Hostname: page-component-788cddb947-jbkpb Total loading time: 0 Render date: 2024-10-07T17:21:20.444Z Has data issue: false hasContentIssue false

Mode coupling between two different interfaces of a gas layer subject to a shock

Published online by Cambridge University Press:  03 April 2024

Chenren Chen
Affiliation:
Advanced Propulsion Laboratory, Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, PR China
He Wang
Affiliation:
Advanced Propulsion Laboratory, Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, PR China
Zhigang Zhai*
Affiliation:
Advanced Propulsion Laboratory, Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, PR China
Xisheng Luo*
Affiliation:
Advanced Propulsion Laboratory, Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, PR China State Key Laboratory of High Temperature Gas Dynamics, Institute of Mechanics, Chinese Academy of Sciences, Beijing 100190, PR China
*
Email addresses for correspondence: [email protected], [email protected]
Email addresses for correspondence: [email protected], [email protected]

Abstract

Shock-tube experiments and theoretical studies have been performed to highlight mode-coupling in an air–SF$_6$–air fluid layer. Initially, the two interfaces of the layer are designed as single mode with different basic modes. It is found that as the two perturbed interfaces become closer, interface coupling induces a different mode from the basic mode on each interface. Then mode coupling further generates new modes. Based on the linear model (Jacobs et al., J. Fluid Mech., vol. 295, 1995, pp. 23–42), a modified model is established by considering the different accelerations of two interfaces and the waves’ effects in the layer, and provides good predictions to the linear growth rates of the basic modes and the modes generated by interface coupling. It is observed that interface coupling behaves differently to the nonlinear growth of the basic modes, which can be characterized generally by the existing or modified nonlinear model. Moreover, a new modal model is established to quantify the mode-coupling effect in the layer. The mode-coupling effect on the amplitude growth is negligible for the basic modes, but is significant for the interface-coupling modes when the initial wavenumber of one interface is twice the wavenumber of the other interface. Finally, amplitude freeze-out of the second single-mode interface is achieved theoretically and experimentally through interface coupling. These findings may be helpful for designing the target in inertial confinement fusion to suppress the hydrodynamic instabilities.

Type
JFM Papers
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

Richtmyer–Meshkov (RM) instability occurs when a perturbed interface separating two different fluids is impacted by a shock wave (Richtmyer Reference Richtmyer1960; Meshkov Reference Meshkov1969). The development of the RM instability is driven mainly by baroclinic vorticity produced by misalignment of the pressure and density gradients. The RM instability plays essential roles in many applications, such as inertial confinement fusion (ICF) (Lindl et al. Reference Lindl, Landen, Edwards, Moses and Team2014; Qiao & Lan Reference Qiao and Lan2021; Zhou et al. Reference Zhou2021), supersonic combustion (Yang, Kubota & Zukoski Reference Yang, Kubota and Zukoski1993) and supernova explosions (Kuranz et al. Reference Kuranz2018). Over the past decades, several reviews have been published to conclude the achievements in the RM instability study (Zabusky Reference Zabusky1999; Brouillette Reference Brouillette2002; Ranjan, Oakley & Bonazza Reference Ranjan, Oakley and Bonazza2011; Zhou Reference Zhou2017a,Reference Zhoub; Zhai et al. Reference Zhai, Zou, Wu and Luo2018).

For an initial single-mode interface, mode coupling can be ignored until the harmonics grow significantly (Zhou Reference Zhou2017a; Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a). A perturbation expansion model with fourth-order accuracy (ZS model) was developed by Zhang & Sohn (Reference Zhang and Sohn1997) to calculate the amplitudes of high-order harmonics. The perturbation expansion method was further simplified by Vandenboomgaerde, Gauthier & Mügler (Reference Vandenboomgaerde, Gauthier and Mügler2002) to obtain higher-order expressions of the harmonics than that in the ZS model. When the initial perturbation is multi-mode, the different basic modes interact with each other, and mode coupling causes the generation of new modes at the early stage (Brouillette Reference Brouillette2002; Zhou Reference Zhou2017b). Mode coupling for a multi-mode interface has been investigated widely (Rikanati, Alon & Shvarts Reference Rikanati, Alon and Shvarts1998; Di Stefano et al. Reference Di Stefano, Malamud, Kuranz, Klein, Stoeckl and Drake2015; McFarland et al. Reference McFarland, Reilly, Black, Greenough and Ranjan2015; Mohaghar et al. Reference Mohaghar, Carter, Pathikonda and Ranjan2019). Theoretically, Haan (Reference Haan1991) proposed a modal model (Haan model) with second-order accuracy to quantify the mode-coupling effect on the Rayleigh–Taylor (RT) instability development (Rayleigh Reference Rayleigh1883; Taylor Reference Taylor1950). Then the applicability range of the Haan model was extended by Ofer et al. (Reference Ofer, Alon, Shvarts, McCrory and Verdon1996). These two models were reformulated recently to apply them to the multi-mode RM instability (Liang et al. Reference Liang, Liu, Zhai, Ding, Si and Luo2021). Experimentally, the mode coupling on a single-interface development has been explored extensively through shock-tube experiments (Luo et al. Reference Luo, Liu, Liang, Ding and Wen2020; Liang et al. Reference Liang, Liu, Zhai, Ding, Si and Luo2021; Xu et al. Reference Xu, Wang, Zhai and Luo2023), and several models have been verified.

In many applications, the RM instability grows on the interfaces of a fluid layer. For example, the RM instability in ICF occurs on both the ablator front and the interface between the ablator and deuterium–tritium ice (Betti & Hurricane Reference Betti and Hurricane2016; Qiao & Lan Reference Qiao and Lan2021). For the RM instability of a fluid layer, most previous studies considered the perturbation of only one mode, i.e. both single-mode interfaces of the fluid layer have the same mode (Jacobs et al. Reference Jacobs, Jenkins, Klein and Benjamin1995; Mikaelian Reference Mikaelian1995; Liang & Luo Reference Liang and Luo2021). Theoretically, a linear analytic theory for the RM instability in a perturbed fluid layer with only one wavenumber was first proposed by Mikaelian (Reference Mikaelian1985, Reference Mikaelian1995), and later verified by numerical simulations (Mikaelian Reference Mikaelian1996). Through solving velocity potential functions of a heavy layer, Jacobs et al. (Reference Jacobs, Jenkins, Klein and Benjamin1995) deduced a linear model (the J-model hereafter) for predicting the perturbation growth rates of both interfaces, and a vortex model for predicting the width growth of the fluid layer. The J-model was modified by Liang & Luo (Reference Liang and Luo2021) to predict the perturbation linear growth rates of a heavy layer. Recently, the effect of the finite thickness of two superimposed fluids on harmonics in the RM instability for arbitrary Atwood numbers was also investigated (Liu et al. Reference Liu, Li, Yu, Fu, Wang, Wang and Ye2018b). Experimentally, the gas curtain technique was used by Jacobs et al. (Reference Jacobs, Klein, Jenkins and Benjamin1993) to produce a thin SF$_6$ layer in air, and the shocked layer evolution was investigated. Based on the gas curtain technique, the width of the shocked gas layer was measured (Jacobs et al. Reference Jacobs, Jenkins, Klein and Benjamin1995), and the circulation in the curtain was calculated (Prestridge et al. Reference Prestridge, Vorobieff, Rightley and Benjamin2000) to verify the vortex model. Recently, the soap-film technique was used widely in creating a gas layer, and extensive studies have been conducted (Liang & Luo Reference Liang and Luo2021, Reference Liang and Luo2023a,Reference Liang and Luob; Zhang et al. Reference Zhang, Ding, Si and Luo2023), focusing on the effects of initial amplitude, width of the gas layer, and perturbation phase on a fluid layer development. When the two interfaces of a fluid layer have the same mode, mode coupling is absent until nonlinearity is significant (Liang & Luo Reference Liang and Luo2021; Liang Reference Liang2022). In real scenarios, however, the interfaces of the fluid layer often have different basic modes. The previous work always considered coupling between modes on the same interface: whether and how the modes on the different interfaces couple; and how to quantify mode coupling between different interfaces. Due to difficulties in forming such a fluid layer with the specific modes on interfaces in experiments and more complicated mechanisms involved, the related studies are rare and these questions remain unclear.

In this work, the soap-film technique is used to create the fluid layer, in which two different modes are present on the two initial interfaces. Seven kinds of disturbed SF$_6$ gas layer with different mode combinations are considered. Through one-dimensional theory and experiments on the undisturbed layer, the wave patterns and the concentration of SF$_6$ inside the layer are obtained. The existing linear and nonlinear models are modified to describe the perturbation evolution of basic modes, and a new modal model is established to quantify the mode-coupling effect. Finally, freeze-out of the amplitude growth of the second single-mode interface is discussed.

2. Experimental methods

The soap-film technique (Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a; Liang et al. Reference Liang, Liu, Zhai, Ding, Si and Luo2021) is used to generate the well-defined discontinuous interfaces of the fluid layer. As shown in figure 1(a), there are three devices, with width 140 mm and height 6 mm. The soap-film interfaces are first generated at both sides of the middle device. Then SF$_6$ is pumped into the middle device through the inflow hole, and air is discharged through the outflow hole. Once the SF$_6$ concentration satisfies the requirement, the inflation stops. Then the three devices are connected, and an SF$_6$ gas layer with two interfaces is formed. For different runs, the inflation rate and duration are the same to ensure a similar concentration of SF$_6$ in the gas layer as far as possible. The concentration of SF$_6$ is determined by comparing the velocities of the incident shock and transmitted shock, and the interface velocity jump, measured from experiments with those predicted by one-dimensional theory.

Figure 1. Schematics of (a) the soap-film interface generation, and (b) the initial interface configuration studied. Here, IS is the incident shock, II$_1$ is the first interface (air–SF$_6$), II$_2$ is the second interface (SF$_6$–air), and $L_0$ is the initial width of the SF$_6$ layer.

The schematic of the initial interface configuration is shown in figure 1(b). The two single-mode interfaces with different wavenumbers are considered. Seven kinds of fluid layer with different wavenumber combinations and relative phase, as shown in table 1, are studied. In this work, the initial phase of the perturbation on the second interface (II$_2$) is defined as positive. In the first six cases, the phase of the first interface (II$_1$) is positive, while the phase of II$_1$ in the seventh case is negative. A dimensionless number $K=k_1/k_2$ (where $k_1$ and $k_2$ are the wavenumbers of II$_1$ and II$_2$) is defined to represent different mode-coupling effects. Here, $k_1a_1^-$ and $k_2a_2^-$ (where $a_1^-$ and $a_2^-$ are the initial amplitudes of II$_1$ and II$_2$, respectively) are all close to 0.1, and the tiny differences among cases are caused by three-dimensionality of the soap-film interface (Wang et al. Reference Wang, Cao, Chen, Zhai and Luo2022). To highlight the mode-coupling and interface-coupling effects, the initial width of the gas layer ($L_0$, defined as the distance between the balanced positions of two interfaces) is fixed as 20 mm. The choice of $L_0$ makes the interface-coupling effect strong enough but avoids the coalescing of two interfaces (Liang & Luo Reference Liang and Luo2021).

Table 1. Interface parameters of perturbed cases: $\lambda _1$ and $\lambda _2$ ($k_1$ and $k_2$, $a_1^-$ and $a_2^-$) are the wavelengths (wavenumbers, initial amplitudes) of II$_1$ and II$_2$, respectively, and $K=k_1/k_2$. Here, ‘IP’ (‘AP’) means the perturbations on the two interfaces have the same (opposite) phase.

