from the sum model calculated with (3.12), and similarly hereinafter.
multi-mode interface and the classical single-mode interface, the predictions from the
impulsive theory and the DR model are calculated, as shown in figures 7(a)and8(a),
respectively. It is evident that the perturbation width growth of a multi-mode interface in
the early stage (τw0.5) is close to or even larger than the single-mode linear growth,
which indicates that the mode-competition effect may enhance the initial growth of the
multi-mode perturbation. Later, the perturbation width growth in all cases is smaller than
the single-mode linear growth when τw>0.7 and smaller than the single-mode nonlinear
growth when τw>1.2. Therefore, the mode-competition effect suppresses the multi-mode
interface development at late stages because it enhances local mixing and reduces the
global mixing. Besides, the w(t)growth curves in all cases deviate from each other, which
indicates that the initial spectrum influences the 2-D multi-mode RM instability until the
late nonlinear stage.
To more clearly distinguish the differences of the perturbation growth with diverse initial
conditions, through the direct differentiation of the experimental data, the perturbation
widthgrowthrate ˙w(t)of a multi-mode interface is acquired, and compared with the
single-mode linear and nonlinear amplitude growth rates, as shown in figures 7(b)and8(b)
for all cases. The perturbation width growth rate is scaled as ζwws(t)/vMR
w. Therefore,
the dimensionless single-mode linear growth rate is ζw=1. In small-w0cases, ˙w(t)
928 A37-16
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
in IP-s and AP-s cases is smaller than that of the single-mode growth rate during the
whole process. The value of ˙w(t)in k3AP-s and k2AP-s cases is larger than that of the
single-mode counterpart in early times (τw0.5), but smaller than that of the single-mode
counterpart in late stages. In large-w0cases, differently, ˙w(t)in IP-h and k2AP-h cases is
smaller than that of the single-mode growth rate during the whole process. The value
of ˙w(t)in k3AP-h and AP-h cases is larger than that of the single-mode counterpart
in early times (τw0.5), but smaller than that of the single-mode counterpart in late
stages. Therefore, both the phase and amplitude of the constituent modes influence the
mode-competition effect on the perturbation width growth of a multi-mode interface.
Besides, the growth rate of a multi-mode interface with all modes in-phase is smaller than
the counterpart in the classical single-mode case and other multi-mode cases. In addition,
under specific conditions, such as in k3AP-s and k3AP-h cases, the mode-competition
effect promotes the perturbation width growth in early stages.
Compared with the reference moving with the post-shock flow velocity, the width
growths of the bubble (wb(t)) and the spike (ws(t)) are separated, as shown in figures 7(c)
and 8(c)forsmall-and large-w0cases, respectively. The bubble/spike width is scaled
as ηb/s=k1[wb/s(t)Zcw0
0b/0s], in which w0b/0sis the initial bubble/spike width, as
listed in table 1. The bubble width of the multi-mode interface in all cases is lower
than that of the single-mode counterpart in late stages (τw0.5), which indicates that
the mode-competition effect advances the saturation of the bubble evolution in the 2-D
multi-mode RM instability. The nonlinear behaviour of the spike of a multi-mode interface
is greatly influenced by the initial spectra. In small-w0cases, ws(t)in IP-s and AP-s
cases are lower while ws(t)in k2AP-s and k3AP-s cases are higher than those of the
single-mode counterpart, which means that the phase of the constituent mode influences
the spike behaviour in the 2-D multi-mode RM instability. In large-w0cases, ws(t)in
k3AP-h, k2AP-h and AP-h cases are close to those of the single-mode counterpart, which
indicates that the initial amplitude of the constituent mode influences the spike behaviour.
The width growth rates of the bubble ˙wb(t)and the spike ˙ws(t)are calculated, as shown in
figures 7(d)and8(d)forsmall-and large-w0cases, respectively. The bubble/spike width
growth rate is scaled as ζb/s=2˙wb/s(t)/vMR
w. It is shown that the bubble width growth
rate of a multi-mode interface is generally lower than that of the single-mode counterpart.
The spike width growth rate reflects whether the mode-competition effect promotes or
suppresses the multi-mode RM instability in early stages.
Since the angle between the incident shock and UP (DP) is zero, no vorticity is deposited
at UP (DP). Therefore, UP and DP on the interface remain single-valued until the late
nonlinear stage before the transition. As a result, the perturbation width growth of a
multi-mode interface can be predicted by superimposing the perturbation growths of
initially constituent modes and the generated harmonics at xUP and xDP. In the present
coordinate system, the time-varying w(t)growth is superimposed by wb(t)and ws(t):
w(t)=wb(t)+ws(t), (3.12)
with
wb(t)=Zcw0b+
nt
0
vln
kncos(knxDP)dt,(3.13)
ws(t)=Zcw0s
nt
0
vln
kncos(knxUP)dt.(3.14)
928 A37-17
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
Case IP-s k3AP-s k2AP-s AP-s IP-h k3AP-h k2AP-h AP-h
˙w0
exp 30 ±232±239±234±237±236±249±239±2
˙w0
theo 28 29 38 32 34 35 52 41
Table 4. Comparison of the experimental initial perturbation width growth rate (˙wexp ) of the multi-mode
interface wit h the theoretical counter part ( ˙wtheo ) calculated with (3.15).
In this work, (3.12)–(3.14) constitute the ‘sum’ model which is used to calculate the
perturbation width by superimposing the amplitudes of three constituent modes and three
generated harmonics. By linear fitting of the perturbation width of a multi-mode interface
before 120 μs in all cases, the experimental initial perturbation width growth rate ( ˙w0
exp)
can be obtained, as listed in table 4. The theoretical initial perturbation width growth rate
of the interface ( ˙w0
theo) can be evaluated by superimposing the amplitude growth rates of
three constituent modes and three generated harmonics:
˙w0
theo =
6
n=1
vln
kn[cos(knxDP)cos(knxUP)].(3.15)
The values of ˙w0
theo in all cases are listed in table 4 and agree well with the experimental
results. Then, the predictions of the sum model for w(t),wb(t)and ws(t)are shown
in figures 7 and 8, and agree well with the experimental counterparts. Meanwhile, the
differentiation of the sum model is calculated and compared with the experimental ˙w(t),
˙wb(t)and ˙ws(t), and a good agreement is also achieved between them. Note that there are
several break points, for example in k3AP-h and AP-h cases, when the sum model is used
to predict the growth rates, which is ascribed to the enforced post-saturation treatment
adopted when ven
kis calculated. After the incident shock wave impacts the interface,
the transverse waves between the transmitted shock and reflected shock interact with the
interface, especially when the initial interface perturbation is prominent. Therefore, the
non-uniform flow influences the linear interface growth (Guo et al. 2020), resulting in the
perturbation width growth varying around the sum model prediction at an early regime.
