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

CONSTRAINED LARGE SOLUTIONS TO LERAY’S PROBLEM IN A DISTORTED STRIP WITH THE NAVIER-SLIP BOUNDARY CONDITION

Zijin Li, Xinghong Pan and Jiaqi Yang
Abstract

In this paper, we solve the Leray’s problem for the stationary Navier-Stokes system in a 2D infinite distorted strip with the Navier-slip boundary condition. The existence, uniqueness, regularity and asymptotic behavior of the solution are investigated. Moreover, we discuss how the friction coefficient affects the well-posedness of the solution. Due to the validity of the Korn’s inequality, all constants in each a priori estimate are independent of the friction coefficient. Thus our method is also valid for the total-slip and no-slip cases. The main novelty is that the total flux of the velocity can be relatively large (proportional to the slip length) when the friction coefficient of the Navier-slip boundary condition is small, which is essentially different from the 3D case.

Keywords: Stationary Navier-Stokes system, Navier-slip boundary condition, Leray’s problem, Distorted strip.

Mathematical Subject Classification 2020: 35Q35, 76D05

1  Introduction

Consider the Navier-Stokes equations

{𝒖𝒖+pΔ𝒖=0,𝒖=0,in𝒮2,\left\{\begin{aligned} &\boldsymbol{u}\cdot\nabla\boldsymbol{u}+\nabla p-\Delta\boldsymbol{u}=0,\\ &\nabla\cdot\boldsymbol{u}=0,\end{aligned}\right.\quad\text{in}\quad\mathcal{S}\subset{\mathbb{R}}^{2}, (1.1)

subject to the Navier-slip boundary condition:

{2(𝕊𝒖𝒏)tan+α𝒖tan=0,𝒖𝒏=0,on𝒮.\left\{\begin{aligned} &2(\mathbb{S}\boldsymbol{u}\cdot\boldsymbol{n})_{\mathrm{tan}}+\alpha\boldsymbol{u}_{\mathrm{tan}}=0,\\ &\boldsymbol{u}\cdot\boldsymbol{n}=0,\\ \end{aligned}\right.\quad\text{on}\quad\partial\mathcal{S}. (1.2)

Here 𝕊𝒖=12(𝒖+T𝒖)\mathbb{S}\boldsymbol{u}=\frac{1}{2}\left(\nabla\boldsymbol{u}+\nabla^{T}\boldsymbol{u}\right) is the stress tensor, where T𝒖\nabla^{T}\boldsymbol{u} represent the transpose of 𝒖\nabla\boldsymbol{u}, and 𝒏\boldsymbol{n} is the unit outer normal vector of 𝒮\partial\mathcal{S}. For a vector field 𝒗\boldsymbol{v}, we denote 𝒗tan\boldsymbol{v}_{\mathrm{tan}} its tangential part:

𝒗tan:=𝒗(𝒗𝒏)𝒏,\boldsymbol{v}_{\mathrm{tan}}:=\boldsymbol{v}-(\boldsymbol{v}\cdot\boldsymbol{n})\boldsymbol{n},

and α0\alpha\geq 0 in (1.2) stands for the friction coefficient which may depend on various elements, such as the property of the boundary and the viscosity of the fluid. When α0+\alpha\to 0_{+}, the boundary condition (1.2) turns to be the total Navier-slip boundary condition, while when α\alpha\to\infty, the boundary condition (1.2) degenerates into the no-slip boundary condition u0u\equiv 0. In this paper, we consider the general case, which assumes 0α+0\leq\alpha\leq+\infty.

The domain 𝒮\mathcal{S} is a two dimensional infinitely distorted smooth strip as follows.

Refer to caption
Figure 1: The infinitely distorted strip 𝒮\mathcal{S}.

Here, 𝒮R\mathcal{S}_{R} and 𝒮L\mathcal{S}_{L} are two semi-infinitely long straight strips. In the cartesian coordinate system x1Ox2x_{1}Ox_{2}, the strip

𝒮R:={x2:(x1,x2)(0,)×(0,1)},\mathcal{S}_{R}:=\{x\in\mathbb{R}^{2}\,:\,(x_{1},x_{2})\in(0,\infty)\times(0,1)\},

while in the cartesian coordinate system y1Oy2y_{1}O^{\prime}y_{2}, the stripe

𝒮L:={y2:(y1,y2)(,0)×(0,c0)}.\mathcal{S}_{L}:=\{y\in\mathbb{R}^{2}\,:\,(y_{1},y_{2})\in(-\infty,0)\times(0,c_{0})\}.

Here c0c_{0} is a fixed constant. They are smoothly connected by the compact distorted part 𝒮0\mathcal{S}_{0} in the middle. We do not insist the domain 𝒮\mathcal{S} to be simply connected, but all obstacles, with their boundaries are smooth Jordan curves, must lie in 𝒮0\mathcal{S}_{0}, and keep away from upper and lower boundaries of 𝒮\mathcal{S}, i.e. 𝒮+\partial\mathcal{S}_{+} and 𝒮\partial\mathcal{S}_{-}, respectively.

Before stating our main results, we give some notations for later convenience.

Notations

Throughout this paper, Ca,b,c,C_{a,b,c,...} denotes a positive constant depending on a,b,c,a,\,b,\,c,\,..., which may be different from line to line. For simplicity, a constant C𝒮C_{\mathcal{S}}, depending on geometry properties of 𝒮\mathcal{S}, is usually abbreviated by CC. Dependence on 𝒮\mathcal{S} is default unless independence of 𝒮\mathcal{S} is particularly stated. We use 𝒆1,𝒆2\boldsymbol{e}_{1},\,\boldsymbol{e}_{2} to denote the unit standard basis in the cartesian coordinate system x1Ox2x_{1}Ox_{2}, and 𝒆1,𝒆2\boldsymbol{e}_{1}^{\prime},\,\boldsymbol{e}_{2}^{\prime} to denote the unit standard basis in the cartesian coordinate system y1Oy2y_{1}O^{\prime}y_{2}. Meanwhile, for any ζ>1\zeta>1, 𝒮R,ζ=(0,ζ)×(0,1)\mathcal{S}_{R,\zeta}=(0,\zeta)\times(0,1) in the cartesian coordinate system x1Ox2x_{1}Ox_{2}, and 𝒮L,ζ=(ζ,0)×(0,c0)\mathcal{S}_{L,\zeta}=(-\zeta,0)\times(0,c_{0}) in the cartesian coordinate system y1Oy2y_{1}O^{\prime}y_{2}. Then the truncated strip is defined by:

𝒮ζ:=𝒮L,ζ𝒮0𝒮R,ζ.\mathcal{S}_{\zeta}:=\mathcal{S}_{L,\zeta}\cup\mathcal{S}_{0}\cup\mathcal{S}_{R,\zeta}. (1.3)

We use Υζ±\Upsilon^{\pm}_{\zeta} to denote the right and left part of 𝒮ζ\𝒮ζ1\mathcal{S}_{\zeta}\backslash\mathcal{S}_{\zeta-1} as follows:

Υζ+:=𝒮R,ζ\𝒮R,ζ1,Υζ:=𝒮L,ζ\𝒮L,ζ1.\Upsilon^{+}_{\zeta}:=\mathcal{S}_{R,\zeta}\backslash\mathcal{S}_{R,\zeta-1},\quad\quad\Upsilon^{-}_{\zeta}:=\mathcal{S}_{L,\zeta}\backslash\mathcal{S}_{L,\zeta-1}.

We also apply ABA\lesssim B to state ACBA\leq CB. Moreover, ABA\simeq B means both ABA\lesssim B and BAB\lesssim A.

For 1p1\leq p\leq\infty and kk\in\mathbb{N}, LpL^{p} denotes the usual Lebesgue space with norm

fLp(D):={(D|f(x)|p𝑑x)1/p,1p<,esssupxD|f(x)|,p=,\|f\|_{L^{p}(D)}:=\left\{\begin{aligned} &\left(\int_{D}|f(x)|^{p}dx\right)^{1/p},\quad&1\leq p<\infty,\\[8.53581pt] &\mathrm{esssup}_{x\in D}|f(x)|,\quad&p=\infty,\\ \end{aligned}\right.

while Wk,pW^{k,p} denotes the usual Sobolev space with its norm

fWk,p(D):=0|L|kLfLp(D),\begin{split}\|f\|_{W^{k,p}(D)}:=&\sum_{0\leq|L|\leq k}\|\nabla^{L}f\|_{L^{p}(D)},\\ \end{split}

where L=(l1,l2)L=(l_{1},l_{2}) is a multi-index. We also simply denote Wk,pW^{k,p} by HkH^{k} provided p=2p=2. Finally, D¯\overline{D} denote the closure of a domain DD. A function gWlock,p(D)g\in W^{k,p}_{\mathrm{loc}}(D) or Wlock,p(D¯)W^{k,p}_{\mathrm{loc}}(\overline{D}) function means gWk,p(D~)g\in W^{k,p}(\tilde{D}), for any D~\tilde{D} compactly contained in DD or D¯\overline{D}.

For the 2D vector-valued function, we define

(𝒮):={𝝋H1(𝒮;2):𝝋𝒏|𝒮=0},σ(𝒮):={𝝋H1(𝒮;2):𝝋=0,𝝋𝒏|𝒮=0},\begin{split}\mathcal{H}(\mathcal{S})&:=\left\{\boldsymbol{\varphi}\in H^{1}({\mathcal{S}};{\mathbb{R}}^{2}):\,\boldsymbol{\varphi}\cdot\boldsymbol{n}\big{|}_{\partial\mathcal{S}}=0\right\},\\ \mathcal{H}_{\sigma}({\mathcal{S}})&:=\left\{\boldsymbol{\varphi}\in H^{1}({\mathcal{S}};\,\mathbb{R}^{2}):\,\nabla\cdot\boldsymbol{\varphi}=0,\ \boldsymbol{\varphi}\cdot\boldsymbol{n}\big{|}_{\partial\mathcal{S}}=0\right\},\end{split}

and

σ,loc(𝒮¯):={𝝋Hloc1(𝒮¯;2):𝝋=0,𝝋𝒏|𝒮=0}.\mathcal{H}_{\sigma,{\mathrm{loc}}}(\overline{\mathcal{S}}):=\left\{\boldsymbol{\varphi}\in H^{1}_{{\mathrm{loc}}}(\overline{\mathcal{S}};\,\mathbb{R}^{2}):\,\nabla\cdot\boldsymbol{\varphi}=0,\ \boldsymbol{\varphi}\cdot\boldsymbol{n}\big{|}_{\partial\mathcal{S}}=0\right\}.

We also denote

𝑿:={𝝋Cc(𝒮¯;2):𝝋=0,𝝋𝒏|𝒮=0}.\boldsymbol{X}:=\big{\{}\boldsymbol{\varphi}\in C_{c}^{\infty}(\overline{\mathcal{S}}\,;\,\mathbb{R}^{2}):\ \nabla\cdot\boldsymbol{\varphi}=0,\ \boldsymbol{\varphi}\cdot\boldsymbol{n}\big{|}_{\partial\mathcal{S}}=0\big{\}}.

Clearly, 𝑿\boldsymbol{X} is dense in σ\mathcal{H}_{\sigma} in H1(𝒮)H^{1}(\mathcal{S}) norm. For matrices 𝚪=(γij)1i,j2\boldsymbol{\Gamma}=(\gamma_{ij})_{1\leq i,j\leq 2} and 𝑲=(κij)1i,j2\boldsymbol{K}=(\kappa_{ij})_{1\leq i,j\leq 2}, we denote

𝚪:𝑲=i,j=12γijκij.\boldsymbol{\Gamma}:\boldsymbol{K}=\sum^{2}_{i,j=1}\gamma_{ij}\kappa_{ij}.

1.1 The generalized Leray’s problem in the distorted strip

For a given flux Φ\Phi which is supposed to be nonnegative without loss of generality, we consider Poiseuille-type flows, 𝑷ΦR=:PΦR(x2)𝒆1\boldsymbol{P}^{R}_{\Phi}=:{P}^{R}_{\Phi}(x_{2})\boldsymbol{e}_{1} and 𝑷ΦL=:PΦL(y2)𝒆1\boldsymbol{P}^{L}_{\Phi}=:{P}^{L}_{\Phi}(y_{2})\boldsymbol{e}_{1}^{\prime}, of (1.1) with the boundary condition (1.2) in 𝒮i\mathcal{S}_{i} (ii denotes RR or LL), then we will find that

{(PΦR(x2))′′=CR,in (0,1),(PΦR(0))=αPΦR(0),(PΦR(1))=αPΦR(1),01PΦR(x2)𝑑x2=Φ,\left\{\begin{aligned} &-\left(P^{R}_{\Phi}(x_{2})\right)^{\prime\prime}=C_{R},\quad\quad\quad\text{in }\quad(0,1),\\ &\left(P^{R}_{\Phi}(0)\right)^{\prime}=\alpha P^{R}_{\Phi}(0),\quad\left(P^{R}_{\Phi}(1)\right)^{\prime}=-\alpha P^{R}_{\Phi}(1),\\ &\int_{0}^{1}P^{R}_{\Phi}(x_{2})dx_{2}=\Phi,\\ \end{aligned}\right. (1.4)

and

{(PΦL(y2))′′=CL,in (0,c0),(PΦL(0))=αPΦL(0),(PΦL(c0))=αPΦL(c0),0c0PΦL(y2)𝑑y2=Φ,\left\{\begin{aligned} &-\left(P^{L}_{\Phi}(y_{2})\right)^{\prime\prime}=C_{L},\quad\quad\quad\text{in }\quad(0,c_{0}),\\ &\left(P^{L}_{\Phi}(0)\right)^{\prime}=\alpha P^{L}_{\Phi}(0),\quad\left(P^{L}_{\Phi}(c_{0})\right)^{\prime}=-\alpha P^{L}_{\Phi}(c_{0}),\\ &\int_{0}^{c_{0}}P^{L}_{\Phi}(y_{2})dy_{2}=\Phi,\\ \end{aligned}\right. (1.5)

where the constants CiC_{i} are uniquely related to Φ\Phi. Direct computation shows that

{PΦR(x2)=6Φ6+α[α(x22+x2)+1];PΦL(y2)=6Φc02(6+c0α)[α(y22+c0y2)+c0].\left\{\begin{split}P_{\Phi}^{R}(x_{2})&=\frac{6\Phi}{6+\alpha}\left[\alpha(-x^{2}_{2}+x_{2})+1\right];\\ P_{\Phi}^{L}(y_{2})&=\frac{6\Phi}{c_{0}^{2}\left(6+c_{0}\alpha\right)}\left[\alpha(-y^{2}_{2}+c_{0}y_{2})+c_{0}\right].\\ \end{split}\right. (1.6)

And

|(PΦi)|CαΦ1+α,i{L,R},\displaystyle|(P^{i}_{\Phi})^{\prime}|\leq C\frac{\alpha\Phi}{1+\alpha},\quad\forall i\in\{L,R\}, (1.7)

where C>0C>0 is a constant independent of both α\alpha and Φ\Phi.

The main objective of this paper is to study the solvability of the following generalized Leray’s problem: For a given flux Φ\Phi, to find a pair (𝒖,p)(\boldsymbol{u},p) such that

{𝒖𝒖+pΔ𝒖=0,𝒖=0,in𝒮,2(𝕊𝒖𝒏)tan+α𝒖tan=0,𝒖𝒏=0,on𝒮,\left\{\begin{aligned} &\boldsymbol{u}\cdot\nabla\boldsymbol{u}+\nabla p-\Delta\boldsymbol{u}=0,\quad\nabla\cdot\boldsymbol{u}=0,\quad\hskip 2.84544pt\text{in}\quad\mathcal{S},\\ &2(\mathbb{S}\boldsymbol{u}\cdot\boldsymbol{n})_{\mathrm{tan}}+\alpha\boldsymbol{u}_{\mathrm{tan}}=0,\quad\boldsymbol{u}\cdot\boldsymbol{n}=0,\quad\text{on}\quad\partial\mathcal{S},\\ \end{aligned}\right. (1.8)

with

01u1(x1,x2)𝑑x2=Φ,for anyx𝒮R\int^{1}_{0}u_{1}(x_{1},x_{2})dx_{2}=\Phi,\quad\text{for any}\quad x\in\mathcal{S}_{R} (1.9)

and

𝒖𝑷Φi,as|x| in 𝒮i.\boldsymbol{u}\rightarrow\boldsymbol{P}^{i}_{\Phi},\quad\text{as}\quad|x|\rightarrow\infty\ \text{ in }\ \mathcal{S}_{i}. (1.10)

To prove the existence of the above generalized Leray’s problem, we first introduce a weak formulation. Multiplying (1.8)1 with 𝝋𝑿\boldsymbol{\varphi}\in\boldsymbol{X} and integration by parts, by using the boundary condition (1.8)2, we can obtain

2𝒮𝕊𝒖:𝕊𝝋dx+α𝒮𝒖tan𝝋tan𝑑S+𝒮𝒖𝝋𝒖dx=0,for all 𝝋𝑿.\begin{split}&2\int_{\mathcal{S}}\mathbb{S}\boldsymbol{u}:\mathbb{S}\boldsymbol{\varphi}dx+\alpha\int_{\partial\mathcal{S}}\boldsymbol{u}_{\mathrm{tan}}\cdot\boldsymbol{\varphi}_{\mathrm{tan}}dS+\int_{\mathcal{S}}\boldsymbol{u}\cdot\nabla\boldsymbol{\varphi}\cdot\boldsymbol{u}dx=0,\quad\text{for all }\ \boldsymbol{\varphi}\in\boldsymbol{X}.\end{split} (1.11)

Now we define the weak solution of the generalized Leary’s problem:

Definition 1.1.

A vector 𝐮:𝒮2\boldsymbol{u}:\mathcal{S}\to{\mathbb{R}}^{2} is called a weak solution of the generalized Leray problem (1.8) to (1.10) if and only if

  • (i).

    𝒖σ,loc(𝒮¯)\boldsymbol{u}\in\mathcal{H}_{\sigma,{\mathrm{loc}}}(\overline{\mathcal{S}});

  • (ii).

    𝒖\boldsymbol{u} satisfies (1.11);

  • (iii).

    𝒖\boldsymbol{u} satisfies (1.9) in the trace sense;

  • (iv).

    𝒖𝑷ΦiH1(𝒮i)\boldsymbol{u}-\boldsymbol{P}^{i}_{\Phi}\in{H}^{1}(\mathcal{S}_{i}), for i=L,Ri=L,R.

Remark 1.2.

The weak solution also satisfies a generalized version of (1.10). Actually, it follows from the trace inequality ([10, Theorem II.4.1]) that for any x𝒮Rx\in\mathcal{S}_{R}:

01|𝒖(x1,x2)𝑷ΦR(x2)|2𝑑x2C𝒖𝑷ΦRH1([x1,+)×[0,1])2,x1>0,\displaystyle\int^{1}_{0}|\boldsymbol{u}(x_{1},x_{2})-\boldsymbol{P}^{R}_{\Phi}(x_{2})|^{2}dx_{2}\leq C\|\boldsymbol{u}-\boldsymbol{P}^{R}_{\Phi}\|^{2}_{H^{1}([x_{1},+\infty)\times[0,1])},\quad\forall x_{1}>0, (1.12)

where the constant CC is independent of x1x_{1}. Using (iv)(iv) in Definition 1.1, the estimate (1.12) implies that

01|𝒖(x1,x2)𝑷ΦR(x2)|2𝑑x20,as x1.\int^{1}_{0}|\boldsymbol{u}(x_{1},x_{2})-\boldsymbol{P}^{R}_{\Phi}(x_{2})|^{2}dx_{2}\to 0,\quad\text{as $x_{1}\to\infty$}.

The case in 𝒮L\mathcal{S}_{L} is similar.

The following result shows that for each weak solution we can associate a corresponding pressure field. See the proof in Section 3.2 below.

Lemma 1.3.

Let 𝐮\boldsymbol{u} be a weak solution to the generalized Leray’s problem defined above. Then there exists a scalar function pLloc2(𝒮¯)p\in L^{2}_{{\mathrm{loc}}}(\overline{\mathcal{S}}) such that

𝒮𝒖:𝝍dx+𝒮𝒖𝒖𝝍dx=𝒮p𝝍𝑑x\int_{\mathcal{S}}\nabla\boldsymbol{u}:\nabla\boldsymbol{\psi}dx+\int_{\mathcal{S}}\boldsymbol{u}\cdot\nabla\boldsymbol{u}\cdot\boldsymbol{\psi}dx=\int_{\mathcal{S}}p\nabla\cdot\boldsymbol{\psi}dx

holds for any 𝛙Cc(𝒮;2)\boldsymbol{\psi}\in C^{\infty}_{c}(\mathcal{S};{\mathbb{R}}^{2}).

1.2 Main results

Now we are ready to state the main theorems of this paper. The first one is existence and uniqueness of the weak solution, the second one addresses the regularity and decay estimates of the weak solution.

Theorem 1.4.

Let 0α+0\leq\alpha\leq+\infty be the friction coefficient given in (1.2). Assume that 𝒮\mathcal{S} is the aforementioned smooth strip. Then there exists C0>0C_{0}>0, independent of α\alpha, such that

  • (i)

    if

    αΦ1+αC0,\frac{\alpha\Phi}{1+\alpha}\leq C_{0}, (1.13)

    then the 2D generalized Leray’s problem (1.8)-(1.9)-(1.10) has a weak solution (𝒖,p)σ,loc1(𝒮¯)×Lloc2(𝒮¯)(\boldsymbol{u},p)\in\mathcal{H}^{1}_{\sigma,{\mathrm{loc}}}(\overline{\mathcal{S}})\times L^{2}_{{\mathrm{loc}}}(\overline{\mathcal{S}}) satisfying

    i=L,R𝒖𝑷ΦiH1(𝒮i)ΦeCΦ,\sum_{i=L,R}\|\boldsymbol{u}-\boldsymbol{P}^{i}_{\Phi}\|_{H^{1}(\mathcal{S}_{i})}\leq\Phi e^{C\Phi}, (1.14)

    for some constant CC independent of α\alpha.

  • (ii)

    Moreover, if 𝒖~\tilde{\boldsymbol{u}} is another weak solution of (1.8)-(1.9)-(1.10) with the flux ΦC0\Phi\leq C_{0}, and satisfies that for ζ\zeta\to\infty,

    𝒖~L2(𝒮ζ)=o(ζ3/2).\|\nabla\tilde{\boldsymbol{u}}\|_{L^{2}(\mathcal{S}_{\zeta})}=o\left(\zeta^{3/2}\right). (1.15)

    then 𝒖~𝒖\tilde{\boldsymbol{u}}\equiv\boldsymbol{u}.

Remark 1.5.

Here we give several remarks.

  • In the existence result (i), noticing that the flux at the cross section Φ\Phi can be relatively large when α<1\alpha<1 is small, since one only needs Φ2C0α1\Phi\leq 2C_{0}\alpha^{-1}. Here α=0\alpha=0 means the flux Φ\Phi can be arbitrarily large.

  • The limiting case α=0\alpha=0 (i.e., the total slip situation) has already been considered in [21], where an extra geometry restriction on the shape of the strip was imposed and the uniqueness was not considered there.

  • The limiting case α=+\alpha=+\infty corresponds to the famous Leray’s problem with the no-slip boundary condition which has been investigated for a long period of time. See a systematic review and study in [10, Chapter XIII].

  • From the uniqueness result in Theorem 1.4, we see that uniqueness can be only guaranteed by assuming that Φ\Phi is small enough, independent of the scale of α\alpha. Actually uniqueness of the weak solution is a more complicated problem than existence. See some discussion and non-uniqueness results in [10, Chapter XII.2] for the stationary 2D exterior problem.

The following Theorem gives the smoothness and the asymptotic behavior of a weak solution, which decays exponentially to the Poiseuille flow 𝑷Φi\boldsymbol{P}^{i}_{\Phi} at each 𝒮i\mathcal{S}_{i} as |x||x| tends to infinity. Only the partial smallness condition (1.13) is imposed.

Theorem 1.6.

Let 𝐮\boldsymbol{u} be a weak solution stated in the item (i)(i) of Theorem 1.4. Then

𝒖C(𝒮¯)\boldsymbol{u}\in C^{\infty}(\overline{\mathcal{S}})

such that: For any m0m\geq 0,

i=L,Rm(𝒖𝑷Φi)L2(𝒮i)+m𝒖L2(𝒮0)CΦ,m.\sum_{i=L,R}\|\nabla^{m}(\boldsymbol{u}-\boldsymbol{P}^{i}_{\Phi})\|_{L^{2}(\mathcal{S}_{i})}+\|\nabla^{m}\boldsymbol{u}\|_{L^{2}(\mathcal{S}_{0})}\leq C_{\Phi,m}. (1.16)

Meanwhile, the following pointwise decay estimates hold for sufficiently large |x||x|:

|m(𝒖𝑷ΦL)(y)|CΦ,meσΦ,my1,for ally1<<1;|m(𝒖𝑷ΦR)(x)|CΦ,meσΦ,mx1,for allx1>>1.\begin{split}|\nabla^{m}(\boldsymbol{u}-\boldsymbol{P}^{L}_{\Phi})(y)|&\leq C_{\Phi,m}e^{\sigma_{\Phi,m}y_{1}},\quad\text{for all}\quad y_{1}<<-1;\\ |\nabla^{m}(\boldsymbol{u}-\boldsymbol{P}^{R}_{\Phi})(x)|&\leq C_{\Phi,m}e^{-\sigma_{\Phi,m}x_{1}},\quad\text{for all}\quad x_{1}>>1.\end{split} (1.17)

Here CΦ,mC_{\Phi,m} and σΦ,m\sigma_{\Phi,m} are positive constants depending only on Φ\Phi and mm.

Also the corresponding pressure pp enjoys estimates akin to (1.16) and (1.17). See Remark 4.3.

Remark 1.7.

After our paper was posted on arXiv, we were informed by Professor Chunjing Xie that their group are also considering 2D Leray’s problem with Navier-slip boundary and two manuscripts on this topic are finished. We are grateful for their kindness of sending us their manuscripts. After checking their manuscripts, though there are partial overlaps of results, the proof between theirs and ours differs in many aspects. Since our two groups’ manuscripts are nearly posted at the same time, we believe that we independently solve this 2D Leray’s problem at almost the same time. Reader can refer their works in [25, 26] for more details.

1.3 Influence of the friction coefficient for the well-posedness

Unlike the 3D generalized Leray’s problem with the Navier-slip boundary condition in our recent work [19], the friction coefficient α\alpha plays an important role for the well-posedness in the 2D problem. Some interesting results different from the 3D problem are presented as follows:

  • (i).

    Largeness of the flux Φ\Phi when we show the existence, regularity and asymptotic behavior of the constructed H1H^{1} weak solution.

  • (ii).

    The α\alpha-independence of all the estimates in Theorem 1.4 and Theorem 1.6 indicates that our results are uniform with the friction coefficient α\alpha and can be applied to the limiting cases α=0\alpha=0 (total slip case) and α=\alpha=\infty (classical Leray’s problem).

The main reason behind the above improvements is the validity of the Korn-type inequality (L2L^{2} norm equivalence between 𝒗\nabla\boldsymbol{v} and 𝕊𝒗\mathbb{S}\boldsymbol{v}) in the 2D strip domain 𝒮\mathcal{S} as displayed in Lemma 2.6 and Corollary 2.7, which fails in the 3D pipe as Remark 2.8.

1.4 Difficulties, outline of the proof and related works

Difficulties and corresponding strategies

In two dimensional case, compared with the no-slip boundary condition, the main difficulties of the problem with Navier-slip boundary condition lie in the following:

  • (i).

    For a given flux, construction of a smooth solenoidal flux carrier, satisfying the Navier-slip boundary condition, and equalling to the Poiseuille flow at a large distance;

  • (ii).

    Achieving Poincaré-type inequalities and Korn-type inequalities in the distorted strip 𝒮\mathcal{S}.

In order to overcome difficulties listed above, our main strategies are as follows:

  • (i).