The experiments are conducted in a shock tube (Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a; Guo et al. Reference Guo, Cong, Si and Luo2022), and the incident shock Mach number is $1.31\pm 0.01$. The flow field is recorded by high-speed schlieren photography. The frame rate of the camera (FASTCAM SA-Z, Photron Ltd) is 50 400 frames per second, and the exposure time is $0.78\,\mathrm {\mu }{\rm s}$. The spatial resolution is ${\sim }0.26\,{\rm mm}\,{\rm pixel}^{-1}$. The ambient pressure and temperature are $101.3\pm 0.1\,{\rm kPa}$ and $295\pm 1.5\,{\rm K}$, respectively. At least three successful experimental runs are performed for each case to ensure the experimental repeatability. The boundary layer effect has already been discussed (Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a), and is ignored here.

3. Results and discussion

3.1. Undisturbed layer

The schlieren images of the undisturbed SF$_6$ layer after the shock impact are presented in figure 2(a). The initial time is defined as the moment when the incident shock (IS) meets the average position of the first initial interface (II$_1$). After IS impacts II$_1$, the first transmitted shock wave (TS$_1$) moving downstream in SF$_6$, the reflected shock (RS) moving upstream in air, and a shocked air–SF$_6$ interface (SI$_1$) are formed ($39\,\mathrm {\mu }{\rm s}$). Then TS$_1$ impacts the second initial interface (II$_2$), which produces a shocked SF$_6$–air interface (SI$_2$), the second transmitted shock wave (TS$_2$) moving downstream in air ($277\,\mathrm {\mu }{\rm s}$), and reflected rarefaction waves (RW) moving upstream in SF$_6$. These RW will interact with SI$_1$, reflecting compression waves (CW) that subsequently interact with SI$_2$. Both RW and CW cannot be observed in the schlieren images of our experiments due to their weak intensity. As studied by Liang & Luo (Reference Liang and Luo2021), RW and CW significantly affect the movements of SI$_1$ and SI$_2$, respectively. During the experimental time studied, after CW interact with SI$_2$, the two undisturbed interfaces move parallel, and no obvious disturbances are generated on the interfaces.

Figure 2. (a) Schlieren images of a planar SF$_6$ layer accelerated by a planar shock, and (b) the trajectories of the interfaces and waves. Here, II$_1$ and II$_2$ (SI$_1$ and SI$_2$) are the first and second initial (shocked) interfaces respectively; TS$_1$ and TS$_2$ are the transmitted shock waves in SF$_6$ and in air, respectively; RS is the reflected shock; RW is reflected rarefaction waves; CW is compression waves; and $x_0$ is the initial position of II$_1$.

The trajectories of shock waves and interfaces extracted from experiment and predicted by one-dimensional theory are compared in figure 2(b), and their velocities are provided in table 2. After CW impact SI$_2$, the two shocked interfaces move steadily, and the effects of other reflected waves can be ignored. Due to the similar initial conditions for different cases, the parameters in table 2 are considered as the initial parameters in the following experiments. In this work, the volume fraction of SF$_6$ in the gas layer is $75\,\%$. The Atwood numbers of pre-shock ($A^-$) and post-shock ($A^+$) SF$_6$ layer ($A$ is defined as $(\rho _1-\rho _2)/(\rho _1+\rho _2)$, with $\rho _1$ and $\rho _2$ being the densities of gases inside and outside the layer, respectively) equal 0.60 and 0.64, respectively. Here, three specific times are defined to facilitate subsequent calculations: $t_1$ denotes the time when TS$_1$ impacts II$_2$ and RW are initially generated at $t=t_1$; $t_{{R}}$ and $t_{{C}}$ denote the times when RW meet SI$_1$ and CW meet SI$_2$, respectively. Here, $t_1$, $t_{{R}}$ and $t_{{C}}$ can be calculated as

(3.1)\begin{equation} \left. \begin{aligned} t_1 & =\frac{L_0}{v_2}, \\ t_{{R}} & =t_1+\frac{L_0\left(1-\dfrac{\Delta u_1}{v_2}\right)}{c_1}, \\ t_{{C}} & =t_{{R}}+\frac{L_0\left(1-\dfrac{\Delta u_1}{v_2}\right) \left(1+\dfrac{\Delta u_2-\Delta u_1}{c_1}\right)}{c_2}, \end{aligned} \right\} \end{equation}

where $v_2$ is the velocity of TS$_1$, $\Delta u_1$ and $\Delta u_2$ are the jump velocities of II$_1$ and II$_2$ induced by shocks, respectively, $c_1$ is the sound speed of the gas between SI$_1$ and the RW front, and $c_2$ is the sound speed of the gas between SI$_2$ and the RW tail.

Table 2. Theoretical parameters in undisturbed case: $M_{{s}}$ is the Mach number of IS; $\Delta u_1$, $\Delta u_2$, $\Delta u_{{R}}$ and $\Delta u_{{C}}$ are the jump velocities of II$_1$ induced by IS, II$_2$ induced by TS$_1$, II$_1$ induced by RW, and II$_2$ induced by CW, respectively; $v_1$ and $v_2$ are the velocities of IS and TS$_1$, respectively; $t_1$, $t_{{R}}$ and $t_{{C}}$ denote the times when TS$_1$ impacts II$_2$, RW impact SI$_1$, and CW impact SI$_2$, respectively; $c_1$ ($c_2$) is the sound speed of the gas between SI$_1$ and the RW front (between SI$_2$ and the RW tail). The units of velocity and time are m s$^{-1}$ and $\mathrm {\mu }{\rm s}$.

3.2. Perturbed layers

3.2.1. Interface morphology

Schlieren images of the developments of the shock-impacted fluid layer for seven cases are shown in figure 3. The wave–interface interaction processes are similar to those in the undisturbed case, and the relevant descriptions are omitted here. Taking case 30-20-IP as an example, when CW impact SI$_2$, the phase inversion process of the perturbation on SI$_2$ has been completed, and TS$_2$ is away from SI$_2$ ($298\,\mathrm {\mu }\textrm {s}$). Just after CW impact SI$_2$, the perturbations on two interfaces can still be considered as single mode. Later, the multi-mode perturbations grow on SI$_1$ and SI$_2$ ($695\,\mathrm {\mu }\textrm {s}$). Finally, spikes and bubbles arise symmetrically ($1091\,\mathrm {\mu }\textrm {s}$). The sizes and shapes of the bubble and spike on SI$_1$ and SI$_2$ differ significantly. From the schlieren images, the interface profile is clear enough to extract its contour, and can still be described as a single-valued function so that the fast Fourier transform can be applied to obtain the amplitude of each mode (Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a). The evolutions of the bubbles and spikes at the two interfaces are very sensitive to the value of $K$. When the first interface is disturbed at a higher mode than the second interface ($K>1$), the perturbation on SI$_2$ remains a quasi-single-mode profile, and the bubbles and spikes develop without inclination. However, the multi-mode perturbations on SI$_1$ appear more significant than those on SI$_2$, and the bubbles and spikes on SI$_1$ develop asymmetrically. Specifically, in case 20-60-IP, the middle bubble on SI$_1$ has a wider spread but a smaller amplitude than the side bubbles, and the spikes on SI$_1$ develop with an apparent inclination to the middle bubble. In case 30-60-IP, the middle spike has a greater scale than the side spikes, and the bubbles are inclined to the side spikes. Besides, vortices have already arisen on the side spikes, which indicate that nonlinearity occurs earlier on the side spike than the middle spike. In case 20-30-IP, the middle bubble has a narrower spread but a larger amplitude than the side bubbles, and the spikes are also inclined to the middle bubble. On the whole, if the bubble (spike) on SI$_1$ faces directly the spike (bubble) on SI$_2$, then it is deformed weakly. Otherwise, the bubble (spike) on SI$_1$ deforms greatly. When $K<1$, the first interface has a lower mode disturbance than the second one, and the multi-mode perturbations on SI$_2$ appear more significant compared with those on SI$_1$. In case 30-20-IP, the middle bubble on SI$_1$ has a larger amplitude than the side bubbles, and two spikes develop equally, with a flattened head. The middle spike on SI$_2$ has a larger amplitude and a narrower head compared with the side spikes, and two bubbles develop similarly. In cases 40-20-IP and 60-20-IP, a small middle spike on SI$_2$ occurs, and the shape of the side spikes on SI$_2$ is similar to the shape of the side bubbles on SI$_1$. For these two cases, the bubbles on SI$_1$ behave quite differently, which illustrates the importance of the wavenumber combination. In case 60-20-AP, in contrast, small side spikes occur on SI$_2$. In general, the shape of SI$_1$ is still dominated by its initial basic mode, while the shape of SI$_2$ obviously presents multi-mode.

Figure 3. Schlieren images of the shock-induced fluid layer evolution for different cases. The symbols have the same meanings as in figure 2. Numbers denote time in $\mathrm {\mu }\textrm {s}$.

3.2.2. Linear growth of amplitude

Previous works (Jacobs et al. Reference Jacobs, Jenkins, Klein and Benjamin1995; Mikaelian Reference Mikaelian1996) have revealed that the amplitudes of a shocked heavy layer grow linearly after the compression stage and start-up process. Considering a planar fluid layer (with thickness $L$) that is composed of a fluid of density $\rho _1$ being surrounded by a fluid of density $\rho _2$, Jacobs et al. (Reference Jacobs, Jenkins, Klein and Benjamin1995) proposed the J-model to predict the perturbation growth rates of the shocked two interfaces. When the fluids are incompressible and irrotational, the equations that govern the growth of infinitesimal disturbances on the interfaces of the layer can be expressed as

(3.2)\begin{equation} \nabla^2\phi=0, \end{equation}

where $\phi$ is the velocity potential of the fluid, and the kinematic and pressure conditions on the interfaces are

(3.3)\begin{equation} \left. \begin{aligned} & \frac{{\partial} \eta}{{\partial} {t}}-\frac{{\partial} \phi}{{\partial}{y}}=0,\\ & \rho\,\frac{{\partial} \phi}{{\partial}{t}}-\rho g\eta+p=P_0. \end{aligned} \right\} \end{equation}

Here, $p$ is the pressure, $g$ is the acceleration, $\eta$ is the perturbation shape on the interfaces, and $P_0$ is a constant equalling the pressure at the interface in the unperturbed rest state.

As shown in figure 4, the periodic perturbations set in the layer are of the form

(3.4)\begin{equation} \left. \begin{aligned} \eta_1 & =a_1(t)\cos(kx),\\ \eta_2 & =a_2(t)\cos(kx), \end{aligned} \right\} \end{equation}

where $a_1(t)$ and $a_2(t)$ are the amplitudes of the first and second interfaces, respectively. Then the velocity potentials in the three fluids can be obtained and expressed as

(3.5)\begin{equation} \left. \begin{aligned} \phi_1 & =b_1(t)\,{\rm e}^{{-}k(\kern0.7pt y-L/2)}\cos(kx), \\ \phi_2 & ={-}\left[\frac{b_1+b_2}{2}\,\frac{\sinh(ky)}{\cosh(kL/2)} +\frac{b_1-b_2}{2}\,\frac{\cosh(ky)}{\sinh(kL/2)}\right]\cos(kx),\\ \phi_3 & ={-}b_2(t)\,{\rm e}^{k(\kern0.7pt y+L/2)}\cos(kx), \end{aligned} \right\} \end{equation}

where $b_1$ and $b_2$ are the functions of time in the velocity potentials. Substituting (3.4) and (3.5) into (3.3) and (3.2), the J-model (Jacobs et al. Reference Jacobs, Jenkins, Klein and Benjamin1995) can then be deduced as

(3.6)\begin{equation} \left. \begin{aligned} \frac{{{\rm d}}^{2}a_1}{{{\rm d}}t^{2}}+\frac{{{\rm d}}^{2}a_2}{{{\rm d}}t^{2}} & =kA_{{t}}g(a_1-a_2),\\ \frac{{{\rm d}}^{2}a_1}{{{\rm d}}t^{2}}-\frac{{{\rm d}}^{2}a_2}{{{\rm d}}t^{2}} & =kA_{{c}}g(a_1+a_2), \end{aligned} \right\} \end{equation}

where

(3.7)\begin{equation} \left. \begin{aligned} A_{{t}} & =\frac{\rho_1-\rho_2}{\rho_1\tanh (kL/2)+\rho_2},\\ A_{{c}} & =\frac{\rho_1-\rho_2}{\rho_1\coth (kL/2)+\rho_2}, \end{aligned} \right\} \end{equation}

are the two modified Atwood numbers.