To further validate the sum model, the experimental results with M(=1.3) and A+
(=0.67) similar to our experiments extracted from figure 4(a)inSadotet al. (1998)
are adopted as shown in figure 9(a). The initial interface is a two-bubble interface,
which is dominated by the first five order modes: a0
k1=1.49 ×103,a0
k2=1.06 ×
103,a0
k3=0.40 ×103,a0
k4=0.21 ×103,a0
k5=0.10 ×103m; k1=180 m1,ki=
ik1with i=2–5; φ1=π,φ2=φ3=φ4=φ5=0; Rk1=0.92, Rk2=0.89, Rk3=0.93,
Rk4=0.95 and Rk5=0.97; β=0.49. The amplitudes of the five constituent modes and
five generated harmonics are calculated by the present model, and then superimposed
according to the sum model. One can see that the predictions of the sum model agree well
with both the large bubble width and the small bubble width. Besides, the experimental
results with M(=1.2) and A+(=0.6) similar to our experiments extracted from
figure 9 in Luo et al. (2019) are shown in figure 9(b). For the CS-1 case (ICS-2 case),
the initial interface is a spike-dominated (bubble-dominated) chevron-shaped interface,
which is dominated by the first five order modes: a0
k1=1.08 ×103,a0
k2=0.81 ×103,
a0
k3=0.48 ×103,a0
k4=0.20 ×103,a0
k5=0.04 ×103m; k1=52 m1,ki=ik1with
i=2–5; for the CS-1 case, φ1=φ3=φ5=0, φ2=φ4=π; for the ICS-2 case, φ1=
928 A37-18
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
0.5 1.0 1.5 2.52.00
0.5
1.0
1.5
Impulsive
theory
DR model
0.5 1.0 1.5 2.52.0
0
0.5
1.0
1.5
k3AP-h
k2AP-h
IP-s
exp Sum
Bubble Bubble
Spike
Spike
AP-h
0.5 1.0 1.5 2.52.0
0
1.5
2.0
2.5
3.0
0.5
1.0
0.5
0
1.5
1.0
0.5 1.0 1.5 2.52.0
0
1.5
2.0
0.5
1.0
0.5
0
1.5
1.0
ηwζw
ηb/sζb/s
τwτw
(a)(b)
(c)(d)
Figure 8. Comparison of the perturbation width (a), growth rate of the perturbation width (b), widths of
bubble and spike (c)and width growth rates of bubble and spike (d)inlarge-w0cases.
φ3=φ5=π,φ2=φ4=0. For these two cases, Rk1=0.98, Rk2=0.97, Rk3=0.98,
Rk4=0.99 and Rk5=1.0; β=0.67. The amplitudes of the five constituent modes and
five generated harmonics are calculated and the sum model well predicts the experimental
results. Moreover, the numerical results with M(=5) and A+(=0.95) much larger
than in our experiments extracted from figure 4 in Pandian et al. (2017) are shown
in figure 9(c). The initial interface consists of two modes: a0
k1=a0
k2=1.1×103m;
k1=1.885 ×103m1,k2=2k1;φ1=π,φ2=0; Rk1=0.59, Rk2=0.42; β=0.26.
The amplitudes of the two constituent modes and two generated harmonics are calculated
and the sum model also well predicts the numerical results. All these agreements achieved
demonstrate the generality of the sum model.
4. Conclusions
In this work, a 2-D complex multi-mode interface constituted of various modes is first
formed by the soap-film technique, and then elaborate shock-tube experiments on the
developments of eight kinds of air–SF6multi-mode interface are performed. Based
on these well-controlled experiments and several theories in the literature, a general
nonlinear theory is established for predicting multi-mode evolution and mixing, and
928 A37-19
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
0 0.2 0.4 0.8
t (ms) k3vl
wt
k3(w(t)-Zcw0)
0.6 1.0 1.2 1.4
–4
0
4
12
8
0.5 1.0 1.5 2.52.0
0
1
3
2
4
exp Sum
Bubble height (mm)
Sadot et al. (1998) Luo et al. (2019)
CS-1
ICS-2
CS-1 (exp)
ICS-2 (exp)
CS-1 (sum)
ICS-2 (sum)
Figure 4(a)
Pandian et al. (2017)
Figure 4
Small bubble
Large bubble
num
Bubble
num
Sum
0.20.1 0.3 0.4 0.5 0.6
0.1
0
0.2
0.3
2πk1w
2πk1Δvt
(a)(b)
(c)
Figure 9. The perturbation width of the multi-mode interface obtained from (a) the experiments and
simulations in Sadot et al. (1998), (b) the experiments in Luo et al. (2019)and (c) the simulations in Pandian
et al. (2017).
finally a quantitative relation between the initial conditions and the perturbation growth
is constructed.
From the schlieren images of the shocked multi-mode interface, it is found that the
phases of the constituent modes greatly affect the initial interface shape and the later
interface evolution. It is also found that the transition from linear to nonlinear may
occur earlier when the initial interface amplitude is larger. By considering different
wavenumbers, initial amplitudes and phases of constituent modes, the dependence of the
perturbation growth on initial spectra is highlighted.
The captured interface morphology is distinct such that the interface contours in all
cases can be easily extracted by an image processing program. Subsequently, spectrum
analysis is performed on the interface contour before the interface becomes multi-valued,
and amplitudes of constituent modes are then acquired. It is first proved that the
mode-competition effect influences the amplitude growth of each mode from the very
beginning (quasi-linear stage), especially when the initial amplitudes of constituent modes
are large. It is interesting that the mode competition starts to play a role at this quasi-linear
stage although the growing of harmonics is so small as to be negligible. Therefore our
findings differ from previous views. The mode-competition effect is closely associated
with the initial spectra.
928 A37-20
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
A nonlinear theory is constructed by considering both the mode-competition effect
and the high-order harmonics effect to predict the amplitude growth of the modes. The
new theory has been validated by our experiments and data in the literature with the
consideration of diverse constituent modes, and a wide range of Mach number and Atwood
number. Further, the nonlinear theory is extended based on the superposition principle to
predict the growths of the total perturbation width and spike/bubble width, and there is
a satisfactory agreement between the predictions and the experimental results. It can be
concluded that the evolution of the shocked multi-mode interface has an evident memory
of the initial conditions from quasi-linear to late nonlinear stages.
The RM instability at the fully turbulent state has been a focus of attention recently.
We will combine the soap-film technique with a time-resolved particle image velocimetry
system to investigate the multi-mode RM instability induced by one shock or two shocks
in the near future. Besides, we look forward to examining the models established in this
work with experiments involving very high Atwood numbers.
Acknowledgements. The authors appreciate the valuable suggestions of the reviewers.
Funding. This work was supported by the Natural Science Foundation of China (nos. 91952205, 12022201,
11772329, 11625211 and 11621202) and Tamkeen under NYU Abu Dhabi Research Institute grant CG002.
Declaration of interests. The authors report no conflict of interest.
Author ORCIDs.
Yu Li a ng https://orcid.org/0000-0002-3254-7073;
Zhigang Zhai https://orcid.org/0000-0002-0094-5210;
Juchun Ding https://orcid.org/0000-0001-6578-1694;
Ting Si https://orcid.org/0000-0001-9071-8646;
Xisheng Luo https://orcid.org/0000-0002-4303- 8290.
REFERENCES
ABARZ HI , S.I. 2008 Coherent structures and pattern formation in Rayleigh–Taylor turbulent mixing. Phys.
Scr. 78 (1), 015401.
ABARZ HI , S.I. 2010 Review of theoretical modelling approaches of Rayleigh–Taylor instabilities and turbulent
mixing. Phil. Trans. R. Soc. A368 (1916), 1809–1828.
ALON,U.,HECHT,J.,MUKAMEL,D.&SHVARTS, D . 1994 Scale invariant mixing rates of hydrodynamically
unstable interface. Phys.Rev.Lett.72, 2867–2870.
ALON,U.,HECHT,J.,OFER,D.&SHVARTS, D. 1995 Power laws and similarity of Rayleigh–Taylor and
Richtmyer-Meshkov mixing fronts. Phys. Rev. Lett. 74, 534–537.
BALASUBRAMANIAN,S.,ORLICZ,G.C.&PRESTRIDGE, K.P. 2013 Experimental study of initial condition
dependence on turbulent mixing in shock-accelerated Richtmyer-Meshkov fluid layers. J. Turbul. 14 (3),
170–196.
BROUILLETTE, M. 2002 The Richtmyer-Meshkov instability. Annu. Rev. Fluid Mech. 34, 445–468.
CHANDRASEKHAR,S.2013Hydrodynamic and Hydromagnetic Stability. Courier Corporation.