    In order to construct the flux carrier, we first introduce a uniform curvilinear coordinates transform near the boundary: T:x(s,t)T:\,x\rightarrow(s,t) to smoothly connect the two semi-infinite long straight strips and the middle distorted part. Under this curvilinear coordinates, a thin strip near the boundary of the middle distorted part is straightened. Under the curvilinear coordinates (s,t)(s,t), the flux carrier is constructed to smoothly connect the Poiseuille flows 𝑷ΦL\boldsymbol{P}_{\Phi}^{L} and 𝑷ΦR\boldsymbol{P}_{\Phi}^{R} at far field with a compact supported divergence-free vector σε(t)𝒆s\sigma^{\prime}_{\varepsilon}(t)\boldsymbol{e}_{s} in 𝒮0\mathcal{S}_{0}. In the intermediate parts, they can be glued smoothly, and the divergence-free property together with the Navier-slip boundary condition keep valid.

  • (ii).

    The Poincaré inequality and Korn’s inequality play important roles during the proof. For the no-slip boundary condition, Poincaré inequality can be applied directly by using zero boundary condition. However, in the case of the Navier-slip boundary condition, the Poincaré inequality is not obvious in the middle distorted part 𝒮0\mathcal{S}_{0}. First, in 𝒮L\mathcal{S}_{L} or 𝒮R\mathcal{S}_{R}, after subtracting the constant flux, 𝒗𝒆𝟏\boldsymbol{v}\cdot\boldsymbol{e_{1}} (or 𝒗𝒆𝟏\boldsymbol{v}\cdot\boldsymbol{e_{1}}^{\prime}) has zero mean value in any cross line, and then combining the impermeable boundary condition, which indicates that 𝒗𝒆𝟐=0\boldsymbol{v}\cdot\boldsymbol{e_{2}}=0 (or 𝒗𝒆𝟐=0\boldsymbol{v}\cdot\boldsymbol{e_{2}}^{\prime}=0) on the boundary, we can achieve Poincaré inequality in the straight strips. Based on the result in the straight strip, we derive the Poincaré inequality in 𝒮0\mathcal{S}_{0} by the trace theorem and a 2D Payne’s identity (2.30). See Lemma 2.4. The α\alpha-independence of constants during the proof of main theorems is creditable to Korn’s inequalities in 2D strips. The Korn’s inequality is proved via a contradiction argument, which is given in Section 2, and is highly dependent on the compactness of the curvature of the boundary. It is not valid in the 3D case. See a counterexample in Remark 2.8.

Outline of the proof

The existence and uniqueness of the solution will be given in Section 3. Before proving the existence theorem, a smooth solenoidal flux carrier 𝒂\boldsymbol{a} will be carefully constructed under the help of the uniformly curvilinear coordinates near the boundary 𝒮\partial\mathcal{S}_{-}. By subtracting this smooth flux carrier, the existence problem of (1.8)-(1.9)-(1.10) is reduced to a related one that the solution approaches zero at spacial infinity, which can be handled by the standard Galerkin method.

The main idea of proving the uniqueness is applying Lemma 2.11, which was originally announced in reference [17] and used to prove the uniqueness of the Leray’s problem with no-slip boundary. Although our idea originates from [17], there are many differences from the previous literature. Some estimates of the present manuscript is much more complicated, involving the Poincaré inequality in a distorted strip as shown in Lemma 2.4 and Korn’s inequality in Lemma 2.6.

Proofs of the asymptotic behavior and smoothness of weak solutions are given in Section 4. The main idea in deriving the exponential decay of H1H^{1} weak solutions is to derive a first order ordinary differential inequality for the L2L^{2} gradient integration in domain 𝒮\𝒮ζ\mathcal{S}\backslash\mathcal{S}_{\zeta}. For the global estimates of higher-order norms, by using a “decomposing-summarizing” technique, H1H^{1} estimate of the vorticity in 𝒮\mathcal{S} will be obtained, and then Biot-Savart law indicates H2H^{2} global estimate of the solution. Using the bootstrapping argument, higher-order global estimates then follow. This also leads to the higher-order exponential decay estimates, by utilizing the H1H^{1}-decay estimate and interpolation inequalities.

Some related works

The well-posedness study of the stationary Navier-Stokes equations in an infinite long pipe (or an infinite strip in the 2D case) with no-slip boundary condition and toward the Poiseuille flow laid down by Ladyzhenskaya in 1950s [15, 16], in which the problem was called Leray’s problem. Later by reducing the problem to the resolution of a variational problem, Amick [3, 4] obtained the existence result of the Leray’s problem with small flux, but the uniqueness was left open. For the planar flow, Amick–Fraenkel [5] studied the Leray’s problem in various types of stripes distinguished by their properties at infinity. An approach to solving the uniqueness of small-flux solution via energy estimate was built by Ladyzhenskaya-Solonnikov [17], in which authors also addressed the existence and asymptotic behavior results. See [2, 12, 24] for more related conclusion, also [10, Chapter XIII] for a systematic review to the Leray’s problem with the no-slip boundary condition. Recently Yang-Yin [32] studied the well-posedness of weak solutions to the steady non-Newtonian fluids in pipe-like domains. Wang-Xie in [28, 29] studied the existence, uniqueness and uniform structural stability of Poiseuille flows for the 3D axially symmetric inhomogeneous Navier-Stokes equations in the 3D regular cylinder, with a force term appearing on the right hand of the equations.

The Navier-slip boundary condition was initialed by Navier [23]. It allows fluid slip on the boundary with a scale proportional to its stress tensor. Different from the no-slip boundary, the Leray’s problem with the Navier-slip boundary condition requires much more complicated mathematical strategies. [21, 22, 14] studied the solvability of the steady Navier-Stokes equations with the perfect Navier-slip condition (α=0\alpha=0). In this case, the solution approaches to a constant vector at the spatial infinity. Authors in [6, 11, 1] studied the properties of solutions to the steady Navier-Stokes equations with the Navier-slip boundary in bounded domains. Wang and Xie [30] showed the uniqueness and uniform structural stability of Poiseuille flows in an infinite straight long pipe with the Navier-slip boundary condition. Authors of the present paper studied the related 3D Leray’s problem with the Naiver-slip boundary condition [19] under more strict smallness of the flux than the recent paper on 2D case. They also proved the characterization of bounded smooth solutions for the 3D axially symmetric Navier-Stokes equations with the perfect Navier-slip boundary condition in the infinitely long cylinder [18].

This paper is arranged as follows: In Section 2, some preliminary work are contained, in which a uniform curvilinear coordinate near the boundary will be introduced and the Navier-slip boundary condition will be written under this curvilinear coordinate frame, and some useful lemmas will be presented. We will concern the existence and uniqueness results in Section 3. Finally, we focus on the higher-order regularity and exponential decay properties of the solution in Section 4.

2  Preliminary

First, we introduce a uniformly curvilinear coordinate near the boundary, which will help to construct the flux carrier. This curvilinear coordinates can be viewed as the straightening of the boundary in the distorted part 𝒮0\mathcal{S}_{0}. Under this curvilinear coordinates, the Navier-slip boundary condition in (1.2) on the boundary of 𝒮0\mathcal{S}_{0} will share almost the same form as that in the semi-infinite straight part of 𝒮L\mathcal{S}_{L} and 𝒮R\mathcal{S}_{R}. See (2.27) below.

2.1 On the uniformly curvilinear coordinates near the boundary

To investigate the delicate feature of the Navier-slip boundary condition, also to construct the flux carrier in the distorted part of the strip, one needs to parameterize the boundary of 𝒮\mathcal{S}. Recalling that

𝒮=𝒮+𝒮𝒮Ob,\partial\mathcal{S}=\partial\mathcal{S}_{+}\cup\partial\mathcal{S}_{-}\cup\partial\mathcal{S}_{Ob},

where 𝒮±\partial\mathcal{S}_{\pm} are upper and lower boundary portions of 𝒮\mathcal{S}, while 𝒮Ob\partial\mathcal{S}_{Ob} denotes the union of boundaries of obstacles in the middle of the strip. For convenience, we only parameterize 𝒮\partial\mathcal{S}_{-} since the others are similar. Besides, under this parameterised curvilinear coordinates, we will construct the divergence-free flux carrier, which is supported in

𝒮(δ):={x𝒮¯:dist(x,𝒮)<δ},\mathcal{S}_{-}(\delta):=\{x\in\overline{\mathcal{S}}:\,\mathrm{dist}(x,\partial\mathcal{S}_{-})<\delta\},

for some suitably small δ\delta in the next section.

Refer to caption
Figure 2: A curvilinear coordinate system (s,t)(s,t) near the boundary portion 𝒮\partial\mathcal{S}_{-}.

Denoting

𝒮={𝖇(s)=(𝔟1(s),𝔟2(s))2:s},\partial\mathcal{S}_{-}=\{\boldsymbol{\mathfrak{b}}(s)=\left(\mathfrak{b}_{1}(s),\mathfrak{b}_{2}(s)\right)\in\mathbb{R}^{2}:\,s\in\mathbb{R}\}, (2.18)

where 𝔟1(s),𝔟2(s)\mathfrak{b}_{1}(s),\mathfrak{b}_{2}(s) are smooth functions of ss. Without loss of generality, we suppose the parameter ss\in\mathbb{R} being the arc length parameter of 𝒮\partial\mathcal{S}_{-}, so that

|𝖇(s)|=(𝔟1(s))2+(𝔟2(s))21,s.|\boldsymbol{\mathfrak{b}}(s)|=\sqrt{(\mathfrak{b}_{1}^{\prime}(s))^{2}+(\mathfrak{b}_{2}^{\prime}(s))^{2}}\equiv 1,\quad\forall s\in\mathbb{R}.

By the definition of 𝒮\mathcal{S} given in Section 1, 𝒮\partial\mathcal{S}_{-} lies on part of straight lines {y2=0}\{y_{2}=0\} or {x2=0}\{x_{2}=0\} except a compact distorted part in the middle, and there exists s0>0s_{0}>0 such that

𝒮𝒮0¯={𝖇(s)=(𝔟1(s),𝔟2(s))2:s[s0,s0]}.\partial\mathcal{S}_{-}\cap\overline{\mathcal{S}_{0}}=\{\boldsymbol{\mathfrak{b}}(s)=\left(\mathfrak{b}_{1}(s),\mathfrak{b}_{2}(s)\right)\in\mathbb{R}^{2}:\,s\in[-s_{0},s_{0}]\}. (2.19)

This indicates 𝖇(s0)=O\boldsymbol{\mathfrak{b}}(s_{0})=O and 𝖇(s0)=O\boldsymbol{\mathfrak{b}}(-s_{0})=O^{\prime}. Meanwhile, all “obstacles” inside 𝒮\mathcal{S} are away from it.

Because of the compact distortion, 𝒮\partial\mathcal{S}_{-} must satisfy the following condition:

Condition 2.1 (Uniform interior sphere).

For any point z𝒮z\in\partial\mathcal{S}_{-}, there exists a disk KzK_{z}, with its radius being RzR_{z}, such that

Kz¯(2𝒮)={z}.\overline{K_{z}}\cap(\mathbb{R}^{2}-\mathcal{S})=\{z\}.

Meanwhile, there exists δ>0\delta>0 such that

Rz2δ,z𝒮.R_{z}\geq 2\delta,\quad\forall z\in\partial\mathcal{S}_{-}. (2.20)

Due to the uniform interior sphere condition, for any x𝒮(δ)x\in\mathcal{S}_{-}(\delta), there exists a unique point z𝒮z\in\partial\mathcal{S}_{-} such that |xz|=dist(x,𝒮)|x-z|=\mathrm{dist}(x,\partial\mathcal{S}_{-}). Recalling (2.18), there exists a unique pair (s,t)×[0,δ)(s,t)\in\mathbb{R}\times[0,\delta) such that y=𝖇(s)y=\boldsymbol{\mathfrak{b}}(s) and t=dist(x,z)t=\mathrm{dist}(x,z). In this way, the following mapping is one-to-one and well-defined:

x(s,t),x𝒮(δ).x\to(s,t),\quad\forall x\in\mathcal{S}_{-}(\delta). (2.21)

Meanwhile, one has

Lemma 2.2.

The mapping defined in (2.21) is smooth.

Proof.

By the construction of this mapping, one deduces

x=ydist(x,z)𝒏z,x=y-\mathrm{dist}(x,z)\boldsymbol{n}_{z},

where 𝒏y\boldsymbol{n}_{y} is the unit outer normal of y𝒮y\in\partial\mathcal{S}_{-}\,. Since y=𝖇(s)y=\boldsymbol{\mathfrak{b}}(s), we define

𝑭(s,t):=𝖇(s)t𝒏𝖇(s)(s,t)×[0,δ).\boldsymbol{F}(s,t):=\boldsymbol{\mathfrak{b}}(s)-t\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)}\quad\forall(s,t)\in\mathbb{R}\times[0,\delta)\,\,.

Clearly 𝑭\boldsymbol{F} is well-defined and smooth in ×[0,δ)\mathbb{R}\times[0,\delta), and its Jacobian matrix writes

J𝑭=(𝔟1(s)t(dds𝒏𝖇(s))1(𝒏𝖇(s))1𝔟2(s)t(dds𝒏𝖇(s))2(𝒏𝖇(s))2).J\boldsymbol{F}=\left(\begin{array}[]{cc}\mathfrak{b}_{1}^{\prime}(s)-t\left(\frac{d}{ds}\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)}\right)_{1}&\quad-(\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)})_{1}\\[5.69054pt] \mathfrak{b}_{2}^{\prime}(s)-t\left(\frac{d}{ds}\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)}\right)_{2}&\quad-(\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)})_{2}\\ \end{array}\right).

Here and below, (𝑿)i(\boldsymbol{X})_{i} with i=1,2i=1,2 means the ii-th component of the vector 𝑿\boldsymbol{X}. Direct calculation shows

det(J𝑭)=(𝖇(s)t(dds𝒏𝖇(s)))𝒏𝖇(s)=1t(dds𝒏𝖇(s))𝖇(s),\mathrm{det}(J\boldsymbol{F})=\left(\boldsymbol{\mathfrak{b}}^{\prime}(s)-t\left(\frac{d}{ds}\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)}\right)\right)\cdot\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)}^{\perp}=1-t\left(\frac{d}{ds}\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)}\right)\cdot\boldsymbol{\mathfrak{b}}^{\prime}(s),

where

𝒏𝖇(s)=((𝒏𝖇(s))2,(𝒏𝖇(s))1)=𝖇(s).\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)}^{\perp}=(-(\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)})_{2},(\boldsymbol{n}_{\boldsymbol{\mathfrak{b}}(s)})_{1})=\boldsymbol{\mathfrak{b}}^{\prime}(s).

This indicates that

det(J𝑭)1t12δ>12,(s,t)×[0,δ)\mathrm{det}(J\boldsymbol{F})\geq 1-t\frac{1}{2\delta}>\frac{1}{2},\quad\forall(s,t)\in\mathbb{R}\times[0,\delta)

due to (2.20) so that 12δ\frac{1}{2\delta} can bound the curvature of 𝒮\partial\mathcal{S}_{-}\,. Recalling the compactness of the distorted part, the lemma is claimed by the inverse mapping theorem. ∎

For any x=(x1,x2)𝒮(δ)x=(x_{1},x_{2})\in\mathcal{S}_{-}(\delta), Condition 2.1 and Lemma 2.2 above guarantee the following well-defined curvilinear coordinate system

(s,t)=(s(x),t(x))×[0,δ).(s,t)=(s(x),t(x))\in\mathbb{R}\times[0,\delta).

Geometrically, t(x)t(x) is the distance of the given point x𝒮(δ)x\in\mathcal{S}_{-}(\delta) to the boundary 𝒮\partial\mathcal{S}_{-}, while s(x)s(x) denotes the parameter coordinate of the unique point y𝒮y\in\partial\mathcal{S}_{-} such that |xy|=dist(x,𝒮)|x-y|=\mathrm{dist}(x,\partial\mathcal{S}_{-}). As it is shown in Figure 2, we denote

𝒆𝒔=(es1,es2)and𝒆𝒕=(et1,et2)\boldsymbol{e_{s}}=(e_{s1}\,,\,e_{s2})\quad\text{and}\quad\boldsymbol{e_{t}}=(e_{t1}\,,\,e_{t2}) (2.22)

are unit tangent vector of ss-curves and tt-curves, respectively. Meanwhile, they are all independent with variable t[0,δ)t\in[0,\delta). Clearly

{𝒆𝒔𝒆𝟏;𝒆𝒕𝒆𝟐,x𝒮R¯;{𝒆𝒔𝒆𝟏;𝒆𝒕𝒆𝟐,x𝒮L¯.\left\{\begin{aligned} \boldsymbol{e_{s}}&\equiv\boldsymbol{e_{1}};\\ \boldsymbol{e_{t}}&\equiv\boldsymbol{e_{2}},\\ \end{aligned}\right.\quad\quad\forall x\in\overline{\mathcal{S}_{R}};\quad\quad\left\{\begin{aligned} \boldsymbol{e_{s}}&\equiv\boldsymbol{e_{1}}^{\prime};\\ \boldsymbol{e_{t}}&\equiv\boldsymbol{e_{2}}^{\prime},\\ \end{aligned}\right.\quad\quad\forall x\in\overline{\mathcal{S}_{L}}. (2.23)

Moreover,

xt=(tx1,tx2)𝒆𝒕,\nabla_{x}t=\left(\frac{\partial t}{\partial x_{1}}\,,\,\frac{\partial t}{\partial x_{2}}\right)\equiv\boldsymbol{e_{t}}, (2.24)

and there exists a smooth function γ(s,t)>γ0>0\gamma(s,t)>\gamma_{0}>0 that

xs=(sx1,sx2)=γ𝒆𝒔.\nabla_{x}s=\left(\frac{\partial s}{\partial x_{1}}\,,\,\frac{\partial s}{\partial x_{2}}\right)=\gamma\boldsymbol{e_{s}}. (2.25)

Thus by denoting

𝑫:=(sx1sx2tx1tx2),\boldsymbol{D}:=\left(\begin{array}[]{cc}\frac{\partial s}{\partial x_{1}}&\frac{\partial s}{\partial x_{2}}\\[5.69054pt] \frac{\partial t}{\partial x_{1}}&\frac{\partial t}{\partial x_{2}}\\ \end{array}\right),

one derives

det𝑫=γ𝒆𝒔𝒆𝒕=γ.\mathrm{det}\,\boldsymbol{D}=-\gamma\boldsymbol{e_{s}}\cdot\boldsymbol{e_{t}}^{\perp}=\gamma.

by (2.24) and (2.25). Moreover, by calculating the inverse matrix of 𝑫\boldsymbol{D}, one deduces

(x1sx1tx2sx2t)=(1γet2es21γet1es1),\left(\begin{array}[]{cc}\frac{\partial x_{1}}{\partial s}&\frac{\partial x_{1}}{\partial t}\\[5.69054pt] \frac{\partial x_{2}}{\partial s}&\frac{\partial x_{2}}{\partial t}\\ \end{array}\right)=\left(\begin{array}[]{cc}\frac{1}{\gamma}e_{t2}&-e_{s2}\\[5.69054pt] -\frac{1}{\gamma}e_{t1}&e_{s1}\\ \end{array}\right),

which indicates

{xs=𝒆𝒔γ;xt=𝒆𝒕.\left\{\begin{aligned} &\frac{\partial x}{\partial s}=\frac{\boldsymbol{e_{s}}}{\gamma};\\ &\frac{\partial x}{\partial t}=\boldsymbol{e_{t}}.\\ \end{aligned}\right.

Since |𝒆𝒔|=|𝒆𝒕|1|\boldsymbol{e_{s}}|=|\boldsymbol{e_{t}}|\equiv 1 and 𝒆𝒔𝒆𝒕0\boldsymbol{e_{s}}\cdot\boldsymbol{e_{t}}\equiv 0, there exists a bounded smooth function κ=κ(s,t)\kappa=\kappa(s,t)\in\mathbb{R}, which denotes the curvature of the boundary, such that

{d𝒆𝒕ds=κ𝒆𝒔γ;d𝒆𝒔ds=κ𝒆𝒕γ,(s,t)×[0,δ).\left\{\begin{split}\frac{d\boldsymbol{e_{t}}}{ds}&=-\frac{\kappa\boldsymbol{e_{s}}}{\gamma};\\[5.69054pt] \frac{d\boldsymbol{e_{s}}}{ds}&=\frac{\kappa\boldsymbol{e_{t}}}{\gamma},\\ \end{split}\right.\quad\quad\forall(s,t)\in\mathbb{R}\times[0,\delta).

By direct calculation, the divergence and curl of a vector field 𝒘=ws(s,t)𝒆𝒔+wt(s,t)𝒆𝒕\boldsymbol{w}=w_{s}(s,t)\boldsymbol{e_{s}}+w_{t}(s,t)\boldsymbol{e_{t}} writes

div𝒘=γsws+twtκwt;curl𝒘=x2w1x1w2=twsγswtκws,\begin{split}\mathrm{div}\,\boldsymbol{w}&=\gamma\partial_{s}w_{s}+\partial_{t}w_{t}-\kappa w_{t};\\ \mathrm{curl}\,\boldsymbol{w}&=\partial_{x_{2}}w_{1}-\partial_{x_{1}}w_{2}=\partial_{t}w_{s}-\gamma\partial_{s}w_{t}-\kappa w_{s},\end{split} (2.26)

under this curvilinear coordinates.

To finish this subsection, let us focus on the Navier-slip boundary condition under the curvilinear coordinates. Writing

𝒖=us𝒆𝒔+ut𝒆𝒕.\boldsymbol{u}=u_{s}\boldsymbol{e_{s}}+u_{t}\boldsymbol{e_{t}}.

Then (1.2) enjoys the following simplified expression:

{us𝒏=tus=(κα)us,ut=0,on𝒮.\left\{\begin{aligned} &\frac{\partial u_{s}}{\partial\boldsymbol{n}}=-\partial_{t}u_{s}=\left(\kappa-\alpha\right)u_{s},\\ &u_{t}=0,\\ \end{aligned}\right.\quad\text{on}\quad\partial\mathcal{S}_{-}. (2.27)

See [31, Proposition 2.1 and Corollary 2.2] for a detailed calculation. Moreover, denoting 𝔴=x2u1x1u2\mathfrak{w}=\partial_{x_{2}}u_{1}-\partial_{x_{1}}u_{2} and applying (2.26)2, one has (2.27)1 is equivalent to

𝔴=(2κ+α)us,on𝒮.\mathfrak{w}=\left(-2\kappa+\alpha\right)u_{s},\quad\text{on}\quad\partial\mathcal{S}_{-}.

2.2 The Poincaré inequality and the Korn’s inequality

The following Poincaré inequalities and Korn’s inequality will play crucial role in the existence and uniqueness results when the no-slip boundary is replaced by the Navier-Slip boundary.

Lemma 2.3 (Poincaré inequality in a straight strip).

Let 𝐠=g1𝐞𝟏+g2𝐞𝟐\boldsymbol{g}=g_{1}\boldsymbol{e_{1}}+g_{2}\boldsymbol{e_{2}} be a H1H^{1} vector field in the box domain S:=[a,b]×[c,d]S:=[a,b]\times[c,d], and satisfies that

cdg1(x1,x2)𝑑x2=g2(x1,x2)|x2=c,d=0,x1[a,b],\int^{d}_{c}g_{1}(x_{1},x_{2})dx_{2}=g_{2}(x_{1},x_{2})\big{|}_{x_{2}=c,d}=0,\quad\forall\ x_{1}\in[a,b],

then we have the following

𝒈L2(S)Cx2𝒈L2(S).\|\boldsymbol{g}\|_{L^{2}(S)}\leq C\|\partial_{x_{2}}\boldsymbol{g}\|_{L^{2}(S)}. (2.28)

where C|cd|C\lesssim|c-d| is a constant depending on the width of the strip.

Proof.   Since g1g_{1} has zero mean and g2g_{2} has zero boundary in the x2x_{2} direction. The classical one dimensional Poincaré inequality leads to

𝒈(x1)Lx22([c,d])2|dc|x2𝒈(x1)Lx22([c,d])2,x1[a,b]\|\boldsymbol{g}(x_{1})\|^{2}_{L^{2}_{x_{2}}([c,d])}\lesssim_{|d-c|}\|\partial_{x_{2}}\boldsymbol{g}(x_{1})\|^{2}_{L^{2}_{x_{2}}([c,d])},\quad\forall\ x_{1}\in[a,b]

Integration on [a,b][a,b] with respect to x1x_{1} variable indicates (2.28).

Lemma 2.4 (Poincaré inequality in the torsion part).

Let ζ>1\zeta>1 and 𝐡=h1𝐞𝟏+h2𝐞𝟐=h~1𝐞𝟏+h~2𝐞𝟐H1(𝒮ζ)\boldsymbol{h}=h_{1}\boldsymbol{e_{1}}+h_{2}\boldsymbol{e_{2}}=\tilde{h}_{1}\boldsymbol{e_{1}}^{\prime}+\tilde{h}_{2}\boldsymbol{e_{2}}^{\prime}\in H^{1}(\mathcal{S}_{\zeta}) be a divergence free vector with zero flux, that is

01h1(x1,x2)𝑑x2=0for anyx𝒮ζ𝒮R.\int_{0}^{1}{h}_{1}(x_{1},x_{2})dx_{2}=0\quad\text{for any}\ x\in\mathcal{S}_{\zeta}\cap\mathcal{S}_{R}.

If we suppose 𝐡𝐧0\boldsymbol{h}\cdot\boldsymbol{n}\equiv 0 on 𝒮𝒮ζ\partial\mathcal{S}\cap\partial\mathcal{S}_{\zeta}, where 𝐧\boldsymbol{n} is the unit outer normal vector on 𝒮\partial\mathcal{S}, then the following Poincaré inequality holds:

𝒉L2(𝒮ζ)C𝒉L2(𝒮ζ).\|\boldsymbol{h}\|_{L^{2}(\mathcal{S}_{\zeta})}\leq C\|\nabla\boldsymbol{h}\|_{L^{2}(\mathcal{S}_{\zeta})}. (2.29)

Here C>0C>0 is a uniform constant, independent of ζ\zeta.

Proof.   Integrating the following identity on 𝒮1\mathcal{S}_{1},

i,j=12[xi(hixjhj)xihixjhj|𝒉|2hixjxihj]=0,\sum^{2}_{i,j=1}\left[\partial_{x_{i}}(h_{i}x_{j}h_{j})-\partial_{x_{i}}h_{i}x_{j}h_{j}-|\boldsymbol{h}|^{2}-h_{i}x_{j}\partial_{x_{i}}h_{j}\right]=0, (2.30)

one deduces

𝒮1|𝒉|2𝑑x=\displaystyle\int_{\mathcal{S}_{1}}|\boldsymbol{h}|^{2}dx= i,j=12𝒮1i(hixjhj)dxJ1i,j=12𝒮1ihixjhjdxi,j=12𝒮1hixjihjdx.\displaystyle\underbrace{\sum^{2}_{i,j=1}\int_{\mathcal{S}_{1}}\partial_{i}(h_{i}x_{j}h_{j})dx}_{J_{1}}-\sum^{2}_{i,j=1}\int_{\mathcal{S}_{1}}\partial_{i}h_{i}x_{j}h_{j}dx-\sum^{2}_{i,j=1}\int_{\mathcal{S}_{1}}h_{i}x_{j}\partial_{i}h_{j}dx. (2.31)

Using the divergence theorem and the boundary condition 𝒉𝒏=0\boldsymbol{h}\cdot{\boldsymbol{n}}=0 on 𝒮1𝒮\partial\mathcal{S}_{1}\cap\partial\mathcal{S}, we can obtain

J1=j=12𝒮1(𝒏𝒉)xjhj𝑑S=01((x𝒉)h1)(1,x2)𝑑x20c0((x𝒉)h~1)(1,y2)𝑑y2.J_{1}=\sum^{2}_{j=1}\int_{\partial\mathcal{S}_{1}}({\boldsymbol{n}}\cdot\boldsymbol{h})x_{j}h_{j}dS=\int_{0}^{1}\left((x\cdot\boldsymbol{h})h_{1}\right)(1,x_{2})dx_{2}-\int_{0}^{c_{0}}\left((x\cdot\boldsymbol{h})\tilde{h}_{1}\right)(-1,y_{2})dy_{2}.

