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

Existence of planar non-symmetric stationary flows
with large flux in an exterior disk

Mitsuo Higaki

Abstract. This paper is concerned with the two-dimensional stationary Navier-Stokes system in the domain exterior to the unit disk. The existence of solutions with critical decay O(|x|1)O(|x|^{-1}) is established around some explicit flows with large flux. The solutions are obtained for non-symmetric external forces, and moreover, are unique in a certain class.

1 Introduction

We consider the two-dimensional stationary Navier-Stokes system in the exterior disk Ω={x=(x1,x2)2||x|>1}\Omega=\{x=(x_{1},x_{2})\in\mathbb{R}^{2}~{}|~{}|x|>1\}:

{Δu+p=uu+finΩdivu=0inΩu=αxγxonΩu(x)0as|x|.\left\{\begin{array}[]{ll}-\Delta u+\nabla p=-u\cdot\nabla u+f&\mbox{in}\ \Omega\\ {\rm div}\,u=0&\mbox{in}\ \Omega\\ u=\alpha x^{\bot}-\gamma x&\mbox{on}\ \partial\Omega\\ u(x)\to 0&\mbox{as}\ |x|\to\infty.\end{array}\right. (NS)

The unknown functions u=(u1(x),u2(x))u=(u_{1}(x),u_{2}(x)) and p=p(x)p=p(x) are respectively the velocity field and the pressure field. The function f=(f1(x),f2(x))f=(f_{1}(x),f_{2}(x)) is a given external force in L2(Ω)2L^{2}(\Omega)^{2}. We assume that both α\alpha and γ\gamma are constants. The vector xx^{\bot} refers to (x2,x1)(-x_{2},x_{1}). The system (NS) describes the motion of a viscous incompressible fluid around the disk rotating at angular velocity α\alpha and whose surface subjects to a normal suction velocity γx-\gamma x.

The existence and uniqueness theories for the problem (NS) are generally open. As for the existence, there are two fundamental difficulties. The first one is lack of certain embeddings. Let ff be smooth and compactly supported in Ω\Omega for brevity. Under smallness conditions, adapting the proof by Leray [15] or relying on Fijita [2], we can actually find weak solutions of (NS) having a finite Dirichlet integral, called DD-solutions. Nevertheless, the bound uL2<\|\nabla u\|_{L^{2}}<\infty itself cannot verify the condition at spatial infinity in (NS) in two-dimensional unbounded domains; see Korobkov, Pileckas and Russo [10, 11, 12] for recent progress on this topic. The other one is the logarithmic growth in the Green function of the exterior Stokes system, which is the source of the famous Stokes paradox; see [1, 3, 4, 14] for descriptions. The uniqueness of solutions will be discussed in Remark 1.2 (iii) below.

These issues illustrate a delicate aspect concerning the zero condition at infinity in (NS). Most existing results treating (NS) assume both smallness and symmetry on the data. The latter is useful for making quantities decay in space by cancellation. The reader is referred to Galdi [3, 4], Russo [20], Yamazaki [22, 23, 24] and Pileckas and Russo [19] for the existence of solutions, and to Nakatsuka [18] and [24] for the uniqueness. Relevant numerical simulations are carried out in Guillod and Wittwer [7]. We note that, if another condition were imposed at infinity, the situation would be quite different. Indeed, the recent work by Korobkov and Ren [13] proves the uniqueness in the class of DD-solutions for plane exterior Navier-Stokes flows converging to a small but non-zero constant vector field at infinity.

In this paper, we examine the problem (NS) from a different angle. Note that, when the external force ff is trivial, there is an explicit solution (αUγW,Pα,γ)(\alpha U-\gamma W,\nabla P_{\alpha,\gamma}) given by

U(x)=x|x|2,W(x)=x|x|2,Pα,γ(x)=(|αU(x)γW(x)|22),\displaystyle\begin{split}U(x)=\frac{x^{\bot}}{|x|^{2}},\qquad W(x)=\frac{x}{|x|^{2}},\qquad\nabla P_{\alpha,\gamma}(x)=-\nabla\Big{(}\frac{|\alpha U(x)-\gamma W(x)|^{2}}{2}\Big{)},\end{split} (1.1)

which is invariant under the scaling of the Navier-Stokes equations. A (non-trivial) solution in this class is called scale-critical and it represents the balance between the nonlinear and linear parts of the equations. Given this nature, it is expected that the Navier-Stokes flows around a scale-critical flow differ quantitatively from those around the trivial flow. In fact, Hillairet and Wittwer [9] consider the Navier-Stokes problem in an exterior disk by perturbating the system around (αU,Pα,0)(\alpha U,\nabla P_{\alpha,0}). A crucial observation is that the decay of solutions to the corresponding linearized system is improved when |α||\alpha| is sufficiently large, more precisely, when |α|>48|\alpha|>\sqrt{48}. This is possible because of the structure of the equation for vorticities. Then, based on iteration to the nonlinear problem with subcritialized nonlinearity, they show the existence of solutions in the form of u(x)=αU(x)+o(|x|1)u(x)=\alpha U(x)+o(|x|^{-1}) when |x||x|\to\infty. These solutions are driven by inhomogeneous boundary data, on which no symmetries are imposed thanks to the mechanism of the proof.

The result in [9] can be read as the scale-critical flow αU\alpha U producing a stabilizing effect in view of spatial decay. In this context, we here consider such an effect of αUγW\alpha U-\gamma W and address the existence of solutions to the problem (NS). Briefly, we will see that the stabilization is effective for any α\alpha if γ>2\gamma>2. As an application, we obtain the Navier-Stokes flows for non-symmetric external forces based on the perturbation. We also describe the asymptotics near spatial infinity in terms of the decay rate of external forces.

Let us introduce some notations to state the main result. For s0s\geq 0, we define

Ls(Ω)={fL(Ω)|fLs<},fLs:=esssupxΩ|x|s|f(x)|,\displaystyle\begin{split}L^{\infty}_{s}(\Omega)&=\{f\in L^{\infty}(\Omega)~{}|~{}\|f\|_{L^{\infty}_{s}}<\infty\},\qquad\|f\|_{L^{\infty}_{s}}:=\operatorname*{ess\,sup}_{x\in\Omega}\,|x|^{s}|f(x)|,\end{split} (1.2)

which is a Banach space under the norm Ls\|\cdot\|_{L^{\infty}_{s}}. Taking advantage of symmetry, we introduce the polar coordinates on Ω\Omega as

x1=rcosθ,x2=rsinθ,r=|x|1,θ[0,2π),\displaystyle x_{1}=r\cos\theta,\qquad x_{2}=r\sin\theta,\quad r=|x|\geq 1,\quad\theta\in[0,2\pi),
𝐞r=x|x|,𝐞θ=x|x|=θ𝐞r.\displaystyle{\bf e}_{r}=\frac{x}{|x|},\qquad{\bf e}_{\theta}=\frac{x^{\bot}}{|x|}=\partial_{\theta}{\bf e}_{r}.

For given vector field v=(v1,v2)v=(v_{1},v_{2}) on Ω\Omega and nn\in\mathbb{Z}, we set

v=vr(r,θ)𝐞r+vθ(r,θ)𝐞θ,vr=v𝐞r,vθ=v𝐞θ\displaystyle v=v_{r}(r,\theta){\bf e}_{r}+v_{\theta}(r,\theta){\bf e}_{\theta},\qquad v_{r}=v\cdot{\bf e}_{r},\qquad v_{\theta}=v\cdot{\bf e}_{\theta}

and denote by 𝒫n\mathcal{P}_{n} the projection on the Fourier mode nn:

𝒫nv(r,θ)=vr,n(r)einθ𝐞r+vθ,n(r)einθ𝐞θ,vr,n(r):=12π02πvr(rcosθ,rsinθ)einθdθ,vθ,n(r):=12π02πvθ(rcosθ,rsinθ)einθdθ.\displaystyle\begin{split}\mathcal{P}_{n}v(r,\theta)&=v_{r,n}(r)e^{in\theta}{\bf e}_{r}+v_{\theta,n}(r)e^{in\theta}{\bf e}_{\theta},\\ v_{r,n}(r)&:=\frac{1}{2\pi}\int_{0}^{2\pi}v_{r}(r\cos\theta,r\sin\theta)e^{-in\theta}\,{\rm d}\theta,\\ v_{\theta,n}(r)&:=\frac{1}{2\pi}\int_{0}^{2\pi}v_{\theta}(r\cos\theta,r\sin\theta)e^{-in\theta}\,{\rm d}\theta.\end{split} (1.3)

Now the result is stated as follows.

Theorem 1.1

Let α\alpha\in\mathbb{R} and γ>2\gamma>2. Then, for any 2<ρ<32<\rho<3 with ρmin{γ,3}\rho\leq\min\{\gamma,3\}, there is a constant ε=ε(α,γ,ρ)>0\varepsilon=\varepsilon(\alpha,\gamma,\rho)>0 such that if fL2ρ1(Ω)2f\in L^{\infty}_{2\rho-1}(\Omega)^{2} satisfies

n𝒫nfL2ρ1ε,\displaystyle\sum_{n\in\mathbb{Z}}\|\mathcal{P}_{n}f\|_{L^{\infty}_{2\rho-1}}\leq\varepsilon, (1.4)

then there is a solution (u,p)(W^1,2(Ω)Wloc2,2(Ω¯)L1(Ω))2×Lloc2(Ω¯)2(u,\nabla p)\in\big{(}\widehat{W}^{1,2}(\Omega)\cap W^{2,2}_{{\rm loc}}(\overline{\Omega})\cap L^{\infty}_{1}(\Omega)\big{)}^{2}\times L^{2}_{{\rm loc}}(\overline{\Omega})^{2} of (NS) unique in a suitable set of functions (see Section 4 for the precise definition). Moreover, when |x||x|\rightarrow\infty, the solution u=u(x)u=u(x) behaves as

u(x)=(αx|x|2γx|x|2)+O(|x|ρ+1),rotu(x)=O(|x|ρ).\displaystyle u(x)=\Big{(}\alpha\frac{x^{\bot}}{|x|^{2}}-\gamma\frac{x}{|x|^{2}}\Big{)}+O(|x|^{-\rho+1}),\qquad{\rm rot}\,u(x)=O(|x|^{-\rho}). (1.5)
Remark 1.2
  1. (i)

    As far as the author knows, Theorem 1.1 is the first result obtaining the plane exterior Navier-Stokes flows zero at spatial infinity, non-symmetric and unique in a certain class. In [9] dealing with the case γ=0\gamma=0, the uniqueness of solutions seems not to be provided. More to the point, the proof should not be easy since the existence of solutions in [9] is verified by the intermediate value theorem. The inconvenience is closely related to the structure of the zero mode of the linearized system (1.8) below.

  2. (ii)

    Contrary to [9] where |α|>48|\alpha|>\sqrt{48} is assumed, we do not need any restrictions on α\alpha in Theorem 1.1. Thus it is interpreted that the stabilization from γW-\gamma W with γ>2\gamma>2 exceeds that from αU\alpha U, and hence, the theorem holds thanks to the presence of γ\gamma.

