1 Introduction

An important and active research direction in mathematical physics/PDE is on the so-called Hilbert’s sixth problem [29] seeking a unified theory of the gas dynamics including different levels of descriptions from a mathematical standpoint by connecting the behavior of solutions to equations from kinetic theory to solutions of other systems that arise in formal limits, such as the N-body problem, the Euler equations, the Navier-Stokes equations, etc. In particular, the hydrodynamic limit of the Boltzmann equation has received a great deal of attention and enthusiasm in the mathematics and physics communities since the pioneering work [30] by Hilbert, which was the first example of his sixth problem. Remarkably, all the basic fluid equations of compressible, incompressible, inviscid, or viscous fluid dynamics can be derived from the Boltzmann equation of rarefied gas dynamics upon the choice of appropriate scalings in a small mean free path limit. Though formal derivations are rather well-understood, as far as mathematical justifications go, despite great progress over the decades (for example see [1,2,3, 13, 19, 23, 50] and the references therein), full understanding of the hydrodynamic limit incorporating important physical applications such as boundary effects or physical phenomena is still far from being complete. The goal of this paper is to make a rigorous connection between the Boltzmann equation and the incompressible Euler equations in the presence of the boundary by bypassing the inviscid limit of the incompressible Navier-Stokes equations.

The dimensionless Boltzmann equation with the Strouhal number and the Knudsen number takes the form of

(1.1)

Here the distribution function of the gas is denoted by \(F(t,x,v) \ge 0\) with the time variable \(t \in {\mathbb {R}}_+: = \{ t \ge 0\}\), the space variable \(x = (x_1,x_2,x_3)\in \Omega \subset {\mathbb {R}}^3\), and the velocity variable \(v = (v_1,v_2,v_3) \in {\mathbb {R}}^3\). The Boltzmann collision operator \(Q(\cdot , \cdot )\) of the hard sphere takes the form of

$$\begin{aligned} \begin{aligned} Q(F,G) =&\frac{1}{2} \int _{{\mathbb {R}}^3} \int _{{\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}} | \{ F(v^\prime ) G(v^\prime _*) + G(v^\prime ) F(v^\prime _*)\\&\qquad \qquad \qquad \qquad \qquad \qquad \quad - F(v ) G(v_* ) - G(v ) F(v_* ) \}\mathrm {d}{\mathfrak {u}} \mathrm {d}v_*, \end{aligned} \end{aligned}$$
(1.2)

where \(v^\prime := v- ((v-v_*) \cdot {\mathfrak {u}} ) {\mathfrak {u}} \) and \(v_*^\prime := v_*+ ((v-v_*) \cdot {\mathfrak {u}} ) {\mathfrak {u}} \). This operator satisfies the so-called collision invariance property: for any \(F(v)\text { and } G(v) \) decaying sufficiently fast as \(|v|\rightarrow 0\),

$$\begin{aligned} \int _{{\mathbb {R}}^3} Q(F,G) (v) \Big ( 1 , v , \frac{| {v } |^2-3}{\sqrt{6}} \Big ) \mathrm {d}v=(0,0,0), \end{aligned}$$
(1.3)

which represents the local conservation laws of mass, momentum and energy. The celebrated Boltzmann’s H-theorem reveals the entropy dissipation:

$$\begin{aligned} \int _{{\mathbb {R}}^3} Q(F,F) (v) \ln F(v) \mathrm {d}v \le 0, \end{aligned}$$
(1.4)

for any \(F(v)>0\) decaying sufficiently fast as \(|v|\rightarrow \infty \). An intrinsic equilibrium, satisfying \(Q(\cdot , \cdot )=0\), is given by a local Maxwellian associated with the density \(R>0\), the macroscopic velocity \(U\in {\mathbb {R}}^3\) and the temperature \(T>0\)

$$\begin{aligned} M_{R, U , T} (v): = \frac{R}{(2\pi T)^{ \frac{3}{2}}}\exp \left\{ - \frac{|v-U|^2}{2T}\right\} , \end{aligned}$$
(1.5)

which is known as the only configuration attaining the equality in (1.4).

In addition to the Strouhal number and Knudsen number we introduce the Mach number . By passing and to zero, one can formally derive PDEs of hydrodynamic variables for the fluctuations around the reference state (1, 0, 1), which are determined as

(1.6)

The famous Reynolds number appears as a ratio between the Knudsen number and Mach number through the von Karman relation:

(1.7)

For instance, the incompressible Navier-Stokes equations with , namely the viscosity of order one, can be derived by setting as \( \varepsilon \downarrow 0\). In this paper we are particularly interested in a scale of large Reynolds number as follows:

(1.8)

through which we will derive the incompressible Euler equations with the no-penetration boundary condition in the limit

$$\begin{aligned} \partial _t {u} _E + {u}_E \cdot \nabla _x {u}_E +\nabla _x p _E =0, \ \nabla _x \cdot {u} _E =0 \ \,&\text {in} \ \ \Omega , \end{aligned}$$
(1.9)
$$\begin{aligned} u _E\cdot n =0 \ \,&\text {on} \ \ \partial \Omega , \end{aligned}$$
(1.10)

with \(\partial _t \theta + u _E \cdot \nabla _x \theta =0\) and \( \nabla _x \theta (t,x) +\nabla _x \rho (t,x)= 0\). Here \(n=n(x)\) denotes an outward normal at x on the boundary \(\partial \Omega \).

For the sake of simplicity we set an initial datum \(\theta _{0}(x)= 0=\rho _{0}(x)\) so that

$$\begin{aligned} \theta (t,x) = 0=\rho (t,x) \text { for all } t \ge 0. \end{aligned}$$
(1.11)

In many important physical applications such as a turbulence theory, it would be relevant to take into account the physical boundary in the hydrodynamic limit. A boundary condition of the Boltzmann equation is determined by the interaction law of the gas with the boundary surface. One of the physical conditions is the so-called diffuse reflection boundary condition, which takes into account an instantaneous thermal equilibration of reflecting gas particle (see [9, 12]): for \( (x,v) \in \{ \partial \Omega \times {\mathbb {R}}^3: n(x) \cdot v<0\},\)

$$\begin{aligned} F (t,x,v) = c_\mu M_{1,0,1}(v) \int _{n(x) \cdot {\mathfrak {v}}>0} F (t,x,{\mathfrak {v}}) (n(x) \cdot {\mathfrak {v}}) \mathrm {d}{\mathfrak {v}} , \end{aligned}$$
(1.12)

where we have taken an isothermal boundary with a rescaled temperature 1 for the sake of simplicity. Here, the normalization constant \(c_\mu := 1/\big (\int _{n(x) \cdot {\mathfrak {v}}>0} M_{1,0,1} ({\mathfrak {v}}) (n(x) \cdot {\mathfrak {v}}) \mathrm {d}{\mathfrak {v}}\big )\) leads to the null flux condition \( \int _{{\mathbb {R}}^3} F(t,x,v) (n(x) \cdot v) \mathrm {d}v=0 \text { on }x \in \partial \Omega \). In particular, it is well-known that the diffuse boundary condition (1.12) is a kinetic boundary condition featuring a mismatch with the no-penetration boundary condition (1.10) of the the Euler flow under (1.8), without any small parameter with respect to or . One can readily see this by expanding F around a local Maxwellian \(M_{1,\varepsilon u_E, 1}(v)\) associated with a flow of the no-penetration boundary condition (1.10) directly. Unfortunately, this local Maxwellian does not honor the diffuse reflection boundary condition when a flow satisfies the no-penetration boundary condition (1.10). In fact a size of the boundary mismatch could be an order of the tangential component of the Euler flow \(u_E\) at the boundary. Therefore a uniform bound to verify the limit (1.6) in a scale of large Reynolds number (1.8) is not expected even at the formal level. This poses a major obstacle in the Euler limit from the Boltzmann equation with the diffuse reflection boundary. It is worth noting that such a mismatch does not appear at least at the formal level when the specular reflection boundary condition is imposed: \(F(t,x,v)= F(t,x,R_x v)\) on \(x \in \partial \Omega \) where \(R_x v= v- 2n(x) (n(x) \cdot v)\); while the mismatch can possess a small factor for the so-called Maxwell boundary condition, a convex combination of the specular reflection and the diffuse reflection boundary conditions, by choosing the coefficient for diffuse reflection known as the accommodation constant to vanish as .

Remarkably, an analogous, better-known boundary mismatch phenomenon exists in the realm of mathematical fluid dynamics, specifically in the inviscid limit problem of the Navier-Stokes equations that addresses the validity of the Euler solutions as the leading order approximation of the Navier-Stokes solutions in the vanishing viscosity limit. The inviscid limit for the no-slip boundary condition features a boundary mismatch between two different boundary conditions for the Navier-Stokes and Euler flows. In fact, whether the solution to the Navier-Stokes equations with a \( \kappa \eta _0 \)-viscosity (a physical constant \(\eta _0\) can be computed explicitly from the Boltzmann theory as in (1.37)) satisfying the no-slip boundary condition

$$\begin{aligned} \partial _t u+ u \cdot \nabla _x u - \kappa \eta _0 \Delta u + \nabla _x p&=0 \ \ \text {in} \ \Omega , \end{aligned}$$
(1.13)
$$\begin{aligned} \nabla _x \cdot u&=0 \ \ \text {in} \ \Omega , \end{aligned}$$
(1.14)
$$\begin{aligned} \ \ u&=0 \ \ \text {on} \ \partial \Omega , \end{aligned}$$
(1.15)

converges to the solution of the Euler equations satisfying the no-penetration boundary condition (1.9)-(1.10) in is an outstanding problem, which is arguably the most relevant and challenging because of the mismatch of two boundary conditions between (1.15) and (1.10) resulting in the formation of boundary layers such as Prandtl layer and unbounded vorticity near the boundary. While the verification of the inviscid limit is still largely open, it holds under certain symmetry assumption on the domain and data or under the flat boundary and strong regularity such as analyticity at least near the boundary [45]. A classical way to tackle the inviscid limit problem is to study the Prandtl expansion [44, 48, 49]: \(u(t,x_1,x_2,x_3)= u_E(t,x_1,x_2,x_3) + u_P (t,x_1,x_2,\frac{x_3}{\sqrt{\kappa }})+ O(\sqrt{\kappa })\). Recently, different frameworks that avoid the boundary layer expansion have become available [38, 47, 54].

The incompressible Euler limit from the Boltzmann equation turns out to be intimately tied to the inviscid limit of the incompressible Navier-Stokes equations, which accounts for the similarity of two boundary mismatches. A beautiful connection stems from the Navier-Stokes solutions of (1.13)-(1.15) in large Reynolds numbers: at least formally, not only they are approximated by the Euler equations (1.9)-(1.10) but also they approximate the Boltzmann equation (1.1) under (1.8) with (1.12), in fact better than the Euler equations (1.9)-(1.10) at each Mach number \(\varepsilon >0\), because the Navier-Stokes equations contain a high order correction term \(\kappa \eta _0 \Delta u\) that captures the dissipative nature of the Boltzmann collision operator (as we will see in Section 1.1). And importantly, a local Maxwellian \(M_{1,\varepsilon u, 1}(v)\) associated with u satisfying the no-slip boundary condition (1.15), satisfies the diffuse reflection boundary condition (1.12) without singular terms. In other words, the Navier-Stokes solutions are compatible with the diffuse reflection boundary condition. Therefore, under the scale (1.8) the Navier-Stokes solution of (1.13)-(1.15) stands in between the Boltzmann solution of (1.1), (1.12) and the Euler solution (1.9)-(1.10).

In this paper, inspired by these observations, we propose to study the Euler limit from the Boltzmann equation through the Navier-Stokes solutions that hold both features of the Euler and the Boltzmann under (1.8) at each Mach number \(\varepsilon >0\). To this end, we expand the Boltzmann solution F around a local Maxwellian associated with a Navier-Stokes flow u to (1.13)-(1.15):

$$\begin{aligned} \mu (v) : = M_{1 , \varepsilon u , 1 }(v), \end{aligned}$$
(1.16)

as

$$\begin{aligned} F = \mu + \varepsilon ^2 f_2 \sqrt{\mu } + \varepsilon ^{3/2} f_R\sqrt{\mu } , \end{aligned}$$
(1.17)

and analyze (1.17) via a new Hilbert expansion presented in Section 1.1. Although the notations \(F^\varepsilon \) and \(f^\varepsilon \) may be more precise for the equation depending on \(\varepsilon \), we will abuse the notations by dropping the superscript \(\varepsilon \) for the sake of simplicity. The next order correction \(f_2\) can be entirely determined by the Navier-Stokes flow and it turns out that its contribution is always smaller than \(f_R\)’s one in our choice of \(\varepsilon \) and \(\kappa \). A choice of the range of the Mach number with respect to the Reynolds number: in \(\varepsilon \downarrow 0\) plays an important role in our analysis and the formal expansion. We will discuss the relation and its role in Section 1.2. With such a choice of the scale, uniform-in-\(\varepsilon \) estimates of the Boltzmann remainder \(f_R\) are achieved by a novel quantitative \(L^p\)-\(L^\infty \) estimate in a setting of the local Maxwellian of the Navier-Stokes approximation (1.16), along with the commutator estimates and the integrability gain of the hydrodynamic part in various spaces.

In order to establish the Euler limit by using the Navier-Stokes solutions of (1.13)-(1.15) as a reference state as \( \varepsilon \downarrow 0\) in a scale of large Reynolds number (1.8), it is imperative to show the uniform-in-\(\kappa \) convergence of the Navier-Stokes solutions to the Euler solutions of (1.9)-(1.10), where the inviscid limit comes into play. In this paper, we take the spatial domain to be the upper-half space with periodic boundary conditions in the horizontal components and analytic data for the Navier-Stokes solutions of (1.13)-(1.15) and obtain uniform-in-\(\kappa \) estimates built upon a recent development on the inviscid limit problem in the half-space based on the Green’s function approach using the boundary vorticity formulation [38, 44, 47, 54].

Our main result concerns a rigorous justification of the passage from the solutions to the dimensionless Boltzmann equation (1.1) of the scale (1.8) with the diffuse reflection boundary condition (1.12) to the solution of the incompressible Euler equation (1.9) with the no-penetration boundary condition (1.10), without introducing any boundary expansion of the Boltzmann equation:

Theorem 1

(Informal statement) We consider a half space in 3D

$$\begin{aligned} \Omega := {\mathbb {T}}^2 \times {\mathbb {R}}_+ \ni (x_1, x_2, x_3), \ \ \text {where } {\mathbb {T}} \text { is a periodic interval of } (-\pi , \pi ). \end{aligned}$$
(1.18)

For some choice of \(\varepsilon \) and \(\kappa (\varepsilon )\), there exists a large set of initial data \(u_{in}\), \(f_{2,in}\) and \(f_{R,in}\) such that a unique solution F(txv) of the form (1.17) to (1.1) and (1.12) with (1.8) exists on [0, T] for some \(T>0\) and satisfies

$$\begin{aligned} \sup _{0 \le t \le T}\left\| \frac{F (t,x,v)- M_{1, \varepsilon u , 1} }{\varepsilon \sqrt{M_{1,\varepsilon u,1}}} \right\| _{L^2(\Omega \times {\mathbb {R}}^3)} \longrightarrow 0 \ \ \text {as} \ \ \varepsilon \downarrow 0, \end{aligned}$$

and

$$\begin{aligned} \sup _{0 \le t \le T}\left\| \frac{F (t,x,v)- M_{1, \varepsilon u_E , 1} }{\varepsilon (1+ |v|)^2\sqrt{M_{1,0,1} } } \right\| _{L^2(\Omega \times {\mathbb {R}}^3)} \longrightarrow 0 \ \ \text {as} \ \ \varepsilon \downarrow 0, \end{aligned}$$

while u and \(u_E\) denote solutions of the Navier-Stokes (1.13)-(1.15) and Euler equations (1.9)-(1.10), respectively.

The precise statement of Theorem 1 is given in Theorem 4 and Corollary 5 in Section 2.3.

Remark 1

To the best of our knowledge our result of this paper appears to be the first rigorous incompressible Euler limit result from the Boltzmann solutions with the sole diffuse reflection (therefore the accommodation constant \(\sim \)1) in the boundary condition! Moreover, our framework captures the inviscid limit of mathematical fluid dynamics from the Boltzmann theory.

Remark 2

Another natural choice of the scale in the study of the Euler limit might be \(\varepsilon ^{{\mathfrak {q}}}=\kappa \) with an integer \({\mathfrak {q}}\ge 1\). Then the second correction \(\frac{1}{\kappa }Lf_2\) is shifted to the next hierarchy (see (1.27)) and as a consequence the Euler equations become the leading approximation with loss of \(\kappa \eta _ 0 \Delta u\). Without the boundary, a higher order expansion \(F=\mu _E+ [\varepsilon f_1 + \varepsilon ^2 f_2 + \varepsilon ^3 f_3 + \cdots + \varepsilon ^r f_R] \sqrt{\mu _E}\) for \(\mu _E=M_{1 , \varepsilon u_E , 1 }\) has been established in [10, 56]. In the presence of the boundary, on the other hand, such an expansion features a boundary mismatch. The usual approach is then drawn on a boundary layer expansion, correcting an interior Hilbert-like expansion at the boundary to satisfy the boundary conditions (for example, see [27, 55]). Our approach is based on an interior expansion up to the second correction \(f_2\) that avoids the boundary layer expansion under our choice of scale \(\varepsilon \ll \kappa \) (see (2.11)).

Before discussing the essence of the methodology and novelty of our result, we shall briefly overview some relevant literatures on the hydrodynamic limit of the Boltzmann equation. One of the first mathematical studies of the limits at the formal level may go back to a work [30] of Hilbert, in which he introduced so-called the Hilbert expansion. Based on the truncated Hilbert expansion rigorous justifications of fluid limits have been shown as long as the solutions of corresponding fluids are bounded in some suitable spaces, for example, in the compressible fluid limits in [7, 53], incompressible fluid limits in [5, 10, 23], diffusive limits from the Vlasov-Maxwell-Boltzmann system in [32], and relativistic fluid limits in [52]. All the derivations mentioned above did not take into account the boundary, while one of the main obstacles to study the Boltzmann solutions with the boundary is its boundary singularity (see [20, 21, 35]). In [22], an \(L^p\)-\(L^\infty \) framework has been developed to construct a unique global solution of the Boltzmann equation with physical boundary conditions. Such a framework has been developed successfully in various problems of the Boltzmann theory (for example [8, 14, 24,25,26,27, 36, 37, 55]). In particular, in [12, 13], the authors have constructed a solution of the Boltzmann equation satisfying the diffuse reflection boundary condition and proved the validity of the hydrodynamic limit toward the incompressible Navier-Stokes-Fourier system in both steady and unsteady settings, based on a novel \(L^6\)-bound of the hydrodynamic part.

Rigorous passage from the renormalized solutions of [11] ([46] with the physical boundary) of the Boltzmann equation toward (weak) solutions of fluid equations has been also extensively explored (see [17, 33, 50] for the references in this direction). In particular, the program of the incompressible Navier-Stokes limit to the Leray-Hopf weak solutions has been developed successfully in [2, 3, 19, 40, 41] without the physical boundary and with the boundary in [33, 42]. As for the incompressible Euler limit in terms of the entropy production, based on the relative entropy method, a dissipative solution of the incompressible Euler equations in [39] has been studied in [40, 41, 51] without the boundary. Notably the results have been extended to the domain with the boundary for the specular reflection boundary condition in [50], and for the Maxwell boundary condition in [4], assuming to set that the accommodation constant (a factor of diffuse reflection) vanishes as \(\varepsilon \downarrow 0\).

For the rest of this section, we present the strategy and key ideas developed in the proof of our result starting with a new (formal) Hilbert expansion followed by the control of the Boltzmann remainder \(f_R\) and higher regularity of Navier-Stokes flows, for the rigorous justification of the formal expansion.

1.1 Hilbert Expansion in a Scale of Large Reynolds Number

Through a new formal Hilbert-type expansion of Boltzmann equation with the diffuse reflection boundary condition we aim to capture the Navier-Stokes equations of vanishing viscosity proportional to and satisfying the no-slip boundary condition.

It is worth pointing out that although more convenient choice of an expansion of F is seemingly the one around the global Maxwellian \(\mu _0:= M_{1,0,1}\) such as \(F=\mu _0 + \varepsilon (u \cdot v) \mu _0 + \varepsilon ^2 {\tilde{f}}_2 \sqrt{\mu _0} + \varepsilon \delta {\tilde{f}}_R \sqrt{\mu _0}\), unfortunately this choice will produce, in the Hilbert expansion (1.26)-(1.30), an unbounded term \(\frac{2}{\kappa \varepsilon } \frac{1}{\sqrt{\mu _0 }}Q(u \cdot v \mu _0 ,f_R \sqrt{\mu _0 })\) even compared to the strongest control in hand, namely a dissipation term (see (1.31))! To achieve a sharper estimate, which provides weaker restriction on \(\kappa \) and \(\varepsilon \), and hence weaker restriction on the initial data, we work on an expansion around the local Maxwellian \(\mu \).

It is conceptually convenient in our analysis to introduce an auxiliary parameter \(\delta =\delta (\varepsilon )\downarrow 0\) as \(\varepsilon \downarrow 0\), which indicates a size of the fluctuation \((F-\mu )/ {\varepsilon }\):

$$\begin{aligned} F = \mu + \varepsilon ^2 f_2 \sqrt{\mu } + \varepsilon \delta f_R\sqrt{\mu } . \end{aligned}$$
(1.19)

In (1.17) we have chosen \(\delta =\sqrt{\varepsilon }\) and in Section 2.3 we will have the same choice such as (2.11), however in Section 2.1, Section 3 and Section 4, \(\delta \) will be regarded as a free parameter and will be chosen at the last step of closing our argument (as (2.11)!).

Interior Expansion. We investigate an expansion (1.19) of the Boltzmann equation (1.1) at the local Maxwellian \(\mu \) in (1.16). Let

$$\begin{aligned} L f = \frac{-2}{\sqrt{\mu }} Q(\mu , \sqrt{\mu }f) ,\ \ \Gamma (f,g)= \frac{1}{ \sqrt{\mu }} Q(\sqrt{\mu }f, \sqrt{\mu }g) . \end{aligned}$$
(1.20)

The operators L and \(\Gamma \) can be read as

$$\begin{aligned}&Lf(v)= \nu f(v) -Kf(v)= \nu (v) f(v)- \int _{{\mathbb {R}}^3} {\mathbf {k}} (v , v_* ) f(v_*) \mathrm {d}v_*, \end{aligned}$$
(1.21)
$$\begin{aligned}&\Gamma (f,g) (t,v) = \Gamma _+ (f,g) (t,v) - \Gamma _- (f,g) (t,v) \nonumber \\&\quad = \iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}}| \sqrt{ \mu (v_* )} \big ( f(t,v ^\prime ) g(t,v_*^\prime ) + g(t,v ^\prime ) f(t,v_*^\prime )\big )\mathrm {d}{\mathfrak {u}} \mathrm {d}v_* \nonumber \\&\qquad - \iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}}| \sqrt{ \mu (v_* )} \big (f(t,v ) g(t,v_* ) +g(t,v ) f(t,v_* )\big ) \mathrm {d}{\mathfrak {u}} \mathrm {d}v_*, \end{aligned}$$
(1.22)

where the precise form of \({\mathbf {k}}\) is delayed to be presented in (3.19). We will demonstrate basic properties of operators L and \(\Gamma \) in Section 3.1. From (1.3) the null space of L, denoted by \({\mathcal {N}}\), is a subspace of \(L^2({\mathbb {R}}^3)\) spanned by orthonormal bases \(\{\varphi _i \sqrt{\mu }\}_{i=0}^4\) with

$$\begin{aligned} \begin{aligned}&\varphi _0 := 1 , \ \ \ \varphi _i: = {v_i -\varepsilon u_i } \ \ \text {for} \ i=1,2,3 , \ \ \ \varphi _4: = ( | {v-\varepsilon u} |^2-3 )/{\sqrt{6}}. \end{aligned} \end{aligned}$$
(1.23)

We define a hydrodynamic projection \({\mathbf {P}}\) as an \(L_v^2\)-projection on \({\mathcal {N}}\) such as

$$\begin{aligned} \begin{aligned} {\mathbf {P}} g:=&\sum ( {P}_j g) \varphi _j \sqrt{\mu }, \ \ {P}_j g:= \langle g ,\varphi _j \sqrt{\mu } \rangle , \ \text {and} \ \ \\ P g:=&(P_0 g, P_1 g, P_2 g, P_3 g, P_4 g ), \end{aligned} \end{aligned}$$
(1.24)

where \(\langle \cdot , \cdot \rangle \) stands for an \(L^2_v\)-inner product. It is well-known that the operators enjoy \({\mathbf {P}}L=L {\mathbf {P}}={\mathbf {P}} \Gamma =0\). Importantly the linear operator L enjoys a coercivity away from the kernel \({\mathcal {N}}\): for \(\nu (v)\ge 0\) defined in (1.21)

$$\begin{aligned} \langle Lf, f\rangle \ge \sigma _0 \Vert \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) f \Vert _{L^2 ({\mathbb {R}}^3)}^2 \ \ \text {for some } \sigma _0>0. \end{aligned}$$
(1.25)

Now we plug the expansion (1.19) into the rescaled equation (1.1) with the scale (1.8). It turns out that by relating \(f_2\) with the flow and locating it carefully in the hierarchy we can exhibit the dissipative nature of the Boltzmann collision operator at the leading order of the fluid approximation. In particular we locate \((v-\varepsilon u) \cdot \nabla _x ({\mathbf {I}} - {\mathbf {P}})f_2\) in \(\frac{1}{\delta }\)-order hierarchy to capture \(\kappa \)-order viscosity in the fluid equation (1.13):

$$\begin{aligned}&\partial _t f_R + \frac{1}{\varepsilon } v\cdot \nabla _x f_R+ \frac{1}{ \varepsilon ^2\kappa } Lf_R + \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_R \end{aligned}$$
(1.26)
$$\begin{aligned}&\quad =- \frac{1}{\varepsilon \delta } \Big \{ \frac{ \varepsilon ^{-1} (v- \varepsilon u )\cdot \nabla _x \mu }{ \sqrt{\mu }} + \frac{ 1}{ \kappa } L f_2\Big \} \end{aligned}$$
(1.27)
$$\begin{aligned}&\qquad - \frac{1}{\delta } \Big \{ \frac{\varepsilon ^{-1} \partial _t \mu }{\sqrt{\mu }} + \frac{ \varepsilon ^{-1} u \cdot \nabla _x \mu }{ \sqrt{\mu }} + ( v- \varepsilon u)\cdot \nabla _x f_2 \Big \} \end{aligned}$$
(1.28)
$$\begin{aligned}&\qquad - \frac{\varepsilon }{\delta } \Big \{ \partial _t f_2 + u \cdot \nabla _x f_2 + \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x ) \sqrt{\mu }}{\sqrt{\mu }}f_2 \Big \} \end{aligned}$$
(1.29)
$$\begin{aligned}&\qquad + \frac{2}{\kappa } \Gamma ({f_2}, f_R) + \frac{\varepsilon }{\delta \kappa } \Gamma ({f_2}, f_2) + \frac{ \delta }{\varepsilon \kappa }\Gamma (f_R,f_R) . \end{aligned}$$
(1.30)

We can readily see an \(L^2\)-energy structure of \(f_R\) with a strong dissipation

$$\begin{aligned} \iint _{\Omega \times {\mathbb {R}}^3} \frac{1}{\varepsilon ^{2} \kappa } L f_R f_R \mathrm {d}v \mathrm {d}x > rsim \Vert \varepsilon ^{-1} \kappa ^{-1/2} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2 (\Omega \times {\mathbb {R}}^3)}^2, \end{aligned}$$
(1.31)

which inherits its lower bound from the coercivity (1.25).

Let us first consider an \(\frac{1}{\varepsilon \delta }\)-hierarchy (1.27). For any non-vanishing term of (1.27) would cause unpleasant unboundedness, we make the term vanish entirely by solving an equation (1.27)\(=0\). By the Fredholm alternative, an inverse map

$$\begin{aligned} L^{-1}: {\mathcal {N}}^{\perp } \rightarrow {\mathcal {N}}^{\perp }, \ \text { where } \ {\mathcal {N}}^\perp \ \text {stands an }L^2_v\text {-orthogonal complement of } {\mathcal {N}},\qquad \quad \end{aligned}$$
(1.32)

is well-defined and hence the solvability condition is given by

$$\begin{aligned} \frac{ \varepsilon ^{-1} (v- \varepsilon u )\cdot \nabla _x \mu }{ \sqrt{\mu }} =\sum _{\ell , m=1}^3 \partial _\ell u_m\varphi _\ell \varphi _m \sqrt{\mu } \in {\mathcal {N}}^\perp . \end{aligned}$$
(1.33)

This condition indeed implies the incompressible condition (1.14).

Once (1.14) holds, we have \(\sum _{\ell , m=1}^3 \partial _\ell u_m\varphi _\ell \varphi _m \sqrt{\mu }= \sum _{\ell , m=1}^3 \partial _\ell u_m (\varphi _\ell \varphi _m - \frac{|v-\varepsilon u|^2}{3} \delta _{\ell m} ) \sqrt{\mu }\). Now we solve (1.27)\(=0\) by setting

$$\begin{aligned} ({\mathbf {I}}-{\mathbf {P}}) f_2= & {} -\kappa \sum _{\ell ,m=1}^3 A_{\ell m} \partial _\ell u_m \ \ \text {with} \nonumber \\ A_{\ell m}:= & {} L^{-1} \Big ( \varphi _\ell \varphi _m \sqrt{\mu } - \frac{|v- \varepsilon u |^2}{3} \delta _{\ell m}\sqrt{\mu } \Big ). \end{aligned}$$
(1.34)

Then we move to an \(\frac{1}{\delta }\)-hierarchy (1.28). The hydrodynamic part of (1.28), unless it vanishes, would induce an unbounded term again. We expand \(\delta \times \) (1.28), using (1.16) and (1.34), as

$$\begin{aligned} \begin{aligned}&- (v- \varepsilon u) \cdot (\partial _t u + u \cdot \nabla _x u ) \sqrt{\mu } + (v- \varepsilon u) \cdot \nabla _x {\mathbf {P}} f_2\\&\quad + \kappa (v- \varepsilon u) \cdot \nabla _x \Big ( \sum _{\ell ,m=1}^3 A_{\ell m} \partial _\ell u_m\Big ) . \end{aligned} \end{aligned}$$
(1.35)

The leading order term of the last term in (1.35) contributes the following to the hydrodynamic part of (1.35) as

$$\begin{aligned} \begin{aligned}&\kappa \sum _{\ell ,m,k=1}^3 \big \langle \varphi _i \varphi _k \sqrt{\mu }, A_{\ell m} \big \rangle \partial _k \partial _\ell u_{ m} \\&\quad = \kappa \sum _{\ell ,m,k=1}^3 \Big \langle \big ( \varphi _i \varphi _k -\frac{|v-\varepsilon u|^2}{3} \delta _{i k} \big )\sqrt{\mu } , A_{\ell m} \Big \rangle \partial _k \partial _\ell u_{ m}\\&\quad = \kappa \sum _{\ell , m, k=1}^3 \left\langle L A_{i k} , A_{\ell m} \right\rangle \partial _{k} \partial _\ell u_{ m} , \end{aligned} \end{aligned}$$
(1.36)

where we have used the fact \(A_{\ell m } \in {\mathcal {N}}^\perp \) and \(\frac{|v-\varepsilon u|^2}{3} \sqrt{\mu } \in {\mathcal {N}}\) at the first step and the definition of \(A_{ik}\) at the last step. It is well-known (e.g. Lemma 4.4 in [3]) that for some constant \(\eta _0 >0\)

$$\begin{aligned} \langle LA_{i k }, A_{\ell m}\rangle = \eta _0 ( \delta _{\ell k} \delta _{mi} + \delta _{\ell i} \delta _{mk} ) - \frac{2}{3} \eta _0 \delta _{\ell m} \delta _{ik}. \end{aligned}$$
(1.37)

Therefore we deduce that (1.36) vanishes for \(i=0,4,\) and the \(\kappa \eta _0\)-viscosity term in (1.13) can be captured:

$$\begin{aligned} \begin{aligned} (1.36)&= \kappa \eta _0 \sum _{\ell ,m,k} \{ ( \delta _{\ell k} \delta _{mi} + \delta _{\ell i} \delta _{mk} ) - \frac{2}{3} \delta _{\ell m} \delta _{ik}\} \partial _{k} \partial _\ell u_{ m} \\&=\kappa \eta _0 \{\Delta u_{i} - \partial _i \nabla \cdot u - \frac{2}{3} \partial _i \nabla \cdot u \}= \kappa \eta _0\Delta u_{i} \ \ \text {for} \ i =1,2,3. \end{aligned} \end{aligned}$$
(1.38)

Here we have used the incompressible condition (1.14) at the last step. On the other hand, a leading order term of the hydrodynamic part of \((v- \varepsilon u) \cdot \nabla _x {\mathbf {P}} f_2\) contributes to the pressure term of (1.13) by choosing a special form of \({\mathbf {P}}f_2\) as in (3.1). Therefore the whole leading order terms of the hydrodynamic part in (1.28) do vanish by solving the Navier-Stokes equations (1.13) and (1.14)! For the sake of brevity we refer to Section 3 for the full expansion of (1.26)-(1.30).

Boundary Conditions. Now we consider a boundary condition of \(f_R\). Noticeably the local Maxwellian \(\mu \) becomes \(M_{1,0,1}\) on the boundary from the no-slip boundary condition (1.15), and hence \(\mu \) satisfies the diffuse reflection boundary condition (1.12). For the detailed study of the boundary condition of \(f_R\) we introduce the incoming and outgoing boundaries

$$\begin{aligned} \gamma _\pm := \{(x,v)\in \partial \Omega \times {\mathbb {R}}^3: n(x) \cdot v > rless 0 \}. \end{aligned}$$

Since \(\mu \) satisfies the diffuse reflection boundary condition (1.12) with a constant wall temperature \( =1\), by plugging (1.19) into the boundary condition, we arrive at

$$\begin{aligned} (\varepsilon ^2 f_2 + \delta \varepsilon f_R )|_{\gamma _-}= c_\mu \sqrt{\mu (v) } \int _{n(x) \cdot {\mathfrak {v}}>0} ( \varepsilon ^2 f_2 + \delta \varepsilon f_R ) \sqrt{\mu ({\mathfrak {v}})}(n(x) \cdot {\mathfrak {v}}) \mathrm {d}{\mathfrak {v}} . \end{aligned}$$

Letting \(P_{\gamma _+}\) be an \(L^2 (\{ v : n(x) \cdot v>0 \})\)-projection of \(\sqrt{c_\mu \mu }\), we derive that

$$\begin{aligned} \begin{aligned} f_R(t,x,v)|_{\gamma _-}&= P_{\gamma _+} f_R(t,x,v)- \frac{\varepsilon }{\delta } (1- P_{\gamma _+}) f_2(t,x,v)\\&:= \sqrt{c_\mu \mu (v)} \int _{n(x) \cdot {\mathfrak {v}}>0} f_R (t,x,{\mathfrak {v}})\sqrt{c_\mu \mu ({\mathfrak {v}})} (n(x) \cdot {\mathfrak {v}}) \mathrm {d}{\mathfrak {v}}\\&\quad \quad - \frac{\varepsilon }{\delta } (1- P_{\gamma _+}) f_2(t,x,v). \end{aligned} \end{aligned}$$
(1.39)

Note that \(\int _{n(x) \cdot v>0} c_\mu \mu (v) (n(x) \cdot v) \mathrm {d}v=1\).

On the other hand, we emphasize that, with the no-penetrate boundary condition of (1.10), the associated local Maxwellian \(M_{1,\varepsilon u_E,1}\) does not satisfy the diffuse reflection boundary condition in general. Therefore the Boltzmann remainder \(f_R\) would have a singularity of an order of \(1/\sqrt{\varepsilon }\) in (1.17).

1.2 Uniform Controls of the Boltzmann Remainder \(f_R\)

For a rigorous justification of the Hilbert expansion (1.19), the major task is to establish uniform-in-\(\varepsilon \) estimates of the Boltzmann remainder \(f_R\) in \(L^2\). The equation of the Boltzmann remainder \(f_R\) in (1.26)-(1.30) with the boundary condition (1.39) features a discrepancy between the behavior of the hydrodynamic part \({\mathbf {P}}f_R\) and pure kinetic part \(({\mathbf {I}}-{\mathbf {P}})f_R\): schematically an \(L^2\)-energy estimate reads

$$\begin{aligned}&\frac{d}{dt} \Vert f_R(t) \Vert _{L^2}^2 + \Vert \varepsilon ^{-1} \kappa ^{-1/2} ({\mathbf {I}}-{\mathbf {P}})f_R\Vert _{L^2}^2 \sim \Vert \nabla _x u\Vert _{L^\infty } \Vert {\mathbf {P}} f_R \Vert _{L^2}^2\\&\quad + \iint _{\Omega \times {\mathbb {R}}^3} \frac{\delta }{\varepsilon \kappa } \Gamma ( {\mathbf {P}}f_R, {\mathbf {P}}f_R) ({\mathbf {I}}-{\mathbf {P}})f_R . \end{aligned}$$

A key difficulty arises from a growth of the hydrodynamic part at least as \( e^{\Vert \nabla _x u\Vert _{L^\infty } } \) which might behave as an exponential of the reciprocal of some power of the viscosity \(\kappa \) due to the unbounded vorticity formed near the boundary, while such strong singularity of the hydrodynamic part enters the nonlinear estimate in turn. In fact such trilinear estimate can be effectively handled only by a point-wise bound of the solutions. Unfortunately as the physical boundary conditions create singularities in general ([35]), the high Sobolev estimates would not be possible. In this paper we develop a quantitative \(L^p\)-\(L^\infty \) estimate solely in the setting of the local Maxwellian associated with the Navier-Stokes flow, in the presence of the diffuse reflection boundary.

Thanks to a strong control of the dissipation from the spectral gap of (1.25), the nonlinear term can be bounded as

$$\begin{aligned} \delta \kappa ^{-\frac{1}{2}} \Vert {P} f_R \Vert _{L^\infty _tL^6_{x }} \Vert {P} f_R \Vert _{L^2_tL^3_{x }} \Vert \varepsilon ^{-1} \kappa ^{-\frac{1}{2}}\sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} . \end{aligned}$$
(1.40)

Notably an integrability gain of the hydrodynamic part \(Pf_R\) should play a role; however a classical velocity average lemma \(Pf_R \in H^{1/2}_x\subset L^3_x\) fails to fulfill the need in 3D. We achieve such a higher integrability by developing a recent \(L^6\)-bound of hydrodynamic part of [13] in the setting of the local Maxwellian on the scale of large Reynolds number. We utilize the micro-macro decomposition and the equation to control \(\kappa ^{1/2}v\cdot \nabla _x {\mathbf {P}} f_R\) mainly by \(\frac{1}{\varepsilon \kappa ^{1/2}} L({\mathbf {I}} - {\mathbf {P}} )f_R\) and \( \varepsilon \kappa ^{1/2}\partial _t f_R\). We invert the operator \(v\cdot \nabla _x {\mathbf {P}}\), employing a recent test function method of [12] in the local Maxwellian setting, to establish a crucial \(L^6\)-bound of the hydrodynamic part, which is controlled by the dissipation plus the a priori \(L^2\)-bound of \(\partial _t f_R\):

$$\begin{aligned} \Vert \kappa ^{1/2} Pf_R(t) \Vert _{L^6_x} \lesssim \Vert \varepsilon ^{-1} \kappa ^{-1/2} ({\mathbf {I}} - {\mathbf {P}}) f_R (t) \Vert _{L^{2}_{x,v}} + \varepsilon \kappa ^{1/2} \Vert \partial _t f_R(t) \Vert _{L^2_{x,v}} + l.o.t.\nonumber \\ \end{aligned}$$
(1.41)

In other words we can achieve the \(L^6\)-estimate of (1.41) as “one spatial derivative gain” through the dissipation provided a temporal derivative being controlled, while the temporal derivative preserves the boundary conditions. It is a critical point in which a temporal derivative gets involved in our analysis of Boltzmann and fluids as well!

New difficulties arise as commutator estimates of \(\frac{1}{\varepsilon ^2 \kappa } \{\partial _t Lf_R- L \partial _t f_R\}\) induce singularities even at the linear level, as well as \(\partial _t(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu } f_R/\sqrt{\mu }\) and the source terms in the equation of \(\partial _t f_R\) possess higher temporal derivatives of the fluid with an initial layer. In fact after a careful analysis we realize such singular terms amount to

$$\begin{aligned} \frac{1}{\kappa ^{{\mathfrak {P}}}} \int ^t_0\Vert Pf_R(s)\Vert _{ L^2_x}^2 \mathrm {d}s , \end{aligned}$$

while \({\mathfrak {P}}\) depends on the singularity of derivatives of the Navier-Stokes flow in large Reynolds numbers.

We establish a unified \(L^\infty \)-estimate in the local Maxwellian setting, devising a special weight function \({\mathfrak {w}}_{\varrho , \ss }(x,v)\) in order to control an extra growth in |v| from \((\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu } f_R/\sqrt{\mu }\) and its temporal derivative. We control \(f_R\) in \(L^\infty _t L^\infty _x\) by the hydrodynamic part \(Pf_R\) in \(L^6\) and the dissipation, studying the particle-trajectory bouncing against the diffuse reflection boundary and geometric change of variables related to the bouncing trajectories. The temporal derivative \(\partial _t f_R\) needs some special attention since the source term of the equation of \(\partial _t f_R\) possesses \(\nabla _x \partial _t^2 u\), which turns out to have an initial-boundary layer. For that we measure \(\partial _t f_R\) using a different time-space norm, namely a weighted \(L^2_tL^\infty _x\), and control it by the hydrodynamic part of \(\partial _t f_R\) in \(L^2_tL^3_x\) (with more singular factor-in-\(\varepsilon \) than the counterpart for \(f_R\)) and the dissipation. Although our estimate of \(\partial _tf_R\) is singular than \(f_R\) due to our choice of different spaces, we are able to balance such extra singularity by the strong dissipation and careful trilinear estimates.

We establish \(L^2_tL^3_x-\)controls for \(Pf_R\) and \(P\partial _t f_R\) via the trajectory rather than the classical average lemma. In fact a direct application of such average lemma has some subtle issue since the source terms of \(f_R\) and \(\partial _t f_R\) equations are known to be bounded within a finite time interval only, while the \(L^2_tL^3_x\)-control enters the nonlinear estimates. In fact it is not clear whether our iteration of estimates would guarantee a nonempty finite time interval of validity. Instead we utilize the Duhamel formula along the trajectories and an extension of solutions in specially designed domains, and employ the \(TT^*\)-method developed in [14, 28, 31]. As a result we achieve \(L^2_tL^P_x\) estimates for \(f_R\) and \(\partial _t f_R\) uniformly for all \(p<3\), which gives us a sufficient bound in \(L^2_tL^3_x\) by interpolating with our \(L^\infty \)-estimates.

Finally upon combining all the estimates above together we are able to bound an energy by the Gronwall’s inequality. The resulting bound is not uniform but growing exponentially as \(e^{1/\kappa ^{{\mathfrak {P}}}}\), in which the power depends on the higher regularity of the fluid. Luckily we are able to find a range of \(\varepsilon \) with respect to \(\kappa \) in a scale of large Reynolds number to absorb the Gronwall growth, and achieve a uniform bound of the Boltzmann remainder, which ensures the rigorous justification of the Hilbert expansion in Section 1.1. The main theorem of the uniform controls of the Boltzmann remainder \(f_R\) is given in Theorem 2.

1.3 Higher Regularity of Navier-Stokes Equations in the Inviscid Limit

The inviscid limit of the Navier-Stokes equations (1.13)-(1.15) is at the heart of our approach. Furthermore, in order to control \(f_R\), as explained in the above, we need to derive quantitative higher regularity estimates of the Navier-Stokes solutions which are not directly available in the usual inviscid limit results. Before discussing new features of our analysis, we briefly discuss some prior works on the inviscid limit most relevant to our result. Due to the formation of boundary layers in the limit caused by the mismatch of boundary conditions (1.15) and (1.10), a classical way to tackle the inviscid limit problem is via the Prandtl expansion, of which rigorous justification was shown in [48, 49] for well-prepared data with analytic regularity and in [44] for the initial datum with Sobolev regularity when the initial vorticity is bounded away from the boundary. In particular, the author of [44] introduced the boundary vorticity formulation of (1.13)-(1.15) (see (2.16)-(2.18)) which prompted subsequent interesting works in the field. Among others, in a recent work [47], the authors proved the inviscid limit in 2D based on the Green’s function approach based on Maekawa’s vorticity formulation without having to construct Prandtl boundary layer corrections but by utilizing the boundary layer weights in the norm. In [38, 54], the inviscid limit was shown for initial data that is analytic only near the boundary and has finite Sobolev regularity in the complement in 2D and 3D respectively.

Our analysis of the Navier-Stokes solutions in the limit is based on the Green’s function approach for the Stokes problem using the vorticity formulation (2.16)-(2.18) in the same spirit of [47]. However, the existing methods [38, 47, 54] do not immediately fulfill the goal of our hydrodynamic limit because the analysis of our remainder \(f_R\) requires higher regularity of Navier-Stokes solutions, more specifically \(L^2\) and \(L^\infty \) bounds for higher order derivatives up to two temporal derivatives of \(\nabla _x u\) and p and two spatial derivatives of \(\partial _tu\), while the existing methods do not decipher any bounds for temporal derivatives and the boundedness of the conormal derivatives in their analytic norms does not rule out \(\frac{1}{x_3}\) singularity of the normal derivative of the vorticity in the boundary layer, which may cause the loss of \(L^2\) integrability. To get around these issues, we pursue new estimates of temporal derivatives of the vorticity \(\omega \) by demanding the compatibility conditions for the initial data. With such conditions, the initial layer is absent for \(\omega \) and \(\partial _t\omega \); we can derive an analogous integral representation formula for \(\partial _t \omega \) so that we may run the same fixed point argument for \(\partial _t\omega \) as in [47] without the initial layer. For the second temporal derivative, we handle the initial-boundary layer for the horizontal part with the initial-boundary weight function. These new features allow us to attain the derivative estimates of the vorticity in the normal direction without \(\frac{1}{x_3}\) singularity near the boundary at the expense of losing a power of \(\sqrt{\kappa }\), which is crucial for the control of \(f_R\). The velocity and pressure estimates are then recovered by utilizing elliptic regularity results and the Biot-Savart law in the analytic setting. The main results of Navier-Stokes solutions to (1.13)-(1.15) are given in Theorem 3.

2 Main Results

For the sake of the readers we present the precise statement of main theorems and their notations in this section. We first present the uniform controls of the Boltzmann remainder \(f_R\) of Theorem 2, and the higher regularity of the Navier-Stokes equations in the inviscid limit of Theorem 3. As a consequence of those two theorems we will show a rigorous justification of kinetic approximation of Navier-Stokes in high Reynolds numbers of Theorem 4. Then using the vorticity estimates in Theorem 3 and the famous Kato’s condition in the inviscid limit, we prove a hydrodynamic limit toward the incompressible Euler equations in Corollary 5.

2.1 Uniform Controls of the Boltzmann Remainder \(f_R\) (Theorem 2)

We recall the expansion of Boltzmann solution \(F = \mu + \varepsilon ^2 f_2 \sqrt{\mu } + \delta \varepsilon f_R\sqrt{\mu } \) in (1.19) around the local Maxwellian \(\mu (v) : = M_{1 , \varepsilon u , 1 }(v)\) for any given flow (up) solving the incompressible Navier-Stokes equation with the no-slip boundary condition (1.13)-(1.15).

Inspired by the energy structure of the PDE and the coercivity of the linear operator L in (1.25), we define an energy and a dissipation as

$$\begin{aligned} {\mathcal {E}} (t):=&\ \Vert f_R (t) \Vert _{L^2 (\Omega \times {\mathbb {R}}^3)}^2 + \Vert \partial _t f_R (t) \Vert _{L^2 (\Omega \times {\mathbb {R}}^3)}^2 ,\nonumber \\ {\mathcal {D}} (t) :=&\ \int ^t_0 \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R (s) \Vert _{L^2 (\Omega \times {\mathbb {R}}^3)}^2\mathrm {d}s \nonumber \\&+\int ^t_0 \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}})\partial _t f_R (s) \Vert _{L^2 (\Omega \times {\mathbb {R}}^3)}^2 \mathrm {d}s \nonumber \\&+ \int ^t_0 \Big ( | \varepsilon ^{-\frac{1}{2}} f_R(s) |_{L^2_\gamma }^2 + | \varepsilon ^{-\frac{1}{2}} \partial _t f_R(s) |_{ L^2_\gamma }^2\Big ) \mathrm {d}s . \end{aligned}$$
(2.1)

As explained in Section 1.2, the temporal derivative gets involved mainly in order to access the \(L^6\)-bound of the hydrodynamic part \({\mathbf {P}}f_R\), while we will control the following auxiliary norm to be used in order to handle the nonlinearity: for \(p<3\) and \(t>0\)

$$\begin{aligned} \begin{aligned} {\mathcal {F}}_p(t):=&\sup _{0 \le s \le t} \Big \{ \Vert \kappa ^{1/2} Pf_R(s)\Vert _{L^6 (\Omega )} ^2 + \Vert \kappa ^{1/2} P f_R\Vert _{L^2((0,s); L^p(\Omega ))}^2\\&+ \Vert \kappa ^{ {\mathfrak {P}}+1/2} P \partial _t f_R\Vert _{L^2((0,s); L^p(\Omega ))}^2 + \Vert \varepsilon ^{1/2}\kappa {\mathfrak {w}}_{\varrho ,\ss } f_R(s) \Vert _{L^{\infty }(\Omega \times {\mathbb {R}}^3) }^2 \\&+ \Vert (\varepsilon \kappa )^{3/p} \kappa ^{ \frac{1}{2}+{\mathfrak {P}}} {\mathfrak {w}}_{\varrho ^\prime ,\ss } f_R(s) \Vert _{L^2((0,s);L^{\infty }(\Omega \times {\mathbb {R}}^3)) }^2 \Big \}. \end{aligned} \end{aligned}$$
(2.2)

Here we have introduced weight functions, in order to control an extra quadratic growth in |v| from \((\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu } f_R/\sqrt{\mu }\)

$$\begin{aligned} {\mathfrak {w}}_{\varrho , \ss }(x,v)={\mathfrak {w}} := \exp \{\varrho |v|^2 - {\mathfrak {z}}_{\ss }(x_3) (x \cdot v) \} \ \ \text {for} \ \ 0< \ss \ll \frac{\varrho }{2\pi } \ \text {and} \ 0< \varrho < \frac{1}{4},\nonumber \\ \end{aligned}$$
(2.3)

where \({\mathfrak {z}}_{\ss }: {\mathbb {R}}_+ \rightarrow {\mathbb {R}}_+\) is defined as, for \(\ss >0\)

$$\begin{aligned} \begin{aligned} {\mathfrak {z}}_{\ss }(x_3)= \ss \ \ \text {for} \ \ x_3 \in \big [0, \frac{1}{\ss }-1 \big ], \ \ \text {and} \ \ {\mathfrak {z}}_{\ss }(x_3)=\frac{1}{1+ x_3} \ \ \text {for} \ \ x_3\in \big [\frac{1}{\ss }-1 , \infty \big ). \end{aligned}\nonumber \\ \end{aligned}$$
(2.4)

We have denoted \({\mathfrak {w}}_{\varrho ', \ss }(x,v) ={\mathfrak {w}}'\) for \(\varrho '<\varrho \). Also we have denoted the boundary norms and integral as

$$\begin{aligned} \begin{aligned} |g|_{L^p_{\gamma }} := \left( \int _{\gamma _+}|g|^p+\int _{\gamma _-}|g|^p \right) ^{1/p}, \ |g|_{L^p_{\gamma _\pm }} := \left( \int _{\gamma _\pm }|g|^p\right) ^{1/p}, \\ \int _{\gamma _\pm } f := \int _{\partial \Omega } \int _{n(x) \cdot v > rless 0}f (x, v) |n(x) \cdot v| \mathrm {d}v \mathrm {d}S_x. \end{aligned} \end{aligned}$$
(2.5)

Next we discuss the initial data of the Boltzmann equation. We note that an initial datum of \(f_2\) is already determined by given flow (up). For given initial data \(f_{R,0}:=f_{R,in}\), inspired by the PDE, we define

$$\begin{aligned} \begin{aligned} \partial _t f_{R,0}:=&- \frac{1}{\varepsilon } v\cdot \nabla _x f_{R,in} - \frac{1}{ \varepsilon ^2\kappa } L_{in} f_{R,in} + \frac{2}{\kappa } \Gamma _{in}({f_2}, f_{R,in}) \\&\quad + \frac{ \sqrt{\varepsilon }}{\varepsilon \kappa }\Gamma _{in}(f_{R,in}, f_{R,in}) \\&- \frac{( \partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu _{in}}}{\sqrt{\mu _{in}}} f_{R,in} + ({\mathbf {I}}- {\mathbf {P}}){\mathfrak {R}}_1 (u,p)|_{t=0} + {\mathfrak {R}}_2(u,p)|_{t=0}, \end{aligned} \end{aligned}$$
(2.6)

where \(({\mathbf {I}}- {\mathbf {P}}){\mathfrak {R}}_1\) and \({\mathfrak {R}}_2\) are defined in (3.2) with \(\delta =\sqrt{\varepsilon }\) and \(\mu _{in}\), \(L_{in}\), \(\Gamma _{in}\) are induced by the initial Naiver-Stokes velocity \(u_{in}\). For the remainder \(f_R\) in (1.17), we will use the norms of the initial data:

$$\begin{aligned} {\mathcal {E}}(0):= & {} {\mathcal {E}} (f_{R,0}):= \ \Vert f_{R,0}\Vert _{L^2 (\Omega \times {\mathbb {R}}^3)}^2 + \Vert \partial _t f_{R,0} \Vert _{L^2(\Omega \times {\mathbb {R}}^3)}^2 , \end{aligned}$$
(2.7)
$$\begin{aligned} {\mathcal {F}}_p(0):= & {} \big \{\kappa ^{\frac{1}{2}} | f_{R,0} |_{L^2_\gamma } + \kappa ^{{\mathfrak {P}}+ \frac{1}{2}} | \partial _t f_{R,0} |_{L^2_\gamma }\nonumber \\&+ \varepsilon ^{\frac{1}{2}} \kappa \Vert {\mathfrak {w}} f_ {R,0}\Vert _{L^\infty ({\bar{\Omega }} \times {\mathbb {R}}^3)} + (\varepsilon \kappa )^{1+ \frac{3}{p}} \kappa ^{{\mathfrak {P}}} \Vert {\mathfrak {w}}^\prime \partial _t f_{R,0}\Vert _{L^\infty ({\bar{\Omega }} \times {\mathbb {R}}^3)}\big \}^2.\quad \end{aligned}$$
(2.8)

Theorem 2

(Uniform controls of the Boltzmann remainder \(f_R\)) Suppose for \(T>0\) and \({\mathfrak {P}}\ge 1/2\)

$$\begin{aligned}&\sum _{\ell =0,1} \Vert \nabla _x\partial _t^\ell u\Vert _{L^\infty ([0,T]\times {\bar{\Omega }})}+ \frac{1}{\kappa ^{1/2}} \sum _{\ell =0,1,2}\Vert \partial _t^\ell u\Vert _{L^\infty ([0,T]\times {\bar{\Omega }})} \nonumber \\&\quad \quad + \frac{1}{\kappa ^{1/2}} \Vert p \Vert _{L^\infty ([0,T]\times {\bar{\Omega }})} \lesssim \frac{1}{ \kappa ^{ {\mathfrak {P}}}}. \end{aligned}$$
(2.9)

We further assume that, for \(0 \le {\mathfrak {P}}^\prime < {\mathfrak {P}}\),

$$\begin{aligned} \begin{aligned}&\sum _{\ell =1,2}\Vert \partial _t ^\ell u \Vert _{L^\infty ([0,T]; L^\infty ({\bar{\Omega }})\cap L^2( {\Omega }) )} +\sum _{ \begin{array}{c} 0 \le \ell \le 1 \\ 1 \le |\beta | \le 2 \end{array} } \Vert \nabla _x^\beta \partial _t^\ell u\Vert _{L^\infty ([0,T]; L^\infty ( {\bar{\Omega }}) \cap L^2( {\Omega }) )}\\&\quad + \sum _{|\beta |=1} \Vert \nabla _x^\beta \partial _t^2 u\Vert _{L^2([0,T]; L^\infty ( {\bar{\Omega }}) \cap L^2( {\Omega }) )} \\&\quad + \Vert \partial _t^2 p \Vert _{L^2 ([0,T]; L^\infty ({\bar{\Omega }}) \cap L^2(\Omega ) )} + \sum _{|\beta |=0,1} \Vert \nabla _x ^\beta \partial _t p \Vert _{L^\infty ([0,T]; L^\infty ({\bar{\Omega }}) \cap L^2(\Omega ))} \\&\lesssim \exp \Big ({\frac{1}{\kappa ^{{\mathfrak {P}}^\prime }}}\Big ) . \end{aligned} \end{aligned}$$
(2.10)

For given such \(T>0\), let us choose \(\varepsilon , \delta \) and \(\kappa \) as, for some \({\mathfrak {C}}\gg 1\),

$$\begin{aligned} \delta = \sqrt{\varepsilon } \ \ \text {and} \ \ \delta =\exp \Big ( \frac{ - {\mathfrak {C}} T}{ \kappa ^{ {\mathfrak {P}}} }\Big ). \end{aligned}$$
(2.11)

Assume that an initial datum for the remainder \(f_{R,in }\) satisfies, for some \(p<3\) and \(|p-3|\ll 1\),

$$\begin{aligned} \sqrt{{\mathcal {E}}(0) }+ \sqrt{{\mathcal {F}} _p (0) } \lesssim 1. \end{aligned}$$
(2.12)

Then we construct a unique solution \(f_R(t,x,v)\) of the form of

$$\begin{aligned} F = M_{1, \varepsilon u ,1 } + \varepsilon ^2 f_{2} \sqrt{M_{1, \varepsilon u ,1 } } + \delta \varepsilon f_{R} \sqrt{M_{1, \varepsilon u ,1 } } \ \ \text {in} \ \ [0,T] \times \Omega \times {\mathbb {R}}^3, \end{aligned}$$

which solves the Boltzmann equation (1.1) and the diffuse reflection boundary condition (1.12) with the scale of (1.8) and (2.11), and satisfies the initial condition \(F|_{t=0}= M_{1, \varepsilon u _{in} ,1 } + \varepsilon ^2 f_{2} \sqrt{M_{1, \varepsilon u ,1 } } |_{t=0} + \delta \varepsilon f_{R,in} \sqrt{M_{1, \varepsilon u ,1 } }|_{t=0}\), in a time interval \(t \in [0,T]\). Moreover, we have

$$\begin{aligned} \delta ^{ \frac{1}{2}- \frac{3}{p} (1- \frac{p}{3})} \sup _{0 \le t \le T} \big \{ \sqrt{{\mathcal {E}}(t)} + \sqrt{{\mathcal {D}}(t)}+ \sqrt{{\mathcal {F}}_{ p } (t)}\big \} \lesssim 1. \end{aligned}$$
(2.13)

Remark 3

The condition (2.11) in the theorem is indeed the largest \(\sqrt{\varepsilon }\) can be allowed. Any smaller \(\sqrt{\varepsilon }\) than \(\exp \Big ( \frac{ - {\mathfrak {C}} T}{ \kappa ^{1/2} }\Big )\) (which means \(\sqrt{\varepsilon }\) decaying faster than \(\exp \Big ( \frac{ - {\mathfrak {C}} T}{ \kappa ^{1/2} }\Big )\) as \(\kappa \downarrow 0\)) will produce the same result. In terms of (1.8) the relation (2.11) implies that the Knudsen number has to vanish only slightly faster than the Mach number :

(2.14)

The proof of Theorem 2 will be given in Section 4.