Thus by (2.31) and the Cauchy-Schwarz inequality, we arrive at

𝒮1|𝒉|2𝑑x12𝒮1|𝒉|2𝑑x+C𝒮1|𝒉|2𝑑x+|01((x𝒉)h1)(1,x2)𝑑x2|+|0c0((x𝒉)h~1)(1,y2)𝑑y2|,\begin{split}\int_{\mathcal{S}_{1}}|\boldsymbol{h}|^{2}dx\leq&\frac{1}{2}\int_{\mathcal{S}_{1}}|\boldsymbol{h}|^{2}dx+C\int_{\mathcal{S}_{1}}|\nabla\boldsymbol{h}|^{2}dx\\ &+\left|\int_{0}^{1}\left((x\cdot\boldsymbol{h})h_{1}\right)(1,x_{2})dx_{2}\right|+\left|\int_{0}^{c_{0}}\left((x\cdot\boldsymbol{h})\tilde{h}_{1}\right)(-1,y_{2})dy_{2}\right|,\end{split}

which indicates

𝒮1|𝒉|2𝑑xC(𝒮1|𝒉|2𝑑x+01|𝒉(1,x2)|2𝑑x2+0c0|𝒉(1,y2)|2𝑑y2).\int_{\mathcal{S}_{1}}|\boldsymbol{h}|^{2}dx\leq C\left(\int_{\mathcal{S}_{1}}|\nabla\boldsymbol{h}|^{2}dx+\int_{0}^{1}\left|\boldsymbol{h}(1,x_{2})\right|^{2}dx_{2}+\int_{0}^{c_{0}}\left|\boldsymbol{h}(-1,y_{2})\right|^{2}dy_{2}\right). (2.32)

Meanwhile, using the trace theorem in 𝒮L,1\mathcal{S}_{L,1}, and Lemma 2.3, one derives

0c0|𝒉(1,y2)|2𝑑y2C((1,0)×(0,c0)|𝒉|2𝑑y1𝑑y2+(1,0)×(0,c0)|𝒉|2𝑑y1𝑑y2)C𝒮L,1|𝒉|2𝑑x.\begin{split}\int_{0}^{c_{0}}\left|\boldsymbol{h}(-1,y_{2})\right|^{2}dy_{2}&\leq C\left(\int_{(-1,0)\times(0,c_{0})}\left|\boldsymbol{h}\right|^{2}dy_{1}dy_{2}+\int_{(-1,0)\times(0,c_{0})}\left|\nabla\boldsymbol{h}\right|^{2}dy_{1}dy_{2}\right)\\ &\leq C\int_{\mathcal{S}_{L,1}}\left|\nabla\boldsymbol{h}\right|^{2}dx.\end{split} (2.33)

Similarly, one deduces that

01|𝒉(1,x2)|2𝑑x2C𝒮R,1|𝒉|2𝑑x.\int_{0}^{1}\left|\boldsymbol{h}(1,x_{2})\right|^{2}dx_{2}\leq C\int_{\mathcal{S}_{R,1}}\left|\nabla\boldsymbol{h}\right|^{2}dx. (2.34)

Substituting (2.33)–(2.34) in the right hand side of (2.32), one deduces

𝒮1|𝒉|2𝑑xC𝒮1|𝒉|2𝑑x.\int_{\mathcal{S}_{1}}|\boldsymbol{h}|^{2}dx\leq C\int_{\mathcal{S}_{1}}|\nabla\boldsymbol{h}|^{2}dx.

Using Lemma 2.3, it is easy to see that

𝒮ζ\𝒮1|𝒉|2𝑑xC𝒮ζ\𝒮1|𝒉|2𝑑x.\int_{\mathcal{S}_{\zeta}\backslash\mathcal{S}_{1}}|\boldsymbol{h}|^{2}dx\leq C\int_{\mathcal{S}_{\zeta}\backslash\mathcal{S}_{1}}|\nabla\boldsymbol{h}|^{2}dx.

Combining the above two inequalities, we finish the proof of (2.35).

After showing Lemma 2.3 and Lemma 2.4, one concludes the following Poincaré inequality in the whole infinite strip:

Corollary 2.5 (Poincaré inequality in 𝒮\mathcal{S}).

Let

𝒈𝒱:={𝒇=(f1,f2)H1(𝒮):(𝒇𝒏)|𝒮=0,div𝒇=0},\boldsymbol{g}\in\mathcal{V}:=\left\{\boldsymbol{f}=(f_{1},f_{2})\in H^{1}(\mathcal{S}):\,\left(\boldsymbol{f}\cdot\boldsymbol{n}\right)\big{|}_{\partial\mathcal{S}}=0,\,\mathrm{div}\,\boldsymbol{f}=0\right\},

then the following Poincaré inequality holds:

𝒈L2(𝒮)C𝒈L2(𝒮).\|\boldsymbol{g}\|_{L^{2}(\mathcal{S})}\leq C\|\nabla\boldsymbol{g}\|_{L^{2}(\mathcal{S})}. (2.35)

Here C>0C>0 is a uniform constant.

Proof.   This is a direct conclusion by gluing results in Lemma 2.3 and Lemma 2.4 together, after we have shown

𝒮{x1=0}g1𝑑x2=0\int_{\mathcal{S}\cap\{x_{1}=0\}}g_{1}dx_{2}=0

holds unconditionally for 𝒈𝒱\boldsymbol{g}\in\mathcal{V}. By the divergence theorem,

𝒮{x1=𝔰}g1𝑑x2𝒮{x1=0}g1𝑑x2=𝒮{0<x1<𝔰}div𝒈𝑑x𝒮{0<x1<𝔰}(𝒈𝒏)𝑑S=0\int_{\mathcal{S}\cap\{x_{1}=\mathfrak{s}\}}g_{1}dx_{2}-\int_{\mathcal{S}\cap\{x_{1}=0\}}g_{1}dx_{2}=\int_{\mathcal{S}\cap\{0<x_{1}<\mathfrak{s}\}}\mathrm{div}\,\boldsymbol{g}dx-\int_{\partial\mathcal{S}\cap\{0<x_{1}<\mathfrak{s}\}}\left(\boldsymbol{g}\cdot\boldsymbol{n}\right)dS=0

for any 𝔰>0\mathfrak{s}>0. Thus if

|𝒮{x1=0}g1𝑑x2|=c0>0,\left|\int_{\mathcal{S}\cap\{x_{1}=0\}}g_{1}dx_{2}\right|=c_{0}>0,

one deduces

c0𝒮{x1=𝔰}|g1|𝑑x2|𝒮{x1=𝔰}|1/2(𝒮{x1=𝔰}|g1|2𝑑x2)1/2.c_{0}\leq\int_{\mathcal{S}\cap\{x_{1}=\mathfrak{s}\}}|g_{1}|dx_{2}\leq\left|\mathcal{S}\cap\{x_{1}=\mathfrak{s}\}\right|^{1/2}\left(\int_{\mathcal{S}\cap\{x_{1}=\mathfrak{s}\}}|g_{1}|^{2}dx_{2}\right)^{1/2}.

This implies that, for any 𝔰>Z0\mathfrak{s}>Z_{0}

𝒮{x1=𝔰}|g1|2𝑑x2c02,\int_{\mathcal{S}\cap\{x_{1}=\mathfrak{s}\}}|g_{1}|^{2}dx_{2}\geq c_{0}^{2},

which results in a paradox with 𝒈H1(𝒮)\boldsymbol{g}\in H^{1}(\mathcal{S}) .

Here goes the Korn’s inequality in the truncated strip:

Lemma 2.6 (Korn’s inequality).

Let 𝒮ζ\mathcal{S}_{\zeta} with ζ>0\zeta>0 be the finite truncated strip given in (1.3). For any

𝒈𝒱ζ:={𝒇=(f1,f2)H1(𝒮ζ;2):(𝒇𝒏)|𝒮𝒮ζ=0,div𝒇=0,𝒮{x1=0}f1𝑑x2=0},\boldsymbol{g}\in\mathcal{V}_{\zeta}:=\left\{\boldsymbol{f}=(f_{1},f_{2})\in H^{1}(\mathcal{S}_{\zeta};\,\mathbb{R}^{2}):\,\left(\boldsymbol{f}\cdot\boldsymbol{n}\right)\big{|}_{\partial\mathcal{S}\cap\partial\mathcal{S}_{\zeta}}=0,\,\mathrm{div}\,\boldsymbol{f}=0,\,\int_{\mathcal{S}\cap\{x_{1}=0\}}f_{1}dx_{2}=0\right\},

there exists C>0C>0, which is independent of 𝐠\boldsymbol{g} or ζ\zeta, such that

𝒈H1(𝒮ζ)2C𝕊𝒈L2(𝒮ζ)2+2{x1=ζ}|𝒈||𝒈|𝑑x2+2{y1=ζ}|𝒈||𝒈|𝑑y2.\|\boldsymbol{g}\|^{2}_{H^{1}(\mathcal{S}_{\zeta})}\leq C\|\mathbb{S}\boldsymbol{g}\|^{2}_{L^{2}(\mathcal{S}_{\zeta})}+2\int_{\{x_{1}=\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dx_{2}+2\int_{\{y_{1}=-\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dy_{2}. (2.36)

Proof.   Noting that

𝒮ζ|𝕊𝒈|2𝑑x\displaystyle\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{g}|^{2}dx =12𝒮ζi,j=12(xjgi+xigj)2dx\displaystyle=\frac{1}{2}\int_{\mathcal{S}_{\zeta}}\sum_{i,j=1}^{2}\left(\partial_{x_{j}}g_{i}+\partial_{x_{i}}g_{j}\right)^{2}dx (2.37)
=𝒮ζi,j=12((xjgi)2+xjgixigj)dx\displaystyle=\int_{\mathcal{S}_{\zeta}}\sum_{i,j=1}^{2}\left(\left(\partial_{x_{j}}g_{i}\right)^{2}+\partial_{x_{j}}g_{i}\partial_{x_{i}}g_{j}\right)dx
=𝒈L2(𝒮ζ)2+𝒮ζi,j=12xjgixigjdxK1.\displaystyle=\|\nabla\boldsymbol{g}\|_{L^{2}(\mathcal{S}_{\zeta})}^{2}+\underbrace{\int_{\mathcal{S}_{\zeta}}\sum_{i,j=1}^{2}\partial_{x_{j}}g_{i}\partial_{x_{i}}g_{j}dx}_{K_{1}}.

For the last term of (2.37), we can assume that 𝒈C2(𝒮ζ)\boldsymbol{g}\in C^{2}(\mathcal{S}_{\zeta}) without loss of generality. By integration by parts, we get

K1=\displaystyle K_{1}= 𝒮ζgixixj2gjdx+𝒮ζ𝒮i,j=12njgixigjdS+𝒮ζ{x1=±ζ}i,j=12njgixigjdx2\displaystyle-\int_{\mathcal{S}_{\zeta}}g_{i}\partial^{2}_{x_{i}x_{j}}g_{j}dx+\int_{\partial\mathcal{S}_{\zeta}\cap\partial\mathcal{S}}\sum_{i,j=1}^{2}n_{j}g_{i}\partial_{x_{i}}g_{j}dS+\int_{\partial\mathcal{S}_{\zeta}\cap\{x_{1}=\pm\zeta\}}\sum_{i,j=1}^{2}n_{j}g_{i}\partial_{x_{i}}g_{j}dx_{2} (2.38)
=\displaystyle= 𝒮(div𝒈)2𝑑x𝒮ζ𝒮i,j=12nigixjgjdS+𝒮ζ𝒮i,j=12njgixigjdS\displaystyle\int_{\mathcal{S}}(\mathrm{div}\,\boldsymbol{g})^{2}dx-\int_{\partial\mathcal{S}_{\zeta}\cap\partial\mathcal{S}}\sum_{i,j=1}^{2}n_{i}g_{i}\partial_{x_{j}}g_{j}dS+\int_{\partial\mathcal{S}_{\zeta}\cap\partial\mathcal{S}}\sum_{i,j=1}^{2}n_{j}g_{i}\partial_{x_{i}}g_{j}dS
{x1=ζ}i,j=12nigixjgjdx2{y1=ζ}i,j=12nigixjgjdy2\displaystyle-\int_{\{x_{1}=\zeta\}}\sum_{i,j=1}^{2}n_{i}g_{i}\partial_{x_{j}}g_{j}dx_{2}-\int_{\{y_{1}=-\zeta\}}\sum_{i,j=1}^{2}n_{i}g_{i}\partial_{x_{j}}g_{j}dy_{2}
+{x1=ζ}i,j=12njgixigjdx2+{y1=ζ}i,j=12njgixigjdy2.\displaystyle+\int_{\{x_{1}=\zeta\}}\sum_{i,j=1}^{2}n_{j}g_{i}\partial_{x_{i}}g_{j}dx_{2}+\int_{\{y_{1}=-\zeta\}}\sum_{i,j=1}^{2}n_{j}g_{i}\partial_{x_{i}}g_{j}dy_{2}.

The first, fourth and fifth terms on the far right of above equation vanish owing to the divergence-free property of 𝒈\boldsymbol{g}, and the second one also vanishes because 𝒈𝒏|𝒮=0\left.\boldsymbol{g}\cdot\boldsymbol{n}\right|_{\partial\mathcal{S}}=0. Meanwhile, the condition 𝒈𝒏|𝒮=0\left.\boldsymbol{g}\cdot\boldsymbol{n}\right|_{\partial\mathcal{S}}=0 also implies that, at the boundary,

i=12gixi(𝒈𝒏)=0 or i,j=12gixigjnj=i,j=12gigjxinj.\sum_{i=1}^{2}g_{i}\partial_{x_{i}}(\boldsymbol{g}\cdot\boldsymbol{n})=0\quad\text{ or }\quad\sum_{i,j=1}^{2}g_{i}\partial_{x_{i}}g_{j}n_{j}=-\sum_{i,j=1}^{2}g_{i}g_{j}\partial_{x_{i}}n_{j}. (2.39)

Using (2.38) to (2.39), we get

K1=\displaystyle K_{1}= 𝒮𝒮ζκ(x)|𝒈|2𝑑S+{x1=ζ}i,j=12njgixigjdx2+{y1=ζ}i,j=12njgixigjdy2.\displaystyle-\int_{\partial\mathcal{S}\cap\partial\mathcal{S}_{\zeta}}\kappa(x)|\boldsymbol{g}|^{2}dS+\int_{\{x_{1}=\zeta\}}\sum_{i,j=1}^{2}n_{j}g_{i}\partial_{x_{i}}g_{j}dx_{2}+\int_{\{y_{1}=-\zeta\}}\sum_{i,j=1}^{2}n_{j}g_{i}\partial_{x_{i}}g_{j}dy_{2}.

where κ(x)\kappa(x) is the curvature of the boundary 𝒮\partial\mathcal{S}. By definition of 𝒮\mathcal{S}, we have κ(x)0\kappa(x)\equiv 0 on 𝒮\𝒮0\partial\mathcal{S}\backslash\partial\mathcal{S}_{0}, and 𝒏=𝒆𝟏\boldsymbol{n}=\boldsymbol{e_{1}} on 𝒮ζ{x1=ζ}\partial\mathcal{S}_{\zeta}\cap\{x_{1}=\zeta\}, while 𝒏=𝒆𝟏\boldsymbol{n}=-\boldsymbol{e_{1}}^{\prime} on 𝒮ζ{y1=ζ}\partial\mathcal{S}_{\zeta}\cap\{y_{1}=-\zeta\}. This guarantees that

|K1|𝒮𝒮0|κ(x)||𝒈|2𝑑S+{x1=ζ}|𝒈||𝒈|𝑑x2+{y1=ζ}|𝒈||𝒈|𝑑y2.|K_{1}|\leq\int_{\partial\mathcal{S}\cap\partial\mathcal{S}_{0}}|\kappa(x)||\boldsymbol{g}|^{2}dS+\int_{\{x_{1}=\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dx_{2}+\int_{\{y_{1}=-\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dy_{2}. (2.40)

Substituting (2.38)–(2.40) in (2.37), one concludes

𝒈L2(𝒮ζ)2𝒮ζ|𝕊𝒈|2𝑑x+𝒮𝒮0|κ(x)||𝒈|2𝑑S+{x1=ζ}|𝒈||𝒈|𝑑x2+{y1=ζ}|𝒈||𝒈|𝑑y2.\|\nabla\boldsymbol{g}\|_{L^{2}(\mathcal{S}_{\zeta})}^{2}\leq\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{g}|^{2}dx+\int_{\partial\mathcal{S}\cap\partial\mathcal{S}_{0}}|\kappa(x)||\boldsymbol{g}|^{2}dS+\int_{\{x_{1}=\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dx_{2}+\int_{\{y_{1}=-\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dy_{2}.

Noting that κ(x)L(𝒮)\|\kappa(x)\|_{L^{\infty}(\partial\mathcal{S})} is uniformly bounded due to the smoothness of 𝒮\partial\mathcal{S} and combining with the Poincaré inequalities in Lemma 2.3 and Lemma 2.4, one deduces that there exists C>0C>0 that

𝒈H1(𝒮ζ)2C(𝒮ζ|𝕊𝒈|2𝑑x+𝒈L2(𝒮𝒮0)2)+{x1=ζ}|𝒈||𝒈|𝑑x2+{y1=ζ}|𝒈||𝒈|𝑑y2.\|\boldsymbol{g}\|_{H^{1}(\mathcal{S}_{\zeta})}^{2}\leq C\left(\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{g}|^{2}dx+\|\boldsymbol{g}\|_{L^{2}(\partial\mathcal{S}\cap\partial\mathcal{S}_{0})}^{2}\right)+\int_{\{x_{1}=\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dx_{2}+\int_{\{y_{1}=-\zeta\}}|\boldsymbol{g}||\nabla\boldsymbol{g}|dy_{2}.

To finish the proof, one only needs to show that there exists C>0C>0 such that:

𝒈L2(𝒮𝒮0)212C𝒈H1(𝒮ζ)2+C𝒮ζ|𝕊𝒈|2𝑑x.\|\boldsymbol{g}\|_{L^{2}(\partial\mathcal{S}\cap\partial\mathcal{S}_{0})}^{2}\leq\frac{1}{2C}\|\boldsymbol{g}\|_{H^{1}(\mathcal{S}_{\zeta})}^{2}+C\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{g}|^{2}dx. (2.41)

We prove this by the method of contradiction. If a number the above CC does not exist, then there exists a bounded sequence {𝒈m}m=0𝒱ζ\left\{\boldsymbol{g}_{m}\right\}_{m=0}^{\infty}\subset\mathcal{V}_{\zeta} such that

𝒈mL2(𝒮𝒮0)212C1𝒈mH1(𝒮ζ)2+m𝒮ζ|𝕊𝒈m|2𝑑x.\left\|\boldsymbol{g}_{m}\right\|_{L_{2}(\partial\mathcal{S}\cap\partial\mathcal{S}_{0})}^{2}\geq\frac{1}{2C_{1}}\left\|\boldsymbol{g}_{m}\right\|_{H^{1}(\mathcal{S}_{\zeta})}^{2}+m\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{g}_{m}|^{2}dx.

Denoting 𝒉m=𝒈m/𝒈mL2(𝒮𝒮0)\boldsymbol{h}_{m}=\boldsymbol{g}_{m}/\left\|\boldsymbol{g}_{m}\right\|_{L^{2}(\partial\mathcal{S}\cap\partial\mathcal{S}_{0})}, one deduces that

𝒉mL2(𝒮𝒮0)=1 and m𝒮ζ|𝕊𝒉m|2𝑑x1.\left\|\boldsymbol{h}_{m}\right\|_{L^{2}(\partial\mathcal{S}\cap\partial\mathcal{S}_{0})}=1\quad\text{ and }\quad m\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{h}_{m}|^{2}dx\leq 1. (2.42)

Since the sequence {𝒈m}\left\{\boldsymbol{g}_{m}\right\} is bounded in 𝒱ζ\mathcal{V}_{\zeta}, we can choose a subsequence {𝒉mk}k=0\left\{\boldsymbol{h}_{m_{k}}\right\}_{k=0}^{\infty} which is weakly convergent in H1(𝒮ζ)H^{1}(\mathcal{S}_{\zeta}) and strongly in L2(𝒮𝒮0)L^{2}(\partial\mathcal{S}\cap\partial\mathcal{S}_{0}) to a vector 𝒉𝒱ζ\boldsymbol{h}_{*}\in\mathcal{V}_{\zeta}. Particularly,

𝕊𝒉mk𝕊𝒉,weakly in L2(𝒮ζ).\mathbb{S}\boldsymbol{h}_{m_{k}}\to\mathbb{S}\boldsymbol{h}_{*},\quad\text{weakly in }L^{2}(\mathcal{S}_{\zeta}).

By (2.42), one knows

𝒮ζ|𝕊𝒉mk|2𝑑x1mk0,ask.\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{h}_{m_{k}}|^{2}dx\leq\frac{1}{m_{k}}\rightarrow 0,\quad\text{as}\quad k\to\infty.

Thus one deduces

𝒮ζ|𝕊𝒉|2dxlim infk𝒮ζ|𝕊𝒉mk|2𝑑x=0,\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{h}_{*}|^{2}\mathrm{~{}d}x\leq\liminf_{k\to\infty}\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{h}_{m_{k}}|^{2}dx=0,

by the Fatou’s lemma for weakly convergent sequences. This concludes 𝕊𝒉0\mathbb{S}\boldsymbol{h}_{*}\equiv 0 in 𝒮ζ\mathcal{S}_{\zeta}. It is well known that 𝒉\boldsymbol{h}_{*} has the form 𝒉=Ax+B\boldsymbol{h}_{*}=Ax+B (see [13, §6]), where AA is a constant skew-symmetric matrix with constant entries and BB is a constant vector, that is,

𝒉=(0aa0)(x1x2)+(b1b2)=(ax2+b1ax1+b2),\boldsymbol{h}_{*}=\begin{pmatrix}0&-a\\ a&0\end{pmatrix}\begin{pmatrix}x_{1}\\ x_{2}\end{pmatrix}+\begin{pmatrix}b_{1}\\ b_{2}\end{pmatrix}=\begin{pmatrix}-a\,x_{2}+b_{1}\\ a\,x_{1}+b_{2}\end{pmatrix}\,,

where a,bia\,,b_{i} (i=1,2i=1\,,2) are some constants. However, by the boundary condition 𝒉𝒏=0\boldsymbol{h}_{*}\cdot\boldsymbol{n}=0 holds everywhere on 𝒮𝒮ζ\partial\mathcal{S}\cap\partial\mathcal{S}_{\zeta}, one has

(h)2=ax1+b20,for all0<x1<ζ,(h_{*})_{2}=ax_{1}+b_{2}\equiv 0,\quad\text{for all}\quad 0<x_{1}<\zeta,

which indicates a=b20a=b_{2}\equiv 0. This indicates (h)1=b1(h_{*})_{1}=b_{1} and thus b1=0b_{1}=0 due to

01(h)1(x1,x2)𝑑x2=0,for all0<x1<ζ.\int_{0}^{1}(h_{*})_{1}(x_{1},x_{2})dx_{2}=0,\quad\text{for all}\quad 0<x_{1}<\zeta.

Therefore one concludes 𝒉0\boldsymbol{h}_{*}\equiv 0 in 𝒮ζ\mathcal{S}_{\zeta}. However, this creates a paradox to the fact

𝒉L2(𝒮𝒮0)=1\|\boldsymbol{h}_{*}\|_{L^{2}(\partial\mathcal{S}\cap\partial\mathcal{S}_{0})}=1

coming from (2.42). This indicates the validity of (2.41) and therefore one concludes (2.36).

If we replace the truncated strip with the infinite strip 𝒮\mathcal{S}, the result in Lemma 2.6 will be simpler with boundary term integrations on the segments {x1=ζ}\{x_{1}=\zeta\} and {y1=ζ}\{y_{1}=-\zeta\} disappearing. We have the following Corollary.

Corollary 2.7.

Let 𝒮\mathcal{S} be the infinite strip given in the previous section. For any 𝐠𝒱\boldsymbol{g}\in\mathcal{V}, there exists C>0C>0, which is independent of 𝐠\boldsymbol{g}, such that

𝒈H1(𝒮)C𝕊𝒈L2(𝒮).\|\boldsymbol{g}\|_{H^{1}(\mathcal{S})}\leq C\|\mathbb{S}\boldsymbol{g}\|_{L^{2}(\mathcal{S})}. (2.43)

Remark 2.8.

Here let us give a brief explanation why this Korn’s inequality fails to be valid in a 3D infinite pipe. Consider the vector

𝒘=(ξ(x3)x2,ξ(x3)x1, 0)\boldsymbol{w}=(-\xi(x_{3})x_{2}\,,\,\xi(x_{3})x_{1}\,,\,0)

given in the cylindrical pipe 𝒟=B×\mathcal{D}=B\times\mathbb{R}, where BB is the unit disk in 2\mathbb{R}^{2}, and ξ\xi is a smooth cut-off function that:

ξ(x3)={1,x3[R,R];0,x3\(R1,R+1),\xi(x_{3})=\left\{\begin{aligned} 1\,,&\quad\quad x_{3}\in[-R,R];\\ 0\,,&\quad\quad x_{3}\in\mathbb{R}\backslash(-R-1,R+1),\end{aligned}\right.

with

|ξ(x3)|2,for anyx3(R1,R)(R,R+1).|\xi^{\prime}(x_{3})|\leq 2,\quad\text{for any}\quad x_{3}\in(-R-1,R)\cup(R,R+1).

One notices that 𝐰\boldsymbol{w} is divergence-free and it satisfies 𝐰𝐧0\boldsymbol{w}\cdot\boldsymbol{n}\equiv 0 on B×\partial B\times\mathbb{R}, also its flux in the cross section B×{x3=0}B\times\{x_{3}=0\} is zero.

For the convenience of calculation, we introduce the cylindrical coordinates:

𝒆𝒓=(x1r,x2r,0),𝒆𝜽=(x2r,x1r,0),𝒆𝒛=(0,0,1),\boldsymbol{e_{r}}=(\frac{x_{1}}{r},\frac{x_{2}}{r},0),\quad\boldsymbol{e_{\theta}}=(-\frac{x_{2}}{r},\frac{x_{1}}{r},0),\quad\boldsymbol{e_{z}}=(0,0,1),

and we find

𝒘=ξ(z)r𝒆𝜽.\boldsymbol{w}=\xi(z)r\boldsymbol{e_{\theta}}.

Using equation (A.4) in [18], one finds

𝕊𝒘=12ξ(z)r(𝒆𝜽𝒆𝒛+𝒆𝒛𝒆𝜽).\mathbb{S}\boldsymbol{w}=\frac{1}{2}\xi^{\prime}(z)r\left(\boldsymbol{e_{\theta}}\otimes\boldsymbol{e_{z}}+\boldsymbol{e_{z}}\otimes\boldsymbol{e_{\theta}}\right).

This indicates

𝒟|𝕊𝒘|2𝑑x=O(1)\int_{\mathcal{D}}|\mathbb{S}\boldsymbol{w}|^{2}dx=O(1)

which is independent with RR. On the other hand

𝒟|𝒘|2𝑑x𝒟|2w1|2𝑑x2πR.\int_{\mathcal{D}}|\nabla\boldsymbol{w}|^{2}dx\geq\int_{\mathcal{D}}|\partial_{2}w_{1}|^{2}dx\geq 2\pi R.

Noting that R>0R>0 is arbitrary, one could not find a uniform constant C>0C>0 such that a “3D version” (2.36) or (2.43) holds.

Remark 2.9.

