Hostname: page-component-788cddb947-2s2w2 Total loading time: 0 Render date: 2024-10-10T02:30:10.065Z Has data issue: false hasContentIssue false

Hydrodynamic instabilities of two successive slow/fast interfaces induced by a weak shock

Published online by Cambridge University Press:  23 January 2023

Yu Liang
Affiliation:
Advanced Propulsion Laboratory, Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, PR China NYUAD Research Institute, New York University Abu Dhabi, Abu Dhabi 129188, UAE
Xisheng Luo*
Affiliation:
Advanced Propulsion Laboratory, Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, PR China
*
Email address for correspondence: [email protected]

Abstract

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.

Type
JFM Papers
Copyright
© The Author(s), 2023. Published by Cambridge University Press

1. Introduction

Rayleigh–Taylor instability (RTI) develops when lighter fluids accelerate heavier fluids (Rayleigh Reference Rayleigh1883; Taylor Reference Taylor1950), then bubbles (lighter fluids penetrating heavier ones) and spikes (heavier fluids penetrating lighter ones) arise, and a flow transition to turbulent mixing may finally appear (Zhou et al. Reference Zhou, Clark, Clark, Glendinning, Skinner, Huntington, Hurricane, Dimits and Remington2019; Livescu Reference Livescu2020). Richtmyer–Meshkov instability (RMI) is a comparable phenomenon when a shock wave accelerates an interface separating two fluids (Richtmyer Reference Richtmyer1960; Meshkov Reference Meshkov1969). As reviewed by Zhou et al. (Reference Zhou2021), both instabilities play essential roles in various industrial and scientific fields, including inertial confinement fusion (ICF), supernova explosions, ejecta, material strength, chemical reactions, solar prominence and ionospheric flows. The RTI and RMI on a semi-infinite single-mode interface (as the simplest mathematical form) have been widely studied (Sharp Reference Sharp1984; Brouillette Reference Brouillette2002; Zhou Reference Zhou2017a,Reference Zhoub; Zhai et al. Reference Zhai, Zou, Wu and Luo2018b). The RTI and RMI, however, are frequently involved in shock-induced multiple interface evolutions (Liang Reference Liang2022a). For example, the RMI occurs when powerful lasers or X-rays interact with the multiple interfaces of an ICF capsule, determining the seed of the RTI during the ICF implosion. The mixing induced by the RTI and RMI significantly reduces and even eliminates the thermonuclear yield (Miles et al. Reference Miles, Edwards, Blue, Hansen, Robey, Drake, Kuranz and Leibrandt2004; Qiao & Lan Reference Qiao and Lan2021). Furthermore, the shocks produced by star collapse in a supernova interact with multiple heavy elements throughout interstellar space. The mixing induced by the RTI and RMI shapes the filament structures in the remnant of the historical supernova. Therefore, studying the hydrodynamic instabilities of multiple interfaces driven by a shock wave is crucial.

In the ICF, the disturbance at the hotspot–fuel interface originates from its initial perturbation, the feedthrough of the perturbation on the ablation surface and the driven inhomogeneity (Hsing & Hoffman Reference Hsing and Hoffman1997; Weir, Chandler & Goodwin Reference Weir, Chandler and Goodwin1998; Shigemori et al. Reference Shigemori, Azechi, Nakai, Endo, Nagaya and Yamanaka2002; Regan et al. Reference Regan2004; Haan et al. Reference Haan2011; Simakov et al. Reference Simakov, Wilson, Yi, Kline, Clark, Milovich, Salmonson and Batha2014; Knapp et al. Reference Knapp2017; Desjardins et al. Reference Desjardins2019). Milovich et al. (Reference Milovich, Amendt, Marinak and Robey2004) showed that the feedthrough of the ablation surface leads to a significant decrease in the implosion efficiency of a double-shell ignition target. Moreover, simulations indicated that the performance of the ignition target obviously depends on the distance between the successive interfaces, and a longer distance reduces the feedthrough between the interfaces (Haan et al. Reference Haan2011; Simakov et al. Reference Simakov, Wilson, Yi, Kline, Clark, Milovich, Salmonson and Batha2014). Therefore, the feedthrough between two successive interfaces is a significant parameter to evaluate in applications. However, it is still an open problem to exactly quantify the feedthrough effect on the hydrodynamic instabilities of two successive interfaces with arbitrary initial conditions. Therefore, it is essential to study the solution and outcome of the feedthrough effect on the hydrodynamic instabilities.

Theoretically, Taylor (Reference Taylor1950) was the first to consider the RTI of two successive interfaces and discovered that the feedthrough is significant when the initial distance between the two interfaces is limited. Ott (Reference Ott1972) derived a nonlinear solution describing the asymmetry of the RTI of a thin massless layer. The linear models for the RMI of successive fast/slow and slow/fast interfaces were separately derived by Mikaelian (Reference Mikaelian1985Reference Mikaelian1995Reference Mikaelian1996) and Jacobs et al. (Reference Jacobs, Jenkins, Klein and Benjamin1995). It was proved that the feedthrough becomes more evident as the initial distance decreases. Based on the point-vortex model, Jacobs et al. (Reference Jacobs, Jenkins, Klein and Benjamin1995) proposed a nonlinear model to quantify the mixing width growth of the fluid layer consisting of two successive interfaces. Recently, the third-order weakly nonlinear solutions for the RTI and RMI of two superimposed fluid layers in a vacuum were separately deduced by Wang et al. (Reference Wang, Guo, Wu, Ye, Liu, Zhang and He2014) and Liu et al. (Reference Liu, Li, Yu, Fu, Wang, Wang and Ye2018b). The density ratio of the two fluid layers has a non-negligible influence on the instability development of the middle interface. It is evident that most previous theoretical studies considered the hydrodynamic instabilities of A/B/A-type successive interfaces, i.e. two successive interfaces separating two kinds of fluids (fluids A and B). However, compared with A/B/A-type successive interfaces, A/B/C-type successive interfaces, i.e. two successive interfaces separating three kinds of fluids (fluids A, B and C), are more general in applications. Therefore, a general, analytical solution for the hydrodynamic instabilities of two successive interfaces separating three kinds of fluids is urgently needed.

Experimentally, the gas curtain technology combined with a particle image velocimetry or/and planar laser-induced fluorescence system was primarily used to investigate the RMI of a thin SF$_6$ gas curtain surrounded by air. The experiments measured the shocked gas curtain's mixing widths, circulation, mixedness and other parameters with the quantitative measurement technique. It was discovered that the shock-induced SF$_6$ gas curtain morphologies are sensitive to the initial curtain shape (Jacobs et al. Reference Jacobs, Klein, Jenkins and Benjamin1993; Budzinski, Benjamin & Jacobs Reference Budzinski, Benjamin and Jacobs1994; Jacobs et al. Reference Jacobs, Jenkins, Klein and Benjamin1995; Rightley, Vorobieff & Benjamin Reference Rightley, Vorobieff and Benjamin1997). Moreover, the initial spectrum of perturbations, shock strength and reshock also influence the late-time mixing (Prestridge et al. Reference Prestridge, Vorobieff, Rightley and Benjamin2000; Tomkins et al. Reference Tomkins, Kumar, Orlicz and Prestridge2008Reference Tomkins, Balakumar, Orlicz, Prestridge and Ristorcelli2013; Balakumar et al. Reference Balakumar, Orlicz, Tomkins and Prestridge2008; Orlicz et al. Reference Orlicz, Balakumar, Tomkins and Prestridge2009; Orlicz, Balasubramanian & Prestridge Reference Orlicz, Balasubramanian and Prestridge2013; Balakumar et al. Reference Balakumar, Orlicz, Ristorcelli, Balasubramanian, Prestridge and Tomkins2012; Balasubramanian et al. Reference Balasubramanian, Orlicz, Prestridge and Balakumar2012). However, the concentration of the test gas inside the gas curtain is non-uniform (Tomkins et al. Reference Tomkins, Kumar, Orlicz and Prestridge2008; Orlicz et al. Reference Orlicz, Balakumar, Tomkins and Prestridge2009; Balasubramanian et al. Reference Balasubramanian, Orlicz, Prestridge and Balakumar2012), and the gas curtain actually consists of an infinite number of gas interfaces. Therefore, it is difficult to analyse the wave patterns and flow features, especially the evolution of each interface, during the shock–gas-curtain interaction. Moreover, the profile of the gas curtain is mainly cylindrical or elliptical (Bai et al. Reference Bai, Zou, Wang, Liu, Huang, Liu, Li, Tan and Liu2010), which means that the perturbations of the upstream and downstream interfaces are corrugated with opposite phases. The other cases, such as two successive interfaces corrugated with the same phase perturbations, have rarely been studied.

The soap film technique was recently used to create a shape-controllable and layer-thickness-controllable SF$_6$ or helium gas layer surrounded by air (Liang et al. Reference Liang, Liu, Zhai, Si and Wen2020a; Liang & Luo Reference Liang and Luo2021a,Reference Liang and LuobReference Liang and Luo2022b). The soap film interfaces are discontinuous, and the number of gas interfaces is definite. It was determined that the reverberating waves between the two successive interfaces of a gas layer induce various additional interfacial instabilities. For example, the reflected rarefaction waves inside an SF$_6$ gas layer impose the additional RTI (Rayleigh Reference Rayleigh1883; Taylor Reference Taylor1950) on the upstream interface, and the compression waves inside an SF$_6$ gas layer impose the additional RTI or Rayleigh–Taylor stabilisation (RTS) on the downstream interface (Liang & Luo Reference Liang and Luo2021a). The reflected shocks inside a helium gas layer also induce the additional RMI on the two interfaces (Liang & Luo Reference Liang and Luo2022b).

Numerically, several independent simulations were performed to analyse the parameters dominating the instabilities of two successive interfaces. It was found that the initial perturbations of the two interfaces, the number of interfaces, the geometry of the computational domain (i.e. planar, cylindrical and spherical geometries) and reshocks have different and significant influences on the hydrodynamic instabilities of the multiple interfaces (Mikaelian Reference Mikaelian1996Reference Mikaelian2005; de Frahan, Movahed & Johnsen Reference de Frahan, Movahed and Johnsen2015; Li, Samtaney & Wheatley Reference Li, Samtaney and Wheatley2018; Qiao & Lan Reference Qiao and Lan2021; Ouellet et al. Reference Ouellet, Rollin, Durant, Koneru and Balachandar2022).

In the weak-shock limit, Liang & Luo (Reference Liang and Luo2022a) developed a linear model for the RMI of two successive fast/slow interfaces which are A/B/C-type successive interfaces. It was proved that the density ratios of the distributed fluids, the amplitude ratio of the two interfaces and the dimensionless distance between them determine the feedthrough effect on the RMI. However, the linear model only quantifies the feedthrough between the two interfaces owning the same wavenumber and the same or opposite phases. The perturbation on the material interfaces in applications should be random. Therefore, extending the linear model by considering arbitrary wavenumber and phase combinations is essential. Moreover, the reverberating waves between two successive slow/fast interfaces differ significantly from those between two fast/slow interfaces. As a result, the additional instabilities caused by the reverberating waves are completely different. Rarefaction waves are expected to reverberate between the shocked two successive interfaces. On comparing with the shock–interface interaction, it is more challenging to quantify the flow field during and after the interaction of rarefaction waves and interfaces. Furthermore, because a shocked slow/fast interface generally experiences phase reversal (Brouillette Reference Brouillette2002), quantifying the waves’ effect on the two successive slow/fast interfaces is more complicated than on the two fast/slow interfaces. In addition, an abnormal phase reversal phenomenon (i.e. phase reversal occurs on a shocked fast/slow interface) was discovered in the shock-tube experiments on the two successive fast/slow interfaces (Liang & Luo Reference Liang and Luo2022a), which is one kind of the abnormal RMI. Nevertheless, it is uncertain whether the abnormal RMI happens on two successive fast/slow interfaces. The necessary and sufficient conditions of the abnormal RMI also need further investigation.

In this work, we shall first extend the soap film technique to generate two successive slow/fast interfaces with the upstream gas of SF$_6$, the middle gas of a mixture of air and SF$_6$, and the downstream gas of air. Second, we use the shock-tube facility to perform three quasi-one-dimensional (1-D) experiments by varying the initial distance between the two interfaces to understand the reverberating rarefaction waves effect on the two interfaces’ movements. We also conduct six quasi-two-dimensional (2-D) experiments by considering various initial distances and interface perturbations to explore the hydrodynamic instabilities. Numerical simulations are performed to provide more quantitative data. Third, an analytical, linear solution is established considering the arbitrary wavenumber and phase combinations and compressibility in the weak-shock limit. Fourth, the influences of the reverberating rarefaction waves on the hydrodynamic instabilities of the two interfaces are quantified. Last, the conditions and outcomes of the freeze-out and abnormal RMI caused by the feedthrough are summarised according to the extended linear model and numerical results.

2. Experimental and numerical methods

2.1. Experimental setup

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 (Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a; Liang et al. Reference Liang, Zhai, Ding and Luo2019Reference Liang, Liu, Zhai, Ding, Si and Luo2021). As shown in figure 1(a), three transparent devices (i.e. left device, middle device and right device) with a width of 140.0 mm and a height of 10.0 mm are first manufactured using transparent acrylic sheets with a thickness of 3.0 mm. Next, the middle device's adjacent boundaries are carefully engraved to be of a sinusoidal shape with a depth of 1.8 mm. Then, on two sides of the middle device, four thin filaments with a height of 2.0 mm are attached to the inner surfaces of the upper and lower plates to restrict the shape of the soap film (marked by green in figure 1a). Thus the filament bulges in the flow field with only 0.2 mm height. Finally, before the interface formation, the filaments are appropriately wetted with a soap solution containing 78 % (in mass fraction) distilled water, 2 % sodium oleate and 20 % glycerine.

Figure 1. Schematics of (a) the soap film interface generation, (b) the shock-tube and schlieren photography, and (c) the initial configuration of two successive slow/fast interfaces, where $L_0$ denotes the initial distance between the two interfaces, II$_1$ denotes the initial upstream interface and II$_2$ denotes the initial downstream interface.

First, a small rectangular frame is drawn along the sinusoidal filaments on both sides of the middle device, with moderate soap solutions dipped on its edges. The middle device is closed, and two soap film interfaces are created. Second, SF$_6$ is pumped into the closed space through an inflow hole to discharge air through an outflow hole. Third, an oxygen concentration detector is placed at the outflow hole to monitor the concentration of SF$_6$ inside the closed space. The inflow and outflow holes are sealed when the volume fraction of oxygen at the outflow hole reduces to 10 %. Fourth, the left and right transparent devices are gently connected to the middle device, and the combined one is inserted into the test section of the shock-tube. Fifth, the air in the driven section of the shock-tube is entirely replaced by SF$_6$, as sketched in figure 1(b). Finally, high-pressure air is pumped into the driver section of the shock-tube, breaking the diaphragm between the driver section and the driven section and generating a shock wave. The two successive slow/fast interfaces are impacted by the shock wave in the open-ended test section.

