1 Introduction
1.1 Presentation of the problem and related results
In this article, we study 2-D incompressible Navier-Stokes equations in a thin domain where the aspect ratio and the Reynolds number have certain constraints:
where $t\geq 0, (x,y)\in \mathcal {S}^{\varepsilon }=\left \{(x, y) \in \mathbb {T}\times \mathbb {R} : 0<y<\varepsilon \right \}.$ Here, $U(t,x,y), P(t,x,y)$ stand for the velocity and pressure function, respectively, and $\eta $ is a positive constant independent of $\varepsilon $ . The width of domain $\mathcal {S}^{\varepsilon }$ is $\varepsilon ,$ and the boundary condition in (1.1) corresponds to the non-slip condition at the walls $y = 0, \varepsilon .$ In addition, the system is prescribed with the initial data of the form
This is a classical model with applications to oceanography, meteorology and geophysical flows, where the vertical dimension of the domain is very small compared with the horizontal dimension of the domain.
To study the process $\varepsilon \to 0$ , we first fix the domain independent of $\varepsilon $ . Here, we rescale the system (1.1) as follows:
We put above relations into (1.1), and then (1.1) is reduced to a scaled anisotropic Navier-Stokes system:
where $(x,y)\in \mathcal {S}=\left \{(x, y) \in \mathbb {T}\times (0,1) \right \}$ .
To simplify the notations, we take $\eta =1$ in this paper and denote $\Delta _{\varepsilon }=\varepsilon ^2\partial _x^2+\partial _y^2$ .
Formally, taking $\varepsilon \to 0$ in (1.3), we derive the hydrostatic Navier-Stokes/Prandtl system (see [Reference Lagrée and Lorthois22, Reference Renardy31]):
This paper aims to justify the limit from the scaled anisotropic Navier-Stokes system (1.3) to the hydrostatic Navier-Stokes/Prandtl system (1.4) for a class of convex data in the optimal Gevrey class with index $\gamma =\frac 32$ .
Before presenting the precise statement of the main result in this paper, we recall some results on system (1.4). If $\eta =0$ in the system (1.4), we get the hydrostatic Euler system:
There is a lot of research on the system (1.5), and readers can refer to [Reference Brenier3, Reference Brenier4, Reference Cao, Ibrahim, Nakanishi and Titi5, Reference Grenier16, Reference Kukavica, Temam, Vicol and Ziane20, Reference Kukavica, Masmoudi, Vicol and Wong19, Reference Masmoudi and Wong26, Reference Renardy31, Reference Wong38]. Renardy [Reference Renardy31] proved the linearization of (1.5) has a growth like $e^{|k|t}$ if the initial data is not uniform convexity (or concavity) with respect to the variable y. Local well-posedness in the analytic setting was established in [Reference Kukavica, Temam, Vicol and Ziane20]. Under the convexity condition, Masmoudi and Wong [Reference Masmoudi and Wong26] got the well-posedness of (1.5) in the Sobolev space.
Next, we recall some results on the well-posedness of the hydrostatic Navier-Stokes/Prandtl system (1.4). Similar to the classical Prandtl equation, (1.4) loses one derivative because of the term $v_p\partial _y u_p$ . Paicu, Zhang and Zhang [Reference Paicu, Zhang and Zhang30] obtained the global well-posedness of system (1.4) when the initial data is small in the analytical space. Meanwhile, Renardy [Reference Renardy31] also proved that the linearization of the hydrostatic Navier-Stokes equations at certain parallel shear flows is ill-posed and may have a growth $e^{|k|t}$ which is the same as (1.5) when the initial data is not convex. Thus, to obtain well-posedness results that break through the analytic space, one may need the convexity condition on the velocity. For that, under the convexity condition, Gérard-Varet, Masmoudi and Vicol proved (1.4) is local well-posedness in the Gevrey class with index $9/8$ in [Reference Gérard-Varet, Masmoudi and Vicol15]. In [Reference Gérard-Varet, Masmoudi and Vicol15], they first derive the vorticity equations $\omega =\partial _y u$ :
where the worst term is $\partial _x v$ leading to one derivative loss. Then, they use the ‘hydrostatic trick’ which means that they take the inner product with $\partial _x \omega /\partial _y \omega $ ( $\partial _y\omega \geq c_0>0$ ) instead of $\partial _x \omega $ to take advantage of the cancellation:
Such an idea was used previously in [Reference Masmoudi and Wong26]. To close the energy estimates, the ‘hydrostatic trick’ is not enough due to the ‘bad’ boundary condition of $\omega $
which loses one derivative too. To overcome that, [Reference Gérard-Varet, Masmoudi and Vicol15] introduce the following decomposition:
where $\omega ^{bl}$ is the boundary corrector which satisfies that
Following the above decomposition, [Reference Gérard-Varet, Masmoudi and Vicol15] obtain the well-posedness results of (1.4) in the Gevrey class with index $\gamma =\frac 98$ .
To search the best functional space for the system (1.4), based on the Tollmien-Schlichting instabilities for Navier-Stokes [Reference Grenier, Guo and Nguyen17], Gérard-Varet, Masmoudi and Vicol also give the following conjecture: ‘Our conjecture - based on a formal parallel with Tollmien-Schlichting instabilities for Navier-Stokes [18] - is that the best exponent possible should be $\gamma =\frac 32$ , but the such result is for the time being out of reach’.
While studying the anisotropic Navier-Stokes system (1.3) and the hydrostatic Navier-Stokes/Prandtl system (1.4), another important problem is to justify the inviscid limit. Under the analytical setting, Paicu, Zhang and Zhang [Reference Paicu, Zhang and Zhang30] justified the limit from (1.3) to (1.4). Based on the work [Reference Gérard-Varet, Masmoudi and Vicol15], we [Reference Wang, Wang and Zhang35] justified the limit in the Gevrey class with index $\gamma =\frac 98$ .
In this paper, we aim to prove the conjecture of Gérard-Varet, Masmoudi and Vicol. To do that, we use some ideas from the classical inviscid limit theory. Next, we recall the recent development of the classical Prandtl equation and the inviscid limit theory.
There are a lot of papers studying the well-posedness of the Prandtl equation in some special functional space. For monotonic initial data, [Reference Oleinik29, Reference Alexandre, Wang, Xu and Yang1, Reference Masmoudi and Wong27] used a different method to get the local existence and uniqueness of classical solutions to the Prandtl equation in Sobolev space. Without monotonic condition, [Reference Lombardo, Cannone and Sammartino24, Reference Sammartino and Caflisch32] proved that the Prandtl equation is well-posedness in the analytic class; [Reference Gérard-Varet and Masmoudi14, Reference Li and Yang23, Reference Chen, Wang and Zhang6] proved the well-posedness of the Prandtl equations in the Gevrey class for a class of concave initial data. Without any structure assumption, Dietert and Gérard-Varet [Reference Dietert and Gérard-Varet8] proved well-posedness in the Gevrey space with index $\gamma =2$ . According to [Reference Gérard-Varet and Dormy10], $\gamma =2$ may be the optimal index for the well-posedness theory. For more results on the Prandtl equation, see [Reference Grenier and Nguyen18, Reference Wang37, Reference Wang, Xin and Zang36, Reference Xiao and Xin39, Reference Xin and Zhang40].