In the 3-dimensional case, the curvature of the domain boundary κ(x)\kappa(x) no longer has compact support. In this case one cannot find a subsequence {𝐡mk}k=0\left\{\boldsymbol{h}_{m_{k}}\right\}_{k=0}^{\infty} which is strongly convergent in L2(𝒟)L^{2}(\partial\mathcal{D}) to a vector 𝐡\boldsymbol{h}_{*}. That is why our method in the proof of Lemma 2.6 fails in the 3-dimensional case.

2.3 Other useful lemmas

The following Brouwer’s fixed point theorem is crucial to establish the existence. See [20] or [10, Lemma IX.3.1].

Lemma 2.10.

Let PP be a continuous operator which maps N\mathbb{R}^{N} into itself, such that for some ρ>0\rho>0

P(ξ)ξ0 for all ξN with |ξ|=ρ.P({\xi})\cdot{\xi}\geq 0\quad\text{ for all }{\xi}\in\mathbb{R}^{N}\text{ with }\,|{\xi}|=\rho.

Then there exists ξ0N{\xi}_{0}\in\mathbb{R}^{N} with |ξ0|ρ|{\xi}_{0}|\leq\rho such that P(ξ0)=0{P}({\xi}_{0})=0.

The following asymptotic estimate of a function that satisfies an ordinary differential inequality will be useful in our further proof. To the best of the authors’ knowledge, it was originally derived by Ladyzhenskaya-Solonnikov in [17]. We also refer readers to [19, Lemma 2.7] for a proof written in a relatively recent format.

Lemma 2.11.

Let Y(ζ)0Y(\zeta)\nequiv 0 be a nondecreasing nonnegative differentiable function satisfying

Y(ζ)Ψ(Y(ζ)),ζ>0.Y(\zeta)\leq\Psi(Y^{\prime}(\zeta)),\quad\forall\zeta>0.

Here Ψ:[0,)[0,)\Psi:\,[0,\infty)\to[0,\infty) is a monotonically increasing function with Ψ(0)=0\Psi(0)=0 and there exists C,τ1>0C,\,\tau_{1}>0, m>1m>1, such that

Ψ(τ)Cτm,τ>τ1.\Psi(\tau)\leq C\tau^{m},\quad\forall\tau>\tau_{1}.

Then

lim infζ+ζmm1Y(ζ)>0.\liminf_{\zeta\to+\infty}\zeta^{-\frac{m}{m-1}}Y(\zeta)>0.

The following two lemmas are essential in creating the pressure field for a weak solution to the Navier-Stokes equations. The first one is a special case of [9, Theorem 17] by De Rham. See also [27, Proposition 1.1].

Lemma 2.12.

For a given open set Ω2\Omega\subset\mathbb{R}^{2}, let 𝓕\boldsymbol{\mathcal{F}} be a distribution in (Cc(Ω))\left(C_{c}^{\infty}(\Omega)\right)^{\prime} which satisfies:

𝓕,ϕ=0,for allϕ{𝒈Cc(Ω;2):div 𝒈=0}.\langle\boldsymbol{\mathcal{F}},\boldsymbol{\phi}\rangle=0,\quad\text{for all}\quad\boldsymbol{\phi}\in\{\boldsymbol{g}\in C_{c}^{\infty}(\Omega;\mathbb{R}^{2}):\,\text{div }\boldsymbol{g}=0\}.

Then there exists a distribution q(Cc(Ω;))q\in\left(C_{c}^{\infty}(\Omega;\mathbb{R})\right)^{\prime} such that

𝓕=q.\boldsymbol{\mathcal{F}}=\nabla q.

The second one states the regularity of the aforementioned field qq:

Lemma 2.13 (See [27], Proposition 1.2).

Let Ω\Omega be a bounded Lipschitz open set in 2\mathbb{R}^{2}. If a distribution qq has all its first derivatives xiq\partial_{x_{i}}q, 1i21\leq i\leq 2, in H1(Ω)H^{-1}(\Omega), then qL2(Ω)q\in L^{2}(\Omega) and

qq¯ΩL2(Ω)CΩqH1(Ω),\left\|q-\bar{q}_{\Omega}\right\|_{L^{2}(\Omega)}\leq C_{\Omega}\|\nabla q\|_{H^{-1}(\Omega)}, (2.44)

where q¯Ω=1|Ω|Ωq𝑑x\bar{q}_{\Omega}=\frac{1}{|\Omega|}\int_{\Omega}qdx. Moreover, if Ω\Omega is any Lipschitz open set in 2\mathbb{R}^{2}, then qLloc2(Ω¯)q\in L_{\mathrm{loc}}^{2}(\overline{\Omega}).

Finally, we state the following lemma, which shows the existence of the solution to problem 𝑽=f\nabla\cdot\boldsymbol{V}=f in a truncated regular stripe.

Lemma 2.14.

For a boxed domain S:=[a,b]×[c,d]S:=[a,b]\times[c,d], if fL2(S)f\in L^{2}(S) with Sf𝑑x=0\int_{S}fdx=0, then there exists a vector valued function 𝐕:S2\boldsymbol{V}:\,S\to\mathbb{R}^{2} belongs to H01(S)H^{1}_{0}(S) such that

𝑽=f,and𝑽L2(S)CfL2(S).\nabla\cdot\boldsymbol{V}=f,\quad\text{and}\quad\|\nabla\boldsymbol{V}\|_{L^{2}(S)}\leq C\|f\|_{L^{2}(S)}. (2.45)

Here C>0C>0 is an absolute constant.

See [7, 8], also [10, Chapter III] for detailed proof of this lemma.

3  Existence and uniqueness of the weak solution

3.1 Construction of the flux carrier

In this subsection, we are devoted to the construction of a flux carrier 𝒂\boldsymbol{a}, which is divergence free, satisfying the Navier-slip boundary condition (1.2), and connects two Poiseuille flows in 𝒮L\mathcal{S}_{L} and 𝒮R\mathcal{S}_{R} smoothly. Meanwhile, the vector 𝒂\boldsymbol{a} will satisfy the following:

Proposition 3.1.

There exists a smooth vector field 𝐚(x)\boldsymbol{a}(x) which enjoys the following properties

  • (i).

    𝒂C(𝒮¯)\boldsymbol{a}\in C^{\infty}(\overline{\mathcal{S}}), and 𝒂=0\nabla\cdot\boldsymbol{a}=0 in 𝒮\mathcal{S};

  • (ii).

    2(𝕊𝒂𝒏)tan+α𝒂tan=02(\mathbb{S}\boldsymbol{a}\cdot\boldsymbol{n})_{\mathrm{tan}}+\alpha\boldsymbol{a}_{\mathrm{tan}}=0, and 𝒂𝒏=0\boldsymbol{a}\cdot\boldsymbol{n}=0 on 𝒮\partial\mathcal{S};

  • (iii).

    For a fixed ε(0,1)\varepsilon\in(0,1) ,

    𝒂={𝑷ΦL(y) in 𝒮{y1e2ε},𝑷ΦR(x) in 𝒮{x1e2ε}.\boldsymbol{a}=\left\{\begin{aligned} &\boldsymbol{P}^{L}_{\Phi}(y)\quad\text{ in }\quad\mathcal{S}\cap\{y_{1}\leq-e^{\frac{2}{\varepsilon}}\},\\ &\boldsymbol{P}^{R}_{\Phi}(x)\quad\text{ in }\quad\mathcal{S}\cap\{x_{1}\geq e^{\frac{2}{\varepsilon}}\}.\end{aligned}\right. (3.46)

    Moreover, for any vector filed 𝒗𝒱\boldsymbol{v}\in\mathcal{V} with

    𝒱:={𝒗H1(𝒮):div𝒗=0,(𝒗𝒏)|𝒮=0},\mathcal{V}:=\left\{\boldsymbol{v}\in H^{1}(\mathcal{S}):\,\operatorname{div}\boldsymbol{v}=0,\ (\boldsymbol{v}\cdot\boldsymbol{n})\big{|}_{\partial\mathcal{S}}=0\right\}, (3.47)

    there exists a constant CC, independent of ε\varepsilon and α\alpha, such that

    |𝒮𝒗𝒂𝒗dx|C𝒮Φ(ε+α1+α)𝒗L2(𝒮)2.\left|\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx\right|\leq C_{\mathcal{S}}\Phi\left(\varepsilon+\frac{\alpha}{1+\alpha}\right)\|\nabla\boldsymbol{v}\|^{2}_{L^{2}(\mathcal{S})}. (3.48)

The following lemma is useful in the construction of 𝒂\boldsymbol{a}.

Lemma 3.2.

There exists a smooth non-decreasing function σε:[0,δ)[Φ,0]\sigma_{\varepsilon}:[0,\delta)\to[-\Phi,0], where 0<ε<<10<\varepsilon<<1, such that

σε(t)={0,fortε;Φ,for0tϵ.\sigma_{\varepsilon}(t)=\left\{\begin{aligned} 0,&\quad\text{for}\quad t\geq\varepsilon;\\ -\Phi,&\quad\text{for}\quad 0\leq t\leq\epsilon.\end{aligned}\right.

Here ϵ:=εe1/ε/3\epsilon:=\varepsilon e^{-1/\varepsilon}/3, and

δ:=dist(𝒮,𝒮ob𝒮+)=inf{|xz|:x𝒮,z𝒮ob𝒮+}>2ε.\delta:=\mathrm{dist}(\partial\mathcal{S}_{-},\partial\mathcal{S}_{ob}\cup\partial\mathcal{S}_{+})=\inf\left\{|x-z|:\,x\in\partial\mathcal{S}_{-},\,z\in\partial\mathcal{S}_{ob}\cup\partial\mathcal{S}_{+}\right\}>2\varepsilon.

Meanwhile, when t[0,ε]t\in\left[0,\varepsilon\right], there exists constant CC such that

{0σε(t)min{Φe1/ε,2Φεt},|σε(k)(t)|CΦe1/ε(ε1e1/ε)k1,for k=2,3.\left\{\begin{aligned} &0\leq\sigma_{\varepsilon}^{\prime}(t)\leq\min\left\{\Phi e^{1/\varepsilon},\frac{2\Phi\varepsilon}{t}\right\},\\[2.84526pt] &\left|\sigma_{\varepsilon}^{(k)}(t)\right|\leq C\Phi e^{1/\varepsilon}(\varepsilon^{-1}e^{1/\varepsilon})^{k-1},\quad\text{for }k=2,3.\\ \end{aligned}\right. (3.49)
Proof.

We start with the piecewise smooth function

τε(t)={0,fortε;εt,forεe1/ε<t<ε;0,for0tεe1/ε.\tau_{\varepsilon}(t)=\left\{\begin{aligned} 0,&\quad\text{for}\quad t\geq\varepsilon;\\ \frac{\varepsilon}{t},&\quad\text{for}\quad\varepsilon e^{-1/\varepsilon}<t<\varepsilon;\\ 0,&\quad\text{for}\quad 0\leq t\leq\varepsilon e^{-1/\varepsilon}.\end{aligned}\right. (3.50)

Then denoting ς\varsigma the classical mollifier with radius equals ϵ\epsilon, the function σε\sigma_{\varepsilon} is given by:

σε(t):={Φ,for0t<ϵ;Φ+C~Φ0t(ςτε)(s)𝑑s,forϵ<t<δϵ;0,forδϵt<δ.\sigma_{\varepsilon}(t):=\left\{\begin{array}[]{ll}-\Phi,&\quad\text{for}\quad 0\leq t<\epsilon;\\[5.69054pt] -\Phi+\tilde{C}\Phi\int_{0}^{t}(\varsigma*\tau_{\varepsilon})(s)ds,&\quad\text{for}\quad\epsilon<t<\delta-\epsilon;\\[5.69054pt] 0,&\quad\text{for}\quad\delta-\epsilon\leq t<\delta.\\ \end{array}\right. (3.51)

Here C~>0\tilde{C}>0 is chosen such that

C~0δϵ(ςτε)(s)𝑑s=1.\tilde{C}\int_{0}^{\delta-\epsilon}(\varsigma*\tau_{\varepsilon})(s)ds=1.

Noting that C~\tilde{C} must be sufficiently close to 11, since 0δτε(s)𝑑s=1\int_{0}^{\delta}\tau_{\varepsilon}(s)ds=1 by (3.50). Finally, (3.49)1 follows directly from (3.50) and (3.51), while the validity of (3.49)2 follows that

σε(k)L=C~Φς(k1)τεLCΦς(k1)L1τεLCΦe1/ε(ε1e1/ε)k1,for k=2,3.\left\|\sigma_{\varepsilon}^{(k)}\right\|_{L^{\infty}}=\tilde{C}\Phi\left\|\varsigma^{(k-1)}*\tau_{\varepsilon}\right\|_{L^{\infty}}\leq C\Phi\left\|\varsigma^{(k-1)}\right\|_{L^{1}}\left\|\tau_{\varepsilon}\right\|_{L^{\infty}}\leq C\Phi e^{1/\varepsilon}(\varepsilon^{-1}e^{1/\varepsilon})^{k-1},\quad\text{for }k=2,3.

Proof of Proposition 3.1 : Given ε<δ2\varepsilon<\frac{\delta}{2}, we define

𝒂=σε(t)𝒆𝒔,in𝒮0,\boldsymbol{a}=\sigma^{\prime}_{\varepsilon}(t)\boldsymbol{e_{s}},\quad\text{in}\quad\mathcal{S}_{0}, (3.52)

where 𝒆𝒔\boldsymbol{e_{s}} is defined around (2.23), while

𝒂={[σε(x2)(1η(x1))+η(x1)PΦR(x2)]𝒆𝟏(η(x1)0x2(PΦR(ξ)σε(ξ))𝑑ξ)𝒆𝟐,in𝒮R;[σε(y2)(1η(y1))+η(y1)PΦL(y2)]𝒆𝟏+(η(y1)0y2(PΦL(ξ)σε(ξ))𝑑ξ)𝒆𝟐,in𝒮L.\boldsymbol{a}=\left\{\begin{array}[]{l}\left[\sigma^{\prime}_{\varepsilon}(x_{2})(1-\eta(x_{1}))+\eta(x_{1})P^{R}_{\Phi}(x_{2})\right]\boldsymbol{e_{1}}-\left(\eta^{\prime}(x_{1})\int_{0}^{x_{2}}\left(P^{R}_{\Phi}(\xi)-\sigma^{\prime}_{\varepsilon}(\xi)\right)d\xi\right)\boldsymbol{e_{2}},\\ \hskip 384.1122pt\text{in}\quad\mathcal{S}_{R};\\[8.53581pt] \left[\sigma^{\prime}_{\varepsilon}(y_{2})(1-\eta(-y_{1}))+\eta(-y_{1})P^{L}_{\Phi}(y_{2})\right]\boldsymbol{e_{1}}^{\prime}+\left(\eta^{\prime}(-y_{1})\int_{0}^{y_{2}}\left(P^{L}_{\Phi}(\xi)-\sigma^{\prime}_{\varepsilon}(\xi)\right)d\xi\right)\boldsymbol{e_{2}}^{\prime},\\ \hskip 384.1122pt\text{in}\quad\mathcal{S}_{L}.\\ \end{array}\right. (3.53)

Here η=η(s)\eta=\eta(s) be the smooth cut-off functions such that

η(s)={1,fors>e2/ε;0,fors<0.\eta(s)=\left\{\begin{aligned} &1,\quad\text{for}\quad s>e^{2/\varepsilon};\\ &0,\quad\text{for}\quad s<0.\\ \end{aligned}\right. (3.54)

and η\eta satisfies

|η|2e2/ε,and|η′′|4e4/ε.|\eta^{\prime}|\leq 2e^{-2/\varepsilon},\quad\text{and}\quad|\eta^{\prime\prime}|\leq 4e^{-4/\varepsilon}.

PΦLP^{L}_{\Phi} and PΦRP^{R}_{\Phi}, which are given in (1.6), are 𝒆𝟏\boldsymbol{e_{1}}-component and 𝒆𝟏\boldsymbol{e_{1}}^{\prime}-component of Poiseuille flows in pipes 𝒮L\mathcal{S}_{L} and 𝒮R\mathcal{S}_{R}, respectively.

Using (2.23) and (2.26)1, the flux carrier 𝒂\boldsymbol{a} constructed in (3.53) is smooth and divergence-free. Meanwhile, since σε(t)=0\sigma_{\varepsilon}(t)=0 near t=0t=0, one has 𝒂\boldsymbol{a} vanishes near 𝒮𝒮0\partial\mathcal{S}\cap\partial\mathcal{S}_{0}. This indicates 𝒂\boldsymbol{a} satisfies the homogeneous Navier-slip boundary condition on 𝒮𝒮0\partial\mathcal{S}\cap\partial\mathcal{S}_{0}.

Now we go to verify that 𝒂\boldsymbol{a} meets the Navier-slip boundary condition on 𝒮(𝒮L𝒮R)\partial\mathcal{S}\cap(\mathcal{S}_{L}\cup\mathcal{S}_{R}). Owing to cases in (3.53)1,2\eqref{Cons}_{1,2} are similar, we only consider (3.53)1\eqref{Cons}_{1} for simplicity. Since in this part, 𝒮R\partial\mathcal{S}_{R} is straight, direct calculation of the Navier-slip boundary condition is to check

{x2a1(x1,0)+αa1(x1,0)=x2a1(x1,1)+αa1(x1,1)=0;a2(x1,0)=a2(x1,1)=0,x1(0,).\left\{\begin{aligned} &-\partial_{x_{2}}a_{1}(x_{1},0)+\alpha a_{1}(x_{1},0)=\partial_{x_{2}}a_{1}(x_{1},1)+\alpha a_{1}(x_{1},1)=0;\\[5.69054pt] &a_{2}(x_{1},0)=a_{2}(x_{1},1)=0,\\ \end{aligned}\right.\quad\quad\forall x_{1}\in(0,\infty).

This could be done by the definition of PΦRP^{R}_{\Phi} in (1.6), the construction of σε\sigma_{\varepsilon} above, and direct calculations.

Then items (i) and item (ii) in Proposition 3.1 is proven and also it is easy to check that (3.46) stands due to the choice of the cutoff function η(x1)\eta(x_{1}). Now it remains to derive (3.48), we define

V:=𝒮𝒗𝒂𝒗dx=𝒮L𝒮0𝒮R𝒗𝒂𝒗dx.V:=\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx=\int_{\mathcal{S}_{L}\cup\mathcal{S}_{0}\cup\mathcal{S}_{R}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx.

Estimates of the 𝒮0\mathcal{S}_{0}-part integration:

In the distorted part 𝒮0\mathcal{S}_{0}, we denote that

𝒗=vs(s,t)𝒆𝒔+vt(s,t)𝒆𝒕,\boldsymbol{v}=v_{s}(s,t)\boldsymbol{e_{s}}+v_{t}(s,t)\boldsymbol{e_{t}},

where the coordinates (s,t)(s,t) and vectors 𝒆𝒔\boldsymbol{e_{s}}, 𝒆𝒕\boldsymbol{e_{t}} are defined in Section 2. Noting that

𝒆𝒕=t,𝒆𝒔=γ(s,t)s,d𝒆𝒔ds=κ(s,t)γ(s,t)𝒆𝒕,\boldsymbol{e_{t}}\cdot\nabla=\partial_{t}\,,\quad\boldsymbol{e_{s}}\cdot\nabla=\gamma(s,t)\partial_{s}\,,\quad\frac{d\boldsymbol{e_{s}}}{ds}=\frac{\kappa(s,t)}{\gamma(s,t)}\boldsymbol{e_{t}}\,,

in 𝒮0\mathcal{S}_{0}, one derives

𝒗𝒂=γ(s,t)vs(s,t)s(σε(t)𝒆𝒔)+vt(s,t)t(σε(t)𝒆𝒔)=κ(s,t)vs(s,t)σε(t)𝒆𝒕+vt(s,t)σε′′(t)𝒆𝒔.\begin{split}\boldsymbol{v}\cdot\nabla\boldsymbol{a}&=\gamma(s,t)v_{s}(s,t)\frac{\partial}{\partial s}(\sigma_{\varepsilon}^{\prime}(t)\boldsymbol{e_{s}})+v_{t}(s,t)\frac{\partial}{\partial t}(\sigma_{\varepsilon}^{\prime}(t)\boldsymbol{e_{s}})\\ &=\kappa(s,t)v_{s}(s,t)\sigma_{\varepsilon}^{\prime}(t)\boldsymbol{e_{t}}+v_{t}(s,t)\sigma_{\varepsilon}^{\prime\prime}(t)\boldsymbol{e_{s}}\,.\end{split}

Hence, we have

𝒗𝒂𝒗=(σε′′(t)+κ(s,t)σε(t))vs(s,t)vt(s,t),in𝒮0.\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}=\left(\sigma_{\varepsilon}^{\prime\prime}(t)+\kappa(s,t)\sigma_{\varepsilon}^{\prime}(t)\right)v_{s}(s,t)v_{t}(s,t),\quad\text{in}\quad\mathcal{S}_{0}.

Recalling (2.19), we deduce

𝒮0𝒗𝒂𝒗dx=0δs0s0(σε′′(t)+κ(s,t)σε(t))vs(s,t)vt(s,t)1γ(s,t)𝑑s𝑑t=0δs0s0σε′′(t)vs(s,t)vt(s,t)1γ(s,t)𝑑s𝑑t+0δs0s0κ(s,t)γ(s,t)σε(t)vs(s,t)vt(s,t)𝑑s𝑑t.\begin{split}\int_{\mathcal{S}_{0}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx=&\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\left(\sigma_{\varepsilon}^{\prime\prime}(t)+\kappa(s,t)\sigma_{\varepsilon}^{\prime}(t)\right)v_{s}(s,t)v_{t}(s,t)\frac{1}{\gamma(s,t)}dsdt\\[2.84526pt] =&\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\sigma_{\varepsilon}^{\prime\prime}(t)v_{s}(s,t)v_{t}(s,t)\frac{1}{\gamma(s,t)}dsdt+\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\frac{\kappa(s,t)}{\gamma(s,t)}\sigma_{\varepsilon}^{\prime}(t)v_{s}(s,t)v_{t}(s,t)dsdt.\end{split}

Noting that σε(t)\sigma^{\prime}_{\varepsilon}(t) vanishes near 𝒮𝒮0\partial\mathcal{S}\cap\partial\mathcal{S}_{0}, integration by parts for the first term of the right hand side of the above equality on tt indicate that

𝒮0𝒗𝒂𝒗dx=0δs0s0σε(t)t(vsγ)vtdsdt0δs0s0σε(t)vstvt1γ(t,s)dsdt+0δs0s0κ(s,t)γ(s,t)σε(t)vs(s,t)vt(s,t)𝑑s𝑑t.\begin{split}\int_{\mathcal{S}_{0}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx=&-\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\sigma_{\varepsilon}^{\prime}(t)\partial_{t}\left(\frac{v_{s}}{\gamma}\right)v_{t}dsdt-\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\sigma_{\varepsilon}^{\prime}(t)v_{s}\partial_{t}v_{t}\frac{1}{\gamma(t,s)}dsdt\\ &+\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\frac{\kappa(s,t)}{\gamma(s,t)}\sigma_{\varepsilon}^{\prime}(t)v_{s}(s,t)v_{t}(s,t)dsdt.\end{split} (3.55)

Recalling (2.26)1, the divergence-free property of 𝒗\boldsymbol{v} in the curvilinear coordinates follows:

γsvs+tvtκvt=0.\gamma\partial_{s}v_{s}+\partial_{t}v_{t}-\kappa v_{t}=0.

Then inserting the divergence-free property into (3.55), we obtain that

𝒮0𝒗𝒂𝒗dx=0δs0s0σε(t)t(vsγ)vtdsdt+0δs0s0σε(t)vssvsdsdt:=V1+V2.\begin{split}\int_{\mathcal{S}_{0}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx&=-\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\sigma_{\varepsilon}^{\prime}(t)\partial_{t}\left(\frac{v_{s}}{\gamma}\right)v_{t}dsdt+\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\sigma_{\varepsilon}^{\prime}(t)v_{s}\partial_{s}v_{s}dsdt\\ &:=V_{1}+V_{2}.\end{split} (3.56)

First, noting that γ(s,t)>γ0>0\gamma(s,t)>\gamma_{0}>0 is smooth, and σε(t)\sigma_{\varepsilon}^{\prime}(t), which is supported on [0,ε][0,\varepsilon], satisfies

|σε(t)|2Φεt,t[0,ε],|\sigma_{\varepsilon}^{\prime}(t)|\leq\frac{2\Phi\varepsilon}{t},\quad\forall t\in[0,\varepsilon],

one bounds V1V_{1} by using the Cauchy-Schwarz inequality and the Poincaré inequality in Lemma 2.4

|V1|CΦε(0δs0s0|𝒗|2𝑑s𝑑t)1/2(0δs0s0|vt|2t2𝑑s𝑑t)1/2V11.\begin{split}|V_{1}|\leq&C\Phi\varepsilon\left(\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}|\nabla\boldsymbol{v}|^{2}dsdt\right)^{1/2}\underbrace{\left(\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\frac{|v_{t}|^{2}}{t^{2}}dsdt\right)^{1/2}}_{V_{11}}.\end{split} (3.57)

Due to vt=0v_{t}=0 on the t=0t=0, the part V11V_{11} can be estimated by the one dimensional Hardy inequality. In fact

0δ|vt(s,t)|2t2𝑑t=0δ|vt(s,t)|2(1t)𝑑t=|vt(s,δ)|2δ0δ|vt(s,t)|2t2𝑑t+20δvt(s,t)ttvt(s,t)dt2(0δ|vt(s,t)|2t2𝑑t)1/2(0δ|tvt(s,t)|2𝑑t)1/2,\begin{split}\int_{0}^{\delta}\frac{|v_{t}(s,t)|^{2}}{t^{2}}dt&=-\int_{0}^{\delta}|v_{t}(s,t)|^{2}\left(\frac{1}{t}\right)^{\prime}dt\\ &=-\frac{|v_{t}(s,\delta)|^{2}}{\delta}\int_{0}^{\delta}\frac{|v_{t}(s,t)|^{2}}{t^{2}}dt+2\int_{0}^{\delta}\frac{v_{t}(s,t)}{t}\partial_{t}v_{t}(s,t)dt\\ &\leq 2\left(\int_{0}^{\delta}\frac{|v_{t}(s,t)|^{2}}{t^{2}}dt\right)^{1/2}\left(\int_{0}^{\delta}|\partial_{t}v_{t}(s,t)|^{2}dt\right)^{1/2},\end{split}

which indicates

0δ|vt(s,t)|2t2𝑑t40δ|tvt(s,t)|2𝑑t.\int_{0}^{\delta}\frac{|v_{t}(s,t)|^{2}}{t^{2}}dt\leq 4\int_{0}^{\delta}|\partial_{t}v_{t}(s,t)|^{2}dt. (3.58)

Thus one concludes

|V1|CΦε𝒗L2(𝒮0)2|V_{1}|\leq C\Phi\varepsilon\|\nabla\boldsymbol{v}\|_{L^{2}(\mathcal{S}_{0})}^{2}

by combining (3.57) and (3.58). For V2V_{2} in (3.56), it follows that

V2\displaystyle V_{2} =120δs0s0σε(t)s(vs)2dsdt\displaystyle=\frac{1}{2}\int_{0}^{\delta}\int_{-s_{0}}^{s_{0}}\sigma_{\varepsilon}^{\prime}(t)\partial_{s}(v_{s})^{2}dsdt (3.59)
=1201σε(x2)(v1)2(0,x2)𝑑x2120c0σε(y2)(v1)2(0,y2)𝑑y2.\displaystyle=\frac{1}{2}\int_{0}^{1}\sigma_{\varepsilon}^{\prime}(x_{2})(v_{1})^{2}(0,x_{2})dx_{2}-\frac{1}{2}\int_{0}^{c_{0}}\sigma_{\varepsilon}^{\prime}(y_{2})(v_{1})^{2}(0,y_{2})dy_{2}.

The second equality above is established due to the Newton-Leibniz formula and fact that the curvilinear coordinates (s,t)(s,t) turns to be Euclidean in 𝒮\𝒮0\mathcal{S}\backslash\mathcal{S}_{0}.

