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

A revisit of the velocity averaging lemma: on the regularity of stationary Boltzmann equation in a bounded convex domain

I-Kun Chen (I.-K. Chen)Institute of Applied Mathematical Sciences, National Taiwan University, No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan ikun.chen@gmail.com Ping-Han Chuang (P.-H. Chuang)Department of Mathematics, National Taiwan University, No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan d09221003@ntu.edu.tw Chun-Hsiung Hsia (C.-H. Hsia)Institute of Applied Mathematical Sciences, National Taiwan University, No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan willhsia@math.ntu.edu.tw  and  Jhe-Kuan Su (J.-K. Su)Department of Mathematics, National Taiwan University, No. 1, Sec. 4, Roosevelt Rd., Taipei 10617, Taiwan hsnu127845@gmail.com
Abstract.

In the present work, we adopt the idea of velocity averaging lemma to establish regularity for stationary linearized Boltzmann equations in a bounded convex domain. Considering the incoming data, with three iterations, we establish regularity in fractional Sobolev space in space variable up to order 11^{-}.
Keywords:Keywords: Boltzmann equation; regularity; averaging lemmas; fractional Sobolev spaces

1. Introduction

The celebrated velocity averaging lemma reveals that the combination of transport and averaging in velocity yields regularity in space variable [15, 16]. This is one of the key features that DiPerna and Lions used to attack the Cauchy problem for Boltzmann equations [12]. It is natural to adopt this technique to the study of regularity problem of linearized Boltzmann equation in the whole space [9]. In [16], in addition to the applications to the whole space domains, the authors also investigate the applications to bounded convex domains for transport equations. By adopting zero extension, they reduce the bounded domain case to the whole space case. In contrast with the whole space case in [9], which the regularity can be improved indefinitely by iterations, when applying the trick in [16] in a bounded domain, one can only proceed for one iteration. Notice that the main tool of velocity averaging lemma, namely, the method of Fourier transform, does not translate well on a bounded domain. In this article, we adopt Slobodeckij semi-norm as an alternative concept of Sobolev function class. This shifts the difficulty to singular integrals. To calculate these integrals, we encounter some estimates related to the geometry of the boundary. The way we do it is to properly compare convex domains with spherical domains in 3\mathbb{R}^{3} so that we can build up estimates for convex domains based on those for spherical domains. Another key feature to prove the boundedness of the singular integrals is the change of variables which will be addressed in Lemma 7.1 and Lemma 7.2. Considering the incoming data, with three iterations, we establish regularity in fractional Sobolev space in space variable up to order 11^{-}.

Recall the velocity averaging lemma in [10, 16]. Suppose uu is an L2L^{2} solution to the transport equation

vxu=G(x,v),(x,v)n×n,v\cdot\nabla_{x}u=G(x,v),\ \ \ \ \ (x,v)\in\mathbb{R}^{n}\times\mathbb{R}^{n},

where GL2.G\in L^{2}. Let

u¯(x):=nu(x,v)ψ(v)𝑑v,\bar{u}(x):=\int_{\mathbb{R}^{n}}u(x,v)\psi(v)\,dv,

where ψ\psi is a bounded function with compact support. Then, we have

u¯(x)H~1/2(n).\bar{u}(x)\in\tilde{H}^{\nicefrac{{1}}{{2}}}(\mathbb{R}^{n}).

Here, the Sobolev space is generalized to non-integer order via the Fourier transform as follows.

Definition 1.1.

We say u:3u:\mathbb{R}^{3}\to\mathbb{R} is in H~xs(3)\tilde{H}^{s}_{x}(\mathbb{R}^{3}) if

(1.1) uH~xs(3)=(3(1+|ξ|2)s|(u)(ξ)|2𝑑ξ)12<,\|u\|_{\tilde{H}^{s}_{x}(\mathbb{R}^{3})}=\left(\int_{\mathbb{R}^{3}}(1+|\xi|^{2})^{s}|\mathcal{F}(u)(\xi)|^{2}\,d\xi\right)^{\frac{1}{2}}<\infty,

where (u)(ξ)\mathcal{F}(u)(\xi) is the Fourier transform of uu, i.e.,

(u)(ξ)=(2π)323u(x)eiξx𝑑x.\mathcal{F}(u)(\xi)=(2\pi)^{-\frac{3}{2}}\int_{\mathbb{R}^{3}}u(x)e^{-i\xi\cdot x}\,dx.

The velocity averaging lemma demonstrates that the regularity in the transport direction can be converted to the regularity in space variable after averaging with weight ψ\psi.

First, we recapitulate the stationary linearized Boltzmann equation in the whole space,

(1.2) vxf(x,v)=L(f),v\cdot\nabla_{x}f(x,v)=L(f),

where ff is the velocity distribution function and LL is the linearized collision operator. The linearized collision operator under consideration can be decomposed into a multiplicative operator and an integral operator.

(1.3) L(f)=ν(v)f+K(f),L(f)=-\nu(v)f+K(f),

where

K(f)=3k(v,v)f(v)𝑑v.K(f)=\int_{\mathbb{R}^{3}}k(v,v_{*})f(v_{*})\,dv_{*}.

Therefore, we can rewrite (1.3) as

(1.4) ν(v)f+vxf=K(f).\nu(v)f+v\cdot\nabla_{x}f=K(f).

Observing that the integral operator KK can serve as an agent of averaging, it is natural to imagine applying velocity averaging lemma to linearized Boltzmann equation. In case the source term Ψ(x,v)\Psi(x,v) is imposed, i.e.,

(1.5) ν(v)f+vxf=K(f)+Ψ(x,v),\nu(v)f+v\cdot\nabla_{x}f=K(f)+\Psi(x,v),

one can derive an integral equation

(1.6) f(x,v)=0eν(v)t[K(f)(xvt,v)+Ψ(xvt,v)]𝑑t=:S(K(f)+Ψ)=SK(f)+S(Ψ),\begin{split}f(x,v)&=\int_{0}^{\infty}e^{-\nu(v)t}[K(f)(x-vt,v)+\Psi(x-vt,v)]\,dt\\ &=:S(K(f)+\Psi)\\ &=SK(f)+S(\Psi),\end{split}

where

(1.7) S(h)(x,v):=0eν(v)th(xvt,t)𝑑t.S(h)(x,v):=\int^{\infty}_{0}e^{-\nu(v)t}h(x-vt,t)\,dt.

Performing the Picard iteration, formally we can derive that

(1.8) f=k=0S(KS)k(Ψ).f=\sum_{k=0}^{\infty}S(KS)^{k}(\Psi).

By carefully adapting the idea of velocity averaging lemma, we find every two iterations improve regularity in space of order 12\frac{1}{2}. More precisely, the following lemma was proved in [9].

Lemma 1.2.

The operator KSK:Lv2(3;H~xs(3))Lv2(3;H~xs+12(3))KSK:L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s}_{x}(\mathbb{R}^{3}))\to L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s+\frac{1}{2}}_{x}(\mathbb{R}^{3})) is bounded for any s0s\geq 0.

Here, the mixed fractional Sobolev space is defined as follows.

Definition 1.3.

We say u:3×3u:\mathbb{R}^{3}\times\mathbb{R}^{3}\to\mathbb{R} is in Lv2(3;H~xs(3))L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s}_{x}(\mathbb{R}^{3})) if

(1.9) uLv2(3;H~xs(3))=(33(1+|ξ|2)s|(u)(ξ,v)|2𝑑ξ𝑑v)12<,\|u\|_{L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s}_{x}(\mathbb{R}^{3}))}=\left(\int_{\mathbb{R}^{3}}\int_{\mathbb{R}^{3}}(1+|\xi|^{2})^{s}|\mathcal{F}(u)(\xi,v)|^{2}\,d\xi dv\right)^{\frac{1}{2}}<\infty,

where (u)(ξ,v)\mathcal{F}(u)(\xi,v) is the Fourier transform of uu with respect to space variable xx.

Motivated by the successful application of velocity averaging lemma to the study of regularity issue for stationary linearized Boltzmann equation in the whole space, we consider to give a similar account for the regularity problem in a bounded domain. However, we immediately notice that Definition 1.3 does not work for bounded space because that the Fourier transform is involved. For bounded domains, we adopt the fractional Sobolev space through the Slobodeckij semi-norm.

Definition 1.4.

Let s(0,1)s\in(0,1), Ω3\Omega\subset\mathbb{R}^{3} open. We say f(x,v)Lv2(3;Hxs(Ω))f(x,v)\in L^{2}_{v}(\mathbb{R}^{3};H^{s}_{x}(\Omega)) if fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}) and

(1.10) 3ΩΩ|f(x,v)f(y,v)|2|xy|3+2s𝑑x𝑑y𝑑v<,\int_{\mathbb{R}^{3}}\int_{\Omega}\int_{\Omega}\frac{|f(x,v)-f(y,v)|^{2}}{|x-y|^{3+2s}}\,dxdydv<\infty,

with

(1.11) fLv2(3;Hxs(Ω))=(fL2(Ω×3)2+3ΩΩ|f(x,v)f(y,v)|2|xy|3+2s𝑑x𝑑y𝑑v)12.\|f\|_{L^{2}_{v}(\mathbb{R}^{3};H^{s}_{x}(\Omega))}=\left(\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})}+\int_{\mathbb{R}^{3}}\int_{\Omega}\int_{\Omega}\frac{|f(x,v)-f(y,v)|^{2}}{|x-y|^{3+2s}}\,dxdydv\right)^{\frac{1}{2}}.

Notice that Definition 1.3 and Definition 1.4 of fractional Sobolev spaces are equivalent on the whole space. In other words, for 0<s<10<s<1, there exist two positive constants C1=C1(s)C_{1}=C_{1}(s) and C2=C2(s)C_{2}=C_{2}(s) such that

(1.12) C1uLv2(3;Hxs(3))uLv2(3;H~xs(3))C2uLv2(3;Hxs(3))C_{1}\|u\|_{L^{2}_{v}(\mathbb{R}^{3};H^{s}_{x}(\mathbb{R}^{3}))}\leq\|u\|_{L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s}_{x}(\mathbb{R}^{3}))}\leq C_{2}\|u\|_{L^{2}_{v}(\mathbb{R}^{3};H^{s}_{x}(\mathbb{R}^{3}))}

for any uLv2(3;H~xs(3))u\in L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s}_{x}(\mathbb{R}^{3})).

Here, we shall first introduce our main result and then explain the multiple obstacles we encounter and how we overcome them. We consider a bounded convex domain which satisfies the following assumption.

Definition 1.5.

We say a C2C^{2} bounded convex domain Ω\Omega in 3\mathbb{R}^{3} satisfies the positive curvature condition if Ω\partial\Omega is of positive Gaussian curvature.

Remark 1.6.

Positive curvature condition implies uniform convexity, which would also imply strict convexity. If the domain is compact, then its being strictly convex is equivalent to being uniformly convex. On the contrary, a uniformly convex domain does not necessarily satisfy the positive curvature condition.

We consider the incoming boundary value problem for linearized Boltzmann equation in Ω\Omega,

(1.13) {vxf(x,v)=L(f)(x,v),for xΩ,v3,f|Γ(q,v)=g(q,v),for (q,v)Γ,\left\{\begin{aligned} v\cdot\nabla_{x}f(x,v)&=L(f)(x,v),&\text{for }&x\in\Omega,\,v\in\mathbb{R}^{3},\\ f|_{\Gamma_{-}}(q,v)&=g(q,v),&\text{for }&(q,v)\in\Gamma_{-},\end{aligned}\right.

where

Γ:={(q,v)Ω×3:n(q)v<0},\Gamma_{-}:=\{(q,v)\in\partial\Omega\times\mathbb{R}^{3}:\,n(q)\cdot v<0\},

and n(q)n(q) is the unit outward normal of Ω\partial\Omega at qq. In this context, LL satisfies one of hard sphere, cutoff hard, and cutoff Maxwellian potentials. The detailed assumption on LL will be addressed in Section 3.

Regarding the existence result of boundary value problem (1.13), it has been studied by Guiraud [18] for convex domains and by Esposito, Guo, Kim, and Marra [14] for general domains. In the paper of Esposito, Guo, Kim, and Marra [14], they proved the solution is continuous away from the grazing set. With stronger assumption on cross-section BB, namely,

(1.14) B(|vv|,θ)=|vv|γβ(θ),0β(θ)Csinθcosθ,\begin{split}&B(|v-v_{*}|,\theta)=|v-v_{*}|^{\gamma}\beta(\theta),\\ &0\leq\beta(\theta)\leq C\sin\theta\cos\theta,\end{split}

the interior Hölder estimate was established in [6] and later improved to interior pointwise estimate for first derivatives [7]. Recently, the nonlinear case was established in [8] for hard sphere potential. Notice that, in [6, 7, 8], the fact KK improves regularity in velocity is a key property used. The idea is to move the regularity in velocity to space through transport and collision. This idea was inspired by the mixture lemma by Liu and Yu [22]. In contrast, in the present result, we do not need the smoothing effect of KK in velocity; the integral operator KK itself provides ”velocity averaging” and therefore regularity. Regarding regularity issues for the time dependent Boltzmann equation, we refer the interested readers to [19, 20].

In this article, we assume the following two conditions on the incoming data gg.

Assumption 1.7.

There are positive constants a,Ca,C such that

(1.15) |g(q1,v)|Cea|v|2|g(q_{1},v)|\leq C\,e^{-a|v|^{2}}

and

(1.16) |g(q1,v)g(q2,v)|C|q1q2|,|g(q_{1},v)-g(q_{2},v)|\leq C\,|q_{1}-q_{2}|,

for any (q1,v)Γ(q_{1},v)\in\Gamma_{-} and (q2,v)Γ(q_{2},v)\in\Gamma_{-}.

The main result of this paper is as follows.

Theorem 1.8.

Let bounded convex domain Ω3\Omega\subset\mathbb{R}^{3} satisfy the positive curvature condition in Definition 1.5 and linearized collision operator LL satisfy angular cutoff assumption (2.5). Suppose the incoming data gg satisfies Assumption 1.7. Then, for any solution fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}) to stationary linearized Boltzmann equation (1.13), we have

(1.17) fLv2(3;Hx1ϵ(Ω)),f\in L^{2}_{v}(\mathbb{R}^{3};{H}^{1-\epsilon}_{x}(\Omega)),

for any 0<ϵ<10<\epsilon<1.

We shall sketch the proof and reveal the difficulties induced by geometry and the method we tackle the problem.

Definition 1.9.

Let xΩx\in\Omega and v3v\in\mathbb{R}^{3}. We define

τ(x,v):=inft>0{t:xvtΩ},\displaystyle\tau_{-}(x,v):=\inf_{t>0}\{t:\,x-vt\notin\Omega\},
q(x,v):=xτ(x,v)v,\displaystyle q_{-}(x,v):=x-\tau_{-}(x,v)v,
τ+(x,v):=inft>0{t:x+vtΩ},\displaystyle\tau_{+}(x,v):=\inf_{t>0}\{t:\,x+vt\notin\Omega\},
q+(x,v):=x+τ+(x,v)v.\displaystyle q_{+}(x,v):=x+\tau_{+}(x,v)v.

With the notations of Definition 1.9, one can rewrite (1.13) as the integral equation

(1.18) f(x,v)=eν(v)τ(x,v)g(q(x,v),v)+0τ(x,v)eν(v)sK(f)(xsv,v)𝑑s.\begin{split}f(x,v)=&e^{-\nu(v)\tau_{-}(x,v)}g(q_{-}(x,v),v)\\ &+\int_{0}^{\tau_{-}(x,v)}e^{-\nu(v)s}K(f)(x-sv,v)\,ds.\end{split}

Hereafter, we define

(1.19) (Jg)(x,v)\displaystyle(Jg)(x,v) :=eν(v)τ(x,v)g(q(x,v),v),\displaystyle:=e^{-\nu(v)\tau_{-}(x,v)}g(q_{-}(x,v),v),
(1.20) (SΩf)(x,v)\displaystyle(S_{\Omega}f)(x,v) :=0τ(x,v)eν(v)sf(xsv,v)𝑑s.\displaystyle:=\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}f(x-sv,v)ds.

Notice that SΩ:Lp(Ω×3)Lp(Ω×3)S_{\Omega}:L^{p}(\Omega\times\mathbb{R}^{3})\to L^{p}(\Omega\times\mathbb{R}^{3}) and J:Lp(Γ;dσ)Lp(Ω×3)J:L^{p}(\Gamma_{-};d\sigma)\to L^{p}(\Omega\times\mathbb{R}^{3}) are bounded for 1p1\leq p\leq\infty with

dσ=|vn(q)|dΣ(q)dv,d\sigma=|v\cdot n(q)|\,d\Sigma(q)dv,

where Σ(q)\Sigma(q) is the surface element on Ω\partial\Omega at qq. Performing Picard iteration, we have

(1.21) f(x,v)=J(g)+SΩK(f)=J(g)+SΩKJ(g)+SΩKSΩK(f)=J(g)+SΩKJ(g)+SΩKSΩKJ(g)+SΩKSΩKSΩK(f)=J(g)+SΩKJ(g)+SΩKSΩKJ(g)+SΩKSΩKSΩKJ(g)+SΩKSΩKSΩKSΩK(f)=g0+g1+g2+g3+f4,\begin{split}f(x,v)=&J(g)+S_{\Omega}K(f)\\ =&J(g)+S_{\Omega}KJ(g)+S_{\Omega}KS_{\Omega}K(f)\\ =&J(g)+S_{\Omega}KJ(g)+S_{\Omega}KS_{\Omega}KJ(g)+S_{\Omega}KS_{\Omega}KS_{\Omega}K(f)\\ =&J(g)+S_{\Omega}KJ(g)+S_{\Omega}KS_{\Omega}KJ(g)+S_{\Omega}KS_{\Omega}KS_{\Omega}KJ(g)\\ &+S_{\Omega}KS_{\Omega}KS_{\Omega}KS_{\Omega}K(f)\\ =&g_{0}+g_{1}+g_{2}+g_{3}+f_{4},\end{split}

where

(1.22) gi\displaystyle g_{i} :=(SΩK)iJ(g),\displaystyle:=(S_{\Omega}K)^{i}J(g),
(1.23) fi\displaystyle f_{i} :=(SΩK)i(f).\displaystyle:=(S_{\Omega}K)^{i}(f).

We observe that each gig_{i} is directly under influence of boundary data and the geometry of the domain. Our strategy is to prove giLv2(3;Hx1ϵ(3))g_{i}\in L^{2}_{v}(\mathbb{R}^{3};{H}^{1-\epsilon}_{x}(\mathbb{R}^{3})). And, concerning the remaining term f4f_{4}, we shall match up the regularity of boundary terms.

We point out the difference between the cases for the whole space and a bounded domain. Let hh be any measurable function defined in Ω×3\Omega\times\mathbb{R}^{3}. We use the notation h~\widetilde{h} to denote its zero extension in 3×3\mathbb{R}^{3}\times\mathbb{R}^{3}. And, let Z:Ω×33×3Z:\Omega\times\mathbb{R}^{3}\to\mathbb{R}^{3}\times\mathbb{R}^{3} be the zero extension operator from Ω×3\Omega\times\mathbb{R}^{3} to 3×3\mathbb{R}^{3}\times\mathbb{R}^{3}, namely,

(1.24) (Zh)(x,v)=h~(x,v)={h(x,v),if xΩ,0,otherwise.(Zh)(x,v)=\widetilde{h}(x,v)=\begin{cases}h(x,v),&\quad\text{if }x\in\Omega,\\ 0,&\quad\text{otherwise.}\end{cases}

Suppose fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}). Notice that

(1.25) SK(f~)|Ω=SΩK(f).\left.SK(\widetilde{f})\right|_{\Omega}=S_{\Omega}K(f).

Therefore, applying Lemma 1.2, we have

Corollary 1.10.

The operator KSΩK:L2(Ω×3)Lv2(3;Hx1/2(Ω))KS_{\Omega}K:L^{2}(\Omega\times\mathbb{R}^{3})\to L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\Omega)) is bounded.

Remark 1.11.

Corollary 1.10 can be viewed as a variant of Theorem 4 in [16]. For the latter, the transport equation under consideration is, for given hL2(Ω×3)h\in L^{2}(\Omega\times\mathbb{R}^{3}),

(1.26) u(x,v)+vxu(x,v)=h(x,v),for xΩ,v3.u(x,v)+v\cdot\nabla_{x}u(x,v)=h(x,v),\,\text{for }x\in\Omega,\,v\in\mathbb{R}^{3}.

On the other hand, the transport equation we consider here is, for given hL2(Ω×3)h\in L^{2}(\Omega\times\mathbb{R}^{3}),

(1.27) ν(v)u(x,v)+vxu(x,v)=Kh(x,v),for xΩ,v3.\nu(v)u(x,v)+v\cdot\nabla_{x}u(x,v)=Kh(x,v),\,\text{for }x\in\Omega,\,v\in\mathbb{R}^{3}.

However, if we want to further iterate, we have to take the geometric structure into consideration. As mentioned earlier, the method of Fourier transform does not apply to bounded domains. In this article, we adopt Slobodeckij semi-norm as an alternative concept of Sobolev function class. As a result, we have

Lemma 1.12.

The operator SΩKSΩK:L2(Ω×3)Lv2(3;Hx1/2(Ω))S_{\Omega}KS_{\Omega}K:L^{2}(\Omega\times\mathbb{R}^{3})\to L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\Omega)) is bounded.

Therefore, if we end iteration at f2f_{2}, we can already claim fLv2(3;Hx1/2(Ω))f\in L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\Omega)).

Considering piling up the regularity, we notice that

(1.28) SKSK(f~)|ΩSΩKSΩK(f).\left.SKSK(\widetilde{f})\right|_{\Omega}\neq S_{\Omega}KS_{\Omega}K(f).

It seemingly comes to the limit of this strategy. Surprisingly, we have

Lemma 1.13.

ZSΩKSΩK:L2(Ω×3)Lv2(3;Hx12ϵ(3))ZS_{\Omega}KS_{\Omega}K:L^{2}(\Omega\times\mathbb{R}^{3})\to L^{2}_{v}(\mathbb{R}^{3};H^{\frac{1}{2}-\epsilon}_{x}(\mathbb{R}^{3})) is bounded for any ϵ(0,12)\epsilon\in(0,\frac{1}{2}). Furthermore, there is a constant CC independent of ϵ\epsilon and ff such that

(1.29) SΩKSΩKf~Lv2(3;Hx12ϵ(3))CϵfL2(Ω×3).\|\widetilde{S_{\Omega}KS_{\Omega}Kf}\|_{L^{2}_{v}(\mathbb{R}^{3};H^{\frac{1}{2}-\epsilon}_{x}(\mathbb{R}^{3}))}\leq\frac{C}{\sqrt{\epsilon}}\,\|f\|_{L^{2}(\Omega\times\mathbb{R}^{3})}.

That is, zero extension only reduces infinitesimal regularity. Therefore, after zero extension, we can repeat our strategy and obtain the desired result.

The rest of the article is organized as follows. Section 2 gives a brief overview of the Boltzmann equation and the details of our setting. In Section 3, we recall several basic properties of the linearized collision operator. In Section 4, we recapitulate the velocity averaging for linearized Boltzmann equation in the whole space. Section 5 provides some geometric properties for bounded convex domains that will be used in our estimates. In Section 6, we study the regularity of transport equation in bounded convex domains. Section 7 is devoted to the regularity via velocity averaging.