On the inviscid limit problem, we refer to [Reference Sammartino and Caflisch33, Reference Wang, Wang and Zhang34, Reference Kukavica, Vicol and Wang21, Reference Nguyen and Nguyen28, Reference Maekawa25, Reference Fei, Tao and Zhang9] for the analytical class. Note that going from analytic to Gevrey data is a challenging problem. The first result in the Gevrey class is given by [Reference Gérard-Varet, Maekawa and Masmoudi12]. Gérard-Varet, Masmoudi and Maekawa [Reference Gérard-Varet, Maekawa and Masmoudi12] proved the stability of the Prandtl expansion for the perturbations in the Gevrey class when $U^{BL}(t, Y)$ is a monotone and concave function where the boundary layer is the shear type like
where $\nu $ is the viscosity coefficient. Later, Chen, Wu and Zhang [Reference Chen, Wu and Zhang7] improved the results in [Reference Gérard-Varet, Maekawa and Masmoudi12] to get the $L^2\cap L^{\infty }$ stability. Very recently, Gérard-Varet, Masmoudi and Maekawa [Reference Gérard-Varet, Maekawa and Masmoudi13] used a very clever decomposition to get the optimal Prandtl expansion around the concave boundary layer. Their results generalized the one obtained in [Reference Gérard-Varet, Maekawa and Masmoudi12, Reference Chen, Wu and Zhang7], which restricted to expansions of shear flow type. In their paper, they decompose the stream function $\phi $ as follows:
where $\phi _{slip}$ enjoys a “good” boundary condition and $\phi _{bc}$ is a corrector which recover the boundary condition. This kind of decomposition is also used in [Reference Chen, Wu and Zhang7]. To estimate $\phi _{bc}$ , they also need the following decomposition
where $\phi _{bc, S}$ satisfies the Stokes equation, $\phi _{bc, T}$ is to correct the stretching term with ‘good’ boundary condition and $\phi _{bc, R}$ solves formally the same system as $\phi _{slip}$ . In this paper, we apply the decomposition in [Reference Gérard-Varet, Maekawa and Masmoudi13] to justify the limit from (1.3) to (1.4).
1.2 Statement of the main results
Before starting the main results, we give some assumptions on the initial data. Assume that initial data belong to the following Gevrey class:
where $H^{r,s}$ is the anisotropic Sobolev space defined by
More precisely, we consider the initial data of the form
which satisfies the compatibility conditions
Moreover, we assume the initial velocity satisfies the convex condition
Now, we are in the position to state the main results of our paper.
Theorem 1.1. Let initial data $u_0$ satisfy (1.6)-(1.9). Then there exist $T>0$ and $C> 0$ independent of $\varepsilon $ such that there exists a unique solution of the scaled anisotropic Navier–Stokes equations (1.3) in $[0,T]$ , which satisfies that for any $t\in [0,T],$ it holds that
where $(u_p, v_p)$ is the solution to (1.4).
Remark 1.2. Although we do not give the proof that the system (1.4) is well-posedness in Gevrey class $\frac 32$ , one can follow the proof of Theorem 1.1 to obtain the well-posedness. To avoid repeatability in the proof, we omit the details. Actually, the main difference between $\varepsilon =0$ and $\varepsilon \neq 0$ is on the construction boundary corrector $\phi _{bc,S}$ , and readers can find more details in Remark 8.1.
Remark 1.3. In the recent work [Reference Gérard-Varet, Iyer and Maekawa11], they established the well-posedness of the linearized Hydrostatic Navier-Stokes system around shear flow in Gevrey class $\frac 32$ . In our present work, we consider the general nonlinear system and focus on the inviscid limit problem.
1.3 Sketch of the proof
In this subsection, we sketch the main ingredients in our proof.
-
(1) Introduce the error equations. In Section 3, we deduce the error equations. We introduce the error
$$ \begin{align*} u^R=u^{\varepsilon}-u^p,\quad v^R=v^{\varepsilon}-v^p, \quad p^R=p^{\varepsilon}-p^p, \end{align*} $$which satisfies
(1.10) $$ \begin{align} \left\{ \begin{aligned} &\partial_t u^R-{\Delta_{\varepsilon} u^R}+v^R\partial_y u^p+\partial_x p^R=\cdots,\\ &\varepsilon^2(\partial_t v^R-\Delta_{\varepsilon} v^R)+\partial_y p^R=\cdots. \end{aligned} \right. \end{align} $$Here, $(u^p, v^p, p^p)$ is an approximate solution given in (3.1). The key point in this paper is to obtain the uniform estimate (in $\varepsilon $ ) of $(u^R, \varepsilon v^R)$ in the Gevrey class with index $\gamma =\frac 32.$ In view of (1.10), since $v^R$ is controlled via the relation $v^R=-\int _0^y \partial _x u^R dy'$ , the main difficulty comes from the term $v^R\partial _y u^p$ , which loses one tangential derivative. In [Reference Wang, Wang and Zhang35], we justify the limit in Gevrey class $\frac 98.$ For the data in the Gevrey class with optimal index $\gamma =\frac 32$ , we need to introduce new ideas.
-
(2) Introduce the vorticity formulation. In order to eliminate $p^R,$ we introduce vorticity $\omega ^R=-\varepsilon ^2\partial _x v^R+\partial _y u^R$ and rewrite the equation of $\omega ^R$ by stream function $\phi $ which satisfies
$$ \begin{align*} v^R=-\partial_x\phi, \quad u^R=\partial_y \phi+C(t), \quad C(t)=\frac1{2\pi}\int_{{\mathcal S}} u^Rdxdy. \end{align*} $$Thus, we get
(1.11) $$ \begin{align} \left\{ \begin{aligned} &(\partial_t-\Delta_{\varepsilon})\Delta_{\varepsilon} \phi-\partial_x\phi\partial_y \omega^p=\cdots,\qquad (x,y)\in\mathcal{S},\\ &\phi|_{y=0,1}=0, \quad \partial_y\phi|_{y=0,1}=C(t),\qquad x\in\mathbb{T}. \end{aligned} \right. \end{align} $$We notice the term $\partial _x\phi \partial _y \omega ^p$ also loses one tangential derivative. But under the convexity condition $\partial _y\omega ^p\geq c_0>0,$ one can use the ‘hydrostatic trick’ to deal with this term. Testing $\frac {\omega ^R}{\partial _y\omega ^p}$ to the (1.11) instead of $\omega ^R,$ we have the following cancellation:
$$ \begin{align*} -\int_{{\mathcal S}}\partial_x\phi \partial_y\omega^p \cdot\frac{\omega^R}{\partial_y\omega^p}dxdy=-\int_{{\mathcal S}}\partial_x\phi \triangle_{\varepsilon} \phi dxdy=\int_{{\mathcal S}}\partial_x|\nabla_{\varepsilon}\phi|^2 dxdy=0, \end{align*} $$where we use $\phi |_{y=0,1}=0.$ However, the boundary condition of $\phi $ is $\partial _y\phi |_{y=0,1}= C(t)$ , which brings an essential difficulty.
By the energy estimates, taking the inner product in $X^{r}$ (the definition is given in section 2) with $-\partial _t\phi ,$ we get
(1.12) $$ \begin{align} &\sup_{s\in[0,t]}(\lambda\|\nabla_{\varepsilon}\phi(s)\|_{X^{\frac73}}^2+\| \Delta_{\varepsilon}\phi(s)\|_{X^2}^2) \end{align} $$$$\begin{align*}\leq &\ C\int_0^t(\varepsilon^{-2}\|\varphi\Delta_{\varepsilon}\phi\|_{X^2}^2+\varepsilon^{-2}\| \nabla_{\varepsilon}\phi\|_{X^2}^2+\cdots)ds,\end{align*}$$where $\varphi (y)=y(1-y).$ All we need to do is to control $\varepsilon ^{-1}\|\varphi \Delta _{\varepsilon }\phi \|_{X^2}$ and $\varepsilon ^{-1}\|\nabla _{\varepsilon }\phi \|_{X^2}$ by the left-hand side of (1.12).
Motivated by [Reference Gérard-Varet, Maekawa and Masmoudi13], we expect to achieve that by a decomposition of stream function, $\phi =\phi _{slip}+\phi _{bc}$ in Gevrey $\frac 32$ regularity. Here, $\phi _{slip}$ enjoys a ‘good’ boundary condition and $\phi _{bc}$ is a corrector which recovers the boundary condition. In the following, we present the decomposition precisely.