In the Cartesian coordinate system, as sketched in figure 1(c), the perturbations on the two interfaces are single-mode. In this work, the perturbation wavenumbers on the two interfaces are fixed as $104.7$ m$^{-1}$, and the initial amplitude of the upstream interface ($a_{1}(0)$) and the downstream interface ($a_{2}(0)$) are fixed as 2.0 mm. To minimise the wall effect of the shock-tube on the interface evolution, a short flat part with 10.0 mm on each side of the two interfaces is adopted (Vandenboomgaerde et al. Reference Vandenboomgaerde, Souffland, Mariani, Biamino, Jourdan and Houas2014). Its influence on interface evolution is limited (Luo et al. Reference Luo, Liang, Si and Zhai2019). Here, $L_0$ is defined as the distance between the average positions of the initial upstream interface (II$_1$) with the initial downstream interface (II$_2$). We shall perform three quasi-1-D experiments and six quasi-2-D experiments. The three cases (i.e. cases L10-IP, L30-IP and L50-IP) in which the initial perturbations on the two interfaces are in-phase are defined as the in-phase cases and the three cases (i.e. cases L10-AP, L30-AP and L50-AP) in which the initial perturbations on the two interfaces are anti-phase are defined as the anti-phase cases. As sketched in figure 1(c), gas A is SF$_6$, gas B is a mixture of air and SF$_6$, and gas C is air. The mass fraction of SF$_6$ in gas B, $MF$, is listed in table 1. Moreover, $R_1$ ($=\rho _{A}/\rho _{B}$) and $R_3$ ($=\rho _{C}/\rho _{B}$) separately denote the density ratios of the fluids on the two sides of the upstream interface and downstream interface, with $\rho _{A}$ ($=6.14$ kg m$^{-3}$), $\rho _{B}$ (as listed in table 1) and $\rho _{C}$ ($=1.20$ kg m$^{-3}$) being the pre-shock densities of gases A, B and C, respectively. In addition, $A_1$ ($=(\rho _{B}-\rho _{A})/(\rho _{B}+\rho _{A})$) and $A_2$ ($=(\rho _{C}-\rho _{B})/(\rho _{C}+\rho _{B})$) are the Atwood numbers of the upstream interface and downstream interface, respectively.

Table 1. Initial physical parameters of two successive slow/fast interfaces, where $L_0$ denotes the initial distance between the two interfaces; $MF$ denotes the mass fraction of SF$_6$ in gas B; $\rho _{B}$ denotes the density of gas B; $R_1$ and $R_3$ denote the density ratios of the fluids on the two sides of the upstream interface and downstream interface, respectively; and $A_1$ and $A_2$ denote the Atwood numbers of the upstream interface and downstream interface, respectively.

The ambient pressure and temperature are 101.3 kPa and $295.5\pm 1.0$ K, respectively. In the experiments, the incident shock wave travels from left to right with a Mach number of $1.24\pm 0.01$ and a velocity ($v_{{IS}}$) of $167\pm 2$ m s$^{-1}$. The velocity of the flow behind the incident shock wave ($v_{{ps}}$) is $56\pm 1$ m s$^{-1}$. We choose schlieren photography combined with a high-speed camera to monitor the flow field and capture the density gradient induced by the reverberating waves between the two interfaces. The frame rate of the high-speed video camera (FASTCAM SA5, Photron Limited) is 60 000 f.p.s., and the shutter time is 1.0 $\mathrm {\mu }$s. The spatial resolution of schlieren images is 0.4 mm pixel$^{-1}$. The flow field visualisation is limited within the range of $x\in [-50,50]\,\mathrm {mm}$, as shown in figure 1(c).

2.2. Numerical scheme

Numerical simulation is performed to obtain quantitative data considering more initial conditions. The process of a planar shock interacting with two successive slow/fast interfaces examined in this study is described by compressible Euler equations, which coincides with the numerical studies focusing on the early to intermediate regimes of RMI with or without reshocks (Grove et al. Reference Grove, Holmes, Sharp, Yang and Zhang1993; Holmes & Grove Reference Holmes and Grove1995; Holmes et al. Reference Holmes, Dimonte, Fryxell, Gittings, Grove, Schneider, Sharp, Velikovich, Weaver and Zhang1999; Herrmann, Moin & Abarzhi Reference Herrmann, Moin and Abarzhi2008; Niederhaus et al. Reference Niederhaus, Greenough, Oakley, Ranjan, Anderson and Bonazza2008; Leinov et al. Reference Leinov2009; Ding et al. Reference Ding, Si, Chen, Zhai, Lu and Luo2017Reference Ding, Liang, Chen, Zhai, Si and Luo2018; Zhai et al. Reference Zhai, Li, Si, Luo, Yang and Lu2017Reference Zhai, Liang, Liu, Ding, Luo and Zou2018a; Zou et al. Reference Zou, Al-Marouf, Cheng, Samtaney, Ding and Luo2019; Igra & Igra Reference Igra and Igra2020). An upwind space–time conservation elements/solution elements (CE/SE) scheme is used with second-order accuracy in both space and time (Shen et al. Reference Shen, Wen, Liu and Zhang2015a; Shen, Wen & Zhang Reference Shen, Wen and Zhang2015b; Shen & Wen Reference Shen and Wen2016). A volume-fraction-based five-equation model (Abgrall Reference Abgrall1996; Shyue Reference Shyue1998) is used to illustrate the different species residing on both sides of the inhomogeneous interface. The contact discontinuity restoring Harten–Lax–van Leer contact Riemann solver (Toro, Spruce & Speares Reference Toro, Spruce and Speares1994) is used to determine the numerical fluxes between the conservation elements. The use of this scheme in capturing shocks and details of complex flow structures for the RMI issues and shock–droplet interactions has been well validated (Shen & Parsani Reference Shen and Parsani2017; Shen et al. Reference Shen, Wen, Parsani and Shu2017; Guan et al. Reference Guan, Liu, Wen and Shen2018; Fan et al. Reference Fan, Guan, Wen and Shen2019; Liang et al. Reference Liang, Zhai, Luo and Wen2020b; Liang Reference Liang2022b). A comprehensive review of the scheme and its extensive applications was recently reported by Jiang, Wen & Zhang (Reference Jiang, Wen and Zhang2020). The initial settings of the 2-D simulation are presented in figure 2. Open boundary conditions are enforced on the left and right boundaries ($y=-40.0$ and $y=200.0$ mm) to eliminate the waves reflected from the left and right boundaries (Liang et al. Reference Liang, Zhai, Luo and Wen2020b; Liang Reference Liang2022b), and reflection conditions are imposed at the top and bottom boundaries ($x=-70.0$ and $x=70.0$ mm), respectively.

Figure 2. Schematics of the initial simulation setup, where $a_1$ and $a_2$ represent the amplitudes of the upstream interface and downstream interface, respectively.

2.3. Code validation

The experimental results of cases L10-1D and L10-AP are used for the code validation. For the data of numerical simulations, the nodes with a mass fraction of SF$_6$ between 0.99 with $(MF+0.01)$ are chosen as the upstream interface, and the nodes with a mass fraction of SF$_6$ between $(MF-0.01)$ with 0.01 are chosen as the downstream interface. Then, the mean value of $y$ of these nodes on each row is taken as the average position of the local interface.

First, the time-varying displacements of the shocked upstream interface (SI$_1$), $y_{{SI1}}$, and the shocked downstream interface (SI$_2$), $y_{{SI2}}$, in the L10-1D case, are extracted from the experiments, as shown with solid symbols for the upstream interface and hollow symbols for the downstream interface in figure 3(a). The moment when the incident shock wave impacts the average position of the II$_1$ at $y_{01}$ is defined as $t=0$. The numerical results with four mesh sizes of 0.40 mm, 0.20 mm, 0.10 mm and 0.05 mm are compared with the experiments, as shown with lines in figure 3(a). It is found that all the numerical results agree well with the experimental ones.

Figure 3. Code validation based on (a) the time-varying displacements of the shocked upstream interface (SI$_1$) and downstream interface (SI$_2$) in the L10-1D case, and (b) the time-varying amplitudes of the SI$_1$ and SI$_2$ in the L10-AP case, where $n=1$ for the upstream interface and $n=2$ for the downstream interface in this study. Solid (or hollow) symbols represent the experimental results for the SI$_1$ (or SI$_2$), and dash–dot (or dashed) lines represent the numerical results for the SI$_1$ (or SI$_2$), considering various mesh sizes.

Second, the time-varying amplitudes of the upstream interface ($a_1$) and downstream interface ($a_2$) are extracted from the experiments in the L10-AP case, as shown in figure 3(b). Here, $a_1$ and $a_2$ are defined as half of the streamwise distances between the bubble tip and spike tip of the upstream interface and downstream interface, respectively, as sketched in figure 2. It is found that the numerical results with mesh sizes of 0.20 mm, 0.10 mm and 0.05 mm quantitatively agree well with the experiments within the experimental measurement errors. Moreover, the values of $a_1$ and $a_2$, acquired from the simulations, separately converge when the mesh size is reduced to 0.10 mm and 0.05 mm in the numerical simulations. Therefore, an initial mesh size of 0.10 mm is adopted for all simulations to ensure accuracy and minimise the computational cost.

3. Quasi-1-D wave pattern and interface movement

We shall first discuss the reverberating waves observed from schlieren images. The interface displacements and velocities are then measured and compared between various initial distance cases. Last, the velocities of the two interfaces are calculated after the reverberating rarefaction waves have impacted them. A general 1-D theory for characterising the interface movement is established in this way.

3.1. Experimental observation

Experimental schlieren images of the evolutions of two successive, quasi-1-D slow/fast interfaces driven by a shock wave are shown in figure 4(ac) for $L_0=10.0$, 30.0 and 50.0 mm, respectively. Taking the L50-1D case as an example, the wave pattern and interface movement are discussed in detail. After the incident shock wave (IS) impacts the II$_1$, the reflected rarefaction waves (RW$_1$) and the transmitted shock (TS$_1$) are immediately generated, and the shocked upstream interface (SI$_1$) begins to move forwards (116 $\mathrm {\mu }$s). Then the TS$_1$ impacts the II$_2$, and the TS$_2$ is refracted downstream, followed by the shocked downstream interface (SI$_2$) (299 $\mathrm {\mu }$s). Meanwhile, the rarefaction waves (rRW$_2$) are reflected upstream since the II$_2$ is a slow/fast interface relative to the movement of the TS$_1$. After the rRW$_2$ impacts the SI$_1$ (449 $\mathrm {\mu }$s), the transmitted rarefaction waves (tRW$_1$) and reflected rarefaction waves (rRW$_1$) are immediately generated since the SI$_1$ is a fast/slow interface relative to the movement of the rRW$_2$. However, due to the limited strength of the rRW$_1$, the density variation induced by the rRW$_1$ is restricted such that it is challenging to distinguish the rRW$_1$ between the two interfaces. Finally, all waves are refracted away from the two interfaces, and the two interfaces move forward at the same velocity.

Figure 4. Schlieren images of evolutions of two successive, quasi-1-D slow/fast interfaces driven by a shock wave in cases (a) L10-1D, (b) L30-1D and (c) L50-1D, where SI$_1$ (or SI$_2$) denotes the shocked upstream (or downstream) interface, RW$_1$ (or TS$_1$) denotes the reflected rarefaction waves (or transmitted shock) after the IS impacts II$_1$, rRW$_2$ (or TS$_2$) denotes the reflected rarefaction waves (or transmitted shock) after the TS$_1$ impacts II$_2$, and tRW$_1$ denotes the transmitted rarefaction waves after the rRW$_2$ impacts the SI$_1$. Numbers indicate time in $\mathrm {\mu }$s.

The interface displacements ($y_{{SIn}}$) and velocities ($v_{{SIn}}$) of the two interfaces, with $n=1$ for the upstream interface and $n=2$ for the downstream interface, are measured from experiments, as shown in figures 5(a) and 5(b), respectively. Time is scaled as $tv_{{t1}}/L_0$ with $v_{{t1}}$ denoting the velocity of the TS$_1$. Interface displacement is scaled as $(y_{{SIn}}-y_{01})/L_0$, and interface velocity is scaled as $v_{{SIn}}/v_{1}^\alpha$ with $v_{1}^\alpha$ denoting the velocity jump of the upstream interface induced by the IS. The values of $v_{{t1}}$ and $v_{1}^\alpha$ in all cases are calculated according to the 1-D gas dynamics theory (Drake Reference Drake2018), as listed in table 2. The dimensionless displacements and velocities of the upstream (or downstream) interface converge in all $L_0$ cases, as shown in figure 5(a,b), indicating that the dimensionless movements of the two interfaces are independent of the initial distance between the two interfaces.

Figure 5. The dimensionless (a) displacements and (b) velocities of the upstream interface (solid symbols) and the downstream interface (hollow symbols). Black, yellow and orange solid (or dashed) lines represent the 1-D theory predictions for the upstream (or downstream) interface movements in Stages I, II and III, respectively. The vertical dash–dot lines represent the specific times calculated with (3.1) and (3.9a,b).

Table 2. Physical parameters of the waves and interfaces in the laboratory reference, where $v_{{t1}}$ denotes the velocity of the TS$_1$; $v_{{rRW}_2}$ (or $v_{{rRW_1}}$) denotes the velocity of the rRW$_2$ head (or rRW$_1$ head); $v_{1}^\alpha$ (or $v_{2}^\alpha$) denotes the velocity of the SI$_1$ (or SI$_2$) in Stage I; $v_{1}^\beta$ (or $v_{2}^\beta$) denotes the velocity of the SI$_1$ (or SI$_2$) in Stage II; $v_{fn}$ denotes the asymptotic velocity of the two interfaces; $t_{2}^\alpha$ denotes the specific time when the TS$_1$ impacts the II$_2$; $t_{1}^\beta$ denotes the time when the rRW$_2$ head impacts the SI$_1$ and $t_{2}^\beta$ denotes the time when the rRW$_1$ impacts the SI$_2$; $t_1^\sigma$ (or $t_2^\sigma$) represents the time when the rRW$_2$ (or rRW$_1$) leaves the SI$_1$ (or SI$_2$). The units for velocity and time are m s$^{-1}$ and $\mathrm {\mu }$s, respectively.

