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

On Well-posed Boundary Conditions for the Linear Non-homogeneous Moment Equations in Half-space

Ruo Li CAPT, LMAM & School of Mathematical Sciences, Peking University, Beijing 100871, China, email: rli@math.pku.edu.cn.    Yichen Yang School of Mathematical Sciences, Peking University, Beijing 100871, China, email: yichenyang@pku.edu.cn.
Abstract

We propose a necessary and sufficient condition for the well-posedness of the linear non-homogeneous Grad moment equations in half-space. The Grad moment system is based on Hermite expansion and regarded as an efficient reduction model of the Boltzmann equation. At a solid wall, the moment equations are commonly equipped with a Maxwell-type boundary condition named the Grad boundary condition. We point out that the Grad boundary condition is unstable for the non-homogeneous half-space problem. Thanks to the proposed criteria, we verify the well-posedness of a class of modified boundary conditions. The technique to make sure the existence and uniqueness mainly includes a well-designed preliminary simultaneous transformation of the coefficient matrices, and Kreiss’ procedure about the linear boundary value problem with characteristic boundaries. The stability is established by a weighted estimate. At the same time, we obtain the analytical expressions of the solution, which may help solve the half-space problem efficiently.

AMS subject classifications 34B40; 35Q35; 76P05; 82B40

Keywords Half-space problem, moment method, well-posed boundary condition, non-homogeneous equations

1 Introduction

Half-space problems are at the center of our understanding of the kinetic boundary layer [29]. The relevant study helps prescribe slip boundary conditions for the fluid-dynamic-type equations [29] and results in the interface coupling condition between kinetic and hydrodynamic equations [1, 21, 11]. For the Boltzmann equation, the theory of linear half-space problems has been well-developed (cf. [2] and references therein).

In this paper, we focus on the linear steady non-homogeneous equations in half-space:

𝑨dW(y)dy=𝑸W(y)+h(y),y[0,+),\displaystyle\boldsymbol{A}\dfrac{\mathrm{d}{W(y)}}{\mathrm{d}{y}}=-\boldsymbol{Q}W(y)+h(y),\quad y\in[0,+\infty), (1.1)
𝑩W(0)=g,\displaystyle\boldsymbol{B}W(0)=g, (1.2)

where eayW(y)L2(+;L2(N))e^{ay}W(y)\in L^{2}(\mathbb{R}_{+};L^{2}(\mathbb{R}^{N})) and eayh(y)L2(+;L2(N))e^{ay}h(y)\in L^{2}(\mathbb{R}_{+};L^{2}(\mathbb{R}^{N})) for some constant a>0a>0. Here 𝑨,𝑸,𝑩\boldsymbol{A},\boldsymbol{Q},\boldsymbol{B} are given N×NN\times N matrices, and gg is an N×1N\times 1 vector, all with constant coefficients. For the considered Grad moment equations, the matrix 𝑨\boldsymbol{A} is symmetric, and 𝑸\boldsymbol{Q} is symmetric positive semi-definite, which individually has a special block structure, as the article will show later.

We can realize the importance of the half-space problem (1.1) from two aspects. First, the moment equations proposed by Grad [14] are regarded as an efficient reduction model [7] of the Boltzmann equation. As an extension of the celebrated Navier-Stokes equations, the moment equations have gained much more attention in recent years [31, 6, 18, 9]. The half-space problem (1.1) could arise from the asymptotic analysis of the moment equations. This problem plays a crucial role in understanding the boundary layer of the moment equations [30, 12]. Second, the moment system (1.1) may serve as an efficient numerical solver of kinetic equations. There are many approximatively analytical and numerical methods to solve half-space kinetic layer equations. Some most famous analytical methods are essentially low-order moment methods, including the Maxwell method [10], Loyalka’s method [24, 25], and the half-range moment method [15]. The arbitrary order moment equations are also widely used to resolve the layer problems [30, 16, 12], which can give formal analytical solutions.

Despite the broad application range, the well-posedness of the moment system with the Grad boundary condition [14] is doubtable. For the linear initial boundary value problem, [28] has pointed out that the Grad boundary condition is unstable. Their method can not apply to the half-space problem directly. We have studied the half-space problem for the linear homogeneous moment equations and proposed several solvability conditions in an earlier paper [23]. The object of this work is to study the well-posed criteria of the linear non-homogeneous Grad moment equations in half-space.

In general, the concept of well-posedness contains three points. First, the solution exists. Second, the solution is unique. Third, the solution is controlled by the boundary data and non-homogeneous term. In the homogeneous case, there is no need to consider the variants of the non-homogeneous item. Due to this significant difference, the non-homogeneous case is not a trivial corollary of the homogeneous situation.

Our method is essentially basic linear algebra. Because the matrices 𝑨\boldsymbol{A} and 𝑸\boldsymbol{Q} may both have zero eigenvalues, the problem has the characteristic boundary, where the characteristic variables have a vanishing speed at the boundary [17]. We first make a simultaneous transformation of the coefficient matrices to reduce the original problem. Then following the characteristic analysis of ODEs, we explicitly write the general solution to the half-space problem. By analysing the stability of the solution, we finally achieve well-posed criteria for the linear non-homogeneous moment system in half-space. The method is similar as Kreiss’ procedure [19] to study linear hyperbolic systems. We find that the Grad boundary condition does not obey the proposed well-posed rules because of the instability of the non-homogeneous term. After minor modification, we can obtain well-posed boundary conditions for the moment system (1.1) in half-space.

A discrete system similar to (1.1) comes from the discrete velocity method (DVM) of the Boltzmann equation, where the matrix 𝑨\boldsymbol{A} is diagonal. The solvability of these discrete equations has been exhaustively studied by Bernhoff [3, 4, 5]. These results can apply to the system (1.1) to obtain the existence and uniqueness of the solution. In comparison, our method utilizes the specific structure of the Grad moment equations, which gives a more subtle description of the solution. We also find the additional stability condition besides the solvability condition. The general abstract theory about the linear boundary value problem can be found in [20, 19, 26, 27].

The paper is arranged as follows: in Section 2, we briefly introduce the linearized Boltzmann equation and the Grad moment equations. Then we state the well-posed conditions without detailed proof. In Section 3, we illustrate the instability of the Grad boundary condition by simple examples and discuss its well-posed modification. In Section 4, we complete the proof of well-posedness. The paper ends with a conclusion.

2 Basic Equations and Main Results

2.1 Basic Equations

Around the equilibrium states, the rarefied gas can be described by the linearized Boltzmann equation (LBE). For single-species monatomic molecules, we consider the LBE with the Maxwell boundary condition

ft+ξdfxd=[f],f=f(t,𝒙,𝝃),𝒙Ω3,𝝃3,\displaystyle\dfrac{\partial{f}}{\partial{t}}+\xi_{d}\dfrac{\partial{f}}{\partial{x_{d}}}=\mathcal{L}[f],\quad f=f(t,\boldsymbol{x},\boldsymbol{\xi}),\ \boldsymbol{x}\in\Omega\subset\mathbb{R}^{3},\ \boldsymbol{\xi}\in\mathbb{R}^{3}, (2.1a)
f(t,𝒙,𝝃)=χfw(t,𝒙,𝝃)+(1χ)f(t,𝒙,𝝃),𝒙Ω,(𝝃𝒖w)𝒏<0,\displaystyle f(t,\boldsymbol{x},\boldsymbol{\xi})=\chi f^{w}(t,\boldsymbol{x},\boldsymbol{\xi})+(1-\chi)f(t,\boldsymbol{x},\boldsymbol{\xi}^{*}),\quad\boldsymbol{x}\in\partial\Omega,\ (\boldsymbol{\xi}-\boldsymbol{u}^{w})\cdot\boldsymbol{n}<0, (2.1b)

where ff is the velocity distribution function, tt denoting the time, 𝒙=(x1,x2,x3)\boldsymbol{x}=(x_{1},x_{2},x_{3}) the spatial coordinates and 𝝃=(ξ1,ξ2,ξ3)\boldsymbol{\xi}=(\xi_{1},\xi_{2},\xi_{3}) the microscopic velocity. Here \mathcal{L} is a linear operator depicting the collision between gas molecules.

The Maxwell model assumes that the boundary is an impermeable wall with a unit normal vector 𝒏\boldsymbol{n} exiting the region, a given velocity 𝒖w\boldsymbol{u}^{w}, and a given temperature θw\theta^{w}. To avoid cumbersome details about the rotation invariance, we assume

Ω={𝒙3:x20},\Omega=\{\boldsymbol{x}\in\mathbb{R}^{3}:\ x_{2}\geq 0\},

with a fixed 𝒏=(0,1,0)\boldsymbol{n}=(0,-1,0). The reflection at the wall is divided into a sum of χ\chi portion of the specular reflection and 1χ1-\chi portion of the diffuse reflection, where χ[0,1]\chi\in[0,1] is the tangential momentum accommodation coefficient. To consider the steady layer equations, we further assume

𝒖w𝒏=0,\boldsymbol{u}^{w}\cdot\boldsymbol{n}=0,

then the velocity from specular reflection is

𝝃=𝝃2[(𝝃𝒖w)𝒏]𝒏=(ξ1,ξ2,ξ3),\boldsymbol{\xi}^{*}=\boldsymbol{\xi}-2[(\boldsymbol{\xi}-\boldsymbol{u}^{w})\cdot\boldsymbol{n}]\boldsymbol{n}=(\xi_{1},-\xi_{2},\xi_{3}),

and

fw(t,𝒙,𝝃)=(𝝃)(ρw+𝒖w𝝃+θw|𝝃|232).f^{w}(t,\boldsymbol{x},\boldsymbol{\xi})=\mathcal{M}(\boldsymbol{\xi})\left(\rho^{w}+\boldsymbol{u}^{w}\cdot\boldsymbol{\xi}+\theta^{w}\frac{|\boldsymbol{\xi}|^{2}-3}{2}\right). (2.2)

The reference distribution \mathcal{M} is a Maxwellian at rest

(𝝃)=1(2π)3/2exp(|𝝃|22),\mathcal{M}(\boldsymbol{\xi})=\frac{1}{(2\pi)^{3/2}}\exp\left(-\frac{|\boldsymbol{\xi}|^{2}}{2}\right),

where ρw\rho^{w} is determined by the no mass flow condition at the wall

3(𝝃𝒖w)𝒏f(t,𝒙,𝝃)d𝝃=0,𝒙Ω.\int_{\mathbb{R}^{3}}\!\!(\boldsymbol{\xi}-\boldsymbol{u}^{w})\cdot\boldsymbol{n}f(t,\boldsymbol{x},\boldsymbol{\xi})\,\mathrm{d}\boldsymbol{\xi}=0,\quad\boldsymbol{x}\in\partial\Omega.

The Grad moment equations with linearized ansatz assume that the distribution function has a Hermite expansion, for a chosen integer M2M\geq 2, as

f(t,𝒙,𝝃)=(𝝃)|𝜶|Mw𝜶(t,𝒙)ϕ𝜶(𝝃),f(t,\boldsymbol{x},\boldsymbol{\xi})=\mathcal{M}(\boldsymbol{\xi})\sum_{|\boldsymbol{\alpha}|\leq M}w_{\boldsymbol{\alpha}}(t,\boldsymbol{x})\phi_{\boldsymbol{\alpha}}(\boldsymbol{\xi}), (2.3)

where 𝜶=(α1,α2,α3)3\boldsymbol{\alpha}=(\alpha_{1},\alpha_{2},\alpha_{3})\in\mathbb{N}^{3} is the multi-index, |𝜶|=α1+α2+α3|\boldsymbol{\alpha}|=\alpha_{1}+\alpha_{2}+\alpha_{3}. The orthonormal Hermite polynomial ϕ𝜶=ϕ𝜶(𝝃)\phi_{\boldsymbol{\alpha}}=\phi_{\boldsymbol{\alpha}}(\boldsymbol{\xi}) is defined [13] by ensuring

ϕ𝜶ϕ𝜷=δ𝜶,𝜷,ϕ𝟎=1,ϕ𝒆i=ξi,:=3d𝝃,\left\langle{\mathcal{M}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle=\delta_{\boldsymbol{\alpha},\boldsymbol{\beta}},\quad\phi_{\boldsymbol{0}}=1,\ \phi_{\boldsymbol{e}_{i}}=\xi_{i},\quad\left\langle{\cdot}\right\rangle:=\int_{\mathbb{R}^{3}}\!\!\cdot\,\mathrm{d}\boldsymbol{\xi},

where 𝒆i3\boldsymbol{e}_{i}\in\mathbb{N}^{3} only has the ii-th component being one. So the ansatz (2.3) gives

w𝜶=w𝜶(t,𝒙)=fϕ𝜶.w_{\boldsymbol{\alpha}}=w_{\boldsymbol{\alpha}}(t,\boldsymbol{x})=\left\langle{f\phi_{\boldsymbol{\alpha}}}\right\rangle.

