This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Central-moment-based discrete Boltzmann modeling of compressible flows

\nameChuandong Lina,b,c, Xianli Sua,d, Linlin Feie and Kai H. Luob,d CONTACT Chuandong Lin Email: linchd3@mail.sysu.edu.cn; Xianli Su Email: suxli@mail2.sysu.edu.cn; Linlin Fei Email: linfei@ethz.ch; Kai Hong Luo Email: k.luo@ucl.ac.uk aSino-French Institute of Nuclear Engineering and Technology, Sun Yat-sen University, Zhuhai, 519082, China;
bCenter for Combustion Energy, Key Laboratory for Thermal Science and Power Engineering of Ministry of Education, Department of Energy and Power Engineering, Tsinghua University, Beijing, 100084, China;
cNational University of Singapore, Department of Mechanical Engineering, National University of Singapore, 119077, Singapore;
dDepartment of Mechanical Engineering, University College London, London, WC1E 7JE, United Kingdom;
eChair of Building Physics, Department of Mechanical and Process Engineering, ETH Zürich, 8092, Switzerland
Abstract

In this work, a central-moment-based discrete Boltzmann method (CDBM) is constructed for fluid flows with variable specific heat ratios. The central kinetic moments are employed to calculate the equilibrium discrete velocity distribution function in the CDBM. In comparison to previous incompressible central-moment-based lattice Boltzmann method, the CDBM possesses the capability of investigating compressible flows with thermodynamic nonequilibrium effects beyond conventional hydrodynamic models. Unlike all existing DBMs which are constructed in raw-moment space, the CDBM stands out by directly providing the nonequilibrium effects related to the thermal fluctuation. The proposed method has been rigorously validated using benchmarks of the Sod shock tube, Lax shock tube, shock wave phenomena, two-dimensional sound wave, and the Taylor-Green vortex flow. The numerical results exhibit an exceptional agreement with theoretical predictions.

keywords:
Compressible flows; nonequilibrium effects; central-moment-based discrete Boltzmann method

1 Introduction

Compressible flows, characterized by both hydrodynamic and thermodynamic nonequilibrium effects, are ubiquitous in natural and engineering contexts, including applications such as inertial confinement fusion, gas pipelines, jet engines, and rocket motors [1]. Notably, modern high-speed airplanes and the jet engines that power them are wonderful examples of the application of compressible flows. Additionally, during the reentry of an aircraft into the atmosphere at high Mach numbers, the surrounding shock wave induces significant acceleration and heating of the air. Therefore, in the modern world of aerospace and mechanical engineering, a deep understanding of the principles of compressible flows is essential [2].

To predict the thermal process and nonequilibrium effects, there are three main categories of numerical methodologies. The first class involves modifying traditional macroscopic models, for example, the Navier-Stokes (NS) equations plus slip boundary conditions [3, 4, 5], the NS-type equations with multi-components and/or multi-temperatures [6, 7], Burnett and super-Burnett equations [8, 9], etc. Since the macroscopic models are constructed under the continuum assumption, the application scope is limited at small Knudsen numbers. The second type encompasses microscopic models, such as the molecular dynamics (MD) [10, 11]. The MD provides details of physical systems, but is only applicable to small temporal and spatial domains due to an excessive computational burden.

Moreover, to address the aforementioned challenges, the third category of methods encompasses mesoscopic approaches based on kinetic theory, such as the direct simulation Monte Carlo (DSMC) method [12], discrete velocity methods (DVM) [13, 14], (discrete) unified gas-kinetic schemes ((D)UGKS) [15, 16, 17, 18], the lattice Boltzmann method (LBM) [19, 20], and the discrete Boltzmann method (DBM) [21, 22], among others. These approaches act as a bridge connecting microscopic and macroscopic scales. The DSMC method, first introduced by G. A. Bird, has been extensively developed and widely applied in rarefied gas dynamics, particularly for high-speed flows [12]. However, DSMC is not well-suited to the simulation of low-speed flows, and has to balance the noise level and computational efficiency [23]. DVM represents another widely used approach, where the continuous particle velocity space is discretized into a finite set of velocity coordinate points, and numerical quadrature is employed to approximate the integration of moments [13, 14]. However, for high-speed compressible flows, especially in the near continuous flow region, the DVM exhibits limited computational efficiency [14]. Besides, in order to solve the discrete velocity Boltzmann equation, Xu et al. [15, 16] and Guo et al. [17, 18] proposed the unified gas-kinetic scheme (UGKS) and the discrete unified gas-kinetic scheme (DUGKS), respectively. In UGKS, the local integration of the discrete velocity Boltzmann equation is used to compute the flux of the distribution function, whereas in DUGKS, the flux is derived directly from the evolution equation.

Among the kinetic methodologies, the LBM stands out [19, 20]. The LBM has emerged as a competitive scheme for simulating various complex flows due to its canonical “collision-streaming” algorithm which disentangles non-linearity and non-locality and is easy to implement. Owing to these inherent advantages, LBM has been successfully applied to simulate various physical problems including multiphase [24], reactive [25], magnetohydrodynamic [26], nano- [27], biomechanics [28], and porous media flow [29]. Although the LBM has achieved significant success in simulating nearly incompressible complex flows, its application to compressible flows continues to present significant challenges.

Alexander et al. devised the first multi-speed lattice Boltzmann model containing 13 discrete velocities, representing the earliest application of the LBM to the compressible NS equations [30]. Shortly after the introduction of Alexander’s model, in 1993, Qian proposed a series of multi-speed models based on the DnQb thermal lattice Bhatnager–Gross–Krook (BGK) isothermal model for thermohydrodynamics, in which a proper internal energy is introduced and the energy equation is obtained [31]. In 1994, Chen et al. introduced a thermal lattice BGK model capable of recovering the standard compressible NS equations through a higher-order velocity expansion of the Maxwellian-type equilibrium distribution [32]. However, this model was constrained by assumptions of zero bulk viscosity and unit Prandtl number [32]. In 1997, Chen et al. constructed a two time relaxation parameters thermal lattice BGK model to achieve adjustable Prandtl number [33]. In the same year, McNamara et al. derived the distribution functions of thermal LBM using a series of moment conservation equations, enabling variability in the Prandtl number within this model [34]. Subsequently, compressible lattice Boltzmann models with a flexible specific heat ratio were introduced by Hu et al. in 1997 [35] and Yan et al. in 1999 [36]. With the aim of simulating compressible flows, Sun et al. developed an adaptive compressible LBM based on a simple delta function, where the lattice velocities vary with mean flow velocity and internal energy [37, 38, 39, 40, 41]. In 2007, Li et al. proposed a coupled double-distribution-function LBM for the NS equations with a flexible specific heat ratio and Prandtl number [42].

In recent years, Yang et al. developed a platform for constructing one-dimensional compressible LBM and subsequently extended the finite volume LBM to simulate compressible flows, including shock waves [43, 44]. Li et al. introduced a novel LBM model designed for fully compressible flows. Building upon a multi-speed model, an extra potential energy distribution function is introduced to recover the full compressible NS equations, featuring a flexible specific heat ratio and Prandtl number [45]. In 2018, Dorschner et al. constructed a particles-on-demand (PonD) method which is suitable for both incompressible and compressible flows [46]. In the PonD method, the discrete particle velocities are defined relative to a local reference frame, determined by the local flow velocity and temperature, which vary in both space and time. Subsequently, Kallikounis et al. further refined PonD method with a revised reference frame transformation using Grad’s projection to enhance stability and accuracy [47]. In 2024, Kallikounis et al. improved the PonD method for variable Prandtl number by employing the quasi-equilibrium relaxation, and the model was validated via simulations of high Mach compressible flows [48]. In the same year, Ji et al proposed an Eulerian realization of the PonD for hypersonic compressible flows, where a kinetic model formulated in the comoving reference frame [49].

