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

On the spectra of the gravity water waves linearized at monotone shear flows

Xiao Liu Department of Mathematics, University of Illinois, 605 E Springfield Ave., Champaign, IL 61820 [email protected]  and  Chongchun Zeng School of Mathematics, Georgia Institute of Technology, Atlanta, GA 30332 [email protected]
Abstract.

We consider the spectra of the 2-dim gravity waves of finite depth linearized at a uniform monotonic shear flow U(x2)U(x_{2}), x2(h,0)x_{2}\in(-h,0), where the wave numbers kk of the horizontal variable x1x_{1} is treated as a parameter. Our main results include a.) a complete branch of non-singular neutral modes c+(k)c^{+}(k) strictly decreasing in k0k\geq 0 and converging to U(0)U(0) as kk\to\infty; b.) another branch of non-singular neutral modes c(k)c_{-}(k), k(k,k)k\in(-k_{-},k_{-}) for some k>0k_{-}>0, with c(±k)=U(h)c_{-}(\pm k_{-})=U(-h); c.) the non-degeneracy and the bifurcation at (k,c=U(h))(k_{-},c=U(-h)); d.) the existence and non-existence of unstable modes for cc near U(0)U(0), U(h)U(-h), and interior inflection values of UU; e.) the complete spectral distribution in the case where U′′U^{\prime\prime} does not change sign or changes sign exactly once non-degenerately. In particular, UU is spectrally stable if UU′′0U^{\prime}U^{\prime\prime}\leq 0 and unstable if UU has a non-degenerate interior inflection value or {UU′′>0}\{U^{\prime}U^{\prime\prime}>0\} accumulate at x2=hx_{2}=-h or 0. Moreover, if UU is an unstable shear flow of the fixed boundary problem in a channel, then strong gravity could cause instability of the linearized gravity waves in all long waves (i.e. |k|1|k|\ll 1).

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

1. Introduction

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

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

or

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

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

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

where g>0g>0 is the gravitational acceleration and the constant density is normalized to be 1. The kinematic boundary condition (1.1c) is equivalent to that t+v\partial_{t}+v\cdot\nabla is tangent to {(t,x)xSt}\{(t,x)\mid x\in S_{t}\} at any xStx\in S_{t}, which means that StS_{t} moves at the velocity restricted to StS_{t}.

Shear flows are a fundamental class of stationary solutions of laminar flows

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

Our goal in this paper is to analyze the spectral distribution (and thus the spectral stability and the linear instability) of the gravity water wave system linearized at a monotonic shear flow satisfying

(H) UCl0([h,0]),l06;U(x2)>0,x2[h,0].U\in C^{l_{0}}([-h,0]),\;l_{0}\geq 6;\quad\;U^{\prime}(x_{2})>0,\ \forall x_{2}\in[-h,0].
Remark 1.1.

Due to the symmetry of horizontal reflection

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

the case of U<0U^{\prime}<0 is completely identical except the sign assumptions on U′′U^{\prime\prime} and U′′′U^{\prime\prime\prime} in the theorems should be reversed.

1.1. Linearization and the spectral problem

We first derive the linearized system of (1.1) at the shear flow (v=(U(x2),0)T,η=0)(v_{*}=(U(x_{2}),0)^{T},\eta_{*}=0) given in (1.2) satisfied and we denote the linearized solutions by (v,η,p)(v,\eta,p). Let (Stϵ,vϵ(t,x),pϵ(t,x))(S_{t}^{\epsilon},v^{\epsilon}(t,x),p^{\epsilon}(t,x)) be a one-parameter family of solutions of (1.1) with (St0,v0(t,x),p0(t,x))=(S,v,p)(S_{t}^{0},v^{0}(t,x),p^{0}(t,x))=(S_{*},v_{*},p_{*}). Following the same procedure as in [17], differentiating the Euler equation system (1.1) with respect to ϵ\epsilon and then evaluating it at ϵ=0\epsilon=0 yield

(1.3a) tv+U(x2)x1v+(U(x2)v2,0)T+p=0,v=0,x2(h,0),\partial_{t}v+U(x_{2})\partial_{x_{1}}v+(U^{\prime}(x_{2})v_{2},0)^{T}+\nabla p=0,\quad\;\;\nabla\cdot v=0,\quad\;x_{2}\in(-h,0),
(1.3b) p=2U(x2)x1v2,x2(h,0), and x2p|x2=h=0.-\triangle p=2U^{\prime}(x_{2})\partial_{x_{1}}v_{2},\quad x_{2}\in(-h,0),\;\text{ and }\;\partial_{x_{2}}p|_{x_{2}=-h}=0.
(1.3c) tη=v2|x2=0U(0)x1η.\partial_{t}\eta=v_{2}|_{x_{2}=0}-U(0)\partial_{x_{1}}\eta.
(1.3d) p=gη, at x2=0.p=g\eta,\quad\text{ at }\ x_{2}=0.

Observe that the variable coefficients in the linear system (1.3) depend only on x2x_{2} and thus the Fourier modes of x1x_{1} evolve in tt independently of each others. The zeroth Fourier mode corresponds to the perturbed shear flow component and can be handled easily, so the main interest lies in the kk-th Fourier modes with k0k\neq 0. To find eigenvalues and eigenfunctions, we consider linearized solutions in the form of

(1.4) v(t,x)=eik(x1ct)(v^1(x2),v^2(x2)),η(t,x1)=eik(x1ct)η^,p(t,x)=eik(x1ct)p^(x)v(t,x)=e^{ik(x_{1}-ct)}\big{(}\hat{v}_{1}(x_{2}),\hat{v}_{2}(x_{2})\big{)},\;\;\eta(t,x_{1})=e^{ik(x_{1}-ct)}\hat{\eta},\;\;p(t,x)=e^{ik(x_{1}-ct)}\hat{p}(x)

where apparently the eigenvalues take the form λ=ikc\lambda=-ikc with the wave speed c=cR+icIc=c_{R}+ic_{I}\in\mathbb{C}. In seeking solutions in the form of (1.4), the wave number kk\in\mathbb{R} is often treated as a parameter. Substituting (1.4) into the linearization (1.3), through straight forward calculations (see [28, 10] as well as [17]), one obtains the boundary value problem of the standard Rayleigh equation

(1.5a) v^2′′+(k2+U′′Uc)v^2=0,-\hat{v}_{2}^{\prime\prime}+\Big{(}k^{2}+\frac{U^{\prime\prime}}{U-c}\Big{)}\hat{v}_{2}=0,
with the boundary conditions
(1.5b) v^2(h)=0,\hat{v}_{2}(-h)=0,
(1.5c) ((Uc)2v^2(U(Uc)+g)v^2)|x2=0=0.\big{(}(U-c)^{2}\hat{v}_{2}^{\prime}-(U^{\prime}(U-c)+g)\hat{v}_{2}\big{)}\big{|}_{x_{2}=0}=0.

Here and throughout the paper denotes the derivative with respect to x2x_{2}. Apparently the last boundary condition at x2=0x_{2}=0 differs it from that of the linearized channel flow for x2(h,0)x_{2}\in(-h,0) with the slip boundary conditions at x2=h,0x_{2}=-h,0. It actually makes a substantial difference as we shall see below. For k0k\neq 0, v^1\hat{v}_{1} and η^\hat{\eta} can be determined by v^2\hat{v}_{2} using the divergence free and the kinematic condition, respectively,

kv^1+v^2=0,η^=v^2(0)ik(U(0)c)).-k\hat{v}_{1}+\hat{v}_{2}^{\prime}=0,\qquad\hat{\eta}=\frac{\hat{v}_{2}(0)}{ik(U(0)-c))}.

So we shall mainly focus on v^2\hat{v}_{2}.

The linear system (1.3) is linearly unstable if there exist solutions (k,c,v^2)(k,c,\hat{v}_{2}) to (1.5) with cI>0c_{I}>0, which appear in conjugate pairs. Obviously the Rayleigh equation becomes singular when cU([h,0])c\in U([-h,0]), otherwise its initial value problem depends on cc and k2k^{2} analytically and thus also even in kk. We recall the following standard terminology.

Definition 1.1.

(k,c)(k,c) is a non-singular mode if cU([h,0])c\in\mathbb{C}\setminus U([-h,0]) and there exists a non-trivial solution y(x2)y(x_{2}) to (1.5) (thus also yields a solution to (1.3) in the form of (1.4)). (k,c)(k,c) is a singular mode if cU([h,0])c\in U([-h,0]) and

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

has a non-trivial Hx22H_{x_{2}}^{2} solution y(x2)y(x_{2}) satisfying the boundary conditions (1.5b) and (1.5c). A non-singular mode is a stable (or unstable) mode if cI0c_{I}\leq 0 (or cI>0c_{I}>0). (k,c)(k,c) is a neutral mode if it is a singular or non-singular mode and cc\in\mathbb{R}.

In all above cases, (1.3) has a solution in the form of (1.4).

We first present our main results and then comment on them along with the literatures.

1.2. Main results

In this paper we focus on the spectral distribution of the gravity waves linearized at a uniformly monotonic shear flow. It serves as the first step for us to understand the linearized flow at monotonic shear flows including the possible linear inviscid damping and the dispersive character (see [17] for capillary gravity waves). This would lay the foundation for the study of the local nonlinear dynamics. The following is a non-technical summary (assuming U>0U^{\prime}>0) of the main results mostly given in the following three main theorems.

  • There is a complete upper branch of neutral modes c+(k)>U(0)c^{+}(k)>U(0) decreasing in k>0k>0 and converging to U(0)U(0) in the same fashion as the dispersion relation of the free gravity wave as k+k\to+\infty.

  • Possibly two more branches of eigen-modes for |k|1|k|\lesssim 1 and |k|1|k|\gg 1, respectively.

    • There is a branch of neutral modes c(k)<U(h)c_{-}(k)<U(-h) increasing in k[0,k]k\in[0,k_{-}] and reaches U(h)U(-h) at some k>0k_{-}>0 and then, near c=U(h)c=U(-h) and for 0<kk10<k-k_{-}\ll 1, it either bifurcates into a (possibly broken) branch of unstable modes whose real parts are given by {U(x2)U(x2)U′′(x2)>0}\{U(x_{2})\mid U^{\prime}(x_{2})U^{\prime\prime}(x_{2})>0\} near U(h)U(-h) or the branch disappears completely if UU′′<0U^{\prime}U^{\prime\prime}<0 near x2=hx_{2}=-h.

    • Unless UU′′<0U^{\prime}U^{\prime\prime}<0 near x2=0x_{2}=0, there is a (possibly broken) branch c(k)c^{-}(k) of "extremely weakly" unstable modes converging to U(0)U(0) as k+k\to+\infty, whose real parts are given by {U(x2)U(x2)U′′(x2)>0}\{U(x_{2})\mid U^{\prime}(x_{2})U^{\prime\prime}(x_{2})>0\} near U(0)U(0).

  • Inflection values of UU in U((h,0))U\big{(}(-h,0)\big{)} always yield one or two singular neutral modes. Bifurcation to unstable modes occurs with cc on one side of any non-degenerate interior inflection value.

  • {U′′U>0}\{U^{\prime\prime}U^{\prime}>0\}\neq\emptyset is necessary and also "almost" sufficient for linear instability.

  • If U′′U>0U^{\prime\prime}U^{\prime}>0 on [h,0][-h,0], then c(k)c^{-}(k) connects to c(k)c_{-}(k) and, along with c+(k)c^{+}(k) and c(k)¯\overline{c^{-}(k)}, give all the eigen-modes (singular and non-singular modes) and (1.3) is spectrally unstable for all |k|>k|k|>k_{-}.

  • If U′′U0U^{\prime\prime}U^{\prime}\leq 0 on [h,0][-h,0], then (1.3) is spectrally stable and c(k)c_{-}(k) for |k|k|k|\leq k_{-}, and c+(k)c^{+}(k), are all the eigen-modes besides those at interior inflection values.

  • If U′′U^{\prime\prime} changes from negative to positive exactly once non-degenerately, then c(k)c^{-}(k) connects to the inflection value at |k|=k0>k|k|=k_{0}>k_{-}, and, along with c+(k)c^{+}(k) and c(k)¯\overline{c^{-}(k)}, give all the eigen-modes. Hence (1.3) is spectrally unstable for all |k|(k,k0)|k|\in(k_{-},k_{0}).

  • If U′′U^{\prime\prime} changes from positive to negative instead, the eigenvalue distribution is also obtained. In particular, in some cases (1.3) with sufficiently strong gravity is spectrally unstable for all long waves (i.e. |k|1|k|\ll 1).

  • Some improvements to those results in [17] on the spectral distribution of the capillary gravity water waves linearized at monotonic shear flows (Proposition 3.16).

The following first main theorem is on the above branches c±(k)c^{\pm}(k) and c(k)c_{-}(k), the singular neutral modes, as well as the bifurcations near singular neutral modes.

Theorem 1.1.

Assume UC6U\in C^{6} and U>0U^{\prime}>0 on [h,0][-h,0], then singular neutral modes (k,c)(k,c) occur only if c=U(h)c=U(-h) or cc is an inflection value of UU. Moreover the following hold for a constant C>0C>0 determined only by UU.

  1. (1)

    There exists k0>0k_{0}>0 such that there exist an analytic function c+(k)(U(0),+)c^{+}(k)\in(U(0),+\infty) and a C4C^{4} function c(k)=cR(k)+icI(k)c^{-}(k)=c_{R}^{-}(k)+ic_{I}^{-}(k) both defined for |k|k0|k|\geq k_{0} and even in kk and the following hold for any |k|k0|k|\geq k_{0}.

    1. (a)

      cI(k)c_{I}^{-}(k) and U′′(U1(cR(k))U^{\prime\prime}(U^{-1}\big{(}c_{R}^{-}(k)\big{)} have the same signs (+,+,-, or 0) and

      lim|k|(c±(k)U(0))/g/|k|=±1,\displaystyle\lim_{|k|\to\infty}\big{(}c^{\pm}(k)-U(0)\big{)}/\sqrt{g/|k|}=\pm 1,
      C1|U′′(U1(cR(k))||k|32e2g|k|U(0)|cI(k)|C|U′′(U1(cR(k))|.\displaystyle C^{-1}\big{|}U^{\prime\prime}(U^{-1}(c_{R}^{-}(k)\big{)}\big{|}\leq|k|^{\frac{3}{2}}e^{\frac{2\sqrt{g|k|}}{U^{\prime}(0)}}|c_{I}^{-}(k)|\leq C\big{|}U^{\prime\prime}(U^{-1}(c_{R}^{-}(k)\big{)}\big{|}.
    2. (b)

      (k,c)(k,c) is a non-singular or singular mode iff i.) c=c+(k)c=c^{+}(k), ii.) c=c(k)c=c^{-}(k) and cI0c_{I}\geq 0, or iii.) c=c(k)¯c=\overline{c^{-}(k)} and cI0c_{I}\leq 0.

    3. (c)

      c+(k)c^{+}(k) can be extended to an even and analytic function of kk\in\mathbb{R} with c+(k)(U(0),c+(0)]c^{+}(k)\in\big{(}U(0),c^{+}(0)\big{]} and (c+)(k)<0(c^{+})^{\prime}(k)<0 for k>0k>0. Moreover, for any kk\in\mathbb{R}, c+(k)c^{+}(k) is the only non-singular mode in (U(0),+)(U(0),+\infty) and it corresponds to an eigenvalue ikc+(k)-ikc^{+}(k) of (1.3).

  2. (2)

    There exist a unique k>0k_{-}>0 and a C1,αC^{1,\alpha} (for any α[0,1)\alpha\in[0,1)) even function c(k)(,U(h)]c_{-}(k)\in(-\infty,U(-h)] defined for |k|k|k|\leq k_{-} such that the following hold.

    1. (a)

      c(k)<U(h)c_{-}(k)<U(-h) is analytic in k(k,k)k\in(-k_{-},k_{-}), c(±k)=U(h)c_{-}(\pm k_{-})=U(-h), and (c)(k)>0(c_{-})^{\prime}(k)>0 for all k(0,k]k\in(0,k_{-}].

    2. (b)

      (k,c)(k,c) is a non-singular mode with cc+(k)c\neq c^{+}(k) outside the disk (1.8) iff |k|<k|k|<k_{-} and c=c(k)c=c_{-}(k).

    3. (c)

      There exist ϵ,ρ>0\epsilon,\rho>0 such that c(k)c_{-}(k) can be extended to [kϵ,k+ϵ][-k_{-}-\epsilon,k_{-}+\epsilon] as a complex valued C1,αC^{1,\alpha} even function satisfying the following.

      1. (i)

        For |k|[k,k+ϵ]|k|\in[k_{-},k_{-}+\epsilon], Rec(k)>0\text{Re}\,c_{-}^{\prime}(k)>0, Imc(k)\text{Im}\,c_{-}(k) and U′′(U1(Rec(k)))U^{\prime\prime}\big{(}U^{-1}(\text{Re}\,c_{-}(k))\big{)} have the same signs (+,+,-, or 0), and

        C1|U′′(U1(Rec(k))|(|k|k)2|Imc(k)|C|U′′(U1(Rec(k))|(|k|k)2.C^{-1}\big{|}U^{\prime\prime}(U^{-1}(\text{Re}\,c_{-}(k)\big{)}\big{|}(|k|-k_{-})^{2}\leq|\text{Im}\,c-(k)|\leq C\big{|}U^{\prime\prime}(U^{-1}(\text{Re}\,c_{-}(k)\big{)}\big{|}(|k|-k_{-})^{2}.
      2. (ii)

        (k,c=cR+icI)(k,c=c_{R}+ic_{I}) is a singular or non-singular mode of (1.3) with |kk|ϵ|k-k_{-}|\leq\epsilon, |cU(h)|ρ|c-U(-h)|\leq\rho, and cI0c_{I}\geq 0 iff c=c(k)c=c_{-}(k) (cI0c_{I}\leq 0 iff c=c(k)¯c=\overline{c_{-}(k)}).

  3. (3)

    Suppose c0=U(x20)c_{0}=U(x_{20}) with x20(h,0)x_{20}\in(-h,0) and U′′(x20)=0U^{\prime\prime}(x_{20})=0, then the following hold.

    1. (a)

      There exists S[0,+)S\subset[0,+\infty) such that |S|=1|S|=1 or 22, maxS>0\max S>0, and (k,c0)(k,c_{0}) is a singular neutral mode of (1.3) iff kSk\in S.

    2. (b)

      If, in addition, U′′′(x20)0U^{\prime\prime\prime}(x_{20})\neq 0, then, for any kS{0}k_{*}\in S\setminus\{0\}, there exist ϵ,ρ>0\epsilon,\rho>0 and a C4C^{4} complex valued function 𝒞(k)\mathcal{C}(k) defined on [kϵ,k+ϵ][-k_{*}-\epsilon,k_{*}+\epsilon] satisfying the following.

      1. (i)

        𝒞I(k)\mathcal{C}_{I}(k) has the same sign as (kk)U′′′(x20)(k-k_{*})U^{\prime\prime\prime}(x_{20}) and

        C1|kk||𝒞I(k)|C|kk|.C^{-1}|k-k_{*}|\leq|\mathcal{C}_{I}(k)|\leq C|k-k_{*}|.
      2. (ii)

        (k,c)(k,c) is a singular or non-singular mode of (1.3) with |kk|ϵ|k-k_{*}|\leq\epsilon, |cc0|ρ|c-c_{0}|\leq\rho, and cI0c_{I}\geq 0 iff (kk)U′′′(x20)0(k-k_{*})U^{\prime\prime\prime}(x_{20})\geq 0 and c=𝒞(k)c=\mathcal{C}(k) (cI0c_{I}\leq 0 iff c=𝒞(k)¯c=\overline{\mathcal{C}(k)} and (kk)U′′′(x20)0(k-k_{*})U^{\prime\prime\prime}(x_{20})\geq 0).

  4. (4)

    The linearized gravity water wave system (1.3) is linearly unstable for some wave number kk\in\mathbb{R} if U′′U^{\prime\prime} has a non-degenerate zero point in (h,0)(-h,0) or there exists a sequence x2,n(h,0)x_{2,n}\in(-h,0) converging to h-h or 0 as n+n\to+\infty such that U′′(x2,n)>0U^{\prime\prime}(x_{2,n})>0 for all nn.

Clearly the above c(k)c^{-}(k), c(k)c_{-}(k), and 𝒞(k)\mathcal{C}(k) are relevant only when their imaginary parts are non-negative, so the subsets of these branches corresponding to eigenvalues are possibly broken. They correspond to unstable modes iff the imaginary parts are positive. From the estimates of the imaginary parts, the strength of the instability (the exponential growth rates) is the strongest near non-degenerate inflection values, and the weakest near U(0)U(0).

Remark 1.2.

a.) Due to the symmetry, the case of U<0U^{\prime}<0 (and also in the following theorems) is completely identical except the signs of U′′U^{\prime\prime} and U′′′U^{\prime\prime\prime} should be reversed.
b.) More detailed asymptotics of c±(k)c^{\pm}(k) can be found in Lemma 3.15.
c.) See Lemma 3.7 for more details on when |S|=1|S|=1 or 22 in statement (3) and its relationship to the singular neutral mode of the linearized channel flow. The statement SS\neq\emptyset had also been proved in [28, 11] and we gave a different proof here. It actually holds for a wider class of shear flows not necessarily monotonic, see Remark 3.2. When |S|=2|S|=2, see Lemma 3.12 for how a closed branch of unstable modes emerges from the smaller wave number and makes the linearized system unstable for |k|1|k|\ll 1.
d.) While U′′′(x20)0U^{\prime\prime\prime}(x_{20})\neq 0 is the most easily checked non-degeneracy condition in statement (3b), a more precise condition and more details of the bifurcation (including the case of k=0k_{*}=0) can be found in Lemma 3.11.
e.) Local bifurcation in statement (4) from an interior inflection value, similar to [17], can be compared to that in [10] for a different class of shear flows. Also see comment below.