2.2 Higher Regularity of Navier-Stokes Equations in the Inviscid Limit (Theorem 3)

For the Navier-Stokes solutions to (1.13)-(1.15), we introduce real analytic norms and function spaces, adopted from [47] and [54] for the 3D counter part with slight modifications.

In this subsection and Section 5, we will use the following notations: \(x=(x_h,x_3)=(x_1,x_2,x_3)\in {\mathbb {T}}^2\times {\mathbb {R}}_+=\Omega \), \(\nabla _x=\nabla =(\nabla _h,\partial _3)=(\partial _{x_1},\partial _{x_2},\partial _{x_3})\); for a vector valued function \(g\in {\mathbb {R}}^3\), \(g=(g_h,g_3)=(g_1,g_2,g_3)\).

We denote the vorticity by

$$\begin{aligned} \omega =\nabla \times u, \ \ \ u =\nabla \times (- \Delta )^{-1} {\omega }, \end{aligned}$$
(2.15)

while the second identity is the famous Biot-Savart law. Here \((- \Delta )^{-1}\) denotes the inverse of \(-\Delta \) with the zero Dirichlet boundary condition on \(\partial \Omega \).

Our analysis of the Navier-Stokes solutions is based on the vorticity formulation in 3D ([43, 44]):

$$\begin{aligned}&\partial _t \omega - \kappa \eta _0 \Delta \omega = - u \cdot \nabla \omega + \omega \cdot \nabla u \ \ \text {in} \ \ \Omega , \quad \end{aligned}$$
(2.16)
$$\begin{aligned}&\omega \,|_{t=0} = \omega _{in} \ \ \text {in} \ \ \Omega , \end{aligned}$$
(2.17)
$$\begin{aligned}&\kappa \eta _0 (\partial _{x_3} + \sqrt{- \Delta _h})\omega _h \, = [\partial _{x_3}(- \Delta )^{-1} (-u \cdot \nabla \omega _h + \omega \cdot \nabla u_h ) ] \, , \ \ \omega _3 =0 \ \ \text {on} \ \ \partial \Omega , \end{aligned}$$
(2.18)

where \(\sqrt{- \Delta _h}=|\nabla _h|\) is defined as

$$\begin{aligned} \sqrt{- \Delta _h}g(x_h,x_3)= \sum _{\xi \in {\mathbb {Z}}^2} |\xi | g_\xi (x_3) e^{i x_h \cdot \xi }. \end{aligned}$$
(2.19)

Here, \(g_{\xi } (x_3)= \frac{1}{(2\pi )^2} \iint _{{\mathbb {T}}^2} e^{-i x_h \cdot \xi } g(x_h,x_3) \mathrm {d}x_h \in {\mathbb {C}} \text { with } \xi = (\xi _1, \xi _2) \in {\mathbb {Z}}^2\) denotes the Fourier transform in the horizontal variables, which satisfies \(g(x_1,x_2,x_3) = \sum _{\xi \in {\mathbb {Z}}^2} g_{\xi } (x_3) e^{i x_h \cdot \xi }.\) The Fourier transform can be regarded as a function \(g_{\xi } (z)\) where z is sitting in a pencil-like complex domain: for any \(\lambda > 0\),

$$\begin{aligned} {\mathcal {H}}_\lambda :=\Big \{ z \in {\mathbb {C}} : \text {Re}\,z\ge 0, \; | \text {Im}\, z| < \lambda \min \{ \text {Re}\,z, 1\} \Big \}. \end{aligned}$$
(2.20)

We define analytic function spaces without the boundary layer, \({\mathfrak {L}}^{p,\lambda }\), for holomorphic functions with a finite norm, for \(p\ge 1\),

$$\begin{aligned} \Vert g \Vert _{p,\lambda } := \sum _{\xi \in {\mathbb {Z}}^2} e^{\lambda |\xi |} \Vert g_{\xi } \Vert _{{\mathcal {L}}^p_\lambda } \ \ \text {where} \ \ \Vert g_{\xi }\Vert _{{\mathcal {L}}^p_\lambda } := \sup _{0\le \sigma \le \lambda } \left( \int _{\partial {\mathcal {H}}_\sigma } | g_{\xi } (z) |^p |\mathrm {d}z| \right) ^{1/p}.\nonumber \\ \end{aligned}$$
(2.21)

Next we introduce an \(L^\infty \)-based analytic boundary layer function space, for \(\lambda >0\) and \(\kappa \ge 0\), that consists of holomorphic functions in \({\mathcal {H}}_\lambda \) with a finite norm

$$\begin{aligned} \begin{aligned} \Vert g\Vert _{\infty ,\lambda ,\kappa } = \sum _{\xi \in {{\mathbb {Z}}^2}} e^{\lambda |\xi |} \Vert g_{\xi } \Vert _{{\mathcal {L}}^\infty _{\lambda , \kappa }} , \end{aligned} \end{aligned}$$
(2.22)

where \(\Vert g_{\xi } \Vert _{{\mathcal {L}}^\infty _{\lambda ,0}} : = \Vert e^{{\bar{\alpha }} \text {Re}\,z} g_{\xi } (z) \Vert _{{\mathcal {L}}^\infty _\lambda } := \sup _{z \in {\mathcal {H}}_\lambda }e^{{\bar{\alpha }} \text {Re}\,z} g_{\xi } (z) \) and

$$\begin{aligned} \Vert g_{\xi } \Vert _{{\mathcal {L}}^\infty _{\lambda , \kappa }} := \bigg \Vert \frac{e^{{\bar{\alpha }} \text {Re}\,z}}{ 1+ \phi _\kappa (z)} g_{\xi } (z) \bigg \Vert _{{\mathcal {L}}^\infty _\lambda } := \sup _{z \in {\mathcal {H}}_\lambda }\frac{e^{{\bar{\alpha }} \text {Re}\,z}}{ 1+ \phi _\kappa (z)} |g_{\xi } (z)| . \end{aligned}$$

Here, a boundary layer weight function is defined as

$$\begin{aligned} \phi _\kappa (z):= \frac{1}{\sqrt{\kappa }} \phi ( \frac{z}{\sqrt{\kappa }}) \ \ \text {with} \ \ \phi (z) = \frac{1}{ 1+|\text {Re}\,z|^{\mathfrak {r}}} \ \text {for some } {\mathfrak {r}}>1. \end{aligned}$$
(2.23)

We define \({\mathfrak {B}}^{\lambda , \kappa }\) for holomorphic functions \(g= (g_1,g_2,g_3)\) with a finite norm

$$\begin{aligned}{}[[g ]]_{ \infty , \lambda ,\kappa }=\sum _{i=1,2} \Vert g_i \Vert _{\infty , \lambda , \kappa } + \Vert g_3 \Vert _{\infty , \lambda , 0}. \end{aligned}$$
(2.24)

We note that \({\mathfrak {B}}^{\lambda , \kappa } \subset {\mathfrak {L}}^{1,\lambda }\), but \({\mathfrak {B}}^{\lambda , 0} \subsetneqq {\mathfrak {L}}^{\infty ,\lambda }\) if \({\bar{\alpha }}>0\).

Due to its singular nature of the Navier-Stokes flow in the inviscid limit, we introduce the conormal derivatives

$$\begin{aligned} D= (D_h, D_3)= (\nabla _h, \zeta (x_3) \partial _3) \ \ \text {where} \ \ \zeta (z) = \frac{z}{1+z}. \end{aligned}$$
(2.25)

With the multi-indices \(\beta =(\beta _h,\beta _3):=(\beta _1,\beta _2,\beta _3)\in {\mathbb {N}}_0^3\), the higher derivatives are denoted by \(D^\beta = \partial _1^{\beta _1} \partial _2 ^{\beta _2} D_3^{\beta _3}\) and \(D^\beta _\xi = (i \xi _1)^{\beta _1} (i \xi _2) ^{\beta _2} D_3^{\beta _3}\).

Now we define, for \(\lambda _0>0\), \(\gamma _0>0\), \(\alpha >0\), \(\kappa \ge 0\), and \(t \in (0, \frac{\lambda _0}{2 \gamma _0})\)

$$\begin{aligned} {\left| \left| \left| g \right| \right| \right| }_{\infty ,\kappa }= & {} \sup _{\lambda <\lambda _0-\gamma _0 t} \bigg \{ \sum _{ 0 \le |\beta | \le 1} [[ D^\beta g ]]_{ \infty , \lambda , \kappa } + \sum _{ |\beta | =2}(\lambda _0-\lambda -\gamma _0 t)^\alpha [[D^\beta g ]]_{ {\infty , \lambda , \kappa } } \bigg \}, \nonumber \\\end{aligned}$$
(2.26)
$$\begin{aligned} {\left| \left| \left| g \right| \right| \right| }_1= & {} \sup _{\lambda <\lambda _0-\gamma _0 t} \bigg \{ \sum _{0\le |\beta |\le 1} \Vert D^\beta (1+|\nabla _h|) g \Vert _{1,\lambda } \nonumber \\&\qquad \qquad \qquad \quad + (\lambda _0-\lambda -\gamma _0 t)^\alpha \sum _{ |\beta |=2} \Vert D^\beta (1+|\nabla _h|) g \Vert _{1,\lambda } \bigg \}. \end{aligned}$$
(2.27)

With an initial-boundary layer weight function as in [47]

$$\begin{aligned} \phi _{\kappa t} (z)= \frac{1}{\sqrt{\kappa t}} \phi ( \frac{z}{\sqrt{\kappa t}}) , \end{aligned}$$
(2.28)

we define an initial-boundary layer function space \({\mathfrak {B}}^{\lambda , \kappa t}\) for holomorphic functions \(g= (g_1,g_2,g_3)\) with a finite norm

$$\begin{aligned}{}[[g ]]_{ \infty , \lambda ,\kappa t}=\sum _{i=1,2} \Vert g_i \Vert _{\infty , \lambda , \kappa t} + \Vert g_3 \Vert _{\infty , \lambda , 0} , \end{aligned}$$
(2.29)

where an \(L^\infty \)-based analytic norm with the initial-boundary layer is defined as

$$\begin{aligned} \Vert g\Vert _{\infty ,\lambda ,\kappa t} = \sum _{\xi \in {\mathbb {Z}}^2} e^{\lambda |\xi |} \Vert g_{\xi } \Vert _{{\mathcal {L}}^\infty _{\lambda , \kappa t}}, \quad \Vert g_{\xi } \Vert _{{\mathcal {L}}^\infty _{\lambda , \kappa t}} = \bigg \Vert \frac{e^{{\bar{\alpha }} \text {Re}\,z}}{ 1+ \phi _\kappa (z)+ \phi _{\kappa t} (z)} g_{\xi } (z)\bigg \Vert _{{\mathcal {L}}^\infty _\lambda } .\nonumber \\ \end{aligned}$$
(2.30)

We finally define, for \(t \in (0, \frac{\lambda _0}{2\gamma _0})\),

$$\begin{aligned} {\left| \left| \left| g \right| \right| \right| }_{\infty , \kappa t}= \sup _{\lambda <\lambda _0-\gamma _0 t} \bigg \{ \sum _{ 0 \le |\beta | \le 1 } [[ D^\beta g]]_{\infty ,\lambda , \kappa t}+ \sum _{|\beta |=2} (\lambda _0-\lambda -\gamma _0 t)^\alpha [[ D^\beta g]]_{\infty ,\lambda , \kappa t} \bigg \}.\nonumber \\ \end{aligned}$$
(2.31)

In this subsection and Section 5, \(\alpha \), \({\bar{\alpha }}\) are given positive small constants, \(\lambda _0\) is a given positive constant, and \(\gamma _0\) is a sufficiently large constant to be determined in Theorem 3.

Next we discuss the initial data of the velocity \(u_{in}\) and the corresponding vorticity \(\omega _{in}= \nabla _x \times u_{in}\). Inspired by the PDEs, let

$$\begin{aligned} \begin{aligned} \omega _0&:=\omega _{in},\quad \partial _t\omega _0:= \kappa \eta _0\Delta \omega _0 - u_0\cdot \nabla \omega _0+ \omega _0 \cdot \nabla u_0,\\ \quad u_0&:= \nabla \times (-\Delta )^{-1} \omega _0 , \quad \partial _t u_0 := \nabla \times (-\Delta )^{-1} \partial _t \omega _0 , \\ \partial _t^2\omega _0&:= \kappa \eta _0 \Delta \partial _t\omega _0 - u_{0} \cdot \nabla \partial _t\omega _0 - \partial _t u_0 \cdot \nabla \omega _0 + \omega _0 \cdot \nabla \partial _t u_0 + \partial _t\omega _0 \cdot \nabla u_0. \end{aligned} \end{aligned}$$
(2.32)

Theorem 3

Let \(\lambda _0>0\) and \(\omega _{in} \in {\mathfrak {B}}^{\lambda _0,\kappa }\) with (2.32) satisfy

$$\begin{aligned} \sum _{0\le |\beta |\le 2} \Vert D^\beta \partial _t^\ell \omega _0 \Vert _{1,\lambda _0}+\sum _{0\le |\beta |\le 2} \Vert D^\beta \partial _t^\ell \omega _0 \Vert _{\infty ,\lambda _0, \kappa } <\infty \ \text { for } \ \ell =0,1,2. \end{aligned}$$
(2.33)

Further assume that \(\omega _{in}=\omega _0\) and (2.32) satisfies the compatibility conditions on \(\partial \Omega \)

$$\begin{aligned} \begin{aligned} \kappa \eta _0 (\partial _{x_3} + \sqrt{- \Delta _h})\omega _{0,h}&= [\partial _{x_3} (-\Delta )^{-1} (-u_0 \cdot \nabla \omega _{0, h} + \omega _0 \cdot \nabla u_{0,h}) ] , \\ \omega _{0,3} =0, \ \ \partial _t\omega _{0,3} = 0 . \end{aligned} \end{aligned}$$
(2.34)

Then there exists a constant \(\gamma _0>0\) and a time \(T>0\) depending only on \(\lambda _0\) and the size of the initial data such that the solution \(\omega (t)\) to the vorticity formulation of the Navier-Stokes equations (2.16)–(2.18) exists in \(C^1([0,T]; {\mathfrak {B}}^{\lambda , \kappa })\) with \(\partial _t^2\omega \) in \(C(0,T; {\mathfrak {B}}^{\lambda , \kappa t})\) for \(0<\lambda <\lambda _0\) satisfying

$$\begin{aligned} \sup _{t\in [0,T]} \left[ \sum _{\ell =0}^2 {\left| \left| \left| \partial _t^\ell \omega (t) \right| \right| \right| }_1 + \sum _{\ell =0}^1 {\left| \left| \left| \partial _t^\ell \omega (t) \right| \right| \right| }_{\infty ,\kappa } + {\left| \left| \left| \partial _t^2\omega (t) \right| \right| \right| }_{\infty ,\kappa t} \right] <\infty . \end{aligned}$$
(2.35)

Furthermore, for each \((t,x)\in [0,T]\times \Omega \),

  1. (1)

    (Bounds on the vorticity and its derivatives) \(\omega (t,x)\) enjoys the following bounds:

    $$\begin{aligned} | \nabla _{h}^i \partial _t^\ell \omega _h (t,x) |&\lesssim e^{-{\bar{\alpha }} x_3} \left( 1 + \phi _\kappa (x_3) \right) , \ \ | \nabla _{h}^i \partial _t^\ell \omega _3 (t,x) | \lesssim e^{-{\bar{\alpha }} x_3} \text { for } i,\ell =0,1, \end{aligned}$$
    (2.36)
    $$\begin{aligned} | \partial _t^2 \omega _h (t,x) |&\lesssim e^{-{\bar{\alpha }} x_3} \left( 1 + \phi _\kappa (x_3) + \phi _{\kappa t} (x_3) \right) , \ \ | \partial _t^2 \omega _3 (t,x) | \lesssim e^{-{\bar{\alpha }} x_3}, \end{aligned}$$
    (2.37)
    $$\begin{aligned} | \partial _{x_3} \partial _t^\ell \omega _h (t,x) |&\lesssim {\kappa }^{-1} e^{-{\bar{\alpha }} x_3}, \ \ | \partial _{x_3} \partial _t^\ell \omega _3 (t,x) | \lesssim e^{-{\bar{\alpha }} x_3} \left( 1 + \phi _\kappa (x_3) \right) \text { for } \ell =0,1. \end{aligned}$$
    (2.38)
  2. (2)

    (Bounds on the velocity and its derivatives) The corresponding velocity field u(tx) satisfies the following:

    $$\begin{aligned} |\partial _t^\ell u (t,x)|&\lesssim 1 \ \ \text {for} \ \ell =0,1,2, \end{aligned}$$
    (2.39)
    $$\begin{aligned} \sum _{1 \le |\beta | \le 2} |\nabla ^{\beta } \partial _t^\ell u(t,x )|&\lesssim \big (1+ \phi _\kappa (x_3) + (|\beta |-1) {\kappa }^{-1} \big ) e^{-\min (1, \frac{{\bar{\alpha }}}{2} )x_3} \ \ \text {for} \ \ell =0,1, \end{aligned}$$
    (2.40)
    $$\begin{aligned} \sum _{ |\beta | =1} |\nabla ^{\beta } \partial _t^2 u(t,x )|&\lesssim \big (1+ \phi _\kappa (x_3) + \phi _{\kappa t} (x_3)\big )e^{-\min (1, \frac{{\bar{\alpha }}}{2} )x_3} . \end{aligned}$$
    (2.41)

    Moreover, we have the decay estimate for \(\partial _t^\ell u\):

    $$\begin{aligned} |\partial _t^\ell u|\lesssim \kappa ^{-\frac{1}{2}} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) x_3} \ \ \text {for} \ \ell =1,2. \end{aligned}$$
    (2.42)
  3. (3)

    (Bounds on the pressure and its derivatives) The pressure defined in (5.73) satisfies the following:

    $$\begin{aligned} |\partial _t^\ell p (t,x)|&\lesssim 1 \ \ \text {for} \ \ell =0,1,2, \end{aligned}$$
    (2.43)
    $$\begin{aligned} \sum _{0 \le |\beta | \le 1} |\nabla ^{\beta } \partial _t^\ell p(t,x )|&\lesssim \kappa ^{-\frac{1}{2}} e^{-\min (1, \frac{{\bar{\alpha }}}{2} )x_3} \ \ \text {for} \ \ell =0,1, \end{aligned}$$
    (2.44)
    $$\begin{aligned} |\partial _t^2 p|&\lesssim ( \kappa ^{-\frac{1}{2}} +\phi _{\kappa t}(x_3) )e^{-\min (1, \frac{{\bar{\alpha }}}{2} )x_3} . \end{aligned}$$
    (2.45)

Remark 4

For simplicity of the presentation, we have taken the analytic data with the same analyticity radius in \(x_1\), \(x_2\) and \(x_3\) with the exponential decay for large \(x_3\). As shown in [38, 54], more general initial data requiring the analyticity only near the boundary can be taken.

Remark 5

The horizontal vorticity \(\omega _h\) and the vertical vorticity \(\omega _3\) obey different boundary conditions (2.18) which enforce different behaviors near the boundary. This is well-reflected in our \(L^\infty \) based norms in (2.24) and (2.29). As noted in [54], such incompatible behaviors of \(\omega _h\) and \(\omega _3\) in 3D are dealt with the \(L^1\) based norm (2.27) which contains one more tangential derivative \((1+|\nabla _h|)\), which is different from 2D analysis [38, 47].

Remark 6

We demand the compatibility conditions in (2.34) in order to avoid singular initial-boundary layers for the temporal derivatives of the vorticity. If the first two conditions in (2.34) were not satisfied, the initial-boundary layers would occur for the first temporal derivative of the vorticity. For the second temporal derivative, we handle the initial-boundary layer for the horizontal part with the initial-boundary layer weight, while for the vertical part we further demand \(\partial _t\omega _{0,3}|_{x_3=0}=0\) in order to rule out a singular initial-boundary layer caused by the Dirichlet boundary condition. This amounts to requiring the second order vanishing condition at the boundary for \(\omega _{0,3}\), which is satisfied by a large class of \(\omega _0\). We remark that the first condition of (2.34) is also satisfied by a large class of \(\omega _0\). In fact, if not, by the result of [47], we can obtain a short time solution \({\tilde{\omega }}(t)\) to (2.16)–(2.18) and may reset the initial data by \(\omega _0 = {\tilde{\omega }}(t=t_0)\) for sufficiently small \(t_0>0\).

The proof of Theorem 3 will be given in Section 5.

2.3 Main Theorem

Now we present the full statement of the main theorems of this paper:

Theorem 4

(Kinetic approximation of Navier-Stokes in large Reynolds numbers) We consider a half space \(\Omega \) in 3D as in (1.18). Suppose an initial datum of the Navier-Stokes flow \(u_{in}\) is divergence-free \(\nabla _x \cdot u_{in}=0\) in \(\Omega \) and the corresponding initial vorticity \(\omega _{in}= \nabla _x \times u_{in}\) belongs to the real analytic space \({\mathfrak {B}}^{\lambda _0,\kappa }\) of (2.24) for some \(\lambda _0>0\) such that (2.33) holds. Further we assume that \(\omega _{in}\) satisfies the compatibility conditions (2.34) on \(\partial \Omega \). Then there exists a unique real analytic solution \((u(t,x), \nabla _x p(t,x))\) to (1.13)–(1.15) in \([0,T] \times \Omega \), while \(T>0\) only depends on \(\lambda _0\) and the size of the initial data as in (2.33).

Choosing a pressure p(tx) such that \(p(t,x)\rightarrow 0\) as \(x_3 \uparrow \infty \), we set the local Maxwellian and the second order correction \(f_2\) as

$$\begin{aligned} \mu:= & {} M_{1,\varepsilon u, 1} =\frac{1}{(2\pi )^{ \frac{3}{2}}}\exp \left\{ - \frac{|v-\varepsilon u|^2}{2}\right\} ,\\ f_2:= & {} {\mathbf {P}}f_2 + ( {\mathbf {I}}-{\mathbf {P}})f_2 = p \varphi _0 \sqrt{\mu }+ (1.34). \end{aligned}$$

For given such \(T>0\), let us choose \(\varepsilon \) and \(\kappa \) in the relation of (2.11).

Assume that an initial datum for the remainder \(f_{R,in }\) satisfies (2.12) for some \(p<3\) and \(|p-3|\ll 1\).

Then we construct a unique solution \(f_R(t,x,v)\) of the form of

$$\begin{aligned} F = M_{1, \varepsilon u ,1 } + \varepsilon ^2 f_{2} \sqrt{M_{1, \varepsilon u ,1 }} + \varepsilon ^{3/2} f_{R} \sqrt{M_{1, \varepsilon u ,1 }} \ \ \text {in} \ \ [0,T] \times \Omega \times {\mathbb {R}}^3, \end{aligned}$$

which solves the Boltzmann equation (1.1) and the diffuse reflection boundary condition (1.12) with the scale of (1.8) and (2.11), and satisfies the initial condition \(F|_{t=0}= M_{1, \varepsilon u _{in} ,1 } + \varepsilon ^2 f_{2} \sqrt{M_{1, \varepsilon u ,1 }} |_{t=0} + \varepsilon ^{3/2} f_{R,in} \sqrt{M_{1, \varepsilon u ,1 }}|_{t=0}\).

Moreover we derive that, for each \(\varepsilon \) and \(\kappa \) of (2.11),

$$\begin{aligned} \sup _{0 \le t \le T}\left\| \frac{F (t,x,v)- M_{1, \varepsilon u(t,x), 1} (v)}{\varepsilon \sqrt{M_{1, \varepsilon u(t,x), 1} (v)} } \right\| _{L^2(\Omega \times {\mathbb {R}}^3)} \lesssim \exp \Big ( \frac{ - {\mathfrak {C}} T}{ 2\kappa ^{1/2} }\Big ) \ \ \text {for} \ \ \kappa \ll 1.\qquad \end{aligned}$$
(2.46)

Proof

The existence of the Navier-Stokes solutions follows from Theorem 3. For the remaining assertions, we note that all the estimates (2.39)–(2.42) of Theorem 3 ensure the conditions of Theorem 2 with \({\mathfrak {P}}=\frac{1}{2}\). Therefore the conclusion follows directly as a consequence of Theorem 2 and Theorem 3. \(\square \)

The incompressible Euler limit follows as a byproduct of the main theorem:

Corollary 5

(Hydrodynamic limit toward the incompressible Euler equation) Let \(u_E(t,x)\) be a (unique) solution of the incompressible Euler equations (1.9)-(1.10) with the initial condition \(u_E|_{t=0}= u_{in}\) in \(\Omega \). Then

$$\begin{aligned} \sup _{0 \le t \le T}\left\| \frac{F (t,x,v)- M_{1, \varepsilon u_E (t,x), 1} (v)}{\varepsilon (1+|v|)^2\sqrt{M_{1,0,1}(v) } } \right\| _{L^2(\Omega \times {\mathbb {R}}^3 )} \longrightarrow 0 \ \ \text {as} \ \ \varepsilon \downarrow 0. \end{aligned}$$

Proof

Note that

$$\begin{aligned} F (t,x,v)- M_{1, \varepsilon u_E(t,x), 1} (v)= & {} \big [F (t,x,v)- M_{1, \varepsilon u(t,x), 1} (v)\big ] \\&+\big [M_{1, \varepsilon u(t,x), 1} (v) - M_{1, \varepsilon u_E(t,x), 1} (v)\big ]. \end{aligned}$$

The first term can be bounded as in (2.46). We bound the second term by an expansion:

$$\begin{aligned} \begin{aligned} |u(t,x)-u_E(t,x)| \int ^\varepsilon _0 |(v-\varepsilon u_E) + a(u_E-u)| e^{- \frac{|(v-\varepsilon u_E) + a(u_E-u)|^2}{2}} \mathrm {d}a . \end{aligned} \end{aligned}$$

Note that \(\Vert \varepsilon u \Vert _{L^\infty } \ll 1 \) and \(\Vert \varepsilon u_E\Vert _{L^\infty }\ll 1\) from Theorem 3. Then we conclude that the second term converges to 0 as \(\kappa \downarrow 0\) from Theorem 3 and the famous Kato’s condition for vanishing viscosity limit in [34].\(\square \)

3 Hilbert Expansion Around a Local Maxwellian and Source Terms

In this section we complete the Hilbert expansion along with the outline of the introduction. As a result we prove

Proposition 6

Suppose that F of (1.19), with a free parameter \(\delta \), solve (1.1) and (1.12) with (1.8) and that (up) solves (1.13)-(1.15). We choose a hydrodynamic part \(f_2\) as

$$\begin{aligned} {\mathbf {P}}f_2 = p \varphi _0 \sqrt{\mu }, \end{aligned}$$
(3.1)

with the pressure p of the Navier-Stokes flow in (1.13), and \(({\mathbf {I}}- {\mathbf {P}})f_2\) has been given in (1.34). Then \(f_{R}\) in (1.19) satisfies that

$$\begin{aligned}&\Big [ \partial _t + \frac{1}{\varepsilon } v\cdot \nabla _x + \frac{1}{ \varepsilon ^2\kappa } L \Big ] f_R = \frac{2}{\kappa } \Gamma ({f_2}, f_R) + \frac{ \delta }{\varepsilon \kappa }\Gamma (f_R, f_R) \nonumber \\&\qquad - \frac{( \partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_{R} + ({\mathbf {I}}- {\mathbf {P}}){\mathfrak {R}}_1 + {\mathfrak {R}}_2, \end{aligned}$$
(3.2)
$$\begin{aligned}&\Big [ \partial _t + \frac{1}{\varepsilon } v\cdot \nabla _x + \frac{1}{ \varepsilon ^2\kappa } L \Big ] \partial _t f_R \nonumber \\&\quad = - \frac{1}{\varepsilon ^2 \kappa } L_t ({\mathbf {I}} - {\mathbf {P}}) f_R +\frac{1}{\varepsilon ^2 \kappa } L({\mathbf {P}}_t f_R)+ \frac{ 2 \delta }{\varepsilon \kappa }\Gamma (f_R,\partial _t f_R)\nonumber \\&\qquad + \frac{2}{\kappa } \Gamma ({f_2},\partial _t f_R) + \frac{2}{\kappa } \Gamma (\partial _t {f_2}, f_R) + \frac{2}{\kappa } \Gamma _t( {f_2}, f_R)+ \frac{\delta }{\varepsilon \kappa } \Gamma _t (f_R,f_R) \nonumber \\&\qquad - \frac{( \partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} \partial _{t} f_{R}- \partial _t \Big ( \frac{( \partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} \Big ) f_{R} \nonumber \\&\qquad +({\mathbf {I}} - {\mathbf {P}}){\mathfrak {R}}_3 + {\mathfrak {R}}_4 , \end{aligned}$$
(3.3)

where the commutators \(L_t\), \({\mathbf {P}}_t\) and \(\Gamma _t\) are given in (3.34), while

$$\begin{aligned}&e^{ \varrho |v-\varepsilon u|^{2}} |({\mathbf {I}} - {\mathbf {P}}) {\mathfrak {R}}_{1} (t,x,v)| \lesssim \frac{1}{\delta } \kappa |\nabla _x^2 u| ,\end{aligned}$$
(3.4)
$$\begin{aligned}&e^{ \varrho |v-\varepsilon u|^{2}} |{\mathfrak {R}}_{2} (t,x,v)| \lesssim \frac{\varepsilon }{\delta } (|p| + \kappa |\nabla _x u|)|\nabla _x u| + \frac{\varepsilon }{\delta } (|\partial _t p| + \kappa |\nabla _x u|) \nonumber \\&\qquad \qquad \qquad \qquad \qquad \qquad + \frac{\varepsilon \kappa }{\delta } (|\nabla _x \partial _t u| + |u| |\nabla _x^2 u|) \nonumber \\&\qquad \qquad \qquad \qquad \qquad \qquad + \frac{\varepsilon ^2}{\delta } (|p| + \kappa | \nabla _x u|) (|\partial _t u| + |u| |\nabla _x u|) , \end{aligned}$$
(3.5)
$$\begin{aligned}&e^{ \varrho |v-\varepsilon u|^{2}} |({\mathbf {I}} - {\mathbf {P}}) {\mathfrak {R}}_{3} (t,x,v)| \lesssim \frac{\kappa }{\delta } |\nabla _x^2 \partial _t u| , \end{aligned}$$
(3.6)
$$\begin{aligned}&e^{ \varrho |v-\varepsilon u|^{2}}|{\mathfrak {R}}_4 (t,x,v)| \nonumber \\&\quad \lesssim \frac{\varepsilon }{\delta } |\partial _t ^2 p| + \frac{\varepsilon \kappa }{\delta } |\nabla _x \partial _t^2 u| + \frac{\varepsilon }{\delta }|\nabla _x \partial _t p||u| + \frac{\varepsilon \kappa }{\delta }|u| |\nabla _x^2 \partial _t u| \nonumber \\&\quad +\frac{\varepsilon }{\delta } \{ (1+ |u|)(|p|+ \kappa |\nabla _x u|) +\kappa \varepsilon |\partial _t u| \} |\nabla _x \partial _t u|\nonumber \\&\quad + \frac{\varepsilon \kappa }{\delta } (1+ \varepsilon \kappa |u|)|\partial _t u ||\nabla _x^2 u| + \frac{\varepsilon ^2}{\delta } \{|p|+\kappa |\nabla _x u|\} |\partial _t^2 u| \nonumber \\&\quad + \frac{\varepsilon }{\delta } \{ (|u| + \varepsilon |p| + \varepsilon ^2 |p||u|)|\partial _t u| + (1+ \varepsilon |u|) |\partial _t p| \} |\nabla _x u|\nonumber \\&\qquad +\frac{\varepsilon ^2\kappa }{\delta } (1+ \varepsilon |u|) |\partial _t u | |\nabla _x u|^2 \nonumber \\&\qquad + \frac{\varepsilon }{\delta } \{ |\partial _t u| + |\nabla _x p| + \varepsilon |\partial _t p| + \frac{\varepsilon }{\kappa } (|p|^2 + \kappa |u| |\nabla _x p| + \varepsilon \kappa |\partial _t u| |p|) \} |\partial _t u| . \end{aligned}$$
(3.7)

At the boundary \(f_R\) and \(\partial _t f_R\) satisfy

$$\begin{aligned} f_R(t,x,v)|_{\gamma _-}= & {} P_{\gamma _+} f_R(t,x,v)- \frac{\varepsilon }{\delta } (1- P_{\gamma _+}) ({\mathbf {I}}- {\mathbf {P}})f_2(t,x,v), \end{aligned}$$
(3.8)
$$\begin{aligned} \partial _t f_R |_{\gamma _-}= & {} P_{\gamma _+} \partial _t f_R- \frac{\varepsilon }{\delta } (1-P_{\gamma _+}) \partial _t ({\mathbf {I}} - {\mathbf {P}})f_2 +r_{\gamma _+} (f_R)- \frac{\varepsilon }{\delta }r_{\gamma _+} (({\mathbf {I}} - {\mathbf {P}}) f_2),\nonumber \\ r _{\gamma _+}(g):= & {} \partial _t \sqrt{c_\mu \mu (v)} \int _{n(x) \cdot {\mathfrak {v}}>0} g \sqrt{c_\mu \mu ({\mathfrak {v}})} n(x) \cdot {\mathfrak {v}} \mathrm {d}{\mathfrak {v}} \nonumber \\&\quad + \sqrt{c_\mu \mu (v)} \int _{n(x) \cdot {\mathfrak {v}}>0}g \partial _t \sqrt{c_\mu \mu ({\mathfrak {v}})} n(x) \cdot {\mathfrak {v}} \mathrm {d}{\mathfrak {v}}. \end{aligned}$$
(3.9)

In addition,

$$\begin{aligned}&e^{ \varrho |v-\varepsilon u|^2}|f_2 (t,x,v)| \lesssim \ |p(t,x)| + \kappa |\nabla _x u(t,x)|, \end{aligned}$$
(3.10)
$$\begin{aligned}&e^{ \varrho |v-\varepsilon u|^2}|\partial _t f_2 (t,x,v)| \lesssim \ |\partial _t p| + \kappa ( |\nabla _x\partial _t u| + \varepsilon |\partial _t u| |\nabla _x u|) + \varepsilon |\partial _t u| |p| , \end{aligned}$$
(3.11)
$$\begin{aligned}&\langle v-\varepsilon u \rangle ^{-2} \Big | \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} \Big | \lesssim \ |\nabla _x u| + \underbrace{\varepsilon |\partial _t u| + \varepsilon |u| |\nabla _x u|}_{(3.12)_*} , \end{aligned}$$
(3.12)
$$\begin{aligned}&\langle v-\varepsilon u \rangle ^{-2}\Big | \partial _t \left( \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} \right) \Big |\nonumber \\&\quad \lesssim \ |\nabla _x \partial _t u| + \underbrace{ \varepsilon \{ |\partial _t^2 u| + |u|| \nabla _x \partial _t u|+| \partial _t u || \nabla _x u| \} + \varepsilon ^2|\partial _t u | (|\partial _t u| + |u| | \nabla _x u|)}_{(3.13)_*} . \end{aligned}$$
(3.13)

Remark 7

We note that due to the choice of (3.1) we remove a contribution of \(p^2\) in \(\frac{\varepsilon }{\delta \kappa } \Gamma (f_2,f_2)\). And also we remark that \({\mathfrak {R}}_4\) is quasi-linear for \(\partial _t^2p\) and \(\nabla _x \partial _t^2 u\).

3.1 Derivatives of \(A_{ij}\) and Commutators in the Local Maxwellian Setting

First we check properties of L and \(\Gamma \) defined in (1.20). Recall the notation of the global Maxwellian \(\mu _0:= M_{1,0,1}\). It is convenient to define

$$\begin{aligned} L_0 f (v) := \frac{-2}{\sqrt{\mu _0 }} Q( \mu _0, \sqrt{\mu _0} {f}) (v), \ \ \Gamma _0 ( {f}, {g}) (v) := \frac{1}{\sqrt{\mu _0 }} Q(\sqrt{\mu _0} {f}, \sqrt{\mu _0} {g}) (v).\nonumber \\ \end{aligned}$$
(3.14)

For a given \(\varepsilon u\), we define \({\tilde{f}}(\cdot ) := f(\cdot + \varepsilon u)\). Then we have

$$\begin{aligned} Lf (v+\varepsilon u)=L_0 {\tilde{f}}(v), \ \ \Gamma (f,g) (v+\varepsilon u)=\Gamma _0 ({\tilde{f}}, {\tilde{g}} )(v). \end{aligned}$$
(3.15)

As in (1.23) a null space of \(L_0\), denoted by \({\mathcal {N}}_0\), is a subspace of \(L^2({\mathbb {R}}^3)\) spanned by orthonormal bases \(\{ {\tilde{\varphi }}_i \sqrt{\mu _0 }\}_{i=0}^4\) with

$$\begin{aligned} \begin{aligned} {\tilde{\varphi }}_0 := 1 , \ \ \ {\tilde{\varphi }}_i: = {v_i } \ \ \text {for} \ i=1,2,3 , \ \ \ {\tilde{\varphi }}_4: = ( | {v } |^2-3 )/{\sqrt{6}}. \end{aligned} \end{aligned}$$
(3.16)

We denote a projection \(\tilde{{\mathbf {P}}}\) on \({\mathcal {N}}_0\) as in (1.24). From standard properties of \(L_0\) and (3.15), we can easily deduce the corresponding properties of L, namely the null space in (1.23), the spectral gap estimate in (1.25), and the existence of a unique inverse \(L^{-1}: {\mathcal {N}}^\perp \rightarrow {\mathcal {N}}^\perp \) in (3.17) which is defined via \(L^{-1}_0: {\mathcal {N}}_0^\perp \rightarrow {\mathcal {N}}_0^\perp \) with the identity

$$\begin{aligned} (L^{-1} f) (v) =( L^{-1}_0 {\tilde{f}}) (v-\varepsilon u). \end{aligned}$$
(3.17)

The inverse enjoys the following bound which turns out useful to prove Lemma 3.

Lemma 1

For \(0< \varrho < \frac{1}{4}\) and \(g \in {\mathcal {N}}_0^\perp \)

$$\begin{aligned} \Vert \nu _0(v)e^{\varrho |v|^2}L_0^{-1} g (v)\Vert _{L^\infty _v} \lesssim \Vert e^{\varrho |v|^2} g(v) \Vert _{L^\infty _v} + \Vert \nu _0(v)^{-1} e^{\varrho |v|^2} g(v) \Vert _{L^2_v}. \end{aligned}$$
(3.18)

The proof is based on the well-known decomposition of \(L_0= \nu _0 -K_0\) and the compactness of \(K_0\): We first recall a standard decomposition

$$\begin{aligned} \begin{aligned} L_0 g(v) =&\ \nu _0(v) g (v) - K_0 g(v)\\ :=&\iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}}| \mu _0(v_*) \mathrm {d}{\mathfrak {u}} \mathrm {d}v_* g(v)\\&- \frac{1}{\sqrt{\mu _0 (v)}} \iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}}|\big \{ \mu _0 (v) \sqrt{\mu _0 (v_*)} g(v_*) \\&-\mu _0 (v^\prime ) \sqrt{\mu _0 (v_*^\prime ) } g(v_*^\prime )-\mu _0 (v^\prime _*) \sqrt{\mu _0 (v^\prime ) } g(v^\prime ) \big \} \mathrm {d}v_* , \end{aligned} \end{aligned}$$
(3.19)

where \( \langle v\rangle \lesssim \nu _0 (v) \lesssim \langle v\rangle \). For (1.21) we have \(\nu (v)= \nu _0 (v-\varepsilon u)\) and \({\mathbf {k}}(v,v_*) = {\mathbf {k}}_0 (v-\varepsilon u, v_*-\varepsilon u)\). It is well-known (see (3.50) and (3.52) in [16]) that one can write \(K_0 g (v)= \int _{{\mathbb {R}}^3} {\mathbf {k}}_0 (v,v_*) g(v_*) \mathrm {d}v_*\) such that for some constants \(C_1, C_2>0\)

$$\begin{aligned} {\mathbf {k}}_0(v,v_*) = C_1 |v-v_*| e^{- \frac{|v|^2 + |v_*|^2}{4}}- \frac{C_{2}}{|v-v_*|} e^{- \frac{|v-v_*|^2}{8} - \frac{1}{8} \frac{(|v|^2 - |v_*|^2)^2}{|v-v_*|^2}}. \end{aligned}$$
(3.20)

It is convenient to introduce a new notation, for \(\vartheta >0\),

$$\begin{aligned} k_{\vartheta }(v,v_*):= \frac{1}{|v-v_*|} e^{- \vartheta {|v-v_*|^2} - \vartheta \frac{(|v|^2 - |v_*|^2)^2}{|v-v_*|^2}}. \end{aligned}$$
(3.21)

Clearly \(|{\mathbf {k}}_0 (v,v_*)| \lesssim k_{\vartheta } (v,v_*)\) for \(0<\vartheta \le 1/8\).

Standard compactness estimates read as follows:

Lemma 2

For \(0<\varrho <2 \vartheta \) and \(C \in {\mathbb {R}}^3\), there exists \(C_{\varrho ,\vartheta }>0\) such that

$$\begin{aligned} \Big | {k}_\vartheta (v,v_*) \frac{e^{\varrho |v|^2+ C \cdot v}}{e^{\varrho |v_*|^2 + C \cdot v_*}} \Big | \lesssim \frac{1}{|v-v_*|} e^{-C_{\varrho } \frac{|v-v_*|^2}{2} } \ \ \text {for} \ 0< \varrho < 2 \vartheta . \end{aligned}$$
(3.22)

Moreover

$$\begin{aligned} \begin{aligned}&\int _{{\mathbb {R}}^3} (1+ |v-v_*|) k_{\vartheta }(v,v_*) \frac{e^{\varrho |v|^2 + C \cdot v }}{e^{\varrho |v_*|^2 + C \cdot v_* }} \mathrm {d}v_*\lesssim _{ \vartheta ,\varrho } \frac{1}{1+|v|} ,\\&\int _{{\mathbb {R}}^3} \frac{1}{|v-v_*| } k_{\vartheta }(v,v_*) \frac{e^{\varrho |v|^2 + C \cdot v }}{e^{\varrho |v_*|^2 + C \cdot v_* }} \mathrm {d}v_*\lesssim _{ \vartheta ,\varrho } 1, \end{aligned} \end{aligned}$$
(3.23)

while the same bounds replacing |v| with \(|v_*|\) hold for integrations over v.

The proof of (3.22) relies on a fact that the exponent has a majorant \(- \vartheta {|v-v_*|^2} - \vartheta \frac{(|v|^2 - |v_*|^2)^2}{|v-v_*|^2}\le -2 \vartheta (|v|+|v_*|) ||v|- |v_*||\) which is a negative definite. Note that an exponent of \(\frac{e^{\varrho |v|^2 }}{e^{\varrho |v_*|^2 }}\) equals \({\varrho (|v|+ |v_*|) ||v| - |v_*|| } \) which can be absorbed as long as \(0< \varrho < 2 \vartheta \). This yields (3.22). We refer to a proof of Lemma 5 in [20] for details to show (3.23).

Proof of Lemma 1

We consider an operator \(g(v) \mapsto \nu _0^{-1} L_0 g (v):= \frac{1}{\nu _0(v)}L_0 g(v)\) on a restricted space of \(\{ g \in L^2({\mathbb {R}}^3): e^{\varrho |v|^2} g(v) \in L^2({\mathbb {R}}^3)\}\). First we claim that

$$\begin{aligned}&\nu _0^{-1} L_0: \{ g \in L^2({\mathbb {R}}^3): e^{\varrho |v|^2} g(v) \in L^2({\mathbb {R}}^3)\} \nonumber \\&\quad \rightarrow \{ g \in L^2({\mathbb {R}}^3): e^{\varrho |v|^2} g(v) \in L^2({\mathbb {R}}^3)\}. \end{aligned}$$
(3.24)

From (3.19) we have \(\nu _0^{-1} L_0 g (v)= g(v) - \nu _0^{-1} e^{-\varrho |v|^2}\int _{{\mathbb {R}}^3} {\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} e^{\varrho |v_*|^2}g(v_*) \mathrm {d}v_*\), and, using (3.23), for \(\varrho < 2 \vartheta \le 1/4\),

$$\begin{aligned}&|e^{\varrho |v|^2} \nu _0^{-1} L_0 g (v)|\\&\quad \le |e^{\varrho |v|^2} g(v)| \\&\qquad + \nu _0(v)^{-1}\sup _v \sqrt{ \int _{{\mathbb {R}}^3} k_{\vartheta } (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}}\mathrm {d}v_*} \sqrt{ \int _{{\mathbb {R}}^3} k_{\vartheta } (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} | e^{\varrho |v_*|^2} g(v_*)|^2 \mathrm {d}v_*}\\&\quad \lesssim |e^{\varrho |v|^2} g(v)| + \sqrt{ \int _{{\mathbb {R}}^3} k_{\vartheta } (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} | e^{\varrho |v_*|^2} g(v_*)|^2 \mathrm {d}v_*}. \end{aligned}$$

Therefore we prove (3.24) from

$$\begin{aligned} \begin{aligned}&\Vert e^{\varrho |v|^2} \nu _0^{-1} L_0 g (v)\Vert _{L^2_v} \\&\quad \lesssim \Vert e^{\varrho |v|^2} g (v)\Vert _{L^2_v} + \sqrt{ \sup _{v_*} \int _{{\mathbb {R}}^3} k_\vartheta (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} \mathrm {d}v \int _{{\mathbb {R}}^3} |e^{\varrho |v_*|^2} g(v_*)|^2 }\\&\quad \lesssim \Vert e^{\varrho |v|^2} g (v)\Vert _{L^2_v} . \end{aligned} \end{aligned}$$
(3.25)

Now we view \(\{ g \in L^2({\mathbb {R}}^3): e^{\varrho |v|^2} g(v) \in L^2({\mathbb {R}}^3)\}\) as the Hilbert space with an inner product \(\langle e^{\varrho |v|^2} \cdot , e^{\varrho |v|^2} \cdot \rangle \). Then the compactness of \(\nu _0^{-1}K_0\) in this space is equivalent to the compactness of \(g \mapsto \int _{{\mathbb {R}}^3} {\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} g(v_*) \mathrm {d}v_*\) in a usual \(L^2_v\). From Lemma 3.5.1 of [16], it suffices to prove that (i) \(\int _{{\mathbb {R}}^3} {\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} \mathrm {d}v\) is bounded in \(v_*\), (ii) \({\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} \in L^2 ( \{|v-v_*| \ge \frac{1}{n} \ \text {and}\ |v| \le n\})\) for all \(n\in {\mathbb {N}}\), and (iii) \(\sup _{v}\int _{{\mathbb {R}}^3} {\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} \{ {\mathbf {1}}_{|v-v_*| \le \frac{1}{n}} + {\mathbf {1}}_{|v|\ge n} \}\mathrm {d}u\rightarrow 0\) as \(n\rightarrow \infty \). Both conditions (i) and (ii) come from the first bound of (3.23) directly. We prove (iii) from (3.22) and the first bound of (3.23). Now applying the Fredholm alternative to \(\nu _0^{-1} L_0=id -\nu _0^{-1}K_0 \) in the Hilbert space, we obtain an inverse map \((\nu _0^{-1} L_0)^{-1}\) which is a bounded operator of the Hilbert space. Note that \(L_0^{-1} (g)= (\nu _0^{-1} L_0)^{-1} (\nu _0^{-1} g)\). Hence we derive that

$$\begin{aligned} \Vert e^{\varrho |v|^2}L_0^{-1} g \Vert _{L^2_v} = \Vert e^{\varrho |v|^2} (\nu _0^{-1} L_0)^{-1} (\nu _0^{-1} g)\Vert _{L^2_v} \lesssim \Vert e^{\varrho |v|^2} \nu _0^{-1}g \Vert _{L^2_v} . \end{aligned}$$
(3.26)

From the decomposition of \(L_0\), we have \(L^{-1}_0 g(v)= \nu _0(v)^{-1} g(v) + \nu _0(v)^{-1} KL^{-1}_0 g(v)\) for \(g \in {\mathcal {N}}_0^\perp \). Then we have

$$\begin{aligned}&|e^{\varrho |v|^2} L_0^{-1} g(v)| \\&\quad \le | \nu _0(v)^{-1}e^{\varrho |v|^2} g (v)| + \Big | \nu _0(v)^{-1} \int _{{\mathbb {R}}^3} {\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} e^{\varrho |v_*|^2} L^{-1}_0 g(v_*) \mathrm {d}v_* \Big |\\&\quad \le \nu _0(v)^{-1} \bigg \{ | e^{\varrho |v|^2} g (v)| + \sqrt{\int _{{\mathbb {R}}^3} \Big | {\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}}\Big |^2 \mathrm {d}v_*} \sqrt{\int _{{\mathbb {R}}^3} | e^{\varrho |v_*|^2} L^{-1}_0 g (v_*)|^2 \mathrm {d}v_* } \bigg \}, \end{aligned}$$

while \(\big |{\mathbf {k}}_0 (v,v_*) \frac{e^{\varrho |v|^2}}{e^{\varrho |v_*|^2}} \big |^2 \lesssim \frac{1}{|v-v_*|^2} e^{-2C_{\varrho } \frac{|v-v_*|^2}{2} } \in L^{\infty }_{v} L^{1}_{v_*}\) from (3.22). Hence we prove (3.18). \(\square \)

Equipped with Lemma 1 we provide bounds of \(A_{ij}\) in (1.34) and its derivatives:

Lemma 3

For \(0< \varrho < \frac{1}{4}\)

$$\begin{aligned} \begin{aligned} |A_{ij}(v)|&\lesssim e^{- \varrho |v-\varepsilon u|^2}, \ |\nabla _x\\ A_{ij}(v)|&\lesssim \varepsilon |\nabla _x u|e^{- \varrho |v-\varepsilon u|^2} , \ |\partial _t A_{ij}(v)|\lesssim \varepsilon |\partial _t u|e^{- \varrho |v-\varepsilon u|^2}, \\ |\nabla _x \partial _t A_{ij}(v)|&\lesssim \varepsilon \{ |\nabla _x \partial _t u| + \varepsilon |\nabla _x u| |\partial _t u| \} e^{- \varrho |v-\varepsilon u|^2}. \end{aligned} \end{aligned}$$
(3.27)

Proof

It is convenient to introduce a notation, with \(L_0\) in (3.14),

$$\begin{aligned} {A}_{0,ij}(v):= L^{-1}_0\Big ((v_i v_j - \frac{|v|^2}{3} \delta _{ij} )\sqrt{\mu _0}\Big )(v). \end{aligned}$$
(3.28)

Then from (3.18) and (3.17) we can immediately prove the first bound in (3.27).