-
(3) Gevrey estimate under artificial boundary conditions. $\phi _{slip}$ enjoying a good boundary condition is defined by
(1.13) $$ \begin{align} \left\{ \begin{aligned} &(\partial_t-\Delta_{\varepsilon})\omega_{slip}-\partial_x\phi_{slip}\partial_y \omega^p=\cdots,\qquad (x,y)\in\mathcal{S}\\ &\phi_{slip}|_{y=0,1}=0, \quad \omega_{slip}|_{y=0,1}=0,\qquad x\in\mathbb{T}, \end{aligned} \right. \end{align} $$where $\omega _{slip}=\Delta _{\varepsilon }\phi _{slip}.$ By ‘hydrostatic trick’ and Navier-slip boundary conditions, we obtain
(1.14) $$ \begin{align} \lambda\int_0^t (\|\omega_{slip}\|_{X^{\frac73}}^2+\|\nabla_{\varepsilon}\phi_{slip}\|_{X^{\frac73}}^2+|\nabla_{\varepsilon}\phi_{slip}|_{y=0,1}|_{X^{\frac73}}^2)ds\leq \frac{C}{\lambda}\int_0^t\| \varepsilon\Delta_{\varepsilon}\phi \|_{X^{2}}^2 ds+\cdots. \end{align} $$The full study of the Orr-Sommerfeld formulation (1.13) with Navier-slip boundary conditions is given in Section 7.
-
(4) Recovery the non-slip boundary condition. In Step (3), we use the slip boundary condition, not the real boundary condition $\partial _y\phi |_{y=0, 1}=C(t).$ To recover the boundary condition, we introduce the following system:
(1.15) $$ \begin{align} \left\{ \begin{aligned} &(\partial_t-\Delta_{\varepsilon})\Delta_{\varepsilon}\phi_{bc}-\partial_x\phi_{bc}\partial_y \omega^p=0, \qquad (x,y)\in\mathcal{S}\\ &\phi_{bc}|_{y=0,1}=0, \quad \partial_y\phi_{bc}|_{y=i}=h^i,\qquad x\in\mathbb{T}, \end{aligned} \right. \end{align} $$where $\omega _{bc}=\Delta _{\varepsilon } \phi _{bc}$ and $i=0,1.$ And we need to choose a suitable $h^i$ such that
$$ \begin{align*} \partial_y \phi_{bc}|_{y=0,1}=-\partial_y\phi_{slip}|_{y=0,1}+C(t). \end{align*} $$Next, we give the main idea for proving the existence of $h^i$ :
-
1. We define $\phi _{bc,S}=\phi _{bc,S}^0+\phi _{bc,S}^1$ , where $\phi _{bc,S}^i$ solve
(1.16) $$ \begin{align} \left\{ \begin{aligned} &(\partial_t -\Delta_{\varepsilon})\Delta_{\varepsilon}\phi^i_{bc, S}=0,\\ &\phi^i_{bc, S}|_{y=i}=0,\quad \partial_y\phi^i_{bc, S}|_{y=i}=h^i,\\ &\phi^i_{bc, S}|_{t=0}=0, \end{aligned} \right. \end{align} $$where $x\in \mathbb {T}$ , $y\in (0,+\infty )$ for $i=0$ and $y\in (-\infty ,1)$ for $i=1.$ Taking Fourier transformation on t and x, we can write the precise expression of the solution to obtain the Gevrey estimate for $\phi _{bc,S}^i$ :
(1.17) $$ \begin{align} &\int_0^t\|\nabla_{\varepsilon}\phi^i_{bc,S}\|_{X^{\frac{5}{2}}_i}^2+\|\varphi^i\Delta_{\varepsilon}\phi^i_{bc,S}\|_{X^{\frac52}_i}^2 + \|\partial_x\phi^i_{bc,S}\|_{X^{\frac53}}^2ds\leq \frac{C}{\lambda^{\frac12}}\int_0^t|h^i|_{X^{\frac73}}^2ds, \end{align} $$where $\varphi ^0(y)=y,~\varphi ^1(y)=1-y.$ Compared with the decomposition in [Reference Wang, Wang and Zhang35], we get more regularity of $\partial _x\phi ^i_{bc,S}$ , which is a key point to get the optimal Gevrey regularity. The details for this step are given in Section 8.1.
-
2. We correct the nonlocal term constructed in the above step by considering the following equations:
(1.18) $$ \begin{align} \left\{ \begin{aligned} &(\partial_t -\Delta_{\varepsilon})\Delta_{\varepsilon}\phi^i_{bc, R}-\partial_x\phi_{bc,R}^i\partial_y\omega^p=\partial_x\phi_{bc,S}^i\partial_y\omega^p,\quad (x,y)\in {\mathcal S}\\ &\phi^i_{bc, R}|_{t=0}=0, \quad (x,y)\in {\mathcal S} \end{aligned} \right. \end{align} $$with Navier-slip conditions. By the same process as Step (3) and combining with the sharp estimate (1.17) to get an estimate for $\phi ^i_{bc,R},$
(1.19) $$ \begin{align} \lambda\int_0^t \|\omega_{bc,R}^i\|_{X^{\frac73}}^2ds&+\int_0^t (\|\nabla_{\varepsilon}\phi_{bc,R}^i\|_{X^{\frac73}}^2+|\partial_y\phi_{bc,R}^i|_{y=0,1}|_{X^{\frac73}}^2)ds \end{align} $$$$\begin{align*}\leq&\frac{C}{\lambda^{\frac12}} \int_0^t|h^i|_{X^{\frac73}}^2ds+\cdots,\nonumber \qquad t\in[0, T]. \end{align*}$$More details are given in Section 8.3.
-
3. We define $\phi _{bc}=\phi _{bc,S}+\phi _{bc,R}$ , where $\phi _{bc,S}=\sum _{i=0, 1}\phi ^i_{bc,S} $ and $\phi _{bc,R}=\sum _{i=0, 1}\phi ^i_{bc,R}$ , which solves system (1.15). To match the boundary condition on the derivative of $\partial _y\phi |_{y=0,1}=C(t)$ , we need
$$ \begin{align*} \partial_y\phi_{bc,S}|_{y=0,1}+\partial_y\phi_{bc,R}|_{y=0,1}=\partial_y\phi_{bc}|_{y=0,1}=- \partial_y\phi_{slip}|_{y=0,1}+C(t). \end{align*} $$On one hand, $\phi _{bc,S}$ and $\phi _{bc,R}$ are defined by $h^i$ . We define a $0$ -order operator $R_{bc}$ given in (8.71) such that
$$ \begin{align*} (1+R_{bc}) h^i= - \partial_y\phi_{slip}|_{y=0,1}+C(t). \end{align*} $$Moreover, by the estimate in Step 1 and Step 2, we can get
$$ \begin{align*} \int_0^t\Big|R_{bc}[h^0, h^1]\Big|_{X^{\frac73}}^2 ds\leq&\frac{C}{\lambda^{\frac12}}\int_0^t|(h^0,h^1)|_{X^{\frac 73}}^2ds, \end{align*} $$which means that $(1+R_{bc})$ is an invertible operator when $\lambda $ is large. That means that $\phi _{bc,S}$ and $\phi _{bc,R}$ are well-defined and (1.15) is well-posedness. Details are given in Section 8.4.
Due to the transport terms, we need to introduce a new auxiliary function $\phi _{bc, T}$ between Step 1 and Step 2. For more details, see Section 8.2.