Generally, lattice Boltzmann methods (LBMs) are broadly classified into the single-relaxation-time (SRT) model and the multiple-relaxation-time (MRT) model. The SRT or BGK operator is the simplest collision operator in the standard LBM, where all distribution functions relax to their local equilibrium values at a constant rate. However, the BGK-LBM may encounter stability issues in the zero-viscosity limit and/or for non-vanishing Mach numbers [20, 50, 51]. To mitigate this issue, numerous strategies have been proposed, including modifications to the numerical discretization, collision model, or both [50, 51]. In 2006, Geier et al. introduced a cascaded operator, conducting collisions in the central-moment space rather than that of raw moments as in the MRT-LBM[52]. Consequently, the corresponding method gradually interpreted as “central-moment-based” lattice Boltzmann method (CLBM) [53, 54, 55, 56, 57, 58]. However, the CLBM does not necessarily provide enhanced stability. For example, when relaxation in the central moment space is implemented using a uniform relaxation parameter for all moments, the approach reduces to a BGK collision operator with an extended equilibrium [59, 60]. To improve the stability of CLBMs, the relaxation frequencies of higher-order moments, including the trace of second-order moments, should be set to one or a value close to one [61, 62].

In this work, we constructed a central-moment-based discrete Boltzmann method (CDBM) tailored specifically for the simulation of compressible flows. The “central-moment-based” indicates that physical quantities and higher-order kinetic moments are computed through central moments. Unlike the majority of previously proposed CLBMs, which are constrained to incompressible flows and neglect the thermodynamic nonequilibrium effects inherent to the Boltzmann equation, the CDBM possesses significantly broader applicability. Based upon the Boltzmann equation, in addition to capturing the general hydrodynamic behaviors described by the NS model, the CDBM provides deeper insights into more detailed thermodynamic nonequilibrium behaviors in various complex fluids. Moreover, distinct from all existing DBMs constructed in raw-moment space [21, 22, 63, 64, 65, 66, 67, 68, 69, 70, 71, 72], the CDBM, using the peculiar velocity, can quantify the nonequilibrium effects related to the thermal fluctuation directly. These distinctive features make the CDBM particularly valuable for investigating compressible flows that exhibit significant thermodynamic nonequilibrium effects.

2 Research methodology

The governing equations for the CDBM are expressed as follows:

fit+𝐯ifirα=1τ(fifieq),\frac{\partial{{f}_{i}}}{\partial t}+{{\mathbf{v}}_{i}}\cdot\frac{\partial{{f}_{i}}}{\partial{{r}_{\alpha}}}=-\frac{1}{\tau}\left({{{{f}}}_{i}}-{f}_{i}^{eq}\right)\tt{,} (1)

where fif_{i} indicate the discrete distribution functions, with the subscript ii denoting the velocity index. The discrete velocities are given by:

𝐯i={va(±1,0),1i4,vb(±1,±1),5i8,vc(±1,0),9i12,vd(±1,±1),13i16,{{\mathbf{v}}_{i}}=\left\{\begin{array}[]{*{35}{l}}\mathit{{v}_{a}}\left(\pm 1,0\right),&1\leq i\leq 4,\\ \mathit{}{{v}_{b}}\left(\pm 1,\pm 1\right),&5\leq i\leq 8,\\ \mathit{{v}_{c}}\left(\pm 1,0\right),&9\leq i\leq 12,\\ \mathit{}{{v}_{d}}\left(\pm 1,\pm 1\right),&13\leq i\leq 16,\end{array}\right. (2)

where vav_{a}, vbv_{b}, vcv_{c} and vdv_{d} are adjustable. Besides, in order to describe the vibrational and/or rotational energies, II is introduced for extra degrees of freedom due to vibration and/or rotation, and the corresponding parameters for degrees of freedom, ηi{\eta}_{i}, are defined as follows,

ηi={ηa,1i4,ηb,5i8,ηc,9i12,ηd,13i16.{{\eta}_{i}}=\left\{\begin{array}[]{*{35}{l}}{{\eta}_{a}},&1\leq i\leq 4,\\ {{\eta}_{b}},&5\leq i\leq 8,\\ {{\eta}_{c}},&9\leq i\leq 12,\\ {{\eta}_{d}},&13\leq i\leq 16.\\ \end{array}\right. (3)

Figure 1 exhibits the schematic of discrete velocities.

Refer to caption
Figure 1: The schematic of discrete velocities.

It is worth noting that the values of these parameters are chosen based on the specific application to ensure numerical stability. In general, the magnitudes of vav_{a}, vbv_{b}, vcv_{c} and vdv_{d} should approximate key physical quantities such as flow velocity 𝐮\mathbf{u}, sound speed vs=γTv_{s}=\sqrt{\gamma T}, and shock speed, etc. Similarly, the magnitudes of ηa{\eta}_{a}, ηb{\eta}_{b}, ηc{\eta}_{c} and ηd{\eta}_{d} should be around the value of IT\sqrt{IT}. This is because, in thermodynamic equilibrium, mη¯2/2=IT/2m\bar{\eta}^{2}/2=IT/2 where m = 1 is the particle mass, and η¯2\bar{\eta}^{2} represents the average value of η2\eta^{2}, according to the equipartition of energy theorem. Hence, η¯=IT\bar{\eta}=\sqrt{IT}, and the values of ηa{\eta}_{a}, ηb{\eta}_{b}, ηc{\eta}_{c} and ηd{\eta}_{d} should be around η¯\bar{\eta}.

The relaxation time τ\tau is a constant which governs the relaxation speed of fi{{{f}}_{i}} towards fieq{f}_{i}^{eq}. fi{{{f}}_{i}} and fieq{f}_{i}^{eq} represent the discrete velocity distributions and the corresponding equilibrium counterparts, respectively. The equilibrium discrete distribution functions, fieq{f}_{i}^{eq}, are uniformly computed via a matrix inversion method [73]. In contrast to prior DBM constructions, this study employs central kinetic moments.

To recover the NS equations, through the Chapman-Enskog expansion, the discrete equilibrium distribution functions should satisfy the following seven central kinetic moments,

feqΨ𝑑𝐯𝑑η=ifieqΨi,\int{\int{{f^{eq}}\Psi d{\bf{v}}d\eta}}=\sum\nolimits_{i}{f_{i}^{eq}{\Psi_{i}}}\tt{,} (4)

