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

Further acceleration of multiscale simulation of rarefied gas flow via a generalized boundary treatment

Wei Liu Yanbing Zhang Jianan Zeng Lei Wu [email protected] Department of Mechanics and Aerospace Engineering, Southern University of Science and Technology, Shenzhen, 518055, China Division of Emerging Interdisciplinary Areas, The Hong Kong University of Science and Technology, Clear Water Bay, Hong Kong, China
Abstract

The recently-developed general synthetic iterative scheme (GSIS) is efficient in simulating multiscale rarefied gas flows due to the coupling of mesoscopic kinetic equation and macroscopic synthetic equation: for linearized Poiseuille flow where the boundary flux is fixed at each iterative step, the steady-state solutions are found within dozens of iterations in solving the gas kinetic equations, while for general nonlinear flows the iteration number is increased by about one order of magnitude, caused by the incompatible treatment of the boundary flux for the macroscopic synthetic equation. In this paper, we propose a generalized boundary treatment (GBT) to further accelerate the convergence of GSIS. The main idea is, the truncated velocity distribution function at the boundary, similar to that used in the Grad 13-moment equation, is reconstructed by the macroscopic conserved quantities from the synthetic equation, and the high-order correction of non-equilibrium stress and heat flux from the kinetic equation; therefore, in each inner iteration solving the synthetic equation, the explicit constitutive relations facilitate real-time updates of the macroscopic boundary flux, driving faster information exchange in the flow field, and consequently achieving quicker convergence. Moreover, the high-order correction derived from the kinetic equation can compensate the approximation by the truncation and ensure the boundary accuracy. The one-dimensional Fourier flow, two-dimensional hypersonic flow around a cylinder, three-dimensional pressure-driven pipe flow and the flow around the hypersonic technology vehicle are simulated. The accuracy of GSIS-GBT is validated by the direct simulation Monte Carlo method, the previous versions of GSIS, and the unified gas-kinetic wave-particle method. For the efficiency, in the near-continuum flow regime and slip regime, GSIS-GBT can be faster than the conventional iteration scheme in the discrete velocity method and the previous versions of GSIS by two- and one-order of magnitude, respectively.

keywords:
Rarefied gas flow , Boltzmann kinetic equation , Finite volume method , general synthetic iterative scheme , hypersonic flow , low-speed internal flow
journal: Elsevier

1 Introduction

Multiscale rarefied gas flows are encountered in many engineering problems, such as the high-altitude aerothermodynamics of space vehicles [1, 2], micro-electro-mechanical systems [3], and gas transportation within ultra-tight shale rock [4, 5]. For instance, supersonic flows around the X38-like vehicle encompass the free-molecular to continuum flow regimes, where the Knudsen number (Kn, the ratio of mean free path of gas molecules to characteristic flow length) traverses a range of five orders of magnitude [6].

While the classic Navier-Stokes (NS) equations are only applicable up to the slip flow regime where Kn0.1Kn\lessapprox 0.1, the Boltzmann equation transcends the limitations of continuum assumption and describes the gas flows from the continuum, slip, transition and free-molecular regimes. The discrete velocity method (DVM) and the direct simulation Monte Carlo (DSMC) method are the two representative numerical methods to solve the Boltzmann equation [7, 8], which are efficient when the Knudsen number is large. However, a significant challenge arises when the rarefaction effects gradually diminish and the gas become denser. Hence, the spatial cell size and the time step must be smaller than the limits imposed by the molecular mean free path and mean collision time, respectively [9], which not only requires huge amount of computer memory, but also converges slowly to the steady-state solution.

The common feature of the DVM and DSMC is that the streaming and collision operators in the Boltzmann equation are handled separately, which results in large numerical dissipation when the Knudsen number is small, so that the cell size and time step have to be small. Based on the integral solution of the Boltzmann-BGK equation, the unified gas-kinetic scheme (UGKS) has been proposed to simultaneously handle the streaming and collision, via the multiscale flux reconstruction in the finite-volume framework [10]. As a result, it has the asymptotic preserving property to overcome the constraints in cell size and time step. This concept and technique have been further explored, and many multiscale schemes are developed [11, 12, 13].

Even though the numerical dissipation in near-continuum flows is solved, the slow convergence remains a persistent problem. In time-dependent solvers, steady-state solutions are usually found after 10510^{5} to 10610^{6} steps [14]. In the implicit solver such as the conventional iterative scheme (CIS, which evolves the temporal derivatives, the streaming term, and the loss part of the collision operator at the current time step, while the gain part of the collision operator is computed based on macroscopic quantities from the previous iteration step), steady-state solutions can be found within tens to hundreds iterations for flows in the free-molecular and transition regimes; however, the iteration number in simulating near-continuum flows experiences a substantial surge, e.g., to about 10610^{6} steps when Kn0.001\text{Kn}\approx 0.001 [14]. This poses a huge computational burden in the DVM since the velocity distribution function in the Boltzmann equation is defined in a six-dimensional phase space. For efficient simulation, the iteration number should be controlled within 10310^{3} or even 10210^{2}, at all Knudsen numbers.

Building on Larsen’s work in accelerating the convergence in neutron transport [15], the concept of mesoscopic-macroscopic coupling for iteration acceleration has been extended to rarefied gas dynamics, including the synthetic iterative scheme [16, 17], the moment-guided Monte Carlo method [18], the high-order/low-order method [19], and the general synthetic iterative scheme (GSIS) [20]. Particularly, the GSIS been successfully applied to gas transportation in porous media [21], polyatomic gas flows [22, 23], and unsteady multiscale flows [24]. Moreover, it has been extended to accelerate the convergence of low-variance DSMC [25]. Rigorous mathematical analysis has shown that the GSIS is able to asymptotically preserve the Navier-Stokes limit when the cell size is much larger than the molecular mean free path, and facilitate rapid convergence to the steady state [26]. The key to the success of GSIS lies in its incorporation of the mesoscopic kinetic equation alongside the macroscopic synthetic equation. The stress and heat flux in the macroscopic synthetic equations are expressed as combinations of the Newton law of viscosity and the Fourier law of heat conduction, as well as the high-order constitutive relations extracted from the kinetic equation. Providing higher-order corrections to the synthetic equations ensures the iterations to converge to accurate results, particularly in strong rarefied flows. The explicit inclusion of the Newton and Fourier constitutive relations enable the macroscopic equations to efficiently exchange the flow information in the near-continuum regime, thereby accelerating convergence, see the Fourier flow in Ref. [20].

However, the coupling acceleration between mesoscopic and macroscopic treatments is applied only in the interior computational domain, e.g., it has been pointed out that “the zeroth-order moment is not accelerated at the boundaries and its estimate is based on the values of the distribution function at the boundaries obtained at the first stage of the iteration” [17]. This does not affect the speed of convergence in the Poiseuille flow since the distribution function reflected from the solid wall is always zero, so that the numerical flux is fixed. For general nonlinear flows, this is not the case, and since the synthetic equation plays an important role in guiding the gas system towards the steady state, its boundary condition (flux) is critical in accelerating the convergence. In the first work to extend the GSIS to nonlinear flows, the numerical fluxes for macroscopic equations at the boundaries are directly calculated from the DVM, which remains unchanged when solve the macroscopic synthetic equations [27]. In the second work of nonlinear GSIS, an incremental feedback of macroscopic variables from the cell centers to wall is used to update the boundary flux [24]. Although both methods are faster than the CIS in near-continuum flow regimes, converged solutions for hypersonic flows are found after hundreds of iterations, which is about 10 times slower than the GSIS for linear flows where the steady-state solutions can be found within dozens of iterations [20]. Thus, for general nonlinear flows, the boundary flux needs to be properly treated, to further reduce the iteration steps.

The boundary conditions are of paramount importance in numerical simulations, and this importance is significant in rarefied gas flows. Many critical phenomena in rarefied flows, such as the velocity slip, temperature jump, and thermal creep [28, 29, 30, 31, 32, 33], appear within a few mean free path away from the wall. The conventional no-slip wall boundary conditions of macroscopic equations are physically inconsistent with these rarefied phenomena, and their direct application can lead to unstable and inaccurate results in macroscopic equations [34]. Fortunately, the macroscopic boundary conditions are solved in the moment equations for slightly-rarefied gas. For instance, Grad proposed the wall boundary condition for the 13-moment equations by integrating the truncated distribution function [35], which is further extended to the regularized 13- and 26-moment equations by Gu et al. [36, 37]. Studies demonstrated that the velocity slip can be automatically obtained by these boundary conditions in the moment equations.

Inspired by the boundary condition in the moment methods [35, 36, 37], we shall propose in the present work a generalized boundary treatment (GBT) to further accelerate the convergence of GSIS. The rest of the paper is organized as follows. In section 2, the mesoscopic kinetic equation and macroscopic synthetic equation are introduced, together with their numerical schemes in the finite-volume framework. In section 3, the GBT is proposed in great details. In section 4, the accuracy and efficiency of GSIS-GBT is assessed in planar Fourier flow, hypersonic flow past a cylinder, pressure-driven flow through three circular pipes, and the multiscale flow around a hypersonic technology vehicle. Finally, conclusions are given in section 5.

2 Methodology

In this section, the mesoscopic kinetic equation for the polyatomic gas with translational and rotational degrees of freedom is firstly introduced, as well as the finite-volume scheme; then, the macroscopic synthetic equation is derived from the kinetic equation, which is also solved by the finite-volume method. Finally, the flowchart of GSIS is outlined.

2.1 The kinetic equation

To simultaneously account for real gas effects while avoiding complex kinetic models, we consider the kinetic model of polyatomic gas with rotational degrees of freedom drd_{r}, excluding vibrational and electronic degrees of freedom. The gas system can be described by the distribution function fT(t,𝒙,𝝃,Ir)f^{T}(t,\bm{x},\bm{\xi},I_{r}), where the variable 𝐱\mathbf{x} denotes spatial position, tt represents time, 𝝃\bm{\xi} represents the molecular velocity, and IrI_{r} is the rotational energy. Two velocity distribution functions (VDFs), f0(𝐱,𝝃,t)=fT𝑑Irf_{0}(\mathbf{x},\bm{\xi},t)=\int f^{T}dI_{r} and f1(𝐱,𝝃,t)=IrfT𝑑Irf_{1}(\mathbf{x},\bm{\xi},t)=\int I_{r}f^{T}dI_{r}, can be introduced to reduce the rotational energy variable, and the final dimensionless kinetic model in the absence of external force can be expressed as [38, 39]

f0t+𝝃f0𝐱=g0tf0τ+g0rg0tZrτ,\displaystyle\frac{\partial f_{0}}{\partial t}+\bm{\xi}\cdot\frac{\partial f_{0}}{\partial\mathbf{x}}=\frac{g_{0t}-f_{0}}{\tau}+\frac{g_{0r}-g_{0t}}{Z_{r}\tau}, (1)
f1t+𝝃f1𝐱=g1tf1τ+g1rg1tZrτ,\displaystyle\frac{\partial f_{1}}{\partial t}+\bm{\xi}\cdot\frac{\partial f_{1}}{\partial\mathbf{x}}=\frac{g_{1t}-f_{1}}{\tau}+\frac{g_{1r}-g_{1t}}{Z_{r}\tau},

where τ\tau and ZrτZ_{r}\tau represent the elastic (where the kinetic energy is conserved) and inelastic (energy exchange between translational and rotational motions) collision times of gas molecules, with ZrZ_{r} being the rotational collision number and the mean collision time is defined as

τ=μpπ2Kn,{\tau=\frac{\mu}{p}\sqrt{\frac{\pi}{2}}\text{Kn},} (2)

where μ\mu is the normalized shear viscosity: if the power-law potential is used, we have

μ=Ttω,\mu=T_{t}^{\omega}, (3)

with ω\omega being the viscosity index.

The function gg denotes the reference distribution function, where the subscripts tt and rr stand for the relaxation processes associated with translational and rotational motions, respectively:

g0t=f0eq[1+2𝐪t𝐜15Ttpt(c22Tt52)],g0r=f1eq[1+2𝐪0𝐜15Tp(c22T52)],g1t=dr2Trg0t+f0eq𝐪r𝐜ρTt,g1r=dr2Tg0r+f1eq𝐪1𝐜ρT,\begin{gathered}g_{0t}=f_{0}^{eq}\left[1+\frac{2\mathbf{q}_{t}\cdot\mathbf{c}}{15T_{t}p_{t}}\left(\frac{c^{2}}{2T_{t}}-\frac{5}{2}\right)\right],\quad g_{0r}=f_{1}^{eq}\left[1+\frac{2\mathbf{q}_{0}\cdot\mathbf{c}}{15{Tp}}\left(\frac{c^{2}}{2T}-\frac{5}{2}\right)\right],\\ g_{1t}=\frac{d_{r}}{2}T_{r}g_{0t}+f_{0}^{eq}\frac{\mathbf{q}_{r}\cdot\mathbf{c}}{\rho T_{t}},\quad g_{1r}=\frac{d_{r}}{2}Tg_{0r}+f_{1}^{eq}\frac{\mathbf{q}_{1}\cdot\mathbf{c}}{\rho T},\end{gathered} (4)

with the equilibrium distribution functions

f0eq=ρ(12πTt)3/2exp(c22Tt),f1eq=ρ(12πT)3/2exp(c22T).\displaystyle f_{0}^{eq}=\rho\left(\frac{1}{2\pi T_{t}}\right)^{3/2}\exp\left(-\frac{c^{2}}{2T_{t}}\right),\quad f_{1}^{eq}=\rho\left(\frac{1}{2\pi T}\right)^{3/2}\exp\left(-\frac{c^{2}}{2T}\right). (5)

It should be noted that, from Eq. (1) we known that the VDFs f0f_{0} and f1f_{1} evolve towards the reference VDFs in Eq. (4). Meanwhile, during the evolution of f0f_{0} and f1f_{1}, the heat fluxes in the reference VDFs gradually go to zero in isolated system. Therefore, in the final steady state, the VDFs f0f_{0} and f1f_{1} will be described by the corresponding equilibrium distribution functions in Eq. (5), which is consistent with the Boltzmann H-theorem.

Macroscopic quantities including the density ρ\rho, velocity 𝐮\mathbf{u}, translational temperature TtT_{t}, rotational temperature TrT_{r}, as well as the stress 𝝈\bm{\sigma}, translational heat flux 𝐪t\mathbf{q}_{t} and rotational heat flux 𝐪r\mathbf{q}_{r} are expressed through the moment integration of the distribution functions as

ρ=f0d𝝃,ρ𝐮=𝝃f0d𝝃,32ρTt=c22f0d𝝃,𝝈=(𝐜𝐜c23I)f0d𝝃,dr2ρTr=f1d𝝃,𝐪t=𝐜c22f0d𝝃,𝐪𝐫=𝐜f1d𝝃.\begin{gathered}\rho=\int f_{0}\mathrm{~{}d}\bm{\xi},\quad\rho\mathbf{u}=\int\bm{\xi}f_{0}\mathrm{~{}d}\bm{\xi},\\ \frac{3}{2}\rho T_{t}=\int\frac{c^{2}}{2}f_{0}\mathrm{~{}d}\bm{\xi},\quad\bm{\sigma}=\int\left(\mathbf{c}\mathbf{c}-\frac{c^{2}}{3}\mathrm{I}\right)f_{0}\mathrm{~{}d}\bm{\xi},\quad\frac{d_{r}}{2}\rho T_{r}=\int f_{1}\mathrm{~{}d}\bm{\xi},\\ \mathbf{q}_{t}=\int\mathbf{c}\frac{c^{2}}{2}f_{0}\mathrm{~{}d}\bm{\xi},\quad\mathbf{q}_{\mathbf{r}}=\int\mathbf{c}f_{1}\mathrm{~{}d}\bm{\xi}.\end{gathered} (6)

Furthermore, relationships pertaining to the total temperature TT, translational pressure ptp_{t}, and total pressure pp are delineated as follows:

T=3Tt+drTr3+dr,pt=ρTt,p=ρT,\displaystyle T=\frac{3T_{t}+d_{r}T_{r}}{3+d_{r}},\quad p_{t}=\rho T_{t},\quad p=\rho T, (7)

where 𝐜=𝝃𝐮\mathbf{c}=\bm{\xi}-\mathbf{u} denotes the peculiar velocity. Here the relaxation rates of the heat flux 𝑨=[Att,Atr,Art,Arr]\bm{A}=\left[A_{tt},A_{tr},A_{rt},A_{rr}\right] are derived from the DSMC simulation to determine 𝐪0\mathbf{q}_{0} and 𝐪1\mathbf{q}_{1} as [39]:

[𝐪0𝐪1]=[(23Att)Zr+13AtrZrArtZrArrZr+1][𝐪t𝐪r].\left[\begin{array}[]{l}\mathbf{q}_{0}\\ \mathbf{q}_{1}\end{array}\right]=\left[\begin{array}[]{cc}\left(2-3A_{tt}\right)Z_{r}+1&-3A_{tr}Z_{r}\\ -A_{rt}Z_{r}&-A_{rr}Z_{r}+1\end{array}\right]\left[\begin{array}[]{l}\mathbf{q}_{t}\\ \mathbf{q}_{r}\end{array}\right]. (8)

It is noted that the derivations mentioned here are based on dimensionless kinetic equations. The dimensionless treatment of macroscopic variables and distribution functions is presented in A for reference.