In the next theorem we consider the case where U′′U^{\prime\prime} does not change sign.

Theorem 1.2.

Assume UC6U\in C^{6} and U>0U^{\prime}>0 on [h,0][-h,0], then the following hold.

  1. (1)

    Suppose U′′(x2)>0U^{\prime\prime}(x_{2})>0 for all x2(h,0)x_{2}\in(-h,0), then the above c(k)c^{-}(k) (given in Theorem 1.1(1)) and c(k)c_{-}(k) (Theorem 1.1(2)) can be extended for all kk\in\mathbb{R} and they coincide, with cI(k)=0c_{I}^{-}(k)=0 for all |k|k|k|\leq k_{-} and cI(k)>0c_{I}^{-}(k)>0 for all |k|>k|k|>k_{-}. Moreover, all singular and non-singular modes of (1.3) are given by (k,c±(k))(k,c^{\pm}(k)) and (k,c(k)¯)(k,\overline{c^{-}(k)}) for kk\in\mathbb{R}.

  2. (2)

    The linearized gravity water wave system (1.3) is spectrally stable for all wave number kk\in\mathbb{R} if U′′0U^{\prime\prime}\leq 0 on [h,0][-h,0]. Moreover (k,c)(k,c) is a neutral mode iff c=c+(k)c=c^{+}(k), c=c(k)c=c_{-}(k) and |k|k|k|\leq k_{-}, or cU((h,0))c\in U((-h,0)) is an inflection value of UU and |k|S|k|\in S given in Theorem 1.1(3).

Remark 1.3.

Under the assumption U′′<0U^{\prime\prime}<0 (and U(0)0U^{\prime}(0)\geq 0), the spectral stability had been established in [28] (see also [10]). The above spectral stability statement (2) can also be proved under the relaxed assumptions U′′0U^{\prime\prime}\leq 0 on [h,0][-h,0] and U(0)>0U^{\prime}(0)>0. See Remark 3.4.

Finally, the last theorem is for the case where UU has exactly one non-degenerate inflection point in (h,0)(-h,0).

Theorem 1.3.

Assume

(1.6) {x2(h,0)U′′(x2)=0}={x20},c0=U(x20),\{x_{2}\in(-h,0)\mid U^{\prime\prime}(x_{2})=0\}=\{x_{20}\},\quad c_{0}=U(x_{20}),

then the following hold.

  1. (1)

    If U′′′(x20)>0U^{\prime\prime\prime}(x_{20})>0, then c(k)c^{-}(k) can be extended evenly for all |k|k0|k|\geq k_{0}, where S={k0}S=\{k_{0}\} (see Theorem 1.1(1)(3)), such that c(k0)=c0c^{-}(k_{0})=c_{0} and cI(k)>0c_{I}^{-}(k)>0 for all |k|>k0|k|>k_{0}. Moreover all singular and non-singular modes of (1.3) are given by (k,c+(k))(k,c^{+}(k)) for kk\in\mathbb{R}, (k,c(k))(k,c^{-}(k)) and (k,c(k)¯)(k,\overline{c^{-}(k)}) for |k|k0|k|\geq k_{0}, and (k,c(k))(k,c_{-}(k)) for |k|k|k|\leq k_{-}.

  2. (2)

    If U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0 and S={k0}S=\{k_{0}\}, then k0>kk_{0}>k_{-} and c(k)c_{-}(k) (given in Theorem 1.1(2)) can be extended evenly for |k|k0|k|\leq k_{0} such that c(k0)=c0c_{-}(k_{0})=c_{0}, Imc(k)>0\text{Im}\,c_{-}(k)>0 for |k|(k,k0)|k|\in(k_{-},k_{0}), and all singular and non-singular modes of (1.3) are given by (k,c+(k))(k,c^{+}(k)) for kk\in\mathbb{R} and (k,c(k))(k,c_{-}(k)) and (k,c(k)¯)(k,\overline{c_{-}(k)}) for |k|k0|k|\leq k_{0}.

  3. (3)

    If U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0, S={k0>k1}S=\{k_{0}>k_{1}\}, and k1kk_{1}\leq k_{-}, then k0>kk_{0}>k_{-} and we have the following.

    1. (a)

      c(k)c_{-}(k) can be extended evenly for |k|k0|k|\leq k_{0} such that c(k0)=c0c_{-}(k_{0})=c_{0} and Imc(k)>0\text{Im}\,c_{-}(k)>0 for |k|(k,k0)|k|\in(k_{-},k_{0}).

    2. (b)

      There exists a C4C^{4} even complex valued function 𝒞(k)\mathcal{C}(k) defined for |k|[0,k1]|k|\in[0,k_{1}] such that 𝒞(±k1)=c0\mathcal{C}(\pm k_{1})=c_{0} and 𝒞I(k)>0\mathcal{C}_{I}(k)>0 for all |k|[0,k1)|k|\in[0,k_{1}).

    3. (c)

      All singular and non-singular modes of (1.3) are given by (k,c+(k))(k,c^{+}(k)) for kk\in\mathbb{R}, (k,c(k))(k,c_{-}(k)) and (k,c(k)¯)(k,\overline{c_{-}(k)}) for |k|k0|k|\leq k_{0}, and (k,𝒞(k))(k,\mathcal{C}(k)) and (k,𝒞(k)¯)(k,\overline{\mathcal{C}(k)}) for |k|k1|k|\leq k_{1}.

  4. (4)

    If U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0, S={k0>k1}S=\{k_{0}>k_{1}\}, and k1>kk_{1}>k_{-}, then k0>kk_{0}>k_{-} and (1.3) is unstable iff |k|<k0|k|<k_{0}. Moreover there exist connected components 𝒮j\mathcal{S}_{j}, j=1,0,1j=-1,0,1, of

    𝒮={non-singular modes (k,c)×cI>0}{(±k,U(h)),(±k0,c0),(±k1,c0)},\mathcal{S}=\{\text{non-singular modes }(k,c)\in\mathbb{R}\times\mathbb{C}\mid c_{I}>0\}\cup\{(\pm k_{-},U(-h)),\,(\pm k_{0},c_{0}),\,(\pm k_{1},c_{0})\},

    such that

    𝒮=𝒮1𝒮0𝒮1,𝒮1={(k,c)(k,c)𝒮1},{(k,c)(k,c)𝒮0}=𝒮0,\mathcal{S}=\mathcal{S}_{-1}\cup\mathcal{S}_{0}\cup\mathcal{S}_{1},\quad\mathcal{S}_{1}=\{(k,c)\mid(-k,c)\in\mathcal{S}_{-1}\},\quad\{(k,c)\mid(-k,c)\in\mathcal{S}_{0}\}=\mathcal{S}_{0},

    and either

    1. (a)

      (k,U(h)),(k0,c0)𝒮1(k_{-},U(-h)),(k_{0},c_{0})\in\mathcal{S}_{1} and (±k1,c0)𝒮0(\pm k_{1},c_{0})\in\mathcal{S}_{0}, or

    2. (b)

      (k,U(h)),(k1,c0)𝒮1(k_{-},U(-h)),(k_{1},c_{0})\in\mathcal{S}_{1} and (±k0,c0)𝒮0(\pm k_{0},c_{0})\in\mathcal{S}_{0}.

Remark 1.4.

a.) An equivalent condition to the existence of k1k_{1} is given in Lemma 3.7.
b.) In this theorem, the case of U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0 corresponds to the intersection of increasing and the so-called class 𝒦+\mathcal{K}^{+} (see [10]) shear flows.
c.) In particular, (1.3) becomes unstable for all long waves (i.e. |k|1|k|\ll 1) if k1>0k_{1}>0.
d.) Assuming U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0 and UU is an unstable shear flow of the channel flow with fixed boundaries. Letting gg increase, the system could deform from case (2) to (3) and then possibly to (4), see Lemma 3.7. In Lemma 3.12 which holds under a weaker assumption, one can see a closed branch of unstable modes bifurcates from c0c_{0} as 1k1>01\gg k_{1}>0. As this branch grows, it may or may not intersect the branch c(k)c_{-}(k) in case (2), resulting in the two possibilities in case (4) where the three subsets may even coincide and be equal to 𝒮\mathcal{S}.

1.3. Backgrounds and discussions

Due to its strong physical and mathematical relevance there have been extensive studies of the Euler equation linearized at shear currents. In particular the linear instability is often viewed as the first step in understanding the transition of the fluid motion from the laminar flows to turbulent ones. Much of the existing analysis was on the fluid in a fixed channel with slip boundary conditions

(1.7) (1.1a)–(1.1b) with g=0,x2(h,0),v2(x1,0)=v2(x1,h)=0,\text{\eqref{E:Euler-1}--\eqref{E:Euler-2} with }g=0,\;\;\;x_{2}\in(-h,0),\quad v_{2}(x_{1},0)=v_{2}(x_{1},-h)=0,

and some of the results have been extended to free boundary problems.

\bullet Linearized channel flows. Classical results on the spectra of the Euler equation (1.7) in a channel linearized at a shear flow include:

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

  • Rayleigh’s necessary condition of instability [23]: unstable eigenvalues do not exist for any kk if U′′0U^{\prime\prime}\neq 0 on [h,0][-h,0] (see also [6]).

  • Howard’s Semicircle Theorem [8]: eigenvalues exist only with cc in the disk

    (1.8) (cR12(Umax+Umin))2+cI214(UmaxUmin)2.\big{(}c_{R}-\tfrac{1}{2}(U_{max}+U_{min})\big{)}^{2}+c_{I}^{2}\leq\tfrac{1}{4}(U_{max}-U_{min})^{2}.
  • Unstable eigenvalues may exist with cc near inflection values of UU (Tollmien [27] formally, also [14]).

Many classical results can be found in books such as [5, 19] etc. For a class of shear flows, the rigorous bifurcation of unstable eigenvalues was proved, e.g., in [7, 15]. In particular, continuation of branches of unstable eigenvalues were obtained by Lin in the latter.

\bullet Linearized gravity water waves. It has been extended to the linearized free boundary problem of the gravity waves at certain classes of shear flows (see [4, 9, 1, 28, 25, 18, 10, 13, 24, 2] etc.) that: a.) the semicircle theorem still holds for unstable modes; b.) the bifurcation and continuation of branches of unstable modes starting from limiting singular neutral modes; c.) the stability if U′′<0U^{\prime\prime}<0 and U(0)0U^{\prime}(0)\geq 0, etc. Compared to channel flows with fixed boundaries, new phenomena of the linearized gravity waves include:

  • In addition to inflection values, critical values of UU where U=0U^{\prime}=0, and c=U(h)c=U(-h) may be limiting singular neutral modes.

  • U′′0U^{\prime\prime}\neq 0 may not ensure the spectral stability.

  • There are non-singular neutral modes with cc in both (,U(h))(-\infty,U(-h)) and (U(0),+)(U(0),+\infty) which are outside the circle (1.8).

  • U(h)U(-h) can be a singular neutral mode for certain wave number.

However, the rigorous bifurcation analysis at U(h)U(-h) or critical values of UU was still missing. This holds the key how Rayleigh’s stability condition may fail in the gravity wave case. Moreover we have not been aware of rigorous studies of the branch of unstable modes near c=U(0)c=U(0) for |k|1|k|\gg 1. These have been addressed in the above Theorem 1.1.

In particular, we would like discuss some of our results (assuming U>0U^{\prime}>0) in relation to those in [28, 10, 11, 24].

* Singular neutral modes and branches from bifurcations. Firstly we do not only prove that U(h)U(-h) and internal inflection values are always singular neutral modes for some wave numbers as in [28, 10, 11], but also obtain the exact numbers of those wave numbers (see Theorem 1.1 and Subsection 3.3). For an internal inflection value c0=U(x20)c_{0}=U(x_{20}), it turns out that whether it becomes a singular neutral mode at one or two wave numbers depends on i.) whether it is a singular neutral mode for the linearized channel flow and also on ii.) whether gg is greater than some threshold. Secondly our bifurcation analysis is carried out in whole neighborhoods of singular neutral modes and allows broken branches, instead of in cones of neighborhoods as in [15, 10]. This is more than a technical improvement, as it gives both the existence and the number of nearby unstable modes which is crucial for the index counting used in the analytic continuation argument. In particular, when the above i) holds, strong gravity could creates a closed branch of unstable modes with |k|k1|k|\leq k_{1} for some k1>0k_{1}>0 (Lemma 3.12 and Theorem 1.3). This is contrary to the common expectation that gravity stabilizes long waves. Finally, when U′′U^{\prime\prime} changes the sign at most once, these ingredients allow us to obtain the complete eigenvalue distribution by identifying how different branches of unstable modes connect to each other. Some of these branches had been observed in numerics [24].

* Comparison to numerics. Three examples of shear flows were computed in [24] which we shall adapt to our notations to avoid unnecessary confusions. In the last two examples, U>0U^{\prime}>0 for x2(2,0)x_{2}\in(-2,0), U′′(x2)=0U^{\prime\prime}(x_{2})=0 iff x2=1x_{2}=-1, and c0=U(1)=0c_{0}=U(-1)=0 is the only inflection value. Their second example is U(x2)=tanha(x2+1)U(x_{2})=\tanh a(x_{2}+1) and a=1/2a=1/2 or 22, where clearly U′′′(1)<0U^{\prime\prime\prime}(-1)<0. The former is stable in the channel flow case, while the latter is unstable. For a=1/2a=1/2, a branch of unstable modes corresponding to c(k)c_{-}(k) in Theorem 1.3(2) was found. For a=2a=2, two branches of unstable modes corresponding to those in either (3) or (4a) of Theorem 1.3 were observed. In their third example, U(x2)=1+x2+(1+x2)3/2U(x_{2})=1+x_{2}+(1+x_{2})^{3}/2 where U′′′(1)>0U^{\prime\prime\prime}(-1)>0, and the branch of unstable modes corresponds to that in Theorem 1.3(1).

\bullet Comparison with the linearized capillary gravity waves at monotonic shear flows. In this paper we follow a strategy to study the spectra of the linearized gravity waves (1.3) similar to that of the linearized capillary gravity waves in [17]. The analysis is based on a detailed understanding of the Rayleigh equation (1.5a), particularly near its singularity where U(x2)c0U(x_{2})-c\sim 0. Subsequently, we consider the solution y(k,c,x2)y_{-}(k,c,x_{2}) to (1.5a) satisfying the normalized initial condition y(h)=0y_{-}(-h)=0 and y(h)=1y_{-}^{\prime}(-h)=1, which are analytic in kk\in\mathbb{R} and cU([h,0])c\in\mathbb{C}\setminus U([-h,0]). Such a solution gives rises to an eigenfunction iff the boundary condition (1.5c) is satisfied. This is formulated into an equation F(k,c)=0F(k,c)=0 where FF is defined in (2.23) whose regularity is carefully studied. In Subsections 2.1 and 2.2, we outline those basic analysis developed in [17] needed in this paper. To identify the roots of FF, the key ingredients are a.) the asymptotic properties of FF as |k|,|c|+|k|,|c|\to+\infty and cI0c_{I}\to 0; b.) the bifurcations for cc near U([h,0])U([-h,0]); and c.) the analytic continuation of the roots of FF. Here kk is treated as a parameter in both b.) and c.).

As revealed in Theorems 1.1 and 1.2, the spectral distributions are significantly different when the surface tension is not present. Moreover, the analysis in the current paper also substantially improves some results for capillary gravity waves (see Proposition 3.16).

On the one hand, while one might expect that the water waves tend to be more unstable when there is no surface tension, particularly in high wave numbers, the main differences in the spectral distributions at monotonic shear flows include the following.

  • (i)

    For |k|1|k|\gg 1, capillary gravity waves are linearly stable, while the gravity waves are unstable if the set {UU′′>0}\{U^{\prime}U^{\prime\prime}>0\} accumulates at x2=0x_{2}=0.

  • (ii)

    While c=U(h)c=U(-h) and interior inflection values of UU may or may not be a singular neutral mode of the linearized capillary gravity waves depending on the combined strength of the gravity and surface tension, they are always singular neutral modes of the linearized gravity waves for one or two k0k\geq 0 (for a unique k>0k>0 in the case of c=U(h)c=U(-h)).

  • (iii)

    As a consequence, {UU′′>0}\{U^{\prime}U^{\prime\prime}>0\}\neq\emptyset is "almost" sufficient for the instability of the linearized gravity waves (Theorem 1.1(4)), which is not always the case for the capillary gravity waves.

  • (iv)

    Contrary to the natural expectation, a.) when the surface tension is very weak with the coefficient 0<σ10<\sigma\ll 1, it actually creates instability near non-degenerate inflection values in relatively large wave numbers kσ1k\sim\sigma^{-1} (Proposition 4.11 in [17]), and b.) strong gravity may also cause instability in all small wave numbers.

Related to (i), one is reminded that, for shear flows, the Rayleigh-Taylor sign pn|x2=0=g>0-\frac{\partial p}{\partial n}\big{|}_{x_{2}=0}=g>0 at any shear flow implies that the nonlinear gravity waves are locally well-posed for initial data near shear flows and there is no unbounded exponential growth rates in high waves numbers. However, when UU′′>0U^{\prime}U^{\prime\prime}>0 near x2=0x_{2}=0, the linearization at the shear flow is still unstable for all |k|1|k|\gg 1, but the exponential growth rates O(|U′′||k|12e2g|k|U(0))O\big{(}|U^{\prime\prime}||k|^{-\frac{1}{2}}e^{-\frac{2\sqrt{g|k|}}{U^{\prime}(0)}}\big{)} of those unstable linear solutions (Theorem 1.1(1a)) are extremely weak.

When the combination of the surface tension and the gravity is sufficiently strong, the linearized capillary gravity wave at a monotonic shear flow is essentially the superposition of a linearized channel flow with fixed boundaries (corresponding to the continuous spectra) and a dispersive system (corresponding to the non-singular modes and the surface motions) with two branches of the dispersion relation given by c±(k)U([h,0])c^{\pm}(k)\notin U([-h,0]) (see [17]). Without surface tension, the complete branches of the non-singular modes alway reach and also possibly converge fast to U([h,0])U([-h,0]). This interaction between the continuous spectra and eigenvalues would make the the linear inviscid damping (see [17]) a very subtle issue except in the simplest case where UU′′<0U^{\prime}U^{\prime\prime}<0 near x2=0x_{2}=0 and the period in x1x_{1} eliminates all the singular neutral modes. We would address the inviscid damping in a future work.

These differences in the spectra could potential result in substantial differences in the nonlinear local dynamics.

On the other hand, the bifurcation analysis of unstable eigenvalues requires certain regularity and non-degeneracy of F(k,c)F(k,c), particular in the most subtle case near c=U(h)c=U(-h). The regularity and non-degeneracy at c=U(h)c=U(-h) in [17] were verified under the additional assumption U′′0U^{\prime\prime}\neq 0 on [h,0][-h,0]. In Subsections 3.1 and 3.4, we prove these regularity and non-degeneracy without any assumptions additional to the monotonicity of UU. Therefore the work in this paper also substantially improves the bifurcation analysis in the capillary gravity wave case, see Proposition 3.16 in Subsection 3.8.

Compared to [17], in this paper we also thoroughly analyzed the case where U′′U^{\prime\prime} changes sign exactly once.

\bullet The spectra of the linearized fluid interface problem. Another related problem is the two phase fluid problem linearized at a shear flow U(x2)=U+(x2)χx2(0,h+)+U(x2)χx2(h,0)U(x_{2})=U^{+}(x_{2})\chi_{x_{2}\in(0,h_{+})}+U^{-}(x_{2})\chi_{x_{2}\in(-h_{-},0)} where the slip boundary condition is assumed at x2=±h±x_{2}=\pm h_{\pm} . The famous Kelvin Helmholtz instability was first identified in the simplest setting where U±=constsU^{\pm}=consts [26, 5, 2]. In the case where the upper fluid is much lighter than the bottom one (the air-water interface, for example), Miles [20, 21, 22, 12] proposed the critical layer theory to model how the wind (shear flow) in the air above a stationary water generates waves through linear instability due to the resonance between the shear flow in the air and the temporal frequency of the linear irrotational capillary gravity water waves. This was later rigorously proved in [3]. More recently, the spectra of the linearized fluid interface problem at monotonic shear flows had been studied more comprehensively and in more details in [16].

Notations. Throughout the paper, wave numbers are always denoted by "kk" and K=k2K=k^{2} is also used. For complex quantities, the subscripts "R" and "I" denote their real and imaginary parts, respectively. Sometimes "Re" and "Im" are also used when the subscripts make the notations too cumbersome.

2. Preliminary analysis

In this section we review some basic results on the homogeneous Rayleigh equation and the equation satisfied by the eigenmodes (k,c,y(x2))(k,c,y(x_{2})). Most of these results have already been obtained in [17] and the readers are referred there for more details.