2. Boltzmann Equation

The Boltzmann equation reads

(2.1) tF+vxF=Q(F,F).\partial_{t}F+v\cdot\nabla_{x}F=Q(F,F).

Here, F=F(t,x,v)0F=F(t,x,v)\geq 0 is the distribution function for the particles located in the position xx with velocity vv at time tt. The collision operator QQ is defined as

(2.2) Q(F,G)=302π0π/2(F(v)G(v)F(v)G(v))B(|vv|,θ)𝑑θ𝑑ϵ𝑑v,Q(F,G)=\int\limits_{\mathbb{R}^{3}}\int\limits^{2\pi}_{0}\int\limits^{\nicefrac{{\pi}}{{2}}}_{0}\biggl{(}F(v^{\prime})G(v^{\prime}_{*})-F(v)G(v_{*})\biggr{)}B(|v-v_{*}|,\theta)\,d\theta d\epsilon dv_{*},

where vv^{\prime} and vv^{\prime}_{*} are the velocities after the elastic collision of two particles whose velocities are vv and vv_{*}, respectively, before the encounter. Here, the cross-section BB is chosen according to the type of interaction between particles. The precise form of cross-section BB depends on the model that is being studied. To specify the properties of the cross-section, we adopt the following coordinates. We set

e1=vv|vv|e_{1}=\frac{v_{*}-v}{|v_{*}-v|}

and choose e2𝕊2e_{2}\in\mathbb{S}^{2} and e3𝕊2e_{3}\in\mathbb{S}^{2} such that {e1,e2,e3}\{e_{1},e_{2},e_{3}\} forms an orthonormal basis for 3\mathbb{R}^{3}, and define

α=cosθe1+sinθcosϵe2+sinθsinϵe3.\alpha=\cos\theta e_{1}+\sin\theta\cos\epsilon e_{2}+\sin\theta\sin\epsilon e_{3}.

Then,

(2.3) v\displaystyle v^{\prime} =v+((vv)α)α,\displaystyle=v+\bigl{(}(v_{*}-v)\cdot\alpha\bigr{)}\alpha,
(2.4) v\displaystyle v^{\prime}_{*} =v((vv)α)α.\displaystyle=v_{*}-\bigl{(}(v_{*}-v)\cdot\alpha\bigr{)}\alpha.

Throughout this article, regarding the cross-section, we apply Grad’s angular cutoff potential [17] by assuming

(2.5) 0B(|vv|,θ)C|vv|γcosθsinθ,0\leq B(|v-v_{*}|,\theta)\leq C|v-v_{*}|^{\gamma}\cos\theta\sin\theta,

where, as mentioned above, γ\gamma depends on the model being studied. Our discussion includes hard sphere model (γ=1\gamma=1), cutoff hard potential (0<γ<10<\gamma<1), and cutoff Maxwellian molecular gases (γ=0\gamma=0). Consider the stationary solution F=F(x,v)F=F(x,v) as a perturbation of the standard Maxwellian,

M(v)=π32e|v|2,M(v)=\pi^{-\frac{3}{2}}e^{-|v|^{2}},

in the form

(2.6) F=M+M12f.F=M+M^{\frac{1}{2}}f.

Plugging the expression (2.6) into (2.1) and discarding the nonlinear term, we arrive at the stationary linearized Boltzmann equation

(2.7) vxf(x,v)=L(f)(x,v),v\cdot\nabla_{x}f(x,v)=L(f)(x,v),

with linearized collision operator LL, which reads

(2.8) L(f)=M12(Q(M,M12f)+Q(M12f,M)).L(f)=M^{-\frac{1}{2}}\bigl{(}Q(M,M^{\frac{1}{2}}f)+Q(M^{\frac{1}{2}}f,M)\bigr{)}.

Under the assumption (2.5), LL can be decomposed into a multiplicative operator and an integral operator, see [17],

(2.9) L(f)=ν(v)f+K(f).L(f)=-\nu(v)f+K(f).

Here, ν\nu is a function of velocity variable vv behaving like (1+|v|)γ(1+|v|)^{\gamma}, i.e., there exist two positive constants ν0\nu_{0} and ν1\nu_{1}, depending only on γ\gamma, such that

(2.10) 0<ν0(1+|v|)γ<ν(v)<ν1(1+|v|)γ,0<\nu_{0}(1+|v|)^{\gamma}<\nu(v)<\nu_{1}(1+|v|)^{\gamma},

for all v3v\in\mathbb{R}^{3}. The integral operator KK reads

K(f)(x,v)=3f(x,v)k(v,v)𝑑v,K(f)(x,v)=\int_{\mathbb{R}^{3}}f(x,v_{*})k(v_{*},v)\,dv_{*},

where the collision kernel kk is symmetric, that is, k(v,v)=k(v,v)k(v,v_{*})=k(v_{*},v). Notice that the assumption of the cross-section here is different from and more general than that in [5, 7]. The significant difference is that the operator KK in the case we consider does not guarantee to have regularity in velocity variables.

Under the decomposition (2.9), we then consider the boundary value problem

(2.11) {ν(v)f(x,v)+vxf(x,v)=K(f)(x,v),for xΩ,v3,f|Γ(q,v)=g(q,v),for (q,v)Γ.\left\{\begin{aligned} \nu(v)f(x,v)+v\cdot\nabla_{x}f(x,v)&=K(f)(x,v),&\text{for }&x\in\Omega,\,v\in\mathbb{R}^{3},\\ f|_{\Gamma_{-}}(q,v)&=g(q,v),&\text{for }&(q,v)\in\Gamma_{-}.\end{aligned}\right.

Let us pause here to make clear that how trace is being defined. We follow [2, 3, 21]. For fLp(Ω×3)f\in L^{p}(\Omega\times\mathbb{R}^{3}) satisfying the equation

(2.12) ν(v)f(x,v)+vxf(x,v)=K(f)(x,v)\nu(v)f(x,v)+v\cdot\nabla_{x}f(x,v)=K(f)(x,v)

in Ω×3\Omega\times\mathbb{R}^{3}, since K(f)Lp(Ω×3)K(f)\in L^{p}(\Omega\times\mathbb{R}^{3}) according to Proposition 3.3, we have

(2.13) vxf(x,v)=ν(v)f(x,v)+K(f)(x,v)Lxp(Ω)v\cdot\nabla_{x}f(x,v)=-\nu(v)f(x,v)+K(f)(x,v)\in L^{p}_{x}(\Omega)

for a.e. v3v\in\mathbb{R}^{3}. For such v3v\in\mathbb{R}^{3}, we fix qΓ(v)q\in\Gamma_{-}(v), where

Γ(v):={qΩ:(q,v)Γ},\Gamma_{-}(v):=\{q\in\partial\Omega:\,(q,v)\in\Gamma_{-}\},

and consider the function

(2.14) hq,v(s)=f(q+sv,v),h_{q,v}(s)=f(q+sv,v),

for 0<s<τ+(q,v)0<s<\tau_{+}(q,v), where τ+(q,v)\tau_{+}(q,v) is as defined in Definition 1.9. Noticing that

(2.15) dhq,v(s)ds=vxf(q+sv,v),\frac{dh_{q,v}(s)}{ds}=v\cdot\nabla_{x}f(q+sv,v),

and the formulas for change of variables,

(2.16) Ω|f(x,v)|p𝑑x\displaystyle\int_{\Omega}|f(x,v)|^{p}\,dx =Γ(v)0τ+(q,v)|f(q+sv,v)|p|vn(q)|𝑑s𝑑Σ(q)\displaystyle=\int_{\Gamma_{-}(v)}\int^{\tau_{+}(q,v)}_{0}|f(q+sv,v)|^{p}|v\cdot n(q)|\,dsd\Sigma(q)
(2.17) Ω|vxf(x,v)|p𝑑x\displaystyle\int_{\Omega}|v\cdot\nabla_{x}f(x,v)|^{p}\,dx =Γ(v)0τ+(q,v)|sf(q+sv,v)|p|vn(q)|𝑑s𝑑Σ(q),\displaystyle=\int_{\Gamma_{-}(v)}\int^{\tau_{+}(q,v)}_{0}|\frac{\partial}{\partial s}f(q+sv,v)|^{p}|v\cdot n(q)|\,dsd\Sigma(q),

we obtain hq,vWs1,p(0,τ+(q,v))h_{q,v}\in W^{1,p}_{s}(0,\tau_{+}(q,v)) for a.e. (q,v)Γ(q,v)\in\Gamma_{-}, which would imply hq,v(s)h_{q,v}(s) is an absolutely continuous function on (0,τ+(q,v))(0,\tau_{+}(q,v)) after possibly being redefined on a set of measure zero. As a result, we define

(2.18) f(q,v):=lims0+hq,v(s)=lims0+f(q+sv,v),f(q,v):=\lim_{s\to 0^{+}}h_{q,v}(s)=\lim_{s\to 0^{+}}f(q+sv,v),

for (q,v)Γ.(q,v)\in\Gamma_{-}.

3. Properties of the linearized collision operator

In this section, we review some basic properties of the linearized collision operator LL, see [1, 17]. As mentioned earlier, under the assumption (2.5), LL can be decomposed into a multiplicative operator and an integral operator,

(3.1) L(f)=ν(v)f+K(f).L(f)=-\nu(v)f+K(f).

There are two constants 0<ν0<ν10<\nu_{0}<\nu_{1} such that the following inequality holds for all v3v\in\mathbb{R}^{3},

(3.2) 0<ν0ν0(1+|v|)γ<ν(v)<ν1(1+|v|)γ.0<\nu_{0}\leq\nu_{0}(1+|v|)^{\gamma}<\nu(v)<\nu_{1}(1+|v|)^{\gamma}.

The integral operator KK is defined as

K(f)(x,v)=3f(x,v)k(v,v)𝑑v,K(f)(x,v)=\int_{\mathbb{R}^{3}}f(x,v_{*})k(v_{*},v)\,dv_{*},

and the collision kernel kk is symmetric. Furthermore, by Caflisch [1], we have an upper bound for kk,

(3.3) |k(v,v)|C1|vv|(1+|v|+|v|)(1γ)e18(|vv|2+(|v|2|v|2)2|vv|2),|k(v,v_{*})|\leq C\frac{1}{|v-v_{*}|}(1+|v|+|v_{*}|)^{-(1-\gamma)}e^{-\frac{1}{8}\left(|v-v_{*}|^{2}+\frac{(|v|^{2}-|v_{*}|^{2})^{2}}{|v-v_{*}|^{2}}\right)},

where C>0C>0 is a constant depending only on γ\gamma. The following lemma by Caflisch [1] is crucial in our estimates.

Lemma 3.1.

For any positive constants ϵ\epsilon, a1a_{1} and a2a_{2}, there exists C=C(ϵ,a1,a2)C=C(\epsilon,a_{1},a_{2}) depending only on ϵ\epsilon, a1a_{1} and a2a_{2} such that

31|vv|3ϵea1|vv|2a2(|v|2|v|2)2|vv|2𝑑v\displaystyle\int_{\mathbb{R}^{3}}\frac{1}{|v-v_{*}|^{3-\epsilon}}e^{-a_{1}|v-v_{*}|^{2}-a_{2}\frac{(|v|^{2}-|v_{*}|^{2})^{2}}{|v-v_{*}|^{2}}}\,dv_{*} C11+|v|\displaystyle\leq C\,\frac{1}{1+|v|}
C.\displaystyle\leq C.

The above inequality combining with (3.3) immediately implies the following facts.

Corollary 3.2.

For any v3v\in\mathbb{R}^{3},

(3.4) 3|k(v,v)|𝑑v\displaystyle\int_{\mathbb{R}^{3}}|k(v,v_{*})|\,dv_{*} C(11+|v|)2γC,\displaystyle\leq C\,\left(\frac{1}{1+|v|}\right)^{2-\gamma}\leq C,
(3.5) 3|k(v,v)|2𝑑v\displaystyle\int_{\mathbb{R}^{3}}|k(v,v_{*})|^{2}\,dv_{*} C(11+|v|)32γC.\displaystyle\leq C\,\left(\frac{1}{1+|v|}\right)^{3-2\gamma}\leq C.
Proposition 3.3.

The integral operator K:Lvp(3)Lvp(3)K:L^{p}_{v}(\mathbb{R}^{3})\to L^{p}_{v}(\mathbb{R}^{3}) is bounded for 1p1\leq p\leq\infty.

Proof.

If 1<p<1<p<\infty, for given hLvp(3)h\in L^{p}_{v}(\mathbb{R}^{3}), we have

(3.6) |Kh(v)|p=(3h(v)k(v,v)𝑑v)p(3|h(v)||k(v,v)|1p|k(v,v)|1p𝑑v)p(3|h(v)|p|k(v,v)|𝑑v)(3|k(v,v)|𝑑v)ppC3|h(v)|p|k(v,v)|𝑑v,\begin{split}|Kh(v)|^{p}&=\left(\int_{\mathbb{R}^{3}}h(v_{*})k(v_{*},v)\,dv_{*}\right)^{p}\\ &\leq\biggl{(}\int_{\mathbb{R}^{3}}|h(v_{*})||k(v_{*},v)|^{\frac{1}{p}}|k(v_{*},v)|^{\frac{1}{p^{\prime}}}\,dv_{*}\biggr{)}^{p}\\ &\leq\left(\int_{\mathbb{R}^{3}}|h(v_{*})|^{p}|k(v_{*},v)|\,dv_{*}\right)\left(\int_{\mathbb{R}^{3}}|k(v_{*},v)|\,dv_{*}\right)^{\frac{p}{p^{\prime}}}\\ &\leq C\int_{\mathbb{R}^{3}}|h(v_{*})|^{p}|k(v_{*},v)|\,dv_{*},\end{split}

where we have applied Corollary 3.2 in the above derivation. Therefore, applying Corollary 3.2 again, we obtain

(3.7) 3|Kh(v)|p𝑑vC33|h(v)|p|k(v,v)|𝑑v𝑑v=C33(|k(v,v)|dv)|h(v)|p𝑑vC3|h(v)|p𝑑v.\begin{split}\int_{\mathbb{R}^{3}}|Kh(v)|^{p}\,dv&\leq C\int_{\mathbb{R}^{3}}\int_{\mathbb{R}^{3}}|h(v_{*})|^{p}|k(v_{*},v)|\,dv_{*}dv\\ &=C\int_{\mathbb{R}^{3}}\int_{\mathbb{R}^{3}}\big{(}|k(v_{*},v)|\,dv\big{)}|h(v_{*})|^{p}\,dv_{*}\\ &\leq C\int_{\mathbb{R}^{3}}|h(v_{*})|^{p}dv_{*}.\end{split}

The proof for the cases where p=1,p=1,\infty is straightforward. We should omit it. ∎

We see from Proposition 3.3 KK is a bounded operator from Lvp(3)L^{p}_{v}(\mathbb{R}^{3}) to Lvp(3)L^{p}_{v}(\mathbb{R}^{3}). In particular, we shall only use the case p=2p=2.

Proposition 3.4.

For any v3v\in\mathbb{R}^{3} and ϵ(0,2)\epsilon\in(0,2), there is a constant CC independent of vv such that

(3.8) 31|v|2ϵ|k(v,v)|𝑑vC.\int_{\mathbb{R}^{3}}\frac{1}{|v_{*}|^{2-\epsilon}}\,|k(v,v_{*})|\,dv_{*}\leq C.
Proof.

Let v3v\in\mathbb{R}^{3} and ϵ(0,2)\epsilon\in(0,2) be given. From (3.3), we have

|k(v,v)|C1|vv|e18|vv|2.|k(v,v_{*})|\leq C\,\frac{1}{|v-v_{*}|}e^{-\frac{1}{8}|v-v_{*}|^{2}}.

It therefore suffices to prove that

(3.9) 31|v|2ϵ|vv|e18|vv|2𝑑vC\int_{\mathbb{R}^{3}}\frac{1}{|v_{*}|^{2-\epsilon}|v-v_{*}|}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}\leq C

for some CC. If |v||vv||v_{*}|\leq|v-v_{*}|, then

1|v|2ϵ|vv|1|v|3ϵ.\frac{1}{|v_{*}|^{2-\epsilon}|v-v_{*}|}\leq\frac{1}{|v_{*}|^{3-\epsilon}}.

If |v|>|vv||v_{*}|>|v-v_{*}|, then

1|v|2ϵ|vv|1|vv|3ϵ.\frac{1}{|v_{*}|^{2-\epsilon}|v-v_{*}|}\leq\frac{1}{|v-v_{*}|^{3-\epsilon}}.

In any case,

(3.10) 31|v|2ϵ|vv|e18|vv|2𝑑v31|v|3ϵe18|vv|2𝑑v+31|vv|3ϵe18|vv|2𝑑v.\begin{split}&\int_{\mathbb{R}^{3}}\frac{1}{|v_{*}|^{2-\epsilon}|v-v_{*}|}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}\\ &\quad\leq\int_{\mathbb{R}^{3}}\frac{1}{|v_{*}|^{3-\epsilon}}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}+\int_{\mathbb{R}^{3}}\frac{1}{|v-v_{*}|^{3-\epsilon}}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}.\end{split}

Clearly,

31|vv|3ϵe18|vv|2𝑑v=31|w|3ϵe18|w|2𝑑wC.\int_{\mathbb{R}^{3}}\frac{1}{|v-v_{*}|^{3-\epsilon}}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}=\int_{\mathbb{R}^{3}}\frac{1}{|w|^{3-\epsilon}}e^{-\frac{1}{8}|w|^{2}}\,dw\leq C.

Notice that

(3.11) 31|v|3ϵe18|vv|2𝑑v{|v|1}1|v|3ϵe18|vv|2𝑑v+{|v|>1}1|v|3ϵe18|vv|2𝑑v{|v|1}1|v|3ϵ𝑑v+3e18|vv|2𝑑vC.\begin{split}&\int_{\mathbb{R}^{3}}\frac{1}{|v_{*}|^{3-\epsilon}}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}\\ &\quad\leq\int_{\{|v_{*}|\leq 1\}}\frac{1}{|v_{*}|^{3-\epsilon}}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}+\int_{\{|v_{*}|>1\}}\frac{1}{|v_{*}|^{3-\epsilon}}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}\\ &\quad\leq\int_{\{|v_{*}|\leq 1\}}\frac{1}{|v_{*}|^{3-\epsilon}}\,dv_{*}+\int_{\mathbb{R}^{3}}e^{-\frac{1}{8}|v-v_{*}|^{2}}\,dv_{*}\\ &\quad\leq C.\end{split}

We deduce (3.9). ∎

In Section 6, we will need the following estimate. The case for hard sphere is proved in [4]. More generally, the proof therein can be extended to the cases for cutoff hard and cutoff Maxwellian potentials.

Lemma 3.5.

Suppose that a function hh satisfies the following estimate for some constant a[0,14)a\in[0,\frac{1}{4}),

(3.12) |h(v)|ea|v|2.|h(v)|\leq e^{-a|v|^{2}}.

Then there exists a positive constant CaC_{a} such that

(3.13) |K(h)(v)|Caea|v|2,|K(h)(v)|\leq C_{a}e^{-a|v|^{2}},

where CaC_{a} is a constant depending on aa.

4. Velocity averaging for linearized Boltzmann equation in the whole space

Lemma 1.2, which was first investigated in [9], plays an important role in our analysis. For readers’ convenience, we recapitulate the proof of Lemma 1.2 in this section. The idea of Lemma 1.2 originated from the velocity averaging lemma in [16], which can be proved by the method of Fourier transform. Roughly speaking, the L2L^{2} averaging lemma can be stated as that if uL2(n×n)u\in L^{2}(\mathbb{R}^{n}\times\mathbb{R}^{n}) is a solution to the transport equation in n×n\mathbb{R}^{n}\times\mathbb{R}^{n} with a source term fL2(n×n)f\in L^{2}(\mathbb{R}^{n}\times\mathbb{R}^{n}),

(4.1) u+vxu=f,u+v\cdot\nabla_{x}u=f,

then, for any ψLc(n)\psi\in L^{\infty}_{c}(\mathbb{R}^{n}), a bounded and compactly supported function, the velocity average of uu satisfies

(4.2) nu(,v)ψ(v)𝑑vH~1/2(n).\int_{\mathbb{R}^{n}}u(\cdot,v)\psi(v)\,dv\in\tilde{H}^{\nicefrac{{1}}{{2}}}(\mathbb{R}^{n}).

By analogy with the above result, we consider the transport equation in 3×3\mathbb{R}^{3}\times\mathbb{R}^{3} with a source term K(f)L2(3×3)K(f)\in L^{2}(\mathbb{R}^{3}\times\mathbb{R}^{3}),

(4.3) ν(v)u+vxu=K(f)\nu(v)u+v\cdot\nabla_{x}u=K(f)

(Notice that fL2(3×3)f\in L^{2}(\mathbb{R}^{3}\times\mathbb{R}^{3}) implies K(f)L2(3×3)K(f)\in L^{2}(\mathbb{R}^{3}\times\mathbb{R}^{3}) according to Proposition 3.3). Recall from (1.7)

(4.4) (Sf)(x,v):=0eν(v)sf(xsv,v)𝑑s.(Sf)(x,v):=\int^{\infty}_{0}e^{-\nu(v)s}f(x-sv,v)\,ds.

A direct calculation shows that SS is a bounded operator from Lp(3×3)L^{p}(\mathbb{R}^{3}\times\mathbb{R}^{3}) to Lp(3×3)L^{p}(\mathbb{R}^{3}\times\mathbb{R}^{3}) for 1p1\leq p\leq\infty and the L2L^{2} solution of (4.3) is expressed as

(4.5) u(x,v)=(SK)(f)(x,v).u(x,v)=(SK)(f)(x,v).

If we regard the integral operator KK as a kind of velocity averaging, then it is natural to anticipate K(u)=KSK(f)K(u)=KSK(f) has certain regularity in the space variable in view of (4.2). And it turns out we indeed have Lemma 1.2. In the following proof, we adopt the idea of Fourier transform in [16] with a careful application of Cauchy-Schwarz inequality. Furthermore, we also take advantage of (3.2) for the function ν\nu to gain some integrability (see (4.13)) since the support of k(v,v)k(v,v_{*}) is not bounded. We also remark that Lemma 1.2 holds for any function ν\nu satisfying (3.2) and any kernel kk satisfying (3.4) and (3.5).

Proof of Lemma 1.2.

We denote the Fourier transform in xx of a function hh by h\mathcal{F}h, the corresponding variable by ξ\xi. Since the operator KK does not involve the space vairable, a direct calculation shows that, for fL1(3×3)f\in L^{1}(\mathbb{R}^{3}\times\mathbb{R}^{3}), we have

(4.6) (Sf)(ξ,v)=\displaystyle\mathcal{F}(Sf)(\xi,v)= f(ξ,v)ν(v)+i(vξ),\displaystyle\frac{\mathcal{F}f(\xi,v)}{\nu(v)+i(v\cdot\xi)},
(4.7) (Kf)(ξ,v)=\displaystyle\mathcal{F}(Kf)(\xi,v)= K(f)(ξ,v).\displaystyle K(\mathcal{F}f)(\xi,v).

