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

Capillary gravity water waves linearized at monotone shear flows: eigenvalues and inviscid damping

Xiao Liu School of Mathematics, Georgia Institute of Technology, Atlanta, GA 30332 xliu458@gatech.edu  and  Chongchun Zeng School of Mathematics, Georgia Institute of Technology, Atlanta, GA 30332 zengch@math.gatech.edu
Abstract.

We consider the 2D capillary gravity waves of finite depth x2(h,0)x_{2}\in(-h,0) linearized at a monotonic shear flow U(x2)U(x_{2}). The focuses are the eigenvalue distribution and linear inviscid damping. Unlike the Euler equation in a fixed channel where eigenvalues exist only in low wave numbers kk of the horizontal variable x1x_{1}, we first prove that the linearized capillary gravity wave has two branches of eigenvalues ikc±(k)-ikc^{\pm}(k), where the wave speeds c±(k)=O(|k|)\mathbb{R}\ni c^{\pm}(k)=O(\sqrt{|k|}) for |k|1|k|\gg 1 are asymptotic to those of the linear irrotational capillary gravity waves. Under the additional assumption U′′0U^{\prime\prime}\neq 0, we obtain the complete continuation of these two branches, which are all the eigenvalues in this (and some other) case(s). In particular, ikc(k)-ikc^{-}(k) could bifurcate into unstable eigenvalues at c(k)=U(h)c^{-}(k)=U(-h). The bifurcation of unstable eigenvalues from inflection values of UU is also obtained. Assuming there are no embedded eigenvalues for any wave number kk, the linearized velocity and surface profile (v(t,x),η(t,x1))(v(t,x),\eta(t,x_{1})) are considered in both periodic-in-x1x_{1} and x1x_{1}\in\mathbb{R} cases. Each solution can be split into (vp,ηp)(v^{p},\eta^{p}) and (vc,ηc)(v^{c},\eta^{c}) whose kk-th Fourier modes in x1x_{1} correspond to the eigenvalues and the continuous spectra of the wave number kk, respectively. The component (vp,ηp)(v^{p},\eta^{p}) is governed by the dispersion relation ω(k)=kc±(k)\omega(k)=-kc^{\pm}(k) in the case of x1x_{1}\in\mathbb{R}. The other component (vc,ηc)(v^{c},\eta^{c}) satisfies the linear inviscid damping as fast as |v1c|Lx2,|ηc|Lx2=O(1|t|)|v_{1}^{c}|_{L_{x}^{2}},|\eta^{c}|_{L_{x}^{2}}=O(\frac{1}{|t|}) and |v2c|Lx2=O(1t2)|v_{2}^{c}|_{L_{x}^{2}}=O(\frac{1}{t^{2}}) as |t||t|\to\infty. Furthermore, additional Lx2LtqL_{x}^{2}L_{t}^{q}, q(2,]q\in(2,\infty], decay of tv1ctv_{1}^{c} and t2v2ct^{2}v_{2}^{c} is obtained after leading asymptotic terms are singled out, which are in various forms of tt-dependent translations in x1x_{1} of certain functions of xx.

CZ is supported in part by the National Science Foundation DMS-1900083.

1. Introduction

Consider the two dimensional capillary gravity water waves in the moving domain of finite depth

𝒰t={(x1,x2)𝕋L×h<x2<η(t,x)},𝕋L:=/L,L>0,\mathcal{U}_{t}=\{(x_{1},x_{2})\in\mathbb{T}_{L}\times\mathbb{R}\mid-h<x_{2}<\eta(t,x)\},\quad\mathbb{T}_{L}:=\mathbb{R}/L\mathbb{Z},\;L>0,

or

𝒰t={(x1,x2)×h<x2<η(t,x)}.\mathcal{U}_{t}=\{(x_{1},x_{2})\in\mathbb{R}\times\mathbb{R}\mid-h<x_{2}<\eta(t,x)\}.

The free surface is given by St={(t,x)x2=η(t,x1)}S_{t}=\{(t,x)\mid x_{2}=\eta(t,x_{1})\}. For x𝒰tx\in\mathcal{U}_{t}, let v=(v1(t,x),v2(t,x))2v=(v_{1}(t,x),v_{2}(t,x))\in\mathbb{R}^{2} denote the fluid velocity and p=p(t,x)p=p(t,x)\in\mathbb{R} the pressure. They satisfy the free boundary problem of the incompressible Euler equation:

(1.1a) tv+(v)v+p+ge2=0,\displaystyle\partial_{t}v+(v\cdot\nabla)v+\nabla p+g\vec{e}_{2}=0, x𝒰t,\displaystyle x\in\mathcal{U}_{t},
(1.1b) v=0,\displaystyle\nabla\cdot v=0, x𝒰t,\displaystyle x\in\mathcal{U}_{t},
(1.1c) tη(t,x1)=v(t,x)(x1η(t,x1),1)T,\displaystyle\partial_{t}\eta(t,x_{1})=v(t,x)\cdot(-\partial_{x_{1}}\eta(t,x_{1}),1)^{T}, xSt,\displaystyle x\in S_{t},
(1.1d) p(t,x)=σκ(t,x),\displaystyle p(t,x)=\sigma\kappa(t,x), xSt,\displaystyle x\in S_{t},
(1.1e) v2(x1,h)=0,\displaystyle v_{2}(x_{1},-h)=0, x2=h,\displaystyle x_{2}=-h,

where σ>0\sigma>0, κ(t,x)=ηx1x1(1+ηx12)32\kappa(t,x)=-\frac{\eta_{x_{1}x_{1}}}{(1+\eta_{x_{1}}^{2})^{\frac{3}{2}}} is the mean curvature of StS_{t} at xx which corresponds to the surface tension, g>0g>0 is the gravitational acceleration, and the constant density is normalized to be 1.

Shear flows are a fundamental class of stationary solutions in the form of

(1.2) v:=(U(x2),0)T,S:={(t,x)|x2=η(x1)0},p=ge2.v_{*}:=\big{(}U(x_{2}),0\big{)}^{T},\quad S_{*}:=\{(t,x)|x_{2}=\eta_{*}(x_{1})\equiv 0\},\quad\nabla p_{*}=-g\vec{e}_{2}.

Our primary goal in this paper is to analyze the capillary gravity water wave system linearized at a monotone shear flow satisfying

(H) UCl0([h,0]),l03,U(x2)0,x2[h,0].U\in C^{l_{0}}([-h,0]),\;\;l_{0}\geq 3,\quad U^{\prime}(x_{2})\neq 0,\;\;\forall x_{2}\in[-h,0].
Remark 1.1.

Due to the symmetry of horizontal reflection

(x1,x2)(x1,x2),(v1,v2,η,p)(v1,v2,η,p),(x_{1},x_{2})\to(-x_{1},x_{2}),\quad(v_{1},v_{2},\eta,p)\to(-v_{1},v_{2},\eta,p),

the case of U<0U^{\prime}<0 is completely identical except the signs of U′′U^{\prime\prime} in Theorem 1.1(3d) should be reversed.

One of the crucial aspect of the linearized problem is the stability/instability, which is also related to the generation of surface and internal waves due to small disturbance. Mathematically, robust instability is often produced by eigenvalues of the linearized system which have positive real parts and lead to solutions with exponential growth in tt, while eigenvalues with negative real parts correspond to linear solutions with exponential decay. Purely imaginary eigenvalues, where there are infinitely many in the free linear capillary gravity waves (namely, linearized at zero), give periodically oscillatory linear solutions. Continuous spectra are also expected to exist, which do exist in the case of the Euler equation in a fixed channel linearized at a shear flow where certain algebraic decay of the linear solutions – linear inviscid damping – had been obtained under certain conditions. See Subsections 1.2 and 2.1 for references and the explicit example of the Couette flow. Hence the two main aspects of the linearized capillary gravity waves that we are focusing on are the eigenvalue distribution and the linear inviscid damping.

1.1. Linearization

We first derive the linearized system of (1.1) at the shear flow (v=(U(x2),0)T,η=0)(v_{*}=(U(x_{2}),0)^{T},\eta_{*}=0) given in (1.2) satisfied by the linearized solutions which we denote by (v,η,p)(v,\eta,p). Let (Stϵ,vϵ(t,x),pϵ(t,x))(S_{t}^{\epsilon},v^{\epsilon}(t,x),p^{\epsilon}(t,x)) be a one-parameter family of solutions of (1.1) with (St0,v0(t,x),p0(t,x))=(S,v,p)(S_{t}^{0},v^{0}(t,x),p^{0}(t,x))=(S_{*},v_{*},p_{*}). Differentiating the Euler equation (1.1a) and (1.1b) with respect to ϵ\epsilon and then evaluating it at ϵ=0\epsilon=0 yield

(1.3a) tv+U(x2)x1v+(U(x2)v2,0)T+p=0,v=0,x2(h,0).\partial_{t}v+U(x_{2})\partial_{x_{1}}v+(U^{\prime}(x_{2})v_{2},0)^{T}+\nabla p=0,\quad\;\;\nabla\cdot v=0,\quad\;x_{2}\in(-h,0).
Taking its divergence and also evaluating the above linearized Euler equation at x2=hx_{2}=-h, we obtain
(1.3b) p=2U(x2)x1v2,x2(h,0), and x2p|x2=h=0.-\triangle p=2U^{\prime}(x_{2})\partial_{x_{1}}v_{2},\quad x_{2}\in(-h,0),\;\text{ and }\;\partial_{x_{2}}p|_{x_{2}=-h}=0.
From the kinematic boundary condition (1.1c), we have
(1.3c) tη=v2|x2=0U(0)x1η.\partial_{t}\eta=v_{2}|_{x_{2}=0}-U(0)\partial_{x_{1}}\eta.
Finally differentiating (1.1d), where the left side is pϵ(t,x1,ηϵ(t,x1))p^{\epsilon}(t,x_{1},\eta^{\epsilon}(t,x_{1})), and using x2p=g\partial_{x_{2}}p_{*}=-g, we obtain
(1.3d) p=gησx12η, at x2=0.p=g\eta-\sigma\partial_{x_{1}}^{2}\eta,\quad\text{ at }\ x_{2}=0.

The above (1.3a1.3d) form the linearization of the capillary gravity water wave problem (1.1) at the shear flow (v,S,p)(v_{*},S_{*},p_{*}) with initial values (v10(x),v20(x),η0(x1))(v_{10}(x),v_{20}(x),\eta_{0}(x_{1})). In fact it can be reduced to an evolutionary problem of the unknowns (v,η)(v,\eta), while pp can be recovered by the boundary value problem of the elliptic system (1.3b) and (1.3d).

1.2. Backgrounds and motivations

Due to its physical and mathematical significance there have been extensive studies of the Euler equation linearized at shear currents. Many of these works were on a fixed channel with slip boundary conditions

(1.4) (1.1a)–(1.1b) with g=0,x2(h,0),v2(x1,0)=v2(x1,h)=0,\text{\eqref{E:Euler-1}--\eqref{E:Euler-2} with }g=0,\;\;\;x_{2}\in(-h,0),\quad v_{2}(x_{1},0)=v_{2}(x_{1},-h)=0,

and some of the results have been extended to free boundary problems such as the gravity waves. The spectral analysis is naturally a crucial part of such linear systems. Eigenvalues yield linear solutions exponential in time, while the continuous spectra often lead to algebraic decay of solutions, the so-called inviscid damping due to the lack of a priori dissipation mechanism of the Euler equation.

\bullet Eigenvalues. Since the variable coefficients in the linearized Euler system depend only on x2x_{2}, the subspace of the kk-th Fourier mode is invariant under the linear evolution for any kk\in\mathbb{R}. Hence it is a common practice to seek eigenvalues and eigenfunctions in the form of

(1.5) v(t,x)=eik(x1ct)(v10(x2),v20(x2)), also η(t,x1)=eik(x1ct)η0(x1)in the free boundary case,v(t,x)=e^{ik(x_{1}-ct)}\big{(}v_{10}(x_{2}),v_{20}(x_{2})\big{)},\;\;\text{ also }\;\eta(t,x_{1})=e^{ik(x_{1}-ct)}\eta_{0}(x_{1})\;\text{in the free boundary case},

where apparently the eigenvalues take the form λ=ikc\lambda=-ikc with the wave speed c=cR+icIc=c_{R}+ic_{I}\in\mathbb{C}. The linear system is spectrally unstable if there exist such cc with cI>0c_{I}>0, which appear in conjugate pairs. Solutions in the above form with cU([h,0])c\in U([-h,0]) are in a subtle situation and are referred to as singular modes (see Definition 2.1 and Remark 4.1 for singular and non-singular modes). In seeking solutions in the form of (1.5), the wave number kk\in\mathbb{R} is often treated as a parameter.

Classical results on the spectra of the Euler equation (1.4) in a channel linearized at a shear flow include:

  • Unstable eigenvalues are isolated for any wave number kk\in\mathbb{R} and do not exist for |k|1|k|\gg 1.

  • Rayleigh’s necessary condition of instability [30]: unstable eigenvalues do not exist for any kk if U′′0U^{\prime\prime}\neq 0 on [h,0][-h,0] (see also [8]).

  • Howard’s Semicircle Theorem [13]: for any k0k\neq 0, eigenvalues exist only with cc in the disk

    (1.6) (cR12(Umax+Umin))2+cI214(UmaxUmin)2.\big{(}c_{R}-\tfrac{1}{2}(U_{max}+U_{min})\big{)}^{2}+c_{I}^{2}\leq\tfrac{1}{4}(U_{max}-U_{min})^{2}.
  • Unstable eigenvalues may exist with cc near inflection values of UU (Tollmien [34] formally, also [21]).

Many classical results can be found in books such as [7, 26] etc. For a class of shear flows, the rigorous bifurcation of unstable eigenvalues was proved, e.g., in [9, 22]. In particular, continuation of branches of unstable eigenvalues were obtained by Lin in the latter.

It has been extended to the linearized free boundary problem of gravity waves (i.e. g>0g>0 and σ=0\sigma=0 in (1.1)) at shear flows (see [39, 14, 31] etc.) that: a.) assuming U>0U^{\prime}>0 and U′′0U^{\prime\prime}\neq 0 on [h,0][-h,0], there are no singular neutral modes in U((h,0))U\big{(}(-h,0)\big{)} (i.e. solution in the form of (1.5) with cU((h,0))c\in U\big{(}(-h,0)\big{)}); b.) the semicircle theorem still holds for unstable eigenvalues; and c.) for a class of shear flows, singular neutral modes may exist at inflection values of UU and the bifurcation and continuation of branches of unstable eigenvalues were also obtained. Compared to channel flows with fixed boundaries, new phenomena of the linearized gravity waves include: a.) in addition, critical values of UU, where U=0U^{\prime}=0, and c=U(h)c=U(-h) may be limiting singular neutral modes; and b.) there are non-singular neutral modes, i.e. cU([h,0])c\in\mathbb{R}\setminus U([-h,0]). Another related result is Miles’ critical layer theory [28, 5] on the instability of shear flows in two-phase fluid interface problem due to the resonance between the temporal frequency of the linear irrotational capillary gravity waves at the completely stationary water and the shear flow in the air in the above.

\bullet Inviscid damping. The analysis of the inviscid damping phenomenon started with the Euler equation in a fixed periodic channel (1.4) linearized at the Couette flow U(x2)=x2U(x_{2})=x_{2}. In 1907, Orr [29] observed that the linearized vertical velocity v2(t,x)v_{2}(t,x) tends to zero as tt\to\infty. Under the assumption v10(x1,x2)𝑑x1=0\int v_{10}(x_{1},x_{2})dx_{1}=0, x2(h,0)\forall x_{2}\in(-h,0), which removes the shear flow component of the linear solutions through an invariant splitting, explicit calculations (see, e.g., [6, 24]) yield, as tt\to\infty,

(1.7) ω0L2|v|L2=o(1),ω0H1|v|L2=O(1|t|),ω0H2|v2|L2=O(1|t|2),\omega_{0}\in L^{2}\implies|v|_{L^{2}}=o(1),\;\;\omega_{0}\in H^{1}\implies|v|_{L^{2}}=O(\tfrac{1}{|t|}),\;\;\omega_{0}\in H^{2}\implies|v_{2}|_{L^{2}}=O(\tfrac{1}{|t|^{2}}),

where ω0\omega_{0} denotes the initial vorticity. More general shear flows in a fixed channel have also been studied extensively. For a class of general stable shear flows, Bouchet and Morita [4] predicted similar decay estimates of the linearized velocity as well as the vorticity depletion phenomenon. For monotone shear flows without infection points, an O(|t|ν)O(|t|^{-\nu}) decay of the stream function was proved in [33] and then the (1.7) type decay in [40, 41] under a smallness assumption of LU′′LU^{\prime\prime} (also ω0|x2=h,0=0\omega_{0}|_{x_{2}=-h,0}=0 in order for the O(t2)O(t^{-2}) decay of v2v_{2}). A significant contribution is [35] by Wei-Zhang-Zhao where the (1.7) type estimates were obtained for general monotone shear flows without singular modes. In the follow-up works [36, 37, 38], vorticity depletion and velocity decay (as well as an Lt2L_{t}^{2} decay if ω0L2\omega_{0}\in L^{2} only) were also obtained for a class of non-monotone shear flows. As the decay rates in (1.7) are basically optimal, some leading order effects from both the interior and the boundary were identified in [40, 20]. In the absence of boundary impact, for compactly supported initial vorticity, linear inviscid damping near a class of monotone shear flows was also obtained in Gevrey spaces [19]. In [12], a different approach using methods from the study of Schrödinger operators was successfully adopted to analyze inviscid damping. See also [2, 18] for important developments for the linear inviscid damping at circular flows in 2\mathbb{R}^{2}.

While this paper focuses on the linearized capillary gravity waves at shear flows, among the rich literatures on the related nonlinear dynamics of the 2-d Euler equation on fixed domains we refer the readers to [1] for nonlinear Lyapunov stability of steady states based on energy-Casimir functions by Arnold; the remarkable asymptotic stability of shear flows in Gevrey class [3, 16, 17, 27] based on the linear inviscid damping; and for nonlinear instability of steady states [10, 11, 23, 25], etc.

\bullet Intuitions and goals on linearized capillary gravity waves. The goal of this paper is to study thoroughly the capillary gravity water waves linearized at a shear flow U(x2)U(x_{2}) under the above monotonicity assumption (H), focusing on the spectral distribution, stability/instability, and, if the eigenvalues are properly separated from the continuous spectra, the spectral projections and the linear inviscid damping.

For an illustration, some explicit computations of the linearized capillary gravity wave system (1.3) at the Couette flow U(x2)=x2U(x_{2})=x_{2} are given in Subsection 2.1. There it is easy to see that, on the one hand, the linear inviscid damping (1.7) holds for the rotational part of the solutions. On the other hand, there exist two branches of neutral modes c±(k)c^{\pm}(k) (see (2.3)) approaching infinity at the same rate as (2.4) of the irrotational capillary gravity waves linearized at zero. They form the two branches of the dispersion relation of the irrotational components in the linearized water wave system at the Couette flow, which is linearly stable.

Based on the above cited existing results on the channel flows with fixed boundaries, as well as those on the gravity water waves, and the explicit calculations of the capillary gravity waves linearized at the Couette flow, the analysis of the linearization (1.3) of the capillary gravity water waves at a general monotonic shear flow U(x2)U(x_{2}) is expected to include the following.

a.) Eigenvalues for wave numbers |k|1|k|\gg 1. Like matrices, eigenvalues (or equivalently, singular and non-singular modes) of (1.3) correspond to roots of a “characteristic" function (the 𝐅(k,c)\mathbf{F}(k,c) defined in (4.1)) analytic in kk\in\mathbb{R} and cU([h,0])c\in\mathbb{C}\setminus U([-h,0]). In contrast to the linearized Euler equation on a fixed domain where no eigenvalues exist for any large wave number kk, as seen in the linearized irrotational solutions of the capillary gravity waves at both the trivial (zero) solution and the Couette flow, likely there exist two non-singular neutral modes c±(k)c^{\pm}(k) for each wave number |k|1|k|\gg 1. These branches behave rather differently compared to the linearized gravity waves since the surface tension is dominant for |k|1|k|\gg 1. This part would be handled by an asymptotic analysis (Section 4).

b.) Analytic continuation and bifurcation of branches of eigenvalues and the spectral stability (Section 4). Each branch of non-singular modes could continue as long as they do not collide with each other or reach U([h,0])U([-h,0]), the boundary of the domain of analyticity of the characteristic function. The bifurcation analysis for cc near U([h,0])U([-h,0]), more specifically near inflection values of UU which possibly generates the instability (compare with the gravity wave case [14, 31, 15]) and U(h)U(-h), would require very careful study of the dependence of the solutions to the classical Rayleigh equation (3.1) on the singular parameter cc (Section 3). Our main tool is a local transformation which isolates the singular part of the solutions.

c.) Spectral projections and the linear inviscid damping (Section 6). Assuming that the eigenvalues are properly separated from the continuous spectra, a decomposition of linear solutions (v,η)=(vp,ηp)+(vc,ηc)(v,\eta)=(v^{p},\eta^{p})+(v^{c},\eta^{c}) into components corresponding to the eigenvalues and the continuous spectra, respectively, is expected. However, the boundedness of this spectral projection still needs to be obtained which is conceptually related to a lower bound of the angles both the two infinite dimensional components. This is a generalization of the Hodge decomposition of the free linear capillary gravity waves (linearized at zero) into the irrotational and the rotational parts. Both this intuition and the calculations of the capillary gravity waves linearized at the Couette flow suggest that (vp,ηp)(v^{p},\eta^{p}) is mostly related to the surface motion and dispersive (possibly with some unstable modes), while the other component (vc,ηc)(v^{c},\eta^{c}) is more determined by the internal rotations and thus by the vorticity. Whether the Euler equation is in a fixed domain or with free boundaries, the vorticity is transported in the same fashion by the fluid flow in the interior of the fluid domain. Hence it is natural to expect the linear inviscid damping (1.7) of (vc,ηc)(v^{c},\eta^{c}). One may further ask whether (1.7) is optimal in general. If so, a deeper question is whether it is possible to identify the leading order parts of (vc,ηc)(v^{c},\eta^{c}) for |t|1|t|\gg 1? These studies would be based on the careful analysis of the spectral contour integrals (Section 6) and the solutions to the homogeneous (Section 3) and non-homogeneous Rayleigh equation (Section 5).

1.3. Main results

We first give the main theorem on the eigenvalue distribution along with its implication on the linear stability. The results on the linear inviscid damping are somewhat more technical and only roughly outlined here. Their more precise statements are given in Theorems 2.1 and 2.2 in Subsection 2.2. See Definition 2.1, Lemma 4.1(5), and Remark 4.1 for what are referred to as singular and non-singular modes. Particularly, by slightly adjusting the same argument as in [13, 39], the Semi-circle Theorem still holds for the unstable modes of the linearized system (1.3) of the capillary gravity water waves at shear flows, namely, any unstable mode satisfies (1.6). We shall take this as granted in the rest of the paper.

Theorem 1.1.

(Eigenvalues.) Suppose UC3U\in C^{3} and U>0U^{\prime}>0 on [h,0][-h,0], then the following hold.

  1. (1)

    There exists k0>0k_{0}>0 such that for any kk\in\mathbb{R} with |k|k0|k|\geq k_{0}, there are no singular modes and exactly two non-singular modes c+(k)(U(0),+)c^{+}(k)\in(U(0),+\infty) and c(k)(,U(h))c^{-}(k)\in(-\infty,U(-h)) which correspond to semi-simple eigenvalues ikc±(k)-ikc^{\pm}(k). Moreover,

    1. (a)

      c±(k)c^{\pm}(k) are even and analytic in kk and c+(k)c^{+}(k) can be extended for all kk\in\mathbb{R} with c+(k)>U(0)c^{+}(k)>U(0);

    2. (b)

      lim|k|c±(k)/σ|k|=±1\lim_{|k|\to\infty}c^{\pm}(k)/\sqrt{\sigma|k|}=\pm 1;

    3. (c)

      if U(h)U(-h) is not a singular mode for any kk\in\mathbb{R}, then c(k)c^{-}(k) can also be extended to be even and analytic in all kk\in\mathbb{R} with c(k)<U(h)c^{-}(k)<U(-h); and

    4. (d)

      if singular modes do not exist k\forall\,k\in\mathbb{R}, then (k,c±(k))(k,c^{\pm}(k)) are the only non-singular modes of (1.3) which is linearly stable.

    5. (e)

      for k>0k>0, c+(k)c^{+}(k) (and c(k)<U(h)c^{-}(k)<U(-h) as well if it can be extended for all kk\in\mathbb{R}) has either none or exactly one critical point depending on whether (4.18) holds. In the latter case, the critical point is non-degenerate.

  2. (2)

    There exists g#0g_{\#}\geq 0 depending only on UU and σ\sigma such that the following hold.

    1. (a)

      If g>g#g>g_{\#}, then the non-singular modes c(k)<U(h)c^{-}(k)<U(-h) can also be extended to be even and analytic in all kk\in\mathbb{R} and ±(c±(k))>0\pm(c^{\pm}(k))^{\prime}>0 for k>0k>0;

    2. (b)

      g#=0g_{\#}=0 if and only if

      (1.8) σh0(U(x2)U(h))2𝑑x2.\sigma\geq\int_{-h}^{0}\big{(}U(x_{2})-U(-h)\big{)}^{2}dx_{2}.
  3. (3)

    If U′′0U^{\prime\prime}\neq 0 on [h,0][-h,0] is also satisfied, then the following hold with the g#g_{\#} given in the above statement (2).

    1. (a)

      The only possible singular mode is c=U(h)c=U(-h).

    2. (b)

      If g>g#g>g_{\#} then there are no singular modes, c±(k)c^{\pm}(k) are the only non-singular modes, and thus (1.3) is linearly stable.

    3. (c)

      If g=g#g=g_{\#} and UC5U\in C^{5}, then there exists k#>0k_{\#}>0 such that c(k)c^{-}(k) can be extended as an even C1,αC^{1,\alpha} function (for any α[0,1)\alpha\in[0,1)) for all kk\in\mathbb{R}. Moreover c(k)<U(h)c^{-}(k)<U(-h) is analytic for all k±k#k\neq\pm k_{\#}, and c(±k#)=U(h)c^{-}(\pm k_{\#})=U(-h). For each kk\in\mathbb{R}, c±(k)c^{\pm}(k) are the only singular or non-singular modes and thus (1.3) is spectrally stable.

    4. (d)

      If g<g#g<g_{\#} and UC5U\in C^{5}, then there exist k#+>k#>0k_{\#}^{+}>k_{\#}^{-}>0 such that we have the following.

      1. (i)

        Assume U′′>0U^{\prime\prime}>0 on [h,0][-h,0], then c(k)c^{-}(k) can be extended as an even C1,αC^{1,\alpha} function (for any α[0,1)\alpha\in[0,1)) for all kk\in\mathbb{R} and analytic except at k=±k#±k=\pm k_{\#}^{\pm} such that

        c(±k#±)=U(h),c(k)<U(h),|k|[k#,k#+],cI(k)>0,|k|(k#,k#+).c^{-}(\pm k_{\#}^{\pm})=U(-h),\quad c^{-}(k)<U(-h),\;\,\forall|k|\notin[k_{\#}^{-},k_{\#}^{+}],\quad c_{I}^{-}(k)>0,\;\,\forall|k|\in(k_{\#}^{-},k_{\#}^{+}).

        Moreover, for each kk, all singular and non-singular modes are exactly c+(k)c^{+}(k), c(k)c^{-}(k), as well as c(k)¯\overline{c^{-}(k)} if |k|(k#,k#+)|k|\in(k_{\#}^{-},k_{\#}^{+}). Consequently, (1.3) is spectrally unstable iff A.) x1x_{1}\in\mathbb{R} or B.) x1𝕋Lx_{1}\in\mathbb{T}_{L} and there exists mm\in\mathbb{N} such that 2πmL(k#,k#+)\frac{2\pi m}{L}\in(k_{\#}^{-},k_{\#}^{+}).

      2. (ii)

        Assume U′′<0U^{\prime\prime}<0 on [h,0][-h,0], then c(k)c^{-}(k) can be extended as an even C1,αC^{1,\alpha} real valued function (for any α[0,1)\alpha\in[0,1)) for |k|(k#,k#+)|k|\notin(k_{\#}^{-},k_{\#}^{+}), analytic in kk if |k|[k#,k#+]|k|\notin[k_{\#}^{-},k_{\#}^{+}], and c(±k#±)=U(h)c^{-}(\pm k_{\#}^{\pm})=U(-h). Moreover, all singular and non-singular modes are exactly c+(k)c^{+}(k) and c(k)c^{-}(k), where the latter is only for |k|(k#,k#+)|k|\notin(k_{\#}^{-},k_{\#}^{+}), and (1.3) is spectrally stable.

  4. (4)

    If UC5U\in C^{5} and U′′(x20)=0U^{\prime\prime}(x_{20})=0 for some x20[h,0)x_{20}\in[-h,0). Let c0=U(x20)c_{0}=U(x_{20}).

    1. (a)

      There exists σ0,k0>0\sigma_{0},k_{0}>0 such that for any σ(0,σ0)\sigma\in(0,\sigma_{0}), there exists a unique k>k0k>k_{0} unique in [k0,)[k_{0},\infty) such that c0c_{0} is a singular neutral mode for ±k\pm k.

    2. (b)

      If x20(h,0)x_{20}\in(-h,0), U′′′(x20)0U^{\prime\prime\prime}(x_{20})\neq 0, and c0c_{0} is a singular neutral mode for some k0>0k_{0}>0, then, under a non-degenerate condition (verified for the one obtained in (4a) for small σ\sigma), there exist unstable modes near c0c_{0} for kk close to k0k_{0} on one side of k0k_{0}.

Remark 1.2.

a.) The linear stability in (1d) and (3b) holds due to the inviscid damping in Theorems 2.1 and 2.2 and the dynamics in the directions of eigenfunctions being only oscillatory.
b.) If U′′0U^{\prime\prime}\neq 0 and (1.8) is satisfied, then the results in the above (3b) hold.
c.) Conceptually both the surface tension and the gravity have stablizing effects. For a given monotonic shear flow, assumption (1.8) is a sufficient condition to ensure that the surface tension itself is strong enough to stabilize the whole branch of the point spectra continued from c()=c^{-}(\infty)=-\infty.
d.) Condition (4.18) is also directly on UU, gg, and σ\sigma, but less explicit, and we leave it in Subsection 4.1.

The existence of the unbounded branches of non-singular neutral modes c±(k)c^{\pm}(k) are in contrast to the gravity waves or the Euler equation on fixed channels. The geometric multiplicity of ikc±(k)-ikc^{\pm}(k) occurs only among different kk. These temporal frequencies kc±(k)-kc^{\pm}(k) are asymptotic to those (see (2.4)) of the irrotational capillary gravity waves linearized at zero. Moreover, after normalizing the L2L^{2} norm of the vv component of the eigenfunction to be 11, the L2L^{2} and H1H^{1} differences in the vv and η\eta components, respectively, between the eigenfunctions of (1.3) and the linearized irrotational capillary gravity waves are of the order O(|k|32)O(|k|^{-\frac{3}{2}}) as |k||k|\to\infty (see Remark 6.1). In the case (happening only if (1.8) is not satisfied) where the branch c(k)c^{-}(k) reaches U(h)U(-h), where the bifurcation equation has the worst regularity, subtle bifurcations occur. This had been pointed out as a possibility in the linearized gravity waves [14, 31, 15], but not analyzed. In particular, it runs out that the sign of U′′U^{\prime\prime} determines whether c(k)c^{-}(k) becomes unstable or disappears at U(h)U(-h).

The spectral stability in the case U′′<0U^{\prime\prime}<0 can also be obtained by directly modifying the usual proof of the Rayleigh theorem in the fixed channel flow case, as done in [39] for the gravity wave. Our proof provides a complete picture of the eigenvalue distribution as in the above theorem, however.

While U(0)U(0) is never a singular mode, just like the Rayleigh’s theorem in the channel flow case the change of sign of U′′U^{\prime\prime} turns out to be necessary for the existence of interior singular modes, which is also sufficient if σ1\sigma\ll 1. In the contrast this may not be sufficient if the stabilizing gravity gg and surface tension σ\sigma are strong, see Remark 4.4.

In the following outline of the linear inviscid damping results, k𝐊=2πLk\in\mathbf{K}=\frac{2\pi}{L}\mathbb{N} if x1𝕋Lx_{1}\in\mathbb{T}_{L}, while k𝐊=k\in\mathbf{K}=\mathbb{R} if x1x_{1}\in\mathbb{R}. The initial velocity v0=(v10,v20)v_{0}=(v_{10},v_{20}) is always assumed to satisfy x1n1x2n2v0L2(𝕋L×[h,0])\partial_{x_{1}}^{n_{1}}\partial_{x_{2}}^{n_{2}}v_{0}\in L^{2}(\mathbb{T}_{L}\times[-h,0]) or L2(×[h,0])L^{2}(\mathbb{R}\times[-h,0]) for some n1n_{1}’s and n2n_{2}’s and similarly x1nη0L2\partial_{x_{1}}^{n}\eta_{0}\in L^{2}. To avoid too much technicality, here we skip the detailed assumptions on their regularity in x1x_{1}, but focus on that of x2x_{2} only. Precise statements are given in Theorems 2.1 and 2.2. The following LtqL_{t}^{q} is always for tt\in\mathbb{R}.

Main results on inviscid damping. Assume (H) and there are no singular modes for all k𝐊k\in\mathbf{K}, then any solution (v(t,x),η(t,x1))(v(t,x),\eta(t,x_{1})) to (1.3) can be decomposed into v=vp+vcv=v^{p}+v^{c} and η=ηp+ηc\eta=\eta^{p}+\eta^{c}, where (vp,ηp)(v^{p},\eta^{p}) belongs to the invariant subspace generated by the eigenfunction of the non-singular modes for all k𝐊k\in\mathbf{K}, while (vc,ηc)(v^{c},\eta^{c}) and its vorticity ωc\omega^{c} satisfy the following estimates.

  1. (1)

    If the initial vorticity ω0Lx22\omega_{0}\in L_{x_{2}}^{2}, then for any q[2,]q\in[2,\infty], vcLx2Ltqv^{c}\in L_{x}^{2}L_{t}^{q} and ηcLtq\eta^{c}\in L_{t}^{q}.

  2. (2)

    If ω0Hx21\omega_{0}\in H_{x_{2}}^{1}, then tv2ctv_{2}^{c}, tηcLx2Ltq1t\eta^{c}\in L_{x}^{2}L_{t}^{q_{1}} for any q1[2,]q_{1}\in[2,\infty]. Moreover there exists Ωc(x)Hx21\Omega^{c}(x)\in H_{x_{2}}^{1} such that for any q2(2,]q_{2}\in(2,\infty],

    tv1cU(x2)1x11Ωc(x1U(x2)t,x2),ωcΩc(x1U(x2)t,x2),x22v2cx1Ωc(x1U(x2)t,x2)Lx2Ltq2.tv_{1}^{c}-U^{\prime}(x_{2})^{-1}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2}),\,\omega^{c}-\Omega^{c}(x_{1}-U(x_{2})t,x_{2}),\,\partial_{x_{2}}^{2}v_{2}^{c}-\partial_{x_{1}}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\in L_{x}^{2}L_{t}^{q_{2}}.
  3. (3)

    If ω0Hx22\omega_{0}\in H_{x_{2}}^{2}, then there exist ΛB(x),ΛT(x)Hx21\Lambda_{B}(x),\Lambda_{T}(x)\in H_{x_{2}}^{1} such that for any q2(2,]q_{2}\in(2,\infty],

    t2v2cU(x2)2x11Ωc(x1U(x2)t,x2)ΛB(x1U(h)t,x2)ΛT(x1U(0)t,x2)Lx2Ltq2.t^{2}v_{2}^{c}-U^{\prime}(x_{2})^{-2}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})-\Lambda_{B}(x_{1}-U(-h)t,x_{2})-\Lambda_{T}(x_{1}-U(0)t,x_{2})\in L_{x}^{2}L_{t}^{q_{2}}.

In the above results, the assumption of the non-existence of singular modes, which is equivalent to the absence of embedded eigenvalues of (1.3) for each wave number kk, turns out to yield the spectral decomposition of the phase space of (1.3) into the invariant subspaces corresponding to the non-singular modes/point spectra and the continuous spectra ikU([h,0])-ikU([-h,0]) for each kk\in\mathbb{R}.

The component (vc,ηc)(v^{c},\eta^{c}) corresponds to the continuous spectra and enjoys temporal algebraic decay as in the case (1.4) of the Euler equation in a fixed channel. For the case of x1x_{1}\in\mathbb{R}, certain stronger decay in |k|1|k|\ll 1 (for long waves) is also assumed on the initial values, see Theorem 2.2 and Remark 2.2. Additional to the above LtqL_{t}^{q} bounds, derivatives-in-tt estimates are given in Theorems 2.1 and 2.2 as well which also imply pointwise-in-tt decay. Compared with (1.7), these additional LtqL_{t}^{q} estimates represent an improvement of decay of roughly an order of O(t1q)O(t^{-\frac{1}{q}}) (after appropriate tt-dependent translations in x1x_{1} of some asymptotic leading terms are identified and singled out in the cases of tv1ctv_{1}^{c}, t2v2ct^{2}v_{2}^{c}, etc.). For the Euler equation in a fixed channel (1.4), a.) when ω0L2\omega_{0}\in L^{2}, the |v|Lt2|v|_{L_{t}^{2}} estimate was also obtained in [36, 38]; b.) comparable asymptotic leading terms were identified in Lemma 3 of [40] and in Lemma 5.1 of [20]. The Fourier transforms (in x1x_{1}) of these leading terms Ωc\Omega^{c}, ΛT\Lambda_{T}, and ΛB\Lambda_{B} are given explicitly in (6.9), (6.16), and (6.15), which represent the impact of the interior flow and the top and bottom boundaries, respectively. See also (2.5) for singular elliptic boundary value problems satisfied by ΛT\Lambda_{T} and ΛB\Lambda_{B}. In particular, the free boundary effect is explicitly reflected in the boundary conditions (2.11c) of the corresponding Rayleigh equation (2.11) and the form (6.15) of ΛT\Lambda_{T}. The error estimates in addition to these leading asymptotic terms also justify that the estimates of tv1tv_{1} and t2v2t^{2}v_{2} in (1.7) are optimal. The precise asymptotic leading terms could be useful for further analysis.

The component (vp,ηp)(v^{p},\eta^{p}) are given by superpositions of the eigenfunctions of those non-singular modes, which is governed by a (possibly unstable) multi-branched dispersion relation given by kkck\to-kc for all non-singular modes cc of the kk-th Fourier modes in x1x_{1}. According to the above spectral analysis, this dispersion relation is asymptotic to that of the linear irrotational capillary gravity wave for |k|1|k|\gg 1. In the case of x1x_{1}\in\mathbb{R}, in the absence of singular modes, all non-singular modes are given by c±(k)c^{\pm}(k) which are neutral/stable. The dispersion of (vp,ηp)(v^{p},\eta^{p}) implies that it should decay if x1x_{1}\in\mathbb{R}, but at a slower rate. Hence the dynamics of (1.3) has two layers: faster inviscid decay of (vc,ηc)(v^{c},\eta^{c}) leaves the remaining (vp,ηp)(v^{p},\eta^{p}) decaying at a slower rate due to the dispersion like a linear irrotational capillary gravity wave.

In the periodic-in-x1x_{1} case, as the non-existence of singular modes is assumed only for k2πLk\in\frac{2\pi}{L}\mathbb{N}, there can still be other non-singular modes besides c±(k)c^{\pm}(k) which may have bifurcated from inflection values of U([h,0])U([-h,0]) (as well as unstable modes from U(h)U(-h)) at some k2πLk\notin\frac{2\pi}{L}\mathbb{N}. In particular instability may appear in finitely many dimensions in low wave numbers.

1.4. Outline of the proofs.

In the preliminary analysis in Subsection 2.3, we first apply the Fourier transform in x1x_{1} to (1.3), resulting in decoupled systems for each wave number kk. The problem can be further reduced to the evolution of v^2(t,k,x2)\hat{v}_{2}(t,k,x_{2}), the Fourier transform of v2v_{2}. The Laplacian transform111Working on the Laplace transform of the unknowns is essentially equivalent to analyzing the resolvent of the linear operator in (1.3). V2(k,c,x2)V_{2}(k,c,x_{2}) of v^2(t,k,x2)\hat{v}_{2}(t,k,x_{2}), where s=ikcs=-ikc is the Laplace transform variable, satisfies a non-homogeneous boundary value problem (2.11) of the Rayleigh equation, solutions to the associated homogeneous problem of which correspond to eigenvalues and eigenfunctions.

A detailed analysis of the homogeneous Rayleigh equation (3.1), carried out in Section 3, lays the foundation of the study of both the eigenvalue distribution and the inviscid damping. The dependence of the estimates of the solutions on the wave number kk is also carefully tracked.

We first study the Rayleigh equation away from the singularity where |U(x2)c|O(μ)|U(x_{2})-c|\geq O(\mu), μ=1k=(1+k2)12\mu=\frac{1}{\langle k\rangle}=(1+k^{2})^{-\frac{1}{2}}. Near the singularity where |U(x2)c|O(μ)|U(x_{2})-c|\leq O(\mu), different from those in, e.g., [35, 20], our approach is an improved version of the technique in [5] based on the ODE blow-up and invariant manifold method. Through a transformation, solutions to the homogeneous Rayleigh equation near U(x2)c=0U(x_{2})-c=0 are expressed in a form involving the explicit log(Uc)\log(U-c) and the heaviside function with coefficients smooth in (k,cR,x2)(k,c_{R},x_{2}).

We focus on a pair of fundamental solutions y±(k,c,x2)y_{\pm}(k,c,x_{2}) to the homogeneous Rayleigh equation which satisfy the corresponding homogeneous boundary conditions (2.11b)-(2.11c) in (2.11) at x2=0,hx_{2}=0,-h, respectively (boundary condition (2.11c) reflects the free boundary setting). For y±y_{\pm}, we establish a.) their a priori bounds; b.) the convergence to their limits y0±(k,cR,x2)y_{0\pm}(k,c_{R},x_{2}) as cI0+c_{I}\to 0+; and c.) the smoothness of y0±y_{0\pm}, particularly, in cRc_{R}. Recall UCl0U\in C^{l_{0}}, we prove y0±y_{0\pm} is Cl02C^{l_{0}-2} in cRc_{R} except at cR=U(h),U(0)c_{R}=U(-h),U(0). Due to the analyticity of y±y_{\pm} in cc with cI>0c_{I}>0, the estimates of y0±y_{0\pm} also yield those of y±y_{\pm} for cI>0c_{I}>0. Eventually general solutions to the non-homogeneous boundary value problem (2.11) of the Rayleigh equation are expressed using y±y_{\pm}. Finally, the quantity Y(k,c)=x2y(k,c,0)/y(k,c,0)Y(k,c)=\partial_{x_{2}}y_{-}(k,c,0)/y_{-}(k,c,0) related to the Reynolds stress is carefully studied, which plays an important role in the analysis of the Rayleigh equation.

The analysis of the Rayleigh equation near the singularity based on a canonical form presented in Section 3 gives the most detailed information of the solutions to this singular ODE. In the representation formula (3.74), additional to the explicit form of the singular part of the solutions, the Taylor expansion of the smooth transformation B()2×2B(\cdot)_{2\times 2} can be carried out to an arbitrary order if needed. The section is a lengthy, but we believe this technique applied to the Rayleigh equation is widely useful for various purposes. In a forthcoming work we are studying the linearized Euler equation at a non-monotonic shear flow with a similar approach.

In Section 4 we prove the results on the eigenvalue distribution based on the detailed analysis in Section 3. We first obtain c±(k)c^{\pm}(k) for |k|1|k|\gg 1, followed by an argument based on analytic continuation and index calculation. Bifurcations may occur at inflection values of UU and particularly subtle at c=U(h)c=U(-h), which are on the boundary of the analyticity of the bifurcation equation F(k,c)=0F(k,c)=0. The regularity obtained in Section 3 implies, when restricted to cI0c_{I}\geq 0, FCl02F\in C^{l_{0}-2} near cU((h,0))c\in U\big{(}(-h,0)\big{)} and FC1,αF\in C^{1,\alpha} near c=U(h)c=U(-h). This makes the bifurcation analysis possible near c=U(h)c=U(-h) and much easier even in the relatively classical case near inflection values of UU.

Among the results in Theorem 1.1, in statement (1), c±(k)c^{\pm}(k) are obtained for large |k||k| in Lemma 4.2(3) with more detailed estimates, the extension of c±(k)c^{\pm}(k) in Corollary 4.3.1, and the semi-simplicity of the eigenvalues ikc±(k)-ikc^{\pm}(k) in Lemma 4.2(3), Proposition 4.4, and Corollaries 4.3.1 and 6.2.1. Under the additional assumption of non-existence of singular modes, the non-existence of other non-singular modes is proved in Proposition 4.4. The analysis of the critical points of c±(k)c^{\pm}(k) is given in Lemmas 4.7. The conjugacy to the linearized irrotational waves is proved in Proposition 6.12. See also Remark 6.5. The existence of g#g_{\#} is proved in Lemma 4.6, along with the existence of k#k_{\#} and/or k#±k_{\#}^{\pm} in statement (3). The rest of statement (3) is proved at the end of Subsection 4.2 after a series of lemmas. Statement (4) is proved in Subsection 4.3 with more details.

Under the assumption of the absence of singular modes, general solutions yB(k,c,x2)y_{B}(k,c,x_{2}) to the non-homogeneous boundary value problem (2.11) of the Rayleigh equation are studied in Section 5, which are expressed in the variation of parameter formula using y±y_{\pm} obtained in Section 3. We establish the basic a priori and convergence (as cI0+c_{I}\to 0+) estimates in Subsection 5.1. The latter is often referred to as the limiting absorption principle (e.g. [36, 20]). For the inviscid damping estimates, it is crucial to obtain the smoothness of yBy_{B} in cc (in Subsection 5.2). Since singularity occurs in the Rayleigh equation along c=U(x2)c=U(x_{2}), cjyB\partial_{c}^{j}y_{B}, j=1,2j=1,2, behaves badly there. Instead we apply a differential operator DcD_{c} to the Rayleigh system (2.11) which differentiates along the direction of cR=U(x2)c_{R}=U(x_{2}), hence DcjyBD_{c}^{j}y_{B} satisfies another boundary value problem of the Rayleigh equation in the form of (2.11) and enjoys reasonable estimates. Essentially this approach is similar to those used in [35, 19, 38] for the Euler equation on fixed channels. The main results of Subsection 5.2 are the estimates of cj1x2yB\partial_{c}^{j_{1}}\partial_{x_{2}}y_{B}, j1=1,2j_{1}=1,2 and j2=0,1j_{2}=0,1, with the most singular terms identified.

The splitting and the linear inviscid damping estimates of solutions (v,η)(v,\eta) to the linearized capillary gravity waves (1.3) are obtained in Section 6. While the vorticity ω\omega is not sufficient to recover the whole solution (which is different from the fixed channel case as in e.g. [41, 35]), the solutions are expressed in terms of the inverse Laplace transform of V2(k,c,x2)V_{2}(k,c,x_{2}), the Laplace transform of v^2(t,k,x2)\hat{v}_{2}(t,k,x_{2}), which is estimated in Section 5. Unlike e.g. [35, 20], technically we do not immediately push the contour integral (in cc) of the inverse Laplace transform to the limit spectra set U([h,0])U([-h,0]), but first keep it along the boundary of a small neighborhood of it in the complex plane. This allows easy integration by parts in cc to establish the decay estimates in tt after the leading asymptotic terms are obtained by applying the Cauchy integral theorem to the most singular terms of cjV2\partial_{c}^{j}V_{2}, j=1,2j=1,2. In fact, in deriving the decay estimates of vv, η\eta, tv2tv_{2}, and tηt\eta where the leading asymptotic terms were not involved, a priori estimates, but not the limiting absorption principle, is sufficient.

The above approach to obtain the inviscid decay also applies to the Euler equation in a fixed channel linearized at a shear flow U(x2)U(x_{2}). Similarly, while the asymptotic leading order terms of tv1tv_{1}, ω\omega, and x22v2\partial_{x_{2}}^{2}v_{2} are all generated by the asymptotic vorticity Ωc\Omega^{c}, that of t2v2t^{2}v_{2} involves two additional functions ΩT\Omega_{T} and ΩB\Omega_{B} due to the contributions from the top and bottom boundaries. We give a brief summary of the results for the channel flow in Subsection 6.4 and see also Remark 6.6.

Notation: Throughout the paper, C>0C>0 denotes a generic constant which might change from line to line, but always independent of kk, cc, and x2x_{2}; δ(x)\delta(x) the delta function; P.V.P.V. (or (P.V.)c(P.V.)_{c}) the principle value (or the principle value with respect to variable cc etc.). The Japanese bracket k=k2+1\langle k\rangle=\sqrt{k^{2}+1} is adopted. For c=cR+icIc=c_{R}+ic_{I} close to U([h,0])U([-h,0]), x2cx_{2}^{c} denotes U1(cR)U^{-1}(c_{R}) after some extension of UU. We always denote μ=k1\mu=\langle k\rangle^{-1} and =x2{}^{\prime}=\partial_{x_{2}}.

2. Main results and preliminaries

In this section we give the precise statements of linear inviscid damping, along with some preliminary analysis. It is well-known that the pressure pp is determined by vv and η\eta (see (1.3b) and (1.3d)), so very often we shall focus only on vv and η\eta.

2.1. A brief motivational study of the Couette flow U(x2)=x2U(x_{2})=x_{2}

We first describe two main relevant properties using the Couette flow as an illustration. The linearized velocity can be decomposed uniquely into the rotational and irrotational/potential parts (see e.g. [32])

v=vir+vrot, where vir,rot=0,v=v^{ir}+v^{rot},\ \text{ where }\,\nabla\cdot v^{ir,rot}=0,\;

where

vir=φ,Δφ=0,x2(h,0), and x2φ|x2=h=0,v^{ir}=\nabla\varphi,\quad\Delta\varphi=0,\;\;x_{2}\in(-h,0),\,\text{ and }\,\partial_{x_{2}}\varphi|_{x_{2}=-h}=0,

and vrotv^{rot} satisfies

vrot=0,v2rot|x2=h,0=0.\nabla\cdot v^{rot}=0,\quad v_{2}^{rot}|_{x_{2}=-h,0}=0.

In particular, the rotational part can almost be determined by the vorticity ω\omega in the same way as in the Euler equation (1.4) in the fixed channel x2(h,0)x_{2}\in(-h,0) with slip boundary condition

(2.1) vrot=(x2,x1)TΔ1ω+(a,0)T, and ω=×v=x1v2x2v1,v^{rot}=(-\partial_{x_{2}},\partial_{x_{1}})^{T}\Delta^{-1}\omega+(a,0)^{T},\,\text{ and }\,\omega=\nabla\times v=\partial_{x_{1}}v_{2}-\partial_{x_{2}}v_{1},

where aa is a constant and Δ1\Delta^{-1} is the inverse Laplacian in the 2-d region x2(h,0)x_{2}\in(-h,0) (LL-periodic in x1x_{1} or x1x_{1}\in\mathbb{R}) under the zero Dirichlet boundary condition along x2=h,0x_{2}=-h,0. In the periodic-in-x1x_{1} case, the constant aa may be non-zero and is determined by the physical quantity circulation.

I. Inviscid damping. For the 2-d Euler equation (1.1a), one often also consider the corresponding vorticity formulation

(2.2) tω+vω=0.\partial_{t}\omega+v\cdot\omega=0.

Linearizing it at ω=1\omega_{*}=-1, which is the vorticity of the Couette flow, yields the linearized vorticity

ω(t,x)=ω0(x1x2t,x2)\omega(t,x)=\omega_{0}(x_{1}-x_{2}t,x_{2})

expressed in term of its initial value ω0\omega_{0}. Since vrotv^{rot} component of the linearized capillary gravity waves (1.3) at the Coutte flow corresponds to the divergence free velocity field determined by its vorticity ω\omega by (2.1) which is the same way as in the fixed boundary problem of the channel flow, the inviscid damping (1.7) of the latter (in the periodic-in-x1x_{1} case) implies

|vrot1L(L2L2v1𝑑x1)𝐞1|L2C(1+|t|)1|ω0|H2,|v2rot|L2C(1+|t|)2|ω0|H2.\Big{|}v^{rot}-\frac{1}{L}\Big{(}\int_{-\frac{L}{2}}^{\frac{L}{2}}v_{1}dx_{1}\Big{)}\vec{\bf e}_{1}\Big{|}_{L^{2}}\leq C(1+|t|)^{-1}|\omega_{0}|_{H^{2}},\quad|v_{2}^{rot}|_{L^{2}}\leq C(1+|t|)^{-2}|\omega_{0}|_{H^{2}}.

II. Singular and non-singular modes. Unlike the Euler equation in a fixed channel, there is the additional surface profile η\eta coupled to the irrotational part virv^{ir} of the velocity, which may not decay. In fact, for any kk\in\mathbb{R}, let

v(t,x)=(1+k2)14eik(x1c±(k)t)|k|h(icoshk(x2+h),sinhk(x2+h))+c.c.v(t,x)=(1+k^{2})^{\frac{1}{4}}e^{ik(x_{1}-c^{\pm}(k)t)-|k|h}\big{(}i\cosh k(x_{2}+h),\sinh k(x_{2}+h)\big{)}+c.c.
η(t,x1)=i(1+k2)14eik(x1c±(k)t)|k|hsinhkh/(kc±(k))+c.c.,\eta(t,x_{1})=i(1+k^{2})^{\frac{1}{4}}e^{ik(x_{1}-c^{\pm}(k)t)-|k|h}\sinh kh/(kc^{\pm}(k))+c.c.,
p(t,x)=\displaystyle p(t,x)= i(1+k2)14eik(x1c±(k)t)|k|h((g+σk2)sinhkhkc±(k)k0x2(x2c±(k))sinhk(x2+h)𝑑x2)+c.c.\displaystyle i(1+k^{2})^{\frac{1}{4}}e^{ik(x_{1}-c^{\pm}(k)t)-|k|h}\Big{(}(g+\sigma k^{2})\frac{\sinh kh}{kc^{\pm}(k)}-k\int_{0}^{x_{2}}(x_{2}^{\prime}-c^{\pm}(k))\sinh k(x_{2}^{\prime}+h)dx_{2}^{\prime}\Big{)}+c.c.
=\displaystyle= i(1+k2)14eik(x1c±(k)t)|k|h((g+σk2)sinhkhkc±(k)(x2c±(k))coshk(x2+h)\displaystyle i(1+k^{2})^{\frac{1}{4}}e^{ik(x_{1}-c^{\pm}(k)t)-|k|h}\Big{(}(g+\sigma k^{2})\frac{\sinh kh}{kc^{\pm}(k)}-(x_{2}-c^{\pm}(k))\cosh k(x_{2}+h)
c±(k)coshkh+k1(sinhk(x2+h)sinhkh))+c.c.\displaystyle-c^{\pm}(k)\cosh kh+k^{-1}(\sinh k(x_{2}+h)-\sinh kh)\Big{)}+c.c.

where “c.c.” denotes “complex conjugates” and

(2.3) c±(k)=1±1+4k(g+σk2)cothkh2kcothkhF(k,c)|c±(k)(c2kcothkh+c(g+σk2))|c±(k)=0.c^{\pm}(k)=\frac{-1\pm\sqrt{1+4k(g+\sigma k^{2})\coth kh}}{2k\coth kh}\implies F(k,c)|_{c^{\pm}(k)}\triangleq(c^{2}k\coth kh+c-(g+\sigma k^{2}))|_{c^{\pm}(k)}=0.

Even though we write down these formulas based on Lemma 2.3 in the below, it is straight forward to verify that they are solutions to (1.3a1.3d) for the Couette flow. Therefore ikc±(k)-ikc^{\pm}(k) are eigenvalues of the linearized systems associated with the above eigenfunctions. As these solutions do not grow or decay as tt\to\infty, c±(k)c^{\pm}(k) are neutral modes.

It is worth paying slightly closer attention to the wave speed c±(k)c^{\pm}(k) and the function F(k,c)F(k,c), all of which are even in kk. We make the following observations.

  1. (1)

    limkc±(k)/(σ|k|)12=±1\lim_{k\to\infty}c^{\pm}(k)/(\sigma|k|)^{\frac{1}{2}}=\pm 1, so for |k|1|k|\gg 1 the dispersion relation kkc±(k)k\to-kc^{\pm}(k) is asymptotic to those of the irrotational capillary gravity waves linearized at zero solution (system (1.3) with U0U\equiv 0 and ×v0\nabla\times v\equiv 0) given by kcir±-kc_{ir}^{\pm} with

    (2.4) cir±(k)=±k1(g+σk2)tanhkh,C1|cir±(k)|C(1+k2)14,c_{ir}^{\pm}(k)=\pm\sqrt{k^{-1}(g+\sigma k^{2})\tanh kh},\quad C^{-1}\leq|c_{ir}^{\pm}(k)|\leq C(1+k^{2})^{\frac{1}{4}},

    which can be obtained through direct calculation based on the Fourier transform.

  2. (2)

    c+(k)>0c^{+}(k)>0 for all kk\in\mathbb{R}, so it is a branch of non-singular neutral modes, namely, wave speeds outside [h,0][-h,0], the range of UU.

  3. (3)

    While c(k)<hc^{-}(k)<-h in (2.3) as seen in the above observation (1) for large kk, it can happen c(k)[h,0]c^{-}(k)\in[-h,0] for 0<g,σ10<g,\sigma\ll 1 and thus becomes singular modes (those in the range of UU).

  4. (4)

    Since kcothkhh1k\coth kh\geq h^{-1} with “==” achieved at k=0k=0, for g,σ1g,\sigma\gg 1, c±(k)cir±(k)=g+σk2kcothkhc^{\pm}(k)\sim c_{ir}^{\pm}(k)=\sqrt{\frac{g+\sigma k^{2}}{k\coth kh}} and thus both c±(k)[h,0]c^{\pm}(k)\notin[-h,0] are non-singular modes. Moreover, one may verify ddk|c±(k)|>0\frac{d}{dk}|c^{\pm}(k)|>0 for all k>0k>0 if σg1\sigma\gg g\gg 1. In particular, in the case of x1x_{1}\in\mathbb{R}, this implies that a.) the dispersion relations kkc±(k)k\to-kc^{\pm}(k) determine a linear dispersive wave system formed by the superposition of these non-singular modes and b.) this dispersive system is conjugate to the irrotational capillary gravity waves linearized at zero, whose the wave speed is given by (2.4). The conjugacy isomorphism can be constructed by associating the modes k1±k_{1}^{\pm} of (2.3) and k2±k_{2}^{\pm} of (2.4) if they have the same temporal frequency k1±c±(k1±)=k2±cir±(k2±)k_{1}^{\pm}c^{\pm}(k_{1}^{\pm})=k_{2}^{\pm}c_{ir}^{\pm}(k_{2}^{\pm}). Moreover, ikc±(k)-ikc^{\pm}(k) would turn out to the only eigenvalues for the linearization at the Couette flow for g,σ1g,\sigma\gg 1 (see Proposition 4.4(2)).

Generalization of the linear analysis to a general shear flow U(x2)U(x_{2})? From the above discussion, one sees that solutions to the capillary gravity water waves linearized at the Couette flow exhibit inviscid damping in their rotational parts while there are infinite many non-singular modes with irrotational eigenfunctions whose evolution is determined by two branches of dispersion relations. However, several complications arise in the linearization at a general shear flow U(x2)U(x_{2}) including at least the following.

  • The crucial function F(k,c)F(k,c) defined in (2.3) which determines the wave speed cc and consequently the dispersion relations, while analytic for cU([h,0])c\in\mathbb{C}\setminus U([-h,0]), may become rather singular for cc approaching U([h,0])U([-h,0]). What regularity of F(k,c)F(k,c) can one expect?

  • Consequently, if a branch of non-singular modes approaches U([h,0])U([-h,0]), possibly very subtle bifurcations may occur at the boundary of analyticity of FF. Can instability be generated?

  • The linear inviscid damping (still of the rotational parts?) becomes much more involved, even in the case of the channel flow (see e.g. [40, 35, 20]).

In the rest of this paper, we address these issues, with some results even more explicit and detailed than the above, through careful analysis starting at rather fundamental level under reasonable assumptions.

2.2. Main theorems on the invariant splitting and linear inviscid damping

In this subsection, assuming there are no singular modes, we present the theorems on the inviscid damping of linearized system (1.3) of the capillary gravity water wave problem (1.1) at the shear flow (v,S,p)(v_{*},S_{*},p_{*}). See Definition 2.1 Lemma 4.1(5), (4.9), and Remark 4.1 for singular and non-singular modes. In this case, we shall prove that any linear solution (v,η)(v,\eta) to (1.3) can be decomposed into the component (vp,ηp)(v^{p},\eta^{p}) corresponding to the non-singular modes and (vc,ηc)(v^{c},\eta^{c}) to the continuous spectra due to U([h,0])U([-h,0]). This splitting is invariant under (1.3) and (vc,ηc)(v^{c},\eta^{c}) is of the order O(|t|1)O(|t|^{-1}) (and the vertical component v2c=O(t2)v_{2}^{c}=O(t^{-2})) as |t||t|\to\infty. In fact, we identify their asymptotic leading order terms so that the remainders decay even faster. These leading order terms are in the form of horizontal translations of three functions Ωc\Omega^{c}, ΛB\Lambda_{B}, and ΛT\Lambda_{T}, which represent the contributions from the interior vorticity and the bottom and top boundary conditions. Their Fourier transforms are given explicitly in (6.9), (6.16), and (6.15), respectively, using the initial vorticity ω0\omega_{0}, the fundamental solutions y±(k,c,x2)y_{\pm}(k,c,x_{2}) to the homogeneous Rayleigh equation, and Ωc\Omega^{c} also by the Laplace transform of v2v_{2}. The results are stated for the cases of x1𝕋Lx_{1}\in\mathbb{T}_{L} and x1x_{1}\in\mathbb{R} separately in the following.

Theorem 2.1.

(Inviscid damping: periodic-in-x1x_{1} case) Suppose x1𝕋Lx_{1}\in\mathbb{T}_{L}. Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, U>0U^{\prime}>0 on [h,0][-h,0], and there are no singular modes (see (4.9) and Lemma 4.1(5)) for any k2πLk\in\frac{2\pi}{L}\mathbb{N}. For any q1[2,]q_{1}\in[2,\infty], q2(2,]q_{2}\in(2,\infty], and ϵ>0\epsilon>0, there exists C>0C>0 depending only on q1q_{1}, q2q_{2}, ϵ\epsilon, and UU, such that, for any n1n_{1}\in\mathbb{R}, integer n00n_{0}\geq 0, and solution (v(t,x),η(t,x1))(v(t,x),\eta(t,x_{1})) of (1.3) with initial value (v0(x),η0(x1))(v_{0}(x),\eta_{0}(x_{1})) and the corresponding initial vorticity ω0(x)\omega_{0}(x), there exist unique solutions (v(t,x),η(t,x1))(v^{\dagger}(t,x),\eta^{\dagger}(t,x_{1})), =p,c\dagger=p,c, to (1.3) and LL-periodic-in-x1x_{1} functions Ωc(x)\Omega^{c}(x), ΛB(x)\Lambda_{B}(x), and ΛT(x)\Lambda_{T}(x) determined by (v0,η0)(v_{0},\eta_{0}) linearly such that

(v,η)=(vc,ηc)+(vp,ηp)(v,\eta)=(v^{c},\eta^{c})+(v^{p},\eta^{p})

and the following hold.

  1. (1)

    (vc,ηc)(v^{c},\eta^{c}) satisfy the following estimates

    |tn0vc|Hx1n1Lx22Ltq1()C(|η0|Hx1n0+n1+121q1+|v10(,0)|Hx1n0+n1321q1+|ω0|Hx1n0+n1121q1+ϵLx22),|\partial_{t}^{n_{0}}v^{c}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}L_{t}^{q_{1}}(\mathbb{R})}\leq C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{0}+n_{1}+\frac{1}{2}-\frac{1}{q_{1}}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{3}{2}-\frac{1}{q_{1}}}}+|\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{1}{2}-\frac{1}{q_{1}}+\epsilon}L_{x_{2}}^{2}}\big{)},
    |tn0ηc|Hx1n1Ltq1()C(|η0|Hx1n0+n111q1+|v10(,0)|Hx1n0+n121q1+|ω0|Hx1n0+n121q1+ϵLx22);|\partial_{t}^{n_{0}}\eta^{c}|_{H_{x_{1}}^{n_{1}}L_{t}^{q_{1}}(\mathbb{R})}\leq C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-1-\frac{1}{q_{1}}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{0}+n_{1}-2-\frac{1}{q_{1}}}}+|\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-2-\frac{1}{q_{1}}+\epsilon}L_{x_{2}}^{2}}\big{)};

    if UC4U\in C^{4},

    |ttn0v2c|Hx1n132Lx22Ltq1()+|ttn0ηc|Hx1n1Ltq1()\displaystyle\big{|}t\partial_{t}^{n_{0}}v_{2}^{c}\big{|}_{H_{x_{1}}^{n_{1}-\frac{3}{2}}L_{x_{2}}^{2}L_{t}^{q_{1}}(\mathbb{R})}+|t\partial_{t}^{n_{0}}\eta^{c}|_{H_{x_{1}}^{n_{1}}L_{t}^{q_{1}}(\mathbb{R})}\leq C(|η0|Hx1n0+n111q1+|v10(,0)|Hx1n0+n131q1\displaystyle C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-1-\frac{1}{q_{1}}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{0}+n_{1}-3-\frac{1}{q_{1}}}}
    +|ω0|Hx1n0+n121q1+ϵLx22+|x2ω0|Hx1n0+n131q1+ϵLx22),\displaystyle+|\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-2-\frac{1}{q_{1}}+\epsilon}L_{x_{2}}^{2}}+|\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-3-\frac{1}{q_{1}}+\epsilon}L_{x_{2}}^{2}}\big{)},
    |tn0(tv1cU(x2)1x11Ωc(x1U(x2)t,x2))|Hx1n1Lx22Ltq2()\displaystyle\big{|}\partial_{t}^{n_{0}}\big{(}tv_{1}^{c}-U^{\prime}(x_{2})^{-1}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    +|tn0(ωcΩc(x1U(x2)t,x2))|Hx1n11Lx22Ltq2()\displaystyle+\big{|}\partial_{t}^{n_{0}}\big{(}\omega^{c}-\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{H_{x_{1}}^{n_{1}-1}L_{x_{2}}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    +|tn0(x22v2cx1Ωc(x1U(x2)t,x2))|Hx1n12Lx22Ltq2()\displaystyle+\big{|}\partial_{t}^{n_{0}}\big{(}\partial_{x_{2}}^{2}v_{2}^{c}-\partial_{x_{1}}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{H_{x_{1}}^{n_{1}-2}L_{x_{2}}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    \displaystyle\leq C(|η0|Hx1n0+n1+121q2+|v10(,0)|Hx1n0+n1321q2+|ω0|Hx1n0+n1121q2+ϵLx22+|x2ω0|Hx1n0+n1321q2+ϵLx22);\displaystyle C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{0}+n_{1}+\frac{1}{2}-\frac{1}{q_{2}}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{3}{2}-\frac{1}{q_{2}}}}+|\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{1}{2}-\frac{1}{q_{2}}+\epsilon}L_{x_{2}}^{2}}+|\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{3}{2}-\frac{1}{q_{2}}+\epsilon}L_{x_{2}}^{2}}\big{)};

    and if, in addition, UC5U\in C^{5}, then

    |tn0(t2v2cU(x2)2x11Ωc(x1U(x2)t,x2)ΛB(x1U(h)t,x2)ΛT(x1U(0)t,x2))|Hx1n1Lx22Ltq2()\displaystyle\big{|}\partial_{t}^{n_{0}}\big{(}t^{2}v_{2}^{c}-U^{\prime}(x_{2})^{-2}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})-\Lambda_{B}(x_{1}-U(-h)t,x_{2})-\Lambda_{T}(x_{1}-U(0)t,x_{2})\big{)}\big{|}_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    \displaystyle\leq C(|η0|Hx1n0+n1+121q2+|v10(,0)|Hx1n0+n1321q2+|ω0|Hx1n0+n1121q2+ϵLx22\displaystyle C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{0}+n_{1}+\frac{1}{2}-\frac{1}{q_{2}}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{3}{2}-\frac{1}{q_{2}}}}+|\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{1}{2}-\frac{1}{q_{2}}+\epsilon}L_{x_{2}}^{2}}
    +|x2ω0|Hx1n0+n1321q2+ϵLx22+|x22ω0|Hx1n0+n1521q2+ϵLx22).\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad+|\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{3}{2}-\frac{1}{q_{2}}+\epsilon}L_{x_{2}}^{2}}+|\partial_{x_{2}}^{2}\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{5}{2}-\frac{1}{q_{2}}+\epsilon}L_{x_{2}}^{2}}\big{)}.
  2. (2)

    Ωc\Omega^{c} and Λ\Lambda_{\dagger}, =B,T\dagger=B,T, satisfy

    |Ωcω0|Hx1n1Lx22C(|η0|Hx1n1+|v10(,0)|Hx1n12+|ω0|Hx1n11+ϵLx22),|\Omega^{c}-\omega_{0}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}}\leq C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{1}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{1}-2}}+|\omega_{0}|_{H_{x_{1}}^{n_{1}-1+\epsilon}L_{x_{2}}^{2}}\big{)},
    |kΛ^B(k,)|Lx2qCk1q|ω^0(k,h)|,|kΛ^T(k,)|Lx2qCk1q(|ω^0(k,0)|+|η^0(k)|),q[1,],|k\hat{\Lambda}_{B}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{-\frac{1}{q}}|\hat{\omega}_{0}(k,-h)|,\quad|k\hat{\Lambda}_{T}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{-\frac{1}{q}}(|\hat{\omega}_{0}(k,0)|+|\hat{\eta}_{0}(k)|),\quad\;\forall q\in[1,\infty],
    |kx2Λ^B(k,)|Lx2qCk11q|ω^0(k,h)|,|kx2Λ^T(k,)|Lx2qCk11q(|ω^0(k,0)|+|η^0(k)|),q[1,),|k\partial_{x_{2}}\hat{\Lambda}_{B}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{1-\frac{1}{q}}|\hat{\omega}_{0}(k,-h)|,\;\;|k\partial_{x_{2}}\hat{\Lambda}_{T}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{1-\frac{1}{q}}(|\hat{\omega}_{0}(k,0)|+|\hat{\eta}_{0}(k)|),\;\;\forall q\in[1,\infty),

    where f^(k,x2)\hat{f}(k,x_{2}) denotes the Fourier transform of a function f(x1,x2)f(x_{1},x_{2}) with respect to x1x_{1}. Moreover, Λ\Lambda_{\dagger}, =B,T\dagger=B,T, satisfy Λ^(k=0,x2)=0\hat{\Lambda}_{\dagger}(k=0,x_{2})=0 and

    (2.5a) {(UU(0))ΔΛT+U′′ΛT=0,x2(h,0),ΛT(x1,h)=0,U(0)2x1ΛT(x1,0)=U′′(0)η0(x1,0)ω0(x1,0);\begin{cases}-(U-U(0))\Delta\Lambda_{T}+U^{\prime\prime}\Lambda_{T}=0,\qquad\qquad\qquad\qquad\qquad\qquad\qquad x_{2}\in(-h,0),\\ \Lambda_{T}(x_{1},-h)=0,\qquad U^{\prime}(0)^{2}\partial_{x_{1}}\Lambda_{T}(x_{1},0)=U^{\prime\prime}(0)\eta_{0}(x_{1},0)-\omega_{0}(x_{1},0);\end{cases}
    (2.5b) {(UU(h))ΔΛB+U′′ΛB=0,x2(h,0),U(h)2x1ΛB(,h)=ω0(x1,h),(U(0)U(h))2x2ΛB(x1,0)(U(0)(U(0)U(h))+gσx12)ΛB(x1,0)=0.\begin{cases}-(U-U(-h))\Delta\Lambda_{B}+U^{\prime\prime}\Lambda_{B}=0,\qquad\qquad\qquad\qquad\qquad\qquad\qquad x_{2}\in(-h,0),\\ U^{\prime}(-h)^{2}\partial_{x_{1}}\Lambda_{B}(\cdot,-h)=-\omega_{0}(x_{1},-h),\\ \big{(}U(0)-U(-h)\big{)}^{2}\partial_{x_{2}}\Lambda_{B}(x_{1},0)-\big{(}U^{\prime}(0)(U(0)-U(-h))+g-\sigma\partial_{x_{1}}^{2}\big{)}\Lambda_{B}(x_{1},0)=0.\end{cases}
    If UC4U\in C^{4}, then
    |x2Ωcx2ω0|Hx1n1Lx22C(|η0|Hx1n1+1+|v10(,0)|Hx1n11+|ω0|Hx1n1+ϵLx22+|x2ω0|Hx1n11+ϵLx22).\displaystyle|\partial_{x_{2}}\Omega^{c}-\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}}\leq C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{1}+1}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{1}-1}}+|\omega_{0}|_{H_{x_{1}}^{n_{1}+\epsilon}L_{x_{2}}^{2}}+|\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{1}-1+\epsilon}L_{x_{2}}^{2}}\big{)}.
  3. (3)

    There exist λ00\lambda_{0}\geq 0 and integer N0N\geq 0 (given in (6.19)) such that, for any n1n_{1}\in\mathbb{R} and integer n2[1,l0]n_{2}\in[1,l_{0}],

    |x1n1+1(v1p(t,)v^10(k=0,))|Lx12Hx2n21+|x1n1v2p(t,)|Lx12Hx2n2\displaystyle\big{|}\partial_{x_{1}}^{n_{1}+1}\big{(}v_{1}^{p}(t,\cdot)-\hat{v}_{10}(k=0,\cdot)\big{)}\big{|}_{L_{x_{1}}^{2}H_{x_{2}}^{n_{2}-1}}+|\partial_{x_{1}}^{n_{1}}v_{2}^{p}(t,\cdot)|_{L_{x_{1}}^{2}H_{x_{2}}^{n_{2}}}
    \displaystyle\leq Ceλ0|t|(1+|t|N1)(|η0|Hx1n1+n2+1+|v10(,0)|Hx1n1+n212+|ω0|Hx1n1+n21Lx22),\displaystyle Ce^{\lambda_{0}|t|}(1+|t|^{N-1})\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{1}+n_{2}+1}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{1}+n_{2}-\frac{1}{2}}}+|\omega_{0}|_{H_{x_{1}}^{n_{1}+n_{2}-1}L_{x_{2}}^{2}}\big{)},
    |ηp(t,)η0(0)|Hx1n1Ceλ0|t|(1+|t|N1)(|η0|Hx1n1+|v10(,0)|Hx1n132+|ω0|Hx1n12Lx22).|\eta^{p}(t,\cdot)-\eta_{0}(0)|_{H_{x_{1}}^{n_{1}}}\leq Ce^{\lambda_{0}|t|}(1+|t|^{N-1})\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{1}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{1}-\frac{3}{2}}}+|\omega_{0}|_{H_{x_{1}}^{n_{1}-2}L_{x_{2}}^{2}}\big{)}.
  4. (4)

    Let

    𝐗={(v,η)|t=0 all (v0,η0)}H1(𝕋L×(h,0))×H2(𝕋L),=c,p,\mathbf{X}^{\dagger}=\{(v^{\dagger},\eta^{\dagger})|_{t=0}\mid\text{ all }(v_{0},\eta_{0})\}\subset H^{1}\big{(}\mathbb{T}_{L}\times(-h,0)\big{)}\times H^{2}(\mathbb{T}_{L}),\quad\dagger=c,p,

    then they are closed invariant subspaces of H1(𝕋L×(h,0))×H2(𝕋L)H^{1}\big{(}\mathbb{T}_{L}\times(-h,0)\big{)}\times H^{2}(\mathbb{T}_{L}) under (1.3). Moreover (1.3) is also well-posed in the L2×H1L^{2}\times H^{1} completion of 𝐗p\mathbf{X}^{p}.

Remark 2.1.

1.) Observe that, for any compactly supported smooth function f(t,x)f(t,x), q1q\geq 1, it holds

|f(t0,x)|qqt0+|f(t,x)|q1|tf(t,x)|𝑑tq|f(,x)|Ltq([t0,+))q1|tf(,x)|Ltq([t0,+)),|f(t_{0},x)|^{q}\leq q\int_{t_{0}}^{+\infty}|f(t^{\prime},x)|^{q-1}|\partial_{t}f(t^{\prime},x)|dt^{\prime}\leq q|f(\cdot,x)|_{L_{t}^{q}([t_{0},+\infty))}^{q-1}|\partial_{t}f(\cdot,x)|_{L_{t}^{q}([t_{0},+\infty))},

which implies

|f(t0,)|Lx2=\displaystyle|f(t_{0},\cdot)|_{L_{x}^{2}}= (|f(t0,x)|2𝑑x)12q1q(|f(,x)|Ltq([t0,+))2(q1)q|tf(,x)|Ltq([t0,+))2q𝑑x)12\displaystyle\Big{(}\int|f(t_{0},x)|^{2}dx\Big{)}^{\frac{1}{2}}\leq q^{\frac{1}{q}}\Big{(}\int|f(\cdot,x)|_{L_{t}^{q}([t_{0},+\infty))}^{\frac{2(q-1)}{q}}|\partial_{t}f(\cdot,x)|_{L_{t}^{q}([t_{0},+\infty))}^{\frac{2}{q}}dx\Big{)}^{\frac{1}{2}}
\displaystyle\leq q1q|f|Lx2Ltq([t0,+))q1q|tf|Lx2Ltq([t0,+))1q.\displaystyle q^{\frac{1}{q}}|f|_{L_{x}^{2}L_{t}^{q}([t_{0},+\infty))}^{\frac{q-1}{q}}|\partial_{t}f|_{L_{x}^{2}L_{t}^{q}([t_{0},+\infty))}^{\frac{1}{q}}.

By the standard density argument, this inequality also holds for any function fLx2Wt1,qf\in L_{x}^{2}W_{t}^{1,q}. Hence the above estimates in statement (1) also imply various pointwise-in-tt decay of vc,ηcLx2v^{c},\eta^{c}\in L_{x}^{2} as tt\to\infty.
2.) The function Ωc(x)\Omega^{c}(x) is referred to as the scattering limit of the vorticity in [41, 35, 20].
3.) The assumption of non-existence of singular modes is satisfied if the horizontal period LL is small (by Theorem 1.1(1) as 2πL\frac{2\pi}{L} is large) or if U′′0U^{\prime\prime}\neq 0 and (1.8) hold (by Theorem 1.1(2b)).

The proof of this theorem is completed in Subsection 6.2.

From (6.16) and (6.15), (4.1), and Lemma 3.19(2), Λ^B,T(0,x2)=0\hat{\Lambda}_{B,T}(0,x_{2})=0 and the elliptic boundary value problem (2.5b) has a unique solution ΛB\Lambda_{B}, while (2.5a) has a unique solution ΛT\Lambda_{T} under the assumption of the non-existence of singular modes. Moreover, according to the definitions (6.15), (4.1), (3.53), (3.83), and Lemma 3.10, x2ΛB\partial_{x_{2}}\Lambda_{B} and x2ΛT\partial_{x_{2}}\Lambda_{T} exhibit logarithmic singularity at x2=hx_{2}=-h and 0, respectively. In particular, ΛB=0\Lambda_{B}=0 vanishes if the initial vorticity ω0|x2=h=0\omega_{0}|_{x_{2}=-h}=0, while ΛT=0\Lambda_{T}=0 if U′′(0)η0ω0|x2=0=0U^{\prime\prime}(0)\eta_{0}-\omega_{0}|_{x_{2}=0}=0. In this paper as we focus on the damping estimates with additional LtqL_{t}^{q} decay of (v,η)(v,\eta) after the leading order terms are singled out, we adopted Lx2L_{x}^{2} based norms to somewhat simplify the calculations. If the decay in other LxrL_{x}^{r} or LxL_{x}^{\infty} based norms is necessary, some basic estimates in these norms are also given in Subsection 5.1 and one may make an attempt following the procedure as in Sections 5 and 6. To avoid more technicality, the assumptions on the regularity of ω0\omega_{0} in x1x_{1} in the theorem may not be close to optimal, particularly when q1q_{1} and q2q_{2} are away from 22, see Remark 6.2(b) as well as Remark 3.8. Moreover, the small ϵ\epsilon may not be necessary, see e.g. [36, 38] in the fixed boundary case. The assumptions on the more essential regularity of ω0\omega_{0} in x2x_{2} are optimal even in the existing results in the fixed boundary case.

In the estimates of the component (vp,ηp)(v^{p},\eta^{p}) which are superpositions of eigenfunctions, the possible exponential growth (if λ0>0\lambda_{0}>0) is caused by unstable modes, where λ0\lambda_{0} is the maximum real parts of the eigenvalues and NN is the maximum multiplicity of those eigenvalues of the maximal real parts. Due to Theorem 1.1(1), growth does not occur for |k|1|k|\gg 1. It is also worth pointing out that, taking n2=0n_{2}=0, the the regularity of ηp\eta^{p} is 32\frac{3}{2} order better than that of vpv^{p} restricted to the surface x2=0x_{2}=0, which is consistent with the regularity results of nonlinear capillary gravity waves in the existing literature. In the contrast, the regularity requirement on ω0\omega_{0} in the damping estimates of (vc,ηc)(v^{c},\eta^{c}) is stronger than that of (vp,ηp)(v^{p},\eta^{p}). Compared with the above example of the linearization at the Couette flow, conceptually these phenomena is due to the fact that the component (vc,ηc)(v^{c},\eta^{c}) is mainly the rotational part of the solution which depends on the vorticity more heavily, while (vp,ηp)(v^{p},\eta^{p}) are more like the irrotational part.

The estimate in statement (3) at t=0t=0 implies the boundedness of the projection onto 𝐗p\mathbf{X}^{p}, whose kernal is 𝐗c\mathbf{X}^{c}. Some more detailed information of this projection can be found in Lemma 6.2 and 6.10. In fact the subspace 𝐗p\mathbf{X}^{p} is generated by the eigenfunction of all non-singular modes for all kk\in\mathbb{R}.

The inviscid decay estimates in the case of x1x_{1}\in\mathbb{R} is slightly subtle due to the presence of small wave number |k|1|k|\ll 1. We use similar notations in the following theorem.

Theorem 2.2.

(Inviscid damping: x1x_{1}\in\mathbb{R} case) Suppose x1x_{1}\in\mathbb{R}. Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, U>0U^{\prime}>0 on [h,0][-h,0], and there are no singular modes (see (4.9) and Lemma 4.1(5)) for any kk\in\mathbb{R}. For any q1[2,]q_{1}\in[2,\infty], q2(2,]q_{2}\in(2,\infty], and ϵ>0\epsilon>0, there exists C>0C>0 depending only on q1q_{1}, q2q_{2}, ϵ\epsilon, and UU, such that, for any n1n_{1}\in\mathbb{R}, integers n00n_{0}\geq 0, and solution (v(t,x),η(t,x1))(v(t,x),\eta(t,x_{1})) of (1.3) with initial value (v0(x),η0(x1))(v_{0}(x),\eta_{0}(x_{1})), there exist solutions (v(t,x),η(t,x1))(v^{\dagger}(t,x),\eta^{\dagger}(t,x_{1})), =p,c\dagger=p,c, to (1.3) and functions Ωc(x)\Omega^{c}(x), ΛB(x)\Lambda_{B}(x), and ΛT(x)\Lambda_{T}(x) determined by (v0,η0)(v_{0},\eta_{0}) linearly such that

(v,η)=(vc,ηc)+(vp,ηp)(v,\eta)=(v^{c},\eta^{c})+(v^{p},\eta^{p})

and the following hold.

  1. (1)

    (vc,ηc)(v^{c},\eta^{c}) satisfy the following estimates

    |tn0x1n1v1c|Lx2Ltq1()+|tn0x1n11(1x12)12\displaystyle|\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}v_{1}^{c}|_{L_{x}^{2}L_{t}^{q_{1}}(\mathbb{R})}+|\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}-1}(1-\partial_{x_{1}}^{2})^{\frac{1}{2}} v2c|Lx2Ltq1()C(||x1|n0+n11q1η0|Hx112\displaystyle v_{2}^{c}|_{L_{x}^{2}L_{t}^{q_{1}}(\mathbb{R})}\leq C\big{(}\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\eta_{0}\big{|}_{H_{x_{1}}^{\frac{1}{2}}}
    +||x1|n0+n11q1v10(,0)|Hx132+||x1|n0+n11q1ω0|Hx1ϵ12Lx22),\displaystyle+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}v_{10}(\cdot,0)\big{|}_{H_{x_{1}}^{-\frac{3}{2}}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{1}{2}}L_{x_{2}}^{2}}\big{)},
    |tn0x1n1ηc|Lx12Ltq1()C(||x1|n0+n11q1η0|Hx11+||x1|n0+n11q1v10(,0)|Hx12+||x1|n0+n11q1ω0|Hx1ϵ2Lx22);|\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}\eta^{c}|_{L_{x_{1}}^{2}L_{t}^{q_{1}}(\mathbb{R})}\leq C\big{(}\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\eta_{0}\big{|}_{H_{x_{1}}^{-1}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}v_{10}(\cdot,0)\big{|}_{H_{x_{1}}^{-2}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-2}L_{x_{2}}^{2}}\big{)};

    if UC4U\in C^{4}, then

    |ttn0x1n1(1\displaystyle\big{|}t\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}(1- x12)14v2c|Lx2Ltq1()+|ttn0x1n1+1ηc|Lx12Ltq1()C(|η0|H˙x1n0+n11q1\displaystyle\partial_{x_{1}}^{2})^{-\frac{1}{4}}v_{2}^{c}\big{|}_{L_{x}^{2}L_{t}^{q_{1}}(\mathbb{R})}+|t\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}+1}\eta^{c}|_{L_{x_{1}}^{2}L_{t}^{q_{1}}(\mathbb{R})}\leq C\big{(}|\eta_{0}|_{\dot{H}_{x_{1}}^{n_{0}+n_{1}-\frac{1}{q_{1}}}}
    +||x1|n0+n11q1v10(,0)|Hx12+||x1|n0+n11q1ω0|Hx1ϵ1Lx22+||x1|n0+n11q1x2ω0|Hx1ϵ2Lx22)\displaystyle+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}v_{10}(\cdot,0)\big{|}_{H_{x_{1}}^{-2}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-1}L_{x_{2}}^{2}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\partial_{x_{2}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-2}L_{x_{2}}^{2}}\big{)}
    |tn0x1n1+1(tv1cU(x2)1x11Ωc(x1U(x2)t,x2))|Lx2Ltq2()\displaystyle\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}+1}\big{(}tv_{1}^{c}-U^{\prime}(x_{2})^{-1}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    +|tn0x1n1(ωcΩc(x1U(x2)t,x2))|Lx2Ltq2()\displaystyle+\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}\big{(}\omega^{c}-\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    +|tn0x1n11(x22v2cx1Ωc(x1U(x2)t,x2))|Lx2Ltq2()\displaystyle+\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}-1}\big{(}\partial_{x_{2}}^{2}v_{2}^{c}-\partial_{x_{1}}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    \displaystyle\leq C(||x1|n0+n11q2η0|Hx132+||x1|n0+n11q2v10(,0)|Hx112+||x1|n0+n11q2ω0|Hx1ϵ+12Lx22\displaystyle C\big{(}\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{2}}}\eta_{0}\big{|}_{H_{x_{1}}^{\frac{3}{2}}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{2}}}v_{10}(\cdot,0)\big{|}_{H_{x_{1}}^{-\frac{1}{2}}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{2}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon+\frac{1}{2}}L_{x_{2}}^{2}}
    +||x1|n0+n11q2x2ω0|Hx1ϵ12Lx22);\displaystyle\qquad+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{2}}}\partial_{x_{2}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{1}{2}}L_{x_{2}}^{2}}\big{)};

    and if, in addition, UC5U\in C^{5}, then

    |tn0x1n1(t2v2cU(x2)2x11Ωc(x1U(x2)t,x2)ΛB(x1U(h)t,x2)\displaystyle\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}\big{(}t^{2}v_{2}^{c}-U^{\prime}(x_{2})^{-2}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})-\Lambda_{B}(x_{1}-U(-h)t,x_{2})
    ΛT(x1U(0)t,x2))|Lx2Ltq2()\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\Lambda_{T}(x_{1}-U(0)t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
    \displaystyle\leq C(||x1|n0+n111q2η0|Hx132+||x1|n0+n111q2v10(,0)|Hx112+||x1|n0+n111q2ω0|Hx1ϵ+12Lx22\displaystyle C\big{(}\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-1-\frac{1}{q_{2}}}\eta_{0}\big{|}_{H_{x_{1}}^{\frac{3}{2}}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-1-\frac{1}{q_{2}}}v_{10}(\cdot,0)\big{|}_{H_{x_{1}}^{-\frac{1}{2}}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-1-\frac{1}{q_{2}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon+\frac{1}{2}}L_{x_{2}}^{2}}
    +||x1|n0+n111q2x2ω0|Hx1ϵ12Lx22+||x1|n0+n111q2x22ω0|Hx1ϵ32Lx22).\displaystyle\qquad+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-1-\frac{1}{q_{2}}}\partial_{x_{2}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{1}{2}}L_{x_{2}}^{2}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-1-\frac{1}{q_{2}}}\partial_{x_{2}}^{2}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{3}{2}}L_{x_{2}}^{2}}\big{)}.
  2. (2)

    ΛT\Lambda_{T} and ΛB\Lambda_{B} satisfy (2.5) and the same estimates as in Theorem 2.1(2). Moreover, for any q[1,)q\in[1,\infty), it holds

    |Ωcω0|Hx1n1Lx22C(|η0|Hx1n1+|v10(,0)|Hx1n12+|ω0|Hx1n11+ϵLx22),|\Omega^{c}-\omega_{0}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}}\leq C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{1}}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{1}-2}}+|\omega_{0}|_{H_{x_{1}}^{n_{1}-1+\epsilon}L_{x_{2}}^{2}}\big{)},

    and if UC4U\in C^{4}, then

    |x2Ωcx2ω0|Hx1n1Lx22C(|η0|Hx1n1+1+|v10(,0)|Hx1n11+|ω0|Hx1n1+ϵLx22+|x2ω0|Hx1n11+ϵLx22).|\partial_{x_{2}}\Omega^{c}-\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}}\leq C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{1}+1}}+|v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{1}-1}}+|\omega_{0}|_{H_{x_{1}}^{n_{1}+\epsilon}L_{x_{2}}^{2}}+|\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{1}-1+\epsilon}L_{x_{2}}^{2}}\big{)}.
  3. (3)

    For any n1n_{1}\in\mathbb{R} and the following integer n2n_{2},

    |x1n1x2n2v1p(t,)|Lx22C(|x1n1η0|Hx1n2+12+|x1n1v10(,0)|Hx1n2122+|x1n1ω0|Hx1n21Lx222),n2[0,l01],|\partial_{x_{1}}^{n_{1}}\partial_{x_{2}}^{n_{2}}v_{1}^{p}(t,\cdot)|_{L_{x}^{2}}^{2}\leq C\big{(}|\partial_{x_{1}}^{n_{1}}\eta_{0}|_{H_{x_{1}}^{n_{2}+1}}^{2}+|\partial_{x_{1}}^{n_{1}}v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{2}-\frac{1}{2}}}^{2}+|\partial_{x_{1}}^{n_{1}}\omega_{0}|_{H_{x_{1}}^{n_{2}-1}L_{x_{2}}^{2}}^{2}\big{)},\quad\forall n_{2}\in[0,l_{0}-1],
    |x1n1x2n2v2p(t,)|Lx22C(|x1n1+1η0|Hx1n22+|x1n1+1v10(,0)|Hx1n2322+|x1n1+1ω0|Hx1n22Lx222),n2[0,l0],|\partial_{x_{1}}^{n_{1}}\partial_{x_{2}}^{n_{2}}v_{2}^{p}(t,\cdot)|_{L_{x}^{2}}^{2}\leq C\big{(}|\partial_{x_{1}}^{n_{1}+1}\eta_{0}|_{H_{x_{1}}^{n_{2}}}^{2}+|\partial_{x_{1}}^{n_{1}+1}v_{10}(\cdot,0)|_{H_{x_{1}}^{n_{2}-\frac{3}{2}}}^{2}+|\partial_{x_{1}}^{n_{1}+1}\omega_{0}|_{H_{x_{1}}^{n_{2}-2}L_{x_{2}}^{2}}^{2}\big{)},\quad\forall n_{2}\in[0,l_{0}],
    |ηp(t,)|H˙x1n12C(|η0|H˙x1n12+|x1n1v10(,0)|Hx1322+|x1n1ω0|Hx12Lx222).|\eta^{p}(t,\cdot)|_{\dot{H}_{x_{1}}^{n_{1}}}^{2}\leq C\big{(}|\eta_{0}|_{\dot{H}_{x_{1}}^{n_{1}}}^{2}+|\partial_{x_{1}}^{n_{1}}v_{10}(\cdot,0)|_{H_{x_{1}}^{-\frac{3}{2}}}^{2}+|\partial_{x_{1}}^{n_{1}}\omega_{0}|_{H_{x_{1}}^{-2}L_{x_{2}}^{2}}^{2}\big{)}.
  4. (4)

    Let

    𝐗={(v,η)|t=0 all (v0,η0)}H1(×(h,0))×H2(),=c,p,\mathbf{X}^{\dagger}=\{(v^{\dagger},\eta^{\dagger})|_{t=0}\mid\text{ all }(v_{0},\eta_{0})\}\subset H^{1}\big{(}\mathbb{R}\times(-h,0)\big{)}\times H^{2}(\mathbb{R}),\quad\dagger=c,p,

    then they are invariant closed subspaces of H1(×(h,0))×H2()H^{1}\big{(}\mathbb{R}\times(-h,0)\big{)}\times H^{2}(\mathbb{R}) under (1.3). Moreover (1.3) is also well-posed in the L2×H1L^{2}\times H^{1} completion of 𝐗p\mathbf{X}^{p}. If, in addition, (4.18) holds for both ±\pm and 0U([h,0])0\in U([-h,0]), then (1.3) restricted to the L2×H1L^{2}\times H^{1} completion of 𝐗p\mathbf{X}^{p}, or 𝐗p(Hn×Hn+1)\mathbf{X}^{p}\cap(H^{n}\times H^{n+1}) with nl01n\leq l_{0}-1, is conjugate through an isomorphism to the irrotational capillary gravity waves linearized at zero (characterized by its wave speed (2.4)).

Remark 2.2.

In the above estimates, for some n0n_{0} and n1n_{1}, the |x1|s|\partial_{x_{1}}|^{-s}, s>0s>0, applied to the initial values indicates some stronger decay assumptions for wave number |k|1|k|\ll 1.

The proof of this theorem is completed in Subsection 6.3. Most of the remarks after Theorem 2.1 are also valid. In particular, there are only two branches of non-singular modes corresponding to eigenvalues ikc±(k)ikc^{\pm}(k) of both algebraic and geometric multiplicity two, hence there is no growth at all. The conjugacy of the dynamics of (vp,ηp)(v^{p},\eta^{p}) to the linear irrotational capillary gravity waves is basically a restatement of Theorem 1.1(2b).

2.3. Preliminary linear analysis

To analyze the linear system (1.3), we first reduce it to an evolution problem of the Fourier transform of v2v_{2} in x1x_{1}, which in turn determines v1v_{1}, η\eta, and pp. We then apply the Laplace transform in tt to obtain a non-homogeneous boundary value problem of the well-known Rayleigh equation in x2(h,0)x_{2}\in(-h,0) with a non-homogeneous Robin type boundary condition at x2=0x_{2}=0 due the boundary conditions at the free boundary. The main analysis will focus on the Rayleigh equation.

Consider the Fourier transforms of the unknowns (v(t,x),η(t,x1),p(t,x))(v(t,x),\eta(t,x_{1}),p(t,x)) in x1x_{1}

v(x)=k2πLv^(k,x2)eikx1,η(x1)=k2πLη^(k)eikx1,p(x)=k2πLp^(k,x2)eikx1,v(x)=\sum_{k\in\tfrac{2\pi}{L}\mathbb{Z}}\hat{v}(k,x_{2})e^{ikx_{1}},\;\;\eta(x_{1})=\sum_{k\in\tfrac{2\pi}{L}\mathbb{Z}}\hat{\eta}(k)e^{ikx_{1}},\;\;p(x)=\sum_{k\in\tfrac{2\pi}{L}\mathbb{Z}}\hat{p}(k,x_{2})e^{ikx_{1}},

in the case of x1𝕋Lx_{1}\in\mathbb{T}_{L} and

v(x)=12πv^(k,x2)eikx1𝑑k,η(x1)=12πη^(k)eikx1𝑑k,p(x)=12πp^(k,x2)eikx1𝑑k,v(x)=\frac{1}{2\pi}\int_{\mathbb{R}}\hat{v}(k,x_{2})e^{ikx_{1}}dk,\;\;\eta(x_{1})=\frac{1}{2\pi}\int_{\mathbb{R}}\hat{\eta}(k)e^{ikx_{1}}dk,\;\;p(x)=\frac{1}{2\pi}\int_{\mathbb{R}}\hat{p}(k,x_{2})e^{ikx_{1}}dk,

in the case of x1x_{1}\in\mathbb{R}, where we skipped the variable tt. The Fourier transform of the linearized system (1.3) takes the form

(2.6) {tv^+ikU(x2)v^+(U(x2)v^2,0)T+(ikp^,p^)T=0,ikv^1+v^2=0,x2(h,0)(k2x22)p=2ikU(x2)v^2,x2(h,0)tη^=ikU(0)η^+v^2(t,k,x2=0),p^(t,k,0)=(g+σk2)η^,v^2(t,k,h)=0,p^(t,k,h)=0,\begin{cases}\partial_{t}\hat{v}+ikU(x_{2})\hat{v}+(U^{\prime}(x_{2})\hat{v}_{2},0)^{T}+(ik\hat{p},\hat{p}^{\prime})^{T}=0,\qquad ik\hat{v}_{1}+\hat{v}_{2}^{\prime}=0,\qquad\qquad&x_{2}\in(-h,0)\\ (k^{2}-\partial_{x_{2}}^{2})p=2ikU^{\prime}(x_{2})\hat{v}_{2},\qquad\qquad\qquad&x_{2}\in(-h,0)\\ \partial_{t}\hat{\eta}=-ikU(0)\hat{\eta}+\hat{v}_{2}(t,k,x_{2}=0),\\ \hat{p}(t,k,0)=(g+\sigma k^{2})\hat{\eta},\\ \hat{v}_{2}(t,k,-h)=0,\quad\hat{p}^{\prime}(t,k,-h)=0,\\ \end{cases}

where denotes the derivative with respect to x2x_{2} as in the rest of the paper. Due to the divergence free condition on vv and the boundary conditions, it is easy to see

(2.7) v^2(t,0,x2)=0,p^(t,0,x2)=g,v^1(t,0,x2)=v10(0,x2),η^(t,0)=η^0(0).\hat{v}_{2}(t,0,x_{2})=0,\quad\hat{p}(t,0,x_{2})=g,\quad\hat{v}_{1}(t,0,x_{2})=v_{10}(0,x_{2}),\quad\hat{\eta}(t,0)=\hat{\eta}_{0}(0).

For k0k\neq 0, v^1\hat{v}_{1} can also be determined by v^2\hat{v}_{2} using the divergence free condition, η^\hat{\eta} by the third equation of (2.6), while p^\hat{p} by v^2\hat{v}_{2} and η^\hat{\eta} by solving the elliptic boundary value problem. So we shall mainly focus on v^2\hat{v}_{2}.