Recall the notations in (3.14) and (3.15). By taking a derivative to \(L_0\) (3.28), it follows that, from the decomposition of \(L_0A_{0,ij} (v) = \, \nu _0(v) A_{0,ij} (v)- \int _{{\mathbb {R}}^3} {\mathbf {k}}_0 (v, v-v_*) A_{0,ij} (v-v_*) \mathrm {d}v_* \) and (3.20),

$$\begin{aligned} \begin{aligned} L_0 \partial _{v_k} A_{0,ij} =\,\,&\partial _{v_k} (v_i v_j - \frac{|v|^2}{3}) \sqrt{\mu _0} + (v_i v_j - \frac{|v|^2}{3}) \partial _{v_k} \sqrt{\mu _0}\\&- \Big \{ \partial _{v_k} \nu _0(v) A_{0,ij}(v) - \int _{{\mathbb {R}}^3} \partial _{v_k} [ {\mathbf {k}}_0 (v,v-v_*)] A_{0,ij} (v-v_*)\mathrm {d}v_* \Big \}. \end{aligned}\nonumber \\ \end{aligned}$$
(3.29)

From (3.20) and \(\nabla _v(|v|^2- |v-v_*|^2)^2= 4 v_{*} (|v| + |v-v_*|) (|v- |v-v_*|)\), it follows \(|\nabla _v [ {\mathbf {k}}_0 (v,v-v_*)]| \lesssim |v_*| \exp \{- \frac{ |v-v_*|^2+ |v*|^2}{8}\} + \frac{1}{|v_*|} \exp \{- \frac{|v_*|^2}{8} - \frac{1}{8} \frac{(|v |^2 - |v-v_*|^2)^2}{| v_*|^2}\}\). From the first bound of (3.23), it follows \(|\int _{{\mathbb {R}}^3} \partial _{v_k} [ {\mathbf {k}}_0 (v,v-v_*)] A_{0,ij} (v-v_*)\mathrm {d}v_*| \lesssim e^{-\varrho |v|^2}\) for any \(0< \varrho < 1/4\). Recall a projection \(\tilde{{\mathbf {P}}}\) on \({\mathcal {N}}_0\). Then \(|({\mathbf {I}}- \tilde{{\mathbf {P}}}) \text { r.h.s. of }(3.29)| \lesssim e^{-\varrho |v|^2}\). Now applying (3.18) to \( \partial _{v_k} A_{0,ij}= L_0^{-1} (({\mathbf {I}}- \tilde{{\mathbf {P}}}) \text { r.h.s. of }\)(3.29)) we derive

$$\begin{aligned} | \nabla _v A_{0,ij} (v)| \lesssim e^{-\varrho |v|^2} \text { for any } 0< \varrho < 1/4. \end{aligned}$$
(3.30)

From (1.34) and (3.17), and the fact \({\tilde{\varphi }}_i=v_i\) for \(i=1,2,3\), and \({\tilde{\varphi }}_4=\frac{|v|^2}{3}\) (the notation \({\tilde{f}}\) is defined in (3.15)), we have

$$\begin{aligned} {A}_{ij}(v) = L^{-1}_0\Big ((v_i v_j - \frac{|v|^2}{3} \delta _{ij} )\sqrt{\mu _0}\Big )(v-\varepsilon u)= {A}_{0,ij}(v-\varepsilon u). \end{aligned}$$
(3.31)

Therefore we prove the second and third bounds in (3.27) using the fact that \(\nabla _{x,t} A_{ij}(v)= -\varepsilon \nabla _{x,t} u \nabla _v A_{0,ij}(v-\varepsilon u)\).

Now we prove

$$\begin{aligned} |\nabla _v^2 A_{0,ij} (v)| \lesssim e^{- \varrho |v|^2} . \end{aligned}$$
(3.32)

By taking one more derivative to (3.29), we derive that

$$\begin{aligned}&L_0 \partial _{v_k}\partial _{v_\ell } A_{0,ij}\\&\quad = \ \partial _{v_k} \partial _{v_\ell } (v_i v_j - \frac{|v|^2}{3}) \sqrt{\mu _0} + \partial _{v_\ell } (v_i v_j - \frac{|v|^2}{3}) \partial _{v_k} \sqrt{\mu _0} \\&\qquad + (v_i v_j - \frac{|v|^2}{3}) \partial _{v_\ell } \partial _{v_k} \sqrt{\mu _0} \\&\qquad - \partial _{v_k} \partial _{v_\ell } \nu _0(v) A_{0,ij}(v)- \partial _{v_k} \nu _0(v) \partial _{v_\ell }A_{0,ij}(v)\\&\qquad +\int _{{\mathbb {R}}^3} \partial _{v_k} \partial _{v_\ell }[ {\mathbf {k}}_0 (v,v-v_*)] A_{0,ij} (v-v_*) \nonumber \\&\qquad + \partial _{v_k} [ {\mathbf {k}}_0 (v,v-v_*)] \partial _{v_\ell } [A_{0,ij} (v-v_*)] \mathrm {d}v_* . \end{aligned}$$

The terms in the first two lines in r.h.s are easily bounded above as \(e^{-\varrho |v|^2}\), recalling the fact \(|\nabla _v\nu _0(v)| + |\nabla ^2_v\nu _0(v)| \lesssim 1\). We only focus on the terms in the last line. From \(|\partial _{v_\ell } \nabla _v (|v|^2 - |v-v_*|^2)^2| \le \big |4 v_* \Big ( \frac{v_\ell }{|v|} + \frac{(v-v_*)_\ell }{|v-v_*|} \Big )(|v| - |v-v_*|) \big | + 4 \big | v_* (|v| + |v-v_*|)\Big ( \frac{v_\ell }{|v|} - \frac{(v-v_*)_\ell }{|v-v_*|} \Big )\big |\lesssim |v_*|^2 + |v_*||v|, \) we have \(| \partial _{v_\ell }\nabla _v [ {\mathbf {k}}_0 (v,v-v_*)]| \lesssim |v_*| \exp \{- \frac{ |v-v_*|^2+ |v*|^2}{8}\} + \frac{1+|v|}{|v_*|^2} \exp \{- \frac{|v_*|^2}{8} - \frac{1}{8} \frac{(|v |^2 - |v-v_*|^2)^2}{| v_*|^2}\}\). Using the second estimate of (3.23) with the first bound of (3.27), we have \(\big |\int _{{\mathbb {R}}^3} \partial _{v_k} \partial _{v_\ell }[ {\mathbf {k}}_0 (v,v-v_*)] A_{0,ij} (v-v_*) \mathrm {d}v_*\big |\lesssim e^{- \varrho |v |^2}\). From (3.30) and the first bound of (3.27), it follows that \(\big |\int _{{\mathbb {R}}^3} \partial _{v_k} [ {\mathbf {k}}_0 (v,v-v_*)] \partial _{v_\ell } [A_{0,ij} (v-v_*)] \mathrm {d}v_*\big | \lesssim e^{- \varrho |v |^2}\). Now we invert the operator \(L_0\) and use (3.18) to conclude (3.32).

Finally from \(\partial _t \nabla _x A_{ij} (v) = -\varepsilon \partial _{t} \nabla _x u \nabla _v A_{0,ij} (v-\varepsilon u) + \varepsilon ^2 \nabla _x u \partial _t u \nabla _v^2 A_{0,ij} (v-\varepsilon u)\), (3.30), and (3.32), we conclude the last estimate of (3.27). \(\square \)

For the estimates of \(\partial _t f_R\) we derive the commutator estimate of \(\partial _t L- L\partial _t \) and the corresponding one for \(\Gamma \) as follows.

Lemma 4

Suppose \(\varepsilon |u|\lesssim 1\) in the definition of \(\mu \) in (1.16). For L and \(\Gamma \) in (1.20) and (1.21),

$$\begin{aligned} \begin{aligned} \partial _t (Lf)&=L\partial _tf + L_t( {\mathbf {I}}-{\mathbf {P}})f - L ({\mathbf {I}} - {\mathbf {P}})( {\mathbf {P}}_t f),\\ \partial _t (\Gamma (f,g))&= \{\Gamma (\partial _t f, g) + \Gamma (f,\partial _t g)\} + \Gamma _t (f,g), \end{aligned} \end{aligned}$$
(3.33)

where

$$\begin{aligned} \begin{aligned} L_t g (t,v)&:= - \varepsilon \partial _t u \cdot \nabla _v \nu _0 (v-\varepsilon u ) g(t, v)\\&\quad \quad + \varepsilon \partial _t u \cdot \int _{{\mathbb {R}}^3} (\nabla _v {\mathbf {k}}_0 + \nabla _{v_*} {\mathbf {k}}_0 ) (v-\varepsilon u, v_* - \varepsilon u)g(t,v_*)\mathrm {d}v_*, \\ ({\mathbf {I}} -{\mathbf {P}}){\mathbf {P}}_t g&:=- \varepsilon \sum _{j=0}^4 ( {P}_j g) ({\mathbf {I}} -{\mathbf {P}})\big ( \partial _tu \cdot \nabla _v( \varphi _j \sqrt{\mu })\big ) ,\\ \Gamma _t(f,g) (t,v)&:=\frac{\varepsilon }{2} \iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}}| \partial _t u \cdot (v_*- \varepsilon u) \sqrt{ \mu (v_* )} \big \{ f(t,v ^\prime ) g(t,v_*^\prime )\\&\qquad \qquad + g(t,v ^\prime ) f(t,v_*^\prime ) \!-\! f(t,v ) g(t,v_* )\!-\! g(t,v ) f(t,v_* ) \big \} \mathrm {d}{\mathfrak {u}} \mathrm {d}v_*. \end{aligned} \end{aligned}$$
(3.34)

We have

$$\begin{aligned} \begin{aligned}&\bigg |\int _{{\mathbb {R}}^3} L_t ({\mathbf {I}} - {\mathbf {P}})f(v) g(v) \mathrm {d}v\bigg | \lesssim \varepsilon |\partial _t u| \Vert \nu ^{1/2} ({\mathbf {I}} -{\mathbf {P}}) f \Vert _{L^2_v} \Vert \nu ^{1/2} g \Vert _{L^2_v} ,\\&\bigg |\int _{{\mathbb {R}}^3} L({\mathbf {P}}_t f) (v) g(v)\mathrm {d}v\bigg | \lesssim \varepsilon |\partial _t u| |Pf| \Vert \nu ^{1/2} ({\mathbf {I}} -{\mathbf {P}}) g \Vert _{L^2_v},\\&\bigg |\int _{{\mathbb {R}}^3}\Gamma _t (f,g)(v) h(v) \mathrm {d}v \bigg | \\&\quad \lesssim \varepsilon |\partial _t u| \Big ( \Vert e^{\varrho |v|^2 + C \cdot v} g\Vert _{L^\infty _v} \Vert \nu ^{1/2} ({\mathbf {I}}- {\mathbf {P}}) f \Vert _{L^2_v}\\&\qquad + \Vert e^{\varrho |v|^2 + C \cdot v} f\Vert _{L^\infty _v} \Vert \nu ^{1/2} ({\mathbf {I}}- {\mathbf {P}}) g \Vert _{L^2_v} + |Pf||Pg| \Big ) \Vert \nu ^{1/2} h \Vert _{L^2_v} . \end{aligned} \end{aligned}$$
(3.35)

Pointwise estimates are given as follows: for \(0< \varrho < 1/4\) and \(C \in {\mathbb {R}}^3\)

$$\begin{aligned} \begin{aligned}&| L_t( {\mathbf {I}}-{\mathbf {P}})f(t, v) - L( {\mathbf {P}}_t f)(t, v) | \\&\quad \lesssim \varepsilon |\partial _t u| \Vert e^{ \varrho |v|^2 + C \cdot v} f(t,v) \Vert _{L^\infty _v} \nu (v)^2 e^{- \varrho |v-\varepsilon u |^2},\\&\quad | \Gamma _t(f,g) (t,v) | \lesssim \varepsilon |\partial _t u| \Vert e^{\varrho |v|^2+ C \cdot v} f(t,v) \Vert _{L^\infty _v} \Vert e^{\varrho |v|^2 + C \cdot v }g(t,v) \Vert _{L^\infty _v} \frac{\nu (v)}{e^{ \varrho |v |^2+ C \cdot v } },\\&\quad \left| \Gamma (f,g)(v)\right| \lesssim \Vert e^{\varrho |v|^{2}+ C \cdot v} f(v) \Vert _{L^{\infty }_{v}}\Vert e^{\varrho |v|^{2}+ C \cdot v} g(v) \Vert _{L^{\infty }_{v}} \frac{\nu (v)}{e^{ \varrho |v |^2+ C \cdot v } } , \end{aligned}\nonumber \\ \end{aligned}$$
(3.36)

and

$$\begin{aligned} |\Gamma (f, g)(v)| \lesssim \Vert e^{\varrho |v|^2+ C \cdot v} f \Vert _\infty \Big ( \nu (v)| g(v)| + \int _{{\mathbb {R}}^3} k_{\vartheta }(v,v_*) |g(v_*)| \mathrm {d}v_* \Big ).\quad \end{aligned}$$
(3.37)

Proof

The decomposition (3.33) with (3.34) comes from a direct computation to (1.21) and \( \partial _t (Lf_R) = \partial _t (L( {\mathbf {I}}-{\mathbf {P}})f_R) =L ( {\mathbf {I}}-{\mathbf {P}})\partial _tf_R + L_t( {\mathbf {I}}-{\mathbf {P}})f_R + L (-{\mathbf {P}}_t f_R). \) On the other hand, from (3.20) it is easy to check that, for any \(0< \varrho < 1/4\),

$$\begin{aligned}&|\nabla _v \nu _0 (v)| = \Big | \iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} {\mathfrak {u}} \frac{(v-v_*) \cdot {\mathfrak {u}}}{|(v-v_*) \cdot {\mathfrak {u}}|}\mu _0 (v_*) \mathrm {d}{\mathfrak {u}} \mathrm {d}v_* \Big |\lesssim 1,\\&\quad |\nabla _v {\mathbf {k}}_0 (v,v_*)| + |\nabla _{v_*} {\mathbf {k}}_0 (v,v_*)| \lesssim \Big (|v-v_*|^{-1}+ \nu _0 (v)^2 |v-v_*| \Big ) k_{\varrho /2} (v,v_*). \end{aligned}$$

These estimates above combining with (3.23) and \({\mathbf {P}}L=0\) yield the first two estimates of (3.35).

We derive \(\Gamma _t\) as in the third identity of (3.34) from a direct computation to (1.22) and \( \partial _t \sqrt{\mu (v_*)} = - \varepsilon \partial _t u \cdot \nabla _v \sqrt{\mu _0 (v_* - \varepsilon u )} = \frac{1}{2} \varepsilon \partial _t u \cdot (v_*-\varepsilon u ) \sqrt{\mu _0 (v_* - \varepsilon u )} . \) Then it is standard (see Lemma 2.13 in [13] for example) to have the last estimate in (3.35).

The first bound of (3.36) is a direct consequence of applying (3.22) and (3.23) to the first identity of (3.34). For the second bound of (3.36) we bound it as

$$\begin{aligned}&\varepsilon |\partial _t u | \Vert e^{\varrho |v|^2+ C \cdot v} f(t,v) \Vert _{L^\infty _v} \Vert e^{\varrho |v|^2+ C \cdot v} g(t,v) \Vert _{L^\infty _v} \frac{1}{e^{ \varrho |v |^2+ C \cdot v } } \\&\qquad \iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}}| e^{- \varrho |v_* |^2- C \cdot v_*} \mathrm {d}{\mathfrak {u}} \mathrm {d}v_*\\&\quad \quad \lesssim \ \frac{\nu (v)}{e^{ \varrho |v |^2+ C \cdot v } }\varepsilon |\partial _t u | \Vert e^{\varrho |v|^2+ C \cdot v} f(t,v) \Vert _{L^\infty _v} \Vert e^{\varrho |v|^2+ C \cdot v} g(t,v) \Vert _{L^\infty _v} \end{aligned}$$

For the last bound of (3.36) we recall a standard estimate (e.g. [16]) that

$$\begin{aligned} |\frac{1}{\nu _{0}(v)} e^{\varrho |v|^{2}+ C \cdot v} \Gamma _{0}(f,g)(v)|\lesssim \Vert e^{\varrho |v|^{2}+ C \cdot v} f(v) \Vert _{L^{\infty }_{v}}\Vert e^{\varrho |v|^{2}+ C \cdot v} g(v) \Vert _{L^{\infty }_{v}} . \end{aligned}$$

From the second equality of (3.15) we deduce the last bound of (3.36). A bound (3.37) is standard. \(\square \)

3.2 Proof of Proposition 6

We verify two statements of Section 1.1 Hilbert Expansion. Firstly, we will show that the solvability condition (1.33) implies the incompressible condition (1.14). From (1.16), (1.23), and direct computations we verify the first identity of (1.33). Then from the oddness of the integrand with respect to the variable \(\varphi _i\) we derive that \(\Big \langle \varphi _i \sqrt{\mu }, \frac{ \varepsilon ^{-1} (v- \varepsilon u )\cdot \nabla _x \mu }{\sqrt{\mu }} \Big \rangle =0\) for \(i=1,2,3\). For \(i=0,4\), we compute that

$$\begin{aligned}&\Big \langle \varphi _i \sqrt{\mu }, \frac{ \varepsilon ^{-1} (v- \varepsilon u )\cdot \nabla _x \mu }{\sqrt{\mu }} \Big \rangle \\&\quad \quad = \sum _{\ell =1}^3 \langle \varphi _i \sqrt{\mu }, \varphi _\ell \varphi _\ell \sqrt{\mu } \rangle \partial _\ell u_\ell = \left\{ \delta _{i0} + \delta _{i4}\sqrt{\frac{2}{3}} \right\} ( \nabla _x \cdot u) \ \ \text {for} \ i=0,4. \end{aligned}$$

This shows that (1.33) implies (1.14).

Secondly, we will verify the following statement of Section 1.1: the leading order terms of the hydrodynamic part in (1.28) vanish by solving the Navier-Stokes equations (1.13)-(1.15). Consider (1.28). We set \({\mathbf {P}}f_2 =\{ {\tilde{\rho }} \varphi _0+ \sum _{\ell =1}^3 {\tilde{u}}_\ell \varphi _\ell +{\tilde{\theta }} \varphi _4 \}\sqrt{\mu }\) whose coefficients will be determined as in (3.1). Then the leading order term of (1.28)= \(\frac{1}{\delta }\) (1.35) can be decomposed as

$$\begin{aligned}&- \frac{1}{\delta } {\mathbf {P}} \Big ( (v- \varepsilon u) \cdot (\partial _t u + u \cdot \nabla _x u ) \sqrt{\mu } \nonumber \\&\quad + \underbrace{ (v- \varepsilon u) \cdot \big ( \nabla _x {\tilde{\rho }} \varphi _0 \sqrt{\mu }- \sum _{\ell =1}^3 \nabla _x {\tilde{u}}_\ell \varphi _\ell \sqrt{\mu } + \nabla _x{\tilde{\theta }} \varphi _4 \sqrt{\mu } \big )}_{ (3.38)_* }\nonumber \\&\quad - \sum _{\ell ,m=1}^3 \kappa (v- \varepsilon u) \cdot A_{\ell m} \nabla _x\partial _\ell u_m \Big ), \end{aligned}$$
(3.38)
$$\begin{aligned}&\quad - \frac{1}{\delta } ({\mathbf {I}}-{\mathbf {P}}) \Big ( (v- \varepsilon u) \cdot \big ( \nabla _x {\tilde{\rho }} \varphi _0 \sqrt{\mu }+ \sum _{\ell =1}^3 \nabla _x {\tilde{u}}_\ell \varphi _\ell \sqrt{\mu } + \nabla _x{\tilde{\theta }} \varphi _4 \sqrt{\mu } \big ) \Big )\nonumber \\&\quad + \frac{1}{\delta } ({\mathbf {I}} - {\mathbf {P}})\Big ( \sum _{\ell ,m=1}^3 \kappa (v- \varepsilon u) \cdot A_{\ell m} \nabla _x\partial _\ell u_m\Big ), \end{aligned}$$
(3.39)

while the lower order term consists of

$$\begin{aligned} \begin{aligned}&-\frac{1}{\delta } (v- \varepsilon u) \cdot \nabla _x \Big ( {\tilde{\rho }} \varphi _0 \sqrt{\mu }+ \sum _{\ell =1}^3 {\tilde{u}}_\ell \varphi _\ell \sqrt{\mu } + {\tilde{\theta }} \varphi _4 \sqrt{\mu } \Big ) \\&\quad +\frac{1}{\delta } (v- \varepsilon u) \cdot \Big ( \nabla _x {\tilde{\rho }} \varphi _0 \sqrt{\mu }+ \sum _{\ell =1}^3 \nabla _x {\tilde{u}}_\ell \varphi _\ell \sqrt{\mu } + \nabla _x{\tilde{\theta }} \varphi _4 \sqrt{\mu } \Big ) \\&\quad + \frac{\kappa }{\delta }\sum _{\ell ,m=1}^3(v-\varepsilon u) \cdot \nabla _x A_{\ell m} \partial _\ell u_m. \end{aligned} \end{aligned}$$
(3.40)

First we focus on a leading order contribution of (3.38)\(_*\) in (3.38). A direct computation yields

$$\begin{aligned} \begin{aligned}&\big \langle \varphi _i \sqrt{\mu }, (3.38)_* \big \rangle \\&\quad = {\left\{ \begin{array}{ll} C_i \nabla _x \cdot {\tilde{u}}, \quad i =0,4\\ \partial _i {\tilde{\rho }} \langle \varphi _i \sqrt{\mu }, \varphi _i \sqrt{\mu }\rangle + \partial _i {\tilde{\theta }} \langle \varphi _i\sqrt{\mu }, \varphi _i \varphi _4 \sqrt{\mu }\rangle = \partial _i \Big ( {\tilde{\rho }} + \sqrt{\frac{2}{3} }{\tilde{\theta }} \Big ), \quad i=1,2,3. \end{array}\right. } \end{aligned}\nonumber \\ \end{aligned}$$
(3.41)

Among many other choices we make a special choice \(({\tilde{\rho }}, {\tilde{u}}, {\tilde{\theta }})=(p,0,0)\) which is equivalent to (3.1). From (1.38), (3.41), and (3.1), it follows that for (up) solving (1.13)

$$\begin{aligned} (3.38)= & {} \frac{1}{\delta } (v- \varepsilon u ) \sqrt{\mu } \cdot \big \{ \partial _t u+ u \cdot \nabla _x u - \kappa \eta _0 \Delta u + \nabla _x p \big \} \nonumber \\= & {} \frac{1}{\delta } (v- \varepsilon u ) \sqrt{\mu } \cdot (1.13) =0, \end{aligned}$$
(3.42)

which verifies the second statement of Section 1.1.

Now we turn to proving the estimates. While the leading order terms vanish in (1.28), the rest of terms of (1.28) are bounded as follows. Upon the choice of (3.1), the first term of (3.39) vanishes and the first line of (3.40) are bounded by

$$\begin{aligned} \Big | \frac{\varepsilon }{2 \delta } (v-\varepsilon u) \cdot \nabla _x u \cdot (v-\varepsilon u) {\mathbf {P}} f_2\Big | \lesssim \frac{\varepsilon }{\delta } |\nabla _x u| |p| \langle v-\varepsilon u \rangle ^2 \sqrt{\mu }. \end{aligned}$$
(3.43)

From (3.27) we deduce that the second term of (3.39) and the second line of (3.40) are bounded respectively by

$$\begin{aligned} \frac{\kappa }{\delta } |\nabla _x^2 u| |v-\varepsilon u| e^{- \varrho |v-\varepsilon u|^2}, \ \frac{ \varepsilon \kappa }{\delta } |\nabla _x u| |v-\varepsilon u| e^{- \varrho |v-\varepsilon u|^2} \ \ \text {for any } 0< \varrho < 1/4.\nonumber \\ \end{aligned}$$
(3.44)

In conclusion we end up with the following result: Assume (up) solves (1.13)-(1.15), and both (1.34) and (3.1) hold. Then

$$\begin{aligned} |(1.28)-(3.39) | \lesssim \frac{\varepsilon }{\delta } \{ |\nabla _x u| |p| + \kappa |\nabla _x u| \} \langle v-\varepsilon u \rangle ^2 e^{- \frac{|v-\varepsilon u|^2}{4}} , \end{aligned}$$
(3.45)
$$\begin{aligned} | ({\mathbf {I}} - {\mathbf {P}}) (3.39) |= | (3.39) | \lesssim \frac{1}{\delta } \kappa |\nabla _x^2 u| e^{- \varrho {|v-\varepsilon u|^2} } . \end{aligned}$$
(3.46)

The term \(\partial _t (1.28)\) can be bounded similarly. The entire leading order term of \(\partial _t (1.28)\) can be decomposed as

$$\begin{aligned}&- \frac{1}{\delta } {\mathbf {P}} \Big ( (v- \varepsilon u) \cdot \partial _t (\partial _t u + u \cdot \nabla _x u ) \sqrt{\mu } + (v- \varepsilon u) \cdot \Big ( \nabla _x \partial _t p \varphi _0 \sqrt{\mu } \Big ) \nonumber \\&\qquad \qquad \quad - \sum _{\ell ,m=1}^3 \kappa (v- \varepsilon u) \cdot A_{\ell m} \nabla _x\partial _\ell \partial _t u_m \Big ), \end{aligned}$$
(3.47)
$$\begin{aligned}&-\frac{1}{\delta } ({\mathbf {I}}-{\mathbf {P}}) \Big ( (v- \varepsilon u) \cdot ( \nabla _x \partial _t p \varphi _0 \sqrt{\mu } ) \Big ) \nonumber \\&\quad + \frac{1}{\delta } ({\mathbf {I}} - {\mathbf {P}})\Big ( \sum _{\ell ,m=1}^3 \kappa (v- \varepsilon u) \cdot A_{\ell m} \nabla _x\partial _\ell \partial _t u_m\Big ). \end{aligned}$$
(3.48)

Following the argument to get (1.38) and (3.41), we derive that

$$\begin{aligned} (3.47)= -\frac{1}{\delta } (v- \varepsilon u) \sqrt{\mu } \cdot \partial _t (1.13)=0. \end{aligned}$$
(3.49)

On the other hand,

$$\begin{aligned} |(3.48)| \lesssim \frac{1}{\delta } \kappa |\nabla _x ^2 \partial _t u| e^{- \varrho |v-\varepsilon u|^2}. \end{aligned}$$
(3.50)

Now the lower order term \(\partial _t (1.28)- (3.47)- (3.48)\) consists of

$$\begin{aligned} \begin{aligned}&\frac{\varepsilon }{\delta } \partial _t u \cdot \Big \{ (\partial _t u + u \cdot \nabla _x u ) \sqrt{\mu } + \nabla _x {\mathbf {P}} f_2 - \kappa \nabla _x \Big ( \sum _{\ell ,m=1}^3 A_{\ell m} \partial _\ell u_m\Big ) \Big \} \\&\quad + \frac{\kappa }{\delta } (v- \varepsilon u ) \cdot \nabla _x \Big ( \sum _{\ell ,m=1}^3 \partial _t A_{\ell m} \partial _\ell u_m\Big ) \\&\quad - \frac{1}{\delta } (v- \varepsilon u) \cdot \nabla _x \partial _t \big ( p \varphi _0 \sqrt{\mu } \big ) +\frac{1}{\delta } (v- \varepsilon u) \cdot \big ( \nabla _x \partial _t p \varphi _0 \sqrt{\mu } \big ) . \end{aligned} \end{aligned}$$
(3.51)

Since the lower order term of \(\partial _t (1.28)\) always contains \(|\partial _t (v-\varepsilon u)| \le \varepsilon |\partial _t u|\), they can be bounded by, from (3.27) and (3.1),

$$\begin{aligned} \begin{aligned} |(3.51)| \lesssim&\ \frac{\varepsilon }{\delta }|\partial _t u| \big \{ |\partial _t u| + |u| |\nabla _x u| + |\nabla _x p| \\&\qquad \quad \quad + \varepsilon |\nabla _x u| |p| + \kappa \varepsilon |\nabla _x u| ^2 + \kappa |\nabla _x ^2 u| \big \} e^{- \varrho |v-\varepsilon u|^2}\\&+ \frac{ \varepsilon }{\delta } \{ |\nabla _x \partial _t u| (1+ \kappa |\nabla _x u|) + |\partial _t p| |\nabla _x u| \} e^{- \varrho |v-\varepsilon u|^2}. \end{aligned} \end{aligned}$$
(3.52)

Now we consider (1.29). From (1.34), (3.1), and (3.27), we derive that

$$\begin{aligned}&| (\partial _t + u \cdot \nabla _x ) {\mathbf {P}}f_2 | \nonumber \\&\quad \lesssim \Big \{ |\partial _t p| + |u| |\nabla _x p| + \varepsilon |p| \{ |\partial _t u| + |u| |\nabla _x u|\} \Big \} \langle v-\varepsilon u\rangle ^{ 2} e^{ -\frac{|v-\varepsilon u|^2}{4}} , \end{aligned}$$
(3.53)
$$\begin{aligned}&| \partial _t (\partial _t + u \cdot \nabla _x ){\mathbf {P}}f_2 | \nonumber \\&\quad \lesssim \ \Big \{ |\partial _t^2 p| + |\partial _t u| |\nabla _x p|+ | u| |\nabla _x \partial _t p| + \varepsilon | \partial _t p| \{ |\partial _t u| + |u| |\nabla _x u|\} \nonumber \\&\qquad + \varepsilon | p| \{ |\partial _t^2 u| + |\partial _t u| |\nabla _x u|+ |u| |\nabla _x \partial _t u|\}\nonumber \\&\qquad + \varepsilon |\partial _t u| \{\text {r.h.s. of } (3.53) \} \Big \} \langle v-\varepsilon u\rangle ^{ 2} e^{ -\frac{|v-\varepsilon u|^2}{4}}, \end{aligned}$$
(3.54)

and, for \(0< \varrho < 1/4\),

$$\begin{aligned}&| (\partial _t + u \cdot \nabla _x ) ({\mathbf {I}} - {\mathbf {P}}) f_2|\nonumber \\&\quad \lesssim \kappa \Big \{ \{ | \nabla _x \partial _t u| + |u| |\nabla _x^2 u|\} + \varepsilon \{|\partial _t u| + |u| |\nabla _x u|\} |\nabla _x u| \Big \} e^{-\varrho |v-\varepsilon u|^2} , \end{aligned}$$
(3.55)
$$\begin{aligned}&| \partial _t (\partial _t + u \cdot \nabla _x ) ({\mathbf {I}} - {\mathbf {P}}) f_2|\nonumber \\&\quad \lesssim \ \kappa \Big \{ \{ | \nabla _x \partial _t^2 u| + |\partial _t u| |\nabla _x^2 u|+ |u| |\nabla _x^2 \partial _t u|\} + \varepsilon \{|\partial _t u| + |u| |\nabla _x u|\} |\nabla _x \partial _t u|\nonumber \\&\qquad \quad + \varepsilon \{|\partial _t^2 u| + |\partial _t u| |\nabla _x u|+ |u| |\nabla _x \partial _t u|\} |\nabla _x u| \nonumber \\&\qquad \quad + \varepsilon |\partial _t u| \{ \text {r.h.s. of } (3.55) \} \Big \} e^{-\varrho |v-\varepsilon u|^2} . \end{aligned}$$
(3.56)

Next we consider the last term in (1.29). From (1.16) and (1.14)

$$\begin{aligned} \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }}&= \frac{1}{2} \big [ (v-\varepsilon u) \cdot \nabla _x u \cdot (v-\varepsilon u) \\&\quad + \varepsilon ( \partial _t u + u \cdot \nabla _x u) \cdot (v-\varepsilon u) \big ] ,\\ \partial _t \left( \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} \right)&= \frac{1}{2} \big [ \varepsilon (\partial _t^2 u + u \nabla _x \partial _t u- \partial _t u \cdot \nabla _x u) \cdot (v-\varepsilon u)\\&\quad - \varepsilon ^2 \partial _t u \cdot (\partial _t u + u \cdot \nabla _x u) \\&\quad + (v-\varepsilon u) \cdot \nabla _x\partial _t u \cdot (v-\varepsilon u) \big ] , \end{aligned}$$

and hence we derive (3.12) and (3.13).

Applying (3.27) to (1.34), it follows that, for \(0< \varrho < \frac{1}{4}\),

$$\begin{aligned} |({\mathbf {I}} - {\mathbf {P}}) f_2 | \lesssim \kappa |\nabla _x u| e^{- \varrho |v-\varepsilon u|^2} , \ \nonumber \\ \ |\partial _t ({\mathbf {I}} - {\mathbf {P}}) f_2 | \lesssim \kappa \{ | \partial _ t \nabla _x u| + \varepsilon |\partial _t u| | \nabla _x u|\} e^{- \varrho |v-\varepsilon u|^2}. \end{aligned}$$
(3.57)

From (3.1)

$$\begin{aligned} |{\mathbf {P}} f_2| \lesssim |p| e^{- \frac{|v-\varepsilon u|^2}{4}}, \ \ |\partial _t {\mathbf {P}} f_2| \lesssim |\partial _t p| e^{- \frac{|v-\varepsilon u|^2}{4}} + \varepsilon |\partial _t u | | p| \langle v-\varepsilon u \rangle e^{- \frac{|v-\varepsilon u|^2}{4}}.\nonumber \\ \end{aligned}$$
(3.58)

These estimates give (3.10) and (3.11).

The last term of (1.29) is bounded as

$$\begin{aligned}&\frac{\varepsilon }{\delta } \Big | \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_2 \Big | \nonumber \\&\qquad \lesssim \frac{\varepsilon }{\delta } \{|p| + \kappa |\nabla _x u| \} \{ |\nabla _x u| + \varepsilon (|\partial _t u| + |u | |\nabla _x u|) \} e^{-\varrho |v-\varepsilon u|^2}, \end{aligned}$$
(3.59)
$$\begin{aligned}&\frac{\varepsilon }{\delta } \Big | \partial _t \bigg ( \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_2 \bigg ) \Big |\nonumber \\&\quad \lesssim \frac{\varepsilon }{\delta } \{|\nabla _x u| + \varepsilon (|\partial _t u| + |u| |\nabla _x u|)\}\nonumber \\&\qquad \times \{ |\partial _t p| + \kappa |\partial _t \nabla _x u| + \varepsilon |\partial _t u| ( |p| +\kappa |\nabla _x u|) \}e^{-\varrho |v-\varepsilon u|^2} \nonumber \\&\qquad \quad + \frac{\varepsilon }{\delta } (|p| + \kappa |\nabla _x u|) \{ |\nabla _x\partial _tu|+\varepsilon (|\partial _t ^2 u | + |u ||\nabla _x \partial _t u| + |\partial _t u| |\nabla _x u|)\nonumber \\&\qquad \qquad \qquad \qquad + \varepsilon ^2 |\partial _t u| (|\partial _t u| + |u| |\nabla _x u|) \}e^{-\varrho |v-\varepsilon u|^2}. \end{aligned}$$
(3.60)

Lastly from (3.36), (3.1), and (1.34)

$$\begin{aligned} \frac{\varepsilon }{\delta \kappa }|\Gamma (f_2,f_2)(v)|= & {} \frac{\varepsilon }{\delta \kappa }|\Gamma (f_2,f_2)(v) - \Gamma ({\mathbf {P}} f_2, {\mathbf {P}}f_2 ) |\nonumber \\\lesssim & {} \frac{\varepsilon }{\delta } (|p|+ \kappa | \nabla _x u|) |\nabla _x u| \nu (v) e^{-\varrho |v-\varepsilon u|^2}, \end{aligned}$$
(3.61)
$$\begin{aligned} \frac{\varepsilon }{\delta \kappa }|\partial _t \Gamma (f_2,f_2)(v)|\lesssim & {} \frac{\varepsilon }{\delta } \{(|p| + \kappa |\nabla _x u| ) |\partial _t \nabla _x u| \nonumber \\&\quad \,+(|\partial _t p| + \kappa |\nabla _x \partial _t u| )|\nabla _x u| \} \nu (v) e^{-\varrho |v-\varepsilon u|^2}\nonumber \\&\quad \,+ \frac{\varepsilon ^2}{\delta }|\partial _t u|(|p| + \kappa |\nabla _x u| )|\nabla _x u| \nu (v) e^{-\varrho |v-\varepsilon u|^2},\qquad \end{aligned}$$
(3.62)

where we have used \(\Gamma ({\mathbf {P}} f_2, {\mathbf {P}}f_2 )=\Gamma (p \sqrt{\mu }, p \sqrt{\mu })=0\) to eliminate the contribution of \(p^2\) in (3.61).

Finally we wrap up the estimates of the source term of (3.2) to show (3.4) and (3.5). The term \(({\mathbf {I}} - {\mathbf {P}}) {\mathfrak {R}}_{1}\) consists of (3.39), which is bounded as (3.46) and hence we prove (3.4). The rest of terms form \({\mathfrak {R}}_{2}\), which can be proved to be bounded as (3.5), from (3.45), (3.53), (3.55), (3.59), and (3.61).

Now we consider the source term of (3.3). The term \(({\mathbf {I}} - {\mathbf {P}}) {\mathfrak {R}}_3\) consists of (3.48), which is bounded as (3.50). From (3.49), (3.52), (3.54)-(3.56), (3.12), (3.13), (3.60), (3.62), and (3.36), we prove (3.7).

4 A Priori Estimates for \(f_R\)

For each \(\varepsilon >0\) an existence of a unique solution in a time interval \([0,\infty )\) can be found in [13]. Thereby we only focus on a priori estimates of \(f_R\) in different spaces. For the sake of simplicity at times we will use simplified notations

$$\begin{aligned}&\Vert g(t,x,v) \Vert _{L^{p_1}_t L^{p_2}_x L^{p_3}_v}\nonumber \\&\quad : = \Big \Vert \big \Vert \Vert g(t,x,v)\Vert _{L^{p_3}_v({\mathbb {R}}^3)} \big \Vert _{L^{p_2}_x(\Omega )} \Big \Vert _{L^{p_1}_t ([0,T])}, \ \ \Vert g \Vert _{L^p_{t,x,v}} := \Vert g \Vert _{L^p_t L^p_x L^p_v}. \end{aligned}$$
(4.1)

Recall the boundary integral and the norms in (2.5). Also recall \({\mathfrak {w}}={\mathfrak {w}}_{\varrho , \ss }(x,v)\) in (2.3) and \({\mathfrak {w}}'={\mathfrak {w}}_{\varrho ', \ss }(x,v) \) for \(0<\varrho '<\varrho \).

4.1 \(L^2\)-Energy Estimate

Our starting point is a basic \(L^2\)-energy estimate for the Boltzmann remainder \(f_R\) and its temporal derivative \(\partial _t f_R\) in which the dissipation (1.31) plays an important role in the nonlinear estimate.

Proposition 7

Under the same assumptions in Proposition 6, we have

$$\begin{aligned} \begin{aligned}&\Vert f_R (t)\Vert _{L^2_{x,v}}^2+ {d}_{2} \int ^t_0 \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}})f_R \Vert _{L^2_{x,v}}^2 + \int ^t_0 | \varepsilon ^{-\frac{1}{2}} f_R |_{ L^2_\gamma }^2 \\&\quad \lesssim \ \Vert f_R (0)\Vert _{L^2_{x,v}}^2 + (1+ \Vert (3.12)\Vert _{L^\infty _{t,x}} )\int ^t_0 \Vert P f_R (s) \Vert _{L^2_x}^2 \mathrm {d}s\\&\qquad + \frac{ \delta ^2}{\kappa ^3 } \Vert \kappa ^{1/2} {P} f_R(s) \Vert _{L^\infty _tL^6_{x }}^2 \Vert \kappa ^{1/2}P f_R \Vert ^2_{L^2_tL^3_{x}} \\&\qquad + \frac{\varepsilon \kappa ^2}{ \delta ^2} | \nabla _x u|^2_{L^2_t L^2 (\partial \Omega )} + \kappa \varepsilon ^{1/8} \Vert \nabla _x u \Vert _{L^2_{t,x }} ^2 + \kappa \varepsilon ^2 \Vert (3.4) \Vert ^2_{L^2_{t,x}} + \Vert (3.5) \Vert ^2_{L^2_{t,x}}, \end{aligned} \end{aligned}$$
(4.2)

where

$$\begin{aligned} {d}_{2}:= & {} \frac{\sigma _0}{2}- \delta \varepsilon \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}} -( \varepsilon ^{\frac{15}{16}} \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}} )^2 - \varepsilon ^2 \Vert (3.10) \Vert _{L^\infty _{t,x}} \nonumber \\&- \frac{\varepsilon ^2}{\kappa } \Vert (3.10) \Vert _{L^\infty _{t,x}}^2 - \varepsilon \kappa ^{1/2} \Vert (3.12)_* \Vert _{L^\infty _{t,x}} . \end{aligned}$$
(4.3)

Here \(L^p_t=L^p_t([0,t])\) in particular. Note that a weighted \(L^\infty \)-bound of \(f_R\) is involved in this energy estimate, where the weight \({\mathfrak {w}}={\mathfrak {w}}_{\varrho , \ss }(x,v)\) is defined in (2.3).

We also have

$$\begin{aligned} \begin{aligned}&\Vert \partial _t f_R (t) \Vert _{L^2_{x,v}}^2 + {d}_{2,t} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R\Vert _{L^2_{t,x,v}}^2 \\&\qquad + |\varepsilon ^{- \frac{1}{2}} \partial _t f_R|_{L^2_tL^2_\gamma }^2 - \varepsilon \Vert \partial _t u \Vert _{L^\infty _{t,x} } | f_R |_{L^2 _t L^2_{\gamma }}^2 \\&\quad \lesssim \ \Vert \partial _t f_R (0) \Vert _{L^2_{x,v}}^2 +\kappa ^{-1} \delta ^2 \Vert Pf_R \Vert _{L^\infty _t L^6_x}^2 \{\Vert P f_R \Vert _{L^2_t L^3_x}^2 + \Vert P \partial _t f_R \Vert _{L^2_t L^3_x}^2 \}\\&\qquad + \Big \{ \kappa ^{-1 } \Vert \partial _t u \Vert _{L^\infty _{t,x }} ^2 + \Vert \nabla _x \partial _t u \Vert _{L^\infty _{t,x}} + \varepsilon \Vert \partial _t^2 u \Vert _{L^2_{t} L^\infty _x} \\&\qquad \qquad + \varepsilon \kappa ^{-1/2} (1+ \Vert \partial _t u \Vert _{L^\infty _{t,x}} )\Vert (3.10) \Vert _{L^\infty _{t,x}} \\&\qquad \qquad +\Vert (3.12) \Vert _{L^\infty _{t,x} } + \Vert (3.13)_* \Vert _{L^\infty _{t,x} } \Big \} \times \int ^t_0 \Vert P \partial _tf_R(s) \Vert _{L^{2}_x }^2 \mathrm {d}s \\&\qquad + \Big \{ \kappa ^{-1} \Vert \partial _t u \Vert _{L^\infty _{t,x}}^2 + \Vert \nabla _x \partial _t u \Vert _{L^\infty _{t,x}} + \varepsilon \Vert \partial _t^2 u \Vert _{L^2_{t} L^\infty _x} + (\varepsilon \Vert (3.10)\Vert _{L^\infty _{t,x}})^2\\&\qquad \qquad + (\varepsilon \kappa ^{-1/2} \Vert (3.11)\Vert _{L^\infty _{t,x}})^2 + \Vert (3.13)_*\Vert _{L^\infty _{t,x}} \Big \} \times \int ^t_0 \Vert P f_R(s) \Vert _{L^{2}_x }^2 \mathrm {d}s \\&\qquad + \Big \{ \varepsilon (1+ \varepsilon \Vert (3.10)\Vert _{L^\infty _{t,x}}) \Vert \partial _t u \Vert _{L^\infty _{t,x}} + \varepsilon \kappa \Vert \nabla _x \partial _t u \Vert _{L^\infty _{t,x}} + \varepsilon ^2 \kappa \Vert \partial _t ^2u \Vert _{L^2_tL^\infty _{ x}} \\&\qquad \qquad + (\varepsilon \kappa ^{1/2} \Vert (3.13)_*\Vert _{L^\infty _{t,x}})^2 +( \varepsilon \delta \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}}) ^2 \Big \} \times \Vert \varepsilon ^{-1}\kappa ^{-1/2} \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}}^2 \\&\qquad + e^{- \frac{\varrho }{4 \varepsilon ^2}} \{\Vert (3.12)\Vert _{L^\infty _{t,x}}^2 + \Vert \nabla _x \partial _t u \Vert _{L^2_t L^\infty _x}^2 \} + (\varepsilon \kappa ^{1/2} \Vert (3.6) \Vert _{L^2_{t,x}})^2 + \Vert (3.7)\Vert _{L^2_{t,x}}^2\\&\qquad + \frac{\varepsilon \kappa ^2}{ \delta ^2} | |\partial _t \nabla _x u| + \varepsilon |\partial _t u| |\nabla _x u| |_{L^2_tL^2 (\partial \Omega ) }^2 + \frac{\varepsilon ^3 \kappa ^2 }{\delta } |\nabla _x u|^2_{L^2_t L^2(\partial \Omega )} \Vert \partial _t u \Vert _{L^\infty _{t,x}} , \end{aligned} \end{aligned}$$
(4.4)

where \(\partial _t f_R (0,x,v) := f_{R,t}(0,x,v)\) is defined in (2.6). Here

$$\begin{aligned} \begin{aligned} {d}_{2,t}:=&\frac{\sigma _0}{2}- \varepsilon ( \kappa ^{-1/2} + \varepsilon \Vert \partial _t u\Vert _{L^\infty _{t,x}}) \Vert (3.10)\Vert _{L^\infty _{t,x}}- \varepsilon \kappa \Vert (3.12) \Vert _{L^\infty _{t,x}}\\&- (\varepsilon \kappa ^{1/2} \Vert (3.13)_* \Vert _{L^\infty _{t,x}})^2\\&- \varepsilon \kappa \Vert \nabla _x \partial _t u \Vert _{L^\infty _{t,x}} - \varepsilon ^2 \kappa \Vert \partial _t^2 u \Vert _{L^2_tL^\infty _{x}} + \varepsilon \Vert \partial _t u \Vert _{L^\infty _{t,x}} (1+ \varepsilon \Vert \partial _t u \Vert _{L^\infty _{t,x}}) \\&- \varepsilon ^{- \frac{\varrho }{4 \varepsilon ^2}} (\varepsilon \kappa ^{1/2} \Vert {\mathfrak {w}}_{\varrho ^\prime , \ss } \partial _t f_R \Vert _{L^2_t L^\infty _{x,v}})^2 - \varepsilon \delta (1+ \varepsilon \Vert \partial _t u \Vert _{L^\infty _{x,v}}) \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x}}\\&- (\varepsilon \kappa ^{1/2} \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}})^2, \end{aligned} \end{aligned}$$
(4.5)

where \(0<\varrho ^\prime < \varrho \).

Remark 8

We utilize several different time-space norms to control the fluid source terms, which possess the initial-boundary and boundary layers as in Theorem 3.

The following trace theorem is useful to control the boundary terms.

Lemma 5

(Trace theorem)

$$\begin{aligned}&\frac{1}{\varepsilon }\int ^t_0 \int _{\gamma _+^N} |h| \mathrm {d}\gamma \mathrm {d}s\nonumber \\&\quad \lesssim _N \iint _{\Omega \times {\mathbb {R}}^3} |h(0)| + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} |h| + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3 } | \partial _t h +\frac{1}{\varepsilon } v\cdot \nabla _x h|, \end{aligned}$$
(4.6)

where \(\gamma _+^N:= \{(x,v) \in \gamma _+: |n(x) \cdot v|> 1/N \ \text {and} \ 1/N<|v|<N \}\).

The proof is standard (for example see Lemma 3.2 in [13] or Lemma 7 in [6]).

Proof of Proposition 7

First we prove (4.2). An energy estimate to (3.2) and (3.8) reads as

$$\begin{aligned}&\frac{1}{2}\Vert f _R(t)\Vert _{L^2_{x,v}}^2 - \frac{1}{2} \Vert f _R (0)\Vert _{L^2_{x,v}}^2 + \frac{1}{ \kappa \varepsilon ^{2 }} \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3}f _R L f_R \end{aligned}$$
(4.7)
$$\begin{aligned}&\qquad + \frac{1}{2\varepsilon } \int ^t_0 \int _{\gamma _+} |f_R|^2 - \frac{1}{2\varepsilon } \int ^t_0 \int _{\gamma _-} | P_{\gamma _+} f_R- \frac{\varepsilon }{\delta } (1- P_{\gamma _+}) ({\mathbf {I}} -{\mathbf {P}}) f_2 |^2 \end{aligned}$$
(4.8)
$$\begin{aligned}&\quad = \frac{\delta }{\kappa \varepsilon } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \Gamma (f_R,f_R) ({\mathbf {I}} - {\mathbf {P}}) f_R \end{aligned}$$
(4.9)
$$\begin{aligned}&\qquad + \frac{2}{ \kappa } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \Gamma (f_2, f_R) ({\mathbf {I}} - {\mathbf {P}}) f_R \end{aligned}$$
(4.10)
$$\begin{aligned}&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} ( {\mathbf {I}} - {\mathbf {P}}) {\mathfrak {R}}_1 ( {\mathbf {I}} - {\mathbf {P}}) f_R \end{aligned}$$
(4.11)
$$\begin{aligned}&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} {\mathfrak {R}}_2 f_R \end{aligned}$$
(4.12)
$$\begin{aligned}&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \frac{- (\partial _t + \varepsilon ^{-1} v\cdot \nabla _x ) \sqrt{\mu }}{\sqrt{\mu }} |f_R|^2. \end{aligned}$$
(4.13)

Among others two terms (4.9) and (4.13) are most problematic.

We start with (4.7). From the spectral gap estimate in (1.25), we have

$$\begin{aligned} (4.7)\ge \frac{1}{2}\Vert f _R(t)\Vert _{L^2_{x,v}}^2 - \frac{1}{2} \Vert f _R (0)\Vert _{L^2_{x,v}}^2 + \sigma _0 \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) f_R\Vert _{L^2_{t,x,v}}^2. \end{aligned}$$
(4.14)

Now we consider (4.9), in which we need integrability gain of \({\mathbf {P}}f_R\) in \(L^6_x\) of the next sections. From decomposition \(f_R= {\mathbf {P}} f_R + ({\mathbf {I}} - {\mathbf {P}}) f_R\) and \(\Gamma =\Gamma _+- \Gamma _-\) in (1.22), we derive

$$\begin{aligned} \begin{aligned} |(4.9)| \lesssim&\ \frac{\delta }{\kappa \varepsilon } \sum _{i=\pm }\int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} | \nu ^{-\frac{1}{2}} \Gamma _{i}( |f_R|, ({\mathbf {I}} -{\mathbf {P}}) f_R)| | \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) f_R|\\&+\frac{\delta }{\kappa \varepsilon } \sum _{i=\pm } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} | \nu ^{-\frac{1}{2}} \Gamma _{i}(|{\mathbf {P}}f_R|, |{\mathbf {P}} f_R|)| | \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) f_R| \\ \lesssim&\ \delta \varepsilon \Vert {\mathfrak {w}}f_R\Vert _{L^\infty _{x,v}} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert ^2_{L^2_{t,x,v}} \\&+ \frac{ \delta }{\kappa ^{3/2}} \Vert \kappa ^{1/2} {P} f_R \Vert _{L^\infty _tL^6_{x }} \Vert \kappa ^{1/2} {P} f_R \Vert _{L^2_tL^3_{x }} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}}. \end{aligned} \end{aligned}$$
(4.15)

From (3.57) and (3.58),

$$\begin{aligned} \begin{aligned}&|(4.10)|\\&\quad \le \frac{\varepsilon }{\kappa ^{1/2}} \Vert (3.10) \Vert _{L^\infty _{t,x}} \{ \Vert P f_R \Vert _{L^2_{t,x}} + \kappa ^{ \frac{1}{2}} \varepsilon \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \}\\&\qquad \times \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \\&\quad \lesssim \{\varepsilon ^2 \Vert (3.10) \Vert _{L^\infty _{t,x}} + \frac{\varepsilon ^2}{\kappa } \Vert (3.10) \Vert _{L^\infty _{t,x}}^2 \} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}}^2 + \Vert P f_R \Vert _{L^2_tL^2_{x}}^2. \end{aligned} \end{aligned}$$
(4.16)

From (3.4) and (3.5) we derive that

$$\begin{aligned} \begin{aligned} |(4.11)|&\lesssim \kappa ^{ 1/2} \varepsilon \Vert (3.4) \Vert _{L^2_{t,x}} \Vert \kappa ^{-1/2} \varepsilon ^{-1} ({\mathbf {I}} - {\mathbf {P}} ) f_R \Vert _{L^2_{t,x,v}}, \\ |(4.12)|&\lesssim \Vert (3.5)\Vert _{L^2_{t,x,v}}^2 + \Vert {\mathbf {P}}f_R\Vert _{L^2_{t,x,v}}^2 \\&\qquad + \kappa ^{ 1/2} \varepsilon \Vert (3.5)\Vert _{L^2_{t,x,v}} \Vert \kappa ^{-1/2} \varepsilon ^{-1} ({\mathbf {I}} -{\mathbf {P}})f_R\Vert _{L^2_{t,x,v}} . \end{aligned} \end{aligned}$$
(4.17)

Next using (3.12) it follows that

$$\begin{aligned}&|(4.13)|\nonumber \\&\quad \lesssim \ \kappa ^{1/2} \varepsilon \Vert {\mathfrak {w}}^{-1}{ \nu ^{3/2} } \nabla _x u \Vert _{L^2_{t,x,v}} \Vert {\mathfrak {w}}f_R\Vert _{ L^\infty _{t,x,v}} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}})f_R\Vert _{L^2_{t,x,v}}\nonumber \\&\qquad + \Vert (3.12)\Vert _{L^\infty _{t,x}} \int ^t_0 \Vert P f_R (s) \Vert _{L^2_x}^2 \mathrm {d}s\nonumber \\&\qquad + \varepsilon \kappa ^{1/2} \Vert (3.12)_* \Vert _{L^\infty _{t,x}} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}})f_R\Vert _{L^2_{t,x,v}}^2\nonumber \\&\quad \lesssim \ \big \{( \varepsilon ^{\frac{15}{16}} \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}} )^2 + \varepsilon \kappa ^{1/2} \Vert (3.12)_* \Vert _{L^\infty _{t,x}} \big \} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}})f_R\Vert _{L^2_{t,x,v}}^2 \nonumber \\&\qquad + \Vert (3.12)\Vert _{L^\infty _{t,x}} \int ^t_0 \Vert P f_R (s) \Vert _{L^2_x}^2 \mathrm {d}s + (\kappa ^{\frac{1}{2}} \varepsilon ^{\frac{1}{16}} \Vert \nabla _x u \Vert _{L^2_{t,x }} )^2. \end{aligned}$$
(4.18)

Finally we control the boundary term (4.8) using a trace theorem (4.6). First we have, from (3.8),

$$\begin{aligned} \begin{aligned} (4.8) =&\frac{1}{2\varepsilon } \int _0^t \int _{\gamma _+} \{ |f_R|^2 - |P_{\gamma _+} f_R|^2 \} - \frac{\varepsilon }{2\delta ^2} \int ^t_0 \int _{\gamma _-} |(1- P_{\gamma _+}) ({\mathbf {I}}-{\mathbf {P}}) f_2|^2\\&- \int ^t_0 \int _{\gamma _-} \frac{1}{\varepsilon ^{1/2}}P_{\gamma _+ } f_R \frac{\varepsilon ^{1/2}}{\delta }(1- {P}_{\gamma _+}) ({ \mathbf {I}}-{\mathbf {P}}) f_2 \\ \ge&\frac{1}{2 } | \varepsilon ^{-\frac{1}{2}} (1- P_{\gamma _+}) f_R |_{L^2_t L^2_{\gamma _+}}^2 -\frac{1}{8C} | \varepsilon ^{-\frac{1}{2}} P_{\gamma _+} f_R |_{L^2_t L^2_{\gamma _+}}^2\\&-( \frac{\varepsilon }{2\delta ^2} + 2C \frac{\varepsilon }{\delta ^2} ) \int ^t_0 \int _{\gamma _+} |(1- P_{\gamma _+}) ({\mathbf {I}}-{\mathbf {P}}) f_2|^2 \ \ \text {for} \ C\gg 1, \end{aligned} \end{aligned}$$
(4.19)

where we have used the fact \(|P_{\gamma _+}f_R|_{L^2_{\gamma _+}}= |P_{\gamma _+}f_R|_{L^2_{\gamma _-}}\) from \(P_{\gamma _+}f_R(t,x,v)\) being a function of (tx, |v|) due to \(u|_{\partial \Omega }=0\).