Figure 4. Schematic showing the configuration analysed using linear stability theory for the double-interface system, where $\rho _1$ and $\rho _2$ are the densities of the fluids inside and outside the layer, respectively, $\eta _1$ and $\eta _2$ represent the perturbation shapes of the two interfaces, $\phi _1$, $\phi _2$ and $\phi _3$ are the velocity potentials, $L$ is the initial layer thickness, and $g$ is the acceleration of the two interfaces.

In this work, $\rho _1=7.98\,\textrm {kg}\,\textrm {m}^{-3}$ ($\rho _2=1.78\,\textrm {kg}\,\textrm {m}^{-3}$) is the fluid density inside (outside) the layer after the reflected compression waves in the layer (CW) interact with the shocked second interface (SI$_2$). Here, $L$ is chosen as the distance between the two shocked interfaces (SI$_1$ and SI$_2$) after CW collide with SI$_2$, and can be expressed as

(3.8)\begin{equation} L=L_0\left(1-\frac{\Delta u_1}{v_2}\right)\left(1+\frac{\Delta u_2-\Delta u_1}{c_1}\right) \left(1-\frac{\Delta u_1+\Delta u_{{R}}-\Delta u_2}{c_2}\right), \end{equation}

where $\Delta u_{{R}}$ is the jump velocity of SI$_1$ induced by the reflected rarefaction waves in the layer (RW). In previous work (Jacobs et al. Reference Jacobs, Jenkins, Klein and Benjamin1995; Liang & Luo Reference Liang and Luo2021), the accelerations $g$ of two interfaces of the layer were considered the same for both RT and RM instabilities. However, when a shock encounters a gas layer with a finite thickness, the accelerations of the two interfaces are different. Assuming that $g_1$ and $g_2$ represent the accelerations of the first and second interfaces, respectively, and then substituting (3.4) and (3.5) into (3.3) and (3.2), yields

(3.9)\begin{equation} \left. \begin{aligned} \frac{{\rm d}{a}_1}{{\rm d}{t}} & ={-}kb_1,\\ \frac{{\rm d}{a}_2}{{\rm d}{t}} & ={-}kb_2, \end{aligned} \right\} \end{equation}
(3.10)\begin{equation} \left. \begin{aligned} \frac{{\rm d}{b}_1}{{\rm d}{t}}+\frac{{\rm d}{b}_2}{{\rm d}{t}} & ={-}A_{{t}}(g_1a_1-g_2a_2),\\ \frac{{\rm d}{b}_1}{{\rm d}{t}}-\frac{{\rm d}{b}_2}{{\rm d}{t}} & ={-}A_{{c}}(g_1a_1+g_2a_2). \end{aligned} \right\} \end{equation}

Combining (3.9) with (3.10), the J-model can be rewritten as

(3.11)\begin{equation} \left. \begin{aligned} \frac{{{\rm d}}^{2}a_1}{{{\rm d}}t^{2}}+\frac{{{\rm d}}^{2}a_2}{{{\rm d}}t^{2}} & =kA_{{t}}(g_1a_1-g_2a_2),\\ \frac{{{\rm d}}^{2}a_1}{{{\rm d}}t^{2}}-\frac{{{\rm d}}^{2}a_2}{{{\rm d}}t^{2}} & =kA_{{c}}(g_1a_1+g_2a_2). \end{aligned} \right\} \end{equation}

For the RM instability in this work, based on the experimental movements of the undisturbed interfaces shown in figure 2, the two interfaces in the layer will, respectively, experience two acceleration processes. For the first interface, the jump velocities induced by the incident shock (IS) and the rarefaction waves (RW) are $\Delta u_1$ and $\Delta u_{{R}}$, respectively. For the second interface, the jump velocities induced by the transmitted shock (TS$_1$) and the compression waves (CW) are $\Delta u_2$ and $\Delta u_{{C}}$, respectively. Due to the very short durations of the waves interacting with the interfaces, the Dirac $\delta$-function is always used to model the shock impacting the gas layer (Jacobs et al. Reference Jacobs, Jenkins, Klein and Benjamin1995; Liang & Luo Reference Liang and Luo2023a). Therefore, the accelerations $g_1$ and $g_2$ can be expressed as

(3.12)\begin{equation} \left. \begin{aligned} g_1 & =\Delta u_1\,\delta t+\Delta u_{{R}}\,\delta (t-t_{{R}}), \\ g_2 & =\Delta u_2\,\delta (t-t_1)+\Delta u_{{C}}\,\delta (t-t_{{C}}), \end{aligned} \right\} \end{equation}

where $t_1$, $t_{{R}}$ and $t_{{C}}$ have been defined in (3.1). Substituting (3.12) into (3.11) and integrating the equation, a new model for predicting the linear growth rate of the heavy layer after CW impact SI$_2$ is established:

(3.13)\begin{equation} \left. \begin{aligned} \frac{{{\rm d}}a_1}{{{\rm d}}t}+\frac{{{\rm d}}a_2}{{{\rm d}}t} & =kA_{{t}}(\Delta u_1\,a_1^++\Delta u_{{R}}\,a_1^{{R}}-\Delta u_2\,a_2^0-\Delta u_{{C}}\,a_2^{{C}}),\\ \frac{{{\rm d}}a_1}{{{\rm d}}t}-\frac{{{\rm d}}a_2}{{{\rm d}}t} & =kA_{{c}}(\Delta u_1\,a_1^++\Delta u_{{R}}\,a_1^{{R}}+\Delta u_2\,a_2^0+\Delta u_{{C}}\,a_2^{{C}}). \end{aligned} \right\} \end{equation}

According to the impulsive model (Richtmyer Reference Richtmyer1960) and MB model (Meyer & Blewett Reference Meyer and Blewett1972), in (3.13), $a_1^+$ is the amplitude of SI$_1$ just after being impacted by IS; $a_1^{{R}}$ is the amplitude of SI$_1$ just after being impacted by RW; $a_2^0$ is the average amplitude of the unshocked and shocked second interface; and $a_2^{{C}}$ is the average amplitude of SI$_2$ before and after being impacted by CW.

Note that (3.13) is established assuming that both interfaces have the same wavenumber. In the linear regime, no additional modes except the initial basic mode are generated even if interface coupling is present. In this work, due to the different initial perturbation wavenumbers on two interfaces ($k_1$ and $k_2$ for the first and second initial interfaces, respectively), the perturbation with $k_1$ ($k_2$) will be imprinted on SI$_2$ (SI$_1$) through such as interface coupling. In other words, both perturbation modes are present on SI$_2$ and SI$_1$ even in the linear regime. Therefore, it is necessary to predict the linear growth rate of $k_1$ on SI$_2$ ($k_2$ on SI$_1$). Because mode coupling in the linear regime is weak (Luo et al. Reference Luo, Liu, Liang, Ding and Wen2020), we attempt to establish a linear model by considering separately the perturbation growth of $k_1$ and $k_2$ on SI$_1$ (SI$_2$).

We first consider the perturbation growths of $k_1$ on SI$_1$ and SI$_2$. As sketched in figure 5(a), a single-mode perturbation with $k_1$ is imposed on II$_1$ and there is no perturbation on II$_2$. Because the jump velocities of the interfaces have been calculated as shown in table 2, here the amplitudes of the two interfaces at the specific times are provided to solve (3.13). After IS interacts with II$_1$, the amplitude of SI$_1$ ($a_1^+$) can be expressed as

(3.14)\begin{equation} a_1^+=a_1^-Z_1, \end{equation}

where $a_1^-$ and $Z_1=1-\Delta u_1/v_1$ are the initial amplitude of II$_1$ and the compression factor of IS, respectively, and $v_1$ is the velocity of IS. In previous work (Ishizaki & Nishihara Reference Ishizaki and Nishihara1997), after a shock interacts with a single-mode interface in the plane geometry, the transmitted shock will be seeded by the initial interface shape. If the initial amplitude of the interface is small, then the transmitted shock has a sinusoidal shape, but its amplitude decreases very quickly because of the pressure perturbation. If the initial amplitude of the interface is high, then the triple-shock configuration will be formed on the transmitted shock front, which is no longer a smooth one (Rikanati et al. Reference Rikanati, Oron, Sadot and Shvarts2003; Wang et al. Reference Wang, Wang, Zhai and Luo2023). The duration of such a transmitted shock recovering to planar will be longer. In the present work, due to the small amplitude of the initial first interface, the amplitude of the transmitted shock when it is generated in the layer (TS$_1$) is also small. It is believed that TS$_1$ will recover quickly to almost planar when it interacts with the second interface. As a result, SI$_2$ is still undisturbed ($a_2^0=0$), and planar RW aregenerated.

Figure 5. Sketches of the interactions of TS$_1$ with II$_1$, RW with SI$_1$, and CW with SI$_2$: (a) II$_2$ is planar and II$_1$ is perturbed; (b) II$_1$ is planar and II$_2$ is perturbed.

The planar RW move upstream and interact with SI$_1$. Due to the stretching of RW, the perturbation amplitude of SI$_1$ ($a_1^{{R}}$) increases and is of the form

(3.15)\begin{equation} a_1^{{R}}=a_1^+R, \end{equation}

where $R=1+\Delta u_{{R}}/(c_1+c_2)$ is the stretching factor of RW. After RW collide with SI$_1$, CW are reflected and move to SI$_2$. Similar to TS$_1$, the amplitude of CW is also ignored in this work when CW encounter the planar SI$_2$ ($a_2^{{C}}=0$). In other words, considering a disturbed II$_1$ and a planar II$_2$ as shown in figure 5(a), because the amplitudes of both TS$_1$ and CW are ignored in this work, both TS$_1$ and CW will not seed the perturbation to the second interface, and SI$_2$ obtains the perturbation only through interface coupling. The specific amplitudes calculated are substituted into (3.13), and the linear growth rates of the perturbations with $k_1$ (the wavenumber of the initial first interface) on SI$_1$ and SI$_2$ can be obtained as

(3.16)\begin{equation} \left. \begin{aligned} \frac{{{\rm d}}a_1}{{{\rm d}}t} & =\frac{k_1[A_{{t}}(\Delta u_1\,a_1^++\Delta u_{{R}}\,a_1^{{R}})+A_{{c}}(\Delta u_1\,a_1^++\Delta u_{{R}}\,a_1^{{R}})]}{2},\\ \frac{{{\rm d}}a_2}{{{\rm d}}t} & =\frac{k_1[A_{{t}}(\Delta u_1\,a_1^++\Delta u_{{R}}\,a_1^{{R}})-A_{{c}}(\Delta u_1\,a_1^++\Delta u_{{R}}\,a_1^{{R}})]}{2}. \end{aligned} \right\} \end{equation}

Figure 6 presents the amplitude developments of $k_1$ on SI$_1$ and SI$_2$ for $K>1$ (figures 6a,b) and $K<1$ (figures 6c,d). Note that the initial time ($t=0$) here is defined as the moment when the linear growth of perturbation starts (after CW compress SI$_2$), which is different from the definition earlier. On SI$_1$, the amplitude of $k_1$ grows fast, and nonlinearity decreases the amplitude growth rate soon. The model can predict the linear growth of the amplitude of $k_1$ on SI$_1$. The slight differences of the linear growth rates among cases are probably ascribed to the slight differences of the initial amplitudes and weak dependence upon $k_1$, as indicated in figure 7. On SI$_2$, due to the small initial amplitude of $k_1$, nonlinearity is weak and the linear model is sufficiently accurate to predict the amplitude evolution, except in cases 30-60-IP and 40-20-IP. In these two cases, the amplitudes of $k_1$ on SI$_2$ at late stages are promoted, as shown in figures 6(b,d), mainly because of mode coupling, i.e. the self-coupling of $k_2$ in case 30-60-IP generates $k_1$, and the mutual coupling of $k_2$ with $k_1$ in case 40-20-IP generates $k_1$. Besides, the linear growth rates of $k_1$ on SI$_2$ differ significantly between cases with different $k_1$, because they are strongly associated with $k_1$, as indicated in figure 7. Specifically, the linear growth rate of $k_1$ on SI$_2$ is greater for a smaller $k_1$. Relatively, the amplitude of $k_1$ on SI$_2$ is small (large) for $K>1$ ($K<1$).