2.1. Results on the homogeneous Rayleigh equation

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

where

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

had been thoroughly analyzed in Section 3 of [17]. In this subsection, we summarize those results which we shall use here. When necessary some modifications will be outlined to address the different needs in this paper. Throughout this subsection we assume

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

As pointed out in Remark 1.1, the case of U<0U^{\prime}<0 can be reduced to the above one. Hence all results under (2.2) hold for all uniformly monotonic U(x2)U(x_{2}), namely those UU satisfying U0U^{\prime}\neq 0 on [h,0][-h,0]. The solutions to the Rayleigh equation (2.1) are obviously even in kk and thus k0k\geq 0 will be assumed mostly. Similarly complex conjugate of solutions also solve (2.1) with cc replaced by c¯\bar{c}, hence we shall consider the case of cI>0c_{I}>0. Most of the estimates are uniform in cc with cI>0c_{I}>0 and thus hold in the limiting case as cI0+c_{I}\to 0+. The dependence on k1k\gg 1 will also be carefully tracked.

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

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

so that

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

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

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

Consider the solution y(k,c,x2)y_{-}(k,c,x_{2}) to (2.1) satisfying the initial conditions

(2.6) y(h)=0,y(h)=1,\begin{split}y_{-}(-h)=0,\quad y_{-}^{\prime}(-h)=1,\end{split}

where, as often in the rest of the paper, the dependence on kk and cc are skipped to make the notations simpler when there is no confusion in the context. Apparently this is a non-trivial solution and the following lemma holds.

Lemma 2.1.

For any kk\in\mathbb{R} and cU([h,0])c\in\mathbb{C}\setminus U([-h,0]), y(k,c,x2)=y(k,c,x2)=y(k,c¯,x2)¯y_{-}(k,c,x_{2})=y_{-}(-k,c,x_{2})=\overline{y_{-}(k,\bar{c},x_{2})} is analytic in cc and k2k^{2}.

To give some basic results on y(k,c,x2)y_{-}(k,c,x_{2}), we first introduce some subintervals of [h,0][-h,0],

(2.7) 2:=(x2l,x2r)={x2[h,0]:1|U(x2)c|>ρ0μ32},1=[h,x2l),3=(x2r,0],μ=k1=(1+k2)12,ρ0=4h0inf[h0h,h0]U.\begin{split}&\mathcal{I}_{2}:=(x_{2l},x_{2r})=\left\{x_{2}\in[-h,0]\,:\,\frac{1}{|U(x_{2})-c|}>\rho_{0}\mu^{-\frac{3}{2}}\right\},\quad\mathcal{I}_{1}=[-h,x_{2l}),\\ &\mathcal{I}_{3}=(x_{2r},0],\quad\mu=\langle k\rangle^{-1}=(1+k^{2})^{-\frac{1}{2}},\quad\rho_{0}=\frac{4}{h_{0}\inf_{[-h_{0}-h,h_{0}]}U^{\prime}}.\end{split}

Clearly [h,0]=123[-h,0]=\mathcal{I}_{1}\cup\mathcal{I}_{2}\cup\mathcal{I}_{3} where any of the three subintervals may be empty. In particular, if 2=\mathcal{I}_{2}=\emptyset, then we consider [h,0][-h,0] as 1\mathcal{I}_{1} in the statement of the following lemma. The choice of the above constant ρ0\rho_{0} and the fact 0μ10\leq\mu\leq 1 ensure

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

The following is a part of Lemma 3.9 in [17].

Lemma 2.2.

For any α(0,12)\alpha\in(0,\frac{1}{2}), there exists C>0C>0 depending only on α\alpha, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that, for any c\U([h,0])c\in\mathbb{C}\backslash U([-h,0]), it holds

(2.9) |μ1y(x2)sinh(μ1(x2+h))|Cμαsinh(μ1(x2+h)),|\mu^{-1}y_{-}(x_{2})-\sinh(\mu^{-1}(x_{2}+h))|\leq C\mu^{\alpha}\sinh(\mu^{-1}(x_{2}+h)),

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

(2.10) |y(x2)cosh(μ1(x2+h))|Cμαsinh(μ1(x2+h)),|y_{-}^{\prime}(x_{2})-\cosh(\mu^{-1}(x_{2}+h))|\leq C\mu^{\alpha}\sinh(\mu^{-1}(x_{2}+h)),

Otherwise if 2\mathcal{I}_{2}\neq\emptyset, then

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

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

(2.12) |y(x2)cosh(μ1(x2+h))U′′(x2c)U(x2c)y(x2l)log|U(x2)c||Cμα(1+μ|log|U(x2)c||)cosh(μ1(x2+h)).\begin{split}\big{|}y_{-}^{\prime}(x_{2})-\cosh(\mu^{-1}(x_{2}+h))&-\frac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{-}(x_{2l})\log|U(x_{2})-c|\big{|}\\ &\leq C\mu^{\alpha}\big{(}1+\mu\big{|}\log|U(x_{2})-c|\big{|}\big{)}\cosh(\mu^{-1}(x_{2}+h)).\end{split}
Remark 2.1.

Even though cU([h,0])c\notin U([-h,0]) is assumed in the lemma, the estimates are uniform in cc and hence they also hold for the limits of solutions as cI0+c_{I}\to 0+, while the limits as cI0c_{I}\to 0- are the conjugates of those as cI0+c_{I}\to 0+.

To study the limits as cc converges to the singular set U([h,0])U([-h,0]), let

(2.13) y0(k,c,x2)=limcI0+y(k,c+icI,x2),cU([h02h,h02]).y_{0-}(k,c,x_{2})=\lim_{c_{I}\to 0+}y_{-}(k,c+ic_{I},x_{2}),\quad c\in U([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]).

The following lemma (a combination of parts of Lemmas 3.10, 3.12, 3.13, 3.14, 3.15, 3.16, and 3.19 in [17]) gives some basic properties of y0(k,c,x2)y_{0-}(k,c,x_{2}) including their dependence in cU([h02h,h02])c\in U([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]).

Lemma 2.3.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3. For c=U(x2c)U([h02h,h02])c=U(x_{2}^{c})\in U([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]) and x2[h,0]x_{2}\in[-h,0], the following hold, where C>0C>0 is determined only by UU.

  1. (1)

    As cI0+c_{I}\to 0+, y(k,c+icI,x2)y0(k,c,x2)y_{-}(k,c+ic_{I},x_{2})\to y_{0-}(k,c,x_{2}) uniformly in x2x_{2} and cc, yy0y_{-}^{\prime}\to y_{0-}^{\prime} locally uniformly in {U(x2)c}\{U(x_{2})\neq c\}.

  2. (2)

    For each cc, y0(x2)y_{0-}(x_{2})\in\mathbb{R} if x2x2cx_{2}\leq x_{2}^{c}. Moreover y0Cα([h02h,h02])y_{0-}\in C^{\alpha}([-\tfrac{h_{0}}{2}-h,\tfrac{h_{0}}{2}]) for any α[0,1)\alpha\in[0,1) and y0y_{0-} is Cl0C^{l_{0}} in x2x2cx_{2}\neq x_{2}^{c}.

  3. (3)

    There exists a unique continuous-in-τ\tau real 2×22\times 2 matrix valued B(μ,c,τ)B(\mu,c,\tau) determined only by UU such that

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

    Here,

    Γ0#(μ,c,τ)=τ+μU′′(x2c)U(x2c)(log|τ|+iπ2(sgn(τ)+1)).\Gamma_{0}^{\#}(\mu,c,\tau)=\tau+\frac{\mu U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\big{(}\log|\tau|+\frac{i\pi}{2}(sgn(\tau)+1)\big{)}.
    b2=limx2x2c(y0(x2)U′′(x2c)U(x2c)y0(x2c)(log(U(x2c)μ|x2x2c|)+iπ2(sgn(x2x2c)+1))) exists,b_{2-}=\lim_{x_{2}\to x_{2}^{c}}\Big{(}y_{0-}^{\prime}(x_{2})-\tfrac{U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}y_{0-}(x_{2}^{c})\big{(}\log\big{(}\tfrac{U^{\prime}(x_{2}^{c})}{\mu}|x_{2}-x_{2}^{c}|\big{)}+\tfrac{i\pi}{2}(sgn(x_{2}-x_{2}^{c})+1)\big{)}\Big{)}\text{ exists},

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

    detB=1,B(μ,c,0)=I2×2,B(0,c,τ)=(coshττsinhτsinhτsinhττcoshτcoshτ).\det B=1,\quad B(\mu,c,0)=I_{2\times 2},\quad B(0,c,\tau)=\begin{pmatrix}\cosh\tau-\tau\sinh\tau&\sinh\tau\\ \sinh\tau-\tau\cosh\tau&\cosh\tau\end{pmatrix}.
  4. (4)

    For any kk\in\mathbb{R}, y0(k,c,x2)>0y_{0-}(k,c,x_{2})>0 for any x2(h,0]x_{2}\in(-h,0] and cU((h,0))c\in\mathbb{R}\setminus U\big{(}(-h,0)\big{)}.

  5. (5)

    For any k,ck,c\in\mathbb{R}, it holds

    (2.14) (Ck)1sinhk(x2+h)y0(k,c,x2)Ck1sinhk(x2+h)(Ck)^{-1}\sinh k(x_{2}+h)\leq y_{0-}(k,c,x_{2})\leq Ck^{-1}\sinh k(x_{2}+h)

    for (x2,c)(x_{2},c) satisfying (U(h)c)(U(x2)c)0\big{(}U(-h)-c\big{)}\big{(}U(x_{2})-c\big{)}\geq 0 and moreover, for cU((h,0))c\in U\big{(}(-h,0)\big{)},

    (2.15) C1|U′′(x2c)||Imy0(k,c,0)|μ2sinhμ1(x2c+h)sinhμ1|x2c|C|U′′(x2c)|.C^{-1}|U^{\prime\prime}(x_{2}^{c})|\leq\frac{|\text{Im}\,y_{0-}(k,c,0)|}{\mu^{2}\sinh\mu^{-1}(x_{2}^{c}+h)\sinh\mu^{-1}|x_{2}^{c}|}\leq C|U^{\prime\prime}(x_{2}^{c})|.
  6. (6)

    y0y_{0-} is locally CαC^{\alpha} in both kk and cc for any α[0,1)\alpha\in[0,1) and (y0,y0)(y_{0-},y_{0-}^{\prime}) are CC^{\infty} in kk at any (k,c,x2)(k,c,x_{2}) except for y0y_{0-}^{\prime} at c=U(x2)c=U(x_{2}).

  7. (7)

    (y0,y0)(y_{0-},y_{0-}^{\prime}) are locally CαC^{\alpha} in both kk and cc for any α[0,1)\alpha\in[0,1) at any (k,c,x2)(k,c,x_{2}) satisfying U(x2)cU(x_{2})\neq c.

  8. (8)

    (y0,y0)(y_{0-},y_{0-}^{\prime}) are Cl02C^{l_{0}-2} in both kk and cc at any (k,c,x2)(k,c,x_{2}) satisfying U(x2)cU(x_{2})\neq c and cU(h)c\neq U(-h) . Moreover, the following estimates hold for cc in a neighborhood of U([h,0])U([-h,0]) (uniform in kk), where ρ0\rho_{0} is given in (2.7).

    1. (a)

      For any j1,j20j_{1},j_{2}\geq 0 with j1+j2>0j_{1}+j_{2}>0 and x2x_{2}\in\mathcal{I}, where [h,0]\mathcal{I}\subset[-h,0] is any interval satisfying h{|Uc|ρ01μ}-h\in\mathcal{I}\subset\{|U-c|\geq\rho_{0}^{-1}\mu\}, we have

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

      where Cj1,j2>0C_{j_{1},j_{2}}>0 depends only on j1j_{1}, j2j_{2}, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}.

    2. (b)

      Suppose [h,0]{|Uc|ρ01μ}[-h,0]\not\subset\{|U-c|\geq\rho_{0}^{-1}\mu\} and l05l_{0}\geq 5, then, for x2[h,0]x_{2}\in[-h,0] satisfying U(x2)cρ01μU(x_{2})-c\geq\rho_{0}^{-1}\mu,

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

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

      1μ|cjy0(x2)|+|cjy0(x2)|C(μ1j+|U(h)c|1j)cosh(μ1(x2+h)).\tfrac{1}{\mu}|\partial_{c}^{j}y_{0-}(x_{2})|+|\partial_{c}^{j}y_{0-}^{\prime}(x_{2})|\leq C\big{(}\mu^{1-j}+|U(-h)-c|^{1-j}\big{)}\cosh(\mu^{-1}(x_{2}+h)).
  9. (9)

    y0(k,c,x2c)y_{0-}(k,c,x_{2}^{c}) is Cl02C^{l_{0}-2} in cc and kk for cU([h,0])c\in U([-h,0]). In addition, for 1jl021\leq j\leq l_{0}-2, it holds

    (2.18) |cj(y0(k,c,x2c))|Cμ1jcosh(μ1(x2c+h)).|\partial_{c}^{j}\big{(}y_{0-}(k,c,x_{2}^{c})\big{)}|\leq C\mu^{1-j}\cosh(\mu^{-1}(x_{2}^{c}+h)).
Remark 2.2.

In many cases, to simplify the notation, we denote y(k,c,x2)=y0(k,c,x2)y_{-}(k,c,x_{2})=y_{0-}(k,c,x_{2}) for cc\in\mathbb{R}. In (2.14), k1sinhkxk^{-1}\sinh kx is understood as xx when k=0k=0. The regularity of y0y_{0-}, the analyticity of yy_{-} and yy_{-}^{\prime} in kk and cc with cI>0c_{I}>0, and its convergence as cI0+c_{I}\to 0+ imply y(k,c,x2)y_{-}(k,c,x_{2}) is also Cl02C^{l_{0}-2} in kk and c{U(x2),U(h)}c\notin\{U(x_{2}),U(-h)\} when restricted to cI0c_{I}\geq 0.

Finally, the following quantity related to the Reynolds stress is crucial for the linearized water wave problem:

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

with the domain

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

Those excluded points (except c=U(0)c=U(0)) exactly correspond to the eigenvalues of the linearized Euler equation in the fixed channel x2(h,0)x_{2}\in(-h,0) at the shear flow U(x2)U(x_{2}). The end point c=U(0)c=U(0) is also excluded due to the singularity of y(k,c=U(0),x2)y_{-}^{\prime}(k,c=U(0),x_{2}) at x2=0x_{2}=0. The following lemma (Lemmas 3.20, and 3.22 and parts – some in the proofs – of Lemmas 3.23, 3.24, and 4.5 in [17]) summarizes some basic properties of Y(k,c)Y(k,c).

Lemma 2.4.

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

  1. (1)

    Y(k,c)Y(k,c)\in\mathbb{R} for all cU((h,0])c\in\mathbb{R}\setminus U\big{(}(-h,0]\big{)} and Y(0,U(h))=U(0)U(0)U(h)Y(0,U(-h))=\frac{U^{\prime}(0)}{U(0)-U(-h)}.

  2. (2)

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

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

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

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

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

    |Y(k,c)kcothkh|Cdist(c,U([h,0])),|Y(k,c)-k\coth kh|\leq\frac{C}{dist(c,U([-h,0]))},

    if kk and cc satisfy

    |k|k and |cU(h)+U(0)2|M+U(0)U(h)2.|k|\leq k_{*}\;\text{ and }\;\Big{|}c-\frac{U(-h)+U(0)}{2}\Big{|}\geq M+\frac{U(0)-U(-h)}{2}.
  5. (5)

    YI(k,c)=0Y_{I}(k,c)=0 for c\U((h,0])c\in\mathbb{R}\backslash U\big{(}(-h,0]\big{)} and

    YI(k,c)=πU′′(x2c)y(k,c,x2c)2U(x2c)|y(k,c,0)|2,c{cU((h,0))y(k,c,0)0}.Y_{I}(k,c)=\frac{\pi U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})^{2}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{2}},\quad c\in\mathcal{I}\triangleq\{c\in U\big{(}(-h,0)\big{)}\mid y_{-}(k,c,0)\neq 0\}.

    Moreover, the above formula implies YI(k,c)Y_{I}(k,c) is CC^{\infty} in kk and Cl02C^{l_{0}-2} in (k,c)D(Y)(×U((h,0)))(k,c)\in D(Y)\cap\big{(}\mathbb{R}\times U\big{(}(-h,0)\big{)}\big{)}.

  6. (6)

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

  7. (7)

    Assume kk\in\mathbb{R} satisfies y(k,c,0)0y_{-}(k,c,0)\neq 0 for all cc\in\mathbb{C}, then, for any c\U([h,0])c\in\mathbb{C}\backslash U([-h,0]),

    (2.20) Y(k,c)=1πU(h)U(0)YI(k,c)cc𝑑c+kcothkh,\begin{split}&Y(k,c)=\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{Y_{I}(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}+k\coth kh,\\ \end{split}

    and for cU([h,0))c\in U\big{(}[-h,0)\big{)},

    (2.21) Y(k,c)=(YI(k,))(c)+iYI(k,c)+kcothkh.\begin{split}&Y(k,c)=-\mathcal{H}\big{(}Y_{I}(k,\cdot)\big{)}(c)+iY_{I}(k,c)+k\coth kh.\end{split}
  8. (8)

    For any kk\in\mathbb{R} and cU((h,0])c\in\mathbb{R}\setminus U\big{(}(-h,0]\big{)}

    K(Y(K,c))>0,K2(Y(K,c))<0,\partial_{K}\big{(}Y\big{(}\sqrt{K},c\big{)}\big{)}>0,\quad\partial_{K}^{2}\big{(}Y\big{(}\sqrt{K},c\big{)}\big{)}<0,
    (U(0)c)2KY(0,c)(U(0)U(h))2KY(0,U(h))=h0(U(x2)U(h))2𝑑x2,(U(0)-c)^{2}\partial_{K}Y(0,c)\leq(U(0)-U(-h))^{2}\partial_{K}Y(0,U(-h))=\int_{-h}^{0}(U(x_{2})-U(-h))^{2}dx_{2},

    where K=k2K=k^{2} and the equal sign in the second inequality happens if and only if c=U(h)c=U(-h).

Here we recall that \mathcal{H} denotes the Hilbert transform, namely,

(f)(τ)=1πP.V.f(s)τs𝑑s,\mathcal{H}(f)(\tau)=\frac{1}{\pi}\text{P.V.}\int_{\mathbb{R}}\frac{f(s)}{\tau-s}ds,

for any function f(s)f(s) defined on \mathbb{R}, where P.V.\int is the principle value of the singular integral.

2.2. Basic properties of eigenvalues

From (1.4), ikc-ikc\in\mathbb{C}, with cU([h,0])c\in\mathbb{C}\setminus U([-h,0]), is an eigenvalue of (1.3) in the kk-th Fourier modes iff

(2.22) 𝐅(k,c)=𝐅R+i𝐅I=(U(0)c)2y(k,c,0)(U(0)(U(0)c)+g)y(k,c,0)=0.\begin{split}&\mathbf{F}(k,c)=\mathbf{F}_{R}+i\mathbf{F}_{I}=\big{(}U(0)-c\big{)}^{2}y_{-}^{\prime}(k,c,0)-\big{(}U^{\prime}(0)\big{(}U(0)-c\big{)}+g\big{)}y_{-}(k,c,0)=0.\end{split}

In the limit as cU([h,0])c\to U([-h,0]), let

𝐅(k,c)=limϵ0+𝐅(k,c+iϵ)=limϵ0+𝐅(k,ciϵ)¯,cU([h,0]).\mathbf{F}(k,c)=\lim_{\epsilon\to 0+}\mathbf{F}(k,c+i\epsilon)=\overline{\lim_{\epsilon\to 0+}\mathbf{F}(k,c-i\epsilon)},\quad c\in U\big{(}[-h,0]\big{)}.

Clearly, y(k,c,x2)y_{-}(k,c,x_{2}) also generates the corresponding eigenfunction of (1.3) if 𝐅(k,c)=0\mathbf{F}(k,c)=0. The roots cc of 𝐅\mathbf{F} with cI>0c_{I}>0 are often referred to as unstable modes, while those roots cc\in\mathbb{R} as neutral modes. We recall that Yih proved that the semicircle theorem also holds for (1.3), i.e., (1.8) holds for all unstable modes [28].

Since 𝐅\mathbf{F} may not be C1C^{1} in cc near c=U(h)c=U(-h) (see Lemma 2.3) which would turn out to be a key bifurcation point, we also consider an actually equivalent equation

(2.23) F(k,c)=y(k,c,0)1𝐅=FR+iFI=Y(k,c)(U(0)c)2U(0)(U(0)c)g=0,F(k,c)=y_{-}(k,c,0)^{-1}\mathbf{F}=F_{R}+iF_{I}=Y(k,c)\big{(}U(0)-c\big{)}^{2}-U^{\prime}(0)\big{(}U(0)-c\big{)}-g=0,

with Y(k,c)Y(k,c) defined in (2.19), and

F(k,c)=limϵ0+F(k,c+iϵ)=limϵ0+F(k,ciϵ)¯,cU([h,0]).F(k,c)=\lim_{\epsilon\to 0+}F(k,c+i\epsilon)=\overline{\lim_{\epsilon\to 0+}F(k,c-i\epsilon)},\quad c\in U\big{(}[-h,0]\big{)}.

