Kinetics of Rayleigh-Taylor instability in van der Waals fluid: the influence of compressibility

Jie Chen1,2, Aiguo Xu2,3,4,5111 Corresponding author. E-mail: [email protected],Yudong Zhang6, Dawei Chen2, Zhihua Chen1 1, National Key Laboratory of Transient Physics, Nan**g University of Science and Technology, Nan**g 210094, China
2, National Key Laboratory of Computational Physics, Institute of Applied Physics and Computational Mathematics, P. O. Box 8009-26, Bei**g 100088, China
3, State Key Laboratory of Explosion Science and Safety Protection, Bei**g Institute of Technology, Bei**g 100081, China
4, HEDPS, Center for Applied Physics and Technology, and College of Engineering, Peking University, Bei**g 100871, China
5, National Key Laboratory of Shock Wave and Detonation Physics, Mianyang 621999, China
6, School of Mechanics and Safety Engineering, Zhengzhou University, Zhengzhou 450001, China
(July 3, 2024)
Abstract

Early studies on Rayleigh-Taylor instability (RTI) primarily relied on the Navier-Stokes (NS) model. As research progresses, it becomes increasingly evident that the kinetic information that the NS model failed to capture is of great value for identifying and even controlling the RTI process; simultaneously, the lack of analysis techniques for complex physical fields results in a significant waste of data information. In addition, early RTI studies mainly focused on the incompressible case and the weakly compressible case. In the case of strong compressibility, the density of the fluid from the upper layer (originally heavy fluid) may become smaller than that of the surrounding (originally light) fluid, thus invalidating the early method of distinguishing light and heavy fluids based on density. In this paper, tracer particles are incorporated into a single-fluid discrete Boltzmann method (DBM) model that considers the van der Waals potential. By using tracer particles to label the matter-particle sources, a careful study of the matter-mixing and energy-mixing processes of the RTI evolution is realized in the single-fluid framework. The effects of compressibility on the evolution of RTI are examined mainly through the analysis of bubble and spike velocities, the ratio of area occupied by heavy fluid, and various entropy generation rates of the system. It is demonstrated that: (i) compressibility has a suppressive effect on the spike velocity, and this suppressive impact diminishes as the Atwood number (At𝐴𝑡Atitalic_A italic_t) increases. The influence of compressibility on bubble velocity shows a staged behavior with increasing At𝐴𝑡Atitalic_A italic_t. (ii) The impact of compressibility on the entropy production rate associated with the heat flow (S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT) is related to the stages of RTI evolution. Moreover, this staged impact of compressibility on S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT varies with At𝐴𝑡Atitalic_A italic_t. Compressibility exhibits an inhibitory effect on the entropy production rate associated with viscous stresses (S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT). (iii) By incorporating the morphological parameter of the proportion of area occupied by heavy fluid (Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT), it is observed that the first minimum point of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t can serve as a criterion for identifying the point at which bubble velocity reaches its first maximum value. The series of physical cognition provides a more accurate understanding of the RTI kinetics and a helpful reference for the development of corresponding regulation techniques.

pacs:
47.11.-j, 47.20.-k, 05.20.Dd
preprint: APS/123-QED

I Introduction

The interface instability that arises when a heavy medium is supported or accelerated by a light-medium is known as Rayleigh-Taylor instability (RTI)  Rayleigh (1882); Taylor (1950). RTI is found in a wide range of natural sciences and engineering applications, such as supernova explosions, inertial confinement fusion (ICF), and super combustion ram engines Zhou (2024, 2017a, 2017b); Zhou et al. (2021). In these processes, the flow is strongly compressible. However, most of the current research on RTI is based on the incompressible or weakly compressible case. At present, research on RTI in the case of strong compressibility remains limited, especially in the later stages of material and energy mixing. It is of great significance to study the influence of compressibility on RTI to control the evolution of interface instability in these processes.

Due to the importance of compressible RTI, this problem has attracted considerable attention from scholars, prompting extensive research in this field. However, the complexity of the issue, coupled with the evolving comprehension of the phenomenon, has resulted in varying or even opposite conclusions about the influence of compressibility on RTI. For example, in the early days, some scholars proposed that compressibility enhances RTI Bernstein and Book (1983); Yang and Zhang (1993), while others argued that compressibility inhibits it Blake (1972). Moreover, there are findings suggesting that both effects coexist Baker (1983).

As research advances, it has become evident that a single parameter cannot determine the influence of compressibility on RTI. For instance, Livescu proposed that the impact of compressibility on RTI depends on the specific heat ratio and the equilibrium pressure at the interface, which have opposite effects on the development of RTI Livescu (2004). Xue and Ye proposed that the influence of compressibility on RTI depends on three factors: density profile, pressure at the interface, and the specific heat ratio Xue and Ye (2010). In fact, compressibility in RTI research can be separated into two categories: fluid compressibility and flow compressibility Livescu (2004); Lafay et al. (2007); Reckinger et al. (2016); Gauthier (2017). The former is related to the inherent properties of fluids (associated with the equation of state and the specific heat ratio difference between the two fluids). The latter, or flow compressibility, is connected to the system’s thermodynamic state. Its static effect leads to background stratification, while its dynamic effect results in an expansion-compression impact on the flow process Luo et al. (2020).

Existing research indicates that flow compressibility shows different effects on bubbles depending on the Atwood numbers (At𝐴𝑡Atitalic_A italic_t). There is a critical value of At𝐴𝑡Atitalic_A italic_t, below which the development of RTI is inhibited by flow compressibility. Conversely, when At𝐴𝑡Atitalic_A italic_t exceeds this critical value, flow compressibility promotes the growth of RTI Reckinger et al. (2016); Luo et al. (2020); Fu et al. (2022). Luo et al. and Fu et al. suggested that this is due to the competition between density stratification and expansion-compression effects Qin et al. (2001); Qin and Wang (2004); Luo et al. (2020); Fu et al. (2022). Fu et al. proposed a modified buoyancy-drag model for stratified compressible RTI. They predicted the critical At𝐴𝑡Atitalic_A italic_t using this model and obtained good agreement with results from direct numerical simulation (DNS). Simultaneously, they discovered that a similar nonlinear saturation is shown for the bubble evolution at At=0.9𝐴𝑡0.9At=0.9italic_A italic_t = 0.9 when the pistonlike effect Olson and Cook (2007); Reckinger et al. (2012), which characterizes the compressive forces exerted by the rising bubble of light fluid on the heavy fluid in front of it, is completely extracted. The piston effect of the rising bubble causes the acceleration of the heavy fluid compressed in front of the bubble at later stages to show an inverse power law of At2.5Ma2𝐴superscript𝑡2.5𝑀superscript𝑎2A{{t}^{2.5}}M{{a}^{2}}italic_A italic_t start_POSTSUPERSCRIPT 2.5 end_POSTSUPERSCRIPT italic_M italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For the RTI of compressible fluid, bubble reacceleration is caused not only by vorticity deposition in bubbles but also by compressibility. Under strong stratification and high At𝐴𝑡Atitalic_A italic_t, flow compressibility dominates bubble reacceleration Fu et al. (2022). Using the vortex transport equation, Wieland et al. explained the asymmetric effect of weak stratification on bubble and spike growth at small At𝐴𝑡Atitalic_A italic_t and the inhibitory effect of strong stratification on RTI growth at small At𝐴𝑡Atitalic_A italic_t Wieland et al. (2019). By introducing dilatation into the classical model, which only considers vortex deposition, Fu et al. proposed a new model that can reliably describe the reacceleration of stratified compressible RTI Fu et al. (2023).

The evolution of hydrodynamic instabilities is a complex process, especially for those exhibiting strong compressibility. In previous studies, the evolution of the RTI system was often described by the velocity and amplitude of bubbles and spikes, as well as the mixing width. These physical quantities are very helpful and intuitive. However, it is far from enough to describe the evolution of the RTI system only based on these physical quantities. They lose a lot of information; that is to say, even if we know these physical quantities, there is still considerable uncertainty in the actual process of material and energy mixing. The incorporation of complex physical field analysis techniques, such as morphological quantities, can be a good complement to the description of RTI evolution. For example, Chen et al. found that the interface length of light and heavy fluid (L𝐿Litalic_L) can be used to measure the ratio of buoyancy to tangential force intensity and to quantitatively judge the primary mechanism in the early development of the coupled Rayleigh–Taylor–Kelvin–Helmholtz instability (RTKHI) system Chen et al. (2020). Actually, the RTI system is a complex, non-equilibrium system, and the physical quantities, such as the velocities and amplitudes of bubbles and spikes, as well as the length of the interface between light and heavy fluids, are only manifestations of the system’s non-equilibrium from a specific perspective. Various perspectives on non-equilibrium evolution are interrelated and complementary, and they can form a more complete image of the system by combining each other.

In recent years, based on statistical physics, Xu et al. have established and developed the discrete Boltzmann method (DBM) Xu and Zhang (2022); Xu et al. (2024). From a historical perspective, DBM is developed from the physical modeling branch of Lattice Boltzmann Method (LBM) Succi (2001); Osborn et al. (1995); Swift et al. (1995); Liang et al. (2014, 2016, 2021). As a method of physical modeling and complex physical field analysis, DBM mainly focuses on constructing physical models and extracting valuable information from extensive datasets. In addition to the three conservation moments of mass, momentum, and energy, which were concerned in the Navier-Stokes (NS) model, DBM describes the evolution of complex systems from multiple perspectives with the help of complex physical field analysis methods, such as the evolution of some non-conservation moments closely related to system evolution and the morphological analysis. This enables DBM to provide some insight that NS methods cannot or are not convenient to obtain. In the study of hydrodynamic instabilities systems, a series of investigations based on DBM have contributed to our understanding of the physical processes and mechanisms involved in the development of hydrodynamic instabilities Chen et al. (2018, 2020); Gan et al. (2019); Lai et al. (2016); Lin et al. (2017). In the study of compressible RTI, with the help of DBM, Lai et al. discovered that the global thermodynamic non-equilibrium (TNE) intensity shows opposite trends in the initial and later stages of the RTI Lai et al. (2016). The local TNE provided a useful physical observable value for tracking the interface of light and heavy fluids. Lin et al. proposed a DBM model for two-component compressible flows with an adjustable specific heat ratio and independent discrete velocity models for each component Lin et al. (2017). Using this model, they studied the invariants of tensors for nonequilibrium effects and the mixed entropy of compressible RTI systems under various Reynolds conditions. These studies describe the evolution of compressible RTI systems from different non-equilibrium perspectives, which are good supplements to the previous research results and help us to understand the physical process in the development of compressible RTI more deeply, but they are all weakly compressible.

In cases of strong compressibility, there may be some novel behaviors and mechanisms in the evolution of the RTI system. For example, the density of the fluid from the upper layer (which was originally heavy fluid) may become lower than that of the fluid around it (which was originally light fluid), thus causing the identity exchange of light and heavy fluids and triggering anti-RTI behavior. The appearance of these novel behaviors and mechanisms makes the early method of judging the source of material particles based on density invalid. It is well known that particle imaging velocimeter (PIV) and planar laser-induced fluorescence (PLIF) technologies have been widely used in experimental fluid mechanics. Inspired by this, under the framework of a single fluid, we can incorporate tracer particles into the numerical simulation to identify and trace the source of material particles in the mixed region. Zhang et al. introduced tracer particles into DBM for the first time. They studied the flow and mixing of compressible RTI, and the description method based on the velocity phase space of tracer particles opened up a new perspective for analyzing and profoundly understanding the flow system Zhang et al. (2021). Later, Li et al. extended the phase space description of tracer particles from simple velocity phase space to position-velocity phase space and studied the evolution of multimode RTI Li et al. (2022). The distribution of tracer particles in position space provides a clear interface for the evolution of the light-heavy fluid interface. The distribution of tracer particles in position-velocity phase space further opens a new perspective for studying complex flow systems.

In previous studies, hydrodynamic instability problems have often been investigated by using ideal gas models to simplify the problems. However, in a real fluid, intermolecular interaction potentials are present. When the intermolecular interaction force is negligible compared to the external force acting on the medium during the physical processes, ideal gas models can help us capture the problem’s primary characteristics and provide a relatively satisfactory answer. Unfortunately, in many cases, intermolecular interactions cannot be negligible compared to external forces. For example, many physical phenomena observed in processes like ICF, such as microjets, tensile failure, melting phase transitions, and perturbation stabilization, are all related to the presence of interaction forces between medium molecules or atoms Miles (1966); Piriz et al. (2009); Li et al. (2021). In such instances, the effects of intermolecular interaction forces must be considered. For compressible RTI systems, the presence of intermolecular interaction potential induces a modification in the fluid media state equation, which directly affects the effect of compressibility on RTI’s evolution. Currently, research on this aspect is relatively limited due to its complexity. The van der Waals model represents the simplest form of intermolecular interaction potential. As a preliminary study, this paper considers the van der Waals interaction potential and investigates the effect of compressibility on the evolution of the van der Waals fluid single-mode RTI system from a statistical physics perspective with the help of tracer particles and morphological analysis. This paper is structured as follows: Section II introduces the numerical method, Section III covers numerical simulation and result analysis, and Section IV presents the conclusion and discussion.

II NUMERCAL METHODS

The DBM is a mesoscale physical model-building method and a complex physical field analysis method that was established and developed in recent yearsXu and Zhang (2022); Xu et al. (2024, 2015, 2018, 2021a, 2021b, 2021c); Gan et al. (2022a); Zhang et al. (2023). Unlike the NS method, DBM is based on the Boltzmann equation, making it convenient for DBM to describe the non-equilibrium behaviors of the system with more non-conservative moments and statistical mechanical quantities. At the same time, since it is not based on the assumption of a continuous medium and the higher-order non-equilibrium moment of the distribution function can be preserved as needed, DBM can study the near-continuous or discontinuous problems or higher-order non-equilibrium problems that the NS methods cannot or are not convenient to study.

It is worth pointing out that as the degree of discretization and non-equilibrium increases, the evolution of fluid instability systems is often accompanied by the emergence of small-scale structures and fast modes. The small-scale structure may lead to the failure of the continuous medium assumption. At this time, the validity of macroscopic hydrodynamic models such as NS is challenged. The flow of small-scale structures often exhibits behavior that appears to be “anomalous” (as opposed to the macroscopic continuum), such as anomalies in heat conductionWu and Liu (2011). Although a wealth of progress has been made on heat and mass transfer in small systemsWang et al. (2010), the small systems extensively studied in the literature of statistical physics and nonlinear science are often much smaller than the whole hydrodynamic instability system. In the face of this dilemma, DBM can better demonstrate its advantages.

It is known that, as illustrated in Fig. 1(a), numerical simulation research can be divided into three parts: (i) physical modeling; (ii) algorithm design; and (iii) numerical experimentation and complex physical field analysis. DBM mainly focuses on (i) and (iii). For algorithm design, the focus of DBM is not to build and design new numerical formats and discrete methods for velocity but to select existing numerical formats and methods for discrete velocity as a user. The construction of a physical model is the basis of DBM. The purpose and core of DBM are to effectively extract and analyze massive amounts of information in complex physical fields. The DBM numerical simulation flow chart is shown in Fig. 1(b).

Refer to caption
Figure 1: (a) DBM simulation study and (b) Flowchart of DBM simulation.