Figure 6. Comparisons of the amplitudes of $k_1$ and $k_2$ between experimental and theoretical results in cases (a,b) $K>1$, and (c,d) $K<1$. For the perturbations with the same wavelength, the predictions of the linear growth rate are the same, and only one theoretical line is provided. Here, the initial time is the moment when the linear growth of perturbation starts (after CW compress SI$_2$), and $a_0$ is the amplitude of the interface at the initial time, and similarly hereafter.

Figure 7. Dependence of linear growth rates of mode amplitudes, predicted by (3.16) and (3.25), on initial wavelength.

If a single-mode perturbation with $k_2$ is imposed on II$_2$, while keeping II$_1$ planar ($a_1^+=0$), then the results will be different. As shown in figure 5(b), because both IS and II$_1$ are planar, TS$_1$ is also planar. Induced by TS$_1$, the average amplitude of II$_2$ and SI$_2$ just being shocked ($a_2^0$) can be calculated:

(3.17)\begin{equation} a_2^0=a_2^-(1+Z_2)/2, \end{equation}

where $a_2^-$ is the amplitude of II$_2$, $v_2$ is the velocity of TS$_1$, and $Z_2=1-\Delta u_2/v_2$ is the compression factor of TS$_1$. The amplitude of SI$_2$ reduces until phase inversion is finished.

The interaction of TS$_1$ with II$_2$ generates perturbed RW. Note that the amplitude of RW increases with time (Velikovich & Phillips Reference Velikovich and Phillips1996). The amplitude of the RW tail before it impacts SI$_1$ ($a_{{RW}}$) can be calculated by the model proposed by Velikovich & Phillips (Reference Velikovich and Phillips1996):

(3.18)\begin{equation} a_{{RW}}=a_2^-Z_2+k_2a_2^-c_1u_r(t_{{R}}-t_1), \end{equation}

where $u_{{r}}=\sqrt {(\gamma +1)(1-{c_2}/{c_1})/(\gamma -1)}$ is a factor of RW, and $\gamma$ is the adiabatic exponent of the gas before the RW front. The time and velocity parameters are all presented in table 2. Due to the RW stretching, the perturbation with $k_2$ arises on SI$_1$ just being impacted by RW, and its amplitude ($a_1^{{R}}$) can be expressed as

(3.19)\begin{equation} a_1^{{R}}=\frac{(a_2^-+a_{{RW}})(R-1)}{2}. \end{equation}

After perturbed RW collide with II$_1$, perturbed CW are formed, but are believed to recover to planar before they interact with SI$_2$ as mentioned before. The amplitude of SI$_2$ just before being impacted by CW ($a_2^{{C}-}$) differs significantly from $a_2^+$ (amplitude of SI$_2$ just after TS$_1$ impact) and can be expressed as

(3.20)\begin{equation} a_2^{{C}-}=a_2^-Z_2+V_2^0(t_{{C}}-t_1), \end{equation}

where $V_2^0$ is the perturbation growth rate of SI$_2$ before being impacted by CW, and can be calculated by the MB model (Meyer & Blewett Reference Meyer and Blewett1972) as

(3.21)\begin{equation} V_2^0={-}k_2a_2^0A^+\,\Delta u_2. \end{equation}

Then the average amplitude of SI$_2$ before and after being impacted by CW ($a_2^{{C}}$) is

(3.22)\begin{equation} a_2^{{C}}=a_2^{{C}-}(1+Z_{{C}})/2, \end{equation}

where $Z_{{C}}=(1-\Delta u_{{C}}/c_2)$ is the compression factor of CW. Substituting $a_1^+$, $a_2^0$, $a_1^{{R}}$ and $a_2^{{C}}$ into (3.13), the linear growth rates of perturbations with $k_2$ on SI$_1$ and SI$_2$ can be predicted. However, the predictions deviate from the experimental results, which is probably ascribed to the assumption of deducing (3.11) that the velocity potentials of the flow in the layer are the same. Although the different accelerations of the two interfaces are considered, the effects of the disturbed RW and CW also need to be considered. In other words, when considering a planar II$_1$ and disturbed II$_2$ as shown in figure 5(b), SI$_1$ obtains a perturbation through both the waves’ effect and interface coupling. Actually, the disturbed RW not only accelerates SI$_1$ but also obviously changes the perturbation growth. The impulsive model (Richtmyer Reference Richtmyer1960) is used to calculate the amplitude growth rate of SI$_1$ induced by the perturbed RW:

(3.23)\begin{equation} \frac{{{\rm d}}a_{1, {{R}}}}{{{\rm d}}t}=k_2a_1^{{R}}\,\Delta u_{{R}}\,A^+. \end{equation}

When CW impacts SI$_2$, the amplitude of $k_2$ develops significantly and the MB model (Meyer & Blewett Reference Meyer and Blewett1972) is used to predict the amplitude growth rate of SI$_2$ induced by CW:

(3.24)\begin{equation} \frac{{{\rm d}}a_{2, {\rm{C}}}}{{{\rm d}}t}={-}k_2a_2^{{C}}\,\Delta u_{{C}}\,A^+. \end{equation}

Combining (3.13), (3.23) with (3.24), the model that can predict the linear growth rate of the perturbation with $k_2$ (the wavenumber of the initial second interface) on SI$_1$ and SI$_2$ can be written as

(3.25)\begin{align} \left. \begin{aligned} \frac{{{\rm d}}a_1}{{{\rm d}}t} & =\frac{k_2[A_{{t}}(\Delta u_{{R}}\,a_1^{{R}}-\Delta u_2\,a_2^0-\Delta u_{{C}}\,a_2^{{C}})+A_{{c}}(\Delta u_{{R}}\,a_1^{{R}}+\Delta u_2\,a_2^0+\Delta u_{{C}}\,a_2^{{C}})]}{2}\\ & \quad +k_2a_1^{{R}}\Delta u_{{R}}\,A^+,\\ \frac{{{\rm d}}a_2}{{{\rm d}}t} & =\frac{k_2[A_{{t}}(\Delta u_{{R}}\,a_1^{{R}}-\Delta u_2\,a_2^0-\Delta u_{{C}}\,a_2^{{C}})-A_{{c}}(\Delta u_{{R}}\,a_1^{{R}}+\Delta u_2\,a_2^0+\Delta u_{{C}}\,a_2^{{C}})]}{2}\\ & \quad -k_2a_2^{{C}}\,\Delta u_{{C}}\,A^+. \end{aligned} \right\} \end{align}

The amplitude developments of $k_2$ on SI$_1$ and SI$_2$ are also presented in figure 6 for $K>1$ (figures 6a,b) and $K<1$ (figures 6c,d). On SI$_2$, the amplitude of $k_2$ grows first linearly and then nonlinearly. On SI$_1$, the linear model can well predict the amplitude growth of $k_2$. From figure 7, the linear growth rates of $k_2$ on both SI$_1$ and SI$_2$ are positively correlated with the wavelength of II$_2$ ($\lambda _2$). As a result, the linear growth rates of $k_2$ on both SI$_1$ and SI$_2$ are greater when $\lambda _2$ is larger. If the phases of two initial perturbations are opposite, then the linear models can also predict the growth rates of both $k_1$ and $k_2$.

3.2.3. Nonlinear growth of amplitude

As shown in figure 6, generally, only the amplitudes of $k_1$ (wavenumber of II$_1$) on SI$_1$, and $k_2$ (wavenumber of II$_2$) on SI$_2$, grow nonlinearly. The amplitude developments of perturbation with $k_1$ on SI$_1$ ($k_2$ on SI$_2$) in dimensionless form are shown in figure 8. On SI$_1$, the time and amplitude of $k_1$ are scaled as $\tau =k_1({{\textrm {d}}a_{1}^l}/{{\textrm {d}}t})t$ and $\alpha =k_1(a_1-a_1^0)$, where ${{\textrm {d}}a_{1}^l}/{{\textrm {d}}t}$ is the experimental linear growth rate of $k_1$, and $a_1^0$ is the amplitude of $k_1$ at $t=0$. On SI$_2$, the time and amplitude of $k_2$ are scaled as $\tau =k_2({{\textrm {d}}a_{2}^{l}}/{{\textrm {d}}t})t$ and $\alpha =k_2(a_2-a_2^0)$, respectively, where ${{\textrm {d}}a_{2}^{l}}/{{\textrm {d}}t}$ is the experimental linear growth rate of $k_2$, and $a_2^0$ is the amplitude of $k_2$ at $t=0$. The nonlinear amplitude growths of $k_1$ on SI$_1$ for cases collapse. However, for $k_2$ on SI$_2$, the amplitude growths in cases with $K>1$ are different from those in cases with $K<1$.

Figure 8. Comparison of the dimensionless amplitudes of the perturbation with (a) $k_1$ on SI$_1$ and (b) $k_2$ on SI$_2$.

Zhang & Guo (Reference Zhang and Guo2016) previously proposed a nonlinear model (the ZG model) to predict the amplitude growth of a single interface. The ZG model can be described as

(3.26)\begin{equation} \frac{{{\rm d}}a_{{{b/s}}}^{{{ZG}}}}{{{\rm d}}t}(t)=\frac{{{\rm d}}a^{l}}{{{\rm d}}t}\,\frac{1}{1+\theta k\,\dfrac{{{\rm d}}a^{l}}{{{\rm d}}t}\,t}, \end{equation}

where $k$ is the wavenumber of the perturbation, ${{\textrm {d}}a^{l}}/{{\textrm {d}}t}$ is the linear growth rate of the perturbation after the waves impact, and

(3.27)\begin{equation} \theta=\frac{3}{4}\,\frac{(1+A^+)(3+A^+)}{3+A^++\sqrt{2(1+A^+)}}\times \frac{4(3+A^+)+\sqrt{2}(9+A^+)(1+A^+)^{1/2}}{(3+A^+)^2+2\sqrt{2}(3-A^+)(1+A^+)^{1/2}} \end{equation}

is a function for the Atwood number ($A^+$) of the shocked interface. The ZG model has been verified widely for a single-mode interface (Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a; Guo et al. Reference Guo, Cong, Si and Luo2022), but it overestimates the nonlinear growth of $k_1$ on SI$_1$, as shown in figure 8(a). This indicates that in a heavy layer, interface coupling promotes nonlinearity of $k_1$ growth on SI$_1$, which is consistent with the observation in figure 3 that interface coupling greatly deforms the morphology of SI$_1$. For $k_2$ on SI$_2$, as shown in figure 8(b), the ZG model slightly overestimates the nonlinear amplitude growth for cases with $K<1$, whereas it generally predicts the nonlinear amplitude growth for cases with $K>1$. These indicate that interface coupling promotes nonlinearity of $k_2$ growth on SI$_2$ for cases with $K<1$, but it has a little effect on the amplitude growth of $k_2$ on SI$_2$ for cases with $K>1$. These conclusions are also consistent with the observations in figure 3.