  3. (iii)

    Theorem 1.1 does not state the uniqueness of DD-solutions of (NS), which is violated even when f=0f=0 by the following Hamel solutions. For γ>2\gamma>2 and tt\in\mathbb{R}, we set

    A(x)=1γ2(1|x|γ+2)U(x),Q(x)=(|αU(x)γW(x)+tA(x)|22)[t|x|γ{αγ+tγ2(1γ12(γ1)|x|γ+2)}].\displaystyle\begin{split}A(x)&=\frac{1}{\gamma-2}(1-|x|^{-\gamma+2})U(x),\\ \nabla Q(x)&=-\nabla\Big{(}\frac{|\alpha U(x)-\gamma W(x)+tA(x)|^{2}}{2}\Big{)}\\ &\quad-\nabla\bigg{[}t|x|^{-\gamma}\Big{\{}\frac{\alpha}{\gamma}+\frac{t}{\gamma-2}\Big{(}\frac{1}{\gamma}-\frac{1}{2(\gamma-1)}|x|^{-\gamma+2}\Big{)}\Big{\}}\bigg{]}.\end{split}

    Then {(αUγW+tA,Q)}t\{(\alpha U-\gamma W+tA,\nabla Q)\}_{t\in\mathbb{R}} is a family of explicit DD-solutions of (NS) with f=0f=0. For more detailed discussion, the reader is referred to [3, Section XII.2].

  4. (iv)

    The stability of the flows in Theorem 1.1 is an open problem. The global L2L^{2}-stability of γW-\gamma W is proved by Guillod [6] and the local L2L^{2}-stability of αUγW\alpha U-\gamma W by Maekawa [16, 17] without any symmetries on initial perturbations. However, these results essentially rely on the smallness of the coefficients and, therefore, cannot be adapted to the flows in Theorem 1.1, even if the Hardy inequality applies to u(αUγW)u-(\alpha U-\gamma W).

The ingredients of the proof of Theorem 1.1 are the following.

(I) Perturbation. We will construct the solution (u,p)(u,\nabla p) of (NS) in the form of

u=αUγW+v,p=(|u|22+q).\displaystyle u=\alpha U-\gamma W+v,\qquad\nabla p=\nabla\Big{(}-\frac{|u|^{2}}{2}+q\Big{)}. (1.6)

The pair (v,q)(v,\nabla q) is understood as the perturbation from (αUγW,Pα,γ)(\alpha U-\gamma W,\nabla P_{\alpha,\gamma}) in response to the external force ff. Inserting the ansatz (1.6) into (NS) and using the relation

uu=urotu+(|u|22),rotu:=1u22u1,\displaystyle u\cdot\nabla u=u^{\bot}{\rm rot}\,u+\nabla\Big{(}\frac{|u|^{2}}{2}\Big{)},\qquad{\rm rot}\,u:=\partial_{1}u_{2}-\partial_{2}u_{1}, (1.7)

as well as rotU=rotW=0{\rm rot}\,U={\rm rot}\,W=0, we see that (v,q)(v,\nabla q) solves the perturbed system

{Δv+(αUγW)rotv+q=vrotv+finΩdivv=0inΩv=0onΩv(x)0as|x|.\left\{\begin{array}[]{ll}-\Delta v+(\alpha U-\gamma W)^{\bot}{\rm rot}\,v+\nabla q=-v^{\bot}{\rm rot}\,v+f&\mbox{in}\ \Omega\\ {\rm div}\,v=0&\mbox{in}\ \Omega\\ v=0&\mbox{on}\ \partial\Omega\\ v(x)\to 0&\mbox{as}\ |x|\to\infty.\end{array}\right. (NS~\widetilde{\mbox{NS}})

Our next task is to prove the existence of solutions to the problem (NS~\widetilde{\mbox{NS}}).

(II) Linear analysis. The linearized system of (NS~\widetilde{\mbox{NS}}) around v=0v=0 is written as

{Δv+(αUγW)rotv+q=finΩdivv=0inΩv=0onΩv(x)0as|x|.\left\{\begin{array}[]{ll}-\Delta v+(\alpha U-\gamma W)^{\bot}{\rm rot}\,v+\nabla q=f&\mbox{in}\ \Omega\\ {\rm div}\,v=0&\mbox{in}\ \Omega\\ v=0&\mbox{on}\ \partial\Omega\\ v(x)\to 0&\mbox{as}\ |x|\to\infty.\end{array}\right. (S~\widetilde{\mbox{S}})

Since (S~\widetilde{\mbox{S}}) is invariant under the action of 𝒫n\mathcal{P}_{n} in (1.3), one can study it in each Fourier mode.

As mentioned in Remark 1.2 (ii), the presence of the parameter γ>2\gamma>2 stabilizes the decay of solutions of (S~\widetilde{\mbox{S}}). A sharp contrast with [9] is that this effect is exerted on the zero mode of the velocity 𝒫0v=vr,0(r)𝐞r+vθ,0(r)𝐞θ\mathcal{P}_{0}v=v_{r,0}(r){\bf e}_{r}+v_{\theta,0}(r){\bf e}_{\theta}. Indeed, while vr,0(r)=0v_{r,0}(r)=0 follows from the second and third lines in (S~\widetilde{\mbox{S}}), vθ,0(r)v_{\theta,0}(r) satisfies the ordinary differential equation

d2vθ,0dr21+γrdvθ,0dr+1γr2vθ,0=fθ,0,r>1,\displaystyle-\frac{\,{\rm d}^{2}v_{\theta,0}}{\,{\rm d}r^{2}}-\frac{1+\gamma}{r}\frac{\,{\rm d}v_{\theta,0}}{\,{\rm d}r}+\frac{1-\gamma}{r^{2}}v_{\theta,0}=f_{\theta,0},\quad r>1, (1.8)

which is independent of α\alpha. Then, since the Green functions of equation are rγ+1r^{-\gamma+1} and r1r^{-1}, the solution vθ,0(r)v_{\theta,0}(r) decays subcritically if the data fθ,0(r)f_{\theta,0}(r) decays fast enough.

The stabilization from γ\gamma displays also in the non-zero modes. However, to prove Theorem 1.1, we need to make precise the relationship between the decay of solutions and that of external forces. For this purpose, we derive the representation formula for the nn-mode of the velocity 𝒫nv=vr,n(r)einθ𝐞r+vθ,n(r)einθ𝐞θ\mathcal{P}_{n}v=v_{r,n}(r)e^{in\theta}{\bf e}_{r}+v_{\theta,n}(r)e^{in\theta}{\bf e}_{\theta} using the streamfunction-vorticity variant of the system in [9] and the Biot-Savart law in [16, 17, 5, 8]. When verifying the formula, we essentially use the decay of the vorticity subcritically improved by γ\gamma.

This paper is organized as follows. In Section 2, we collect preliminary results from vector calculus in the polar coordinates. In Section 3, we study the linear problem (S~\widetilde{\mbox{S}}) decomposed into the Fourier modes. In Section 4, we prove Theorem 1.1.

Notations. We denote by CC the constant and by C(a,b,c,)C(a,b,c,\ldots) the constant depending on a,b,c,a,b,c,\ldots. Both of these may vary from line to line. We use the function spaces

W^1,2(Ω)={pLloc2(Ω¯)|pL2(Ω)2},C0,σ(Ω)={φC0(Ω)2|divφ=0}\widehat{W}^{1,2}(\Omega)=\{p\in L^{2}_{{\rm loc}}(\overline{\Omega})~{}|~{}\nabla p\in L^{2}(\Omega)^{2}\},\quad C^{\infty}_{0,\sigma}(\Omega)=\{\varphi\in C^{\infty}_{0}(\Omega)^{2}~{}|~{}{\rm div}\,\varphi=0\}

and Lσ2(Ω)L^{2}_{\sigma}(\Omega) which is the completion of C0,σ(Ω)C^{\infty}_{0,\sigma}(\Omega) in the L2L^{2}-norm. If there is no confusion, we use the same notation to denote the quantities concerning scalar-, vector- or tensor-valued functions. For example, ,\langle\cdot,\cdot\rangle denotes to the inner product on L2(Ω)L^{2}(\Omega), L2(Ω)2L^{2}(\Omega)^{2} or L2(Ω)2×2L^{2}(\Omega)^{2\times 2}.

2 Preliminaries

This section collects useful facts about the vector calculus in the exterior disk Ω\Omega.

2.1 Operators in the polar coordinates

The following formulas will be used:

divv=1v1+2v2=1rr(rvr)+1rθvθ,rotv=1v22v1=1rr(rvθ)1rθvr,|v|2=|rvr|2+|rvθ|2+1r2(|θvrvθ|2+|vr+θvθ|2),\displaystyle\begin{split}{\rm div}\,v&=\partial_{1}v_{1}+\partial_{2}v_{2}=\frac{1}{r}\partial_{r}(rv_{r})+\frac{1}{r}\partial_{\theta}v_{\theta},\\ {\rm rot}\,v&=\partial_{1}v_{2}-\partial_{2}v_{1}=\frac{1}{r}\partial_{r}(rv_{\theta})-\frac{1}{r}\partial_{\theta}v_{r},\\ |\nabla v|^{2}&=|\partial_{r}v_{r}|^{2}+|\partial_{r}v_{\theta}|^{2}+\frac{1}{r^{2}}(|\partial_{\theta}v_{r}-v_{\theta}|^{2}+|v_{r}+\partial_{\theta}v_{\theta}|^{2}),\end{split} (2.1)

and

Δv={r(1rr(rvr))1r2θ2vr+2r2θvθ}𝐞r+{r(1rr(rvθ))1r2θ2vθ2r2θvr}𝐞θ.\displaystyle\begin{split}-\Delta v&=\Big{\{}-\partial_{r}\Big{(}\frac{1}{r}\partial_{r}(rv_{r})\Big{)}-\frac{1}{r^{2}}\partial_{\theta}^{2}v_{r}+\frac{2}{r^{2}}\partial_{\theta}v_{\theta}\Big{\}}{\bf e}_{r}\\ &\quad+\Big{\{}-\partial_{r}\Big{(}\frac{1}{r}\partial_{r}(rv_{\theta})\Big{)}-\frac{1}{r^{2}}\partial_{\theta}^{2}v_{\theta}-\frac{2}{r^{2}}\partial_{\theta}v_{r}\Big{\}}{\bf e}_{\theta}.\end{split}

2.2 Fourier series

Let nn\in\mathbb{Z}. We define, for a vector field v=v(r,θ)v=v(r,\theta) on Ω\Omega,

vn(r,θ)=𝒫nv(r,θ),\displaystyle v_{n}(r,\theta)=\mathcal{P}_{n}v(r,\theta), (2.2)

where 𝒫n\mathcal{P}_{n} is the projection defined in (1.3), for a scalar function ω=ω(r,θ)\omega=\omega(r,\theta) on Ω\Omega,

𝒫nω(r,θ)=(12π02πω(rcoss,rsins)einsds)einθ,ωn(r)=(𝒫nω)einθ,\displaystyle\begin{split}\mathcal{P}_{n}\omega(r,\theta)&=\bigg{(}\frac{1}{2\pi}\int_{0}^{2\pi}\omega(r\cos s,r\sin s)e^{-ins}\,{\rm d}s\bigg{)}e^{in\theta},\\ \omega_{n}(r)&=(\mathcal{P}_{n}\omega)e^{-in\theta},\end{split} (2.3)

and for a function space X(Ω)Lloc1(Ω¯)2X(\Omega)\subset L^{1}_{\rm loc}(\overline{\Omega})^{2} or X(Ω)Lloc1(Ω¯)X(\Omega)\subset L^{1}_{\rm loc}(\overline{\Omega}),

𝒫nX(Ω)={𝒫nf|fX(Ω)}.\displaystyle\mathcal{P}_{n}X(\Omega)=\big{\{}\mathcal{P}_{n}f~{}\big{|}~{}f\in X(\Omega)\big{\}}.