Combining the equation of v^2\hat{v}_{2} acted by k2x22k^{2}-\partial_{x_{2}}^{2} and the one of p^\hat{p} acted by x2\partial_{x_{2}}, we obtain

(2.8a) (t+ikU)(k2x22)v^2+ikU′′v^2=0,x2(h,0),(\partial_{t}+ikU)(k^{2}-\partial_{x_{2}}^{2})\hat{v}_{2}+ikU^{\prime\prime}\hat{v}_{2}=0,\quad x_{2}\in(-h,0),
which is actually the linearized transport equation of the vorticity (as defined in (2.1))
ω^=ikv^2v^1=ik(k2x22)v^2\hat{\omega}=ik\hat{v}_{2}-\hat{v}_{1}^{\prime}=\tfrac{i}{k}(k^{2}-\partial_{x_{2}}^{2})\hat{v}_{2}
in its Fourier transform. In addition to the above equation, we need its boundary information to completely determine v^2\hat{v}_{2}. Applying x2\partial_{x_{2}} to the first equation of (2.6), then evaluating at x2=0x_{2}=0, and using the equation of p^\hat{p}, we have
(t+ikU(0))v^2(t,k,0)ikU(0)v^2(t,k,0)+k2(g+σk2)η^(t,k)=0.(\partial_{t}+ikU(0))\hat{v}_{2}^{\prime}(t,k,0)-ikU^{\prime}(0)\hat{v}_{2}(t,k,0)+k^{2}(g+\sigma k^{2})\hat{\eta}(t,k)=0.
Finally applying t+ikU(0)\partial_{t}+ikU(0) to the above equation and using the third equation of (2.6), we obtain
(2.8b) ((t+ikU)2v^2ikU(t+ikU)v^2+k2(g+σk2)v^2)|x2=0=0,v^2|x2=h=0,\big{(}(\partial_{t}+ikU)^{2}\hat{v}_{2}^{\prime}-ikU^{\prime}(\partial_{t}+ikU)\hat{v}_{2}+k^{2}(g+\sigma k^{2})\hat{v}_{2}\big{)}\big{|}_{x_{2}=0}=0,\quad\hat{v}_{2}|_{x_{2}=-h}=0,

where we also included the boundary value of v^2\hat{v}_{2} at x2=hx_{2}=-h.

To analyze the evolutionary problem, we apply the Laplace transform \mathcal{L} to the unknowns

(2.9) V(s)=(V1(s),V2(s)):={v^}(s),P(s):={p}(s),η~:={η^}(s).V(s)=(V_{1}(s),V_{2}(s)):=\mathcal{L}\{\hat{v}\}(s),\quad P(s):=\mathcal{L}\{p\}(s),\quad\tilde{\eta}:=\mathcal{L}\{\hat{\eta}\}(s).

An often used change of variable for k0k\neq 0 is

(2.10) c:=is/k=cR+icIc:=is/k=c_{R}+ic_{I}

with cRc_{R} and cIc_{I} being the real and imaginary parts. From (2.8), our main unknown V2(k,c,x2)V_{2}(k,c,x_{2}) satisfies the following non-homogeneous Rayleigh equation

(2.11a) V2′′+(k2+U′′Uc)V2=(k2x22)v^20ik(Uc)=ω^0Uc,x2(h,0),-V_{2}^{\prime\prime}+(k^{2}+\frac{U^{\prime\prime}}{U-c})V_{2}=\frac{(k^{2}-\partial_{x_{2}}^{2})\hat{v}_{20}}{ik(U-c)}=-\frac{\hat{\omega}_{0}}{U-c},\qquad\qquad x_{2}\in(-h,0),
where ω^=ω^0(k,x2)\hat{\omega}=\hat{\omega}_{0}(k,x_{2}) is the Fourier transform of the initial vorticity, with the obvious boundary condition
(2.11b) V2(h)=0.V_{2}(-h)=0.
Here we skipped the kk and cc variables of V2V_{2}. Similarly, the Laplace transform applied to the boundary equation (2.8b) implies
((Uc)2V2(U(Uc)+(g+σk2))V2)|x2=0=1k2(tv^2ickv^2+2ikUv^2ikUv^2)|t=x2=0\displaystyle\big{(}(U-c)^{2}V_{2}^{\prime}-(U^{\prime}(U-c)+(g+\sigma k^{2}))V_{2}\big{)}\big{|}_{x_{2}=0}=-\tfrac{1}{k^{2}}\big{(}\partial_{t}\hat{v}_{2}^{\prime}-ick\hat{v}_{2}^{\prime}+2ikU\hat{v}_{2}^{\prime}-ikU^{\prime}\hat{v}_{2}\big{)}\big{|}_{t=x_{2}=0}
=\displaystyle= 1k2((t+ikU)v^2ikUv^2+ik(Uc)v^2)|t=x2=0.\displaystyle-\tfrac{1}{k^{2}}\big{(}(\partial_{t}+ikU)\hat{v}_{2}^{\prime}-ikU^{\prime}\hat{v}_{2}+ik(U-c)\hat{v}_{2}^{\prime}\big{)}\big{|}_{t=x_{2}=0}.
Therefore we obtain
(2.11c) ((Uc)2V2(U(Uc)+(g+σk2))V2)|x2=0=(g+σk2)η^0ik(U(0)c)v^20(0).\big{(}(U-c)^{2}V_{2}^{\prime}-(U^{\prime}(U-c)+(g+\sigma k^{2}))V_{2}\big{)}\big{|}_{x_{2}=0}=(g+\sigma k^{2})\hat{\eta}_{0}-\tfrac{i}{k}(U(0)-c)\hat{v}_{20}^{\prime}(0).

The last boundary condition can be viewed as determining the dispersion relation which is highly nonlocal. The Laplace transforms of V1V_{1} and η~\tilde{\eta} of v^1\hat{v}_{1} and η^\hat{\eta} can be recovered from the divergence free condition and the third equation of (2.6)

(2.12) V1=ikV2,η~(c,k)=V2(c,k,0)+η^0(k)ik(U(0)c).V_{1}=\tfrac{i}{k}V_{2}^{\prime},\quad\tilde{\eta}(c,k)=\frac{V_{2}(c,k,0)+\hat{\eta}_{0}(k)}{ik\big{(}U(0)-c\big{)}}.

Hence in most of the paper we shall focus on the non-homogeneous boundary value problem (2.11) of the Rayleigh equation and then use it to obtain the eigenvalue distribution of (1.3) and the inviscid damping of its solutions.

System (2.11) is a boundary value problem of a non-homogeneous second order ODE with coefficients analytic in kk\in\mathbb{R} and cU([h,0])c\in\mathbb{C}\setminus U([-h,0]), so it has a unique solution analytic in kk and cc except for those (k,c)(k,c) for which the corresponding homogeneous system of (2.11), where v^20=0\hat{v}_{20}=0 and η^0=0\hat{\eta}_{0}=0, has non-trivial solutions. Such singular (k,c)(k,c) also give the eigenvalues of (2.11) in the form of ick-ick. In fact we have the following lemma.

Lemma 2.3.

For k\{0}k\in\mathbb{R}\backslash\{0\}, there exists a non-trivial solution (c,V2(x2))\big{(}c,V_{2}(x_{2})\big{)} with cU([h,0])c\notin U([-h,0]) to the corresponding homogeneous problem of (2.11) (namely, with v^20=0\hat{v}_{20}=0 and η^0=0\hat{\eta}_{0}=0) if and only if ikc-ikc is an eigenvalue of the linearized capillary-gravity wave system (1.3) associated with the linear solution in the form of (1.5) given by

(2.13) v1(t,x)=ikeik(x1ct)V2(x2),v2(t,x)=eik(x1ct)V2(x2),η(t,x1)=eik(x1ct)V2(0)ik(U(0)c),p(t,x)=eik(x1ct)(g+σk2ik(U(0)c)V2(0)ik0x2(Uc)V2𝑑x2).\begin{split}&v_{1}(t,x)=\frac{i}{k}e^{ik(x_{1}-ct)}V_{2}^{\prime}(x_{2}),\quad v_{2}(t,x)=e^{ik(x_{1}-ct)}V_{2}(x_{2}),\quad\eta(t,x_{1})=e^{ik(x_{1}-ct)}\frac{V_{2}(0)}{ik(U(0)-c)},\\ &p(t,x)=e^{ik(x_{1}-ct)}\Big{(}\frac{g+\sigma k^{2}}{ik(U(0)-c)}V_{2}(0)-ik\int_{0}^{x_{2}}(U-c)V_{2}dx_{2}^{\prime}\Big{)}.\end{split}
Proof.

On the one hand, it is straight forward to verify that the above vv, η\eta, and pp satisfy (1.3c), (1.3d), x2p|x2=h=0\partial_{x_{2}}p|_{x_{2}=-h}=0, and v=0\nabla\cdot v=0. The Poisson equation of pp in (1.3b) is a consequence of the linearized Euler equation in (1.3a), the v2v_{2} equation of which is also easily verified. Hence we only need to consider the v1v_{1} equation in (1.3a). In fact, that equation holds for the above (v,η,p)(v,\eta,p) if

(Uc)V2+UV2+g+σk2U(0)cV2(0)+k20x2(Uc)V2𝑑x2=0.-(U-c)V_{2}^{\prime}+U^{\prime}V_{2}+\frac{g+\sigma k^{2}}{U(0)-c}V_{2}(0)+k^{2}\int_{0}^{x_{2}}(U-c)V_{2}dx_{2}^{\prime}=0.

The x2x_{2}-derivative of this function is equal to 0 due to the Rayleigh equation (2.11a) and its boundary value equal is to 0 at x2=0x_{2}=0 due to the boundary condition (2.11c).

On the other hand, suppose (k,c,v2(t,x),η(t,x1),p(t,x))\big{(}k,c,v_{2}(t,x),\eta(t,x_{1}),p(t,x)\big{)} is a solution to (1.3) in the form of (1.5) with k0k\neq 0 and cU([h,0])c\notin U([-h,0]). Equation (2.8a) implies that V2V_{2} must be a solution to the corresponding homogeneous equation of (2.11a), while (2.8b) yields the homogeneous boundary conditions of the types of (2.11b-2.11c). Therefore (c,V2(x2))(c,V_{2}(x_{2})) have to be homogeneous solutions to (2.11). Subsequently, v1v_{1} is obtained from v=0\nabla\cdot v=0, η\eta from the third equation in (2.6), and pp from the v2v_{2} equation in (2.6) along with its boundary value at x2=0x_{2}=0. ∎

Definition 2.1.

(k,c)(k,c) is a non-singular mode if cU([h,0])c\in\mathbb{C}\setminus U([-h,0]) and there exists a non-trivial solution V2(x2)V_{2}(x_{2}) to the corresponding homogeneous problem of (2.11) (thus also yields a solution to (1.3) in the form of (1.5)). (k,c)(k,c) is a singular mode if cU([h,0])c\in U([-h,0]) and there exists a Hx22H_{x_{2}}^{2} solution y(x2)y(x_{2}) to

(2.14) (Uc)(y′′+k2y)+U′′y=0(U-c)(-y^{\prime\prime}+k^{2}y)+U^{\prime\prime}y=0

along with the corresponding homogeneous boundary conditions of (2.11b2.11c). (See also Remark 4.1.)

After acquiring good understanding on the homogeneous problem of the Rayleigh equation (2.11) (Section 3) and its eigenvalues (Section 4), we proceed to analyze the general non-homogeneous problem of (2.11) (Section 5), in particular, the dependence of solutions on cc. Finally in Section 6 we apply the inverse Laplace transform to estimate the solution to the linear system (1.3). Recall the inverse Laplace transform

(2.15) f(t)=12πiγiγ+iestF(s)𝑑s=|k|2π+iγk++iγkeikctF(ikc)𝑑c,f(t)=\frac{1}{2\pi i}\int_{\gamma-i\infty}^{\gamma+i\infty}e^{st}F(s)ds=\frac{|k|}{2\pi}\int_{-\infty+\frac{i\gamma}{k}}^{+\infty+\frac{i\gamma}{k}}e^{-ikct}F(-ikc)dc,

where γ\gamma is a real number so that F(s)F(s) is analytic in the region Res>γ\text{Re}\,s>\gamma and the change of variable (2.10) was used in the second equality. Due to the analyticity, the integral can be eventually carried out along contours enclosing U([h,0])U([-h,0])\subset\mathbb{C} and the non-singular modes of (1.3). Assuming there is no singular modes in U([h,0])U([-h,0]), we shall eventually obtain the decay in tt of the component of the linear solution corresponding to the integral along the contour surrounding U([h,0])U([-h,0]) by integration by parts in cc.

3. Analysis of the Rayleigh equation

In this section, we shall thoroughly analyze the homogeneous Rayleigh equation

(3.1) y′′(x2)+(k2+U′′(x2)U(x2)c)y(x2)=0,x2[h,0],-y^{\prime\prime}(x_{2})+\big{(}k^{2}+\tfrac{U^{\prime\prime}(x_{2})}{U(x_{2})-c}\big{)}y(x_{2})=0,\quad x_{2}\in[-h,0],

where

k,c=cR+icI,=x2.k\in\mathbb{R},\quad c=c_{R}+ic_{I}\in\mathbb{C},\quad^{\prime}=\partial_{x_{2}}.

Throughout this section (except for some lemmas in Subsection 3.6), we assume

(3.2) U(x2)>0,x2[h,0].U^{\prime}(x_{2})>0,\quad\forall x_{2}\in[-h,0].

As pointed out in the introduction, due to the symmetry of the reflection in x1x_{1} variable, the case of U<0U^{\prime}<0 can be reduced to the above one. Hence all results under (3.2) hold for all uniformly monotonic U(x2)U(x_{2}), namely those UU satisfying U0U^{\prime}\neq 0 on [h,0][-h,0].

To some extent, we will also consider the non-homogeneous Rayleigh equation

(3.3) y′′(x2)+(k2+U′′(x2)U(x2)c)y(x2)=ϕ(k,c,x2),x2[h,0].-y^{\prime\prime}(x_{2})+\big{(}k^{2}+\tfrac{U^{\prime\prime}(x_{2})}{U(x_{2})-c}\big{)}y(x_{2})=\phi\big{(}k,c,x_{2}\big{)},\quad x_{2}\in[-h,0].

More detailed forms and conditions of ϕ(k,c,x2)\phi(k,c,x_{2}) will be specified when we obtained detailed estimates in Sections 5 and 6. As in typical problems of linear estimates based on density argument, we shall mostly work on ϕ\phi with sufficient regularity, but carefully tracking its norms involved in the estimates.

The solutions to the Rayleigh equation (3.1) are obviously even in kk and thus k0k\geq 0 will be assumed mostly. Similarly complex conjugate of solutions also solve (3.1) with cc replaced by c¯\bar{c}, so we will restrict our consideration to cI0c_{I}\geq 0. We have to consider the cases of cc\in\mathbb{C} away from U([h,0])U([-h,0]), near U([h,0])U([-h,0]), and then finally cU([h,0])c\in U([-h,0]), separately. Due to small scales in x2x_{2} created by k1k\gg 1, the dependence of the estimates on k1k\gg 1 will be carefully tracked.

Recall UCl0U\in C^{l_{0}}. For technical convenience we extend UU to be a Cl0C^{l_{0}} function on a neighborhood [h0h,h0][-h_{0}-h,h_{0}] of [h,0][-h,0], where

(3.4) h0=min{h2,inf[h,0]U4|U′′|C0([h,0])}>0,h_{0}=\min\Big{\{}\frac{h}{2},\,\frac{\inf_{[-h,0]}U^{\prime}}{4|U^{\prime\prime}|_{C^{0}([-h,0])}}\Big{\}}>0,

such that, on [h0h,h0][-h_{0}-h,h_{0}],

(3.5) U12inf[h,0]U(x2),|U|Cl01([h0h,h0])2|U|Cl01([h,0]).U^{\prime}\geq\tfrac{1}{2}\inf_{[-h,0]}U^{\prime}(x_{2}),\quad|U^{\prime}|_{C^{l_{0}-1}([-h_{0}-h,h_{0}])}\leq 2|U^{\prime}|_{C^{l_{0}-1}([-h,0])}.

In the analysis of the most singular case of cc close to the range U([h,0])U([-h,0]), we let x2cx_{2}^{c} be such that

(3.6) cR=U(x2c), if cRU([h0h,h0]).c_{R}=U(x_{2}^{c}),\;\text{ if }c_{R}\in U([-h_{0}-h,h_{0}]).

We also extend the non-homogeneous term ϕ(k,c,x2)\phi(k,c,x_{2}) for x2[h0h,h0]x_{2}\in[-h_{0}-h,h_{0}] while keeping its relevant bounds comparable.

3.1. Rayleigh equation in the regular region

In the initial step we consider the rather regular case where k2|Uc|k^{2}|U-c| is bounded from below. For not so small kk, we first transform the homogeneous Rayleigh equation (3.1) into a system of first order (complex valued) ODEs. Let

z±=y±|k|y,z_{\pm}=y^{\prime}\pm|k|y,

and then (3.1) takes the form of the coupled equations

(3.7) z±=±|k|z±+12β(k,c,x2)(z+z),β(k,c,x2)=U′′|k|(Uc).z_{\pm}^{\prime}=\pm|k|z_{\pm}+\tfrac{1}{2}\beta(k,c,x_{2})(z_{+}-z_{-}),\qquad\beta(k,c,x_{2})=\tfrac{U^{\prime\prime}}{|k|(U-c)}.
Lemma 3.1.

There exists C>0C>0 depending only on |U|C1|U^{\prime}|_{C^{1}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any ρ(0,1]\rho\in(0,1], k0k\neq 0, and =[x2l,x2r][h0h,h0]\mathcal{I}=[x_{2l},x_{2r}]\subset[-h_{0}-h,h_{0}] satisfying

(3.8) |1Uc|ρk2(1+|U′′|C0([h0h,h0]))1,x2,\left|\frac{1}{U-c}\right|\leq\rho k^{2}(1+|U^{\prime\prime}|_{C^{0}([-h_{0}-h,h_{0}])})^{-1},\quad\forall\,x_{2}\in\mathcal{I},

and any solution z=(z+,z)Tz=(z_{+},z_{-})^{T} to (3.7) with

(3.9) |z+(x2l)||z(x2l)|,|z_{+}(x_{2l})|\geq|z_{-}(x_{2l})|,

it holds, for x2x_{2}\in\mathcal{I}, |z+(x2)||z(x2)||z_{+}(x_{2})|\geq|z_{-}(x_{2})| and

(3.10) |z+(x2)e|k|(x2x2l)z+(x2l)|+|z(x2)e|k|(x2x2l)z(x2l)|C|k|1log(1+Cρk2(x2x2l))e|k|(x2x2l)|z+(x2l)|.\begin{split}&\left|z_{+}(x_{2})-e^{|k|(x_{2}-x_{2l})}z_{+}(x_{2l})\right|+\left|z_{-}(x_{2})-e^{-|k|(x_{2}-x_{2l})}z_{-}(x_{2l})\right|\\ \leq&C|k|^{-1}\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)}e^{|k|(x_{2}-x_{2l})}|z_{+}(x_{2l})|.\end{split}

Moreover, for any solution with

(3.11) |z+(x2r)||z(x2r)|,|z_{+}(x_{2r})|\leq|z_{-}(x_{2r})|,

we have, for x2x_{2}\in\mathcal{I}, |z+(x2)||z(x2)||z_{+}(x_{2})|\leq|z_{-}(x_{2})| and

(3.12) |z+(x2)e|k|(x2x2r)z+(x2r)|+|z(x2)e|k|(x2x2r)z(x2r)|C|k|1log(1+Cρk2(x2rx2))e|k|(x2rx2)|z(x2r)|.\begin{split}&\left|z_{+}(x_{2})-e^{|k|(x_{2}-x_{2r})}z_{+}(x_{2r})\right|+\left|z_{-}(x_{2})-e^{-|k|(x_{2}-x_{2r})}z_{-}(x_{2r})\right|\\ \leq&C|k|^{-1}\log\big{(}1+C\rho k^{2}(x_{2r}-x_{2})\big{)}e^{|k|(x_{2r}-x_{2})}|z_{-}(x_{2r})|.\end{split}

While (3.9) provides some technical convenience, indeed some assumption of this type on the initial values is needed to ensure estimates of solutions such as (3.10). For example, if |β|k|\beta|\ll k, the standard ODE theory implies that there are two solutions behaving like e±k(x2x2l)e^{\pm k(x_{2}-x_{2l})} corresponding to the Lyapunov exponents close to ±k\pm k, then the decaying solution may not satisfy (3.10) with CC uniform in k1k\gg 1.

Proof.

We start with the observation of a simple consequence of (3.8). Namely, one may compute straight forwardly

(3.13) (|z+|2|z|2)=2|k|(|z+|2+|z|2)+Reβ|z+z|20.(|z_{+}|^{2}-|z_{-}|^{2})^{\prime}=2|k|(|z_{+}|^{2}+|z_{-}|^{2})+\text{Re}\beta|z_{+}-z_{-}|^{2}\geq 0.

This monotonicity along with boundary conditions yields an order relation between |z±||z_{\pm}| which can be used to control terms in (3.7).

We shall focus on the case under assumption (3.9), which ensures

(3.14) |z+||z|,x2.|z_{+}|\geq|z_{-}|,\quad\forall\,x_{2}\in\mathcal{I}.

By factorizing z+z_{+} on the right side of (3.7), its solutions satisfy

(3.15) z+(x2)e|k|(x2x2l)z+(x2l)=(e12x2lx2β(k,c,x2)(1z(x2)z+(x2))𝑑x21)e|k|(x2x2l)z+(x2l).z_{+}(x_{2})-e^{|k|(x_{2}-x_{2l})}z_{+}(x_{2l})=\big{(}e^{\frac{1}{2}\int_{x_{2l}}^{x_{2}}\beta(k,c,x_{2}^{\prime})\big{(}1-\frac{z_{-}(x_{2}^{\prime})}{z_{+}(x_{2}^{\prime})}\big{)}dx_{2}^{\prime}}-1\big{)}e^{|k|(x_{2}-x_{2l})}z_{+}(x_{2l}).

If cRU([h0h,h0])c_{R}\in U([-h_{0}-h,h_{0}]), let x2cx_{2}^{c} be defined as in (3.6) and we use (3.8) to estimate

x2lx2|β(k,c,x2)|𝑑x2\displaystyle\int_{x_{2l}}^{x_{2}}|\beta(k,c,x_{2}^{\prime})|dx_{2}^{\prime}\leq C|k|x2lx2(|x2x2c|2+cI2)12𝑑x2=C|k||logx2x2c+(x2x2c)2+cI2x2lx2c+(x2lx2c)2+cI2|,\displaystyle\frac{C}{|k|}\int_{x_{2l}}^{x_{2}}(|x_{2}^{\prime}-x_{2}^{c}|^{2}+c_{I}^{2})^{-\frac{1}{2}}dx_{2}^{\prime}=\frac{C}{|k|}\left|\log\frac{x_{2}-x_{2}^{c}+\sqrt{(x_{2}-x_{2}^{c})^{2}+c_{I}^{2}}}{x_{2l}-x_{2}^{c}+\sqrt{(x_{2l}-x_{2}^{c})^{2}+c_{I}^{2}}}\right|,

where the last equality is the exact integral. If x2cx2lx2x_{2}^{c}\leq x_{2l}\leq x_{2}, then the numerator in the logarithm is greater than the denominator. Applying the triangle inequality to x2x_{2}, x2lx_{2l} and cc, we obtain

|logx2x2c+(x2x2c)2+cI2x2lx2c+(x2lx2c)2+cI2|\displaystyle\left|\log\frac{x_{2}-x_{2}^{c}+\sqrt{(x_{2}-x_{2}^{c})^{2}+c_{I}^{2}}}{x_{2l}-x_{2}^{c}+\sqrt{(x_{2l}-x_{2}^{c})^{2}+c_{I}^{2}}}\right|\leq |log(1+C(x2x2l)x2lx2c+|U(x2l)c|)|log(1+Cρk2(x2x2l)).\displaystyle\left|\log\left(1+\frac{C(x_{2}-x_{2l})}{x_{2l}-x_{2}^{c}+|U(x_{2l})-c|}\right)\right|\leq\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)}.

If x2lx2x2cx_{2l}\leq x_{2}\leq x_{2}^{c}, multiplying the top and bottom of the quotient by their conjugates and proceeding much as in the previous case, we have

|logx2x2c+(x2x2c)2+cI2x2lx2c+(x2lx2c)2+cI2|=\displaystyle\left|\log\frac{x_{2}-x_{2}^{c}+\sqrt{(x_{2}-x_{2}^{c})^{2}+c_{I}^{2}}}{x_{2l}-x_{2}^{c}+\sqrt{(x_{2l}-x_{2}^{c})^{2}+c_{I}^{2}}}\right|= |logx2cx2l+(x2lx2c)2+cI2x2cx2+(x2x2c)2+cI2|log(1+Cρk2(x2x2l)).\displaystyle\left|\log\frac{x_{2}^{c}-x_{2l}+\sqrt{(x_{2l}-x_{2}^{c})^{2}+c_{I}^{2}}}{x_{2}^{c}-x_{2}+\sqrt{(x_{2}-x_{2}^{c})^{2}+c_{I}^{2}}}\right|\leq\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)}.

Finally, in the case x2l<x2c<x2x_{2l}<x_{2}^{c}<x_{2}, by splitting the interval at x2cx_{2}^{c} and applying the above estimates on the two subintervals, we obtain

|logx2x2c+(x2x2c)2+cI2x2lx2c+(x2lx2c)2+cI2|\displaystyle\left|\log\frac{x_{2}-x_{2}^{c}+\sqrt{(x_{2}-x_{2}^{c})^{2}+c_{I}^{2}}}{x_{2l}-x_{2}^{c}+\sqrt{(x_{2l}-x_{2}^{c})^{2}+c_{I}^{2}}}\right|
=\displaystyle= |logx2x2c+(x2x2c)2+cI2|cI|+log|cI|x2lx2c+(x2lx2c)2+cI2|\displaystyle\left|\log\frac{x_{2}-x_{2}^{c}+\sqrt{(x_{2}-x_{2}^{c})^{2}+c_{I}^{2}}}{|c_{I}|}+\log\frac{|c_{I}|}{x_{2l}-x_{2}^{c}+\sqrt{(x_{2l}-x_{2}^{c})^{2}+c_{I}^{2}}}\right|
\displaystyle\leq log(1+Cρk2(x2x2c))+log(1+Cρk2(x2cx2l))2log(1+Cρk2(x2x2l)).\displaystyle\log\big{(}1+C\rho k^{2}(x_{2}-x_{2}^{c})\big{)}+\log\big{(}1+C\rho k^{2}(x_{2}^{c}-x_{2l})\big{)}\leq 2\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)}.

Therefore the desired estimate (3.10) on z+z_{+} follows from (3.15) and (3.14) and

|e12x2lx2β(k,c,x2)(1z(x2)z+(x2))𝑑x21|Cx2lx2|β(k,c,x2)|𝑑x2C|k|1log(1+Cρk2(x2x2l)),\displaystyle\left|e^{\frac{1}{2}\int_{x_{2l}}^{x_{2}}\beta(k,c,x_{2}^{\prime})\big{(}1-\frac{z_{-}(x_{2}^{\prime})}{z_{+}(x_{2}^{\prime})}\big{)}dx_{2}^{\prime}}-1\right|\leq C\int_{x_{2l}}^{x_{2}}|\beta(k,c,x_{2}^{\prime})|dx_{2}^{\prime}\leq C|k|^{-1}\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)},

as C|k|1log(1+Cρk2(x2x2l))C|k|^{-1}\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)} is bounded uniformly in all k0k\neq 0. If cRU([h0h,h0])c_{R}\notin U([-h_{0}-h,h_{0}]), one can bound |β||\beta| by C|k|min{1,ρk2}\tfrac{C}{|k|}\min\{1,\rho k^{2}\} which is also bounded for all k0k\neq 0. If ρk21\rho k^{2}\leq 1, then ρk2(x2x2l)\rho k^{2}(x_{2}-x_{2l}) is bounded by Clog(1+ρk2(x2x2l))C\log\big{(}1+\rho k^{2}(x_{2}-x_{2l})\big{)}. If 1ρk21\leq\rho k^{2}, then

x2x2lClog(1+x2x2l)Clog(1+ρk2(x2x2l)).x_{2}-x_{2l}\leq C\log\big{(}1+x_{2}-x_{2l}\big{)}\leq C\log\big{(}1+\rho k^{2}(x_{2}-x_{2l})\big{)}.

Therefore in both cases we have

x2lx2|β(k,c,x2)|𝑑x2C|k|min{1,ρk2}(x2x2l)C|k|log(1+ρk2(x2x2l))\int_{x_{2l}}^{x_{2}}|\beta(k,c,x_{2}^{\prime})|dx_{2}^{\prime}\leq\tfrac{C}{|k|}\min\{1,\rho k^{2}\}(x_{2}-x_{2l})\leq\tfrac{C}{|k|}\log\big{(}1+\rho k^{2}(x_{2}-x_{2l})\big{)}

and thus (3.10) for z+z_{+} follows from (3.15) and (3.14).

Turning attention to zz_{-}, from the variation of parameter formula, we have

(3.16) z(x2)e|k|(x2x2l)z(x2l)=12x2lx2e|k|(x2x2)β(k,c,x2)(z+(x2)z(x2))𝑑x2,z_{-}(x_{2})-e^{-|k|(x_{2}-x_{2l})}z_{-}(x_{2l})=\frac{1}{2}\int_{x_{2l}}^{x_{2}}e^{-|k|(x_{2}-x_{2}^{\prime})}\beta(k,c,x_{2}^{\prime})\big{(}z_{+}(x_{2}^{\prime})-z_{-}(x_{2}^{\prime})\big{)}dx_{2}^{\prime},

which along with (3.8), (3.10) for z+z_{+}, and (3.14), implies

|z(x2)e|k|(x2x2l)z(x2l)|Ce|k|(x2x2l)|z+(x2l)|x2lx2|β(k,c,x2)|𝑑x2.\displaystyle|z_{-}(x_{2})-e^{-|k|(x_{2}-x_{2l})}z_{-}(x_{2l})|\leq Ce^{|k|(x_{2}-x_{2l})}|z_{+}(x_{2l})|\int_{x_{2l}}^{x_{2}}|\beta(k,c,x_{2}^{\prime})|dx_{2}^{\prime}.

The desired estimate on zz_{-} follows from the above inequality on |β|\int|\beta|. The estimates on z±(x2)z_{\pm}(x_{2}) with initial condition z±(x2r)z_{\pm}(x_{2r}) satisfying (3.11) can be derived in exactly the same fashion. ∎

In the following we use the above lemma to analyze some solutions to the homogeneous and non-homogeneous Rayleigh equations (3.1) and (3.3).

Lemma 3.2.

Consider

(Θ1,Θ2){sinh,cosh}2{(cosh,sinh)}.(\Theta_{1},\Theta_{2})\in\{\sinh,\cosh\}^{2}\setminus\{(\cosh,\sinh)\}.

There exists C>0C>0 depending only on |U|C1|U^{\prime}|_{C^{1}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any k0k\neq 0, ρ(0,1]\rho\in(0,1], C00C_{0}\geq 0, and interval =[x2l,x2r][h,0]\mathcal{I}=[x_{2l},x_{2r}]\subset[-h,0] satisfying (3.8),

  1. (1)

    if a solution y(x2)y(x_{2}) to (3.1) satisfies

    (3.17) ||k|y(x2l)sinh|k|s|C0Θ1(|k|s),|y(x2l)coshks|C0Θ2(|k|s),s0,\big{|}|k|y(x_{2l})-\sinh|k|s\big{|}\leq C_{0}\Theta_{1}(|k|s),\;\;|y^{\prime}(x_{2l})-\cosh ks|\leq C_{0}\Theta_{2}(|k|s),\quad s\geq 0,

    then it holds that, for all x2x_{2}\in\mathcal{I},

    ||k|y(x2)sinh|k|(x2x2l+s)|C(C0+(1+C0)(ρ+|k|1log(1+Cρk2)))Θ1(|k|(x2x2l+s)),\big{|}|k|y(x_{2})-\sinh|k|(x_{2}-x_{2l}+s)|\leq C\big{(}C_{0}+(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}\big{)}\Theta_{1}(|k|(x_{2}-x_{2l}+s)),
    |y(x2)coshk(x2x2l+s)|C(C0+(1+C0)(ρ+|k|1log(1+Cρk2)))Θ2(|k|(x2x2l+s));|y^{\prime}(x_{2})-\cosh k(x_{2}-x_{2l}+s)|\leq C\big{(}C_{0}+(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}\big{)}\Theta_{2}(|k|(x_{2}-x_{2l}+s));
  2. (2)

    if a solution y(x2)y(x_{2}) to (3.1) satisfies

    (3.18) ||k|y(x2r)sinh|k|s|C0Θ1(|ks|),|y(x2r)coshks|C0Θ2(|ks|),s0,\big{|}|k|y(x_{2r})-\sinh|k|s\big{|}\leq C_{0}\Theta_{1}(|ks|),\;\;|y^{\prime}(x_{2r})-\cosh ks|\leq C_{0}\Theta_{2}(|ks|),\quad s\leq 0,

    then it holds that, for all x2x_{2}\in\mathcal{I},

    ||k|y(x2)sinh|k|(x2x2r+s)|C(C0+(1+C0)(ρ+|k|1log(1+Cρk2)))Θ1(|k(x2x2r+s)|),\big{|}|k|y(x_{2})-\sinh|k|(x_{2}-x_{2r}+s)\big{|}\leq C\big{(}C_{0}+(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}\big{)}\Theta_{1}(|k(x_{2}-x_{2r}+s)|),
    |y(x2)coshk(x2x2r+s)|C(C0+(1+C0)(ρ+|k|1log(1+Cρk2)))Θ2(|k(x2x2r+s)|).|y^{\prime}(x_{2})-\cosh k(x_{2}-x_{2r}+s)|\leq C\big{(}C_{0}+(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}\big{)}\Theta_{2}(|k(x_{2}-x_{2r}+s)|).
  3. (3)

    Moreover, the solution y(x2)y(x_{2}) to (3.3) with y(x20)=y(x20)=0y(x_{20})=y^{\prime}(x_{20})=0 for some x20x_{20}\in\mathcal{I} satisfies

    (3.19) ||k|y(x2)x20x2ϕ(k,c,x2)sinh|k(x2x2)|dx2|+|y(x2)x20x2ϕ(k,c,x2)coshk(x2x2)𝑑x2|C(ρ+|k|1log(1+Cρk2))|x20x2ϕ(k,c,x2)sinh|k(x2x2)|dx2|.\begin{split}&\Big{|}|k|y(x_{2})-\int_{x_{20}}^{x_{2}}\phi(k,c,x_{2}^{\prime})\sinh|k(x_{2}-x_{2}^{\prime})|dx_{2}^{\prime}\Big{|}\\ &+\Big{|}y^{\prime}(x_{2})-\int_{x_{20}}^{x_{2}}\phi(k,c,x_{2}^{\prime})\cosh k(x_{2}-x_{2}^{\prime})dx_{2}^{\prime}\Big{|}\\ \leq&C\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}\Big{|}\int_{x_{20}}^{x_{2}}\phi\big{(}k,c,x_{2}^{\prime}\big{)}\sinh|k(x_{2}-x_{2}^{\prime})|dx_{2}^{\prime}\Big{|}.\end{split}
Proof.

We first consider the special solution y(x2)y(x_{2}) to the homogeneous (3.1) satisfying (3.17) with C0=0C_{0}=0, namely, with the initial values

y(x2l)=|k|1sinh|k|s,y(x2l)=cosh|k|s,s0,y(x_{2l})=|k|^{-1}\sinh|k|s,\;\;y^{\prime}(x_{2l})=\cosh|k|s,\quad s\geq 0,

whose corresponding form in terms of z±z_{\pm} with initial values z±(x2l)=e±|k|sz_{\pm}(x_{2l})=e^{\pm|k|s} satisfies the assumptions of Lemma 3.1. On the one hand, for |k|(x2x2l)1|k|(x_{2}-x_{2l})\leq 1, it holds

|k|1e|k|(x2x2l)log(1+Cρk2(x2x2l))Cρ|k|(x2x2l)Cρsinh|k|(x2x2l),\displaystyle|k|^{-1}e^{|k|(x_{2}-x_{2l})}\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)}\leq C\rho|k|(x_{2}-x_{2l})\leq C\rho\sinh|k|(x_{2}-x_{2l}),

while, for |k|(x2x2l)1|k|(x_{2}-x_{2l})\geq 1, we have

|k|1e|k|(x2x2l)log(1+Cρk2(x2x2l))C|k|1log(1+Cρk2)sinh|k|(x2x2l).\displaystyle|k|^{-1}e^{|k|(x_{2}-x_{2l})}\log\big{(}1+C\rho k^{2}(x_{2}-x_{2l})\big{)}\leq C|k|^{-1}\log(1+C\rho k^{2})\sinh|k|(x_{2}-x_{2l}).

Therefore Lemma 3.1 and ρ(0,1]\rho\in(0,1] imply

|z+(x2)e|k|(x2x2l+s)|\displaystyle|z_{+}(x_{2})-e^{|k|(x_{2}-x_{2l}+s)}| +|z(x2)e|k|(x2x2l+s)|\displaystyle+|z_{-}(x_{2})-e^{-|k|(x_{2}-x_{2l}+s)}|
\displaystyle\leq C(ρ+|k|1log(1+Cρk2))e|k|ssinh|k|(x2x2l).\displaystyle C\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}e^{|k|s}\sinh|k|(x_{2}-x_{2l}).

Recovering y(x2)y(x_{2}) and y(x2)y^{\prime}(x_{2}) from z±(x2)z_{\pm}(x_{2}), we obtain the desired estimates in the case of Θ1=Θ2=sinh\Theta_{1}=\Theta_{2}=\sinh under the additional assumption C0=0C_{0}=0.

In the following we prove the estimates for a homogeneous solution y(x2)y(x_{2}) to (3.1) under (3.17) with general C00C_{0}\geq 0. Let Y1(x2)Y_{1}(x_{2}) and Y2(x2)Y_{2}(x_{2}) be solution to (3.1) with initial values

Y1(x2l)=|k|1sinh1,Y1(x2l)=cosh1;Y2(x2l)=0,Y2(x2l)=1.Y_{1}(x_{2l})=|k|^{-1}\sinh 1,\;\;Y_{1}^{\prime}(x_{2l})=\cosh 1;\quad Y_{2}(x_{2l})=0,\;\;Y_{2}^{\prime}(x_{2l})=1.

Clearly Y1Y_{1} and Y2Y_{2} satisfy the above estimates with s=|k|1s=|k|^{-1} and s=0s=0, respectively, and

y(x2)=|k|(sinh1)1y(x2l)Y1(x2)+(y(x2l)|k|(coth1)y(x2l))Y2(x2).y(x_{2})=|k|(\sinh 1)^{-1}y(x_{2l})Y_{1}(x_{2})+\big{(}y^{\prime}(x_{2l})-|k|(\coth 1)y(x_{2l})\big{)}Y_{2}(x_{2}).

Therefore, for x2x_{2}\in\mathcal{I},

||k|y\displaystyle\big{|}|k|y (x2)sinh|k|(x2x2l+s)|=|(sinh1)1(|k|y(x2l)(|k|Y1(x2)sinh(|k|(x2x2l)+1))\displaystyle(x_{2})-\sinh|k|(x_{2}-x_{2l}+s)\big{|}=\Big{|}(\sinh 1)^{-1}\Big{(}|k|y(x_{2l})\big{(}|k|Y_{1}(x_{2})-\sinh(|k|(x_{2}-x_{2l})+1)\big{)}
+(|k|y(x2l)sinh|k|s)sinh(|k|(x2x2l)+1))+(sinh1)1sinh|k|ssinh(|k|(x2x2l)+1)\displaystyle+(|k|y(x_{2l})-\sinh|k|s)\sinh(|k|(x_{2}-x_{2l})+1)\Big{)}+(\sinh 1)^{-1}\sinh|k|s\sinh(|k|(x_{2}-x_{2l})+1)
+(y(x2l)(coth1)|k|y(x2l))(|k|Y2(x2)sinh|k|(x2x2l))\displaystyle+\big{(}y^{\prime}(x_{2l})-(\coth 1)|k|y(x_{2l})\big{)}\big{(}|k|Y_{2}(x_{2})-\sinh|k|(x_{2}-x_{2l})\big{)}
+(y(x2l)cosh|k|s(coth1)(|k|y(x2l)sinh|k|s))sinh|k|(x2x2l)\displaystyle+\big{(}y^{\prime}(x_{2l})-\cosh|k|s-(\coth 1)(|k|y(x_{2l})-\sinh|k|s)\big{)}\sinh|k|(x_{2}-x_{2l})
+(cosh|k|s(coth1)sinh|k|s)sinh|k|(x2x2l)sinh|k|(x2x2l+s)|.\displaystyle+(\cosh|k|s-(\coth 1)\sinh|k|s)\sinh|k|(x_{2}-x_{2l})-\sinh|k|(x_{2}-x_{2l}+s)\Big{|}.

In the above summation, all the hyperbolic trigonometric combinations without y(x2l)y(x_{2l}) or Y1,2(x2)Y_{1,2}(x_{2}) are eventually cancelled and the remaining terms can be estimated by the using the assumptions on the initial values and the already obtained estimates on Y1Y_{1} and Y2Y_{2}. We have

||k|y(x2)sinh|k|(x2x2l+s)|((1+C0)(ρ+|k|1log(1+Cρk2))+C0)\displaystyle\big{|}|k|y(x_{2})-\sinh|k|(x_{2}-x_{2l}+s)\big{|}\leq\big{(}(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}+C_{0}\big{)}
×(Θ1(|k|s)sinh(|k|(x2x2l)+1)+cosh|k|ssinh|k|(x2x2l))\displaystyle\qquad\qquad\qquad\qquad\times\big{(}\Theta_{1}(|k|s)\sinh(|k|(x_{2}-x_{2l})+1)+\cosh|k|s\sinh|k|(x_{2}-x_{2l})\big{)}
\displaystyle\leq ((1+C0)(ρ+|k|1log(1+Cρk2))+C0)Θ1|k|(x2x2l+s),\displaystyle\big{(}(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}+C_{0}\big{)}\Theta_{1}|k|(x_{2}-x_{2l}+s),

where the last inequality was obtained by considering the two possible cases of Θ1\Theta_{1} spearately. The inequality on y(x2)y^{\prime}(x_{2}) can be obtained similarly as

|y\displaystyle\big{|}y^{\prime} (x2)cosh|k|(x2x2l+s)||(sinh1)1(|k|y(x2l)(|k|Y1(x2)cosh(|k|(x2x2l)+1))\displaystyle(x_{2})-\cosh|k|(x_{2}-x_{2l}+s)\big{|}\leq\Big{|}(\sinh 1)^{-1}\Big{(}|k|y(x_{2l})\big{(}|k|Y_{1}^{\prime}(x_{2})-\cosh(|k|(x_{2}-x_{2l})+1)\big{)}
+(|k|y(x2l)sinh|k|s)cosh(|k|(x2x2l)+1))+(sinh1)1sinh|k|scosh(|k|(x2x2l)+1)\displaystyle+(|k|y(x_{2l})-\sinh|k|s)\cosh(|k|(x_{2}-x_{2l})+1)\Big{)}+(\sinh 1)^{-1}\sinh|k|s\cosh(|k|(x_{2}-x_{2l})+1)
+(y(x2l)(coth1)|k|y(x2l))(|k|Y2(x2)cosh|k|(x2x2l))\displaystyle+\big{(}y^{\prime}(x_{2l})-(\coth 1)|k|y(x_{2l})\big{)}\big{(}|k|Y_{2}^{\prime}(x_{2})-\cosh|k|(x_{2}-x_{2l})\big{)}
+(y(x2l)cosh|k|s(coth1)(|k|y(x2l)sinh|k|s))cosh|k|(x2x2l)\displaystyle+\big{(}y^{\prime}(x_{2l})-\cosh|k|s-(\coth 1)(|k|y(x_{2l})-\sinh|k|s)\big{)}\cosh|k|(x_{2}-x_{2l})
+(cosh|k|s(coth1)sinh|k|s)cosh|k|(x2x2l)cosh|k|(x2x2l+s)|\displaystyle+(\cosh|k|s-(\coth 1)\sinh|k|s)\cosh|k|(x_{2}-x_{2l})-\cosh|k|(x_{2}-x_{2l}+s)\Big{|}

and thus

|y(x2)cosh|k|(x2x2l+s)|((1+C0)(ρ+|k|1log(1+Cρk2))+C0)\displaystyle\big{|}y^{\prime}(x_{2})-\cosh|k|(x_{2}-x_{2l}+s)\big{|}\leq\big{(}(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}+C_{0}\big{)}
×(Θ1(|k|s)cosh(|k|(x2x2l)+1)+cosh|k|ssinh|k|(x2x2l)+Θ2(|k|s)cosh|k|(x2x2l))\displaystyle\quad\times\big{(}\Theta_{1}(|k|s)\cosh(|k|(x_{2}-x_{2l})+1)+\cosh|k|s\sinh|k|(x_{2}-x_{2l})+\Theta_{2}(|k|s)\cosh|k|(x_{2}-x_{2l})\big{)}
\displaystyle\leq ((1+C0)(ρ+|k|1log(1+Cρk2))+C0)Θ2|k|(x2x2l+s).\displaystyle\big{(}(1+C_{0})\big{(}\rho+|k|^{-1}\log(1+C\rho k^{2})\big{)}+C_{0}\big{)}\Theta_{2}|k|(x_{2}-x_{2l}+s).

This proves the desired estimates under the assumption (3.17). The proofs of the inequalities under assumption (3.18) are similar and we omit the details.

Using the variation of parameter formula, we can write the solution y(x2)y(x_{2}) with y(x20)=y(x20)=0y(x_{20})=y^{\prime}(x_{20})=0 to the non-homogeneous Rayleigh equation (3.3) as

(yy)(x2)=x20x2ϕ(k,c,x2)S(x2,x2)(01)𝑑x2\begin{pmatrix}y\\ y^{\prime}\end{pmatrix}(x_{2})=\int_{x_{20}}^{x_{2}}\phi\big{(}k,c,x_{2}^{\prime}\big{)}S(x_{2},x_{2}^{\prime})\begin{pmatrix}0\\ 1\end{pmatrix}dx_{2}^{\prime}

where S(x2,x2)S(x_{2},x_{2}^{\prime}) is the 2×22\times 2 fundamental matrix of the homogeneous equation (3.1) with initial value S(x2,x2)=IS(x_{2}^{\prime},x_{2}^{\prime})=I. Therefore,

S(x2,x2)(01)=(y~(x2,x2)y~(x2,x2))S(x_{2},x_{2}^{\prime})\begin{pmatrix}0\\ 1\end{pmatrix}=\begin{pmatrix}\tilde{y}(x_{2},x_{2}^{\prime})\\ \tilde{y}^{\prime}(x_{2},x_{2}^{\prime})\end{pmatrix}

where y~(,x2)\tilde{y}(\cdot,x_{2}^{\prime}) is the solution to (3.1) whose initial value is given by y~(x2,x2)=0\tilde{y}(x_{2}^{\prime},x_{2}^{\prime})=0 and y~(x2,x2)=1\tilde{y}(x_{2}^{\prime},x_{2}^{\prime})=1. The desired estimates follow from applying the above estimates in the homogeneous case with s=0=C0s=0=C_{0} and Θ1=Θ2=sinh\Theta_{1}=\Theta_{2}=\sinh. ∎

Practically the above estimates are more effective for kk bounded from below. To end this subsection, we give the following simple estimate of the Rayleigh equation for kk bounded from above, which compares y(x2)y(x_{2}) to the free solution (where the UU term is removed)

yF(x2)=(coshk(x2x20))y(x20)+k1(sinhk(x2x20))y(x20).y_{F}(x_{2})=\big{(}\cosh k(x_{2}-x_{20})\big{)}y(x_{20})+k^{-1}\big{(}\sinh k(x_{2}-x_{20})\big{)}y^{\prime}(x_{20}).

Here k1sinhks|k=0=sk^{-1}\sinh ks|_{k=0}=s is understood.

Lemma 3.3.

For any k,M>0k^{*},M>0, there exists C>0C>0 depending only on hh, kk^{*}, and MM such that for any |k|k|k|\leq k^{*}, C0>0C_{0}>0, x20=[x2l,x2r][h,0]x_{20}\in\mathcal{I}=[x_{2l},x_{2r}]\subset[-h,0] satisfying

|1Uc|C0M,x2,\left|\frac{1}{U-c}\right|\leq C_{0}\leq M,\quad\forall\,x_{2}\in\mathcal{I},

and any solution y(x2)y(x_{2}) to (3.3), it holds

|y(x2)yF(x2)|+|y(x2)yF(x2)|C(\displaystyle|y(x_{2})-y_{F}(x_{2})|+|y^{\prime}(x_{2})-y_{F}^{\prime}(x_{2})|\leq C\Big{(} C0(|y(x20)||x2x20|+|y(x20)||x2x20|2)\displaystyle C_{0}\big{(}|y(x_{20})||x_{2}-x_{20}|+|y^{\prime}(x_{20})||x_{2}-x_{20}|^{2}\big{)}
+|x20x2|ϕ(k,c,x2)|dx2|).\displaystyle+\Big{|}\int_{x_{20}}^{x_{2}}|\phi(k,c,x_{2}^{\prime})|dx_{2}^{\prime}\Big{|}\Big{)}.
Proof.

The proof is based on some straight forward elementary argument and we shall only outline it. Let y~=yyF\tilde{y}=y-y_{F}. We can write the solution y(x2)y(x_{2}) using the variation of constant formula

(y~(x2)y~(x2))=\displaystyle\begin{pmatrix}\tilde{y}(x_{2})\\ \tilde{y}^{\prime}(x_{2})\end{pmatrix}= x20x2(U′′yFUcϕ)(x2)(k1sinhk(x2x2)coshk(x2x2))𝑑x2\displaystyle\int_{x_{20}}^{x_{2}}\left(\frac{U^{\prime\prime}y_{F}}{U-c}-\phi\right)(x_{2}^{\prime})\begin{pmatrix}k^{-1}\sinh k(x_{2}-x_{2}^{\prime})\\ \cosh k(x_{2}-x_{2}^{\prime})\end{pmatrix}dx_{2}^{\prime}
+x20x2(U′′y~Uc)(x2)(k1sinhk(x2x2)coshk(x2x2))𝑑x2.\displaystyle+\int_{x_{20}}^{x_{2}}\left(\frac{U^{\prime\prime}\tilde{y}}{U-c}\right)(x_{2}^{\prime})\begin{pmatrix}k^{-1}\sinh k(x_{2}-x_{2}^{\prime})\\ \cosh k(x_{2}-x_{2}^{\prime})\end{pmatrix}dx_{2}^{\prime}.

It implies

|y~(x2)|+|y~(x2)|\displaystyle|\tilde{y}(x_{2})|+|\tilde{y}^{\prime}(x_{2})|\leq C(C0(|y(x20)||x2x20|+|y(x20)||x2x20|2)+|x20x2|ϕ(k,c,x2)|dx2|)\displaystyle C\Big{(}C_{0}\big{(}|y(x_{20})||x_{2}-x_{20}|+|y^{\prime}(x_{20})||x_{2}-x_{20}|^{2}\big{)}+\left|\int_{x_{20}}^{x_{2}}|\phi(k,c,x_{2}^{\prime})|dx_{2}^{\prime}\right|\Big{)}
+CC0|x20x2|y~(x2)|dx2|\displaystyle+CC_{0}\left|\int_{x_{20}}^{x_{2}}|\tilde{y}(x_{2}^{\prime})|dx_{2}^{\prime}\right|

and the estimates on yyFy-y_{F} and yyFy^{\prime}-y_{F}^{\prime} follow immediately from the Gronwall inequality. ∎

3.2. Rayleigh equation near singularity and its convergence as cI0+c_{I}\to 0+

In the rest of the section, we shall mostly focus on the case when (1+k2)12|Uc|(1+k^{2})^{\frac{1}{2}}|U-c| is small, so

(3.20) cR=U(x2c),x2c[12h0h,12h0],c_{R}=U(x_{2}^{c}),\quad x_{2}^{c}\in[-\tfrac{1}{2}h_{0}-h,\tfrac{1}{2}h_{0}],

will always be assumed, while the domains of UU and ϕ\phi have been extended to [h0h,h0][-h_{0}-h,h_{0}]. Due to complex conjugacy, we only need to consider cI0c_{I}\geq 0. In particular, if x2c(h,x2)x_{2}^{c}\in(-h,x_{2}), the strong singularity in (3.1) will lead to y(cR+i(0+),k,x2)y\big{(}c_{R}+i(0+),k,x_{2}\big{)}\notin\mathbb{R} even if y(h),y(h)y(-h),y^{\prime}(-h)\in\mathbb{R}. Even though some estimates are stated for cI>0c_{I}>0, most of the inequalities are mostly uniform as cI0+c_{I}\to 0+ and thus hold for the limits.

In order to obtain estimates uniform in kk\in\mathbb{R}, rescale

(3.21) μ=k1=1k2+1,x2=x2c+μτ,cI=μϵ,w=(w1,w2)T=(μ1y,y)T2,\mu=\langle k\rangle^{-1}=\tfrac{1}{\sqrt{k^{2}+1}},\quad x_{2}=x_{2}^{c}+\mu\tau,\quad c_{I}=\mu\epsilon,\quad w=(w_{1},w_{2})^{T}=(\mu^{-1}y,y^{\prime})^{T}\in\mathbb{C}^{2},

where x2cx_{2}^{c} satisfies (3.20) as well as in the above. Equation (3.1) becomes

(3.22) wτ=(011μ2+μ2U′′(x2c+μτ)U(x2c+μτ)c0)w(0ϕ~(μ,c,τ)),w_{\tau}=\begin{pmatrix}0&1\\ 1-\mu^{2}+\frac{\mu^{2}U^{\prime\prime}(x_{2}^{c}+\mu\tau)}{U(x_{2}^{c}+\mu\tau)-c}&0\end{pmatrix}w-\begin{pmatrix}0\\ \tilde{\phi}(\mu,c,\tau)\end{pmatrix},

where

ϕ~(μ,c,τ)=μϕ(k,c,x2c+μτ).\tilde{\phi}(\mu,c,\tau)=\mu\phi\big{(}k,c,x_{2}^{c}+\mu\tau\big{)}.

We shall consider this ODE on intervals τ[M,M]\tau\in[-M,M] such that

(3.23) [x2cμM,x2c+μM][h0h,h0],[x_{2}^{c}-\mu M,x_{2}^{c}+\mu M]\subset[-h_{0}-h,h_{0}],

is that UU is well-defined when |τ|M|\tau|\leq M. As cI0+c_{I}\to 0+, one would naturally expect w(τ)w(\tau) to converge to solutions to

(3.24) Wτ=(011μ2+μ2U′′(x2c+μτ)U(x2c+μτ)cR0)W(0ϕ~(μ,cR,τ)).W_{\tau}=\begin{pmatrix}0&1\\ 1-\mu^{2}+\frac{\mu^{2}U^{\prime\prime}(x_{2}^{c}+\mu\tau)}{U(x_{2}^{c}+\mu\tau)-c_{R}}&0\end{pmatrix}W-\begin{pmatrix}0\\ \tilde{\phi}\big{(}\mu,c_{R},\tau\big{)}\end{pmatrix}.

However, this limit equation becomes singular at τ=0\tau=0 and conditions have to be specified there.

\bullet Fundamental matrix of the homogeneous Rayleigh equation. Its construction is adapted from the one used in [5]. Let

(3.25) Γ(μ,cR,ϵ,τ)=(1μ2)τ+μU′′(x2c)2U(x2c)log(U~2+ϵ2)+γ(μ,cR,ϵ,τ)+MτiμϵU2(τ)U~(τ)2+ϵ2𝑑τ,\Gamma(\mu,c_{R},\epsilon,\tau)=(1-\mu^{2})\tau+\frac{\mu U^{\prime\prime}(x_{2}^{c})}{2U^{\prime}(x_{2}^{c})}\log(\tilde{U}^{2}+\epsilon^{2})+\gamma(\mu,c_{R},\epsilon,\tau)+\int_{-M}^{\tau}\frac{i\mu\epsilon U_{2}(\tau^{\prime})}{\tilde{U}(\tau^{\prime})^{2}+\epsilon^{2}}d\tau^{\prime},

where, for j=1,2,3j=1,2,3,

(3.26) Uj(cR,μ,τ)=(djdx2jU)(x2c+μτ),U~(cR,μ,τ)=1μ(U(x2c+μτ)cR)=1μ(U(x2c+μτ)U(x2c)),U_{j}(c_{R},\mu,\tau)=(\tfrac{d^{j}}{dx_{2}^{j}}U)(x_{2}^{c}+\mu\tau),\;\;\tilde{U}(c_{R},\mu,\tau)=\tfrac{1}{\mu}\big{(}U(x_{2}^{c}+\mu\tau)-c_{R}\big{)}=\tfrac{1}{\mu}\big{(}U(x_{2}^{c}+\mu\tau)-U(x_{2}^{c})\big{)},

and the remainder γ\gamma of Γ\Gamma is given by

(3.27) γ(μ,cR,ϵ,0)=0,γτ=μ(U1(0)U2U2(0)U1)U~U1(0)(U~2+ϵ2),Γτ=1μ2+μU2U~iϵ.\gamma(\mu,c_{R},\epsilon,0)=0,\quad\gamma_{\tau}=\frac{\mu\big{(}U_{1}(0)U_{2}-U_{2}(0)U_{1}\big{)}\tilde{U}}{U_{1}(0)(\tilde{U}^{2}+\epsilon^{2})},\;\Longrightarrow\;\Gamma_{\tau}=1-\mu^{2}+\frac{\mu U_{2}}{\tilde{U}-i\epsilon}.

It is not hard to see that γ(μ,cR,0,τ)\gamma(\mu,c_{R},0,\tau) is Cl02C^{l_{0}-2} in τ\tau and μ\mu and Cl03C^{l_{0}-3} in cRc_{R}. We often skip writing the explicit dependence on those variables other than τ\tau. Denote

(3.28) Γ0(μ,cR,τ)=limϵ0+Γ(μ,cR,ϵ,τ)=(1μ2)τ+μU′′(x2c)U(x2c)log|U~(τ)|+γ(μ,cR,0,τ)+iπμU′′(x2c)2U(x2c)(sgn(τ)+1),\begin{split}\Gamma_{0}(\mu,c_{R},\tau)=&\lim_{\epsilon\to 0+}\Gamma(\mu,c_{R},\epsilon,\tau)\\ =&(1-\mu^{2})\tau+\frac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\log|\tilde{U}(\tau)|+\gamma(\mu,c_{R},0,\tau)+\frac{i\pi\mu U^{\prime\prime}(x_{2}^{c})}{2U^{\prime}(x_{2}^{c})}\big{(}sgn(\tau)+1\big{)},\end{split}

where we note that the integrand of the imaginary part of Γ\Gamma converges to a delta mass as ϵ0+\epsilon\to 0+ and produces a jump in Γ0\Gamma_{0} at τ=0\tau=0 (see Lemma 3.4 in the below). Let B~(μ,cR,ϵ,τ)\tilde{B}(\mu,c_{R},\epsilon,\tau) be a 2×22\times 2 matrix given by

(3.29) B~τ=(Γ(μ,cR,ϵ,τ)1Γ(μ,cR,ϵ,τ)2Γ(μ,cR,ϵ,τ))B~,B~(μ,cR,ϵ,0)=I2×2,\tilde{B}_{\tau}=\begin{pmatrix}\Gamma(\mu,c_{R},\epsilon,\tau)&1\\ -\Gamma(\mu,c_{R},\epsilon,\tau)^{2}&-\Gamma(\mu,c_{R},\epsilon,\tau)\end{pmatrix}\tilde{B},\quad\tilde{B}(\mu,c_{R},\epsilon,0)=I_{2\times 2},

and

(3.30) Φ~(μ,c,τ)=(Φ~1(μ,c,τ)Φ~2(μ,c,τ))=Mτϕ~(μ,c,τ)B~(μ,cR,cIμ,τ)1(01)𝑑τ.\tilde{\Phi}(\mu,c,\tau)=\begin{pmatrix}\tilde{\Phi}_{1}(\mu,c,\tau)\\ \tilde{\Phi}_{2}(\mu,c,\tau)\end{pmatrix}=\int_{-M}^{\tau}\tilde{\phi}(\mu,c,\tau^{\prime})\tilde{B}(\mu,c_{R},\tfrac{c_{I}}{\mu},\tau^{\prime})^{-1}\begin{pmatrix}0\\ 1\end{pmatrix}d\tau^{\prime}.

It is worth pointing out that Γ0\Gamma_{0} is real for τ<0\tau<0 and imaginary for τ>0\tau>0. To keep the notations simple we often skip the arguments other than τ\tau. In the following lemma we collect some basic estimates of Γ\Gamma and B~\tilde{B} where we often bound the log|τ|\log|\tau| singularity in Γ\Gamma by |τ|α|\tau|^{-\alpha}, α>0\alpha>0, for simplicity.

Lemma 3.4.

For any M>0M>0 satisfying (3.23) and α,α(0,1)\alpha,\alpha^{\prime}\in(0,1) with α+α<1\alpha+\alpha^{\prime}<1, there exists C>0C>0 depending only on MM, α\alpha, α\alpha^{\prime}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any 0<ϵ<M0<\epsilon<M, the following hold for |τ|M|\tau|\leq M,

(3.31) detB~=1,|B~I|e|τ|+C(|τ|3+μ2|τ|α)1,|B~1I|4(e|τ|+C(|τ|3+μ2|τ|α)1).\det\tilde{B}=1,\quad|\tilde{B}-I|\leq e^{|\tau|+C(|\tau|^{3}+\mu^{2}|\tau|^{\alpha})}-1,\quad|\tilde{B}^{-1}-I|\leq 4(e^{|\tau|+C(|\tau|^{3}+\mu^{2}|\tau|^{\alpha})}-1).
(3.32) |Γ(ϵ,τ)Γ0(τ)|Cμ(μϵ|logϵ|+ϵϵ+|τ|+log(1+Cϵ2τ2))|\Gamma(\epsilon,\tau)-\Gamma_{0}(\tau)|\leq C\mu\big{(}\mu\epsilon|\log\epsilon|+\tfrac{\epsilon}{\epsilon+|\tau|}+\log(1+\tfrac{C\epsilon^{2}}{\tau^{2}})\big{)}
(3.33) |B~(ϵ,τ)B~0(τ)|Cμmin{ϵα(|τ|1α+μ|τ|α),ϵ(1+|logϵ|+μlog2ϵ)}\left|\tilde{B}(\epsilon,\tau)-\tilde{B}_{0}(\tau)\right|\leq C\mu\min\{\epsilon^{\alpha}(|\tau|^{1-\alpha}+\mu|\tau|^{\alpha^{\prime}}),\ \epsilon(1+|\log\epsilon|+\mu\log^{2}\epsilon)\}

where B~0(μ,cR,τ)=limϵ0+B~(μ,cR,ϵ,τ)\tilde{B}_{0}(\mu,c_{R},\tau)=\lim_{\epsilon\to 0+}\tilde{B}(\mu,c_{R},\epsilon,\tau). Moreover, general solutions of (3.22) with cI>0c_{I}>0 is given by

(3.34) (μ1y(x2)y(x2))=w(τ)=(10Γ(μ,cR,ϵ,τ)1)B~(μ,cR,ϵ,τ)(bΦ~(μ,c,τ)),b=(b1b2)2.\begin{pmatrix}\mu^{-1}y(x_{2})\\ y^{\prime}(x_{2})\end{pmatrix}=w(\tau)=\begin{pmatrix}1&0\\ \Gamma(\mu,c_{R},\epsilon,\tau)&1\end{pmatrix}\tilde{B}(\mu,c_{R},\epsilon,\tau)\big{(}b-\tilde{\Phi}(\mu,c,\tau)\big{)},\;b=\begin{pmatrix}b_{1}\\ b_{2}\end{pmatrix}\in\mathbb{C}^{2}.
Remark 3.1.

Even though ϵ>0\epsilon>0 is assumed in the above and the remaining statements in this and the next subsections, as C>0C>0 is independent of ϵ=kcI(0,M]\epsilon=\langle k\rangle c_{I}\in(0,M] in a priori estimates and thus they hold even as ϵ0+\epsilon\to 0+.

Expression (3.34) essentially is the variation of parameter formula including the fundamental matrix of the Rayleigh equation. Due to detB~=1\det\tilde{B}=1, it is possible to extend the definition of B~\tilde{B} to include all x2[h0h,h0]x_{2}\in[-h_{0}-h,h_{0}], but its bound would be non-uniform in k1k\gg 1 for |x2x2c|μ|x_{2}-x_{2}^{c}|\gg\mu.

Proof.

Since Γ\Gamma has a logarithmic singularity at the worst (even for ϵ=0\epsilon=0), B~\tilde{B} is obviously well-defined. The zero trace value of the coefficient matrix in (3.29) yields detB=1\det B=1. The form (3.34) of general solutions of (3.22) for cI>0c_{I}>0 follows from straightforward verifications.

Equation (3.29) implies

|(B~I)τ|1+Γ2+(1+Γ2)|B~I|,|(\tilde{B}-I)_{\tau}|\leq 1+\Gamma^{2}+(1+\Gamma^{2})|\tilde{B}-I|,

where 1+Γ21+\Gamma^{2} is the operator norm of the coefficient matrix. From Gronwall inequality, we obtain

|B~I|e|τ+0τΓ2𝑑τ|1.|\tilde{B}-I|\leq e^{|\tau+\int_{0}^{\tau}\Gamma^{2}d\tau^{\prime}|}-1.

It is clear from the definition of γ\gamma that

|γτ|Cμ2,|MτiμϵU2(τ)U~(τ)2+ϵ2𝑑τ|Cμ.|\gamma_{\tau}|\leq C\mu^{2},\quad\left|\int_{-M}^{\tau}\frac{i\mu\epsilon U_{2}(\tau^{\prime})}{\tilde{U}(\tau^{\prime})^{2}+\epsilon^{2}}d\tau^{\prime}\right|\leq C\mu.

The definition of Γ\Gamma, the boundedness of |U~||\tilde{U}|, and the estimate on γτ\gamma_{\tau} imply, for τ[M,M]\tau\in[-M,M],

|0τΓ2𝑑τ|C(|τ|2+μ2)|τ|+Cμ2|0τlog2(|τ|+ϵ)𝑑τ|C(|τ|3+μ2|τ|α),\displaystyle\left|\int_{0}^{\tau}\Gamma^{2}d\tau^{\prime}\right|\leq C(|\tau|^{2}+\mu^{2})|\tau|+C\mu^{2}\left|\int_{0}^{\tau}\log^{2}(|\tau^{\prime}|+\epsilon)d\tau^{\prime}\right|\leq C(|\tau|^{3}+\mu^{2}|\tau|^{\alpha}),

where CC is a generic constant determined by MM and kk_{*} and the Hölder inequality was used to obtain |τ|α|\tau|^{\alpha}, for any α(0,1)\alpha\in(0,1). The desired estimate in (3.31) on B~I\tilde{B}-I follows immediately which along with detB=1\det B=1 in turn yields the estimate on B~1I\tilde{B}^{-1}-I.

The definition of γ\gamma implies

|γ(ϵ,τ)γ(0,τ)||0τCμ2ϵ2(τ)2+ϵ2𝑑τ|=Cμ2ϵtan1|τ|ϵ.|\gamma(\epsilon,\tau)-\gamma(0,\tau)|\leq\left|\int_{0}^{\tau}\frac{C\mu^{2}\epsilon^{2}}{(\tau^{\prime})^{2}+\epsilon^{2}}d\tau^{\prime}\right|=C\mu^{2}\epsilon\tan^{-1}\frac{|\tau|}{\epsilon}.

Regarding the imaginary part of Γ\Gamma, we observe

Mτ|U2(τ)U~(τ)2+ϵ2U2(0)U1(0)2(τ)2+ϵ2|𝑑τ\displaystyle\int_{-M}^{\tau}\left|\frac{U_{2}(\tau^{\prime})}{\tilde{U}(\tau^{\prime})^{2}+\epsilon^{2}}-\frac{U_{2}(0)}{U_{1}(0)^{2}(\tau^{\prime})^{2}+\epsilon^{2}}\right|d\tau^{\prime}
\displaystyle\leq Mτ|U2(τ)(U1(0)2(τ)2+ϵ2)U2(0)(U~(τ)2+ϵ2)|(U~(τ)2+ϵ2)(U1(0)2(τ)2+ϵ2)𝑑τ\displaystyle\int_{-M}^{\tau}\frac{\left|U_{2}(\tau^{\prime})\big{(}U_{1}(0)^{2}(\tau^{\prime})^{2}+\epsilon^{2}\big{)}-U_{2}(0)\big{(}\tilde{U}(\tau^{\prime})^{2}+\epsilon^{2}\big{)}\right|}{\big{(}\tilde{U}(\tau^{\prime})^{2}+\epsilon^{2}\big{)}\big{(}U_{1}(0)^{2}(\tau^{\prime})^{2}+\epsilon^{2}\big{)}}d\tau^{\prime}
\displaystyle\leq CμMτ|τ|(τ)2+ϵ2𝑑τCμ(1+|logϵ|),\displaystyle C\mu\int_{-M}^{\tau}\frac{|\tau^{\prime}|}{(\tau^{\prime})^{2}+\epsilon^{2}}d\tau^{\prime}\leq C\mu(1+|\log\epsilon|),

where we used the smoothness of U1U_{1} and U2U_{2} in μτ\mu\tau. It implies

(3.35) |MτμϵU2(τ)U~(τ)2+ϵ2𝑑τμU2(0)U1(0)(tan1U1(0)τϵ+π2)|Cμϵ(1+μ|logϵ|),\left|\int_{-M}^{\tau}\frac{\mu\epsilon U_{2}(\tau^{\prime})}{\tilde{U}(\tau^{\prime})^{2}+\epsilon^{2}}d\tau^{\prime}-\frac{\mu U_{2}(0)}{U_{1}(0)}\Big{(}\tan^{-1}\frac{U_{1}(0)\tau}{\epsilon}+\frac{\pi}{2}\Big{)}\right|\leq C\mu\epsilon(1+\mu|\log\epsilon|),

and thus

|MτμϵU2(τ)U~(τ)2+ϵ2𝑑τπμU2(0)2U1(0)(sgn(τ)+1)|\displaystyle\left|\int_{-M}^{\tau}\frac{\mu\epsilon U_{2}(\tau^{\prime})}{\tilde{U}(\tau^{\prime})^{2}+\epsilon^{2}}d\tau^{\prime}-\frac{\pi\mu U_{2}(0)}{2U_{1}(0)}\big{(}sgn(\tau)+1\big{)}\right|\leq Cμ(μϵ|logϵ|+min{1,ϵ|τ|1})\displaystyle C\mu\big{(}\mu\epsilon|\log\epsilon|+\min\{1,\,\epsilon|\tau|^{-1}\}\big{)}
\displaystyle\leq Cμ(μϵ|logϵ|+11+|τ|ϵ).\displaystyle C\mu\big{(}\mu\epsilon|\log\epsilon|+\frac{1}{1+\tfrac{|\tau|}{\epsilon}}\big{)}.

The error estimate (3.32) follows consequently.

Proceeding to consider B~(ϵ,τ)B~0(τ)\tilde{B}(\epsilon,\tau)-\tilde{B}_{0}(\tau) where B~0(μ,cR,τ)=B~(μ,cR,0,τ)\tilde{B}_{0}(\mu,c_{R},\tau)=\tilde{B}(\mu,c_{R},0,\tau), we have

τ(B~(ϵ,τ)B~0(τ))(Γ0(τ)1Γ0(τ)2Γ0(τ))(B~(ϵ,τ)B~0(τ))\displaystyle\partial_{\tau}\big{(}\tilde{B}(\epsilon,\tau)-\tilde{B}_{0}(\tau)\big{)}-\begin{pmatrix}\Gamma_{0}(\tau)&1\\ -\Gamma_{0}(\tau)^{2}&-\Gamma_{0}(\tau)\end{pmatrix}\big{(}\tilde{B}(\epsilon,\tau)-\tilde{B}_{0}(\tau)\big{)}
=\displaystyle= (Γ(ϵ,τ)Γ0(τ)0Γ(ϵ,τ)2+Γ0(τ)2Γ(ϵ,τ)+Γ0(τ))B~(ϵ,τ).\displaystyle\begin{pmatrix}\Gamma(\epsilon,\tau)-\Gamma_{0}(\tau)&0\\ -\Gamma(\epsilon,\tau)^{2}+\Gamma_{0}(\tau)^{2}&-\Gamma(\epsilon,\tau)+\Gamma_{0}(\tau)\end{pmatrix}\tilde{B}(\epsilon,\tau).

Recalling that B~0(τ)\tilde{B}_{0}(\tau) is the elementary fundamental matrix of the above corresponding homogeneous ODE system, the variation of parameter formula implies

|B~(ϵ,τ)B~0(τ)|=|0τB~0(τ)B~0(τ)1(Γ(ϵ,τ)Γ0(τ)0Γ(ϵ,τ)2+Γ0(τ)2Γ(ϵ,τ)+Γ0(τ))B~(ϵ,τ)𝑑τ|\displaystyle\left|\tilde{B}(\epsilon,\tau)-\tilde{B}_{0}(\tau)\right|=\left|\int_{0}^{\tau}\tilde{B}_{0}(\tau)\tilde{B}_{0}(\tau^{\prime})^{-1}\begin{pmatrix}\Gamma(\epsilon,\tau^{\prime})-\Gamma_{0}(\tau^{\prime})&0\\ -\Gamma(\epsilon,\tau^{\prime})^{2}+\Gamma_{0}(\tau^{\prime})^{2}&-\Gamma(\epsilon,\tau^{\prime})+\Gamma_{0}(\tau^{\prime})\end{pmatrix}\tilde{B}(\epsilon,\tau^{\prime})d\tau^{\prime}\right|
\displaystyle\leq C|0τ(1+|Γ(ϵ,τ)|+|Γ0(τ)|)|Γ(ϵ,τ)Γ0(τ)|𝑑τ|\displaystyle C\left|\int_{0}^{\tau}\big{(}1+|\Gamma(\epsilon,\tau^{\prime})|+|\Gamma_{0}(\tau^{\prime})|\big{)}|\Gamma(\epsilon,\tau^{\prime})-\Gamma_{0}(\tau^{\prime})|d\tau^{\prime}\right|
\displaystyle\leq C|0τ(1+μ|log|τ||)|Γ(ϵ,τ)Γ0(τ)|dτ|C|1+μ|log()||L11α|Γ(ϵ,)Γ0()|L1α.\displaystyle C\left|\int_{0}^{\tau}\big{(}1+\mu\big{|}\log|\tau^{\prime}|\big{|}\big{)}|\Gamma(\epsilon,\tau^{\prime})-\Gamma_{0}(\tau^{\prime})|d\tau^{\prime}\right|\leq C\big{|}1+\mu|\log(\cdot)|\big{|}_{L^{\frac{1}{1-\alpha}}}|\Gamma(\epsilon,\cdot)-\Gamma_{0}(\cdot)|_{L^{\frac{1}{\alpha}}}.

The second desired upper bound in (3.33) of B~B~0\tilde{B}-\tilde{B}_{0} follows from direct estimating the above integral without using the Hölder inequality. For the first upper bound there we use, for any |τ1|,|τ2|M|\tau_{1}|,|\tau_{2}|\leq M,

(3.36) |Γ(ϵ,)Γ0()|Lρ[τ1,τ2]Cμϵ1ρ,ρ(1,+);|Γ(ϵ,)Γ0()|L1[τ1,τ2]Cμϵ(1+|logϵ|),|\Gamma(\epsilon,\cdot)-\Gamma_{0}(\cdot)|_{L^{\rho}[\tau_{1},\tau_{2}]}\leq C\mu\epsilon^{\frac{1}{\rho}},\;\rho\in(1,+\infty);\quad|\Gamma(\epsilon,\cdot)-\Gamma_{0}(\cdot)|_{L^{1}[\tau_{1},\tau_{2}]}\leq C\mu\epsilon(1+|\log\epsilon|),

which can be verified by straight forward computation. The proof of the lemma is complete. ∎

\bullet A priori estimates. A direct corollary of the form (3.34) of the general solution to the Rayleigh equation (3.22) is an estimate of w(τ)w(\tau) in terms of bb and Φ~\tilde{\Phi}. Let Γ~(τ)\tilde{\Gamma}(\tau) denote

Γ~(τ)=μU′′(x2c)U(x2c)(12log(U~(τ)2+ϵ2)+i(tan1U(x2c)τϵ+π2)).\tilde{\Gamma}(\tau)=\frac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\frac{1}{2}\log(\tilde{U}(\tau)^{2}+\epsilon^{2})+i\big{(}\tan^{-1}\frac{U^{\prime}(x_{2}^{c})\tau}{\epsilon}+\frac{\pi}{2}\big{)}\big{)}.
Corollary 3.4.1.

For b2b\in\mathbb{C}^{2} and |τ|M|\tau|\leq M, let

b~(τ)=(10Γ(τ)1)B~(τ)b,b~0(τ)=(10Γ0(τ)1)B~0(τ)b,\tilde{b}(\tau)=\begin{pmatrix}1&0\\ \Gamma(\tau)&1\end{pmatrix}\tilde{B}(\tau)b,\quad\tilde{b}_{0}(\tau)=\begin{pmatrix}1&0\\ \Gamma_{0}(\tau)&1\end{pmatrix}\tilde{B}_{0}(\tau)b,

then under the same assumptions of Lemma 3.4, it holds, for any α1[0,1α)\alpha_{1}\in[0,1-\alpha),

|b~1(τ)b1|C(|τ|+μ2|τ|α)|b|,|b~2(τ)(b2+b1Γ~(τ))|C(|τ|+μ(|τ|α+ϵα))|b|\displaystyle|\tilde{b}_{1}(\tau)-b_{1}|\leq C(|\tau|+\mu^{2}|\tau|^{\alpha})|b|,\quad\big{|}\tilde{b}_{2}(\tau)-\big{(}b_{2}+b_{1}\tilde{\Gamma}(\tau)\big{)}\big{|}\leq C\big{(}|\tau|+\mu(|\tau|^{\alpha}+\epsilon^{\alpha})\big{)}|b|
|b~1(τ)b~01(τ)|Cμϵα(|τ||b|+min{|τ|1α,ϵ1α(1+|logϵ|)}|b1|),|\tilde{b}_{1}(\tau)-\tilde{b}_{01}(\tau)|\leq C\mu\epsilon^{\alpha}\big{(}|\tau||b|+\min\{|\tau|^{1-\alpha},\,\epsilon^{1-\alpha}(1+|\log\epsilon|)\}|b_{1}|\big{)},
|b~2(τ)b~02(τ)|Cμ(ϵα|τ|α1|b|+(ϵϵ+|τ|+log(1+Cϵ2τ2))|b1|).|\tilde{b}_{2}(\tau)-\tilde{b}_{02}(\tau)|\leq C\mu\big{(}\epsilon^{\alpha}|\tau|^{\alpha_{1}}|b|+\big{(}\tfrac{\epsilon}{\epsilon+|\tau|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau^{2}}\big{)}\big{)}|b_{1}|\big{)}.
Proof.

The estimates on b~\tilde{b} follows from straight forward calculation based on (3.35) and the bound on B~I\tilde{B}-I given in Lemma 3.4 and we omit the details.

Regrading b~(τ)b~0(τ)\tilde{b}(\tau)-\tilde{b}_{0}(\tau), let B~jl\tilde{B}_{jl} denote the entries of B~\tilde{B}. Using Lemma 3.4 where the estimates are uniform in ϵ>0\epsilon>0, we have

|b~2(τ)b~02(τ)|(1+|Γ0|)|B~B~0||b|+|ΓΓ0|(|B~11||b1|+|B~12||b2|)\displaystyle|\tilde{b}_{2}(\tau)-\tilde{b}_{02}(\tau)|\leq(1+|\Gamma_{0}|)|\tilde{B}-\tilde{B}_{0}||b|+|\Gamma-\Gamma_{0}|(|\tilde{B}_{11}||b_{1}|+|\tilde{B}_{12}||b_{2}|)
\displaystyle\leq C((1+μ|log|τ||)|B~B~0||b|+|ΓΓ0|(|b1|+(|τ|+μ|τ|α)|b2|))\displaystyle C\big{(}(1+\mu\big{|}\log|\tau|\big{|})|\tilde{B}-\tilde{B}_{0}||b|+|\Gamma-\Gamma_{0}|(|b_{1}|+(|\tau|+\mu|\tau|^{\alpha^{\prime}})|b_{2}|)\big{)}
\displaystyle\leq Cμ(ϵα|τ|α1|b|+(ϵϵ+|τ|+log(1+Cϵ2τ2))(|b1|+|τ|α|b2|)).\displaystyle C\mu\big{(}\epsilon^{\alpha}|\tau|^{\alpha_{1}}|b|+\big{(}\tfrac{\epsilon}{\epsilon+|\tau|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau^{2}}\big{)}\big{)}(|b_{1}|+|\tau|^{\alpha^{\prime}}|b_{2}|)\big{)}.

Since

(3.37) |τ|β(ϵϵ+|τ|+log(1+Cϵ2τ2))Cϵβ,β(0,1],|\tau|^{\beta}\big{(}\tfrac{\epsilon}{\epsilon+|\tau|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau^{2}}\big{)}\big{)}\leq C\epsilon^{\beta},\quad\beta\in(0,1],

the upper on b~2(τ)b~02(τ)\tilde{b}_{2}(\tau)-\tilde{b}_{02}(\tau) follows accordingly.

To derive the estimate on b~1(τ)b~01(τ)\tilde{b}_{1}(\tau)-\tilde{b}_{01}(\tau), we notice b~1(0)=b~01(0)=b1\tilde{b}_{1}(0)=\tilde{b}_{01}(0)=b_{1} and the desired estimate follows from integrating τ(b~1b~01)=b~2b~02\partial_{\tau}(\tilde{b}_{1}-\tilde{b}_{01})=\tilde{b}_{2}-\tilde{b}_{02} using (3.36). ∎

Remark 3.2.

The above estimates imply, that for any solution w(τ)w(\tau) to (3.22)

(3.38) |w1(τ)(b1Φ~1(τ))|C(|τ|+μ2|τ|α)|bΦ~(τ)|,\displaystyle\big{|}w_{1}(\tau)-\big{(}b_{1}-\tilde{\Phi}_{1}(\tau)\big{)}\big{|}\leq C(|\tau|+\mu^{2}|\tau|^{\alpha})\big{|}b-\tilde{\Phi}(\tau)\big{|},
(3.39) |w2(τ)(b2Φ~2(τ)+Γ~(τ)(b1Φ~1(τ)))|C(|τ|+μ(|τ|α+ϵα))|bΦ~(τ)|.\displaystyle\Big{|}w_{2}(\tau)-\Big{(}b_{2}-\tilde{\Phi}_{2}(\tau)+\tilde{\Gamma}(\tau)\big{(}b_{1}-\tilde{\Phi}_{1}(\tau)\big{)}\Big{)}\Big{|}\leq C\big{(}|\tau|+\mu(|\tau|^{\alpha}+\epsilon^{\alpha})\big{)}\big{|}b-\tilde{\Phi}(\tau)\big{|}.

The following lemma gives another estimate of w(τ)w(\tau) in terms of some initial value w(τ0)w(\tau_{0}) which we shall use mainly for τ0\tau_{0} away from 0.

Lemma 3.5.

For any M>0M>0 satisfying (3.23) and α(0,1)\alpha\in(0,1), there exists C>0C>0 depending only on MM, α\alpha, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any 0<ϵ<M0<\epsilon<M, and τ0,τ[M,M]\tau_{0},\tau\in[-M,M], the following hold for any solution w(τ)w(\tau) to (3.22):

(3.40a) |w1(τ)w1(τ0)|C|ττ0|(|w(τ0)|+μ|log(τ02+ϵ2)||w1(τ0)|)+Cμ|ττ0|α(|w(τ0)|+|Φ~1()Φ~1(τ0)|L[τ0,τ])+C|Φ~()Φ~(τ0)|L1[τ0,τ],\begin{split}\big{|}w_{1}(\tau)-w_{1}&(\tau_{0})\big{|}\leq C|\tau-\tau_{0}|\big{(}|w(\tau_{0})|+\mu\big{|}\log\big{(}\tau_{0}^{2}+\epsilon^{2}\big{)}\big{|}|w_{1}(\tau_{0})|\big{)}\\ &+C\mu|\tau-\tau_{0}|^{\alpha}\big{(}|w(\tau_{0})|+\big{|}\tilde{\Phi}_{1}(\cdot)-\tilde{\Phi}_{1}(\tau_{0})\big{|}_{L^{\infty}[\tau_{0},\tau]}\big{)}+C\big{|}\tilde{\Phi}(\cdot)-\tilde{\Phi}(\tau_{0})\big{|}_{L^{1}[\tau_{0},\tau]},\end{split}
(3.40b) |w2(τ)(w2(τ0)+Φ~2(τ0)Φ~2(τ)Γ~(τ0)w1(τ0)+Γ~(τ)(w1(τ0)+Φ~1(τ0)Φ~1(τ)))|C((|τ|α+μϵα)|Φ~(τ0)Φ~(τ)|+(μϵα+|τ|α+|τ0|α(1+μ|log(τ2+ϵ2)|))|w(τ0)|+μ|τ|α|log(τ02+ϵ2)||w1(τ0)|).\begin{split}&\big{|}w_{2}(\tau)-\big{(}w_{2}(\tau_{0})+\tilde{\Phi}_{2}(\tau_{0})-\tilde{\Phi}_{2}(\tau)-\tilde{\Gamma}(\tau_{0})w_{1}(\tau_{0})+\tilde{\Gamma}(\tau)(w_{1}(\tau_{0})+\tilde{\Phi}_{1}(\tau_{0})-\tilde{\Phi}_{1}(\tau))\big{)}\big{|}\\ \leq&C\Big{(}(|\tau|^{\alpha}+\mu\epsilon^{\alpha})\big{|}\tilde{\Phi}(\tau_{0})-\tilde{\Phi}(\tau)\big{|}+\big{(}\mu{\epsilon^{\alpha}}+|\tau|^{\alpha}+|\tau_{0}|^{\alpha}(1+\mu|\log(\tau^{2}+\epsilon^{2})|)\big{)}|w(\tau_{0})|\\ &+\mu|\tau|^{\alpha}|\log(\tau_{0}^{2}+\epsilon^{2})||w_{1}(\tau_{0})|\Big{)}.\end{split}
Proof.

We shall first estimate bΦ~(τ0)b-\tilde{\Phi}(\tau_{0}) based on w(τ0)w(\tau_{0}) and then apply Corollary 3.4.1. From (3.34) and detB~=1\det\tilde{B}=1 which allows us to write B~1\tilde{B}^{-1} explicitly, we have

(3.41) bΦ~(τ0)=(B~22B~12B~21B~11)(10Γ1)w|τ0=(B~22+ΓB~12B~12B~21ΓB~11B~11)w|τ0.b-\tilde{\Phi}(\tau_{0})=\begin{pmatrix}\tilde{B}_{22}&-\tilde{B}_{12}\\ -\tilde{B}_{21}&\tilde{B}_{11}\end{pmatrix}\begin{pmatrix}1&0\\ -\Gamma&1\end{pmatrix}w\Big{|}_{\tau_{0}}=\begin{pmatrix}\tilde{B}_{22}+\Gamma\tilde{B}_{12}&-\tilde{B}_{12}\\ -\tilde{B}_{21}-\Gamma\tilde{B}_{11}&\tilde{B}_{11}\end{pmatrix}w\Big{|}_{\tau_{0}}.

Using Lemma 3.4, one may estimate

(3.42) |b1Φ~1(τ0)w1(τ0)|C(|τ0|+μ|τ0|α)|w(τ0)|,|b_{1}-\tilde{\Phi}_{1}(\tau_{0})-w_{1}(\tau_{0})|\leq C(|\tau_{0}|+\mu|\tau_{0}|^{\alpha})|w(\tau_{0})|,
(3.43) |b2Φ~2(τ0)+Γ~(τ0)w1(τ0)w2(τ0)|C(|τ0|+μ(|τ0|α+ϵα))|w(τ0)|,|b_{2}-\tilde{\Phi}_{2}(\tau_{0})+\tilde{\Gamma}(\tau_{0})w_{1}(\tau_{0})-w_{2}(\tau_{0})|\leq C\big{(}|\tau_{0}|+\mu(|\tau_{0}|^{\alpha}+\epsilon^{\alpha})\big{)}|w(\tau_{0})|,

where we also used (3.35). Combining these inequalities and Corollary 3.4.1, we obtain

|w2(τ)(w2(τ0)+Φ~2(τ0)Φ~2(τ)Γ~(τ0)w1(τ0)+Γ~(τ)(w1(τ0)+Φ~1(τ0)Φ~1(τ)))|\displaystyle\big{|}w_{2}(\tau)-\big{(}w_{2}(\tau_{0})+\tilde{\Phi}_{2}(\tau_{0})-\tilde{\Phi}_{2}(\tau)-\tilde{\Gamma}(\tau_{0})w_{1}(\tau_{0})+\tilde{\Gamma}(\tau)(w_{1}(\tau_{0})+\tilde{\Phi}_{1}(\tau_{0})-\tilde{\Phi}_{1}(\tau))\big{)}\big{|}
\displaystyle\leq C(|τ|+μ(|τ|α+ϵα))|bΦ~(τ)|+C(|τ0|+μ(|τ0|α+ϵα)+|Γ~(τ)|(|τ0|+μ|τ0|α))|w(τ0)|\displaystyle C\big{(}|\tau|+\mu(|\tau|^{\alpha^{\prime}}+\epsilon^{\alpha^{\prime}})\big{)}\big{|}b-\tilde{\Phi}(\tau)\big{|}+C\big{(}|\tau_{0}|+\mu(|\tau_{0}|^{\alpha}+\epsilon^{\alpha})+|\tilde{\Gamma}(\tau)|(|\tau_{0}|+\mu|\tau_{0}|^{\alpha})\big{)}|w(\tau_{0})|
\displaystyle\leq C(|τ|α+μϵα)(|Φ~(τ0)Φ~(τ)|+|w(τ0)|+|Γ~(τ0)||w1(τ0)|)\displaystyle C(|\tau|^{\alpha^{\prime}}+\mu\epsilon^{\alpha^{\prime}})\big{(}\big{|}\tilde{\Phi}(\tau_{0})-\tilde{\Phi}(\tau)\big{|}+|w(\tau_{0})|+|\tilde{\Gamma}(\tau_{0})||w_{1}(\tau_{0})|\big{)}
+C(μϵα+|τ0|α(1+μ|log(τ2+ϵ2)|))|w(τ0)|\displaystyle+C\big{(}\mu\epsilon^{\alpha}+|\tau_{0}|^{\alpha}(1+\mu|\log(\tau^{2}+\epsilon^{2})|)\big{)}|w(\tau_{0})|
\displaystyle\leq C((|τ|α+μϵα)|Φ~(τ0)Φ~(τ)|+(μϵα+|τ|α+|τ0|α(1+μ|log(τ2+ϵ2)|))|w(τ0)|\displaystyle C\Big{(}(|\tau|^{\alpha}+\mu\epsilon^{\alpha})\big{|}\tilde{\Phi}(\tau_{0})-\tilde{\Phi}(\tau)\big{|}+\big{(}\mu\epsilon^{\alpha}+|\tau|^{\alpha}+|\tau_{0}|^{\alpha}(1+\mu|\log(\tau^{2}+\epsilon^{2})|)\big{)}|w(\tau_{0})|
+μ|τ|α|log(τ02+ϵ2)||w1(τ0)|),\displaystyle+\mu|\tau|^{\alpha}|\log(\tau_{0}^{2}+\epsilon^{2})||w_{1}(\tau_{0})|\Big{)},

where α(α,1)\alpha^{\prime}\in(\alpha,1). This yields inequality (3.40b) of w2(τ)w_{2}(\tau). The estimate of w1(τ)w_{1}(\tau) is obtained through integrating that of w2(τ)=w1τ(τ)w_{2}(\tau)=w_{1\tau}(\tau). ∎

\bullet Convergence estimates as cI0+c_{I}\to 0+. As ϵ=μ1cI=kcI0+\epsilon=\mu^{-1}c_{I}=\langle k\rangle c_{I}\to 0+, from Lemma 3.4, it is natural to expect that the limit of solutions to the non-homogenous Rayleigh equation (3.3) is also given by formula (3.34) with Γ\Gamma, B~\tilde{B}, and Φ~\tilde{\Phi} replaced by Γ0\Gamma_{0}, B~0\tilde{B}_{0}, and Φ~0=limϵ0+Φ~\tilde{\Phi}_{0}=\lim_{\epsilon\to 0+}\tilde{\Phi}.

With the above preparations, we are ready to obtain the convergence and error estimates of solutions to the Rayleigh equation (3.24). While the limits of non-homogeneous Rayleigh equation under appropriate assumptions on ϕ(k,c,x2)\phi(k,c,x_{2}) can be studied in the framework in this section, we shall just focus on the homogeneous case, i.e. with ϕ0\phi\equiv 0, and leave the non-homogeneous one to Section 5. In fact, (3.28) and Lemma 3.4 imply that, as cI0c_{I}\to 0, w1(τ)w_{1}(\tau) would converge to a Hölder continuous limit, while w2(τ)w_{2}(\tau) develops a jump proportional to w1(0)w_{1}(0) and a logarithmic singularity at τ=0\tau=0. More precisely, the limit W(τ)W(\tau) of solutions should (see the proposition in the below) satisfy the Rayleigh equation (3.24) with cc\in\mathbb{R} for τ0\tau\neq 0 and satisfy at τ=0\tau=0,

(3.44) {W1C0([M,M]),W2C0([M,M]\{0}),limτ0+(W2(τ)W2(τ))=iπμU′′(x2c)U(x2c)W1(0).\begin{cases}&W_{1}\in C^{0}\big{(}[-M,M]\big{)},\quad W_{2}\in C^{0}\big{(}[-M,M]\backslash\{0\}\big{)},\\ &\lim_{\tau\to 0+}\big{(}W_{2}(\tau)-W_{2}(-\tau)\big{)}=\frac{i\pi\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}W_{1}(0).\end{cases}

It is worth pointing out that the existence of the limit of W2(τ)W2(τ)W_{2}(\tau)-W_{2}(-\tau) does not imply a simple jump discontinuity of W2W_{2}, which actually has a symmetric logarithmic singularity. In the distribution sense, the limit homogeneous Rayleigh equation (3.24) (with ϕ=0\phi=0) along with (3.44) can be written as

(3.45) Wτ=(P.V.)τ(011μ2+μ2U′′(x2c+μτ)U(x2c+μτ)c0)W+(0iπμU′′(x2c)U(x2c)W1(0))δ(τ).\begin{split}W_{\tau}=&(P.V.)_{\tau}\begin{pmatrix}0&1\\ 1-\mu^{2}+\frac{\mu^{2}U^{\prime\prime}(x_{2}^{c}+\mu\tau)}{U(x_{2}^{c}+\mu\tau)-c}&0\end{pmatrix}W+\begin{pmatrix}0\\ \frac{i\pi\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}W_{1}(0)\end{pmatrix}\delta(\tau).\end{split}

Here δ(τ)\delta(\tau) denotes the delta function of τ\tau and “(P.V.)τ(P.V.)_{\tau}” indicate the principle value when the corresponding distributions are applied to test functions of τ\tau. They occur in W2τW_{2\tau} only. In terms of the original unknown y(x2)y(x_{2}), the limit of (3.3) as cI0+c_{I}\to 0+ is

(3.46) y′′+k2y+(P.V.)x2(U′′yUc)=iπU′′(x2c)U(x2c)y(x2c)δx2(x2x2c),-y^{\prime\prime}+k^{2}y+(P.V.)_{x_{2}}\big{(}\tfrac{U^{\prime\prime}y}{U-c}\big{)}=-\tfrac{i\pi U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y(x_{2}^{c})\delta_{x_{2}}(x_{2}-x_{2}^{c}),

where the subscript x2\cdot_{x_{2}} indicates the distributions as generalized functions of x2x_{2}. For cI0c_{I}\to 0-, the parallel results hold except with the complex conjugate. It also means that homogeneous Rayleigh equation takes different limit as cI±0c_{I}\to\pm 0.

Lemma 3.6.

General solutions of homogeneous (3.24) (with ϕ=0\phi=0) along with (3.44) are

(3.47) W(τ)=(10Γ0(μ,cR,τ)1)B~0(μ,cR,τ)b0,b0=(b01b02)2.W(\tau)=\begin{pmatrix}1&0\\ \Gamma_{0}(\mu,c_{R},\tau)&1\end{pmatrix}\tilde{B}_{0}(\mu,c_{R},\tau)b_{0},\quad b_{0}=\begin{pmatrix}b_{01}\\ b_{02}\end{pmatrix}\in\mathbb{C}^{2}.

Moreover, W(τ)C0W(\tau)\in C^{0} if W1(0)=0W_{1}(0)=0.

Proof.

On [M,0)[-M,0) and (0,M](0,M], (3.24) is regular and thus Lemma 3.4, in particular the form (3.34) of the general solutions implies the above (3.47) with parameters b0±=(b01±,b02±)T2b_{0}^{\pm}=(b_{01}^{\pm},b_{02}^{\pm})^{T}\in\mathbb{C}^{2}. The continuity of W1(τ)W_{1}(\tau) and the estimates of Γ\Gamma and B~\tilde{B} in Lemma 3.4 immediately yields b01+=b01b_{01}^{+}=b_{01}^{-}. Finally b02+=b02b_{02}^{+}=b_{02}^{-} follows from the jump condition of W2(τ)W_{2}(\tau) at τ=0\tau=0 after writing b02±b_{02}^{\pm} using (3.41) and again using the estimates of Γ\Gamma and B~\tilde{B}.

Finally, the continuity of W(τ)W(\tau) under the assumptions W1(0)=0W_{1}(0)=0 follows from (3.47), the Hölder continuity of B~\tilde{B}, and the logarithmic upper bound of Γ0\Gamma_{0}. ∎

The following proposition provides the convergence estimates.

Proposition 3.7.

For any M>0M>0 satisfying (3.23) and α,α(0,1)\alpha,\alpha^{\prime}\in(0,1), there exists C>0C>0 depending only on MM, α\alpha, α\alpha^{\prime}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any 0<ϵ<M0<\epsilon<M, τ[M,M]\tau\in[-M,M], and solutions w(τ)w(\tau) and W(τ)W(\tau) to (3.22) and (3.24) (with ϕ=0\phi=0) in the forms (3.34) and (3.47) with parameters b,b02b,b_{0}\in\mathbb{C}^{2}, respectively, the following hold:

(3.48) |w1(τ)W1(τ)(w1(0)W1(0))|C(|τ|(|b2b02|+μϵα|b02|)+(|τ|+μ|τ|α)|w1(0)W1(0)|+ϵαμ|τ|1α|W1(0)|),\begin{split}|w_{1}(\tau)-W_{1}(\tau)-(w_{1}(0)-&W_{1}(0))|\leq C\big{(}|\tau|(|b_{2}-b_{02}|+\mu\epsilon^{\alpha}|b_{02}|)\\ &+(|\tau|+\mu|\tau|^{\alpha})|w_{1}(0)-W_{1}(0)|+\epsilon^{\alpha^{\prime}}\mu|\tau|^{1-\alpha^{\prime}}|W_{1}(0)|\big{)},\end{split}
(3.49) |w2(τ)W2(τ)|C(μϵα|τ|1α2(|W1(0)|+|b02|)+(1+μ|log(|τ|+ϵ)|)|w1(0)W1(0)|+|b2b02|+μ(ϵϵ+|τ|+log(1+Cϵ2τ2))|W1(0)|).\begin{split}|w_{2}(\tau)-W_{2}(\tau)|\leq&C\Big{(}\mu\epsilon^{\alpha}|\tau|^{\frac{1-\alpha}{2}}(|W_{1}(0)|+|b_{02}|)+(1+\mu\big{|}\log(|\tau|+\epsilon)\big{|})|w_{1}(0)-W_{1}(0)|\\ &+|b_{2}-b_{02}|+\mu\big{(}\tfrac{\epsilon}{\epsilon+|\tau|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau^{2}}\big{)}\big{)}|W_{1}(0)|\Big{)}.\end{split}

Moreover, for any τ,τ0[M,M]\tau,\tau_{0}\in[-M,M], let τ=min{|τ|,|τ0|}>0\tau_{*}=\min\{|\tau|,|\tau_{0}|\}>0, we have

(3.50) |w1(τ)W1(τ)(w1(τ0)W1(τ0))|Cμϵα|ττ0||W(τ0)|+C|ττ0|α|w(τ0)W(τ0)|+Cμ(|ττ0|(ϵϵ+|τ0|+log(1+Cϵ2τ02))+ϵα|ττ0|1α)×(|W1(τ0)|+|τ0|α|W(τ0)|),\begin{split}|w_{1}(\tau)-W_{1}(\tau)-(w_{1}(\tau_{0})-&W_{1}(\tau_{0}))|\leq C\mu\epsilon^{\alpha}|\tau-\tau_{0}||W(\tau_{0})|+C|\tau-\tau_{0}|^{\alpha}|w(\tau_{0})-W(\tau_{0})|\\ &+C\mu\Big{(}|\tau-\tau_{0}|\big{(}\tfrac{\epsilon}{\epsilon+|\tau_{0}|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau_{0}^{2}}\big{)}\big{)}+\epsilon^{\alpha^{\prime}}|\tau-\tau_{0}|^{1-\alpha^{\prime}}\Big{)}\\ &\times\big{(}|W_{1}(\tau_{0})|+|\tau_{0}|^{\alpha}|W(\tau_{0})|\big{)},\end{split}
(3.51) |w2(τ)W2(τ)|C((1+μ|log(ϵ+|τ|)|)|w(τ0)W(τ0)|+μϵα|W(τ0)|+μ(ϵϵ+τ+log(1+Cϵ2τ2))(|W1(τ0)|+|τ0|α|W(τ0)|)).\begin{split}|w_{2}(\tau)-W_{2}(\tau)|\leq&C\Big{(}\big{(}1+\mu|\log(\epsilon+|\tau|)\big{|}\big{)}|w(\tau_{0})-W(\tau_{0})|\\ &+\mu\epsilon^{\alpha}|W(\tau_{0})|+\mu\big{(}\tfrac{\epsilon}{\epsilon+\tau_{*}}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau_{*}^{2}}\big{)}\big{)}\big{(}|W_{1}(\tau_{0})|+|\tau_{0}|^{\alpha}|W(\tau_{0})|\big{)}\Big{)}.\end{split}
Remark 3.3.