Now we estimate \(P_{\gamma _+}f_R\). Since \(P_{\gamma _+}\) in (3.8) is a projection of \(c_\mu \sqrt{\mu }\) on \(\gamma _+\), it follows \(\int _{\gamma _+} |P_{\gamma _+}f|^2 \le 2 \int _{\gamma _+^N} |P_{\gamma _+}f|^2\) for large enough \(N>0\), where \(\gamma _+^N:= \{(x,v) \in \gamma _+: |n(x) \cdot v|> 1/N \ \text {and} \ 1/N<|v|<N \}\). Setting \(h=|f|^2\) in (4.6) and using (3.2), (3.4), and (3.5) we derive

$$\begin{aligned} \begin{aligned}&\frac{1}{\varepsilon }\int ^t_0 \int _{\gamma _+ } |f_R|^2 \mathrm {d}\gamma \mathrm {d}s\\&\quad \le C_N \iint _{\Omega \times {\mathbb {R}}^3} |f_R(0)|^2 + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} |f_R|^2 \\&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3 }\Big |\Big [ - \frac{1}{\varepsilon ^2 \kappa } Lf_R + \frac{1}{\kappa } \Gamma ({f_2}, f_R) + \frac{ \delta }{ \varepsilon \kappa }\Gamma (f_R, f_R) \\&\qquad \qquad \qquad \qquad \qquad \quad - \frac{( \partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_{R} + ({\mathbf {I}}- {\mathbf {P}}){\mathfrak {R}}_1 + {\mathfrak {R}}_2\Big ] f_R\Big |\\&\quad \le C_N \big \{ \Vert f_R(0)\Vert ^2_{L^2_{x,v}} + \Vert {\mathbf {P}} f_R \Vert _{L^{2}_{t,x,v}}^2 + \Vert \varepsilon ^{-1} \kappa ^{-1/2} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^{2}_{t,x,v}}^2\\&\qquad \qquad \quad + (4.15) + (4.16) +(4.17) + (4.18)\big \}. \end{aligned} \end{aligned}$$
(4.20)

Furthermore from (3.8) and (4.20)

$$\begin{aligned} | f_R|_{L^2_t L^2_{\gamma _-} }^2\lesssim & {} | f_R|_{L^2_t L^2_{\gamma _+} }^2 + \frac{\varepsilon ^2}{\delta ^2} | (1- P_{\gamma _+}) ({\mathbf {I}}-{\mathbf {P}}) f_2| _{L^2_t L^2_{\gamma _-} } ^2 \nonumber \\= & {} |f_R|_{L^2_t L^2_{\gamma _+} }^2 + \frac{\varepsilon ^2\kappa ^2}{\delta ^2} | \nabla _x u| _{L^2_t L^2 (\partial \Omega ) } ^2 . \end{aligned}$$
(4.21)

Finally we collect the terms as

$$\begin{aligned} \begin{aligned}&\text {r.h.s of} \ (4.14)+ (4.19) + \frac{1}{4C} | \varepsilon ^{-\frac{1}{2}} P_{\gamma _+} f_R |_{L^2_t L^2_{\gamma _+}}^2 + \frac{\varepsilon ^{-1}}{16C} | f_R|_{ L^2_t L^2_{\gamma _-}}^2\\&\quad \le \text {r.h.s of} \ (4.15) + (4.16) + (4.17) + (4.18) + \frac{1}{4C} \times \text {r.h.s of} \ (4.20) \\&\qquad + \frac{\varepsilon ^{-1}}{16C} \times \text {r.h.s of} \ (4.21). \end{aligned} \end{aligned}$$

We choose large N and then large C so that \(\frac{C_N}{4C}\ll \sigma _0\). Using Young’s inequality for products, and then moving contributions of \( \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}}^2\) to l.h.s., we derive (4.2).

Next we prove (4.4). An energy estimate to (3.3) and (3.9) lead to (4.4)

$$\begin{aligned}&\frac{1}{2}\Vert \partial _t f _R(t)\Vert _2^2 - \frac{1}{2} \Vert \partial _t f _R (0)\Vert _2^2 + \frac{1}{ \kappa \varepsilon ^{2 }} \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3}\partial _t f _R L \partial _t f_R \end{aligned}$$
(4.22)
$$\begin{aligned}&\qquad + \frac{1}{2\varepsilon } \int ^t_0 \int _{\gamma _+} |\partial _t f_R|^2 - \frac{1}{2\varepsilon } \int ^t_0 \int _{\gamma _-} | P_{\gamma _+} \partial _t f_R\nonumber \\&\qquad - \frac{\varepsilon }{\delta } (1-P_{\gamma _+}) \partial _t ({\mathbf {I}} -{\mathbf {P}}) f_2 +r_{\gamma _+} (f_R)- \frac{\varepsilon }{\delta }r_{\gamma _+} ( ({\mathbf {I}} -{\mathbf {P}}) f_2) |^2 \end{aligned}$$
(4.23)
$$\begin{aligned}&\quad = - \frac{1}{\varepsilon ^2 \kappa } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} L_t ({\mathbf {I}} - {\mathbf {P}})f_R \partial _t f_R +\frac{1}{\varepsilon ^2 \kappa } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} L({\mathbf {P}}_t f_R)({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \end{aligned}$$
(4.24)
$$\begin{aligned}&\qquad + \frac{2\delta }{\kappa \varepsilon } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \Gamma (f_R,\partial _t f_R) ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \nonumber \\&\qquad + \frac{2 }{ \kappa } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \Gamma ( f_2,\partial _t f_R) ({\mathbf {I}} - {\mathbf {P}})\partial _t f_R \end{aligned}$$
(4.25)
$$\begin{aligned}&\qquad + \frac{2}{ \kappa } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \Gamma (\partial _t f_2, f_R) ({\mathbf {I}} - {\mathbf {P}})\partial _t f_R \end{aligned}$$
(4.26)
$$\begin{aligned}&\qquad + \frac{2}{\kappa } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \Gamma _t( {f_2}, f_R) \partial _t f_R+ \frac{\delta }{\varepsilon \kappa } \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \Gamma _t (f_R,f_R)\partial _tf_R \end{aligned}$$
(4.27)
$$\begin{aligned}&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} ( {\mathbf {I}} - {\mathbf {P}}) {\mathfrak {R}}_3 ( {\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \end{aligned}$$
(4.28)
$$\begin{aligned}&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} {\mathfrak {R}}_4 \partial _t f_R \end{aligned}$$
(4.29)
$$\begin{aligned}&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3} \frac{- (\partial _t + \varepsilon ^{-1} v\cdot \nabla _x ) \sqrt{\mu }}{\sqrt{\mu }} |\partial _t f_R|^2 \nonumber \\&\qquad + \int ^t_0 \iint _{\Omega \times {\mathbb {R}}^3}\partial _t \Big ( \frac{ -(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x ) \sqrt{\mu }}{\sqrt{\mu }} \Big )f_R \partial _t f_R . \end{aligned}$$
(4.30)

We consider the first term of (4.30). We decompose \(\partial _t f_R = {\mathbf {P}} \partial _t f_R + ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R\). The contribution of \({\mathbf {P}} \partial _t f_R\) can be bounded above as, from (3.12),

$$\begin{aligned} \Vert (3.12) \Vert _{L^\infty _{t,x} } \int ^t_0 \Vert P \partial _tf_R(s) \Vert _{L^{2}_x }^2 \mathrm {d}s . \end{aligned}$$
(4.31)

For the contribution of \(({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R\) we utilize an extra decomposition \({\mathbf {1}}_{|v|\le \varepsilon ^{-1}} + {\mathbf {1}}_{|v|\ge \varepsilon ^{-1}}\) Then it is bounded as

$$\begin{aligned} \begin{aligned}&\Vert (3.12) \Vert _\infty \Big \{ \iiint {\mathbf {1}}_{|v|\le \varepsilon ^{-1}} |v| \nu (v) |({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R|^2\\&\qquad + \iiint {\mathbf {1}}_{|v|\ge \varepsilon ^{-1}} \frac{|v|^{3/2}}{{\mathfrak {w}}^\prime (v)} {\mathfrak {w}}^\prime (v) \partial _t f_R(v) \sqrt{\nu (v)} |({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R | \Big \}\\&\quad \lesssim \ \Vert (3.12) \Vert _\infty \Big \{ \varepsilon ^{-1} \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}}^2\\&\qquad + e^{-\frac{\varrho }{4\varepsilon ^2}} \Vert {\mathfrak {w}}^\prime \partial _t f_R \Vert _{L^2_t L^\infty _{x,v}} \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \Big \}. \end{aligned} \end{aligned}$$
(4.32)

For the second term of (4.30) using (3.13) we bound it by

$$\begin{aligned} \begin{aligned}&\Vert (3.13)_* \Vert _{L^\infty _{t,x}} \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}}\\&\quad + e^{-\frac{\varrho }{4\varepsilon ^2}} \Vert \nabla _x \partial _t u \Vert _{L^2_t L^\infty _{ x}} \Vert {\mathfrak {w}} f_R \Vert _{ L^\infty _{t,x,v}} \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \\&\quad +\{ \Vert \nabla _x \partial _t u \Vert _{L^\infty _{t,x}}+ \Vert (3.13)_*\Vert _{L^\infty _{t,x}}\} \Big \{ \int ^t_0 \Vert Pf_R (s)\Vert _{L^2_x}^2 \mathrm {d}s + \int ^t_0 \Vert P\partial _t f_R (s)\Vert _{L^2_x}^2 \mathrm {d}s \Big \}. \end{aligned} \end{aligned}$$
(4.33)

Using (3.35) we bound (4.24) and (4.27) as

$$\begin{aligned} |(4.24)|\lesssim & {} \kappa ^{-\frac{1}{2}} \Vert \partial _t u\Vert _{L^\infty _{t,x}} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \nonumber \\&\times \{\Vert P \partial _t f_R \Vert _{L^2_{t,x}} + \kappa ^{ \frac{1}{2}} \varepsilon \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \}\nonumber \\&+ \kappa ^{-\frac{1}{2}} \Vert \partial _t u\Vert _{L^\infty _{t,x}} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}})\partial _t f_R \Vert _{L^2_{t,x,v}} \Vert P f_R \Vert _{L^2_{t,x}} , \end{aligned}$$
(4.34)
$$\begin{aligned} |(4.27)|\lesssim & {} \kappa ^{-\frac{1}{2}}\varepsilon \Vert \partial _t u \Vert _{L^\infty _{t,x}} \Vert (3.10) \Vert _{L^\infty _{t,x}} \{ \Vert \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \nonumber \\&\qquad + \Vert P f_R \Vert _{L^2_{t,x}}\} \Vert \partial _t f_R \Vert _{L^2_{t,x,v}}\nonumber \\&+ \delta \kappa ^{-1} \Vert \partial _t u\Vert _{L^\infty _{t,x}} \{ \Vert P \partial _t f_R \Vert _{L^2_{t,x}}+ \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \}\nonumber \\&\times \{ \Vert P f_R \Vert _{L^\infty _tL^6_x} \Vert Pf_R \Vert _{L^2_t L^3_x} \nonumber \\&\qquad + \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \kappa ^{ \frac{1}{2}} \varepsilon \Vert {\mathfrak {w}}f _R\Vert _{L^\infty _{t,x,v}} \}. \end{aligned}$$
(4.35)

The rest of terms can be controlled similarly as in the proof of (4.2):

$$\begin{aligned} (4.22)\ge&\ \frac{1}{2}\Vert \partial _t f _R(t)\Vert _{L^2_{x,v}}^2 - \frac{1}{2} \Vert \partial _t f _R (0)\Vert _{L^2_{x,v}}^2 + \sigma _0 \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu }({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R\Vert _{L^2_{t,x,v}}^2, \end{aligned}$$
(4.36)
$$\begin{aligned} |(4.25)| \lesssim&\ \{ \delta \varepsilon \Vert {\mathfrak {w}}f_R\Vert _{L^\infty _{t,x,v}} + \varepsilon ^2 \Vert (3.10)\Vert _{L^\infty _{t,x }} \} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert ^2_{L^2_{t,x,v}} \nonumber \\&+ \frac{ \delta }{\kappa ^{3/2}} \Vert \kappa ^{1/2} {P} f_R \Vert _{L^\infty _tL^6_{x }} \Vert \kappa ^{1/2} {P} \partial _t f_R \Vert _{L^2_tL^3_{x }} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \end{aligned}$$
(4.37)
$$\begin{aligned}&+ \frac{\varepsilon }{\kappa ^{1/2}} \Vert {P} f_2 \Vert _{L^\infty _{t,x} } \Vert {P} \partial _t f_R \Vert _{L^2_{t,x} } \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}}, \end{aligned}$$
(4.38)
$$\begin{aligned} |(4.26)| \lesssim&\ \kappa ^{-\frac{1}{2}}\varepsilon \Vert \sqrt{\nu } \partial _t f_2 \Vert _{L^\infty _{t,x,v}} \{ \Vert P f_R \Vert _{L^2_{t,x}} \nonumber \\&+ \Vert \sqrt{\nu }({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \} \Vert \kappa ^{-\frac{1}{2}} \varepsilon ^{-1} ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} , \end{aligned}$$
(4.39)
$$\begin{aligned} |(4.28)| \lesssim&\ \kappa ^{ 1/2} \varepsilon \Vert (3.6) \Vert _{L^2_{t,x}} \Vert \kappa ^{-1/2} \varepsilon ^{-1} ({\mathbf {I}} - {\mathbf {P}} )\partial _t f_R \Vert _{L^2_{t,x,v}}, \end{aligned}$$
(4.40)
$$\begin{aligned} |(4.29)| \lesssim&\ \Vert (3.7) \Vert _{L^2_{t,x}} \big \{ \Vert {P } \partial _t f_R\Vert _{L^2_{t,x}}+ \kappa ^{1/2}\varepsilon \Vert \kappa ^{-1/2} \varepsilon ^{-1} ({\mathbf {I}} -{\mathbf {P}}) \partial _tf_R\Vert _{L^2_{t,x,v}} \big \}. \end{aligned}$$
(4.41)

Lastly we estimate (4.23) and the first term of (4.30). As in (4.19) we derive that (4.23) is bounded from below by

$$\begin{aligned} \begin{aligned}&\frac{1}{2 } | \varepsilon ^{-\frac{1}{2}} (1- P_{\gamma _+}) \partial _t f_R |_{L^2((0,T); L^2_{\gamma _+})}^2 -\frac{1}{8C} | \varepsilon ^{-\frac{1}{2}} P_{\gamma _+} \partial _t f_R |_{L^2 ((0,T); L^2_{\gamma _+})}^2\\&\qquad - C \Big \{ \frac{\varepsilon }{ \delta ^2} |(1- P_{\gamma _+}) \partial _t ({\mathbf {I}} -{\mathbf {P}}) f_2|_{L^2 ((0,T); L^2_{\gamma _-})}^2 + \varepsilon \Vert \partial _t u \Vert _{\infty } | f_R |_{L^2 ((0,T); L^2_{\gamma _+})}^2\\&\qquad + \Vert \partial _t u \Vert _{\infty }\frac{\varepsilon ^3}{\delta } | ({\mathbf {I}} -{\mathbf {P}}) f_2 |_{L^2 ((0,T); L^2_{\gamma _+})}^2 \Big \}\\&\quad \ge \frac{1}{2 } | \varepsilon ^{-\frac{1}{2}} (1- P_{\gamma _+}) \partial _t f_R |_{L^2((0,T); L^2_{\gamma _+})}^2 -\frac{1}{8C} | \varepsilon ^{-\frac{1}{2}} P_{\gamma _+} \partial _t f_R |_{L^2 ((0,T); L^2_{\gamma _+})}^2\\&\qquad - C \Big \{ \frac{\varepsilon \kappa ^2}{ \delta ^2} | |\partial _t \nabla _x u| + \varepsilon |\partial _t u| |\nabla _x u| |_{L^2_tL^2 (\partial \Omega ) }^2 + \varepsilon \Vert \partial _t u \Vert _{\infty } | f_R |_{L^2 ((0,T); L^2_{\gamma _+})}^2\\&\qquad + \Vert \partial _t u \Vert _{\infty }\frac{\varepsilon ^3}{\delta } \kappa ^2 |\nabla _x u|^2_{L^2_t L^2(\partial \Omega )} \Big \} \ \ \text {for} \ C\gg 1, \end{aligned} \end{aligned}$$
(4.42)

where we have used \(|r_{\gamma _+} (g)|_{L^2(\gamma _-)}\lesssim \varepsilon \Vert \partial _t u \Vert _\infty | g|_{L^2(\gamma _-)}\) from (3.9). Now we bound \(P_{\gamma _+} \partial _t f_R\) using (4.6). Following the argument arriving at (4.20) and setting \(h=|\partial _t f|^2\) we derive

$$\begin{aligned} \begin{aligned}&\frac{1}{\varepsilon } \int ^t_0 \int _{\gamma _+} | \partial _t f_R |^2 \mathrm {d}\gamma \mathrm {d}s \\&\quad \lesssim _N \Vert \partial _t f_R(0) \Vert _{L^2_{x,v}} + \Vert \partial _t f_R \Vert _{L^2_{t,x,v}}\\&\qquad + \int _0^t \iint _{\Omega \times {\mathbb {R}}^3} \Big |\Big ( - \frac{1}{\varepsilon ^2 \kappa } L \partial _t f_R + \text {r.h.s of } (3.3)\Big )\partial _t f_R \Big |\\&\quad \lesssim _N \Vert \partial _t f_R(0) \Vert _{L^2_{x,v}} ^2+ \Vert P \partial _t f_R \Vert _{L^2_{t,x}} ^2 + \Vert \varepsilon ^{-1} \kappa ^{-1/2} \sqrt{\nu } ({\mathbf {I}}-{\mathbf {P}}) \partial _t f _R \Vert _{L^2_{t,x,v}} ^2\\&\qquad + (4.31) + \cdots + (4.35)+ (3.37)+ \cdots + (4.41). \end{aligned} \end{aligned}$$
(4.43)

We conclude (4.4) by collecting the terms.\(\square \)

4.2 \(L^6_x\)-Integrability Gain for \({\mathbf {P}}f_R\)

Proposition 8

Under the same assumptions in Proposition 6, we have for all \(t \in [0,T]\)

$$\begin{aligned} \begin{aligned}&d_6 \Vert {P} f_R(t) \Vert _{L^6_x} \\&\quad \lesssim \ (\varepsilon \Vert (3.12)\Vert _{L^\infty _{t,x}} + \varepsilon \kappa ^{-1} \Vert (3.10)\Vert _{L^\infty _{t,x}} ) \Vert f_R(t) \Vert _{L^2_{x,v}} + \varepsilon \Vert \partial _t f_R(t) \Vert _{ {L^2_{x,v}}} \\&\qquad +o(1) (\kappa \varepsilon )^{1/2} \Vert {\mathfrak {w}} f_R (t) \Vert _{L^\infty _{x,v}} + \frac{\varepsilon }{\delta }| (3.10)|_{L^4(\partial \Omega )} + \varepsilon \Vert (3.4) \Vert _{L^2_{x,v}}+ \varepsilon \Vert (3.5) \Vert _{L^2_{x,v}} \\&\qquad + \Big ( \frac{1}{\varepsilon \kappa }+ \frac{\delta }{\kappa } \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R(t) \Vert _{L^\infty _{x,v}} \Big ) \big \{ \Vert ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{ {L^2_{t,x,v}}} + \Vert ({\mathbf {I}} - {\mathbf {P}})\partial _t f_R \Vert _{ {L^2_{t,x,v}}} \\&\qquad + \varepsilon \Vert \partial _t u \Vert _{L^\infty _{t,x}} \Vert P f_R \Vert _{L^2_{t,x}} \big \} \\&\qquad + \Vert {\mathfrak {w}}_{\varrho , \ss } f_R(t) \Vert _{L^\infty _{x,v}}^{1/2} { \big \{ | f_R |_{L^2_tL^2({\gamma _+})}^{1/2} + | \partial _t f_R |_{L^2_tL^2({\gamma _+})}^{1/2} \big \} } , \end{aligned}\nonumber \\ \end{aligned}$$
(4.44)

where

$$\begin{aligned} d_6:=1-\Big [ \frac{\delta }{\kappa } \Vert Pf_R(t)\Vert _{L^6_x}^{1/2} \Vert Pf_R (t) \Vert _{L^2_x}^{1/2} + \varepsilon \Vert u(t) \Vert _{L^\infty _x}\Big ]^{1/6}. \end{aligned}$$
(4.45)

Proof

For the sake of simplicity we use notations (4.1) throughout this subsection.

We view (3.2) as a weak formulation for a test function \(\psi \)

$$\begin{aligned} \begin{aligned}&\underbrace{ \iint _{\Omega \times {\mathbb {R}}^3} f_R v\cdot \nabla _x \psi }_{(4.46)_1} - \underbrace{ \int _\gamma f_R \psi }_{(4.46)_2} -\underbrace{ \iint _{\Omega \times {\mathbb {R}}^3} \varepsilon \partial _t f_R \psi }_{(4.46)_3} \\&\quad = \iint _{\Omega \times {\mathbb {R}}^3} \psi \left\{ \frac{1}{ \varepsilon \kappa } L f_R - \frac{2\varepsilon }{\kappa } \Gamma ({f_2}, f_R) - \frac{ \delta }{ \kappa }\Gamma (f_R, f_R) \right. \\&\qquad \left. + \frac{( \varepsilon \partial _t + v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_{R} - \varepsilon ({\mathbf {I}}- {\mathbf {P}}){\mathfrak {R}}_1 - \varepsilon {\mathfrak {R}}_2\right\} . \end{aligned} \end{aligned}$$
(4.46)

The proof of the lemma is based on a recent test function method in the weak formulation ([12, 13]). We define

$$\begin{aligned} {\tilde{\mathbf {P}}}f_R := \Big \{ a + b \cdot v + c \frac{|v|^2-3}{\sqrt{6}} \Big \} \sqrt{\mu _0} \ \text {and} \ {\tilde{P}}f_R := (a,b,c), \end{aligned}$$
(4.47)

where \(a:= \langle f_R, \sqrt{\mu _0} \rangle , b:= \langle f_R,v \sqrt{\mu _0} \rangle \), and \(c:= \langle f_R, \frac{|v|^2-3}{\sqrt{6}}\sqrt{\mu _0} \rangle \). We choose a family of test functions as

$$\begin{aligned} \psi _{a }&:= (|v|^{2}-\beta _{a})v\sqrt{ \mu _0}\cdot \nabla _{x}\varphi _{a} , \end{aligned}$$
(4.48)
$$\begin{aligned} \psi _{b,1 }^{i,j}&:= (v_{i}^{2}-\beta _{b})\sqrt{\mu _0}\partial _{j} \varphi _{b }^{j},\quad i,j=1,2,3, \end{aligned}$$
(4.49)
$$\begin{aligned} \psi ^{i,j}_{b,2}&:=|v|^{2}v_{i}v_{j}\sqrt{\mu _0}\partial _{j}\varphi _{b }^{i} ,\quad i\ne j, \end{aligned}$$
(4.50)
$$\begin{aligned} \psi _{c }&:= (|v|^{2}-\beta _{c})v\sqrt{\mu _0}\cdot \nabla _{x}\varphi _{c } , \end{aligned}$$
(4.51)

where we choose \(\beta _a=10, \beta _b=1, \beta _c=5\) such that

$$\begin{aligned} 0= & {} \int _{{\mathbb {R}}^3} (|v|^2 - \beta _a) \frac{|v|^2-3}{\sqrt{6}} (v_1)^2 \mu _0 (v) \mathrm {d}v = \int _{{\mathbb {R}}} (v_1^2 - \beta _b) \mu _0 (v_1) \mathrm {d}v_1 \nonumber \\= & {} \int _{{\mathbb {R}}^3} (|v|^2-\beta _c) v_i^2 \mu _0 (v) \mathrm {d}v . \end{aligned}$$
(4.52)

Here,

$$\begin{aligned} -\Delta _{x}\varphi _{a } = a^{5}&\ \ \text {with} \ \ \frac{\partial \varphi _{a }}{ \partial n}\Big |_{\partial \Omega }=0, \end{aligned}$$
(4.53)
$$\begin{aligned} -\Delta _{x}\varphi _{b }^{j} =b^{5}_{j}&\ \ \text {with} \ \ \varphi _{b}^{j}|_{\partial \Omega }=0, \end{aligned}$$
(4.54)
$$\begin{aligned} -\Delta _{x}\varphi _{c } =c^{5}&\ \ \text {with} \ \ \varphi _{c }|_{\partial \Omega }=0. \end{aligned}$$
(4.55)

A unique solvability to the above Poisson equations when \((a,b,c) \in L^6(\Omega )\) and an estimate

$$\begin{aligned}&\Vert \nabla _x^2 \varphi _{(a,b,c)}\Vert _{L^{6/5}(\Omega )} + \Vert \nabla _x \varphi _{(a,b,c)}\Vert _{L^2(\Omega )} + \Vert \varphi _{(a,b,c)}\Vert _{ L^6(\Omega ) }\nonumber \\&\quad \lesssim \Vert |{\tilde{P}}f_R|^5\Vert _{L^{6/5}(\Omega )} \lesssim \Vert {\tilde{P}}f_R\Vert _{L^6(\Omega )}^5 . \end{aligned}$$
(4.56)

is a direct consequence of Lax-Milgram and suitable extension (extend \(a^5\) of (4.53) evenly in \(x_3 \in {\mathbb {R}}\), and \(b^5\) and \(c^5\) of (4.54) and (4.55) oddly in \(x_3 \in {\mathbb {R}}\), then solve the Poisson equation, and then restrict the whole space solutions to the half space \(x_3>0\)) and a standard elliptic estimate \((L^{\frac{6}{5}}(\Omega ) \rightarrow {\dot{W}}^{2,\frac{6}{5}} (\Omega ) \cap {\dot{W}}^{1,2} (\Omega ) \cap L^6(\Omega ))\).

From \(M_{1,\varepsilon u , 1}(v) = M_{1,0 , 1}(v) + O( \varepsilon ) |u| | v-\varepsilon u | M_{1,\varepsilon u , 1}(v) \) we can easily check that

$$\begin{aligned} |{\mathbf {P}} f_R (t,x,v)- {\tilde{\mathbf {P}}}f_R (t,x,v)|\lesssim \varepsilon |u(t,x)| |v-\varepsilon u| \sqrt{\mu } | f_R(t,x,v)|. \end{aligned}$$
(4.57)

Therefore we have

$$\begin{aligned} \begin{aligned} \Vert P f_R(t) \Vert _{ L^6_x} \lesssim&\Vert {\mathbf {P}} f_R (t) \Vert _{L^6_{x,v}} \lesssim \Vert {\tilde{\mathbf {P}}} f_R (t) \Vert _{L^6_{x,v}} \\&+ \varepsilon \Vert u(t) \Vert _{\infty } \{ \Vert {P} f_R (t) \Vert _{L^6_{x }} + \Vert ( {\mathbf {I}}-{\mathbf {P}}) f_R (t) \Vert _{L^6_{x,v}} \}\\ \lesssim&(1+ \varepsilon \Vert u \Vert _\infty ) \Vert {{\tilde{P}}} f_R (t) \Vert _{L^6_{x }} + \varepsilon \Vert u(t) \Vert _{\infty } \Vert ( {\mathbf {I}}-{\mathbf {P}}) f_R (t) \Vert _{L^6_{x,v}} . \end{aligned} \end{aligned}$$
(4.58)

Note that \( \Vert ( {\mathbf {I}}-{\mathbf {P}}) f_R (t) \Vert _{L^6_{x,v}} \le \Vert ( {\mathbf {I}}-{\mathbf {P}}) f_R (t) \Vert _{L^\infty _{x,v}} ^{2/3} \Vert ( {\mathbf {I}}-{\mathbf {P}}) f_R (t) \Vert _{L^2_{x,v}} ^{1/3} \lesssim o(1) (\kappa \varepsilon )^{1/2} \Vert {\mathfrak {w}} f_R (t) \Vert _{L^\infty _{x,v}} + (\kappa \varepsilon )^{-1} \Vert ( {\mathbf {I}}-{\mathbf {P}}) f_R (t) \Vert _{L^2_{x,v}}\). Hence to prove the lemma and (4.44) it suffices to prove the same bound for \(\Vert {\tilde{P}}f_R \Vert _{L^6_{x,v}}:=\Vert ( a,b,c) \Vert _{L^6_{x }} \).

Following the direct computations in the proof of Lemma 2.12 in [13] we derive that

$$\begin{aligned} (4.46)_1 = {\left\{ \begin{array}{ll} -5 \Vert a(t) \Vert _{6}^6 +o(1) \Vert {\tilde{\mathbf {P}}} f_R(t) \Vert _6^6 +O(1) \Vert ({\mathbf {I}} - \mathbf { {P}}) f_R(t) \Vert _6^6 &{} \text {if } \psi =\psi _a,\\ -2 \int _\Omega b_i \partial _i \partial _j \varphi _b^j +o(1) \Vert {\tilde{\mathbf {P}}} f_R(t) \Vert _6^6 +O(1) \Vert ({\mathbf {I}} - \mathbf { {P}}) f_R(t) \Vert _6^6 &{} \text {if } \psi =\psi _{b,1}^{i,j},\\ \int _\Omega b_j \partial _{i} \partial _j \varphi _b^i + \int _\Omega b_i \partial _{j} \partial _j \varphi _b^i +O(1) \Vert ({\mathbf {I}} - \mathbf { {P}}) f_R(t) \Vert _6^6 &{} \text {if } \psi =\psi _{b,2}^{i,j} \ \text {and} \ i \ne j,\\ 5 \Vert c(t)\Vert _6^6 +o(1) \Vert {\tilde{\mathbf {P}}} f_R(t) \Vert _6^6 + O(1) \Vert ({\mathbf {I}} - \mathbf { {P}}) f_R(t) \Vert _6^6 &{} \text {if } \psi =\psi _c. \end{array}\right. }\nonumber \\ \end{aligned}$$
(4.59)

For \(\Vert b_i\Vert _{6}^6\), using the second and third estimate of (4.59) we deduce that

$$\begin{aligned} \begin{aligned} \Vert b_i\Vert _{L^6(\Omega )}^6&= - \int _\Omega b_i \Delta _x\varphi _b^i \mathrm {d}x = - \int _\Omega b_i \partial _i^2\varphi _b^i \mathrm {d}x - \sum _{j(\ne i)} \int _\Omega b_i \partial _j^2\varphi _b^i \mathrm {d}x \\&= \frac{1}{2}\sum _{j}(4.46)_1|_{\psi ^{j,i}_{b,1}} - \sum _{j(\ne i)}{(4.46)_1|_{\psi ^{i,j}_{b,2}}} + o(1) \Vert {\tilde{\mathbf {P}}} f_R(t) \Vert _6^6 \\&\quad +O(1) \Vert ({\mathbf {I}} - \mathbf { {P}}) f_R(t) \Vert _6^6. \end{aligned} \end{aligned}$$
(4.60)

Now we consider the boundary term \((4.46)_2\). From (4.48)-(4.51) and (4.52)

$$\begin{aligned} \int _{\gamma } \psi P_{\gamma _+} f_R= {\left\{ \begin{array}{ll} \int _{\partial \Omega }\partial _n \varphi _a \int _{{\mathbb {R}}^3} (|v|^2 - \beta _a) (v\cdot n)^2 \mu _0 \mathrm {d}v \mathrm {d}S_x =0 &{} \text {if } \psi =\psi _a,\\ 0&{} \text {if } \psi =\psi _{b,1}^{i,j} \ \text {or} \ \psi _{b,2}^{i,j},\\ \int _{\partial \Omega }\partial _n \varphi _c \int _{{\mathbb {R}}^3} (|v|^2 - \beta _c) (v\cdot n)^2 \mu _0 \mathrm {d}v \mathrm {d}S_x =0 &{} \text {if } \psi =\psi _c. \end{array}\right. }\nonumber \\ \end{aligned}$$
(4.61)

Here we have used the Neumann boundary condition of (4.53) for \(\psi _a\), and the last identity in (4.52) for \(\psi _c\). For \(\psi _{b,1}^{i,j}\) or \(\psi _{b,2}^{i,j}\) we used the fact that the integrands are odd in v. From (3.8), we decompose \(f|_{\gamma }= P_{\gamma _+} f + {\mathbf {1}}_{\gamma _+} (1- P_{\gamma _+})f - {\mathbf {1}}_{\gamma _-}\frac{\varepsilon }{\delta } (1- P_{\gamma _+}) f_2\). From (4.61) together with (3.57) and (3.58) we have

(4.62)

where we have used \(|\int _{\gamma _+} \psi (1-P_{\gamma _+}) f|\lesssim |\nabla _x \varphi |_{L^{ {4} / {3}}(\partial \Omega ) } |(1-P_{\gamma _+}) f|_{4, \gamma _+}\) at the last line. Here \(\varphi \in \{ \varphi _a, \varphi _b, \varphi _c\}\). For the first term of (4.62) we interpolate

$$\begin{aligned} |(1-P_{\gamma _+}) f_R|_{4, \gamma _+}\lesssim | \varepsilon ^{-\frac{1}{2}}(1-P_{\gamma _+}) f_R|_{2, \gamma _+}^{1/2} \varepsilon ^{\frac{1}{4}} \Vert {\mathfrak {w}}_{\varrho , \ss } f_R\Vert _\infty ^{1/2}. \end{aligned}$$
(4.63)

For the second term of (4.62), we use (4.56) and a trace theorem \(({\dot{W}}^{1,\frac{6}{5}} ({\mathbb {T}}^2\times {\mathbb {R}}_+) \cap L^2({\mathbb {T}}^2\times {\mathbb {R}}_+) \rightarrow {W}^{1- \frac{1}{6/5},\frac{6}{5}} ({\mathbb {T}}^2 )) \), and the Sobolev embedding \((W^{\frac{1}{6},\frac{6}{5}}( {\mathbb {T}}^2) \rightarrow L^{4/3} (\mathbb {T}^2))\) to conclude that

$$\begin{aligned} |\nabla _x \varphi |_{L^{\frac{4}{3}}( {\mathbb {T}}^2)}\lesssim |\nabla _x \varphi |_{W^{\frac{1}{6},\frac{6}{5}}( {\mathbb {T}}^2)}\lesssim \Vert \nabla _x \varphi \Vert _{{\dot{W}}^{1, \frac{6}{5}}( {\mathbb {T}}^2 \times {\mathbb {R}}_+) \cap L^2(\Omega ) } \lesssim \Vert {\tilde{P}}f_R \Vert _{L^6({\mathbb {T}}^2 \times {\mathbb {R}}_+)}^5 .\nonumber \\ \end{aligned}$$
(4.64)

Next we consider \((4.46)_3\). For \(\psi \) of (4.48)-(4.51) and \(\varphi \) of (4.53)-(4.55), using (4.56), it follows that

$$\begin{aligned} \begin{aligned} |(4.46)_3|&\lesssim \varepsilon \Vert \partial _t f_R \Vert _{L^{2}_{x,v}} \Vert \psi \Vert _{L^{2}_{x,v}} \lesssim \varepsilon \Vert \partial _t f_R \Vert _{L^{2}_{x,v}}\Vert \nabla _x \varphi \Vert _{L^{2}_{x }} \lesssim \varepsilon \Vert \partial _t f_R \Vert _{L^{2}_{x,v}} \Vert {\tilde{P}}f_R \Vert _{L^6_x}^{5}\\&\le O(1) [\varepsilon \Vert \partial _t f_R \Vert _{L^{2}_{x,v}}]^6 + o(1) \Vert {\tilde{P}}f_R \Vert _{L^6_x}^{6}. \end{aligned}\nonumber \\ \end{aligned}$$
(4.65)

Lastly we consider the right hand side of (4.46). From (1.21), (3.23), (3.37), and (4.56), it follows

$$\begin{aligned} \begin{aligned}&\Big |\iint _{\Omega \times {\mathbb {R}}^3} \psi \frac{1}{\varepsilon \kappa } Lf_R\Big | =\Big | \iint _{\Omega \times {\mathbb {R}}^3} \psi \frac{1}{\varepsilon \kappa } L ({\mathbf {I}} - {\mathbf {P}})f_R\Big |\\&\quad \lesssim \frac{1}{\varepsilon \kappa } \int _\Omega \int _{{\mathbb {R}}^3} |\nabla _x \varphi _{(a,b,c)}(x)| \mu (v)^{1/4} \Big [ \nu (v ) |({\mathbf {I}} - {\mathbf {P}})f_R (x,v ) |\\&\qquad +\int _{{\mathbb {R}}^3} k_\vartheta (v,v_*) |({\mathbf {I}} - {\mathbf {P}})f_R (x,v_*) | \mathrm {d}v_*\Big ] \mathrm {d}v \mathrm {d}x \\&\quad \lesssim \frac{1}{\varepsilon \kappa } \Vert \nabla _x \varphi _{(a,b,c)} \Vert _{L^2_x} \Vert ({\mathbf {I}} - {\mathbf {P}}) f_R\Vert _{L^2_{x,v}} \lesssim \frac{1}{\varepsilon \kappa } \Vert {\tilde{P}}f \Vert _{L^6_x}^5\Vert ({\mathbf {I}} - {\mathbf {P}}) f_R\Vert _{L^2_{x,v}}\\&\quad \le o(1) \Vert {\tilde{P}}f _R\Vert _{L^6_x}^6+ \big [ \varepsilon ^{-1} \kappa ^{-1} \Vert ({\mathbf {I}} - {\mathbf {P}}) f_R\Vert _{L^2_{x,v}}\big ]^6 . \end{aligned} \end{aligned}$$
(4.66)

Note that, from (3.37), \(|\Gamma (\frac{\varepsilon }{\kappa } {f_2}, f_R) |\lesssim \frac{\varepsilon }{\kappa } \Vert {\mathfrak {w}}_{\varrho , \ss } f_2 \Vert _{\infty } {\mathfrak {w}}_{\varrho , \ss }(v)^{-1} \big [\nu (v) f_R(v)+ \int _{{\mathbb {R}}^3} k_{\vartheta } (v,v_*) f_R(v_*) \mathrm {d}v_*\big ].\) Then from (3.57) and (3.58)

$$\begin{aligned} \begin{aligned} \Big |\iint _{\Omega \times {\mathbb {R}}^3} \psi \frac{\varepsilon }{\kappa } \Gamma ({f_2}, f_R) \Big |&\lesssim \Vert \nabla _x \varphi _{(a,b,c)} \Vert _{L^2_x} \frac{\varepsilon }{\kappa } \Vert (3.10)\Vert _\infty \Vert f_R \Vert _{L^2_{x,v}}\\&\le o(1) \Vert {\tilde{P}}f_R \Vert _{L^6_x}^6+ \big [ \varepsilon \kappa ^{-1} \Vert (3.10)\Vert _\infty \Vert f_R\Vert _{L^2_{x,v}}\big ]^6. \end{aligned}\nonumber \\ \end{aligned}$$
(4.67)

For the contribution of \(\Gamma (f_R,f_R)\) we decompose \(f_R= {\mathbf {P}} f_R + ({\mathbf {I}} - {\mathbf {P}}) f_R\). From (3.37) (or (3.36))

$$\begin{aligned} \begin{aligned}&|\Gamma (f_R,f_R)(v)|\\&\quad \lesssim |\Gamma ({\mathbf {P}}f_R,{\mathbf {P}}f_R)(v)| + |\Gamma (( {\mathbf {I}}-{\mathbf {P}})f_R,( {\mathbf {I}}-{\mathbf {P}})f_R)(v)|\\&\quad \lesssim \nu (v) |Pf_R |^2+ \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R \Vert _\infty \Big \{ \nu (v) |( {\mathbf {I}}-{\mathbf {P}})f_R)(v)| \\&\ \ \ \ + \int _{{\mathbb {R}}^3} k_\vartheta (v,v_*) |( {\mathbf {I}}-{\mathbf {P}})f_R)(v_*)|\mathrm {d}v_* \Big \}. \end{aligned} \end{aligned}$$
(4.68)

Then from (4.48)-(4.51), (3.23), and the Hölder’s inequality (\(1=1/2+ 1/3+ 1/6\))

$$\begin{aligned} \begin{aligned}&\Big |\iint _{\Omega \times {\mathbb {R}}^3} \psi \frac{ \delta }{ \kappa }\Gamma (f_R, f_R) \Big | \\&\quad \lesssim \frac{\delta }{ \kappa } \Vert \nabla _x \varphi _{(a,b,c)}\Vert _{L^2_x}\Big \{ \Vert Pf_R \Vert _{L^3_x} \Vert Pf_R \Vert _{L^6_x} + \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R \Vert _{L^\infty _{x,v}} \Vert ( {\mathbf {I}}-{\mathbf {P}})f_R \Vert _{L^2_{x,v}} \Big \}\\&\quad \lesssim \frac{\delta }{ \kappa } \Vert {\tilde{P}}f_R \Vert _{L^6_x}^{5} \Vert Pf_R\Vert _{L_x^6}^{3/2} \Vert Pf_R \Vert _{L_x^2}^{1/2} \\&\qquad + \frac{\varepsilon \delta }{ \kappa ^{1/2}} \Vert {\tilde{P}}f_R \Vert _{L^6_x}^{5} \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R \Vert _{L^\infty _{x,v}} \Vert \varepsilon ^{-1} \kappa ^{-1/2}( {\mathbf {I}}-{\mathbf {P}})f_R \Vert _{L^2_{x,v}}, \end{aligned}\nonumber \\ \end{aligned}$$
(4.69)

where we have used an interpolation \(\Vert Pf_R\Vert _{L^3}\le \Vert Pf_R\Vert _{L^6}^{1/2} \Vert Pf_R \Vert _{L^2}^{1/2}\) and (4.56) at the last step. A contribution of the rest of terms in the r.h.s of (4.46) can be easily bounded as, from (3.4) and (3.5),

$$\begin{aligned} \begin{aligned}&\iint _{\Omega \times {\mathbb {R}}^3} |\psi | \left| \frac{( \varepsilon \partial _t + v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_{R} - \varepsilon ({\mathbf {I}}- {\mathbf {P}}){\mathfrak {R}}_1 - \varepsilon {\mathfrak {R}}_2\right| \\&\quad \lesssim \ \Vert Pf_R \Vert _{L^6_x}^{5}\Big \{ \varepsilon \Vert (3.12)\Vert _{\infty } \Vert f_R \Vert _{L^2_{x,v}} + \varepsilon \Vert (3.4) + (3.5) \Vert _{L^2_{x,v}} \Big \}. \end{aligned} \end{aligned}$$
(4.70)

In conclusion, collecting the terms from (4.59) with (4.60), (4.62) with (4.63) and (4.64), (4.65), (4.66), (4.67), (4.69), (4.70), and utilizing (4.58), and two facts from (A.1):

$$\begin{aligned} \begin{aligned} \sup _{0 \le s \le t} \Vert ({\mathbf {I}} -{\mathbf {P}}) f_R(s ) \Vert _{L^2_{x,v}}&\lesssim \Vert ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} + \Vert ({\mathbf {I}} -{\mathbf {P}})\partial _t f_R \Vert _{L^2_{t,x,v}}\\&\quad + \varepsilon \Vert \partial _t u \Vert _{L^\infty _{t,x}}\Vert {P} f_R \Vert _{L^2_{t,x}} ,\\ \sup _{0 \le s \le t} |(1- P_{\gamma _+}) f_R(s)|_{L^2({\gamma _+})}&\lesssim \sup _{0 \le s \le t} | f_R(s)|_{L^2({\gamma _+})}\\&\lesssim | f_R |_{L^2_tL^2({\gamma _+})} + |\partial _t f_R |_{L^2_tL^2({\gamma _+})}, \end{aligned} \end{aligned}$$
(4.71)

we prove (4.44).\(\square \)

4.3 Average in Velocity

We prove a version of velocity lemma when a suitable bound for source terms is only known in a finite time interval. In this section we often specify domains in which an \(L^p\)-norm is taken while the simplified notation (4.1) will be used only when the domain is \([0,T] \times \Omega \times {\mathbb {R}}^3\).

Proposition 9

Assume the same assumptions in Proposition 6. Then we have, for \(2<p<3\),

$$\begin{aligned} \begin{aligned}&d_3\big \Vert {P} f_R \big \Vert _{L^2_t L^p_x } \\&\quad \lesssim \ (1+ \varepsilon \Vert (3.12) \Vert _{L^2_t L^\infty _x } ) \Vert f_R \Vert _{L^\infty _t L^2 _{x,v}} \\&\qquad +\Big \{ \frac{1}{\varepsilon \kappa } + \frac{\delta }{\kappa } \Vert {\mathfrak {w}}_{\varrho , \ss } f_R \Vert _{L^\infty _{t,x,v} } + \Vert {\mathfrak {w}}_{\varrho , \ss } f_R \Vert _{ L^\infty _{t,x,v}}^{\frac{p-2}{p}} \Big \} \Vert \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}})f_R \Vert _{L^2_{t,x,v} } \\&\qquad + \Vert f_R (0) \Vert _{L^2_\gamma } + \varepsilon \Vert (3.6)\Vert _{L^2_{t,x} }+\varepsilon \Vert (3.7) \Vert _{L^2_{t,x} }, \end{aligned} \end{aligned}$$
(4.72)

with

$$\begin{aligned} d_{3}:= 1- O(\varepsilon ) \Vert u\Vert _{L^\infty _{t,x}} - \frac{\varepsilon }{\kappa } \Vert (3.10) \Vert _{L^\infty _t L_x^{\frac{2p}{p-2}} } - \frac{\delta }{\kappa } \Vert Pf_R \Vert _{L^\infty _tL_x^{6} }^{\frac{3(p-2)}{p}} \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R \Vert _{L_{t,x,v}^{\infty } }^{\frac{6-2p}{p}},\nonumber \\ \end{aligned}$$
(4.73)

and for \(\varrho '<\varrho \)

$$\begin{aligned} \begin{aligned}&d_{3,t} \big \Vert {P} \partial _t f_R \big \Vert _{L^2_t L^p_x } \\&\quad \lesssim \frac{1}{\kappa } \Vert \partial _t u \Vert _{ L^\infty _{t, x} }\big ( 1+ \varepsilon ^2 \Vert (3.10) \Vert _{L^\infty _{t,x}} \big )\Vert Pf_R \Vert _{ L^2_{t, x} } \\&\qquad + \frac{\delta \varepsilon }{\kappa } \Vert \partial _t u \Vert _{L^\infty _{t,x} }\Vert {P} f_R \Vert _{L^\infty _tL^6_{x }}^{\frac{3(p-2)}{p}}\Vert {\mathfrak {w}} f_R\Vert _{L^\infty _{t,x,v}}^{\frac{6-2p}{p}}\Vert {P} f_R \Vert _{L^2_tL^p_{x }} \\&\qquad + \frac{\varepsilon }{\kappa } \Vert \partial _t u \Vert _{L^\infty _{t,x} } (\delta \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}} + \Vert (3.10) \Vert _{L^\infty _{t,x}}) \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \\&\qquad + (\kappa \varepsilon ) ^{\frac{2}{p-2}} \Vert {\mathfrak {w}}_{\varrho ^\prime ,\ss } \partial _t f_R \Vert _{L^2_t L^\infty _{x,v} }+ \Vert \partial _t f_R \Vert _{L^\infty _t L^2_{x,v}}+ \varepsilon \Vert (3.13)\Vert _{L^2_tL^\infty _{x}} \Vert f_R\Vert _{L^\infty _t L^2_{x,v}} \\&\qquad + \Big \{\frac{1}{\kappa \varepsilon } + \frac{\delta }{\kappa } \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R\Vert _{L^\infty _{t,x,v}} + \frac{\varepsilon }{\kappa } \Vert (3.10)\Vert _{L^\infty _{t,x}} \\&\qquad + \varepsilon \Vert (3.12) \Vert _{L^\infty _{t,x} } \Big \} \Vert \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \\&\qquad + \Vert \partial _t f_R (0) \Vert _{L^2_\gamma } + \frac{\varepsilon }{\kappa } \Vert (3.11) \Vert _{L^2_{t,x,v}} \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R \Vert _{L^\infty _{t,x,v}} + \varepsilon \Vert (3.6)\Vert _{L^2_{t,x}} + \varepsilon \Vert (3.7)\Vert _{L^2_{t,x}}, \end{aligned} \end{aligned}$$
(4.74)

with

$$\begin{aligned} d_{3,t}:= & {} 1- O(\varepsilon ) \Vert u\Vert _{L^\infty _{t,x}} - \frac{\varepsilon }{\kappa } \Vert (3.10) \Vert _{L^\infty _t L_x^{\frac{2p}{p-2}} }-\varepsilon \Vert (3.12)\Vert _{L^\infty _{t } L^{\frac{2p}{p-2}}_x } \nonumber \\&- \frac{\delta }{\kappa } \Vert P f_R \Vert _{L^\infty _t L_x^{6} }^{\frac{3(p-2)}{p}} \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R \Vert _{L^{\infty } _{t,x,v}}^{\frac{6-2p}{p}}, \end{aligned}$$
(4.75)

where both bounds are uniform-in-p for \(2<p<3\).

We prove the proposition by several steps.

Step 1: Extension. We define a subset

$$\begin{aligned} {\tilde{\Omega }} := (0,2\pi )\times (0,2\pi ) \times (0,\infty ) \subset {\mathbb {R}}^3. \end{aligned}$$
(4.76)

We regard \({\tilde{\Omega }}\) as an open subset but not a periodic domain as \(\Omega \). Without loss of generality we may assume that \(f_R(0,x,v)\) is defined in \({\mathbb {R}}^3 \times {\mathbb {R}}^3\) and \(\Vert f_R(0) \Vert _{L^p({\mathbb {R}}^3 \times {\mathbb {R}}^3)} \lesssim \Vert f_R(0) \Vert _{L^p({\tilde{\Omega }} \times {\mathbb {R}}^3)}\) for all \(1 \le p \le \infty \). Then we extend a solution for whole time \(t \in {\mathbb {R}}\) as

$$\begin{aligned} f_I (t,x,v) : = {\mathbf {1}}_{t \ge 0 } f_R (t,x,v) + {\mathbf {1}}_{t \le 0} \chi _1 (t) f_R (0, x ,v ), \end{aligned}$$
(4.77)

where a smooth non-negative function \(\chi _1\) satisfies \(\chi _1 (t)\equiv 1\) for \(t \in [-1,0]\), \(\chi _1 (t) \equiv 0\) for \(t<-2\), and \(0 \le \frac{d}{dt}\chi _1 \le 4\).

A closure of \({\tilde{\Omega }}\) is given as \(cl({\tilde{\Omega }})= [0,2\pi ]\times [0,2\pi ] \times [0,\infty )\). Let us define \({\tilde{t}}_B(x,v) \in {\mathbb {R}}\) for \((x,v) \in ({\mathbb {R}}^3 \backslash {\tilde{\Omega }}) \times {\mathbb {R}}^3\). We consider \({\tilde{B}}(x,v):=\{s \in {\mathbb {R}}:x+ s v \in {\mathbb {R}}^3 \backslash cl({\tilde{\Omega }}) \}\). Clearly if \({\tilde{B}}(x,v) \ne \emptyset \) then \( \{s>0\}\subset {\tilde{B}} (x,v)\) or \( \{s<0\}\subset {\tilde{B}} (x,v)\) exclusively. If \( \{s>0\} \subset {\tilde{B}} (x,v)\), let \( I_+\) be the largest interval such that \( \{s > 0 \} \subset I_+ \subset {{\tilde{B}}}(x,v) \). And if \( \{s < 0\} \subset {\tilde{B}} (x,v)\), let \( I_-\) be the largest interval such that \( \{s > 0 \} \subset I_- \subset {{\tilde{B}}}(x,v)\).

We define

$$\begin{aligned} {\tilde{t}}_B(x,v) = {\left\{ \begin{array}{ll} 0 &{} \text {if} \ \ x \in \partial {\tilde{\Omega }}, \\ \inf {\tilde{B}}(x,v) &{} \text {if} \ \ x \in {\mathbb {R}}^3 \backslash cl({\tilde{\Omega }}) \ \text {and} \ {\tilde{B}}(x,v) \ne \emptyset \ \text {and} \ \{s>0\} \subset I_+\subset {\tilde{B}}(x,v) ,\\ \sup {\tilde{B}}(x,v) &{} \text {if} \ \ x \in {\mathbb {R}}^3 \backslash cl({\tilde{\Omega }}) \ \text {and} \ {\tilde{B}}(x,v) \ne \emptyset \ \text {and} \ \{s<0\} \subset I_-\subset {\tilde{B}}(x,v) ,\\ - \infty &{} \text {if} \ \ {\tilde{B}}(x,v) = \emptyset \ \text {and} \ x \notin \partial {\tilde{\Omega }}. \end{array}\right. }\nonumber \\ \end{aligned}$$
(4.78)

Using (4.78) we define

$$\begin{aligned}&f_E (t,x,v): = {\mathbf {1}}_{(x,v) \in ({\mathbb {R}}^3 \backslash {\tilde{\Omega }}) \times {\mathbb {R}}^3 } f_I (t+ \varepsilon {\tilde{t}}_B(x,v), {\tilde{x}}_B(x,v),v) \ \ \nonumber \\&\quad \text {with} \ \ {\tilde{x}}_B(x,v):= x+{\tilde{t}}_B(x,v)v. \end{aligned}$$
(4.79)

It is easy to see that \(\varepsilon \partial _t f_E + v\cdot \nabla _x f_E=0\) in the sense of distributions.