Our definition of fnf_{n} differs according to whether ff is vectorial or scalar. The former and latter are respectively defined in (2.2) as fn=𝒫nff_{n}=\mathcal{P}_{n}f and in (2.3) as fn=(𝒫nf)einθf_{n}=(\mathcal{P}_{n}f)e^{-in\theta}.

2.3 Biot-Savart law

For a given ωL2(Ω)\omega\in L^{\infty}_{2}(\Omega), we consider the Poisson equation

{Δψ=ωinΩψ=0onΩ.\left\{\begin{array}[]{ll}-\Delta\psi=\omega&\mbox{in}\ \Omega\\ \psi=0&\mbox{on}\ \partial\Omega.\end{array}\right.

The solution ψ\psi is called the streamfunction. Let ω𝒫nL2(Ω)\omega\in\mathcal{P}_{n}L^{\infty}_{2}(\Omega) with |n|1|n|\geq 1 and set ψn=(𝒫nψ)einθ\psi_{n}=(\mathcal{P}_{n}\psi)e^{-in\theta} and ωn=(𝒫nω)einθ\omega_{n}=(\mathcal{P}_{n}\omega)e^{-in\theta}. In the polar coordinates, ψn=ψn(r)\psi_{n}=\psi_{n}(r) solves the ordinary differential equation

d2ψndr21rdψndr+n2r2ψn=ωn,r>1,ψn(1)=0.\displaystyle-\frac{\,{\rm d}^{2}\psi_{n}}{\,{\rm d}r^{2}}-\frac{1}{r}\frac{\,{\rm d}\psi_{n}}{\,{\rm d}r}+\frac{n^{2}}{r^{2}}\psi_{n}=\omega_{n},\quad r>1,\qquad\psi_{n}(1)=0. (2.4)

The decaying solution ψn=ψn[ωn]\psi_{n}=\psi_{n}[\omega_{n}] of (2.4) is given by

ψn[ωn](r)=12|n|(dn[ωn]r|n|+r|n|1rs|n|+1ωn(s)ds+r|n|rs|n|+1ωn(s)ds),dn[ωn]:=1s|n|+1ωn(s)ds.\displaystyle\begin{split}\psi_{n}[\omega_{n}](r)&=\frac{1}{2|n|}\bigg{(}-d_{n}[\omega_{n}]r^{-|n|}\\ &\qquad\qquad+r^{-|n|}\int_{1}^{r}s^{|n|+1}\omega_{n}(s)\,{\rm d}s+r^{|n|}\int_{r}^{\infty}s^{-|n|+1}\omega_{n}(s)\,{\rm d}s\bigg{)},\\ d_{n}[\omega_{n}]&:=\int_{1}^{\infty}s^{-|n|+1}\omega_{n}(s)\,{\rm d}s.\end{split} (2.5)

Then the following vector field Vn[ωn]V_{n}[\omega_{n}] is called the Biot-Savart law:

Vn[ωn](r,θ)=Vr,n[ωn](r)einθ𝐞r+Vθ,n[ωn](r)einθ𝐞θ,Vr,n[ωn]=inrψn[ωn],Vθ,n[ωn]=ddrψn[ωn].\displaystyle\begin{split}&V_{n}[\omega_{n}](r,\theta)=V_{r,n}[\omega_{n}](r)e^{in\theta}{\bf e}_{r}+V_{\theta,n}[\omega_{n}](r)e^{in\theta}{\bf e}_{\theta},\\ &V_{r,n}[\omega_{n}]=\frac{in}{r}\psi_{n}[\omega_{n}],\qquad V_{\theta,n}[\omega_{n}]=-\frac{\,{\rm d}}{\,{\rm d}r}\psi_{n}[\omega_{n}].\end{split} (2.6)

It is straightforward to see that

divVn[ωn]=0,rotVn[ωn]=ωneinθ,(𝐞rVn[ωn])|Ω=0.\displaystyle\begin{split}&{\rm div}\,V_{n}[\omega_{n}]=0,\qquad{\rm rot}\,V_{n}[\omega_{n}]=\omega_{n}e^{in\theta},\qquad({\bf e}_{r}\cdot V_{n}[\omega_{n}])|_{\partial\Omega}=0.\end{split} (2.7)

If additionally ωLρ(Ω)\omega\in L^{\infty}_{\rho}(\Omega) with ρ>2\rho>2, one can check that Vn[ωn]W1,2(Ω)2V_{n}[\omega_{n}]\in W^{1,2}(\Omega)^{2}.

We state two propositions related to the Biot-Savart law. The first one is implicitly contained in [16, Proposition 2.6 and Lemma 3.1] and the second one has the same content as [16, Corollary 2.7]. However, we provide slightly more concise proofs for completeness.

Proposition 2.1

Let |n|1|n|\geq 1 and vn𝒫nW01,2(Ω)2v_{n}\in\mathcal{P}_{n}W^{1,2}_{0}(\Omega)^{2}. Set ωn=(rotvn)n\omega_{n}=({\rm rot}\,v_{n})_{n}. Then, if divvn=0{\rm div}\,v_{n}=0 and ωnLρ(Ω)\omega_{n}\in L^{\infty}_{\rho}(\Omega) for some ρ>2\rho>2, we have vn=Vn[ωn]v_{n}=V_{n}[\omega_{n}] and dn[ωn]=0d_{n}[\omega_{n}]=0 in (2.5).

Proof.

Define u=vnVn[ωn]u=v_{n}-V_{n}[\omega_{n}]. Then one has u𝒫nW1,2(Ω)2u\in\mathcal{P}_{n}W^{1,2}(\Omega)^{2} and

divu=0,rotu=0,(𝐞ru)|Ω=0.\displaystyle\begin{split}&{\rm div}\,u=0,\qquad{\rm rot}\,u=0,\qquad({\bf e}_{r}\cdot u)|_{\partial\Omega}=0.\end{split}

In particular, we have Δu=0\Delta u=0 in the sense of distributions and hence uL2=0\|\nabla u\|_{L^{2}}=0. Thus uu is a constant. Then ur=0u_{r}=0 follows from (𝐞ru)|Ω=0({\bf e}_{r}\cdot u)|_{\partial\Omega}=0 and uθ=0u_{\theta}=0 from divu=1rr(rur)+inrθuθ=0{\rm div}\,u=\frac{1}{r}\partial_{r}(ru_{r})+\frac{in}{r}\partial_{\theta}u_{\theta}=0. Consequently, we have vn=Vn[ωn]v_{n}=V_{n}[\omega_{n}] and, from (2.5) and (2.6),

dn[ωn]=ddrψn[ωn](1)=Vθ,n[ωn](1)=vθ,n(1)=0.\displaystyle\begin{split}d_{n}[\omega_{n}]=\frac{\,{\rm d}}{\,{\rm d}r}\psi_{n}[\omega_{n}](1)=-V_{\theta,n}[\omega_{n}](1)=-v_{\theta,n}(1)=0.\end{split}

The proof is complete. ∎

Proposition 2.2

Let |n|1|n|\geq 1 and fn𝒫nL2(Ω)2f_{n}\in\mathcal{P}_{n}L^{2}(\Omega)^{2}. Then, if rotfn=0{\rm rot}\,f_{n}=0 in the sense of distributions, we have f=𝒫npf=\nabla\mathcal{P}_{n}p for some 𝒫np𝒫nW^1,2(Ω)\mathcal{P}_{n}p\in\mathcal{P}_{n}\widehat{W}^{1,2}(\Omega).

Proof.

We only need to show that fn,φn=0\langle f_{n},\varphi_{n}\rangle=0 for all φn𝒫nC0,σ(Ω)\varphi_{n}\in\mathcal{P}_{n}C^{\infty}_{0,\sigma}(\Omega). Then the assertion follows from the Helmholtz decomposition in L2(Ω)2L^{2}(\Omega)^{2}. Note that, for fn𝒫nC0(Ω)2f_{n}\in\mathcal{P}_{n}C^{\infty}_{0}(\Omega)^{2} and 𝒫nψ𝒫nC0(Ω)\mathcal{P}_{n}\psi\in\mathcal{P}_{n}C^{\infty}_{0}(\Omega), by integration by parts we have

rotfn,ψn=02π1(1rddr(rfθ,n(r))inrfr,n(r))ψn(r)¯rdrdθ=02π1(fr,n(r)inrψn(r)¯fθ,n(r)ddrψn(r)¯)rdrdθ.\displaystyle\begin{split}\langle{\rm rot}\,f_{n},\psi_{n}\rangle&=\int_{0}^{2\pi}\int_{1}^{\infty}\Big{(}\frac{1}{r}\frac{\,{\rm d}}{\,{\rm d}r}(rf_{\theta,n}(r))-\frac{in}{r}f_{r,n}(r)\Big{)}\overline{\psi_{n}(r)}r\,{\rm d}r\,{\rm d}\theta\\ &=\int_{0}^{2\pi}\int_{1}^{\infty}\Big{(}f_{r,n}(r)\overline{\frac{in}{r}\psi_{n}(r)}-f_{\theta,n}(r)\overline{\frac{\,{\rm d}}{\,{\rm d}r}\psi_{n}(r)}\Big{)}r\,{\rm d}r\,{\rm d}\theta.\end{split}

Let φn𝒫nC0,σ(Ω)\varphi_{n}\in\mathcal{P}_{n}C^{\infty}_{0,\sigma}(\Omega) and set ωn=(rotφn)n\omega_{n}=({\rm rot}\,\varphi_{n})_{n}. By Proposition 2.1, φn\varphi_{n} is represented by the Biot-Savart law as φn=Vn[ωn]\varphi_{n}=V_{n}[\omega_{n}]. Then the definition of Vr,n[ωn](r)V_{r,n}[\omega_{n}](r) in (2.6) implies that the streamfunction ψ(r,θ)=ψn[ωn](r)einθ\psi(r,\theta)=\psi_{n}[\omega_{n}](r)e^{in\theta} is smooth and compactly supported in Ω\Omega. Thus the condition that rotfn=0{\rm rot}\,f_{n}=0 in the sense of distributions leads to

02π1(fr,n(r)inrψn[ωn](r)¯fθ,n(r)ddrψn[ωn](r)¯)rdrdθ=0.\displaystyle\begin{split}\int_{0}^{2\pi}\int_{1}^{\infty}\Big{(}f_{r,n}(r)\overline{\frac{in}{r}\psi_{n}[\omega_{n}](r)}-f_{\theta,n}(r)\overline{\frac{\,{\rm d}}{\,{\rm d}r}\psi_{n}[\omega_{n}](r)}\Big{)}r\,{\rm d}r\,{\rm d}\theta=0.\end{split}

The left-hand side can be rewritten as fn,Vn[ωn]=fn,φn\langle f_{n},V_{n}[\omega_{n}]\rangle=\langle f_{n},\varphi_{n}\rangle. Consequently, we have fn,φn=0\langle f_{n},\varphi_{n}\rangle=0 for all φn𝒫nC0,σ(Ω)\varphi_{n}\in\mathcal{P}_{n}C^{\infty}_{0,\sigma}(\Omega). This completes the proof. ∎

