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

Stability of planar exterior stationary flows with suction

Mitsuo Higaki
Department of Mathematics, Graduate School of Science, Kobe University,
1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan
E-mail: [email protected]

Abstract. We consider the two-dimensional Navier-Stokes system in a domain exterior to a disk. The system admits a stationary solution with critical decay O(|x|1)O(|x|^{-1}) written as a linear combination of the pure rotating flow and the flux carrier. We prove its nonlinear stability in large time for initial disturbances in L2L^{2} under smallness conditions, assuming that there is suction across the boundary, namely that the sign of coefficients of the flux carrier is negative. This result partially solves an open problem in the literature.

Keywords. Navier-Stokes system, Two-dimensional exterior domains, Stability of stationary solutions, Scale-critical decay.

2020 MSC. 35Q30, 35B35, 76D05, 76D17.

1 Introduction

We consider the two-dimensional Navier-Stokes system in an exterior disk

{tuΔu+p=uuin(0,)×Ωdivu=0in[0,)×Ωu(x)=αxδxon(0,)×Ωu|t=0=u0inΩ.\left\{\begin{array}[]{ll}\partial_{t}u-\Delta u+\nabla p=-u\cdot\nabla u&\mbox{in}\ (0,\infty)\times\Omega\\ \operatorname{div}u=0&\mbox{in}\ [0,\infty)\times\Omega\\ u(x)=\alpha x^{\bot}-\delta x&\mbox{on}\ (0,\infty)\times\partial\Omega\\ u|_{t=0}=u_{0}&\mbox{in}\ \Omega.\end{array}\right. (NS)

The unknown functions u=(u1(t,x),u2(t,x))u=(u_{1}(t,x),u_{2}(t,x)) and p=p(t,x)p=p(t,x) are respectively the velocity field of the fluid and the pressure field. The function u0=(u0,1(x),u0,2(x))u_{0}=(u_{0,1}(x),u_{0,2}(x)) is a given initial data. The set Ω\Omega denotes the exterior unit disk {x=(x1,x2)2||x|>1}\{x=(x_{1},x_{2})\in\mathbb{R}^{2}~{}|~{}|x|>1\} where |x|=x12+x22|x|=\sqrt{x_{1}^{2}+x_{2}^{2}}. We assume that both α\alpha and δ\delta are real number constants. The vector xx^{\bot} refers to (x2,x1)(-x_{2},x_{1}). The system (NS) describes the time evolution of viscous incompressible fluids around the disk rotating at angular velocity α\alpha on whose surface there is suction in the orthogonal direction when δ>0\delta>0 and injection when δ<0\delta<0.

The system (NS) admits an explicit stationary solution (αUδW,Pα,δ)(\alpha U-\delta W,\nabla P_{\alpha,\delta}) where

U(x)=x|x|2,W(x)=x|x|2,\displaystyle\begin{split}U(x)=\frac{x^{\bot}}{|x|^{2}},\qquad W(x)=\frac{x}{|x|^{2}},\end{split} (1.1)

and

Pα,δ(x)=(|αU(x)δW(x)|22).\displaystyle\begin{split}\nabla P_{\alpha,\delta}(x)=-\nabla\Big{(}\frac{|\alpha U(x)-\delta W(x)|^{2}}{2}\Big{)}.\end{split} (1.2)

This velocity is a linear combination of the vector field UU denoting the pure rotating flow in Ω\Omega and WW the flux carrier. To lighten notation, in the following, we write

V=V(α,δ)=αUδW.\displaystyle V=V(\alpha,\delta)=\alpha U-\delta W. (1.3)

The solution VV is invariant under the scaling of the Navier-Stokes equations. A (non-trivial) solution having this property is said to be scale-critical and it represents the balance between the nonlinear and linear parts of the equations. Therefore, investigating the properties of the scale-critical solutions is a fundamental and important issue in understanding the typical behavior of the Navier-Stokes flows. Let us mention that VV is an element of the family of stationary solutions of (NS) found by Hamel [20]. This family is known to be an example showing the non-uniqueness of the DD-solutions; see Galdi [15, Section XII.2]. The Hamel solutions are generalized by Guillod and Wittwer [19] in view of rotation symmetries.

In this paper, we study the nonlinear stability of VV in large time. More precisely, assuming that an initial disturbance around VV belongs to the Lebesgue spaces, we consider the time evolution of the disturbance in the nonlinear system (NS). Particularly, we are interested in the large-time decay estimate. By using the relation

uv+vu=urotv+vrotu+(|u+v|2|u|2|v|22)\displaystyle u\cdot\nabla v+v\cdot\nabla u=u^{\bot}{\rm rot}\,v+v^{\bot}{\rm rot}\,u+\nabla\Big{(}\frac{|u+v|^{2}-|u|^{2}-|v|^{2}}{2}\Big{)} (1.4)

and rotV=0\operatorname{rot}V=0, we see that the pair of new unknown functions

v=uVandq=(p+|u|22)v=u-V\qquad\text{and}\qquad\nabla q=\nabla\Big{(}p+\frac{|u|^{2}}{2}\Big{)}

formally solves the nonlinear problem

{tvΔv+Vrotv+q=vrotvin(0,)×Ωdivv=0in[0,)×Ωv=0on(0,)×Ωv|t=0=v0:=u|t=0VinΩ.\left\{\begin{array}[]{ll}\partial_{t}v-\Delta v+V^{\bot}\operatorname{rot}v+\nabla q=-v^{\bot}\operatorname{rot}v&\mbox{in}\ (0,\infty)\times\Omega\\ \operatorname{div}v=0&\mbox{in}\ [0,\infty)\times\Omega\\ v=0&\mbox{on}\ (0,\infty)\times\partial\Omega\\ v|_{t=0}=v_{0}:=u|_{t=0}-V&\mbox{in}\ \Omega.\end{array}\right. (NP)

The linearized problem of (NP) is given by

{tvΔv+Vrotv+q=0in(0,)×Ωdivv=0in[0,)×Ωv=0on(0,)×Ωv|t=0=v0inΩ.\left\{\begin{array}[]{ll}\partial_{t}v-\Delta v+V^{\bot}\operatorname{rot}v+\nabla q=0&\mbox{in}\ (0,\infty)\times\Omega\\ \operatorname{div}v=0&\mbox{in}\ [0,\infty)\times\Omega\\ v=0&\mbox{on}\ (0,\infty)\times\partial\Omega\\ v|_{t=0}=v_{0}&\mbox{in}\ \Omega.\end{array}\right. (LP)

Our main aim in this paper is to obtain large-time decay estimates of the solutions of (LP), by studying the operators associated with (LP). We will provide the LpL^{p}-LqL^{q} estimates sufficient to prove the nonlinear stability of the stationary solution VV in large time.

In order to make the framework clearer, we recall some notations and basic facts about the linear system (LP). We let C0,σ(Ω)C^{\infty}_{0,\sigma}(\Omega) denote {φC0(Ω)2|divφ=0}\{\varphi\in C^{\infty}_{0}(\Omega)^{2}~{}|~{}\operatorname{div}\varphi=0\}, Lσ2(Ω)L^{2}_{\sigma}(\Omega) the closure of C0,σ(Ω)C^{\infty}_{0,\sigma}(\Omega) in L2(Ω)2L^{2}(\Omega)^{2}, and :L2(Ω)2Lσ2(Ω){\mathbb{P}}:L^{2}(\Omega)^{2}\to L^{2}_{\sigma}(\Omega) the orthogonal projection. The operator {\mathbb{P}} is called the Helmholtz projection and satisfies p=0{\mathbb{P}}\nabla p=0 for pLloc2(Ω¯)p\in L^{2}_{{\rm loc}}(\overline{\Omega}) with pL2(Ω)2\nabla p\in L^{2}(\Omega)^{2}. The operator, called the Stokes operator, is defined by

𝔸=Δ,D(𝔸)=Lσ2(Ω)W01,2(Ω)2W2,2(Ω)2.{\mathbb{A}}=-{\mathbb{P}}\Delta,\qquad D({\mathbb{A}})=L^{2}_{\sigma}(\Omega)\cap W^{1,2}_{0}(\Omega)^{2}\cap W^{2,2}(\Omega)^{2}.

It is well known that 𝔸{\mathbb{A}} is nonnegative and self-adjoint in Lσ2(Ω)L^{2}_{\sigma}(\Omega) and that 𝔸-{\mathbb{A}} generates the C0C_{0}-analytic semigroup; see Sohr [39]. Moreover, the spectrum of 𝔸-{\mathbb{A}} is the set of nonpositive real numbers σ(𝔸)=0={x|x0}\sigma(-{\mathbb{A}})=\mathbb{R}_{\leq 0}=\{x\in\mathbb{R}~{}|~{}x\leq 0\}; see Section 3 for the references. With these notations, we define the operator associated with (LP) by

𝔸Vv=𝔸v+Vrotv,D(𝔸V)=D(𝔸),{\mathbb{A}}_{V}v={\mathbb{A}}v+{\mathbb{P}}V^{\bot}\operatorname{rot}v,\qquad D({\mathbb{A}}_{V})=D({\mathbb{A}}),

and write (LP) equivalently with the evolution system

dvdt+𝔸Vv=0in(0,),v|t=0=v0.\frac{\,{\rm d}v}{\,{\rm d}t}+{\mathbb{A}}_{V}v=0\mkern 9.0mu\mbox{in}\ (0,\infty),\qquad v|_{t=0}=v_{0}. (1.5)

We aim at proving the properties of solutions of (1.5) by studying the operator 𝔸V-{\mathbb{A}}_{V}.

One basic way to study the properties of 𝔸V-{\mathbb{A}}_{V} is to consider the equation

(λ+𝔸V)v=f(\lambda+{\mathbb{A}}_{V})v=f (R)

for given λ\lambda\in\mathbb{C} and fLσ2(Ω)f\in L^{2}_{\sigma}(\Omega). This equation can be obtained by formal application of the Laplace transform to (1.5). From the general theory of functional analysis, we find the following two facts. First, as the operator Vrot{\mathbb{P}}V^{\bot}\operatorname{rot} is lower order with respect to 𝔸{\mathbb{A}}, from theory for sectorial operators, we see that 𝔸V-{\mathbb{A}}_{V} is also sectorial in Lσ2(Ω)L^{2}_{\sigma}(\Omega) and generates the C0C_{0}-analytic semigroup, denoted by {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0}; see Lunardi [32, Proposition 2.4.3]. Second, as Vrot{\mathbb{P}}V^{\bot}\operatorname{rot} is relatively compact with respect to 𝔸{\mathbb{A}}, from the perturbation theory of operators, we see that σ(𝔸V)=0σdisc(𝔸V)\sigma(-{\mathbb{A}}_{V})=\mathbb{R}_{\leq 0}\cup\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}) where σdisc(𝔸V)\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}) denotes the discrete spectrum of 𝔸V-{\mathbb{A}}_{V}; see Section 3 for details. These two facts, however, are not sufficient to obtain the large-time estimate of {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0} since σ(𝔸V)\sigma(-{\mathbb{A}}_{V}) contains 0\mathbb{R}_{\leq 0}. We need a precise estimate of the resolvent (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1} when λ\lambda is close to the origin.

The fundamental difficulty in analyzing (R) when |λ|1|\lambda|\ll 1 is that the Hardy inequality

xf(x)|x|L2CfL2,fW˙01,2(Ω)d=(C0(Ω)¯L2)d\displaystyle\Big{\|}x\mapsto\frac{f(x)}{|x|}\Big{\|}_{L^{2}}\leq C\|\nabla f\|_{L^{2}},\quad f\in\dot{W}^{1,2}_{0}(\Omega)^{d}=\Big{(}\overline{C^{\infty}_{0}(\Omega)}^{\|\nabla\,\cdot\,\|_{L^{2}}}\Big{)}^{d} (1.6)

does not hold in exterior domains Ωd\Omega\subset\mathbb{R}^{d} when d=2d=2. If (1.6) holds when d=2d=2, the term Vrotv{\mathbb{P}}V^{\bot}\operatorname{rot}v in (R) can be controlled by the dissipation from Δv-\Delta v if |α|+|δ||\alpha|+|\delta| is small. Nevertheless, one needs a logarithmic correction in the left-hand side of (1.6) to obtain the correspondence; see [15, Theorem II.6.1]. This implies that energy method does not work well in general in deriving estimates for (R) when |λ|1|\lambda|\ll 1. One way to recover inequalities of the type (1.6) when d=2d=2 is to assume symmetries both on Ω\Omega and ff; see Galdi and Yamazaki [16], Yamazaki [41], and Guillod [18] for the stability results of symmetric flows under symmetries. As we do not assume any symmetries on initial data in (NP), unlike [16, 41, 18], such inequalities are not applicable to (LP) nor (R). This is in stark contrast to the three-dimensional stability results by Heywood [21] and by Borchers and Miyakawa [4, 5] in which the Hardy inequality (1.6) with d=3d=3 is an essential tool. As a recent monograph of the three-dimensional results, we refer to Brandolese and Schonbek [7].

Therefore, even for the flow V=V(α,δ)V=V(\alpha,\delta) explicitly given in (1.3), the stability analysis in two-dimensional exterior domains requires specific considerations depending on the parameters α\alpha and δ\delta. The known results are summarized as follows.

  • The case α=0\alpha=0 and δ0\delta\neq 0 is treated in Guillod [18]. This case is tractable and similar to the three-dimensional cases if |δ||\delta| is sufficiently small. In fact, for general exterior domains Ω2\Omega\subset\mathbb{R}^{2}, Russo [38, Lemma 3] proves the Hardy-type inequality

    |uu,W|CuL22,uW˙0,σ1,2(Ω)=C0,σ(Ω)¯L2.\displaystyle|\langle u\cdot\nabla u,W\rangle|\leq C\|\nabla u\|_{L^{2}}^{2},\quad u\in\dot{W}^{1,2}_{0,\sigma}(\Omega)=\overline{C^{\infty}_{0,\sigma}(\Omega)}^{\|\nabla\,\cdot\,\|_{L^{2}}}. (1.7)

    The reader is referred to [15, Remark X.4.2] and [18, Lemma 3] for further discussions. Combining (1.7) with the relation (1.4), we obtain the control

    |(δW)rotv,v|C|δ|vL22,vD(𝔸V).\displaystyle|\langle{\mathbb{P}}(\delta W)^{\bot}\operatorname{rot}v,v\rangle|\leq C|\delta|\|\nabla v\|_{L^{2}}^{2},\quad v\in D({\mathbb{A}}_{V}). (1.8)

    This observation implies that, by a simple energy estimate applied to (R), we can obtain the LpL^{p}-LqL^{q} estimates for the system (LP) and prove the nonlinear stability of V=δWV=\delta W. Alternatively, as is done in [18], one can prove the stability by considering L2L^{2}-estimates of the semigroup generated by the adjoint of the operator 𝔸δW-{\mathbb{A}}_{\delta W}. A similar idea is also used in Karch and Pilarczyk [27].

  • The case α0\alpha\neq 0 and δ=0\delta=0 is treated in Maekawa [33]. In this case, energy method is not useful for (R). Indeed, [18, Lemma 4] points out that the Hardy-type inequality (1.7) does not hold if WW is replaced by UU. To relax the situation, [33] considers the problem in an exterior disk and performs explicit computations. The LpL^{p}-LqL^{q} estimates for (LP) are obtained when |α||\alpha| is sufficiently small by an explicit formula for the resolvent (λ+𝔸αU)1(\lambda+{\mathbb{A}}_{\alpha U})^{-1}. Also, the nonlinear stability of αU\alpha U is proved when both |α||\alpha| and the L2L^{2}-norm of initial data in (NP) are sufficiently small. This stability result is extended by the author in [22] to a certain class of non-symmetric domains where the domains are assumed to be small perturbations of the exterior unit disk, and in [24] for three-dimensional initial disturbances around an infinite cylinder.

  • The case α0\alpha\neq 0 and δ0\delta\neq 0 is treated in Maekawa [34]. The problem is considered on an exterior disk as in [33]. The idea of the proof is to regard the term (δW)rotv{\mathbb{P}}(\delta W)^{\bot}\operatorname{rot}v in (R) as an external force and to utilize the estimate of (λ+𝔸αU)1(\lambda+{\mathbb{A}}_{\alpha U})^{-1} in [33]. The LpL^{p}-LqL^{q} estimates for (LP) are obtained when |α|+|δ||\alpha|+|\delta| is sufficiently small, under the restriction that initial data belong to a subcritical space L2LqL^{2}\cap L^{q} for some 1<q<21<q<2. Also, the nonlinear stability of VV is proved when both |α|+|δ||\alpha|+|\delta| and the (L2LqL^{2}\cap L^{q})-norm of initial data in (NP) are sufficiently small. This restriction on exponents is essentially needed when estimating (λ+𝔸αU)1(δW)rotv(\lambda+{\mathbb{A}}_{\alpha U})^{-1}{\mathbb{P}}(\delta W)^{\bot}\operatorname{rot}v. As is mentioned in [34, Remark 2], it is not clear if the condition q<2q<2 can be removed in this method.

1.1 Main results

This paper addresses large-time estimates for the system (LP), namely, of the semigroup {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0}, when α0\alpha\neq 0 and δ0\delta\neq 0 as in [34]. Our particular interest is the LpL^{p}-L2L^{2} estimates left as open problems in [34]. The following theorem solves it affirmatively under a condition on the sign of δ\delta. This condition is discussed in Remark 1.2 (iii) below.

Theorem 1.1

Let α,δ\alpha,\delta\in\mathbb{R} satisfy α0\alpha\neq 0 and δ0\delta\geq 0 and let |α|+δ|\alpha|+\delta be sufficiently small. For fLσ2(Ω)f\in L^{2}_{\sigma}(\Omega), we have

et𝔸VfL2CfL2,t>0,et𝔸VfL2Ct12fL2,t>0.\displaystyle\begin{split}\|e^{-t{\mathbb{A}}_{V}}f\|_{L^{2}}&\leq C\|f\|_{L^{2}},\quad t>0,\\ \|\nabla e^{-t{\mathbb{A}}_{V}}f\|_{L^{2}}&\leq Ct^{-\frac{1}{2}}\|f\|_{L^{2}},\quad t>0.\end{split} (1.9)

The constant CC depends on α,δ,p\alpha,\delta,p.

Remark 1.2
  1. (i)

    By combining Theorem 1.1 with the LpL^{p}-LqL^{q} estimates in [34] and by applying the Gagliardo-Nirenberg inequality, we obtain

    et𝔸VfLpCt1q+1pfLq,t>0,et𝔸VfL2Ct1qfLq,t>0\displaystyle\begin{split}\|e^{-t{\mathbb{A}}_{V}}f\|_{L^{p}}&\leq Ct^{-\frac{1}{q}+\frac{1}{p}}\|f\|_{L^{q}},\quad t>0,\\ \|\nabla e^{-t{\mathbb{A}}_{V}}f\|_{L^{2}}&\leq Ct^{-\frac{1}{q}}\|f\|_{L^{q}},\quad t>0\end{split} (1.10)

    for 1<q2p<1<q\leq 2\leq p<\infty and fLσ2(Ω)Lq(Ω)2f\in L^{2}_{\sigma}(\Omega)\cap L^{q}(\Omega)^{2} with a constant C=C(α,δ,q,p)C=C(\alpha,\delta,q,p).

  2. (ii)

    The proof of Theorem 1.1 is based on an analysis of the operator 𝔸V-{\mathbb{A}}_{V}. The estimate (1.9) for {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0} is deduced by the Dunford integral of the resolvent (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1}. Inspired by [33], we determine the spectrum of 𝔸V-{\mathbb{A}}_{V} and estimate (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1} by explicit computations. It is shown in Section 3 that the function characterizing the discrete spectrum of 𝔸V-{\mathbb{A}}_{V} crucially depends on both α\alpha and δ\delta. Therefore, it is suggested that, when estimating (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1} for |λ|1|\lambda|\ll 1, one cannot regard (δW)rotv{\mathbb{P}}(\delta W)^{\bot}\operatorname{rot}v in (R) as an external force even if |δ||\delta| is small, in spite of the control (1.8).

  3. (iii)

    It is an open problem whether the same estimate as in (1.9) can be obtained for the case δ<0\delta<0. Actually, by following the argument in Section 4, one can prove (1.9) if δ\delta is chosen to depend on a given α\alpha, but the general case is still open. It might be meaningful to recall here that the case δ<0\delta<0 corresponds to the situation where there is injection into fluids at the boundary. We mention Drazin and Reid [11, Problem 3.7] and Drazin and Riley [12, Section 3.1] as the references related to this topic.

  4. (iv)

    It is important to extend the LpL^{p}-LqL^{q} estimates in (1.10) to general exterior domains. However, this is a difficult problem because of the dependence of constants on α,δ\alpha,\delta. The problem when δ=0\delta=0 is tackled in [22] and it is shown that, if the domain Ω\Omega is a perturbation from the exterior unit disk in algebraic order of |α||\alpha|, then the LpL^{p}-LqL^{q} estimates can be obtained by energy method combined with explicit formulas. The restriction to a class of domains is due to singularity in the operator norm of the resolvent (λ+𝔸αU)1(\lambda+{\mathbb{A}}_{\alpha U})^{-1} for small |α||\alpha|. It is observed that, in explicit computations, cancellation of the effects from the two terms λv\lambda v and αUrotv\alpha U^{\bot}\operatorname{rot}v in (R) (with δ=0\delta=0) occurs for λ\lambda in a certain domain, dubbed the “nearly-resonance regime” in [22]. This cancellation causes the singularity at algebraic order of |α||\alpha|, which in energy method restricts the shape of domains, more precisely the lengths between domains and the exterior unit disk. Such singularity also appears in the operator norm of (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1} for small |α|+δ|\alpha|+\delta in the present problem and is an obstacle to the extension.

By using Theorem 1.1, we can prove the nonlinear stability of VV. Using the semigroup {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0}, we consider the mild solutions of (NP) solving

v(t)=et𝔸Vv00te(ts)𝔸V(vrotv)(s)ds,t>0.\displaystyle v(t)=e^{-t{\mathbb{A}}_{V}}v_{0}-\int_{0}^{t}e^{-(t-s){\mathbb{A}}_{V}}\mathbb{P}(v^{\bot}\operatorname{rot}v)(s)\,{\rm d}s,\quad t>0. (1.11)

The following theorem can be shown by a simple application of the Banach fixed point theorem and thus is omitted in this paper. For details, see [33] treating the case δ=0\delta=0.

Theorem 1.3

Let α,δ\alpha,\delta\in\mathbb{R} satisfy α0\alpha\neq 0 and δ0\delta\geq 0 and let |α|+δ|\alpha|+\delta be sufficiently small. Let v0v_{0} belong to Lσ2(Ω)L^{2}_{\sigma}(\Omega) and let v0L2\|v_{0}\|_{L^{2}} be sufficiently small depending on α,δ\alpha,\delta. There is a unique mild solution vC([0,);Lσ2(Ω))C((0,);W01,2(Ω)2)v\in C\big{(}[0,\infty);L^{2}_{\sigma}(\Omega)\big{)}\cap C\big{(}(0,\infty);W^{1,2}_{0}(\Omega)^{2}\big{)} of (1.11) satisfying

limttk2kv(t)L2=0,k=0,1.\displaystyle\lim_{t\to\infty}t^{\frac{k}{2}}\|\nabla^{k}v(t)\|_{L^{2}}=0,\quad k=0,1. (1.12)

1.2 Related results

Let us refer to the results that are closely related to the present study.