Estimates in 𝒮L\mathcal{S}_{L} and 𝒮R\mathcal{S}_{R}.

The cases in subsets 𝒮L\mathcal{S}_{L} and 𝒮R\mathcal{S}_{R} are similar, thus we only discuss the latter one for simplicity. At the beginning, we denote that

𝒮R:=𝒮R1𝒮R2,\mathcal{S}_{R}:=\mathcal{S}_{R1}\cup\mathcal{S}_{R2},

where

{𝒮R1=𝒮{x1[0,e2/ε]};𝒮R2=𝒮{x1(e2/ε,)}.\left\{\begin{aligned} &\mathcal{S}_{R1}=\mathcal{S}\cap\{x_{1}\in[0,\,e^{2/\varepsilon}]\};\\ &\mathcal{S}_{R2}=\mathcal{S}\cap\{x_{1}\in(e^{2/\varepsilon},\infty)\}.\end{aligned}\right.

Direct calculation shows

𝒮R𝒗𝒂𝒗dx=\displaystyle\int_{\mathcal{S}_{R}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx= 𝒮R1η(x1)(PΦR(x2)σε(x2))((v1)2(v2)2)𝑑x\displaystyle\int_{\mathcal{S}_{R1}}\eta^{\prime}(x_{1})\left(P^{R}_{\Phi}(x_{2})-\sigma^{\prime}_{\varepsilon}(x_{2})\right)\left((v_{1})^{2}-(v_{2})^{2}\right)dx
+𝒮R1η′′(x1)(0x2(σε(ξ)PΦR(ξ))𝑑ξ)v1v2𝑑x\displaystyle+\int_{\mathcal{S}_{R1}}\eta^{\prime\prime}(x_{1})\left(\int_{0}^{x_{2}}\left(\sigma^{\prime}_{\varepsilon}(\xi)-P^{R}_{\Phi}(\xi)\right)d\xi\right)v_{1}v_{2}dx
+𝒮R1σε′′(x2)(1η(x1))v1v2𝑑x+𝒮R2η(x1)(PΦR)(x2)v1v2𝑑x\displaystyle+\int_{\mathcal{S}_{R1}}\sigma_{\varepsilon}^{\prime\prime}(x_{2})\left(1-\eta(x_{1})\right)v_{1}v_{2}dx+\int_{\mathcal{S}_{R2}}\eta(x_{1})(P_{\Phi}^{R})^{\prime}(x_{2})v_{1}v_{2}dx
:=\displaystyle:= J1+J2+J3+J4.\displaystyle J_{1}+J_{2}+J_{3}+J_{4}.

Here, noticing that |η|2e2/ε|\eta^{\prime}|\leq 2e^{-2/\varepsilon}, |η′′|4e4ε|\eta^{\prime\prime}|\leq 4e^{-4\varepsilon} and

|σε(t)|Φe1/εt[0,δ)|\sigma_{\varepsilon}^{\prime}(t)|\leq\Phi e^{1/\varepsilon}\quad\forall t\in[0,\delta)

which follows from (3.49), one concludes that

|J1|+|J2|CΦe1/ε𝒮R1|𝒗|2𝑑xCΦε𝒗L2(𝒮)2|J_{1}|+|J_{2}|\leq C\Phi e^{-1/\varepsilon}\int_{\mathcal{S}_{R1}}|\boldsymbol{v}|^{2}dx\leq C\Phi\varepsilon\|\nabla\boldsymbol{v}\|_{L^{2}(\mathcal{S})}^{2}

by applying the Cauchy-Schwarz inequality and the Poincaré inequality in Corollary 2.5. Moreover, adopting integrating by parts, one deduces

J3=𝒮R1σε(x2)(1η(x1))x2v1v2dx𝒮R1σε(x2)(1η(x1))v1x2v2dx:=J31+J32.\begin{split}J_{3}=&-\int_{\mathcal{S}_{R1}}\sigma_{\varepsilon}^{\prime}(x_{2})\left(1-\eta(x_{1})\right)\partial_{x_{2}}v_{1}v_{2}dx-\int_{\mathcal{S}_{R1}}\sigma_{\varepsilon}^{\prime}(x_{2})\left(1-\eta(x_{1})\right)v_{1}\partial_{x_{2}}v_{2}dx\\ :=&J_{31}+J_{32}.\end{split}

Via an analogous route as we go through for V1V_{1} in (3.56) above, one deduces

|J31|CΦε𝒗L2(𝒮)2.|J_{31}|\leq C\Phi\varepsilon\|\nabla\boldsymbol{v}\|_{L^{2}(\mathcal{S})}^{2}.

For the term J32J_{32}, applying the divergence-free property of 𝒗\boldsymbol{v} and using integration by parts, one arrives

J32=\displaystyle J_{32}= 12𝒮R1σε(x2)(1η(x1))x1(v1)2dx\displaystyle\frac{1}{2}\int_{\mathcal{S}_{R1}}\sigma_{\varepsilon}^{\prime}(x_{2})\left(1-\eta(x_{1})\right)\partial_{x_{1}}(v_{1})^{2}dx
=\displaystyle= 12𝒮R1σε(x2)η(x1)(v1)2𝑑x1201σε(x2)(v1)2(0,x2)𝑑x2\displaystyle\frac{1}{2}\int_{\mathcal{S}_{R1}}\sigma_{\varepsilon}^{\prime}(x_{2})\eta^{\prime}(x_{1})(v_{1})^{2}dx-\frac{1}{2}\int_{0}^{1}\sigma_{\varepsilon}^{\prime}(x_{2})(v_{1})^{2}(0,x_{2})dx_{2}
:=\displaystyle:= J321+J322.\displaystyle J_{321}+J_{322}.

Noting that J321J_{321} can be estimated in the same way as we do on J1J_{1} and J2J_{2}, that is

|J321|CΦε𝒗L2(𝒮)2.|J_{321}|\leq C\Phi\varepsilon\|\nabla\boldsymbol{v}\|_{L^{2}(\mathcal{S})}^{2}.

Due to J322J_{322} being cancelled out with the first term in (3.59)2, it remains only to estimate J4J_{4}. Recall the LL^{\infty} bound of (PΦR)(P_{\Phi}^{R})^{\prime} in (1.7), one concludes that

J4CαΦ1+α𝒗NL2(𝒮)2.J_{4}\leq C\frac{\alpha\Phi}{1+\alpha}\|\nabla\boldsymbol{v}^{N}\|_{L^{2}(\mathcal{S})}^{2}.

Collecting the above estimates and cancellations, we derive that

|V|=|𝒮𝒗𝒂𝒗dx|CΦ(ε+α1+α)𝒗L2(𝒮)2,|V|=\left|\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}dx\right|\leq C\Phi\left(\varepsilon+\frac{\alpha}{1+\alpha}\right)\|\nabla\boldsymbol{v}\|_{L^{2}(\mathcal{S})}^{2},

which concludes (3.48). This completes the proof of Proposition 3.1.

3.2 Existence of the weak solution

We will look for a solution to (1.1)-(1.10) of the form

𝒖=𝒗+𝒂.\boldsymbol{u}=\boldsymbol{v}+\boldsymbol{a}. (3.60)

Thus, our problem turns to the following equivalent form:

Problem 3.3.

Find (𝐯,p)(\boldsymbol{v},p) such that

{𝒗𝒗+𝒂𝒗+𝒗𝒂+pΔ𝒗=Δ𝒂𝒂𝒂,𝒗=0,in 𝒮,\left\{\begin{aligned} &\boldsymbol{v}\cdot\nabla\boldsymbol{v}+\boldsymbol{a}\cdot\nabla\boldsymbol{v}+\boldsymbol{v}\cdot\nabla\boldsymbol{a}+\nabla p-\Delta\boldsymbol{v}=\Delta\boldsymbol{a}-\boldsymbol{a}\cdot\nabla\boldsymbol{a},\\ &\nabla\cdot\boldsymbol{v}=0,\\ \end{aligned}\right.\quad\text{in }\quad\mathcal{S},

subject to the Navier-slip boundary condition

{2(𝕊𝒗𝒏)tan+α𝒗tan=0,𝒗𝒏=0,on 𝒮,\left\{\begin{aligned} &2(\mathbb{S}\boldsymbol{v}\cdot\boldsymbol{n})_{\mathrm{tan}}+\alpha\boldsymbol{v}_{\mathrm{tan}}=0,\\ &\boldsymbol{v}\cdot\boldsymbol{n}=0,\\ \end{aligned}\right.\quad\text{on }\quad\partial\mathcal{S}, (3.61)

with the asymptotic behavior as |x||x|\to\infty

𝒗(x)𝟎,as|x|.\boldsymbol{v}(x)\to\boldsymbol{0},\quad\text{as}\quad|x|\to\infty.

From the weak formulation (1.11), we have that 𝒗\boldsymbol{v} satisfies the following weak formulation:

Definition 3.4.

Let 𝐚\boldsymbol{a} be a smooth vector satisfying the properties stated in the above. We say that 𝐯σ(𝒮)\boldsymbol{v}\in\mathcal{H}_{\sigma}(\mathcal{S}) is a weak solution of Problem 3.3 if

2𝒮𝕊𝒗:𝕊𝝋dx+α𝒮𝒗tan𝝋tan𝑑S+𝒮𝒗𝒗𝝋dx+𝒮𝒗𝒂𝝋dx+𝒮𝒂𝒗𝝋dx=𝒮(Δ𝒂𝒂𝒂)𝝋𝑑x\begin{split}&2\int_{\mathcal{S}}\mathbb{S}\boldsymbol{v}:\mathbb{S}\boldsymbol{\varphi}dx+\alpha\int_{\partial\mathcal{S}}\boldsymbol{v}_{\mathrm{tan}}\cdot\boldsymbol{\varphi}_{\mathrm{tan}}dS+\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\boldsymbol{v}\cdot\boldsymbol{\varphi}dx+\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{\varphi}dx\\ +&\int_{\mathcal{S}}\boldsymbol{a}\cdot\nabla\boldsymbol{v}\cdot\boldsymbol{\varphi}dx=\int_{\mathcal{S}}\big{(}\Delta\boldsymbol{a}-\boldsymbol{a}\cdot\nabla\boldsymbol{a}\big{)}\cdot\boldsymbol{\varphi}dx\end{split} (3.62)

holds for any vector-valued function 𝛗σ(𝒮)\boldsymbol{\varphi}\in\mathcal{H}_{\sigma}(\mathcal{S}).

Now we state our main result of this part.

Theorem 3.5.

There is a constant Φ0>0\Phi_{0}>0 depending on the curvature of 𝒮\partial\mathcal{S} such that if αΦ1+α<Φ0\frac{\alpha\Phi}{1+\alpha}<\Phi_{0}, then Problem 3.3 admits at least one weak solution (𝐯,p)σ(𝒮)×Lloc2(𝒮¯)(\boldsymbol{v},p)\in\mathcal{H}_{\sigma}(\mathcal{S})\times L^{2}_{\mathrm{loc}}(\overline{\mathcal{S}}), with

𝒗H1(𝒮)C(𝒂𝒂L2(𝒮eCΦ)+Δ𝒂L2(𝒮eCΦ))ΦeCΦ.\|\boldsymbol{v}\|_{H^{1}(\mathcal{S})}\leq C(\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{C\Phi}})}+\|\Delta\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{C\Phi}})})\leq\Phi e^{C\Phi}\,. (3.63)

Proof.

Set

𝑿:=Cσ,c(𝒮¯;2)={𝝋Cc(𝒮¯;2):𝝋=0,𝝋𝒏|𝒮=0},\boldsymbol{X}:=C^{\infty}_{\sigma,c}(\overline{\mathcal{S}};\,\mathbb{R}^{2})=\big{\{}\boldsymbol{\varphi}\in C^{\infty}_{c}(\overline{\mathcal{S}}\,;\,\mathbb{R}^{2}):\,\nabla\cdot\boldsymbol{\varphi}=0,\ \boldsymbol{\varphi}\cdot\boldsymbol{n}\big{|}_{\partial\mathcal{S}}=0\big{\}},

and {𝝋k}k=1𝑿\{\boldsymbol{\varphi}_{k}\}_{k=1}^{\infty}\subset\boldsymbol{X} be an unit orthonormal basis of σ(𝒮)\mathcal{H}_{\sigma}(\mathcal{S}), that is:

𝝋i,𝝋jH1(𝒮)={1,if i=j;0,if ij,\langle\boldsymbol{\varphi}_{i},\boldsymbol{\varphi}_{j}\rangle_{H^{1}(\mathcal{S})}=\begin{cases}1,&\quad\text{if }\quad i=j;\\ 0,&\quad\text{if }\quad i\neq j,\\ \end{cases}

i,j\forall i,j\in\mathbb{N}. We look for an approximation of 𝒗\boldsymbol{v} of the form

𝒗N(x)=i=1NciN𝝋i(x).\boldsymbol{v}^{N}(x)=\sum_{i=1}^{N}c_{i}^{N}\boldsymbol{\varphi}_{i}(x).

Testing the weak formulation (3.62) by 𝝋i\boldsymbol{\varphi}_{i}, with i=1,2,,Ni=1,2,...,N, one has

2i=1NciN𝒮𝕊𝝋i:𝕊𝝋jdx+αi=1NciN𝒮(𝝋i)tan(𝝋j)tan𝑑S+i,k=1NciNckN𝒮𝝋i𝝋k𝝋jdx+i=1N𝒮𝝋i𝒂𝝋jdx+i=1NciN𝒮𝒂𝝋i𝝋jdx=𝒮(Δ𝒂𝒂𝒂)𝝋j𝑑x,j=1,2,,N.\begin{split}&2\sum_{i=1}^{N}c_{i}^{N}\int_{\mathcal{S}}\mathbb{S}\boldsymbol{\varphi}_{i}:\mathbb{S}\boldsymbol{\varphi}_{j}dx+\alpha\sum_{i=1}^{N}c_{i}^{N}\int_{\partial\mathcal{S}}(\boldsymbol{\varphi}_{i})_{\mathrm{tan}}(\boldsymbol{\varphi}_{j})_{\mathrm{tan}}dS+\sum_{i,k=1}^{N}c_{i}^{N}c_{k}^{N}\int_{\mathcal{S}}\boldsymbol{\varphi}_{i}\cdot\nabla\boldsymbol{\varphi}_{k}\cdot\boldsymbol{\varphi}_{j}dx\\ +&\sum_{i=1}^{N}\int_{\mathcal{S}}\boldsymbol{\varphi}_{i}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{\varphi}_{j}dx+\sum_{i=1}^{N}c_{i}^{N}\int_{\mathcal{S}}\boldsymbol{a}\cdot\nabla\boldsymbol{\varphi}_{i}\cdot\boldsymbol{\varphi}_{j}dx=\int_{\mathcal{S}}\big{(}\Delta\boldsymbol{a}-\boldsymbol{a}\cdot\nabla\boldsymbol{a}\big{)}\cdot\boldsymbol{\varphi}_{j}dx,\quad\forall j=1,2,...,N.\end{split}

This is a system of nonlinear algebraic equations of NN-dimensional vector

𝒄N:=(c1N,c2N,,cNN).\boldsymbol{c}^{N}:=(c_{1}^{N},c_{2}^{N},...,c_{N}^{N}).

We denote P:NN{P}:\,\mathbb{R}^{N}\to\mathbb{R}^{N} such that

(P(𝒄N))j=2i=1NciN𝒮𝕊𝝋i:𝕊𝝋jdx+αi=1NciN𝒮(𝝋i)tan(𝝋j)tan𝑑S+i,k=1NciNckN𝒮𝝋i𝝋k𝝋jdx+i=1N𝒮𝝋i𝒂𝝋jdx+i=1NciN𝒮𝒂𝝋i𝝋jdx𝒮(Δ𝒂𝒂𝒂)𝝋j𝑑x,j=1,2,,N.\begin{split}\big{(}{P}(\boldsymbol{c}^{N})\big{)}_{j}=&2\sum_{i=1}^{N}c_{i}^{N}\int_{\mathcal{S}}\mathbb{S}\boldsymbol{\varphi}_{i}:\mathbb{S}\boldsymbol{\varphi}_{j}dx+\alpha\sum_{i=1}^{N}c_{i}^{N}\int_{\partial\mathcal{S}}(\boldsymbol{\varphi}_{i})_{\mathrm{tan}}\cdot(\boldsymbol{\varphi}_{j})_{\mathrm{tan}}dS+\sum_{i,k=1}^{N}c_{i}^{N}c_{k}^{N}\int_{\mathcal{S}}\boldsymbol{\varphi}_{i}\cdot\nabla\boldsymbol{\varphi}_{k}\cdot\boldsymbol{\varphi}_{j}dx\\ &+\sum_{i=1}^{N}\int_{\mathcal{S}}\boldsymbol{\varphi}_{i}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{\varphi}_{j}dx+\sum_{i=1}^{N}c_{i}^{N}\int_{\mathcal{S}}\boldsymbol{a}\cdot\nabla\boldsymbol{\varphi}_{i}\cdot\boldsymbol{\varphi}_{j}dx-\int_{\mathcal{S}}\big{(}\Delta\boldsymbol{a}-\boldsymbol{a}\cdot\nabla\boldsymbol{a}\big{)}\cdot\boldsymbol{\varphi}_{j}dx,\\ &\hskip 284.52756pt\quad\forall j=1,2,...,N.\end{split}

It is easy to check that

P(𝒄N)𝒄N=2𝒮|𝕊𝒗N|2𝑑x+α𝒮|(𝒗N)tan|2𝑑SA1+𝒮((𝒗N+𝒂)(𝒗N+𝒂))𝒗N𝑑xA2𝒮𝒗NΔ𝒂𝑑xA3.{P}(\boldsymbol{c}^{N})\cdot\boldsymbol{c}^{N}=\underbrace{2\int_{\mathcal{S}}|\mathbb{S}\boldsymbol{v}^{N}|^{2}dx+\alpha\int_{\partial\mathcal{S}}|(\boldsymbol{v}^{N})_{\mathrm{tan}}|^{2}dS}_{A_{1}}+\underbrace{\int_{\mathcal{S}}\left((\boldsymbol{v}^{N}+\boldsymbol{a})\cdot\nabla(\boldsymbol{v}^{N}+\boldsymbol{a})\right)\cdot\boldsymbol{v}^{N}dx}_{A_{2}}-\underbrace{\int_{\mathcal{S}}\boldsymbol{v}^{N}\cdot\Delta\boldsymbol{a}dx}_{A_{3}}.

By Lemma 2.6, we have

A1C0𝒮|𝒗N|2𝑑x.A_{1}\geq C_{0}\int_{\mathcal{S}}|\nabla\boldsymbol{v}^{N}|^{2}dx.

Next, by using integration by parts, together with the divergence-free property of 𝒗N\boldsymbol{v}^{N} and 𝒂\boldsymbol{a}, one knows that

A2=𝒮𝒗N𝒂𝒗NdxV+𝒮𝒂𝒂𝒗NdxK.A_{2}=\underbrace{\int_{\mathcal{S}}\boldsymbol{v}^{N}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}^{N}dx}_{V}+\underbrace{\int_{\mathcal{S}}\boldsymbol{a}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}^{N}dx}_{K}.

We now focus on the term VV. Applying (iii) in Proposition 3.1 to 𝒗N\boldsymbol{v}^{N}, one deduces

|V|C1Φ(ε+α1+α)𝒗NL2(𝒮)2,|V|\leq C_{1}\Phi\left(\varepsilon+\frac{\alpha}{1+\alpha}\right)\|\nabla\boldsymbol{v}^{N}\|_{L^{2}(\mathcal{S})}^{2}, (3.64)

where the constant C1C_{1} is independent with NN. For the term KK, since 𝒂\boldsymbol{a} equals to the Poiseuille flow 𝑷ΦL\boldsymbol{P}^{L}_{\Phi} or 𝑷ΦR\boldsymbol{P}^{R}_{\Phi} in 𝒮𝒮e2/ε\mathcal{S}-\mathcal{S}_{e^{2/\varepsilon}}, we have 𝒂𝒂0\boldsymbol{a}\cdot\nabla\boldsymbol{a}\equiv 0 in 𝒮𝒮e2/ε\mathcal{S}-\mathcal{S}_{e^{2/\varepsilon}}. Using the Cauchy-Schwarz inequality and the Poincaré inequality, one arrives at

|K|=|𝒮𝒂𝒂𝒗Ndx|C𝒂𝒂L2(𝒮e2/ε)𝒗NL2(𝒮).\begin{split}|K|=\left|\int_{\mathcal{S}}\boldsymbol{a}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{v}^{N}dx\right|\leq C\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}\|\nabla\boldsymbol{v}^{N}\|_{L^{2}(\mathcal{S})}\,.\end{split} (3.65)

Thus, by combining (3.64) and (3.65), one deduces

|A2|C1Φ(ε+α1+α)𝒗NL2(𝒮)2+C𝒂𝒂L2(𝒮e2/ε)𝒗NL2(𝒮).|A_{2}|\leq C_{1}\Phi\left(\varepsilon+\frac{\alpha}{1+\alpha}\right)\|\nabla\boldsymbol{v}^{N}\|_{L^{2}(\mathcal{S})}^{2}+C\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}\|\nabla\boldsymbol{v}^{N}\|_{L^{2}(\mathcal{S})}.

Finally, by the construction of the Poiseuille flow 𝑷ΦR\boldsymbol{P}_{\Phi}^{R} and 𝑷ΦL\boldsymbol{P}_{\Phi}^{L}, we have

(,e2/ε)×(0,c0)𝒗NΔ𝒂𝑑y=CLe2/ε0c0𝒗N𝒆𝟏𝑑y=0\int_{(-\infty,-e^{2/\varepsilon})\times(0,c_{0})}\boldsymbol{v}^{N}\cdot\Delta\boldsymbol{a}dy=-C_{L}\int^{-e^{2/\varepsilon}}_{-\infty}\int_{0}^{c_{0}}\boldsymbol{v}^{N}\cdot\boldsymbol{e_{1}}^{\prime}dy=0

and

(e2/ε,+)×(0,1)𝒗NΔ𝒂𝑑x=CRe2/ε01𝒗N𝒆𝟏𝑑x=0.\int_{(e^{2/\varepsilon},+\infty)\times(0,1)}\boldsymbol{v}^{N}\cdot\Delta\boldsymbol{a}dx=-C_{R}\int^{\infty}_{e^{2/\varepsilon}}\int_{0}^{1}\boldsymbol{v}^{N}\cdot\boldsymbol{e_{1}}dx=0\,.

Thus, by the Cauchy-Schwarz inequality and the Poincaré inequality, we deduce that

|A3|=|𝒮e2/ε𝒗NΔ𝒂𝑑x|CΔ𝒂L2(𝒮e2/ε)𝒗NL2(𝒮).|A_{3}|=\left|\int_{\mathcal{S}_{e^{2/\varepsilon}}}\boldsymbol{v}^{N}\cdot\Delta\boldsymbol{a}dx\right|\leq C\|\Delta\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}\|\nabla\boldsymbol{v}^{N}\|_{L^{2}(\mathcal{S})}.

Substituting the above estimates for A1A_{1}A3A_{3}, and choosing ε>0\varepsilon>0 being sufficiently small such that C1εΦ<C02C_{1}\varepsilon\Phi<\frac{C_{0}}{2}, one derives

P(𝒄N)𝒄N𝒗NH1(𝒮)((C02C1αΦ1+α)𝒗NH1(𝒮)C(𝒂𝒂L2(𝒮e2/ε)+Δ𝒂L2(𝒮e2/ε))),{P}(\boldsymbol{c}^{N})\cdot\boldsymbol{c}^{N}\geq\|\boldsymbol{v}^{N}\|_{H^{1}(\mathcal{S})}\left(\left(\frac{C_{0}}{2}-C_{1}\frac{\alpha\Phi}{1+\alpha}\right)\|\boldsymbol{v}^{N}\|_{H^{1}(\mathcal{S})}-C(\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}+\|\Delta\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})})\right),

which guarantees

P(𝒄N)𝒄N0,{P}(\boldsymbol{c}^{N})\cdot\boldsymbol{c}^{N}\geq 0,

provided

αΦ1+α<Φ0:=12C11C0and|𝒄N|=𝒗NH1(𝒮)C(𝒂𝒂L2(𝒮e2/ε)+Δ𝒂L2(𝒮e2/ε))C0/2C1αΦ/(1+α):=ρ.\frac{\alpha\Phi}{1+\alpha}<\Phi_{0}:=\frac{1}{2}C_{1}^{-1}C_{0}\quad\text{and}\quad|\boldsymbol{c}^{N}|=\|\boldsymbol{v}^{N}\|_{H^{1}(\mathcal{S})}\geq\frac{C\left(\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}+\|\Delta\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}\right)}{C_{0}/2-C_{1}\alpha\Phi/(1+\alpha)}:=\rho.

Using Lemma 2.10, there exists

(𝒗N)span{𝝋1,𝝋2,,𝝋N},and(𝒗N)H1(𝒮)C(𝒂𝒂L2(𝒮e2/ε)+Δ𝒂L2(𝒮e2/ε))C0/2C1αΦ/(1+α),(\boldsymbol{v}^{N})^{*}\in\text{span}\left\{\boldsymbol{\varphi}_{1},\boldsymbol{\varphi}_{2},...,\boldsymbol{\varphi}_{N}\right\},\quad\text{and}\quad\|(\boldsymbol{v}^{N})^{*}\|_{H^{1}(\mathcal{S})}\leq\frac{C\left(\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}+\|\Delta\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}\right)}{C_{0}/2-C_{1}\alpha\Phi/(1+\alpha)}, (3.66)

such that

2𝒮𝕊(𝒗N):𝕊ϕNdx+α𝒮(𝒗N)tan(ϕN)tandS+𝒮(𝒗N)(𝒗N)ϕNdx+𝒮(𝒗N)𝒂ϕNdx+𝒮𝒂(𝒗N)ϕNdx=𝒮(Δ𝒂𝒂𝒂)ϕNdx,ϕNspan{𝝋1,𝝋2,,𝝋N}.\begin{split}&2\int_{\mathcal{S}}\mathbb{S}(\boldsymbol{v}^{N})^{*}:\mathbb{S}\boldsymbol{\phi}_{N}dx+\alpha\int_{\partial\mathcal{S}}(\boldsymbol{v}^{N})^{*}_{\mathrm{tan}}\cdot(\boldsymbol{\phi}_{N})_{\mathrm{tan}}dS+\int_{\mathcal{S}}(\boldsymbol{v}^{N})^{*}\cdot\nabla(\boldsymbol{v}^{N})^{*}\cdot\boldsymbol{\phi}_{N}dx+\int_{\mathcal{S}}(\boldsymbol{v}^{N})^{*}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{\phi}_{N}dx\\ +&\int_{\mathcal{S}}\boldsymbol{a}\cdot\nabla(\boldsymbol{v}^{N})^{*}\cdot\boldsymbol{\phi}_{N}dx=\int_{\mathcal{S}}\big{(}\Delta\boldsymbol{a}-\boldsymbol{a}\cdot\nabla\boldsymbol{a}\big{)}\cdot\boldsymbol{\phi}_{N}dx,\quad\forall\boldsymbol{\phi}_{N}\in\text{span}\left\{\boldsymbol{\varphi}_{1},\boldsymbol{\varphi}_{2},...,\boldsymbol{\varphi}_{N}\right\}.\end{split} (3.67)

The above bound (3.66) and Rellich-Kondrachov embedding theorem imply the existence of a field 𝒗σ(𝒮)\boldsymbol{v}\in\mathcal{H}_{\sigma}(\mathcal{S}) and a subsequence, which we will always denote by (𝒗N)(\boldsymbol{v}^{N})^{*}, such that

(𝒗N)𝒗weakly in σ(𝒮)(\boldsymbol{v}^{N})^{*}\to\boldsymbol{v}\quad\text{weakly in $\mathcal{H}_{\sigma}(\mathcal{S})$}

and

(𝒗N)𝒗strongly in L2(𝒮), for all bounded 𝒮𝒮.(\boldsymbol{v}^{N})^{*}\to\boldsymbol{v}\quad\text{strongly in $L^{2}(\mathcal{S}^{\prime})$, for all bounded $\mathcal{S}^{\prime}\subset\mathcal{S}$}\,.