3.2. Characteristics of reverberating rarefaction waves

The reverberating rarefaction waves between the two successive slow/fast interfaces are sketched in figure 6. Here, we define the time when the TS$_1$ impacts the II$_2$ as $t_2^\alpha$, the time when the rRW$_2$ impacts the SI$_1$ as $t_1^\beta$, and the time when the rRW$_1$ impacts the SI$_2$ as $t_2^\beta$. Then $t_2^\alpha$, $t_1^\beta$ and $t_2^\beta$ can be derived as

(3.1)\begin{equation} \left. \begin{aligned} t_2^\alpha & =\frac{L_0}{v_{{t1}}},\\ t_1^\beta & =t_2^\alpha+\frac{L_0-t_2^\alpha v_1^\alpha}{v_1^\alpha-v_{{rRW}_2}},\\ t_2^\beta & =t_1^\beta+\frac{L_0-t_2^\alpha v_1^\alpha+(t_1^\beta-t_2^\alpha)(v_2^\alpha-v_1^\alpha)}{v_{{rRW_1}}-v_2^\alpha}, \end{aligned} \right\} \end{equation}

in which $v_2^{\alpha }$ represents the velocity jump of the downstream interface induced by the TS$_1$; $v_{{rRW}_2}$ and $v_{{rRW_1}}$ separately represent the velocity of the rRW$_2$ head and that of the rRW$_1$ head, and they can be stated as

(3.2a,b)\begin{equation} v_{{rRW}_2}=v_1^{\alpha}-c_B^{\alpha},\quad v_{{rRW}_1}=v_2^{\alpha}+c_B^{\beta}, \end{equation}

where $c_B^{\alpha }$ denotes the sound speed of gas B between the TS$_1$ and SI$_1$, and $c_B^{\beta }$ denotes the sound speed of gas B between the rRW$_2$ tail and SI$_2$, as sketched in figure 6(a,b). The values of $v_2^{\alpha }$, $v_{{rRW}_2}$ and $v_{{rRW}_1}$ are calculated according to the 1-D gas dynamics theory (Drake Reference Drake2018) and (3.2a,b), as listed in table 2. Then $t_2^\alpha$, $t_1^\beta$ and $t_2^\beta$ are calculated based on (3.1), as listed in table 2. As $L_0$ increases, $t_2^\alpha$, $t_1^\beta$ and $t_2^\beta$ increase.

Figure 6. Sketches of (a) the interaction of the TS$_1$ and II$_2$, (b) the interaction of the rRW$_2$ and SI$_1$, (c) the interaction of the rRW$_1$ and SI$_2$, and (d) the movements of the SI$_1$ and SI$_2$ after all waves are refracted away, where $\gamma _A$, $\gamma _B$ and $\gamma _C$ denote the specific heat ratios of gas A, gas B and gas C, respectively; $c_A$ denotes the sound speed of gas A after the IS impacts the II$_1$, $c_B^\alpha$ denotes the sound speed of gas B between the TS$_1$ with the II$_2$, $c_B^\beta$ denotes the sound speed of gas B between the rRW$_2$ tail and the SI$_2$, $c_B^\sigma$ denotes the sound speed of gas B between the rRW$_1$ tail and the SI$_1$, $c_C$ denotes the sound speed of gas C behind TS$_2$; $p_1$ denotes the pressure of gas B behind the TS$_1$, $p_2$ denotes the pressure of gas B behind the rRW$_2$ tail, $p_3$ denotes the pressure of gas B behind the rRW$_1$ tail, and $p_4$ denotes the pressure of gas B after the interaction of the rRW$_1$ and SI$_2$. (a) $t=t_2^\alpha$, (b) $t=t_1^\beta$, (c) $t=t_2^\beta$, (d) $t>t_2^\beta$.

After the rRW$_2$ impacts the SI$_1$, the velocity of the SI$_1$ increases from $v_1^\alpha$ to $v_1^\beta$ since the pressure behind the rRW$_2$ tail ($p_2$) is lower than the pressure in front of the rRW$_2$ head ($p_1$), as sketched in figure 6(a,b). Differently, after the rRW$_1$ impacts the SI$_2$, the velocity of the SI$_2$ reduces from $v_2^\alpha$ to $v_2^\beta$ since the pressure behind the rRW$_1$ tail ($p_3$) is lower than $p_2$, as sketched in figure 6(b,c). The solutions of $v_1^\beta$ and $v_2^\beta$ are discussed in detail.

First, based on the 1-D gas dynamics theory for the interaction of rarefaction waves and a fast/slow interface (Drake Reference Drake2018; Liang & Luo Reference Liang and Luo2022a), $v_1^{\beta }$ can be solved as

(3.3)\begin{equation} v_1^{\beta}=v_1^{\alpha}-\frac{2c_A}{\gamma_A-1}\left[\,p_{31}^{(\gamma_A-1)/2\gamma_A}-1\right], \end{equation}

where $\gamma _A$ is the specific heat ratio of gas A, and $c_A$ is the sound speed of gas A after the IS impacts the II$_1$, as sketched in figure 6(a). The only unknown parameter $\,p_{31}$, i.e. the ratio of the $p_3$ to $p_1$, satisfies

(3.4)\begin{equation} \frac{c_A}{c_B^{\alpha}}=\frac{(\gamma_A-1)\left[\,p_{31}^{(\gamma_B-1)/2\gamma_B}-2\,p_{21}^{(\gamma_B-1)/2\gamma_B}+1\right]}{(\gamma_B-1)\left[1-\,p_{31}^{(\gamma_A-1)/2\gamma_A}\right]}, \end{equation}

where $\gamma _B$ is the specific heat ratio of gas B, $c_B^\alpha$ is the sound speed of gas B between the TS$_1$ with the II$_2$, and $\,p_{21}$ is the ratio of $p_2$ to $p_1$. Then $v_1^\beta$ in all cases can be calculated by solving (3.3)–(3.4), as listed in table 2.

Second, based on the 1-D gas dynamics theory for the interaction of rarefaction waves and a slow/fast interface (Drake Reference Drake2018; Liang & Luo Reference Liang and Luo2021a), $v_2^{\beta }$ can be calculated as

(3.5)\begin{equation} v_{2}^\beta=v_{2}^\alpha+\frac{2c_B^\beta}{\gamma_B-1}\left[(\,p_{32})^{(\gamma_B-1)/2\gamma_B}-1\right]-\frac{\sqrt{2}c_B^\beta(\,p_{42}/p_{32}-1)}{\sqrt{\gamma_B(\gamma_B+1)\,p_{42}/p_{32}}}, \end{equation}

where $c_B^\beta$ is the sound speed of gas B between the rRW$_2$ tail and the SI$_2$, and $\,p_{32}$ is the ratio of $p_3$ to $p_2$. The only unknown parameter $\,p_{42}$, i.e. the ratio of the pressure of the flow after the rRW$_1$ leaves the SI$_2$ ($p_4$) to $p_2$, satisfies

(3.6)\begin{equation} \frac{c_C}{c_B^\beta}=\frac{\gamma_C-1}{\,p_{42}^{(\gamma_C-1)/2\gamma_C}-1} \left\{\frac{2}{\gamma_B-1}\left[\,p_{32}^{(\gamma_B-1)/2\gamma_B}-1\right] -\frac{\sqrt{2}(\,p_{42}/p_{32}-1)\,p_{32}^{(\gamma_B-1)/2\gamma_B}} {\sqrt{(\gamma_B+1)\gamma_B\,p_{42}/p_{32}+(\gamma_B-1)\gamma_B}}\right\}, \end{equation}

where $c_C$ is the sound speed of gas C after the TS$_1$ impacts the II$_2$, and $\gamma _C$ is the specific heat ratio of gas C. Then $v_{2}^\beta$ in all cases can be calculated by solving (3.5)–(3.6), as listed in table 2.

We note that the number of reverberating waves between the two interfaces should be infinite. However, the consecutive pressure and velocity changes imparted by these reverberations decrease, eventually converging to a well-defined asymptotic 1-D post-shock state. The influences of the reverberating waves on the 1-D and 2-D dynamics of the two interfaces should be considered until the two interfaces enter the asymptotic post-shock state (Aglitskiy et al. Reference Aglitskiy2006; Liang & Luo Reference Liang and Luo2022a). A simple and advantageous method is adopted to judge whether the higher-order reverberations should be accounted for or not. The interaction of an incident shock wave with a semi-infinite interface separating two fluids with $\rho _A$ and $\rho _C$ is considered, ignoring the intermediate layer of gas B. The post-shock final velocity, $v_{{fn}}$, is calculated according to the 1-D gas dynamics theory (Drake Reference Drake2018), as listed in table 2. The differences among the $v_1^\beta$, $v_2^\beta$ and $v_{{fn}}$ are limited: $v_{{fn}}$ is only 0.2 % smaller than $v_1^\beta$ and 0.1 % larger than $v_2^\beta$. Therefore, it is reasonable to regard that after the rRW$_1$ impacts the SI$_2$, the movements of the two interfaces enter the asymptotic state.

The interaction between rarefaction waves and an interface is a continuous process (Li & Book Reference Li and Book1991; Li, Kailasanath & Book Reference Li, Kailasanath and Book1991; Morgan, Likhachev & Jacobs Reference Morgan, Likhachev and Jacobs2016; Morgan et al. Reference Morgan, Cabot, Greenough and Jacobs2018; Liang et al. Reference Liang, Zhai, Luo and Wen2020b; Wang et al. Reference Wang, Song, Ma, Ma, Wang and Wang2022). The length of the rRW$_2$ at $t_1^\beta$ ($L_{rRW_2}$), the interaction time of the rRW$_2$ with the SI$_1$ ($T_{{rRW_2}}$), and the average acceleration imposed on the SI$_1$ by the rRW$_2$ ($\bar {g}_{{rRW_2}}$) are respectively deduced as

(3.7)\begin{equation} \left. \begin{aligned} L_{rRW_2} & =(c_B^\alpha+v_2^\alpha-c_B^\beta-v_1^\alpha)(t_1^\beta-t_2^\alpha),\\ T_{{rRW_2}} & =\frac{2L_{rRW_2}}{2(c_B^\beta-v_2^\alpha)+v_1^\alpha+v_1^\beta},\\ \bar{g}_{{rRW_2}} & =\frac{v_1^\beta-v_1^\alpha}{T_{{rRW_2}}}. \end{aligned} \right\} \end{equation}

Similarly, the length of the rRW$_1$ at $t_2^\beta$ ($L_{rRW_1}$), the interaction time of the rRW$_1$ with the SI$_2$ ($T_{{rRW_1}}$), and the average deceleration imposed on the SI$_2$ by the rRW$_1$ ($\bar {g}_{{rRW_1}}$) are respectively derived as

(3.8)\begin{equation} \left. \begin{aligned} L_{rRW_1} & =(c_B^\beta+v_2^\alpha-c_B^\sigma-v_1^\beta)(t_2^\beta-t_1^\beta),\\ T_{{rRW_1}} & =\frac{2L_{rRW_1}}{2(c_B^\sigma+v_1^\beta)-v_2^\alpha-v_2^\beta},\\ \bar{g}_{{rRW_1}} & =\frac{v_2^\beta-v_2^\alpha}{T_{{rRW_1}}}, \end{aligned} \right\} \end{equation}

in which $c_B^\sigma$ ($=c_B^\beta -(\gamma _B-1)(u_2^\alpha -u_1^\beta )/2$) is the sound speed of gas B behind the rRW$_1$, as sketched in figure 6(c). The values of $L_{rRW_2}$, $T_{{rRW_2}}$ and $\bar {g}_{{rRW_2}}$ in all cases are calculated based on (3.7), and the values of $L_{rRW_1}$, $T_{{rRW_1}}$ and $\bar {g}_{{rRW_1}}$ in all cases are calculated according to (3.8), as listed in table 3. As $L_0$ increases, $L_{rRW_2}$, $L_{rRW_1}$, $T_{{rRW_2}}$ and $T_{{rRW_1}}$ increase, whereas both $\bar {g}_{{rRW_2}}$ and $|\bar {g}_{{rRW_1}}|$ decrease.

Table 3. Physical parameters of the reverberating rarefaction waves, where $L_{rRW_2}$ denotes the length of the rRW$_2$ at $t_1^\beta$, $T_{{rRW_2}}$ denotes the interaction time of the rRW$_2$ and the SI$_1$, $\bar {g}_{{rRW_2}}$ denotes the average acceleration imposed on the SI$_1$ by the rRW$_2$, $L_{rRW_1}$ denotes the length of the rRW$_1$ at $t_2^\beta$, $T_{{rRW_1}}$ denotes the interaction time of the rRW$_1$ and the SI$_2$, and $\bar {g}_{{rRW_1}}$ denotes the average deceleration imposed on the SI$_2$ by the rRW$_1$. The units for length, time and acceleration are mm, $\mathrm {\mu }$s and $10^6\ \textrm {m}\ \textrm {s}^{-2}$, respectively.

In summary, the shocked two successive slow/fast interfaces’ movements can be separated into Stages I, II and III, as listed in table 4, where $t_1^\sigma$ (or $t_2^\sigma$) represents the time when the rRW$_2$ (or rRW$_1$) leaves the SI$_1$ (or SI$_2$),

(3.9a,b)\begin{equation} t_1^\sigma=t_1^\beta+T_{{rRW_2}},\quad t_2^\sigma=t_2^\beta+T_{{rRW_1}}. \end{equation}

As a result, a general 1-D theory is established to describe the movements of the two interfaces based on the derived specific times ((3.1) and (3.9a,b)) and interface velocities ((3.2a,b)–(3.8)). The predictions of the 1-D theory for the interface movements in Stages I, II and III are marked with black, yellow and orange lines, respectively, as shown with solid lines for the upstream interface and dashed lines for the downstream interface in figure 5(a,b), and they agree well with the experimental results. Significantly, the specific times when the reverberating waves impact the two interfaces (e.g. $t_1^\beta$, $t_1^\sigma$, $t_2^\beta$, and $t_2^\sigma$) and the velocity jumps at these moments predicted by the 1-D gas dynamic theory agree well with the experiments.

Table 4. The movements of the shocked two successive slow/fast interfaces in different stages, where ‘Uni.’ is short for ‘Uniform movement’, ‘Acc.’ is short for ‘Acceleration movement’ and ‘Dec.’ is short for ‘Deceleration movement’.

4. Quasi-2-D hydrodynamic instabilities

We shall first examine the deformations of the two interfaces observed from experimental and numerical schlieren images. The amplitude growth rates are then acquired from the experiments and simulations. Next, an analytical, linear model is further extended by considering the arbitrary wavenumber and phase combinations and compressibility, which successfully describes the RMI of the two successive slow/fast interfaces. Later, the influences of the reverberating rarefaction waves on the hydrodynamic instabilities of the two interfaces are quantified. Last, the conditions and outcomes of the freeze-out and abnormal RMI are summarised.