Clearly it holds

(2.24) 𝐅(k,c)=𝐅(k,c)=𝐅(k,c¯)¯,cU((h,0));\mathbf{F}(-k,c)=\mathbf{F}(k,c)=\overline{\mathbf{F}(k,\bar{c})},\;\forall c\notin U\big{(}(-h,0)\big{)};
(2.25) F(k,c)=F(k,c)=F(k,c¯)¯,cD(Y)U((h,0)).F(-k,c)=F(k,c)=\overline{F(k,\bar{c})},\;c\in D(Y)\setminus U\big{(}(-h,0)\big{)}.

According to the following Lemma 3.1, which is an improvement of Lemma 3.24 in [17], FF is C1,αC^{1,\alpha} near c=U(h)c=U(-h) (without additional assumptions, unlike Lemma 3.24 in [17]). Obviously 𝐅\mathbf{F} and FF are analytic in their domains except for cU([h,0])c\in U([-h,0]).

Our strategy to analyze the eigenvalue distribution includes the following key ingredients:

  • asymptotic analysis of eigenvalues for |k|1|k|\gg 1, which turn out to accumulate at U(0)U(0);

  • the existence and bifurcation of (possibly unstable) eigenvalues for cc near singular neutral modes at c0=U(h)c_{0}=U(-h) or interior inflection values of UU;

  • an analytic continuation argument on the extension of the non-singular modes c(k)c(k) away from the above bifurcation points.

This strategy had been successfully applied to analyze the eigenvalue distribution of the linearized capillary gravity water waves at monotonic shear flows in Section 4 in [17]. The absence of the surface tension does not affect some basic properties of 𝐅\mathbf{F} and FF, some of the bifurcation analysis of unstable eigenvalues, or the continuation argument. In the rest of this subsection, we shall outline the basic properties and the continuation argument obtained in [17]. The bifurcation analysis will be given in Section 3 with a similar approach, but under considerably relaxed assumptions. The existence of singular neutral modes and the eigenvalues for |k|1|k|\gg 1 will also be studied in Section 3, where we shall see the phenomena would turn out to be substantially different from the capillary gravity waves.

\bullet We first give some elementary properties of 𝐅\mathbf{F} and FF for a monotonic shear flow UU, starting with some relatively qualitative properties.

Lemma 2.5.

Assume UCl0U\in C^{l_{0}}, l03l_{0}\geq 3, then for any kk\in\mathbb{R}, the following hold.

  1. (1)

    𝐅\mathbf{F} is well defined for (k,c)×\forall(k,c)\in\mathbb{R}\times\mathbb{C}. When restricted to cI0c_{I}\geq 0, 𝐅\mathbf{F} is CC^{\infty} in kk and is Cl02C^{l_{0}-2} in both kk and c{U(h),U(0)}c\notin\{U(-h),U(0)\} and 𝐅\mathbf{F} is also CαC^{\alpha} in both kk and cc with cI0c_{I}\geq 0. The same holds for F(k,c)F(k,c) except at (k,c)(k,c) where y(k,c,0)=0y_{-}(k,c,0)=0.

  2. (2)

    𝐅(k,c)0\mathbf{F}(k,c)\neq 0 if y(k,c,0)=0y_{-}(k,c,0)=0 and thus {𝐅(k,c)=0}={F(k,c)=0}\{\mathbf{F}(k,c)=0\}=\{F(k,c)=0\}.

  3. (3)

    𝐅(k,c)=0\mathbf{F}(k,c)=0 iff there exists a C2C^{2} solution y(x2)y(x_{2}) to (2.1) satisfying the corresponding homogeneous boundary conditions of (1.5b-1.5c).

  4. (4)

    For any x2c(h,0)x_{2}^{c}\in(-h,0) and c=U(x2c)c=U(x_{2}^{c}),

    FI(k,c)=(U(0)c)2YI(k,c)=π(U(0)c)2U′′(x2c)y(k,c,x2c)2U(x2c)|y(k,c,0)|2,F_{I}(k,c)=\big{(}U(0)-c\big{)}^{2}Y_{I}(k,c)=\frac{\pi(U(0)-c)^{2}U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})^{2}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{2}},

    and thus 𝐅I(k,U(x2))0\mathbf{F}_{I}(k,U(x_{2}))\neq 0 if x2(h,0)x_{2}\in(-h,0) and U′′(x2)0U^{\prime\prime}(x_{2})\neq 0.

These statements are mainly contained in Lemma 4.1 in [17] where one could easily see that the lack of surface tension (namely, σ=0\sigma=0) does not affect the proof. The next lemma is on some quantitative properties of 𝐅\mathbf{F} and FF.

Lemma 2.6.

Suppose UC3U\in C^{3}. The following hold for any kk\in\mathbb{R}.

  1. (1)

    F(k,c)F(k,c) is CC^{\infty} in kk and is well-defined for cc close to U(h)U(-h) and U(0)U(0), C1C^{1} near c=U(0)c=U(0), and

    F(k,U(h)),F(0,U(h))=F(k,U(0))=g,cF(k,U(0))=U(0),F(k,U(-h))\in\mathbb{R},\quad F\big{(}0,U(-h)\big{)}=F\big{(}k,U(0)\big{)}=-g,\quad\partial_{c}F\big{(}k,U(0)\big{)}=U^{\prime}(0),
    F(0,c)=1h0(Uc)2𝑑x2g=(U(0)c)(U(h)c)y(0,c,0)g,cU((h,0)).F(0,c)=\frac{1}{\int_{-h}^{0}(U-c)^{-2}dx_{2}}-g=\frac{(U(0)-c)(U(-h)-c)}{y_{-}(0,c,0)}-g,\quad c\in\mathbb{C}\setminus U((-h,0)).
  2. (2)

    Let K=k2K=k^{2}, we have

    K(F(K,c))>0,K2(F(K,c))<0,k,cU((h,0]),\partial_{K}\big{(}F\big{(}\sqrt{K},c\big{)}\big{)}>0,\quad\partial_{K}^{2}\big{(}F\big{(}\sqrt{K},c\big{)}\big{)}<0,\;\ \forall k\in\mathbb{R},\;c\in\mathbb{R}\setminus U\big{(}(-h,0]\big{)},
    KF(0,c)h0(U(x2)c)2𝑑x2,cU((h,0)),\partial_{K}F(0,c)\leq\int_{-h}^{0}\big{(}U(x_{2})-c\big{)}^{2}dx_{2},\quad\forall c\in\mathbb{R}\setminus U((-h,0)),

    where “=” occurs only at c=U(h)c=U(-h).

Again, as the missing surface has no impact on the proofs, the results in this lemma are mostly contained in the statements and the proofs of Lemmas 4.1 and 4.5 in [17], except KF>0\partial_{K}F>0 in (2) follows directly from Lemma 2.4(8). In particular, the form of F(0,c)F(0,c) for cU([h,0])c\in\mathbb{C}\setminus U([-h,0]) was obtained in Section 4 in [17] based on the explicit formulas

(2.26) y(0,c,x2)=(U(x2)c)hx2U(h)c(U(x2)c)2𝑑x2,Y(0,c)=U(0)h0(Uc)2𝑑x2+(U(0)c)1(U(0)c)h0(Uc)2𝑑x2,cU([h,0]).\begin{split}&y_{-}(0,c,x_{2})=(U(x_{2})-c)\int_{-h}^{x_{2}}\frac{U(-h)-c}{(U(x_{2}^{\prime})-c)^{2}}dx_{2}^{\prime},\\ &Y(0,c)=\frac{U^{\prime}(0)\int_{-h}^{0}(U-c)^{-2}dx_{2}+(U(0)-c)^{-1}}{(U(0)-c)\int_{-h}^{0}(U-c)^{-2}dx_{2}},\end{split}\qquad c\in\mathbb{C}\setminus U([-h,0]).

\bullet As 𝐅\mathbf{F} and FF are analytic functions in their domains outside cU([h,0])c\in U([-h,0]), the analytic continuation argument is a standard tool in the study of the spectra of the linearized Euler equation at shear flows. The following lemma also applies to F(k,c)F(k,c) due to Lemma 2.5(2).

Lemma 2.7.

Assume UC3U\in C^{3}. Suppose k0k_{0}\in\mathbb{R} and c0U([h,0])c_{0}\in\mathbb{C}\setminus U([-h,0]) satisfy 𝐅(k0,c0)=0\mathbf{F}(k_{0},c_{0})=0 and c𝐅(k0,c0)0\partial_{c}\mathbf{F}(k_{0},c_{0})\neq 0, then the following hold.

  1. (1)

    There exists an analytic function c(k)U([h,0])c(k)\in\mathbb{C}\setminus U([-h,0]) defined on a max interval (k,k+)k0(k_{-},k_{+})\ni k_{0} such that 𝐅(k,c(k))=0\mathbf{F}\big{(}k,c(k)\big{)}=0 and c𝐅(k,c(k))0\partial_{c}\mathbf{F}\big{(}k,c(k)\big{)}\neq 0.

  2. (2)

    c(k)c(k)\in\mathbb{R} for all k(k,k+)k\in(k_{-},k_{+}) if and only if c0c_{0}\in\mathbb{R}.

  3. (3)

    If k+<k_{+}<\infty (or k>k_{-}>-\infty), then

    1. (a)

      limk(k+)dist(c(k),U([h,0]))=0\lim_{k\to(k_{+})-}dist(c(k),U([-h,0]))=0 (or limk(k)+dist(c(k),U([h,0]))=0\lim_{k\to(k_{-})+}dist(c(k),U([-h,0]))=0 if k>k_{-}>-\infty), or

    2. (b)

      lim infk(k+)min{|c(k)c|:c s. t. 𝐅(k,c)=0,cc(k)}=0\liminf_{k\to(k_{+})-}\min\{|c(k)-c|\,:\,\forall c\text{ s. t. }\mathbf{F}(k,c)=0,\,c\neq c(k)\}=0 (or lim infk(k)+min{|c(k)c|:𝐅(k,c)=0,cc(k)}=0\liminf_{k\to(k_{-})+}\min\{|c(k)-c|\,:\,\mathbf{F}(k,c)=0,\,c\neq c(k)\}=0 if k>k_{-}>-\infty).

Remark 2.3.

This is exactly Lemma 4.3 in [17] where the absence of the surface tension does not affect the proof. It is based on the non-negative integer valued index of 𝐅\mathbf{F} (or any general complex analytic functions) on various appropriate bounded piecewise smooth domains ΩU([h,0])\Omega\subset\mathbb{C}\setminus U([-h,0]) satisfying that 𝐅(k,)\mathbf{F}(k,\cdot) is holomorphic in Ω\Omega, C0C^{0} in Ω¯\bar{\Omega}, and 𝐅(k,)0\mathbf{F}(k,\cdot)\neq 0 on Ω\partial\Omega,

(2.27) Ind(𝐅(k,),Ω):=12πiΩc𝐅(k,c)𝐅(k,c)𝑑c{0},\text{Ind}\big{(}\mathbf{F}(k,\cdot),\Omega\big{)}:=\frac{1}{2\pi i}\oint_{\partial\Omega}\frac{\partial_{c}\mathbf{F}(k,c)}{\mathbf{F}(k,c)}dc\in\mathbb{N}\cup\{0\},

which is equal to the total number of zeros of 𝐅(k,)\mathbf{F}(k,\cdot) inside Ω\Omega, counting their multiplicities. In particular, given a bounded domain Ω\Omega\subset\mathbb{C} with piecewise smooth Ω\partial\Omega and k1<k2k_{1}<k_{2} such that 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for any k[k1,k2]k\in[k_{1},k_{2}] and cΩc\in\partial\Omega, Ind(𝐅(k,),Ω)\big{(}\mathbf{F}(k,\cdot),\Omega\big{)} is a constant in k[k1,k2]k\in[k_{1},k_{2}]. If Ind(𝐅(k,),Ω)=1\big{(}\mathbf{F}(k,\cdot),\Omega\big{)}=1, then the unique root c(k)c(k) of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω\Omega is simple and analytic in kk. In addition, if k2=k1k_{2}=-k_{1}, then c(k)c(k) is even in kk due to the evenness of 𝐅\mathbf{F} in kk and the uniqueness of its root in Ω\Omega. The proof of statement also needs Lemma 3.2(2) which ensures the roots to be bounded for kk in any compact interval.

3. Eigenvalue distribution of the linearized gravity water waves

In this section, we study the eigenvalues of the linearized gravity waves in details. A complete eigenvalue distribution will be obtained under certain conditions. On the one hand, we shall focus on those aspects which are different from the linearized capillary gravity waves, e.g. the instability in wave number |k|1|k|\gg 1. On the other hand, some similar results as in [17] will be obtained under substantially weaker assumptions, which requires some further detailed analysis on Y(k,c)Y(k,c) and F(k,c)F(k,c) given in the next two subsections.

3.1. Further analysis of Y(k,c)Y(k,c)

In this subsection, we extend the analysis of Y(k,c)Y(k,c) defined in (2.19). In particular, instead of using Lemma 3.24 in [17], we shall obtain the same improved regularity of Y(k,)Y(k,\cdot) near c=U(h)c=U(-h) without any assumption additional to the monotonicity of U(x2)U(x_{2}).

Lemma 3.1.

Suppose UCl0U\in C^{l_{0}}. Assume l03l_{0}\geq 3, then for any M>0M>0 there exists C>0C>0 depending on UU and MM such that for any j0j\geq 0 and cc\in\mathbb{C} satisfying

(3.1) |c12(U(h)+U(0))|12(U(0)U(h))+M,\big{|}c-\tfrac{1}{2}\big{(}U(-h)+U(0)\big{)}\big{|}\geq\tfrac{1}{2}\big{(}U(0)-U(-h)\big{)}+M,

it holds

(3.2) |cj(Y(k,c)kcothkh)|Cj!μMj1,μ=(1+k2)12.|\partial_{c}^{j}\big{(}Y(k,c)-k\coth kh\big{)}|\leq Cj!\mu M^{-j-1},\quad\mu=(1+k^{2})^{-\frac{1}{2}}.

Moreover, if l04l_{0}\geq 4 then for any q[1,)q\in[1,\infty), j1,j20j_{1},j_{2}\geq 0, j22j_{2}\leq 2, and j1+j2l04j_{1}+j_{2}\leq l_{0}-4, kj1cRj2Y(k,c)\partial_{k}^{j_{1}}\partial_{c_{R}}^{j_{2}}Y(k,c) are LkLcRqL_{k}^{\infty}L_{c_{R}}^{q} locally in kk and cRc_{R} in the domain D(Y)D(Y).

Proof.

We first work on the regularity of YY near c=U(h)c=U(-h). Fix k0k_{0}\in\mathbb{R}. Our proof is based on an integral formula of Y(k,c)Y(k,c) for kk near k0k_{0} and cD(Y(k,))U([h,0])c\in D(Y(k,\cdot))\setminus U([-h,0]).

The first step is to identify a domain where there are only finitely many zero points of y(k,,0)y_{-}(k,\cdot,0). Due to the Semi-circle Theorem for the linearized Euler equation at a shear flow in a fixed 2-dim channel x2(h,0)x_{2}\in(-h,0), its unstable and stable modes, which correspond to the zero points of y(k,,0)y_{-}(k,\cdot,0), are contained inside the disk (1.8) with a diameter U([h,0])U([-h,0]). The analyticity of yy_{-} in cc with cI>0c_{I}>0 yields all zero points of y(k0,,0)y_{-}(k_{0},\cdot,0) with cI>0c_{I}>0 are isolated. Due to the continuity of yy_{-} in cc restricted to cI0c_{I}\geq 0, Lemma 2.3(4)(5), and the Semi-circle Theorem, the set 𝒦\mathcal{K} of all accumulation points of the zero points of y(k0,,0)y_{-}(k_{0},\cdot,0) is compact and

𝒦U((U′′)1(0))U((h,0)).\mathcal{K}\subset\subset U\big{(}(U^{\prime\prime})^{-1}(0)\big{)}\cap U\big{(}(-h,0)\big{)}.

Therefore, for any ϵ0>0\epsilon_{0}>0, there exists a function ϕC(,[0,ϵ0])\phi\in C^{\infty}(\mathbb{R},[0,\epsilon_{0}]) such that

n{0},ϵ(0,ϵ0],inf{|cc~Riϕ(c~R)|:y(k0,c,0)=0,c~R}>2ϵ,ϕ|[U(h)+ϵ,U(0)ϵ]0,{c𝒰y(k0,c,0)=0}={cj=cjR+icjIj=1,,n},where 𝒰{cR+icI±cIϕ(cR)}.\begin{split}&\exists n\in\mathbb{N}\cup\{0\},\epsilon\in(0,\epsilon_{0}],\;\;\inf\{|c-\tilde{c}_{R}-i\phi(\tilde{c}_{R})|:\ y_{-}(k_{0},c,0)=0,\,\tilde{c}_{R}\in\mathbb{R}\}>2\epsilon,\\ &\phi|_{\mathbb{R}\setminus[U(-h)+\epsilon,U(0)-\epsilon]}\equiv 0,\;\;\{c\in\mathcal{U}\mid y_{-}(k_{0},c,0)=0\}=\{c_{j}=c_{jR}+ic_{jI}\mid j=1,\ldots,n\},\\ &\text{where }\;\mathcal{U}\triangleq\{c_{R}+ic_{I}\mid\pm c_{I}\geq\phi(c_{R})\}.\end{split}

Roughly ϕ(cR)[0,ϵ0]\phi(c_{R})\in[0,\epsilon_{0}] is supported inside U((h,0))U\big{(}(-h,0)\big{)} and the region bounded by the graphs of ±ϕ\pm\phi contains all except finitely many roots of y(k0,,0)y_{-}(k_{0},\cdot,0). Clearly ϕ\phi can be constructed so that it is supported in any prescribed neighborhood of 𝒦\mathcal{K}\subset\mathbb{C}. Let

𝒦~={cR+icIcR[U(h),U(0)],|cI|ϕ(cR)}[U(h),U(0)],\widetilde{\mathcal{K}}=\{c_{R}+ic_{I}\mid c_{R}\in[U(-h),U(0)],\,|c_{I}|\leq\phi(c_{R})\}\supset[U(-h),U(0)],

which is a compact subset of \mathbb{C} with its smooth upper and lower boundaries given by the graph of ±ϕ\pm\phi restricted to [U(h),U(0)][U(-h),U(0)]. For kk close to k0k_{0}, Y(k,c)Y(k,c) is holomorphic in c(𝒦~(j=1nB(cj,ϵ))(j=1nB(cj¯,ϵ)))c\in\mathbb{C}\setminus\big{(}\widetilde{\mathcal{K}}\cup(\cup_{j=1}^{n}B(c_{j},\epsilon))\cup(\cup_{j=1}^{n}B(\overline{c_{j}},\epsilon))\big{)}, where B(c,ϵ)B(c,\epsilon) denotes the open ball in \mathbb{C} centered at cc with radius ϵ\epsilon.

Our next step is to derive an integral formula of Y(k,c)Y(k,c). For any r>0r>0, let

(3.3) 𝒦~r={cR+icIcR[U(h)r,U(0)+r],|cI|ϕ(cR)+r},\widetilde{\mathcal{K}}_{r}=\{c_{R}+ic_{I}\mid c_{R}\in[U(-h)-r,U(0)+r],\,|c_{I}|\leq\phi(c_{R})+r\}\subset\mathbb{C},

which is a neighborhood of 𝒦~\widetilde{\mathcal{K}} roughly with the margin rr. For

r21r1>0 and c𝒦~r2(𝒦~r1(j=1nB(cj,ϵ))(j=1nB(cj¯,ϵ))),r_{2}\gg 1\gg r_{1}>0\;\text{ and }\;c\in\widetilde{\mathcal{K}}_{r_{2}}\setminus\big{(}\widetilde{\mathcal{K}}_{r_{1}}\cup(\cup_{j=1}^{n}B(c_{j},\epsilon))\cup(\cup_{j=1}^{n}B(\overline{c_{j}},\epsilon))\big{)},

the Cauchy Integral Theorem yields

Y(k,c)=\displaystyle Y(k,c)= 12πi(𝒦~r2𝒦~r1)Y(k,c)ccdc12πij=1n(B(cj,2ϵ)+B(cj¯,2ϵ))Y(k,c)ccdc.\displaystyle\frac{1}{2\pi i}\Big{(}\oint_{\partial\widetilde{\mathcal{K}}_{r_{2}}}-\oint_{\partial\widetilde{\mathcal{K}}_{r_{1}}}\Big{)}\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}-\frac{1}{2\pi i}\sum_{j=1}^{n}\Big{(}\oint_{\partial B(c_{j},2\epsilon)}+\oint_{\partial B(\overline{c_{j}},2\epsilon)}\Big{)}\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}.