Analysis of (NP) and (LP) when V0V\equiv 0. For (NP), the estimate (1.12) for V0V\equiv 0, which can be viewed as the nonlinear stability of the trivial solution, is classical; see Masuda [37] for the proof when k=0k=0 and Kozono and Ogawa [29] when k=1k=1. These results do not require smallness on the initial data in Lσ2(Ω)L^{2}_{\sigma}(\Omega). For (LP), the LpL^{p}-LqL^{q} estimates of the Stokes semigroup {et𝔸}t0\{e^{-t{\mathbb{A}}}\}_{t\geq 0} are established by Maremonti and Solonnikov [36] and by Dan and Shibata [9, 10]. We note that all of the results above hold in general exterior domains Ω2\Omega\subset\mathbb{R}^{2}. It is pointed out in [33, Remark 1.4] that the logarithmic singularity of the resolvent (λ+𝔸)1(\lambda+{\mathbb{A}})^{-1} for small |λ||\lambda|, observed in [9, §3], disappears in (λ+𝔸αU)1(\lambda+{\mathbb{A}_{\alpha U}})^{-1} if α0\alpha\neq 0. As compensation, however, singularity appears in the operator norm of (λ+𝔸αU)1(\lambda+{\mathbb{A}_{\alpha U}})^{-1} for small |α||\alpha|. Such singularity, as discussed in Remark 1.2 (iv), also appears in the operator norm of (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1}, and is an obstacle when generalizing the LpL^{p}-LqL^{q} estimates in (1.10). Let us mention the study of the boundedness of {et𝔸}t0\{e^{-t{\mathbb{A}}}\}_{t\geq 0} in spaces Lσp(Ω)=C0,σ(Ω)¯LpL^{p}_{\sigma}(\Omega)=\overline{C^{\infty}_{0,\sigma}(\Omega)}^{\|\,\cdot\,\|_{L^{p}}} pioneered by Borchers and Varnhorn [6]. See Abe [1, 2] for the recent progress.

Non-symmetric stationary solutions around VV. We consider the stationary problem of (NS), which also admits the explicit solution VV. It is known that, for suitably chosen α,δ\alpha,\delta, the fundamental solution for the linearized problem around VV, namely for the stationary problem of (LP), has a better spatial decay compared to the one for the problem linearized around the trivial solution V0V\equiv 0. This improvement is due to the vorticity transport by VV and implies the resolution of the famous Stokes paradox; see [8, 14, 15, 30, 26] for descriptions. Furthermore, these new fundamental solutions allow us to construct non-symmetric solutions for the nonlinear problem decaying in the order O(|x|1)O(|x|^{-1}). This is done in Hillairet and Wittwer [25] when |α|>48|\alpha|>\sqrt{48} and δ=0\delta=0 for given zero-flux boundary data in a suitable class, and in [23] when α\alpha\in\mathbb{R} and δ>2\delta>2 for given external forces with suitable spatial decay. The solutions in [23] are compatible with the Liouville-type theorem in Guillod [17, Proposition 4.6]. We emphasize that the results in [25, 23] do not require any symmetries on the given data. Interestingly, such improvement in the fundamental solutions occurs even for small α,δ\alpha,\delta. Indeed, Maekawa and Tsurumi [35] constructs non-symmetric solutions for the nonlinear problem in the whole space 2\mathbb{R}^{2}, whose principal part at spatial infinity is cUcU with a small but nonzero constant cc. This result is contrasting with [25] in view of the size of coefficients, and the reason is that, as there are no boundaries in 2\mathbb{R}^{2}, the terms needed to match the no-slip boundary condition in exterior domains do not appear in the problem.

1.3 Outlined proof

We describe the proof of Theorem 1.1. However, the estimate (1.9) is almost a direct consequence of the estimate of the resolvent (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1} in Proposition 5.1. Hence we give in Appendix C the proof that derives Theorem 1.1 from Proposition 5.1, and outline here the proof of Proposition 5.1. As noted in Remark 1.2 (ii), it consists of two steps:

(I) Spectral analysis of 𝔸V-{\mathbb{A}}_{V}. Recall that σ(𝔸V)=0σdisc(𝔸V)\sigma(-{\mathbb{A}}_{V})=\mathbb{R}_{\leq 0}\cup\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}). Thus we identify the location of the discrete spectrum σdisc(𝔸V)\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}) to obtain the large-time decay of {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0}. For this purpose, we consider the homogeneous equation of (R) and its general solutions, by using the streamfunction-vorticity equations. We see that the no-slip boundary condition imposes that λ\lambda belongs to σdisc(𝔸V)\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}) if and only if λ\lambda belongs to

n{λ0|Fn(λ)=0}.\bigcup_{n\in\mathbb{Z}}\{\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}~{}|~{}F_{n}(\sqrt{\lambda})=0\}.

Here Fn=Fn(z)F_{n}=F_{n}(z) is the analytic function defined in (3.7) in Section 3. For |n|1|n|\neq 1, one can show that Fn()F_{n}(\sqrt{\cdot}) has no zeros in the sector Σ34π\Sigma_{\frac{3}{4}\pi} by energy method if |α|+|δ||\alpha|+|\delta| is sufficiently small; see Propositions 3.3 and 3.4. However, for |n|=1|n|=1, we need to deal with the function FnF_{n} directly to determine the location of its zeros, which reflects the fact that the Hardy inequality does not hold in two-dimensional exterior domains. We will prove that FnF_{n} with |n|=1|n|=1 has no zeros in sectors Σ34πϵ\Sigma_{\frac{3}{4}\pi-\epsilon} for ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}) if δ0\delta\geq 0 and |α|+δ|\alpha|+\delta is sufficiently small depending on ϵ\epsilon. The proof is the most tricky part of this paper and will take the whole of Section 4. We perform an asymptotic analysis of FnF_{n} that refines the methods in [33, 22]. Interestingly, the analysis is highly dependent on the sign of δ\delta being positive or negative. Furthermore, we observe that the condition δ>0\delta>0, which is also an assumption of Theorem 1.1, provides a certain stabilizing effect compared to the case δ=0\delta=0; see Remark 4.8.

(II) Estimate of the resolvent (λ+𝔸V)1(\lambda+{\mathbb{A}}_{V})^{-1}. In the next step, we estimate the solution of (R) for λ\lambda belonging to the resolvent set. We derive and estimate an explicit formula for the solution using the streamfunction-vorticity equations. The computations are lengthy ones estimating the formulas involving the modified Bessel functions, but the approach itself is broadly the same as that used in [33, 22]. Thus we omit some details; see Section 5.

This paper is organized as follows. In Section 2, we collect the items used in this paper. In Sections 3 and 4, we study the spectrum of the operator 𝔸V-{\mathbb{A}}_{V}. We apply the perturbation theory of operators in Section 3 and perform an asymptotic analysis of FnF_{n} with |n|=1|n|=1 in Section 4. In Section 5, we provide the estimate of the resolvent. Some facts about the modified Bessel functions and technical supplements are given in Appendices A, B and C.

Notations. We let CC denote a constant and 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 denote ={x|x0}\mathbb{R}^{\ast}=\{x\in\mathbb{R}~{}|~{}x\neq 0\}, 0={x|x0}\mathbb{R}_{\geq 0}=\{x\in\mathbb{R}~{}|~{}x\geq 0\}, 0={x|x0}\mathbb{R}_{\leq 0}=\{x\in\mathbb{R}~{}|~{}x\leq 0\} and Σϕ={z{0}||argz|<ϕ}\Sigma_{\phi}=\{z\in\mathbb{C}\setminus\{0\}~{}|~{}|\operatorname{arg}z|<\phi\}. For zz\in\mathbb{C}, let z\Re z and z\Im z denote the real and imaginary parts of zz, respectively. For z0z\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}, let zμz^{\mu} denote eμLogze^{\mu\operatorname{Log}z} where z=|z|eiargzz=|z|e^{i\operatorname{arg}z}, argz(π,π)\operatorname{arg}z\in(-\pi,\pi) and Logz=log|z|+iargz\operatorname{Log}z=\log|z|+i\operatorname{arg}z. We take the square root z\sqrt{z} so that z>0\Re\sqrt{z}>0. We use the function spaces

W^1,2(Ω)={pLloc2(Ω¯)|pL2(Ω)2}\widehat{W}^{1,2}(\Omega)=\{p\in L^{2}_{{\rm loc}}(\overline{\Omega})~{}|~{}\nabla p\in L^{2}(\Omega)^{2}\}

and

C0,σ(Ω)={φC0(Ω)2|divφ=0},Lσ2(Ω)=C0,σ(Ω)¯L2.C^{\infty}_{0,\sigma}(\Omega)=\{\varphi\in C^{\infty}_{0}(\Omega)^{2}~{}|~{}\operatorname{div}\varphi=0\},\qquad L^{2}_{\sigma}(\Omega)=\overline{C^{\infty}_{0,\sigma}(\Omega)}^{\|\,\cdot\,\|_{L^{2}}}.

Not to burden notation, we use the same symbols to denote the quantities for scalar-, vector- or tensor-valued functions, e.g., ,\langle\cdot,\cdot\rangle is 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 the items used throughout the paper.

2.1 Vectors in the polar coordinates

The polar coordinates on the exterior unit disk Ω\Omega are written as

x1=rcosθ,x2=rsinθ,r[1,),θ[0,2π),\displaystyle x_{1}=r\cos\theta,\qquad x_{2}=r\sin\theta,\quad r\in[1,\infty),\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}.

Let a vector field v=(v1,v2)v=(v_{1},v_{2}) on Ω\Omega be given. 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 for a given nn\in\mathbb{Z},

𝒫nv(r,θ)=vr,n(r)einθ𝐞r+vθ,n(r)einθ𝐞θ,vr,n(r):=12π02πvr(rcosσ,rsinσ)einσdσ,vθ,n(r):=12π02πvθ(rcosσ,rsinσ)einσdσ.\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\sigma,r\sin\sigma)e^{-in\sigma}\,{\rm d}\sigma,\\ v_{\theta,n}(r)&:=\frac{1}{2\pi}\int_{0}^{2\pi}v_{\theta}(r\cos\sigma,r\sin\sigma)e^{-in\sigma}\,{\rm d}\sigma.\end{split} (2.1)

We will use the formulas

|v|2=|rvr|2+|rvθ|2+1r2(|θvrvθ|2+|vr+θvθ|2),divv=1v1+2v2=1r(r(rvr)+θvθ),rotv=1v22v1=1r(r(rvθ)θvr),\displaystyle\begin{split}|\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}),\\ \operatorname{div}v&=\partial_{1}v_{1}+\partial_{2}v_{2}=\frac{1}{r}\Big{(}\partial_{r}(rv_{r})+\partial_{\theta}v_{\theta}\Big{)},\\ \operatorname{rot}v&=\partial_{1}v_{2}-\partial_{2}v_{1}=\frac{1}{r}\Big{(}\partial_{r}(rv_{\theta})-\partial_{\theta}v_{r}\Big{)},\end{split} (2.2)

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} (2.3)

2.2 Fourier series and decomposition

Let nn\in\mathbb{Z} and 𝒫n\mathcal{P}_{n} be defined in (2.1). We set, 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), (2.4)

for a scalar function ω=ω(r,θ)\omega=\omega(r,\theta) on Ω\Omega,

𝒫nω(r,θ)=(12π02πω(rcosσ,rsinσ)einσdσ)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\sigma,r\sin\sigma)e^{-in\sigma}\,{\rm d}\sigma\bigg{)}e^{in\theta},\\ \omega_{n}(r)&=(\mathcal{P}_{n}\omega)e^{-in\theta},\end{split} (2.5)

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{\}}.

The definition of fnf_{n} differs according to whether ff is vectorial or scalar. The former and latter are defined in (2.4) as fn=𝒫nff_{n}=\mathcal{P}_{n}f and in (2.5) as fn=(𝒫nf)einθf_{n}=(\mathcal{P}_{n}f)e^{-in\theta}, respectively.

By definition, any vector field vL2(Ω)2v\in L^{2}(\Omega)^{2} is expanded into the convergent series

v=n𝒫nv=nvn,v=\sum_{n\in\mathbb{Z}}\mathcal{P}_{n}v=\sum_{n\in\mathbb{Z}}v_{n},

and 𝒫n\mathcal{P}_{n} is an orthogonal projection of L2(Ω)2L^{2}(\Omega)^{2} onto 𝒫nL2(Ω)2\mathcal{P}_{n}L^{2}(\Omega)^{2}. Moreover, the following orthogonal decomposition of the subspace Lσ2(Ω)L2(Ω)2L^{2}_{\sigma}(\Omega)\subset L^{2}(\Omega)^{2} holds:

Lσ2(Ω)=nLσ,n2(Ω),Lσ,n2(Ω):=𝒫nLσ2(Ω).\displaystyle L^{2}_{\sigma}(\Omega)=\bigoplus_{n\in\mathbb{Z}}L^{2}_{\sigma,n}(\Omega),\qquad L^{2}_{\sigma,n}(\Omega):=\mathcal{P}_{n}L^{2}_{\sigma}(\Omega). (2.6)

From (2.2), we have

vL22\displaystyle\|\nabla v\|_{L^{2}}^{2} =n𝒫nvL22,\displaystyle=\sum_{n\in\mathbb{Z}}\|\nabla\mathcal{P}_{n}v\|_{L^{2}}^{2},
|𝒫nv|2\displaystyle|\nabla\mathcal{P}_{n}v|^{2} =|rvr,n|2+|rvθ,n|2+1+n2r2(|vr,n|2+|vθ,n|2)4nr2(vθ,nvr,n¯).\displaystyle=|\partial_{r}v_{r,n}|^{2}+|\partial_{r}v_{\theta,n}|^{2}+\frac{1+n^{2}}{r^{2}}(|v_{r,n}|^{2}+|v_{\theta,n}|^{2})-\frac{4n}{r^{2}}\Im(v_{\theta,n}\overline{v_{r,n}}).

In particular,

|rvr,n|2+|rvθ,n|2+(|n|1)2r2(|vr,n|2+|vθ,n|2)|𝒫nv|2.|\partial_{r}v_{r,n}|^{2}+|\partial_{r}v_{\theta,n}|^{2}+\frac{(|n|-1)^{2}}{r^{2}}(|v_{r,n}|^{2}+|v_{\theta,n}|^{2})\leq|\nabla\mathcal{P}_{n}v|^{2}.

Therefore, if |n|1|n|\neq 1, the Hardy-type inequality

(r,θ)𝒫nv(r,θ)rL2𝒫nvL2\Big{\|}(r,\theta)\mapsto\frac{\mathcal{P}_{n}v(r,\theta)}{r}\Big{\|}_{L^{2}}\leq\|\nabla\mathcal{P}_{n}v\|_{L^{2}}

holds. Thus it is convenient to set

v=v|n|=1𝒫nv,v_{\neq}=v-\sum_{|n|=1}\mathcal{P}_{n}v,

for which we have

(r,θ)v(r,θ)rL2vL2.\displaystyle\Big{\|}(r,\theta)\mapsto\frac{v_{\neq}(r,\theta)}{r}\Big{\|}_{L^{2}}\leq\|\nabla v_{\neq}\|_{L^{2}}. (2.7)

Again from (2.2), we have, for vW1,2(Ω)2v\in W^{1,2}(\Omega)^{2},

𝒫ndivv=div𝒫nv,𝒫nrotv=rot𝒫nv\mathcal{P}_{n}\operatorname{div}v=\operatorname{div}\mathcal{P}_{n}v,\qquad\mathcal{P}_{n}\operatorname{rot}v=\operatorname{rot}\mathcal{P}_{n}v

and, from (2.3), for vW2,2(Ω)2v\in W^{2,2}(\Omega)^{2},

𝒫nΔv=Δ𝒫nv.\mathcal{P}_{n}\Delta v=\Delta\mathcal{P}_{n}v.

Since the condition 𝐞rv=0{\bf e}_{r}\cdot v=0 on Ω\partial\Omega is preserved under 𝒫n\mathcal{P}_{n}, it can be shown that

𝒫n=𝒫n,Lσ,n2(Ω)=𝒫nC0,σ(Ω)¯L2.\displaystyle\mathcal{P}_{n}\mathbb{P}=\mathbb{P}\mathcal{P}_{n},\qquad L^{2}_{\sigma,n}(\Omega)=\overline{\mathcal{P}_{n}C^{\infty}_{0,\sigma}(\Omega)}^{\|\,\cdot\,\|_{L^{2}}}.

We refer to Farwig and Neustupa [13, Lemma 3.1] for a more detailed proof. Although the proof in [13] is for the three-dimensional cases, a similar argument is applicable.

Now we define the closed linear operator 𝔸n\mathbb{A}_{n} on Lσ,n2(Ω)L^{2}_{\sigma,n}(\Omega) in (2.6) by

𝔸n=𝔸|Lσ,n2(Ω)D(𝔸),D(𝔸n)=Lσ,n2(Ω)D(𝔸).\mathbb{A}_{n}=\mathbb{A}|_{L^{2}_{\sigma,n}(\Omega)\cap D(\mathbb{A})},\qquad D(\mathbb{A}_{n})=L^{2}_{\sigma,n}(\Omega)\cap D(\mathbb{A}).

It is not hard to see that 𝔸n\mathbb{A}_{n} is nonnegative and self-adjoint. Also, keeping the relation

𝒫nVrotv=Vrotvn,vW1,2(Ω)2\mathcal{P}_{n}{\mathbb{P}}V^{\bot}\operatorname{rot}v={\mathbb{P}}V^{\bot}\operatorname{rot}v_{n},\quad v\in W^{1,2}(\Omega)^{2}

in mind, we define the closed linear operator 𝔸V,n\mathbb{A}_{V,n} on Lσ,n2(Ω)L^{2}_{\sigma,n}(\Omega) by

𝔸V,n=𝔸V|Lσ,n2(Ω)D(𝔸V),D(𝔸V,n)=D(𝔸n).\mathbb{A}_{V,n}=\mathbb{A}_{V}|_{L^{2}_{\sigma,n}(\Omega)\cap D(\mathbb{A}_{V})},\qquad D(\mathbb{A}_{V,n})=D(\mathbb{A}_{n}).

2.3 Equations in the polar coordinates

To study the operator 𝔸V,n\mathbb{A}_{V,n}, we consider

(λ+𝔸V,n)vn=fn(\lambda+\mathbb{A}_{V,n})v_{n}=f_{n} (Rn)

for given λ0\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0} and fnLσ,n2(Ω)f_{n}\in L^{2}_{\sigma,n}(\Omega). The equation is equivalent to the system

{λvnΔvn+Vrotvn+qn=fninΩdivvn=0inΩvn=0onΩ,\left\{\begin{array}[]{ll}\lambda v_{n}-\Delta v_{n}+V^{\bot}\operatorname{rot}v_{n}+\nabla q_{n}=f_{n}&\mbox{in}\ \Omega\\ \operatorname{div}v_{n}=0&\mbox{in}\ \Omega\\ v_{n}=0&\mbox{on}\ \partial\Omega,\end{array}\right. (2.8)

with some pressure qn\nabla q_{n}. Operating rot\operatorname{rot} to the first line, we see that rotvn\operatorname{rot}v_{n} solves

λ(rotvn)Δ(rotvn)+V(rotvn)=rotfn.\lambda(\operatorname{rot}v_{n})-\Delta(\operatorname{rot}v_{n})+V\cdot\nabla(\operatorname{rot}v_{n})=\operatorname{rot}f_{n}. (2.9)

In the polar coordinates on Ω\Omega where vn=vn(r,θ)v_{n}=v_{n}(r,\theta) is written as

vn(r,θ)=vr,n(r)einθ𝐞r+vθ,n(r)einθ𝐞θ,v_{n}(r,\theta)=v_{r,n}(r)e^{in\theta}{\bf e}_{r}+v_{\theta,n}(r)e^{in\theta}{\bf e}_{\theta},

we see from (2.8) that (vr,n(r),vθ,n(r))(v_{r,n}(r),v_{\theta,n}(r)) and qn(r)q_{n}(r) satisfy

λvr,nddr(1rddr(rvr,n))+n2r2vr,n+2inr2vθ,nαr2(ddr(rvθ,n)invr,n)+dqndr=fr,n,r>1,\displaystyle\begin{aligned} &\lambda v_{r,n}-\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{)}+\frac{\,{\rm d}q_{n}}{\,{\rm d}r}=f_{r,n},\quad r>1,\\ \end{aligned} (2.10)
λvθ,nddr(1rddr(rvθ,n))+n2r2vθ,n2inr2vr,nδr2(ddr(rvθ,n)invr,n)+inrqn=fθ,n,r>1\displaystyle\begin{aligned} &\lambda v_{\theta,n}-\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{\delta}{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} (2.11)

and the divergence-free and the no-slip boundary conditions

ddr(rvr,n)+invθ,n=0,r>1,vr,n(1)=vθ,n(1)=0.\displaystyle\frac{\,{\rm d}}{\,{\rm d}r}(rv_{r,n})+inv_{\theta,n}=0,\quad r>1,\qquad v_{r,n}(1)=v_{\theta,n}(1)=0. (2.12)

Moreover, from (2.9), ωn(r):=(rotvn)n(r)\omega_{n}(r):=(\operatorname{rot}v_{n})_{n}(r) satisfies

d2ωndr21+δrdωndr+(λ+n2+iαnr2)ωn=(rotfn)n,r>1.\displaystyle-\frac{\,{\rm d}^{2}\omega_{n}}{\,{\rm d}r^{2}}-\frac{1+\delta}{r}\frac{\,{\rm d}\omega_{n}}{\,{\rm d}r}+\Big{(}\lambda+\frac{n^{2}+i\alpha n}{r^{2}}\Big{)}\omega_{n}=(\operatorname{rot}f_{n})_{n},\quad r>1. (2.13)

2.4 Biot-Savart law

To simplify the explanation, only in this subsection, we use the function space

Ls(Ω)={fL(Ω)|fLs<},fLs:=esssupxΩ|x|s|f(x)|.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)|.

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.

Let ω𝒫nL2(Ω)\omega\in\mathcal{P}_{n}L^{\infty}_{2}(\Omega) with |n|1|n|\geq 1 and let ψ\psi be the decaying solution, called the streamfunction. Applying the notation in (2.5), we find that ψn=ψn(r)\psi_{n}=\psi_{n}(r) satisfies

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. (2.14)

By elementary computation, we see that ψn=ψn[ωn]\psi_{n}=\psi_{n}[\omega_{n}] 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} (2.15)

The following vector field

𝒱n[ωn](r,θ)=𝒱r,n[ωn](r)einθ𝐞r+𝒱θ,n[ωn](r)einθ𝐞θ,𝒱r,n[ωn](r):=inrψn[ωn](r),𝒱θ,n[ωn](r):=ddrψn[ωn](r)\displaystyle\begin{split}&\mathcal{V}_{n}[\omega_{n}](r,\theta)=\mathcal{V}_{r,n}[\omega_{n}](r)e^{in\theta}{\bf e}_{r}+\mathcal{V}_{\theta,n}[\omega_{n}](r)e^{in\theta}{\bf e}_{\theta},\\ &\mathcal{V}_{r,n}[\omega_{n}](r):=\frac{in}{r}\psi_{n}[\omega_{n}](r),\qquad\mathcal{V}_{\theta,n}[\omega_{n}](r):=-\frac{\,{\rm d}}{\,{\rm d}r}\psi_{n}[\omega_{n}](r)\end{split} (2.16)

is called the Biot-Savart law. It is straightforward to see that

div𝒱n[ωn]=0,rot𝒱n[ωn](r,θ)=ωn(r)einθ,(𝐞r𝒱n[ωn])|Ω=0.\displaystyle\begin{split}&\operatorname{div}\mathcal{V}_{n}[\omega_{n}]=0,\qquad\operatorname{rot}\mathcal{V}_{n}[\omega_{n}](r,\theta)=\omega_{n}(r)e^{in\theta},\qquad({\bf e}_{r}\cdot\mathcal{V}_{n}[\omega_{n}])|_{\partial\Omega}=0.\end{split} (2.17)

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

Here are useful two propositions in the subsequent sections. The reader is referred to [33, Proposition 2.6 and Lemma 3.1], [23, Proposition 2.1] for the proof of the first proposition and [33, Corollary 2.7], [23, Proposition 2.2] for the proof of the second.

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}=(\operatorname{rot}v_{n})_{n}. If divvn=0\operatorname{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}=\mathcal{V}_{n}[\omega_{n}] and dn[ωn]=0d_{n}[\omega_{n}]=0 in (2.15).

Proposition 2.2

Let |n|1|n|\geq 1 and fn𝒫nL2(Ω)2f_{n}\in\mathcal{P}_{n}L^{2}(\Omega)^{2}. If rotfn=0\operatorname{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).