II.1 Physical modeling

Unlike the NS method, the model established by DBM is a mesoscale model based on the Boltzmann equation. The original Boltzmann equation is a high-dimensional differential integral equation, as shown in Eq.(1)Xu and Zhang (2022):

ft+𝐯f𝐫+𝐚f𝐯=+04π(ff1ff1)vrσ𝑑Ω𝑑𝐯1.𝑓𝑡𝐯𝑓𝐫𝐚𝑓𝐯superscriptsubscriptsuperscriptsubscript04𝜋superscript𝑓superscriptsubscript𝑓1𝑓subscript𝑓1subscript𝑣𝑟𝜎differential-dΩdifferential-dsubscript𝐯1.\frac{\partial f}{\partial t}+\mathbf{v}\cdot\frac{\partial f}{\partial\mathbf% {r}}+\mathbf{a}\cdot\frac{\partial f}{\partial\mathbf{v}}=\int_{-\infty}^{+% \infty}{\int_{0}^{4\pi}{\left({{f}^{*}}f_{1}^{*}-f{{f}_{1}}\right){{v}_{r}}% \sigma d\Omega d{{\mathbf{v}}_{1}}}}\text{.}divide start_ARG ∂ italic_f end_ARG start_ARG ∂ italic_t end_ARG + bold_v ⋅ divide start_ARG ∂ italic_f end_ARG start_ARG ∂ bold_r end_ARG + bold_a ⋅ divide start_ARG ∂ italic_f end_ARG start_ARG ∂ bold_v end_ARG = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 italic_π end_POSTSUPERSCRIPT ( italic_f start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - italic_f italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_σ italic_d roman_Ω italic_d bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT . (1)

Evidently, the collision term in the original Boltzmann equation is quite complex, and the direct solution presents significant challenges. Furthermore, the original Boltzmann equation only considers the hardball collision between molecules, a premise suited only for ideal gases. Thus, there arises a necessity to modify and simplify the original Boltzmann equation for a convenient and efficient description of fluid systems. For the van der Waals fluid, by introducing the intermolecular interaction potential and simplifying the collision term with the BGK model, the Boltzmann equation of the van der Waals fluid system in BGK form can be obtained:

ft+𝐯f𝐫+𝐚f𝐯=1τ(ffeq)+I,𝑓𝑡𝐯𝑓𝐫𝐚𝑓𝐯1𝜏𝑓superscript𝑓𝑒𝑞𝐼,\frac{\partial f}{\partial t}+\mathbf{v}\cdot\frac{\partial f}{\partial\mathbf% {r}}+\mathbf{a}\cdot\frac{\partial f}{\partial\mathbf{v}}=-\frac{1}{\tau}(f-{{% f}^{eq}})+I\text{,}divide start_ARG ∂ italic_f end_ARG start_ARG ∂ italic_t end_ARG + bold_v ⋅ divide start_ARG ∂ italic_f end_ARG start_ARG ∂ bold_r end_ARG + bold_a ⋅ divide start_ARG ∂ italic_f end_ARG start_ARG ∂ bold_v end_ARG = - divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG ( italic_f - italic_f start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT ) + italic_I , (2)

wheref,t,𝐯,𝐫,𝐚𝑓𝑡𝐯𝐫𝐚f,t,\mathbf{v},\mathbf{r},\mathbf{a}italic_f , italic_t , bold_v , bold_r , bold_a and τ𝜏\tauitalic_τ are molecular distribution function, time, velocity, spatial position, acceleration, and relaxation time, respectively. feq=ρ(12πRT)D/2exp[(𝐯𝐮)22RT]superscript𝑓𝑒𝑞𝜌superscript12𝜋𝑅𝑇𝐷2superscript𝐯𝐮22𝑅𝑇{{f}^{eq}}=\rho{{\left(\frac{1}{2\pi RT}\right)}^{{D}/{2}\;}}\exp\left[-\frac{% {{\left(\mathbf{v}-\mathbf{u}\right)}^{2}}}{2RT}\right]italic_f start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT = italic_ρ ( divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_R italic_T end_ARG ) start_POSTSUPERSCRIPT italic_D / 2 end_POSTSUPERSCRIPT roman_exp [ - divide start_ARG ( bold_v - bold_u ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_R italic_T end_ARG ] is the Maxwell distribution function. I=[A+𝐁(𝐯𝐮)+(C+Cq)|𝐯𝐮|2]feq𝐼delimited-[]𝐴𝐁𝐯𝐮𝐶subscript𝐶𝑞superscript𝐯𝐮2superscript𝑓𝑒𝑞I=-\left[A+\mathbf{B}\cdot\left(\mathbf{v}-\mathbf{u}\right)+\left(C+{{C}_{q}}% \right){{\left|\mathbf{v}-\mathbf{u}\right|}^{2}}\right]{{f}^{eq}}italic_I = - [ italic_A + bold_B ⋅ ( bold_v - bold_u ) + ( italic_C + italic_C start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ) | bold_v - bold_u | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_f start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT represents the intermolecular interactionGonnella et al. (2007); Xu and Zhang (2022), where

A=2(C+Cq)T,𝐴2𝐶+subscript𝐶𝑞𝑇,A=-2\left(C\text{+}{{C}_{q}}\right)T\text{,}italic_A = - 2 ( italic_C + italic_C start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ) italic_T , (3)
𝐁=1ρT[(PρT)𝐈+𝚲],𝐁1𝜌𝑇delimited-[]𝑃𝜌𝑇𝐈𝚲,\mathbf{B}=\frac{1}{\rho T}\nabla\cdot\left[\left(P-\rho T\right)\mathbf{I}+{% \bm{\Lambda}}\right]\text{,}bold_B = divide start_ARG 1 end_ARG start_ARG italic_ρ italic_T end_ARG ∇ ⋅ [ ( italic_P - italic_ρ italic_T ) bold_I + bold_Λ ] , (4)
C=𝐶absent\displaystyle C=italic_C = 12ρT2[(PρT)𝐮+𝚲:𝐮+aρ2𝐮\displaystyle\frac{1}{2\rho{{T}^{2}}}\left[\left(P-\rho T\right)\nabla\cdot% \mathbf{u}\right.+\bm{\Lambda}:\nabla\mathbf{u}+a{{\rho}^{2}}\nabla\cdot% \mathbf{u}divide start_ARG 1 end_ARG start_ARG 2 italic_ρ italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ ( italic_P - italic_ρ italic_T ) ∇ ⋅ bold_u + bold_Λ : ∇ bold_u + italic_a italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∇ ⋅ bold_u (5)
\displaystyle-- M(12ρρ𝐮+ρ(𝐮)+ρ𝐮ρ)],\displaystyle\left.M(\frac{1}{2}\nabla\rho\cdot\nabla\rho\nabla\cdot\mathbf{u}% +\nabla\rho\cdot\nabla\left(\nabla\cdot\mathbf{u}\right)+\nabla\rho\cdot\nabla% \mathbf{u}\cdot\nabla\rho)\right]\text{,}italic_M ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∇ italic_ρ ⋅ ∇ italic_ρ ∇ ⋅ bold_u + ∇ italic_ρ ⋅ ∇ ( ∇ ⋅ bold_u ) + ∇ italic_ρ ⋅ ∇ bold_u ⋅ ∇ italic_ρ ) ] ,
Cq=1ρT2(qρTT).subscript𝐶𝑞1𝜌superscript𝑇2𝑞𝜌𝑇𝑇.{{C}_{q}}=\frac{1}{\rho{{T}^{2}}}\nabla\cdot\left(q\rho T\nabla T\right)\text{.}italic_C start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_ρ italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∇ ⋅ ( italic_q italic_ρ italic_T ∇ italic_T ) . (6)

Where

𝚲=p1𝐈+Mρρ,𝚲subscript𝑝1𝐈𝑀𝜌𝜌,{\bm{\Lambda}}={{p}_{1}}\mathbf{I}+M\nabla\rho\nabla\rho\text{,}bold_Λ = italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT bold_I + italic_M ∇ italic_ρ ∇ italic_ρ , (7)
p1=ρTMTρρM2ρ+(ρMM)12|ρ|2,subscript𝑝1𝜌𝑇𝑀𝑇𝜌𝜌𝑀superscript2𝜌𝜌superscript𝑀𝑀12superscript𝜌2,{{p}_{1}}=-\rho T\nabla\frac{M}{T}\cdot\nabla\rho-\rho M{{\nabla}^{2}}\rho+(% \rho{M}^{\prime}-M)\frac{1}{2}{{\left|\nabla\rho\right|}^{2}}\text{,}italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = - italic_ρ italic_T ∇ divide start_ARG italic_M end_ARG start_ARG italic_T end_ARG ⋅ ∇ italic_ρ - italic_ρ italic_M ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ + ( italic_ρ italic_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_M ) divide start_ARG 1 end_ARG start_ARG 2 end_ARG | ∇ italic_ρ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (8)
P=ρT1bρaρ2.𝑃𝜌𝑇1𝑏𝜌𝑎superscript𝜌2.P=\frac{\rho T}{1-b\rho}-a{{\rho}^{2}}\text{.}italic_P = divide start_ARG italic_ρ italic_T end_ARG start_ARG 1 - italic_b italic_ρ end_ARG - italic_a italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (9)

Where M=K+HT𝑀𝐾𝐻𝑇M=K+HTitalic_M = italic_K + italic_H italic_T, M=(Mρ)Tsuperscript𝑀subscript𝑀𝜌𝑇{M}^{\prime}={{\left(\frac{\partial M}{\partial\rho}\right)}_{T}}italic_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = ( divide start_ARG ∂ italic_M end_ARG start_ARG ∂ italic_ρ end_ARG ) start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT, H𝐻Hitalic_H and K𝐾Kitalic_K are functions of ρ𝜌\rhoitalic_ρ, a𝑎aitalic_a and b𝑏bitalic_b are parameters related to the intermolecular interaction potential. It should be pointed out that the BGK-like linearization models in a series of current kinetic methods are not the models with the same name that were directly simplified by adding constraints to Boltzmann equation, they can only be the products of combining the kinetic theory and the mean field theory according to specific problemsXu and Zhang (2022); Xu et al. (2024); Gan et al. (2022b). After discretizing the velocity space of Eq.(2), the discrete model can be obtained asChen et al. (2022):

tfji+𝐯jifji𝐚(𝐯ji𝐮)RTfjieq=1τ[fjifjieq]+Iji.subscript𝑡subscript𝑓𝑗𝑖subscript𝐯𝑗𝑖subscript𝑓𝑗𝑖𝐚subscript𝐯𝑗𝑖𝐮𝑅𝑇superscriptsubscript𝑓𝑗𝑖𝑒𝑞1𝜏delimited-[]subscript𝑓𝑗𝑖superscriptsubscript𝑓𝑗𝑖𝑒𝑞subscript𝐼𝑗𝑖.{{\partial}_{t}}{{f}_{ji}}+{{\mathbf{v}}_{ji}}\cdot\nabla{{f}_{ji}}-\frac{% \mathbf{a}\cdot\left({{\mathbf{v}}_{ji}}-\mathbf{u}\right)}{RT}f_{ji}^{eq}=-% \frac{1}{\tau}\left[{{f}_{ji}}-f_{ji}^{eq}\right]+{{I}_{ji}}\text{.}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT + bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT ⋅ ∇ italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT - divide start_ARG bold_a ⋅ ( bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT - bold_u ) end_ARG start_ARG italic_R italic_T end_ARG italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT = - divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG [ italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT ] + italic_I start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT . (10)

Where fjisubscript𝑓𝑗𝑖{{f}_{ji}}italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT and fjieqsuperscriptsubscript𝑓𝑗𝑖𝑒𝑞f_{ji}^{eq}italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT are discrete distribution functions and discrete Maxwell distribution functions, respectively. Here, the distribution function f𝑓fitalic_f in the third term at the left end of Eq.(2) is approximately considered as feqsuperscript𝑓𝑒𝑞{{f}^{eq}}italic_f start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT during the discretization processSong et al. (2024). The schematic of discrete velocities used in this paper is shown in Fig. 23.

II.2 Complex physical field analysis

Establishing an accurate and satisfactory physical model is just the first step in numerical simulation. We can obtain a wealth of information and outcomes by accurately calculating the specific problems using the established model. How to extract useful information from massive results and conduct in-depth analyses of them is critical for us to know and understand physical problems. Unfortunately, due to the non-equilibrium system’s complexity, very little information has been effectively used, and it can be said that most of the information is dormant.

For the analysis of complex physical fields, the task of DBM is to extract more and more effective information from massive amounts of information. In addition to the well-known macroscopic quantities such as the pressure gradient p𝑝\nabla p∇ italic_p, temperature gradient T𝑇\nabla T∇ italic_T, density gradient ρ𝜌\nabla\rho∇ italic_ρ, and Knudsen number Kn𝐾𝑛Knitalic_K italic_n in macroscopic fluid mechanics, DBM describes how and to what extent the system deviates from equilibrium by using more non-conservative kinetic moments of distribution functions:

𝚫m,n=𝐌m,n(f)𝐌m,n(fEQ),subscript𝚫𝑚𝑛subscript𝐌𝑚𝑛𝑓subscript𝐌𝑚𝑛superscript𝑓𝐸𝑄,{{\bm{\Delta}}_{m,n}}={{\mathbf{M}}_{m,n}}\left(f\right)-{{\mathbf{M}}_{m,n}}% \left({{f}^{EQ}}\right)\text{,}bold_Δ start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT = bold_M start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f ) - bold_M start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f start_POSTSUPERSCRIPT italic_E italic_Q end_POSTSUPERSCRIPT ) , (11)
𝚫m,n=𝐌m,n(f)𝐌m,n(fEQ).subscriptsuperscript𝚫bold-′𝑚𝑛subscriptsuperscript𝐌𝑚𝑛𝑓subscriptsuperscript𝐌𝑚𝑛superscript𝑓𝐸𝑄.{{\bm{{\Delta}^{\prime}}}_{m,n}}={{\mathbf{{M}^{\prime}}}_{m,n}}\left(f\right)% -{{\mathbf{{M}^{\prime}}}_{m,n}}\left({{f}^{EQ}}\right)\text{.}bold_Δ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT = bold_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f ) - bold_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f start_POSTSUPERSCRIPT italic_E italic_Q end_POSTSUPERSCRIPT ) . (12)

In Eq.(11), 𝐌m,n(f)subscript𝐌𝑚𝑛𝑓{{\mathbf{M}}_{m,n}}\left(f\right)bold_M start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f ) and 𝐌m,n(fEQ)subscript𝐌𝑚𝑛superscript𝑓𝐸𝑄{{\mathbf{M}}_{m,n}}\left({{f}^{EQ}}\right)bold_M start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f start_POSTSUPERSCRIPT italic_E italic_Q end_POSTSUPERSCRIPT ) are the n-order tensors contracted from the m-order kinetic moments of the distribution function f𝑓fitalic_f and the equilibrium distribution function fEQsuperscript𝑓𝐸𝑄{{f}^{EQ}}italic_f start_POSTSUPERSCRIPT italic_E italic_Q end_POSTSUPERSCRIPT (is equal to feqsuperscript𝑓𝑒𝑞{{f}^{eq}}italic_f start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT for ideal gas) for the molecular velocity 𝐯𝐯\mathbf{v}bold_v, respectively. 𝚫m,nsubscript𝚫𝑚𝑛{{\bm{\Delta}}_{m,n}}bold_Δ start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT is called the thermo-hydrodynamic non-equilibrium (THNE) characteristic quantity, which describes the THNE characteristics of the system. 𝐌m,n(f)subscriptsuperscript𝐌𝑚𝑛𝑓{{\mathbf{{M}^{\prime}}}_{m,n}}\left(f\right)bold_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f ) and 𝐌m,n(fEQ)subscriptsuperscript𝐌𝑚𝑛superscript𝑓𝐸𝑄{{\mathbf{{M}^{\prime}}}_{m,n}}\left({{f}^{EQ}}\right)bold_M start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( italic_f start_POSTSUPERSCRIPT italic_E italic_Q end_POSTSUPERSCRIPT ) in Eq.(12) are the n-order tensors contracted from the m-order kinetic central moments of f𝑓fitalic_f and fEQsuperscript𝑓𝐸𝑄{{f}^{EQ}}italic_f start_POSTSUPERSCRIPT italic_E italic_Q end_POSTSUPERSCRIPT for the thermal fluctuation velocity 𝐯𝐮𝐯𝐮\mathbf{v}-\mathbf{u}bold_v - bold_u, respectively. 𝚫m,nsubscriptsuperscript𝚫bold-′𝑚𝑛{{\bm{{\Delta}^{\prime}}}_{m,n}}bold_Δ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT is called the thermodynamic non-equilibrium(TNE) characteristic quantity, which describes the TNE characteristics of the system. The TNE behavior neglected by NS is attracting more attention with time and is becoming one of the current hot research topicsXu and Zhang (2022); Xu et al. (2024); Cai et al. (2021); Shan et al. (2018); Qiu et al. (2024).

The non-conservative moments mentioned above describe the non-equilibrium behaviors and characteristics of the system from their individual perspectives. These non-conservative moments are related and complementary. Together, they form a relatively more complete picture of the system. However, because of the complexity of non-equilibrium systems, DBM incorporates morphological and thermodynamic characteristics, in addition to non-conservative moments, to study and describe complex physical fields.