The nonlinear model proposed by Sadot et al. (Reference Sadot, Erez, Alon, Oron, Levin, Erez, Ben-Dor and Shvarts1998) (the SEA model) has been modified empirically (Dimonte & Ramaprabhu Reference Dimonte and Ramaprabhu2010; Mansoor et al. Reference Mansoor, Dalton, Martinez, Desjardins, Charonko and Prestridge2020) to predict the nonlinear growth of RM instability. The SEA model can be expressed as

(3.28)\begin{equation} \frac{{\rm d}a_{b/s}^{{{SEA}}}}{{\rm d}t}(t)=\frac{{\rm d}a^{l}}{{\rm d}t}\, \frac{1+k\,\dfrac{{\rm d}a^{l}}{{\rm d}t}\,t}{1+(1\pm A^+)k\,\dfrac{{\rm d}a^{l}}{{\rm d}t}\,t +C(1\pm A^+)/(1+A^+)\left(k\dfrac{{\rm d}a^{l}}{{\rm d}t}t\right)^2}, \end{equation}

where $C=1$ for $|A| \rightarrow 0$, and $C=1.5$ for $|A| > 0.5$. Based on the experimental results, the SEA model is modified empirically in the current work to predict the amplitude growths of two interfaces. The expression of the modified SEA model (mSEA model) can be described as

(3.29)\begin{equation} \frac{{{\rm d}}a_{{{b/s}}}^{{{mSEA}}}}{{{\rm d}}t}(t)=\frac{{{\rm d}}a^{l}}{{{\rm d}}t}\,\frac{1+\dfrac{(1\mp A^+)}{2}k\,\dfrac{{{\rm d}}a^{l}}{{{\rm d}}t}\,t}{1+(1\pm A^+)k\,\dfrac{{{\rm d}}a^{l}}{{{\rm d}}t}\,t+5(1+A^+)(1-A^+)\left(k\,\dfrac{{{\rm d}}a^{l}}{{{\rm d}}t}\,t\right)^2}. \end{equation}

The predictions of the mSEA model for $k_1$ on SI$_1$ and $k_2$ on SI$_2$ are shown in figure 8. For $k_2$ on SI$_2$ in cases $K<1$, and $k_1$ on SI$_1$, the mSEA model provides reasonable predictions.

3.2.4. Modal evolution

As described earlier, new modes are generated due to interface coupling on the interfaces except for the initial basic modes, then mode coupling arises. For the evolution of a multi-mode interface, a modal model with second-order accuracy was proposed by Haan (Reference Haan1991) – referred to as the Haan model hereafter – to calculate the amplitude growth of the coupling modes resulting from the addition or subtraction of the initial basic modes. The Haan model was then simplified by Remington et al. (Reference Remington, Weber, Marinak, Haar, Kilkenny, Wallace and Dimonte1994) with second-order accuracy to quantify mode coupling in the RT instability (denoted as the REM model) and modified by Luo et al. (Reference Luo, Liu, Liang, Ding and Wen2020) to predict the evolution of the multi-mode RM instability. However, all the previous models were solved and modified only for a single multi-mode interface but not for a gas layer. To quantify the mode-coupling effect in a shocked heavy layer, a new model is established by combining the REM model with (3.11)–(3.12). For the two different basic modes $k_1$ and $k_2$ on the initial interfaces considered in this work, the REM model (Remington et al. Reference Remington, Weber, Marinak, Haar, Kilkenny, Wallace and Dimonte1994) can be expressed as

(3.30)\begin{equation} a_{k_1 \pm k_2}={\mp} \frac{1}{2}(k_1\pm k_2)A^+(a_{k_1}^la_{k_2}^l), \end{equation}

where $a_{k_1}$ ($a_{k_2}$) is the amplitude of the perturbation with mode $k_1$ ($k_2$), $a_{k_1 \pm k_2}$ is the amplitude of the perturbation with coupling modes $k_1 \pm k_2$, and the superscript $l$ denotes the linear regime. The right-hand side of (3.30) represents the generation of coupling mode $k_1\pm k_2$ from initial modes $k_1$ and $k_2$. Taking the second derivative of (3.30) with time as in our previous work (Luo et al. Reference Luo, Liu, Liang, Ding and Wen2020), one can obtain

(3.31)\begin{equation} \frac{{{\rm d}}^{2}a_{k_1\pm k_2}}{{{\rm d}}t^{2}}={\mp} \frac{1}{2}\,A^+(k_1\pm k_2)\left(\frac{{{\rm d}}^{2}a_{k_1}^l}{{{\rm d}}t^{2}}\,a_{k_2}^l +\frac{{{\rm d}}^{2}a_{k_2}^l}{{{\rm d}}t^{2}}\,a_{k_1}^l +2\,\frac{{{\rm d}}a_{k_1}^l}{{{\rm d}}t}\,\frac{{{\rm d}}a_{k_2}^l}{{{\rm d}}t}\right). \end{equation}

For a fluid layer, to quantify the evolution of ${{\textrm {d}}^{2}a_{k_1}^l}/{{\textrm {d}}t^{2}}$ and ${{\textrm {d}}^{2}a_{k_2}^l}/{{\textrm {d}}t^{2}}$, (3.11) can be written as follows: for the first interface,

(3.32)\begin{equation} \left. \begin{aligned} \frac{{{\rm d}}^{2}a_{1,k_1}^l}{{{\rm d}}t^{2}} & =\frac{k_1}{2}\left[A_{{t},k_1}(g_1a_{1,k_1}^l-g_2a_{2,k_1}^l)+A_{{c},k_1}(g_1a_{1,k_1}^l+g_2a_{2,k_1}^l)\right],\\ \frac{{{\rm d}}^{2}a_{1,k_2}^l}{{{\rm d}}t^{2}} & =\frac{k_2}{2}\left[A_{{t},k_2}(g_1a_{1,k_2}^l-g_2a_{2,k_2}^l)+A_{{c},k_2}(g_1a_{1,k_2}^l+g_2a_{2,k_2}^l)\right]; \end{aligned} \right\} \end{equation}

for the second interface,

(3.33)\begin{equation} \left. \begin{aligned} \frac{{{\rm d}}^{2}a_{2,k_1}^l}{{{\rm d}}t^{2}} & =\frac{k_1}{2}\left[A_{{t},k_1}(g_1a_{1,k_1}^l-g_2a_{2,k_1}^l)-A_{{c},k_1}(g_1a_{1,k_1}^l+g_2a_{2,k_1}^l)\right],\\ \frac{{{\rm d}}^{2}a_{2,k_2}^l}{{{\rm d}}t^{2}} & =\frac{k_2}{2}\left[A_{{t},k_2}(g_1a_{1,k_2}^l-g_2a_{2,k_2}^l)-A_{{c},k_2}(g_1a_{1,k_2}^l+g_2a_{2,k_2}^l)\right]. \end{aligned} \right\} \end{equation}

Substituting (3.32)–(3.33) into (3.31), the following expressions can be obtained: for the first interface,

(3.34) \begin{align} \frac{{{\rm d}}^{2}a_{1, k_1\pm k_2}}{{{\rm d}}t^{2}}&={\mp} \frac{1}{2}\,A^+(|k_1\pm k_2|)\nonumber\\ &\qquad \left\{\vphantom{\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}}\frac{1}{2}\left[k_1A_{{{t}}, k_1}\left(g_1a_{1, k_1}^l-g_2a_{2, k_1}^l\right)+k_1A_{{{c}}, k_1}(g_1a_{1, k_1}^l+g_2a_{2, k_1}^l)\right]a_{1, k_2}^l\right.\nonumber\\ &\quad +\frac{1}{2}\left[k_2A_{{{t}}, k_2}(g_1a_{1, k_2}^l-g_2a_{2, k_2}^l)+k_2A_{{{c}}, k_2}(g_1a_{1,k_2}^l+g_2a_{2, k_2}^l)\right]a_{1, k_1}^l\nonumber\\ &\left.\quad {}+2\,\frac{{{\rm d}}a_{1, k_1}^l}{{{\rm d}}t}\,\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}\right\}; \end{align}

for the second interface,

(3.35) \begin{align} \frac{{{\rm d}}^{2}a_{2, k_1\pm k_2}}{{{\rm d}}t^{2}}&={\mp} \frac{1}{2}\,A^+(|k_1\pm k_2|)\nonumber\\ &\qquad \left\{\vphantom{\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}}\frac{1}{2}\left[k_1A_{{{t}}, k_1}(g_1a_{1, k_1}^l-g_2a_{2, k_1}^l)-k_1A_{{{c}}, k_1}(g_1a_{1, k_1}^l+g_2a_{2, k_1}^l)\right]a_{2, k_2}^l\right.\nonumber\\ &\quad +\frac{1}{2}\left[k_2A_{{{t}}, k_2}(g_1a_{1, k_2}^l-g_2a_{2, k_2}^l)-k_2A_{{{c}}, k_2}(g_1a_{1, k_2}^l+g_2a_{2, k_2}^l)\right]a_{2, k_1}^l\nonumber\\ &\left.\quad {}+2\,\frac{{{\rm d}}a_{2, k_1}^l}{{{\rm d}}t}\,\frac{{{\rm d}}a_{2, k_2}^l}{{{\rm d}}t}\right\}. \end{align}

Here, $a_{1, k_1}$ ($a_{2, k_1}$), $a_{1, k_2}$ ($a_{2, k_2}$) and $a_{1, k_1\pm k_2}$ ($a_{2, k_1\pm k_2}$) are the amplitudes of modes $k_1$, $k_2$ and $k_1\pm k_2$ on SI$_1$ (SI$_2$), respectively; $A_{{{t}}, k_1}$ ($A_{{{t}}, k_2}$) and $A_{{{c}}, k_1}$ ($A_{{{c}}, k_2}$) are the two modified Atwood numbers of perturbation with $k_1$ ($k_2$), which can be expressed by (3.7), and $g_1$ ($g_2$) is the acceleration of the first (second) interface calculated by (3.12). Substituting $g_1$ and $g_2$ into (3.34)–(3.35) and then integrating (3.34)–(3.35), a model that can quantify the mode-coupling effect in a shocked layer is obtained and expressed as follows: for the first interface,