When the above convergence estimate is applied in the rest of the manuscript, it always holds that |W1(τ0)|M|τ0|α0|W_{1}(\tau_{0})|\leq M|\tau_{0}|^{\alpha_{0}} for some α0>0\alpha_{0}>0 which makes the right sides of (3.50) and (3.51) converging to 0 as ϵ0\epsilon\to 0 locally uniformly in τ0\tau\neq 0.

Proof.

We first work on the error estimates in terms of W1(0)W_{1}(0) and b2b_{2}. Let

w~(τ)=(10Γ(μ,cR,ϵ,τ)1)B~(μ,cR,ϵ,τ)b0.\tilde{w}(\tau)=\begin{pmatrix}1&0\\ \Gamma(\mu,c_{R},\epsilon,\tau)&1\end{pmatrix}\tilde{B}(\mu,c_{R},\epsilon,\tau)b_{0}.

Controlling w2w~2w_{2}-\tilde{w}_{2} and w~2W2\tilde{w}_{2}-W_{2} by Corollary 3.4.1 (w2w~2w_{2}-\tilde{w}_{2} by (3.38) and (3.39) in particular), where we recall the estimates are uniform in ϵ>0\epsilon>0, we have

(3.52) |w2(τ)W2(τ)||w2(τ)w~2(τ)|+|w~2(τ)W2(τ)|C(|bb0|+|Γ~(τ)||b1b01|+μϵα|τ|α1|b0|+μ(ϵ|τ|+ϵ+log(1+Cϵ2τ2))|b01|)C(|b2b02|+(1+μ|log(|τ|+ϵ)|)|w1(0)W1(0)|+μ(ϵϵ+|τ|+log(1+Cϵ2τ2))|W1(0)|+μϵα|τ|α1(|W1(0)|+|b02|)).\begin{split}|w_{2}(\tau)-W_{2}&(\tau)|\leq|w_{2}(\tau)-\tilde{w}_{2}(\tau)|+|\tilde{w}_{2}(\tau)-W_{2}(\tau)|\\ \leq&C\Big{(}|b-b_{0}|+|\tilde{\Gamma}(\tau)||b_{1}-b_{01}|+\mu\epsilon^{\alpha}|\tau|^{\alpha_{1}}|b_{0}|+\mu\big{(}\tfrac{\epsilon}{|\tau|+\epsilon}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau^{2}}\big{)}\big{)}|b_{01}|\Big{)}\\ \leq&C\Big{(}|b_{2}-b_{02}|+(1+\mu\big{|}\log(|\tau|+\epsilon)\big{|})|w_{1}(0)-W_{1}(0)|\\ &+\mu\big{(}\tfrac{\epsilon}{\epsilon+|\tau|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau^{2}}\big{)}\big{)}|W_{1}(0)|+\mu\epsilon^{\alpha}|\tau|^{\alpha_{1}}\big{(}|W_{1}(0)|+|b_{02}|\big{)}\Big{)}.\end{split}

where α1[0,1α)\alpha_{1}\in[0,1-\alpha) and we also used

W1(0)=b01,w1(0)=b1.W_{1}(0)=b_{01},\quad w_{1}(0)=b_{1}.

This completes the proof of inequality (3.49). The estimate (3.48) on w1W1w_{1}-W_{1} is derived by integrating τ(w1W1)=w2W2\partial_{\tau}(w_{1}-W_{1})=w_{2}-W_{2} and using (3.36) and (3.32).

In the following, based on (3.52) we establish the error estimates in terms of initial values given at some τ00\tau_{0}\neq 0. From formula (3.41) we have

bb0(Γ0(τ0)W1(τ0)Γ(τ0)w1(τ0))(0, 1)T\displaystyle b-b_{0}-\big{(}\Gamma_{0}(\tau_{0})W_{1}(\tau_{0})-\Gamma(\tau_{0})w_{1}(\tau_{0})\big{)}(0,\,1)^{T}
=\displaystyle= (B~1(wW)+(B~1B~01)W+(Γ0(W1w1)(B~01I)+Γ0w1(B~01B~1)\displaystyle\Big{(}\tilde{B}^{-1}(w-W)+(\tilde{B}^{-1}-\tilde{B}_{0}^{-1})W+\Big{(}\Gamma_{0}(W_{1}-w_{1})(\tilde{B}_{0}^{-1}-I)+\Gamma_{0}w_{1}(\tilde{B}_{0}^{-1}-\tilde{B}^{-1})
+(Γ0Γ)w1(B~1I))(0, 1)T)|τ0.\displaystyle+(\Gamma_{0}-\Gamma)w_{1}(\tilde{B}^{-1}-I)\Big{)}(0,\,1)^{T}\Big{)}\Big{|}_{\tau_{0}}.

From (3.37) and Lemma 3.4, one may estimate

|Γ0(τ0)||B~(τ0)1I|C(1+μ|log|τ0||)(|τ0|+μ2|τ0|α)C(|τ0|+μ|τ0|α),|\Gamma_{0}(\tau_{0})||\tilde{B}(\tau_{0})^{-1}-I|\leq C\big{(}1+\mu\big{|}\log|\tau_{0}|\big{|}\big{)}(|\tau_{0}|+\mu^{2}|\tau_{0}|^{\alpha^{\prime}})\leq C(|\tau_{0}|+\mu|\tau_{0}|^{\alpha}),
|Γ0||B~01B~1||τ0C(1+μ|log|τ0||)μϵα(|τ0|1α+μ|τ0|α1)Cμϵα|τ0|α1,|\Gamma_{0}||\tilde{B}_{0}^{-1}-\tilde{B}^{-1}|\big{|}_{\tau_{0}}\leq C\big{(}1+\mu\big{|}\log|\tau_{0}|\big{|}\big{)}\mu\epsilon^{\alpha}(|\tau_{0}|^{1-\alpha}+\mu|\tau_{0}|^{\alpha_{1}})\leq C\mu\epsilon^{\alpha}|\tau_{0}|^{\alpha_{1}},
|Γ0Γ||B~1I||τ0Cμ(|τ0|+μ2|τ0|α)(μϵ|logϵ|+ϵϵ+|τ0|+log(1+Cϵ2τ02))Cμϵα|τ0|α1,|\Gamma_{0}-\Gamma||\tilde{B}^{-1}-I|\big{|}_{\tau_{0}}\leq C\mu(|\tau_{0}|+\mu^{2}|\tau_{0}|^{\alpha^{\prime}})\big{(}\mu\epsilon|\log\epsilon|+\tfrac{\epsilon}{\epsilon+|\tau_{0}|}+\log(1+\tfrac{C\epsilon^{2}}{\tau_{0}^{2}})\big{)}\leq C\mu\epsilon^{\alpha}|\tau_{0}|^{\alpha_{1}},

where α1[0,1α)\alpha_{1}\in[0,1-\alpha). Therefore we obtain

|b2\displaystyle\big{|}b_{2}- b02(Γ0(τ0)W1(τ0)Γ(τ0)w1(τ0))|+|b1b01|C(|(wW)(τ0)|+μϵα|τ0|α1|W(τ0)|).\displaystyle b_{02}-\big{(}\Gamma_{0}(\tau_{0})W_{1}(\tau_{0})-\Gamma(\tau_{0})w_{1}(\tau_{0})\big{)}\big{|}+\big{|}b_{1}-b_{01}\big{|}\leq C\big{(}|(w-W)(\tau_{0})|+\mu\epsilon^{\alpha}|\tau_{0}|^{\alpha_{1}}|W(\tau_{0})|\big{)}.

Applying (3.42) and (3.43) to control b0b_{0} in (3.52), we can estimate

|w2(τ)W2(τ)|\displaystyle|w_{2}(\tau)-W_{2}(\tau)|\leq C((1+μ|log(ϵ+|τ|)|)(|w(τ0)W(τ0)|+μϵα|τ0|α1|W(τ0)|)\displaystyle C\Big{(}\big{(}1+\mu\big{|}\log(\epsilon+|\tau|)\big{|}\big{)}\big{(}|w(\tau_{0})-W(\tau_{0})|+\mu\epsilon^{\alpha}|\tau_{0}|^{\alpha_{1}}|W(\tau_{0})|\big{)}
+|(Γ0W1Γw1)(τ0)|+μϵα(|W(τ0)|+μ|log|τ0|||W1(τ0)|)\displaystyle+|(\Gamma_{0}W_{1}-\Gamma w_{1})(\tau_{0})|+\mu\epsilon^{\alpha}\big{(}|W(\tau_{0})|+\mu\big{|}\log|\tau_{0}|\big{|}|W_{1}(\tau_{0})|\big{)}
+μ(ϵϵ+|τ|+log(1+Cϵ2τ2))(|W1(τ0)|+|τ0|α|W(τ0)|))\displaystyle+\mu\big{(}\tfrac{\epsilon}{\epsilon+|\tau|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau^{2}}\big{)}\big{)}\big{(}|W_{1}(\tau_{0})|+|\tau_{0}|^{\alpha}|W(\tau_{0})|\big{)}\Big{)}

Inequality (3.51) is obtained by simplifying the above. In particular, we used

ϵα|logτ|Cϵα if τmin{1,ϵ2} and log(1+Cϵ2τ2)log(ϵ2τ1τ)|logτ| if τmin{1,ϵ2},\epsilon^{\alpha}|\log\tau_{*}|\leq C\epsilon^{\alpha^{\prime}}\,\text{ if }\,\tau_{*}\geq\min\{1,\epsilon^{2}\}\,\text{ and }\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau_{*}^{2}}\big{)}\geq\log\big{(}\tfrac{\epsilon^{2}}{\tau_{*}}\tfrac{1}{\tau_{*}}\big{)}\geq|\log\tau_{*}|\,\text{ if }\,\tau_{*}\leq\min\{1,\epsilon^{2}\},

to absorb the term μ2ϵα|log|τ0|||W1(τ0)|\mu^{2}\epsilon^{\alpha}\big{|}\log|\tau_{0}|\big{|}|W_{1}(\tau_{0})|.

Again we integrate (3.51) to derive (3.50). The only non-trivial terms are those involving τ\tau_{*}

|τ0τϵϵ+min{|τ|,|τ0|}+log(1+Cϵ2min{|τ|,|τ0|}2)dτ|\displaystyle\left|\int_{\tau_{0}}^{\tau}\tfrac{\epsilon}{\epsilon+\min\{|\tau^{\prime}|,|\tau_{0}|\}}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\min\{|\tau^{\prime}|,|\tau_{0}|\}^{2}}\big{)}d\tau^{\prime}\right|\leq Cϵα|ττ0|1α+||τ||τ0||(ϵϵ+|τ0|+log(1+Cϵ2τ02))\displaystyle C\epsilon^{\alpha}|\tau-\tau_{0}|^{1-\alpha}+\big{|}|\tau|-|\tau_{0}|\big{|}\big{(}\tfrac{\epsilon}{\epsilon+|\tau_{0}|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau_{0}^{2}}\big{)}\big{)}
\displaystyle\leq Cϵα|ττ0|1α+|ττ0|(ϵϵ+|τ0|+log(1+Cϵ2τ02)),\displaystyle C\epsilon^{\alpha}|\tau-\tau_{0}|^{1-\alpha}+|\tau-\tau_{0}|\big{(}\tfrac{\epsilon}{\epsilon+|\tau_{0}|}+\log\big{(}1+\tfrac{C\epsilon^{2}}{\tau_{0}^{2}}\big{)}\big{)},

which are obtained by considering whether |τ||τ0||\tau^{\prime}|\geq|\tau_{0}| and using (3.36). ∎

3.3. A priori bounds on the two fundamental solutions y±(k,c,x2)y_{\pm}(k,c,x_{2}) to the homogeneous Rayleigh equation with cI0c_{I}\geq 0

In this subsection, we analyze and derive the basic estimates of of two fixed solutions y±(k,c,x2)y_{\pm}(k,c,x_{2}) to the homogeneous equation (3.1) with initial values

(3.53) y(h)=0,y(h)=1, and y+(0)=(U(0)c)2g+σk2,y+(0)=1+U(0)(U(0)c)g+σk2,y_{-}(-h)=0,\quad y_{-}^{\prime}(-h)=1,\quad\text{ and }\quad y_{+}(0)=\frac{(U(0)-c)^{2}}{g+\sigma k^{2}},\quad y_{+}^{\prime}(0)=1+\frac{U^{\prime}(0)(U(0)-c)}{g+\sigma k^{2}},

which also depend on parameters kk and cc\in\mathbb{C}. The initial condition of y+y_{+} at x2=0x_{2}=0 is motivated by the linearized capillary gravity water wave problem (2.11). (If it had been the linearized Euler equation at a shear flow in the channel, then naturally the boundary condition would be y+(0)=0y_{+}(0)=0 and y+(0)=1y_{+}^{\prime}(0)=1.) As throughout this section, we often skip the arguments rather than x2x_{2}. Particularly when working near x2c=U1(cR)x_{2}^{c}=U^{-1}(c_{R}), we shall continue using the notations introduced in Subsection 3.2, like cR,μ,ϵc_{R},\mu,\epsilon, etc. The following lemma is standard. Due to conjugacy, we only consider cI0c_{I}\geq 0.

Lemma 3.8.

For cU([h,0])c\notin U([-h,0]) and x2[h,0]x_{2}\in[-h,0], the solutions y±(k,c,x2)y_{\pm}(k,c,x_{2}) are even in kk, analytic in k2k^{2} and cc, and is Cl0+2C^{l_{0}+2} in x2x_{2}. Moreover y±(k,c¯,x2)=y±(k,c,x2)¯y_{\pm}(k,\bar{c},x_{2})=\overline{y_{\pm}(k,c,x_{2})}.

In the next step we give a priori estimates of y±(k,c,x2)y_{\pm}(k,c,x_{2}). In particular, we consider up to three subintervals,

(3.54) 2:=(x2l,x2r)={x2[h,0]:1|U(x2)c|>ρ0μ32},ρ0=4h0inf[h0h,h0]U,\mathcal{I}_{2}:=(x_{2l},x_{2r})=\left\{x_{2}\in[-h,0]\,:\,\frac{1}{|U(x_{2})-c|}>\rho_{0}\mu^{-\frac{3}{2}}\right\},\quad\rho_{0}=\frac{4}{h_{0}\inf_{[-h_{0}-h,h_{0}]}U^{\prime}},
(3.55) 1=[h,x2l),3=(x2r,0].\mathcal{I}_{1}=[-h,x_{2l}),\quad\mathcal{I}_{3}=(x_{2r},0].

Here μ=k1\mu=\langle k\rangle^{-1} as in (3.21). Clearly [h,0]=123[-h,0]=\mathcal{I}_{1}\cup\mathcal{I}_{2}\cup\mathcal{I}_{3} and any of these subintervals may be empty. If 2=\mathcal{I}_{2}=\emptyset, then [h,0][-h,0] is considered as 1\mathcal{I}_{1} for yy_{-} and as 3\mathcal{I}_{3} for y+y_{+} in the statement of the following lemma. The choice of the above constant ρ0\rho_{0} and the fact 0μ10\leq\mu\leq 1 ensure

(3.56) cRU([14h0h,14h0]) if 2.c_{R}\in U([-\tfrac{1}{4}h_{0}-h,\tfrac{1}{4}h_{0}])\;\text{ if }\;\mathcal{I}_{2}\neq\emptyset.
Lemma 3.9.

For any α(0,12)\alpha\in(0,\frac{1}{2}), there exists C>0C>0 depending only on α\alpha, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} (also on gg and σ\sigma for the estimates of y+y_{+}), such that, for any c\U([h,0])c\in\mathbb{C}\backslash U([-h,0]), the following hold:

(3.57) |μ1y(x2)sinh(μ1(x2+h))|Cμαsinh(μ1(x2+h)),|\mu^{-1}y_{-}(x_{2})-\sinh(\mu^{-1}(x_{2}+h))|\leq C\mu^{\alpha}\sinh(\mu^{-1}(x_{2}+h)),
(3.58) |μ1y+(x2)sinh(μ1x2)|C(μα+μ|c|2)cosh(μ1x2),|\mu^{-1}y_{+}(x_{2})-\sinh(\mu^{-1}x_{2})|\leq C\big{(}\mu^{\alpha}+\mu|c|^{2}\big{)}\cosh(\mu^{-1}x_{2}),

for all x2[h,0]x_{2}\in[-h,0]. Moreover, if 2=\mathcal{I}_{2}=\emptyset, then for all x2[h,0]x_{2}\in[-h,0],

(3.59) |y(x2)cosh(μ1(x2+h))|Cμαsinh(μ1(x2+h)),|y_{-}^{\prime}(x_{2})-\cosh(\mu^{-1}(x_{2}+h))|\leq C\mu^{\alpha}\sinh(\mu^{-1}(x_{2}+h)),
(3.60) |y+(x2)cosh(μ1x2)|C(μα+μ|c|2)cosh(μ1x2).|y_{+}^{\prime}(x_{2})-\cosh(\mu^{-1}x_{2})|\leq C\big{(}\mu^{\alpha}+\mu|c|^{2}\big{)}\cosh(\mu^{-1}x_{2}).

If otherwise 2\mathcal{I}_{2}\neq\emptyset, then

(3.61) |y(x2)cosh(μ1(x2+h))|{Cμαsinh(μ1(x2+h)),x21Cμαcosh(μ1(x2+h)),x23|y_{-}^{\prime}(x_{2})-\cosh(\mu^{-1}(x_{2}+h))|\leq\begin{cases}C\mu^{\alpha}\sinh(\mu^{-1}(x_{2}+h)),\qquad&x_{2}\in\mathcal{I}_{1}\\ C\mu^{\alpha}\cosh(\mu^{-1}(x_{2}+h)),&x_{2}\in\mathcal{I}_{3}\end{cases}
(3.62) |y+(x2)cosh(μ1x2)|Cμαcosh(μ1x2),x213,|y_{+}^{\prime}(x_{2})-\cosh(\mu^{-1}x_{2})|\leq C\mu^{\alpha}\cosh(\mu^{-1}x_{2}),\qquad x_{2}\in\mathcal{I}_{1}\cup\mathcal{I}_{3},

and for x22x_{2}\in\mathcal{I}_{2},

(3.63) |y(x2)cosh(μ1(x2+h))U(x2c)U(x2c)y(x2l)log|U(x2)c||Cμα(1+μ|log|U(x2)c||)cosh(μ1(x2+h)),\begin{split}\big{|}y_{-}^{\prime}(x_{2})-\cosh(\mu^{-1}(x_{2}+h))&-\frac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{-}(x_{2l})\log|U(x_{2})-c|\big{|}\\ &\leq C\mu^{\alpha}\big{(}1+\mu\big{|}\log|U(x_{2})-c|\big{|}\big{)}\cosh(\mu^{-1}(x_{2}+h)),\end{split}
(3.64) |y+(x2)cosh(μ1x2)U(x2c)U(x2c)y+(x2r)log|U(x2)c||Cμα(1+μ|log|U(x2)c||)cosh(μ1x2).\begin{split}\big{|}y_{+}^{\prime}(x_{2})-\cosh(\mu^{-1}x_{2})&-\frac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{+}(x_{2r})\log|U(x_{2})-c|\big{|}\\ &\leq C\mu^{\alpha}\big{(}1+\mu\big{|}\log|U(x_{2})-c|\big{|}\big{)}\cosh(\mu^{-1}x_{2}).\end{split}
Remark 3.4.

Even though the lemma assumes c\U([h,0])c\in\mathbb{C}\backslash U([-h,0]), the estimates are uniform in cc and thus they also hold for the limits of solutions as cI0+c_{I}\to 0+, while the limits as cI0c_{I}\to 0- are the conjugates of those as cI0+c_{I}\to 0+. Moreover, the constant CC does not depend on σ>0\sigma>0, and in particular, CC for yy_{-} does not depend on gg either.

It is possible that x2c[h,0]x_{2}^{c}\notin[-h,0] as the domain of UU has been extended. However, the constant CC in (3.59), (3.60), (3.61), and (3.62) are independent of the extensions of UU satisfying (3.5).

Proof.

The estimates of y±y_{\pm} can be derived in exactly the same procedure by reversing the direction of the variable x2x_{2}. We shall focus on y(k,c,x2)y_{-}(k,c,x_{2}) and give a brief description on the argument for y+y_{+} afterwards. The cases of x2x_{2} close to and away from x2cx_{2}^{c} will be considered differently based on Lemma 3.2 and Proposition 3.7, respectively.

Step 1. Assume 1\mathcal{I}_{1}\neq\emptyset. We consider kk in two cases. The first on is for those larger |k||k| such that

(3.65) ρ:=ρ0μ32k2(1+|U|C0([h0h,h0]))min{1,Cμ12},\rho:=\rho_{0}\mu^{-\frac{3}{2}}k^{-2}(1+|U^{\prime\prime}|_{C^{0}([-h_{0}-h,h_{0}])})\leq\min\{1,C\mu^{\frac{1}{2}}\},

where (3.8) is satisfied and Lemma 3.2 is applicable. Observe

(3.66) μ1k=1+k2k=1μ1+k(0,μ),\mu^{-1}-k=\sqrt{1+k^{2}}-k=\tfrac{1}{\mu^{-1}+k}\in(0,\mu),

and

(3.67) |sinh(μ1(x2+h))sinh(k(x2+h))|=2sinhx2+hμ1+kcosh(12(μ1+k)(x2+h))Cμsinh(μ1(x2+h)),\begin{split}|\sinh(\mu^{-1}(x_{2}+h))-\sinh(k(x_{2}+h))|=&2\sinh\tfrac{x_{2}+h}{\mu^{-1}+k}\cosh(\tfrac{1}{2}(\mu^{-1}+k)(x_{2}+h))\\ \leq&C\mu\sinh(\mu^{-1}(x_{2}+h)),\end{split}

where the last inequality could be derived by considering whether μ1(x2+h)1\mu^{-1}(x_{2}+h)\geq 1. The same upper bound also holds for cosh\cosh. Therefore applying Lemma 3.2 on 1\mathcal{I}_{1} with s=0s=0 and C0=0C_{0}=0, we immediately obtain the desired estimates (3.57), (3.59), (3.61) on yy_{-} and yy_{-}^{\prime} on 1\mathcal{I}_{1}, respectively. Otherwise in the case of smaller |k||k|, the desired estimates follows from Lemma 3.3 with ϕ=0\phi=0.

Step 2. Assume 2\mathcal{I}_{2}\neq\emptyset and x2r>x2lx_{2r}>x_{2l} otherwise step 1 has completed the proof. In this case, x2c[h04h,h04]x_{2}^{c}\in[-\tfrac{h_{0}}{4}-h,\tfrac{h_{0}}{4}] due to (3.56). Let

(3.68) M=ρ01|(U)1|C0=14h0,M=\rho_{0}^{-1}|(U^{\prime})^{-1}|_{C^{0}}=\tfrac{1}{4}h_{0},

which implies

(3.69) 2[x2cμM,x2c+μM][x2c2μM,x2c+2μM][h0h,h0].\mathcal{I}_{2}\subset[x_{2}^{c}-\mu M,x_{2}^{c}+\mu M]\subset[x_{2}^{c}-2\mu M,x_{2}^{c}+2\mu M]\subset[-h_{0}-h,h_{0}].

Therefore results in Subsection 3.2 in the corresponding rescaled variables w1,2(τ)w_{1,2}(\tau) and x2=x2c+μτx_{2}=x_{2}^{c}+\mu\tau given in (3.21) are applicable. Moreover the definition of 2\mathcal{I}_{2} further yields

|τ|=μ1|x2x2c|Cμ12,x22.|\tau|=\mu^{-1}|x_{2}-x_{2}^{c}|\leq C\mu^{\frac{1}{2}},\quad\forall\,x_{2}\in\mathcal{I}_{2}.

Let

τ0=μ1(x2lx2c).\tau_{0}=\mu^{-1}(x_{2l}-x_{2}^{c}).

Lemma 3.5 (with ϕ=0\phi=0) implies that, for any x22x_{2}\in\mathcal{I}_{2}

|y(x2)+U(x2c)U(x2c)y(x2l)log|U(x2l)c||U(x2)c|\displaystyle\Big{|}y_{-}^{\prime}(x_{2})+\tfrac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{-}(x_{2l})\log\tfrac{|U(x_{2l})-c|}{|U(x_{2})-c|} y(x2l)|C(1+μα2|log|U(x2l)c||)|y(x2l)|\displaystyle-y_{-}^{\prime}(x_{2l})\Big{|}\leq C\big{(}1+\mu^{\frac{\alpha^{\prime}}{2}}\big{|}\log|U(x_{2l})-c|\big{|}\big{)}|y_{-}(x_{2l})|
+C(μα2+μ|τ0|α|log|U(x2)c||)(μ1|y(x2l)|+|y(x2l)|),\displaystyle+C\big{(}\mu^{\frac{\alpha^{\prime}}{2}}+\mu|\tau_{0}|^{\alpha^{\prime}}\big{|}\log|U(x_{2})-c|\big{|}\big{)}\big{(}\mu^{-1}|y_{-}(x_{2l})|+|y_{-}^{\prime}(x_{2l})|\big{)},

for any α(0,1)\alpha^{\prime}\in(0,1). Moving the log|U(x2l)c|\log|U(x_{2l})-c| term to the right side, we obtain

(3.70) |y(x2)U(x2c)U(x2c)y(x2l)log|U(x2)c|y(x2l)|C(1+|log|U(x2l)c||)|y(x2l)|+C(μα2+μ|τ0|α|log|U(x2)c||)(μ1|y(x2l)|+|y(x2l)|).\begin{split}\Big{|}y_{-}^{\prime}(x_{2})-\tfrac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{-}(x_{2l})&\log|U(x_{2})-c|-y_{-}^{\prime}(x_{2l})\Big{|}\leq C\big{(}1+\big{|}\log|U(x_{2l})-c|\big{|}\big{)}|y_{-}(x_{2l})|\\ &\quad+C\big{(}\mu^{\frac{\alpha^{\prime}}{2}}+\mu|\tau_{0}|^{\alpha^{\prime}}\big{|}\log|U(x_{2})-c|\big{|}\big{)}\big{(}\mu^{-1}|y_{-}(x_{2l})|+|y_{-}^{\prime}(x_{2l})|\big{)}.\end{split}

Notice that, no matter whether 1=\mathcal{I}_{1}=\emptyset or not, (3.57) and (3.59) are satisfied at x2lx_{2l} due to either the initial condition of y(x2)y_{-}(x_{2}) or the above step 1. On the one hand, regarding the above first term on the right side, it holds that either y(x2l)=0y_{-}(x_{2l})=0 if x2l=hx_{2l}=-h or μ32C|x2cx2l|\mu^{\frac{3}{2}}\leq C|x_{2}^{c}-x_{2l}| if x2l>hx_{2l}>-h, hence this term would only contribute an error term of at most O(μα|y(x2l)|)O(\mu^{-\alpha^{\prime\prime}}|y_{-}(x_{2l})|), for any α>0\alpha^{\prime\prime}>0, in the upper bounds. On the other hand, 0x2x2lCμ320\leq x_{2}-x_{2l}\leq C\mu^{\frac{3}{2}} implies that replacing the above μ1y(x2l)\mu^{-1}y_{-}(x_{2l}), y(x2l)y_{-}^{\prime}(x_{2l}) and cosh(μ1(x2l+h))\cosh(\mu^{-1}(x_{2l}+h)) by cosh(μ1(x2+h))\cosh(\mu^{-1}(x_{2}+h)) would also only produce an error terms of at most O(μα2cosh(μ1(x2+h)))O\big{(}\mu^{\frac{\alpha^{\prime}}{2}}\cosh(\mu^{-1}(x_{2}+h))\big{)} in the upper bounds. Therefore we have

(3.71) |y(x2)cosh(μ1(x2+h))U(x2c)U(x2c)y(x2l)log|U(x2)c||C(μα2+μ|τ0|α|log|U(x2)c||)cosh(μ1(x2+h)),\begin{split}\big{|}y_{-}^{\prime}(x_{2})-\cosh(\mu^{-1}(x_{2}+h))&-\tfrac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{-}(x_{2l})\log|U(x_{2})-c|\big{|}\\ &\leq C\big{(}\mu^{\frac{\alpha^{\prime}}{2}}+\mu|\tau_{0}|^{\alpha^{\prime}}\big{|}\log|U(x_{2})-c|\big{|}\big{)}\cosh(\mu^{-1}(x_{2}+h)),\end{split}

and thus (3.63) follows by letting α=2α\alpha^{\prime}=2\alpha.

Integrating (3.71) over [x2l,x2]2[x_{2l},x_{2}]\subset\mathcal{I}_{2}, we have, for α(2α,1)\alpha^{\prime}\in(2\alpha,1),

|μ1y(x2)sinh(μ1\displaystyle|\mu^{-1}y_{-}(x_{2})-\sinh(\mu^{-1} (x2+h))|Cμαsinh(μ1(x2l+h))+Cμx2lx2|y(x2l)|(1+|log|x2x2c||)\displaystyle(x_{2}+h))|\leq C\mu^{\alpha}\sinh(\mu^{-1}(x_{2l}+h))+\frac{C}{\mu}\int_{x_{2l}}^{x_{2}}|y_{-}(x_{2l})|\big{(}1+\big{|}\log|x_{2}^{\prime}-x_{2}^{c}|\big{|}\big{)}
+(μα2+μ|τ0|α|log|x2x2c||)cosh(μ1(x2l+h))dx2\displaystyle\qquad\qquad\quad+\big{(}\mu^{\frac{\alpha^{\prime}}{2}}+\mu|\tau_{0}|^{\alpha^{\prime}}\big{|}\log|x_{2}^{\prime}-x_{2}^{c}|\big{|}\big{)}\cosh(\mu^{-1}(x_{2l}+h))dx_{2}^{\prime}
\displaystyle\leq Cμαsinh(μ1(x2+h))+C|τ0|αcosh(μ1(x2l+h))x2lx2|log|x2x2c||dx2.\displaystyle C\mu^{\alpha}\sinh(\mu^{-1}(x_{2}+h))+C|\tau_{0}|^{\alpha^{\prime}}\cosh(\mu^{-1}(x_{2l}+h))\int_{x_{2l}}^{x_{2}}\big{|}\log|x_{2}^{\prime}-x_{2}^{c}|\big{|}dx_{2}^{\prime}.

where we used (3.57), |x2x2l|Cμ32|x_{2}-x_{2l}|\leq C\mu^{\frac{3}{2}}, and and the first term of the right side of (3.70) was incorporated into others as remarked just below (3.70). For |x2x2l|12|x2lx2c||x_{2}-x_{2l}|\leq\frac{1}{2}|x_{2l}-x_{2}^{c}|, we have

|τ0|αx2lx2|log|x2x2c||dx2μα|x2lx2c|α|x2x2l|(1+|log|x2lx2c||)Cμα|x2x2l|,|\tau_{0}|^{\alpha^{\prime}}\int_{x_{2l}}^{x_{2}}\big{|}\log|x_{2}^{\prime}-x_{2}^{c}|\big{|}dx_{2}^{\prime}\leq\mu^{-\alpha^{\prime}}|x_{2l}-x_{2}^{c}|^{\alpha^{\prime}}|x_{2}-x_{2l}|\big{(}1+\big{|}\log|x_{2l}-x_{2}^{c}|\big{|}\big{)}\leq C\mu^{\alpha}|x_{2}-x_{2l}|,

while for |x2x2l|12|x2lx2c||x_{2}-x_{2l}|\geq\frac{1}{2}|x_{2l}-x_{2}^{c}|,

|τ0|αx2lx2|log|x2x2c||dx2Cμα|x2lx2c|α|x2x2l|113(α2α)Cμα|x2x2l|.|\tau_{0}|^{\alpha^{\prime}}\int_{x_{2l}}^{x_{2}}\big{|}\log|x_{2}^{\prime}-x_{2}^{c}|\big{|}dx_{2}^{\prime}\leq C\mu^{-\alpha^{\prime}}|x_{2l}-x_{2}^{c}|^{\alpha^{\prime}}|x_{2}-x_{2l}|^{1-\frac{1}{3}(\alpha^{\prime}-2\alpha)}\leq C\mu^{\alpha}|x_{2}-x_{2l}|.

Therefore we obtain

|μ1y(x2)sinh(μ1(x2+h))|\displaystyle|\mu^{-1}y_{-}(x_{2})-\sinh(\mu^{-1}(x_{2}+h))|\leq Cμα(sinh(μ1(x2+h))+|x2x2l|cosh(μ1(x2l+h)))\displaystyle C\mu^{\alpha}\big{(}\sinh(\mu^{-1}(x_{2}+h))+|x_{2}-x_{2l}|\cosh(\mu^{-1}(x_{2l}+h))\big{)}
\displaystyle\leq Cμαsinh(μ1(x2+h))\displaystyle C\mu^{\alpha}\sinh(\mu^{-1}(x_{2}+h))

which proves (3.57) on 2\mathcal{I}_{2}.

Step 3. Assume 3=[x2r,0]\mathcal{I}_{3}=[x_{2r},0]\neq\emptyset, which implies x2r>hx_{2r}>-h. In this case, surely 2\mathcal{I}_{2}\neq\emptyset either and μ32C|U(x2r)c|\mu^{\frac{3}{2}}\leq C|U(x_{2r})-c|. With (3.57) for yy_{-} and (3.63) for yy_{-}^{\prime} established at x2=x2rx_{2}=x_{2r}, y(x2)y_{-}(x_{2}) satisfies assumption (3.17) for the interval 3\mathcal{I}_{3} with Θ1=sinh\Theta_{1}=\sinh, Θ2=cosh\Theta_{2}=\cosh, and C0=CμαC_{0}=C\mu^{\alpha}.

As in the step 1, for larger |k||k| so that (3.65) holds, the desired estimates (3.57) and (3.63) in 3\mathcal{I}_{3} follow directly from (3.66), (3.67), and Lemma 3.2.

For smaller kk, say, |k|k1|k|\leq k_{1}, we express y(x2)y_{-}(x_{2}) and y(x2)y_{-}^{\prime}(x_{2}) in terms of w(τ)w(\tau), τ[μ1(hx2c),μ1x2c]\tau\in[\mu^{-1}(-h-x_{2}^{c}),-\mu^{-1}x_{2}^{c}], as in (3.21). Let

M=(1+k12)12(2h0+h),τ0=μ1(hx2c).M=(1+k_{1}^{2})^{\frac{1}{2}}(2h_{0}+h),\quad\tau_{0}=\mu^{-1}(-h-x_{2}^{c}).

Since 2\mathcal{I}_{2}\neq\emptyset, otherwise [h,0]=1[-h,0]=\mathcal{I}_{1} for y(x2)y_{-}(x_{2}), it along with (3.56) and |k|k1|k|\leq k_{1} implies

x2c[h0h,h0]|h+x2c|,|x2c|2h0+h[μ1(hx2c),μ1x2c][M,M].x_{2}^{c}\in[-h_{0}-h,h_{0}]\Longrightarrow|h+x_{2}^{c}|,\,|x_{2}^{c}|\leq 2h_{0}+h\Longrightarrow[\mu^{-1}(-h-x_{2}^{c}),-\mu^{-1}x_{2}^{c}]\subset[-M,M].

Namely, the domain of w(τ)w(\tau) is contained in [M,M][-M,M]. Applying (3.40b) (with ϕ=0\phi=0), using w1(τ0)=0w_{1}(\tau_{0})=0, w2(τ0)=1w_{2}(\tau_{0})=1, and

3ρ01μ32=|U(x2r)c||U(x2)c|,x23,\mathcal{I}_{3}\neq\emptyset\Longrightarrow\rho_{0}^{-1}\mu^{\frac{3}{2}}=|U(x_{2r})-c|\leq|U(x_{2})-c|,\quad\forall x_{2}\in\mathcal{I}_{3},

we obtain |y(x2)|C|y_{-}^{\prime}(x_{2})|\leq C on 3\mathcal{I}_{3}. It in turn implies

|μ1y(x2)sinh(μ1(x2+h))|\displaystyle|\mu^{-1}y_{-}(x_{2})-\sinh(\mu^{-1}(x_{2}+h))|\leq μ1|y(x2)y(x2r)|+|μ1y(x2r)sinh(μ1(x2r+h))|\displaystyle\mu^{-1}|y_{-}(x_{2})-y_{-}(x_{2r})|+|\mu^{-1}y_{-}(x_{2r})-\sinh(\mu^{-1}(x_{2r}+h))|
+|sinh(μ1(x2r+h))sinh(μ1(x2+h))|\displaystyle+|\sinh(\mu^{-1}(x_{2r}+h))-\sinh(\mu^{-1}(x_{2}+h))|
\displaystyle\leq C(|x2x2r|+sinh(μ1(x2r+h)))Csinh(μ1(x2+h)).\displaystyle C\big{(}|x_{2}-x_{2r}|+\sinh(\mu^{-1}(x_{2r}+h))\big{)}\leq C\sinh(\mu^{-1}(x_{2}+h)).

Therefore (3.57) and (3.61) hold on 3\mathcal{I}_{3} due to |k|k1|k|\leq k_{1}.

Estimating y+y_{+} Finally, we give a brief sketch of the argument for y+y_{+}, for which we proceed from 3\mathcal{I}_{3} to 1\mathcal{I}_{1}.

Suppose 3\mathcal{I}_{3}\neq\emptyset. The initial values of y+y_{+} at x2=0x_{2}=0 satisfy (3.18) with Θ1=Θ2=cosh\Theta_{1}=\Theta_{2}=\cosh and C0=C(1+|c|2)μC_{0}=C(1+|c|^{2})\mu. For larger |k||k| so that (3.65) holds, the desired estimates (3.58) and (3.60) in 3\mathcal{I}_{3} follow directly from (3.66), (3.67), and Lemma 3.2. The estimates for smaller kk is again a consequence of Lemma 3.3.

Suppose 2\mathcal{I}_{2}\neq\emptyset which implies |c|C|c|\leq C. Inequality (3.70) with x2lx_{2l} replaced by x2rx_{2r} still follows from exactly the same argument, namely, for x22x_{2}\in\mathcal{I}_{2} and any α[0,1)\alpha^{\prime}\in[0,1),

|y+(x2)U(x2c)U(x2c)y+(x2r)\displaystyle\Big{|}y_{+}^{\prime}(x_{2})-\tfrac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{+}(x_{2r}) log|U(x2)c|y+(x2r)|C(1+|log|U(x2r)c||)|y+(x2r)|\displaystyle\log|U(x_{2})-c|-y_{+}^{\prime}(x_{2r})\Big{|}\leq C\big{(}1+\big{|}\log|U(x_{2r})-c|\big{|}\big{)}|y_{+}(x_{2r})|
+C(μα2+μ|τ0|α|log|U(x2)c||)(μ1|y+(x2r)|+|y+(x2r)|).\displaystyle\quad+C\big{(}\mu^{\frac{\alpha^{\prime}}{2}}+\mu|\tau_{0}|^{\alpha^{\prime}}\big{|}\log|U(x_{2})-c|\big{|}\big{)}\big{(}\mu^{-1}|y_{+}(x_{2r})|+|y_{+}^{\prime}(x_{2r})|\big{)}.

If x2r=0x_{2r}=0, then

|log|U(x2r)c|||y+(x2r)|=|log|U(0)c|||y+(0)|Cμ2coshμ1x2.\big{|}\log|U(x_{2r})-c|\big{|}|y_{+}(x_{2r})|=\big{|}\log|U(0)-c|\big{|}|y_{+}(0)|\leq C\mu^{2}\cosh\mu^{-1}x_{2}.

Otherwise, x2r<0x_{2r}<0 and thus, for any α(0,1)\alpha^{\prime}\in(0,1),

|U(x2r)c|=ρ01μ32|log|U(x2r)c|||y+(x2r)|μαcoshμ1x2,|U(x_{2r})-c|=\rho_{0}^{-1}\mu^{\frac{3}{2}}\,\Longrightarrow\big{|}\log|U(x_{2r})-c|\big{|}|y_{+}(x_{2r})|\leq\mu^{\alpha^{\prime}}\cosh\mu^{-1}x_{2},

where (3.58) at x2=x2rx_{2}=x_{2r} was also used. These estimates, along with (3.58) and (3.60) at x2=x2rx_{2}=x_{2r} yield (3.64) on 2\mathcal{I}_{2}. Inequality (3.58) follows from direct integrating the estimate on y+y_{+}^{\prime}, actually without going through the technical argument at the end of step 2 for yy_{-} since the cosh\cosh, instead of sinh\sinh, is in the upper bound in (3.58).

Suppose 1\mathcal{I}_{1}\neq\emptyset where it must hold 2\mathcal{I}_{2}\neq\emptyset and |c|C|c|\leq C. From step 2, y+(x2)y_{+}(x_{2}) satisfies assumption (3.18) for the interval 1\mathcal{I}_{1} with Θ1=Θ2=cosh\Theta_{1}=\Theta_{2}=\cosh, and C0=CμαC_{0}=C\mu^{\alpha}. For larger |k||k|, the desired estimates (3.58) and (3.62) follow from Lemma 3.2 and for smaller |k||k| from Lemma 3.3. ∎

3.4. Limits of solutions to the homogeneous Rayleigh equation with cI=0+c_{I}=0+

Now that the convergence of solutions of the Rayleigh equation as cI0+c_{I}\to 0+ has been established in Proposition 3.7, in this subsection, we shall focus on the analysis of the limit equation (3.24) along with the jump condition (3.44) at the singularity τ=0\tau=0. In this subsection we consider c=U(x2c)U([12h0h,12h0])c=U(x_{2}^{c})\in U\big{(}[-\tfrac{1}{2}h_{0}-h,\tfrac{1}{2}h_{0}]\big{)} unless otherwise specified. As transformation (3.34) was rather helpful in the proof of Proposition 3.7, its limit would also turn out to be an effective tool in the study of (3.24). However B~(τ)\tilde{B}(\tau) as well as B~0(τ)\tilde{B}_{0}(\tau) appears only Hölder in τ\tau, or equivalently in x2x_{2}. In the notations given in (3.21) in Subsection 3.2, we first prove the following lemma to isolate the singularity in B~0\tilde{B}_{0}. Recall UCl0U\in C^{l_{0}}, x2cx_{2}^{c} and cRc_{R} correspond to each other via (3.6), U~,U1Cl01\tilde{U},U_{1}\in C^{l_{0}-1}, and U2Cl02U_{2}\in C^{l_{0}-2} are defined in (3.26), and Γ0(μ,c,τ)=Γ(μ,c,ϵ=0,τ)\Gamma_{0}(\mu,c,\tau)=\Gamma(\mu,c,\epsilon=0,\tau) in (3.28).

Lemma 3.10.

Assume l03l_{0}\geq 3. There exists a unique continuous-in-τ\tau real 2×22\times 2 matrix valued B(μ,c,τ)B(\mu,c,\tau) satisfying

(3.72) Bτ=(011μ2+μU2U~0)BB(001+μU2(0)U~(0)τ0),τ0;B(μ,c,0)=I2×2.B_{\tau}=\begin{pmatrix}0&1\\ 1-\mu^{2}+\frac{\mu U_{2}}{\tilde{U}}&0\end{pmatrix}B-B\begin{pmatrix}0&0\\ 1+\frac{\mu U_{2}(0)}{\tilde{U}(0)\tau}&0\end{pmatrix},\;\;\tau\neq 0;\;\quad B(\mu,c,0)=I_{2\times 2}.

Moreover the following hold.

  1. (1)

    The matrix B(μ,c,τ)B(\mu,c,\tau) is Cl02C^{l_{0}-2} in cU([12h0h,12h0])c\in U\big{(}[-\tfrac{1}{2}h_{0}-h,\tfrac{1}{2}h_{0}]\big{)}, τ\tau, and μ\mu and

    detB=1,Bτ(μ,c,0)=(μU(x2c)U(x2c)1μ2(1+U(x2c)U(x2c)5U(x2c)22U(x2c)2)μU(x2c)U(x2c)),\det B=1,\;\;B_{\tau}(\mu,c,0)=\begin{pmatrix}-\frac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}&1\\ \mu^{2}\big{(}-1+\frac{U^{\prime\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}-\frac{5U^{\prime\prime}(x_{2}^{c})^{2}}{2U^{\prime}(x_{2}^{c})^{2}}\big{)}&\frac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\end{pmatrix},
    B(0,c,τ)=(coshττsinhτsinhτsinhττcoshτcoshτ)=(coshτsinhτsinhτcoshτ)(10τ1).B(0,c,\tau)=\begin{pmatrix}\cosh\tau-\tau\sinh\tau&\sinh\tau\\ \sinh\tau-\tau\cosh\tau&\cosh\tau\end{pmatrix}=\begin{pmatrix}\cosh\tau&\sinh\tau\\ \sinh\tau&\cosh\tau\end{pmatrix}\begin{pmatrix}1&0\\ -\tau&1\end{pmatrix}.

    Moreover for any M>0M>0 satisfying (3.23), there exists C>0C>0 depending only on |U|Cl01|U^{\prime}|_{C^{l_{0}-1}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that |B|Cl02C|B|_{C^{l_{0}-2}}\leq C.

  2. (2)

    BB and B~0\tilde{B}_{0} are conjugate, namely,

    (3.73) B(μ,c,τ)=(10Γ0(μ,c,τ)1)B~0(μ,c,τ)(10Γ0#(μ,c,τ)1),B(\mu,c,\tau)=\begin{pmatrix}1&0\\ \Gamma_{0}(\mu,c,\tau)&1\end{pmatrix}\tilde{B}_{0}(\mu,c,\tau)\begin{pmatrix}1&0\\ -\Gamma_{0}^{\#}(\mu,c,\tau)&1\end{pmatrix},

    where

    Γ0#(μ,c,τ)=τ+μU(x2c)U(x2c)(log|τ|+iπ2(sgn(τ)+1)).\Gamma_{0}^{\#}(\mu,c,\tau)=\tau+\frac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tau|+\frac{i\pi}{2}(sgn(\tau)+1)\big{)}.
  3. (3)

    General solutions to (3.24) satisfying (3.44) are

    (3.74) W(τ)=(W1(τ)W2(τ))=B(μ,c,τ)(10Γ0#(μ,c,τ)1)(bΦ~0(μ,c,τ)),b=(b1b2)2,=((B11+Γ0#B12)(b1Φ~01)+B12(b2Φ~02)(B21+Γ0#B22)(b1Φ~01)+B22(b2Φ~02)),\begin{split}W(\tau)=&\begin{pmatrix}W_{1}(\tau)\\ W_{2}(\tau)\end{pmatrix}=B(\mu,c,\tau)\begin{pmatrix}1&0\\ \Gamma_{0}^{\#}(\mu,c,\tau)&1\end{pmatrix}\big{(}b-\tilde{\Phi}_{0}(\mu,c,\tau)\big{)},\quad b=\begin{pmatrix}b_{1}\\ b_{2}\end{pmatrix}\in\mathbb{C}^{2},\\ =&\begin{pmatrix}(B_{11}+\Gamma_{0}^{\#}B_{12})(b_{1}-\tilde{\Phi}_{01})+B_{12}(b_{2}-\tilde{\Phi}_{02})\\ (B_{21}+\Gamma_{0}^{\#}B_{22})(b_{1}-\tilde{\Phi}_{01})+B_{22}(b_{2}-\tilde{\Phi}_{02})\end{pmatrix},\end{split}

    where Bj1j2B_{j_{1}j_{2}} are the entries of BB and Φ~0=(Φ~01,Φ~02)T=limϵ0+Φ~\tilde{\Phi}_{0}=(\tilde{\Phi}_{01},\tilde{\Phi}_{02})^{T}=\lim_{\epsilon\to 0+}\tilde{\Phi} with Φ~\tilde{\Phi} given in (3.30).

  4. (4)

    If ϕ0\phi\equiv 0, the general solution W(τ)W(\tau) to (3.24)–(3.44) with b2b\in\mathbb{C}^{2} as in (3.74) satisfies W1ClocαW_{1}\in C_{loc}^{\alpha^{\prime}} for any α(0,1)\alpha^{\prime}\in(0,1), W1(0)=b1W_{1}(0)=b_{1}, and

    limτ0(W2(τ)b2W1(0)μU(x2c)U(x2c)(log(U(x2c)|τ|)+iπ2(sgn(τ)+1))=0.\lim_{\tau\to 0}\Big{(}W_{2}(\tau)-b_{2}-W_{1}(0)\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log(U^{\prime}(x_{2}^{c})|\tau|)+\frac{i\pi}{2}(sgn(\tau)+1)\Big{)}=0.
  5. (5)

    Finally, W(τ)W(\tau) are Cl02C^{l_{0}-2} in μ\mu, cRc_{R}, and τ\tau if ϕ0\phi\equiv 0 and b1=W1(0)=0b_{1}=W_{1}(0)=0.

Remark 3.5.

If needed, higher order Taylor expansions of BB can be obtained based on (3.75) through rather standard calculations in the analysis of local invariant manifolds.

One is reminded that both Γ0(τ)\Gamma_{0}(\tau) has a logarithmic singularity and a jump at τ=0\tau=0 which leads to such singularities of W2(τ)W_{2}(\tau) there even in the homogeneous case. Since Γ0\Gamma_{0}\notin\mathbb{R} for τ>0\tau>0, B~0\tilde{B}_{0} should not be real for τ>0\tau>0. Hence it is a non-obvious statement that this conjugate matrix BB is real. The above lemma isolates the singularity of B~0\tilde{B}_{0} into the explicit Γ0\Gamma_{0} along with the smooth BB. Conceptually, the smoothness of BB in cRc_{R} is related to the smoothness of the spectral resolution of the identity with respect to the spectral parameter, and thus would play crucial role in proving the partial inviscid damping to the linearized Euler equation at the shear flow U(x2)U(x_{2}).

Proof.

The construction of B(μ,c,τ)B(\mu,c,\tau) is adapted from the one in [5], where the main issue is to handle the singularity caused by U~(μ,c,0)=0\tilde{U}(\mu,c,0)=0. We first make (3.72) autonomous by changing the independent variable an auxiliary one ss such that τs=τ\tau_{s}=\tau and thus we have

(3.75) {Bs=(0τ(1μ2)τ+μτU2U~0)BB(00τ+μU2(0)U1(0)0),τs=τ.\begin{cases}B_{s}=\begin{pmatrix}0&\tau\\ (1-\mu^{2})\tau+\frac{\mu\tau U_{2}}{\tilde{U}}&0\end{pmatrix}B-B\begin{pmatrix}0&0\\ \tau+\frac{\mu U_{2}(0)}{U_{1}(0)}&0\end{pmatrix},\\ \tau_{s}=\tau.\end{cases}

Obviously solutions to (3.72) correspond (up to a translation in ss) to those to the Cl02C^{l_{0}-2} ODE system (3.75) of 5-dim which converge to (I2×2,0)(I_{2\times 2},0) as ss\to-\infty, namely those on the unstable manifold of the steady state (I2×2,0)(I_{2\times 2},0). The linearized system of (3.75) at (I2×2,0)(I_{2\times 2},0) is given by

{Bs=μU(x2c)U(x2c)𝒜B+τ(01μ2(1+U3(0)U1(0)U2(0)22U1(0)2)0), where 𝒜B=(0010)BB(0010),τs=τ.\begin{cases}B_{s}=\frac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\mathcal{A}B+\tau\begin{pmatrix}0&1\\ \mu^{2}\big{(}-1+\frac{U_{3}(0)}{U_{1}(0)}-\frac{U_{2}(0)^{2}}{2U_{1}(0)^{2}}\big{)}&0\end{pmatrix},\quad\text{ where }\mathcal{A}B=\begin{pmatrix}0&0\\ 1&0\end{pmatrix}B-B\begin{pmatrix}0&0\\ 1&0\end{pmatrix},\\ \tau_{s}=\tau.\end{cases}

It is easy to compute that, on the one hand, an eigenvector associated to the eigenvalue 11 is

(B1,1),B1=(μU2(0)U1(0)1μ2(1+U3(0)U1(0)5U2(0)22U1(0)2)μU2(0)U1(0)).(B_{1},1),\quad B_{1}=\begin{pmatrix}-\frac{\mu U_{2}(0)}{U_{1}(0)}&1\\ \mu^{2}\big{(}-1+\frac{U_{3}(0)}{U_{1}(0)}-\frac{5U_{2}(0)^{2}}{2U_{1}(0)^{2}}\big{)}&\frac{\mu U_{2}(0)}{U_{1}(0)}\end{pmatrix}.

On the other hand, one may verify

es𝒜B=(10s1)B(10s1)e^{s\mathcal{A}}B=\begin{pmatrix}1&0\\ s&1\end{pmatrix}B\begin{pmatrix}1&0\\ -s&1\end{pmatrix}

which implies that in the 4-dim center subspace {τ=0}\{\tau=0\} there is not any decay backward in ss. Therefore there exists a unique Cl02C^{l_{0}-2} unstable manifold of 1-dim which corresponds a unique solution B(μ,c,τ)B(\mu,c,\tau) satisfying B(μ,c,0)=IB(\mu,c,0)=I and Bτ(μ,c,0)=B1B_{\tau}(\mu,c,0)=B_{1} and Cl02C^{l_{0}-2} in all its variables. In fact, the 4-dim center subspace {τ=0}\{\tau=0\} is also invariant under the nonlinear system (3.75), where the flow is given by the above non-decaying linear flow of conjugation. Therefore this B(μ,cR,τ)B(\mu,c_{R},\tau) is the only solution to (3.72) decaying to II as ss\to-\infty, or equivalently τ0+\tau\to 0+. Even though this construction is local in τ\tau, the domain of BB can be extended due to the linearity of equation (3.72).

With the existence of the Cl02C^{l_{0}-2} solution B(μ,c,τ)B(\mu,c,\tau) to (3.72) established through (3.75), letting μ=0\mu=0 in (3.75) and then transforming back to (3.72), we have

Bτ(0,c,τ)=(0110)B(0,c,τ)B(0,c,τ)(0010),τ0;B(0,c,0)=I2×2.B_{\tau}(0,c,\tau)=\begin{pmatrix}0&1\\ 1&0\end{pmatrix}B(0,c,\tau)-B(0,c,\tau)\begin{pmatrix}0&0\\ 1&0\end{pmatrix},\;\;\tau\neq 0;\quad B(0,c,0)=I_{2\times 2}.

This equation can be solved explicitly to yield

B(0,c,τ)=(coshτsinhτsinhτcoshτ)(10τ1)=(coshττsinhτsinhτsinhττcoshτcoshτ).B(0,c,\tau)=\begin{pmatrix}\cosh\tau&\sinh\tau\\ \sinh\tau&\cosh\tau\end{pmatrix}\begin{pmatrix}1&0\\ -\tau&1\end{pmatrix}=\begin{pmatrix}\cosh\tau-\tau\sinh\tau&\sinh\tau\\ \sinh\tau-\tau\cosh\tau&\cosh\tau\end{pmatrix}.

The conjugation relation is the consequence of the facts that both BB and the right side of (3.73) a.) are equal to II at τ=0\tau=0, b.) satisfy the same ODE system (3.72) for τ0\tau\neq 0, c.) are continuous in τ\tau due to the construction of BB and (3.31) in Lemma 3.4, and d.) the uniqueness of solutions to (3.72) satisfying a.)–c.), which is obtained in the above construction based on the local invariant manifold theory. The property detB=1\det B=1 follows directly from (3.73) and (3.31).

Formula (3.74) of the general solutions follows from the conjugacy relation (3.73) and Lemma 3.6. Under the assumption ϕ0\phi\equiv 0, since W2(τ)W_{2}(\tau) has at most logarithmic singularity at τ=0\tau=0 and W1τ=W2W_{1\tau}=W_{2}, the Hölder continuity of W1W_{1} in τ\tau follows. From formula (3.74) and |B(μ,c,0)I|=O(|τ|)|B(\mu,c,0)-I|=O(|\tau|), we obtain W1(0)=b1W_{1}(0)=b_{1}. The limit property of W2(τ)b2W_{2}(\tau)-b_{2} also follows from similar calculation. Finally, the Cl02C^{l_{0}-2} smoothness of W(τ)W(\tau) under the assumptions ϕ0\phi\equiv 0 and b1=W1(0)=0b_{1}=W_{1}(0)=0 is again obvious from the representation of the solution (3.74). The proof of the lemma is complete. ∎

For cU([h02h,h02])c\in U([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]), with the help of B(μ,c,τ)B(\mu,c,\tau) and Lemma 3.10 we shall analyze the 2×22\times 2 fundamental matrices in two different forms of the homogeneous problem (3.24) with the condition (3.44) at τ=0\tau=0

(3.76) S0(μ,c,τ)=B(μ,c,τ)(10Γ0#(μ,c,τ)1),S(μ,c,τ,τ0)=B(μ,c,τ)(10Γ0#(μ,c,τ)Γ0#(μ,c,τ0)1)B(μ,c,τ0)1,\begin{split}&S^{0}(\mu,c,\tau)=B(\mu,c,\tau)\begin{pmatrix}1&0\\ \Gamma_{0}^{\#}(\mu,c,\tau)&1\end{pmatrix},\\ &S(\mu,c,\tau,\tau_{0})=B(\mu,c,\tau)\begin{pmatrix}1&0\\ \Gamma_{0}^{\#}(\mu,c,\tau)-\Gamma_{0}^{\#}(\mu,c,\tau_{0})&1\end{pmatrix}B(\mu,c,\tau_{0})^{-1},\end{split}

where τ0\tau_{0} in SS is the initial value of the independent variable and hence S(μ,c,τ0,τ0)=IS(\mu,c,\tau_{0},\tau_{0})=I. To analyze S0S^{0} and SS, let

(3.77) S~0(μ,c,τ)=B(μ,c,τ)(0010)=(B12(μ,c,τ)0B22(μ,c,τ)0),S~(μ,c,τ,τ0)=B(μ,c,τ)(0010)B(μ,c,τ0)1=(B12(μ,c,τ)B22(μ,c,τ0)B12(μ,c,τ)B12(μ,c,τ0)B22(μ,c,τ)B22(μ,c,τ0)B22(μ,c,τ)B12(μ,c,τ0)),\begin{split}\tilde{S}^{0}(\mu,c,\tau)=&B(\mu,c,\tau)\begin{pmatrix}0&0\\ 1&0\end{pmatrix}=\begin{pmatrix}B_{12}(\mu,c,\tau)&0\\ B_{22}(\mu,c,\tau)&0\end{pmatrix},\\ \tilde{S}(\mu,c,\tau,\tau_{0})=&B(\mu,c,\tau)\begin{pmatrix}0&0\\ 1&0\end{pmatrix}B(\mu,c,\tau_{0})^{-1}\\ =&\begin{pmatrix}B_{12}(\mu,c,\tau)B_{22}(\mu,c,\tau_{0})&-B_{12}(\mu,c,\tau)B_{12}(\mu,c,\tau_{0})\\ B_{22}(\mu,c,\tau)B_{22}(\mu,c,\tau_{0})&-B_{22}(\mu,c,\tau)B_{12}(\mu,c,\tau_{0})\end{pmatrix},\end{split}

where detB=1\det B=1 was used to compute the more explicit form of S~\tilde{S} in the above, and

(3.78) Serr0=S0(coshτsinhτsinhτcoshτ)μU(x2c)U(x2c)(log|τ|+iπ2(sgn(τ)+1))S~0,Serr=S(cosh(ττ0)sinh(ττ0)sinh(ττ0)cosh(ττ0))μU(x2c)U(x2c)(log|ττ0|+iπ2(sgn(τ)sgn(τ0)))S~.\begin{split}&S_{err}^{0}=S^{0}-\begin{pmatrix}\cosh\tau&\sinh\tau\\ \sinh\tau&\cosh\tau\end{pmatrix}-\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tau|+\tfrac{i\pi}{2}\big{(}sgn(\tau)+1\big{)}\big{)}\tilde{S}^{0},\\ &S_{err}=S-\begin{pmatrix}\cosh(\tau-\tau_{0})&\sinh(\tau-\tau_{0})\\ \sinh(\tau-\tau_{0})&\cosh(\tau-\tau_{0})\end{pmatrix}-\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}\big{(}sgn(\tau)-sgn(\tau_{0})\big{)}\big{)}\tilde{S}.\end{split}

The following lemma provides some very basic estimates on SS. More detailed ones on SjlS_{jl} will be derived when needed.

Lemma 3.11.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3. The fundamental matrices S0(μ,c,τ)S^{0}(\mu,c,\tau) and S(μ,c,τ,τ0)S(\mu,c,\tau,\tau_{0}) and their entries Sjl0S_{jl}^{0} and Sj1j2S_{j_{1}j_{2}} satisfy the following for any α(0,1)\alpha\in(0,1).

  1. (1)

    S0S^{0} is Cl02C^{l_{0}-2} in its variables if τ0\tau\neq 0 and SS is Cl02C^{l_{0}-2} in its variables if τ0\tau\neq 0 and τ00\tau_{0}\neq 0.

  2. (2)

    S110S_{11}^{0}, S120S_{12}^{0}, and S220S_{22}^{0} are CαC^{\alpha} in τ\tau and Cl02C^{l_{0}-2} in μ\mu and cc. If τ00\tau_{0}\neq 0, then S11S_{11} and S12S_{12} are CαC^{\alpha} in τ\tau and Cl02C^{l_{0}-2} in μ\mu, cc, and τ0\tau_{0}.

  3. (3)

    If τ0\tau\neq 0, then S12S_{12} and S22S_{22} are CαC^{\alpha} in τ0\tau_{0} and Cl02C^{l_{0}-2} in μ\mu, cc, and τ\tau.

  4. (4)

    S12S_{12} and τ0S11\tau_{0}S_{11} are CαC^{\alpha} in τ0\tau_{0} and τ\tau and Cl02C^{l_{0}-2} in μ\mu and cc.

  5. (5)

    SerrS_{err} and Serr0S_{err}^{0} are Cl02C^{l_{0}-2} in μ[0,1]\mu\in[0,1], cU([12h0h,12h0])c\in U([-\frac{1}{2}h_{0}-h,\frac{1}{2}h_{0}]), and τ,τ0[M,M]\tau,\tau_{0}\in[-M,M].

  6. (6)

    For any MM satisfying (3.23), there exists C>0C>0 depending only on MM, |U|Cl01|U^{\prime}|_{C^{l_{0}-1}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} such that for any τ,τ0[M,M]\tau,\tau_{0}\in[-M,M],

    |μj1D1Dj2Serr|C|ττ0|,|D1Dj1Serr|Cμ,|μj1cj2Serr0|C|τ|,|\partial_{\mu}^{j_{1}}D_{1}\ldots D_{j_{2}}S_{err}|\leq C|\tau-\tau_{0}|,\quad|D_{1}\ldots D_{j_{1}}S_{err}|\leq C\mu,\quad|\partial_{\mu}^{j_{1}}\partial_{c}^{j_{2}}S_{err}^{0}|\leq C|\tau|,

    for 0j1l030\leq j_{1}\leq l_{0}-3, 0j2l03j10\leq j_{2}\leq l_{0}-3-j_{1}, and D1,Dj{c,1U(x2c)(τ+τ0)}D_{1},\ldots D_{j}\in\{\partial_{c},\tfrac{1}{U^{\prime}(x_{2}^{c})}(\partial_{\tau}+\partial_{\tau_{0}})\}, and for l04l_{0}\geq 4,

    |D1Dj2Serr|Cμ|ττ0|,|cj2Serr0|Cμ|τ|,|D_{1}\ldots D_{j_{2}}S_{err}|\leq C\mu|\tau-\tau_{0}|,\quad|\partial_{c}^{j_{2}}S_{err}^{0}|\leq C\mu|\tau|,

    for 0j2l040\leq j_{2}\leq l_{0}-4.

The reason we consider τ+τ0\partial_{\tau}+\partial_{\tau_{0}} of SS instead of individual τ\partial_{\tau} or τ0\partial_{\tau_{0}} is not only that it yields better estimate. Recall the change of variables τ=μ1(x2x2c)\tau=\mu^{-1}(x_{2}-x_{2}^{c}). The above fundamental matrix is in the form of S(μ,c,μ1(x2x2c),μ1(x20x2c))S\big{(}\mu,c,\mu^{-1}(x_{2}-x_{2}^{c}),\mu^{-1}(x_{20}-x_{2}^{c})\big{)}. Therefore cτ+τ0μU(x2c)\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})} corresponds to the partial differentiation with respect to cc in the (c,x2)(c,x_{2}) coordinates. Here we also used

(3.79) cx2c=1U(x2c).\partial_{c}x_{2}^{c}=\tfrac{1}{U^{\prime}(x_{2}^{c})}.
Proof.

The argument for S0S^{0} and SS are very similar and we shall mainly focus on SS. Let

(3.80) S#(μ,c,τ,τ0)=B(μ,c,τ)(10ττ01)B(μ,c,τ0)1.S^{\#}(\mu,c,\tau,\tau_{0})=B(\mu,c,\tau)\begin{pmatrix}1&0\\ \tau-\tau_{0}&1\end{pmatrix}B(\mu,c,\tau_{0})^{-1}.

Clearly we have

(3.81) S=S#+μU(x2c)U(x2c)(log|ττ0|+iπ2(sgn(τ)sgn(τ0)))S~.S=S^{\#}+\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\Big{(}\log\big{|}\tfrac{\tau}{\tau_{0}}\big{|}+\tfrac{i\pi}{2}\big{(}sgn(\tau)-sgn(\tau_{0})\big{)}\Big{)}\tilde{S}.

All the Cl02C^{l_{0}-2} smoothness follows from that of BB. The CαC^{\alpha} Hölder regularity in τ\tau and τ0\tau_{0} in statements (2)–(4) is due to B12(μ,c,0)=0B_{12}(\mu,c,0)=0.

Straight forward computation based on Lemma 3.10 yields

S#(0,c,τ,τ0)=(cosh(ττ0)sinh(ττ0)sinh(ττ0)cosh(ττ0)),S#(μ,c,τ0,τ0)=I.S^{\#}(0,c,\tau,\tau_{0})=\begin{pmatrix}\cosh(\tau-\tau_{0})&\sinh(\tau-\tau_{0})\\ \sinh(\tau-\tau_{0})&\cosh(\tau-\tau_{0})\end{pmatrix},\quad S^{\#}(\mu,c,\tau_{0},\tau_{0})=I.

Therefore

Serr(μ,c,τ,τ0)=S#(μ,c,τ,τ0)S#(0,c,τ,τ0).S_{err}(\mu,c,\tau,\tau_{0})=S^{\#}(\mu,c,\tau,\tau_{0})-S^{\#}(0,c,\tau,\tau_{0}).

It follows immediately that SerrS_{err} and its derivatives in cc, τ\tau, and τ0\tau_{0} are of the order O(|μ|)O(|\mu|). By mimicking f(μ,s)=f(0,s)+μ01fμ(θ1μ,0)+s01fμs(θ1μ,θ2s)dθ2dθ1f(\mu,s)=f(0,s)+\mu\int_{0}^{1}f_{\mu}(\theta_{1}\mu,0)+s\int_{0}^{1}f_{\mu s}(\theta_{1}\mu,\theta_{2}s)d\theta_{2}d\theta_{1}, we have

|S#(μ,c,τ,τ0)S#(0,c,τ,τ0)|=\displaystyle|S^{\#}(\mu,c,\tau,\tau_{0})-S^{\#}(0,c,\tau,\tau_{0})|= μ|ττ0||0101μτS#(θ1μ,c,τ0+θ2(ττ0),τ0)dθ2dθ1|\displaystyle\mu|\tau-\tau_{0}|\left|\int_{0}^{1}\int_{0}^{1}\partial_{\mu}\partial_{\tau}S^{\#}(\theta_{1}\mu,c,\tau_{0}+\theta_{2}(\tau-\tau_{0}),\tau_{0})d\theta_{2}d\theta_{1}\right|
\displaystyle\leq Cμ|ττ0|.\displaystyle C\mu|\tau-\tau_{0}|.

Moreover, for 1j2l041\leq j_{2}\leq l_{0}-4 and D1,Dj2{c,1U(x2c)(τ+τ0)}D_{1},\ldots D_{j_{2}}\in\{\partial_{c},\tfrac{1}{U^{\prime}(x_{2}^{c})}(\partial_{\tau}+\partial_{\tau_{0}})\}, we have

D1Dj2S#=0, if μ=0,μD1Dj2S#=0, if τ=τ0.D_{1}\ldots D_{j_{2}}S^{\#}=0,\text{ if }\mu=0,\qquad\partial_{\mu}D_{1}\ldots D_{j_{2}}S^{\#}=0,\text{ if }\tau=\tau_{0}.

A similar procedure yields

|D1Dj2S#(μ,c,τ,τ0)|=\displaystyle|D_{1}\ldots D_{j_{2}}S^{\#}(\mu,c,\tau,\tau_{0})|= μ|ττ0||0101μτD1Dj2S#(θ1μ,c,τ0+θ2(ττ0),τ0)dθ2dθ1|\displaystyle\mu|\tau-\tau_{0}|\left|\int_{0}^{1}\int_{0}^{1}\partial_{\mu}\partial_{\tau}D_{1}\ldots D_{j_{2}}S^{\#}(\theta_{1}\mu,c,\tau_{0}+\theta_{2}(\tau-\tau_{0}),\tau_{0})d\theta_{2}d\theta_{1}\right|
\displaystyle\leq Cμ|ττ0|.\displaystyle C\mu|\tau-\tau_{0}|.

Finally, since S#S^{\#} is Cl02C^{l_{0}-2} in all variables, for l04l_{0}\geq 4, 1j1l031\leq j_{1}\leq l_{0}-3, and 0j2l0j130\leq j_{2}\leq l_{0}-j_{1}-3, the estimate on μj1D1Dj2Serr\partial_{\mu}^{j_{1}}D_{1}\ldots D_{j_{2}}S_{err} follows from its C1C^{1} smoothness and vanishing at τ=τ0\tau=\tau_{0}.

To analyze S0S^{0}, parallelly we consider

S0#(μ,c,τ)=B(μ,c,τ)(10τ1).S_{0}^{\#}(\mu,c,\tau)=B(\mu,c,\tau)\begin{pmatrix}1&0\\ \tau&1\end{pmatrix}.

Subsequently we have

S0=S0#+μU(x2c)U(x2c)(log|τ|+iπ2(sgn(τ)+1))S~0,S0#|μ=0=(coshτsinhτsinhτcoshτ),S0#|τ=0=I.S^{0}=S_{0}^{\#}+\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tau|+\tfrac{i\pi}{2}(sgn(\tau)+1)\big{)}\tilde{S}^{0},\;\;S_{0}^{\#}|_{\mu=0}=\begin{pmatrix}\cosh\tau&\sinh\tau\\ \sinh\tau&\cosh\tau\end{pmatrix},\;\;S_{0}^{\#}|_{\tau=0}=I.

The rest of the proof follows exactly as in the case of SS. ∎

Recall the expressions (3.74) of a solution W(τ)W(\tau) to the non-homogeneous Rayleigh equation (3.24) along with (3.44) at τ=0\tau=0. We can use this formula to solve for the parameter bb from W(τ0)W(\tau_{0}) for some τ0[M,M]\tau_{0}\in[-M,M] and then rewrite W(τ)W(\tau) using the fundamental matrix S(μ,c,τ,τ0)S(\mu,c,\tau,\tau_{0}) as

(3.82) W(τ)=S(μ,c,τ,τ0)W(τ0)B(μ,c,τ)(10Γ0#(μ,c,τ)1)(Φ~0(μ,c,τ)Φ~0(μ,c,τ0)).\begin{split}W(\tau)=&S(\mu,c,\tau,\tau_{0})W(\tau_{0})-B(\mu,c,\tau)\begin{pmatrix}1&0\\ \Gamma_{0}^{\#}(\mu,c,\tau)&1\end{pmatrix}\big{(}\tilde{\Phi}_{0}(\mu,c,\tau)-\tilde{\Phi}_{0}(\mu,c,\tau_{0})\big{)}.\end{split}

3.5. Dependence in cc and kk of the fundamental solutions to the Homogeneous Rayleigh equation (3.1) with cI=0+c_{I}=0+

In this subsection we revisit the two fundamental solutions

(3.83) y0±(k,c,x2)=limcI0+y±(k,c+icI,x2),x2[h,0],y_{0\pm}(k,c,x_{2})=\lim_{c_{I}\to 0+}y_{\pm}(k,c+ic_{I},x_{2}),\quad x_{2}\in[-h,0],

of the homogeneous Rayleigh equation (3.1) for cU([h02h,h02])c\in U([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]) satisfying initial conditions (3.53). We often skip the dependence on cc and kk (or equivalently, on μ=(1+k2)12\mu=(1+k^{2})^{-\frac{1}{2}}) when there is no confusion. The following lemma is a summary of results from Proposition 3.7, Lemmas 3.10, and Remark 3.2, where x2cx_{2}^{c} is defined in (3.6).

Lemma 3.12.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3. For cU([h02h,h02])c\in U([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]) and x2[h,0]x_{2}\in[-h,0], the following hold.

  1. (1)

    As cI0+c_{I}\to 0+, y±(k,c+icI,x2)y0±(k,c,x2)y_{\pm}(k,c+ic_{I},x_{2})\to y_{0\pm}(k,c,x_{2}) uniformly in x2x_{2} and cc.

  2. (2)

    As cI0+c_{I}\to 0+, y±y0±y_{\pm}^{\prime}\to y_{0\pm}^{\prime} locally uniformly in {U(x2)c}\{U(x_{2})\neq c\} and also in LcLx2rL_{c}^{\infty}L_{x_{2}}^{r} and Lx2LcrL_{x_{2}}^{\infty}L_{c}^{r} for any r[1,)r\in[1,\infty).

  3. (3)

    For each cc, y0(x2)y_{0-}(x_{2})\in\mathbb{R} if U(x2)cU(x_{2})\leq c, y0+(x2)y_{0+}(x_{2})\in\mathbb{R} if U(x2)cU(x_{2})\geq c, y0±Cα([h02h,h02])y_{0\pm}\in C^{\alpha}([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]) for any α[0,1)\alpha\in[0,1) and Cl0C^{l_{0}} in x2x2cx_{2}\neq x_{2}^{c}.

  4. (4)

    Moreover,

    (1μy0±(x2)y0±(x2))=B(μ,c,1μ(x2x2c))(10Γ0#(μ,c,1μ(x2x2c))1)(1μy0±(x2c)b2±),\begin{pmatrix}\tfrac{1}{\mu}y_{0\pm}(x_{2})\\ y_{0\pm}^{\prime}(x_{2})\end{pmatrix}=B\big{(}\mu,c,\tfrac{1}{\mu}(x_{2}-x_{2}^{c})\big{)}\begin{pmatrix}1&0\\ \Gamma_{0}^{\#}\big{(}\mu,c,\tfrac{1}{\mu}(x_{2}-x_{2}^{c})\big{)}&1\end{pmatrix}\begin{pmatrix}\tfrac{1}{\mu}y_{0\pm}(x_{2}^{c})\\ b_{2\pm}\end{pmatrix},

    where

    b2±=limx2x2c(y0±(x2)U(x2c)U(x2c)y0±(x2c)(log(U(x2c)μ|x2x2c|)+iπ2(sgn(x2x2c)+1))) exists.b_{2\pm}=\lim_{x_{2}\to x_{2}^{c}}\Big{(}y_{0\pm}^{\prime}(x_{2})-\tfrac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{0\pm}(x_{2}^{c})\big{(}\log\big{(}\tfrac{U^{\prime}(x_{2}^{c})}{\mu}|x_{2}-x_{2}^{c}|\big{)}+\tfrac{i\pi}{2}(sgn(x_{2}-x_{2}^{c})+1)\big{)}\Big{)}\text{ exists}.
Remark 3.6.

When cc takes the end point values U(h)U(-h), according to the above representation formula and the smoothness of BB, actually y0Cl0([h0h,h0])y_{0-}\in C^{l_{0}}([-h_{0}-h,h_{0}]).

Remark 3.7.

Suppose cU((h,0))c\in U\big{(}(-h,0)\big{)} and y(k,c,x2)=limϵ0+y(k,c+iϵ,x2)y(k,c,x_{2})=\lim_{\epsilon\to 0+}y(k,c+i\epsilon,x_{2}) where y(k,c+iϵ,x2)y(k,c+i\epsilon,x_{2}) is a solution to the homogeneous Rayleigh equation (3.1) with y(h),y(h)y(-h),y^{\prime}(-h)\in\mathbb{R}. The above analysis in Subsection 3.2 implies that a.) y(k,c,x2)y(k,c,x_{2})\in\mathbb{R} for x2[h,x2c]x_{2}\in[-h,x_{2}^{c}]; and b.) if U(x2c)0U^{\prime\prime}(x_{2}^{c})\neq 0, an imaginary part Imy(k,c,x2)\text{Im}\,y(k,c,x_{2}) occurs for x2>x2cx_{2}>x_{2}^{c} which satisfies the homogeneous Rayleigh equation (3.1) for x2[x2c,0]x_{2}\in[x_{2}^{c},0] with initial condition

Imy(x2c)=0,Imy(x2c)=πU(x2c)U(x2c)y(x2c).\text{Im}\,y(x_{2}^{c})=0,\quad\text{Im}\,y^{\prime}(x_{2}^{c})=\tfrac{\pi U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y(x_{2}^{c}).

The main goal of this subsection is to analyze the differentiation of y0y_{0-} in cc. Even though most of the results also hold for y0+y_{0+}, the proof is slightly more technical. We shall skip those analysis of y+y_{+} as they are not necessary for the rest of the paper. See Remark 3.10.

The proof of the following lemma would be embedded in those of the four subsequent lemmas, actually mainly Lemma 3.15.

Lemma 3.13.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3. For k,ck,c\in\mathbb{R}, it holds that
a.) y0y_{0-} is locally CαC^{\alpha} in both kk and cc for any α[0,1)\alpha\in[0,1);
b.) (y0,y0)(y_{0-},y_{0-}^{\prime}) are locally CαC^{\alpha} in both kk and cc for any α[0,1)\alpha\in[0,1) at any (k,c,x2)(k,c,x_{2}) satisfying U(x2)cU(x_{2})\neq c;
c.) (y0,y0)(y_{0-},y_{0-}^{\prime}) are Cl02C^{l_{0}-2} in both kk and cc at any (k,c,x2)(k,c,x_{2}) satisfying U(x2)cU(x_{2})\neq c and cU(h)c\neq U(-h);
d.) y0(k,c,x2c)y_{0-}(k,c,x_{2}^{c}) is Cl02C^{l_{0}-2} in cc and kk if cU([h,0])c\in U([-h,0]);
e.) (y0,y0)(y_{0-},y_{0-}^{\prime}) are Cl02C^{l_{0}-2} in kk, at any (k,c,x2)(k,c,x_{2}) except for y0y_{0-}^{\prime} at c=U(x2)c=U(x_{2});
f.) assume l04l_{0}\geq 4, then, for any l=0,1l=0,1, j1,j20j_{1},j_{2}\geq 0, j1+j2l04j_{1}+j_{2}\leq l_{0}-4, r[1,)r\in[1,\infty), and x2[h,0]x_{2}\in[-h,0],

(U(h)c)j2kj1cj2x2ly0(k,c,x2),(U(-h)-c)^{j_{2}}\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}\partial_{x_{2}}^{l}y_{0-}(k,c,x_{2}),

is locally LkWc1,rL_{k}^{\infty}W_{c}^{1,r} for cc near U(h)U(-h).

To obtain the estimates, for fixed cc\in\mathbb{R} near U([h,0])U([-h,0]), as in Lemma 3.9, we divide [h,0][-h,0] into subintervals

(3.84) 2:=(x2l,x2r)={x2[h,0]:1|U(x2)c|>ρ0μ},1=[h,x2l],3=[x2r,0],\mathcal{I}_{2}:=(x_{2l},x_{2r})=\left\{x_{2}\in[-h,0]\,:\tfrac{1}{|U(x_{2})-c|}>\tfrac{\rho_{0}}{\mu}\right\},\quad\mathcal{I}_{1}=[-h,x_{2l}],\quad\mathcal{I}_{3}=[x_{2r},0],

where ρ0\rho_{0} is defined as in (3.54). 2\mathcal{I}_{2} is an interval due to the monotonic assumption of UU. Clearly [h,0]=123[-h,0]=\mathcal{I}_{1}\cup\mathcal{I}_{2}\cup\mathcal{I}_{3} and any of these subintervals may be empty. If 2=\mathcal{I}_{2}=\emptyset, then [h,0][-h,0] is considered as 1\mathcal{I}_{1} for y0y_{0-} and as 3\mathcal{I}_{3} for y0+y_{0+}. If 2\mathcal{I}_{2}\neq\emptyset, then (3.56) holds and x2c[12h0h,12h0]x_{2}^{c}\in[-\tfrac{1}{2}h_{0}-h,\tfrac{1}{2}h_{0}] is well defined. In the next three lemmas, we obtain the estimates on y0y_{0-} on subintervals in the order of 1\mathcal{I}_{1}, 2\mathcal{I}_{2}, and 3\mathcal{I}_{3}. The proof of Lemma 3.13 is mainly contained in that of Lemma 3.15 as the smooth dependence of solutions to the Rayleigh equation on kk and cc and the initial values is trivial on 1\mathcal{I}_{1} and 3\mathcal{I}_{3}. While we mainly focus on y0y_{0-} in the following lemmas, we shall also just outline the estimates on cy0+\partial_{c}y_{0+}, which would be enough for the rest of the paper.

Lemma 3.14.

Assume l02l_{0}\geq 2 and 1\mathcal{I}_{1}\neq\emptyset. For any kk\in\mathbb{R} and any cc\in\mathbb{R}, the following estimates hold for x21x_{2}\in\mathcal{I}_{1} and j1,j20j_{1},j_{2}\geq 0 with j1+j2>0j_{1}+j_{2}>0,

(3.85) μ1|kj1cj2y0(x2)|+|kj1cj2y0(x2)|Cj1,j2μ(|U(x2)c|j2+|U(h)c|j2)×(1+μj1(x2+h)j1)sinh(μ1(x2+h))Cj1,j2μ1j1j2sinh(μ1(x2+h)),\begin{split}\mu^{-1}|\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}(x_{2})|+|\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}^{\prime}(x_{2})|\leq&C_{j_{1},j_{2}}\mu\big{(}|U(x_{2})-c|^{-j_{2}}+|U(-h)-c|^{-j_{2}}\big{)}\\ &\times\big{(}1+\mu^{-j_{1}}(x_{2}+h)^{j_{1}}\big{)}\sinh(\mu^{-1}(x_{2}+h))\\ \leq&C_{j_{1},j_{2}}\mu^{1-j_{1}-j_{2}}\sinh(\mu^{-1}(x_{2}+h)),\end{split}

where Cj1,j2>0C_{j_{1},j_{2}}>0 depends only on j1j_{1}, j2j_{2}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}. Moreover, it also holds, for any x23x_{2}\in\mathcal{I}_{3}

μ1|cy0+(x2)|+|cy0+(x2)|C(sinhμ1|x2|+μ(1+|c|)coshμ1x2).\mu^{-1}|\partial_{c}y_{0+}(x_{2})|+|\partial_{c}y_{0+}^{\prime}(x_{2})|\leq C(\sinh\mu^{-1}|x_{2}|+\mu(1+|c|)\cosh\mu^{-1}x_{2}).

The above estimate holds in a neighborhood of 1\mathcal{I}_{1} actually.

Proof.

It is obvious that, for x21x_{2}\in\mathcal{I}_{1}, y0y_{0-} is analytic in cc and kk. Let K=k2=μ21K=k^{2}=\mu^{-2}-1. One may compute that Kj1cj2y0(x2)\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}(x_{2}) satisfies the non-homogeneous Rayleigh equation (3.3) in the form of

(3.86) Kj1cj2y0+(K+UUc)Kj1cj2y0=j1Kj11cj2y0j=0j21mj2,jU(Uc)j2+1jKj1cjy0,-\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}^{\prime\prime}+\big{(}K+\frac{U^{\prime\prime}}{U-c}\big{)}\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}=-j_{1}\partial_{K}^{j_{1}-1}\partial_{c}^{j_{2}}y_{0-}-\sum_{j^{\prime}=0}^{j_{2}-1}\frac{m_{j_{2},j^{\prime}}U^{\prime\prime}}{(U-c)^{j_{2}+1-j^{\prime}}}\partial_{K}^{j_{1}}\partial_{c}^{j^{\prime}}y_{0-},

with some constants mj2,jm_{j_{2},j^{\prime}}. Note that the definition of 2\mathcal{I}_{2} implies that (3.8) is satisfied on 1\mathcal{I}_{1} with

(3.87) ρ=ρ0μ1k2(1+|U|C0([h0h,h0]))=ρ0k11+k2(1+|U|C0([h0h,h0])).\rho=\rho_{0}\mu^{-1}k^{-2}(1+|U^{\prime\prime}|_{C^{0}([-h_{0}-h,h_{0}])})=\rho_{0}k^{-1}\sqrt{1+k^{-2}}(1+|U^{\prime\prime}|_{C^{0}([-h_{0}-h,h_{0}])}).

We shall estimate the derivatives of y0y_{0-} with respect to cc and kk for large kk and small kk separately.

For any k1k_{*}\geq 1 sufficiently large so that ρ1\rho\leq 1, we shall apply (3.19) with x02=hx_{02}=-h to prove

(3.88) μ1|Kj1cj2y0(x2)|+|Kj1cj2y0(x2)|Cj1,j2μ(x2+h)j1(|U(x2)c|j2+|U(h)c|j2)sinh(μ1(x2+h)),\begin{split}\mu^{-1}|\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}(x_{2})|+&|\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}^{\prime}(x_{2})|\\ \leq&C_{j_{1},j_{2}}\mu(x_{2}+h)^{j_{1}}\big{(}|U(x_{2})-c|^{-j_{2}}+|U(-h)-c|^{-j_{2}}\big{)}\sinh(\mu^{-1}(x_{2}+h)),\end{split}

for any |k|k|k|\geq k_{*}, j1,j20j_{1},j_{2}\geq 0 with j1+j2>0j_{1}+j_{2}>0. The proof is a simple mathematical induction in j1+j2j_{1}+j_{2}.

Since (3.88) does not include the case j1=j2=0j_{1}=j_{2}=0, there are two base cases (j1,j2)=(0,1)(j_{1},j_{2})=(0,1) and (j1,j2)=(1,0)(j_{1},j_{2})=(1,0), which we have to consider separately. For cy0\partial_{c}y_{0-}, from (3.86), (3.19), Lemma 3.9, and the definition of 2\mathcal{I}_{2}, we have, for any x21x_{2}\in\mathcal{I}_{1},

k|cy0(x2)|+|cy0(x2)|\displaystyle k|\partial_{c}y_{0-}(x_{2})|+|\partial_{c}y_{0-}^{\prime}(x_{2})|\leq Chx2cosh(μ1(x2x2))μsinh(μ1(x2+h))(U(x2)c)2dx2\displaystyle C\int_{-h}^{x_{2}}\cosh(\mu^{-1}(x_{2}-x_{2}^{\prime}))\frac{\mu\sinh(\mu^{-1}(x_{2}^{\prime}+h))}{(U(x_{2}^{\prime})-c)^{2}}dx_{2}^{\prime}
\displaystyle\leq Cμsinh(μ1(x2+h))hx21(U(x2)c)2dx2\displaystyle C\mu\sinh(\mu^{-1}(x_{2}+h))\int_{-h}^{x_{2}}\frac{1}{(U(x_{2}^{\prime})-c)^{2}}dx_{2}^{\prime}
\displaystyle\leq Cμ(|U(x2)c|1+|U(h)c|1)sinh(μ1(x2+h)),\displaystyle C\mu\big{(}|U(x_{2})-c|^{-1}+|U(-h)-c|^{-1}\big{)}\sinh(\mu^{-1}(x_{2}+h)),

where (3.66) and (3.67) are also used for kkk\geq k_{*} to convert the estimates in terms of kk into those in terms of μ\mu. Similarly, Ky0\partial_{K}y_{0-} satisfies

k|Ky0(x2)|+|Ky0(x2)|\displaystyle k|\partial_{K}y_{0-}(x_{2})|+|\partial_{K}y_{0-}^{\prime}(x_{2})|\leq Cμhx2cosh(μ1(x2x2))sinh(μ1(x2+h))dx2\displaystyle C\mu\int_{-h}^{x_{2}}\cosh(\mu^{-1}(x_{2}-x_{2}^{\prime}))\sinh(\mu^{-1}(x_{2}^{\prime}+h))dx_{2}^{\prime}
\displaystyle\leq Cμ(x2+h)sinh(μ1(x2+h)).\displaystyle C\mu(x_{2}+h)\sinh(\mu^{-1}(x_{2}+h)).

With the estimates in the base cases established, for j1+j2>1j_{1}+j_{2}>1, using the induction assumption (and Lemma 3.9 for j1=j=0j_{1}=j^{\prime}=0 in (3.86)) and proceeding much as in the above, we obtain

k|Kj1cj2y0(x2)|+|Kj1cj2y0(x2)|\displaystyle k|\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}(x_{2})|+|\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}y_{0-}^{\prime}(x_{2})|
\displaystyle\leq Cμsinh(μ1(x2+h))hx2j1(x2+h)j11(|U(x2)c|j2+|U(h)c|j2)\displaystyle C\mu\sinh(\mu^{-1}(x_{2}+h))\int_{-h}^{x_{2}}j_{1}(x_{2}^{\prime}+h)^{j_{1}-1}\big{(}|U(x_{2}^{\prime})-c|^{-j_{2}}+|U(-h)-c|^{-j_{2}}\big{)}
+(x2+h)j1|U(x2)c|2(|U(x2)c|1j2+|U(h)c|1j2)dx2,\displaystyle+(x_{2}^{\prime}+h)^{j_{1}}|U(x_{2}^{\prime})-c|^{-2}\big{(}|U(x_{2}^{\prime})-c|^{1-j_{2}}+|U(-h)-c|^{1-j_{2}}\big{)}dx_{2}^{\prime},

and (3.88) follows consequently.

For |k|k|k|\leq k_{*}, as μ1\mu\sim 1, we apply Lemma 3.3 to (3.86) on [h,x2][-h,x_{2}] with

C0=max{(U(h)c)1,(U(x2)c)1}ρ0μ1C.C_{0}=\max\{(U(-h)-c)^{-1},\ (U(x_{2})-c)^{-1}\}\leq\rho_{0}\mu^{-1}\leq C.

Following a similar induction procedure and using Lemma 3.3, we obtain, for x21x_{2}\in\mathcal{I}_{1}, l=0,1l=0,1, and j1,j20j_{1},j_{2}\geq 0 with j1+j2>0j_{1}+j_{2}>0,

|Kj1cj2x2ly0(x2)|\displaystyle|\partial_{K}^{j_{1}}\partial_{c}^{j_{2}}\partial_{x_{2}}^{l}y_{0-}(x_{2})|\leq Cj1,j2(x2+h)j1(|U(x2)c|j2+|U(h)c|j2).\displaystyle C_{j_{1},j_{2}}(x_{2}+h)^{j_{1}}\big{(}|U(x_{2})-c|^{-j_{2}}+|U(-h)-c|^{-j_{2}}\big{)}.

Therefore (3.88) holds for all kk\in\mathbb{R}.

Since

k=2kKkj=0lj2m~j,lkj2lKjl\partial_{k}=2k\partial_{K}\Longrightarrow\partial_{k}^{j}=\sum_{0\leq l\leq\frac{j}{2}}\tilde{m}_{j,l}k^{j-2l}\partial_{K}^{j-l}

for some constants m~j,l\tilde{m}_{j,l}, (3.88) implies (3.85) on 1\mathcal{I}_{1} (actually in a neighborhood of 1\mathcal{I}_{1}).

Estimating cy0+\partial_{c}y_{0+} on 3\mathcal{I}_{3}. Let y1(x2)y_{1}(x_{2}) be solutions to the homogeneous Rayleigh equation (3.1) with initial values

y1(0)=2(U(0)c)/(g+σk2),y1(0)=U(0)/(g+σk2),y_{1}(0)=-2(U(0)-c)/(g+\sigma k^{2}),\quad y_{1}^{\prime}(0)=-U^{\prime}(0)/(g+\sigma k^{2}),

and y2(x2)y_{2}(x_{2}) be the solution to the initial value problem of the non-homogeneous Rayleigh equation

y2+(k2+UUc)y0+=U(Uc)2y0+,y2(0)=y2(0)=0.-y_{2}^{\prime\prime}+\big{(}k^{2}+\tfrac{U^{\prime\prime}}{U-c}\big{)}y_{0+}=-\tfrac{U^{\prime\prime}}{(U-c)^{2}}y_{0+},\quad y_{2}(0)=y_{2}^{\prime}(0)=0.

On 3\mathcal{I}_{3}, y2y_{2} can be estimated much as y0y_{0-} on 1\mathcal{I}_{1}, while y1y_{1} much as in the proof of Lemma 3.9. When Lemma 3.2 is used to estimate y1y_{1} for large |k||k|, we set s=0s=0, Θ1=Θ2=cosh\Theta_{1}=\Theta_{2}=\cosh, and C0=Cμ(1+|c|)C_{0}=C\mu(1+|c|). The desired inequality on cy0+\partial_{c}y_{0+} follows from cy0+=y1+y2\partial_{c}y_{0+}=y_{1}+y_{2}. ∎

Lemma 3.15.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, and 2\mathcal{I}_{2}\neq\emptyset, then Lemma 3.13a.)–e.) hold for x22x_{2}\in\mathcal{I}_{2}. Moreover, there exists C>0C>0 depending only on |U|Cl01|U^{\prime}|_{C^{l_{0}-1}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any kk\in\mathbb{R} and any cc\in\mathbb{R}, the following estimates hold.

  1. (1)

    For 1jl021\leq j\leq l_{0}-2,

    (3.89) |cj(y0(k,c,x2c))|Cμ1jcosh(μ1(x2c+h)), if x2c[h,0];c(y0(k,c,x2c))|c=U(h)=U(h)1.\begin{split}&|\partial_{c}^{j}\big{(}y_{0-}(k,c,x_{2}^{c})\big{)}|\leq C\mu^{1-j}\cosh(\mu^{-1}(x_{2}^{c}+h)),\;\text{ if }x_{2}^{c}\in[-h,0];\\ &\partial_{c}\big{(}y_{0-}(k,c,x_{2}^{c})\big{)}|_{c=U(-h)}=U^{\prime}(-h)^{-1}.\end{split}
  2. (2)

    If l05l_{0}\geq 5, then, for any x22x_{2}\in\mathcal{I}_{2}, we have

    (3.90) μ1|cy0(x2)|C(1+|log|U(x2)c||U(x2l)c||)sinh(μ1(x2+h)),\begin{split}\mu^{-1}|\partial_{c}y_{0-}(x_{2})|\leq&C\left(1+\left|\log\frac{|U(x_{2})-c|}{|U(x_{2l})-c|}\right|\right)\sinh(\mu^{-1}(x_{2}+h)),\end{split}
    (3.91) |cy0(x2)+U(x2c)U(x2c)((P.V.)c(1U(x2)c)+iπδc(U(x2)c))y0(x2c)|C(1+|log|U(x2)c||U(x2l)c||)cosh(μ1(x2+h)),\begin{split}\Big{|}\partial_{c}y_{0-}^{\prime}(x_{2})+\frac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}(P.V.)_{c}(\frac{1}{U(x_{2})-c})&+i\pi\delta_{c}(U(x_{2})-c)\big{)}y_{0-}(x_{2}^{c})\Big{|}\\ \leq&C\Big{(}1+\Big{|}\log\frac{|U(x_{2})-c|}{|U(x_{2l})-c|}\Big{|}\Big{)}\cosh(\mu^{-1}(x_{2}+h)),\end{split}
    μ1|cy0+(x2)|C(1+|log|U(x2)c||U(x2r)c||)cosh(μ1x2),\mu^{-1}|\partial_{c}y_{0+}(x_{2})|\leq C\left(1+\left|\log\frac{|U(x_{2})-c|}{|U(x_{2r})-c|}\right|\right)\cosh(\mu^{-1}x_{2}),
    |cy0+(x2)+U(x2c)U(x2c)((P.V.)c(1U(x2)c)\displaystyle\Big{|}\partial_{c}y_{0+}^{\prime}(x_{2})+\frac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}(P.V.)_{c}(\frac{1}{U(x_{2})-c}) +iπδc(U(x2)c))y0+(x2c)|\displaystyle+i\pi\delta_{c}(U(x_{2})-c)\big{)}y_{0+}(x_{2}^{c})\Big{|}
    \displaystyle\leq C(1+|log|U(x2)c||U(x2r)c||)cosh(μ1x2),\displaystyle C\Big{(}1+\Big{|}\log\frac{|U(x_{2})-c|}{|U(x_{2r})-c|}\Big{|}\Big{)}\cosh(\mu^{-1}x_{2}),

    and for 2jl042\leq j\leq l_{0}-4 and cU(x2)c\neq U(x_{2}) and cU(h)c\neq U(-h),

    (3.92) μ1|cjy0(x2)|C(|U(x2)c|1j+|U(h)c|1j)sin(μ1(x2+h)),\begin{split}\mu^{-1}|\partial_{c}^{j}y_{0-}(x_{2})|\leq&C(|U(x_{2})-c|^{1-j}+|U(-h)-c|^{1-j})\sin(\mu^{-1}(x_{2}+h)),\end{split}
    (3.93) |cjy0(x2)|Cμ|U(x2)c|1(|U(x2)c|1j+|U(h)c|1j)sinh(μ1(x2+h)).\begin{split}|\partial_{c}^{j}y_{0-}^{\prime}(x_{2})|\leq&C\mu|U(x_{2})-c|^{-1}(|U(x_{2})-c|^{1-j}+|U(-h)-c|^{1-j})\sinh(\mu^{-1}(x_{2}+h)).\end{split}

In the above lemma δ()\delta(\cdot) denotes the delta mass supported at 0 and (P.V.)c(P.V.)_{c} and δc\delta_{c} emphasize them as distributions of the variable cc. Near U(x2)=cU(x_{2})=c or U(h)=cU(-h)=c, singular distributions of cjy0\partial_{c}^{j}y_{0-} and cjy0\partial_{c}^{j}y_{0-}^{\prime} at the level comparable to those negative exponents in (3.92) and (3.93) would occur. The quantities with loglog upper bounds are LpL^{p} functions for any p[1,)p\in[1,\infty).

Remark 3.8.

Statement (2) also holds for l03l_{0}\geq 3 with slightly weaker upper bounds. From the proof, it is easy to see that if l04l_{0}\geq 4, then (3.90), (3.91), and (3.92) and (3.93) for jl03j\leq l_{0}-3 hold with an additional μ1\mu^{-1} or all sinh\sinh replaced by cosh\cosh on the right sides. If l03l_{0}\geq 3, then these inequality hold for jl02j\leq l_{0}-2 with all sinh\sinh on the right sides replaced by cosh\cosh besides the additional μ1\mu^{-1}.

Proof.