Since the term U′′(x2)U(x2)c0\frac{U^{\prime\prime}(x_{2})}{U(x_{2})-c^{\prime}}\to 0 in the Rayleigh equation (2.1) as |c||c^{\prime}|\to\infty, one may prove y(k,c,x2)k1sinhk(x2+h)y_{-}(k,c^{\prime},x_{2})\to k^{-1}\sinh k(x_{2}+h) and Y(k,c)kcothkhY(k,c^{\prime})\to k\coth kh as |c||c^{\prime}|\to\infty (see Lemma 3.3 and the proof of Lemma 3.21 in [17]), which also holds even if k=0k=0. Therefore the outer integral along 𝒦~r2\partial\widetilde{\mathcal{K}}_{r_{2}} converges to kcothkhk\coth kh as r2+r_{2}\to+\infty and we obtain

Y(k,c)=\displaystyle Y(k,c)= kcothkh12πi𝒦~r1Y(k,c)cc𝑑c12πij=1n(B(cj,2ϵ)+B(cj¯,2ϵ))Y(k,c)ccdc.\displaystyle k\coth kh-\frac{1}{2\pi i}\oint_{\partial\widetilde{\mathcal{K}}_{r_{1}}}\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}-\frac{1}{2\pi i}\sum_{j=1}^{n}\Big{(}\oint_{\partial B(c_{j},2\epsilon)}+\oint_{\partial B(\overline{c_{j}},2\epsilon)}\Big{)}\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}.

We observe that 𝒦~r1\partial\widetilde{\mathcal{K}}_{r_{1}} is the union of the two graphs of ±(r1+ϕ)\pm(r_{1}+\phi) over [U(h),U(0)][U(-h),U(0)], the left half of the boundary of the square centered at U(h)U(-h) with the horizontal and vertical side length 2r12r_{1}, and the right half boundary of such a square centered at U(0)U(0). As r10+r_{1}\to 0+, due to the continuity of YY at cU(0)c\neq U(0) and its logarithmic upper bound near U(0)U(0) (Lemma 2.4(2)), the Cauchy integrals along the half boundaries of the squares converge to zero as r10+r_{1}\to 0+. Therefore the above integral formula yields

(3.4) Y(k,c)=kcothkh12πij=1n(B(cj,2ϵ)+B(cj¯,2ϵ))Y(k,c)ccdc+I1(k,c),Y(k,c)=k\coth kh-\frac{1}{2\pi i}\sum_{j=1}^{n}\Big{(}\oint_{\partial B(c_{j},2\epsilon)}+\oint_{\partial B(\overline{c_{j}},2\epsilon)}\Big{)}\frac{Y(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}+I_{1}(k,c),

for

c(𝒦~(j=1nB(cj,ϵ))(j=1nB(cj¯,ϵ))),c\in\mathbb{C}\setminus\big{(}\widetilde{\mathcal{K}}\cup(\cup_{j=1}^{n}B(c_{j},\epsilon))\cup(\cup_{j=1}^{n}B(\overline{c_{j}},\epsilon))\big{)},

where, for c𝒦~c\notin\widetilde{\mathcal{K}},

(3.5) I1(k,c)=limr0+12πi±(±U(h)U(0)Y(k,c±iϕ(c)±ir)c±iϕ(c)±ircd(c±iϕ(c)))=12πi±(±csupp(ϕ)Y(k,c±iϕ(c)±0i)c±iϕ(c)cd(c±iϕ(c)))+1πsupp(ϕ)YI(k,c)cc𝑑c.\begin{split}I_{1}(k,c)=&\lim_{r\to 0+}\frac{1}{2\pi i}\sum_{\pm}\Big{(}\pm\int_{U(-h)}^{U(0)}\frac{Y(k,c^{\prime}\pm i\phi(c^{\prime})\pm ir)}{c^{\prime}\pm i\phi(c^{\prime})\pm ir-c}d(c^{\prime}\pm i\phi(c^{\prime}))\Big{)}\\ =&\frac{1}{2\pi i}\sum_{\pm}\Big{(}\pm\int_{c^{\prime}\in supp(\phi)}\frac{Y(k,c^{\prime}\pm i\phi(c^{\prime})\pm 0i)}{c^{\prime}\pm i\phi(c^{\prime})-c}d(c^{\prime}\pm i\phi(c^{\prime}))\Big{)}\\ &+\frac{1}{\pi}\int_{\mathbb{R}\setminus supp(\phi)}\frac{Y_{I}(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}.\end{split}

Here we used the regularity of Y(k,c)Y(k,c) (Lemma 2.4) and the property supp(YI)[U(h),U(0)]supp(Y_{I})\cap\mathbb{R}\subset[U(-h),U(0)]. For csupp(ϕ)c\in\mathbb{R}\setminus supp(\phi) and cI>0c_{I}>0, the above formula applies to c+icI𝒦~c+ic_{I}\notin\widetilde{\mathcal{K}}. By taking cI0+c_{I}\to 0+, we obtain

(3.6) I1(k,c)=12πi±(±csupp(ϕ)Y(k,c±iϕ(c)±0i)c±iϕ(c)cd(c±iϕ(c)))(χsupp(ϕ)YI(k,))(c)+iYI(k,c),csupp(ϕ).\begin{split}I_{1}(k,c)=&\frac{1}{2\pi i}\sum_{\pm}\Big{(}\pm\int_{c^{\prime}\in supp(\phi)}\frac{Y(k,c^{\prime}\pm i\phi(c^{\prime})\pm 0i)}{c^{\prime}\pm i\phi(c^{\prime})-c}d(c^{\prime}\pm i\phi(c^{\prime}))\Big{)}\\ &-\mathcal{H}\big{(}\chi_{\mathbb{R}\setminus supp(\phi)}Y_{I}(k,\cdot)\big{)}(c)+iY_{I}(k,c),\qquad\qquad c\in\mathbb{R}\setminus supp(\phi).\end{split}

For any cc\in\mathbb{C} such that (k0,c)D(Y)(k_{0},c)\in D(Y) (namely, y(k0,c,0)0y_{-}(k_{0},c,0)\neq 0), by taking sufficiently small ϵ0\epsilon_{0} in the choice of ϕ\phi, the above formulas (3.4)–(3.6) apply to Y(k,c)Y(k,c) for kk close to k0k_{0}.

While analytic in (k,c)D(Y)×U([h,0])(k,c)\in D(Y)\setminus\mathbb{R}\times U([-h,0]) and, when restricted to cI0c_{I}\geq 0, YY is Cl02C^{l_{0}-2} in cD(Y(k,))×{U(h)}c\in D(Y(k,\cdot))\setminus\mathbb{R}\times\{U(-h)\} (Lemma 2.4), so we only need to focus on cc near U(h)U(-h) restricted to cI0c_{I}\geq 0. In the above formulas, the integrals along B(cj,2ϵ)\partial B(c_{j},2\epsilon) and its conjugate are smooth in cc near U(h)U(-h) and thus we only need to consider the regularity of the term I1(k,c)I_{1}(k,c). We recall U(h)U(-h) is an interior point of supp(ϕ)csupp(\phi)^{c}\subset\mathbb{R} according to the definition of ϕ\phi. From Lemma 2.4(6), even though YIY_{I} is W3,qW^{3,q} locally in cU([h,0))c\in U\big{(}[-h,0)\big{)} if l04l_{0}\geq 4, when viewed as a function of cc in a whole neighborhood of U(h)U(-h), only kj1cR2YIL\partial_{k}^{j_{1}}\partial_{c_{R}}^{2}Y_{I}\in L^{\infty} holds due to the jump of cR2YI\partial_{c_{R}}^{2}Y_{I} at c=U(h)c=U(-h). The desired regularity of YY follows from that of YIY_{I}, the above representation formula of YY, and the boundedness in LqL^{q} of the convolution by 1c+icI\frac{1}{c^{\prime}+ic_{I}} uniform in the parameter cI0c_{I}\geq 0.

Finally, it remains to prove inequality (3.2). According to (2.9), there exists k1>0k_{1}>0 such that for any |k|k1|k|\geq k_{1}, it holds

|y(k,c,0)|12μsinhμ1h,c,|y_{-}(k,c,0)|\geq\tfrac{1}{2}\mu\sinh\mu^{-1}h,\quad\forall\,c\in\mathbb{C},

and thus Y(k,c)Y(k,c) is well-defined for all cc\in\mathbb{C}. In this case, (2.20) for cU([h,0])c\notin U([-h,0]) becomes

Y(k,c)kcothkh=\displaystyle Y(k,c)-k\coth kh= 1πU(h)U(0)YI(k,c)cc𝑑c=1πU(h)U(0)πU′′(U1(c))y(k,c,U1(c))2U(U1(c))|y(k,c,0)|2(cc)𝑑c.\displaystyle\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{Y_{I}(k,c^{\prime})}{c^{\prime}-c}dc^{\prime}=\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{\pi U^{\prime\prime}(U^{-1}(c^{\prime}))y_{-}(k,c^{\prime},U^{-1}(c^{\prime}))^{2}}{U^{\prime}(U^{-1}(c^{\prime}))|y_{-}(k,c^{\prime},0)|^{2}(c^{\prime}-c)}dc^{\prime}.

Again, (2.9) implies, for |k|k1|k|\geq k_{1} and c=U(x~2c)c^{\prime}=U(\tilde{x}_{2}^{c}) with x~2c[h,c]\tilde{x}_{2}^{c}\in[-h,c],

|y(k,c,x~2c)|2μsinhμ1(x~2c+h)|YI(k,c)|Ce2μx~2c.|y_{-}(k,c^{\prime},\tilde{x}_{2}^{c})|\leq 2\mu\sinh\mu^{-1}(\tilde{x}_{2}^{c}+h)\implies|Y_{I}(k,c^{\prime})|\leq Ce^{\frac{2}{\mu}\tilde{x}_{2}^{c}}.

Therefore

|cj(Y(k,c)kcothkh)|C(j!)dist(c,U([h,0]))j+1U(h)U(0)e2μx~2c𝑑c\displaystyle|\partial_{c}^{j}\big{(}Y(k,c)-k\coth kh\big{)}|\leq\frac{C(j!)}{dist(c,U([-h,0]))^{j+1}}\int_{U(-h)}^{U(0)}e^{\frac{2}{\mu}\tilde{x}_{2}^{c}}dc^{\prime}
\displaystyle\leq C(j!)dist(c,U([h,0]))j+1h0e2x2μ𝑑x2C(j!)μdist(c,U([h,0]))j+1,\displaystyle\frac{C(j!)}{dist(c,U([-h,0]))^{j+1}}\int_{-h}^{0}e^{\frac{2x_{2}}{\mu}}dx_{2}\leq\frac{C(j!)\mu}{dist(c,U([-h,0]))^{j+1}},

where the substitution c=U(x~2c)c^{\prime}=U(\tilde{x}_{2}^{c}) was used in the above integration. Hence inequality (3.2) follows for |k|k1|k|\geq k_{1}. For |k|k1|k|\leq k_{1}, let BMB_{M}\subset\mathbb{C} be the open disk centered at 12(U(h)+U(0))\tfrac{1}{2}\big{(}U(-h)+U(0)\big{)} with radius 12(U(0)U(h))+M2\tfrac{1}{2}\big{(}U(0)-U(-h)\big{)}+\frac{M}{2}. From the Semi-circle Theorem, y(k,c,0)0y_{-}(k,c,0)\neq 0 for all kk and cBMc\notin B_{M} and Y(k,c)Y(k,c) is analytic. From the same procedure in deriving (3.4), but replacing the inner contour by BM\partial B_{M}, we obtain

cj(Y(k,c)kcothkh)=j!2πiBMY(k,c)(cc)j+1𝑑c.\partial_{c}^{j}\big{(}Y(k,c)-k\coth kh\big{)}=-\frac{j!}{2\pi i}\oint_{\partial B_{M}}\frac{Y(k,c^{\prime})}{(c^{\prime}-c)^{j+1}}dc^{\prime}.

Since max{|Y(k,c)||k|k1,cBM}<\max\{|Y(k,c^{\prime})|\mid|k|\leq k_{1},\,c^{\prime}\in\partial B_{M}\}<\infty, due to the Semi-circle Theorem and the regularity of YY, (3.2) follows immediately. ∎

The following corollary is a direct consequence of the above lemma, the definition (2.23) of F(k,c)F(k,c), and Lemma 2.4.

Corollary 3.1.1.

Assume UCl0U\in C^{l_{0}}, l04l_{0}\geq 4, then, when restricted to cI0c_{I}\geq 0 and near c=U(h)c=U(-h), FF is CC^{\infty} in kk and kj1cj2F\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}F is CαC^{\alpha} for any α[0,1)\alpha\in[0,1), j2=0,1j_{2}=0,1, and 0j1l04j20\leq j_{1}\leq l_{0}-4-j_{2}.

3.2. Basic properties of 𝐅(k,c)\mathbf{F}(k,c) and F(k,c)F(k,c)

Recall that eigenvalues of the linearized gravity waves at the shear flow U(x2)U(x_{2}) are given in the form of ikc-ikc where (k,c)(k,c) satisfy 𝐅(k,c)=F(k,c)=0\mathbf{F}(k,c)=F(k,c)=0 with 𝐅\mathbf{F} and FF defined in (2.22) and (2.23). In this subsection, we derive some basic estimates of them. In particular, in Lemma 3.10 we prove a non-degenerate sign property of cF\partial_{c}F at c=U(h)c=U(-h), crucial for the bifurcation of eigenvalues, without assuming the convexity/concavity of UU as in Lemma 4.9 in [17].

Lemma 3.2.

Assume UC3U\in C^{3}, then we have the following for any α(0,12)\alpha\in(0,\frac{1}{2}).

  1. (1)

    There exists C>0C>0 depending on α\alpha, |U|C2|U^{\prime}|_{C^{2}}, and |(U)1|C0|(U^{\prime})^{-1}|_{C^{0}}, such that

    |𝐅(U(0)c)2coshμ1h|C(μ+μα|U(0)c|2)coshμ1h,\displaystyle|\mathbf{F}-(U(0)-c)^{2}\cosh\mu^{-1}h|\leq C\big{(}\mu+\mu^{\alpha}|U(0)-c|^{2}\big{)}\cosh\mu^{-1}h,

    where we recall μ=(1+k2)12\mu=(1+k^{2})^{-\frac{1}{2}}.

  2. (2)

    For any k,M>0k_{*},M>0, there exists C>0C>0 depending only on kk_{*}, MM, and |U′′|C0|U^{\prime\prime}|_{C^{0}}, such that, for any |k|k|k|\leq k_{*} and cc satisfying dist(c,U(h,0])Mdist(c,U(-h,0])\geq M,

    |𝐅(U(0)c)2coshkh|C(1+|c|+|U(0)c|2dist(c,U([h,0]))1).|\mathbf{F}-(U(0)-c)^{2}\cosh kh|\leq C\big{(}1+|c|+|U(0)-c|^{2}dist\big{(}c,U([-h,0])\big{)}^{-1}\big{)}.

Statement (1) is mainly applied for |k|1|k|\gg 1 while statement (2) mainly for |c|1|c|\gg 1.

Proof.

From Lemma 2.2, one observes that the logarithmic singularity in y(k,c,0)y_{-}^{\prime}(k,c,0) is significant only when |c|C|c|\leq C and it becomes bounded when multiplied by U(0)cU(0)-c. Therefore

|𝐅(U(0)c)2coshμ1h|\displaystyle|\mathbf{F}-(U(0)-c)^{2}\cosh\mu^{-1}h|\leq C(1+|U(0)c|)|y(k,c,0)|\displaystyle C(1+|U(0)-c|)|y_{-}(k,c,0)|
+|U(0)c|2|y(k,c,0)coshμ1h|\displaystyle\qquad+|U(0)-c|^{2}|y_{-}^{\prime}(k,c,0)-\cosh\mu^{-1}h|
\displaystyle\leq C(μ+μ|U(0)c|+μα|U(0)c|2)coshμ1h,\displaystyle C(\mu+\mu|U(0)-c|+\mu^{\alpha}|U(0)-c|^{2})\cosh\mu^{-1}h,

which implies statement (1) in the lemma.

When |k|k|k|\leq k_{*}, dist(c,U([h,0]))Mdist\big{(}c,U([-h,0])\big{)}\geq M, and x2[h,0]x_{2}\in[-h,0], it is easy to estimate the Rayleigh equation (2.1) and derive that y(k,c,x2)y_{-}(k,c,x_{2}) and y(k,c,x2)y_{-}^{\prime}(k,c,x_{2}) are uniformly bounded by some C>0C>0 depending on kk_{*}, MM, and |U′′|C0|U^{\prime\prime}|_{C^{0}}. Hence by regarding U′′Ucy\frac{U^{\prime\prime}}{U-c}y_{-} in (2.1) as a perturbation term, from the variation of parameter formula, it is straight forward to estimate

|y(k,c,0)k1sinhkh|+|y(k,c,0)coshkh|Cdist(c,U([h,0]))1.|y_{-}(k,c,0)-k^{-1}\sinh kh|+|y_{-}^{\prime}(k,c,0)-\cosh kh|\leq Cdist\big{(}c,U([-h,0])\big{)}^{-1}.

Statement (2) follows immediately. ∎

The next lemma states that U(h)U(-h) becomes a singular neutral mode for a unique kk_{-}.

Lemma 3.3.

Assume UC3U\in C^{3}, then y(k,U(h),0)>0y_{-}(k,U(-h),0)>0 for any kk\in\mathbb{R} and there exists k>0k_{-}>0, unique among k[0,)k\in[0,\infty), such that F(k,U(h))=0F(k_{-},U(-h))=0.

Proof.

The statement y(k,U(h),0)>0y_{-}(k,U(-h),0)>0 had been given Lemma 2.3(4) and thus F(k,U(h))F(k,U(-h))\in\mathbb{R} is well-defined for all kk\in\mathbb{R} and even in kk. The existence and uniqueness of kk_{-} is a direct corollary of Lemma 2.6, which says F(0,U(h))=gF(0,U(-h))=-g and F(,U(h))F(\cdot,U(-h)) is increasing in k>0k>0, and Lemma 3.2(1), which implies F(k,U(h))>0F(k,U(-h))>0 for k1k\gg 1. ∎

3.3. Interior singular neutral modes

Lemma 2.7 on the analytic continuation and the continuity of 𝐅\mathbf{F} (Lemma 2.5) when restricted to cI0c_{I}\geq 0 imply that any branch of roots of 𝐅\mathbf{F} can be continued until it either collides with another branch or approaches singular neutral modes – roots of 𝐅\mathbf{F} with cU([h,0])c\in U([-h,0]) which are at the boundary of analyticity of 𝐅\mathbf{F}. While c=U(0)c=U(0) is impossible (Lemma 2.6(1)), besides (k,U(h))(k_{-},U(-h)) obtained in lemma 3.3, in the following we give some basic properties of singular neutral modes with cU((h,0))c\in U((-h,0)).

Lemma 3.4.

Assume UC3U\in C^{3}. The following are equivalent for c0=U(x20)U((h,0])c_{0}=U(x_{20})\in U((-h,0]).

  1. (1)

    There exists k0k_{0}\in\mathbb{R} such that 𝐅(k0,c0)=0\mathbf{F}(k_{0},c_{0})=0.

  2. (2)

    There exists a k00k_{0}\geq 0 such that F(k0,c0)=0F(k_{0},c_{0})=0.

  3. (3)

    U′′(x20)=0U^{\prime\prime}(x_{20})=0 and the Sturm-Liouville operator \mathcal{R} has a non-positive eigenvalue, where :=x22+U′′Uc0\mathcal{R}:=-\partial_{x_{2}}^{2}+\tfrac{U^{\prime\prime}}{U-c_{0}} with boundary conditions

    y(h)=0,(U(0)c0)2y(0)(U(0)(U(0)c0)+g)y(0)=0.y(-h)=0,\quad(U(0)-c_{0})^{2}y^{\prime}(0)-(U^{\prime}(0)(U(0)-c_{0})+g)y(0)=0.

Moreover, it is also satisfied that KF(K,c0)|K=k2>0\partial_{K}F(\sqrt{K},c_{0})|_{K=k^{2}}>0 for such c0c_{0} and any kk\in\mathbb{R} with y(k,c0,0)0y_{-}(k,c_{0},0)\neq 0.

We observe that the assumption U′′(x20)=0U^{\prime\prime}(x_{20})=0 implies that \mathcal{R} along with its boundary conditions is a self-adjoint operator with C0C^{0} coefficient.

Remark 3.1.

The last statement also applies to c0=U(h)c_{0}=U(-h). Namely (k)2-(k_{-})^{2} is the eigenvalue of \mathcal{R} at c=U(h)c=U(-h).

Proof.

Before we prove the equivalence of the above statements, we first extend the monotonicity, as well as the concavity under certain conditions, of Y(k,c)Y(k,c) and F(k,c)F(k,c) with respect to K=k2K=k^{2} (originally for cU((h,0])c\in\mathbb{R}\setminus U((-h,0]) in Lemma 2.4(8) and 2.6(2), respectively) also to any c0=U(x20)U((h,0))c_{0}=U(x_{20})\in U((-h,0)) satisfying U′′(x20)=0U^{\prime\prime}(x_{20})=0. Here we observe that U′′(x20)=0U^{\prime\prime}(x_{20})=0 implies y(k,c0,x2)y_{-}(k,c_{0},x_{2})\in\mathbb{R} and is C2C^{2} even for x2x_{2} near x20x_{20} for such c0c_{0} and any kk\in\mathbb{R} and thus Y(k,c0),F(k,c0)Y(k,c_{0}),F(k,c_{0})\in\mathbb{R} as well. The proof is similar to that of Lemma 4.5 in [17], starting with the following claim.