For the case where fL2(3×3)f\in L^{2}(\mathbb{R}^{3}\times\mathbb{R}^{3}), one can approximate ff by a sequence of functions in L1L2(3×3)L^{1}\cap L^{2}(\mathbb{R}^{3}\times\mathbb{R}^{3}) to obtain (4.6)-(4.7). To prove Lemma 1.2, it suffices to show the boundedness of integral

(4.8) I:=33|ξ|2s+1|KSKf(ξ,v)|2𝑑ξ𝑑vI:=\int_{\mathbb{R}^{3}}\int_{\mathbb{R}^{3}}|\xi|^{2s+1}\left|\mathcal{F}KSKf(\xi,v)\right|^{2}\,d\xi dv

for given fLv2(3;H~xs(3))f\in L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s}_{x}(\mathbb{R}^{3})). Using identities (4.6), (4.7) and Cauchy-Schwarz inequality yields

(4.9) I=33|ξ|2s+1|3SKf(ξ,v)k(v,v)𝑑v|2𝑑ξ𝑑v=33|ξ|2s+1|3Kf(ξ,v)ν(v)+i(vξ)k(v,v)𝑑v|2𝑑ξ𝑑v|ξ|2s+1(|Kf(ξ,v)|2(1+|v|)32γ|k(v,v)|𝑑v)×(|k(v,v)|(ν(v)2+(vξ)2)(1+|v|)32γ𝑑v)dξdv.\begin{split}I&=\int_{\mathbb{R}^{3}}\int_{\mathbb{R}^{3}}|\xi|^{2s+1}\left|\int_{\mathbb{R}^{3}}\mathcal{F}SKf(\xi,v_{*})k(v_{*},v)\,dv_{*}\right|^{2}\,d\xi dv\\ &=\int_{\mathbb{R}^{3}}\int_{\mathbb{R}^{3}}|\xi|^{2s+1}\left|\int_{\mathbb{R}^{3}}\frac{\mathcal{F}Kf(\xi,v_{*})}{\nu(v_{*})+i(v_{*}\cdot\xi)}k(v_{*},v)\,dv_{*}\right|^{2}\,d\xi dv\\ &\leq\iint|\xi|^{2s+1}\left(\int|\mathcal{F}Kf(\xi,v_{*})|^{2}(1+|v_{*}|)^{3-2\gamma}|k(v_{*},v)|\,dv_{*}\right)\\ &\quad\times\left(\int\frac{|k(v_{*},v)|}{\bigl{(}\nu(v_{*})^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}(1+|v_{*}|)^{3-2\gamma}}\,dv_{*}\right)d\xi dv.\end{split}

By Corollary 3.2, we have

(4.10) |Kf(ξ,v)|2=|3f(ξ,w)k(w,v)𝑑w|2(3|f(ξ,w)|2𝑑w)(3|k(w,v)|2𝑑w)C3|f(ξ,w)|2𝑑w(1+|v|)32γ,\begin{split}|\mathcal{F}Kf(\xi,v_{*})|^{2}=&\left|\int_{\mathbb{R}^{3}}\mathcal{F}f(\xi,w)k(w,v_{*})\,dw\right|^{2}\\ \leq&\left(\int_{\mathbb{R}^{3}}|\mathcal{F}f(\xi,w)|^{2}\,dw\right)\left(\int_{\mathbb{R}^{3}}|k(w,v_{*})|^{2}\,dw\right)\\ \leq&C\,\frac{\int_{\mathbb{R}^{3}}|\mathcal{F}f(\xi,w)|^{2}\,dw}{(1+|v_{*}|)^{3-2\gamma}},\end{split}

which leads to

(4.11) |Kf(ξ,v)|2(1+|v|)32γ|k(v,v)|𝑑vC|f(ξ,w)|2|k(v,v)|𝑑w𝑑vC|f(ξ,w)|2𝑑w,\begin{split}\int|\mathcal{F}Kf(\xi,v_{*})|^{2}(1+|v_{*}|)^{3-2\gamma}|k(v_{*},v)|\,dv_{*}\leq&C\,\iint|\mathcal{F}f(\xi,w)|^{2}|k(v_{*},v)|\,dwdv_{*}\\ \leq&C\,\int|\mathcal{F}f(\xi,w)|^{2}\,dw,\end{split}

where the second inequality follows from Corollary 3.2. Therefore, it follows that

(4.12) IC|ξ|2s+1|f(ξ,w)|2|k(v,v)|(ν(v)2+(vξ)2)(1+|v|)32γ𝑑v𝑑v𝑑ξ𝑑w|ξ|2s+1|f(ξ,w)|21(ν(v)2+(vξ)2)(1+|v|)53γ𝑑v𝑑ξ𝑑w,\begin{split}I\leq&C\,\iiiint|\xi|^{2s+1}|\mathcal{F}f(\xi,w)|^{2}\frac{|k(v_{*},v)|}{\bigl{(}\nu(v_{*})^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}(1+|v_{*}|)^{3-2\gamma}}\,dvdv_{*}d\xi dw\\ \leq&\iiint|\xi|^{2s+1}|\mathcal{F}f(\xi,w)|^{2}\frac{1}{\bigl{(}\nu(v_{*})^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}(1+|v_{*}|)^{5-3\gamma}}\,dv_{*}d\xi dw,\end{split}

where we have used Corollary 3.2 again in the last line. We observe that

(4.13) 1ν(v)2+(vξ)2=1(ν(v)2+(vξ)2)2/31(ν(v)2+(vξ)2)1/31(ν02+(vξ)2)2/31ν(v)2/3C1(ν02+(vξ)2)2/31(1+|v|)2γ/3.\begin{split}\frac{1}{\nu(v_{*})^{2}+(v_{*}\cdot\xi)^{2}}=&\frac{1}{\bigl{(}\nu(v_{*})^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}^{2/3}}\cdot\frac{1}{\bigl{(}\nu(v_{*})^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}^{1/3}}\\ \leq&\frac{1}{\bigl{(}\nu_{0}^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}^{2/3}}\cdot\frac{1}{\nu(v_{*})^{\nicefrac{{2}}{{3}}}}\\ \leq&C\,\frac{1}{\bigl{(}\nu_{0}^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}^{2/3}}\cdot\frac{1}{(1+|v_{*}|)^{\nicefrac{{2\gamma}}{{3}}}}.\end{split}

As a result, we have

(4.14) IC|ξ|2s+1|f(ξ,w)|2(ν02+(vξ)2)2/3(1+|v|)57γ/3𝑑v𝑑ξ𝑑w.I\leq C\iiint|\xi|^{2s+1}\frac{|\mathcal{F}f(\xi,w)|^{2}}{\bigl{(}\nu_{0}^{2}+(v_{*}\cdot\xi)^{2}\bigr{)}^{2/3}(1+|v_{*}|)^{5-\nicefrac{{7\gamma}}{{3}}}}\,dv_{*}d\xi dw.

Denote the component of vv_{*} parallel to ξ\xi and the component perpendicular to ξ\xi respectively by

(4.15) {vξ=vξ|ξ|vξ=v(vξ)ξ|ξ|.\left\{\begin{array}[]{l}v_{*\parallel\xi}=v_{*}\cdot\frac{\xi}{|\xi|}\\ v_{*\perp\xi}=v_{*}-(v_{*\parallel\xi})\frac{\xi}{|\xi|}.\end{array}\right.

Consequently, we deduce

(4.16) IC|ξ|2s+1|f(ξ,w)|2(ν02+|ξ|2vξ2)2/3(1+|vξ|)57γ/3𝑑vξdξ𝑑ξ𝑑wC|ξ|2s+1|f(ξ,w)|2(ν02+|ξ|2vξ2)2/3𝑑vξ𝑑ξ𝑑wC|ξ|2s|f(ξ,w)|2𝑑ξ𝑑wCfLv2(3;H~xs(3))2,\begin{split}I\leq&C\,\iiiint|\xi|^{2s+1}\frac{|\mathcal{F}f(\xi,w)|^{2}}{\bigl{(}\nu_{0}^{2}+|\xi|^{2}v^{2}_{*\parallel\xi}\bigr{)}^{2/3}(1+|v_{*\perp\xi}|)^{5-\nicefrac{{7\gamma}}{{3}}}}\,dv_{*\perp\xi}d_{*\parallel\xi}d\xi dw\\ \leq&C\,\iiint|\xi|^{2s+1}\frac{|\mathcal{F}f(\xi,w)|^{2}}{\bigl{(}\nu_{0}^{2}+|\xi|^{2}v^{2}_{*\parallel\xi}\bigr{)}^{2/3}}\,dv_{*\parallel\xi}d\xi dw\\ \leq&C\,\iint|\xi|^{2s}|\mathcal{F}f(\xi,w)|^{2}\,d\xi dw\\ \leq&C\|f\|^{2}_{L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{s}_{x}(\mathbb{R}^{3}))},\end{split}

where the second inequality follows since 57γ3>25-\frac{7\gamma}{3}>2 for 0γ10\leq\gamma\leq 1 and vξv_{*\perp\xi} is two-dimensional plane. This completes the proof. ∎

5. Geometric properties of bounded convex domains

In this section, we introduce some auxiliary geometric results involving bounded convex domains. Throughout this section, Ω3\Omega\subset\mathbb{R}^{3} denotes a C2C^{2} bounded strictly convex domain. Our strategy is to properly compare convex domain Ω\Omega with spherical domains in 3\mathbb{R}^{3} so that we can build up estimates for convex domains based on those for spherical domains.

Let n(q)n(q) denote the unit outward normal of Ω\partial\Omega at qΩq\in\partial\Omega, and v^=v|v|\hat{v}=\frac{v}{|v|} denote the unit vector with the direction v3v\in\mathbb{R}^{3}. We start with several geometric notations we shall frequently use.

Definition 5.1.

Let diamΩ\operatorname{diam}\Omega be the diameter of bounded domain Ω\Omega. Namely, we define

(5.1) diamΩ:=sup(x1,x2)Ω×Ω|x1x2|.\operatorname{diam}\Omega:=\sup_{(x_{1},x_{2})\in\Omega\times\Omega}|x_{1}-x_{2}|.

For an interior point xΩx\in\Omega, let

(5.2) dx:=d(x,Ω)=infqΩ|xq|\begin{split}d_{x}:&=\operatorname{d}(x,\partial\Omega)\\ &=\inf_{q\in\partial\Omega}|x-q|\end{split}

be the distance from xx to Ω\partial\Omega. For xΩ¯x\in\bar{\Omega} and v3v\in\mathbb{R}^{3}, we define τ(x,v)\tau_{-}(x,v) to be the backward exit time for xx leaving Ω\Omega with velocity v-v, and we define q(x,v)q_{-}(x,v) to be the corresponding point that the backward trajectory touches Ω\partial\Omega. More precisely,

τ(x,v)\displaystyle\tau_{-}(x,v) :=inf{t>0:xtvΩ},\displaystyle:=\inf\{t>0:\,x-tv\notin\Omega\},
q(x,v)\displaystyle q_{-}(x,v) :=xτ(x,v)v.\displaystyle:=x-\tau_{-}(x,v)v.

Furthermore, the absolute value of the component of velocity vv passing through the surface Ω\partial\Omega at q(x,v)q_{-}(x,v) is denoted by

(5.3) N(x,v):=|n(q(x,v))v^|.N_{-}(x,v):=|n(q_{-}(x,v))\cdot\hat{v}|.

In a similar fashion, we can define the corresponding forward concept by

τ+(x,v)\displaystyle\tau_{+}(x,v) :=inf{t>0:x+tvΩ},\displaystyle:=\inf\{t>0:\,x+tv\notin\Omega\},
q+(x,v)\displaystyle q_{+}(x,v) :=x+τ+(x,v)v,\displaystyle:=x+\tau_{+}(x,v)v,
N+(x,v)\displaystyle N_{+}(x,v) :=|n(q+(x,v))v^|.\displaystyle:=|n(q_{+}(x,v))\cdot\hat{v}|.

The following two propositions from [7] concern estimates for backward trajectory.

Proposition 5.2.

Let xx be an interior point of Ω\Omega and v3v\in\mathbb{R}^{3}. Then

|xq(x,v)|dxN(x,v).|x-q_{-}(x,v)|\geq\frac{d_{x}}{N_{-}(x,v)}.
Proposition 5.3.

Let x,yx,y be interior points of Ω\Omega and v3v\in\mathbb{R}^{3}. If
|xq(x,v)||yq(y,v)||x-q_{-}(x,v)|\leq|y-q_{-}(y,v)|, then

(5.4) |q(x,v)q(y,v)||xy|N(x,v),\displaystyle|q_{-}(x,v)-q_{-}(y,v)|\leq\frac{|x-y|}{N_{-}(x,v)},
(5.5) ||xq(x,v)||yq(y,v)||2|xy|N(x,v).\displaystyle\biggl{|}|x-q_{-}(x,v)|-|y-q_{-}(y,v)|\biggr{|}\leq\frac{2|x-y|}{N_{-}(x,v)}.
Definition 5.4.

We say a bounded convex domain Ω\Omega satisfies uniform interior sphere condition (resp. uniform sphere-enclosing condition) if there is a constant r1=r1(Ω)>0r_{1}=r_{1}(\Omega)>0 (resp. R1=R1(Ω)>0R_{1}=R_{1}(\Omega)>0) such that for any boundary point qΩq\in\partial\Omega, there is a sphere Si(q)=Br1(y)S_{i}(q)=\partial B_{r_{1}}(y) (resp. So(q)=BR1(y)S_{o}(q)=\partial B_{R_{1}}(y^{\prime})) such that Br1(y)ΩB_{r_{1}}(y)\subset\Omega (resp. ΩBR1(y)\Omega\subset B_{R_{1}}(y^{\prime})) with qSi(q)q\in S_{i}(q) (resp. qSo(q)q\in S_{o}(q)). (See Figure 1.)

Refer to caption
Figure 1.
Proposition 5.5.

Let 𝒟2\mathcal{D}\subset\mathbb{R}^{2} be a disk of radius rr and A,BA,B two points on 𝒟\partial\mathcal{D}. Suppose there is a regular parametrized curve α(s)\alpha(s) connecting α(s0)=A\alpha(s_{0})=A to α(s1)=B\alpha(s_{1})=B contained in the circular segment bounded by chord AB¯\overline{AB} and minor arc AB\stackrel{{\scriptstyle\frown}}{{AB}}. We further assume that α(s)\alpha(s) is convex, i.e. the curvature k(s)k(s) at α(s)\alpha(s) is positive everywhere, α(s)\alpha(s) is tangent to 𝒟\mathcal{D} at AA, and α(s)\alpha(s) leaves 𝒟\mathcal{D} at BB as Figure 2 shows. Then, there exists s(s0,s1)s_{*}\in(s_{0},s_{1}) such that k(s)1rk(s_{*})\leq\frac{1}{r}.

Refer to caption
Figure 2.
Proof.

We may assume α(s)\alpha(s) is parametrized by arc length. Let e1e_{1} denote the vector AB|AB|\frac{\overrightarrow{AB}}{|A-B|}. We choose e2𝕊1e_{2}\in\mathbb{S}^{1} such that {e1,e2}\{e_{1},e_{2}\} forms a positively oriented orthonormal basis. Denote the signed angle from vector e1e_{1} to tangent vector α(s)\alpha^{\prime}(s) by θ(s)\theta(s). Therefore, |AB|=2r|sinθ(s0)||A-B|=2r|\sin\theta(s_{0})|. On the other hand, since α(s)=cosθ(s)e1+sinθ(s)e2\alpha^{\prime}(s)=\cos\theta(s)e_{1}+\sin\theta(s)e_{2}, we obtain

(5.6) 2r|sinθ(s0)|=|AB|=s0s1α(s)e1𝑑s=s0s1cosθ(s)𝑑s.2r|\sin\theta(s_{0})|=|A-B|=\int^{s_{1}}_{s_{0}}\alpha^{\prime}(s)\cdot e_{1}\,ds=\int^{s_{1}}_{s_{0}}\cos\theta(s)\,ds.

Suppose k(s)>1rk(s)>\frac{1}{r} for any s(s0,s1)s\in(s_{0},s_{1}). Then, for any s(s0,s1)s\in(s_{0},s_{1}),

(5.7) dθ(s)ds=k(s)>1r,\frac{d\theta(s)}{ds}=k(s)>\frac{1}{r},

or

(5.8) ds(θ)dθ<r.\frac{ds(\theta)}{d\theta}<r.

Hence, we have

(5.9) 2r|sinθ(s0)|=s0s1cosθ(s)𝑑s=θ0θ1cosθds(θ)dθdθ<rθ0θ1cosθdθ=r(sinθ1sinθ0),\begin{split}2r|\sin\theta(s_{0})|=&\int^{s_{1}}_{s_{0}}\cos\theta(s)\,ds\\ =&\int^{\theta_{1}}_{\theta_{0}}\cos\theta\frac{ds(\theta)}{d\theta}\,d\theta\\ <&r\,\int^{\theta_{1}}_{\theta_{0}}\cos\theta\,d\theta\\ =&r\,(\sin\theta_{1}-\sin\theta_{0}),\end{split}

where θ0=θ(s0)\theta_{0}=\theta(s_{0}) and θ1=θ(s1)\theta_{1}=\theta(s_{1}). Since π2|θ0|θ1>0\frac{\pi}{2}\geq|\theta_{0}|\geq\theta_{1}>0, we notice that

(5.10) r(sinθ1sinθ0)2r|sinθ0|,r\,(\sin\theta_{1}-\sin\theta_{0})\leq 2r|\sin\theta_{0}|,

which leads to a contradiction. ∎

Proposition 5.6.

If Ω\Omega satisfies the positive curvature condition in Definition 1.5, then it also satisfies both uniform interior sphere condition and uniform sphere-enclosing condition.

Proof.

Since Ω\partial\Omega is C2C^{2} and compact, there is a tubular neighborhood near Ω\partial\Omega. That is, there exists a number ϵ>0\epsilon>0 such that whenever q1,q2Ωq_{1},q_{2}\in\partial\Omega the segments of the normal lines of length 2ϵ2\epsilon, centered at q1q_{1} and q2q_{2}, are disjoint. This fact can be found in [13], for example. Let r1=ϵ2r_{1}=\frac{\epsilon}{2}. For a given point qΩq\in\partial\Omega, therefore we can consider the sphere Br1(y)ΩB_{r_{1}}(y)\subset\Omega such that qBr1(y)q\in\partial B_{r_{1}}(y). Clearly, Si(q)=Br1(y)S_{i}(q)=\partial B_{r_{1}}(y) satisfies the desired property.

Concerning uniform sphere-enclosing condition, we shall proceed the proof by contradiction. We denote the principal curvatures of Ω\partial\Omega at qq by k1(q)k_{1}(q) and k2(q)k_{2}(q) (k1(q)k2(q)k_{1}(q)\leq k_{2}(q)). By positive curvature property of Ω\partial\Omega, we can choose k0k_{0} such that 0<k0<minqΩk1(q)0<k_{0}<\min_{q\in\partial\Omega}k_{1}(q). We claim that R1=1k0R_{1}=\frac{1}{k_{0}} is a desired radius for the uniform sphere-enclosing condition. For a given point qΩq\in\partial\Omega, there is yy^{\prime} such that yq//n(q)\overrightarrow{y^{\prime}q}//n(q) and |yq|=R1|y^{\prime}-q|=R_{1}. Under suitable rotations and translations, Ω\partial\Omega can be locally expressed as the graph of a function z=h(x,y)z=h(x,y) at q=(0,0,0)q=(0,0,0), where

(5.11) h(x,y)=12(k1(q)x2+k2(q)y2)+o(x2+y2).h(x,y)=\frac{1}{2}(k_{1}(q)x^{2}+k_{2}(q)y^{2})+o(x^{2}+y^{2}).

Since BR1(y)B_{R_{1}}(y^{\prime}) has constant normal curvature k0k_{0}, BR1(y)¯\overline{B_{R_{1}}(y^{\prime})} contains a neighborhood of qq in Ω\partial\Omega by (5.11). Suppose ΩBR1(y)\Omega\not\subset B_{R_{1}}(y^{\prime}), then there exists q1ΩBR1(y)q_{1}\in\partial\Omega\cap\partial B_{R_{1}}(y^{\prime}) such that the plane EE containing q,y,q1q,y^{\prime},q_{1} has the intersection curve with Ω\partial\Omega satisfies the condition of Proposition 5.5 with A=qA=q, B=q1B=q_{1}, and r=R1r=R_{1}. Then Proposition 5.5 implies that there is a point qΩq_{*}\in\partial\Omega such that k(s)1R1k(s_{*})\leq\frac{1}{R_{1}}, where q=α(s)q_{*}=\alpha(s_{*}) under the notation from Proposition 5.5. Hence, we note that the normal curvature kn(q)k_{n}(q_{*}) of Ω\partial\Omega at qq_{*} satisfies

(5.12) kn(q)k(q)1R1=k0,k_{n}(q_{*})\leq k(q_{*})\leq\frac{1}{R_{1}}=k_{0},

which is a contradiction. Therefore, we define So(q)=BR1(y)S_{o}(q)=\partial B_{R_{1}}(y^{\prime}). This completes the proof. ∎

Remark 5.7.

The above proof shows that 1minqΩk1(q)ϵ\frac{1}{\min\limits_{q\in\partial\Omega}k_{1}(q)-\epsilon} is a uniform radius for sphere-enclosing condition for every small ϵ>0\epsilon>0. Letting ϵ0+\epsilon\to 0^{+}, we obtain the smallest and optimal radius R1=1minqΩk1(q)R_{1}=\frac{1}{\min\limits_{q\in\partial\Omega}k_{1}(q)}.

The following lemma in [7] is useful in our estimates.

Lemma 5.8.

Suppose Ω\Omega satisfies the positive curvature condition in Definition 1.5. Then there exists a constant C=C(Ω)C=C(\Omega) such that, for any interior point xΩx\in\Omega, we have

(5.13) Ω1|xq|2𝑑Σ(q)C(|log(dx)|+1),\int_{\partial\Omega}\frac{1}{|x-q|^{2}}\,d\Sigma(q)\leq C(|\log(d_{x})|+1),

where Σ(q)\Sigma(q) is the surface element of Ω\partial\Omega at point qΩq\in\partial\Omega.

The following proposition concerns an estimate for chords in a bounded convex domain.

Proposition 5.9.

For a given bounded convex domain Ω\Omega satisfying positive curvature condition in Definition 1.5, there exists a constant C=C(Ω)C=C(\Omega) such that for any xΩ¯x\in\bar{\Omega} and v3v\in\mathbb{R}^{3}, we have

(5.14) |q(x,v)q+(x,v)|CN(x,v).|q_{-}(x,v)-q_{+}(x,v)|\leq CN_{-}(x,v).
Proof.

For given xΩ¯x\in\bar{\Omega} and v3v\in\mathbb{R}^{3}, write q=q(x,v)q_{-}=q_{-}(x,v), q+=q+(x,v)q_{+}=q_{+}(x,v), and θ=arccosN(x,v)\theta_{-}=\arccos N_{-}(x,v) for simplicity. In view of Proposition 5.6, there exists sphere So(q)S_{o}(q_{-}) as defined in Definition 5.4. Denote the other intersection of half-line qq+\overrightarrow{q_{-}q_{+}} and So(q)S_{o}(q_{-}) by q1q_{1}. Therefore, we have

(5.15) |qq+||qq1|=2R1cosθ=2R1N(x,v).\begin{split}|q_{-}-q_{+}|&\leq|q_{-}-q_{1}|\\ &=2R_{1}\cos\theta_{-}\\ &=2R_{1}N_{-}(x,v).\end{split}

Remark 5.10.