Entropy is a crucial physical quantity in compression science, ICF, aerospace, and other fields and has an essential influence on the system’s evolution. In ICF, for instance, an increase in entropy will, on the one hand, make it harder to compact the target pill; however, an increase in entropy can prevent instability from develo** on the other hand. As a result, it is critical to control the entropy increase reasonably for ignition to succeed. With DBM, we can conveniently analyze the system’s entropy. Considering the contribution of the density gradient to entropy and internal energy, we can get the generalized form of the total entropy of the systemGonnella et al. (2007); Onuki (2005):

Sb=[ns(n,e)12H|n|2]𝑑𝐫,subscript𝑆𝑏delimited-[]𝑛𝑠𝑛𝑒12𝐻superscript𝑛2differential-d𝐫,{{S}_{b}}=\int{\left[ns\left(n,e\right)-\frac{1}{2}H{{\left|\nabla n\right|}^{% 2}}\right]}d\mathbf{r}\text{,}italic_S start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = ∫ [ italic_n italic_s ( italic_n , italic_e ) - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_H | ∇ italic_n | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_d bold_r , (13)

and the total internal energy of the system:

Eb=(e+12K|n|2)𝑑𝐫,subscript𝐸𝑏𝑒12𝐾superscript𝑛2differential-d𝐫,{{E}_{b}}=\int{(e+\frac{1}{2}K|\nabla n{{|}^{2}})}d\mathbf{r}\text{,}italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = ∫ ( italic_e + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_K | ∇ italic_n | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_d bold_r , (14)

where n𝑛nitalic_n is the particle-number density, s𝑠sitalic_s is the entropy per particle, and e𝑒eitalic_e is the internal energy per particle. The gradient terms in Eqs. (13) and (14) represent a decrease in entropy and an increase in internal energy because of the inhomogeneity of n𝑛nitalic_n, respectively; the space integrals extend to the entire computational domain. For van der Waals fluids, we can getXu and Zhang (2022):

dSbdt=(1T𝐣+1T𝚷:𝐮)d𝐫,\frac{d{{S}_{b}}}{dt}=\int{\left(\frac{1}{T}\nabla\cdot\mathbf{j}+\frac{1}{T}% \bm{\Pi}:\nabla\mathbf{u}\right)}d\mathbf{r}\text{,}divide start_ARG italic_d italic_S start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG = ∫ ( divide start_ARG 1 end_ARG start_ARG italic_T end_ARG ∇ ⋅ bold_j + divide start_ARG 1 end_ARG start_ARG italic_T end_ARG bold_Π : ∇ bold_u ) italic_d bold_r , (15)

where 𝐣𝐣\mathbf{j}bold_j and 𝚷𝚷\bm{\Pi}bold_Π are heat flux and viscous stress, respectively. Through the relationship between the material derivative and the temporal partial derivative, we can get the temporal partial derivative of the entropy density:

sbt=(𝐮sb1T𝐣)𝐣(1T)+1T𝚷:𝐮.:subscript𝑠𝑏𝑡𝐮subscript𝑠𝑏1𝑇𝐣𝐣1𝑇1𝑇𝚷𝐮.\frac{\partial{{s}_{b}}}{\partial t}=-\nabla\cdot\left(\mathbf{u}{{s}_{b}}-% \frac{1}{T}\mathbf{j}\right)-\mathbf{j}\cdot\nabla\left(\frac{1}{T}\right)+% \frac{1}{T}\bm{\Pi}:\nabla\mathbf{u}\text{.}divide start_ARG ∂ italic_s start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG = - ∇ ⋅ ( bold_u italic_s start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG italic_T end_ARG bold_j ) - bold_j ⋅ ∇ ( divide start_ARG 1 end_ARG start_ARG italic_T end_ARG ) + divide start_ARG 1 end_ARG start_ARG italic_T end_ARG bold_Π : ∇ bold_u . (16)

It can be seen that the entropy production rate can be divided into two parts. One part is contributed by heat flux:

S˙NOEF=𝐣1Td𝐫;subscript˙𝑆𝑁𝑂𝐸𝐹𝐣1𝑇𝑑𝐫;{{\dot{S}}_{NOEF}}=\int{\mathbf{j}\cdot\nabla\frac{1}{T}}d\mathbf{r}\text{;}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT = ∫ bold_j ⋅ ∇ divide start_ARG 1 end_ARG start_ARG italic_T end_ARG italic_d bold_r ; (17)

the other part is contributed by viscous stress:

S˙NOMF=1T𝚷:𝐮d𝐫.:subscript˙𝑆𝑁𝑂𝑀𝐹1𝑇𝚷𝐮𝑑𝐫.{{\dot{S}}_{NOMF}}=\int{-\frac{1}{T}\bm{\Pi}:\nabla\mathbf{u}}d\mathbf{r}\text% {.}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT = ∫ - divide start_ARG 1 end_ARG start_ARG italic_T end_ARG bold_Π : ∇ bold_u italic_d bold_r . (18)

Under the continuous limit, 𝐣=κT𝐣=𝜅𝑇\mathbf{j}\text{=}-\kappa\nabla Tbold_j = - italic_κ ∇ italic_T and 𝚷=μ[𝐮+(𝐮)T(𝐮)𝐈]𝚷𝜇delimited-[]𝐮superscript𝐮𝑇𝐮𝐈\bm{\Pi}=-\mu\left[\nabla\mathbf{u}+{{\left(\nabla\mathbf{u}\right)}^{T}}-% \left(\nabla\cdot\mathbf{u}\right)\mathbf{I}\right]bold_Π = - italic_μ [ ∇ bold_u + ( ∇ bold_u ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT - ( ∇ ⋅ bold_u ) bold_I ], where κ𝜅\kappaitalic_κ is the heat conductivity and μ𝜇\muitalic_μ is the viscosity coefficient. Substituting them into Eq.(17) and Eq.(18), respectively, we can get

S˙NOEF=κ|T|2T2𝑑𝐫,subscript˙𝑆𝑁𝑂𝐸𝐹𝜅superscript𝑇2superscript𝑇2differential-d𝐫,{{\dot{S}}_{NOEF}}=\int{\frac{\kappa{{\left|\nabla T\right|}^{2}}}{{{T}^{2}}}}% d\mathbf{r}\text{,}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT = ∫ divide start_ARG italic_κ | ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_d bold_r , (19)
S˙NOMF=μ[𝐮:𝐮+(𝐮)T:𝐮|𝐮|2]T𝑑𝐫.{{\dot{S}}_{NOMF}}=\int{\frac{\mu\left[\nabla\mathbf{u}:\nabla\mathbf{u}+{{% \left(\nabla\mathbf{u}\right)}^{T}}:\nabla\mathbf{u}-{{\left|\nabla\cdot% \mathbf{u}\right|}^{2}}\right]}{T}}d\mathbf{r}\text{.}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT = ∫ divide start_ARG italic_μ [ ∇ bold_u : ∇ bold_u + ( ∇ bold_u ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT : ∇ bold_u - | ∇ ⋅ bold_u | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] end_ARG start_ARG italic_T end_ARG italic_d bold_r . (20)

The morphological analysis method has gradually become an effective data analysis and information extraction technology for complex physical fields. By incorporating different morphological quantities into DBM, a series of new insights into the study of hydrodynamic instability systems have been obtainedGan et al. (2019); Chen et al. (2020). In this paper, we introduce the ratio of the area occupied by heavy fluids (Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT) and the length of the interface between light and heavy fluids (L𝐿Litalic_L) to provide a supplementary description of the evolution of the RTI.

In actuality, rather than relying only on variables with the same unit or dimension to describe the non-equilibrium state and behavior of the system together, DBM defines a more generalized non-equilibrium intensity, 𝐃𝐃\mathbf{D}bold_D, to describe the system’s non-equilibrium for the extraction and analysis of complex physical field information. This generalized non-equilibrium intensity can include various non-equilibrium parameters of the system, such as:

𝐃=𝐃absent\displaystyle\mathbf{D}=bold_D = {Kn,|ρ|,|T|,|p|,L,Ah,|𝚫m,n|,\displaystyle\{Kn,\lvert\nabla\rho\rvert,\lvert\nabla T\rvert,\lvert\nabla p% \rvert,L,{{A}_{h}},\lvert{{\bm{\Delta}}_{m,n}}\rvert,{ italic_K italic_n , | ∇ italic_ρ | , | ∇ italic_T | , | ∇ italic_p | , italic_L , italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT , | bold_Δ start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT | , (21)
|𝚫m,n|,|S˙NOEF|,|S˙NOMF|,}.\displaystyle\lvert{{{\bm{{\Delta}^{\prime}}}}_{m,n}}\rvert,\lvert{{{\dot{S}}}% _{NOEF}}\rvert,\lvert{{{\dot{S}}}_{NOMF}}\rvert,\cdots\}\text{.}| bold_Δ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT | , | over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT | , | over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT | , ⋯ } .

The specific parameters and the number of parameters contained in 𝐃𝐃\mathbf{D}bold_D can be selected according to the needs of specific problems.

II.3 Introduction of tracer particles

In cases of strong compressibility, the early approach of determining the fluid source based on density is invalid because the density of fluid from the upper layer (original weight) may be smaller than that of the surrounding (original light) fluid. Taking inspiration from PIV and PILF technologies in experimental fluid mechanics, introducing tracer particles into numerical simulation allows for the tracking and identification of fluid sources in mixed regions within the framework of a single fluid.

As is well known, the motion state of moving particles in fluid media can be described by the Stokes number (Stk𝑆𝑡𝑘Stkitalic_S italic_t italic_k):

Stk=u0τPl0.𝑆𝑡𝑘subscript𝑢0subscript𝜏𝑃subscript𝑙0.Stk=\frac{{{u}_{0}}\cdot{{\tau}_{P}}}{{{l}_{0}}}\text{.}italic_S italic_t italic_k = divide start_ARG italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⋅ italic_τ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT end_ARG start_ARG italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG . (22)

Where u0subscript𝑢0{{u}_{0}}italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the local flow velocity, τPsubscript𝜏𝑃{{\tau}_{P}}italic_τ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT is the relaxation time of particles, and l0subscript𝑙0{{l}_{0}}italic_l start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the characteristic size of particles. The Stokes number represents the inertia of particles moving in a fluid medium. The smaller the τPsubscript𝜏𝑃{{\tau}_{P}}italic_τ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT is, the easier it is for particles to follow the surrounding fluid. When τPsubscript𝜏𝑃{{\tau}_{P}}italic_τ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT is close to 0, Stk<<1much-less-than𝑆𝑡𝑘1Stk<<1italic_S italic_t italic_k < < 1, the particles completely follow the fluid’s movement. Therefore, if we set the volume and mass of moving particles very small in numerical simulation, the momentum exchange between particles and fluid medium will be completed instantly. The particle velocity will be completely determined by the local flow velocityZhang et al. (2021), as shown in Eq.(23). Then, the function of tracing particles can be realized numerically by marking the moving particles to distinguish the source fluid where the particles come from.

𝐮P(𝐫k)=Z𝐮(𝐫)δ(|𝐫𝐫k|)𝑑𝐫,subscript𝐮𝑃subscript𝐫𝑘subscript𝑍𝐮𝐫𝛿𝐫subscript𝐫𝑘differential-d𝐫,{{\mathbf{u}}_{P}}\left({{\mathbf{r}}_{k}}\right)=\int_{Z}{\mathbf{u}\left(% \mathbf{r}\right)}\cdot\delta\left(\left|\mathbf{r}-{{\mathbf{r}}_{k}}\right|% \right)d\mathbf{r}\text{,}bold_u start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = ∫ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT bold_u ( bold_r ) ⋅ italic_δ ( | bold_r - bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | ) italic_d bold_r , (23)

where 𝐮psubscript𝐮𝑝{{\mathbf{u}}_{p}}bold_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the particle velocity, 𝐫ksubscript𝐫𝑘{{\mathbf{r}}_{k}}bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the position of the particle, 𝐮(𝐫)𝐮𝐫\mathbf{u}\left(\mathbf{r}\right)bold_u ( bold_r ) is the local velocity at the spatial position 𝐫𝐫\mathbf{r}bold_r, Z𝑍Zitalic_Z is the region where the particle is affected by the local velocity of the fluid medium, δ𝛿\deltaitalic_δ is the Dirac function, and its discrete form ψ𝜓\psiitalic_ψ is usually used in numerical simulation. In the two-dimensional case, ψ𝜓\psiitalic_ψ can be expressed as:

ψ(𝐫i,j,𝐫k)=ψ(|𝐫i,j𝐫k|)=φ(Δrx)φ(Δry),𝜓subscript𝐫𝑖𝑗subscript𝐫𝑘𝜓subscript𝐫𝑖𝑗subscript𝐫𝑘𝜑Δsubscript𝑟𝑥𝜑Δsubscript𝑟𝑦,\psi\left({{\mathbf{r}}_{i,j}},{{\mathbf{r}}_{k}}\right)=\psi\left(\left|{{% \mathbf{r}}_{i,j}}-{{\mathbf{r}}_{k}}\right|\right)=\varphi(\Delta{{r}_{x}})% \cdot\varphi(\Delta{{r}_{y}})\text{,}italic_ψ ( bold_r start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT , bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = italic_ψ ( | bold_r start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | ) = italic_φ ( roman_Δ italic_r start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ) ⋅ italic_φ ( roman_Δ italic_r start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) , (24)

where φ𝜑\varphiitalic_φ is the weight function. In this paper, φ𝜑\varphiitalic_φ is selected asZhang et al. (2021):

φ(Δrx)={{1+cos[(Δrx/Δx)π/2]}/4,Δrx2Δx0,Δrx>2Δx.\varphi(\Delta{{r}_{x}})=\left\{\begin{matrix}\left\{1+\cos\left[(\Delta{{r}_{% x}}/\Delta x)\cdot\pi/2\right]\right\}/4,&\Delta{{r}_{x}}\leq 2\Delta x\\ 0,&\Delta{{r}_{x}}>2\Delta x\\ \end{matrix}\right.\text{.}italic_φ ( roman_Δ italic_r start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ) = { start_ARG start_ROW start_CELL { 1 + roman_cos [ ( roman_Δ italic_r start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT / roman_Δ italic_x ) ⋅ italic_π / 2 ] } / 4 , end_CELL start_CELL roman_Δ italic_r start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ≤ 2 roman_Δ italic_x end_CELL end_ROW start_ROW start_CELL 0 , end_CELL start_CELL roman_Δ italic_r start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT > 2 roman_Δ italic_x end_CELL end_ROW end_ARG . (25)

III
SIMULATIONS AND RESULTS

By introducing tracer particles into the DBM model with van der Waals potential, we study the evolution of a two-dimensional single-mode compressible RTI system under different compressibility conditions in a single fluid framework. The research details and analysis of the results are presented in this section.

III.1 Initialization and numerical specifications

Figure 2(a) shows the two-dimensional single-mode RTI system studied in this paper. In a computational domain with the width of [0,1]01\left[0,1\right][ 0 , 1 ] and the height of [4,4]44\left[-4,4\right][ - 4 , 4 ], the upper part is heavy fluid, and the lower part is light fluid. At the initial time, an initial disturbance yc(x)=y0cos(kx)subscript𝑦𝑐𝑥subscript𝑦0𝑘𝑥{{y}_{c}}\left(x\right)={{y}_{0}}\cos\left(kx\right)italic_y start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_x ) = italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos ( italic_k italic_x ) of amplitude y0=0.02subscript𝑦00.02{{y}_{0}}=0.02italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.02 exists at the light-heavy fluid interface that positioned at y=0𝑦0y=0italic_y = 0. k=2π/λ𝑘2𝜋𝜆k=2\pi/\lambdaitalic_k = 2 italic_π / italic_λ is the wave number, and λ=1𝜆1\lambda=1italic_λ = 1 is the wavelength of the disturbance. The density of heavy fluid at the interface is

ρ0(0,yc+(0))=1.subscript𝜌00superscriptsubscript𝑦c01.{{\rho}_{0}}\left(0,y_{\text{c}}^{+}\left(0\right)\right)=1\text{.}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 , italic_y start_POSTSUBSCRIPT c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( 0 ) ) = 1 . (26)

The density of light fluid at the interface is

ρ0(0,yc(0))=1×[(1At)/(1+At)].subscript𝜌00superscriptsubscript𝑦𝑐01delimited-[]1𝐴𝑡1𝐴𝑡.{{\rho}_{0}}\left(0,y_{c}^{-}\left(0\right)\right)=1\times\left[{\left(1-At% \right)}/{\left(1+At\right)}\;\right]\text{.}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 , italic_y start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) ) = 1 × [ ( 1 - italic_A italic_t ) / ( 1 + italic_A italic_t ) ] . (27)

yc+superscriptsubscript𝑦cy_{\text{c}}^{+}italic_y start_POSTSUBSCRIPT c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT represents the upper side of the disturbance interface and ycsuperscriptsubscript𝑦cy_{\text{c}}^{-}italic_y start_POSTSUBSCRIPT c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT represents the lower side of the disturbance interface. The temperatures of the light and heavy fluids are set to be:

T0(y)=subscript𝑇0𝑦absent\displaystyle\text{ }{{T}_{0}}\left(y\right)=italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_y ) = 1,yyc(x),1,𝑦subscript𝑦𝑐𝑥,\displaystyle 1\text{,}y\geq{{y}_{c}}\left(x\right)\text{,}1 , italic_y ≥ italic_y start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_x ) , (28)
T0(y)=subscript𝑇0𝑦absent\displaystyle\text{ }{{T}_{0}}\left(y\right)=italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_y ) = [1+aρ0(0,yc(0))][1bρ0(0,yc(0))]delimited-[]1𝑎subscript𝜌00superscriptsubscript𝑦𝑐0delimited-[]1𝑏subscript𝜌00superscriptsubscript𝑦𝑐0\displaystyle{\left[1+a\cdot{{\rho}_{0}}\left(0,y_{c}^{-}\left(0\right)\right)% \right]\cdot}{\left[1-b\cdot{{\rho}_{0}}\left(0,y_{c}^{-}\left(0\right)\right)% \right]}[ 1 + italic_a ⋅ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 , italic_y start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) ) ] ⋅ [ 1 - italic_b ⋅ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 , italic_y start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) ) ]
/ρ0(0,yc(0)),y<yc(x).\displaystyle/{{{\rho}_{0}}\left(0,y_{c}^{-}\left(0\right)\right)}\text{,}y<{{% y}_{c}}\left(x\right)\text{.}/ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 , italic_y start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( 0 ) ) , italic_y < italic_y start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_x ) .