Next we define two cutoff functions. For any \(N>0\) we define smooth non-negative functions as

$$\begin{aligned} \begin{aligned}&\chi _2 (x) \equiv 1 \ \text {for} \ x \in [-\pi , 3\pi ] \times [-\pi , 3\pi ] \times [-\pi , \infty ), \\&\chi _2 (x ) \equiv 0 \ \text {for} \ x \notin [-2\pi , 4\pi ] \times [-2\pi , 4\pi ] \times [-2\pi , \infty ) , \ \ |\nabla _{x}\chi _2 | \le 10 ,\\ \end{aligned}\end{aligned}$$
(4.80)
$$\begin{aligned} \begin{aligned}&\chi _3 (v) \equiv 1 \ \text {for} \ |v| \le N-1, \ \text {and} \ |v_i|\ge 2/N \ \text {for all } i=1,2,3,\\&\chi _3 (v) \equiv 0 \ \text {for} \ |v| \ge N \ \text {or} \ |v_i| \le 1/N \ \text {for any } i=1,2,3, \ \ |\nabla _v \chi _3 | \le 10 . \end{aligned} \end{aligned}$$
(4.81)

We denote

$$\begin{aligned}&U : = [-2\pi , 4\pi ] \times [-2\pi , 4\pi ] \times [-2\pi , \infty ), \ \ \nonumber \\&V: = \{ v\in {\mathbb {R}}^3: |v| \le N \} \cap \bigcap _{i=1,2,3} \{ v\in {\mathbb {R}}^3: |v_i| \ge 1/N \} \end{aligned}$$
(4.82)

We define an extension of cut-offed solutions

$$\begin{aligned}&{\bar{f}}_{R}(t,x,v) := \chi _2 (x) \chi _3 (v) \big \{ {\mathbf {1}}_{{\tilde{\Omega }}}(x) f_I(t,x,v) + f_E (t,x,v)\big \} \ \ \nonumber \\&\quad \text {for} \ (t,x,v) \in (-\infty ,T] \times {\mathbb {R}}^3 \times {\mathbb {R}}^3 . \end{aligned}$$
(4.83)

We note that in the sense of distributions \({\bar{f}}_R\) solves

$$\begin{aligned} \begin{aligned}&\varepsilon \partial _t {\bar{f}}_R + v\cdot \nabla _x {\bar{f}}_R = {\bar{g}} \ \ \text {in} \ (-\infty ,T] \times {\mathbb {R}}^3 \times {\mathbb {R}}^3,\\&{\bar{g}}:= \frac{v\cdot \nabla _x \chi _2 }{\chi _2 } {\bar{f}}_R + {\mathbf {1}}_{t \ge 0} {\mathbf {1}}_{{\tilde{\Omega }}} (x) \chi _2(x) \chi _3(v) [\varepsilon \partial _t + v\cdot \nabla _x] f_R\\&\qquad +{\mathbf {1}}_{t \le 0} \{ \varepsilon \partial _t \chi _1(t) f_R (0,x,v) + \chi _1 (t) v\cdot \nabla _x f_R (0,x,v) \} \end{aligned} \end{aligned}$$
(4.84)

Here we have used the fact that \({\bar{f}}_R\) in (4.84) is continuous along the characteristics across \(\partial {\tilde{\Omega }}\) and \(\{t=0\}\). We derive that, using (4.84),

$$\begin{aligned} {\bar{f}}_R(t,x,v) = \frac{1}{\varepsilon } \int ^t_{-\infty } {\bar{g}} (s, x- \frac{t-s}{\varepsilon } v, v) \mathrm {d}s \ \ \text {for} \ (t,x,v) \in (-\infty ,T] \times {\mathbb {R}}^3 \times {\mathbb {R}}^3.\nonumber \\ \end{aligned}$$
(4.85)

Recall \({\tilde{\varphi }}_i \in \{{\tilde{\varphi }}_0, \cdots {\tilde{\varphi }}_4\}\) in (4.47). From (4.83) we note that

$$\begin{aligned} \begin{aligned}&\left\| \int _{{\mathbb {R}}^3} {\bar{f}}_R(t,x,v) {\tilde{\varphi }}_i (v) \sqrt{\mu _0(v)} \mathrm {d}v \right\| _{L^2_t ((0,T); L^p_x ( {\tilde{\Omega }}))} \\&\quad = \left\| \int _{{\mathbb {R}}^3} \chi _2 (x) \chi _3 (v) f_R(t,x,v) {\tilde{\varphi }}_i (v) \sqrt{\mu _0(v)} \mathrm {d}v \right\| _{L^2_t ((0,T); L^p_x ({\tilde{\Omega }}))} \end{aligned} \end{aligned}$$
(4.86)

From (1.24), we decompose

$$\begin{aligned} \begin{aligned} (4.86)\ge&\ \Big \Vert \sum _j \chi _2 (x) {\tilde{P}}_jf_R(t,x) \int _{{\mathbb {R}}^3} \chi _3 (v) {\tilde{\varphi }}_j (v) {\tilde{\varphi }} _i (v) \mu _0 (v) \mathrm {d}v \Big \Vert _{L^2_t ((0,T); L^p_x ({\tilde{\Omega }}))} \\&- \left\| \int _{{\mathbb {R}}^3} \chi _3 (v) ({\mathbf {I}} - {\tilde{\mathbf {P}}}) f_R(t,x,v) {\tilde{\varphi }}_i (v) \sqrt{\mu _0(v)} \mathrm {d}v \right\| _{L^2_t ((0,T); L^p_x ({\tilde{\Omega }}))}. \end{aligned} \end{aligned}$$

We consider the right hand side of above terms. From (4.57), \(\int {\tilde{\varphi }}_i{\tilde{\varphi }}_j \mu _0 = \delta _{ij}\), and (4.81), the first term can be bounded below by \(\big (1-O( \varepsilon ) \Vert u\Vert _\infty - O(\frac{1}{N})\big ) \big \Vert \chi _2 {P} f_R \big \Vert _{L^2_t ((0,T); L^p_x ({\tilde{\Omega }}))}\). For the second term we use (4.57), \(L^2_t (0,T) \subset L^p_t( 0,T )\), and \(L^1(\{|v|\le N\}) \subset L^p(\{|v|\le N\})\) to bound it above by \(C_{T,N} \Vert ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^p((0, T) \times {\tilde{\Omega }} \times {\mathbb {R}}^3)}+ \big ( O( \varepsilon ) \Vert u\Vert _\infty + O(\frac{1}{N})\big ) \big \Vert {P} f_R \big \Vert _{L^2_t( (0,T) ;L^p_x({\tilde{\Omega }}))}\). Hence we derive

$$\begin{aligned} \begin{aligned}&(4.86) \\&\quad \ge \ \big (1- O(\varepsilon )\Vert u\Vert _\infty - O(\frac{1}{N})\big ) \big \Vert {P} f_R \big \Vert _{L^2_t( (0,T) ;L^p_x({\tilde{\Omega }}))} \\&\qquad - C_{T,N} \Vert ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^p((0, T) \times {\tilde{\Omega }} \times {\mathbb {R}}^3)} \\&\quad \ge \ \big (1- O(\varepsilon ) \Vert u\Vert _\infty - O(\frac{1}{N})\big ) \big \Vert {P} f_R \big \Vert _{L^2_t((0,T);L^p_x({\tilde{\Omega }}))}\\&\qquad -C_{T,N} \Vert {\mathfrak {w}}_{\varrho , \ss } f_R (t) \Vert _{ L^\infty ((0,T) \times {\tilde{\Omega }} \times {\mathbb {R}}^3)}^{\frac{p-2}{p}} \Vert ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2((0, T) \times {\tilde{\Omega }} \times {\mathbb {R}}^3)}^{\frac{2}{p}} . \end{aligned} \end{aligned}$$
(4.87)

Step 2: Average lemma. Recall \({\tilde{\varphi }}_i \in \{{\tilde{\varphi }}_0, \cdots {\tilde{\varphi }}_4\}\) in (4.47). We choose \({\tilde{\varphi }}(v)\) such that

$$\begin{aligned} \begin{aligned}&\chi _3 (v) |{\tilde{\varphi }}_i(v)| \sqrt{\mu _0 (v)} \le {\tilde{\varphi }}(v) , \ \ {\tilde{\varphi }}(v) \in C^\infty _c ({\mathbb {R}}^3) \\&\quad \text {and} \ \ {\tilde{\varphi }}(v)\equiv 0 \ \ \text {for} \ \ |v|\ge N \ \ \text {or} \ \ |v_i| \le 1/N \ \text {for any } i=1,2,3. \end{aligned} \end{aligned}$$
(4.88)

Lemma 6

We define

$$\begin{aligned} S({\bar{g}})(t,x):= \frac{1}{\varepsilon } \int ^t_{-\infty }\int _{{\mathbb {R}}^3} | {\bar{g}}(s, x- \frac{t-s}{\varepsilon } v, v) | {\tilde{\varphi }} (v) \mathrm {d}v \mathrm {d}s \ \ \text {for} \ (t,x) \in (-\infty , T ] \times {\mathbb {R}}^3.\nonumber \\ \end{aligned}$$
(4.89)

Then, for \(p<3\) and \(1\ll N\),

$$\begin{aligned} \Vert S({\bar{g}}) \Vert _{L^2_t((0,T); L^p_x({\mathbb {T}}^2 \times {\mathbb {R}}))} \lesssim _{N} \Vert {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} {\bar{g}}\Vert _{L^2 ((0,T)\times ({\mathbb {T}}^2 \times {\mathbb {R}}) \times \{|v| \le N \})}, \end{aligned}$$
(4.90)

where the bound (4.90) only depends on N but can be independent on \(p<3\).

We remark that from (4.85) and (4.89) \( \int _{{\mathbb {R}}^3}{\bar{f}}_R(t,x,v) {\tilde{\varphi }}_i (v) \mathrm {d}v \le S({\bar{g}})(t,x)\).

Proof of Lemma 6

We prove (4.90) by a \(TT^*\)(\(SS^*\) for our case) method. First we derive a dual of S in the following equalities:

$$\begin{aligned} \begin{aligned}&\int ^T_{-\infty } \int _{ {\mathbb {R}}^3 } S({\bar{g}}) (t,x) h(t,x) \mathrm {d}x \mathrm {d}t \\&\quad =\int ^T_{-\infty } \int _{{\mathbb {R}}^3 } \frac{1}{\varepsilon } \int ^t_{-\infty }\int _{{\mathbb {R}}^3} |{\bar{g}}(s, x- \frac{t-s}{\varepsilon } v, v) | {\tilde{\varphi }} (v) h(t,x) \mathrm {d}v \mathrm {d}s \mathrm {d}x \mathrm {d}t \\&\quad =\int ^T_{-\infty } \int _{{\mathbb {R}}^3 }\int _{{\mathbb {R}}^3} |{\bar{g}}(s, x , v) | \left[ \frac{1}{\varepsilon } \int ^T_{s} h(t,x+ \frac{t-s}{\varepsilon } v) {\tilde{\varphi }} (v) \mathrm {d}t \right] \mathrm {d}v \mathrm {d}x \mathrm {d}s \\&\quad =\int ^T_{-\infty } \iint _{{\mathbb {R}}^3\times {\mathbb {R}}^3} |{\bar{g}}(t, x , v) | \left[ \frac{1}{\varepsilon } \int ^T_{t} h(s,x+ \frac{s-t}{\varepsilon } v) {\tilde{\varphi }} (v) \mathrm {d}s \right] \mathrm {d}v \mathrm {d}x \mathrm {d}t \\&\quad = \int ^T_{-\infty } \iint _{{\mathbb {R}}^3\times {\mathbb {R}}^3} | {\bar{g}}(t, x , v) | S^*(h) (t,x,v) \mathrm {d}v \mathrm {d}x \mathrm {d}t , \end{aligned} \end{aligned}$$
(4.91)

where we have defined

$$\begin{aligned} S^*(h) (t,x,v) : = \frac{1}{\varepsilon } \int ^T_{t} h(\tau ,x+ \frac{\tau -t}{\varepsilon } v) {\tilde{\varphi }} (v) \mathrm {d}\tau . \end{aligned}$$
(4.92)

Here, in the second equality of (4.91) we have used the Fubini theorem for changing order of s and t integrations, and then used a change of variables \(x \mapsto x- \frac{t-s}{\varepsilon } v\). In the third equality of (4.91) we have used a change of variable \((t,s) \mapsto (s,t)\) and the fact \(\text {Supp} (g) \subset (-\infty ,T] \times U \times V\).

On the other hand, for \(1/p+1/q=1\) , following the argument of (4.91) with \(h(t,x) = {\mathbf {1}}_{x \in {\tilde{\Omega }}} h(t,x)\) we derive that

$$\begin{aligned} \begin{aligned}&\Vert S({\bar{g}}) \Vert _{L^2_t((-1,T]; L^p_x({\tilde{\Omega }}))} \\&\quad = \sup _{\Vert h \Vert _{L^2_t((-1,T]; L^q_x({\tilde{\Omega }}))} \le 1} \int ^T_{-1} \int _{{\tilde{\Omega }}} S({\bar{g}}) (t,x) h(t,x) \mathrm {d}x \mathrm {d}t\\&\quad = \sup _{\Vert h \Vert _{L^2_t((-1,T]; L^q_x({\tilde{\Omega }}))} \le 1} \int ^T_{-1} \iint _{U\times V} |{\bar{g}} (t,x,v) |S^*(h)(t,x,v) \mathrm {d}v \mathrm {d}x \mathrm {d}t. \end{aligned} \end{aligned}$$
(4.93)

It is important to check the integral region in space of the last term of (4.93). From (4.92), we note that if \(x + \frac{\tau -t}{\varepsilon } v \notin cl( {\tilde{\Omega }})\) for all \(\tau \in [t,T]\) then the last term would vanish since \(\text {supp} (h) \subset (-\infty ,T] \times {\tilde{\Omega }}\). Therefore we can exclude (txv) from the last integration in (4.93) if \(L(t,x,v) \cap {\tilde{\Omega }} = \emptyset \) for \(L(t,x,v):=\{x + \frac{\tau -t}{\varepsilon } v : \tau \in [t,T] \}\). Now we define

$$\begin{aligned} {\mathfrak {D}}_T:= \big \{ (t,x,v) \in (-1, T] \times U \times V: L(t,x,v) \cap {\tilde{\Omega }} \ne \emptyset \big \}. \end{aligned}$$
(4.94)

Then we can write

$$\begin{aligned} \begin{aligned} (4.93)&= \sup _{\Vert h \Vert _{L^2_t((-1,T]; L^q_x({\tilde{\Omega }}))} \le 1} \int ^T_{-1} \iint _{U\times V} {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} |{\bar{g}}(t, x , v) | S^*(h) (t,x,v) \mathrm {d}v \mathrm {d}x \mathrm {d}t\\&\le \Vert {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} {\bar{g}}\Vert _{L^2 ((-1,T]\times U \times V)}\\&\qquad \sup _{\Vert h \Vert _{L^2_t((-1,T]; L^q_x({\tilde{\Omega }}))} \le 1} \big \Vert S^* (h)(t,x,v) \big \Vert _{L^2 ((-1,T]\times U \times V)}. \end{aligned} \end{aligned}$$
(4.95)

Therefore to prove (4.90) it suffices to show that

$$\begin{aligned} \Vert S^* (h) \Vert _{L^2 ((-1,T]\times U \times V)} \lesssim \Vert h \Vert _{L^2_t((-1,T]; L^q_x({\tilde{\Omega }}))}. \end{aligned}$$
(4.96)

Note that since \(\text {supp}(h) \subset (-1,T] \times U\) and \(\text {supp} ({\tilde{\varphi }}) = V\) for \((x,v) \in U\times V\), we have, with \(x= (x_1,x_2,x_3), v=(v_1,v_2,v_3)\)

$$\begin{aligned} |x_1 + \frac{\tau - t}{\varepsilon } v_1|\ge \frac{|\tau -t|}{\varepsilon }|v_1| - |x_1| \ge \frac{10 \pi N \varepsilon }{\varepsilon } \frac{1}{N} - 4\pi > 4\pi \ \ \text {if} \ \tau \ge t+ 10 \pi N \varepsilon . \end{aligned}$$

Hence we can rewrite (4.92) as

$$\begin{aligned} \begin{aligned} S^*(h) (t,x,v) =&\frac{1}{\varepsilon } \int ^{\min \{T, t+ 10 \pi N\varepsilon \}}_{t} h(\tau ,x+ \frac{\tau -t}{\varepsilon } v) {\tilde{\varphi }} (v) \mathrm {d}\tau \\&\ \ \text {for} \ (x,v) \in U\times V, \ \ \text {if} \ \text {supp}(h) \subset (-1,T] \times U. \end{aligned} \end{aligned}$$
(4.97)

On the other hand, from (4.91), we have for \(\text {supp}(h) \in (-1,T] \times {\tilde{\Omega }}\),

$$\begin{aligned} \begin{aligned} \Vert S^* (h) \Vert _{L^2 ((-1,T]\times U \times V)}^2 =&\int ^T_{-1} \iint _{U\times V} S^* (h)(t,x,v) S^* (h)(t,x,v) \mathrm {d}v \mathrm {d}x \mathrm {d}t\\ =&\int ^T_{-1} \iint _{U\times V} SS^* (h) (t,x) h(t,x) \mathrm {d}x \mathrm {d}t \\ \le&\ \Vert SS^* (h)\Vert _{L^2_t((-1,T); L^p_x(U))} \Vert h\Vert _{L^2_t((-1,T]; L^q_x({\tilde{\Omega }}))}. \end{aligned} \end{aligned}$$

Therefore to show (4.96) (which will imply (4.90)) we only need to prove that, for \(\text {supp}(h) \subset (-1,T] \times {\tilde{\Omega }}\),

$$\begin{aligned} \Vert SS^* (h)\Vert _{L^2_t((-1,T]; L^p_x(U))}\lesssim \Vert h\Vert _{L^2_t((-1,T]; L^q_x({\tilde{\Omega }}))}. \end{aligned}$$
(4.98)

Now we prove (4.98). From (4.89) and (4.97), we read

$$\begin{aligned} \begin{aligned}&SS^*(h) (t,x)\\&\quad = \ \frac{1}{\varepsilon } \int ^t_{-1} \int _{{\mathbb {R}}^3} S^*(h) (s, x- \frac{t-s}{\varepsilon } v,v ) {\tilde{\varphi }} (v) \mathrm {d}v \mathrm {d}s\\&\quad = \ \frac{1}{\varepsilon ^2} \int ^{t}_{-1} \int _{{\mathbb {R}}^3} \int ^{\min \{T, s+ 10 \pi N\varepsilon \}} _s h(\tau , x- \frac{t-s}{\varepsilon } v + \frac{\tau -s}{\varepsilon } v ) \mathrm {d}\tau ({\tilde{\varphi }} (v)) ^2\mathrm {d}v \mathrm {d}s\\&\quad = \frac{1}{\varepsilon ^2} \int ^{t}_{-1} \int ^{\min \{T, s+ 10 \pi N\varepsilon \}} _s\int _{{\mathbb {R}}^3} h(\tau , x + \frac{\tau -t}{\varepsilon } v ) ({\tilde{\varphi }} (v)) ^2 \mathrm {d}v \mathrm {d}\tau \mathrm {d}s. \end{aligned} \end{aligned}$$

Now for the same reason to restrict \(\tau \)-integration in (4.97) we rewrite the above expression as

$$\begin{aligned}&SS^*(h) (t,x) \nonumber \\&\quad =\frac{1}{\varepsilon ^2} \int ^{t}_{ \max \{-1, t- 10\pi N \varepsilon \} } \int ^{\min \{T, s+ 10 \pi N\varepsilon \}} _s\int _{{\mathbb {R}}^3} h(\tau , x + \frac{\tau -t}{\varepsilon } v ) ({\tilde{\varphi }} (v)) ^2 \mathrm {d}v \mathrm {d}\tau \mathrm {d}s.\nonumber \\ \end{aligned}$$
(4.99)

We consider a map with the change of variables

$$\begin{aligned} v \in V \mapsto y:=x+ \frac{\tau -t}{\varepsilon } v \in {\mathbb {R}}^3, \ \ \mathrm {d}v= {\mathrm {d}y}\Big /{\left| \frac{\partial y}{\partial v}\right| } =\frac{ \varepsilon ^3}{|\tau - t|^3}\mathrm {d}y. \end{aligned}$$
(4.100)

Now we apply (4.100) to (4.99) and derive that

$$\begin{aligned} \begin{aligned}&|SS^*(h) (t,x)| \\&\quad \le \frac{1}{\varepsilon ^2} \int ^t_{t- 10 \pi N\varepsilon }\int ^{ \min \{ s+ 10 \pi N\varepsilon \} }_s \int _{{\tilde{\Omega }}} | {\mathbf {1}}_{ \tau \in [ -1, T] } h(\tau , y )| \frac{\varepsilon ^3}{|\tau -t|^3} {\tilde{\varphi }} \\&\qquad \Big (\varepsilon \frac{|y-x| }{|\tau -t|} \Big )^2 \mathrm {d}y \mathrm {d}\tau \mathrm {d}s. \end{aligned} \end{aligned}$$
(4.101)

First using the Minkowski’s inequality and the Young’s inequality to a convolution in y with \(1+ 1/p=1/q + 1/(p/2)\) we have

$$\begin{aligned} \begin{aligned}&\left\| \frac{1}{\varepsilon ^2} \int ^t_{t- 10\pi N \varepsilon }\int ^{s+ 10 \pi N \varepsilon }_s \int _{{\tilde{\Omega }}} {\mathbf {1}}_{ \tau \in [ -1, T] } |h(\tau , y )| \frac{\varepsilon ^3}{|\tau -t|^3} {\tilde{\varphi }} \Big (\varepsilon \frac{|y-x|}{|\tau -t|}\Big )^2 \mathrm {d}y\mathrm {d}\tau \mathrm {d}s\right\| _{L^p_x({\tilde{\Omega }})}\\&\quad \le \ \frac{1}{\varepsilon ^2} \int ^t_{t- 10\pi N \varepsilon }\int ^{s+ 10 \pi N \varepsilon }_s \\&\qquad \left\| \int _{{\tilde{\Omega }}} {\mathbf {1}}_{ \tau \in [ -1, T] } |h(\tau , y )| \frac{\varepsilon ^3}{|\tau -t|^3} {\tilde{\varphi }} \Big (\varepsilon \frac{|y-x|}{|\tau -t|}\Big )^2 \mathrm {d}y \right\| _{L^p_x({\tilde{\Omega }})} \mathrm {d}\tau \mathrm {d}s\\&\quad \le \ \frac{1}{\varepsilon ^2} \int ^t_{t- 10\pi N \varepsilon }\int ^{s+ 10 \pi N \varepsilon }_s \Vert {\mathbf {1}}_{ \tau \in [ -1, T] } h(\tau , \cdot )\Vert _{L^q_x( {\tilde{\Omega }})}\\&\qquad \underbrace{\Big \Vert \frac{\varepsilon ^3}{|\tau -t|^3} {\tilde{\varphi }} \Big (\varepsilon \frac{|\cdot |}{|\tau -t|}\Big )^2 \Big \Vert _{L^{p/2}_x({\tilde{\Omega }})}}_{(4.102)_*} \mathrm {d}\tau \mathrm {d}s. \end{aligned} \end{aligned}$$
(4.102)

From the properties of \({\tilde{\varphi }} \in C_c^\infty \), it follows that

$$\begin{aligned} (4.102)_* \le \frac{\varepsilon ^3}{|\tau -t|^3} \left( \frac{|\tau -t|^3}{\varepsilon ^3}\right) ^{\frac{2}{p}} \left[ \int _{ {\mathbb {R}}^3}\left| {\tilde{\varphi }}({\tilde{y}}) \right| ^{p } \mathrm {d}{\tilde{y}} \right] ^{\frac{1}{p/2}} \lesssim \bigg (\frac{\varepsilon }{|\tau -t|}\bigg )^{3-\frac{6}{p}}, \end{aligned}$$

where \({\tilde{y}}= \frac{\varepsilon }{|\tau -t|}(y-x)\) with \(\mathrm {d}{\tilde{y}}= \frac{\varepsilon ^3 }{|\tau -t|^3}\mathrm {d}y\). Therefore we derive that

$$\begin{aligned} \begin{aligned}&\Vert SS^*(h)(t, \cdot ) \Vert _{L^p_x} \\&\quad \lesssim \frac{1}{\varepsilon ^2} \int ^t_{t- 10 \pi N \varepsilon } \int ^{s+ 10 \pi N \varepsilon }_s \Vert {\mathbf {1}}_{ \tau \in [ -1, T] } h(\tau , \cdot )\Vert _{L^q_x({\tilde{\Omega }})} \bigg (\frac{\varepsilon }{|\tau -t|}\bigg )^{3-\frac{6}{p}} \mathrm {d}\tau \mathrm {d}s . \end{aligned}\nonumber \\ \end{aligned}$$
(4.103)

Using the Minkowski’s inequality and the Young’s inequality, finally we prove (4.98) as

$$\begin{aligned}&\big \Vert \Vert SS^*(h)(t, \cdot ) \Vert _{L^p_x} \big \Vert _{L^2_t (0, T)} \\&\quad \lesssim \ \frac{1}{\varepsilon ^2} \Big \Vert {\mathbf {1}}_{[t- 10 \pi N \varepsilon , t]} (s) \Big \Vert _{L^1_s } \left\| \sup _{s \in [t- 10 \pi N \varepsilon , t]} \int ^{s+ 10 \pi N \varepsilon }_{s} \right. \\&\qquad \left. \Vert {\mathbf {1}}_{ \tau \in [ -1, T] } h(\tau , \cdot )\Vert _{L^q_x({\tilde{\Omega }})} \bigg (\frac{\varepsilon }{|\tau -t|}\bigg )^{3-\frac{6}{p}} \mathrm {d}\tau \right\| _{L^2_t}\\&\quad \lesssim \ \frac{1}{\varepsilon ^2} 10 \pi N \varepsilon \left\| \int ^{t+ 10 \pi N \varepsilon }_{t- 10 \pi N \varepsilon } \Vert {\mathbf {1}}_{ \tau \in [ -1, T] } h(\tau , \cdot )\Vert _{L^q_x({\tilde{\Omega }})} \bigg (\frac{\varepsilon }{|\tau -t|}\bigg )^{3-\frac{6}{p}} \mathrm {d}\tau \right\| _{L^2_t}\\&\quad \lesssim \ \frac{1}{\varepsilon ^2} 10 \pi N \varepsilon \big \Vert \Vert h(\tau , \cdot )\Vert _{L^q_x({\tilde{\Omega }})}\big \Vert _{L^2_\tau ((-1, T])} \left\| \bigg (\frac{\varepsilon }{| t|}\bigg )^{3-\frac{6}{p}} \right\| _{L^1_t ((0, 10 \pi N \varepsilon ))}\\&\quad \lesssim \ N^{-1 + \frac{6}{p}} \Vert h \Vert _{L^2_t ((-1 , T];L^q_x({\tilde{\Omega }}) )} . \end{aligned}$$

\(\square \)

Step 3: Applying Lemma 6. Now we apply Lemma 6 to (4.85) and derive that

$$\begin{aligned}&\left\| \int _{{\mathbb {R}}^3}{\bar{f}}_R (t,x,v) {\tilde{\varphi }} (v) \mathrm {d}v \right\| _{L^2_t( (-1,T]; L^p_x ({\tilde{\Omega }}))}\nonumber \\&\quad \lesssim \ \Vert {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} {\bar{g}} \Vert _{L^2 ((-1,T]\times U \times V)}\nonumber \\&\quad \lesssim \ \Vert f_R (t,x,v)\Vert _{L^2 ((0,T]\times {\tilde{\Omega }} \times V)} + \Vert f_R (0,x,v) \Vert _{L^2 ( {\tilde{\Omega }} \times V)} \nonumber \\&\qquad + \Vert {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} f_I (t+ \varepsilon {\tilde{t}}_B(x,v), {\tilde{x}}_B(x,v) ,v) \Vert _{L^2 ((-1,T]\times (U \backslash {\tilde{\Omega }})\times V)} \end{aligned}$$
(4.104)
$$\begin{aligned}&\qquad + \Vert [\varepsilon \partial _t + v\cdot \nabla _x ] f_R \Vert _{L^2 ((0,T]\times {\tilde{\Omega }} \times V)}, \end{aligned}$$
(4.105)

where we have used (4.83), (4.77), (4.79), and the fact that \(|v\cdot \nabla _x \chi _2 (x)| \lesssim _N 1\) on \(v \in V\).

First we consider (4.104). We split the cases of (4.104) according to (4.78). For \(x \in \partial {\tilde{\Omega }}\), which has a zero measure in \(L^2 ((-1,T]\times (U \backslash {\tilde{\Omega }})\times V)\), we have \({\tilde{t}}_B(x,v)=0\) from the first line of (4.78). If \({\tilde{B}}(x,v) = \emptyset \) and \(x \notin \partial {\tilde{\Omega }}\) then \({\tilde{t}}_B(x,v)=-\infty \) from the last line of (4.78) and hence \({\bar{f}}_R(-\infty )=0\) since \(\chi _1(-\infty )=0\) in (4.77). Therefore we derive that

$$\begin{aligned} (4.104)\le&\ \Vert {\mathbf {1}}_{\{s<0\} \subset {\tilde{B}} (x,v)} {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} f_I \nonumber \\&\quad (t+ \varepsilon {\tilde{t}}_B(x,v), {\tilde{x}}_B(x,v) ,v) \Vert _{L^2 ((-1,T]\times (U \backslash {\tilde{\Omega }})\times V)} \end{aligned}$$
(4.106)
$$\begin{aligned}&+ \Vert {\mathbf {1}}_{\{s>0\} \subset {\tilde{B}} (x,v)} {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} f_I \nonumber \\&\quad (t+ \varepsilon {\tilde{t}}_B(x,v), {\tilde{x}}_B(x,v),v) \Vert _{L^2 ((-1,T]\times (U \backslash {\tilde{\Omega }})\times V)}. \end{aligned}$$
(4.107)

We need a special attention to (4.106). Since \((t,x,v) \in {\mathfrak {D}}_T\) we know that \(\inf \{\tau \ge t: x+ \frac{\tau - t}{\varepsilon } v \in cl({\tilde{\Omega }})\} \le T.\) If \(\{s<0\} \subset {\tilde{B}} (x,v)\) then, from the third line of (4.78), \({\tilde{t}}_B(x,v) = \sup {\tilde{B}} (x,v)= \sup \{s \in {\mathbb {R}}: x+ sv \in {\mathbb {R}}^3 \backslash cl ({\tilde{\Omega }})\} \le (T-t)/\varepsilon \). Therefore the argument of \(f_I\) in (4.106) is confined as

$$\begin{aligned} (t+ \varepsilon {\tilde{t}}_B(x,v), {\tilde{x}}_B(x,v),v) \in (- \infty ,T] \times \partial {\tilde{\Omega }} \times V . \end{aligned}$$
(4.108)

For (4.107), from the second line of (4.78), \({\tilde{t}}_B (x,v)= \inf {\tilde{B}}(x,v) = \inf \{s \in {\mathbb {R}}: x+ sv \in {\mathbb {R}}^3 \backslash cl ({\tilde{\Omega }})\} \le 0\). Therefore \(t+ \varepsilon {\tilde{t}}_B (x,v) \le t \le T\) and hence the argument of \(f_I\) in (4.107) is confined as in (4.108). Now we apply the Minkowski’s inequality in time, change of variables \(t+ \varepsilon {\tilde{t}}_B (x,v) \mapsto t\), and use (4.108) to derive that

$$\begin{aligned} (4.106) + (4.107) \lesssim \Big \Vert \Vert f_I(t, {\tilde{x}}_B(x,v),v)\Vert _{L^2_t ((-1, T])} \Big \Vert _{L^2_{x,v}((U \backslash {\tilde{\Omega }}) \times V)}. \end{aligned}$$
(4.109)

Let us define an outward normal \({\tilde{n}}(x)\) on \(\partial {\tilde{\Omega }}\). More precisely

$$\begin{aligned} {\tilde{n}}(x)= {\left\{ \begin{array}{ll} (0,0,-1) &{}\ \ \text {if} \ x_3=0 \ \text {and} \ x \in \partial {\tilde{\Omega }},\\ ((-1)^{\frac{x_1}{2\pi }+1},0,0) &{}\ \ \text {if} \ x_1 \in \{0, 2\pi \} \ \text {and} \ x \in \partial {\tilde{\Omega }},\\ (0,(-1)^{\frac{x_2}{2\pi }+1},0) &{}\ \ \text {if} \ x_2 \in \{0, 2\pi \} \ \text {and} \ x \in \partial {\tilde{\Omega }}. \end{array}\right. } \end{aligned}$$
(4.110)

From (4.82) we have therefore \((x ,v) \in (U \backslash {\tilde{\Omega }}) \times V\) then \(|{\tilde{n}}( {\tilde{x}}_B (x,v) ) \cdot v|\ge 1/N\). We consider maps

$$\begin{aligned} \begin{aligned} (x_1,x_3)&\mapsto {\tilde{x}}_B (x,v) \in (0,2\pi ) \times (0, 2\pi ) \times \{x_3=0\},\\&\ \ \text {with} \ \ \Big |\det \Big (\frac{\partial ( {\tilde{x}}_{B,1} (x,v) , {\tilde{x}}_{B,2} (x,v) )}{\partial (x_1,x_3)}\Big )\Big |= \Big |\frac{ v_2}{v \cdot {\tilde{n}}}\Big |,\\ (x_i,x_3)&\mapsto ({\tilde{x}}_{B,i} (x,v), {\tilde{x}}_{B,3} (x,v)) \in (0,2\pi ) \times (0,\infty ), \\&\ \ \text {with} \ \ \Big |\det \Big (\frac{\partial ( {\tilde{x}}_{B,i} (x,v) , {\tilde{x}}_{B,3} (x,v) )}{\partial (x_1,x_3)}\Big )\Big |= \Big |\frac{ v_i }{v \cdot {\tilde{n}}}\Big |, \ \ \text {for} \ i =1,2. \end{aligned} \end{aligned}$$
(4.111)

Note that if \(v \in V\) of (4.82) then \(|v_i| \ge 1/N\) for all \(i=1,2,3.\) We define

$$\begin{aligned} {\tilde{\gamma }} := \partial {\tilde{\Omega }} \times {\mathbb {R}}^3, \ \ \ {\tilde{\gamma }}^{N} := \partial {\tilde{\Omega }} \times ({\mathbb {R}}^3\backslash V). \end{aligned}$$
(4.112)

We apply the change of variables (4.111) to (4.109):

$$\begin{aligned} \begin{aligned} (4.109)&= \bigg \Vert \bigg [ \int _{-2\pi }^{4\pi } \int _{-2\pi }^\infty \int _{-2\pi }^{4\pi } \Vert f_I (t, {\tilde{x}}_B(x,v),v)\Vert _{L^2_t ((-1, T])}^2 \mathrm {d}x_1 \mathrm {d}x_3 \mathrm {d}x_2 \bigg ]^{1/2} \bigg \Vert _{L^2_{v}(V)}\\&\le \bigg \Vert \bigg [ 5 \times 6 \pi N \int _{\partial {\tilde{\Omega }}}\int _{-1}^T |f_I (t, y,v) |^2 |v \cdot {\tilde{n}}(y)| \mathrm {d}t \mathrm {d}y\bigg ]^{1/2} \bigg \Vert _{L^2_{v}(V)}\\&\lesssim \Vert f_R\Vert _{L^2((0,T) \times {\tilde{\gamma }} \backslash {\tilde{\gamma }}^{ N } )} + \Vert f_R(0)\Vert _{L^2( {\tilde{\gamma }} \backslash {\tilde{\gamma }}^{ N } )}. \end{aligned}\nonumber \\ \end{aligned}$$
(4.113)

We recall the trace theorem:

$$\begin{aligned} \begin{aligned} \int ^T_{0} \int _{{\tilde{\gamma }}\backslash {\tilde{\gamma }}^{1/N}} |h| \mathrm {d}\gamma \mathrm {d}s \lesssim&\sup _{t \in [0,T]} \Vert h(t) \Vert _{L^1({\tilde{\Omega }} \times V)} + \int ^T_{0} \Vert h(s) \Vert _{L^1({\tilde{\Omega }} \times V)} \mathrm {d}s \\&+ \int ^T_{0} \Vert [\varepsilon \partial _t + v\cdot \nabla _x ] h \Vert _{L^1({\tilde{\Omega }} \times V)} \mathrm {d}s. \end{aligned} \end{aligned}$$
(4.114)

We apply (5.33) with \(h=f^2\) and derive an estimate

$$\begin{aligned} \begin{aligned}&\Vert f_R\Vert _{L^2((0,T) \times {\tilde{\gamma }} \backslash {\tilde{\gamma }}^{ N } )}^2 \\&\quad \lesssim \sup _{t \in [0,T]} \Vert f_R(t) \Vert _{L^2({\tilde{\Omega }} \times V)}^2 + \int ^T_{0} \Vert f_R(s) \Vert _{L^2({\tilde{\Omega }} \times V)}^2 \mathrm {d}s\\&\qquad + \int ^T_{0} \iint _{{\tilde{\Omega }}\times V}\big | f_R[\varepsilon \partial _t + v\cdot \nabla _x ] f_R \big | \mathrm {d}x \mathrm {d}v \mathrm {d}s\\&\quad \lesssim _T \Vert f_R \Vert _{L^\infty ([0,T];L^2( {\Omega } \times {\mathbb {R}}^3))}^2 + \big \Vert [\varepsilon \partial _t+ v\cdot \nabla _x ] f_R \big \Vert _{L^2( [0,T] \times {\Omega } \times {\mathbb {R}}^3)}. \end{aligned} \end{aligned}$$
(4.115)

Finally we conclude a bound of (4.104) as below via (4.106), (4.107), (4.109), (4.113), and (4.115)

$$\begin{aligned} (4.104)\lesssim & {} \Vert f_R (0)\Vert _{L^2_\gamma } + \Vert f_R \Vert _{L^\infty ([0,T];L^2( {\Omega } \times {\mathbb {R}}^3))} \nonumber \\&+ \underbrace{\big \Vert [\varepsilon \partial _t+ v\cdot \nabla _x ] f_R \big \Vert _{L^2( [0,T] \times {\Omega } \times {\mathbb {R}}^3)}}_{(4.116)_*}. \end{aligned}$$
(4.116)

Next we estimate (4.105) (and \((4.116)_*\)). Using (4.84) and (3.2) we conclude that

$$\begin{aligned} \begin{aligned}&(4.105)+ (4.116)_* \\&\quad \lesssim \bigg \Vert - \frac{1}{\varepsilon \kappa } L({\mathbf {I}} - {\mathbf {P}}) f_R + \frac{\varepsilon }{\kappa } \Gamma (f_2, f_R) + \frac{\delta }{\kappa } \Gamma (f_R, f_R) \\&\qquad - \frac{( \varepsilon \partial _t + v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} f_{R} + \varepsilon ({\mathbf {I}}- {\mathbf {P}}){\mathfrak {R}}_1 +\varepsilon {\mathfrak {R}}_2 \bigg \Vert _{L^2 ((0,T]\times \Omega \times V)}. \end{aligned} \end{aligned}$$

Following the arguments of (4.15)-(4.18), and (3.4), (3.5), we derive that

$$\begin{aligned}&(4.105)+(4.116)_*\nonumber \\&\quad \lesssim \ \Big \{ \frac{\varepsilon }{\kappa } \Vert (3.10) \Vert _{L^\infty _t ((0,T); L_x^{\frac{2p}{p-2}}(\Omega )) } + \frac{\delta }{\kappa } \Vert Pf_R \Vert _{L^\infty _t ((0,T); L_x^{\frac{2p}{p-2}}(\Omega )) } \Big \}\nonumber \\&\quad \Vert P f_R \Vert _{L^2_t ((0,T); L_x^{p}(\Omega ))}\nonumber \\&\qquad +\Big \{ \frac{1}{\varepsilon \kappa } + \frac{\delta }{\kappa } \Vert {\mathfrak {w}}_{\varrho , \ss } f_R \Vert _{L^\infty _t ((0,T) \times \Omega \times {\mathbb {R}}^3) } \Big \} \Vert ({\mathbf {I}} - {\mathbf {P}})f_R \Vert _{L^2_t ((0,T) \times \Omega \times {\mathbb {R}}^3) } \nonumber \\&\qquad +\varepsilon \big \Vert (3.12) \big \Vert _{L^2_t ((0,T); L^\infty _x(\Omega ))} \Vert f_R(t) \Vert _{L^\infty _t ((0,T);L^2(\Omega \times {\mathbb {R}}^3))} \nonumber \\&\qquad + \varepsilon \{ \Vert (3.4)\Vert _{L^2_t ((0,T); L_x^{2}(\Omega ))} + \Vert (3.5)\Vert _{L^2_t ((0,T); L_x^{2}(\Omega ))} \} , \end{aligned}$$
(4.117)

where we further bound

$$\begin{aligned} \Vert Pf_R \Vert _{ L_x^{\frac{2p}{p-2}}(\Omega ) } \le \Vert Pf_R \Vert _{L_x^{6}(\Omega )}^{\frac{3(p-2)}{p}} \Vert {\mathfrak {w}}_{\varrho ,\ss } f_R \Vert _{L_x^{\infty }(\Omega )}^{\frac{6-2p}{p}}. \end{aligned}$$
(4.118)

Step 4. Proof of (4.72). First we use (4.87) and then (4.104) and (4.105). We bound (4.104) via (4.109) and (4.113), which are bounded by (4.115) and (4.117) respectively. These conclude that, for \(p<3\),

$$\begin{aligned} \begin{aligned}&\big (1- O(\varepsilon ) \Vert u\Vert _\infty - O(\frac{1}{N})\big ) \big \Vert {P} f_R \big \Vert _{L^2_t((0,T);L^p_x({\tilde{\Omega }}))}\\&\qquad -C_{T,N} \Vert {\mathfrak {w}}_{\varrho , \ss } f_R (t) \Vert _{ L^\infty ((0,T) \times {\tilde{\Omega }} \times {\mathbb {R}}^3)}^{\frac{p-2}{p}} \Vert ({\mathbf {I}} - {\mathbf {P}}) f_R \Vert _{L^2((0, T) \times {\tilde{\Omega }} \times {\mathbb {R}}^3)}^{\frac{2}{p}}\\&\quad \le \ \left\| \int _{{\mathbb {R}}^3} {\bar{f}}_R(t,x,v) {\tilde{\varphi }}_i (v) \sqrt{\mu _0(v)} \mathrm {d}v \right\| _{L^2_t ((0,T); L^p_x ( {\tilde{\Omega }}))}\\&\quad \le \ \left\| \int _{{\mathbb {R}}^3} {\bar{f}}_R(t,x,v) {\tilde{\varphi }} (v) \mathrm {d}v \right\| _{L^2_t ((0,T); L^p_x ( {\tilde{\Omega }}))}\\&\quad \lesssim \ \Vert f_R \Vert _{L^\infty ([0,T];L^2( {\Omega } \times {\mathbb {R}}^3))} + \Vert f_R (0) \Vert _{L^2_\gamma }+ \text {r.h.s. of } (4.117) \text { with } (4.118). \end{aligned} \end{aligned}$$
(4.119)

Then we move a contribution of \(\Vert P f_R \Vert _{L^2_t ((0,T); L_x^{p}(\Omega ))}\) to the l.h.s and use (4.118). This concludes (4.72).

Step 5: Sketch of proof for (4.74). We follow the same argument for (4.72). Thereby we only pin point the difference of the proof of (4.74). Recall \(\partial _t f_R (0,x,v) = f_{R,t}(0,x,v)\) from (2.6). We regard \({\tilde{\Omega }}\) as an open subset but not a periodic domain as \(\Omega \). Without loss of generality we may assume that \(f_{R,t}(0,x,v)\) is defined in \({\mathbb {R}}^3 \times {\mathbb {R}}^3\) and \(\Vert f_{R,t}(0) \Vert _{L^p({\mathbb {R}}^3) \times {\mathbb {R}}^3} \lesssim \Vert f_{R,t}(0) \Vert _{L^p({\tilde{\Omega }}) \times {\mathbb {R}}^3}\) for all \(1 \le p \le \infty \). Then we extend a solution for whole time \(t \in {\mathbb {R}}\) as

$$\begin{aligned} f_{ I,t} (t,x,v) : = {\mathbf {1}}_{t \ge 0 } \partial _t f_R (t,x,v) + {\mathbf {1}}_{t \le 0} \chi _1 (t) f_{R,t} (0, x ,v ). \end{aligned}$$
(4.120)

Using \({\tilde{t}}_B(x,v) \) in (4.78) we define

$$\begin{aligned} f_{E,t} (t,x,v): = {\mathbf {1}}_{(x,v) \in ({\mathbb {R}}^3 \backslash {\tilde{\Omega }}) \times {\mathbb {R}}^3 } f_{I,t} (t+ \varepsilon {\tilde{t}}_B(x,v), {\tilde{x}}_B(x,v),v). \end{aligned}$$
(4.121)

We define an extension of cut-offed solutions

$$\begin{aligned}&{\bar{f}}_{R,t}(t,x,v) := \chi _2 (x) \chi _3 (v) \big \{ {\mathbf {1}}_{{\tilde{\Omega }}}(x) f_{I,t}(t,x,v) + f_{E,t} (t,x,v)\big \} \ \ \nonumber \\&\quad \text {for} \ (t,x,v) \in (-\infty ,T] \times {\mathbb {R}}^3 \times {\mathbb {R}}^3 . \end{aligned}$$
(4.122)

We note that in the sense of distributions \({\bar{f}}_{R,t}\) solves

$$\begin{aligned} \begin{aligned}&\varepsilon \partial _t {\bar{f}}_{R,t} + v\cdot \nabla _x {\bar{f}}_{R,t} = \ {\bar{g}}_t \ \ \text {in} \ (-\infty ,T] \times {\mathbb {R}}^3 \times {\mathbb {R}}^3, \ \ \text {where}\\&{\bar{g}}_t : = \frac{v\cdot \nabla _x \chi _2 }{\chi _2 } {\bar{f}}_{R,t} + {\mathbf {1}}_{t \ge 0} {\mathbf {1}}_{{\tilde{\Omega }}} (x) \chi _2(x) \chi _3(v) [\varepsilon \partial _t + v\cdot \nabla _x] \partial _t f_R \\&\quad \ \ +{\mathbf {1}}_{t \le 0} \chi _2 (x) \chi _3 (v) \{ \varepsilon \partial _t \chi _1(t) f_{R,t} (0,x,v) + \chi _1 (t) v\cdot \nabla _x f_{R,t} (0,x,v) \}. \end{aligned} \end{aligned}$$
(4.123)

Here we have used the fact that \({\bar{f}}_{R,t}\) in (4.123) is continuous along the characteristics across \(\partial {\tilde{\Omega }}\) and \(\{t=0\}\). We derive that, using (4.123),

$$\begin{aligned} {\bar{f}}_{R,t}(t,x,v) = \frac{1}{\varepsilon } \int ^t_{-\infty } {\bar{g}}_t (s, x- \frac{t-s}{\varepsilon } v, v) \mathrm {d}s \ \ \text {for} \ (t,x,v) \in (-\infty ,T] \times {\mathbb {R}}^3 \times {\mathbb {R}}^3.\nonumber \\ \end{aligned}$$
(4.124)

Now we apply Lemma 6 to (4.124) and derive that, for \(p<3\),

$$\begin{aligned} \begin{aligned}&\Vert S(\bar{g_t}) \Vert _{L^2_t((0,T); L^p_x({\mathbb {T}}^2 \times {\mathbb {R}}))} \\&\quad \lesssim \Vert {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} {\bar{g}}_t\Vert _{L^2 ((0,T)\times ({\mathbb {T}}^2 \times {\mathbb {R}}) \times \{|v| \le N \})}\\&\quad \lesssim \ \Vert f_{R,t} (0)\Vert _{L^2(\Omega \times {\mathbb {R}}^3)} + \Vert \varepsilon \partial _t f_{R,t} + v\cdot \nabla _x f_{R,t} \Vert _{L^2 ((0,T) \times {\tilde{\Omega }} \times V)} \\&\qquad + \Vert {\mathbf {1}}_{(t,x,v) \in {\mathfrak {D}}_T} f_{I,t} (t+ \varepsilon {\tilde{t}}_B(x,v), {\tilde{x}}_B(x,v) ,v) \Vert _{L^2 ((-1,T]\times (U \backslash {\tilde{\Omega }})\times V)}. \end{aligned} \end{aligned}$$
(4.125)

Following the same argument of (4.116)-(4.117) we deduce that

$$\begin{aligned} \begin{aligned} (4.125) \lesssim&\Vert \partial _t f_R \Vert _{L^\infty ([0,T];L^2( {\Omega } \times {\mathbb {R}}^3))} + \Vert \partial _t f_R (0)\Vert _{L^2_\gamma } \\&+\big \Vert - \frac{1}{\varepsilon \kappa } L({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R + \varepsilon \times \text {r.h.s. of } (3.3) \big \Vert _{L^2 ((0,T]\times \Omega \times V)}. \end{aligned} \end{aligned}$$
(4.126)

From (4.31)-(4.33), the last term of (4.126) is bounded above by

$$\begin{aligned} \begin{aligned}&\Big \{\frac{1}{\kappa } \Vert \partial _t u \Vert _{L^\infty _{t,x}} \Big ( 1+ \delta \varepsilon \Vert {\mathfrak {w}}f_R \Vert _{L^\infty _{t,x,v}} \Big ) + \frac{\varepsilon ^2}{\kappa } \Vert (3.10) \Vert _{L^\infty _{t,x}} \Big \} \Big \{ \Vert Pf_R \Vert _{L^2_{t,x}}\\&\quad + \Vert \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \Big \} \\&\quad + \Big \{\frac{1}{\kappa \varepsilon } + \frac{\delta }{\kappa } \Vert {\mathfrak {w}}f_R\Vert _{L^\infty _{t,x,v}} + \frac{\varepsilon }{\kappa } \Vert (3.10) \Vert _{L^\infty _{t,x}} + \varepsilon \Vert (3.12) \Vert _{L^\infty _{t,x} } \Big \} \Vert \sqrt{\nu } ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \\&\quad + \Big \{ \frac{\delta }{\kappa }\Vert {P} f_R \Vert _{L^\infty _tL^6_{x }}^{\frac{3(p-2)}{p}} \Vert {\mathfrak {w}} f_R\Vert _{L^\infty _{t,x,v}}^{\frac{6-2p}{p}} + \frac{\varepsilon }{\kappa }\Vert (3.10) \Vert _{L^\infty _tL^{\frac{2p}{p-2}}_{x }} \\&\quad + \varepsilon \Vert (3.12) \Vert _{L^\infty _{t } L^\frac{2p}{p-2}_x } \Big \} \Vert {P} \partial _t f_R \Vert _{L^2_tL^p_{x }} \\&\quad + \frac{\varepsilon }{\kappa } \Vert (3.11) \Vert _{L^2_{t,x,v}} \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}} +\varepsilon \Vert (3.13)\Vert _{L^2_tL^\infty _{x}} \Vert f_R\Vert _{L^\infty _t L^2_{x,v}} \\&\quad + \varepsilon \{ \Vert (3.6) \Vert _{L^2_{t,x}} + \Vert (3.7) \Vert _{L^2_{t,x}} \}. \end{aligned} \end{aligned}$$
(4.127)

Here the most singular term comes from \(\frac{1}{\varepsilon ^2 \kappa } L({\mathbf {P}}_t f_R)\) in the r.h.s. of (3.3) .

On the other hand from (4.122) and the argument of (4.86) we derive

$$\begin{aligned} \begin{aligned}&\Vert S(\bar{g_t}) \Vert _{L^2_t((0,T); L^p_x({\mathbb {T}}^2 \times {\mathbb {R}}))} > rsim \left\| \int _{{\mathbb {R}}^3} {\bar{f}}_{R,t}(t,x,v) {\tilde{\varphi }}_i (v) \sqrt{\mu _0(v)} \mathrm {d}v \right\| _{L^2_t ((0,T); L^p_x ( {\tilde{\Omega }}))} \\&\quad > rsim \ \big (1- O(\varepsilon )\Vert u\Vert _\infty - O(\frac{1}{N})\big ) \big \Vert {P} \partial _t f_R \big \Vert _{L^2_t( (0,T) ;L^p_x({\tilde{\Omega }}))}\\&\qquad - (\kappa \varepsilon ) ^{\frac{2}{p-2}} \Vert {\mathfrak {w}}^\prime \partial _t f_R \Vert _{L^2_t ((0,T);L^\infty _{x,v} ( \Omega \times {\mathbb {R}}^3))}-\frac{1}{\kappa \varepsilon }\Vert ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2 ((0,T) \times \Omega \times {\mathbb {R}}^3)}. \end{aligned} \end{aligned}$$
(4.128)

Here we have used

$$\begin{aligned} \begin{aligned}&\left\| \int _{{\mathbb {R}}^3} \chi _2 (x) \chi _3 (v) ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R(t,x,v) {\tilde{\varphi }}_i (v) \sqrt{\mu _0(v)} \mathrm {d}v \right\| _{L^2_t ((0,T); L^p_x ({\tilde{\Omega }}))} \\&\quad \le \left\| ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R(t,x,v) \right\| _{L^2_t ((0,T); L^p_{x,v} ({\tilde{\Omega }} \times {\mathbb {R}}^3))} \\&\quad \lesssim \ \Big \Vert \Vert {\mathfrak {w}}^\prime \partial _t f_R \Vert _{L^\infty _{x,v} ( \Omega \times {\mathbb {R}}^3)}^{\frac{p-2}{p}} \Vert ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{x,v} ( \Omega \times {\mathbb {R}}^3)}^{\frac{2}{p}} \Big \Vert _{L^2_t((0,T))}\\&\quad \lesssim \ \Big \Vert \Vert {\mathfrak {w}} ^\prime \partial _t f_R \Vert _{L^\infty _{x,v} ( \Omega \times {\mathbb {R}}^3)}^{\frac{p-2}{p}}\Big \Vert _{L^{\frac{2p}{p-2}}_t((0,T))} \Big \Vert \Vert ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2_{x,v} ( \Omega \times {\mathbb {R}}^3)}^{\frac{2}{p}}\Big \Vert _{L^p_t((0,T))} \\&\quad \lesssim \ (\kappa \varepsilon )^{\frac{2}{p}} \Vert {\mathfrak {w}}^\prime \partial _t f_R \Vert _{L^2_t ((0,T);L^\infty _{x,v} ( \Omega \times {\mathbb {R}}^3))}^{\frac{p-2}{p}} (\kappa \varepsilon )^{-\frac{2}{p}} \Vert ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2 ((0,T) \times \Omega \times {\mathbb {R}}^3)}^{\frac{2}{p}}\\&\quad \lesssim \ (\kappa \varepsilon ) ^{\frac{2}{p-2}} \Vert {\mathfrak {w}}^\prime \partial _t f_R \Vert _{L^2_t ((0,T);L^\infty _{x,v} ( \Omega \times {\mathbb {R}}^3))}+ (\kappa \varepsilon )^{-1} \Vert ({\mathbf {I}} - {\mathbf {P}}) \partial _t f_R \Vert _{L^2 ((0,T) \times \Omega \times {\mathbb {R}}^3)}. \end{aligned} \end{aligned}$$
(4.129)

Combining (4.128), (4.125), (4.126), and (4.127) and choosing \(N\gg 1\) we conclude (4.74).