where Ψ=1\Psi=1, 𝐯{{\mathbf{v}}^{*}}, 𝐯𝐯+η2{{\mathbf{v}}^{*}}\cdot{{\mathbf{v}}^{*}}+{{\eta}^{2}}, 𝐯𝐯{{\mathbf{v}}^{*}}\cdot{{\mathbf{v}}^{*}}, (𝐯𝐯+η2)𝐯\left({{\mathbf{v}}^{*}}\cdot{{\mathbf{v}}^{*}}+{{\eta}^{2}}\right){{\mathbf{v}}^{*}}, 𝐯𝐯𝐯{{\mathbf{v}}^{*}}{{\mathbf{v}}^{*}}{{\mathbf{v}}^{*}}, (𝐯𝐯+η2)𝐯𝐯\left({{\mathbf{v}}^{*}}\cdot{{\mathbf{v}}^{*}}+{{\eta}^{2}}\right){{\mathbf{v}}^{*}}{{\mathbf{v}}^{*}}, correspondingly, Ψi=1{{\Psi}_{i}}=1, 𝐯i\mathbf{v}_{i}^{*}, 𝐯i𝐯i+ηi2\mathbf{v}_{i}^{*}\cdot\mathbf{v}_{i}^{*}+\eta_{i}^{2}, 𝐯i𝐯i\mathbf{v}_{i}^{*}\cdot\mathbf{v}_{i}^{*}, (𝐯i𝐯i+ηi2)𝐯i\left(\mathbf{v}_{i}^{*}\cdot\mathbf{v}_{i}^{*}+\eta_{i}^{2}\right)\mathbf{v}_{i}^{*}, 𝐯i𝐯i𝐯i\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}, (𝐯i𝐯i+ηi2)𝐯i𝐯i\left(\mathbf{v}_{i}^{*}\cdot\mathbf{v}_{i}^{*}+\eta_{i}^{2}\right)\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}. Here 𝐯=𝐯𝐮{{\mathbf{v}}^{*}}=\mathbf{v}-\mathbf{u} and 𝐯i=𝐯i𝐮\mathbf{v}_{i}^{*}={{\mathbf{v}}_{i}}-\mathbf{u}, with 𝐮\mathbf{u} the flow velocity. In addition, the equilibrium distribution function feqf^{eq} is

feq=n(m2πT)D/2(m2πIT)1/2exp[m|𝐯𝐮|22Tmη22IT],{{f}^{eq}}=n{{\left(\frac{m}{2\pi T}\right)}^{D/2}}{{\left(\frac{m}{2\pi IT}\right)}^{1/2}}\exp\left[-\frac{m{{\left|\mathbf{v}-\mathbf{u}\right|}^{2}}}{2T}-\frac{m{{\eta}^{2}}}{2IT}\right]\tt{,} (5)

where DD denotes the dimensional translational degree of freedom, II stands for extra degrees of freedom due to vibration and/or rotation, and η\eta is used to describe the corresponding vibrational and/or rotational energies. The other parameters include nn as the particle number density, TT as the temperature, mm = 1 as the particle mass, and ρ=nm\rho=nm as the mass density. Besides, the specific heat ratio is γ=(D+I+2)/(D+I)\gamma=\left(D+I+2\right)/\left(D+I\right), the dynamic viscosity μ=ρTτ\mu={\rho T}{\tau}, kinematic viscosity ν=μ/ρ=Tτ\nu={\mu}/{\rho}={T}{\tau}, and bulk viscosity μB=μ(2/D2/(D+I)){{\mu}_{B}}=\mu\left({2}/{D}-{2}/({D+I})\right).

The central kinetic moments can be expressed in a unified form

𝐟¯𝐞𝐪=𝐌𝐟𝐞𝐪.{{\bf{\bar{f}}}^{{\bf{eq}}}}={\bf{M}}{{\bf{f}}^{{\bf{eq}}}}\tt{.} (6)

Here, 𝐟¯𝐞𝐪=(f¯1eqf¯2eqf¯16eq)T{{\bf{\bar{f}}}^{{\bf{eq}}}}={\left({\begin{matrix}{\bar{f}_{1}^{eq}}\>{\bar{f}_{2}^{eq}}\>\cdots\>{\bar{f}_{16}^{eq}}\end{matrix}}\right)^{\rm{T}}} and 𝐟eq=(f1eqf2eqf16eq)T{{\bf{f}}^{eq}}={\left({\begin{matrix}{f_{1}^{eq}}\>{f_{2}^{eq}}\>\cdots\>{f_{16}^{eq}}\end{matrix}}\right)^{\rm{T}}} represent the equilibrium velocity distribution functions in the central moment space and velocity space, respectively. The matrix 𝐌=(𝐌1𝐌2𝐌16)T{\bf{M}}={{\left({{\mathbf{M}}_{1}}\;{{\mathbf{M}}_{2}}\;\cdots\;{{\mathbf{M}}_{16}}\right)}^{\mathrm{T}}} acts as the bridge for transforming the velocity distribution function between the moment space and the discrete velocity space, which contains the blocks, 𝐌i=(Mi1Mi2Mi16){{\mathbf{M}}_{i}}=\left(\begin{matrix}{{M}_{i1}}\>{{M}_{i2}}\>\cdots\>{{M}_{i16}}\end{matrix}\right), with elements M1i=1{M}_{1i}=1, M2i=vix{M}_{2i}=v_{ix}^{*}, M3i=viy{M}_{3i}=v_{iy}^{*}, M4i=vi2+ηi2{M}_{4i}=v_{i}^{{*}2}+\eta_{i}^{2}, M5i=vix2{M}_{5i}=v_{ix}^{{*}2}, M6i=vixviy{M}_{6i}=v_{ix}^{*}v_{iy}^{*}, M7i={M}_{7i}= viy2v_{iy}^{{*}2}, M8i=(vi2+ηi2)vix{M}_{8i}=(v_{i}^{{*}2}+\eta_{i}^{2})v_{ix}^{*}, M9i=(vi2+ηi2)viy{M}_{9i}=(v_{i}^{{*}2}+\eta_{i}^{2})v_{iy}^{*}, M10i=vix3{M}_{10i}=v_{ix}^{{*}3}, M11i=vix2viy{M}_{11i}=v_{ix}^{{*}2}v_{iy}^{*}, M12i=vixviy2{M}_{12i}=v_{ix}^{*}v_{iy}^{{*}2}, M13i=viy3{M}_{13i}=v_{iy}^{{*}3}, M14i=(vi2+ηi2)vix2{M}_{14i}=(v_{i}^{{*}2}+\eta_{i}^{2})v_{ix}^{{*}2}, M15i=(vi2+ηi2)vixviy{M}_{15i}=(v_{i}^{{*}2}+\eta_{i}^{2})v_{ix}^{*}v_{iy}^{*}, M16i=(vi2+ηi2)viy2{M}_{16i}=(v_{i}^{{*}2}+\eta_{i}^{2})v_{iy}^{{*}2}. Consequently, Eq. 6 can be expressed as

𝐟𝐞𝐪=𝐌𝟏𝐟¯𝐞𝐪.{\bf{f}}^{{\bf{eq}}}={\bf{M}}^{{\bf{-1}}}{\bf{\bar{f}}}^{{\bf{eq}}}\tt{.} (7)

It should be mentioned that the mass, momentum and energy conservation are described by the first three kinetic moments in Eq. (4), where fieqf_{i}^{eq} can be replaced by fif_{i}. In other words, the density, flow velocity and temperature are obtained from the kinetic moments of fif_{i}. Replacing fieqf_{i}^{eq} with fif_{i} results in the imbalance in the last four kinetic moments. The difference between the results calculated by fieqf_{i}^{eq} and fif_{i} can be used to describe the deviation of the system from equilibrium state. Consequently, the CDBM contains the following nonequilibrium manifestations:

𝚫2=i(fifieq)𝐯i𝐯i,{\mathbf{\Delta}}_{2}^{*}=\sum_{i}\Big{(}f_{i}-f_{i}^{eq}\Big{)}\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}\tt{,} (8)
𝚫3,1=i(fifieq)(𝐯i𝐯i+ηi2)𝐯i,{\mathbf{\Delta}}_{3,1}^{*}=\sum_{i}\Big{(}f_{i}-f_{i}^{eq}\Big{)}\Big{(}\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}+\eta_{i}^{2}\Big{)}\mathbf{v}_{i}^{*}\tt{,} (9)
𝚫3=i(fifieq)𝐯i𝐯i𝐯i,{\mathbf{\Delta}}_{3}^{*}=\sum\nolimits_{i}{\left({{f}_{i}}-f_{i}^{eq}\right)}\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}\tt{,} (10)
𝚫4,2=i(fifieq)(𝐯i𝐯i+ηi2)𝐯i𝐯i.{\mathbf{\Delta}}_{4,2}^{*}=\sum\nolimits_{i}{\left({{f}_{i}}-f_{i}^{eq}\right)\left(\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}+\eta_{i}^{2}\right)}\mathbf{v}_{i}^{*}\mathbf{v}_{i}^{*}\tt{.} (11)