3 Spectral analysis

In this section, we study the spectrum of the operator 𝔸V-{\mathbb{A}}_{V}. The main result is Proposition 3.4 which characterizes the discrete spectrum σdisc(𝔸V)\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}) as zeros of certain analytic functions. We are aware that the presentation in Subsections 3.1 and 3.2 has similarity to [33] treating the case δ=0\delta=0. This is quite natural because, in analysis in the L2L^{2}-framework, especially in computation of the numerical ranges, one can control terms involving δW\delta W by using the Hardy-type inequality (1.7). Consequently, for example, the statement of Proposition 3.3 holds independently of sufficiently small δ\delta. However, the difference from the case δ=0\delta=0 appears when one studies the spectrum of 𝔸V-{\mathbb{A}}_{V}. Indeed, in Proposition 3.4, the functions characterizing the discrete spectrum depend both on α\alpha and δ\delta. These functions will be studied in detail quantitatively in the next section.

3.1 Notation

Let us recall the standard notation in the perturbation theory. Our main reference is Kato [28]. Let XX be a Banach space and 𝕃:D(𝕃)XX\mathbb{L}:D(\mathbb{L})\subset X\to X be a closed linear operator. We let N(𝕃)N(\mathbb{L}) denote the null space of 𝕃\mathbb{L}, R(𝕃)R(\mathbb{L}) its range, and X/R(𝕃)X/R(\mathbb{L}) the quotient space of XX by R(𝕃)R(\mathbb{L}). Moreover, ρ(𝕃)\rho(\mathbb{L}) denotes the resolvent set of 𝕃\mathbb{L}, σ(𝕃)\sigma(\mathbb{L}) its spectrum, and σdisc(𝕃)\sigma_{{\rm disc}}(\mathbb{L}) its discrete spectrum, namely, the set of isolated eigenvalues of 𝕃\mathbb{L} with finite multiplicity. The operator 𝕃\mathbb{L} is said to be semi-Fredholm if R(𝕃)R(\mathbb{L}) is closed and at least one of dimN(𝕃)\operatorname{dim}N(\mathbb{L}) or dimX/R(𝕃)\operatorname{dim}X/R(\mathbb{L}) is finite. If 𝕃\mathbb{L} is semi-Fredholm, the index of 𝕃\mathbb{L}

ind(𝕃)=dimN(𝕃)dimX/R(𝕃)\operatorname{ind}(\mathbb{L})=\operatorname{dim}N(\mathbb{L})-\operatorname{dim}X/R(\mathbb{L})

is well-defined, taking values in [,][-\infty,\infty]. Finally, let us set

ρsf(𝕃)={λ|λ𝕃 is semi-Fredholm},σess(𝕃)=ρsf(𝕃)\rho_{{\rm sf}}(\mathbb{L})=\{\lambda\in\mathbb{C}~{}|~{}\text{$\lambda-\mathbb{L}$ is semi-Fredholm}\},\qquad\sigma_{{\rm ess}}(\mathbb{L})=\mathbb{C}\setminus\rho_{{\rm sf}}(\mathbb{L})

and call the semi-Fredholm domain of 𝕃\mathbb{L} and the essential spectrum of 𝕃\mathbb{L}, respectively.

Generally, ρsf(𝕃)\rho_{{\rm sf}}(\mathbb{L}) is the union of a countable (at most) family of connected open sets. From the argument in [28, Chapter IV\mathrm{I}\mathrm{V} §5 6], we see that ind(λ𝕃)\operatorname{ind}(\lambda-\mathbb{L}) is a constant function of λ\lambda in each component GG of ρsf(𝕃)\rho_{{\rm sf}}(\mathbb{L}). Moreover, both dimN(λ𝕃)\operatorname{dim}N(\lambda-\mathbb{L}) and dimX/R(λ𝕃)\operatorname{dim}X/R(\lambda-\mathbb{L}) are constants in each GG except for an isolated set of values of λ\lambda. Therefore, if these constants are zero in particular, then GG is contained in ρ(𝕃)\rho(\mathbb{L}) with possible exception of isolated points of σ(𝕃)\sigma(\mathbb{L}), which are, isolated eigenvalues of finite algebraic multiplicity.

3.2 Perturbation theory

We start with the perturbation theory of operators.

Proposition 3.1

Let α,δ\alpha,\delta\in\mathbb{R}. We have the following.

  1. (1)

    σess(𝔸V)=0\sigma_{{\rm ess}}(-{\mathbb{A}}_{V})=\mathbb{R}_{\leq 0} and σdisc(𝔸V)ρ(𝔸V)=0\sigma_{{\rm disc}}(-{\mathbb{A}}_{V})\sqcup\rho(-\mathbb{A}_{V})=\mathbb{C}\setminus\mathbb{R}_{\leq 0}.

  2. (2)

    The same statement with 𝔸V\mathbb{A}_{V} replaced by 𝔸V,n\mathbb{A}_{V,n} holds for nn\in\mathbb{Z}.

  3. (3)

    σdisc(𝔸V)=nσdisc(𝔸V,n)\sigma_{{\rm disc}}(-{\mathbb{A}}_{V})=\bigcup_{n\in\mathbb{Z}}\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,n}) and ρ(𝔸V)=nρ(𝔸V,n)\rho(-{\mathbb{A}}_{V})=\bigcap_{n\in\mathbb{Z}}\rho(-{\mathbb{A}}_{V,n}).

Proof.

(1) The fact that σ(𝔸)=0\sigma(-{\mathbb{A}})=\mathbb{R}_{\leq 0} is well-known and essentially due to Ladyzhenskaya [31]. Based on this fact, one can prove that σess(𝔸)=σ(𝔸)\sigma_{{\rm ess}}(-{\mathbb{A}})=\sigma(-{\mathbb{A}}) by showing the non-existence of eigenvalues in a similar manner as in [13, Lemma 2.6], or by using the property of the index ind(λ+𝔸)\operatorname{ind}(\lambda+\mathbb{A}) as is done in [33, Proof of Proposition 2.12]. Because of the regularity and decay of VV, the operator 𝔸V+𝔸=Vrot-{\mathbb{A}}_{V}+{\mathbb{A}}=-{\mathbb{P}}V^{\bot}\operatorname{rot} is relatively compact with respect to 𝔸-{\mathbb{A}}. The proof is quite similar to the one in [33, Section 2.4] for the case δ=0\delta=0 and thus we omit the details. Hence, from [28, Chapter IV\mathrm{I}\mathrm{V}, Theorem 5.35], we see that 𝔸V-{\mathbb{A}}_{V} and 𝔸-{\mathbb{A}} have the same essential spectrum. This implies the first statement.

For the second statement, we first observe that the equality

ind(λ+𝔸V)=ind(λ+𝔸)=0,λρsf(𝔸V)=0\operatorname{ind}(\lambda+\mathbb{A}_{V})=\operatorname{ind}(\lambda+\mathbb{A})=0,\quad\lambda\in\rho_{{\rm sf}}(-\mathbb{A}_{V})=\mathbb{C}\setminus\mathbb{R}_{\leq 0}

holds by [28, Chapter IV\mathrm{I}\mathrm{V}, Theorems 5.26 and 5.35]. Hence, since 0\mathbb{C}\setminus\mathbb{R}_{\leq 0} has only one component, by the argument in Subsection 3.1, we only need to prove that dimN(λ+𝔸V)=0\operatorname{dim}N(\lambda+\mathbb{A}_{V})=0 for at least one point λ0\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}. For this purpose, we consider

Θ(𝔸V)={𝔸Vv,v|vD(𝔸V) with vL2=1},\Theta(-\mathbb{A}_{V})=\{\langle-\mathbb{A}_{V}v,v\rangle~{}|~{}\text{$v\in D(\mathbb{A}_{V})$ with $\|v\|_{L^{2}}=1$}\},

which is called the numerical range of 𝔸V-\mathbb{A}_{V}; see [28, Chapter V\mathrm{V} §3 2].

Let vD(𝔸V)v\in D(\mathbb{A}_{V}). From the relation

𝔸Vv,v=vL22Vrotv,v,\langle-\mathbb{A}_{V}v,v\rangle=-\|\nabla v\|_{L^{2}}^{2}-\langle V^{\bot}\operatorname{rot}v,v\rangle,

we have

|𝔸Vv,v|+𝔸Vv,vvL22+2|Vrotv,v|.\displaystyle|\Im\langle-\mathbb{A}_{V}v,v\rangle|+\Re\langle-\mathbb{A}_{V}v,v\rangle\leq-\|\nabla v\|_{L^{2}}^{2}+2|\langle V^{\bot}\operatorname{rot}v,v\rangle|. (3.1)

Now let vD(𝔸V)v\in D(\mathbb{A}_{V}) and vL2=1\|v\|_{L^{2}}=1. The term 2|Vrotv,v|2|\langle V^{\bot}\operatorname{rot}v,v\rangle| is estimated as

2|Vrotv,v|\displaystyle 2|\langle V^{\bot}\operatorname{rot}v,v\rangle| 2(|α|+|δ|)vL2rotvL2\displaystyle\leq 2(|\alpha|+|\delta|)\|v\|_{L^{2}}\|\operatorname{rot}v\|_{L^{2}}
=2(|α|+|δ|)vL2\displaystyle=2(|\alpha|+|\delta|)\|\nabla v\|_{L^{2}}
(|α|+|δ|)2+vL22.\displaystyle\leq(|\alpha|+|\delta|)^{2}+\|\nabla v\|_{L^{2}}^{2}.

We have used rotvL2=vL2\|\operatorname{rot}v\|_{L^{2}}=\|\nabla v\|_{L^{2}} for vW01,2(Ω)2Lσ2(Ω)v\in W^{1,2}_{0}(\Omega)^{2}\cap L^{2}_{\sigma}(\Omega) in the second line and the Young inequality in the third line. Hence we obtain

|𝔸Vv,v|+𝔸Vv,v(|α|+|δ|)20,|\Im\langle-\mathbb{A}_{V}v,v\rangle|+\Re\langle-\mathbb{A}_{V}v,v\rangle-(|\alpha|+|\delta|)^{2}\leq 0,

which leads to the inclusion

Θ(𝔸V)¯{λ||λ|+λ(|α|+|δ|)20}.\overline{\Theta(-{\mathbb{A}}_{V})}\subset\{\lambda\in\mathbb{C}~{}|~{}|\Im\lambda|+\Re\lambda-(|\alpha|+|\delta|)^{2}\leq 0\}.

From [28, Chapter V\mathrm{V}, Theorem 3.2], we know that dimN(λ+𝔸V)=0\operatorname{dim}N(\lambda+\mathbb{A}_{V})=0 for any λ\lambda belonging to the complement of the right-hand side

{λ||λ|+λ(|α|+|δ|)2>0}.\{\lambda\in\mathbb{C}~{}|~{}|\Im\lambda|+\Re\lambda-(|\alpha|+|\delta|)^{2}>0\}.

This set is obviously a subset of 0\mathbb{C}\setminus\mathbb{R}_{\leq 0} and thus the second statement follows.

(2) The fact that σ(𝔸n)=0\sigma(-\mathbb{A}_{n})=\mathbb{R}_{\leq 0} can be proved in a similar manner as in [13, Lemma 3.3], and σess(𝔸n)=σ(𝔸n)\sigma_{{\rm ess}}(-{\mathbb{A}}_{n})=\sigma(-{\mathbb{A}}_{n}) follows by the property of ind(λ+𝔸n)\operatorname{ind}(\lambda+\mathbb{A}_{n}). Hence the first statement σess(𝔸n)=0\sigma_{{\rm ess}}(-{\mathbb{A}}_{n})=\mathbb{R}_{\leq 0} follows from the relative compactness of 𝔸V,n+𝔸n-{\mathbb{A}}_{V,n}+{\mathbb{A}}_{n} with respect to 𝔸n-{\mathbb{A}}_{n}. The second statement σdisc(𝔸V,n)ρ(𝔸V,n)=0\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,n})\sqcup\rho(-\mathbb{A}_{V,n})=\mathbb{C}\setminus\mathbb{R}_{\leq 0} can be deduced from the same discussion as above with 𝔸V\mathbb{A}_{V} replaced by 𝔸V,n\mathbb{A}_{V,n}.

(3) It suffices to prove the first statement σdisc(𝔸V)=nσdisc(𝔸V,n)\sigma_{{\rm disc}}(-{\mathbb{A}}_{V})=\bigcup_{n\in\mathbb{Z}}\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,n}). If λσdisc(𝔸V)\lambda\in\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}), there is a nonzero vD(𝔸V)v\in D({\mathbb{A}}_{V}) such that (λ+𝔸V)v=0(\lambda+{\mathbb{A}}_{V})v=0. Choosing nn\in\mathbb{Z} such that vn=𝒫nv0v_{n}=\mathcal{P}_{n}v\neq 0, we have vnD(𝔸V,n)v_{n}\in D({\mathbb{A}}_{V,n}) and (λ+𝔸V,n)vn=0(\lambda+{\mathbb{A}}_{V,n})v_{n}=0. Then we see that λσdisc(𝔸V,n)\lambda\in\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,n}) and hence λnσdisc(𝔸V,n)\lambda\in\bigcup_{n\in\mathbb{Z}}\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,n}). Oppositely, if λnσdisc(𝔸V,n)\lambda\in\bigcup_{n\in\mathbb{Z}}\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,n}), then there are nn\in\mathbb{Z} and nonzero vnD(𝔸V,n)v_{n}\in D({\mathbb{A}}_{V,n}) such that (λ+𝔸V,n)vn=0(\lambda+{\mathbb{A}}_{V,n})v_{n}=0. Then we have vnD(𝔸V)v_{n}\in D({\mathbb{A}}_{V}) and (λ+𝔸V)vn=0(\lambda+{\mathbb{A}}_{V})v_{n}=0 and hence λσdisc(𝔸V)\lambda\in\sigma_{{\rm disc}}(-{\mathbb{A}}_{V}). The proof is complete. ∎

The estimate of the numerical range Θ(𝔸V)\Theta(-\mathbb{A}_{V}) in the proof of Proposition 3.1 is quite rough. We consider its refinement in Lemma 3.2 to prove Proposition 3.3 below.

Lemma 3.2

Let α,δ\alpha,\delta\in\mathbb{R}. For vW01,2(Ω)2Lσ2(Ω)v\in W^{1,2}_{0}(\Omega)^{2}\cap L^{2}_{\sigma}(\Omega), we have

|Vrotv,v||α|||n|=1Urotvn,vn|+|α|vL22+|δ|vL22.\displaystyle|\langle V^{\bot}\operatorname{rot}v,v\rangle|\leq|\alpha|\Big{|}\sum_{|n|=1}\langle U^{\bot}\operatorname{rot}v_{n},v_{n}\rangle\Big{|}+|\alpha|\|\nabla v_{\neq}\|_{L^{2}}^{2}+|\delta|\|\nabla v\|_{L^{2}}^{2}. (3.2)

Moreover, for any T>0T>0,

||n|=1Urotvn,vn|2h(T)(|n|=1vnL22)+14T2h(T)(|n|=1vnL22).\displaystyle\Big{|}\sum_{|n|=1}\langle U^{\bot}\operatorname{rot}v_{n},v_{n}\rangle\Big{|}\leq 2h(T)\Big{(}\sum_{|n|=1}\|\nabla v_{n}\|_{L^{2}}^{2}\Big{)}+\frac{1}{4T^{2}h(T)}\Big{(}\sum_{|n|=1}\|v_{n}\|_{L^{2}}^{2}\Big{)}. (3.3)

Here the function h=h(T)h=h(T) is defined by

h(T)=0T1τe1τdτ,T>0,h(T)=\int_{0}^{T}\frac{1}{\tau}e^{-\frac{1}{\tau}}\,{\rm d}\tau,\quad T>0,

which satisfies

e1logTh(T)logT,T>e.\displaystyle e^{-1}\log T\leq h(T)\leq\log T,\quad T>e. (3.4)
Proof.

By the definition of VV, we see that

Vrotv,v=αUrotv,vδWrotv,v.\langle V^{\bot}\operatorname{rot}v,v\rangle=\alpha\langle U^{\bot}\operatorname{rot}v,v\rangle-\delta\langle W^{\bot}\operatorname{rot}v,v\rangle.

The Fourier series expansion leads to

Urotv,v=|n|=1Urotvn,vn+Urotv,v.\langle U^{\bot}\operatorname{rot}v,v\rangle=\sum_{|n|=1}\langle U^{\bot}\operatorname{rot}v_{n},v_{n}\rangle+\langle U^{\bot}\operatorname{rot}v_{\neq},v_{\neq}\rangle.

We have

|Urotv,v|\displaystyle|\langle U^{\bot}\operatorname{rot}v_{\neq},v_{\neq}\rangle| (r,θ)v(r,θ)rL2rotvL2\displaystyle\leq\Big{\|}(r,\theta)\mapsto\frac{v_{\neq}(r,\theta)}{r}\Big{\|}_{L^{2}}\|\operatorname{rot}v_{\neq}\|_{L^{2}}
vL22,\displaystyle\leq\|\nabla v_{\neq}\|_{L^{2}}^{2},

where the Hardy-type inequality (2.7) and rotuL2=uL2\|\operatorname{rot}u\|_{L^{2}}=\|\nabla u\|_{L^{2}} for uW01,2(Ω)2Lσ2(Ω)u\in W^{1,2}_{0}(\Omega)^{2}\cap L^{2}_{\sigma}(\Omega) are applied. Also, from (1.4) and (1.7) in the introduction,

|Wrotv,v|vL22.|\langle W^{\bot}\operatorname{rot}v,v\rangle|\leq\|\nabla v\|_{L^{2}}^{2}.

Combining all the estimates so far, we obtain (3.2).

Next let |n|=1|n|=1. We compute

|Urotvn,vn|02π11r|(rotvn)n(r)||vr,n(r)|rdrdθ.|\langle U^{\bot}\operatorname{rot}v_{n},v_{n}\rangle|\leq\int_{0}^{2\pi}\int_{1}^{\infty}\frac{1}{r}|(\operatorname{rot}v_{n})_{n}(r)||v_{r,n}(r)|r\,{\rm d}r\,{\rm d}\theta.

As is shown in [33, Proof of Lemma 3.26], we have

02π11r|(rotvn)n(r)||vr,n(r)|rdrdθ\displaystyle\int_{0}^{2\pi}\int_{1}^{\infty}\frac{1}{r}|(\operatorname{rot}v_{n})_{n}(r)||v_{r,n}(r)|r\,{\rm d}r\,{\rm d}\theta
h(T)vnL22+1TvnL2vnL2,T>0.\displaystyle\leq h(T)\|\nabla v_{n}\|_{L^{2}}^{2}+\frac{1}{T}\|v_{n}\|_{L^{2}}\ \|\nabla v_{n}\|_{L^{2}},\quad T>0.

The Young inequality yields

1TvnL2vnL2h(T)vnL22+14T2h(T)vnL22.\frac{1}{T}\|v_{n}\|_{L^{2}}\ \|\nabla v_{n}\|_{L^{2}}\leq h(T)\|\nabla v_{n}\|_{L^{2}}^{2}+\frac{1}{4T^{2}h(T)}\|v_{n}\|_{L^{2}}^{2}.

These estimates imply (3.3) after summation. The proof is complete. ∎

Proposition 3.3

Let α,δ\alpha,\delta\in\mathbb{R} be sufficiently small. We have the following.

  1. (1)

    The set

    Σ34π+4eα2e14|α|={λ||λ|+λ4eα2e14|α|>0}\Sigma_{\frac{3}{4}\pi}+4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}=\Big{\{}\lambda\in\mathbb{C}~{}\Big{|}~{}|\Im\lambda|+\Re\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}>0\Big{\}}

    is contained in ρ(𝔸V)\rho(-\mathbb{A}_{V}).

  2. (2)

    The same statement with 𝔸V\mathbb{A}_{V} replaced by 𝔸V,n\mathbb{A}_{V,n} holds for each |n|=1|n|=1.

  3. (3)

    The set Σ34π\Sigma_{\frac{3}{4}\pi} is contained in ρ(𝔸V,n)\rho(-\mathbb{A}_{V,n}) for each |n|1|n|\neq 1.

Proof.

Let us consider the numerical range as in the proof of Proposition 3.1.

(1) We first estimate 2|Vrotv,v|2|\langle V^{\bot}\operatorname{rot}v,v\rangle|. Let vD(𝔸V)v\in D(\mathbb{A}_{V}). Using Lemma 3.2, we have

2|Vrotv,v|\displaystyle 2|\langle V^{\bot}\operatorname{rot}v,v\rangle| 2|α|(2h(T)vL22+14T2h(T)vL22)\displaystyle\leq 2|\alpha|\bigg{(}2h(T)\|\nabla v\|_{L^{2}}^{2}+\frac{1}{4T^{2}h(T)}\|v\|_{L^{2}}^{2}\bigg{)}
+2(|α|+|δ|)vL22\displaystyle\quad+2(|\alpha|+|\delta|)\|\nabla v\|_{L^{2}}^{2}
2(2|α|h(T)+|α|+|δ|)vL22+|α|2T2h(T)vL22.\displaystyle\leq 2(2|\alpha|h(T)+|\alpha|+|\delta|)\|\nabla v\|_{L^{2}}^{2}+\frac{|\alpha|}{2T^{2}h(T)}\|v\|_{L^{2}}^{2}.

Let us choose T=e18|α|T=e^{\frac{1}{8|\alpha|}}. From (3.4), we see that

2(2|α|h(T)+|α|+|δ|)12+2(|α|+|δ|)2(2|\alpha|h(T)+|\alpha|+|\delta|)\leq\frac{1}{2}+2(|\alpha|+|\delta|)

and that

|α|2T2h(T)e|α|2T2logT=4eα2e14|α|.\frac{|\alpha|}{2T^{2}h(T)}\leq\frac{e|\alpha|}{2T^{2}\log T}=4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}.

Hence we obtain

2|Vrotv,v|{12+2(|α|+|δ|)}vL22+4eα2e14|α|vL22.\displaystyle 2|\langle V^{\bot}\operatorname{rot}v,v\rangle|\leq\Big{\{}\frac{1}{2}+2(|\alpha|+|\delta|)\Big{\}}\|\nabla v\|_{L^{2}}^{2}+4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\|v\|_{L^{2}}^{2}. (3.5)

Note that up to this point the smallness of α,δ\alpha,\delta is not needed.

Now let vD(𝔸V)v\in D(\mathbb{A}_{V}) and vL2=1\|v\|_{L^{2}}=1. From (3.1) and (3.5), we have

|𝔸Vv,v|+𝔸Vv,v4eα2e14|α|{12+2(|α|+|δ|)}vL22.\displaystyle\begin{split}&|\Im\langle-\mathbb{A}_{V}v,v\rangle|+\Re\langle-\mathbb{A}_{V}v,v\rangle-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\\ &\leq\Big{\{}-\frac{1}{2}+2(|\alpha|+|\delta|)\Big{\}}\|\nabla v\|_{L^{2}}^{2}.\end{split}

Therefore, for sufficiently small α,δ\alpha,\delta, we obtain the inclusion

Θ(𝔸V)¯{λ||λ|+λ4eα2e14|α|0}.\overline{\Theta(-{\mathbb{A}}_{V})}\subset\{\lambda\in\mathbb{C}~{}|~{}|\Im\lambda|+\Re\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\leq 0\}.

Then the statement follows from the same argument as in the proof of Proposition 3.1 (2).

(2) A similar proof as above leads to the statement.

(3) Let vD(𝔸V,n)v\in D(\mathbb{A}_{V,n}) with vL2=1\|v\|_{L^{2}}=1. Using Lemma 3.2, we estimate

|𝔸V,nv,v|+𝔸V,nv,v\displaystyle|\Im\langle-\mathbb{A}_{V,n}v,v\rangle|+\Re\langle-\mathbb{A}_{V,n}v,v\rangle vL22+2|Vrotv,v|\displaystyle\leq-\|\nabla v\|_{L^{2}}^{2}+2|\langle V^{\bot}\operatorname{rot}v,v\rangle|
{1+2(|α|+|δ|)}vL22.\displaystyle\leq\{-1+2(|\alpha|+|\delta|)\}\|\nabla v\|_{L^{2}}^{2}.