-
-
(5) Close the energy estimates (1.12). Summing estimates (1.17) and (1.19) in Step (4), we get an estimate for $\phi _{bc}:$
$$ \begin{align*} \int_0^t\|\nabla_{\varepsilon}\phi_{bc}\|_{X^{\frac{7}{3}}}^2+\|\varphi\Delta_{\varepsilon}\phi_{bc}\|_{X^{\frac 73}}^2ds\leq &\ \frac{C}{\lambda^{\frac12}}\int_0^t|(h^0, h^1)|_{X^{\frac73}}^2ds\\ \leq &\ C\int_0^t |\nabla_{\varepsilon}\phi_{slip}|_{y=0,1}|^2_{X^{\frac73}}ds+\cdots, \end{align*} $$which along with (1.14), we have$$ \begin{align*} \int_0^t (\|\varphi\omega\|_{X^{\frac73}}^2+\|\nabla_{\varepsilon}\phi\|_{X^{\frac73}}^2)ds\leq \frac{C}{\lambda}\int_0^t\| \varepsilon\Delta_{\varepsilon}\phi \|_{X^{2}}^2 ds+\cdots, \end{align*} $$and then we put the above estimate into (1.12) to close the estimate for system (1.11).
1.4 Notations
-
– $\mathcal {S}^{\varepsilon }=\left \{(x, y) \in \mathbb {T}\times \mathbb {R} : 0<y<\varepsilon \right \}$ and $\mathcal {S}=\left \{(x, y) \in \mathbb {T}\times \mathbb {R} : 0<y<1\right \}.$
-
– $\nabla _{\varepsilon }=(\varepsilon \partial _x, \partial _y)$ and $\Delta _{\varepsilon }=\varepsilon ^2 \partial _x^2+\partial _y^2$ .
-
– Vorticity of Prandtl part $\omega ^p$ is defined by $ \omega ^p=\partial _y u^p$ .
-
– Vorticity of reminder part $\omega ^R=\Delta _{\varepsilon } \phi $ is defined by $\omega ^R=\varepsilon ^2\partial _x v^R-\partial _y u^R$ . In this paper, we also define $\omega ^i_{bc, j}=\Delta _{\varepsilon } \phi _{bc, j}^i$ , where $i=0, 1$ and $j\in \{R, T\}$
-
– Cut-off functions $\varphi (y)=y(1-y)$ and $\varphi ^i(y)=i+(-1)^iy$ .
-
– $C(t)=\frac 1{2\pi }\int _{{\mathcal S}} u^Rdxdy$ .
-
– The Fourier transform of $f_{\Phi }$ is defined by $e^{(1-\lambda t)\langle k\rangle ^{\frac 23}} \widehat {f}(k)$ .
2 Gevrey norms and preliminary lemmas
At the beginning of this section, we define the functional space $X^r$ and the Gevrey class. First, we define
where $\tau (t)\ge 0$ . Moreover, it is easy to get that $\Phi (t,k)$ satisfies the subadditive inequality
Now, we are in the position to define $X^r_{\tau }$ , which is defined by
We say that a function f belongs to the Gevrey class $\frac 32$ if $\|f\|_{X^r_{\tau }}<+\infty $ .
Moreover, we need to deal with some Gevrey class functions defined on the boundary. Thus, we introduce the following functional space:
where f depends on variable x.
By the definition of $X^r_{\tau }$ , it is easy to see that if $r'\geq r,$ then $\|\cdot \|_{X^{r'}_{\tau }}\geq \|\cdot \|_{X^{r}_{\tau }}.$ For simplicity, we drop subscript $\tau $ in the notations $\|f\|_{X^r_{\tau }}, |f|_{X^r_{\tau }}$ etc. In the sequel, we always take
with $\lambda \geq 1$ determined later. Thus, if we take t small enough, we have $\tau>0$ .
In the following, we present some lemmas on product estimates in the Gevrey class and the readers can refer to Lemmas 2.1–2.3 in [Reference Wang, Wang and Zhang34] for details. The first lemma is the commutator estimate in Sobolev space:
Lemma 2.1. Let $r\ge 0,~s_1>\frac 32$ , $s>\frac 12$ and $0\leq \delta \leq 1$ . Then it holds that
In the Gevrey class, we have the following:
Lemma 2.2. Let $r\ge 0$ and $s> \frac 12$ . Then it holds that
For the commutator in the Gevrey class, we have the following:
Lemma 2.3. Let $r\ge 0,~s_1>\frac 32$ , $s>\frac 12$ and $0\leq \delta \leq 1$ . Then it holds that
3 Approximate equations and Error equations
3.1 Approximate equations
By the Hilbert asymptotic method, we can obtain the approximate solutions. We define approximate solutions as the following:
where $(u_p^0, v_p^0, p_p^0)$ satisfies equation (1.4) and $(u_p^2, v_p^2, p_p^2)$ satisfies equation
We point here that $(u_p^1, v_p^1, p_p^1)=0$ by matching the equation of order $\varepsilon $ . Based on the equation of $(u_p^0, v_p^0, p_p^0)$ and $(u_p^2, v_p^2, p_p^2)$ , we deduce the approximate solution $(u^p, v^p, p^p)$ which satisfies the following equation:
where reminder $(R_1, R_2)$ is given by
By the definition of $R_1$ and $R_2$ , it is easy to get that
3.2 Equations of error functions
We define error functions $(u^R, v^R, p^R)$ :
It is easy to deduce the system of error functions:
For convenience, we rewrite (3.6) as
Here, $(\mathcal {N}_u,\mathcal {N}_v)$ is a nonlinear term given by
Based on the above system, we get the equations of the vorticity $\omega ^R=\partial _y u^R-\varepsilon ^2 \partial _x v^R$ :
where $f_1, f_2$ are defined by
Moreover, following the calculations in [Reference Wang, Wang and Zhang34], we can obtain the boundary conditions of $\omega ^R$ :
where
3.3 Equations of stream function
Thanks to $\partial _x u^R+\partial _yv^R=0$ and $v^R|_{y=0,1}=0$ , there exists a stream function $\phi $ satisfying the following system:
Since $\int _{\mathbb {T}} v^Rdx=0,$ the function $\phi $ is periodic in x. Thanks to $\partial _x\phi |_{y=0,1}=0$ and $\phi (1,x)-\phi (0,x)=0$ , we may assume that $\phi |_{y=0,1}=0$ . Thus, there holds that
Taking (3.18) and (3.19) into (3.9) and using the boundary condition $(u^R, v^R)|_{y=0,1}=0$ , we obtain
where $C(t)=\frac 1{2\pi }\int _{{\mathcal S}} u^Rdxdy$ and $(\mathcal {N}_u, \mathcal {N}_v), ~f_1, ~f_2$ are given in (3.8), (3.10) and (3.11).
At the end of the subsection, we state some elliptic estimates which can be obtained by classical theory. First, by elliptic estimate and Hardy inequality, we have
where $\varphi (y)=y(1-y)$ and $\nabla _{\varepsilon }=(\partial _y,\varepsilon \partial _x).$ Indeed, taking the $L^2$ inner product with $-\phi $ in (3.19) and integrating by parts, we use boundary condition $\phi |_{y={0,1}}=0$ and the Hardy inequality to have
Since $(u^R,v^R)$ satisfies the elliptic equations
we arrive at
for any $r\geq 0.$
4 Estimate of $\nabla _{\varepsilon }\phi $ and $\Delta _{\varepsilon }\phi $ in the Gevrey space
Before giving the estimate of $\nabla _{\varepsilon }\phi $ and $\Delta _{\varepsilon }\phi $ , we need the estimates of the reminder terms $R_1$ and $R_2$ which are defined by the approximate solution $u^p$ and $v^p$ . For $(u^p, v^p),$ we have the following bound:
Lemma 4.1. Let initial data $u_0$ of (1.4) satisfy (1.6)-(1.9). There exists a time $T_p$ such that $(u_p^i, v_p^i),~i=0,2$ defined in (1.4) and (3.2) have the following estimates:
for $t\in [0, T_p]$ .
Moreover, according to (3.1), it holds that
and
Proof. Here, the key of this lemma is to prove that (1.4) is well-posedness in the Gevrey class $\frac 32$ which is the conjecture in [Reference Gérard-Varet, Masmoudi and Vicol15]. If we set $\varepsilon =0$ and follow the step-by-step process in this paper, we can get the conjecture proved. Here, to avoid repeatability, we leave the proof to the readers.