2.2 The conventional iterative scheme

We briefly introduce the CIS to solve the kinetic equation by the implicit finite-volume method. Considering that the evolution equations for translational and rotational distribution functions in Eq. (1) exhibit identical structural formulations, here we employ the notations ff and gg to represent a unified form of translational/rotational distribution functions. The discrete form of the mesoscopic kinetic equation within the control volume cell and a time step Δtf=tn+1tn\Delta t_{f}=t^{n+1}-t^{n} from the nn-th step to the (n+1)(n+1)-th step is shown as follows

fin+1finΔtf+1ΩijN(i)ξnfijn+1Sij=gt,infin+1τin+gr,ingt,inZrτin,\frac{f_{i}^{n+1}-f_{i}^{n}}{\Delta t_{f}}+\frac{1}{\Omega_{i}}\sum_{j\in N(i)}{\xi}_{n}f_{ij}^{n+1}S_{ij}=\frac{g_{t,i}^{n}-f_{i}^{n+1}}{\tau_{i}^{n}}+\frac{g_{r,i}^{n}-g_{t,i}^{n}}{Z_{r}\tau_{i}^{n}}, (9)

where the subscript ii denotes the average variables at the cell center, while the subscript ijij represents the cell interface located between cells ii and jj. SijS_{ij} and Ωi\Omega_{i} respectively denote the area of the cell interface and the volume of the cell. ξn=𝝃n𝐧ij{\xi}_{n}=\bm{{\xi}}_{n}\cdot\mathbf{n}_{ij} represents the molecular velocity component along the interface normal vector 𝐧ij\mathbf{n}_{ij}.

To facilitate matrix-free implicit iterations, an incremental form of the distribution function is introduced:

Δfin=fin+1fin\Delta f_{i}^{n}=f_{i}^{n+1}-f_{i}^{n} (10)

which allows the rewriting of Eq. (9) as

(1Δtf+1τin)Δfin+1ΩijN(i)ξnΔfijnSij=ginfinτin1ΩijN(i)ξnfijnSijRHSi.\left(\frac{1}{\Delta t_{f}}+\frac{1}{\tau_{i}^{n}}\right)\Delta f_{i}^{n}+\frac{1}{\Omega_{i}}\sum_{j\in N(i)}\xi_{n}\Delta f_{ij}^{n}S_{ij}=\underbrace{\frac{g_{i}^{n}-f_{i}^{n}}{\tau_{i}^{n}}-\frac{1}{\Omega_{i}}\sum_{j\in N(i)}\xi_{n}f_{ij}^{n}S_{ij}}_{\mathrm{RHS}_{i}}. (11)

where RHSi\mathrm{RHS}_{i} is the mesoscopic residual. The mesoscopic time step Δtf\Delta t_{f} can be determined based on the Courant–Friedrichs–Lewy (CFL) condition as [40]:

Δtf=ϑfminiΩi(|ui,x|+3Ti)Si,x+(|ui,y|+3Ti)Si,y+(|uz|+3Ti)Si,z,\Delta t_{f}=\vartheta_{f}\min_{i}\frac{\Omega_{i}}{\left(\left|u_{i,x}\right|+3\sqrt{T_{i}}\right)S_{i,x}+\left(\left|u_{i,y}\right|+3\sqrt{T_{i}}\right)S_{i,y}+\left(\left|u_{z}\right|+3\sqrt{T_{i}}\right)S_{i,z}}, (12)

in which SiS_{i} is the projections of the control volume on the corresponding direction and ϑf\vartheta_{f} is the mesoscopic CFL number.

The CIS is found to be efficient when the Knudsen number is large, but is highly dissipative and inefficient when the Knudsen number is small [14]. To expedite the convergence of the kinetic equations, the VDF is firstly evaluated at the intermediate step, i.e., n+1n+1 in Eqs. (9) and (10) is replaced by n+1/2n+1/2. Subsequently, macroscopic synthetic equations are introduced in GSIS which will guide the evolution of the VDF towards steady-state solutions.

2.3 The macroscopic synthetic equation in GSIS

The proper design of macroscopic synthetic equation is crucial in GSIS, which the guidelines are: (i) the synthetic equation should be derived exactly from the kinetic equation, (ii) the synthetic equation should facilitate efficient exchange of flow information in the whole computational domain, and (iii) the synthetic equation should be as simple as possible, so that it is compatible with the numerical techniques in the computational fluid dynamics. In this case, the second version of GSIS (which contains the evolution equations for the density, velocity, and temperature only) is used [27, 24], rather than the first version (which not only contains the evolution equations for the density, velocity, and temperature, but also the evolution equations for the stress and heat flux) [20].

Exploiting the mesoscopic-macroscopic relationships in Eq. (6) and applying them to Eq. (1), the governing equations for macroscopic quantities can be derived as

ρt+(ρ𝐮)=0,\displaystyle\frac{\partial\rho}{\partial t}+\nabla\cdot(\rho\mathbf{u})=0, (13)
t(ρ𝐮)+(ρ𝐮𝐮)+(ρTtI+𝝈)=0,\displaystyle\frac{\partial}{\partial t}(\rho\mathbf{u})+\nabla\cdot(\rho\mathbf{u}\mathbf{u})+\nabla\cdot\left(\rho T_{t}\mathrm{I}+\bm{\sigma}\right)=0,
t(ρE)+(ρE𝐮)+(ρTt𝐮+𝝈𝐮+𝐪t+𝐪r)=0,\displaystyle\frac{\partial}{\partial t}(\rho E)+\nabla\cdot(\rho E\mathbf{u})+\nabla\cdot\left(\rho T_{t}\mathbf{u}+\bm{\sigma}\cdot\mathbf{u}+\mathbf{q}_{t}+\mathbf{q}_{r}\right)=0,
t(ρEr)+(ρEr𝐮)+𝐪r=drρ2TTrZτ,\displaystyle\frac{\partial}{\partial t}\left(\rho E_{r}\right)+\nabla\cdot\left(\rho E_{r}\mathbf{u}\right)+\nabla\cdot\mathbf{q}_{r}=\frac{d_{r}\rho}{2}\frac{T-T_{r}}{Z\tau},

where E=(3Tt+drTr)/2+u2/2E=\left(3T_{t}+d_{r}T_{r}\right)/2+u^{2}/2 and Er=drTr/2E_{r}=d_{r}T_{r}/2 are the total and rotational energies, respectively.

When the Knudsen number is small, the stress and heat flux in Eq. (13) can be approximated via the first-order Chapman-Enskog expansion of the kinetic equation as [41]:

𝝈\displaystyle\bm{\sigma}\approx 𝝈M=μ(𝐮+𝐮T23𝐮𝐈),\displaystyle\bm{\sigma}^{\mathrm{M}}=-\mu\left(\nabla\mathbf{u}+\nabla\mathbf{u}^{\mathrm{T}}-\frac{2}{3}\nabla\cdot\mathbf{uI}\right), (14)
𝐪t\displaystyle\mathbf{q}_{t}\approx 𝐪tM=κtTt,𝐪r𝐪rM=κrTr,\displaystyle\mathbf{q}_{t}^{\mathrm{M}}=-\kappa_{t}\nabla T_{t},\quad\mathbf{q}_{r}\approx\mathbf{q}_{r}^{\mathrm{M}}=-\kappa_{r}\nabla T_{r},

where κt\kappa_{t} and κr\kappa_{r} are the translational and rotational thermal conductivities:

[κtκr]=ptτ2[AttAtrArtArr]1[5dr].\left[\begin{array}[]{l}\kappa_{t}\\ \kappa_{r}\end{array}\right]=\frac{p_{t}\tau}{2}\left[\begin{array}[]{ll}A_{tt}&A_{tr}\\ A_{rt}&A_{rr}\end{array}\right]^{-1}\left[\begin{array}[]{c}5\\ d_{r}\end{array}\right]. (15)

The macroscopic synthetic equation is constructed as follows:

ρt+(ρ𝐮)=0,\displaystyle\frac{\partial\rho}{\partial t}+\nabla\cdot(\rho\mathbf{u})=0, (16)
t(ρ𝐮)+(ρ𝐮𝐮)+(ρTtI+𝝈M)=(Δ𝝈),\displaystyle\frac{\partial}{\partial t}(\rho\mathbf{u})+\nabla\cdot(\rho\mathbf{u}\mathbf{u})+\nabla\cdot\left(\rho T_{t}\mathrm{I}+\bm{\sigma}^{\mathrm{M}}\right)=-\nabla\cdot(\Delta\bm{\sigma}),
t(ρE)+(ρE𝐮)+(ρTt𝐮+𝝈M𝐮+𝐪tM+𝐪rM)=(Δ𝝈𝐮+Δ𝐪t+Δ𝐪r),\displaystyle\frac{\partial}{\partial t}(\rho E)+\nabla\cdot(\rho E\mathbf{u})+\nabla\cdot\left(\rho T_{t}\mathbf{u}+\bm{\sigma}^{\mathrm{M}}\cdot\mathbf{u}+\mathbf{q}^{\mathrm{M}}_{t}+\mathbf{q}^{\mathrm{M}}_{r}\right)=-\nabla\cdot(\Delta\bm{\sigma}\cdot\mathbf{u}+\Delta\mathbf{q}_{t}+\Delta\mathbf{q}_{r}),
t(ρEr)+(ρEr𝐮)+𝐪rMdrρ2TTrZτ=Δ𝐪r,\displaystyle\frac{\partial}{\partial t}\left(\rho E_{r}\right)+\nabla\cdot\left(\rho E_{r}\mathbf{u}\right)+\nabla\cdot\mathbf{q}^{\mathrm{M}}_{r}-\frac{d_{r}\rho}{2}\frac{T-T_{r}}{Z\tau}=-\nabla\cdot\Delta\mathbf{q}_{r},

where

Δ𝝈=𝝈n+1/2𝝈M,n+1/2=(𝐜c23I)f0n+1/2dξ𝝈M,n+1/2,Δ𝐪t=𝐪tn+1/2𝐪tM,n+1/2=𝐜c22f0n+1/2dξ𝐪tM,n+1/2,Δ𝐪r=𝐪rn+1/2𝐪rM,n+1/2=𝐜f1n+1/2dξ𝐪rM,n+1/2.\begin{gathered}\Delta\bm{\sigma}=\bm{\sigma}^{n+1/2}-\bm{\sigma}^{\mathrm{M},n+1/2}=\int\left(\mathbf{c}-\frac{c^{2}}{3}\mathrm{I}\right)f_{0}^{n+1/2}\mathrm{~{}d}\xi-\bm{\sigma}^{\mathrm{M},n+1/2},\\ \Delta\mathbf{q}_{t}=\mathbf{q}_{t}^{n+1/2}-\mathbf{q}_{t}^{\mathrm{M},n+1/2}=\int\mathbf{c}\frac{c^{2}}{2}f_{0}^{n+1/2}\mathrm{~{}d}\xi-\mathbf{q}_{t}^{\mathrm{M},n+1/2},\\ \Delta\mathbf{q}_{r}=\mathbf{q}_{r}^{n+1/2}-\mathbf{q}_{r}^{\mathrm{M},n+1/2}=\int\mathbf{c}f_{1}^{n+1/2}\mathrm{~{}d}\xi-\mathbf{q}_{r}^{\mathrm{M},n+1/2}.\end{gathered} (17)

It should be emphasized that, in the numerical simulation, the variables at the left-hand-side of Eq. (16) are evaluated at the (n+1)(n+1)-th time step, while the one at the right-hand-side are evaluated at the (n+1/2)(n+1/2)-th time step. Only when the solution finally converges (the solution at the (n+1/2)(n+1/2)-th time step is the same as that at the (n+1)(n+1)-th time step) can the synthetic equation turns back to Eq. (13) which is exactly derived from the kinetic equation. In the intermediate iteration step, Eq. (16) is not the same as the exact one (13), but such treatment plays an important role in boosting the convergence to the steady state.

Compared to the original macroscopic equation (13), the synthetic equation introduces high-order corrections in the stress and heat flux Δ𝝈,Δ𝐪t,Δ𝐪r\Delta\bm{\sigma},\Delta\mathbf{q}_{t},\Delta\mathbf{q}_{r} on the right-hand-side of equations, ensuring the accuracy of macroscopic equation iterations in the presence of strong rarefaction effects. On the other hand, the explicit inclusion of the Newton law of viscosity and the Fourier law of thermal conductivity in the left-hand-side of Eq. (16) could lead to fast exchange of the flow field, since in the steady state the diffusion-type equations for the velocity and temperature can be obtained, where the information propagation speed is infinite.

2.4 Finite-volume method for the synthetic equations

Under the finite-volume discretization, the macroscopic synthetic equation within the control volume cell ii and a inner time step ΔtW=tm+1tm\Delta t_{W}=t^{m+1}-t^{m} can be formulated as follows:

𝐖im+1𝐖imΔtW+SijΩijN(i)[𝐅c,ijm+1+𝐅v,ijm+1(𝝈M,𝐪M)]=𝐐im+1SijΩijN(i)𝐅v,ijHoT(Δ𝝈,Δ𝐪),\frac{\mathbf{W}_{i}^{m+1}-\mathbf{W}_{i}^{m}}{\Delta t_{W}}+\frac{S_{ij}}{\Omega_{i}}\sum_{j\in N(i)}\left[\mathbf{F}_{c,ij}^{m+1}+\mathbf{F}_{v,ij}^{m+1}\left(\bm{\sigma}^{\mathrm{M}},\mathbf{q}^{\mathrm{M}}\right)\right]=\mathbf{Q}_{i}^{m+1}-\frac{S_{ij}}{\Omega_{i}}\sum_{j\in N(i)}\mathbf{F}_{v,ij}^{\mathrm{HoT}}(\Delta\bm{\sigma},\Delta\mathbf{q}), (18)

Here, the superscript mm corresponds to the iteration step of the macroscopic equations, distinguishing it from the iteration step of the mesoscopic kinetic equation; N(i)N(i) denotes the set of neighbouring cells. The macroscopic time step ΔtW\Delta t_{W} for the inner iteration can be determined based on the newly macroscopic variables 𝐖im\mathbf{W}_{i}^{m} in conjunction with Eq. (12) by replacing the mesoscopic CFL number ϑf\vartheta_{f} with macroscopic CFL number ϑW\vartheta_{W}. Vectors related to the macroscopic quantities 𝐖\mathbf{W}, the convective numerical flux 𝐅c\mathbf{F}_{c}, and the source term 𝐐\mathbf{Q} are defined as

𝐖=[ρρuxρuyρuzρEρEr],𝐅c=[ρunρuxun+nxptρuyun+nyptρuzun+nzptun(ρE+pt)unρEr],𝐐=[00000dr2ρTTrZrτ],\mathbf{W}=\left[\begin{array}[]{c}\rho\\ \rho u_{x}\\ \rho u_{y}\\ \rho u_{z}\\ \rho E\\ \rho E_{r}\end{array}\right],\quad\mathbf{F}_{c}=\left[\begin{array}[]{c}\rho u_{n}\\ \rho u_{x}u_{n}+n_{x}p_{t}\\ \rho u_{y}u_{n}+n_{y}p_{t}\\ \rho u_{z}u_{n}+n_{z}p_{t}\\ u_{n}\left(\rho E+p_{t}\right)\\ u_{n}\rho E_{r}\end{array}\right],\quad\mathbf{Q}=\left[\begin{array}[]{c}0\\ 0\\ 0\\ 0\\ 0\\ \frac{d_{r}}{2}\rho\frac{T-T_{r}}{Z_{r}\tau}\end{array}\right], (19)

where un=𝐮𝐧iju_{n}=\mathbf{u}\cdot\mathbf{n}_{ij} is the velocity along the normal direction of the interface. The viscous numerical flux 𝐅v(𝝈,𝐪)\mathbf{F}_{v}(\bm{\sigma},\mathbf{q}) is given by

𝐅v(𝝈,𝐪)=[0nxσxx+nyσxy+nzσxznxσyx+nyσyy+nzσyznxσzx+nyσzy+nzσzznxΘx+nyΘy+nzΘznxqx,r+nyqy,r+nzqz,r],\mathbf{F}_{v}(\bm{\sigma},\mathbf{q})=\left[\begin{array}[]{c}0\\ n_{x}\sigma_{xx}+n_{y}\sigma_{xy}+n_{z}\sigma_{xz}\\ n_{x}\sigma_{yx}+n_{y}\sigma_{yy}+n_{z}\sigma_{yz}\\ n_{x}\sigma_{zx}+n_{y}\sigma_{zy}+n_{z}\sigma_{zz}\\ n_{x}\Theta_{x}+n_{y}\Theta_{y}+n_{z}\Theta_{z}\\ n_{x}q_{x,r}+n_{y}q_{y,r}+n_{z}q_{z,r}\end{array}\right], (20)

where Θx(𝝈,𝐪)=uxσxx+uyσxy+uzσxz+qx,t+qr,t\Theta_{x}(\bm{\sigma},\mathbf{q})=u_{x}\sigma_{xx}+u_{y}\sigma_{xy}+u_{z}\sigma_{xz}+q_{x,t}+q_{r,t}, Θy(𝝈,𝐪)=uxσyx+uyσyy+uzσyz+qy,t+qy,r\Theta_{y}(\bm{\sigma},\mathbf{q})=u_{x}\sigma_{yx}+u_{y}\sigma_{yy}+u_{z}\sigma_{yz}+q_{y,t}+q_{y,r} and Θz(𝝈,𝐪)=uxσzx+uyσzy+uzσzz+qz,t+qz,r\Theta_{z}(\bm{\sigma},\mathbf{q})=u_{x}\sigma_{zx}+u_{y}\sigma_{zy}+u_{z}\sigma_{zz}+q_{z,t}+q_{z,r}.