Since 2\mathcal{I}_{2}\neq\emptyset, it is easy to prove that (3.56) holds and x2c[14h0h,14h0]x_{2}^{c}\in[-\tfrac{1}{4}h_{0}-h,\tfrac{1}{4}h_{0}] is well defined. Let MM be defined as in (3.68) and (3.69) still holds. This allows us to work in the τ=μ1(x2x2c)\tau=\mu^{-1}(x_{2}-x_{2}^{c}) coordinate and apply Lemma 3.10, 3.11, and 3.12. It is natural to express y0y_{0-} using the fundamental matrix S(μ,c,τ,τ0)S(\mu,c,\tau,\tau_{0}) defined in (3.76). One is reminded that x2lx_{2l} depends on cc. To study the regularity of y0y_{0-} and y0y_{0-}^{\prime} with respect to cc at some cU([12h0h,12h0])c_{*}\in U([-\tfrac{1}{2}h_{0}-h,\tfrac{1}{2}h_{0}]), we fix some x20[h,x2l(c)]x_{20}\in[-h,x_{2l}(c_{*})] (so independent of cc) in a O(μ)O(\mu) neighborhood of x2l(c)x_{2l}(c_{*}). For cc near cc_{*}, x22x_{2}\in\mathcal{I}_{2}, we can write

(3.94) (μ1y0(k,c,x2)y0(k,c,x2))=S(μ,c,τ,τ0)(μ1y0(k,c,x20)y0(k,c,x20)),τ=x2x2cμ,τ0=x20x2cμ.\begin{pmatrix}\mu^{-1}y_{0-}(k,c,x_{2})\\ y_{0-}^{\prime}(k,c,x_{2})\end{pmatrix}=S\big{(}\mu,c,\tau,\tau_{0}\big{)}\begin{pmatrix}\mu^{-1}y_{0-}(k,c,x_{20})\\ y_{0-}^{\prime}(k,c,x_{20})\end{pmatrix},\quad\tau=\frac{x_{2}-x_{2}^{c}}{\mu},\;\;\tau_{0}=\frac{x_{20}-x_{2}^{c}}{\mu}.

Note that τ=μ1(x2x2c)=0\tau=\mu^{-1}(x_{2}-x_{2}^{c})=0 iff U(x2)=cU(x_{2})=c and τ0=μ1(x20x2c)=0\tau_{0}=\mu^{-1}(x_{20}-x_{2}^{c})=0 iff U(x20)=cU(x_{20})=c, the latter of which happens iff U(h)=cU(-h)=c_{*}. Clearly y(x20)y_{-}(x_{20}) and y(x20)y_{-}^{\prime}(x_{20}) are smooth in cc and kk either due to the initial conditions or due to the smoothness of the Rayleigh equation on 1\mathcal{I}_{1}. Hence the regularity statement (c) of Lemma 3.13 follows from statement (1) in Lemma 3.11. If cU(x2)c\neq U(x_{2}) is close to U(h)U(-h), then we could fix x20=hx_{20}=-h. In this case, y0y_{0-} and y0y_{0-}^{\prime} involve only S12S_{12} and S22S_{22} due to y0(h)=0y_{0-}(-h)=0, and thus statement (b) follows from statement (3) in Lemma 3.11. When cc is close to U(x2)U(x_{2}), the CαC^{\alpha} regularity of y0y_{0-} in kk and cc is a consequence of statement (2) in Lemma 3.11, unless c=U(x2)=U(x20)=U(h)c=U(x_{2})=U(x_{20})=U(-h). Near the last exceptional case, the CαC^{\alpha} regularity of yy_{-} in kk and cc is due to (4) of Lemma 3.11. Statement (e) of the Cl02C^{l_{0}-2} smoothness in kk of (y0,y0)(y_{0-},y_{0-}^{\prime}) also following from the properties of SS given in Lemma 3.11.

We shall derive the estimates of the differentiation by c\partial_{c} at cc_{*} in two cases.

* Case 1: x2l(c)μhx_{2l}(c_{*})\geq\mu-h. In this case, fix x20=x2l(c)x_{20}=x_{2l}(c_{*}) which implies Cτ01-C\tau_{0}\geq 1. Hence sgn(τ0)=1sgn(\tau_{0})=-1 and log|τ0|\log|\tau_{0}| as well as its derivatives are of order O(1)O(1) when cc varies slightly. Therefore the τ0\tau_{0} related terms can be estimated easily. From the estimate at x20x_{20} derived in Lemma 3.14 (or from the initial condition at x2=hx_{2}=-h), (3.79), and Lemma 3.11, for 1jl031\leq j\leq l_{0}-3, it holds on 2\mathcal{I}_{2},

cj(μ1y0(x2)y0(x2))=\displaystyle\partial_{c}^{j}\begin{pmatrix}\mu^{-1}y_{0-}(x_{2})\\ y_{0-}^{\prime}(x_{2})\end{pmatrix}= j=0j(cτ+τ0μU(x2c))j(μU(x2c)U(x2c)(log|ττ0|+iπ2(sgn(τ)sgn(τ0)))S~(τ,τ0))\displaystyle\sum_{j^{\prime}=0}^{j}\big{(}\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})}\big{)}^{j^{\prime}}\Big{(}\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}(sgn(\tau)-sgn(\tau_{0}))\big{)}\tilde{S}(\tau,\tau_{0})\Big{)}
×cjj(μ1y0(x20)y0(x20))+O(μ1jsinh(μ1(x2+h)))\displaystyle\times\partial_{c}^{j-j^{\prime}}\begin{pmatrix}\mu^{-1}y_{0-}(x_{20})\\ y_{0-}^{\prime}(x_{20})\end{pmatrix}+O\big{(}\mu^{1-j}\sinh(\mu^{-1}(x_{2}+h))\big{)}
=\displaystyle= j=0j(cτ+τ0μU(x2c))j(μU(x2c)U(x2c)(log|τ|+iπ2sgn(τ))S~(τ,τ0))cjj(μ1y0(x20)y0(x20))\displaystyle\sum_{j^{\prime}=0}^{j}\big{(}\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})}\big{)}^{j^{\prime}}\Big{(}\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tau|+\tfrac{i\pi}{2}sgn(\tau)\big{)}\tilde{S}(\tau,\tau_{0})\Big{)}\partial_{c}^{j-j^{\prime}}\begin{pmatrix}\mu^{-1}y_{0-}(x_{20})\\ y_{0-}^{\prime}(x_{20})\end{pmatrix}
+O(μ1jsinh(μ1(x2+h))),\displaystyle+O\big{(}\mu^{1-j}\sinh(\mu^{-1}(x_{2}+h))\big{)},
=\displaystyle= j=0jcj(μU(x2c)U(x2c)(log|U(x2)c|μ+iπ2sgn(U(x2)c))S~(U(x2)cμ,U(x20)cμ))\displaystyle\sum_{j^{\prime}=0}^{j}\partial_{c}^{j^{\prime}}\Big{(}\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log\tfrac{|U(x_{2})-c|}{\mu}+\tfrac{i\pi}{2}sgn(U(x_{2})-c)\big{)}\tilde{S}\big{(}\tfrac{U(x_{2})-c}{\mu},\tfrac{U(x_{20})-c}{\mu}\big{)}\Big{)}
×cjj(μ1y0(x20)y0(x20))+O(μ1jsinh(μ1(x2+h))),\displaystyle\times\partial_{c}^{j-j^{\prime}}\begin{pmatrix}\mu^{-1}y_{0-}(x_{20})\\ y_{0-}^{\prime}(x_{20})\end{pmatrix}+O\big{(}\mu^{1-j}\sinh(\mu^{-1}(x_{2}+h))\big{)},

where S~\tilde{S} is given in (3.77) and the constant CC in the O()O(\cdot) terms depends only on |U|Cl01|U^{\prime}|_{C^{l_{0}-1}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}. We also used that sinh\sinh and cosh\cosh are comparable at x2+hμ\tfrac{x_{2}+h}{\mu} for x22x_{2}\in\mathcal{I}_{2} in this case.

For j=1j=1, keeping the most singular terms arising from the derivatives of log\log and sgnsgn in the distribution sense, we have

c(μ1y0(x2)y0(x2))=\displaystyle\partial_{c}\begin{pmatrix}\mu^{-1}y_{0-}(x_{2})\\ y_{0-}^{\prime}(x_{2})\end{pmatrix}= μU(x2c)U(x2c)((P.V.)c(1U(x2)c)+iπδc(U(x2)c))S~(τ,τ0)(μ1y0(x20)y0(x20))\displaystyle-\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}(P.V.)_{c}(\tfrac{1}{U(x_{2})-c})+i\pi\delta_{c}(U(x_{2})-c)\big{)}\tilde{S}(\tau,\tau_{0})\begin{pmatrix}\mu^{-1}y_{0-}(x_{20})\\ y_{0-}^{\prime}(x_{20})\end{pmatrix}
+O((1+|log|U(x2)c|μ|)sinh(μ1(x2+h))).\displaystyle+O\big{(}\big{(}1+\big{|}\log\tfrac{|U(x_{2})-c|}{\mu}\big{|}\big{)}\sinh(\mu^{-1}(x_{2}+h))\big{)}.

Using (3.77), (3.74), the smoothness of BB and B(μ,c,0)=IB(\mu,c,0)=I, one may compute

(3.95) (B22(τ0),B12(τ0))(μ1y0(x20),y0(x20))T=μ1y0(x2c),\big{(}B_{22}(\tau_{0}),-B_{12}(\tau_{0})\big{)}\big{(}\mu^{-1}y_{0-}(x_{20}),y_{0-}^{\prime}(x_{20})\big{)}^{T}=\mu^{-1}y_{0-}(x_{2}^{c}),
(3.96) S~(τ,τ0)(μ1y0(x20)y0(x20))=μ1y0(x2c)(B12(μ,c,τ)B22(μ,c,τ))=μ1y0(x2c)((01)+O(|τ|)),\tilde{S}(\tau,\tau_{0})\begin{pmatrix}\mu^{-1}y_{0-}(x_{20})\\ y_{0-}^{\prime}(x_{20})\end{pmatrix}=\mu^{-1}y_{0-}(x_{2}^{c})\begin{pmatrix}B_{12}(\mu,c,\tau)\\ B_{22}(\mu,c,\tau)\end{pmatrix}=\mu^{-1}y_{0-}(x_{2}^{c})\left(\begin{pmatrix}0\\ 1\end{pmatrix}+O(|\tau|)\right),

and it yields the desired estimates for j=1j=1 in this case.

Similarly, at x2x2cx_{2}\neq x_{2}^{c} for 2jl032\leq j\leq l_{0}-3, keeping the worst term and using (3.96), we have

cj(μ1y0(x2)y0(x2))=\displaystyle\partial_{c}^{j}\begin{pmatrix}\mu^{-1}y_{0-}(x_{2})\\ y_{0-}^{\prime}(x_{2})\end{pmatrix}= ((τμU(x2c))jlog|τ|)μU(x2c)U(x2c)S~(τ,τ0)(μ1y0(x20)y0(x20))\displaystyle\big{(}\big{(}-\tfrac{\partial_{\tau}}{\mu U^{\prime}(x_{2}^{c})}\big{)}^{j}\log|\tau|\big{)}\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\tilde{S}(\tau,\tau_{0})\begin{pmatrix}\mu^{-1}y_{0-}(x_{20})\\ y_{0-}^{\prime}(x_{20})\end{pmatrix}
+O(μ1j|τ|1jsinh(μ1(x2+h)))\displaystyle+O\big{(}\mu^{1-j}|\tau|^{1-j}\sinh(\mu^{-1}(x_{2}+h))\big{)}
=\displaystyle= μ1jsinh(μ1(x2+h))(O(|τ|1j)O(|τ|j)).\displaystyle\mu^{1-j}\sinh(\mu^{-1}(x_{2}+h))\begin{pmatrix}O(|\tau|^{1-j})\\ O(|\tau|^{-j})\end{pmatrix}.

The desired inequality (3.92) in case 1 follows.

To finish the analysis in this case, we consider y0(k,c,x2c)y_{0-}(k,c,x_{2}^{c}). From (3.95) we obtain Cl02C^{l_{0}-2} smoothness in kk and cc. Differentiating (3.95) in cc and using Lemma 3.10 and 3.14, one may estimate, for 1jl021\leq j\leq l_{0}-2,

|cj(y(c,x2c))|\displaystyle\big{|}\partial_{c}^{j}\big{(}y(c,x_{2}^{c})\big{)}\big{|}\leq |j=0j(((cτμU(x2c))jjB22)(c,τ0)cjy0(c,x20)\displaystyle\left|\sum_{j^{\prime}=0}^{j}\Big{(}\big{(}(\partial_{c}-\tfrac{\partial_{\tau}}{\mu U^{\prime}(x_{2}^{c})})^{j-j^{\prime}}B_{22}\big{)}(c,\tau_{0})\partial_{c}^{j^{\prime}}y_{0-}(c,x_{20})\right.
μ((cτμU(x2c))jjB12)(c,τ0)cjy0(c,x20))|Cμ1jcosh(μ1(x2c+h)),\displaystyle\quad\left.-\mu\big{(}(\partial_{c}-\tfrac{\partial_{\tau}}{\mu U^{\prime}(x_{2}^{c})})^{j-j^{\prime}}B_{12}\big{)}(c,\tau_{0})\partial_{c}^{j^{\prime}}y_{0-}^{\prime}(c,x_{20})\Big{)}\right|\leq C\mu^{1-j}\cosh(\mu^{-1}(x_{2}^{c}+h)),

which proves (3.89) in case 1.

* Case 2: hx2l(c)μh-h\leq x_{2l}(c_{*})\leq\mu-h. In this case, let x20=hx_{20}=-h. While we have to deal with possibly very small τ0\tau_{0} in (3.94), the initial values (y0(x20),y0(x20))=(0,1)(y_{0-}(x_{20}),y_{0-}^{\prime}(x_{20}))=(0,1). Hence from Lemma 3.11 we obtain, for 0jl040\leq j\leq l_{0}-4, x22x_{2}\in\mathcal{I}_{2},

(3.97) cj(μ1y0(x2)y0(x2))=(cτ+τ0μU(x2c))j(S12(τ,τ0)S22(τ,τ0))=(cτ+τ0μU(x2c))j(μU(x2c)U(x2c)(log|ττ0|+iπ2(sgn(τ)sgn(τ0)))(S~12(τ,τ0)S~22(τ,τ0)))+O(μ1j|ττ0|).\begin{split}&\partial_{c}^{j}\begin{pmatrix}\mu^{-1}y_{0-}(x_{2})\\ y_{0-}^{\prime}(x_{2})\end{pmatrix}=\big{(}\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})}\big{)}^{j}\begin{pmatrix}S_{12}(\tau,\tau_{0})\\ S_{22}(\tau,\tau_{0})\end{pmatrix}\\ =&\big{(}\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})}\big{)}^{j}\Big{(}\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}(sgn(\tau)-sgn(\tau_{0}))\big{)}\begin{pmatrix}\tilde{S}_{12}(\tau,\tau_{0})\\ \tilde{S}_{22}(\tau,\tau_{0})\end{pmatrix}\Big{)}\\ &+O(\mu^{1-j}|\tau-\tau_{0}|).\end{split}

From (3.77) and Lemma 3.10, S~12ττ0\tfrac{\tilde{S}_{12}}{\tau\tau_{0}} and S~22τ0\tfrac{\tilde{S}_{22}}{\tau_{0}} are Cl03C^{l_{0}-3} functions, which could be used to reduce some singularity. As μ|ττ0|=|x2+h|\mu|\tau-\tau_{0}|=|x_{2}+h|, one can compute for j=1j=1,

μ1cy0(x2)=\displaystyle\mu^{-1}\partial_{c}y_{0-}(x_{2})= U(x2c)U(x2c)2S~12ττ0(τ+τ0)(ττ0(log|ττ0|+iπ2(sgn(τ)sgn(τ0))))\displaystyle-\tfrac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})^{2}}\tfrac{\tilde{S}_{12}}{\tau\tau_{0}}(\partial_{\tau}+\partial_{\tau_{0}})\Big{(}\tau\tau_{0}\big{(}\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}(sgn(\tau)-sgn(\tau_{0}))\big{)}\Big{)}
+O(|ττ0(log|ττ0|+iπ2(sgn(τ)sgn(τ0)))|)+O(μ1|x2+h|).\displaystyle+O\Big{(}\big{|}\tau\tau_{0}\big{(}\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}(sgn(\tau)-sgn(\tau_{0}))\big{)}\big{|}\Big{)}+O(\mu^{-1}|x_{2}+h|).

We use the following elementary inequalities to handle the above log\log terms:

|log|ττ0||=||τ||τ0|1τdτ|||τ||τ0||min{|τ|,|τ0|},|τ|+|τ0||ττ0|+2min{|τ|,|τ0|},\big{|}\log|\tfrac{\tau}{\tau_{0}}|\big{|}=\left|\int_{|\tau|}^{|\tau_{0}|}\tfrac{1}{\tau^{\prime}}d\tau^{\prime}\right|\leq\frac{\big{|}|\tau|-|\tau_{0}|\big{|}}{\min\{|\tau|,|\tau_{0}|\}},\quad|\tau|+|\tau_{0}|\leq|\tau-\tau_{0}|+2\min\{|\tau|,|\tau_{0}|\},

which also imply

|ττ0log|ττ0||C|ττ0|,(|τ|+|τ0|)|log|ττ0|||ττ0|(2+|log|ττ0||).\big{|}\tau\tau_{0}\log|\tfrac{\tau}{\tau_{0}}|\big{|}\leq C|\tau-\tau_{0}|,\quad(|\tau|+|\tau_{0}|)\big{|}\log|\tfrac{\tau}{\tau_{0}}|\big{|}\leq|\tau-\tau_{0}|\big{(}2+\big{|}\log|\tfrac{\tau}{\tau_{0}}|\big{|}\big{)}.

The delta functions produced by differentiating sgnsgn are cancelled by ττ0\tau\tau_{0}. Finally sgn(τ)sgn(τ0)0sgn(\tau)-sgn(\tau_{0})\neq 0 only when hx2cx2-h\leq x_{2}^{c}\leq x_{2} which implies μ(|τ0|+|τ|)=x2+h\mu(|\tau_{0}|+|\tau|)=x_{2}+h. Summarizing these estimates we obtain

|μ1cy0(x2)|Cμ1|x2+h|(1+|log|ττ0||).|\mu^{-1}\partial_{c}y_{0-}(x_{2})|\leq C\mu^{-1}|x_{2}+h|\big{(}1+\big{|}\log|\tfrac{\tau}{\tau_{0}}|\big{|}\big{)}.

If μhx2l(c)>h\mu-h\geq x_{2l}(c_{*})>-h, then 1Cμx2c(c)x2l(c)μτ0Cμ\tfrac{1}{C}\mu\leq x_{2}^{c}(c_{*})-x_{2l}(c_{*})\leq-\mu\tau_{0}\leq C\mu, while μ|τ0|=|x2l(c)x2c(c)|\mu|\tau_{0}|=|x_{2l}(c_{*})-x_{2}^{c}(c_{*})| if x2l(c)=hx_{2l}(c_{*})=-h. Hence |log|μτ0|log|x2l(c)x2c(c)||C\big{|}\log|\mu\tau_{0}|-\log|x_{2l}(c_{*})-x_{2}^{c}(c_{*})|\big{|}\leq C, which along with the estimate in case 1 yields (3.90).

Much as in the above, we estimate cy0(x2)\partial_{c}y_{0-}^{\prime}(x_{2}) in case 2 using (3.77) and Lemma 3.10

cy0(x2)=\displaystyle\partial_{c}y_{0-}^{\prime}(x_{2})= μU(x2c)U(x2c)S~22τ0U(x20)cμc(log|U(x2)c||U(x20c|+iπ2(sgn(U(x2)c)sgn(U(x20)c)))+O(1+|log|ττ0||)\displaystyle\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\tfrac{\tilde{S}_{22}}{\tau_{0}}\tfrac{U(x_{20})-c}{\mu}\partial_{c}\big{(}\log\tfrac{|U(x_{2})-c|}{|U(x_{20}-c|}+\tfrac{i\pi}{2}(sgn(U(x_{2})-c)-sgn(U(x_{20})-c))\big{)}+O\big{(}1+\big{|}\log|\tfrac{\tau}{\tau_{0}}|\big{|}\big{)}
=\displaystyle= μU(x2c)U(x2)S~22((P.V.)c(1U(x2)c)+iπδc(U(x2)c))+O(1+|log|ττ0||)\displaystyle-\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2})}\tilde{S}_{22}\big{(}(P.V.)_{c}(\tfrac{1}{U(x_{2})-c})+i\pi\delta_{c}(U(x_{2})-c)\big{)}+O\big{(}1+\big{|}\log|\tfrac{\tau}{\tau_{0}}|\big{|}\big{)}
=\displaystyle= μU(x2c)U(x2c)B12(τ0)((P.V.)c(1U(x2)c)+iπδc(U(x2)c))+O(1+|log|ττ0||).\displaystyle\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}B_{12}(\tau_{0})\big{(}(P.V.)_{c}(\tfrac{1}{U(x_{2})-c})+i\pi\delta_{c}(U(x_{2})-c)\big{)}+O\big{(}1+\big{|}\log|\tfrac{\tau}{\tau_{0}}|\big{|}\big{)}.

It along with (3.95) implies (3.91).

Similarly, for τ0\tau\neq 0, τ00\tau_{0}\neq 0, and 2jl042\leq j\leq l_{0}-4, where sgnsgn are constants, one may compute

cj(μ1y0(x2)y0(x2))\displaystyle\partial_{c}^{j}\begin{pmatrix}\mu^{-1}y_{0-}(x_{2})\\ y_{0-}^{\prime}(x_{2})\end{pmatrix}\leq Cμ1j(j=0j|(τ+τ0)j((ττ0τ0)log|ττ0|)|+|ττ0|).\displaystyle C\mu^{1-j}\Big{(}\sum_{j^{\prime}=0}^{j}\left|(\partial_{\tau}+\partial_{\tau_{0}})^{j^{\prime}}\Big{(}\begin{pmatrix}\tau\tau_{0}\\ \tau_{0}\end{pmatrix}\log|\tfrac{\tau}{\tau_{0}}|\Big{)}\right|+|\tau-\tau_{0}|\Big{)}.

For j3j\geq 3, we have

|(τ+τ0)j(ττ0log|ττ0|)C(|τ2jτ02j|+|τ+τ0||τ1jτ01j|+|ττ0||τjτ0j|)\displaystyle|(\partial_{\tau}+\partial_{\tau_{0}})^{j}\big{(}\tau\tau_{0}\log|\tfrac{\tau}{\tau_{0}}|\big{)}\leq C(|\tau^{2-j}-\tau_{0}^{2-j}|+|\tau+\tau_{0}||\tau^{1-j}-\tau_{0}^{1-j}|+|\tau\tau_{0}||\tau^{-j}-\tau_{0}^{-j}|)
\displaystyle\leq C|ττ0||τ1τ01|(|τ|1j+|τ|2j|τ0|1++|τ0|1j)C|ττ0|(|τ|1j+|τ0|1j).\displaystyle C|\tau\tau_{0}||\tau^{-1}-\tau_{0}^{-1}|(|\tau|^{1-j}+|\tau|^{2-j}|\tau_{0}|^{-1}+\ldots+|\tau_{0}|^{1-j})\leq C|\tau-\tau_{0}|(|\tau|^{1-j}+|\tau_{0}|^{1-j}).

If j=2j=2, the first term on the right side of the first inequality would be log|ττ0|\log|\tfrac{\tau}{\tau_{0}}| which as shown previously also satisfies the above final estimate. Similarly, one can also calculate, for τ0\tau\neq 0, τ00\tau_{0}\neq 0, and j2j\geq 2,

|(τ+τ0)j(τ0log|ττ0|)C(|τ1jτ01j|+|τ0||τjτ0j|)\displaystyle|(\partial_{\tau}+\partial_{\tau_{0}})^{j}\big{(}\tau_{0}\log|\tfrac{\tau}{\tau_{0}}|\big{)}\leq C(|\tau^{1-j}-\tau_{0}^{1-j}|+|\tau_{0}||\tau^{-j}-\tau_{0}^{-j}|)
\displaystyle\leq C|τ0||τ1τ01|(|τ|1j+|τ|2j|τ0|1++|τ0|1j)C|τ|1|ττ0|(|τ|1j+|τ0|1j).\displaystyle C|\tau_{0}||\tau^{-1}-\tau_{0}^{-1}|(|\tau|^{1-j}+|\tau|^{2-j}|\tau_{0}|^{-1}+\ldots+|\tau_{0}|^{1-j})\leq C|\tau|^{-1}|\tau-\tau_{0}|(|\tau|^{1-j}+|\tau_{0}|^{1-j}).

The cases of j=0,1j=0,1 have been considered earlier and would only make minor contributions. Therefore (3.92) and (3.93) are satisfied in case 2 as well.

Regarding y0(k,c,x2c)y_{0-}(k,c,x_{2}^{c}), much as in case 1, but with much simpler initial value at τ0=hx2cμ\tau_{0}=\tfrac{-h-x_{2}^{c}}{\mu}, we have

y0(k,c,x2c)=μB12(μ,c,τ0)y_{0-}(k,c,x_{2}^{c})=-\mu B_{12}(\mu,c,\tau_{0})

which also yields its Cl02C^{l_{0}-2} smoothness. Differentiating in cc and using Lemma 3.10 and 3.14, one may estimate, for 1jl021\leq j\leq l_{0}-2,

|cj(y0(k,c,x2c))|=\displaystyle\big{|}\partial_{c}^{j}\big{(}y_{0-}(k,c,x_{2}^{c})\big{)}\big{|}= μ|((cτμU(x2c))jB12)(μ,c,τ0)|Cμ1j,\displaystyle\mu\big{|}\big{(}(\partial_{c}-\tfrac{\partial_{\tau}}{\mu U^{\prime}(x_{2}^{c})})^{j}B_{12}\big{)}(\mu,c,\tau_{0})\big{|}\leq C\mu^{1-j},

which proves the inequality in (3.89) in case 2. Finally, from Lemma 3.10,

c(y0(k,c,x2c))|c=U(h)=μ((cτμU(h))B12)(μ,U(h),0)=U(h)1.\partial_{c}\big{(}y_{0-}(k,c,x_{2}^{c})\big{)}|_{c=U(-h)}=-\mu\big{(}(\partial_{c}-\tfrac{\partial_{\tau}}{\mu U^{\prime}(-h)})B_{12}\big{)}(\mu,U(-h),0)=U^{\prime}(-h)^{-1}.

Estimating cy0+\partial_{c}y_{0+} on 2\mathcal{I}_{2}. In this case 2\mathcal{I}_{2}\neq\emptyset implies |c|C|c|\leq C. Much as in the above argument for y0y_{0-}, we consider the estimates related to cy0+\partial_{c}y_{0+} at some c[12h0h,12h0]c_{*}\in[-\frac{1}{2}h_{0}-h,\frac{1}{2}h_{0}]. Observe that, as an expression of solution to the homogeneous Rayleigh equation, (3.94) also applies to y0+y_{0+} on 2\mathcal{I}_{2} with x20x_{20} chosen near x2r(c)x_{2r}(c_{*}). In the case of x2r(c)μx_{2r}(c_{*})\leq-\mu, the same arguments yields the desired estimates of cy+\partial_{c}y_{+}.

In the case of x2r(c)[μ,0]x_{2r}(c_{*})\in[-\mu,0], we take x20=0x_{20}=0 and proceed roughly as in the above case 2. Due to the initial condition (3.53), equation (3.97) is replaced by

c(μ1y0+(x2)y0+(x2))=(cτ+τ0μU(x2c))(S(τ,τ0)(μ1y0+(0)y0+(0)))\displaystyle\partial_{c}\begin{pmatrix}\mu^{-1}y_{0+}(x_{2})\\ y_{0+}^{\prime}(x_{2})\end{pmatrix}=\big{(}\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})}\big{)}\left(S(\tau,\tau_{0})\begin{pmatrix}\mu^{-1}y_{0+}(0)\\ y_{0+}^{\prime}(0)\end{pmatrix}\right)
=\displaystyle= (cτ+τ0μU(x2c))(μU(x2c)U(x2c)(log|ττ0|+iπ2(sgn(τ)sgn(τ0)))S~(τ,τ0)(μ1y0+(0)y0+(0)))+O(|ττ0|),\displaystyle\big{(}\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})}\big{)}\left(\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}(sgn(\tau)-sgn(\tau_{0}))\big{)}\tilde{S}(\tau,\tau_{0})\begin{pmatrix}\mu^{-1}y_{0+}(0)\\ y_{0+}^{\prime}(0)\end{pmatrix}\right)+O(|\tau-\tau_{0}|),

where (3.78) and Lemma 3.11 are used. Let

W(τ,τ0)=(W1,W2)T=S~(τ,τ0)(μ1y0+(0),y0+(0))T.W(\tau,\tau_{0})=(W_{1},W_{2})^{T}=\tilde{S}(\tau,\tau_{0})\big{(}\mu^{-1}y_{0+}(0),y_{0+}^{\prime}(0)\big{)}^{T}.

Recall from initial condition (3.53)

|y0+(0)|Cμ4τ02,|y0+(0)1|Cμ3|τ0|.|y_{0+}(0)|\leq C\mu^{4}\tau_{0}^{2},\quad|y_{0+}^{\prime}(0)-1|\leq C\mu^{3}|\tau_{0}|.

On the one hand, from (3.77), and Lemma 3.10, we have that

(W1ττ0,W2τ0)=(B12(τ)τ,B22(μ))(B22(τ0)y0+(0)μτ0B12(τ0)τ0y0(0))\Big{(}\frac{W_{1}}{\tau\tau_{0}},\frac{W_{2}}{\tau_{0}}\Big{)}=\Big{(}\frac{B_{12}(\tau)}{\tau},B_{22}(\mu)\Big{)}\Big{(}B_{22}(\tau_{0})\frac{y_{0+}(0)}{\mu\tau_{0}}-\frac{B_{12}(\tau_{0})}{\tau_{0}}y_{0-}^{\prime}(0)\Big{)}

are Cl03C^{l_{0}-3} function with bounds uniform in cc and μ\mu. Hence the estimate on cy0+(x2)\partial_{c}y_{0+}(x_{2}) is obtained much as that of cy0(x2)\partial_{c}y_{0-}(x_{2}). On the other hand, as (3.96) and (3.95) also apply to y0+y_{0+}, it holds

W2=B22(τ)μ1y0+(x2c)=(1+O(|τ|))μ1y0+(x2c).W_{2}=B_{22}(\tau)\mu^{-1}y_{0+}(x_{2}^{c})=(1+O(|\tau|))\mu^{-1}y_{0+}(x_{2}^{c}).

With these estimates, the desired estimate on cy0+(x2)\partial_{c}y_{0+}^{\prime}(x_{2}) follows much as that of cy0(x2)\partial_{c}y_{0-}^{\prime}(x_{2}). This completes the proof. ∎

Lemma 3.16.

Assume l03l_{0}\geq 3, 2\mathcal{I}_{2}\neq\emptyset, and 3\mathcal{I}_{3}\neq\emptyset, then Lemma 3.13a.)–e.) hold for x23x_{2}\in\mathcal{I}_{3}. Moreover, if l05l_{0}\geq 5, then there exists C>0C>0 depending only on |U|Cl01|U^{\prime}|_{C^{l_{0}-1}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any kk\in\mathbb{R} and any cc\in\mathbb{R}, the following estimates hold for x23x_{2}\in\mathcal{I}_{3}

(3.98) μ1|cy0(x2)|+|cy0(x2)|C(1+logμmin{μ,|U(h)c|})cosh(μ1(x2+h)),\mu^{-1}|\partial_{c}y_{0-}(x_{2})|+|\partial_{c}y_{0-}^{\prime}(x_{2})|\leq C\Big{(}1+\log\frac{\mu}{\min\{\mu,|U(-h)-c|\}}\Big{)}\cosh(\mu^{-1}(x_{2}+h)),

and for 2jl042\leq j\leq l_{0}-4,

μ1|cjy0(x2)|+|cjy0(x2)|C(μ1j+|U(h)c|1j)cosh(μ1(x2+h)).\mu^{-1}|\partial_{c}^{j}y_{0-}(x_{2})|+|\partial_{c}^{j}y_{0-}^{\prime}(x_{2})|\leq C\big{(}\mu^{1-j}+|U(-h)-c|^{1-j}\big{)}\cosh(\mu^{-1}(x_{2}+h)).

Moreover, if 2\mathcal{I}_{2}\neq\emptyset and 1\mathcal{I}_{1}\neq\emptyset, then it also holds for x21x_{2}\in\mathcal{I}_{1},

μ1|cy0+(x2)|+|cy0+(x2)|C(1+logμmin{μ,|U(0)c|})cosh(μ1x2).\mu^{-1}|\partial_{c}y_{0+}(x_{2})|+|\partial_{c}y_{0+}^{\prime}(x_{2})|\leq C\Big{(}1+\log\frac{\mu}{\min\{\mu,|U(0)-c|\}}\Big{)}\cosh(\mu^{-1}x_{2}).
Remark 3.9.

Using Remark 3.8, the above estimates with an an additional μ1\mu^{-1} on the right sides also hold for l04l_{0}\geq 4 and jl03j\leq l_{0}-3.

Proof.

The assumption 2\mathcal{I}_{2}\neq\emptyset and 3=[x2r,0]\mathcal{I}_{3}=[x_{2r},0]\neq\emptyset imply x2r>hx_{2r}>-h and μ1(U(x2r)c)\mu^{-1}\big{(}U(x_{2r})-c\big{)} is uniformly bounded from above and below away from 0. The regularity of y0y_{0-} and y0y_{0-}^{\prime} in cc and kk for x23x_{2}\in\mathcal{I}_{3} follow directly from such smoothness at x2rx_{2r} obtained in Lemma 3.15. Their estimates at x2rx_{2r} can be summarized into

μ1|cy0(x2r)|+|cy0(x2r)|C(1+logμmin{μ,|U(h)c|})cosh(μ1(x2r+h)),\mu^{-1}|\partial_{c}y_{0-}(x_{2r})|+|\partial_{c}y_{0-}^{\prime}(x_{2r})|\leq C\Big{(}1+\log\frac{\mu}{\min\{\mu,|U(-h)-c|\}}\Big{)}\cosh(\mu^{-1}(x_{2r}+h)),

and for 2jl042\leq j\leq l_{0}-4

μ1|cjy0(x2r)|+|cjy0(x2r)|C(μ1j+|U(h)c|1j)cosh(μ1(x2r+h))\mu^{-1}|\partial_{c}^{j}y_{0-}(x_{2r})|+|\partial_{c}^{j}y_{0-}^{\prime}(x_{2r})|\leq C\big{(}\mu^{1-j}+|U(-h)-c|^{1-j}\big{)}\cosh(\mu^{-1}(x_{2r}+h))

where we also used 1μ1(x2r+h)1\leq\mu^{-1}(x_{2r}+h) as x2l>hx_{2l}>-h. Much as the proof of Lemma 3.14, we shall obtain the estimates inductively in jj by considering the cases of small and large kk separately.

As ρ0>0\rho_{0}>0, we take k1k_{*}\geq 1 such that ρ<1\rho<1 (defined in (3.87)) for |k|k|k|\geq k_{*} and thus (3.8) is satisfied on 3\mathcal{I}_{3} with ρ<min{1,Cμ}\rho<\min\{1,C\mu\}. We shall obtain the estimates for this case of |k|k|k|\geq k_{*} by splitting cy0\partial_{c}y_{0-} into homogeneous and non-homogeneous parts. For j1j\geq 1, let y1(x2)y_{1}(x_{2}) be the solution to the homogeneous Rayleigh equation (3.1) with initial condition

y1(x2r)=cjy0(x2r),y1(x2r)=cjy0(x2r),y_{1}(x_{2r})=\partial_{c}^{j}y_{0-}(x_{2r}),\;\;y_{1}^{\prime}(x_{2r})=\partial_{c}^{j}y_{0-}^{\prime}(x_{2r}),

and y2(x2)y_{2}(x_{2}) be the solution to the non-homogeneous Rayleigh equation (3.3) with the zero initial conditions at x2=x2rx_{2}=x_{2r} and the non-homogeneous term given by the right side of (3.86) (with j1=0j_{1}=0 and j2=jj_{2}=j). Clearly it holds

(3.99) cjy0=y1+y2, on 3.\partial_{c}^{j}y_{0-}=y_{1}+y_{2},\;\text{ on }\;\mathcal{I}_{3}.

Using the the above estimates on cy0\partial_{c}y_{0-} at x2rx_{2r}, we apply Lemma 3.2 to y1y_{1} with

Θ1=Θ2=cosh,s=0,C0=μ1|cjy0(x2r)|+|cjy0(x2r)|+1\Theta_{1}=\Theta_{2}=\cosh,\quad s=0,\quad C_{0}=\mu^{-1}|\partial_{c}^{j}y_{0-}(x_{2r})|+|\partial_{c}^{j}y_{0-}^{\prime}(x_{2r})|+1

to obtain, for x23x_{2}\in\mathcal{I}_{3},

μ1|y1(x2)|+|y1(x2)|C(μ1|cjy0(x2r)|+|cjy0(x2r)|+1)coshμ1(x2x2r).\displaystyle\mu^{-1}|y_{1}(x_{2})|+|y_{1}^{\prime}(x_{2})|\leq C\big{(}\mu^{-1}|\partial_{c}^{j}y_{0-}(x_{2r})|+|\partial_{c}^{j}y_{0-}^{\prime}(x_{2r})|+1\big{)}\cosh\mu^{-1}(x_{2}-x_{2r}).

Concerning y2(x2)y_{2}(x_{2}), Lemmas 3.2 and the same computation as in the proof of Lemma 3.14 implies, for any x23x_{2}\in\mathcal{I}_{3},

|μ1y2(x2)|+|y2(x2)|Cj=0j1x2rx2cosh(μ1(x2x2))|U(x2)c|j+1j|cjy0(x2)|dx2.|\mu^{-1}y_{2}(x_{2})|+|y_{2}^{\prime}(x_{2})|\leq C\sum_{j^{\prime}=0}^{j-1}\int_{x_{2r}}^{x_{2}}\frac{\cosh(\mu^{-1}(x_{2}-x_{2}^{\prime}))}{|U(x_{2}^{\prime})-c|^{j+1-j^{\prime}}}|\partial_{c}^{j^{\prime}}y_{0-}(x_{2}^{\prime})|dx_{2}^{\prime}.

The desired estimate for j=1j=1 follows from (3.99), Lemma 3.9, and direct integration. For j2j\geq 2, one may compute inductively using the above estimates and (3.94),

|μ1y2(x2)|+|y2(x2)|Cj=0j1(U(x2r)c)jj|cjy0(x2r)|coshμ1(x2r+h)coshμ1(x2+h)\displaystyle|\mu^{-1}y_{2}(x_{2})|+|y_{2}^{\prime}(x_{2})|\leq C\sum_{j^{\prime}=0}^{j-1}(U(x_{2r})-c)^{j^{\prime}-j}\frac{|\partial_{c}^{j^{\prime}}y_{0-}(x_{2r})|}{\cosh\mu^{-1}(x_{2r}+h)}\cosh\mu^{-1}(x_{2}+h)
\displaystyle\leq C(μ1j+μ1|U(h)c|2j+μ1jlogμmin{μ,|U(h)c|})coshμ1(x2+h).\displaystyle C\Big{(}\mu^{1-j}+\mu^{-1}|U(-h)-c|^{2-j}+\mu^{1-j}\log\frac{\mu}{\min\{\mu,|U(-h)-c|\}}\Big{)}\cosh\mu^{-1}(x_{2}+h).

If |U(h)c|μ|U(-h)-c|\geq\mu, the desired estimate follows immediately, otherwise it follows from the fact logxx\log x\leq x for any x1x\geq 1.

In the case kkk\leq k_{*}, μ1\mu\sim 1 and Lemma 3.3 yields the estimates through a similar induction.

The estimates on cy0+\partial_{c}y_{0+} is also obtained much as cy0\partial_{c}y_{0-} using Lemmas 3.2 and 3.3 based on the estimates of cy0+\partial_{c}y_{0+} at x2lx_{2l} obtained in Lemma 3.15. In particular, the fact that 2\mathcal{I}_{2}\neq\emptyset also implies |c|<C|c|<C is also used. We skip the details. ∎

The following lemma proves Lemma 3.13(f) and (the case of x2=0x_{2}=0) will be used in analyzing the eigenvalues.

Lemma 3.17.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3. For any kk\in\mathbb{R} and x2[h,0]x_{2}\in[-h,0], there exist R,C~>0R,\tilde{C}>0 such that

(3.100) |U(h)c|j|kj1cj2x2ly0(k,c,x2)|{C~(1+|log|U(h)c||),j+1=j2{0,1},C~,j+1=j22,|U(-h)-c|^{j}|\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}\partial_{x_{2}}^{l}y_{0-}(k,c,x_{2})|\leq\begin{cases}\tilde{C}\big{(}1+\big{|}\log|U(-h)-c|\big{|}\big{)},&j+1=j_{2}\in\{0,1\},\\ \tilde{C},&j+1=j_{2}\geq 2,\end{cases}

for any |cU(h)|R|c-U(-h)|\leq R, l=0,1l=0,1, j1,j20j_{1},j_{2}\geq 0, j1+j2l03j_{1}+j_{2}\leq l_{0}-3. Here C~\tilde{C} can be taken independent of kk for kk in any bounded set.

Unlike in most other lemmas, the constants RR and C~\tilde{C} may depend on kk and x2x_{2}.

Proof.

The lemma is trivial if x2=hx_{2}=-h, so we assume x2>hx_{2}>-h. Since the lemma is concerned with cc close to U(h)U(-h) where RR and C~\tilde{C} may depend on x2x_{2} and kk, we consider

c=U(x2c),x2c[h0h,(h+x2)/2]τ=μ1(x2x2c)μ1(x2+h)/2>0.c=U(x_{2}^{c}),\;\;x_{2}^{c}\in[-h_{0}-h,(-h+x_{2})/2]\,\Longrightarrow\,\tau=\mu^{-1}(x_{2}-x_{2}^{c})\geq\mu^{-1}(x_{2}+h)/2>0.

If x2x2c>μx_{2}-x_{2}^{c}>\mu, then (3.100) clearly holds as x2x_{2} is away from the singularity of y0y_{0-}. Otherwise, let τ0=μ1(x2c+h)\tau_{0}=-\mu^{-1}(x_{2}^{c}+h). From (3.78) and the Cl02C^{l_{0}-2} smoothness of SerrS_{err} due to Lemma 3.11, we have

τ0jkj1cj2(μ1y0(x2)y0(x2))\displaystyle\tau_{0}^{j}\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}\begin{pmatrix}\mu^{-1}y_{0-}(x_{2})\\ y_{0-}^{\prime}(x_{2})\end{pmatrix}
=\displaystyle= τ0j(cτ+τ0μU(x2c))j2[(log|ττ0|+iπ2(sgn(τ)sgn(τ0)))kj1(μU(x2c)U(x2c)(S~12(μ,c,τ,τ0)S~22(μ,c,τ,τ0)))]+O(1)\displaystyle\tau_{0}^{j}\big{(}\partial_{c}-\tfrac{\partial_{\tau}+\partial_{\tau_{0}}}{\mu U^{\prime}(x_{2}^{c})}\big{)}^{j_{2}}\left[\big{(}\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}(sgn(\tau)-sgn(\tau_{0}))\big{)}\partial_{k}^{j_{1}}\Big{(}\tfrac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\begin{pmatrix}\tilde{S}_{12}(\mu,c,\tau,\tau_{0})\\ \tilde{S}_{22}(\mu,c,\tau,\tau_{0})\end{pmatrix}\Big{)}\right]+O(1)

where we also used that log|ττ0|+iπ2(sgn(τ)sgn(τ0))\log|\tfrac{\tau}{\tau_{0}}|+\tfrac{i\pi}{2}(sgn(\tau)-sgn(\tau_{0})) is independent of kk. The desired inequality (3.100) follows from straight forward calculations using the Cl03C^{l_{0}-3} smoothness of S~12τ0\tfrac{\tilde{S}_{12}}{\tau_{0}}, S~22τ0\tfrac{\tilde{S}_{22}}{\tau_{0}}, and U(h)cτ0\tfrac{U(-h)-c}{\tau_{0}}. ∎

Remark 3.10.

Most of the above regularity results and estimates also hold for y0+(k,c,x2)y_{0+}(k,c,x_{2}). Since y+y_{+} plays a less substantial role as yy_{-} in the rest of the paper, we only gave the basic estimates on y+y_{+}.

In the above cy±\partial_{c}y_{\pm} was considered only for cU([h02h,h02])c\in U([-\frac{h_{0}}{2}-h,\frac{h_{0}}{2}]). To end this subsection, we extend some estimate for cc\in\mathbb{C} using the analyticity of y±y_{\pm} in cc in the following lemma.

Lemma 3.18.

Assume UC5U\in C^{5}. The following hold.

  1. (1)

    For any cc\in\mathbb{C} with cI>0c_{I}>0, it holds

    cy(k,c,x2)=12πicy0(k,c,x2)ccdc,\partial_{c}y_{-}(k,c,x_{2})=\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{\partial_{c}y_{0-}(k,c^{\prime},x_{2})}{c^{\prime}-c}dc^{\prime},
    (U(x2)c)cy(k,c,x2)=12πi(U(x2)c)cy0(k,c,x2)ccdc.(U(x_{2})-c)\partial_{c}y_{-}^{\prime}(k,c,x_{2})=\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{(U(x_{2})-c^{\prime})\partial_{c}y_{0-}^{\prime}(k,c^{\prime},x_{2})}{c^{\prime}-c}dc^{\prime}.
    cy+(k,c,x2)(U(x2)ci)2=12πicy0+(k,c,x2)(U(x2)ci)2(cc)dc,\frac{\partial_{c}y_{+}(k,c,x_{2})}{(U(x_{2})-c-i)^{2}}=\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{\partial_{c}y_{0+}(k,c^{\prime},x_{2})}{(U(x_{2})-c^{\prime}-i)^{2}(c^{\prime}-c)}dc^{\prime},
    (U(x2)c)cy+(k,c,x2)(U(x2)ci)3=12πi(U(x2)c)cy0+(k,c,x2)(U(x2)ci)3(cc)dc.\frac{(U(x_{2})-c)\partial_{c}y_{+}^{\prime}(k,c,x_{2})}{(U(x_{2})-c-i)^{3}}=\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{(U(x_{2})-c^{\prime})\partial_{c}y_{0+}(k,c^{\prime},x_{2})}{(U(x_{2})-c^{\prime}-i)^{3}(c^{\prime}-c)}dc^{\prime}.
  2. (2)

    For any r(1,)r\in(1,\infty),

    1. (a)

      there exists C>0C>0 depending only on rr, |U|C4|U^{\prime}|_{C^{4}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any kk\in\mathbb{R}, x2[h,0]x_{2}\in[-h,0], cI>0c_{I}>0,

      μ1|cy(k,c,x2)|LcRr()+|(U(x2)c)cy(k,c,x2)|LcRr()Ccoshμ1(x2+h);\mu^{-1}|\partial_{c}y_{-}(k,c,x_{2})|_{L_{c_{R}}^{r}(\mathbb{R})}+|(U(x_{2})-c)\partial_{c}y_{-}^{\prime}(k,c,x_{2})|_{L_{c_{R}}^{r}(\mathbb{R})}\leq C\cosh\mu^{-1}(x_{2}+h);
    2. (b)

      as cI0+c_{I}\to 0+, cy\partial_{c}y_{-} and (Uc)cy(U-c)\partial_{c}y_{-}^{\prime} converge to cy0\partial_{c}y_{0-} and (UcR)cy0(U-c_{R})\partial_{c}y_{0-}^{\prime} in LcRr()L_{c_{R}}^{r}(\mathbb{R}), respectively, for any x2[h,0]x_{2}\in[-h,0]. Moreover, the convergence also holds in LcR,x2r(×[h,0])L_{c_{R},x_{2}}^{r}(\mathbb{R}\times[-h,0]).

  3. (3)

    For any r(1,)r\in(1,\infty) and compact interval \mathcal{I}\subset\mathbb{R},

    1. (a)

      there exists C>0C>0 depending only on rr, \mathcal{I}, |U|C4|U^{\prime}|_{C^{4}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any kk\in\mathbb{R}, x2[h,0]x_{2}\in[-h,0], cI>0c_{I}>0,

      μ1|cy+(k,c,x2)|LcRr()+|(U(x2)c)cy+(k,c,x2)|LcRr()Ccoshμ1(x2+h);\mu^{-1}|\partial_{c}y_{+}(k,c,x_{2})|_{L_{c_{R}}^{r}(\mathcal{I})}+|(U(x_{2})-c)\partial_{c}y_{+}^{\prime}(k,c,x_{2})|_{L_{c_{R}}^{r}(\mathcal{I})}\leq C\cosh\mu^{-1}(x_{2}+h);
    2. (b)

      as cI0+c_{I}\to 0+, cy+\partial_{c}y_{+} and (Uc)cy+(U-c)\partial_{c}y_{+}^{\prime} converge to cy0+\partial_{c}y_{0+} and (UcR)cy0+(U-c_{R})\partial_{c}y_{0+}^{\prime} in LcRr()L_{c_{R}}^{r}(\mathcal{I}), respectively, for any x2[h,0]x_{2}\in[-h,0]. Moreover, the convergence also holds in LcR,x2r(×[h,0])L_{c_{R},x_{2}}^{r}(\mathcal{I}\times[-h,0]).

The multiplier U(x2)cU(x_{2})-c in front of cy±\partial_{c}y_{\pm}^{\prime} is added to regularize their singularities near U(x2)=cU(x_{2})=c and the denominators (U(x2)ci)n(U(x_{2})-c-i)^{n}, n=2,3n=2,3, in the expressions related to cy+\partial_{c}y_{+} are to make it decay as |c||c|\to\infty (recall the initial conditions (3.53) of y+y_{+} involving cc).

Proof.

Let Bh0B_{h_{0}}\subset\mathbb{C} be the open disk with diameter segment U([h02h,h02])U([-\frac{h_{0}}{2}-h,\frac{h_{0}}{2}]). For any cBh0c\notin B_{h_{0}}, let

ρ=k2(1+|U|C0)max[h,0]|Uc|1Ck2(1+|c|)1.\rho=k^{-2}(1+|U^{\prime\prime}|_{C^{0}})\max_{[-h,0]}|U-c|^{-1}\leq Ck^{-2}(1+|c|)^{-1}.

There exists k>0k_{*}>0 such that ρ<1\rho<1 for any |k|k|k|\geq k_{*}. Lemma 3.2 (with x2l=hx_{2l}=-h, =[h,0]\mathcal{I}=[-h,0], C0=0C_{0}=0, and Θ1=Θ2=sinh\Theta_{1}=\Theta_{2}=\sinh) implies, for |k|k|k|\geq k_{*} and cBh0c\notin B_{h_{0}},

(3.101) |y(k,c,x2)|k|1sinh|k|(x2+h)|+μ|y(k,c,x2)cosh|k|(x2+h)|Cμk1(1+|c|)1sinh|k|(x2+h).\begin{split}\big{|}y_{-}(k,c,x_{2})-|k|^{-1}\sinh|k|(x_{2}+h)\big{|}+\mu\big{|}y_{-}^{\prime}(k,c,&x_{2})-\cosh|k|(x_{2}+h)\big{|}\\ \leq&C\mu k^{-1}(1+|c|)^{-1}\sinh|k|(x_{2}+h).\end{split}

For |k|<k|k|<k_{*}, Lemma 3.3 implies that the above inequality still holds for cBh0c\notin B_{h_{0}}.

From equation (3.86) (j1=0j_{1}=0 and j2=1j_{2}=1) of cy\partial_{c}y_{-}, applying (3.19) with ϕ=U(uc)2y\phi=-\frac{U^{\prime\prime}}{(u-c)^{2}}y_{-} and using Lemma 3.9, we have for |k|k|k|\geq k_{*} and cBh0c\notin B_{h_{0}},

(3.102) μ1|cy(k,c,x2)|+|cy(k,c,x2)|\displaystyle\mu^{-1}|\partial_{c}y_{-}(k,c,x_{2})|+|\partial_{c}y_{-}^{\prime}(k,c,x_{2})|\leq Cμ(1+|c|)2sinhμ1(x2+h).\displaystyle C\mu(1+|c|)^{-2}\sinh\mu^{-1}(x_{2}+h).

For |k|k|k|\leq k_{*}, Lemma 3.3 implies that the above inequality still holds for cBh0c\notin B_{h_{0}}.

For any cc\in\mathbb{C} with cI>0c_{I}>0, the analyticity of cy\partial_{c}y_{-} and its O(|c|2)O(|c|^{-2}) decay as |c||c|\to\infty imply, for any β(0,cI)\beta\in(0,c_{I}),

cy(k,c,x2)=12πi+iβcy(k,c,x2)ccdc=12πi+iβy(k,c,x2)(cc)2dc,\partial_{c}y_{-}(k,c,x_{2})=\frac{1}{2\pi i}\int_{\mathbb{R}+i\beta}\frac{\partial_{c}y_{-}(k,c^{\prime},x_{2})}{c^{\prime}-c}dc^{\prime}=\frac{1}{2\pi i}\int_{\mathbb{R}+i\beta}\frac{y_{-}(k,c^{\prime},x_{2})}{(c^{\prime}-c)^{2}}dc^{\prime},

where the boundary terms at infinity in the above integration by parts vanish due to the uniform-in-cc bound on |y||y_{-}| given in (3.101). Letting β0+\beta\to 0+, the same bound and Lemma 3.12 yield

cy(k,c,x2)=12πiy0(k,c,x2)(cc)2dc=12πicy0(k,c,x2)ccdc=12πicy0(k,c,x2)(cRc)+icIdc,\partial_{c}y_{-}(k,c,x_{2})=\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{y_{0-}(k,c^{\prime},x_{2})}{(c^{\prime}-c)^{2}}dc^{\prime}=\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{\partial_{c}y_{0-}(k,c^{\prime},x_{2})}{c^{\prime}-c}dc^{\prime}=-\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{\partial_{c}y_{0-}(k,c^{\prime},x_{2})}{(c_{R}-c^{\prime})+ic_{I}}dc^{\prime},

where we integrated by parts again. The desired estimate on |cy|LcRr|\partial_{c}y_{-}|_{L_{c_{R}}^{r}} follows from the boundedness of the convolution kernel 1c+icI\frac{1}{c^{\prime}+ic_{I}} on Lr()L^{r}(\mathbb{R}), (3.102) for |c|1|c|\gg 1, and Lemmas 3.143.16.

The results for cy\partial_{c}y_{-}^{\prime} are derived in the same manner. In fact

(U(x2)c)cy(k,c,x2)=12πi+iβ(U(x2)c)cy(k,c,x2)ccdc\displaystyle(U(x_{2})-c)\partial_{c}y_{-}^{\prime}(k,c,x_{2})=\frac{1}{2\pi i}\int_{\mathbb{R}+i\beta}\frac{(U(x_{2})-c^{\prime})\partial_{c}y_{-}^{\prime}(k,c^{\prime},x_{2})}{c^{\prime}-c}dc^{\prime}
=\displaystyle= 12πi+iβ(U(x2)c)cy(k,c,x2)cccy(k,c,x2)dc\displaystyle\frac{1}{2\pi i}\int_{\mathbb{R}+i\beta}\frac{(U(x_{2})-c)\partial_{c}y_{-}^{\prime}(k,c^{\prime},x_{2})}{c^{\prime}-c}-\partial_{c}y_{-}^{\prime}(k,c^{\prime},x_{2})dc^{\prime}
=\displaystyle= 12πi+iβ(U(x2)c)y(k,c,x2)(cc)2dc=12πi(U(x2)c)y0(k,c,x2)(cc)2dc\displaystyle\frac{1}{2\pi i}\int_{\mathbb{R}+i\beta}\frac{(U(x_{2})-c)y_{-}^{\prime}(k,c^{\prime},x_{2})}{(c^{\prime}-c)^{2}}dc^{\prime}=\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{(U(x_{2})-c)y_{0-}^{\prime}(k,c^{\prime},x_{2})}{(c^{\prime}-c)^{2}}dc^{\prime}
=\displaystyle= 12πi(U(x2)c)cy0(k,c,x2)ccdc,\displaystyle\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{(U(x_{2})-c^{\prime})\partial_{c}y_{0-}^{\prime}(k,c^{\prime},x_{2})}{c^{\prime}-c}dc^{\prime},

where we used (3.101) to cancel the two boundary terms at infinity in the above both integrations by parts and also used the integrability of (U(x2)c)cy(k,c,x2)(U(x_{2})-c)\partial_{c}y_{-}^{\prime}(k,c,x_{2}) near U(x2)=cU(x_{2})=c given in Lemma 3.15. The latter also yields the estimate on (Uc)cy(U-c)\partial_{c}y_{-}^{\prime}.

In statement (2b), the pointwise-in-x2x_{2} convergence in LcRrL_{c_{R}}^{r} is standard due to the convergence of the convolution kernel 1c+icI\frac{1}{c^{\prime}+ic_{I}} on Lr()L^{r}(\mathbb{R}) as cI0+c_{I}\to 0+, as well as the analyticity of yy_{-} for cI>0c_{I}>0. The convergence in LcR,x2rL_{c_{R},x_{2}}^{r} follows from the pointwise-in-x2x_{2} convergence in LcRrL_{c_{R}}^{r}, the Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r} bounds in statement (2a), and the dominant convergence theorem.

Finally, cy+\partial_{c}y_{+} can be analyzed similar. However, the initial values (3.53) induce an O(|c|2)O(|c|^{2}) growth in y+y_{+} and y+y_{+}^{\prime} and an O(|c|)O(|c|) growth of cy+\partial_{c}y_{+} and cy+\partial_{c}y_{+}^{\prime} for |c|1|c|\gg 1 (Lemma 3.3). Instead we consider, for cI>0c_{I}>0,

cy+(k,c,x2)(U(x2)ci)2=12πi+iβcy+(k,c,x2)(U(x2)ci)2(cc)dc,\frac{\partial_{c}y_{+}(k,c,x_{2})}{(U(x_{2})-c-i)^{2}}=\frac{1}{2\pi i}\int_{\mathbb{R}+i\beta}\frac{\partial_{c}y_{+}(k,c^{\prime},x_{2})}{(U(x_{2})-c^{\prime}-i)^{2}(c^{\prime}-c)}dc^{\prime},

which holds for any β(0,cI)\beta\in(0,c_{I}). From this Cauchy integral formula we proceed much as in the above and obtain the integral representation in term of cy0+\partial_{c}y_{0+}. The derivation of the corresponding formula of cy+\partial_{c}y_{+}^{\prime} is also similar. The desired convergence and estimates of cy+\partial_{c}y_{+} and cy+\partial_{c}y_{+}^{\prime} in LcRr()L_{c_{R}}^{r}(\mathcal{I}) on a compact interval \mathcal{I} again follow from the properties of the convolution by the kernel 1c+icI\frac{1}{c^{\prime}+ic_{I}}. ∎

3.6. An important quantity Y=y(0)/y(0)Y=y_{-}^{\prime}(0)/y_{-}(0)

To end this section, we analyze a quantity related to the Reynolds stress, which is crucial for the linearized water wave problem:

(3.103) Y(k,c)=YR(k,c)+iYI(k,c):=y(k,c,0)y(k,c,0),c=cR+icIU([h,0]),Y(k,c)=limϵ0+Y(k,c+iϵ)=y0(k,c,0)y0(k,c,0),cU([h,0)),\begin{split}&Y(k,c)=Y_{R}(k,c)+iY_{I}(k,c):=\frac{y_{-}^{\prime}(k,c,0)}{y_{-}(k,c,0)},\quad c=c_{R}+ic_{I}\in\mathbb{C}\setminus U([-h,0]),\\ &Y(k,c)=\lim_{\epsilon\to 0+}Y(k,c+i\epsilon)=\frac{y_{0-}^{\prime}(k,c,0)}{y_{0-}(k,c,0)},\quad c\in U\big{(}[-h,0)\big{)},\end{split}

where y(k,c,x2)y_{-}(k,c,x_{2}) is the solution to homogeneous Rayleigh equation (3.1) satisfying y(h)=0y_{-}(-h)=0 and y(h)=1y_{-}^{\prime}(-h)=1 defined in Subsection 3.3 and y0(k,c,x2)=limϵ0+y(k,c+iϵ,x2)y_{0-}(k,c,x_{2})=\lim_{\epsilon\to 0+}y_{-}(k,c+i\epsilon,x_{2}) for cc\in\mathbb{R}. Due to Remark 3.4, y(k,cR+icI,x2)y_{-}(k,c_{R}+ic_{I},x_{2}) satisfies estimates uniform in 0<ϵ10<\epsilon\ll 1. With slight abuse of notations, we would not distinguish y0y_{0-} from yy_{-} in the rest of this section. Apparently the domain of Y(k,c)Y(k,c) is given by

D(Y)={(k,c)×cU(0),y(k,c,0)0},D(Y)=\{(k,c)\in\mathbb{R}\times\mathbb{C}\mid c\neq U(0),\ y_{-}(k,c,0)\neq 0\},

and those excluded points (except c=U(0)c=U(0)) exactly are the eigenvalues of of the linearized Euler equation in the fixed channel x2[h,0]x_{2}\in[-h,0] at the shear flow (U(x2),0)(U(x_{2}),0). YY is not defined at c=U(0)c=U(0) since y(x2)y_{-}^{\prime}(x_{2}) has singularity at x2=0x_{2}=0. We first summarize some basic or standard properties of y(k,c,0)y_{-}(k,c,0) in the following lemma.

Lemma 3.19.

Assume UC3U\in C^{3}. The following hold.

  1. (1)

    For any kk\in\mathbb{R}, y(k,c,x2)>0y_{-}(k,c,x_{2})>0 for any x2(h,0]x_{2}\in(-h,0] and cU((h,0))c\in\mathbb{R}\setminus U\big{(}(-h,0)\big{)}.

  2. (2)

    There exists C>0C>0 depending on UU such that, for any k,ck,c\in\mathbb{R}, it holds

    (3.104) (Ck)1sinhk(x2+h)y(k,c,x2)Ck1sinhk(x2+h) if (U(x2)c)(U(h)c)0.(Ck)^{-1}\sinh k(x_{2}+h)\leq y_{-}(k,c,x_{2})\leq Ck^{-1}\sinh k(x_{2}+h)\;\text{ if }\;(U(x_{2})-c)(U(-h)-c)\geq 0.
  3. (3)

    There exists C>0C>0 depending only on UU such that, for any c=U(x2c)c=U(x_{2}^{c}), x2c[h,0)x_{2}^{c}\in[-h,0), it holds, for any kk\in\mathbb{R},

    C1μ2|U(x2c)|sinhμ1(x2c+h)sinhμ1|x2c||Imy(k,c,0)|Cμ2|U(x2c)|sinhμ1(x2c+h)sinhμ1|x2c|.C^{-1}\mu^{2}|U^{\prime\prime}(x_{2}^{c})|\sinh\mu^{-1}(x_{2}^{c}+h)\sinh\mu^{-1}|x_{2}^{c}|\leq|\text{Im}\,y_{-}(k,c,0)|\leq C\mu^{2}|U^{\prime\prime}(x_{2}^{c})|\sinh\mu^{-1}(x_{2}^{c}+h)\sinh\mu^{-1}|x_{2}^{c}|.
  4. (4)

    There exists k>0k_{*}>0 and C>0C>0 depending only on M>0M>0, |U|C2|U^{\prime}|_{C^{2}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} such that, if |k|k|k|\geq k_{*} or |c(U(h)+U(0))/2|M+(U(0)U(h))/2|c-(U(-h)+U(0))/2|\geq M+(U(0)-U(-h))/2 then

    (3.105) |y(k,c,0)|(Ck)1sinhkh.|y_{-}(k,c,0)|\geq(Ck)^{-1}\sinh kh.
  5. (5)

    Suppose a closed subset SS\subset\mathbb{C} satisfies y(k,c,0)0y_{-}(k,c,0)\neq 0 for all cSc\in S and k𝐊k\in\mathbf{K} where 𝐊=\mathbf{K}=\mathbb{R} or 2πL\frac{2\pi}{L}\mathbb{Z}, then there exists C>0C>0 depending only on SS and UU such that (3.105) holds for all k𝐊k\in\mathbf{K} and cSc\in S.

Remark 3.11.

According to Lemmas 3.3 and 3.9, the assumption y(k,c,0)0y_{-}(k,c,0)\neq 0 on SS in Statement (5) is automatically satisfied except possibly a compact set of (k,c)𝐊×S(k,c)\subset\mathbf{K}\times S. In particular, due to statement (3), it is satisfied for S=S=\mathbb{C} if U0U^{\prime\prime}\neq 0 on [h,0][-h,0]. We also recall y(k,c,0)=0y_{-}(k,c,0)=0 is equivalent to that ikc-ikc is an eigenvalue of the linearized Euler equation at the shear flow UU on the fixed channel x2(h,0)x_{2}\in(-h,0) associated with an eigenfunction v2(x2)=eikx1y(k,c,x2)v_{2}(x_{2})=e^{ikx_{1}}y_{-}(k,c,x_{2}).

Proof.

We first claim the following standard result.

Claim. Let y(x2)y(x_{2}) is a solution to the homogeneous Rayleigh equation (3.1) on an interval =(x2l,x2r)[h,0]\mathcal{I}=(x_{2l},x_{2r})\subset[-h,0] with cU()c\in\mathbb{R}\setminus U(\mathcal{I}) such that (y(x20),y(x20)){0}×({0})(y(x_{20}),y^{\prime}(x_{20}))\in\{0\}\times(\mathbb{R}\setminus\{0\}) at some x20¯x_{20}\in\overline{\mathcal{I}}, then y(x2){0}y(x_{2})\in\mathbb{R}\setminus\{0\} at any x2¯{x20}x_{2}\in\overline{\mathcal{I}}\setminus\{x_{20}\}.

If U(x20)cU(x_{20})\neq c, then the claim y(x2)y(x_{2})\in\mathbb{R} and yy is in CαC^{\alpha} on ¯\overline{\mathcal{I}} are obvious since the coefficients of (3.1) are real. If U(x20)=cU()U(x_{20})=c\notin U(\mathcal{I}), then it must hold x20{x2l,x2r}x_{20}\in\{x_{2l},x_{2r}\} and Lemma 3.10 implies that yC1(¯)y\in C^{1}(\overline{\mathcal{I}}) and W=(μ1y,y)(+x20)W=(\mu^{-1}y,y^{\prime})(\cdot+x_{20}) satisfies (3.74) with W1(0)=0W_{1}(0)=0 and Φ~00\tilde{\Phi}_{0}\equiv 0. This formula yields yy\in\mathbb{R}. Finally, suppose y(x21)=0y(x_{21})=0 at some x21¯{x20}x_{21}\in\overline{\mathcal{I}}\setminus\{x_{20}\}. Let y=(Uc)ξy=(U-c)\xi. Again ξC1({x20,x21})\xi\in C^{1}(\mathcal{I}\cup\{x_{20},x_{21}\}) due to Lemma 3.10 and it is standard to verify

(3.106) ((Uc)2ξ)+k2(Uc)2ξ=0,x2.-\big{(}(U-c)^{2}\xi^{\prime}\big{)}^{\prime}+k^{2}(U-c)^{2}\xi=0,\quad x_{2}\in\mathcal{I}.

Multiplying it by ξ\xi and integrating it between x20x_{20} and x21x_{21} leads to a contradiction. Hence the claim is proved.

For cc\in\mathbb{R}, applying the above claim to yy_{-} on the interval [h,0][-h,0] if cU((h,0))c\notin U\big{(}(-h,0)\big{)} and on [h,x2c][-h,x_{2}^{c}] if cU((h,0))c\in U\big{(}(-h,0)\big{)}, respectively, implies that y(x2)y_{-}(x_{2})\in\mathbb{R} does not change signs on these intervals. Hence we obtain statement (1) and y(x2)>0y_{-}(x_{2})>0 for x2(h,x2c]x_{2}\in(-h,x_{2}^{c}] if cU((h,0))c\in U\big{(}(-h,0)\big{)}. Along with (3.57), the continuity of y(k,c,x2)μsinhμ1(x2+h)\frac{y_{-}(k,c,x_{2})}{\mu\sinh\mu^{-1}(x_{2}+h)}, and Lemma 3.3, it also yields Statement (2).

In the view of Lemma 3.12, Remark 3.7, and statement (1), y(x2)=Imy(x2)y(x_{2})=\text{Im}\,y_{-}(x_{2}) is also a solution on [x2c,0][x_{2}^{c},0] satisfying y(x2c)=0y(x_{2}^{c})=0 and y(x2c)=πU(x2c)U(x2c)y(x2c)y^{\prime}(x_{2}^{c})=\frac{\pi U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{-}(x_{2}^{c}). Statement (3) follows from statement (2) applied to yy_{-} on [h,x2c][-h,x_{2}^{c}] and to Imy\text{Im}\,y_{-} on [x2c,0][x_{2}^{c},0].

From (3.57) and Remark 3.4, there exists k>0k_{*}>0 such that (3.105) holds for all |k|k|k|\geq k_{*} and cc\in\mathbb{C}. For |k|k|k|\leq k_{*}, the restriction on cc involving M>0M>0 ensures y(k,c,0)0y_{-}(k,c,0)\neq 0 due to the semicircle theorem (of the channel flow) and thus we obtain (3.105) from Lemma 3.3, which completes the proof of statement (4).

Finally assume y(k,c,0)0y_{-}(k,c,0)\neq 0 for all k𝐊k\in\mathbf{K} and cSc\in S. Recalling the convergence estimates (3.50) and the locally Hölder continuity of yy_{-} in cc\in\mathbb{R} (Lemma 3.13), we obtain the continuity of yy_{-} in cc\in\mathbb{C} for cI0c_{I}\geq 0. Lemmas 3.3 and 3.9 along with the continuity of y(k,c,0)y_{-}(k,c,0) and the non-vanishing assumption imply that (3.105) holds for all k𝐊k\in\mathbf{K} and cSc\in S with cI0c_{I}\geq 0. As y(k,c¯,x2)=y(k,c,x2)¯y_{-}(k,\bar{c},x_{2})=\overline{y_{-}(k,c,x_{2})}, statement (5) follows and it completes the proof of the lemma. ∎

In the following we give some basic properties of Y(k,c)Y(k,c).

Lemma 3.20.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3. It holds that Y(k,c¯)=Y(k,c)¯Y(k,\bar{c})=\overline{Y(k,c)} and YY is a.) analytic in both (k,c)D(Y)(×U([h,0]))(k,c)\in D(Y)\setminus(\mathbb{R}\times U([-h,0])), and, when restricted to cI0c_{I}\geq 0, b.) Cl02C^{l_{0}-2} in (k,c)D(Y)(×{U(h)})(k,c)\in D(Y)\setminus(\mathbb{R}\times\{U(-h)\}), and c.) Cl02C^{l_{0}-2} in kk and locally CαC^{\alpha} in (k,c)D(Y)(k,c)\in D(Y) for any α[0,1)\alpha\in[0,1). Moreover,

  1. (1)

    Y(k,U(h))Y(k,U(-h))\in\mathbb{R} and Y(0,U(h))=U(0)U(0)U(h)Y(0,U(-h))=\frac{U^{\prime}(0)}{U(0)-U(-h)}.

  2. (2)

    There exists C,ρ>0C,\rho>0 depending only on UU such that

    |Y(k,c)|C(μ1+|logmin{1,|U(0)c|}|),k,|cU(0)|ρ.|Y(k,c)|\leq C\big{(}\mu^{-1}+\big{|}\log\min\big{\{}1,|U(0)-c|\big{\}}\big{|}\big{)},\;\;\forall k\in\mathbb{R},\;|c-U(0)|\leq\rho.
  3. (3)

    For any α(0,12)\alpha\in(0,\frac{1}{2}), there exist k0>0k_{0}>0 and C>0C>0 depending only on α\alpha, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} such that,

    |Y(k,c)kcothkh|C(μα1+|logmin{1,|U(0)c|}|),|k|k0,cU(0).|Y(k,c)-k\coth kh|\leq C(\mu^{\alpha-1}+|\log\min\{1,\,|U(0)-c|\}|),\quad\forall|k|\geq k_{0},\;c\neq U(0).
  4. (4)

    For any M>0M>0 and k>0k_{*}>0, there exists C>0C>0 depending only on kk_{*} and MM such that

    |Y(k,c)kcothkh|Cdist(c,U([h,0])),|k|k,|cU(h)+U(0)2|M+U(0)U(h)2.|Y(k,c)-k\coth kh|\leq\frac{C}{dist(c,U([-h,0]))},\quad\forall|k|\leq k_{*},\;\Big{|}c-\frac{U(-h)+U(0)}{2}\Big{|}\geq M+\frac{U(0)-U(-h)}{2}.
Proof.

The analyticity and the conjugacy property of YY are obvious from its definition. The property Y(k,U(h))Y(k,U(-h))\in\mathbb{R} is a direct corollary of Lemma 3.19(1). The Cl02C^{l_{0}-2} smoothness of YY away from U(h)U(-h) and U(0)U(0) follows from Lemma 3.13 and the analyticity of YY in cc with cI>0c_{I}>0. The Hölder continuity of YY is again a corollary of Lemma 3.13 for cc varying along \mathbb{R} and Proposition 3.7 for cc varying along ii\mathbb{R}. The explicit form of Y(0,U(h))Y\big{(}0,U(-h)\big{)} is a direct consequence of the observation

(3.107) y(0,U(h),x2)=(U(x2)U(h))/U(h).y_{-}\big{(}0,U(-h),x_{2}\big{)}=\big{(}U(x_{2})-U(-h)\big{)}/U^{\prime}(-h).

To end the proof of the lemma, we obtain the quantitive estimate on Y(k,c)Y(k,c). From Lemma 3.19, y(k,U(0),0)0y_{-}(k,U(0),0)\neq 0 for any kk\in\mathbb{R}. Along with Lemma 3.9, it implies that (3.105) holds for |cU(0)|ρ|c-U(0)|\leq\rho for some ρ>0\rho>0 depending only on UU. Statement (2) follows from the upper bound of |y(k,c,0)||y_{-}^{\prime}(k,c,0)| given in Lemma 3.9. Statement (3) is also a direct consequence of Lemma 3.9 where k0k_{0} is involved to ensure y(k,c,0)0y_{-}(k,c,0)\neq 0. In statement (4), the restriction on cc guarantees y(k,c,0)0y_{-}(k,c,0)\neq 0 due to the semicircle theorem and the desired inequality follows Lemma 3.3. ∎

The analyticity of YY in cc allows us to use the Cauchy integral to analyze Y(k,c)Y(k,c). For r>0r>0, let

(3.108) 𝒟r=B(U([h,0]),r)\mathcal{D}_{r}=B\big{(}U([-h,0]),r\big{)}\subset\mathbb{C}

be the rr-neighborhood of U([h,0])U([-h,0])\subset\mathbb{C}.

Lemma 3.21.

Assume UC3U\in C^{3}, kk\in\mathbb{R}, and r>0r>0 satisfy y(k,,0)0y_{-}(k,\cdot,0)\neq 0 on 𝒟r\mathbb{C}\setminus\mathcal{D}_{r}, then for any c𝒟r¯c\in\mathbb{C}\setminus\overline{\mathcal{D}_{r}} and n1n\geq 1 we have

(3.109) Y(k,c)=kcothkh12πi𝒟rY(k,c)ccdc,cnY(k,c)=n!2πi𝒟rY(k,c)(cc)n+1dc,Y(k,c)=k\coth kh-\frac{1}{2\pi i}\oint_{\partial\mathcal{D}_{r}}\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime},\quad\partial_{c}^{n}Y(k,c)=-\frac{n!}{2\pi i}\oint_{\partial\mathcal{D}_{r}}\frac{Y(k,c^{\prime})}{(c^{\prime}-c)^{n+1}}dc^{\prime},

where \oint denote the integral along the contours counterclockwisely.

Here cY=12(cRicI)Y\partial_{c}Y=\frac{1}{2}(\partial_{c_{R}}-i\partial_{c_{I}})Y denotes the derivative of YY as a function of the complex variable cc and thus cY=cRY=icIY\partial_{c}Y=\partial_{c_{R}}Y=-i\partial_{c_{I}}Y due to its analyticity.

Proof.

The assumption implies Y(k,)Y(k,\cdot) is analytic in 𝒟r¯\mathbb{C}\setminus\overline{\mathcal{D}_{r}} and continuous in 𝒟r\mathbb{C}\setminus\mathcal{D}_{r}. For any r1r^{\prime}\gg 1, the analyticity of YY yields

(3.110) Y(k,c)=12πi(𝒟r𝒟r)Y(k,c)ccdc.Y(k,c)=\frac{1}{2\pi i}\left(\oint_{\partial\mathcal{D}_{r^{\prime}}}-\oint_{\partial\mathcal{D}_{r}}\right)\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}.

Applying Lemma 3.20(4) with M=1M=1 and k=1+|k|k_{*}=1+|k|, we have

|Y(k,c)kcothkh|C/dist(c,U([h,0])),c𝒟r.|Y(k,c^{\prime})-k\coth kh|\leq C/dist\big{(}c^{\prime},U([-h,0])\big{)},\quad\forall c^{\prime}\in\partial\mathcal{D}_{r^{\prime}}.

Therefore

limr+(12πi𝒟rY(k,c)ccdckcothkh)=0\lim_{r^{\prime}\to+\infty}\left(\frac{1}{2\pi i}\oint_{\partial\mathcal{D}_{r^{\prime}}}\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}-k\coth kh\right)=0

and thus the desired integral formula of Y(k,c)Y(k,c) follows. The representation of cnY\partial_{c}^{n}Y simply follows from direct differentiation. ∎

Remark 3.12.

Though not needed in the rest of the paper, this lemma could be modified for general kk and cU([h,0])c\notin U([-h,0]). In this case, 0<r<dist(c,U([h,0]))0<r<dist\big{(}c,U([-h,0])\big{)} should be chosen so that y(k,,0)0y_{-}(k,\cdot,0)\neq 0 along 𝒟r\partial\mathcal{D}_{r}. The integral representation formula would involve the residue at those roots of y(k,,0)y_{-}(k,\cdot,0) outside 𝒟r\mathcal{D}_{r}. The estimates should also be modified accordingly.

To analyze the remaining integral in (3.109), we start with the imaginary part YIY_{I} of YY.

Lemma 3.22.

YI(k,c)=0Y_{I}(k,c)=0 for c\U((h,0])c\in\mathbb{R}\backslash U\big{(}(-h,0]\big{)}. Assume UC3U\in C^{3}, c=U(x2c)U((h,0))c=U(x_{2}^{c})\in U\big{(}(-h,0)\big{)}, and y(k,c,0)0y_{-}(k,c,0)\neq 0, then

YI(k,c)=πU(x2c)y(k,c,x2c)2U(x2c)|y(k,c,0)|2.Y_{I}(k,c)=\frac{\pi U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})^{2}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{2}}.
Proof.

The vanishing of YI(k,c)Y_{I}(k,c) for c\U((h,0])c\in\mathbb{R}\backslash U((-h,0]) is obvious from its definition and Lemma 3.19(1). To derive the expression of YI(k,c)Y_{I}(k,c) for c=U(x2c)c=U(x_{2}^{c}) with x2c(h,0)x_{2}^{c}\in(-h,0) and y(k,c,0)0y_{-}(k,c,0)\neq 0, we may consider y(ϵ,x2)=y(k,c+iϵ,x2)y(k,c+iϵ,0)y(\epsilon,x_{2})=\frac{y_{-}(k,c+i\epsilon,x_{2})}{y_{-}(k,c+i\epsilon,0)}, ϵ>0\epsilon>0, which is also a solution to the homogeneous Rayleigh equation with y(ϵ,h)=0y(\epsilon,-h)=0 and y(ϵ,0)=1y(\epsilon,0)=1. It is straight forward to calculate

Imy(ϵ,0)=12ih0x2(yy¯yy¯)dx2=h0ϵU|y|2|Uc|2+ϵ2dx2.\text{Im}\,y^{\prime}(\epsilon,0)=\frac{1}{2i}\int_{-h}^{0}\partial_{x_{2}}(y^{\prime}\bar{y}-y\bar{y}^{\prime})dx_{2}=\int_{-h}^{0}\frac{\epsilon U^{\prime\prime}|y|^{2}}{|U-c|^{2}+\epsilon^{2}}dx_{2}.

Applying the convergence estimates (3.50) and the Hölder continuity of y0(k,c,x2)y_{0-}(k,c,x_{2})\in\mathbb{R} in x2x_{2}, we obtain the desired

YI(k,c)=limϵ0+Imy(ϵ,0)=πU(x2c)y(k,c,x2c)2U(x2c)|y(k,c,0)|2.Y_{I}(k,c)=\lim_{\epsilon\to 0+}\text{Im}\,y^{\prime}(\epsilon,0)=\frac{\pi U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})^{2}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{2}}.

This completes the proof of the lemma. ∎

The above formula yields some refined estimates of YIY_{I} for cU([h,0])c\in U([-h,0]).

Lemma 3.23.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, then the following hold for YI(k,c)Y_{I}(k,c).

  1. (1)

    YI(k,c)Y_{I}(k,c) is Cl02C^{l_{0}-2} in (k,c)D(Y)(×U((h,0)))(k,c)\in D(Y)\cap\big{(}\mathbb{R}\times U\big{(}(-h,0)\big{)}\big{)} and it satisfies

    limcU(0)YI(k,c)=πU(0)U(0),limcU(h)+YI(k,c)(cU(h))2=πU(h)U(h)3|y(k,U(h),0)|2.\lim_{c\to U(0)-}Y_{I}(k,c)=\frac{\pi U^{\prime\prime}(0)}{U^{\prime}(0)},\quad\lim_{c\to U(-h)+}\frac{Y_{I}(k,c)}{(c-U(-h))^{2}}=\frac{\pi U^{\prime\prime}(-h)}{U^{\prime}(-h)^{3}|y_{-}(k,U(-h),0)|^{2}}.

    Moreover, if l04l_{0}\geq 4, then, for any q[1,)q\in[1,\infty), j1,j20j_{1},j_{2}\geq 0, j22j_{2}\leq 2, and j1+j2l04j_{1}+j_{2}\leq l_{0}-4, kj1cRj2YI\partial_{k}^{j_{1}}\partial_{c_{R}}^{j_{2}}Y_{I} is LkWcR1,qL_{k}^{\infty}W_{c_{R}}^{1,q} locally in (k,c)D(Y)(×U([h,0)))(k,c)\in D(Y)\cap\big{(}\mathbb{R}\times U\big{(}[-h,0)\big{)}\big{)}.

  2. (2)

    Assume UC5U\in C^{5}, then there exists C>0C>0 depending only on |U|C4|U^{\prime}|_{C^{4}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} such that, for any cU([h,0))D(Y(k,))c\in U\big{(}[-h,0)\big{)}\cap D\big{(}Y(k,\cdot)\big{)}, we have

    μ2sinh2(μ1(x2c+h))C|y(k,c,0)|2YI(k,c)Cμ2sinh2(μ1(x2c+h))|y(k,c,0)|2,\frac{\mu^{2}\sinh^{2}(\mu^{-1}(x_{2}^{c}+h))}{C|y_{-}(k,c,0)|^{2}}\leq Y_{I}\big{(}k,c\big{)}\leq\frac{C\mu^{2}\sinh^{2}(\mu^{-1}(x_{2}^{c}+h))}{|y_{-}(k,c,0)|^{2}},
    |cRYI(k,c)|\displaystyle|\partial_{c_{R}}Y_{I}(k,c)|\leq Cμsinh(2μ1(x2c+h))|y(k,c,0)|2\displaystyle C\frac{\mu\sinh(2\mu^{-1}(x_{2}^{c}+h))}{|y_{-}(k,c,0)|^{2}}
    +Cμ3sinh(μ1h)sinh(2μ1(x2c+h))|y(k,c,0)|3(1+|logmin{1,|μ1(U(0)c)|}|),\displaystyle+C\frac{\mu^{3}\sinh(\mu^{-1}h)\sinh(2\mu^{-1}(x_{2}^{c}+h))}{|y_{-}(k,c,0)|^{3}}\Big{(}1+\big{|}\log\min\{1,|\mu^{-1}(U(0)-c)|\}\big{|}\Big{)},

    where μ=k1=(1+k2)12\mu=\langle k\rangle^{-1}=(1+k^{2})^{-\frac{1}{2}}.

Proof.

Lemma 3.13 implies the Cl02C^{l_{0}-2} smoothness of y(k,c,0)y_{-}(k,c,0) in kk and cU((h,0))c\in U\big{(}(-h,0)\big{)} and that of y(k,c,x2c)y_{-}(k,c,x_{2}^{c}) in cU([h,0])c\in U([-h,0]), hence YIY_{I} is Cl02C^{l_{0}-2} in (k,c)D(Y)×U((h,0))(k,c)\in D(Y)\cap\mathbb{R}\times U\big{(}(-h,0)\big{)}. Moreover, y(k,c,0)>0y_{-}(k,c,0)>0 for c{(U(h),U(0)}c\in\{(U(-h),U(0)\} due to Lemma 3.19(1), which along with Lemma 3.13a.) implies that y(k,c,0)0y_{-}(k,c,0)\neq 0 for cU([h,0])c\in U([-h,0]) near U(h)U(-h) or U(0)U(0) and thus YI(k,c)Y_{I}(k,c) is Hölder continuous for cU([h,0])c\in U([-h,0]) near U(h)U(-h) and U(0)U(0). The local regularity of kj1cRj2YI\partial_{k}^{j_{1}}\partial_{c_{R}}^{j_{2}}Y_{I} follows from Lemma 3.13(f).

The upper bound estimate of YIY_{I} and its limits as cc approaches U(0)U(0)- and U(h)+U(-h)+ are direct corollaries of Lemmas 3.9 and 3.19(5) and Remark 3.4, as well as (3.89), (3.104) and (3.105). In

cRYI(k,c)=\displaystyle\partial_{c_{R}}Y_{I}(k,c)= c(πU(x2c)U(x2c))y(k,c,x2c)2|y(k,c,0)|2+2πU(x2c)y(k,c,x2c)c(y(k,c,x2c))U(x2c)|y(k,c,0)|2\displaystyle\partial_{c}\left(\frac{\pi U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\right)\frac{y_{-}(k,c,x_{2}^{c})^{2}}{|y_{-}(k,c,0)|^{2}}+\frac{2\pi U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})\partial_{c}\big{(}y_{-}(k,c,x_{2}^{c})\big{)}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{2}}
2πU(x2c)y(k,c,x2c)2(y(k,c,0)cy(k,c,0))U(x2c)|y(k,c,0)|4,\displaystyle-\frac{2\pi U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})^{2}\big{(}y_{-}(k,c,0)\cdot\partial_{c}y_{-}(k,c,0)\big{)}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{4}},

c(y(k,c,x2c))\partial_{c}\big{(}y_{-}(k,c,x_{2}^{c})\big{)} is estimated by (3.89). The other key term cy(k,c,0)\partial_{c}y_{-}(k,c,0) will be considered in three possible cases of cU([h,0])c\in U([-h,0]) according to the division of [h,0]=123[-h,0]=\mathcal{I}_{1}\cup\mathcal{I}_{2}\cup\mathcal{I}_{3} defined in (3.84) in Subsection 3.5. Observing cU([h,0])c\in U([-h,0]) implies 2\mathcal{I}_{2}\neq\emptyset and x2=023x_{2}=0\in\mathcal{I}_{2}\cup\mathcal{I}_{3}.

* Case 1: x2=03x_{2}=0\in\mathcal{I}_{3} and x2l=hx_{2l}=-h. The former happens if and only if cU(0)ρ01μc\leq U(0)-\rho_{0}^{-1}\mu, while x2l=hx_{2l}=-h if and only if cU(h)+ρ01μc\leq U(-h)+\rho_{0}^{-1}\mu. Lemma 3.16 implies

|cy(k,c,0)|Cμ(1+|log(μ1(cU(h)))|)coshμ1h.|\partial_{c}y_{-}(k,c,0)|\leq C\mu\big{(}1+\big{|}\log\big{(}\mu^{-1}(c-U(-h))\big{)}\big{|}\big{)}\cosh\mu^{-1}h.

* Case 2: x2=03x_{2}=0\in\mathcal{I}_{3} and x2l>hx_{2l}>-h which occurs if and only if U(h)+ρ01μcU(0)ρ01μU(-h)+\rho_{0}^{-1}\mu\leq c\leq U(0)-\rho_{0}^{-1}\mu. Also from Lemma 3.16, we have

|cy(k,c,0)|Cμcoshμ1h.|\partial_{c}y_{-}(k,c,0)|\leq C\mu\cosh\mu^{-1}h.

* Case 3: x2=02x_{2}=0\in\mathcal{I}_{2} which happens iff U(0)cρ01μU(0)-c\leq\rho_{0}^{-1}\mu and thus x2r=0x_{2r}=0. From the definitions (3.84) of 2\mathcal{I}_{2}, (3.54) of ρ0\rho_{0}, and (3.4) of h0h_{0}, it holds

0U(x2r)U(x2l)2ρ01μ12h0infUx2l=x2rx2l12h0x2l>h.0\leq U(x_{2r})-U(x_{2l})\leq 2\rho_{0}^{-1}\mu\leq\tfrac{1}{2}h_{0}\inf U^{\prime}\Longrightarrow-x_{2l}=x_{2r}-x_{2l}\leq\tfrac{1}{2}h_{0}\Longrightarrow x_{2l}>-h.

This in turn implies cU(x2l)=ρ01μc-U(x_{2l})=\rho_{0}^{-1}\mu and thus Lemma 3.15 yields

|cy(k,c,0)|Cμ(1+|log(μ1(U(0)c))|)coshμ1h.|\partial_{c}y_{-}(k,c,0)|\leq C\mu\big{(}1+\big{|}\log\big{(}\mu^{-1}(U(0)-c)\big{)}\big{|}\big{)}\cosh\mu^{-1}h.

The desired estimates on cRYI\partial_{c_{R}}Y_{I} follow from (3.89), Lemmas 3.19 and 3.9, and the above estimates. In particular, in the above case 1, we also used

μsinh|μ1(x2c+h)||log|μ1(cU(h))||Cμcosh(μ1(x2c+h)),\mu\sinh|\mu^{-1}(x_{2}^{c}+h)|\big{|}\log|\mu^{-1}\big{(}c-U(-h)\big{)}|\big{|}\leq C\mu\cosh(\mu^{-1}(x_{2}^{c}+h)),

which can be shown by considering whether |μ1(cU(h))|1|\mu^{-1}(c-U(-h))|\leq 1 separately. ∎

In the following we analyze Y(k,c)Y(k,c) by writing it as a Cauchy integral of YIY_{I}.

Lemma 3.24.

Assume UCl0U\in C^{l_{0}}, l05l_{0}\geq 5, and kk\in\mathbb{R} satisfy that y(k,c,0)0y_{-}(k,c,0)\neq 0 for all cU([h,0])c\in U([-h,0]), then Y(k,c)Y(k,c) and kj1cRj2Y(k,c)\partial_{k}^{j_{1}}\partial_{c_{R}}^{j_{2}}Y(k,c) are LkLcRqL_{k}^{\infty}L_{c_{R}}^{q} locally in kk\in\mathbb{R} and cRc_{R} in the domain D(Y){cI0}D(Y)\cap\{c_{I}\geq 0\} for any q(1,)q\in(1,\infty), 0j220\leq j_{2}\leq 2, and 0j1l03j20\leq j_{1}\leq l_{0}-3-j_{2}. Assume, in addition, y(k,c,0)0y_{-}(k,c,0)\neq 0 for all cc\in\mathbb{C}, then, for any cU([h,0])c\notin U([-h,0]),

(3.111) Y(k,c)=1πU(h)U(0)YI(k,c)ccdc+kcothkh,Y(k,c)=\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{Y_{I}(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}+k\coth kh,

and for cU([h,0))c\in U\big{(}[-h,0)\big{)},

(3.112) Y(k,c)=(YI(k,))(c)+iYI(k,c)+kcothkh.Y(k,c)=-\mathcal{H}\big{(}Y_{I}(k,\cdot)\big{)}(c)+iY_{I}(k,c)+k\coth kh.

Here \mathcal{H} denotes the Hilbert transform in cc\in\mathbb{R}, namely,

(YI(k,))(c)=1πP.V.YI(k,c)ccdc=1πP.V.U(h)U(0)YI(k,c)ccdc,\mathcal{H}\big{(}Y_{I}(k,\cdot)\big{)}(c)=\frac{1}{\pi}\text{P.V.}\int_{\mathbb{R}}\frac{Y_{I}(k,c^{\prime})}{c-c^{\prime}}dc^{\prime}=\frac{1}{\pi}\text{P.V.}\int_{U(-h)}^{U(0)}\frac{Y_{I}(k,c^{\prime})}{c-c^{\prime}}dc^{\prime},

where P.V.\int represent the principle value of the singular integral. We also recall Y(k,c)=Y(k,c+i0)Y(k,c)=Y(k,c+i0) and Y(k,ci0)=Y(k,c+i0)¯Y(k,c-i0)=\overline{Y(k,c+i0)} for cc\in\mathbb{R}.

Proof.

Let us first assume y(k,c,0)0y_{-}(k,c,0)\neq 0 for all cc\in\mathbb{C}, then Y(k,c)Y(k,c) is well-defined for all cU(0)c\neq U(0). We shall apply Lemma 3.21. The contour 𝒟r\partial\mathcal{D}_{r} is the union of two segments [U(h),U(0)]±ir[U(-h),U(0)]\pm ir, the left half circle centered at U(h)U(-h) with radius rr, and the right half circle centered at U(0)U(0) with radius rr. As r0+r\to 0+, due to the continuity of YY (when restricted to {cI0}\{c_{I}\geq 0\}) at cU(0)c\neq U(0) and its logarithmic upper bound near U(0)U(0) given in Lemma 3.20, the Cauchy integrals along the two half circles converge to zero as r0+r\to 0+. Hence the integral form (3.111) of Y(k,c)Y(k,c) follows from taking the limit of (3.109) as r0+r\to 0+ and the conjugacy Y(k,c¯)=Y(k,c)¯Y(k,\overline{c^{\prime}})=\overline{Y(k,c^{\prime})}.

For cI0c_{I}\neq 0, the integral form (3.111) can be rewritten as

Y(k,c)=1πU(h)U(0)(ccR+icI)(ccR)2+cI2YI(k,c)dc+kcothkh.Y(k,c)=\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{(c^{\prime}-c_{R}+ic_{I})}{(c^{\prime}-c_{R})^{2}+c_{I}^{2}}Y_{I}(k,c^{\prime})dc^{\prime}+k\coth kh.

A standard treatment of the above singular integral as cI0+c_{I}\to 0+, along with the regularity of YI(k,c)Y_{I}(k,c^{\prime}) in cU([h,0))c^{\prime}\in U\big{(}[-h,0)\big{)} given in Lemma 3.22 and 3.23, yields (3.112).

The regularity of YY follows from that of YIY_{I} and the boundedness in LqL^{q} of the convolution by 1c+icI\frac{1}{c^{\prime}+ic_{I}} with the parameter cI0c_{I}\geq 0. Here the singularity of YIY_{I} near c=U(0)c=U(0) does not affect the regularity of YY away from U(0)U(0) due to the localization property of this convolution operator.

Finally, if we only assume y(k,c,0)0y_{-}(k,c,0)\neq 0 for cU([h,0])c\in U([-h,0]), due to its analyticity in U([h,0])\mathbb{C}\setminus U([-h,0]), its continuity when restricted to {cI0}\{c_{I}\geq 0\}, and lim|c|y(k,c,0)=k1sinhkh\lim_{|c|\to\infty}y_{-}(k,c,0)=k^{-1}\sinh kh (Lemma 3.3), Y(k,)Y(k,\cdot) has at most finitely many singular points U([h,0])\mathbb{C}\setminus U([-h,0]). Hence there would be at most finitely many additional contour integrals of YY in (3.111) along contours in U([h,0])\mathbb{C}\setminus U([-h,0]) enclosing the roots of y(k,,0)y_{-}(k,\cdot,0). Those integrals in the analytic region of YY would not affect the regularity of YY. The proof of the lemma is complete. ∎

With the representation of YY in terms of Cauchy integrals, we may also calculate its derivatives in more details.

Corollary 3.24.1.

It holds, for cU([h,0])c\notin U([-h,0]),

(3.113) cY(k,c)=1πU(h)U(0)YI(k,c)(cc)2dc=1πU(h)U(0)cRYI(k,c)ccdcU(0)U(0)(U(0)c),\partial_{c}Y(k,c)=\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{Y_{I}(k,c^{\prime})}{(c^{\prime}-c)^{2}}dc^{\prime}=\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{\partial_{c_{R}}Y_{I}(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}-\frac{U^{\prime\prime}(0)}{U^{\prime}(0)\big{(}U(0)-c\big{)}},

and for cU([h,0))c\in U\big{(}[-h,0)\big{)},

(3.114) cY(k,c)=(cRYI(k,))(c)+icRYI(k,c)U(0)U(0)(U(0)c).\begin{split}\partial_{c}Y(k,c)=-\mathcal{H}\big{(}\partial_{c_{R}}Y_{I}(k,\cdot)\big{)}(c)&+i\partial_{c_{R}}Y_{I}(k,c)-\frac{U^{\prime\prime}(0)}{U^{\prime}(0)\big{(}U(0)-c\big{)}}.\end{split}

Using the regularity of and estimates on YIY_{I} and cRYI\partial_{c_{R}}Y_{I} given in Lemma 3.23 , (3.113) follows from direct differentiation and integration by parts, along with the explicit form of YI(k,U(0))Y_{I}\big{(}k,U(0)-\big{)}. Equality (3.114) is obtained by taking the limit of (3.113) as cI0+c_{I}\to 0+. We omit the details of these straight forward calculations.

4. Eigenvalues of the linearization of the water wave at shear flows

In this section, we shall discuss the distribution of eigenvalues of the linearized gravity-capillary water wave system (1.3) at the shear flow (U(x2),0)T\big{(}U(x_{2}),0\big{)}^{T}. As (1.3) preserves Fourier mode eikx1e^{ikx_{1}} for any kk, the wave number kk\in\mathbb{R} would be treated as a parameter in this section. According to Lemma 2.3, ikc-ikc\in\mathbb{C}, cU([h,0])c\in\mathbb{C}\setminus U([-h,0]), is an eigenvalue of (1.3) with parameter kk if and only if

(4.1) 𝐅(k,c)=𝐅R+i𝐅I:=(g+σk2)y+(k,c,h)=(g+σk2)(y+yy+y)(k,c,0)=(U(0)c)2y(k,c,0)(U(0)(U(0)c)+g+σk2)y(k,c,0)=0,\begin{split}&\mathbf{F}(k,c)=\mathbf{F}_{R}+i\mathbf{F}_{I}:=(g+\sigma k^{2})y_{+}(k,c,-h)=(g+\sigma k^{2})(y_{+}y_{-}^{\prime}-y_{+}^{\prime}y_{-})(k,c,0)\\ =&\big{(}U(0)-c\big{)}^{2}y_{-}^{\prime}(k,c,0)-\big{(}U^{\prime}(0)\big{(}U(0)-c\big{)}+g+\sigma k^{2}\big{)}y_{-}(k,c,0)=0,\end{split}

where the last equal sign in the first row is due to the conservation of the Wronskian of (3.1). Let

𝐅(k,c)=limϵ0+𝐅(k,c+iϵ)=limϵ0+𝐅(k,ciϵ)¯,cU([h,0]).\mathbf{F}(k,c)=\lim_{\epsilon\to 0+}\mathbf{F}(k,c+i\epsilon)=\overline{\lim_{\epsilon\to 0+}\mathbf{F}(k,c-i\epsilon)},\quad c\in U\big{(}[-h,0]\big{)}.

It is easy to see that, if 𝐅(k,c)=0\mathbf{F}(k,c)=0, then y(k,c,x2)y_{-}(k,c,x_{2}) also generates the associated eigenfunction of (1.3). In the literatures, those zero point cc of 𝐅\mathbf{F} with cI>0c_{I}>0 are often referred to as unstable modes, while those zero point cc\in\mathbb{R} as neutral modes. We recall that Yih proved that the semicircle theorem also holds for free boundary problem [39], namely, (1.6) holds for all unstable modes.

From the analysis in Subsection 3.5, it is not clear whether 𝐅\mathbf{F} is C1C^{1} at c=U(h)c=U(-h) which would be crucial for the bifurcation analysis of eigenvalues. We also consider an almost equivalent quantity

(4.2) F(k,c)=y(k,c,0)1𝐅=FR+iFI=Y(k,c)(U(0)c)2U(0)(U(0)c)(g+σk2)=0,F(k,c)=y_{-}(k,c,0)^{-1}\mathbf{F}=F_{R}+iF_{I}=Y(k,c)\big{(}U(0)-c\big{)}^{2}-U^{\prime}(0)\big{(}U(0)-c\big{)}-(g+\sigma k^{2})=0,

where Y(k,c)Y(k,c) is defined in (3.103), and

F(k,c)=limϵ0+F(k,c+iϵ)=limϵ0+F(k,ciϵ)¯,cU([h,0]).F(k,c)=\lim_{\epsilon\to 0+}F(k,c+i\epsilon)=\overline{\lim_{\epsilon\to 0+}F(k,c-i\epsilon)},\quad\forall c\in U\big{(}[-h,0]\big{)}.