4.1. Experimental and numerical results

Schlieren images of the shock-induced two successive slow/fast interface evolutions acquired from experiments and simulations are shown in figure 7(af). The magnitude of the density gradient ($\boldsymbol {\nabla }\rho$) field in the simulations is calculated as (Quirk & Karni Reference Quirk and Karni1996; Sembian, Liverts & Apazidis Reference Sembian, Liverts and Apazidis2018)

(4.1)\begin{equation} |\boldsymbol{\nabla}\rho|=\left[\left(\frac{\partial\rho}{\partial x}\right)^2+\left(\frac{\partial\rho}{\partial y}\right)^2\right]^{{1}/{2}}. \end{equation}

Taking the L50-IP case as an example, the deformations of the two interfaces are discussed in detail. After the IS impacts the perturbed II$_1$, the rippled RW$_1$ is reflected upstream, and the rippled TS$_1$ moves towards the II$_2$ (121 $\mathrm {\mu }$s). Meanwhile, the perturbation on the SI$_1$ decreases due to phase reversal (Brouillette Reference Brouillette2002). After the TS$_1$ impacts the perturbed II$_2$, the rippled TS$_2$ is refracted downstream, and the rippled rRW$_2$ is reflected towards the SI$_1$ (304 $\mathrm {\mu }$s). Meanwhile, the perturbation on the SI$_2$ decreases due to phase reversal (Brouillette Reference Brouillette2002). After the rRW$_2$ impacts the SI$_1$, the tRW$_1$ is refracted upstream (454 $\mathrm {\mu }$s) and, theoretically, the rRW$_1$ is reflected towards the SI$_2$. Finally, the perturbations on the two interfaces increase gradually with phases opposite to their initial perturbations.

Figure 7. Schlieren images of evolutions of quasi-2-D two successive slow/fast interfaces driven by a shock wave acquired from experiments (upper) and simulations (below) in cases (a) L10-IP, (b) L10-AP, (c) L30-IP, (d) L30-AP, (e) L50-IP and (f) L50-AP. The upstream interface in panel (b) is highlighted with a white dashed line. Although the upstream interface in panel (b) is a slow/fast one, it does not experience phase reversal. Numbers denote time in $\mathrm {\mu }$s.

In addition, we observe that the SI$_1$ in the L10-AP case does not experience phase reversal (see the white dashed lines in figure 7b), which is different from the other cases and the RMI of a semi-infinite slow/fast interface reported before (Jourdan & Houas Reference Jourdan and Houas2005; Mariani et al. Reference Mariani, Vandenboomgaerde, Jourdan, Souffland and Houas2008). The phenomena, including the perturbation of a shocked slow/fast interface growing with the same phase as its initial state and the perturbation of a shocked fast/slow interface growing with the opposite phase as its initial state, are called ‘abnormal RMI’. This study is the first to provide observational evidence of the abnormal RMI.

The amplitude growth rate of the upstream interface obtained from experiments ($\dot {a}_1^{{exp}}$) and simulations ($\dot {a}_1^{{num}}$), and that of the downstream interface obtained from experiments ($\dot {a}_2^{{exp}}$) and simulations ($\dot {a}_2^{{num}}$) in Stage I are acquired by linearly fitting the quantitative data, as listed in table 5. First, it is found that $\dot {a}_1^{{num}}$ and $\dot {a}_2^{{num}}$ separately agree well with $\dot {a}_1^{{exp}}$ and $\dot {a}_2^{{exp}}$ within the experimental measurement errors, further validating the code adopted in this study. Second, in the three in-phase cases, as $L_0$ increases, both $|\dot {a}_1^{{exp}}|$ and $|\dot {a}_2^{{exp}}|$ decrease. However, in the three anti-phase cases, as $L_0$ increases, both $|\dot {a}_1^{{exp}}|$ and $|\dot {a}_2^{{exp}}|$ increase. The results indicate that the RMI of two successive interfaces is influenced by the initial distance and perturbations of the two interfaces, i.e. the feedthrough between the two interfaces. Third, due to the feedthrough effect, the sign of $\dot {a}_1^{{exp}}$ is the same as the sign of $a_1(0)$ in the L10-AP case, which goes against the classical RMI where the sign of the amplitude growth rate of a semi-infinite slow/fast interface is opposite to the sign of its initial perturbation (Jourdan & Houas Reference Jourdan and Houas2005; Mariani et al. Reference Mariani, Vandenboomgaerde, Jourdan, Souffland and Houas2008). As we mentioned above, this is called abnormal RMI. Unlike the present results, an opposite amplitude growth rate to the initial perturbation amplitude of a fast/slow interface was found by Liang & Luo (Reference Liang and Luo2022a). Therefore, the outcomes of the abnormal RMI are dependent on the fluid distribution.

Table 5. The linear amplitude growth rates of the two interfaces, where $\dot {a}_1^{{exp}}$, $\dot {a}_1^{{num}}$ and $\dot {a}_1^{{lin}}$ denote the experimental, numerical and theoretical results of the upstream interface, respectively; $\dot {a}_2^{{exp}}$, $\dot {a}_2^{{num}}$ and $\dot {a}_2^{{lin}}$ denote the experimental, numerical and theoretical results of the downstream interface, respectively; and $\psi _1$ and $\psi _2$ denote the feedthrough effect on the RMI of the upstream and downstream interfaces, respectively.

4.2. The feedthrough effect on the RMI

Because the influences of the feedthrough and reverberating rarefaction waves on the hydrodynamic instabilities of the two interfaces are coupled, it is essential to quantify the two effects separately and analytically. Based on the linear stability analysis, Liang & Luo (Reference Liang and Luo2022a) deduced a linear model for evaluating the amplitude growth rates of the upstream interface ($\dot {a}_1^{{lin}}$) and downstream interface ($\dot {a}_2^{{lin}}$) owning the same wavenumber ($k$) as

(4.2)\begin{equation} \left. \begin{aligned} & \dot{a}_1^{{lin}}=\frac{kv_1^{\alpha} \left[Z_1a_1(0)(1-R_1)(2R_3\xi+\xi^2+1)+Z_2a_2(0)(R_3-1)(1-\xi^2)\right]}{2\xi(R_1R_3+1)+(R_1+R_3)(\xi^2+1)},\\ & \dot{a}_2^{{lin}}=\frac{kv_2^{\alpha} \left[Z_1a_1(0)(R_1-1)(\xi^2-1)+Z_2a_2(0)(R_3-1)(2R_1\xi+\xi^2+1)\right]}{2\xi(R_1R_3+1)+(R_1+R_3)(\xi^2+1)}, \end{aligned} \right\} \end{equation}

in which three compression factors $Z_1$ ($=1-v_1^\alpha /v_{s}$), $Z_2$ ($=1-v_2^\alpha /v_{{t1}}$) and $Z_L$ ($=1-v_1^\alpha /v_{{t1}}$) are introduced; $\xi =\tanh (kh_0)$ and $h_0=Z_LL_0/2$ are included. However, this model can only quantify the RMI of two successive fast/slow interfaces with the same wavenumber and initially in-phase or anti-phase perturbations. In this work, we extend the linear model by considering the arbitrary wavenumber and phase combinations and compressibility.

We set the single-mode perturbation on the initial upstream interface as $\eta _1(0)=a_1(0)\cos (k_1x)$ and the single-mode perturbation on the initial downstream interface as $\eta _2(0)=a_2(0)\cos (k_2x+\theta )$, where $\theta$ represents the phase difference between the two interfaces’ perturbations. If there is no perturbation on the downstream interface, then the amplitude growth rates of the two interfaces at $x=0$ are

(4.3)\begin{equation} \left. \begin{aligned} & \dot{a}_1^{{lin}}(a_1(0),0)=\frac{k_1v_1^{\alpha}Z_1a_1(0)(1-R_1)(2R_3\xi+\xi^2+1)}{2\xi(R_1R_3+1)+(R_1+R_3)(\xi^2+1)},\\ & \dot{a}_2^{{lin}}(a_1(0),0)=\frac{k_1v_2^{\alpha}Z_1a_1(0)(R_1-1)(\xi^2-1)\cos(\theta)}{2\xi(R_1R_3+1)+(R_1+R_3)(\xi^2+1)}. \end{aligned} \right\} \end{equation}

If there is no perturbation on the upstream interface, then the amplitude growth rates of the two interfaces at $x=-\theta /k_2$ are

(4.4)\begin{equation} \left. \begin{aligned} & \dot{a}_1^{{lin}}(0,a_2(0))=\frac{k_2v_1^{\alpha}Z_2a_2(0)(R_3-1)(1-\xi^2)\cos(\theta)}{2\xi(R_1R_3+1)+(R_1+R_3)(\xi^2+1)},\\ & \dot{a}_2^{{lin}}(0,a_2(0))=\frac{k_2v_2^{\alpha}Z_2a_2(0)(R_3-1)(2R_1\xi+\xi^2+1)}{2\xi(R_1R_3+1)+(R_1+R_3)(\xi^2+1)}. \end{aligned} \right\} \end{equation}

Based on the linear superposition principle, if there are perturbations imposed on the two initial interfaces simultaneously, then the amplitude growth rates of the two interfaces can be derived by superposing (4.3) and (4.4) as

(4.5)\begin{equation} \dot{a}_1^{{lin}}=\psi_1\dot{a}_1^{SI},~\dot{a}_2^{{lin}}=\psi_2\dot{a}_2^{SI}, \end{equation}

in which $\dot {a}_1^{SI}$ and $\dot {a}_2^{SI}$ denote the linear amplitude growth rates of the upstream interface and downstream interface when they are semi-infinite, respectively; and $\psi _1$ and $\psi _2$ quantify the feedthrough effect on the RMI of the upstream and downstream interfaces, respectively,

(4.6)\begin{equation} \left. \begin{aligned} & \psi_1 = \frac{(R_1 + 1)(2R_3\xi + \xi^2 + 1) - (R_1 + 1)(R_3 - 1)(1 - \xi^2)\cos(\theta)/[\kappa(R_1 - 1)]}{2\xi(R_1R_3 + 1) + (R_1 + R_3)(\xi^2 + 1)},\\ & \psi_2 = \frac{\kappa(R_1 - 1)(R_3 + 1)(\xi^2 - 1)\cos(\theta)/(R_3 - 1) + (R_3 + 1)(2R_1\xi + \xi^2 + 1)}{2\xi(R_1R_3 + 1) + (R_1 + R_3)(\xi^2 + 1)}, \end{aligned} \right\} \end{equation}

with $\kappa =k_1Z_1a_1(0)/k_2Z_2a_2(0)$.

Liang & Luo (Reference Liang and Luo2022a) adopted the original prescription introduced by Richtmyer (Reference Richtmyer1960) to evaluate $\dot {a}_1^{SI}$ and $\dot {a}_2^{SI}$ considering the shock-reflected case in incompressible flow,

(4.7a,b)\begin{equation} \dot{a}_1^{SI}=k_1Z_1a_1(0)v_1^{\alpha}A_1,\quad \dot{a}_2^{SI}=k_2Z_2a_2(0)v_2^{\alpha}A_2. \end{equation}

Therefore, the original linear model only applies to two successive fast/slow interfaces in the weak-shock limit. In this work, we extend the linear model considering the compressibility by using the linearised compressible solution for a shocked slow/fast interface introduced by Wouchuk & Nishihara (Reference Wouchuk and Nishihara1996Reference Wouchuk and Nishihara1997). Then, the expressions of $\dot {a}_1^{SI}$ and $\dot {a}_2^{SI}$ can be written as

(4.8)\begin{equation} \left. \begin{aligned} & \dot{a}_1^{SI}=k_1a_1(0)v_1^{\alpha}\frac{v_{{IS}}-v_{{t1}} -R_1(1-v_{{ps}})(v_{{IS}}+c_A-v_1^{\alpha}) }{(R_1+1)v_{{IS}}},\\ & \dot{a}_2^{SI}=k_2a_2(0)v_2^{\alpha}\frac{v_{{t1}}-v_{{t2}} -R_3(1-v_1^{\alpha})(v_{{t1}}+c_B^{\beta}-v_2^{\alpha}) }{(R_3+1)v_{{t1}}}, \end{aligned} \right\} \end{equation}

in which $v_{{t2}}$ denotes the velocity of the TS$_2$. Equation (4.8) not only quantifies the baroclinic contribution to the vorticity but also considers the tangential velocity perturbations along the contact surface separately generated by the corrugated transmitted shock and the corrugated rarefaction waves.

Moreover, Liang & Luo (Reference Liang and Luo2022a) concluded that the feedthrough effect on RMI is partially related to the amplitude ratio of the two interfaces, which is not rigorous. From (4.6), it is noted that the feedthrough $\psi$ is not only related to the density ratios $R_1$ and $R_3$, the dimensionless distance $\xi$ and the phase difference $\theta$, but is also related to the ratio of the products of the wavenumber and amplitude of the two interfaces $\kappa$.

The values of $\dot {a}_1^{{lin}}$ and $\dot {a}_2^{{lin}}$ are calculated with the extended linear model (i.e. (4.5)–(4.8)) and listed in table 5. It can be seen that they agree well with all experimental and numerical results. Notably, the extended linear model successfully predicts the abnormal RMI of the upstream interface in the L10-AP case. Therefore, the extended linear model is proved to be applicable for describing the RMI of two successive slow/fast interfaces with various initial conditions.

The values of $\psi _1$ and $\psi _2$ in all experimental cases are listed in table 5. In the L10-AP case, $\psi _1<0$ indicates the abnormal RMI of the upstream interface caused by the feedthrough. Moreover, it is found that $|\psi _1|>1$ and $|\psi _2|>1$ in the three in-phase cases and $|\psi _1|<1$ and $|\psi _2|<1$ in the three anti-phase cases, indicating that the feedthrough leads to the two in-phase slow/fast interfaces being more unstable and the two anti-phase slow/fast interfaces being more stable in the six experimental cases. However, this conclusion is not rigorously applied to all possibilities. Some exceptions are found under specific initial conditions. The values of $\psi$ versus $kh_0$ under different $\kappa$ conditions when the two interfaces are initially in-phase and anti-phase are calculated based on (4.6), as shown in figures 8(a) and 8(b), respectively. It is found that when $kh_0>2$, the feedthrough has negligible influence on the RMI of the two interfaces.

Figure 8. The values of $\psi$ versus $kh_0$ under different $\kappa$ conditions when (a) $\theta =0$ and (b) $\theta ={\rm \pi}$. Hollow circles represent the instability freeze-out of the upstream or downstream interface.

In figure 8(a), we plot $\psi _1$ and $\psi _2$ as functions of $kh_0$ for the case $\kappa$ varying from 0.1 to 10.0, $R_1=1.5$, $R_3=0.5$ and $\theta =0$. It is found that when $kh_0<2$, except for the $\kappa =0.1$ case, $\psi _1>1$ and $\psi _2>1$, indicating that the feedthrough results in the two interfaces being more unstable. However, when $\kappa$ and $kh_0$ are limited (e.g. $\kappa =0.1$ and $kh_0<1$), $\psi _1>1$ and $1>\psi _2>0$, demonstrating that the feedthrough leads to the upstream interface being more unstable and the downstream interface being more stable. Therefore, the two successive interfaces with initially in-phase perturbations are generally destabilised by the feedthrough, but the downstream interface might be stabilised if $\kappa$ is limited.

