Capillary gravity water waves linearized at monotone shear flows: eigenvalues and inviscid damping
Abstract.
We consider the 2D capillary gravity waves of finite depth linearized at a monotonic shear flow . 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 of the horizontal variable , we first prove that the linearized capillary gravity wave has two branches of eigenvalues , where the wave speeds for are asymptotic to those of the linear irrotational capillary gravity waves. Under the additional assumption , we obtain the complete continuation of these two branches, which are all the eigenvalues in this (and some other) case(s). In particular, could bifurcate into unstable eigenvalues at . The bifurcation of unstable eigenvalues from inflection values of is also obtained. Assuming there are no embedded eigenvalues for any wave number , the linearized velocity and surface profile are considered in both periodic-in- and cases. Each solution can be split into and whose -th Fourier modes in correspond to the eigenvalues and the continuous spectra of the wave number , respectively. The component is governed by the dispersion relation in the case of . The other component satisfies the linear inviscid damping as fast as and as . Furthermore, additional , , decay of and is obtained after leading asymptotic terms are singled out, which are in various forms of -dependent translations in of certain functions of .
1. Introduction
Consider the two dimensional capillary gravity water waves in the moving domain of finite depth
or
The free surface is given by . For , let denote the fluid velocity and the pressure. They satisfy the free boundary problem of the incompressible Euler equation:
(1.1a) | ||||
(1.1b) | ||||
(1.1c) | ||||
(1.1d) | ||||
(1.1e) |
where , is the mean curvature of at which corresponds to the surface tension, 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) |
Our primary goal in this paper is to analyze the capillary gravity water wave system linearized at a monotone shear flow satisfying
(H) |
Remark 1.1.
Due to the symmetry of horizontal reflection
the case of is completely identical except the signs of 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 , 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 given in (1.2) satisfied by the linearized solutions which we denote by . Let be a one-parameter family of solutions of (1.1) with . Differentiating the Euler equation (1.1a) and (1.1b) with respect to and then evaluating it at yield
(1.3a) | |||
Taking its divergence and also evaluating the above linearized Euler equation at , we obtain | |||
(1.3b) | |||
From the kinematic boundary condition (1.1c), we have | |||
(1.3c) | |||
Finally differentiating (1.1d), where the left side is , and using , we obtain | |||
(1.3d) |
The above (1.3a – 1.3d) form the linearization of the capillary gravity water wave problem (1.1) at the shear flow with initial values . In fact it can be reduced to an evolutionary problem of the unknowns , while 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) |
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.
Eigenvalues. Since the variable coefficients in the linearized Euler system depend only on , the subspace of the -th Fourier mode is invariant under the linear evolution for any . Hence it is a common practice to seek eigenvalues and eigenfunctions in the form of
(1.5) |
where apparently the eigenvalues take the form with the wave speed . The linear system is spectrally unstable if there exist such with , which appear in conjugate pairs. Solutions in the above form with 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 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 and do not exist for .
- •
-
•
Howard’s Semicircle Theorem [13]: for any , eigenvalues exist only with in the disk
(1.6) - •
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. and in (1.1)) at shear flows (see [39, 14, 31] etc.) that: a.) assuming and on , there are no singular neutral modes in (i.e. solution in the form of (1.5) with ); 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 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 , where , and may be limiting singular neutral modes; and b.) there are non-singular neutral modes, i.e. . 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.
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 . In 1907, Orr [29] observed that the linearized vertical velocity tends to zero as . Under the assumption , , which removes the shear flow component of the linear solutions through an invariant splitting, explicit calculations (see, e.g., [6, 24]) yield, as ,
(1.7) |
where 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 decay of the stream function was proved in [33] and then the (1.7) type decay in [40, 41] under a smallness assumption of (also in order for the decay of ). 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 decay if 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 .
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.
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 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 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 (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 is expected to include the following.
a.) Eigenvalues for wave numbers . Like matrices, eigenvalues (or equivalently, singular and non-singular modes) of (1.3) correspond to roots of a “characteristic" function (the defined in (4.1)) analytic in and . In contrast to the linearized Euler equation on a fixed domain where no eigenvalues exist for any large wave number , 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 for each wave number . These branches behave rather differently compared to the linearized gravity waves since the surface tension is dominant for . 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 , the boundary of the domain of analyticity of the characteristic function. The bifurcation analysis for near , more specifically near inflection values of which possibly generates the instability (compare with the gravity wave case [14, 31, 15]) and , would require very careful study of the dependence of the solutions to the classical Rayleigh equation (3.1) on the singular parameter (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 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 is mostly related to the surface motion and dispersive (possibly with some unstable modes), while the other component 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 . 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 for ? 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 and on , then the following hold.
-
(1)
There exists such that for any with , there are no singular modes and exactly two non-singular modes and which correspond to semi-simple eigenvalues . Moreover,
-
(a)
are even and analytic in and can be extended for all with ;
-
(b)
;
-
(c)
if is not a singular mode for any , then can also be extended to be even and analytic in all with ; and
-
(d)
if singular modes do not exist , then are the only non-singular modes of (1.3) which is linearly stable.
-
(e)
for , (and as well if it can be extended for all ) has either none or exactly one critical point depending on whether (4.18) holds. In the latter case, the critical point is non-degenerate.
-
(a)
-
(2)
There exists depending only on and such that the following hold.
-
(a)
If , then the non-singular modes can also be extended to be even and analytic in all and for ;
-
(b)
if and only if
(1.8)
-
(a)
-
(3)
If on is also satisfied, then the following hold with the given in the above statement (2).
-
(a)
The only possible singular mode is .
-
(b)
If then there are no singular modes, are the only non-singular modes, and thus (1.3) is linearly stable.
-
(c)
If and , then there exists such that can be extended as an even function (for any ) for all . Moreover is analytic for all , and . For each , are the only singular or non-singular modes and thus (1.3) is spectrally stable.
-
(d)
If and , then there exist such that we have the following.
-
(i)
Assume on , then can be extended as an even function (for any ) for all and analytic except at such that
Moreover, for each , all singular and non-singular modes are exactly , , as well as if . Consequently, (1.3) is spectrally unstable iff A.) or B.) and there exists such that .
-
(ii)
Assume on , then can be extended as an even real valued function (for any ) for , analytic in if , and . Moreover, all singular and non-singular modes are exactly and , where the latter is only for , and (1.3) is spectrally stable.
-
(i)
-
(a)
-
(4)
If and for some . Let .
-
(a)
There exists such that for any , there exists a unique unique in such that is a singular neutral mode for .
-
(b)
If , , and is a singular neutral mode for some , then, under a non-degenerate condition (verified for the one obtained in (4a) for small ), there exist unstable modes near for close to on one side of .
-
(a)
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 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 .
d.) Condition (4.18) is also directly on , , and , but less explicit, and we leave it in Subsection 4.1.
The existence of the unbounded branches of non-singular neutral modes are in contrast to the gravity waves or the Euler equation on fixed channels. The geometric multiplicity of occurs only among different . These temporal frequencies are asymptotic to those (see (2.4)) of the irrotational capillary gravity waves linearized at zero. Moreover, after normalizing the norm of the component of the eigenfunction to be , the and differences in the and components, respectively, between the eigenfunctions of (1.3) and the linearized irrotational capillary gravity waves are of the order as (see Remark 6.1). In the case (happening only if (1.8) is not satisfied) where the branch reaches , 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 determines whether becomes unstable or disappears at .
The spectral stability in the case 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 is never a singular mode, just like the Rayleigh’s theorem in the channel flow case the change of sign of turns out to be necessary for the existence of interior singular modes, which is also sufficient if . In the contrast this may not be sufficient if the stabilizing gravity and surface tension are strong, see Remark 4.4.
In the following outline of the linear inviscid damping results, if , while if . The initial velocity is always assumed to satisfy or for some ’s and ’s and similarly . To avoid too much technicality, here we skip the detailed assumptions on their regularity in , but focus on that of only. Precise statements are given in Theorems 2.1 and 2.2. The following is always for .
Main results on inviscid damping. Assume (H) and there are no singular modes for all , then any solution to (1.3) can be decomposed into and , where belongs to the invariant subspace generated by the eigenfunction of the non-singular modes for all , while and its vorticity satisfy the following estimates.
-
(1)
If the initial vorticity , then for any , and .
-
(2)
If , then , for any . Moreover there exists such that for any ,
-
(3)
If , then there exist such that for any ,
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 , 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 for each .
The component 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 , certain stronger decay in (for long waves) is also assumed on the initial values, see Theorem 2.2 and Remark 2.2. Additional to the above bounds, derivatives-in- estimates are given in Theorems 2.1 and 2.2 as well which also imply pointwise-in- decay. Compared with (1.7), these additional estimates represent an improvement of decay of roughly an order of (after appropriate -dependent translations in of some asymptotic leading terms are identified and singled out in the cases of , , etc.). For the Euler equation in a fixed channel (1.4), a.) when , the 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 ) of these leading terms , , and 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 and . 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 . The error estimates in addition to these leading asymptotic terms also justify that the estimates of and in (1.7) are optimal. The precise asymptotic leading terms could be useful for further analysis.
The component 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 for all non-singular modes of the -th Fourier modes in . According to the above spectral analysis, this dispersion relation is asymptotic to that of the linear irrotational capillary gravity wave for . In the case of , in the absence of singular modes, all non-singular modes are given by which are neutral/stable. The dispersion of implies that it should decay if , but at a slower rate. Hence the dynamics of (1.3) has two layers: faster inviscid decay of leaves the remaining decaying at a slower rate due to the dispersion like a linear irrotational capillary gravity wave.
In the periodic-in- case, as the non-existence of singular modes is assumed only for , there can still be other non-singular modes besides which may have bifurcated from inflection values of (as well as unstable modes from ) at some . 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 to (1.3), resulting in decoupled systems for each wave number . The problem can be further reduced to the evolution of , the Fourier transform of . The Laplacian transform111Working on the Laplace transform of the unknowns is essentially equivalent to analyzing the resolvent of the linear operator in (1.3). of , where 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 is also carefully tracked.
We first study the Rayleigh equation away from the singularity where , . Near the singularity where , 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 are expressed in a form involving the explicit and the heaviside function with coefficients smooth in .
We focus on a pair of fundamental solutions to the homogeneous Rayleigh equation which satisfy the corresponding homogeneous boundary conditions (2.11b)-(2.11c) in (2.11) at , respectively (boundary condition (2.11c) reflects the free boundary setting). For , we establish a.) their a priori bounds; b.) the convergence to their limits as ; and c.) the smoothness of , particularly, in . Recall , we prove is in except at . Due to the analyticity of in with , the estimates of also yield those of for . Eventually general solutions to the non-homogeneous boundary value problem (2.11) of the Rayleigh equation are expressed using . Finally, the quantity 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 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 for , followed by an argument based on analytic continuation and index calculation. Bifurcations may occur at inflection values of and particularly subtle at , which are on the boundary of the analyticity of the bifurcation equation . The regularity obtained in Section 3 implies, when restricted to , near and near . This makes the bifurcation analysis possible near and much easier even in the relatively classical case near inflection values of .
Among the results in Theorem 1.1, in statement (1), are obtained for large in Lemma 4.2(3) with more detailed estimates, the extension of in Corollary 4.3.1, and the semi-simplicity of the eigenvalues 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 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 is proved in Lemma 4.6, along with the existence of and/or 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 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 obtained in Section 3. We establish the basic a priori and convergence (as ) 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 in (in Subsection 5.2). Since singularity occurs in the Rayleigh equation along , , , behaves badly there. Instead we apply a differential operator to the Rayleigh system (2.11) which differentiates along the direction of , hence 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 , and , with the most singular terms identified.
The splitting and the linear inviscid damping estimates of solutions to the linearized capillary gravity waves (1.3) are obtained in Section 6. While the vorticity 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 , the Laplace transform of , which is estimated in Section 5. Unlike e.g. [35, 20], technically we do not immediately push the contour integral (in ) of the inverse Laplace transform to the limit spectra set , but first keep it along the boundary of a small neighborhood of it in the complex plane. This allows easy integration by parts in to establish the decay estimates in after the leading asymptotic terms are obtained by applying the Cauchy integral theorem to the most singular terms of , . In fact, in deriving the decay estimates of , , , and 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 . Similarly, while the asymptotic leading order terms of , , and are all generated by the asymptotic vorticity , that of involves two additional functions and 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, denotes a generic constant which might change from line to line, but always independent of , , and ; the delta function; (or ) the principle value (or the principle value with respect to variable etc.). The Japanese bracket is adopted. For close to , denotes after some extension of . We always denote and .
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 is determined by and (see (1.3b) and (1.3d)), so very often we shall focus only on and .
2.1. A brief motivational study of the Couette flow
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])
where
and satisfies
In particular, the rotational part can almost be determined by the vorticity in the same way as in the Euler equation (1.4) in the fixed channel with slip boundary condition
(2.1) |
where is a constant and is the inverse Laplacian in the 2-d region (-periodic in or ) under the zero Dirichlet boundary condition along . In the periodic-in- case, the constant 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) |
Linearizing it at , which is the vorticity of the Couette flow, yields the linearized vorticity
expressed in term of its initial value . Since component of the linearized capillary gravity waves (1.3) at the Coutte flow corresponds to the divergence free velocity field determined by its vorticity 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- case) implies
II. Singular and non-singular modes. Unlike the Euler equation in a fixed channel, there is the additional surface profile coupled to the irrotational part of the velocity, which may not decay. In fact, for any , let
where “c.c.” denotes “complex conjugates” and
(2.3) |
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 (1.3a–1.3d) for the Couette flow. Therefore are eigenvalues of the linearized systems associated with the above eigenfunctions. As these solutions do not grow or decay as , are neutral modes.
It is worth paying slightly closer attention to the wave speed and the function , all of which are even in . We make the following observations.
-
(1)
, so for the dispersion relation is asymptotic to those of the irrotational capillary gravity waves linearized at zero solution (system (1.3) with and ) given by with
(2.4) which can be obtained through direct calculation based on the Fourier transform.
-
(2)
for all , so it is a branch of non-singular neutral modes, namely, wave speeds outside , the range of .
-
(3)
While in (2.3) as seen in the above observation (1) for large , it can happen for and thus becomes singular modes (those in the range of ).