The second order tensor 𝚫2\mathbf{\Delta}_{2}^{*} corresponds to the viscous stress tensor and twice the nonorganized energy. The vector 𝚫3,1\mathbf{\Delta}_{3,1}^{*} is associated with the heat flux and twice the nonorganized energy flux. 𝚫3\mathbf{\Delta}_{3}^{*} and 𝚫4,2\mathbf{\Delta}_{4,2}^{*} are higher-order nonequilibrium quantities beyond traditional NS models [74]. It is crucial to note that, in comparison with previous DBMs constructed in raw-moment space, the CDBM can provide the nonequilibrium effects related to the thermal fluctuation directly.

In addition, in the basic lattice Boltzmann model, spatial and temporal discretizations are coupled, which restricts the choice of discrete velocities and also affects the construction of the discrete equilibrium distribution function. By contrast, the discrete Boltzmann method retains the use of discrete velocities but eliminates dependence on specific discretization schemes. Instead, it directly solves the continuous Boltzmann equation, allowing the CDBM to adopt spatial and temporal discretization schemes flexibly. In the subsequent simulations, the temporal derivatives are computed using the second-order Runge-Kutta scheme [75], while the spatial discretization employs the second-order non-oscillatory, non-free-parameter dissipation finite difference (NND) scheme [76]. The details are as follows,

virfir=1Δr[H(ir+12)H(ir12)],{{v}_{ir}}\frac{\partial{{f}_{i}}}{\partial r}=-\frac{1}{\Delta r}[H(ir+\frac{1}{2})-H(ir-\frac{1}{2})]\tt{,} (12)
H(ir+12)=HL(ir+12)+HR(ir+12),H(ir+\frac{1}{2})={{H}_{L}}(ir+\frac{1}{2})+{{H}_{R}}(ir+\frac{1}{2})\tt{,} (13)
HL(ir+12)=fi+(ir)+12minmod[Δfi+(ir+12),Δfi+(ir12)],{{H}_{L}}(ir+\frac{1}{2})=f_{i}^{+}(ir)+\frac{1}{2}{\rm{minmod}}[\Delta f_{i}^{+}(ir+\frac{1}{2}),\Delta f_{i}^{+}(ir-\frac{1}{2})]\\ \tt{,} (14)
HR(ir+12)=fi(ir+1)12minmod[Δfi(ir+12),Δfi(ir+32)].{{H}_{R}}(ir+\frac{1}{2})=f_{i}^{-}(ir+1)-\frac{1}{2}{\rm{minmod}}[\Delta f_{i}^{-}(ir+\frac{1}{2}),\Delta f_{i}^{-}(ir+\frac{3}{2})]\tt{.} (15)

The function minmod{\rm{minmod}} is also a type of flux limiter, defined as follows:

minmod[X,Y]={0,Y=0orXY0X,|XY|1Y,|XY|>1.{\rm{minmod}}\left[X,Y\right]=\begin{cases}0,&Y=0\quad{\rm{or}}\quad XY\leqslant 0\\[4.30554pt] X,&\left|\frac{X}{Y}\right|\leqslant 1\\[4.30554pt] Y,&\left|\frac{X}{Y}\right|>1\end{cases}\tt{.} (16)

In addition,

Δfi+(ir+12)=fi+(ir+1)fi+(ir),\Delta f_{i}^{+}(ir+\frac{1}{2})=f_{i}^{+}(ir+1)-f_{i}^{+}(ir)\tt{,} (17)
Δfi(ir+12)=fi(ir+1)fi(ir),\Delta f_{i}^{-}(ir+\frac{1}{2})=f_{i}^{-}(ir+1)-f_{i}^{-}(ir)\tt{,} (18)
fi+(ir)=12(vir+|vir|)fi,f_{i}^{+}(ir)=\frac{1}{2}({{v}_{ir}}+\left|{{v}_{ir}}\right|){{f}_{i}}\tt{,} (19)
fi(ir)=12(vir|vir|)fi,f_{i}^{-}(ir)=\frac{1}{2}({{v}_{ir}}-\left|{{v}_{ir}}\right|){{f}_{i}}\tt{,} (20)

where irir denotes the ii-th grid point in the rr direction. The NND scheme achieves second-order spatial accuracy. This scheme can suppress odd-even decoupling oscillations and effectively capture strong discontinuities.

3 Numerical validation

In this section, let us verify that the CDBM is not only applicable for high-speed compressible flows, but can also capture the nonequilibrium effects accurately. To this end, five representative benchmarks are considered: the Sod shock tube and Lax shock tube validate the ability of the CDBM to capture discontinuities under varying specific heat ratios, the shock wave test demonstrates the suitability of the CDBM for hypervelocity compressible flows, the two-dimensional sound wave and decaying Taylor-Green vortex flow are used to verify the efficiency of CDBM in two-dimensional cases.

3.1 Sod shock tube

First, we consider two typical Riemann problems, i.e., the Sod shock tube and the Lax shock tube. For the Sod shock tube, the initial conditions are given by

{(ρ,ux,uy,T)L=(1,0,0,1),(ρ,ux,uy,T)R=(0.125,0,0,0.8).\left\{\begin{array}[]{*{35}{l}}{{\left(\rho,{u}_{x},{u}_{y},T\right)}_{L}}=\left(1,0,0,1\right)\tt{,}\\ {{\left(\rho,{u}_{x},{u}_{y},T\right)}_{R}}=\left(0.125,0,0,0.8\right)\tt{.}\end{array}\right.

Here, LL and RR denote the regions 0<x0.0750<x\leq 0.075 and 0.075<x0.150.075<x\leq 0.15, respectively. The simulation parameters are specified as follows: the grid mesh is Nx×Ny=1500×1N_{x}\times N_{y}=1500\times 1, with a spatial resolution of Δx=Δy=1×104\Delta x=\Delta y=1\times 10^{-4}, a time step of Δt=2×105\Delta t=2\times 10^{-5}, and a specific heat ratio γ=2\gamma=2. The discrete velocity set is (va,vb,vc,vd)=(2.5,2.5,1.2,1.1)(v_{a},v_{b},v_{c},v_{d})=(2.5,2.5,1.2,1.1) and (ηa,ηb,ηc,ηd)=(0,0,0,1.2)(\eta_{a},\eta_{b},\eta_{c},\eta_{d})=(0,0,0,1.2). Additionally, the relaxation time is set to τ=1×104\tau=1\times 10^{-4}.

Refer to caption
Figure 2: Profiles of the density ρ{\rho} (a) and the nonequilibrium quantity Δ2,xx{\Delta}_{2,xx}^{*} (b) in the Sod shock tube.

Figure 2 (a) illustrates the evolution of density in the Sod shock tube. The stars, circles, and solid line correspond to the DBM results, CDBM results, and the Riemann solutions, respectively. The simulation results exhibit good agreement with the analytical Riemann solutions. Some discrepancies are observed between the numerical results and the Riemann solutions, particularly near the rarefaction wave, material interface, and shock front. These arise from the fact that the CDBM simulation results incorporate essential thermodynamic nonequilibrium effects, which are neglected by the Riemann solutions.