4.4 \(L^\infty \)-Estimate

In this section we develop a unified \(L^\infty \)-estimate in the local Maxwellian setting. We devise the weight functions to control an extra growth in |v| comes from \(\frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }}\) and its temporal derivative:

$$\begin{aligned} {\mathfrak {w}}_{\varrho , \ss }(x,v)={\mathfrak {w}} := \exp \{\varrho |v|^2 - {\mathfrak {z}}_{\ss }(x_3) (x \cdot v) \} \ \ \text {for} \ \ 0< \ss \ll \frac{\varrho }{2\pi } \ \text {and} \ 0< \varrho < \frac{1}{4}, \end{aligned}$$

where \({\mathfrak {z}}_{\ss }: {\mathbb {R}}_+ \rightarrow {\mathbb {R}}_+\) is defined as, for \(\ss >0\)

$$\begin{aligned} \begin{aligned} {\mathfrak {z}}_{\ss }(x_3)= \ss \ \ \text {for} \ \ x_3 \in \big [0, \frac{1}{\ss }-1 \big ], \ \ \text {and} \ \ {\mathfrak {z}}_{\ss }(x_3)=\frac{1}{1+ x_3} \ \ \text {for} \ \ x_3\in \big [\frac{1}{\ss }-1 , \infty \big ). \end{aligned} \end{aligned}$$

We often abuse the notation of \({\mathfrak {w}}_{\varrho , \ss }\) and \({\mathfrak {w}}\). We compute to have

$$\begin{aligned}&\frac{v\cdot \nabla _x {\mathfrak {w}}_{\varrho , \ss }(x,v) }{{\mathfrak {w}}_{\varrho , \ss }(x,v)}\\&\quad = - {\mathfrak {z}}_{\ss } (x_3) |v |^2 - v_3 \partial _{x_3}{\mathfrak {z}}_{\ss }(x_3) ( x_1 v_1+ x_2 v_2 + x_3 v_3 ) \\&\quad = - {\mathfrak {z}}_{\ss } (x_3) |v_3|^2 - x_3 \partial _{x_3}{\mathfrak {z}}_{\ss }(x_3) |v_3|^2 - {\mathfrak {z}}_{\ss } (x_3) (|v_1|^2 + |v_2|^2) \\&\qquad - \partial _{x_3} {\mathfrak {z}}_{\ss } (x_3)(x_1v_1 + x_2 v_2) v_3 \\&\quad = -\ss {\mathbf {1}}_{[0, \ss ^{-1}-1]} (x_3) |v |^2 - {\mathbf {1}}_{[ \ss ^{-1}-1, \infty )} (x_3) (1+ x_3)^{-2}|v_3|^2 \\&\qquad - {\mathbf {1}}_{[ \ss ^{-1}-1, \infty )} (x_3) \frac{1}{1+ x_3} (|v_1|^2 + |v_2|^2) \\&\qquad - \partial _{x_3} {\mathfrak {z}}_{\ss } (x_3)(x_1v_1 + x_2 v_2) v_3 , \end{aligned}$$

where we have used \(\partial _{x_3} {\mathfrak {z}}_{\ss } (x_3)= {\mathbf {1}}_{[ \ss ^{-1}-1,\infty )} (x_3) \frac{-1}{(1+ x_3)^2}\). The last term, the sole term without a sign, can be bounded as

$$\begin{aligned}&|- \partial _{x_3} {\mathfrak {z}}_{\ss } (x_3)(x_1v_1 + x_2 v_2) v_3 |\\&\quad \le \ 2\sqrt{2 } \pi {\mathbf {1}}_{[ \ss ^{-1}-1,\infty )} (x_3) (1+ x_3)^{-2} (|v_1|^2 + |v_2|^2)^{1/2} |v_3| \\&\quad \le \ 4 \pi ^2 {\mathbf {1}}_{[ \ss ^{-1}-1,\infty )} (x_3) (1+ x_3)^{-2} (|v_1|^2 + |v_2|^2) \nonumber \\&\qquad + \frac{1}{2}{\mathbf {1}}_{[ \ss ^{-1}-1,\infty )} (x_3) (1+ x_3)^{-2} |v_3|^2 . \end{aligned}$$

Therefore we conclude that

$$\begin{aligned} \begin{aligned} - v\cdot \nabla _x {\mathfrak {w}}_{\varrho , \ss }(x,v) \ge&\ \Big \{ \ss {\mathbf {1}}_{[0, \ss ^{-1}-1]} (x_3) |v |^2 + \frac{1}{2} {\mathbf {1}}_{[ \ss ^{-1}-1, \infty )} (x_3) (1+ x_3)^{-2}|v_3|^2 \\&+(1- 4 \pi ^2 \ss ) {\mathbf {1}}_{[\ss ^{-1}-1, \infty )} (x_3) \frac{1}{1+ x_3} (|v_1|^2 + |v_2|^2) \Big \} {\mathfrak {w}}_{\varrho , \ss }(x,v) \\ \ge&\ \frac{{\mathfrak {z}}_{\ss } (x_3)}{2} |v|^2 {\mathfrak {w}}_{\varrho , \ss }(x,v). \end{aligned}\nonumber \\ \end{aligned}$$
(4.130)

We consider

$$\begin{aligned} h(t,x,v) = {\mathfrak {w}}_{\varrho , \ss }(x,v) f_R(t,x,v). \end{aligned}$$
(4.131)

An equation for h can be written from (3.2) and (3.8) as

$$\begin{aligned} \partial _t h+ \frac{1}{\varepsilon } v\cdot \nabla _x h+ \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } h = \frac{1}{\varepsilon ^2 \kappa }K_{{\mathfrak {w}}} h + {\mathcal {S}}_h , \end{aligned}$$
(4.132)
$$\begin{aligned} h|_{\gamma _-} = {\mathfrak {w}} P_{\gamma _+} \Big (\frac{h}{{\mathfrak {w}}}\Big )+ r. \end{aligned}$$
(4.133)

For (4.131), we have \(r=- \frac{\varepsilon }{\delta } {\mathfrak {w}} (1- P_{\gamma _+}) f_2\) and \({\mathcal {S}}_h: = \frac{\delta }{\kappa \varepsilon } \Gamma _{ {{\mathfrak {w}} }}( {h} , {h} ) +\frac{2}{\kappa } \Gamma _{{\mathfrak {w}}} ( {\mathfrak {w}} f_2,h) + {\mathfrak {w}} ({\mathbf {I}} - {\mathbf {P}}) {\mathfrak {R}}_1 +{\mathfrak {w}} {\mathbf {R}}_2, \) and

$$\begin{aligned} \nu _{\ss } := \nu (v) - \varepsilon \kappa \frac{v \cdot \nabla _x {\mathfrak {w}}_{\varrho , \ss }}{ {\mathfrak {w}}_{\varrho , \ss }} + \varepsilon ^2 \kappa \frac{(\partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} , \end{aligned}$$
(4.134)

where we denote \(\Gamma _{{\mathfrak {w}}}(\cdot , \cdot )(v) := {\mathfrak {w}} (v)\Gamma (\frac{\cdot }{{\mathfrak {w}}}, \frac{\cdot }{{\mathfrak {w}}})(v)\) and \(K_{{\mathfrak {w}}}(\cdot ):= {\mathfrak {w}} K( \frac{\cdot }{{\mathfrak {w}}})\).

If we have

$$\begin{aligned} \varepsilon ^{5/2} \kappa |\partial _t u| + \varepsilon ^{1/2}\sup _{x \in \Omega }(1+ x_3) |\nabla _x u(t,x)| <\infty , \end{aligned}$$
(4.135)

then for sufficiently small \(\varepsilon , \kappa >0\), from (4.130),

$$\begin{aligned} \nu _{\ss }\ge & {} \nu (v) +\frac{ \varepsilon \kappa }{2} {\mathfrak {z}}_{\ss } (x_3) |v|^2 - \varepsilon ^2\kappa \{ \varepsilon |\partial _t u| + |\nabla _x u| |v| \}|v-\varepsilon u| \nonumber \\\ge & {} \frac{\nu (v) }{2} + \frac{ \varepsilon \kappa }{4} {\mathfrak {z}}_{\ss } (x_3) |v|^2 . \end{aligned}$$
(4.136)

From (1.20), (1.22), and (2.3)

$$\begin{aligned} \begin{aligned}&|{\mathfrak {w}}(v) \Gamma ( \frac{h}{{\mathfrak {w}}}, \frac{h}{{\mathfrak {w}}}) (v)| \\&\quad \le \iint _{{\mathbb {R}}^3 \times {\mathbb {S}}^2} |(v-v_*) \cdot {\mathfrak {u}}| \sqrt{ \mu (v_* )} e^{- \varrho |v_*|^2+ \frac{\varrho }{2 } |v_*|}\\&\qquad \qquad \times \big \{ |h( v ^\prime )|| h( v_*^\prime )| +| h( v )||h( v_* )| \big \} \mathrm {d}{\mathfrak {u}} \mathrm {d}v_* \\&\quad \lesssim _\varrho \nu (v) \Vert h\Vert _{L^\infty _v}^2. \end{aligned} \end{aligned}$$
(4.137)

From (3.20) clearly we have

$$\begin{aligned} {\mathbf {k}}(v,v_*) \frac{{\mathfrak {w}}_{\varrho , \ss }(v)}{{\mathfrak {w}}_{\varrho , \ss }(v _*)} \le {\mathbf {k}}_{{\mathfrak {w}}} (v,v_*): = \frac{2C_{2}}{|v-v_*|} e^{- \frac{|v-v_*|^2}{8} - \frac{1}{8} \frac{(|v-\varepsilon u|^2 - |v_*- \varepsilon u|^2)^2}{|v-v_*|^2}} \frac{{\mathfrak {w}}_{\varrho , \ss }(v)}{{\mathfrak {w}}_{\varrho , \ss }(v _*)} .\nonumber \\ \end{aligned}$$
(4.138)

As in (3.23) we derive

$$\begin{aligned} \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}} (v,v_*) \mathrm {d}v_* \lesssim \frac{1}{1+ |v|}. \end{aligned}$$
(4.139)

Proposition 10

Recall \({\mathfrak {w}}_{\varrho ,\ss }\) in (2.3). Assume the same assumptions in Proposition 6. In addition we assume (4.135), and the conditions of \(\varrho \) and \(\ss \) in (2.3). Then

$$\begin{aligned} \begin{aligned}&d_\infty \Vert {\mathfrak {w}}_{\varrho , \ss } f _R \Vert _ {L^\infty _{t,x,v}} \\&\quad \lesssim \ \Vert {\mathfrak {w}}_{\varrho , \ss } f (0)\Vert _ {L^\infty _{ x,v}} +\frac{\varepsilon }{\delta } \Vert (3.10) \Vert _{L^\infty _{t,x}} + \varepsilon ^2 \kappa ( \Vert (3.4)\Vert _{L^\infty _{t,x}}+ \Vert (3.5)\Vert _{L^\infty _{t,x}}) \\&\qquad + \frac{1}{\varepsilon ^{1/2} \kappa ^{1/2}} \Vert P f_R \Vert _{L^\infty _tL^6_ {x}} + \frac{1}{\varepsilon ^{3/2} \kappa ^{3/2}}\Big \{ \Vert \sqrt{\nu }({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \\&\qquad + \Vert \sqrt{\nu }({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \Big \}\\&\qquad { +\frac{1}{\varepsilon ^{1/2}\kappa ^{3/2}} \Vert \partial _t u \Vert _{L^\infty _{t,x}} \Vert Pf_R \Vert _{L^2_{t,x}} } , \end{aligned} \end{aligned}$$
(4.140)

where

$$\begin{aligned} d_{\infty }:= 1- \varepsilon ^2 \Vert (3.10) \Vert _{L^\infty _{t,x}} - \varepsilon \delta \Vert {\mathfrak {w}}_{\varrho , \ss } f_R \Vert _{L^\infty _{t,x,v}} . \end{aligned}$$
(4.141)

Proposition 11

Assume the same assumptions of Proposition 10. We denote

$$\begin{aligned} {\mathfrak {w}}^\prime (x,v) : = {\mathfrak {w}}_{\varrho ^\prime , \ss }(x,v) \ \ \text {for} \ \varrho ^\prime <\varrho . \end{aligned}$$
(4.142)

Let \(p<3\). Then

$$\begin{aligned} \begin{aligned}&d_{\infty , t} \Vert {\mathfrak {w}} ^\prime \partial _t f_R \Vert _{L^2_t ((0,T);L^{\infty }_{x,v} (\Omega \times {\mathbb {R}}^3))}\\&\quad \lesssim \ \varepsilon \kappa ^{1/2} \Vert {\mathfrak {w}} ^\prime \partial _t f_R(0 ) \Vert _{ L^{\infty }_{x,v}}+ \frac{1}{\varepsilon ^{3/p} \kappa ^{3/p}} \Vert P \partial _t f \Vert _{L^2_t L^p_{x} } \\&\qquad + \frac{1}{\varepsilon ^{3/2} \kappa ^{3/2} } \Vert \sqrt{\nu } ( {\mathbf {I}} - {\mathbf {P}}) \partial _t f \Vert _{L^2 _{t,x,v}} \\&\qquad + \frac{\varepsilon }{\delta } \Vert (3.11)\Vert _{ L^{\infty }_{x,v}} + \frac{\varepsilon ^2}{\delta } \Vert \partial _t u\Vert _{ L^{\infty }_{x,v}} \Vert (3.10) \Vert _{ L^{\infty }_{x,v}}\\&\qquad + \varepsilon ^2 \kappa \Vert (3.6)\Vert _{ L^{\infty }_{x,v}}+ \varepsilon ^2 \kappa \Vert (3.7) \Vert _{ L^2_tL^{\infty }_{x}}\\&\qquad +\varepsilon \big ( \Vert \partial _t u \Vert _{L^\infty _{t,x}}+ \varepsilon \Vert (3.11)\Vert _{L^\infty _{t,x}}+ \varepsilon \kappa \Vert (3.13) \Vert _{L^2_tL^\infty _{x}} \big ) \Vert {\mathfrak {w}} f_R\Vert _{L^\infty _{t,x,v}}\\&\qquad + \varepsilon \Big ( \varepsilon \Vert (3.11)\Vert _{L^\infty _{t,x}} + \varepsilon \kappa \Vert (3.13)\Vert _{L^2_tL^\infty _{x}} \\&\qquad + \Vert \partial _t u\Vert _{L^\infty _{t,x}} \big (1+\varepsilon ^2\Vert (3.10) \Vert _{L^\infty _{t,x}} + \varepsilon \delta \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}} \big ) \Big ) \Vert {\mathfrak {w}} f_R\Vert _{L^\infty _{t,x,v}}, \end{aligned} \end{aligned}$$
(4.143)

with

$$\begin{aligned} d_{\infty , t}: = 1- \varepsilon ^2 \Vert (3.10)\Vert _{L^\infty _{t,x}} - \varepsilon \delta \Vert {\mathfrak {w}} f_R\Vert _{L^\infty _{t,x,v}}. \end{aligned}$$
(4.144)

In the proof of propositions, for simplicity, we often use \(\Vert \ \cdot \ \Vert _\infty \) for \(\Vert \ \cdot \ \Vert _{L^\infty _{t,x,v}}\), \(\Vert \ \cdot \ \Vert _{L^\infty _{x,v}}\) or \(\Vert \ \cdot \ \Vert _{L^\infty _{x}}\) if there would be no confusion.

Proof of Proposition 10

We define backward exit time and position as

$$\begin{aligned} t_{{\mathbf {b}}}(x,v) : = \varepsilon \frac{x_3}{v_3}, \ \ \ x_{{\mathbf {b}}}(x,v) := x- \frac{x_3}{v_3} v \ \ \text {for} \ \ (x,v) \in \Omega \times {\mathbb {R}}^3. \end{aligned}$$
(4.145)

Since the characteristics for (4.132) are given by \((x- \frac{t-s}{\varepsilon }v, v)\), we have, for \(0 \le t-s < t_{{\mathbf {b}}}(x,v)\),

$$\begin{aligned} \frac{d}{ds}\Big \{ e^{-\int ^t_s \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } } h(s,x- \frac{t-s}{\varepsilon }v, v ) \Big \}=e^{-\int ^t_s \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } } \Big \{ \frac{1}{\varepsilon ^2 \kappa }K_{{\mathfrak {w}}} h +{\mathcal {S}}_h \Big \}(s,x- \frac{t-s}{\varepsilon }v, v ).\nonumber \\ \end{aligned}$$
(4.146)

Here \(e^{-\int ^t_s \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } }=e^{-\int ^t_s \frac{\nu _{\ss } (\tau , x- \frac{t-\tau }{\varepsilon }v, v )}{\varepsilon ^2 \kappa } \mathrm {d}\tau }\). We regard \((x_1- \frac{t-s}{\varepsilon } v_1, x_2- \frac{t-s}{\varepsilon } v_2) \in {\mathbb {R}}^2\) belongs to \({\mathbb {T}}^2\) without redefining them in \([- \pi , \pi ]^2\).

Now we represent h using (4.146) and (4.133) as

$$\begin{aligned} h(t,x,v) =&{\mathbf {1}}_{t-t_{{\mathbf {b}}}(x,v)<0}e^{-\int ^t_0 \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } } h(0,x- \frac{t }{\varepsilon }v, v ) \nonumber \\&+ \int ^t_{\max \{0, t-t_{{\mathbf {b}}}(x,v)\}} e^{-\int ^t_s \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } } \frac{1}{\varepsilon ^2 \kappa }K_{{\mathfrak {w}}} h (s,x- \frac{t-s}{\varepsilon }v, v ) \mathrm {d}s \end{aligned}$$
(4.147)
$$\begin{aligned}&+ \int ^t_{\max \{0, t-t_{{\mathbf {b}}}(x,v)\}} e^{-\int ^t_s \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } } {\mathcal {S}}_h(s,x- \frac{t-s}{\varepsilon }v, v ) \mathrm {d}s \nonumber \\&+ {\mathbf {1}}_{t-t_{{\mathbf {b}}}(x,v)\ge 0}e^{-\int ^t_{t-t_{{\mathbf {b}}}(x,v)} \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } } h(t-t_{{\mathbf {b}}}(x,v),x_{{\mathbf {b}}}(x,v), v ) . \end{aligned}$$
(4.148)

Since the integrand of (4.148) reads on the boundary, using the boundary condition (4.133) and (4.146) again, we represent it as

$$\begin{aligned}&h(t-t_{{\mathbf {b}}}(x,v),x_{{\mathbf {b}}}(x,v), v )\nonumber \\&\quad = {\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} h(t-t_{{\mathbf {b}}}(x,v),x_{{\mathbf {b}}}(x,v), {\mathfrak {v}}) \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), {\mathfrak {v}}) }\mathrm {d}{\mathfrak {v}}\nonumber \\&\qquad + r(t-t_{{\mathbf {b}}}(x,v),x_{{\mathbf {b}}}(x,v), v ) \nonumber \\&\quad = {\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} e^{ - \int ^{t-t_{{\mathbf {b}}}(x,v)}_0 \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } }\nonumber \\&\qquad \times h(0,x_{{\mathbf {b}}}(x,v)- \frac{t-t_{{\mathbf {b}}}(x,v)}{\varepsilon } {\mathfrak {v}}, {\mathfrak {v}}) \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), {\mathfrak {v}}) }\mathrm {d}{\mathfrak {v}} \nonumber \\&\qquad +{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} \int ^{t-t_{{\mathbf {b}}}(x,v)}_0 e^{- \int ^{t-t_{{\mathbf {b}}}(x,v)}_s \frac{\nu _{\ss }}{\varepsilon ^2 \kappa } } \nonumber \\&\qquad \times \frac{1}{\varepsilon ^2 \kappa } K_{{\mathfrak {w}}} h (s, x_{{\mathbf {b}}}(x,v) - \frac{t-t_{{\mathbf {b}}}(x,v)-s}{\varepsilon } {\mathfrak {v}}, {\mathfrak {v}} ) \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), {\mathfrak {v}}) } \mathrm {d}s \mathrm {d}{\mathfrak {v}}\nonumber \\&\qquad +{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} \int ^{t-t_{{\mathbf {b}}}(x,v)}_0 e^{- \int ^{t-t_{{\mathbf {b}}}(x,v)}_s \frac{\nu _{\ss }}{\varepsilon ^2 \kappa } } \nonumber \\&\qquad \times {\mathcal {S}}_h (s, x_{{\mathbf {b}}}(x,v) - \frac{t-t_{{\mathbf {b}}}(x,v)-s}{\varepsilon } {\mathfrak {v}}, {\mathfrak {v}} ) \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), {\mathfrak {v}}) } \mathrm {d}s \mathrm {d}{\mathfrak {v}}\nonumber \\&\qquad + r (t-t_{{\mathbf {b}}}(x,v),x_{{\mathbf {b}}}(x,v), v ) , \end{aligned}$$
(4.149)

where \(r= - \frac{\varepsilon }{\delta } {\mathfrak {w}} (1- P_{\gamma _+}) f_2\) and \(e^{- \int ^{t-t_{{\mathbf {b}}}(x,v)}_0 \frac{\nu _{\ss } }{\varepsilon ^2 \kappa } }:= e^{-\int ^{t-t_{{\mathbf {b}}}(x,v)}_0 \frac{1}{\varepsilon ^2 \kappa } {\nu _{\ss } (\tau , x- \frac{ t_{{\mathbf {b}}}(x,v) }{\varepsilon }v - \frac{t-t_{{\mathbf {b}}}(x,v) -s }{\varepsilon } {\mathfrak {v}} , {\mathfrak {v}} )}\mathrm {d}\tau }\).

Note that, from (3.4), (3.5), (3.57), (3.58), and (4.137),

$$\begin{aligned} \begin{aligned} |{\mathcal {S}}_h(s, x- \frac{t-s}{\varepsilon } v,v)|&\lesssim \nu (v)\frac{\delta }{\kappa \varepsilon }\Vert h \Vert _\infty ^2 + \frac{\nu (v)}{\kappa } \Vert (3.10)\Vert _\infty \Vert h \Vert _\infty \\&\quad +\Vert (3.4)\Vert _\infty + \Vert (3.5)\Vert _\infty ,\\ | {\mathfrak {w}} (1- P_{\gamma _+}) f_2|&\lesssim \Vert (3.10)\Vert _\infty . \end{aligned} \end{aligned}$$
(4.150)

We derive a preliminary estimate as

$$\begin{aligned}&|h(t,x,v) | \lesssim e^{- \frac{\nu }{2\varepsilon ^2 \kappa }t} \Vert h(0)\Vert _\infty \nonumber \\&\quad + \varepsilon \delta \sup _{0 \le s \le t}\Vert h(s) \Vert _\infty ^2 + \varepsilon ^2 \sup _{0 \le s \le t} \Vert (3.10)\Vert _\infty \Vert h(s) \Vert _\infty \nonumber \\&\quad +\frac{\varepsilon }{\delta } \sup _{0 \le s \le t} \Vert (3.10)\Vert _\infty + {\varepsilon ^2 \kappa } ( \Vert (3.4)\Vert _\infty + \Vert (3.5)\Vert _\infty ) \end{aligned}$$
(4.151)
$$\begin{aligned}&\qquad + \int ^t_0 \frac{e^{- \frac{\nu }{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}} (v,v_*) |h(s,x- \frac{t-s}{\varepsilon }, v_*)| \mathrm {d}v_* \mathrm {d}s \end{aligned}$$
(4.152)
$$\begin{aligned}&\quad +{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} \int ^{t-t_{{\mathbf {b}}}(x,v)}_0 \frac{ e^{- \frac{\nu }{ 2\varepsilon ^2 \kappa } (t-s) } }{\varepsilon ^2 \kappa }\nonumber \\&\quad \times \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}}({\mathfrak {v}} ,v_*) |h (s, x_{{\mathbf {b}}}(x,v) - \frac{t-t_{{\mathbf {b}}}(x,v)-s}{\varepsilon } {\mathfrak {v}}, v_* )| \mathrm {d}v_* \mathrm {d}s \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}} (x_{{\mathbf {b}}}(x,v), {\mathfrak {v}}) }\mathrm {d}{\mathfrak {v}}. \end{aligned}$$
(4.153)

We note that \(|h(s,x- \frac{t-s}{\varepsilon }, v_*)|\) has the same upper bound. Then we bound (4.152) by a summation of (4.151) and

$$\begin{aligned} \begin{aligned}&\sup _{ \begin{array}{c} (x_{{\mathbf {b}}}, v) \in \partial \Omega \times {\mathbb {R}}^3 \\ t-t_{{\mathbf {b}}}\ge 0 \end{array}} {\mathfrak {w}} (x_{{\mathbf {b}}}, v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} \int ^{t-t_{{\mathbf {b}}}}_0 \frac{ e^{- \frac{\nu }{ 2\varepsilon ^2 \kappa } (t-s) } }{\varepsilon ^2 \kappa }\\&\quad \times \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}}({\mathfrak {v}} ,v_*) |h (s, x_{{\mathbf {b}}}- \frac{t-t_{{\mathbf {b}}}-s}{\varepsilon } {\mathfrak {v}}, v_* )| \mathrm {d}v_* \mathrm {d}s \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}} (x_{{\mathbf {b}}}, {\mathfrak {v}}) }\mathrm {d}{\mathfrak {v}}, \end{aligned} \end{aligned}$$
(4.154)

and importantly

$$\begin{aligned} \begin{aligned}&\int ^t_0 \frac{e^{- \frac{\nu (v)}{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}} (v,v_*) \int ^s_0 \frac{e^{- \frac{\nu (v_*)}{2\varepsilon ^2 \kappa }(s-\tau )}}{\varepsilon ^2 \kappa }\\&\quad \times \int _{{\mathbb {R}}^3}{\mathbf {k}}_{{\mathfrak {w}}} (v_*,v_{**}) |h(s,x- \frac{t-s}{\varepsilon }v - \frac{s-\tau }{\varepsilon }v_* , v_{**})| \mathrm {d}v_{**} \mathrm {d}\tau \mathrm {d}v_* \mathrm {d}s. \end{aligned} \end{aligned}$$
(4.155)

We consider (4.155). We decompose the integration of \(\tau \in [0,s] = [0, s- o(1)\varepsilon ^2 \kappa ] \cup [s- o(1)\varepsilon ^2 \kappa , s]\). The contribution of \(\int ^s_{s-o(1) \varepsilon ^2 \kappa } \cdots \mathrm {d}\tau \) is bounded as

$$\begin{aligned}&\frac{2}{\nu (v)}\big (1- e^{- \frac{\nu (v)}{2 \varepsilon ^2 \kappa }}\big ) \Vert {\mathbf {k}}_{{\mathfrak {w}}}(v , \cdot )\Vert _{L^1} \frac{o(1) \varepsilon ^2 \kappa }{\varepsilon ^2 \kappa } \Vert {\mathbf {k}}_{{\mathfrak {w}}}(v_*, \cdot )\Vert _{L^1} \sup _{0 \le s \le t}\Vert h(s) \Vert _\infty \nonumber \\&\quad \le o(1) \sup _{0 \le s \le t}\Vert h(s) \Vert _\infty . \end{aligned}$$
(4.156)

For the rest of term we decompose \({\mathbf {k}}_{{\mathfrak {w}} }(v_*,v_{**})= {\mathbf {k}}_{{\mathfrak {w}},N}(v_*,v_{**}) + \{ {\mathbf {k}}_{{\mathfrak {w}} }(v_*,v_{**})- {\mathbf {k}}_{{\mathfrak {w}},N}(v_*,v_{**})\}\) where \({\mathbf {k}}_{{\mathfrak {w}},N}(v_*,v_{**}):={\mathbf {k}}_{{\mathfrak {w}}}(v_*,v_{**})\) \( \times {\mathbf {1}}_{ \frac{1}{N}<|v_*- v_{**}|< N \ \& \ |v_*|< N }. \) From (4.139), \(\int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}}(v_*,v_{**}) {\mathbf {1}}_{|v_*|\ge N} \mathrm {d}v_{**} \lesssim 1/N\). Also from the fact \({\mathbf {k}}_{{\mathfrak {w}}}(v_*,v_{**}) \le \frac{e^{-C |v_*- v_{**}|^2}}{|v_*- v_{**}|} \in L^1(\{ v_*- v_{**} \in {\mathbb {R}}^3 \})\), \(\sup _{v_*}\int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}}(v_*,v_{**}) \{{\mathbf {1}}_{\frac{1}{N}\ge |v_*- v_{**}| } + {\mathbf {1}}_{ |v_*- v_{**}|\ge N } \}\mathrm {d}v_{**} \downarrow 0\) as \(N \rightarrow \infty \). Hence for \(N\gg 1\)

$$\begin{aligned} \begin{aligned} (4.155)&\le \int ^t_0 \frac{e^{- \frac{\nu (v)}{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}} ,N} (v,v_*) \int ^{s- o(1) \varepsilon ^2 \kappa }_0 \frac{e^{- \frac{\nu (v_*)}{2\varepsilon ^2 \kappa }(s-\tau )}}{\varepsilon ^2 \kappa }\\&\ \ \ \ \times \int _{{\mathbb {R}}^3}{\mathbf {k}}_{{\mathfrak {w}},N } (v_*,v_{**}) |h(s,x- \frac{t-s}{\varepsilon }v - \frac{s-\tau }{\varepsilon }v_* , v_{**})| \mathrm {d}v_{**} \mathrm {d}\tau \mathrm {d}v_* \mathrm {d}s \\&\le C_N \int ^t_0 \frac{e^{- \frac{\nu (v)}{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int _{|v_*| \le 2N} \int ^{s- o(1) \varepsilon ^2 \kappa }_0 \frac{e^{- \frac{\nu (v_*)}{2\varepsilon ^2 \kappa }(s-\tau )}}{\varepsilon ^2 \kappa } \\&\quad \times \int _{|v_{**}| < 2N} |f_R(s,x- \frac{t-s}{\varepsilon }v - \frac{s-\tau }{\varepsilon }v_* , v_{**})| \mathrm {d}v_{**} \mathrm {d}\tau \mathrm {d}v_* \mathrm {d}s \\&\quad + o(1) \sup _{0 \le s \le t} \Vert h(s)\Vert _{L^\infty _{x,v}}, \end{aligned} \end{aligned}$$
(4.157)

where we have used the fact \(\sup _{x}{\mathbf {k}}_{{\mathfrak {w}} } (v_*, v_{**}) {\mathfrak {w}}_{\varrho , \ss }(v_{**}) \le C_N<\infty \) when \(\frac{1}{N}<|v_*- v_{**}|< N\) and \(|v_*|< N\) (then \(|v_{**}|< 2N\)).

Now we decompose \(f_R={\mathbf {P}}f_R+ ({\mathbf {I}} -{\mathbf {P}})f_R\). We first take integrations (4.157) over \(v_{*}\) and \(v_{**}\) and use Holder’s inequality with \(p=6, p=2\) in \(1/p+ 1/p^\prime =1\) for \({\mathbf {P}}f_R , ({\mathbf {I}} -{\mathbf {P}})f_R\) respectively to derive

$$\begin{aligned} \begin{aligned}&(4.157)\\&\quad \le \ (4N)^3C_N \frac{1}{\nu (v)} \sup _{\begin{array}{c} 0 \le s \le t\\ 0 \le \tau \le s- o(1) \varepsilon ^2 \kappa \end{array}} \left( \iint _{|v_*|\le N,|v_{**}| \le 2 N} \right. \\&\qquad \qquad \qquad \left. |{\mathbf {P}}f_R(s, x- \frac{t-s}{\varepsilon } v- \frac{s-\tau }{\varepsilon } v_*, v_{**})|^6 \mathrm {d}v_{**}\mathrm {d}v_*\right) ^{1/6} \\&\qquad + (4N)^3C_N \frac{1}{\nu (v)} \sup _{\begin{array}{c} 0 \le s \le t\\ 0 \le \tau \le s- o(1) \varepsilon ^2 \kappa \end{array}} \left( \iint _{|v_*|\le N,|v_{**}| \le 2 N} \right. \\&\qquad \qquad \qquad \left. | ( {\mathbf {I}}-{\mathbf {P}})f_R(s, x- \frac{t-s}{\varepsilon } v- \frac{s-\tau }{\varepsilon } v_*, v_{**})|^2 \mathrm {d}v_{**}\mathrm {d}v_*\right) ^{1/2} . \end{aligned} \end{aligned}$$
(4.158)

Now we consider a map

$$\begin{aligned}&v_* \in \{{\mathbb {R}}^3: |v_*| \le N\} \mapsto y\nonumber \\&\quad :=x- \frac{t-s}{\varepsilon } v - \frac{s-\tau }{\varepsilon } v_* \in \Omega , \ \ \text {where} \ \ \Big |\frac{\partial y}{\partial v_*}\Big |= \Big |\frac{s-\tau }{\varepsilon }\Big |^3 > rsim \varepsilon ^3 \kappa ^3.\qquad \end{aligned}$$
(4.159)

We note that this mapping is not one-to-one and the image can cover \(\Omega \) at most N times. Therefore we have

$$\begin{aligned}&\left( \iint _{|v_*|\le N,|v_{**}| \le N} |{\mathbf {P}}f_R(s, x- \frac{t-s}{\varepsilon } v - \frac{s-\tau }{\varepsilon } v_*, v_{**})|^6 \mathrm {d}v_{**}\mathrm {d}v_*\right) ^{1/6}\\&\quad \le N^{1/6} \left( \iint _{|v_*|\le N,|v_{**}| \le N} |{\mathbf {P}}f_R(s, y, v_{**})|^6 \mathrm {d}v_{**} \frac{\mathrm {d}y}{\varepsilon ^3 \kappa ^3}\right) ^{1/6} \\&\quad \le \frac{N^{1/6}}{\varepsilon ^{1/2} \kappa ^{1/2}} \Vert {\mathbf {P}} f_R(s) \Vert _{L^6_{x,v}}, \\&\left( \iint _{|v_*|\le N,|v_{**}| \le N} |( {\mathbf {I}}-{\mathbf {P}})f_R(s, x- \frac{t-s}{\varepsilon } v - \frac{s-\tau }{\varepsilon } v_*, v_{**})|^2 \mathrm {d}v_{**}\mathrm {d}v_*\right) ^{1/2}\\&\quad \le \frac{N^{1/2}}{\varepsilon ^{3/2} \kappa ^{3/2}} \Vert ( {\mathbf {I}}-{\mathbf {P}})f_R(s) \Vert _{L^2_{x,v}}. \end{aligned}$$

Therefore we conclude that

$$\begin{aligned}&(4.155)\nonumber \\&\quad \le (4N)^3C_N(4.158) + o(1)\sup _{0 \le s \le t} \Vert h(s)\Vert _{L^\infty _{x,v}} \nonumber \\&\quad \le (4N)^4C_N\left\{ \frac{1}{\varepsilon ^{1/2}\kappa ^{1/2}} \sup _{0 \le s \le t}\Vert {\mathbf {P}} f_R(s ) \Vert _{L^6_{x,v}} + \frac{1}{\varepsilon ^{3/2}\kappa ^{3/2}} \sup _{0 \le s \le t} \Vert ({\mathbf {I}} -{\mathbf {P}}) f_R(s ) \Vert _{L^2_{x,v}} \right\} \nonumber \\&\qquad + o(1)\sup _{0 \le s \le t} \Vert h(s)\Vert _{L^\infty _{x,v}} \nonumber \\&\quad \lesssim _N \frac{1}{\varepsilon ^{1/2}\kappa ^{1/2}} \sup _{0 \le s \le t}\Vert {\mathbf {P}} f_R(s ) \Vert _{L^6_{x,v}} + \frac{1}{\varepsilon ^{3/2}\kappa ^{3/2}} \Big \{ \Vert ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}} \nonumber \\&\qquad + \Vert ({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \Big \} \nonumber \\&\qquad { + \frac{1}{\varepsilon ^{1/2} \kappa ^{3/2}} \Vert \partial _t u \Vert _{L^\infty _{t,x}}\Vert {P} f_R \Vert _{L^2_{t,x}} } + o(1)\sup _{0 \le s \le t} \Vert h(s)\Vert _{L^\infty _{x,v}}, \end{aligned}$$
(4.160)

where we have used (A.1) the Sobolev embedding in 1D at the last line.

Now we consider (4.153) and (4.154). We decompose \(s \in [0, t-t_{{\mathbf {b}}}] = [0, t-t_{{\mathbf {b}}}- o(1) \varepsilon ^2 \kappa ] \cup [t-t_{{\mathbf {b}}}- o(1) \varepsilon ^2 \kappa , t-t_{{\mathbf {b}}}]\). The contribution of \(\int ^{t-t_{{\mathbf {b}}}}_{t-t_{{\mathbf {b}}}- o(1) \varepsilon ^2 \kappa } \cdots \) is bounded as

$$\begin{aligned} \frac{o(1) \varepsilon ^2 \kappa }{\varepsilon ^2 \kappa } \Vert {\mathbf {k}}_{{\mathfrak {w}}}({\mathfrak {v}}, \cdot ) \Vert _{L^1} \sup _{0 \le s \le t} \Vert h(s) \Vert _\infty \le o(1)\sup _{0 \le s \le t} \Vert h(s) \Vert _\infty . \end{aligned}$$
(4.161)

For \(s \in [0, t-t_{{\mathbf {b}}}- o(1) \varepsilon ^2 \kappa ]\) we consider a map as (4.159)

$$\begin{aligned}&{\mathfrak {v}} \in \{{\mathfrak {v}} \in {\mathbb {R}}^3: {\mathfrak {v}} _3<0 \} \mapsto y\nonumber \\&\quad :=x_{{\mathbf {b}}}- \frac{t-t_{{\mathbf {b}}}-s}{\varepsilon } {\mathfrak {v}} \in \Omega , \ \ \text {where} \ \ \left| \frac{\partial y}{\partial {\mathfrak {v}}}\right| = \left| \frac{t-t_{{\mathbf {b}}}-s}{\varepsilon }\right| ^3 > rsim \varepsilon ^3 \kappa ^3. \end{aligned}$$
(4.162)

Following the argument to have (4.158) we bound

$$\begin{aligned} \begin{aligned}&\text {the contribution of} \ \int ^{t-t_{{\mathbf {b}}}- o(1) \varepsilon ^2 \kappa }_0 \cdots \ \text {of } \ (4.154)\\&\quad \lesssim _N \frac{1}{\varepsilon ^{1/2} \kappa ^{1/2}} \Vert {\mathbf {P}} f_R(s) \Vert _{L^6_ {x,v}} + \frac{1}{\varepsilon ^{3/2} \kappa ^{3/2}} \Big \{ \Vert ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}}\\&\qquad + \Vert ({\mathbf {I}} -{\mathbf {P}}) \partial _t f_R \Vert _{L^2_{t,x,v}} \Big \} { + \frac{1}{\varepsilon ^{1/2} \kappa ^{3/2}} \Vert \partial _t u \Vert _{L^\infty _{t,x}}\Vert {P} f_R \Vert _{L^2_{t,x}} } . \end{aligned} \end{aligned}$$
(4.163)

In conclusion, we bound |h(txv)| by (4.151), (4.160), (4.161), (4.163) and conclude (4.140) by choosing small enough o(1) in (4.160) and (4.161). \(\square \)

Proof of Proposition 11

Since many parts of the proof are overlapped with the proof of Proposition 10 we only pin point the differences. An equation for \({\mathfrak {w}}^\prime \partial _t f_R\) takes the similar form of (4.132) and (4.133). We can read (3.3) for

$$\begin{aligned} h(t,x,v) = {\mathfrak {w}} ^\prime (x,v) \partial _t f_R(t,x,v), \ \ \text {for} \ \varrho ^\prime <\varrho , \end{aligned}$$
(4.164)

as (4.132) and (4.133) replacing

$$\begin{aligned} \begin{aligned} {\mathcal {S}}_h&= \frac{2}{\kappa } \Gamma _{{\mathfrak {w}}^\prime }( {{\mathfrak {w}}^\prime }{f_2},h) + \frac{ 2 \delta }{\varepsilon \kappa }\Gamma _{{\mathfrak {w}}^\prime }( {{\mathfrak {w}}^\prime } f_R,h)+ \frac{2}{\kappa } \Gamma _{{\mathfrak {w}}^\prime }( {{\mathfrak {w}}^\prime }\partial _t {f_2}, {{\mathfrak {w}}^\prime }f_R) \\&\quad - \partial _t \Big ( \frac{( \partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} \Big ) \frac{{\mathfrak {w}}^\prime }{{\mathfrak {w}} } {\mathfrak {w}} f_{R} + {{\mathfrak {w}}^\prime } ({\mathbf {I}} - {\mathbf {P}}){\mathfrak {R}}_3 + {{\mathfrak {w}}^\prime } {\mathfrak {R}}_4\\&\quad - \frac{1}{\varepsilon ^2 \kappa } {\mathfrak {w}}^\prime L_t ({\mathbf {I}} - {\mathbf {P}}) f_R +\frac{1}{\varepsilon ^2 \kappa } {\mathfrak {w}}^\prime L({\mathbf {P}}_t f_R)\\&\quad + \frac{2}{\kappa } {\mathfrak {w}}^\prime \Gamma _t( {f_2}, f_R)+ \frac{\delta }{\varepsilon \kappa } {\mathfrak {w}}^\prime \Gamma _t (f_R,f_R) , \\ r&= - \frac{\varepsilon }{\delta } {{\mathfrak {w}}^\prime } (1-P_{\gamma _+}) \partial _tf_2 +{{\mathfrak {w}}^\prime } r_{\gamma _+} (f_R)-{{\mathfrak {w}}^\prime } \frac{\varepsilon }{\delta }r_{\gamma _+} (f_2), \end{aligned} \end{aligned}$$
(4.165)

where \(r _{\gamma _+}(g) \) has been defined in (3.9).

We have the same equality of (4.147), (4.148) with (4.149) for h of (4.164) but replacing \(S_h\) and r of (4.165). Note that \(\frac{{\mathfrak {w}}^\prime (x,v)}{{\mathfrak {w}} (x,v)}\lesssim e^{- (\varrho - \varrho ^\prime )|v|^2}\) and hence \(\Big |\partial _t \Big ( \frac{( \partial _t + \varepsilon ^{-1} v\cdot \nabla _x) \sqrt{\mu }}{\sqrt{\mu }} \Big ) \frac{{\mathfrak {w}}^\prime }{{\mathfrak {w}} }\Big | \lesssim (3.13) \) from (3.13). From (3.36), (3.6), (3.7), (3.57), (3.12), (3.13), (3.58), and (4.137), we bound terms of (4.165)

$$\begin{aligned} |S_h|\lesssim & {} \nu (v) \big \{\frac{1}{\kappa } | (3.10) | +\frac{\delta }{\kappa \varepsilon } \Vert {\mathfrak {w}} f_R\Vert _\infty \big \} \Vert h\Vert _\infty + (3.6) + (3.7)\nonumber \\&+ \Big ( \frac{\nu (v)}{\kappa } (3.11) + (3.13) + |\partial _t u| \big (\frac{1}{\varepsilon \kappa } + \frac{\varepsilon }{\kappa } (3.10) + \frac{\delta }{ \kappa }\Vert {\mathfrak {w}} f_R \Vert _\infty \big ) \Big ) \Vert {\mathfrak {w}} f_R \Vert _\infty , \nonumber \\\end{aligned}$$
(4.166)
$$\begin{aligned} |r|\lesssim & {} \ \frac{\varepsilon }{\delta } (3.11) + \frac{\varepsilon ^2}{\delta } |\partial _t u| (3.10) + \varepsilon |\partial _t u |\Vert {\mathfrak {w}} f_R\Vert _\infty . \end{aligned}$$
(4.167)

Then as in (4.151)-(4.155) we derive a preliminary estimate as

$$\begin{aligned}&|h(t,x,v) | \nonumber \\&\quad \lesssim \ e^{- \frac{\nu }{2\varepsilon ^2 \kappa }t} \Vert h(0)\Vert _\infty + \frac{\varepsilon ^2 \kappa }{\nu (v)} (4.166) + (4.167) \end{aligned}$$
(4.168)
$$\begin{aligned}&\qquad + \int ^t_0 \frac{e^{- \frac{\nu }{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}^\prime } (v,v_*) |h(s,x- \frac{t-s}{\varepsilon }, v_*)| \mathrm {d}v_* \mathrm {d}s \end{aligned}$$
(4.169)
$$\begin{aligned}&\qquad +{\mathfrak {w}}^\prime (x_{{\mathbf {b}}}(x,v), v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} \int ^{t-t_{{\mathbf {b}}}(x,v)}_0 \frac{ e^{- \frac{\nu }{ 2\varepsilon ^2 \kappa } (t-s) } }{\varepsilon ^2 \kappa }\nonumber \\&\qquad \times \int _{{\mathbb {R}}^3} {\mathbf {k}}_{{\mathfrak {w}}^\prime }({\mathfrak {v}} ,v_*) |h (s, x_{{\mathbf {b}}}(x,v) \nonumber \\&\qquad - \frac{t-t_{{\mathbf {b}}}(x,v)-s}{\varepsilon } {\mathfrak {v}}, v_* )| \mathrm {d}v_* \mathrm {d}s \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}}^\prime (x_{{\mathbf {b}}}(x,v), {\mathfrak {v}}) }\mathrm {d}{\mathfrak {v}}. \end{aligned}$$
(4.170)

As (4.154) and (4.155), we bound (4.169) by a summation of (4.168) and

$$\begin{aligned}&\int ^t_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int ^{s-o(1) \varepsilon ^2 \kappa }_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }(s-\tau )}}{\varepsilon ^2 \kappa } \int _{|v_*| \le 2N} \nonumber \\&\quad \times \int _{|v_{**} | \le 2N} |h(s,x- \frac{t-s}{\varepsilon }v - \frac{s-\tau }{\varepsilon }v_* , v_{**})| \mathrm {d}v_{**} \mathrm {d}v_* \mathrm {d}\tau \mathrm {d}s , \end{aligned}$$
(4.171)
$$\begin{aligned}&\quad +\sup _{ \begin{array}{c} (x_{{\mathbf {b}}}, v) \in \partial \Omega \times {\mathbb {R}}^3 \\ t-t_{{\mathbf {b}}}\ge 0 \end{array}} {\mathfrak {w}}^\prime (x_{{\mathbf {b}}}, v ) c_\mu \sqrt{\mu (v)} \int _{{\mathfrak {v}}_3<0} \int ^{t-t_{{\mathbf {b}}}-o(1) \varepsilon ^2 \kappa }_0 \frac{ e^{- \frac{\nu }{ 2\varepsilon ^2 \kappa } (t-s) } }{\varepsilon ^2 \kappa }\nonumber \\&\quad \times \int _{|v_*| \le 2N} |h (s, x_{{\mathbf {b}}}- \frac{t-t_{{\mathbf {b}}}-s}{\varepsilon } {\mathfrak {v}}, v_* )| \mathrm {d}v_* \mathrm {d}s \frac{ \sqrt{\mu ({\mathfrak {v}})} |{\mathfrak {v}}_3| }{{\mathfrak {w}}^\prime (x_{{\mathbf {b}}}, {\mathfrak {v}}) }\mathrm {d}{\mathfrak {v}} \end{aligned}$$
(4.172)
$$\begin{aligned}&\quad + o(1) \sup _{0 \le s \le t} \Vert h(s) \Vert _{L^{\infty }_{x,v}} . \end{aligned}$$
(4.173)

Then we follow the argument of (4.158)-(4.160) to derive that, for \(p<3\),

$$\begin{aligned} |(4.171)| \lesssim&\ \int ^t_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int ^{s-o(1) \varepsilon ^2 \kappa }_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }(s-\tau )}}{\varepsilon ^2 \kappa } \frac{N^{1/3}}{\varepsilon ^{3/p} \kappa ^{3/p}} \Vert {\mathbf {P}} \partial _t f (\tau ) \Vert _{L^p_{x,v}} \mathrm {d}\tau \mathrm {d}s \end{aligned}$$
(4.174)
$$\begin{aligned}&+ \int ^t_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }(t-s)}}{\varepsilon ^2 \kappa } \int ^{s-o(1) \varepsilon ^2 \kappa }_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }(s-\tau )}}{\varepsilon ^2 \kappa } \frac{N^{1/2}}{\varepsilon ^{3/2} \kappa ^{3/2}} \Vert {\mathbf {P}} \partial _t f (\tau ) \Vert _{L^2_{x,v}} \mathrm {d}\tau \mathrm {d}s. \end{aligned}$$
(4.175)

Now we use the Young’s inequality for temporal convolution twice to derive that, for \(p<3\),

$$\begin{aligned} \begin{aligned}&\Vert (4.171)\Vert _{L^2_t(0,T)}\\&\quad \lesssim \ \bigg \Vert \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }|s |}}{\varepsilon ^2 \kappa } \bigg \Vert _{L^1_s({\mathbb {R}})} \bigg \Vert \int ^s_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }(s-\tau )}}{\varepsilon ^2 \kappa } \bigg ( \frac{N^{1/3}}{\varepsilon ^{3/p} \kappa ^{3/p}} \Vert {\mathbf {P}} \partial _t f (\tau ) \Vert _{L^p_{x,v}} \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad + \frac{N^{1/2}}{\varepsilon ^{3/2} \kappa ^{3/2}} \Vert ({\mathbf {I}}- {\mathbf {P}}) \partial _t f (\tau ) \Vert _{L^2_{x,v}} \bigg ) \mathrm {d}\tau \bigg \Vert _{L^2_s ({\mathbb {R}})} \\&\quad \lesssim \ \bigg \Vert \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }|s |}}{\varepsilon ^2 \kappa } \bigg \Vert _{L^1_s({\mathbb {R}})} \bigg \Vert \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }|\tau |}}{\varepsilon ^2 \kappa } \bigg \Vert _{L^1_\tau ({\mathbb {R}})} \\&\qquad \quad \times \bigg ( \frac{N^{1/3}}{\varepsilon ^{3/p} \kappa ^{3/p} } \Vert P \partial _t f \Vert _{L^2_t ((0,T);L^p_{x}(\Omega ))} + \frac{N^{1/2}}{\varepsilon ^{3/2} \kappa ^{3/2} } \Vert ({\mathbf {I}}- {\mathbf {P}}) \partial _t f \Vert _{L^2 ((0,T) \times \Omega \times {\mathbb {R}}^3)} \bigg )\\&\quad \lesssim _N \ \frac{1}{\varepsilon ^{3/p} \kappa ^{3/p} } \Vert P \partial _t f \Vert _{L^2_t ((0,T);L^p_{x}(\Omega ))} + \frac{1}{\varepsilon ^{3/2} \kappa ^{3/2} } \Vert ( {\mathbf {I}} - {\mathbf {P}}) \partial _t f \Vert _{L^2 ((0,T) \times \Omega \times {\mathbb {R}}^3)} . \end{aligned} \end{aligned}$$
(4.176)

As in (4.163), for (4.172) we use (4.162) to derive that, for \(p<3\),

$$\begin{aligned} \begin{aligned}&\Vert (4.172)\Vert _{L^2_t (0,T)}\\&\quad \lesssim \ \bigg \Vert \int ^t_0 \frac{e^{- \frac{C_\nu }{2\varepsilon ^2\kappa } (t-s)}}{\varepsilon ^2 \kappa } \bigg ( \frac{1}{\varepsilon ^{3/p} \kappa ^{3/p} } \Vert {\mathbf {P}} \partial _t f (s) \Vert _{L^p_{x,v}} \\&\qquad + \frac{1}{\varepsilon ^{3/2} \kappa ^{3/2}} \Vert ({\mathbf {I}}- {\mathbf {P}}) \partial _t f (s) \Vert _{L^2_{x,v}} \bigg ) \mathrm {d}s\bigg \Vert _{L^2_t (0,T)} \\&\quad \lesssim \ \bigg \Vert \frac{e^{- \frac{C_\nu }{2\varepsilon ^2 \kappa }|s |}}{\varepsilon ^2 \kappa } \bigg \Vert _{L^1_s({\mathbb {R}})} \Big \{\frac{1}{\varepsilon ^{3/p} \kappa ^{3/p} } \Vert P \partial _t f \Vert _{L^2_t ((0,T);L^p_{x}(\Omega ))} \\&\qquad + \frac{1}{\varepsilon ^{3/2} \kappa ^{3/2} } \Vert ( {\mathbf {I}} - {\mathbf {P}}) \partial _t f \Vert _{L^2 ((0,T) \times \Omega \times {\mathbb {R}}^3)} \Big \}\\&\quad \lesssim \ \frac{1}{\varepsilon ^{3/p} \kappa ^{3/p} } \Vert P \partial _t f \Vert _{L^2_t ((0,T);L^p_{x}(\Omega ))} + \frac{1}{\varepsilon ^{3/2} \kappa ^{3/2} } \Vert ( {\mathbf {I}} - {\mathbf {P}}) \partial _t f \Vert _{L^2 ((0,T) \times \Omega \times {\mathbb {R}}^3)}, \end{aligned} \end{aligned}$$
(4.177)

where we have used the Young’s inequality for temporal convolution.

In conclusion, we bound \(\Vert h\Vert _{L^2_t L^\infty _{x,v}}\) by \(\Vert (4.168)\Vert _{L^2_t L^\infty _{x,v}},\) (4.176), (4.173), (4.177) and conclude (4.143) by choosing small enough o(1) in (4.173). \(\square \)

4.5 Proof of Theorem 2

An existence of a unique global solution F for each \(\varepsilon >0\) can be found in [12,13,14,15]. Thereby we only focus on the (a priori) estimates (2.13).

Step 1. Define \(T_*>0\) as

$$\begin{aligned} \begin{aligned} T_*=&\sup \Big \{ t\ge 0: \ \min \{d_2, d_{2,t}, d_6, d_3, d_{3,t}, d_\infty , d_{\infty , t}\} \ge \frac{\sigma _0}{4}\\&\ \ \text {and} \ \ \frac{\delta \varepsilon ^{1/2}}{\kappa } \sqrt{{\mathcal {D}}(s)} + \varepsilon \delta \Vert {\mathfrak {w}}_{\varrho , \ss } f(s) \Vert _{L^\infty _{x,v}} + \frac{\varepsilon ^{1/2} \delta }{\kappa ^{1+ {\mathfrak {P}} }} \Vert Pf_R(s) \Vert _{L^2_x} \ll 1 \\&\text { for all} \ 0 \le s \le t \Big \}, \end{aligned}\qquad \end{aligned}$$
(4.178)

where \(d_2, d_{2,t}, d_6, d_3, d_{3,t}, d_\infty , d_{\infty , t}\) are defined in (4.3), (4.5), (4.45), (4.73), (4.75), (4.141) and (4.144).

From (2.10) and (4.178) we read all the estimates of Proposition 7, Proposition 8, Proposition 10, Proposition 9, and Proposition 11 in terms of \({\mathcal {E}} (t)\) and \({\mathcal {D}} (t)\) as follows.