-
(4)
Since with “” achieved at , for , and thus both are non-singular modes. Moreover, one may verify for all if . In particular, in the case of , this implies that a.) the dispersion relations 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 of (2.3) and of (2.4) if they have the same temporal frequency . Moreover, would turn out to the only eigenvalues for the linearization at the Couette flow for (see Proposition 4.4(2)).
Generalization of the linear analysis to a general shear flow ? 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 including at least the following.
-
•
The crucial function defined in (2.3) which determines the wave speed and consequently the dispersion relations, while analytic for , may become rather singular for approaching . What regularity of can one expect?
-
•
Consequently, if a branch of non-singular modes approaches , possibly very subtle bifurcations may occur at the boundary of analyticity of . Can instability be generated?
- •
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 . 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 to (1.3) can be decomposed into the component corresponding to the non-singular modes and to the continuous spectra due to . This splitting is invariant under (1.3) and is of the order (and the vertical component ) as . 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 , , and , 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 , the fundamental solutions to the homogeneous Rayleigh equation, and also by the Laplace transform of . The results are stated for the cases of and separately in the following.
Theorem 2.1.
(Inviscid damping: periodic-in- case) Suppose . Assume , , on , and there are no singular modes (see (4.9) and Lemma 4.1(5)) for any . For any , , and , there exists depending only on , , , and , such that, for any , integer , and solution of (1.3) with initial value and the corresponding initial vorticity , there exist unique solutions , , to (1.3) and -periodic-in- functions , , and determined by linearly such that
and the following hold.
-
(1)
satisfy the following estimates
if ,
and if, in addition, , then
-
(2)
and , , satisfy
where denotes the Fourier transform of a function with respect to . Moreover, , , satisfy and
(2.5a) (2.5b) If , then - (3)
- (4)
Remark 2.1.
1.) Observe that, for any compactly supported smooth function , , it holds
which implies
By the standard density argument, this inequality also holds for any function .
Hence the above estimates in statement (1) also imply various pointwise-in- decay of as .
2.) The function 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 is small (by Theorem 1.1(1) as is large) or if 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), and the elliptic boundary value problem (2.5b) has a unique solution , while (2.5a) has a unique solution 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, and exhibit logarithmic singularity at and , respectively. In particular, vanishes if the initial vorticity , while if . In this paper as we focus on the damping estimates with additional decay of after the leading order terms are singled out, we adopted based norms to somewhat simplify the calculations. If the decay in other or 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 in in the theorem may not be close to optimal, particularly when and are away from , see Remark 6.2(b) as well as Remark 3.8. Moreover, the small may not be necessary, see e.g. [36, 38] in the fixed boundary case. The assumptions on the more essential regularity of in are optimal even in the existing results in the fixed boundary case.
In the estimates of the component which are superpositions of eigenfunctions, the possible exponential growth (if ) is caused by unstable modes, where is the maximum real parts of the eigenvalues and is the maximum multiplicity of those eigenvalues of the maximal real parts. Due to Theorem 1.1(1), growth does not occur for . It is also worth pointing out that, taking , the the regularity of is order better than that of restricted to the surface , which is consistent with the regularity results of nonlinear capillary gravity waves in the existing literature. In the contrast, the regularity requirement on in the damping estimates of is stronger than that of . Compared with the above example of the linearization at the Couette flow, conceptually these phenomena is due to the fact that the component is mainly the rotational part of the solution which depends on the vorticity more heavily, while are more like the irrotational part.
The estimate in statement (3) at implies the boundedness of the projection onto , whose kernal is . Some more detailed information of this projection can be found in Lemma 6.2 and 6.10. In fact the subspace is generated by the eigenfunction of all non-singular modes for all .
The inviscid decay estimates in the case of is slightly subtle due to the presence of small wave number . We use similar notations in the following theorem.
Theorem 2.2.
(Inviscid damping: case) Suppose . Assume , , on , and there are no singular modes (see (4.9) and Lemma 4.1(5)) for any . For any , , and , there exists depending only on , , , and , such that, for any , integers , and solution of (1.3) with initial value , there exist solutions , , to (1.3) and functions , , and determined by linearly such that
and the following hold.
-
(1)
satisfy the following estimates
if , then
and if, in addition, , then
- (2)
-
(3)
For any and the following integer ,
-
(4)
Let
then they are invariant closed subspaces of under (1.3). Moreover (1.3) is also well-posed in the completion of . If, in addition, (4.18) holds for both and , then (1.3) restricted to the completion of , or with , 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 and , the , , applied to the initial values indicates some stronger decay assumptions for wave number .
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 of both algebraic and geometric multiplicity two, hence there is no growth at all. The conjugacy of the dynamics of 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 in , which in turn determines , , and . We then apply the Laplace transform in to obtain a non-homogeneous boundary value problem of the well-known Rayleigh equation in with a non-homogeneous Robin type boundary condition at due the boundary conditions at the free boundary. The main analysis will focus on the Rayleigh equation.
Consider the Fourier transforms of the unknowns in
in the case of and
in the case of , where we skipped the variable . The Fourier transform of the linearized system (1.3) takes the form
(2.6) |
where ′ denotes the derivative with respect to as in the rest of the paper. Due to the divergence free condition on and the boundary conditions, it is easy to see
(2.7) |
For , can also be determined by using the divergence free condition, by the third equation of (2.6), while by and by solving the elliptic boundary value problem. So we shall mainly focus on .
Combining the equation of acted by and the one of acted by , we obtain
(2.8a) | |||
which is actually the linearized transport equation of the vorticity (as defined in (2.1)) | |||
in its Fourier transform. In addition to the above equation, we need its boundary information to completely determine . Applying to the first equation of (2.6), then evaluating at , and using the equation of , we have | |||
Finally applying to the above equation and using the third equation of (2.6), we obtain | |||
(2.8b) |
where we also included the boundary value of at .
To analyze the evolutionary problem, we apply the Laplace transform to the unknowns
(2.9) |
An often used change of variable for is
(2.10) |
with and being the real and imaginary parts. From (2.8), our main unknown satisfies the following non-homogeneous Rayleigh equation
(2.11a) | ||||
where is the Fourier transform of the initial vorticity, with the obvious boundary condition | ||||
(2.11b) | ||||
Here we skipped the and variables of . Similarly, the Laplace transform applied to the boundary equation (2.8b) implies | ||||
Therefore we obtain | ||||
(2.11c) |
The last boundary condition can be viewed as determining the dispersion relation which is highly nonlocal. The Laplace transforms of and of and can be recovered from the divergence free condition and the third equation of (2.6)
(2.12) |
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 and , so it has a unique solution analytic in and except for those for which the corresponding homogeneous system of (2.11), where and , has non-trivial solutions. Such singular also give the eigenvalues of (2.11) in the form of . In fact we have the following lemma.
Lemma 2.3.
Proof.
On the one hand, it is straight forward to verify that the above , , and satisfy (1.3c), (1.3d), , and . The Poisson equation of in (1.3b) is a consequence of the linearized Euler equation in (1.3a), the equation of which is also easily verified. Hence we only need to consider the equation in (1.3a). In fact, that equation holds for the above if
The -derivative of this function is equal to due to the Rayleigh equation (2.11a) and its boundary value equal is to at due to the boundary condition (2.11c).
On the other hand, suppose is a solution to (1.3) in the form of (1.5) with and . Equation (2.8a) implies that 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 have to be homogeneous solutions to (2.11). Subsequently, is obtained from , from the third equation in (2.6), and from the equation in (2.6) along with its boundary value at . ∎
Definition 2.1.
is a non-singular mode if and there exists a non-trivial solution to the corresponding homogeneous problem of (2.11) (thus also yields a solution to (1.3) in the form of (1.5)). is a singular mode if and there exists a solution to
(2.14) |
along with the corresponding homogeneous boundary conditions of (2.11b–2.11c). (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 . 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) |
where is a real number so that is analytic in the region 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 and the non-singular modes of (1.3). Assuming there is no singular modes in , we shall eventually obtain the decay in of the component of the linear solution corresponding to the integral along the contour surrounding by integration by parts in .
3. Analysis of the Rayleigh equation
In this section, we shall thoroughly analyze the homogeneous Rayleigh equation
(3.1) |
where
Throughout this section (except for some lemmas in Subsection 3.6), we assume
(3.2) |
As pointed out in the introduction, due to the symmetry of the reflection in variable, the case of can be reduced to the above one. Hence all results under (3.2) hold for all uniformly monotonic , namely those satisfying on .
To some extent, we will also consider the non-homogeneous Rayleigh equation
(3.3) |
More detailed forms and conditions of 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 with sufficient regularity, but carefully tracking its norms involved in the estimates.
The solutions to the Rayleigh equation (3.1) are obviously even in and thus will be assumed mostly. Similarly complex conjugate of solutions also solve (3.1) with replaced by , so we will restrict our consideration to . We have to consider the cases of away from , near , and then finally , separately. Due to small scales in created by , the dependence of the estimates on will be carefully tracked.
Recall . For technical convenience we extend to be a function on a neighborhood of , where
(3.4) |
such that, on ,
(3.5) |
In the analysis of the most singular case of close to the range , we let be such that
(3.6) |
We also extend the non-homogeneous term for 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 is bounded from below. For not so small , we first transform the homogeneous Rayleigh equation (3.1) into a system of first order (complex valued) ODEs. Let
and then (3.1) takes the form of the coupled equations
(3.7) |
Lemma 3.1.
There exists depending only on , and , such that for any , , and satisfying
(3.8) |
and any solution to (3.7) with
(3.9) |
it holds, for , and
(3.10) |
Moreover, for any solution with
(3.11) |
we have, for , and
(3.12) |
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 , the standard ODE theory implies that there are two solutions behaving like corresponding to the Lyapunov exponents close to , then the decaying solution may not satisfy (3.10) with uniform in .
Proof.
We start with the observation of a simple consequence of (3.8). Namely, one may compute straight forwardly
(3.13) |
This monotonicity along with boundary conditions yields an order relation between which can be used to control terms in (3.7).
We shall focus on the case under assumption (3.9), which ensures
(3.14) |
By factorizing on the right side of (3.7), its solutions satisfy
(3.15) |
If , let be defined as in (3.6) and we use (3.8) to estimate
where the last equality is the exact integral. If , then the numerator in the logarithm is greater than the denominator. Applying the triangle inequality to , and , we obtain
If , multiplying the top and bottom of the quotient by their conjugates and proceeding much as in the previous case, we have
Finally, in the case , by splitting the interval at and applying the above estimates on the two subintervals, we obtain
Therefore the desired estimate (3.10) on follows from (3.15) and (3.14) and
as is bounded uniformly in all . If , one can bound by which is also bounded for all . If , then is bounded by . If , then
Therefore in both cases we have
Turning attention to , from the variation of parameter formula, we have
(3.16) |
which along with (3.8), (3.10) for , and (3.14), implies
The desired estimate on follows from the above inequality on . The estimates on with initial condition 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.
Proof.
We first consider the special solution to the homogeneous (3.1) satisfying (3.17) with , namely, with the initial values
whose corresponding form in terms of with initial values satisfies the assumptions of Lemma 3.1. On the one hand, for , it holds
while, for , we have
Therefore Lemma 3.1 and imply
Recovering and from , we obtain the desired estimates in the case of under the additional assumption .
In the following we prove the estimates for a homogeneous solution to (3.1) under (3.17) with general . Let and be solution to (3.1) with initial values
Clearly and satisfy the above estimates with and , respectively, and
Therefore, for ,
In the above summation, all the hyperbolic trigonometric combinations without or 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 and . We have
where the last inequality was obtained by considering the two possible cases of spearately. The inequality on can be obtained similarly as
and thus
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 with to the non-homogeneous Rayleigh equation (3.3) as
where is the fundamental matrix of the homogeneous equation (3.1) with initial value . Therefore,
where is the solution to (3.1) whose initial value is given by and . The desired estimates follow from applying the above estimates in the homogeneous case with and . ∎
Practically the above estimates are more effective for bounded from below. To end this subsection, we give the following simple estimate of the Rayleigh equation for bounded from above, which compares to the free solution (where the term is removed)
Here is understood.
Lemma 3.3.
For any , there exists depending only on , , and such that for any , , satisfying
and any solution to (3.3), it holds
Proof.
The proof is based on some straight forward elementary argument and we shall only outline it. Let . We can write the solution using the variation of constant formula
It implies
and the estimates on and follow immediately from the Gronwall inequality. ∎
3.2. Rayleigh equation near singularity and its convergence as
In the rest of the section, we shall mostly focus on the case when is small, so
(3.20) |
will always be assumed, while the domains of and have been extended to . Due to complex conjugacy, we only need to consider . In particular, if , the strong singularity in (3.1) will lead to even if . Even though some estimates are stated for , most of the inequalities are mostly uniform as and thus hold for the limits.
In order to obtain estimates uniform in , rescale
(3.21) |
where satisfies (3.20) as well as in the above. Equation (3.1) becomes
(3.22) |
where
We shall consider this ODE on intervals such that
(3.23) |
is that is well-defined when . As , one would naturally expect to converge to solutions to
(3.24) |
However, this limit equation becomes singular at and conditions have to be specified there.
Fundamental matrix of the homogeneous Rayleigh equation. Its construction is adapted from the one used in [5]. Let
(3.25) |
where, for ,
(3.26) |
and the remainder of is given by
(3.27) |
It is not hard to see that is in and and in . We often skip writing the explicit dependence on those variables other than . Denote
(3.28) |
where we note that the integrand of the imaginary part of converges to a delta mass as and produces a jump in at (see Lemma 3.4 in the below). Let be a matrix given by
(3.29) |
and
(3.30) |
It is worth pointing out that is real for and imaginary for . To keep the notations simple we often skip the arguments other than . In the following lemma we collect some basic estimates of and where we often bound the singularity in by , , for simplicity.
Lemma 3.4.
Remark 3.1.
Even though is assumed in the above and the remaining statements in this and the next subsections, as is independent of in a priori estimates and thus they hold even as .
Expression (3.34) essentially is the variation of parameter formula including the fundamental matrix of the Rayleigh equation. Due to , it is possible to extend the definition of to include all , but its bound would be non-uniform in for .
Proof.