COHEN, R.D. 1991 Shattering of a liquid drop due to impact. Proc. R. Soc. Lond. A435, 483–503.
COLLINS,B.D.&JACOBS, J.W. 2002 PLIF flow visualization and measurements of the Richtmyer-Meshkov
instability of an air/SF6interface. J. Fluid Mech. 464, 113–136.
DELL, Z., STELLINGWERF,R.F.&ABA RZHI, S.I. 2015 Effect of initial perturbation amplitude on
Richtmyer-Meshkov flows induced by strong shocks. Phys. Plasmas 22 (9), 092711.
DELL, Z.R., PANDIAN,A.,BHOWMICK, A.K., SWISHER, N.C., STANIC,M.,STELLINGWERF,R.F.&
ABARZ HI , S.I. 2017 Maximum initial growth-rate of strong-shock-driven Richtmyer-Meshkov instability.
Phys. Plasmas 24 (9), 090702.
DISTEFANO, C.A., MALAMUD,G.,KURANZ, C.C., KLEIN,S.R.&DRAKE,R.P.2015aMeasurement of
Richtmyer-Meshkov mode coupling under steady shock conditions and at high energy density. High Energy
Density Phys. 17, 263–269.
928 A37-21
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
DISTEFANO, C.A., MALAMUD,G.,KURANZ, C.C., KLEIN, S.R., STOECKL,C.&DRAKE,R.P.2015b
Richtmyer-Meshkov evolution under steady shock conditions in the high-energy-density regime. Appl.
Phys. Lett. 106 (11), 114103.
DIMONTE,G.,FRERKING, C.E., SCHNEIDER,M.&REMINGTON, B. 1996 Richtmyer-Meshkov instability
with strong radiatively driven shocks. Phys. Plasmas 3(2), 614–630.
DIMONTE,G.&RAMAPRABHU, P. 2010 Simulations and model of the nonlinear Richtmyer-Meshkov
instability. Phys. Fluids 22,014104.
DIMONTE,G.&SCHNEIDER, M. 2000 Density ratio dependence of Rayleigh–Taylor mixing for sustained
and impulsive acceleration histories. Phys. Fluids 12, 304–321.
DING, J.C., SI,T.,CHEN, M.J., ZHAI, Z.G., LU,X.Y.&LUO, X.S. 2017 On the interaction of a planar
shock with a three-dimensional light gas cylinder. J. Fluid Mech. 828, 289–317.
DRAZIN,P.G.&REID, W.H. 2004 Hydrodynamic Stability. Cambridge University Press.
ELBAZ,Y.&SHVARTS, D. 2018 Modal model mean field self-similar solutions to the asymptotic evolution
of Rayleigh–Taylor and Richtmyer-Meshkov instabilities and its dependence on the initial conditions. Phys.
Plasmas 25 (6), 062126.
GONCHAROV, V.N. 1999 Theory of the ablative Richtmyer-Meshkov instability. Phys. Rev. Lett. 82 (10),
2091.
GROOM,M.&THORNBER, B. 2020 The influence of initial perturbation power spectra on the growth of a
turbulent mixing layer induced by Richtmyer-Meshkov instability. Physica D407, 132463.
GUO,X.,ZHAI, Z., DING,J.,SI,T.&LUO, X. 2020 Effects of transverse shock waves on early evolution of
multi-mode chevron interface. Phys. Fluids 32 (10), 106101.
HAAN, S.W. 1989 Onset of nonlinear saturation for Rayleigh–Taylor growth in the presence of a full spectrum
of modes. Phys. Rev. A39 (11), 5812.
HAAN, S.W. 1991 Weakly nonlinear hydrodynamic instabilities in inertial fusion. Phys. Fluids B3,
2349–2355.
HECHT,J.,ALON,U.&SHVARTS, D. 1994 Potential flow models of Rayleigh–Taylor and
Richtmyer-Meshkov bubble fronts. Phys. Fluids 6, 4019–4030.
HOSSEINI, S.H.R. & TAKAYAMA, K. 2005 Experimental study of Richtmyer-Meshkov instability induced by
cylindrical shock waves. Phys. Fluids 17, 084101.
HURRICANE, O.A., BURKE, E., MAPLES,S.&VISWANATHAN, M. 2000 Saturation of Richtmyer’s
impulsive model. Phys. Fluids 12 (8), 2148–2151.
JACOB S,J.W.&KRIVETS, V.V. 2005 Experiments on the late-time development of single-mode
Richtmyer-Meshkov instability. Phys. Fluids 17, 034105.
JACOB S,J.W.&SHEELEY, J.M. 1996 Experimental study of incompressible Richtmyer-Meshkov instability.
Phys. Fluids 8, 405–415.
KURANZ,C.C.,et al. 2018 How high energy fluxes may affect Rayleigh–Taylor instability growth in young
supernova remnants. Nat. Commun. 9, 1564.
LAYZER, D. 1955 On the instability of superposed fluids in a gravitational field. Astrophys. J. 122, 1–12.
LEI,F.,DING,J.,SI,T.,ZHAI,Z.&LUO, X. 2017 Experimental study on a sinusoidal air/SF6interface
accelerated by a cylindrically converging shock. J. Fluid Mech. 826, 819–829.
LIANG,Y.,LIU, L., ZHAI, Z., SI,T.&LUO, X. 2021 Universal perturbation growth of Richtmyer-Meshkov
instability for minimum-surface featured interface induced by weak shock waves. Phys. Fluids 33 (3),
032110.
LIANG,Y.,ZHAI, Z., DING,J.&LUO, X. 2019 Richtmyer-Meshkov instability on a quasi-single-mode
interface. J. Fluid Mech. 872, 729–751.
LINDL,J.,LANDEN,O.,EDWARDS,J.,MOSES,E.&TEAM, N. 2014 Review of the national ignition
campaign 2009–2012. Phys. Plasmas 21, 020501.
LIU,H.&XIAO, Z. 2016 Scale-to-scale energy transfer in mixing flow induced by the Richtmyer-Meshkov
instability. Phys. Rev. E93 (5), 053112.
LIU, L., LIANG,Y.,DING,J.,LIU,N.&LUO, X. 2018 An elaborate experiment on the single-mode
Richtmyer-Meshkov instability. J. Fluid Mech. 853,R2.
LUO,X.,LIANG,Y.,SI,T.&ZHAI, Z. 2019 Effects of non-periodic portions of interface on
Richtmyer-Meshkov instability. J. Fluid Mech. 861, 309–327.
LUO,X.,LIU, L., LIANG,Y.,DING,J.&WEN, C.Y. 2020 Richtmyer-Meshkov instability on a dual-mode
interface. J. Fluid Mech. 905,A5.
LUO,X.,WANG,X.&SI, T. 2013 The Richtmyer-Meshkov instability of a three-dimensional air/SF6interface
with a minimum-surface feature. J. Fluid Mech. 722,R2.
LUO,X.,WANG,M.,SI,T.&ZHAI, Z. 2015 On the interaction of a planar shock with an SF6polygon.
J. Fluid Mech. 773 (2), 366–394.
928 A37-22
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
RM instability on 2-D multi-mode interfaces
MANSOOR, M.M., DALTON, S.M., MARTINEZ, A.A., DESJARDINS,T.,CHARONKO,J.J.&PRESTRIDGE,
K.P. 2020 The effect of initial conditions on mixing transition of the Richtmyer-Meshkov instability.
J. Fluid Mech. 904,A3.
MARIANI,C.,VANDENBOOMGAERDE,M.,JOURDAN,G.,SOUFFLAND,D.&HOUAS, L. 2008
Investigation of the Richtmyer-Meshkov instability with stereolithographed interfaces. Phys. Rev. Lett. 100,
254503.