Substituting the ansatz into the linearized Boltzmann equation and matching the coefficients before basis functions, we have the MM-th order moment equations

w𝜶t+ξdϕ𝜶ϕ𝜷w𝜷xd=[ϕ𝜷]ϕ𝜶w𝜷,|𝜶|M,\dfrac{\partial{w_{\boldsymbol{\alpha}}}}{\partial{t}}+\left\langle{\mathcal{M}\xi_{d}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle\dfrac{\partial{w_{\boldsymbol{\beta}}}}{\partial{x_{d}}}=\left\langle{\mathcal{L}[\mathcal{M}\phi_{\boldsymbol{\beta}}]\phi_{\boldsymbol{\alpha}}}\right\rangle w_{\boldsymbol{\beta}},\quad|\boldsymbol{\alpha}|\leq M, (2.4)

where Einstein’s convention is used to omit the summation notation about |𝜷|M.|\boldsymbol{\beta}|\leq M.

Intuitively, the moment variables w𝜶w_{\boldsymbol{\alpha}} can be related to the macroscopic variables such as the density, the macroscopic velocity, and the temperature. So (2.4) would contain the linearized Euler equations (around the equilibrium state given by \mathcal{M}). We can regard (2.4) as an extension of hydrodynamic equations [14].

It’s not difficult to check [22] that (2.4) is symmetric hyperbolic. More precisely, the coefficient matrices are all symmetric. To ensure the correct number of boundary conditions for the hyperbolic system [17], Grad [14] suggested testing the Maxwell boundary condition (2.1b) with odd polynomials (about the argument (𝝃𝒖w)𝒏=ξ2(\boldsymbol{\xi}-\boldsymbol{u}^{w})\cdot\boldsymbol{n}=-\xi_{2}) no higher than the MM-th degree.

The Grad boundary condition has some equivalent representations [14, 30, 8, 12]. To exhibit the continuity of fluxes at the boundary, we extract the factor ξ2\xi_{2} of these odd polynomials and write the Grad boundary condition in an equivalent form:

20+ξ2ϕ𝜶f(t,𝒙,𝝃)d𝝃\displaystyle\int_{\mathbb{R}^{2}}\!\!\int_{0}^{+\infty}\!\!\xi_{2}\phi_{\boldsymbol{\alpha}}f(t,\boldsymbol{x},\boldsymbol{\xi})\,\mathrm{d}\boldsymbol{\xi} =\displaystyle= χ20+ξ2ϕ𝜶fw(t,𝒙,𝝃)d𝝃+\displaystyle\chi\int_{\mathbb{R}^{2}}\!\!\int_{0}^{+\infty}\!\!\xi_{2}\phi_{\boldsymbol{\alpha}}f^{w}(t,\boldsymbol{x},\boldsymbol{\xi})\,\mathrm{d}\boldsymbol{\xi}+
(1χ)20+ξ2ϕ𝜶f(t,𝒙,𝝃)d𝝃,\displaystyle(1-\chi)\int_{\mathbb{R}^{2}}\!\!\int_{0}^{+\infty}\!\!\xi_{2}\phi_{\boldsymbol{\alpha}}f(t,\boldsymbol{x},\boldsymbol{\xi}^{*})\,\mathrm{d}\boldsymbol{\xi},

where α2\alpha_{2} is even and |𝜶|M1.|\boldsymbol{\alpha}|\leq M-1. Plugging the ansatz (2.3) into the above formula, we can utilize the even-odd parity of Hermite polynomials [23] to get

(1χ2)β2 oddξ2ϕ𝜶ϕ𝜷w𝜷=χ2β2 even|ξ2|ϕ𝜶ϕ𝜷(w𝜷b𝜷),\left(1-\frac{\chi}{2}\right)\sum_{\beta_{2}\text{\ odd}}\left\langle{\xi_{2}\mathcal{M}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle w_{\boldsymbol{\beta}}=-\frac{\chi}{2}\sum_{\beta_{2}\text{\ even}}\left\langle{|\xi_{2}|\mathcal{M}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle(w_{\boldsymbol{\beta}}-b_{\boldsymbol{\beta}}), (2.5)

where the entries b𝟎=ρw,b𝒆i=uiw,b2𝒆i=θw/2b_{\boldsymbol{0}}=\rho^{w},\ b_{\boldsymbol{e}_{i}}=u_{i}^{w},\ b_{2\boldsymbol{e}_{i}}=\theta^{w}/\sqrt{2} and otherwise b𝜶=0.b_{\boldsymbol{\alpha}}=0.

In conclusion, the MM-th order linear Grad moment equations (in the plane geometry) are (2.4) with the Grad boundary condition (2.5), where ρw\rho^{w} is determined by the no mass flow condition

w𝒆2u2w=0,atx2=0.w_{\boldsymbol{e}_{2}}-u_{2}^{w}=0,\quad\text{at}\ x_{2}=0.

Assume there is a boundary layer near Ω={x2=0}.\partial\Omega=\{x_{2}=0\}. Then we may introduce the fast variables, which vary dramatically in the normal direction of the wall and vanish outside the boundary layer. According to the asymptotic analysis [29] of (2.4), these fast variables may satisfy the linear steady moment equations in half-space:

ξ2ϕ𝜶ϕ𝜷dw𝜷dy=[ϕ𝜷]ϕ𝜶w𝜷+h𝜶,|𝜶|M,\left\langle{\mathcal{M}\xi_{2}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle\dfrac{\mathrm{d}{w_{\boldsymbol{\beta}}}}{\mathrm{d}{y}}=\left\langle{\mathcal{L}[\mathcal{M}\phi_{\boldsymbol{\beta}}]\phi_{\boldsymbol{\alpha}}}\right\rangle w_{\boldsymbol{\beta}}+h_{\boldsymbol{\alpha}},\quad|\boldsymbol{\alpha}|\leq M,

where w𝜷=w𝜷(y),y[0,+),w_{\boldsymbol{\beta}}=w_{\boldsymbol{\beta}}(y),\ y\in[0,+\infty), represents the fast variables, with w𝜷(+)=0w_{\boldsymbol{\beta}}(+\infty)=0. The non-homogeneous term h𝜶=h𝜶(y)h_{\boldsymbol{\alpha}}=h_{\boldsymbol{\alpha}}(y) is given, arising from the other contributions in the boundary layer.

The Grad boundary condition becomes

(1χ2)β2 oddξ2ϕ𝜶ϕ𝜷(w𝜷+w¯𝜷)=χ2β2 even|ξ2|ϕ𝜶ϕ𝜷(w𝜷+w¯𝜷b𝜷),\left(1-\frac{\chi}{2}\right)\sum_{\beta_{2}\text{\ odd}}\left\langle{\xi_{2}\mathcal{M}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle(w_{\boldsymbol{\beta}}+\bar{w}_{\boldsymbol{\beta}})=-\frac{\chi}{2}\sum_{\beta_{2}\text{\ even}}\left\langle{|\xi_{2}|\mathcal{M}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle(w_{\boldsymbol{\beta}}+\bar{w}_{\boldsymbol{\beta}}-b_{\boldsymbol{\beta}}),

where α2\alpha_{2} is even with |𝜶|M1,|\boldsymbol{\alpha}|\leq M-1, and w¯𝜷\bar{w}_{\boldsymbol{\beta}} is given by the bulk flow outside the boundary layer.

2.2 Main Results

For the classical linearized Boltzmann operator [10], we shall write the Grad moment equations in half-space as an abstract form

𝑨dWdy=𝑸W+h,W=W(y),y[0,+),\displaystyle\boldsymbol{A}\dfrac{\mathrm{d}{W}}{\mathrm{d}{y}}=-\boldsymbol{Q}W+h,\quad W=W(y),\ y\in[0,+\infty), (2.6)
W(+)=0,\displaystyle W(+\infty)=0,

where 𝑨=𝑨TN×N,𝑸0\boldsymbol{A}=\boldsymbol{A}^{T}\in\mathbb{R}^{N\times N},\ \boldsymbol{Q}\geq 0 and WN.W\in\mathbb{R}^{N}. Here N=#{𝜶3:|𝜶|M}N=\#\{\boldsymbol{\alpha}\in\mathbb{N}^{3}:\ |\boldsymbol{\alpha}|\leq M\} and we assume

Null(𝑨)Null(𝑸)={0}.\mathrm{Null}(\boldsymbol{A})\cap\mathrm{Null}(\boldsymbol{Q})=\{0\}. (2.7)

In this paper, we will discuss the well-posed conditions of (2.6) with h0,h\neq 0, i.e., which boundary condition should be prescribed at y=0y=0 to ensure the well-posedness of (2.6). Roughly, we only need to write general solutions for the ODEs and analyze their stability.

If we can solve the generalized eigenvalue problem of (𝑨,𝑸)(\boldsymbol{A},\boldsymbol{Q}), we may deal with the characteristic equations

λidvidy=vi+hi,vi(+)=0,\lambda_{i}\dfrac{\mathrm{d}{v_{i}}}{\mathrm{d}{y}}=-v_{i}+h_{i},\quad v_{i}(+\infty)=0,

where λi{}\lambda_{i}\in\mathbb{R}\cup\{\infty\} and hi=hi(y)h_{i}=h_{i}(y) is a given 1D function.

If λi=0,\lambda_{i}=0, we have vi=hiv_{i}=h_{i} and vi(0)=hi(0)v_{i}(0)=h_{i}(0), which means that there is no need to prescribe extra boundary conditions at y=0.y=0. If λ<0\lambda<0, we have

vi(y)=eλi1yvi(0)+λi10yeλi1(ys)hi(s)ds.v_{i}(y)=e^{-\lambda_{i}^{-1}y}v_{i}(0)+\lambda_{i}^{-1}\int_{0}^{y}\!\!e^{-\lambda_{i}^{-1}(y-s)}h_{i}(s)\,\mathrm{d}s.

So to make sure vi(+)=0,v_{i}(+\infty)=0, there must be

vi(0)=λi10+eλi1shi(s)ds,v_{i}(0)=-\lambda_{i}^{-1}\int_{0}^{+\infty}\!\!e^{\lambda_{i}^{-1}s}h_{i}(s)\,\mathrm{d}s,

if the integral exists. If λi>0,\lambda_{i}>0, from the above formula we can see that vi(0)v_{i}(0) can be arbitrarily given when

limy+0yeλi1(ys)hi(s)ds=0.\lim_{y\rightarrow+\infty}\int_{0}^{y}\!\!e^{-\lambda_{i}^{-1}(y-s)}h_{i}(s)\,\mathrm{d}s=0.

Finally, if λi=\lambda_{i}=\infty, formally viv_{i} should be a constant and must be zero.

Our results are based on these simple observations. Due to the structure of coefficient matrices, the role of the generalized eigenvalue decomposition can be replaced by an interpretable simultaneous transformation of the matrices 𝑨\boldsymbol{A} and 𝑸\boldsymbol{Q}. We note that a similar transformation appears in [3, 5] in the language of projection operators.

For this purpose, we assume that 𝑮N×p\boldsymbol{G}\in\mathbb{R}^{N\times p} is the orthonormal basis matrix of Null(𝑸),\mathrm{Null}(\boldsymbol{Q}), and 𝑿p×r\boldsymbol{X}\in\mathbb{R}^{p\times r} is the orthonormal basis matrix of Null(𝑮T𝑨𝑮),\mathrm{Null}(\boldsymbol{G}^{T}\boldsymbol{A}\boldsymbol{G}), where

p=dimNull(𝑸),r=dimNull(𝑮T𝑨𝑮).p=\mathrm{dim}\ \mathrm{Null}(\boldsymbol{Q}),\quad r=\mathrm{dim}\ \mathrm{Null}(\boldsymbol{G}^{T}\boldsymbol{A}\boldsymbol{G}).

Since Null(𝑨)Null(𝑸)={0}\mathrm{Null}(\boldsymbol{A})\cap\mathrm{Null}(\boldsymbol{Q})=\{0\}, we have

rank(𝑨𝑮)=rank(𝑮)=p.\mathrm{rank}(\boldsymbol{A}\boldsymbol{G})=\mathrm{rank}(\boldsymbol{G})=p.

So we can let 𝑽1=𝑮𝑿N×r\boldsymbol{V}_{1}=\boldsymbol{G}\boldsymbol{X}\in\mathbb{R}^{N\times r} and construct 𝑽2N×p\boldsymbol{V}_{2}\in\mathbb{R}^{N\times p} by a Gram-Schmidt orthogonalization of 𝑨𝑮\boldsymbol{A}\boldsymbol{G}. Then

span{𝑽2}=span{𝑨𝑮}\mathrm{span}\{\boldsymbol{V}_{2}\}=\mathrm{span}\{\boldsymbol{A}\boldsymbol{G}\}

and 𝑽2T𝑽1=𝟎.\boldsymbol{V}_{2}^{T}\boldsymbol{V}_{1}=\boldsymbol{0}. Let 𝑽3\boldsymbol{V}_{3} be the orthogonal complement of [𝑽1,𝑽2][\boldsymbol{V}_{1},\boldsymbol{V}_{2}]. So 𝑽=[𝑽1,𝑽2,𝑽3]\boldsymbol{V}=[\boldsymbol{V}_{1},\boldsymbol{V}_{2},\boldsymbol{V}_{3}] is orthogonal.