Since has a logarithmic singularity at the worst (even for ), is obviously well-defined. The zero trace value of the coefficient matrix in (3.29) yields . The form (3.34) of general solutions of (3.22) for follows from straightforward verifications.
Equation (3.29) implies
where is the operator norm of the coefficient matrix. From Gronwall inequality, we obtain
It is clear from the definition of that
The definition of , the boundedness of , and the estimate on imply, for ,
where is a generic constant determined by and and the Hölder inequality was used to obtain , for any . The desired estimate in (3.31) on follows immediately which along with in turn yields the estimate on .
The definition of implies
Regarding the imaginary part of , we observe
where we used the smoothness of and in . It implies
(3.35) |
and thus
The error estimate (3.32) follows consequently.
Proceeding to consider where , we have
Recalling that is the elementary fundamental matrix of the above corresponding homogeneous ODE system, the variation of parameter formula implies
The second desired upper bound in (3.33) of follows from direct estimating the above integral without using the Hölder inequality. For the first upper bound there we use, for any ,
(3.36) |
which can be verified by straight forward computation. The proof of the lemma is complete. ∎
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 in terms of and . Let denote
Corollary 3.4.1.
Proof.
The estimates on follows from straight forward calculation based on (3.35) and the bound on given in Lemma 3.4 and we omit the details.
Regrading , let denote the entries of . Using Lemma 3.4 where the estimates are uniform in , we have
Since
(3.37) |
the upper on follows accordingly.
To derive the estimate on , we notice and the desired estimate follows from integrating using (3.36). ∎
Remark 3.2.
The above estimates imply, that for any solution to (3.22)
(3.38) | |||
(3.39) |
The following lemma gives another estimate of in terms of some initial value which we shall use mainly for away from .
Lemma 3.5.
Proof.
We shall first estimate based on and then apply Corollary 3.4.1. From (3.34) and which allows us to write explicitly, we have
(3.41) |
Using Lemma 3.4, one may estimate
(3.42) |
(3.43) |
where we also used (3.35). Combining these inequalities and Corollary 3.4.1, we obtain
where . This yields inequality (3.40b) of . The estimate of is obtained through integrating that of . ∎
Convergence estimates as . As , 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 , , and replaced by , , and .
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 can be studied in the framework in this section, we shall just focus on the homogeneous case, i.e. with , and leave the non-homogeneous one to Section 5. In fact, (3.28) and Lemma 3.4 imply that, as , would converge to a Hölder continuous limit, while develops a jump proportional to and a logarithmic singularity at . More precisely, the limit of solutions should (see the proposition in the below) satisfy the Rayleigh equation (3.24) with for and satisfy at ,
(3.44) |
It is worth pointing out that the existence of the limit of does not imply a simple jump discontinuity of , which actually has a symmetric logarithmic singularity. In the distribution sense, the limit homogeneous Rayleigh equation (3.24) (with ) along with (3.44) can be written as
(3.45) |
Here denotes the delta function of and “” indicate the principle value when the corresponding distributions are applied to test functions of . They occur in only. In terms of the original unknown , the limit of (3.3) as is
(3.46) |
where the subscript indicates the distributions as generalized functions of . For , the parallel results hold except with the complex conjugate. It also means that homogeneous Rayleigh equation takes different limit as .
Lemma 3.6.
Proof.
On and , (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 . The continuity of and the estimates of and in Lemma 3.4 immediately yields . Finally follows from the jump condition of at after writing using (3.41) and again using the estimates of and .
Finally, the continuity of under the assumptions follows from (3.47), the Hölder continuity of , and the logarithmic upper bound of . ∎
The following proposition provides the convergence estimates.
Proposition 3.7.
Remark 3.3.
Proof.
We first work on the error estimates in terms of and . Let
Controlling and by Corollary 3.4.1 ( by (3.38) and (3.39) in particular), where we recall the estimates are uniform in , we have
(3.52) |
where and we also used
This completes the proof of inequality (3.49). The estimate (3.48) on is derived by integrating 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 . From formula (3.41) we have
From (3.37) and Lemma 3.4, one may estimate
where . Therefore we obtain
Applying (3.42) and (3.43) to control in (3.52), we can estimate
Inequality (3.51) is obtained by simplifying the above. In particular, we used
to absorb the term .
3.3. A priori bounds on the two fundamental solutions to the homogeneous Rayleigh equation with
In this subsection, we analyze and derive the basic estimates of of two fixed solutions to the homogeneous equation (3.1) with initial values
(3.53) |
which also depend on parameters and . The initial condition of at 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 and .) As throughout this section, we often skip the arguments rather than . Particularly when working near , we shall continue using the notations introduced in Subsection 3.2, like , etc. The following lemma is standard. Due to conjugacy, we only consider .
Lemma 3.8.
For and , the solutions are even in , analytic in and , and is in . Moreover .
In the next step we give a priori estimates of . In particular, we consider up to three subintervals,
(3.54) |
(3.55) |
Here as in (3.21). Clearly and any of these subintervals may be empty. If , then is considered as for and as for in the statement of the following lemma. The choice of the above constant and the fact ensure
(3.56) |
Lemma 3.9.
For any , there exists depending only on , , and (also on and for the estimates of ), such that, for any , the following hold:
(3.57) |
(3.58) |
for all . Moreover, if , then for all ,
(3.59) |
(3.60) |
If otherwise , then
(3.61) |
(3.62) |
and for ,
(3.63) |
(3.64) |
Remark 3.4.
Even though the lemma assumes , the estimates are uniform in and thus they also hold for the limits of solutions as , while the limits as are the conjugates of those as . Moreover, the constant does not depend on , and in particular, for does not depend on either.
It is possible that as the domain of has been extended. However, the constant in (3.59), (3.60), (3.61), and (3.62) are independent of the extensions of satisfying (3.5).
Proof.
The estimates of can be derived in exactly the same procedure by reversing the direction of the variable . We shall focus on and give a brief description on the argument for afterwards. The cases of close to and away from will be considered differently based on Lemma 3.2 and Proposition 3.7, respectively.
Step 1. Assume . We consider in two cases. The first on is for those larger such that
(3.65) |
where (3.8) is satisfied and Lemma 3.2 is applicable. Observe
(3.66) |
and
(3.67) |
where the last inequality could be derived by considering whether . The same upper bound also holds for . Therefore applying Lemma 3.2 on with and , we immediately obtain the desired estimates (3.57), (3.59), (3.61) on and on , respectively. Otherwise in the case of smaller , the desired estimates follows from Lemma 3.3 with .
Step 2. Assume and otherwise step 1 has completed the proof. In this case, due to (3.56). Let
(3.68) |
which implies
(3.69) |
Therefore results in Subsection 3.2 in the corresponding rescaled variables and given in (3.21) are applicable. Moreover the definition of further yields
Let
Lemma 3.5 (with ) implies that, for any
for any . Moving the term to the right side, we obtain
(3.70) |
Notice that, no matter whether or not, (3.57) and (3.59) are satisfied at due to either the initial condition of or the above step 1. On the one hand, regarding the above first term on the right side, it holds that either if or if , hence this term would only contribute an error term of at most , for any , in the upper bounds. On the other hand, implies that replacing the above , and by would also only produce an error terms of at most in the upper bounds. Therefore we have
(3.71) |
and thus (3.63) follows by letting .
Integrating (3.71) over , we have, for ,
where we used (3.57), , and and the first term of the right side of (3.70) was incorporated into others as remarked just below (3.70). For , we have
while for ,
Therefore we obtain
which proves (3.57) on .
Step 3. Assume , which implies . In this case, surely either and . With (3.57) for and (3.63) for established at , satisfies assumption (3.17) for the interval with , , and .
As in the step 1, for larger so that (3.65) holds, the desired estimates (3.57) and (3.63) in follow directly from (3.66), (3.67), and Lemma 3.2.
For smaller , say, , we express and in terms of , , as in (3.21). Let
Since , otherwise for , it along with (3.56) and implies
Namely, the domain of is contained in . Applying (3.40b) (with ), using , , and
we obtain on . It in turn implies
Estimating Finally, we give a brief sketch of the argument for , for which we proceed from to .
Suppose . The initial values of at satisfy (3.18) with and . For larger so that (3.65) holds, the desired estimates (3.58) and (3.60) in follow directly from (3.66), (3.67), and Lemma 3.2. The estimates for smaller is again a consequence of Lemma 3.3.
Suppose which implies . Inequality (3.70) with replaced by still follows from exactly the same argument, namely, for and any ,
If , then
Otherwise, and thus, for any ,
where (3.58) at was also used. These estimates, along with (3.58) and (3.60) at yield (3.64) on . Inequality (3.58) follows from direct integrating the estimate on , actually without going through the technical argument at the end of step 2 for since the , instead of , is in the upper bound in (3.58).
3.4. Limits of solutions to the homogeneous Rayleigh equation with
Now that the convergence of solutions of the Rayleigh equation as 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 . In this subsection we consider 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 as well as appears only Hölder in , or equivalently in . In the notations given in (3.21) in Subsection 3.2, we first prove the following lemma to isolate the singularity in . Recall , and correspond to each other via (3.6), , and are defined in (3.26), and in (3.28).
Lemma 3.10.
Assume . There exists a unique continuous-in- real matrix valued satisfying
(3.72) |
Moreover the following hold.
-
(1)
The matrix is in , , and and
Moreover for any satisfying (3.23), there exists depending only on and , such that .
-
(2)
and are conjugate, namely,
(3.73) where
- (3)
- (4)
-
(5)
Finally, are in , , and if and .
Remark 3.5.
If needed, higher order Taylor expansions of can be obtained based on (3.75) through rather standard calculations in the analysis of local invariant manifolds.
One is reminded that both has a logarithmic singularity and a jump at which leads to such singularities of there even in the homogeneous case. Since for , should not be real for . Hence it is a non-obvious statement that this conjugate matrix is real. The above lemma isolates the singularity of into the explicit along with the smooth . Conceptually, the smoothness of in 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 .
Proof.
The construction of is adapted from the one in [5], where the main issue is to handle the singularity caused by . We first make (3.72) autonomous by changing the independent variable an auxiliary one such that and thus we have
(3.75) |
Obviously solutions to (3.72) correspond (up to a translation in ) to those to the ODE system (3.75) of 5-dim which converge to as , namely those on the unstable manifold of the steady state . The linearized system of (3.75) at is given by
It is easy to compute that, on the one hand, an eigenvector associated to the eigenvalue is
On the other hand, one may verify
which implies that in the 4-dim center subspace there is not any decay backward in . Therefore there exists a unique unstable manifold of 1-dim which corresponds a unique solution satisfying and and in all its variables. In fact, the 4-dim center subspace 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 is the only solution to (3.72) decaying to as , or equivalently . Even though this construction is local in , the domain of can be extended due to the linearity of equation (3.72).
With the existence of the solution to (3.72) established through (3.75), letting in (3.75) and then transforming back to (3.72), we have
This equation can be solved explicitly to yield
The conjugation relation is the consequence of the facts that both and the right side of (3.73) a.) are equal to at , b.) satisfy the same ODE system (3.72) for , c.) are continuous in due to the construction of 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 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 , since has at most logarithmic singularity at and , the Hölder continuity of in follows. From formula (3.74) and , we obtain . The limit property of also follows from similar calculation. Finally, the smoothness of under the assumptions and is again obvious from the representation of the solution (3.74). The proof of the lemma is complete. ∎
For , with the help of and Lemma 3.10 we shall analyze the fundamental matrices in two different forms of the homogeneous problem (3.24) with the condition (3.44) at
(3.76) |
where in is the initial value of the independent variable and hence . To analyze and , let
(3.77) |
where was used to compute the more explicit form of in the above, and
(3.78) |
The following lemma provides some very basic estimates on . More detailed ones on will be derived when needed.
Lemma 3.11.
Assume , . The fundamental matrices and and their entries and satisfy the following for any .
-
(1)
is in its variables if and is in its variables if and .
-
(2)
, , and are in and in and . If , then and are in and in , , and .
-
(3)
If , then and are in and in , , and .
-
(4)
and are in and and in and .
-
(5)
and are in , , and .
-
(6)
For any satisfying (3.23), there exists depending only on , , and such that for any ,
for , , and , and for ,
for .
The reason we consider of instead of individual or is not only that it yields better estimate. Recall the change of variables . The above fundamental matrix is in the form of . Therefore corresponds to the partial differentiation with respect to in the coordinates. Here we also used
(3.79) |
Proof.
The argument for and are very similar and we shall mainly focus on . Let
(3.80) |
Clearly we have
(3.81) |
All the smoothness follows from that of . The Hölder regularity in and in statements (2)–(4) is due to .
Straight forward computation based on Lemma 3.10 yields
Therefore
It follows immediately that and its derivatives in , , and are of the order . By mimicking , we have
Moreover, for and , we have
A similar procedure yields
Finally, since is in all variables, for , , and , the estimate on follows from its smoothness and vanishing at .
To analyze , parallelly we consider
Subsequently we have
The rest of the proof follows exactly as in the case of . ∎
3.5. Dependence in and of the fundamental solutions to the Homogeneous Rayleigh equation (3.1) with
In this subsection we revisit the two fundamental solutions
(3.83) |
of the homogeneous Rayleigh equation (3.1) for satisfying initial conditions (3.53). We often skip the dependence on and (or equivalently, on ) 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 is defined in (3.6).
Lemma 3.12.
Assume , . For and , the following hold.
-
(1)
As , uniformly in and .
-
(2)
As , locally uniformly in and also in and for any .
-
(3)
For each , if , if , for any and in .
-
(4)
Moreover,
where
Remark 3.6.
When takes the end point values , according to the above representation formula and the smoothness of , actually .
Remark 3.7.
The main goal of this subsection is to analyze the differentiation of in . Even though most of the results also hold for , the proof is slightly more technical. We shall skip those analysis of 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 , . For , it holds that
a.) is locally in both and for any ;
b.) are locally in both and for any at any satisfying ;
c.)
are in both and at any satisfying and ;
d.) is in and if ;
e.) are in , at any except for at ;
f.) assume , then,
for any , , , , and ,
is locally for near .