From (4.140), (4.178), and (2.10)

$$\begin{aligned}&\sup _{0 \le s \le t} \Vert {\mathfrak {w}}_{\varrho , \ss } f _R(s) \Vert _{L^\infty _{x,v}}\nonumber \\&\quad \lesssim \ \frac{1}{\varepsilon ^{1/2} \kappa ^{1/2} } \sup _{0 \le s \le t} \Vert {\mathbf {P}} f_R(s) \Vert _{L^6_ {x,v}} \nonumber \\&\qquad + \frac{1}{\varepsilon ^{1/2} \kappa } \sqrt{ {\mathcal {D}}(t) } {+ \frac{1}{\varepsilon ^{1/2}\kappa ^{1+ {\mathfrak {P}}}} \Vert P f _R\Vert _{L^2_{t,x}} } \nonumber \\&\qquad + \Vert {\mathfrak {w}}_{\varrho , \ss } f (0)\Vert _\infty + \varepsilon ^{1/2} \exp \Big (\frac{3}{\kappa ^{{\mathfrak {P}}^\prime }}\Big ). \end{aligned}$$
(4.179)

Now applying (4.179) to (4.44) we derive that

$$\begin{aligned} \begin{aligned} \sup _{0 \le s \le t}\Vert {P} f_R(s) \Vert _{L^6_x}\lesssim&\ \frac{\varepsilon }{\kappa } \exp \Big (\frac{1}{\kappa ^{{\mathfrak {P}}^\prime }} \Big )\sup _{0 \le s \le t} \sqrt{{\mathcal {E}}(s)}+ \frac{1}{\kappa ^{1/2}} \sqrt{{\mathcal {D}}(t)} {+ \frac{1}{ \kappa ^{1/2 + {\mathfrak {P}}} } \Vert P f _R\Vert _{L^2_{t,x}} }\\&+ ( \varepsilon \kappa )^{\frac{1}{2}} \Vert {\mathfrak {w}}_{\varrho , \ss } f (0)\Vert _{L^\infty _{x,v}} +\varepsilon ^{1/2} \exp \Big (\frac{3}{\kappa ^{{\mathfrak {P}}^\prime }}\Big ) . \end{aligned}\nonumber \\ \end{aligned}$$
(4.180)

From (4.179), (4.180), and (4.178) and (2.10) we conclude that

$$\begin{aligned} \begin{aligned}&\sup _{0 \le s \le t} \big \{ \kappa ^{\frac{1}{2}} \Vert {P} f_R(s) \Vert _{L^6_x} + \varepsilon ^{\frac{1}{2}}\kappa \Vert {\mathfrak {w}}_{\varrho , \ss } f _R(s) \Vert _{L^\infty _{x,v}}\big \} \\&\quad \lesssim \underbrace{ \exp \Big (\frac{3}{\kappa ^{{\mathfrak {P}}^\prime }}\Big ) + \sqrt{{\mathcal {F}}_{ p } (0)} + \sup _{0 \le s \le t} \big \{\sqrt{{\mathcal {E}}(s)} + \sqrt{{\mathcal {D}}(s)} \big \} {+ \frac{1}{\kappa ^{{\mathfrak {P}}}} \Vert P f_R \Vert _{L^2_{t,x}} } }_{(4.181)_*} . \end{aligned}\nonumber \\ \end{aligned}$$
(4.181)

From (4.72), (4.181), (4.178) and (2.10)

$$\begin{aligned} \begin{aligned}&\kappa ^{\frac{1}{2}} \Vert {P} f_R \Vert _{L^2_t((0,t);L^p_x )}\\&\quad \lesssim \underbrace{ (4.181)_* \Big \{1+ \frac{\varepsilon ^{1/2} \delta }{\kappa }(4.181)_* + \Big ( \varepsilon ^{\frac{p+2}{2(p-2)} }\kappa ^{\frac{2}{p-2}} (4.181)_* \Big )^{\frac{p-2}{p}} \Big \}}_{(4.182)_*}. \end{aligned} \end{aligned}$$
(4.182)

Using (4.181) and (2.9), from (4.74) and (4.143), we deduce that, for \(p<3\) and \(\varrho ^\prime <\varrho \),

$$\begin{aligned} \begin{aligned}&\kappa ^{ \frac{1}{2}+\ss } \big \Vert {P} \partial _t f_R \big \Vert _{L^2_t((0,t); L^p_x) } + (\varepsilon \kappa )^{3/p} \kappa ^{ \frac{1}{2}+\ss } \Vert {\mathfrak {w}} _{\varrho ^\prime ,\ss } \partial _t f_R \Vert _{L^2_t((0,t); L^{\infty } _{x,v})} \\&\lesssim \underbrace{ (4.182)_* \Big \{ 1 + \varepsilon ^{1- \frac{3-p}{p}} \delta \kappa ^{ - \frac{3}{p}} \{(4.181)_* +(4.182)_* \}\Big \}}_{(4.183)_*}. \end{aligned} \end{aligned}$$
(4.183)

Step 2. Using the estimates of the previous step we will close the estimate ultimately in the basic energy estimates (4.2) and (4.4) via the Gronwall’s inequality. We note that from (2.9) the multipliers of \(\int ^t_0 \Vert Pf_R(s) \Vert _{L^2_x}^2 \mathrm {d}s\) in (4.2) and \(\int ^t_0 \Vert P \partial _t f_R(s) \Vert _{L^2_x}^2 \mathrm {d}s\) in (4.4) are bounded above by

$$\begin{aligned} O(1) \kappa ^{-2{\mathfrak {P}}} \big ( 1+ \varepsilon \kappa ^{\frac{1}{2}-{\mathfrak {P}}} + ( \varepsilon \kappa ^{\frac{1}{2}-{\mathfrak {P}}})^2 \big ) \lesssim \kappa ^{-2{\mathfrak {P}}} , \end{aligned}$$
(4.184)

where we have used (2.11).

In (4.2) and (4.4) we bound

$$\begin{aligned} \begin{aligned} \Vert \kappa ^{1/2} Pf_R \Vert _{L^2_t L^3_x}&\lesssim \kappa ^{\frac{1}{2}(1- \frac{p}{3})} \Vert P f_R \Vert _{L^2_t L^\infty _x}^{1- \frac{p}{3}} \Vert \kappa ^{1/2} P f_R \Vert _{L^2_t L^p_x}^{\frac{p}{3}} \\&\lesssim _T (\varepsilon \kappa )^{-\frac{1}{2} (1- \frac{p}{3})} |(4.181)_*|^{1- \frac{p}{3}} |(4.183)_*|^{\frac{p}{3}} ,\\ \Vert P\partial _t f_R \Vert _{L^2_t L^3_x}&\lesssim \Vert P\partial _t f_R \Vert _{L^2_t L^\infty _x}^{1- \frac{p}{3}} \Vert P\partial _t f_R \Vert _{L^2_t L^p_x}^{\frac{p}{3}} \\&\lesssim \varepsilon ^{-\frac{3}{p}(1-\frac{p}{3}) } \kappa ^{ - \frac{1}{2} - {\mathfrak {P}} - \frac{3}{p} (1- \frac{p}{3}) } |(4.183)_*|. \end{aligned} \end{aligned}$$
(4.185)

We can check that the multiplier of \( \Vert \varepsilon ^{-1}\kappa ^{-1/2} \sqrt{\nu } ({\mathbf {I}} -{\mathbf {P}}) f_R \Vert _{L^2_{t,x,v}}^2\) in (4.4) is bounded as, from (2.10) and (4.181),

$$\begin{aligned} \begin{aligned}&\Big \{ \varepsilon (1+ \varepsilon \Vert (3.10)\Vert _{L^\infty _{t,x}}) \Vert \partial _t u \Vert _{L^\infty _{t,x}} + \varepsilon \kappa \Vert \nabla _x \partial _t u \Vert _{L^\infty _{t,x}} + (\varepsilon \kappa ^{1/2} \Vert (3.13)_*\Vert _{L^\infty _{t,x}})^2\\&\qquad +( \varepsilon \delta \Vert {\mathfrak {w}} f_R \Vert _{L^\infty _{t,x,v}}) ^2 \Big \}\\&\quad \lesssim \varepsilon \kappa ^{1/2-{\mathfrak {P}}}+ \varepsilon \delta ^2 \kappa ^{-2} |(4.181)_*|^2. \end{aligned} \end{aligned}$$

Applying (4.181), (4.182), (4.183), (4.185) to (4.2)+o(1)(4.4), using the above bound and (2.11), and collecting the terms, we derive that

$$\begin{aligned} \begin{aligned}&\sup _{0 \le s \le t}{\mathcal {E}} (s) + (1- \varepsilon \delta ^2 \kappa ^{-2} |(4.181)_*|^2) {\mathcal {D}}(t) \\&\quad \lesssim \ {\mathcal {E}}(0) + {\mathcal {F}} (0) + \exp \Big ( \frac{6}{\kappa ^{{\mathfrak {P}}^\prime }}\Big ) + (4.184) \int ^T_0 {\mathcal {E}} (s) \mathrm {d}s \\&\qquad + \delta ^2 \varepsilon ^{- (1- \frac{p}{3})}\kappa ^{- 4 + \frac{p}{3}} |(4.181)_*|^{4- \frac{2p}{3}} |(4.183)_*|^{\frac{2p}{3} }\\&\qquad + \delta ^2 \varepsilon ^{- \frac{6}{p} (1- \frac{p}{3}) } \kappa ^{-3 - 2{\mathfrak {P}} - \frac{6}{p} (1- \frac{p}{3})} |(4.181)_*|^{2 } |(4.183)_*|^{2}. \end{aligned} \end{aligned}$$
(4.186)

Under the assumption of

$$\begin{aligned} \begin{aligned} \varepsilon ^{1/2} \delta \kappa ^{-1}(4.181)_*&\ll 1 ,\\ \varepsilon ^{\frac{p+2}{2(p-2)} }\kappa ^{\frac{2}{p-2}} (4.181)_* \ll 1, \ \ \big [\varepsilon ^{1- \frac{3-p}{p}} \delta \kappa ^{ - \frac{3}{p} } \big ]^{1/2} (4.181)_*&\ll 1,\\ {[}\delta ^2 \varepsilon ^{- (1- \frac{p}{3})}\kappa ^{- 4 + \frac{p}{3}}]^{1/4} (4.181)_*&\ll 1, \ \\ [\delta ^2 \varepsilon ^{- \frac{6}{p} (1- \frac{p}{3}) } \kappa ^{-3 - 2{\mathfrak {P}} - \frac{6}{p} (1- \frac{p}{3})} ]^{1/4} (4.181)_*&\ll 1, \end{aligned} \end{aligned}$$
(4.187)

we derive that, for some constants \({\mathfrak {C}}_1>0\) and \({\mathfrak {C}}_2>0\),

$$\begin{aligned} \sup _{0 \le s\le t}{\mathcal {E}}(s) + {\mathcal {D}}(t) \le {\mathfrak {C}}_1 \Big ({\mathcal {E}}(0) + {\mathcal {F}}_p(0) +\exp \Big (\frac{6}{\kappa ^{{\mathfrak {P}}^\prime }}\Big ) \Big )+ {\mathfrak {C}}_2 \kappa ^{-{\mathfrak {P}}} \int ^t_0{\mathcal {E}}(s) \mathrm {d}s .\nonumber \\ \end{aligned}$$
(4.188)

Note that among others the last condition condition is the strongest in (4.187), which can be read as, from \(\delta =\sqrt{\varepsilon }\) of (2.11),

$$\begin{aligned} \delta ^{\frac{1}{2}- \frac{3}{p} \big (1- \frac{p}{3}\big ) } \kappa ^{- \frac{3}{4} - \frac{{\mathfrak {P}}}{2} - \frac{3}{2 p} (1- \frac{p}{3}) } (4.181)_* \ll 1. \end{aligned}$$
(4.189)

Applying the Gronwall’s inequality to (4.188) (we may redefine \({\mathcal {E}}(t)\) as \(\sup _{0 \le s \le t}{\mathcal {E}}(s)\) if necessary), we derive that

$$\begin{aligned} \begin{aligned} \sup _{0 \le s \le t}{\mathcal {E}} (s)&\le {\mathfrak {C}} _1\Big ({\mathcal {E}}(0) + {\mathcal {F}}_p(0)+ \exp \Big (\frac{6}{\kappa ^{{\mathfrak {P}}^\prime }}\Big )\Big ) \Big \{ 1+ \frac{{\mathfrak {C}}_2 t }{\kappa ^{{\mathfrak {P}}}} \exp \Big ( \frac{{\mathfrak {C}}_2 t }{\kappa ^{{\mathfrak {P}}}} \Big )\Big \} . \end{aligned} \end{aligned}$$

Applying this estimate to the last term of (4.188) and using the fact \({\mathfrak {P}}^\prime < {\mathfrak {P}}\) we derive that, after redefining \({\mathfrak {C}}_1\) if necessary,

$$\begin{aligned} \sup _{0 \le s \le t}{\mathcal {E}} (s) + {\mathcal {D}}(t) + {\mathcal {F}}_p (t) \le {\mathfrak {C}}_1 \big ( {\mathcal {E}}(0) + {\mathcal {F}}_p(0)+ 1 \big ) \exp \Big ( \frac{ 2{\mathfrak {C}}_2 t}{ \kappa ^{ {\mathfrak {P}}} }\Big ) \ \ \text {for all } \ t \le T_*, \nonumber \\ \end{aligned}$$
(4.190)

under the assumptions of (2.10), (4.178), and (4.187).

Step 3. Now we find out the ranges of \(\delta , \kappa , \varepsilon \) satisfying the assumptions of (4.178) and (4.187). From (2.12) and (4.190), if we choose \(\delta \) as

$$\begin{aligned} \delta \le \left[ \frac{ \kappa ^{ \frac{3}{2} + {\mathfrak {P}} + \frac{3}{ p} (1- \frac{p}{3}) } }{{\mathfrak {C}}_1 \big ( {\mathcal {E}}(0) + {\mathcal {F}}_p(0)+ 1 \big ) } \exp \Big ( \frac{ -2{\mathfrak {C}}_2 T}{ \kappa ^{ {\mathfrak {P}}} }\Big ) \right] ^{\frac{1}{1- \frac{6}{p} (1- \frac{p}{3})}}. \end{aligned}$$
(4.191)

then we can achieve (4.189) and hence all conditions of (4.187). Clearly (2.11) and (2.12) ensure (4.191).

Now from (4.190) and (2.11) we derive (2.13), which implies

$$\begin{aligned} \begin{aligned}&\sup _{0 \le s \le t} \Big \{ \Vert \kappa ^{1/2} Pf_R(s)\Vert _{L^6_x} + \Vert \varepsilon ^{1/2}\kappa {\mathfrak {w}}_{\varrho ,\ss } f_R(s) \Vert _{L_{x,v}^{\infty } } \\&\qquad + \Vert (\varepsilon \kappa )^{3/p} \kappa ^{ \frac{1}{2}+{\mathfrak {P}}} {\mathfrak {w}}_{\varrho ^\prime ,\ss } f_R(s) \Vert _{L^2((0,s);L_{x,v}^{\infty }) } \Big \}\\&\quad \lesssim \delta ^{-\frac{1}{2}+ \frac{3}{p} (1- \frac{p}{3})}. \end{aligned} \end{aligned}$$

These imply \(\min \{d_2, d_{2,t}, d_6, d_3, d_{3,t}, d_\infty , d_{\infty , t}\} \ge \frac{1}{4}\) and \(\frac{\delta \varepsilon ^{1/2}}{\kappa } \sqrt{{\mathcal {D}}(t)}\ll 1\) from (4.3), (4.5), (4.45), (4.73), (4.75), (4.141) and (4.144).

Then by the standard continuation argument we can verify all assumptions (4.178) up to \(t\le T\) and \(T=T_*\). The estimate (2.13) follows easily.

5 Navier-Stokes Approximations of the Euler Equations

In this section we prove Theorem 3. The proof of the theorem relies on the integral representation of the solution to the Navier-Stokes equations using the Green’s function for the Stokes problem in the same spirit of [47].

5.1 Elliptic Estimates and Nonlinear Estimates

In this section, we prove the estimates of the solutions of incompressible Navier-Stokes equations in large Reynolds numbers with the no slip boundary condition satisfying (1.13)-(1.15) based on recent Green’s function approach using the vorticity formulation of (2.16)-(2.18) applied to the inviscid limit problem [38, 44, 47, 54]. An advantage of working with analytic function spaces is the Cauchy estimates useful for recovery of the loss of derivatives. We recall the spaces, norms, and terminology we have defined in Section 2.

Lemma 7

([47, 54], Embeddings and Cauchy estimates) The following holds

  1. (1)

    \({\mathfrak {B}}^{\lambda , \kappa t} \subset {\mathfrak {L}}^{1,\lambda }\) and \({\mathfrak {B}}^{\lambda , \kappa } \subset {\mathfrak {L}}^{1,\lambda }\).

  2. (2)

    \(\Vert g_1 g_2\Vert _{*,\lambda }\lesssim \Vert g_1\Vert _{\infty ,\lambda } \Vert g_2\Vert _{*,\lambda }\).

  3. (3)

    \( \sum _{|\beta |=1} \Vert D^\beta g\Vert _{*,\lambda } \lesssim \frac{\Vert g\Vert _{*,{\tilde{\lambda }}}}{{{\tilde{\lambda }}} -\lambda }\), for any \(0< \lambda <{{\tilde{\lambda }}}\).

For (2) and (3), \(\Vert \cdot \Vert _{*,\lambda }\) can be either \(\Vert \cdot \Vert _{\infty ,\lambda ,\kappa }\) or \( \Vert \cdot \Vert _{\infty ,\lambda ,\kappa t}\) or \(\Vert \cdot \Vert _{\infty , \lambda , 0}\) or \( \Vert \cdot \Vert _{1,\lambda }\).

Lemma 8

([47, 54], Elliptic estimates) Let \(\phi \) be the solution of \(-\Delta \phi = \omega \) with the zero Dirichlet boundary condition, and let \(u=\nabla \times \phi \). Then

$$\begin{aligned} \begin{aligned} \Vert u \Vert _{\infty ,\lambda } + \Vert \nabla u \Vert _{1,\lambda }&\lesssim \Vert \omega \Vert _{1,\lambda } , \\ \Vert \nabla _h u \Vert _{\infty ,\lambda } + \Vert \nabla u_3\Vert _{\infty ,\lambda }&\lesssim \sum _{0 \le |\beta | \le 1} \Vert \nabla _h^\beta \omega \Vert _{1,\lambda } , \\ \Vert \partial _3 u _h\Vert _{\infty , \lambda }&\lesssim \sum _{0 \le |\beta |\le 1} \Vert \nabla _h^\beta \omega \Vert _{1,\lambda } + \Vert \omega _h \Vert _{\infty , \lambda },\\ \Vert \zeta ^{-1} \nabla _h^{\beta ^\prime } u_3 \Vert _{\infty , \lambda }&\lesssim \sum _{0 \le |\beta |\le 1} \Vert \nabla _h ^{\beta + \beta ^\prime } \omega _h \Vert _{1, \lambda }. \end{aligned} \end{aligned}$$
(5.1)

Proof

Here we only sketch the proofs. For full justification we refer to Proposition 2.3 in [47] for 2D and Section 4 of [54] for 3D and the proofs therein. From \((|\xi |^2 - \partial _z^2 )\phi _\xi = \omega _\xi \) and \(\phi _\xi (0)=0\) we write

$$\begin{aligned} \begin{aligned} \phi _\xi (z)=&\int ^z_0 G_- (y,z) \omega _\xi (y) \mathrm {d}y + \int ^\infty _z G_+ (y,z) \omega _\xi (y) \mathrm {d}y, \\&\text {with} \ \ G_{\pm } (y,z) := \frac{-1}{2 |\xi |} \Big ( e^{\pm |\xi | (z-y)} -e^{- |\xi | (y+z)}\Big ). \end{aligned} \end{aligned}$$
(5.2)

The first two estimates of (5.1) can be easily derived from this explicit form. For the third estimate of (5.1), we write \(u_1 = \partial _2 (-\Delta )^{-1}\omega _3 - \partial _3 (-\Delta )^{-1} \omega _2\) and \(\partial _3u_1 =\partial _3 \partial _2 (-\Delta )^{-1}\omega _3 -\partial _3 \partial _3 (-\Delta )^{-1} \omega _2\). Then the third estimate of (5.1) follows from the identity

$$\begin{aligned} \begin{aligned} \partial _z (\partial _3 (-\Delta )^{-1} \omega _2)_\xi =&\frac{1}{2} \bigg ( \int ^z_0 |\xi | e^{-|\xi | (z-y)} (1- e^{-2 |\xi | y}) \omega _{ \xi ,2} (s, y) \mathrm {d}y \\&+ \int ^\infty _z |\xi | e^{-|\xi | (y-z)} (1+ e^{-2 |\xi | z}) \omega _{ \xi ,2 } (x, y) \mathrm {d}y \\&+ \int ^\infty _z (-2 |\xi |) e^{- |\xi | (y-x)} e^{-2 |\xi |z } \omega _{ \xi ,2 } (s, y) \mathrm {d}y \bigg )- \omega _{\xi ,2}(z). \end{aligned} \end{aligned}$$

Next we prove the last estimate. Note that

$$\begin{aligned} \begin{aligned}&\frac{1+z}{z} \nabla _ h u_3 (z) \\&\quad = \frac{1 }{z} \int ^z_0 \partial _y \nabla _h u_3 (x_h, y) \mathrm {d}y + \nabla _h u_3 (z)\\&\quad = \frac{1 }{z} \int ^z_0 \nabla _h \big (\partial _1\partial _3 (-\Delta )^{-1} \omega _2 - \partial _2 \partial _3 (-\Delta )^{-1}\omega _1\big ) (x_h, y) \mathrm {d}y + \nabla _h u_3 (z). \end{aligned} \end{aligned}$$

From (5.2) we read that for \(i=1,2\)

$$\begin{aligned} \begin{aligned}&\Big ||\xi |^{|\beta |}(\partial _3 (-\Delta )^{-1} \omega _i)_\xi (s,z)\Big |\\&\quad \le \frac{1}{2} \Big ( \int ^z_0 e^{-|\xi | (z-y)} (1- e^{- 2|\xi | y})|\xi |^{|\beta |}| \omega _{\xi , i} (s,y)| \mathrm {d}y \\&\qquad + \int ^\infty _z e^{-|\xi | (y-z)} (1+ e^{- 2 |\xi | z}) |\xi |^{|\beta |}| \omega _{\xi , i} (s,y) |\mathrm {d}y \Big ) \\&\quad \lesssim \sup _{0 \le \sigma < \lambda } \big \Vert |\xi |^{|\beta |} \omega _{\xi ,h} \big \Vert _{L^1 (\partial {\mathcal {H}}_\sigma )}. \end{aligned} \end{aligned}$$

From the identity and estimate above we conclude the last bound of (5.1). \(\square \)

As a consequence of Lemma 8, we have the following nonlinear estimates.

Lemma 9

([47, 54]) Let u and \({{\tilde{u}}}\) be the velocity field associated with \(\omega = \nabla _x \times u\) and \({{\tilde{\omega }}}= \nabla _x \times {{\tilde{u}}}\) respectively. Then

$$\begin{aligned} \begin{aligned} \Vert u \cdot \nabla {{\tilde{\omega }}} \Vert _{1,\lambda }&\lesssim \Vert \omega \Vert _{1,\lambda } \Vert \nabla _h {{\tilde{\omega }}} \Vert _{1,\lambda } + \Vert (1+ |\nabla _h|) \omega \Vert _{1, \lambda } \Vert \zeta \partial _z {{\tilde{\omega }}} \Vert _{1,\lambda }, \\ \Vert \omega \cdot \nabla {{\tilde{u}}}_3\Vert _{1, \lambda }&\lesssim \Vert \omega _h \Vert _{1, \lambda } \Vert \nabla _h {\tilde{u}}_3 \Vert _{\infty , \lambda } + \Vert \omega _3 \Vert _{1,\lambda } \Vert \partial _3 {\tilde{u}}_3 \Vert _{\infty , \lambda }\\&\lesssim \Vert \omega \Vert _{1,\lambda } \Vert (1+ |\nabla _h|) {{\tilde{\omega }}} \Vert _{1, \lambda } , \\ \Vert \omega \cdot \nabla {{\tilde{u}}}_h\Vert _{1, \lambda }&\lesssim \Vert \omega _h \Vert _{1, \lambda } \Vert \nabla _h {\tilde{u}}_h \Vert _{\infty , \lambda } + \Vert \omega _3 \Vert _{\infty ,\lambda } \Vert \partial _3 {\tilde{u}}_h \Vert _{1, \lambda }\\&\lesssim \Vert \omega \Vert _{1,\lambda } \big ( \Vert {{\tilde{\omega }}}_3 \Vert _{\infty , \lambda } + \Vert (1+ |\nabla _h|) \omega \Vert _{1, \lambda } \big ) . \end{aligned} \end{aligned}$$
(5.3)

Moreover

$$\begin{aligned} \begin{aligned} \Vert u \cdot \nabla {{\tilde{\omega }}}_h \Vert _{*,\lambda }&\lesssim \Vert \omega \Vert _{1,\lambda } \Vert \nabla _h {{\tilde{\omega }}}_h\Vert _{*,\lambda } + \big ( \Vert (1+ |\nabla _h|) \omega \Vert _{1, \lambda } + \Vert \zeta \partial _z \omega _3 \Vert _{\infty , \lambda } \big ) \Vert \zeta \partial _z {{\tilde{\omega }}}_h\Vert _{*,\lambda }, \\ \Vert \omega \cdot \nabla {{\tilde{u}}}_h \Vert _{*,\lambda }&\lesssim \Vert \omega _3 \Vert _{\infty , \lambda , 0} \big ( \Vert (1+ |\nabla _h|){{\tilde{\omega }}} \Vert _{1, \lambda } + \Vert {{\tilde{\omega }}}_h \Vert _{*, \lambda } \big )\\&\quad + \Vert \omega _h \Vert _{*,\lambda } \sum _{0\le |\beta |\le 1} \Vert \nabla _h^\beta {{\tilde{\omega }}}\Vert _{1,\lambda } , \end{aligned} \end{aligned}$$
(5.4)

where \(\Vert \cdot \Vert _{*, \lambda }\) can be either \(\Vert \cdot \Vert _{\infty , \lambda , \kappa }\) or \(\Vert \cdot \Vert _{\infty , \lambda , \kappa t}\).

Furthermore

$$\begin{aligned} \begin{aligned} \Vert u \cdot \nabla {{\tilde{\omega }}}_3 \Vert _{\infty ,\lambda ,0}&\lesssim \Vert \omega \Vert _{1,\lambda } \Vert \nabla _h{{\tilde{\omega }}}_3\Vert _{\infty ,\lambda ,0} + \Vert (1+ |\nabla _h|) \omega \Vert _{1, \lambda } \Vert \zeta \partial _3 {{\tilde{\omega }}}_3\Vert _{\infty ,\lambda ,0 },\\ \Vert \omega \cdot \nabla {{\tilde{u}}}_3 \Vert _{\infty ,\lambda ,0}&\lesssim \Vert \omega _h \Vert _{*, \lambda } \Vert (1+ |\nabla _h|^2) {{\tilde{\omega }}}_h \Vert _{1, \lambda } + \Vert \omega _3 \Vert _{\infty , \lambda , 0} \Vert (1+ |\nabla _h|) {{\tilde{\omega }}}_h \Vert _{1, \lambda }, \end{aligned}\nonumber \\ \end{aligned}$$
(5.5)

where \(\Vert (1+ |\nabla _h|^k) g \Vert _* = \sum _{\ell =0}^k \Vert \nabla _h^\ell g \Vert _*\).

Proof

Again we refer to Proposition 2.3 in [47] for 2D and Section 4 of [54] for the full justification. The bounds (5.3) and (5.4) directly follow from Lemma 8. The proof of the first estimate of (5.5) is an outcome of applying (5.1) to an easy bound

$$\begin{aligned} \Vert u \cdot \nabla {{\tilde{\omega }}}_3 \Vert _{\infty ,\lambda ,0} \lesssim \Vert u_h \Vert _{\infty , \lambda } \Vert \nabla _h {\tilde{\omega }}_3 \Vert _{\infty , \lambda ,0} + \Vert \zeta (z) ^{-1} u_3 \Vert _{\infty , \lambda } \Vert \zeta (z) \partial _3 {\tilde{\omega }}_3 \Vert _{\infty , \lambda ,0}. \end{aligned}$$

For the second estimate of (5.5) it suffices to prove the bound for \(\omega _h \cdot \nabla _h {{\tilde{u}}}_3\). From \(|\zeta (z)(1+ \phi _\kappa (z) )|\lesssim 1\) or \(|\zeta (z)(1+ \phi _\kappa (z) + \phi _{\kappa t} (z))|\lesssim 1\),

$$\begin{aligned} \Vert \omega _h \cdot \nabla _h {{\tilde{u}}}_3 \Vert _{\infty , \lambda , 0}&\lesssim \Vert \omega _h \Vert _{*, \lambda } \Big \Vert \zeta (z)(1+ \phi _\kappa (z) + \phi _{\kappa t} (z) ) \frac{\nabla _h {{\tilde{u}}}_3}{\zeta (z)} \Big \Vert _{\infty , \lambda } \\&\lesssim \Vert \omega _h \Vert _{*, \lambda } \Vert \zeta ^{-1}{\nabla _h {{\tilde{u}}}_3} \Vert _{\infty , \lambda }. \end{aligned}$$

Then we use the last bound of (5.1) to finish the proof. \(\square \)

We finally record the crucial estimate of nonlinear forcing terms \(N=-u\cdot \nabla \omega + \omega \cdot \nabla u\), as an outcome of Lemma 9, that will be also crucially used to control \(B= [\partial _{x_3} (-\Delta )^{-1} (-u \cdot \nabla \omega + \omega \cdot \nabla u ) ] \, |_{x_3=0}\) in the vorticity formulation (2.16) and (2.18).

Lemma 10

([47, 54], Nonlinear estimate) Let \(\lambda \in (0,\lambda _0-\gamma s)\) be given. We have the following:

$$\begin{aligned}&\Vert (1+ |\nabla _h |) N \Vert _{1, \lambda }\lesssim \big ( \Vert (1+ |\nabla _h| ) \omega \Vert _{1, \lambda } + \Vert (1+ |\nabla _h| ) \omega _3\Vert _{\infty , \lambda , 0}\big ) \Vert (1+ |\nabla _h|^2 ) \omega \Vert _{1, \lambda } \nonumber \\&\qquad + \sum _{|\beta |=1} \Vert (1+ |\nabla _h|) D^\beta \omega \Vert _{1, \lambda } \Vert (1+ |\nabla _h|^2) \omega \Vert _{1, \lambda } , \end{aligned}$$
(5.6)
$$\begin{aligned}&\sum _{|\beta |=1}\Vert D^\beta (1+ |\nabla _h |) N \Vert _{1, \lambda }\nonumber \\&\quad \lesssim \sum _{|\beta | \le 1} \Vert D^\beta (1+ |\nabla _h| ) \omega \Vert _{1, \lambda } \bigg ( \sum _{|\beta | \le 2}\Vert D^\beta (1+ |\nabla _h| ) \omega \Vert _{1,\lambda }) + \Vert (1+ |\nabla _h|) \omega \Vert _{\infty , \lambda , 0} \bigg ) \nonumber \\&\qquad + \sum _{|\beta | \le 1} \Vert D^\beta (1+ |\nabla _h| ) \omega _3 \Vert _{\infty , \lambda , 0} \Vert (1+ |\nabla _h|)^2 \omega \Vert _{1,\lambda }. \end{aligned}$$
(5.7)

For \([[ \ \cdot \ ]]_{*,\lambda }\) to be either \([[ \ \cdot \ ]]_{\infty , \lambda , \kappa }\) or \([[ \ \cdot \ ]]_{\infty , \lambda , \kappa t}\),

$$\begin{aligned}&{[}[N]]_{*,\lambda } \lesssim \Vert (1+ |\nabla _h|^2) \omega \Vert _{1, \lambda } [[\omega ]]_{*,\lambda } + \Vert (1+ |\nabla _h| )\omega \Vert _{1, \lambda } [[ D \omega ]]_{*,\lambda }, \end{aligned}$$
(5.8)
$$\begin{aligned}&\sum _{|\beta |=1} [[ D^\beta N ]]_{*,\lambda } \lesssim \sum _{|\beta |=1} \Vert (1+ |\nabla _h|^{|\beta _h|+2}) \omega \Vert _{1, \lambda } [[ \omega ]]_{*,\lambda }\nonumber \\&\quad + \sum _{|\beta |=1} [[D^\beta \omega ]]_{*,\lambda } (\Vert (1+ |\nabla _h|^2) \omega \Vert _{1, \lambda } + \beta _3[[D_3^{\beta _3} \omega ]]_{*,\lambda }) \nonumber \\&\quad + \sum _{|\beta |=2} [[ D^\beta \omega ]]_{*,\lambda } \Vert (1+ |\nabla _h| )\omega \Vert _{1,\lambda } . \end{aligned}$$
(5.9)

The proof relies on Lemma 9. We refer to Lemma 4.2 and Lemma 4.5 in [54] for the detailed proof.

5.2 Green’s Function and Integral Representation for the Vorticity Formulation

By taking the Fourier transform of (2.16)-(2.18) in \(x_h \in {\mathbb {T}}^2\), we obtain

$$\begin{aligned} \partial _t \omega _\xi - \kappa \eta _0 \Delta _\xi \omega _\xi&= N_\xi \quad \text {in }{\mathbb {R}}_+, \end{aligned}$$
(5.10)
$$\begin{aligned} \kappa \eta _0 (\partial _{x_3} +|\xi |)\omega _{\xi ,h}&= B_\xi , \ \ \omega _{\xi ,3} =0 \quad \text {on } x_3=0 , \end{aligned}$$
(5.11)

with \(\omega _\xi |_{t=0} = {\omega _0}_\xi \) for \(\xi \in {{\mathbb {Z}}^2}\). Here

$$\begin{aligned} \Delta _\xi = - |\xi |^2 + \partial _{x_3}^2, \end{aligned}$$
(5.12)

and

$$\begin{aligned} N_\xi= & {} N_\xi (t,x_3):= \left( -u \cdot \nabla \omega + \omega \cdot \nabla u \right) _\xi (t,x_3), \quad \nonumber \\ B_\xi= & {} B_\xi (t):= (\partial _{x_3} (-\Delta _\xi )^{-1} N_{\xi ,h} (t))|_{x_3=0}. \end{aligned}$$
(5.13)

Here \((-\Delta _\xi )^{-1}\) denotes the inverse of \(-\Delta _\xi \) with the zero Dirichlet boundary condition at \(x_3=0\).

We give the integral representation and present key estimates on Green’s function for the Stokes problem. As shown in [47, 54], letting \(G_{\xi }(t,x_3,y)\) be the Green’s function for (5.10)-(5.11), the solution can be represented by the integral formula via Duhamel’s principle:

$$\begin{aligned} \begin{aligned} \omega _\xi (t,x_3) =&{\int ^\infty _0 G_{\xi } (t,x_3, y) \omega _{0\xi } (y) \mathrm {d}y} + {\int ^t_0\int ^\infty _0 G_{\xi } (t-s, x_3, y) N_\xi (s,y ) \mathrm {d}y \mathrm {d}s}\\&- {\int ^t_0 G_{\xi } (t-s, x_3, 0) (B_\xi (s) ,0)\mathrm {d}s} , \end{aligned} \end{aligned}$$
(5.14)

where

$$\begin{aligned} G_{\xi } = \begin{bmatrix} G_{\xi h} &{} 0 &{} 0\\ 0 &{} G_{\xi h} &{} 0 \\ 0&{} 0&{} G_{\xi 3} \end{bmatrix}, \end{aligned}$$
(5.15)

with \(G_{\xi h}\) of (5.19) and \(G_{\xi 3}\) of (5.22): for \(G_{\xi *}\) can be either \(G_{\xi h}\) or \(G_{\xi 3}\)

$$\begin{aligned} \partial _ t G_{\xi *} (t, x_3, y) - \kappa \eta _0 \Delta _\xi G_{\xi *} (t, x_3, y) =0, \ \ x_3>0, \end{aligned}$$
(5.16)
$$\begin{aligned} \kappa \eta _0 (\partial _{x_3} + |\xi |) G_{\xi h} (t, x_3, y)=0, \ \ x_3=0, \end{aligned}$$
(5.17)
$$\begin{aligned} G_{\xi 3} (t, x_3, y)=0, \ \ x_3=0. \end{aligned}$$
(5.18)

The following estimates and properties for \(G_{\xi }\) will be useful to show the propagation of analytic norms of \(\omega \), \(\partial _t\omega \) and \(\partial _t^2\omega \).

Lemma 11

([47, 54])

  1. (1)

    (Bounds on \(G_{\xi h}\)) The Green’s function \(G_{\xi h}\) for the Stokes problem (5.16) and (5.17) is given by

    $$\begin{aligned} G_{\xi h} = {\tilde{H}}_\xi + R_\xi , \end{aligned}$$
    (5.19)

    where \({\tilde{H}}_\xi \) is the one dimensional Heat kernel in the half-space with the homogeneous Neumann boundary condition which takes the form of

    $$\begin{aligned} {\tilde{H}}_\xi (t,x_3,y)= & {} H_\xi (t,x_3-y) + H_\xi (t,x_3+y) \nonumber \\= & {} \frac{1}{\sqrt{\kappa \eta _0 t}} \bigg ( e^{- \frac{|x_3-y|^2}{4 \kappa \eta _0 t}} + e^{- \frac{|x_3+y|^2}{4 \kappa \eta _0 t}} \bigg ) e^{- \kappa \eta _0 |\xi |^2 t}, \end{aligned}$$
    (5.20)

    and the residual kernel \(R_\xi \) due to the boundary condition satisfies

    $$\begin{aligned} |\partial _{x_3}^k R_\xi (t,x_3, y)| \lesssim b^{k+1} e^{- \theta _0 b (x_3 + y)} + \frac{1}{(\kappa \eta _0 t)^{(k+1)/2}} e^{- \theta _0 \frac{|x_3+y|^2}{\kappa \eta _0 t}} e^{- \frac{\kappa \eta _0 |\xi |^2 t}{8}} ,\nonumber \\ \end{aligned}$$
    (5.21)

    with \(b= |\xi | + \frac{1}{\sqrt{\kappa \eta _0}}\) and \(R_\xi (t,x_3,y) = R_\xi (t, x_3 + y)\).

  2. (2)

    (Formula of \(G_{\xi 3}\)) The Green’s function \(G_{\xi 3}\) for the Stokes problem (5.16) and (5.18) is given by one dimensional Heat kernel in the half-space with the homogeneous Dirichlet boundary condition as

    $$\begin{aligned} G_{\xi 3} (t,x_3,y)= & {} H_\xi (t,x_3-y) -H_\xi (t,x_3+y)\nonumber \\= & {} \frac{1}{\sqrt{\kappa \eta _0 t}} \bigg ( e^{- \frac{|x_3-y|^2}{4 \kappa \eta _0 t}} - e^{- \frac{|x_3+y|^2}{4 \kappa \eta _0 t}} \bigg ) e^{- \kappa \eta _0 |\xi |^2 t}. \end{aligned}$$
    (5.22)
  3. (3)

    (Complex extension) The Green’s function \(G_{\xi }\) has a natural extension to the complex domain \({\mathcal {H}}_\lambda \) for small \(\lambda >0\) with similar bounds in terms of \(\text {Re}\, y\) and \(\text {Re}\, z\) (cf. (3.16) in [47]). The solution \(\omega _\xi \) to (5.10)-(5.11) in \({\mathcal {H}}_\lambda \) has a similar representation: for any \(z\in {\mathcal {H}}_\lambda \), let \(\sigma \) be the positive constant so that \(z\in \partial {\mathcal {H}}_\lambda \), then \(\omega _\xi \) satisfies

    $$\begin{aligned} \begin{aligned} \omega _\xi (t,z) =&{\int _{\partial {\mathcal {H}}_\lambda } G_{\xi } (t,z, y) \omega _{0\xi } (y) \mathrm {d}y} + {\int ^t_0\int _{\partial {\mathcal {H}}_\lambda } G_{\xi } (t-s, z, y) N_\xi (s,y ) \mathrm {d}y \mathrm {d}s}\\&- {\int ^t_0 G_{\xi } (t-s, z, 0) (B_\xi (s), 0) \mathrm {d}s} . \end{aligned} \end{aligned}$$

The proof of Lemma 11 can be found in Proposition 3.3 and Section 3.3 of [47]. The next lemma concerns the convolution estimates.

Lemma 12

Let \(T>0\) be given. Recall the norms defined in Section 2. For any \(0\le s < t\le T\) and \(k\ge 0\), there exists a constant \(C_T>0\) so that the following estimates hold: for \(G_{\xi *}\) can be either \(G_{\xi h}\) or \(G_{\xi 3}\)

  1. (1)

    (\({\mathcal {L}}^1_\lambda \) estimates)

    $$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty G_{\xi *}(t, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^1_\lambda } \le C_T \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^1_\lambda }, \end{aligned}$$
    (5.23)
    $$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty G_{\xi *}(t-s,z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^1_\lambda } \le C_T \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^1_\lambda }. \end{aligned}$$
    (5.24)
  2. (2)

    (\({\mathcal {L}}^\infty _{\lambda ,\kappa t}\) estimates)

    $$\begin{aligned}&\sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty G_{\xi *} (t, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa t}} \nonumber \\&\quad \le C_T \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}, \end{aligned}$$
    (5.25)
    $$\begin{aligned}&\sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty G_{\xi *}(t-s, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa t}} \nonumber \\&\quad \le C_T \sum _{j=0}^k \sqrt{\frac{t}{s}} \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa s}}. \end{aligned}$$
    (5.26)
  3. (3)

    (\({\mathcal {L}}^\infty _{\lambda ,\kappa }\) estimates) For either \(\kappa =0\) or \(\kappa >0\)

    $$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty G_{\xi *}(t,z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \le C_T \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}, \end{aligned}$$
    (5.27)
    $$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty G_{\xi *}(t-s, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \le C_T \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}. \end{aligned}$$
    (5.28)

Proof

We only give a proof for \(G_{\xi h}\) since \(G_{\xi 3}\) can be handled easier than the other. The proof of (1) and (2) can be found in Propositions 3.7 and 3.8 of [47]. Here we present the detail for (3), the second inequality. We consider real values \(y,z\in {\mathbb {R}}_+\) only as the complex extension follows similarly (cf. (3) in Lemma 11). Note that in view of (5.19), (5.20) and (5.21), it suffices to show

$$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty R (t-s, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \le C_T \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} , \end{aligned}$$
(5.29)
$$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j \int _0^\infty H (t-s, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \le C_T \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}, \end{aligned}$$
(5.30)

where \(R(t,z,y)= b e^{- b (y+z)}\) and \(H(t,z,y) = \frac{1}{\sqrt{\kappa t}} e^{-\frac{|y-z|^2}{M\kappa t}}\) for some \(M>0\). We start with (5.29). Let \(k=0\) first. First note that

$$\begin{aligned} \left| \int _0^\infty R (t-s, z, y ) g_\xi (y) \mathrm {d}y \right|&= \left| e^{-bz}\int _0^\infty b e^{-({{\bar{\alpha }}}+b)y} (1+\phi _\kappa (y)) \frac{e^{{{\bar{\alpha }}} y}}{1+\phi _\kappa (y)} g_\xi (y) \mathrm {d}y\right| \\&\le e^{-bz} (1+\phi _\kappa (0)) \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \int _0^\infty b e^{-({{\bar{\alpha }}}+b)y} \mathrm {d}y , \end{aligned}$$

since \(\phi _\kappa \) is a decreasing function. The last integral is uniformly finite for all \(|\xi |\) and \(\kappa \). Hence,

$$\begin{aligned} \left\| \int _0^\infty R (t-s, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}&\lesssim \sup _{z}\left( \frac{1+\phi _\kappa (0)}{1+\phi _\kappa (z)}e^{({{\bar{\alpha }}}-b) z} \right) \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} . \end{aligned}$$

For \({{\bar{\alpha }}}>0\), if \(\kappa < \frac{1}{4 \eta _0 {{\bar{\alpha }}}^2}\) then

$$\begin{aligned}&\sup _{z}\left( \frac{1+\phi _\kappa (0)}{1+\phi _\kappa (z)}e^{({{\bar{\alpha }}}-b) z} \right) \nonumber \\&\quad \lesssim \frac{ \sqrt{\kappa }+1 }{ \inf _{z}\left[ (\sqrt{\kappa } + \frac{1}{1+|\frac{z}{\sqrt{\kappa }}|^p} )e^{\frac{z}{2\sqrt{ \eta _0\kappa }}} \right] } < \min \Big \{1,\frac{\sqrt{\kappa } +1}{\sqrt{\kappa } + \frac{1}{ (2 \sqrt{\eta _0})^pp!} }\Big \}, \end{aligned}$$
(5.31)

where the last bound follows from the fact that

$$\begin{aligned} \frac{1}{1+|\frac{z}{\sqrt{\kappa }}|^p} e^{\frac{z}{2\sqrt{\eta _0\kappa }}} \ge \frac{1}{1+|\frac{z}{\sqrt{\kappa }}|^p} \{1+ \frac{1}{p!} | \frac{z}{2\sqrt{\eta _0\kappa }}|^p\} \ge \min \big \{1, \frac{1}{ (2 \sqrt{\eta _0})^pp!}\big \} . \end{aligned}$$

For \(k\ge 1\), since \(|\zeta (z)\partial _z R |\lesssim be^{-\frac{b z}{2}}\), the derivative estimates follow analogously. Therefore, (5.29) holds true.

We move onto (5.30). Let \(k=0\) first. Note that, for \(0 \le s < t \le T\) and \(\kappa \lesssim 1\)

$$\begin{aligned}&e^{-\frac{|y-z|^2}{2M\kappa (t-s)}} e^{-{{\bar{\alpha }}} y} = e^{-\frac{1}{2}|\frac{y-z}{\sqrt{M\kappa (t-s)}} + {{\bar{\alpha }}} \sqrt{M\kappa (t-s)} |^2}\nonumber \\&e^{\frac{M}{2}{{\bar{\alpha }}}^2\kappa (t-s)} e^{-{{\bar{\alpha }}} z} \le e^{\frac{M}{2}{{\bar{\alpha }}}^2\kappa (t-s)} e^{-{{\bar{\alpha }}} z} \lesssim e^{-{{\bar{\alpha }}} z}, \end{aligned}$$
(5.32)

and thus

$$\begin{aligned}&\left| \int _0^\infty H (t-s, z, y ) g_\xi (y) \mathrm {d}y \right|&\\&\quad = \left| \int _0^\infty \frac{1}{\sqrt{\kappa (t-s)}} e^{-\frac{|y-z|^2}{M\kappa (t-s)}} e^{-{{\bar{\alpha }}} y}(1+\phi _\kappa (y)) \frac{e^{{{\bar{\alpha }}} y}}{1+\phi _\kappa (y)} g_\xi (y) \mathrm {d}y\right| \\&\lesssim e^{-{{\bar{\alpha }}} z} \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \int _0^\infty \frac{1}{\sqrt{\kappa (t-s)}} e^{-\frac{|y-z|^2}{2M\kappa (t-s)}} (1+\phi _\kappa (y)) \mathrm {d}y. \end{aligned}$$

For the last integral, we divide the integral into two: \(\int _0^\infty = \int _0^{\frac{z}{2}} + \int _{\frac{z}{2}}^\infty \). For the latter, since \(\phi _\kappa \) is decreasing and the kernel is in \(L^1_y\), we deduce

$$\begin{aligned} \int _{\frac{z}{2}}^\infty \frac{1}{\sqrt{\kappa (t-s)}} e^{-\frac{|y-z|^2}{2M\kappa (t-s)}} (1+\phi _\kappa (y)) \mathrm {d}y \lesssim 1+\phi _\kappa ({z}) . \end{aligned}$$

For \( \int _0^{\frac{z}{2}} \mathrm {d}y\), note \(|y-z|\ge \frac{z}{2}\) and \(1+\phi _\kappa (y) \le 1+\phi _\kappa (0)\) for \(y\in (0,\frac{z}{2})\). Hence

$$\begin{aligned} \int ^{\frac{z}{2}}_0\frac{1}{\sqrt{\kappa (t-s)}} e^{-\frac{|y-z|^2}{2M\kappa (t-s)}} (1+\phi _\kappa (y)) \mathrm {d}y \lesssim e^{-\frac{|z|^2}{16M\kappa (t-s)}} (1+\phi _\kappa (0)) . \end{aligned}$$

Then

$$\begin{aligned} \left\| \int _0^\infty H (t-s, z, y ) g_\xi (y) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}&\lesssim \sup _{z}\left( \frac{1+\phi _\kappa (0)}{1+\phi _\kappa (z)}e^{-\frac{|z|^2}{16M\kappa (t-s)}}+1 \right) \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} . \end{aligned}$$

A similar argument as in (5.31) shows that \(\frac{1+\phi _\kappa (0)}{1+\phi _\kappa (z)}e^{-\frac{|z|^2}{16M\kappa (t-s)}}\) is uniformly finite in \(\kappa \). This shows (5.30) for \(k=0\).

For the derivative estimate, by splitting the integral into two parts and using \(\partial _z H (t,z,y)= - \partial _y H(t,z,y)\), we rewrite

$$\begin{aligned}&\int _0^\infty \zeta (z)\partial _z H (t-s, z, y ) g_\xi (y) \mathrm {d}y \\&\quad = \int _0^\frac{z}{2} \zeta (z)\partial _z H (t-s, z, y ) g_\xi (y) \mathrm {d}y - \zeta (z) H(t-s, z, \frac{z}{2}) g_\xi (\frac{z}{2}) \\&\qquad + \int _\frac{z}{2}^\infty \zeta (z) H (t-s, z, y ) \partial _y g_\xi (y) \mathrm {d}y . \end{aligned}$$

For the first integral, since \( |y-z|\ge \frac{z}{2}\) for \(y\in (0,\frac{z}{2})\),

$$\begin{aligned} | \zeta (z)\partial _z H (t-s, z, y ) |&\lesssim \frac{z}{1+z}\frac{1}{\kappa (t-s)} e^{-\frac{|y-z|^2}{2M\kappa (t-s)}}\lesssim |y-z| \frac{1}{\kappa (t-s)} e^{-\frac{|y-z|^2}{2M\kappa (t-s)}}\\&\lesssim \frac{1}{\sqrt{\kappa (t-s)}} e^{-\frac{|y-z|^2}{4M\kappa (t-s)}}. \end{aligned}$$

Hence, by the same argument as in \(k=0\) leads to the desired bound. For the second term,

$$\begin{aligned} |\zeta (z) H(t-s, z, \frac{z}{2}) g_\xi (\frac{z}{2}) |&\lesssim \frac{z}{\sqrt{\kappa (t-s)}} e^{-\frac{|z|^2}{4M\kappa (t-s)}} e^{-{{\bar{\alpha }}}\frac{z}{2}}(1+\phi _\kappa (\frac{z}{2})) \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \\&\lesssim e^{-\frac{|z|^2}{8M\kappa (t-s)}} e^{-{{\bar{\alpha }}}\frac{z}{2}}(1+\phi _\kappa ({z})) \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \\&= e^{-\frac{1}{2}| \frac{z}{2\sqrt{M\kappa (t-s)}} -{{\bar{\alpha }}} \sqrt{M\kappa (t-s)} |^2} e^{\frac{M}{2}{{\bar{\alpha }}}^2\kappa (t-s)} e^{-{{\bar{\alpha }}} z}\\&\quad (1+\phi _\kappa ({z})) \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \\&\lesssim e^{-{{\bar{\alpha }}} z}(1+\phi _\kappa ({z})) \left\| g_\xi \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}, \end{aligned}$$

which leads to the desired bound. For the last integral, note that \(\zeta (z) \le 2 \zeta (y)\) for \(y\ge \frac{z}{2}\). Therefore the corresponding integral can be treated in the same way as in \(k=0\) with \(g_\xi (y)\) replaced by \(\zeta (y)\partial _yg_\xi (y)\). This shows (5.30) for \(k=1\). Other \(k\ge 2\) can be estimated analogously. \(\square \)

The next result concerns the estimates for the trace kernel.

Lemma 13

Let \(a_\xi (s)= [\partial _{z}(- \Delta _\xi )^{-1} g_\xi ] \, |_{z=0}\). Then for any \(0\le s<t\le T\) and \(k\ge 0\), we have the following

$$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j G_{\xi h }(t-s,z, 0 ) a_\xi (s) \right\| _{{\mathcal {L}}^1_\lambda }&\lesssim \left\| g_\xi \right\| _{{\mathcal {L}}^1_\lambda }, \end{aligned}$$
(5.33)
$$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j G_{\xi h}(t-s,z, 0 ) a_\xi (s) \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}&\lesssim \frac{1}{\sqrt{t-s}} \left\| g_\xi \right\| _{{\mathcal {L}}^1_\lambda }. \end{aligned}$$
(5.34)

Proof

Note that from (5.19), (5.20) and (5.21), the conormal derivatives \((\zeta (z)\partial _z)^j\) of \(G_{\xi h}(t-s,z,0)\) enjoy the same bounds as \(G_{\xi h}(t-s,z,0)\): for some small constant \(c_0\),

$$\begin{aligned} | (\zeta (z)\partial _{z})^j G_{\xi }(t-s,z, 0 ) | \lesssim b e^{-c_0 b z} + \frac{1}{\sqrt{\kappa (t-s)}} e^{-c_0 \frac{|z|^2}{\kappa (t-s)}}. \end{aligned}$$
(5.35)

Therefore, it suffices to show the bounds for \(k=0\). We first recall the representation formula for \(a_\xi \) (cf. (4.29) of [38] or (4.2) of [54]):

$$\begin{aligned} a_\xi (s) = \int _0^\infty e^{-|\xi | y} g_\xi (y) \mathrm {d}y , \end{aligned}$$

from which we have \(\Vert a_\xi \Vert _{{\mathcal {L}}^\infty _\lambda } \lesssim \left\| g_\xi \right\| _{{\mathcal {L}}^1_\lambda }\). Since the above upper bound of \(G(t-s,z,0)\) is integrable in z, (5.33) follows. To show (5.34), we compute \(\Vert G_{\xi h}(t-s,z, 0 ) \Vert _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}\):

$$\begin{aligned} \Vert G_{\xi h}(t-s,z, 0 ) \Vert _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \lesssim \sup _z \left[ \frac{ be^{({{\bar{\alpha }}}-c_0 b) z}}{1+\phi _\kappa (z)}\right] + \frac{1}{\sqrt{t-s}} \sup _z\left[ \frac{e^{ {{\bar{\alpha }}} z-c_0 \frac{|z|^2}{\kappa (t-s)}}}{\sqrt{\kappa }+\sqrt{\kappa }\phi _\kappa (z)} \right] . \end{aligned}$$

It is a routine to check that both supremum norms are uniformly bounded in \(\kappa \) and \(|\xi |\). Therefore (5.34) is obtained. \(\square \)

5.3 Proof of Theorem 3

Our goal is to show that \(\omega (t)\) indeed belongs to \(C^1([0,T];{\mathfrak {B}}^{\lambda , \kappa })\) without the initial layer under the compatibility condition (2.34), and that \(\partial _t^2\omega \) in \({\mathfrak {B}}^{\lambda , \kappa t}\) with the initial layer. The existence of \(\omega (t)\) in \(C^1([0,T];{\mathfrak {B}}^{\lambda , \kappa t})\) under the assumption of Theorem 3 can be proved by following the argument of [47] and [54]. For the 2D case, Theorem 1.1 of [47] indeed ensures the existence of \(\omega (t)\) in \(C^1([0,T];{\mathfrak {B}}^{\lambda , \kappa t})\) under the assumption of Theorem 3. Such a result follows from Lemma 7, Lemma 11, Lemma 12, Lemma 8, Lemma 9. A 3D result can be obtained analogously. Hence, it suffices to show the propagation of the analytic norms in (2.35).

Step 1: Propagation of analytic norms for \(\omega \). It is convenient to define

$$\begin{aligned} {\left| \left| \left| \omega (t) \right| \right| \right| }_t:= {\left| \left| \left| \omega (t) \right| \right| \right| }_{\infty , \kappa } + {\left| \left| \left| \omega (t) \right| \right| \right| }_{1}. \end{aligned}$$
(5.36)

The estimation of \(\omega \) follows from the nonlinear iteration using the representation formula (5.14).