Apparently 𝐅\mathbf{F} and FF satisfy

(4.3) 𝐅(k,c)=𝐅(k,c)=𝐅(k,c¯)¯,cU((h,0));\mathbf{F}(-k,c)=\mathbf{F}(k,c)=\overline{\mathbf{F}(k,\bar{c})},\;\forall c\notin U\big{(}(-h,0)\big{)};
(4.4) F(k,c)=F(k,c)=F(k,c¯)¯,cD(Y)U((h,0)).F(-k,c)=F(k,c)=\overline{F(k,\bar{c})},\;c\in D(Y)\setminus U\big{(}(-h,0)\big{)}.

From Lemma 3.24 FF is C1,αC^{1,\alpha} near c0=U(h)c_{0}=U(-h) if y(k,c,0)0y_{-}(k,c,0)\neq 0 for all cU([h,0])c\in U([-h,0]), which is crucial for the bifurcation analysis.

4.1. Basic properties of eigenvalues

Apparently it holds that

(4.5) 𝐅 is analytic in k&cU([h,0]) and F analytic in k&cD(Y)U([h,0]),\mathbf{F}\text{ is analytic in }k\in\mathbb{R}\ \&\ c\notin U([-h,0])\text{ and }\,F\text{ analytic in }k\in\mathbb{R}\ \&\ c\in D(Y)\setminus U([-h,0]),
𝐅(k,c)=0c is a non-singular or singular mode of (2.11).\mathbf{F}(k,c)=0\Longleftrightarrow\;c\text{ is a non-singular or singular mode of \eqref{E:Ray}}.

In the following we first give some basic properties of 𝐅\mathbf{F} under minimal assumptions.

Lemma 4.1.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, then for any kk\in\mathbb{R}, the following hold.

  1. (1)

    𝐅\mathbf{F} is well defined for all kk\in\mathbb{R} and cc\in\mathbb{C}. When restricted to cI0c_{I}\geq 0, 𝐅\mathbf{F} is Cl02C^{l_{0}-2} in kk and c{U(h),U(0)}c\notin\{U(-h),U(0)\} and 𝐅\mathbf{F} is also CαC^{\alpha} in both kk and cc with cI0c_{I}\geq 0.

  2. (2)

    F(k,c)F(k,c) is well-defined for cc close to U(h)U(-h) and U(0)U(0), C1C^{1} near c=U(0)c=U(0), and

    F(k,U(h)),F(0,U(h))=g,F(k,U(0))=gσk2,cF(k,U(0))=U(0).F(k,U(-h))\in\mathbb{R},\quad F\big{(}0,U(-h)\big{)}=-g,\quad F\big{(}k,U(0)\big{)}=-g-\sigma k^{2},\quad\partial_{c}F\big{(}k,U(0)\big{)}=U^{\prime}(0).
  3. (3)

    Assume UC5U\in C^{5}, then for any r(1,)r\in(1,\infty), there exists C>0C>0 determined only by rr, |U|C4|U^{\prime}|_{C^{4}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any cI0c_{I}\geq 0 and kk\in\mathbb{R},

    |c𝐅(k,+icI)|LcRrCμ1eμ1h,limcI0+|𝐅(k,+icI)𝐅(k,)|WcR1,r=0,|\partial_{c}\mathbf{F}(k,\cdot+ic_{I})|_{L_{c_{R}}^{r}}\leq C\mu^{-1}e^{\mu^{-1}h},\quad\lim_{c_{I}\to 0+}|\mathbf{F}(k,\cdot+ic_{I})-\mathbf{F}(k,\cdot)|_{W_{c_{R}}^{1,r}}=0,

    where the norm is taken on cR[12h0h,12h0]c_{R}\in[-\frac{1}{2}h_{0}-h,\frac{1}{2}h_{0}] and we recall μ=(1+k2)12\mu=(1+k^{2})^{-\frac{1}{2}}.

  4. (4)

    𝐅(k,c)0\mathbf{F}(k,c)\neq 0 if y(k,c,0)=0y_{-}(k,c,0)=0. Hence {c𝐅(k,c)=0}={cF(k,c)=0}\{c\mid\mathbf{F}(k,c)=0\}=\{c\mid F(k,c)=0\} for any kk\in\mathbb{R}.

  5. (5)

    𝐅(k,c)=0\mathbf{F}(k,c)=0 iff there exists a C2C^{2} solution y(x2)y(x_{2}) to (2.14) satisfying the corresponding homogeneous boundary conditions of (2.11b-2.11c).

  6. (6)

    For any x2(0,h)x_{2}\in(0,-h), FI(k,U(x2))0F_{I}(k,U(x_{2}))\neq 0 if U(x2)0U^{\prime\prime}(x_{2})\neq 0.

Proof.

For cRU([h,0))c_{R}\in U\big{(}[-h,0)\big{)}, the convergence of 𝐅(k,cR+icI)\mathbf{F}(k,c_{R}+ic_{I}) as cI0+c_{I}\to 0+ follows from the convergence estimates given in Proposition 3.7. For cc near U(0)U(0), the logarithmic singularity in y(k,c,0)y_{-}^{\prime}(k,c,0) is cancelled by (U(0)c)2(U(0)-c)^{2} and thus the convergence of 𝐅(k,U(0)+icI)\mathbf{F}(k,U(0)+ic_{I}) and the continuity of 𝐅\mathbf{F} at c=U(0)c=U(0) follow. The CαC^{\alpha} and Cl02C^{l_{0}-2} smoothness of 𝐅\mathbf{F} is obtained from those of y(k,c,0)y_{-}(k,c,0) and y(k,c,0)y_{-}^{\prime}(k,c,0) (Lemmas 3.13 and 3.15) as well as using the factor (U(0)c)2(U(0)-c)^{2} multiplied to y(k,c,0)y_{-}^{\prime}(k,c,0).

From Lemma 3.19(1), y(k,U(h),0),y(k,U(0),0)>0y_{-}(k,U(-h),0),y_{-}(k,U(0),0)>0 and thus FF is well-defined near c=U(h),U(0)c=U(-h),U(0). The property F(k,U(h))F(k,U(-h))\in\mathbb{R} and the value of F(0,U(h))F(0,U(-h)) are due to those of YY given in Lemma 3.20(1). The C1C^{1} smoothness of FF for cc near U(0)U(0) follows from Lemma 3.20(2) and the definition of FF. The values of FF and cF\partial_{c}F at (k,U(0))(k,U(0)) is obtained by direct computation.

Statement (3) is a corollary of Proposition 3.7, Lemma 3.18(3), and the definition of 𝐅\mathbf{F}.

Suppose y(k,c,0)=0y_{-}(k,c,0)=0. Lemma 3.19(1) implies cU(0)c\neq U(0). As a non-trivial solution to the homogeneous Rayleigh equation (3.1), it must hold y(k,c,0)0y_{-}^{\prime}(k,c,0)\neq 0. Therefore 𝐅(k,c)0\mathbf{F}(k,c)\neq 0.

To prove statement (5), we first observe that 𝐅(k,c)=0\mathbf{F}(k,c)=0 iff yy_{-} satisfies the corresponding homogeneous boundary conditions of (2.11c), which happens only if y(k,c,0)0y_{-}(k,c,0)\neq 0 and thus Y(k,c)Y(k,c) and F(k,c)F(k,c) are well-defined. Moreover the statement is obvious for cU([h,0])c\notin U\big{(}[-h,0]\big{)} and also for c=U(h)c=U(-h) due to the smoothness of yy_{-} (Lemma 3.10), while F(k,U(0))0F(k,U(0))\neq 0 due to statement (2). Hence we focus on cU((h,0))c\in U\big{(}(-h,0)\big{)} only. “\Longrightarrow”: As cU((h,0))c\in U\big{(}(-h,0)\big{)}, F(k,c)=0F(k,c)=0 implies YI(k,c)=0Y_{I}(k,c)=0 and consequently U(x2c)=0U^{\prime\prime}(x_{2}^{c})=0 according to Lemma 3.22. Consequently Lemma 3.10, particularly formula (3.74), and the definition of Γ0\Gamma_{0} yield the smoothness of yy_{-} which apparently satisfies (2.14). “\Longleftarrow”: This solution y(x2)y(x_{2}) has to be proportional to yy_{-} on [h,x2c][-h,x_{2}^{c}] which yields y(x2c)0y(x_{2}^{c})\neq 0 due to 3.19(2). Hence the smoothness of y(x2)y(x_{2}) and equation (2.14) imply U(x2c)=0U^{\prime\prime}(x_{2}^{c})=0. Consequently both (2.14) and the homogeneous Rayleigh equation (3.1) are regular on [h,0][-h,0] and are equivalent to each other. Therefore y(x2)y_{-}(x_{2}) and y(x2)y(x_{2}) are proportional on [h,0][-h,0] and thus yy_{-} satisfies the boundary condition at x2=0x_{2}=0.

To prove the last statement, let c=U(x2)c=U(x_{2}), x2(h,0)x_{2}\in(-h,0). According to Lemma 3.19, Imy(k,c,0)0\text{Im}\,y_{-}(k,c,0)\neq 0 if U(x2)0U^{\prime\prime}(x_{2})\neq 0 and thus Y(k,c)Y(k,c) is well-defined. Lemma 3.22 yields

(4.6) FI(k,c)=(U(0)c)2YI(k,c)=π(U(0)c)2U(x2)y(k,c,x2)2U(x2)|y(k,c,0)|20,F_{I}(k,c)=\big{(}U(0)-c\big{)}^{2}Y_{I}(k,c)=\frac{\pi(U(0)-c)^{2}U^{\prime\prime}(x_{2})y_{-}(k,c,x_{2})^{2}}{U^{\prime}(x_{2})|y_{-}(k,c,0)|^{2}}\neq 0,

which prove statement (5). This is the same argument as in [39] in the case of gravity waves. ∎

Remark 4.1.

The monotonicity assumption on UU is used in the above proof of statement (5). If UU is not monotonic, U1(c)U^{-1}(c) may contain several points in [h,0][-h,0] for a root of F(k,)F(k,\cdot) and the corresponding solution y(k,c,x2)y_{-}(k,c,x_{2}) may not be in Hx22H_{x_{2}}^{2}. Therefore the set of roots of F(k,)F(k,\cdot), which is what really matters, may be larger than those defined as singular modes in Definition 2.1.

In the next step, we consider 𝐅\mathbf{F} for |k|1|k|\gg 1. Unlike the linearized Euler equation on a fixed channel where no eigenvalues exist for large kk. Eigenvalues do exist for each large kk for the linearized water wave system. According to Lemma 4.2(2), we often consider F(k,c)F(k,c) as well.

Lemma 4.2.

Assume UC3U\in C^{3}, then the following hold for any α(0,12)\alpha\in(0,\frac{1}{2}).

  1. (1)

    There exists C>0C>0 depending only on α\alpha, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that

    |𝐅+σk2μsinhμ1h(U(0)c)2coshμ1h|C(μα1+|c|2μα)coshμ1h,\displaystyle|\mathbf{F}+\sigma k^{2}\mu\sinh\mu^{-1}h-(U(0)-c)^{2}\cosh\mu^{-1}h|\leq C\big{(}\mu^{\alpha-1}+|c|^{2}\mu^{\alpha}\big{)}\cosh\mu^{-1}h,

    where we recall μ=(1+k2)12\mu=(1+k^{2})^{-\frac{1}{2}}.

  2. (2)

    For any k,M>0k_{*},M>0, there exists C>0C>0 depending only on MM, kk_{*}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any |k|k|k|\leq k_{*} and cc satisfying dist(c,U([h,0])Mdist(c,U([-h,0])\geq M,

    |𝐅(U(0)c)2coshkh|C(1+|c|+|U(0)c|2dist(c,U([h,0]))1).|\mathbf{F}-(U(0)-c)^{2}\cosh kh|\leq C\big{(}1+|c|+|U(0)-c|^{2}dist(c,U([-h,0])\big{)}^{-1}\big{)}.
  3. (3)

    There exist k0>0k_{0}>0 and C>0C>0 depending only on |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any |k|k0|k|\geq k_{0}, (4.1) has exactly two solutions c±(k)c^{\pm}(k) which are contained in U([h,0])\mathbb{C}\setminus U([-h,0]) and depend on kk analytically. Moreover they satisfy

    c±(k),c±(k)=c±(k),|c±(k)σ|k|U(0)|C,|cF(k,c±(k))2σ|k|32|C|k|.c^{\pm}(k)\in\mathbb{R},\quad c^{\pm}(-k)=c^{\pm}(k),\quad\left|c^{\pm}(k)\mp\sqrt{\sigma|k|}-U(0)\right|\leq C,\quad\big{|}\partial_{c}F\big{(}k,c^{\pm}(k)\big{)}\mp 2\sqrt{\sigma}|k|^{\frac{3}{2}}\big{|}\leq C|k|.
Proof.

The first statement follows directly from Lemma 3.9, where the factor (U(0)c)2(U(0)-c)^{2} is used to cancel the logarithmic singularity in the estimate of yy_{-}^{\prime}, and the second from Lemma 3.3 with C0=dist(c,U([h,0])1C_{0}=dist(c,U([-h,0])^{-1}. We focus on the roots of 𝐅\mathbf{F}. From Lemma 3.9,

k0>0, s. t. |y(k,c,0)|(1/2)μsinhμ1h>0,|k|k0,c,\exists k_{0}>0,\text{ s. t. }|y_{-}(k,c,0)|\geq(1/2)\mu\sinh\mu^{-1}h>0,\quad\forall|k|\geq k_{0},\,c\in\mathbb{C},

and thus we can work with F(k,c)F(k,c) and Y(k,c)Y(k,c). Let

Sk={c|c|σ|k|/2}.S_{k}=\{c\in\mathbb{C}\mid|c|\geq\sqrt{\sigma|k|}/2\}.

From statement (1) and Lemma 3.20(3), it holds that there exist k0>0k_{0}>0 such that, for any |k|k0|k|\geq k_{0}, F(k,c)=0F(k,c)=0 only if cSkc\in S_{k}. We may take larger k0>0k_{0}>0 if necessary such that dist(Sk,U([h,0]))1dist\big{(}S_{k},U([-h,0])\big{)}\geq 1. From Lemma 3.24 and 3.23 and Corollary 3.24.1, there exists C>0C>0 depending only on UU such that, for all |k|k0|k|\geq k_{0}, cSkc\in S_{k},

|Y(k,c)kcothkh|C(1+|c|)sinh2hμU(h)U(0)sinh21μ(U1(c)+h)dc,|Y(k,c)-k\coth kh|\leq\frac{C}{(1+|c|)\sinh^{2}\frac{h}{\mu}}\int_{U(-h)}^{U(0)}\sinh^{2}\tfrac{1}{\mu}(U^{-1}(c^{\prime})+h)dc^{\prime},
|cY(k,c)|C1+|c|+C(1+|c|)μsinh2hμU(h)U(0)sinh2μ(U1(c)+h)dc.|\partial_{c}Y(k,c)|\leq\frac{C}{1+|c|}+\frac{C}{(1+|c|)\mu\sinh^{2}\frac{h}{\mu}}\int_{U(-h)}^{U(0)}\sinh\tfrac{2}{\mu}\big{(}U^{-1}(c^{\prime})+h\big{)}dc^{\prime}.

By a substitution τ=1μ(U1(c)+h)\tau=\tfrac{1}{\mu}\big{(}U^{-1}(c^{\prime})+h\big{)} we obtain

|Y(k,c)kcothkh|C(|k|+1)1(1+|c|)1,|cY(k,c)|C(1+|c|)1,|k|k0.|Y(k,c)-k\coth kh|\leq C(|k|+1)^{-1}(1+|c|)^{-1},\quad|\partial_{c}Y(k,c)|\leq C(1+|c|)^{-1},\quad\forall|k|\geq k_{0}.

On the other hand, viewing F(k,c)=0F(k,c)=0 as a quadratic equation of U(0)cU(0)-c, its roots also satisfy

c=f±(k,c), where f±(k,c)=U(0)U(0)2Y(k,c)±U(0)24Y(k,c)2+g+σk2Y(k,c).c=f_{\pm}(k,c),\;\text{ where }\;f_{\pm}(k,c)=U(0)-\frac{U^{\prime}(0)}{2Y(k,c)}\pm\sqrt{\frac{U^{\prime}(0)^{2}}{4Y(k,c)^{2}}+\frac{g+\sigma k^{2}}{Y(k,c)}}.

Using the above estimates on YY and coths=1+2e2s1\coth s=1+\tfrac{2}{e^{2s}-1}, it is straight forward to verify that for any |k|k0|k|\geq k_{0} and cSkc\in S_{k},

|f±(k,c)σ|k|U(0)|C,\left|f_{\pm}(k,c)\mp\sqrt{\sigma|k|}-U(0)\right|\leq C,

and

|cf±(k,c)|=\displaystyle|\partial_{c}f_{\pm}(k,c)|= |U(0)2Y212(U(0)24Y2+g+σk2Y)12(U(0)22Y3+g+σk2Y2)||cY|C|k|.\displaystyle\left|\frac{U^{\prime}(0)}{2Y^{2}}\mp\frac{1}{2}\left(\frac{U^{\prime}(0)^{2}}{4Y^{2}}+\frac{g+\sigma k^{2}}{Y}\right)^{-\frac{1}{2}}\left(\frac{U^{\prime}(0)^{2}}{2Y^{3}}+\frac{g+\sigma k^{2}}{Y^{2}}\right)\right||\partial_{c}Y|\leq\frac{C}{|k|}.

Therefore f±(k,)f_{\pm}(k,\cdot) are contractions acting on SkS_{k}. Their fixed points c±(k)c^{\pm}(k), analytic in kk, are the only solutions to (4.1), or equivalently (4.2). These c±(k)c^{\pm}(k)\in\mathbb{R} since f±(k,c)f_{\pm}(k,c)\in\mathbb{R} for cc\in\mathbb{R} which allows the iteration to be taken in \mathbb{R}. Finally, one may compute

(4.7) cF=(U(0)c)2cY+2(cU(0))Y+U(0).\partial_{c}F=(U(0)-c)^{2}\partial_{c}Y+2(c-U(0))Y+U^{\prime}(0).

Using the above estimates on Y|k|Y-|k|, cY\partial_{c}Y, and c±(k)c^{\pm}(k), one may compute

|cF(k,c±(k))2σ|k|32|=|2Y(cU(0))2σ|k|32+cY(U(0)c)2+U(0)|c=c±(k)C|k|.\big{|}\partial_{c}F\big{(}k,c^{\pm}(k)\big{)}\mp 2\sqrt{\sigma}|k|^{\frac{3}{2}}\big{|}=\big{|}2Y\big{(}c-U(0)\big{)}\mp 2\sqrt{\sigma}|k|^{\frac{3}{2}}+\partial_{c}Y\big{(}U(0)-c\big{)}^{2}+U^{\prime}(0)\big{|}_{c=c^{\pm}(k)}\leq C|k|.

The evenness of c±(k)c^{\pm}(k) in kk is due to that of 𝐅(k,c)\mathbf{F}(k,c) and the uniqueness of the fixed points of the above contractions. This completes the proof of the lemma. ∎

We shall track the two roots c±(k)c^{\pm}(k) of the analytic function F(k,)F(k,\cdot) as |k||k| decreases, based on a standard analytic continuation argument.

Lemma 4.3.

Assume UC3U\in C^{3}. Suppose k0k_{0}\in\mathbb{R} and c0U([h,0])c_{0}\in\mathbb{C}\setminus U([-h,0]) satisfy 𝐅(k0,c0)=0\mathbf{F}(k_{0},c_{0})=0 and c𝐅(k0,c0)0\partial_{c}\mathbf{F}(k_{0},c_{0})\neq 0, then the following hold.

  1. (1)

    There exists an analytic function c(k)U([h,0])c(k)\in\mathbb{C}\setminus U([-h,0]) defined on a maximal interval (k,k+)k0(k_{-},k_{+})\ni k_{0} such that 𝐅(k,c(k))=0\mathbf{F}\big{(}k,c(k)\big{)}=0 and c𝐅(k,c(k))0\partial_{c}\mathbf{F}\big{(}k,c(k)\big{)}\neq 0.

  2. (2)

    c(k)c(k)\in\mathbb{R} for all k(k,k+)k\in(k_{-},k_{+}) if and only if c0c_{0}\in\mathbb{R}.

  3. (3)

    If k+<k_{+}<\infty (or k>k_{-}>-\infty), then

    1. (a)

      limk(k+)dist(c(k),U([h,0]))=0\lim_{k\to(k_{+})-}dist(c(k),U([-h,0]))=0 (or limk(k)+dist(c(k),U([h,0]))=0\lim_{k\to(k_{-})+}dist(c(k),U([-h,0]))=0 if k>k_{-}>-\infty), or

    2. (b)

      lim infk(k+)min{|c(k)c|:c s. t. 𝐅(k,c)=0,cc(k)}=0\liminf_{k\to(k_{+})-}\min\{|c(k)-c|\,:\,\forall c\text{ s. t. }\mathbf{F}(k,c)=0,\,c\neq c(k)\}=0 (or lim infk(k)+min{|c(k)c|:𝐅(k,c)=0,cc(k)}=0\liminf_{k\to(k_{-})+}\min\{|c(k)-c|\,:\,\mathbf{F}(k,c)=0,\,c\neq c(k)\}=0 if k>k_{-}>-\infty).

Proof.

We start the proof with a simple and standard consideration of the index of complex analytic functions. Suppose 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 at any cΩc\in\partial\Omega where ΩU([h,0])\Omega\subset\mathbb{C}\setminus U([-h,0]) is a domain with piecewise smooth boundary Ω\partial\Omega, then the index

(4.8) Ind(𝐅(k,),Ω):=12πiΩc𝐅(k,c)𝐅(k,c)dc{0}\text{Ind}\big{(}\mathbf{F}(k,\cdot),\Omega\big{)}:=\frac{1}{2\pi i}\oint_{\partial\Omega}\frac{\partial_{c}\mathbf{F}(k,c)}{\mathbf{F}(k,c)}dc\in\mathbb{N}\cup\{0\}

is equal to the number of zeros of 𝐅(k,)\mathbf{F}(k,\cdot) inside Ω\Omega, counting their multiplicities. Therefore the analyticity of FF in kk and cc implies that Ind(𝐅(k,),Ω)\big{(}\mathbf{F}(k,\cdot),\Omega\big{)} is a constant in kk as long as 𝐅(k,c)=0\mathbf{F}(k,c)=0 does not occur on Ω\partial\Omega.

As a consequence, starting with the simple root c0U([h,0])c_{0}\in\mathbb{C}\setminus U([-h,0]) of 𝐅(k0,)\mathbf{F}(k_{0},\cdot), a unique continuation of c(k)U([h,0])c(k)\subset\mathbb{C}\setminus U([-h,0]) of simple roots of 𝐅(k,)\mathbf{F}(k,\cdot) exists and is analytic in kk. The simplicity of c(k)c(k) is due to the fact Ind(𝐅(k,),Ω)=1\big{(}\mathbf{F}(k,\cdot),\Omega\big{)}=1 for any sufficiently small neighborhood Ω\Omega of c(k)c(k) in the continuation procedure. For any cU([h,0])c\in\mathbb{R}\setminus U([-h,0]), we have 𝐅(k,c)\mathbf{F}(k,c)\in\mathbb{R} and cR𝐅R(k,c)=c𝐅(k,c)0\partial_{c_{R}}\mathbf{F}_{R}(k,c)=\partial_{c}\mathbf{F}(k,c)\neq 0. Therefore if c(k1)U([h,0])c(k_{1})\in\mathbb{R}\setminus U([-h,0]) for some k1k_{1} along the continuation curve, then the unique extension c(k)c(k) coincides with the (real) root of 𝐅R(k,cR)\mathbf{F}_{R}(k,c_{R}) obtained by applying the Implicit Function Theorem to the real function 𝐅R(k,cR)\mathbf{F}_{R}(k,c_{R}). Hence c(k)c(k)\in\mathbb{R} if and only if c0c_{0}\in\mathbb{R}.

Let (k,k+)(k_{-},k_{+}) be the max interval of the continuation c(k)U([h,0])c(k)\subset\mathbb{C}\setminus U([-h,0]) as simple roots of 𝐅(k,)\mathbf{F}(k,\cdot) and we shall prove statement (3). Suppose k>k_{-}>-\infty, while the other case k+<+k_{+}<+\infty can be analyzed similarly. As k(k)+k\to(k_{-})+, the solution curve c(k)c(k) is bounded due to Lemma 4.2(2). Therefore there exists a sequence (kj)j=1(k,k+)(k_{j})_{j=1}^{\infty}\subset(k_{-},k_{+}) such that limjkj=k\lim_{j\to\infty}k_{j}=k_{-} and c=limjc(kj)c_{-}=\lim_{j\to\infty}c(k_{j})\in\mathbb{C} exists. Statement (2) implies that c(k)c(k) stays in the closure of either the upper or lower half of \mathbb{C} and thus 𝐅(k,c)=0\mathbf{F}(k_{-},c_{-})=0. Assume statement (3)(a) does not hold, then such a subsequence can be chosen such that cU([h,0])c_{-}\notin U([-h,0]). Therefore cc_{-} is a root in the domain of analyticity of 𝐅(k,)\mathbf{F}(k_{-},\cdot). Clearly cc_{-} is not a simple zero of 𝐅(k,)\mathbf{F}(k_{-},\cdot), otherwise c(k)c(k) can be extended beyond kk_{-}. Recall cc_{-} has to be an isolated root of 𝐅(k,)\mathbf{F}(k_{-},\cdot) since all roots of non-trivial analytic functions are isolated. Therefore, there exists a small neighborhood Ω\Omega of cc_{-} such that, for any kkk\geq k_{-} sufficiently close to kk_{-}, it hold Ind(𝐅(k,),Ω)2\big{(}\mathbf{F}(k,\cdot),\Omega\big{)}\geq 2. Consequently, for each kjk_{j} close to kk_{-}, there exists at least another root cc of 𝐅(kj,)\mathbf{F}(k_{j},\cdot) in Ω\Omega and thus (3)(b) holds. ∎

The semicircle theorem of Yih [39] states that all imaginary roots cc of 𝐅(k,)\mathbf{F}(k,\cdot) are contained in the circle with the diameter segment U([h,0])U([-h,0]), so the only possibility for the branches c±(k)c^{\pm}(k) of simple roots of 𝐅(k,)\mathbf{F}(k,\cdot) obtained in Lemma 4.2 can not be extended for all kk\in\mathbb{R} is when they reaches U(0)U(0) or U(h)U(-h), respectively. As a corollary of 𝐅(k,U(0))0\mathbf{F}\big{(}k,U(0)\big{)}\neq 0 and we have

Corollary 4.3.1.

(1) The branch c+(k)c^{+}(k) can be extended for all kk\in\mathbb{R}. Moreover c+(k)c^{+}(k)\in\mathbb{R} is even in kk, cF(k,c+(k))>0\partial_{c}F\big{(}k,c^{+}(k)\big{)}>0, and c+(k)>U(0)+ρ0c^{+}(k)>U(0)+\rho_{0} for all kk\in\mathbb{R}, for some ρ0>0\rho_{0}>0 independent of kk.
(2) If 𝐅(k,U(h))0\mathbf{F}(k,U(-h))\neq 0 for all kk\in\mathbb{R}, then c(k)c^{-}(k) of simple roots of 𝐅(k,)\mathbf{F}(k,\cdot) obtained in can also be extended for all kk\in\mathbb{R}. Moreover c(k)c^{-}(k)\in\mathbb{R} is even in kk, cF(k,c(k))<0\partial_{c}F\big{(}k,c^{-}(k)\big{)}<0, and c(k)<U(h)ρ0c^{-}(k)<U(-h)-\rho_{0} for all kk\in\mathbb{R}, for some ρ0>0\rho_{0}>0 independent of kk.

Proof.

Let k0k_{0} be given in Lemma 4.2(3) and we only need to focus on |k|k0|k|\leq k_{0}. We may assume k0k_{0} is sufficiently large such that c+(k0)>U(0)c^{+}(k_{0})>U(0) and c(k0)<U(h)c^{-}(k_{0})<U(-h). From Lemma 4.2(2), there exists R>0R>0 such that 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all k[k0,k0]k\in[-k_{0},k_{0}] and |c|R|c|\geq R. Hence c+(k0)(U(0),R)c^{+}(k_{0})\in(U(0),R) and c(k0)(R,U(h))c^{-}(k_{0})\in(-R,U(-h)) are the only roots of 𝐅(±k0,)\mathbf{F}(\pm k_{0},\cdot), which are also simple with ±cF(k0,c±(k0))>0\pm\partial_{c}F(k_{0},c^{\pm}(k_{0}))>0.

We first consider c+(k)c^{+}(k). Let

Ω={ccR(U(0),R),cI(1,1)}.\Omega=\{c\in\mathbb{C}\mid c_{R}\in(U(0),R),\,c_{I}\in(-1,1)\}.

According to Lemma 4.1(2), 𝐅(k,U(0))0\mathbf{F}(k,U(0))\neq 0 for any kk. Hence the semicircle theorem and the choice of RR imply that a.) c+(k0)Ωc^{+}(k_{0})\in\Omega and b.) 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all |k|k0|k|\leq k_{0} and cΩc\in\partial\Omega, and thus

Ind(𝐅(k,),Ω)=Ind(𝐅(k0,),Ω)=1,|k|k0.\text{Ind}(\mathbf{F}(k,\cdot),\Omega)=\text{Ind}(\mathbf{F}(k_{0},\cdot),\Omega)=1,\forall|k|\leq k_{0}.

Therefore none of the possibilities in Lemma 4.3(3ab) can happen to the extension c+(k)Ωc^{+}(k)\in\Omega starting from k=k0k=k_{0}, so this branch of simple root of 𝐅(k,c)˙\mathbf{F}(k,c\dot{)} can be uniquely extended for all k[k0,k0]k\in[-k_{0},k_{0}] with c+(k)(U(0),R)c^{+}(k)\in(U(0),R) as the only root of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω\Omega. The value of this extension at k=k0k=-k_{0} has to coincide with c+(k0)=c+(k0)c^{+}(-k_{0})=c^{+}(k_{0}) as c±(k0)c^{\pm}(-k_{0}) are the only roots of 𝐅(k0,)\mathbf{F}(-k_{0},\cdot) while c(k0)<U(h)c^{-}(-k_{0})<U(-h). Therefore the extensions starting from c+(±k0)c^{+}(\pm k_{0}) have to coincide. The evenness of c+(k)c^{+}(k) in k[k0,k0]k\in[-k_{0},k_{0}] follows from that of 𝐅\mathbf{F} and the uniqueness of its root in Ω\Omega. The sign of cF(k,c+(k))\partial_{c}F\big{(}k,c^{+}(k)\big{)} remains positive from k=k0k=k_{0} as c+(k)c^{+}(k) is always simple. The existence of ρ0>0\rho_{0}>0 is simple due to the continuity of 𝐅\mathbf{F}. The same argument applies to c(k)c^{-}(k) under the assumption 𝐅(k,U(h))0\mathbf{F}(k,U(-h))\neq 0 all for kk. The proof is complete. ∎

Based on the above analysis, we shall conclude that ikc±(k)-ikc^{\pm}(k) are the only eigenvalues of the linearized capillary gravity wave under the additonal assumption of the absence of singular modes

(4.9) 𝐅(k,U(x2))0,k𝐊,x2[h,0],\mathbf{F}(k,U(x_{2}))\neq 0,\quad\forall k\in\mathbf{K},\;x_{2}\in[-h,0],

where 𝐊=\mathbf{K}=\mathbb{R} or 2πL\frac{2\pi}{L}\mathbb{N} and LL is the period of the water wave in the x1x_{1} direction.

Proposition 4.4.

Assume UC3U\in C^{3} and (4.9) for 𝐊=\mathbf{K}=\mathbb{R} or 2πL\frac{2\pi}{L}\mathbb{N}, then there exists ρ>0\rho>0 such that

  1. (1)

    F0inf{(1+k2)12ehμ|𝐅(k,c)|k𝐊,cR[U(h)ρ,U(0)+ρ],cI[ρ,ρ]}>0F_{0}\triangleq\inf\{(1+k^{2})^{-\frac{1}{2}}e^{-\frac{h}{\mu}}|\mathbf{F}(k,c)|\mid k\in\mathbf{K},\,c_{R}\in[U(-h)-\rho,U(0)+\rho],\,c_{I}\in[-\rho,\rho]\}>0.

  2. (2)

    Assume 𝐊=\mathbf{K}=\mathbb{R}, then {c𝐅(k,c)=0}={c±(k)}\{c\mid\mathbf{F}(k,c)=0\}=\{c^{\pm}(k)\}.

Proof.

The first statement is a direct corollary of the continuity of 𝐅\mathbf{F}, its analyticity outside U([h,0])U([-h,0]), assumption (4.9), and Lemma 4.2.

Let us consider statement (2). Corollary 4.3.1 and (4.9) imply that both c+(k)(U(0),+)c^{+}(k)\in(U(0),+\infty) and c(k)(,U(h))c^{-}(k)\in(-\infty,U(-h)) can be extended as even analytic functions of kk\in\mathbb{R}. Let k0,R>0k_{0},R>0 be taken as in the proof of Corollary 4.3.1 and we only need to focus on |k|k0|k|\leq k_{0}. Assumption (4.9) also yields ρ>0\rho>0 such that

𝐅(k,c)0,dist(c,U([h,0]))=ρ,|k|k0.\mathbf{F}(k,c)\neq 0,\;\;\forall dist\big{(}c,U([-h,0])\big{)}=\rho,\;|k|\leq k_{0}.

Let

Ω={c|c|<R,dist(c,U([h,0]))>ρ},\Omega=\{c\in\mathbb{C}\mid|c|<R,\,dist\big{(}c,U([-h,0])\big{)}>\rho\},

then we have 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all |k|k0|k|\leq k_{0} and cΩc\in\partial\Omega. Therefore

Ind(𝐅(k,),Ω)=Ind(𝐅(k0,),Ω)=2,|k|k0,\text{Ind}(\mathbf{F}(k,\cdot),\Omega)=\text{Ind}(\mathbf{F}(k_{0},\cdot),\Omega)=2,\forall|k|\leq k_{0},

and 𝐅(k,)\mathbf{F}(k,\cdot) does not have any other roots. ∎

In order to obtain a more complete picture of the eigenvalue distribution we shall derive some sign properties in the following lemma, where FF and YY are viewed as function of cc and K=k20K=k^{2}\geq 0. According to Lemma 3.19(1), FF is well-defined for cc in a neighborhood of U((h,0))\mathbb{R}\setminus U\big{(}(-h,0)\big{)}.

Lemma 4.5.

Assume UC3U\in C^{3}, then we have

K2(F(K,c))<0,k,cU((h,0]),\partial_{K}^{2}\big{(}F\big{(}\sqrt{K},c\big{)}\big{)}<0,\;\ \forall k\in\mathbb{R},\;c\in\mathbb{R}\setminus U\big{(}(-h,0]\big{)},
KF(0,c)<σ+h0(U(x2)c)2dx2,cU([h,0]),\partial_{K}F(0,c)<-\sigma+\int_{-h}^{0}\big{(}U(x_{2})-c\big{)}^{2}dx_{2},\quad\forall c\in\mathbb{R}\setminus U([-h,0]),
KF(0,U(h))=σ+h0(U(x2)U(h))2dx2.\partial_{K}F\big{(}0,U(-h)\big{)}=-\sigma+\int_{-h}^{0}\big{(}U(x_{2})-U(-h)\big{)}^{2}dx_{2}.
Proof.

For K0K\geq 0 and cc\in\mathbb{C} with y(k,c,0)0y_{-}(k,c,0)\neq 0 and cI0c_{I}\geq 0, let

(4.10) =(K,c)=x22+K+U(x2)U(x2)c,y~(K,c,x2)=y(K,c,x2)y(K,c,0),x2[h,0],\mathcal{R}=\mathcal{R}(K,c)=-\partial_{x_{2}}^{2}+K+\frac{U^{\prime\prime}(x_{2})}{U(x_{2})-c},\;\;\tilde{y}(K,c,x_{2})=\frac{y_{-}(\sqrt{K},c,x_{2})}{y_{-}(\sqrt{K},c,0)},\quad\;x_{2}\in[-h,0],

be the differential operator in the Rayleigh equation (3.1) and the normalization of the fundamental solution yy_{-} defined in (3.53) and (3.83). Clearly

y~(h)=0,y~(h)=y(K,c,0)1,y~(x2)>0,x2(h,0),y~(0)=1,Y(K,c)=y~(0),\tilde{y}(-h)=0,\quad\tilde{y}^{\prime}(-h)=y_{-}(\sqrt{K},c,0)^{-1},\quad\tilde{y}(x_{2})>0,\;x_{2}\in(-h,0),\quad\tilde{y}(0)=1,\quad Y(\sqrt{K},c)=\tilde{y}^{\prime}(0),

where the sign properties follows from Lemma 3.19(1). It is straight forward to compute, for cU((h,0))c\in\mathbb{R}\setminus U\big{(}(-h,0)\big{)} and x2(h,0)x_{2}\in(-h,0),

Ky~=y~<0,K2y~=2Ky~,\mathcal{R}\partial_{K}\tilde{y}=-\tilde{y}<0,\quad\mathcal{R}\partial_{K}^{2}\tilde{y}=-2\partial_{K}\tilde{y},

where the smoothness of y~\tilde{y} in KK is ensured by Lemma 3.10. The following claim is used to analyze these and some other functions.

Claim. Suppose yC2((h,0))C0([h,0])y\in C^{2}((-h,0))\cap C^{0}([-h,0]) is a solution to (y)(x2)=f(x2)(\mathcal{R}y)(x_{2})=f(x_{2}) and y(h)=y(0)=0y(-h)=y(0)=0 with cU((h,0))c\in\mathbb{R}\setminus U\big{(}(-h,0)\big{)}, where ff is C0C^{0} on [h,0][-h,0], then we have the following through direct computations

(4.11) (y~yy~y)=y~fy(0)=h0y~fdx2,y(x2)=y~(x2)x201y~(x2)2hx2y~(x2)f(x2)dx2dx2.(\tilde{y}^{\prime}y-\tilde{y}y^{\prime})^{\prime}=\tilde{y}f\Rightarrow y^{\prime}(0)=-\int_{-h}^{0}\tilde{y}fdx_{2},\;\;y(x_{2})=\tilde{y}(x_{2})\int_{x_{2}}^{0}\frac{1}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\tilde{y}(x_{2}^{\prime\prime})f(x_{2}^{\prime\prime})dx_{2}^{\prime\prime}dx_{2}^{\prime}.

Applying this claim to Ky~\partial_{K}\tilde{y} and KKy~\partial_{KK}\tilde{y} implies, for cU((h,0])c\in\mathbb{R}\setminus U\big{(}(-h,0]\big{)},

(4.12) KY=Ky~(0)=h0y~2dx2>0,K2Y=2h0y~(x2)2x20y~(x2)2hx2y~(x2)2dx2dx2dx2<0.\begin{split}&\partial_{K}Y=\partial_{K}\tilde{y}^{\prime}(0)=\int_{-h}^{0}\tilde{y}^{2}dx_{2}>0,\\ &\partial_{K}^{2}Y=-2\int_{-h}^{0}\tilde{y}(x_{2})^{2}\int_{x_{2}}^{0}\tilde{y}(x_{2}^{\prime})^{-2}\int_{-h}^{x_{2}^{\prime}}\tilde{y}(x_{2}^{\prime\prime})^{2}dx_{2}^{\prime\prime}dx_{2}^{\prime}dx_{2}<0.\end{split}

The definition of FF implies K2F<0\partial_{K}^{2}F<0.

For k=0k=0, through direct calculation, one may verify, for cU([h,0])c\notin U([-h,0]),

(4.13) y(0,c,x2)=(U(x2)c)hx2U(h)c(U(x2)c)2dx2.y_{-}(0,c,x_{2})=(U(x_{2})-c)\int_{-h}^{x_{2}}\frac{U(-h)-c}{(U(x_{2}^{\prime})-c)^{2}}dx_{2}^{\prime}.

For cU([h,0])c\in\mathbb{R}\setminus U([-h,0]), from (4.12), we have

KY(0,c)=\displaystyle\partial_{K}Y(0,c)= h0y~2dx2=h0(Uc)2(U(0)c)2(hx2dx2(U(x2)c)2)2dx2(h0dx2(U(x2)c)2)2,\displaystyle\int_{-h}^{0}\tilde{y}^{2}dx_{2}=\int_{-h}^{0}\frac{(U-c)^{2}}{(U(0)-c)^{2}}\Big{(}\int_{-h}^{x_{2}}\frac{dx_{2}^{\prime}}{(U(x_{2}^{\prime})-c)^{2}}\Big{)}^{2}dx_{2}\Big{(}\int_{-h}^{0}\frac{dx_{2}^{\prime}}{(U(x_{2}^{\prime})-c)^{2}}\Big{)}^{-2},

and thus

(4.14) KF(0,c)=(U(0)c)2KY(0,c)σ=h0(Uc)2(hx2dx2(U(x2)c)2)2dx2(h0dx2(U(x2)c)2)2σ<h0(Uc)2dx2σ.\begin{split}\partial_{K}F(0,c)=&(U(0)-c)^{2}\partial_{K}Y(0,c)-\sigma\\ =&\int_{-h}^{0}(U-c)^{2}\Big{(}\int_{-h}^{x_{2}}\frac{dx_{2}^{\prime}}{(U(x_{2}^{\prime})-c)^{2}}\Big{)}^{2}dx_{2}\Big{(}\int_{-h}^{0}\frac{dx_{2}^{\prime}}{(U(x_{2}^{\prime})-c)^{2}}\Big{)}^{-2}-\sigma\\ <&\int_{-h}^{0}(U-c)^{2}dx_{2}-\sigma.\end{split}

For k=0k=0 and c=U(h)c=U(-h), we can use (3.107) to compute

(4.15) y~(0,U(h),x2)=(U(x2)U(h))/(U(0)U(h)).\tilde{y}\big{(}0,U(-h),x_{2}\big{)}=\big{(}U(x_{2})-U(-h)\big{)}/\big{(}U(0)-U(-h)\big{)}.

Consequently, one obtains explicitly

KY(0,U(h))=h0(U(x2)U(h))2(U(0)U(h))2dx2,\partial_{K}Y\big{(}0,U(-h)\big{)}=\int_{-h}^{0}\frac{\big{(}U(x_{2})-U(-h)\big{)}^{2}}{\big{(}U(0)-U(-h)\big{)}^{2}}dx_{2},

which in turn yields the desired formula of KF(0,U(h))\partial_{K}F\big{(}0,U(-h)\big{)}. ∎

The information on the derivatives of FF leads to the following properties of the roots of FF.

Lemma 4.6.

Assume UC3U\in C^{3}, the following hold.

  1. (1)

    If

    (4.16) σh0(U(x2)U(h))2dx2KF(0,U(h))0,\sigma\geq\int_{-h}^{0}\big{(}U(x_{2})-U(-h)\big{)}^{2}dx_{2}\Longleftrightarrow\partial_{K}F\big{(}0,U(-h)\big{)}\leq 0,

    then F(k,U(h))g=F(0,U(h))F\big{(}k,U(-h)\big{)}\leq-g=F(0,U(-h)) for all kk\in\mathbb{R}.

  2. (2)

    Let

    g#=\displaystyle g_{\#}= max{Y(k,U(h))(U(0)U(h))2U(0)(U(0)U(h))σk2k}\displaystyle\max\big{\{}Y\big{(}k,U(-h)\big{)}\big{(}U(0)-U(-h)\big{)}^{2}-U^{\prime}(0)\big{(}U(0)-U(-h)\big{)}-\sigma k^{2}\mid k\in\mathbb{R}\big{\}}
    =\displaystyle= max{F(k,U(h))+gk},\displaystyle\max\big{\{}F\big{(}k,U(-h)\big{)}+g\mid k\in\mathbb{R}\big{\}},

    then we have

    1. (a)

      g#F(0,U(h))+g=0g_{\#}\geq F\big{(}0,U(-h)\big{)}+g=0 and “=” in the “\leq" holds if and only if (4.16) holds.

    2. (b)

      If g>g#g>g_{\#}, then F(k,U(h))<0F\big{(}k,U(-h)\big{)}<0 for all kk\in\mathbb{R}.

    3. (c)

      If 0<g=g#0<g=g_{\#} , then there exists a unique k#>0k_{\#}>0 such that F(±k#,U(h))=0F\big{(}\pm k_{\#},U(-h)\big{)}=0 and F(k,U(h))<0F\big{(}k,U(-h)\big{)}<0 for all |k|k#|k|\neq k_{\#}.

    4. (d)

      If 0<g<g#0<g<g_{\#}, then there exist k#+>k#>0k_{\#}^{+}>k_{\#}^{-}>0 such that

      F(k,U(h))<0,|k|(k#,k#+);F(k,U(h))>0,|k|(k#,k#+);kF(k#±,U(h))>0.F\big{(}k,U(-h)\big{)}<0,\;\;|k|\notin(k_{\#}^{-},k_{\#}^{+});\quad F\big{(}k,U(-h)\big{)}>0,\;\;|k|\in(k_{\#}^{-},k_{\#}^{+});\quad\mp\partial_{k}F(k_{\#}^{\pm},U(-h))>0.
Proof.

Statement (1) is a direct consequence of the concavity of F(k,U(h))F\big{(}k,U(-h)\big{)} in K=k2K=k^{2} and F(0,U(h))=g<0F\big{(}0,U(-h)\big{)}=-g<0. Statement (2) is also an immediate implication of this concavity and Lemma 4.2(1). ∎

Along with statement (2b ) and Corollary 4.3.1, (4.16) provides an explicit sufficient condition ensuring that the branch c(k)c^{-}(k) does not reach U([h,0])U([-h,0]) and thus staying in (,U(h))(-\infty,U(-h)) for all kk\in\mathbb{R}.

To end this subsection we prove the following monotonicity of the even functions c±(k)c^{\pm}(k) which will be used in obtaining the conjugacy between the irrotational linearized capillary gravity water waves and the component of the solutions linearized at the shear U(x2)U(x_{2}). From the definition of FF and (4.13), we first compute, for cU([h,0])c\notin U([-h,0]),

Y(0,c)=U(0)h0(Uc)2dx2+(U(0)c)1(U(0)c)h0(Uc)2dx2Y(0,c)=\frac{U^{\prime}(0)\int_{-h}^{0}(U-c)^{-2}dx_{2}+(U(0)-c)^{-1}}{(U(0)-c)\int_{-h}^{0}(U-c)^{-2}dx_{2}}

and thus

F(0,c)=(U(0)c)2Y(0,c)U(0)(U(0)c)g=1h0(Uc)2dx2g,F(0,c)=(U(0)-c)^{2}Y(0,c)-U^{\prime}(0)(U(0)-c)-g=\frac{1}{\int_{-h}^{0}(U-c)^{-2}dx_{2}}-g,

which is uniformly increasing on (,U(h))(-\infty,U(-h)) and uniformly decreasing on (U(0),+)(U(0),+\infty). Therefore F(0,)F(0,\cdot) has two real roots

(4.17) c0+(U(0),+),c0(,U(h)), s. t. F(0,c0±)=1h0(Uc0±)2dx2g=0,c_{0}^{+}\in(U(0),+\infty),\;\;c_{0}^{-}\in(-\infty,U(-h)),\;\text{ s. t. }\;F(0,c_{0}^{\pm})=\frac{1}{\int_{-h}^{0}(U-c_{0}^{\pm})^{-2}dx_{2}}-g=0,

which are unique in the above intervals.

Lemma 4.7.

Assume UC3U\in C^{3}, then the following hold.

  1. (1)

    For {+,}\dagger\in\{+,-\}, suppose c(k)U([h,0])c^{\dagger}(k)\in\mathbb{R}\setminus U([-h,0]) can be extended as simple roots of F(k,)F(k,\cdot) for all kk0k\geq k_{*}\geq 0, then (c)(k)=0(c^{\dagger})^{\prime}(k)=0 has most one solution on (k,+)(k_{*},+\infty), where (c)(k)0(c^{\dagger})^{\prime\prime}(k)\neq 0 is also satisfied.

  2. (2)

    For {+,}\dagger\in\{+,-\}, suppose c(k)U([h,0])c^{\dagger}(k)\in\mathbb{R}\setminus U([-h,0]) can be extended as simple roots of F(k,)F(k,\cdot) for all kk\in\mathbb{R}, then (c)(k)0(c^{\dagger})^{\prime}(k)\neq 0 for all k0k\neq 0 if and only if

    (4.18) σg2h0(Uc0)2(hx2dx2(U(x2)c0)2)2dx2,\sigma\geq g^{2}\int_{-h}^{0}(U-c_{0}^{\dagger})^{2}\Big{(}\int_{-h}^{x_{2}}\frac{dx_{2}^{\prime}}{(U(x_{2}^{\prime})-c_{0}^{\dagger})^{2}}\Big{)}^{2}dx_{2},

    with c0c_{0}^{\dagger} defined in (4.17).

Proof.

We shall work with c(k)c^{-}(k), while the same proof works for c+(k)c^{+}(k). Suppose there exists k0>k0k_{0}>k_{*}\geq 0 such that (c)(k0)=0(c^{-})^{\prime}(k_{0})=0, then

2k0(KF)(k0,c(k0))=kF(k0,c(k0))=cF(k0,c(k0))(c)(k0)=0.2k_{0}(\partial_{K}F)(k_{0},c^{-}(k_{0}))=\partial_{k}F(k_{0},c^{-}(k_{0}))=-\partial_{c}F(k_{0},c^{-}(k_{0}))(c^{-})^{\prime}(k_{0})=0.

Computing the second order derivative at k0k_{0}, we have

(c)(k0)=k2FcF|(k0,c(k0))=4k02(K2F)+2(KF)cF|(k0,c(k0))=4k02(K2F)cF|(k0,c(k0)),(c^{-})^{\prime\prime}(k_{0})=-\frac{\partial_{k}^{2}F}{\partial_{c}F}\Big{|}_{(k_{0},c^{-}(k_{0}))}=-\frac{4k_{0}^{2}(\partial_{K}^{2}F)+2(\partial_{K}F)}{\partial_{c}F}\Big{|}_{(k_{0},c^{-}(k_{0}))}=-\frac{4k_{0}^{2}(\partial_{K}^{2}F)}{\partial_{c}F}\Big{|}_{(k_{0},c^{-}(k_{0}))},

which along with Lemma 4.5 and cF(k,c(k0))<0\partial_{c}F(k,c^{-}(k_{0}))<0 (Corollary 4.3.1) implies (c)(k0)<0(c^{-})^{\prime\prime}(k_{0})<0. Hence k0>kk_{0}>k_{*} has to be the only positive critical point of c(k)c^{-}(k).

To prove Statement (2) where k=0k_{*}=0, on the one hand, we first observe that since c0c_{0}^{-} is the unique root of F(0,)F(0,\cdot) in (,U(h))(-\infty,U(-h)) and c(0)c^{-}(0) is also such a root, so c(0)=c0c^{-}(0)=c_{0}^{-}. Moreover, (4.14) implies that (4.18) is equivalent to KF(0,c(0))0\partial_{K}F(0,c^{-}(0))\leq 0. On the other hand, observe that the evenness of c(k)c^{-}(k) yields (c)(0)=0(c^{-})^{\prime}(0)=0. One may compute

(KF)(0,c(0))=k2F(0,c(0))/2=cF(0,c(0))((c)(0))/2.(\partial_{K}F)(0,c^{-}(0))=\partial_{k}^{2}F(0,c^{-}(0))/2=-\partial_{c}F(0,c^{-}(0))\big{(}(c^{-})^{\prime\prime}(0)\big{)}/2.

From Lemma 4.2(3), (c)(k)<0(c^{-})^{\prime}(k)<0 for some k1k\gg 1. Hence, on the one hand, if (4.18) does not hold, then cF(0,c(0))<0\partial_{c}F(0,c^{-}(0))<0 (Lemma 4.2(3) and 4.3) and the above identity imply (c)(0)>0(c^{-})^{\prime\prime}(0)>0. Along with (c)(0)=0(c^{-})^{\prime}(0)=0, it yields that cc^{-} has a critical point k0>0k_{0}>0. On the other hand, through the same argument, (4.18) yields (c)(0)0(c^{-})^{\prime\prime}(0)\leq 0 while (c)(0)=0(c^{-})^{\prime}(0)=0. Therefore, if KF(0,c(0))<0\partial_{K}F(0,c^{-}(0))<0 which implies (c)(0)<0(c-)^{\prime\prime}(0)<0, it is impossible that there exists a unique critical point of cc^{-} where (c)<0(c^{-})^{\prime\prime}<0. In the borderline case of KF(0,c(0))=0\partial_{K}F(0,c^{-}(0))=0 which implies (c)(0)<0(c-)^{\prime\prime}(0)<0, further Taylor expansions of the even-in-kk functions F(k,c)F(k,c) and c(k)c^{-}(k) yields

k4c(0)=12K2F(0,c(0))/cF(0,c(0))<0.\partial_{k}^{4}c^{-}(0)=-12\partial_{K}^{2}F(0,c^{-}(0))/\partial_{c}F(0,c^{-}(0))<0.

From the same reasoning, we obtain that (c)0(c^{-})^{\prime}\neq 0 for k>0k>0. The proof of the lemma is complete. ∎

4.2. Eigenvalue distribution of convex/concave shear flow UU

To analyze eigenvalues under less implicit assumptions than (4.9), particularly the generation of unstable modes from c=U(h)c=U(-h), we further assume U0U^{\prime\prime}\neq 0 on [h,0][-h,0]. Due to Lemma 4.1(6), this rules out the possibility of roots of 𝐅\mathbf{F} on U((h,0])U\big{(}(-h,0]) and provides better smoothness of FF for the bifurcation analysis.

Lemma 4.8.

Assume UCl0U\in C^{l_{0}}, l05l_{0}\geq 5, and U0U^{\prime\prime}\neq 0 on [h,0][-h,0], then F(k,c)F(k,c) is well defined for all kk\in\mathbb{R} and cc\in\mathbb{C} and
a.) FF is analytic in both kk\in\mathbb{R} and cU([h,0])c\notin U([-h,0]) and, when restricted to cI0c_{I}\geq 0, is Cl02C^{l_{0}-2} in both kk\in\mathbb{R} and c{U(h),U(0)}c\notin\{U(-h),U(0)\},
b.) kj1cj2F\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}F is locally CαC^{\alpha} in both kk and cU(0)c\neq U(0) with cI0c_{I}\geq 0 for any α[0,1)\alpha\in[0,1), j2=0,1j_{2}=0,1, and 0j1l04j20\leq j_{1}\leq l_{0}-4-j_{2},
c.) FF is C1C^{1} in kk and cc with cI0c_{I}\geq 0.

Remark 4.2.

Note that, in the above statement, for fixed cU([h,0))c\in U([-h,0)), FF is Cl02C^{l_{0}-2} in kk. This stronger regularity in kk follows from that of (y0,y0)(y_{0-},y_{0-}^{\prime}) and YY (see Lemmas 3.12 and 3.24). Moreover, one could prove that FF and kF\partial_{k}F are also C1,αC^{1,\alpha} near c=U(0)c=U(0) with cI0c_{I}\geq 0 by estimating cR2YI(k,c)=O(|cU(0)|1)\partial_{c_{R}}^{2}Y_{I}(k,c)=O\big{(}|c-U(0)|^{-1}\big{)} using Lemmas 3.143.16 and 3.24 as well as Corollary 3.24.1.

Proof.

The assumption U0U^{\prime\prime}\neq 0 implies that y(k,c,0)0y_{-}(k,c,0)\neq 0 for all kk and cc (Lemma 3.19(5)) and thus FF is well defined. The analyticity and the Cl02C^{l_{0}-2} and C1,αC^{1,\alpha} (restricted to cI0c_{I}\geq 0 for the latter two) regularity of FF follow directly from those of YY given in Lemma 3.24 except at c=U(0)c=U(0). Near cU(0)c\in U(0), the regularity and estimates on YY (Lemma 3.20, 3.23, 3.24) and cY\partial_{c}Y (Lemma 3.23 and Corollary 3.24.1) yield the regularity of FF. ∎

As a corollary of the Lemmas 4.3, 4.6 and 4.8 and the semicircle theorem, we obtain a sufficient condition for (4.9) to hold for 𝐊=\mathbf{K}=\mathbb{R}.

Corollary 4.8.1.

Suppose U0U^{\prime\prime}\neq 0 on [h,0][-h,0] and (4.16) hold, then (4.9) is true for all kk\in\mathbb{R}.

Assuming U0U^{\prime\prime}\neq 0, in general c=U(h)c=U(-h) is the only point outside the domain of analyticity of F(k,)F(k,\cdot) which might happen to be a root and also might be the end point of branches of roots of F(k,)F(k,\cdot), it is a crucial step to analyze zeros of FF around U(h)U(-h).

Lemma 4.9.

Assume UC5U\in C^{5}, then (a) cF(k,U(h))<0\partial_{c}F(k,U(-h))<0 for all kk\in\mathbb{R} if U>0U^{\prime\prime}>0 on [h,0][-h,0]; and (b) if U<0U^{\prime\prime}<0 on [h,0][-h,0], then cF(k,U(h))<0\partial_{c}F(k,U(-h))<0 if F(k,U(h))=0F(k,U(-h))=0, where FF is understood as restricted to cI0c_{I}\geq 0.

Proof.

We shall use the notations \mathcal{R} and y~\tilde{y} defined in the proof of Lemma 4.5 and FF and YY are also viewed as function of cc and K=k20K=k^{2}\geq 0. It is straight forward to compute, for c<U(h)c<U(-h) and x2(h,0)x_{2}\in(-h,0),

cy~=U(Uc)2y~,Kcy~=cy~U(Uc)2Ky~.\mathcal{R}\partial_{c}\tilde{y}=-\tfrac{U^{\prime\prime}}{(U-c)^{2}}\tilde{y},\quad\mathcal{R}\partial_{Kc}\tilde{y}=-\partial_{c}\tilde{y}-\tfrac{U^{\prime\prime}}{(U-c)^{2}}\partial_{K}\tilde{y}.

Applying (4.11) we obtain

(4.19) U(0)cY=U(0)cy~(0)=U(0)h0Uy~2(Uc)2dx2>0,U(0)KcY<0,c<U(h).\begin{split}&U^{\prime\prime}(0)\partial_{c}Y=U^{\prime\prime}(0)\partial_{c}\tilde{y}^{\prime}(0)=U^{\prime\prime}(0)\int_{-h}^{0}\frac{U^{\prime\prime}\tilde{y}^{2}}{\big{(}U-c\big{)}^{2}}dx_{2}>0,\;\;U^{\prime\prime}(0)\partial_{Kc}Y<0,\quad\forall c<U(-h).\end{split}

These integral representation of cY\partial_{c}Y still holds as cU(h)c\to U(-h)-, and thus also its sign. For k=0k=0 and c=U(h)c=U(-h), we can use (4.15) to compute

cY(0,U(h))=U(0)U(h)(U(0)U(h))2cF(0,U(h))=U(h)<0.\partial_{c}Y\big{(}0,U(-h)\big{)}=\frac{U^{\prime}(0)-U^{\prime}(-h)}{\big{(}U(0)-U(-h)\big{)}^{2}}\implies\partial_{c}F\big{(}0,U(-h)\big{)}=-U^{\prime}(-h)<0.

Finally we obtain the sign of cF(k,U(h))\partial_{c}F(k,U(-h)) in two cases separately, based on the sign of UU^{\prime\prime}. Suppose U>0U^{\prime\prime}>0. The above (4.12) and (4.19) implies that, for cU(h)c\leq U(-h), Y(K,c)Y(\sqrt{K},c) is strictly increasing in KK and cY(K,c)\partial_{c}Y(\sqrt{K},c) is strictly deceasing in KK, and thus

cF=(U(0)c)2cY2(U(0)c)Y+U(0)\partial_{c}F=(U(0)-c)^{2}\partial_{c}Y-2(U(0)-c)Y+U^{\prime}(0)

is also strictly decreasing in KK. Letting cU(h)c\to U(-h)-, this monotonicity yields

cF(k,U(h))cF(0,U(h))=U(h)<0.\partial_{c}F(k,U(-h))\leq\partial_{c}F(0,U(-h))=-U^{\prime}(-h)<0.

In the other case of U<0U^{\prime\prime}<0, suppose F(k,U(h))=0F(k,U(-h))=0 for some kk\in\mathbb{R}, which implies

Y(k,U(h))=g+σk2(U(0)U(h))2+U(0)U(0)U(h).Y(k,U(-h))=\frac{g+\sigma k^{2}}{(U(0)-U(-h))^{2}}+\frac{U^{\prime}(0)}{U(0)-U(-h)}.

Therefore

cF(k,U(h))=\displaystyle\partial_{c}F(k,U(-h))= (U(0)U(h))2cY(k,U(h))2(U(0)U(h))Y(k,U(h))+U(0)\displaystyle(U(0)-U(-h))^{2}\partial_{c}Y(k,U(-h))-2(U(0)-U(-h))Y(k,U(-h))+U^{\prime}(0)
=\displaystyle= (U(0)U(h))2cY(k,U(h))U(0)2(g+σk2)/(U(0)U(h)).\displaystyle(U(0)-U(-h))^{2}\partial_{c}Y(k,U(-h))-U^{\prime}(0)-2(g+\sigma k^{2})/(U(0)-U(-h)).

We also have cY(k,U(h))<0\partial_{c}Y(k,U(-h))<0 from taking the limit of (4.19). Hence we obtain cF(k,U(h))<0\partial_{c}F(k,U(-h))<0 and the proof of the lemma is complete. ∎

In the next step we shall study the roots of F(k,)F(k,\cdot) near c=U(h)c=U(-h).

Lemma 4.10.

Assume UC5U\in C^{5}, and U0U^{\prime\prime}\neq 0 on [h,0][-h,0]. Suppose F(k0,U(h))=0F\big{(}k_{0},U(-h)\big{)}=0, then there exist ϵ>0\epsilon>0, ρ(0,U(0)U(h))\rho\in\big{(}0,U(0)-U(-h)\big{)}, and 𝒞C1,α([k0ϵ,k0+ϵ],)\mathcal{C}\in C^{1,\alpha}\big{(}[k_{0}-\epsilon,k_{0}+\epsilon],\mathbb{C}\big{)} for any α[0,1)\alpha\in[0,1) such that

𝒞(k0)=U(h),𝒞(k)U([h,0]),  0<|kk0|ϵ,U𝒞I(k)0,|kk0|ϵ,\mathcal{C}(k_{0})=U(-h),\;\;\mathcal{C}(k)\notin U([-h,0]),\;\;0<|k-k_{0}|\leq\epsilon,\quad U^{\prime\prime}\mathcal{C}_{I}(k)\geq 0,\;\;|k-k_{0}|\leq\epsilon,
cF(k,𝒞(k))0, if 𝒞I(k)0,\partial_{c}F(k,\mathcal{C}(k))\neq 0,\;\text{ if }\;\mathcal{C}_{I}(k)\geq 0,

and for cI[0,ρ]c_{I}\in[0,\rho] and |cRU(h)|ρ|c_{R}-U(-h)|\leq\rho,

F(k,c)=0 with k[k0ϵ,k0+ϵ], iff c=𝒞(k)=𝒞R(k)+i𝒞I(k).F(k,c)=0\text{ with }k\in[k_{0}-\epsilon,k_{0}+\epsilon],\;\text{ iff }c=\mathcal{C}(k)=\mathcal{C}_{R}(k)+i\mathcal{C}_{I}(k).

Moreover, without loss of generality assume k0>0k_{0}>0 (Lemma 4.8 implies k00k_{0}\neq 0) and this branch of roots of FF satisfies

  1. (1)

    If kF(k0,U(h))=0\partial_{k}F\big{(}k_{0},U(-h)\big{)}=0, then 𝒞(k0)=0\mathcal{C}^{\prime}(k_{0})=0, 𝒞I0\mathcal{C}_{I}\equiv 0 and 𝒞(k)<U(h)\mathcal{C}(k)<U(-h) for all 0<|kk0|ϵ0<|k-k_{0}|\leq\epsilon.

  2. (2)

    If ±kF(k0,U(h))>0\pm\partial_{k}F\big{(}k_{0},U(-h)\big{)}>0, then ±𝒞R(k0)>0\pm\mathcal{C}_{R}^{\prime}(k_{0})>0 and

    𝒞R(k)<U(h),𝒞I(k)=0, 0<±(k0k)ϵ,\mathcal{C}_{R}(k)<U(-h),\;\;\mathcal{C}_{I}(k)=0,\quad\forall\,0<\pm(k_{0}-k)\leq\epsilon,

    and for some C~>0\tilde{C}>0 determined by k0k_{0} and UU,

    𝒞R(k)>U(h),|𝒞I(k)YI(k,𝒞R(k))+((U(0)U(h))2cF(k0,U(h))|C~|kk0|α, 0<±(kk0)ϵ,\mathcal{C}_{R}(k)>U(-h),\;\;\left|\frac{\mathcal{C}_{I}(k)}{Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}}+\frac{\big{(}\big{(}U(0)-U(-h)\big{)}^{2}}{\partial_{c}F\big{(}k_{0},U(-h)\big{)}}\right|\leq\tilde{C}|k-k_{0}|^{\alpha},\quad\forall\,0<\pm(k-k_{0})\leq\epsilon,

    which implies

    0<|𝒞I(k)|C~(kk0)2,U(0)𝒞I(k)>0, 0<±(kk0)ϵ.0<|\mathcal{C}_{I}(k)|\leq\tilde{C}(k-k_{0})^{2},\;\;U^{\prime\prime}(0)\mathcal{C}_{I}(k)>0,\quad\forall\,0<\pm(k-k_{0})\leq\epsilon.

In the generic case kF(k0,U(h))0\partial_{k}F\big{(}k_{0},U(-h)\big{)}\neq 0, locally the roots of F(k,c)F(k,c) consists of the intersection of the graph of 𝒞(k)\mathcal{C}(k) and the closure of the upper half complex plane, along with its complex conjugate. In this case, however, one observes that d𝒞I(k)/d𝒞R(k)=0d\mathcal{C}_{I}(k)/d\mathcal{C}_{R}(k)=0 at k=k0k=k_{0} means 𝒞I\mathcal{C}_{I} is very weak when it is nonzero. The following proof is based on both the Implicit Function Theorem and the Intermediate Value Theorem.

Proof.

According to Lemma 4.8, FF is C1,αC^{1,\alpha} in kk and cc in the region cI0c_{I}\geq 0. As FIF_{I} is not continuous at cU((h,0])c\in U\big{(}(-h,0]\big{)}\subset\mathbb{C} in general, let F~(k,c)=F~R+iF~I\tilde{F}(k,c)=\tilde{F}_{R}+i\tilde{F}_{I}\in\mathbb{C} be a C1,αC^{1,\alpha} extension of FF into a neighborhood of (k0,U(h))×\big{(}k_{0},U(-h)\big{)}\in\mathbb{R}\times\mathbb{C} which coincides with FF for cI0c_{I}\geq 0. From Lemma 4.9, the 2×22\times 2 Jacobian matrix of DcF~D_{c}\tilde{F} satisfies

DcF~(k0,U(h))=(cRF~RcIF~RcRF~IcIF~I)|(k0,U(h))=cF(k0,U(h))I2×2,cF(k0,U(h))<0,D_{c}\tilde{F}\big{(}k_{0},U(-h)\big{)}=\begin{pmatrix}\partial_{c_{R}}\tilde{F}_{R}&\partial_{c_{I}}\tilde{F}_{R}\\ \partial_{c_{R}}\tilde{F}_{I}&\partial_{c_{I}}\tilde{F}_{I}\end{pmatrix}\Big{|}_{\big{(}k_{0},U(-h)\big{)}}=\partial_{c}F\big{(}k_{0},U(-h)\big{)}I_{2\times 2},\quad\partial_{c}F\big{(}k_{0},U(-h)\big{)}<0,

where we used the Cauchy-Riemann equation and the fact F(k,c)F(k,c)\in\mathbb{R} for all c<U(h)c<U(-h). Therefore the Implicit Function Theorem implies that all roots of F~(k,c)\tilde{F}(k,c) near (k0,U(h))\big{(}k_{0},U(-h)\big{)} form the graph of a C1,αC^{1,\alpha} complex-valued function 𝒞(k)\mathcal{C}(k) which contains (k0,U(h))\big{(}k_{0},U(-h)\big{)}. To complete the proof of the lemma, we only need to prove that 𝒞(k)\mathcal{C}(k) satisfies properties (1) and (2).

Firstly we prove 𝒞(k)\mathcal{C}(k)\in\mathbb{R} if 𝒞R(k)U(h)\mathcal{C}_{R}(k)\leq U(-h) and thus F(k,𝒞(k))=F~(k,𝒞(k))=0F\big{(}k,\mathcal{C}(k)\big{)}=\tilde{F}\big{(}k,\mathcal{C}(k)\big{)}=0 as well. When restricted to \mathbb{R}, FRC1F_{R}\in C^{1} and cRFR(k0,U(h))=cF(k0,U(h))<0\partial_{c_{R}}F_{R}\big{(}k_{0},U(-h)\big{)}=\partial_{c}F(k_{0},U(-h))<0. The Implicit Function Theorem yields a C1C^{1} real-valued function 𝒞~(k)\tilde{\mathcal{C}}(k) for kk near k0k_{0} such that

(4.20) 𝒞~(k0)=U(h),FR(k,𝒞~(k))=0.\tilde{\mathcal{C}}(k_{0})=U(-h),\quad F_{R}\big{(}k,\tilde{\mathcal{C}}(k))=0.

Since FI(k,c)=0F_{I}(k,c)=0 if cU(h)c\leq U(-h), the uniqueness of solutions ensured by the Implicit Function Theorem implies that 𝒞(k)=𝒞~(k)\mathcal{C}(k)=\tilde{\mathcal{C}}(k)\in\mathbb{R} if 𝒞~(k)U(h)\tilde{\mathcal{C}}(k)\leq U(-h).

Next we consider the case kF(k0,U(h))=0\partial_{k}F\big{(}k_{0},U(-h)\big{)}=0. Along with

cRFR(k0,U(h)),KKFR(k0,U(h))<0, where K=k2,\partial_{c_{R}}F_{R}\big{(}k_{0},U(-h)\big{)},\,\partial_{KK}F_{R}\big{(}k_{0},U(-h)\big{)}<0,\,\text{ where }K=k^{2},

it implies

FR(k0,c)>0,0<U(h)c1,FR(k,U(h))<0,k+{k0}.F_{R}(k_{0},c)>0,\;\forall 0<U(-h)-c\ll 1,\quad F_{R}\big{(}k,U(-h)\big{)}<0,\;\forall k\in\mathbb{R}^{+}\setminus\{k_{0}\}.

From the Intermediate Value Theorem, for kk near k0k_{0}, there exist real roots of FR(k,)F_{R}(k,\cdot) slightly smaller than U(h)U(-h), which must belong to 𝒞~(k)\tilde{\mathcal{C}}(k) due to the uniqueness of solutions ensured by the Implicit Function Theorem. Therefore along with the last step, we conclude 𝒞(k)=𝒞~(k)<U(h)\mathcal{C}(k)=\tilde{\mathcal{C}}(k)<U(-h) for kk0k\neq k_{0} close to k0k_{0}.

Finally, we consider the case of kF(k0,U(h))>0\partial_{k}F\big{(}k_{0},U(-h)\big{)}>0, while the opposite case can be handled similarly. The fact cF(k0,U(h))<0\partial_{c}F\big{(}k_{0},U(-h)\big{)}<0 yields

k𝒞(k0)=k𝒞~(k0)=kF(k0,U(h))cF(k0,U(h))>0,\partial_{k}\mathcal{C}(k_{0})=\partial_{k}\tilde{\mathcal{C}}(k_{0})=-\frac{\partial_{k}F\big{(}k_{0},U(-h)\big{)}}{\partial_{c}F\big{(}k_{0},U(-h)\big{)}}>0,

where in the calculation of 𝒞~(k0)\tilde{\mathcal{C}}(k_{0}) we also used FI(k,c)=0F_{I}(k,c)=0 for cU(h)c\leq U(-h) and the smoothness of FF. Hence we obtain 𝒞(k)=𝒞~(k)<U(h)\mathcal{C}(k)=\tilde{\mathcal{C}}(k)<U(-h) for kk slightly smaller than k0k_{0}. In the following we shall focus on k>k0k>k_{0} where 𝒞R(k)>U(h)\mathcal{C}_{R}(k)>U(-h). In this case, apparently F~I(k,𝒞R(k))=FI(k,𝒞R(k))0\tilde{F}_{I}(k,\mathcal{C}_{R}(k))=F_{I}(k,\mathcal{C}_{R}(k))\neq 0 and thus 𝒞I(k)0\mathcal{C}_{I}(k)\neq 0. From the Mean Value Theorem, there exists θ\theta between 0 and 𝒞I(k)\mathcal{C}_{I}(k) such that

0=F~I(k,𝒞(k))=FI(k,𝒞R(k))+𝒞I(k)cIF~I(k,𝒞R(k)+iθ).0=\tilde{F}_{I}\big{(}k,\mathcal{C}(k)\big{)}=F_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}+\mathcal{C}_{I}(k)\partial_{c_{I}}\tilde{F}_{I}\big{(}k,\mathcal{C}_{R}(k)+i\theta\big{)}.

The C1,αC^{1,\alpha} regularity of FF and 𝒞(k)\mathcal{C}(k) implies

𝒞I(k)=\displaystyle\mathcal{C}_{I}(k)= FI(k,𝒞R(k))cIF~I(k,𝒞R(k)+iθ)=YI(k,𝒞R(k))(U(0)𝒞R(k))2cIF~I(k,𝒞R(k))+O(|𝒞I(k)|α)\displaystyle-\frac{F_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}}{\partial_{c_{I}}\tilde{F}_{I}\big{(}k,\mathcal{C}_{R}(k)+i\theta\big{)}}=-\frac{Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}\big{(}U(0)-\mathcal{C}_{R}(k)\big{)}^{2}}{\partial_{c_{I}}\tilde{F}_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}+O\big{(}|\mathcal{C}_{I}(k)|^{\alpha}\big{)}}
=\displaystyle= YI(k,𝒞R(k))(U(0)𝒞R(k))2cIFI(k,𝒞R(k))+O(|𝒞I(k)|α)=YI(k,𝒞R(k))(U(0)U(h)+O(|kk0|))2cF(k0,U(h))+O(|kk0|α).\displaystyle-\frac{Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}\big{(}U(0)-\mathcal{C}_{R}(k)\big{)}^{2}}{\partial_{c_{I}}F_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}+O\big{(}|\mathcal{C}_{I}(k)|^{\alpha}\big{)}}=-\frac{Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}\big{(}U(0)-U(-h)+O(|k-k_{0}|)\big{)}^{2}}{\partial_{c}F\big{(}k_{0},U(-h)\big{)}+O\big{(}|k-k_{0}|^{\alpha}\big{)}}.

The proof of the lemma is complete. ∎

While the branch c+(k)(U(0),+)c^{+}(k)\in(U(0),+\infty) of neutral modes is global in kk\in\mathbb{R} and contained in (U(0),)\big{(}U(0),\infty\big{)} as addressed in Corollary 4.3.1, in the following we completes the picture of the other branch c(k)c^{-}(k) by combining Lemmas 4.34.10 and finish the proof of Theorem 1.1.

Proof of Theorem 1.1(3). Let g#0g_{\#}\geq 0, k#k_{\#}, and/or k#±k_{\#}^{\pm} be the thresholds given in Lemma 4.6.

Case 1. g>g#g>g_{\#}. The desired result follows from Lemma 4.6 and Corollary 4.3.1 immediately.

We start the rest of the proof much as in that of Corollary 4.3.1 and Proposition 4.4. Namely, let k0k_{0} be given by Lemma 4.2(3) and we only need to focus on c(k)c^{-}(k) for |k|k0|k|\leq k_{0}. From Lemma 4.2(2), there exists R>0R>0 such that F(k,c)0F(k,c)\neq 0 for all k[k01,k0+1]k\in[-k_{0}-1,k_{0}+1] and |c|R|c|\geq R, which also implies c+(k)(U(0),R)c^{+}(k)\in\big{(}U(0),R) for all |k|k0+1|k|\leq k_{0}+1 and c(k)(R,U(h))c^{-}(k)\in\big{(}-R,U(-h)\big{)} for all |k|[k0,k0+1]|k|\in[k_{0},k_{0}+1].

Case 2. g=g#g=g_{\#}. One the one hand, for any k1(k#,k0]k_{1}\in(k_{\#},k_{0}], Lemmas 4.2, 4.8, and 4.6 imply that there exists r0>0r_{0}>0 such that

F(k,c)0,k[k1,k0],cΩ1𝒟r0, where Ω1={c|c|<R,c𝒟r0¯},F(k,c)\neq 0,\;\forall k\in[k_{1},k_{0}],\,c\in\partial\Omega_{1}\cup\mathcal{D}_{r_{0}},\;\text{ where }\Omega_{1}=\{c\in\mathbb{C}\mid|c|<R,\,c\notin\overline{\mathcal{D}_{r_{0}}}\},

where 𝒟r\mathcal{D}_{r} is the rr-neighborhood of U([h,0])U([-h,0]) (see also (3.108)). Hence for all k[k1,k0]k\in[k_{1},k_{0}], we have Ind(F(k,),Ω1)=(F(k0,),Ω1)=2\big{(}F(k,\cdot),\Omega_{1}\big{)}=\big{(}F(k_{0},\cdot),\Omega_{1}\big{)}=2, which is equal to the number of roots of F(k,)F(k,\cdot) in Ω1\Omega_{1}. According to Corollary 4.3.1, c+(k)(U(0),R)c^{+}(k)\in\big{(}U(0),R\big{)}, |k|k0+1\forall|k|\leq k_{0}+1, is one of them. Therefore neither cases in Lemma 4.3(3) can happen to the branch c(k)c^{-}(k) and the simple root c(k)(,U(h))c^{-}(k)\in\big{(}-\infty,U(-h)\big{)} can be extended analytically for all k[k1,k0]k\in[k_{1},k_{0}]. Therefore c(k)c^{-}(k) can be extended to at least (k#,)(k_{\#},\infty) which along with c+(k)c^{+}(k) are the only roots of F(k,)F(k,\cdot) for k(k#,)k\in(k_{\#},\infty). On the other hand, according to Lemma 4.10, there exists a C1,αC^{1,\alpha} branch 𝒞(k)\mathcal{C}(k) of the only roots of F(k,c)F(k,c) for |kk#|,|cU(h)|1|k-k_{\#}|,|c-U(-h)|\ll 1. Moreover 𝒞(k)<U(h)\mathcal{C}(k)<U(-h) for 0<|kk#|10<|k-k_{\#}|\ll 1. Therefore c(k)=𝒞(k)c^{-}(k)=\mathcal{C}(k) for 0<kk#10<k-k_{\#}\ll 1 as c±(k)c^{\pm}(k) are the only roots of F(k,)F(k,\cdot) for k>k#k>k_{\#}. In particular, c(k)c^{-}(k) is thus extended to |k#k|1|k_{\#}-k|\ll 1 as a C1,αC^{1,\alpha} function with c(k)(R,U(h))c^{-}(k)\in(-R,U(-h)) for 0<|k#k|10<|k_{\#}-k|\ll 1.

Moreover, on the one hand, c(k)c^{-}(k) is the only root of F(k,)F(k,\cdot) near U([h,0])U([-h,0]) for kk near k#k_{\#} and it satisfies c(k#)=U(h)c^{-}(k_{\#})=U(-h). On the other hand, the continuity of c(k)c^{-}(k) implies that there exists ϵ1,r1>0\epsilon_{1},r_{1}>0 such that F(k,c)0F(k,c)\neq 0 for any |kk#|ϵ1|k-k_{\#}|\leq\epsilon_{1} and dist(c,U([h,0]))=r1dist\big{(}c,U([-h,0])\big{)}=r_{1}. It implies

Ind(F(k,),Ω2)=Ind(F(k#+ϵ1,),Ω2)=1,|kk#|ϵ1,\text{Ind}\big{(}F(k,\cdot),\Omega_{2}\big{)}=\text{Ind}\big{(}F(k_{\#}+\epsilon_{1},\cdot),\Omega_{2}\big{)}=1,\ \forall|k-k_{\#}|\leq\epsilon_{1},

due to the root c+(k)c^{+}(k), where

Ω2={c|c|<R,dist(c,U([h,0]))>r1}.\Omega_{2}=\{c\in\mathbb{C}\mid|c|<R,\,dist\big{(}c,U([-h,0])\big{)}>r_{1}\}.

Therefore, c±(k)c^{\pm}(k) are the only root of F(k,)F(k,\cdot) for kk near k#k_{\#}, which are also simple. As c(k)(,U(h))c^{-}(k)\in(-\infty,U(-h)) is away from c+(k)c^{+}(k), Lemma 4.3 implies that the branch c(k)c^{-}(k) of simple roots can be extended at least to (k#,+)(-k_{\#},+\infty) and remains in (,U(h))\big{(}-\infty,U(-h)\big{)}. As FF is even in kk, we have c±(k)c^{\pm}(k) are also the only roots of F(k,)F(-k,\cdot) for k(,k#)k\in(-\infty,k_{\#}). Therefore the extension c(k)c^{-}(k) must be even on (k#,k#)(-k_{\#},k_{\#}) and we obtain the whole branch c(k)c^{-}(k) for kk\in\mathbb{R}.

Case 3a. g<g#g<g_{\#} and U>0U^{\prime\prime}>0. Following the same arguments as in case 2, we obtain that c(k)=cR(k)+icI(k)c^{-}(k)=c_{R}^{-}(k)+ic_{I}^{-}(k) can be extended to a C1,αC^{1,\alpha} function on (k1,+)(k_{1},+\infty) for some k1<k#+k_{1}<k_{\#}^{+}, such that c±(k)c^{\pm}(k) and c(k)¯\overline{c^{-}(k)} are the only roots of F(k,)F(k,\cdot) for all k(k1,+)k\in(k_{1},+\infty) and cI(k)>0c_{I}^{-}(k)>0 for k(k1,k#+)k\in(k_{1},k_{\#}^{+}). Let (k1,k#+)(k_{1},k_{\#}^{+}) also denote the maximal interval of the analytic extension of c(k)c^{-}(k) as a simple root of F(k,)F(k,\cdot) inside U([h,0])\mathbb{C}\setminus U([-h,0]). The same above index based argument (in case 1) applied to [k,k#+ϵ][k,k_{\#}^{+}-\epsilon] for any k(max{k1,k#},k#+)k\in(\max\{k_{1},k_{\#}^{-}\},k_{\#}^{+}) and 0ϵ<k#+k0\ll\epsilon<k_{\#}^{+}-k also implies that c±(k)c^{\pm}(k) and c(k)¯\overline{c^{-}(k)} are the only roots of F(k,)F(k,\cdot) for all k(max{k1,k#},k#+)k\in(\max\{k_{1},k_{\#}^{-}\},k_{\#}^{+}). According to Lemma 4.2 we have k1k0>k_{1}\geq-k_{0}>-\infty. For k(k1,k#+)k\in(k_{1},k_{\#}^{+}), the semicircle theorem implies that c(k)c^{-}(k) lies in the closed upper semi-disk with the boundary diameter U([h,0])U([-h,0]) and thus |c(k)c+(k)|>ρ0|c^{-}(k)-c^{+}(k)|>\rho_{0} where ρ0>0\rho_{0}>0 is given in Corollary 4.3.1. Moreover, since F(k,c)0F(k,c)\neq 0 for any cU((h,0])c\in U\big{(}(-h,0]\big{)} (Lemmas 4.1 and 4.8), we obtain from Lemma 4.3

limkk1+c(k)=U(h)F(k2,U(h))=0k2{k#,k#,k#+}.\lim_{k\to k_{1}+}c^{-}(k)=U(-h)\Longrightarrow F\big{(}k_{2},U(-h)\big{)}=0\Longrightarrow k_{2}\in\{k_{\#}^{-},\,-k_{\#}^{-},\,-k_{\#}^{+}\}.

It must hold k2=k#k_{2}=k_{\#}^{-}, otherwise c(k#)U(h)c^{-}(k_{\#}^{-})\neq U(-h), F(k#,U(h))=0F\big{(}k_{\#}^{-},U(-h)\big{)}=0, kF(k#,U(h))>0\partial_{k}F\big{(}k_{\#}^{-},U(-h)\big{)}>0, and Lemma 4.10 imply that there exists the fourth root near U(h)U(-h) for 0<kk#10<k-k_{\#}^{-}\ll 1. This contradicts that F(k,)F(k,\cdot) has exactly three roots for all k(max{k1,k#},k#+)k\in(\max\{k_{1},k_{\#}^{-}\},k_{\#}^{+}) and thus k2=k#k_{2}=k_{\#}^{-} and c(k#)=U(h)c^{-}(k_{\#}^{-})=U(-h). For 0<k#k10<k_{\#}^{-}-k\ll 1, Lemma 4.10 yields the further extension of c(k)c^{-}(k) back into (,U(h))\big{(}-\infty,U(-h)\big{)}. From a similar argument, we can extend this branch to k=k#k=-k_{\#}^{-} with c(k#)=U(h)c^{-}(-k_{\#}^{-})=U(-h). Finally, the whole branch c(k)c^{-}(k) for kk\in\mathbb{R} is obtained by the evenness c(k)=c(k)c^{-}(-k)=c^{-}(-k).

Case 3b. g<g#g<g_{\#} and U<0U^{\prime\prime}<0. Following the same arguments as in case 2, we obtain that c(k)=cR(k)+icI(k)c^{-}(k)=c_{R}^{-}(k)+ic_{I}^{-}(k) can be extended to a C1,αC^{1,\alpha} function on [k#+,+)[k_{\#}^{+},+\infty) and c(k#+)=U(h)c^{-}(k_{\#}^{+})=U(-h). However, for 0<k#+k10<k_{\#}^{+}-k\ll 1, Lemma 4.10 implies that there does not exist any roots of F(k,)F(k,\cdot) near U(h)U(-h) (as 𝒞I<0\mathcal{C}_{I}<0 due to U<0U^{\prime\prime}<0). The same index argument further yields that c+(k)c^{+}(k) is the only root for k(k#,k#+)k\in(k_{\#}^{-},k_{\#}^{+}). From Lemma 4.10, we obtain another branch of roots in (,U(h))(-\infty,U(-h)) of F(k,)F(k,\cdot) for k(k#,k#)k\in(-k_{\#}^{-},k_{\#}^{-}) which along with the c+(k)c^{+}(k) are the only roots. The final conclusion again follows from the even symmetry as in the above cases. \square

Remark 4.3.

As in [39] for the gravity wave, the spectral stability in the case U<0U^{\prime\prime}<0 can also be obtained by directly modifying the usual proof of the Rayleigh theorem in the fixed boundary case. Namely, multiplying (3.1) by y¯\bar{y}, integrating on [h,0][-h,0], using the homogeneous boundary condition as in (2.11b) and (2.11c), and the semicircle theorem, a contradiction occurs if an unstable mode cc exists. Our above proof provides a complete picture of the eigenvalue distribution, however.

4.3. Singular neutral modes at inflection values

To end this section, we discuss the spectrum near inflection values of UU, which are the only possible singular neutral modes other than U(h)U(-h) according to Lemma 4.1(6).

Proposition 4.11.

Assume UC5U\in C^{5}, x20[h,0)x_{20}\in[-h,0), and U(x20)=0U^{\prime\prime}(x_{20})=0, then the following hold for c0=U(x20)c_{0}=U(x_{20}).

  1. (1)

    For any α(0,12)\alpha\in(0,\frac{1}{2}), there exist C>0C>0 depending only on UU, gg, and α\alpha, such that, with

    k=Cmax{1,(U(0)c0)2},σ0=(U(0)c0)2/(2k),k_{*}=C\max\{1,(U(0)-c_{0})^{-2}\},\quad\sigma_{0}=(U(0)-c_{0})^{2}/(2k_{*}),

    for any σ(0,σ0)\sigma\in(0,\sigma_{0}), there exists a unique k0kk_{0}\geq k_{*} such that F(k0,c0)=0F(k_{0},c_{0})=0. Moreover it satisfies

    |k0(U(0)c0)2/σ|C(U(0)c0)2,|kF(k0,c0)+(U(0)c0)2|C(U(0)c0)2ασα.|k_{0}-(U(0)-c_{0})^{2}/\sigma|\leq C(U(0)-c_{0})^{-2},\quad|\partial_{k}F(k_{0},c_{0})+(U(0)-c_{0})^{2}|\leq C(U(0)-c_{0})^{-2\alpha}\sigma^{\alpha}.
  2. (2)

    In addition, suppose x20hx_{20}\neq-h and

    𝐅(k0,c0)=0,k0>0,kF(k0,c0)0,U(x20)0,\mathbf{F}(k_{0},c_{0})=0,\;k_{0}>0,\;\partial_{k}F(k_{0},c_{0})\neq 0,\;U^{\prime\prime\prime}(x_{20})\neq 0,

    then there exist C~>0\tilde{C}>0, δ>0\delta>0, and a C1C^{1} function c(k)c(k) defined for

    0|kk0|δ,(kk0)U(x20)kF(k0,c0)>0,0\leq|k-k_{0}|\leq\delta,\quad(k-k_{0})U^{\prime\prime\prime}(x_{20})\partial_{k}F(k_{0},c_{0})>0,

    such that c(k0)=c0c(k_{0})=c_{0}, cI(k)>0c_{I}(k)>0 for the above kk0k\neq k_{0}, and

    F(k,c)=0,|kk0|δ and |cc0|C~δ iff c{c(k),c(k)¯}.F(k,c)=0,\;|k-k_{0}|\leq\delta\text{ and }|c-c_{0}|\leq\tilde{C}\delta\;\text{ iff }\;c\in\{c(k),\overline{c(k)}\}.

In the above statement (2), note that 𝐅(k0,c0)=0\mathbf{F}(k_{0},c_{0})=0 and Lemma 4.1(4) imply y(k0,c0,0)0y_{-}(k_{0},c_{0},0)\neq 0 and thus Y(k,c0)Y(k,c_{0}) is well-defined which is actually real due to U(x20)=0U^{\prime\prime}(x_{20})=0 and Lemma 3.22. Therefore it makes sense to talk about the sign of kF(k0,c0)\partial_{k}F(k_{0},c_{0}). Statement (1) also implies that assumptions of statement (2) may be satisfied at inflection values of UU with |k|1|k|\gg 1 if σ\sigma is small.

Proof.

From Lemma 3.9 and Remark 3.4, there exists C0>0C_{0}>0 such that

ky(k,c,x2)(1/2)sinhμ1(x2+h)y(k,c,0)0,|k|C0,c,ky_{-}(k,c,x_{2})\geq(1/2)\sinh\mu^{-1}(x_{2}+h)\implies y_{-}(k,c,0)\neq 0,\quad\forall|k|\geq C_{0},\,c\in\mathbb{C},

and thus F(k,c)F(k,c) and Y(k,c)Y(k,c) are defined for all |k|C0|k|\geq C_{0}. According to (4.6), FI(k,c0)=0F_{I}(k,c_{0})=0 for all kk\in\mathbb{R} and thus F(k,c0)F(k,c_{0})\in\mathbb{R}. Lemmas 3.24 and 3.22 imply, for |k|C0|k|\geq C_{0} and cU([h,0))c\in U([-h,0)),

|YI(k,c)|C|U(x2c)|e2μ1x2c|Y(k,c0)|k||Cμ.|Y_{I}(k,c)|\leq C|U^{\prime\prime}(x_{2}^{c})|e^{2\mu^{-1}x_{2}^{c}}\implies\big{|}Y(k,c_{0})-|k|\big{|}\leq C\mu.

Therefore, for |k|C0|k|\geq C_{0}, it holds

||k|1F(k,c0)(U(0)c0)2+σ|k||Cμ.\big{|}|k|^{-1}F(k,c_{0})-(U(0)-c_{0})^{2}+\sigma|k|\big{|}\leq C\mu.

Let

k=max{C0,3C(U(0)c0)2}Ck1(U(0)c0)2/3.k_{*}=\max\{C_{0},3C(U(0)-c_{0})^{-2}\}\implies C\langle k_{*}\rangle^{-1}\leq(U(0)-c_{0})^{2}/3.

From the Intermediate Value Theorem, for every 0<σσ00<\sigma\leq\sigma_{0}, there exists a root k0[k,+)k_{0}\in[k_{*},+\infty) of F(,c0)F(\cdot,c_{0}) close to (U(0)c0)2/σ(U(0)-c_{0})^{2}/\sigma.

To estimate kF(k0,c0)\partial_{k}F(k_{0},c_{0}) and obtain the uniqueness of k0k_{0}, we analyze kY(k0,c0)\partial_{k}Y(k_{0},c_{0}) using the same standard method used in the proof of Lemma 4.5. Let

y(k,x2)=y0(k,c0,x2)y0(k,c0,0)y+(k2+UUc0)y=0,y(h)=0,y(0)=1,Y(k,c0)=y(0),y(k,x_{2})=\frac{y_{0-}(k,c_{0},x_{2})}{y_{0-}(k,c_{0},0)}\implies-y^{\prime\prime}+\Big{(}k^{2}+\frac{U^{\prime\prime}}{U-c_{0}}\Big{)}y=0,\;\;y(-h)=0,\;y(0)=1,\;Y(k,c_{0})=y^{\prime}(0),

where UUc0C3([h,0])\frac{U^{\prime\prime}}{U-c_{0}}\in C^{3}([-h,0]). Differentiating the above equation with respect to kk yields

ky+(k2+UUc0)ky=2ky,ky(h)=ky(0)=0,kY(k,c0)=ky(0),\displaystyle-\partial_{k}y^{\prime\prime}+\Big{(}k^{2}+\frac{U^{\prime\prime}}{U-c_{0}}\Big{)}\partial_{k}y=-2ky,\;\;\partial_{k}y(-h)=\partial_{k}y(0)=0,\;\partial_{k}Y(k,c_{0})=\partial_{k}y^{\prime}(0),
\displaystyle\implies kY(k,c0)=h0(kyykyy)dx2=2kh0y(x2)2dx2.\displaystyle\partial_{k}Y(k,c_{0})=\int_{-h}^{0}(\partial_{k}y^{\prime}y-\partial_{k}yy^{\prime})^{\prime}dx_{2}=2k\int_{-h}^{0}y(x_{2})^{2}dx_{2}.

From Lemma 3.9, we can estimate, for any α(0,12)\alpha\in(0,\frac{1}{2}) and |k|>k|k|>k_{*},

|kY(k,c0)2kh0(sinhμ1(x2+h)sinhμ1h)2dx2|Cμα1h0(sinhμ1(x2+h)sinhμ1h)2dx2\displaystyle\Big{|}\partial_{k}Y(k,c_{0})-2k\int_{-h}^{0}\Big{(}\frac{\sinh\mu^{-1}(x_{2}+h)}{\sinh\mu^{-1}h}\Big{)}^{2}dx_{2}\Big{|}\leq C\mu^{\alpha-1}\int_{-h}^{0}\Big{(}\frac{\sinh\mu^{-1}(x_{2}+h)}{\sinh\mu^{-1}h}\Big{)}^{2}dx_{2}
\displaystyle\implies |kY(k,c0)sgn(k)|Cμα.\displaystyle|\partial_{k}Y(k,c_{0})-sgn(k)|\leq C\mu^{\alpha}.

Therefore we obtain

kF(k,c0)=(U(0)c0)2kY(k,c0)2σk=(U(0)c0)2sgn(k)2σk+O(|k|α),\partial_{k}F(k,c_{0})=(U(0)-c_{0})^{2}\partial_{k}Y(k,c_{0})-2\sigma k=(U(0)-c_{0})^{2}sgn(k)-2\sigma k+O(|k|^{-\alpha}),

which implies

kF(k0,c0)=(U(0)c0)2+O(k0α) if k0(k,) and F(k0,c0)=0.\partial_{k}F(k_{0},c_{0})=-(U(0)-c_{0})^{2}+O(k_{0}^{-\alpha})\;\text{ if }\;k_{0}\in(k_{*},\infty)\text{ and }F(k_{0},c_{0})=0.

The desired estimate on kF(k0,c0)\partial_{k}F(k_{0},c_{0}) follows immediately, whose always negative sign also implies the uniqueness of such k0(k,)k_{0}\in(k_{*},\infty).

Under the assumption in statement (2) of the proposition, Lemma 4.1(4) implies y(k0,c0,0)0y_{-}(k_{0},c_{0},0)\neq 0 and thus F(k,c)F(k,c) is C1C^{1} in (k,c)(k,c) near (k0,c0)(k_{0},c_{0}) with cI0c_{I}\geq 0. Much as in the proof of Lemma 4.10, statement (2) can be proved by applying the Implicit Function Theorem to F~(k,c)\tilde{F}(k,c), an extension of F(k,c)F(k,c) which is C1C^{1} in (k,c)(k,c) in ×\mathbb{R}\times\mathbb{C} near (k0,c0)(k_{0},c_{0}). The Jacobi matrix of F~\tilde{F} is

DcF~(k0,c0)=(cRF~RcIF~RcRF~IcIF~I)|(k0,c0)=(cRFRcRFIcRFIcRFR)|(k0,c0),D_{c}\tilde{F}(k_{0},c_{0})=\begin{pmatrix}\partial_{c_{R}}\tilde{F}_{R}&\partial_{c_{I}}\tilde{F}_{R}\\ \partial_{c_{R}}\tilde{F}_{I}&\partial_{c_{I}}\tilde{F}_{I}\end{pmatrix}\Big{|}_{(k_{0},c_{0})}=\begin{pmatrix}\partial_{c_{R}}F_{R}&-\partial_{c_{R}}F_{I}\\ \partial_{c_{R}}F_{I}&\partial_{c_{R}}F_{R}\end{pmatrix}\Big{|}_{(k_{0},c_{0})},

where we also used the Cauchy-Riemann equation. According to Lemma 3.22, Y(k,c0)Y(k,c_{0})\in\mathbb{R} and

cRFI(k0,c0)=(U(0)c0)2cRYI(k0,c0)=(U(0)c0)2πU(x20)y0(k0,c0,x20)2U(x20)2y0(k0,c0,0)20,\partial_{c_{R}}F_{I}(k_{0},c_{0})=(U(0)-c_{0})^{2}\partial_{c_{R}}Y_{I}(k_{0},c_{0})=(U(0)-c_{0})^{2}\frac{\pi U^{\prime\prime\prime}(x_{20})y_{0-}(k_{0},c_{0},x_{20})^{2}}{U^{\prime}(x_{20})^{2}y_{0-}(k_{0},c_{0},0)^{2}}\neq 0,

and has the same sign as U(x20)U^{\prime\prime\prime}(x_{20}). Therefore DcF~(k0,c0)D_{c}\tilde{F}(k_{0},c_{0}) is invertible and thus there exist δ>0\delta>0 and a C1C^{1} function c(k)=cR(k)+icI(k)c(k)=c_{R}(k)+ic_{I}(k) defined for all |kk0|δ|k-k_{0}|\leq\delta such that F~(k,c)=0\tilde{F}(k,c)=0 for (k,c)×(k,c)\in\mathbb{R}\times\mathbb{C} iff c=c(k)c=c(k). Consequently F(k,c)=0F(k,c)=0 for (k,c)(k,c) near (k0,c0)(k_{0},c_{0}) iff c{c(k),c(k)¯}c\in\{c(k),\overline{c(k)}\} and cI(k)0c_{I}(k)\geq 0. Identifying complex numbers with 2-d column vectors, since

kc(k0)=(DcF~(k0,c0))1kF~(k0,c0)=kF(k0,c0)/cF(k0,c0)\displaystyle\partial_{k}c(k_{0})=-(D_{c}\tilde{F}(k_{0},c_{0}))^{-1}\partial_{k}\tilde{F}(k_{0},c_{0})=-\partial_{k}F(k_{0},c_{0})/\partial_{c}F(k_{0},c_{0})

implies cI(k)(kk0)kF(k0,c0)U(x20)>0c_{I}(k)(k-k_{0})\partial_{k}F(k_{0},c_{0})U^{\prime\prime\prime}(x_{20})>0 for kk near k0k_{0}, statement (2) follows readily. ∎

Remark 4.4.