In figure 8(b), we plot $\psi _1$ and $\psi _2$ as functions of $kh_0$ for the case $\kappa$ varying from 0.1 to 10.0, $R_1=1.5$, $R_3=0.5$ and $\theta ={\rm \pi}$. First, $\psi _1=0$ or/and $\psi _2=0$ under specific $kh_0$ conditions demonstrate that the feedthrough might cause the RMI of the upstream interface or/and the downstream interface to be frozen. Notably, the freeze-out RMI concerned in this study is caused by the feedthrough between the two interfaces, not the compressibility as discussed by previous studies (Fraley Reference Fraley1986; Mikaelian Reference Mikaelian1993Reference Mikaelian1994; Wouchuk & Nishihara Reference Wouchuk and Nishihara1997Reference Wouchuk and Nishihara2004; Wouchuk & Sano Reference Wouchuk and Sano2015). Moreover, $\psi _1<0$ or $\psi _2<0$ demonstrates that the abnormal RMI of the upstream interface or downstream interface occurs. Although the two slow/fast interfaces are initially anti-phase, $|\psi _1|>1$ under the small $\kappa$ condition (e.g. $\kappa =0.1$ and 0.5) indicates the RMI of the upstream interface is promoted, and $|\psi _2|>1$ under the large $\kappa$ condition (e.g. $\kappa =10.0$) indicates the RMI of the downstream interface is promoted.

Overall, the feedthrough generally leads to the two in-phase slow/fast interfaces being more unstable and the two anti-phase slow/fast interfaces being more stable, but there are exceptions under specific conditions: first, the downstream one of the two initially in-phase interfaces might be stabilised if $\kappa$ is limited; second, the upstream one of the two initially anti-phase interfaces might be destabilised if $\kappa$ is limited; third, the downstream one of the two initially anti-phase interfaces might be destabilised if $\kappa$ is large. It is evident that the freeze-out and abnormal RMI complicate the analysis of the hydrodynamic instabilities of the two interfaces and will be discussed in detail later.

4.3. The reverberating rarefaction waves effect

The time-varying amplitudes of the upstream interface and downstream interface are scaled to eliminate the feedthrough effect and highlight the reverberating rarefaction waves effect, as shown in figures 9(a) and 9(b), respectively. For the upstream interface, time is scaled as $\tau _1=|k_1\dot {a}_1^{{lin}}t|$ and amplitude is scaled as $\eta _1=|k_1a_1|$. For the downstream interface, time is scaled as $\tau _2=|k_2\dot {a}_2^{{lin}}(t-t_2^\alpha )|$ and amplitude is scaled as $\eta _2=|k_2a_2|$. Except for $\eta _1$ in the L10-AP case, both $\eta _1$ and $\eta _2$ decrease to zero at first due to phase reversal (Brouillette Reference Brouillette2002) and then increase. The rarefaction waves have short-time influences on the hydrodynamic instabilities due to the short duration times of the interaction of the rarefaction waves and the two interfaces in this study (see $T_{{rRW}_2}$ and $T_{{rRW}_1}$ in table 3). However, it is found that the perturbation of the upstream interface in the L10-AP case is suddenly stretched by the rRW$_2$, as seen in the inset of figure 9(a). The rarefaction waves influence the hydrodynamic instabilities of the two interfaces in two ways: on the one hand, the rarefaction waves stretch the two interfaces in the streamwise direction during the interaction; on the other hand, the rarefaction waves impose the short-time RTI or RTS on the two interfaces since the interaction of the rarefaction waves and an interface is not instantaneous but a process (Li & Book Reference Li and Book1991; Li et al. Reference Li, Kailasanath and Book1991; Morgan et al. Reference Morgan, Likhachev and Jacobs2016Reference Morgan, Cabot, Greenough and Jacobs2018; Liang et al. Reference Liang, Zhai, Luo and Wen2020b; Wang et al. Reference Wang, Song, Ma, Ma, Wang and Wang2022).

Figure 9. The time-varying dimensionless amplitudes of the (a) upstream interface and (b) downstream interface obtained from experiments.

To simplify the analysis, the interaction of 1-D rarefaction waves and 2-D interfaces are considered. For the upstream interface, since the rRW$_2$ accelerates the spike then the bubble of the upstream interface, the rRW$_2$ stretches the perturbation of the upstream interface during the interaction. The amplitude of the upstream interface when the rRW$_2$ head impacts the upstream interface is $a_1^\beta$ ($=Z_1a_1(0)+\dot {a}_1^{{lin}}t_1^\beta$), the relative velocity difference between the upstream interface and the rRW$_2$ is $c_B^\alpha$, and the upstream interface velocity increment induced by the rRW$_2$ is ($v_1^\beta -v_1^\alpha$). Therefore, the time-varying stretching effect of rRW$_2$ on the upstream interface amplitude ($a_{{rRW_2}}$) is

(4.9)\begin{equation} a_{{rRW_2}}=\frac{(v_1^\beta-v_1^\alpha)(t-t_1^\beta+a_1^\beta/c_B^\alpha)}{2}, \end{equation}

with $(t_1^\beta +a_1^\beta /c_B^\alpha )>t>(t_1^\beta -a_1^\beta /c_B^\alpha )$.

For the downstream interface, since the rRW$_1$ decelerates the bubble tip then the spike tip of the downstream interface, the rRW$_1$ also stretches the perturbation of the downstream interface during the interaction. The amplitude of the downstream interface when the rRW$_1$ head impacts the downstream interface is $a_2^\beta$ ($=Z_2a_2(0)+\dot {a}_2^{{lin}}(t_2^\beta -t_2^\alpha )$), the relative velocity difference between the downstream interface and the rRW$_1$ is $c_B^\beta$, and the downstream interface velocity decrement induced by the rRW$_1$ is ($v_2^\alpha -v_2^\beta$). Therefore, the time-varying stretching effect of rRW$_1$ on the downstream interface amplitude ($a_{{rRW_1}}$) is

(4.10)\begin{equation} a_{{rRW_1}}=\frac{(v_2^\alpha-v_1^\beta)(t-t_2^\beta+a_2^\beta/c_B^\beta)}{2}, \end{equation}

with $(t_2^\beta +a_2^\beta /c_B^\beta )>t>(t_2^\beta -a_2^\beta /c_B^\beta )$.

Except for the freeze-out and abnormal RMI cases, the two slow/fast interfaces generally experience phase reversal, resulting in the difficulty of modelling the additional instabilities imposed by the rarefaction waves on the two interfaces. Since the post-shock amplitude of the upstream (or downstream) interface is $Z_1a_1(0)$ (or $Z_2a_2(0)$) and the perturbation amplitude of the upstream (or downstream) interface decreases at a linear rate ${\dot {a}}_1^{{lin}}$ (or ${\dot {a}}_2^{{lin}}$), the time when the perturbation of the upstream (or downstream) interface decreases to zero, i.e. the end time of the phase reversal of the upstream (or downstream) interface $t_1^{{pr}}$ (or $t_2^{{pr}}$), can be evaluated as

(4.11a,b)\begin{equation} t_1^{{pr}}=|Z_1a_1(0)/\dot{a}_1^{{lin}}|,\quad t_2^{{pr}}=t_2^\alpha+|Z_2a_2(0)/\dot{a}_2^{{lin}}|. \end{equation}

The values of $t_1^{{pr}}$ and $t_2^{{pr}}$ in all cases are listed in table 6. On comparing $t_1^{{pr}}$ with $t_1^{\beta }$ in table 2, it is found $t_1^{{pr}}>t_1^{\beta }$ in all cases except for the case L10-AP, indicating that the rRW$_2$ impacts the upstream interface before the end of the upstream interface's phase reversal. Because of the baroclinic vorticity created by the misalignment of the density gradient ($\boldsymbol {\nabla }\rho$) and the pressure gradient ($\boldsymbol {\nabla } p$), the rRW$_2$ continuously deposits the vorticity with the same direction as the vorticity deposited by the IS on the upstream interface, as sketched in figure 10(a). Therefore, the rRW$_2$ imposes the short-time RTI on the upstream interface, resulting in the upstream interface being more unstable. Due to the abnormal RMI in the L10-AP case, the vorticity deposited on the upstream interface by the IS is opposite to that in other cases, as sketched in figure 10(b). As a result, the rRW$_2$ imposes the short-time RTS on the upstream interface in the L10-AP case, leading to the upstream interface being more stable.

Table 6. The end times of the upstream (or downstream) interface's phase reversal $t_1^{{pr}}$ (or $t_2^{{pr}}$) calculated with (4.11a,b). The unit of time is $\mathrm {\mu }$s.

Figure 10. Sketches of (a) the interaction of the rRW$_2$ and SI$_1$ in all cases (except for the L10-AP case), (b) the interaction of the rRW$_2$ and SI$_1$ in the L10-AP case, (c) the interaction of the rRW$_1$ and SI$_2$ under $L_0=10.0$ and 30.0 mm conditions, and (d) the interaction of the rRW$_1$ and SI$_2$ under $L_0=50.0$ mm conditions. The blue (or orange) arcs with an arrow in panels (a,b) illustrate the vorticity deposited by the IS (or rRW$_2$) on the upstream interface. The blue (or orange) arcs with an arrow in panels (c,d) represent the vorticity deposited by the TS$_1$ (or rRW$_1$) on the downstream interface. The purple (or green) arrows represent the pressure (or density) gradient $\boldsymbol {\nabla } p$ (or $\boldsymbol {\nabla }\rho$).

Similarly, by comparing $t_2^{{pr}}$ with $t_2^{\beta }$ in table 2, it is found that $t_2^{{pr}}>t_2^{\beta }$ under $L_0=10.0$ and 30.0 mm conditions and $t_2^{{pr}}< t_2^{\beta }$ under $L_0=50.0$ mm conditions. In other words, the rRW$_1$ impacts the downstream interface before the end of the downstream interface's phase reversal under $L_0=10.0$ and 30.0 mm conditions, depositing the vorticity in the opposite direction to the vorticity deposited by the TS$_1$ on the upstream interface, as sketched in figure 10(c). In contrast, the rRW$_1$ impacts the downstream interface after the downstream interface's phase reversal under $L_0=50.0$ mm conditions, depositing the vorticity with the same direction as the vorticity deposited by the TS$_1$ on the downstream interface, as sketched in figure 10(d). As a result, the rRW$_1$ imposes the short-time RTS on the downstream interface, resulting in the downstream interface being more stable under $L_0=10.0$ and 30.0 mm conditions. In contrast, the rRW$_1$ imposes the short-time RTI on the downstream interface, leading to the upstream interface being more unstable under $L_0=50.0$ mm conditions.

Here, we modify the nonlinear model proposed by Zhang & Guo (Reference Zhang and Guo2016), considering the universal curves of all spikes and bubbles at any density ratio, to quantify the short-time RTI or RTS imposed by the rRW$_2$ on the upstream interface and that imposed by the rRW$_1$ on the downstream interface. The expressions of the modified nonlinear model for the bubble/spike amplitude growth rate of the upstream interface ($\dot {a}_{1b/1s}$) and the downstream interface ($\dot {a}_{2b/2s}$) in Stage II are

(4.12)\begin{equation} \ddot{a}_{ns/nb}=({-}1)^i\alpha_{ns/nb}k_n\left[(\dot{a}_{ns/nb})^{2}-(\dot{a}_{ns/nb}^{qs})^2\right]. \end{equation}

For the upstream interface, $i=0$ when $t_1^{{pr}}>t_1^{\beta }$ or $\psi _1<0$, and $i=1$ when $t_1^{{pr}}< t_1^{\beta }$; for the downstream interface, $i=0$ when $t_2^{{pr}}< t_2^{\beta }$ or $\psi _2<0$, and $i=1$ when $t_2^{{pr}}>t_2^{\beta }$;

(4.13)\begin{equation} \dot{a}_{ns/nb}^{qs}=\sqrt{\frac{|A_n\bar{g}_{{rRW_{(3-n)}}}|}{3k_n}\frac{8}{(1\pm A_n)(3\pm A_n)}\frac{\left[3\pm A_n+\sqrt{2(1\pm A_n)}\right]^{2}}{\left[4(3\pm A_n)+\sqrt{2(1\pm A_n)}(9\pm A_n)\right]}}, \end{equation}

and

(4.14)\begin{equation} \alpha_{ns/nb}=\frac{3}{4}\frac{(1\pm A_n)(3\pm A_n)}{\left[3\pm A_n+\sqrt{2(1\pm A_n)}\right]}\frac{\left[4(3\pm A_n)+\sqrt{2(1\pm A_n)}(9\pm A_n)\right]}{\left[\sqrt{3\pm A_n}+2\sqrt{2(1\pm A_n)}(3\mp A_n)\right]}. \end{equation}

The upper (or lower) signs of $\pm$ and $\mp$ in all equations apply to the spike (or bubble). Note that the feedthrough is considered since $\dot {a}_{ns/nb}$ equals $\dot {a}_n^{{lin}}$ at $t_n^\beta$.

After the rRW$_2$ (or rRW$_1$) leaves the upstream (or downstream) interface, only the RMI dominates the perturbation growths of the upstream (or downstream) interface in Stage III. The upstream (or downstream) interface evolves linearly with the amplitude growth rate almost the same as in Stage I. The predictions of the models established in this work for the upstream (or downstream) interface amplitude in Stages I, II and III are marked as solid blue (or dashed orange) lines in figure 11(af) in all cases, and they agree well with the experimental and numerical results. Notably, the stretching effect imposed by the rarefaction waves on the upstream interface perturbation growth is well described in the inset of figure 11(b).

Figure 11. Comparisons of the amplitudes of the upstream interface and downstream interface measured from experiments and simulations with theories in cases (a) L10-IP, (b) L10-AP, (c) L30-IP, (d) L30-AP, (e) L50-IP and (f) L50-AP. The solid and hollow symbols represent the experimental results for the upstream interface and downstream interface, respectively; the solid red and dashed green lines represent the simulation results for the upstream interface and downstream interface, respectively; and the solid blue and dashed orange lines represent the theoretical predictions for the upstream interface and downstream interface, respectively.

4.4. The freeze-out and abnormal RMI

The freeze-out and abnormal RMI will be discussed in detail using the extended linear model and numerical simulations. It is evident that the sufficient condition for the perturbation growth freeze-out of the upstream interface is $\psi _1=0$, and that of the downstream interface is $\psi _2=0$. Based on (4.6), the freeze-out RMI of the upstream interface occurs when

(4.15)\begin{equation} (\xi^2-1)\cos(\theta)=\frac{\kappa(2R_3\xi+\xi^2+1)(R_1-1)}{1-R_3}, \end{equation}

and the freeze-out RMI of the downstream interface occurs when

(4.16)\begin{equation} (\xi^2-1)\cos(\theta)=\frac{(1-R_3)(2R_1\xi+\xi^2+1)}{\kappa(R_1-1)}. \end{equation}

For two successive slow/fast interfaces, $R_1>1$ and $R_3<1$, indicating that the right-hand side of the equality signs in (4.15) and (4.16) are larger than 0. Since $\xi \in (0,1)$, the terms $(\xi ^2-1)$ must be negative on the left-hand side of the equality signs, and, therefore, the term $\cos (\theta )$ must also be negative. As a result, the phase difference between the perturbations of the two interfaces $\theta$ should be within the range of $(0.5{\rm \pi}, {\rm \pi}]$. Suppose the fluid distribution and perturbations of the two interfaces are known. In that case, the values of $R_1$, $R_3$, $\kappa$ and $\theta$ are given. Then, we can deduce the value of the initial distance between the two interfaces $L_0$ for the freeze-out RMI of the upstream interface or the downstream interface by solving (4.15) or (4.16).