Fluid in the flow field satisfies hydrostatic equilibrium, i.e.,

yP0(y)=gρ0(y).subscript𝑦subscript𝑃0𝑦𝑔subscript𝜌0𝑦.{{\partial}_{y}}{{P}_{0}}\left(y\right)=-g{{\rho}_{0}}\left(y\right)\text{.}∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_y ) = - italic_g italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_y ) . (29)

Substituting the van der Waals state equation, Eq.(9), into Eq.(29) gives:

yρ0(y)=g/2a12aTgT2aρ0(y)[1bρ0(y)]2,subscript𝑦subscript𝜌0𝑦𝑔2𝑎12𝑎𝑇𝑔𝑇2𝑎subscript𝜌0𝑦superscriptdelimited-[]1𝑏subscript𝜌0𝑦2,{{\partial}_{y}}{{\rho}_{0}}\left(y\right)={g}/{2a}\;-\frac{1}{2a}\cdot\frac{% Tg}{T-2a{{\rho}_{0}}\left(y\right){{\left[1-b{{\rho}_{0}}\left(y\right)\right]% }^{2}}}\text{,}∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_y ) = italic_g / 2 italic_a - divide start_ARG 1 end_ARG start_ARG 2 italic_a end_ARG ⋅ divide start_ARG italic_T italic_g end_ARG start_ARG italic_T - 2 italic_a italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_y ) [ 1 - italic_b italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_y ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (30)

where a=98𝑎98a=\frac{9}{8}italic_a = divide start_ARG 9 end_ARG start_ARG 8 end_ARG and b=13𝑏13b=\frac{1}{3}italic_b = divide start_ARG 1 end_ARG start_ARG 3 end_ARG in this paper.

Refer to caption
Figure 2: (a) The computational domain with [0,1]×[4,4]0144\left[0,1\right]\times\left[-4,4\right][ 0 , 1 ] × [ - 4 , 4 ]. Red represents heavy fluid, and blue represents light fluid. (b)--(e) The density profile at x=0𝑥0x=0italic_x = 0 across various Ma𝑀𝑎Maitalic_M italic_a conditions for At=0.1,0.25,0.5𝐴𝑡0.10.250.5At=0.1,0.25,0.5italic_A italic_t = 0.1 , 0.25 , 0.5, and 0.70.70.70.7, respectively.

Following the previous studiesFu et al. (2023, 2022); Wieland et al. (2019), the parameter settings and research results presented in this paper are dimensionless. The characteristic scales selected are the perturbation wave length λsuperscript𝜆{{\lambda}^{*}}italic_λ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, initial temperature TIsuperscript𝑇𝐼{{T}^{I*}}italic_T start_POSTSUPERSCRIPT italic_I ∗ end_POSTSUPERSCRIPT, density ρIsuperscript𝜌𝐼{{\rho}^{I*}}italic_ρ start_POSTSUPERSCRIPT italic_I ∗ end_POSTSUPERSCRIPT, and the isothermal speed of sound UPI/ρIsuperscript𝑈superscript𝑃𝐼superscript𝜌𝐼{{U}^{*}}\equiv\sqrt{{{{P}^{I*}}}/{{{\rho}^{I*}}}\;}italic_U start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≡ square-root start_ARG italic_P start_POSTSUPERSCRIPT italic_I ∗ end_POSTSUPERSCRIPT / italic_ρ start_POSTSUPERSCRIPT italic_I ∗ end_POSTSUPERSCRIPT end_ARG at the interface. The superscript ‘*’ means the dimensional physical quantities and the superscript ‘I𝐼Iitalic_I’ means the quantities at the initial interface. Several dimensionless numbers used in this paper are defined as followsFu et al. (2023, 2022):

At=ρhIρlIρhI+ρlI,Ma=gλPI/ρ,Re=ρUλμ,\displaystyle At=\frac{\rho{{{}_{h}^{I}}^{*}}-\rho{{{}_{l}^{I}}^{*}}}{\rho{{{}% _{h}^{I}}^{*}}+\rho{{{}_{l}^{I}}^{*}}},Ma=\sqrt{\frac{{{g}^{*}}{{\lambda}^{*}}% }{{{{P}^{I}}^{*}}/{{{\rho}^{*}}}\;}},Re=\frac{{{\rho}^{*}}{{U}^{*}}{{\lambda}^% {*}}}{{{\mu}^{*}}},italic_A italic_t = divide start_ARG italic_ρ start_FLOATSUBSCRIPT italic_h end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - italic_ρ start_FLOATSUBSCRIPT italic_l end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_ρ start_FLOATSUBSCRIPT italic_h end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT + italic_ρ start_FLOATSUBSCRIPT italic_l end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG , italic_M italic_a = square-root start_ARG divide start_ARG italic_g start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_P start_POSTSUPERSCRIPT italic_I end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / italic_ρ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG end_ARG , italic_R italic_e = divide start_ARG italic_ρ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_μ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG , (31)
ReP=ρλAt/(1+At)gλ/μ,Pr=CPμk.formulae-sequence𝑅subscript𝑒𝑃superscript𝜌superscript𝜆𝐴𝑡1𝐴𝑡superscript𝑔superscript𝜆superscript𝜇𝑃𝑟subscript𝐶𝑃superscript𝜇superscript𝑘\displaystyle{{Re}_{P}}={{{\rho}^{*}}{{\lambda}^{*}}\sqrt{{At}/{\left(1+At% \right){{g}^{*}}{{\lambda}^{*}}}\;}}/{{{\mu}^{*}}}\;,Pr={{C}_{P}}\frac{{{\mu}^% {*}}}{{{k}^{*}}}.italic_R italic_e start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = italic_ρ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT square-root start_ARG italic_A italic_t / ( 1 + italic_A italic_t ) italic_g start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG / italic_μ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_P italic_r = italic_C start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT divide start_ARG italic_μ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG .

Here, At𝐴𝑡Atitalic_A italic_t represents the light and heavy fluid density ratio at the interface, and Ma𝑀𝑎Maitalic_M italic_a characterizes the strength of the isothermal background stratificationReckinger et al. (2016); Livescu (2004).

Our current study examined the evolution of RTI across a range of At𝐴𝑡Atitalic_A italic_t from 0.1 to 0.7 and Ma𝑀𝑎Maitalic_M italic_a from 0.3 to 0.9. Figs. 2(b)--2(e) represent the density profile for different At𝐴𝑡Atitalic_A italic_t and Ma𝑀𝑎Maitalic_M italic_a at x=0𝑥0x=0italic_x = 0. It can be seen that with the increase in At𝐴𝑡Atitalic_A italic_t, the density difference between light and heavy fluids at the interface increases, and the density stratification becomes less pronounced. Other parameters are set as follows: Prandtl number Pr=0.72𝑃𝑟0.72Pr=0.72italic_P italic_r = 0.72 and disturbed Reynolds number ReP=1500𝑅subscript𝑒𝑃1500{{Re}_{P}}=1500italic_R italic_e start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = 1500. The heights of the bubble and spike are defined as the distance from the tips of the bubble and spike to y=0𝑦0y=0italic_y = 0, respectively. The bubble and spike velocities are defined as the derivatives of the bubble height and spike height over time, respectively. As done by Fu et al. and Bian et al., the time scale λ/(Atg)𝜆𝐴𝑡𝑔\sqrt{{\lambda}/{\left(Atg\right)}\;}square-root start_ARG italic_λ / ( italic_A italic_t italic_g ) end_ARG and the saturated velocity Atgλ/(1+At)𝐴𝑡𝑔𝜆1𝐴𝑡\sqrt{{Atg\lambda}/{\left(1+At\right)}\;}square-root start_ARG italic_A italic_t italic_g italic_λ / ( 1 + italic_A italic_t ) end_ARG are used to scale time and the bubble and spike velocities, respectivelyFu et al. (2023, 2022); Bian et al. (2020).

In this paper, the first-order forward Euler finite difference scheme is used for the discretization of the temporal derivative. The second-order non-oscillatory non-free dissipative (NND) scheme and the nine-point stencil schemeZhang et al. (2019); Tiribocchi et al. (2009) are used to discretize spatial derivative and the intermolecular interaction item I𝐼Iitalic_I, respectively. The variables ρ𝜌\rhoitalic_ρ, 𝐮𝐮\mathbf{u}bold_u, and e𝑒eitalic_e are fixed at their initial values at the borders of the y-direction, and periodic boundary conditions are applied along the x-directionFu et al. (2023); Luo et al. (2020); Reckinger et al. (2016). The grid length is set to be Δx=Δy=0.0039Δ𝑥Δ𝑦0.0039\Delta x=\Delta y=0.0039roman_Δ italic_x = roman_Δ italic_y = 0.0039.

III.2 RESULTS AND DISCUSSION

III.2.1 Bubble and spike velocity

Figures 3(a)--3(d) show the spike and bubble velocity under different Ma𝑀𝑎Maitalic_M italic_a conditions for At=0.1,0.25,0.5,𝐴𝑡0.10.250.5At=0.1,0.25,0.5,italic_A italic_t = 0.1 , 0.25 , 0.5 , and 0.7, respectively. The horizontal gray dashed line in each panel represents the bubble-saturated velocity predicted by the potential flow model.

Refer to caption
Figure 3: The spike and bubble velocity under various Ma𝑀𝑎Maitalic_M italic_a conditions for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7. The horizontal gray dashed line represents the bubble-saturated velocity predicted by potential flow models.

As indicated in Fig. 3(a), when At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, all bubble velocities are lower than the value predicted by the potential flow model. Following the initial acceleration, the velocities of the bubble and spike did not reach and maintain a saturated velocity but instead began to decrease continuously. As Ma𝑀𝑎Maitalic_M italic_a increases, the velocities of the bubble and spike decrease noticeably. Compressibility shows a significant inhibitory effect on bubble and spike velocity. When At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, the bubble and spike velocities are greater than those under the same Ma condition and the condition of At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, as depicted in Fig. 3(b). When Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, both bubbles and spikes are observed to reaccelerate. The inhibitory effect of compressibility on bubble and spike velocity diminished with the increase in At𝐴𝑡Atitalic_A italic_t. With the further increase of At𝐴𝑡Atitalic_A italic_t to 0.5, except for the case of Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9, both bubble and spike exhibit reacceleration behavior, as shown in Fig. 3(c). When At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7, the spike reaccelerates under all the considered Ma𝑀𝑎Maitalic_M italic_a values. It is noteworthy that the bubble only exhibits reacceleration at Ma=0.5𝑀𝑎0.5Ma=0.5italic_M italic_a = 0.5 and Ma=0.7𝑀𝑎0.7Ma=0.7italic_M italic_a = 0.7. The bubble has no reacceleration at a greater Ma𝑀𝑎Maitalic_M italic_a value of 0.9 and a lower Ma𝑀𝑎Maitalic_M italic_a value of 0.3. Before the RTI enters the bubble reacceleration stage, compressibility exhibits an inhibitory effect on spike velocity but a facilitating effect on bubble velocity.

In a compressible system, velocity can be divided into two parts: the compressible (irrotational) component caused by the fluid’s expansion and compression, and the solenoidal component, independent of the fluid’s density change. To determine the reason for the variation in the effect of compressibility on bubble and spike velocity, we employ the Helmholtz decomposition of the velocity fieldWang et al. (2013):

𝐮=𝐮C+𝐮S.𝐮subscript𝐮𝐶subscript𝐮𝑆.\mathbf{u}={{\mathbf{u}}_{C}}+{{\mathbf{u}}_{S}}\text{.}bold_u = bold_u start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT + bold_u start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT . (32)

Where 𝐮=(U,V)𝐮𝑈𝑉\mathbf{u}=\left(U,V\right)bold_u = ( italic_U , italic_V ), U𝑈Uitalic_U represents the velocity component in the x𝑥xitalic_x direction, and V𝑉Vitalic_V represents the velocity component in the y𝑦yitalic_y direction. 𝐮C=(UC,VC)subscript𝐮𝐶subscript𝑈𝐶subscript𝑉𝐶{{\mathbf{u}}_{C}}=\left({{U}_{C}},{{V}_{C}}\right)bold_u start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = ( italic_U start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) is the compressible (irrotational) component, and 𝐮S=(US,VS)subscript𝐮𝑆subscript𝑈𝑆subscript𝑉𝑆{{\mathbf{u}}_{S}}=\left({{U}_{S}},{{V}_{S}}\right)bold_u start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT = ( italic_U start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ) is the solenoidal component.

Figure 4 shows the contour plots of the compressive (VCsubscript𝑉𝐶{{V}_{C}}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, left half) and solenoidal (VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT, right half) components of vertical velocity under the condition of At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1. It can be seen that VCsubscript𝑉𝐶{{V}_{C}}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is two orders of magnitude smaller than VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT. This indicates that in the current movement, divergence-free movement is dominant. The velocity change caused by the fluid medium’s expansion and compression is minimal. It can be seen that with the increase in Ma𝑀𝑎Maitalic_M italic_a, VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT decreases significantly. Compressibility exerts a significant inhibitory effect on VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT. From a stress perspective, the primary reason for this phenomenon is that under the condition of At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, the density difference between the light and heavy fluids at the interface is slight. As the heavy fluid moves into the light fluid, the density of the light fluid around the spike quickly exceeds that of the heavy fluid inside the spike. The buoyancy of the heavy fluid increases rapidly. With the increase in Ma𝑀𝑎Maitalic_M italic_a, the density stratification intensifies, causing the spike to experience a stronger upward force while descending an equivalent distance. This resulted in a lower VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT at a higher Ma𝑀𝑎Maitalic_M italic_a. It can be observed that at higher Ma𝑀𝑎Maitalic_M italic_a, the pressure severely distorts the bubble and the spike, limiting their movement to a narrow region near the original interface.

Refer to caption
Figure 4: The contour plots of the compressive (VCsubscript𝑉𝐶{{V}_{C}}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, left half) and solenoidal (VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT, right half) components of vertical velocity for At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1. From (a) to (f): t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6666, respectively. From (1) to (4): Ma=0.3,0.5,0.7𝑀𝑎0.30.50.7Ma=0.3,0.5,0.7italic_M italic_a = 0.3 , 0.5 , 0.7, and 0.90.90.90.9, respectively. The dark lines represent the interface of the heavy and light fluid.

Interestingly, although VCsubscript𝑉𝐶{{V}_{C}}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is caused by the fluid’s expansion and compression, it decreases with an increase in Ma𝑀𝑎Maitalic_M italic_a at At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1. This is primarily due to the fact that the range of motion for both light and heavy fluids is significantly limited. Because of this limitation, the pressure changes of light and heavy fluids that occur during movement are also restricted, resulting in the restricted expansion and compression of both fluids.

Refer to caption
Figure 5: The contour plots of the compressive (VCsubscript𝑉𝐶{{V}_{C}}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, left half) and solenoidal (VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT, right half) components of vertical velocity for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7. From (a) to (f): t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6666, respectively. From (1) to (4): Ma=0.3,0.5,0.7𝑀𝑎0.30.50.7Ma=0.3,0.5,0.7italic_M italic_a = 0.3 , 0.5 , 0.7, and 0.90.90.90.9, respectively. The dark lines represent the interface of the heavy and light fluid.

With the increase in At𝐴𝑡Atitalic_A italic_t, the inhibitory effect of compressibility on the bubble and spike velocity is weakened. When At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7, compressibility even shows a promoting effect on bubble velocity. Fig. 5 shows the VCsubscript𝑉𝐶{{V}_{C}}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT contour plots for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7. It is demonstrated that VCsubscript𝑉𝐶{{V}_{C}}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is one order of magnitude smaller than VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT. This means that the movement of the bubble and spike is still dominated by VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT. However, with the increase in Ma𝑀𝑎Maitalic_M italic_a, the spike and bubble are not restricted in a narrow range near the initial interface under the condition of At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7. And there was no obvious difference in the spike amplitude across various Ma𝑀𝑎Maitalic_M italic_a conditions before t=3𝑡3t=3italic_t = 3.

We can try to understand why the inhibitory effect of compressibility was weakened with the help of Fig. 2(e). As shown in Fig. 2(e), there is a significant disparity in density between the light and heavy fluids at the interface at the initial time. The density of the light fluid at the interface is significantly lower than that of the heavy fluid at the interface. Thus, compared with the density of the heavy fluid at the initial interface, the density increase of light fluid with the decrease in height is slight, and the difference in density stratification of light fluid across various Ma𝑀𝑎Maitalic_M italic_a conditions is not apparent. Therefore, as the heavy fluid moves downward, the pressure increase is slight compared with the gravity of the spike. The difference in the blocking effect of light fluid on spikes under different Ma𝑀𝑎Maitalic_M italic_a conditions is not apparent, especially in the early stages. This is why the early movements of the spikes under different conditions of Ma𝑀𝑎Maitalic_M italic_a are almost the same, and the movements of the bubble and spike are not limited to a narrow range near the initial interface with the increase in Ma𝑀𝑎Maitalic_M italic_a.