MESHKOV, E.E. 1969 Instability of the interface of two gases accelerated by a shock wave. Fluid Dyn. 4,
101–104.
MEYER,K.A.&BLEWETT, P.J. 1972 Numerical investigation of the stability of a shock-accelerated interface
between two fluids. Phys. Fluids 15, 753–759.
MIKAELIAN, K.O. 1998 Analytic approach to nonlinear Rayleigh–Taylor and Richtmyer-Meshkov
instabilities. Phys. Rev. Lett. 80, 508–511.
MIKAELIAN, K. O. 2003 Explicit expressions for the evolution of single-mode Rayleigh–Taylor and
Richtmyer-Meshkov instabilities at arbitrary Atwood numbers. Phys. Rev. E67, 026319.
MIKAELIAN, K. O. 2005 Richtmyer-Meshkov instability of arbitrary shapes. Phys. Fluids 17, 034101.
MILES, A.R., EDWARDS, M.J., BLUE,B.,HANSEN, J.F., ROBEY, H.F., DRAKE, R.P., KURANZ,C.&
LEIBRANDT, D.R. 2004 The effects of a short-wavelength mode on the evolution of a long-wavelength
perturbatoin driven by a strong blast wave. Phys. Plasmas 11, 5507–5519.
MOHAGHAR,M.,CARTER,J.,MUSCI,B.,REILLY,D.,MCFARLAND,J.A.&RANJAN, D. 2017 Evaluation
of turbulent mixing transition in a shock-driven variable-density flow. J. Fluid Mech. 831, 779–825.
MOHAGHAR,M.,CARTER,J.,PATHIKONDA,G.&RANJAN, D. 2019 The transition to turbulence in
shock-driven mixing: effects of Mach number and initial conditions. J. Fluid Mech. 871, 595–635.
MÜGLER,C.&GAUTHIER, S. 1998 Numerical simulations of single-mode Richtmyer-Meshkov experiments.
Phys. Rev. E58 (4), 4548.
NIEDERHAUS, C.E. & JACO BS, J.W. 2003 Experimental study of the Richtmyer-Meshkov instability of
incompressible fluids. J. Fluid Mech. 485, 243–277.
NISHIHARA,K.,WOUCHUK, J.G., MATSUOKA,C.,ISHIZAKI,R.&ZHAKHOVSKY,V.V.2010
Richtmyer-Meshkov instability: theory of linear and nonlinear evolution. Phil. Trans. R. Soc. A368,
1769–1807.
OFER,D.,ALON,U.,SHVARTS,D.,MCCRORY, R.L. & VERDON, C. P. 1996 Modal model for the nonlinear
multimode Rayleigh–Taylor instability. Phys. Plasmas 3(8), 3073–3090.
ORON,D.,ARAZI, L., KARTOON,D.,RIKANATI,A.,ALON,U.&SHVARTS, D. 2001 Dimensionality
dependence of the Rayleigh–Taylor and Richtmyer-Meshkov instability late-time scaling laws. Phys.
Plasmas 8, 2883–2889.
PANDIAN,A.,STELLINGWERF,R.F.&ABARZHI, S .I. 2017 Effect of a relative phase of waves
constituting the initial perturbation and the wave interference on the dynamics of strong-shock-driven
Richtmyer-Meshkov flows. Phys. Rev. Fluids 2(7), 073903.
RANJAN,D.,ANDERSON,M.,OAKLEY,J.&BONAZZA, R. 2005 Experimental investigation of a strongly
shocked gas bubble. Phys. Rev. Lett. 94, 184507.
RAYLEIGH, L. 1883 Investigation of the character of the equilibrium of an incompressible heavy fluid of
variable density. Proc. Lond. Math. Soc. 14, 170–177.
REMINGTON, B.A., WEBER, S.V., MARINAK, M.M., HAAN, S.W., KILKENNY, J.D., WALLACE,R.J.&
DIMONTE, G 1995 Single-mode and multimode Rayleigh–Taylor experiments on nova. Phys. Plasmas 2
(1), 241–255.
RICHTMYER, R.D. 1960 Taylor instability in shock acceleration of compressible fluids. Commun. Pure Appl.
Maths 13, 297–319.
RIKANATI,A.,ALON,U.&SHVARTS, D. 1998 Vortex model for the nonlinear evolution of the multimode
Richtmyer-Meshkov instability at low Atwood numbers. Phys. Rev. E58, 7410–7418.
RIKANATI,A.,ORON,D.,SADOT,O.&SHVARTS, D. 2003 High initial amplitude and high Mach number
effects on the evolution of the single-mode Richtmyer-Meshkov instability. Phys. Rev. E67, 026307.
SADOT,O.,EREZ, L., ALON,U.,ORON ,D.,LEVIN, L.A., EREZ,G.,BEN-DOR,G.&SHVARTS,D.
1998 Study of nonlinear evolution of single-mode and two-bubble interaction under Richtmyer-Meshkov
instability. Phys.Rev.Lett.80, 1654–1657.
SEWELL, E.G., FERGUSON, K.J., KRIVETS,V.V.&JACO BS , J.W. 2021 Time-resolved particle image
velocimetry measurements of the turbulent Richtmyer–Meshkov instability. J. Fluid Mech. 917,A41.
SI,T.,LONG,T.,ZHAI,Z.&LUO, X. 2015 Experimental investigation of cylindrical converging shock waves
interacting with a polygonal heavy gas cylinder. J. Fluid Mech. 784, 225–251.
TAYL OR, G. 1950 The instability of liquid surfaces when accelerated in a direction perpendicular to their
planes. I. Proc. R. Soc. Lond. A201 (1065), 192–196.
928 A37-23
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Y. Liang, L. Liu, Z. Zhai, J. Ding, T. Si and X. Luo
THORNBER, B. 2016 Impact of domain size and statistical errors in simulations of homogeneous decaying
turbulence and the Richtmyer-Meshkov instability. Phys. Fluids 28 (4), 045106.
THORNBER,B.,DRIKAKIS,D.,YOUNGS, D.L. & WILLIAMS, R.J.R. 2010 The influence of initial condition
on turbulent mixing due to Richtmyer-Meshkov instability. J. Fluid Mech. 654, 99–139.
VANDENBOOMGAERDE,M.,GAUTHIER,S.&MÜGLER, C. 2002 Nonlinear regime of a multimode
Richtmyer-Meshkov instability: a simplified perturbation theory. Phys. Fluids 14 (3), 1111–1122.
VANDENBOOMGAERDE,M.,SOUFFLAND,D.,MARIANI,C.,BIAMINO, L., JOURDAN,G.&HOUAS,
L. 2014 An experimental and numerical investigation of the dependency on the initial conditions of the
Richtmyer-Meshkov instability. Phys. Fluids 26, 024109.
VELIKOVICH,A.,HERRMANN,M.&ABAR ZHI, S. 2014 Perturbation theory and numerical modelling of
weakly and moderately nonlinear dynamics of the incompressible Richtmyer-Meshkov instability. J. Fluid
Mech. 751, 432–479.
VELIKOVICH, A.L. & DIMONTE, G. 1996 Nonlinear perturbation theory of the incompressible
Richtmyer-Meshkov instability. Phys.Rev.Lett.76 (17), 3112.
ZHAI, Z., DONG,P.,SI,T.&LUO, X. 2016 The Richtmyer-Meshkov instability of a V shaped air/helium
interface subjected to a weak shock. Phys. Fluids 28 (8), 082104.
ZHAI, Z., WANG,M.,SI,T.&LUO, X . 2014 On the interaction of a planar shock with a light SF6polygonal
interface. J. Fluid Mech. 757 (2), 800–816.