In part (1) of the proposition, one may also seek k0k_{0} satisfying 𝐅(k0,c0)=0\mathbf{F}(k_{0},c_{0})=0 using the Intermediate Value Theorem instead. It is easy to see 𝐅(k,c0)\mathbf{F}(k,c_{0})\in\mathbb{R} approaches -\infty as kk\to\infty. Therefore such k0k_{0} exists if supk0𝐅(k,c0)>0\sup_{k\geq 0}\mathbf{F}(k,c_{0})>0 and only if supk0𝐅(k,c0)0\sup_{k\geq 0}\mathbf{F}(k,c_{0})\geq 0, which may not the case if gg and σ\sigma are sufficiently large. This is different from the gravity waves (i.e. σ=0\sigma=0), see [39, 14, 15]. It is also worth pointing out that the smoothness of FF for cI0c_{I}\geq 0 based on Section 3 made the analysis using the Implicit Function Theorem in part (2) easier, compared with, e.g. [14].

5. Boundary value problems of the non-homogeneous Rayleigh equation

In this section, using the fundamental solutions y±(k,c,x2)y_{\pm}(k,c,x_{2}) to the homogeneous Rayleigh equation (3.1), we study the boundary value problem of the non-homogeneous Rayleigh equation

(5.1a) y+(k2+UUc)y=ψ(c,x2)Uc,x2(h,0);-y^{\prime\prime}+\big{(}k^{2}+\frac{U^{\prime\prime}}{U-c}\big{)}y=\frac{\psi(c,x_{2})}{U-c},\quad x_{2}\in(-h,0);
(5.1b) y(h)=ζ(c),(U(0)c)2y(0)(U(0)(U(0)c)+g+σk2)y(0)=ζ+(c),y(-h)=\zeta_{-}(c),\quad\big{(}U(0)-c\big{)}^{2}y^{\prime}(0)-\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}y(0)=\zeta_{+}(c),

where the boundary conditions are from the linearized water wave system (2.11).

Using the two fundamental solutions y±y_{\pm} to the homogeneous equation with zero boundary values, for cU([h,0])c\in\mathbb{C}\setminus U([-h,0]) it is standard to compute the solution to (5.1) in the form

(5.2) yB(k,c,x2)=ζ+(c)𝐅(k,c)y(k,c,x2)+ζ(c)y+(k,c,h)y+(k,c,x2)+ynh(k,c,x2),y_{B}(k,c,x_{2})=\frac{\zeta_{+}(c)}{\mathbf{F}(k,c)}y_{-}(k,c,x_{2})+\frac{\zeta_{-}(c)}{y_{+}(k,c,-h)}y_{+}(k,c,x_{2})+y_{nh}(k,c,x_{2}),

where ynhy_{nh} is the solution to (5.1a) with zero boundary values in (5.1b) given by

(5.3) ynh(k,c,x2)=y+(k,c,x2)y+(k,c,h)hx2(yψ)(k,c,x2)U(x2)cdx2+y(k,c,x2)y+(k,c,h)x20(y+ψ)(k,c,x2)U(x2)cdx2.y_{nh}(k,c,x_{2})=\frac{y_{+}(k,c,x_{2})}{y_{+}(k,c,-h)}\int_{-h}^{x_{2}}\frac{(y_{-}\psi)(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime}+\frac{y_{-}(k,c,x_{2})}{y_{+}(k,c,-h)}\int_{x_{2}}^{0}\frac{(y_{+}\psi)(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime}.

Its derivative in x2x_{2} is given by

(5.4) ynh(k,c,x2)=y+(k,c,x2)y+(k,c,h)hx2(yψ)(k,c,x2)U(x2)cdx2+y(k,c,x2)y+(k,c,h)x20(y+ψ)(k,c,x2)U(x2)cdx2.y_{nh}^{\prime}(k,c,x_{2})=\frac{y_{+}^{\prime}(k,c,x_{2})}{y_{+}(k,c,-h)}\int_{-h}^{x_{2}}\frac{(y_{-}\psi)(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime}+\frac{y_{-}^{\prime}(k,c,x_{2})}{y_{+}(k,c,-h)}\int_{x_{2}}^{0}\frac{(y_{+}\psi)(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime}.

Here the unique solvability condition of (5.1) is 𝐅(k,c)0\mathbf{F}(k,c)\neq 0, where 𝐅\mathbf{F} is defined in (4.1), as the Wronskian of the fundamental solutions y±y_{\pm}, which is a constant in x2x_{2}, is given by

(5.5) y+(k,c,h)=(g+σk2)1𝐅(k,c)=(y+yy+y)(k,c,x2).y_{+}(k,c,-h)=(g+\sigma k^{2})^{-1}\mathbf{F}(k,c)=(y_{+}y_{-}^{\prime}-y_{+}^{\prime}y_{-})(k,c,x_{2}).

Throughout this section, we consider

c=cR+icI,cR=U([hρ0,ρ0]),|cI|ρ0,c=c_{R}+ic_{I},\quad c_{R}\in\mathcal{I}=U([-h-\rho_{0},\rho_{0}]),\quad|c_{I}|\leq\rho_{0},

where ρ0[0,h0]\rho_{0}\in[0,h_{0}]. By choosing ρ0\rho_{0} smaller, we also have that, for some C>0C>0 depending only on |U|C1|U|_{C^{1}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}},

(5.6) Re(g+σk2+U(0)(U(0)c))(1+k2)/C,k,c+i[ρ0,ρ0].\text{Re}\,\big{(}g+\sigma k^{2}+U^{\prime}(0)(U(0)-c)\big{)}\geq(1+k^{2})/C,\quad\forall k\in\mathbb{R},\;c\in\mathcal{I}+i[-\rho_{0},\rho_{0}].

This and boundary condition (5.1b) imply

(5.7) |y(0)|Cμ2(|U(0)c|2|y(0)|+|ζ+|),|y(0)|\leq C\mu^{2}(|U(0)-c|^{2}|y^{\prime}(0)|+|\zeta_{+}|),

which will be used repeatedly to control y(0)y(0) in terms of y(0)y^{\prime}(0).

Throughout this section, we assume that, there exists ρ0>0\rho_{0}>0 such that

(5.8) F0=inf{(1+k2)12ehμ|𝐅(k,c)|cR=U([U(h)ρ0,U(0)+ρ0]),|cI|[ρ0,ρ0]}>0.F_{0}=\inf\{(1+k^{2})^{-\frac{1}{2}}e^{-\frac{h}{\mu}}|\mathbf{F}(k,c)|\mid c_{R}\in\mathcal{I}=U([U(-h)-\rho_{0},U(0)+\rho_{0}]),\,|c_{I}|\in[-\rho_{0},\rho_{0}]\}>0.

In this subsection, mostly we shall not vary kk\in\mathbb{R}, but carefully track the dependence of the estimates on kk, or equivalently μ=(1+k2)12\mu=(1+k^{2})^{-\frac{1}{2}}. From Lemma 3.9, it is easy to compute that, for any r1[1,]r_{1}\in[1,\infty], r2[1,)r_{2}\in[1,\infty), and |cI|ρ0|c_{I}|\leq\rho_{0},

μ(1+1r1)|y±|Lx2r1LcR+μ1r1|y±|Lx2r1LcRr2+μ1r2|y±|LcRLx2r2+|y+(h)|LcRr2+|y(0)|LcRr2Ceμ1h,\mu^{-(1+\frac{1}{r_{1}})}|y_{\pm}|_{L_{x_{2}}^{r_{1}}L_{c_{R}}^{\infty}}+\mu^{-\frac{1}{r_{1}}}|y_{\pm}^{\prime}|_{L_{x_{2}}^{r_{1}}L_{c_{R}}^{r_{2}}}+\mu^{-\frac{1}{r_{2}}}|y_{\pm}^{\prime}|_{L_{c_{R}}^{\infty}L_{x_{2}}^{r_{2}}}+|y_{+}^{\prime}(-h)|_{L_{c_{R}}^{r_{2}}}+|y_{-}^{\prime}(0)|_{L_{c_{R}}^{r_{2}}}\leq Ce^{\mu^{-1}h},

where x2[h,0]x_{2}\in[-h,0] and cRc_{R}\in\mathcal{I}. This inequality will be used repeatedly in the rest of the paper.

Solutions to this system are rather smooth away from c{U(x2),U(0),U(h)}c\in\{U(x_{2}),\,U(0),\,U(-h)\} and their singular behaviors near this set could be analyzed rather detailedly following the approach in Section 3, based on (3.34) and (3.74) and the estimates on B~\tilde{B} and BB. However, for the purpose of this paper, it is sufficient just to obtain certain bounds of the solutions based on the properties of the homogeneous solutions y±y_{\pm}, which is carried out in this section.

As a preparation, in Subsection 5.1 we shall first consider (5.1) with zero boundary conditions ζ±=0\zeta_{\pm}=0 in (5.1b). Subsequently in Subsection 5.2, we study the non-homogenous Rayleigh system (5.1) with ζ±\zeta_{\pm} linear in cc, particularly focusing on the derivatives of the solutions on c+i[ρ0,ρ0]c\in\mathcal{I}+i[-\rho_{0},\rho_{0}]. We sometimes skip writing parameters kk and cc explicitly.

5.1. Non-homogeneous Rayleigh system (5.1) with zero boundary conditions ζ±=0\zeta_{\pm}=0

The formulas (5.3) and (5.4) of ynh(k,c,x2)y_{nh}(k,c,x_{2}) and ynh(k,c,x2)y_{nh}^{\prime}(k,c,x_{2}) are actually consistent with (3.34) for x2x_{2} near x2cx_{2}^{c}. In fact, (3.34) implies that (10Γ1)B~\begin{pmatrix}1&0\\ \Gamma&1\end{pmatrix}\tilde{B} is a fundamental matrix of (3.1) and hence B~\tilde{B} can be rewritten in terms of y±y_{\pm} and Γ\Gamma. A straight forward calculation using (3.30) and (3.34) also yields (5.3). This solution also satisfy

ynh(k,c¯,x2)¯=ynh(k,c,x2)=ynh(k,c,x2),\overline{y_{nh}(k,\bar{c},x_{2})}=y_{nh}(k,c,x_{2})=y_{nh}(-k,c,x_{2}),

so we mainly focus on cI0c_{I}\geq 0. Assume ψ(cR+icI,x2)ψ0(cR,x2)\psi(c_{R}+ic_{I},x_{2})\to\psi_{0}(c_{R},x_{2}) as cI0+c_{I}\to 0+. Due to the singularity of the non-homogeneous term at x2=x2cx_{2}=x_{2}^{c} (as defined in (3.20) by U(x2c)=cRU(x_{2}^{c})=c_{R}) as cI0+c_{I}\to 0+, the limits of ynhy_{nh} and ynhy_{nh}^{\prime} involve P.V.P.V. of integrals and delta masses

(5.9) ynh0(x2)=P.V.h0ψ0(x2)y0+(x2)y0(x2)χ{x2<x2}+y0(x2)y0+(x2)χ{x2>x2}y0+(h)(U(x2)cR)dx2+iπψ0(x2c)U(x2c)(y0+(x2)y0(x2c)y0+(h)χ{U(x2)>cR>U(h)}+y0(x2)y0+(x2c)y0+(h)χ{U(0)>cR>U(x2)}),\begin{split}y_{nh0}(x_{2})&=P.V.\int_{-h}^{0}\psi_{0}(x_{2}^{\prime})\frac{y_{0+}(x_{2})y_{0-}(x_{2}^{\prime})\chi_{\{x_{2}^{\prime}<x_{2}\}}+y_{0-}(x_{2})y_{0+}(x_{2}^{\prime})\chi_{\{x_{2}^{\prime}>x_{2}\}}}{y_{0+}(-h)(U(x_{2}^{\prime})-c_{R})}dx_{2}^{\prime}\\ &+\frac{i\pi\psi_{0}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\Big{(}\frac{y_{0+}(x_{2})y_{0-}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(x_{2})>c_{R}>U(-h)\}}+\frac{y_{0-}(x_{2})y_{0+}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(0)>c_{R}>U(x_{2})\}}\Big{)},\end{split}
(5.10) ynh0(x2)=P.V.h0ψ0(x2)y0+(x2)y0(x2)χ{x2<x2}+y0(x2)y0+(x2)χ{x2>x2}y0+(h)(U(x2)cR)dx2+iπψ0(x2c)U(x2c)(y0+(x2)y0(x2c)y0+(h)χ{U(x2)>cR>U(h)}+y0(x2)y0+(x2c)y0+(h)χ{U(0)>cR>U(x2)}),\begin{split}y_{nh0}^{\prime}(x_{2})&=P.V.\int_{-h}^{0}\psi_{0}(x_{2}^{\prime})\frac{y_{0+}^{\prime}(x_{2})y_{0-}(x_{2}^{\prime})\chi_{\{x_{2}^{\prime}<x_{2}\}}+y_{0-}^{\prime}(x_{2})y_{0+}(x_{2}^{\prime})\chi_{\{x_{2}^{\prime}>x_{2}\}}}{y_{0+}(-h)(U(x_{2}^{\prime})-c_{R})}dx_{2}^{\prime}\\ &+\frac{i\pi\psi_{0}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\Big{(}\frac{y_{0+}^{\prime}(x_{2})y_{0-}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(x_{2})>c_{R}>U(-h)\}}+\frac{y_{0-}^{\prime}(x_{2})y_{0+}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(0)>c_{R}>U(x_{2})\}}\Big{)},\end{split}

where χ\chi is the characteristic function and we skipped the dependence on cRc_{R} of ψ0\psi_{0}, y0±y_{0\pm}, and ynh0y_{nh0}. Naturally, in the above the P.V.P.V. is taken only when there are singularities in the integral.

We consider a priori and convergence estimates of ynhy_{nh} as cI0+c_{I}\to 0+ in the following two cases of ψ(c,x2)\psi(c,x_{2}), motivated by the non-homogenous Rayleigh system (2.11) and its differentiation in cc.

\bullet Case 1: ψ(c,)Lx2r,r(1,)\psi^{\prime}(c,\cdot)\in L_{x_{2}}^{r},\;r\in(1,\infty). While this case occurs in the linearized capillary gravity wave (2.11) when some regularity is assumed on the initial vorticity, it is also a crucial part of the analysis when (2.11) is differentiated in cc.

Lemma 5.1.

Assume (5.8). For any ϵ>0\epsilon>0222Like the generic upper bound C>0C>0, the small constant ϵ>0\epsilon>0 in this and the next section may change from line to line., there exists C>0C>0 depending only on rr, ϵ\epsilon, F0F_{0}, ρ0\rho_{0}, |U|C2|U^{\prime}|_{C^{2}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that the following hold.

  1. (1)

    For any kk\in\mathbb{R}, x2[h,0]x_{2}\in[-h,0], cI(0,ρ0]c_{I}\in(0,\rho_{0}], and cRc_{R}\in\mathcal{I} it holds

    |ynh(k,c,x2)|C\displaystyle|y_{nh}(k,c,x_{2})|\leq C μ11rϵ(μ|ψ|Lx2r+|ψ|Lx2r),\displaystyle\mu^{1-\frac{1}{r}-\epsilon}(\mu|\psi^{\prime}|_{L_{x_{2}}^{r}}+|\psi|_{L_{x_{2}}^{r}}),
    |ynh(k,c,x2)|Cμ1rϵ\displaystyle|y_{nh}^{\prime}(k,c,x_{2})|\leq C\mu^{-\frac{1}{r}-\epsilon} (1+|log|U(x2)c||)(μ|ψ|Lx2r+|ψ|Lx2r).\displaystyle\big{(}1+\big{|}\log|U(x_{2})-c|\big{|}\big{)}(\mu|\psi^{\prime}|_{L_{x_{2}}^{r}}+|\psi|_{L_{x_{2}}^{r}}).
  2. (2)

    Assume ψ(+icI,)ψ(,)\psi(\cdot+ic_{I},\cdot)\to\psi(\cdot,\cdot) in LcRr1Wx21,rL_{c_{R}}^{r_{1}}W_{x_{2}}^{1,r} as cI0+c_{I}\to 0+ with r1(1,)r_{1}\in(1,\infty) and r(1,)r\in(1,\infty), then

    1. (a)

      ynhynh0y_{nh}\to y_{nh0} in LcRq1Lx2L_{c_{R}}^{q_{1}}L_{x_{2}}^{\infty} for any q1[1,r1)q_{1}\in[1,r_{1}) and ynhynh0y_{nh}^{\prime}\to y_{nh0}^{\prime} in LcRq1Lx2q2L_{c_{R}}^{q_{1}}L_{x_{2}}^{q_{2}} for any q1[1,r1)q_{1}\in[1,r_{1}) and q2[1,)q_{2}\in[1,\infty);

    2. (b)

      at x~2=h\tilde{x}_{2}=-h and x~2=0\tilde{x}_{2}=0, ynh(+icI,x~2)ynh0(,x~2)y_{nh}(\cdot+ic_{I},\tilde{x}_{2})\to y_{nh0}(\cdot,\tilde{x}_{2}) and ynh(+icI,x~2)ynh0(,x~2)y_{nh}^{\prime}(\cdot+ic_{I},\tilde{x}_{2})\to y_{nh0}^{\prime}(\cdot,\tilde{x}_{2}) in LcRq1L_{c_{R}}^{q_{1}} for any q1[1,r1)q_{1}\in[1,r_{1}). Moreover, and for any ϵ>0\epsilon>0, for any kk\in\mathbb{R}, cI[0,1]c_{I}\in[0,1],

      |ynh(c,x~2)|C(μ1rϵ(μ|ψ|Lx2r+|ψ|Lx2r)+(1+|log|U(x~2)c||)|ψ(x~2)|).|y_{nh}^{\prime}(c,\tilde{x}_{2})|\leq C\big{(}\mu^{-\frac{1}{r}-\epsilon}(\mu|\psi^{\prime}|_{L_{x_{2}}^{r}}+|\psi|_{L_{x_{2}}^{r}})+\big{(}1+\big{|}\log|U(\tilde{x}_{2})-c|\big{|}\big{)}|\psi(\tilde{x}_{2})|\big{)}.

Even though the above formulas of ynh0y_{nh0} involve some subtlety at x2=x2cx_{2}=x_{2}^{c}, the regularity of ynh0y_{nh0}^{\prime} in x2x_{2} implies that ynh0y_{nh0} is Hölder continuous. In fact, the continuity of ynh0y_{nh0} at x2=x2cx_{2}=x_{2}^{c} can also be seen directly by using the rather precise local form of y0±y_{0\pm} near x2cx_{2}^{c} given in Lemma 3.10. Moreover, while the convergence is given in the integral norms, one could attempt to obtain more detailed convergence estimates near x2cx_{2}^{c} using the tools given in Lemma 3.4 and Proposition 3.7.

Proof.

Since cI>0c_{I}>0, no singularity is involved in (5.3) and (5.4), one can compute via integration by parts

hx2yψUcdx2=hx2yψU(log(Uc))dx2=(yψUlog(Uc))(x2)hx2(yψU)log(Uc)dx2.\int_{-h}^{x_{2}}\frac{y_{-}\psi}{U-c}dx_{2}^{\prime}=\int_{-h}^{x_{2}}\frac{y_{-}\psi}{U^{\prime}}\big{(}\log(U-c)\big{)}^{\prime}dx_{2}^{\prime}=\big{(}\frac{y_{-}\psi}{U^{\prime}}\log(U-c)\big{)}(x_{2})-\int_{-h}^{x_{2}}\big{(}\frac{y_{-}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}.

The other integral can be handled similarly,

x20y+ψUcdx2=(y+ψUlog(Uc))|x20x20(y+ψU)log(Uc)dx2.\int_{x_{2}}^{0}\frac{y_{+}\psi}{U-c}dx_{2}^{\prime}=\big{(}\frac{y_{+}\psi}{U^{\prime}}\log(U-c)\big{)}\Big{|}_{x_{2}}^{0}-\int_{x_{2}}^{0}\big{(}\frac{y_{+}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}.

Observing that the boundary terms at x2x_{2} are canceled and we have

(5.11) ynh(x2)=y+(x2)y+(h)hx2(yψU)log(Uc)dx2y(x2)y+(h)x20(y+ψU)log(Uc)dx2+y(x2)y+(h)(y+ψUlog(Uc))(0).\begin{split}y_{nh}(x_{2})=&-\frac{y_{+}(x_{2})}{y_{+}(-h)}\int_{-h}^{x_{2}}\big{(}\frac{y_{-}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}-\frac{y_{-}(x_{2})}{y_{+}(-h)}\int_{x_{2}}^{0}\big{(}\frac{y_{+}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}\\ &+\frac{y_{-}(x_{2})}{y_{+}(-h)}\big{(}\frac{y_{+}\psi}{U^{\prime}}\log(U-c)\big{)}(0).\end{split}

The above two integrals can be estimated similarly and we shall focus on the first one only. Lemma 3.9 implies

|(yψU)log(Uc)|=\displaystyle\big{|}\big{(}\frac{y_{-}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)\big{|}= |U|2|(yψU+yψUyψU)log(Uc)|\displaystyle|U^{\prime}|^{-2}\big{|}(y_{-}^{\prime}\psi U^{\prime}+y_{-}\psi^{\prime}U^{\prime}-y_{-}\psi U^{\prime\prime})\log(U-c)\big{|}
\displaystyle\leq Ccosh(μ1(x2+h))(μ|ψ|+(1+μ|log|Uc||)|ψ|)(1+|log|Uc||).\displaystyle C\cosh(\mu^{-1}(x_{2}+h))\Big{(}\mu|\psi^{\prime}|+\big{(}1+\mu\big{|}\log|U-c|\big{|}\big{)}|\psi|\Big{)}\big{(}1+\big{|}\log|U-c|\big{|}\big{)}.

Using the Hölder inequality we obtain

|hx2(yψU)log(Uc)dx2|\displaystyle\Big{|}\int_{-h}^{x_{2}}\big{(}\frac{y_{-}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}\Big{|}
\displaystyle\leq C(μ|ψ|Lx2r+|ψ|Lx2r)|cosh(μ1(x2+h))(1+|log|U(x2)c||2)|Lx2rr1([h,x2])\displaystyle C(\mu|\psi^{\prime}|_{L_{x_{2}}^{r}}+|\psi|_{L_{x_{2}}^{r}})\big{|}\cosh(\mu^{-1}(x_{2}^{\prime}+h))\big{(}1+\big{|}\log|U(x_{2}^{\prime})-c|\big{|}^{2}\big{)}\big{|}_{L_{x_{2}^{\prime}}^{\frac{r}{r-1}}([-h,x_{2}])}
\displaystyle\leq Cμ11rϵ(|ψ|Lx2r+μ|ψ|Lx2r)coshμ1(x2+h).\displaystyle C\mu^{1-\frac{1}{r}-\epsilon}(|\psi|_{L_{x_{2}}^{r}}+\mu|\psi^{\prime}|_{L_{x_{2}}^{r}})\cosh\mu^{-1}(x_{2}+h).

From the initial condition (3.53) (in particular y+(0)=O(μ2|cU(0)|2)y_{+}(0)=O(\mu^{2}|c-U(0)|^{2})) and (5.5), the remaining boundary term can be estimated as

|y(x2)y+(h)(y+ψUlog(Uc))(0)|Cμ|𝐅(k,c)||ψ(0)|sinhμ1(x2+h).\Big{|}\frac{y_{-}(x_{2})}{y_{+}(-h)}\big{(}\frac{y_{+}\psi}{U^{\prime}}\log(U-c)\big{)}(0)\Big{|}\leq\frac{C\mu}{|\mathbf{F}(k,c)|}|\psi(0)|\sinh\mu^{-1}(x_{2}+h).

The desired estimate on ynhy_{nh} follows from (5.8), (5.5), Lemma 3.9, the above inequalities, and the standard Sobolev inequality

(5.12) |ψ|Lx2C(μ11r|ψ|Lx2r+μ1r|ψ|Lx2r).|\psi|_{L_{x_{2}}^{\infty}}\leq C(\mu^{1-\frac{1}{r}}|\psi^{\prime}|_{L_{x_{2}}^{r}}+\mu^{-\frac{1}{r}}|\psi|_{L_{x_{2}}^{r}}).

The estimate of ynhy_{nh}^{\prime} can be obtained much as in the above. Integrating by parts and using (5.5) to handle the boundary terms at x2x_{2}, we have

(5.13) ynh(x2)=y+(x2)y+(h)hx2(yψU)log(Uc)dx2y(x2)y+(h)x20(y+ψU)log(Uc)dx2(ψUlog(Uc))(x2)+y(x2)y+(h)(y+ψUlog(Uc))(0).\begin{split}y_{nh}^{\prime}(x_{2})=&-\frac{y_{+}^{\prime}(x_{2})}{y_{+}(-h)}\int_{-h}^{x_{2}}\big{(}\frac{y_{-}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}-\frac{y_{-}^{\prime}(x_{2})}{y_{+}(-h)}\int_{x_{2}}^{0}\big{(}\frac{y_{+}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}\\ &-\big{(}\frac{\psi}{U^{\prime}}\log(U-c)\big{)}(x_{2})+\frac{y_{-}^{\prime}(x_{2})}{y_{+}(-h)}\big{(}\frac{y_{+}\psi}{U^{\prime}}\log(U-c)\big{)}(0).\end{split}

The desired estimate on ynhy_{nh}^{\prime} follows from (5.4), (5.12), the above estimate on the integrals, and Lemma 3.9.

To consider the convergence of ynhy_{nh}, we first note that, for cI>0c_{I}>0, the imaginary part of log(U(x2)c)\log(U(x_{2})-c) belongs to (π,0)(-\pi,0) and as cI0+c_{I}\to 0+,

(5.14) log(U(x2)c)log|U(x2)cR|+iπ2(sgn(U(x2)cR)1) in LcRLx2q,q[1,).\log(U(x_{2})-c)\to\log|U(x_{2})-c_{R}|+\tfrac{i\pi}{2}\big{(}sgn(U(x_{2})-c_{R})-1\big{)}\;\text{ in }\;L_{c_{R}}^{\infty}L_{x_{2}}^{q},\;\;\forall q\in[1,\infty).

Using expression (5.11), the estimates thereafter, bounds on y±y_{\pm} in Lemmas 3.9, and the convergence of y±y_{\pm} to y0±y_{0\pm} as cI0c_{I}\to 0 in Lemma 3.12, it is straight forward to obtain

ynh(x2)\displaystyle y_{nh}(x_{2})\to y0+(x2)y0+(h)hx2(y0ψ0U)log|UcR|dx2y0(x2)y0+(h)x20(y0+ψ0U)log|UcR|dx2\displaystyle-\frac{y_{0+}(x_{2})}{y_{0+}(-h)}\int_{-h}^{x_{2}}\big{(}\frac{y_{0-}\psi_{0}}{U^{\prime}}\big{)}^{\prime}\log|U-c_{R}|dx_{2}^{\prime}-\frac{y_{0-}(x_{2})}{y_{0+}(-h)}\int_{x_{2}}^{0}\big{(}\frac{y_{0+}\psi_{0}}{U^{\prime}}\big{)}^{\prime}\log|U-c_{R}|dx_{2}^{\prime}
+iπψ0(x2c)U(x2c)(y0+(x2)y0(x2c)y0+(h)χ{U(x2)>cR>U(h)}+y0(x2)y0+(x2c)y0+(h)χ{U(0)>cR>U(x2)})\displaystyle+\frac{i\pi\psi_{0}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\Big{(}\frac{y_{0+}(x_{2})y_{0-}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(x_{2})>c_{R}>U(-h)\}}+\frac{y_{0-}(x_{2})y_{0+}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(0)>c_{R}>U(x_{2})\}}\Big{)}
+y0(x2)y0+(h)(y0+ψ0Ulog|UcR|)(0),\displaystyle+\frac{y_{0-}(x_{2})}{y_{0+}(-h)}\big{(}\frac{y_{0+}\psi_{0}}{U^{\prime}}\log|U-c_{R}|\big{)}(0),

in LcRq1Lx2L_{c_{R}}^{q_{1}}L_{x_{2}}^{\infty} for any q1[1,r1)q_{1}\in[1,r_{1}), where, for cR>U(0)c_{R}>U(0), two other terms involving sgn(UcR)sgn(U-c_{R}) (one from upper limit term from the second integral and the other from the boundary term in (5.11)) cancelled each other. Here the loss of the integrability in cRc_{R} in the convergence is due to the last logarithmic term. Since (log|UcR|)=P.V.UUcR(\log|U-c_{R}|)^{\prime}=P.V.\tfrac{U^{\prime}}{U-c_{R}} in the distribution sense, the above limit is equal to ynh0y_{nh0} after integration by parts. The convergence of ynhy_{nh}^{\prime} is obtained using (5.13) along with (5.5) in a similar fashion

ynh(x2\displaystyle y_{nh}^{\prime}(x_{2} )y0+(x2)y0+(h)hx2(y0ψ0U)log|UcR|dx2y0(x2)y0+(h)x20(y0+ψ0U)log|UcR|dx2\displaystyle)\to-\frac{y_{0+}^{\prime}(x_{2})}{y_{0+}(-h)}\int_{-h}^{x_{2}}\big{(}\frac{y_{0-}\psi_{0}}{U^{\prime}}\big{)}^{\prime}\log|U-c_{R}|dx_{2}^{\prime}-\frac{y_{0-}^{\prime}(x_{2})}{y_{0+}(-h)}\int_{x_{2}}^{0}\big{(}\frac{y_{0+}\psi_{0}}{U^{\prime}}\big{)}^{\prime}\log|U-c_{R}|dx_{2}^{\prime}
+iπ(y0+(x2)y0(x2c)ψ0(x2c)y0+(h)U(x2c)χ{U(x2)>cR>U(h)}ψ0(x2)U(x2)χ{cR>U(x2)}\displaystyle+i\pi\Big{(}\frac{y_{0+}^{\prime}(x_{2})y_{0-}(x_{2}^{c})\psi_{0}(x_{2}^{c})}{y_{0+}(-h)U^{\prime}(x_{2}^{c})}\chi_{\{U(x_{2})>c_{R}>U(-h)\}}-\frac{\psi_{0}(x_{2})}{U^{\prime}(x_{2})}\chi_{\{c_{R}>U(x_{2})\}}
+y0(x2)y0+(x2c)ψ0(x2c)y0+(h)U(x2c)χ{U(0)>cR>U(x2)})((ψ0Ulog|UcR|)(x2)iπψ0(x2)U(x2)χ{cR>U(x2)})\displaystyle+\frac{y_{0-}^{\prime}(x_{2})y_{0+}(x_{2}^{c})\psi_{0}(x_{2}^{c})}{y_{0+}(-h)U^{\prime}(x_{2}^{c})}\chi_{\{U(0)>c_{R}>U(x_{2})\}}\Big{)}-\big{(}\big{(}\frac{\psi_{0}}{U^{\prime}}\log|U-c_{R}|\big{)}(x_{2})-i\pi\frac{\psi_{0}(x_{2})}{U^{\prime}(x_{2})}\chi_{\{c_{R}>U(x_{2})\}}\big{)}
+y0(x2)y0+(h)(y0+ψ0Ulog|UcR|)(0)\displaystyle+\frac{y_{0-}^{\prime}(x_{2})}{y_{0+}(-h)}\big{(}\frac{y_{0+}\psi_{0}}{U^{\prime}}\log|U-c_{R}|\big{)}(0)

where again two other terms involving sgn(UcR)sgn(U-c_{R}) cancelled each other for cR>U(0)c_{R}>U(0). Here the convergence in the slightly weaker norm LcRq1Lx2q2L_{c_{R}}^{q_{1}}L_{x_{2}}^{q_{2}}, for any q1[1,r1)q_{1}\in[1,r_{1}) and q2[1,)q_{2}\in[1,\infty) is due to the logarithmic singularity both explicitly outside the integrals and in y±y_{\pm}^{\prime} (see also Lemma 3.12). The limit can be simplified to

y0+(x2)y0+(h)hx2(y0ψ0U)log|UcR|dx2y0(x2)y0+(h)x20(y0+ψ0U)log|UcR|dx2\displaystyle-\frac{y_{0+}^{\prime}(x_{2})}{y_{0+}(-h)}\int_{-h}^{x_{2}}\big{(}\frac{y_{0-}\psi_{0}}{U^{\prime}}\big{)}^{\prime}\log|U-c_{R}|dx_{2}^{\prime}-\frac{y_{0-}^{\prime}(x_{2})}{y_{0+}(-h)}\int_{x_{2}}^{0}\big{(}\frac{y_{0+}\psi_{0}}{U^{\prime}}\big{)}^{\prime}\log|U-c_{R}|dx_{2}^{\prime}
+iπψ0(x2c)U(x2c)(y0+(x2)y0(x2c)y0+(h)χ{U(x2)>cR>U(h)}+y0(x2)y0+(x2c)y0+(h)χ{U(0)>cR>U(x2)})\displaystyle+\frac{i\pi\psi_{0}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\Big{(}\frac{y_{0+}^{\prime}(x_{2})y_{0-}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(x_{2})>c_{R}>U(-h)\}}+\frac{y_{0-}^{\prime}(x_{2})y_{0+}(x_{2}^{c})}{y_{0+}(-h)}\chi_{\{U(0)>c_{R}>U(x_{2})\}}\Big{)}
(ψ0Ulog|UcR|)(x2)+y0(x2)y0+(h)(y0+ψ0Ulog|UcR|)(0),\displaystyle-\big{(}\frac{\psi_{0}}{U^{\prime}}\log|U-c_{R}|\big{)}(x_{2})+\frac{y_{0-}^{\prime}(x_{2})}{y_{0+}(-h)}\big{(}\frac{y_{0+}\psi_{0}}{U^{\prime}}\log|U-c_{R}|\big{)}(0),

which is equal to ynh0y_{nh0}^{\prime} after an integration by parts.

At the end point x2=h,0x_{2}=-h,0, ynh(x~2)y_{nh}(\tilde{x}_{2}) and ynh(x~2)y_{nh}^{\prime}(\tilde{x}_{2}) have only one integrals and, unlike for general x2(h,0)x_{2}\in(-h,0), the terms y+(0)y_{+}(0), y+(0)y_{+}^{\prime}(0) and y(h)y_{-}^{\prime}(-h) outside the integrals are prescribed in (3.53) without any singularity. Hence the same above argument yields slightly better estimates and convergence. One may make the following computations using (5.3) and (5.4),

ynh(0)=y+(0)y+(h)h0yψUcdx2=y+(0)y+(h)h0(yψU)log(Uc)dx2+(y+yψ)(0)U(0)y+(h)log(U(0)c),y_{nh}^{\prime}(0)=\frac{y_{+}^{\prime}(0)}{y_{+}(-h)}\int_{-h}^{0}\frac{y_{-}\psi}{U-c}dx_{2}^{\prime}=-\frac{y_{+}^{\prime}(0)}{y_{+}(-h)}\int_{-h}^{0}\big{(}\frac{y_{-}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}+\frac{(y_{+}^{\prime}y_{-}\psi)(0)}{U^{\prime}(0)y_{+}(-h)}\log(U(0)-c),
ynh(h)=1y+(h)h0y+ψUcdx2=\displaystyle y_{nh}^{\prime}(-h)=\frac{1}{y_{+}(-h)}\int_{-h}^{0}\frac{y_{+}\psi}{U-c}dx_{2}^{\prime}= 1y+(h)h0(y+ψU)log(Uc)dx2\displaystyle-\frac{1}{y_{+}(-h)}\int_{-h}^{0}\big{(}\frac{y_{+}\psi}{U^{\prime}}\big{)}^{\prime}\log(U-c)dx_{2}^{\prime}
+1y+(h)(y+ψUlog(Uc))|h0.\displaystyle+\frac{1}{y_{+}(-h)}\Big{(}\frac{y_{+}\psi}{U^{\prime}}\log(U-c)\Big{)}\Big{|}_{-h}^{0}.

The desired inequalities follow from (3.53) and the above estimates, which completes the proof of the lemma. ∎

Assuming ψLcR2Hx21\psi\in L_{c_{R}}^{2}H_{x_{2}}^{1}, in the following we estimate ynhy_{nh} and ynhy_{nh}^{\prime} as well as their derivatives in x2x_{2} in LcR,x22L_{c_{R},x_{2}}^{2}, in particular their dependence on kk, by an energy estimate approach.

Lemma 5.2.

Assume (5.8). For any ϵ(0,1)\epsilon\in(0,1), there exists C>0C>0 depending only on ϵ\epsilon, F0F_{0}, ρ0\rho_{0}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any cI0c_{I}\geq 0 and kk\in\mathbb{R}, it holds

(5.15) |ynh|LcR,x222+μ2|ynh|LcR,x222C(|ψ|LcR,x222+μ2ϵ|ψ|LcR,x222),|y_{nh}^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}+\mu^{-2}|y_{nh}|_{L_{c_{R},x_{2}}^{2}}^{2}\leq C(|\psi|_{L_{c_{R},x_{2}}^{2}}^{2}+\mu^{2-\epsilon}|\psi^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}),

where the norms are taken for cRc_{R}\in\mathcal{I} and x2[h,0]x_{2}\in[-h,0].

Proof.

We first assume cI>0c_{I}>0 and drop the subscript nh\cdot_{nh} for notation simplification. Multiplying the Rayleigh equation (5.1a) by y¯\bar{y} and integrating in both cRc_{R} and x2x_{2}, we have

h0|y|2+k2|y|2dx2dcR=yy¯dcR|x2=0+h0ψy¯U|y|2Ucdx2dcR\displaystyle\quad\int_{\mathcal{I}}\int_{-h}^{0}|y^{\prime}|^{2}+k^{2}|y|^{2}dx_{2}dc_{R}=\int_{\mathcal{I}}y^{\prime}\bar{y}dc_{R}\Big{|}_{x_{2}=0}+\int_{\mathcal{I}}\int_{-h}^{0}\frac{\psi\bar{y}-U^{\prime\prime}|y|^{2}}{U-c}dx_{2}dc_{R}
=yy¯dcR|x2=0+h0UUc((ψy¯U|y|2U)(c,x2)(ψy¯U|y|2U)(c,x2c))dx2dcR\displaystyle=\int_{\mathcal{I}}y^{\prime}\bar{y}dc_{R}\Big{|}_{x_{2}=0}+\int_{\mathcal{I}}\int_{-h}^{0}\frac{U^{\prime}}{U-c}\Big{(}\big{(}\frac{\psi\bar{y}-U^{\prime\prime}|y|^{2}}{U^{\prime}}\big{)}\big{(}c,x_{2})-\big{(}\frac{\psi\bar{y}-U^{\prime\prime}|y|^{2}}{U^{\prime}}\big{)}\big{(}c,x_{2}^{c})\Big{)}dx_{2}dc_{R}
+(U(h12ρ0)U(12h)+U(12h)U(12ρ0))(ψy¯U|y|2U)(c,x2c)(log(U(0)c)log(U(h)c))dcRj=14Aj.\displaystyle\;+\left(\int_{U(-h-\frac{1}{2}\rho_{0})}^{U(-\frac{1}{2}h)}+\int_{U(-\frac{1}{2}h)}^{U(\frac{1}{2}\rho_{0})}\right)\big{(}\frac{\psi\bar{y}-U^{\prime\prime}|y|^{2}}{U^{\prime}}\big{)}\big{(}c,x_{2}^{c})\big{(}\log(U(0)-c)-\log(U(-h)-c)\big{)}dc_{R}\triangleq\sum_{j=1}^{4}A_{j}.

The first term A1A_{1} of boundary contribution can be estimated by Lemma 5.1(2b) and (5.7) with ζ±=0\zeta_{\pm}=0, as well as (5.6), (5.8) and (5.12),

|A1||yy¯dcR|x2=0|Cμ2||U(0)c|2|y(0)|2dcR|Cμ1ϵ(μ|ψ|LcR,x22+|ψ|LcR,x22)2.|A_{1}|\leq\Big{|}\int_{\mathcal{I}}y^{\prime}\bar{y}dc_{R}\Big{|}_{x_{2}=0}\Big{|}\leq C\mu^{2}\Big{|}\int_{\mathcal{I}}|U(0)-c|^{2}|y^{\prime}(0)|^{2}dc_{R}\Big{|}\leq C\mu^{1-\epsilon}(\mu|\psi^{\prime}|_{L_{c_{R},x_{2}}^{2}}+|\psi|_{L_{c_{R},x_{2}}^{2}})^{2}.

Concerning the last integral A4A_{4}, we first split it as

|A4|U(12h)U(12ρ0)(|(ψy¯U|y|2)(c,)|Cx2α|x2c|α+|(ψy¯U|y|2)(c,0)|)(1+|log|U(0)c||)dcR.|A_{4}|\leq\int_{U(-\frac{1}{2}h)}^{U(\frac{1}{2}\rho_{0})}\Big{(}\big{|}(\psi\bar{y}-U^{\prime\prime}|y|^{2})(c,\cdot)\big{|}_{C_{x_{2}}^{\alpha}}|x_{2}^{c}|^{\alpha}+\big{|}(\psi\bar{y}-U^{\prime\prime}|y|^{2})(c,0)\big{|}\Big{)}\big{(}1+\big{|}\log|U(0)-c|\big{|}\big{)}dc_{R}.

The above terms at x2=0x_{2}=0 can estimated much as A1A_{1} and we obtain

U(12h)U(12ρ0)|(ψy¯U|y|2)(c,0)|(1+|log|U(0)c||)dcR\displaystyle\int_{U(-\frac{1}{2}h)}^{U(\frac{1}{2}\rho_{0})}\big{|}(\psi\bar{y}-U^{\prime\prime}|y|^{2})(c,0)\big{|}\big{(}1+\big{|}\log|U(0)-c|\big{|}\big{)}dc_{R}
\displaystyle\leq CU(12h)U(12ρ0)(μ2|ψ|2+μ2|y|2)(c,0)|U(0)c|dcRCμ1ϵ(μ|ψ|LcR,x22+|ψ|LcR,x22)2.\displaystyle C\int_{U(-\frac{1}{2}h)}^{U(\frac{1}{2}\rho_{0})}(\mu^{2}|\psi|^{2}+\mu^{2}|y^{\prime}|^{2})(c,0)|U(0)-c|dc_{R}\leq C\mu^{1-\epsilon}(\mu|\psi^{\prime}|_{L_{c_{R},x_{2}}^{2}}+|\psi|_{L_{c_{R},x_{2}}^{2}})^{2}.

We shall estimate all the remaining terms using the Hölder norms of ψy¯\psi\bar{y} and |y|2|y|^{2}. For any H1H^{1} function f(x)f(x) on an interval, it holds

(5.16) |f|CαC|f|L212α|f|H112+α,α[0,12],|f|_{C^{\alpha}}\leq C|f|_{L^{2}}^{\frac{1}{2}-\alpha}|f|_{H^{1}}^{\frac{1}{2}+\alpha},\quad\alpha\in[0,\tfrac{1}{2}],

which applies to ψy¯\psi\bar{y} and |y|2|y|^{2}. In the f|H1f|_{H^{1}} can be replaced by |f|L2|f^{\prime}|_{L^{2}} if ff vanishes somewhere in the interval. Hence for each fixed cc with cI>0c_{I}>0 and cRc_{R}\in\mathcal{I},

||y|2|Cx2αC|y|Cx2α|y¯|Cx20C|y|Lx221α|y|Lx221+α,\big{|}|y|^{2}\big{|}_{C_{x_{2}}^{\alpha}}\leq C|y|_{C_{x_{2}}^{\alpha}}|\bar{y}|_{C_{x_{2}}^{0}}\leq C|y|_{L_{x_{2}}^{2}}^{1-\alpha}|y^{\prime}|_{L_{x_{2}}^{2}}^{1+\alpha},
|ψy¯|Cx2αC(|ψ|Lx2212α|ψ|Hx2112+α|y|Lx2212|y|Lx2212+|ψ|Lx2212|ψ|Hx2112|y|Lx2212α|y|Lx2212+α).|\psi\bar{y}|_{C_{x_{2}}^{\alpha}}\leq C\big{(}|\psi|_{L_{x_{2}}^{2}}^{\frac{1}{2}-\alpha}|\psi|_{H_{x_{2}}^{1}}^{\frac{1}{2}+\alpha}|y|_{L_{x_{2}}^{2}}^{\frac{1}{2}}|y^{\prime}|_{L_{x_{2}}^{2}}^{\frac{1}{2}}+|\psi|_{L_{x_{2}}^{2}}^{\frac{1}{2}}|\psi|_{H_{x_{2}}^{1}}^{\frac{1}{2}}|y|_{L_{x_{2}}^{2}}^{\frac{1}{2}-\alpha}|y^{\prime}|_{L_{x_{2}}^{2}}^{\frac{1}{2}+\alpha}\big{)}.

For any α(0,12]\alpha\in(0,\tfrac{1}{2}] and k>0k>0, using y(c,h)=0y(c,-h)=0 and the above estimates, we obtain

|y|LcR,x222+k2|y|LcR,x222Ch0|(ψy¯U|y|2)(c,)|Cx2α|x2x2c|α1dx2dcR\displaystyle|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}+k^{2}|y|_{L_{c_{R},x_{2}}^{2}}^{2}\leq C\int_{\mathcal{I}}\int_{-h}^{0}\big{|}(\psi\bar{y}-U^{\prime\prime}|y|^{2})(c,\cdot)\big{|}_{C_{x_{2}}^{\alpha}}|x_{2}-x_{2}^{c}|^{\alpha-1}dx_{2}dc_{R}
+CU(h12ρ0)U(12h)|(ψy¯U|y|2)(c,)|Cx2α|x2c+h|α(1+|log|U(h)c||)dcR\displaystyle\qquad\qquad+C\int_{U(-h-\frac{1}{2}\rho_{0})}^{U(-\frac{1}{2}h)}\big{|}(\psi\bar{y}-U^{\prime\prime}|y|^{2})(c,\cdot)\big{|}_{C_{x_{2}}^{\alpha}}|x_{2}^{c}+h|^{\alpha}\big{(}1+\big{|}\log|U(-h)-c|\big{|}\big{)}dc_{R}
+CU(12h)U(12ρ0)|(ψy¯U|y|2)(c,)|Cx2α|x2c|α(1+|log|U(0)c||)dcR\displaystyle\qquad\qquad+C\int_{U(-\frac{1}{2}h)}^{U(\frac{1}{2}\rho_{0})}\big{|}(\psi\bar{y}-U^{\prime\prime}|y|^{2})(c,\cdot)\big{|}_{C_{x_{2}}^{\alpha}}|x_{2}^{c}|^{\alpha}\big{(}1+\big{|}\log|U(0)-c|\big{|}\big{)}dc_{R}
+Cμ1ϵ(μ|ψ|LcR,x22+|ψ|LcR,x22)2\displaystyle\qquad\qquad+C\mu^{1-\epsilon}(\mu|\psi^{\prime}|_{L_{c_{R},x_{2}}^{2}}+|\psi|_{L_{c_{R},x_{2}}^{2}})^{2}
\displaystyle\leq C(|ψ|Lx2212α|ψ|Hx2112+α|y|Lx2212|y|Lx2212+|ψ|Lx2212|ψ|Hx2112|y|Lx2212α|y|Lx2212+α+|y|Lx221α|y|Lx221+α)|cdcR\displaystyle C\int_{\mathcal{I}}\big{(}|\psi|_{L_{x_{2}}^{2}}^{\frac{1}{2}-\alpha}|\psi|_{H_{x_{2}}^{1}}^{\frac{1}{2}+\alpha}|y|_{L_{x_{2}}^{2}}^{\frac{1}{2}}|y^{\prime}|_{L_{x_{2}}^{2}}^{\frac{1}{2}}+|\psi|_{L_{x_{2}}^{2}}^{\frac{1}{2}}|\psi|_{H_{x_{2}}^{1}}^{\frac{1}{2}}|y|_{L_{x_{2}}^{2}}^{\frac{1}{2}-\alpha}|y^{\prime}|_{L_{x_{2}}^{2}}^{\frac{1}{2}+\alpha}+|y|_{L_{x_{2}}^{2}}^{1-\alpha}|y^{\prime}|_{L_{x_{2}}^{2}}^{1+\alpha}\big{)}\big{|}_{c}dc_{R}
+Cμ1ϵ(μ|ψ|LcR,x22+|ψ|LcR,x22)2\displaystyle+C\mu^{1-\epsilon}(\mu|\psi^{\prime}|_{L_{c_{R},x_{2}}^{2}}+|\psi|_{L_{c_{R},x_{2}}^{2}})^{2}
\displaystyle\leq C(|ψ|LcR,x2212α|ψ|LcR2Hx2112+α|y|LcR,x2212|y|LcR,x2212+|ψ|LcR,x2212|ψ|LcR2Hx2112|y|LcR,x2212α|y|LcR,x2212+α+|y|LcR,x221α|y|LcR,x221+α)\displaystyle C\big{(}|\psi|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}-\alpha}|\psi|_{L_{c_{R}}^{2}H_{x_{2}}^{1}}^{\frac{1}{2}+\alpha}|y|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}}+|\psi|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}}|\psi|_{L_{c_{R}}^{2}H_{x_{2}}^{1}}^{\frac{1}{2}}|y|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}-\alpha}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}+\alpha}+|y|_{L_{c_{R},x_{2}}^{2}}^{1-\alpha}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{1+\alpha}\big{)}
+Cμ1ϵ(μ|ψ|LcR,x22+|ψ|LcR,x22)2\displaystyle+C\mu^{1-\epsilon}(\mu|\psi^{\prime}|_{L_{c_{R},x_{2}}^{2}}+|\psi|_{L_{c_{R},x_{2}}^{2}})^{2}
\displaystyle\leq 12|y|LcR,x222+(C+12k2)|y|LcR,x222+C(|ψ|LcR,x222+k2(12α)|ψ|LcR,x222).\displaystyle\tfrac{1}{2}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}+(C+\tfrac{1}{2}k^{2})|y|_{L_{c_{R},x_{2}}^{2}}^{2}+C\big{(}|\psi|_{L_{c_{R},x_{2}}^{2}}^{2}+k^{-2(1-2\alpha)}|\psi^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}\big{)}.

By choosing α=ϵ/2\alpha=\epsilon/2, we have that, there exists k0>0k_{0}>0 such that for any |k|k0|k|\geq k_{0} and cI>0c_{I}>0, y(+icI,)y(\cdot+ic_{I},\cdot) satisfies (5.15). To obtain the estimates for ynh0y_{nh0} and ynh0y_{nh0}^{\prime} in the limiting case cI=0+c_{I}=0+, for cI>0c_{I}>0, let y(c,x2)y(c,x_{2}) and y(c,x2)y^{\prime}(c,x_{2}) be defined by (5.3) and (5.4), which satisfy the desired estimates uniform in cI>0c_{I}>0. For |k|k0|k|\leq k_{0} and cI>0c_{I}>0, the desired estimates simply follows from the estimates and convergence obtained in Lemma 5.1.

Finally we consider the case cI=0c_{I}=0. Given ψ(cR,x2)LcR2Hx21\psi(c_{R},x_{2})\in L_{c_{R}}^{2}H_{x_{2}}^{1}, let ynh(k,cR+icI,x2)y_{nh}(k,c_{R}+ic_{I},x_{2}) be given by (5.3) with c=cR+icIc=c_{R}+ic_{I} with 1cI>01\gg c_{I}>0, which solves (5.1a). From Lemma 5.1, it holds that y(+icI,)ynh0y(\cdot+ic_{I},\cdot)\to y_{nh0} and y(+icI,)ynh0y^{\prime}(\cdot+ic_{I},\cdot)\to y_{nh0}^{\prime} in LcR32Lx22L_{c_{R}}^{\frac{3}{2}}L_{x_{2}}^{2} as cI0+c_{I}\to 0+. Therefore ynh0y_{nh0} and ynh0y_{nh0}^{\prime} are also the weak limit of yy and yy^{\prime} in LcR,x22L_{c_{R},x_{2}}^{2} as cI0+c_{I}\to 0+ and thus also satisfy (5.15). ∎

\bullet Case 2:

(5.17) ψ(c,x2)=f(c,x2)ψ0(x2),f(+icI,)LcRr1Cx2α,ψ0Lr,r>1,r1[rr1,],α>0.\psi(c,x_{2})=f(c,x_{2})\psi_{0}(x_{2}),\quad f(\cdot+ic_{I},\cdot)\in L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha},\;\psi_{0}\in L^{r},\;r>1,\;r_{1}\in[\tfrac{r}{r-1},\infty],\;\alpha>0.

Again we start with rough estimates on ynhy_{nh} and ynhy_{nh}^{\prime}.

Lemma 5.3.

Assume (5.8) and (5.17). For any q[1,rr1r+r1)q\in[1,\tfrac{rr_{1}}{r+r_{1}}), the following hold for x2[h,0]x_{2}\in[-h,0] and cRc_{R}\in\mathcal{I}.

  1. (1)

    There exists C>0C>0 depending only on rr, r1r_{1}, qq, α\alpha, F0F_{0}, ρ0\rho_{0}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any kk\in\mathbb{R} and cI(0,ρ0]c_{I}\in(0,\rho_{0}], it holds

    |ynh(k,+icI,)|Lx2LcRq+μ1|ynh(k,+icI,)|Lx2LcRrr1r+r1Cμα|f(+icI,)|LcRr1Cx2α|ψ|Lr.|y_{nh}^{\prime}(k,\cdot+ic_{I},\cdot)|_{L_{x_{2}}^{\infty}L_{c_{R}}^{q}}+\mu^{-1}|y_{nh}(k,\cdot+ic_{I},\cdot)|_{L_{x_{2}}^{\infty}L_{c_{R}}^{\frac{rr_{1}}{r+r_{1}}}}\leq C\mu^{-\alpha}|f(\cdot+ic_{I},\cdot)|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha}}|\psi|_{L^{r}}.
  2. (2)

    Assume f(+icI,)f0(,)f(\cdot+ic_{I},\cdot)\to f_{0}(\cdot,\cdot) in LcRr1Cx2αL_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha} as cI0+c_{I}\to 0+, then

    1. (a)

      ynhynh0y_{nh}\to y_{nh0} in Lx2LcRrr1r+r1L_{x_{2}}^{\infty}L_{c_{R}}^{\frac{rr_{1}}{r+r_{1}}} and ynhynh0y_{nh}^{\prime}\to y_{nh0}^{\prime} in Lx2LcRqL_{x_{2}}^{\infty}L_{c_{R}}^{q}, where ynh0y_{nh0} and ynh0y_{nh0}^{\prime} are given by (5.9) and (5.10) with ψ0\psi_{0} replaced by f0ψ0f_{0}\psi_{0};

    2. (b)

      at x~2=h,0\tilde{x}_{2}=-h,0, ynh(k,+icI,x~2)ynh0(k,,x~2)y_{nh}^{\prime}(k,\cdot+ic_{I},\tilde{x}_{2})\to y_{nh0}^{\prime}(k,\cdot,\tilde{x}_{2}) in LcRrr1r+r1L_{c_{R}}^{\frac{rr_{1}}{r+r_{1}}}. Moreover, and for any ϵ(0,1r)\epsilon\in(0,\frac{1}{r}) with ϵα\epsilon\leq\alpha, for any kk\in\mathbb{R}, cI0c_{I}\geq 0, it holds

      |ynh(+icI,x~2)|LcRrr1r+r1Cμϵ|f|LcRr1Cx2α|ψ|Lr,|y_{nh}^{\prime}(\cdot+ic_{I},\tilde{x}_{2})|_{L_{c_{R}}^{\frac{rr_{1}}{r+r_{1}}}}\leq C\mu^{-\epsilon}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha}}|\psi|_{L^{r}},

      where CC also depends on ϵ>0\epsilon>0.

Proof.

Since the desired estimates are stronger and with weaker assumptions if α(0,1)\alpha\in(0,1) is smaller (with possibly greater C>0C>0), without loss of generality, we may assume α<1r\alpha<\tfrac{1}{r}. In the following we shall need the modification x~2c\tilde{x}_{2}^{c} determined by cRc_{R}\in\mathcal{I}:

(5.18) x~2c(c,x2)={min{x2,x2c},if cR>U(h),h,if cRU(h),,x~2+c(c,x2)={max{x2,x2c},if cR<U(0),0,if cRU(0).\tilde{x}_{2-}^{c}(c,x_{2})=\begin{cases}\min\{x_{2},x_{2}^{c}\},&\text{if }c_{R}>U(-h),\\ -h,&\text{if }c_{R}\leq U(-h),\end{cases},\;\,\tilde{x}_{2+}^{c}(c,x_{2})=\begin{cases}\max\{x_{2},x_{2}^{c}\},&\text{if }c_{R}<U(0),\\ 0,&\text{if }c_{R}\geq U(0).\end{cases}

For cI>0c_{I}>0, we first split ynhy_{nh} into

y1(x2)=\displaystyle y_{1}(x_{2})= y+(x2)y+(h)(yfU)(x~2c)hx2ψ0UUcdx2+y(x2)y+(h)(y+fU)(x~2+c)x20ψ0UUcdx2,\displaystyle\frac{y_{+}(x_{2})}{y_{+}(-h)}\big{(}\frac{y_{-}f}{U^{\prime}}\big{)}(\tilde{x}_{2-}^{c})\int_{-h}^{x_{2}}\frac{\psi_{0}U^{\prime}}{U-c}dx_{2}^{\prime}+\frac{y_{-}(x_{2})}{y_{+}(-h)}\big{(}\frac{y_{+}f}{U^{\prime}}\big{)}(\tilde{x}_{2+}^{c})\int_{x_{2}}^{0}\frac{\psi_{0}U^{\prime}}{U-c}dx_{2}^{\prime},

and

y2(x2)=\displaystyle y_{2}(x_{2})= y+(x2)y+(h)hx2((yfU)(x2)(yfU)(x~2c))ψ0UUcdx2\displaystyle\frac{y_{+}(x_{2})}{y_{+}(-h)}\int_{-h}^{x_{2}}\Big{(}\big{(}\frac{y_{-}f}{U^{\prime}}\big{)}(x_{2}^{\prime})-\big{(}\frac{y_{-}f}{U^{\prime}}\big{)}(\tilde{x}_{2-}^{c})\Big{)}\frac{\psi_{0}U^{\prime}}{U-c}dx_{2}^{\prime}
+y(x2)y+(h)x20((y+fU)(x2)(y+fU)(x~2+c))ψ0UUcdx2,\displaystyle+\frac{y_{-}(x_{2})}{y_{+}(-h)}\int_{x_{2}}^{0}\Big{(}\big{(}\frac{y_{+}f}{U^{\prime}}\big{)}(x_{2}^{\prime})-\big{(}\frac{y_{+}f}{U^{\prime}}\big{)}(\tilde{x}_{2+}^{c})\Big{)}\frac{\psi_{0}U^{\prime}}{U-c}dx_{2}^{\prime},

where we skipped all the dependence on cc and kk. Clearly ynh=y1+y2y_{nh}=y_{1}+y_{2}.

To estimate y1y_{1}, we can rewrite its integral part as

hx2ψ0UUcdx2=χU([h,x2])(ψ0U1)τcRicIdτ=((1τ+icI)ψ~(x2,))(cR).\int_{-h}^{x_{2}}\frac{\psi_{0}U^{\prime}}{U-c}dx_{2}^{\prime}=\int_{\mathbb{R}}\frac{\chi_{U([-h,x_{2}])}(\psi_{0}\circ U^{-1})}{\tau-c_{R}-ic_{I}}d\tau=-\Big{(}\big{(}\frac{1}{\tau+ic_{I}}\big{)}*\tilde{\psi}_{-}(x_{2},\cdot)\Big{)}(c_{R}).

where

ψ~(x2,τ)=χU([h,x2])(ψ0U1)(τ),ψ~+(x2,τ)=χU([x2,0])(ψ0U1)(τ).\tilde{\psi}_{-}(x_{2},\tau)=\chi_{U([-h,x_{2}])}(\psi_{0}\circ U^{-1})(\tau),\quad\tilde{\psi}_{+}(x_{2},\tau)=\chi_{U([x_{2},0])}(\psi_{0}\circ U^{-1})(\tau).

The operator of convolution by 1τ+icI\tfrac{1}{\tau+ic_{I}} is bounded on LrL^{r} uniformly in cI>0c_{I}>0 and converges to π(+iI)\pi(\mathcal{H}+iI) strongly in LrL^{r} as cI0+c_{I}\to 0+, where \mathcal{H} is the Hilbert transform and II is the identity. The other integral can be treated similarly and we obtain from (5.8) and Lemma 3.9

|y1|Lx2LcRrr1r+r1\displaystyle|y_{1}|_{L_{x_{2}}^{\infty}L_{c_{R}}^{\frac{rr_{1}}{r+r_{1}}}}\leq C(|y+(x2)y(x~2c)y+(h)|LcR,x2+|y(x2)y+(x~2+)y+(h)|LcR,x2)|f|LcRr1Lx2|ψ0|LrCμ|f|LcRr1Lx2|ψ0|Lr.\displaystyle C\big{(}\big{|}\tfrac{y_{+}(x_{2})y_{-}(\tilde{x}_{2-}^{c})}{y_{+}(-h)}\big{|}_{L_{c_{R},x_{2}}^{\infty}}+\big{|}\tfrac{y_{-}(x_{2})y_{+}(\tilde{x}_{2+})}{y_{+}(-h)}\big{|}_{L_{c_{R},x_{2}}^{\infty}}\big{)}|f|_{L_{c_{R}}^{r_{1}}L_{x_{2}}^{\infty}}|\psi_{0}|_{L^{r}}\leq C\mu|f|_{L_{c_{R}}^{r_{1}}L_{x_{2}}^{\infty}}|\psi_{0}|_{L^{r}}.

Moreover, since x2ψ~±(x2,)x_{2}\to\tilde{\psi}_{\pm}(x_{2},\cdot) are two uniformly continuous mapping from [h,0][-h,0] to Lr()L^{r}(\mathbb{R}) and the above convolution (1τ+icI)\big{(}\frac{1}{\tau+ic_{I}}\big{)}* is bounded on LcRr()L_{c_{R}}^{r}(\mathbb{R}) uniformly in cI>0c_{I}>0, we have that (1τ+icI)ψ~±(x2,)\big{(}\frac{1}{\tau+ic_{I}}\big{)}*\tilde{\psi}_{\pm}(x_{2},\cdot) are two families (with parameter cIc_{I}) of equicontinuous functions (of x2x_{2}) from [h,0][-h,0] to LcRrL_{c_{R}}^{r}. As cI0+c_{I}\to 0+, they converge pointwisely (in x2x_{2}) to π(+iI)ψ~±(x2,)LcRr\pi(\mathcal{H}+iI)\tilde{\psi}_{\pm}(x_{2},\cdot)\in L_{c_{R}}^{r} which are also uniformly continuous in x2x_{2}. The equicontinuity and the compactness of [h,0][-h,0] imply that the convergence is uniform in x2x_{2}. Therefore, along with the LcR,x2L_{c_{R},x_{2}}^{\infty} convergence of y±y_{\pm} as cI0+c_{I}\to 0+ (Lemma 3.12), we obtain that, as cI0+c_{I}\to 0+,

y1(cR+icI,x2)\displaystyle y_{1}(c_{R}+ic_{I},x_{2})\to πy0+(cR,x2)y0+(cR,h)(y0f0U)(cR,x~2c)((+iI)ψ~(x2,))(cR)\displaystyle\pi\frac{y_{0+}(c_{R},x_{2})}{y_{0+}(c_{R},-h)}\big{(}\frac{y_{0-}f_{0}}{U^{\prime}}\big{)}(c_{R},\tilde{x}_{2-}^{c})\big{(}(\mathcal{H}+iI)\tilde{\psi}_{-}(x_{2},\cdot)\big{)}(c_{R})
+πy0(cR,x2)y0+(cR,h)(y0+f0U)(cR,x~2+c)((+iI)ψ~+(x2,))(cR) in Lx2LcRrr1r+r1.\displaystyle+\pi\frac{y_{0-}(c_{R},x_{2})}{y_{0+}(c_{R},-h)}\big{(}\frac{y_{0+}f_{0}}{U^{\prime}}\big{)}(c_{R},\tilde{x}_{2+}^{c})\big{(}(\mathcal{H}+iI)\tilde{\psi}_{+}(x_{2},\cdot)\big{)}(c_{R})\quad\text{ in }L_{x_{2}}^{\infty}L_{c_{R}}^{\frac{rr_{1}}{r+r_{1}}}.

The other part y2y_{2} can be estimated by the Hölder continuity of ff and y±y_{\pm} in x2x_{2} as

|y2(c,x2)|\displaystyle|y_{2}(c,x_{2})|\leq C(|y+(x2)y+(h)||yf|Cx2α([h,x2])hx2|UU(x~2c)|α|Uc||ψ0|Udx2\displaystyle C\Big{(}\Big{|}\frac{y_{+}(x_{2})}{y_{+}(-h)}\Big{|}|y_{-}f|_{C_{x_{2}^{\prime}}^{\alpha}([-h,x_{2}])}\int_{-h}^{x_{2}}\frac{|U-U(\tilde{x}_{2-}^{c})|^{\alpha}}{|U-c|}|\psi_{0}|U^{\prime}dx_{2}^{\prime}
+|y(x2)y+(h)||y+f|Cx2α([x2,0])x20|UU(x~2+c)|α|Uc||ψ0|Udx2)\displaystyle\quad+\Big{|}\frac{y_{-}(x_{2})}{y_{+}(-h)}\Big{|}|y_{+}f|_{C_{x_{2}^{\prime}}^{\alpha}([x_{2},0])}\int_{x_{2}}^{0}\frac{|U-U(\tilde{x}_{2+}^{c})|^{\alpha}}{|U-c|}|\psi_{0}|U^{\prime}dx_{2}^{\prime}\Big{)}
\displaystyle\leq Cμ1α|f|Cx2α|τcR|α|τc|(χU([h,0])(|ψ0|U1))(τ)dτ,\displaystyle C\mu^{1-\alpha}|f|_{C_{x_{2}}^{\alpha}}\int_{\mathbb{R}}\frac{|\tau-c_{R}|^{\alpha}}{|\tau-c|}\big{(}\chi_{U([-h,0])}(|\psi_{0}|\circ U^{-1})\big{)}(\tau)d\tau,

where we also used

|yf|Cx2α[h,x2]C|y|Cx2α([h,x2])|f|Cx2αCμ1αeμ1(x2+h)|f|Cx2α.|y_{-}f|_{C_{x_{2}^{\prime}}^{\alpha}[-h,x_{2}]}\leq C|y_{-}|_{C_{x_{2}^{\prime}}^{\alpha}([-h,x_{2}])}|f|_{C_{x_{2}}^{\alpha}}\leq C\mu^{1-\alpha}e^{\mu^{-1}(x_{2}+h)}|f|_{C_{x_{2}}^{\alpha}}.

and a similar estimate for |y+f|Cx2α([x2,0])|y_{+}f|_{C_{x_{2}^{\prime}}^{\alpha}([x_{2},0])} due to Lemma 3.9. Since |τ|α|τ+icI|\frac{|\tau|^{\alpha}}{|\tau+ic_{I}|} is a weak-L11αL^{\frac{1}{1-\alpha}} function of τ\tau with norm uniformly bounded in cI>0c_{I}>0, the weak Young’s inequality yield

|y2|Lx2LcRr2Cμ1α|f|LcRr1Cx2α|ψ0|Lr, where 1r2=1r1+1rα<1r1+1r.\displaystyle|y_{2}|_{L_{x_{2}}^{\infty}L_{c_{R}}^{r_{2}}}\leq C\mu^{1-\alpha}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha}}|\psi_{0}|_{L^{r}},\;\text{ where }\;\tfrac{1}{r_{2}}=\tfrac{1}{r_{1}}+\tfrac{1}{r}-\alpha<\tfrac{1}{r_{1}}+\tfrac{1}{r}.

To obtain the convergence of y2y_{2} as cI0c_{I}\to 0, using the LcR,x2L_{c_{R},x_{2}}^{\infty} convergence of y±y_{\pm} and the LcRLx2q~L_{c_{R}}^{\infty}L_{x_{2}}^{\tilde{q}} and Lx2LcRq~L_{x_{2}}^{\infty}L_{c_{R}}^{\tilde{q}}, q~(1,)\forall\tilde{q}\in(1,\infty), convergence of y±y_{\pm}^{\prime} (Lemma 3.12), one may easily reduce the problem to the convergence of

Δ~=\displaystyle\tilde{\Delta}= |y0+(x2)y0+(h)hx2((y0f0U)(x2)(y0f0U)(x~2c))(1Uc1UcR)ψ0Udx2|Lx2LcRr2\displaystyle\Big{|}\frac{y_{0+}(x_{2})}{y_{0+}(-h)}\int_{-h}^{x_{2}}\Big{(}\big{(}\frac{y_{0-}f_{0}}{U^{\prime}}\big{)}(x_{2}^{\prime})-\big{(}\frac{y_{0-}f_{0}}{U^{\prime}}\big{)}(\tilde{x}_{2-}^{c})\Big{)}\big{(}\frac{1}{U-c}-\frac{1}{U-c_{R}}\big{)}\psi_{0}U^{\prime}dx_{2}^{\prime}\Big{|}_{L_{x_{2}}^{\infty}L_{c_{R}}^{r_{2}}}
\displaystyle\leq Cμ1α||f0|LCx2α||τcR|α|τc||τcR|α1|(χU([h,0])(|ψ0|U1))(τ)dτ|LcRr2\displaystyle C\mu^{1-\alpha}\Big{|}|f_{0}|_{L_{C_{x_{2}}^{\alpha}}}\int_{\mathbb{R}}\big{|}\frac{|\tau-c_{R}|^{\alpha}}{|\tau-c|}-|\tau-c_{R}|^{\alpha-1}\Big{|}\big{(}\chi_{U([-h,0])}(|\psi_{0}|\circ U^{-1})\big{)}(\tau)d\tau\Big{|}_{L_{c_{R}}^{r_{2}}}

and that of a similar term of the other integral. It is easy to see via a rescaling that, for s[1,11α)s\in[1,\frac{1}{1-\alpha}),

||τ|α|τ+icI||τ|α1|Ls=|cI|α1|γ(τcI)|Ls=|cI|1s1+α|γ|Ls, where γ(τ)=|τ|α|τ+i||τ|α1,\Big{|}\frac{|\tau|^{\alpha}}{|\tau+ic_{I}|}-|\tau|^{\alpha-1}\Big{|}_{L^{s}}=|c_{I}|^{\alpha-1}\big{|}\gamma\big{(}\frac{\tau}{c_{I}}\big{)}\big{|}_{L^{s}}=|c_{I}|^{\frac{1}{s}-1+\alpha}|\gamma|_{L^{s}},\;\;\text{ where }\;\gamma(\tau)=\frac{|\tau|^{\alpha}}{|\tau+i|}-|\tau|^{\alpha-1},

while with the weak-L11αL^{\frac{1}{1-\alpha}} norm equal to |γ|wL11α|\gamma|_{w-L^{\frac{1}{1-\alpha}}}. Hence

|||τ|α|τ+icI||τ|α1|φ|L11rα0, as cI0,\left|\Big{|}\frac{|\tau|^{\alpha}}{|\tau+ic_{I}|}-|\tau|^{\alpha-1}\Big{|}*\varphi\right|_{L^{\frac{1}{\frac{1}{r}-\alpha}}}\to 0,\quad\;\text{ as }c_{I}\to 0,

for any φLr~\varphi\in L^{\tilde{r}} with r~>r\tilde{r}>r. Through a standard density argument and using the above uniform bound on the weak-L11αL^{\frac{1}{1-\alpha}} norm of the convolution kernel, this convergence also holds for any φLr\varphi\in L^{r}. Therefore, we obtain Δ~0\tilde{\Delta}\to 0 and thus

y2(cR+icI,x2)\displaystyle y_{2}(c_{R}+ic_{I},x_{2})\to y0+(cR,x2)y0+(cR,h)hx2((y0f0U)(cR,x2)(y0f0U)(cR,x~2c))ψ0UUcRdx2\displaystyle\frac{y_{0+}(c_{R},x_{2})}{y_{0+}(c_{R},-h)}\int_{-h}^{x_{2}}\Big{(}\big{(}\frac{y_{0-}f_{0}}{U^{\prime}}\big{)}(c_{R},x_{2}^{\prime})-\big{(}\frac{y_{0-}f_{0}}{U^{\prime}}\big{)}(c_{R},\tilde{x}_{2-}^{c})\Big{)}\frac{\psi_{0}U^{\prime}}{U-c_{R}}dx_{2}^{\prime}
+y0(cR,x2)y0+(cR,h)x20((y0+f0U)(cR,x2)(y0+f)U)(cR,x~2+c))ψ0UUcRdx2.\displaystyle+\frac{y_{0-}(c_{R},x_{2})}{y_{0+}(c_{R},-h)}\int_{x_{2}}^{0}\Big{(}\big{(}\frac{y_{0+}f_{0}}{U^{\prime}}\big{)}(c_{R},x_{2}^{\prime})-\big{(}\frac{y_{0+}f_{)}}{U^{\prime}}\big{)}(c_{R},\tilde{x}_{2+}^{c})\Big{)}\frac{\psi_{0}U^{\prime}}{U-c_{R}}dx_{2}^{\prime}.

The above estimates of y1y_{1} and y2y_{2} together yield the desired estimates of ynhy_{nh} and its convergence as cI0c_{I}\to 0. The analysis on ynhy_{nh}^{\prime} also follows from the above estimates with minor modifications, mostly replacing some |y±|LcR,x2|y_{\pm}|_{L_{c_{R},x_{2}}^{\infty}} by |y±|Lx2LcRs|y_{\pm}^{\prime}|_{L_{x_{2}}^{\infty}L_{c_{R}}^{s}} or |y±|LcRLx2s|y_{\pm}^{\prime}|_{L_{c_{R}}^{\infty}L_{x_{2}}^{s}} outside the integrals, needed to control its logarithmic singularity caused by y±y_{\pm}^{\prime}. We omit the details.

Finally, as in Lemma 5.1, stronger estimates and convergence can be obtained at x2=h,0x_{2}=-h,0 due to prescribed boundary values (3.53). In fact,

ynh(0)=\displaystyle y_{nh}^{\prime}(0)= y+(0)y+(h)h0yfψ0Ucdx2\displaystyle\frac{y_{+}^{\prime}(0)}{y_{+}(-h)}\int_{-h}^{0}\frac{y_{-}f\psi_{0}}{U-c}dx_{2}^{\prime}
=\displaystyle= y+(0)y+(h)(yfU)(x~2c)h0ψ0UUcdx2+y+(0)y+(h)h0((yfU)(x2)(yfU)(x~2c))ψ0Ucdx2\displaystyle\frac{y_{+}^{\prime}(0)}{y_{+}(-h)}\big{(}\frac{y_{-}f}{U^{\prime}}\big{)}(\tilde{x}_{2-}^{c})\int_{-h}^{0}\frac{\psi_{0}U^{\prime}}{U-c}dx_{2}^{\prime}+\frac{y_{+}^{\prime}(0)}{y_{+}(-h)}\int_{-h}^{0}\Big{(}\big{(}\frac{y_{-}f}{U^{\prime}}\big{)}(x_{2}^{\prime})-\big{(}\frac{y_{-}f}{U^{\prime}}\big{)}(\tilde{x}_{2-}^{c})\Big{)}\frac{\psi_{0}}{U-c}dx_{2}^{\prime}

implies

|ynh(0)|LcRrr1r+r1\displaystyle|y_{nh}^{\prime}(0)|_{L_{c_{R}}^{\frac{rr_{1}}{r+r_{1}}}}\leq C(|f|LcRr1Lx2|U(h)U(0)ψ0τcdτ|LcRr\displaystyle C\Big{(}|f|_{L_{c_{R}}^{r_{1}}L_{x_{2}}^{\infty}}\Big{|}\int_{U(-h)}^{U(0)}\frac{\psi_{0}}{\tau-c}d\tau\Big{|}_{L_{c_{R}}^{r}}
+μ1eμh|y|LcR1ϵCx2ϵ|f|LcRr1Cx2ϵ|U(h)U(0)|ψ0||τc|1ϵdτ|LcRr1ϵr).\displaystyle+\mu^{-1}e^{-\mu h}|y_{-}|_{L_{c_{R}}^{\frac{1}{\epsilon}}C_{x_{2}}^{\epsilon}}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\epsilon}}\Big{|}\int_{U(-h)}^{U(0)}\frac{|\psi_{0}|}{|\tau-c|^{1-\epsilon}}d\tau\Big{|}_{L_{c_{R}}^{\frac{r}{1-\epsilon r}}}\Big{)}.

From the same procedure as in estimating y1y_{1} and y2y_{2} in the above, we obtain the desired estimate. Its convergence follows much as that of ynhy_{nh}. The same argument applies to ynh(c,h)y_{nh}^{\prime}(c,-h) and the proof of the lemma is complete. ∎

The following is an estimate ynh0y_{nh0} and ynh0y_{nh0}^{\prime} in LcR,x22L_{c_{R},x_{2}}^{2} and their dependence on kk.

Lemma 5.4.

In addition to (5.8) and (5.17), assume 121r+1r1\frac{1}{2}\geq\frac{1}{r}+\tfrac{1}{r_{1}}. For any ϵ(0,1)\epsilon\in(0,1), there exists C>0C>0 depending only on ϵ\epsilon, rr, r1r_{1}, F0F_{0}, ρ0\rho_{0}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any kk\in\mathbb{R} and cI0c_{I}\geq 0, it holds

|ynh|LcR,x222+μ2|ynh|LcR,x222Cμ1ϵ|f|LcRr1Cx2α2|ψ0|Lr2.|y_{nh}^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}+\mu^{-2}|y_{nh}|_{L_{c_{R},x_{2}}^{2}}^{2}\leq C\mu^{1-\epsilon}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha}}^{2}|\psi_{0}|_{L^{r}}^{2}.

where the norms are taken for cRc_{R}\in\mathcal{I} and x2[h,0]x_{2}\in[-h,0].

Proof.

As in the proof of Lemma 5.2, we first consider for cI>0c_{I}>0 and drop the subscript nh\cdot_{nh} for notation simplification. Multiplying the Rayleigh equation (5.1a) by y¯\bar{y} and integrating in both cRc_{R} and x2x_{2}, we have

h0|y|2+k2|y|2dx2dcR=h0fψ0y¯U|y|2Ucdx2dcR+yy¯dcR|x2=0\displaystyle\int_{\mathcal{I}}\int_{-h}^{0}|y^{\prime}|^{2}+k^{2}|y|^{2}dx_{2}dc_{R}=\int_{\mathcal{I}}\int_{-h}^{0}\frac{f\psi_{0}\bar{y}-U^{\prime\prime}|y|^{2}}{U-c}dx_{2}dc_{R}+\int_{\mathcal{I}}y^{\prime}\bar{y}dc_{R}\Big{|}_{x_{2}=0}
=\displaystyle= h0Uψ0Uc((fy¯U)(c,x2)(fy¯U)(c,x2c))dx2dcR+(fy¯U)(c,x2c)h0Uψ0Ucdx2dcR\displaystyle\int_{\mathcal{I}}\int_{-h}^{0}\frac{U^{\prime}\psi_{0}}{U-c}\Big{(}\big{(}\frac{f\bar{y}}{U^{\prime}}\big{)}\big{(}c,x_{2})-\big{(}\frac{f\bar{y}}{U^{\prime}}\big{)}\big{(}c,x_{2}^{c})\Big{)}dx_{2}dc_{R}+\int_{\mathcal{I}}\big{(}\frac{f\bar{y}}{U^{\prime}}\big{)}\big{(}c,x_{2}^{c})\int_{-h}^{0}\frac{U^{\prime}\psi_{0}}{U-c}dx_{2}^{\prime}dc_{R}
h0U|y|2Ucdx2dcR+yy¯dcR|x2=0I1+I2+I3+I4.\displaystyle-\int_{\mathcal{I}}\int_{-h}^{0}\frac{U^{\prime\prime}|y|^{2}}{U-c}dx_{2}dc_{R}+\int_{\mathcal{I}}y^{\prime}\bar{y}dc_{R}\Big{|}_{x_{2}=0}\triangleq I_{1}+I_{2}+I_{3}+I_{4}.

The term I4I_{4} can be estimated much as in the proof of Lemma 5.2 using Lemmas 3.9 and 5.3(2b)

|I4|Cμ2||U(0)c|2|y(0)|2dcR|Cμ2ϵ|f|LcRr1Cx2α2|ψ0|Lr2.|I_{4}|\leq C\mu^{2}\Big{|}\int_{\mathcal{I}}|U(0)-c|^{2}|y^{\prime}(0)|^{2}dc_{R}\Big{|}\leq C\mu^{2-\epsilon}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha}}^{2}|\psi_{0}|_{L^{r}}^{2}.

Choose α1\alpha_{1} and r2r_{2} such that

0<α1max{ϵ2,α,1r+1r1},1r2=1+α11r1r1(12,1],0<\alpha_{1}\leq\max\{\tfrac{\epsilon}{2},\alpha,\tfrac{1}{r}+\tfrac{1}{r_{1}}\},\;\;\tfrac{1}{r_{2}}=1+\alpha_{1}-\tfrac{1}{r}-\tfrac{1}{r_{1}}\in(\tfrac{1}{2},1],

which is possible due to our assumption on α\alpha, rr, and r1r_{1}. The integral I1I_{1} can be controlled by the Hölder continuity of ff and yy in x2x_{2}, the weak Young’s inequality, and the (5.16) type interpolation inequality as

|I1|\displaystyle|I_{1}|\leq C(χ|(fy¯)(cR,)|Cx2α1)|τcR|α11|(χU([h,0])ψ0U1)(τ)|dτdcR\displaystyle C\int_{\mathbb{R}}\int_{\mathbb{R}}\big{(}\chi_{\mathcal{I}}|(f\bar{y})(c_{R},\cdot)|_{C_{x_{2}}^{\alpha_{1}}}\big{)}|\tau-c_{R}|^{\alpha_{1}-1}|(\chi_{U([-h,0])}\psi_{0}\circ U^{-1})(\tau)|d\tau dc_{R}
\displaystyle\leq C|fy¯|LcR11+α11rCx2α1|ψ0|LrC|f|LcRr1Cx2α1|y|LcRr2Cx2α1|ψ0|Lr\displaystyle C|f\bar{y}|_{L_{c_{R}}^{\frac{1}{1+\alpha_{1}-\frac{1}{r}}}C_{x_{2}}^{\alpha_{1}}}|\psi_{0}|_{L^{r}}\leq C|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha_{1}}}|y|_{L_{c_{R}}^{r_{2}}C_{x_{2}}^{\alpha_{1}}}|\psi_{0}|_{L^{r}}
\displaystyle\leq C|f|LcRr1Cx2α1||y|Lx2212+α1|y|Lx2212α1|LcRr2|ψ0|LrC|f|LcRr1Cx2α1|y|LcR,x2212+α1|y|LcRr3Lx2212α1|ψ0|Lr\displaystyle C|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha_{1}}}\big{|}|y^{\prime}|_{L_{x_{2}}^{2}}^{\frac{1}{2}+\alpha_{1}}|y|_{L_{x_{2}}^{2}}^{\frac{1}{2}-\alpha_{1}}\big{|}_{L_{c_{R}}^{r_{2}}}|\psi_{0}|_{L^{r}}\leq C|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha_{1}}}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}+\alpha_{1}}|y|_{L_{c_{R}}^{r_{3}}L_{x_{2}}^{2}}^{\frac{1}{2}-\alpha_{1}}|\psi_{0}|_{L^{r}}

where r3<2r_{3}<2 is determined by 12+α12+12α1r3=1r2\frac{\frac{1}{2}+\alpha_{1}}{2}+\frac{\frac{1}{2}-\alpha_{1}}{r_{3}}=\frac{1}{r_{2}}. Therefore we obtain

|I1|14(|y|LcR,x222+k2|y|LcR,x222)+Ck(12α1)|f|LcRr1Cx2α12|ψ0|Lr2.|I_{1}|\leq\tfrac{1}{4}\big{(}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}+k^{2}|y|_{L_{c_{R},x_{2}}^{2}}^{2}\big{)}+Ck^{-(1-2\alpha_{1})}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha_{1}}}^{2}|\psi_{0}|_{L^{r}}^{2}.

The estimate of I2I_{2} is much as in the proof of Lemma 5.3 based on the boundedness of the convolution operator on LrL^{r}

|I2|\displaystyle|I_{2}|\leq C|ψ|Lr|(fy¯)(cR,x2c)|LcRrr1C|ψ|Lr||f|Lx2|y|Lx2212|y|Lx2212|LcRrr1\displaystyle C|\psi|_{L^{r}}|(f\bar{y})(c_{R},x_{2}^{c})|_{L_{c_{R}}^{\frac{r}{r-1}}}\leq C|\psi|_{L^{r}}\big{|}|f|_{L_{x_{2}}^{\infty}}|y|_{L_{x_{2}}^{2}}^{\frac{1}{2}}|y^{\prime}|_{L_{x_{2}}^{2}}^{\frac{1}{2}}\big{|}_{L_{c_{R}}^{\frac{r}{r-1}}}
\displaystyle\leq C|ψ|Lr|f|LcRr4Lx2|y|LcR,x2212|y|LcR,x2212,\displaystyle C|\psi|_{L^{r}}|f|_{L_{c_{R}}^{r_{4}}L_{x_{2}}^{\infty}}|y|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}},

where r4=2rr2r1r_{4}=\frac{2r}{r-2}\leq r_{1}. Hence

|I2|14(|y|LcR,x222+k2|y|LcR,x222)+Ck1|f|LcRr1Cx2α12|ψ|Lr2.|I_{2}|\leq\tfrac{1}{4}\big{(}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}+k^{2}|y|_{L_{c_{R},x_{2}}^{2}}^{2}\big{)}+Ck^{-1}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha_{1}}}^{2}|\psi|_{L^{r}}^{2}.

Finally I3I_{3} can be estimated exactly as in the proof of Lemma 5.2 (and also applying Lemma 5.3(2b)) and we have

|I3|14|y|LcR,x222+C(|y|LcR,x222+μ4ϵ|f|LcRr1Cx2α2|ψ0|Lr2).|I_{3}|\leq\tfrac{1}{4}|y^{\prime}|_{L_{c_{R},x_{2}}^{2}}^{2}+C\big{(}|y|_{L_{c_{R},x_{2}}^{2}}^{2}+\mu^{4-\epsilon}|f|_{L_{c_{R}}^{r_{1}}C_{x_{2}}^{\alpha}}^{2}|\psi_{0}|_{L^{r}}^{2}\big{)}.

Therefore, there exists k0>0k_{0}>0 such that yy and yy^{\prime} satisfy the desired estimates for |k|k0|k|\geq k_{0} and cI>0c_{I}>0. For those |k|k0|k|\leq k_{0}, the |y|LcR,x222|y|_{L_{c_{R},x_{2}}^{2}}^{2} term in the upper bound of I3I_{3} can be controlled by Lemma 5.3 directly and thus the desired estimates are also satisfied by yy and yy^{\prime}. The estimate in the limiting case of cI=0+c_{I}=0+ can be obtained through the same weak convergence argument as in the proof of Lemma 5.2. ∎

Remark 5.1.

In some sense the LcR,x22L_{c_{R},x_{2}}^{2} assumption on ψ\psi and ψ\psi^{\prime} in the Lemma 5.2 is the (unreachable) borderline case of Lemma 5.4. In fact, ψ(cR,x2)\psi(c_{R},x_{2}) can be written as ψ1\psi\cdot 1, where the former belongs to LcR2Cx212L_{c_{R}}^{2}C_{x_{2}}^{\frac{1}{2}} with r1=2r_{1}=2. As r<r<\infty and 1r+1r1=12\frac{1}{r}+\frac{1}{r_{1}}=\frac{1}{2} are assumed in (5.17) and Lemma 5.4, it does not apply in this case.

5.2. Differentiation in cc of solutions to non-homogeneous Rayleigh system

Based on the analysis of the non-homogeneous Rayleigh equation (5.1) with zero boundary conditions, in this subsection we shall mainly consider (2.11c) type non-zero boundary conditions, in particular the estimates of the derivative of solutions yB(k,c,x2)y_{B}(k,c,x_{2}) given in (5.2) with respect to cc.

Through straight forward calculations and applying Lemma 3.9, we obtain

Lemma 5.5.

Assume (5.8) and c+i[ρ0,ρ0]c\in\mathcal{I}+i[-\rho_{0},\rho_{0}]. For any 1<r1<r2<1<r_{1}<r_{2}<\infty, there exists C>0C>0 depending only on r1r_{1}, r2r_{2}, F0F_{0}, ρ0\rho_{0}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that for any |cI|ρ0|c_{I}|\leq\rho_{0}, the unique solution yB(k,c,x2)y_{B}(k,c,x_{2}) to (5.1) satisfies

|yB|LcR,x22C(|ynh|LcR,x22+μ52|ζ+|LcR2+μ12|ζ|LcR2),|y_{B}|_{L_{c_{R},x_{2}}^{2}}\leq C\big{(}|y_{nh}|_{L_{c_{R},x_{2}}^{2}}+\mu^{\frac{5}{2}}|\zeta_{+}|_{L_{c_{R}}^{2}}+\mu^{\frac{1}{2}}|\zeta_{-}|_{L_{c_{R}}^{2}}\big{)},
|yB|LcR,x22C(|ynh|LcR,x22+μ32|ζ+|LcR2+μ12|ζ|LcR2),|y_{B}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\leq C\big{(}|y_{nh}^{\prime}|_{L_{c_{R},x_{2}}^{2}}+\mu^{\frac{3}{2}}|\zeta_{+}|_{L_{c_{R}}^{2}}+\mu^{-\frac{1}{2}}|\zeta_{-}|_{L_{c_{R}}^{2}}\big{)},
|yB(h)|LcRr1C(|ynh(h)|LcRr1+μ1|ζ|LcRr1+|ζ|LcRr2+μeμ1h|ζ+|LcRr1),|y_{B}^{\prime}(-h)|_{L_{c_{R}}^{r_{1}}}\leq C\big{(}|y_{nh}^{\prime}(-h)|_{L_{c_{R}}^{r_{1}}}+\mu^{-1}|\zeta_{-}|_{L_{c_{R}}^{r_{1}}}+|\zeta_{-}|_{L_{c_{R}}^{r_{2}}}+\mu e^{-\mu^{-1}h}|\zeta_{+}|_{L_{c_{R}}^{r_{1}}}\big{)},
|yB(0)|LcRr1C(|ynh(0)|LcRr1+μ|ζ+|LcRr1+μ2|ζ+|LcRr2+μ1eμ1h|ζ|LcRr1)),|y_{B}^{\prime}(0)|_{L_{c_{R}}^{r_{1}}}\leq C\big{(}|y_{nh}^{\prime}(0)|_{L_{c_{R}}^{r_{1}}}+\mu|\zeta_{+}|_{L_{c_{R}}^{r_{1}}}+\mu^{2}|\zeta_{+}|_{L_{c_{R}}^{r_{2}}}+\mu^{-1}e^{-\mu^{-1}h}|\zeta_{-}|_{L_{c_{R}}^{r_{1}}})\big{)},

where the norm is taken on cRc_{R}\in\mathcal{I} and x2[h,0]x_{2}\in[-h,0].

We shall also consider the limit

(5.19) yB0=yB|cI=0+=limcI0+yB=b0y0+b0+y0++ynh0,y_{B0}=y_{B}|_{c_{I}=0+}=\lim_{c_{I}\to 0+}y_{B}=b_{0-}y_{0-}+b_{0+}y_{0+}+y_{nh0},

which exists for appropriate ψ(c,x2)\psi(c,x_{2}) and satisfies the same estimates as yBy_{B} (see Subsection 5.1).

In the rest of the subsection, we shall focus on the special case motivated by (2.11):

(5.20) ψ=ψ0(x2),ζ(c)=ξ,ζ+(c)=ξ1+(U(0)c)ξ2,\psi=\psi_{0}(x_{2}),\quad\zeta_{-}(c)=\xi_{-},\quad\zeta_{+}(c)=\xi_{1}+(U(0)-c)\xi_{2},

where ψ0\psi_{0}, ξ\xi_{-}, ξ1\xi_{1}, and ξ2\xi_{2} are all independent of cc. Our goal is to obtain the estimates of the derivatives of the solution yB(k,c,x2)y_{B}(k,c,x_{2}) to (5.1) in cRc_{R}.

Proposition 5.6.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, (5.8), and (5.20). For any ϵ(0,1)\epsilon\in(0,1), r(1,)r\in(1,\infty), there exists C>0C>0 depending on ϵ\epsilon, rr, F0F_{0}, ρ0\rho_{0}, |U|Cl01|U^{\prime}|_{C^{l_{0}-1}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} such that the solution yB(k,c,x2)y_{B}(k,c,x_{2}) to (5.1) satisfies that for any |cI|ρ0|c_{I}|\leq\rho_{0} and kk\in\mathbb{R},

|yB|LcR,x22+μ|yB|LcR,x22+μ32|yB(0)|LcR2+μ12|yB(0)|LcR2Cμ52(μ1ϵ|ψ0|L2+|ξ1|+|ξ2|+μ2|ξ|);|y_{B}|_{L_{c_{R},x_{2}}^{2}}+\mu|y_{B}^{\prime}|_{L_{c_{R},x_{2}}^{2}}+\mu^{\frac{3}{2}}|y_{B}^{\prime}(0)|_{L_{c_{R}}^{2}}+\mu^{\frac{1}{2}}|y_{B}(0)|_{L_{c_{R}}^{2}}\leq C\mu^{\frac{5}{2}}\big{(}\mu^{-1-\epsilon}|\psi_{0}|_{L^{2}}+|\xi_{1}|+|\xi_{2}|+\mu^{-2}|\xi_{-}|\big{)};

if l04l_{0}\geq 4, then

|cRyB|LcR,x22+μ|cRyB+1U(x2c)yB|LcR,x22+μ32|(cRyB+1U(x2c)yB)(0)|LcR2+μ12|cRyB(0)|LcR2\displaystyle|\partial_{c_{R}}y_{B}|_{L_{c_{R},x_{2}}^{2}}+\mu|\partial_{c_{R}}y_{B}^{\prime}+\tfrac{1}{U^{\prime}(x_{2}^{c})}y_{B}^{\prime\prime}|_{L_{c_{R},x_{2}}^{2}}+\mu^{\frac{3}{2}}|(\partial_{c_{R}}y_{B}^{\prime}+\tfrac{1}{U^{\prime}(x_{2}^{c})}y_{B}^{\prime\prime})(0)|_{L_{c_{R}}^{2}}+\mu^{\frac{1}{2}}|\partial_{c_{R}}y_{B}(0)|_{L_{c_{R}}^{2}}
\displaystyle\leq Cμ32(μ1ϵ|ψ0|L2+μϵ|ψ0|L2+|ξ1|+|ξ2|+μ2|ξ|);\displaystyle C\mu^{\frac{3}{2}}\big{(}\mu^{-1-\epsilon}|\psi_{0}|_{L^{2}}+\mu^{-\epsilon}|\psi_{0}^{\prime}|_{L^{2}}+|\xi_{1}|+|\xi_{2}|+\mu^{-2}|\xi_{-}|\big{)};

and, if l05l_{0}\geq 5, then

|y~B|LcR,x22Cμ12(μ1ϵ|ψ0|L2+μϵ|ψ0|L2+μ1ϵ|ψ0|L2+|ξ1|+|ξ2|+μ2|ξ|),|\tilde{y}_{B}|_{L_{c_{R},x_{2}}^{2}}\leq C\mu^{\frac{1}{2}}\big{(}\mu^{-1-\epsilon}|\psi_{0}|_{L^{2}}+\mu^{-\epsilon}|\psi_{0}^{\prime}|_{L^{2}}+\mu^{1-\epsilon}|\psi_{0}^{\prime\prime}|_{L^{2}}+|\xi_{1}|+|\xi_{2}|+\mu^{-2}|\xi_{-}|\big{)},

where

y~B=cR2yB+1U(x2c)2(yB+g+σk2𝐅(k,c)(yB(h)y+yB(0)y)),\tilde{y}_{B}=\partial_{c_{R}}^{2}y_{B}+\frac{1}{U^{\prime}(x_{2}^{c})^{2}}\Big{(}-y_{B}^{\prime\prime}+\frac{g+\sigma k^{2}}{\mathbf{F}(k,c)}\big{(}y_{B}^{\prime\prime}(-h)y_{+}-y_{B}^{\prime\prime}(0)y_{-}\big{)}\Big{)},

and all the norms are taken on (cR,x2)×[h,0](c_{R},x_{2})\in\mathcal{I}\times[-h,0]. Moreover, as cI0+c_{I}\to 0+, the following hold.

  1. (1)

    Assume ψ0L2\psi_{0}\in L^{2} and UC3U\in C^{3}, then for any r[1,2)r\in[1,2), yByB0y_{B}\to y_{B0} in Lx2LcR2L_{x_{2}}^{\infty}L_{c_{R}}^{2}, yByB0y_{B}^{\prime}\to y_{B0}^{\prime} in Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r}, and yB(0)yB0(0)y_{B}^{\prime}(0)\to y_{B0}^{\prime}(0) in LcR2L_{c_{R}}^{2}.

  2. (2)

    Assume ψ0H1\psi_{0}\in H^{1} and UC4U\in C^{4}, then for any r[1,2)r\in[1,2) and q[1,)q\in[1,\infty), cRyBcRyB0\partial_{c_{R}}y_{B}\to\partial_{c_{R}}y_{B0} in Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r}, cyB+1U(x2c)yBcyB0+1U(x2c)yB0\partial_{c}y_{B}^{\prime}+\tfrac{1}{U^{\prime}(x_{2}^{c})}y_{B}^{\prime\prime}\to\partial_{c}y_{B0}^{\prime}+\tfrac{1}{U^{\prime}(x_{2}^{c})}y_{B0}^{\prime\prime} in Lx2qLcRrL_{x_{2}}^{q}L_{c_{R}}^{r}, and (cyB+1U(x2c)yB)(0)(cyB0+1U(x2c)yB0)(0)(\partial_{c}y_{B}^{\prime}+\tfrac{1}{U^{\prime}(x_{2}^{c})}y_{B}^{\prime\prime})(0)\to(\partial_{c}y_{B0}^{\prime}+\tfrac{1}{U^{\prime}(x_{2}^{c})}y_{B0}^{\prime\prime})(0) in LcRrL_{c_{R}}^{r}.

  3. (3)

    Assume ψ0H2\psi_{0}\in H^{2} and UC5U\in C^{5}, then for any r[1,2)r\in[1,2), y~B\tilde{y}_{B} also converges in Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r} to its limit y~B0\tilde{y}_{B0}.

Since yBy_{B} is holomorphic in cU([h,0])c\notin U([-h,0]), cyB=cRyB\partial_{c}y_{B}=\partial_{c_{R}}y_{B}. From the Rayleigh equation, singularity at the level of delta mass appears in yBy_{B}^{\prime\prime} along U(x2)=cRU(x_{2})=c_{R}, x2[h,0]x_{2}\in[-h,0], as cI0+c_{I}\to 0+. Therefore c2yB\partial_{c}^{2}y_{B} and cyB\partial_{c}y_{B}^{\prime} also display such singularities which are singled out in the above estimates. The yBy_{B}^{\prime\prime} involved in the singular terms will be substituted by using the Rayleigh equation (5.1a) whenever necessary.

Proof.

The LcR,x22L_{c_{R},x_{2}}^{2} estimates on yBy_{B} and yBy_{B}^{\prime}, as well as the LcRrL_{c_{R}}^{r} estimate of yB(0)y_{B}^{\prime}(0) with r(1,)r\in(1,\infty), follow readily from (5.8), (5.20), Lemmas 3.9, 5.4, 5.3 (with r=2r=2, r1=r_{1}=\infty, and f0=1f_{0}=1), and 5.5. The estimate of yB(0)y_{B}(0) is simply obtained from those of yBy_{B} and yBy_{B}^{\prime}. Moreover, for the rest of the proof of the proposition we shall also need the following inequality for r(1,)r\in(1,\infty) which is also derived form Lemma 5.3 and Lemma 5.5 and uniform in cI[0,ρ0]c_{I}\in[0,\rho_{0}]

(5.21) |yB(0)|LcRrC(μϵ|ψ0|Lr+μ(|ξ1|+|ξ2|)+μ1eμ1h|ξ|),|yB(h)|LcRrC(μϵ|ψ0|Lr+μeμ1h(|ξ1|+|ξ2|)+μ1|ξ|).\begin{split}&|y_{B}^{\prime}(0)|_{L_{c_{R}}^{r}}\leq C(\mu^{-\epsilon}|\psi_{0}|_{L^{r}}+\mu(|\xi_{1}|+|\xi_{2}|)+\mu^{-1}e^{-\mu^{-1}h}|\xi_{-}|),\\ &|y_{B}^{\prime}(-h)|_{L_{c_{R}}^{r}}\leq C(\mu^{-\epsilon}|\psi_{0}|_{L^{r}}+\mu e^{-\mu^{-1}h}(|\xi_{1}|+|\xi_{2}|)+\mu^{-1}|\xi_{-}|).\end{split}

The convergence of yBy_{B}, yBy_{B}^{\prime}, and yB(0)y_{B}^{\prime}(0) follow directly from the continuity of 𝐅\mathbf{F} (Lemma 4.1) and the convergence of y±y_{\pm} and y±y_{\pm}^{\prime} (Lemma 3.12) and ynhy_{nh} (Lemma 5.3). Moreover, we also have the convergence of yB(h)y_{B}^{\prime}(-h) in LcR2L_{c_{R}}^{2}.