3 Linearized problem

In this section, we study the linearized problem

{Δv+(αUγW)rotv+q=finΩdivv=0inΩv=0onΩv(x)0as|x|.\left\{\begin{array}[]{ll}-\Delta v+(\alpha U-\gamma W)^{\bot}{\rm rot}\,v+\nabla q=f&\mbox{in}\ \Omega\\ {\rm div}\,v=0&\mbox{in}\ \Omega\\ v=0&\mbox{on}\ \partial\Omega\\ v(x)\to 0&\mbox{as}\ |x|\to\infty.\end{array}\right. (S~\widetilde{\mbox{S}})

Let nn\in\mathbb{Z}. Applying 𝒫n\mathcal{P}_{n} in (1.3) to (S~\widetilde{\mbox{S}}), we see that (vr,n(r),vθ,n(r))(v_{r,n}(r),v_{\theta,n}(r)) and qn(r)q_{n}(r) satisfy

ddr(1rddr(rvr,n))+n2r2vr,n+2inr2vθ,nαr2(ddr(rvθ,n)invr,n)+rqn=fr,n,r>1,\displaystyle\begin{aligned} &-\frac{\,{\rm d}}{\,{\rm d}r}\Big{(}\frac{1}{r}\frac{\,{\rm d}}{\,{\rm d}r}(rv_{r,n})\Big{)}+\frac{n^{2}}{r^{2}}v_{r,n}+\frac{2in}{r^{2}}v_{\theta,n}\\ &\qquad\qquad-\frac{\alpha}{r^{2}}\Big{(}\frac{\,{\rm d}}{\,{\rm d}r}(rv_{\theta,n})-inv_{r,n}\Big{)}+\partial_{r}q_{n}=f_{r,n},\quad r>1,\\ \end{aligned} (3.1)
ddr(1rddr(rvθ,n))+n2r2vθ,n2inr2vr,nγr2(ddr(rvθ,n)invr,n)+inrqn=fθ,n,r>1,\displaystyle\begin{aligned} &-\frac{\,{\rm d}}{\,{\rm d}r}\Big{(}\frac{1}{r}\frac{\,{\rm d}}{\,{\rm d}r}(rv_{\theta,n})\Big{)}+\frac{n^{2}}{r^{2}}v_{\theta,n}-\frac{2in}{r^{2}}v_{r,n}\\ &\qquad\qquad-\frac{\gamma}{r^{2}}\Big{(}\frac{\,{\rm d}}{\,{\rm d}r}(rv_{\theta,n})-inv_{r,n}\Big{)}+\frac{in}{r}q_{n}=f_{\theta,n},\quad r>1,\end{aligned} (3.2)

the divergence-free and the no-slip boundary conditions

ddr(rvr,n)+invθ,n=0,vr,n(1)=vθ,n(1)=0,\displaystyle\frac{\,{\rm d}}{\,{\rm d}r}(rv_{r,n})+inv_{\theta,n}=0,\qquad v_{r,n}(1)=v_{\theta,n}(1)=0, (3.3)

and the condition at infinity

|vr,n(r)|+|vθ,n(r)|0,r.\displaystyle|v_{r,n}(r)|+|v_{\theta,n}(r)|\to 0,\quad r\to\infty. (3.4)

3.1 Zero mode

Proposition 3.1

For α\alpha\in\mathbb{R}, γ>2\gamma>2, 2<ργ2<\rho\leq\gamma and f=f0𝒫0L2ρ1(Ω)2f=f_{0}\in\mathcal{P}_{0}L^{\infty}_{2\rho-1}(\Omega)^{2}, there is a unique solution (v0,𝒫0q)(v_{0},\nabla\mathcal{P}_{0}q) of (S~\widetilde{\mbox{S}}) with v0𝒫0Lσ2(Ω)W01,2(Ω)2W2,2(Ω)2v_{0}\in\mathcal{P}_{0}L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap W^{2,2}(\Omega)^{2} and 𝒫0q𝒫0W^1,2(Ω)\mathcal{P}_{0}q\in\mathcal{P}_{0}\widehat{W}^{1,2}(\Omega) satisfying v0(r,θ)=vθ,0(r)𝐞θv_{0}(r,\theta)=v_{\theta,0}(r){\bf e}_{\theta} and

rρ|ω0(r)|\displaystyle r^{\rho}|\omega_{0}(r)| Cρ2f0L2ρ1,\displaystyle\leq\frac{C}{\rho-2}\|f_{0}\|_{L^{\infty}_{2\rho-1}}, (3.5)
rρ1|v0(r,θ)|+1γ1rρ|v0(r,θ)|\displaystyle r^{\rho-1}|v_{0}(r,\theta)|+\frac{1}{\gamma-1}r^{\rho}|\nabla v_{0}(r,\theta)| C(γ2)(ρ2)f0L2ρ1.\displaystyle\leq\frac{C}{(\gamma-2)(\rho-2)}\|f_{0}\|_{L^{\infty}_{2\rho-1}}. (3.6)

Here ω0:=(rotv0)0\omega_{0}:=({\rm rot}\,v_{0})_{0}. The constant CC is independent of α\alpha, γ\gamma and ρ\rho.

Proof.

Let n=0n=0 in (LABEL:eq.polar.vr)–(3.4). The first condition in (3.3) leads to that vr,0(r)=Crv_{r,0}(r)=\frac{C}{r} with some constant CC. Then the second condition leads to C=0C=0, which yields that vr,0=0v_{r,0}=0.

Thus we focus on the angular part vθ,0=vθ,0(r)v_{\theta,0}=v_{\theta,0}(r). From (LABEL:eq.polar.vtheta) and (3.3), we find that

d2vθ,0dr21+γrdvθ,0dr+1γr2vθ,0=fθ,0,r>1,vθ,0(1)=0.\displaystyle-\frac{\,{\rm d}^{2}v_{\theta,0}}{\,{\rm d}r^{2}}-\frac{1+\gamma}{r}\frac{\,{\rm d}v_{\theta,0}}{\,{\rm d}r}+\frac{1-\gamma}{r^{2}}v_{\theta,0}=f_{\theta,0},\quad r>1,\qquad v_{\theta,0}(1)=0. (3.7)

The linearly independent solutions of the homogeneous equation of (3.7) are

rγ+1andr1,\displaystyle r^{-\gamma+1}\quad\text{and}\quad r^{-1},

and their Wronskian is (γ2)rγ1(\gamma-2)r^{-\gamma-1}. Hence the solution of (3.7) in L2(Ω)L^{2}(\Omega) is given by

vθ,0(r)=1γ2{(1s2fθ,0(s)ds)rγ+1+rγ+11rsγfθ,0(s)ds+r1rs2fθ,0(s)ds}.\begin{split}v_{\theta,0}(r)&=\frac{1}{\gamma-2}\bigg{\{}-\Big{(}\int_{1}^{\infty}s^{2}f_{\theta,0}(s)\,{\rm d}s\Big{)}r^{-\gamma+1}\\ &\qquad\qquad\qquad+r^{-\gamma+1}\int_{1}^{r}s^{\gamma}f_{\theta,0}(s)\,{\rm d}s+r^{-1}\int_{r}^{\infty}s^{2}f_{\theta,0}(s)\,{\rm d}s\bigg{\}}.\end{split}

Then we have

ω0(r)=1rddr(rvθ,0)(r)=(1s2fθ,0(s)ds)rγrγ1rsγfθ,0(s)ds.\displaystyle\begin{split}\omega_{0}(r)=\frac{1}{r}\frac{\,{\rm d}}{\,{\rm d}r}(rv_{\theta,0})(r)=\Big{(}\int_{1}^{\infty}s^{2}f_{\theta,0}(s)\,{\rm d}s\Big{)}r^{-\gamma}-r^{-\gamma}\int_{1}^{r}s^{\gamma}f_{\theta,0}(s)\,{\rm d}s.\end{split}

By the assumption f𝒫0L2ρ1(Ω)2f\in\mathcal{P}_{0}L^{\infty}_{2\rho-1}(\Omega)^{2} with 2<ργ2<\rho\leq\gamma and the computations

rγ+11rsγ2ρ+1ds1ρ2rρ+1,r1rs2ρ+3ds1ρ2rρ+1,\displaystyle r^{-\gamma+1}\int_{1}^{r}s^{\gamma-2\rho+1}\,{\rm d}s\leq\frac{1}{\rho-2}r^{-\rho+1},\qquad r^{-1}\int_{r}^{\infty}s^{-2\rho+3}\,{\rm d}s\leq\frac{1}{\rho-2}r^{-\rho+1},

one can check that ω0\omega_{0} satisfies (3.5) and that v0=vθ,0𝐞θv_{0}=v_{\theta,0}{\bf e}_{\theta} belongs to 𝒫0Lσ2(Ω)W01,2(Ω)2W2,2(Ω)2\mathcal{P}_{0}L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap W^{2,2}(\Omega)^{2} and satisfies (3.6). The pressure 𝒫0qW^1,2(Ω)\mathcal{P}_{0}q\in\widehat{W}^{1,2}(\Omega) is obtained by (LABEL:eq.polar.vr). Clearly, (v0,𝒫0q)(v_{0},\nabla\mathcal{P}_{0}q) is the unique solution of (LABEL:eq.polar.vr)–(3.4). The proof is complete. ∎

3.2 Non-zero modes

By Proposition 3.1, the zero mode of the solution of (S~\widetilde{\mbox{S}}) decays as fast as desired, if the external force does correspondingly. On the other hand, for the non-zero mode, the decay rate is governed by the Biot-Savart law (2.6). Taking this into account, we will build a solution vnv_{n} of the non-zero mode of (S~\widetilde{\mbox{S}}) satisfying, for 2<ρmin{γ,3}2<\rho\leq\min\{\gamma,3\},

rotvn=(|x|ρ),vn=O(|x|ρ+1),\displaystyle{\rm rot}\,v_{n}=(|x|^{-\rho}),\qquad v_{n}=O(|x|^{-\rho+1}), (3.8)

under suitable assumptions on the external force fnf_{n}.

For |n|1|n|\geq 1, we define

nγ={n2+(γ2)2}12,ζn=(nγ2+iαn)12.\displaystyle n_{\gamma}=\Big{\{}n^{2}+\Big{(}\frac{\gamma}{2}\Big{)}^{2}\Big{\}}^{\frac{1}{2}},\qquad\zeta_{n}=(n_{\gamma}^{2}+i\alpha n)^{\frac{1}{2}}. (3.9)

This definition of ζn\zeta_{n} coincides with the one in [9] if γ=0\gamma=0. We compute

|ζn|=nγ{1+(αnnγ2)2}14,(ζn)=nγ2[{1+(αnnγ2)2}12+1]12,(ζn)=sgn(αn)nγ2[{1+(αnnγ2)2}121]12.\displaystyle\begin{split}|\zeta_{n}|&=n_{\gamma}\Big{\{}1+\Big{(}\frac{\alpha n}{n_{\gamma}^{2}}\Big{)}^{2}\Big{\}}^{\frac{1}{4}},\\ \Re(\zeta_{n})&=\frac{n_{\gamma}}{\sqrt{2}}\bigg{[}\Big{\{}1+\Big{(}\frac{\alpha n}{n_{\gamma}^{2}}\Big{)}^{2}\Big{\}}^{\frac{1}{2}}+1\bigg{]}^{\frac{1}{2}},\\ \Im(\zeta_{n})&={\rm sgn}(\alpha n)\frac{n_{\gamma}}{\sqrt{2}}\bigg{[}\Big{\{}1+\Big{(}\frac{\alpha n}{n_{\gamma}^{2}}\Big{)}^{2}\Big{\}}^{\frac{1}{2}}-1\bigg{]}^{\frac{1}{2}}.\end{split} (3.10)

Let us set ξn=(ζn)\xi_{n}=\Re(\zeta_{n}) for simplicity. Then we have

ξn|ζn|2ξn,ξn|n|(|α|12+γ),0<(ξnγ2)1<2γ\displaystyle\xi_{n}\leq|\zeta_{n}|\leq\sqrt{2}\xi_{n},\qquad\frac{\xi_{n}}{|n|}\leq(|\alpha|^{\frac{1}{2}}+\gamma),\qquad 0<\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}<2\gamma (3.11)

with CC independent of nn, α\alpha and γ\gamma.