(3.36) \begin{align} \frac{{{\rm d}}a_{1, k_1\pm k_2}}{{{\rm d}}t}&={\mp} \frac{1}{2}\,A^+(|k_1\pm k_2|)\nonumber\\ &\qquad \left\{\vphantom{\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}}\frac{1}{2}\left[k_1A_{{{t}}, k_1}\left(\Delta u_1\,a_{1, k_1}^+a_{1, k_2}^++\Delta u_{{R}}\,a_{1, k_1}^{{R}}a_{1, k_2}^{{R}}-\Delta u_2\,a_{2, k_1}^0a_{1, k_2}^+{-}\Delta u_{{C}}\,a_{2, k_1}^{{C}}a_{1, k_2}^{{C}}\right)\right.\right.\nonumber\\ &\left.\quad {}+k_1A_{{{c}}, k_1}\left(\Delta u_1\,a_{1, k_1}^+a_{1, k_2}^++\Delta u_{{R}}\,a_{1, k_1}^{{R}}a_{1, k_2}^{{R}}+\Delta u_2\,a_{2, k_1}^0a_{1, k_2}^++\Delta u_{{C}}\,a_{2, k_1}^{{C}}a_{1, k_2}^{{C}}\right)\right]\nonumber\\ &\quad +\frac{1}{2}\left[k_2A_{{{t}}, k_2}\left(\Delta u_1\,a_{1, k_2}^+a_{1, k_1}^++\Delta u_{{R}}\,a_{1, k_2}^{{R}}a_{1, k_1}^{{R}}-\Delta u_2\,a_{2, k_2}^0a_{1, k_1}^+{-}\Delta u_{{C}}\,a_{2, k_2}^{{C}}a_{1, k_1}^{{C}}\right)\right.\nonumber\\ &\left.\quad{}+k_2A_{{{c}}, k_2}\left(\Delta u_1\,a_{1, k_2}^+a_{1, k_1}^++\Delta u_{{R}}\,a_{1, k_2}^{{R}}a_{1, k_1}^{{R}}+\Delta u_2\,a_{2, k_2}^0a_{1, k_1}^++\Delta u_{{C}}\,a_{2, k_2}^{{C}}a_{1, k_1}^{{C}}\right)\right]\nonumber\\ &\left.\quad{}+2\int \frac{{{\rm d}}a_{1, k_1}^{{{mSEA}}}}{{{\rm d}}t}\,\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}\,{\rm d}t\right\}; \end{align}

for the second interface,

(3.37) \begin{align} \frac{{{\rm d}}a_{2, k_1\pm k_2}}{{{\rm d}}t}&={\mp} \frac{1}{2}\,A^+(|k_1\pm k_2|)\nonumber\\ &\qquad \left\{\vphantom{\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}}\frac{1}{2}\left[k_1A_{{{t}}, k_1}\left(\Delta u_1\,a_{1, k_1}^+a_{2,k_2}^0+\Delta u_{{R}}\,a_{1,k_1}^{{R}}a_{2,k_2}^{{R}}-\Delta u_2\,a_{2, k_1}^0a_{2,k_2}^0-\Delta u_{{C}}\,a_{2,k_1}^{{C}}a_{2,k_2}^{{C}}\right)\right.\right.\nonumber\\ &\left.\quad -k_1A_{{{c}}, k_1}\left(\Delta u_1\,a_{1, k_1}^+a_{2,k_2}^0+\Delta u_{{R}}\,a_{1,k_1}^{{R}}a_{2,k_2}^{{R}}+\Delta u_2\,a_{2, k_1}^0a_{2,k_2}^0+\Delta u_{{C}}\,a_{2,k_1}^{{C}}a_{2,k_2}^{{C}}\right)\right]\nonumber\\ &\quad +\frac{1}{2}\left[k_2A_{{{t}}, k_2}\left(\Delta u_1\,a_{1, k_2}^+a_{2,k_1}^0+\Delta u_{{R}}\,a_{1,k_2}^{{R}}a_{2,k_1}^{{R}}-\Delta u_2\,a_{2, k_2}^0a_{2,k_1}^0-\Delta u_{{C}}\,a_{2,k_2}^{{C}}a_{2,k_1}^{{C}}\right)\right.\nonumber\\ &\left.\quad -k_2A_{{{c}}, k_2}\left(\Delta u_1\,a_{1, k_2}^+a_{2,k_1}^0+\Delta u_{{R}}\,a_{1,k_2}^{{R}}a_{2,k_1}^{{R}}+\Delta u_2\,a_{2, k_2}^0a_{2,k_1}^0+\Delta u_{{C}}\,a_{2,k_2}^{{C}}a_{2,k_1}^{{C}}\right)\right]\nonumber\\ &\left.\quad {}+2\int\frac{{{\rm d}}a_{2, k_1}^l}{{{\rm d}}t}\,\frac{{{\rm d}}a_{2, k_2}^{{{ZG}}}}{{{\rm d}}t}\,{\rm d}t\right\}. \end{align}

Here, $a_1^+$ ($a_1^{{R}}$) represents the amplitude of the first interface just being impacted by IS (RW), and $a_2^0$ ($a_2^{{C}}$) is the average amplitude of the second interface before and after being impacted by TS$_1$ (CW), as defined before. In this work, $a_{1, k_2}^+$, $a_{2, k_1}^0$, $a_{2, k_1}^{{R}}$ and $a_{2, k_1}^{{C}}$ are all equal to zero, and other specific amplitudes can be calculated through the linear theory mentioned in § 3.2.2. In the integral terms on the right-hand sides of (3.34) and (3.35), ${{\textrm {d}}a_{1, k_2}^l}/{{\textrm {d}}t}$ and ${{\textrm {d}}a_{2, k_1}^l}/{{\textrm {d}}t}$ can be calculated by (3.16) and (3.25), but ${{\textrm {d}}a_{1, k_1}^l}/{{\textrm {d}}t}$ and ${{\textrm {d}}a_{2, k_2}^l}/{{\textrm {d}}t}$ are substituted respectively with ${{\textrm {d}}a_{1, k_1}^{{{mSEA}}}}/{{\textrm {d}}t}$ expressed in (3.29) and ${{\textrm {d}}a_{2, k_2}^{{{ZG}}}}/{{\textrm {d}}t}$ expressed in (3.26), because the growths of $a_{1, k_1}$ and $a_{2, k_2}$ quickly enter the nonlinear stage.

For cases 20-30-IP and 30-20-IP, modes $k_1\pm k_2$ and harmonics have no influence on the growths of $a_{2, k_1}$ and $a_{1, k_2}$. As a result, the linear models can provide accurate predictions for these two amplitude developments. From figure 3, a nearly perfect single-mode profile of SI$_2$ is observed in case 20-30-IP, which means that mode coupling on SI$_2$ is very weak, i.e. $a_{2, k_1\pm k_2}$ should be very small and the predictions are ignored. Besides, $a_{2, k_1\pm k_2}$ in case 30-20-IP are also very limited. Therefore, only the developments of $a_{1, k_1\pm k_2}$ in these two cases are predicted by (3.36) and discussed here. As presented in figure 9, the theories provide good predictions for the amplitude developments of these coupling modes. Note that the deviation between experiment and theoretical model for the developments of $a_{1,k_1-k_2}$ for the 30-20 case seems significant, which is caused mainly by the relatively small range of the vertical axis in figure 9. Considering the error bars, the model can reasonably predict the experimental results for $a_{1,k_1-k_2}$.

Figure 9. Comparison of the developments of $a_{1, k_1\pm k_2}$ between experimental measurements and theoretical predictions in cases 20-30-IP and 30-20-IP.

For cases 60-20-IP (20-60-IP) and 60-20-AP, the coupling mode $k_2-k_1$ ($k_1-k_2$) can be described as the second-order harmonic of $k_1$ ($k_2$). For case 40-20-IP (30-60-IP), $k_2=2k_1$ ($k_1=2k_2$). To fully quantify the modal evolutions, the amplitude growths of the harmonics are also considered after constructing (3.36) and (3.37) for $k_1\pm k_2$. Based on the perturbation expansion method, the ZS model was proposed (Zhang & Sohn Reference Zhang and Sohn1997) to calculate the amplitude growth of the harmonics for a single-mode interface. Here, the ZS model is considered for the amplitude growth of the second-order harmonics: for the first interface,

(3.38)\begin{align} \left. \begin{aligned} a_{1, 2k_1} & =\frac{1}{2}\,A^+k_1\,\frac{{{\rm d}}a_{1, k_1}^l}{{{\rm d}}t}\,t^2-\frac{1}{12}\,k_1^3\left(\frac{{{\rm d}}a_{1, k_1}^l}{{{\rm d}}t}\right)^2\left[4(A^+)^3\left(\frac{{{\rm d}}a_{1, k_1}^l}{{{\rm d}}t}\right)^2t^4+3A^+a_{1,k_1}^{{R}}t^2\right],\\ a_{1, 2k_2} & =\frac{1}{2}\,A^+k_2\,\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}\,t^2-\frac{1}{12}\,k_2^3\left(\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}\right)^2\left[4(A^+)^3\left(\frac{{{\rm d}}a_{1, k_2}^l}{{{\rm d}}t}\right)^2t^4+3A^+a_{1,k_2}^{{R}}t^2\right]; \end{aligned} \right\} \end{align}

for the second interface,

(3.39)\begin{align} \left. \begin{aligned} a_{2, 2k_1} & =\frac{1}{2}\,A^+k_1\,\frac{{{\rm d}}a_{2, k_1}^l}{{{\rm d}}t}\,t^2-\frac{1}{12}\,k_1^3\left(\frac{{{\rm d}}a_{2, k_1}^l}{{{\rm d}}t}\right)^2\left[4(A^+)^3\left(\frac{{{\rm d}}a_{2, k_1}^l}{{{\rm d}}t}\right)^2t^4+3A^+a_{2,k_1}^{{C}}t^2\right],\\ a_{2, 2k_2} & =\frac{1}{2}\,A^+k_2\,\frac{{{\rm d}}a_{2, k_2}^l}{{{\rm d}}t}\,t^2-\frac{1}{12}\,k_2^3\left(\frac{{{\rm d}}a_{2, k_2}^l}{{{\rm d}}t}\right)^2\left[4(A^+)^3\left(\frac{{{\rm d}}a_{2, k_2}^l}{{{\rm d}}t}\right)^2t^4+3A^+a_{2,k_2}^{{C}}t^2\right]. \end{aligned} \right\} \end{align}

For the single-mode RM instability, previous work (Liu et al. Reference Liu, Liang, Ding, Liu and Luo2018a) has shown that the amplitudes of harmonics higher than third order grow slowly and can be ignored. Therefore, only the second-order harmonic is considered in this work. The linear and nonlinear developments of amplitudes of the basic modes ($a_{1, k_1}$ and $a_{2, k_2}$) can be well predicted through (3.16), (3.25), (3.26) and (3.29). Considering the amplitudes of other modes, (3.36)–(3.39) are used to provide predictions for modal evolutions.

For cases 20-60-IP, 60-20-IP and 60-20-AP, the linear models can also well predict the growths of $a_{1, k_2}$ and $a_{2, k_1}$. The experimental amplitudes of mode $k_1+k_2$ agree with the predictions by (3.36)–(3.37), as given in figures 10(a,b). However, for predicting the amplitude of mode $|k_1-k_2|$, the effect of harmonics needs to be considered because $|k_1-k_2|=2k_1$ or $2k_2$. Through adding (3.38)–(3.39) to (3.36)–(3.37), the amplitude growths of $|k_1-k_2|$ can be well predicted. For cases 30-60-IP and 40-20-IP, because $k_1=2k_2$ and $k_2=2k_1$, the second-order harmonics and mode-coupling mode ($|k_1-k_2|$) directly affect the growth of $a_{2, k_1}$ and $a_{1, k_2}$, therefore the linear models fail to predict the growths of $a_{1, k_2}$ and $a_{2,k_1}$ at late stages. Combining (3.16), (3.25) with (3.36)–(3.39), the predictions for $a_{1, k_2}$ and $a_{2, k_1}$ in these two cases are provided. In case 40-20-IP, the amplitude growth of $k_2$ on SI$_1$ even experiences a phase inversion process because the harmonic induced by $k_1$ develops faster than the linear growth of $k_2$ at late stages. To predict the developments of $a_{1, k_1\pm k_2}$ and $a_{2, k_1\pm k_2}$, (3.36)–(3.37) are adopted and ${{\textrm {d}}a_{1, k_2}^l}/{{\textrm {d}}t}$ (${{\textrm {d}}a_{2, k_1}^l}/{{\textrm {d}}t}$) is replaced with ${{\textrm {d}}a_{1, k_2}}/{{\textrm {d}}t}$ (${{\textrm {d}}a_{2, k_1}}/{{\textrm {d}}t}$), which is calculated by combining the linear model with the mode-coupling model and harmonic model.

Figure 10. Comparisons of the amplitudes of $k_1\pm k_2$ measured from experiments with models in (a,b) cases 20-60, 60-20 and 60-20-AP, and (c,d) cases 30-60 and 40-20.

In summary, through combining the linear models (3.16) and (3.25), and nonlinear models (3.26) and (3.29), with modal models (3.36)–(3.39), a complete theory that can reasonably predict the amplitude growths of the basic modes ($a_{1, k_1}$ and $a_{2,k_2}$), the modes generated by interface coupling ($a_{2, k_1}$ and $a_{1, k_2}$), and the modes generated by mode coupling ($a_{k_1\pm k_2}$) is established. By comparing the results in figures 6 and 10, we see that mode coupling is significant (ignorable) for the growth of $a_{2, k_1}$ and $a_{1, k_2}$ ($a_{1, k_1}$ and $a_{2,k_2}$).