The estimates for the \(L^1\)-based norm \({\left| \left| \left| \omega (t) \right| \right| \right| }_1\) are already available in Section 5 of [54] (for 2D see Section 4.1 of [47]): From (5.23), (5.24) and (5.33), we have that for \(k=0,1,2\)

$$\begin{aligned}&\sum _{j=0}^k \Vert (\zeta (x_3)\partial _{x_3})^j \omega _{\xi } \Vert _{{\mathcal {L}}^1_{\lambda }} \\&\quad \le \sum _{j=0}^k \left\| (\zeta (x_3)\partial _{x_3})^j{\int ^\infty _0 G_{\xi } (t,x_3, y) \omega _{0\xi } (y) \mathrm {d}y} \right\| _{{\mathcal {L}}^1_{\lambda }} \\&\qquad + \sum _{j=0}^k {\int ^t_0 \left\| (\zeta (x_3)\partial _{x_3})^j\int ^\infty _0 G_{\xi } (t-s, x_3, y) N_{\xi } (s,y ) \mathrm {d}y \right\| _{{\mathcal {L}}^1_{\lambda }} \mathrm {d}s}\\&\qquad +\sum _{j=0}^k {\int ^t_0 \left\| (\zeta (x_3)\partial _{x_3})^j G_{\xi } (t-s, x_3, 0) (B_{\xi } (s),0) \right\| _{{\mathcal {L}}^1_{\lambda }}\mathrm {d}s} \\&\quad \lesssim \sum _{j=0}^k \left\| (\zeta (x_3)\partial _{x_3})^j \omega _{0\xi } \right\| _{{\mathcal {L}}^1_{\lambda }} + \sum _{j=0}^k \int _0^t \left\| (\zeta (x_3)\partial _{x_3})^j N_{\xi } (s) \right\| _{{\mathcal {L}}^1_{\lambda }} \mathrm {d}s + \int _0^t \Vert N_{\xi }(s) \Vert _{{\mathcal {L}}^1_\lambda } \mathrm {d}s. \end{aligned}$$

For \(k=1\), after summing up over \(\xi \in {\mathbb {Z}}^2\), we deduce that

$$\begin{aligned} \begin{aligned} \sum _{0 \le |\beta | \le 1} \Vert D^\beta (1+ |\nabla _h|) \omega (s)\Vert _{1,\lambda } \lesssim&\sum _{0 \le |\beta | \le 1} \Vert D^\beta (1+ |\nabla _h|) \omega _0\Vert _{1,\lambda }\\&+ \int ^t_0 \sum _{0 \le |\beta | \le 1} \Vert D^\beta (1+ |\nabla _h|) N(s) \Vert _{1,\lambda } \mathrm {d}s. \end{aligned} \end{aligned}$$
(5.37)

Using (5.6), (5.7), and the definition of \({\left| \left| \left| \ \cdot \ \right| \right| \right| }_s\) in (5.36) we derive that

$$\begin{aligned} \begin{aligned} \int ^t_0 \sum _{0 \le |\beta | \le 1} \Vert D^\beta (1+ |\nabla _h|) N(s) \Vert _{1,\lambda } \mathrm {d}s&\lesssim \int ^t_0 {\left| \left| \left| \omega (s) \right| \right| \right| }_s^2 \big [1+(\lambda _0-\lambda -\gamma _0 s)^{- \alpha }\big ] \mathrm {d}s\\&\lesssim \Big (t+ \frac{1}{\gamma _0}\Big ) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_s . \end{aligned} \end{aligned}$$
(5.38)

The second order derivatives can be treated similarly except for the contributions of N for which we apply the analyticity recovery estimate using (3) of Lemma 7 while other terms are estimated in the same way. More precisely, we have

$$\begin{aligned}&\sum _{|\beta |=2}\Vert D^\beta (1+ |\nabla _h|) N (s) \Vert _{1,\lambda } \nonumber \\&\quad \lesssim \frac{1}{{{\tilde{\lambda }}}-\lambda } \sum _{0 \le |\beta |\le 1}\Vert D^\beta (1+ |\nabla _h|) N (s) \Vert _{1,{{\tilde{\lambda }}}} \ \text {for any} \ {{\tilde{\lambda }}} >\lambda , \end{aligned}$$
(5.39)

while we choose \({{\tilde{\lambda }}}=\frac{\lambda +\lambda _0-\gamma _0 s}{2}\) in particular. We note that still \({\tilde{\lambda }}< \lambda _0 - \gamma _0 s\) if \(\lambda < \lambda _0 -\gamma _0 s\) and hence from (5.6) and (5.7)

$$\begin{aligned} \begin{aligned}&\sum _{0 \le |\beta | \le 1} \Vert D^\beta (1+ |\nabla _h|) N (s) \Vert _{1, {\tilde{\lambda }}(s) }\\&\quad \lesssim \Big ( \sum _{0 \le |\beta |\le 1} \Vert D^\beta (1+ |\nabla _h|) \omega (s) \Vert _{1,{{\tilde{\lambda }}}(s) }\Big ) \Big (\sum _{0 \le |\beta |\le 2}[[ D^\beta \omega (s) ]]_{\infty , {\tilde{\lambda }}(s), \kappa } \Big )\\&\qquad +\Big ( \sum _{0 \le |\beta |\le 2} \Vert D^\beta (1+ |\nabla _h|) \omega (s) \Vert _{1,{{\tilde{\lambda }}}(s) }\Big ) \Big (\sum _{0 \le |\beta |\le 1}\Vert D^\beta (1+ |\nabla _h|) \omega (s)\Vert _{1,{{\tilde{\lambda }}}(s) }\Big )\\&\quad \lesssim \big [1+ (\lambda _0 -\lambda -\gamma _0 s)^{-\alpha } \big ] {\left| \left| \left| \omega (s) \right| \right| \right| }_s ^2. \end{aligned} \end{aligned}$$

Therefore we derive that for \(t< \frac{\lambda _0}{2\gamma _0}\) and \(\lambda < \lambda _0 - \gamma _0 t\)

$$\begin{aligned} \begin{aligned}&\sum _{|\beta |=2} \Vert D^\beta (1+ |\nabla _h|) \omega (t) \Vert _{1, \lambda } \\&\quad \lesssim \sum _{|\beta |=2} \Vert D^\beta (1+ |\nabla _h|) \omega _{0} \Vert _{1, \lambda _0 } + \int _0^t \big [1+ (\lambda _0-\lambda -\gamma _0 s)^{ -(\alpha + 1)} \big ] {\left| \left| \left| \omega (s) \right| \right| \right| }_{s }^2 \mathrm {d}s \\&\quad \lesssim \sum _{|\beta |=2} \Vert D^\beta (1+ |\nabla _h|) \omega _{0} \Vert _{1, \lambda _0 } + \Big ( (\lambda _0 - \lambda - \gamma _0 t)^{-\alpha } \frac{1}{\gamma _0} +{t} \Big ) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_s^2. \end{aligned}\nonumber \\ \end{aligned}$$
(5.40)

Therefore, we conclude that, from (5.37) with (5.38), and (5.40)

$$\begin{aligned} {\left| \left| \left| \omega (t) \right| \right| \right| }_1\lesssim \sum _{0 \le |\beta |\le 2} \Vert D^\beta (1+ |\nabla _h|) \omega _{0} \Vert _{1, \lambda _0 } + (t+ \frac{1}{\gamma _0}) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_s^2 \ \ \text {for} \ t< \frac{\lambda _0}{2 \gamma _0}.\nonumber \\ \end{aligned}$$
(5.41)

The propagation of the boundary layer norm \({\left| \left| \left| \omega (t) \right| \right| \right| }_{\infty ,\kappa }\) can be shown analogously using \({\mathcal {L}}^\infty _{\lambda , \kappa }\) estimates of Lemma 12 and Lemma 13: For \(k=0,1,2\) and \(\kappa >0\) for \(i=1,2\) and \(\kappa =0\) for \(i=3\) we have

$$\begin{aligned}&\sum _{j=0}^k \Vert (\zeta (x_3)\partial _{x_3})^j \omega _{\xi , i} \Vert _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \\&\quad \le \sum _{j=0}^k \left\| (\zeta (x_3)\partial _{x_3})^j{\int ^\infty _0 G_{\xi i} (t,x_3, y) \omega _{0\xi ,i} (y) \mathrm {d}y} \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \\&\qquad + \sum _{j=0}^k {\int ^t_0 \left\| (\zeta (x_3)\partial _{x_3})^j\int ^\infty _0 G_{\xi ,i} (t-s, x_3, y) N_{\xi , i} (s,y ) \mathrm {d}y \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \mathrm {d}s}\\&\qquad +\sum _{j=0}^k {\int ^t_0 \left\| (\zeta (x_3)\partial _{x_3})^j G_{\xi , i} (t-s, x_3, 0) B_{\xi ,i} (s) \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }}\mathrm {d}s} \\&\quad \lesssim \sum _{j=0}^k \left\| (\zeta (x_3)\partial _{x_3})^j \omega _{0\xi , i } \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} + \sum _{j=0}^k \int _0^t \left\| (\zeta (x_3)\partial _{x_3})^j N_{\xi ,i} (s) \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa }} \mathrm {d}s\\&\qquad +(1- \delta _{i3}) \int _0^t \frac{1}{\sqrt{t-s}} \Vert N_{\xi , i} \Vert _{{\mathcal {L}}^1_\lambda }. \end{aligned}$$

Let \(k=1\). After summing up over \(\xi \in {\mathbb {Z}}\) and \(i=1,2\) (with \(\kappa >0\)) and \(i=3\) (with \(\kappa =0\)) , we deduce that

$$\begin{aligned} \sum _{0\le |\beta |\le 1} [[D^\beta \omega (t) ]]_{\infty , \lambda , \kappa }\lesssim & {} \sum _{0\le |\beta |\le 1} [[ D^\beta \omega _{0} ]]_{\infty , \lambda _0, \kappa } + \int ^t_0 \sum _{0\le |\beta |\le 1} [[ D^\beta N(s) ]]_{\infty , \lambda ,\kappa } \mathrm {d}s \\&+ \int _0^t \frac{1}{\sqrt{t-s}} {\left| \left| \left| \omega (s) \right| \right| \right| }_1^2 \mathrm {d}s. \end{aligned}$$

Using the definition of \({\left| \left| \left| \ \cdot \ \right| \right| \right| }_s\) in (5.36), and applying Lemma 10 with (5.6), (5.8), and (5.9), we derive

$$\begin{aligned} \begin{aligned} \sum _{0\le |\beta |\le 1} [[ D^\beta N(s) ]]_{\infty , \lambda ,\kappa }&\lesssim \Big ( \sum _{0 \le |\beta | \le 2} \Vert D^\beta (1+ |\nabla _h| )\omega (s) \Vert _{1,\lambda } \Big )\\&\quad \Big (\sum _{0 \le |\beta | \le 1} [[ D^\beta \omega (s) \Vert _{\infty , \lambda , \kappa }\Big ) \\&\quad + \Vert (1+ |\nabla _h| )\omega (s) \Vert _{1,\lambda } \sum _{|\beta |=2}[[ D^\beta \omega (s) ]]_{\infty ,\lambda , \kappa }\\&\lesssim \big [1+(\lambda _0 -\lambda -\gamma _0 s)^{- \alpha } \big ] {\left| \left| \left| \omega (s) \right| \right| \right| }_s ^2. \end{aligned} \end{aligned}$$

Therefore we derive that

$$\begin{aligned} \begin{aligned}&\sum _{0\le |\beta |\le 1} [[D^\beta \omega (t) ]]_{\infty , \lambda , \kappa } \\&\quad \lesssim \sum _{0\le |\beta |\le 1} [[ D^\beta \omega _{0} ]]_{\infty , \lambda _0, \kappa } + \int _0^t {\left| \left| \left| \omega (s) \right| \right| \right| }_{s }^2 \big [1+ (\lambda _0-\lambda -\gamma _0 s)^{ - \alpha } \big ] \mathrm {d}s \\&\qquad + \int _0^t \frac{1}{\sqrt{t-s}} {\left| \left| \left| \omega (s) \right| \right| \right| }_s^2 ds \\&\lesssim \sum _{0\le |\beta |\le 1} [[ D^\beta \omega _{0} ]]_{\infty , \lambda _0, \kappa } + \Big ( \sqrt{t} + \frac{1}{\gamma _0}\Big ) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_s^2. \end{aligned} \end{aligned}$$
(5.42)

Now we control the second order derivatives similarly except for the N. As in (5.39) we use the analyticity recovery estimate using Lemma 7

$$\begin{aligned} \sum _{|\beta |=2}[[ D^\beta N (s) ]]_{\infty ,\lambda ,\kappa } \lesssim \frac{1}{{{\tilde{\lambda }}}-\lambda } \sum _{0 \le |\beta |\le 1}[[ D^\beta N (s) ]]_{\infty ,{{\tilde{\lambda }}},\kappa } \ \text {for any} \ {{\tilde{\lambda }}} >\lambda , \end{aligned}$$
(5.43)

while again we choose \({{\tilde{\lambda }}}=\frac{\lambda +\lambda _0-\gamma _0 s}{2}\) in particular. We note that still \({\tilde{\lambda }}< \lambda _0 - \gamma _0 s\) if \(\lambda < \lambda _0 -\gamma _0 s\) and hence from (5.8) and (5.9)

$$\begin{aligned} \sum _{0 \le |\beta | \le 1} [[ D^\beta N (s) ]]_{\infty , {\tilde{\lambda }}(s), \kappa }\lesssim (\lambda _0 -\lambda -\gamma _0 s)^{- \alpha } {\left| \left| \left| \omega (s) \right| \right| \right| }_s ^2. \end{aligned}$$

Therefore we derive that for \(t< \frac{\lambda _0}{2\gamma _0}\) and \(\lambda < \lambda _0 - \gamma _0 t\)

$$\begin{aligned} \begin{aligned}&\sum _{ |\beta |=2}[[ D^\beta \omega (t) ]]_{\infty , \lambda , \kappa }\\&\quad \lesssim \sum _{ |\beta |=2} \Vert D^\beta \omega _{0} \Vert _{\infty , \lambda _0, \kappa } + \int ^t_0 ( \lambda _0 - \lambda -\gamma _0 s)^{- (\alpha + \frac{3}{2}) } {\left| \left| \left| \omega (s) \right| \right| \right| }_s^2 \mathrm {d}s \\&\qquad + \int _0^t \frac{1}{\sqrt{t-s}} {\left| \left| \left| \omega (s) \right| \right| \right| }_1^2 \mathrm {d}s \\&\quad \lesssim \sum _{ |\beta |=2} \Vert D^\beta \omega _{0} \Vert _{\infty , \lambda _0, \kappa } +(\lambda _0- \lambda - \gamma _0 t)^{-\alpha }\\&\qquad \Big (\int ^t_0 (\lambda _0 - \lambda - \gamma _0 s)^{- \frac{3}{2}} \mathrm {d}s \Big ) \sup _{0 \le s\le t} {\left| \left| \left| \omega (s) \right| \right| \right| }_s^2\\&\qquad + \sqrt{t} \sup _{0 \le s \le t} {\left| \left| \left| \omega (s) \right| \right| \right| }_s^2 \\&\quad \lesssim \sum _{ |\beta |=2} \Vert D^\beta \omega _{0} \Vert _{\infty , \lambda _0, \kappa } +\Big ( (\lambda _0-\lambda -\gamma _0 t)^{-\alpha } \frac{1}{\gamma _0} + \sqrt{t}\Big ) \sup _{0 \le s \le t} {\left| \left| \left| \omega (s) \right| \right| \right| }_{s}^2. \end{aligned} \end{aligned}$$
(5.44)

Therefore we conclude that, from (5.42) and (5.44),

$$\begin{aligned} {\left| \left| \left| \omega (t) \right| \right| \right| }_{\infty ,\kappa } \lesssim \sum _{0 \le |\beta | \le 2} \Vert D^\beta \omega _{0} \Vert _{\infty , \lambda _0, \kappa } +\Big ( \sqrt{t}+ \frac{1}{\gamma _0}\Big ) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_{s}^2 \ \ \text {for} \ t < \frac{\lambda _0}{2\gamma _0}.\nonumber \\ \end{aligned}$$
(5.45)

In conclusion, from (5.41), (5.45), and by a standard continuity argument we obtain for sufficiently large \(\gamma _0\)

$$\begin{aligned} \sup _{0 \le t <\frac{\lambda _0}{2\gamma _0}}{\left| \left| \left| \omega (t) \right| \right| \right| }_{t}\lesssim \sum _{0 \le |\beta | \le 2} \Vert D^\beta \omega _{0} \Vert _{\infty , \lambda _0, \kappa }+ \sum _{0 \le |\beta |\le 2} \Vert D^\beta (1+ |\nabla _h|) \omega _{0} \Vert _{1, \lambda _0 }.\nonumber \\ \end{aligned}$$
(5.46)

Step 2: Propagation of analytic norms for \(\partial _t\omega \). The continuity of \(\omega (t)\) in t follows from the mild solution form (5.14) of \(\omega _\xi (t)\). We claim that \(\omega (t) \in C^1([0,T]; {\mathfrak {B}}^{\lambda ,\kappa })\) and moreover \({\left| \left| \left| \partial _t\omega (t) \right| \right| \right| }_{t} \) is bounded. To this end, we first derive the mild form of \(\partial _t\omega _\xi \) from (5.14):

$$\begin{aligned} \begin{aligned} \partial _t\omega _\xi (t,x_3) =&{\int ^\infty _0 G_\xi (t,x_3, y) \partial _t \omega _{0 \xi } (y) \mathrm {d}y}\\&\qquad + {\int ^t_0\int ^\infty _0 G_\xi (t-s, x_3, y) \partial _s N_\xi (s,y ) \mathrm {d}y \mathrm {d}s}\\&- {\int ^t_0 G_\xi (t-s, x_3, 0) (\partial _s B_\xi (s), 0) \mathrm {d}s} , \end{aligned} \end{aligned}$$
(5.47)

where we recall \(\partial _t \omega _0\) in (2.32). To justify this formula, we first recall (5.16)-(5.18). We start with the horizontal part of the formula (5.47) for \(\partial _t\omega _{\xi ,h}\). From Lemma 11, \(G_{\xi h} (t,x_3,y) = H_\xi (t, x_3-y) +H_\xi (t,x_3+y)+ R_\xi (t,x_3+y)\). Then by using the fact that \(H^\prime _\xi (t, \cdot )\) is an odd function, we see \( \partial _{x_3} G_{\xi h} (t,x_3,y)|_{x_3=0} = R^\prime _\xi (t, y)\). Now we read (5.17) as

$$\begin{aligned} \begin{aligned} \kappa \eta _0 R^\prime _\xi (t,y ) + \kappa \eta _0 |\xi | G_{\xi h} (t,0,y)=0, \quad \kappa \eta _0 R^\prime _\xi (t,x_3) + \kappa \eta _0 |\xi | G_{\xi h} (t,x_3,0)=0 , \end{aligned}\nonumber \\ \end{aligned}$$
(5.48)

where we have used that \(H_\xi (t, \cdot )\) is an even function for the second relation. On the other hand, since we also have \( \partial _{y_3} G_{\xi h} (t,x_3,y)|_{y=0} = R^\prime _\xi (t, x_3)\), we deduce that

$$\begin{aligned} \kappa \eta _0 (\partial _{y_3} +|\xi |) G_{\xi h} (t,x_3,y_3) =0, \ \ y_3=0. \end{aligned}$$
(5.49)

It is straightforward to see \(\Delta _\xi G_{\xi h} = \partial _{x_3}^2G_{\xi h} - |\xi |^2G_{\xi h} = \partial _y^2 G_{\xi h} - |\xi |^2 G_{\xi h}.\)

We now take \(\partial _t\) of (5.14):

$$\begin{aligned}&\partial _t \int ^\infty _0 G_{\xi h} (t,x_3,y)\omega _{0\xi ,h}(y) \mathrm {d}y = \int ^\infty _0 \partial _t G_{\xi , h} (t,x_3,y)\omega _{0\xi , h}(y) \mathrm {d}y \\&\quad = \int ^\infty _0 \kappa \eta _0 (\partial _{y}^2 - |\xi |^2) G_{\xi h} (t,x_3,y)\omega _{0\xi ,h}(y) \mathrm {d}y \\&\quad = \big [ \kappa \eta _0 \partial _y G_{\xi h} (t,x_3,y)\omega _{0\xi , h}(y) \big ]_{y=0}^{y=\infty } - \int ^\infty _0 \kappa \eta _0 |\xi |^2 G_{\xi h} (t,x_3,y)\omega _{0\xi , h }(y) \mathrm {d}y\\&\qquad - \int ^\infty _0 \kappa \eta _0 \partial _y G_{\xi h} (t,x_3,y)\partial _y\omega _{0\xi , h}(y) \mathrm {d}y \\&\quad = \big [ \kappa \eta _0 \partial _y G_{\xi h} (t,x_3,y)\omega _{0\xi , h }(y) \big ]_{y=0}^{y=\infty } - \big [ \kappa \eta _0 G_{\xi h} (t,x_3,y)\partial _y\omega _{0\xi , h}(y) \big ]_{y=0}^{y=\infty } \\&\qquad + \int ^\infty _0 G_{\xi h} (t,x_3,y)\kappa \eta _0 \Delta _\xi \omega _{0\xi , h}(y) \mathrm {d}y\\&\quad =- \kappa \eta _0 \partial _y G_{\xi h} (t,x_3,0)\omega _{0\xi ,h}(0) + \kappa \eta _0 G_{\xi h} (t,x_3,0)\partial _y\omega _{0\xi ,h}(0) \\&\qquad + \int ^\infty _0 G_{\xi h} (t,x_3,y)\kappa \eta _0 \Delta _{\xi h} \omega _{0\xi ,h}(y) \mathrm {d}y, \end{aligned}$$

and

$$\begin{aligned}&\partial _t {\int ^t_0\int ^\infty _0 G_{\xi h} (t-s, x_3, y) N_{\xi , h} (s,y ) \mathrm {d}y \mathrm {d}s}\\&\quad =\int _0^\infty G_{\xi h} (t,x_3,y) N_{\xi ,h} (0,y)\mathrm {d}y\\&\qquad + \int ^t_0 \int ^\infty _0 G_{\xi h} (s,x_3, y) \partial _t N_{\xi , h} (t-s,y) \mathrm {d}y \mathrm {d}s,\\&\partial _t {\int ^t_0 G_{\xi h} (t-s, x_3, 0) B_\xi (s) \mathrm {d}s} = G_{\xi h} (t,x_3,0) B_\xi (0) \\&\qquad +\int ^t_0 G_{\xi h} (t-s,x_3,0) \partial _s B_\xi ( s) \mathrm {d}s \end{aligned}$$

Therefore we obtain

$$\begin{aligned} \begin{aligned} \partial _t \omega _{\xi , h} (t,x_3) =&- \kappa \eta _0 \partial _y G_{\xi h} (t,x_3,0)\omega _{0\xi ,h}(0) + \kappa \eta _0 G_{\xi h} (t,x_3,0)\partial _y\omega _{0\xi , h}(0) \\&- G_{\xi h} (t,x_3,0) B_\xi (0) \\&+ \int ^\infty _0 G_{\xi h} (t,x_3,y) \{ \kappa \eta _0\Delta _\xi \omega _{0\xi ,h}(y) + N_{\xi , h} (0,y) \} \mathrm {d}y\\&+ \int ^t_0 \int ^\infty _0 G_{\xi h} (t-s,x_3, y) \partial _s N_{\xi , h} (s,y) \mathrm {d}y \mathrm {d}s \\&- \int ^t_0 G_{\xi h} (t-s,x_3,0) \partial _s B_\xi (s) \mathrm {d}s. \end{aligned} \end{aligned}$$
(5.50)

Next we show that the first line in the right-hand side is 0. From (5.49)

$$\begin{aligned}&- \kappa \eta _0 \partial _y G_{\xi h} (t,x_3,0)\omega _{0\xi , h}(0) + \kappa \eta _0 G_{\xi h} (t,x_3,0)\partial _y\omega _{0\xi ,h}(0)\\&\quad =G_{\xi h} (t,x_3,0) \kappa \eta _0(|\xi | + \partial _y) \omega _{0\xi , h}(0), \end{aligned}$$

and hence the first line of (5.50) reads

$$\begin{aligned} G_{\xi h} (t,x_3,0) \left[ \kappa \eta _0(|\xi | + \partial _{x_3}) \omega _{0\xi , h}(0) - B_\xi (0) \right] , \end{aligned}$$
(5.51)

which is zero due to the first compatibility condition of (2.34). Recalling \(\partial _t \omega _0\) in (2.32), the formula (5.47) for \(\partial _t\omega _{\xi ,h}\) has been established. We may follow the same procedure to verify the vertical part of the formula (5.47) for \(\partial _t\omega _{\xi ,3}\) by noting that the second compatibility condition of (2.34) removes the term \(- \kappa \eta _0 \partial _y G_{\xi 3} (t,x_3,0)\omega _{0\xi ,3}(0)\) which would create the initial layer otherwise because \( \partial _y G_{\xi 3} (t,x_3,0)\) does not vanish.

We may now repeat Step 1 for \(\partial _t\omega \) using the representation formula (5.47). The estimates are obtained in the same fashion. For the nonlinear terms, since \(\partial _t N = - u \cdot \nabla \partial _t \omega - \partial _t u\cdot \nabla \omega + \omega \cdot \nabla \partial _t u+\partial _t \omega \cdot \nabla u \), the structure of \(\partial _t N\) with respect to \(\partial _t\omega \) is consistent with the one of N with respect to \(\omega \) and we can use the bilinear estimates (5.3) and (5.4). In summary, one can derive that for \(t < \frac{\lambda _0}{2 \gamma _0}\)

$$\begin{aligned} {\left| \left| \left| \partial _t\omega (t) \right| \right| \right| }_{1}&\lesssim \sum _{0\le |\beta |\le 2} \Vert D^\beta (1+|\nabla _h|) \partial _t \omega _0 \Vert _{1, \lambda _0} \nonumber \\&\quad +\left( t+ \frac{1}{\gamma _0}\right) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_{s} \sup _{0 \le s \le t} {\left| \left| \left| \partial _t \omega (s) \right| \right| \right| }_s, \end{aligned}$$
(5.52)
$$\begin{aligned} {\left| \left| \left| \partial _t\omega (t) \right| \right| \right| }_{\infty ,\kappa }&\lesssim \sum _{0\le |\beta |\le 2} \Vert D^\beta \partial _t \omega _{0} \Vert _{\infty , \lambda _0, \kappa } \nonumber \\&\quad +\left( \sqrt{t}+ \frac{1}{\gamma _0}\right) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_{s} \sup _{0 \le s \le t} {\left| \left| \left| \partial _t \omega (s) \right| \right| \right| }_s, \end{aligned}$$
(5.53)

which lead to the desired bounds for \(\partial _t\omega (t)\) by choosing sufficiently large \(\gamma _0\).

Step 3: Propagation of analytic norms for \(\partial _t^2\omega _\xi \). As a consequence of Step 2, \(\partial _t\omega _\xi (t,x_3)\) solves the following system

$$\begin{aligned} \partial _t^2 \omega _\xi - \kappa \eta _0 \Delta _\xi \partial _t \omega _\xi =\partial _t N_\xi \quad&\text {in }{\mathbb {R}}_+, \end{aligned}$$
(5.54)
$$\begin{aligned} \kappa \eta _0 (\partial _{x_3} + |\xi | )\partial _t\omega _{\xi ,h} = \partial _tB_\xi \quad&\text {on } x_3=0, \end{aligned}$$
(5.55)
$$\begin{aligned} \partial _t \omega _{\xi , 3} = 0 \quad&\text {on } x_3=0, \end{aligned}$$
(5.56)

with \(\partial _t\omega _\xi |_{t=0} = \partial _t \omega _{0 \xi }\) for \(\xi \in {\mathbb {Z}}^2\) where \(\partial _t \omega _0\) is defined in (2.32). Then as done in Step 2, by using the properties of \(G_\xi \) and integration by parts and by the last compatibility condition of (2.34), we can derive the representation formula for \(\partial _t^2\omega \):

$$\begin{aligned} \begin{aligned}&\partial _t^2\omega _\xi (t,x_3) =( G_{\xi h} (t,x_3,0) \left[ \kappa \eta _0(|\xi | + \partial _{x_3}) \partial _t \omega _{ 0\xi , h}(0) - \partial _tB_\xi (0) \right] , 0 ) \\&\quad \quad \quad + {\int ^\infty _0 G_\xi (t,x_3, y) \partial _t^2 \omega _{0\xi } (y) \mathrm {d}y} + {\int ^t_0\int ^\infty _0 G_\xi (t-s, x_3, y) \partial _s^2 N_\xi (s,y ) \mathrm {d}y \mathrm {d}s}\\&\quad \quad \quad - {\int ^t_0 G_\xi (t-s, x_3, 0)( \partial _s^2 B_\xi (s),0) \mathrm {d}s} , \end{aligned}\nonumber \\ \end{aligned}$$
(5.57)

where we recall \(\partial _t^2\omega _0\) in (2.32). As we do not require higher order compatibility condition for the horizontal vorticity, a new term representing the initial-boundary layer emerges. We first examine \(G_\xi (t,z,0)\). Recall (5.35).

Similar to Lemma 13, we have for \(C_0<\infty \)

$$\begin{aligned} \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j G_\xi (t,z, 0 ) \right\| _{{\mathcal {L}}^1_\lambda } \lesssim C_0 , \quad \sum _{j=0}^k \left\| (\zeta (z)\partial _{z})^j G_\xi (t,z, 0 ) \right\| _{{\mathcal {L}}^\infty _{\lambda ,\kappa t}} \lesssim C_0. \end{aligned}$$
(5.58)

From (5.58), (5.33) and (5.3)

$$\begin{aligned}&\sum _{0 \le |\beta | \le 2} \sum _{\xi \in {\mathbb {Z}}^2} e^{\lambda |\xi |} \left\| D^\beta _\xi (1+ |\xi |) \right. \\&\qquad \left. \Big [ ( G_{\xi h} (t,x_3,0) \left( \kappa \eta _0(|\xi | + \partial _{x_3}) \partial _t \omega _{ 0\xi , h}(0) - \partial _tB_\xi (0) \right) , 0 )\Big ]\right\| _{{\mathcal {L}}^1_\lambda }\\&\quad \lesssim \kappa \eta _0 \sum _{0 \le |\beta | \le 2} \Vert \nabla _h^\beta (1+ |\nabla _h|) \nabla \partial _t \omega _{0, h}\Vert _{1,\lambda } + \sum _{0 \le |\beta | \le 2} \Vert \nabla ^\beta _h (1+ |\nabla _h|) \partial _t N(0) \Vert _{1,\lambda }\\&\quad \lesssim \kappa \eta _0 \Vert (1+ |\nabla _h|^3) \nabla \partial _t \omega _{0 } \Vert _{1,\lambda }\\&\qquad + \Vert (1+|\nabla _h|^4) \partial _t \omega _{0} \Vert _{1,\lambda } \sum _{0 \le |\beta | \le 1} \Vert D^\beta (1+ |\nabla _h|^3) \partial _t \omega _{0}\Vert _{1,\lambda }. \end{aligned}$$

Hence an \({L}^1\)-based analytic norm is easily obtained as

$$\begin{aligned} \begin{aligned}&{\left| \left| \left| \partial _t^2 \omega (t) \right| \right| \right| } _1\\&\quad \lesssim \kappa \eta _0 \Vert (1+ |\nabla _h|^3) \nabla \partial _t \omega _{0 } \Vert _{1, \lambda } + \Vert (1+|\nabla _h|^4) \partial _t \omega _{0} \Vert _{1, \lambda } \sum _{0 \le |\beta | \le 1} \Vert D^\beta (1\\&\qquad + |\nabla _h|^3) \partial _t \omega _{0}\Vert _{1, \lambda } \\&\qquad + \sum _{0\le |\beta |\le 2} \Vert D^\beta (1+|\nabla _h|) \partial ^2_t \omega _0 \Vert _{1, \lambda _0} +( t+ \frac{1}{\gamma _0}) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_{s} \sup _{0 \le s \le t} {\left| \left| \left| \partial _t^2 \omega (s) \right| \right| \right| }_s \\&\qquad +( t+ \frac{1}{\gamma _0}) \sup _{0 \le s \le t}{\left| \left| \left| \partial _t \omega (s) \right| \right| \right| }_{s} ^2. \end{aligned}\nonumber \\ \end{aligned}$$
(5.59)

Now we move to the \(L^\infty \)-based analytic norm bound. We compute \(\Vert G_\xi (t,z, 0 ) \Vert _{{\mathcal {L}}^\infty _{\lambda ,\kappa t}}\):

$$\begin{aligned} \Vert (\zeta (z) \partial _z)^jG_\xi (t,z, 0 ) \Vert _{{\mathcal {L}}^\infty _{\lambda ,\kappa t}} \lesssim&\sup _z \left[ \frac{ be^{({{\bar{\alpha }}}-c_0 b) z}}{1+\phi _\kappa (z) +\phi _{\kappa t} (z)}\right] \\&+ \sup _z\left[ \frac{e^{ {{\bar{\alpha }}} z-c_0 \frac{|z|^2}{\kappa t}}}{\sqrt{\kappa t}+\sqrt{\kappa t}\phi _\kappa (z) +\sqrt{\kappa t}\phi _{\kappa t}(z) } \right] . \end{aligned}$$

It is a routine to check that both supremum norms are uniformly bounded in \(\kappa \) and \(|\xi |\). Hence (5.58) shows that the kernel \(G_\xi (t,z,0)\) is well-behaved in \({\mathcal {L}}^1_\lambda \) and the initial-boundary layer analytic space \({\mathcal {L}}^\infty _{\lambda ,\kappa t}\). Then we run the same argument as in Step 2 but with \({\mathcal {L}}^\infty _{\lambda ,\kappa t}\) in place of \({\mathcal {L}}^\infty _{\lambda ,\kappa }\). Thanks to (5.58), the estimates of the first term in (5.57) are bounded by the initial norm (2.33):

$$\begin{aligned}&\sum _{0 \le |\beta | \le 2} \sum _{\zeta \in {\mathbb {Z}}^2} e^{\lambda |\xi |} \left\| D^{\beta }_\xi \Big [( G_{\xi h} (t,x_3,0) \left( \kappa \eta _0(|\xi | + \partial _{x_3}) \partial _t \omega _{ 0\xi , h}(0) - \partial _t B_\xi (0) \right) , 0 ) \Big ]\right\| _{{\mathcal {L}}^\infty _{\lambda , \kappa t}}\\&\quad \lesssim \kappa \eta _0 \sum _{0 \le |\beta | \le 2} \Vert \nabla _h^\beta \nabla \partial _t \omega _{0} \Vert _{\infty , \lambda } + \sum _{0 \le |\beta | \le 2} \Vert \nabla ^\beta _h \partial _t N(0) \Vert _{1,\lambda }\\&\quad \lesssim \kappa \eta _0 \sum _{0 \le |\beta | \le 2} \Vert \nabla _h^\beta \nabla \partial _t \omega _{0} \Vert _{\infty , \lambda }\\&\qquad + \Vert (1+|\nabla _h|^3) \partial _t \omega _{0} \Vert _{1,\lambda } \sum _{0 \le |\beta | \le 1} \Vert D^\beta (1 + |\nabla _h|^2) \partial _t \omega _{0}\Vert _{1,\lambda }. \end{aligned}$$

Other three terms are estimated in the same way as in [47] or [54] and we arrive that

$$\begin{aligned} \begin{aligned}&{\left| \left| \left| \partial _t^2 \omega (t) \right| \right| \right| } _{\infty , \kappa t}\\&\quad \lesssim \kappa \eta _0\! \sum _{0 \le |\beta | \le 2} \Vert \nabla _h^\beta \nabla \partial _t \omega _{0} \Vert _{\infty , \lambda } \\&\qquad + \Vert (1+|\nabla _h|^3) \partial _t \omega _{0} \Vert _{1,\lambda } \sum _{0 \le |\beta | \le 1} \Vert D^\beta (1+ |\nabla _h|^2) \partial _t \omega _{0}\Vert _{1,\lambda } \\&\qquad + \sum _{0\le |\beta |\le 2} \Vert D^\beta \partial ^2_t \omega _0 \Vert _{\infty , \lambda _0, \kappa t} +( \sqrt{t}+ \frac{1}{\gamma _0}) \sup _{0 \le s \le t}{\left| \left| \left| \omega (s) \right| \right| \right| }_{s} \sup _{0 \le s \le t} {\left| \left| \left| \partial _t^2 \omega (s) \right| \right| \right| }_s \\&\qquad +(\sqrt{ t}+ \frac{1}{\gamma _0}) \sup _{0 \le s \le t}{\left| \left| \left| \partial _t \omega (s) \right| \right| \right| }_{s} ^2. \end{aligned}\nonumber \\ \end{aligned}$$
(5.60)

Finally combining (5.59) and (5.60) and then choosing sufficiently large \(\gamma _0\) we derive a desired estimate for \({\left| \left| \left| \partial _t^2 \omega (t) \right| \right| \right| }_t\) for \(t \in (0, \frac{\lambda _0}{2\gamma _0})\).

Altogether from (5.46), (5.52), (5.53), (5.59), and (5.60), we finish the proof of the estimate (2.35).

Step 4: Estimate (1), vorticity estimates. Both (2.36) and (2.37) are direct consequences of (2.35). To show (2.38), we first note that the boundedness of \(\omega (t)\) norms implies \(|\partial _{x_3}\omega _\xi (t,x_3)|\lesssim e^{-{\bar{\alpha }} x_3}e^{-\lambda |\xi |}\) for all \(|\xi |\) and \(x_3 \ge 1\) (away from the boundary). When \(x_3\le 1\), we draw on the equation (5.10) to rewrite \( \partial _{x_3}^2\omega _{\xi ,h} = \frac{1}{\kappa \eta _0}\{ \partial _t \omega _{\xi ,h} + \kappa \eta _0 |\xi |^2 \omega _{\xi ,h} - N_{\xi ,h}\}\) and the boundary condition (5.11):

$$\begin{aligned} \begin{aligned} \partial _{x_3} \omega _{\xi ,h} (t, x_3)&= \partial _{x_3} \omega _{\xi ,h} (t,0) + \int _0^{x_3} \partial _{x_3}^2\omega _{\xi ,h}(t,y) \mathrm {d}y\\&=- |\xi | \omega _{\xi ,h} (t,0) + \frac{1}{\kappa \eta _0} B_\xi (t) + \int _0^{x_3} \frac{1}{\kappa \eta _0}[ \partial _t \omega _{\xi ,h} + \kappa \eta _0 |\xi |^2 \omega _{\xi ,h} \\&\quad - N_{\xi ,h}](t,y)\mathrm {d}y. \end{aligned}\nonumber \\ \end{aligned}$$
(5.61)

We now appeal to \(|B_\xi (t)| \le \Vert N_\xi (t)\Vert _{{\mathcal {L}}^1_\lambda }\) and \(\sum _{0\le \ell \le 1} ( {\left| \left| \left| \partial _t^\ell \omega (t) \right| \right| \right| }_{\infty ,\kappa }+ {\left| \left| \left| \partial _t^\ell \omega (t) \right| \right| \right| }_{1}) <\infty \) to obtain that for all \(x_3\in {\mathbb {R}}_+\)

$$\begin{aligned} | \partial _{x_3} \omega _{\xi , h} (t, x_3)|\lesssim \frac{1}{\kappa } e^{-{\bar{\alpha }} x_3}e^{-\lambda |\xi |} \text { for } 0<\lambda <\lambda _0, \end{aligned}$$
(5.62)

which proves (2.38) for \(\omega _h\) and \(\ell =0\). The remaining case can be estimated similarly. Near O(1) boundary, from (5.54) and (5.55), we derive

$$\begin{aligned} \begin{aligned}&\partial _{x_3} \partial _t\omega _{\xi ,h} (t, x_3)\\&\quad =- |\xi | \partial _t\omega _{\xi ,h} (t,0) + \frac{1}{\kappa \eta _0} \partial _tB_\xi (t) \\&\qquad + \int _0^{x_3} \frac{1}{\kappa \eta _0}[\partial _t^2 \omega _{\xi ,h} + \kappa \eta _0 |\xi |^2 \partial _t\omega _{\xi ,h} - \partial _tN_{\xi ,h}](t,y)\mathrm {d}y. \end{aligned} \end{aligned}$$
(5.63)

Together with \(\sum _{0\le \ell \le 1} {\left| \left| \left| \partial _t^\ell \omega (t) \right| \right| \right| }_{\infty ,\kappa }+ \sum _{0\le \ell \le 2} {\left| \left| \left| \partial _t^\ell \omega (t) \right| \right| \right| }_{1}<\infty \) we deduce (2.38) for \(\omega _h\) and \(\ell =1\). For \(\omega _3\) we use \(\nabla \cdot \omega =0\) to write \(\partial _3 \omega _3 = - \partial _1\omega _1- \partial _2\omega _2\). Now (2.38) for \(\omega _3\) follows from (2.36).

Step 5: Estimate (2), velocity estimates, except (2.42). From (5.2)

$$\begin{aligned} |\xi |^{\beta _h} | \partial _{z}^{\beta _3}\partial _t^\ell \phi _\xi (t,z)| \lesssim \int _{\partial {\mathcal {H}}_\lambda } |\xi |^{|\beta |-1} e^{-|\xi | |y-z|} | \partial _t^\ell \omega _\xi (t,y)| |\mathrm {d}y| \ \ \text {for} \ \beta _3\le 1.\qquad \end{aligned}$$
(5.64)

For \(|\beta |= |\beta _h| + \beta _3=1\) we bound (5.64) by \(e^{-\lambda |\xi |} \Vert \partial _t^\ell \omega (t) \Vert _{1,\lambda }\). Then from (2.35) we conclude (2.39).

For \(|\beta |\ge 2\) and \(\beta _3 \le 1\), we bound (5.64) by

$$\begin{aligned} \begin{aligned} (5.64)&\lesssim \int _{\partial {\mathcal {H}}_\lambda } |\xi |^{|\beta |-2} |\xi |e^{-|\xi | |y-z|} e^{- {{\bar{\alpha }}} \text {Re}\,y}e^{-\lambda |\xi |} \big (1+ \phi _\kappa (y)+ \phi _{\kappa t} (y)\big ) |\mathrm {d}y| \\&\lesssim |\xi |^{|\beta |-2} e^{-\lambda |\xi |} e^{- \min (1, \frac{{\bar{\alpha }}}{2})x_3} \int _{\partial {\mathcal {H}}_\lambda } e^{- \frac{{\bar{\alpha }}}{2} \text {Re} \, y}\big ( 1+ \phi _\kappa (y)+ \phi _{\kappa t} (y)\big ) |\mathrm {d}y|\\&\lesssim |\xi |^{|\beta |-2} e^{-\lambda |\xi |} e^{- \min (1, \frac{{\bar{\alpha }}}{2})x_3} \ \ \text {for} \ |\beta | \ge 2, \ \text {and} \ \beta _3 \le 1, \\&\quad \text {and} \ \ell =0,1,2, \ \text {and} \ t \in [0,T], \end{aligned}\nonumber \\ \end{aligned}$$
(5.65)

where we have used \(|\xi ||y-z| + \frac{{\bar{\alpha }}}{2} \text {Re}\, y\ge \min (1, \frac{{\bar{\alpha }}}{2} ) x_3\) for \(|\xi |\ge 1\) and (2.35).

For \(\beta _3=2,3\) we use \(\partial _z^2 \partial _t^\ell \phi _\xi = |\xi |^2 \partial _t^\ell \phi _\xi + \partial _t^\ell \omega _\xi \). Then following the same argument of (5.65), we derive

$$\begin{aligned} \begin{aligned}&|\xi |^{\beta _h} | \partial _{z}^{\beta _3}\partial _t^\ell \phi _\xi (t,z)|\\&\quad \lesssim |\xi |^{|\beta _h|+2} | \partial _{z}^{\beta _3-2}\partial _t^\ell \phi _\xi (t,z)| +|\xi |^{\beta _h} | \partial _{z}^{\beta _3-2}\partial _t^\ell \omega _\xi (t,z)|\\&\quad \lesssim {\left\{ \begin{array}{ll} (|\xi |^{|\beta |-2}+|\xi |^{\beta _h}) e^{- \lambda |\xi |} e^{- \min (1, \frac{{\bar{\alpha }}}{2}) \text {Re}\, z} (1+ \phi _\kappa (z)) \ \ \text {for} \ \ell =0,1, \ \text {and} \ \beta _3=2,\\ (|\xi |^{|\beta |-2}+|\xi |^{\beta _h}) e^{- \lambda |\xi |} e^{- \min (1, \frac{{\bar{\alpha }}}{2})\text {Re}\, z}\kappa ^{-1} \ \ \text {for} \ \ell =0,1, \ \text {and} \ \beta _3=3, \\ (|\xi |^{|\beta |-2}+|\xi |^{\beta _h}) e^{- \lambda |\xi |} e^{- \min (1, \frac{{\bar{\alpha }}}{2})\text {Re}\, z} (1+ \phi _\kappa (z)+ \phi _{\kappa t} (z)) \ \ \text {for} \ \ell =2, \ \text {and} \ \beta _3=2. \end{array}\right. } \end{aligned}\nonumber \\ \end{aligned}$$
(5.66)

Finally from (5.65) and (5.66) we conclude (2.40) and (2.41).

Step 6: Estimate (3), pressure estimates and (2.42). We next turn to the pressure. Taking the divergence to (1.13) and using (1.14), we deduce

$$\begin{aligned} - \Delta p = \sum _{\ell ,m=1}^3 \partial _{ \ell } u_m \partial _{ m} u_\ell . \end{aligned}$$
(5.67)

We obtain the boundary condition of p by reading the third component of (1.13), and then using (1.14) and (1.15),

$$\begin{aligned} \begin{aligned} \partial _{ 3} p&= \kappa \eta _0 \Delta u_3= \kappa \eta _0 \partial _{ 3} \partial _{ 3} u_3 = - \kappa \eta _0 \partial _1 \partial _{ 3} u_1 - \kappa \eta _0 \partial _2\partial _{ 3} u_2\\&= - \kappa \eta _0 \partial _1 (\omega _2 + \partial _1 u_3) - \kappa \eta _0 \partial _2 (-\omega _1 + \partial _2 u_3)\\&= - \kappa \eta _0 \partial _1 \omega _2 + \kappa \eta _0 \partial _2 \omega _1 \ \ \text {for} \ x_3=0, \end{aligned} \end{aligned}$$
(5.68)

where \(\omega _1= \partial _2 u_3 - \partial _3 u_2\) and \(\omega _2 = - \partial _1 u_3 + \partial _3 u_1\).

In the Fourier side we read the problem as

$$\begin{aligned} \begin{aligned} (|\xi |^2 - \partial _3^2) p_\xi (t,x_3) = g_\xi (t,x_3):= \sum _{\ell , m=1}^3 (\partial _\ell u_m \partial _m u_\ell )_\xi (t,x_3) \ \ \text {for} \ x_3 \in {\mathbb {R}}_+,\\ \partial _3 p_\xi (t,0) =- i \kappa \eta _0 \xi _1 \omega _{\xi ,2} (t,0)+ i \kappa \eta _0 \xi _2 \omega _{\xi , 1}(t,0). \end{aligned}\nonumber \\ \end{aligned}$$
(5.69)

A representation of \(p_\xi (t,x_3)\) is given by

$$\begin{aligned} \begin{aligned} p_\xi (t,x_3)&= -\int ^{x_3}_0 \frac{1}{2|\xi |} e^{-|\xi | (x_3-y)} g_\xi (y) \mathrm {d}y - \int _{x_3}^\infty \frac{1}{2|\xi |} e^{-|\xi | (y-x_3)} g_\xi (y) \mathrm {d}y\\&\quad - \int ^\infty _0 \frac{1}{2|\xi |} e^{-|\xi | (y+x_3)} g_\xi (y) \mathrm {d}y \\&\quad - \frac{1}{|\xi |} e^{-|\xi | x_3} (- i \kappa \eta _0 \xi _1 \omega _{\xi ,2} (t,0)+ i \kappa \eta _0 \xi _2 \omega _{\xi , 1}(t,0)), \end{aligned} \end{aligned}$$
(5.70)

which is valid for all \(\xi \ne 0\). When \(\xi =0\), by integrating (5.69) and by using the boundary conditions \(\partial _3 p_0(t,0)=0\), \(u(t,x_h,0)=0\) and the divergence free condition \(\nabla \cdot u=0\), we first obtain

$$\begin{aligned} \begin{aligned} \partial _3 p_0(t,x_3)&= - \frac{1}{(2\pi )^2} \int _0^{x_3}\iint _{{\mathbb {T}}^2} \sum _{\ell ,m=1}^3 \partial _{ \ell } u_m \partial _{ m} u_\ell \mathrm {d}x_h \mathrm {d}y_3 \\&= - \frac{1}{(2\pi )^2}\iint _{{\mathbb {T}}^2} (u\cdot \nabla u_3)(t,x_h,x_3) \mathrm {d}x_h \\&= - \frac{2}{(2\pi )^2}\iint _{{\mathbb {T}}^2} (u_3\partial _3 u_3)(t,x_h,x_3) \mathrm {d}x_h , \end{aligned} \end{aligned}$$
(5.71)

where we have used the integration by parts and \(\nabla \cdot u=0\) at the last step.

Observe that \(\partial _3p_0\) decays exponentially in \(x_3\), and in particular \(\int _0^\infty | \partial _3 p_0(t,x_3) | \mathrm {d}x_3 <\infty \). The integration yields

$$\begin{aligned} p_0(t,x_3) =p_0(t,0) - \int _0^{x_3} \frac{2}{(2\pi )^2}\iint _{{\mathbb {T}}^2} (u_3\partial _3 u_3)(t,x_h,y_3) \mathrm {d}x_h \mathrm {d}y_3 . \end{aligned}$$

Since \(p_0(t,0)\) is a free constant in \(x_3\), we fix \(p_0(t,x_3)\) by choosing

$$\begin{aligned} p_0(t,0)= \frac{2}{(2\pi )^2}\int _0^{\infty } \iint _{{\mathbb {T}}^2} (u_3\partial _3 u_3)(t,x_h,y_3) \mathrm {d}x_h \mathrm {d}y_3 <\infty , \end{aligned}$$

such that

$$\begin{aligned} \begin{aligned} p_0(t,x_3)&= \frac{2}{(2\pi )^2}\int _{x_3}^\infty \iint _{{\mathbb {T}}^2} (u_3\partial _3 u_3)(t,x_h,y_3) \mathrm {d}x_h \mathrm {d}y_3 . \end{aligned} \end{aligned}$$
(5.72)

The pressure p is then recovered by

$$\begin{aligned} p(t,x_h,x_3) = p_0 (t,x_3) + \sum _{|\xi |\ge 1, \xi \in {\mathbb {Z}}^2} p_\xi (t,x_3) e^{ix_h\cdot \xi } , \end{aligned}$$
(5.73)

where \( p_0 (t,x_3) \) and \(p_\xi (t,x_3)\) are given in (5.72) and (5.70).

Now the pressure estimate follows readily from the velocity and vorticity estimates. To show (2.43), we first note from (2.39) and (2.40) \(|p_0(t,x_3)|\lesssim |u_3(t,x)| \int _\Omega |\partial _3 u_3 (t,x)| \mathrm {d}x \lesssim 1\) and from Lemma 8

$$\begin{aligned} \begin{aligned} | g_\xi |&\lesssim e^{-\lambda |\xi |}\sum _{i=1}^2 \left( \Vert \partial _i u_h \Vert ^2_{\infty ,\lambda }+ \Vert \zeta ^{-1}\partial _i u_3\Vert _{\infty ,\lambda }\Vert \zeta \partial _3 u_i\Vert _{\infty ,\lambda } \right) \\&\lesssim e^{-\lambda |\xi |} \left[ \sum _{0\le |\beta |\le 1} \Vert \nabla _h^\beta \omega \Vert ^2_{1,\lambda }+\sum _{1\le |\beta |\le 2} \Vert \nabla _h^\beta \omega _h\Vert _{1,\lambda } \right. \\&\qquad \left. \times \left( \sum _{0\le |\beta |\le 1} \Vert \nabla _h^\beta \omega \Vert _{1,\lambda }+ \Vert \zeta \omega _h\Vert _{\infty ,\lambda }\right) \right] , \end{aligned} \end{aligned}$$

from which we deduce \(|p(t,x_h,x_3)|\lesssim 1\). The estimation of \(\partial _t p \) and \(\partial _t^2p\) follows analogously.

For the decay estimates (2.44), we start with \(\ell =0\) and \(\beta =0\). Due to our choice of \(p_0(t,x_3)\) in (5.72), using (2.39) and (2.40), we have the spatial decay for \(p_0(t,x_3)\):

$$\begin{aligned} |p_0(t,x_3)|\lesssim \int _{x_3}^\infty \iint _{{\mathbb {T}}^2} (1+\phi _\kappa (y_3)) e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) y_3 } \mathrm {d}x_h \mathrm {d}y_3\lesssim \kappa ^{-\frac{1}{2}} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) x_3 }. \end{aligned}$$

For \(\xi \ne 0\), we use another estimate for \(|g_\xi |\) and Lemma 8

$$\begin{aligned} \begin{aligned} |g_\xi (y)|&\lesssim \sum _{\ell ,m=1}^3\sum _{\eta \in {\mathbb {Z}}^2} e^{-\lambda |\xi -\eta |} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) y}(1+\phi _\kappa (y)) |(\partial _m u_\ell )_\eta (y)| \\&\lesssim \kappa ^{-\frac{1}{2}} e^{-\lambda |\xi |} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) y} \sum _{\ell ,m=1}^3 \sum _{\eta \in {\mathbb {Z}}^2} e^{\lambda |\eta | } |(\partial _m u_\ell )_\eta (y)| , \end{aligned} \end{aligned}$$
(5.74)

from which we deduce that \(| p_\xi (t,x_3) |\lesssim \kappa ^{-\frac{1}{2}} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) x_3} \). Hence (2.44) holds for \(\ell =0\) and \(\beta =0\). For the pressure gradient estimate when \(|\beta |=1\), from (5.71) and (2.40) we first note

$$\begin{aligned} |\partial _3 p_0(t,x_3) | \lesssim \sup _{x_h \in {\mathbb {T}}^2}( |u_3| |\partial _3 u_3|) \lesssim (1+\phi _\kappa (x_3)) e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) x_3}. \end{aligned}$$

For \(\xi \ne 0\), by (5.74) it is easy to see that \(|\xi p_\xi (t,x_3) |\lesssim \kappa ^{-\frac{1}{2}} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) x_3} \). Note that \(\partial _3 p_\xi (t,x_3)\) has a similar integral form as \(|\xi | p_\xi (t,x_3)\) and the estimate follows in the same way, which results in \( |\partial _3 p_\xi (t,x_3) |\lesssim \kappa ^{-\frac{1}{2}} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) x_3}\). This finishes (2.44) for \(\ell =0\) and \(|\beta |=1\). The remaining cases for \(\ell =1\) and \(|\beta |=0,1\) can be treated in the same way.

For the decay estimate of \(\partial _t^2p\), we take into account the initial layer which occurs at \(\partial _t^2 \omega \) and \(\nabla \partial _t^2 u\). First using (2.39), (2.40) and (2.41) we have

$$\begin{aligned} \begin{aligned} |\partial _t^2 p_0(t,x_3) |&\lesssim \left| \int _{x_3}^\infty \iint _{{\mathbb {T}}^2} (u_3\partial _3\partial _t^2 u_3+ \partial _t^2u_3\partial _3 u_3+2\partial _tu_3\partial _3\partial _t u_3)(t,x_h,y_3) \mathrm {d}x_h \mathrm {d}y_3 \right| \\&\lesssim \big (1+ \phi _\kappa (x_3) + \phi _{\kappa t} (x_3)\big )e^{-\min (1, \frac{{\bar{\alpha }}}{2} )x_3} , \end{aligned} \end{aligned}$$

while for \(|\xi |\ne 0\) we have

$$\begin{aligned} \begin{aligned} |\partial _t^2 g_\xi (y)|&\lesssim \sum _{\ell ,m=1}^3\sum _{\eta \in {\mathbb {Z}}^2} e^{-\lambda |\xi -\eta |} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) y}(1+\phi _\kappa (y)) |(\partial _m u_\ell )_\eta (y)| \\&\lesssim \kappa ^{-\frac{1}{2}} e^{-\lambda |\xi |} e^{-\min (1,\frac{{{\bar{\alpha }}}}{2}) y} \sum _{i=1}^2 \sum _{\ell ,m=1}^3 \sum _{\eta \in {\mathbb {Z}}^2} e^{\lambda |\eta | } |(\partial _m \partial _t^i u_\ell )_\eta (y)| , \end{aligned} \end{aligned}$$

from which we deduce (2.45).

The last estimate for \(\partial _t^\ell u\) for \(\ell =1,2\) follows from the equation: \(\partial _t u = \kappa \eta _0 \Delta u - u \cdot \nabla u - \nabla p \) and \(\partial _t^2 u = \kappa \eta _0 \Delta \partial _t u - u \cdot \nabla \partial _t u - \partial _t u \cdot \nabla u - \nabla \partial _t p \).