Proposition 3.2

For |n|1|n|\geq 1, α\alpha\in\mathbb{R}, γ>2\gamma>2, 2<ρ32<\rho\leq 3 with 2<ρmin{γ,3}2<\rho\leq\min\{\gamma,3\} and f=fn𝒫nL2ρ1(Ω)2f=f_{n}\in\mathcal{P}_{n}L^{\infty}_{2\rho-1}(\Omega)^{2}, there is a unique solution (vn,𝒫nq)(v_{n},\nabla\mathcal{P}_{n}q) of (S~\widetilde{\mbox{S}}) with vn𝒫nLσ2(Ω)W01,2(Ω)2Wloc2,2(Ω¯)2v_{n}\in\mathcal{P}_{n}L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap W^{2,2}_{{\rm loc}}(\overline{\Omega})^{2} and 𝒫nq𝒫nWloc1,2(Ω¯)\mathcal{P}_{n}q\in\mathcal{P}_{n}W^{1,2}_{{\rm loc}}(\overline{\Omega}) satisfying the following.

  1. (1)

    For |n|1|n|\geq 1,

    rρ|ωn(r)|C|n|(ξn+γ2)(ξnγ2)1fnL2ρ1.\displaystyle r^{\rho}|\omega_{n}(r)|\leq\frac{C}{|n|}\Big{(}\xi_{n}+\frac{\gamma}{2}\Big{)}\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\|f_{n}\|_{L^{\infty}_{2\rho-1}}. (3.12)
  2. (2)

    If 2<ργ<32<\rho\leq\gamma<3 or 2<ρ<3γ2<\rho<3\leq\gamma, for |n|1|n|\geq 1,

    rρ1|vn(r,θ)|+rρ|n||vn(r,θ)|C|n|(|n|ρ+2)(ξn+γ2)(ξnγ2)1fnL2ρ1.\displaystyle\begin{split}&r^{\rho-1}|v_{n}(r,\theta)|+\frac{r^{\rho}}{|n|}|\nabla v_{n}(r,\theta)|\\ &\leq\frac{C}{|n|(|n|-\rho+2)}\Big{(}\xi_{n}+\frac{\gamma}{2}\Big{)}\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\|f_{n}\|_{L^{\infty}_{2\rho-1}}.\end{split} (3.13)
  3. (3)

    If γ3\gamma\geq 3 and ρ=3\rho=3, for |n|=1|n|=1,

    r2(logr)1|vn(r,θ)|+r3(logr)1|vn(r,θ)|C(ξn+γ2)(ξnγ2)1fnL5,\displaystyle\begin{split}&r^{2}(\log r)^{-1}|v_{n}(r,\theta)|+r^{3}(\log r)^{-1}|\nabla v_{n}(r,\theta)|\\ &\leq C\Big{(}\xi_{n}+\frac{\gamma}{2}\Big{)}\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\|f_{n}\|_{L^{\infty}_{5}},\end{split} (3.14)

    and for |n|>1|n|>1,

    r2|vn(r,θ)|+r3|n||vn(r,θ)|C|n|2(ξn+γ2)(ξnγ2)1fnL5.\displaystyle\begin{split}r^{2}|v_{n}(r,\theta)|+\frac{r^{3}}{|n|}|\nabla v_{n}(r,\theta)|\leq\frac{C}{|n|^{2}}\Big{(}\xi_{n}+\frac{\gamma}{2}\Big{)}\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\|f_{n}\|_{L^{\infty}_{5}}.\end{split} (3.15)

Here ωn:=(rotvn)n\omega_{n}:=({\rm rot}\,v_{n})_{n}. The constant CC is independent of nn, α\alpha, γ\gamma and ρ\rho.

Proof.

The proof is divided into two parts. First we show the existence of solutions using the representation formula. Second we verify the uniqueness by Proposition 2.1.

(Existence) Initially, let us assume that fn𝒫nC0(Ω)2f_{n}\in\mathcal{P}_{n}C^{\infty}_{0}(\Omega)^{2}. Operating rot{\rm rot}\, and 𝒫n\mathcal{P}_{n} to the first line of (S~\widetilde{\mbox{S}}), we see that ωn=ωn(r)\omega_{n}=\omega_{n}(r) satisfies the ordinary differential equation

d2ωndr21+γrdωndr+n2+iαnr2ωn=(rotfn)n,r>1.\displaystyle-\frac{\,{\rm d}^{2}\omega_{n}}{\,{\rm d}r^{2}}-\frac{1+\gamma}{r}\frac{\,{\rm d}\omega_{n}}{\,{\rm d}r}+\frac{n^{2}+i\alpha n}{r^{2}}\omega_{n}=({\rm rot}\,f_{n})_{n},\quad r>1. (3.16)

By the transformation

ωn(r)=rγ2ω~n(r),\displaystyle\omega_{n}(r)=r^{-\frac{\gamma}{2}}\tilde{\omega}_{n}(r),

we find that ω~n\tilde{\omega}_{n} solves

d2ω~ndr21rdω~ndr+ζn2r2ω~n=rγ2(rotfn)n,r>1.\displaystyle-\frac{\,{\rm d}^{2}\tilde{\omega}_{n}}{\,{\rm d}r^{2}}-\frac{1}{r}\frac{\,{\rm d}\tilde{\omega}_{n}}{\,{\rm d}r}+\frac{\zeta_{n}^{2}}{r^{2}}\tilde{\omega}_{n}=r^{\frac{\gamma}{2}}({\rm rot}\,f_{n})_{n},\quad r>1. (3.17)

The linearly independent solutions of the homogeneous equation of (3.17) are

rζnandrζn,\displaystyle r^{-\zeta_{n}}\quad\text{and}\quad r^{\zeta_{n}},

and their Wronskian is 2ζnr12\zeta_{n}r^{-1}. Hence the decaying solution of (3.17) is given by

ω~n(r)\displaystyle\tilde{\omega}_{n}(r) =cn[fn]rζn+Ψn[fn](r),\displaystyle=c_{n}[f_{n}]r^{-\zeta_{n}}+\Psi_{n}[f_{n}](r), (3.18)

where the constant cn[fn]c_{n}[f_{n}] is to be determined later and Ψn[fn]\Psi_{n}[f_{n}] is defined by

Ψn[fn](r)=rζnζn1rsζn+γ2+1(rotfn)n(s)ds+rζnζnrsζn+γ2+1(rotfn)n(s)ds=rζnζn1rsζn+γ2{(ζn+γ2)fθ,n(s)+infr,n(s)}ds+rζnζnrsζn+γ2{(ζnγ2)fθ,n(s)infr,n(s)}ds.\displaystyle\begin{split}\Psi_{n}[f_{n}](r)&=\frac{r^{-\zeta_{n}}}{\zeta_{n}}\int_{1}^{r}s^{\zeta_{n}+\frac{\gamma}{2}+1}({\rm rot}\,f_{n})_{n}(s)\,{\rm d}s\\ &\quad+\frac{r^{\zeta_{n}}}{\zeta_{n}}\int_{r}^{\infty}s^{-\zeta_{n}+\frac{\gamma}{2}+1}({\rm rot}\,f_{n})_{n}(s)\,{\rm d}s\\ &=-\frac{r^{-\zeta_{n}}}{\zeta_{n}}\int_{1}^{r}s^{\zeta_{n}+\frac{\gamma}{2}}\Big{\{}\Big{(}\zeta_{n}+\frac{\gamma}{2}\Big{)}f_{\theta,n}(s)+inf_{r,n}(s)\Big{\}}\,{\rm d}s\\ &\quad+\frac{r^{\zeta_{n}}}{\zeta_{n}}\int_{r}^{\infty}s^{-\zeta_{n}+\frac{\gamma}{2}}\Big{\{}\Big{(}\zeta_{n}-\frac{\gamma}{2}\Big{)}f_{\theta,n}(s)-inf_{r,n}(s)\Big{\}}\,{\rm d}s.\end{split} (3.19)

Here we performed integration by parts using (rotfn)n(r)=1rddr(rfθ,n)(r)inrfr,n(r)({\rm rot}\,f_{n})_{n}(r)=\frac{1}{r}\frac{\,{\rm d}}{\,{\rm d}r}(rf_{\theta,n})(r)-\frac{in}{r}f_{r,n}(r). Going back to the equation (3.16), we see that the decaying solution is given by

ωn(r)=cn[fn]rζnγ2+Φn[fn](r),Φn[fn](r):=rγ2Ψn[fn](r).\displaystyle\omega_{n}(r)=c_{n}[f_{n}]r^{-\zeta_{n}-\frac{\gamma}{2}}+\Phi_{n}[f_{n}](r),\qquad\Phi_{n}[f_{n}](r):=r^{-\frac{\gamma}{2}}\Psi_{n}[f_{n}](r). (3.20)

Let us determine cn[fn]c_{n}[f_{n}] in (3.18). From fn𝒫nC0(Ω)2f_{n}\in\mathcal{P}_{n}C^{\infty}_{0}(\Omega)^{2}, we have |Φn[fn](r)|C(f,n)rξnγ2|\Phi_{n}[f_{n}](r)|\leq C(f,n)r^{-\xi_{n}-\frac{\gamma}{2}}. We choose cn[fn]c_{n}[f_{n}] so that dn[ωn]d_{n}[\omega_{n}] in (2.5) is zero, namely,

cn[fn]=(ζn+|n|+γ22)1s|n|+1Φn[fn](s)ds.\displaystyle\begin{split}c_{n}[f_{n}]=-\Big{(}\zeta_{n}+|n|+\frac{\gamma}{2}-2\Big{)}\int_{1}^{\infty}s^{-|n|+1}\Phi_{n}[f_{n}](s)\,{\rm d}s.\end{split} (3.21)

Then the Biot-Savart law Vn[ωn]=:vnV_{n}[\omega_{n}]=:v_{n} in (2.6) is written as

vn(r,θ)=vr,n(r)einθ𝐞r+vr,n(r)einθ𝐞θ,vr,n(r)=in2|n|(r|n|11rs|n|+1ωn(s)ds+r|n|1rs|n|+1ωn(s)ds),vθ,n(r)=12(r|n|11rs|n|+1ωn(s)dsr|n|1rs|n|+1ωn(s)ds).\displaystyle\begin{split}v_{n}(r,\theta)&=v_{r,n}(r)e^{in\theta}{\bf e}_{r}+v_{r,n}(r)e^{in\theta}{\bf e}_{\theta},\\ v_{r,n}(r)&=\frac{in}{2|n|}\Big{(}r^{-|n|-1}\int_{1}^{r}s^{|n|+1}\omega_{n}(s)\,{\rm d}s+r^{|n|-1}\int_{r}^{\infty}s^{-|n|+1}\omega_{n}(s)\,{\rm d}s\Big{)},\\ v_{\theta,n}(r)&=\frac{1}{2}\Big{(}r^{-|n|-1}\int_{1}^{r}s^{|n|+1}\omega_{n}(s)\,{\rm d}s-r^{|n|-1}\int_{r}^{\infty}s^{-|n|+1}\omega_{n}(s)\,{\rm d}s\Big{)}.\end{split} (3.22)