To obtain the estimates, for fixed near , as in Lemma 3.9, we divide into subintervals
(3.84) |
where is defined as in (3.54). is an interval due to the monotonic assumption of . Clearly and any of these subintervals may be empty. If , then is considered as for and as for . If , then (3.56) holds and is well defined. In the next three lemmas, we obtain the estimates on on subintervals in the order of , , and . 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 and and the initial values is trivial on and . While we mainly focus on in the following lemmas, we shall also just outline the estimates on , which would be enough for the rest of the paper.
Lemma 3.14.
Assume and . For any and any , the following estimates hold for and with ,
(3.85) |
where depends only on , , , and . Moreover, it also holds, for any
The above estimate holds in a neighborhood of actually.
Proof.
It is obvious that, for , is analytic in and . Let . One may compute that satisfies the non-homogeneous Rayleigh equation (3.3) in the form of
(3.86) |
with some constants . Note that the definition of implies that (3.8) is satisfied on with
(3.87) |
We shall estimate the derivatives of with respect to and for large and small separately.
For any sufficiently large so that , we shall apply (3.19) with to prove
(3.88) |
for any , with . The proof is a simple mathematical induction in .
Since (3.88) does not include the case , there are two base cases and , which we have to consider separately. For , from (3.86), (3.19), Lemma 3.9, and the definition of , we have, for any ,
where (3.66) and (3.67) are also used for to convert the estimates in terms of into those in terms of . Similarly, satisfies
With the estimates in the base cases established, for , using the induction assumption (and Lemma 3.9 for in (3.86)) and proceeding much as in the above, we obtain
and (3.88) follows consequently.
For , as , we apply Lemma 3.3 to (3.86) on with
Following a similar induction procedure and using Lemma 3.3, we obtain, for , , and with ,
Therefore (3.88) holds for all .
Estimating on . Let be solutions to the homogeneous Rayleigh equation (3.1) with initial values
and be the solution to the initial value problem of the non-homogeneous Rayleigh equation
On , can be estimated much as on , while much as in the proof of Lemma 3.9. When Lemma 3.2 is used to estimate for large , we set , , and . The desired inequality on follows from . ∎
Lemma 3.15.
Assume , , and , then Lemma 3.13a.)–e.) hold for . Moreover, there exists depending only on , and , such that, for any and any , the following estimates hold.
-
(1)
For ,
(3.89) -
(2)
If , then, for any , we have
(3.90) (3.91) and for and and ,
(3.92) (3.93)
In the above lemma denotes the delta mass supported at and and emphasize them as distributions of the variable . Near or , singular distributions of and at the level comparable to those negative exponents in (3.92) and (3.93) would occur. The quantities with upper bounds are functions for any .
Remark 3.8.
Statement (2) also holds for with slightly weaker upper bounds. From the proof, it is easy to see that if , then (3.90), (3.91), and (3.92) and (3.93) for hold with an additional or all replaced by on the right sides. If , then these inequality hold for with all on the right sides replaced by besides the additional .
Proof.
Since , it is easy to prove that (3.56) holds and is well defined. Let be defined as in (3.68) and (3.69) still holds. This allows us to work in the coordinate and apply Lemma 3.10, 3.11, and 3.12. It is natural to express using the fundamental matrix defined in (3.76). One is reminded that depends on . To study the regularity of and with respect to at some , we fix some (so independent of ) in a neighborhood of . For near , , we can write
(3.94) |
Note that iff and iff , the latter of which happens iff . Clearly and are smooth in and either due to the initial conditions or due to the smoothness of the Rayleigh equation on . Hence the regularity statement (c) of Lemma 3.13 follows from statement (1) in Lemma 3.11. If is close to , then we could fix . In this case, and involve only and due to , and thus statement (b) follows from statement (3) in Lemma 3.11. When is close to , the regularity of in and is a consequence of statement (2) in Lemma 3.11, unless . Near the last exceptional case, the regularity of in and is due to (4) of Lemma 3.11. Statement (e) of the smoothness in of also following from the properties of given in Lemma 3.11.
We shall derive the estimates of the differentiation by at in two cases.
* Case 1: . In this case, fix which implies . Hence and as well as its derivatives are of order when varies slightly. Therefore the related terms can be estimated easily. From the estimate at derived in Lemma 3.14 (or from the initial condition at ), (3.79), and Lemma 3.11, for , it holds on ,
where is given in (3.77) and the constant in the terms depends only on and . We also used that and are comparable at for in this case.
For , keeping the most singular terms arising from the derivatives of and in the distribution sense, we have
Using (3.77), (3.74), the smoothness of and , one may compute
(3.95) |
(3.96) |
and it yields the desired estimates for in this case.
Similarly, at for , keeping the worst term and using (3.96), we have
The desired inequality (3.92) in case 1 follows.
To finish the analysis in this case, we consider . From (3.95) we obtain smoothness in and . Differentiating (3.95) in and using Lemma 3.10 and 3.14, one may estimate, for ,
which proves (3.89) in case 1.
* Case 2: . In this case, let . While we have to deal with possibly very small in (3.94), the initial values . Hence from Lemma 3.11 we obtain, for , ,
(3.97) |
From (3.77) and Lemma 3.10, and are functions, which could be used to reduce some singularity. As , one can compute for ,
We use the following elementary inequalities to handle the above terms:
which also imply
The delta functions produced by differentiating are cancelled by . Finally only when which implies . Summarizing these estimates we obtain
If , then , while if . Hence , which along with the estimate in case 1 yields (3.90).
Similarly, for , , and , where are constants, one may compute
For , we have
If , the first term on the right side of the first inequality would be which as shown previously also satisfies the above final estimate. Similarly, one can also calculate, for , , and ,
The cases of have been considered earlier and would only make minor contributions. Therefore (3.92) and (3.93) are satisfied in case 2 as well.
Regarding , much as in case 1, but with much simpler initial value at , we have
which also yields its smoothness. Differentiating in and using Lemma 3.10 and 3.14, one may estimate, for ,
which proves the inequality in (3.89) in case 2. Finally, from Lemma 3.10,
Estimating on . In this case implies . Much as in the above argument for , we consider the estimates related to at some . Observe that, as an expression of solution to the homogeneous Rayleigh equation, (3.94) also applies to on with chosen near . In the case of , the same arguments yields the desired estimates of .
In the case of , we take and proceed roughly as in the above case 2. Due to the initial condition (3.53), equation (3.97) is replaced by
where (3.78) and Lemma 3.11 are used. Let
Recall from initial condition (3.53)
On the one hand, from (3.77), and Lemma 3.10, we have that
are function with bounds uniform in and . Hence the estimate on is obtained much as that of . On the other hand, as (3.96) and (3.95) also apply to , it holds
With these estimates, the desired estimate on follows much as that of . This completes the proof. ∎
Lemma 3.16.
Assume , , and , then Lemma 3.13a.)–e.) hold for . Moreover, if , then there exists depending only on and , such that, for any and any , the following estimates hold for
(3.98) |
and for ,
Moreover, if and , then it also holds for ,
Remark 3.9.
Using Remark 3.8, the above estimates with an an additional on the right sides also hold for and .
Proof.
The assumption and imply and is uniformly bounded from above and below away from 0. The regularity of and in and for follow directly from such smoothness at obtained in Lemma 3.15. Their estimates at can be summarized into
and for
where we also used as . Much as the proof of Lemma 3.14, we shall obtain the estimates inductively in by considering the cases of small and large separately.
As , we take such that (defined in (3.87)) for and thus (3.8) is satisfied on with . We shall obtain the estimates for this case of by splitting into homogeneous and non-homogeneous parts. For , let be the solution to the homogeneous Rayleigh equation (3.1) with initial condition
and be the solution to the non-homogeneous Rayleigh equation (3.3) with the zero initial conditions at and the non-homogeneous term given by the right side of (3.86) (with and ). Clearly it holds
(3.99) |
Using the the above estimates on at , we apply Lemma 3.2 to with
to obtain, for ,
Concerning , Lemmas 3.2 and the same computation as in the proof of Lemma 3.14 implies, for any ,
The desired estimate for follows from (3.99), Lemma 3.9, and direct integration. For , one may compute inductively using the above estimates and (3.94),
If , the desired estimate follows immediately, otherwise it follows from the fact for any .
In the case , and Lemma 3.3 yields the estimates through a similar induction.
The following lemma proves Lemma 3.13(f) and (the case of ) will be used in analyzing the eigenvalues.
Lemma 3.17.
Assume , . For any and , there exist such that
(3.100) |
for any , , , . Here can be taken independent of for in any bounded set.
Unlike in most other lemmas, the constants and may depend on and .
Proof.
The lemma is trivial if , so we assume . Since the lemma is concerned with close to where and may depend on and , we consider
If , then (3.100) clearly holds as is away from the singularity of . Otherwise, let . From (3.78) and the smoothness of due to Lemma 3.11, we have
where we also used that is independent of . The desired inequality (3.100) follows from straight forward calculations using the smoothness of , , and . ∎
Remark 3.10.
Most of the above regularity results and estimates also hold for . Since plays a less substantial role as in the rest of the paper, we only gave the basic estimates on .
In the above was considered only for . To end this subsection, we extend some estimate for using the analyticity of in in the following lemma.
Lemma 3.18.
Assume . The following hold.
-
(1)
For any with , it holds
-
(2)
For any ,
-
(a)
there exists depending only on , , and , such that for any , , ,
-
(b)
as , and converge to and in , respectively, for any . Moreover, the convergence also holds in .
-
(a)
-
(3)
For any and compact interval ,
-
(a)
there exists depending only on , , , and , such that for any , , ,
-
(b)
as , and converge to and in , respectively, for any . Moreover, the convergence also holds in .
-
(a)
The multiplier in front of is added to regularize their singularities near and the denominators , , in the expressions related to are to make it decay as (recall the initial conditions (3.53) of involving ).
Proof.
Let be the open disk with diameter segment . For any , let
There exists such that for any . Lemma 3.2 (with , , , and ) implies, for and ,
(3.101) |
For , Lemma 3.3 implies that the above inequality still holds for .
From equation (3.86) ( and ) of , applying (3.19) with and using Lemma 3.9, we have for and ,
(3.102) |
For , Lemma 3.3 implies that the above inequality still holds for .
For any with , the analyticity of and its decay as imply, for any ,
where the boundary terms at infinity in the above integration by parts vanish due to the uniform-in- bound on given in (3.101). Letting , the same bound and Lemma 3.12 yield
where we integrated by parts again. The desired estimate on follows from the boundedness of the convolution kernel on , (3.102) for , and Lemmas 3.14–3.16.
The results for are derived in the same manner. In fact
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 near given in Lemma 3.15. The latter also yields the estimate on .
In statement (2b), the pointwise-in- convergence in is standard due to the convergence of the convolution kernel on as , as well as the analyticity of for . The convergence in follows from the pointwise-in- convergence in , the bounds in statement (2a), and the dominant convergence theorem.
Finally, can be analyzed similar. However, the initial values (3.53) induce an growth in and and an growth of and for (Lemma 3.3). Instead we consider, for ,
which holds for any . From this Cauchy integral formula we proceed much as in the above and obtain the integral representation in term of . The derivation of the corresponding formula of is also similar. The desired convergence and estimates of and in on a compact interval again follow from the properties of the convolution by the kernel . ∎
3.6. An important quantity
To end this section, we analyze a quantity related to the Reynolds stress, which is crucial for the linearized water wave problem:
(3.103) |
where is the solution to homogeneous Rayleigh equation (3.1) satisfying and defined in Subsection 3.3 and for . Due to Remark 3.4, satisfies estimates uniform in . With slight abuse of notations, we would not distinguish from in the rest of this section. Apparently the domain of is given by
and those excluded points (except ) exactly are the eigenvalues of of the linearized Euler equation in the fixed channel at the shear flow . is not defined at since has singularity at . We first summarize some basic or standard properties of in the following lemma.
Lemma 3.19.
Assume . The following hold.
-
(1)
For any , for any and .
-
(2)
There exists depending on such that, for any , it holds
(3.104) -
(3)
There exists depending only on such that, for any , , it holds, for any ,
-
(4)
There exists and depending only on , and such that, if or then
(3.105) -
(5)
Suppose a closed subset satisfies for all and where or , then there exists depending only on and such that (3.105) holds for all and .
Remark 3.11.
According to Lemmas 3.3 and 3.9, the assumption on in Statement (5) is automatically satisfied except possibly a compact set of . In particular, due to statement (3), it is satisfied for if on . We also recall is equivalent to that is an eigenvalue of the linearized Euler equation at the shear flow on the fixed channel associated with an eigenfunction .
Proof.
We first claim the following standard result.
Claim. Let is a solution to the homogeneous Rayleigh equation (3.1) on an interval with such that at some , then at any .
If , then the claim and is in on are obvious since the coefficients of (3.1) are real. If , then it must hold and Lemma 3.10 implies that and satisfies (3.74) with and . This formula yields . Finally, suppose at some . Let . Again due to Lemma 3.10 and it is standard to verify
(3.106) |
Multiplying it by and integrating it between and leads to a contradiction. Hence the claim is proved.
For , applying the above claim to on the interval if and on if , respectively, implies that does not change signs on these intervals. Hence we obtain statement (1) and for if . Along with (3.57), the continuity of , and Lemma 3.3, it also yields Statement (2).
In the view of Lemma 3.12, Remark 3.7, and statement (1), is also a solution on satisfying and . Statement (3) follows from statement (2) applied to on and to on .
From (3.57) and Remark 3.4, there exists such that (3.105) holds for all and . For , the restriction on involving ensures 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 for all and . Recalling the convergence estimates (3.50) and the locally Hölder continuity of in (Lemma 3.13), we obtain the continuity of in for . Lemmas 3.3 and 3.9 along with the continuity of and the non-vanishing assumption imply that (3.105) holds for all and with . As , statement (5) follows and it completes the proof of the lemma. ∎
In the following we give some basic properties of .
Lemma 3.20.
Assume , . It holds that and is a.) analytic in both , and, when restricted to , b.) in , and c.) in and locally in for any . Moreover,
-
(1)
and .
-
(2)
There exists depending only on such that
-
(3)
For any , there exist and depending only on , , and such that,
-
(4)
For any and , there exists depending only on and such that
Proof.