Claim. Assume c0U((h,0])c_{0}\in\mathbb{C}\setminus U\big{(}(-h,0]\big{)} or c0=U(x20)U((h,0))c_{0}=U(x_{20})\in U\big{(}(-h,0)\big{)} with U′′(x20)=0U^{\prime\prime}(x_{20})=0 and y~,yC0([h,0])C2((h,0))\tilde{y},y\in C^{0}([-h,0])\cap C^{2}((-h,0)) are solutions to

(+k2)y~=0,y~(h)=0,y~(0)=1;(+k2)y=fC0([h,0]),y(h)=y(0)=0;(\mathcal{R}+k^{2})\tilde{y}=0,\;\tilde{y}(-h)=0,\;\tilde{y}(0)=1;\quad(\mathcal{R}+k^{2})y=f\in C^{0}([-h,0]),\;y(-h)=y(0)=0;

then we have

(3.7) y(0)=h0y~f𝑑x2.y^{\prime}(0)=-\int_{-h}^{0}\tilde{y}fdx_{2}.

Moreover, if y~0\tilde{y}\neq 0 on (h,0)(-h,0) which in particular is true if c0U((h,0))c_{0}\in\mathbb{R}\setminus U\big{(}(-h,0)\big{)} due to lemma 2.3(4), then it also holds

(3.8) y(x2)=y~(x2)x201y~(x2)2hx2y~(x2′′)f(x2′′)𝑑x2′′𝑑x2.y(x_{2})=\tilde{y}(x_{2})\int_{x_{2}}^{0}\frac{1}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\tilde{y}(x_{2}^{\prime\prime})f(x_{2}^{\prime\prime})dx_{2}^{\prime\prime}dx_{2}^{\prime}.

Here the assumption on c0c_{0} insures that \mathcal{R} has C0C^{0} coefficients over [h,0][-h,0]. The claim follows through direct computations using (y~yy~y)=y~f(\tilde{y}^{\prime}y-\tilde{y}y^{\prime})^{\prime}=\tilde{y}f.

If y(k,c0,0)0y_{-}(k,c_{0},0)\neq 0, then apparently y~=yy(0)\tilde{y}=\tfrac{y_{-}}{y_{-}(0)}\in\mathbb{R} and Y(k,c0)=y~(0)Y(k,c_{0})=\tilde{y}^{\prime}(0). It is straight forward to compute by differentiating with respect to K=k2K=k^{2},

KY(k,c0)=Ky~(0),K2Y(k,c0)=K2y~(0),(+k2)Ky~=y~,(+k2)K2y~=2Ky~,\partial_{K}Y(k,c_{0})=\partial_{K}\tilde{y}^{\prime}(0),\;\;\partial_{K}^{2}Y(k,c_{0})=\partial_{K}^{2}\tilde{y}^{\prime}(0),\;\;(\mathcal{R}+k^{2})\partial_{K}\tilde{y}=-\tilde{y},\;\;(\mathcal{R}+k^{2})\partial_{K}^{2}\tilde{y}=-2\partial_{K}\tilde{y},

and Ky~\partial_{K}\tilde{y} and K2y~\partial_{K}^{2}\tilde{y} satisfy the zero Dirichlet boundary conditions assumed in the claim. Applying (3.7) to Ky~\partial_{K}\tilde{y} and K2y~\partial_{K}^{2}\tilde{y}, along with Lemma 2.3(4)(5), implies, for c0U((h,0])c_{0}\in\mathbb{R}\setminus U\big{(}(-h,0]\big{)} or c0=U(x20)U([h,0))c_{0}=U(x_{20})\in U\big{(}[-h,0)\big{)} with U′′(x20)=0U^{\prime\prime}(x_{20})=0 as long as y(k,c0,0)0y_{-}(k,c_{0},0)\neq 0,

(3.9) KY(k,c0)=Ky~(0)=h0y~2𝑑x2>0,KF(k,c0)>0.\partial_{K}Y(k,c_{0})=\partial_{K}\tilde{y}^{\prime}(0)=\int_{-h}^{0}\tilde{y}^{2}dx_{2}>0,\quad\partial_{K}F(k,c_{0})>0.

Moreover, if y(k,c0,x2)0y_{-}(k,c_{0},x_{2})\neq 0 on (h,0)(-h,0), then

(3.10) K2Y(k,c0)=2h0y~(x2)2x20y~(x2)2hx2y~(x2′′)2𝑑x2′′𝑑x2𝑑x2<0,K2F(k,c0)<0.\partial_{K}^{2}Y(k,c_{0})=-2\int_{-h}^{0}\tilde{y}(x_{2})^{2}\int_{x_{2}}^{0}\tilde{y}(x_{2}^{\prime})^{-2}\int_{-h}^{x_{2}^{\prime}}\tilde{y}(x_{2}^{\prime\prime})^{2}dx_{2}^{\prime\prime}dx_{2}^{\prime}dx_{2}<0,\quad\partial_{K}^{2}F(k,c_{0})<0.

We are ready to prove the lemma. According to Lemmas 2.5(2)(3) and 2.6(1), F(k0,c0)=0F(k_{0},c_{0})=0 iff 𝐅(±k0,c0)=0\mathbf{F}(\pm k_{0},c_{0})=0 where y(k0,c0,0)0y_{-}(k_{0},c_{0},0)\neq 0 and x200x_{20}\neq 0 are necessary. Consequently (3.9) immediately yields the equivalence of statements (1) and (2). Moreover Lemmas 2.5(4) and 2.3(5) imply U′′(x20)=0U^{\prime\prime}(x_{20})=0 or c0=U(h)c_{0}=U(-h), where y(k,c0,x2)y_{-}(k,c_{0},x_{2})\in\mathbb{R} in both cases. We obtain the equivalence with statement (3) simply by observing the associated Sturm-Liouville problem structure for the neutral modes. ∎

Next we prove that any interior inflection value of UU is a singular neutral mode for some wave number, which is through a different proof as in [28, 11].

Lemma 3.5.

Suppose UC3U\in C^{3} and c0=U(x20)U((h,0))c_{0}=U(x_{20})\in U((-h,0)) satisfy U′′(x20)=0U^{\prime\prime}(x_{20})=0, then there exists k0>0k_{0}>0 such that F(k0,c0)=0F(k_{0},c_{0})=0 with its eigenfunction y(k0,c0,x2)>0y_{-}(k_{0},c_{0},x_{2})>0 for all x2(h,0]x_{2}\in(-h,0].

Proof.

According to Lemma 3.4, F(k,c0)=0F(k,c_{0})=0 iff y(k,c0,x2)y_{-}(k,c_{0},x_{2}) is an eigenfunction with the eigenvalue k2-k^{2} of the Sturm-Liouville operator \mathcal{R} along with boundary conditions in (1.5). It corresponds to the variational functional

I(c0,y)=12h0|y(x2)|2+U′′(x2)U(x2)c0|y(x2)|2dx2U(0)(U(0)c0)+g2(U(0)c0)2|y(0)|2,I(c_{0},y)=\frac{1}{2}\int_{-h}^{0}|y^{\prime}(x_{2})|^{2}+\frac{U^{\prime\prime}(x_{2})}{U(x_{2})-c_{0}}|y(x_{2})|^{2}dx_{2}-\frac{U^{\prime}(0)(U(0)-c_{0})+g}{2(U(0)-c_{0})^{2}}|y(0)|^{2},

for yH1((h,0))y\in H^{1}\big{(}(-h,0)\big{)} with y(h)=0y(-h)=0. Consider

y(x2)=0, if x2[h,x20];y(x2)=U(x2)c0, if x2[x20,0],y(x_{2})=0,\;\text{ if }\;x_{2}\in[-h,x_{20}];\quad y(x_{2})=U(x_{2})-c_{0},\;\text{ if }\;x_{2}\in[x_{20},0],

which is clearly a qualified test function. It is straight forward to verify

I(c0,y)=\displaystyle I(c_{0},y)= 12x200U(x2)2+U′′(x2)(U(x2)c0)dx212(U(0)(U(0)c0)+g)\displaystyle\frac{1}{2}\int_{x_{20}}^{0}U^{\prime}(x_{2})^{2}+U^{\prime\prime}(x_{2})(U(x_{2})-c_{0})dx_{2}-\frac{1}{2}\big{(}U^{\prime}(0)(U(0)-c_{0})+g\big{)}
=\displaystyle= 12U(x2)(U(x2)c0)|x20012(U(0)(U(0)c0)+g)=g2<0.\displaystyle\frac{1}{2}U^{\prime}(x_{2})(U(x_{2})-c_{0})\big{|}_{x_{20}}^{0}-\frac{1}{2}\big{(}U^{\prime}(0)(U(0)-c_{0})+g\big{)}=-\frac{g}{2}<0.

It yields the negative sign of the first eigenvalue k02<0-k_{0}^{2}<0 of \mathcal{R} with its eigenfunction y(k0,c0,x2)>0y_{-}(k_{0},c_{0},x_{2})>0. ∎

Remark 3.2.

The proof of Lemma 3.5, which is different from that in [28, 10, 11], actually implies the existence of singular neutral modes at inflection values without assuming the monotonicity of UU. More precisely, it holds that
"Suppose UC3U\in C^{3}, U1(c0)={x2,1<<x2,n}[h,0)U^{-1}(c_{0})=\{x_{2,1}<\ldots<x_{2,n}\}\subset[-h,0), n1n\geq 1, and U′′(x2)=0U^{\prime\prime}(x_{2})=0 for all x2U1(c0)x_{2}\in U^{-1}(c_{0}), then F(,c0)F(\cdot,c_{0}) has at least nn distinct roots in [0,+)[0,+\infty) and at least max{1,n1}\max\{1,n-1\} roots in +\mathbb{R}^{+}."
To see this, one may consider the test functions yj(x2)=(Uc0)χ(x2,j,x2,j+1)y_{j}(x_{2})=(U-c_{0})\chi_{(x_{2,j},x_{2,j+1})} where x2,n+1=0x_{2,n+1}=0 is understood. Clearly they are L2L^{2}-orthogonal and I(y1)==I(yn1)=0I(y_{1})=\ldots=I(y_{n-1})=0 and I(yn)=g/2I(y_{n})=-g/2. Hence the non-positive subspace is at least nn-dim. The statement follows immediately since each eigenvalue of this Sturm-Liouville problem has only one linear independent eigenfunction. The above assumptions on UU is more general than those on the so-called class 𝒦+\mathcal{K}^{+} in [10, 11].

Given an interior inflection value c0c_{0}, the property KF(k,c0)>0\partial_{K}F(k,c_{0})>0, whenever y(k,c0,0)0y_{-}(k,c_{0},0)\neq 0, does not imply the uniqueness of the root of F(,c0)=0F(\cdot,c_{0})=0 due to the possibility of y(k,c0,0)=0y_{-}(k,c_{0},0)=0 for some k>0k>0. In the following we address the number of wave numbers which make c0c_{0} a singular neutral mode. We first prove a lemma on the relationship among several critical wave numbers.

Lemma 3.6.

Suppose UC3U\in C^{3} and c0=U(x20)U((h,0))c_{0}=U(x_{20})\in U((-h,0)) satisfy U′′(x20)=0U^{\prime\prime}(x_{20})=0 and k0>0k_{0}>0 is the maximal root of F(,c0)F(\cdot,c_{0}), then the following hold.

  1. (1)

    If kC0k_{C}\geq 0 satisfies y(kC,c0,0)=0y_{-}(k_{C},c_{0},0)=0, then k0>kCk_{0}>k_{C}.

  2. (2)

    If U′′Uc00\frac{U^{\prime\prime}}{U-c_{0}}\leq 0 on (h,0)(-h,0), then k0>kk_{0}>k_{-} where kk_{-} is the unique root of F(,U(h))F(\cdot,U(-h)) given in Lemma 3.3.

Clearly the above (kC,c0)(k_{C},c_{0}) coincides with a singular neutral mode for the linearized channel flow at the monotonic UU.

Proof.

It is clear that y(k,c0,0)=0y_{-}(k,c_{0},0)=0 iff y(k,c0,x2)y_{-}(k,c_{0},x_{2}) is an eigenfunction with the eigenvalue k2-k^{2} of the Sturm-Liouville problem

(3.11) y=λy,y(h)=y(0)=0,\mathcal{R}y=\lambda y,\quad y(-h)=y(0)=0,

which corresponds to the variational functional

IC(c0,y)=12h0|y(x2)|2+U′′(x2)U(x2)c0|y(x2)|2dx2,yH01((h,0)).I_{C}(c_{0},y)=\frac{1}{2}\int_{-h}^{0}|y^{\prime}(x_{2})|^{2}+\frac{U^{\prime\prime}(x_{2})}{U(x_{2})-c_{0}}|y(x_{2})|^{2}dx_{2},\quad y\in H_{0}^{1}\big{(}(-h,0)\big{)}.

Since obviously IICI\leq I_{C} and ICI_{C} has one more restriction y(0)=0y(0)=0 on its domain, so the first eigenvalue k02-k_{0}^{2} of I(c0,y)I(c_{0},y) must satisfy k02<kC2-k_{0}^{2}<-k_{C}^{2}.

One may verify

I(c0,y)I(U(h),y)=\displaystyle I(c_{0},y)-I(U(-h),y)= 12h0(c0U(h))U′′y2(Uc0)(UU(h))dx2(U(0)(c0U(h))(U(0)c0)(U(0)U(h))\displaystyle\frac{1}{2}\int_{-h}^{0}\frac{(c_{0}-U(-h))U^{\prime\prime}y^{2}}{(U-c_{0})(U-U(-h))}dx_{2}-\Big{(}\frac{U^{\prime}(0)(c_{0}-U(-h))}{(U(0)-c_{0})(U(0)-U(-h))}
+g(c0U(h))(2U(0)c0U(h))(U(0)c0)2(U(0)U(h))2)|y(0)|22.\displaystyle+\frac{g(c_{0}-U(-h))(2U(0)-c_{0}-U(-h))}{(U(0)-c_{0})^{2}(U(0)-U(-h))^{2}}\Big{)}\frac{|y(0)|^{2}}{2}.

Hence the assumption U′′Uc00\frac{U^{\prime\prime}}{U-c_{0}}\leq 0 on (h,0)(-h,0) implies I(c0,y)I(U(h),y)I(c_{0},y)\leq I(U(-h),y). Since y(k,U(h),0)0y_{-}(k_{-},U(-h),0)\neq 0, we obtain that the first eigenvalue k02-k_{0}^{2} of I(c0,y)I(c_{0},y) satisfy k02<(k)2-k_{0}^{2}<-(k_{-})^{2}. ∎

Finally we show how whether an interior inflection value c0c_{0} has either one or two critical wave numbers depends on kCk_{C} and gg.

Lemma 3.7.

Suppose UC3U\in C^{3} and c0=U(x20)U((h,0))c_{0}=U(x_{20})\in U((-h,0)) satisfy U′′(x20)=0U^{\prime\prime}(x_{20})=0, then the following hold.

  1. (1)

    There exists at most one kC0k_{C}\geq 0 such that y(kC,c0,0)=0y_{-}(k_{C},c_{0},0)=0.

  2. (2)

    Such kC>0k_{C}>0 exists iff F0(c0)>0F^{0}(c_{0})>0, where

    (3.12) F0(c0)F(0,c0)+g=(U(0)c0)2Y(0,c0)U(0)(U(0)c0).F^{0}(c_{0})\triangleq F(0,c_{0})+g=(U(0)-c_{0})^{2}Y(0,c_{0})-U^{\prime}(0)(U(0)-c_{0}).
  3. (3)

    Let k=kCk_{*}=k_{C} if kCk_{C} exists or k=0k_{*}=0 otherwise, there exists a unique k0>kk_{0}>k_{*} such that F(k0,c0)=0F(k_{0},c_{0})=0. It also holds that the corresponding eigenfunction y(k0,c0,x2)>0y_{-}(k_{0},c_{0},x_{2})>0 on (h,0](-h,0] and K2F(k0,c0)<0\partial_{K}^{2}F(k_{0},c_{0})<0.

  4. (4)

    If kC>0k_{C}>0, then

    1. (a)

      if gF0(c0)g\geq F^{0}(c_{0}), then there exists k1[0,kC)k_{1}\in[0,k_{C}) such that {k0F(k,c0)=0}={k1,k0}\{k\geq 0\mid F(k,c_{0})=0\}=\{k_{1},k_{0}\}, and

    2. (b)

      if g<F0(c0)g<F^{0}(c_{0}), then k0k_{0} is the only root of F(,c0)F(\cdot,c_{0}) on [0,)[0,\infty).

Proof.

Recall that y(k,c0,0)=0y_{-}(k,c_{0},0)=0 iff y(k,c0,x2)y_{-}(k,c_{0},x_{2}) is an eigenfunction with the eigenvalue k2-k^{2} of (3.11) corresponding to the variational functional IC(y)I_{C}(y) on H01((h,0))H_{0}^{1}\big{(}(-h,0)\big{)}. Suppose λ0-\lambda\leq 0 is an eigenvalue with an eigenfunction y(x2)y(x_{2}), then Lemma 2.3(5) applied to initial conditions at both x2=hx_{2}=-h and x2=0x_{2}=0 implies y0y\neq 0 on (h,x20](-h,x_{20}] and [x20,0)[x_{20},0), and thus y0y\neq 0 on both (h,0)(-h,0). Hence it is the first eigenvalue and therefore the only non-negative one of (3.11), which also implies kC2=λ0k_{C}^{2}=\lambda\geq 0 is unique if it exists. This proves statement (1).

On the one hand, suppose such kC>0k_{C}>0 does not exist, then F(k,c0)F(k,c_{0}) is well-defined for all k>0k>0. The fact kF(k,c0)>0\partial_{k}F(k,c_{0})>0 for all k>0k>0 (Lemma 3.4) and F(k0,c0)=0F(k_{0},c_{0})=0 with k0>0k_{0}>0 (Lemma 3.5) imply F(0,c0)<0F(0,c_{0})<0 if it is well-defined. Since this conclusion is independent of g>0g>0, we obtain F0(c0)0F^{0}(c_{0})\leq 0 if F(0,c0)F(0,c_{0}) is well-defined. On the other hand, assume kC>0k_{C}>0 exists. From the Sturm-Liouville theory (or one may easily prove it directly), y(0,c0,x2)y_{-}(0,c_{0},x_{2}) has a root x~2(h,0)\tilde{x}_{2}\in(-h,0) with y(0,c0,x~2)<0y_{-}^{\prime}(0,c_{0},\tilde{x}_{2})<0. Moreover Lemma 2.3(5) yields x~2>x20\tilde{x}_{2}>x_{20}. As in (2.26), it is straight forward to verify, for x2>x~2x_{2}>\tilde{x}_{2},

y(0,c0,x2)=y(0,c0,x~2)(U(x2)c0)x~2x2U(x~2)c0(Uc0)2𝑑x2.y_{-}(0,c_{0},x_{2})=y_{-}^{\prime}(0,c_{0},\tilde{x}_{2})(U(x_{2})-c_{0})\int_{\tilde{x}_{2}}^{x_{2}}\frac{U(\tilde{x}_{2})-c_{0}}{(U-c_{0})^{2}}dx_{2}.

Consequently, one may compute

F0(c0)=(x~201(Uc0)2𝑑x2)1>0.F^{0}(c_{0})=\Big{(}\int_{\tilde{x}_{2}}^{0}\frac{1}{(U-c_{0})^{2}}dx_{2}\Big{)}^{-1}>0.

This completes the proof of statement (2).

Statements (3) and (4) follow from Lemmas 3.5, 3.6, and kF(k,c0)>0\partial_{k}F(k,c_{0})>0 for all kkCk\neq k_{C} if kC0k_{C}\geq 0 exists (Lemma 3.4). ∎

3.4. Non-degeneracy of FF at (k,c=U(h))(k_{-},c=U(-h)).

As indicated in Lemmas 2.5(4) and 2.6(1), c=U(h)c=U(-h) is the only singular neutral mode which is not an inflection value of UU. It would be one of the key bifurcation points where the instability occurs. Our goal in the following is to verify a non-degeneracy of FF at c=U(h)c=U(-h) to be used in the bifurcation analysis. We need the following two lemmas of technical preparation. The first is an estimate related to the fundamental solution y(k,c,x2)y_{-}(k,c,x_{2}).

Lemma 3.8.

Suppose UCl0U\in C^{l_{0}}, l04l_{0}\geq 4, and let y~(k,c,x2)=y(k,c,x2)y(k,c,0)\tilde{y}(k,c,x_{2})=\frac{y_{-}(k,c,x_{2})}{y_{-}(k,c,0)} if y(k,c,0)0y_{-}(k,c,0)\neq 0, then, when restricted to cI0c_{I}\geq 0 and x2(h,0]x_{2}\in(-h,0], for j1,j20j_{1},j_{2}\geq 0, j1+j2l04j_{1}+j_{2}\leq l_{0}-4, j22j_{2}\leq 2, and q[1,)q\in[1,\infty), it holds that kj1cj2y~,kj1cj2y~LkLcRq\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}\tilde{y},\,\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}\tilde{y}^{\prime}\in L_{k}^{\infty}L_{c_{R}}^{q} locally in kk and cRc_{R}. Moreover, at c=U(h)c=U(-h), it holds,

limx2(h)+1(x2+h)2(y~(x2)U(x2)y~(x2)U(x2)U(h))=13k2y~(h),\lim_{x_{2}\to(-h)+}\frac{1}{(x_{2}+h)^{2}}\Big{(}\tilde{y}^{\prime}(x_{2})-\frac{U^{\prime}(x_{2})\tilde{y}(x_{2})}{U(x_{2})-U(-h)}\Big{)}=\frac{1}{3}k^{2}\tilde{y}^{\prime}(-h),

and there exists C~>0\tilde{C}>0 depending on UU and kk such that, for x2(h,0]x_{2}\in(-h,0],

|y~(x2)|C~,C~(x2+h)y~(x2)C~1(x2+h),|\tilde{y}^{\prime}(x_{2})|\leq\tilde{C},\quad\tilde{C}(x_{2}+h)\geq\tilde{y}(x_{2})\geq\tilde{C}^{-1}(x_{2}+h),
|cy~(x2)|C~(x2+h)(1+|log(x2+h)|),|cy~(x2)|C~(1+|log(x2+h)|).|\partial_{c}\tilde{y}(x_{2})|\leq\tilde{C}(x_{2}+h)(1+|\log(x_{2}+h)|),\;\;|\partial_{c}\tilde{y}^{\prime}(x_{2})|\leq\tilde{C}(1+|\log(x_{2}+h)|).