ZHAI, Z., ZOU, L., WU,Q.&LUO, X. 2018 Review of experimental Richtmyer-Meshkov instability in shock
tube: from simple to complex. Proc. Inst. Mech. Engrs C232, 2830–2849.
ZHANG,Q.&SOHN, S.I . 1997 Nonlinear theory of unstable fluid mixing driven by shock wave. Phys. Fluids
9, 1106–1124.
ZHOU, Y. 2007 Unification and extension of the similarity scaling criteria and mixing transition for studying
astrophysics using high energy density laboratory experiments or numerical simulations. Phys. Plasmas 14
(8), 082701.
ZHOU,Y.2017aRayleigh–Taylor and Richtmyer-Meshkov instability induced flow, turbulence, and mixing. I.
Phys. Rep. 720–722, 1–136.
ZHOU,Y.2017bRayleigh–Taylor and Richtmyer-Meshkov instability induced flow, turbulence, and mixing. II.
Phys. Rep. 723–725, 1–160.
ZHOU,Y.,CLARK, T.T., CLARK, D.S., GLENDINNING, S.S., SKINNER, A.A., HUNTINGTON,C.,
HURRICANE, O.A., DIMITS,A.M.&REMINGTON, B.A. 2019 Turbulent mixing and transition criteria
of flows induced by hydrodynamic instabilities. Phys. Plasmas 26 (8), 080901.
ZHOU,Y.,ROBEY,H.F.&BUCKINGHAM, A.C. 2003 Onset of turbulence in accelerated
high-Reynolds-number flow. Phys. Rev. E67 (5), 056305.
928 A37-24
https://doi.org/10.1017/jfm.2021.849
Downloaded from https://www.cambridge.org/core. NYU Grossman School of Medicine, on 15 Oct 2021 at 13:21:41, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
... In te rf ac e S tr ip s SF 6 Val ve complicated mechanisms are involved. In our previous work, a smoothly convergent shock was generated based on shock dynamics theory (Zhai et al. 2010;Zhan et al. 2018;Luo et al. 2019), and a well-defined initial multi-mode interface was formed by the soap-film technique (Liang et al. 2021a;Guo et al. 2022). These provide us an opportunity to explore the developments of multi-mode interfaces accelerated by a convergent shock. ...
... The soap-film technique (Luo et al. 2016;Liang et al. 2021a;Guo et al. 2022) is used to form the well-defined dual-mode interfaces. Before interface formation, the transparent devices (devices A and B) shown in figure 1(b) are manufactured by combining two transparent acrylic plates (3.0 mm in thickness) with pedestals (7.0 mm in height). ...
... A similar conclusion was also drawn by Miles et al. (2004). Liang et al. (2021a) reported in their planar multi-mode RM instability study that although the mode-coupling effects reduce global mixing, they enhance local mixing. The perturbation width seems to be affected less significantly by initial phase difference. ...
Article
Full-text available
We report the first shock-tube experiments on two-dimensional dual-mode air–SF $_6$ interfaces with different initial spectra subjected to a convergent shock wave. The convergent shock tube is specially designed with a tail opening to highlight the Bell–Plesset (BP) and mode-coupling effects on amplitude development of fundamental mode (FM). The results show that the BP effect promotes the occurrence of mode coupling, and the feedback of high-order modes to the FM also arises earlier in convergent geometry than that in its planar counterpart. Relatively, the amplitude growth of the FM with a higher mode number is inhibited by the feedback, and saturates earlier. The FM with a lower mode number is affected more heavily by the BP effect, and finally dominates the flow. A new model is proposed to well predict the amplitude growths of the FM and high-order modes in convergent geometry. In particular, for FM that reaches its saturation amplitude, the post-saturation relation is introduced in the model to achieve a better prediction.
... (i) Theoretically, Haan (1991) first proposed a modal model with second-order accuracy to quantify the influence of the mode-coupling between the multiple modes on the RT instability. The modal model and its extended types have achieved a wide range of validation in the issues of RT instability (Remington et al. 1995;Ofer et al. 1996;Elbaz & Shvarts 2018) and RM instability (Liang et al. 2021). Assuming that the mode-coupling is absent before each bubble of a multimode RM unstable interface reaches its asymptotic growth, Alon et al. (1994) proposed a statistical potential flow model to predict the eventual average bubble distribution and bubble amplitude growth rate. ...
... Earlier mixing transitions for higher amplitude-to-wavelength ratio cases were noted from the experiments. Liang et al. (2021) investigated the RM instability on a multimode air-SF 6 interface initially dominated by three modes. It was revealed that the mode-coupling is closely related to the initial spectrum and plays an essential role in RM flows from the very beginning if the initial amplitudes of the constituent modes are large. ...
... Benefiting from the open test section, as sketched in figure 1(a), a removable interface formation device can be efficiently designed in which a shape-controllable gaseous interface can be generated using the extended soap film technique. The soap film technique can essentially eliminate the additional short-wavelength perturbations, diffusion layer and three-dimensionality Liang et al. 2021). As shown in figures 1(b) and 1(c), a device consisting of two semicircular transparent acrylic sheets with a spacing of 5.0 mm is fixed on a rectangular base. ...
Article
Full-text available
Shock-tube experiments are performed on the convergent Richtmyer-Meshkov (RM) instability of a multimode interface. The temporal growth of each Fourier mode perturbation is measured. The hydrodynamic instabilities, including the RM instability and the additional Rayleigh-Taylor (RT) effect, imposed by the convergent shock wave on the dual-mode interface, are investigated. The mode-coupling effect on the convergent RM instability coupled with the RT effect is quantified. It is evident that the amplitude growths of all first-order modes and second-order harmonics and their couplings depend on the variance of the interface radius, and are influenced by the mode-coupling from the very beginning. It is confirmed that the mode-coupling mechanism is closely related to the initial spectrum, including azimuthal wavenumbers, relative phases and initial amplitudes of the constituent modes. Different from the conclusion in previous studies on the convergent single-mode RM instability that the additional RT effect always suppresses the perturbation growth, the mode-coupling might result in the additional RT effect promoting the instability of the constituent Fourier mode. By considering the geometry convergence, the mode-coupling effect and other physical mechanisms, second-order nonlinear solutions are established to predict the RM instability and the additional RT effect in the cylindrical geometry, reasonably quantifying the amplitude growths of each mode, harmonic and coupling. The nonlinear solutions are further validated by simulations considering various initial spectra. Last, the temporal evolutions of the mixed mass and normalized mixed mass of a shocked multimode interface are calculated numerically to quantify the mixing of two fluids in the cylindrical geometry.
... Shock waves accelerating a density interface cause complex physical phenomena, including the interface amplitude growth, wave pattern evolution, and vorticity generation [1][2][3]. Richtmyer introduced the linear theory for this process in 1960 [4], which was later confirmed by Meshkov's shock tube experiments [5]. This impulsive version of the Rayleigh-Taylor instability (RTI) [6,7] is known as the Richtmyer-Meshkov instability (RMI) [8]. ...
... III are based on the simulations using the experimental configuration outlined in Ding et al. [37], with a specific emphasis on investigating the Richtmyer-Meshkov instability induced by a single perturbed cylindrical interface in a shocked medium. The interface shape is modeled after the air-SF 6 3], where the cylindrical interface is centered at (0, 0). The initial density interface (DI) has a radius of R = 1 and is perturbed with an amplitude of a = 0.04 using the coordinate r = R + a sin(nθ + π/2), where n is the azimuthal mode number and θ is the azimuthal angle. ...