For technical reasons, we further define 𝑼=[𝑼1,𝑼2,𝑼3]\boldsymbol{U}=[\boldsymbol{U}_{1},\boldsymbol{U}_{2},\boldsymbol{U}_{3}], where

𝑼1=𝑮N×p,𝑼2=𝑨𝑮𝑿N×r,𝑼3=𝑽3.\boldsymbol{U}_{1}=\boldsymbol{G}\in\mathbb{R}^{N\times p},\ \boldsymbol{U}_{2}=\boldsymbol{A}\boldsymbol{G}\boldsymbol{X}\in\mathbb{R}^{N\times r},\ \boldsymbol{U}_{3}=\boldsymbol{V}_{3}.

We claim that 𝑼\boldsymbol{U} is invertible. Since 𝑽3T𝑽2=𝟎,\boldsymbol{V}_{3}^{T}\boldsymbol{V}_{2}=\boldsymbol{0}, we have 𝑽3T𝑼2=𝟎\boldsymbol{V}_{3}^{T}\boldsymbol{U}_{2}=\boldsymbol{0}. To make 𝑼\boldsymbol{U} invertible, it’s enough to show that

span{𝑽3}span{𝑮}={0}.\mathrm{span}\{\boldsymbol{V}_{3}\}\cap\mathrm{span}\{\boldsymbol{G}\}=\{0\}. (2.8)

In fact, if 𝑮c1=𝑽3c2\boldsymbol{G}c_{1}=\boldsymbol{V}_{3}c_{2} for some c1pc_{1}\in\mathbb{R}^{p} and c2(Npr)c_{2}\in\mathbb{R}^{(N-p-r)}, then (𝑮c1)T𝑨𝑮=0(\boldsymbol{G}c_{1})^{T}\boldsymbol{A}\boldsymbol{G}=0 since 𝑽3T𝑽2=𝟎\boldsymbol{V}_{3}^{T}\boldsymbol{V}_{2}=\boldsymbol{0}. So c1span{𝑿}c_{1}\in\mathrm{span}\{\boldsymbol{X}\} and 𝑽3c2span{𝑮𝑿}=span{𝑽1}\boldsymbol{V}_{3}c_{2}\in\mathrm{span}\{\boldsymbol{G}\boldsymbol{X}\}=\mathrm{span}\{\boldsymbol{V}_{1}\}. Thus, from 𝑽3T𝑽1=𝟎\boldsymbol{V}_{3}^{T}\boldsymbol{V}_{1}=\boldsymbol{0} we have c2=0c_{2}=0, which implies (2.8).

Lemma 2.1.

The transformed matrices 𝐀ij=𝐔iT𝐀𝐕j\boldsymbol{A}_{ij}=\boldsymbol{U}_{i}^{T}\boldsymbol{A}\boldsymbol{V}_{j} and 𝐐ij=𝐔iT𝐐𝐕j\boldsymbol{Q}_{ij}=\boldsymbol{U}_{i}^{T}\boldsymbol{Q}\boldsymbol{V}_{j} satisfy

  • 𝑸1j=𝟎,𝑸i1=𝟎,𝑸33>0,\boldsymbol{Q}_{1j}=\boldsymbol{0},\ \boldsymbol{Q}_{i1}=\boldsymbol{0},\ \boldsymbol{Q}_{33}>0, for i,j=1,2,3.i,j=1,2,3.

  • 𝑨31=𝟎,𝑨33T=𝑨33\boldsymbol{A}_{31}=\boldsymbol{0},\ \boldsymbol{A}_{33}^{T}=\boldsymbol{A}_{33} and rank(𝑨21)=rank(𝑼2).\mathrm{rank}(\boldsymbol{A}_{21})=\mathrm{rank}(\boldsymbol{U}_{2}).

Proof.

By definition, 𝑨33\boldsymbol{A}_{33} is symmetric. Since 𝑸𝑮=𝟎,\boldsymbol{Q}\boldsymbol{G}=\boldsymbol{0}, we have 𝑸1j=𝟎\boldsymbol{Q}_{1j}=\boldsymbol{0} and 𝑸i1=𝟎.\boldsymbol{Q}_{i1}=\boldsymbol{0}. From (2.8), we have 𝑸33>0.\boldsymbol{Q}_{33}>0. We have rank(𝑨21)=rank(𝑼2T𝑼2)=rank(𝑼2)\mathrm{rank}(\boldsymbol{A}_{21})=\mathrm{rank}(\boldsymbol{U}_{2}^{T}\boldsymbol{U}_{2})=\mathrm{rank}(\boldsymbol{U}_{2}). Meanwhile, 𝑨31=𝑽3T𝑼2=𝟎\boldsymbol{A}_{31}=\boldsymbol{V}_{3}^{T}\boldsymbol{U}_{2}=\boldsymbol{0}. ∎

Thanks to the above transformation, we may consider the generalized eigenvalue problem of (𝑨33,𝑸33)(\boldsymbol{A}_{33},\boldsymbol{Q}_{33}) rather than (𝑨,𝑸),(\boldsymbol{A},\boldsymbol{Q}), where 𝑸33\boldsymbol{Q}_{33} is symmetric positive definite while 𝑸\boldsymbol{Q} is symmetric positive semi-definite. Here 𝑨33\boldsymbol{A}_{33} and 𝑨\boldsymbol{A} are both symmetric matrices which may have zero eigenvalues. The choices of 𝑽2\boldsymbol{V}_{2} and 𝑽3\boldsymbol{V}_{3} are not unique, but they do not affect the results of Lemma 2.1.

The general solutions of ODEs should be written with the aid of the eigenvalue decomposition. According to Sylvester’s law of inertia, 𝑸331𝑨33\boldsymbol{Q}_{33}^{-1}\boldsymbol{A}_{33} should have the same number of positive, negative and zero eigenvalues as 𝑨33\boldsymbol{A}_{33}. Assume the Cholesky decomposition

𝑸33=𝑳𝑳T,\boldsymbol{Q}_{33}=\boldsymbol{L}\boldsymbol{L}^{T},

then we must have an orthogonal eigenvalue decomposition of the real symmetric matrix:

𝑳1𝑨33𝑳T𝑹=𝑹𝚲.\boldsymbol{L}^{-1}\boldsymbol{A}_{33}\boldsymbol{L}^{-T}\boldsymbol{R}=\boldsymbol{R}\boldsymbol{\Lambda}. (2.9)

We assume the diagonal matrix

𝚲=[𝚲+𝟎𝚲],\boldsymbol{\Lambda}=\begin{bmatrix}\boldsymbol{\Lambda}_{+}&&\\ &\boldsymbol{0}&\\ &&\boldsymbol{\Lambda}_{-}\end{bmatrix},

where 𝚲+n+×n+\boldsymbol{\Lambda}_{+}\in\mathbb{R}^{n_{+}\times n_{+}} has postive entries and 𝚲n×n\boldsymbol{\Lambda}_{-}\in\mathbb{R}^{n_{-}\times n_{-}} has negative entries. The matrix 𝑹=[𝑹+,𝑹0,𝑹]\boldsymbol{R}=[\boldsymbol{R}_{+},\boldsymbol{R}_{0},\boldsymbol{R}_{-}] is assumed orthogonal, where the columns of 𝑹+,𝑹0,𝑹\boldsymbol{R}_{+},\boldsymbol{R}_{0},\boldsymbol{R}_{-} are individually the number of postive, zero and negative eigenvalues of 𝑨33\boldsymbol{A}_{33}, i.e., n+,n0,nn_{+},n_{0},n_{-}. Suppose

𝑻=𝑳T𝑹=[𝑻+,𝑻0,𝑻],\boldsymbol{T}=\boldsymbol{L}^{-T}\boldsymbol{R}=[\boldsymbol{T}_{+},\boldsymbol{T}_{0},\boldsymbol{T}_{-}],

where the columns of 𝑻+,𝑻0,𝑻\boldsymbol{T}_{+},\boldsymbol{T}_{0},\boldsymbol{T}_{-} are individually n+,n0,nn_{+},n_{0},n_{-}. Then 𝑻\boldsymbol{T} is invertible and we have

𝑸331𝑨33[𝑻+,𝑻0,𝑻]=[𝑻+,𝑻0,𝑻][𝚲+𝟎𝚲].\boldsymbol{Q}_{33}^{-1}\boldsymbol{A}_{33}[\boldsymbol{T}_{+},\boldsymbol{T}_{0},\boldsymbol{T}_{-}]=[\boldsymbol{T}_{+},\boldsymbol{T}_{0},\boldsymbol{T}_{-}]\begin{bmatrix}\boldsymbol{\Lambda}_{+}&&\\ &\boldsymbol{0}&\\ &&\boldsymbol{\Lambda}_{-}\end{bmatrix}. (2.10)

For a positive number aa, we define the norm in L2(+,L2(N))L^{2}(\mathbb{R}_{+},L^{2}(\mathbb{R}^{N})) as

ha=(0+e2ayhTh(y)dy)1/2,\|h\|_{a}=\left(\int_{0}^{+\infty}\!\!e^{2ay}h^{T}h(y)\,\mathrm{d}y\right)^{1/2},

and the vector norm as

g=gTg.\|g\|=\sqrt{g^{T}g}.

Then we have the following well-posed theorem, whose complete proof is put in Section 4.

Theorem 2.1.

Assume a>0a>0 is a given constant such that a<1/λmaxa<1/\lambda_{max}, where λmax\lambda_{max} is the maximal eigenvalue of 𝐐331𝐀33\boldsymbol{Q}_{33}^{-1}\boldsymbol{A}_{33}. We propose the boundary condition

𝑩𝑽3TW(0)=g\boldsymbol{B}\boldsymbol{V}_{3}^{T}W(0)=g (2.11)

for the system (2.6), where 𝐁n+×(n++n0+n)\boldsymbol{B}\in\mathbb{R}^{n_{+}\times(n_{+}+n_{0}+n_{-})} and gn+g\in\mathbb{R}^{n_{+}} are given, with constant coefficients.

For any hh with ha<+\|h\|_{a}<+\infty and gg with g<+\|g\|<+\infty, the system (2.6) has a unique solution and there exists a positive constant Ca,MC_{a,M} independent of the non-homogeneous term hh such that the following estimation holds:

WaCa,M(ha+g),\|W\|_{a}\leq C_{a,M}\left(\|h\|_{a}+\|g\|\right), (2.12)

if and only if the boundary condition satisfies the following conditions

  1. 1.

    rank(𝑩𝑻+)=n+.\mathrm{rank}(\boldsymbol{B}\boldsymbol{T}_{+})=n_{+}.

  2. 2.

    𝑩𝑻0=𝟎.\boldsymbol{B}\boldsymbol{T}_{0}=\boldsymbol{0}.

What’s more, the analytical expressions of the solution are given by (4), (4.1) and (4.3).

Remark 2.1.

According to the proof in Section 4, we can write 𝐕3TW(0)\boldsymbol{V}_{3}^{T}W(0) as

𝑽3TW(0)=𝑻+z+(0)+𝑻0z0(0)+𝑻z(0),\boldsymbol{V}_{3}^{T}W(0)=\boldsymbol{T}_{+}z_{+}(0)+\boldsymbol{T}_{0}z_{0}(0)+\boldsymbol{T}_{-}z_{-}(0),

where z0z_{0} and zz_{-} are determined by hh. So the first condition in Theorem 2.1 makes sure the unique solvability of z+(0)n+z_{+}(0)\in\mathbb{R}^{n_{+}}. The second condition shows that z+(0)z_{+}(0) will not be affected by z0(0)z_{0}(0), i.e.,

z+(0)=(𝑩𝑻+)1(g𝑩𝑻z(0)).z_{+}(0)=(\boldsymbol{B}\boldsymbol{T}_{+})^{-1}(g-\boldsymbol{B}\boldsymbol{T}_{-}z_{-}(0)).

Because z0(0)\|z_{0}(0)\| can not be controlled by g\|g\| and ha\|h\|_{a}, removing it from the solution ensures the estimation (2.12). In the homogeneous case, we have h=0h=0 and z0=0,z=0z_{0}=0,\ z_{-}=0, which means that there is no need to consider the second condition.

Remark 2.2.