The analyticity and the conjugacy property of are obvious from its definition. The property is a direct corollary of Lemma 3.19(1). The smoothness of away from and follows from Lemma 3.13 and the analyticity of in with . The Hölder continuity of is again a corollary of Lemma 3.13 for varying along and Proposition 3.7 for varying along . The explicit form of is a direct consequence of the observation
(3.107) |
To end the proof of the lemma, we obtain the quantitive estimate on . From Lemma 3.19, for any . Along with Lemma 3.9, it implies that (3.105) holds for for some depending only on . Statement (2) follows from the upper bound of given in Lemma 3.9. Statement (3) is also a direct consequence of Lemma 3.9 where is involved to ensure . In statement (4), the restriction on guarantees due to the semicircle theorem and the desired inequality follows Lemma 3.3. ∎
The analyticity of in allows us to use the Cauchy integral to analyze . For , let
(3.108) |
be the -neighborhood of .
Lemma 3.21.
Assume , , and satisfy on , then for any and we have
(3.109) |
where denote the integral along the contours counterclockwisely.
Here denotes the derivative of as a function of the complex variable and thus due to its analyticity.
Proof.
The assumption implies is analytic in and continuous in . For any , the analyticity of yields
(3.110) |
Applying Lemma 3.20(4) with and , we have
Therefore
and thus the desired integral formula of follows. The representation of simply follows from direct differentiation. ∎
Remark 3.12.
Though not needed in the rest of the paper, this lemma could be modified for general and . In this case, should be chosen so that along . The integral representation formula would involve the residue at those roots of outside . The estimates should also be modified accordingly.
To analyze the remaining integral in (3.109), we start with the imaginary part of .
Lemma 3.22.
for . Assume , , and , then
Proof.
The vanishing of for is obvious from its definition and Lemma 3.19(1). To derive the expression of for with and , we may consider , , which is also a solution to the homogeneous Rayleigh equation with and . It is straight forward to calculate
Applying the convergence estimates (3.50) and the Hölder continuity of in , we obtain the desired
This completes the proof of the lemma. ∎
The above formula yields some refined estimates of for .
Lemma 3.23.
Assume , , then the following hold for .
-
(1)
is in and it satisfies
Moreover, if , then, for any , , , and , is locally in .
-
(2)
Assume , then there exists depending only on and such that, for any , we have
where .
Proof.
Lemma 3.13 implies the smoothness of in and and that of in , hence is in . Moreover, for due to Lemma 3.19(1), which along with Lemma 3.13a.) implies that for near or and thus is Hölder continuous for near and . The local regularity of follows from Lemma 3.13(f).
The upper bound estimate of and its limits as approaches and 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
is estimated by (3.89). The other key term will be considered in three possible cases of according to the division of defined in (3.84) in Subsection 3.5. Observing implies and .
* Case 1: and . The former happens if and only if , while if and only if . Lemma 3.16 implies
* Case 2: and which occurs if and only if . Also from Lemma 3.16, we have
In the following we analyze by writing it as a Cauchy integral of .
Lemma 3.24.
Assume , , and satisfy that for all , then and are locally in and in the domain for any , , and . Assume, in addition, for all , then, for any ,
(3.111) |
and for ,
(3.112) |
Here denotes the Hilbert transform in , namely,
where P.V. represent the principle value of the singular integral. We also recall and for .
Proof.
Let us first assume for all , then is well-defined for all . We shall apply Lemma 3.21. The contour is the union of two segments , the left half circle centered at with radius , and the right half circle centered at with radius . As , due to the continuity of (when restricted to ) at and its logarithmic upper bound near given in Lemma 3.20, the Cauchy integrals along the two half circles converge to zero as . Hence the integral form (3.111) of follows from taking the limit of (3.109) as and the conjugacy .
For , the integral form (3.111) can be rewritten as
A standard treatment of the above singular integral as , along with the regularity of in given in Lemma 3.22 and 3.23, yields (3.112).
The regularity of follows from that of and the boundedness in of the convolution by with the parameter . Here the singularity of near does not affect the regularity of away from due to the localization property of this convolution operator.
Finally, if we only assume for , due to its analyticity in , its continuity when restricted to , and (Lemma 3.3), has at most finitely many singular points . Hence there would be at most finitely many additional contour integrals of in (3.111) along contours in enclosing the roots of . Those integrals in the analytic region of would not affect the regularity of . The proof of the lemma is complete. ∎
With the representation of in terms of Cauchy integrals, we may also calculate its derivatives in more details.
Corollary 3.24.1.
It holds, for ,
(3.113) |
and for ,
(3.114) |
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 . As (1.3) preserves Fourier mode for any , the wave number would be treated as a parameter in this section. According to Lemma 2.3, , , is an eigenvalue of (1.3) with parameter if and only if
(4.1) |
where the last equal sign in the first row is due to the conservation of the Wronskian of (3.1). Let
It is easy to see that, if , then also generates the associated eigenfunction of (1.3). In the literatures, those zero point of with are often referred to as unstable modes, while those zero point 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 is at which would be crucial for the bifurcation analysis of eigenvalues. We also consider an almost equivalent quantity
(4.2) |
where is defined in (3.103), and
Apparently and satisfy
(4.3) |
(4.4) |
From Lemma 3.24 is near if for all , which is crucial for the bifurcation analysis.
4.1. Basic properties of eigenvalues
Apparently it holds that
(4.5) |
In the following we first give some basic properties of under minimal assumptions.
Lemma 4.1.
Assume , , then for any , the following hold.
-
(1)
is well defined for all and . When restricted to , is in and and is also in both and with .
-
(2)
is well-defined for close to and , near , and
-
(3)
Assume , then for any , there exists determined only by , , and , such that, for any and ,
where the norm is taken on and we recall .
-
(4)
if . Hence for any .
- (5)
-
(6)
For any , if .
Proof.
For , the convergence of as follows from the convergence estimates given in Proposition 3.7. For near , the logarithmic singularity in is cancelled by and thus the convergence of and the continuity of at follow. The and smoothness of is obtained from those of and (Lemmas 3.13 and 3.15) as well as using the factor multiplied to .
From Lemma 3.19(1), and thus is well-defined near . The property and the value of are due to those of given in Lemma 3.20(1). The smoothness of for near follows from Lemma 3.20(2) and the definition of . The values of and at is obtained by direct computation.
Suppose . Lemma 3.19(1) implies . As a non-trivial solution to the homogeneous Rayleigh equation (3.1), it must hold . Therefore .
To prove statement (5), we first observe that iff satisfies the corresponding homogeneous boundary conditions of (2.11c), which happens only if and thus and are well-defined. Moreover the statement is obvious for and also for due to the smoothness of (Lemma 3.10), while due to statement (2). Hence we focus on only. “”: As , implies and consequently according to Lemma 3.22. Consequently Lemma 3.10, particularly formula (3.74), and the definition of yield the smoothness of which apparently satisfies (2.14). “”: This solution has to be proportional to on which yields due to 3.19(2). Hence the smoothness of and equation (2.14) imply . Consequently both (2.14) and the homogeneous Rayleigh equation (3.1) are regular on and are equivalent to each other. Therefore and are proportional on and thus satisfies the boundary condition at .
Remark 4.1.
The monotonicity assumption on is used in the above proof of statement (5). If is not monotonic, may contain several points in for a root of and the corresponding solution may not be in . Therefore the set of roots of , which is what really matters, may be larger than those defined as singular modes in Definition 2.1.
In the next step, we consider for . Unlike the linearized Euler equation on a fixed channel where no eigenvalues exist for large . Eigenvalues do exist for each large for the linearized water wave system. According to Lemma 4.2(2), we often consider as well.
Lemma 4.2.
Assume , then the following hold for any .
-
(1)
There exists depending only on , , and , such that
where we recall .
-
(2)
For any , there exists depending only on , , , and , such that, for any and satisfying ,
-
(3)
There exist and depending only on , and , such that for any , (4.1) has exactly two solutions which are contained in and depend on analytically. Moreover they satisfy
Proof.
The first statement follows directly from Lemma 3.9, where the factor is used to cancel the logarithmic singularity in the estimate of , and the second from Lemma 3.3 with . We focus on the roots of . From Lemma 3.9,
and thus we can work with and . Let
From statement (1) and Lemma 3.20(3), it holds that there exist such that, for any , only if . We may take larger if necessary such that . From Lemma 3.24 and 3.23 and Corollary 3.24.1, there exists depending only on such that, for all , ,
By a substitution we obtain
On the other hand, viewing as a quadratic equation of , its roots also satisfy
Using the above estimates on and , it is straight forward to verify that for any and ,
and
Therefore are contractions acting on . Their fixed points , analytic in , are the only solutions to (4.1), or equivalently (4.2). These since for which allows the iteration to be taken in . Finally, one may compute
(4.7) |
Using the above estimates on , , and , one may compute
The evenness of in is due to that of and the uniqueness of the fixed points of the above contractions. This completes the proof of the lemma. ∎
We shall track the two roots of the analytic function as decreases, based on a standard analytic continuation argument.
Lemma 4.3.
Assume . Suppose and satisfy and , then the following hold.
-
(1)
There exists an analytic function defined on a maximal interval such that and .
-
(2)
for all if and only if .
-
(3)
If (or ), then
-
(a)
(or if ), or
-
(b)
(or if ).
-
(a)
Proof.
We start the proof with a simple and standard consideration of the index of complex analytic functions. Suppose at any where is a domain with piecewise smooth boundary , then the index
(4.8) |
is equal to the number of zeros of inside , counting their multiplicities. Therefore the analyticity of in and implies that Ind is a constant in as long as does not occur on .
As a consequence, starting with the simple root of , a unique continuation of of simple roots of exists and is analytic in . The simplicity of is due to the fact Ind for any sufficiently small neighborhood of in the continuation procedure. For any , we have and . Therefore if for some along the continuation curve, then the unique extension coincides with the (real) root of obtained by applying the Implicit Function Theorem to the real function . Hence if and only if .
Let be the max interval of the continuation as simple roots of and we shall prove statement (3). Suppose , while the other case can be analyzed similarly. As , the solution curve is bounded due to Lemma 4.2(2). Therefore there exists a sequence such that and exists. Statement (2) implies that stays in the closure of either the upper or lower half of and thus . Assume statement (3)(a) does not hold, then such a subsequence can be chosen such that . Therefore is a root in the domain of analyticity of . Clearly is not a simple zero of , otherwise can be extended beyond . Recall has to be an isolated root of since all roots of non-trivial analytic functions are isolated. Therefore, there exists a small neighborhood of such that, for any sufficiently close to , it hold Ind. Consequently, for each close to , there exists at least another root of in and thus (3)(b) holds. ∎
The semicircle theorem of Yih [39] states that all imaginary roots of are contained in the circle with the diameter segment , so the only possibility for the branches of simple roots of obtained in Lemma 4.2 can not be extended for all is when they reaches or , respectively. As a corollary of and we have
Corollary 4.3.1.
(1) The branch
can be extended for all . Moreover is even in , , and for all , for some independent of .
(2) If for all , then of simple roots of obtained in can also be extended for all . Moreover is even in , , and for all , for some independent of .
Proof.
Let be given in Lemma 4.2(3) and we only need to focus on . We may assume is sufficiently large such that and . From Lemma 4.2(2), there exists such that for all and . Hence and are the only roots of , which are also simple with .
We first consider . Let
According to Lemma 4.1(2), for any . Hence the semicircle theorem and the choice of imply that a.) and b.) for all and , and thus
Therefore none of the possibilities in Lemma 4.3(3ab) can happen to the extension starting from , so this branch of simple root of can be uniquely extended for all with as the only root of in . The value of this extension at has to coincide with as are the only roots of while . Therefore the extensions starting from have to coincide. The evenness of in follows from that of and the uniqueness of its root in . The sign of remains positive from as is always simple. The existence of is simple due to the continuity of . The same argument applies to under the assumption all for . The proof is complete. ∎
Based on the above analysis, we shall conclude that are the only eigenvalues of the linearized capillary gravity wave under the additonal assumption of the absence of singular modes
(4.9) |
where or and is the period of the water wave in the direction.
Proposition 4.4.
Assume and (4.9) for or , then there exists such that
-
(1)
.
-
(2)
Assume , then .
Proof.
The first statement is a direct corollary of the continuity of , its analyticity outside , assumption (4.9), and Lemma 4.2.
Let us consider statement (2). Corollary 4.3.1 and (4.9) imply that both and can be extended as even analytic functions of . Let be taken as in the proof of Corollary 4.3.1 and we only need to focus on . Assumption (4.9) also yields such that
Let
then we have for all and . Therefore
and 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 and are viewed as function of and . According to Lemma 3.19(1), is well-defined for in a neighborhood of .
Lemma 4.5.
Assume , then we have
Proof.
For and with and , let
(4.10) |
be the differential operator in the Rayleigh equation (3.1) and the normalization of the fundamental solution defined in (3.53) and (3.83). Clearly
where the sign properties follows from Lemma 3.19(1). It is straight forward to compute, for and ,
where the smoothness of in is ensured by Lemma 3.10. The following claim is used to analyze these and some other functions.
Claim. Suppose is a solution to and with , where is on , then we have the following through direct computations
(4.11) |
Applying this claim to and implies, for ,
(4.12) |
The definition of implies .
For , through direct calculation, one may verify, for ,
(4.13) |
For , from (4.12), we have
and thus
(4.14) |
For and , we can use (3.107) to compute
(4.15) |
Consequently, one obtains explicitly
which in turn yields the desired formula of . ∎
The information on the derivatives of leads to the following properties of the roots of .
Lemma 4.6.
Assume , the following hold.
-
(1)
If
(4.16) then for all .
-
(2)
Let
then we have
-
(a)
and “=” in the “" holds if and only if (4.16) holds.
-
(b)
If , then for all .
-
(c)
If , then there exists a unique such that and for all .
-
(d)
If , then there exist such that
-
(a)
Proof.
Statement (1) is a direct consequence of the concavity of in and . 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 does not reach and thus staying in for all .
To end this subsection we prove the following monotonicity of the even functions 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 . From the definition of and (4.13), we first compute, for ,
and thus
which is uniformly increasing on and uniformly decreasing on . Therefore has two real roots
(4.17) |
which are unique in the above intervals.
Lemma 4.7.
Assume , then the following hold.
-
(1)
For , suppose can be extended as simple roots of for all , then has most one solution on , where is also satisfied.
-
(2)
For , suppose can be extended as simple roots of for all , then for all if and only if
(4.18) with defined in (4.17).