Figure 2 (b) shows the profile of twice the nonorganized energy along the xx direction. The stars and circles denote the simulation results obtained from the DBM and CDBM, while the solid line represents the exact solution [73]

Δ2,xx=2ρTτ(1DID+Iuxx+1D+Iuyy).\Delta_{2,xx}^{*}={2\rho T}{\tau}\biggl{(}\frac{1-D-I}{D+I}\frac{\partial u_{x}}{\partial x}+\frac{1}{D+I}\frac{\partial u_{y}}{\partial y}\biggr{)}\tt{.} (21)

Obviously, the nonorganized energy near rarefaction wave exhibits negativity which is illustrated in the upper subplot, and a peak emerges around the shock wave as depicted in the lower subplot. Furthermore, the nonorganized energy approaches zero in other regions, with minor numerical oscillations observed at the contact wave. In fact, the minor undershoot observed in the Sod shock tube arises due to the sharp discontinuity in the physical field inherent in the initial configuration. This artificial discontinuity deviates from the characteristics of a natural interface, which typically exhibits a smooth transition layer. Consequently, a minor undershoot develops near this discontinuity during the early stages and gradually dissipates over time. Importantly, the CDBM results align more closely with the exact solution than the DBM results across the entire profile, especially in the peak region. This provides evidence that the CDBM is highly effective in capturing nonequilibrium effects in various regions of the Sod shock tube.

3.2 Lax shock tube

The initial condition for the Lax shock tube is

{(ρ,ux,uy,T)L=(0.445,0.698,0,7.928),(ρ,ux,uy,T)R=(0.5,0,0,1.142),\left\{\begin{array}[]{*{35}{l}}{{\left(\rho,{u}_{x},{u}_{y},T\right)}_{L}}=\left(0.445,0.698,0,7.928\right)\tt{,}\\ {{\left(\rho,{u}_{x},{u}_{y},T\right)}_{R}}=\left(0.5,0,0,1.142\right)\tt{,}\end{array}\right.

where L[0,1)L\in[0,1) and R[1,2]R\in[1,2] stand for the left and right sides, respectively. The computational grid is specified as Nx×Ny=1000×1N_{x}\times N_{y}=1000\times 1, with a spatial resolution of Δx=Δy=2×103\Delta x=\Delta y=2\times 10^{-3} and a time step of Δt=2×105\Delta t=2\times 10^{-5}. In addition, the specific heat ratio is γ=1.4\gamma=1.4, and the discrete velocity set is (va,vb,vc,vd)=(4.4,4.4,3,1.8)(v_{a},v_{b},v_{c},v_{d})=(4.4,4.4,3,1.8) and (ηa,ηb,ηc,ηd)=(0,0,5,0)(\eta_{a},\eta_{b},\eta_{c},\eta_{d})=(0,0,5,0). The relaxation time is τ=1×104\tau=1\times 10^{-4}.

Refer to caption
Figure 3: Profiles of the density ρ{\rho} (a), temperature TT (b), Mach number Ma\rm{Ma} (c), and pressure PP (d) in the Lax shock tube.

Figure 3 presents the results of the Lax problem at t=0.2t=0.2. The Riemann solutions are represented by lines, while the symbols denote the computed quantities obtained using the CDBM. It can be observed that the simulation results exhibit a good agreement with the analytical solutions. To be specific, there is a rarefaction wave propagating to the left, a material interface in the center, and a shock wave traveling to the right. Consequently, the CDBM successfully captures all three distinct structures.

3.3 Shock wave

The CDBM is suitable for high-speed compressible flow. To assess its performance at high Mach numbers, we simulate a shock wave with a Mach number of Ma=15\rm{Ma}=15. The initial conditions are specified as follows,

{(ρ,ux,uy,T,P)L=(5.8696,14.7245,0,44.6938,262.3333),(ρ,ux,uy,T,P)R=(1,0,0,1,1).\left\{\begin{array}[]{*{35}{l}}{{\left(\rho,{u}_{x},{u}_{y},T,P\right)}_{L}}=\left(5.8696,14.7245,0,44.6938,262.3333\right)\tt{,}\\ {{\left(\rho,{u}_{x},{u}_{y},T,P\right)}_{R}}=\left(1,0,0,1,1\right)\tt{.}\end{array}\right.

Here, LL and RR stand for 0<x0.020<x\leq 0.02 and 0.02<x1.50.02<x\leq 1.5, respectively. The simulation parameters are as follows: the grid mesh is Nx×Ny=10000×1N_{x}\times N_{y}=10000\times 1, the spatial resolution is Δx=Δy=2×104\Delta x=\Delta y=2\times 10^{-4}, the time step is Δt=2×106\Delta t=2\times 10^{-6}, the specific heat ratio is γ=1.4\gamma=1.4. The discrete velocity set is given as (va,vb,vc,vd)=(21.8,21.8,17.8,9.8)(v_{a},v_{b},v_{c},v_{d})=(21.8,21.8,17.8,9.8) and (ηa,ηb,ηc,ηd)=(18.8,0,0,0)(\eta_{a},\eta_{b},\eta_{c},\eta_{d})=(18.8,0,0,0).

Figures 4 (a)-(d) illustrate the profiles of density, pressure, horizontal velocity and temperature, respectively. The symbols represent the CDBM results, The symbols represent the CDBM results, while the solid lines denote the Riemann solutions. Clearly, the CDBM results align closely with the Riemann solutions. Therefore, the CDBM successfully captures the shock wave at a Mach number of Ma=15\rm{Ma}=15. In other words, the model is suitable for a wide range of flow regimes, from incompressible flows to hypervelocity compressible flows.

Refer to caption
Figure 4: Profiles of the density (a), pressure (b), horizontal velocity (c) and temperature (d) in the shock wave.

3.4 Sound wave

Then, the sound wave simulation verifies the suitability of the CDBM for compressible flows. Figure 5 illustrates the initial condition. A perturbation is introduced at position x0x_{0}, spreading at the speed of sound. Simultaneously, the perturbation propagates rightward at twice the speed of sound. Therefore, the angle of propagation can be utilized to verify the accuracy of our model.

Refer to caption
Figure 5: The sketch of the propagation for the two-dimensional sound wave.

The grid mesh size is Nx×Ny=1600×1200N_{x}\times N_{y}=1600\times 1200, with a spatial resolution of Δx=Δy=5×104\Delta x=\Delta y=5\times 10^{-4}, and a time step of Δt=2×105\Delta t=2\times 10^{-5}. In addition, the relaxation time is τ=1×104\tau=1\times 10^{-4}, the kinematic viscosity is μ=1×104\mu=1\times 10^{-4}, the specific heat ratio is γ=1.4\gamma=1.4, and the discrete velocity set is (va,vb,vc,vd)=(1.4,1.1,1.1,1)(v_{a},v_{b},v_{c},v_{d})=(1.4,1.1,1.1,1) and (ηa,ηb,ηc,ηd)=(3,0,0,0)(\eta_{a},\eta_{b},\eta_{c},\eta_{d})=(3,0,0,0).

Figure 6 illustrates the propagation of two-dimensional sound waves at different times: (a) tt = 0.02, (b) tt=0.1, and (c) tt=0.2. At tt=0.2, the exact solution is sinθ=0.5\sin\theta=0.5, while the simulation result is sinθ=0.50106\sin\theta=0.50106. The relative error between the simulation result and the exact solution is 0.212%0.212\%, indicating satisfactory agreement. These results demonstrate that the CDBM is well-suited for modeling two-dimensional compressible waves.