From the above proof, we have

(5.16) |qq+|2R1cosθ.|q_{-}-q_{+}|\leq 2R_{1}\cos\theta_{-}.

Replacing vv with v-v in Proposition 5.9 yields

(5.17) |qq+|2R1cosθ+,|q_{-}-q_{+}|\leq 2R_{1}\cos\theta_{+},

where θ+=arccosN+(x,v)\theta_{+}=\arccos N_{+}(x,v).

Proposition 5.11.

For a given circle 𝒞\mathcal{C} in 2\mathbb{R}^{2} centered at OO with radius rr and two given points A,BA,B on 𝒞\mathcal{C}. Let NN be the arc midpoint of AB\stackrel{{\scriptstyle\frown}}{{AB}} (the minor arc) and MM be the midpoint of AB¯\overline{AB}. Then for any YAM¯Y\in\overline{AM} (resp. YBM¯Y\in\overline{BM}), we have

(5.18) d(Y,𝒞)12|ZY|,\operatorname{d}(Y,\mathcal{C})\geq\frac{1}{\sqrt{2}}|Z-Y|,

where ZZ is the point on AN¯\overline{AN} (resp. BN¯\overline{BN}) such that (ZY)(AB)(Z-Y)\perp(A-B) (See Figure 3).

Proof.

Without loss of generality, we may assume YAM¯Y\in\overline{AM}. Let θ\theta denote the angle OAM\angle{OAM}. Suppose half-line OY\overrightarrow{OY} meets 𝒞,AN¯\mathcal{C},\overline{AN} at P,QP,Q, respectively. Thus, |PY|=d(Y,𝒞)|P-Y|=\operatorname{d}(Y,\mathcal{C}). We also notice that

(5.19) YZQ=ONA=π4+θ2.\angle{YZQ}=\angle{ONA}=\frac{\pi}{4}+\frac{\theta}{2}.

Therefore, sinYZQ12\sin{\angle{YZQ}}\geq\frac{1}{\sqrt{2}}. By the law of sines, we have

(5.20) |PY||QY|=sinYZQsinYQZ|ZY|12|ZY|.\begin{split}|P-Y|\geq|Q-Y|&=\frac{\sin\angle{YZQ}}{\sin\angle{YQZ}}\,|Z-Y|\\ &\geq\frac{1}{\sqrt{2}}|Z-Y|.\end{split}
Refer to caption
Figure 3.

We are now in a position to prove the following lemma.

Lemma 5.12.

Suppose Ω3\Omega\subset\mathbb{R}^{3} satisfies the positive curvature condition in Definition 1.5. Then, there exists a constant C=C(Ω)C=C(\Omega) such that for any yΩy\in\Omega and v^𝕊2\hat{v}\in\mathbb{S}^{2}, we have

(5.21) 0|q+(y,v^)y|dy+rv^12+ϵ𝑑rC,ϵ[0,12),\int^{|q_{+}(y,\hat{v})-y|}_{0}d_{y+r\hat{v}}^{-\frac{1}{2}+\epsilon}\,dr\leq C,\quad\forall\epsilon\in[0,\frac{1}{2}),

and

(5.22) 0|q+(y,v^)y|dy+rv^1+ϵ𝑑rCϵdy12+ϵ,ϵ(0,12).\int^{|q_{+}(y,\hat{v})-y|}_{0}d_{y+r\hat{v}}^{-1+\epsilon}\,dr\leq\frac{C}{\epsilon}d_{y}^{-\frac{1}{2}+\epsilon},\quad\forall\epsilon\in(0,\frac{1}{2}).
Proof.

The idea of proof is that we first show the lemma holds for balls. For general cases, we employ Proposition 5.6 to compare the convex domain Ω\Omega with balls. For given ϵ[0,12)\epsilon\in[0,\frac{1}{2}), yΩy\in\Omega and v^𝕊2\hat{v}\in\mathbb{S}^{2}. We denote q=q(y,v^)q_{-}=q_{-}(y,\hat{v}) and q+=q+(y,v^)q_{+}=q_{+}(y,\hat{v}) for simplicity. We claim

(5.23) 0|q+q|dq+rv^12+ϵ𝑑rC.\int^{|q_{+}-q_{-}|}_{0}d_{q_{-}+r\hat{v}}^{-\frac{1}{2}+\epsilon}\,dr\leq C.

First, we shall prove that (5.23) holds for the case of balls in 3\mathbb{R}^{3}. Let 3\mathcal{B}\subset\mathbb{R}^{3} be an open ball with radius ρ\rho centered at OO and A,BA,B be two points on \partial\mathcal{B}. For v^=AB|AB|\hat{v}=\frac{\overrightarrow{AB}}{|A-B|}, the following integral is bounded by some constant C=C(ρ)C=C(\rho),

(5.24) 0|AB|d(A+rv^,)12+ϵdrC.\int^{|A-B|}_{0}\operatorname{d}(A+r\hat{v},\partial\mathcal{B})^{-\frac{1}{2}+\epsilon}\,dr\leq C.

Let 𝒞\mathcal{C} be the intersection circle of \partial\mathcal{B} and the plane passing through A,BA,B and OO. Denote the midpoint of AB¯\overline{AB} by MM and write θ=OAB\theta=\angle{OAB}. For 0r|AM|0\leq r\leq|A-M|, it follows by Proposition 5.11 that

(5.25) d(A+rv^,𝒞)12|ZY|=12tan(π4θ2)r=121sinθcosθr,\begin{split}\operatorname{d}(A+r\hat{v},\mathcal{C})&\geq\frac{1}{\sqrt{2}}|Z-Y|\\ &=\frac{1}{\sqrt{2}}\tan\big{(}\frac{\pi}{4}-\frac{\theta}{2}\big{)}\,r\\ &=\frac{1}{\sqrt{2}}\frac{1-\sin\theta}{\cos\theta}\,r,\end{split}

where Y=A+rv^Y=A+r\hat{v} and ZZ is as defined in Proposition 5.11. Therefore, we obtain

(5.26) 0|AM|d(A+rv^,)12+ϵdr=0|AM|d(A+rv^,𝒞)12+ϵdrC0ρcosθ(cosθ1sinθ)12ϵr12+ϵ𝑑rCρ12+ϵcosθ(1sinθ)12ϵCmax{ρ,1},\begin{split}\int^{|A-M|}_{0}\operatorname{d}(A+r\hat{v},\partial\mathcal{B})^{-\frac{1}{2}+\epsilon}\,dr&=\int^{|A-M|}_{0}\operatorname{d}(A+r\hat{v},\mathcal{C})^{-\frac{1}{2}+\epsilon}\,dr\\ &\leq C\int^{\rho\cos\theta}_{0}\left(\frac{\cos\theta}{1-\sin\theta}\right)^{\frac{1}{2}-\epsilon}r^{-\frac{1}{2}+\epsilon}\,dr\\ &\leq C\rho^{\frac{1}{2}+\epsilon}\,\frac{\cos\theta}{(1-\sin\theta)^{\frac{1}{2}-\epsilon}}\\ &\leq C\max\{\rho,1\},\end{split}

where the last inequality follows from the fact cosθ2(1sinθ)\cos\theta\leq\sqrt{2(1-\sin\theta)}. The situation |AM|r|AB||A-M|\leq r\leq|A-B| can be treated similarly. Therefore, (5.24) follows.

For general cases, we make a comparison with sphere cases. According to Proposition 5.6, there are spheres Si(q)S_{i}(q_{-}) and Si(q+)S_{i}(q_{+}) with radii r1=r1(Ω)r_{1}=r_{1}(\Omega) as defined in Definition 5.4. Denote the other intersection of Si(q)S_{i}(q_{-}) and qq+¯\overline{q_{-}q_{+}} by q1q_{1}, the other intersection of Si(q+)S_{i}(q_{+}) and qq+¯\overline{q_{-}q_{+}} by q2q_{2}. For 0r|q1q|0\leq r\leq|q_{1}-q_{-}|, we notice that

(5.27) d(q+rv^,Ω)d(q+rv^,Si(q)).\operatorname{d}(q_{-}+r\hat{v},\partial\Omega)\geq\operatorname{d}(q_{-}+r\hat{v},S_{i}(q_{-})).

For |q2q|r|q+q||q_{2}-q_{-}|\leq r\leq|q_{+}-q_{-}|, similarly we have

(5.28) d(q+rv^,Ω)d(q+rv^,Si(q+)).\operatorname{d}(q_{-}+r\hat{v},\partial\Omega)\geq\operatorname{d}(q_{-}+r\hat{v},S_{i}(q_{+})).

Therefore, in the case of qq1¯q2q+¯=qq+¯\overline{q_{-}q_{1}}\cup\overline{q_{2}q_{+}}=\overline{q_{-}q_{+}}, by (5.24) there exists a constant C=C(r1)C=C(r_{1}) such that

(5.29) 0|q+q|dq+rv^12+ϵ𝑑r0|q1q|dq+rv^12+ϵ𝑑r+|q2q||q+q|dq+rv^12+ϵ𝑑r0|q1q|d(q+rv^,Si(q))12+ϵdr+|q2q||q+q|d(q+rv^,Si(q+))12+ϵdrC.\begin{split}\int^{|q_{+}-q_{-}|}_{0}d_{q_{-}+r\hat{v}}^{-\frac{1}{2}+\epsilon}\,dr&\leq\int^{|q_{1}-q_{-}|}_{0}d_{q_{-}+r\hat{v}}^{-\frac{1}{2}+\epsilon}\,dr+\int^{|q_{+}-q_{-}|}_{|q_{2}-q_{-}|}d_{q_{-}+r\hat{v}}^{-\frac{1}{2}+\epsilon}\,dr\\ &\leq\int^{|q_{1}-q_{-}|}_{0}\operatorname{d}(q_{-}+r\hat{v},S_{i}(q_{-}))^{-\frac{1}{2}+\epsilon}\,dr\\ &\quad\quad+\int^{|q_{+}-q_{-}|}_{|q_{2}-q_{-}|}\operatorname{d}(q_{-}+r\hat{v},S_{i}(q_{+}))^{-\frac{1}{2}+\epsilon}\,dr\\ &\leq C.\end{split}

If qq1¯q2q+¯\overline{q_{-}q_{1}}\cup\overline{q_{2}q_{+}} does not cover qq+¯\overline{q_{-}q_{+}}, we consider the convex hull of Si(q)Si(q+)S_{i}(q_{-})\cup S_{i}(q_{+}), denoted by SS. Clearly, one can see that

(5.30) d(q+rv^,Ω)d(q+rv^,S).\operatorname{d}(q_{-}+r\hat{v},\partial\Omega)\geq\operatorname{d}(q_{-}+r\hat{v},\partial S).

From solid geometry, we can see the following property.

Claim.

For |q1+q2q|r|q2+q+2q|\left|\frac{q_{1}+q_{-}}{2}-q_{-}\right|\leq r\leq\left|\frac{q_{2}+q_{+}}{2}-q_{-}\right|, if N(y,v^)N+(y,v^)N_{-}(y,\hat{v})\leq N_{+}(y,\hat{v}) (resp. N(y,v^)>N+(y,v^)N_{-}(y,\hat{v})>N_{+}(y,\hat{v})) or equivalently θθ+\theta_{-}\geq\theta_{+} (resp. θ<θ+\theta_{-}<\theta_{+}), where θ±=arccosN±(y,v^)\theta_{\pm}=\arccos N_{\pm}(y,\hat{v}), then we have the inequality

(5.31) d(q+rv^,S)d(q1+q2,Si(q))=r1(1sinθ)\displaystyle\operatorname{d}(q_{-}+r\hat{v},\partial S)\geq\operatorname{d}(\frac{q_{1}+q_{-}}{2},S_{i}(q_{-}))=r_{1}(1-\sin\theta_{-})
(5.32) (resp.\displaystyle\bigg{(}\text{resp. } d(q+rv^,S)d(q2+q+2,Si(q+))=r1(1sinθ+)).\displaystyle\operatorname{d}(q_{-}+r\hat{v},\partial S)\geq\operatorname{d}(\frac{q_{2}+q_{+}}{2},S_{i}(q_{+}))=r_{1}(1-\sin\theta_{+})\bigg{)}.
Refer to caption
Figure 4.

See Figure 4 for two-dimensional cases.

Recalling that |qq+|2R1cosθ|q_{-}-q_{+}|\leq 2R_{1}\cos\theta_{-} from Remark 5.10, thus we deduce that, if N(y,v^)N+(y,v^)N_{-}(y,\hat{v})\leq N_{+}(y,\hat{v}),

(5.33) 0|q+q|dq+rv^12+ϵ𝑑r0|q1q|+|q1q||q2q|+|q2q||q+q|dq+rv^12+ϵ𝑑rC(r1)+2R1r112+ϵcosθ(1sinθ)12+ϵ+C(r1)C(r1,R1).\begin{split}\int^{|q_{+}-q_{-}|}_{0}d_{q_{-}+r\hat{v}}^{-\frac{1}{2}+\epsilon}\,dr&\leq\int^{|q_{1}-q_{-}|}_{0}+\int^{|q_{2}-q_{-}|}_{|q_{1}-q_{-}|}+\int^{|q_{+}-q_{-}|}_{|q_{2}-q_{-}|}d_{q_{-}+r\hat{v}}^{-\frac{1}{2}+\epsilon}\,dr\\ &\leq C(r_{1})+2R_{1}r_{1}^{-\frac{1}{2}+\epsilon}\cos\theta_{-}\,(1-\sin\theta_{-})^{-\frac{1}{2}+\epsilon}+C(r_{1})\\ &\leq C(r_{1},R_{1}).\end{split}

The inequality holds for N(y,v^)>N+(y,v^)N_{-}(y,\hat{v})>N_{+}(y,\hat{v}) similarly.

Let us now look at (5.22). We mimic the above proof for (5.23), since the steps proceed in the same way. For given ϵ(0,12)\epsilon\in(0,\frac{1}{2}), yΩy\in\Omega and v^𝕊2\hat{v}\in\mathbb{S}^{2}, this time we claim

(5.34) 0|q+q|dq+rv^1+ϵ𝑑rCϵdy12+ϵ\int^{|q_{+}-q_{-}|}_{0}d_{q_{-}+r\hat{v}}^{-1+\epsilon}\,dr\leq\frac{C}{\epsilon}d_{y}^{-\frac{1}{2}+\epsilon}

for some constant C=C(r1,R1)C=C(r_{1},R_{1}). For the cases of open balls in 3\mathbb{R}^{3}, we adopt the above notations and deduce that

(5.35) 0|AM|d(A+rv^,)1+ϵdr=0|AM|d(A+rv^,𝒞)1+ϵdrC0ρcosθ(cosθ1sinθ)1ϵr1+ϵ𝑑rCρϵcosθϵ(1sinθ)1ϵCmax{ρ,1}ϵ(1sinθ)12+ϵCϵd(M,𝒞)12+ϵ(Cϵd(Y,𝒞)12+ϵfor any YAB¯),\begin{split}\int^{|A-M|}_{0}\operatorname{d}(A+r\hat{v},\mathcal{\partial}\mathcal{B})^{-1+\epsilon}\,dr&=\int^{|A-M|}_{0}\operatorname{d}(A+r\hat{v},\mathcal{C})^{-1+\epsilon}\,dr\\ &\leq C\int^{\rho\cos\theta}_{0}\left(\frac{\cos\theta}{1-\sin\theta}\right)^{1-\epsilon}r^{-1+\epsilon}\,dr\\ &\leq C\,\frac{\rho^{\epsilon}\cos\theta}{\epsilon(1-\sin\theta)^{1-\epsilon}}\\ &\leq C\,\frac{\max\{\rho,1\}}{\epsilon(1-\sin\theta)^{-\frac{1}{2}+\epsilon}}\\ &\leq\frac{C}{\epsilon}\,\operatorname{d}(M,\mathcal{C})^{-\frac{1}{2}+\epsilon}\\ (&\leq\frac{C}{\epsilon}\,\operatorname{d}(Y,\mathcal{C})^{-\frac{1}{2}+\epsilon}\quad\text{for any }Y\in\overline{AB}),\end{split}

where we have used cosθ2(1sinθ)\cos\theta_{-}\leq\sqrt{2(1-\sin\theta_{-})}. For general cases, we denote q3=q+q12q_{3}=\frac{q_{-}+q_{1}}{2} and q4=q++q22q_{4}=\frac{q_{+}+q_{2}}{2} for convenience. By considering the convex hull SS of Si(q)Si(q+)S_{i}(q_{-})\cup S_{i}(q_{+}) again, if N(y,v^)N+(y,v^)N_{-}(y,\hat{v})\leq N_{+}(y,\hat{v}), we deduce

(5.36) 0|q+q|dq+rv^1+ϵ𝑑r0|q1q|+|q1q||q2q|+|q2q||q+q|dq+rv^1+ϵ𝑑rCϵd((q3,Si(q))12+ϵ+2R1r11+ϵcosθ(1sinθ)1+ϵ+Cϵd(q4,Si(q+))12+ϵCϵ(d(q3,Si(q))12+ϵ+d(q4,Si(q+))12+ϵ),\begin{split}\int^{|q_{+}-q_{-}|}_{0}d_{q_{-}+r\hat{v}}^{-1+\epsilon}\,dr&\leq\int^{|q_{1}-q_{-}|}_{0}+\int^{|q_{2}-q_{-}|}_{|q_{1}-q_{-}|}+\int^{|q_{+}-q_{-}|}_{|q_{2}-q_{-}|}d_{q_{-}+r\hat{v}}^{-1+\epsilon}\,dr\\ &\leq\frac{C}{\epsilon}\,\operatorname{d}\bigl{(}(q_{3},S_{i}(q_{-})\bigr{)}^{-\frac{1}{2}+\epsilon}\\ &\quad+2R_{1}r_{1}^{-1+\epsilon}\cos\theta_{-}\,(1-\sin\theta_{-})^{-1+\epsilon}\\ &\quad+\frac{C}{\epsilon}\,\operatorname{d}\bigl{(}q_{4},S_{i}(q_{+})\bigr{)}^{-\frac{1}{2}+\epsilon}\\ &\leq\frac{C}{\epsilon}\,\left(\operatorname{d}\bigl{(}q_{3},S_{i}(q_{-})\bigr{)}^{-\frac{1}{2}+\epsilon}+\operatorname{d}\bigl{(}q_{4},S_{i}(q_{+})\bigr{)}^{-\frac{1}{2}+\epsilon}\right),\end{split}

where we have used cosθ2(1sinθ)\cos\theta_{-}\leq\sqrt{2(1-\sin\theta_{-})} and d(q3,Si(q))=r1(1sinθ)\operatorname{d}\bigl{(}q_{3},S_{i}(q_{-})\bigr{)}=r_{1}(1-\sin\theta_{-}) in the last inequality. To complete the proof for (5.34), we consider sphere So(q)S_{o}(q_{-}) defined in Definition 5.4. Denote the other intersection of half-line qq+\overrightarrow{q_{-}q_{+}} and So(q)S_{o}(q_{-}) by q5q_{5}, the midpoint of the line segment qq5¯\overline{q_{-}q_{5}} by q6q_{6}. Clearly, we have

(5.37) dy=d(y,Ω)d(y,So(q))d(q6,So(q))=2R1(1sinθ).d_{y}=\operatorname{d}(y,\partial\Omega)\leq\operatorname{d}\bigl{(}y,S_{o}(q_{-})\bigr{)}\leq\operatorname{d}\bigl{(}q_{6},S_{o}(q_{-})\bigr{)}=2R_{1}(1-\sin\theta_{-}).

Hence, it follows that

(5.38) d(q3,Si(q))12+ϵ=r112+ϵ(1sinθ)12+ϵ=(r1R1)12+ϵd(q6,So(q))12+ϵCdy12+ϵ.\begin{split}\operatorname{d}\bigl{(}q_{3},S_{i}(q_{-})\bigr{)}^{-\frac{1}{2}+\epsilon}&=r_{1}^{-\frac{1}{2}+\epsilon}(1-\sin\theta_{-})^{-\frac{1}{2}+\epsilon}\\ &=\left(\frac{r_{1}}{R_{1}}\right)^{-\frac{1}{2}+\epsilon}\operatorname{d}\bigl{(}q_{6},S_{o}(q_{-})\bigr{)}^{-\frac{1}{2}+\epsilon}\\ &\leq C\,d_{y}^{-\frac{1}{2}+\epsilon}.\end{split}

Similarly, we have

(5.39) d(q4,Si(q+))12+ϵCdy12+ϵ.\operatorname{d}\bigl{(}q_{4},S_{i}(q_{+})\bigr{)}^{-\frac{1}{2}+\epsilon}\leq C\,d_{y}^{-\frac{1}{2}+\epsilon}.

We conclude that

(5.40) 0|q+q|dq+rv^1+ϵ𝑑rCϵdy12+ϵ.\int^{|q_{+}-q_{-}|}_{0}d_{q_{-}+r\hat{v}}^{-1+\epsilon}\,dr\leq\frac{C}{\epsilon}\,d_{y}^{-\frac{1}{2}+\epsilon}.

The inequality holds for N(y,v^)>N+(y,v^)N_{-}(y,\hat{v})>N_{+}(y,\hat{v}) similarly. This completes the proof. ∎

Remark 5.13.

One can see that the behavior of integral 0|q+q|dq+rv^s𝑑r\int^{|q_{+}-q_{-}|}_{0}d^{-s}_{q_{-}+r\hat{v}}\,dr depends heavily on boundedness of Θ(θ;s)=cosθ(1sinθ)s\Theta(\theta;s)=\frac{\cos\theta}{(1-\sin\theta)^{s}} for θ(0,π2)\theta\in(0,\frac{\pi}{2}). When s12s\leq\frac{1}{2}, there is a constant CC independent of θ\theta such that Θ(θ;s)C\Theta(\theta;s)\leq C. On the other hand, there is no such constant whenever s>12s>\frac{1}{2}. We also notice that the integral blows up as s1s\to 1-.

Corollary 5.14.

Suppose Ω3\Omega\subset\mathbb{R}^{3} satisfies the positive curvature condition in Definition 1.5. Then, there exists a constant C=C(Ω)C=C(\Omega) such that, for any small ϵ>0\epsilon>0, we have

(5.41) Ω1dx1ϵ𝑑xCϵ.\int_{\Omega}\frac{1}{d_{x}^{1-\epsilon}}\,dx\leq\frac{C}{\epsilon}.
Proof.