In Theorem 2.1, we restrict the shape of the coefficient matrix 𝐁\boldsymbol{B} to be n+×(n++n0+n)n_{+}\times(n_{+}+n_{0}+n_{-}). Then since 𝐓+(n++n0+n)×n+\boldsymbol{T}_{+}\in\mathbb{R}^{(n_{+}+n_{0}+n_{-})\times n_{+}}, the first condition in Theorem 2.1 shows that 𝐁𝐓+\boldsymbol{B}\boldsymbol{T}_{+} is invertible. The assumption is mainly for ease of exposition. If we give the boundary condition

𝑩3𝑽3TW(0)=g~,\boldsymbol{B}_{3}\boldsymbol{V}_{3}^{T}W(0)=\tilde{g},

where 𝐁3n~×(n++n0+n)\boldsymbol{B}_{3}\in\mathbb{R}^{\tilde{n}\times(n_{+}+n_{0}+n_{-})}, then the solvability requires rank(𝐁3𝐓+)=n+\mathrm{rank}(\boldsymbol{B}_{3}\boldsymbol{T}_{+})=n_{+} and

g~𝑩3𝑻0z0(0)𝑩3𝑻z(0)span{𝑩3𝑻+}.\tilde{g}-\boldsymbol{B}_{3}\boldsymbol{T}_{0}z_{0}(0)-\boldsymbol{B}_{3}\boldsymbol{T}_{-}z_{-}(0)\in\mathrm{span}\{\boldsymbol{B}_{3}\boldsymbol{T}_{+}\}. (2.13)

In that case, g~\tilde{g} may depend on hh due to the condition (2.13) and g~\|\tilde{g}\| may not have a upper bound. So it’s a little wordy to obtain an analogous estimation as (2.12). The key point is that if we can find a matrix 𝐂n~×n+\boldsymbol{C}\in\mathbb{R}^{\tilde{n}\times n_{+}} such that 𝐂T𝐁3𝐓+\boldsymbol{C}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{+} is invertible and 𝐂T𝐁3𝐓0=𝟎\boldsymbol{C}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{0}=\boldsymbol{0}, then Theorem 2.1 helps to give the estimation.

3 Modified Boundary Conditions

3.1 Instability of the Grad Boundary Condition

For initial-boundary value problems of the linear hyperbolic system, Majda and Osher [26] emphasize that the linear space determined by the boundary condition should contain the null space of the boundary matrix. Otherwise, the boundary condition may be unstable. Based on the similar observation, [28] points out that the linear Grad boundary condition is unstable for the initial-boundary value problem.

For the half-space problem, we will use a simple example to illustrate the instability of the Grad boundary condition. From the example, we can see that the instability not only comes from the half-space problem.

If we consider Kramers’ problem with the BGK collision term [12], the simplest moment system when M=3M=3 can read as

dσ12dy=0,\displaystyle\dfrac{\mathrm{d}{\sigma_{12}}}{\mathrm{d}{y}}=0,
du1dy+2df3dy=νσ12,\displaystyle\dfrac{\mathrm{d}{u_{1}}}{\mathrm{d}{y}}+\sqrt{2}\dfrac{\mathrm{d}{f_{3}}}{\mathrm{d}{y}}=-\nu\sigma_{12}, (3.1)
2dσ12dy=νf3+h3,\displaystyle\sqrt{2}\dfrac{\mathrm{d}{\sigma_{12}}}{\mathrm{d}{y}}=-\nu f_{3}+h_{3},

where ν>0\nu>0 is a constant and we write the moment variables as σ12,u1,f3\sigma_{12},u_{1},f_{3}. The term h3=h3(y)h_{3}=h_{3}(y) is the given non-homogeneous term. The Grad boundary condition reads as

σ¯+σ12+χ^(u1+u¯+22f3)=0,\bar{\sigma}+\sigma_{12}+\hat{\chi}\left(u_{1}+\bar{u}+\frac{\sqrt{2}}{2}f_{3}\right)=0, (3.2)

where χ^=2χ2χ12π\hat{\chi}=\displaystyle\frac{2\chi}{2-\chi}\frac{1}{\sqrt{2\pi}} and σ¯,u¯\bar{\sigma},\ \bar{u} are given by the far field. Since the moment variables vanish when y=+y=+\infty, we can solve from (3.1) that

σ12=0,u1=2f3,f3=h3/ν.\sigma_{12}=0,\ u_{1}=-\sqrt{2}f_{3},\ f_{3}=h_{3}/\nu. (3.3)

So the Grad boundary condition gives

u¯=1χ^σ¯+22h3(0)ν.\bar{u}=-\frac{1}{\hat{\chi}}\bar{\sigma}+\frac{\sqrt{2}}{2}\frac{h_{3}(0)}{\nu}.

Only when u¯\bar{u} and σ¯\bar{\sigma} satisfy the above relation can the half-space problem has a unique solution. However, the term h3(0)h_{3}(0) can not be controlled by the weighted L2L^{2} norm h3a\|h_{3}\|_{a}. In this sense, we claim that u¯\bar{u} is unstable because u¯\|\bar{u}\| can not be controlled by h3a\|h_{3}\|_{a}.

Using the notations in Theorem 2.1, in this example, we have

𝑨=[001002120],𝑸=[0νν],h=[0h30],W=[u1f3σ12].\boldsymbol{A}=\begin{bmatrix}0&0&1\\ 0&0&\sqrt{2}\\ 1&\sqrt{2}&0\end{bmatrix},\ \boldsymbol{Q}=\begin{bmatrix}0&&\\ &\nu&\\ &&\nu\end{bmatrix},\ h=\begin{bmatrix}0\\ h_{3}\\ 0\end{bmatrix},\ W=\begin{bmatrix}u_{1}\\ f_{3}\\ \sigma_{12}\end{bmatrix}.

The previous procedure will give

𝑽1=[100],𝑽3=[010],n+=0,n0=1.\boldsymbol{V}_{1}=\begin{bmatrix}1\\ 0\\ 0\end{bmatrix},\ \boldsymbol{V}_{3}=\begin{bmatrix}0\\ 1\\ 0\end{bmatrix},\ n_{+}=0,\ n_{0}=1.

Since n+=0,n_{+}=0, Theorem 2.1 tells that the half-space problem does not need boundary conditions at y=0y=0, where the moment system directly gives the solution (3.3). But the Grad boundary condition gives one more condition. So this condition asks for relations of the given values, e.g., u¯\bar{u} and σ¯\bar{\sigma}. We should also consider the stability of these additional solutions.

When M=5M=5, we can check that n+>0n_{+}>0 and the Grad boundary condition does not satisfy

𝑩𝑻0=𝟎,\boldsymbol{B}\boldsymbol{T}_{0}=\boldsymbol{0},

which shows the instability of the Grad boundary condition from Theorem 2.1. An analogous example is given in [28].

To overcome this drawback, one may change (3.2) as

c(σ¯+σ12)+χ^(u1+u¯+2f3)=0,c(\bar{\sigma}+\sigma_{12})+\hat{\chi}\left(u_{1}+\bar{u}+\sqrt{2}f_{3}\right)=0,

where c>0c>0 is an arbitrary positive constant. Then the modified boundary condition will give

u¯=1cχ^σ¯,\bar{u}=-\frac{1}{c\hat{\chi}}\bar{\sigma},

which is stable about h3h_{3}. The boundary conditions involving higher-order moment variables can be modified in the same way. Authors of [28] suggest choosing c=1c=1, such that the modified boundary condition differs from the Grad boundary condition only by the coefficients before the highest order moment variables, i.e., f3f_{3} in this example. However, it’s not clear whether the modification is optimal in the sense of providing the most accurate solutions.

3.2 Modified Boundary Conditions

We systematically introduce a class of modified boundary conditions and prove their well-posedness. Assume

𝕀e={𝜶3:α2 even,|𝜶|M},𝕀o={𝜶3:α2 odd,|𝜶|M},\mathbb{I}_{e}=\{\boldsymbol{\alpha}\in\mathbb{N}^{3}:\ \alpha_{2}\text{\ even},\ |\boldsymbol{\alpha}|\leq M\},\quad\mathbb{I}_{o}=\{\boldsymbol{\alpha}\in\mathbb{N}^{3}:\ \alpha_{2}\text{\ odd},\ |\boldsymbol{\alpha}|\leq M\},

and m=#𝕀e,n=#𝕀o.m=\#\mathbb{I}_{e},\ n=\#\mathbb{I}_{o}. So we have m+n=Nm+n=N and mn.m\geq n. Suppose the multi-indices with the even second component are always ordered before the ones with the odd second component, e.g., (a1,0,a3)(a_{1},0,a_{3}) is ordered before (b1,1,b3)(b_{1},1,b_{3}) for any a1,a3,b1,b3a_{1},a_{3},b_{1},b_{3}. Then we can write the Grad boundary condition (2.5) as

𝑬𝑴(Wobo)+χ^𝑬𝑺(Webe)=0,\boldsymbol{E}\boldsymbol{M}(W_{o}-b_{o})+\hat{\chi}\boldsymbol{E}\boldsymbol{S}(W_{e}-b_{e})=0, (3.4)

where χ^=2χ2χ12π\hat{\chi}=\displaystyle\frac{2\chi}{2-\chi}\frac{1}{\sqrt{2\pi}}. The matrix 𝑴m×n\boldsymbol{M}\in\mathbb{R}^{m\times n} has entries ξ2ϕ𝜶ϕ𝜷\left\langle{\mathcal{M}\xi_{2}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle for 𝜶𝕀e\boldsymbol{\alpha}\in\mathbb{I}_{e} and 𝜷𝕀o.\boldsymbol{\beta}\in\mathbb{I}_{o}. And the matrix 𝑺m×m\boldsymbol{S}\in\mathbb{R}^{m\times m} has entries |ξ2|ϕ𝜶ϕ𝜷\left\langle{\mathcal{M}|\xi_{2}|\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle for 𝜶,𝜷𝕀e\boldsymbol{\alpha},\ \boldsymbol{\beta}\in\mathbb{I}_{e}. The variables are divided as

W=[WeWo],b=[bebo],W=\begin{bmatrix}W_{e}\\ W_{o}\end{bmatrix},\ b=\begin{bmatrix}b_{e}\\ b_{o}\end{bmatrix},

where the elements of WemW_{e}\in\mathbb{R}^{m} are w𝜶,𝜶𝕀ew_{\boldsymbol{\alpha}},\ \boldsymbol{\alpha}\in\mathbb{I}_{e} and the elements of WonW_{o}\in\mathbb{R}^{n} are w𝜶,𝜶𝕀o.w_{\boldsymbol{\alpha}},\ \boldsymbol{\alpha}\in\mathbb{I}_{o}. Every row of 𝑬n×m\boldsymbol{E}\in\mathbb{R}^{n\times m} is a unit vector with only one component being one, such that the entries of 𝑬𝑴\boldsymbol{E}\boldsymbol{M} are ξ2ϕ𝜶ϕ𝜷\left\langle{\mathcal{M}\xi_{2}\phi_{\boldsymbol{\alpha}}\phi_{\boldsymbol{\beta}}}\right\rangle ,𝜶𝕀e,|𝜶|M1,,\boldsymbol{\alpha}\in\mathbb{I}_{e},\ |\boldsymbol{\alpha}|\leq M-1, and 𝜷𝕀o.\boldsymbol{\beta}\in\mathbb{I}_{o}.

Due to the recursion relation and orthogonality of Hermite polynomials [6], we can write

𝑨=[𝟎𝑴𝑴T𝟎],𝑸=[𝑸e𝟎𝟎𝑸o],\boldsymbol{A}=\begin{bmatrix}\boldsymbol{0}&\boldsymbol{M}\\ \boldsymbol{M}^{T}&\boldsymbol{0}\end{bmatrix},\quad\boldsymbol{Q}=\begin{bmatrix}\boldsymbol{Q}_{e}&\boldsymbol{0}\\ \boldsymbol{0}&\boldsymbol{Q}_{o}\end{bmatrix},

where 𝑸em×m\boldsymbol{Q}_{e}\in\mathbb{R}^{m\times m} and 𝑸on×n.\boldsymbol{Q}_{o}\in\mathbb{R}^{n\times n}. By definition, we also know that 𝑴\boldsymbol{M} is of full column rank and 𝑺\boldsymbol{S} is symmetric positive definite [23].

The modified boundary condition for the initial-boundary value problem is

𝑯(Wobo)+χ^𝑴T(Webe)=0,\boldsymbol{H}(W_{o}-b_{o})+\hat{\chi}\boldsymbol{M}^{T}(W_{e}-b_{e})=0, (3.5)

where 𝑯n×n\boldsymbol{H}\in\mathbb{R}^{n\times n} is an arbitrary symmetric positive definite matrix. For the half-space problem, the boundary condition should write as

𝑯(Wo+W¯obo)+χ^𝑴T(We+W¯ebe)=0,\boldsymbol{H}(W_{o}+\bar{W}_{o}-b_{o})+\hat{\chi}\boldsymbol{M}^{T}(W_{e}+\bar{W}_{e}-b_{e})=0, (3.6)

where W¯e\bar{W}_{e} and W¯o\bar{W}_{o} are given by the flow outside the boundary layer.