By passing to the limit in (3.67), one obtains

2𝒮𝕊𝒗:𝕊𝝋dx+α𝒮𝒗tan𝝋tan𝑑S+𝒮𝒗𝒗𝝋dx+𝒮𝒗𝒂𝝋dx+𝒮𝒂𝒗𝝋dx=𝒮(Δ𝒂𝒂𝒂)𝝋𝑑x,for any𝝋σ(𝒮).\begin{split}&2\int_{\mathcal{S}}\mathbb{S}\boldsymbol{v}:\mathbb{S}\boldsymbol{\varphi}dx+\alpha\int_{\partial\mathcal{S}}\boldsymbol{v}_{\mathrm{tan}}\cdot\boldsymbol{\varphi}_{\mathrm{tan}}dS+\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\boldsymbol{v}\cdot\boldsymbol{\varphi}dx+\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\boldsymbol{a}\cdot\boldsymbol{\varphi}dx\\ +&\int_{\mathcal{S}}\boldsymbol{a}\cdot\nabla\boldsymbol{v}\cdot\boldsymbol{\varphi}dx=\int_{\mathcal{S}}\big{(}\Delta\boldsymbol{a}-\boldsymbol{a}\cdot\nabla\boldsymbol{a}\big{)}\cdot\boldsymbol{\varphi}dx,\quad\text{for any}\quad\boldsymbol{\varphi}\in\mathcal{H}_{\sigma}(\mathcal{S}).\end{split} (3.68)

It follows from (3.66) and the Fatou lemma for weakly convergent sequences that

𝒗H1(𝒮)C(𝒂𝒂L2(𝒮e2/ε)+Δ𝒂L2(𝒮e2/ε)).\|\boldsymbol{v}\|_{H^{1}(\mathcal{S})}\leq C\left(\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}+\|\Delta\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}\right)\,. (3.69)

Now it remains to verify (3.63). From the construction of 𝒂\boldsymbol{a} in (3.52)–(3.53) and the estimate of σε\sigma_{\varepsilon} in (3.49), we have

|k𝒂|CΦe1ε(ε1e1ε)k,for k=0,1,2.\left|\nabla^{k}\boldsymbol{a}\right|\leq C\Phi e^{\frac{1}{\varepsilon}}(\varepsilon^{-1}e^{\frac{1}{\varepsilon}})^{k},\quad\text{for }k=0,1,2.

According to the construction of 𝒗\boldsymbol{v} given before, it is legal to choose ε=min{C04C1Φ,δ2}\varepsilon=\min\left\{\frac{C_{0}}{4C_{1}\Phi},\frac{\delta}{2}\right\}. This indicates that

𝒂𝒂L2(𝒮e2/ε)+Δ𝒂L2(𝒮e2/ε)Ce1/ε(Φ2ε1e3/ε+Φε2e3/ε)CΦ(1+Φ2eCΦ),\|\boldsymbol{a}\cdot\nabla\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}+\|\Delta\boldsymbol{a}\|_{L^{2}(\mathcal{S}_{e^{2/\varepsilon}})}\leq Ce^{1/\varepsilon}\left(\Phi^{2}\varepsilon^{-1}e^{3/\varepsilon}+\Phi\varepsilon^{-2}e^{3/\varepsilon}\right)\leq C\Phi\left(1+\Phi^{2}e^{C\Phi}\right),

which gives

𝒗H1(𝒮)ΦeCΦ.\|\boldsymbol{v}\|_{H^{1}(\mathcal{S})}\leq\Phi e^{C\Phi}\,.

Now we focus on the pressure. Let 𝒗\boldsymbol{v} be a weak solution of (3.62) constructed in the above. Using (3.68), one has 𝒖=𝒗+𝒂\boldsymbol{u}=\boldsymbol{v}+\boldsymbol{a} satisfies

𝒮𝒖ϕdx+𝒮𝒖𝒖ϕdx=0,for allϕ{𝒈Cc(𝒮;2):div 𝒈=0}.\int_{\mathcal{S}}\nabla\boldsymbol{u}\cdot\nabla\boldsymbol{\phi}\,dx+\int_{\mathcal{S}}\boldsymbol{u}\cdot\nabla\boldsymbol{u}\cdot\boldsymbol{\phi}\,dx=0,\quad\text{for all}\quad\boldsymbol{\phi}\in\{\boldsymbol{g}\in C_{c}^{\infty}(\mathcal{S};\mathbb{R}^{2}):\,\text{div }\boldsymbol{g}=0\}.

Thus by Lemma 2.12, there exists p(Cc(𝒮;))p\in\left(C_{c}^{\infty}(\mathcal{S};\mathbb{R})\right)^{\prime}, such that

Δ𝒖𝒖𝒖=p\Delta\boldsymbol{u}-\boldsymbol{u}\cdot\nabla\boldsymbol{u}=\nabla p (3.70)

in the sense of distribution. Furthermore, we have that (3.70) is equivalent to

div(𝒗𝒗𝒗𝒂𝒗𝒗𝒂)+Δ𝒂+CRη(x1)𝒆𝟏+CLη(y1)𝒆𝟏𝒂𝒂=Π,\text{div}\big{(}\nabla\boldsymbol{v}-\boldsymbol{v}\otimes\boldsymbol{v}-\boldsymbol{a}\otimes\boldsymbol{v}-\boldsymbol{v}\otimes\boldsymbol{a}\big{)}+\Delta\boldsymbol{a}+{C_{R}}{\eta(x_{1})}\boldsymbol{e_{1}}+{C_{L}}{\eta(-y_{1})}\boldsymbol{e_{1}}^{\prime}-\boldsymbol{a}\cdot\nabla\boldsymbol{a}=\nabla\Pi, (3.71)

with

Π=p+CRx1η(s)𝑑sCLy1η(s)𝑑s,\Pi=p+{C_{R}}{\int_{-\infty}^{x_{1}}\eta(s)ds}-{C_{L}}{\int_{-\infty}^{-y_{1}}\eta(s)ds}\,, (3.72)

where CLC_{L} and CRC_{R} are Poiseuille constants defined in (1.5)1 and (1.4)1, respectively. By the definition of 𝒂\boldsymbol{a}, one has both

Δ𝒂+CRη(x1)𝒆𝟏+CLη(y1)𝒆𝟏\Delta\boldsymbol{a}+{C_{R}}{\eta(x_{1})}\boldsymbol{e_{1}}+{C_{L}}{\eta(-y_{1})}\boldsymbol{e_{1}}^{\prime}

and 𝒂𝒂\boldsymbol{a}\cdot\nabla\boldsymbol{a} are smooth and have compact support. Since 𝒗H1(𝒮)\boldsymbol{v}\in H^{1}(\mathcal{S}) and 𝒂\boldsymbol{a} is uniformly bounded, one deduces

𝒗𝒗𝒗𝒂𝒗𝒗𝒂L2(𝒮),\nabla\boldsymbol{v}-\boldsymbol{v}\otimes\boldsymbol{v}-\boldsymbol{a}\otimes\boldsymbol{v}-\boldsymbol{v}\otimes\boldsymbol{a}\in L^{2}(\mathcal{S}),

directly by the Sobolev embedding and Hölder’s inequality. Therefore one concludes the left hand side of (3.71) belongs to H1(𝒮)H^{-1}(\mathcal{S}). Then applying Lemma 2.13, we have ΠLloc2(𝒮¯)\Pi\in L^{2}_{\mathrm{loc}}(\overline{\mathcal{S}}), which leads to pLloc2(𝒮¯)p\in L^{2}_{\mathrm{loc}}(\overline{\mathcal{S}}) by (3.72). ∎

3.3 Uniqueness result

The rest part of this section is devoted to the proof of uniqueness. We will show that the solution (𝒖,p)(\boldsymbol{u},{p}) constructed earlier in this section with its flux being Φ\Phi is unique for Φ\Phi being sufficiently small and independent of α\alpha.

3.3.1 Estimate of the pressure

Below, we give a proposition to show that an integration estimate related to the pressure in the truncated strip ΥZ+:=(𝒮Z\𝒮Z1){x1>0}\Upsilon_{Z}^{+}:=\left(\mathcal{S}_{Z}\backslash\mathcal{S}_{Z-1}\right)\cap\{x_{1}>0\} or ΥZ:=(𝒮Z\𝒮Z1){y1<0}\Upsilon_{Z}^{-}:=\left(\mathcal{S}_{Z}\backslash\mathcal{S}_{Z-1}\right)\cap\{y_{1}<0\}.

Proposition 3.6.

Let (𝐮~,p~)(\tilde{\boldsymbol{u}},\tilde{p}) be an alternative weak solution of (1.1) in the strip 𝒮\mathcal{S}, subject to the Navier-slip boundary condition (1.2). If the total flux

𝒮{x1=s}𝒖~(s,x2)𝒆𝟏𝑑x2=Φ=𝒮{x1=s}𝒖(s,x2)𝒆𝟏𝑑x2,for any s1,\int_{\mathcal{S}\cap\{x_{1}=s\}}\tilde{\boldsymbol{u}}(s,x_{2})\cdot\boldsymbol{e_{1}}dx_{2}=\Phi=\int_{\mathcal{S}\cap\{x_{1}=s\}}{\boldsymbol{u}}(s,x_{2})\cdot\boldsymbol{e_{1}}dx_{2},\quad\text{for any }s\geq 1,

then the following estimate of 𝐰:=𝐮~𝐮\boldsymbol{w}:=\tilde{\boldsymbol{u}}-{\boldsymbol{u}} and the pressure holds

|ΥK±(p~p)w1𝑑x|C(𝒖L4(ΥK±)𝒘L2(ΥK±)2+𝒘L2(ΥK±)2+𝒘L2(ΥK±)3),K2,\left|\int_{\Upsilon^{\pm}_{K}}(\tilde{p}-p)w_{1}dx\right|\leq C\left(\|\boldsymbol{u}\|_{L^{4}(\Upsilon^{\pm}_{K})}\|\nabla\boldsymbol{w}\|^{2}_{L^{2}(\Upsilon^{\pm}_{K})}+\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon^{\pm}_{K})}^{2}+\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon^{\pm}_{K})}^{3}\right),\forall K\geq 2, (3.73)

where C>0C>0 is a constant independent of KK.

Proof.   We only show (3.73) on the Υ+\Upsilon^{+} since the rest part is similar. During the proof, we cancel the upper index “++” of the domain for simplicity. Noticing

𝒮{x1=s}w1(s,x2)𝑑x20,s1,\int_{\mathcal{S}\cap\{x_{1}=s\}}w_{1}(s,x_{2})dx_{2}\equiv 0,\quad\forall s\geq 1,

by integrating the above equality for variable ss from K1K-1 to KK, we deduce that

ΥKw1𝑑x=0,K2.\int_{\Upsilon_{K}}w_{1}dx=0,\quad\forall K\geq 2.

Using Lemma 2.14, one derives the existence of a vector field VV satisfying (2.45) with f=w1f=w_{1}. Applying equation (1.1)1, one arrives

ΥK(p~p)w1𝑑x=\displaystyle\int_{\Upsilon_{K}}(\tilde{p}-{p})w_{1}dx= ΥK(p~p)𝑽𝑑x\displaystyle\int_{\Upsilon_{K}}(\tilde{p}-{p})\nabla\cdot\boldsymbol{V}dx
=\displaystyle= ΥK(p~p)𝑽𝑑x=ΥK(𝒘𝒘+𝒖𝒘+𝒘𝒖Δ𝒘)𝑽𝑑x.\displaystyle-\int_{\Upsilon_{K}}\nabla(\tilde{p}-{p})\cdot\boldsymbol{V}dx=\int_{\Upsilon_{K}}\left(\boldsymbol{w}\cdot\nabla\boldsymbol{w}+\boldsymbol{u}\cdot\nabla\boldsymbol{w}+\boldsymbol{w}\cdot\nabla\boldsymbol{u}-\Delta\boldsymbol{w}\right)\cdot\boldsymbol{V}dx.

Using integration by parts, one deduces

ΥK(p~p)w1𝑑x=i,j=12ΥK(iwjwiwjuiwjujwi)iVjdx.\int_{\Upsilon_{K}}(\tilde{p}-p)w_{1}dx=\sum_{i,j=1}^{2}\int_{\Upsilon_{K}}(\partial_{i}w_{j}-w_{i}w_{j}-u_{i}w_{j}-u_{j}w_{i})\partial_{i}V_{j}dx.

By applying Hölder’s inequality and (2.45) in Lemma 2.14, one deduces that

|ΥK(p~p)w1𝑑x|C(𝒘L2(ΥK)+𝒘L4(ΥK)2+𝒖L4(ΥK)𝒘L4(ΥK))w1L2(ΥK).\left|\int_{\Upsilon_{K}}(\tilde{p}-p)w_{1}dx\right|\leq C\left(\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}+\|\boldsymbol{w}\|_{L^{4}(\Upsilon_{K})}^{2}+\|\boldsymbol{u}\|_{L^{4}(\Upsilon_{K})}\|\boldsymbol{w}\|_{L^{4}(\Upsilon_{K})}\right)\|w_{1}\|_{L^{2}(\Upsilon_{K})}. (3.74)

Since w1w_{1} has a zero mean value on each cross section {x1=s}\{x_{1}=s\} for s1s\geq 1 and w2w_{2} has zero boundary on in the x2x_{2} direction, then Poincaré inequality in x2x_{2} direction implies that

𝒘L2(ΥK)Cx2𝒘L2(ΥK).\|\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}\leq C\|\partial_{x_{2}}\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}. (3.75)

Substituting (3.75) in (3.74), also noting the Gagliardo-Nirenberg inequality

𝒘L4(ΥK)2C(𝒘L2(ΥK)𝒘L2(ΥK)+𝒘L2(ΥK)2),\|\boldsymbol{w}\|_{L^{4}(\Upsilon_{K})}^{2}\leq C\left(\|\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}+\|\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}^{2}\right),

one concludes

|ΥK(p~p)w1𝑑x|C(𝒖L4(ΥK)𝒘L2(ΥK)2+𝒘L2(ΥK)2+𝒘L2(ΥK)3).\left|\int_{\Upsilon_{K}}(\tilde{p}-p)w_{1}dx\right|\leq C\left(\|\boldsymbol{u}\|_{L^{4}(\Upsilon_{K})}\|\nabla\boldsymbol{w}\|^{2}_{L^{2}(\Upsilon_{K})}+\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}^{2}+\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K})}^{3}\right).

3.3.2 Main estimates of the uniqueness result

Subtracting the equation of 𝒖\boldsymbol{u} from the equation of 𝒖~\tilde{\boldsymbol{u}}, one finds

𝒘𝒘+𝒖𝒘+𝒘𝒖+(p~p)Δ𝒘=0.\boldsymbol{w}\cdot\nabla\boldsymbol{w}+\boldsymbol{u}\cdot\nabla\boldsymbol{w}+\boldsymbol{w}\cdot\nabla\boldsymbol{u}+\nabla(\tilde{p}-p)-\Delta\boldsymbol{w}=0. (3.76)

Multiplying 𝒘\boldsymbol{w} on both sides of (3.76), and integrating on 𝒮ζ\mathcal{S}_{\zeta}, one derives

𝒮ζ𝒘Δ𝒘𝑑x=𝒮ζ𝒘(𝒘𝒘+𝒖𝒘+𝒘𝒖+(p~p))𝑑x.-\int_{\mathcal{S}_{\zeta}}\boldsymbol{w}\cdot\Delta\boldsymbol{w}dx=-\int_{\mathcal{S}_{\zeta}}\boldsymbol{w}\big{(}\boldsymbol{w}\cdot\nabla\boldsymbol{w}+\boldsymbol{u}\cdot\nabla\boldsymbol{w}+\boldsymbol{w}\cdot\nabla\boldsymbol{u}+\nabla(\tilde{p}-p)\big{)}dx. (3.77)

Using the divergence-free property and the Navier-slip boundary condition of 𝒖\boldsymbol{u} and 𝒖~\tilde{\boldsymbol{u}}, one deduces

𝒮ζ𝒘Δ𝒘𝑑x=𝒮ζwixj(xjwi+xiwj)dx=i,j=12𝒮ζxjwi(xjwi+xiwj)dxi,j=12𝒮ζwinj(xjwi+xiwj)𝑑x=2𝒮ζ|𝕊𝒘|2𝑑x+α𝒮ζ𝒮|wτ|2𝑑Si=12{x1=ζ}wi(x1wi+xiw1)𝑑x2+i=12{y1=ζ}wi(x1wi+xiw1)𝑑y2.\begin{split}-\int_{\mathcal{S}_{\zeta}}\boldsymbol{w}\cdot\Delta\boldsymbol{w}dx&=-\int_{\mathcal{S}_{\zeta}}w_{i}\partial_{x_{j}}(\partial_{x_{j}}w_{i}+\partial_{x_{i}}w_{j})dx\\ &=\sum_{i,j=1}^{2}\int_{\mathcal{S}_{\zeta}}\partial_{x_{j}}w_{i}(\partial_{x_{j}}w_{i}+\partial_{x_{i}}w_{j})dx-\sum_{i,j=1}^{2}\int_{\partial\mathcal{S}_{\zeta}}w_{i}n_{j}(\partial_{x_{j}}w_{i}+\partial_{x_{i}}w_{j})dx\\ &=2\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{w}|^{2}dx+\alpha\int_{\partial\mathcal{S}_{\zeta}\cap\partial\mathcal{S}}|w_{\tau}|^{2}dS\\ &\hskip 28.45274pt-\sum_{i=1}^{2}\int_{\{x_{1}=\zeta\}}w_{i}(\partial_{x_{1}}w_{i}+\partial_{x_{i}}w_{1})dx_{2}+\sum_{i=1}^{2}\int_{\{y_{1}=-\zeta\}}w_{i}(\partial_{x_{1}}w_{i}+\partial_{x_{i}}w_{1})dy_{2}.\end{split}

Here 𝒏=(n1,n2)\boldsymbol{n}=(n_{1},n_{2}) is the unit outer normal vector on 𝒮\partial\mathcal{S}. Then one concludes that

𝒮ζ𝒘Δ𝒘𝑑x+{x1=ζ}|𝒘||𝒘|𝑑x2+{y1=ζ}|𝒘||𝒘|𝑑y22𝒮ζ|𝕊𝒘|2𝑑x+α𝒮ζ𝒮|𝒘tan|2𝑑S.-\int_{\mathcal{S}_{\zeta}}\boldsymbol{w}\cdot\Delta\boldsymbol{w}dx+\int_{\{x_{1}=\zeta\}}|\boldsymbol{w}||\nabla\boldsymbol{w}|dx_{2}+\int_{\{y_{1}=-\zeta\}}|\boldsymbol{w}||\nabla\boldsymbol{w}|dy_{2}\geq 2\int_{\mathcal{S}_{\zeta}}|\mathbb{S}\boldsymbol{w}|^{2}dx+\alpha\int_{\partial\mathcal{S}_{\zeta}\cap\partial\mathcal{S}}|\boldsymbol{w}_{\mathrm{tan}}|^{2}dS.

Then using the Korn inequality (2.36) in Lemma 2.6, we can achieve that

𝒮ζ|𝒘|2𝑑xC(𝒮ζ𝒘Δ𝒘𝑑x+{x1=ζ}|𝒘||𝒘|𝑑x2+{y1=ζ}|𝒘||𝒘|𝑑y2).\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dx\leq C\left(-\int_{\mathcal{S}_{\zeta}}\boldsymbol{w}\cdot\Delta\boldsymbol{w}dx+\int_{\{x_{1}=\zeta\}}|\boldsymbol{w}||\nabla\boldsymbol{w}|dx_{2}+\int_{\{y_{1}=-\zeta\}}|\boldsymbol{w}||\nabla\boldsymbol{w}|dy_{2}\right). (3.78)

Now we focus on the right hand side of (3.77). Applying integration by parts, one derives

𝒮ζ𝒘(𝒘𝒘+(p~p))𝑑x=𝒮{x1=ζ}𝒘𝒆𝟏(12|𝒘|2+(p~p))𝑑x2+𝒮{y1=ζ}𝒘𝒆𝟏(12|𝒘|2+(p~p))𝑑y2.\begin{split}-\int_{\mathcal{S}_{\zeta}}\boldsymbol{w}\big{(}\boldsymbol{w}\cdot\nabla\boldsymbol{w}+\nabla(\tilde{p}-p)\big{)}dx=&-\int_{\mathcal{S}\cap\{x_{1}=\zeta\}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}\left(\frac{1}{2}|\boldsymbol{w}|^{2}+(\tilde{p}-p)\right)dx_{2}\\ &+\int_{\mathcal{S}\cap\{y_{1}=-\zeta\}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}^{\prime}\left(\frac{1}{2}|\boldsymbol{w}|^{2}+(\tilde{p}-p)\right)dy_{2}.\end{split} (3.79)

Applying Hölder’s inequality, noting that 𝒖=𝒗+𝒂\boldsymbol{u}=\boldsymbol{v}+\boldsymbol{a}, where 𝒂\boldsymbol{a} is the flux carrier constructed in Proposition 3.1, while 𝒗\boldsymbol{v} is the H1H^{1}-weak solution given in Section 3.2, one has

|𝒮ζ(𝒘𝒖𝒘+𝒖𝒘𝒘)𝑑x|𝒗L2(𝒮ζ)𝒘L4(𝒮ζ)2+𝒗L4(𝒮ζ)𝒘L2(𝒮ζ)𝒘L4(𝒮ζ)+𝒂L(𝒮ζ)𝒘L2(𝒮ζ)2+𝒂L(𝒮ζ)𝒘L2(𝒮ζ)𝒘L2(𝒮ζ)C(𝒗H1(𝒮ζ)+𝒂W1,(𝒮ζ))𝒮ζ|𝒘|2𝑑xΦeCΦ𝒮ζ|𝒘|2𝑑x.\begin{split}\left|-\int_{\mathcal{S}_{\zeta}}\big{(}\boldsymbol{w}\cdot\nabla\boldsymbol{u}\cdot\boldsymbol{w}+\boldsymbol{u}\cdot\nabla\boldsymbol{w}\cdot\boldsymbol{w}\big{)}dx\right|\leq&\,\|\nabla\boldsymbol{v}\|_{L^{2}(\mathcal{S}_{\zeta})}\|\boldsymbol{w}\|_{L^{4}(\mathcal{S}_{\zeta})}^{2}+\|\boldsymbol{v}\|_{L^{4}(\mathcal{S}_{\zeta})}\|\nabla\boldsymbol{w}\|_{L^{2}(\mathcal{S}_{\zeta})}\|\boldsymbol{w}\|_{L^{4}(\mathcal{S}_{\zeta})}\\ &+\|\nabla\boldsymbol{a}\|_{L^{\infty}(\mathcal{S}_{\zeta})}\|\boldsymbol{w}\|_{L^{2}(\mathcal{S}_{\zeta})}^{2}+\|\boldsymbol{a}\|_{L^{\infty}(\mathcal{S}_{\zeta})}\|\nabla\boldsymbol{w}\|_{L^{2}(\mathcal{S}_{\zeta})}\|\boldsymbol{w}\|_{L^{2}(\mathcal{S}_{\zeta})}\\ \leq&\,C\left(\|\boldsymbol{v}\|_{H^{1}(\mathcal{S}_{\zeta})}+\|\boldsymbol{a}\|_{W^{1,\infty}(\mathcal{S}_{\zeta})}\right)\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dx\\ \leq&\,\Phi e^{C\Phi}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dx.\end{split} (3.80)

Here in the second inequality, we have applied the Gagliardo-Nirenberg inequality and the Poincaré inequality (2.35) in Lemma 2.4, which indicate

𝒘L4(𝒮ζ)C(𝒘L2(𝒮ζ)1/2𝒘L2(𝒮ζ)1/2+𝒘L2(𝒮ζ))C(𝒮ζ|𝒘|2𝑑x)1/2.\|\boldsymbol{w}\|_{L^{4}(\mathcal{S}_{\zeta})}\leq C\left(\|\boldsymbol{w}\|^{1/2}_{L^{2}(\mathcal{S}_{\zeta})}\|\nabla\boldsymbol{w}\|^{1/2}_{L^{2}(\mathcal{S}_{\zeta})}+\|\boldsymbol{w}\|_{L^{2}(\mathcal{S}_{\zeta})}\right)\leq C\left(\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dx\right)^{1/2}.

Meanwhile, the third inequality in (3.80) is guaranteed by (3.63) and Estimates for 𝒂\boldsymbol{a}. Substituting (3.78), (3.79) and (3.80) in (3.77), one arrives

𝒮ζ|𝒘|2𝑑xC({x1=ζ}|𝒘|(|𝒘|+|𝒘|2)dx2+{y1=ζ}|𝒘|(|𝒘|+|𝒘|2)dy2+ΦeCΦ𝒮ζ|𝒘|2dx{x1=ζ}𝒘𝒆𝟏(p~p)dx2+{y1=ζ}𝒘𝒆𝟏(p~p)dy2).\begin{split}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dx\leq&C\left(\int_{\{x_{1}=\zeta\}}|\boldsymbol{w}|(|\nabla\boldsymbol{w}|+|\boldsymbol{w}|^{2})dx_{2}+\int_{\{y_{1}=-\zeta\}}|\boldsymbol{w}|(|\nabla\boldsymbol{w}|+|\boldsymbol{w}|^{2})dy_{2}+\Phi e^{C\Phi}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dx\right.\\ &\left.-\int_{\{x_{1}=\zeta\}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}\left(\tilde{p}-p\right)dx_{2}+\int_{\{y_{1}=-\zeta\}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}^{\prime}\left(\tilde{p}-p\right)dy_{2}\right).\end{split}

Now one concludes that if Φ<<1\Phi<<1 being small enough such that

ΦeCΦ<12,\Phi e^{C\Phi}<\frac{1}{2},

then we achieve

𝒮ζ|𝒘|2𝑑xC({x1=ζ}|𝒘|(|𝒘|+|𝒘|2)dx2+{y1=ζ}|𝒘|(|𝒘|+|𝒘|2)dy2{x1=ζ}𝒘𝒆𝟏(p~p)dx2+{y1=ζ}𝒘𝒆𝟏(p~p)dy2).\begin{split}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dx\leq C&\left(\int_{\{x_{1}=\zeta\}}|\boldsymbol{w}|(|\nabla\boldsymbol{w}|+|\boldsymbol{w}|^{2})dx_{2}+\int_{\{y_{1}=-\zeta\}}|\boldsymbol{w}|(|\nabla\boldsymbol{w}|+|\boldsymbol{w}|^{2})dy_{2}\right.\\ &\left.-\int_{\{x_{1}=\zeta\}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}\left(\tilde{p}-p\right)dx_{2}+\int_{\{y_{1}=-\zeta\}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}^{\prime}\left(\tilde{p}-p\right)dy_{2}\right).\end{split}

Therefore, one derives the following estimate by integrating with ζ\zeta on [K1,K][K-1,K], where K2K\geq 2:

K1K𝒮ζ|𝒘|2𝑑x𝑑ζC(ΥK+|𝒘|(|𝒘|+|𝒘|2)dx+ΥK|𝒘|(|𝒘|+|𝒘|2)dy+|ΥK+𝒘𝒆𝟏(p~p)dx|+|ΥK𝒘𝒆𝟏(p~p)dy|).\begin{split}\int_{K-1}^{K}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dxd\zeta\leq C&\left(\int_{\Upsilon_{K}^{+}}|\boldsymbol{w}|(|\nabla\boldsymbol{w}|+|\boldsymbol{w}|^{2})dx+\int_{\Upsilon_{K}^{-}}|\boldsymbol{w}|(|\nabla\boldsymbol{w}|+|\boldsymbol{w}|^{2})dy\right.\\ &\left.+\Big{|}\int_{\Upsilon_{K}^{+}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}\left(\tilde{p}-p\right)dx\Big{|}+\Big{|}\int_{\Upsilon_{K}^{-}}\boldsymbol{w}\cdot\boldsymbol{e_{1}}^{\prime}\left(\tilde{p}-p\right)dy\Big{|}\right).\end{split} (3.81)

Now we only handle integrations on ΥK+\Upsilon_{K}^{+} since the cases of ΥK\Upsilon_{K}^{-} are similar. Using the Cauchy-Schwarz inequality and the Poincaré inequality Lemma 2.3, one has

ΥK+|𝒘||𝒘|𝑑x𝒘L2(ΥK+)𝒘L2(ΥK+)C𝒘L2(ΥK+)2.\int_{\Upsilon_{K}^{+}}|\boldsymbol{w}||\nabla\boldsymbol{w}|dx\leq\|\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}\leq C\|\nabla\boldsymbol{w}\|^{2}_{L^{2}(\Upsilon_{K}^{+})}. (3.82)

Moreover, by Hölder’s inequality and the Gagliardo-Nirenberg inequality, one writes

ΥK+|𝒘|3𝑑xC(𝒘L2(ΥK+)2𝒘L2(ΥK+)+𝒘L2(ΥK+)3),\int_{\Upsilon_{K}^{+}}|\boldsymbol{w}|^{3}dx\leq C\left(\|\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}^{2}\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}+\|\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}^{3}\right),

which follows by the Poincaré inequality that

ΥK+|𝒘|3𝑑xC𝒘L2(ΥK+)3.\int_{\Upsilon_{K}^{+}}|\boldsymbol{w}|^{3}dx\leq C\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}^{3}.