Let yy be an interior point of Ω\Omega and {u,v,w}\{u,v,w\} be an orthonormal basis for 3\mathbb{R}^{3}. We note that

(5.42) Ω1dx1ϵ𝑑x=|q(y,u)y||q+(y,u)y||q(y+ru,v)(y+ru)||q+(y+ru,v)(y+ru)||q(y+ru+sv,w)(y+ru+sv)||q+(y+ru+sv,w)(y+ru+sv)|dy+ru+sv+tw1+ϵ𝑑t𝑑s𝑑r.\begin{split}&\int_{\Omega}\frac{1}{d_{x}^{1-\epsilon}}\,dx\\ =&\int^{|q_{+}(y,u)-y|}_{-|q_{-}(y,u)-y|}\int^{|q_{+}(y+ru,v)-(y+ru)|}_{-|q_{-}(y+ru,v)-(y+ru)|}\int^{|q_{+}(y+ru+sv,w)-(y+ru+sv)|}_{-|q_{-}(y+ru+sv,w)-(y+ru+sv)|}d_{y+ru+sv+tw}^{-1+\epsilon}\,dtdsdr.\end{split}

According to Lemma 5.12, it follows that

(5.43) |q(y,u)y||q+(y,u)y||q(y+ru,v)(y+ru)||q+(y+ru,v)(y+ru)||q(y+ru+sv,w)(y+ru+sv)||q+(y+ru+sv,w)(y+ru+sv)|dy+ru+sv+tw1+ϵ𝑑t𝑑s𝑑r|q(y,u)y||q+(y,u)y||q(y+ru,v)(y+ru)||q+(y+ru,v)(y+ru)|Cϵdy+ru+sv12+ϵ𝑑s𝑑rCϵ|q(y,u)y||q+(y,u)y|C𝑑rCϵ.\begin{split}&\int^{|q_{+}(y,u)-y|}_{-|q_{-}(y,u)-y|}\int^{|q_{+}(y+ru,v)-(y+ru)|}_{-|q_{-}(y+ru,v)-(y+ru)|}\int^{|q_{+}(y+ru+sv,w)-(y+ru+sv)|}_{-|q_{-}(y+ru+sv,w)-(y+ru+sv)|}d_{y+ru+sv+tw}^{-1+\epsilon}\,dtdsdr\\ &\,\leq\int^{|q_{+}(y,u)-y|}_{-|q_{-}(y,u)-y|}\int^{|q_{+}(y+ru,v)-(y+ru)|}_{-|q_{-}(y+ru,v)-(y+ru)|}\frac{C}{\epsilon}\,d_{y+ru+sv}^{-\frac{1}{2}+\epsilon}\,dsdr\\ &\,\leq\frac{C}{\epsilon}\int^{|q_{+}(y,u)-y|}_{-|q_{-}(y,u)-y|}C\,dr\\ &\,\leq\frac{C}{\epsilon}.\end{split}

6. Regularity of transport equation in a convex domain

Hereafter, Ω\Omega denotes a bounded convex domain satisfying positive curvature condition in Definition 1.5 and gg denotes an incoming data satisfying Assumption 1.7.

Lemma 6.1.

Let JgJg as defined by (1.19). Then JgLv2(3;Hx1ϵ(Ω))Jg\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) for any small ϵ>0\epsilon>0.

Proof.

To prove the lemma, for given ϵ>0\epsilon>0 and an incoming data gg satisfying Assumption 1.7, it suffices to show the boundedness of integral

(6.1) 3ΩΩ|Jg(x,v)Jg(y,v)|2|xy|52ϵ𝑑x𝑑y𝑑v.\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega}\frac{|Jg(x,v)-Jg(y,v)|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv.

We partition the domain of integration D:=3×Ω×ΩD:=\mathbb{R}^{3}\times\Omega\times\Omega into D1,D2D_{1},D_{2} as below.

(6.2) D1\displaystyle D_{1} :={(x,y,v)D:|xq(x,v)||yq(y,v)|},\displaystyle:=\{(x,y,v)\in D:\,|x-q_{-}(x,v)|\leq|y-q_{-}(y,v)|\},
(6.3) D2\displaystyle D_{2} :={(x,y,v)D:|xq(x,v)|>|yq(y,v)|}.\displaystyle:=\{(x,y,v)\in D:\,|x-q_{-}(x,v)|>|y-q_{-}(y,v)|\}.

In view of symmetry of D1,D2D_{1},D_{2}, we only need to calculate the integral (6.1) over D1D_{1}. Notice that

(6.4) |Jg(x,v)Jg(y,v)|22e2ν(v)τ(x,v)|g(q(x,v),v)g(q(y,v),v)|2+2|g(q(y,v),v)|2|eν(v)τ(x,v)eν(v)τ(y,v)|2.\begin{split}|Jg(x,v)-Jg(y,v)|^{2}\leq&2e^{-2\nu(v)\tau_{-}(x,v)}|g(q_{-}(x,v),v)-g(q_{-}(y,v),v)|^{2}\\ &+2|g(q_{-}(y,v),v)|^{2}|e^{-\nu(v)\tau_{-}(x,v)}-e^{-\nu(v)\tau_{-}(y,v)}|^{2}.\end{split}

Consequently, we have

(6.5) D1|Jg(x,v)Jg(y,v)|2|xy|52ϵ𝑑x𝑑y𝑑vI1+I2,\int_{D_{1}}\frac{|Jg(x,v)-Jg(y,v)|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv\leq I_{1}+I_{2},

where

(6.6) I1\displaystyle I_{1} :=D12e2ν(v)τ(x,v)|g(q(x,v),v)g(q(y,v),v)|2|xy|52ϵ𝑑x𝑑y𝑑v,\displaystyle:=\int_{D_{1}}\frac{2e^{-2\nu(v)\tau_{-}(x,v)}|g(q_{-}(x,v),v)-g(q_{-}(y,v),v)|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv,
(6.7) I2\displaystyle I_{2} :=D12|g(q(y,v),v)|2|eν(v)τ(x,v)eν(v)τ(y,v)|2|xy|52ϵ𝑑x𝑑y𝑑v.\displaystyle:=\int_{D_{1}}\frac{2|g(q_{-}(y,v),v)|^{2}|e^{-\nu(v)\tau_{-}(x,v)}-e^{-\nu(v)\tau_{-}(y,v)}|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv.

To estimate I1I_{1}, by Hölder continuity (1.16) and the fact from condition (1.15) that

(6.8) |g(q(x,v),v)g(q(y,v),v)|Cea|v|2,|g(q_{-}(x,v),v)-g(q_{-}(y,v),v)|\leq Ce^{-a|v|^{2}},

we obtain

(6.9) e2ν(v)τ(x,v)|g(q(x,v),v)g(q(y,v),v)|2=e2ν(v)τ(x,v)|g(q(x,v),v)g(q(y,v),v)|2ϵ×|g(q(x,v),v)g(q(y,v),v)|ϵCe2ν(v)τ(x,v)|q(x,v)q(y,v)|2ϵeϵa|v|2.\begin{split}&e^{-2\nu(v)\tau_{-}(x,v)}|g(q_{-}(x,v),v)-g(q_{-}(y,v),v)|^{2}\\ &\qquad=e^{-2\nu(v)\tau_{-}(x,v)}|g(q_{-}(x,v),v)-g(q_{-}(y,v),v)|^{2-\epsilon}\\ &\qquad\quad\times|g(q_{-}(x,v),v)-g(q_{-}(y,v),v)|^{\epsilon}\\ &\qquad\leq C\,e^{-2\nu(v)\tau_{-}(x,v)}|q_{-}(x,v)-q_{-}(y,v)|^{2-\epsilon}e^{-\epsilon a|v|^{2}}.\end{split}

According to Proposition 5.3, we have

(6.10) |q(x,v)q(y,v)|1N(x,v)|xy|,|q_{-}(x,v)-q_{-}(y,v)|\leq\frac{1}{N_{-}(x,v)}|x-y|,

whenever |xq(x,v)||yq(y,v)||x-q_{-}(x,v)|\leq|y-q_{-}(y,v)|. Proposition 5.2 implies that

(6.11) e2ν(v)τ(x,v)1ν(v)1ϵτ(x,v)1ϵ1ν(v)1ϵτ(x,v)1ϵ(N(x,v)|xq(x,v)|dx)1ϵN(x,v)1ϵ|v|1ϵν(v)1ϵdx1ϵ.\begin{split}e^{-2\nu(v)\tau_{-}(x,v)}\leq&\frac{1}{\nu(v)^{1-\epsilon}{\tau_{-}(x,v)^{1-\epsilon}}}\\ \leq&\frac{1}{\nu(v)^{1-\epsilon}{\tau_{-}(x,v)^{1-\epsilon}}}\cdot\left(\frac{N_{-}(x,v)|x-q_{-}(x,v)|}{d_{x}}\right)^{1-\epsilon}\\ \leq&\frac{N_{-}(x,v)^{1-\epsilon}|v|^{1-\epsilon}}{\nu(v)^{1-\epsilon}d_{x}^{1-\epsilon}}.\end{split}

Combining (6.8),(LABEL:eq:jgI1estimate2),(6.10), and (6.11), we have

(6.12) I1CD1|v|1ϵeϵa|v|2ν(v)1ϵdx1ϵN(x,v)|xy|3ϵ𝑑x𝑑y𝑑vCΩ3Ω|v|1ϵeϵa|v|2ν(v)1ϵdx1ϵN(x,v)|xy|3ϵ𝑑y𝑑v𝑑xCΩ3|v|1ϵeϵa|v|2dx1ϵN(x,v)𝑑v𝑑x.\begin{split}I_{1}&\leq C\,\int\limits_{D_{1}}\frac{|v|^{1-\epsilon}e^{-\epsilon a|v|^{2}}}{\nu(v)^{1-\epsilon}d_{x}^{1-\epsilon}N_{-}(x,v)|x-y|^{3-\epsilon}}\,dxdydv\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\frac{|v|^{1-\epsilon}e^{-\epsilon a|v|^{2}}}{\nu(v)^{1-\epsilon}d_{x}^{1-\epsilon}N_{-}(x,v)|x-y|^{3-\epsilon}}\,dydvdx\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|v|^{1-\epsilon}e^{-\epsilon a|v|^{2}}}{d_{x}^{1-\epsilon}{N_{-}(x,v)}}\,dvdx.\end{split}

For fixed xΩx\in\Omega, we introduce a change of variable v=(xz)lv=(x-z)l with zΩz\in\partial\Omega and 0l<0\leq l<\infty. For any local chart of Ω\partial\Omega, say z=ϕ(α,β)z=\phi(\alpha,\beta), we note that

v(α,β,l)=(xϕ(α,β))lv(\alpha,\beta,l)=\bigl{(}x-\phi(\alpha,\beta)\bigr{)}l

and the Jacobian for vv is given by

(6.13) |det𝐉v(α,β,l)|=|lv(αv×βv)|=l2|(xϕ(α,β))(ϕα×ϕβ)|=l2|(xϕ(α,β))n(ϕ(α,β))||ϕα×ϕβ|.\begin{split}\left|\det\mathbf{J}_{v}(\alpha,\beta,l)\right|&=\left|\partial_{l}v\cdot(\partial_{\alpha}v\times\partial_{\beta}v)\right|\\ &=l^{2}\left|(x-\phi(\alpha,\beta))\cdot(\phi_{\alpha}\times\phi_{\beta})\right|\\ &=l^{2}\left|(x-\phi(\alpha,\beta))\cdot n(\phi(\alpha,\beta))\right||\phi_{\alpha}\times\phi_{\beta}|.\end{split}

Covering Ω\partial\Omega by finitely many such coordinate charts, we obtain the formula for change of variables

(6.14) 3h(v)𝑑v=Ω0h((xz)l)l2|(xz)n(z)|𝑑l𝑑Σ(z),\int_{\mathbb{R}^{3}}h(v)\,dv=\int_{\partial\Omega}\int^{\infty}_{0}h((x-z)l)\,l^{2}\left|(x-z)\cdot n(z)\right|\,dld\Sigma(z),

for any hL1(3)h\in L^{1}(\mathbb{R}^{3}). Therefore, by the change of variables, we obtain

(6.15) Ω3|v|1ϵeϵa|v|2dx1ϵN(x,v)𝑑v𝑑x=ΩΩ0eϵa|xz|2l2|xz|1ϵl1ϵdx1ϵ|(xz)n(z)||xz|l2|(xz)n(z)|𝑑l𝑑Σ(z)𝑑x=ΩΩ0eϵa|xz|2l2|xz|2ϵl3ϵdx1ϵ𝑑l𝑑Σ(z)𝑑x.\begin{split}&\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|v|^{1-\epsilon}e^{-\epsilon a|v|^{2}}}{d_{x}^{1-\epsilon}{N_{-}(x,v)}}\,dvdx\\ &\qquad=\int\limits_{\Omega}\int\limits_{\partial\Omega}\int\limits^{\infty}_{0}\frac{e^{-\epsilon a|x-z|^{2}l^{2}}|x-z|^{1-\epsilon}l^{1-\epsilon}}{d_{x}^{1-\epsilon}\cdot\frac{|(x-z)\cdot n(z)|}{|x-z|}}\,l^{2}|(x-z)\cdot n(z)|dld\Sigma(z)dx\\ &\qquad=\int\limits_{\Omega}\int\limits_{\partial\Omega}\int\limits^{\infty}_{0}\frac{e^{-\epsilon a|x-z|^{2}l^{2}}|x-z|^{2-\epsilon}l^{3-\epsilon}}{d_{x}^{1-\epsilon}}\,dld\Sigma(z)dx.\end{split}

Letting s=|xz|ls=|x-z|\,l yields

(6.16) I1CΩΩ0eϵa|xz|2l2|xz|2ϵl3ϵdx1ϵ𝑑l𝑑Σ(z)𝑑x=CΩΩ0eϵas2s3ϵdx1ϵ|xz|2𝑑s𝑑Σ(z)𝑑xCΩΩ1dx1ϵ|xz|2𝑑Σ(z)𝑑xCΩ|log(dx)|+1dx1ϵ𝑑xC.\begin{split}I_{1}&\leq C\,\int\limits_{\Omega}\int\limits_{\partial\Omega}\int\limits^{\infty}_{0}\frac{e^{-\epsilon a|x-z|^{2}l^{2}}|x-z|^{2-\epsilon}l^{3-\epsilon}}{d_{x}^{1-\epsilon}}\,dld\Sigma(z)dx\\ &=C\,\int\limits_{\Omega}\int\limits_{\partial\Omega}\int\limits^{\infty}_{0}\frac{e^{-\epsilon as^{2}}s^{3-\epsilon}}{d_{x}^{1-\epsilon}|x-z|^{2}}\,dsd\Sigma(z)dx\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\partial\Omega}\frac{1}{d_{x}^{1-\epsilon}|x-z|^{2}}\,d\Sigma(z)dx\\ &\leq C\,\int_{\Omega}\frac{|\log(d_{x})|+1}{d_{x}^{1-\epsilon}}\,dx\\ &\leq C.\end{split}

The third inequality above follows from Lemma 5.8 and the last inequality follows from Corollary 5.14 and the fact

(6.17) |log(dx)|dx1ϵ1dx1ϵ2\frac{|\log(d_{x})|}{d_{x}^{1-\epsilon}}\leq\frac{1}{d_{x}^{1-\frac{\epsilon}{2}}}

for small dx>0d_{x}>0.

Concerning I2I_{2}, the condition (1.15) implies that

(6.18) |g(q(y,v),v)|2|eν(v)τ(x,v)eν(v)τ(y,v)|2Ce2a|v|2|eν(v)τ(x,v)eν(v)τ(y,v)|2ϵ.\begin{split}&|g(q_{-}(y,v),v)|^{2}|e^{-\nu(v)\tau_{-}(x,v)}-e^{-\nu(v)\tau_{-}(y,v)}|^{2}\\ &\qquad\leq Ce^{-2a|v|^{2}}|e^{-\nu(v)\tau_{-}(x,v)}-e^{-\nu(v)\tau_{-}(y,v)}|^{2-\epsilon}.\end{split}

By the mean value theorem and Proposition 5.3, we obtain

(6.19) |eν(v)τ(x,v)eν(v)τ(y,v)|ν(v)eν(v)τ(x,v)|τ(x,v)τ(y,v)|2ν(v)eν(v)τ(x,v)|xy|N(x,v)|v|.\begin{split}|e^{-\nu(v)\tau_{-}(x,v)}-e^{-\nu(v)\tau_{-}(y,v)}|&\leq\nu(v)e^{-\nu(v)\tau_{-}(x,v)}|\tau_{-}(x,v)-\tau_{-}(y,v)|\\ &\leq\frac{2\nu(v)e^{-\nu(v)\tau_{-}(x,v)}|x-y|}{N_{-}(x,v)|v|}.\end{split}

Proposition 5.2 implies that

(6.20) e(2ϵ)ν(v)τ(x,v)1ν(v)1ϵτ(x,v)1ϵN(x,v)1ϵ|v|1ϵν(v)1ϵdx1ϵ.\begin{split}e^{-(2-\epsilon)\nu(v)\tau_{-}(x,v)}\leq&\frac{1}{\nu(v)^{1-\epsilon}\tau_{-}(x,v)^{1-\epsilon}}\\ \leq&\frac{N_{-}(x,v)^{1-\epsilon}|v|^{1-\epsilon}}{\nu(v)^{1-\epsilon}d_{x}^{1-\epsilon}}.\end{split}

Taking (LABEL:eq:jgI2estimate1),(6.19), and (6.20) into consideration, we obtain

(6.21) I2CD1e2a|v|2ν(v)dxdydv|xy|3ϵN(x,v)|v|dx1ϵ.I_{2}\leq C\,\int_{D_{1}}\frac{e^{-2a|v|^{2}}\nu(v)\,dxdydv}{|x-y|^{3-\epsilon}N_{-}(x,v)|v|d_{x}^{1-\epsilon}}.

Comparing (6.21) with (6.12), one can repeat the steps in (6.12), (LABEL:eq:jgIntegral1ChangeOfVariable1), and (6.16) to obtain the following boundedness,

I2C.I_{2}\leq C.

This completes the proof. ∎

Since K:L2(v2)L2(v2)K:L^{2}(\mathbb{R}^{2}_{v})\to L^{2}(\mathbb{R}^{2}_{v}) is a bounded operator regarding velocity variable, we can see that KK would preserve regularity in space variable.

Proposition 6.2.

The operator K:Lv2(3;Hx1ϵ(Ω))Lv2(3;Hx1ϵ(Ω))K:L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega))\to L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) is bounded for any small ϵ>0\epsilon>0.

Next, we deal with regularity preservation of SΩS_{\Omega}. Recall

(6.22) SΩh(x,v)=0τ(x,v)eν(v)sh(xsv,v)𝑑s.S_{\Omega}h(x,v)=\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}h(x-sv,v)\,ds.
Lemma 6.3.

Suppose hLv2(3;Hx1ϵ(Ω))h\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) and there exist positive constants aa and CC such that |h(x,v)|Cea|v|2|h(x,v)|\leq Ce^{-a|v|^{2}}. Then SΩhLv2(3;Hx1ϵ(Ω))S_{\Omega}h\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) for any small ϵ>0\epsilon>0.

Proof.

To prove the lemma, for given ϵ>0\epsilon>0 and hLv2(3;Hx1ϵ(Ω))h\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) with |h(x,v)|Cea|v|2|h(x,v)|\leq Ce^{-a|v|^{2}}, it suffices to show the boundedness of integral

(6.23) D1|SΩh(x,v)SΩh(y,v)|2|xy|52ϵ𝑑x𝑑y𝑑v,\int\limits_{D_{1}}\frac{|S_{\Omega}h(x,v)-S_{\Omega}h(y,v)|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv,

where D1D_{1} as defined by (6.2). In domain D1D_{1}, we have

(6.24) |SΩh(x,v)SΩh(y,v)|22|0τ(x,v)eν(v)s(h(xsv,v)h(ysv,v))𝑑s|2+2|τ(x,v)τ(y,v)eν(v)sh(ysv,v)𝑑s|2.\begin{split}|S_{\Omega}h(x,v)-S_{\Omega}h(y,v)|^{2}\leq&2\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(h(x-sv,v)-h(y-sv,v))\,ds\right|^{2}\\ &+2\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}h(y-sv,v)\,ds\right|^{2}.\end{split}

Therefore, we have

(6.25) D1|SΩh(x,v)SΩh(y,v)|2|xy|52ϵ𝑑x𝑑y𝑑vI1+I2,\int\limits_{D_{1}}\frac{|S_{\Omega}h(x,v)-S_{\Omega}h(y,v)|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv\leq I_{1}+I_{2},

where

(6.26) I1\displaystyle I_{1} :=D12|0τ(x,v)eν(v)s(h(xsv,v)h(ysv,v))𝑑s|2|xy|52ϵ𝑑x𝑑y𝑑v,\displaystyle:=\int_{D_{1}}\frac{2\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(h(x-sv,v)-h(y-sv,v))\,ds\right|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv,
(6.27) I2\displaystyle I_{2} :=D12|τ(x,v)τ(y,v)eν(v)sh(ysv,v)𝑑s|2|xy|52ϵ𝑑x𝑑y𝑑v.\displaystyle:=\int_{D_{1}}\frac{2\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}h(y-sv,v)\,ds\right|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv.

Concerning I1I_{1}, by Cauchy-Schwarz inequality, we obtain

(6.28) |0τ(x,v)eν(v)s(h(xsv,v)h(ysv,v))𝑑s|2(0eν(v)s𝑑s)(0τ(x,v)eν(v)s(h(xsv,v)h(ysv,v))2𝑑s)1ν00τ(x,v)eν(v)s(h(xsv,v)h(ysv,v))2𝑑s1ν00eν(v)s(h¯(xsv,v)h¯(ysv,v))2𝑑s,\begin{split}&\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(h(x-sv,v)-h(y-sv,v))\,ds\right|^{2}\\ &\quad\leq\left(\int^{\infty}_{0}e^{-\nu(v)s}\,ds\right)\left(\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}\bigl{(}h(x-sv,v)-h(y-sv,v)\bigr{)}^{2}\,ds\right)\\ &\quad\leq\frac{1}{\nu_{0}}\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}\bigl{(}h(x-sv,v)-h(y-sv,v)\bigr{)}^{2}\,ds\\ &\quad\leq\frac{1}{\nu_{0}}\int^{\infty}_{0}e^{-\nu(v)s}\bigl{(}\bar{h}(x-sv,v)-\bar{h}(y-sv,v)\bigr{)}^{2}\,ds,\end{split}

where h¯Lv2(3;Hx1ϵ(3))\bar{h}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\mathbb{R}^{3})) is an extension of hLv2(3;Hx1ϵ(Ω))h\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) as defined in [11, Theorem 5.4].

Remark 6.4.

By Theorem 5.4 from [11], we know Hs(Ω)H^{s}(\Omega) can be continuously embedded in Hs(3)H^{s}(\mathbb{R}^{3}). That is, there exists a constant C=C(s,Ω)C=C(s,\Omega) such that, for given uHs(Ω)u\in H^{s}(\Omega), there is an extension u¯Hs(3)\overline{u}\in H^{s}(\mathbb{R}^{3}) of uu, i.e., u¯|Ω=u\left.\overline{u}\right|_{\Omega}=u, satisfying

u¯Hs(3)CuHs(Ω).\|\overline{u}\|_{H^{s}(\mathbb{R}^{3})}\leq C\,\|u\|_{H^{s}(\Omega)}.

For a.e. fixed v3v\in\mathbb{R}^{3}, we then define h¯(x,v)=h(,v)¯(x)Lv2(3;Hx1ϵ(3))\bar{h}(x,v)=\overline{h(\cdot,v)}(x)\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\mathbb{R}^{3})).