In the following differentiations in cRc_{R} are all carried out for cI>0c_{I}>0. The convergence analysis based on the convergence results of y±y_{\pm} and those of ynhy_{nh} in Subsection 5.1 ensure that the estimates hold also for cI=0+c_{I}=0+. Directly differentiating the Rayleigh equation (5.1a) in cRc_{R} directly would cause worse singularity in the equation. Instead we first consider

(5.22) Dc=U(x2c)cR+x2,cR=U(x2c)1(Dcx2),[Dc,x2]=0,D_{c}=U^{\prime}(x_{2}^{c})\partial_{c_{R}}+\partial_{x_{2}},\qquad\partial_{c_{R}}=U^{\prime}(x_{2}^{c})^{-1}(D_{c}-\partial_{x_{2}}),\qquad[D_{c},\partial_{x_{2}}]=0,

where x2cx_{2}^{c} is defined by U(x2c)=cRU(x_{2}^{c})=c_{R} as in (3.6). It satisfies

(5.23) Dc(1U(x2)c)=U(x2)U(x2c)(U(x2)c)2,Dc2(1U(x2)c)=2(U(x2)U(x2c))2(U(x2)c)3U(x2)U(x2c)(U(x2)c)2,D_{c}\big{(}\tfrac{1}{U(x_{2})-c}\big{)}=-\tfrac{U^{\prime}(x_{2})-U^{\prime}(x_{2}^{c})}{(U(x_{2})-c)^{2}},\quad D_{c}^{2}\big{(}\tfrac{1}{U(x_{2})-c}\big{)}=\tfrac{2(U^{\prime}(x_{2})-U^{\prime}(x_{2}^{c}))^{2}}{(U(x_{2})-c)^{3}}-\tfrac{U^{\prime\prime}(x_{2})-U^{\prime\prime}(x_{2}^{c})}{(U(x_{2})-c)^{2}},

where the singularity remains at the same level.

\bullet Estimating cRyB\partial_{c_{R}}y_{B}. Applying DcD_{c} to (5.1a) and simplifying, we obtain

(5.24a) (DcyB)+(k2+UUc)DcyB=ψ0(x2)+f1(c,x2)ψ0(x2)+ψ1(c,x2)Uc;\begin{split}-(D_{c}y_{B})^{\prime\prime}+&\big{(}k^{2}+\frac{U^{\prime\prime}}{U-c}\big{)}D_{c}y_{B}=\frac{\psi_{0}^{\prime}(x_{2})+f_{1}(c,x_{2})\psi_{0}(x_{2})+\psi_{1}(c,x_{2})}{U-c};\end{split}
where
ψ1=(U(UU(x2c))UcU)yB,f1=(Uc)Dc(1Uc)=UU(x2c)Uc,\psi_{1}=\big{(}\frac{U^{\prime\prime}(U^{\prime}-U^{\prime}(x_{2}^{c}))}{U-c}-U^{\prime\prime\prime}\big{)}y_{B},\quad f_{1}=(U-c)D_{c}\big{(}\frac{1}{U-c}\big{)}=-\frac{U^{\prime}-U^{\prime}(x_{2}^{c})}{U-c},
and boundary conditions
(5.24b) DcyB(h)=ζ1yB(h);D_{c}y_{B}(-h)=\zeta_{1-}\triangleq y_{B}^{\prime}(-h);
(5.24c) (U(0)c)2(DcyB)(0)(U(0)(U(0)c)+g+σk2)DcyB(0)=ζ1+(c)\big{(}U(0)-c\big{)}^{2}(D_{c}y_{B})^{\prime}(0)-\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}D_{c}y_{B}(0)=\zeta_{1+}(c)
where
ζ1+=\displaystyle\zeta_{1+}= ξ2U(x2c)(U(0)c)ψ0(0)+((2U(x2c)U(0))(U(0)c)gσk2)yB(0)\displaystyle-\xi_{2}U^{\prime}(x_{2}^{c})-(U(0)-c)\psi_{0}(0)+\big{(}(2U^{\prime}(x_{2}^{c})-U^{\prime}(0))(U(0)-c)-g-\sigma k^{2}\big{)}y_{B}^{\prime}(0)
+(k2(U(0)c)2+U(0)(U(0)c)U(x2c)U(0))yB(0).\displaystyle+\big{(}k^{2}(U(0)-c)^{2}+U^{\prime\prime}(0)(U(0)-c)-U^{\prime}(x_{2}^{c})U^{\prime}(0)\big{)}y_{B}(0).

Let y~1(c,x2)\tilde{y}_{1}(c,x_{2}) and y~2(c,x2)\tilde{y}_{2}(c,x_{2}) be the solution to the non-homogeneous Rayleigh equation (5.1a), but with zero boundary values in (5.1b), with ψ(c,x2)\psi(c,x_{2}) replaced by ψ1\psi_{1} and ψ0+f1ψ0\psi_{0}^{\prime}+f_{1}\psi_{0}, respectively. Both are given by the formula (5.3). Using the estimates of yBy_{B} derived in the above and apply Lemmas 5.2, we have

|y~1|LcR,x22+μ|y~1|LcR,x22Cμ(|yB|LcR,x22+μ1ϵ4|yB|LcR,x22)Cμ52ϵ(μ|ξ1|+μ|ξ2|+|ψ0|L2+μ1|ξ|).|\tilde{y}_{1}|_{L_{c_{R},x_{2}}^{2}}+\mu|\tilde{y}_{1}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\leq C\mu\big{(}|y_{B}|_{L_{c_{R},x_{2}}^{2}}+\mu^{1-\frac{\epsilon}{4}}|y_{B}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\big{)}\leq C\mu^{\frac{5}{2}-\epsilon}\big{(}\mu|\xi_{1}|+\mu|\xi_{2}|+|\psi_{0}|_{L^{2}}+\mu^{-1}|\xi_{-}|\big{)}.

Moreover, from Lemma 5.1(2b) and (5.1b), (5.7), and (5.20), one can compute

|y~1(c,0)|\displaystyle|\tilde{y}_{1}^{\prime}(c,0)|\leq Cμ12(1+ϵ)(|yB|Lx22+μ|yB|Lx22)+Cμ2(1+|log|U(0)c||)(|ζ+|+|U(0)c|2|yB(c,0)|),\displaystyle C\mu^{-\frac{1}{2}(1+\epsilon)}(|y_{B}|_{L_{x_{2}}^{2}}+\mu|y_{B}^{\prime}|_{L_{x_{2}}^{2}})+C\mu^{2}\big{(}1+\big{|}\log|U(0)-c|\big{|}\big{)}(|\zeta_{+}|+|U(0)-c|^{2}|y_{B}^{\prime}(c,0)|),

where yB(0)y_{B}(0) was substituted by using (5.1b). It along with the above estimates on yBy_{B} implies

|y~1(0)|LcR2\displaystyle|\tilde{y}_{1}^{\prime}(0)|_{L_{c_{R}}^{2}}\leq Cμ12(1+ϵ)(|yB|LcR,x22+μ|yB|LcR,x22)+Cμ2(|ξ1|+|ξ2|+|yB(0)|LcR2)\displaystyle C\mu^{-\frac{1}{2}(1+\epsilon)}(|y_{B}|_{L_{c_{R},x_{2}}^{2}}+\mu|y_{B}^{\prime}|_{L_{c_{R},x_{2}}^{2}})+C\mu^{2}(|\xi_{1}|+|\xi_{2}|+|y_{B}^{\prime}(0)|_{L_{c_{R}}^{2}})
\displaystyle\leq Cμ1ϵ(μ|ξ1|+μ|ξ2|+|ψ0|L2+μ1|ξ|).\displaystyle C\mu^{1-\epsilon}\big{(}\mu|\xi_{1}|+\mu|\xi_{2}|+|\psi_{0}|_{L^{2}}+\mu^{-1}|\xi_{-}|\big{)}.

The estimate at x=hx=-h based on Lemma 5.1(2b) is similar

|y~1(c,h)|Cμ12(1+ϵ)(|yB|Lx22+μ|yB|Lx22)+C(1+|log|U(h)c||)|ξ|,|\tilde{y}_{1}^{\prime}(c,-h)|\leq C\mu^{-\frac{1}{2}(1+\epsilon)}\big{(}|y_{B}|_{L_{x_{2}}^{2}}+\mu|y_{B}^{\prime}|_{L_{x_{2}}^{2}}\big{)}+C\big{(}1+\big{|}\log|U(-h)-c|\big{|}\big{)}|\xi_{-}|,

which yields

|y~1(h)|LcR2\displaystyle|\tilde{y}_{1}^{\prime}(-h)|_{L_{c_{R}}^{2}}\leq Cμ12(1+ϵ)(|yB|LcR,x22+μ|yB|LcR,x22)+C|ξ|\displaystyle C\mu^{-\frac{1}{2}(1+\epsilon)}(|y_{B}|_{L_{c_{R},x_{2}}^{2}}+\mu|y_{B}^{\prime}|_{L_{c_{R},x_{2}}^{2}})+C|\xi_{-}|
\displaystyle\leq Cμ1ϵ(μ|ξ1|+μ|ξ2|+|ψ0|L2+μ1|ξ|).\displaystyle C\mu^{1-\epsilon}\big{(}\mu|\xi_{1}|+\mu|\xi_{2}|+|\psi_{0}|_{L^{2}}+\mu^{-1}|\xi_{-}|\big{)}.

From the convergence of yBy_{B} and Lemma 5.1, as cI0+c_{I}\to 0+, we have the convergence of y~1\tilde{y}_{1} in LcRrLx2L_{c_{R}}^{r}L_{x_{2}}^{\infty}, y~1\tilde{y}_{1}^{\prime} in LcRrLx2qL_{c_{R}}^{r}L_{x_{2}}^{q}, and y~1(0)\tilde{y}_{1}^{\prime}(0) in LcRrL_{c_{R}}^{r}, for any r[1,2)r\in[1,2) and q[1,)q\in[1,\infty).

Due to the smoothness of f1f_{1}, we apply Lemmas 5.4 and 5.3 instead to estimate y~2\tilde{y}_{2}

|y~2|LcR,x22+μ|y~2|LcR,x22Cμ32ϵ|ψ0|H1,|y~2(h)|LcR2+|y~2(0)|LcR2Cμϵ|ψ0|H1.|\tilde{y}_{2}|_{L_{c_{R},x_{2}}^{2}}+\mu|\tilde{y}_{2}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\leq C\mu^{\frac{3}{2}-\epsilon}|\psi_{0}|_{H^{1}},\quad|\tilde{y}_{2}^{\prime}(-h)|_{L_{c_{R}}^{2}}+|\tilde{y}_{2}^{\prime}(0)|_{L_{c_{R}}^{2}}\leq C\mu^{-\epsilon}|\psi_{0}|_{H^{1}}.

Again from Lemma 5.3, as cI0+c_{I}\to 0+, we have the convergence of y~2\tilde{y}_{2} in Lx2LcR2L_{x_{2}}^{\infty}L_{c_{R}}^{2}, y~2\tilde{y}_{2}^{\prime} in Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r}, for any r[1,2)r\in[1,2), and y~2(0)\tilde{y}_{2}^{\prime}(0) in LcR2L_{c_{R}}^{2}.

Finally, from (5.21) and (5.7), we have, for any r(1,)r\in(1,\infty),

|ζ1|LcRrC(μϵ|ψ0|Lr+μeμ1h(|ξ1|+|ξ2|)+μ1|ξ|),|\zeta_{1-}|_{L_{c_{R}}^{r}}\leq C\big{(}\mu^{-\epsilon}|\psi_{0}|_{L^{r}}+\mu e^{-\mu^{-1}h}(|\xi_{1}|+|\xi_{2}|)+\mu^{-1}|\xi_{-}|\big{)},
|ζ1+|LcRrCμ1(|ξ1|+|ξ2|+μ|ψ0(0)|+μ1ϵ|ψ0|Lr+μ2eμ1h|ξ|),|\zeta_{1+}|_{L_{c_{R}}^{r}}\leq C\mu^{-1}\big{(}|\xi_{1}|+|\xi_{2}|+\mu|\psi_{0}(0)|+\mu^{-1-\epsilon}|\psi_{0}|_{L^{r}}+\mu^{-2}e^{-\mu^{-1}h}|\xi_{-}|\big{)},

where again we substituted yB(0)y_{B}(0) by (5.1b) and (5.7). Moreover, from the convergence of yBy_{B} and yBy_{B}^{\prime}, we have the convergence of ζ1±\zeta_{1\pm} in LcR2L_{c_{R}}^{2}.

As y~1+y~2\tilde{y}_{1}+\tilde{y}_{2} plays the role of "ynhy_{nh}" in the representation of DcyBD_{c}y_{B} as given in Lemma 5.5, the above estimates imply

(5.25) |DcyB|LcR,x22+μ|DcyB|LcR,x22C(μ32|ξ1|+μ32|ξ2|+μ12|ξ|+μ12ϵ|ψ0|L2+μ32ϵ|ψ0|L2),|D_{c}y_{B}|_{L_{c_{R},x_{2}}^{2}}+\mu|D_{c}y_{B}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\leq C\big{(}\mu^{\frac{3}{2}}|\xi_{1}|+\mu^{\frac{3}{2}}|\xi_{2}|+\mu^{-\frac{1}{2}}|\xi_{-}|+\mu^{\frac{1}{2}-\epsilon}|\psi_{0}|_{L^{2}}+\mu^{\frac{3}{2}-\epsilon}|\psi_{0}^{\prime}|_{L^{2}}\big{)},

where the ψ0(0)\psi_{0}(0) term was bounded by the other norms of ψ0\psi_{0} via interpolation. The desired LcR,x22L_{c_{R},x_{2}}^{2} estimates on cRyB\partial_{c_{R}}y_{B} and cRyB\partial_{c_{R}}y_{B}^{\prime} follow from that of yBy_{B}^{\prime}, (5.22), and the above inequality. We also obtain the LcR2L_{c_{R}}^{2} estimate of DcyB(0)D_{c}y_{B}(0) from (5.25) which in turn yields the LcR2L_{c_{R}}^{2} bound on cRyB(0)\partial_{c_{R}}y_{B}(0). The convergence of cRyB\partial_{c_{R}}y_{B} is a direct consequence of those of y~1\tilde{y}_{1}, y~2\tilde{y}_{2}, ζ1±\zeta_{1\pm}, and the representation formula given in Lemma 5.5. Moreover, we also have the convergence of DcyB|x2=0,hD_{c}y_{B}^{\prime}|_{x_{2}=0,-h} in LcRrL_{c_{R}}^{r} for any r[1,2)r\in[1,2).

To complete the estimates on cRyB\partial_{c_{R}}y_{B} and also for the next step, we also need the following inequalities which are also derived from the above estimates and Lemma 5.5

|DcyB(h)|LcR2\displaystyle|D_{c}y_{B}^{\prime}(-h)|_{L_{c_{R}}^{2}}\leq C(μ1|ψ0|L2+|ψ0|Lr+|ψ0|L2+μ2ϵ(|ξ1|+|ξ2|))+Cμ2|ξ|\displaystyle C\big{(}\mu^{-1}|\psi_{0}|_{L^{2}}+|\psi_{0}|_{L^{r}}+|\psi_{0}^{\prime}|_{L^{2}}+\mu^{2-\epsilon}(|\xi_{1}|+|\xi_{2}|)\big{)}+C\mu^{-2}|\xi_{-}|
\displaystyle\leq Cμϵ(μ2(|ξ1|+|ξ2|)+μ1|ψ0|L2+|ψ0|L2)+Cμ2|ξ|,\displaystyle C\mu^{-\epsilon}\big{(}\mu^{2}(|\xi_{1}|+|\xi_{2}|)+\mu^{-1}|\psi_{0}|_{L^{2}}+|\psi_{0}^{\prime}|_{L^{2}}\big{)}+C\mu^{-2}|\xi_{-}|,
|DcyB(0)|LcR2\displaystyle|D_{c}y_{B}^{\prime}(0)|_{L_{c_{R}}^{2}}\leq C(μ1|ψ0|L2+|ψ0|Lr+μ|ψ0(0)|+|ψ0|L2+|ξ1|+|ξ2|)\displaystyle C\big{(}\mu^{-1}|\psi_{0}|_{L^{2}}+|\psi_{0}|_{L^{r}}+\mu|\psi_{0}(0)|+|\psi_{0}^{\prime}|_{L^{2}}+|\xi_{1}|+|\xi_{2}|\big{)}
\displaystyle\leq C(|ξ1|+|ξ2|+μ1ϵ|ψ0|L2+μϵ|ψ0|L2+μϵ|ξ|),\displaystyle C\big{(}|\xi_{1}|+|\xi_{2}|+\mu^{-1-\epsilon}|\psi_{0}|_{L^{2}}+\mu^{-\epsilon}|\psi_{0}^{\prime}|_{L^{2}}+\mu^{-\epsilon}|\xi_{-}|\big{)},

where the terms involving |ψ0(0)||\psi_{0}(0)| and |ψ0|Lr|\psi_{0}|_{L^{r}}, r>2r>2, are bounded by other norms of ψ0\psi_{0}.

\bullet Estimating cR2yB\partial_{c_{R}}^{2}y_{B}. In order to analyze cR2yB\partial_{c_{R}}^{2}y_{B}, we still first apply DcD_{c} to (5.24). Due to the commutativity (5.22) between DcD_{c} and x2\partial_{x_{2}}, the Rayleigh equation (5.1a) and (5.20) imply

(Dc2yB)+(k2+UUc)Dc2yB=\displaystyle-(D_{c}^{2}y_{B})^{\prime\prime}+\big{(}k^{2}+\frac{U^{\prime\prime}}{U-c}\big{)}D_{c}^{2}y_{B}= ψ0U(4)yB2UDcyBUc+2Dc(1Uc)(ψ0Dc(UyB))\displaystyle\frac{\psi_{0}^{\prime\prime}-U^{(4)}y_{B}-2U^{\prime\prime\prime}D_{c}y_{B}}{U-c}+2D_{c}\big{(}\frac{1}{U-c}\big{)}\big{(}\psi_{0}^{\prime}-D_{c}(U^{\prime\prime}y_{B})\big{)}
+Dc2(1Uc)(ψ0UyB).\displaystyle+D_{c}^{2}\big{(}\frac{1}{U-c}\big{)}(\psi_{0}-U^{\prime\prime}y_{B}).
We can write
(5.26a) (Dc2yB)+(k2+UUc)Dc2yB=ψ0(x2)+f2(c,x2)ψ0(x2)+2f1(c,x2)ψ0(x2)+ψ2(c,x2)Uc,-(D_{c}^{2}y_{B})^{\prime\prime}+\big{(}k^{2}+\frac{U^{\prime\prime}}{U-c}\big{)}D_{c}^{2}y_{B}=\frac{\psi_{0}^{\prime\prime}(x_{2})+f_{2}(c,x_{2})\psi_{0}(x_{2})+2f_{1}(c,x_{2})\psi_{0}^{\prime}(x_{2})+\psi_{2}(c,x_{2})}{U-c},
where f1f_{1} was defined in (5.24) and
f2=(Uc)Dc2(1Uc),ψ2=(2U+Uf1)DcyB(U(4)+Uf1+Uf2)yB.f_{2}=(U-c)D_{c}^{2}\big{(}\frac{1}{U-c}\big{)},\quad\psi_{2}=-(2U^{\prime\prime\prime}+U^{\prime\prime}f_{1})D_{c}y_{B}-(U^{(4)}+U^{\prime\prime\prime}f_{1}+U^{\prime\prime}f_{2})y_{B}.
From (5.23) and the assumption UC4U\in C^{4}, it holds f2f_{2} and f3f_{3} are C1C^{1} in x2x_{2} and cRc_{R} with bounds uniform in |cI|ρ0|c_{I}|\leq\rho_{0}. At x2=hx_{2}=-h, one can compute using (3.79),
(Dc2yB)(h)=(U(x2c)2cR2yB+U(x2c)cRyB+U(x2c)cRyB+(DcyB))|x2=h.(D_{c}^{2}y_{B})(-h)=\big{(}U^{\prime}(x_{2}^{c})^{2}\partial_{c_{R}}^{2}y_{B}+U^{\prime\prime}(x_{2}^{c})\partial_{c_{R}}y_{B}+U^{\prime}(x_{2}^{c})\partial_{c_{R}}y_{B}^{\prime}+(D_{c}y_{B})^{\prime}\big{)}\big{|}_{x_{2}=-h}.
From (5.20) and (5.1a), we can write
(5.26b) (Dc2yB)(c,h)=ζ2(c)(2(DcyB)yB)(h)=2(DcyB)(h)+ψ0(h)U(h)c.(D_{c}^{2}y_{B})(c,-h)=\zeta_{2-}(c)\triangleq\big{(}2(D_{c}y_{B})^{\prime}-y_{B}^{\prime\prime}\big{)}(-h)=2(D_{c}y_{B})^{\prime}(-h)+\frac{\psi_{0}(-h)}{U(-h)-c}.
At x2=0x_{2}=0, we write
(5.26c) (U(0)c)2(Dc2yB)(0)(U(0)(U(0)c)+g+σk2)Dc2yB(0)=ζ2+(c).\big{(}U(0)-c\big{)}^{2}(D_{c}^{2}y_{B})^{\prime}(0)-\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}D_{c}^{2}y_{B}(0)=\zeta_{2+}(c).

One may compute ζ2+\zeta_{2+} using (5.22) and (5.24c)

ζ2+=\displaystyle\zeta_{2+}= U(x2c)(cRζ1++2(U(0)c)(DcyB)(0)U(0)DcyB(0))\displaystyle U^{\prime}(x_{2}^{c})\big{(}\partial_{c_{R}}\zeta_{1+}+2(U(0)-c)(D_{c}y_{B})^{\prime}(0)-U^{\prime}(0)D_{c}y_{B}(0)\big{)}
+(U(0)c)2(DcyB)(0)(U(0)(U(0)c)+g+σk2)(DcyB)(0).\displaystyle+\big{(}U(0)-c\big{)}^{2}(D_{c}y_{B})^{\prime\prime}(0)-\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}(D_{c}y_{B})^{\prime}(0).

On the one hand, the U(x2c)cRζ1+U^{\prime}(x_{2}^{c})\partial_{c_{R}}\zeta_{1+} turns out to involve some of the most singular terms in ζ2+\zeta_{2+},

U(x2c)cRζ1+=\displaystyle U^{\prime}(x_{2}^{c})\partial_{c_{R}}\zeta_{1+}= ξ2U(x2c)+U(x2c)ψ0(0)+(2U(x2c)(U(0)c)U(x2c)(2U(x2c)U(0)))yB(0)\displaystyle-\xi_{2}U^{\prime\prime}(x_{2}^{c})+U^{\prime}(x_{2}^{c})\psi_{0}(0)+\big{(}2U^{\prime\prime}(x_{2}^{c})(U(0)-c)-U^{\prime}(x_{2}^{c})(2U^{\prime}(x_{2}^{c})-U^{\prime}(0))\big{)}y_{B}^{\prime}(0)
+U(x2c)((2U(x2c)U(0))(U(0)c)gσk2)cRyB(0)\displaystyle+U^{\prime}(x_{2}^{c})\big{(}(2U^{\prime}(x_{2}^{c})-U^{\prime}(0))(U(0)-c)-g-\sigma k^{2}\big{)}\partial_{c_{R}}y_{B}^{\prime}(0)
+U(x2c)(k2(U(0)c)2+U(0)(U(0)c)U(x2c)U(0))cRyB(0)\displaystyle+U^{\prime}(x_{2}^{c})\big{(}k^{2}(U(0)-c)^{2}+U^{\prime\prime}(0)(U(0)-c)-U^{\prime}(x_{2}^{c})U^{\prime}(0)\big{)}\partial_{c_{R}}y_{B}(0)
+(U(x2c)(2k2(cU(0))U(0))U(x2c)U(0))yB(0).\displaystyle+\big{(}U^{\prime}(x_{2}^{c})(2k^{2}(c-U(0))-U^{\prime\prime}(0))-U^{\prime\prime}(x_{2}^{c})U^{\prime}(0)\big{)}y_{B}(0).

We shall use (5.22) to replace cRyB\partial_{c_{R}}y_{B} and cRyB\partial_{c_{R}}y_{B}^{\prime} by DcyBD_{c}y_{B} and (DcyB)(D_{c}y_{B})^{\prime}, the latter of which would produce yB(0)y_{B}^{\prime\prime}(0). All those yB(0)y_{B}^{\prime\prime}(0) multiplied by U(0)cU(0)-c can be substituted by (5.1a), but we keep other yB(0)y_{B}^{\prime\prime}(0) terms in the expression. On the other hand, we use (5.24a) to substitute (DcyB)(0)(D_{c}y_{B})^{\prime\prime}(0) in ζ2+\zeta_{2+}, which turns out to be rather regular due to the multiplier (U(0)c)2(U(0)-c)^{2}. Finally, we can write

ζ2+=\displaystyle\zeta_{2+}= ξ2U(x2c)+f3(c)ψ0(0)+f4(c)ψ0(0)+f5(k,c)yB(0)+f6(k,c)yB(0)+f7(k,c)DcyB(0)\displaystyle-\xi_{2}U^{\prime\prime}(x_{2}^{c})+f_{3}(c)\psi_{0}(0)+f_{4}(c)\psi_{0}^{\prime}(0)+f_{5}(k,c)y_{B}(0)+f_{6}(k,c)y_{B}^{\prime}(0)+f_{7}(k,c)D_{c}y_{B}(0)
+f8(k,c)(DcyB)(0)+(g+σk2)yB(0),\displaystyle+f_{8}(k,c)(D_{c}y_{B})^{\prime}(0)+(g+\sigma k^{2})y_{B}^{\prime\prime}(0),

where the functions fj(k,c,x2)f_{j}(k,c,x_{2}), j=3,,8j=3,\ldots,8, are

f3=(cU(0))f1+3U(x2c)U(0),f4=cU(0),\displaystyle f_{3}=(c-U(0))f_{1}+3U^{\prime}(x_{2}^{c})-U^{\prime}(0),\quad f_{4}=c-U(0),
f5=k2((4U(x2c)U(0))(cU(0)))+U(0)(U(0)c)2U(0)U(x2c)U(0)U(x2c),\displaystyle f_{5}=k^{2}\big{(}(4U^{\prime}(x_{2}^{c})-U^{\prime}(0))(c-U(0))\big{)}+U^{\prime\prime\prime}(0)(U(0)-c)-2U^{\prime\prime}(0)U^{\prime}(x_{2}^{c})-U^{\prime}(0)U^{\prime\prime}(x_{2}^{c}),
f6=k2(U(0)c)2+(2U(x2c)U(0))(U(0)c)2U(x2c)(U(x2c)U(0)),\displaystyle f_{6}=-k^{2}(U(0)-c)^{2}+(2U^{\prime\prime}(x_{2}^{c})-U^{\prime\prime}(0))(U(0)-c)-2U^{\prime}(x_{2}^{c})(U^{\prime}(x_{2}^{c})-U^{\prime}(0)),
f7=2(k2(U(0)c)2+U(0)(U(0)c)U(x2c)U(0)),\displaystyle f_{7}=2\big{(}k^{2}(U(0)-c)^{2}+U^{\prime\prime}(0)(U(0)-c)-U^{\prime}(x_{2}^{c})U^{\prime}(0)\big{)},
f8=2((2U(x2c)U(0))(U(0)c)gσk2),\displaystyle f_{8}=2\big{(}(2U^{\prime}(x_{2}^{c})-U^{\prime}(0))(U(0)-c)-g-\sigma k^{2}\big{)},

and are at least C1C^{1} in cRc_{R} and x2x_{2}.

The terms yB(h)y_{B}^{\prime\prime}(-h) in ζ2\zeta_{2-} and yB(0)y_{B}^{\prime\prime}(0) in ζ2+\zeta_{2+} generate the most singular part of Dc2yBD_{c}^{2}y_{B} which, based on Lemma 5.5, takes the form

yS(x2)=yB(h)y+(h)y+(x2)+(g+σk2)yB(0)𝐅(k,c)y(x2)=g+σk2𝐅(k,c)(yB(h)y+(x2)+yB(0)y(x2)).y_{S}(x_{2})=-\frac{y_{B}^{\prime\prime}(-h)}{y_{+}(-h)}y_{+}(x_{2})+\frac{(g+\sigma k^{2})y_{B}^{\prime\prime}(0)}{\mathbf{F}(k,c)}y_{-}(x_{2})=\frac{g+\sigma k^{2}}{\mathbf{F}(k,c)}\big{(}-y_{B}^{\prime\prime}(-h)y_{+}(x_{2})+y_{B}^{\prime\prime}(0)y_{-}(x_{2})\big{)}.

Let

y~=Dc2yByS.\tilde{y}=D_{c}^{2}y_{B}-y_{S}.

Clearly, it satisfies the same non-homogeneous Rayleigh equation (5.26a) and boundary conditions

(5.27) y~(c,h)=ζ~2(c)2(DcyB)(c,h)\tilde{y}(c,-h)=\tilde{\zeta}_{2-}(c)\triangleq 2(D_{c}y_{B})^{\prime}(c,-h)
(5.28) (U(0)c)2y~(0)(U(0)(U(0)c)+g+σk2)y~(0)=ζ~2+(c)ζ2+(g+σk2)yB(0).\big{(}U(0)-c\big{)}^{2}\tilde{y}^{\prime}(0)-\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}\tilde{y}(0)=\tilde{\zeta}_{2+}(c)\triangleq\zeta_{2+}-(g+\sigma k^{2})y_{B}^{\prime\prime}(0).

Let y~3\tilde{y}_{3} and y~4\tilde{y}_{4} be the solutions to (5.26a) with zero boundary values in (5.1b) and non-homogeneous terms

ψ2Uc,ψ0+f2ψ0+2f1ψ0Uc,\frac{\psi_{2}}{U-c},\quad\frac{\psi_{0}^{\prime\prime}+f_{2}\psi_{0}+2f_{1}\psi_{0}^{\prime}}{U-c},

respectively. Using the above estimates of yBy_{B} and DcyBD_{c}y_{B} and applying Lemma 5.2, we obtain

|y~3|LcR,x22+μ|y~3|LcR,x22\displaystyle|\tilde{y}_{3}|_{L_{c_{R},x_{2}}^{2}}+\mu|\tilde{y}_{3}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\leq Cμ(|yB|LcR,x22+|DcyB|LcR,x22+μ1ϵ|yB|LcR,x22+μ1ϵ|(DcyB)|LcR,x22)\displaystyle C\mu\big{(}|y_{B}|_{L_{c_{R},x_{2}}^{2}}+|D_{c}y_{B}|_{L_{c_{R},x_{2}}^{2}}+\mu^{1-\epsilon}|y_{B}^{\prime}|_{L_{c_{R},x_{2}}^{2}}+\mu^{1-\epsilon}|(D_{c}y_{B})^{\prime}|_{L_{c_{R},x_{2}}^{2}}\big{)}
\displaystyle\leq Cμ52ϵ(|ξ1|+|ξ2|+μ1|ψ0|L2+|ψ0|L2+μ2|ξ|).\displaystyle C\mu^{\frac{5}{2}-\epsilon}\big{(}|\xi_{1}|+|\xi_{2}|+\mu^{-1}|\psi_{0}|_{L^{2}}+|\psi_{0}^{\prime}|_{L^{2}}+\mu^{-2}|\xi_{-}|\big{)}.

As cI0+c_{I}\to 0+, the convergence of yBy_{B} and DcyBD_{c}y_{B} implies that of ψ2\psi_{2} in LcRrWx21,rL_{c_{R}}^{r}W_{x_{2}}^{1,r} for any r[1,2)r\in[1,2). From Lemma 5.1(2a), we obtain the convergence of y~3\tilde{y}_{3} in LcRrLx2L_{c_{R}}^{r}L_{x_{2}}^{\infty}.

Again we apply Lemma 5.4 to estimate y~4\tilde{y}_{4}

|y~4|LcR,x22+μ|y~4|LcR,x22Cμ32ϵ(|ψ0|L2+|ψ0|L2+|ψ0|L2).|\tilde{y}_{4}|_{L_{c_{R},x_{2}}^{2}}+\mu|\tilde{y}_{4}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\leq C\mu^{\frac{3}{2}-\epsilon}\big{(}|\psi_{0}|_{L^{2}}+|\psi_{0}^{\prime}|_{L^{2}}+|\psi_{0}^{\prime\prime}|_{L^{2}}\big{)}.

As cI0+c_{I}\to 0+, Lemma 5.4(2a) implies that y~4\tilde{y}_{4} converges in Lx2LcR2L_{x_{2}}^{\infty}L_{c_{R}}^{2}.

The boundary values of y~\tilde{y} satisfy

|ζ~2|LcR2Cμϵ(μ2(|ξ1|+|ξ2|)+μ1|ψ0|L2+|ψ0|L2)+Cμ2|ξ|,|\tilde{\zeta}_{2-}|_{L_{c_{R}}^{2}}\leq C\mu^{-\epsilon}\big{(}\mu^{2}(|\xi_{1}|+|\xi_{2}|)+\mu^{-1}|\psi_{0}|_{L^{2}}+|\psi_{0}^{\prime}|_{L^{2}}\big{)}+C\mu^{-2}|\xi_{-}|,
|ζ~2+|LcR2\displaystyle|\tilde{\zeta}_{2+}|_{L_{c_{R}}^{2}}\leq C(|ξ2|+|ψ0(0)|+|ψ0(0)|+μ2(|yB|LcR2+|DcyB|LcR2+|yB|LcR2+|(DcyB)|LcR2)|x2=0)\displaystyle C\big{(}|\xi_{2}|+|\psi_{0}(0)|+|\psi_{0}^{\prime}(0)|+\mu^{-2}\big{(}|y_{B}|_{L_{c_{R}}^{2}}+|D_{c}y_{B}|_{L_{c_{R}}^{2}}+|y_{B}^{\prime}|_{L_{c_{R}}^{2}}+|(D_{c}y_{B})^{\prime}|_{L_{c_{R}}^{2}}\big{)}\big{|}_{x_{2}=0}\big{)}
\displaystyle\leq C(|ψ0(0)|+μ2(|ξ1|+|ξ2|+μ1ϵ|ψ0|L2+μϵ|ψ0|L2+μϵ|ξ|)),\displaystyle C\big{(}|\psi_{0}^{\prime}(0)|+\mu^{-2}\big{(}|\xi_{1}|+|\xi_{2}|+\mu^{-1-\epsilon}|\psi_{0}|_{L^{2}}+\mu^{-\epsilon}|\psi_{0}^{\prime}|_{L^{2}}+\mu^{-\epsilon}|\xi_{-}|\big{)}\big{)},

where we also used the boundary conditions of yBy_{B} and DcyBD_{c}y_{B} to express them in terms of yBy_{B}^{\prime} and DcyBD_{c}y_{B}^{\prime} at x2=0x_{2}=0. As cI0+c_{I}\to 0+, the convergence of yBy_{B} and DcyBD_{c}y_{B} at x2=0,hx_{2}=0,-h implies that of ξ±\xi_{\pm} in LcRrL_{c_{R}}^{r} for any r[1,2)r\in[1,2).

As y~3+y~4\tilde{y}_{3}+\tilde{y}_{4} plays the role of "ynhy_{nh}" in the representation of Dc2yBD_{c}^{2}y_{B} as given in Lemma 5.5, the above estimates and Lemma 5.5 imply

|y~|LcR,x22+μ|y~|LcR,x22C(μ12|ξ1|+μ12|ξ2|+μ12ϵ|ψ0|L2+μ12ϵ|ψ0|L2+μ32ϵ|ψ0|L2),|\tilde{y}|_{L_{c_{R},x_{2}}^{2}}+\mu|\tilde{y}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\leq C\big{(}\mu^{\frac{1}{2}}|\xi_{1}|+\mu^{\frac{1}{2}}|\xi_{2}|+\mu^{-\frac{1}{2}-\epsilon}|\psi_{0}|_{L^{2}}+\mu^{\frac{1}{2}-\epsilon}|\psi_{0}^{\prime}|_{L^{2}}+\mu^{\frac{3}{2}-\epsilon}|\psi_{0}^{\prime\prime}|_{L^{2}}\big{)},

where the ψ0(0)\psi_{0}^{\prime}(0) term was bounded by the other norms of ψ0\psi_{0} via interpolation. Finally, using (5.22) one can compute

(5.29) cR2=U(x2c)2(Dc22x2Dc+x22)(U(x2c))3U(x2c)(Dcx2).\partial_{c_{R}}^{2}=U^{\prime}(x_{2}^{c})^{-2}(D_{c}^{2}-2\partial_{x_{2}}D_{c}+\partial_{x_{2}}^{2})-(U^{\prime}(x_{2}^{c}))^{-3}U^{\prime\prime}(x_{2}^{c})(D_{c}-\partial_{x_{2}}).

This relationship and the definition of y~B\tilde{y}_{B} and y~\tilde{y} yield

y~B=\displaystyle\tilde{y}_{B}= cR2yB+1U(x2c)2(yB+g+σk2𝐅(k,c)(yB(h)y+yB(0)y))\displaystyle\partial_{c_{R}}^{2}y_{B}+\frac{1}{U^{\prime}(x_{2}^{c})^{2}}\Big{(}-y_{B}^{\prime\prime}+\frac{g+\sigma k^{2}}{\mathbf{F}(k,c)}\big{(}y_{B}^{\prime\prime}(-h)y_{+}-y_{B}^{\prime\prime}(0)y_{-}\big{)}\Big{)}
=\displaystyle= 1U(x2c))2(y~2(DcyB)U(x2c)U(x2c)(DcyByB)).\displaystyle\frac{1}{U^{\prime}(x_{2}^{c}))^{2}}\Big{(}\tilde{y}-2(D_{c}y_{B})^{\prime}-\frac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}(D_{c}y_{B}-y_{B}^{\prime})\Big{)}.

Therefore the desired estimate on y~B\tilde{y}_{B} follows from those of y~\tilde{y}, yBy_{B}, and DcyBD_{c}y_{B}. The convergence of y~B\tilde{y}_{B} is also obtained much as that of DcyBD_{c}y_{B}. ∎

6. Solutions to the Euler equation linearized at shear flows

In this section, we finally return to the linearized flow of the capillary gravity water waves at the shear flow U(x2)U(x_{2}) in both the horizontally LL-periodic (in x1x_{1}) case and the x1x_{1}\in\mathbb{R} case. Under the assumption (4.9) of the absence of singular modes for all kk, we shall show that a.) inviscid damping occurs to a large component (remotely related to the rotational part) of the solutions and b.) what is left in the solutions are superpositions of non-singular modes (smooth eigenfunctions). The latter is a linear dispersive flow which is asymptotic to the linear irrotational flow for high spatial wave numbers kk.

6.1. Estimating each Fourier mode of the linear solutions

Based on (2.11) and the formula of the inverse Laplace transform, we first derive some integral representation formulas of the linear solution (v^(t,k,x2),η^(t,k,x2))\big{(}\hat{v}(t,k,x_{2}),\hat{\eta}(t,k,x_{2})\big{)} of (2.6) for a fixed wave number k0k\neq 0 satisfying (5.8). This procedure is essentially obtaining the linear solution group from contour integrals of the resolvents of the linear operator defined by the linearized water wave problem at the shear flow. Subsequently estimates of solutions are obtained using these formulas. Due to the conjugacy relation v^(t,k,x2)=v^(t,k,x2)¯\hat{v}(t,-k,x_{2})=\overline{\hat{v}(t,k,x_{2})} and η^(t,k)=η^(t,k)¯\hat{\eta}(t,-k)=\overline{\hat{\eta}(t,k)}, we shall mostly work on estimates for k>0k>0 in this subsection, unless otherwise specified.

Recall 𝐅\mathbf{F} defined in (4.1). Denote the set of non-singular modes

(6.1) R(k)={cU([h,0])𝐅(k,c)=0}R(k)=\{c\notin U([-h,0])\mid\mathbf{F}(k,c)=0\}

Throughout this subsection, we fix k0k\neq 0 and assume (5.8). We shall also use (5.6), possibly after choosing smaller ρ0\rho_{0}. The continuity of 𝐅\mathbf{F} and (5.8) imply that R(k)R(k) is a finite set, which consists of only simple roots c±(k)c^{\pm}(k) for large kk due to Lemma 4.2(3). We shall work on the following type of neighborhoods of U([h,0])U([-h,0])\subset\mathbb{C}

(6.2) 𝒟r1,r2=[r1+U(h),U(0)+r1]+i[r2,r2]R(k)c,r1,r2(0,ρ0),\mathcal{D}_{r_{1},r_{2}}=[-r_{1}+U(-h),U(0)+r_{1}]+i[-r_{2},r_{2}]\subset R(k)^{c},\quad r_{1},r_{2}\in(0,\rho_{0}),

where ρ0\rho_{0} is given in (5.8).

Recall the Laplace transform V2(k,c,x2)V_{2}(k,c,x_{2}) of v^2(t,k,x2)\hat{v}_{2}(t,k,x_{2}), defined by (2.9) and (2.10), is the solution of the boundary value problem (2.11) of the Rayleigh equation, or equivalently, the solution to (5.1) and (5.20) with

(6.3) ψ=ω^0(k,x2)=ik1(k2x22)v^20,ξ=0,ξ1=(g+σk2)η^0(k),ξ2=ik1v^20(k,0),\psi=-\hat{\omega}_{0}(k,x_{2})=-ik^{-1}(k^{2}-\partial_{x_{2}}^{2})\hat{v}_{20},\;\;\xi_{-}=0,\;\;\xi_{1}=(g+\sigma k^{2})\hat{\eta}_{0}(k),\;\;\xi_{2}=-ik^{-1}\hat{v}_{20}^{\prime}(k,0),

and ω^0(k,x2)\hat{\omega}_{0}(k,x_{2}), η^0(k)\hat{\eta}_{0}(k) and v^20(k,x2)\hat{v}_{20}(k,x_{2}) are the Fourier transforms with respect to x1x_{1} of the initial values ω0(x)\omega_{0}(x), η0(x1)\eta_{0}(x_{1}) and v20(x)v_{20}(x). The solution V2(k,c,x2)V_{2}(k,c,x_{2}) to (2.11) is still given by Lemma 5.5 along with (5.20) and (6.3). More explicitly, if 𝐅(k,c)0\mathbf{F}(k,c)\neq 0, then

(6.4) V2(k,c,x2)=(g+σk2)η^0(k)ik(U(0)c)v^20(k,0)𝐅(k,c)y(k,c,x2)+ynh(k,c,x2),\begin{split}V_{2}(k,c,x_{2})=&\frac{(g+\sigma k^{2})\hat{\eta}_{0}(k)-\frac{i}{k}(U(0)-c)\hat{v}_{20}^{\prime}(k,0)}{\mathbf{F}(k,c)}y_{-}(k,c,x_{2})+y_{nh}(k,c,x_{2}),\end{split}

where y±y_{\pm} are solutions to the homogeneous Rayleigh equation (3.24) satisfying initial conditions (3.53) and ynhy_{nh} the solution to (5.1) given by (5.3) with ζ±=0\zeta_{\pm}=0 and ψ=ω^0(k,x2)\psi=\hat{\omega}_{0}(k,x_{2}). The Laplace transform η~(k,c)\tilde{\eta}(k,c) of η^(t,k)\hat{\eta}(t,k) can be computed by using (2.12) and the boundary condition (5.1b) along with (5.20), (6.3), and (5.6)

(6.5) η~(k,c)=V2(k,c,0)+η^0(k)ik(U(0)c)=V2(k,c,0)(U(0)c)+U(0)η^0(k)+ikv^20(k,0)ik(U(0)(U(0)c)+g+σk2),k0.\begin{split}\tilde{\eta}(k,c)=&\frac{V_{2}(k,c,0)+\hat{\eta}_{0}(k)}{ik(U(0)-c)}=\frac{V_{2}^{\prime}(k,c,0)(U(0)-c)+U^{\prime}(0)\hat{\eta}_{0}(k)+\frac{i}{k}\hat{v}_{20}^{\prime}(k,0)}{ik\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}},\quad k\neq 0.\end{split}

We shall also need the following quantities

(6.6) 𝐛(t,k,c,x2)=(ik)Res(V2eik(cc)t,c),𝐛S(t,k,c)=(ik)Res(η~eik(cc)t,c)=Res(V2(k,c,0)eik(cc)t/(U(0)c),c)\begin{split}&\mathbf{b}(t,k,c_{*},x_{2})=-(ik)Res\big{(}V_{2}e^{-ik(c-c_{*})t},c_{*}\big{)},\\ &\mathbf{b}_{S}(t,k,c_{*})=-(ik)Res\big{(}\tilde{\eta}e^{-ik(c-c_{*})t},c_{*}\big{)}=-Res\big{(}V_{2}(k,c,0)e^{-ik(c-c_{*})t}/(U(0)-c),c_{*}\big{)}\end{split}

where Res(f(z),z)Res(f(z),z_{*}) is the residue of a meromorphic function f(z)f(z) at zz_{*}. Apparently 𝐛=𝐛S=0\mathbf{b}=\mathbf{b}_{S}=0 unless 𝐅(k,c)=0\mathbf{F}(k,c_{*})=0, or equivalently cR(k)c_{*}\in R(k). The following lemma is obtained from applying the inverse Laplace transform.

Lemma 6.1.

Assume UC3U\in C^{3} and k>0k>0 satisfies (5.8), then for any r1,r2(0,ρ0)r_{1},r_{2}\in(0,\rho_{0}), we have

v^2(t,k,x2)=v^2c+v^2pk2π𝒟r1,r2eikctV2(k,c,x2)dc+cR(k)eickt𝐛(t,k,c,x2),\hat{v}_{2}(t,k,x_{2})=\hat{v}_{2}^{c}+\hat{v}_{2}^{p}\triangleq-\frac{k}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}e^{-ikct}V_{2}(k,c,x_{2})dc+\sum_{c_{*}\in R(k)}e^{-ic_{*}kt}\mathbf{b}(t,k,c_{*},x_{2}),
η^(t,k)=η^c+η^pk2π𝒟r1,r2eikctη~(k,c)dc+cR(k)eickt𝐛S(k,c).\hat{\eta}(t,k)=\hat{\eta}^{c}+\hat{\eta}^{p}\triangleq-\frac{k}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}e^{-ikct}\tilde{\eta}(k,c)dc+\sum_{c_{*}\in R(k)}e^{-ic_{*}kt}\mathbf{b}_{S}(k,c_{*}).

From Lemma 4.2, cR(k)c_{*}\in R(k) implies y(k,c,0)0y_{-}(k,c,0)\neq 0 and thus F(k,c)F(k,c) is well-defined for cc near cc_{*}. In part (2), similar types of formula and estimates of 𝐛S\mathbf{b}_{S} can be obtained from those of 𝐛\mathbf{b} and (6.6). In the subsequent analysis, the limits of the above contour integrals as 𝒟r1,r2\mathcal{D}_{r_{1},r_{2}} shrinks to U([h,0])U([-h,0]) will be taken and estimated whenever needed.

Proof.

From the definition (2.9) and the inverse Laplace transform formula (2.15), we have

v^2(t,k,x2)=k2π+iγ++iγeikctV2(k,c,x2)dc,\hat{v}_{2}(t,k,x_{2})=\frac{k}{2\pi}\int_{-\infty+i\gamma}^{+\infty+i\gamma}e^{-ikct}V_{2}(k,c,x_{2})dc,

where γ>0\gamma>0 is chosen such that the above integrand is analytic for cI>γc_{I}>\gamma. Apparently V2V_{2} is analytic in c(U([h,0]){𝐅=0})c\notin\big{(}U([-h,0])\cup\{\mathbf{F}=0\}\big{)}. In order to analyze V2V_{2} for |c|1|c|\gg 1, we first consider y+y_{+} and then ynhy_{nh} for |c|1|c|\gg 1. From Lemma 3.3 and initial conditions (3.53), it holds that

0<lim inf|c||y+(k,c,x2)|/(1+|c|2)lim sup|c||y+(k,c,x2)|/(1+|c|2)<.0<\liminf_{|c|\to\infty}|y_{+}(k,c,x_{2})|/(1+|c|^{2})\leq\limsup_{|c|\to\infty}|y_{+}(k,c,x_{2})|/(1+|c|^{2})<\infty.

Along with (6.3) and Lemma 3.9 which yields the boundedness of yy_{-} for |c|1|c|\gg 1, it implies

lim sup|c||c||ynh(k,c,x2)|<.\limsup_{|c|\to\infty}|c||y_{nh}(k,c,x_{2})|<\infty.

From Lemma 4.2(2) and again Lemma 3.9, we obtain333Through a more careful analysis we may obtain a Taylor expansion of V2V_{2} in terms of 1c\frac{1}{c} as |c||c|\to\infty.

lim sup|c||c||V2(k,c,x2)|<.\limsup_{|c|\to\infty}|c||V_{2}(k,c,x_{2})|<\infty.

As |eikct|=ektImc|e^{-ikct}|=e^{kt\text{Im}\,c}, the Cauchy integral theorem yields

iγ+iγeikctV2(k,c,x2)dc=0 and v^2(t,k,x2)=k2π(+iγ++iγiγ+iγ)eikctV2(k,c,x2)dc.\displaystyle\int_{-\infty-i\gamma}^{+\infty-i\gamma}e^{-ikct}V_{2}(k,c,x_{2})dc=0\;\text{ and }\;\hat{v}_{2}(t,k,x_{2})=\frac{k}{2\pi}\Big{(}\int_{-\infty+i\gamma}^{+\infty+i\gamma}-\int_{-\infty-i\gamma}^{+\infty-i\gamma}\Big{)}e^{-ikct}V_{2}(k,c,x_{2})dc.

The desired expression of v^2\hat{v}_{2} follows immediately from the residue calculation.

Concerning η^\hat{\eta}, one first obtains

η^(t,k)=k2π(+iγ++iγiγ+iγ)eikctη~(k,c)dc.\hat{\eta}(t,k)=\frac{k}{2\pi}\Big{(}\int_{-\infty+i\gamma}^{+\infty+i\gamma}-\int_{-\infty-i\gamma}^{+\infty-i\gamma}\Big{)}e^{-ikct}\tilde{\eta}(k,c)dc.

Using the first expression in (6.5), the desired formula for η^\hat{\eta} is derived via the same arguments as in the above. In particular, the η^0\hat{\eta}_{0} term does not contribute to the residue as R(k)R(k) is away from U([h,0])U([-h,0]) due to assumption (5.8). ∎

From the divergence free condition on the velocity, it holds that the Fourier transform (in x1x_{1}) of the velocity field satisfies ikv^1=v^2ik\hat{v}_{1}=-\hat{v}_{2}^{\prime}. Therefore, we have

Corollary 6.1.1.

Under the assumptions of Lemma 6.1, we have

v^1(t,k,x2)=v^1c+v^1pi2π𝒟r1,r2eikctV2(k,c,x2)dc+cR(k)ikeickt𝐛(t,k,c,x2).\hat{v}_{1}(t,k,x_{2})=\hat{v}_{1}^{c}+\hat{v}_{1}^{p}\triangleq-\frac{i}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}e^{-ikct}V_{2}^{\prime}(k,c,x_{2})dc+\sum_{c_{*}\in R(k)}\frac{i}{k}e^{-ic_{*}kt}\mathbf{b}^{\prime}(t,k,c_{*},x_{2}).

In the following lemma, we give some basic properties of 𝐛(t,k,c,x2)\mathbf{b}(t,k,c,x_{2}) and 𝐛S(t,k,c)\mathbf{b}_{S}(t,k,c) at some cR(k)c_{*}\in R(k). Since cc_{*} is away from U([h,0])U([-h,0]) and 𝐅(k,)\mathbf{F}(k,\cdot) and F(k,)F(k,\cdot) are analytic in a neighborhood of cc_{*}, the assumption (5.8) is not needed.

Lemma 6.2.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, and k>0k>0. Let cR(k)c_{*}\in R(k) be a root of 𝐅(k,)\mathbf{F}(k,\cdot) (or equivalently, of F(k,)F(k,\cdot) defined in (4.2)) of degree n1n\geq 1, then the following hold.

  1. (1)

    eikct𝐛(t,k,c,x2)e^{-ikc_{*}t}\mathbf{b}(t,k,c_{*},x_{2}) is a solution to (2.8).

  2. (2)

    𝐛(t,k,c,x2)\mathbf{b}(t,k,c_{*},x_{2}) is a linear combination of tl1cl2y(k,c,x2)t^{l_{1}}\partial_{c}^{l_{2}}y_{-}(k,c_{*},x_{2}), 0l1+l2l=n10\leq l_{1}+l_{2}\leq l=n-1, and 𝐛S(t,k,c)\mathbf{b}_{S}(t,k,c_{*}) a linear combination of tlt^{l}, 0ln10\leq l\leq n-1, with coefficients depending on kk and cc_{*}. The leading terms of 𝐛(t,k,c,x2)\mathbf{b}(t,k,c_{*},x_{2}) with l1+l2=n1l_{1}+l_{2}=n-1 are given by

    (6.7) (n!)(ik)l1+1l1!l2!cnF(k,c)((g+σk2)η^0(k)ik(U(0)c)v^20(k,0)+(U(0)c)2y(k,c,0)h0(yω^0)(k,c,x2)U(x2)cdx2)tl1cl2y(k,c,x2)y(k,c,0),\begin{split}\frac{(n!)(-ik)^{l_{1}+1}}{l_{1}!l_{2}!\partial_{c}^{n}F(k,c_{*})}\Big{(}&(g+\sigma k^{2})\hat{\eta}_{0}(k)-\frac{i}{k}(U(0)-c_{*})\hat{v}_{20}^{\prime}(k,0)\\ &+\frac{(U(0)-c_{*})^{2}}{y_{-}(k,c_{*},0)}\int_{-h}^{0}\frac{(y_{-}\hat{\omega}_{0})(k,c_{*},x_{2}^{\prime})}{U(x_{2}^{\prime})-c_{*}}dx_{2}^{\prime}\Big{)}\frac{t^{l_{1}}\partial_{c}^{l_{2}}y_{-}(k,c_{*},x_{2})}{y_{-}(k,c_{*},0)},\end{split}

    and the leading terms of ik(U(0)c)𝐛S(t,k,c)ik(U(0)-c_{*})\mathbf{b}_{S}(t,k,c_{*}) is given by the above expression evaluated at x2=0x_{2}=0.

  3. (3)

    If cc_{*} is a simple root of 𝐅(k,)\mathbf{F}(k,\cdot), i.e., n=1n=1, then 𝐛\mathbf{b} and ik(U(0)c)𝐛Sik(U(0)-c_{*})\mathbf{b}_{S} are given by the above expression and there exists C>0C>0 determined only by |U|Cl01|U^{\prime}|_{C^{l_{0}-1}} and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} such that

    |x2n2𝐛(k,c,x2)|C|cF(k,c)|1(\displaystyle|\partial_{x_{2}}^{n_{2}}\mathbf{b}(k,c_{*},x_{2})|\leq C|\partial_{c}F(k,c_{*})|^{-1}\Big{(} |k|μ2|η^0(k)|+(1+|c|)|v^20(k,0)|\displaystyle|k|\mu^{-2}|\hat{\eta}_{0}(k)|+(1+|c_{*}|)|\hat{v}_{20}^{\prime}(k,0)|
    +|k|μ32eμ1h(1+|c|2)|ω^0(k)|Lx22dist(c,U([h,0]))|y(k,c,0)|)|μ1n2eμ1(x2+h)y(k,c,0)|,\displaystyle+\frac{|k|\mu^{\frac{3}{2}}e^{\mu^{-1}h}(1+|c_{*}|^{2})|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}}{dist(c_{*},U([-h,0]))|y_{-}(k,c_{*},0)|}\Big{)}\Big{|}\frac{\mu^{1-n_{2}}e^{\mu^{-1}(x_{2}+h)}}{y_{-}(k,c_{*},0)}\Big{|},

    for any n2[0,l0]n_{2}\in[0,l_{0}], where we recall μ=(1+k2)12\mu=(1+k^{2})^{-\frac{1}{2}}.

Proof.

According to Lemma 4.1(4), 𝐅(k,c)=0\mathbf{F}(k,c_{*})=0 implies y(k,c,0)0y_{-}(k,c_{*},0)\neq 0 and thus F(k,c)F(k,c) is analytic in cc for cc near cc_{*} and the degree of cc_{*} as a root of both F(k,)F(k,\cdot) and 𝐅(k,)\mathbf{F}(k,\cdot) is n1n\geq 1. By the definition of R(k)R(k) and the analyticity of F(k,)F(k,\cdot), cR(k)c_{*}\in R(k) is an isolated root of F(k,)F(k,\cdot). Let 1R>01\gg R>0 such that there are no other roots of F(k,)F(k,\cdot) in the disk B(c,R)B(c_{*},R) centered at cc_{*} with radius RR. Using the fact that V2(k,c,x2)V_{2}(k,c,x_{2}) solves (2.11), one may compute

(t+ikU)(k2x22)(eikct𝐛(t,k,c,x2))=t+ikU2πB(c,R)keikct(k2x22)V2(k,c,x2)dc\displaystyle(\partial_{t}+ikU)(k^{2}-\partial_{x_{2}}^{2})(e^{-ikc_{*}t}\mathbf{b}(t,k,c_{*},x_{2}))=\frac{\partial_{t}+ikU}{2\pi}\oint_{\partial B(c_{*},R)}ke^{-ikct}(k^{2}-\partial_{x_{2}}^{2})V_{2}(k,c,x_{2})dc
=\displaystyle= ik22πB(c,R)eikct(Uc)(k2x22)V2(k,c,x2)dc=ik2U2πB(c,R)eikctV2(k,c,x2)dc\displaystyle\frac{ik^{2}}{2\pi}\oint_{\partial B(c_{*},R)}e^{-ikct}(U-c)(k^{2}-\partial_{x_{2}}^{2})V_{2}(k,c,x_{2})dc=-\frac{ik^{2}U^{\prime\prime}}{2\pi}\oint_{\partial B(c_{*},R)}e^{-ikct}V_{2}(k,c,x_{2})dc
=\displaystyle= ikUeikct𝐛(t,k,c,x2),\displaystyle-ikU^{\prime\prime}e^{-ikc_{*}t}\mathbf{b}(t,k,c_{*},x_{2}),

and thus (2.8a) is satisfied. Similar calculation also proves the boundary condition (2.8b) at x2=0x_{2}=0. The zero boundary value at x2=hx_{2}=-h is obvious from that of V2V_{2} at x2=hx_{2}=-h. Therefore statement (1) is proved.

To analyze 𝐛\mathbf{b} in more details, let

F1(c)=(cc)nF(k,c)F1(c)=cnF(k,c)/(n!)0,F_{1}(c)=(c-c_{*})^{-n}F(k,c)\implies F_{1}(c_{*})=\partial_{c}^{n}F(k,c_{*})/(n!)\neq 0,

and

y~(c,x2)=y+(k,c,x2)(U(0)c)2(g+σk2)y(k,c,0)y(k,c,x2).\tilde{y}(c,x_{2})=y_{+}(k,c,x_{2})-\frac{(U(0)-c)^{2}}{(g+\sigma k^{2})y_{-}(k,c,0)}y_{-}(k,c,x_{2}).

From the initial conditions (3.53) of y±y_{\pm}, it is straight forward to verify

y~(c,0)=0,y~(c,0)=F(k,c)g+σk2=O(|cc|n)y+(x2)=(U(0)c)2y(x2)(g+σk2)y(0)+O(|cc|n).\tilde{y}(c,0)=0,\quad\tilde{y}^{\prime}(c,0)=-\frac{F(k,c)}{g+\sigma k^{2}}=O(|c-c_{*}|^{n})\implies y_{+}(x_{2})=\frac{(U(0)-c)^{2}y_{-}(x_{2})}{(g+\sigma k^{2})y_{-}(0)}+O(|c-c_{*}|^{n}).

Using the above expression to substitute y+(k,c,x2)y_{+}(k,c,x_{2}) in the residue (in the definition of 𝐛\mathbf{b}) and observing that the O(|cc|n)O(|c-c_{*}|^{n}) term cancels the singularity of y+(k,c,h)y_{+}(k,c,-h) for |cc|1|c-c_{*}|\ll 1 which results in an analytic function contributing nothing to the residue, we have

Res(ynh(k,c,x2)eik(cc)t,c)=Res((U(0)c)2y(k,c,x2)eik(cc)t(g+σk2)y(k,c,0)y+(k,c,h)h0(yω^0)(k,c,x2)U(x2)cdx2,c).Res(y_{nh}(k,c,x_{2})e^{-ik(c-c_{*})t},c_{*})=Res\Big{(}\frac{(U(0)-c)^{2}y_{-}(k,c,x_{2})e^{-ik(c-c_{*})t}}{(g+\sigma k^{2})y_{-}(k,c,0)y_{+}(k,c,-h)}\int_{-h}^{0}\frac{(y_{-}\hat{\omega}_{0})(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime},c_{*}\Big{)}.

From definitions (5.3), (4.1), and (4.2), of ynhy_{nh}, 𝐅\mathbf{F}, and FF, (6.3), we have

Res(ynh(k,c,x2)eik(cc)t,c)=Res((U(0)c)2y(k,c,x2)eik(cc)t(cc)nF1(c)y(k,c,0)2h0(yω^0)(k,c,x2)U(x2)cdx2,c)\displaystyle Res(y_{nh}(k,c,x_{2})e^{-ik(c-c_{*})t},c_{*})=Res\Big{(}\frac{(U(0)-c)^{2}y_{-}(k,c,x_{2})e^{-ik(c-c_{*})t}}{(c-c_{*})^{n}F_{1}(c)y_{-}(k,c,0)^{2}}\int_{-h}^{0}\frac{(y_{-}\hat{\omega}_{0})(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime},c_{*}\Big{)}
=\displaystyle= 1(n1)!cn1((U(0)c)2y(k,c,x2)eik(cc)tF1(c)y(k,c,0)2h0(yω^0)(k,c,x2)U(x2)cdx2)|c=c\displaystyle\frac{1}{(n-1)!}\partial_{c}^{n-1}\Big{(}\frac{(U(0)-c)^{2}y_{-}(k,c,x_{2})e^{-ik(c-c_{*})t}}{F_{1}(c)y_{-}(k,c,0)^{2}}\int_{-h}^{0}\frac{(y_{-}\hat{\omega}_{0})(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime}\Big{)}\Big{|}_{c=c_{*}}
=\displaystyle= l=0n11l!(nl1)!cnl1((U(0)c)2eik(cc)tF1(c)y(k,c,0)2h0(yω^0)(k,c,x2)U(x2)cdx2)|c=ccly(k,c,x2).\displaystyle\sum_{l=0}^{n-1}\frac{1}{l!(n-l-1)!}\partial_{c}^{n-l-1}\Big{(}\frac{(U(0)-c)^{2}e^{-ik(c-c_{*})t}}{F_{1}(c)y_{-}(k,c,0)^{2}}\int_{-h}^{0}\frac{(y_{-}\hat{\omega}_{0})(k,c,x_{2}^{\prime})}{U(x_{2}^{\prime})-c}dx_{2}^{\prime}\Big{)}\Big{|}_{c=c_{*}}\partial_{c}^{l}y_{-}(k,c_{*},x_{2}).

Therefore this residue is a linear combination of tl1cl2y(k,c,x2)t^{l_{1}}\partial_{c}^{l_{2}}y_{-}(k,c_{*},x-2), 0l1+l2n10\leq l_{1}+l_{2}\leq n-1, with coefficients depending on kk and cc_{*}. The coefficients for l1+l2=n1l_{1}+l_{2}=n-1 are given by

(cl1eik(cc)t)|c=cl1!l2!tl1(U(0)c)2F1(c)y(k,c,0)2h0(yω^0)(k,c,x2)U(x2)cdx2\displaystyle\frac{(\partial_{c}^{l_{1}}e^{-ik(c-c_{*})t})|_{c=c_{*}}}{l_{1}!l_{2}!t^{l_{1}}}\frac{(U(0)-c_{*})^{2}}{F_{1}(c_{*})y_{-}(k,c_{*},0)^{2}}\int_{-h}^{0}\frac{(y_{-}\hat{\omega}_{0})(k,c_{*},x_{2}^{\prime})}{U(x_{2}^{\prime})-c_{*}}dx_{2}^{\prime}
=\displaystyle= (n!)(ik)l1(U(0)c)2l1!l2!cnF(k,c)y(k,c,0)2h0(yω^0)(k,c,x2)U(x2)cdx2.\displaystyle\frac{(n!)(-ik)^{l_{1}}(U(0)-c_{*})^{2}}{l_{1}!l_{2}!\partial_{c}^{n}F(k,c_{*})y_{-}(k,c_{*},0)^{2}}\int_{-h}^{0}\frac{(y_{-}\hat{\omega}_{0})(k,c_{*},x_{2}^{\prime})}{U(x_{2}^{\prime})-c_{*}}dx_{2}^{\prime}.

The contributions of the terms involving η0(k)\eta_{0}(k) and v^20(k,0)\hat{v}_{20}^{\prime}(k,0) can be analyzed similarly (actually simpler as y+y_{+} is not involved) and we obtain the desired statement (2) on the form of 𝐛\mathbf{b} and 𝐛S\mathbf{b}_{S}.

If cR(k)c_{*}\in R(k) is a simple root of 𝐅(k,)\mathbf{F}(k,\cdot), i.e. n=1n=1, then 𝐛\mathbf{b} and 𝐛s\mathbf{b}_{s} have only one term with l1=l2=0l_{1}=l_{2}=0 and are constants in tt as given in statement (2). It along with Lemma 3.9 readily leads to its estimate. ∎

Corollary 6.2.1.

v^2c\hat{v}_{2}^{c} is also a solution to (2.8). Moreover if cc_{*} is a simple root of F(k,)F(k,\cdot), then the corresponding eigenvalue ikc-ikc_{*} is algebraically simple in the subspace of the kk-th Fourier modes.

Based on the above lemmas, it is natural to define

(6.8) 𝐏(k,c):(v^0,η^0)(i𝐛(0,k,c,)/k,𝐛(0,k,c,),𝐛S(0,k,c)),𝐗(k,c)=range(𝐏(k,c)).\mathbf{P}(k,c_{*}):(\hat{v}_{0},\hat{\eta}_{0})\to\big{(}i\mathbf{b}(0,k,c_{*},\cdot)^{\prime}/k,\mathbf{b}(0,k,c_{*},\cdot),\mathbf{b}_{S}(0,k,c_{*})\big{)},\;\mathbf{X}(k,c_{*})=range(\mathbf{P}(k,c_{*})).

The following lemma gives that 𝐏(k,c)\mathbf{P}(k,c_{*}) defines the invariant spectral projection to the eigenspace 𝐗(k,c)\mathbf{X}(k,c_{*}) of ikc-ikc_{*} spanned by cly(k,c,)\partial_{c}^{l}y_{-}(k,c_{*},\cdot), 0ln10\leq l\leq n-1.

Lemma 6.3.

Assume the same conditions as in Lemma 6.2, then

𝐗(k,c)=span{(icly(k,c,)/k,cly(k,c,),cly(k,c,0)/(ik(U(0)c))l=0,,n1},\mathbf{X}(k,c_{*})=span\big{\{}\big{(}i\partial_{c}^{l}y_{-}^{\prime}(k,c_{*},\cdot)/k,\partial_{c}^{l}y_{-}(k,c_{*},\cdot),\partial_{c}^{l}y_{-}(k,c_{*},0)/(ik(U(0)-c_{*})\big{)}\mid l=0,\ldots,n-1\},

is an invariant subspace of (2.6) and

𝐏(k,c):(v^0,η^0)(i𝐛(0,k,c,)/k,𝐛(0,k,c,),𝐛S(0,k,c))\mathbf{P}(k,c_{*}):(\hat{v}_{0},\hat{\eta}_{0})\to\big{(}i\mathbf{b}(0,k,c_{*},\cdot)^{\prime}/k,\mathbf{b}(0,k,c_{*},\cdot),\mathbf{b}_{S}(0,k,c_{*})\big{)}

is an invariant projection operator of (2.6) to 𝐗(k,c)\mathbf{X}(k,c_{*}) with

ker(ΣcR(k)𝐏(k,c))={(v^c(0,k,),η^c(0,k)) all initial values v^0(k,),η0(k)}.\ker\big{(}\Sigma_{c_{*}\in R(k)}\mathbf{P}(k,c_{*})\big{)}=\big{\{}\big{(}\hat{v}^{c}(0,k,\cdot),\hat{\eta}^{c}(0,k)\big{)}\mid\text{ all initial values }\hat{v}_{0}(k,\cdot),\eta_{0}(k)\big{\}}.
Proof.

The statement of the lemma is rather standard in the operator calculus and Laplace transform, while constructing solutions to (2.6) using Laplace transform is equivalent to using contour integrals of the resolvent operators in the complex spectral plane. We shall only outline the proof and skip some details.

Due to the translation invariance in tt of solutions to (2.6), the t=0t=0 in the definition of 𝐗(k,c)\mathbf{X}(k,c_{*}) can be replaced by any tt\in\mathbb{R}. From Lemma 6.2, all solutions (i𝐛/k,𝐛,𝐛S)(i\mathbf{b}^{\prime}/k,\mathbf{b},\mathbf{b}_{S}) are polynomials of tt of degree no more than n1n-1. It is standard to show inductively that 𝐗(k,c)\mathbf{X}(k,c_{*}) consists of all possible coefficients of tlt^{l}, which can be computed to be generated by cly(k,c,)\partial_{c}^{l}y_{-}(k,c_{*},\cdot), 0ln10\leq l\leq n-1, using (6.7) and the relationship between 𝐛\mathbf{b} and 𝐛S\mathbf{b}_{S}. The invariance of 𝐗(k,c)\mathbf{X}(k,c_{*}) under (2.6) is due to the fact that (i𝐛/k,𝐛,𝐛s)(i\mathbf{b}/k,\mathbf{b},\mathbf{b}_{s}) are solutions to (2.6). To show 𝐏(k,c)2=𝐏(k,c)\mathbf{P}(k,c_{*})^{2}=\mathbf{P}(k,c_{*}), let (u^0,ν^0)=𝐏(k,c)(v^0,η^0)𝐗(k,c)(\hat{u}_{0},\hat{\nu}_{0})=\mathbf{P}(k,c_{*})(\hat{v}_{0},\hat{\eta}_{0})\in\mathbf{X}(k,c_{*}). With this initial value, the solution (u^(t),ν^(t))(\hat{u}(t),\hat{\nu}(t)) is simply the (i𝐛/k,𝐛,𝐛S)(i\mathbf{b}/k,\mathbf{b},\mathbf{b}_{S}) component of the solution with the initial value (v^0,η^0)(\hat{v}_{0},\hat{\eta}_{0}). Hence (u^(t),ν^(t))(\hat{u}(t),\hat{\nu}(t)) takes the form given in Lemma 6.2(2). Its Laplace transform is analytic at all ccc\neq c_{*} and thus the (i𝐛/k,𝐛,𝐛S)(i\mathbf{b}/k,\mathbf{b},\mathbf{b}_{S}) component of (u^(t),ν^(t))(\hat{u}(t),\hat{\nu}(t)) is equal to itself. Therefore we obtain 𝐏(k,c)(u^0,ν^0)=(u^0,ν^0)\mathbf{P}(k,c_{*})(\hat{u}_{0},\hat{\nu}_{0})=(\hat{u}_{0},\hat{\nu}_{0}). Finally the description of the kernel of cR(k)𝐏(k,c)\sum_{c_{*}\in R(k)}\mathbf{P}(k,c_{*}) is obvious due to the fact that both (i𝐛/k,𝐛,𝐛S)(i\mathbf{b}/k,\mathbf{b},\mathbf{b}_{S}) and (v^c(t),η^c(t))(\hat{v}^{c}(t),\hat{\eta}^{c}(t)) are solutions. ∎

Remark 6.1.

In particular, if

v^20(k,x2)=y(k,c,x2),η^0=y(k,c,0)/(ik(U(0)c),\hat{v}_{20}(k,x_{2})=y_{-}(k,c_{*},x_{2}),\quad\hat{\eta}_{0}=y_{-}(k,c_{*},0)/\big{(}ik(U(0)-c_{*}\big{)},

then straight forward verification yields

V2(k,c,x2)=y(k,c,x2)ik(cc),v^c=0,𝐛=y(k,c,x2)=v^20,𝐛S=y(k,c,0)ik(U(0)c)=η^0.V_{2}(k,c,x_{2})=\frac{y_{-}(k,c_{*},x_{2})}{ik(c_{*}-c)},\quad\hat{v}^{c}=0,\quad\mathbf{b}=y_{-}(k,c_{*},x_{2})=\hat{v}_{20},\quad\mathbf{b}_{S}=\frac{y_{-}(k,c_{*},0)}{ik(U(0)-c_{*})}=\hat{\eta}_{0}.

From Lemma 2.3, ick-ic_{*}k is an eigenvalue (with the above eigenfunctions generated by y(k,c,x2)y_{-}(k,c_{*},x_{2})) of the linearized capillary gravity water wave at the shear flow, which has to be geometrically simple when restricted to the kk-th Fourier mode in x1x_{1}. Its algebraic multiplicity is equal to the degree of the root cc_{*} of 𝐅(k,)\mathbf{F}(k,\cdot). The eigenfunctions of the linearized irrotational capillary gravity wave are generated by 1ksinhk(x2+h)\frac{1}{k}\sinh k(x_{2}+h). From Lemmas 3.2(1) (with ρ=O(k52)\rho=O(k^{-\frac{5}{2}}), s=0s=0, C0=0C_{0}=0, and Θ1=Θ2=sinh\Theta_{1}=\Theta_{2}=\sinh) and 4.2(3), it is straight forward to estimate that, after normalizing the L2L^{2} norm of v2v_{2} to be 11, the L2L^{2} and H1H^{1} differences in the vv and η\eta components, respectively, between the eigenfunctions of (2.6) and the irrotational capillary gravity waves linearized at zero is of order O(k32)O(k^{-\frac{3}{2}}) as |k||k|\to\infty.

In the rest of this subsection we consider v^c(t,k,x2)\hat{v}^{c}(t,k,x_{2}) and η^c(t,k)\hat{\eta}^{c}(t,k). We shall always work on c[U(h)ρ0,U(0)+ρ0]+i[ρ0,ρ0]c\in[U(-h)-\rho_{0},U(0)+\rho_{0}]+i[-\rho_{0},\rho_{0}]. We first present some properties of V2V_{2} and η~\tilde{\eta}. Let us keep in mind that for analytic functions, c\partial_{c} and cR\partial_{c_{R}} are equivalent.

Lemma 6.4.

It holds that V2V_{2} and η~\tilde{\eta} are analytic in c(U([h,0])R(k))c\in\mathbb{C}\setminus\big{(}U([-h,0])\cup R(k)\big{)} and satisfy

V2(k,c¯,x2)=V2(k,c,x2)¯,η~(k,c¯,x2)=η~(k,c,x2)¯.V_{2}(-k,\bar{c},x_{2})=\overline{V_{2}(k,c,x_{2})},\quad\tilde{\eta}(-k,\bar{c},x_{2})=\overline{\tilde{\eta}(k,c,x_{2})}.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, and (5.8), then the following hold for some .

  1. (1)

    For any ϵ>0\epsilon>0, there exists C>0C>0 determined only by ϵ\epsilon, F0F_{0}, ρ0\rho_{0}, |U|Cl01|U^{\prime}|_{C^{l_{0}-1}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} (independent of kk\in\mathbb{R}) such that for any cI[0,ρ0]c_{I}\in[0,\rho_{0}],

    |V2|LcR,x22+μ|V2|LcR,x22+μ32|V2(0)|LcR2+μ12|V2(0)|LcR2\displaystyle|V_{2}|_{L_{c_{R},x_{2}}^{2}}+\mu|V_{2}^{\prime}|_{L_{c_{R},x_{2}}^{2}}+\mu^{\frac{3}{2}}|V_{2}^{\prime}(0)|_{L_{c_{R}}^{2}}+\mu^{\frac{1}{2}}|V_{2}(0)|_{L_{c_{R}}^{2}}
    \displaystyle\leq C(μ12|η^0(k)|+|k|1μ52|v^20(k,0)|+μ32ϵ|ω^0(k)|Lx22);\displaystyle C\big{(}\mu^{\frac{1}{2}}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{\frac{5}{2}}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{\frac{3}{2}-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)};
    |η~|LcR2C(|k|1μ|η^0(k)|+|k|2μ2|v^20(k,0)|+|k|1μ2ϵ|ω^0(k)|Lx22),|\tilde{\eta}|_{L_{c_{R}}^{2}}\leq C\big{(}|k|^{-1}\mu|\hat{\eta}_{0}(k)|+|k|^{-2}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|k|^{-1}\mu^{2-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)},

    if l04l_{0}\geq 4, then

    |cRη~|LcR2C(|k|1|η^0(k)|+|k|2μ2|v^20(k,0)|+|k|1μ1ϵ|ω^0(k)|Lx22+|k|1μ2ϵ|ω^0(k)|Lx22),|\partial_{c_{R}}\tilde{\eta}|_{L_{c_{R}}^{2}}\leq C\big{(}|k|^{-1}|\hat{\eta}_{0}(k)|+|k|^{-2}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|k|^{-1}\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+|k|^{-1}\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)},
    |cRV2|LcR,x22+μ|cRV2+V2U|LcR,x22+μ32|(cRV2+V2U)(0)|LcR2+μ12|cRV2(0)|LcR2\displaystyle|\partial_{c_{R}}V_{2}|_{L_{c_{R},x_{2}}^{2}}+\mu|\partial_{c_{R}}V_{2}^{\prime}+\tfrac{V_{2}^{\prime\prime}}{U^{\prime}}|_{L_{c_{R},x_{2}}^{2}}+\mu^{\frac{3}{2}}\big{|}\big{(}\partial_{c_{R}}V_{2}^{\prime}+\tfrac{V_{2}^{\prime\prime}}{U^{\prime}}\big{)}(0)\big{|}_{L_{c_{R}}^{2}}+\mu^{\frac{1}{2}}|\partial_{c_{R}}V_{2}(0)|_{L_{c_{R}}^{2}}
    \displaystyle\leq C(μ12|η^0(k)|+|k|1μ32|v^20(k,0)|+μ12ϵ|ω^0(k)|Lx22+μ32ϵ|ω^0(k)|Lx22);\displaystyle C\big{(}\mu^{-\frac{1}{2}}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{\frac{3}{2}}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{\frac{1}{2}-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{\frac{3}{2}-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)};

    and if UC5U\in C^{5}, then

    |V~2|LcR,x22C(μ32|η^0(k)|+|k|1μ12|v^20(k,0)|+μ12ϵ|ω^0(k)|Lx22+μ12ϵ|ω^0(k)|Lx22+μ32ϵ|ω^0(k)|Lx22),|\tilde{V}_{2}|_{L_{c_{R},x_{2}}^{2}}\leq C\big{(}\mu^{-\frac{3}{2}}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{\frac{1}{2}}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{-\frac{1}{2}-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{\frac{1}{2}-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}+\mu^{\frac{3}{2}-\epsilon}|\hat{\omega}_{0}^{\prime\prime}(k)|_{L_{x_{2}}^{2}}\big{)},

    where

    V~2(x2)=cR2V2(x2)V2(x2)U(x2)2+g+σk2𝐅(k,c)(V2(h)U(h)2y+(x2)V2(0)U(0)2y(x2)),\tilde{V}_{2}(x_{2})=\partial_{c_{R}}^{2}V_{2}(x_{2})-\frac{V_{2}^{\prime\prime}(x_{2})}{U^{\prime}(x_{2})^{2}}+\frac{g+\sigma k^{2}}{\mathbf{F}(k,c)}\Big{(}\frac{V_{2}^{\prime\prime}(-h)}{U^{\prime}(-h)^{2}}y_{+}(x_{2})-\frac{V_{2}^{\prime\prime}(0)}{U^{\prime}(0)^{2}}y_{-}(x_{2})\Big{)},

    and all the norms are taken on (cR,x2)[U(h)ρ0,U(0)+ρ0]×[h,0](c_{R},x_{2})\in[U(-h)-\rho_{0},U(0)+\rho_{0}]\times[-h,0].

  2. (2)

    As cI0+c_{I}\to 0+, on [r1+U(h),r1][-r_{1}+U(-h),r_{1}]

    V20(k,cR,x2)limcI0+V2(k,cR+icI,x2),η~0(k,cR,x2)limcI0+η~(k,cR+icI,x2)V_{20}(k,c_{R},x_{2})\triangleq\lim_{c_{I}\to 0+}V_{2}(k,c_{R}+ic_{I},x_{2}),\quad\tilde{\eta}_{0}(k,c_{R},x_{2})\triangleq\lim_{c_{I}\to 0+}\tilde{\eta}(k,c_{R}+ic_{I},x_{2})

    exist and the following hold.

    1. (a)

      Assume ω^0(k)L2\hat{\omega}_{0}(k)\in L^{2} and UC3U\in C^{3}, then for any r[1,2)r\in[1,2), V2V20V_{2}\to V_{20} in Lx2LcR2L_{x_{2}}^{\infty}L_{c_{R}}^{2}, V2V20V_{2}^{\prime}\to V_{20}^{\prime} in Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r}, and V2(0)V20(0)V_{2}^{\prime}(0)\to V_{20}^{\prime}(0) and η~η~0\tilde{\eta}\to\tilde{\eta}_{0} in LcR2L_{c_{R}}^{2}.

    2. (b)

      Assume ω^0(k)H1\hat{\omega}_{0}(k)\in H^{1} and UC4U\in C^{4}, then for any r[1,2)r\in[1,2) and q[1,)q\in[1,\infty), cRV2cRV20\partial_{c_{R}}V_{2}\to\partial_{c_{R}}V_{20} in Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r}, cRV2+V2UcRV20+V20U(x2)\partial_{c_{R}}V_{2}^{\prime}+\tfrac{V_{2}^{\prime\prime}}{U^{\prime}}\to\partial_{c_{R}}V_{20}^{\prime}+\tfrac{V_{20}^{\prime\prime}}{U^{\prime}(x_{2})} in Lx2qLcRrL_{x_{2}}^{q}L_{c_{R}}^{r}, and (cRV2+V2U)(0)(cRV20+V20U)(0)\big{(}\partial_{c_{R}}V_{2}^{\prime}+\tfrac{V_{2}^{\prime\prime}}{U^{\prime}}\big{)}(0)\to\big{(}\partial_{c_{R}}V_{20}^{\prime}+\tfrac{V_{20}^{\prime\prime}}{U^{\prime}}\big{)}(0) and cRη~cRη~0\partial_{c_{R}}\tilde{\eta}\to\partial_{c_{R}}\tilde{\eta}_{0} in LcRrL_{c_{R}}^{r}.

    3. (c)

      Assume ω^0(k)H2\hat{\omega}_{0}(k)\in H^{2} and UC5U\in C^{5}, then for any r[1,2)r\in[1,2), V~2\tilde{V}_{2} converges to its limit V~20\tilde{V}_{20} in Lx2LcRrL_{x_{2}}^{\infty}L_{c_{R}}^{r}.

Compared to Proposition 5.6, the modifications in the definition of V~2\tilde{V}_{2} is to make it analytic in cc which will make it more convenient in applying Lemma 6.5 in the below.

Proof.

The estimates of V2V_{2}, V2V_{2}^{\prime}, V2(0)V_{2}^{\prime}(0), cRV2\partial_{c_{R}}V_{2}, and their convergences are all direct corollaries of (6.3) and Proposition 5.6. The estimate of η~\tilde{\eta} and its convergence follows from the second expression of (6.5) and the above properties of V2V_{2}.

We also notice that, compared to Propositions 5.6, in the definition of V~2\tilde{V}_{2} as well as in the estimate related to cRV2\partial_{c_{R}}V_{2}^{\prime}, the U(x2c)U^{\prime}(x_{2}^{c}) in front of V2V_{2}^{\prime\prime}, V2(h)V_{2}^{\prime\prime}(-h), and V2(0)V_{2}^{\prime\prime}(0) had been replaced by U(x2)U^{\prime}(x_{2}), U(h)U^{\prime}(-h), and U(0)U^{\prime}(0), respectively. This modification brings at most minor changes to the upper bounds. In fact,

|(U(x2c)nU(x2)n)V2|LcR,x22C||U(x2)c||V2||LcR,x22C(μ2|V2|LcR,x22+|ω^0(k)|Lx22),\big{|}\big{(}U^{\prime}(x_{2}^{c})^{-n}-U^{\prime}(x_{2})^{-n}\big{)}V_{2}^{\prime\prime}\big{|}_{L_{c_{R},x_{2}}^{2}}\leq C\big{|}|U(x_{2})-c||V_{2}^{\prime\prime}|\big{|}_{L_{c_{R},x_{2}}^{2}}\leq C\big{(}\mu^{-2}|V_{2}|_{L_{c_{R},x_{2}}^{2}}+|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)},

for n=1,2n=1,2, where the Rayleigh equation was also used. This error bound and the estimate on V2V_{2} are then used to obtain the desired inequality on cRV2\partial_{c_{R}}V_{2}^{\prime}. The term V2U(x2)2\frac{V_{2}^{\prime\prime}}{U^{\prime}(x_{2})^{2}} in V~2\tilde{V}_{2} is handled by the same argument. Similarly,

|(U(x2c)1U(0)1)V2(0)|LcR2C||U(0)c||V2(0)||LcR2C(μ2|V2(0)|LcR2+|ω^0(k,0)|),\big{|}\big{(}U^{\prime}(x_{2}^{c})^{-1}-U^{\prime}(0)^{-1}\big{)}V_{2}^{\prime\prime}(0)\big{|}_{L_{c_{R}}^{2}}\leq C\big{|}|U(0)-c||V_{2}^{\prime\prime}(0)|\big{|}_{L_{c_{R}}^{2}}\leq C\big{(}\mu^{-2}|V_{2}(0)|_{L_{c_{R}}^{2}}+|\hat{\omega}_{0}(k,0)|\big{)},

and this along with the estimate on V2(0)V_{2}(0) yields the estimate on (cRV2+V2U)(0)\big{(}\partial_{c_{R}}V_{2}^{\prime}+\tfrac{V_{2}^{\prime\prime}}{U^{\prime}}\big{)}(0). It remain the consider the modifications to the correction terms in V~2\tilde{V}_{2} at x2=hx_{2}=-h and x2=0x_{2}=0. Similarly,

|(U(x2c)2U(0)2)V2(0)y(x2)|LcR,x22C(μ2|V2(0)|LcR2+|ω^0(k,0)|)|y|LcRLx22\displaystyle\big{|}\big{(}U^{\prime}(x_{2}^{c})^{-2}-U^{\prime}(0)^{-2}\big{)}V_{2}^{\prime\prime}(0)y_{-}(x_{2})\big{|}_{L_{c_{R},x_{2}}^{2}}\leq C\big{(}\mu^{-2}|V_{2}(0)|_{L_{c_{R}}^{2}}+|\hat{\omega}_{0}(k,0)|\big{)}|y_{-}|_{L_{c_{R}}^{\infty}L_{x_{2}}^{2}}

which is controlled using |y|LcRLx22Cμ32ehμ|y_{-}|_{L_{c_{R}}^{\infty}L_{x_{2}}^{2}}\leq C\mu^{\frac{3}{2}}e^{\frac{h}{\mu}} due to lemma 3.9. The last remaining modification from U(x2c)2U^{\prime}(x_{2}^{c})^{-2} to U(h)2U^{\prime}(-h)^{-2} can be justified by the same argument (even easier as V2(h)=0V_{2}(-h)=0.)

Finally we consider cRη~\partial_{c_{R}}\tilde{\eta} for in c𝒟ρ0,ρ0c\in\mathcal{D}_{\rho_{0},\rho_{0}}. From (6.5), one may compute

cRη~=1ik(U(0)(U(0)c)+g+σk2)2(cRV2(0)(U(0)c)(U(0)(U(0)c)+g+σk2)\displaystyle\partial_{c_{R}}\tilde{\eta}=\tfrac{1}{ik}\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}^{-2}\Big{(}\partial_{c_{R}}V_{2}^{\prime}(0)(U(0)-c)\big{(}U^{\prime}(0)(U(0)-c)+g+\sigma k^{2}\big{)}
(g+σk2)V2(0)+U(0)(U(0)η^0(k)+ikv^20(k,0)))\displaystyle-(g+\sigma k^{2})V_{2}^{\prime}(0)+U^{\prime}(0)\big{(}U^{\prime}(0)\hat{\eta}_{0}(k)+\tfrac{i}{k}\hat{v}_{20}^{\prime}(k,0)\big{)}\Big{)}
=\displaystyle= 1ik(U(Uc)+g+σk2)2((cRV2+V2U1U(k2V2+UV2+ω^0Uc))(Uc)(U(Uc)+g+σk2)\displaystyle\tfrac{1}{ik}\big{(}U^{\prime}(U-c)+g+\sigma k^{2}\big{)}^{-2}\Big{(}\big{(}\partial_{c_{R}}V_{2}^{\prime}+\tfrac{V_{2}^{\prime\prime}}{U^{\prime}}-\tfrac{1}{U^{\prime}}\big{(}k^{2}V_{2}+\tfrac{U^{\prime\prime}V_{2}+\hat{\omega}_{0}}{U-c}\big{)}\big{)}(U-c)\big{(}U^{\prime}(U-c)+g+\sigma k^{2}\big{)}
(g+σk2)V2+U(Uη^0(k)+ikv^20(k,0)))|x2=0\displaystyle-(g+\sigma k^{2})V_{2}^{\prime}+U^{\prime}\big{(}U^{\prime}\hat{\eta}_{0}(k)+\tfrac{i}{k}\hat{v}_{20}^{\prime}(k,0)\big{)}\Big{)}\Big{|}_{x_{2}=0}

where we used the Rayleigh equation (5.1a) in the last step. Therefore from Proposition 5.6 we have, for any kk\in\mathbb{R} and cI[0,r2]c_{I}\in[0,r_{2}],

|cRη~|LcR2C(|k|1μ4|η^0(k)|+|k|2μ4|v^20(k,0)|+|k|1μ2|V2(0)|LcR2+|k|1μ2|(cRV2+V2U)(0)|LcR2\displaystyle|\partial_{c_{R}}\tilde{\eta}|_{L_{c_{R}}^{2}}\leq C\big{(}|k|^{-1}\mu^{4}|\hat{\eta}_{0}(k)|+|k|^{-2}\mu^{4}|\hat{v}_{20}^{\prime}(k,0)|+|k|^{-1}\mu^{2}|V_{2}^{\prime}(0)|_{L_{c_{R}}^{2}}+|k|^{-1}\mu^{2}\big{|}\big{(}\partial_{c_{R}}V_{2}^{\prime}+\tfrac{V_{2}^{\prime\prime}}{U^{\prime}}\big{)}(0)\big{|}_{L_{c_{R}}^{2}}
+|k|1μ2|(k2(U(0)c)+U(0))V2(0)+ω^(k,0)|LcR2)\displaystyle\qquad\qquad\quad+|k|^{-1}\mu^{2}|(k^{2}(U(0)-c)+U^{\prime\prime}(0))V_{2}(0)+\hat{\omega}(k,0)|_{L_{c_{R}}^{2}}\big{)}
\displaystyle\leq C(|k|1|η^0(k)|+|k|2μ2|v^20(k,0)|+|k|1μ1ϵ|ω^0(k)|Lx22+|k|1μ2ϵ|ω^0(k)|Lx22+|k|1μ2|ω^0(k,0)|).\displaystyle C\big{(}|k|^{-1}|\hat{\eta}_{0}(k)|+|k|^{-2}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|k|^{-1}\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+|k|^{-1}\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}+|k|^{-1}\mu^{2}|\hat{\omega}_{0}(k,0)|\big{)}.

The last terms can be controlled by the previous two terms, which completes the estimate on cRη~\partial_{c_{R}}\tilde{\eta}. The convergence of cRη\partial_{c_{R}}\eta also follows from those of V2(0)V_{2}(0), V2(0)V_{2}^{\prime}(0) and (cRV2+1U(x2c)V2)(0)\big{(}\partial_{c_{R}}V_{2}^{\prime}+\tfrac{1}{U^{\prime}(x_{2}^{c})}V_{2}^{\prime\prime}\big{)}(0). ∎

The following lemma will be used in the decay estimates.

Lemma 6.5.

Suppose n0n\geq 0 is an integer, q[2,]q\in[2,\infty], f(c)f(c) and f1(c)f_{1}(c) are analytic functions on

𝒟0, where 𝒟=+i[ρ,ρ],0(b1,b2),=[b1,b2],ρ>0,\mathcal{D}\setminus\mathcal{I}_{0}\subset\mathbb{C},\;\text{ where }\;\mathcal{D}=\mathcal{I}+i[-\rho,\rho],\;\;\mathcal{I}_{0}\subset(b_{1},b_{2}),\;\;\mathcal{I}=[b_{1},b_{2}]\subset\mathbb{R},\;\;\rho>0,

and there exists M>0M>0 such that |(f(n)f1)(+icI)|Lqq1()M|(f^{(n)}-f_{1})(\cdot+ic_{I})|_{L^{\frac{q}{q-1}}(\mathcal{I})}\leq M for all 0<|cI|ρ0<|c_{I}|\leq\rho, then there exists C>0C>0 depending only on b2b1b_{2}-b_{1} such that, for any k0k\neq 0,

|𝒟eickt(tnf(c)(ik)nf1(c))dc|Ltq()C|k|n1qM.\Big{|}\oint_{\partial\mathcal{D}}e^{-ickt}\big{(}t^{n}f(c)-(ik)^{-n}f_{1}(c)\big{)}dc\Big{|}_{L_{t}^{q}(\mathbb{R})}\leq C|k|^{-n-\frac{1}{q}}M.
Proof.

Integrating by parts we have, for any 0<|r|ρ0<|r|\leq\rho,

+irtneicktf(c)dc=(ik)nerkteicRktf(n)(cR+ir)dcReicktl=1ntnl(ik)lf(l1)(c)|c=b1+irc=b2+ir.\int_{\mathcal{I}+ir}t^{n}e^{-ickt}f(c)dc=(ik)^{-n}e^{rkt}\int_{\mathcal{I}}e^{-ic_{R}kt}f^{(n)}(c_{R}+ir)dc_{R}-e^{-ickt}\sum_{l=1}^{n}t^{n-l}(ik)^{-l}f^{(l-1)}(c)\big{|}_{c=b_{1}+ir}^{c=b_{2}+ir}.

For any T>0T>0, the Lqq1LqL^{\frac{q}{q-1}}\to L^{q} boundedness (for q[2,]q\in[2,\infty]) of the Fourier transform implies

|eicRkt(f(n)f1)(cR+ir)dcR|Ltq([T,T])C|k|1q|(f(n)f1)(+ir)|Lqq1()C|k|1qM.\Big{|}\int_{\mathcal{I}}e^{-ic_{R}kt}(f^{(n)}-f_{1})(c_{R}+ir)dc_{R}\Big{|}_{L_{t}^{q}([-T,T])}\leq C|k|^{-\frac{1}{q}}|(f^{(n)}-f_{1})(\cdot+ir)|_{L^{\frac{q}{q-1}}(\mathcal{I})}\leq C|k|^{-\frac{1}{q}}M.

From this inequality and the Cauchy integral theorem, we obtain, for any r(0,ρ]r\in(0,\rho],

|𝒟eickt(tnf(c)(ik)nf1(c))dc|Ltq([T,T])\displaystyle\Big{|}\oint_{\partial\mathcal{D}}e^{-ickt}\big{(}t^{n}f(c)-(ik)^{-n}f_{1}(c)\big{)}dc\Big{|}_{L_{t}^{q}([-T,T])}
=\displaystyle= |(+i[r,r])eickt(tnf(c)(ik)nf1(c))dc|Ltq([T,T])\displaystyle\Big{|}\oint_{\partial(\mathcal{I}+i[-r,r])}e^{-ickt}\big{(}t^{n}f(c)-(ik)^{-n}f_{1}(c)\big{)}dc\Big{|}_{L_{t}^{q}([-T,T])}
\displaystyle\leq C|k|n1qer|k|TM+|(b1+irb1ir+b2irb2+ir)eickt(tnf(c)(ik)nf1(c))dc\displaystyle C|k|^{-n-\frac{1}{q}}e^{r|k|T}M+\Big{|}\Big{(}\int_{b_{1}+ir}^{b_{1}-ir}+\int_{b_{2}-ir}^{b_{2}+ir}\Big{)}e^{-ickt}\big{(}t^{n}f(c)-(ik)^{-n}f_{1}(c)\big{)}dc
+l=1ntnl(ik)l(eicktf(l1)(c)|c=b1+irc=b2+ireicktf(l1)(c)|c=b1irc=b2ir)|Ltq([T,T]).\displaystyle+\sum_{l=1}^{n}t^{n-l}(ik)^{-l}\big{(}e^{-ickt}f^{(l-1)}(c)\big{|}_{c=b_{1}+ir}^{c=b_{2}+ir}-e^{-ickt}f^{(l-1)}(c)\big{|}_{c=b_{1}-ir}^{c=b_{2}-ir}\big{)}\Big{|}_{L_{t}^{q}([-T,T])}.

Letting r0r\to 0, the analyticity assumption of ff and f1f_{1} implies all those terms on the vertical boundary of 𝒟\mathcal{D} vanish and the above estimates on the integrals along the horizontal edges yield

|𝒟tneickt(tnf(c)(ik)nf1(c))dc|Ltq([T,T])C|k|n1qM.\Big{|}\oint_{\partial\mathcal{D}}t^{n}e^{-ickt}\big{(}t^{n}f(c)-(ik)^{-n}f_{1}(c)\big{)}dc\Big{|}_{L_{t}^{q}([-T,T])}\leq C|k|^{-n-\frac{1}{q}}M.

The lemma follows by letting T+T\to+\infty. ∎

Remark 6.2.

In the following applications of this lemma, we often use the L2L^{2} norm to control the Lqq1L^{\frac{q}{q-1}} norm. This leads to fact that the regularity requirements in x1x_{1} (i.e. the exponents of kk) may not be close to optimal.

Applying the above lemma, we first obtain the decay of v^c(t,k,x2)\hat{v}^{c}(t,k,x_{2}) and η^c(t,k)\hat{\eta}^{c}(t,k).

Lemma 6.6.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, and (5.8), then for any ϵ(0,1)\epsilon\in(0,1), q[2,]q\in[2,\infty], and integer m0m\geq 0, there exists C>0C>0 determined only by ϵ\epsilon, qq, mm, F0F_{0}, ρ0\rho_{0}, |U|Cl01|U^{\prime}|_{C^{l_{0}-1}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} (independent of k0k\neq 0) such that

|tmv^2c(k)|Lx22Ltq()\displaystyle|\partial_{t}^{m}\hat{v}_{2}^{c}(k)|_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})} +|k|μ|tmv^1c(k)|Lx22Ltq()+μ|tm(v^2c)(k)|Lx22Ltq()\displaystyle+|k|\mu|\partial_{t}^{m}\hat{v}_{1}^{c}(k)|_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}+\mu|\partial_{t}^{m}(\hat{v}_{2}^{c})^{\prime}(k)|_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
\displaystyle\leq C|k|m+11q(μ12|η^0(k)|+|k|1μ52|v^20(k,0)|+μ32ϵ|ω^0(k)|Lx22),\displaystyle C|k|^{m+1-\frac{1}{q}}\big{(}\mu^{\frac{1}{2}}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{\frac{5}{2}}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{\frac{3}{2}-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)},
|tmη^c(k)|Ltq()C|k|m11q(|k|μ|η^0(k)|+μ2|v^20(k,0)|+|k|μ2ϵ|ω^0(k)|Lx22);|\partial_{t}^{m}\hat{\eta}^{c}(k)|_{L_{t}^{q}(\mathbb{R})}\leq C|k|^{m-1-\frac{1}{q}}\big{(}|k|\mu|\hat{\eta}_{0}(k)|+\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|k|\mu^{2-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)};

and if l04l_{0}\geq 4,

|ttmv^2c(k)|Lx22Ltq()C|k|m1q(μ12|η^0(k)|+|k|1μ32|v^20(k,0)|+μ12ϵ|ω^0(k)|Lx22+μ32ϵ|ω^0(k)|Lx22),|t\partial_{t}^{m}\hat{v}_{2}^{c}(k)|_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}\leq C|k|^{m-\frac{1}{q}}\big{(}\mu^{-\frac{1}{2}}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{\frac{3}{2}}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{\frac{1}{2}-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{\frac{3}{2}-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)},
|ttmη^c(k)|Ltq()C|k|m21q(|k||η^0(k)|+μ2|v^20(k,0)|+|k|μ1ϵ|ω^0(k)|Lx22+|k|μ2ϵ|ω^0(k)|Lx22).|t\partial_{t}^{m}\hat{\eta}^{c}(k)|_{L_{t}^{q}(\mathbb{R})}\leq C|k|^{m-2-\frac{1}{q}}\big{(}|k||\hat{\eta}_{0}(k)|+\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|k|\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+|k|\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.
Proof.