Then, by the definition of $(R_1, R_2)$ in (3.4)-(3.5), using Lemma 2.2, we get the following:
Lemma 4.2. It holds that
Now, we state our main result in this section:
Proposition 4.3. There exist $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ and $\lambda _0\geq 1,$ such that for any $t\in [0,T]$ and $\lambda \geq \lambda _0,$ it holds that
where $\Delta _{\varepsilon }\phi =\omega ^R$ , $\varphi (y)=y(1-y)$ and C is a constant independent of $\varepsilon $ .
Proof. Acting $e^{\Phi (t, D_x)}$ on the both sides of the first equation of (3.20), we get
Taking $H^{2,0}$ inner product with $-\partial _t\phi _{\Phi }$ and using boundary conditions
we integrate by parts to arrive at
First, let’s estimate $I_i,~i=1,\cdots , 5$ term by term.
$\underline {\mathrm {Estimate \ of} \ I_1.}$ Since divergence free condition $\partial _x u^p+\partial _y v^p=0,$ we get
According to $(u^p, v^p)|_{y=0,1}=0,$ we use integration by parts and Lemma 2.2 to have
$\underline {\mathrm {Estimate \ of} \ I_2.}$ Similarly, we write
Then, along with $\phi |_{y=0,1}=0$ , we use integration by parts and Lemma 2.2 to deduce
$\underline {\mathrm {Estimate \ of} \ I_3.}$ Due to $\phi |_{y=0,1}=0,$ taking integration by parts, it yields that
$\underline {\mathrm {Estimate \ of} \ I_4.}$ Recall $f_1$ and $f_2$ in (3.10)-(3.11). According to (3.22) and Lemma 4.2, we have
$\underline {\mathrm {Estimate \ of} \ I_5.}$ Poincaré inequality implies
for $\phi |_{y=0,1}=0.$ Since
we get
Collecting $I_1-I_5$ together, it holds that
Next, we focus on the boundary term $\Big \langle \Delta _{\varepsilon }\phi _{\Phi },\partial _t\partial _y\phi _{\Phi }\Big \rangle _{H^{2}_x}\Big |_{y=0}^{y=1}$ . First, we give the estimate of $C'(t)$ :
which gives
Owing to
where $\delta (k)$ is a Dirac function and $k\in \mathbb {Z},$ we have
where we used (4.3) in the last step.
Putting the above estimate and (4.2) into (4.1), we get
Next, we give the estimates of $\|\partial _y\omega ^R\|^2_{L^2}$ . First, we recall the equation of $\omega ^R$ :
with boundary conditions
where $f_1,~f_2, f$ and $\mathcal {N}$ are given in (3.10)-(3.16).
Taking the $L^2$ inner product with $\omega ^R$ on (4.5) and integration by parts, it follows from $(\mathcal {N}_u,\varepsilon \mathcal {N}_v)|_{y=0,1}=0$ and $(u^p, v^p)|_{y=0,1}=0$ to obtain
For the boundary term, we use (4.6)-(4.7) to write
Let $y_0\in [0,1]$ so that
Then, along with Gagliardo-Nirenberg inequality
it infers that
Similarly, we use (4.9) and $|C(t)|\leq C\|u^R\|_{L^2}\leq C\|\omega ^R\|_{L^2}$ to have
All we have left is to do $B_3.$ With the fact that operator $\partial _y(\triangle _{\varepsilon ,D})^{-1},~\partial _y(\triangle _{\varepsilon ,D})^{-1}(\partial _y,\varepsilon \partial _x)$ and $\partial _y^2(\triangle _{\varepsilon ,D})^{-1}$ are bounded from $L^2\to L^2,$ we have
According to the definition of (3.16) and (3.15), we have
and
which give that
Summarizing $B_1-B_3$ together, we obtain
Substituting (4.10) into (4.8), we take $\varepsilon $ small enough to arrive at
Bring the above estimate into (4.4) and integrate time from $0$ to t to get the desired results.
5 Sketch the proof to Theorem 1.1
In this section, we shall sketch the proof of Theorem 1.1. In the paper, we use the continued argument. Here, we define
5.1 The key a priori estimates
In this subsection, we shall present the key a priori estimates used in the proof of Theorem 1.1.
By Proposition 4.3, we need the estimates of $\int _0^t (\|\nabla _{\varepsilon }\phi \|_{X^2}^2+\|\varphi \Delta _{\varepsilon }\phi \|_{X^2}^2)ds$ to close the energy.
Proposition 5.1. Let $\phi $ be the solution of (3.20). Then there exists $\lambda _0\geq 1$ and $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ such that for $\lambda \geq \lambda _0$ and $t\in [0, T]$ , it holds that
with $t\in [0, T]$ .
The proof of the above proposition is the main part of this paper, and we prove it in Section 6.
5.2 Proof of Theorem 1.1
Before we prove Theorem 1.1, we first give the estimates for the nonlinear terms:
Proposition 5.2. Under the assumption (5.1), there holds that
where $t\in [0, T^*]$ .
Proof. By the definition of $\mathcal {N}_u,$ we have
It follows from Lemma 2.2 and (3.22) that
where we use the Gagliardo-Nirenberg inequality (4.9) in the second step.
Similarly, we use Lemma 2.2 and (3.22) to deduce
where we use $v^R=-\int _0^y \partial _x u^R dy'$ in the second step.
Collecting $I_1$ and $I_2$ together and using (5.1), it holds that
The estimate for $\varepsilon \mathcal {N}_v$ is obtained by changing $u^R$ into $\varepsilon v^R$ in the above argument and we omit details. Thus, we obtain (5.3).
For (5.4), we use the definition of $\mathcal {N}$ to have
With Proposition 5.1 and Proposition 5.2 in hand, we are in the position to prove Theorem 1.1. By Proposition 4.3, Proposition 5.1 and Proposition 5.2, we get
for $t\in [0,T].$ By the Gronwall inequality and choosing a small $T<\min \{T_p, \frac {1}{2\lambda }\}$ , we get that
By the Sobolev embedding theorem and Lemma 4.1, we get Theorem 1.1 proved.
6 The proof of Proposition 5.1
All we have left is to prove Proposition 5.1. To prove that, we first give the following decomposition of $\phi $ :
where $\phi _{slip}$ satisfies that
and $\phi _{bc}$ satisfies that
To prove Proposition 5.1, we need the estimates of $\phi _{slip}$ and $\phi _{bc}$ . First, we notice that $\phi _{slip}$ has a good boundary condition. We use the ‘hydrostatic trick’ method to get its estimates. The proof of the following proposition is given in Section 7.
Proposition 6.1. There exists $\lambda _0>1$ and $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ such that for all $t\in [0,T]$ , $\lambda \geq \lambda _0$ , there holds that
The estimates of $\phi _{bc}$ are much more difficult. Here, we state the main results on it:
Proposition 6.2. There exists $\lambda _0>1$ and $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ such that for all $t\in [0,T]$ , $\lambda \geq \lambda _0$ , there holds that
where C is a universal constant.
The proof of Proposition 6.2 is given in Section 8.
Based on the above two propositions, we are in the position to prove Proposition 5.1. First, we give the estimates of $\|u^R\|_{L^2}$ which are used to control the $C(t)$ .
Lemma 6.3. There exist $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ and $\lambda _0\geq 1$ such that for $t\in [0,T]$ and $\lambda \geq \lambda _0,$ it holds that
Remark 6.4. We use weighted quantity $\|e^{(1-\lambda t)}(u^R,\varepsilon v^R)\|_{L^2}$ instead of $\|(u^R,\varepsilon v^R)\|_{L^2}$ to obtain a small constant factor in front of $\int _0^t\|\nabla _{\varepsilon }\phi \|_{X^2}^2ds$ in (6.5).