Thus the statement follows. The proof of Proposition 3.3 is complete. ∎

3.3 Analysis by explicit computation

Proposition 3.3 does not provide information on the discrete spectrum of 𝔸V-{\mathbb{A}}_{V} near the origin. This is a consequence of the fact that the Hardy inequality fails to hold in two-dimensional exterior domains. Therefore, we investigate the homogeneous equation of (Rn) by a more explicit computation, exploiting the symmetry of the exterior disk Ω\Omega.

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

ξn=ξn(α,δ)=[{n2+(δ2)2}12+iαn]12\displaystyle\xi_{n}=\xi_{n}(\alpha,\delta)=\bigg{[}\Big{\{}n^{2}+\Big{(}\frac{\delta}{2}\Big{)}^{2}\Big{\}}^{\frac{1}{2}}+i\alpha n\bigg{]}^{\frac{1}{2}} (3.6)

and

Fn(λ)=Fn(λ;α,δ)=1s|n|+1δ2Kξn(λs)ds,λ0.\displaystyle F_{n}(\sqrt{\lambda})=F_{n}(\sqrt{\lambda};\alpha,\delta)=\int_{1}^{\infty}s^{-|n|+1-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}s)\,{\rm d}s,\quad\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}. (3.7)
Proposition 3.4

Let α,δ\alpha,\delta\in\mathbb{R}. We have the following.

  1. (1)

    σdisc(𝔸V,0)=\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,0})=\emptyset.

  2. (2)

    σdisc(𝔸V,n)={λ0|Fn(λ)=0}\sigma_{{\rm disc}}(-{\mathbb{A}}_{V,n})=\{\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}~{}|~{}F_{n}(\sqrt{\lambda})=0\} for |n|1|n|\geq 1.

Proof.

(1) Let λ0\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}. In view of Proposition 3.1 (2) and ind(λ+𝔸V,0)=0\operatorname{ind}(\lambda+\mathbb{A}_{V,0})=0, we will show that the equation (λ+𝔸V,0)v0=0(\lambda+{\mathbb{A}}_{V,0})v_{0}=0 has only the trivial solution in D(𝔸V,0)D({\mathbb{A}}_{V,0}). Put n=0n=0 in (LABEL:eq.polar.vr)–(2.12) with f0=0f_{0}=0. The conditions in (2.12) imply that vr,0(r)=0v_{r,0}(r)=0 and hence that v0=vθ,0(r)𝐞θv_{0}=v_{\theta,0}(r){\bf e}_{\theta}. From (LABEL:eq.polar.vtheta)–(2.12), we see that vθ,0(r)v_{\theta,0}(r) satisfies

d2vθ,0dr21+δrdvθ,0dr+(λ+1δr2)vθ,0=0,r>1,vθ,0(1)=0.-\frac{\,{\rm d}^{2}v_{\theta,0}}{\,{\rm d}r^{2}}-\frac{1+\delta}{r}\frac{\,{\rm d}v_{\theta,0}}{\,{\rm d}r}+\Big{(}\lambda+\frac{1-\delta}{r^{2}}\Big{)}v_{\theta,0}=0,\quad r>1,\qquad v_{\theta,0}(1)=0.

By summability, the solution is given by, with some constant c0c_{0},

vθ,0(r)=c0rδ2K|1δ2|(λr).v_{\theta,0}(r)=c_{0}r^{-\frac{\delta}{2}}K_{|1-\frac{\delta}{2}|}(\sqrt{\lambda}r).

Then the boundary condition leads to c0=0c_{0}=0 since Kν()K_{\nu}(\cdot) has no zeros in Σπ2\Sigma_{\frac{\pi}{2}} if ν0\nu\geq 0; see [40, Chapter XV\mathrm{X}\mathrm{V} 15\cdot7]. Hence we obtain that v0=vθ,0(r)𝐞θ=0v_{0}=v_{\theta,0}(r){\bf e}_{\theta}=0, which is to be shown.

(2) Let λ0\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}. In view of Proposition 3.1 (2) and ind(λ+𝔸V,n)=0\operatorname{ind}(\lambda+\mathbb{A}_{V,n})=0, we will show that the equation (λ+𝔸V,n)vn=0(\lambda+{\mathbb{A}}_{V,n})v_{n}=0 admits a nontrivial solution in D(𝔸V,n)D({\mathbb{A}}_{V,n}) if and only if Fn(λ)=0F_{n}(\sqrt{\lambda})=0. Let vnD(𝔸V,n)v_{n}\in D({\mathbb{A}}_{V,n}) be nontrivial and solve (λ+𝔸V,n)vn=0(\lambda+{\mathbb{A}}_{V,n})v_{n}=0. Notice that vnv_{n} is smooth by the elliptic regularity of the Stokes system. Setting ωn(r)=(rotvn)n(r)\omega_{n}(r)=(\operatorname{rot}v_{n})_{n}(r), we see that ωn\omega_{n} satisfies the homogeneous equation of (2.13). Its linearly independent solutions are (B.1) in Appendix B. By the summability of vnv_{n}, we must have, with some constant cnc_{n},

ωn(r)=cnrδ2Kζn(λr).\omega_{n}(r)=c_{n}r^{-\frac{\delta}{2}}K_{\zeta_{n}}(\sqrt{\lambda}r).

Since ωn(r)\omega_{n}(r) decays exponentially as rr\to\infty, Proposition 2.1 leads to that

vn=𝒱n[ωn]=cn𝒱n[rrδ2Kξn(λr)]anddn[ωn]=cndn[rrδ2Kξn(λr)]=0,\displaystyle\begin{split}&v_{n}=\mathcal{V}_{n}[\omega_{n}]=c_{n}\mathcal{V}_{n}\big{[}r\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\big{]}\\ &\text{and}\qquad d_{n}[\omega_{n}]=c_{n}d_{n}\big{[}r\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\big{]}=0,\end{split}

with the notations in (2.15)–(2.16). The former condition implies that cnc_{n} is nonzero since vnv_{n} is assumed to be nontrivial. The latter one can be written equivalently to

cnFn(λ)=0.c_{n}F_{n}(\sqrt{\lambda})=0.

Thus we have that Fn(λ)=0F_{n}(\sqrt{\lambda})=0 since cn0c_{n}\neq 0. This completes the proof of the only if part.

For the if part, let Fn(λ)=0F_{n}(\sqrt{\lambda})=0. Then, for any nonzero cnc_{n}, the vector field

vn=cn𝒱n[rrδ2Kξn(λr)]v_{n}=c_{n}\mathcal{V}_{n}\big{[}r\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\big{]}

gives a nontrivial solution of (λ+𝔸V,n)vn=0(\lambda+{\mathbb{A}}_{V,n})v_{n}=0. Indeed, from the proof of the only if part, we ensure that vnv_{n} is smooth and belongs to D(𝔸V,n)D({\mathbb{A}}_{V,n}). Note that the no-slip condition vn|Ω=0v_{n}{|_{\partial\Omega}}=0 is verified by the assumption that Fn(λ)=0F_{n}(\sqrt{\lambda})=0. Moreover, setting

fn=λvnΔvn+Vrotvn,f_{n}=\lambda v_{n}-\Delta v_{n}+V^{\bot}\operatorname{rot}v_{n},

from (2.17), we see that

rotfn=λ(rotvn)Δ(rotvn)+V(rotvn)=0.\operatorname{rot}f_{n}=\lambda(\operatorname{rot}v_{n})-\Delta(\operatorname{rot}v_{n})+V\cdot\nabla(\operatorname{rot}v_{n})=0.

Thus Proposition 2.2 yields that there is a function pW^1,2(Ω)p\in\widehat{W}^{1,2}(\Omega) such that fn=pf_{n}=-\nabla p. Operating the Helmholtz projection {\mathbb{P}} to this equality, we find that (λ+𝔸V,n)vn=0(\lambda+{\mathbb{A}}_{V,n})v_{n}=0. This completes the proof of the if part. The proof of Proposition 3.4 is complete. ∎

The following is a corollary of Propositions 3.1 (3), 3.3 (3) and 3.4.

Corollary 3.5

Let α,δ\alpha,\delta\in\mathbb{R} be sufficiently small. We have

σdisc(𝔸V)Σ34π=|n|=1{λΣ34π|Fn(λ)=0}.\sigma_{{\rm disc}}(-{\mathbb{A}}_{V})\cap\Sigma_{\frac{3}{4}\pi}=\bigcup_{|n|=1}\Big{\{}\lambda\in\Sigma_{\frac{3}{4}\pi}~{}\Big{|}~{}F_{n}(\sqrt{\lambda})=0\Big{\}}.

4 Quantitative analysis of discrete spectrum

In this section, keeping Corollary 3.5 in mind, we analyze zeros of the analytic function Fn(λ)F_{n}(\sqrt{\lambda}) with |n|=1|n|=1 defined in (3.7). Thanks to Proposition 3.3, it suffices to consider the zeros in disks centered at the origin with radius exponentially small in |α||\alpha|. The main result is Proposition 4.7. The proof is based on asymptotic analysis under the smallness of α,δ\alpha,\delta.

Note that one can recover the results in [33, 22] by putting δ=0\delta=0 in the statements of this section. However, this observation is not useful in the proof since we need to describe precisely the zeros of functions having multiple parameters. A continuity argument is not enough and quantitative analysis is needed. In fact, it is revealed that situations are different depending on the sign of δ\delta, and that the case δ<0\delta<0 seems to be more delicate.

4.1 Expansion of the order

When |n|=1|n|=1, we denote

1δ={1+(δ2)2}12,ηn=ξn1.\displaystyle 1_{\delta}=\Big{\{}1+\Big{(}\frac{\delta}{2}\Big{)}^{2}\Big{\}}^{\frac{1}{2}},\qquad\eta_{n}=\xi_{n}-1. (4.1)

Here ξn\xi_{n} is defined in (3.6). A direct computation shows that

(ξn)=1δ2[{1+(α1δ2)2}12+1]12,(ξn)=sgn(αn)1δ2[{1+(α1δ2)2}121]12.\displaystyle\begin{split}\Re(\xi_{n})&=\frac{1_{\delta}}{\sqrt{2}}\bigg{[}\Big{\{}1+\Big{(}\frac{\alpha}{1_{\delta}^{2}}\Big{)}^{2}\Big{\}}^{\frac{1}{2}}+1\bigg{]}^{\frac{1}{2}},\\ \Im(\xi_{n})&=\operatorname{sgn}(\alpha n)\frac{1_{\delta}}{\sqrt{2}}\bigg{[}\Big{\{}1+\Big{(}\frac{\alpha}{1_{\delta}^{2}}\Big{)}^{2}\Big{\}}^{\frac{1}{2}}-1\bigg{]}^{\frac{1}{2}}.\end{split}

We need the following expansion of ηn\eta_{n} in the next subsection.

Lemma 4.1

Let |n|=1|n|=1. For sufficiently small α,δ\alpha,\delta\in\mathbb{R}, we have

(ηn)\displaystyle\Re(\eta_{n}) =α2+δ28+O(α4+δ4),\displaystyle=\frac{\alpha^{2}+\delta^{2}}{8}+O(\alpha^{4}+\delta^{4}), (4.2)
(ηn)\displaystyle\Im(\eta_{n}) =sgn(αn)|α|2+O(|α|(α2+δ2)).\displaystyle=\operatorname{sgn}(\alpha n)\frac{|\alpha|}{2}+O\big{(}|\alpha|(\alpha^{2}+\delta^{2})\big{)}. (4.3)

All the implicit constants in O()O(\cdot) are independent of α,δ\alpha,\delta.

Proof.

The proof is done by the Taylor theorem. For (4.2), from

(ηn)=(ξn)1=1δ{1+18(α1δ2)25128(α1δ2)4+O(α6)}1,\Re(\eta_{n})=\Re(\xi_{n})-1=1_{\delta}\bigg{\{}1+\frac{1}{8}\Big{(}\frac{\alpha}{1_{\delta}^{2}}\Big{)}^{2}-\frac{5}{128}\Big{(}\frac{\alpha}{1_{\delta}^{2}}\Big{)}^{4}+O(\alpha^{6})\bigg{\}}-1,

we see that

(ηn)=1δ1+α28+α28(11δ31)5α41285α4128(11δ71)+O(α6).\Re(\eta_{n})=1_{\delta}-1+\frac{\alpha^{2}}{8}+\frac{\alpha^{2}}{8}\Big{(}\frac{1}{1_{\delta}^{3}}-1\Big{)}-\frac{5\alpha^{4}}{128}-\frac{5\alpha^{4}}{128}\Big{(}\frac{1}{1_{\delta}^{7}}-1\Big{)}+O(\alpha^{6}).

Hence (4.2) is obtained by

1δ1=δ28δ4128+O(δ6)\displaystyle 1_{\delta}-1=\frac{\delta^{2}}{8}-\frac{\delta^{4}}{128}+O(\delta^{6})

and

11δ31=3δ28+O(δ4),11δ71=7δ28+O(δ4).\displaystyle\frac{1}{1_{\delta}^{3}}-1=-\frac{3\delta^{2}}{8}+O(\delta^{4}),\qquad\frac{1}{1_{\delta}^{7}}-1=-\frac{7\delta^{2}}{8}+O(\delta^{4}).

For (4.3), from

(ηn)=(ξn)=sgn(αn)1δ{12|α1δ2|116|α1δ2|3+O(|α|5)},\Im(\eta_{n})=\Im(\xi_{n})=\operatorname{sgn}(\alpha n)1_{\delta}\bigg{\{}\frac{1}{2}\Big{|}\frac{\alpha}{1_{\delta}^{2}}\Big{|}-\frac{1}{16}\Big{|}\frac{\alpha}{1_{\delta}^{2}}\Big{|}^{3}+O(|\alpha|^{5})\bigg{\}},

we see that

(ηn)=sgn(αn){|α|2+|α|2(11δ1)|α|316|α|316(11δ51)+O(|α|5)}.\Im(\eta_{n})=\operatorname{sgn}(\alpha n)\bigg{\{}\frac{|\alpha|}{2}+\frac{|\alpha|}{2}\Big{(}\frac{1}{1_{\delta}}-1\Big{)}-\frac{|\alpha|^{3}}{16}-\frac{|\alpha|^{3}}{16}\Big{(}\frac{1}{1_{\delta}^{5}}-1\Big{)}+O(|\alpha|^{5})\bigg{\}}.

Hence (4.3) is obtained by

11δ1=δ28+O(δ4),11δ51=5δ28+O(δ4).\displaystyle\frac{1}{1_{\delta}}-1=-\frac{\delta^{2}}{8}+O(\delta^{4}),\qquad\frac{1}{1_{\delta}^{5}}-1=-\frac{5\delta^{2}}{8}+O(\delta^{4}).

This completes the proof. ∎

4.2 Asymptotic analysis

We consider Fn(λ)F_{n}(\sqrt{\lambda}) in (3.7) with |n|=1|n|=1, namely, the function

Fn(z)=1sδ2K1+ηn(zs)ds,zΣπ2.\displaystyle F_{n}(z)=\int_{1}^{\infty}s^{-\frac{\delta}{2}}K_{1+\eta_{n}}(zs)\,{\rm d}s,\quad z\in\Sigma_{\frac{\pi}{2}}. (4.4)
Lemma 4.2

Let |n|=1|n|=1. For α,δ\alpha,\delta\in\mathbb{R}, we have

(δ2+ηn)Fn(z)=K1+ηn(z)z1s1δ2Kηn(zs)ds,zΣπ2.\displaystyle\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}F_{n}(z)=K_{1+\eta_{n}}(z)-z\int_{1}^{\infty}s^{1-\frac{\delta}{2}}K_{\eta_{n}}(zs)\,{\rm d}s,\quad z\in\Sigma_{\frac{\pi}{2}}. (4.5)
Proof.

By the recurrence relation (see [40, Chapter III\mathrm{I}\mathrm{I}\mathrm{I} 3\cdot71 (3)])

μKμ(z)=zdKμdz(z)zKμ1(z),\mu K_{\mu}(z)=-z\frac{\,{\rm d}K_{\mu}}{\,{\rm d}z}(z)-zK_{\mu-1}(z),

we have

(1+ηn)K1+ηn(zs)=sddsK1+ηn(zs)zsKηn(zs).(1+\eta_{n})K_{1+\eta_{n}}(zs)=-s\frac{\,{\rm d}}{\,{\rm d}s}K_{1+\eta_{n}}(zs)-zsK_{\eta_{n}}(zs).

Thus the definition (4.4) and integration by parts give

(1+ηn)Fn(z)=1sδ2(1+ηn)K1+ηn(zs)ds=1sδ2(sddsK1+ηn(zs)zsKηn(zs))ds=K1+ηn(z)+(1δ2)Fn(z)z1s1δ2Kηn(zs)ds,\displaystyle\begin{split}(1+\eta_{n})F_{n}(z)&=\int_{1}^{\infty}s^{-\frac{\delta}{2}}(1+\eta_{n})K_{1+\eta_{n}}(zs)\,{\rm d}s\\ &=\int_{1}^{\infty}s^{-\frac{\delta}{2}}\Big{(}-s\frac{\,{\rm d}}{\,{\rm d}s}K_{1+\eta_{n}}(zs)-zsK_{\eta_{n}}(zs)\Big{)}\,{\rm d}s\\ &=K_{1+\eta_{n}}(z)+\Big{(}1-\frac{\delta}{2}\Big{)}F_{n}(z)-z\int_{1}^{\infty}s^{1-\frac{\delta}{2}}K_{\eta_{n}}(zs)\,{\rm d}s,\end{split}

which implies the assertion of the lemma. ∎

Using the relation (4.5), we investigate zeros of Fn(z)F_{n}(z) near the origin. We perform asymptotic analysis when |z||z| is sufficiently small. Since the asymptotics of K1+ηn(z)K_{1+\eta_{n}}(z) is already obtained in Lemma A.2 (1), we focus on the second term on the right-hand side of (4.5). In what follows in this section, we assume smallness of α,δ\alpha,\delta. Although some estimates can be proved under weaker assumptions, we will not give the details for simplicity.

Lemma 4.3

Let |n|=1|n|=1. For sufficiently small α,δ\alpha,\delta\in\mathbb{R}, we have

z1s1δ2Kηn(zs)ds\displaystyle z\int_{1}^{\infty}s^{1-\frac{\delta}{2}}K_{\eta_{n}}(zs)\,{\rm d}s =Δ(δ,ηn)2(z2)1+δ2+Rn(2)(z),zΣπ2{|z|<1}.\displaystyle=\frac{\Delta(\delta,\eta_{n})}{2}\Big{(}\frac{z}{2}\Big{)}^{-1+\frac{\delta}{2}}+R^{(2)}_{n}(z),\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}. (4.6)

Here Δ(δ,ηn)\Delta(\delta,\eta_{n}) is defined by

Δ(δ,ηn)=Γ(1δ4ηn2)Γ(1δ4+ηn2),\displaystyle\Delta(\delta,\eta_{n})=\Gamma\Big{(}1-\frac{\delta}{4}-\frac{\eta_{n}}{2}\Big{)}\Gamma\Big{(}1-\frac{\delta}{4}+\frac{\eta_{n}}{2}\Big{)}, (4.7)

where Γ(z)\Gamma(z) is the Gamma function, and Rn(2)(z)R^{(2)}_{n}(z) is the remainder and satisfies

|Rn(2)(z)|C|z|1ηn(1+|log|z||),zΣπ2{|z|<1}.\displaystyle|R^{(2)}_{n}(z)|\leq C|z|^{1-\Re\eta_{n}}\big{(}1+\big{|}\log|z|\big{|}\big{)},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}. (4.8)

The constant CC is independent of α,δ\alpha,\delta.

Proof.

If we show that

z0s1δ2Kηn(zs)ds=Δ(δ,ηn)2(z2)1+δ2,\displaystyle\begin{split}z\int_{0}^{\infty}s^{1-\frac{\delta}{2}}K_{\eta_{n}}(zs)\,{\rm d}s=\frac{\Delta(\delta,\eta_{n})}{2}\Big{(}\frac{z}{2}\Big{)}^{-1+\frac{\delta}{2}},\end{split} (4.9)

the assertion follows. Indeed, it is not hard to check that

Rn(2)(z):=z01s1δ2Kηn(zs)dsR^{(2)}_{n}(z):=-z\int_{0}^{1}s^{1-\frac{\delta}{2}}K_{\eta_{n}}(zs)\,{\rm d}s

satisfies (4.8) using the estimates in Lemma A.2 (2).

By the representation (A.6) and by the Fubini theorem, we have

z0s1δ2Kηn(zs)ds=z0s1δ2(120ezs2(t+1t)tηn1dt)ds=z20t1ηn(0s1δ2ez2(t+1t)sds)dt.\displaystyle\begin{split}&z\int_{0}^{\infty}s^{1-\frac{\delta}{2}}K_{\eta_{n}}(zs)\,{\rm d}s\\ &=z\int_{0}^{\infty}s^{1-\frac{\delta}{2}}\bigg{(}\frac{1}{2}\int_{0}^{\infty}e^{-\frac{zs}{2}(t+\frac{1}{t})}t^{-\eta_{n}-1}\,{\rm d}t\bigg{)}\,{\rm d}s\\ &=\frac{z}{2}\int_{0}^{\infty}t^{-1-\eta_{n}}\bigg{(}\int_{0}^{\infty}s^{1-\frac{\delta}{2}}e^{-\frac{z}{2}(t+\frac{1}{t})s}\,{\rm d}s\bigg{)}\,{\rm d}t.\end{split}

Observing that

0s1δ2easds=Γ(2δ2)a2+δ2,aΣπ2,\displaystyle\int_{0}^{\infty}s^{1-\frac{\delta}{2}}e^{-as}\,{\rm d}s=\Gamma\Big{(}2-\frac{\delta}{2}\Big{)}a^{-2+\frac{\delta}{2}},\quad a\in\Sigma_{\frac{\pi}{2}},

we have

z20t1ηn(0s1δ2ez2(t+1t)sds)dt=Γ(2δ2)(z2)1+δ20t1ηn(t+1t)2+δ2dt=Γ(2δ2)(z2)1+δ20t1δ2ηn(t2+1)2δ2dt.\displaystyle\begin{split}&\frac{z}{2}\int_{0}^{\infty}t^{-1-\eta_{n}}\bigg{(}\int_{0}^{\infty}s^{1-\frac{\delta}{2}}e^{-\frac{z}{2}(t+\frac{1}{t})s}\,{\rm d}s\bigg{)}\,{\rm d}t\\ &=\Gamma\Big{(}2-\frac{\delta}{2}\Big{)}\Big{(}\frac{z}{2}\Big{)}^{-1+\frac{\delta}{2}}\int_{0}^{\infty}t^{-1-\eta_{n}}\Big{(}t+\frac{1}{t}\Big{)}^{-2+\frac{\delta}{2}}\,{\rm d}t\\ &=\Gamma\Big{(}2-\frac{\delta}{2}\Big{)}\Big{(}\frac{z}{2}\Big{)}^{-1+\frac{\delta}{2}}\int_{0}^{\infty}\frac{t^{1-\frac{\delta}{2}-\eta_{n}}}{(t^{2}+1)^{2-\frac{\delta}{2}}}\,{\rm d}t.\end{split}

The change of variable t=τ12t=\tau^{\frac{1}{2}} leads to

0t1δ2ηn(t2+1)2δ2dt=120τδ4ηn2(τ+1)2δ2dτ=12B(1δ4ηn2,1δ4+ηn2),\displaystyle\begin{split}\int_{0}^{\infty}\frac{t^{1-\frac{\delta}{2}-\eta_{n}}}{(t^{2}+1)^{2-\frac{\delta}{2}}}\,{\rm d}t&=\frac{1}{2}\int_{0}^{\infty}\frac{\tau^{-\frac{\delta}{4}-\frac{\eta_{n}}{2}}}{(\tau+1)^{2-\frac{\delta}{2}}}\,{\rm d}\tau\\ &=\frac{1}{2}B\Big{(}1-\frac{\delta}{4}-\frac{\eta_{n}}{2},1-\frac{\delta}{4}+\frac{\eta_{n}}{2}\Big{)},\end{split}

where B(p,q)B(p,q) is the Beta function. Then the well-known formulas

zΓ(z)=Γ(z+1),B(p,q)=Γ(p)Γ(q)Γ(p+q)\displaystyle z\Gamma(z)=\Gamma(z+1),\qquad B(p,q)=\frac{\Gamma(p)\Gamma(q)}{\Gamma(p+q)}

imply (4.9). This completes the proof. ∎

Corollary 4.4

Let |n|=1|n|=1. For sufficiently small α,δ\alpha,\delta\in\mathbb{R}, we have

(δ2+ηn)Fn(z)=Γ(1+ηn)2(z2)1ηnΔ(δ,ηn)2(z2)1+δ2+Rn(3)(z),zΣπ2{|z|<1}.\displaystyle\begin{split}\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}F_{n}(z)&=\frac{\Gamma(1+\eta_{n})}{2}\Big{(}\frac{z}{2}\Big{)}^{-1-\eta_{n}}-\frac{\Delta(\delta,\eta_{n})}{2}\Big{(}\frac{z}{2}\Big{)}^{-1+\frac{\delta}{2}}\\ &\quad+R^{(3)}_{n}(z),\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}.\end{split} (4.10)

Here Rn(3)R^{(3)}_{n} is the remainder and satisfies

|Rn(3)(z)|C|z|1ηn(1+|log|z||),zΣπ2{|z|<1}.\displaystyle|R^{(3)}_{n}(z)|\leq C|z|^{1-\Re\eta_{n}}\big{(}1+\big{|}\log|z|\big{|}\big{)},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}. (4.11)

The constant CC is independent of α,δ\alpha,\delta.