Refer to caption
Figure 6: Profiles of the two-dimensional sound waves at various time instants, (a) tt = 0.02, (b) tt=0.1, and (c) tt=0.2.

3.5 Taylor-Green vortex flow

Finanlly, the CDBM was validated by a two-dimensional Taylor-Green vortex flow. The solution of this flow problem can be given analytically as

ux(x,y,t)=u0cos(πx/L)sin(πy/L)e2π2u0t/(ReL),u_{x}(x,y,t)=-u_{0}\cos(\pi x/L)\sin(\pi y/L)e^{-2\pi^{2}u_{0}t/({\mathop{\rm Re}\nolimits}L)}\tt{,} (22)
uy(x,y,t)=+u0sin(πx/L)cos(πy/L)e2π2u0t/(ReL),u_{y}(x,y,t)=+u_{0}\sin(\pi x/L)\cos(\pi y/L)e^{-2\pi^{2}u_{0}t/({\mathop{\rm Re}\nolimits}L)}\tt{,} (23)
p(x,y,t)=p0p0u024[cos(2πx/L)+cos(2πy/L)]e4π2u0t/(ReL),p\bigl{(}x,y,t\bigr{)}=p_{0}-\frac{p_{0}u_{0}^{2}}{4}\Bigl{[}\cos\bigl{(}2\pi x/L\bigr{)}+\cos\bigl{(}2\pi y/L\bigr{)}\Bigr{]}e^{-4\pi^{2}u_{0}t/\bigl{(}{\mathop{\rm Re}\nolimits}L\bigr{)}}\tt{,} (24)

where u0=0.01u_{0}=0.01 denotes the reference velocity, p0=1.0p_{0}=1.0 represents the reference pressure, L=0.05L=0.05 signifies the reference length, and the Reynolds number, Re{\mathop{\rm Re}\nolimits}, is defined as Re=ρ0u0L/μ=1{\mathop{\rm Re}\nolimits}=\rho_{0}u_{0}L/\mu=1, where kinematic viscosity is μ=5×104\mu=5\times 10^{-4}. The computational domain for this flow problem extends over [0,2L][0,2L], with the grid discretization performed using a mesh of size Nx×Ny=100×100N_{x}\times N_{y}=100\times 100. Additional parameters for this simulation include the spatial resolution Δx=Δy=1×103\Delta x=\Delta y=1\times{10^{-3}}, the temporal step Δt=2×105\Delta t=2\times{10^{-5}}, a specific heat ratio of γ=1.4\gamma=1.4, a discrete velocity set of (va,vb,vc,vd)=(0.2,0.2,2,2)(v_{a},v_{b},v_{c},v_{d})=(0.2,0.2,2,2), and (ηa,ηb,ηc,ηd)=(0,5,0,2.6)(\eta_{a},\eta_{b},\eta_{c},\eta_{d})=(0,5,0,2.6). Moreover, the relaxation time is τ=5×104\tau=5\times 10^{-4}.

Refer to caption
Figure 7: Comparison of CDBM simulation and analytical solutions for Taylor-Green vortex velocity fields at t=0.1t=0.1. (a) Horizontal velocity of CDBM, (b) vertical velocity of CDBM, (c) Horizontal velocity of analytical solution, and (d) vertical velocity of analytical solution.

In this section, we compare the simulation results obtained from the CDBM with the exact analytical solutions for the Taylor-Green vortex flow. Figures 7 (a) and (b) depict the horizontal and vertical velocity fields obtained from the CDBM simulation, whereas Figs. 7 (c) and (d) illustrate the corresponding analytical results. Both sets of results exhibit a high degree of qualitative agreement, demonstrating similar periodic structures and symmetry in the velocity fields. The close alignment of velocity magnitudes, as indicated by the color bars, further confirms the accuracy of the CDBM simulation.

In addition, the relative error of velocity component between the numerical result and the analytical solution is measured using the L2L_{2} norm, defined as follows,

L2(ux)relative=1Nx×Nyi,j(ux(i,j)numericalux(i,j)exactu0)2.{{L}_{2}}{{\left(\begin{matrix}{{u}_{x}}\\ \end{matrix}\right)}_{relative}}=\sqrt{\frac{1}{{{N}_{x}}\times{{N}_{y}}}\sum\limits_{i,j}{{{\left(\frac{{{u}_{x}}_{\left(i,j\right)}^{numerical}-{{u}_{x}}_{\left(i,j\right)}^{exact}}{{{u}_{0}}}\right)}^{2}}}}\tt{.} (25)

Five different uniform grid meshes with sizes of 20×2020\times 20, 30×3030\times 30, 40×4040\times 40, 50×5050\times 50, and 60×6060\times 60 are used to discretize the domain. The corresponding spatial steps are 5×1035\times 10^{-3}, 3.34×1033.34\times 10^{-3}, 2.5×1032.5\times 10^{-3}, 2×1032\times 10^{-3}, and 1.67×1031.67\times 10^{-3}, respectively.

Refer to caption
Figure 8: L2L_{2} norm of the relative error of horizontal velocity versus space step Δx\Delta x for decaying vortex flow.

Figure 8 showes the numerical results of L2L_{2} norms, where the space step Δx\Delta x is plotted on a logarithmic scale. The slope of the line is 1.69, which is near the second order in theory. These findings suggest that the CDBM method effectively captures the dynamics of the Taylor-Green vortex flow, validating its reliability for studying complex two-dimensional fluid flow phenomena.

4 Conclusion

In summary, a CDBM is developed for kinetic modeling of compressible flows with both hydrodynamic and thermodynamic nonequilibrium effects. The central kinetic moments are employed for calculating the equilibrium discrete distribution functions. Conservation moments of the discrete velocity distribution function are utilized to derive macroscopic physical quantities, while higher-order central kinetic moments are employed to characterize nonequilibrium effects. Its ability to capture nonequilibrium effects is demonstrated through the simulation of the Sod shock tube. The results exhibit excellent agreement with exact solutions. Furthermore, the Lax shock tube benchmark confirms that the CDBM can accurately capture discontinuities under different specific heat ratios. Additionally, the simulation of shock waves at Mach number Ma=15\rm{Ma}=15 verifies the capability of the CDBM for hypervelocity compressible flows. The results of two-dimensional sound wave propagation and decaying Taylor-Green vortex flow further showcase the efficiency of the CDBM in two-dimensional cases. Overall, the proposed CDBM proves to be a robust and promising tool for investigating complex compressible fluid systems, particularly those with significant thermodynamic nonequilibrium effects at the mesoscopic level.

Acknowledgements

This work is supported by Guangdong Basic and Applied Basic Research Foundation (under Grant No. 2022A1515012116), China Scholarship Council (Nos. 202306380179 and 202306380288), and Fundamental Research Funds for the Central Universities, Sun Yat-sen University (under Grant No. 24qnpy044). Support from the UK Engineering and Physical Sciences Research Council under the project “UK Consortium on Mesoscale Engineering Sciences (UKCOMES)” (Grant No. EP/X035875/1) is gratefully acknowledged. This work made use of computational support by CoSeC, the Computational Science Centre for Research Communities, through UKCOMES.