Article
Full-text available
The investigation of the converging shock-induced Richtmyer-Meshkov instability, which arises from the interaction of converging shocks with the interface between materials of differing densities in cylindrical capsules, is of significant importance in the field of inertial confinement fusion (ICF). The use of converging shocks, which exhibit higher efficiency than planar shocks in the development of the RMI due to the Bell-Plesset effects, is particularly relevant to energy production in the ICF. Moreover, external magnetic fields are often utilized to mitigate the development of the RMI. This paper presents a systematic investigation of the anisotropic nature of the Richtmyer-Meshkov instability in magnetohydrodynamic induced by the interaction between converging shocks and perturbed semicylindrical density interfaces (DI) based on numerical simulations using Athena++. The results reveal that magnetic fields with β=1000, 100, and 10 (β is defined as the ratio of the plasma pressure to the magnetic pressure) lead to an anisotropic intensification of magnetic fields, anisotropic accelerations of various shock waves [including the converging incident shock (CIS), transmitted shock (TS), and reflected shock (RS)], and anisotropic growth of the DI with subsequent anisotropic vorticity distribution. Upon closer inspection, it becomes evident that these phenomena are strongly interconnected. In particular, the region where the wave front of the CIS impacts the middle point of semicylindrical DI, where the magnetic field is more perpendicular to the fluid motion, experiences a more significant amplification of the magnetic fields. This generates higher-pressure jumps, which in turn accelerates the shock wave near this region. Furthermore, the anisotropic amplification of the magnetic fields reduces the movement of the RMI near the middle point of semicylindrical DI and leads to the anisotropic formation of RMI-induced bubbles and spikes, as well as vortices. By examining vorticity distributions, the results underscore the crucial role of magnetic tension forces in inhibiting fluid rotation.
... These grids block the flow and influence the shock movement and interface deformation. Recently, the soap-film technique has been proposed to create single-mode, multi-mode, and polygonal interfaces (Zhai et al. 2014;Liu et al. 2018;Liang et al. 2021). Note that the soap-film interface has a discontinuous feature relative to membrane-free interface (Motl et al. 2009), and it is easily broken compared to interfaces formed by jelly (Meshkov and Abarzhi 2019) or nitrocellulose membranes Mariani et al. 2008). ...
... To maintain the initial interface shape, the soap-film interface must be constrained, for example, by filaments. In the previous experiments (Liu et al. 2018;Liang et al. 2021), two constraint filaments were used to form the soap-film interface. The effect of filaments on post-shock flow features has been investigated recently . ...
Article
Full-text available
We propose a new interface formation method for shock–interface interaction studies by using the super-hydrophobic–oleophobic surface instead of filaments to constrain the soap–film interface. To verify this method, developments of a single-mode air–SF6\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$_6$$\end{document} interface and a heavy gas layer accelerated by shock waves are experimentally investigated and compared with the previous studies. For single-mode interface developments, experimental schlieren images show that the interfaces are more fully developed, and the thickness of the interface profile reduces more than 60%. For shock-induced heavy gas layer instability, the interface profile is more distinct, and the mixing width of the upstream interface after it passes through the initial position of the downstream interface is largely weakened. Quantitative comparison shows that the filaments used to constrain the soap–film interface have a significant effect on the movement and amplitude growth of the upstream interface, and the superiority of the present method is well demonstrated. Graphical abstract
... The RTI and RMI have been extensively studied on semiinfinite gas interfaces due to their fundamental significance [7][8][9][10][11][12][13][14][15]. However, in practical applications such as ICF, these instabilities occur in multi-layer fluids. ...
Article
Full-text available
Hydrodynamic instabilities induced by a shock wave can be observed in both natural phenomena and engineering applications, and are frequently employed to study gas dynamics, vortex dynamics, and turbulence. Controlling these instabilities is very desirable, but remains a challenge in applications such as inertial confinement fusion. The field of "shock-gas-layer interac-tion" has experienced rapid development, driven by advances in experimental and numerical techniques as well as theoretical understanding. This domain has uncovered a diverse array of wave patterns and hydrodynamic instabilities, such as reverberating waves, feedthrough, abnormal and freeze-out Richtmyer-Meshkov instability, among others. Studies have shown that it is possible to suppress these instabilities by appropriately configuring a gas layer. Here we review the recent progress in theories, experiments, and simulations of shock-gas-layer interactions, and the feedthrough mechanism, the reverberating waves and their induced additional instabilities, as well as the convergent geometry and reshock effects, are focused. The conditions for suppressing hydrodynamic instabilities are summarized. The review concludes by highlighting the challenges and prospects for future research in this area.
... In terms of the flow cross-section area, RM instability can be categorized into two types: area-invariant RM instability and area-varied RM instability. The former usually refers to planar shock-induced RM instability, which has been extensively studied by experimentalists (Biamino et al. 2015;Reese et al. 2018;Liang et al. 2021;Sewell et al. 2021), theorists (Richtmyer 1960;Zhang & Sohn 1997;Dimonte & Ramaprabhu 2010;Zhang & Guo 2016) and numerical experts (Schilling & Latini 2010;Lombardini, Pullin & Meiron 2014;Wonga & Lelea 2017;Li et al. 2022). It is widely accepted that pressure disturbance (caused by pressure waves behind the refracted shock) and baroclinic vorticity (caused by the misalignment of pressure and density gradients) are the major mechanisms for the growth of area-invariant RM instability (Brouillette 2002;Ranjan, Oakley & Bonazza 2011;Zhou 2017). ...
Article
Full-text available
Experiments on divergent Richtmyer–Meshkov (RM) instability at a heavy gas layer are performed in a specially designed shock tube. A novel soap-film technique is extended to generate gas layers with controllable thicknesses and shapes. An unperturbed gas layer is first examined and its two interfaces are found to move uniformly at the early stage and be decelerated later. A general one-dimensional theory applicable to an arbitrary-thickness layer is established, which gives a good prediction of the layer motion in divergent geometry. Then, six kinds of perturbed SF $_6$ layers with various thicknesses and shapes surrounded by air are examined. At the early stage, the amplitude growths of the inner interface for various-thickness layers collapse quite well and also can be predicted by the Bell model for cylindrical RM instability at a single interface, which indicates a negligible interface coupling effect. Later, a rarefaction wave accelerates the inner interface, causing a dramatic rise in the growth rate. It is found that a thicker gas layer will result in a larger extent that the rarefaction wave can promote the instability growth. A modified Bell model accounting for both Rayleigh–Taylor (RT) instability and interface stretching caused by a rarefaction wave is established, which well reproduces the quick instability growth. At late stages, reverberating waves inside the layer are negligibly weak such that the inner interface growth is dominated by RM instability and RT stability. The major factors driving the outer interface development are a compression wave and interface coupling. A new interface coupling phenomenon existing uniquely in divergent geometry caused by the gradual thinning of the gas layer is observed and also modelled.
... The soap film technique is extended to form two shape-controllable, discontinuous slow/fast interfaces, mainly reducing the additional short-wavelength disturbances, interface diffusion and three-dimensionality Liang et al. 2019Liang et al. , 2021. As shown in figure 1(a), three transparent devices (i.e. ...
Article
Full-text available
Shock-induced instability developments of two successive interfaces have attracted much attention, but remain a difficult problem to solve. The feedthrough and reverberating waves between two successive interfaces significantly influence the hydrodynamic instabilities of the two interfaces. The evolutions of two successive slow/fast interfaces driven by a weak shock wave are examined experimentally and numerically. First, a general one-dimensional theory is established to describe the movements of the two interfaces by studying the rarefaction waves reflected between the two interfaces. Second, an analytical, linear model is established by considering the arbitrary wavenumber and phase combinations and compressibility to quantify the feedthrough effect on the Richtmyer-Meshkov instability (RMI) of two successive slow/fast interfaces. The feedthrough significantly influences the RMI of the two interfaces, and even leads to abnormal RMI (i.e. phase reversal of a shocked slow/fast interface is inhibited) which is the first observational evidence of the abnormal RMI provided by the present study. Moreover, the stretching effect and short-time Rayleigh-Taylor instability or Rayleigh-Taylor stabilisation imposed by the rarefaction waves on the two interfaces are quantified considering the two interfaces' phase reversal. The conditions and outcomes of the freeze-out and abnormal RMI caused by the feedthrough are summarised based on the theoretical model and numerical simulation. A specific requirement for the simultaneously freeze-out of the instability of the two interfaces is proposed, which can potentially be used in the applications to suppress the hydrodynamic instabilities.