Similarly, an incremental form of macroscopic quantities is introduced as Δ𝑾im=𝑾im+1𝑾im\Delta\bm{W}_{i}^{m}=\bm{W}_{i}^{m+1}-\bm{W}_{i}^{m}, and Eq. (18) can be reformulated as follows:

[1ΔtW(𝐐𝐖)m]Δ𝐖im+SijΩijN(i)Δ𝐅ijm+1=𝐑im+𝐑iHoT,𝐑im=SijΩijN(i)[𝐅c,ijm+1+𝐅v,ijm+1(𝝈m,𝐪tm,𝐪rm)]+𝐐im,𝐑iHoT=SijΩijN(i)𝐅v,ijHoT(Δ𝝈n+1/2,Δ𝐪tn+1/2,Δ𝐪rn+1/2).\begin{gathered}{\left[\frac{1}{\Delta t_{W}}-\left(\frac{\partial\mathbf{Q}}{\partial\mathbf{W}}\right)^{m}\right]\Delta\mathbf{W}_{i}^{m}+\frac{S_{ij}}{\Omega_{i}}\sum_{j\in N(i)}\Delta\mathbf{F}_{ij}^{m+1}=\mathbf{R}_{i}^{m}+\mathbf{R}_{i}^{\mathrm{HoT}},}\\ \mathbf{R}_{i}^{m}=-\frac{S_{ij}}{\Omega_{i}}\sum_{j\in N(i)}\left[\mathbf{F}_{c,ij}^{m+1}+\mathbf{F}_{v,ij}^{m+1}(\bm{\sigma}^{m},\mathbf{q}_{t}^{m},\mathbf{q}_{r}^{m})\right]+\mathbf{Q}_{i}^{m},\\ \mathbf{R}_{i}^{\mathrm{HoT}}=-\frac{S_{ij}}{\Omega_{i}}\sum_{j\in N(i)}\mathbf{F}_{v,ij}^{\mathrm{HoT}}(\Delta\bm{\sigma}^{{n+1/2}},\Delta\mathbf{q}_{t}^{{n+1/2}},\Delta\mathbf{q}_{r}^{{n+1/2}}).\end{gathered} (21)

Here, 𝐑im\mathbf{R}_{i}^{m} and 𝐑iHoT\mathbf{R}_{i}^{\mathrm{HoT}} represent the macroscopic residual and high-order residual, respectively. In order to incorporate the high-order corrections including Δ𝝈\Delta\bm{\sigma}, Δ𝐪t\Delta\mathbf{q}_{t}, Δ𝐪r\Delta\mathbf{q}_{r} into the macroscopic iterations, the corresponding viscous flux 𝐅v,ijHoT\mathbf{F}_{v,ij}^{\mathrm{HoT}} is computed following the same form as 𝐅v\mathbf{F}_{v} in Eq. (20), but the stress and heat flux therein need to be replaced by higher-order corrections Δ𝝈,Δ𝐪t,Δ𝐪r\Delta\bm{\sigma},\Delta\mathbf{q}_{t},\Delta\mathbf{q}_{r}.

It is worth noting that the high-order corrections in Eq. (17) are computed based on the VDF and macroscopic quantities at the time step n+1/2n+1/2, and they remain constant during inner iterations. Therefore, the high-order residual terms 𝐑iHoT\mathbf{R}_{i}^{\mathrm{HoT}} in Eq. (21) should be treated as source terms in the computation.

The Rusanov scheme and the central-difference scheme are employed to compute the convective flux 𝐅c,ij\mathbf{F}_{c,ij} and viscous flux 𝐅v,ij\mathbf{F}_{v,ij} in the residual 𝐑im\mathbf{R}_{i}^{m} of Eq. (21), respectively. Meanwhile, the incremental form of the macroscopic flux Δ𝐅ij\Delta\mathbf{F}_{ij} can be approximated by the Eulerian form of the convective flux. Equations (11) and (21) can be iteratively solved using the classical Lower Upper Symmetric Gauss–Seidel (LU-SGS) technology. Given the widespread application of LU-SGS, specific details are omitted here for the sake of brevity. Relevant literature specifying the settings for implicitly solving these equations using LU-SGS are provided in the reference [27, 24].

2.5 The flowchart of GSIS

The framework of the GSIS algorithm, aimed at accelerating convergence by alternately solving mesoscopic kinetic equation and macroscopic synthetic equation within one iteration step, is summarized below:

  • 1)

    Based on the macroscopic quantities 𝐖in\mathbf{W}_{i}^{n} and mesoscopic CFL number ϑf\vartheta_{f}, the time step Δtf\Delta t_{f} for the mesoscopic kinetic equations can be set according to Eq. (12).

  • 2)

    The implicit solution of the mesoscopic kinetic equations evolves the VDF to fin+1/2f_{i}^{n+1/2} according to Eq. (11). The kinetic gas-surface boundary condition is used.

  • 3)

    The corresponding macroscopic quantities 𝐖in+1/2\mathbf{W}_{i}^{n+1/2}, as well as the stress 𝝈n+1/2\bm{\sigma}^{n+1/2} and the heat flux 𝐪tn+1/2,𝐪rn+1/2\mathbf{q}_{t}^{n+1/2},\mathbf{q}_{r}^{n+1/2}, are obtained through moment integration of the distribution function as presented in Eq. (6).

  • 4)

    Calculate the derivatives of the macroscopic quantities by the least square method. Then, the high-order correction including Δ𝝈n+1/2,Δ𝐪n+1/2,Δ𝐪n+1/2\Delta\bm{\sigma}^{n+1/2},\Delta\mathbf{q}^{n+1/2},\Delta\mathbf{q}^{n+1/2} can be obtained from Eq. (17).

  • 5)

    The LU-SGS technology is employed for the implicit iteration of the macroscopic synthetic equations (21). In contrast to the kinetic equations, which are computed only once in each step, the synthetic equations involve hundreds of inner iterations to expedite the flow information exchange. Note that the traditional no-velocity-slip and no-temperature boundary conditions do not work any more in rarefied gas flows [29], therefore the boundary condition should be carefully designed; this will be elaborated in section 3.

  • 6)

    Based on the macroscopic quantities 𝐖ik\mathbf{W}_{i}^{k} from the inner iterations (kk is the total steps of inner iterations), the next-step VDF fin+1f_{i}^{n+1} are computed as

    f0,in+1=f0,in+1/2+[f0eq(𝐖ik)f0eq(𝐖in+1/2)],\displaystyle f_{0,i}^{n+1}=f_{0,i}^{n+1/2}+\left[f_{0}^{\mathrm{eq}}\left(\mathbf{W}_{i}^{k}\right)-f_{0}^{\mathrm{eq}}\left(\mathbf{W}_{i}^{n+1/2}\right)\right], (22)
    f1,in+1=f1,in+1/2+[f1eq(𝐖ik)f1eq(𝐖in+1/2)].\displaystyle f_{1,i}^{n+1}=f_{1,i}^{n+1/2}+\left[f_{1}^{\mathrm{eq}}\left(\mathbf{W}_{i}^{k}\right)-f_{1}^{\mathrm{eq}}\left(\mathbf{W}_{i}^{n+1/2}\right)\right].

    Then the macroscopic quantities 𝐖in+1\mathbf{W}_{i}^{n+1} are consistent with those obtained through moment integration of fin+1f_{i}^{n+1} in Eq. (6).

  • 7)

    Repeat the steps (1-7) until the following convergence criterion is satisfied:

    i(ψin+1ψin)2Ωii(ψin)2Ωi|max<106, for ψ{ρ,𝐮,T,Tr}.\left.\frac{\sqrt{\sum_{i}\left(\psi_{i}^{n+1}-\psi_{i}^{n}\right)^{2}\Omega_{i}}}{\sqrt{\sum_{i}\left(\psi_{i}^{n}\right)^{2}\Omega_{i}}}\right|_{\max}<10^{-6},\quad\text{ for }\quad\psi\in\left\{\rho,\mathbf{u},T,T_{r}\right\}. (23)

It is noteworthy that, in the simulation of hypersonic and complex three-dimensional flows, it is a common practice to perform thousands steps of macroscopic iterations by first-order Euler equations to initialize the flow field [42, 43] and 10 steps of CIS iterations to initialize the VDF [44] prior to GSIS iterations.

3 Generalized boundary treatment

3.1 General considerations

The fast convergence of GSIS lies in the use of macroscopic synthetic equation (16) to guide the evolution of gas systems towards the final steady state [20, 26, 22]. Since the deterministic mesoscopic solver is much more time-consuming to solve than the macroscopic synthetic equation, in the outer iteration, Eq. (11) is only solved once at each iteration, while in the inner iteration, Eq. (16) is solved multiple times or till convergence. Since the macroscopic synthetic equation can be solved quickly and for many steps, the flow field has the time to exchange information globally, and hence faster convergence to the steady state is achieved in GSIS than solving the kinetic equation solely.

However, the macroscopic synthetic equation needs boundary conditions, but in general rarefied gas flows the no-velocity-slip and no-temperature-jump boundary conditions are not valid anymore. But, since the kinetic solver is already slow to converge when the Knudsen number is small, directly feeding its boundary information into macroscopic synthetic equation is not very efficient [27]. In order to further accelerate the GSIS, we need to figure out a way to provide the boundary condition to the macroscopic synthetic equation at each inner iteration [24], and this boundary condition should be compatible with that in the kinetic equation.

In the moment methods [35, 36, 37], the boundary conditions are introduced in three steps. First, the VDF is re-constructed from conserved macroscopic quantities and high-order moments in the moment equation. Second, the reflected VDF from the wall is calculated by applying the kinetic gas-surface boundary condition. Third, the macroscopic boundary conditions are explicitly derived for the wall VDF. The similar idea can be borrowed to GSIS. To be specific, the truncated distribution function applicable to polyatomic gases with rotational degrees of freedom is derived to approximate the unknown distribution function at the wall for an explicit integration. Then, the numerical flux can be directly reconstructed recursively based on the conserved macroscopic quantities, as well as the stress and heat flux at the interface. The stress and heat flux embedded in the truncated distribution function can be closed based on explicit constitutive relations of the Newton and Fourier. Additionally, the concept of high-order correction to the constitutive equations, which is employed in GSIS, is also incorporated into this macroscopic boundary condition, ensuring stability and accuracy in hypersonic flows and strong rarefied flows.

3.2 Overview of boundary conditions in previous versions of GSIS

Within the finite-volume framework, the boundary condition can be treated as an special cell interface. Therefore, the implementation of boundary conditions requires the VDF and macroscopic numerical flux at the interface based on the gas-surface kinetic boundary condition. Taking the classical Maxwell boundary condition as an example, assuming the left side of the interface ijij is the wall and the right side is the interior cell jj, the distribution function at the interface ijij can be determined as follows

fk,ij=[Hn+(αMfk,ijW+(1αM)fk,ijR(𝝃))+(1Hn+)fk,ijR(𝝃)],k=0,1,\begin{gathered}f_{k,ij}=\left[\mathrm{H}_{n}^{+}\left(\alpha_{M}f_{k,ij}^{W}+\left(1-\alpha_{M}\right)f_{k,ij}^{R}(-\bm{{\xi}})\right)+\left(1-\mathrm{H}_{n}^{+}\right)f_{k,ij}^{R}(\bm{{\xi}})\right],\quad k=0,1,\end{gathered} (24)

where αM\alpha_{M} is the accommodation coefficient representing the proportion of gas molecule would follow the diffuse reflection, while the remaining proportion follows the specular reflection; Hn+=[1+sign(𝝃𝐧ij)]\mathrm{H}_{n}^{+}=\left[1+\operatorname{sign}\left(\bm{{\xi}}\cdot\mathbf{n}_{ij}\right)\right] is the Heaviside function and fk,ijR(𝝃)f_{k,ij}^{R}(\bm{{\xi}}) represents the incident distribution function; the latter can be obtained by interpolating the VDFs at the cell center to the position of cell interface as

fk,ijR(𝝃)=fk,j+fk,j(𝐱ij𝐱i),k=0,1.\begin{gathered}f_{k,ij}^{R}(\bm{{\xi}})=f_{k,j}+\nabla f_{k,j}\cdot\left(\mathbf{x}_{ij}-\mathbf{x}_{i}\right),\quad k=0,1.\end{gathered} (25)

The αM\alpha_{M} proportion of distribution function at the left side obeys diffuse reflection and thus becomes an equilibrium distribution function as

f0,ijW=ρW(12πTtW)3/2exp(|𝝃𝐮W|22TtW),f1,ijW=dr2TrWρW(12πTtW)3/2exp(|𝝃𝐮W|22TtW),\begin{gathered}f_{0,ij}^{W}=\rho^{W}\left(\frac{1}{2\pi T_{t}^{W}}\right)^{3/2}\exp\left(-\frac{\left|\bm{\xi}-\mathbf{u}^{W}\right|^{2}}{2T_{t}^{W}}\right),\\ f_{1,ij}^{W}=\frac{d_{r}}{2}T_{r}^{W}\rho^{W}\left(\frac{1}{2\pi T_{t}^{W}}\right)^{3/2}\exp\left(-\frac{\left|\bm{\xi}-\mathbf{u}^{W}\right|^{2}}{2T_{t}^{W}}\right),\end{gathered} (26)

while the rest of gas molecules undergo specular reflection and the corresponding VDFs follow fk,ijR(𝝃)f_{k,ij}^{R}(-\bm{{\xi}}). The superscript WW in Eq. (26) denotes that the corresponding macroscopic quantities need to be determined based on the wall boundary conditions.

In the first GSIS for nonlinear flows [27], the numerical fluxes of the macroscopic equations at the interface 𝐅ij\mathbf{F}_{ij} are calculated by the VDFs at the interface ijij as

𝐅ij=𝐅c,ij+𝐅v,ij=ξn[f0,ijξxf0,ijξyf0,ijξzf0,ijc22f0,ij+f1,ijf1,ij]𝑑𝝃.\mathbf{F}_{ij}=\mathbf{F}_{c,ij}+\mathbf{F}_{v,ij}=\int\xi_{n}\left[\begin{array}[]{c}f_{0,ij}\\ \xi_{x}f_{0,ij}\\ \xi_{y}f_{0,ij}\\ \xi_{z}f_{0,ij}\\ \frac{c^{2}}{2}f_{0,ij}+f_{1,ij}\\ f_{1,ij}\end{array}\right]d\bm{\xi}. (27)

While this boundary treatment ensures that the macroscopic equations attain consistent boundary conditions with the mesoscopic kinetic equations, it fails to maintain updates during macroscopic iterations, which does not realize the best possible performance of GSIS. In the second GSIS for nonlinear flows [24], we introduce a correction for macroscopic boundaries based on the increment of macroscopic quantities at the cell center. That is, in the inner iteration, after a single LU-SGS iteration, we obtain the global macro quantity increment and use a method similar to the Roe scheme to modify the interface flux using the physical quantity increment of the boundary element:

𝐅ij=𝐅ij+Fc(Δ𝐖i)2.\mathbf{F}_{ij}^{\star}=\mathbf{F}_{ij}+\frac{F_{c}(\Delta\mathbf{W}_{i})}{2}. (28)

This approach involves updating the boundary flux based on the updates in the interior flow, ensuring that the macroscopic wall flux evolves in tandem with the interior field to boost convergence. Nevertheless, the macroscopic boundary, unable to evolve independently, would result in a delay in the evolution of the macroscopic equations at the boundary during the inner iterations. This delay hinders the acceleration of convergence for the kinetic equations at the boundary.