Let us show that, for vnv_{n} in (3.22), there is a pressure 𝒫nq𝒫nW^1,2(Ω)\mathcal{P}_{n}q\in\mathcal{P}_{n}\widehat{W}^{1,2}(\Omega) such that the pair (vn,𝒫nq)(v_{n},\nabla\mathcal{P}_{n}q) is a solution of (S~\widetilde{\mbox{S}}). From fn𝒫nC0(Ω)2f_{n}\in\mathcal{P}_{n}C^{\infty}_{0}(\Omega)^{2} and ξn>γ2\xi_{n}>\frac{\gamma}{2} implied by (3.11), we see that ωn\omega_{n} in (3.20) is smooth and satisfies |kωn(r)|C(f,n,k)rmin{γ,3}|\nabla^{k}\omega_{n}(r)|\leq C(f,n,k)r^{-\min\{\gamma,3\}} for any k0k\in\mathbb{Z}_{\geq 0}. This, combined with the choice of cn[fn]c_{n}[f_{n}] in (3.21) and Lemma A.1, ensures that vnv_{n} is smooth and satisfies vn|Ω=0v_{n}|_{\partial\Omega}=0 and kvnL2<\|\nabla^{k}v_{n}\|_{L^{2}}<\infty for any k0k\in\mathbb{Z}_{\geq 0}. Accordingly, we can apply Proposition 2.2 because of (rotvn)n=ωn({\rm rot}\,v_{n})_{n}=\omega_{n} and

(rot(Δvn+(αUγW)rotvnfn))n=d2ωndr21+γrdωndr+n2+iαnr2ωn(rotfn)n=0.\displaystyle\begin{split}&\big{(}{\rm rot}\,(-\Delta v_{n}+(\alpha U-\gamma W)^{\bot}{\rm rot}\,v_{n}-f_{n})\big{)}_{n}\\ &=-\frac{\,{\rm d}^{2}\omega_{n}}{\,{\rm d}r^{2}}-\frac{1+\gamma}{r}\frac{\,{\rm d}\omega_{n}}{\,{\rm d}r}+\frac{n^{2}+i\alpha n}{r^{2}}\omega_{n}-({\rm rot}\,f_{n})_{n}=0.\end{split}

Hence there is a pressure 𝒫nq𝒫nW^1,2(Ω)\mathcal{P}_{n}q\in\mathcal{P}_{n}\widehat{W}^{1,2}(\Omega) such that (vn,𝒫nq)(v_{n},\nabla\mathcal{P}_{n}q) is a solution of (S~\widetilde{\mbox{S}}).

Now, let fn𝒫nL2ρ1(Ω)2f_{n}\in\mathcal{P}_{n}L^{\infty}_{2\rho-1}(\Omega)^{2} in (3.19)–(3.22). We will prove the estimates (3.12)–(3.15). For Φn[fn]\Phi_{n}[f_{n}] in (3.20) and cn[fn]c_{n}[f_{n}] in (3.21), using Lemma A.1, we have

Φn[fn]Lρ+|cn[fn]|C(1+|ζn|+|n|+γ22|n|+ρ2)Φn[fn]LρC|n|(ξn+γ2)(ξnγ2)1fnL2ρ1,\displaystyle\begin{split}\|\Phi_{n}[f_{n}]\|_{L^{\infty}_{\rho}}+|c_{n}[f_{n}]|&\leq C\bigg{(}1+\frac{|\zeta_{n}|+|n|+\dfrac{\gamma}{2}-2}{|n|+\rho-2}\bigg{)}\|\Phi_{n}[f_{n}]\|_{L^{\infty}_{\rho}}\\ &\leq\frac{C}{|n|}\Big{(}\xi_{n}+\frac{\gamma}{2}\Big{)}\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\|f_{n}\|_{L^{\infty}_{2\rho-1}},\end{split}

where (3.11) is used in the second inequality and CC is independent of nn, α\alpha, γ\gamma and ρ\rho. Hence we see from (3.20) that ωn\omega_{n} satisfies (3.12). For vnv_{n} in (3.22), using (2.1) and (3.12), we have

|vn(r,θ)|+r|n||vn(r,θ)|C|n|(ξn+γ2)(ξnγ2)1fnL2ρ1×(r|n|11rs|n|ρ+1ds+r|n|1rs|n|ρ+1ds).\displaystyle\begin{split}&|v_{n}(r,\theta)|+\frac{r}{|n|}|\nabla v_{n}(r,\theta)|\\ &\leq\frac{C}{|n|}\Big{(}\xi_{n}+\frac{\gamma}{2}\Big{)}\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\|f_{n}\|_{L^{\infty}_{2\rho-1}}\\ &\qquad\times\bigg{(}r^{-|n|-1}\int_{1}^{r}s^{|n|-\rho+1}\,{\rm d}s+r^{|n|-1}\int_{r}^{\infty}s^{-|n|-\rho+1}\,{\rm d}s\bigg{)}.\end{split}

Hence we see from Lemma A.1 (2) that vnv_{n} satisfies (3.13)–(3.15).

It remains to verify that vnv_{n} given by (3.22) is a solution of (S~\widetilde{\mbox{S}}) for general fn𝒫nL2ρ1(Ω)2f_{n}\in\mathcal{P}_{n}L^{\infty}_{2\rho-1}(\Omega)^{2}. The condition vnW01,2(Ω)2v_{n}\in W^{1,2}_{0}(\Omega)^{2} follows from (3.13)–(3.15) and the choice of cn[fn]c_{n}[f_{n}] in (3.21). By Lemma A.2 and direct computation, one has

Φn[fn]L2+|cn[fn]|+ωnL2+vnL2C(n,α,γ)μfnL2,μ(x):=|x|2.\displaystyle\begin{split}&\|\Phi_{n}[f_{n}]\|_{L^{\infty}_{2}}+|c_{n}[f_{n}]|+\|\omega_{n}\|_{L^{\infty}_{2}}+\|\nabla v_{n}\|_{L^{2}}\\ &\leq C(n,\alpha,\gamma)\|\mu f_{n}\|_{L^{2}},\qquad\mu(x):=|x|^{2}.\end{split} (3.23)

To show that vnv_{n} satisfies the weak formulation of (S~\widetilde{\mbox{S}}), we take {ψn(m)}m=1𝒫nC0(Ω)2\{\psi_{n}^{(m)}\}_{m=1}^{\infty}\subset\mathcal{P}_{n}C^{\infty}_{0}(\Omega)^{2} such that limmμ(fnψn(m))L2=0\displaystyle{\lim_{m\to\infty}\|\mu(f_{n}-\psi_{n}^{(m)})\|_{L^{2}}}=0. Let vn(m)v_{n}^{(m)} denote the smooth solution given by (3.22) replacing fnf_{n} by ψn(m)\psi_{n}^{(m)}. Also, let 𝒫nqn(m)𝒫nW^1,2(Ω)\mathcal{P}_{n}q_{n}^{(m)}\in\mathcal{P}_{n}\widehat{W}^{1,2}(\Omega) be an associated pressure. Then, using (3.23) and linearity of the equations, we have, for any φC0,σ(Ω)\varphi\in C^{\infty}_{0,\sigma}(\Omega),

vn,φ+(αUγW)rotvnfn,φ=limmΔvn(m)+(αUγW)rotvn(m)ψn(m),φ=limm𝒫nqn(m),φ=0.\displaystyle\begin{split}&\langle\nabla v_{n},\nabla\varphi\rangle+\langle(\alpha U-\gamma W)^{\bot}{\rm rot}\,v_{n}-f_{n},\varphi\rangle\\ &=\lim_{m\to\infty}\langle-\Delta v_{n}^{(m)}+(\alpha U-\gamma W)^{\bot}{\rm rot}\,v_{n}^{(m)}-\psi_{n}^{(m)},\varphi\rangle\\ &=\lim_{m\to\infty}\langle\nabla\mathcal{P}_{n}q_{n}^{(m)},\varphi\rangle=0.\end{split}

Here we performed integration by parts in the first and last equalities. Consequently, we see that vnv_{n} is a weak solution of (S~\widetilde{\mbox{S}}). Then, by regarding (αUγW)rotvnfn(\alpha U-\gamma W)^{\bot}{\rm rot}\,v_{n}-f_{n} as the external force, one can prove the local regularity vnWloc2,2(Ω¯)2v_{n}\in W^{2,2}_{{\rm loc}}(\overline{\Omega})^{2} and the existence of an associated pressure 𝒫nq𝒫nWloc1,2(Ω¯)\mathcal{P}_{n}q\in\mathcal{P}_{n}W^{1,2}_{{\rm loc}}(\overline{\Omega}) by standard theory for the Stokes system; see Sohr [21, Chapter III] for example. This completes the existence part of the proof.

(Uniqueness) Assume that (vn,𝒫nq)(𝒫nLσ2(Ω)W01,2(Ω)2Wloc2,2(Ω¯)2)×𝒫nLloc2(Ω¯)2(v_{n},\nabla\mathcal{P}_{n}q)\in\big{(}\mathcal{P}_{n}L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap W^{2,2}_{{\rm loc}}(\overline{\Omega})^{2}\big{)}\times\mathcal{P}_{n}L^{2}_{{\rm loc}}(\overline{\Omega})^{2} is a solution of (S~\widetilde{\mbox{S}}) with f=0f=0. By the elliptic regularity, vnv_{n} is smooth in Ω\Omega. Since ωn=(rotvn)n\omega_{n}=({\rm rot}\,v_{n})_{n} satisfies the homogeneous equation of (3.16), due to the summability vnL2(Ω)2×2\nabla v_{n}\in L^{2}(\Omega)^{2\times 2}, we have ωn=c~nrζnγ2\omega_{n}=\tilde{c}_{n}r^{-\zeta_{n}-\frac{\gamma}{2}} with some constant c~n\tilde{c}_{n}. Then Proposition 2.1 implies that vn=Vn[ωn]v_{n}=V_{n}[\omega_{n}] and dn[ωn]=0d_{n}[\omega_{n}]=0. The latter condition is rewritten as

0=dn[ωn]=c~n1sζn|n|γ2+1ds=c~nζn+|n|+γ22.\displaystyle 0=d_{n}[\omega_{n}]=\tilde{c}_{n}\int_{1}^{\infty}s^{-\zeta_{n}-|n|-\frac{\gamma}{2}+1}\,{\rm d}s=\frac{\tilde{c}_{n}}{\zeta_{n}+|n|+\dfrac{\gamma}{2}-2}.