First, we take case F-1 as an example to freeze the RMI of the upstream interface. The initial parameters in case F-1 are the same as the experimental case L10-AP, including $R_1=1.37$, $R_3=0.33$, $\kappa =1.0$ and $\theta ={\rm \pi}$. Then, we can deduce that when the initial distance $L_0$ equals 11.8 mm, the RMI of the upstream interface is frozen. The time-varying amplitudes of the two interfaces in the F-1 case are shown in figure 12(a). The perturbation growth of the upstream interface is frozen by the feedthrough (see the solid symbols), which can be well described by the models established in this work (see the solid lines).

Figure 12. The time-varying amplitudes of the upstream interface and downstream interface in cases (a) F-1 and (b) F-2. The solid and hollow symbols represent the numerical results for the upstream interface and downstream interface, respectively; the solid red and dashed blue lines represent the theoretical predictions for the upstream interface and downstream interface, respectively, and similarly hereinafter.

Second, we take case F-2 as an example to freeze the RMI of the downstream interface. Most of the initial parameters in case F-2 are the same as in case F-1, including $R_1=1.37$, $R_3=0.33$ and $\theta ={\rm \pi}$. Differently, we adopt $\kappa =4$ in the F-2 case. Then, we can deduce that when the initial distance $L_0$ equals 9.1 mm, the RMI of the downstream interface is frozen. The time-varying amplitudes of the two interfaces in the F-2 case are shown in figure 12(b). The perturbation growth of the downstream interface is frozen by the feedthrough (see the hollow symbols), which can be well described by the models established in this work (see the dashed lines). However, we note that $a_1$ does not decrease to 0 when the upstream interface finishes its phase reversal (see $a_1$ at approximately 1160 $\mathrm {\mu }$s), which goes against the theoretical prediction. We conjecture that this phenomenon is ascribed to the effect of high-order harmonics introduced by the feedthrough on the RMI of the upstream interface at a later time. The feedthrough accompanied by the nonlinearity effect on the RMI of two interfaces will be investigated in further studies.

In the ICF, we prefer the perturbations on all surfaces to be frozen by the feedthrough. Based on (4.15) and (4.16), it is easy to write the sufficient condition for the simultaneously freeze-out RMI of the two interfaces as

(4.17)\begin{equation} \kappa(2R_3\xi+\xi^2+1)(R_1-1)^2=(2R_1\xi+\xi^2+1)(1-R_3)^2. \end{equation}

It can be found that the simultaneously freeze-out RMI can be achieved by carefully choosing the combinations of $R_1$, $R_3$, $\kappa$ and $\xi$. Moreover, we can find a specific solution for (4.17): $\kappa =1$, $R_1+R_3=2$ and $\xi \ll 1$. Here, we take case F-3 as an example to freeze the RMI of both the upstream and downstream interfaces and validate the specific solution. The initial parameters in case F-3 include the incident shock Mach number of 1.24, $\rho _A=6.14$ kg m$^{-3}$, $\rho _B=4.10$ kg m$^{-3}$, $\rho _C=2.05$ kg m$^{-3}$, $L_0=4.0$ mm, $k_1=k_2=20.9$ m$^{-1}$, $a_1(0)=a_2(0)=0.5$ mm and $\theta ={\rm \pi}$. In other words, we set $R_1=1.5$, $R_3=0.5$, $\kappa =1.0$ and $\xi =0.03$ in case F-3. The time-varying amplitudes of the two interfaces in the F-3 case are shown in figure 13. The feedthrough freezes the perturbation growths of both the upstream and downstream interfaces (see the solid and hollow symbols), which can be well described by the models established in this work (see the solid and dashed lines).

Figure 13. The time-varying amplitudes of the upstream and downstream interfaces in case F-3.

Next, we focus on the abnormal RMI of the upstream or downstream interface. It is evident that the sufficient condition for the abnormal RMI of the upstream interface is $\psi _1<0$, and that of the downstream interface is $\psi _2<0$. Because the denominators on the right-hand side of (4.6) are larger than 1, the numerators on the right-hand side of (4.6) should be negative if the abnormal RMI occurs. Therefore, for two successive slow/fast interfaces, the abnormal RMI of the upstream interface occurs when

(4.18)\begin{equation} (\xi^2-1)\cos(\theta)>\frac{\kappa(2R_3\xi+\xi^2+1)(R_1-1)}{1-R_3}, \end{equation}

and the abnormal RMI of the downstream interface occurs when

(4.19)\begin{equation} (\xi^2-1)\cos(\theta)>\frac{(1-R_3)(2R_1\xi+\xi^2+1)}{\kappa(R_1-1)}. \end{equation}

Similar to the abnormal RMI of the two interfaces, the right-hand sides of the equality signs in (4.18) and (4.19) are larger than 0. Therefore, the term $\cos (\theta )$ must be negative. As a result, the phase difference $\theta \in (0.5{\rm \pi}, {\rm \pi}]$ is not only the necessary condition for the freeze-out RMI but also the necessary condition for the abnormal RMI.

We plot $\psi _1$ and $\psi _2$ as functions of $kh_0$ for the case $\theta$ varying from 0 to ${\rm \pi}$, $R_1=1.5$, $R_3=0.5$, and $\kappa =0.1$ and 10.0 in figures 14(a) and 14(b), respectively. It is again confirmed that the abnormal RMI cannot occur when $\theta \leq 0.5{\rm \pi}$. Furthermore, if the abnormal RMI occurs on the upstream (or downstream) interface, then the downstream (or upstream) interface must experience phase reversal. In other words, the abnormal RMI cannot occur on the two interfaces simultaneously. For example, if the abnormal RMI occurs on the downstream interface, then (4.19) holds. Because $\xi \in (0,1)$ and $\cos (\theta )\in [-1,1]$, the product of the two terms on the left-hand side of (4.19) is smaller than 1, and, therefore, the product of the terms on the right-hand side of (4.19) must be smaller than 1, too. As a result, we can derive an inequality as

(4.20)\begin{equation} \kappa(R_1-1)>(1-R_3)(2R_1\xi+\xi^2+1). \end{equation}

Then, we substitute (4.20) to the right-hand side of (4.18) and acquire

(4.21)\begin{equation} (\xi^2-1)\cos(\theta)>(2R_1\xi+\xi^2+1)(2R_3\xi+\xi^2+1). \end{equation}

However, (4.21) cannot hold because the product of the terms on the left-hand side is less than 1, but the product of the terms on the right-hand side is larger than 1. Therefore, the abnormal RMI of the upstream interface cannot occur. As a result, we have proved that the abnormal RMI cannot occur on the two interfaces simultaneously.

Figure 14. The values of $\psi$ versus $kh_0$ under different $\theta$ conditions when (a) $\kappa =0.1$ and (b) $\kappa =10.0$.

5. Conclusions

Shock-tube experiments and simulations on two successive slow/fast interfaces are performed to investigate wave patterns, interface movements and hydrodynamic instabilities of the two interfaces. Three quasi-1-D cases with different initial distances between the two interfaces and six quasi-2-D cases with various initial distances and perturbations are created using the extended soap film approach in the experiments. Schlieren photography combined with a high-speed camera is used to capture legible experimental pictures, providing the first observational evidence of the abnormal RMI. The reverberating rarefaction waves between the two interfaces are focused. The rarefaction waves, on the one hand, stretch the two interface perturbations; on the other hand, they impose the short-time RTI or RTS on the two interfaces depending on the phases of the two interfaces when the rarefaction waves impact them. Theoretical models for quantifying the various hydrodynamic instabilities are examined. For the first time, the conditions and outcomes of the freeze-out and abnormal RMI caused by the feedthrough between the two interfaces are summarised.

First, the wave patterns and interface movements of two successive slow/fast interfaces are studied. The rarefaction waves reflected from the downstream interface accelerate the upstream interface, and the rarefaction waves reflected from the upstream interface decelerate the downstream interface. The movements of the two interfaces finally enter the asymptotic state. A general 1-D theory is adopted to describe the two interfaces’ movements in three stages.

Second, the hydrodynamic instabilities of two successive slow/fast interfaces induced by a shock wave are investigated. The final phases of the two interfaces are generally opposite to the two interfaces’ initial perturbations. However, the abnormal RMI, that is, phase reversal of a shocked slow/fast interface is inhibited, is observed when the initial distance is limited, and the two interfaces are initially anti-phase. The hydrodynamic instabilities can be separated into three stages: Stage I. The RMI affected by the feedthrough dominates the two interface perturbation growths. Compared with the semi-infinite RMI, the feedthrough destabilises the two in-phase interfaces but stabilises the two anti-phase interfaces in the experiments. Stage II. On the one hand, the rarefaction waves stretch the two interfaces’ streamwise perturbations. On the other hand, if the end time of the upstream interface's phase reversal is larger (or smaller) than the time when the rarefaction waves impact the upstream interface, i.e. $t_1^{{pr}}>t_1^{\beta }$ (or $t_1^{{pr}}< t_1^{\beta }$), the rarefaction waves impose the short-time RTI (or RTS) on the upstream interface, destabilising (or stabilising) the upstream interface; if the end time of the downstream interface's phase reversal is larger (or smaller) than the time when rarefaction waves impact the downstream interface, i.e. $t_2^{{pr}}>t_2^{\beta }$ (or $t_2^{{pr}}< t_2^{\beta }$), the rarefaction waves impose the short-time RTS (or RTI) on the downstream interface, stabilising (or destabilising) the downstream interface. Stage III. After the rarefaction waves are refracted away from the two interfaces, only the RMI dominates the two interfaces’ developments.

Third, the linear solution proposed by Liang & Luo (Reference Liang and Luo2022a) is extended by considering arbitrary wavenumber and phase combinations and the compressibility, which is proved to apply to two successive slow/fast interfaces and successfully quantify the feedthrough. It is proved that the feedthrough generally leads to the two in-phase slow/fast interfaces being more unstable and the two anti-phase slow/fast interfaces being more stable, but there are exceptions under specific conditions. During the interaction of the reverberating rarefaction waves and the two interfaces, on the one hand, the stretching effect imposed by the rarefaction waves on the two interfaces is quantified; on the other hand, the nonlinear model proposed by Zhang & Guo (Reference Zhang and Guo2016) is modified to describe the RTI or RTS imposed by the rarefaction waves on the two interfaces with the consideration of the two interfaces’ phase reversal.

The freeze-out and abnormal RMI caused by the feedthrough are discussed in detail. The necessary condition for both the freeze-out and abnormal RMI is $\theta \in (0.5{\rm \pi}, {\rm \pi}]$. The sufficient conditions for the instability freeze-out of the upstream interface or/and the downstream interface are $\psi _1=0$ or/and $\psi _2=0$. The sufficient condition of the abnormal RMI of the upstream interface or the downstream interface is $\psi _1<0$ or $\psi _2<0$. It has been proved that the simultaneous freeze-out RMI of the two interfaces occurs under specific conditions, but the abnormal RMI cannot occur on the two interfaces simultaneously. Our study suggests that when the initial distance between two successive surfaces is limited, and the perturbations on the two surfaces are comparable, it is potentially favourable to adopt the fluid distribution satisfying ($R_1+R_3=2$) in the ICF capsule design to freeze the instability of the two interfaces simultaneously.

In the ICF, the initial perturbations on the surfaces are essentially multi-mode with wavenumbers spanning many orders of magnitude. It is believed that the coupling occurs not only between the constituent modes of each interface but also between the constituent modes of all interfaces. We look forward to performing the experiments and simulations on two successive interfaces consisting of multi-mode perturbations and proposing a general nonlinear theory to quantify the complicated mode coupling mechanism.

Funding

This work was supported by the Natural Science Foundation of China (nos. 91952205 and 11625211), and Tamkeen under the NYU Abu Dhabi Research Institute grant CG002.

Declaration of interests

The authors report no conflict of interest.