Proof.

This is a consequence of the previous proposition and Lemma A.2 (1). ∎

Proposition 4.7 below, giving a lower bound of |Fn(z)||F_{n}(z)|, is proved based on the expansion (4.10). In the proof, we need precise estimates of the coefficients appearing in (4.10).

Lemma 4.5

Let |n|=1|n|=1. For sufficiently small α,δ\alpha,\delta\in\mathbb{R}, we have

LogΓ(1+ηn)\displaystyle\operatorname{Log}\Gamma(1+\eta_{n}) =γηn+O(|ηn|2),\displaystyle=-\gamma\eta_{n}+O(|\eta_{n}|^{2}), (4.12)
LogΔ(δ,ηn)\displaystyle\operatorname{Log}\Delta(\delta,\eta_{n}) =γ(δ2)+O((δ2)2+|ηn|2),\displaystyle=\gamma\Big{(}\frac{\delta}{2}\Big{)}+O\Big{(}\Big{(}\frac{\delta}{2}\Big{)}^{2}+|\eta_{n}|^{2}\Big{)}, (4.13)

where γ=0.5772\gamma=0.5772\ldots is the Euler constant. Moreover, if δ0\delta\geq 0,

LogΔ(δ,ηn)LogΓ(1+ηn)\displaystyle\operatorname{Log}\Delta(\delta,\eta_{n})-\operatorname{Log}\Gamma(1+\eta_{n}) =γ(δ2+ηn)+O(|δ2+ηn|2).\displaystyle=\gamma\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}+O\Big{(}\Big{|}\frac{\delta}{2}+\eta_{n}\Big{|}^{2}\Big{)}. (4.14)

All the implicit constants in O()O(\cdot) are independent of α,δ\alpha,\delta.

Proof.

We may apply the Taylor series expansion of LogΓ(1+z)\operatorname{Log}\Gamma(1+z)

LogΓ(1+z)=γ(z)+k=2ζ(k)k(z)k,{|z|<1}.\displaystyle\operatorname{Log}\Gamma(1+z)=\gamma(-z)+\sum_{k=2}^{\infty}\frac{\zeta(k)}{k}(-z)^{k},\quad\{|z|<1\}. (4.15)

Here ζ(k)=m=1mk\zeta(k)=\sum_{m=1}^{\infty}m^{-k} is the Riemann zeta function. One can prove (4.15) using

zΓ(z)=Γ(z+1),1Γ(z)=zeγzm=1(1+zm)ezm.\displaystyle z\Gamma(z)=\Gamma(z+1),\qquad\frac{1}{\Gamma(z)}=ze^{\gamma z}\prod_{m=1}^{\infty}\Big{(}1+\frac{z}{m}\Big{)}e^{-\frac{z}{m}}.

Indeed, from

LogΓ(1+z)\displaystyle\operatorname{Log}\Gamma(1+z) =LogzLog1Γ(z)\displaystyle=\operatorname{Log}z-\operatorname{Log}\frac{1}{\Gamma(z)}
=γ(z)m=1(Log(1+zm)zm)\displaystyle=\gamma(-z)-\sum_{m=1}^{\infty}\bigg{(}\operatorname{Log}\Big{(}1+\frac{z}{m}\Big{)}-\frac{z}{m}\bigg{)}

and the Taylor series expansion

Log(1+z)=k=11k(z)k,{|z|<1},\displaystyle\operatorname{Log}(1+z)=-\sum_{k=1}^{\infty}\frac{1}{k}(-z)^{k},\quad\{|z|<1\},

we see that

LogΓ(1+z)=γ(z)+m=1k=21k(zm)k,\displaystyle\operatorname{Log}\Gamma(1+z)=\gamma(-z)+\sum_{m=1}^{\infty}\sum_{k=2}^{\infty}\frac{1}{k}\Big{(}-\frac{z}{m}\Big{)}^{k},

which leads to (4.15) after change of order of summations.

The expansion (4.12) is a direct consequence of (4.15). Also, by

LogΓ(1δ4ηn2)\displaystyle\operatorname{Log}\Gamma\Big{(}1-\frac{\delta}{4}-\frac{\eta_{n}}{2}\Big{)} =γ2(δ2+ηn)+O(|δ2+ηn|2),\displaystyle=\frac{\gamma}{2}\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}+O\Big{(}\Big{|}\frac{\delta}{2}+\eta_{n}\Big{|}^{2}\Big{)},
LogΓ(1δ4+ηn2)\displaystyle\operatorname{Log}\Gamma\Big{(}1-\frac{\delta}{4}+\frac{\eta_{n}}{2}\Big{)} =γ2(δ2ηn)+O(|δ2ηn|2)\displaystyle=\frac{\gamma}{2}\Big{(}\frac{\delta}{2}-\eta_{n}\Big{)}+O\Big{(}\Big{|}\frac{\delta}{2}-\eta_{n}\Big{|}^{2}\Big{)}

and the definition of Δ(δ,ηn)\Delta(\delta,\eta_{n}) in (4.7), we have

LogΔ(δ,ηn)\displaystyle\operatorname{Log}\Delta(\delta,\eta_{n}) =LogΓ(1δ4ηn2)+LogΓ(1δ4+ηn2)\displaystyle=\operatorname{Log}\Gamma\Big{(}1-\frac{\delta}{4}-\frac{\eta_{n}}{2}\Big{)}+\operatorname{Log}\Gamma\Big{(}1-\frac{\delta}{4}+\frac{\eta_{n}}{2}\Big{)}
=γ(δ2)+O(|δ2+ηn|2+|δ2ηn|2),\displaystyle=\gamma\Big{(}\frac{\delta}{2}\Big{)}+O\Big{(}\Big{|}\frac{\delta}{2}+\eta_{n}\Big{|}^{2}+\Big{|}\frac{\delta}{2}-\eta_{n}\Big{|}^{2}\Big{)},

which implies (4.13). If δ0\delta\geq 0, we see from Lemma 4.1 that, for sufficiently small α,δ\alpha,\delta,

(δ2)2+|ηn|2=|δ2+ηn|2δηn|δ2+ηn|2,\displaystyle\Big{(}\frac{\delta}{2}\Big{)}^{2}+|\eta_{n}|^{2}=\Big{|}\frac{\delta}{2}+\eta_{n}\Big{|}^{2}-\delta\Re\eta_{n}\leq\Big{|}\frac{\delta}{2}+\eta_{n}\Big{|}^{2},

which implies (4.14). This completes the proof. ∎

The following is the key technical lemma in the proof of Proposition 4.7 below.

Lemma 4.6

Let ϵ(0,π2)\epsilon\in(0,\frac{\pi}{2}). Suppose that ζ\zeta\in\mathbb{C} with |ζ|1|\zeta|\ll 1 satisfies

ζ>0,|ζ|>0\displaystyle\Re\zeta>0,\qquad|\Im\zeta|>0 (4.16)

and

{ζ+(1+κ)(ζ)2ζ}(π2ϵ)<π\displaystyle\Big{\{}\Re\zeta+(1+\kappa)\frac{(\Im\zeta)^{2}}{\Re\zeta}\Big{\}}\Big{(}\frac{\pi}{2}-\epsilon\Big{)}<\pi (4.17)

with some constant κ=κ(ϵ)(0,12)\kappa=\kappa(\epsilon)\in(0,\frac{1}{2}) independent of ζ\zeta. Then, by defining

K(ζ)=min{{(ζ)2|ζ|+|ζ|},ζ},\displaystyle K(\zeta)=\min\bigg{\{}\Big{\{}\frac{(\Re\zeta)^{2}}{|\Im\zeta|}+|\Im\zeta|\Big{\}},\Re\zeta\bigg{\}}, (4.18)

one has

|1wζ|Cmin{1,K(ζ)|log|w||},wΣπ2ϵ{|z|<1}.\displaystyle|1-w^{\zeta}|\geq C\min\Big{\{}1,K(\zeta)\big{|}\log|w|\big{|}\Big{\}},\quad w\in\Sigma_{\frac{\pi}{2}-\epsilon}\cap\{|z|<1\}. (4.19)

The constant CC depends only on ϵ\epsilon and κ\kappa.

Proof.

By setting

μ\displaystyle\mu =(ζ)log|w|(ζ)argw,\displaystyle=(\Re\zeta)\log|w|-(\Im\zeta)\operatorname{arg}w, (4.20)
θ\displaystyle\theta =(ζ)log|w|+(ζ)argw,\displaystyle=(\Im\zeta)\log|w|+(\Re\zeta)\operatorname{arg}w, (4.21)

we denote

1wζ=1eμeiθ.1-w^{\zeta}=1-e^{\mu}e^{i\theta}.

From

log|w|=ζζargw+1ζμ,\log|w|=\frac{\Im\zeta}{\Re\zeta}\operatorname{arg}w+\frac{1}{\Re\zeta}\mu,

we compute

θ={ζ+(ζ)2ζ}argw+ζζμ.\displaystyle\theta=\Big{\{}\Re\zeta+\frac{(\Im\zeta)^{2}}{\Re\zeta}\Big{\}}\operatorname{arg}w+\frac{\Im\zeta}{\Re\zeta}\mu. (4.22)

Before going into details, let us explain the difficulties. When μ\mu is close to zero, one essentially needs to provide lower bounds of |1eiθ||1-e^{i\theta}|. However, such bounds require good control of θ\theta, since 1eiθ1-e^{i\theta} vanishes when θ=2mπ\theta=2m\pi with mm\in\mathbb{Z}. The reason why the conditions (4.16)–(4.17) are needed is to control the range of θ\theta when μ\mu is close to zero.

We will consider two cases:

(i) Case |μ|κ|ζ||argw||\mu|\leq\kappa|\Im\zeta||\operatorname{arg}w|. In this case, we have

12eμ32.\frac{1}{2}\leq e^{\mu}\leq\frac{3}{2}.

In addition, by (4.22) and the assumption (4.17),

|θ|{ζ+(1+κ)(ζ)2ζ}|argw|<π,wΣπ2ϵ{|z|<1}.|\theta|\leq\Big{\{}\Re\zeta+(1+\kappa)\frac{(\Im\zeta)^{2}}{\Re\zeta}\Big{\}}|\operatorname{arg}w|<\pi,\quad w\in\Sigma_{\frac{\pi}{2}-\epsilon}\cap\{|z|<1\}.

Thus eiθe^{i\theta} is equal to 11 if and only if θ=0\theta=0. If 0|θ|<π20\leq|\theta|<\frac{\pi}{2}, the imaginary part gives

|1eμeiθ|eμ|sinθ|eμ2π|θ|1π|θ|.\displaystyle|1-e^{\mu}e^{i\theta}|\geq e^{\mu}|\sin\theta|\geq e^{\mu}\frac{2}{\pi}|\theta|\geq\frac{1}{\pi}|\theta|. (4.23)

If π2|θ|<π\frac{\pi}{2}\leq|\theta|<\pi, the real part gives

|1eμeiθ||1eμcosθ|1>1π|θ|.\displaystyle|1-e^{\mu}e^{i\theta}|\geq|1-e^{\mu}\cos\theta|\geq 1>\frac{1}{\pi}|\theta|. (4.24)

Hence we estimate |θ||\theta|. Combining |μ|κ|ζ||argw||\mu|\leq\kappa|\Im\zeta||\operatorname{arg}w| with (4.22), we have

|θ|{ζ+(1κ)(ζ)2ζ}|argw|.|\theta|\geq\Big{\{}\Re\zeta+(1-\kappa)\frac{(\Im\zeta)^{2}}{\Re\zeta}\Big{\}}|\operatorname{arg}w|.

Combining with (4.20),

ζ|log|w|||ζ||argw|+|μ|(1+κ)|ζ||argw|.\displaystyle\Re\zeta\big{|}\log|w|\big{|}\leq|\Im\zeta||\operatorname{arg}w|+|\mu|\leq(1+\kappa)|\Im\zeta||\operatorname{arg}w|.

By these two estimates, we obtain

|θ|{ζ+(1κ)(ζ)2ζ}ζ|log|w||(1+κ)|ζ|1κ1+κ{(ζ)2|ζ|+|ζ|}|log|w||.\displaystyle\begin{split}|\theta|&\geq\Big{\{}\Re\zeta+(1-\kappa)\frac{(\Im\zeta)^{2}}{\Re\zeta}\Big{\}}\frac{\Re\zeta\big{|}\log|w|\big{|}}{(1+\kappa)|\Im\zeta|}\\ &\geq\frac{1-\kappa}{1+\kappa}\Big{\{}\frac{(\Re\zeta)^{2}}{|\Im\zeta|}+|\Im\zeta|\Big{\}}\big{|}\log|w|\big{|}.\end{split}

Therefore, from (4.23) and (4.24), we see that

|1eμeiθ|1π1κ1+κ{(ζ)2|ζ|+|ζ|}|log|w||.\displaystyle|1-e^{\mu}e^{i\theta}|\geq\frac{1}{\pi}\frac{1-\kappa}{1+\kappa}\Big{\{}\frac{(\Re\zeta)^{2}}{|\Im\zeta|}+|\Im\zeta|\Big{\}}\big{|}\log|w|\big{|}. (4.25)

(ii) Case |μ|>κ|ζ||argw||\mu|>\kappa|\Im\zeta||\operatorname{arg}w|. In this case, we may rely on

|1eμeiθ||1eμ|e1min{1,|μ|},μ,θ.\displaystyle|1-e^{\mu}e^{i\theta}|\geq|1-e^{\mu}|\geq e^{-1}\min\{1,|\mu|\},\quad\mu,\theta\in\mathbb{R}. (4.26)

We deduce that if |argw|>12ζ|ζ||log|w|||\operatorname{arg}w|>\frac{1}{2}\frac{\Re\zeta}{|\Im\zeta|}\big{|}\log|w|\big{|},

|μ|>κ2ζ|log|w|||\mu|>\frac{\kappa}{2}\Re\zeta\big{|}\log|w|\big{|}

by |μ|>κ|ζ||argw||\mu|>\kappa|\Im\zeta||\operatorname{arg}w|, and that if |argw|12ζ|ζ||log|w|||\operatorname{arg}w|\leq\frac{1}{2}\frac{\Re\zeta}{|\Im\zeta|}\big{|}\log|w|\big{|},

|μ|ζ|log|w|||ζ||argw|12ζ|log|w||.|\mu|\geq\Re\zeta\big{|}\log|w|\big{|}-|\Im\zeta||\operatorname{arg}w|\geq\frac{1}{2}\Re\zeta\big{|}\log|w|\big{|}.

by (4.20). Combining these two with (4.26), we obtain

|1eμeiθ|e1min{1,κ2ζ|log|w||}.\displaystyle|1-e^{\mu}e^{i\theta}|\geq e^{-1}\min\Big{\{}1,\frac{\kappa}{2}\Re\zeta\big{|}\log|w|\big{|}\Big{\}}. (4.27)

The assertion follows from (4.25) and (4.27). The proof is complete. ∎

Proposition 4.7

Let |n|=1|n|=1 and ϵ(0,π2)\epsilon\in(0,\frac{\pi}{2}). Let K(ζ)K(\zeta) be defined in (4.18). For sufficiently small (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}, we have

|(δ2+ηn)Fn(z)|C|z|1ηnmin{1,K(δ2+ηn)|log|z||},zΣπ2ϵ{|z|<K(δ2+ηn)}.\displaystyle\begin{split}\Big{|}\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}F_{n}(z)\Big{|}\geq C|z|^{-1-\Re\eta_{n}}\min\Big{\{}1,K\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}\big{|}\log|z|\big{|}\Big{\}},&\\ z\in\Sigma_{\frac{\pi}{2}-\epsilon}\cap\Big{\{}|z|<K\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}\Big{\}}.&\end{split} (4.28)

The constant CC depends only on ϵ\epsilon.

Remark 4.8

One observes a sort of stabilizing effect by the flow δW\delta W from this proposition. By the definition (4.18) and Lemma 4.1, we have a simple (but rough) estimate from below

K(δ2+ηn)min{|(δ2+ηn)|,(δ2+ηn)}18min{|α|,δ+α2}.\displaystyle\begin{split}K\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}&\geq\min\Big{\{}\Big{|}\Im\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}\Big{|},\Re\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}\Big{\}}\\ &\geq\frac{1}{8}\min\{|\alpha|,\delta+\alpha^{2}\}.\end{split} (4.29)

The second inequality is valid for sufficiently small α,δ\alpha,\delta. Therefore, the radius of the disks on which Fn(z)F_{n}(z) has no zeros is greater than that for δ=0\delta=0. This is interpreted as a stabilizing effect by δW\delta W in time frequency near zero related to large-time behavior of flows.

Proof.

Let zΣπ2ϵ{|z|<12}z\in\Sigma_{\frac{\pi}{2}-\epsilon}\cap\{|z|<\frac{1}{2}\} first. Using Corollary 4.4, we write

(δ2+ηn)Fn(z)\displaystyle\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}F_{n}(z) =Γ(1+ηn)2(z2)1ηn{1Δ(δ,ηn)Γ(1+ηn)(z2)δ2+ηn+Rn(z)}.\displaystyle=\frac{\Gamma(1+\eta_{n})}{2}\Big{(}\frac{z}{2}\Big{)}^{-1-\eta_{n}}\Big{\{}1-\frac{\Delta(\delta,\eta_{n})}{\Gamma(1+\eta_{n})}\Big{(}\frac{z}{2}\Big{)}^{\frac{\delta}{2}+\eta_{n}}+R_{n}(z)\Big{\}}. (4.30)

Here

Rn(z)=2Γ(1+ηn)(z2)1+ηnRn(3)(z)R_{n}(z)=\frac{2}{\Gamma(1+\eta_{n})}\Big{(}\frac{z}{2}\Big{)}^{1+\eta_{n}}R^{(3)}_{n}(z)

is the remainder and satisfies

|Rn(z)|C|z|2|log|z||,zΣπ2{|z|<12}.\displaystyle|R_{n}(z)|\leq C|z|^{2}\big{|}\log|z|\big{|},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\Big{\{}|z|<\frac{1}{2}\Big{\}}. (4.31)

The condition δ0\delta\geq 0 and Lemma 4.5 imply

Δ(δ,ηn)Γ(1+ηn)=eLogΔ(δ,ηn)LogΓ(1+ηn)=eγn(δ2+ηn),\frac{\Delta(\delta,\eta_{n})}{\Gamma(1+\eta_{n})}=e^{\operatorname{Log}\Delta(\delta,\eta_{n})-\operatorname{Log}\Gamma(1+\eta_{n})}=e^{\gamma_{n}(\frac{\delta}{2}+\eta_{n})},

where γn=γn(δ,ηn)\gamma_{n}=\gamma_{n}(\delta,\eta_{n}) is the function satisfying

γn=γ+O(|δ2+ηn|).\gamma_{n}=\gamma+O\Big{(}\Big{|}\frac{\delta}{2}+\eta_{n}\Big{|}\Big{)}.

Setting

ζn=δ2+ηn,wn=eγn2z,\zeta_{n}=\frac{\delta}{2}+\eta_{n},\qquad w_{n}=\frac{e^{\gamma_{n}}}{2}z,

we will derive a lower bound of

1Δ(δ,ηn)Γ(1+ηn)(z2)δ2+ηn=1wnζn.1-\frac{\Delta(\delta,\eta_{n})}{\Gamma(1+\eta_{n})}\Big{(}\frac{z}{2}\Big{)}^{\frac{\delta}{2}+\eta_{n}}=1-w_{n}^{\zeta_{n}}.

To apply Lemma 4.6, we check that all the conditions are fulfilled by ζn,wn\zeta_{n},w_{n}. We have

|wn|1,|argwn|π2ϵ2|w_{n}|\leq 1,\qquad|\operatorname{arg}w_{n}|\leq\frac{\pi}{2}-\frac{\epsilon}{2}

for sufficiently small α,δ\alpha,\delta. We also have (4.16) by Lemma 4.1. By the same lemma, there are constants C1,C2C_{1},C_{2} independent of α,δ\alpha,\delta such that

(ζn)2ζn{|α|2+C1|α|(α2+δ2)}2{α2+δ28C2(α4+δ4)}1.\frac{(\Im\zeta_{n})^{2}}{\Re\zeta_{n}}\leq\Big{\{}\frac{|\alpha|}{2}+C_{1}|\alpha|(\alpha^{2}+\delta^{2})\Big{\}}^{2}\Big{\{}\frac{\alpha^{2}+\delta^{2}}{8}-C_{2}(\alpha^{4}+\delta^{4})\Big{\}}^{-1}.

Thus, for sufficiently small α,δ\alpha,\delta, we have

(ζn)2ζn=2+O(α2+δ2)\frac{(\Im\zeta_{n})^{2}}{\Re\zeta_{n}}=2+O(\alpha^{2}+\delta^{2})

and, with a constant κn=κn(ϵ)(0,12)\kappa_{n}=\kappa_{n}(\epsilon)\in(0,\frac{1}{2}) independent of α,δ\alpha,\delta,

{ζn+(1+κn)(ζn)2ζn}(π2ϵ2)<π,\Big{\{}\Re\zeta_{n}+(1+\kappa_{n})\frac{(\Im\zeta_{n})^{2}}{\Re\zeta_{n}}\Big{\}}\Big{(}\frac{\pi}{2}-\frac{\epsilon}{2}\Big{)}<\pi,

which is (4.17) with ϵ\epsilon replaced by ϵ2\frac{\epsilon}{2}. Now, applying Lemma 4.6, we see that

|1wnζn|\displaystyle|1-w_{n}^{\zeta_{n}}| Cmin{1,K(ζn)|log|wnζn||}\displaystyle\geq C\min\Big{\{}1,K(\zeta_{n})\big{|}\log|w_{n}^{\zeta_{n}}|\big{|}\Big{\}}
Cmin{1,K(ζn)|log|z||},zΣπ2ϵ{|z|<12},\displaystyle\geq C\min\Big{\{}1,K(\zeta_{n})\big{|}\log|z|\big{|}\Big{\}},\quad z\in\Sigma_{\frac{\pi}{2}-\epsilon}\cap\Big{\{}|z|<\frac{1}{2}\Big{\}},

for sufficiently small α,δ\alpha,\delta. The constant CC depends only on ϵ\epsilon.