Article
Shock-tube experiments and theoretical studies have been performed to highlight mode-coupling in an air–SF $_6$ –air fluid layer. Initially, the two interfaces of the layer are designed as single mode with different basic modes. It is found that as the two perturbed interfaces become closer, interface coupling induces a different mode from the basic mode on each interface. Then mode coupling further generates new modes. Based on the linear model (Jacobs et al. , J. Fluid Mech. , vol. 295, 1995, pp. 23–42), a modified model is established by considering the different accelerations of two interfaces and the waves’ effects in the layer, and provides good predictions to the linear growth rates of the basic modes and the modes generated by interface coupling. It is observed that interface coupling behaves differently to the nonlinear growth of the basic modes, which can be characterized generally by the existing or modified nonlinear model. Moreover, a new modal model is established to quantify the mode-coupling effect in the layer. The mode-coupling effect on the amplitude growth is negligible for the basic modes, but is significant for the interface-coupling modes when the initial wavenumber of one interface is twice the wavenumber of the other interface. Finally, amplitude freeze-out of the second single-mode interface is achieved theoretically and experimentally through interface coupling. These findings may be helpful for designing the target in inertial confinement fusion to suppress the hydrodynamic instabilities.
Article
The Richtmyer–Meshkov instability causes perturbations to grow after a shock traverses a fluid density interface. This increases the mixing rate between fluid from either side of the interface. We use the Flash Eulerian hydrodynamic code to investigate alterations when a thin third layer of intermediate density is placed along the interface, effectively creating two adjacent unstable interfaces. This is a common occurrence in engineering applications where a thin barrier initially separates two materials. We find that the width of the mixing layer is similar or slightly reduced; however, the total mass of mixed material can actually increase. The mixing layer becomes more compact and efficient. However, the normalized mixed mass decreases, meaning that finger entrainment becomes more important than in the simple two-layer case. The effect of adding the central layer appears to decrease when the Atwood number is decreased. The Flash results are also benchmarked against two-layer experimental data from a shock tube at the University of Arizona.
Article
Full-text available
Shock-tube experiments on Richtmyer–Meshkov (RM) instability at a perturbed SF $_6$ layer surrounded by air, induced by a cylindrical divergent shock, are reported. To explore the effects of reverberating waves and interface coupling on instability growth, gas layers with various shapes are created: unperturbed inner interface and sinusoidal outer interface (case US); sinusoidal inner and outer interfaces that have identical phase (case IP); sinusoidal inner and outer interfaces that have opposite phase (case AP). For each case, three thicknesses are considered. Results show that reverberating waves inside the layer dominate the early-stage instability growth, while interface coupling dominates the late-stage growth. The influences of waves on divergent RM instability are more pronounced than the planar and convergent counterparts, which are estimated accurately based on gas dynamics theory. Both the wave influence and interface coupling depend heavily on the layer shape, leading to diverse growth rates: the quickest growth for case AP, medium growth for case US, the slowest growth for case IP. In particular, for the IP case, there exists a critical thickness below which the instability growth is suppressed by both the reverberating waves and interface coupling. This provides an efficient way to modulate the growth of divergent RM instability. It is found that divergent RM instability involves weak nonlinearity and strong interface coupling such that the linear theory of Mikaelian ( Phys. Fluids , vol. 17, 2005, 094105) can well reproduce the instability growth at late stages for all cases. This constitutes the first experimental confirmation of the Mikaelian theory.
Article
Full-text available
Experimental and theoretical investigations are performed to explore the development of Richtmyer–Meshkov (RM) instability for a minimum-surface featured (3D-) interface. The exact mathematical expression of 3D-interface perturbation is obtained for the first time by the spectrum analysis, describing as a superposition of transverse two-dimensional (2D) single-mode with three-dimensional (3D) multi-mode. In particular, the normalized 3D-interface profile is found to be solely determined by one dimensionless parameter related to the 3D-interface initial spectrum. The shock tube experiments are performed by varying the interface height to change the mode-composition of 3D-interfaces under weak shock conditions. It is found that the 3D multi-mode component of a 3D-interface promotes/suppresses the RM instability at the transverse boundary/symmetry plane in comparison with the classical 2D single-mode case. At the linear regime, the 3D perturbation growth can be well predicted by combining the amplitude growth of a 2D single-mode and a 3D dual-mode. At the nonlinear regime, as the interface height reduces, the nonlinear effect on the RM instability at the boundary plane becomes stronger. A generalized nonlinear model is established to predict the interface amplitude by considering the interface spectrum and the mode-coupling of 3D modes. It is found that the mode-coupling has an evident influence on the bubble evolution, and the first-order 3D mode leads to different behaviors for the bubble and spike width growths. This work may provide great insight into the physical mechanism of the 3D RM instability existing in practical applications.
Article
Full-text available
Deep neural networks for nonlinear model order reduction of unsteady flows Physics of Fluids 32, 105104 (2020); https://doi.org/10.1063/5.0020526 Experimental investigation of the transonic shock-wave/boundary-layer interaction over a shock-generation bump Physics of Fluids 32, 106102 (2020); https://doi.org/10.1063/5.0018763 Ultrafast tomographic particle image velocimetry investigation on hypersonic boundary layers Physics of Fluids 32, 094103 (2020); https://doi. ABSTRACT Effects of transverse shock waves are important in the evolution of a multi-mode interface. However, the related experimental studies are scarce due to the difficulty in creating a well-defined interface. In the present work, we realized such an experimental study by using the soap film technique to form a multi-mode chevron air/SF 6 interface. By changing the shock Mach number and the initial amplitude of the interface, the intensity of the transverse shock waves is varied. It is found that the impact of transverse shock waves together with the shock proximity effects flattens the bubble front and reduces the amplitude growth rate. For small initial amplitudes where the transverse shock waves are weak enough, the interface deforms little and the mode coupling is proven to be weak. For high initial amplitudes, the inverse cascade of modes causes the amplitude increase (decrease) of the first mode (high-order modes) at low Mach numbers. As the Mach number increases, the transverse shock waves and the shock proximity effects introduce external forces to the flow, resulting in the generation of additional high-order modes and the reduction in the first mode amplitude. Specifically, the augment of the second harmonic mode amplitude is crucial to flattening the bubble front.
Article
Full-text available
This paper investigates the influence of different broadband perturbations on the evolution of a Richtmyer–Meshkov turbulent mixing layer initiated by a Mach 1.84 shock traversing a perturbed interface separating gases with a density ratio of 3:1. Both the bandwidth of modes in the interface perturbation, as well as their relative amplitudes, are varied in a series of carefully designed numerical simulations at grid resolutions up to 3.2×109 cells. Three different perturbations are considered, characterised by a power spectrum of the form P(k)∝km where m=−1, −2 and −3. The growth of the mixing layer is shown to strongly depend on the initial conditions, with the growth rate exponent θ found to be 0.5, 0.63 and 0.75 for each value of m at the highest grid resolution. The asymptotic values of the molecular mixing fraction Θ are also shown to vary significantly with m; at the latest time considered Θ is 0.56, 0.39 and 0.20 respectively. Turbulent kinetic energy (TKE) is also analysed in both the temporal and spectral domains. The temporal decay rate of TKE is found not to match the predicted value of n=2−3θ, which is shown to be due to a time-varying normalised dissipation rate Cϵ. In spectral space, the data follow the theoretical scaling of k(m+2)∕2 at low wavenumbers and tend towards k−3∕2 and k−5∕3 scalings at high wavenumbers for the spectra of transverse and normal velocity components respectively. The results represent a significant extension of previous work on the Richtmyer–Meshkov instability evolving from broadband initial perturbations and provide useful benchmarks for future research.