Consequently, we deduce that

(6.29) I1C3330eν(v)s(h¯(xsv,v)h¯(ysv,v))2|xy|52ϵ𝑑s𝑑x𝑑y𝑑vC3330eν(v)s(h¯(x,v)h¯(y,v))2|xy|52ϵ𝑑s𝑑x𝑑y𝑑vCh¯Lv2(3;Hx1ϵ(3))2ChLv2(3;Hx1ϵ(Ω))2,\begin{split}I_{1}&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\int^{\infty}_{0}\frac{e^{-\nu(v)s}\bigl{(}\bar{h}(x-sv,v)-\bar{h}(y-sv,v)\bigr{)}^{2}}{|x-y|^{5-2\epsilon}}\,dsdxdydv\\ &\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\int^{\infty}_{0}\frac{e^{-\nu(v)s^{\prime}}\bigl{(}\bar{h}(x^{\prime},v^{\prime})-\bar{h}(y^{\prime},v^{\prime})\bigr{)}^{2}}{|x^{\prime}-y^{\prime}|^{5-2\epsilon}}\,ds^{\prime}dx^{\prime}dy^{\prime}dv^{\prime}\\ &\leq C\,\|\bar{h}\|^{2}_{L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\mathbb{R}^{3}))}\\ &\leq C\,\|h\|^{2}_{L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega))},\end{split}

where we have used the change of variables

(6.30) {v=v,y=ysv,x=xsv,s=s.\left\{\begin{array}[]{l}v^{\prime}=v,\\ y^{\prime}=y-sv,\\ x^{\prime}=x-sv,\\ s^{\prime}=s.\end{array}\right.

Regarding I2I_{2}, we notice that

(6.31) |τ(x,v)τ(y,v)eν(v)sh(ysv,v)𝑑s|2C|τ(x,v)τ(y,v)eν(v)sea|v|2𝑑s|2=Ce2a|v|21ν(v)2|eν(v)τ(x,v)eν(v)τ(y,v)|2Ce2a|v|21ν(v)2|eν(v)τ(x,v)eν(v)τ(y,v)|2ϵCe2a|v|21ν(v)ϵe(2ϵ)ν(v)τ(x,v)|τ(x,v)τ(y,v)|2ϵ.\begin{split}&\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}h(y-sv,v)\,ds\right|^{2}\\ &\qquad\leq C\,\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}e^{-a|v|^{2}}\,ds\right|^{2}\\ &\qquad=C\,e^{-2a|v|^{2}}\frac{1}{\nu(v)^{2}}|e^{-\nu(v)\tau_{-}(x,v)}-e^{-\nu(v)\tau_{-}(y,v)}|^{2}\\ &\qquad\leq C\,e^{-2a|v|^{2}}\frac{1}{\nu(v)^{2}}|e^{-\nu(v)\tau_{-}(x,v)}-e^{-\nu(v)\tau_{-}(y,v)}|^{2-\epsilon}\\ &\qquad\leq C\,e^{-2a|v|^{2}}\frac{1}{\nu(v)^{\epsilon}}\,e^{-(2-\epsilon)\nu(v)\tau_{-}(x,v)}|\tau_{-}(x,v)-\tau_{-}(y,v)|^{2-\epsilon}.\end{split}

Proposition 5.2 implies that

(6.32) e(2ϵ)ν(v)τ(x,v)C1ν(v)1ϵτ(x,v)1ϵCN(x,v)1ϵ|v|1ϵν(v)1ϵdx1ϵ.\begin{split}e^{-(2-\epsilon)\nu(v)\tau_{-}(x,v)}\leq&C\,\frac{1}{\nu(v)^{1-\epsilon}\tau_{-}(x,v)^{1-\epsilon}}\\ \leq&C\,\frac{N_{-}(x,v)^{1-\epsilon}|v|^{1-\epsilon}}{\nu(v)^{1-\epsilon}d_{x}^{1-\epsilon}}.\end{split}

By Proposition 5.3, we have

(6.33) |τ(x,v)τ(y,v)|2ϵC|xy|2ϵN(x,v)2ϵ|v|2ϵ.|\tau_{-}(x,v)-\tau_{-}(y,v)|^{2-\epsilon}\leq C\,\frac{|x-y|^{2-\epsilon}}{N_{-}(x,v)^{2-\epsilon}|v|^{2-\epsilon}}.

Combining (6.31)-(6.33), we deduce

(6.34) I2CD1e2a|v|2N(x,v)|xy|3ϵ|v|dx1ϵ𝑑x𝑑y𝑑v.I_{2}\leq C\,\int\limits_{D_{1}}\frac{e^{-2a|v|^{2}}}{N_{-}(x,v)|x-y|^{3-\epsilon}|v|d_{x}^{1-\epsilon}}\,dxdydv.

To show the boundedness of above integral on the right, comparing (6.34) with (6.12), one can repeat the steps in (6.12), (LABEL:eq:jgIntegral1ChangeOfVariable1), and (6.16) to conclude SΩhLv2(3;Hx1ϵ(Ω))S_{\Omega}h\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)). ∎

Corollary 6.5.

Let gig_{i} be as defined by (1.22). Then,
giLv2(3;Hx1ϵ(Ω))g_{i}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) for each i0i\geq 0.

Proof.

We prove giLv2(3;Hx1ϵ(Ω))g_{i}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) by induction on i0i\geq 0.
Step 1 .
For i=0i=0, since g0=Jgg_{0}=Jg, g0Lv2(3;Hx1ϵ(Ω))g_{0}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) follows by Lemma 6.1. For i=1i=1, we notice that

(6.35) |Jg(x,v)||g(q(x,v),v)|Cea|v|2.|Jg(x,v)|\leq|g(q_{-}(x,v),v)|\leq Ce^{-a|v|^{2}}.

In view of Lemma 3.5, we have

(6.36) |KJg(x,v)|Cea|v|2,|KJg(x,v)|\leq Ce^{-a|v|^{2}},

where we may assume a<14a<\frac{1}{4}. With this in mind, Lemma 6.3 guarantees that g1Lv2(3;Hx1ϵ(Ω))g_{1}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)). Moreover, by noticing that

(6.37) SΩ(ea|v|2)(x,v)=0τ(x,v)eν(v)sea|v|2𝑑s1ν0ea|v|2,S_{\Omega}(e^{-a|v|^{2}})(x,v)=\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}e^{-a|v|^{2}}\,ds\leq\frac{1}{\nu_{0}}e^{-a|v|^{2}},

we see that

(6.38) |g1(x,v)|Cea|v|2.|g_{1}(x,v)|\leq Ce^{-a|v|^{2}}.

Step 2.
Suppose for some i2i\geq 2, we have

(6.39) gi1=(SΩK)i1JgLv2(3;Hx1ϵ(Ω))g_{i-1}=(S_{\Omega}K)^{i-1}Jg\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega))

and

(6.40) |gi1(x,v)|Cea|v|2.|g_{i-1}(x,v)|\leq Ce^{-a|v|^{2}}.

Then Lemma 3.5 implies

(6.41) |Kgi1(x,v)|Cea|v|2.|Kg_{i-1}(x,v)|\leq Ce^{-a|v|^{2}}.

Applying Lemma 6.3 again yields

(6.42) giLv2(3;Hx1ϵ(Ω)).g_{i}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)).

Finally, we also notice that (6.37) implies

(6.43) |gi(x,v)|Cea|v|2.|g_{i}(x,v)|\leq Ce^{-a|v|^{2}}.

We conclude that giLv2(3;Hx1ϵ(Ω))g_{i}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) for each i0i\geq 0 by induction. ∎

7. Regularity via velocity averaging

This section is devoted to the regularity due to velocity averaging. In contrast to Section 4, we address the velocity averaging effect in bounded convex domains instead of the whole space. As mentioned in the introduction, the method of Fourier transform does not apply to bounded domains. To remedy this crux, we adopt Slobodeckij semi-norm as an alternative concept of Sobolev function class, see Definition 1.4. Hence, the difficulty shifts to estimates of singular integrals. To carry out this strategy, we introduce changes of coordinates, see Lemma 7.1 and Lemma 7.2, to obtain the boundedness of the aforementioned singular integrals.

We begin with the proof of Corollary 1.10. Recall from (1.24) that the notation h~\widetilde{h} denotes the zero extension of hh from Ω×3\Omega\times\mathbb{R}^{3} to 3×3\mathbb{R}^{3}\times\mathbb{R}^{3} and Z:Ω×33×3Z:\Omega\times\mathbb{R}^{3}\to\mathbb{R}^{3}\times\mathbb{R}^{3} is the zero extension operator from Ω×3\Omega\times\mathbb{R}^{3} to 3×3\mathbb{R}^{3}\times\mathbb{R}^{3}.

Proof of Corollary 1.10.

For given fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}), according to Lemma 1.2, we have KSKf~Lv2(3;H~x1/2(3))KSK\widetilde{f}\in L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{\nicefrac{{1}}{{2}}}_{x}(\mathbb{R}^{3})). Noticing (1.12) and

(7.1) .(KSKf~)|Ω×3=KSΩKf,\biggl{.}\bigl{(}KSK\widetilde{f}\bigr{)}\biggr{|}_{\Omega\times\mathbb{R}^{3}}=KS_{\Omega}Kf,

we see that

(7.2) KSΩKfLv2(3;Hx1/2(Ω))(KSKf~)Lv2(3;Hx1/2(3))C(KSKf~)Lv2(3;H~x1/2(3))Cf~L2(3×3)=CfL2(Ω×3).\begin{split}\|KS_{\Omega}Kf\|_{L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\Omega))}&\leq\left\|\bigl{(}KSK\widetilde{f}\bigr{)}\right\|_{L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\mathbb{R}^{3}))}\\ &\leq C\,\left\|\bigl{(}KSK\widetilde{f}\bigr{)}\right\|_{L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{\nicefrac{{1}}{{2}}}_{x}(\mathbb{R}^{3}))}\\ &\leq C\,\|\widetilde{f}\|_{L^{2}(\mathbb{R}^{3}\times\mathbb{R}^{3})}\\ &=C\,\|f\|_{L^{2}(\Omega\times\mathbb{R}^{3})}.\end{split}

Lemma 7.1.

We have the following formula for change of variables for any non-negative measurable function h=h(v,y,r)h=h(v,y,r).

(7.3) 3Ω0|yq(y,v)|h(v,y,r)𝑑r𝑑y𝑑v=3Ω0|yq+(y,v)|h(v,y+rv^,r)𝑑r𝑑y𝑑v.\begin{split}&\int_{\mathbb{R}^{3}}\int_{\Omega}\int^{|y-q_{-}(y,v)|}_{0}h(v,y,r)\,drdydv\\ &\qquad=\int_{\mathbb{R}^{3}}\int_{\Omega}\int^{|y^{\prime}-q_{+}(y^{\prime},v^{\prime})|}_{0}h(v^{\prime},y^{\prime}+r^{\prime}\hat{v^{\prime}},r^{\prime})\,dr^{\prime}dy^{\prime}dv^{\prime}.\end{split}
Proof.

Let AA and BB denote the domain of integration on the left hand side and right hand side of (LABEL:eq:ChangeOfVariable1), respectively. We consider the change of variables

(7.4) {v=v,y=yrv^,r=r.\left\{\begin{array}[]{l}v^{\prime}=v,\\ y^{\prime}=y-r\hat{v},\\ r^{\prime}=r.\end{array}\right.

Let X:ABX:A\to B be defined by X(v,y,r)=(v,yrv^,r)X(v,y,r)=(v,y-r\hat{v},r) and Y:BAY:B\to A be defined by Y(v,y,r)=(v,y+rv^,r)Y(v^{\prime},y^{\prime},r^{\prime})=(v^{\prime},y^{\prime}+r^{\prime}\hat{v^{\prime}},r^{\prime}).
Step 1 . X:ABX:A\to B is well-defined.
To verify well-definedness of XX, for given (v,y,r)A(v,y,r)\in A, yrv^Ωy-r\hat{v}\in\Omega clearly. We also notice that

(7.5) 0<r<r+|yq+(y,v)|=|(yrv^)q+(yrv^,v)|.0<r<r+|y-q_{+}(y,v)|=|(y-r\hat{v})-q_{+}(y-r\hat{v},v)|.

Therefore, X(v,y,r)BX(v,y,r)\in B and XX is well-defined.
Step 2 . Y:BAY:B\to A is well-defined.
To verify well-definedness of YY, for given (v,y,r)B(v^{\prime},y^{\prime},r^{\prime})\in B, y+rv^Ωy^{\prime}+r^{\prime}\hat{v^{\prime}}\in\Omega clearly. We note that

(7.6) 0<r<r+|yq(y,v)|=|(y+rv^)q(y+rv^,v)|.0<r^{\prime}<r^{\prime}+|y^{\prime}-q_{-}(y^{\prime},v^{\prime})|=|(y^{\prime}+r^{\prime}\hat{v^{\prime}})-q_{-}(y^{\prime}+r^{\prime}\hat{v^{\prime}},v^{\prime})|.

Therefore, Y(v,y,r)AY(v^{\prime},y^{\prime},r^{\prime})\in A and YY is well-defined.
Since XY=YX=idX\circ Y=Y\circ X=\text{id} and the Jacobian for the change of variables is clearly constant 11, we have (LABEL:eq:ChangeOfVariable1). ∎

Lemma 7.2.

We have the following formula for change of variables for any non-negative measurable function h=h(v,y,x,r)h=h(v,y,x,r).

(7.7) D1|xq(x,v)||yq(y,v)|h(v,y,x,r)𝑑r𝑑x𝑑y𝑑v=3ΩΩv,y|xq1(x,v)|min{|xq2(x,v)|,|yq+(y,v)|}h(v,y+rv^,x+rv^,r)drdxdydv,\begin{split}&\int_{D_{1}}\int^{|y-q_{-}(y,v)|}_{|x-q_{-}(x,v)|}h(v,y,x,r)\,drdxdydv\\ &=\int_{\mathbb{R}^{3}}\int_{\Omega}\int_{\Omega_{v^{\prime},y^{\prime}}}\int^{\min\{|x^{\prime}-q^{2}(x^{\prime},v^{\prime})|,|y^{\prime}-q_{+}(y^{\prime},v^{\prime})|\}}_{|x^{\prime}-q^{1}(x^{\prime},v^{\prime})|}\\ &\qquad\qquad\qquad h(v^{\prime},y^{\prime}+r^{\prime}\hat{v^{\prime}},x^{\prime}+r^{\prime}\hat{v^{\prime}},r^{\prime})\,dr^{\prime}dx^{\prime}dy^{\prime}dv^{\prime},\end{split}

where D1D_{1} is as defined by (6.2) and Ωv,y\Omega_{v^{\prime},y^{\prime}} is defined by (see Figure 5)

(7.8) Ωv,y={xΩ:there exist ti(x,v),i=1,2,such that 0<t1(x,v)<t2(x,v),qi(x,v):=x+ti(x,v)v^Ω,t1(x,v)<|yq+(y,v)|}.\begin{split}\Omega_{v^{\prime},y^{\prime}}=&\large\{x^{\prime}\notin\Omega:\,\text{there exist }t^{i}(x^{\prime},v^{\prime}),\,i=1,2,\\ &\text{such that }0<t^{1}(x^{\prime},v^{\prime})<t^{2}(x^{\prime},v^{\prime}),\\ &q^{i}(x^{\prime},v^{\prime}):=x^{\prime}+t^{i}(x^{\prime},v^{\prime})\hat{v^{\prime}}\in\partial\Omega,\\ &t^{1}(x^{\prime},v^{\prime})<|y^{\prime}-q_{+}(y^{\prime},v^{\prime})|\large\}.\end{split}
Refer to caption
Figure 5.
Remark 7.3.

An alternative way to characterize Ωv,y\Omega_{v^{\prime},y^{\prime}} is as follows. We first define

(7.9) Ωv,y={ztv^:zΩ, 0<t<|yq+(y,v)|}.\Omega^{\prime}_{v^{\prime},y^{\prime}}=\large\{z-t\hat{v^{\prime}}:\,z\in\Omega,\,0<t<|y^{\prime}-q_{+}(y^{\prime},v^{\prime})|\large\}.

Then, Ωv,y\Omega_{v^{\prime},y^{\prime}} can be expressed as

(7.10) Ωv,y=Ωv,yΩ¯.\Omega_{v^{\prime},y^{\prime}}=\Omega^{\prime}_{v^{\prime},y^{\prime}}\setminus\bar{\Omega}.
Proof.

We denote the domain of integration on the left hand side and right hand side by Aˇ\check{A} and Bˇ\check{B}, respectively. This time we consider

(7.11) {v=v,y=yrv^,x=xrv^,r=r.\left\{\begin{array}[]{l}v^{\prime}=v,\\ y^{\prime}=y-r\hat{v},\\ x^{\prime}=x-r\hat{v},\\ r^{\prime}=r.\end{array}\right.

Let Xˇ:AˇBˇ\check{X}:\check{A}\to\check{B} be defined by Xˇ(v,y,x,r)=(v,yrv^,xrv^,r)\check{X}(v,y,x,r)=(v,y-r\hat{v},x-r\hat{v},r) and Yˇ:BˇAˇ\check{Y}:\check{B}\to\check{A} be defined by Yˇ(v,y,x,r)=(v,y+rv^,x+rv^,r)\check{Y}(v^{\prime},y^{\prime},x^{\prime},r^{\prime})=(v^{\prime},y^{\prime}+r^{\prime}\hat{v^{\prime}},x^{\prime}+r^{\prime}\hat{v^{\prime}},r^{\prime}).
Step 1 . Xˇ:AˇBˇ\check{X}:\check{A}\to\check{B} is well-defined.
For given (v,y,x,r)Aˇ(v,y,x,r)\in\check{A}, clearly we have yrv^Ωy-r\hat{v}\in\Omega and xrv^Ωx-r\hat{v}\notin\Omega and there are two positive numbers t1(xrv^,v)<t2(xrv^,v)t^{1}(x-r\hat{v},v)<t^{2}(x-r\hat{v},v) such that

qi(xrv^,v)=xrv^+ti(xrv^,v)v^Ω.q^{i}(x-r\hat{v},v)=x-r\hat{v}+t^{i}(x-r\hat{v},v)\hat{v}\in\partial\Omega.

Moreover, comparing the distances results in

(7.12) t1(xrv^,v)=|(xrv^)q1(xrv^,v)|<|(xrv^)x|=r=|(yrv^)y|<|(yrv^)q+(yrv^,v)|.\begin{split}t^{1}(x-r\hat{v},v)&=|(x-r\hat{v})-q^{1}(x-r\hat{v},v)|\\ &<|(x-r\hat{v})-x|\\ &=r\\ &=|(y-r\hat{v})-y|\\ &<|(y-r\hat{v})-q_{+}(y-r\hat{v},v)|.\end{split}

On the other hand, we notice

(7.13) r=|(xrv^)x|<|(xrv^)q2(xrv^,v)|.\begin{split}r&=|(x-r\hat{v})-x|\\ &<|(x-r\hat{v})-q^{2}(x-r\hat{v},v)|.\end{split}

We conlcude that Xˇ(v,y,x,r)Bˇ\check{X}(v,y,x,r)\in\check{B} and therefore Xˇ\check{X} is well-defined.
Step 2 . Yˇ:BˇAˇ\check{Y}:\check{B}\to\check{A} is well-defined.
For given (v,y,x,r)Bˇ(v^{\prime},y^{\prime},x^{\prime},r^{\prime})\in\check{B}, since r|yq+(y,v)|r^{\prime}\leq|y^{\prime}-q_{+}(y^{\prime},v^{\prime})|, we have y+rv^Ωy^{\prime}+r^{\prime}\hat{v^{\prime}}\in\Omega. Likewise, since

|xq1(x,v)|<r<|xq2(x,v)|,|x^{\prime}-q^{1}(x^{\prime},v^{\prime})|<r<|x^{\prime}-q^{2}(x^{\prime},v^{\prime})|,

we have x+rv^Ωx^{\prime}+r^{\prime}\hat{v^{\prime}}\in\Omega. Comparing the distances yields

(7.14) |(x+rv^)q(x+rv^,v)|<|(x+rv^)x|=r=|(y+rv^)y|<|(y+rv^)q(y+rv^,v)|.\begin{split}|(x^{\prime}+r^{\prime}\hat{v^{\prime}})-q_{-}(x^{\prime}+r^{\prime}\hat{v^{\prime}},v^{\prime})|&<|(x^{\prime}+r^{\prime}\hat{v^{\prime}})-x^{\prime}|\\ &=r^{\prime}\\ &=|(y^{\prime}+r^{\prime}\hat{v^{\prime}})-y^{\prime}|\\ &<|(y^{\prime}+r^{\prime}\hat{v^{\prime}})-q_{-}(y^{\prime}+r^{\prime}\hat{v^{\prime}},v^{\prime})|.\end{split}

Therefore, it follows that Yˇ(v,y,x,r)Aˇ\check{Y}(v^{\prime},y^{\prime},x^{\prime},r^{\prime})\in\check{A} and therefore Yˇ\check{Y} is well-defined.
Since XˇYˇ=YˇXˇ=id\check{X}\circ\check{Y}=\check{Y}\circ\check{X}=\text{id} and the Jacobian for the change of variables is clearly constant 11, we conclude (LABEL:eq:ChangeOfVariable2). This completes the proof. ∎

Proposition 7.4.

Let hh be a function belonging to L2(Ω×3)L^{2}(\Omega\times\mathbb{R}^{3}). Then, there exists a constant CC such that

(7.15) |SΩKh(y,v)|2\displaystyle|S_{\Omega}Kh(y,v)|^{2}\leq C1|v|30|yq(y,v)||k(v,v)||h(yrv^,v)|2𝑑r𝑑v,\displaystyle C\,\frac{1}{|v|}\int_{\mathbb{R}^{3}}\int^{|y-q_{-}(y,v)|}_{0}|k(v,v_{*})||h(y-r\hat{v},v_{*})|^{2}\,drdv_{*},
(7.16) |KSΩh(y,v)|2\displaystyle|KS_{\Omega}h(y,v)|^{2}\leq C30|yq(y,v)|1|v||k(v,v)||h(yrv^,v)|2𝑑r𝑑v.\displaystyle C\,\int_{\mathbb{R}^{3}}\int^{|y-q_{-}(y,v_{*})|}_{0}\frac{1}{|v_{*}|}|k(v,v_{*})||h(y-r\hat{v_{*}},v_{*})|^{2}\,drdv_{*}.
Proof.