Therefore, combining this estimate with (4.30) and (4.31), we obtain a lower bound

|ζnFn(z)|C|z|1ηn(min{1,K(ζn)|log|z||}|z|2|log|z||),\displaystyle\begin{split}|\zeta_{n}F_{n}(z)|\geq C|z|^{-1-\Re\eta_{n}}\Big{(}\min\Big{\{}1,K(\zeta_{n})\big{|}\log|z|\big{|}\Big{\}}-|z|^{2}\big{|}\log|z|\big{|}\Big{)},\end{split} (4.32)

which implies the desired estimate (4.28). The proof is complete. ∎

We state two corollaries to this proposition. The first one gives a simpler version of (4.28) useful for later calculation. The second one uses the results in Section 3.

Corollary 4.9

Let |n|=1|n|=1 and ϵ(0,π)\epsilon\in(0,\pi). For sufficiently small (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}, we have

|(δ2+ηn)Fn(λ)|C|λ|ξn2min{1,α2|log|λ||},\displaystyle\Big{|}\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}F_{n}(\sqrt{\lambda})\Big{|}\geq C|\lambda|^{-\frac{\Re\xi_{n}}{2}}\min\big{\{}1,\alpha^{2}\big{|}\log|\lambda|\big{|}\big{\}},
λΣπϵ{|z|<α4}.\displaystyle\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<\alpha^{4}\}.

In particular,

1|Fn(λ)|C|1α(δ2+ηn)|1|λ|ξn2,λΣπϵ{|z|<e14|α|}.\frac{1}{|F_{n}(\sqrt{\lambda})|}\leq C\Big{|}\frac{1}{\alpha}\Big{(}\frac{\delta}{2}+\eta_{n}\Big{)}\Big{|}^{-1}|\lambda|^{\frac{\Re\xi_{n}}{2}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\Big{\{}|z|<e^{-\frac{1}{4|\alpha|}}\Big{\}}.

The constant depends only on ϵ\epsilon.

Proof.

The assertion follows from (4.32) combined with the simple lower bound (4.29). ∎

Corollary 4.10

Let ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}). For sufficiently small (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}, we have

Σ34πϵρ(𝔸V).\Sigma_{\frac{3}{4}\pi-\epsilon}\subset\rho(-\mathbb{A}_{V}).
Proof.

In view of Proposition 3.3 and Corollary 4.9, we set

𝒮1(α)=(Σ34π+4eα2e14|α|){|z|>8eα2e14|α|},𝒮2(α)=Σ34π{|z|<e14|α|}.\displaystyle\begin{split}\mathcal{S}_{1}(\alpha)&=\Big{(}\Sigma_{\frac{3}{4}\pi}+4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{)}\cap\Big{\{}|z|>8e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{\}},\\ \mathcal{S}_{2}(\alpha)&=\Sigma_{\frac{3}{4}\pi}\cap\Big{\{}|z|<e^{-\frac{1}{4|\alpha|}}\Big{\}}.\end{split}

From Propositions 3.1 (1) and 3.3, and Corollary 3.5, we see that both 𝒮1(α)\mathcal{S}_{1}(\alpha) and 𝒮2(α)\mathcal{S}_{2}(\alpha) are contained in ρ(𝔸V)\rho(-\mathbb{A}_{V}) for sufficiently small α,δ\alpha,\delta. For a given ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}), by an easy geometric consideration, we find that Σ34πϵ\Sigma_{\frac{3}{4}\pi-\epsilon} is contained in 𝒮1(α)𝒮1(α)\mathcal{S}_{1}(\alpha)\cup\mathcal{S}_{1}(\alpha) if α\alpha is small enough depending on ϵ\epsilon. This implies the assertion. ∎

5 Resolvent estimate

In this section, we estimate the solutions of

(λ+𝔸V)v=f(\lambda+\mathbb{A}_{V})v=f (R)

for given λρ(𝔸V)\lambda\in\rho(-{\mathbb{A}}_{V}) and fLσ2(Ω)f\in L^{2}_{\sigma}(\Omega). The main result is the following.

Proposition 5.1

Let ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0} be sufficiently small. We have, for q(1.2]q\in(1.2] and fLσ2(Ω)Lq(Ω)2f\in L^{2}_{\sigma}(\Omega)\cap L^{q}(\Omega)^{2},

(λ+𝔸V)1fL2C|λ|32+1qfLq,λΣ34πϵ\begin{split}\|(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}\leq C|\lambda|^{-\frac{3}{2}+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\Sigma_{\frac{3}{4}\pi-\epsilon}\end{split} (5.1)

and, for fLσ2(Ω)f\in L^{2}_{\sigma}(\Omega),

(λ+𝔸V)1fL2C|λ|12fL2,λΣ34πϵ.\begin{split}\|\nabla(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}\leq C|\lambda|^{-\frac{1}{2}}\|f\|_{L^{2}},\quad\lambda\in\Sigma_{\frac{3}{4}\pi-\epsilon}.\end{split} (5.2)

The constant CC depends only on α,δ,ϵ,q\alpha,\delta,\epsilon,q.

Once Proposition 5.1 is proved, it is routine to prove Theorem 1.1 by representing {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0} in the Dunford integral of the resolvent. Thus the detail will be given in Appendix C.

We prove Proposition 5.1 in Subsection 5.3 by a combination of energy method and explicit formulas for the solution. Note that the estimate (5.1) cannot be obtained by energy method alone, due to the absence of the Hardy inequality. However, this is not the case when λ\lambda belongs to sectors shifted exponentially small in |α||\alpha|; see Proposition 5.3 for details. Therefore, all that remains is to prove the estimate when λ\lambda belongs to the intersection of sectors and the disks centered at the origin whose radius is exponentially small in |α||\alpha|. This proof is done by explicit formulas; see Proposition 5.4 for details.

5.1 Energy method

We start with a priori estimates for (R) using energy method.

Lemma 5.2

Let α,δ\alpha,\delta\in\mathbb{R}. For λ\lambda\in\mathbb{C}, q(1,2]q\in(1,2] and fLσ2(Ω)Lq(Ω)2f\in L^{2}_{\sigma}(\Omega)\cap L^{q}(\Omega)^{2}, suppose that there is a solution vD(𝔸V)v\in D(\mathbb{A}_{V}) of (R). Then we have the following.

  1. (1)

    For vnv_{n} with |n|=1|n|=1,

    (|λ|+λ4eα2e14|α|)vnL22+{142(|α|+|δ|)}vnL22CfLq2q3q2vnL24(q1)3q2.\displaystyle\begin{split}&\Big{(}|\Im\lambda|+\Re\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{)}\|v_{n}\|_{L^{2}}^{2}+\Big{\{}\frac{1}{4}-2(|\alpha|+|\delta|)\Big{\}}\|\nabla v_{n}\|_{L^{2}}^{2}\\ &\leq C\|f\|_{L^{q}}^{\frac{2q}{3q-2}}\|v_{n}\|_{L^{2}}^{\frac{4(q-1)}{3q-2}}.\end{split}
  2. (2)

    For v=v|n|=1vnv_{\neq}=v-\sum_{|n|=1}v_{n},

    (|λ|+λ)vL22+{342(|α|+|δ|)}vL22CfLq2q3q2vL24(q1)3q2.\displaystyle\begin{split}&(|\Im\lambda|+\Re\lambda)\|v_{\neq}\|_{L^{2}}^{2}+\Big{\{}\frac{3}{4}-2(|\alpha|+|\delta|)\Big{\}}\|\nabla v_{\neq}\|_{L^{2}}^{2}\\ &\leq C\|f\|_{L^{q}}^{\frac{2q}{3q-2}}\|v_{\neq}\|_{L^{2}}^{\frac{4(q-1)}{3q-2}}.\end{split}

The constant CC depends only on qq.

Proof.

(1) Taking the inner product of (R) with vnv_{n}, we see that

λvnL22𝔸Vvn,vn=f,vn\lambda\|v_{n}\|_{L^{2}}^{2}-\langle-\mathbb{A}_{V}v_{n},v_{n}\rangle=\langle f,v_{n}\rangle

and hence that

(|λ|+λ)vnL22|𝔸Vvn,vn|+𝔸Vvn,vn+2|f,vn|.\displaystyle\begin{split}(|\Im\lambda|+\Re\lambda)\|v_{n}\|_{L^{2}}^{2}\leq|\Im\langle-\mathbb{A}_{V}v_{n},v_{n}\rangle|+\Re\langle-\mathbb{A}_{V}v_{n},v_{n}\rangle+2|\langle f,v_{n}\rangle|.\end{split} (5.3)

From (3.1) and (3.5) in Section 3, we have

|𝔸Vvn,vn|+𝔸Vvn,vnvnL22+2|Vrotvn,vn|{12+2(|α|+|δ|)}vnL22+4eα2e14|α|vnL22.\displaystyle\begin{split}&|\Im\langle-\mathbb{A}_{V}v_{n},v_{n}\rangle|+\Re\langle-\mathbb{A}_{V}v_{n},v_{n}\rangle\\ &\leq-\|\nabla v_{n}\|_{L^{2}}^{2}+2|\langle V^{\bot}\operatorname{rot}v_{n},v_{n}\rangle|\\ &\leq\Big{\{}-\frac{1}{2}+2(|\alpha|+|\delta|)\Big{\}}\|\nabla v_{n}\|_{L^{2}}^{2}+4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\|v_{n}\|_{L^{2}}^{2}.\end{split} (5.4)

One has, by the Hölder and the Gagliardo-Nirenberg inequalities,

|f,u|fLquLqq1CfLquL22(11q)uL22q1CfLq2q3q2uL24(q1)3q2+18uL22,uW1,2(Ω)2.\displaystyle\begin{split}|\langle f,u\rangle|&\leq\|f\|_{L^{q}}\|u\|_{L^{\frac{q}{q-1}}}\\ &\leq C\|f\|_{L^{q}}\|u\|_{L^{2}}^{2(1-\frac{1}{q})}\|\nabla u\|_{L^{2}}^{\frac{2}{q}-1}\\ &\leq C\|f\|_{L^{q}}^{\frac{2q}{3q-2}}\|u\|_{L^{2}}^{\frac{4(q-1)}{3q-2}}+\frac{1}{8}\|\nabla u\|_{L^{2}}^{2},\quad u\in W^{1,2}(\Omega)^{2}.\end{split} (5.5)

The Young inequality is applied in the last line. The statement follows from (5.3)–(5.5).

(2) In a similar manner as above, we see that

(|λ|+λ)vL22|𝔸Vv,v|+𝔸Vv,v+2|f,v|.\displaystyle\begin{split}(|\Im\lambda|+\Re\lambda)\|v_{\neq}\|_{L^{2}}^{2}\leq|\Im\langle-\mathbb{A}_{V}v_{\neq},v_{\neq}\rangle|+\Re\langle-\mathbb{A}_{V}v_{\neq},v_{\neq}\rangle+2|\langle f,v_{\neq}\rangle|.\end{split} (5.6)

From (3.1) and (3.2) in Section 3, we have

|𝔸Vv,v|+𝔸Vv,vvL22+2|Vrotv,v|{1+2(|α|+|δ|)}vL22.\displaystyle\begin{split}&|\Im\langle-\mathbb{A}_{V}v_{\neq},v_{\neq}\rangle|+\Re\langle-\mathbb{A}_{V}v_{\neq},v_{\neq}\rangle\\ &\leq-\|\nabla v_{\neq}\|_{L^{2}}^{2}+2|\langle V^{\bot}\operatorname{rot}v_{\neq},v_{\neq}\rangle|\\ &\leq\{-1+2(|\alpha|+|\delta|)\}\|\nabla v_{\neq}\|_{L^{2}}^{2}.\end{split} (5.7)

The statement follows from (5.6)–(5.7) combined with (5.5). The proof is complete. ∎

Lemma 5.2 gives the following estimate of the resolvent.

Proposition 5.3

Let ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}) and let α,δ\alpha,\delta\in\mathbb{R} be sufficiently small. Set

𝒮1ϵ(α)=(Σ34πϵ+4eα2e14|α|){|z|>8eα2e14|α|}ρ(𝔸V).\mathcal{S}_{1}^{\epsilon}(\alpha)=\Big{(}\Sigma_{\frac{3}{4}\pi-\epsilon}+4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{)}\cap\Big{\{}|z|>8e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{\}}\subset\rho(-\mathbb{A}_{V}).

For q(1,2]q\in(1,2] and fLσ2(Ω)Lq(Ω)2f\in L^{2}_{\sigma}(\Omega)\cap L^{q}(\Omega)^{2}, we have

(λ+𝔸V)1fL2C|λ|32+1qfLq,λ𝒮1ϵ(α),(λ+𝔸V)1fL2C|λ|1+1qfLq,λ𝒮1ϵ(α).\begin{split}\|(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}&\leq C|\lambda|^{-\frac{3}{2}+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\mathcal{S}_{1}^{\epsilon}(\alpha),\\ \|\nabla(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}&\leq C|\lambda|^{-1+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\mathcal{S}_{1}^{\epsilon}(\alpha).\end{split} (5.8)

The constant CC depends only on ϵ,q\epsilon,q.

Proof.

Since 𝒮1ϵ(α)Σ34πϵρ(𝔸V)\mathcal{S}_{1}^{\epsilon}(\alpha)\subset\Sigma_{\frac{3}{4}\pi-\epsilon}\subset\rho(-\mathbb{A}_{V}) by Corollary 4.10, we see that (λ+𝔸V)1f(\lambda+{\mathbb{A}}_{V})^{-1}f exists for any λ𝒮1ϵ(α)\lambda\in\mathcal{S}_{1}^{\epsilon}(\alpha). Observe that, if λ𝒮1ϵ(α)\lambda\in\mathcal{S}_{1}^{\epsilon}(\alpha), we have both

|λ|+λ4eα2e14|α|=|(λ4eα2e14|α|)|+(λ4eα2e14|α|)C|λ4eα2e14|α||,\displaystyle\begin{split}|\Im\lambda|+\Re\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}&=\Big{|}\Im\Big{(}\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{)}\Big{|}+\Re\Big{(}\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{)}\\ &\geq C\Big{|}\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{|},\end{split}

with a constant C=C(ϵ)C=C(\epsilon), and

|λ4eα2e14|α||>|λ||λ|2=|λ|2.\displaystyle\begin{split}\Big{|}\lambda-4e\alpha^{2}e^{-\frac{1}{4|\alpha|}}\Big{|}>|\lambda|-\frac{|\lambda|}{2}=\frac{|\lambda|}{2}.\end{split}

Hence, under the smallness on α,δ\alpha,\delta, Lemma 5.2 gives

|λ|vnL22+vnL22CfLq2q3q2vnL24(q1)3q2,|n|=1,λ𝒮1ϵ(α),|λ|vL22+vL22CfLq2q3q2vL24(q1)3q2,λ𝒮1ϵ(α)\displaystyle\begin{split}|\lambda|\|v_{n}\|_{L^{2}}^{2}+\|\nabla v_{n}\|_{L^{2}}^{2}&\leq C\|f\|_{L^{q}}^{\frac{2q}{3q-2}}\|v_{n}\|_{L^{2}}^{\frac{4(q-1)}{3q-2}},\quad|n|=1,\quad\lambda\in\mathcal{S}_{1}^{\epsilon}(\alpha),\\ |\lambda|\|v_{\neq}\|_{L^{2}}^{2}+\|\nabla v_{\neq}\|_{L^{2}}^{2}&\leq C\|f\|_{L^{q}}^{\frac{2q}{3q-2}}\|v_{\neq}\|_{L^{2}}^{\frac{4(q-1)}{3q-2}},\quad\lambda\in\mathcal{S}_{1}^{\epsilon}(\alpha)\end{split}

for the solution of (R), namely, for v=(λ+𝔸V)1fv=(\lambda+{\mathbb{A}}_{V})^{-1}f. This implies the assertion (5.8). ∎

5.2 Explicit formulas

Energy method can not lead to Proposition 5.1 due to the absence of the Hardy inequality. Instead, we employ explicit formulas and prove the following proposition.

Proposition 5.4

Let |n|=1|n|=1 and ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0} be sufficiently small. Set

𝒮2ϵ(α)=Σ34πϵ{|z|<e14|α|}ρ(𝔸V).\mathcal{S}_{2}^{\epsilon}(\alpha)=\Sigma_{\frac{3}{4}\pi-\epsilon}\cap\Big{\{}|z|<e^{-\frac{1}{4|\alpha|}}\Big{\}}\subset\rho(-\mathbb{A}_{V}).

We have, for q(1.2]q\in(1.2] and fLσ2(Ω)Lq(Ω)2f\in L^{2}_{\sigma}(\Omega)\cap L^{q}(\Omega)^{2},

𝒫n(λ+𝔸V)1fL2C|λ|32+1qfLq,λ𝒮2ϵ(α)\begin{split}\|\mathcal{P}_{n}(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}\leq C|\lambda|^{-\frac{3}{2}+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha)\end{split} (5.9)

and, for fLσ2(Ω)f\in L^{2}_{\sigma}(\Omega),

𝒫n(λ+𝔸V)1fL2C|λ|12fL2,λ𝒮2ϵ(α).\begin{split}\|\nabla\mathcal{P}_{n}(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}\leq C|\lambda|^{-\frac{1}{2}}\|f\|_{L^{2}},\quad\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha).\end{split} (5.10)

The constant CC depends only on α,δ,ϵ,q\alpha,\delta,\epsilon,q.

The derivation of the formula is as follows. Let λρ(𝔸V)\lambda\in\rho(-{\mathbb{A}}_{V}) and assume first fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega) in (R). Then the solution v=(λ+𝔸V)1fv=(\lambda+{\mathbb{A}}_{V})^{-1}f is smooth in Ω\Omega thanks to the elliptic regularity of the Stokes system, and ωn(r):=(rotvn)n(r)\omega_{n}(r):=(\operatorname{rot}v_{n})_{n}(r) solves the equation (2.13) in Subsection 2.3. Since the linearly independent solutions of its homogeneous equation are (B.1) in Appendix B and the Wronskian is r1δr^{-1-\delta}, we see that ωn(r)\omega_{n}(r) is given by

ωn(r)=c~λ,n[fn]rδ2Kξn(λr)+Φλ,n[fn](r).\displaystyle\omega_{n}(r)=\tilde{c}_{\lambda,n}[f_{n}]r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)+\Phi_{\lambda,n}[f_{n}](r). (5.11)

The constant c~λ,n[fn]\tilde{c}_{\lambda,n}[f_{n}] is determined later and Φλ,n[fn]\Phi_{\lambda,n}[f_{n}] is defined by

Φλ,n[fn](r)\displaystyle\Phi_{\lambda,n}[f_{n}](r) =rδ2Kξn(λr)1rs1+δ2Iξn(λs)(rotfn)n(s)ds\displaystyle=r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\int_{1}^{r}s^{1+\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}s)(\operatorname{rot}f_{n})_{n}(s)\,{\rm d}s
+rδ2Iξn(λr)rs1+δ2Kξn(λs)(rotfn)n(s)ds.\displaystyle\quad+r^{-\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}r)\int_{r}^{\infty}s^{1+\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}s)(\operatorname{rot}f_{n})_{n}(s)\,{\rm d}s.

Using integration by parts and setting

gn(1)=(ξn+δ2)fθ,n+infr,n,gn(2)=(ξnδ2)fθ,ninfr,n,\displaystyle g^{(1)}_{n}=\Big{(}\xi_{n}+\frac{\delta}{2}\Big{)}f_{\theta,n}+inf_{r,n},\qquad g^{(2)}_{n}=\Big{(}\xi_{n}-\frac{\delta}{2}\Big{)}f_{\theta,n}-inf_{r,n}, (5.12)

we have

Φλ,n[fn](r)=rδ2Kξn(λr)1rsδ2Iξn(λs)gn(1)(s)dsλrδ2Kξn(λr)1rs1+δ2Iξn+1(λs)fθ,n(s)ds+rδ2Iξn(λr)rsδ2Kξn(λs)gn(2)(s)ds+λrδ2Iξn(λr)rs1+δ2Kξn1(λs)fθ,n(s)ds.\displaystyle\begin{split}\Phi_{\lambda,n}[f_{n}](r)&=-r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\int_{1}^{r}s^{\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}s)g^{(1)}_{n}(s)\,{\rm d}s\\ &\quad-\sqrt{\lambda}r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\int_{1}^{r}s^{1+\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}s)f_{\theta,n}(s)\,{\rm d}s\\ &\quad+r^{-\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}r)\int_{r}^{\infty}s^{\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}s)g^{(2)}_{n}(s)\,{\rm d}s\\ &\quad+\sqrt{\lambda}r^{-\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}r)\int_{r}^{\infty}s^{1+\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}s)f_{\theta,n}(s)\,{\rm d}s.\end{split} (5.13)

Since ωn(r)\omega_{n}(r) decays exponentially, we see from Proposition 2.1 that vnv_{n} is uniquely represented by the Biot–Savart law as, with the notations in (2.15)–(2.16),

vn=𝒱n[ωn]=c~λ,n[fn]𝒱n[rrδ2Kξn(λr)]+𝒱n[Φλ,n[fn]].\displaystyle\begin{split}v_{n}=\mathcal{V}_{n}[\omega_{n}]=\tilde{c}_{\lambda,n}[f_{n}]\mathcal{V}_{n}\big{[}r\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\big{]}+\mathcal{V}_{n}\big{[}\Phi_{\lambda,n}[f_{n}]\big{]}.\end{split} (5.14)

This formula is implemented with the constraint dn[ωn]=0d_{n}[\omega_{n}]=0, which we write

c~n,λ[fn]Fn(λ)+dn[Φn,λ[fn]]=0,\displaystyle\tilde{c}_{n,\lambda}[f_{n}]F_{n}(\sqrt{\lambda})+d_{n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]}=0, (5.15)

by using Fn(λ)F_{n}(\sqrt{\lambda}) in (3.7). This relation determines c~n,λ[fn]\tilde{c}_{n,\lambda}[f_{n}]. We set

cn,λ[fn]=dn[Φn,λ[fn]]=1s1|n|Φn,λ[fn](s)ds.\displaystyle c_{n,\lambda}[f_{n}]=d_{n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]}=\int_{1}^{\infty}s^{1-|n|}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s. (5.16)

Collecting (5.11)–(5.16), we find that

𝒫n(λ+𝔸V)1f=cn,λ[fn]Fn(λ)𝒱n[rrδ2Kξn(λr)]+𝒱n[Φn,λ[fn]]\displaystyle\begin{split}\mathcal{P}_{n}(\lambda+{\mathbb{A}}_{V})^{-1}f&=-\frac{c_{n,\lambda}[f_{n}]}{F_{n}(\sqrt{\lambda})}\mathcal{V}_{n}\big{[}r\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\big{]}+\mathcal{V}_{n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]}\end{split} (5.17)

and that

(rot𝒫n(λ+𝔸V)1f)(r,θ)=cn,λ[fn]Fn(λ)rδ2Kξn(λr)einθ+Φn,λ[fn]einθ.\displaystyle\begin{split}&\big{(}\operatorname{rot}\mathcal{P}_{n}(\lambda+{\mathbb{A}}_{V})^{-1}f\big{)}(r,\theta)=-\frac{c_{n,\lambda}[f_{n}]}{F_{n}(\sqrt{\lambda})}r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)e^{in\theta}+\Phi_{n,\lambda}[f_{n}]e^{in\theta}.\end{split} (5.18)

For general fLσ2(Ω)f\in L^{2}_{\sigma}(\Omega), one should understand the formulas (5.17)–(5.18) by density argument. This understanding is possible thanks to the estimates in Proposition 5.4. Note that the uniqueness of representation is guaranteed by Proposition 2.1.