Recalling Proposition 3.6, one arrives at

|ΥK+w3(p~p)𝑑x|C(𝒖L4(ΥK+)𝒘L2(ΥK+)2+𝒘L2(ΥK+)2+𝒘L2(ΥK+)3).\left|\int_{\Upsilon_{K}^{+}}w_{3}\left(\tilde{p}-p\right)dx\right|\leq C\left(\|\boldsymbol{u}\|_{L^{4}(\Upsilon_{K}^{+})}\|\nabla\boldsymbol{w}\|^{2}_{L^{2}(\Upsilon_{K}^{+})}+\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}^{2}+\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+})}^{3}\right). (3.83)

Substituting (3.82)–(3.83), together with their related inequality on domain ΥK\Upsilon_{K}^{-}, in (3.81), one concludes

K1K𝒮ζ|𝒘|2𝑑x𝑑ζC(𝒘L2(ΥK+ΥK)2+𝒘L2(ΥK+ΥK)3).\int_{K-1}^{K}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dxd\zeta\leq C\left(\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+}\cup\Upsilon_{K}^{-})}^{2}+\|\nabla\boldsymbol{w}\|_{L^{2}(\Upsilon_{K}^{+}\cup\Upsilon_{K}^{-})}^{3}\right). (3.84)

3.3.3 End of proof

Finally, by defining

Y(K):=K1K𝒮ζ|𝒘|2𝑑x𝑑ζ,Y(K):=\int_{K-1}^{K}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dxd\zeta,

(3.84) indicates

Y(K)C(Y(K)+(Y(K))3/2),K1.Y(K)\leq C\left(Y^{\prime}(K)+\left(Y^{\prime}(K)\right)^{3/2}\right),\quad\forall K\geq 1.

By Lemma 2.11, we derive

lim infζK3Y(K)>0,\liminf_{\zeta\to\infty}K^{-3}Y(K)>0,

that is, there exists C0>0C_{0}>0 such that

K1K𝒮ζ|𝒘|2𝑑x𝑑ζC0K3.\int_{K-1}^{K}\int_{\mathcal{S}_{\zeta}}|\nabla\boldsymbol{w}|^{2}dxd\zeta\geq C_{0}K^{3}.

However, this leads to a paradox with the condition (1.14). Thus Y(K)0Y(K)\equiv 0 for all K1K\geq 1, which proves 𝒖𝒖~\boldsymbol{u}\equiv\tilde{\boldsymbol{u}}. This concludes the uniqueness.

4  Asymptotic and regularity of the weak solution

4.1 Decay estimate of the weak solution

In this subsection we will show the weak solution constructed in the previous section decays exponentially to Poiseuille flows (1.6) as |x||x|\to\infty. Our proof is also valid for stationary Navier-Stokes problem on domains which is less regular, say an infinite pipe only with a C1,1C^{1,1} boundary.

For the convenience of our further statement, we localize the problem in the following way: Denoting

𝒮=k𝔖k,\mathcal{S}=\bigcup_{k\in\mathbb{Z}}\mathfrak{S}_{k}, (4.85)

where

𝔖k:={𝒮{x2:(3k21)ZΦx1(3k2+1)ZΦ},k>0;𝒮ZΦ,k=0;𝒮{x2:(3k21)ZΦy1(3k2+1)ZΦ},k<0,\mathfrak{S}_{k}:=\left\{\begin{array}[]{ll}\mathcal{S}\cap\left\{x\in\mathbb{R}^{2}:\,\left(\frac{3k}{2}-1\right)Z_{\Phi}\leq x_{1}\leq\left(\frac{3k}{2}+1\right)Z_{\Phi}\right\},&\quad k>0;\\[4.2679pt] \mathcal{S}_{Z_{\Phi}},&\quad k=0;\\[4.2679pt] \mathcal{S}\cap\left\{x\in\mathbb{R}^{2}:\,\left(\frac{3k}{2}-1\right)Z_{\Phi}\leq y_{1}\leq\left(\frac{3k}{2}+1\right)Z_{\Phi}\right\},&\quad k<0,\\ \end{array}\right.

where ZΦ=e2/εeCΦZ_{\Phi}=e^{2/\varepsilon}\leq e^{C\Phi}, while ε>0\varepsilon>0 is a fixed small constant given in the construction of 𝒂\boldsymbol{a}. Here is the main result of this subsection:

Proposition 4.1.

Let the conditions of the item (ii) in Theorem 1.4 be satisfied and (𝐯,Π)(\boldsymbol{v},\,\Pi) is given in (3.60) and (3.72). Then there exist positive constants CC, σ\sigma, depending only on Φ\Phi, such that

𝒖𝑷ΦLH1(𝒮L\𝒮ζ)+𝒖𝑷ΦRH1(𝒮R\𝒮ζ)C𝒗H1(𝒮)exp(σζ),\begin{split}\left\|\boldsymbol{u}-\boldsymbol{P}^{L}_{\Phi}\right\|_{H^{1}(\mathcal{S}_{L}\backslash\mathcal{S}_{\zeta})}+\left\|\boldsymbol{u}-\boldsymbol{P}^{R}_{\Phi}\right\|_{H^{1}(\mathcal{S}_{R}\backslash\mathcal{S}_{\zeta})}&\leq C\|\boldsymbol{v}\|_{H^{1}(\mathcal{S})}\exp(-\sigma\zeta),\\ \end{split} (4.86)

for any ζ\zeta being large enough.

During the proof of Proposition 4.1, we need the following refined estimate of the pressure field:

Lemma 4.2.

The reformulated pressure field Π\Pi given in (3.72) enjoys the following uniform estimate:

kΠΠ¯𝔖kL2(𝔖k)2Φ2eCΦ<.\sum_{k\in\mathbb{Z}}\|\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}\|^{2}_{L^{2}(\mathfrak{S}_{k})}\leq\Phi^{2}e^{C\Phi}<\infty.

Proof.   Applying (2.44) in Lemma 2.13, one deduces

ΠΠ¯𝔖kL2(𝔖k)CkΠH1(𝔖k).\|\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}\|_{L^{2}(\mathfrak{S}_{k})}\leq C_{k}\|\nabla\Pi\|_{H^{-1}(\mathfrak{S}_{k})}. (4.87)

Notice that, each 𝔖k\mathfrak{S}_{k} (kk\in\mathbb{Z}) is congruent to an element in {𝔖1,𝔖0,𝔖1}\{\mathfrak{S}_{-1},\,\mathfrak{S}_{0},\,\mathfrak{S}_{1}\}. This indicates constants CkC_{k} in estimates (4.87) above could be chosen uniformly with respect to kk\in\mathbb{Z}. By equation

Π=div(𝒗𝒗𝒗𝒂𝒗𝒗𝒂)+Δ𝒂+CRη(x1)𝒆𝟏+CLη(y1)𝒆𝟏𝒂𝒂,\nabla\Pi=\text{div}\big{(}\nabla\boldsymbol{v}-\boldsymbol{v}\otimes\boldsymbol{v}-\boldsymbol{a}\otimes\boldsymbol{v}-\boldsymbol{v}\otimes\boldsymbol{a}\big{)}+\Delta\boldsymbol{a}+{C_{R}}{\eta(x_{1})}\boldsymbol{e_{1}}+{C_{L}}{\eta(-y_{1})}\boldsymbol{e_{1}}^{\prime}-\boldsymbol{a}\cdot\nabla\boldsymbol{a},

with both Δ𝒂+CRη(x1)𝒆𝟏+CLη(y1)𝒆𝟏\Delta\boldsymbol{a}+{C_{R}}{\eta(x_{1})}\boldsymbol{e_{1}}+{C_{L}}{\eta(-y_{1})}\boldsymbol{e_{1}}^{\prime} and 𝒂𝒂\boldsymbol{a}\cdot\nabla\boldsymbol{a} vanish in 𝔖k\mathfrak{S}_{k} with |k|2|k|\geq 2, one concludes from (4.87) that

ΠΠ¯𝔖kL2(𝔖k)C(𝒗L2(𝔖k)+𝒗L4(𝔖k)2+ΦeCΦ𝒗L2(𝔖k))+ΦeCΦχ|k|1C𝒗H1(𝔖k)(1+ΦeCΦ+𝒗H1(𝔖k))+ΦeCΦχ|k|1.\begin{split}\|\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}\|_{L^{2}(\mathfrak{S}_{k})}&\leq C\left(\|\nabla\boldsymbol{v}\|_{L^{2}(\mathfrak{S}_{k})}+\|\boldsymbol{v}\|_{L^{4}(\mathfrak{S}_{k})}^{2}+\Phi e^{C\Phi}\|\boldsymbol{v}\|_{L^{2}(\mathfrak{S}_{k})}\right)+\Phi e^{C\Phi}\chi_{|k|\leq 1}\\ &\leq C\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\left(1+\Phi e^{C\Phi}+\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\right)+\Phi e^{C\Phi}\chi_{|k|\leq 1}.\end{split}

Here we have applied the Sobolev imbedding theorem and interpolations of LpL^{p} spaces. This completes the proof of Lemma 4.2.

Proof of Proposition 4.1: We only prove the estimate of term 𝒖𝑷ΦRH1(𝒮R\𝒮ζ)\|\boldsymbol{u}-\boldsymbol{P}^{R}_{\Phi}\|_{H^{1}(\mathcal{S}_{R}\backslash\mathcal{S}_{\zeta})} since the rest term is essentially identical. For ζ>ZΦ\zeta>Z_{\Phi}, in 𝒮R\𝒮ζ\mathcal{S}_{R}\backslash\mathcal{S}_{\zeta}, the equation of 𝒗=𝒖𝒂\boldsymbol{v}=\boldsymbol{u}-\boldsymbol{a} reads

𝒗𝒗+𝒂𝒗+𝒗𝒂+ΠΔ𝒗=0.\boldsymbol{v}\cdot\nabla\boldsymbol{v}+\boldsymbol{a}\cdot\nabla\boldsymbol{v}+\boldsymbol{v}\cdot\nabla\boldsymbol{a}+\nabla\Pi-\Delta\boldsymbol{v}=0. (4.88)

This is because

Δ𝒂+CRη(x3)𝒆𝟏+CLη(y1)𝒆𝟏𝒂𝒂=(ΔPΦR+CL)𝒆𝟏=0,in𝒮R\𝒮ζ.\Delta\boldsymbol{a}+{C_{R}}{\eta(x_{3})}\boldsymbol{e_{1}}+{C_{L}}{\eta(-y_{1})}\boldsymbol{e_{1}}^{\prime}-\boldsymbol{a}\cdot\nabla\boldsymbol{a}=\left(\Delta{P}^{R}_{\Phi}+{C_{L}}\right)\boldsymbol{e_{1}}=0,\quad\text{in}\quad\mathcal{S}_{R}\backslash\mathcal{S}_{\zeta}.

In the following proof, we will drop (upper or lower) indexes “RR” for convenience. For any ZΦ<ζζ<ζ1Z_{\Phi}<\zeta\leq\zeta^{\prime}<\zeta_{1}, taking inner product with 𝒗\boldsymbol{v} on both sides of (4.88) and integrating on 𝒮R(𝒮ζ1\𝒮ζ)\mathcal{S}_{R}\cap(\mathcal{S}_{\zeta_{1}}\backslash\mathcal{S}_{\zeta^{\prime}}), one has

01ζζ1𝒗Δ𝒗𝑑x1𝑑x2LHS=01ζζ1(𝒗𝒗+𝒂𝒗+𝒗𝒂+Π)𝒗𝑑x1𝑑x2RHS.\underbrace{\int_{0}^{1}\int_{\zeta^{\prime}}^{\zeta_{1}}\boldsymbol{v}\cdot\Delta\boldsymbol{v}dx_{1}dx_{2}}_{LHS}=\underbrace{\int_{0}^{1}\int_{\zeta^{\prime}}^{\zeta_{1}}\big{(}\boldsymbol{v}\cdot\nabla\boldsymbol{v}+\boldsymbol{a}\cdot\nabla\boldsymbol{v}+\boldsymbol{v}\cdot\nabla\boldsymbol{a}+\nabla\Pi\big{)}\cdot\boldsymbol{v}dx_{1}dx_{2}}_{RHS}. (4.89)

To handle the left hand side of (4.89), one first recalls the derivation of (3.78) that

01ζζ1𝒗Δ𝒗𝑑x1𝑑x2=201ζζ1|𝕊𝒗|2𝑑x1𝑑x2αζζ1|𝒗tan|2|x2=1dx1αζζ1|𝒗tan|2|x2=0dx1i=1201vi(x1vi+xiv1)|x1=ζdx2+i=1201vi(x1vi+xiv1)|x1=ζ1dx2.\begin{split}\int_{0}^{1}\int_{\zeta^{\prime}}^{\zeta_{1}}\boldsymbol{v}\cdot\Delta\boldsymbol{v}dx_{1}dx_{2}=&-2\int_{0}^{1}\int_{\zeta^{\prime}}^{\zeta_{1}}|\mathbb{S}\boldsymbol{v}|^{2}dx_{1}dx_{2}-\alpha\int_{\zeta^{\prime}}^{\zeta_{1}}|\boldsymbol{v}_{tan}|^{2}\Big{|}_{x_{2}=1}dx_{1}-\alpha\int_{\zeta^{\prime}}^{\zeta_{1}}|\boldsymbol{v}_{tan}|^{2}\Big{|}_{x_{2}=0}dx_{1}\\ &-\sum_{i=1}^{2}\int_{0}^{1}v_{i}(\partial_{x_{1}}v_{i}+\partial_{x_{i}}v_{1})\Big{|}_{x_{1}=\zeta^{\prime}}dx_{2}+\sum_{i=1}^{2}\int_{0}^{1}v_{i}(\partial_{x_{1}}v_{i}+\partial_{x_{i}}v_{1})\Big{|}_{x_{1}=\zeta_{1}}dx_{2}.\end{split}

Applying Lemma 2.6, the Korn’s inequality in a truncated stripe, one deduces the left hand side of (4.89) satisfies

LHSC(01ζζ1|𝒗|2𝑑x1𝑑x2+01|𝒗||𝒗||x1=ζdx2+01|𝒗||𝒗||x1=ζ1dx2)\begin{split}LHS\leq&C\left(-\int_{0}^{1}\int_{\zeta^{\prime}}^{\zeta_{1}}|\nabla\boldsymbol{v}|^{2}dx_{1}dx_{2}+\int_{0}^{1}|\boldsymbol{v}||\nabla\boldsymbol{v}|\Big{|}_{x_{1}=\zeta^{\prime}}dx_{2}+\int_{0}^{1}|\boldsymbol{v}||\nabla\boldsymbol{v}|\Big{|}_{x_{1}=\zeta_{1}}dx_{2}\right)\end{split} (4.90)

Using integration by parts for the right hand side of (4.89), one arrives

RHS=01(12(v1+PΦ)|𝒗|2+v1Π+PΦ(v1)2)|x1=ζ1dx201(12(v1+PΦ)|𝒗|2+v1Π+PΦ(v1)2)|x1=ζdx201ζζ1𝒗𝒗𝒂dx1dx2.\begin{split}RHS=&\int_{0}^{1}\left(\frac{1}{2}\left(v_{1}+P_{\Phi}\right)|\boldsymbol{v}|^{2}+v_{1}\Pi+P_{\Phi}(v_{1})^{2}\right)\Big{|}_{x_{1}=\zeta_{1}}dx_{2}\\ &-\int_{0}^{1}\left(\frac{1}{2}\left(v_{1}+P_{\Phi}\right)|\boldsymbol{v}|^{2}+v_{1}\Pi+P_{\Phi}(v_{1})^{2}\right)\Big{|}_{x_{1}=\zeta^{\prime}}dx_{2}\\ &-\int_{0}^{1}\int_{\zeta^{\prime}}^{\zeta_{1}}\boldsymbol{v}\cdot\nabla\boldsymbol{v}\cdot\boldsymbol{a}dx_{1}dx_{2}.\end{split} (4.91)

Now we are ready to perform ζ1\zeta_{1}\to\infty. To do this, one must be careful with the integrations on {x1=ζ1}×(0,1)\{x_{1}=\zeta_{1}\}\times(0,1) in both (4.90) and (4.91). Recalling estimates of (𝒗,Π)(\boldsymbol{v},\Pi) in Theorem 3.5 and Lemma 4.2, one derives

𝒗H1(𝒮)2+𝒗L4(𝒮)4+kΠΠ¯𝔖kL2(𝔖k)2Φ2eCΦ<.\|\boldsymbol{v}\|^{2}_{H^{1}(\mathcal{S})}+\|\boldsymbol{v}\|^{4}_{L^{4}(\mathcal{S})}+\sum_{k\in\mathbb{Z}}\|\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}\|^{2}_{L^{2}(\mathfrak{S}_{k})}\leq\Phi^{2}e^{C\Phi}<\infty. (4.92)

Choosing M:=Φ2eCΦZΦM:=\frac{\Phi^{2}e^{C\Phi}}{Z_{\Phi}}, one concludes that for any k>1k>1, there exists a slice {x1=ζ1,k}×(0,1)\{x_{1}=\zeta_{1,k}\}\times(0,1) which satisfies

{x1=ζ1,k}×(0,1)𝒮{x2:(3k212)ZΦx1(3k2+12)ZΦ}𝔖k,\{x_{1}=\zeta_{1,k}\}\times(0,1)\subset\mathcal{S}\cap\left\{x\in\mathbb{R}^{2}:\,\left(\frac{3k}{2}-\frac{1}{2}\right)Z_{\Phi}\leq x_{1}\leq\left(\frac{3k}{2}+\frac{1}{2}\right)Z_{\Phi}\right\}\subset\mathfrak{S}_{k},

and it holds that

01(|𝒗|2+|𝒗|4+|ΠΠ¯𝔖k|2)|x1=ζ1,kdx2M.\int_{0}^{1}\left(|\nabla\boldsymbol{v}|^{2}+|\boldsymbol{v}|^{4}+|\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}|^{2}\right)\Big{|}_{{x_{1}}=\zeta_{1,k}}dx_{2}\leq M.

Otherwise, one has

𝒗H1(𝔖k)2+𝒗L4(𝔖k)4+ΠΠ¯𝔖kL2(𝔖k)2>ZΦM=Φ2eCΦ,\|\boldsymbol{v}\|^{2}_{H^{1}(\mathfrak{S}_{k})}+\|\boldsymbol{v}\|^{4}_{L^{4}(\mathfrak{S}_{k})}+\|\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}\|^{2}_{L^{2}(\mathfrak{S}_{k})}{>Z_{\Phi}M=\Phi^{2}e^{C\Phi}},

which creates a paradox to (4.92). Choosing k0>0k_{0}>0 being sufficiently large such that the sequence {ζ1,k}k=k0[ζ,)\{\zeta_{1,k}\}_{k=k_{0}}^{\infty}\subset[\zeta^{\prime},\infty), clearly one has ζ1,k\zeta_{1,k}\nearrow\infty as kk\to\infty. Moreover, using the trace theorem of functions in the Sobolev space H1H^{1}, one has

01|𝒗(x1,x2)|2𝑑x2Cz>x101(|𝒗|2+|𝒗|2)(z,x2)𝑑x2𝑑z0,asx1.\int_{0}^{1}|\boldsymbol{v}(x_{1},x_{2})|^{2}dx_{2}\leq C\int_{z>x_{1}}\int_{0}^{1}(|\boldsymbol{v}|^{2}+|\nabla\boldsymbol{v}|^{2})(z,x_{2})dx_{2}dz\to 0,\quad\text{as}\quad x_{1}\to\infty.

Noting that 01v1(ζ1,k,x2)𝑑x2=0\int_{0}^{1}v_{1}(\zeta_{1,k},x_{2})dx_{2}=0 for kk0k\geq k_{0}, we deduce the following by the Poincaré inequality:

|01v3Π|x1=ζ1,kdx2|=|01v3(ΠΠ¯𝔖k)|x1=ζ1,kdx2|(01|𝒗|2|x1=ζ1,kdx2)1/2(01|ΠΠ¯𝔖k|2|x1=ζ1,kdx2)1/20,ask.\begin{split}\left|\int_{0}^{1}v_{3}\Pi\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\right|&=\left|\int_{0}^{1}v_{3}\left(\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}\right)\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\right|\\ &\leq\left(\int_{0}^{1}|\boldsymbol{v}|^{2}\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\right)^{1/2}\left(\int_{0}^{1}|\Pi-\overline{\Pi}_{\mathfrak{S}_{k}}|^{2}\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\right)^{1/2}\to 0,\quad\text{as}\quad k\to\infty.\end{split}

Meanwhile, one finds

Σ×{x3=ζ1,k}|𝒗|(|𝒗|+|𝒗|2)|x1=ζ1,kdx2(01(|𝒗|2+|𝒗|4)|x1=ζ1,kdx2)1/2(01|𝒗|2|x1=ζ1,kdx2)1/20,ask;\begin{split}&\int_{\Sigma\times\{x_{3}=\zeta_{1,k}\}}|\boldsymbol{v}|\left(|\nabla\boldsymbol{v}|+|\boldsymbol{v}|^{2}\right)\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\\ &\leq\left(\int_{0}^{1}\left(|\nabla\boldsymbol{v}|^{2}+|\boldsymbol{v}|^{4}\right)\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\right)^{1/2}\left(\int_{0}^{1}|\boldsymbol{v}|^{2}\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\right)^{1/2}\to 0,\quad\text{as}\quad k\to\infty;\end{split}

and

01|PΦ||𝒗|2|x1=ζ1,kdx2PΦL(𝒮R)01|𝒗|2|x1=ζ1,kdx20,ask.\int_{0}^{1}|P_{\Phi}||\boldsymbol{v}|^{2}\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\leq\|P_{\Phi}\|_{L^{\infty}(\mathcal{S}_{R})}\int_{0}^{1}|\boldsymbol{v}|^{2}\Big{|}_{x_{1}=\zeta_{1,k}}dx_{2}\to 0,\quad\text{as}\quad k\to\infty.

Choosing ζ1=ζ1,k\zeta_{1}=\zeta_{1,k} (kk0k\geq k_{0}) in (4.90) and (4.91), respectively, and performing kk\to\infty, one can deduce that

01ζ|𝒗|2𝑑xC01ζ𝒗𝒗𝒂dxR1+C01(|𝒗|(|𝒗|2+|PΦ||𝒗|+|𝒗|)+v1Π)|x1=ζdx2.\begin{split}\int_{0}^{1}\int_{\zeta^{\prime}}^{\infty}|\nabla\boldsymbol{v}|^{2}dx\leq&\,\underbrace{C\int_{0}^{1}\int_{\zeta^{\prime}}^{\infty}\boldsymbol{v}\cdot\nabla\boldsymbol{v}\cdot\boldsymbol{a}dx}_{R_{1}}\\ &+C\int_{0}^{1}\Big{(}|\boldsymbol{v}|\left(|\boldsymbol{v}|^{2}+|P_{\Phi}||\boldsymbol{v}|+|\nabla\boldsymbol{v}|\right)+v_{1}\Pi\Big{)}\Big{|}_{x_{1}=\zeta^{\prime}}dx_{2}.\end{split}

Using the Cauchy-Schwarz inequality, the Poincaré inequality in Lemma 2.3, and the construction of profile vector 𝒂\boldsymbol{a}, one derives

R1CPΦL(𝒮R)(01ζ|𝒗|2𝑑x)1/2(01ζ|𝒗|2𝑑x)1/2CαΦ1+α01ζ|𝒗|2𝑑x,R_{1}\leq C\|P_{\Phi}\|_{L^{\infty}(\mathcal{S}_{R})}\left(\int_{0}^{1}\int_{\zeta^{\prime}}^{\infty}|\nabla\boldsymbol{v}|^{2}dx\right)^{1/2}\left(\int_{0}^{1}\int_{\zeta^{\prime}}^{\infty}|\boldsymbol{v}|^{2}dx\right)^{1/2}\leq\frac{C\alpha\Phi}{1+\alpha}\int_{0}^{1}\int_{\zeta^{\prime}}^{\infty}|\nabla\boldsymbol{v}|^{2}dx,

which indicates the following estimate provided αΦ\alpha\Phi is small enough such that CαΦ1+α<1\frac{C\alpha\Phi}{1+\alpha}<1:

01ζ|𝒗|2𝑑xC01(|𝒗|(|𝒗|2+|PΦ||𝒗|+|𝒗|)+v3Π)|x1=ζdx2.\int_{0}^{1}\int_{\zeta^{\prime}}^{\infty}|\nabla\boldsymbol{v}|^{2}dx\leq C\int_{0}^{1}\Big{(}|\boldsymbol{v}|\left(|\boldsymbol{v}|^{2}+|P_{\Phi}||\boldsymbol{v}|+|\nabla\boldsymbol{v}|\right)+v_{3}\Pi\Big{)}\Big{|}_{x_{1}=\zeta^{\prime}}dx_{2}. (4.93)

Denoting

𝒢(ζ):=01ζ|𝒗|2𝑑x,\mathcal{G}(\zeta^{\prime}):=\int_{0}^{1}\int_{\zeta^{\prime}}^{\infty}|\nabla\boldsymbol{v}|^{2}dx, (4.94)

and integrating (4.93) with ζ\zeta^{\prime} on (ζ,)(\zeta,\infty), one arrives

ζ𝒢(ζ)𝑑ζC(01ζ(|𝒗|(|𝒗|2+|PΦ||𝒗|+|𝒗|))𝑑x+|01ζv1Π𝑑x|).\int_{\zeta}^{\infty}\mathcal{G}(\zeta^{\prime})d\zeta^{\prime}\leq C\left(\int_{0}^{1}\int_{\zeta}^{\infty}\Big{(}|\boldsymbol{v}|\left(|\boldsymbol{v}|^{2}+|P_{\Phi}||\boldsymbol{v}|+|\nabla\boldsymbol{v}|\right)\Big{)}dx+\left|\int_{0}^{1}\int_{\zeta}^{\infty}v_{1}\Pi dx\right|\right). (4.95)

Applying the Poincaré inequality in Lemma 2.3, one deduces

01ζ|𝒗|(|𝒗|2+|PΦ||𝒗|+|𝒗|)𝑑xC01ζ|𝒗|2𝑑x.\int_{0}^{1}\int_{\zeta}^{\infty}|\boldsymbol{v}|\left(|\boldsymbol{v}|^{2}+|P_{\Phi}||\boldsymbol{v}|+|\nabla\boldsymbol{v}|\right)dx\leq C\int_{0}^{1}\int_{\zeta}^{\infty}|\nabla\boldsymbol{v}|^{2}dx. (4.96)

Moreover, using a similar approach as in the proof of Proposition 3.6, one notices that

|01ζv1Π𝑑x|m=1|Υζ+m+v1Π𝑑x|Cm=1(PΦL(Υζ+m+)𝒗L2(Υζ+m+)2+𝒗L2(Υζ+m+)2+𝒗L2(Υζ+m+)3)C01ζ|𝒗|2𝑑x.\begin{split}\left|\int_{0}^{1}\int_{\zeta}^{\infty}v_{1}\Pi dx\right|&\leq\sum_{m=1}^{\infty}\left|\int_{\Upsilon_{\zeta+m}^{+}}v_{1}\Pi dx\right|\\ &\leq C\sum_{m=1}^{\infty}\left(\|P_{\Phi}\|_{L^{\infty}(\Upsilon^{+}_{\zeta+m})}\|\nabla\boldsymbol{v}\|^{2}_{L^{2}(\Upsilon^{+}_{\zeta+m})}+\|\nabla\boldsymbol{v}\|_{L^{2}(\Upsilon^{+}_{\zeta+m})}^{2}+\|\nabla\boldsymbol{v}\|_{L^{2}(\Upsilon^{+}_{\zeta+m})}^{3}\right)\\ &\leq C\int_{0}^{1}\int_{\zeta}^{\infty}|\nabla\boldsymbol{v}|^{2}dx.\end{split} (4.97)

Substituting (4.96) and (4.97) in (4.95), one arrives at

ζ𝒢(ζ)𝑑ζC𝒢(ζ),for anyζ>ZΦ.\int_{\zeta}^{\infty}\mathcal{G}(\zeta^{\prime})d\zeta^{\prime}\leq C\mathcal{G}(\zeta),\quad\text{for any}\quad\zeta>Z_{\Phi}.