By Cauchy-Schwarz inequality, we see that

(7.17) |SΩKh(y,v)|2=(0τ(y,v)eν(v)sKh(ysv,v)𝑑s)2(0τ(y,v)e2ν(v)s𝑑s)(0τ(y,v)|Kh(ysv,v)|2𝑑s)C0τ(y,v)|Kh(ysv,v)|2𝑑s=C0|yq(y,v)|1|v||Kh(yrv^,v)|2𝑑r,\begin{split}|S_{\Omega}Kh(y,v)|^{2}=&\left(\int^{\tau_{-}(y,v)}_{0}e^{-\nu(v)s}Kh(y-sv,v)\,ds\right)^{2}\\ \leq&\left(\int^{\tau_{-}(y,v)}_{0}e^{-2\nu(v)s}\,ds\right)\left(\int^{\tau_{-}(y,v)}_{0}|Kh(y-sv,v)|^{2}\,ds\right)\\ \leq&C\,\int^{\tau_{-}(y,v)}_{0}|Kh(y-sv,v)|^{2}\,ds\\ =&C\,\int^{|y-q_{-}(y,v)|}_{0}\frac{1}{|v|}|Kh(y-r\hat{v},v)|^{2}\,dr,\end{split}

where r=s|v|r=s|v|. Furthermore, Corollary 3.2 implies that

(7.18) |Kh(yrv^,v)|2(3|k(v,v)|𝑑v)(3|k(v,v)||h(yrv^,v)|2𝑑v)C3|k(v,v)||h(yrv^,v)|2𝑑v.\begin{split}|Kh(y-r\hat{v},v)|^{2}\leq&\left(\int\limits_{\mathbb{R}^{3}}|k(v,v_{*})|\,dv_{*}\right)\left(\int\limits_{\mathbb{R}^{3}}|k(v,v_{*})||h(y-r\hat{v},v_{*})|^{2}\,dv_{*}\right)\\ \leq&C\,\int\limits_{\mathbb{R}^{3}}|k(v,v_{*})||h(y-r\hat{v},v_{*})|^{2}\,dv_{*}.\end{split}

Combining (7.17) and (7.18) yields (7.15).

To prove (7.16), we apply Cauchy-Schwarz inequality again to deduce that

(7.19) |KSΩh(y,v)|2=(30τ(y,v)eν(v)sk(v,v)h(ysv,v)𝑑s𝑑v)2(3|k(v,v)|𝑑v)(30τ(y,v)|k(v,v)||h(ysv,v)|2𝑑s𝑑v)C30τ(y,v)|k(v,v)||h(ysv,v)|2𝑑s𝑑v=C30|yq(y,v)|1|v||k(v,v)||h(yrv^,v)|2𝑑r𝑑v,\begin{split}|KS_{\Omega}h(y,v)|^{2}=&\left(\int_{\mathbb{R}^{3}}\int^{\tau_{-}(y,v_{*})}_{0}e^{-\nu(v_{*})s}k(v,v_{*})h(y-sv_{*},v_{*})\,dsdv_{*}\right)^{2}\\ \leq&\left(\int_{\mathbb{R}^{3}}|k(v,v_{*})|\,dv_{*}\right)\left(\int_{\mathbb{R}^{3}}\int^{\tau_{-}(y,v_{*})}_{0}|k(v,v_{*})||h(y-sv_{*},v_{*})|^{2}\,dsdv_{*}\right)\\ \leq&C\,\int_{\mathbb{R}^{3}}\int^{\tau_{-}(y,v_{*})}_{0}|k(v,v_{*})||h(y-sv_{*},v_{*})|^{2}\,dsdv_{*}\\ =&C\,\int_{\mathbb{R}^{3}}\int^{|y-q_{-}(y,v_{*})|}_{0}\frac{1}{|v_{*}|}|k(v,v_{*})||h(y-r\hat{v_{*}},v_{*})|^{2}\,drdv_{*},\end{split}

where r=s|v|r=s|v_{*}|. ∎

We are now in a position to prove Lemma 1.12.

Proof of Lemma 1.12.

For given fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}), we recall fif_{i} denotes the function (SΩK)if(S_{\Omega}K)^{i}f. To prove the lemma, it suffices to show

(7.20) D1|f2(x,v)f2(y,v)|2|xy|4𝑑x𝑑y𝑑vCfL2(Ω×3)2,\int_{D_{1}}\frac{|f_{2}(x,v)-f_{2}(y,v)|^{2}}{|x-y|^{4}}\,dxdydv\leq C\,\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})},

where D1D_{1} is as defined by (6.2). According to Corollary 1.10, we have

(7.21) Kf1Lv2(3;Hx1/2(Ω))CfL2(Ω×3).\|Kf_{1}\|_{L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\Omega))}\leq C\,\|f\|_{L^{2}(\Omega\times\mathbb{R}^{3})}.

In domain D1D_{1}, we notice that τ(x,v)τ(y,v)\tau_{-}(x,v)\leq\tau_{-}(y,v). Therefore,

(7.22) |SΩKf1(x,v)SΩKf1(y,v)|22|0τ(x,v)eν(v)s(Kf1(xsv,v)Kf1(ysv,v))𝑑s|2+2|τ(x,v)τ(y,v)eν(v)sKf1(ysv,v)𝑑s|2.\begin{split}&|S_{\Omega}Kf_{1}(x,v)-S_{\Omega}Kf_{1}(y,v)|^{2}\\ &\qquad\leq 2\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(Kf_{1}(x-sv,v)-Kf_{1}(y-sv,v))\,ds\right|^{2}\\ &\qquad\quad+2\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}Kf_{1}(y-sv,v)\,ds\right|^{2}.\end{split}

Hence, we have

(7.23) D1|f2(x,v)f2(y,v)|2|xy|4𝑑x𝑑y𝑑vI1+I2,\int\limits_{D_{1}}\frac{|f_{2}(x,v)-f_{2}(y,v)|^{2}}{|x-y|^{4}}\,dxdydv\leq I_{1}+I_{2},

where

(7.24) I1\displaystyle I_{1} :=D12|0τ(x,v)eν(v)s(Kf1(xsv,v)Kf1(ysv,v))𝑑s|2|xy|4𝑑x𝑑y𝑑v,\displaystyle:=\int_{D_{1}}\frac{2\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(Kf_{1}(x-sv,v)-Kf_{1}(y-sv,v))\,ds\right|^{2}}{|x-y|^{4}}\,dxdydv,
(7.25) I2\displaystyle I_{2} :=D12|τ(x,v)τ(y,v)eν(v)sKf1(ysv,v)𝑑s|2|xy|4𝑑x𝑑y𝑑v.\displaystyle:=\int_{D_{1}}\frac{2\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}Kf_{1}(y-sv,v)\,ds\right|^{2}}{|x-y|^{4}}\,dxdydv.

To estimate I1I_{1}, taking ϵ=12\epsilon=\frac{1}{2} and h=Kf1h=Kf_{1} in the steps of (LABEL:eq:SOextension1) and (6.29), in the same fashion, we conclude

(7.26) D1|0τ(x,v)eν(v)s(Kf1(xsv,v)Kf1(ysv,v))𝑑s|2|xy|4𝑑x𝑑y𝑑vCKf1Lv2(3;Hx1/2(Ω))2CfL2(Ω×3)2,\begin{split}&\int\limits_{D_{1}}\frac{\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(Kf_{1}(x-sv,v)-Kf_{1}(y-sv,v))\,ds\right|^{2}}{|x-y|^{4}}\,dxdydv\\ &\qquad\leq C\,\|Kf_{1}\|^{2}_{L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\Omega))}\\ &\qquad\leq C\,\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})},\end{split}

where we have used (7.21). To deal with I2I_{2}, by Cauchy-Schwarz inequality again, we have

(7.27) |τ(x,v)τ(y,v)eν(v)sKf1(ysv,v)𝑑s|2(τ(x,v)τ(y,v)eν(v)s𝑑s)(τ(x,v)τ(y,v)eν(v)s|Kf1(ysv,v)|2𝑑s).\begin{split}&\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}Kf_{1}(y-sv,v)\,ds\right|^{2}\\ &\qquad\leq\left(\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}\,ds\right)\left(\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}|Kf_{1}(y-sv,v)|^{2}\,ds\right).\end{split}

For small ϵ>0\epsilon>0,

(7.28) τ(x,v)τ(y,v)eν(v)s𝑑s(0eν(v)s𝑑s)ϵ(τ(x,v)τ(y,v)eν(v)s𝑑s)1ϵ1ν0ϵ|τ(x,v)τ(y,v)|1ϵC|xy|1ϵN(x,v)1ϵ|v|1ϵ,\begin{split}\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}\,ds\leq&\left(\int^{\infty}_{0}e^{-\nu(v)s}\,ds\right)^{\epsilon}\left(\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}\,ds\right)^{1-\epsilon}\\ &\leq\frac{1}{\nu_{0}^{\epsilon}}|\tau_{-}(x,v)-\tau_{-}(y,v)|^{1-\epsilon}\\ &\leq C\,\frac{|x-y|^{1-\epsilon}}{N_{-}(x,v)^{1-\epsilon}|v|^{1-\epsilon}},\end{split}

where the last inequality follows from Proposition 5.3. On the other hand, the change of variable r=s|v|r=s|v| results in

(7.29) τ(x,v)τ(y,v)eν(v)s|Kf1(ysv,v)|2𝑑s=1|v||xq(x,v)||yq(y,v)|eν(v)|v|r|Kf1(yrv^,v)|2𝑑r1|v||xq(x,v)||yq(y,v)||Kf1(yrv^,v)|2𝑑r.\begin{split}\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}|Kf_{1}(y-sv,v)|^{2}\,ds=&\frac{1}{|v|}\int^{|y-q_{-}(y,v)|}_{|x-q_{-}(x,v)|}e^{-\frac{\nu(v)}{|v|}r}|Kf_{1}(y-r\hat{v},v)|^{2}\,dr\\ \leq&\frac{1}{|v|}\int^{|y-q_{-}(y,v)|}_{|x-q_{-}(x,v)|}|Kf_{1}(y-r\hat{v},v)|^{2}\,dr.\end{split}

Combining (7.27)-(7.29), we have

(7.30) |τ(x,v)τ(y,v)eν(v)sKf1(ysv,v)𝑑s|2C|xy|1ϵN(x,v)1ϵ|v|2ϵ|xq(x,v)||yq(y,v)||Kf1(yrv^,v)|2𝑑r.\begin{split}&\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}Kf_{1}(y-sv,v)\,ds\right|^{2}\\ &\qquad\leq C\,\frac{|x-y|^{1-\epsilon}}{N_{-}(x,v)^{1-\epsilon}|v|^{2-\epsilon}}\int^{|y-q_{-}(y,v)|}_{|x-q_{-}(x,v)|}|Kf_{1}(y-r\hat{v},v)|^{2}\,dr.\end{split}

Therefore, it follows that

(7.31) I2CD1|xq(x,v)||yq(y,v)||Kf1(yrv^,v)|2N(x,v)1ϵ|v|2ϵ|xy|3+ϵ𝑑r𝑑x𝑑y𝑑v.I_{2}\leq C\,\int_{D_{1}}\int^{|y-q_{-}(y,v)|}_{|x-q_{-}(x,v)|}\frac{|Kf_{1}(y-r\hat{v},v)|^{2}}{N_{-}(x,v)^{1-\epsilon}|v|^{2-\epsilon}|x-y|^{3+\epsilon}}\,drdxdydv.

Letting x=xrv^x^{\prime}=x-r\hat{v} and y=yrv^y^{\prime}=y-r\hat{v}, by Lemma 7.2, we obtain

(7.32) D1|xq(x,v)||yq(y,v)||Kf1(yrv^,v)|2N(x,v)1ϵ|v|2ϵ|xy|3+ϵ𝑑r𝑑x𝑑y𝑑v=3ΩΩv,y|q1(x,v)x|min{|q2(x,v)x|,|q+(y,v)y|}|Kf1(y,v)|2N(q1(x,v),v)1ϵ|v|2ϵ|xy|3+ϵdrdxdydv3ΩΩv,y|Kf1(y,v)|2|q2(x,v)q1(x,v)|N(q1(x,v),v)1ϵ|v|2ϵ|xy|3+ϵ𝑑x𝑑y𝑑vC3ΩΩv,y|Kf1(y,v)|2|v|2ϵ|xy|3+ϵ𝑑x𝑑y𝑑v,\begin{split}&\int_{D_{1}}\int^{|y-q_{-}(y,v)|}_{|x-q_{-}(x,v)|}\frac{|Kf_{1}(y-r\hat{v},v)|^{2}}{N_{-}(x,v)^{1-\epsilon}|v|^{2-\epsilon}|x-y|^{3+\epsilon}}\,drdxdydv\\ &\quad=\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega_{v,y^{\prime}}}\int^{\min\{|q^{2}(x^{\prime},v)-x^{\prime}|,|q_{+}(y^{\prime},v)-y^{\prime}|\}}_{|q^{1}(x^{\prime},v)-x^{\prime}|}\\ &\qquad\qquad\frac{|Kf_{1}(y^{\prime},v)|^{2}}{N_{-}(q^{1}(x^{\prime},v),v)^{1-\epsilon}|v|^{2-\epsilon}|x^{\prime}-y^{\prime}|^{3+\epsilon}}\,drdx^{\prime}dy^{\prime}dv\\ &\quad\leq\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega_{v,y^{\prime}}}\frac{|Kf_{1}(y^{\prime},v)|^{2}|q^{2}(x^{\prime},v)-q^{1}(x^{\prime},v)|}{N_{-}(q^{1}(x^{\prime},v),v)^{1-\epsilon}|v|^{2-\epsilon}|x^{\prime}-y^{\prime}|^{3+\epsilon}}\,dx^{\prime}dy^{\prime}dv\\ &\quad\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega_{v,y^{\prime}}}\frac{|Kf_{1}(y^{\prime},v)|^{2}}{|v|^{2-\epsilon}|x^{\prime}-y^{\prime}|^{3+\epsilon}}\,dx^{\prime}dy^{\prime}dv,\end{split}

where we have utilized Proposition 5.9 by identifying q1(x,v)=q(x,v)q^{1}(x^{\prime},v)=q_{-}(x,v) and q2(x,v)=q+(x,v)q^{2}(x^{\prime},v)=q_{+}(x,v). Noticing that from (7.8), Ωv,y\Omega_{v,y^{\prime}} lies outside of Ω\Omega, we have

Ωv,y(3Ω)(3Bdy(y)).\Omega_{v,y^{\prime}}\subset\left(\mathbb{R}^{3}\setminus\Omega\right)\subset\left(\mathbb{R}^{3}\setminus B_{d_{y^{\prime}}}(y^{\prime})\right).

With this in mind, we deduce

(7.33) I2C3ΩΩv,y|Kf1(y,v)|2|v|2ϵ|xy|3+ϵ𝑑x𝑑y𝑑v3Ω3Bdy(y)|Kf1(y,v)|2|v|2ϵ|xy|3+ϵ𝑑x𝑑y𝑑vC3Ω|Kf1(y,v)|2|v|2ϵdyϵ𝑑y𝑑v.\begin{split}I_{2}&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega_{v,y^{\prime}}}\frac{|Kf_{1}(y^{\prime},v)|^{2}}{|v|^{2-\epsilon}|x^{\prime}-y^{\prime}|^{3+\epsilon}}\,dx^{\prime}dy^{\prime}dv\\ &\leq\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int_{\mathbb{R}^{3}\setminus B_{d_{y^{\prime}}}(y^{\prime})}\frac{|Kf_{1}(y^{\prime},v)|^{2}}{|v|^{2-\epsilon}|x^{\prime}-y^{\prime}|^{3+\epsilon}}\,dx^{\prime}dy^{\prime}dv\\ &\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\frac{|Kf_{1}(y^{\prime},v)|^{2}}{|v|^{2-\epsilon}d^{\epsilon}_{y^{\prime}}}\,dy^{\prime}dv.\\ \end{split}

Proposition 7.4 and Lemma 7.1 imply

(7.34) 3Ω|Kf1(y,v)|2|v|2ϵdyϵ𝑑y𝑑vC3Ω30|yq(y,v)||k(v,v)||Kf(yrv^,v)|2|v||v|2ϵdyϵ𝑑r𝑑v𝑑y𝑑v=C3Ω30|y′′q+(y′′,v)||k(v,v)||Kf(y′′,v)|2|v||v|2ϵdy′′+rv^ϵ𝑑r𝑑v𝑑y′′𝑑vC3Ω3|k(v,v)||Kf(y′′,v)|2|v||v|2ϵ𝑑v𝑑y′′𝑑v,\begin{split}&\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\frac{|Kf_{1}(y^{\prime},v)|^{2}}{|v|^{2-\epsilon}d^{\epsilon}_{y^{\prime}}}\,dy^{\prime}dv\\ &\quad\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int^{|y^{\prime}-q_{-}(y^{\prime},v_{*})|}_{0}\frac{|k(v,v_{*})||Kf(y^{\prime}-r\hat{v_{*}},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}d_{y^{\prime}}^{\epsilon}}\,drdv_{*}dy^{\prime}dv\\ &\quad=C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int^{|y^{\prime\prime}-q_{+}(y^{\prime\prime},v_{*})|}_{0}\frac{|k(v,v_{*})||Kf(y^{\prime\prime},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}d_{y^{\prime\prime}+r\hat{v_{*}}}^{\epsilon}}\,drdv_{*}dy^{\prime\prime}dv\\ &\quad\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|k(v,v_{*})||Kf(y^{\prime\prime},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}}\,dv_{*}dy^{\prime\prime}dv,\end{split}

where y′′=yrv^y^{\prime\prime}=y^{\prime}-r\hat{v_{*}} and we have used Lemma 5.12 in the last line. Utilizing Proposition 3.4, we conclude

(7.35) I2CΩ33|k(v,v)||Kf(y′′,v)|2|v||v|2ϵ𝑑v𝑑v𝑑y′′CΩ3|Kf(y′′,v)|2|v|𝑑v𝑑y′′CΩ31|v|(3|k(v,w)|𝑑w)(3|k(v,w)||f(y′′,w)|2𝑑w)𝑑v𝑑y′′CΩ33|k(v,w)||f(y′′,w)|2|v|𝑑v𝑑w𝑑y′′CΩ3|f(y′′,w)|2𝑑w𝑑y′′.\begin{split}I_{2}&\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\frac{|k(v,v_{*})||Kf(y^{\prime\prime},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}}\,dvdv_{*}dy^{\prime\prime}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|Kf(y^{\prime\prime},v_{*})|^{2}}{|v_{*}|}\,dv_{*}dy^{\prime\prime}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{1}{|v_{*}|}\left(\int\limits_{\mathbb{R}^{3}}|k(v_{*},w)|\,dw\right)\left(\int\limits_{\mathbb{R}^{3}}|k(v_{*},w)||f(y^{\prime\prime},w)|^{2}\,dw\right)\,dv_{*}dy^{\prime\prime}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\frac{|k(v_{*},w)||f(y^{\prime\prime},w)|^{2}}{|v_{*}|}\,dv_{*}dwdy^{\prime\prime}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}|f(y^{\prime\prime},w)|^{2}dwdy^{\prime\prime}.\end{split}

This completes the proof. ∎

Let us elaborate on Lemma 1.13 for a while. For fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}), in view of Lemma 1.12, the function SΩKSΩKfS_{\Omega}KS_{\Omega}Kf belongs to Lv2(3;Hx1/2(Ω))L^{2}_{v}(\mathbb{R}^{3};H^{\nicefrac{{1}}{{2}}}_{x}(\Omega)). Now consider its zero extension SΩKSΩKf~\widetilde{S_{\Omega}KS_{\Omega}Kf} in the whole space 3×3\mathbb{R}^{3}\times\mathbb{R}^{3}. We claim that SΩKSΩKf~\widetilde{S_{\Omega}KS_{\Omega}Kf} belongs to Lv2(3;Hx12ϵ(3))L^{2}_{v}(\mathbb{R}^{3};H^{\frac{1}{2}-\epsilon}_{x}(\mathbb{R}^{3})) for any ϵ(0,12)\epsilon\in(0,\frac{1}{2}) and

(7.36) SΩKSΩKf~Lv2(3;Hx12ϵ(3))2CϵfL2(Ω×3)2\|\widetilde{S_{\Omega}KS_{\Omega}Kf}\|^{2}_{L^{2}_{v}(\mathbb{R}^{3};H^{\frac{1}{2}-\epsilon}_{x}(\mathbb{R}^{3}))}\leq\frac{C}{\epsilon}\,\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})}

for some constant CC.

Proof of Lemma 1.13.

Recall f1:=SΩKff_{1}:=S_{\Omega}Kf and f2:=SΩKSΩKff_{2}:=S_{\Omega}KS_{\Omega}Kf. Since f2~(,v)\widetilde{f_{2}}(\cdot,v) vanishes outside of Ω\Omega, we have

(7.37) 333|f2~(x,v)f2~(y,v)|2|xy|42ϵ𝑑x𝑑y𝑑v=I1+I2+I3,\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\frac{|\widetilde{f_{2}}(x,v)-\widetilde{f_{2}}(y,v)|^{2}}{|x-y|^{4-2\epsilon}}\,dxdydv=I_{1}+I_{2}+I_{3},

where

(7.38) I1:=\displaystyle I_{1}:= 3ΩΩ|f2(x,v)f2(y,v)|2|xy|42ϵ𝑑x𝑑y𝑑v,\displaystyle\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega}\frac{|f_{2}(x,v)-f_{2}(y,v)|^{2}}{|x-y|^{4-2\epsilon}}\,dxdydv,
(7.39) I2:=\displaystyle I_{2}:= 3Ω3Ω|f2(y,v)|2|xy|42ϵ𝑑x𝑑y𝑑v,\displaystyle\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}\setminus\Omega}\frac{|f_{2}(y,v)|^{2}}{|x-y|^{4-2\epsilon}}\,dxdydv,
(7.40) I3:=\displaystyle I_{3}:= 33ΩΩ|f2(x,v)|2|xy|42ϵ𝑑x𝑑y𝑑v.\displaystyle\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}\setminus\Omega}\int\limits_{\Omega}\frac{|f_{2}(x,v)|^{2}}{|x-y|^{4-2\epsilon}}\,dxdydv.

Regarding the boundedness of I1I_{1}, by Lemma 1.12, we have

(7.41) I1=3ΩΩ|f2(x,v)f2(y,v)|2|xy|4|xy|2ϵ𝑑x𝑑y𝑑vmax{diamΩ,1}3ΩΩ|f2(x,v)f2(y,v)|2|xy|4𝑑x𝑑y𝑑vCfL2(Ω×3)2.\begin{split}I_{1}&=\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega}\frac{|f_{2}(x,v)-f_{2}(y,v)|^{2}}{|x-y|^{4}}\,|x-y|^{2\epsilon}\,dxdydv\\ &\leq\max\{\operatorname{diam}\Omega,1\}\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega}\frac{|f_{2}(x,v)-f_{2}(y,v)|^{2}}{|x-y|^{4}}\,dxdydv\\ &\leq C\,\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})}.\end{split}

By symmetry, one can see that I2=I3I_{2}=I_{3}. Consequently, the remaining task is to deal with I2I_{2}. We observe that, for yΩy\in\Omega,

(7.42) 3Ω1|xy|42ϵ𝑑x3Bdy(y)1|xy|42ϵ𝑑xCdy1+2ϵ.\int\limits_{\mathbb{R}^{3}\setminus\Omega}\frac{1}{|x-y|^{4-2\epsilon}}\,dx\leq\int\limits_{\mathbb{R}^{3}\setminus B_{d_{y}}(y)}\frac{1}{|x-y|^{4-2\epsilon}}\,dx\leq C\,d_{y}^{-1+2\epsilon}.