Proof. Taking the $L^2$ inner product with $e^{2(1-\lambda t)}u^R$ in the first equation of (3.7) and with $e^{2(1-\lambda t)}v^R$ in the second equation of (3.7), we use the fact
and integrate by parts by boundary condition $(u^R, v^R)|_{y=0,1}=0$ to yield
where we write $v^R=-\int _0^y \partial _x u^Rdy'$ and use the fact $\partial _xu^p+\partial _y v^p=0 $ to eliminate the transport term and $\partial _xu^R+\partial _y v^R=0 $ to eliminate the pressure term, respectively.
Afterwards, integrating time from $0$ to t and using $\partial _x u^R=-\partial _x \partial _y\phi $ , we obtain
Finally, we use $\|e^{(1-\lambda s)}\partial _x\partial _y\phi \|_{L^2}\leq C\|\nabla _{\varepsilon }\phi \|_{X^2}$ to complete the proof.
Proof of Proposition 5.1
Now, we give the proof Proposition 5.1. We divide this proof into two parts.
$\underline {\mathrm {Estimates \ of } \ \int _0^t \|\nabla _{\varepsilon }\phi \|_{X^2}^2.}$ Since
by Lemma 6.3, we ensure
By the definition of $f_1$ and $f_2$ , we obtain that
and we get
Then, it follows $\phi =\phi _{slip}+\phi _{bc}$ and (6.4) to deduce
Plusing (6.6) and above estimates together and taking $\lambda $ large enough, we get
$\underline {\mathrm {Estimates \ of } \ \int _0^t\|\varphi \triangle _{\varepsilon }\phi \|_{X^2}^2.}$
It follows from (6.7) and (6.9) that
Applying Proposition 6.2 again, we get
Combining the above two estimates, we get
By now, we get the desired results.
7 Vorticity estimates under artificial boundary condition: Proof of Proposition 6.1
In this section, we give the proof of Proposition 6.1. To simplify the notation, we drop the subscript in the system (6.2):
The goal in this section is to establish a uniform (in $\varepsilon $ ) estimate of vorticity $\omega =\triangle _{\varepsilon }\phi .$
Proposition 7.1. There exists $\lambda _0>0$ and $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ such that for all $t\in [0,T]$ , $\lambda \geq \lambda _0$ , the following holds:
Proof. By Lemma 4.1, we have
Hence, we use the ‘hydrostatic trick’ to get the desired results. First, acting operator $e^{\Phi (t,D_x)}$ on the first equation of (6.2), we get
In view of (7.2), the terrible term comes from $\partial _x\phi _{\Phi }\partial _y \omega ^p,$ which loses one tangential derivative. In order to overcome the derivative loss, we take $\langle D_x\rangle ^2$ on the (7.2) and then take the $L^2$ inner product with $\frac {\langle D_x\rangle ^2\omega _{\Phi }}{\partial _y\omega ^p}$ to obtain that
The boundary term is zero due to artificial boundary condition $\omega |_{y=0,1}=\Delta _{\varepsilon }\phi |_{y=0,1}=0.$ Integrating on $[0,t)$ with $t\leq T$ and using $\partial _y\omega ^p\geq c_0$ , we obtain
Now, we estimate $T^i, i=0,\cdots , 10$ one by one.
$\underline {\mathrm {Estimate \ of } \ T^0 \ \mathrm {and} \ T^1.}$ Since $\partial _y\omega ^p\geq c_0>0$ and Lemma 4.1 imply
it is easy to see
$\underline {\mathrm {Estimate \ of} \ T^2 \ \mathrm {and} \ T^4.}$ By using Lemma 2.1, we get
where we used the Poincaré inequality and $\phi |_{y=0,1}=0$ to ensure
in the last step.
According to
classical elliptic estimate and (7.3) imply
which gives
Therefore, it follows from $\partial _y\omega ^p\geq c_0>0$ to get
$\underline {\mathrm {Estimate \ of} \ T_3.}$ Using Lemma 2.2 and (7.5), it shows
$\underline {\mathrm {Estimate \ of} \ T^.}$ This term is a troubling term because it loses one tangential derivative. However, the hydrostatic trick implies
by using $\phi |_{y=0,1}=0.$
$\underline {\mathrm {Estimate \ of} \ T^6, T^7 \ \mathrm {and}\ T^8.}$ Let’s estimate commutators by Lemma 2.3. Since $\partial _y\omega ^p\geq c_0>0,$ we use Lemma 2.3 to ensure that
Here, we use (7.3) and (7.5) in the last estimate.
$\underline {\mathrm {Estimate \ of} \ T^9 \ \mbox {and} \ T^{10}.}$ Integration by parts and boundary condition $\omega |_{y=0,1}=0$ give that
However, using the Hölder inequality, we get
Collecting $T^0-T^{10}$ together, we finally obtain
Taking $\lambda $ large enough, we deduce
However, (7.5) gives
Calderon-Zygmund inequality and Gagliardo-Nirenberg inequality (4.9) imply
8 Construction of the boundary corrector: Proof of Proposition 6.2
In the previous section, we construct a solution to the Orr-Sommerfeld equation with artificial boundary conditions: we replace condition $\partial _y\phi |_{y=0,1}=0$ by $\Delta _{\varepsilon }\phi |_{y=0,1}=0.$ To go back to the original system, we need to correct Neumann’s condition. Thus, we define $\phi _{bc}$ which satisfies the following system:
To estimate $\phi _{bc}$ , we use the following decomposition:
The definitions and estimates of $\phi _{bc, S}, \phi _{bc, T}$ and $\phi _{bc, R}$ are given in the following subsections.
8.1 The estimates of $\phi _{bc, S}$ : Stokes equation
In this subsection, we deal with $\phi _{bc, S}$ .
Because of two boundaries $y=0$ and $y=1$ , we define
where $\phi ^0_{bc, S}$ satisfies the following Stokes equation:
and $\phi ^1_{bc, S}$ satisfies the following Stokes equation:
where $t\in [0,T]$ . Here, $(h^0,h^1)$ is a given boundary data satisfying $(h^0(t), h^1(t))=0$ for $t=0$ and $t\geq T.$ Here, we point out that $h^i$ is defined by
where the operator $\mathcal {A}$ is a zero-order operator which is defined later.
In the following, we only give the process for $\phi ^0_{bc, S}.$ The case of $\phi ^1_{bc, S}$ is almost the same and we leave details to readers.
At first, we give zero extension of $\phi ^0_{bc, S}$ and $h^0$ with $t\leq 0$ such that we can take a Fourier transform in $t.$ Let $\widehat {\phi ^0_{bc, S}}=\widehat {\phi ^0_{bc, S}}(\zeta , k, y)$ be the Fourier transform of $\phi ^0_{bc, S}$ on x and t. Then $\widehat {(\phi ^0_{bc, S})_{\Phi }}$ satisfies the ODE:
where $\zeta \in \mathbb {R}$ and $k\in \mathbb {Z}.$ Assuming the decay of $(|k|\phi ^0_{bc, S},\partial _y\phi ^0_{bc, S})$ and the boundedness of $\partial _y\phi ^0_{bc, S},$ we obtain the formula
where the square root is taken so that the real part is positive, and it follows that
This inequality will be used frequently. It is easy to calculate that
The formula (8.8) will be used in estimating velocity and (8.9) will be used in estimating vorticity. With the same process above, we get the formula for $\widehat {(\phi ^1_{bc, S})_{\Phi }}:$
with $\gamma $ given in (8.6). It is easy to see
Remark 8.1. For $\varepsilon =0$ in (8.2), $\triangle _0=\partial _y^2.$
solves (8.4) with $\varepsilon =0$ and $\widehat {(\phi ^0_{bc, S})_{\Phi }}$ holds $\lim _{y\to +\infty }=\frac {\widehat {h^0_{\Phi }}}{\gamma _0}.$ Though $\widehat {(\phi ^0_{bc, S})_{\Phi }}$ does not tend to zero as y tends to infinity, the solution $\widehat {(\phi ^0_{bc, S})_{\Phi }}$ is only used to correct the boundary condition near $y=0$ , and we do not care about its value at infinity. It is easy to deduce
and we find these two terms decay to zero as y tends to infinity. By the same method, we can get another solution near $y=1:$
These constructions are the main difference between $\varepsilon =0$ and $\varepsilon \neq 0,$ but they enjoy the same properties stated below.