This implies

𝒩(ζ):=ζ𝒢(ζ)𝑑ζ\mathcal{N}(\zeta):=\int_{\zeta}^{\infty}\mathcal{G}(\zeta^{\prime})d\zeta^{\prime}

is well-defined for all ζ>ZΦ\zeta>Z_{\Phi}, and

𝒩(ζ)C𝒩(ζ),for anyζ>ZΦ.\mathcal{N}(\zeta)\leq-C\mathcal{N}^{\prime}(\zeta),\quad\text{for any}\quad\zeta>Z_{\Phi}. (4.98)

Multiplying the factor eC1ζe^{C^{-1}\zeta} on both sides of (4.98) and integrating on [ZΦ,ζ][Z_{\Phi},\zeta], one deduces

𝒩(ζ)Cexp(C1ζ),for anyζ>ZΦ.\mathcal{N}(\zeta)\leq C\exp\left(-C^{-1}\zeta\right),\quad\text{for any}\quad\zeta>Z_{\Phi}.

According to the definition (4.94), one has 𝒢\mathcal{G} is both non-negative and non-increasing. Thus

𝒢(ζ)ζ1ζ𝒢(ζ)𝑑ζ𝒩(ζ1)Cexp(C1ζ),for anyζ>ZΦ+1.\mathcal{G}(\zeta)\leq\int_{\zeta-1}^{\zeta}\mathcal{G}(\zeta^{\prime})d\zeta^{\prime}\leq\mathcal{N}(\zeta-1)\leq C\exp\left(-C^{-1}\zeta\right),\quad\text{for any}\quad\zeta>Z_{\Phi}+1.

This completes the proof of the (4.86) by choosing σ=C1\sigma=C^{-1}.

4.2 Higher-order regularity of weak solutions

4.2.1 𝑯𝒎\boldsymbol{H^{m}}-estimates of weak solutions

Given an arbitrary ϕCc(𝒮¯,)\phi\in C_{c}^{\infty}(\overline{\mathcal{S}}\,,\,\mathbb{R}), with ϕ=0\phi=0 on 𝒮\partial\mathcal{S}, direct calculation shows 𝝋:=(x2ϕ,x1ϕ)\boldsymbol{\varphi}:=(-\partial_{x_{2}}\phi\,,\,\partial_{x_{1}}\phi) defines a well-defined test function in Definition 3.4. By replacing 𝝋\boldsymbol{\varphi} with (x2ϕ,x1ϕ)(-\partial_{x_{2}}\phi\,,\,\partial_{x_{1}}\phi) in (3.62), and denoting ω=x2v1x1v2\omega=\partial_{x_{2}}v_{1}-\partial_{x_{1}}v_{2}, one deduces

𝒮ωΔϕ𝑑x+𝒮(α2κ)vtanϕ𝒏𝑑S+𝒮𝒗ωϕdx=𝒮(Δb𝒂b)ϕ𝑑x+𝒮(𝒗𝒂+𝒂𝒗)ϕdx,\begin{split}-\int_{\mathcal{S}}\omega\Delta\phi dx+\int_{\partial\mathcal{S}}(\alpha-2\kappa)v_{\mathrm{tan}}\frac{\partial\phi}{\partial\boldsymbol{n}}dS+\int_{\mathcal{S}}\boldsymbol{v}\cdot\nabla\omega\cdot\phi dx=&\int_{\mathcal{S}}\big{(}\Delta b-\boldsymbol{a}\cdot\nabla b\big{)}\cdot\phi dx\\ &+\int_{\mathcal{S}}\left(\boldsymbol{v}\cdot\nabla\boldsymbol{a}+\boldsymbol{a}\cdot\nabla\boldsymbol{v}\right)^{\perp}\cdot\nabla\phi dx,\end{split}

where b=x2a1x1a2b=\partial_{x_{2}}a_{1}-\partial_{x_{1}}a_{2}. This implies ω\omega solves the following linear elliptic problem weakly:

{Δω+𝒗ω=(Δb𝒂b)(𝒗𝒂+𝒂𝒗),in𝒮;ω=(2κ+α)vtan,on𝒮.\left\{\begin{array}[]{ll}-\Delta\omega+\boldsymbol{v}\cdot\nabla\omega=(\Delta b-\boldsymbol{a}\cdot\nabla b)-\nabla\cdot\left(\boldsymbol{v}\cdot\nabla\boldsymbol{a}+\boldsymbol{a}\cdot\nabla\boldsymbol{v}\right)^{\perp},&\text{in}\quad\mathcal{S};\\[5.69054pt] \omega=\left(-2\kappa+\alpha\right)v_{\mathrm{tan}},&\text{on}\quad\partial\mathcal{S}.\end{array}\right. (4.99)

Here 𝒗H1(𝒮)\boldsymbol{v}\in H^{1}(\mathcal{S}) is treated as a known function solved in Section 3.2, while 𝒂\boldsymbol{a} is the smooth divergence-free flux carrier constructed in Section 3.1.

To study bounds of higher-order norms of the solution, we split the problem (4.99) into a sequence of problems on bounded smooth domains. Recall the definition of 𝔖k\mathfrak{S}_{k} in (4.85), we denote the related cut-off function

ψk={ψ(x13kZΦ2),fork>0;ψ(y13kZΦ2),fork<0,\psi_{k}=\left\{\begin{array}[]{ll}\psi\left(x_{1}-\frac{3kZ_{\Phi}}{2}\right),&\quad\text{for}\quad k>0;\\[2.84526pt] \psi\left(y_{1}-\frac{3kZ_{\Phi}}{2}\right),&\quad\text{for}\quad k<0,\end{array}\right.

where ψ\psi is a smooth 1D cut-off function that satisfies:

{suppψ[9ZΦ/10,9ZΦ/10];ψ1,in [4ZΦ/5,4ZΦ/5];0ψ1,in [ZΦ,ZΦ];|ψ(m)|C/ZΦmC,for m=1,2.\left\{\begin{array}[]{*{2}{ll}}\mathrm{supp}\,\psi\subset\left[-9Z_{\Phi}/10\,,9Z_{\Phi}/10\right];\\ \psi\equiv 1,&\quad\text{in }\left[-{4Z_{\Phi}}/{5}\,,{4Z_{\Phi}}/{5}\right];\\ 0\leq\psi\leq 1,&\quad\text{in }\left[-Z_{\Phi}\,,Z_{\Phi}\right];\\ |\psi^{(m)}|\leq C/Z_{\Phi}^{m}\leq C,&\quad\text{for }m=1,2.\\ \end{array}\right.

Meanwhile, ψ0\psi_{0} is a 2D smooth cut-off function that enjoys

{suppψ0𝒮9ZΦ/10;ψ1,in 𝒮4ZΦ/5;0ψ1,in 𝒮ZΦ;|ψ(m)|C/ZΦmC,for m=1,2.\left\{\begin{array}[]{*{2}{ll}}\mathrm{supp}\,\psi_{0}\subset\mathcal{S}_{9Z_{\Phi}/10};\\ \psi\equiv 1,&\quad\text{in }\mathcal{S}_{4Z_{\Phi}/5};\\ 0\leq\psi\leq 1,&\quad\text{in }\mathcal{S}_{Z_{\Phi}};\\ |\psi^{(m)}|\leq C/Z_{\Phi}^{m}\leq C,&\quad\text{for }m=1,2.\\ \end{array}\right.

However, domains 𝔖k\mathfrak{S}_{k} given in previous subsection are only Lipschitzian, which may cause unnecessary difficulty in deriving higher-order regularity of ω\omega. To this end, we introduce 𝔖~k\tilde{\mathfrak{S}}_{k}, a bounded smooth domain which contains 𝔖k\mathfrak{S}_{k}, with its boundary 𝔖~k𝔖k𝒮\partial\tilde{\mathfrak{S}}_{k}\supset\partial\mathfrak{S}_{k}\cap\partial\mathcal{S}. In order to make the constants of specific inequalities (i.e. imbedding inequalities, trace inequalities, Biot-Savart law) on each 𝔖~k\tilde{\mathfrak{S}}_{k} (kk\in\mathbb{Z}) being uniform, one chooses every 𝔖~k\tilde{\mathfrak{S}}_{k} with k>0k>0 to be congruent to 𝔖~1\tilde{\mathfrak{S}}_{1}, and every 𝔖~k\tilde{\mathfrak{S}}_{k} with k<0k<0 to be congruent to 𝔖~1\tilde{\mathfrak{S}}_{-1}. This can be guaranteed by the definition of 𝔖k\mathfrak{S}_{k}. By the splitting and constructions above, the “distorted part” in the middle of the stripe is totally contained in 𝔖0𝔖~0\mathfrak{S}_{0}\subset\tilde{\mathfrak{S}}_{0}, and ψk\nabla\psi_{k} are totally supported away from this“distorted part” for each kk\in\mathbb{Z}.

Multiplying (4.99)1 by ψk\psi_{k}, we can convert the problem (4.99) to related problem in domain 𝔖~k\tilde{\mathfrak{S}}_{k}, with kk\in\mathbb{Z}:

{Δωk+𝒗ωk=𝑭k+𝒇k,in𝔖~k;ωk=gk,on𝔖~k.\left\{\begin{array}[]{ll}-\Delta\omega_{k}+\boldsymbol{v}\cdot\nabla\omega_{k}=\nabla\cdot\boldsymbol{F}_{k}+\boldsymbol{f}_{k},&\text{in}\quad\tilde{\mathfrak{S}}_{k};\\[5.69054pt] \omega_{k}=g_{k},&\text{on}\quad\partial\tilde{\mathfrak{S}}_{k}.\end{array}\right.

Here ωk=ψkω\omega_{k}=\psi_{k}\omega, while

𝑭k=ψk(𝒗𝒂+𝒂𝒗)2ωψk;𝒇k=ψk(Δb𝒂b)+ψk(𝒗𝒂+𝒂𝒗)+ω(Δψk+𝒗ψk);gk=(2κ+α)ψk𝒗tan.\begin{split}\boldsymbol{F}_{k}=&-\psi_{k}\left(\boldsymbol{v}\cdot\nabla\boldsymbol{a}+\boldsymbol{a}\cdot\nabla\boldsymbol{v}\right)^{\perp}-2\omega\nabla\psi_{k};\\[5.69054pt] \boldsymbol{f}_{k}=&\,\,\psi_{k}\left(\Delta b-\boldsymbol{a}\cdot\nabla b\right)+\nabla\psi_{k}\cdot\left(\boldsymbol{v}\cdot\nabla\boldsymbol{a}+\boldsymbol{a}\cdot\nabla\boldsymbol{v}\right)^{\perp}+\omega\left(\Delta\psi_{k}+\boldsymbol{v}\cdot\nabla\psi_{k}\right);\\[5.69054pt] g_{k}=&\left(-2\kappa+\alpha\right)\psi_{k}\boldsymbol{v}_{\mathrm{tan}}.\end{split}

Using Gagliardo-Nirenberg interpolation together with the trace theorem, it is not hard to derive

𝑭kL2(𝔖~k)+𝒇kL2(𝔖~k)+gkH1/2(𝔖~k)C𝒗H1(𝔖k)(1+ΦeCΦ+𝒗H1(𝔖k)),k.\|\boldsymbol{F}_{k}\|_{L^{2}(\tilde{\mathfrak{S}}_{k})}+\|\boldsymbol{f}_{k}\|_{L^{2}(\tilde{\mathfrak{S}}_{k})}+\|g_{k}\|_{H^{1/2}(\partial\tilde{\mathfrak{S}}_{k})}\leq C\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\left(1+\Phi e^{C\Phi}+\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\right),\quad\forall k\in\mathbb{Z}. (4.100)

Noting that the constant CC above is independent with kk, due to congruent property of domains {𝔖~k}k\{\tilde{\mathfrak{S}}_{k}\}_{k\in\mathbb{Z}}. Therefore, using the classical theory of elliptic equations and (4.100), one derives

ωkH1(𝔖~k)C𝒗H1(𝔖k)(1+ΦeCΦ+𝒗H1(𝔖k)),k.\|\omega_{k}\|_{H^{1}(\tilde{\mathfrak{S}}_{k})}\leq C\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\left(1+\Phi e^{C\Phi}+\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\right),\quad\forall k\in\mathbb{Z}.

Applying the Biot-Savart law, one derives

𝒗H2(𝔖k)C(ωkH1(𝔖~k)+𝒗L2(𝔖k))C𝒗H1(𝔖k)(1+ΦeCΦ+𝒗H1(𝔖k)),k,\|\boldsymbol{v}\|_{H^{2}(\mathfrak{S}_{k}^{\prime})}\leq C\left(\|\omega_{k}\|_{H^{1}(\tilde{\mathfrak{S}}_{k})}+\|\boldsymbol{v}\|_{L^{2}(\mathfrak{S}_{k}^{\prime})}\right)\leq C\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\left(1+\Phi e^{C\Phi}+\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\right),\quad\forall k\in\mathbb{Z},

where

𝔖k={x𝔖k:ψk=1}.\mathfrak{S}_{k}^{\prime}=\{x\in\mathfrak{S}_{k}:\,\psi_{k}=1\}.

This implies, by summing over kk\in\mathbb{Z}, that

𝒗H2(𝒮)2k𝒗H2(𝔖k)2Ck𝒗H1(𝔖k)2(1+ΦeCΦ+𝒗H1(𝔖k))2C𝒗H1(𝒮)2(1+Φ2eCΦ+𝒗H1(𝒮)2).\begin{split}\|\boldsymbol{v}\|^{2}_{H^{2}(\mathcal{S})}\leq&\sum_{k\in\mathbb{Z}}\|\boldsymbol{v}\|^{2}_{H^{2}(\mathfrak{S}_{k}^{\prime})}\\ \leq&C\sum_{k\in\mathbb{Z}}\|\boldsymbol{v}\|^{2}_{H^{1}(\mathfrak{S}_{k})}\left(1+\Phi e^{C\Phi}+\|\boldsymbol{v}\|_{H^{1}(\mathfrak{S}_{k})}\right)^{2}\\ \leq&C\|\boldsymbol{v}\|^{2}_{H^{1}(\mathcal{S})}\left(1+\Phi^{2}e^{C\Phi}+\|\boldsymbol{v}\|^{2}_{H^{1}(\mathcal{S})}\right).\\ \end{split}

This concludes the global H2H^{2}-regularity estimate of 𝒗\boldsymbol{v}. From this, similarly as we derive (4.100), one achieves a “one-order upper” regularity of 𝑭k\boldsymbol{F}_{k}, 𝒇k\boldsymbol{f}_{k} and gkg_{k}, for any kk\in\mathbb{Z}, that is

𝑭kH1(𝔖~k)+𝒇kH1(𝔖~k)+gkH3/2(𝔖~k)CΦ,\|\boldsymbol{F}_{k}\|_{H^{1}(\tilde{\mathfrak{S}}_{k})}+\|\boldsymbol{f}_{k}\|_{H^{1}(\tilde{\mathfrak{S}}_{k})}+\|g_{k}\|_{H^{3/2}(\partial\tilde{\mathfrak{S}}_{k})}\leq C_{\Phi},

which indicates the H3H^{3}-regularity of 𝒗\boldsymbol{v}. Following this bootstrapping argument, one deduces 𝒗\boldsymbol{v} is smooth and

𝒗Hm(𝒮)CΦ,m,m.\|\boldsymbol{v}\|_{H^{m}(\mathcal{S})}\leq C_{\Phi,m},\quad\forall m\in\mathbb{N}.

This finished the proof of the regularity part of Theorem 1.6.

4.2.2 Exponential decay of higher-order norms

Finally, the higher-order regularity and the H1H^{1}-exponential decay estimate in previous subsection, indicates the higher-order exponential decay. In fact, using Sobolev imbedding, we first need to show the following decay of the solution in HmH^{m} norms, with m2m\geq 2:

𝒗Hm(𝒮L\𝒮ζ))+𝒗Hm(𝒮R\𝒮ζ)CΦ,m(𝒗H1(𝒮L\𝒮ζZΦ)+𝒗H1(𝒮R\𝒮ζZΦ)),\begin{split}&\|\boldsymbol{v}\|_{H^{m}(\mathcal{S}_{L}\backslash\mathcal{S}_{\zeta}))}+\|\boldsymbol{v}\|_{H^{m}(\mathcal{S}_{R}\backslash\mathcal{S}_{\zeta})}\leq C_{\Phi,m}\left(\|\boldsymbol{v}\|_{H^{1}(\mathcal{S}_{L}\backslash\mathcal{S}_{\zeta-Z_{\Phi}})}+\|\boldsymbol{v}\|_{H^{1}(\mathcal{S}_{R}\backslash\mathcal{S}_{\zeta-Z_{\Phi}})}\right),\\[2.84526pt] \end{split}

for all ζ>2ZΦ\zeta>2Z_{\Phi}. This is derived by using the method in the proof of Section 4.2.1, but summing over kk\in\mathbb{Z} such that

suppψk(𝒮\𝒮ζ).\mathrm{supp}\,\psi_{k}\cap\left(\mathcal{S}\backslash\mathcal{S}_{\zeta}\right)\neq\varnothing.

Then, the proof is completed by the H1H^{1} decay estimate (4.86). This finishes the proof of Theorem 1.6.

Remark 4.3.

For the pressure pp, there exists two constants CL,CR>0C_{L},\,C_{R}>0 (See (1.5) and (1.4)) , and a smooth cut-off function η\eta given in (3.54) such that: For any m0m\geq 0,

m(p+CRx1η(s)𝑑sCLy1η(s)𝑑s)L2(𝒮)CΦ,m.\left\|\nabla^{m}\nabla\left(p+{C_{R}\int_{-\infty}^{x_{1}}\eta(s)ds}-{C_{L}\int_{-\infty}^{-y_{1}}\eta(s)ds}\right)\right\|_{L^{2}(\mathcal{S})}\leq C_{\Phi,m}.

Meanwhile, the following pointwise decay estimate holds: for all |x|>>1|x|>>1,

|m(p+CRx1η(s)𝑑sCLy1η(s)𝑑s)(x)|CΦ,mexp{σΦ,m|x|},\left|\nabla^{m}\nabla\left(p+{C_{R}\int_{-\infty}^{x_{1}}\eta(s)ds}-{C_{L}\int_{-\infty}^{-y_{1}}\eta(s)ds}\right)(x)\right|\leq C_{\Phi,m}\exp\left\{-\sigma_{\Phi,m}|x|\right\},

where CΦ,mC_{\Phi,m} and σΦ,m\sigma_{\Phi,m} are positive constants depending on Φ\Phi and mm. The subtracted term

π𝑷:=CRx1η(s)𝑑s+CLy1η(s)𝑑s\pi_{\boldsymbol{P}}:=-{C_{R}\int_{-\infty}^{x_{1}}\eta(s)ds}+{C_{L}\int_{-\infty}^{-y_{1}}\eta(s)ds}

is set to balance the pressure of the Poiseuille flows.

Data availability statement

Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.

Conflict of interest statement

The authors declare that they have no conflict of interest.

Acknowledgments

Z. Li is supported by Natural Science Foundation of Jiangsu Province (No. BK20200803) and National Natural Science Foundation of China (No. 12001285). X. Pan is supported by National Natural Science Foundation of China (No. 11801268, 12031006). J. Yang is supported by National Natural Science Foundation of China (No. 12001429).

References

  • [1] T. Acevedo, C. Amrouche, C. Conca and A. Ghosh: Stokes and Navier-Stokes equations with Navier boundary conditions. J. Differential Equations 285 (2021), 258–320.
  • [2] K. Ames and L. Payne: Decay estimates in steady pipe flow. SIAM J. Math. Anal. 20 (1989), no. 4, 789–815.
  • [3] C. Amick: Steady solutions of the Navier-Stokes equations in unbounded channels and pipes. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 4 (1977), no. 3, 473–513.
  • [4] C. Amick: Properties of steady Navier-Stokes solutions for certain unbounded channels and pipes. Nonlinear Anal. 2 (1978), no. 6, 689–720.
  • [5] C. Amick and L. Fraenkel: Steady solutions of the Navier-Stokes equations representing plane flow in channels of various types. Acta Math. 2 (1980), no. 1–2, 83–151.
  • [6] C. Amrouche and A. Rejaiba: LpL^{p}-theory for Stokes and Navier-Stokes equations with Navier boundary condition, J. Differential Equations (256) (2014), no. 4, 1515–1547.
  • [7] M. Bogovskiĭ: Solution of the first boundary value problem for an equation of continuity of an incompressible medium, (Russian) Dokl. Akad. Nauk SSSR (248) 5 (1979), no. 3, 1037–1040.
  • [8] M. Bogovskiĭ: Solutions of some problems of vector analysis, associated with the operators div and grad, (Russian) Theory of cubature formulas and the application of functional analysis to problems of mathematical physics, pp. 5–40, 149, Trudy Sem. S. L. Soboleva, No. 1, 1980, Akad. Nauk SSSR Sibirsk. Otdel., Inst. Mat., Novosibirsk, 1980.
  • [9] G. De Rham: Differentiable manifolds. Forms, currents, harmonic forms. Translated from the French by F. R. Smith. With an introduction by S. S. Chern, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 266. Springer-Verlag, Berlin, 1984.
  • [10] G. Galdi: An introduction to the mathematical theory of the Navier-Stokes equations: Steady-state problems, Second edition. Springer Monographs in Mathematics. Springer, New York, 2011.
  • [11] A. Ghosh: Navier-Stokes equations with Navier boundary condition. Thesis (Ph.D.) Université de Pauet des Pays de Adour, France and Universidad del Paĺs Vasco, Bilbao. 2018, 168 pp.
  • [12] C. Horgan and L. Wheeler: Spatial decay estimates for the Navier-Stokes equations with application to the problem of entry flow. SIAM J. Appl. Math. 35 (1978), no. 1, 97–116.
  • [13] V.A. Kondrat’ev and O.A. Olenik, Boundary value problems for a system in elasticity theory in unbounded domains. Korn inequalities. Russ. Math. Surv. 43 (1988), 65–119.
  • [14] P. Konieczny: On a steady flow in a three-dimensional infinite pipe. Colloq. Math. 104 (2006), no. 1, 33–56.
  • [15] O. Ladyženskaya: Investigation of the Navier-Stokes equation for stationary motion of an incompressible fluid. (Russian) Uspehi Mat. Nauk 14 (1959), no. 3, 75–97.
  • [16] O. Ladyženskaya: Stationary motion of viscous incompressible fluids in pipes. Soviet Physics. Dokl. 124 (4) (1959), 68–70 (551–553 Dokl. Akad. Nauk SSSR).
  • [17] O. Ladyženskaya and V. Solonnikov: Determination of the solutions of boundary value problems for stationary Stokes and Navier-Stokes equations having an unbounded Dirichlet integral, Translated from Zapiski Nauchnykh Seminarov Leningradskogo Otdeleniya Matematicheskogo Instituta im. V. A. Steklova Akad. Nauk SSSR, Vol. 96, pp. 117-160, 1980.
  • [18] Z. Li X. Pan and J. Yang: Characterization of bounded smooth solutions to the axially symmetric Navier-Stokes equations in an infinite pipe with Navier-slip boundary. arXiv: 2110.02445.
  • [19] Z. Li X. Pan and J. Yang: On Leray’s problem in an infinite-long pipe with the Navier-slip boundary condition. arXiv: 2204.10578.
  • [20] J. Lions: Quelques methodes de résolution des problèmes aux limites non linéaires, Dunod, Paris, 1969.
  • [21] P. Mucha: On Navier-Stokes equations with silp boundary conditions in an infinite pipe. Acta Appl. Math. (76) (2003), 1–15.
  • [22] P. Mucha: Asymptotic behavior of a steady flow in a two-dimensional pipe. Studia Math. 158 (2003), no. 1, 39–58.
  • [23] C. Navier: Sur les lois du mouvement des fuides. Mem. Acad. R. Sci. Inst. France, 6, (1823), 389–440.
  • [24] K. Piletskas: On the asymptotic behavior of solutions of a stationary system of Navier-Stokes equations in a domain of layer type. (Russian) Mat. Sb. 193 (2002), no. 12, 69–104; translation in Sb. Math. 193 (2002), no. 11–12, 1801–1836.
  • [25] K. Sha, Y. Wang and C. Xie: On the Steady Navier-Stokes system with Navier slip boundary conditions in two-dimensional channels arXiv: 2210.15204.
  • [26] K. Sha, Y. Wang and C. Xie: On the Leray problem for steady flows in two-dimensional infinitely long channels with slip boundary conditions. arXiv: 2210.16833.
  • [27] R. Temam: Navier-Stokes equations: Theory and numerical analysis, AMS Chelsea edition. American Mathematical Society \cdot Providence, Rhode Island, 2001.
  • [28] Y. Wang and C. Xie: Uniform structural stability of Hagen-Poiseuille flows in a pipe. Comm. Math. Phys. 393 (2022), no. 3, 1347–1410.
  • [29] Y. Wang and C. Xie: Existence and asymptotic behavior of large axisymmetric solutions for steady Navier-Stokes system in a pipe. Arch. Ration. Mech. Anal. 243 (2022), no. 3, 1325–1360.
  • [30] Y. Wang and C. Xie: Uniqueness and uniform structural stability of Poiseuille flows in an infinitely long pipe with Navier boundary conditions. arXiv: 2111.08200.
  • [31] J. Watanabe: On incompressible viscous fluid flows with slip boundary conditions. J. Comput. Appl. Math. 159 (2003), no. 1, 161–172.
  • [32] J. Yang and H. Yin: On the steady non-Newtonian fluids in domains with noncompact boundaries. SIAM J. Math. Anal. 50 (2018), no. 1, 283–338.

Z. Li: School of Mathematics and Statistics, Nanjing University of Information Science and Technology, Nanjing 210044, China

E-mail address: [email protected]

X. Pan: College of Mathematics and Key Laboratory of MIIT, Nanjing University of Aeronautics and Astronautics, Nanjing 211106, China

E-mail address: [email protected]

J. Yang: School of Mathematics and Statistics, Northwestern Polytechnical University, Xi’an 710129, China

E-mail address: [email protected]