3.3 New treatment

Hence, with the hinder-sight, we propose a GBT for the macroscopic synthetic equation. The fundamental concept lies in approximating the incident distribution function at the interface using the truncated distribution function, thereby achieving an explicit reconstruction of the macroscopic numerical flux at the interface. In order to approximate the distribution function and close the higher-order terms in the macroscopic equations, Grad proposed a series of distribution functions based on Hermite orthogonal polynomials. Following this expansion, we present the truncated distribution function fijT(t,𝒙ij,𝝃,Ir)f_{ij}^{T}(t,\bm{x}_{ij},\bm{\xi},I_{r}) applicable to polyatomic gas with rotational energy IrI_{r} as

fijT(t,𝐱ij,ξ,Ir)=f0,ijeqfIr,ijeq[1+𝝈𝐜𝐜2ρ(Tt)2+𝐪t𝐜ρ(Tt)2(c25Tt1)+𝐪r𝐜dr2ρTtTr(IrTrdr2)],f_{ij}^{T}\left(t,\mathbf{x}_{ij},\xi,I_{r}\right)=f_{0,ij}^{eq}f_{I_{r},ij}^{eq}\left[1+\frac{\bm{\sigma}\cdot\mathbf{cc}}{2\rho\left(T_{t}\right)^{2}}+\frac{\mathbf{q}_{t}\cdot\mathbf{c}}{\rho\left(T_{t}\right)^{2}}\left(\frac{c^{2}}{5T_{t}}-1\right)+\frac{\mathbf{q}_{r}\cdot\mathbf{c}}{\frac{d_{r}}{2}\rho T_{t}T_{r}}\left(\frac{I_{r}}{T_{r}}-\frac{d_{r}}{2}\right)\right], (29)

where fIr,ijeqf_{I_{r},ij}^{eq} is defined as

fIr,ijeq=Irdr/21Γ(dr/2)(Tr)dr/2exp(IrTr).f_{I_{r},ij}^{eq}=\frac{I_{r}^{d_{r}/2-1}}{\Gamma\left(d_{r}/2\right)\left(T_{r}\right)^{d_{r}/2}}\exp\left(-\frac{I_{r}}{T_{r}}\right). (30)

To reduce the rotational energy space for the distribution function, we have

f0,ijT(t,𝐱ij,𝝃)\displaystyle f_{0,ij}^{T}(t,\mathbf{x}_{ij},\bm{\xi}) =fijT(t,𝐱ij,𝝃,Ir)Ir\displaystyle=\left\langle f_{ij}^{T}\left(t,\mathbf{x}_{ij},\bm{\xi},I_{r}\right)\right\rangle_{I_{r}} (31)
=f0,ijeqfIr,ijeq[1+𝝈𝐜𝐜2ρ(Tt)2+𝐪t𝐜ρ(Tt)2(c25Tt1)+𝐪r𝐜dr2ρTtTr(IrTrdr2)]Ir\displaystyle=\left\langle f_{0,ij}^{eq}f_{I_{r},ij}^{eq}\left[1+\frac{\bm{\sigma}\cdot\mathbf{cc}}{2\rho\left(T_{t}\right)^{2}}+\frac{\mathbf{q}_{t}\cdot\mathbf{c}}{\rho\left(T_{t}\right)^{2}}\left(\frac{c^{2}}{5T_{t}}-1\right)+\frac{\mathbf{q}_{r}\cdot\mathbf{c}}{\frac{d_{r}}{2}\rho T_{t}T_{r}}\left(\frac{I_{r}}{T_{r}}-\frac{d_{r}}{2}\right)\right]\right\rangle_{I_{r}}
=f0,ijeq[1+𝝈𝐜𝐜2ρTt2+𝐪t𝐜ρTt2(c25Tt1)],\displaystyle=f_{0,ij}^{eq}\left[1+\frac{\bm{\sigma}\cdot\mathbf{cc}}{2\rho T_{t}^{2}}+\frac{\mathbf{q}_{t}\cdot\mathbf{c}}{\rho T_{t}^{2}}\left(\frac{c^{2}}{5T_{t}}-1\right)\right],

and

f1,ijT(t,𝐱ij,𝝃)\displaystyle f_{1,ij}^{T}(t,\mathbf{x}_{ij},\bm{\xi}) =IrfijT(t,𝐱ij,𝝃,Ir)Ir\displaystyle=\left\langle I_{r}f_{ij}^{T}\left(t,\mathbf{x}_{ij},\bm{\xi},I_{r}\right)\right\rangle_{I_{r}} (32)
=Irf0,ijeqfIr,ijeq[1+𝝈𝐜𝐜2ρ(Tt)2+𝐪t𝐜ρ(Tt)2(c25Tt1)+𝐪r𝐜dr2ρTtTr(IrTrdr2)]Ir\displaystyle=\left\langle I_{r}f_{0,ij}^{eq}f_{I_{r},ij}^{eq}\left[1+\frac{\bm{\sigma}\cdot\mathbf{cc}}{2\rho\left(T_{t}\right)^{2}}+\frac{\mathbf{q}_{t}\cdot\mathbf{c}}{\rho\left(T_{t}\right)^{2}}\left(\frac{c^{2}}{5T_{t}}-1\right)+\frac{\mathbf{q}_{r}\cdot\mathbf{c}}{\frac{d_{r}}{2}\rho T_{t}T_{r}}\left(\frac{I_{r}}{T_{r}}-\frac{d_{r}}{2}\right)\right]\right\rangle_{I_{r}}
=(dr2Tr)f0,ijeq[1+𝝈𝐜𝐜2ρTt2+𝐪t𝐜ρTt2(c25Tt1)]+f0,ijeq𝐪r𝐜ρTt,\displaystyle=\left(\frac{d_{r}}{2}T_{r}\right)f_{0,ij}^{eq}\left[1+\frac{\bm{\sigma}\cdot\mathbf{cc}}{2\rho T_{t}^{2}}+\frac{\mathbf{q}_{t}\cdot\mathbf{c}}{\rho T_{t}^{2}}\left(\frac{c^{2}}{5T_{t}}-1\right)\right]+f_{0,ij}^{eq}\frac{\mathbf{q}_{r}\cdot\mathbf{c}}{\rho T_{t}},

where Ir=0+()𝑑Ir\langle\cdots\rangle_{I_{r}}=\int_{0}^{+\infty}(\cdots)dI_{r} is the integration over rotational energy space. A detailed derivation of the truncated distribution function is provided in B. It is observed that the truncated distribution functions fijTf_{ij}^{T} in Eqs. (31) and (32) are expressed as polynomials in terms of macroscopic conserved variables, stress and heat flux and fijTf_{ij}^{T} would degenerate to fijeqf_{ij}^{eq} when the stress and heat flux are zero.

With the approximation of the incident distribution function fk,ijRf_{k,ij}^{R} by the truncated distribution function fijTf_{ij}^{T}, the macroscopic numerical flux at the interface can be explicitly reconstructed. Taking the case of complete diffuse reflection (αM\alpha_{M} = 1) in Eq. (24) as an example, the macroscopic numerical flux at the interface can be expressed as follows

𝐅ij=𝐅c,ij+𝐅v,ij=ρL[F1LnxF2L+mxF3L+oxF3LnyF2L+myF3L+oyF3LnzF2L+mzF3L+ozF3LF5LF6L]+ρR[F1RnxF2R+mxF3R+oxF3RnyF2R+myF3R+oyF3RnzF2R+mzF3R+ozF3RF5RF6R].\mathbf{F}_{ij}=\mathbf{F}_{c,ij}+\mathbf{F}_{v,ij}=\rho^{L}\left[\begin{array}[]{c}F_{1}^{L}\\ n_{x}F_{2}^{L}+m_{x}F_{3}^{L}+o_{x}F_{3}^{L}\\ n_{y}F_{2}^{L}+m_{y}F_{3}^{L}+o_{y}F_{3}^{L}\\ n_{z}F_{2}^{L}+m_{z}F_{3}^{L}+o_{z}F_{3}^{L}\\ F_{5}^{L}\\ F_{6}^{L}\end{array}\right]+\rho^{R}\left[\begin{array}[]{c}F_{1}^{R}\\ n_{x}F_{2}^{R}+m_{x}F_{3}^{R}+o_{x}F_{3}^{R}\\ n_{y}F_{2}^{R}+m_{y}F_{3}^{R}+o_{y}F_{3}^{R}\\ n_{z}F_{2}^{R}+m_{z}F_{3}^{R}+o_{z}F_{3}^{R}\\ F_{5}^{R}\\ F_{6}^{R}\end{array}\right]. (33)

Here, 𝐧ij=[nx,ny,nz]\mathbf{n}_{ij}=\left[n_{x},n_{y},n_{z}\right], 𝐦ij=[mx,my,mz]\mathbf{m}_{ij}=\left[m_{x},m_{y},m_{z}\right], and 𝐨ij=[ox,oy,oz]\mathbf{o}_{ij}=\left[o_{x},o_{y},o_{z}\right] constitute a set of orthogonal vectors determining the normal and two tangential directions of the interface. The definitions of the integral coefficients F1F_{1} to F6F_{6} at the interface are defined as follows

F1L=ξn1ξτ0ξs0f0,ijT>0,F1R=ξn1ξτ0ξs0f0,ijT<0,\displaystyle F_{1}^{L}=\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{>0},\quad F_{1}^{R}=\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{<0}, (34)
F2L=ξn2ξτ0ξs0f0,ijT>0,F2R=ξn2ξτ0ξs0f0,ijT<0,\displaystyle F_{2}^{L}=\left\langle\xi_{n}^{2}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{>0},\quad F_{2}^{R}=\left\langle\xi_{n}^{2}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{<0},
F3L=ξn1ξτ1ξs0f0,ijT>0,F3R=ξn1ξτ1ξs0f0,ijT<0,\displaystyle F_{3}^{L}=\left\langle\xi_{n}^{1}\xi_{\tau}^{1}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{>0},\quad F_{3}^{R}=\left\langle\xi_{n}^{1}\xi_{\tau}^{1}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{<0},
F4L=ξn1ξτ0ξs1f0,ijT>0,F4R=ξn1ξτ0ξs1f0,ijT<0,\displaystyle F_{4}^{L}=\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{1}f_{0,ij}^{T}\right\rangle_{>0},\quad F_{4}^{R}=\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{1}f_{0,ij}^{T}\right\rangle_{<0},

and

F5L=12(ξn3ξτ0ξs0f0,ijT>0+ξn1ξτ2ξs0f0,ijT>0+ξn1ξτ0ξs2f0,ijT>0)+F6L,F5R=12(ξn3ξτ0ξs0f0,ijT<0+ξn1ξτ2ξs0f0,ijT<0+ξn1ξτ0ξs2f0,ijT<0)+F6R,F6L=dr2Trξn1ξτ0ξs0f0,ijT>0+ξn1ξτ0ξs0f1,ijT>0,F6R=dr2Trξn1ξτ0ξs0f0,ijT<0+ξn1ξτ0ξs0f1,ijT<0,\begin{gathered}F_{5}^{L}=\frac{1}{2}\left(\left\langle\xi_{n}^{3}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{>0}+\left\langle\xi_{n}^{1}\xi_{\tau}^{2}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{>0}+\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{2}f_{0,ij}^{T}\right\rangle_{>0}\right)+F_{6}^{L},\\ F_{5}^{R}=\frac{1}{2}\left(\left\langle\xi_{n}^{3}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{<0}+\left\langle\xi_{n}^{1}\xi_{\tau}^{2}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{<0}+\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{2}f_{0,ij}^{T}\right\rangle_{<0}\right)+F_{6}^{R},\\ F_{6}^{L}=\frac{d_{r}}{2}T_{r}\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{>0}+\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{0}f_{1,ij}^{T}\right\rangle_{>0},\\ F_{6}^{R}=\frac{d_{r}}{2}T_{r}\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{<0}+\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{0}f_{1,ij}^{T}\right\rangle_{<0},\end{gathered} (35)

where the subscripts nn, τ\tau, and ss correspond to the directions of molecular velocity after the transformation into the local coordinate system. The symbol <0\langle\cdot\rangle_{<0} represents the integral with the interval from negative infinity to zero and the >0\langle\cdot\rangle_{>0} represents the integral with the interval from zero to infinity:

fT<0=1ρ0++()fT𝑑ξn𝑑ξτ𝑑ξs\displaystyle\langle\cdots f^{T}\rangle_{<0}=\frac{1}{\rho}\int_{-\infty}^{0}\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}(\cdots)f^{T}d\xi_{n}d\xi_{\tau}d\xi_{s} (36)
fT>0=1ρ0+++()fT𝑑ξn𝑑ξτ𝑑ξs.\displaystyle\langle\cdots f^{T}\rangle_{>0}=\frac{1}{\rho}\int_{0}^{+\infty}\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}(\cdots)f^{T}d\xi_{n}d\xi_{\tau}d\xi_{s}.

The next task is computing the moment coefficients including ξnoξτpξsqfijT>0\left\langle\xi_{n}^{o}\xi_{\tau}^{p}\xi_{s}^{q}f_{ij}^{T}\right\rangle_{>0} and ξnoξτpξsqfijT<0\left\langle\xi_{n}^{o}\xi_{\tau}^{p}\xi_{s}^{q}f_{ij}^{T}\right\rangle_{<0}. Based on the binomial theorem, the coefficients associated with fijTf_{ij}^{T} on the right side of interface can be expressed through a two-step recursion as

ξniξτjξskfT<0=m=0kk!m!(km)!ξniξτjcskfT<0uskm,\displaystyle\left\langle\xi_{n}^{i}\xi_{\tau}^{j}\xi_{s}^{k}f^{T}\right\rangle_{<0}=\sum_{m=0}^{k}\frac{k!}{m!(k-m)!}\left\langle\xi_{n}^{i}\xi_{\tau}^{j}c_{s}^{k}f^{T}\right\rangle_{<0}u_{s}^{k-m}, (37)
ξniξτjcskfT<0=m=0jj!m!(jm)!ξnicτmcskfT<0uτjm.\displaystyle\left\langle\xi_{n}^{i}\xi_{\tau}^{j}c_{s}^{k}f^{T}\right\rangle_{<0}=\sum_{m=0}^{j}\frac{j!}{m!(j-m)!}\left\langle\xi_{n}^{i}c_{\tau}^{m}c_{s}^{k}f^{T}\right\rangle_{<0}u_{\tau}^{j-m}.

Here, ii, jj, and kk represent non-negative integers and fTf^{T} can represent both f0Tf_{0}^{T} and f1Tf_{1}^{T}. In accordance with the definition of the truncated distribution function, we can express ξnicτjcskf0T<0\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{0}^{T}\right\rangle_{<0} and ξnicτjcskf1T<0\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{1}^{T}\right\rangle_{<0} as the linear combinations of integral coefficients associated with the equilibrium distribution function as follows