Inspired by the illustrative examples, we can first find a matrix 𝑪n×n+\boldsymbol{C}\in\mathbb{R}^{n\times n_{+}} and multiply (3.6) left by 𝑪T\boldsymbol{C}^{T}. Then, we can use Theorem 2.1 to check the well-posedness of the half-space problem. Finally, we may solve the remaining part of (3.6) to obtain relations between the given vectors, e.g., W¯o\bar{W}_{o} and W¯e\bar{W}_{e}.

Following this way, we first calculate the value of n+n_{+} by the special block structure of 𝑨\boldsymbol{A} and 𝑸\boldsymbol{Q}. We may as well write

𝑮=[𝑮e𝑮o],𝑿=[𝑿e𝑿o],\boldsymbol{G}=\begin{bmatrix}\boldsymbol{G}_{e}&\\ &\boldsymbol{G}_{o}\end{bmatrix},\ \boldsymbol{X}=\begin{bmatrix}\boldsymbol{X}_{e}&\\ &\boldsymbol{X}_{o}\end{bmatrix},\

where 𝑮em×p1\boldsymbol{G}_{e}\in\mathbb{R}^{m\times p_{1}}, 𝑮on×p2\boldsymbol{G}_{o}\in\mathbb{R}^{n\times p_{2}}, 𝑿em×r1\boldsymbol{X}_{e}\in\mathbb{R}^{m\times r_{1}}, 𝑿on×r2\boldsymbol{X}_{o}\in\mathbb{R}^{n\times r_{2}}. And

𝑽1=[𝒀1𝒁1],𝑽2=[𝒀2𝒁2],𝑽3=[𝒀3𝒁3],\boldsymbol{V}_{1}=\begin{bmatrix}\boldsymbol{Y}_{1}&\\ &\boldsymbol{Z}_{1}\end{bmatrix},\ \boldsymbol{V}_{2}=\begin{bmatrix}\boldsymbol{Y}_{2}&\\ &\boldsymbol{Z}_{2}\end{bmatrix},\ \boldsymbol{V}_{3}=\begin{bmatrix}\boldsymbol{Y}_{3}&\\ &\boldsymbol{Z}_{3}\end{bmatrix},

where 𝒀1=𝑮e𝑿e,𝒁1=𝑮o𝑿o\boldsymbol{Y}_{1}=\boldsymbol{G}_{e}\boldsymbol{X}_{e},\ \boldsymbol{Z}_{1}=\boldsymbol{G}_{o}\boldsymbol{X}_{o},

span{𝒀2}=span{𝑴𝑮o},span{𝒁2}=span{𝑴T𝑮e},\mathrm{span}\{\boldsymbol{Y}_{2}\}=\mathrm{span}\{\boldsymbol{M}\boldsymbol{G}_{o}\},\ \mathrm{span}\{\boldsymbol{Z}_{2}\}=\mathrm{span}\{\boldsymbol{M}^{T}\boldsymbol{G}_{e}\},

𝒀3m×(mr1p2)\boldsymbol{Y}_{3}\in\mathbb{R}^{m\times(m-r_{1}-p_{2})} and 𝒁3n×(nr2p1)\boldsymbol{Z}_{3}\in\mathbb{R}^{n\times(n-r_{2}-p_{1})}. Under these assumptions, we claim that

Lemma 3.1.

n+=nr2p1.n_{+}=n-r_{2}-p_{1}.

Proof.

We may as well write the Cholesky decomposition 𝑸33=𝑳𝑳T\boldsymbol{Q}_{33}=\boldsymbol{L}\boldsymbol{L}^{T} as

𝑳=[𝑳e𝑳o],𝑳e(mr1p2)×(mr1p2),𝑳o(nr2p1)×(nr2p1).\boldsymbol{L}=\begin{bmatrix}\boldsymbol{L}_{e}&\\ &\boldsymbol{L}_{o}\end{bmatrix},\ \boldsymbol{L}_{e}\in\mathbb{R}^{(m-r_{1}-p_{2})\times(m-r_{1}-p_{2})},\ \boldsymbol{L}_{o}\in\mathbb{R}^{(n-r_{2}-p_{1})\times(n-r_{2}-p_{1})}.

Then we have

𝑳1𝑨33𝑳T=[𝟎𝑳e1𝒀3T𝑴𝒁3𝑳oT𝑳o1𝒁3T𝑴T𝒀3𝑳eT𝟎].\boldsymbol{L}^{-1}\boldsymbol{A}_{33}\boldsymbol{L}^{-T}=\begin{bmatrix}\boldsymbol{0}&\boldsymbol{L}_{e}^{-1}\boldsymbol{Y}_{3}^{T}\boldsymbol{M}\boldsymbol{Z}_{3}\boldsymbol{L}_{o}^{-T}\\ \boldsymbol{L}_{o}^{-1}\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}\boldsymbol{Y}_{3}\boldsymbol{L}_{e}^{-T}&\boldsymbol{0}\end{bmatrix}.

We will show that 𝒀3T𝑴𝒁3\boldsymbol{Y}_{3}^{T}\boldsymbol{M}\boldsymbol{Z}_{3} is of full column rank, and the positive as well as negative eigenvalues of 𝑳1𝑨33𝑳T\boldsymbol{L}^{-1}\boldsymbol{A}_{33}\boldsymbol{L}^{-T} should appear in pair. These facts will lead to n+=nr2p1.n_{+}=n-r_{2}-p_{1}.

Suppose 𝒀3T𝑴𝒁3x=0\boldsymbol{Y}_{3}^{T}\boldsymbol{M}\boldsymbol{Z}_{3}x=0 for some xnr2p1x\in\mathbb{R}^{n-r_{2}-p_{1}}. Since [𝒀1,𝒀2,𝒀3][\boldsymbol{Y}_{1},\boldsymbol{Y}_{2},\boldsymbol{Y}_{3}] is orthogonal, there exists x1r1x_{1}\in\mathbb{R}^{r_{1}} and x2p2x_{2}\in\mathbb{R}^{p_{2}} such that

𝑴𝒁3x=𝒀1x1+𝒀2x2.\boldsymbol{M}\boldsymbol{Z}_{3}x=\boldsymbol{Y}_{1}x_{1}+\boldsymbol{Y}_{2}x_{2}. (3.7)

Since 𝒁3T𝑴T𝒀1=𝟎\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}\boldsymbol{Y}_{1}=\boldsymbol{0} and 𝒀2T𝒀1=𝟎\boldsymbol{Y}_{2}^{T}\boldsymbol{Y}_{1}=\boldsymbol{0}, we have x1=0.x_{1}=0. The relation (2.8) implies that

span{𝒁3}span{𝑮o}={0}.\mathrm{span}\{\boldsymbol{Z}_{3}\}\cap\mathrm{span}\{\boldsymbol{G}_{o}\}=\{0\}.

So there must be x=0x=0 since 𝑴\boldsymbol{M} and 𝒁3\boldsymbol{Z}_{3} are of full column rank. This shows that 𝒀3T𝑴𝒁3\boldsymbol{Y}_{3}^{T}\boldsymbol{M}\boldsymbol{Z}_{3} is of full column rank.

For eigenvalues and eigenvectors of the symmetric matrix, we introduce a general conclusion. Let 𝑫α×β\boldsymbol{D}\in\mathbb{R}^{\alpha\times\beta} and rank(𝑫)=γ\mathrm{rank}(\boldsymbol{D})=\gamma. Then

𝑫~:=[𝟎𝑫𝑫T𝟎]\boldsymbol{\tilde{D}}:=\begin{bmatrix}\boldsymbol{0}&\boldsymbol{D}\\ \boldsymbol{D}^{T}&\boldsymbol{0}\end{bmatrix}

has α+β2γ\alpha+\beta-2\gamma zero eigenvalues, γ\gamma positive eigenvalues and γ\gamma negative eigenvalues.

In fact, the symmetric matrix 𝑫~\boldsymbol{\tilde{D}} must have α+β\alpha+\beta real eigenvalues. Assume λ\lambda\in\mathbb{R} is an eigenvalue, then there exists xαx\in\mathbb{R}^{\alpha} and yβy\in\mathbb{R}^{\beta} such that

[𝟎𝑫𝑫T𝟎][xy]=λ[xy][𝟎𝑫𝑫T𝟎][xy]=λ[xy].\begin{bmatrix}\boldsymbol{0}&\boldsymbol{D}\\ \boldsymbol{D}^{T}&\boldsymbol{0}\end{bmatrix}\begin{bmatrix}x\\ y\end{bmatrix}=\lambda\begin{bmatrix}x\\ y\end{bmatrix}\quad\Rightarrow\quad\begin{bmatrix}\boldsymbol{0}&\boldsymbol{D}\\ \boldsymbol{D}^{T}&\boldsymbol{0}\end{bmatrix}\begin{bmatrix}x\\ -y\end{bmatrix}=-\lambda\begin{bmatrix}x\\ -y\end{bmatrix}.

So λ-\lambda is also an eigenvalue, which implies that 𝑫~\boldsymbol{\tilde{D}} has the same number of positive and negative eigenvalues. Since rank(𝑫)=rank(𝑫T)=γ\mathrm{rank}(\boldsymbol{D})=\mathrm{rank}(\boldsymbol{D}^{T})=\gamma, there must be α+β2γ\alpha+\beta-2\gamma zero eigenvalues.

Applying the above result to 𝑳1𝑨33𝑳T\boldsymbol{L}^{-1}\boldsymbol{A}_{33}\boldsymbol{L}^{-T}, we then have n+=nr2p1.n_{+}=n-r_{2}-p_{1}.

Incidentally, applying the above result in Lemma 3.1 to 𝑮T𝑨𝑮\boldsymbol{G}^{T}\boldsymbol{A}\boldsymbol{G}, we have r1+p2=r2+p1.r_{1}+p_{2}=r_{2}+p_{1}. Then the required matrix 𝑪n×n+\boldsymbol{C}\in\mathbb{R}^{n\times n_{+}} can be found.

Lemma 3.2.

Let 𝐁3=[χ^𝐌T,𝐇]𝐕3\boldsymbol{B}_{3}=[\hat{\chi}\boldsymbol{M}^{T},\boldsymbol{H}]\boldsymbol{V}_{3} and 𝐂=𝐙3\boldsymbol{C}=\boldsymbol{Z}_{3}. Then when χ[0,1]\chi\in[0,1], we have

  • i.

    𝑪T𝑩3𝑻+\boldsymbol{C}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{+} is invertible.

  • ii.

    𝑪T𝑩3𝑻0=𝟎.\boldsymbol{C}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{0}=\boldsymbol{0}.

Proof.

Thanks to the block structure of 𝑳1𝑨33𝑳T\boldsymbol{L}^{-1}\boldsymbol{A}_{33}\boldsymbol{L}^{-T}, we can assume

𝑹+=[𝑹e𝑹o],𝑹e(mr1p2)×n+,𝑹on+×n+,\boldsymbol{R}_{+}=\begin{bmatrix}\boldsymbol{R}_{e}\\ \boldsymbol{R}_{o}\end{bmatrix},\quad\boldsymbol{R}_{e}\in\mathbb{R}^{(m-r_{1}-p_{2})\times n_{+}},\quad\boldsymbol{R}_{o}\in\mathbb{R}^{n_{+}\times n_{+}},

where 𝑹eT𝑹e=𝑹oT𝑹o=12𝑰n+\displaystyle\boldsymbol{R}_{e}^{T}\boldsymbol{R}_{e}=\boldsymbol{R}_{o}^{T}\boldsymbol{R}_{o}=\frac{1}{2}\boldsymbol{I}_{n_{+}} with the identity matrix 𝑰n+\boldsymbol{I}_{n_{+}}. We have

𝒁3T𝑩3𝑻+\displaystyle\boldsymbol{Z}_{3}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{+} =\displaystyle= χ^𝒁3T𝑴T𝒀3𝑳eT𝑹e+𝒁3T𝑯𝒁3𝑳oT𝑹o\displaystyle\hat{\chi}\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}\boldsymbol{Y}_{3}\boldsymbol{L}_{e}^{-T}\boldsymbol{R}_{e}+\boldsymbol{Z}_{3}^{T}\boldsymbol{H}\boldsymbol{Z}_{3}\boldsymbol{L}_{o}^{-T}\boldsymbol{R}_{o}
=\displaystyle= χ^𝑳o𝑹o𝚲++𝒁3T𝑯𝒁3𝑳oT𝑹o.\displaystyle\hat{\chi}\boldsymbol{L}_{o}\boldsymbol{R}_{o}\boldsymbol{\Lambda}_{+}+\boldsymbol{Z}_{3}^{T}\boldsymbol{H}\boldsymbol{Z}_{3}\boldsymbol{L}_{o}^{-T}\boldsymbol{R}_{o}.