Thus we have c~n=0\tilde{c}_{n}=0 and hence ωn=0\omega_{n}=0. Then the uniqueness follows from vn=Vn[ωn]=0v_{n}=V_{n}[\omega_{n}]=0. This completes the uniqueness part of the proof. Hence we conclude. ∎

4 Nonlinear problem

In this section we prove Theorem 1.1 by showing the unique solvability of the system

{Δv+(αUγW)rotv+q=vrotv+finΩdivv=0inΩv=0onΩv(x)0as|x|.\left\{\begin{array}[]{ll}-\Delta v+(\alpha U-\gamma W)^{\bot}{\rm rot}\,v+\nabla q=-v^{\bot}{\rm rot}\,v+f&\mbox{in}\ \Omega\\ {\rm div}\,v=0&\mbox{in}\ \Omega\\ v=0&\mbox{on}\ \partial\Omega\\ v(x)\to 0&\mbox{as}\ |x|\to\infty.\end{array}\right. (NS~\widetilde{\mbox{NS}})

Let us collect two lemmas needed in the proof. For ρ0\rho\geq 0, we define the Banach space

l1(Lρ(Ω))\displaystyle l^{1}\big{(}L^{\infty}_{\rho}(\Omega)\big{)} ={f=n𝒫nf|fl1Lρ:=n𝒫nfLρ<}.\displaystyle=\bigg{\{}f=\sum_{n\in\mathbb{Z}}\mathcal{P}_{n}f~{}\bigg{|}~{}\|f\|_{l^{1}L^{\infty}_{\rho}}:=\sum_{n\in\mathbb{Z}}\|\mathcal{P}_{n}f\|_{L^{\infty}_{\rho}}<\infty\bigg{\}}.
Lemma 4.1

For α\alpha\in\mathbb{R}, γ>2\gamma>2, 2<ρ<32<\rho<3 with 2<ρmin{γ,3}2<\rho\leq\min\{\gamma,3\} and fl1(L2ρ1(Ω))2f\in l^{1}\big{(}L^{\infty}_{2\rho-1}(\Omega)\big{)}^{2}, there is a unique solution (v,q)(v,\nabla q) of (S~\widetilde{\mbox{S}}) with vLσ2(Ω)W01,2(Ω)2Wloc2,2(Ω¯)2v\in L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap W^{2,2}_{{\rm loc}}(\overline{\Omega})^{2} and qWloc1,2(Ω¯)q\in W^{1,2}_{{\rm loc}}(\overline{\Omega}) satisfying

vl1Lρ1+vl1Lρκfl1L2ρ1,κ=κ(α,γ,ρ):=C0(|α|12+γ)γ(ρ2)2(3ρ).\displaystyle\begin{split}\|v\|_{l^{1}L^{\infty}_{\rho-1}}+\|\nabla v\|_{l^{1}L^{\infty}_{\rho}}\leq\kappa\|f\|_{l^{1}L^{\infty}_{2\rho-1}},\qquad\kappa=\kappa(\alpha,\gamma,\rho):=\frac{C_{0}(|\alpha|^{\frac{1}{2}}+\gamma)\gamma}{(\rho-2)^{2}(3-\rho)}.\end{split} (4.1)

The constant C0C_{0} is independent of α\alpha, γ\gamma and ρ\rho.

Proof.

This follows from Propositions 3.1 and 3.2 and an estimate obtained from (3.11):

1|n|(ξn+γ2)(ξnγ2)14(|α|12+γ)γ.\displaystyle\begin{split}\frac{1}{|n|}\Big{(}\xi_{n}+\frac{\gamma}{2}\Big{)}\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\leq 4(|\alpha|^{\frac{1}{2}}+\gamma)\gamma.\end{split}

Indeed, we see from the propositions that there is a weak solution vLσ2(Ω)W01,2(Ω)2v\in L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2} of (S~\widetilde{\mbox{S}}) satisfying (4.1). Then, as in the proof of Proposition 3.2, we have the local regularity vWloc2,2(Ω¯)2v\in W^{2,2}_{{\rm loc}}(\overline{\Omega})^{2} and an associate pressure qWloc1,2(Ω¯)q\in W^{1,2}_{{\rm loc}}(\overline{\Omega}). The uniqueness of (v,q)(v,\nabla q) follows from that of (vn,𝒫nq)(v_{n},\nabla\mathcal{P}_{n}q) for nn\in\mathbb{Z}. This completes the proof. ∎

Lemma 4.2

For vl1(Lγ1(Ω))2v\in l^{1}\big{(}L^{\infty}_{\gamma_{1}}(\Omega)\big{)}^{2} and ωl1(Lγ2(Ω))\omega\in l^{1}\big{(}L^{\infty}_{\gamma_{2}}(\Omega)\big{)}, we have

vωl1Lγ1+γ2vl1Lγ1ωl1Lγ2.\displaystyle\begin{split}\|v\omega\|_{l^{1}L^{\infty}_{\gamma_{1}+\gamma_{2}}}\leq\|v\|_{l^{1}L^{\infty}_{\gamma_{1}}}\|\omega\|_{l^{1}L^{\infty}_{\gamma_{2}}}.\end{split}
Proof.

The identity

𝒫n(vω)\displaystyle\mathcal{P}_{n}(v\omega) =m(𝒫mv)(𝒫nmω),n\displaystyle=\sum_{m\in\mathbb{Z}}(\mathcal{P}_{m}v)(\mathcal{P}_{n-m}\omega),\quad n\in\mathbb{Z}

and the Young inequality for sequences imply

vωl1Lγ1+γ2\displaystyle\|v\omega\|_{l^{1}L^{\infty}_{\gamma_{1}+\gamma_{2}}} nm𝒫mvLγ1𝒫nmωLγ2\displaystyle\leq\sum_{n\in\mathbb{Z}}\sum_{m\in\mathbb{Z}}\|\mathcal{P}_{m}v\|_{L^{\infty}_{\gamma_{1}}}\|\mathcal{P}_{n-m}\omega\|_{L^{\infty}_{\gamma_{2}}}
vl1Lγ1ωl1Lγ2.\displaystyle\leq\|v\|_{l^{1}L^{\infty}_{\gamma_{1}}}\|\omega\|_{l^{1}L^{\infty}_{\gamma_{2}}}.

This completes the proof. ∎


Proof of Theorem 1.1: Firstly we show the unique solvability of (NS~\widetilde{\mbox{NS}}) under smallness conditions on ff. We define the Banach space

𝒳ρ={wLσ2(Ω)W01,2(Ω)2l1(Lρ1(Ω))2|wl1(Lρ(Ω))2×2},\displaystyle\mathcal{X}_{\rho}=\Big{\{}w\in L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap l^{1}\big{(}L^{\infty}_{\rho-1}(\Omega)\big{)}^{2}~{}\Big{|}~{}\nabla w\in l^{1}\big{(}L^{\infty}_{\rho}(\Omega)\big{)}^{2\times 2}\Big{\}},

equipped with the norm w𝒳ρ:=wl1Lρ1+wl1Lρ\|w\|_{\mathcal{X}_{\rho}}:=\|w\|_{l^{1}L^{\infty}_{\rho-1}}+\|\nabla w\|_{l^{1}L^{\infty}_{\rho}}, and consider the closed subset

ρ(δ)={w𝒳ρ|w𝒳ρδ},δ>0.\displaystyle\mathcal{B}_{\rho}(\delta)=\{w\in\mathcal{X}_{\rho}~{}|~{}\|w\|_{\mathcal{X}_{\rho}}\leq\delta\},\quad\delta>0.

For any w𝒳ρw\in\mathcal{X}_{\rho}, by Lemma 4.1, there is a unique solution (vw,qw)(v_{w},\nabla q_{w}) to the problem

{Δvw+(αUγW)rotvw+qw=wrotw+finΩdivvw=0inΩvw=0onΩvw(x)0as|x|\left\{\begin{array}[]{ll}-\Delta v_{w}+(\alpha U-\gamma W)^{\bot}{\rm rot}\,v_{w}+\nabla q_{w}=-w^{\bot}{\rm rot}\,w+f&\mbox{in}\ \Omega\\ {\rm div}\,v_{w}=0&\mbox{in}\ \Omega\\ v_{w}=0&\mbox{on}\ \partial\Omega\\ v_{w}(x)\to 0&\mbox{as}\ |x|\to\infty\end{array}\right.

satisfying

vw𝒳ρ=vwl1Lρ1+vwl1Lρκ(wrotwl1L2ρ1+fl1L2ρ1)κ(wl1Lρ1rotwl1Lρ+fl1L2ρ1),\displaystyle\begin{split}\|v_{w}\|_{\mathcal{X}_{\rho}}=\|v_{w}\|_{l^{1}L^{\infty}_{\rho-1}}+\|\nabla v_{w}\|_{l^{1}L^{\infty}_{\rho}}&\leq\kappa\big{(}\|w^{\bot}{\rm rot}\,w\|_{l^{1}L^{\infty}_{2\rho-1}}+\|f\|_{l^{1}L^{\infty}_{2\rho-1}}\big{)}\\ &\leq\kappa\big{(}\|w\|_{l^{1}L^{\infty}_{\rho-1}}\|{\rm rot}\,w\|_{l^{1}L^{\infty}_{\rho}}+\|f\|_{l^{1}L^{\infty}_{2\rho-1}}\big{)},\end{split} (4.2)

where Lemma 4.2 is applied in the second inequality. Hence, by denoting vwv_{w} by T(w)T(w), we see that TT defines a linear map from 𝒳ρ\mathcal{X}_{\rho} to itself.

Let us show that TT is a contraction on ρ(δ)𝒳ρ\mathcal{B}_{\rho}(\delta)\subset\mathcal{X}_{\rho} if both fl1L2ρ1\|f\|_{l^{1}L^{\infty}_{2\rho-1}} and δ\delta are sufficiently small depending on κ=κ(α,γ,ρ)\kappa=\kappa(\alpha,\gamma,\rho). Then the unique existence of solutions of (NS~\widetilde{\mbox{NS}}) follows from the Banach fixed-point theorem. For any wρ(δ)w\in\mathcal{B}_{\rho}(\delta), by (4.2), we have

T(w)𝒳ρκ(δ2+fl1L2ρ1).\displaystyle\begin{split}\|T(w)\|_{\mathcal{X}_{\rho}}&\leq\kappa\big{(}\delta^{2}+\|f\|_{l^{1}L^{\infty}_{2\rho-1}}\big{)}.\end{split} (4.3)

On the other hand, for any w1,w2ρ(δ)w_{1},w_{2}\in\mathcal{B}_{\rho}(\delta), by Lemma 4.2 again, we have

T(w2)T(w1)𝒳ρκw2rotw2w1rotw1l1L2ρ12κδw2w1𝒳ρ.\displaystyle\begin{split}\|T(w_{2})-T(w_{1})\|_{\mathcal{X}_{\rho}}&\leq\kappa\|w_{2}^{\bot}{\rm rot}\,w_{2}-w_{1}^{\bot}{\rm rot}\,w_{1}\|_{l^{1}L^{\infty}_{2\rho-1}}\\ &\leq 2\kappa\delta\|w_{2}-w_{1}\|_{\mathcal{X}_{\rho}}.\end{split} (4.4)

Therefore, fixing δ\delta and fl1L2ρ1\|f\|_{l^{1}L^{\infty}_{2\rho-1}} so that

δ<12κ~andfl1L2ρ1δ2,\displaystyle\begin{split}\delta<\frac{1}{2\widetilde{\kappa}}\quad\text{and}\quad\|f\|_{l^{1}L^{\infty}_{2\rho-1}}\leq\delta^{2},\end{split} (4.5)

we see that T:ρ(δ)ρ(δ)T:\mathcal{B}_{\rho}(\delta)\to\mathcal{B}_{\rho}(\delta) is contractive. Then the Banach fixed-point theorem yields that there is a unique element vρ(δ)v\in\mathcal{B}_{\rho}(\delta) such that T(v)=vT(v)=v. Thus we obtain a solution (v,q)(Lσ2(Ω)W01,2(Ω)2Wloc2,2(Ω¯)2)×Lloc2(Ω¯)2(v,\nabla q)\in\big{(}L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap W^{2,2}_{{\rm loc}}(\overline{\Omega})^{2}\big{)}\times L^{2}_{{\rm loc}}(\overline{\Omega})^{2} of (NS~\widetilde{\mbox{NS}}) with vv unique in ρ(δ)\mathcal{B}_{\rho}(\delta).