ξnicτjcskf0T<0=\displaystyle\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{0}^{T}\right\rangle_{<0}= ξnicn0<0eqcτjeqcskeq+σnnξnicn2<0eqcτjeqcskeq\displaystyle\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}+\sigma_{nn}^{*}\left\langle\xi_{n}^{i}c_{n}^{2}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq} (38)
+σττξnicn0<0eqcτj+2eqcskeq+σssξnicn0<0eqcτjeqcsk+2eq\displaystyle+\sigma_{\tau\tau}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+2}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}+\sigma_{ss}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k+2}\right\rangle^{eq}
+2σnτξnicn1<0eqcτj+1eqcskeq+2σnsξnicn1<0eqcτjeqcsk+1eq\displaystyle+2\sigma_{n\tau}^{*}\left\langle\xi_{n}^{i}c_{n}^{1}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+1}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}+2\sigma_{ns}^{*}\left\langle\xi_{n}^{i}c_{n}^{1}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k+1}\right\rangle^{eq}
+2στsξnicn0<0eqcτj+1eqcsk+1eqqt,nξnicn1<0eqcτjeqcskeq\displaystyle+2\sigma_{\tau s}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+1}\right\rangle^{eq}\left\langle c_{s}^{k+1}\right\rangle^{eq}-q_{t,n}^{*}\left\langle\xi_{n}^{i}c_{n}^{1}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}
qt,τξnicn0<0eqcτj+1eqcskeqqt,sξnicn0<0eqcτjeqcsk+1eq\displaystyle-q_{t,\tau}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+1}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}-q_{t,s}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k+1}\right\rangle^{eq}
+0.4λtqt,nξnicn3<0eqcτjeqcskeq+0.4λtqt,nξnicn1<0eqcτjeqcsk+2eq\displaystyle+0.4\lambda_{t}q_{t,n}^{*}\left\langle\xi_{n}^{i}c_{n}^{3}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}+0.4\lambda_{t}q_{t,n}^{*}\left\langle\xi_{n}^{i}c_{n}^{1}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k+2}\right\rangle^{eq}
+0.4λtqt,nξnicn1<0eqcτj+2eqcskeq+0.4λtqt,τξnicn2<0eqcτj+1eqcskeq\displaystyle+0.4\lambda_{t}q_{t,n}^{*}\left\langle\xi_{n}^{i}c_{n}^{1}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+2}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}+0.4\lambda_{t}q_{t,\tau}^{*}\left\langle\xi_{n}^{i}c_{n}^{2}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+1}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}
+0.4λtqt,τξnicn0<0eqcτj+3eqcskeq+0.4λtqt,τξnicn0<0eqcτj+1eqcsk+2eq\displaystyle+0.4\lambda_{t}q_{t,\tau}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+3}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}+0.4\lambda_{t}q_{t,\tau}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+1}\right\rangle^{eq}\left\langle c_{s}^{k+2}\right\rangle^{eq}
+0.4λtqt,sξnicn2<0eqcτjeqcsk+1eq+0.4λtqt,sξnicn0<0eqcτj+2eqcsk+1eq\displaystyle+0.4\lambda_{t}q_{t,s}^{*}\left\langle\xi_{n}^{i}c_{n}^{2}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k+1}\right\rangle^{eq}+0.4\lambda_{t}q_{t,s}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+2}\right\rangle^{eq}\left\langle c_{s}^{k+1}\right\rangle^{eq}
+0.4λtqt,sξnicn0<0eqcτjeqcsk+3eq,\displaystyle+0.4\lambda_{t}q_{t,s}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k+3}\right\rangle^{eq},

and

ξnicτjcskf1T<0=\displaystyle\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{1}^{T}\right\rangle_{<0}= dr2Trξnicτjcskf0T<0+qr,nξnicn1<0eqcτjeqcskeq\displaystyle\frac{d_{r}}{2}T_{r}\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{0}^{T}\right\rangle_{<0}+q_{r,n}^{*}\left\langle\xi_{n}^{i}c_{n}^{1}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq} (39)
+qr,τξnicn0<0eqcτj+1eqcskeq+qr,sξnicn0<0eqcτjeqcsk+1eq,\displaystyle+q_{r,\tau}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j+1}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}+q_{r,s}^{*}\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{<0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k+1}\right\rangle^{eq},

where λt=1/2Tt\lambda_{t}=1/2T_{t}, and the stress and heat flux marked with an asterisk superscript are given by

σnn=σnn/(2ptTt),σnτ=σnτ/(2ptTt),σns=σns/(2ptTt),σττ=σττ/(2ptTt),στs=στs/(2ptTt),qt,n=qt,n/(ptTt),qt,τ=qt,τ/(ptTt),qt,s=qt,s/(ptTt),qr,n=qr,n/pt,qr,τ=qr,τ/pt,qr,s=qr,s/pt.\begin{gathered}\sigma_{nn}^{*}=\sigma_{nn}/\left(2p_{t}T_{t}\right),\quad\sigma_{n\tau}^{*}=\sigma_{n\tau}/\left(2p_{t}T_{t}\right),\quad\sigma_{ns}^{*}=\sigma_{ns}/\left(2p_{t}T_{t}\right),\\ \sigma_{\tau\tau}^{*}=\sigma_{\tau\tau}/\left(2p_{t}T_{t}\right),\quad\sigma_{\tau s}^{*}=\sigma_{\tau s}/\left(2p_{t}T_{t}\right),\\ q_{t,n}^{*}=q_{t,n}/\left(p_{t}T_{t}\right),\quad q_{t,\tau}^{*}=q_{t,\tau}/\left(p_{t}T_{t}\right),\quad q_{t,s}^{*}=q_{t,s}/\left(p_{t}T_{t}\right),\\ q_{r,n}^{*}=q_{r,n}/p_{t},\quad q_{r,\tau}^{*}=q_{r,\tau}/p_{t},\quad q_{r,s}^{*}=q_{r,s}/p_{t}.\end{gathered} (40)

The integral coefficients associated with the equilibrium distribution function are defined as

eq=1ρ+++()f0eq𝑑ξn𝑑ξτ𝑑ξs\displaystyle\langle\cdots\rangle^{eq}=\frac{1}{\rho}\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}(\cdots)f_{0}^{eq}d\xi_{n}d\xi_{\tau}d\xi_{s} (41)
<0eq=1ρ0++()f0eq𝑑ξn𝑑ξτ𝑑ξs\displaystyle\langle\cdots\rangle_{<0}^{eq}=\frac{1}{\rho}\int_{-\infty}^{0}\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}(\cdots)f_{0}^{eq}d\xi_{n}d\xi_{\tau}d\xi_{s}
>0eq=1ρ0+++()f0eq𝑑ξn𝑑ξτ𝑑ξs.\displaystyle\langle\cdots\rangle_{>0}^{eq}=\frac{1}{\rho}\int_{0}^{+\infty}\int_{-\infty}^{+\infty}\int_{-\infty}^{+\infty}(\cdots)f_{0}^{eq}d\xi_{n}d\xi_{\tau}d\xi_{s}.

To obtain the coefficient of ξnicnj<0eq\left\langle\xi_{n}^{i}c_{n}^{j}\right\rangle_{<0}^{eq}, we have the following equation

ξnicnj<0eq=(1)jmm=0jj!m!(jm)!ξnm+i<0eq(un)jm.\left\langle\xi_{n}^{i}c_{n}^{j}\right\rangle_{<0}^{eq}=(-1)^{j-m}\sum_{m=0}^{j}\frac{j!}{m!(j-m)!}\left\langle\xi_{n}^{m+i}\right\rangle_{<0}^{eq}\left(u_{n}\right)^{j-m}. (42)

While our previous description only provides the calculation procedure for the integration coefficients on the right side of interface, i.e., ξniξτjξskfT<0\left\langle\xi_{n}^{i}\xi_{\tau}^{j}\xi_{s}^{k}f^{T}\right\rangle_{<0}, the coefficients on the left side follows a similar procedure. The only difference is that the expressions <0\langle\cdot\cdot\cdot\rangle_{<0} in Eqs. (37), (38) and (39) should be replaced by >0\langle\cdot\cdot\cdot\rangle_{>0} and the macroscopic variables in the right side of the interface should be replaced by that in the left side. Taking the diffuse reflection boundary condition as an example, the integration coefficients on the left side of the interface corresponding to Eqs. (38) and (39) can be simplified to

ξnicτjcskf0T>0=\displaystyle\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{0}^{T}\right\rangle_{>0}= ξnicn0>0eqcτjeqcskeq,\displaystyle\left\langle\xi_{n}^{i}c_{n}^{0}\right\rangle_{>0}^{eq}\left\langle c_{\tau}^{j}\right\rangle^{eq}\left\langle c_{s}^{k}\right\rangle^{eq}, (43)

and

ξnicτjcskf1T>0=\displaystyle\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{1}^{T}\right\rangle_{>0}= dr2Trξnicτjcskf0T>0.\displaystyle\frac{d_{r}}{2}T_{r}\left\langle\xi_{n}^{i}c_{\tau}^{j}c_{s}^{k}f_{0}^{T}\right\rangle_{>0}. (44)

The remaining task is to calculate the terms ξnk>0eq\left\langle\xi_{n}^{k}\right\rangle_{>0}^{eq}, ξnk<0eq\left\langle\xi_{n}^{k}\right\rangle_{<0}^{eq}, cτkeq\left\langle c_{\tau}^{k}\right\rangle^{eq} and cskeq\left\langle c_{s}^{k}\right\rangle^{eq}, which are related to the moment integration of equilibrium state, see C for details.

Prior to computing the macroscopic numerical flux, it is essential to obtain the macroscopic variables on both sides of the interface 𝐖\mathbf{W}. It is worth noting that the expression of macroscopic numerical flux presented here is of a general form, without specifying whether the wall is located to the left or right of the interface. We continue to use the example where the wall is located to the left of the interface. The values 𝐖R\mathbf{W}^{R} on the right side of the interface can be interpolated to the interface by the values at the centers of the nearest cells adjacent to the wall as

𝐖R(𝐱ij)=𝐖(𝐱j)+L(𝐖j)𝐖(𝐱j)(𝐱ij𝐱j),\mathbf{W}^{R}\left(\mathbf{x}_{ij}\right)=\mathbf{W}\left(\mathbf{x}_{j}\right)+L\left(\mathbf{W}_{j}\right)\nabla\mathbf{W}\left(\mathbf{x}_{j}\right)\cdot\left(\mathbf{x}_{ij}-\mathbf{x}_{j}\right), (45)

where L(𝐖)L\left(\mathbf{W}\right) is the Venkatakrishnan’s limiter [45] functions of the variable 𝐖\mathbf{W}. The gradient 𝐖(𝐱)n\nabla\mathbf{W}(\mathbf{x})^{n} can be computed by the the least squares method.

In addition, the numerical integration based on the truncated distribution function requires the determination of stress and heat flux. Since the boundary conditions here are intended for the macroscopic synthetic equation, the stress and heat flux should be obtained explicitly based on the constitutive relations and taking into account the high-order corrections applied in GSIS as

𝝈R=μ(𝐮+𝐮T23uI)+Δ𝝈j,𝐪tR=κtTt+Δ𝐪t,j,𝐪rR=κrTr+Δ𝐪r,j,\begin{gathered}\bm{\sigma}^{R}=-\mu\left(\nabla\mathbf{u}+\nabla\mathbf{u}^{\mathrm{T}}-\frac{2}{3}\nabla\cdot u\mathrm{I}\right)+\Delta\bm{\sigma}_{j},\\ \mathbf{q}_{t}^{R}=-\kappa_{t}\nabla T_{t}+\Delta\mathbf{q}_{t,j},\\ \mathbf{q}_{r}^{R}=-\kappa_{r}\nabla T_{r}+\Delta\mathbf{q}_{r,j},\end{gathered} (46)

where the explicit constitutive relations facilitate real-time updates of the macroscopic boundary conditions during inner iterations, boosting faster information exchange in the flow field, and consequently achieving quicker convergence. Furthermore, the introduction of high-order corrections ensures that the macroscopic numerical flux at the boundary remains accurate and stable even under strong rarefied flow conditions.

The macroscopic variables on the left side of the interface, denoted as 𝐖L\mathbf{W}^{L}, are determined based on the boundary conditions, i.e., 𝐖L=𝐖W\mathbf{W}^{L}=\mathbf{W}^{W}, and the stress and heat flux on the left side of the wall interface need to be set to zero. Taking isothermal wall boundary conditions as an example, the velocity at the interface is set to 𝐮L=𝐮W=0\mathbf{u}^{L}=\mathbf{u}^{W}=0, while the temperature is set to the wall temperature TL=TWT^{L}=T^{W}. The density is determined according to the non-penetration boundary condition as follows:

ρW=2πTtWξn1ξτ0ξs0f0,ijT<0.\rho_{W}=-\sqrt{\frac{2\pi}{T_{t}^{W}}}\left\langle\xi_{n}^{1}\xi_{\tau}^{0}\xi_{s}^{0}f_{0,ij}^{T}\right\rangle_{<0}. (47)

4 Numerical experiments

In this section, we assess the performance of the GSIS-GBT. Our tests focus on two key issues: (i) whether the GBT can yield stable and accurate results, especially in hypersonic and strong rarefied flow scenarios, and (ii) whether the GBT can assist the further acceleration of GSIS. The planar Fourier flow, hypersonic flow past a cylinder, pressure-driven flow in a variable-diameter circular pipe, and the multiscale flow around a hypersonic technology vehicle, are considered.

The viscosity index ω\omega in Eq. (3) are chosen as 0.81, 0.75 and 0.71 for the monatomic gas, air and nitrogen, respectively. Following the previous study [39], the thermal relaxation rates in Eq. (8) are given as Atr=0.059A_{tr}=-0.059, Att=0.786A_{tt}=0.786, Art=0.059A_{rt}=-0.059 and Arr=0.842A_{rr}=0.842 for polyatomic gas. The convergence criterion is determined based on Eq. (23) to ensure the convergence of all macroscopic quantities.

4.1 Planar Fourier flow

Consider two stationary parallel plates separated by a distance HH, with temperatures T0+ΔTT_{0}+\Delta T and T0+ΔTT_{0}+\Delta T on the upper and lower plates, respectively. Due to the temperature difference, the gas temperature and density distribution exhibit gradients. Furthermore, with the enhancement of rarefaction effects, a notable temperature jump will be observed.

The computational settings are as follows: the plate spacing HH is set to 1.0 as the reference length, and a mesh of 50 points is employed to discretize the computational domain of y[0,1]y\in[0,1] uniformly. The monatomic gas is initially set to a state with the reference density ρ0\rho_{0} and temperature T0T_{0} both equal to 1.0. The temperature difference at the wall is set to ΔT=0.25\Delta T=0.25, resulting in temperatures of 1.25 and 0.75 for the upper and lower surfaces, respectively. The wall boundary condition is subjected to complete diffuse reflection. The Gauss–Hermite quadrature with 28×2828\times 28 points is adopted in the discretization of velocity domain. The collision time τ\tau and shear viscosity μ\mu are determined according to Eqs. (2) and (3), with the rotational collision number set to 0.001 to recover the monatomic gas model. To expedite computations and facilitate rapid convergence to steady state, the number of inner iterations in Eq. (18) is set to m=100m=100, and the CFL numbers for the mesoscopic and macroscopic levels are designated as 1000 and 500, respectively.

Refer to caption
Refer to caption
Figure 1: Temperature and density profiles of the heat transfer between two plates at different Kn.
Refer to caption
Refer to caption
Refer to caption
Figure 2: Convergence history of the CIS, GSIS and GSIS-GBT in the simulation of heat transfer between two plates at different Kn. The dashed lines represent the density and temperature profiles at different iteration steps, while the circles indicate the accurate results serving as references.

To validate the accuracy of GSIS-GBT, the temperature and velocity profiles between two plates at Kn = 0.01, 0.1, and 1.0 are compared to those obtained from the CIS [27]. Figure 1 shows that, as the Knudsen number increases, the temperature profiles become fatter, and the temperature jump near the wall increases. The agreement between the results obtained from GSIS-GBT and CIS at different Knudsen number confirms the accuracy of GBT in different flow regimes.

Figure 2 compares the iteration steps used in the CIS, GSIS [24], and GSIS-GBT. When Kn=1.0, the three methods complete the computation within 100 steps. However, as the Knudsen number decreases, the iteration steps in CIS significantly increase. At Kn = 0.01, CIS requires more than 1800 steps; the GSIS, due to the coupled macroscopic inner iterations, significantly expedites the information exchange within the flow field, achieving convergence in as few as 85 steps; the GSIS-GBT further reduces the steps to 34, where the additional acceleration is attributed to the GBT on the evolution of the flow field.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Density and temperature at different iteration steps from the GSIS (left column) and GSIS-GBT (right column) for the planar Fourier heat transfer when Kn = 0.01. Reference solutions from [27].

To confirm that the GSIS-GBT further accelerates the evolution of the flow field at the boundaries, we compare in Fig. 3 the density and temperature profiles for the GSIS [24] and GSIS-GBT at intermediate iteration steps. After the first iteration, both methods produce the density and temperature close to their initial states. However, after just 3 iterations, the density and temperature in GSIS-GBT quickly converge towards the steady-state solutions, while GSIS evolves obviously slower. After the seventh iteration, the GSIS-GBT results closely align with the reference ones, while GSIS still has discrepancies with the reference solution, mostly near the wall. Hence, the relative error in GSIS-GBT exhibits a rapid decrease after the seventh iteration, whereas that of the GSIS performs relatively slower. This example provides a strong support that the GBT aids in accelerating the flow field evolution near the boundaries, thereby expediting the overall convergence.

4.2 Supersonic flow passing through the cylinder

Compared to the flat-plate heat transfer, the hypersonic flow around a cylinder poses greater challenges. The primary challenge arises from that the shock waves can be generated by the interaction between the hypersonic incoming flow and the cylinder surface. The computational parameters are configured as follows: the cylinder radius R=0.5R=0.5 is served as the reference length, and the flow domain is defined as a circular region with a radius of 6, discretized by 128 radial and 100 axial structural grid points. The gas is selected as nitrogen with the rotational collision number set to 2.226. The grid height near the cylinder surface is set to 0.002. The reduced VDFs are introduced to further reduce the molecular velocity space ξz\xi_{z} [24] and a grid of 64 ×\times 64 velocity points is applied, discretized uniformly over the range of [12,12]2\left[-12,12\right]^{2}. The reference temperature for the flow is 273 K. At the initial state, both the dimensionless temperature of the flow field and the wall are set to 1.0. The CFL numbers for the mesoscopic and macroscopic equations are set to 10,000 and 1000, respectively.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Comparison of the velocity (1st column), translational temperature (2nd column) and rotational temperature (3rd column) for supersonic flows around cylinder at Ma = 5.0 and Kn = 1 (1st row) and 0.1 (2nd row). Background contour: GSIS-GBT; White dashed line: GSIS; Black solid line: DSMC.