By Proposition 7.4, we have

(7.43) |f2(y,v)|2=|SΩKf1(y,v)|2C30|yq(y,v)|1|v||k(v,v)||f1(yrv^,v)|2𝑑r𝑑v.\begin{split}|f_{2}(y,v)|^{2}=&|S_{\Omega}Kf_{1}(y,v)|^{2}\\ \leq&C\,\int_{\mathbb{R}^{3}}\int^{|y-q_{-}(y,v)|}_{0}\frac{1}{|v|}|k(v,v_{*})||f_{1}(y-r\hat{v},v_{*})|^{2}\,drdv_{*}.\end{split}

Therefore, first combining (7.42) and (7.43) and performing the change of variable y=yrv^y^{\prime}=y-r\hat{v} as in Lemma 7.1, we deduce that

(7.44) I2C3Ω30|yq(y,v)|1|v||k(v,v)||f1(yrv^,v)|2dy1+2ϵ𝑑r𝑑y𝑑v𝑑v=C3Ω30|yq+(yv)|1|v||k(v,v)||f1(y,v)|2dy+rv^1+2ϵ𝑑r𝑑y𝑑v𝑑vCϵ3Ω3(1|v||k(v,v)|dv)|f1(y,v)|2dy12+2ϵ𝑑y𝑑vCϵ3Ω|f1(y,v)|2dy12+2ϵ𝑑y𝑑v,\begin{split}I_{2}&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int^{|y-q_{-}(y,v)|}_{0}\frac{1}{|v|}|k(v,v_{*})||f_{1}(y-r\hat{v},v_{*})|^{2}d_{y}^{-1+2\epsilon}\,drdydvdv_{*}\\ &=C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int^{|y^{\prime}-q_{+}(y^{\prime}v)|}_{0}\frac{1}{|v|}|k(v,v_{*})||f_{1}(y^{\prime},v_{*})|^{2}d_{y^{\prime}+r\hat{v}}^{-1+2\epsilon}\,drdy^{\prime}dvdv_{*}\\ &\leq\frac{C}{\epsilon}\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\left(\frac{1}{|v|}|k(v,v_{*})|\,dv\right)\,|f_{1}(y^{\prime},v_{*})|^{2}d_{y^{\prime}}^{-\frac{1}{2}+2\epsilon}\,dy^{\prime}dv_{*}\\ &\leq\frac{C}{\epsilon}\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}|f_{1}(y^{\prime},v_{*})|^{2}d_{y^{\prime}}^{-\frac{1}{2}+2\epsilon}\,dy^{\prime}dv_{*},\end{split}

where the third inequality follows from Lemma 5.12 and the last inequality follows from Proposition 3.4. Applying Proposition 7.4 again, we have

(7.45) |f1(y,v)|2C30|yq(y,v)|1|v||k(v,w)||f(yrv^,w)|2𝑑r𝑑w.|f_{1}(y^{\prime},v_{*})|^{2}\leq C\,\int_{\mathbb{R}^{3}}\int^{|y^{\prime}-q_{-}(y^{\prime},v_{*})|}_{0}\frac{1}{|v_{*}|}|k(v_{*},w)||f(y^{\prime}-r\hat{v_{*}},w)|^{2}\,drdw.

Then, it follows that

(7.46) 3Ω|f1(y,v)|2dy12+2ϵ𝑑y𝑑vC3Ω30|yq(y,v)|1|v||k(v,w)||f(yrv^,w)|2dy12+2ϵ𝑑r𝑑w𝑑y𝑑v=C33Ω0|y′′q+(y′′,v)|1|v||k(v,w)||f(y′′,w)|2dy′′+rv^12+2ϵ𝑑r𝑑y′′𝑑v𝑑w,\begin{split}&\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}|f_{1}(y^{\prime},v_{*})|^{2}d_{y^{\prime}}^{-\frac{1}{2}+2\epsilon}\,dy^{\prime}dv_{*}\\ &\quad\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int^{|y^{\prime}-q_{-}(y^{\prime},v_{*})|}_{0}\frac{1}{|v_{*}|}|k(v_{*},w)||f(y^{\prime}-r\hat{v_{*}},w)|^{2}d_{y^{\prime}}^{-\frac{1}{2}+2\epsilon}\,drdwdy^{\prime}dv_{*}\\ &\quad=C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int^{|y^{\prime\prime}-q_{+}(y^{\prime\prime},v_{*})|}_{0}\frac{1}{|v_{*}|}|k(v_{*},w)||f(y^{\prime\prime},w)|^{2}d_{y^{\prime\prime}+r\hat{v_{*}}}^{-\frac{1}{2}+2\epsilon}\,drdy^{\prime\prime}dv_{*}dw,\end{split}

where y′′=yrv^y^{\prime\prime}=y^{\prime}-r\hat{v_{*}}. Similarly, by Lemma 5.12 and Proposition 3.4, we conclude that

(7.47) I2Cϵ3Ω|f1(y,v)|2dy12+2ϵ𝑑y𝑑vCϵ3Ω31|v||k(v,w)||f(y′′,w)|2𝑑v𝑑y′′𝑑wCϵ3Ω|f(y′′,w)|2𝑑y′′𝑑w.\begin{split}I_{2}&\leq\frac{C}{\epsilon}\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}|f_{1}(y^{\prime},v_{*})|^{2}d_{y^{\prime}}^{-\frac{1}{2}+2\epsilon}\,dy^{\prime}dv_{*}\\ &\leq\frac{C}{\epsilon}\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{1}{|v_{*}|}|k(v_{*},w)||f(y^{\prime\prime},w)|^{2}\,dv_{*}dy^{\prime\prime}dw\\ &\leq\frac{C}{\epsilon}\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}|f(y^{\prime\prime},w)|^{2}\,dy^{\prime\prime}dw.\end{split}

Combining (7.41) and (7.47) completes the proof. ∎

Lemma 1.13 allows us to improve regularity results via Lemma 1.2.

Corollary 7.5.

The operator KSΩKSΩKSΩK:L2(Ω×3)Lv2(3;Hx1ϵ(Ω))KS_{\Omega}KS_{\Omega}KS_{\Omega}K:L^{2}(\Omega\times\mathbb{R}^{3})\to L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) is bounded for any ϵ(0,12)\epsilon\in(0,\frac{1}{2}).

Proof.

For given fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}) and ϵ(0,12)\epsilon\in(0,\frac{1}{2}), Lemma 1.13 implies

(7.48) SΩKSΩKf~Lv2(3;Hx12ϵ(3))=Lv2(3;H~x12ϵ(3)).\widetilde{S_{\Omega}KS_{\Omega}Kf}\in L^{2}_{v}(\mathbb{R}^{3};H^{\frac{1}{2}-\epsilon}_{x}(\mathbb{R}^{3}))=L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{\frac{1}{2}-\epsilon}_{x}(\mathbb{R}^{3})).

By Lemma 1.2, we obtain

(7.49) KSK(SΩKSΩKf~)Lv2(3;H~x1ϵ(3)).KSK(\widetilde{S_{\Omega}KS_{\Omega}Kf})\in L^{2}_{v}(\mathbb{R}^{3};\tilde{H}^{1-\epsilon}_{x}(\mathbb{R}^{3})).

Since

(7.50) (KSK(SΩKSΩKf~))|Ω×3=KSΩKSΩKSΩKf,\left.\biggl{(}KSK(\widetilde{S_{\Omega}KS_{\Omega}Kf})\biggr{)}\right|_{\Omega\times\mathbb{R}^{3}}=KS_{\Omega}KS_{\Omega}KS_{\Omega}Kf,

consequently KSΩKSΩKSΩKfKS_{\Omega}KS_{\Omega}KS_{\Omega}Kf belongs to Lv2(3;Hx1ϵ(Ω))L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)). The boundedness of the operator is due to boundedness of KSKKSK and ZSΩKSΩKZS_{\Omega}KS_{\Omega}K. ∎

We are now ready to prove the last lemma in this section.

Lemma 7.6.

The operator SΩKSΩKSΩKSΩK:L2(Ω×3)Lv2(3;Hx1ϵ(Ω))S_{\Omega}KS_{\Omega}KS_{\Omega}KS_{\Omega}K:L^{2}(\Omega\times\mathbb{R}^{3})\to L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)) is bounded for any ϵ(0,12)\epsilon\in(0,\frac{1}{2}). Therefore, f4Lv2(3;Hx1ϵ(Ω))f_{4}\in L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega)).

Proof.

We shall prove this lemma in a similar fashion as we prove Lemma 1.12. To do so, for given fL2(Ω×3)f\in L^{2}(\Omega\times\mathbb{R}^{3}) and ϵ(0,12)\epsilon\in(0,\frac{1}{2}), it suffices to show

(7.51) D1|f4(x,v)f4(y,v)|2|xy|52ϵ𝑑x𝑑y𝑑vCfL2(Ω×3)2,\int_{D_{1}}\frac{|f_{4}(x,v)-f_{4}(y,v)|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv\leq C\,\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})},

where D1D_{1} is as defined by (6.2). According to Corollary 7.5, it follows that

(7.52) Kf3Lv2(3;Hx1ϵ(Ω))CfL2(Ω×3).\|Kf_{3}\|_{L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega))}\leq C\,\|f\|_{L^{2}(\Omega\times\mathbb{R}^{3})}.

In domain D1D_{1}, we have

(7.53) |SΩKf3(x,v)SΩKf3(y,v)|22|0τ(x,v)eν(v)s(Kf3(xsv,v)Kf3(ysv,v))𝑑s|2+2|τ(x,v)τ(y,v)eν(v)sKf3(ysv,v)𝑑s|2.\begin{split}&|S_{\Omega}Kf_{3}(x,v)-S_{\Omega}Kf_{3}(y,v)|^{2}\\ &\qquad\leq 2\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(Kf_{3}(x-sv,v)-Kf_{3}(y-sv,v))\,ds\right|^{2}\\ &\qquad\quad+2\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}Kf_{3}(y-sv,v)\,ds\right|^{2}.\end{split}

Therefore, we have

(7.54) D1|f4(x,v)f4(y,v)|2|xy|52ϵ𝑑x𝑑y𝑑vI1+I2,\int\limits_{D_{1}}\frac{|f_{4}(x,v)-f_{4}(y,v)|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv\leq I_{1}+I_{2},

where

(7.55) I1\displaystyle I_{1} :=D12|0τ(x,v)eν(v)s(Kf3(xsv,v)Kf3(ysv,v))𝑑s|2|xy|52ϵ𝑑x𝑑y𝑑v,\displaystyle:=\int_{D_{1}}\frac{2\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(Kf_{3}(x-sv,v)-Kf_{3}(y-sv,v))\,ds\right|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv,
(7.56) I2\displaystyle I_{2} :=D12|τ(x,v)τ(y,v)eν(v)sKf3(ysv,v)𝑑s|2|xy|52ϵ𝑑x𝑑y𝑑v.\displaystyle:=\int_{D_{1}}\frac{2\left|\int^{\tau_{-}(y,v)}_{\tau_{-}(x,v)}e^{-\nu(v)s}Kf_{3}(y-sv,v)\,ds\right|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv.

To estimate I1I_{1}, taking h=Kf3h=Kf_{3} in the steps of (LABEL:eq:SOextension1) and (6.29), in the same fashion, we conclude

(7.57) D1|0τ(x,v)eν(v)s(Kf3(xsv,v)Kf3(ysv,v))𝑑s|2|xy|52ϵ𝑑x𝑑y𝑑vCKf3Lv2(3;Hx1ϵ(Ω))2CfL2(Ω×3)2,\begin{split}&\int\limits_{D_{1}}\frac{\left|\int^{\tau_{-}(x,v)}_{0}e^{-\nu(v)s}(Kf_{3}(x-sv,v)-Kf_{3}(y-sv,v))\,ds\right|^{2}}{|x-y|^{5-2\epsilon}}\,dxdydv\\ &\qquad\leq C\,\|Kf_{3}\|^{2}_{L^{2}_{v}(\mathbb{R}^{3};H^{1-\epsilon}_{x}(\Omega))}\\ &\qquad\leq C\,\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})},\end{split}

where we have used (7.52). Concerning I2I_{2}, we proceed as in (7.30)-(7.34) to deduce

(7.58) I2C3ΩΩ|xq(x,v)||yq(y,v)||Kf3(yrv^,v)|2N(x,v)1ϵ|v|2ϵ|xy|4ϵ𝑑r𝑑x𝑑y𝑑v=C3ΩΩv,y|q1(x,v)x|min{|q2(x,v)x|,|q+(y,v)y|}|Kf3(y,v)|2N(q1(x,v),v)1ϵ|v|2ϵ|xy|4ϵdrdxdydvC3ΩΩv,y|Kf3(y,v)|2|v|2ϵ|xy|4ϵ𝑑x𝑑y𝑑v\begin{split}I_{2}&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega}\int^{|y-q_{-}(y,v)|}_{|x-q_{-}(x,v)|}\frac{|Kf_{3}(y-r\hat{v},v)|^{2}}{N_{-}(x,v)^{1-\epsilon}|v|^{2-\epsilon}|x-y|^{4-\epsilon}}\,drdxdydv\\ &=C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega_{v,y^{\prime}}}\int^{\min\{|q^{2}(x^{\prime},v)-x^{\prime}|,|q_{+}(y^{\prime},v)-y^{\prime}|\}}_{|q^{1}(x^{\prime},v)-x^{\prime}|}\\ &\qquad\quad\frac{|Kf_{3}(y^{\prime},v)|^{2}}{N_{-}(q^{1}(x^{\prime},v),v)^{1-\epsilon}|v|^{2-\epsilon}|x^{\prime}-y^{\prime}|^{4-\epsilon}}\,drdx^{\prime}dy^{\prime}dv\\ &\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\Omega_{v,y^{\prime}}}\frac{|Kf_{3}(y^{\prime},v)|^{2}}{|v|^{2-\epsilon}|x^{\prime}-y^{\prime}|^{4-\epsilon}}\,dx^{\prime}dy^{\prime}dv\end{split}
(7.59) C3Ω|Kf3(y,v)|2|v|2ϵdy1ϵ𝑑y𝑑vC3Ω30|yq(y,v)||k(v,v)||Kf2(yrv^,v)|2|v||v|2ϵdy1ϵ𝑑r𝑑v𝑑y𝑑v=C3Ω30|y′′q+(y′′,v)||k(v,v)||Kf2(y′′,v)|2|v||v|2ϵdy′′+rv^1ϵ𝑑r𝑑v𝑑y′′𝑑vC3Ω3|k(v,v)||Kf2(y′′,v)|2|v||v|2ϵdy′′12ϵ𝑑v𝑑y′′𝑑v,\begin{split}\quad&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\frac{|Kf_{3}(y^{\prime},v)|^{2}}{|v|^{2-\epsilon}d^{1-\epsilon}_{y^{\prime}}}\,dy^{\prime}dv\\ \quad&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int^{|y^{\prime}-q_{-}(y^{\prime},v_{*})|}_{0}\frac{|k(v,v_{*})||Kf_{2}(y^{\prime}-r\hat{v_{*}},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}d_{y^{\prime}}^{1-\epsilon}}\,drdv_{*}dy^{\prime}dv\\ \quad&=C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int^{|y^{\prime\prime}-q_{+}(y^{\prime\prime},v_{*})|}_{0}\frac{|k(v,v_{*})||Kf_{2}(y^{\prime\prime},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}d_{y^{\prime\prime}+r\hat{v_{*}}}^{1-\epsilon}}\,drdv_{*}dy^{\prime\prime}dv\\ \quad&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|k(v,v_{*})||Kf_{2}(y^{\prime\prime},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}d_{y^{\prime\prime}}^{\frac{1}{2}-\epsilon}}\,dv_{*}dy^{\prime\prime}dv,\end{split}

where we utilized (5.22) of Lemma 5.12 in the last line instead. Continuing as in (7.35) yields

(7.60) I2C3Ω3|k(v,v)||Kf2(y′′,v)|2|v||v|2ϵdy′′12ϵ𝑑v𝑑y′′𝑑vCΩ3|Kf2(y′′,v)|2|v|dy′′12ϵ𝑑v𝑑y′′CΩ31|v|dy′′12ϵ(3|k(v,w)|𝑑w)(3|k(v,w)||f2(y′′,w)|2𝑑w)𝑑v𝑑y′′CΩ33|k(v,w)||f2(y′′,w)|2|v|dy′′12ϵ𝑑v𝑑w𝑑y′′CΩ3|f2(y′′,w)|2dy′′12ϵ𝑑w𝑑y′′.\begin{split}I_{2}&\leq C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|k(v,v_{*})||Kf_{2}(y^{\prime\prime},v_{*})|^{2}}{|v_{*}||v|^{2-\epsilon}d_{y^{\prime\prime}}^{\frac{1}{2}-\epsilon}}\,dvdy^{\prime\prime}dv_{*}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|Kf_{2}(y^{\prime\prime},v_{*})|^{2}}{|v_{*}|d_{y^{\prime\prime}}^{\frac{1}{2}-\epsilon}}\,dv_{*}dy^{\prime\prime}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{1}{|v_{*}|d_{y^{\prime\prime}}^{\frac{1}{2}-\epsilon}}\left(\int\limits_{\mathbb{R}^{3}}|k(v_{*},w)|\,dw\right)\left(\int\limits_{\mathbb{R}^{3}}|k(v_{*},w)||f_{2}(y^{\prime\prime},w)|^{2}\,dw\right)\,dv_{*}dy^{\prime\prime}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\int\limits_{\mathbb{R}^{3}}\frac{|k(v_{*},w)||f_{2}(y^{\prime\prime},w)|^{2}}{|v_{*}|d_{y^{\prime\prime}}^{\frac{1}{2}-\epsilon}}\,dv_{*}dwdy^{\prime\prime}\\ &\leq C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|f_{2}(y^{\prime\prime},w)|^{2}}{d_{y^{\prime\prime}}^{\frac{1}{2}-\epsilon}}dwdy^{\prime\prime}.\end{split}

The last line above is (LABEL:eq:fdintegral) with f2f_{2} in place of f1f_{1}. In the same fashion, we can conclude that

(7.61) I2CΩ3|f2(y′′,w)|2dy′′12ϵ𝑑w𝑑y′′C3Ω|f1(z,w)|2𝑑z𝑑wCfL2(Ω×3)2.\begin{split}I_{2}\leq&C\,\int\limits_{\Omega}\int\limits_{\mathbb{R}^{3}}\frac{|f_{2}(y^{\prime\prime},w)|^{2}}{d_{y^{\prime\prime}}^{\frac{1}{2}-\epsilon}}dwdy^{\prime\prime}\\ \leq&C\,\int\limits_{\mathbb{R}^{3}}\int\limits_{\Omega}|f_{1}(z,w_{*})|^{2}\,dzdw_{*}\\ \leq&C\,\|f\|^{2}_{L^{2}(\Omega\times\mathbb{R}^{3})}.\\ \end{split}

This completes the proof. ∎

Remark 7.7.

From our calculation, we are not able to show that f4~Lv2(3;Hx12(3))\widetilde{f_{4}}\in L^{2}_{v}(\mathbb{R}^{3};H^{\frac{1}{2}}_{x}(\mathbb{R}^{3})). As a result, we can not further improve regularity by the same method as in Corollary 7.5.

Acknowledgements: The first author is supported in part by MOST grant 108-2628-M-002 -006 -MY4 and 106-2115-M-002 -011 -MY2. The third author is supported by NCTS and MOST grant 104-2628-M-002-007-MY3.

References

  • [1] Caflisch, R. E. (1980). The Boltzmann equation with a soft potential. Communications in Mathematical Physics, 74(1), 71-95.
  • [2] Cessenat, M. (1984). Théoremes de trace Lp pour des espaces de fonctions de la neutronique. Comptes rendus des séances de l’Académie des sciences. Série 1, Mathématique, 299(16), 831-834.
  • [3] Cessenat, M., & DAUTRAY, R. (1985). Théoremes de trace pour des espaces de fonctions de la neutronique. Comptes rendus des séances de l’Académie des sciences. Série 1, Mathématique, 300(3), 89-92.
  • [4] Chen, I. K., Liu, T. P., & Takata, S. (2014). Boundary singularity for thermal transpiration problem of the linearized Boltzmann equation. Archive for Rational Mechanics and Analysis, 212(2), 575-595.
  • [5] Chen, I. K., & Hsia, C. H. (2015). Singularity of macroscopic variables near boundary for gases with cutoff hard potential. SIAM Journal on Mathematical Analysis, 47(6), 4332-4349.
  • [6] Chen, I. K. (2018). Regularity of stationary solutions to the linearized Boltzmann equations. SIAM Journal on Mathematical Analysis, 50(1), 138-161.
  • [7] Chen, I. K., Hsia, C. H., & Kawagoe, D. (2019, May). Regularity for diffuse reflection boundary problem to the stationary linearized Boltzmann equation in a convex domain. In Annales de l’Institut Henri Poincaré C, Analyse non linéaire (Vol. 36, No. 3, pp. 745-782). Elsevier Masson.
  • [8] Chen, H., & Kim, C. (2020). Regularity of Stationary Boltzmann equation in Convex Domains. arXiv preprint arXiv:2006.09279.
  • [9] Chuang, P. H. (2019). Velocity Averaging Lemmas and Their Application to Boltzmann Equation (Mater’s thesis). Available from airiti Library. (U0001-2107201920343100) doi:10.6342/NTU201901751
  • [10] Desvillettes, L. (2003). About the use of the Fourier transform for the Boltzmann equation. Riv. Mat. Univ. Parma, 7(2), 1-99.
  • [11] Di Nezzaa, E., Palatuccia, G., & Valdinocia, E. (2012). Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math, 136(5), 512-573.
  • [12] DiPerna, R. J., & Lions, P. L. (1989). On the Cauchy problem for Boltzmann equations: global existence and weak stability. Annals of Mathematics, 321-366.
  • [13] Do Carmo, M. P. (2016). Differential geometry of curves and surfaces: revised and updated second edition. Courier Dover Publications.
  • [14] Esposito, R., Guo, Y., Kim, C., & Marra, R. (2013). Non-isothermal boundary in the Boltzmann theory and Fourier law. Communications in Mathematical Physics, 323(1), 177-239.
  • [15] Golse, F., Perthame, B., & Sentis, R. Un rsultat de compacité pour les équations de transport. CR Acad. Sci. Paris I Math. 301 (1985) 341, 344.
  • [16] Golse, F., Lions, P. L., Perthame, B., & Sentis, R. (1988). Regularity of the moments of the solution of a transport equation. Journal of functional analysis, 76(1), 110-125.
  • [17] Grad, H. Asymptotic theory of the Boltzmann equation. II, 1963 Rarefied Gas Dynamics (Proc. 3rd Internat. Sympos., Palais de l’UNESCO, Paris, 1962). Vol. I, 26-59.
  • [18] Guiraud, J. P. (1970). Probleme aux limites intérieur pour l’équation de Boltzmann linéaire. J. Mécanique, 9, 443-490.
  • [19] Guo, Y., Kim, C., Tonon, D., & Trescases, A. (2016). BV-regularity of the Boltzmann equation in non-convex domains. Archive for Rational Mechanics and Analysis, 220(3), 1045-1093.
  • [20] Guo, Y., Kim, C., Tonon, D., & Trescases, A. (2017). Regularity of the Boltzmann equation in convex domains. Inventiones mathematicae, 207(1), 115-290.
  • [21] Kawagoe, D. (2018). Regularity of solutions to the stationary transport equation with the incoming boundary data.
  • [22] Liu, T. P., & Yu, S. H. (2004). The Green’s function and large‐time behavior of solutions for the one‐dimensional Boltzmann equation. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 57(12), 1543-1608.