With the further evolution of RTI, the heavy fluid keeps converging to the spike. As shown in Fig. 2(e), it is evident that with the increase in Ma𝑀𝑎Maitalic_M italic_a, there is a noticeable decrease in the density of the heavy fluid as the height increases. This causes the total mass of the heavy fluid accumulated within the spike to decrease with the increase in Ma𝑀𝑎Maitalic_M italic_a. Thus, during the downward motion of the spike, while the buoyancy of the spike remains relatively constant as Ma𝑀𝑎Maitalic_M italic_a increases, the spike’s gravitational force noticeably reduces as Ma𝑀𝑎Maitalic_M italic_a increases. This is the main reason there is an obvious decrease in spike velocity in the later period as Ma𝑀𝑎Maitalic_M italic_a increases.

Interestingly, in the case of At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7, the compressibility shows a promotion effect on the bubble velocity before the bubble enters the reacceleration stage. The bubble’s reacceleration phenomenon is only observed under the middle Ma𝑀𝑎Maitalic_M italic_a conditions of Ma=0.5𝑀𝑎0.5Ma=0.5italic_M italic_a = 0.5 and Ma=0.7𝑀𝑎0.7Ma=0.7italic_M italic_a = 0.7, while it is not observed when Ma𝑀𝑎Maitalic_M italic_a decreases to 0.3 and increases to 0.9. Fig. 5 illustrates that when At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7, the motion of the bubble and spike is still dominated by VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT. To conduct a more comprehensive examination of the reasons contributing to this intriguing phenomenon, we graph the VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT curves along the vertical lines across the tips of the bubble (x=1𝑥1x=1italic_x = 1, denoted by the dashed line) and the spike (x=0.5𝑥0.5x=0.5italic_x = 0.5, denoted by the solid line) at t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6, as shown in Fig. 6. The circles on the VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT profile indicate the vertical position of the bubble and spike tips at each moment.

Refer to caption
Figure 6: The VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT profiles along the vertical lines across the bubble (x=1𝑥1x=1italic_x = 1 , denoted by dash lines) and spike (x=0.5𝑥0.5x=0.5italic_x = 0.5 , denoted by solid lines) tips for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7. From (a) to (f): t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6666, respectively. The circles marked on the VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT profile denote the vertical position of the bubble and spike tips at each moment.

As illustrated in Fig. 6(a), at t=1𝑡1t=1italic_t = 1, the bubble and spike VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT increase as Ma𝑀𝑎Maitalic_M italic_a increases. This can be understood with the help of Fig. 7. Fig. 7 shows the pressure contour at the initial time for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 under various Ma𝑀𝑎Maitalic_M italic_a conditions. The dark line represents the interface of the heavy and light fluids. Initially, the pressure of the heavy fluid above the interface is uniform in the x direction. However, below the interface, the pressure in the middle of the light fluid is higher than that on both sides. This causes the heavy fluid in the spike and the light fluid below the spike to press and move toward the light fluid on both sides. With the increase in Ma𝑀𝑎Maitalic_M italic_a, this pressure difference increases, and the tendency of the middle zone fluid to press and move to both sides of the fluid increases. This causes the spike’s VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT to rise as Ma𝑀𝑎Maitalic_M italic_a increases. Simultaneously, the fluids on both sides move upward due to compression from the middle zone fluid. So, as Ma𝑀𝑎Maitalic_M italic_a increases, the VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT below the bubble also increases.

Refer to caption
Figure 7: The pressure contour at the initial time for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7. From (a) to (d): Ma=0.3,0.5,0.7𝑀𝑎0.30.50.7Ma=0.3,0.5,0.7italic_M italic_a = 0.3 , 0.5 , 0.7, and 0.9, respectively. The dark lines represent the interface of the heavy and light fluids

At t=2𝑡2t=2italic_t = 2 and t=3𝑡3t=3italic_t = 3, the bubble VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT still increases as Ma𝑀𝑎Maitalic_M italic_a increases, as shown in Figs. 6(b) and 6(c). However, unlike at time t=1𝑡1t=1italic_t = 1, the VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT below the tip of the bubble under different Ma𝑀𝑎Maitalic_M italic_a conditions is nearly the same at t=2𝑡2t=2italic_t = 2 and t=3𝑡3t=3italic_t = 3. Therefore, the increase in VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT at the bubble’s tip with an increase in Ma𝑀𝑎Maitalic_M italic_a at this time is not due to a difference in the compression of the bubble by the fluid below. In Fig. 2(e), we can see that as Ma𝑀𝑎Maitalic_M italic_a increases, the density of heavy fluid decreases rapidly with height. So, the pressure on the bubble tip decreases rapidly during the bubble-rising process. This is the primary cause of the growth in bubble VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT as Ma𝑀𝑎Maitalic_M italic_a increases.

After t=4𝑡4t=4italic_t = 4, the bubble VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT increases first, then decreases as Ma𝑀𝑎Maitalic_M italic_a increases. At this period, the bubble VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT is gradually influenced by the vorticity deposition at the tail of the Kelvin-Helmholtz (KH) vortex created by the upward curl of the spike. Compressibility has two effects on the transport of vorticity deposited at the tail of the KH vortex to the tip of the bubble. On the one hand, with the increase in Ma𝑀𝑎Maitalic_M italic_a, the length of the KH vortex (the distance from the tip of the spike to the tail of the KH vortex) decreases, which causes the tail of the KH vortex to be farther away from the tip of the bubble, as shown in Figs. 5(d)--5(f). On the other hand, with the increase in Ma𝑀𝑎Maitalic_M italic_a, the downward motion distance of the spike decreases, which causes the tail of the KH vortex to be closer to the tip of the bubble. Due to the influence of these two factors, when Ma𝑀𝑎Maitalic_M italic_a is high or low, the vorticity deposited at the tail of the KH vortex cannot be effectively transported to the tip of the bubble and thus fails to cause the reacceleration of the bubble. As a result, the bubble has no reacceleration in cases of high compressibility (Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9) and low compressibility (Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3). This is the reason why the bubble VSsubscript𝑉𝑆{{V}_{S}}italic_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT initially rises and subsequently falls as Ma𝑀𝑎Maitalic_M italic_a increases.

III.2.2 Proportion of area occupied by the heavy fluid

The amplitude and velocity of the bubble and spike are a convenient and intuitive way to describe the evolution of the RTI. However, they represent the evolution of RTI in a single direction and dimension; their depiction of the system is constrained. The non-equilibrium evolution of RTI systems, especially compressible RTI systems, is complex and multidimensional. There will be complex physical processes in the evolution of the compressible RTI system, such as the transformation among the expansion and compression work, the internal energy, the kinetic energy, and the potential energy. As a result, the fluid’s density, temperature, and velocity changed. By paying attention to the evolution of fluid density, we can better understand the physical mechanics of these complex processes. In the two-dimensional case, we can use the proportion of area occupied by the heavy fluid to describe the density variation of the heavy and light fluids.

Figures 8(a)--8(d) illustrate the evolution of the proportion of area occupied by the heavy fluid Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT under varying Ma𝑀𝑎Maitalic_M italic_a for At=0.1,0.25,0.5𝐴𝑡0.10.250.5At=0.1,0.25,0.5italic_A italic_t = 0.1 , 0.25 , 0.5, and 0.7, respectively. To intuitively correlate the evolution of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT with the evolution stages of RTI, we take the case of Ma=0.7𝑀𝑎0.7Ma=0.7italic_M italic_a = 0.7 as an example and insert the flow field contour at the corresponding time point in Figs. 8(a)--8(d).

Refer to caption
Figure 8: The evolution of the proportion of area occupied by the heavy fluid Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

As shown in Fig. 8, the evolution of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT over time presents specific stages across all At𝐴𝑡Atitalic_A italic_t conditions. The effect of compressibility on Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT shows clear differences across various At𝐴𝑡Atitalic_A italic_t conditions. To provide a clearer depiction of the temporal evolution stages of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, the evolution of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t is derived by taking the derivative of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT to time, as illustrated in Fig. 9. Similarly, we inserted the flow field contour under Ma=0.7𝑀𝑎0.7Ma=0.7italic_M italic_a = 0.7 at the corresponding time point in the corresponding positions in Figs. 9(b)--9(d).

Refer to caption
Figure 9: The evolution of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

In the following, with the help of Figs. 8 and 9, let’s try to understand the physical process of the compressible RTI’s evolution under various At𝐴𝑡Atitalic_A italic_t and Ma𝑀𝑎Maitalic_M italic_a conditions from the Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT perspective. Fig. 8(a) shows the evolution of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT under different Ma𝑀𝑎Maitalic_M italic_a conditions when At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1. It can be seen that when At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, the variation of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT over time is minimal. This is because the fluid’s expansion and compression are closely related to the mutual flow between light and heavy fluids. During RTI development, the heavy fluid moves downward into the light fluid, causing an increase in pressure that compresses the heavy fluid. Conversely, the light fluid flows upward into the heavy fluid, causing a gradual decrease in pressure and an outward expansion of the light fluid. In Fig. 2(b), we can see that when At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, the density difference between the light and heavy fluids at the initial interface is slight. Thus, during the downward movement of the heavy fluid, the density of the light fluid around the spike quickly exceeds that of the heavy fluid inside the spike. The spike’s buoyant force will soon exceed the gravitational force, forming an upward resultant force that inhibits the spike’s downward movement. As Ma𝑀𝑎Maitalic_M italic_a increases, the density and pressure gradients intensify. This led to a stronger inhibition effect on the spike’s downward movement. As shown in Fig. 4, under the conditions of Ma=0.5,0.7𝑀𝑎0.50.7Ma=0.5,0.7italic_M italic_a = 0.5 , 0.7, and 0.9, the bubble’s motion and the spike’s motion are confined to a very narrow range near the initial interface. Thus, the pressure change of the spike and bubble during the movement is minimal. Therefore, the expansion and compression of light and heavy fluids are limited. This explains why the Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT variation is so small. In the case of Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, it is evident that the spike moves downward a certain distance. However, due to the light fluid’s weak density stratification, the change in pressure within the range of the spike’s movement is still relatively small. So, under these circumstances, the expansion and compression of light and heavy fluids are also small.

When At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, the density difference between the light and heavy fluids at the initial interface increases, causing the bubble and spike’s motion range to no longer be as severely suppressed as when At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1. As a result, the Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT variation range increases.

As shown in Fig. 8(b), the effect of compressibility on Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT varies over time. This can be understood with the help of Fig. 9(b). According to Fig. 9(b), dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t is positive and increases with increasing Ma𝑀𝑎Maitalic_M italic_a before t=1𝑡1t=1italic_t = 1. This is mainly because, in the early stage, the heavy fluid within the spike expands to the surrounding light fluid due to the higher pressure, as shown in Fig. 7. This led to a positive value of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t during this process. Furthermore, the higher the Ma𝑀𝑎Maitalic_M italic_a, the greater the pressure difference and expansion. So, dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t increases with increasing Ma𝑀𝑎Maitalic_M italic_a. With the development of RTI, between t=1𝑡1t=1italic_t = 1 and 2, the spike accelerates downward into the light fluid. During this process, the pressure surrounding the spike accelerates to become larger, which makes the heavy fluid compressed and the Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT smaller. Moreover, as Ma𝑀𝑎Maitalic_M italic_a increases, so does the downward pressure gradient. As a result, the dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t is negative during this period and decreases as Ma𝑀𝑎Maitalic_M italic_a increases. After t=2𝑡2t=2italic_t = 2, an upward-curling KH vortex starts forming at the spike’s head, as shown in the flow field contour in Fig. 9(b). A portion of the heavy fluid expands through the upward-curling KH vortex, and its density decreases. At this stage, the evolution of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT is impacted by two factors. On the one hand, the spike’s downward movement causes the heavy fluid to be constantly compressed, reducing Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT. On the other hand, the formation of the KH vortex causes heavy fluids to curl and expand upward, which increases Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT. There is a competition between the two influences. During t=2𝑡2t=2italic_t = 2 to 3, the spike velocity gradually decreases, as shown in Fig. 3(b). This leads to a decreased influence of compression on the heavy fluid. So, dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t gradually increases at this stage, mainly due to the upward expansion of the heavy fluid through the KH vortex. After t=3𝑡3t=3italic_t = 3, the evolutionary trend of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t under different Ma𝑀𝑎Maitalic_M italic_a conditions exhibits differences. For Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t begins to decrease. However, in the cases of Ma=0.5,0.7𝑀𝑎0.50.7Ma=0.5,0.7italic_M italic_a = 0.5 , 0.7, and 0.9, dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t continues to increase. This is because, after t=3𝑡3t=3italic_t = 3, the evolution of the spike velocity under different Ma𝑀𝑎Maitalic_M italic_a conditions appears different. In the case of Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, the spike velocity starts to increase again, and the effect of the compression on the spike due to its accelerating downward motion gradually increases so that dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t decreases again. For cases of Ma=0.5,0.7𝑀𝑎0.50.7Ma=0.5,0.7italic_M italic_a = 0.5 , 0.7, and 0.9, the spike velocity keeps decelerating. This caused dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t to be primarily affected by the expansion of the heavy fluid through the KH vortex, resulting in a gradual increase. After t=4𝑡4t=4italic_t = 4, for the cases of Ma=0.7𝑀𝑎0.7Ma=0.7italic_M italic_a = 0.7 and 0.9, the spike and bubble velocity becomes negative. This is because the motion of the bubble and spike is limited to a specific range in these circumstances, as depicted in the flow field diagrams for t=4𝑡4t=4italic_t = 4 and 5 in Fig. 9(b). Therefore, the pressure variation on the spike and bubble over time is minimal, and dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t is nearly zero.

As At𝐴𝑡Atitalic_A italic_t increases further, the density difference between the light and heavy fluids at the initial interface becomes more pronounced. As a result, the restricted impact of compressibility on bubble and spike movement is further reduced. It can be seen that when At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, only in the case of Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9 does dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t gradually increase after t=3𝑡3t=3italic_t = 3. This is because the spike is inhibited from reaccelerating after t=3𝑡3t=3italic_t = 3, only in the case of Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9, as shown in Fig. 3. When At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7, under all Ma𝑀𝑎Maitalic_M italic_a values considered here, dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t gradually decreases at the later stages due to the reacceleration of the spike. Through the investigation of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t, it was found that the first minimum value of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t can serve as a criterion for the bubble velocity to reach the first maximal value. Here, we take the case under the conditions of At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 and Ma=0.7𝑀𝑎0.7Ma=0.7italic_M italic_a = 0.7 as an example to show the evolution of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t and bubble velocity, as depicted in Fig. 10.

Refer to caption
Figure 10: Evolution of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t and Vbubblesubscript𝑉𝑏𝑢𝑏𝑏𝑙𝑒{{V}_{bubble}}italic_V start_POSTSUBSCRIPT italic_b italic_u italic_b italic_b italic_l italic_e end_POSTSUBSCRIPT for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 and Ma=0.7𝑀𝑎0.7Ma=0.7italic_M italic_a = 0.7. The vertical grey dash-dot line represents the moment t=2.0𝑡2.0t=2.0italic_t = 2.0.

The evolution of non-equilibrium systems is multifaceted; as the saying goes, "It’s a range viewed from the face and peaks viewed from the side, assuming different shapes viewed from far and wide." The descriptions of the evolution of non-equilibrium systems from different perspectives are often complementary and related, but not interchangeable. The evolution of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t is closely linked to the developmental stage of RTI. Although the evolution of the bubble and spike velocity and the evolution of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t are interrelated and consistent, it can be seen that the evolution of Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t enhanced our understanding of the physical process of the system’s evolution.

III.2.3 Entropy production of the system

An increase in entropy is an important product of the evolution of a system from a non-equilibrium state to an equilibrium state. The entropy production rate is closely related to the non-equilibrium state of the system and its evolution. As shown in Eq.(16), the entropy production rate of the system can be divided into two parts: the part related to heat flow and the part related to viscous stress, denoted by S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT and S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT, respectively.

Refer to caption
Figure 11: The evolution of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.
Refer to caption
Figure 12: The evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

Figures 11(a)--11(d) show the evolution of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT under different Ma𝑀𝑎Maitalic_M italic_a for At=0.1,0.25,0.5𝐴𝑡0.10.250.5At=0.1,0.25,0.5italic_A italic_t = 0.1 , 0.25 , 0.5, and 0.7, respectively. It can be seen that the effect of At𝐴𝑡Atitalic_A italic_t and Ma𝑀𝑎Maitalic_M italic_a on the evolution of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT is complicated. As At𝐴𝑡Atitalic_A italic_t increases, the evolutionary trend of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT over time changes, and so does the effect of Ma𝑀𝑎Maitalic_M italic_a on S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT. As shown in Eq.(19), S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT can be written as a function of the square of the temperature gradient mode |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. So the evolution of the total flow field |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT can, to a certain extent, represent the evolution of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT. Figs. 12(a)--12(d) show the evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT under different Ma𝑀𝑎Maitalic_M italic_a for At=0.1,0.25,0.5𝐴𝑡0.10.250.5At=0.1,0.25,0.5italic_A italic_t = 0.1 , 0.25 , 0.5, and 0.7, respectively. It is evident that the evolution of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT is very consistent with the evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Therefore, we can try to figure out what causes the evolutionary pattern of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT by analyzing the evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Refer to caption
Figure 13: The contour plots of |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 under two Ma𝑀𝑎Maitalic_M italic_a conditions: (a) Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and (b) Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9. From (1) to (6): t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6, respectively. The dark lines represent the interface of the heavy and light fluids.