For the inflow boundary condition, the macroscopic quantities 𝐖=[ρ,𝐮,T]\mathbf{W}_{\infty}=\left[\rho_{\infty},\mathbf{u}_{\infty},T_{\infty}\right] in the far field is prescribed, and the corresponding equilibrium VDFs in Eq. (5) are used. For the outflow boundary condition, the gradients of macroscopic quantities along the flow direction are set to zero.

We first consider the flow with the Mach number 5.0 and Knudsen numbers 1.0, 0.1 and 0.01. Figure 4 compare the density, velocity, and temperature profiles obtained from the GSIS-GBT, GSIS [24], and DSMC. The shown velocity are normalized by the sound speed

cs=γRT0,withγ=5+dr3+dr.c_{s}=\sqrt{\gamma RT_{0}},\quad\text{with}\quad\gamma=\frac{5+d_{r}}{3+d_{r}}. (48)

Results from GSIS and GSIS-GBT align well with those obtained from DSMC, confirming the accuracy of GSIS in rarefied gas flows. Furthermore, the GSIS-GBT exhibits comparable performance to the GSIS in highly rarefied flows, supporting the assertion that GBT accurately provides boundary conditions. When the Knudsen number decreases to 0.01, DSMC would necessitate an exceedingly dense meshes for accurate calculations, rendering the computational cost high. Given that the flow is approaching the slip regime and the rarefaction effects are less pronounced, we present the contours of GSIS-GBT in Fig. 5, along with the results from the Navier-Stokes equations (NS-GBT)111Note that the NS-GBT utilizes the governing equations (13) and (14) to compute the internal interfaces of the flow field, while employing GBT to simulate the velocity slip and temperature jump boundary conditions. In contrast to the macroscopic inner iterations in GSIS-GBT, NS-GBT does not introduce higher-order corrections for stress and heat flux, relying solely on the constitutive relations. As a result, it is applicable primarily to rarefied flows near the slip regime.. The results in Fig. 5 indicate a close agreement between the results from the GSIS-GBT and NS-GBT, particularly in the frontal region of the cylinder.

Refer to caption
Refer to caption
Refer to caption
Figure 5: Comparison of the (left) velocity, (middle) translational temperature and (right) rotational temperature for the supersonic flow around cylinder at Ma = 5.0 and Kn = 0.1. Background contour: GSIS-GBT; White dashed lines: GSIS; Red solid lines: NS-GBT.

To validate the performance of GSIS-GBT in hypersonic flows, results along the stagnation line obtained from the GSIS [24], GSIS-GBT, DSMC, and NS-GBT are compared. Figures 6(a-f) shows that the results from the GSIS and GSIS-GBT exhibit a favorable agreement with DSMC. When the Knudsen number decreases to 0.01 and 0.001, slight discrepancies emerge in the results of GSIS and NS equations, especially in the temperature results presented in Figs. 6(i-l): the temperature profile obtained from the GSIS at the shock wave is slightly higher than the that from the NS equations. This is due to the stronger local rarefaction effects in the shock wave, making the NS equations loss the accuracy. However, outside the shock wave region, results obtained from the two versions of GSIS align very well with those from the NS equations.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Comparison of density, velocity, translational temperature and rotational temperature along the stagnation line for the supersonic cylinder flow at Ma = 5.0 and different Kn.
Table 1: Convergence step and computational time in CIS, GSIS and GSIS-GBT for the supersonic cylinder flow at Ma = 5.0 and various Kn. The simulations are conducted on a parallel computer of AMD Epyc 7742 processor with 20 cores.
CIS GSIS GSIS-GBT Speedup
Kn\mathrm{Kn} Steps Times (s)(\mathrm{s}) Steps Times (s)(\mathrm{s}) Steps Times (s)(\mathrm{s}) Ratio
0.1 462 356 91 138.2 73 109.3 3.3/1.3
0.01 3952 3082 133 199.6 43 64.5 47/3.1
0.001 7865 5876 184 280.5 34 61.1 98/4.6
  • The steps contains the total steps for the 10-step CIS iterations and the subsequent GSIS iterations to convergence. Each iteration of GSIS comprises solving one time mesoscopic equation and 200 times of macroscopic inner iterations.

  • The computational time contains the total time for the 10-step CIS iterations and the subsequent GSIS iterations to convergence.

  • The speed-up ratio comprises the acceleration ratio of GSIS-GBT relative to CIS and the acceleration ratio relative to the original GSIS.

Table 1 presents the convergence steps for hypersonic cylinder flows. Both GSIS [24] and GSIS-GBT exhibit accelerated convergence compared to the CIS across various Knudsen numbers, where the acceleration becomes more pronounced as the Knudsen number decreases. For instance, at Kn = 0.01, CIS require nearly an hour of computational time, whereas GSIS-GBT converges in approximately one minute. This acceleration can be attributed to two factors: (i) the macroscopic inner iterations facilitates the fast exchange of information within the flow field, and (ii) the GBT in the macroscopic equations accelerates the evolution of the flow field at the boundaries, resulting in a nearly five-fold acceleration over the GSIS. In Fig. 7, we present a comparison of velocity and temperature profiles between the GSIS and GSIS-GBT at the same iteration steps. It is seen that the flow evolution near the wall (x=0.5x=-0.5) is more rapid in GSIS-GBT. Specifically, after the 10-th iteration, results from GSIS-GBT closely approach the reference solution, while GSIS exhibits a larger discrepancy to the reference results even after 15 iterations.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Density and temperature profiles at different iteration steps, obtained from the GSIS (lst column) and GSIS-GBT (2nd column) for the supersonic cylinder flow when Ma = 5.0 and Kn = 0.01.
Refer to caption
Refer to caption
Figure 8: Convergence history of GSIS (left) and GSIS-GBT (right) in the simulation of planar Fourier heat transfer at Kn = 0.01 with different numbers of macroscopic inner iterations, e.g., NS800 means that the maximum value of mm in Eq. (18) is 800.

In each iteration of the kinetic equation in GSIS, the macroscopic synthetic equations need to be solved mm times to accelerate the exchange of information within the flow field, see Eq. (18). Intuitively, increasing the value of mm boosts the convergence of GSIS. However, this is not be the case in practice. Taking the case of Kn=0.01\text{Kn}=0.01 as an example, we compare the convergence steps of the GSIS and GSIS-GBT under different numbers of macroscopic iterations. Figure 8 shows that, increasing the solution steps of the NS equations in GSIS [24] has a negligible effect on accelerating convergence: by increasing the inner iterations from 100 to 400 steps, the overall convergence steps decrease only from 138 to 129. However, when GBT is applied to the boundaries, increasing the inner iteration steps significantly reduces the convergence steps. To be specific, when the inner iterations increases from 100 to 400 steps, the convergence steps of GSIS-GBT decrease from 68 to 28, resulting in an acceleration of convergence by nearly 2.5 times. We believe this phenomenon stems from the constrained convergence of the macroscopic equations in the GSIS due to the limitations imposed by integral boundary conditions. With the application of GBT, the macroscopic equations can acquire more real-time information from the boundary, thus reducing the number of outer iteration, see nn in Eq. (11), which is the most time-consuming part.

Finally, the hypersonic flow around a cylinder with the Mach number of 20 is simulated to investigate the efficiency and stability of GBT. Figure 9 illustrates the meshes in physical and velocity spaces. To minimize the number of discrete meshes in the vast velocity space of [36,36]2\left[-36,36\right]^{2}, an technique of unstructured discrete velocity mesh [46] is introduced, with 5550 discrete velocity meshes and local refinement at the positions where the molecular velocities are equal to 0 and 20. In Fig. 9, the velocity and temperature contours obtained from the GSIS-GBT are presented at Kn=0.001, 0.01, and 0.1. In contrast to the hypersonic flow with Ma = 5, the case of Ma = 20 not only results in higher wall temperature and larger velocity variations, but also higher density variations, leading to significantly higher local Knudsen numbers in the wake region. Even at Kn = 0.01, strong rarefied flow occurs within the wake region, posing a considerable challenge for numerical algorithms due to the multiscale nature of the flow [47].

Table 2 provides a comparison of the convergence steps in the CIS, GSIS, and GSIS-GBT. All methods employed the flow field obtained from 4000 steps of the first-order Euler equation for initialization. Prior to iteration, GSIS and GSIS-GBT adopt 20 steps of CIS iterations to initialize the VDFs. Due to the numerical dissipation and slow convergence, CIS costs over 12 hours to converge at Kn = 0.01, while GSIS takes around 1 hour. Furthermore, benefiting from the independent evolution of macroscopic boundary conditions, GSIS-GBT finishes the computation by 51 iterations, and within 5 minutes. These results demonstrate the excellent stability and significant acceleration of GSIS-GBT in hypersonic flow.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: First row: physical and velocity mesh in the supersonic cylinder flow at Ma = 20. (2nd, 3rd and 4th rows) Contours of velocity, translational temperature and rotational temperature when Kn = 0.001 (2nd row), 0.01 (3rd row) and 0.1 (4th row).
Table 2: Convergence step and computational time in CIS, GSIS and GSIS-GBT for the supersonic cylinder flow at Ma = 20.0 and various Kn. The simulations are conducted on a parallel computer of AMD Epyc 7742 with 20 cores.
CIS GSIS GSIS-GBT Speedup
Kn\mathrm{Kn} Steps Times (s)(\mathrm{s}) Steps Times (s)(\mathrm{s}) Steps Times (s)(\mathrm{s}) Ratio
0.1 1778 5569 330 1647 83 490 11/3.3
0.01 15209 46615 655 3340 51 294 158/11
0.001 - - 820 3736 49 284 -/13.2
  • The steps contains the total steps for the 20-step CIS iterations and the subsequent GSIS iterations to convergence. Each iteration of GSIS comprises solving one time mesoscopic equation and 400 times of macroscopic inner iterations.

  • The computational time contains the total time for the 20-step CIS iterations and the subsequent GSIS iterations to convergence.

  • The speed-up ratio comprises the acceleration ratio of GSIS-GBT relative to CIS and the acceleration ratio relative to the original GSIS.

4.3 Pressure-driven flow in a variable-diameter circular pipe

To assess the performance of GSIS-GBT in three-dimensional internal flow, the pressure-driven gas flow through a variable-diameter circular pipe is simulated. Figure 10 illustrates the global geometry and the cross-sectional mesh. As shown in the figure, two cylindrical pipes with radii and lengths of 5 are connected to a small cylindrical pipe with a radius and length of 1. The diffuse gas-surface interaction is adopted. The reference temperature T0T_{0} at all boundaries and the inner domain is set to 1.0. Inlet and outlet boundary conditions are applied at planes of z=0z=0 and z=11z=11, corresponding to the inlet and outlet pressures of pin=1.0p_{in}=1.0 and pout=0.5p_{out}=0.5, respectively. Since it is difficult to give information about macroscopic quantities such as velocity and density in micro-flow problems, the pressure boundary conditions use characteristic relations to determine the variables at the boundary, assuming that the given boundary conditions are the inlet pressure pinp_{in}, the outlet pressure poutp_{out} and the inlet temperature TinT_{in}. For the pressure inlet boundary conditions, the velocity is obtained by interpolation, the temperature is TL=pin/ρLT_{L}=p_{in}/\rho_{L}, where the density is

ρL=ρi+pinpiρiai,\rho_{L}=\rho_{i}+\frac{p_{in}-p_{i}}{\rho_{i}a_{i}}, (49)

where the subscript ii is the cell index on the right side of the boundary and aa is local sound speed. A similar treatment is applied to the pressure outlet boundary:

ρR=ρi+poutpiρiai.\rho_{R}=\rho_{i}+\frac{p_{out}-p_{i}}{\rho_{i}a_{i}}. (50)
Refer to caption
(a) global mesh
Refer to caption
(b) cross-sectional mesh when y=0y=0
Refer to caption
(c) δrp=1000\delta_{rp}=1000
Refer to caption
(d) δrp=100\delta_{rp}=100
Refer to caption
(e) δrp=10\delta_{rp}=10
Refer to caption
(f) δrp=1\delta_{rp}=1
Figure 10: (a,b) Spatial mesh for the pressure-driven pipe flow. (c-f) Contours of WW-velocity in the pressure-driven pipe flow. Background contour: GSIS-GBT; White dashed lines: GSIS.
Table 3: Reduced flow rates computed by different methods for the pressure-driven pipe flow.
      δrp\delta_{rp}       TVD1D [48]       DSMC [49]       GSIS-GBT
      1       0.419       0.405       0.397
      10       0.871       0.866       0.858
      100       1.292       1.29       1.292
      1000       -       1.40       1.405

For cross-validation with other numerical methods, the hard-sphere model is adopted, where the viscosity index ω\omega in Eq. (3) is set to 0.5 and the rotational collision number set to 0.001 to recover the hard-sphere model. The reference length is chosen as the radius of the middle circular pipe R=1R=1. Gas flow is simulated under various rarefaction parameters

δrp=π2Kn.\delta_{rp}=\frac{\sqrt{\pi}}{2\text{Kn}}. (51)

When the rarefaction parameter is greater than or equal to 10, the molecular velocity space is discretized by a 16 ×\times 16 ×\times 16 points by Gauss-Hermite quadrature. In the case of a rarefaction parameter of 1, a uniform Newton-Cotes quadrature with 41 ×\times 41 ×\times 41 points is employed to discretize the molecular velocity space in the range of [6,6]3\left[-6,6\right]^{3}. The computational domain is discretized by 143,370 hexahedral cells.

Table 4: Convergence step and computational time in CIS, GSIS and GSIS-GBT for the pressure-driven pipe flow at various δrp\delta_{rp}. The simulations are conducted on a parallel computer of AMD Epyc 7742 processor with 600 cores.
CIS GSIS GSIS-GBT Speedup
δrp\delta_{rp} Steps Times (s)(\mathrm{s}) Steps Times (s)(\mathrm{s}) Steps Times (s)(\mathrm{s}) Ratio
1 194 345.3 61 277.9 40 167.3 2.06/1.66
10 1395 2316.3 65 280.9 34 135.8 17.1/2.07
100 6793 10990.3 140 666.3 42 173.1 63.5/3.85
1000 36225 60154.3 298 1355.4 24 100.5 598.5/13.5
  • The steps contains the total steps for the 10-step CIS iterations and the subsequent GSIS iterations to convergence. Each iteration of GSIS comprises solving one time mesoscopic equation and 600 times of macroscopic inner iterations.

  • The computational time contains the total time for the 4000-step iterations of first-order Euler solver, 10-step CIS iterations and the subsequent GSIS iterations to convergence.

  • The speed-up ratio comprises the acceleration ratio of GSIS-GBT relative to CIS and the acceleration ratio relative to the original GSIS.

The contours of the WW-velocity at the plane of y=0y=0 obtained from the GSIS-GBT are illustrated in Figure 10, which agree well with the results from the GSIS [24]. Furthermore, the reduced mass flow rate M^\hat{M} is calculated within the circular pipe:

M^=2πA(z)ρuz𝑑x𝑑y,\hat{M}=\sqrt{\frac{2}{\pi}}\int_{A(z)}\rho u_{z}dxdy, (52)

where A(z){A(z)} is the cross-section area at position zz. Table 3 presents a comparison of the reduced mass flow rates calculated by the locally structured reconstruction method [48], DSMC [49], and GSIS-GBT. The GSIS-GBT results exhibit good agreement with both numerical methods. Specifically, the errors of GSIS-GBT, compared to the results from DSMC, remain within 1%.

Table 4 presents the convergence step and computational time in CIS, GSIS and GSIS-GBT. The GSIS-GBT achieves convergence within 40 iterations in this low-speed pressure-driven flows across different rarefaction parameters. Specifically, when δrp=1000\delta_{rp}=1000, the GSIS-GBT is faster than the CIS and GSIS by 598.5 and 13.5 times, respectively, demonstrating excellent acceleration performance.

Refer to caption
Refer to caption
Figure 11: Meshes in the physical (left) and velocity (right) spaces for the flow around hypersonic technology vehicle at Ma = 6.0.
Refer to caption
Refer to caption
Figure 12: Comparison of (left) UU-velocity normalized by the sound speed and (right) translational temperature along the stagnation line (y=0.02y=0.02) at Ma = 6.0 and Kn = 0.005 under different mesh numbers.
Refer to caption
Refer to caption
Figure 13: Comparison of (left) UU-velocity and (right) VV-velocity, normalized by the sound speed, for the flow around the HTV at Ma = 6.0 and Kn = 0.005. Background contour with white solid lines: GSIS-GBT; Black dashed lines: UGKWP [43].