Proof.
We shall work with , while the same proof works for . Suppose there exists such that , then
Computing the second order derivative at , we have
which along with Lemma 4.5 and (Corollary 4.3.1) implies . Hence has to be the only positive critical point of .
To prove Statement (2) where , on the one hand, we first observe that since is the unique root of in and is also such a root, so . Moreover, (4.14) implies that (4.18) is equivalent to . On the other hand, observe that the evenness of yields . One may compute
From Lemma 4.2(3), for some . Hence, on the one hand, if (4.18) does not hold, then (Lemma 4.2(3) and 4.3) and the above identity imply . Along with , it yields that has a critical point . On the other hand, through the same argument, (4.18) yields while . Therefore, if which implies , it is impossible that there exists a unique critical point of where . In the borderline case of which implies , further Taylor expansions of the even-in- functions and yields
From the same reasoning, we obtain that for . The proof of the lemma is complete. ∎
4.2. Eigenvalue distribution of convex/concave shear flow
To analyze eigenvalues under less implicit assumptions than (4.9), particularly the generation of unstable modes from , we further assume on . Due to Lemma 4.1(6), this rules out the possibility of roots of on and provides better smoothness of for the bifurcation analysis.
Lemma 4.8.
Assume , , and on , then
is well defined for all and and
a.) is analytic in both and and, when restricted to , is in both and ,
b.) is locally in both and with for any , , and ,
c.) is in and with .
Remark 4.2.
Proof.
The assumption implies that for all and (Lemma 3.19(5)) and thus is well defined. The analyticity and the and (restricted to for the latter two) regularity of follow directly from those of given in Lemma 3.24 except at . Near , the regularity and estimates on (Lemma 3.20, 3.23, 3.24) and (Lemma 3.23 and Corollary 3.24.1) yield the regularity of . ∎
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 .
Assuming , in general is the only point outside the domain of analyticity of which might happen to be a root and also might be the end point of branches of roots of , it is a crucial step to analyze zeros of around .
Lemma 4.9.
Assume , then (a) for all if on ; and (b) if on , then if , where is understood as restricted to .
Proof.
We shall use the notations and defined in the proof of Lemma 4.5 and and are also viewed as function of and . It is straight forward to compute, for and ,
Applying (4.11) we obtain
(4.19) |
These integral representation of still holds as , and thus also its sign. For and , we can use (4.15) to compute
Finally we obtain the sign of in two cases separately, based on the sign of . Suppose . The above (4.12) and (4.19) implies that, for , is strictly increasing in and is strictly deceasing in , and thus
is also strictly decreasing in . Letting , this monotonicity yields
In the other case of , suppose for some , which implies
Therefore
We also have from taking the limit of (4.19). Hence we obtain and the proof of the lemma is complete. ∎
In the next step we shall study the roots of near .
Lemma 4.10.
Assume , and on . Suppose , then there exist , , and for any such that
and for and ,
Moreover, without loss of generality assume (Lemma 4.8 implies ) and this branch of roots of satisfies
-
(1)
If , then , and for all .
-
(2)
If , then and
and for some determined by and ,
which implies
In the generic case , locally the roots of consists of the intersection of the graph of and the closure of the upper half complex plane, along with its complex conjugate. In this case, however, one observes that at means 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, is in and in the region . As is not continuous at in general, let be a extension of into a neighborhood of which coincides with for . From Lemma 4.9, the Jacobian matrix of satisfies
where we used the Cauchy-Riemann equation and the fact for all . Therefore the Implicit Function Theorem implies that all roots of near form the graph of a complex-valued function which contains . To complete the proof of the lemma, we only need to prove that satisfies properties (1) and (2).
Firstly we prove if and thus as well. When restricted to , and . The Implicit Function Theorem yields a real-valued function for near such that
(4.20) |
Since if , the uniqueness of solutions ensured by the Implicit Function Theorem implies that if .
Next we consider the case . Along with
it implies
From the Intermediate Value Theorem, for near , there exist real roots of slightly smaller than , which must belong to due to the uniqueness of solutions ensured by the Implicit Function Theorem. Therefore along with the last step, we conclude for close to .
Finally, we consider the case of , while the opposite case can be handled similarly. The fact yields
where in the calculation of we also used for and the smoothness of . Hence we obtain for slightly smaller than . In the following we shall focus on where . In this case, apparently and thus . From the Mean Value Theorem, there exists between and such that
The regularity of and implies
The proof of the lemma is complete. ∎
While the branch of neutral modes is global in and contained in as addressed in Corollary 4.3.1, in the following we completes the picture of the other branch by combining Lemmas 4.3 – 4.10 and finish the proof of Theorem 1.1.
We start the rest of the proof much as in that of Corollary 4.3.1 and Proposition 4.4. Namely, let be given by Lemma 4.2(3) and we only need to focus on for . From Lemma 4.2(2), there exists such that for all and , which also implies for all and for all .
Case 2. . One the one hand, for any , Lemmas 4.2, 4.8, and 4.6 imply that there exists such that
where is the -neighborhood of (see also (3.108)). Hence for all , we have Ind, which is equal to the number of roots of in . According to Corollary 4.3.1, , , is one of them. Therefore neither cases in Lemma 4.3(3) can happen to the branch and the simple root can be extended analytically for all . Therefore can be extended to at least which along with are the only roots of for . On the other hand, according to Lemma 4.10, there exists a branch of the only roots of for . Moreover for . Therefore for as are the only roots of for . In particular, is thus extended to as a function with for .
Moreover, on the one hand, is the only root of near for near and it satisfies . On the other hand, the continuity of implies that there exists such that for any and . It implies
due to the root , where
Therefore, are the only root of for near , which are also simple. As is away from , Lemma 4.3 implies that the branch of simple roots can be extended at least to and remains in . As is even in , we have are also the only roots of for . Therefore the extension must be even on and we obtain the whole branch for .
Case 3a. and . Following the same arguments as in case 2, we obtain that can be extended to a function on for some , such that and are the only roots of for all and for . Let also denote the maximal interval of the analytic extension of as a simple root of inside . The same above index based argument (in case 1) applied to for any and also implies that and are the only roots of for all . According to Lemma 4.2 we have . For , the semicircle theorem implies that lies in the closed upper semi-disk with the boundary diameter and thus where is given in Corollary 4.3.1. Moreover, since for any (Lemmas 4.1 and 4.8), we obtain from Lemma 4.3
It must hold , otherwise , , , and Lemma 4.10 imply that there exists the fourth root near for . This contradicts that has exactly three roots for all and thus and . For , Lemma 4.10 yields the further extension of back into . From a similar argument, we can extend this branch to with . Finally, the whole branch for is obtained by the evenness .
Case 3b. and . Following the same arguments as in case 2, we obtain that can be extended to a function on and . However, for , Lemma 4.10 implies that there does not exist any roots of near (as due to ). The same index argument further yields that is the only root for . From Lemma 4.10, we obtain another branch of roots in of for which along with the are the only roots. The final conclusion again follows from the even symmetry as in the above cases.
Remark 4.3.
As in [39] for the gravity wave, the spectral stability in the case can also be obtained by directly modifying the usual proof of the Rayleigh theorem in the fixed boundary case. Namely, multiplying (3.1) by , integrating on , using the homogeneous boundary condition as in (2.11b) and (2.11c), and the semicircle theorem, a contradiction occurs if an unstable mode 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 , which are the only possible singular neutral modes other than according to Lemma 4.1(6).
Proposition 4.11.
Assume , , and , then the following hold for .
-
(1)
For any , there exist depending only on , , and , such that, with
for any , there exists a unique such that . Moreover it satisfies
-
(2)
In addition, suppose and
then there exist , , and a function defined for
such that , for the above , and
In the above statement (2), note that and Lemma 4.1(4) imply and thus is well-defined which is actually real due to and Lemma 3.22. Therefore it makes sense to talk about the sign of . Statement (1) also implies that assumptions of statement (2) may be satisfied at inflection values of with if is small.
Proof.
From Lemma 3.9 and Remark 3.4, there exists such that
and thus and are defined for all . According to (4.6), for all and thus . Lemmas 3.24 and 3.22 imply, for and ,
Therefore, for , it holds
Let
From the Intermediate Value Theorem, for every , there exists a root of close to .
To estimate and obtain the uniqueness of , we analyze using the same standard method used in the proof of Lemma 4.5. Let
where . Differentiating the above equation with respect to yields
From Lemma 3.9, we can estimate, for any and ,
Therefore we obtain
which implies
The desired estimate on follows immediately, whose always negative sign also implies the uniqueness of such .
Under the assumption in statement (2) of the proposition, Lemma 4.1(4) implies and thus is in near with . Much as in the proof of Lemma 4.10, statement (2) can be proved by applying the Implicit Function Theorem to , an extension of which is in in near . The Jacobi matrix of is
where we also used the Cauchy-Riemann equation. According to Lemma 3.22, and
and has the same sign as . Therefore is invertible and thus there exist and a function defined for all such that for iff . Consequently for near iff and . Identifying complex numbers with 2-d column vectors, since
implies for near , statement (2) follows readily. ∎
Remark 4.4.
In part (1) of the proposition, one may also seek satisfying using the Intermediate Value Theorem instead. It is easy to see approaches as . Therefore such exists if and only if , which may not the case if and are sufficiently large. This is different from the gravity waves (i.e. ), see [39, 14, 15]. It is also worth pointing out that the smoothness of for 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 to the homogeneous Rayleigh equation (3.1), we study the boundary value problem of the non-homogeneous Rayleigh equation
(5.1a) | |||
(5.1b) |
where the boundary conditions are from the linearized water wave system (2.11).
Using the two fundamental solutions to the homogeneous equation with zero boundary values, for it is standard to compute the solution to (5.1) in the form
(5.2) |
where is the solution to (5.1a) with zero boundary values in (5.1b) given by
(5.3) |
Its derivative in is given by
(5.4) |
Here the unique solvability condition of (5.1) is , where is defined in (4.1), as the Wronskian of the fundamental solutions , which is a constant in , is given by
(5.5) |
Throughout this section, we consider
where . By choosing smaller, we also have that, for some depending only on and ,
(5.6) |
This and boundary condition (5.1b) imply
(5.7) |
which will be used repeatedly to control in terms of .
Throughout this section, we assume that, there exists such that
(5.8) |
In this subsection, mostly we shall not vary , but carefully track the dependence of the estimates on , or equivalently . From Lemma 3.9, it is easy to compute that, for any , , and ,
where and . This inequality will be used repeatedly in the rest of the paper.
Solutions to this system are rather smooth away from 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 and . 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 , which is carried out in this section.
As a preparation, in Subsection 5.1 we shall first consider (5.1) with zero boundary conditions in (5.1b). Subsequently in Subsection 5.2, we study the non-homogenous Rayleigh system (5.1) with linear in , particularly focusing on the derivatives of the solutions on . We sometimes skip writing parameters and explicitly.
5.1. Non-homogeneous Rayleigh system (5.1) with zero boundary conditions
The formulas (5.3) and (5.4) of and are actually consistent with (3.34) for near . In fact, (3.34) implies that is a fundamental matrix of (3.1) and hence can be rewritten in terms of and . A straight forward calculation using (3.30) and (3.34) also yields (5.3). This solution also satisfy
so we mainly focus on . Assume as . Due to the singularity of the non-homogeneous term at (as defined in (3.20) by ) as , the limits of and involve of integrals and delta masses
(5.9) |
(5.10) |
where is the characteristic function and we skipped the dependence on of , , and . Naturally, in the above the is taken only when there are singularities in the integral.
We consider a priori and convergence estimates of as in the following two cases of , motivated by the non-homogenous Rayleigh system (2.11) and its differentiation in .
Case 1: . 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 .
Lemma 5.1.
Assume (5.8). For any 222Like the generic upper bound , the small constant in this and the next section may change from line to line., there exists depending only on , , , , and , such that the following hold.
-
(1)
For any , , , and it holds
-
(2)
Assume in as with and , then
-
(a)
in for any and in for any and ;
-
(b)
at and , and in for any . Moreover, and for any , for any , ,
-
(a)
Even though the above formulas of involve some subtlety at , the regularity of in implies that is Hölder continuous. In fact, the continuity of at can also be seen directly by using the rather precise local form of near 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 using the tools given in Lemma 3.4 and Proposition 3.7.
Proof.
Since , no singularity is involved in (5.3) and (5.4), one can compute via integration by parts
The other integral can be handled similarly,
Observing that the boundary terms at are canceled and we have
(5.11) |
The above two integrals can be estimated similarly and we shall focus on the first one only. Lemma 3.9 implies
Using the Hölder inequality we obtain
From the initial condition (3.53) (in particular ) and (5.5), the remaining boundary term can be estimated as
The desired estimate on follows from (5.8), (5.5), Lemma 3.9, the above inequalities, and the standard Sobolev inequality
(5.12) |
The estimate of can be obtained much as in the above. Integrating by parts and using (5.5) to handle the boundary terms at , we have
(5.13) |
The desired estimate on follows from (5.4), (5.12), the above estimate on the integrals, and Lemma 3.9.
To consider the convergence of , we first note that, for , the imaginary part of belongs to and as ,
(5.14) |
Using expression (5.11), the estimates thereafter, bounds on in Lemmas 3.9, and the convergence of to as in Lemma 3.12, it is straight forward to obtain
in for any , where, for , two other terms involving (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 in the convergence is due to the last logarithmic term. Since in the distribution sense, the above limit is equal to after integration by parts. The convergence of is obtained using (5.13) along with (5.5) in a similar fashion
where again two other terms involving cancelled each other for . Here the convergence in the slightly weaker norm , for any and is due to the logarithmic singularity both explicitly outside the integrals and in (see also Lemma 3.12). The limit can be simplified to
which is equal to after an integration by parts.
At the end point , and have only one integrals and, unlike for general , the terms , and 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),
The desired inequalities follow from (3.53) and the above estimates, which completes the proof of the lemma. ∎
Assuming , in the following we estimate and as well as their derivatives in in , in particular their dependence on , by an energy estimate approach.
Lemma 5.2.
Assume (5.8). For any , there exists depending only on , , , , and , such that for any and , it holds
(5.15) |
where the norms are taken for and .
Proof.