At the initial time, the heavy fluid is at a uniformly lower temperature, while the light fluid is at a uniformly higher temperature. The temperature gradient is only present at the interface of the light and heavy fluids. As RTI grows, the temperature gradient develops with the development of the interface of the light and heavy fluids. Fig. 13 shows the contour plots of |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 under Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9 conditions. The dark lines represent the interface of the heavy and light fluids. It can be seen that |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is mainly present at the interface, which indicates the tight association between |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and the evolution of the interface.

Refer to caption
Figure 14: The evolution of L𝐿Litalic_L for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

In addition to the interfacial length L𝐿Litalic_L, the heat transfer between light and heavy fluids, as well as the energy transfer between the internal energy (Ebsubscript𝐸𝑏{{E}_{b}}italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT) and the expansion and compression work (WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT), will also influence the evolution of |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. These impacts are coupled. Figs. 14(a)--14(d) illustrate the evolution of L𝐿Litalic_L under different Ma𝑀𝑎Maitalic_M italic_a for At=0.1,0.25,0.5,𝐴𝑡0.10.250.5At=0.1,0.25,0.5,italic_A italic_t = 0.1 , 0.25 , 0.5 , and 0.7, respectively. The increase in L𝐿Litalic_L has two effects on the evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. On the one hand, it increases the area of the system where the temperature gradient exists, which tends to increase |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT; on the other hand, it also increases the area of the system where the heat transfer exists, which increases the total heat transfer rate of the system and thus tends to decrease |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT results from the competition between increasing temperature gradient area and increasing total heat transfer rate. To better understand how the increase in L𝐿Litalic_L affects the evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, we calculate the time derivatives for both variables, as depicted in Figs. 15 and 16, respectively.

Refer to caption
Figure 15: The evolution of dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

Figures 14(a) and 15(a) show the evolution of L𝐿Litalic_L and dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t under different Ma𝑀𝑎Maitalic_M italic_a conditions for At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, respectively. It can be seen that before t=1𝑡1t=1italic_t = 1, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t is small, and L𝐿Litalic_L grows slowly with time. During this time, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is mainly reduced due to the heat transfer at the interface of the light and heavy fluids. Thus, d|T|total2/dt𝑑superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2𝑑𝑡{d\left|\nabla T\right|_{total}^{2}}/{dt}italic_d | ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_d italic_t is negative and gradually increases with time, as shown in Fig. 16(a).

Refer to caption
Figure 16: The evolution of d|T|total2/dt𝑑superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2𝑑𝑡{d\left|\nabla T\right|_{total}^{2}}/{dt}italic_d | ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_d italic_t for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

Between t=1𝑡1t=1italic_t = 1 and 2, in the case of lower compressibility with Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t increases rapidly. This led to a rapid increase in area where the temperature gradient exists. Therefore, d|T|total2/dt𝑑superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2𝑑𝑡{d\left|\nabla T\right|_{total}^{2}}/{dt}italic_d | ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_d italic_t is positive and increases rapidly at this time. This is why |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT experiences high growth following the initial period of decline. However, for the cases of Ma=0.5,0.7𝑀𝑎0.50.7Ma=0.5,0.7italic_M italic_a = 0.5 , 0.7, and 0.9, the growth of dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t is restricted by compressibility. As shown in Fig. 15(a), the growth of dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t decreases significantly as Ma𝑀𝑎Maitalic_M italic_a increases. Because the increase in L𝐿Litalic_L is insufficient to induce d|T|total2/dt𝑑superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2𝑑𝑡{d\left|\nabla T\right|_{total}^{2}}/{dt}italic_d | ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_d italic_t to increase to a positive value, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is still reduced, mainly due to heat transfer. When Ma=0.5,0.7𝑀𝑎0.50.7Ma=0.5,0.7italic_M italic_a = 0.5 , 0.7, and 0.9, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t gradually decreases between t=2𝑡2t=2italic_t = 2 and 6. Thus, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is primarily affected by heat transfer and decreases during this time interval.

For the lower compressibility case with Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, during the time interval t=2𝑡2t=2italic_t = 2 and 4, the growth of dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t gradually slows down and eventually reaches a certain value. During this period, d|T|total2/dt𝑑superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2𝑑𝑡{d\left|\nabla T\right|_{total}^{2}}/{dt}italic_d | ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_d italic_t is mainly affected by the increasing rate of heat transfer and decreases gradually. Between t=4𝑡4t=4italic_t = 4 and 6, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t gradually decreases. During this period, d|T|total2/dt𝑑superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2𝑑𝑡{d\left|\nabla T\right|_{total}^{2}}/{dt}italic_d | ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_d italic_t is primarily affected by heat transfer and is still a negative value. However, the gradual decline in dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t also results in a gradual weakening of the trend of increasing the overall heat transfer rate of the system. The growth of the overall heat transfer rate is mainly reduced by the gradual decrease in the temperature gradient. Thus, d|T|total2/dt𝑑superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2𝑑𝑡{d\left|\nabla T\right|_{total}^{2}}/{dt}italic_d | ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_d italic_t gradually increases to zero.

With the increase in At𝐴𝑡Atitalic_A italic_t, the inhibitory effect of compressibility on RTI’s development diminishes. From the perspective of the morphological quantity L𝐿Litalic_L (dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t), this is reflected by the increasing growth rate of L𝐿Litalic_L (dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t) with the increase in At𝐴𝑡Atitalic_A italic_t and the diminishing gap between L𝐿Litalic_L (dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t) under different Ma𝑀𝑎Maitalic_M italic_a conditions with increasing At𝐴𝑡Atitalic_A italic_t, as shown in Figs. 14 and 15. This can be used to explain why the evolutionary trend of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT varies with the increase in At𝐴𝑡Atitalic_A italic_t. As shown in Fig. 12, in the case of At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, only under the condition of Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 does |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT exhibit an increase after the initial decrease. With the increase in At𝐴𝑡Atitalic_A italic_t, the growth rate of L𝐿Litalic_L (dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t) increases obviously. Thus, in the cases of At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5 and 0.7, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT exhibits an increase after the initial decrease across all Ma𝑀𝑎Maitalic_M italic_a conditions.

Figures 12(b) and 12(c) demonstrate that the impact of compressibility on |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT can be categorized into two distinct stages in the cases of At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25 and 0.5. In the early stage, when |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT decreases with time, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT increases as Ma𝑀𝑎Maitalic_M italic_a increases, and after that, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT decreases as Ma𝑀𝑎Maitalic_M italic_a increases. Before t=1𝑡1t=1italic_t = 1, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t is very small. Thus, L𝐿Litalic_L grows slowly and is almost the same across various Ma𝑀𝑎Maitalic_M italic_a conditions. The difference in |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT under different Ma𝑀𝑎Maitalic_M italic_a conditions is mainly due to the difference in |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at the interface. In the previous analysis, we knew that the pressure of the fluid inside and below the spike was higher than that of the light fluid on both sides at the initial moment, as shown in Fig. 7. This causes the fluid in the spike and below the spike to press the light fluid on both sides. The greater the Ma𝑀𝑎Maitalic_M italic_a, the greater the pressure difference and the stronger the tendency to press. The heavy fluid inside the spike expands and presses the surrounding light fluid, and the degree of expansion and press increases with the increase in Ma𝑀𝑎Maitalic_M italic_a. The expansion of the heavy fluid resulted in a decrease in its temperature, whereas the compression of the light fluid led to an increase in its temperature. Thus, |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at the interface increases with increasing Ma𝑀𝑎Maitalic_M italic_a, as shown in Figs. 13(a1) and 13(b1). This is why |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT increases as Ma𝑀𝑎Maitalic_M italic_a increases before t=1𝑡1t=1italic_t = 1. After t=1𝑡1t=1italic_t = 1, due to the development of RTI, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t increases rapidly. Affected by the rapid increase in dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT begins to increase. At this time, due to the inhibitory effect of compressibility on dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t, |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT decreases with increasing Ma𝑀𝑎Maitalic_M italic_a.

In fact, as At𝐴𝑡Atitalic_A italic_t increases, the bubble’s and spike’s motion range increases. This led to an increase in WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT of the heavy and light fluids. This, in turn, makes the effect of the energy transformation between Ebsubscript𝐸𝑏{{E}_{b}}italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT and WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT on the evolution of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT evident. As shown in Figs. 13(a3)--13(a5) and Figs. 13(b3)--13(b5), it can be seen that the magnitude of |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at the interface is obviously different. This resulted from the difference in the transformation between Ebsubscript𝐸𝑏{{E}_{b}}italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT and WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT in the cases of Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9. The initial rise and subsequent decline of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT with Ma𝑀𝑎Maitalic_M italic_a during t=2.5𝑡2.5t=2.5italic_t = 2.5 to 4.5, as depicted in Fig. 12(d), is also due to the difference in the transformations between Ebsubscript𝐸𝑏{{E}_{b}}italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT and WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT under different Ma𝑀𝑎Maitalic_M italic_a conditions. To figure out how Ma𝑀𝑎Maitalic_M italic_a affects the value of |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, let’s use the scenario depicted in Fig. 12(d) at t=4𝑡4t=4italic_t = 4 as an example. This is when |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT shows a pronounced non-monotonic variation evolution with Ma𝑀𝑎Maitalic_M italic_a.

Refer to caption
Figure 17: The contour plots of (a) |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and (b) T𝑇Titalic_T for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 at t=4𝑡4t=4italic_t = 4. From (1) to (3): Ma=0.3,0.5𝑀𝑎0.30.5Ma=0.3,0.5italic_M italic_a = 0.3 , 0.5, and 0.9, respectively.

Figures 17(a1)--17(a3) show the contour plots of |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at moment t=4𝑡4t=4italic_t = 4 for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 under the conditions of Ma=0.3,0.5𝑀𝑎0.30.5Ma=0.3,0.5italic_M italic_a = 0.3 , 0.5, and 0.9, respectively. It can be seen that the length L𝐿Litalic_L of the interface is almost the same for Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9. The difference in |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is mainly caused by the difference in |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at the interface, as shown in the red boxes in Figs. 17(a1) and 17(a2). Figs. 17(b1)--17(b3) show the contour plots of T𝑇Titalic_T at t=4𝑡4t=4italic_t = 4 for Ma=0.3,0.5𝑀𝑎0.30.5Ma=0.3,0.5italic_M italic_a = 0.3 , 0.5, and 0.9, respectively. It is observed that as Ma𝑀𝑎Maitalic_M italic_a increases, the compression of the heavy fluid inside the spike head increases during the downward movement of the spike. This leads to a higher internal energy and temperature for the heavy fluid inside the spike head. Meanwhile, the heavy fluid in the spike head curls and expands upward through the KH vortex. The upward expansion of heavy fluid through the KH vortex is primarily governed by two causes. On the one hand, compressibility has an inhibitory effect on the development of the KH vortex, which restricts the upward expansion; on the other hand, as compressibility increases, the degree of upward expansion of the light fluid below the bubble increases, which promotes the expansion of the heavy fluid within the KH vortex into the surrounding light fluid.

When Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, compressibility has a relatively weak inhibitory effect on the development of the KH vortex; however, at the same time, the degree of upward expansion of the light fluid through the bubble is also limited. Thus, during the upward expansion, the heavy fluid is compressed by the light fluid. As a result, the temperature of the heavy fluid in the middle and tail of the KH vortex increases. As Ma𝑀𝑎Maitalic_M italic_a increases to 0.5, the upward expansion of the light fluid through the bubble intensifies. This results in the light fluid exerting less compression on the upwardly expanded heavy fluid. As a result, the temperature increase of the heavy fluid in the middle and tail of the KH vortex decreases relative to that under the condition of Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3. This lower increase in temperature leads to an increase in |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at the interface. As Ma𝑀𝑎Maitalic_M italic_a increases to 0.9, the expansion degree of the light fluid in the bubble further increases. Thus, the compression effect of the light fluid on the upwardly expanded heavy fluid further decreases, which results in a further decrease in the temperature increase of the heavy fluid in the middle and tail of the KH vortex. However, the expansion of the light fluid leads to a more noticeable decrease in its temperature. This results in lower |T|2superscript𝑇2{{\left|\nabla T\right|}^{2}}| ∇ italic_T | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at the KH vortex interface compared to the case with Ma=0.5𝑀𝑎0.5Ma=0.5italic_M italic_a = 0.5. In addition, at Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9, it is evident that increasing compressibility greatly suppresses the formation of KH vortices, leading to a significant reduction in the size of the KH vortex. This also contributes to the considerable decrease in |T|total2superscriptsubscript𝑇𝑡𝑜𝑡𝑎𝑙2\left|\nabla T\right|_{total}^{2}| ∇ italic_T | start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Refer to caption
Figure 18: The evolution of S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

Figures 18(a)--18(d) show the evolution of S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT under different Ma𝑀𝑎Maitalic_M italic_a conditions for At=0.1,0.25,0.5𝐴𝑡0.10.250.5At=0.1,0.25,0.5italic_A italic_t = 0.1 , 0.25 , 0.5, and 0.7, respectively. It can be seen that for At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT across various Ma𝑀𝑎Maitalic_M italic_a conditions all exhibit an evolutionary pattern of first increasing and then decreasing with time. As At𝐴𝑡Atitalic_A italic_t increases, the evolutionary pattern of S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT gradually changes to increase all the time. However, with the increase in At𝐴𝑡Atitalic_A italic_t, the effect of Ma𝑀𝑎Maitalic_M italic_a on S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT does not change; S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT decreases with the increase in Ma𝑀𝑎Maitalic_M italic_a for all At𝐴𝑡Atitalic_A italic_t.

Refer to caption
Figure 19: The evolution of (𝐮:𝐮)total{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

As shown in Eq.(20), S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT can be written as a function of the square of the second-order reduction of the velocity gradient, denoted as 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u. So, the evolution of the total flow field 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u can, to a certain extent, represent the evolution of S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT. Figs 19(a)--19(d) show the evolution of (𝐮:𝐮)total{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT under different Ma𝑀𝑎Maitalic_M italic_a conditions for At=0.1,0.25,0.5𝐴𝑡0.10.250.5At=0.1,0.25,0.5italic_A italic_t = 0.1 , 0.25 , 0.5, and 0.7, respectively. The correlation between the evolution of S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT and the evolution of (𝐮:𝐮)total{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT. To gain a better understanding of the evolution of (𝐮:𝐮)total{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT, we calculate the derivative of (𝐮:𝐮)total{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT with respect to time, as depicted in Fig. 20.

Refer to caption
Figure 20: The evolution of d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t for (a) At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, (b) At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25, (c) At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5, and (d) At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7.

The velocity gradient in the evolution of RTI is mostly a result of the relative motion between the light and heavy fluids, which is predominantly observed at the interface between the two fluids. Therefore, in addition to the magnitude of the relative motion velocity, the evolution of L𝐿Litalic_L also has an important effect on the evolution of (𝐮:𝐮)total{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT. Firstly, we take the case with At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1 and Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 as an example to provide a preliminary understanding of the physical process involved in the evolution of (𝐮:𝐮)total{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT.

Figs. 21(a)--21(f) show the contour plots of 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u under the conditions of At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1 and Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 at t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6, respectively. It is clear that 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u exists primarily where there is a large tangential velocity difference between the light and heavy fluids at the interface. Between t = 0 and 1, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t grows very slowly. Hence, the gradual increase in 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u is mainly due to the slow increase in tangential velocity difference between the light and heavy fluids at the interface, as shown in Fig. 21(a). During the time interval of t=1𝑡1t=1italic_t = 1 and 2, the heavy fluid undergoes a downward acceleration to enter the light fluid. In contrast, the light fluid accelerates upward to enter the heavy fluid. Consequently, the tangential velocity difference between light and heavy fluids experiences a rapid increase, and simultaneously, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t also experiences a rapid increase, resulting in an acceleration increase of d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t, as depicted in Fig. 20(a).

After t=2𝑡2t=2italic_t = 2, both the bubble and spike decelerate, leading to a weakening of the relative motion between the light and heavy fluids. Consequently, there is a gradual decrease in d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t with time. However, as can be seen in Fig. 15(a), dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t continues to increase during t=2𝑡2t=2italic_t = 2 to 3 due to the upward curling of the KH vortex at the spike, which causes d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t to remain positive at this period, although it gradually decreases with time. Between t=3𝑡3t=3italic_t = 3 and 4, dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t ceases to increase, and the motion of the bubble and spike continues to diminish, which causes d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t to continue to decrease to a value less than zero. After t=4𝑡4t=4italic_t = 4, d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t gradually increases again. This is mainly attributed to the formation of numerous small-scale vortex structures near the interface, as shown in Figs. 21(e)--21(f). The decreasing scale of the vortex structures retards the decay of the 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u.