Now we let |n|=1|n|=1 and estimate (5.17)–(5.18). Firstly we estimate

𝒱n[Φn,λ[fn]]=𝒱r,n[Φn,λ[fn]](r)einθ𝐞r+𝒱θ,n[Φn,λ[fn]](r)einθ𝐞θ\mathcal{V}_{n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]}=\mathcal{V}_{r,n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]}(r)e^{in\theta}{\bf e}_{r}+\mathcal{V}_{\theta,n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]}(r)e^{in\theta}{\bf e}_{\theta}

in (5.17), where

𝒱r,n[Φn,λ[fn]]\displaystyle\mathcal{V}_{r,n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]} =in2r(cn,λ[fn]r1r1rs2Φn,λ[fn](s)dsrrΦn,λ[fn](s)ds),\displaystyle=-\frac{in}{2r}\bigg{(}\frac{c_{n,\lambda}[f_{n}]}{r}-\frac{1}{r}\int_{1}^{r}s^{2}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s-r\int_{r}^{\infty}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s\bigg{)},
𝒱θ,n[Φn,λ[fn]]\displaystyle\mathcal{V}_{\theta,n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]} =12r(cn,λ[fn]r1r1rs2Φn,λ[fn](s)ds+rrΦn,λ[fn](s)ds).\displaystyle=\frac{1}{2r}\bigg{(}\frac{c_{n,\lambda}[f_{n}]}{r}-\frac{1}{r}\int_{1}^{r}s^{2}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s+r\int_{r}^{\infty}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s\bigg{)}.
Lemma 5.5

Let |n|=1|n|=1 and let α,δ\alpha,\delta\in\mathbb{R}. For λ0\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0} and fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega), we have

1r1rs2Φn,λ[fn](s)ds\displaystyle\frac{1}{r}\int_{1}^{r}s^{2}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s =l=19Jl[fn](r),\displaystyle=\sum_{l=1}^{9}J_{l}[f_{n}](r), (5.19)
rrΦn,λ[fn](s)ds\displaystyle r\int_{r}^{\infty}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s =l=1017Jl[fn](r),\displaystyle=\sum_{l=10}^{17}J_{l}[f_{n}](r), (5.20)

and

cn,λ[fn]\displaystyle c_{n,\lambda}[f_{n}] =l=11,13,14,15,17Jl[fn](1),\displaystyle=\sum_{l=11,13,14,15,17}J_{l}[f_{n}](1), (5.21)

where

J1[fn](r)\displaystyle J_{1}[f_{n}](r) =1r1rτδ2Iξn(λτ)gn(1)(τ)τrs2δ2Kξn(λs)dsdτ,\displaystyle=-\frac{1}{r}\int_{1}^{r}\tau^{\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}\tau)g^{(1)}_{n}(\tau)\int_{\tau}^{r}s^{2-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J2[fn](r)\displaystyle J_{2}[f_{n}](r) =(ξn+1δ2)1r1rτ1+δ2Iξn+1(λτ)fθ,n(τ)τrs1δ2Kξn1(λs)dsdτ,\displaystyle=-\Big{(}\xi_{n}+1-\frac{\delta}{2}\Big{)}\frac{1}{r}\int_{1}^{r}\tau^{1+\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\int_{\tau}^{r}s^{1-\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J3[fn](r)\displaystyle J_{3}[f_{n}](r) =1r1rτδ2Kξn(λτ)gn(2)(τ)1τs2δ2Iξn(λs)dsdτ,\displaystyle=\frac{1}{r}\int_{1}^{r}\tau^{\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}\tau)g^{(2)}_{n}(\tau)\int_{1}^{\tau}s^{2-\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J4[fn](r)\displaystyle J_{4}[f_{n}](r) =(ξn1+δ2)1r1rτ1+δ2Kξn1(λτ)fθ,n(τ)1τs1δ2Iξn+1(λs)dsdτ,\displaystyle=\Big{(}\xi_{n}-1+\frac{\delta}{2}\Big{)}\frac{1}{r}\int_{1}^{r}\tau^{1+\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\int_{1}^{\tau}s^{1-\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J5[fn](r)\displaystyle J_{5}[f_{n}](r) =1rrsδ2Kξn(λs)gn(2)(s)ds1rs2δ2Iξn(λs)ds,\displaystyle=\frac{1}{r}\int_{r}^{\infty}s^{\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}s)g^{(2)}_{n}(s)\,{\rm d}s\int_{1}^{r}s^{2-\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}s)\,{\rm d}s,
J6[fn](r)\displaystyle J_{6}[f_{n}](r) =(ξn1+δ2)1rrs1+δ2Kξn1(λs)fθ,n(s)ds1rs1δ2Iξn+1(λs)ds,\displaystyle=\Big{(}\xi_{n}-1+\frac{\delta}{2}\Big{)}\frac{1}{r}\int_{r}^{\infty}s^{1+\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}s)f_{\theta,n}(s)\,{\rm d}s\int_{1}^{r}s^{1-\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}s)\,{\rm d}s,
J7[fn](r)\displaystyle J_{7}[f_{n}](r) =r1δ2Kξn1(λr)1rτ1+δ2Iξn+1(λτ)fθ,n(τ)dτ,\displaystyle=r^{1-\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}r)\int_{1}^{r}\tau^{1+\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\,{\rm d}\tau,
J8[fn](r)\displaystyle J_{8}[f_{n}](r) =r1δ2Iξn+1(λr)rτ1+δ2Kξn1(λτ)fθ,n(τ)dτ,\displaystyle=r^{1-\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}r)\int_{r}^{\infty}\tau^{1+\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\,{\rm d}\tau,
J9[fn](r)\displaystyle J_{9}[f_{n}](r) =Iξn+1(λ)1r1τ1+δ2Kξn1(λτ)fθ,n(τ)dτ,\displaystyle=-I_{\xi_{n}+1}(\sqrt{\lambda})\frac{1}{r}\int_{1}^{\infty}\tau^{1+\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\,{\rm d}\tau,
J10[fn](r)\displaystyle J_{10}[f_{n}](r) =r1rsδ2Iξn(λs)gn(1)(s)dsrsδ2Kξn(λs)ds,\displaystyle=-r\int_{1}^{r}s^{\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}s)g^{(1)}_{n}(s)\,{\rm d}s\int_{r}^{\infty}s^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}s)\,{\rm d}s,
J11[fn](r)\displaystyle J_{11}[f_{n}](r) =rrτδ2Iξn(λτ)gn(1)(τ)τsδ2Kξn(λs)dsdτ,\displaystyle=-r\int_{r}^{\infty}\tau^{\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}\tau)g^{(1)}_{n}(\tau)\int_{\tau}^{\infty}s^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J12[fn](r)\displaystyle J_{12}[f_{n}](r) =(ξn1δ2)r1rs1+δ2Iξn+1(λs)fθ,n(s)dsrs1δ2Kξn1(λs)ds,\displaystyle=-\Big{(}\xi_{n}-1-\frac{\delta}{2}\Big{)}r\int_{1}^{r}s^{1+\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}s)f_{\theta,n}(s)\,{\rm d}s\int_{r}^{\infty}s^{-1-\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}s)\,{\rm d}s,
J13[fn](r)\displaystyle J_{13}[f_{n}](r) =(ξn1δ2)rrτ1+δ2Iξn+1(λτ)fθ,n(s)τs1δ2Kξn1(λs)dsdτ,\displaystyle=-\Big{(}\xi_{n}-1-\frac{\delta}{2}\Big{)}r\int_{r}^{\infty}\tau^{1+\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}\tau)f_{\theta,n}(s)\int_{\tau}^{\infty}s^{-1-\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J14[fn](r)\displaystyle J_{14}[f_{n}](r) =rrτδ2Kξn(λτ)gn(2)(τ)rτsδ2Iξn(λs)dsdτ,\displaystyle=r\int_{r}^{\infty}\tau^{\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}\tau)g^{(2)}_{n}(\tau)\int_{r}^{\tau}s^{-\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J15[fn](r)\displaystyle J_{15}[f_{n}](r) =(ξn+1+δ2)rrτ1+δ2Kξn1(λτ)fθ,n(τ)rτs1δ2Iξn+1(λs)dsdτ,\displaystyle=\Big{(}\xi_{n}+1+\frac{\delta}{2}\Big{)}r\int_{r}^{\infty}\tau^{1+\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\int_{r}^{\tau}s^{-1-\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}s)\,{\rm d}s\,{\rm d}\tau,
J16[fn](r)\displaystyle J_{16}[f_{n}](r) =r1δ2Kξn1(λr)1rτ1+δ2Iξn+1(λτ)fθ,n(τ)dτ,\displaystyle=-r^{1-\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}r)\int_{1}^{r}\tau^{1+\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\,{\rm d}\tau,
J17[fn](r)\displaystyle J_{17}[f_{n}](r) =r1δ2Iξn+1(λr)rτ1+δ2Kξn1(λτ)fθ,n(τ)dτ.\displaystyle=-r^{1-\frac{\delta}{2}}I_{\xi_{n}+1}(\sqrt{\lambda}r)\int_{r}^{\infty}\tau^{1+\frac{\delta}{2}}K_{\xi_{n}-1}(\sqrt{\lambda}\tau)f_{\theta,n}(\tau)\,{\rm d}\tau.
Remark 5.6
  1. (1)

    From (5.19)–(5.21), we see that

    cn,λ[fn]r1r1rs2Φn,λ[fn](s)ds=1rl=11,13,14,15Jl[fn](1)l=18Jl[fn](r).\frac{c_{n,\lambda}[f_{n}]}{r}-\frac{1}{r}\int_{1}^{r}s^{2}\Phi_{n,\lambda}[f_{n}](s)\,{\rm d}s=\frac{1}{r}\sum_{l=11,13,14,15}J_{l}[f_{n}](1)-\sum_{l=1}^{8}J_{l}[f_{n}](r).

    Thus we do not take the term J9[fn]J_{9}[f_{n}] into account when estimating 𝒱n[Φn,λ[fn]]\mathcal{V}_{n}\big{[}\Phi_{n,\lambda}[f_{n}]\big{]}.

  2. (2)

    Observe that J7[fn]=J16[fn]J_{7}[f_{n}]=-J_{16}[f_{n}] and that J8[fn]=J17[fn]J_{8}[f_{n}]=-J_{17}[f_{n}].


Proof of Lemma 5.5: The equalities (5.19)–(5.20) can be proved by change of the order of integration, the recurrence relations (see [40, Chapter III\mathrm{I}\mathrm{I}\mathrm{I} 3\cdot71 (3), (4)])

zKμ(z)\displaystyle zK_{\mu}(z) =(μ1)Kμ1(z)zdKμ1dz(z),\displaystyle=(\mu-1)K_{\mu-1}(z)-z\frac{\,{\rm d}K_{\mu-1}}{\,{\rm d}z}(z),
zIμ(z)\displaystyle zI_{\mu}(z) =(μ+1)Iμ+1(z)+zdIμ+1dz(z),\displaystyle=(\mu+1)I_{\mu+1}(z)+z\frac{\,{\rm d}I_{\mu+1}}{\,{\rm d}z}(z),

and integration by parts. We omit the details since they are analogous to those in the proof of [33, Lemmas 3.6 and 3.9] corresponding to the case δ=0\delta=0. The equality (5.21) follows from the definition (5.16) and (5.20) with r=1r=1. The proof is complete.  \Box

Lemma 5.7

Let |n|=1|n|=1 and ϵ(0,π)\epsilon\in(0,\pi) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}. We have the following.

  1. (1)

    Let l{1,,17}{7,8,9,16,17}l\in\{1,\ldots,17\}\setminus\{7,8,9,16,17\}. For q[1.)q\in[1.\infty) and fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega),

    supr1r2q|r1Jl[fn](r)|\displaystyle\sup_{r\geq 1}r^{\frac{2}{q}}|r^{-1}J_{l}[f_{n}](r)| C|λ|1fnLq,λΣπϵ{|z|<1},\displaystyle\leq C|\lambda|^{-1}\|f_{n}\|_{L^{q}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\},
    supr1|r1Jl[fn](r)|\displaystyle\sup_{r\geq 1}|r^{-1}J_{l}[f_{n}](r)| C|λ|1fnL1,λΣπϵ{|z|<1}.\displaystyle\leq C|\lambda|^{-1}\|f_{n}\|_{L^{1}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}.
  2. (2)

    Let l{7,8,16,17}l\in\{7,8,16,17\}. For fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega),

    1|r1Jl[fn](r)|rdr\displaystyle\int_{1}^{\infty}|r^{-1}J_{l}[f_{n}](r)|r\,{\rm d}r C|λ|1fnL1,λΣπϵ{|z|<1},\displaystyle\leq C|\lambda|^{-1}\|f_{n}\|_{L^{1}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\},
    supr1|r1Jl[fn](r)|\displaystyle\sup_{r\geq 1}|r^{-1}J_{l}[f_{n}](r)| C|λ|1fnL,λΣπϵ{|z|<1},\displaystyle\leq C|\lambda|^{-1}\|f_{n}\|_{L^{\infty}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\},
    supr1|r1Jl[fn](r)|\displaystyle\sup_{r\geq 1}|r^{-1}J_{l}[f_{n}](r)| CfnL1,λΣπϵ{|z|<1}.\displaystyle\leq C\|f_{n}\|_{L^{1}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}.

The constant CC depends only on α,δ,ϵ,q\alpha,\delta,\epsilon,q.

Proof.

Each of the estimates can be proved by Lemmas A.3 and A.4 in Appendix A. We omit the calculations since they are analogous to the ones in the proof of [33, Lemmas 3.7 and 3.10] corresponding to the case δ=0\delta=0. The proof is complete. ∎

Lemma 5.8

Let |n|=1|n|=1 and ϵ(0,π)\epsilon\in(0,\pi) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}. We have the following.

  1. (1)

    Let l{1,,17}{9}l\in\{1,\ldots,17\}\setminus\{9\}. For 1q<p1\leq q<p\leq\infty or 1<qp<1<q\leq p<\infty and fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega),

    (r,θ)r1Jl[fn](r)Lp\displaystyle\big{\|}(r,\theta)\mapsto r^{-1}J_{l}[f_{n}](r)\big{\|}_{L^{p}} C|λ|1+1q1pfLq,λΣπϵ{|z|<1}.\displaystyle\leq C|\lambda|^{-1+\frac{1}{q}-\frac{1}{p}}\|f\|_{L^{q}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}.
  2. (2)

    For q(1.)q\in(1.\infty) and fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega),

    |cn,λ[fn]|\displaystyle|c_{n,\lambda}[f_{n}]| C|λ|1+1qfLq,λΣπϵ{|z|<1}.\displaystyle\leq C|\lambda|^{-1+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}.

The constant CC depends only on α,δ,ϵ,q,p\alpha,\delta,\epsilon,q,p.

Proof.

(1) The estimate can be proved by Lemma 5.7 and interpolation theorems. We omit the details since they are analogous to those in the proof of [33, Corollary 3.12] and [22, Corollary 3.8] corresponding to the case δ=0\delta=0.

(2) The estimate follows from (5.21) and (1) with p=p=\infty. The proof is complete. ∎

Next we estimate 𝒱n[rrδ2Kξn(λr)]\mathcal{V}_{n}\big{[}r\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\big{]} in (5.17) and the terms in (5.18).

Lemma 5.9

Let |n|=1|n|=1 and ϵ(0,π)\epsilon\in(0,\pi) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}. We have the following.

  1. (1)

    For p(1,]p\in(1,\infty],

    𝒱n[rrδ2Kξn(λr)]LpC|λ|ξn21p,λΣπϵ{|z|<1}.\big{\|}\mathcal{V}_{n}\big{[}r\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\big{]}\big{\|}_{L^{p}}\leq C|\lambda|^{-\frac{\Re\xi_{n}}{2}-\frac{1}{p}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}.
  2. (2)

    For p[1,2]p\in[1,2],

    (r,θ)rδ2Kξn(λr)einθLpC|λ|ξn21p+12,λΣπϵ{|z|<1},\big{\|}(r,\theta)\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)e^{in\theta}\big{\|}_{L^{p}}\leq C|\lambda|^{-\frac{\Re\xi_{n}}{2}-\frac{1}{p}+\frac{1}{2}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\},

    and for p[2,)p\in[2,\infty),

    (r,θ)rδ2Kξn(λr)einθLpC|λ|ξn2,λΣπϵ{|z|<1}.\big{\|}(r,\theta)\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)e^{in\theta}\big{\|}_{L^{p}}\leq C|\lambda|^{-\frac{\Re\xi_{n}}{2}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}.
  3. (3)

    For p[1,]p\in[1,\infty] and fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega),

    (r,θ)Φn,λ[fn](r)einθLpC|λ|12fnLp,λΣπϵ{|z|<1}.\big{\|}(r,\theta)\mapsto\Phi_{n,\lambda}[f_{n}](r)e^{in\theta}\big{\|}_{L^{p}}\leq C|\lambda|^{-\frac{1}{2}}\|f_{n}\|_{L^{p}},\quad\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}.

The constant CC depends only on α,δ,ϵ,p\alpha,\delta,\epsilon,p.

Proof.

(1) The estimate can be proved by Lemma A.3 and interpolation theorems. We omit the details since they are analogous to those in the proof of [33, Proposition 3.17] and [22, Proposition 3.9] corresponding to the case δ=0\delta=0.

(2) The estimate follows from the inequality

(r,θ)rδ2Kξn(λr)einθLp(r,θ)Kξn(λr)einθLp.\big{\|}(r,\theta)\mapsto r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)e^{in\theta}\big{\|}_{L^{p}}\leq\big{\|}(r,\theta)\mapsto K_{\xi_{n}}(\sqrt{\lambda}r)e^{in\theta}\big{\|}_{L^{p}}.

and the estimate of the right-hand in [33, Lemma 3.22] and [22, Lemma B.4].

(3) The estimate can be proved by Lemma A.5 and interpolation theorems. We omit the details since they are analogous to those in the proof of [33, Lemma 3.21] corresponding to the case δ=0\delta=0. The proof is complete. ∎


Proof of Proposition 5.4: Since 𝒮2ϵ(α)Σ34πϵρ(𝔸V)\mathcal{S}_{2}^{\epsilon}(\alpha)\subset\Sigma_{\frac{3}{4}\pi-\epsilon}\subset\rho(-\mathbb{A}_{V}) by Corollary 4.10, we see that (λ+𝔸V)1f(\lambda+{\mathbb{A}}_{V})^{-1}f exists for any λ𝒮2ϵ(α)\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha). Let λ𝒮2ϵ(α)\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha). By density argument, it suffices to prove (5.9)–(5.10) for fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega). From Corollaries 4.9 and 5.8 (2), we have

|cn,λ[fn]Fn(λ)|C|λ|1+1q+ξn2fLq.\Big{|}\frac{c_{n,\lambda}[f_{n}]}{F_{n}(\sqrt{\lambda})}\Big{|}\leq C|\lambda|^{-1+\frac{1}{q}+\frac{\Re\xi_{n}}{2}}\|f\|_{L^{q}}.

Thus, from (5.17) and Lemma 5.5, putting p=2p=2 in Lemmas 5.8 and 5.9, we see that

𝒫n(λ+𝔸V)1fL2C|λ|32+1qfLq,\begin{split}\|\mathcal{P}_{n}(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}\leq C|\lambda|^{-\frac{3}{2}+\frac{1}{q}}\|f\|_{L^{q}},\end{split}

which is (5.9). Also, from (5.18), putting p=2p=2 in Lemma 5.9, we see that

rot𝒫n(λ+𝔸V)1fL2C|λ|12fL2,\begin{split}\|\operatorname{rot}\mathcal{P}_{n}(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}\leq C|\lambda|^{-\frac{1}{2}}\|f\|_{L^{2}},\end{split}

which leads to (5.10) since rotuL2=uL2\|\operatorname{rot}u\|_{L^{2}}=\|\nabla u\|_{L^{2}} for uW01,2(Ω)2Lσ2(Ω)u\in W^{1,2}_{0}(\Omega)^{2}\cap L^{2}_{\sigma}(\Omega). All the constants CC above are independent of λ\lambda. This completes the proof.  \Box

5.3 Proof of Proposition 5.1

Proposition 5.1 is a consequence of Lemma 5.2 and Propositions 5.3 and 5.4.

Proof of Proposition 5.1: Let ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}) be given. The same consideration as in the proof of Corollary 4.10 shows that Σ34πϵ𝒮1ϵ2(α)𝒮2ϵ2(α)\Sigma_{\frac{3}{4}\pi-\epsilon}\subset\mathcal{S}_{1}^{\frac{\epsilon}{2}}(\alpha)\cup\mathcal{S}_{2}^{\frac{\epsilon}{2}}(\alpha) for sufficiently small α,δ\alpha,\delta depending on ϵ\epsilon. In view of Proposition 5.3, the desired estimates (5.1)–(5.2) follow if we prove

(λ+𝔸V)1fL2C|λ|32+1qfLq,λ𝒮2ϵ(α),(λ+𝔸V)1fL2C|λ|1+1qfL2,λ𝒮2ϵ(α)\begin{split}\|(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}&\leq C|\lambda|^{-\frac{3}{2}+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha),\\ \|\nabla(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}&\leq C|\lambda|^{-1+\frac{1}{q}}\|f\|_{L^{2}},\quad\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha)\end{split} (5.22)

for fC0,σ(Ω)f\in C^{\infty}_{0,\sigma}(\Omega) and ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}). Let λ𝒮2ϵ(α)\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha) and set v=(λ+𝔸V)1fv=(\lambda+{\mathbb{A}}_{V})^{-1}f. Thanks to Proposition 5.4, we only need to estimate v=v|n|=1vnv_{\neq}=v-\sum_{|n|=1}v_{n}. Lemma 5.2 (2) implies that

|λ|vL22+vL22CfLq2q3q2vL24(q1)3q2,λ𝒮2ϵ(α)\displaystyle\begin{split}|\lambda|\|v_{\neq}\|_{L^{2}}^{2}+\|\nabla v_{\neq}\|_{L^{2}}^{2}&\leq C\|f\|_{L^{q}}^{\frac{2q}{3q-2}}\|v_{\neq}\|_{L^{2}}^{\frac{4(q-1)}{3q-2}},\quad\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha)\end{split}

with a constant C=C(ϵ)C=C(\epsilon), and hence that

vL2C|λ|32+1qfLq,λ𝒮2ϵ(α),vL2C|λ|1+1qfLq,λ𝒮2ϵ(α).\begin{split}\|v_{\neq}\|_{L^{2}}&\leq C|\lambda|^{-\frac{3}{2}+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha),\\ \|\nabla v_{\neq}\|_{L^{2}}&\leq C|\lambda|^{-1+\frac{1}{q}}\|f\|_{L^{q}},\quad\lambda\in\mathcal{S}_{2}^{\epsilon}(\alpha).\end{split}

Hence the proof is complete.  \Box

Appendix A Modified Bessel function

We summarize the facts about the modified Bessel functions. Our main references are [40, 3]. The modified Bessel function of the first kind Iμ(z)I_{\mu}(z) of order μ\mu is defined by

Iμ(z)=(z2)μm=01m!Γ(μ+m+1)(z2)2m,z0,\displaystyle I_{\mu}(z)=\Big{(}\frac{z}{2}\Big{)}^{\mu}\sum_{m=0}^{\infty}\frac{1}{m!\Gamma(\mu+m+1)}\Big{(}\frac{z}{2}\Big{)}^{2m},\quad z\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}, (A.1)

where Γ(z)\Gamma(z) is the Gamma function, the second kind Kμ(z)K_{\mu}(z) of order μ\mu\notin\mathbb{Z} is by

Kμ(z)=π2Iμ(z)Iμ(z)sin(μπ),z0,\displaystyle K_{\mu}(z)=\frac{\pi}{2}\frac{I_{-\mu}(z)-I_{\mu}(z)}{\sin(\mu\pi)},\quad z\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}, (A.2)

and Kn(z)K_{n}(z) of order nn\in\mathbb{Z} is by the limit of Kμ(z)K_{\mu}(z) in (A.2) as μn\mu\to n. In this paper, we exclusively consider the case where the order μ\mu satisfies μ\mu\notin\mathbb{Z} and μ>0\Re\mu>0.