We first assume and drop the subscript for notation simplification. Multiplying the Rayleigh equation (5.1a) by and integrating in both and , we have
The first term of boundary contribution can be estimated by Lemma 5.1(2b) and (5.7) with , as well as (5.6), (5.8) and (5.12),
Concerning the last integral , we first split it as
The above terms at can estimated much as and we obtain
We shall estimate all the remaining terms using the Hölder norms of and . For any function on an interval, it holds
(5.16) |
which applies to and . In the can be replaced by if vanishes somewhere in the interval. Hence for each fixed with and ,
For any and , using and the above estimates, we obtain
By choosing , we have that, there exists such that for any and , satisfies (5.15). To obtain the estimates for and in the limiting case , for , let and be defined by (5.3) and (5.4), which satisfy the desired estimates uniform in . For and , the desired estimates simply follows from the estimates and convergence obtained in Lemma 5.1.
Case 2:
(5.17) |
Again we start with rough estimates on and .
Lemma 5.3.
Proof.
Since the desired estimates are stronger and with weaker assumptions if is smaller (with possibly greater ), without loss of generality, we may assume . In the following we shall need the modification determined by :
(5.18) |
For , we first split into
and
where we skipped all the dependence on and . Clearly .
To estimate , we can rewrite its integral part as
where
The operator of convolution by is bounded on uniformly in and converges to strongly in as , where is the Hilbert transform and is the identity. The other integral can be treated similarly and we obtain from (5.8) and Lemma 3.9
Moreover, since are two uniformly continuous mapping from to and the above convolution is bounded on uniformly in , we have that are two families (with parameter ) of equicontinuous functions (of ) from to . As , they converge pointwisely (in ) to which are also uniformly continuous in . The equicontinuity and the compactness of imply that the convergence is uniform in . Therefore, along with the convergence of as (Lemma 3.12), we obtain that, as ,
The other part can be estimated by the Hölder continuity of and in as
where we also used
and a similar estimate for due to Lemma 3.9. Since is a weak- function of with norm uniformly bounded in , the weak Young’s inequality yield
To obtain the convergence of as , using the convergence of and the and , , convergence of (Lemma 3.12), one may easily reduce the problem to the convergence of
and that of a similar term of the other integral. It is easy to see via a rescaling that, for ,
while with the weak- norm equal to . Hence
for any with . Through a standard density argument and using the above uniform bound on the weak- norm of the convolution kernel, this convergence also holds for any . Therefore, we obtain and thus
The above estimates of and together yield the desired estimates of and its convergence as . The analysis on also follows from the above estimates with minor modifications, mostly replacing some by or outside the integrals, needed to control its logarithmic singularity caused by . We omit the details.
Finally, as in Lemma 5.1, stronger estimates and convergence can be obtained at due to prescribed boundary values (3.53). In fact,
implies
From the same procedure as in estimating and in the above, we obtain the desired estimate. Its convergence follows much as that of . The same argument applies to and the proof of the lemma is complete. ∎
The following is an estimate and in and their dependence on .
Lemma 5.4.
Proof.
As in the proof of Lemma 5.2, we first consider for and drop the subscript for notation simplification. Multiplying the Rayleigh equation (5.1a) by and integrating in both and , we have
The term can be estimated much as in the proof of Lemma 5.2 using Lemmas 3.9 and 5.3(2b)
Choose and such that
which is possible due to our assumption on , , and . The integral can be controlled by the Hölder continuity of and in , the weak Young’s inequality, and the (5.16) type interpolation inequality as
where is determined by . Therefore we obtain
The estimate of is much as in the proof of Lemma 5.3 based on the boundedness of the convolution operator on
where . Hence
Finally can be estimated exactly as in the proof of Lemma 5.2 (and also applying Lemma 5.3(2b)) and we have
Therefore, there exists such that and satisfy the desired estimates for and . For those , the term in the upper bound of can be controlled by Lemma 5.3 directly and thus the desired estimates are also satisfied by and . The estimate in the limiting case of can be obtained through the same weak convergence argument as in the proof of Lemma 5.2. ∎
5.2. Differentiation in 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 given in (5.2) with respect to .
Through straight forward calculations and applying Lemma 3.9, we obtain
Lemma 5.5.
We shall also consider the limit
(5.19) |
which exists for appropriate and satisfies the same estimates as (see Subsection 5.1).
In the rest of the subsection, we shall focus on the special case motivated by (2.11):
(5.20) |
where , , , and are all independent of . Our goal is to obtain the estimates of the derivatives of the solution to (5.1) in .
Proposition 5.6.
Assume , , (5.8), and (5.20). For any , , there exists depending on , , , , , and such that the solution to (5.1) satisfies that for any and ,
if , then
and, if , then
where
and all the norms are taken on . Moreover, as , the following hold.
-
(1)
Assume and , then for any , in , in , and in .
-
(2)
Assume and , then for any and , in , in , and in .
-
(3)
Assume and , then for any , also converges in to its limit .
Since is holomorphic in , . From the Rayleigh equation, singularity at the level of delta mass appears in along , , as . Therefore and also display such singularities which are singled out in the above estimates. The involved in the singular terms will be substituted by using the Rayleigh equation (5.1a) whenever necessary.
Proof.
The estimates on and , as well as the estimate of with , follow readily from (5.8), (5.20), Lemmas 3.9, 5.4, 5.3 (with , , and ), and 5.5. The estimate of is simply obtained from those of and . Moreover, for the rest of the proof of the proposition we shall also need the following inequality for which is also derived form Lemma 5.3 and Lemma 5.5 and uniform in
(5.21) |
The convergence of , , and follow directly from the continuity of (Lemma 4.1) and the convergence of and (Lemma 3.12) and (Lemma 5.3). Moreover, we also have the convergence of in .
In the following differentiations in are all carried out for . The convergence analysis based on the convergence results of and those of in Subsection 5.1 ensure that the estimates hold also for . Directly differentiating the Rayleigh equation (5.1a) in directly would cause worse singularity in the equation. Instead we first consider
(5.22) |
where is defined by as in (3.6). It satisfies
(5.23) |
where the singularity remains at the same level.
Estimating . Applying to (5.1a) and simplifying, we obtain
(5.24a) | ||||
where | ||||
and boundary conditions | ||||
(5.24b) | ||||
(5.24c) | ||||
where | ||||
Let and be the solution to the non-homogeneous Rayleigh equation (5.1a), but with zero boundary values in (5.1b), with replaced by and , respectively. Both are given by the formula (5.3). Using the estimates of derived in the above and apply Lemmas 5.2, we have
Moreover, from Lemma 5.1(2b) and (5.1b), (5.7), and (5.20), one can compute
where was substituted by using (5.1b). It along with the above estimates on implies
The estimate at based on Lemma 5.1(2b) is similar
which yields
From the convergence of and Lemma 5.1, as , we have the convergence of in , in , and in , for any and .
Due to the smoothness of , we apply Lemmas 5.4 and 5.3 instead to estimate
Again from Lemma 5.3, as , we have the convergence of in , in , for any , and in .
Finally, from (5.21) and (5.7), we have, for any ,
where again we substituted by (5.1b) and (5.7). Moreover, from the convergence of and , we have the convergence of in .
As plays the role of "" in the representation of as given in Lemma 5.5, the above estimates imply
(5.25) |
where the term was bounded by the other norms of via interpolation. The desired estimates on and follow from that of , (5.22), and the above inequality. We also obtain the estimate of from (5.25) which in turn yields the bound on . The convergence of is a direct consequence of those of , , , and the representation formula given in Lemma 5.5. Moreover, we also have the convergence of in for any .
To complete the estimates on and also for the next step, we also need the following inequalities which are also derived from the above estimates and Lemma 5.5
where the terms involving and , , are bounded by other norms of .
Estimating . In order to analyze , we still first apply to (5.24). Due to the commutativity (5.22) between and , the Rayleigh equation (5.1a) and (5.20) imply
We can write | ||||
(5.26a) | ||||
where was defined in (5.24) and | ||||
From (5.23) and the assumption , it holds and are in and with bounds uniform in . At , one can compute using (3.79), | ||||
From (5.20) and (5.1a), we can write | ||||
(5.26b) | ||||
At , we write | ||||
(5.26c) |
One may compute using (5.22) and (5.24c)
On the one hand, the turns out to involve some of the most singular terms in ,
We shall use (5.22) to replace and by and , the latter of which would produce . All those multiplied by can be substituted by (5.1a), but we keep other terms in the expression. On the other hand, we use (5.24a) to substitute in , which turns out to be rather regular due to the multiplier . Finally, we can write
where the functions , , are
and are at least in and .
The terms in and in generate the most singular part of which, based on Lemma 5.5, takes the form
Let
Clearly, it satisfies the same non-homogeneous Rayleigh equation (5.26a) and boundary conditions
(5.27) |
(5.28) |
Let and be the solutions to (5.26a) with zero boundary values in (5.1b) and non-homogeneous terms
respectively. Using the above estimates of and and applying Lemma 5.2, we obtain
As , the convergence of and implies that of in for any . From Lemma 5.1(2a), we obtain the convergence of in .
The boundary values of satisfy
where we also used the boundary conditions of and to express them in terms of and at . As , the convergence of and at implies that of in for any .
As plays the role of "" in the representation of as given in Lemma 5.5, the above estimates and Lemma 5.5 imply
where the term was bounded by the other norms of via interpolation. Finally, using (5.22) one can compute
(5.29) |
This relationship and the definition of and yield
Therefore the desired estimate on follows from those of , , and . The convergence of is also obtained much as that of . ∎
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 in both the horizontally -periodic (in ) case and the case. Under the assumption (4.9) of the absence of singular modes for all , 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 .
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 of (2.6) for a fixed wave number 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 and , we shall mostly work on estimates for in this subsection, unless otherwise specified.
Recall defined in (4.1). Denote the set of non-singular modes
(6.1) |
Throughout this subsection, we fix and assume (5.8). We shall also use (5.6), possibly after choosing smaller . The continuity of and (5.8) imply that is a finite set, which consists of only simple roots for large due to Lemma 4.2(3). We shall work on the following type of neighborhoods of
(6.2) |
where is given in (5.8).
Recall the Laplace transform of , 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) |
and , and are the Fourier transforms with respect to of the initial values , and . The solution to (2.11) is still given by Lemma 5.5 along with (5.20) and (6.3). More explicitly, if , then
(6.4) |
where are solutions to the homogeneous Rayleigh equation (3.24) satisfying initial conditions (3.53) and the solution to (5.1) given by (5.3) with and . The Laplace transform of can be computed by using (2.12) and the boundary condition (5.1b) along with (5.20), (6.3), and (5.6)
(6.5) |
We shall also need the following quantities
(6.6) |
where is the residue of a meromorphic function at . Apparently unless , or equivalently . The following lemma is obtained from applying the inverse Laplace transform.
Lemma 6.1.
From Lemma 4.2, implies and thus is well-defined for near . In part (2), similar types of formula and estimates of can be obtained from those of and (6.6). In the subsequent analysis, the limits of the above contour integrals as shrinks to will be taken and estimated whenever needed.
Proof.
From the definition (2.9) and the inverse Laplace transform formula (2.15), we have
where is chosen such that the above integrand is analytic for . Apparently is analytic in . In order to analyze for , we first consider and then for . From Lemma 3.3 and initial conditions (3.53), it holds that
Along with (6.3) and Lemma 3.9 which yields the boundedness of for , it implies
From Lemma 4.2(2) and again Lemma 3.9, we obtain333Through a more careful analysis we may obtain a Taylor expansion of in terms of as .
As , the Cauchy integral theorem yields
The desired expression of follows immediately from the residue calculation.
From the divergence free condition on the velocity, it holds that the Fourier transform (in ) of the velocity field satisfies . Therefore, we have
Corollary 6.1.1.
Under the assumptions of Lemma 6.1, we have
In the following lemma, we give some basic properties of and at some . Since is away from and and are analytic in a neighborhood of , the assumption (5.8) is not needed.
Lemma 6.2.
Assume , , and . Let be a root of (or equivalently, of defined in (4.2)) of degree , then the following hold.
-
(1)
is a solution to (2.8).
-
(2)
is a linear combination of , , and a linear combination of , , with coefficients depending on and . The leading terms of with are given by
(6.7) and the leading terms of is given by the above expression evaluated at .
-
(3)
If is a simple root of , i.e., , then and are given by the above expression and there exists determined only by and such that
for any , where we recall .
Proof.
According to Lemma 4.1(4), implies and thus is analytic in for near and the degree of as a root of both and is . By the definition of and the analyticity of , is an isolated root of . Let such that there are no other roots of in the disk centered at with radius . Using the fact that solves (2.11), one may compute
and thus (2.8a) is satisfied. Similar calculation also proves the boundary condition (2.8b) at . The zero boundary value at is obvious from that of at . Therefore statement (1) is proved.
To analyze in more details, let
and
From the initial conditions (3.53) of , it is straight forward to verify
Using the above expression to substitute in the residue (in the definition of ) and observing that the term cancels the singularity of for which results in an analytic function contributing nothing to the residue, we have
From definitions (5.3), (4.1), and (4.2), of , , and , (6.3), we have
Therefore this residue is a linear combination of , , with coefficients depending on and . The coefficients for are given by
The contributions of the terms involving and can be analyzed similarly (actually simpler as is not involved) and we obtain the desired statement (2) on the form of and .
If is a simple root of , i.e. , then and have only one term with and are constants in as given in statement (2). It along with Lemma 3.9 readily leads to its estimate. ∎
Corollary 6.2.1.
is also a solution to (2.8). Moreover if is a simple root of , then the corresponding eigenvalue is algebraically simple in the subspace of the -th Fourier modes.
Based on the above lemmas, it is natural to define
(6.8) |
The following lemma gives that defines the invariant spectral projection to the eigenspace of spanned by , .
Lemma 6.3.
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 of solutions to (2.6), the in the definition of can be replaced by any . From Lemma 6.2, all solutions are polynomials of of degree no more than . It is standard to show inductively that consists of all possible coefficients of , which can be computed to be generated by , , using (6.7) and the relationship between and . The invariance of under (2.6) is due to the fact that are solutions to (2.6). To show , let . With this initial value, the solution is simply the component of the solution with the initial value . Hence takes the form given in Lemma 6.2(2). Its Laplace transform is analytic at all and thus the component of is equal to itself. Therefore we obtain . Finally the description of the kernel of is obvious due to the fact that both and are solutions. ∎
Remark 6.1.