4.4 Multiscale flow around hypersonic technology vehicle

The flow around a Hypersonic Technology Vehicle (HTV) is considered to assess the performance of GSIS-GBT in complex geometry, where the schematic representations of the physical and velocity meshes are illustrated in Fig. 11. The length and width of the HTV are 2 m and 0.3 m, respectively, and the diameter of the front edge is 0.05 m. The characteristic length is chosen as the length of the spacecraft. The HTV is surrounded by the air with a reference temperature of 188.8 K and a reference velocity of 1653 m/s, so the three-dimensional molecular velocity space is discretized using 14806 unstructured meshes, ranging from [36,36]3\left[-36,36\right]^{3}, with local refinement at positions where the molecular velocity is 0 and 6. The wall of HTV is characterized by a fully diffuse boundary condition with constant-temperature of 300 K. The incoming flow has a attack angle of zero-degree. Inlet and outlet boundaries are positioned at far-field regions. Again, the velocity results in this section are normalized with respect to the speed of sound.

To validate the mesh independence, 288864, 598000, and 969696 cells in physical space are simulated. The velocity and temperature profiles along the stagnation line (y=0.02y=0.02, this is the shock region where refined mesh is highly demanded) for the three mesh sets overlap with each other, see Fig. 12, demonstrating the mesh independence has been achieved, due to the asymptotic preserving property of the GSIS [26].

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 14: Contours of the density, UU-velocity, VV-velocity and translational temperature for the supersonic cylinder flow around hypersonic technology vehicle Ma = 6.0 and Kn = 0.005.

Figure 13 presents the velocity contours at the tip of the HTV. The black dashed lines are obtained from the unified gas-kinetic wave-particle (UGKWP) method [43]. By considering the collision and streaming of gas molecules simultaneously in the flux reconstruction, UGKWP has been demonstrated to possess the capability to overcome the limitations imposed by the mean free path while preserving the advantages of stochastic particle methods in hypersonic flow [47, 12, 42]. Due to the small diameter of the tip relative to the characteristic length of the aircraft, a pronounced rarefaction effect occurs at the tip. The study of Fan et al. indicates that at a global Knudsen number of 0.005 and Mach number of 6, the local Knudsen number at the tip can reach 0.44 [43]. Results obtained from the NS equations with first-order slip boundary conditions exhibit significant deviations under these conditions. In contrast, the GSIS-GBT results agree well with those from the UGKWP. Note that in UGKWP 2 million spatial cells are used, while due to the nice asymptotic preserving property of the GSIS [26], 288,864 cells are adequate to get the converged solution.

Figures 14-16 offer a comprehensive visualization of the hypersonic flow over the HTV at Kn = 0.005, 0.05 and 0.5, respectively. With the increasing influence of rarefaction effects, noticeable enhancements are observed in both velocity slip and temperature jump at the wall. Additionally, the thickness of shock waves continues to grow.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 15: Contours of the density, UU-velocity, VV-velocity and translational temperature for the supersonic cylinder flow around hypersonic technology vehicle Ma = 6.0 and Kn = 0.05.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 16: Contours of the density, UU-velocity, VV-velocity and translational temperature for the supersonic cylinder flow around hypersonic technology vehicle Ma = 6.0 and Kn = 0.5.

Table 3 summarizes the convergence steps of different numerical methods. The UGKWP, GSIS [24], and GSIS-GBT all utilized results from the Euler equations (solved by first-order scheme) as the initial field. Note that the UGKWP uses a total of 50,000 steps of evolution to reach the steady-state, followed by an additional 50,000 steps for statistical averaging. The convergence steps and times for GSIS and GSIS-GBT encompass 10 iterations of CIS to initialize the distribution function. Benefiting from the dual acceleration at the boundaries and in the internal field, GSIS-GBT achieves convergence within 50 steps across different Knudsen numbers.

The corresponding computational time of UGKWP based on MPI parallelization employing 92 cores and 2,208,000 cells in physical space is also given in Table 5. Neglecting variations in CPU performance and considering the linear scalability, the computational time of the CPU-UGKWP, when normalized for the same number of parallel cores (600 cores) and cell number of 288,864 in physical space, is approximately 5.71 hours at Knudsen number 0.005; the computational time for GSIS-GBT is only 0.06 hours, which is 96 times faster than CPU-UGKWP and 6.33 times faster than GSIS.

Table 5: Convergence step and computational time in GSIS, GSIS-GBT and UGKWP for the supersonic HTV flow at Ma = 6.0 and various Kn. The GSIS and GSIS-GBT simulations are conducted on a parallel computer of AMD Epyc 7742 processor with 600 cores.
CPU-UGKWP GSIS GSIS-GBT Speedup
Kn\mathrm{Kn} Steps Times (h)(\mathrm{h}) Steps Times (h)(\mathrm{h}) Steps Times (h)(\mathrm{h}) Ratio
0.5 - - 49 0.14 46 0.13 1.07
0.05 - - 77 0.23 37 0.07 3.28
0.005 100000 212.7 129 0.38 23 0.06 6.33
  • The computational time contains the total time for the 50,000-step implicit NS solver and the subsequent UGKWP iterations to convergence. The simulations fo UGKWP are conducted on a parallel computer of AMD Epyc 7k83 with 92 cores.

  • The steps contains the total steps for the 10-step CIS iterations and the subsequent GSIS iterations to convergence. Each iteration of GSIS comprises solving one time mesoscopic equation and 400 times of macroscopic inner iterations.

  • The computational time contains the total time for the 2,000-step iterations of first-order Euler equations, 10-step CIS iterations and the subsequent GSIS iterations to convergence.

  • The speed-up ratio comprises the acceleration ratio of GSIS-GBT relative to the GSIS.

5 Conclusions

In summary, we have proposed a GBT to further accelerate the convergence to the steady state in GSIS. Inspired by the moment methods for handling boundary conditions, we have derived the truncated distribution function to approximate the incident distribution function at the wall, which contains the information of the conserved quantities and the Newton law of viscosity and the Fourier law of heat conductivity from the synthetic equation, as well as the higher-order stress and heat flux from the kinetic equation. Then, we have derived the macroscopic numerical flux at the boundary condition by explicitly integrating the truncated distribution function. Thus, the boundary conditions in the macroscopic synthetic equation is compatible with that in the mesoscopic kinetic equation. It has been found that, the explicit Newton and Fourier constitutive relations facilitate the macroscopic boundary in accelerating the evolution of flow, while higher-order corrections ensure accuracy of boundary treatment from the continuum to rarefied flow regimes.

Representative and challenging numerical cases, ranging from one-dimensional to three-dimensional flows, and from low-speed to hypersonic flows, have been examined to asses the performance of GSIS-GBT. Results in the hypersonic cylinder flows have shown the accuracy and stability of GBT in strong non-equilibrium flows. The GSIS-GBT has also been applied to the Fourier heat transfer and pressure-driven pipe flow, demonstrating its superior efficiency in handling temperature jump phenomena and low-speed internal flows.

More importantly, in the Fourier heat transfer and hypersonic cylinder flow, in addition to the previous versions of GSIS in accelerating the flow evolution in interior computational domain, it has been revealed that GSIS-GBT further accelerates the evolution of the flow field at the boundaries. Specifically, GSIS-GBT can yield speedup ratios ranging from 98 to 159 compared to the conventional iterative scheme in hypersonic cylinder flows with Mach numbers of 5 and 20. Compared to the previous versions of GSIS, a further acceleration of 5 to 13 times can be achieved by GSIS-GBT. This further-acceleration has also been seen in pressure-driven pipe flow, supporting the effectiveness of GBT on low-speed internal flows. Furthermore, GSIS-GBT has been applied to the flow around a hypersonic technology with three-dimensional complex geometries. The GBT enables the GSIS to converge within 50 steps across various flow regimes, thus giving a positive answer to the title of our initial work “Can we find steady-state solutions to multiscale rarefied gas flows within dozens of iterations” [20], even for general nonlinear flows.

Although only steady-state problems have been investigated to achieve accelerated convergence, the extension of GBT to unsteady GSIS is straightforward [23]. Besides, the derivation of GBT is based on a polyatomic gas model with only the rotational degrees of freedom. Within the current framework, further extension to kinetic models with vibration and radiation is not overly challenging. We believe that this work presents significant potential for simulating practical engineering problems involving multiscale rarefied gas flows.

Acknowledgments

This work is supported by the National Natural Science Foundation of China (12172162). Special thanks are given to the Center for Computational Science and Engineering at the Southern University of Science and Technology. The authors thank G. C. Fan in the Zhejiang University for sharing the mesh of hypersonic technology vehicle.

Appendix A Nondimensionalization for the kinetic equation

We provide the dimensionless process for the kinetic theory equations. Variables with a colon superscript denote dimensional quantities. The specific gas constant RR, reference length L^\hat{L}, reference density ρ^0\hat{\rho}_{0}, and reference temperature T^0\hat{T}_{0} are utilized in the dimensionless calculations:

f0=f^0ρ^0/(RT^0)3/2,f1=f^1ρ^0RT^0/(RT^0)3/2,x=x^L^,ρ=ρ^ρ^0,t=t^L^/(RT^0)1/2,(𝝃,𝐜,𝐮)=(𝝃^,𝐜^,𝐮^)(RT^0)1/2,(T,Tt,Tr)=(T^,T^t,T^r)T^0,(𝝈,p,pt)=(𝝈^,p^,p^t)ρ^0RT^0,(qt,qr)=(q^t,q^r)ρ^0RT^0(RT^0)1/2.\begin{gathered}f_{0}=\frac{\hat{f}_{0}}{\hat{\rho}_{0}/\left(R\hat{T}_{0}\right)^{3/2}},\quad f_{1}=\frac{\hat{f}_{1}}{\hat{\rho}_{0}R\hat{T}_{0}/\left(R\hat{T}_{0}\right)^{3/2}},\quad x=\frac{\hat{x}}{\hat{L}},\quad\rho=\frac{\hat{\rho}}{\hat{\rho}_{0}},\quad t=\frac{\hat{t}}{\hat{L}/\left(R\hat{T}_{0}\right)^{1/2}},\\ (\bm{\xi},\mathbf{c},\mathbf{u})=\frac{(\hat{\bm{\xi}},\hat{\mathbf{c}},\hat{\mathbf{u}})}{\left(R\hat{T}_{0}\right)^{1/2}},\quad\left(T,T_{t},T_{r}\right)=\frac{\left(\hat{T},\hat{T}_{t},\hat{T}_{r}\right)}{\hat{T}_{0}},\\ \left(\bm{\sigma},p,p_{t}\right)=\frac{\left(\hat{\bm{\sigma}},\hat{p},\hat{p}_{t}\right)}{\hat{\rho}_{0}R\hat{T}_{0}},\quad\left(q_{t},q_{r}\right)=\frac{\left(\hat{q}_{t},\hat{q}_{r}\right)}{\hat{\rho}_{0}R\hat{T}_{0}\left(R\hat{T}_{0}\right)^{1/2}}.\end{gathered} (53)

The equilibrium distribution functions in Eqs. (4) and (5) are normalized in the same way as f^0\hat{f}_{0} and f^1\hat{f}_{1}.

The Knudsen number is defined as the ratio of the molecular mean free path λ\lambda to the reference length L^\hat{L}:

Kn=λL^,λ=μ(T^0)ρ^0RT^0πRT^02,\text{Kn}=\frac{\lambda}{\hat{L}},\quad\lambda=\frac{\mu(\hat{T}_{0})}{\hat{\rho}_{0}R\hat{T}_{0}}\sqrt{\frac{\pi R\hat{T}_{0}}{2}}, (54)

where μ(T^0)\mu(\hat{T}_{0}) is the gas shear viscosity at the reference temperature T^0\hat{T}_{0}.

Appendix B Truncated distribution function for the non-vibrating polyatomic gas with rotational energy

The truncation of distribution functions fT(t,𝒙,𝝃,Ir)f^{T}(t,\bm{x},\bm{\xi},I_{r}) with rotational energy IrI_{r} is expressed as a linear combination of different orders of peculiar velocity as follows

fT(t,𝒙,𝝃,Ir)=f0eqfIreq(a0+(a11cx+a12cy+a13cz)+(a211cx2+a222cy2+a233cz2)+(a212cxcy+a213cxcz+a223cycz)+(a31cxc2+a32cyc2+a33czc2)+b01Ir+(b11cx+b12cy+b13cz)Ir),f^{T}(t,\bm{x},\bm{\xi},I_{r})=f_{0}^{eq}f_{I_{r}}^{eq}\left(\begin{array}[]{l}a_{0}+\left(a_{11}c_{x}+a_{12}c_{y}+a_{13}c_{z}\right)\\ +\left(a_{211}c_{x}^{2}+a_{222}c_{y}^{2}+a_{233}c_{z}^{2}\right)\\ +\left(a_{212}c_{x}c_{y}+a_{213}c_{x}c_{z}+a_{223}c_{y}c_{z}\right)\\ +\left(a_{31}c_{x}c^{2}+a_{32}c_{y}c^{2}+a_{33}c_{z}c^{2}\right)\\ {\color[rgb]{0,0,0}+b_{01}I_{r}+\left(b_{11}c_{x}+b_{12}c_{y}+b_{13}c_{z}\right)I_{r}}\end{array}\right), (55)

The factors aa and bb are the coefficients need to be determined in the distribution function based on the moment relationship as follows

ρ=fd𝝃dIr,0=𝐜fd𝝃dIr,3ρTt=c2fd𝝃dIr,\displaystyle\rho=\iint f\mathrm{~{}d}\bm{\xi}\mathrm{d}I_{r},\quad 0=\iint\mathbf{c}f\mathrm{~{}d}\bm{\xi}\mathrm{d}I_{r},\quad 3\rho T_{t}=\iint c^{2}f\mathrm{~{}d}\bm{\xi}\mathrm{d}I_{r}, (56)
𝝈=(𝐜𝐜c23I)fd𝝃dIr,𝐪t=𝐜c22fd𝝃dIr,\displaystyle\bm{\sigma}=\iint\left(\mathbf{c}\mathbf{c}-\frac{c^{2}}{3}\mathrm{I}\right)f\mathrm{~{}d}\bm{\xi}\mathrm{d}I_{r},\quad\mathbf{q}_{t}=\iint\mathbf{c}\frac{c^{2}}{2}f\mathrm{~{}d}\bm{\xi}\mathrm{d}I_{r},
dr2ρTr=Irfd𝝃dIr,𝐪r=𝐜Irfd𝝃dIr.\displaystyle\frac{d_{r}}{2}\rho T_{r}=\iint I_{r}f\mathrm{~{}d}\bm{\xi}\mathrm{d}I_{r},\quad\mathbf{q}_{r}=\iint\mathbf{c}I_{r}f\mathrm{~{}d}\bm{\xi}\mathrm{d}I_{r}.