Since χ^0\hat{\chi}\geq 0 when χ[0,1]\chi\in[0,1], for any xn+x\in\mathbb{R}^{n_{+}}, we have

xT𝑹oT𝑳o1𝒁3T𝑩3𝑻+x\displaystyle x^{T}\boldsymbol{R}_{o}^{T}\boldsymbol{L}_{o}^{-1}\boldsymbol{Z}_{3}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{+}x =\displaystyle= 12χ^xT𝚲+x+xT𝑹oT𝑳o1𝒁3T𝑯𝒁3𝑳oT𝑹ox\displaystyle\frac{1}{2}\hat{\chi}x^{T}\boldsymbol{\Lambda}_{+}x+x^{T}\boldsymbol{R}_{o}^{T}\boldsymbol{L}_{o}^{-1}\boldsymbol{Z}_{3}^{T}\boldsymbol{H}\boldsymbol{Z}_{3}\boldsymbol{L}_{o}^{-T}\boldsymbol{R}_{o}x (3.8)
\displaystyle\geq xT𝑹oT𝑳o1𝒁3T𝑯𝒁3𝑳oT𝑹ox0,\displaystyle x^{T}\boldsymbol{R}_{o}^{T}\boldsymbol{L}_{o}^{-1}\boldsymbol{Z}_{3}^{T}\boldsymbol{H}\boldsymbol{Z}_{3}\boldsymbol{L}_{o}^{-T}\boldsymbol{R}_{o}x\geq 0,

where the equality holds if and only if x=0.x=0. This shows that 𝑹oT𝑳o1𝒁3T𝑩3𝑻+\boldsymbol{R}_{o}^{T}\boldsymbol{L}_{o}^{-1}\boldsymbol{Z}_{3}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{+} is symmetric positive definite. So 𝒁3T𝑩3𝑻+\boldsymbol{Z}_{3}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{+} is invertible.

Then we show 𝒁3T𝑩3𝑻0=𝟎\boldsymbol{Z}_{3}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{0}=\boldsymbol{0}. By definition, we have 𝑸331𝑨33𝑻0=𝟎,\boldsymbol{Q}_{33}^{-1}\boldsymbol{A}_{33}\boldsymbol{T}_{0}=\boldsymbol{0}, which shows that 𝑨33𝑻0=𝟎.\boldsymbol{A}_{33}\boldsymbol{T}_{0}=\boldsymbol{0}. Due to the block structure of 𝑨33\boldsymbol{A}_{33}, the matrix 𝑻0\boldsymbol{T}_{0} can write as

𝑻0=[𝑻𝟎],𝑻(mr1p2)×n0,\boldsymbol{T}_{0}=\begin{bmatrix}\boldsymbol{T}^{*}\\ \boldsymbol{0}\end{bmatrix},\ \boldsymbol{T}^{*}\in\mathbb{R}^{(m-r_{1}-p_{2})\times n_{0}},

since 𝒀3T𝑴𝒁3\boldsymbol{Y}_{3}^{T}\boldsymbol{M}\boldsymbol{Z}_{3} is of full column rank. This gives 𝒁3T𝑴T𝒀3𝑻=𝟎\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}\boldsymbol{Y}_{3}\boldsymbol{T}^{*}=\boldsymbol{0} and we have

𝒁3T𝑩3𝑻0=χ^𝒁3T𝑴T𝒀3𝑻=𝟎,\boldsymbol{Z}_{3}^{T}\boldsymbol{B}_{3}\boldsymbol{T}_{0}=\hat{\chi}\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}\boldsymbol{Y}_{3}\boldsymbol{T}^{*}=\boldsymbol{0},

which completes the proof. ∎

With the aid of the above lemmas, we state the well-posedness theorem as follows:

Theorem 3.1.

Suppose χ^>0\hat{\chi}>0 and r2=0r_{2}=0. For any hh with ha<+\|h\|_{a}<+\infty and g1=W¯obog_{1}=\bar{W}_{o}-b_{o}g2=(𝐈m𝐆e𝐆eT)(W¯ebe)g_{2}=(\boldsymbol{I}_{m}-\boldsymbol{G}_{e}\boldsymbol{G}_{e}^{T})(\bar{W}_{e}-b_{e}) with g1<+,g2<+\|g_{1}\|<+\infty,\ \|g_{2}\|<+\infty, the moment system (2.6) with the boundary condition (3.6) has a unique solution of WW and 𝐆eT(W¯ebe)\boldsymbol{G}_{e}^{T}(\bar{W}_{e}-b_{e}), and the solution satisfies the estimation

Waha+g1+g2,\displaystyle\|W\|_{a}\lesssim\|h\|_{a}+\|g_{1}\|+\|g_{2}\|,
𝑮eT(W¯ebe)ha+g1+g2,\displaystyle\|\boldsymbol{G}_{e}^{T}(\bar{W}_{e}-b_{e})\|\lesssim\|h\|_{a}+\|g_{1}\|+\|g_{2}\|,

where aba\lesssim b represents that there exists a constant c>0c>0 such that acb.a\leq cb.

Proof.

The assumption r2=0r_{2}=0 makes the matrices 𝑿o\boldsymbol{X}_{o} and 𝒁1\boldsymbol{Z}_{1} vanish, which simplifies the proof. Note that 𝑽\boldsymbol{V} is orthogonal. Plugging 𝑽𝑽T\boldsymbol{V}\boldsymbol{V}^{T} before the variables WeW_{e} and WoW_{o} in (3.6), and decomposing W¯ebe\bar{W}_{e}-b_{e} as

W¯ebe=𝑮e𝑮eT(W¯ebe)+(𝑰m𝑮e𝑮eT)(W¯ebe),\bar{W}_{e}-b_{e}=\boldsymbol{G}_{e}\boldsymbol{G}_{e}^{T}(\bar{W}_{e}-b_{e})+(\boldsymbol{I}_{m}-\boldsymbol{G}_{e}\boldsymbol{G}_{e}^{T})(\bar{W}_{e}-b_{e}),

the boundary condition can be equivalently written as

χ^𝑴T(𝒀1𝒀1TWe(0)+𝑮e𝑮eT(W¯ebe))+𝑩3𝑽3TW(0)\displaystyle\hat{\chi}\boldsymbol{M}^{T}\left(\boldsymbol{Y}_{1}\boldsymbol{Y}_{1}^{T}W_{e}(0)+\boldsymbol{G}_{e}\boldsymbol{G}_{e}^{T}(\bar{W}_{e}-b_{e})\right)+\boldsymbol{B}_{3}\boldsymbol{V}_{3}^{T}W(0) (3.9)
=\displaystyle= 𝑯(W¯obo)χ^𝑴T(𝑰m𝑮e𝑮eT)(W¯ebe)[χ^𝑴T,𝑯]𝑽2𝑽2TW(0),\displaystyle-\boldsymbol{H}(\bar{W}_{o}-b_{o})-\hat{\chi}\boldsymbol{M}^{T}(\boldsymbol{I}_{m}-\boldsymbol{G}_{e}\boldsymbol{G}_{e}^{T})(\bar{W}_{e}-b_{e})-[\hat{\chi}\boldsymbol{M}^{T},\boldsymbol{H}]\boldsymbol{V}_{2}\boldsymbol{V}_{2}^{T}W(0),

where 𝑩3=[χ^𝑴T,𝑯]𝑽3\boldsymbol{B}_{3}=[\hat{\chi}\boldsymbol{M}^{T},\boldsymbol{H}]\boldsymbol{V}_{3} as in Lemma 3.2.

We extract the lower block of 𝑼\boldsymbol{U} as 𝑼~=[𝑮o,𝑴T𝑮e𝑿e,𝒁3]n×n.\boldsymbol{\tilde{U}}=[\boldsymbol{G}_{o},\boldsymbol{M}^{T}\boldsymbol{G}_{e}\boldsymbol{X}_{e},\boldsymbol{Z}_{3}]\in\mathbb{R}^{n\times n}. Then 𝑼~\boldsymbol{\tilde{U}} is invertible since 𝑼\boldsymbol{U} is invertible. Multiplying (3.9) left by 𝑼~T\boldsymbol{\tilde{U}}^{T}, the boundary condition can be divided into three parts.

The first part corresponds to multiply (3.9) left by 𝒁3T\boldsymbol{Z}_{3}^{T}. Since span{𝑴T𝑮e}=span{𝒁2}\mathrm{span}\{\boldsymbol{M}^{T}\boldsymbol{G}_{e}\}=\mathrm{span}\{\boldsymbol{Z}_{2}\} and 𝒀1=𝑮e𝑿e\boldsymbol{Y}_{1}=\boldsymbol{G}_{e}\boldsymbol{X}_{e}, we have 𝒁3T𝑴T𝒀1=𝒁3T𝑴T𝑮e=𝟎\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}\boldsymbol{Y}_{1}=\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}\boldsymbol{G}_{e}=\boldsymbol{0} because 𝒁3T𝒁2=𝟎.\boldsymbol{Z}_{3}^{T}\boldsymbol{Z}_{2}=\boldsymbol{0}. Now the boundary condition gives

𝒁3T𝑩3𝑽3TW(0)=𝒁3T𝑯g1χ^𝒁3T𝑴Tg2𝒁3T[χ^𝑴T,𝑯]𝑽2𝑽2TW(0),\displaystyle\boldsymbol{Z}_{3}^{T}\boldsymbol{B}_{3}\boldsymbol{V}_{3}^{T}W(0)=-\boldsymbol{Z}_{3}^{T}\boldsymbol{H}g_{1}-\hat{\chi}\boldsymbol{Z}_{3}^{T}\boldsymbol{M}^{T}g_{2}-\boldsymbol{Z}_{3}^{T}[\hat{\chi}\boldsymbol{M}^{T},\boldsymbol{H}]\boldsymbol{V}_{2}\boldsymbol{V}_{2}^{T}W(0), (3.10)

which satisfies the well-posed conditions in Theorem 2.1 due to Lemma 3.2. From the proof of Theorem 2.1, we know

𝑽2TW(0)ha.\|\boldsymbol{V}_{2}^{T}W(0)\|\lesssim\|h\|_{a}.

So from Theorem 2.1, the half-space system (2.6) with the boundary condition (3.10) has a unique solution of WW and

Waha+g1+g2.\displaystyle\|W\|_{a}\lesssim\|h\|_{a}+\|g_{1}\|+\|g_{2}\|.

The remaining part of the boundary condition (3.9) should be compatible. Note that r1+p2=r2+p1=p1.r_{1}+p_{2}=r_{2}+p_{1}=p_{1}. We write 𝑼0=[𝑮o,𝑴T𝑮e𝑿e]n×p1\boldsymbol{U}_{0}=[\boldsymbol{G}_{o},\boldsymbol{M}^{T}\boldsymbol{G}_{e}\boldsymbol{X}_{e}]\in\mathbb{R}^{n\times p_{1}}. Then the remaining part of the boundary condition corresponds to multiply (3.9) left by 𝑼0T.\boldsymbol{U}_{0}^{T}. We claim that

rank(𝑼0T𝑴T𝑮e)=p1.\mathrm{rank}(\boldsymbol{U}_{0}^{T}\boldsymbol{M}^{T}\boldsymbol{G}_{e})=p_{1}.

Otherwise, there exists xp1x\in\mathbb{R}^{p_{1}} such that xspan{𝑼0}x\in\mathrm{span}\{\boldsymbol{U}_{0}\} and xT𝑴T𝑮e=0.x^{T}\boldsymbol{M}^{T}\boldsymbol{G}_{e}=0. Since 𝒁1\boldsymbol{Z}_{1} vanishes and [𝒁2,𝒁3][\boldsymbol{Z}_{2},\boldsymbol{Z}_{3}] is orthogonal, we have xspan{𝒁3}x\in\mathrm{span}\{\boldsymbol{Z}_{3}\}. But [𝑼0,𝒁3][\boldsymbol{U}_{0},\boldsymbol{Z}_{3}] is invertible. So x=0x=0 and rank(𝑼0T𝑴T𝑮e)=p1.\mathrm{rank}(\boldsymbol{U}_{0}^{T}\boldsymbol{M}^{T}\boldsymbol{G}_{e})=p_{1}.

Thus, when WW is known, we can solve a unique 𝑮eT(W¯ebe)\boldsymbol{G}_{e}^{T}(\bar{W}_{e}-b_{e}) by multiplying (3.9) left with 𝑼0T.\boldsymbol{U}_{0}^{T}. From the proof of Theorem 2.1, only the trace z0(0)z_{0}(0) included in 𝑽3TW(0)\boldsymbol{V}_{3}^{T}W(0) can not be controlled by ha\|h\|_{a}. Fortunately, from (4), the term 𝑨21𝑽1TW(0)+𝑨23𝑽3TW(0)\boldsymbol{A}_{21}\boldsymbol{V}_{1}^{T}W(0)+\boldsymbol{A}_{23}\boldsymbol{V}_{3}^{T}W(0) will not contain z0(0)z_{0}(0). So we have the estimation

𝑮eT(W¯ebe)g1+g2+ha.\|\boldsymbol{G}_{e}^{T}(\bar{W}_{e}-b_{e})\|\lesssim\|g_{1}\|+\|g_{2}\|+\|h\|_{a}.