In particular, if
then straight forward verification yields
From Lemma 2.3, is an eigenvalue (with the above eigenfunctions generated by ) of the linearized capillary gravity water wave at the shear flow, which has to be geometrically simple when restricted to the -th Fourier mode in . Its algebraic multiplicity is equal to the degree of the root of . The eigenfunctions of the linearized irrotational capillary gravity wave are generated by . From Lemmas 3.2(1) (with , , , and ) and 4.2(3), it is straight forward to estimate that, after normalizing the norm of to be , the and differences in the and components, respectively, between the eigenfunctions of (2.6) and the irrotational capillary gravity waves linearized at zero is of order as .
In the rest of this subsection we consider and . We shall always work on . We first present some properties of and . Let us keep in mind that for analytic functions, and are equivalent.
Lemma 6.4.
It holds that and are analytic in and satisfy
Assume , , and (5.8), then the following hold for some .
-
(1)
For any , there exists determined only by , , , , and (independent of ) such that for any ,
if , then
and if , then
where
and all the norms are taken on .
-
(2)
As , on
exist and the following hold.
-
(a)
Assume and , then for any , in , in , and and in .
-
(b)
Assume and , then for any and , in , in , and and in .
-
(c)
Assume and , then for any , converges to its limit in .
-
(a)
Compared to Proposition 5.6, the modifications in the definition of is to make it analytic in which will make it more convenient in applying Lemma 6.5 in the below.
Proof.
The estimates of , , , , and their convergences are all direct corollaries of (6.3) and Proposition 5.6. The estimate of and its convergence follows from the second expression of (6.5) and the above properties of .
We also notice that, compared to Propositions 5.6, in the definition of as well as in the estimate related to , the in front of , , and had been replaced by , , and , respectively. This modification brings at most minor changes to the upper bounds. In fact,
for , where the Rayleigh equation was also used. This error bound and the estimate on are then used to obtain the desired inequality on . The term in is handled by the same argument. Similarly,
and this along with the estimate on yields the estimate on . It remain the consider the modifications to the correction terms in at and . Similarly,
which is controlled using due to lemma 3.9. The last remaining modification from to can be justified by the same argument (even easier as .)
Finally we consider for in . From (6.5), one may compute
where we used the Rayleigh equation (5.1a) in the last step. Therefore from Proposition 5.6 we have, for any and ,
The last terms can be controlled by the previous two terms, which completes the estimate on . The convergence of also follows from those of , and . ∎
The following lemma will be used in the decay estimates.
Lemma 6.5.
Suppose is an integer, , and are analytic functions on
and there exists such that for all , then there exists depending only on such that, for any ,
Proof.
Integrating by parts we have, for any ,
For any , the boundedness (for ) of the Fourier transform implies
From this inequality and the Cauchy integral theorem, we obtain, for any ,
Letting , the analyticity assumption of and implies all those terms on the vertical boundary of vanish and the above estimates on the integrals along the horizontal edges yield
The lemma follows by letting . ∎
Remark 6.2.
In the following applications of this lemma, we often use the norm to control the norm. This leads to fact that the regularity requirements in (i.e. the exponents of ) may not be close to optimal.
Applying the above lemma, we first obtain the decay of and .
Lemma 6.6.
Assume , , and (5.8), then for any , , and integer , there exists determined only by , , , , , , and (independent of ) such that
and if ,
Proof.
In the following we shall focus on , , and , where is the Fourier transform (in ) of the vorticity of . In order to characterize their asymptotic behavior, define
(6.9) |
In the above expression, exactly one of and is equal to and the other equal to . The dependence of on is only through its sign, so we skipped specifying the dependence. We also notice that may not be at . The available conjugacy properties of are not sufficient to imply . We shall see that provides the asymptotic profile of the vorticity . We first give the following some basic properties of .
Lemma 6.7.
Assume and (5.8), then and, for any , there exists determined only by , , , , and (independent of ) such that
Proof.
The conjugacy relation of is clear from its definition. According to Lemma 6.4, satisfies the same estimates as for . We have, for and ,
(6.10) |
which implies the estimate of . Apparently the estimate of depends on that of
where was defined in (5.22). From (5.25) and (6.3), we have
which yields and completes the proof of the lemma. ∎
In the following lemma, we obtain the leading order terms of , , and .
Lemma 6.8.
Assume and (5.8), then, for any , , and integer , there exists determined only by , , , , , , and (independent of ) such that
Remark 6.3.
This lemma also implies, for any integer ,
while for , there is another term on the left side. The form in the lemma is more consistent with other estimates including that of to be given in the following, however.
Proof.
The definition of implies, for each and ,
Applying Lemma 6.5 with and and and Lemma 6.4, we obtain
(6.11) |
In the rest of the proof, we shall focus on the integral involving which also yields the other desired estimates. Substituting the term by the Rayleigh equation (2.11a) and applying the Cauchy Integral Theorem yield
Since is bounded by on , we can control those terms using Lemma 6.5 and obtain
(6.12) |
where we also changed the variable in the last integral and
(6.13) |
It remains to handle this integral term and we shall identify its leading terms.
Fix . We first let
for any , where , Lemma 6.4 was used. In the rest of the proof of this lemma, we use , , to denote the left, right, top, bottom sides of the rectangle with the counterclockwise orientation. For any and , and , integrating by parts and using the boundedness (for ) of the Fourier transform, we obtain
where is taken along which in the upper half plane. Next from Lemma 6.4 we have
The above error analysis implies that the main contribution of the integral along would come from the product
For any , on the one hand,
which is useful for . On the other hand,
The first integral can be evaluated as by using the Cauchy Integral Theorem. Integrating the second integral (in the way opposite to the above) we obtain
Therefore
Along with (6.10), we have
The integrals along the vertical sides of converge to as as all the integrands are smooth there. The integrals along , , can be treated much as in the above. Recall . Letting , the Cauchy Integral Theorem and the above error analysis imply
Taking , it follows from the above inequality and (6.12)
(6.14) |
Along with (6.11) and Lemma 6.6 it implies the desired estimate of . The estimates on and are also obtained from the above inequality and Lemma 6.6. ∎
Finally we consider .
Lemma 6.9.
Assume and (5.8), then, for any , , and integer , there exists determined only by , , , , , and (independent of ) such that
where
(6.15) |
(6.16) |
Remark 6.4.
In the above lemmas, we also notice , . The leading order terms and represent the contribution from the rigid bottom and the water surface, while the asymptotic vorticity from the fluid interior. In the fixed boundary problem for with slip boundary condition on both horizontal boundaries, and would take similar forms and would be similar to . See Subsection 6.4.
Proof.
The definition of implies, for each and ,
Let and
with defined in Lemma 6.4. Applying Lemma 6.5 with and Lemma 6.4, we obtain
Substituting in by using the Rayleigh equation (2.11a) yields
where
Again the terms involving not being divided by can be estimated by using assumption (5.8) and Lemmas 6.5, 3.9, and 6.4 and we have
We shall identify the principle contributions from the terms , , and following a similar strategy and use the same notations , , as in the proof of Lemma 6.8, with necessary modifications to treat the contributions from the .
Fix . We start with by letting
From assumption (5.8), Lemmas 3.12, 3.18(2b), 4.2, and 6.4, we have, for any ,
The next step is the same argument via integrating by parts in as in the proof of Lemma 6.8,
From (5.8) and Lemmas 3.9, 3.14–3.16, 4.1(3), and 4.2(1), one may estimate,
Along with Lemma 6.4, it implies, for any ,
where and were bounded by the norms of , , , and . Consequently, for any
As in the proof of Lemma 6.8, by considering contour integrals, we have
Again, since is bounded for , the above estimate implies
The contributions from the integral along can be treated similarly and using the conjugacy relation, while the integrals along the vertical boundaries of vanish as . Using the Cauchy Integral Theorem, combining the above analysis, letting , and then , we obtain
where
We give closer look at . From boundary condition (5.1b), (5.20), and (6.3),
and thus
Since for , we obtain and hence leads to the desired form (6.15) of . The term involving can be analyzed similarly (actually slightly simpler due to ) using Lemmas 3.14–3.16 and 3.18. The term involving can be estimated much as in (6.14). Summarizing this estimates we obtain
Combining it with Lemma 6.6, the desired estimate on follows. ∎
6.2. Linearized capillary gravity waves in the horizontally periodic case of
In this subsection, we consider the case where the system is periodic in with wave length . In this case
where the latter properties is due to the divergence free condition on . For , let
where , , and are defined in Lemma 6.1 and Corollary 6.1.1. Here we used (2.7) that the zeroth modes and are invariant in . Throughout this subsection, we assume (4.9) holds for .
We first give the decay estimates of based on Lemma 6.6–6.9. In particular, for the estimates of and , recall and , defined in (6.9), (6.16), and (6.15), respective. Let
(6.17) |
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 implies (5.8) holds for all with constants and uniform in . Therefore from the definition of and Lemma 6.6, it is straight forward to estimate
The desired inequality follows from . The estimates on , and are obtained similarly. The inequalities on and are obtained by applying Lemma 6.8 and 6.9 through a similar procedure. The estimates on and , , follow directly from the formula and estimates of their each Fourier modes given in those lemmas and
(6.18) |
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 , satisfied by in . The boundary conditions of and are direct corollaries of their definitions and the boundary conditions (3.53) of .
Next we consider the part of the linear solution . Let
(6.19) |
where the lower bounds are obtained due to the roots for large (Lemma 4.2(3)).
Proof of Theorem 2.1(3). On the one hand, according to Lemma 4.2(3), there exists such that with simple roots for all . On the other hand, (4.9) and Proposition 4.4 imply that (5.8) holds for all . Along with Lemma 4.2(2), we obtain that, for all with , the set of roots is contained in a subset in the domain of analyticity of uniformly in such . Hence is a discrete set and the total algebraic multiplicity of for all with is finite. This proves .
For any and , let denote the degree of as a root of , then and are polynomials of of degree (Lemma 6.2). Hence to prove the regularity estimates, we only need to consider with where all roots of are simple. For such , and Lemma 4.2(3) implies that there exists such that
From the homogeneous Rayleigh equation (3.1), (5.8), and Lemma 3.9, it holds,
(6.20) |
Hence Lemmas 6.1 and 6.2 and the definition of imply, for any and ,
The desired inequality follows from the divergence free condition. The expression of involves and thus it can be differentiated in at most times. The procedure to obtain the estimates of and are similar and we skip the details.
Finally we give the invariant decomposition of the phase space which proves Theorem 2.1(4).
Lemma 6.10.
The boundedness of follows from the estimates in Theorem 2.1 at . The invariance of and is due to Lemma 6.2 and Corollary 6.2.1. The well-posedness of (2.6) on the completion of is due to the fact that except for finitely many . Here we did not set and in is due to the issue that we can not ensure for .
6.3. Linearized capillary gravity waves in the horizontally infinite case of
In this subsection, we consider the case where and thus . Throughout this subsection, we assume (4.9) for . For , let
(6.21) |
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 implies that (5.8) holds and with all these simple roots of away from for all . Moreover, Lemma 4.2 yields
Like in the periodic-in- case, the proof of the decay of is also a direct verification using Lemmas 6.6–6.9 along with (6.18) and the divergence free condition. We omit the details.
From Lemmas 6.1 and 6.2(3), we obtain and are independent of and satisfy, for any ,
The desired estimates follow from (6.20), , and direct computations.
Similar to the periodic case, we also have the decomposition by invariant subspaces.
Lemma 6.11.
To end this subsection we show that, under assumptions (4.9) for and (4.18) for , due to the monotonicity of in (Lemma 4.7) and the asymptotics of for (Lemma 4.2(3)), the dynamics of the non-singular modes is conjugate to that of linear irrotational capillary gravity waves.
For , let
where is the wave speed of the free linear capillary gravity wave (system (1.3) with and ) given in (2.4). Here correspond to the two non-singular modes in the -th Fourier modes in , while the modes of irrotational linear capillary gravity waters waves. Define
Clearly is equal to the completion of and and parametrize and by . The following proposition finishes the proof of Theorem 1.1(2b) and Theorem 2.2(4).
Proposition 6.12.
Assume and (4.9) for , then the following hold.
-
(1)
The mappings and are isomorphisms. Moreover there exists depending only on such that
- (2)
- (3)
Proof.
The estimates on and are derived from direct computations based on Lemma 3.9. In particular, since , formula (4.13) of for and the bound on are used in obtaining the lower bounds of for close to . The estimates of and follow from those of and and the Parseval’s identity. Statement (2) is a direct consequence of Lemma 2.3 and the definition of and .
Since and , under the additional assumptions (4.18) and , Proposition 4.4 and Lemma 4.7 imply that a.) both and are odd in , b.) both and have positive derivative for , and c.) both are of the order for and of the order for . Hence exist and satisfy the estimates, which implies the boundedness of . The conjugacy relation (6.25) can be verified directly using (6.23), (6.24), and the definition of . ∎
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 with slip boundary condition at . Let be a shear flow and we assume
() |
As in the literatures, singular modes mean linearized solutions in the form of with and .
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 , the main modifications are: a.) replacing and by and which solve the homogeneous and non-homogeneous Rayleigh equations satisfying boundary conditions
respectively, and b.) replacing by . For the simplification of notations, we also use , , and to denote their limits as . In this case of channel flow with fixed boundary, obviously the set of non-singular modes (roots of outside ) for all is finite, actually empty if . Assuming (), through the same procedure as in Lemma 6.1, the solution to the linearized Euler equation at the shear flow can also be split into
associated to the continuous spectra and point spectra. Under assumption (), belongs to the eigenspace of unstable modes which is finite dimensional if . Let
and and , , be defined as in (6.17) for the -periodic-in- case and in (6.22) for the case of .
Theorem 6.13.
Assume , , and () holds for all where or , then, for any , , , , and integer , there exists depending only on , , , and such that any solution with satisfy
if , then
and if , then
Moreover,
Finally, , , satisfy and
Remark 6.6.
In the case of the Couette flow , assumption () is satisfied. Obviously , which in fact gives the whole linearized vorticity and the leading asymptotic terms of and . However, does also include contributions and 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 -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.