Refer to caption
Figure 21: The contour plots of 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u for At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1 and Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3. From (a) to (f): t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6, respectively.

As At𝐴𝑡Atitalic_A italic_t increases to 0.25, the evolution of d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t under the condition of Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 shows some differences in the later stages, as shown in Fig. 20(b). It can be seen that in the case of At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25 and Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t gradually increases again during the time interval of t=3𝑡3t=3italic_t = 3 and 4, after a brief decline during t=2𝑡2t=2italic_t = 2 to 3. This can be understood with the assistance of Fig. 3(b), which shows that after a short period of decrease in the spike velocity during t=2𝑡2t=2italic_t = 2 to 3, there is a reacceleration behavior of the spike during t=3𝑡3t=3italic_t = 3 to 4. The downward reacceleration of the spike causes the velocity difference between the heavy and light fluids to increase again, which is the main reason why d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t increases again during this period.

Refer to caption
Figure 22: The contour plots of 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 under two Ma𝑀𝑎Maitalic_M italic_a conditions: (a) Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, (b) Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9. From (1)--(6): t=1,2,3,4,5𝑡12345t=1,2,3,4,5italic_t = 1 , 2 , 3 , 4 , 5, and 6, respectively. The dark lines represent the interface of the heavy and light fluids.

Figures 20(c) and 20(d) show the evolution of d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t for At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5 and 0.7, respectively. It can be seen that as At𝐴𝑡Atitalic_A italic_t further increases, d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t appears to increase rapidly again at later times at low Ma𝑀𝑎Maitalic_M italic_a, as shown in Fig. 20(c) for Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and in Fig. 20(d) for Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and Ma=0.5𝑀𝑎0.5Ma=0.5italic_M italic_a = 0.5. From Figs. 3 and 15, it can be seen that when At=0.5𝐴𝑡0.5At=0.5italic_A italic_t = 0.5 and 0.7, the reacceleration of the spike and a rapid increase in dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t do not only appear in the case under the lower Ma𝑀𝑎Maitalic_M italic_a conditions. For example, when At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7, there is a reacceleration of the spike and a rapid increase in dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t under all Ma𝑀𝑎Maitalic_M italic_a conditions. However, the rapid increase in d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t at later times appears only under the conditions of Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and 0.5. So, in this circumstance, the rapid increase in d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t at a later stage is not only determined by the reacceleration of the spike and the rapid increase in dL/dt𝑑𝐿𝑑𝑡{dL}/{dt}italic_d italic_L / italic_d italic_t. To understand the underlying reason for this behavior, we make the contour plots of 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u for At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 under Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3 and Ma=0.9𝑀𝑎0.9Ma=0.9italic_M italic_a = 0.9 conditions, as depicted in Fig. 22. It can be seen that, although the difference in L𝐿Litalic_L between the two cases is small, it is evident that there is a significant difference in the value of 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u at the interface after t=3𝑡3t=3italic_t = 3. Due to the inhibition effect of compressibility on the deposition of vorticity at the interface, the value of 𝐮:𝐮:𝐮𝐮\nabla\mathbf{u}:\nabla\mathbf{u}∇ bold_u : ∇ bold_u under higher Ma𝑀𝑎Maitalic_M italic_a is smaller than that under lower Ma𝑀𝑎Maitalic_M italic_a. This phenomenon explains the absence of reacceleration of d(𝐮:𝐮)total/dt{d{{\left(\nabla\mathbf{u}:\nabla\mathbf{u}\right)}_{total}}}/{dt}italic_d ( ∇ bold_u : ∇ bold_u ) start_POSTSUBSCRIPT italic_t italic_o italic_t italic_a italic_l end_POSTSUBSCRIPT / italic_d italic_t under high Ma𝑀𝑎Maitalic_M italic_a conditions.

IV Conclusions and discussions

The relatively little research on strongly compressible fluid systems within the field of RTI has resulted in a deficiency in understanding their physical processes and mechanics. In the case of strong compressibility, the density of the fluid from the upper layer (originally heavy fluid) may become smaller than that of the surrounding (originally light) fluid, thus invalidating the early method of distinguishing light and heavy fluids based on density. In this paper, tracer particles are incorporated into a single-fluid discrete Boltzmann method (DBM) model that considers the van der Waals potential. By using tracer particles to label the matter-particle sources (where the matter particle comes from: the originally heavy fluid or the originally light fluid), the study of the compressible RTI’s evolution is realized in a single-fluid framework. The focus of DBM is on physical modeling before simulation and complex physical field analysis after simulation. Our simulation results contain two parts: one that can be given by the NS model and one that cannot be given or is not conveniently given by the NS model. A series of methods for analyzing complex physical fields are provided by DBM, allowing the study to be carried out in depth.

It is found that compressibility has an inhibitory effect on the spike velocity, and the inhibitory effect decreases with increasing At𝐴𝑡Atitalic_A italic_t. Unlike spike velocity, the influence of compressibility on bubble velocity shows a staged behavior with increasing At𝐴𝑡Atitalic_A italic_t. There exists a critical Atwood number AtC𝐴subscript𝑡𝐶{At}_{C}italic_A italic_t start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. For At𝐴𝑡Atitalic_A italic_t values below AtC𝐴subscript𝑡𝐶{At}_{C}italic_A italic_t start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, compressibility exhibits an inhibitory effect on bubble velocity, and the strength of the inhibition decreases with increasing At𝐴𝑡Atitalic_A italic_t. However, when At=AtC𝐴𝑡𝐴subscript𝑡𝐶At={At}_{C}italic_A italic_t = italic_A italic_t start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, compressibility is observed to promote the bubble velocity before the bubble reacceleration stage. Interestingly, due to the competition between the inhibitory effect of compressibility on the downward motion of the spike and its inhibitory effect on the development of the KH vortex, the bubble does not exhibit reaccelerating behavior for either higher or lower compressibility. For the specific scenario studied in this paper, the AtC𝐴subscript𝑡𝐶{At}_{C}italic_A italic_t start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is about 0.7. The amplitude and velocity of the bubble and spike are a convenient and intuitive way to describe the evolution of the RTI. However, they represent the evolution of RTI in a single direction and dimension; their depiction of the system is constrained. It is worth pointing out that these results can also be given using the NS model.

The non-equilibrium evolution of RTI systems, especially compressible RTI systems, is complex and multidimensional. It has been discovered that for compressible RTI systems, the variation in fluid density is closely related to the stages of RTI evolution. In the two-dimensional case, we can express the change in the densities of the light and heavy fluids in relation to the evolution of the area proportion occupied by the heavy fluid, denoted as Ahsubscript𝐴{{A}_{h}}italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT. It is found that the first minimum point of dAh/dt𝑑subscript𝐴𝑑𝑡{d{{A}_{h}}}/{dt}italic_d italic_A start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_d italic_t can serve as a criterion for identifying the moment when the bubble velocity reaches its initial maximum value. The evolution of entropy production rate related to heat conduction (S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT) is a complex process influenced by multiple factors, including (i) length of the interface between light and heavy fluids L𝐿Litalic_L, (ii) heat transfer rate, and (iii) the interconversion between the work performed by expansion and compression of light and heavy fluids WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT and internal energy Ebsubscript𝐸𝑏{{E}_{b}}italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. The At𝐴𝑡Atitalic_A italic_t, Ma𝑀𝑎Maitalic_M italic_a, and the stage of the evolution of RTI collectively determine the evolution of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT. Under conditions of low At𝐴𝑡Atitalic_A italic_t and high Ma𝑀𝑎Maitalic_M italic_a, the bubble and spike movements are confined to a narrow region around the initial interface, so S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT is mainly affected by heat transfer and gradually decreases. As At𝐴𝑡Atitalic_A italic_t increases, the inhibitory effect of compressibility on the motions of the bubble and spike diminishes. This results in an increase in L𝐿Litalic_L and WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT, which in turn causes an increase in the impact of WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT and the variation of WCEsubscript𝑊𝐶𝐸{{W}_{CE}}italic_W start_POSTSUBSCRIPT italic_C italic_E end_POSTSUBSCRIPT on S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT. Under these circumstances, the evolutionary trend of the rate of S˙NOEFsubscript˙𝑆𝑁𝑂𝐸𝐹{{\dot{S}}_{NOEF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_E italic_F end_POSTSUBSCRIPT is characterized by a decrease, followed by an increase, and then another decrease over time. Entropy production rate related to viscous stress (S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT) is mainly influenced by the interface length and the tangential velocity difference between light and heavy fluids. Compressibility exerts a suppressive effect on S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT for all At𝐴𝑡Atitalic_A italic_t values. For low At𝐴𝑡Atitalic_A italic_t, due to the inhibitory effect on bubble and spike movements, S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT demonstrates an evolutionary trend of initially growing and subsequently declining over time. As the value of At𝐴𝑡Atitalic_A italic_t grows, the movements of the bubble and spike grow more pronounced; thus, S˙NOMFsubscript˙𝑆𝑁𝑂𝑀𝐹{{\dot{S}}_{NOMF}}over˙ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_N italic_O italic_M italic_F end_POSTSUBSCRIPT tends to exhibit continuous growth over time. The RTI system is a complex, non-equilibrium system, various perspectives on non-equilibrium evolution are interrelated and complementary, and they can form a more complete image of the system by combining each other. The incorporation of these complex physical field analysis techniques is a good complement to the description of RTI evolution. However, these results cannot be or are not conveniently given by the NS model. The DBM’s complex physical field analysis method makes this part of the study possible or easier.

In this work, we started with a relatively simple case, a two-dimensional single-mode RTI. The two-dimensional problem can be considered a special case of the three-dimensional problem, in which all physical quantities are uniform and unchanged in the third dimension (Z). Due to the additional third dimension, there will be some differences between the three- and two-dimensional cases. In three-dimensional problems, the change in densities of light and heavy fluids will be expressed as the change in volume occupied by the heavy fluid. The interface between light and heavy fluids is described by the interface area, rather than the interface length. For a three-dimensional single-mode RTI problem, there should also be a turning point in the change rate of the proportion of the volume occupied by the heavy fluid, which can serve as a criterion for the bubble velocity reaching its initial maximum value. Non-equilibrium quantities, such as the system’s temperature gradient and velocity gradient and their change rates, will also show different variation patterns at various stages of RTI development. The factors that determine their evolutionary patterns are the same as in the two-dimensional case. These are consistent with the two-dimensional, single-mode RTI. However, even in simple three-dimensional, single-mode scenarios, the coupling between perturbations in various directions makes the evolution of the heavy fluid volume and the interface more complex, necessitating further study.

The evolution of RTI systems in late time may exhibit different behavior from that in early time, but late-time behavior studies require either using sufficiently large systems or considering the effects of various boundaries. The former significantly increases the demands on computation resources and workload, while the latter requires the design of reasonable kinetic boundary conditions for different scenarios. The simulation results presented in this paper are not sufficient to answer the characteristics of the RTI systems’ late-time behavior, such as under what conditions the RTI will enter the turbulent stage. Furthermore, they cannot address how material properties, environmental, and other factors influence the mixing of matter and energy in the turbulence stage if it occurs. The simulation result in this paper shows that with the increase in At𝐴𝑡Atitalic_A italic_t, the inhibitory effect of compressibility on the spike velocity diminishes. Under the condition of At=0.7𝐴𝑡0.7At=0.7italic_A italic_t = 0.7 and Ma=0.3𝑀𝑎0.3Ma=0.3italic_M italic_a = 0.3, at t=6𝑡6t=6italic_t = 6, the spike nearly reaches the lower boundary. When t>6𝑡6t>6italic_t > 6, the spike will hit the lower boundary, which will significantly impact the flow field. To avoid considering the influence of the boundary on the flow field, the results after t=6𝑡6t=6italic_t = 6 were not analyzed in our work. For convenient comparison in all cases, we have only taken the results before t=6𝑡6t=6italic_t = 6 for all the calculations in our work. Nevertheless, we can make some remarks for extended late-time calculations for different Atwood number conditions. In cases with At=0.1𝐴𝑡0.1At=0.1italic_A italic_t = 0.1, compressibility strongly inhibits bubble and spike motion due to the small density difference between the light and heavy fluids. After t=5𝑡5t=5italic_t = 5, the velocities of the bubbles and spikes remain near zero. The bubbles and spikes are confined near the initial interface due to the effect of compressibility. As time evolves, the systems will slowly reach equilibrium as heat transfers between the heavy and light fluids and vorticity dissipates. This is consistent with the evolutionary pattern observed at t=6𝑡6t=6italic_t = 6. Under this circumstance, no major changes are expected in the evolution during extended later-time calculations. However, for cases with At=0.25𝐴𝑡0.25At=0.25italic_A italic_t = 0.25 and 0.5, it is evident that the relative motion of the bubble and the spike remains strong at t=6𝑡6t=6italic_t = 6. Therefore, further studies are needed to determine whether the vortex’s deposition occurring after t=6𝑡6t=6italic_t = 6 and before the spike reaches the bottom boundary will cause the bubble and spike to accelerate again, and whether these reaccelerations will alter the evolution of the system’s non-equilibrium qualities.

In addition to the DBM modeling and analysis method for complex fluids with strong compressibility, the series of physical cognitions in this paper provides a more accurate understanding of the RTI kinetics and a helpful reference for the development of corresponding regulation techniques.

Acknowledgements

The authors thank Yanbiao Gan, Feng Chen, Chuandong Lin, Huilin Lai, Zhipeng Liu, Ge Zhang, Yiming Shan, Dejia Zhang, Jiahui Song, Hanwei Li, Yingqi Jia, and Xuan Zhang for helpful discussions on DBM. This work was supported by the National Natural Science Foundation of China (under Grant Nos. 12172061,12102397 and 12101064), the Opening Project of State Key Laboratory of Explosion Science and Safety Protection (Bei**g Institute of Technology) (Grant No. KFJJ23-02M), the Foundation of National Key Laboratory of Shock Wave and Detonation Physics (Grant No. JCKYS2023212003), and the 2023 Computational Physics Key Laboratory Youth Fund Sponsored Project (under Grant No. 6241A05QN23001).

Appendix

Figure 23 shows the schematic of discrete velocities used in this work, which contains a zero velocity and four sets of 8 velocities in each direction. Each set of discrete velocities has the same velocity value vjsubscript𝑣𝑗{{v}_{j}}italic_v start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, and discrete velocity 𝐯jisubscript𝐯𝑗𝑖{{\mathbf{v}}_{ji}}bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT can be written as:

𝐯ji=vj[cos(i14π),sin(i14π)],subscript𝐯𝑗𝑖subscript𝑣𝑗𝑖14𝜋𝑖14𝜋,{{\mathbf{v}}_{ji}}={{v}_{j}}\left[\cos(\frac{i-1}{4}\pi),\sin(\frac{i-1}{4}% \pi)\right]\text{,}bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT [ roman_cos ( divide start_ARG italic_i - 1 end_ARG start_ARG 4 end_ARG italic_π ) , roman_sin ( divide start_ARG italic_i - 1 end_ARG start_ARG 4 end_ARG italic_π ) ] , (33)

where j=0,1,2,3,4;i=1,2,,8formulae-sequence𝑗01234𝑖128j=0,1,2,3,4;i=1,2,\cdots,8italic_j = 0 , 1 , 2 , 3 , 4 ; italic_i = 1 , 2 , ⋯ , 8, j=0𝑗0j=0italic_j = 0 is the zero velocity. The equilibrium distribution function corresponding to the discrete velocity can be written as:

fjieq=ρFj[(1u22T+u48T2)+𝐯ji𝐮T(1u22T)+(𝐯ji𝐮)22T2(1u22T)+(𝐯ji𝐮)36T3+(𝐯ji𝐮)424T4],superscriptsubscript𝑓𝑗𝑖𝑒𝑞𝜌subscript𝐹𝑗delimited-[]1superscript𝑢22𝑇superscript𝑢48superscript𝑇2subscript𝐯𝑗𝑖𝐮𝑇1superscript𝑢22𝑇superscriptsubscript𝐯𝑗𝑖𝐮22superscript𝑇21superscript𝑢22𝑇superscriptsubscript𝐯𝑗𝑖𝐮36superscript𝑇3superscriptsubscript𝐯𝑗𝑖𝐮424superscript𝑇4,f_{ji}^{eq}=\rho{{F}_{j}}\left[\left(1-\frac{{{u}^{2}}}{2T}+\frac{{{u}^{4}}}{8% {{T}^{2}}}\right)+\frac{{{\mathbf{v}}_{ji}}\cdot\mathbf{u}}{T}\left(1-\frac{{{% u}^{2}}}{2T}\right)+\frac{{{({{\mathbf{v}}_{ji}}\cdot\mathbf{u})}^{2}}}{2{{T}^% {2}}}\left(1-\frac{{{u}^{2}}}{2T}\right)\right.\left.+\frac{{{({{\mathbf{v}}_{% ji}}\cdot\mathbf{u})}^{3}}}{6{{T}^{3}}}+\frac{{{({{\mathbf{v}}_{ji}}\cdot% \mathbf{u})}^{4}}}{24{{T}^{4}}}\right]\text{,}italic_f start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_e italic_q end_POSTSUPERSCRIPT = italic_ρ italic_F start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT [ ( 1 - divide start_ARG italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_T end_ARG + divide start_ARG italic_u start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG start_ARG 8 italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) + divide start_ARG bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT ⋅ bold_u end_ARG start_ARG italic_T end_ARG ( 1 - divide start_ARG italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_T end_ARG ) + divide start_ARG ( bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT ⋅ bold_u ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( 1 - divide start_ARG italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_T end_ARG ) + divide start_ARG ( bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT ⋅ bold_u ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG 6 italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG + divide start_ARG ( bold_v start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT ⋅ bold_u ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG start_ARG 24 italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG ] , (34)

where Fjsubscript𝐹𝑗{{F}_{j}}italic_F start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the weighting factor:

F1=48T46(v22+v32+v42)T3+(v22v32+v22v42+v32v42)T214v22v32v42Tv12(v12v22)(v12v32)(v12v42),subscript𝐹148superscript𝑇46superscriptsubscript𝑣22superscriptsubscript𝑣32superscriptsubscript𝑣42superscript𝑇3superscriptsubscript𝑣22superscriptsubscript𝑣32superscriptsubscript𝑣22superscriptsubscript𝑣42superscriptsubscript𝑣32superscriptsubscript𝑣42superscript𝑇214superscriptsubscript𝑣22superscriptsubscript𝑣32superscriptsubscript𝑣42𝑇superscriptsubscript𝑣12superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣12superscriptsubscript𝑣32superscriptsubscript𝑣12superscriptsubscript𝑣42,{{F}_{1}}=\frac{48{{T}^{4}}-6(v_{2}^{2}+v_{3}^{2}+v_{4}^{2}){{T}^{3}}+(v_{2}^{% 2}v_{3}^{2}+v_{2}^{2}v_{4}^{2}+v_{3}^{2}v_{4}^{2}){{T}^{2}}-\frac{1}{4}v_{2}^{% 2}v_{3}^{2}v_{4}^{2}T}{v_{1}^{2}(v_{1}^{2}-v_{2}^{2})(v_{1}^{2}-v_{3}^{2})(v_{% 1}^{2}-v_{4}^{2})}\text{,}italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG 48 italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 6 ( italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + ( italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T end_ARG start_ARG italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG , (35)
F2=48T46(v12+v32+v42)T3+(v12v32+v12v42+v32v42)T214v12v32v42Tv22(v22v12)(v22v32)(v22v42),subscript𝐹248superscript𝑇46superscriptsubscript𝑣12superscriptsubscript𝑣32superscriptsubscript𝑣42superscript𝑇3superscriptsubscript𝑣12superscriptsubscript𝑣32superscriptsubscript𝑣12superscriptsubscript𝑣42superscriptsubscript𝑣32superscriptsubscript𝑣42superscript𝑇214superscriptsubscript𝑣12superscriptsubscript𝑣32superscriptsubscript𝑣42𝑇superscriptsubscript𝑣22superscriptsubscript𝑣22superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣32superscriptsubscript𝑣22superscriptsubscript𝑣42,{{F}_{2}}=\frac{48{{T}^{4}}-6(v_{1}^{2}+v_{3}^{2}+v_{4}^{2}){{T}^{3}}+(v_{1}^{% 2}v_{3}^{2}+v_{1}^{2}v_{4}^{2}+v_{3}^{2}v_{4}^{2}){{T}^{2}}-\frac{1}{4}v_{1}^{% 2}v_{3}^{2}v_{4}^{2}T}{v_{2}^{2}(v_{2}^{2}-v_{1}^{2})(v_{2}^{2}-v_{3}^{2})(v_{% 2}^{2}-v_{4}^{2})}\text{,}italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG 48 italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 6 ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T end_ARG start_ARG italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG , (36)
F3=48T46(v12+v22+v42)T3+(v12v22+v12v42+v22v42)T214v12v22v42Tv32(v32v12)(v32v22)(v12v42),subscript𝐹348superscript𝑇46superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣42superscript𝑇3superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣12superscriptsubscript𝑣42superscriptsubscript𝑣22superscriptsubscript𝑣42superscript𝑇214superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣42𝑇superscriptsubscript𝑣32superscriptsubscript𝑣32superscriptsubscript𝑣12superscriptsubscript𝑣32superscriptsubscript𝑣22superscriptsubscript𝑣12superscriptsubscript𝑣42,{{F}_{3}}=\frac{48{{T}^{4}}-6(v_{1}^{2}+v_{2}^{2}+v_{4}^{2}){{T}^{3}}+(v_{1}^{% 2}v_{2}^{2}+v_{1}^{2}v_{4}^{2}+v_{2}^{2}v_{4}^{2}){{T}^{2}}-\frac{1}{4}v_{1}^{% 2}v_{2}^{2}v_{4}^{2}T}{v_{3}^{2}(v_{3}^{2}-v_{1}^{2})(v_{3}^{2}-v_{2}^{2})(v_{% 1}^{2}-v_{4}^{2})}\text{,}italic_F start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = divide start_ARG 48 italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 6 ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T end_ARG start_ARG italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG , (37)
F4=48T46(v12+v22+v32)T3+(v12v22+v12v32+v22v32)T214v12v22v32Tv42(v42v12)(v42v22)(v42v32),subscript𝐹448superscript𝑇46superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣32superscript𝑇3superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣12superscriptsubscript𝑣32superscriptsubscript𝑣22superscriptsubscript𝑣32superscript𝑇214superscriptsubscript𝑣12superscriptsubscript𝑣22superscriptsubscript𝑣32𝑇superscriptsubscript𝑣42superscriptsubscript𝑣42superscriptsubscript𝑣12superscriptsubscript𝑣42superscriptsubscript𝑣22superscriptsubscript𝑣42superscriptsubscript𝑣32,{{F}_{4}}=\frac{48{{T}^{4}}-6(v_{1}^{2}+v_{2}^{2}+v_{3}^{2}){{T}^{3}}+(v_{1}^{% 2}v_{2}^{2}+v_{1}^{2}v_{3}^{2}+v_{2}^{2}v_{3}^{2}){{T}^{2}}-\frac{1}{4}v_{1}^{% 2}v_{2}^{2}v_{3}^{2}T}{v_{4}^{2}(v_{4}^{2}-v_{1}^{2})(v_{4}^{2}-v_{2}^{2})(v_{% 4}^{2}-v_{3}^{2})}\text{,}italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = divide start_ARG 48 italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT - 6 ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_T end_ARG start_ARG italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_v start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG , (38)
F0=18(F1+F2+F3+F4).subscript𝐹018subscript𝐹1subscript𝐹2subscript𝐹3subscript𝐹4.{{F}_{0}}=1-8({{F}_{1}}+{{F}_{2}}+{{F}_{3}}+{{F}_{4}})\text{.}italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 - 8 ( italic_F start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_F start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + italic_F start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) . (39)
Refer to caption
Figure 23: Schematic of discrete velocities.

References

References

  • Rayleigh (1882) L. Rayleigh, Proceedings of the London mathematical society 1, 170 (1882).
  • Taylor (1950) G. Taylor, Proceedings of the Royal Society of London.Series A.Mathematical and Physical Sciences 201, 192 (1950).
  • Zhou (2024) Y. Zhou, in Hydrodynamic Instabilities and Turbulence: Rayleigh-Taylor, Richtmyer-Meshkov, and Kelvin-Helmholtz Mixing (Cambridge University Press, Cambridge, 2024).
  • Zhou (2017a) Y. Zhou, Physics Reports 720-722, 1 (2017a).
  • Zhou (2017b) Y. Zhou, Physics Reports 723, 1 (2017b).
  • Zhou et al. (2021) Y. Zhou, R. J. R. Williams, P. Ramaprabhu, M. Groom, B. Thornber, A. Hillier, W. Mostert, B. Rollin, S. Balachandar, P. D. Powell, et al., Physica D 423, 132838 (2021).
  • Bernstein and Book (1983) I. B. Bernstein and D. L. Book, Physics of Fluids 26, 453 (1983).
  • Yang and Zhang (1993) Y. M. Yang and Q. Zhang, Physics of Fluids A: Fluid Dynamics 5, 1167 (1993).
  • Blake (1972) G. M. Blake, Monthly Notices of the Royal Astronomical Society 156, 67 (1972).
  • Baker (1983) L. Baker, The Physics of Fluids 26, 950 (1983).
  • Livescu (2004) D. Livescu, Physics of fluids 16, 118 (2004).
  • Xue and Ye (2010) C. Xue and W. H. Ye, Physics of Plasmas 17 (2010).
  • Lafay et al. (2007) M. A. Lafay, B. L. Creurer, and S. Gauthier, Europhysics Letters 79, 64002 (2007).
  • Reckinger et al. (2016) S. J. Reckinger, D. Livescu, and O. V. Vasilyev, Journal of Computational Physics 313, 181 (2016).
  • Gauthier (2017) S. Gauthier, Journal of Fluid Mechanics 830, 211 (2017).
  • Luo et al. (2020) T. F. Luo, J. C. Wang, C. Y. Xie, M. P. Wan, and S. Y. Chen, Physics of fluids 32 (2020).
  • Fu et al. (2022) C. Q. Fu, Z. Y. Zhao, X. Xu, P. Wang, N. S. Liu, Z. H. Wang, and X. Y. Lu, Physical Review Fluids 7, 023902 (2022).
  • Qin et al. (2001) C. S. Qin, F. G. Zhang, and Y. Li, EXPLOSION AND SHOCK WAVES 21, 5 (2001).
  • Qin and Wang (2004) C. S. Qin and P. Wang, EXPLOSION AND SHOCK WAVES 24, 1 (2004).
  • Olson and Cook (2007) B. J. Olson and A. W. Cook, Physics of Fluids 19, 128108 (2007).
  • Reckinger et al. (2012) S. J. Reckinger, D. Livescu, and O. V. Vasilyev, in Seventh International Conference on Computational Fluid Dynamics (ICCFD7), Big Island, Hawaii (2012).
  • Wieland et al. (2019) S. A. Wieland, P. E. Hamlington, S. J. Reckinger, and D. Livescu, Physical review fluids 4, 093905 (2019).
  • Fu et al. (2023) C. G. Fu, Z. Y. Zhao, P. Wang, N. S. Liu, Z. H. Wan, and X. Y. Lu, Journal of Fluid Mechanics 954, A16 (2023).
  • Chen et al. (2020) F. Chen, A. G. Xu, Y. D. Zhang, and Q. K. Zeng, Physics of Fluids 32, 104111 (2020).
  • Xu and Zhang (2022) A. G. Xu and Y. D. Zhang, in Complex Media Kinetics (Science Press, Bei**g, 2022).
  • Xu et al. (2024) A. G. Xu, J. D. Zhang, and Y. B. Gan, Frontiers of Physics 19, 42500 (2024).
  • Succi (2001) S. Succi, in The lattice Boltzmann equation: for fluid dynamics and beyond (Oxford university press, Oxford, 2001).
  • Osborn et al. (1995) W. Osborn, E. Orlandini, M. R. Swift, J. Yeomans, and J. R. Banavar, Physical review letters 75, 4031 (1995).
  • Swift et al. (1995) M. R. Swift, W. Osborn, and J. Yeomans, Physical review letters 75, 830 (1995).
  • Liang et al. (2014) H. Liang, B. C. Shi, Z. L. Guo, and Z. H. Chai, Physical review E 89, 053320 (2014).
  • Liang et al. (2016) H. Liang, Q. X. Li, B. C. Shi, and Z. H. Chai, Physical review E 93, 033113 (2016).
  • Liang et al. (2021) H. Liang, Z. H. Xia, and H. W. Huang, Physics of Fluids 33, 082103 (2021).
  • Chen et al. (2018) F. Chen, A. G. Xu, and G. C. Zhang, Physics of Fluids 30, 102105 (2018).
  • Gan et al. (2019) Y. B. Gan, A. G. Xu, G. C. Zhang, C. D. Lin, H. L. Lai, and Z. P. Liu, Frontiers of Physics 14, 1 (2019).
  • Lai et al. (2016) H. L. Lai, A. G. Xu, G. C. Zhang, Y. B. Gan, Y. J. Ying, and S. Succi, Physical Review E 94, 023106 (2016).
  • Lin et al. (2017) C. D. Lin, A. G. Xu, G. C. Zhang, K. H. Luo, and Y. J. Li, Physical Review E 96, 053305 (2017).
  • Zhang et al. (2021) G. Zhang, A. G. Xu, D. J. Zhang, Y. J. Li, H. L. Lai, and X. M. Hu, Physics of Fluids 33, 076105 (2021).
  • Li et al. (2022) H. W. Li, A. G. Xu, G. Zhang, and Y. M. Shan, Communications in Theoretical Physics 74, 115601 (2022).
  • Miles (1966) J. W. Miles, Tech. Rep., General Dynamics San Diego Ca General Atomic Div (1966).
  • Piriz et al. (2009) A. R. Piriz, J. J. L. Cela, and N. A. Tahir, Physical Review E 80, 046305 (2009).
  • Li et al. (2021) B. Y. Li, J. X. Peng, Y. Gu, and L. H. He, Acta Physica Sinica 70, 114701 (2021).
  • Xu et al. (2015) A. G. Xu, G. C. Zhang, and Y. J. Ying, Acta Physica Sinica 64 (2015).
  • Xu et al. (2018) A. G. Xu, G. C. Zhang, and Y. D. Zhang, in Kinetic Theory, edited by G. Kyzas and A. Mitropoulos (InTech, Rijeka, 2018), chap. 02.
  • Xu et al. (2021a) A. G. Xu, J. Chen, J. H. Song, D. W. Chen, and Z. H. Chen, Acta Aerodyn.Sin 39, 138 (2021a).
  • Xu et al. (2021b) A. G. Xu, Y. M. Shan, F. Chen, Y. B. Gan, and C. D. Lin, Acta Aeronauticaet Astronautica Sinica 42, 46 (2021b).
  • Xu et al. (2021c) A. G. Xu, J. H. Song, F. Chen, K. Xie, and Y. J. Ying, Chinese Journal of Computational Physics 38, 631 (2021c).
  • Gan et al. (2022a) Y. B. Gan, A. G. Xu, H. L. Lai, W. Li, G. L. Sun, and S. Succi, Journal of Fluid Mechanics 951, A8 (2022a).
  • Zhang et al. (2023) D. J. Zhang, A. G. Xu, Y. B. Gan, Y. D. Zhang, J. H. Song, and Y. J. Li, Physics of Fluids 35, 106113 (2023).
  • Wu and Liu (2011) X. Wu and Z. H. Liu, Complex systems and complexity scienc 8, 0039 (2011).
  • Wang et al. (2010) L. Wang, D. H. He, and B. B. Hu, Phys. Rev. Lett 105, 160601 (2010).
  • Gonnella et al. (2007) G. Gonnella, A. Lamura, and V. Sofonea, Physical Review E 76, 036703 (2007).
  • Gan et al. (2022b) Y. B. Gan, A. G. Xu, H. L. Lai, W. Li, G. L. Sun, and S. Succi, J. Fluid Mech. 951, A8 (2022b).
  • Chen et al. (2022) J. Chen, A. G. Xu, D. W. Chen, Y. D. Zhang, and Z. H. Chena, Physical Review E 106, 015102 (2022).
  • Song et al. (2024) J. H. Song, A. G. Xu, L. Miao, F. Chen, Z. P. Liu, L. F. Wang, and X. Hou, Physics of Fluids 36, 016107 (2024).
  • Cai et al. (2021) H. B. Cai, X. X. Yan, P. L. Yao, and S. P. Zhu, Matter and Radiation at Extremes 6, 9 (2021).
  • Shan et al. (2018) L. Q. Shan, H. B. Cai, W. S. Zhang, Q. Tang, F. Zhang, Z. F. Song, B. Bi, F. J. Ge, J. B. Chen, D. X. Liu, et al., Physical review letters 120, 195001 (2018).
  • Qiu et al. (2024) R. F. Qiu, X. Y. Yan, Y. Bao, Y. C. You, and H. J, Entropy 26, 200 (2024).
  • Onuki (2005) A. Onuki, Physical review letters 94, 054501 (2005).
  • Bian et al. (2020) X. Bian, H. Aluie, D. X. Zhao, H. S. Zhang, and D. Livescu, Physica D: Nonlinear Phenomena 403, 132250 (2020).
  • Zhang et al. (2019) Y. D. Zhang, A. G. Xu, G. C. Zhang, Y. B. Gan, Z. H. Chen, and S. Succi, Soft matter 15, 2245 (2019).
  • Tiribocchi et al. (2009) A. Tiribocchi, N. Stella, G. Gonnella, and A. Lamura, Physical Review E 80, 026701 (2009).
  • Wang et al. (2013) J. C. Wang, Y. T. Yang, Y. P. Shi, Z. L. Xiao, X. T. He, and Y. J. Li, Physical review letters 110, 214505 (2013).