This completes the proof. ∎

In conclusion, the boundary condition for the moment equations in half-space is different from the case of initial-boundary value problems. Only when the variables given by the outside flow, e.g., W¯e\bar{W}_{e} and W¯o,\bar{W}_{o},\ satisfy some relations, can the half-space problem has a unique solution. What’s more, in the initial-boundary value problem, the boundary condition (3.5) should be equipped with the no mass flow condition w𝒆2=0w_{\boldsymbol{e}_{2}}=0 to determine the extra variable ρw\rho^{w} in beb_{e}. While in the half-space problem, w𝒆2=0w_{\boldsymbol{e}_{2}}=0 is automatically ensured by the system (2.6) and we do not need extra conditions at y=0y=0 other than (3.6).

Another possible choice of 𝑯\boldsymbol{H} is 𝑯=𝑴T𝑺1𝑴\boldsymbol{H}=\boldsymbol{M}^{T}\boldsymbol{S}^{-1}\boldsymbol{M}. The physical meaning of this construction is imposing alternative continuity of fluxes at the boundary. The condition can be put into the framework of Grad’s, i.e., testing the Maxwell boundary condition by some odd polynomials. But in this case the test functions are different from the ones in the Grad boundary condition. To see this, we first test the Maxwell boundary condition with ξ2ϕ𝜶,𝜶𝕀e\xi_{2}\phi_{\boldsymbol{\alpha}},\boldsymbol{\alpha}\in\mathbb{I}_{e} to get

𝑴(Wo+W¯obo)+χ^𝑺(We+W¯ebe)=0.\boldsymbol{M}(W_{o}+\bar{W}_{o}-b_{o})+\hat{\chi}\boldsymbol{S}(W_{e}+\bar{W}_{e}-b_{e})=0.

The test functions in the Grad boundary condition are ξ2ϕ𝜶,𝜶𝕀e,|𝜶|M1.\xi_{2}\phi_{\boldsymbol{\alpha}},\boldsymbol{\alpha}\in\mathbb{I}_{e},|\boldsymbol{\alpha}|\leq M-1. Alternatively, we can combine the polynomials ξ2ϕ𝜶,𝜶𝕀e,\xi_{2}\phi_{\boldsymbol{\alpha}},\boldsymbol{\alpha}\in\mathbb{I}_{e}, linearly to get

𝑴T𝑺1𝑴(Wo+W¯obo)+χ^𝑴T(We+W¯ebe)=0,\boldsymbol{M}^{T}\boldsymbol{S}^{-1}\boldsymbol{M}(W_{o}+\bar{W}_{o}-b_{o})+\hat{\chi}\boldsymbol{M}^{T}(W_{e}+\bar{W}_{e}-b_{e})=0,

which gives 𝑯=𝑴T𝑺1𝑴.\boldsymbol{H}=\boldsymbol{M}^{T}\boldsymbol{S}^{-1}\boldsymbol{M}. The construction is convenient to use for arbitrary order moment equations.

4 Details in the Proof

The section aims to fill in the missing details in the proof of Theorem 2.1. The proof mainly contains two steps, where we first explicitly write the analytical solution of (2.6) and then estimate it.

Lemma 4.1.

Under the assumption of Theorem 2.1, the system has a unique solution given by (4), (4.1) and (4.3), if and only if

rank(𝑩𝑻+)=n+.\mathrm{rank}(\boldsymbol{B}\boldsymbol{T}_{+})=n_{+}.
Proof.

Using Lemma 2.1, the system (2.6) is equivalent to

𝑼T𝑨𝑽d(𝑽TW)dy=𝑼T𝑸𝑽(𝑽TW)+𝑼Th.\boldsymbol{U}^{T}\boldsymbol{A}\boldsymbol{V}\dfrac{\mathrm{d}{(\boldsymbol{V}^{T}W)}}{\mathrm{d}{y}}=-\boldsymbol{U}^{T}\boldsymbol{Q}\boldsymbol{V}(\boldsymbol{V}^{T}W)+\boldsymbol{U}^{T}h.

Since 𝑸1j=𝟎\boldsymbol{Q}_{1j}=\boldsymbol{0}, the first pp lines of the system would give

𝑮T𝑨W(y)=y+𝑮Th(s)ds.\boldsymbol{G}^{T}\boldsymbol{A}W(y)=-\int_{y}^{+\infty}\!\!\boldsymbol{G}^{T}h(s)\,\mathrm{d}s. (4.1)

Since span{𝑽2}=span{𝑨𝑮}\mathrm{span}\{\boldsymbol{V}_{2}\}=\mathrm{span}\{\boldsymbol{A}\boldsymbol{G}\}, the formula (4.1) can uniquely determine the value of 𝑽2TW\boldsymbol{V}_{2}^{T}W.

Since rank(𝑨21)=rank(𝑼2)=r\mathrm{rank}(\boldsymbol{A}_{21})=\mathrm{rank}(\boldsymbol{U}_{2})=r, the matrix 𝑨21\boldsymbol{A}_{21} is invertible and the next rr lines of the system give

𝑽1TW(y)\displaystyle\boldsymbol{V}_{1}^{T}W(y) =\displaystyle= 𝑨211(𝑨22(𝑽2TW(y))+𝑨23(𝑽3TW(y)))\displaystyle-\boldsymbol{A}_{21}^{-1}\left(\boldsymbol{A}_{22}(\boldsymbol{V}_{2}^{T}W(y))+\boldsymbol{A}_{23}(\boldsymbol{V}_{3}^{T}W(y))\right)
+y+𝑨211(𝑸22(𝑽2TW(s))+𝑸23(𝑽3TW(s))𝑼2Th(s))ds.\displaystyle+\int_{y}^{+\infty}\boldsymbol{A}_{21}^{-1}\left(\boldsymbol{Q}_{22}(\boldsymbol{V}_{2}^{T}W(s))+\boldsymbol{Q}_{23}(\boldsymbol{V}_{3}^{T}W(s))-\boldsymbol{U}_{2}^{T}h(s)\right)\,\mathrm{d}s.

The formula (4) uniquely determines 𝑽1TW\boldsymbol{V}_{1}^{T}W if 𝑽3TW\boldsymbol{V}_{3}^{T}W and 𝑽2TW\boldsymbol{V}_{2}^{T}W are known.

The last NprN-p-r equations are separated alone as

𝑨33ddy(𝑽3TW)=𝑸33(𝑽3TW)+𝑽3Th𝑸32𝑽2TW𝑨32d𝑽2TWdy,\boldsymbol{A}_{33}\dfrac{\mathrm{d}}{\mathrm{d}{y}}(\boldsymbol{V}_{3}^{T}W)=-\boldsymbol{Q}_{33}(\boldsymbol{V}_{3}^{T}W)+\boldsymbol{V}_{3}^{T}h-\boldsymbol{Q}_{32}\boldsymbol{V}_{2}^{T}W-\boldsymbol{A}_{32}\dfrac{\mathrm{d}{\boldsymbol{V}_{2}^{T}W}}{\mathrm{d}{y}},

where 𝑽2TW\boldsymbol{V}_{2}^{T}W is given by (4.1). Denote by

h3=𝑸331(𝑽3Th𝑸32𝑽2TW𝑨32d𝑽2TWdy),h_{3}=\boldsymbol{Q}_{33}^{-1}\left(\boldsymbol{V}_{3}^{T}h-\boldsymbol{Q}_{32}\boldsymbol{V}_{2}^{T}W-\boldsymbol{A}_{32}\dfrac{\mathrm{d}{\boldsymbol{V}_{2}^{T}W}}{\mathrm{d}{y}}\right),

then we have

𝑸331𝑨33d𝑽3TWdy=𝑽3TW+h3,\boldsymbol{Q}_{33}^{-1}\boldsymbol{A}_{33}\dfrac{\mathrm{d}{\boldsymbol{V}_{3}^{T}W}}{\mathrm{d}{y}}=-\boldsymbol{V}_{3}^{T}W+h_{3},

where h3h_{3} is a given vector relying on hh.

From the eigenvalue decomposition (2.10) of 𝑸331𝑨33\boldsymbol{Q}_{33}^{-1}\boldsymbol{A}_{33}, we assume 𝑻1h3=[h+T,h0T,hT]T,\boldsymbol{T}^{-1}h_{3}=[h_{+}^{T},h_{0}^{T},h_{-}^{T}]^{T}, where the rows of h+,h0h_{+},h_{0} and hh_{-} are individually n+,n0n_{+},n_{0} and nn_{-}. We write

𝑽3TW=𝑻+z++𝑻0z0+𝑻z,\boldsymbol{V}_{3}^{T}W=\boldsymbol{T}_{+}z_{+}+\boldsymbol{T}_{0}z_{0}+\boldsymbol{T}_{-}z_{-}, (4.3)

then the characteristic equations give the solution

z0=h0(y),\displaystyle z_{0}=h_{0}(y), (4.4)
z=𝚲1y+exp(𝚲1(sy))h(s)ds,\displaystyle z_{-}=-\boldsymbol{\Lambda}_{-}^{-1}\int_{y}^{+\infty}\!\!\exp(\boldsymbol{\Lambda}_{-}^{-1}(s-y))h_{-}(s)\,\mathrm{d}s, (4.5)

and

z+=exp(𝚲+1y)z+(0)+𝚲+10yexp(𝚲+1(sy))h+(s)ds,z_{+}=\exp(-\boldsymbol{\Lambda}_{+}^{-1}y)z_{+}(0)+\boldsymbol{\Lambda}_{+}^{-1}\int_{0}^{y}\!\!\exp(\boldsymbol{\Lambda}_{+}^{-1}(s-y))h_{+}(s)\,\mathrm{d}s, (4.6)

where only z+(0)n+z_{+}(0)\in\mathbb{R}^{n_{+}} should be prescribed by the boundary condition.

The boundary condition (2.11) leads to

𝑩𝑻+z+(0)=g𝑩(𝑻0z0(0)+𝑻z(0)).\boldsymbol{B}\boldsymbol{T}_{+}z_{+}(0)=g-\boldsymbol{B}(\boldsymbol{T}_{0}z_{0}(0)+\boldsymbol{T}_{-}z_{-}(0)). (4.7)

According to unique solvability of the linear algebraic system, we can solve a unique z+(0)z_{+}(0) from (4.7) if and only if the first condition in Theorem 2.1 holds.

In a word, when the system is given, the variable 𝑽1TW\boldsymbol{V}_{1}^{T}W is determined. Then 𝑽3TW\boldsymbol{V}_{3}^{T}W is uniquely solvable if and only if the first condition in this lemma holds. Finally, 𝑽2TW\boldsymbol{V}_{2}^{T}W can be determined. ∎

To estimate the formal solutions (4), (4.1) and (4.3), we introduce some Poincare type inequalities. For any hh with ha<+\|h\|_{a}<+\infty, we define

r(y)=y+h(s)ds.r(y)=-\int_{y}^{+\infty}\!\!h(s)\,\mathrm{d}s. (4.8)

Then we have r(y)=h(y)r^{\prime}(y)=h(y) and for any component rir_{i} of rr, we have

|ri(y)|2y+e2asdsy+e2ashThds<+.|r_{i}(y)|^{2}\leq\int_{y}^{+\infty}\!\!e^{-2as}\,\mathrm{d}s\int_{y}^{+\infty}\!\!e^{2as}h^{T}h\,\mathrm{d}s<+\infty.
Lemma 4.2.

There exists a constant c0>0c_{0}>0 such that the Poincare inequality holds

rac0ra=c0ha,\|r\|_{a}\leq c_{0}\|r^{\prime}\|_{a}=c_{0}\|h\|_{a}, (4.9)

and the trace inequality holds

r(0)c0ha.\|r(0)\|\leq c_{0}\|h\|_{a}. (4.10)
Proof.

By definition, we have

e2ayrT(y)r(y)\displaystyle e^{2ay}r^{T}(y)r(y) =\displaystyle= y+s(e2asrTr(s))ds\displaystyle-\int_{y}^{+\infty}\!\!\partial_{s}(e^{2as}r^{T}r(s))\,\mathrm{d}s
=\displaystyle= 2ay+e2asrTr(s)ds2y+e2as(r)Tr(s)ds.\displaystyle-2a\int_{y}^{+\infty}\!\!e^{2as}r^{T}r(s)\,\mathrm{d}s-2\int_{y}^{+\infty}\!\!e^{2as}(r^{\prime})^{T}r(s)\,\mathrm{d}s.