References

  • [1] P. Balachandran, Fundamentals of compressible fluid dynamics, PHI Learning Pvt. Ltd., 2006.
  • [2] J.D. Anderson, Modern compressible flow: with historical perspective, Vol. 12, McGraw-Hill New York, 2021.
  • [3] S.K. Aliabadi and T.E. Tezduyar, Space-time finite element computation of compressible flows involving moving boundaries and interfaces, Comput. Methods Appl. Mech. Eng. 107 (1993), pp. 209–223.
  • [4] T.J. Hughes, G. Scovazzi, and T.E. Tezduyar, Stabilized methods for compressible flows, J. Sci. Comput. 43 (2010), pp. 343–368.
  • [5] K. Aoki, C. Baranger, M. Hattori, S. Kosuge, G. Martalò, J. Mathiaud, and L. Mieussens, Slip boundary conditions for the compressible Navier-Stokes equations, J. Stat. Phys. 169 (2017), pp. 744–781.
  • [6] J.B. Pedro Costa, On the numerical simulation of compressible flows (2019).
  • [7] A.A. Shershnev, A.N. Kudryavtsev, A.V. Kashkovsky, G.V. Shoev, S.P. Borisov, T.Y. Shkredov, D.P. Polevshchikov, A.A. Korolev, D.V. Khotyanovsky, and Y.V. Kratova, A Numerical Code for a Wide Range of Compressible Flows on Hybrid Computational Architectures, Supercomput. Front. Innov. 9 (2022), pp. 85–99.
  • [8] R.K. Agarwal and K.Y. Yun, Burnett equations for simulation of transitional flows, Appl. Mech. Rev. 55 (2002), pp. 219–240.
  • [9] R.S. Jadhav, N. Singh, and A. Agrawal, Force-driven compressible plane Poiseuille flow by Onsager-Burnett equations, Phys. Fluids 29 (2017).
  • [10] M. Sun and C. Ebner, Molecular-dynamics simulation of compressible fluid flow in two-dimensional channels, Phys. Rev. A 46 (1992), p. 4813.
  • [11] I. Kandemir and A.M. Kaya, Molecular dynamics simulation of compressible hot/cold moving lid-driven microcavity flow, Microfluid. Nanofluid. 12 (2012), pp. 509–520.
  • [12] G.A. Bird, Molecular gas dynamics and the direct simulation of gas flows, Oxford university press, 1994.
  • [13] L. Mieussens, Discrete-velocity models and numerical schemes for the Boltzmann-BGK equation in plane and axisymmetric geometries, J. Comput. Phys. 162 (2000), pp. 429–466.
  • [14] L. Yang, C. Shu, J. Wu, and Y. Wang, Comparative study of discrete velocity method and high-order lattice Boltzmann method for simulation of rarefied flows, Comput. Fluids 146 (2017), pp. 125–142.
  • [15] K. Xu and J.C. Huang, A unified gas-kinetic scheme for continuum and rarefied flows, J. Comput. Phys. 229 (2010), pp. 7747–7764.
  • [16] S. Chen, K. Xu, C. Lee, and Q. Cai, A unified gas kinetic scheme with moving mesh and velocity space adaptation, J. Comput. Phys. 231 (2012), pp. 6643–6664.
  • [17] Z. Guo, R. Wang, and K. Xu, Discrete unified gas kinetic scheme for all Knudsen number flows. II. Thermal compressible case, Phys. Rev. E 91 (2015), p. 033313.
  • [18] Z. Guo and K. Xu, Progress of discrete unified gas-kinetic scheme for multiscale flows, Adv. Aerodyn. 3 (2021), pp. 1–42.
  • [19] S. Succi, The Lattice Boltzmann Equation for Fluid Dynamics and Beyond, Oxford University Press, New York, 2001.
  • [20] Z. Guo and C. Shu, Lattice Boltzmann method and its applications in engineering, World Scientific, Singapore, 2013.
  • [21] A. Xu, G. Zhang, Y. Gan, F. Chen, and X. Yu, Lattice Boltzmann modeling and simulation of compressible flows, Front. Phys. 7 (2012), pp. 582–600.
  • [22] B. Yan, A. Xu, G. Zhang, Y. Ying, and H. Li, Lattice Boltzmann model for combustion and detonation, Front. Phys. 8 (2013), pp. 94–110.
  • [23] N.G. Hadjiconstantinou, A.L. Garcia, M.Z. Bazant, and G. He, Statistical error in particle simulations of hydrodynamic phenomena, J. Comput. Phys. 187 (2003), pp. 274–297.
  • [24] R. Haghani Hassan Abadi, A. Fakhari, and M.H. Rahimian, Numerical simulation of three-component multiphase flows at high density and viscosity ratios using lattice Boltzmann methods, Phys. Rev. E 97 (2018), p. 033312.
  • [25] A.F. Di Rienzo, P. Asinari, E. Chiavazzo, N. Prasianakis, and J. Mantzaras, Lattice Boltzmann model for reactive flow simulations, Europhys. Lett. 98 (2012), p. 34001.
  • [26] I. Cherkaoui, S. Bettaibi, A. Barkaoui, and F. Kuznik, Magnetohydrodynamic blood flow study in stenotic coronary artery using lattice Boltzmann method, Comput. Methods Programs Biomed. 221 (2022), p. 106850.
  • [27] A. Afzalabadi, V. Bordbar, and M.A. Amooie, Lattice Boltzmann simulation of magneto-hydrodynamic double-diffusive natural convection in an enclosure with internal heat and solute source using nanofluids, SN Appl. Sci 2 (2020), pp. 1–19.
  • [28] D. Xu, Y. Luo, and Y. Xu, Study on deposition characteristics of microparticles in terminal pulmonary acini by IB-LBM, Micromachines 12 (2021), p. 957.
  • [29] X. Liu, Z. Tong, and Y. He, Enthalpy-based immersed boundary-lattice Boltzmann model for solid-liquid phase change in porous media under local thermal non-equilibrium condition, Int. J. Therm. Sci. 182 (2022), p. 107786.
  • [30] F.J. Alexander, S. Chen, and J. Sterling, Lattice boltzmann thermohydrodynamics, Phys. Rev. E 47 (1993), p. R2249.
  • [31] Y. Qian, Simulating thermohydrodynamics with lattice BGK models, J. Sci. Comput. 8 (1993), pp. 231–242.
  • [32] Y. Chen, H. Ohashi, and M. Akiyama, Thermal lattice Bhatnagar-Gross-Krook model without nonlinear deviations in macrodynamic equations, Phys. Rev. E 50 (1994), p. 2776.
  • [33] Y. Chen, H. Ohashi, and M. Akiyama, Two-parameter thermal lattice BGK model with a controllable Prandtl number, J. Sci. Comput. 12 (1997), pp. 169–185.
  • [34] G.R. McNamara, A.L. Garcia, and B.J. Alder, A hydrodynamically correct thermal lattice Boltzmann model, J. Stat. Phys. 87 (1997), pp. 1111–1121.
  • [35] H. Shouxin, Y. Guangwu, and S. Weiping, A lattice Boltzmann model for compressible perfect gas, Acta Mech. Sin. 13 (1997), pp. 218–226.
  • [36] Y. Guangwu, C. Yaosong, and H. Shouxin, Simple lattice Boltzmann model for simulating flows with shock wave, Phys. Rev. E 59 (1999), p. 454.
  • [37] C. Sun, Adaptive lattice Boltzmann model for compressible flows: viscous and conductive properties, Phys. Rev. E 61 (2000), p. 2645.
  • [38] C. Sun, Simulations of compressible flows with strong shocks by an adaptive lattice Boltzmann model, J. Comput. Phys. 161 (2000), pp. 70–84.
  • [39] S. Chenghai, Lattice-Boltzmann model for compressible perfect gases, Acta Mech. Sin. 16 (2000), pp. 289–300.