Then, we can get all coefficients aa and bb as

a0=1,b01=0,a211=σxx2ρ(Tt)2,a212=σxy2ρ(Tt)2,a213=σxz2ρ(Tt)2,a212=σxyρ(Tt)2,a213=σxzρ(Tt)2,a223=σyzρ(Tt)2,\displaystyle\begin{gathered}a_{0}=1,\quad b_{01}=0,\\ a_{211}=\frac{\sigma_{xx}}{2\rho\left(T_{t}\right)^{2}},\quad a_{212}=\frac{\sigma_{xy}}{2\rho\left(T_{t}\right)^{2}},\quad a_{213}=\frac{\sigma_{xz}}{2\rho\left(T_{t}\right)^{2}},\\ a_{212}=\frac{\sigma_{xy}}{\rho\left(T_{t}\right)^{2}},\quad a_{213}=\frac{\sigma_{xz}}{\rho\left(T_{t}\right)^{2}},\quad a_{223}=\frac{\sigma_{yz}}{\rho\left(T_{t}\right)^{2}},\end{gathered} (57)

and

a11=qt,xρ(Tt)2qr,xρTt(Tr),\displaystyle a_{11}=-\frac{q_{t,x}}{\rho\left(T_{t}\right)^{2}}-\frac{q_{r,x}}{\rho T_{t}\left(T_{r}\right)}, (58)
a12=qt,yρ(Tt)2qr,yρTt(Tr),\displaystyle a_{12}=-\frac{q_{t,y}}{\rho\left(T_{t}\right)^{2}}-\frac{q_{r,y}}{\rho T_{t}\left(T_{r}\right)},
a13=q3,tρ(Tt)2q3,rρTt(Tr),\displaystyle a_{13}=-\frac{q_{3,t}}{\rho\left(T_{t}\right)^{2}}-\frac{q_{3,r}}{\rho T_{t}\left(T_{r}\right)},
a31=qt,x5ρ(Tt)3,a32=qt,y5ρ(Tt)3,a33=q3,t5ρ(Tt)3,\displaystyle a_{31}=\frac{q_{t,x}}{5\rho\left(T_{t}\right)^{3}},\quad a_{32}=\frac{q_{t,y}}{5\rho\left(T_{t}\right)^{3}},\quad a_{33}=\frac{q_{3,t}}{5\rho\left(T_{t}\right)^{3}}, (59)
b11=qr,xdr2(Tr)2ρTt,b12=qr,ydr2(Tr)2ρTt,b13=q3,rdr2(Tr)2ρTt.\displaystyle b_{11}=\frac{q_{r,x}}{\frac{d_{r}}{2}\left(T_{r}\right)^{2}\rho T_{t}},\quad b_{12}=\frac{q_{r,y}}{\frac{d_{r}}{2}\left(T_{r}\right)^{2}\rho T_{t}},\quad b_{13}=\frac{q_{3,r}}{\frac{d_{r}}{2}\left(T_{r}\right)^{2}\rho T_{t}}. (60)

Finally, the expression of distribution functions fT(t,𝒙,𝝃,Ir)f^{T}(t,\bm{x},\bm{\xi},I_{r}) is given as

fT(t,𝐱,ξ,Ir)=f0eqfIreq[1+𝝈𝐜𝐜2ρ(Tt)2+𝐪t𝐜ρ(Tt)2(c25Tt1)+𝐪r𝐜dr2ρTtTr(IrTrdr2)].f^{T}\left(t,\mathbf{x},\xi,I_{r}\right)=f_{0}^{eq}f_{I_{r}}^{eq}\left[1+\frac{\bm{\sigma}\cdot\mathbf{cc}}{2\rho\left(T_{t}\right)^{2}}+\frac{\mathbf{q}_{t}\cdot\mathbf{c}}{\rho\left(T_{t}\right)^{2}}\left(\frac{c^{2}}{5T_{t}}-1\right)+\frac{\mathbf{q}_{r}\cdot\mathbf{c}}{\frac{d_{r}}{2}\rho T_{t}T_{r}}\left(\frac{I_{r}}{T_{r}}-\frac{d_{r}}{2}\right)\right]. (61)

Appendix C Integration coefficients related to the equilibrium state

It should be noted that the coefficients of ξnk>0eq\left\langle\xi_{n}^{k}\right\rangle_{>0}^{eq} and ξnk<0eq\left\langle\xi_{n}^{k}\right\rangle_{<0}^{eq} have different expressions. The parameters of ξnk>0eq\left\langle\xi_{n}^{k}\right\rangle_{>0}^{eq} and ξnk<0eq\left\langle\xi_{n}^{k}\right\rangle_{<0}^{eq} could be given as

ξn0>0eq=12[1+erf(λtLunL)],\left\langle\xi_{n}^{0}\right\rangle_{>0}^{eq}=\frac{1}{2}\left[1+\operatorname{erf}\left(\sqrt{\lambda_{t}^{L}}u_{n}^{L}\right)\right], (62)
ξn1>0eq=unLξn0>0eq+12eλtL(unL)2λtLπ,\left\langle\xi_{n}^{1}\right\rangle_{>0}^{eq}=u_{n}^{L}\left\langle\xi_{n}^{0}\right\rangle_{>0}^{eq}+\frac{1}{2}\frac{e^{-\lambda_{t}^{L}\left(u_{n}^{L}\right)^{2}}}{\sqrt{\lambda_{t}^{L}\pi}}, (63)
ξnk+2>0eq=unLξnk+1>0e+k+12λtLξnk>0eq,k=0,1,2,,\left\langle\xi_{n}^{k+2}\right\rangle_{>0}^{eq}=u_{n}^{L}\left\langle\xi_{n}^{k+1}\right\rangle_{>0}^{e}+\frac{k+1}{2\lambda_{t}^{L}}\left\langle\xi_{n}^{k}\right\rangle_{>0}^{eq},\quad k=0,1,2,\cdots, (64)

and

ξn0<0eq=12erfc(λtRunR),\left\langle\xi_{n}^{0}\right\rangle_{<0}^{eq}=\frac{1}{2}\operatorname{erfc}\left(\sqrt{\lambda_{t}^{R}}u_{n}^{R}\right), (65)
ξn1<0eq=unRξn0<0eq12eλtR(unR)2unRπ,\left\langle\xi_{n}^{1}\right\rangle_{<0}^{eq}=u_{n}^{R}\left\langle\xi_{n}^{0}\right\rangle_{<0}^{eq}-\frac{1}{2}\frac{e^{-\lambda_{t}^{R}\left(u_{n}^{R}\right)^{2}}}{\sqrt{u_{n}^{R}\pi}}, (66)
ξnk+2<0eq=unRξnk+1<0eq+k+12λtRξnk<0eq,k=0,1,2,.\left\langle\xi_{n}^{k+2}\right\rangle_{<0}^{eq}=u_{n}^{R}\left\langle\xi_{n}^{k+1}\right\rangle_{<0}^{eq}+\frac{k+1}{2\lambda_{t}^{R}}\left\langle\xi_{n}^{k}\right\rangle_{<0}^{eq},\quad k=0,1,2,\cdots. (67)

Following the binomial theory, part of even order of integration parameters cτkeq\left\langle c_{\tau}^{k}\right\rangle^{eq} and cskeq\left\langle c_{s}^{k}\right\rangle^{eq} could be computed as

cτ0eq=cs0eq=1,\left\langle c_{\tau}^{0}\right\rangle^{eq}=\left\langle c_{s}^{0}\right\rangle^{eq}=1, (68)
cτ2eq=cs2eq=12λt,\left\langle c_{\tau}^{2}\right\rangle^{eq}=\left\langle c_{s}^{2}\right\rangle^{eq}=\frac{1}{2\lambda_{t}}, (69)
cτ4eq=cs4eq=34λt2,\left\langle c_{\tau}^{4}\right\rangle^{eq}=\left\langle c_{s}^{4}\right\rangle^{eq}=\frac{3}{4\lambda_{t}^{2}}, (70)
cτ6eq=cs6eq=158λt3,\left\langle c_{\tau}^{6}\right\rangle^{eq}=\left\langle c_{s}^{6}\right\rangle^{eq}=\frac{15}{8\lambda_{t}^{3}}, (71)
cτ8eq=cs8eq=10516λt4.\left\langle c_{\tau}^{8}\right\rangle^{eq}=\left\langle c_{s}^{8}\right\rangle^{eq}=\frac{105}{16\lambda_{t}^{4}}. (72)

When the order of kk is odd, the moment integrals of cτkeq\left\langle c_{\tau}^{k}\right\rangle^{eq} and cskeq\left\langle c_{s}^{k}\right\rangle^{eq} are equal to zero.

References

  • Ivanov and Gimelshein [1998] M. S. Ivanov, S. F. Gimelshein, Computational Hypersonic Rarefied Flows, Annual Review of Fluid Mechanics 30 (1998) 469–505.
  • Blanchard et al. [2012] R. C. Blanchard, R. G. Wilmoth, J. N. Moss, Aerodynamic Flight Measurements and Rarefied-Flow Simulations of Mars Entry Vehicles, Journal of Spacecraft and Rockets (2012).
  • Gad-el Hak [2001] M. Gad-el Hak, Flow physics in MEMS, Mécanique & Industries 2 (2001) 313–341.
  • Klinkenberg [1941] J. Klinkenberg, The permeability of porous media to liquids and gases, Am. Petrol. Inst., Drilling and Production Practice 2 (1941) 200–213.
  • Darabi et al. [2012] H. Darabi, A. Ettehad, F. Javadpour, K. Sepehrnoori, Gas flow in ultra-tight shale strata, Journal of Fluid Mechanics 710 (2012) 641–658. Publisher: Cambridge University Press.
  • Jiang et al. [2019] D. Jiang, M. Mao, J. Li, X. Deng, An implicit parallel UGKS solver for flows covering various regimes, Advances in Aerodynamics 1 (2019) 8.
  • Goldstein et al. [1989] D. Goldstein, B. Sturtevant, J. E. Broadwell, Investigations of the motion of discrete-velocity gases, in: Progress in Astronautics and Aeronautics, vol. 118, AIAA, Washington, 1989.
  • Bird [1994] G. Bird, Molecular Gas Dynamics and the Direct Simulation of Gas Flows, Oxford University Press, 1994.
  • Neumann et al. [2014] P. Neumann, W. Eckhardt, H.-J. Bungartz, Hybrid molecular–continuum methods: From prototypes to coupling software, Computers & Mathematics with Applications 67 (2014) 272–281.
  • Xu and Huang [2010] K. Xu, J.-C. Huang, A unified gas-kinetic scheme for continuum and rarefied flows, Journal of Computational Physics 229 (2010) 7747–7764.
  • Guo et al. [2013] Z. Guo, K. Xu, R. Wang, Discrete unified gas kinetic scheme for all Knudsen number flows: Low-speed isothermal case, Physical Review E 88 (2013) 033305. Publisher: American Physical Society.
  • Zhu et al. [2019] Y. Zhu, C. Liu, C. Zhong, K. Xu, Unified gas-kinetic wave-particle methods. II. Multiscale simulation on unstructured mesh, Physics of Fluids 31 (2019) 067105.
  • Fei [2023] F. Fei, A time-relaxed Monte Carlo method preserving the Navier-Stokes asymptotics, Journal of Computational Physics 486 (2023) 112128.
  • Wang et al. [2018] P. Wang, M. T. Ho, L. Wu, Z. Guo, Y. Zhang, A comparative study of discrete velocity methods for low-speed rarefied gas flows, Computers & Fluids 161 (2018) 33–46.
  • Larsen [1983] E. W. Larsen, On Numerical Solutions of Transport Problems in the Diffusion Limit, Nuclear Science and Engineering 83 (1983) 90–99.
  • Valougeorgis and Naris [2003] D. Valougeorgis, S. Naris, Acceleration schemes of the discrete velocity method: Gaseous flows in rectangular microchannels, SIAM J. Sci. Comput. 25 (2003) 534–552.
  • Szalmás and Valougeorgis [2010] L. Szalmás, D. Valougeorgis, A fast iterative model for discrete velocity calculations on triangular grids, Journal of Computational Physics 229 (2010) 4315–4326.
  • Degond et al. [2011] P. Degond, G. Dimarco, L. Pareschi, The moment-guided Monte Carlo method, International Journal for Numerical Methods in Fluids 67 (2011) 189–213.
  • Chacón et al. [2017] L. Chacón, G. Chen, D. Knoll, C. Newman, H. Park, W. Taitano, J. Willert, G. Womeldorff, Multiscale high-order/low-order (HOLO) algorithms and applications, Journal of Computational Physics 330 (2017) 21–45.
  • Su et al. [2020a] W. Su, L. Zhu, P. Wang, Y. Zhang, L. Wu, Can we find steady-state solutions to multiscale rarefied gas flows within dozens of iterations?, Journal of Computational Physics 407 (2020a) 109245.
  • Su et al. [2020b] W. Su, M. T. Ho, Y. Zhang, L. Wu, GSIS: An efficient and accurate numerical method to obtain the apparent gas permeability of porous media, Computers & Fluids 206 (2020b) 104576.
  • Su et al. [2021] W. Su, Y. Zhang, L. Wu, Multiscale simulation of molecular gas flows by the general synthetic iterative scheme, Computer Methods in Applied Mechanics and Engineering 373 (2021) 113548.
  • Zeng et al. [2023a] J. Zeng, R. Yuan, Y. Zhang, Q. Li, L. Wu, General synthetic iterative scheme for polyatomic rarefied gas flows, Computers & Fluids 265 (2023a) 105998.
  • Zeng et al. [2023b] J. Zeng, W. Su, L. Wu, General Synthetic Iterative Scheme for Unsteady Rarefied Gas Flows, Communications in Computational Physics 34 (2023b) 173–207.
  • Luo et al. [2023] L. Luo, Q. Li, L. Wu, Boosting the convergence of low-variance DSMC by GSIS, Advances in Aerodynamics 5 (2023) 10.
  • Su et al. [2020] W. Su, L. Zhu, L. Wu, Fast Convergence and Asymptotic Preserving of the General Synthetic Iterative Scheme, SIAM Journal on Scientific Computing 42 (2020) B1517–B1540.
  • Zhu et al. [2021] L. Zhu, X. Pi, W. Su, Z.-H. Li, Y. Zhang, L. Wu, General synthetic iterative scheme for nonlinear gas kinetic simulation of multi-scale rarefied gas flows, Journal of Computational Physics 430 (2021) 110091.
  • Maxwell [1997] J. C. Maxwell, VII. On stresses in rarified gases arising from inequalities of temperature, Philosophical Transactions of the Royal Society of London 170 (1997) 231–256. Publisher: Royal Society.
  • Sharipov [2011] F. Sharipov, Data on the Velocity Slip and Temperature Jump on a Gas-Solid Interface, Journal of Physical and Chemical Reference Data 40 (2011) 023101.
  • Smoluchowski von Smolan [1898] M. Smoluchowski von Smolan, Ueber Wärmeleitung in verdünnten Gasen, Annalen der Physik 300 (1898) 101–130.
  • Ohwada et al. [1989] T. Ohwada, Y. Sone, K. Aoki, Numerical analysis of the Poiseuille and thermal transpiration flows between two parallel plates on the basis of the Boltzmann equation for hard‐sphere molecules, Physics of Fluids A: Fluid Dynamics 1 (1989) 2042–2049. Publisher: American Institute of Physics.
  • Sone [1972] Y. Sone, A Flow Induced by Thermal Stress in Rarefied Gas, Journal of the Physical Society of Japan 33 (1972) 232–236. Publisher: The Physical Society of Japan.
  • Sone et al. [2000] Y. Sone, C. Bardos, F. Golse, H. Sugimoto, Asymptotic theory of the Boltzmann system, for a steady flow of a slightly rarefied gas with a finite Mach number: General theory, European Journal of Mechanics - B/Fluids 19 (2000) 325–360.
  • Zhao et al. [2023] G. Zhao, C. Zhong, S. Liu, Y. Wang, C. Zhuo, Application of Gas-Kinetic Scheme for Continuum and Near-Continuum Flow on Unstructured Mesh, International Journal of Computational Fluid Dynamics 0 (2023) 1–23.
  • Grad [1949] H. Grad, On the kinetic theory of rarefied gases, Communications on Pure and Applied Mathematics 2 (1949) 331–407.
  • Gu and Emerson [2007] X. Gu, D. Emerson, A computational strategy for the regularized 13 moment equations with enhanced wall-boundary conditions, Journal of Computational Physics 225 (2007) 263–283.
  • Gu and Emerson [2009] X. J. Gu, D. R. Emerson, A high-order moment approach for capturing non-equilibrium phenomena in the transition regime, Journal of Fluid Mechanics 636 (2009) 177–216.
  • Rykov [1976] V. A. Rykov, A model kinetic equation for a gas with rotational degrees of freedom, Fluid Dynamics 10 (1976) 959–966.
  • Li et al. [2021] Q. Li, J. Zeng, W. Su, L. Wu, Uncertainty quantification in rarefied dynamics of molecular gas: rate effect of thermal relaxation, Journal of Fluid Mechanics 917 (2021) A58.
  • Blazek [2015] J. Blazek, Computational Fluid Dynamics: Principles and Applications, Butterworth-Heinemann, 2015.
  • Chapman and Cowling [1962] S. Chapman, T. G. Cowling, The Mathematical Theory of Non-Uniform Gases, American Journal of Physics 30 (1962) 389–389.
  • Chen et al. [2020] Y. Chen, Y. Zhu, K. Xu, A three-dimensional unified gas-kinetic wave-particle solver for flow computation in all regimes, Physics of Fluids 32 (2020) 096108.
  • Fan et al. [2023] G. Fan, W. Zhao, S. Yao, Z. Jiang, W. Chen, The implementation of the three-dimensional unified gas-kinetic wave-particle method on multiple graphics processing units, Physics of Fluids 35 (2023) 086108.
  • Zhang et al. [2023] Y. Zhang, J. Zeng, R. Yuan, W. Liu, Q. Li, L. Wu, Efficient parallel solver for high-speed rarefied gas flow using GSIS, 2023. ArXiv:2310.18916 [physics].
  • Venkatakrishnan [1995] V. Venkatakrishnan, Convergence to Steady State Solutions of the Euler Equations on Unstructured Grids with Limiters, Journal of Computational Physics 118 (1995) 120–130.
  • Yuan et al. [2021] R. Yuan, S. Liu, C. Zhong, A multi-prediction implicit scheme for steady state solutions of gas flow in all flow regimes, Communications in Nonlinear Science and Numerical Simulation 92 (2021) 105470.
  • Liu et al. [2020] C. Liu, Y. Zhu, K. Xu, Unified gas-kinetic wave-particle methods I: Continuum and rarefied gas flow, Journal of Computational Physics 401 (2020) 108977.
  • Titarev et al. [2014] V. Titarev, M. Dumbser, S. Utyuzhnikov, Construction and comparison of parallel implicit kinetic solvers in three spatial dimensions, Journal of Computational Physics 256 (2014) 17–33.
  • Varoutis et al. [2009] S. Varoutis, D. Valougeorgis, F. Sharipov, Simulation of gas flow through tubes of finite length over the whole range of rarefaction for various pressure drop ratios, Journal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films 27 (2009) 1377–1391.