The functions Kμ(z)K_{\mu}(z) and Iμ(z)I_{\mu}(z) are linearly independent solutions of

d2ωdz21zdωdz+(1+μ2z2)ω=0,z0,-\frac{\,{\rm d}^{2}\omega}{\,{\rm d}z^{2}}-\frac{1}{z}\frac{\,{\rm d}\omega}{\,{\rm d}z}+\Big{(}1+\frac{\mu^{2}}{z^{2}}\Big{)}\omega=0,\quad z\in\mathbb{C}\setminus\mathbb{R}_{\leq 0},

with the Wronskian

det(Kμ(z)Iμ(z)dKμdz(z)dIμdz(z))=1z.\displaystyle\det\left(\begin{array}[]{cc}K_{\mu}(z)&I_{\mu}(z)\\ \displaystyle{\frac{\,{\rm d}K_{\mu}}{\,{\rm d}z}(z)}&\displaystyle{\frac{\,{\rm d}I_{\mu}}{\,{\rm d}z}(z)}\end{array}\right)=\frac{1}{z}. (A.5)

It is well known that Iμ(z)I_{\mu}(z) grows exponentially and Kμ(z)K_{\mu}(z) decays exponentially as |z||z|\to\infty in Σπ2\Sigma_{\frac{\pi}{2}}; see [3, Section 4.12]. As an integral representation useful in Section 4, we have

Kμ(z)=120ez2(t+1t)tμ1dt,zΣπ2,\displaystyle K_{\mu}(z)=\frac{1}{2}\int_{0}^{\infty}e^{-\frac{z}{2}(t+\frac{1}{t})}t^{-\mu-1}\,{\rm d}t,\quad z\in\Sigma_{\frac{\pi}{2}}, (A.6)

which can be deduced by the formula [40, Chapter VI\mathrm{V}\mathrm{I} 6\cdot22 (5)] and change of variables.

Collected below are the estimates involving Kμ(z)K_{\mu}(z) and Iμ(z)I_{\mu}(z) used in this paper. Each of them can be found in the references [3, 40] or follows from a simple calculation using Lemma A.1. and hence we omit the proof. For the details when δ=0\delta=0, we refer to [33, Lemma 3.31 and Appendix A] and to [22] studying the dependence on α\alpha in the estimates.

We recall that the constants ηn\eta_{n} and ξn\xi_{n} are defined in (4.1) and (3.6), respectively.

Lemma A.1

Let μ>0\Re\mu>0, ϵ(0,π2)\epsilon\in(0,\frac{\pi}{2}) and M>0M>0. We have

|Kμ(z)|\displaystyle|K_{\mu}(z)| C|z|μ,zΣπ2{|z|<M},\displaystyle\leq C|z|^{-\Re\mu},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<M\},
|Kμ(z)|\displaystyle|K_{\mu}(z)| C|z|12ez,zΣπ2{|z|M},\displaystyle\leq C|z|^{-\frac{1}{2}}e^{-\Re z},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|\geq M\},
|Iμ(z)|\displaystyle|I_{\mu}(z)| C|z|μ,zΣπ2{|z|<M},\displaystyle\leq C|z|^{\Re\mu},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<M\},
|Iμ(z)|\displaystyle|I_{\mu}(z)| C|z|12ez,zΣπ2ϵ{|z|M}.\displaystyle\leq C|z|^{-\frac{1}{2}}e^{\Re z},\quad z\in\Sigma_{\frac{\pi}{2}-\epsilon}\cap\{|z|\geq M\}.

The constant CC depends on μ,ϵ,M\mu,\epsilon,M.

Lemma A.2

Let |n|=1|n|=1. We have the following.

  1. (1)

    For sufficiently small α,δ\alpha,\delta\in\mathbb{R},

    K1+ηn(z)=Γ(1+ηn)2(z2)1ηn+Rn(1)(z),zΣπ2{|z|<1}.K_{1+\eta_{n}}(z)=\frac{\Gamma(1+\eta_{n})}{2}\Big{(}\frac{z}{2}\Big{)}^{-1-\eta_{n}}+R^{(1)}_{n}(z),\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}.

    Here Rn(1)(z)R^{(1)}_{n}(z) is the remainder and satisfies

    |Rn(1)(z)|C|z|1ηn(1+|log|z||),zΣπ2{|z|<1}.|R^{(1)}_{n}(z)|\leq C|z|^{1-\Re\eta_{n}}\big{(}1+\big{|}\log|z|\big{|}\big{)},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}.
  2. (2)

    For sufficiently small α,δ\alpha,\delta\in\mathbb{R},

    Kηn(z)=π2sin(ηnπ)(1Γ(1ηn)(z2)ηn1Γ(1+ηn)(z2)ηn)+R~n(1)(z),\displaystyle K_{\eta_{n}}(z)=\frac{\pi}{2\sin(\eta_{n}\pi)}\bigg{(}\frac{1}{\Gamma(1-\eta_{n})}\Big{(}\frac{z}{2}\Big{)}^{-\eta_{n}}-\frac{1}{\Gamma(1+\eta_{n})}\Big{(}\frac{z}{2}\Big{)}^{\eta_{n}}\bigg{)}+\tilde{R}^{(1)}_{n}(z),
    zΣπ2{|z|<1}.\displaystyle\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}.

    Here R~n(1)(z)\tilde{R}^{(1)}_{n}(z) is the remainder and satisfies

    |R~n(1)(z)|C|z|2ηn(1+|log|z||),zΣπ2{|z|<1}.|\tilde{R}^{(1)}_{n}(z)|\leq C|z|^{2-\Re\eta_{n}}\big{(}1+\big{|}\log|z|\big{|}\big{)},\quad z\in\Sigma_{\frac{\pi}{2}}\cap\{|z|<1\}.

The constant CC is independent of α,δ\alpha,\delta.

Lemma A.3

Let |n|=1|n|=1 and ϵ(0,π)\epsilon\in(0,\pi) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}. For λΣπϵ{|z|<1}\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}, we have the following.

  1. (1)

    For 1τr(λ)11\leq\tau\leq r\leq(\Re\sqrt{\lambda})^{-1} and k=0,1k=0,1,

    τrs2kδ2|Kξnk(λs)|dsC|λ|ξn2+k2r3ξnδ2.\int_{\tau}^{r}s^{2-k-\frac{\delta}{2}}|K_{\xi_{n}-k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{\Re\xi_{n}}{2}+\frac{k}{2}}r^{3-\Re\xi_{n}-\frac{\delta}{2}}.
  2. (2)

    For 1τ(λ)1r1\leq\tau\leq(\Re\sqrt{\lambda})^{-1}\leq r\leq\infty and k=0,1k=0,1,

    τrs2kδ2|Kξnk(λs)|dsC|λ|32+k2+δ4.\int_{\tau}^{r}s^{2-k-\frac{\delta}{2}}|K_{\xi_{n}-k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{2}+\frac{k}{2}+\frac{\delta}{4}}.
  3. (3)

    For (λ)1τr(\Re\sqrt{\lambda})^{-1}\leq\tau\leq r\leq\infty and k=0,1k=0,1,

    τrs2kδ2|Kξnk(λs)|dsC|λ|34τ32kδ2e(λ)τ.\int_{\tau}^{r}s^{2-k-\frac{\delta}{2}}|K_{\xi_{n}-k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{4}}\tau^{\frac{3}{2}-k-\frac{\delta}{2}}e^{-(\Re\sqrt{\lambda})\tau}.
  4. (4)

    For 1τ(λ)11\leq\tau\leq(\Re\sqrt{\lambda})^{-1} and k=0,1k=0,1,

    τskδ2|Kξnk(λs)|dsC|λ|ξn2+k2τ1ξnδ2.\int_{\tau}^{\infty}s^{-k-\frac{\delta}{2}}|K_{\xi_{n}-k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{\Re\xi_{n}}{2}+\frac{k}{2}}\tau^{1-\Re\xi_{n}-\frac{\delta}{2}}.
  5. (5)

    For τ(λ)1\tau\geq(\Re\sqrt{\lambda})^{-1} and k=0,1k=0,1,

    τskδ2|Kξnk(λs)|dsC|λ|34τ12kδ2e(λ)τ.\int_{\tau}^{\infty}s^{-k-\frac{\delta}{2}}|K_{\xi_{n}-k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{4}}\tau^{-\frac{1}{2}-k-\frac{\delta}{2}}e^{-(\Re\sqrt{\lambda})\tau}.

The constant CC depends only on α,δ,ϵ\alpha,\delta,\epsilon.

Lemma A.4

Let |n|=1|n|=1 and ϵ(0,π)\epsilon\in(0,\pi) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}. For λΣπϵ{|z|<1}\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}, we have the following.

  1. (1)

    For 1τ(λ)11\leq\tau\leq(\Re\sqrt{\lambda})^{-1} and k=0,1k=0,1,

    1τs2kδ2|Iξn+k(λs)|dsC|λ|ξn2+k2τ3+ξnδ2.\int_{1}^{\tau}s^{2-k-\frac{\delta}{2}}|I_{\xi_{n}+k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{\frac{\Re\xi_{n}}{2}+\frac{k}{2}}\tau^{3+\Re\xi_{n}-\frac{\delta}{2}}.
  2. (2)

    For τ(λ)1\tau\geq(\Re\sqrt{\lambda})^{-1} and k=0,1k=0,1,

    1τs2kδ2|Iξn+k(λs)|dsC|λ|34τ32kδ2e(λ)τ.\int_{1}^{\tau}s^{2-k-\frac{\delta}{2}}|I_{\xi_{n}+k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{4}}\tau^{\frac{3}{2}-k-\frac{\delta}{2}}e^{(\Re\sqrt{\lambda})\tau}.
  3. (3)

    For 1rτ(λ)11\leq r\leq\tau\leq(\Re\sqrt{\lambda})^{-1} and k=0,1k=0,1,

    rτskδ2|Iξn+k(λs)|dsC|λ|ξn2+k2τ1+ξnδ2.\int_{r}^{\tau}s^{-k-\frac{\delta}{2}}|I_{\xi_{n}+k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{\frac{\Re\xi_{n}}{2}+\frac{k}{2}}\tau^{1+\Re\xi_{n}-\frac{\delta}{2}}.
  4. (4)

    For 1r(λ)1τ1\leq r\leq(\Re\sqrt{\lambda})^{-1}\leq\tau and k=0,1k=0,1,

    rτskδ2|Iξn+k(λs)|dsC|λ|34τ12kδ2e(λ)τ.\int_{r}^{\tau}s^{-k-\frac{\delta}{2}}|I_{\xi_{n}+k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{4}}\tau^{-\frac{1}{2}-k-\frac{\delta}{2}}e^{(\Re\sqrt{\lambda})\tau}.
  5. (5)

    For (λ)1rτ(\Re\sqrt{\lambda})^{-1}\leq r\leq\tau and k=0,1k=0,1,

    rτskδ2|Iξn+k(λs)|dsC|λ|34τ12kδ2e(λ)τ.\int_{r}^{\tau}s^{-k-\frac{\delta}{2}}|I_{\xi_{n}+k}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{4}}\tau^{-\frac{1}{2}-k-\frac{\delta}{2}}e^{(\Re\sqrt{\lambda})\tau}.

The constant CC depends only on α,δ,ϵ\alpha,\delta,\epsilon.

Lemma A.5

Let |n|=1|n|=1 and ϵ(0,π)\epsilon\in(0,\pi) and let (α,δ)×0(\alpha,\delta)\in\mathbb{R}^{\ast}\times\mathbb{R}_{\geq 0}. For λΣπϵ{|z|<1}\lambda\in\Sigma_{\pi-\epsilon}\cap\{|z|<1\}, we have the following.

  1. (1)

    For 1τ(λ)11\leq\tau\leq(\Re\sqrt{\lambda})^{-1},

    1τs1δ2|Iξn(λs)|dsC|λ|ξn2τ2+ξnδ2.\int_{1}^{\tau}s^{1-\frac{\delta}{2}}|I_{\xi_{n}}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{\frac{\Re\xi_{n}}{2}}\tau^{2+\Re\xi_{n}-\frac{\delta}{2}}.
  2. (2)

    For τ(λ)1\tau\geq(\Re\sqrt{\lambda})^{-1},

    1τs1δ2|Iξn(λs)|dsC|λ|34τ12δ2e(λ)τ.\int_{1}^{\tau}s^{1-\frac{\delta}{2}}|I_{\xi_{n}}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{4}}\tau^{\frac{1}{2}-\frac{\delta}{2}}e^{(\Re\sqrt{\lambda})\tau}.
  3. (3)

    For 1τ(λ)11\leq\tau\leq(\Re\sqrt{\lambda})^{-1},

    τs1δ2|Kξn(λs)|dsC|λ|ξn212τ1ξnδ2+C|λ|1+δ4.\int_{\tau}^{\infty}s^{1-\frac{\delta}{2}}|K_{\xi_{n}}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{\Re\xi_{n}}{2}-\frac{1}{2}}\tau^{1-\Re\xi_{n}-\frac{\delta}{2}}+C|\lambda|^{-1+\frac{\delta}{4}}.
  4. (4)

    For τ(λ)1\tau\geq(\Re\sqrt{\lambda})^{-1},

    τs1δ2|Kξn(λs)|dsC|λ|34τ12δ2e(λ)τ.\int_{\tau}^{\infty}s^{1-\frac{\delta}{2}}|K_{\xi_{n}}(\sqrt{\lambda}s)|\,{\rm d}s\leq C|\lambda|^{-\frac{3}{4}}\tau^{\frac{1}{2}-\frac{\delta}{2}}e^{-(\Re\sqrt{\lambda})\tau}.

The constant CC depends only on α,δ,ϵ\alpha,\delta,\epsilon.

Appendix B Homogeneous equation for vorticity

For λ0\lambda\in\mathbb{C}\setminus\mathbb{R}_{\leq 0}, we consider the homogeneous equation of (2.13)

d2ωndr21+δrdωndr+(λ+n2+iαnr2)ωn=0,r>1.-\frac{\,{\rm d}^{2}\omega_{n}}{\,{\rm d}r^{2}}-\frac{1+\delta}{r}\frac{\,{\rm d}\omega_{n}}{\,{\rm d}r}+\Big{(}\lambda+\frac{n^{2}+i\alpha n}{r^{2}}\Big{)}\omega_{n}=0,\quad r>1.

We will prove that its linearly independent solutions are, with ξn\xi_{n} defined in (3.6),

rδ2Kξn(λr)andrδ2Iξn(λr),\displaystyle r^{-\frac{\delta}{2}}K_{\xi_{n}}(\sqrt{\lambda}r)\quad\text{and}\quad r^{-\frac{\delta}{2}}I_{\xi_{n}}(\sqrt{\lambda}r), (B.1)

and the Wronskian is r1δr^{-1-\delta}. The proof is as follows. Applying the transformation

ωn(r)=rδ2ω~n(r),\displaystyle\omega_{n}(r)=r^{-\frac{\delta}{2}}\tilde{\omega}_{n}(r), (B.2)

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

d2ω~ndr21rdω~ndr+(λ+ξn2r2)ω~n=0,r>1.-\frac{\,{\rm d}^{2}\tilde{\omega}_{n}}{\,{\rm d}r^{2}}-\frac{1}{r}\frac{\,{\rm d}\tilde{\omega}_{n}}{\,{\rm d}r}+\Big{(}\lambda+\frac{\xi_{n}^{2}}{r^{2}}\Big{)}\tilde{\omega}_{n}=0,\quad r>1.

By Appendix A, its linearly independent solutions are

Kξn(λr)andIξn(λr).K_{\xi_{n}}(\sqrt{\lambda}r)\quad\text{and}\quad I_{\xi_{n}}(\sqrt{\lambda}r).

Hence, by the inverse transformation of (B.2), we see that the desired solutions are (B.1). One can easily compute the Wronskian using (A.5). The proof is complete.

Appendix C Proof of Theorem 1.1

Let ϵ(0,π4)\epsilon\in(0,\frac{\pi}{4}) and fix a number ϕ(π2,34πϵ)\phi\in(\frac{\pi}{2},\frac{3}{4}\pi-\epsilon). Taking b(0,1)b\in(0,1) and a curve γb\gamma_{b} in \mathbb{C}

γb={|argz|=ϕ,|z|b}{|argz|ϕ,|z|=b}\gamma_{b}=\{|\operatorname{arg}z|=\phi,\mkern 9.0mu|z|\geq b\}\cup\{|\operatorname{arg}z|\leq\phi,\mkern 9.0mu|z|=b\}

oriented counterclockwise, we use a representation of {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0} in the Dunford integral

et𝔸V=12πiγbetλ(λ+𝔸V)1dλ,t>0.e^{-t{\mathbb{A}}_{V}}=\frac{1}{2\pi i}\int_{\gamma_{b}}e^{t\lambda}(\lambda+{\mathbb{A}}_{V})^{-1}\,{\rm d}\lambda,\quad t>0.

From (5.1) for q=2q=2 in Proposition 5.1, we see that {et𝔸V}t0\{e^{-t{\mathbb{A}}_{V}}\}_{t\geq 0} is bounded in Lσ2(Ω)L^{2}_{\sigma}(\Omega), which implies the first line of (1.9). From (5.2), letting t>0t>0 and fLσ2(Ω)f\in L^{2}_{\sigma}(\Omega),

et𝔸VfL2\displaystyle\|\nabla e^{-t{\mathbb{A}}_{V}}f\|_{L^{2}} lim¯b0γbetλ(λ+𝔸V)1fL2|dλ|\displaystyle\leq\varlimsup_{b\to 0}\int_{\gamma_{b}}\|e^{t\lambda}\nabla(\lambda+{\mathbb{A}}_{V})^{-1}f\|_{L^{2}}|\,{\rm d}\lambda|
CfL20s12e(cosϕ)tsds,\displaystyle\leq C\|f\|_{L^{2}}\int_{0}^{\infty}s^{-\frac{1}{2}}e^{(\cos\phi)ts}\,{\rm d}s,

which implies the second line of (1.9). This completes the proof of Theorem 1.1.

Acknowledgements

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

References

  • [1] Ken Abe. Liouville theorems for the Stokes equations with applications to large time estimates. J. Funct. Anal., 278(2):108321, 30, 2020.
  • [2] Ken Abe. On the large time LL^{\infty}-estimates of the Stokes semigroup in two-dimensional exterior domains. J. Differential Equations, 300:337–355, 2021.
  • [3] George E. Andrews, Richard Askey, and Ranjan Roy. Special functions, volume 71 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1999.
  • [4] Wolfgang Borchers and Tetsuro Miyakawa. L2L^{2}-decay for Navier-Stokes flows in unbounded domains, with application to exterior stationary flows. Arch. Rational Mech. Anal., 118(3):273–295, 1992.
  • [5] Wolfgang Borchers and Tetsuro Miyakawa. On stability of exterior stationary Navier-Stokes flows. Acta Math., 174(2):311–382, 1995.
  • [6] Wolfgang Borchers and Werner Varnhorn. On the boundedness of the Stokes semigroup in two-dimensional exterior domains. Math. Z., 213(2):275–299, 1993.
  • [7] Lorenzo Brandolese and Maria E. Schonbek. Large time behavior of the Navier-Stokes flow. In Handbook of mathematical analysis in mechanics of viscous fluids, pages 579–645. Springer, Cham, 2018.
  • [8] 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.
  • [9] Wakako Dan and Yoshihiro Shibata. On the LqL_{q}LrL_{r} estimates of the Stokes semigroup in a two-dimensional exterior domain. J. Math. Soc. Japan, 51(1):181–207, 1999.
  • [10] Wakako Dan and Yoshihiro Shibata. Remark on the LqL_{q}-LL_{\infty} estimate of the Stokes semigroup in a 22-dimensional exterior domain. Pacific J. Math., 189(2):223–239, 1999.
  • [11] P. G. Drazin and W. H. Reid. Hydrodynamic stability. Cambridge Mathematical Library. Cambridge University Press, Cambridge, second edition, 2004. With a foreword by John Miles.
  • [12] P. G. Drazin and N. Riley. The Navier-Stokes equations: a classification of flows and exact solutions, volume 334 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2006.
  • [13] Reinhard Farwig and Jiří Neustupa. On the spectrum of a Stokes-type operator arising from flow around a rotating body. Manuscripta Math., 122(4):419–437, 2007.
  • [14] 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.
  • [15] 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.
  • [16] Giovanni P. Galdi and Masao Yamazaki. Stability of stationary solutions of two-dimensional Navier-Stokes exterior problem. In The proceedings on Mathematical Fluid Dynamics and Nonlinear Wave, volume 37 of GAKUTO Internat. Ser. Math. Sci. Appl., pages 135–162. Gakkōtosho, Tokyo, 2015.
  • [17] Julien Guillod. Steady solutions of the Navier–Stokes equations in the plane. arXiv:1511.03938, 2015.
  • [18] 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.
  • [19] Julien Guillod and Peter Wittwer. Generalized scale-invariant solutions to the two-dimensional stationary Navier-Stokes equations. SIAM J. Math. Anal., 47(1):955–968, 2015.
  • [20] Georg Hamel. Spiralförmige bewegungen zäher flüssigkeiten. Jahresbericht der Deutschen Mathematiker-Vereinigung, 25:34–60, 1917.
  • [21] John G. Heywood. On stationary solutions of the Navier-Stokes equations as limits of nonstationary solutions. Arch. Rational Mech. Anal., 37:48–60, 1970.
  • [22] 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.
  • [23] Mitsuo Higaki. Existence of planar non-symmetric stationary flows with large flux in an exterior disk. arXiv:2207.11922, 2022.
  • [24] Mitsuo Higaki. Stability of a two-dimensional stationary rotating flow in an exterior cylinder. Math. Nachr., 296(1):314–338, 2023.
  • [25] 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.
  • [26] Toshiaki Hishida. Asymptotic structure of steady Stokes flow around a rotating obstacle in two dimensions. In Mathematical fluid dynamics, present and future, volume 183 of Springer Proc. Math. Stat., pages 95–137. Springer, Tokyo, 2016.
  • [27] Grzegorz Karch and Dominika Pilarczyk. Asymptotic stability of Landau solutions to Navier-Stokes system. Arch. Ration. Mech. Anal., 202(1):115–131, 2011.
  • [28] Tosio Kato. Perturbation theory for linear operators. Grundlehren der Mathematischen Wissenschaften, Band 132. Springer-Verlag, Berlin-New York, second edition, 1976.
  • [29] Hideo Kozono and Takayoshi Ogawa. Two-dimensional Navier-Stokes flow in unbounded domains. Math. Ann., 297(1):1–31, 1993.
  • [30] 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.
  • [31] O. A. Ladyzhenskaya. The mathematical theory of viscous incompressible flow. Mathematics and its Applications, Vol. 2. Gordon and Breach Science Publishers, New York-London-Paris, 1969. Second English edition, revised and enlarged, Translated from the Russian by Richard A. Silverman and John Chu.
  • [32] Alessandra Lunardi. Analytic semigroups and optimal regularity in parabolic problems, volume 16 of Progress in Nonlinear Differential Equations and their Applications. Birkhäuser Verlag, Basel, 1995.
  • [33] Yasunori Maekawa. On stability of steady circular flows in a two-dimensional exterior disk. Arch. Ration. Mech. Anal., 225(1):287–374, 2017.
  • [34] 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.
  • [35] Yasunori Maekawa and Hiroyuki Tsurumi. Existence of the stationary navier-stokes flow in 2\mathbb{R}^{2} around a radial flow. Journal of Differential Equations, 350:202–227, 2023.
  • [36] P. Maremonti and V. A. Solonnikov. On nonstationary Stokes problem in exterior domains. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 24(3):395–449, 1997.
  • [37] Kyūya Masuda. Weak solutions of Navier-Stokes equations. Tohoku Math. J. (2), 36(4):623–646, 1984.
  • [38] Antonio Russo. On the existence of D-solutions of the steady-state navier-stokes equations in plane exterior domains. arXiv:1101.1243, 2011.
  • [39] Hermann Sohr. The Navier-Stokes equations: An elementary functional analytic approach. Modern Birkhäuser Classics. Birkhäuser/Springer Basel AG, Basel, 2001.
  • [40] G. N. Watson. A Treatise on the Theory of Bessel Functions. Cambridge University Press, Cambridge, England; The Macmillan Company, New York, 1944.
  • [41] Masao Yamazaki. Rate of convergence to the stationary solution of the Navier-Stokes exterior problem. In Recent developments of mathematical fluid mechanics, Adv. Math. Fluid Mech., pages 459–482. Birkhäuser/Springer, Basel, 2016.

M. Higaki

Department of Mathematics, Graduate School of Science, Kobe University, 1-1 Rokkodai, Nada-ku, Kobe 657-8501, Japan. Email: [email protected]