3.2.5. Freeze-out of the second interface growth

In previous work, freeze-out of amplitude growth (i.e. the amplitude growth stagnates) was realized by compressibility (Mikaelian Reference Mikaelian1994), the second successive shock impact (Charakhch'yan Reference Charakhch'yan2001) or the weak reflected shock/rarefaction wave impact (Chen et al. Reference Chen, Wang, Zhai and Luo2023a,Reference Chen, Xing, Wang, Zhai and Luob). In this work, if the first interface is planar and the second interface is single-mode, then amplitude freeze-out of the second interface may be realized through interface coupling according to (3.25). In other words, amplitude freeze-out of a single-mode heavy–light perturbation could be realized through interface coupling by adding a planar light–heavy interface in front of the perturbed interface. Evidently, the condition for the amplitude freeze-out of the second interface is ${{\textrm {d}}a_2}/{{\textrm {d}}t}=0$, i.e.

(3.40) \begin{align} 0&=\frac{1}{2}\left\{A_{{t}}\left(\Delta u_{{R}}\,\frac{1+Z_2+k_2c_1u_{{r}}(t_{{R}}-t_1)}{2}\,(R-1)-\Delta u_2\,\frac{1+Z_2}{2}\right.\right.\nonumber\\ &\left.\quad {}-\Delta u_{{C}} \left(Z_2-k_2\,\frac{1+Z_2}{2}\,A^+\,\Delta u_2\,(t_{{C}}-t_1)\right)(1+Z_{{C}})/2\right)\nonumber\\ &\quad -A_{{c}}\left(\Delta u_{{R}}\,\frac{1+Z_2+k_2c_1u_{{r}}(t_{{R}}-t_1)}{2}\,(R-1)+\Delta u_2\,\frac{1+Z_2}{2}\right. \nonumber\\ &\left.\left.\quad {}+\Delta u_{{C}} \left(Z_2-k_2\,\frac{1+Z_2}{2}\,A^+\,\Delta u_2\,(t_{{C}}-t_1)\right) (1+Z_{{C}})/2\right)\right\}\nonumber\\ &\quad-\Delta u_{{C}}\,A^+\left(Z_2-k_2\,\frac{1+Z_2}{2}\,A^+\,\Delta u_2\,(t_{{C}}-t_1)\right)(1+Z_{{C}})/2. \end{align}

Here, all the parameters except $k_2$ are dependent on the shock intensity, gas properties and $L_0$, but are independent of the initial amplitude. Based on (3.40), when the shock intensities and gas properties are fixed, $k_2$ and $L_0$ satisfy the specific relationship to achieve the amplitude freeze-out of the second interface. According to this relationship, an experiment with $k_2=20$ mm, $k_2a_2^-=0.083$ and $L_0$ = 27 mm is designed and conducted. The shock intensity and gas properties are the same as the other cases in this work. The typical schlieren images of the perturbed SF$_6$–air interface are provided in figure 11(a), in which SI$_1$ is not shown after $453\,\mathrm {\mu }\textrm {s}$. After CW impacts SI$_2$ ($453\unicode{x2013} 969\,\mathrm {\mu }\textrm {s}$), the amplitude of SI$_2$ changes very slowly, and amplitude freeze-out is almost achieved. Note that if the first interface is perturbed and the second interface is planar, then it is impossible to achieve the amplitude freeze-out of the first interface through changing $L_0$ based on (3.16).

Figure 11. (a) Schlieren images showing the evolution of the SF$_6$–air interface. (b) The temporal variations of the amplitude of SI$_2$ after the phase inversion is finished at $t=t_0$.

The amplitude development of the second interface after the phase inversion is shown in figure 11(b), in which $t_0$ is the time when the phase inversion is completed. The amplitude of SI$_2$ grows linearly and can be well predicted by the MB model before CW arrive. This also proves that when II$_1$ is planar, interface coupling does not affect the perturbation evolution on SI$_2$ before CW impact. After CW impact SI$_2$, the amplitude is compressed first and then rises for a short while because interface coupling needs a start-up process. When the start-up process is finished, the amplitude almost settles in a steady-state value, and freeze-out is almost realized (the tiny increase of the amplitude may be attributed to the negligible error of the theory). As a result, amplitude freeze-out of a single-mode SF$_6$–air interface is almost realized experimentally through generating a planar air–SF$_6$ interface upstream with a suitable distance.

Through interface coupling between two successive heavy–light interfaces, the amplitude freeze-out of a heavy–light interface was realized numerically by Liang & Luo (Reference Liang and Luo2023a). However, the initial amplitudes of two successive interfaces need to satisfy a specific relationship that is difficult to realize in ICF. In the present work, the first light–heavy interface is planar, and amplitude freeze-out of the second perturbed heavy–light interface is achieved just by simply manipulating the distance of two interfaces. In previous work for the target designs in ICF (Qiao & Lan Reference Qiao and Lan2021), the glow discharge polymer plastic (CH, with density 1.05 g cm$^{-3}$) was used as the outermost ablator layer, while the high-density carbon (HDC, with density 3.48 g cm$^{-3}$) was kept as the main ablator. Through constructing the outer CH-HDC interface, the hydrodynamic instabilities on the interface between HDC and the deuterium–tritium ice (DT ice, with density 0.26 g cm$^{-3}$) can be reduced significantly. The CH-HDC interface (light–heavy) and the HDC-DT interface (heavy–light) construct a heavy layer. The decreasing instabilities on the inner HDC-DT interface coincide with the finding in this work that the amplitude freeze-out of the second interface can be realized in a shocked heavy layer. These findings suggest that in ICF, through generating a light–heavy interface with a suitable distance upstream from the interface separating the ablator material and DT ice, the RM instability may be effectively suppressed.

4. Conclusions

We performed the shock-tube experiments and theoretical studies to highlight mode coupling between two single-mode interfaces with different basic modes. Experimentally, the soap-film technique is used to create planar and single-mode interfaces of the layers. Experimental results are presented and analysed for an initially undisturbed layer, and the seven different kinds of disturbed layers. The schlieren images clearly show the interface morphologies, and the distinct profiles facilitate the modal analysis.

The results indicate that the rarefaction waves (RW) and compression waves (CW) generated in the layer, respectively, accelerate the first and second interface movements. Generally, after CW impact, both interfaces move steadily, which means that the effects of other waves can be ignored. For the perturbed layers, just after CW impact, the perturbations on two interfaces still present single-mode shapes with the basic modes. As two interfaces become closer, interface coupling induces a different mode from the basic mode on each interface. Then mode coupling further generates new modes and complicates the interface profile. The interface with a smaller initial wavenumber generally remains a quasi-single-mode profile, and the bubbles (spikes) develop symmetrically. The multi-mode features become prominent on the interface with a larger initial wavenumber, and the bubbles (spikes) grow with an inclination and size difference.

In the linear regime, the mode-coupling effect is weak, and the linear growth of the basic mode is considered separately. The linear model proposed by Jacobs et al. (Reference Jacobs, Jenkins, Klein and Benjamin1995) is modified by considering the different accelerations of two interfaces and the waves’ effects in the layer. The modified linear model provides good predictions to the linear growth rates of the basic modes and the modes generated by interface coupling. In the nonlinear regime, the amplitude growths of the basic modes can be characterized generally by the existing or modified nonlinear model. It is found that interface coupling promotes nonlinearity of the basic mode growth on the first interface, but behaves differently to the nonlinear growth of the basic mode on the second interface. Moreover, a new modal model is established based on the model proposed by Remington et al. (Reference Remington, Weber, Marinak, Haar, Kilkenny, Wallace and Dimonte1994) to quantify the mode-coupling effect in a gas layer. Besides, the model proposed by Zhang & Sohn (Reference Zhang and Sohn1997) (the ZS model) with fourth-order accuracy is used to characterize the growth of second-order harmonics. Combining the linear model, new modal model with the ZS model, the amplitude growths of modes generated by both mode coupling and interface coupling are well predicted. The mode-coupling effect on the amplitude growth is negligible for the basic modes, but is significant for the interface-coupling modes when the initial one wavenumber is twice the other wavenumber.

The present studies show that if a planar light–heavy interface is added in front of the single-mode heavy–light interface, then amplitude freeze-out of the heavy–light interface is possible, caused by interface coupling. Theoretically, the conditions for amplitude freeze-out of the second single-mode interface are independent of initial amplitude but dependent on initial wavelength and width of the layer. According to the predictions of initial conditions, the specific experiment is designed and conducted. It is verified that amplitude freeze-out of the second interface can be realized after CW impact. These findings suggest that in ICF, through generating a planar light–heavy interface with a suitable distance upstream from the interface separating the ablator material and deuterium–tritium ice, the hydrodynamic instabilities may be suppressed effectively.

Funding

This work was supported by the National Natural Science Foundation of China (nos 12372281, 12022201, 12388101 and 12102425) and Youth Innovation Promotion Association CAS.

Declaration of interests

The authors report no conflict of interest.