According to Lemmas 2.2 and 2.3(4)(5), it holds y(k,c,x2)0y_{-}(k,c,x_{2})\neq 0 for any x2(h,0]x_{2}\in(-h,0] and (k,c)×(k,c)\in\mathbb{R}\times\mathbb{C} in an open set containing ×{U(h)}\mathbb{R}\times\{U(-h)\}, and thus y~\tilde{y} is well defined for such (k,c)(k,c). For c=U(h)c=U(-h), Lemma 2.3(3) and y(h)=0y_{-}(-h)=0 imply y~\tilde{y} is C4C^{4} in x2x_{2} and y~Uy~/(UU(h))\tilde{y}^{\prime}-U^{\prime}\tilde{y}/(U-U(-h)) vanishes at x2=hx_{2}=-h. The above lemma shows that the latter is actually of quadratic order and gives the leading order coefficient.

Proof.

From the Rayleigh equation (2.1), y~\tilde{y} satisfies

(3.13) y~′′+(k2+U′′Uc)y~=0,y~(h)=0,y~(0)=1,Y(k,c)=y~(0).-\tilde{y}^{\prime\prime}+\big{(}k^{2}+\frac{U^{\prime\prime}}{U-c}\big{)}\tilde{y}=0,\quad\tilde{y}(-h)=0,\;\tilde{y}(0)=1,\;Y(k,c)=\tilde{y}^{\prime}(0).

For x2(h,0]x_{2}\in(-h,0], y~(k,c,x2)\tilde{y}(k,c,x_{2}) obeys the above Rayleigh equation with smooth coefficients and initial values (1,Y(k,c))(1,Y(k,c)) given at x2=0x_{2}=0. Lemma 3.1 implies the regularity of y~\tilde{y} and y~\tilde{y}^{\prime}.

The desired estimates of y~\tilde{y} and |y~||\tilde{y}^{\prime}| at c=U(h)c=U(-h) follow immediately from Lemma 2.3(3)(4) along with y~(h)0\tilde{y}^{\prime}(-h)\neq 0. With the differentiability in cc near U(h)U(-h) had been obtained, differentiating (3.13) in cc yields

(3.14) cy~′′+(k2+U′′Uc)cy~=c(U′′Uc)y~,cy~(h)=cy~(0)=0,cY(k,c)=cy~(0).-\partial_{c}\tilde{y}^{\prime\prime}+\big{(}k^{2}+\tfrac{U^{\prime\prime}}{U-c}\big{)}\partial_{c}\tilde{y}=-\partial_{c}\big{(}\tfrac{U^{\prime\prime}}{U-c}\big{)}\tilde{y},\quad\partial_{c}\tilde{y}(-h)=\partial_{c}\tilde{y}(0)=0,\quad\partial_{c}Y(k,c)=\partial_{c}\tilde{y}^{\prime}(0).

For cU(h)c\leq U(-h), the claim in the proof of Lemma 3.4 applies and Lemma 2.3(4) and (3.8) imply

(3.15) cy~(x2)=x20y~(x2)y~(x2)2hx2c(U′′Uc)y~2dx2′′dx2=x20y~(x2)y~(x2)2hx2U′′y~2(Uc)2𝑑x2′′𝑑x2\begin{split}\partial_{c}\tilde{y}(x_{2})=&-\int_{x_{2}}^{0}\frac{\tilde{y}(x_{2})}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\partial_{c}\big{(}\frac{U^{\prime\prime}}{U-c}\big{)}\tilde{y}^{2}dx_{2}^{\prime\prime}dx_{2}^{\prime}=-\int_{x_{2}}^{0}\frac{\tilde{y}(x_{2})}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\frac{U^{\prime\prime}\tilde{y}^{2}}{(U-c)^{2}}dx_{2}^{\prime\prime}dx_{2}^{\prime}\end{split}

Using the estimates on y~\tilde{y} and y~\tilde{y}^{\prime} we obtain, at c=U(h)c=U(-h),

|cy~(x2)|\displaystyle\big{|}\partial_{c}\tilde{y}(x_{2})\big{|}\leq x20y~(x2)y~(x2)2hx2|U′′|y~2(UU(h))2𝑑x2′′𝑑x2\displaystyle\int_{x_{2}}^{0}\frac{\tilde{y}(x_{2})}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\frac{|U^{\prime\prime}|\tilde{y}^{2}}{(U-U(-h))^{2}}dx_{2}^{\prime\prime}dx_{2}^{\prime}
\displaystyle\leq C~x20x2+hx2+h𝑑x2C~(x2+h)(1+|log(x2+h)|),\displaystyle\tilde{C}\int_{x_{2}}^{0}\frac{x_{2}+h}{x_{2}^{\prime}+h}dx_{2}^{\prime}\leq\tilde{C}(x_{2}+h)\big{(}1+|\log(x_{2}+h)|\big{)},
|cy~(x2)|=\displaystyle\big{|}\partial_{c}\tilde{y}^{\prime}(x_{2})\big{|}= |x20y~(x2)y~(x2)2hx2U′′y~2(UU(h))2𝑑x2′′𝑑x2+1y~(x2)hx2U′′y~2(UU(h))2𝑑x2|\displaystyle\Big{|}-\int_{x_{2}}^{0}\frac{\tilde{y}^{\prime}(x_{2})}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\frac{U^{\prime\prime}\tilde{y}^{2}}{(U-U(-h))^{2}}dx_{2}^{\prime\prime}dx_{2}^{\prime}+\frac{1}{\tilde{y}(x_{2})}\int_{-h}^{x_{2}}\frac{U^{\prime\prime}\tilde{y}^{2}}{(U-U(-h))^{2}}dx_{2}^{\prime}\Big{|}
\displaystyle\leq C~(1+|log(x2+h)|).\displaystyle\tilde{C}\big{(}1+|\log(x_{2}+h)|\big{)}.

To complete the proof of the lemma, for c=U(h)c=U(-h), we consider the property of y~Uy~/(UU(h))\tilde{y}^{\prime}-U^{\prime}\tilde{y}/(U-U(-h)) near x2=hx_{2}=-h. From (3.13), it holds

((UU(h))y~Uy~)=k2(UU(h))y~.\big{(}(U-U(-h))\tilde{y}^{\prime}-U^{\prime}\tilde{y}\big{)}^{\prime}=k^{2}(U-U(-h))\tilde{y}.

Hence

y~(x2)U(x2)y~(x2)U(x2)U(h)=k2U(x2)U(h)hx2(U(x2)U(h))y~(x2)𝑑x2,\tilde{y}^{\prime}(x_{2})-\frac{U^{\prime}(x_{2})\tilde{y}(x_{2})}{U(x_{2})-U(-h)}=\frac{k^{2}}{U(x_{2})-U(-h)}\int_{-h}^{x_{2}}(U(x_{2}^{\prime})-U(-h))\tilde{y}(x_{2}^{\prime})dx_{2}^{\prime},

which is clearly of the quadratic order in x2+hx_{2}+h. We obtain the desired asymptotics using the leading order expansions of UU and y~\tilde{y} near x2=hx_{2}=-h. ∎

The next lemma is a property of families of real analytic functions whose roots obey certain properties related to the Semi-circle Theorem (1.8).

Lemma 3.9.

Let a,ρ0,q0>0a,\rho_{0},q_{0}>0 and f(ρ,q,z=z1+iz2)f(\rho,q,z=z_{1}+iz_{2}) be a real analytic function of zΩ={z=zR+izI|z+a/2|<a/2}z\in\Omega=\{z=z_{R}+iz_{I}\in\mathbb{C}\mid|z+a/2|<a/2\} (i.e. additionally it satisfies f(z¯)=f(z)¯f(\bar{z})=\overline{f(z)}) and is also a C1C^{1} function of ρ,q,zR,zI\rho,q,z_{R},z_{I} for |ρ|ρ0|\rho|\leq\rho_{0}, |q|q0|q|\leq q_{0}, and zΩ¯z\in\overline{\Omega}. Moreover assume

(3.16) f(ρ,q,z)0 if z[a,0];f(0,0,z)=0z=0;qf(0,0,0)0;f(\rho,q,z)\neq 0\;\text{ if }\;z\notin[-a,0];\quad f(0,0,z)=0\Leftrightarrow z=0;\quad\partial_{q}f(0,0,0)\neq 0;

then there exists ρ~(0,ρ0]\tilde{\rho}\in(0,\rho_{0}] and a C1C^{1} function ψ(ρ)\psi(\rho) defined for |ρ|ρ~|\rho|\leq\tilde{\rho} such that, for any such that ρ\rho,

f(ρ,ψ(ρ),z)=0z=0,f(\rho,\psi(\rho),z)=0\Leftrightarrow z=0,

which clearly implies

f(ρ,ψ(ρ),a)zf(ρ,ψ(ρ),0)0.f(\rho,\psi(\rho),-a)\partial_{z}f(\rho,\psi(\rho),0)\leq 0.

Note that ff being a real analytic function of zΩz\in\Omega and C1C^{1} up to Ω\partial\Omega yields ff\in\mathbb{R} for z[a,0]z\in[-a,0], hence qf(ρ,q,z),zf(ρ,q,z)\partial_{q}f(\rho,q,z),\partial_{z}f(\rho,q,z)\in\mathbb{R} for any z[a,0]z\in[-a,0]. The first assumption in (3.16) is crucial. Otherwise a counter example is f=qz2+2ρzf=q-z^{2}+2\rho z where clearly it has to be ψ(ρ)=0\psi(\rho)=0, but f(ρ,0,)f(\rho,0,\cdot) has another root z=2ρ(a,0)z=2\rho\in(-a,0) for ρ<0\rho<0.

Proof.

It is easy to see how ψ\psi should be defined. In fact, from q(Ref)(0,0,0)=qf(0,0,0)0\partial_{q}(\text{Re}\,f)(0,0,0)=\partial_{q}f(0,0,0)\neq 0 and the Implicit Function Theorem applied to Ref\text{Re}\,f, there exists δ,q1,ρ1>0\delta,q_{1},\rho_{1}>0 and a C1C^{1} function ϕ(ρ,z)\phi(\rho,z) defined for |ρ|ρ1|\rho|\leq\rho_{1} and z[δ,0]z\in[-\delta,0] such that ϕ(0,0)=0\phi(0,0)=0 and

(3.17) f(ρ,q,z)=Ref(ρ,q,z)=0,|ρ|ρ1,|q|q1,z[δ,0]q=ϕ(ρ,z).f(\rho,q,z)=\text{Re}\,f(\rho,q,z)=0,\;|\rho|\leq\rho_{1},\;|q|\leq q_{1},\;z\in[-\delta,0]\Longleftrightarrow q=\phi(\rho,z).

Let ψ(ρ)=ϕ(ρ,0)\psi(\rho)=\phi(\rho,0). As we assumed f(0,0,z)=0f(0,0,z)=0 iff z=0z=0 and f=0f=0 only if z[δ,0]z\in[-\delta,0], by the compactness of Ω¯{z|z+δ/3|<2δ/3}\overline{\Omega}\setminus\{z\in\mathbb{C}\mid|z+\delta/3|<2\delta/3\}, there exist ρ~(0,ρ1]\tilde{\rho}\in(0,\rho_{1}] and ϵ>0\epsilon>0 such that

(3.18) f(ρ,q,z)=0,|ρ|ρ~,|q|max|ρ|ρ~|ψ(ρ)|+ϵz[δ,0]q=ϕ(ρ,z).f(\rho,q,z)=0,\;|\rho|\leq\tilde{\rho},\;|q|\leq\max_{|\rho^{\prime}|\leq\tilde{\rho}}|\psi(\rho^{\prime})|+\epsilon\implies z\in[-\delta,0]\implies q=\phi(\rho,z).

It remains to prove f(ρ,ψ(ρ),z)0f(\rho,\psi(\rho),z)\neq 0 if z[δ,0)z\in[-\delta,0) and |ρ|ρ~|\rho|\leq\tilde{\rho}, which we proceed by an argument by contradiction.

Assume

(3.19) ρ[ρ~,ρ~],z[δ,0) such that f(ρ,q,z)=0, where q=ψ(ρ).\exists\rho_{*}\in[-\tilde{\rho},\tilde{\rho}],\;z_{*}\in[-\delta,0)\;\text{ such that }\;f(\rho_{*},q_{*},z_{*})=0,\;\text{ where }\;q_{*}=\psi(\rho_{*}).

The definition of ϕ\phi yields ϕ(ρ,z)=q=ϕ(ρ,0)\phi(\rho_{*},z_{*})=q_{*}=\phi(\rho_{*},0). When ρ~\tilde{\rho} is sufficiently small, the function f(ρ,q,)f(\rho_{*},q_{*},\cdot) of zz is not identically zero, which, along with its analyticity, implies that it is not identically zero on [z,0][z_{*},0]. Therefore ϕ(ρ,)\phi(\rho_{*},\cdot) is not identically equal to qq_{*} for z[z,0]z\in[z_{*},0]. Without loss of generality, suppose

z#(z,0),q#=ϕ(ρ,z#)=maxz[z,0]ϕ(ρ,z)>q.z_{\#}\in(z_{*},0),\quad q_{\#}=\phi(\rho_{*},z_{\#})=\max_{z\in[z_{*},0]}\phi(\rho_{*},z)>q_{*}.

Let

Ω1={z|zz/2|<|z|/2}Ω.\Omega_{1}=\{z\in\mathbb{C}\mid|z-z_{*}/2|<|z_{*}|/2\}\subset\Omega.

From (3.17), (3.18), and ϕ(ρ,0)=q=ϕ(ρ,z)\phi(\rho_{*},0)=q_{*}=\phi(\rho_{*},z_{*}), the index Ind(f(ρ,q,),Ω1)\big{(}f(\rho_{*},q,\cdot),\Omega_{1}\big{)} is well defined and takes a constant non-negative integer value for q>qq>q_{*}. Since f(ρ,q#,z#)=0f(\rho_{*},q_{\#},z_{\#})=0 with z#(z,0)Ω1z_{\#}\in(z_{*},0)\subset\Omega_{1} and q#qq_{\#}\geq q_{*}, it holds Ind(f(ρ,q,),Ω1)1\big{(}f(\rho_{*},q,\cdot),\Omega_{1}\big{)}\geq 1 for q>qq>q_{*}. However, f(ρ,q,z)0f(\rho_{*},q,z)\neq 0 for any z(z,0)z\in(z_{*},0) and q>q#q>q_{\#}. It implies Ind(f(ρ,q,),Ω1)=0\big{(}f(\rho_{*},q,\cdot),\Omega_{1}\big{)}=0 for q>q#q>q_{\#}, which is a contradiction. Therefore (3.19) can not be true and we complete the proof of the lemma. ∎

We are ready to prove the main lemma of the non-degeneracy of FF at c=U(h)c=U(-h). In order for this result to be also applicable to the capillary gravity waves to improve results in [17], we include the surface tension in the consideration. Let

(3.20) σ(k,c)=F(k,c)σk2=Y(k,c)(U(0)c)2U(0)(U(0)c)gσk2,σ0,\mathcal{F}_{\sigma}(k,c)=F(k,c)-\sigma k^{2}=Y(k,c)\big{(}U(0)-c\big{)}^{2}-U^{\prime}(0)\big{(}U(0)-c\big{)}-g-\sigma k^{2},\quad\sigma\geq 0,

whose zero points correspond to the eigenvalues of the capillary gravity waves linearized at the shear flow U(x2)U(x_{2}) if σ>0\sigma>0 (see [17]).

Lemma 3.10.

Suppose UC6U\in C^{6}, k0k_{0}\in\mathbb{R}, and c1<U(h)c_{1}<U(-h) satisfy

σ(k0,U(h))=0,σ(k0,c)0,c[c1,U(h)),\mathcal{F}_{\sigma}(k_{0},U(-h))=0,\quad\mathcal{F}_{\sigma}(k_{0},c)\neq 0,\;c\in[c_{1},U(-h)),

then

σ(k0,c1)cσ(k0,U(h))<0.\mathcal{F}_{\sigma}(k_{0},c_{1})\partial_{c}\mathcal{F}_{\sigma}(k_{0},U(-h))<0.
Proof.

According to Lemma 2.6(1), it must hold k00k_{0}\neq 0. Due to the evenness of σ\mathcal{F}_{\sigma} in kk, without loss of generality, we assume k0>0k_{0}>0. Essentially we need to prove cσ(k0,U(h))0\partial_{c}\mathcal{F}_{\sigma}(k_{0},U(-h))\neq 0, which will be done by an argument by contradiction based on Lemma 3.9 and a carefully constructed localized perturbation to UU. We shall first prove the lemma under the assumption

kσ(k0,U(h))0\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))\neq 0

and then remove it at the end of the proof.

Let γC(,[0,1])\gamma\in C^{\infty}(\mathbb{R},[0,1]) be an auxiliary function satisfying

(3.21) γ′′(s)0,s,γ′′(s)=0,|s|1,11γ′′(s)𝑑s=1,γ(1)=γ(1)=0,\gamma^{\prime\prime}(s)\geq 0,\;\forall s\in\mathbb{R},\;\;\gamma^{\prime\prime}(s)=0,\;\forall|s|\geq 1,\;\;\int_{-1}^{1}\gamma^{\prime\prime}(s^{\prime})ds^{\prime}=1,\;\;\gamma(-1)=\gamma^{\prime}(-1)=0,

which implies

(3.22) |γ(s)sχs>1|C,γ[0,1],|\gamma(s)-s\chi_{s>-1}|\leq C,\quad\gamma^{\prime}\in[0,1],

where χ\chi is the characteristic function. For |ρ|1|\rho|\ll 1 and

(3.23) x20(h,0),0<δ<L(x20)/2min{x20,x20+h}/2,x_{20}\in(-h,0),\quad 0<\delta<L(x_{20})/2\triangleq\min\{-x_{20},\,x_{20}+h\}/2,

to be determined later, let

𝕌(ρ,x2)=U(x2)+ρδγ((x2x20)/δ),\mathbb{U}(\rho,x_{2})=U(x_{2})+\rho\delta\gamma\big{(}(x_{2}-x_{20})/\delta\big{)},

which coincides with U(x2)U(x_{2}) in the δ\delta-neighborhood of h-h. Sometimes we skip some of the variables to prevent the notations from being overly complicated. Clearly,

𝕌(ρ,)C6,U+1𝕌U.\mathbb{U}(\rho,\cdot)\in C^{6},\quad U^{\prime}+1\geq\mathbb{U}^{\prime}\geq U^{\prime}.

Let y(ρ,k,c,x2)y_{-}(\rho,k,c,x_{2}) be the solution to the homogeneous Rayleigh equation (2.1) at the shear flow 𝕌(ρ,)\mathbb{U}(\rho,\cdot) with the initial condition (2.6) and 𝕐(ρ,k,c)\mathbb{Y}(\rho,k,c) and 𝔽(ρ,k,c)\mathbb{F}(\rho,k,c) be defined in terms of y(ρ,k,c,x2)y_{-}(\rho,k,c,x_{2}) as in (2.19) and (3.20), respectively, which also depend on x20x_{20} and δ\delta.

To see the regularity of 𝕐\mathbb{Y} and 𝔽\mathbb{F} in ρ\rho as well as cc near 𝕌(ρ,h)=U(h)\mathbb{U}(\rho,-h)=U(-h), let u(ρ,k,c,x2)u(\rho,k,c,x_{2}) be the solution to the homogeneous Rayleigh equation (2.1) with the initial conditions

u(h+L(x20)/2)=1,u(h+L(x20)/2)=Y~(k,c)y(ρ,k,c,h+L(x20)/2)y(ρ,k,c,h+L(x20)/2),u\big{(}-h+L(x_{20})/2\big{)}=1,\quad u^{\prime}\big{(}-h+L(x_{20})/2\big{)}=\tilde{Y}(k,c)\triangleq\tfrac{y_{-}^{\prime}(\rho,k,c,-h+L(x_{20})/2)}{y_{-}(\rho,k,c,-h+L(x_{20})/2)},

which clearly implies

𝕐(ρ,k,c)=u(ρ,k,c,0)u(ρ,k,c,0).\mathbb{Y}(\rho,k,c)=\tfrac{u^{\prime}(\rho,k,c,0)}{u(\rho,k,c,0)}.