  • [40] C. Sun and A.T. Hsu, Three-dimensional lattice Boltzmann model for compressible flows, Phys. Rev. E 68 (2003), p. 016303.
  • [41] C. Sun and A. Hsu, Multi-level lattice Boltzmann model on square lattice for compressible flows, Comput. Fluids 33 (2004), pp. 1363–1385.
  • [42] Q. Li, Y. He, Y. Wang, and W. Tao, Coupled double-distribution-function lattice Boltzmann method for the compressible Navier-Stokes equations, Phys. Rev. E 76 (2007), p. 056705.
  • [43] L. Yang, C. Shu, and J. Wu, Development and comparative studies of three non-free parameter lattice Boltzmann models for simulation of compressible flows, Adv. Appl. Math. Mech. 4 (2012), pp. 454–472.
  • [44] L. Yang, C. Shu, and J. Wu, A moment conservation-based non-free parameter compressible lattice Boltzmann model and its application for flux evaluation at cell interface, Comput. Fluids 79 (2013), pp. 190–199.
  • [45] K. Li and C. Zhong, A lattice Boltzmann model for simulation of compressible flows, Int. J. Numer. Methods Fluids 77 (2015), pp. 334–357.
  • [46] B. Dorschner, F. Bösch, and I.V. Karlin, Particles on demand for kinetic theory, Phys. Rev. Lett. 121 (2018), p. 130602.
  • [47] N.G. Kallikounis, B. Dorschner, and I.V. Karlin, Particles on demand for flows with strong discontinuities, Phys. Rev. E 106 (2022), p. 015301.
  • [48] N. Kallikounis and I. Karlin, Particles on demand method: Theoretical analysis, simplification techniques, and model extensions, Phys. Rev. E 109 (2024), p. 015304.
  • [49] Y. Ji, S.A. Hosseini, B. Dorschner, K.H. Luo, and I.V. Karlin, Eulerian discrete kinetic framework in comoving reference frame for hypersonic flows, J. Fluid Mech 983 (2024), p. A11.
  • [50] C. Coreixas, B. Chopard, and J. Latt, Comprehensive comparison of collision models in the lattice Boltzmann framework: Theoretical investigations, Phys. Rev. E 100 (2019), p. 033305.
  • [51] C. Coreixas, G. Wissocq, B. Chopard, and J. Latt, Impact of collision models on the physical properties and the stability of lattice Boltzmann methods, Philos. Trans. R. Soc. A 378 (2020), p. 20190397.
  • [52] M. Geier, A. Greiner, and J.G. Korvink, Cascaded digital lattice Boltzmann automata for high Reynolds number flow, Phys. Rev. E 73 (2006), p. 066705.
  • [53] D. Lycett-Brown, K.H. Luo, R. Liu, and P. Lv, Binary droplet collision simulations by a multiphase cascaded lattice Boltzmann method, Phys. Fluids 26 (2014).
  • [54] D. Lycett-Brown and K.H. Luo, Multiphase cascaded lattice Boltzmann method, Comput. Math. Appl. 67 (2014), pp. 350–362.
  • [55] Y. Ning, K.N. Premnath, and D.V. Patil, Numerical study of the properties of the central moment lattice Boltzmann method, Int. J. Numer. Methods Fluids 82 (2016), pp. 59–90.
  • [56] L. Fei and K.H. Luo, Consistent forcing scheme in the cascaded lattice Boltzmann method, Phys. Rev. E 96 (2017), p. 053307.
  • [57] L. Fei, K.H. Luo, C. Lin, and Q. Li, Modeling incompressible thermal flows using a central-moments-based lattice Boltzmann method, Int. J. Heat Mass Transfer 120 (2018), pp. 624–634.
  • [58] X. Liu, Z. Tong, Y. He, S. Du, and M. Li, Enthalpy-based cascaded lattice Boltzmann method for conjugate heat transfer, Int. Commun. Heat Mass Transfer 159 (2024), p. 107956.
  • [59] P. Asinari, Generalized local equilibrium in the cascaded lattice Boltzmann method, Phys. Rev. E 78 (2008), p. 016701.
  • [60] A. De Rosis and K.H. Luo, Role of higher-order Hermite polynomials in the central-moments-based lattice Boltzmann framework, Phys. Rev. E 99 (2019), p. 013301.
  • [61] M. Geier, A. Pasquali, and M. Schönherr, Parametrization of the cumulant lattice Boltzmann method for fourth order accurate diffusion part II: Application to flow around a sphere at drag crisis, J. Comput. Phys. 348 (2017), pp. 889–898.
  • [62] G. Wissocq and P. Sagaut, Hydrodynamic limits and numerical errors of isothermal lattice Boltzmann schemes, J. Comput. Phys. 450 (2022), p. 110858.
  • [63] C. Lin, A. Xu, G. Zhang, Y. Li, and S. Succi, Polar-coordinate lattice Boltzmann modeling of compressible flows, Phys. Rev. E 89 (2014), p. 013307.
  • [64] H. Lai, A. Xu, G. Zhang, Y. Gan, Y. Ying, and S. Succi, Nonequilibrium thermohydrodynamic effects on the Rayleigh–Taylor instability in compressible flows, Phys. Rev. E 94 (2016), p. 023106.
  • [65] Y. Gan, A. Xu, G. Zhang, Y. Zhang, and S. Succi, Discrete Boltzmann trans-scale modeling of high-speed compressible flows, Phys. Rev. E 97 (2018), p. 053312.
  • [66] C. Lin, X. Su, and Y. Zhang, Hydrodynamic and Thermodynamic Nonequilibrium Effects around Shock Waves: Based on a Discrete Boltzmann Method, Entropy 22 (2020), p. 1397.
  • [67] L. Chen, H. Lai, C. Lin, and D. Li, Specific heat ratio effects of compressible Rayleigh—Taylor instability studied by discrete Boltzmann method, Front. Phys. 16 (2021), pp. 1–12.
  • [68] X. Su and C. Lin, Nonequilibrium effects of reactive flow based on gas kinetic theory, Commun. Theor. Phys. 74 (2022), p. 035604.
  • [69] X. Su and C. Lin, Unsteady detonation with thermodynamic nonequilibrium effect based on the kinetic theory, Commun. Theor. Phys. 75 (2023).
  • [70] Y. Gan, A. Xu, G. Zhang, and S. Succi, Discrete Boltzmann modeling of multiphase flows: hydrodynamic and thermodynamic non-equilibrium effects, Soft Matter 11 (2015), pp. 5336–5345.
  • [71] C. Lin, A. Xu, G. Zhang, and Y. Li, Double-distribution-function discrete Boltzmann model for combustion, Combust. Flame 164 (2016), pp. 137–151.
  • [72] Z. Liu, J. Song, A. Xu, Y. Zhang, and K. Xie, Discrete Boltzmann modeling of plasma shock wave, Proc. Inst. Mech. Eng., Part C: J. Mech. Eng. Sci. 237 (2023), pp. 2532–2548.
  • [73] C. Lin and K.H. Luo, Discrete Boltzmann modeling of unsteady reactive flows with nonequilibrium effects, Phys. Rev. E 99 (2019), p. 012142.
  • [74] Y. Zhang, A. Xu, G. Zhang, C. Zhu, and C. Lin, Kinetic modeling of detonation and effects of negative temperature coefficient, Combust. Flame 173 (2016), pp. 483–492.
  • [75] C. Lin, Simplified two-dimensional discrete Boltzmann model of high-speed compressible reactive flows, Acta Aerodynamica Sinica 40 (2022), pp. 98–108.
  • [76] H. Zhang and F. Zhuang, NND schemes and their applications to numerical simulation of two-and three-dimensional flows, Adv. Appl. Mech. 29 (1991), pp. 193–256.