Lemma 8.2. Let $\phi ^i_{bc,S}$ be the solution of (8.2). It holds that
where $i=0,1$ and $L^2_{\zeta , y}=l^2_{\zeta }(L^2_{y}(0,+\infty ))$ for $i=0$ and $L^2_{\zeta , y}=l^2_{\zeta }(L^2_y(-\infty ,1))$ for $i=1.$
It is also held that
where $i=0,1$ and $L^2_{\zeta , y}=l^2_{\zeta }(L^2_{y}(0,1)).$
Proof. We only give the proof for $i=0.$ The case $i=1$ is almost the same, and we omit details to readers.
(8.16) follows from (8.5), (8.8) and the Plancherel theorem by observing the estimate for multipliers
The estimate (8.18) is a direct consequence of
For (8.19), we divide it into two cases: 1. $\varepsilon |k|\leq \frac 12\lambda ^{\frac 12}{\langle k\rangle }^{\frac 13}$ , and 2. $\varepsilon |k|\geq \frac 12\lambda ^{\frac 12}{\langle k\rangle }^{\frac 13}.$ In case 1,
which implies
In case 2, we use the bound
for ${\mathbf {Re}}(z)>0,$ which implies that
Combining cases 1–2, we complete (8.19), which yields (8.16). The estimate (8.17) is proved by using (8.5), Placherel theorem and
Indeed, note that the integral interval is $y\in (0,1)$ and we also divide it into $\varepsilon |k|\leq \frac 12\lambda ^{\frac 12}{\langle k\rangle }^{\frac 13}$ and $\varepsilon |k|\geq \frac 12\lambda ^{\frac 12}{\langle k\rangle }^{\frac 13}.$ When $\varepsilon |k|\geq \frac 12\lambda ^{\frac 12}{\langle k\rangle }^{\frac 13},$ the similar argument above gives that
When $\varepsilon |k|\leq \frac 12\lambda ^{\frac 12}{\langle k\rangle }^{\frac 13}$ ( with $\varepsilon |k|\ll 1$ ), we compute as
The finite interval $(0,1)$ is essential here. Thus, we complete this lemma.
To express this clearly, we introduce norms related to $y>0$ and $y<1$ , respectively. For any function f, we define
where $I_0=(0,+\infty )$ and $I_1=(-\infty ,1).$ It is obvious to see $\|\cdot \|_{X^r}\leq \|\cdot \|_{X^r_i}$ for any $i=0,1.$ Using Lemma 8.2 above, we can deduce the estimate for $\nabla _{\varepsilon }\phi ^i,$ where $i=0,1.$
Proposition 8.3. Let $\phi ^i_{bc,S}$ be the solution of (8.2). It holds that
Next, we give the estimate to the boundary term $\phi ^0_{bc,S}|_{y=1}$ and $\phi ^1_{bc,S}|_{y=0.}$
Lemma 8.4. For any $M\geq 0$ and $i=0,1,$ it holds that
and
for any $r\geq 0.$
Proof. We only give the proof for the case $i=0$ ; the case $i=1$ is similar and we omit details to readers. Taking $y=1$ in (8.5) and using
we get
However, we refer to (8.8) and take $y=1$ in it by noticing
for any $N\geq 0$ , and combining with (8.28) to deduce
Thus, we finish our proof.
At the end of this subsection, we give some weight estimates of vorticity $\omega ^i_{bc,S}=\Delta _{\varepsilon }\phi ^i_{bc,S}$ . Denote
Proposition 8.5. It holds that
As a consequence, we get for $\theta '\in [-\frac 12,2]$
Proof. The result is obtained by using formula (8.6), (8.9) and (8.12), the Plancherel theorem and by observing that multiplier $\varphi ^i(y)$ gains $\frac {1}{\lambda ^{\frac 12}{\langle k\rangle }^{\frac 13}}$ . More precisely,
Thus, we complete the proof.
Based on the above proposition, we have more estimates on $\omega ^i_{bc,S}$ :
Proposition 8.6. Let $\theta \in [0,2].$ It holds that
Proof. (8.32) is a direct result of Lemma 8.5 by taking $\theta =0.$ It is easy to check
by taking $\theta '=\theta +\frac 12$ in Lemma 8.5 and completing (8.33). Similarly, we check
by taking $\theta '=\theta -\frac 12$ in Lemma 8.5 to completing (8.34).
8.2 The estimates of $\phi _{bc, T}$ : vorticity transport estimate
$\phi _{bc, T}$ is defined by
where $\phi ^0_{bc, T}$ is defined by
and $\phi ^1_{bc, T}$ is defined by
We need to emphasize that we extend $(u^p, v^p)$ to $y\in \mathbb {R}$ by zero, which means that $(u^p, v^p)=0$ when $y\in \mathbb {R}\setminus [0,1].$
Before we give the estimates of $\phi ^i_{bc, T}$ , we use Proposition 8.6 and $(u^p, v^p)|_{y=0,1}=0$ to get that
where $\theta =0,1,2$ .
We are in the position to give the estimates of $\phi ^i_{bc,T}$ :
Proposition 8.7. Let $\theta =0,1,2$ and $i=0,1$ . There exists $\lambda _0>1$ and $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ such that for all $t\in [0,T]$ , $\lambda \geq \lambda _0$ , it holds that
where $\Delta _{\varepsilon }\phi ^i_{bc,T}=\omega ^i_{bc,T}$ and $\varphi ^i$ is given in (8.30).
Proof. Acting $e^{\Phi (t,D_x)}$ on the first equation of (8.35), we obtain
Then, taking the $L^2_y(I_i;H^{\frac {11}{6}+\frac {\theta }3}_x)$ inner product with $(\varphi ^i)^{2\theta }(\omega ^i_{bc,T})_{\Phi }$ , we get by using $\omega ^i_{bc,T}|_{y=i}=0,~\partial _xu^p+\partial _y v^p=0$ and integrating by parts that
where $\mathcal {S}_i=\mathbb {T}\times I_i.$ Integrating on $[0,t)$ with $t\leq T$ , we obtain
Now, we estimate $I_j^i,~j=1,\cdots ,6$ term by term.
$\underline {\mathrm {Estimate \ of} \ I_1^i.}$ It follows from Lemma 2.1 that
which deduces that
$\underline {\mathrm {Estimate \ of} \ I_2^i.}$ Thanks to
by using $v^p|_{y=i}=0,$ it is obvious to see
$\underline {\mathrm {Estimate \ of} \ I_3^i.}$ Applying Lemma 2.3, we find
$\underline {\mathrm {Estimate \ of} \ I_4^i.}$ Applying Lemma 2.2, we get
$\underline {\mathrm {Estimate \ of} \ I_5^i.}$ By the fact
we have
$\underline {\mathrm {Estimate \ of} \ I_6^i.}$ It follows from
for $\theta =1,2$ and
by using Hardy inequality for $\theta =0.$ Therefore, we get for $\theta =0,1,2$ that
Putting $I_1^i-I_6^i$ together, we have
Then we insert them into (8.39) and take $\lambda $ large enough to obtain
where we use (8.37) in the last step.
All we have left to do is to control the last term of the above inequality. For that, we rewrite it as follows:
Combing all the above estimates, we get the desired results.
Based on estimates of $\omega _{bc, T}^i$ , we use the elliptic equation to get the estimates of $\phi ^i_{bc,T}$ .