Let y=0y=0 and use the Cauchy-Schwarz inequality, then we have

rT(0)r(0)+2ara22raha.\displaystyle r^{T}(0)r(0)+2a\|r\|_{a}^{2}\leq 2\|r\|_{a}\|h\|_{a}.

Since rT(0)r(0)0,r^{T}(0)r(0)\geq 0, we have

ra1aha.\|r\|_{a}\leq\frac{1}{a}\|h\|_{a}.

Analogously, we can ignore the term ra2\|r\|_{a}^{2} to get

rT(0)r(0)2aha.\sqrt{r^{T}(0)r(0)}\leq\sqrt{\frac{2}{a}}\|h\|_{a}.

This completes the proof. ∎

For convenience, we use aba\lesssim b to represent that there exists a constant c>0c>0 such that acb.a\leq cb. Because all the matrices considered here are finite-dimensional, with constant coefficients, they should have a uniform upper bound on matrix norms.

Lemma 4.3.

Under the assumption of Theorem 2.1, we suppose the first condition in Theorem 2.1 holds. Then the estimation (2.12) holds if and only if the second condition in Theorem 2.1 holds, i.e., 𝐁𝐓0=𝟎.\boldsymbol{B}\boldsymbol{T}_{0}=\boldsymbol{0}.

Proof.

Utilizing the boundedness of the matrices and Poincare inequality, from (4.1), we have

𝑽2TWa𝑮T𝑨Wa𝑮Thaha.\|\boldsymbol{V}_{2}^{T}W\|_{a}\lesssim\|\boldsymbol{G}^{T}\boldsymbol{A}W\|_{a}\lesssim\|\boldsymbol{G}^{T}h\|_{a}\lesssim\|h\|_{a}.

Analogously, the formula (4) gives

𝑽1TWa𝑽2TWa+𝑽3TWa+haha+𝑽3TWa.\|\boldsymbol{V}_{1}^{T}W\|_{a}\lesssim\|\boldsymbol{V}_{2}^{T}W\|_{a}+\|\boldsymbol{V}_{3}^{T}W\|_{a}+\|h\|_{a}\lesssim\|h\|_{a}+\|\boldsymbol{V}_{3}^{T}W\|_{a}.

The formula (4.3) gives

𝑽3TWaz0a+za+z+a,\|\boldsymbol{V}_{3}^{T}W\|_{a}\lesssim\|z_{0}\|_{a}+\|z_{-}\|_{a}+\|z_{+}\|_{a},

where z0z_{0} is given by (4.4),

z0ah3aha,\|z_{0}\|_{a}\lesssim\|h_{3}\|_{a}\lesssim\|h\|_{a},

and from (4.5), we have

zah3aha.\|z_{-}\|_{a}\lesssim\|h_{3}\|_{a}\lesssim\|h\|_{a}.

Due to the condition a<1/λmaxa<1/\lambda_{max}, the term exp(𝚲+1y)a\|\exp\left(-\boldsymbol{\Lambda}_{+}^{-1}y\right)\|_{a} is bounded and we have

limy+𝚲+10yexp(𝚲+1(sy))h+(s)ds=0.\lim_{y\rightarrow+\infty}\boldsymbol{\Lambda}_{+}^{-1}\int_{0}^{y}\!\!\exp(\boldsymbol{\Lambda}_{+}^{-1}(s-y))h_{+}(s)\,\mathrm{d}s=0.

So from (4.6), we have

z+az+(0)exp(𝚲+1y)a+haz+(0)+ha.\|z_{+}\|_{a}\lesssim\|z_{+}(0)\|\|\exp\left(-\boldsymbol{\Lambda}_{+}^{-1}y\right)\|_{a}+\|h\|_{a}\lesssim\|z_{+}(0)\|+\|h\|_{a}.

Now the key point lies in the estimation of z+(0)\|z_{+}(0)\|. We can inverse 𝑩𝑻+\boldsymbol{B}\boldsymbol{T}_{+} and solve z+(0)z_{+}(0) as

z+(0)=(𝑩𝑻+)1(g𝑩𝑻0z0(0)𝑩𝑻z(0)).z_{+}(0)=(\boldsymbol{B}\boldsymbol{T}_{+})^{-1}\left(g-\boldsymbol{B}\boldsymbol{T}_{0}z_{0}(0)-\boldsymbol{B}\boldsymbol{T}_{-}z_{-}(0)\right). (4.11)

If the second condition in Theorem 2.1 holds, then we have

z+(0)=(𝑩𝑻+)1(g𝑩𝑻z(0)),z_{+}(0)=(\boldsymbol{B}\boldsymbol{T}_{+})^{-1}\left(g-\boldsymbol{B}\boldsymbol{T}_{-}z_{-}(0)\right),

and

z+(0)g+z(0).\|z_{+}(0)\|\lesssim\|g\|+\|z_{-}(0)\|.

Due to the trace inequality, we can see from (4.5) that z(0)ha\|z_{-}(0)\|\lesssim\|h\|_{a}. Thus, we finally get the estimation (2.12).

On the contrary, if 𝑩𝑻0𝟎,\boldsymbol{B}\boldsymbol{T}_{0}\neq\boldsymbol{0}, we consider z0(0)\|z_{0}(0)\|. By definition, we choose h=𝑽3ch=\boldsymbol{V}_{3}c with c=c(y)n++n0+nc=c(y)\in\mathbb{R}^{n_{+}+n_{0}+n_{-}} such that

c=𝑸33𝑻[0c00],c0n0,c=\boldsymbol{Q}_{33}\boldsymbol{T}\begin{bmatrix}0\\ c_{0}\\ 0\end{bmatrix},\quad c_{0}\in\mathbb{R}^{n_{0}},

then we have z0=h0=c0.z_{0}=h_{0}=c_{0}. So z0(0)\|z_{0}(0)\| can not be controlled by ha\|h\|_{a}, i.e., for any M0>0M_{0}>0, there exists c0c_{0} such that ha=𝑽3ca=1\|h\|_{a}=\|\boldsymbol{V}_{3}c\|_{a}=1 but z0(0)=c0(0)>M0.\|z_{0}(0)\|=\|c_{0}(0)\|>M_{0}. From (4.11), since gg and z(0)z_{-}(0) are bounded, we can always find c0c_{0} such that ha=𝑽3ca=1\|h\|_{a}=\|\boldsymbol{V}_{3}c\|_{a}=1 but z+(0)>M0.\|z_{+}(0)\|>M_{0}. So the estimation (2.12) does not hold when 𝑩𝑻0𝟎.\boldsymbol{B}\boldsymbol{T}_{0}\neq\boldsymbol{0}.

In a word, we complete the proof. ∎

Theorem 2.1 is the combination of Lemma 4.1 and Lemma 4.3. From the proof of Lemma 4.3, we can see that 𝑽3TW(0)\boldsymbol{V}_{3}^{T}W(0) generally can not be controlled by ha\|h\|_{a} due to the contribution of z0(0)z_{0}(0). So is 𝑽1TW(0)\boldsymbol{V}_{1}^{T}W(0) from the formula (4). But from (4), we have

𝑨21𝑽1TW(0)+𝑨23𝑽3TW(0)ha\|\boldsymbol{A}_{21}\boldsymbol{V}_{1}^{T}W(0)+\boldsymbol{A}_{23}\boldsymbol{V}_{3}^{T}W(0)\|\lesssim\|h\|_{a}

because of the Poincare inequality.

5 Conclusions

We extended the well-posed conditions of the linear moment system in half-space to the non-homogeneous case. The stability of the non-homogeneous equations would ask for additional criteria compared to the homogeneous case. The Grad boundary condition was shown unstable and thus unreliable in reality. Some modified boundary conditions were proved well-posed, while the accuracy of these conditions should be studied further. Thanks to the analytical expressions of solutions, we can solve the non-homogeneous half-space problem efficiently.

Acknowledgments This work is financially supported by the National Key R&D Program of China, Project Number 2020YFA0712000.

Data Availibility This theoretical work uses no external dataset.

References

  • [1] A. Arnold and U. Giering. An analysis of the Marshak conditions for matching Boltzmann and Euler equations. Math. Models Methods Appl. Sci., 07(04):557–577, 1997.
  • [2] C. Bardos, F. Golse, and Y. Sone. Half-space problems for the Boltzmann equation: A survey. J. Stat. Phys., 124:275–300, 2006.
  • [3] N. Bernhoff. On half-space problems for the linearized discrete Boltzmann equation. Riv. Mat. Univ. Parma, 9:73–124, 2008.
  • [4] N. Bernhoff. On half-space problems for the discrete Boltzmann equation. Il Nuovo Cimento C, 33(1):47–54, 2010.
  • [5] N. Bernhoff. On half-space problems for the weakly non-linear discrete Boltzmann equation. Kinet. Relat. Models, 3:195 – 222, 2010.
  • [6] Z. Cai, Y. Fan, and R. Li. Globally hyperbolic regularization of Grad’s moment system. Comm. Pure Appl. Math., 67(3):464–518, 2014.
  • [7] Z. Cai, Y. Fan, and R. Li. A framework on moment model reduction for kinetic equation. SIAM J. Appl. Math., 75(5):2001–2023, 2015.
  • [8] Z. Cai, R. Li, and Z. Qiao. NRxxxx simulation of microflows with Shakhov model. SIAM J. Sci. Comput., 34(1):339–369, 2011.
  • [9] Z. Cai and Y. Wang. Regularized 13-moment equations for inverse power law models. J. Fluid Mech., 894:A12, 2020.
  • [10] C. Cercignani. The Boltzmann Equation and Its Applications. Springer-Verlag, 1989.
  • [11] H. Chen, Q. Li, and J. Lu. A numerical method for coupling the BGK model and Euler equations through the linearized Knudsen layer. J. Comput. Phys., 398:108893, 2019.
  • [12] Y. Fan, J. Li, R. Li, and Z. Qiao. Resolving Knudsen layer by high order moment expansion. Contin. Mech. Thermodyn., 31(5):1313–1337, 2019.
  • [13] H. Grad. Note on N-dimensional Hermite polynomials. Comm. Pure Appl. Math., 2(4):325–330, 1949.
  • [14] H. Grad. On the kinetic theory of rarefied gases. Comm. Pure Appl. Math., 2(4):331–407, 1949.
  • [15] E. Gross and S. Ziering. Kinetic theory of linear shear flow. Phys. Fluids, 1(3):215–224, 1958.
  • [16] X. Gu and D. Emerson. Linearized-moment analysis of the temperature jump and temperature defect in the Knudsen layer of a rarefied gas. Phys. Rev. E, 89(6):063020, 2014.
  • [17] D. Hilditch. An introduction to well-posedness and free-evolution. Int. J. Mod. Phys. A, 28(22n23):1340015, 2013.
  • [18] Z. Hu and G. Hu. An efficient steady-state solver for microflows with high-order moment model. J. Comput. Phys., 392:462–482, 2019.
  • [19] H. Kreiss. Initial boundary value problems for hyperbolic systems. Comm. Pure Appl. Math., 23(3):277–298, 1970.
  • [20] P. Lax and R. Phillips. Local boundary conditions for dissipative symmetric linear differential operators. Comm. Pure Appl. Math., 13(3):427–455, 1960.
  • [21] Q. Li, J. Lu, and W. Sun. A convergent method for linear half-space kinetic equations. ESAIM: Math. Model. Numer. Anal., 51(5):1583–1615, 2017.
  • [22] R. Li, W. Li, and L. Zheng. Direct flux gradient approximation to moment closure of kinetic equations. SIAM J. Appl. Math., 81(5):2153–2179, 2021.
  • [23] R. Li and Y. Yang. Linear moment models to approximate Knudsen layers. arXiv:2204.04941, 2022.
  • [24] S. Loyalka and J. Ferziger. Model dependence of the slip coefficient. Phys. Fluids, 10(8):1833–1839, 1967.
  • [25] S. Loyalka and J. Ferziger. Model dependence of the temperature slip coefficient. Phys. Fluids, 11(8):1668–1671, 1968.
  • [26] A. Majda and S. Osher. Initial-boundary value problems for hyperbolic equations with uniformly characteristic boundary. Comm. Pure Appl. Math., 28(5):607–675, 1975.
  • [27] J. Rauch. Symmetric positive systems with boundary characteristic of constant multiplicity. Trans. Amer. Math. Soc., 291(1):167–187, 1985.
  • [28] N. Sarna and M. Torrilhon. On stable wall boundary conditions for the Hermite discretization of the linearised Boltzmann equation. J. Stat. Phys., 170:101–126, 2018.
  • [29] Y. Sone. Molecular gas dynamics: Theory, techniques, and applications. Birkhäuser, Boston, 2007.
  • [30] H. Struchtrup. Linear kinetic heat transfer: Moment equations, boundary conditions, and Knudsen layers. Physica A, 387(8-9):1750–1766, 2008.
  • [31] M. Torrilhon. Special issues on moment methods in kinetic gas theory. Contin. Mech. Thermodyn., 21(5):341–343, 2009.