By the argument so far, for a given fl1(L2ρ1(Ω))2f\in l^{1}\big{(}L^{\infty}_{2\rho-1}(\Omega)\big{)}^{2} satisfying (4.5), the pair

u:=αUγW+v,p:=(|u|22+q)\displaystyle u:=\alpha U-\gamma W+v,\qquad\nabla p:=\nabla\Big{(}-\frac{|u|^{2}}{2}+q\Big{)}

is a solution of (NS) in (W^1,2(Ω)Wloc2,2(Ω¯)L1(Ω))2×Lloc2(Ω¯)2\big{(}\widehat{W}^{1,2}(\Omega)\cap W^{2,2}_{{\rm loc}}(\overline{\Omega})\cap L^{\infty}_{1}(\Omega)\big{)}^{2}\times L^{2}_{{\rm loc}}(\overline{\Omega})^{2} unique in the set

{(u,p)|u=αUγW+v,vρ(δ)}.\displaystyle\{(u,\nabla p)~{}|~{}u=\alpha U-\gamma W+v,\ \ v\in\mathcal{B}_{\rho}(\delta)\}.

The asymptotics (1.5) can be checked easily. The proof of Theorem 1.1 is complete.  \Box

Appendix A Integral estimates

We summarize the integral estimates used in the proof of Proposition 3.2. The proof only uses elementary calculations and thus we will just state the results. Recall that

nγ={n2+(γ2)2}12,ξn=nγ2[{1+(αnnγ2)2}12+1]12.\displaystyle n_{\gamma}=\Big{\{}n^{2}+\Big{(}\frac{\gamma}{2}\Big{)}^{2}\Big{\}}^{\frac{1}{2}},\qquad\xi_{n}=\frac{n_{\gamma}}{\sqrt{2}}\bigg{[}\Big{\{}1+\Big{(}\frac{\alpha n}{n_{\gamma}^{2}}\Big{)}^{2}\Big{\}}^{\frac{1}{2}}+1\bigg{]}^{\frac{1}{2}}.
Lemma A.1

For α\alpha\in\mathbb{R}, we have the following.

  1. (1)

    Let 2<ργ2<\rho\leq\gamma. For |n|1|n|\geq 1,

    rξnγ21rsξn+γ22ρ+1ds\displaystyle r^{-\xi_{n}-\frac{\gamma}{2}}\int_{1}^{r}s^{\xi_{n}+\frac{\gamma}{2}-2\rho+1}\,{\rm d}s min{1ρ2,(ξnγ2)1}rρ,\displaystyle\leq\min\Big{\{}\frac{1}{\rho-2},\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}\Big{\}}r^{-\rho},
    rξnγ2rsξn+γ22ρ+1ds\displaystyle r^{\xi_{n}-\frac{\gamma}{2}}\int_{r}^{\infty}s^{-\xi_{n}+\frac{\gamma}{2}-2\rho+1}\,{\rm d}s (ξnγ2)1rρ.\displaystyle\leq\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-1}r^{-\rho}.
  2. (2)

    Let 2<ρ32<\rho\leq 3.

    1. (i)

      If 2<ρ<32<\rho<3, for |n|1|n|\geq 1,

      r|n|11rs|n|ρ+1ds\displaystyle r^{-|n|-1}\int_{1}^{r}s^{|n|-\rho+1}\,{\rm d}s 1|n|ρ+2rρ+1.\displaystyle\leq\frac{1}{|n|-\rho+2}r^{-\rho+1}.
    2. (ii)

      If ρ=3\rho=3, for |n|=1|n|=1,

      r|n|11rs|n|ρ+1ds\displaystyle r^{-|n|-1}\int_{1}^{r}s^{|n|-\rho+1}\,{\rm d}s =r2logr,\displaystyle=r^{-2}\log r,

      and for |n|>1|n|>1,

      r|n|11rs|n|ρ+1ds\displaystyle r^{-|n|-1}\int_{1}^{r}s^{|n|-\rho+1}\,{\rm d}s 1|n|1r2.\displaystyle\leq\frac{1}{|n|-1}r^{-2}.
    3. (iii)

      For |n|1|n|\geq 1,

      r|n|1rs|n|ρ+1ds\displaystyle r^{|n|-1}\int_{r}^{\infty}s^{-|n|-\rho+1}\,{\rm d}s 1|n|+ρ2rρ+1.\displaystyle\leq\frac{1}{|n|+\rho-2}r^{-\rho+1}.
Lemma A.2

Under the assumption in Proposition 3.2,

rξnγ21rsξn+γ2|fn|(s)ds\displaystyle r^{-\xi_{n}-\frac{\gamma}{2}}\int_{1}^{r}s^{\xi_{n}+\frac{\gamma}{2}}|f_{n}|(s)\,{\rm d}s (ξnγ2)12(1rs4|fn|2(s)sds)12r2,\displaystyle\leq\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-\frac{1}{2}}\bigg{(}\int_{1}^{r}s^{4}|f_{n}|^{2}(s)s\,{\rm d}s\bigg{)}^{\frac{1}{2}}r^{-2},
rξnγ2rsξn+γ2|fn|(s)ds\displaystyle r^{\xi_{n}-\frac{\gamma}{2}}\int_{r}^{\infty}s^{-\xi_{n}+\frac{\gamma}{2}}|f_{n}|(s)\,{\rm d}s (ξnγ2)12(rs4|fn|2(s)sds)12r2.\displaystyle\leq\Big{(}\xi_{n}-\frac{\gamma}{2}\Big{)}^{-\frac{1}{2}}\bigg{(}\int_{r}^{\infty}s^{4}|f_{n}|^{2}(s)s\,{\rm d}s\bigg{)}^{\frac{1}{2}}r^{-2}.

Acknowledgements

The author is partially supported by JSPS KAKENHI Grant Number JP 20K14345.

References

  • [1] I-Dee Chang and Robert Finn. On the solutions of a class of equations occurring in continuum mechanics, with application to the Stokes paradox. Arch. Rational Mech. Anal., 7:388–401, 1961.
  • [2] Hiroshi Fujita. On the existence and regularity of the steady-state solutions of the Navier-Stokes theorem. J. Fac. Sci. Univ. Tokyo Sect. I, 9:59–102 (1961), 1961.
  • [3] Giovanni P. Galdi. Stationary Navier-Stokes problem in a two-dimensional exterior domain. In Stationary partial differential equations. Vol. I, Handb. Differ. Equ., pages 71–155. North-Holland, Amsterdam, 2004.
  • [4] Giovanni P. Galdi. An introduction to the mathematical theory of the Navier-Stokes equations. Springer Monographs in Mathematics. Springer, New York, second edition, 2011. Steady-state problems.
  • [5] Isabelle Gallagher, Mitsuo Higaki, and Yasunori Maekawa. On stationary two-dimensional flows around a fast rotating disk. Math. Nachr., 292(2):273–308, 2019.
  • [6] Julien Guillod. On the asymptotic stability of steady flows with nonzero flux in two-dimensional exterior domains. Comm. Math. Phys., 352(1):201–214, 2017.
  • [7] Julien Guillod and Peter Wittwer. Asymptotic behaviour of solutions to the stationary Navier-Stokes equations in two-dimensional exterior domains with zero velocity at infinity. Math. Models Methods Appl. Sci., 25(2):229–253, 2015.
  • [8] Mitsuo Higaki. Note on the stability of planar stationary flows in an exterior domain without symmetry. Adv. Differential Equations, 24(11-12):647–712, 2019.
  • [9] Matthieu Hillairet and Peter Wittwer. On the existence of solutions to the planar exterior Navier Stokes system. J. Differential Equations, 255(10):2996–3019, 2013.
  • [10] Mikhail V. Korobkov, Konstantin Pileckas, and Remigio Russo. On convergence of arbitrary DD-solution of steady Navier-Stokes system in 2D exterior domains. Arch. Ration. Mech. Anal., 233(1):385–407, 2019.
  • [11] Mikhail V. Korobkov, Konstantin Pileckas, and Remigio Russo. On the steady Navier-Stokes equations in 2D2D exterior domains. J. Differential Equations, 269(3):1796–1828, 2020.
  • [12] Mikhail V. Korobkov, Konstantin Pileckas, and Remigio Russo. Leray’s plane steady state solutions are nontrivial. Adv. Math., 376:Paper No. 107451, 20, 2021.
  • [13] Mikhail V. Korobkov and Xiao Ren. Uniqueness of plane stationary Navier-Stokes flow past an obstacle. Arch. Ration. Mech. Anal., 240(3):1487–1519, 2021.
  • [14] Hideo Kozono and Hermann Sohr. On a new class of generalized solutions for the Stokes equations in exterior domains. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 19(2):155–181, 1992.
  • [15] Jean Leray. Étude de diverses équations intégrales non linéaires et de quelques problèmes que pose l’hydrodynamique. J. Math. Pures Appl., 12:1–82, 1933.
  • [16] Yasunori Maekawa. On stability of steady circular flows in a two-dimensional exterior disk. Arch. Ration. Mech. Anal., 225(1):287–374, 2017.
  • [17] Yasunori Maekawa. Remark on stability of scale-critical stationary flows in a two-dimensional exterior disk. In Mathematics for nonlinear phenomena—analysis and computation, volume 215 of Springer Proc. Math. Stat., pages 105–130. Springer, Cham, 2017.
  • [18] Tomoyuki Nakatsuka. On uniqueness of symmetric Navier-Stokes flows around a body in the plane. Adv. Differential Equations, 20(3-4):193–212, 2015.
  • [19] Konstantin Pileckas and Remigio Russo. On the existence of vanishing at infinity symmetric solutions to the plane stationary exterior Navier-Stokes problem. Math. Ann., 352(3):643–658, 2012.
  • [20] Antonio Russo. A note on the exterior two-dimensional steady-state Navier-Stokes problem. J. Math. Fluid Mech., 11(3):407–414, 2009.
  • [21] Hermann Sohr. The Navier-Stokes equations: An elementary functional analytic approach. Modern Birkhäuser Classics. Birkhäuser/Springer Basel AG, Basel, 2001.
  • [22] Masao Yamazaki. Unique existence of stationary solutions to the two-dimensional Navier-Stokes equations on exterior domains. In Mathematical analysis on the Navier-Stokes equations and related topics, past and future, volume 35 of GAKUTO Internat. Ser. Math. Sci. Appl., pages 220–241. Gakkotosho, Tokyo, 2011.
  • [23] Masao Yamazaki. Two-dimensional stationary Navier-Stokes equations with 4-cyclic symmetry. Math. Nachr., 289(17-18):2281–2311, 2016.
  • [24] Masao Yamazaki. Existence and uniqueness of weak solutions to the two-dimensional stationary Navier-Stokes exterior problem. J. Math. Fluid Mech., 20(4):2019–2051, 2018.

M. Higaki

Department of Mathematics, Graduate School of Science, Kobe University, 1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan.

Email: [email protected]