References

REFERENCES

Abgrall, R. 1996 How to prevent pressure oscillations in multicomponent flow calculations: a quasi conservative approach. J. Comput. Phys. 125 (1), 150160.CrossRefGoogle Scholar
Aglitskiy, Y., et al. 2006 Perturbation evolution started by Richtmyer–Meshkov instability in planar laser targets. Phys. Plasmas 13 (8), 080703.CrossRefGoogle Scholar
Bai, J.S., Zou, L.Y., Wang, T., Liu, K., Huang, W.B., Liu, J.H., Li, P., Tan, D.W. & Liu, C.L. 2010 Experimental and numerical study of shock-accelerated elliptic heavy gas cylinders. Phys. Rev. E 82, 056318.CrossRefGoogle ScholarPubMed
Balakumar, B.J., Orlicz, G.C., Ristorcelli, J.R., Balasubramanian, S., Prestridge, K.P. & Tomkins, C.D. 2012 Turbulent mixing in a Richtmyer–Meshkov fluid layer after reshock: velocity and density statistics. J. Fluid Mech. 696, 6793.CrossRefGoogle Scholar
Balakumar, B.J., Orlicz, G.C., Tomkins, C.D. & Prestridge, K. 2008 Simultaneous particle-image velocimetry-planar laser-induced fluorescence measurements of Richtmyer–Meshkov instability growth in a gas curtain with and without reshock. Phys. Fluids 20, 124103.CrossRefGoogle Scholar
Balasubramanian, S., Orlicz, G.C., Prestridge, K.P. & Balakumar, B.J. 2012 Experimental study of initial condition dependence on Richtmyer–Meshkov instability in the presence of reshock. Phys. Fluids 24 (3), 034103.CrossRefGoogle Scholar
Brouillette, M. 2002 The Richtmyer–Meshkov instability. Annu. Rev. Fluid Mech. 34, 445468.CrossRefGoogle Scholar
Budzinski, J.M., Benjamin, R.F. & Jacobs, J.W. 1994 Influence of initial conditions on the flow patters of a shock-accelerated thin fluid layer. Phys. Fluids 6, 35103512.CrossRefGoogle Scholar
Desjardins, T.R., et al. 2019 A platform for thin-layer Richtmyer–Meshkov at OMEGA and the NIF. High Energy Density Phys. 33, 100705.CrossRefGoogle Scholar
Ding, J., Liang, Y., Chen, M., Zhai, Z., Si, T. & Luo, X. 2018 Interaction of planar shock wave with three-dimensional heavy cylindrical bubble. Phys. Fluids 30 (10), 106109.Google Scholar
Ding, J., Si, T., Chen, M., Zhai, Z., Lu, X. & Luo, X. 2017 On the interaction of a planar shock with a three-dimensional light gas cylinder. J. Fluid Mech. 828, 289317.CrossRefGoogle Scholar
Drake, R.P. 2018 High-Energy-Density Physics: Foundation of Inertial Fusion and Experimental Astrophysics. Springer.CrossRefGoogle Scholar
Fan, E., Guan, B., Wen, C.Y. & Shen, H. 2019 Numerical study on the jet formation of simple-geometry heavy gas inhomogeneities. Phys. Fluids 31, 026103.CrossRefGoogle Scholar
de Frahan, M.T.H., Movahed, P. & Johnsen, E. 2015 Numerical simulations of a shock interacting with successive interfaces using the discontinuous galerkin method: the multilayered Richtmyer–Meshkov and Rayleigh–Taylor instabilities. Shock Waves 25 (4), 329345.CrossRefGoogle Scholar
Fraley, G. 1986 Rayleigh–Taylor stability for a normal shock wave-density discontinuity interaction. Phys. Fluids 29, 376386.CrossRefGoogle Scholar
Grove, J.W., Holmes, R., Sharp, D.H., Yang, Y. & Zhang, Q. 1993 Quantitative theory of Richtmyer–Meshkov instability. Phys. Rev. Lett. 71, 34733476.CrossRefGoogle ScholarPubMed
Guan, B., Liu, Y., Wen, C.Y. & Shen, H. 2018 Numerical study on liquid droplet internal flow under shock impact. AIAA J. 56 (9), 33823387.CrossRefGoogle Scholar
Haan, S.W., et al. 2011 Point design targets, specifications, and requirements for the 2010 ignition campaign on the National Ignition Facility. Phys. Plasmas 18 (5), 051001.CrossRefGoogle Scholar
Herrmann, M., Moin, P. & Abarzhi, S.I. 2008 Nonlinear evolution of the Richtmyer–Meshkov instability. J. Fluid Mech. 612, 311338.CrossRefGoogle Scholar
Holmes, R.L., Dimonte, G., Fryxell, B., Gittings, M.L., Grove, J.W., Schneider, M., Sharp, D.H., Velikovich, A.L., Weaver, R.P. & Zhang, Q. 1999 Richtmyer–Meshkov instability growth: experiment, simulation and theory. J. Fluid Mech. 389, 5579.CrossRefGoogle Scholar
Holmes, R.L. & Grove, J.W. 1995 Numerical investigation of Richtmyer–Meshkov instability using front tracking. J. Fluid Mech. 301, 5164.CrossRefGoogle Scholar
Hsing, W.W. & Hoffman, N.M. 1997 Measurement of feedthrough and instability growth in radiation-driven cylindrical implosions. Phys. Rev. Lett. 78 (20), 3876.CrossRefGoogle Scholar
Igra, D. & Igra, O. 2020 Shock wave interaction with a polygonal bubble containing two different gases, a numerical investigation. J. Fluid Mech. 889, A26.CrossRefGoogle Scholar
Jacobs, J.W., Jenkins, D.G., Klein, D.L. & Benjamin, R.F. 1995 Nonlinear growth of the shock-accelerated instability of a thin fluid layer. J. Fluid Mech. 295, 2342.CrossRefGoogle Scholar
Jacobs, J.W., Klein, D.L., Jenkins, D.G. & Benjamin, R.F. 1993 Instability growth patterns of a shock-accelerated thin fluid layer. Phys. Rev. Lett. 70, 583586.CrossRefGoogle ScholarPubMed
Jiang, Y., Wen, C.Y. & Zhang, D. 2020 Space-time conservation element and solution element method and its applications. AIAA J. 58 (12), 54085430.CrossRefGoogle Scholar
Jourdan, G. & Houas, L. 2005 High-amplitude single-mode perturbation evolution at the Richtmyer–Meshkov instability. Phys. Rev. Lett. 95, 204502.CrossRefGoogle ScholarPubMed
Knapp, P.F., et al. 2017 Direct measurement of the inertial confinement time in a magnetically driven implosion. Phys. Plasmas 24 (4), 042708.CrossRefGoogle Scholar
Leinov, E., et al. 2009 Experimental and numerical investigation of the Richtmyer–Meshkov instability under re-shock conditions. J. Fluid Mech. 626, 449475.CrossRefGoogle Scholar
Li, C. & Book, D.L. 1991 Instability generated by acceleration due to rarefaction waves. Phys. Rev. A 43 (6), 3153.CrossRefGoogle ScholarPubMed
Li, C., Kailasanath, K. & Book, D.L. 1991 Mixing enhancement by expansion waves in supersonic flows of different densities. Phys. Fluids A 3 (5), 13691373.CrossRefGoogle Scholar
Li, Y., Samtaney, R. & Wheatley, V. 2018 The Richtmyer–Meshkov instability of a double-layer interface in convergent geometry with magnetohydrodynamics. Matter Radiat. Extrem. 3 (4), 207218.CrossRefGoogle Scholar
Liang, Y. 2022 a Fundamental Studies of Shock-Driven Hydrodynamic Instabilities. Springer Nature.CrossRefGoogle Scholar
Liang, Y. 2022 b The phase effect on the Richtmyer–Meshkov instability of a fluid layer. Phys. Fluids 34 (3), 034106.CrossRefGoogle Scholar
Liang, Y., Liu, L., Zhai, Z., Ding, J., Si, T. & Luo, X. 2021 Richtmyer–Meshkov instability on two-dimensional multi-mode interfaces. J. Fluid Mech. 928, A37.CrossRefGoogle Scholar
Liang, Y., Liu, L., Zhai, Z., Si, T. & Wen, C.-Y. 2020 a Evolution of shock-accelerated heavy gas layer. J. Fluid Mech. 886, A7.CrossRefGoogle Scholar
Liang, Y. & Luo, X. 2021 a On shock-induced heavy-fluid-layer evolution. J. Fluid Mech. 920, A13.CrossRefGoogle Scholar
Liang, Y. & Luo, X. 2021 b Shock-induced dual-layer evolution. J. Fluid Mech. 929, R3.CrossRefGoogle Scholar
Liang, Y. & Luo, X. 2022 a On shock-induced evolution of a gas layer with two fast/slow interfaces. J. Fluid Mech. 939, A16.CrossRefGoogle Scholar
Liang, Y. & Luo, X. 2022 b On shock-induced light-fluid-layer evolution. J. Fluid Mech. 933, A10.CrossRefGoogle Scholar
Liang, Y., Zhai, Z., Ding, J. & Luo, X. 2019 Richtmyer–Meshkov instability on a quasi-single-mode interface. J. Fluid Mech. 872, 729751.CrossRefGoogle Scholar
Liang, Y., Zhai, Z., Luo, X. & Wen, C. 2020 b Interfacial instability at a heavy/light interface induced by rarefaction waves. J. Fluid Mech. 885, A42.CrossRefGoogle Scholar
Liu, W., Li, X., Yu, C., Fu, Y., Wang, P., Wang, L. & Ye, W. 2018 b Theoretical study on finite-thickness effect on harmonics in Richtmyer–Meshkov instability for arbitrary Atwood numbers. Phys. Plasmas 25 (12), 122103.CrossRefGoogle Scholar
Liu, L., Liang, Y., Ding, J., Liu, N. & Luo, X. 2018 a An elaborate experiment on the single-mode Richtmyer–Meshkov instability. J. Fluid Mech. 853, R2.CrossRefGoogle Scholar
Livescu, D. 2020 Turbulence with large thermal and compositional density variations. Annu. Rev. Fluid Mech. 52, 309341.CrossRefGoogle Scholar
Luo, X., Liang, Y., Si, T. & Zhai, Z. 2019 Effects of non-periodic portions of interface on Richtmyer–Meshkov instability. J. Fluid Mech. 861, 309327.CrossRefGoogle Scholar
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.CrossRefGoogle ScholarPubMed
Meshkov, E.E. 1969 Instability of the interface of two gases accelerated by a shock wave. Fluid Dyn. 4, 101104.CrossRefGoogle Scholar
Mikaelian, K.O. 1985 Richtmyer–Meshkov instabilities in stratified fluids. Phys. Rev. A 31, 410419.CrossRefGoogle ScholarPubMed
Mikaelian, K.O. 1993 Growth rate of the Richtmyer–Meshkov instability at shocked interfaces. Phys. Rev. Lett. 71 (18), 2903.CrossRefGoogle ScholarPubMed
Mikaelian, K.O. 1994 Freeze-out and the effect of compressibility in the Richtmyer–Meshkov instability. Phys. Fluids 6, 356368.CrossRefGoogle Scholar
Mikaelian, K.O. 1995 Rayleigh–Taylor and Richtmyer–Meshkov instabilities in finite-thickness fluid layers. Phys. Fluids 7 (4), 888890.CrossRefGoogle Scholar
Mikaelian, K.O. 1996 Numerical simulations of Richtmyer–Meshkov instabilities in finite-thickness fluid layers. Phys. Fluids 8 (5), 12691292.CrossRefGoogle Scholar
Mikaelian, K.O. 2005 Rayleigh–Taylor and Richtmyer–Meshkov instabilities and mixing in stratified cylindrical shells. Phys. Fluids 17 (9), 094105.CrossRefGoogle Scholar
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, 55075519.CrossRefGoogle Scholar
Milovich, J.L., Amendt, P., Marinak, M. & Robey, H. 2004 Multimode short-wavelength perturbation growth studies for the National Ignition Facility double-shell ignition target designs. Phys. Plasmas 11 (4), 15521568.CrossRefGoogle Scholar
Morgan, R.V., Cabot, W.H., Greenough, J.A. & Jacobs, J.W. 2018 Rarefaction-driven Rayleigh–Taylor instability. Part 2. Experiments and simulations in the nonlinear regime. J. Fluid Mech. 838, 320355.CrossRefGoogle Scholar
Morgan, R.V., Likhachev, O.A. & Jacobs, J.W. 2016 Rarefaction-driven Rayleigh–Taylor instability. Part 1. Diffuse-interface linear stability measurements and theory. J. Fluid Mech. 791, 3460.CrossRefGoogle Scholar
Niederhaus, J.H.J., Greenough, J.A., Oakley, J.G., Ranjan, D., Anderson, M.H. & Bonazza, R. 2008 A computational parameter study for the three-dimensional shock-bubble interaction. J. Fluid Mech. 594, 85124.CrossRefGoogle Scholar
Orlicz, G.C., Balakumar, B.J., Tomkins, C.D. & Prestridge, K.P. 2009 A Mach number study of the Richtmyer–Meshkov instability in a varicose, heavy-gas curtain. Phys. Fluids 21 (6), 064102.CrossRefGoogle Scholar
Orlicz, G.C., Balasubramanian, S. & Prestridge, K.P. 2013 Incident shock mach number effects on Richtmyer–Meshkov mixing in a heavy gas layer. Phys. Fluids 25 (11), 114101.CrossRefGoogle Scholar
Ott, E. 1972 Nonlinear evolution of the Rayleigh–Taylor instability of a thin layer. Phys. Rev. Lett. 29 (21), 1429.CrossRefGoogle Scholar
Ouellet, F., Rollin, B., Durant, B., Koneru, R.B. & Balachandar, S. 2022 Shock-driven dispersal of a corrugated finite-thickness particle layer. Phys. Fluids 34 (8), 083301.CrossRefGoogle Scholar
Prestridge, K., Vorobieff, P., Rightley, P.M. & Benjamin, R.F. 2000 Validation of an instability growth model using Particle Image Velocimtery measurement. Phys. Rev. Lett. 84, 43534356.CrossRefGoogle Scholar
Qiao, X. & Lan, K. 2021 Novel target designs to mitigate hydrodynamic instabilities growth in inertial confinement fusion. Phys. Rev. Lett. 126 (18), 185001.CrossRefGoogle ScholarPubMed
Quirk, J.J. & Karni, S. 1996 On the dynamics of a shock-bubble interaction. J. Fluid Mech. 318, 129163.CrossRefGoogle Scholar
Rayleigh, Lord 1883 Investigation of the character of the equilibrium of an incompressible heavy fluid of variable density. Proc. Lond. Math. Soc. 14, 170177.Google Scholar
Regan, S.P., et al. 2004 Dependence of shell mix on feedthrough in direct drive inertial confinement fusion. Phys. Rev. Lett. 92 (18), 185002.CrossRefGoogle ScholarPubMed
Richtmyer, R.D. 1960 Taylor instability in shock acceleration of compressible fluids. Commun. Pure Appl. Maths 13, 297319.CrossRefGoogle Scholar
Rightley, P.M., Vorobieff, P. & Benjamin, R.F. 1997 Evolution of a shock-accelerated thin fluid layer. Phys. Fluids 9 (6), 17701782.CrossRefGoogle Scholar
Sembian, S., Liverts, M. & Apazidis, N. 2018 Plane blast wave interaction with an elongated straight and inclined heat-generated inhomogeneity. J. Fluid Mech. 851, 245267.CrossRefGoogle Scholar
Sharp, D.H. 1984 An overview of Rayleigh–Taylor instability. Physica D 12 (1), 318.CrossRefGoogle Scholar
Shen, H. & Parsani, M. 2017 The role of multidimensional instabilities in direct initiation of gaseous detonations in free space. J. Fluid Mech. 813, R4.CrossRefGoogle Scholar
Shen, H. & Wen, C.Y. 2016 A characteristic space-time conservation element and solution element method for conservation laws. II. Multidimensional extension. J. Comput. Phys. 305 (C), 775792.CrossRefGoogle Scholar
Shen, H., Wen, C.Y., Liu, K.X. & Zhang, D.L. 2015 a Robust high-order space-time conservative schemes for solving conservation laws on hybrid meshes. J. Comput. Phys. 281, 375402.CrossRefGoogle Scholar
Shen, H., Wen, C.Y., Parsani, M. & Shu, C.W. 2017 Maximum-principle-satisfying space-time conservation element and solution element scheme applied to compressible multifluids. J. Comput. Phys. 330, 668692.CrossRefGoogle Scholar
Shen, H., Wen, C.Y. & Zhang, D.L. 2015 b A characteristic space-time conservation element and solution element method for conservation laws. J. Comput. Phys. 288, 101118.CrossRefGoogle Scholar
Shigemori, K., Azechi, H., Nakai, M., Endo, T., Nagaya, T. & Yamanaka, T. 2002 Perturbation transfer from the front to rear surface of laser-irradiated targets. Phys. Rev. E 65 (4), 045401.CrossRefGoogle ScholarPubMed
Shyue, K.M. 1998 An efficient shock-capturing algorithm for compressible multicomponent problems. J. Comput. Phys. 142 (1), 208242.CrossRefGoogle Scholar
Simakov, A.N., Wilson, D.C., Yi, S.A., Kline, J.L., Clark, D.S., Milovich, J.L., Salmonson, J.D. & Batha, S.H. 2014 Optimized beryllium target design for indirectly driven inertial confinement fusion experiments on the national ignition facility. Phys. Plasmas 21 (2), 022701.CrossRefGoogle Scholar
Taylor, G. 1950 The instability of liquid surfaces when accelerated in a direction perpendicular to their planes. I. Proc. R. Soc. Lond. A 201 (1065), 192196.Google Scholar
Tomkins, C.D., Balakumar, B.J., Orlicz, G., Prestridge, K.P. & Ristorcelli, J.R. 2013 Evolution of the density self-correlation in developing Richtmyer–Meshkov turbulence. J. Fluid Mech. 735, 288306.CrossRefGoogle Scholar
Tomkins, C., Kumar, S., Orlicz, G. & Prestridge, K. 2008 An experimental investigation of mixing mechanisms in shock-accelerated flow. J. Fluid Mech. 611, 131150.CrossRefGoogle Scholar
Toro, E.F., Spruce, M. & Speares, W. 1994 Restoration of the contact surface in the HLL-Riemann solver. Shock Waves 4 (1), 2534.CrossRefGoogle Scholar
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.CrossRefGoogle Scholar
Wang, L., Guo, H., Wu, J., Ye, W., Liu, J., Zhang, W. & He, X. 2014 Weakly nonlinear Rayleigh–Taylor instability of a finite-thickness fluid layer. Phys. Plasmas 21 (12), 122710.Google Scholar
Wang, R., Song, Y., Ma, Z., Ma, D., Wang, L. & Wang, P. 2022 The transition to turbulence in rarefaction-driven Rayleigh–Taylor mixing: effects of diffuse interface. Phys. Fluids 34 (1), 015125.CrossRefGoogle Scholar
Weir, S.T., Chandler, E.A. & Goodwin, B.T. 1998 Rayleigh–Taylor instability experiments examining feedthrough growth in an incompressible, convergent geometry. Phys. Rev. Lett. 80 (17), 3763.CrossRefGoogle Scholar
Wouchuk, J.G. & Nishihara, K. 1996 Linear perturbation growth at a shocked interface. Phys. Plasmas 3 (10), 37613776.CrossRefGoogle Scholar
Wouchuk, J.G. & Nishihara, K. 1997 Asymptotic growth in the linear Richtmyer–Meshkov instability. Phys. Plasmas 4 (4), 10281038.CrossRefGoogle Scholar
Wouchuk, J.G. & Nishihara, K. 2004 Normal velocity freeze-out of the Richtmyer–Meshkov instability when a shock is reflected. Phys. Rev. E 70 (2), 026305.CrossRefGoogle ScholarPubMed
Wouchuk, J.G. & Sano, T. 2015 Normal velocity freeze-out of the Richtmyer–Meshkov instability when a rarefaction is reflected. Phys. Rev. E 91 (2), 023005.CrossRefGoogle ScholarPubMed
Zhai, Z., Li, W., Si, T., Luo, X., Yang, J. & Lu, X. 2017 Refraction of cylindrical converging shock wave at an air/helium gaseous interface. Phys. Fluids 29 (1), 016102.CrossRefGoogle Scholar
Zhai, Z., Liang, Y., Liu, L., Ding, J., Luo, X. & Zou, L. 2018 a Interaction of rippled shock wave with flat fast-slow interface. Phys. Fluids 30 (4), 046104.CrossRefGoogle Scholar
Zhai, Z., Zou, L., Wu, Q. & Luo, X. 2018 b Review of experimental Richtmyer–Meshkov instability in shock tube: from simple to complex. Proc. Inst. Mech. Engng 232, 28302849.Google Scholar
Zhang, Q. & Guo, W. 2016 Universality of finger growth in two-dimensional Rayleigh–Taylor and Richtmyer–Meshkov instabilities with all density ratios. J. Fluid Mech. 786, 4761.CrossRefGoogle Scholar
Zhou, Y. 2017 a Rayleigh–Taylor and Richtmyer–Meshkov instability induced flow, turbulence, and mixing. I. Phys. Rep. 720–722, 1136.Google Scholar
Zhou, Y. 2017 b Rayleigh–Taylor and Richtmyer–Meshkov instability induced flow, turbulence, and mixing. II. Phys. Rep. 723–725, 1160.Google Scholar
Zhou, Y., et al. 2021 Rayleigh–Taylor and Richtmyer–Meshkov instabilities: a journey through scales. Physica D 423, 132838.CrossRefGoogle Scholar
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.CrossRefGoogle Scholar
Zou, L., Al-Marouf, M., Cheng, W., Samtaney, R., Ding, J. & Luo, X. 2019 Richtmyer–Meshkov instability of an unperturbed interface subjected to a diffracted convergent shock. J. Fluid Mech. 879, 448467.CrossRefGoogle Scholar
Figure 0