The estimates of tmv^2c\partial_{t}^{m}\hat{v}_{2}^{c}, ttmv^2ct\partial_{t}^{m}\hat{v}_{2}^{c}, tm(v^2c)\partial_{t}^{m}(\hat{v}_{2}^{c})^{\prime}, tmv^1c\partial_{t}^{m}\hat{v}_{1}^{c}, tmη^c\partial_{t}^{m}\hat{\eta}^{c}, and ttmη^ct\partial_{t}^{m}\hat{\eta}^{c} are based on the definitions of v^c(t,k,x2)\hat{v}^{c}(t,k,x_{2}) and η^c(t,k)\hat{\eta}^{c}(t,k) from direct application of Lemma 6.4 and Lemma 6.5 on 𝒟ρ0,ρ0\mathcal{D}_{\rho_{0},\rho_{0}} with f1=0f_{1}=0 and ff being cmV2c^{m}V_{2} (with n=0,1n=0,1), cmV2c^{m}V_{2}^{\prime}, cmη~c^{m}\tilde{\eta} (with n=0,1n=0,1), respectively. We omit the details. ∎

In the following we shall focus on ttmv^1c(t,k,x2)t\partial_{t}^{m}\hat{v}_{1}^{c}(t,k,x_{2}), t2tmv^2c(t,k,x2)t^{2}\partial_{t}^{m}\hat{v}_{2}^{c}(t,k,x_{2}), and tmω^c(t,k,x2)\partial_{t}^{m}\hat{\omega}^{c}(t,k,x_{2}), where ω^c\hat{\omega}^{c} is the Fourier transform (in x1x_{1}) of the vorticity ωc=x1v2cx2v1c\omega^{c}=\partial_{x_{1}}v_{2}^{c}-\partial_{x_{2}}v_{1}^{c} of vc(t,x)v^{c}(t,x). In order to characterize their asymptotic behavior, define

(6.9) Ω^c(k,x2)=ω^0(k,x2)+12U(x2)((1+sgn(kt))V20(k,U(x2),x2)+(1sgn(kt))V20(k,U(x2),x2)¯).\begin{split}\hat{\Omega}^{c}(k,x_{2})=&\hat{\omega}_{0}(k,x_{2})\\ &+\tfrac{1}{2}U^{\prime\prime}(x_{2})\big{(}(1+sgn(kt))V_{20}(k,U(x_{2}),x_{2})+(1-sgn(kt))\overline{V_{20}(-k,U(x_{2}),x_{2})}\big{)}.\end{split}

In the above expression, exactly one of 1+sgn(kt)1+sgn(kt) and 1sgn(kt)1-sgn(kt) is equal to 22 and the other equal to 0. The dependence of Ω^c\hat{\Omega}^{c} on tt is only through its sign, so we skipped specifying the tt dependence. We also notice that V2V_{2} may not be C0C^{0} at cU([h,0])c\in U([-h,0])\subset\mathbb{C}. The available conjugacy properties of V2V_{2} are not sufficient to imply V20(k,U(x2),x2)¯=V20(k,U(x2),x2)\overline{V_{20}(-k,U(x_{2}),x_{2})}=V_{20}(k,U(x_{2}),x_{2}). We shall see that Ω^c\hat{\Omega}^{c} provides the asymptotic profile of the vorticity ω^c\hat{\omega}^{c}. We first give the following some basic properties of Ω^c\hat{\Omega}^{c}.

Lemma 6.7.

Assume UC4U\in C^{4} and (5.8), then Ω^c(k,x2)=Ω^c(k,x2)¯\hat{\Omega}^{c}(-k,x_{2})=\overline{\hat{\Omega}^{c}(k,x_{2})} and, for any ϵ(0,1)\epsilon\in(0,1), there exists C>0C>0 determined only by ϵ\epsilon, F0F_{0}, ρ0\rho_{0}, |U|C3|U^{\prime}|_{C^{3}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} (independent of k0k\neq 0) such that

|Ω^cω^0|Lx22C(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22),|\hat{\Omega}^{c}-\hat{\omega}_{0}|_{L_{x_{2}}^{2}}\leq C\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}),
|(Ω^c)ω^0|Lx22Cμ1(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22).|(\hat{\Omega}^{c})^{\prime}-\hat{\omega}_{0}^{\prime}|_{L_{x_{2}}^{2}}\leq C\mu^{-1}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.
Proof.

The conjugacy relation of Ω^c\hat{\Omega}^{c} is clear from its definition. According to Lemma 6.4, V20V_{20} satisfies the same estimates as V2V_{2} for |cI|(0,ρ0]|c_{I}|\in(0,\rho_{0}]. We have, for x2[h,0]x_{2}\in[-h,0] and cU([h,0])c\in U([-h,0]),

(6.10) |V20(k,U(),)|Lx22C|V20|LcR,x2212|V20|LcR2Hx2112C(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22),\begin{split}|V_{20}(k,U(\cdot),\cdot)|_{L_{x_{2}}^{2}}\leq&C|V_{20}|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}}|V_{20}|_{L_{c_{R}}^{2}H_{x_{2}}^{1}}^{\frac{1}{2}}\\ \leq&C\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}),\end{split}

which implies the estimate of Ω^c\hat{\Omega}^{c}. Apparently the estimate of (Ω^c)(\hat{\Omega}^{c})^{\prime} depends on that of

x2(V20(k,U(x2),x2))=(DcV20)(k,U(x2),x2),\partial_{x_{2}}\big{(}V_{20}(k,U(x_{2}),x_{2})\big{)}=(D_{c}V_{20})(k,U(x_{2}),x_{2}),

where DcD_{c} was defined in (5.22). From (5.25) and (6.3), we have

|x2(V20(k,U(),))|Lx22\displaystyle\big{|}\partial_{x_{2}}\big{(}V_{20}(k,U(\cdot),\cdot)\big{)}\big{|}_{L_{x_{2}}^{2}}\leq C|DcV20|LcR,x2212|DcV20|LcR2Hx2112\displaystyle C|D_{c}V_{20}|_{L_{c_{R},x_{2}}^{2}}^{\frac{1}{2}}|D_{c}V_{20}|_{L_{c_{R}}^{2}H_{x_{2}}^{1}}^{\frac{1}{2}}
\displaystyle\leq Cμ1(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22),\displaystyle C\mu^{-1}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)},

which yields and completes the proof of the lemma. ∎

In the following lemma, we obtain the leading order terms of tv^1ct\hat{v}_{1}^{c}, (v^2c)(\hat{v}_{2}^{c})^{\prime\prime}, and ω^c\hat{\omega}^{c}.

Lemma 6.8.

Assume UC4U\in C^{4} and (5.8), then, for any ϵ(0,1)\epsilon\in(0,1), q(2,]q\in(2,\infty], and integer m0m\geq 0, there exists C>0C>0 determined only by ϵ\epsilon, qq, mm, F0F_{0}, ρ0\rho_{0}, |U|C3|U^{\prime}|_{C^{3}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} (independent of k0k\neq 0) such that

k2|tm(tv^1c(t,k,x2)+ik1U(x2)1eikU(x2)tΩ^c(k,x2))|Lx22Ltq()\displaystyle k^{2}\big{|}\partial_{t}^{m}\big{(}t\hat{v}_{1}^{c}(t,k,x_{2})+ik^{-1}U^{\prime}(x_{2})^{-1}e^{-ikU(x_{2})t}\hat{\Omega}^{c}(k,x_{2})\big{)}\big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
+|k||tm(ω^c(t,k,x2)eikU(x2)tΩ^c(k,x2))|Lx22Ltq()\displaystyle+|k|\big{|}\partial_{t}^{m}\big{(}\hat{\omega}^{c}(t,k,x_{2})-e^{-ikU(x_{2})t}\hat{\Omega}^{c}(k,x_{2})\big{)}\big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
+|tm((v^2c)(t,k,x2)ikeikU(x2)tΩ^c(k,x2))|Lx22Ltq()\displaystyle+\big{|}\partial_{t}^{m}\big{(}(\hat{v}_{2}^{c})^{\prime\prime}(t,k,x_{2})-ike^{-ikU(x_{2})t}\hat{\Omega}^{c}(k,x_{2})\big{)}\big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
\displaystyle\leq C|k|m+11qμ32(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22).\displaystyle C|k|^{m+1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.
Remark 6.3.

This lemma also implies, for any integer m1m\geq 1,

|ttm(eikU(x2)tv^1c(t,k,x2))|Lx22Ltq()C|k|m11qμ32(\displaystyle|t\partial_{t}^{m}\big{(}e^{ikU(x_{2})t}\hat{v}_{1}^{c}(t,k,x_{2})\big{)}\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}\leq C|k|^{m-1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(} |η^0(k)|+|k|1μ2|v^20(k,0)|\displaystyle|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|
+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22),\displaystyle+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)},

while for m=0m=0, there is another term ik1U(x2)1eikU(x2)tΩ^c(k,x2)ik^{-1}U^{\prime}(x_{2})^{-1}e^{-ikU(x_{2})t}\hat{\Omega}^{c}(k,x_{2}) on the left side. The form in the lemma is more consistent with other estimates including that of t2v^2ct^{2}\hat{v}_{2}^{c} to be given in the following, however.

Proof.

The definition of v^c\hat{v}^{c} implies, for each x2[h,0]x_{2}\in[-h,0] and r1,r2(0,ρ0]r_{1},r_{2}\in(0,\rho_{0}],

ttmv^1c(t,k,x2)=i(ik)m2π𝒟r1,r2teikctcmV2(k,c,x2)dc,\displaystyle t\partial_{t}^{m}\hat{v}_{1}^{c}(t,k,x_{2})=\frac{-i(-ik)^{m}}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}te^{-ikct}c^{m}V_{2}^{\prime}(k,c,x_{2})dc,
tm((ikω^c+k2v^2c)(t,k,x2))=tm(v^2c)(t,k,x2)=k(ik)m2π𝒟r1,r2eikctcmV2(k,c,x2)dc.\displaystyle\partial_{t}^{m}\big{(}(ik\hat{\omega}^{c}+k^{2}\hat{v}_{2}^{c})(t,k,x_{2})\big{)}=\partial_{t}^{m}(\hat{v}_{2}^{c})^{\prime\prime}(t,k,x_{2})=\frac{-k(-ik)^{m}}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}e^{-ikct}c^{m}V_{2}^{\prime\prime}(k,c,x_{2})dc.

Applying Lemma 6.5 with n=1n=1 and f=cmV2f=c^{m}V_{2}^{\prime} and f1=cmU(x2)V2f_{1}=-\frac{c^{m}}{U^{\prime}(x_{2})}V_{2}^{\prime\prime} and Lemma 6.4, we obtain

(6.11) |ttmv^1c(t,k,x2)𝒟r1,r2(i)mkm1cm2πU(x2)eikctV2(k,c,x2)dc|Lx22Ltq()C|k|m11qsup|cI|(0,r2](|cRV2+U(x2)1V2|LcR,x22+|V2|LcR,x22)C|k|m11qμ32(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22).\begin{split}&\Big{|}t\partial_{t}^{m}\hat{v}_{1}^{c}(t,k,x_{2})-\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}\frac{(-i)^{m}k^{m-1}c^{m}}{2\pi U^{\prime}(x_{2})}e^{-ikct}V_{2}^{\prime\prime}(k,c,x_{2})dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}\\ \leq&C|k|^{m-1-\frac{1}{q}}\sup_{|c_{I}|\in(0,r_{2}]}\big{(}\big{|}\partial_{c_{R}}V_{2}^{\prime}+U^{\prime}(x_{2})^{-1}V_{2}^{\prime\prime}\big{|}_{L_{c_{R},x_{2}}^{2}}+|V_{2}^{\prime}|_{L_{c_{R},x_{2}}^{2}}\big{)}\\ \leq&C|k|^{m-1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.\end{split}

In the rest of the proof, we shall focus on the integral involving V2V_{2}^{\prime\prime} which also yields the other desired estimates. Substituting the term V2V_{2}^{\prime\prime} by the Rayleigh equation (2.11a) and applying the Cauchy Integral Theorem yield

𝒟r1,r2cmeikctU(x2)V2dc=𝒟r1,r2cmeikctU(x2)(k2V2+U(x2)V2+ω^0(k,x2)U(x2)c)dc\displaystyle\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}\frac{c^{m}e^{-ikct}}{U^{\prime}(x_{2})}V_{2}^{\prime\prime}dc=\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}\frac{c^{m}e^{-ikct}}{U^{\prime}(x_{2})}\big{(}k^{2}V_{2}+\frac{U^{\prime\prime}(x_{2})V_{2}+\hat{\omega}_{0}(k,x_{2})}{U(x_{2})-c}\big{)}dc
=\displaystyle= 𝒟r1,r2eikctU(x2)(k2cm+(U(x2)mU(x2)c+cmU(x2)mU(x2)c)U(x2))V2dc2πiU(x2)mω^0(k,x2)U(x2)eikU(x2)t\displaystyle\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}\frac{e^{-ikct}}{U^{\prime}(x_{2})}\Big{(}k^{2}c^{m}+\big{(}\frac{U(x_{2})^{m}}{U(x_{2})-c}+\frac{c^{m}-U(x_{2})^{m}}{U(x_{2})-c}\big{)}U^{\prime\prime}(x_{2})\Big{)}V_{2}dc-\frac{2\pi iU(x_{2})^{m}\hat{\omega}_{0}(k,x_{2})}{U^{\prime}(x_{2})}e^{-ikU(x_{2})t}

Since k2cm+cmUmUcUk^{2}c^{m}+\frac{c^{m}-U^{m}}{U-c}U^{\prime\prime} is bounded by Cμ2C\mu^{-2} on 𝒟r1,r2\mathcal{D}_{r_{1},r_{2}}, we can control those terms using Lemma 6.5 and obtain

(6.12) |𝒟r1,r2(i)mkm1cm2πU(x2)eikctV2(k,c,x2)dc+(ik)m1U(x2)mU(x2)eikU(x2)t(ω^0(k,x2)iU(x2)2π𝒟r1,r2x21ceikctV2(k,c+U(x2),x2)dc)|Lx22Ltq()C|k|m11qμ32(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22),\begin{split}&\Big{|}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}\frac{(-i)^{m}k^{m-1}c^{m}}{2\pi U^{\prime}(x_{2})}e^{-ikct}V_{2}^{\prime\prime}(k,c,x_{2})dc+\frac{(-ik)^{m-1}U(x_{2})^{m}}{U^{\prime}(x_{2})}e^{-ikU(x_{2})t}\Big{(}\hat{\omega}_{0}(k,x_{2})\\ &\qquad\qquad\qquad\qquad\qquad\qquad-\frac{iU^{\prime\prime}(x_{2})}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}^{x_{2}}}\frac{1}{c}e^{-ikct}V_{2}(k,c+U(x_{2}),x_{2})dc\Big{)}\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}\\ \leq&C|k|^{m-1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)},\end{split}

where we also changed the variable cU(x2)cc-U(x_{2})\to c in the last integral and

(6.13) 𝒟r1,r2x2=𝒟r1,r2U(x2).\mathcal{D}_{r_{1},r_{2}}^{x_{2}}=\mathcal{D}_{r_{1},r_{2}}-U(x_{2}).

It remains to handle this integral term and we shall identify its leading terms.

Fix T>0T>0. We first let

w(k,c,x2)=V2(k,c+U(x2),x2)V20(k,cR+U(x2),x2)limcI0+|w(k,+icI,)|Lx2WcR1,q1=0,w(k,c,x_{2})=V_{2}(k,c+U(x_{2}),x_{2})-V_{20}(k,c_{R}+U(x_{2}),x_{2})\ \Longrightarrow\ \lim_{c_{I}\to 0+}|w(k,\cdot+ic_{I},\cdot)|_{L_{x_{2}}^{\infty}W_{c_{R}}^{1,q_{1}}}=0,

for any q1[1,2)q_{1}\in[1,2), where ω^0Hx21\hat{\omega}_{0}\in H_{x_{2}}^{1}, Lemma 6.4 was used. In the rest of the proof of this lemma, we use 𝒟r1,r2x2\partial^{\dagger}\mathcal{D}_{r_{1},r_{2}}^{x_{2}}, =L,R,T,B\dagger=L,R,T,B, to denote the left, right, top, bottom sides of the rectangle 𝒟r1,r2x2\mathcal{D}_{r_{1},r_{2}}^{x_{2}} with the counterclockwise orientation. For any r(0,r2]r\in(0,r_{2}] and k0k\neq 0, and 1qq1<q1<21\leq\frac{q}{q-1}<q_{1}<2, integrating by parts and using the Lqq1LqL^{\frac{q}{q-1}}\to L^{q} boundedness (for q[2,]q\in[2,\infty]) of the Fourier transform, we obtain

|T𝒟r1,rx21ceikctwdc|Lx22Ltq([T,T])=|(eikctwlogc)|U(0)U(x2)+r1+irU(h)U(x2)r1+ir\displaystyle\Big{|}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}\frac{1}{c}e^{-ikct}wdc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}=\Big{|}(e^{-ikct}w\log c)\big{|}_{U(0)-U(x_{2})+r_{1}+ir}^{U(-h)-U(x_{2})-r_{1}+ir}
ekrtT𝒟r1,rx2eikcRt(iktw+cRw)logcdc|Lx22Ltq([T,T])\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad-e^{krt}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}e^{-ikc_{R}t}(-iktw+\partial_{c_{R}}w)\log c\,dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}
\displaystyle\leq Ce|k|rT(T1q(1+|logr1|)|w(k,+ir,)|Lx22LcR+|k|1q(1+|k|T)|w(k,+ir,)|Lx22WcR1,q1),\displaystyle Ce^{|k|rT}\big{(}T^{\frac{1}{q}}(1+|\log r_{1}|)|w(k,\cdot+ir,\cdot)|_{L_{x_{2}}^{2}L_{c_{R}}^{\infty}}+|k|^{-\frac{1}{q}}(1+|k|T)|w(k,\cdot+ir,\cdot)|_{L_{x_{2}}^{2}W_{c_{R}}^{1,q_{1}}}\big{)},

where log\log is taken along T𝒟r1,rx2\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}} which in the upper half plane. Next from Lemma 6.4 we have

|T𝒟r1,rx21ceikct(V20(k,cR+U(x2),x2)V20(k,U(x2),x2))dc|Lx22Ltq([T,T])\displaystyle\Big{|}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}\frac{1}{c}e^{-ikct}\big{(}V_{20}(k,c_{R}+U(x_{2}),x_{2})-V_{20}(k,U(x_{2}),x_{2})\big{)}dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}
\displaystyle\leq C|k|1qe|k|rT|cRV20|Lx22LcRq1\displaystyle C|k|^{-\frac{1}{q}}e^{|k|rT}|\partial_{c_{R}}V_{20}|_{L_{x_{2}}^{2}L_{c_{R}}^{q_{1}}}
\displaystyle\leq C|k|1qμ12e|k|rT(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22).\displaystyle C|k|^{-\frac{1}{q}}\mu^{-\frac{1}{2}}e^{|k|rT}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.

The above error analysis implies that the main contribution of the integral along T𝒟r1,rx2\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}} would come from the product

V20(k,U(x2),x2)f(r,x2,kt), where f(r,x2,τ)=T𝒟r1,rx2eiτccdc.V_{20}(k,U(x_{2}),x_{2})f(r,x_{2},kt),\;\text{ where }\;f(r,x_{2},\tau)=\oint_{\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}\frac{e^{-i\tau c}}{c}dc.

For any r(0,r2]r\in(0,r_{2}], on the one hand,

|f(r,x2,τ)|=erτ|(eiτcRlogc)|U(h)U(x2)r1U(0)U(x2)+r1+iτT𝒟r1,rx2eiτcRlogcdc|C(1+|τ|)erτ,|f(r,x_{2},\tau)|=e^{r\tau}\big{|}-(e^{-i\tau c_{R}}\log c)\big{|}_{U(-h)-U(x_{2})-r_{1}}^{U(0)-U(x_{2})+r_{1}}+i\tau\oint_{\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}e^{-i\tau c_{R}}\log c\,dc\big{|}\leq C(1+|\tau|)e^{r\tau},

which is useful for |τ|1|\tau|\leq 1. On the other hand,

f(r,x2,τ)=(+ir(+ir)T𝒟r1,rx2)1ceiτcdc.f(r,x_{2},\tau)=-\Big{(}\int_{\mathbb{R}+ir}-\int_{(\mathbb{R}+ir)\setminus\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}\Big{)}\frac{1}{c}e^{-i\tau c}dc.

The first integral can be evaluated as iπ(sgn(τ)+1)i\pi(sgn(\tau)+1) by using the Cauchy Integral Theorem. Integrating the second integral (in the way opposite to the above) we obtain

|(+ir)T𝒟r1,rx21ceiτcdc|=erτ|τ||eiτcRc|U(0)U(x2)+r1U(h)U(x2)r1+(+ir)T𝒟r1,rx2eiτcRc2dc|Cerτ|τ|.\Big{|}\int_{(\mathbb{R}+ir)\setminus\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}\frac{1}{c}e^{-i\tau c}dc\Big{|}=\frac{e^{r\tau}}{|\tau|}\Big{|}\frac{e^{-i\tau c_{R}}}{c}\big{|}_{U(0)-U(x_{2})+r_{1}}^{U(-h)-U(x_{2})-r_{1}}+\int_{(\mathbb{R}+ir)\setminus\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}\frac{e^{-i\tau c_{R}}}{c^{2}}dc\Big{|}\leq C\frac{e^{r\tau}}{|\tau|}.

Therefore

|f(r,x2,τ)iπ(sgn(τ)+1)|C(1+|τ|)1er|τ|,τ.|f(r,x_{2},\tau)-i\pi(sgn(\tau)+1)|\leq C(1+|\tau|)^{-1}e^{r|\tau|},\quad\forall\tau\in\mathbb{R}.

Along with (6.10), we have

|V20(k,U(x2),x2)(T𝒟r1,rx21ceikctdciπ(sgn(kt)+1))|Lx22Ltq([T,T])\displaystyle\Big{|}V_{20}(k,U(x_{2}),x_{2})\Big{(}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}^{x_{2}}}\frac{1}{c}e^{-ikct}dc-i\pi(sgn(kt)+1)\Big{)}\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}
\displaystyle\leq C|k|1qer|k|T(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22).\displaystyle C|k|^{-\frac{1}{q}}e^{r|k|T}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)}.

The integrals along the vertical sides of 𝒟r1,rx2\partial\mathcal{D}_{r_{1},r}^{x_{2}} converge to 0 as r0+r\to 0+ as all the integrands are smooth there. The integrals along B𝒟r1,r\partial^{B}\mathcal{D}_{r_{1},r}, r(0,r2]r\in(0,r_{2}], can be treated much as in the above. Recall V2(k,c¯,x2)=V2(k,c,x2)¯V_{2}(k,\bar{c},x_{2})=\overline{V_{2}(-k,c,x_{2})}. Letting r0+r\to 0+, the Cauchy Integral Theorem and the above error analysis imply

|𝒟r1,r2x21ceikctV2(k,c+U(x2),x2)dc\displaystyle\Big{|}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}^{x_{2}}}\frac{1}{c}e^{-ikct}V_{2}(k,c+U(x_{2}),x_{2})dc
iπ((1+sgn(kt))V20(k,U(x2),x2)+(1sgn(kt))V20(k,U(x2),x2)¯)|Lx22Ltq([T,T])\displaystyle\qquad\quad-i\pi\big{(}(1+sgn(kt))V_{20}(k,U(x_{2}),x_{2})+(1-sgn(kt))\overline{V_{20}(-k,U(x_{2}),x_{2})}\big{)}\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}
\displaystyle\leq C|k|1qμ12(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22).\displaystyle C|k|^{-\frac{1}{q}}\mu^{-\frac{1}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.

Taking TT\to\infty, it follows from the above inequality and (6.12)

(6.14) |𝒟r1,r2(i)mkm1cm2πU(x2)eikctV2(k,c,x2)dc+(ik)m1U(x2)mU(x2)eikU(x2)tΩ^c(k,x2)|Lx22Ltq()C|k|m11qμ32(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22).\begin{split}&\Big{|}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}\frac{(-i)^{m}k^{m-1}c^{m}}{2\pi U^{\prime}(x_{2})}e^{-ikct}V_{2}^{\prime\prime}(k,c,x_{2})dc+\frac{(-ik)^{m-1}U(x_{2})^{m}}{U^{\prime}(x_{2})}e^{-ikU(x_{2})t}\hat{\Omega}^{c}(k,x_{2})\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}\\ \leq&C|k|^{m-1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.\end{split}

Along with (6.11) and Lemma 6.6 it implies the desired estimate of tm(tv^1c)\partial_{t}^{m}(t\hat{v}_{1}^{c}). The estimates on tm(v^2c)\partial_{t}^{m}(\hat{v}_{2}^{c})^{\prime\prime} and tmω^c\partial_{t}^{m}\hat{\omega}^{c} are also obtained from the above inequality and Lemma 6.6. ∎

Finally we consider t2v^2t^{2}\hat{v}_{2}.

Lemma 6.9.

Assume UC5U\in C^{5} and (5.8), then, for any ϵ(0,1)\epsilon\in(0,1), q(2,]q\in(2,\infty], and integer m0m\geq 0, there exists C>0C>0 determined only by ϵ\epsilon, qq, mm, F0F_{0}, |U|C4|U^{\prime}|_{C^{4}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}} (independent of k0k\neq 0) such that

|tm(t2v^2c(t,k,x2)(ieikU(x2)tkU(x2)2Ω^c(k,x2)+eikU(0)tΛ^T(k,x2)+eikU(h)tΛ^B(k,x2)))|Lx22Ltq()\displaystyle\Big{|}\partial_{t}^{m}\Big{(}t^{2}\hat{v}_{2}^{c}(t,k,x_{2})-\Big{(}-\frac{ie^{-ikU(x_{2})t}}{kU^{\prime}(x_{2})^{2}}\hat{\Omega}^{c}(k,x_{2})+e^{-ikU(0)t}\hat{\Lambda}_{T}(k,x_{2})+e^{-ikU(-h)t}\hat{\Lambda}_{B}(k,x_{2})\Big{)}\Big{)}\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
\displaystyle\leq C|k|m11qμ32(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22+μ3ϵ|ω^0(k)|Lx22),\displaystyle C|k|^{m-1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}+\mu^{3-\epsilon}|\hat{\omega}_{0}^{\prime\prime}(k)|_{L_{x_{2}}^{2}}\big{)},

where

(6.15) Λ^T(k,x2)=ikU(0)η^0(k)ω^0(k,0)U(0)2y0(k,U(0),0)y0(k,U(0),x2),\hat{\Lambda}_{T}(k,x_{2})=-\frac{i}{k}\frac{U^{\prime\prime}(0)\hat{\eta}_{0}(k)-\hat{\omega}_{0}(k,0)}{U^{\prime}(0)^{2}y_{0-}(k,U(0),0)}y_{0-}(k,U(0),x_{2}),
(6.16) Λ^B(k,x2)=iω^0(k,h)y0+(k,U(h),x2)kU(h)2y0+(k,U(h),h).\hat{\Lambda}_{B}(k,x_{2})=\frac{i\hat{\omega}_{0}(k,-h)y_{0+}(k,U(-h),x_{2})}{kU^{\prime}(-h)^{2}y_{0+}(k,U(-h),-h)}.
Remark 6.4.

In the above lemmas, we also notice Λ^(k,x2)=Λ^(k,x2)¯\hat{\Lambda}_{\dagger}(-k,x_{2})=\overline{\hat{\Lambda}_{\dagger}(k,x_{2})}, =T,B\dagger=T,B. The leading order terms Λ^B\hat{\Lambda}_{B} and Λ^T\hat{\Lambda}_{T} represent the contribution from the rigid bottom and the water surface, while the asymptotic vorticity Ω^c\hat{\Omega}^{c} from the fluid interior. In the fixed boundary problem for x2[h,0]x_{2}\in[-h,0] with slip boundary condition on both horizontal boundaries, Ω^c\hat{\Omega}^{c} and Λ^B\hat{\Lambda}_{B} would take similar forms and Λ^T\hat{\Lambda}_{T} would be similar to Λ^B\hat{\Lambda}_{B}. See Subsection 6.4.

Proof.

The definition of v^2c\hat{v}_{2}^{c} implies, for each x2[h,0]x_{2}\in[-h,0] and r1,r2(0,ρ0]r_{1},r_{2}\in(0,\rho_{0}],

t2tmv^2c(t,k,x2)=(i)mkm+12π𝒟r1,r2t2eikctcmV2(k,c,x2)dc.\displaystyle t^{2}\partial_{t}^{m}\hat{v}_{2}^{c}(t,k,x_{2})=\frac{-(-i)^{m}k^{m+1}}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}t^{2}e^{-ikct}c^{m}V_{2}(k,c,x_{2})dc.

Let f=cmV2f=c^{m}V_{2} and

f1=cm(V2(x2)U(x2)2g+σk2𝐅(k,c)(V2(h)U(h)2y+(x2)V2(0)U(0)2y(x2)))=cm(cR2V2V~2),f_{1}=c^{m}\Big{(}\frac{V_{2}^{\prime\prime}(x_{2})}{U^{\prime}(x_{2})^{2}}-\frac{g+\sigma k^{2}}{\mathbf{F}(k,c)}\Big{(}\frac{V_{2}^{\prime\prime}(-h)}{U^{\prime}(-h)^{2}}y_{+}(x_{2})-\frac{V_{2}^{\prime\prime}(0)}{U^{\prime}(0)^{2}}y_{-}(x_{2})\Big{)}\Big{)}=c^{m}(\partial_{c_{R}}^{2}V_{2}-\tilde{V}_{2}),

with V~2\tilde{V}_{2} defined in Lemma 6.4. Applying Lemma 6.5 with n=2n=2 and Lemma 6.4, we obtain

|t2tmv^2c(t,k,x2)(i)mkm12π𝒟r1,r2eikctf1(k,c,x2)dc|Lx22Ltq()\displaystyle\Big{|}t^{2}\partial_{t}^{m}\hat{v}_{2}^{c}(t,k,x_{2})-\frac{(-i)^{m}k^{m-1}}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}e^{-ikct}f_{1}(k,c,x_{2})dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
\displaystyle\leq C|k|m11qsupcI(0,r2](|V~2|LcR,x22+|cRV2|LcR,x22+|V2|LcR,x22)\displaystyle C|k|^{m-1-\frac{1}{q}}\sup_{c_{I}\in(0,r_{2}]}\big{(}\big{|}\tilde{V}_{2}|_{L_{c_{R},x_{2}}^{2}}+|\partial_{c_{R}}V_{2}|_{L_{c_{R},x_{2}}^{2}}+|V_{2}|_{L_{c_{R},x_{2}}^{2}}\big{)}
\displaystyle\leq C|k|m11qμ32(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22+μ3ϵ|ω^0(k)|Lx22).\displaystyle C|k|^{m-1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}+\mu^{3-\epsilon}|\hat{\omega}_{0}^{\prime\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.

Substituting V2V_{2}^{\prime\prime} in f1f_{1} by using the Rayleigh equation (2.11a) yields

f1=cm(V2(x2)U(x2)2+f1BU(h)c+f1TU(0)c+(g+σk2)y(x2)U(0)2𝐅(k,c)k2V2(0)),f_{1}=c^{m}\Big{(}\frac{V_{2}^{\prime\prime}(x_{2})}{U^{\prime}(x_{2})^{2}}+\frac{f_{1B}}{U(-h)-c}+\frac{f_{1T}}{U(0)-c}+\frac{(g+\sigma k^{2})y_{-}(x_{2})}{U^{\prime}(0)^{2}\mathbf{F}(k,c)}k^{2}V_{2}(0)\Big{)},

where

f1B=(g+σk2)ω^0(h)U(h)2𝐅(k,c)y+(x2)=ω^0(h)y+(x2)U(h)2y+(h),f_{1B}=-\frac{(g+\sigma k^{2})\hat{\omega}_{0}(-h)}{U^{\prime}(-h)^{2}\mathbf{F}(k,c)}y_{+}(x_{2})=-\frac{\hat{\omega}_{0}(-h)y_{+}(x_{2})}{U^{\prime}(-h)^{2}y_{+}(-h)},
f1T=(g+σk2)(U(0)V2(0)+ω^0(0))U(0)2𝐅(k,c)y(x2).f_{1T}=\frac{(g+\sigma k^{2})\big{(}U^{\prime\prime}(0)V_{2}(0)+\hat{\omega}_{0}(0)\big{)}}{U^{\prime}(0)^{2}\mathbf{F}(k,c)}y_{-}(x_{2}).

Again the terms involving k2V2(0)k^{2}V_{2}(0) not being divided by UcU-c can be estimated by using assumption (5.8) and Lemmas 6.5, 3.9, and 6.4 and we have

|t2tmv^2c(t,k,x2)(i)mkm12π𝒟r1,r2eikctcm(V2(x2)U(x2)2+f1BU(h)c+f1TU(0)c)dc|Lx22Ltq()\displaystyle\Big{|}t^{2}\partial_{t}^{m}\hat{v}_{2}^{c}(t,k,x_{2})-\frac{(-i)^{m}k^{m-1}}{2\pi}\oint_{\partial\mathcal{D}_{r_{1},r_{2}}}e^{-ikct}c^{m}\Big{(}\frac{V_{2}^{\prime\prime}(x_{2})}{U^{\prime}(x_{2})^{2}}+\frac{f_{1B}}{U(-h)-c}+\frac{f_{1T}}{U(0)-c}\Big{)}dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
\displaystyle\leq C|k|m+11qμ12(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22+μ3ϵ|ω^0(k)|Lx22).\displaystyle C|k|^{m+1-\frac{1}{q}}\mu^{\frac{1}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}+\mu^{3-\epsilon}|\hat{\omega}_{0}^{\prime\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.

We shall identify the principle contributions from the terms V2(x2)V_{2}^{\prime\prime}(x_{2}), f1Bf_{1B}, and f1Tf_{1T} following a similar strategy and use the same notations 𝒟r1,r2\partial^{\dagger}\mathcal{D}_{r_{1},r_{2}}, =T,B,L,R\dagger=T,B,L,R, as in the proof of Lemma 6.8, with necessary modifications to treat the contributions from the x2=0,hx_{2}=0,-h.

Fix T>0T>0. We start with f1Tf_{1T} by letting

f1T0(k,cR,x2)=limcI0+f1T(k,c,x2)=(g+σk2)(U(0)V20(0)+ω^0(0))U(0)2𝐅(k,cR)y0(x2).f_{1T}^{0}(k,c_{R},x_{2})=\lim_{c_{I}\to 0+}f_{1T}(k,c,x_{2})=\frac{(g+\sigma k^{2})\big{(}U^{\prime\prime}(0)V_{20}(0)+\hat{\omega}_{0}(0)\big{)}}{U^{\prime}(0)^{2}\mathbf{F}(k,c_{R})}y_{0-}(x_{2}).

From assumption (5.8), Lemmas 3.12, 3.18(2b), 4.2, and 6.4, we have, for any q1[1,2)q_{1}\in[1,2),

|(f1Tf1T0)(k,+cI,)|Lx2WcR1,q10, as cI0+.|(f_{1T}-f_{1T}^{0})(k,\cdot+c_{I},\cdot)|_{L_{x_{2}}^{\infty}W_{c_{R}}^{1,q_{1}}}\to 0,\;\text{ as }c_{I}\to 0+.

The next step is the same argument via integrating by parts in cRc_{R} as in the proof of Lemma 6.8,

|T𝒟r1,reikctcmf1T(k,c,x2)f1T0(k,cR,x2)U(0)cdc|Lx22Ltq([T,T])\displaystyle\Big{|}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}}e^{-ikct}c^{m}\frac{f_{1T}(k,c,x_{2})-f_{1T}^{0}(k,c_{R},x_{2})}{U(0)-c}dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}
=\displaystyle= |(eikctcm(f1Tf1T0)log(U(0)c))|U(0)+r1+irU(h)r1+ir\displaystyle\Big{|}\big{(}e^{-ikct}c^{m}(f_{1T}-f_{1T}^{0})\log(U(0)-c)\big{)}\big{|}_{U(0)+r_{1}+ir}^{U(-h)-r_{1}+ir}
T𝒟r1,reikct(ikt+cR)(cm(f1Tf1T0))log(U(0)c)dc|Lx22Ltq([T,T])0 as r0+.\displaystyle\quad\;-\oint_{\partial^{T}\mathcal{D}_{r_{1},r}}e^{-ikct}(-ikt+\partial_{c_{R}})\big{(}c^{m}(f_{1T}-f_{1T}^{0})\big{)}\log(U(0)-c)dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}\to 0\;\text{ as }r\to 0+.

From (5.8) and Lemmas 3.9, 3.143.16, 4.1(3), and 4.2(1), one may estimate,

|y0/𝐅|Lx22LcR+|cR(y0/𝐅)|Lx22LcRq1Cμ52,q1[1,).|y_{0-}/\mathbf{F}|_{L_{x_{2}}^{2}L_{c_{R}}^{\infty}}+|\partial_{c_{R}}(y_{0-}/\mathbf{F})|_{L_{x_{2}}^{2}L_{c_{R}}^{q_{1}}}\leq C\mu^{\frac{5}{2}},\;\;\forall q_{1}\in[1,\infty).

Along with Lemma 6.4, it implies, for any q2[1,2)q_{2}\in[1,2),

|f1T0|Lx22LcR+|cRf1T0|Lx22LcRq2Cμ12(|η^0(k)|+|k|1μ2|v^20(k,0)|+|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22|),|f_{1T}^{0}|_{L_{x_{2}}^{2}L_{c_{R}}^{\infty}}+|\partial_{c_{R}}f_{1T}^{0}|_{L_{x_{2}}^{2}L_{c_{R}}^{q_{2}}}\leq C\mu^{-\frac{1}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}|\big{)},

where |ω^0(0)||\hat{\omega}_{0}(0)| and |V2(0)|LcR|V_{2}(0)|_{L_{c_{R}}^{\infty}} were bounded by the L2L^{2} norms of ω^0(k)\hat{\omega}_{0}(k), ω^0(k)\hat{\omega}_{0}^{\prime}(k), |V2(0)|LcR2|V_{2}(0)|_{L_{c_{R}}^{2}}, and |cRV2(0)|LcR2|\partial_{c_{R}}V_{2}(0)|_{L_{c_{R}}^{2}}. Consequently, for any r(0,r2]r\in(0,r_{2}]

|T𝒟r1,reikctcmU(0)cf1T0(k,cR,x2)dcf1T0(k,U(0),x2)T𝒟r1,reikctcmU(0)cdc|Lx22Ltq([T,T])\displaystyle\Big{|}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}}\frac{e^{-ikct}c^{m}}{U(0)-c}f_{1T}^{0}(k,c_{R},x_{2})dc-f_{1T}^{0}(k,U(0),x_{2})\oint_{\partial^{T}\mathcal{D}_{r_{1},r}}\frac{e^{-ikct}c^{m}}{U(0)-c}dc\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}
\displaystyle\leq Cer|k|T|k|1qμ12(|η^0(k)|+|k|1μ2|v^20(k,0)|+|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22).\displaystyle Ce^{r|k|T}|k|^{-\frac{1}{q}}\mu^{-\frac{1}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.

As in the proof of Lemma 6.8, by considering contour integrals, we have

|T𝒟r1,reikctU(0)cdc+iπ(1+sgn(kt))eikU(0)t|Cer|kt|1+|kt|.\Big{|}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}}\frac{e^{-ikct}}{U(0)-c}dc+i\pi(1+sgn(kt))e^{-ikU(0)t}\Big{|}\leq\frac{Ce^{r|kt|}}{1+|kt|}.

Again, since cmU(0)mcU(0)\frac{c^{m}-U(0)^{m}}{c-U(0)} is bounded for m1m\geq 1, the above |f1T0|Lx22LcR|f_{1T}^{0}|_{L_{x_{2}}^{2}L_{c_{R}}^{\infty}} estimate implies

|f1T0(k,U(0),x2)(T𝒟r1,reikctcmU(0)cdc+iπ(1+sgn(kt))eikU(0)tU(0)m)|Lx22Ltq([T,T])\displaystyle\Big{|}f_{1T}^{0}(k,U(0),x_{2})\Big{(}\oint_{\partial^{T}\mathcal{D}_{r_{1},r}}\frac{e^{-ikct}c^{m}}{U(0)-c}dc+i\pi(1+sgn(kt))e^{-ikU(0)t}U(0)^{m}\Big{)}\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}([-T,T])}
\displaystyle\leq Cer|k|T|k|1qμ12(|η^0(k)|+|k|1μ2|v^20(k,0)|+|ω^0(k)|Lx22+μ2|ω^0(k)|Lx22).\displaystyle Ce^{r|k|T}|k|^{-\frac{1}{q}}\mu^{-\frac{1}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.

The contributions from the integral along B𝒟r1,r2\partial^{B}\mathcal{D}_{r_{1},r_{2}} can be treated similarly and using the conjugacy relation, while the integrals along the vertical boundaries of 𝒟r1,r2\partial\mathcal{D}_{r_{1},r_{2}} vanish as r0+r\to 0+. Using the Cauchy Integral Theorem, combining the above analysis, letting r0+r\to 0+, and then T0+T\to 0+, we obtain

|T𝒟r1,r2eikctcmf1T(k,c,x2)dc2πkeikU(0)tU(0)mΛ^T(k,x2)|Lx22Ltq()\displaystyle\Big{|}\oint_{\partial^{T}\mathcal{D}_{r_{1},r_{2}}}e^{-ikct}c^{m}f_{1T}(k,c,x_{2})dc-2\pi ke^{-ikU(0)t}U(0)^{m}\hat{\Lambda}_{T}(k,x_{2})\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
\displaystyle\leq C|k|1qμ12(|η^0(k)|+|k|1μ2|v^20(k,0)|+|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22),\displaystyle C|k|^{-\frac{1}{q}}\mu^{-\frac{1}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}\big{)},

where

Λ^T(k,x2)=i2k((1+sgn(kt))f1T0(k,U(0),x2)+(1sgn(kt))f1T0(k,U(0),x2)¯).\hat{\Lambda}_{T}(k,x_{2})=-\tfrac{i}{2k}\big{(}(1+sgn(kt))f_{1T}^{0}(k,U(0),x_{2})+(1-sgn(kt))\overline{f_{1T}^{0}(-k,U(0),x_{2})}\big{)}.

We give closer look at Λ^T\hat{\Lambda}_{T}. From boundary condition (5.1b), (5.20), and (6.3),

V2(k,U(0),0)=ζ+(U(0))/(g+σk2)=η^0(k),V_{2}(k,U(0),0)=-\zeta_{+}(U(0))/(g+\sigma k^{2})=-\hat{\eta}_{0}(k),

and thus

f1T0(k,U(0),x2)=U(0)η^0(k)ω^0(k,0)U(0)2y0(k,U(0),0)y0(k,U(0),x2).f_{1T}^{0}(k,U(0),x_{2})=\frac{U^{\prime\prime}(0)\hat{\eta}_{0}(k)-\hat{\omega}_{0}(k,0)}{U^{\prime}(0)^{2}y_{0-}(k,U(0),0)}y_{0-}(k,U(0),x_{2}).

Since y0(k,U(0),x2)y_{0-}(k,U(0),x_{2})\in\mathbb{R} for x2[h,0]x_{2}\in[-h,0], we obtain f1T0(k,U(0),x2)¯=f1T0(k,U(0),x2)\overline{f_{1T}^{0}(k,U(0),x_{2})}=f_{1T}^{0}(-k,U(0),x_{2}) and hence leads to the desired form (6.15) of ΛT\Lambda_{T}. The term involving f1Bf_{1B} can be analyzed similarly (actually slightly simpler due to V2(h)=0V_{2}(-h)=0) using Lemmas 3.143.16 and 3.18. The term involving V2V_{2}^{\prime\prime} can be estimated much as in (6.14). Summarizing this estimates we obtain

|t2tmv^2c(t,k,x2)(ik)m(iU(x2)mkU(x2)2eikU(x2)tΩ^c(k,x2)\displaystyle\Big{|}t^{2}\partial_{t}^{m}\hat{v}_{2}^{c}(t,k,x_{2})-(-ik)^{m}\Big{(}-\frac{iU(x_{2})^{m}}{kU^{\prime}(x_{2})^{2}}e^{-ikU(x_{2})t}\hat{\Omega}^{c}(k,x_{2})
+U(0)meikU(0)tΛ^T(k,x2)+U(h)meikU(h)tΛ^B(k,x2))|Lx22Ltq()\displaystyle\qquad\qquad+U(0)^{m}e^{-ikU(0)t}\hat{\Lambda}_{T}(k,x_{2})+U(-h)^{m}e^{-ikU(-h)t}\hat{\Lambda}_{B}(k,x_{2})\Big{)}\Big{|}_{L_{x_{2}}^{2}L_{t}^{q}(\mathbb{R})}
\displaystyle\leq C|k|m11qμ32(|η^0(k)|+|k|1μ2|v^20(k,0)|+μ1ϵ|ω^0(k)|Lx22+μ2ϵ|ω^0(k)|Lx22+μ3ϵ|ω^0(k)|Lx22).\displaystyle C|k|^{m-1-\frac{1}{q}}\mu^{-\frac{3}{2}}\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{2}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{1-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}+\mu^{2-\epsilon}|\hat{\omega}_{0}^{\prime}(k)|_{L_{x_{2}}^{2}}+\mu^{3-\epsilon}|\hat{\omega}_{0}^{\prime\prime}(k)|_{L_{x_{2}}^{2}}\big{)}.

Combining it with Lemma 6.6, the desired estimate on tm(t2v^2c)\partial_{t}^{m}(t^{2}\hat{v}_{2}^{c}) follows. ∎

6.2. Linearized capillary gravity waves in the horizontally periodic case of x1𝕋Lx_{1}\in\mathbb{T}_{L}

In this subsection, we consider the case where the system is periodic in x1x_{1} with wave length L>0L>0. In this case

k2πL,v^2(t,k=0,x2)=0,k\in\tfrac{2\pi}{L}\mathbb{Z},\quad\hat{v}_{2}(t,k=0,x_{2})=0,

where the latter properties is due to the divergence free condition on vv. For =c,p\dagger=c,p, let

v2(t,x)=|k|2πLv^2(t,k,x2)eikx1,ηc(t,x1)=|k|2πLη^c(t,k)eikx1,v1c(t,x)=|k|2πLv^1c(t,k,x2)eikx1,v_{2}^{\dagger}(t,x)=\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\hat{v}_{2}^{\dagger}(t,k,x_{2})e^{ikx_{1}},\;\;\eta^{c}(t,x_{1})=\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\hat{\eta}^{c}(t,k)e^{ikx_{1}},\;\;v_{1}^{c}(t,x)=\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\hat{v}_{1}^{c}(t,k,x_{2})e^{ikx_{1}},
ηp(t,x1)=η^0(0)+|k|2πLη^p(t,k)eikx1,v1p(t,x)=v^1(0,x2)+|k|2πLv^1p(t,k,x2)eikx1,v=(v1,v2),\eta^{p}(t,x_{1})=\hat{\eta}_{0}(0)+\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\hat{\eta}^{p}(t,k)e^{ikx_{1}},\;\;v_{1}^{p}(t,x)=\hat{v}_{1}(0,x_{2})+\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\hat{v}_{1}^{p}(t,k,x_{2})e^{ikx_{1}},\quad v^{\dagger}=(v_{1}^{\dagger},v_{2}^{\dagger}),

where v^1\hat{v}_{1}^{\dagger}, v^2\hat{v}_{2}^{\dagger}, and η^\hat{\eta}^{\dagger} are defined in Lemma 6.1 and Corollary 6.1.1. Here we used (2.7) that the zeroth modes v^1(k=0)\hat{v}_{1}(k=0) and η^(0)\hat{\eta}(0) are invariant in tt. Throughout this subsection, we assume (4.9) holds for 𝐊=2πL\mathbf{K}=\frac{2\pi}{L}\mathbb{N}.

We first give the decay estimates of (vc,ηc)(v^{c},\eta^{c}) based on Lemma 6.66.9. In particular, for the estimates of tv1ctv_{1}^{c} and t2v2ct^{2}v_{2}^{c}, recall Ω^c(k,x2)\hat{\Omega}^{c}(k,x_{2}) and Λ^(k,x2)\hat{\Lambda}_{\dagger}(k,x_{2}), =B,T\dagger=B,T defined in (6.9), (6.16), and (6.15), respective. Let

(6.17) Ωc(x1,x2)=|k|2πLΩ^c(k,x2)eikx1,Λ(x1,x2)=|k|2πLΛ^(k,x2)eikx1.\Omega^{c}(x_{1},x_{2})=\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\hat{\Omega}^{c}(k,x_{2})e^{ikx_{1}},\quad\Lambda_{\dagger}(x_{1},x_{2})=\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\hat{\Lambda}_{\dagger}(k,x_{2})e^{ikx_{1}}.

Proof of Theorem 2.1(1–2). The assumption of the non-existence of singular modes is given in the form of (4.9). According to Proposition 4.4, (4.9) for 𝐊=2πL\mathbf{K}=\frac{2\pi}{L}\mathbb{N} implies (5.8) holds for all kk with constants ρ0\rho_{0} and F0F_{0} uniform in kk. Therefore from the definition of v2cv_{2}^{c} and Lemma 6.6, it is straight forward to estimate

|tn0vc|Hx1n1Lx22Ltq1()2\displaystyle|\partial_{t}^{n_{0}}v^{c}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}L_{t}^{q_{1}}(\mathbb{R})}^{2}\leq C|k|2πLμ2n1|k|2n0+22q1(μ12|η^0(k)|+|k|1μ52|v^20(k,0)|+μ32ϵ|ω^0(k)|Lx22)2\displaystyle C\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}\mu^{-2n_{1}}|k|^{2n_{0}+2-\frac{2}{q_{1}}}\big{(}\mu^{\frac{1}{2}}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{\frac{5}{2}}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{\frac{3}{2}-\epsilon}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)}^{2}
\displaystyle\leq C|k|2πL|k|2(n0+n1+11q1)(|k|1|η^0(k)|2+|k|7|v^20(k,0)|2+|k|2ϵ3|ω^0(k)|Lx222)\displaystyle C\sum_{|k|\in\tfrac{2\pi}{L}\mathbb{N}}|k|^{2(n_{0}+n_{1}+1-\frac{1}{q_{1}})}\big{(}|k|^{-1}|\hat{\eta}_{0}(k)|^{2}+|k|^{-7}|\hat{v}_{20}^{\prime}(k,0)|^{2}+|k|^{2\epsilon-3}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}^{2}\big{)}
\displaystyle\leq C(|η0|Hx1n0+n1+121q12+|x2v20(,0)|Hx1n0+n1521q12+|ω0|Hx1n0+n1121q1+ϵLx222).\displaystyle C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{0}+n_{1}+\frac{1}{2}-\frac{1}{q_{1}}}}^{2}+|\partial_{x_{2}}v_{20}(\cdot,0)|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{5}{2}-\frac{1}{q_{1}}}}^{2}+|\omega_{0}|_{H_{x_{1}}^{n_{0}+n_{1}-\frac{1}{2}-\frac{1}{q_{1}}+\epsilon}L_{x_{2}}^{2}}^{2}\big{)}.

The desired inequality follows from x2v20=x1v10\partial_{x_{2}}v_{20}=-\partial_{x_{1}}v_{10}. The estimates on tn0ηc\partial_{t}^{n_{0}}\eta^{c}, ttn0v2ct\partial_{t}^{n_{0}}v_{2}^{c} and ttn0ηct\partial_{t}^{n_{0}}\eta^{c} are obtained similarly. The inequalities on tn0(tv1c)\partial_{t}^{n_{0}}(tv_{1}^{c}) and tn0(t2v2c)\partial_{t}^{n_{0}}(t^{2}v_{2}^{c}) are obtained by applying Lemma 6.8 and 6.9 through a similar procedure. The estimates on Ωc\Omega^{c} and Λ\Lambda_{\dagger}, =B,T\dagger=B,T, follow directly from the formula and estimates of their each Fourier modes given in those lemmas and

(6.18) |y0±(k,c,)/y0±(k,c,0)|Lx2qCμ1q,q[1,];|x2y0±(k,c,)/y0±(k,c,0)|Lx2qCμ1q1,q[1,);\begin{split}&|y_{0\pm}(k,c,\cdot)/y_{0\pm}(k,c,0)|_{L_{x_{2}}^{q}}\leq C\mu^{\frac{1}{q}},\;\forall q\in[1,\infty];\\ &|\partial_{x_{2}}y_{0\pm}(k,c,\cdot)/y_{0\pm}(k,c,0)|_{L_{x_{2}}^{q}}\leq C\mu^{\frac{1}{q}-1},\;\forall q\in[1,\infty);\end{split}

which is obtained using (5.8) and Lemma 3.9. The singular elliptic equations in (2.5) are simply from the homogeneous Rayleigh equation with c=U(h),U(0)c=U(-h),U(0), satisfied by y0±y_{0\pm} in (h,0)(-h,0). The boundary conditions of ΛB\Lambda_{B} and ΛT\Lambda_{T} are direct corollaries of their definitions and the boundary conditions (3.53) of y0±y_{0\pm}. \square

Next we consider the (vp(t,x),ηp(t,x1))(v^{p}(t,x),\eta^{p}(t,x_{1})) part of the linear solution (v,η)(v,\eta). Let

(6.19) λ0=max{Re(ick)k2πL,cR(k)}0,N=max{degree of root c of 𝐅(k,)k2πL,cR(k),Re(ikc)=λ0}1,\begin{split}&\lambda_{0}=\max\{\text{Re}\,(-ic_{*}k)\mid k\in\tfrac{2\pi}{L}\mathbb{N},\ c_{*}\in R(k)\}\geq 0,\\ &N=\max\{\text{degree of root }c_{*}\text{ of }\mathbf{F}(k,\cdot)\mid k\in\tfrac{2\pi}{L}\mathbb{N},\ c_{*}\in R(k),\ \text{Re}\,(-ikc_{*})=\lambda_{0}\}\geq 1,\end{split}

where the lower bounds are obtained due to the roots c±(k)c^{\pm}(k) for large kk (Lemma 4.2(3)).

Proof of Theorem 2.1(3). On the one hand, according to Lemma 4.2(3), there exists k0>0k_{0}>0 such that R(k)={c±(k)}R(k)=\{c^{\pm}(k)\} with simple roots c±(k)c^{\pm}(k) for all |k|k0|k|\geq k_{0}. On the other hand, (4.9) and Proposition 4.4 imply that (5.8) holds for all k2πLk\in\frac{2\pi}{L}\mathbb{N}. Along with Lemma 4.2(2), we obtain that, for all k2πLk\in\frac{2\pi}{L}\mathbb{N} with |k|<k0|k|<k_{0}, the set of roots R(k)R(k) is contained in a subset in the domain of analyticity of 𝐅(k,)\mathbf{F}(k,\cdot) uniformly in such kk. Hence R(k)R(k) is a discrete set and the total algebraic multiplicity of cR(k)c_{*}\in R(k) for all k2πLk\in\frac{2\pi}{L}\mathbb{N} with |k|<k0|k|<k_{0} is finite. This proves λ0,N<\lambda_{0},N<\infty.

For any k2πLk\in\frac{2\pi}{L}\mathbb{N} and cR(k)c_{*}\in R(k), let nn denote the degree of cc_{*} as a root of 𝐅(k,)\mathbf{F}(k,\cdot), then 𝐛\mathbf{b} and 𝐛S\mathbf{b}_{S} are polynomials of tt of degree n1n-1 (Lemma 6.2). Hence to prove the regularity estimates, we only need to consider k2πLk\in\frac{2\pi}{L}\mathbb{N} with |k|k0|k|\geq k_{0} where all roots of 𝐅(k,)\mathbf{F}(k,\cdot) are simple. For such kk, R(k)={c±(k)}R(k)=\{c^{\pm}(k)\} and Lemma 4.2(3) implies that there exists C>0C>0 such that

|c|1C|k|12,|cF(k,c)|1C|k|32,cR(k),k0|k|2πL.|c_{*}|\geq\tfrac{1}{C}|k|^{\frac{1}{2}},\;\;|\partial_{c}F(k,c_{*})|\geq\tfrac{1}{C}|k|^{\frac{3}{2}},\quad\forall c_{*}\in R(k),\;k_{0}\leq|k|\in\tfrac{2\pi}{L}\mathbb{N}.

From the homogeneous Rayleigh equation (3.1), (5.8), and Lemma 3.9, it holds,

(6.20) |x2sy(k,c,)|Lx22Cμ32s,s[0,l0],k,cR(k).|\partial_{x_{2}}^{s}y_{-}(k,c_{*},\cdot)|_{L_{x_{2}}^{2}}\leq C\mu^{\frac{3}{2}-s},\quad\forall\,s\in[0,l_{0}],\ k\in\mathbb{R},\ c_{*}\in R(k).

Hence Lemmas 6.1 and 6.2 and the definition of v2pv_{2}^{p} imply, for any n1n_{1}\in\mathbb{R} and n2[0,l0]n_{2}\in[0,l_{0}],

k0|k|2πLμ2n1|v^2p(t,k,)|Hx2n22Ck0|k|2πLc=c±(k)μ2n1|𝐛(k,c,)|Hx2n22\displaystyle\sum_{k_{0}\leq|k|\in\tfrac{2\pi}{L}\mathbb{N}}\mu^{-2n_{1}}|\hat{v}_{2}^{p}(t,k,\cdot)|_{H_{x_{2}}^{n_{2}}}^{2}\leq C\sum_{k_{0}\leq|k|\in\tfrac{2\pi}{L}\mathbb{N}}\sum_{c_{*}=c^{\pm}(k)}\mu^{-2n_{1}}|\mathbf{b}(k,c_{*},\cdot)|_{H_{x_{2}}^{n_{2}}}^{2}
\displaystyle\leq Ck0|k|2πL|k|2(n1+n2)4(|k|3|η^0(k)|+|k|12|v^20(k,0)|+|k||ω^0(k,)|Lx22)2\displaystyle C\sum_{k_{0}\leq|k|\in\tfrac{2\pi}{L}\mathbb{N}}|k|^{2(n_{1}+n_{2})-4}\big{(}|k|^{3}|\hat{\eta}_{0}(k)|+|k|^{\frac{1}{2}}|\hat{v}_{20}^{\prime}(k,0)|+|k||\hat{\omega}_{0}(k,\cdot)|_{L_{x_{2}}^{2}}\big{)}^{2}
\displaystyle\leq C(|η0|Hx1n1+n2+12+|x2v20(,0)|Hx1n1+n2322+|ω0|Hx1n1+n21Lx222).\displaystyle C\big{(}|\eta_{0}|_{H_{x_{1}}^{n_{1}+n_{2}+1}}^{2}+|\partial_{x_{2}}v_{20}(\cdot,0)|_{H_{x_{1}}^{n_{1}+n_{2}-\frac{3}{2}}}^{2}+|\omega_{0}|_{H_{x_{1}}^{n_{1}+n_{2}-1}L_{x_{2}}^{2}}^{2}\big{)}.

The desired inequality follows from the divergence free condition. The expression of v1pv_{1}^{p} involves yy_{-}^{\prime} and thus it can be differentiated in x2x_{2} at most l01l_{0}-1 times. The procedure to obtain the estimates of v1pv_{1}^{p} and ηp\eta^{p} are similar and we skip the details. \square

Finally we give the invariant decomposition of the phase space which proves Theorem 2.1(4).

Lemma 6.10.

Let

𝐗p=span{range(eikx1𝐏(k,c))cR(k),k2πL}¯H1(𝕋L×(h,0))×H2(𝕋L),\mathbf{X}^{p}=\overline{span\{range(e^{ikx_{1}}\mathbf{P}(k,c_{*}))\mid c_{*}\in R(k),\,k\in\tfrac{2\pi}{L}\mathbb{Z}\}}\subset H^{1}(\mathbb{T}_{L}\times(-h,0))\times H^{2}(\mathbb{T}_{L}),
𝐏(v,η)=cR(k),k2πLeikx1𝐏(k,c)(v^(k),η^(k)),𝐗c=ker𝐏H1(𝕋L×(h,0))×H2(𝕋L).\mathbf{P}(v,\eta)=\oplus_{c_{*}\in R(k),k\in\tfrac{2\pi}{L}\mathbb{Z}}e^{ikx_{1}}\mathbf{P}(k,c_{*})\big{(}\hat{v}(k),\hat{\eta}(k)\big{)},\quad\mathbf{X}^{c}=\ker\mathbf{P}\subset H^{1}(\mathbb{T}_{L}\times(-h,0))\times H^{2}(\mathbb{T}_{L}).

where 𝐏(k,c)\mathbf{P}(k,c_{*}) was defined in (6.8), then the following hold.

  1. (1)

    𝐏\mathbf{P} is a bounded projection operator from Hn(𝕋L×(h,0))×Hn+1(𝕋L)H^{n}(\mathbb{T}_{L}\times(-h,0))\times H^{n+1}(\mathbb{T}_{L}) to 𝐗p(Hn(𝕋L×(h,0))×Hn+1(𝕋L))\mathbf{X}^{p}\cap\big{(}H^{n}(\mathbb{T}_{L}\times(-h,0))\times H^{n+1}(\mathbb{T}_{L})\big{)} for any integer n[1,l01]n\in[1,l_{0}-1].

  2. (2)

    𝐗p\mathbf{X}^{p} and 𝐗c\mathbf{X}^{c} are both invariant subspaces of (2.6).

  3. (3)

    Moreover (2.6) is also well-posed on the L2×H1L^{2}\times H^{1} completion of 𝐗p\mathbf{X}^{p} and is a (possibly unstable) dispersive equation with the (multi-branches of) dispersion relation given by kkck\to-kc_{*} including all cR(k)c_{*}\in R(k).

The boundedness of 𝐏\mathbf{P} follows from the estimates in Theorem 2.1 at t=0t=0. The invariance of 𝐗p\mathbf{X}^{p} and 𝐗c\mathbf{X}^{c} is due to Lemma 6.2 and Corollary 6.2.1. The well-posedness of (2.6) on the L2×H1L^{2}\times H^{1} completion of 𝐗p\mathbf{X}^{p} is due to the fact that R(k)={c±(k)}U([h,0])R(k)=\{c^{\pm}(k)\}\subset\mathbb{R}\setminus U([-h,0]) except for finitely many k2πLk\in\tfrac{2\pi}{L}\mathbb{Z}. Here we did not set 𝐗p\mathbf{X}^{p} and 𝐗c\mathbf{X}^{c} in L2×H1L^{2}\times H^{1} is due to the issue that we can not ensure v1(,0)Hx112v_{1}(\cdot,0)\in H_{x_{1}}^{-\frac{1}{2}} for vL2v\in L^{2}.

6.3. Linearized capillary gravity waves in the horizontally infinite case of x1x_{1}\in\mathbb{R}

In this subsection, we consider the case where x1x_{1}\in\mathbb{R} and thus kk\in\mathbb{R}. Throughout this subsection, we assume (4.9) for 𝐊=\mathbf{K}=\mathbb{R}. For =c,p\dagger=c,p, let

(6.21) v(t,x)=v^(t,k,x2)eikx1dk,η(t,x1)=η^(t,k)eikx1dk,v=(v1,v2),v^{\dagger}(t,x)=\int_{\mathbb{R}}\hat{v}^{\dagger}(t,k,x_{2})e^{ikx_{1}}dk,\;\;\eta^{\dagger}(t,x_{1})=\int_{\mathbb{R}}\hat{\eta}^{\dagger}(t,k)e^{ikx_{1}}dk,\quad v^{\dagger}=(v_{1}^{\dagger},v_{2}^{\dagger}),

where v^1\hat{v}_{1}^{\dagger}, v^2\hat{v}_{2}^{\dagger}, and η^\hat{\eta}^{\dagger} are defined in Lemma 6.1 and Corollary 6.1.1.

Again we first carry out the decay estimates of (vc,ηc)(v^{c},\eta^{c}) based on Lemma 6.66.9. Let

(6.22) Ωc(x1,x2)=Ω^c(k,x2)eikx1dk,Λ(x1,x2)=Λ^(k,x2)eikx1dk.\Omega^{c}(x_{1},x_{2})=\int_{\mathbb{R}}\hat{\Omega}^{c}(k,x_{2})e^{ikx_{1}}dk,\quad\Lambda_{\dagger}(x_{1},x_{2})=\int_{\mathbb{R}}\hat{\Lambda}_{\dagger}(k,x_{2})e^{ikx_{1}}dk.

Proof of Theorem 2.2(1–3). Again the assumption of the non-existence of singular modes is given in the form of (4.9). According to Proposition 4.4, assumption (4.9) for 𝐊=\mathbf{K}=\mathbb{R} implies that (5.8) holds and R(k)={c±(k)}R(k)=\{c^{\pm}(k)\} with all these simple roots c±(k)c^{\pm}(k) of 𝐅(k,)\mathbf{F}(k,\cdot) away from U([h,0])U([-h,0]) for all kk\in\mathbb{R}. Moreover, Lemma 4.2 yields

|dist(c±(k),U([h,0]))|1Cμ12,|cF(k,c±(k))|1Cμ32,k.|dist(c^{\pm}(k),U([-h,0]))|\geq\tfrac{1}{C}\mu^{-\frac{1}{2}},\;\;|\partial_{c}F(k,c^{\pm}(k))|\geq\tfrac{1}{C}\mu^{-\frac{3}{2}},\quad\forall k\in\mathbb{R}.

Like in the periodic-in-x1x_{1} case, the proof of the decay of (vc,ηc)(v^{c},\eta^{c}) is also a direct verification using Lemmas 6.66.9 along with (6.18) and the divergence free condition. We omit the details.

From Lemmas 6.1 and 6.2(3), we obtain 𝐛\mathbf{b} and 𝐛S\mathbf{b}_{S} are independent of tt and satisfy, for any n2[0,l0]n_{2}\in[0,l_{0}],

|x2n2𝐛(k,c±(k),x2)|C(|k|μ12|η^0(k)|+μ|v^20(k,0)|+|k|μ32|ω^0(k)|Lx22)|μ1n2eμ1(x2+h)y(k,c±(k),0)|,|\partial_{x_{2}}^{n_{2}}\mathbf{b}(k,c^{\pm}(k),x_{2})|\leq C\big{(}|k|\mu^{-\frac{1}{2}}|\hat{\eta}_{0}(k)|+\mu|\hat{v}_{20}^{\prime}(k,0)|+|k|\mu^{\frac{3}{2}}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)}\big{|}\tfrac{\mu^{1-n_{2}}e^{\mu^{-1}(x_{2}+h)}}{y_{-}(k,c^{\pm}(k),0)}\big{|},
|𝐛S(k,c±(k))|C(|η^0(k)|+|k|1μ32|v^20(k,0)|+μ2|ω^0(k)|Lx22).|\mathbf{b}_{S}(k,c^{\pm}(k))|\leq C\big{(}|\hat{\eta}_{0}(k)|+|k|^{-1}\mu^{\frac{3}{2}}|\hat{v}_{20}^{\prime}(k,0)|+\mu^{2}|\hat{\omega}_{0}(k)|_{L_{x_{2}}^{2}}\big{)}.

The desired estimates follow from (6.20), ikv^1=v^2ik\hat{v}_{1}=-\hat{v}_{2}, and direct computations. \square

Similar to the periodic case, we also have the decomposition by invariant subspaces.

Lemma 6.11.

Let

𝐏(v,η)=𝐏(k,c+(k))(v,η)dk+𝐏(k,c(k))(v,η)dk,\mathbf{P}(v,\eta)=\int_{\mathbb{R}}\mathbf{P}(k,c^{+}(k))(v,\eta)dk+\int_{\mathbb{R}}\mathbf{P}(k,c^{-}(k))(v,\eta)dk,
𝐗p=range(𝐏)H1(×(h,0))×H2(),𝐗c=ker𝐏H1(×(h,0))×H2(),\mathbf{X}^{p}=range(\mathbf{P})\subset H^{1}(\mathbb{R}\times(-h,0))\times H^{2}(\mathbb{R}),\quad\mathbf{X}^{c}=\ker\mathbf{P}\subset H^{1}(\mathbb{R}\times(-h,0))\times H^{2}(\mathbb{R}),

where 𝐏(k,c±(k))\mathbf{P}(k,c^{\pm}(k)) was defined in (6.8), then the following hold.

  1. (1)

    𝐏\mathbf{P} is a bounded projection operator from Hn(×(h,0))×Hn+1()H^{n}(\mathbb{R}\times(-h,0))\times H^{n+1}(\mathbb{R}) to 𝐗p(Hn(×(h,0))×Hn+1())\mathbf{X}^{p}\cap\big{(}H^{n}(\mathbb{R}\times(-h,0))\times H^{n+1}(\mathbb{R})\big{)} for any integer n[1,l01]n\in[1,l_{0}-1].

  2. (2)

    𝐗p\mathbf{X}^{p} and 𝐗c\mathbf{X}^{c} are both invariant subspaces of (2.6).

  3. (3)

    In fact (2.6) is also well-posed on the L2×H1L^{2}\times H^{1} completion of 𝐗p\mathbf{X}^{p} and is a dispersive equation with the dispersion relation given by kkc±(k)k\to-kc^{\pm}(k).

To end this subsection we show that, under assumptions (4.9) for 𝐊=\mathbf{K}=\mathbb{R} and (4.18) for c±(k)c^{\pm}(k), due to the monotonicity of c±(k)c^{\pm}(k) in k>0k>0 (Lemma 4.7) and the asymptotics of c±(k)c^{\pm}(k) for |k|1|k|\gg 1 (Lemma 4.2(3)), the dynamics of the non-singular modes is conjugate to that of linear irrotational capillary gravity waves.

For kk\in\mathbb{R}, let

e±(k,x2)=(v1,v2,η)=e|k|h(μ12y(k,c±(k),x2),ikμ12y(k,c±(k),x2),y(k,c±(k),0)μ12(U(0)c±(k))),e^{\pm}(k,x_{2})=(v_{1},v_{2},\eta)=e^{-|k|h}\Big{(}\mu^{-\frac{1}{2}}y_{-}^{\prime}(k,c^{\pm}(k),x_{2}),-ik\mu^{-\frac{1}{2}}y_{-}(k,c^{\pm}(k),x_{2}),-\frac{y_{-}(k,c^{\pm}(k),0)}{\mu^{\frac{1}{2}}(U(0)-c^{\pm}(k))}\Big{)},
eir±(k,x2)=(v1,v2,η)=e|k|h(μ12coshk(x2+h),iμ12sinhk(x2+h),sinhkhkμ12cir±(k)),e_{ir}^{\pm}(k,x_{2})=(v_{1},v_{2},\eta)=e^{-|k|h}\Big{(}\mu^{-\frac{1}{2}}\cosh k(x_{2}+h),-i\mu^{-\frac{1}{2}}\sinh k(x_{2}+h),\frac{\sinh kh}{k\mu^{\frac{1}{2}}c_{ir}^{\pm}(k)}\Big{)},\;

where cir±(k)c_{ir}^{\pm}(k) is the wave speed of the free linear capillary gravity wave (system (1.3) with U0U\equiv 0 and ×v0\nabla\times v\equiv 0) given in (2.4). Here e±(k)e^{\pm}(k) correspond to the two non-singular modes in the kk-th Fourier modes in x1x_{1}, while eir±e_{ir}^{\pm} the modes of irrotational linear capillary gravity waters waves. Define

±(f)=f(k)eikx1e±(k)dk,ir±(f)=f(k)eikx1eir±(k)dk,\mathcal{E}^{\pm}\big{(}f\big{)}=\int_{\mathbb{R}}f(k)e^{ikx_{1}}e^{\pm}(k)dk,\quad\mathcal{E}_{ir}^{\pm}\big{(}f\big{)}=\int_{\mathbb{R}}f(k)e^{ikx_{1}}e_{ir}^{\pm}(k)dk,
𝐗±={±(f)fL2()},𝐗ir±={ir±(f)fL2()}.\mathbf{X}^{\pm}=\{\mathcal{E}^{\pm}(f)\mid f\in L^{2}(\mathbb{R})\},\quad\mathbf{X}_{ir}^{\pm}=\{\mathcal{E}_{ir}^{\pm}(f)\mid f\in L^{2}(\mathbb{R})\}.

Clearly 𝐗+𝐗\mathbf{X}^{+}\oplus\mathbf{X}^{-} is equal to the L2×H1L^{2}\times H^{1} completion of 𝐗p\mathbf{X}^{p} and ±:L2()𝐗±\mathcal{E}^{\pm}:L^{2}(\mathbb{R})\to\mathbf{X}^{\pm} and ir±:L2()𝐗ir±\mathcal{E}_{ir}^{\pm}:L^{2}(\mathbb{R})\to\mathbf{X}_{ir}^{\pm} parametrize 𝐗±\mathbf{X}^{\pm} and 𝐗ir±\mathbf{X}_{ir}^{\pm} by L2L^{2}. The following proposition finishes the proof of Theorem 1.1(2b) and Theorem 2.2(4).

Proposition 6.12.

Assume UC3U\in C^{3} and (4.9) for 𝐊=\mathbf{K}=\mathbb{R}, then the following hold.

  1. (1)

    The mappings ±\mathcal{E}^{\pm} and ir±\mathcal{E}_{ir}^{\pm} are isomorphisms. Moreover there exists C>0C>0 depending only on UU such that

    C1|e±(k)|L2,|eir±(k)|L2C,C1|f|L2|±(f)|L2,|ir±(f)|C|f|L2,k,fL2().C^{-1}\leq|e^{\pm}(k)|_{L^{2}},|e_{ir}^{\pm}(k)|_{L^{2}}\leq C,\;\;\ C^{-1}|f|_{L^{2}}\leq|\mathcal{E}^{\pm}(f)|_{L^{2}},|\mathcal{E}_{ir}^{\pm}(f)|\leq C|f|_{L^{2}},\quad\forall k\in\mathbb{R},\,f\in L^{2}(\mathbb{R}).
  2. (2)

    For any solution (v(t,x),η(t,x1))(v(t,x),\eta(t,x_{1})) to the capillary gravity wave linearized at the shear flow U(x2)U(x_{2}), if its component (vp,ηp)(v^{p},\eta^{p}) as defined in (6.21) belongs to 𝐗+𝐗\mathbf{X}^{+}\oplus\mathbf{X}^{-}, then it takes the form

    (6.23) (vp,ηp)=+(eikc+(k)tf+(k))+(eikc(k)tf(k)),(v^{p},\eta^{p})=\mathcal{E}^{+}(e^{-ikc^{+}(k)t}f_{+}(k))+\mathcal{E}^{-}(e^{-ikc^{-}(k)t}f_{-}(k)),

    for some unique f±L2()f_{\pm}\in L^{2}(\mathbb{R}). Similarly, any solution (v(t,x),η(t,x1))L2(v(t,x),\eta(t,x_{1}))\in L^{2} to the free linear capillary gravity wave (system (1.3) with U0U\equiv 0), then it takes the form

    (6.24) (v,η)=ir+(eikcir+(k)tf+(k))+ir(eikcir(k)tf(k)),f±L2.(v,\eta)=\mathcal{E}_{ir}^{+}(e^{-ikc_{ir}^{+}(k)t}f_{+}(k))+\mathcal{E}_{ir}^{-}(e^{-ikc_{ir}^{-}(k)t}f_{-}(k)),\quad f_{\pm}\in L^{2}.
  3. (3)

    In addition, assume (4.18) for c±(k)c^{\pm}(k) and 0U([h,0])0\in U\big{(}[-h,0]\big{)}, then there exist odd C1C^{1} functions φ±(k)\varphi_{\pm}(k) and C>0C>0 depending only on UU such that

    φ±(k)c±(φ±(k))=kcir±(k),C1|k|1|φ±(k)|,(φ±)(k)C,k.\varphi^{\pm}(k)c^{\pm}(\varphi^{\pm}(k))=kc_{ir}^{\pm}(k),\quad C^{-1}\leq|k|^{-1}|\varphi^{\pm}(k)|,(\varphi^{\pm})^{\prime}(k)\leq C,\quad\forall k\in\mathbb{R}.

    Define Φ±:𝐗±𝐗ir±\Phi^{\pm}:\mathbf{X}^{\pm}\to\mathbf{X}_{ir}^{\pm} as

    Φ±(±(f))=ir±(fφ±)\Phi^{\pm}\big{(}\mathcal{E}^{\pm}(f)\big{)}=\mathcal{E}_{ir}^{\pm}(f\circ\varphi^{\pm})

    for any ±(f)𝐗±\mathcal{E}^{\pm}(f)\in\mathbf{X}^{\pm}, then Φ++Φ\Phi^{+}+\Phi^{-} is an isomorphism from (𝐗+𝐗)(Hn×Hn+1)(\mathbf{X}^{+}\oplus\mathbf{X}^{-})\cap(H^{n}\times H^{n+1}) to (𝐗ir+𝐗ir)(Hn×Hn+1)(\mathbf{X}_{ir}^{+}\oplus\mathbf{X}_{ir}^{-})\cap(H^{n}\times H^{n+1}) for any n[0,l01]n\in[0,l_{0}-1]. Moreover flows (6.23) and (6.24) are conjugate through Φ++Φ\Phi^{+}+\Phi^{-}. Namely, for any f±L2f_{\pm}\in L^{2}, it holds

    (6.25) Φ+(+(eikc+(k)tf+(k)))+Φ((eikc(k)tf(k)))=ir+(eikcir+(k)tf+(φ+(k)))+ir(eikcir(k)tf(φ(k))).\begin{split}&\Phi^{+}\big{(}\mathcal{E}^{+}(e^{-ikc^{+}(k)t}f_{+}(k))\big{)}+\Phi^{-}\big{(}\mathcal{E}^{-}(e^{-ikc^{-}(k)t}f_{-}(k))\big{)}\\ =&\mathcal{E}_{ir}^{+}(e^{-ikc_{ir}^{+}(k)t}f_{+}(\varphi^{+}(k)))+\mathcal{E}_{ir}^{-}(e^{-ikc_{ir}^{-}(k)t}f_{-}(\varphi^{-}(k))).\end{split}
Proof.

The estimates on |e±(k)|L2|e^{\pm}(k)|_{L^{2}} and |eir±(k)|L2|e_{ir}^{\pm}(k)|_{L^{2}} are derived from direct computations based on Lemma 3.9. In particular, since c±(k)U([h,0])c^{\pm}(k)\in\mathbb{R}\setminus U([-h,0]), formula (4.13) of yy_{-} for k=0k=0 and the bound on ky\partial_{k}y_{-} are used in obtaining the lower bounds of |e±(k)|L2|e^{\pm}(k)|_{L^{2}} for |k||k| close to 0. The estimates of |±(f)|L2|\mathcal{E}^{\pm}(f)|_{L^{2}} and |ir±(f)|L2|\mathcal{E}_{ir}^{\pm}(f)|_{L^{2}} follow from those of e±(f)e^{\pm}(f) and eir±(f)e_{ir}^{\pm}(f) and the Parseval’s identity. Statement (2) is a direct consequence of Lemma 2.3 and the definition of c±(k)c^{\pm}(k) and cir±(k)c_{ir}^{\pm}(k).

Since cir±(0)=gh0c_{ir}^{\pm}(0)=\sqrt{gh}\neq 0 and c±(0)U([h,0])c^{\pm}(0)\notin U([-h,0]), under the additional assumptions (4.18) and 0U((h,0))0\in U\big{(}(-h,0)\big{)}, Proposition 4.4 and Lemma 4.7 imply that a.) both kc±(k)kc^{\pm}(k) and kcir±(k)kc_{ir}^{\pm}(k) are odd in kk, b.) both ±kc±(k)\pm kc^{\pm}(k) and ±kcir±(k)\pm kc_{ir}^{\pm}(k) have positive derivative for k>0k>0, and c.) both are of the order O(|k|32)O(|k|^{\frac{3}{2}}) for |k|1|k|\gg 1 and of the order O(|k|)O(|k|) for |k|1|k|\ll 1. Hence φ±\varphi^{\pm} exist and satisfy the estimates, which implies the boundedness of Φ\Phi. The conjugacy relation (6.25) can be verified directly using (6.23), (6.24), and the definition of φ±\varphi^{\pm}. ∎

Remark 6.5.

Under (4.18), 0U([h,0])0\in U([-h,0]), and F(k,U(h))0F(k,U(-h))\neq 0 for all kk\in\mathbb{R}, without assuming (4.9), 𝐗+𝐗\mathbf{X}^{+}\oplus\mathbf{X}^{-} may only be a closed subspace of 𝐗p\mathbf{X}^{p}, but c±(k)U([h,0])c^{\pm}(k)\in\mathbb{R}\setminus U([-h,0]) are still monotonic and isolated from the rest of the singular or non-singular modes. The exactly same argument implies that the conclusions of the above proposition still hold on 𝐗+𝐗\mathbf{X}^{+}\oplus\mathbf{X}^{-}.

6.4. A remark on the linearized Euler equation on a fixed 2-d channel

We briefly comment on the 2-d Euler equation on a fixed channel x2(h,0)x_{2}\in(-h,0) with slip boundary condition v2=0v_{2}=0 at x2=h,0x_{2}=-h,0. Let U(x2)U(x_{2}) be a shear flow and we assume

(𝐇\mathbf{H}) U>0 and there are no singular modes.U^{\prime}>0\text{ and there are no singular modes.}

As in the literatures, singular modes mean linearized solutions in the form of eik(x1ct)v(x2)e^{ik(x_{1}-ct)}v(x_{2}) with vHx21v\in H_{x_{2}}^{1} and cU([h,0])c\in U([-h,0]).

The approach in this paper can be easily adapted to analyze this problem. While the non-homogeneous term in the Rayleigh equation (2.11a) is still ω^0(k,x2)U(x2)c-\frac{\hat{\omega}_{0}(k,x_{2})}{U(x_{2})-c}, the main modifications are: a.) replacing y+(k,c,x2)y_{+}(k,c,x_{2}) and V2(k,c,x2)V_{2}(k,c,x_{2}) by y~+(k,c,x2)\tilde{y}_{+}(k,c,x_{2}) and yE(k,c,x2)y_{E}(k,c,x_{2}) which solve the homogeneous and non-homogeneous Rayleigh equations satisfying boundary conditions

y~+(0)=yE(0)=yE(h)=0,y~+(0)=1,\tilde{y}_{+}(0)=y_{E}(0)=y_{E}(-h)=0,\quad\tilde{y}_{+}^{\prime}(0)=1,

respectively, and b.) replacing 𝐅(k,c)\mathbf{F}(k,c) by y(k,c,0)y_{-}(k,c,0). For the simplification of notations, we also use yy_{-}, y~+\tilde{y}_{+}, and yEy_{E} to denote their limits as cI0+c_{I}\to 0+. In this case of channel flow with fixed boundary, obviously the set of non-singular modes (roots of y(k,c,0)y_{-}(k,c,0) outside U([h,0])U([-h,0])) for all kk\in\mathbb{R} is finite, actually empty if U0U^{\prime\prime}\neq 0. Assuming (𝐇\mathbf{H}), through the same procedure as in Lemma 6.1, the solution v(t,x)v(t,x) to the linearized Euler equation at the shear flow U(x2)U(x_{2}) can also be split into

v(t,x)=vc(t,x)+vp(t,x)v(t,x)=v^{c}(t,x)+v^{p}(t,x)

associated to the continuous spectra and point spectra. Under assumption (𝐇\mathbf{H}), vp(t,)v^{p}(t,\cdot) belongs to the eigenspace of unstable modes which is finite dimensional if x1Lx_{1}\in\mathbb{Z}_{L}. Let

Ω^c(k,x2)=\displaystyle\hat{\Omega}^{c}(k,x_{2})= ω^0(k,x2)\displaystyle\hat{\omega}_{0}(k,x_{2})
+12U(x2)((1+sgn(kt))yE(k,U(x2),x2)+(1sgn(kt))yE(k,U(x2),x2)¯),\displaystyle+\tfrac{1}{2}U^{\prime\prime}(x_{2})\big{(}(1+sgn(kt))y_{E}(k,U(x_{2}),x_{2})+(1-sgn(kt))\overline{y_{E}(-k,U(x_{2}),x_{2})}\big{)},
Λ^T(k,x2)=\displaystyle\hat{\Lambda}_{T}(k,x_{2})= iω^0(k,0)y(k,U(0),x2)kU(0)2y(k,U(0),0),Λ^B(k,x2)=iω^0(k,h)y~+(k,U(h),x2)kU(h)2y~+(k,U(h),h),\displaystyle\frac{i\hat{\omega}_{0}(k,0)y_{-}(k,U(0),x_{2})}{kU^{\prime}(0)^{2}y_{-}(k,U(0),0)},\quad\hat{\Lambda}_{B}(k,x_{2})=\frac{i\hat{\omega}_{0}(k,-h)\tilde{y}_{+}(k,U(-h),x_{2})}{kU^{\prime}(-h)^{2}\tilde{y}_{+}(k,U(-h),-h)},

and Ωc\Omega^{c} and Λ\Lambda_{\dagger}, =B,T\dagger=B,T, be defined as in (6.17) for the LL-periodic-in-x1x_{1} case and in (6.22) for the case of x1x_{1}\in\mathbb{R}.

Theorem 6.13.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, and (𝐇\mathbf{H}) holds for all kKk\in K where K=2πLK=\frac{2\pi}{L}\mathbb{N} or K=K=\mathbb{R}, then, for any q1[2,]q_{1}\in[2,\infty], q2(2,]q_{2}\in(2,\infty], ϵ>0\epsilon>0, n1n_{1}\in\mathbb{R}, and integer n00n_{0}\geq 0, there exists C>0C>0 depending only on q1q_{1}, q2q_{2}, ϵ\epsilon, and UU such that any solution with v^10(0,x2)=0\hat{v}_{10}(0,x_{2})=0 satisfy

|tn0x1n1v1c|Lx2Ltq1()+|tn0x1n11(1x12)12\displaystyle|\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}v_{1}^{c}|_{L_{x}^{2}L_{t}^{q_{1}}(\mathbb{R})}+|\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}-1}(1-\partial_{x_{1}}^{2})^{\frac{1}{2}} v2c|Lx2Ltq1()C||x1|n0+n11q1ω0|Hx1ϵ12Lx22;\displaystyle v_{2}^{c}|_{L_{x}^{2}L_{t}^{q_{1}}(\mathbb{R})}\leq C\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{1}{2}}L_{x_{2}}^{2}};

if l04l_{0}\geq 4, then

|ttn0x1n1(1x12)12v2c|Lx2Ltq1()+|tn0x1n1+1(tv1cU(x2)1x11Ωc(x1U(x2)t,x2))|Lx2Ltq2()\displaystyle\big{|}t\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}(1-\partial_{x_{1}}^{2})^{\frac{1}{2}}v_{2}^{c}\big{|}_{L_{x}^{2}L_{t}^{q_{1}}(\mathbb{R})}+\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}+1}\big{(}tv_{1}^{c}-U^{\prime}(x_{2})^{-1}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
+|tn0x1n1(ωcΩc(x1U(x2)t,x2))|Lx2Ltq2()\displaystyle+\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}}\big{(}\omega^{c}-\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
+|tn0x1n11(x22v2cx1Ωc(x1U(x2)t,x2))|Lx2Ltq2()\displaystyle+\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}-1}\big{(}\partial_{x_{2}}^{2}v_{2}^{c}-\partial_{x_{1}}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
\displaystyle\leq C(||x1|n0+n11q1ω0|Hx1ϵ+12Lx22+||x1|n0+n11q1x2ω0|Hx1ϵ12Lx22);\displaystyle C\big{(}\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon+\frac{1}{2}}L_{x_{2}}^{2}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\partial_{x_{2}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{1}{2}}L_{x_{2}}^{2}}\big{)};

and if UC5U\in C^{5}, then

|tn0x1n1+1(t2v2cU(x2)2x11Ωc(x1U(x2)t,x2)ΛB(x1U(h)t,x2)\displaystyle\big{|}\partial_{t}^{n_{0}}\partial_{x_{1}}^{n_{1}+1}\big{(}t^{2}v_{2}^{c}-U^{\prime}(x_{2})^{-2}\partial_{x_{1}}^{-1}\Omega^{c}(x_{1}-U(x_{2})t,x_{2})-\Lambda_{B}(x_{1}-U(-h)t,x_{2})
ΛT(x1U(0)t,x2))|Lx2Ltq2()\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\Lambda_{T}(x_{1}-U(0)t,x_{2})\big{)}\big{|}_{L_{x}^{2}L_{t}^{q_{2}}(\mathbb{R})}
\displaystyle\leq C(||x1|n0+n11q1ω0|Hx1ϵ+12Lx22+||x1|n0+n11q1x2ω0|Hx1ϵ12Lx22+||x1|n0+n11q1x22ω0|Hx1ϵ32Lx22).\displaystyle C\big{(}\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon+\frac{1}{2}}L_{x_{2}}^{2}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\partial_{x_{2}}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{1}{2}}L_{x_{2}}^{2}}+\big{|}|\partial_{x_{1}}|^{n_{0}+n_{1}-\frac{1}{q_{1}}}\partial_{x_{2}}^{2}\omega_{0}\big{|}_{H_{x_{1}}^{\epsilon-\frac{3}{2}}L_{x_{2}}^{2}}\big{)}.

Moreover,

|Ωcω0|Hx1n1Lx22C|ω0|Hx1n11+ϵLx22,|x2Ωcx2ω0|Hx1n1Lx22C(|ω0|Hx1n1+ϵLx22+|x2ω0|Hx1n11+ϵLx22).|\Omega^{c}-\omega_{0}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}}\leq C|\omega_{0}|_{H_{x_{1}}^{n_{1}-1+\epsilon}L_{x_{2}}^{2}},\quad|\partial_{x_{2}}\Omega^{c}-\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{1}}L_{x_{2}}^{2}}\leq C\big{(}|\omega_{0}|_{H_{x_{1}}^{n_{1}+\epsilon}L_{x_{2}}^{2}}+|\partial_{x_{2}}\omega_{0}|_{H_{x_{1}}^{n_{1}-1+\epsilon}L_{x_{2}}^{2}}\big{)}.
|kΛ^B(k,)|Lx2qCk1q|ω^0(,h)|,|kΛ^T(k,)|Lx2qCk1q|ω^0(k,0)|,q[1,],|k\hat{\Lambda}_{B}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{-\frac{1}{q}}|\hat{\omega}_{0}(\cdot,-h)|,\quad|k\hat{\Lambda}_{T}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{-\frac{1}{q}}|\hat{\omega}_{0}(k,0)|,\quad\;\forall q\in[1,\infty],
|kx2Λ^B(k,)|Lx2qCk11q|ω^0(,h)|,|kx2Λ^T(k,)|Lx2qCk11q|ω^0(k,0)|,q[1,).|k\partial_{x_{2}}\hat{\Lambda}_{B}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{1-\frac{1}{q}}|\hat{\omega}_{0}(\cdot,-h)|,\;\;|k\partial_{x_{2}}\hat{\Lambda}_{T}(k,\cdot)|_{L_{x_{2}}^{q}}\leq C\langle k\rangle^{1-\frac{1}{q}}|\hat{\omega}_{0}(k,0)|,\;\;\forall q\in[1,\infty).

Finally, Λ\Lambda_{\dagger}, =B,T\dagger=B,T, satisfy Λ^(k=0,x2)=0\hat{\Lambda}_{\dagger}(k=0,x_{2})=0 and

{(UU(0))ΔΛT+UΛT=0,x2(h,0),ΛT(x1,h)=0,x1ΛT(x1,0)=U(0)2ω0(x1,0);\begin{cases}-(U-U(0))\Delta\Lambda_{T}+U^{\prime\prime}\Lambda_{T}=0,\qquad\qquad\qquad\qquad\qquad\qquad\qquad x_{2}\in(-h,0),\\ \Lambda_{T}(x_{1},-h)=0,\qquad\partial_{x_{1}}\Lambda_{T}(x_{1},0)=-U^{\prime}(0)^{-2}\omega_{0}(x_{1},0);\end{cases}
{(UU(h))ΔΛB+UΛB=0,x2(h,0),x1ΛB(,h)=U(h)2ω0(x1,h),ΛB(x1,0)=0.\begin{cases}-(U-U(-h))\Delta\Lambda_{B}+U^{\prime\prime}\Lambda_{B}=0,\qquad\qquad\qquad\qquad\qquad\qquad\qquad x_{2}\in(-h,0),\\ \partial_{x_{1}}\Lambda_{B}(\cdot,-h)=-U^{\prime}(-h)^{-2}\omega_{0}(x_{1},-h),\qquad\Lambda_{B}(x_{1},0)=0.\end{cases}
Remark 6.6.

In the case of the Couette flow U(x2)=x2U(x_{2})=x_{2}, assumption (𝐇\mathbf{H}) is satisfied. Obviously Ωc=ω0\Omega^{c}=\omega_{0}, which in fact gives the whole linearized vorticity ω(t,x)=ω0(x1x2t,x2)\omega(t,x)=\omega_{0}(x_{1}-x_{2}t,x_{2}) and the leading asymptotic terms of tv1tv_{1} and x22v2\partial_{x_{2}}^{2}v_{2}. However, t2v2t^{2}v_{2} does also include contributions ΛT\Lambda_{T} and ΛB\Lambda_{B} from the top and bottom boundaries. These asymptotic leading order terms are essentially same as those obtained in [20] (after simplifications of (5.1) in Lemma 5.1 there), see also Lemma 3 in [40].

Acknowledgement

The second author would like to thank Zhiwu Lin, Hao Jia, and Zhifei Zhang for helpful discussions during the completion of the paper.

References

  • [1] V. I. Arnold and B. A. Khesin. Topological methods in hydrodynamics, volume 125 of Applied Mathematical Sciences. Springer-Verlag, New York, 1998.
  • [2] J. Bedrossian, M. Coti Zelati, and V. Vicol. Vortex axisymmetrization, inviscid damping, and vorticity depletion in the linearized 2D Euler equations. Ann. PDE, 5(1):Paper No. 4, 192, 2019.
  • [3] J. Bedrossian and N. Masmoudi. Inviscid damping and the asymptotic stability of planar shear flows in the 2D Euler equations. Publ. Math. Inst. Hautes Études Sci., 122:195–300, 2015.
  • [4] F. Bouchet and H. Morita. Large time behavior and asymptotic stability of the 2D Euler and linearized Euler equations. Phys. D, 239(12):948–966, 2010.
  • [5] O. Buhler, J. Shatah, S. Walsh, and C. Zeng. On the wind generation of water waves. Arch. Ration. Mech. Anal., 222:827 – 878, 2016.
  • [6] K. M. Case. Stability of inviscid plane Couette flow. Phys. Fluids, 3:143–148, 1960.
  • [7] P. Drazin and W. Reid. Hydrodynamic stability. Cambridge: Cambridge University Press., 2004.
  • [8] R. Fjø rtoft. Application of integral theorems in deriving criteria of stability for laminar flows and for the baroclinic circular vortex. Geofys. Publ. Norske Vid.-Akad. Oslo, 17(6):52, 1950.
  • [9] S. Friedlander and L. Howard. Instability in parallel flows revisited. Stud. Appl. Math., 101(1):1–21, 1998.
  • [10] S. Friedlander, W. Strauss, and M. Vishik. Nonlinear instability in an ideal fluid. Ann. Inst. H. Poincaré Anal. Non Linéaire, 14(2):187–209, 1997.
  • [11] E. Grenier. On the nonlinear instability of euler and prandtl equations. Communications on Pure and Applied Mathematics, 53:1067–1091, 2000.
  • [12] E. Grenier, T. T. Nguyen, F. Rousset, and A. Soffer. Linear inviscid damping and enhanced viscous dissipation of shear flows by using the conjugate operator method. J. Funct. Anal., 278(3):108339, 27, 2020.
  • [13] L. Howard. Note on a paper of john w. miles. J. Fluid. Mech., 10:509–512, 1961.
  • [14] V. Hur and Z. Lin. Unstable surface waves in running water. Comm. Math Phys., 282:733 – 796, 2008.
  • [15] V. M. Hur and Z. Lin. Erratum to: Unstable surface waves in running water [mr2426143]. Comm. Math. Phys., 318(3):857–861, 2013.
  • [16] A. D. Ionescu and H. Jia. Inviscid damping near the Couette flow in a channel. Comm. Math. Phys., 374(3):2015–2096, 2020.
  • [17] A. D. Ionescu and H. Jia. Nonlinear inviscid damping near monotonic shear flows. arXiv:2001.03087, 2020.
  • [18] A. D. Ionescu and H. Jia. Linear vortex symmetrization: the spectral density function. arXiv:2109.12815, 2021.
  • [19] H. Jia. Linear inviscid damping in Gevrey spaces. Arch. Ration. Mech. Anal., 235(2):1327–1355, 2020.
  • [20] H. Jia. Linear inviscid damping near monotone shear flows. SIAM J. Math. Anal., 52:623–652, 2020.
  • [21] C. C. Lin. The theory of hydrodynamic stability. Cambridge, at the University Press, 1955.
  • [22] Z. Lin. Instability of some ideal plane flows. SIAM J. Math. Anal., 35(2):318–356, 2003.
  • [23] Z. Lin. Some stability and instability criteria for ideal plane flows. Commun.Math.Phys., 246:87–112, 2004.
  • [24] Z. Lin and C. Zeng. Inviscid dynamical structures near couette flow. Arch Rational Mech Anal 200., page 1075–1097, 2011.
  • [25] Z. Lin and C. Zeng. Unstable manifolds of euler equations. Communications on Pure and Applied Mathematics, 66:1803–1836, 2013.
  • [26] C. Marchioro and M. Pulvirenti. Mathematical theory of incompressible nonviscous fluids, volume 96 of Applied Mathematical Sciences. Springer-Verlag, New York, 1994.
  • [27] N. Masmoudi and W. Zhao. Nonlinear inviscid damping for a class of monotone shear flows in finite channel. arXiv:2001.08564, 2020.
  • [28] J. Miles. On the generation of surface waves by shear flows. J. Fluid Mech., 3:185–204, 1957.
  • [29] W. M. F. Orr. Stability and instability of steady motions of a perfect liquid. Proc. Ir. Acad. Sect. A, Math Astron. Phys. Sci., 27:9–66, 1907.
  • [30] L. Rayleigh. On the stability or instability of certain fluid motions. Pro. London Math. Soc., 9:57–70, 1880.
  • [31] M. Renardy and Y. Renardy. On the stability of inviscid parallel shear flows with a free surface. Math. Fluid Mech., 15:129–137, 2013.
  • [32] J. Shatah and C. Zeng. Geometry and a priori estimates for free boundary problems of the Euler equation. Comm. Pure Appl. Math., 61(5):698–744, 2008.
  • [33] S. A. Stepin. The nonselfadjoint Friedrichs model in the theory of hydrodynamic stability. Funktsional. Anal. i Prilozhen., 29(2):22–35, 95, 1995.
  • [34] W. Tollmien. Ein allgemeines kriterium der instabititat laminarer geschwindigkeitsverteilungen. Nachr. Ges. Wiss. Gottingen Math. Phys., 50:79–114, 1935.
  • [35] D. Wei, Z. Zhang, and W. Zhao. Linear inviscid damping for a class of monotone shear flow in sobolev spaces. Commun. on pure and applied math, 71:617–687, 2018.
  • [36] D. Wei, Z. Zhang, and W. Zhao. Linear inviscid damping and vorticity depletion for shear flows. Ann. PDE, 5(1):Paper No. 3, 101, 2019.
  • [37] D. Wei, Z. Zhang, and W. Zhao. Linear inviscid damping and enhanced dissipation for the Kolmogorov flow. Adv. Math., 362:106963, 103, 2020.
  • [38] D. Wei, Z. Zhang, and H. Zhu. Linear inviscid damping for the β\beta-plane equation. Comm. Math. Phys., 375(1):127–174, 2020.
  • [39] C.-S. Yih. Surface waves in flowing water. J. Fluid. Mech., 51:209–220, 1972.
  • [40] C. Zillinger. Linear inviscid damping for monotone shear flows in a finite periodic channel, boundary effects, blow-up and critical Sobolev regularity. Arch. Ration. Mech. Anal., 221(3):1449–1509, 2016.
  • [41] C. Zillinger. Linear inviscid damping for monotone shear flows. Trans. Amer. Math. Soc., 369(12):8799–8855, 2017.