References

Betti, R. & Hurricane, O.A. 2016 Inertial-confinement fusion with lasers. Nat. Phys. 12, 435448.CrossRefGoogle Scholar
Brouillette, M. 2002 The Richtmyer–Meshkov instability. Annu. Rev. Fluid Mech. 34, 445468.CrossRefGoogle Scholar
Charakhch'yan, A.A. 2001 Reshocking at the non-linear stage of Richtmyer–Meshkov instability. Plasma Phys. Control. Fusion 43, 11691179.CrossRefGoogle Scholar
Chen, C., Wang, H., Zhai, Z. & Luo, X. 2023 a Attenuation of perturbation growth of single-mode $\textrm {SF}_6$–air interface through reflected rarefaction waves. J. Fluid Mech. 969, R1.CrossRefGoogle Scholar
Chen, C., Xing, Y., Wang, H., Zhai, Z. & Luo, X. 2023 b Freeze-out of perturbation growth of single-mode helium–air interface through reflected shock in Richtmyer–Meshkov flows. J. Fluid Mech. 956, R2.CrossRefGoogle Scholar
Di Stefano, C., Malamud, G., Kuranz, C., Klein, S., Stoeckl, C. & Drake, R. 2015 Richtmyer–Meshkov evolution under steady shock conditions in the high-energy-density regime. Appl. Phys. Lett. 106, 114103.CrossRefGoogle Scholar
Dimonte, G. & Ramaprabhu, P. 2010 Simulations and model of the nonlinear Richtmyer–Meshkov instability. Phys. Fluids 22, 014104.CrossRefGoogle Scholar
Guo, X., Cong, Z., Si, T. & Luo, X. 2022 Shock-tube studies of single- and quasi-single-mode perturbation growth in Richtmyer–Meshkov flows with reshock. J. Fluid Mech. 941, A65.CrossRefGoogle Scholar
Haan, S.W. 1991 Weakly nonlinear hydrodynamic instabilities in inertial fusion. Phys. Fluids B 3, 23492355.CrossRefGoogle Scholar
Ishizaki, R. & Nishihara, K. 1997 Propagation of a rippled shock wave driven by nonuniform laser ablation. Phys. Rev. Lett. 78, 19201923.CrossRefGoogle Scholar
Jacobs, J., Jenkins, D., Klein, D. & Benjamin, R. 1995 Nonlinear growth of the shock-accelerated instability of a thin fluid layer. J. Fluid Mech. 295, 2342.CrossRefGoogle Scholar
Jacobs, J., Klein, D., Jenkins, D. & Benjamin, R. 1993 Instability growth patterns of a shock-accelerated thin fluid layer. Phys. Rev. Lett. 70, 583586.CrossRefGoogle ScholarPubMed
Kuranz, C., et al. 2018 How high energy fluxes may affect Rayleigh–Taylor instability growth in young supernova remnants. Nat. Commun. 9, 1564.CrossRefGoogle ScholarPubMed
Liang, Y. 2022 The phase effect on the Richtmyer–Meshkov instability of a fluid layer. Phys. Fluids 34, 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. & Luo, X. 2021 On shock-induced heavy-fluid-layer evolution. J. Fluid Mech. 920, A13.CrossRefGoogle Scholar
Liang, Y. & Luo, X. 2023 a Hydrodynamic instabilities of two successive slow/fast interfaces induced by a weak shock. J. Fluid Mech. 955, A40.CrossRefGoogle Scholar
Liang, Y. & Luo, X. 2023 b Review on hydrodynamic instabilities of a shocked gas layer. Sci. China-Phys. Mech. Astron. 66, 104701.CrossRefGoogle Scholar
Lindl, J., Landen, O., Edwards, J., Moses, E. & Team, N. 2014 Review of the national ignition campaign 2009–2012. Phys. Plasmas 21, 020501.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
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, 122103.CrossRefGoogle Scholar
Luo, X., Liu, L., Liang, Y., Ding, J. & Wen, C. 2020 Richtmyer–Meshkov instability on a dual-mode interface. J. Fluid Mech. 905, A5.CrossRefGoogle Scholar
Mansoor, M.M., Dalton, S.M., Martinez, A.A., Desjardins, T., Charonko, J.J. & Prestridge, K.P. 2020 The effect of initial conditions on mixing transition of the Richtmyer–Meshkov instability. J. Fluid Mech. 904, A3.CrossRefGoogle Scholar
McFarland, J., Reilly, D., Black, W., Greenough, J. & Ranjan, D. 2015 Modal interactions between a large-wavelength inclined interface and small-wavelength multimode perturbations in a Richtmyer–Meshkov instability. Phys. Rev. E 92, 013023.CrossRefGoogle Scholar
Meshkov, E.E. 1969 Instability of the interface of two gases accelerated by a shock wave. Fluid Dyn. 4, 101104.CrossRefGoogle Scholar
Meyer, K.A. & Blewett, P.J. 1972 Numerical investigation of the stability of a shock-accelerated interface between two fluids. Phys. Fluids 15, 753759.CrossRefGoogle Scholar
Mikaelian, K.O. 1985 Richtmyer–Meshkov instabilities in stratified fluids. Phys. Rev. A 31, 410419.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, 888890.CrossRefGoogle Scholar
Mikaelian, K.O. 1996 Numerical simulations of Richtmyer–Meshkov instabilities in finite-thickness fluid layers. Phys. Fluids 8, 12691292.CrossRefGoogle Scholar
Mohaghar, M., Carter, J., Pathikonda, G. & Ranjan, D. 2019 The transition to turbulence in shock-driven mixing: effects of Mach number and initial conditions. J. Fluid Mech. 871, 595635.CrossRefGoogle Scholar
Ofer, D., Alon, U., Shvarts, D., McCrory, R. & Verdon, C. 1996 Modal model for the nonlinear multimode Rayleigh–Taylor instability. Phys. Plasmas 3, 30733090.CrossRefGoogle Scholar
Prestridge, K., Vorobieff, P., Rightley, P. & Benjamin, R. 2000 Validation of an instability growth model using particle image velocimetry measurements. Phys. Rev. Lett. 84, 43534356.CrossRefGoogle ScholarPubMed
Qiao, X. & Lan, K. 2021 Novel target designs to mitigate hydrodynamic instabilities growth in inertial confinement fusion. Phys. Rev. Lett. 126, 185001.CrossRefGoogle ScholarPubMed
Ranjan, D., Oakley, J. & Bonazza, R. 2011 Shock–bubble interactions. Annu. Rev. Fluid Mech. 43, 117140.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
Remington, B., Weber, S., Marinak, M., Haar, S., Kilkenny, J., Wallace, R. & Dimonte, G. 1994 Multimode Rayleigh–Taylor experiments on Nova. Phys. Rev. Lett. 73, 545548.CrossRefGoogle ScholarPubMed
Richtmyer, R.D. 1960 Taylor instability in shock acceleration of compressible fluids. Commun. Pure Appl. Maths 13, 297319.CrossRefGoogle Scholar
Rikanati, A., Alon, U. & Shvarts, D. 1998 Vortex model for the nonlinear evolution of the multimode Richtmyer–Meshkov instability at low Atwood numbers. Phys. Rev. E 58, 74107418.CrossRefGoogle Scholar
Rikanati, A., Oron, D., Sadot, O. & Shvarts, D. 2003 High initial amplitude and high Mach number effects on the evolution of the single-mode Richtmyer–Meshkov instability. Phys. Rev. E 67, 026307.CrossRefGoogle ScholarPubMed
Sadot, O., Erez, L., Alon, U., Oron, D., Levin, L.A., Erez, G., Ben-Dor, G. & Shvarts, D. 1998 Study of nonlinear evolution of single-mode and two-bubble interaction under Richtmyer–Meshkov instability. Phys. Rev. Lett. 80, 16541657.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, 192196.Google Scholar
Vandenboomgaerde, M., Gauthier, S. & Mügler, C. 2002 Nonlinear regime of a multimode Richtmyer–Meshkov instability: a simplified perturbation theory. Phys. Fluids 14, 11111122.CrossRefGoogle Scholar
Velikovich, A. & Phillips, L. 1996 Instability of a plane centered rarefaction wave. Phys. Fluids 8, 11071118.CrossRefGoogle Scholar
Wang, H., Cao, Q., Chen, C., Zhai, Z. & Luo, X. 2022 Experimental study on a light–heavy interface evolution induced by two successive shock waves. J. Fluid Mech. 953, A15.CrossRefGoogle Scholar
Wang, H., Wang, H., Zhai, Z. & Luo, X. 2023 High-amplitude effect on single-mode Richtmyer–Meshkov instability of a light–heavy interface. Phys. Fluids 35, 016106.CrossRefGoogle Scholar
Xu, J., Wang, H., Zhai, Z. & Luo, X. 2023 Convergent Richtmyer–Meshkov instability on two-dimensional dual-mode interfaces. J. Fluid Mech. 965, A8.CrossRefGoogle Scholar
Yang, J., Kubota, T. & Zukoski, E. 1993 Applications of shock-induced mixing to supersonic combustion. AIAA J. 31, 854862.CrossRefGoogle Scholar
Zabusky, N.J. 1999 Vortex paradigm for accelerated inhomogeneous flows: visiometrics for the Rayleigh–Taylor and Richtmyer–Meshkov environments. Annu. Rev. Fluid Mech. 31, 495536.CrossRefGoogle Scholar
Zhai, Z., Zou, L., Wu, Q. & Luo, X. 2018 Review of experimental Richtmyer–Meshkov instability in shock tube: from simple to complex. Proc. Inst. Mech. Engng C 232, 28302849.CrossRefGoogle Scholar
Zhang, D., Ding, J., Si, T. & Luo, X. 2023 Divergent Richtmyer–Meshkov instability on a heavy gas layer. J. Fluid Mech. 959, A37.CrossRefGoogle 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
Zhang, Q. & Sohn, S.I. 1997 Nonlinear theory of unstable fluid mixing driven by shock wave. Phys. Fluids 9, 11061124.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
Figure 0

Figure 1. Schematics of (a) the soap-film interface generation, and (b) the initial interface configuration studied. Here, IS is the incident shock, II$_1$ is the first interface (air–SF$_6$), II$_2$ is the second interface (SF$_6$–air), and $L_0$ is the initial width of the SF$_6$ layer.

Figure 1

Table 1. Interface parameters of perturbed cases: $\lambda _1$ and $\lambda _2$ ($k_1$ and $k_2$, $a_1^-$ and $a_2^-$) are the wavelengths (wavenumbers, initial amplitudes) of II$_1$ and II$_2$, respectively, and $K=k_1/k_2$. Here, ‘IP’ (‘AP’) means the perturbations on the two interfaces have the same (opposite) phase.

Figure 2

Figure 2. (a) Schlieren images of a planar SF$_6$ layer accelerated by a planar shock, and (b) the trajectories of the interfaces and waves. Here, II$_1$ and II$_2$ (SI$_1$ and SI$_2$) are the first and second initial (shocked) interfaces respectively; TS$_1$ and TS$_2$ are the transmitted shock waves in SF$_6$ and in air, respectively; RS is the reflected shock; RW is reflected rarefaction waves; CW is compression waves; and $x_0$ is the initial position of II$_1$.

Figure 3

Table 2. Theoretical parameters in undisturbed case: $M_{{s}}$ is the Mach number of IS; $\Delta u_1$, $\Delta u_2$, $\Delta u_{{R}}$ and $\Delta u_{{C}}$ are the jump velocities of II$_1$ induced by IS, II$_2$ induced by TS$_1$, II$_1$ induced by RW, and II$_2$ induced by CW, respectively; $v_1$ and $v_2$ are the velocities of IS and TS$_1$, respectively; $t_1$, $t_{{R}}$ and $t_{{C}}$ denote the times when TS$_1$ impacts II$_2$, RW impact SI$_1$, and CW impact SI$_2$, respectively; $c_1$ ($c_2$) is the sound speed of the gas between SI$_1$ and the RW front (between SI$_2$ and the RW tail). The units of velocity and time are m s$^{-1}$ and $\mathrm {\mu }{\rm s}$.

Figure 4

Figure 3. Schlieren images of the shock-induced fluid layer evolution for different cases. The symbols have the same meanings as in figure 2. Numbers denote time in $\mathrm {\mu }\textrm {s}$.

Figure 5

Figure 4. Schematic showing the configuration analysed using linear stability theory for the double-interface system, where $\rho _1$ and $\rho _2$ are the densities of the fluids inside and outside the layer, respectively, $\eta _1$ and $\eta _2$ represent the perturbation shapes of the two interfaces, $\phi _1$, $\phi _2$ and $\phi _3$ are the velocity potentials, $L$ is the initial layer thickness, and $g$ is the acceleration of the two interfaces.

Figure 6

Figure 5. Sketches of the interactions of TS$_1$ with II$_1$, RW with SI$_1$, and CW with SI$_2$: (a) II$_2$ is planar and II$_1$ is perturbed; (b) II$_1$ is planar and II$_2$ is perturbed.

Figure 7

Figure 6. Comparisons of the amplitudes of $k_1$ and $k_2$ between experimental and theoretical results in cases (a,b) $K>1$, and (c,d) $K<1$. For the perturbations with the same wavelength, the predictions of the linear growth rate are the same, and only one theoretical line is provided. Here, the initial time is the moment when the linear growth of perturbation starts (after CW compress SI$_2$), and $a_0$ is the amplitude of the interface at the initial time, and similarly hereafter.

Figure 8

Figure 7. Dependence of linear growth rates of mode amplitudes, predicted by (3.16) and (3.25), on initial wavelength.

Figure 9

Figure 8. Comparison of the dimensionless amplitudes of the perturbation with (a) $k_1$ on SI$_1$ and (b) $k_2$ on SI$_2$.

Figure 10

Figure 9. Comparison of the developments of $a_{1, k_1\pm k_2}$ between experimental measurements and theoretical predictions in cases 20-30-IP and 30-20-IP.

Figure 11

Figure 10. Comparisons of the amplitudes of $k_1\pm k_2$ measured from experiments with models in (a,b) cases 20-60, 60-20 and 60-20-AP, and (c,d) cases 30-60 and 40-20.

Figure 12

Figure 11. (a) Schlieren images showing the evolution of the SF$_6$–air interface. (b) The temporal variations of the amplitude of SI$_2$ after the phase inversion is finished at $t=t_0$.