Corollary 8.8. There exists $\lambda _0>1$ and $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ such that for all $t\in [0,T]$ , $\lambda \geq \lambda _0$ , it holds that
Proof. Here, we only give the proof of the case $i=0$ . The case $i=1$ is the same.
We recall the elliptic equation
for $y>0$ . Then we take the $X^{r}_0$ inner product with $\phi ^0_{bc,T}$ and use the Hardy inequality to get
which implies
for $r\geq 0.$ By Proposition 8.7, we get
For the boundary term, using the interpolation inequality, we get
where we use (8.43) and the Calderon-Zygmund inequality in the last step. Along with Proposition 8.7, we arrive at
Next, we deal with the term $\|\partial _x\phi ^0_{bc,T}\|_{X^{\frac {5}{3}}}$ . Taking the Fourier transform in x to (8.42), we write the solution
Then we have
Decomposing the integral $\int _0^y$ into $\int _0^{\min \{y,{\langle k\rangle }^{-\frac 13}\}}$ and $\int _{\min \{y,{\langle k\rangle }^{-\frac 13}\}}^y$ , it follows from the Hölder inequality that
We take summation $\sum _{k\in \mathbb {Z}}$ and use the Plancherel theorem to deduce
Thus, we get that
Collecting (8.44), (8.45) and (8.48) together, we get the corollary proved.
8.3 The estimates of $\phi _{bc, R}$ : full construction of boundary corrector
All we have left is the term $\phi _{bc, R}$ . Like the previous argument, we define
where $\phi ^i_{bc, R}$ satisfies that
and
For simplicity, denote $\omega _{bc,R}^i=\Delta _{\varepsilon } \phi ^i_{bc,R}$ , which has the following relationship:
and
where
To homogenize boundary conditions, we introduce
where $\widetilde {\phi }_{bc,R}^0$ satisfies
where
Similarly, we introduce
and $\widetilde {\phi }_2^1$ satisfies
where
First, we give some elliptic estimates.
Lemma 8.9. Let $(f^0, f^1)$ , $(g^0, g^1)$ – introduced in (8.53)–(8.54), (8.55) and (8.58). It holds that
for $i=0,1.$
Moreover, $\phi _{bc,R}^i~(i=0,1)$ has the following estimates:
Proof. Here we only prove the case $i=0.$ The case $i=1$ is almost the same, and we omit details to readers. We first give the proof for $f^0$ . By the definition of $f^0$ , we get
where we used Lemma 8.4 and Corollary 8.8.
For $g^0$ , by Corollary 8.8, Proposition 8.3, we have
On one hand, using Proposition 8.3, Proposition 8.5, Corollary 8.8 and Proposition 8.7, we get
which implies that
At last, we prove (8.62). Taking the $X^{\frac 73}$ inner product with $\widetilde {\phi }_{bc,R}^0$ toward (8.56), we use integration by parts and then integrate time from $0$ to t that
Due to $ \widetilde {\phi }_{bc,R}^0|_{y=0,1}=0,$ we use the Poincaré inequality to imply
According to (8.61), we get
Combining (8.63) and (8.64), we deduce
Bringing $\nabla _{\varepsilon }\phi _{bc,R}^0=\nabla _{\varepsilon }\widetilde {\phi }_{bc,R}^0-\nabla _{\varepsilon } g^0$ into the above inequality, we obtain
By now, we finish the proof.
In order to estimate the right-hand side of (8.49) and (8.50) and the boundary term, we need the following results:
Lemma 8.10. For $i=0,1,$ we have that
Proof. Here, we only prove the case $i=0.$ The case $i=1$ is almost the same, and we omit details to readers.
By Proposition 8.3 and Corollary 8.8, we get (8.65) proved.
Next, we deal with the boundary term. A direct calculation gives that
and
due to $\phi _{bc, S}^0|_{y=0}=\phi _{bc, T}^0|_{y=0}=0.$
By Corollary 8.8, we get
where we use an elliptic estimate and the Calderon-Zygmund inequality
For the last estimate, we use (8.66) and (8.61) to imply
Here, we complete this lemma.
We are coming to the main part of this section. We shall give the estimate for the system (8.49) and (8.50).
Proposition 8.11. Let $\phi _{bc,R}^0$ and $\phi _{bc,R}^1$ be the solution of (8.49) and (8.50), respectively, and $\omega _{bc,R}^i=\Delta _{\varepsilon } \phi _{bc,R}^i$ for $i=0,1.$ Then, for every $i=0,1$ , it holds that
where $0<T<\min \{T_p, \frac {1}{2\lambda }\}$ .
Proof. The result mainly comes from the process of Proposition 7.1. Here we take $(\mathcal {N}_u, \varepsilon \mathcal {N}_v)=0$ and $\varepsilon ^2 f_1+f_2-C(t)\partial _x\omega ^p$ is replaced by $G^i=-\partial _y(\phi _{bc,S}^i+\phi _{bc,T}^i)\partial _x\omega ^p+\partial _x(\phi _{bc,S}^i+\phi _{bc,T}^i)\partial _y\omega ^p$ for $i=0,1.$ In order to estimate the source term $\int _0^t \|G^i\|_{X^{\frac 53}}^2ds$ , using Lemma 8.10 and product estimate in Lemma 2.2, we get
The only difference comes from the boundary conditions
which are not zero compared with equation (6.2). We review $T^5$ in Propoposition 7.1. After integration by parts, the boundary term is left. More precisely, we need to estimate $\int _0^t \Big \langle \partial _x(\phi _{bc,R}^i)_{\Phi },\partial _y(\phi _{bc,R}^i)_{\Phi } \Big \rangle _{H^2_x}\Big |_{y=0}^{y=1}ds.$ According to Lemma 8.10, we have
Here, we take $\lambda $ large enough to complete the proof.
8.4 Proof of Proposition 6.2
In this subsection, we combine all the above estimates to finish the proof of Proposition 6.2. Recalling the definition of $\phi _{bc}$ :
we get that
Here, $R_{bc}^{ji}~ (j=0,1,~i=0,1)$ are linear operators and are defined by
Compared with the system (8.1), we need to find $(h^0, h^1)$ such that
hold. To do that, we define an operator $R_{bc}[h^0, h^1],$ which is defined by
which is a $2\times 2$ matrix operator and is well-defined on the Banach space
Proposition 8.12. There exists $\lambda _0\geq 1$ such that if $\lambda \geq \lambda _0$ , the map $R_{bc}: Z_{bc}\to Z_{bc}$ defined by (8.71) satisfies
Hence, the operator $I+R_{bc}$ is invertible in $Z_{bc}$ . Moreover, there exists $(h_0, h_1)\in Z_{bc}$ such that (8.70) holds and $(h_0, h_1)$ is defined by
Proof. First, by Lemma 8.4, Proposition 8.8, Corollary 8.7 and Proposition 8.11, it is easy to get
Taking $\lambda $ large enough, we get that the operator $I+R_{bc}$ is invertible in $Z_{bc}$ . Thus, there exists $(h_0, h_1)\in Z_{bc}$ such that (8.70) holds.
Let’s continue to prove Proposition 6.2. According to Proposition 8.3, Proposition 8.6, Corollary 8.8 and Proposition 8.11, we get by (8.68) that
and
which imply
Due to Proposition 8.12 and taking $\mathcal {A}=(I+R_{bc})^{-1},$ we know $\mathcal {A}$ is a zero-order bounded operator in $Z_{bc}$ and obtain
which finishes this proposition.
Acknowledgements
The authors would like to thank Professor Zhifei Zhang for the valuable discussions and suggestions. C. Wang is partially supported by the NSF of China under Grant 12071008. Y. Wang is partially supported by NSF of China under Grant 12101431 and the Sichuan Youth Science and Technology Foundation No. 21CXTD0076.
Competing Interests
The author declares none.