Article
Full-text available
In diverse areas of science and technology, including inertial confinement fusion (ICF), astrophysics, geophysics, and engineering processes, turbulent mixing induced by hydrodynamic instabilities is of scientific interest as well as practical significance. Because of the fundamental roles they often play in ICF and other applications, three classes of hydrodynamic instability-induced turbulent flows—those arising from the Rayleigh-Taylor, Richtmyer-Meshkov, and Kelvin-Helmholtz instabilities—have attracted much attention. ICF implosions, supernova explosions, and other applications illustrate that these phases of instability growth do not occur in isolation, but instead are connected so that growth in one phase feeds through to initiate growth in a later phase. Essentially, a description of these flows must encompass both the temporal and spatial evolution of the flows from their inception. Hydrodynamic instability will usually start from potentially infinitesimal spatial perturbations, will eventually transition to a turbulent flow, and then will reach a final state of a true multiscale problem. Indeed, this change in the spatial scales can be vast, with hydrodynamic instability evolving from just a few microns to thousands of kilometers in geophysical or astrophysical problems. These instabilities will evolve through different stages before transitioning to turbulence, experiencing linear, weakly, and highly nonlinear states. The challenges confronted by researchers are enormous. The inherent difficulties include characterizing the initial conditions of such flows and accurately predicting the transitional flows. Of course, fully developed turbulence, a focus of many studies because of its major impact on the mixing process, is a notoriously difficult problem in its own right. In this pedagogical review, we will survey challenges and progress, and also discuss outstanding issues and future directions.
Article
Full-text available
Experiments on Richtmyer–Meshkov instability of quasi-single-mode interfaces are performed. Four quasi-single-mode air/ $\text{SF}_{6}$ interfaces with different deviations from the single-mode one are generated by the soap film technique to evaluate the effects of high-order modes on amplitude growth in the linear and weakly nonlinear stages. For each case, two different initial amplitudes are considered to highlight the high-amplitude effect. For the single-mode and saw-tooth interfaces with high initial amplitude, a cavity is observed at the spike head, providing experimental evidence for the previous numerical results for the first time. For the quasi-single-mode interfaces, the fundamental mode is the dominant one such that it determines the amplitude linear growth, and subsequently the impulsive theory gives a reasonable prediction of the experiments by introducing a reduction factor. The discrepancy in linear growth rates between the experiment and the prediction is amplified as the quasi-single-mode interface deviates more severely from the single-mode one. In the weakly nonlinear stage, the nonlinear model valid for a single-mode interface with small amplitude loses efficacy, which indicates that the effects of high-order modes on amplitude growth must be considered. For the saw-tooth interface with small amplitude, the amplitudes of the first three harmonics are extracted from the experiment and compared with the previous theory. The comparison proves that each initial mode develops independently in the linear and weakly nonlinear stages. A nonlinear model proposed by Zhang & Guo ( J. Fluid Mech. , vol. 786, 2016, pp. 47–61) is then modified by considering the effects of high-order modes. The modified model is proved to be valid in the weakly nonlinear stage even for the cases with high initial amplitude. More high-order modes are needed to match the experiment for the interfaces with a more severe deviation from the single-mode one.
Article
Full-text available
The effects of incident shock strength on the mixing transition in the Richtmyer–Meshkov instability (RMI) are experimentally investigated using simultaneous density–velocity measurements. This effort uses a shock with an incident Mach number of 1.9, in concert with previous work at Mach 1.55 (Mohaghar et al., J. Fluid Mech., vol. 831, 2017 pp. 779–825) where each case is followed by a reshock wave. Single- and multi-mode interfaces are used to quantify the effect of initial conditions on the evolution of the RMI. The interface between light and heavy gases (N2/CO2, Atwood number, A~0.22; amplitude to wavelength ratio of 0.088) is created in an inclined shock tube at 80 degree relative to the horizontal, resulting in a predominantly single-mode perturbation. To investigate the effects of initial perturbations on the mixing transition, a multi-mode inclined interface is also created via shear and buoyancy superposed on the dominant inclined perturbation. The evolution of mixing is investigated via the density fields by computing mixed mass and mixed-mass thickness, along with mixing width, mixedness and the density self-correlation (DSC). It is shown that the amount of mixing is dependent on both initial conditions and incident shock Mach number. Evolution of the density self-correlation is discussed and the relative importance of different DSC terms is shown through fields and spanwise-averaged profiles. The localized distribution of vorticity and the development of roll-up features in the flow are studied through the evolution of interface wrinkling and length of the interface edge, which indicate that the vorticity concentration shows a strong dependence on the Mach number. The contribution of different terms in the Favre-averaged Reynolds stress is shown, and while the mean density-velocity fluctuation correlation term, <𝜌><u'iu'j>, is dominant, a high dependency on the initial condition and reshock is observed for the turbulent mass-flux term. Mixing transition is analysed through two criteria: the Reynolds number (Dimotakis, J. Fluid Mech., vol. 409, 2000, pp. 69–98) for mixing transition and Zhou (Phys. Plasmas, vol. 14 (8), 2007, 082701 for minimum state) and the time-dependent length scales (Robey et al., Phys. Plasmas, vol. 10 (3), 2003, 614622; Zhou et al., Phys. Rev. E, vol. 67 (5), 2003, 056305). The Reynolds number threshold is surpassed in all cases after reshock. In addition, the Reynolds number is around the threshold range for the multi-mode, high Mach number case (M~1.9) before reshock. However, the time-dependent length-scale threshold is surpassed by all cases only at the latest time after reshock, while all cases at early times after reshock and the high Mach number case at the latest time before reshock fall around the threshold. The scaling analysis of the turbulent kinetic energy spectra after reshock at the latest time, at which mixing transition analysis suggests that an inertial range has formed, indicates power scaling of -1.8+-0.05 for the low Mach number case and -2.1+-0.1 for the higher Mach number case. This could possibly be related to the high anisotropy observed in this flow resulting from strong, large-scale streamwise fluctuations produced by large-scale shear.
Article
Full-text available
The development of a non-periodic $\text{air}\text{/}\text{SF}_{6}$ gaseous interface subjected to a planar shock wave is investigated experimentally and theoretically to evaluate the effects of the non-periodic portions of the interface on the Richtmyer–Meshkov instability. Experimentally, five kinds of discontinuous chevron-shaped interfaces with or without non-periodic portions are created by the extended soap film technique. The post-shock flows and the interface morphologies are captured by schlieren photography combined with a high-speed video camera. A periodic chevron-shaped interface, which is multi-modal (81 % fundamental mode and 19 % high-order modes), is first considered to evaluate the impulsive linear model and several typical nonlinear models. Then, the non-periodic chevron-shaped interfaces are investigated and the results show that the existence of non-periodic portions significantly changes the balanced position of the initial interface, and subsequently disables the nonlinear model which is applicable to the periodic chevron-shaped interface. A modified nonlinear model is proposed to consider the effects of the non-periodic portions. It turns out that the new model can predict the growth of the shocked non-periodic interface well. Finally, a method is established using spectrum analysis on the initial shape of the interface to separate its bubble structure and spike structure such that the new model can apply to any random perturbed interface. These findings can facilitate the understanding of the evolution of non-periodic interfaces which are more common in reality.