Figure 1. Schematics of (a) the soap film interface generation, (b) the shock-tube and schlieren photography, and (c) the initial configuration of two successive slow/fast interfaces, where $L_0$ denotes the initial distance between the two interfaces, II$_1$ denotes the initial upstream interface and II$_2$ denotes the initial downstream interface.

Figure 1

Table 1. Initial physical parameters of two successive slow/fast interfaces, where $L_0$ denotes the initial distance between the two interfaces; $MF$ denotes the mass fraction of SF$_6$ in gas B; $\rho _{B}$ denotes the density of gas B; $R_1$ and $R_3$ denote the density ratios of the fluids on the two sides of the upstream interface and downstream interface, respectively; and $A_1$ and $A_2$ denote the Atwood numbers of the upstream interface and downstream interface, respectively.

Figure 2

Figure 2. Schematics of the initial simulation setup, where $a_1$ and $a_2$ represent the amplitudes of the upstream interface and downstream interface, respectively.

Figure 3

Figure 3. Code validation based on (a) the time-varying displacements of the shocked upstream interface (SI$_1$) and downstream interface (SI$_2$) in the L10-1D case, and (b) the time-varying amplitudes of the SI$_1$ and SI$_2$ in the L10-AP case, where $n=1$ for the upstream interface and $n=2$ for the downstream interface in this study. Solid (or hollow) symbols represent the experimental results for the SI$_1$ (or SI$_2$), and dash–dot (or dashed) lines represent the numerical results for the SI$_1$ (or SI$_2$), considering various mesh sizes.

Figure 4

Figure 4. Schlieren images of evolutions of two successive, quasi-1-D slow/fast interfaces driven by a shock wave in cases (a) L10-1D, (b) L30-1D and (c) L50-1D, where SI$_1$ (or SI$_2$) denotes the shocked upstream (or downstream) interface, RW$_1$ (or TS$_1$) denotes the reflected rarefaction waves (or transmitted shock) after the IS impacts II$_1$, rRW$_2$ (or TS$_2$) denotes the reflected rarefaction waves (or transmitted shock) after the TS$_1$ impacts II$_2$, and tRW$_1$ denotes the transmitted rarefaction waves after the rRW$_2$ impacts the SI$_1$. Numbers indicate time in $\mathrm {\mu }$s.

Figure 5

Figure 5. The dimensionless (a) displacements and (b) velocities of the upstream interface (solid symbols) and the downstream interface (hollow symbols). Black, yellow and orange solid (or dashed) lines represent the 1-D theory predictions for the upstream (or downstream) interface movements in Stages I, II and III, respectively. The vertical dash–dot lines represent the specific times calculated with (3.1) and (3.9a,b).

Figure 6

Table 2. Physical parameters of the waves and interfaces in the laboratory reference, where $v_{{t1}}$ denotes the velocity of the TS$_1$; $v_{{rRW}_2}$ (or $v_{{rRW_1}}$) denotes the velocity of the rRW$_2$ head (or rRW$_1$ head); $v_{1}^\alpha$ (or $v_{2}^\alpha$) denotes the velocity of the SI$_1$ (or SI$_2$) in Stage I; $v_{1}^\beta$ (or $v_{2}^\beta$) denotes the velocity of the SI$_1$ (or SI$_2$) in Stage II; $v_{fn}$ denotes the asymptotic velocity of the two interfaces; $t_{2}^\alpha$ denotes the specific time when the TS$_1$ impacts the II$_2$; $t_{1}^\beta$ denotes the time when the rRW$_2$ head impacts the SI$_1$ and $t_{2}^\beta$ denotes the time when the rRW$_1$ impacts the SI$_2$; $t_1^\sigma$ (or $t_2^\sigma$) represents the time when the rRW$_2$ (or rRW$_1$) leaves the SI$_1$ (or SI$_2$). The units for velocity and time are m s$^{-1}$ and $\mathrm {\mu }$s, respectively.

Figure 7

Figure 6. Sketches of (a) the interaction of the TS$_1$ and II$_2$, (b) the interaction of the rRW$_2$ and SI$_1$, (c) the interaction of the rRW$_1$ and SI$_2$, and (d) the movements of the SI$_1$ and SI$_2$ after all waves are refracted away, where $\gamma _A$, $\gamma _B$ and $\gamma _C$ denote the specific heat ratios of gas A, gas B and gas C, respectively; $c_A$ denotes the sound speed of gas A after the IS impacts the II$_1$, $c_B^\alpha$ denotes the sound speed of gas B between the TS$_1$ with the II$_2$, $c_B^\beta$ denotes the sound speed of gas B between the rRW$_2$ tail and the SI$_2$, $c_B^\sigma$ denotes the sound speed of gas B between the rRW$_1$ tail and the SI$_1$, $c_C$ denotes the sound speed of gas C behind TS$_2$; $p_1$ denotes the pressure of gas B behind the TS$_1$, $p_2$ denotes the pressure of gas B behind the rRW$_2$ tail, $p_3$ denotes the pressure of gas B behind the rRW$_1$ tail, and $p_4$ denotes the pressure of gas B after the interaction of the rRW$_1$ and SI$_2$. (a) $t=t_2^\alpha$, (b) $t=t_1^\beta$, (c) $t=t_2^\beta$, (d) $t>t_2^\beta$.

Figure 8

Table 3. Physical parameters of the reverberating rarefaction waves, where $L_{rRW_2}$ denotes the length of the rRW$_2$ at $t_1^\beta$, $T_{{rRW_2}}$ denotes the interaction time of the rRW$_2$ and the SI$_1$, $\bar {g}_{{rRW_2}}$ denotes the average acceleration imposed on the SI$_1$ by the rRW$_2$, $L_{rRW_1}$ denotes the length of the rRW$_1$ at $t_2^\beta$, $T_{{rRW_1}}$ denotes the interaction time of the rRW$_1$ and the SI$_2$, and $\bar {g}_{{rRW_1}}$ denotes the average deceleration imposed on the SI$_2$ by the rRW$_1$. The units for length, time and acceleration are mm, $\mathrm {\mu }$s and $10^6\ \textrm {m}\ \textrm {s}^{-2}$, respectively.

Figure 9

Table 4. The movements of the shocked two successive slow/fast interfaces in different stages, where ‘Uni.’ is short for ‘Uniform movement’, ‘Acc.’ is short for ‘Acceleration movement’ and ‘Dec.’ is short for ‘Deceleration movement’.

Figure 10

Figure 7. Schlieren images of evolutions of quasi-2-D two successive slow/fast interfaces driven by a shock wave acquired from experiments (upper) and simulations (below) in cases (a) L10-IP, (b) L10-AP, (c) L30-IP, (d) L30-AP, (e) L50-IP and (f) L50-AP. The upstream interface in panel (b) is highlighted with a white dashed line. Although the upstream interface in panel (b) is a slow/fast one, it does not experience phase reversal. Numbers denote time in $\mathrm {\mu }$s.

Figure 11

Table 5. The linear amplitude growth rates of the two interfaces, where $\dot {a}_1^{{exp}}$, $\dot {a}_1^{{num}}$ and $\dot {a}_1^{{lin}}$ denote the experimental, numerical and theoretical results of the upstream interface, respectively; $\dot {a}_2^{{exp}}$, $\dot {a}_2^{{num}}$ and $\dot {a}_2^{{lin}}$ denote the experimental, numerical and theoretical results of the downstream interface, respectively; and $\psi _1$ and $\psi _2$ denote the feedthrough effect on the RMI of the upstream and downstream interfaces, respectively.

Figure 12

Figure 8. The values of $\psi$ versus $kh_0$ under different $\kappa$ conditions when (a) $\theta =0$ and (b) $\theta ={\rm \pi}$. Hollow circles represent the instability freeze-out of the upstream or downstream interface.

Figure 13

Figure 9. The time-varying dimensionless amplitudes of the (a) upstream interface and (b) downstream interface obtained from experiments.

Figure 14

Table 6. The end times of the upstream (or downstream) interface's phase reversal $t_1^{{pr}}$ (or $t_2^{{pr}}$) calculated with (4.11a,b). The unit of time is $\mathrm {\mu }$s.

Figure 15

Figure 10. Sketches of (a) the interaction of the rRW$_2$ and SI$_1$ in all cases (except for the L10-AP case), (b) the interaction of the rRW$_2$ and SI$_1$ in the L10-AP case, (c) the interaction of the rRW$_1$ and SI$_2$ under $L_0=10.0$ and 30.0 mm conditions, and (d) the interaction of the rRW$_1$ and SI$_2$ under $L_0=50.0$ mm conditions. The blue (or orange) arcs with an arrow in panels (a,b) illustrate the vorticity deposited by the IS (or rRW$_2$) on the upstream interface. The blue (or orange) arcs with an arrow in panels (c,d) represent the vorticity deposited by the TS$_1$ (or rRW$_1$) on the downstream interface. The purple (or green) arrows represent the pressure (or density) gradient $\boldsymbol {\nabla } p$ (or $\boldsymbol {\nabla }\rho$).

Figure 16

Figure 11. Comparisons of the amplitudes of the upstream interface and downstream interface measured from experiments and simulations with theories in cases (a) L10-IP, (b) L10-AP, (c) L30-IP, (d) L30-AP, (e) L50-IP and (f) L50-AP. The solid and hollow symbols represent the experimental results for the upstream interface and downstream interface, respectively; the solid red and dashed green lines represent the simulation results for the upstream interface and downstream interface, respectively; and the solid blue and dashed orange lines represent the theoretical predictions for the upstream interface and downstream interface, respectively.

Figure 17

Figure 12. The time-varying amplitudes of the upstream interface and downstream interface in cases (a) F-1 and (b) F-2. The solid and hollow symbols represent the numerical results for the upstream interface and downstream interface, respectively; the solid red and dashed blue lines represent the theoretical predictions for the upstream interface and downstream interface, respectively, and similarly hereinafter.

Figure 18

Figure 13. The time-varying amplitudes of the upstream and downstream interfaces in case F-3.

Figure 19

Figure 14. The values of $\psi$ versus $kh_0$ under different $\theta$ conditions when (a) $\kappa =0.1$ and (b) $\kappa =10.0$.