Recall that y(ρ,k,c,x2)y_{-}(\rho,k,c,x_{2}) is independent of ρ\rho for x2[h,h+L(x20)/2]x_{2}\in[-h,-h+L(x_{20})/2] due to the definition of 𝕌\mathbb{U}. In particular, y(ρ,k,c,h+L(x20)/2)0y_{-}(\rho,k,c,-h+L(x_{20})/2)\neq 0 as cU(h)c\leq U(-h) (Lemma 2.3(4)), which implies that Y~(k,c)\tilde{Y}(k,c) is well-defined for cc near U(h)U(-h). Through the same proof of Lemma 3.1 and Corollary 3.1.1 (where x2=0x_{2}=0 being replaced by x2=h+L(x20)/2x_{2}=-h+L(x_{20})/2 does not affect the arguments), when restricted to cI0c_{I}\geq 0, kj1cj2Y~(k,c)LkLcRq\partial_{k}^{j_{1}}\partial_{c}^{j_{2}}\tilde{Y}(k,c)\in L_{k}^{\infty}L_{c_{R}}^{q} locally in kk and cRc_{R}, for j1,j20j_{1},j_{2}\geq 0, j22j_{2}\leq 2, j1+j23j_{1}+j_{2}\leq 3, and q[1,)q\in[1,\infty). Apparently u(ρ,k,c,0)u(\rho,k,c,0) is smooth in kk, ρ\rho, Y~\tilde{Y}, and cc near U(h)U(-h). Therefore 𝕐\mathbb{Y} and 𝔽\mathbb{F} are smooth in ρ\rho and satisfy the same regularity in kk and cc. Due to the assumptions on k0k_{0} and kσ(k0,U(h))0\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))\neq 0 and the semi-circle Theorem of the unstable modes of the water wave problems, the hypotheses of Lemma 3.9 are satisfied. Hence there exist ρ0>0\rho_{0}>0 and a C1,αC^{1,\alpha} function k(ρ)k_{*}(\rho) defined for |ρ|ρ0|\rho|\ll\rho_{0} such that

(3.24) 𝔽(ρ,k(ρ),U(h))=0,𝔽(ρ,k0,c1)c𝔽(ρ,k(ρ),U(h))0,|ρ|ρ0.\mathbb{F}(\rho,k_{*}(\rho),U(-h))=0,\quad\mathbb{F}(\rho,k_{0},c_{1})\partial_{c}\mathbb{F}(\rho,k_{*}(\rho),U(-h))\leq 0,\quad\forall|\rho|\leq\rho_{0}.

To prove the lemma by an argument by contradiction, we assume

(3.25) cσ(k0,U(h))=0,\partial_{c}\mathcal{F}_{\sigma}(k_{0},U(-h))=0,

and then prove that there exist x20x_{20} and δ\delta satisfying (3.23) such that

(3.26) ρ(c𝔽(ρ,k(ρ),U(h)))|ρ=00.\partial_{\rho}\big{(}\partial_{c}\mathbb{F}(\rho,k_{*}(\rho),U(-h))\big{)}\big{|}_{\rho=0}\neq 0.

This would immediately lead to a contradiction to (3.24) for some small ρ0\rho\neq 0.

The definition (3.20) and 𝔽(ρ,k(ρ),U(h))=0\mathbb{F}(\rho,k_{*}(\rho),U(-h))=0 yield

(3.27) 𝕐(k(ρ),U(h))=(𝕌(0)U(h))1𝕌(0)+(𝕌(0)U(h))2(g+σk(ρ)2),\begin{split}\mathbb{Y}(k_{*}(\rho),U(-h))=(\mathbb{U}(0)-U(-h))^{-1}\mathbb{U}^{\prime}(0)+(\mathbb{U}(0)-U(-h))^{-2}(g+\sigma k_{*}(\rho)^{2}),\end{split}

where we skipped some ρ\rho arguments. Subsequently we can compute

(3.28) c𝔽(ρ,k(ρ),U(h))=((𝕌(0)c)2c𝕐2(𝕌(0)c)𝕐+𝕌(0))|(k,c)=(k(ρ),U(h))=((𝕌(0)c)2c𝕐𝕌(0)2(g+σk2)𝕌(0)c)|(k,c)=(k(ρ),U(h)).\begin{split}\partial_{c}\mathbb{F}(\rho,k_{*}(\rho),U(-h))=&\big{(}(\mathbb{U}(0)-c)^{2}\partial_{c}\mathbb{Y}-2(\mathbb{U}(0)-c)\mathbb{Y}+\mathbb{U}^{\prime}(0)\big{)}\big{|}_{(k,c)=(k_{*}(\rho),U(-h))}\\ =&\big{(}(\mathbb{U}(0)-c)^{2}\partial_{c}\mathbb{Y}-\mathbb{U}^{\prime}(0)-\tfrac{2(g+\sigma k^{2})}{\mathbb{U}(0)-c}\big{)}\big{|}_{(k,c)=(k_{*}(\rho),U(-h))}.\end{split}

The negation assumption (3.25) also yields

(3.29) cY(k0,U(h))=(U(0)U(h))2U(0)+2(U(0)U(h))3(g+σk02).\partial_{c}Y(k_{0},U(-h))=(U(0)-U(-h))^{-2}U^{\prime}(0)+2(U(0)-U(-h))^{-3}(g+\sigma k_{0}^{2}).

Meanwhile, from the Implicit Function Theorem, the definition of 𝕌\mathbb{U}, (3.21), and (3.23),

(3.30) ρk(0)=ρ𝔽(0,k0,U(h))kσ(k0,U(h))=(kσ(k0,U(h)))1((U(0)U(h))2ρ𝕐(0,k0,U(h))+δγ(x20δ)(2(U(0)U(h))Y(k0,U(h))U(0))(U(0)U(h)))=(kσ(k0,U(h)))1((U(0)U(h))2ρ𝕐(0,k0,U(h))+δγ(x20δ)(U(0)+2(g+σk02)U(0)U(h))(U(0)U(h))).\begin{split}\partial_{\rho}k_{*}(0)=&-\frac{\partial_{\rho}\mathbb{F}(0,k_{0},U(-h))}{\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))}\\ =&-\big{(}\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))\big{)}^{-1}\Big{(}(U(0)-U(-h))^{2}\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h))\\ &+\delta\gamma(-\tfrac{x_{20}}{\delta})\big{(}2(U(0)-U(-h))Y(k_{0},U(-h))-U^{\prime}(0)\big{)}-(U(0)-U(-h))\Big{)}\\ =&-\big{(}\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))\big{)}^{-1}\Big{(}(U(0)-U(-h))^{2}\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h))\\ &+\delta\gamma(-\tfrac{x_{20}}{\delta})\big{(}U^{\prime}(0)+\tfrac{2(g+\sigma k_{0}^{2})}{U(0)-U(-h)}\big{)}-(U(0)-U(-h))\Big{)}.\end{split}

Moreover, from (3.28) one may compute

ρ(c𝔽(ρ,k(ρ),\displaystyle\partial_{\rho}\big{(}\partial_{c}\mathbb{F}(\rho,k_{*}(\rho), U(h)))|ρ=0=14σk0U(0)U(h)ρk(0)\displaystyle U(-h))\big{)}\big{|}_{\rho=0}=-1-\frac{4\sigma k_{0}}{U(0)-U(-h)}\partial_{\rho}k_{*}(0)
+(U(0)U(h))2(cρ𝕐(0,k0,U(h))+ckY(k0,U(h))ρk(0))\displaystyle+(U(0)-U(-h))^{2}\Big{(}\partial_{c\rho}\mathbb{Y}(0,k_{0},U(-h))+\partial_{ck}Y(k_{0},U(-h))\partial_{\rho}k_{*}(0)\Big{)}
+2δγ(x20δ)(g+σk02(U(0)U(h))2+(U(0)U(h))cY(k0,U(h))),\displaystyle+2\delta\gamma(-\tfrac{x_{20}}{\delta})\Big{(}\frac{g+\sigma k_{0}^{2}}{(U(0)-U(-h))^{2}}+(U(0)-U(-h))\partial_{c}Y(k_{0},U(-h))\Big{)},

where ρk(0)\partial_{\rho}k_{*}(0) and cY(k0,U(h))\partial_{c}Y(k_{0},U(-h)) should be substituted by (3.30) and (3.29), respectively. According to (3.22), δγ(x20δ)\delta\gamma(-\tfrac{x_{20}}{\delta}) can be approximated by x20-x_{20}. Along with (3.30) and (3.29), we obtain

(3.31) |ρ(c𝔽(ρ,k(ρ),U(h)))|ρ=0A0+A1x20A2ρ𝕐(0,k0,U(h))(U(0)U(h))2cρ𝕐(0,k0,U(h))|C~δ,\begin{split}\big{|}\partial_{\rho}\big{(}\partial_{c}\mathbb{F}(\rho,k_{*}(\rho),U(-h))\big{)}\big{|}_{\rho=0}-&A_{0}+A_{1}x_{20}-A_{2}\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h))\\ &-(U(0)-U(-h))^{2}\partial_{c\rho}\mathbb{Y}(0,k_{0},U(-h))\big{|}\leq\tilde{C}\delta,\end{split}

where

A0=1+(kσ(k0,U(h)))1(4σk0+(U(0)U(h))3ckY(k0,U(h))),A_{0}=-1+\big{(}\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))\big{)}^{-1}\big{(}-4\sigma k_{0}+(U(0)-U(-h))^{3}\partial_{ck}Y(k_{0},U(-h))\big{)},
A1=\displaystyle A_{1}= (U(0)U(h))2kσ(k0,U(h))(4σk0(U(0)U(h))3+ckY(k0,U(h)))\displaystyle-\frac{(U(0)-U(-h))^{2}}{\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))}\Big{(}-\frac{4\sigma k_{0}}{(U(0)-U(-h))^{3}}+\partial_{ck}Y(k_{0},U(-h))\Big{)}
×(U(0)+2(g+σk02)U(0)U(h))+2(U(0)U(0)U(h)+3(g+σk02)(U(0)U(h))2),\displaystyle\times\Big{(}U^{\prime}(0)+\frac{2(g+\sigma k_{0}^{2})}{U(0)-U(-h)}\Big{)}+2\Big{(}\frac{U^{\prime}(0)}{U(0)-U(-h)}+\frac{3(g+\sigma k_{0}^{2})}{(U(0)-U(-h))^{2}}\Big{)},
A2=(U(0)U(h))4kσ(k0,U(h))(4σk0(U(0)U(h))3+ckY(k0,U(h))),A_{2}=-\frac{(U(0)-U(-h))^{4}}{\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))}\Big{(}-\frac{4\sigma k_{0}}{(U(0)-U(-h))^{3}}+\partial_{ck}Y(k_{0},U(-h))\Big{)},

and C~\tilde{C} is proportional to |A1||A_{1}| depending on UU, gg, σ\sigma, γ\gamma, k0k_{0}, but independent of x20x_{20} and δ\delta satisfying (3.23). In the rest of the proof, we will estimate ρ𝕐(0,k0,U(h))\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h)) and cρ𝕐(0,k0,U(h))\partial_{c\rho}\mathbb{Y}(0,k_{0},U(-h)) carefully to show that there exists x20(h,0)x_{20}\in(-h,0) such that (3.26) holds for 0<δ10<\delta\ll 1 based on (3.31). In the following the generic constant C~\tilde{C} depends on UU, gg, σ\sigma, γ\gamma, and k0k_{0}, but always independent of x20x_{20} and 0<δ10<\delta\ll 1.

\bullet Estimates of ρ𝕐(0,k0,U(h))\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h)). For c(,U(h)]c\in(-\infty,U(-h)], let

y~(ρ,k,c,x2)=y(ρ,k,c,x2)y(ρ,k,c,0)>0,y~0(x2)=y~(0,k0,U(h),x2),x2(h,0],\tilde{y}(\rho,k,c,x_{2})=\tfrac{y_{-}(\rho,k,c,x_{2})}{y_{-}(\rho,k,c,0)}>0,\;\;\tilde{y}_{0}(x_{2})=\tilde{y}(0,k_{0},U(-h),x_{2}),\quad\forall x_{2}\in(-h,0],

which are well-defined and positive due to Lemma 2.3(4) and satisfies (3.13) and those estimates given in Lemma 3.8. Differentiating (3.13) with respect to ρ\rho, we have

ρy~′′+(k2+𝕌′′𝕌c)ρy~=ρ(𝕌′′𝕌c)y~,ρy~(h)=ρy~(0)=0.-\partial_{\rho}\tilde{y}^{\prime\prime}+\big{(}k^{2}+\tfrac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\partial_{\rho}\tilde{y}=-\partial_{\rho}\big{(}\tfrac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y},\quad\partial_{\rho}\tilde{y}(-h)=\partial_{\rho}\tilde{y}(0)=0.

For cU(h)c\leq U(-h), the claim in the proof of Lemma 3.4 applies and yields

(3.32) ρy~(x2)=x20y~(x2)y~(x2)2hx2ρ(𝕌′′(ρ,x2′′)𝕌(ρ,x2′′)c)y~(x2′′)2dx2′′dx2,\partial_{\rho}\tilde{y}(x_{2})=-\int_{x_{2}}^{0}\frac{\tilde{y}(x_{2})}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\partial_{\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}(\rho,x_{2}^{\prime\prime})}{\mathbb{U}(\rho,x_{2}^{\prime\prime})-c}\big{)}\tilde{y}(x_{2}^{\prime\prime})^{2}dx_{2}^{\prime\prime}dx_{2}^{\prime},
ρ𝕐(0,k0,U(h))=ρy~(0,k0,U(h),0)=h0ρ(𝕌′′𝕌U(h))|ρ=0y~02dx2.\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h))=\partial_{\rho}\tilde{y}^{\prime}(0,k_{0},U(-h),0)=\int_{-h}^{0}\partial_{\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-U(-h)}\big{)}\big{|}_{\rho=0}\tilde{y}_{0}^{2}dx_{2}.

One may compute

(3.33) ρ(𝕌′′(ρ,x2)𝕌(ρ,x2)c)|ρ=0=1δγ′′(x2x20δ)U(x2)cδγ(x2x20δ)U′′(x2)(U(x2)c)2.\begin{split}&\partial_{\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}(\rho,x_{2})}{\mathbb{U}(\rho,x_{2})-c}\big{)}\Big{|}_{\rho=0}=\frac{\frac{1}{\delta}\gamma^{\prime\prime}(\frac{x_{2}-x_{20}}{\delta})}{U(x_{2})-c}-\frac{\delta\gamma(\frac{x_{2}-x_{20}}{\delta})U^{\prime\prime}(x_{2})}{(U(x_{2})-c)^{2}}.\end{split}

From (3.21), Lemma 3.8, and

|(y~02UU(h))|2|y~0||y~0UU(h)|+U|y~02(UU(h))2|C~,\big{|}\big{(}\tfrac{\tilde{y}_{0}^{2}}{U-U(-h)}\big{)}^{\prime}\big{|}\leq 2|\tilde{y}_{0}^{\prime}|\big{|}\tfrac{\tilde{y}_{0}}{U-U(-h)}\big{|}+U^{\prime}\big{|}\tfrac{\tilde{y}_{0}^{2}}{(U-U(-h))^{2}}\big{|}\leq\tilde{C},

along with (3.22) we obtain an estimate on ρ𝕐(0,k0,U(h))\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h))

(3.34) |ρ𝕐(0,k0,U(h))y~0(x20)2U(x20)U(h)+x200(x2x20)U′′(x2)y~0(x2)2(U(x2)U(h))2𝑑x2|C~δ+x20δx20+δ1δγ′′(x2x20δ)|y~02UU(h)y~0(x20)2U(x20)U(h)|𝑑x2C~δ,\begin{split}&\Big{|}\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h))-\frac{\tilde{y}_{0}(x_{20})^{2}}{U(x_{20})-U(-h)}+\int_{x_{20}}^{0}\frac{(x_{2}-x_{20})U^{\prime\prime}(x_{2})\tilde{y}_{0}(x_{2})^{2}}{(U(x_{2})-U(-h))^{2}}dx_{2}\Big{|}\\ \leq&\tilde{C}\delta+\int_{x_{20}-\delta}^{x_{20}+\delta}\frac{1}{\delta}\gamma^{\prime\prime}(\frac{x_{2}-x_{20}}{\delta})\Big{|}\frac{\tilde{y}_{0}^{2}}{U-U(-h)}-\frac{\tilde{y}_{0}(x_{20})^{2}}{U(x_{20})-U(-h)}\Big{|}dx_{2}\leq\tilde{C}\delta,\end{split}

where we used the fact that γ′′0\gamma^{\prime\prime}\geq 0 is supported in [1,1][-1,1] with total integral equal to 1.

\bullet Estimates of cρ𝕐(0,k0,U(h))\partial_{c\rho}\mathbb{Y}(0,k_{0},U(-h)). Differentiating (3.13) with respect to ρ\rho and cc yields

cρy~′′+(k2+𝕌′′𝕌c)cρy~=cρ(𝕌′′𝕌c)y~ρ(𝕌′′𝕌c)cy~c(𝕌′′𝕌c)ρy~,-\partial_{c\rho}\tilde{y}^{\prime\prime}+\big{(}k^{2}+\tfrac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\partial_{c\rho}\tilde{y}=-\partial_{c\rho}\big{(}\tfrac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}-\partial_{\rho}\big{(}\tfrac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\partial_{c}\tilde{y}-\partial_{c}\big{(}\tfrac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\partial_{\rho}\tilde{y},

with zero boundary values at x2=h,0x_{2}=-h,0. Again from (3.7), we have

(3.35) cρ𝕐(0,k0,U(h))=cρy~(0,k0,U(h),0)=I1+I2+I3h0cρ(𝕌′′𝕌c)y~2+ρ(𝕌′′𝕌c)y~cy~+c(𝕌′′𝕌c)y~ρy~dx2|(ρ,k,c)=(0,k0,U(h)),\begin{split}&\partial_{c\rho}\mathbb{Y}(0,k_{0},U(-h))=\partial_{c\rho}\tilde{y}^{\prime}(0,k_{0},U(-h),0)=I_{1}+I_{2}+I_{3}\\ \triangleq&\int_{-h}^{0}\partial_{c\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}^{2}+\partial_{\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}\partial_{c}\tilde{y}+\partial_{c}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}\partial_{\rho}\tilde{y}dx_{2}\Big{|}_{(\rho,k,c)=(0,k_{0},U(-h))},\end{split}

where cy~\partial_{c}\tilde{y} and ρy~\partial_{\rho}\tilde{y} were given in (3.15) and (3.32), respectively. Using (3.15) and through an integration by parts, it follows

I2=\displaystyle I_{2}= h0ρ(𝕌′′𝕌c)y~cy~dx2|(ρ,k,c)=(0,k0,U(h))\displaystyle\int_{-h}^{0}\partial_{\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}\partial_{c}\tilde{y}dx_{2}\Big{|}_{(\rho,k,c)=(0,k_{0},U(-h))}
=\displaystyle= h0(ρ(𝕌′′𝕌c)y~2)(x2)x201y~(x2)2hx2(c(𝕌′′𝕌c)y~2)(x2′′)𝑑x2′′𝑑x2𝑑x2\displaystyle-\int_{-h}^{0}\Big{(}\partial_{\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}^{2}\Big{)}(x_{2})\int_{x_{2}}^{0}\frac{1}{\tilde{y}(x_{2}^{\prime})^{2}}\int_{-h}^{x_{2}^{\prime}}\Big{(}\partial_{c}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}^{2}\Big{)}(x_{2}^{\prime\prime})dx_{2}^{\prime\prime}dx_{2}^{\prime}dx_{2}
=\displaystyle= h01y~(x2)2hx2(ρ(𝕌′′𝕌c)y~2)(x2)𝑑x2hx2(c(𝕌′′𝕌c)y~2)(x2)𝑑x2𝑑x2,\displaystyle-\int_{-h}^{0}\frac{1}{\tilde{y}(x_{2})^{2}}\int_{-h}^{x_{2}}\Big{(}\partial_{\rho}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}^{2}\Big{)}(x_{2}^{\prime})dx_{2}^{\prime}\int_{-h}^{x_{2}}\Big{(}\partial_{c}\big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\big{)}\tilde{y}^{2}\Big{)}(x_{2}^{\prime})dx_{2}^{\prime}dx_{2},

evaluated at (ρ,k,c)=(0,k0,U(h))(\rho,k,c)=(0,k_{0},U(-h)). Through another integration by parts in a similar fashion applied to I3I_{3}, we obtain I2=I3I_{2}=I_{3}.

Like (3.34) for ρ𝕐(0,k0,U(h))\partial_{\rho}\mathbb{Y}(0,k_{0},U(-h)), we will also identify the leading terms in I1,2,3I_{1,2,3} of cρ𝕐\partial_{c\rho}\mathbb{Y}. From Lemma 3.8, we have

|((y~cy~)(0,k0,U(h),x2)U(x2)U(h))|\displaystyle\big{|}\big{(}\tfrac{(\tilde{y}\partial_{c}\tilde{y})(0,k_{0},U(-h),x_{2})}{U(x_{2})-U(-h)}\big{)}^{\prime}\big{|}\leq C~(1+|log(x2+h)|),\displaystyle\tilde{C}\big{(}1+|\log(x_{2}+h)|\big{)},

which implies the CαC^{\alpha} continuity of y~cy~Uc|(ρ,k)=(0,k0)\frac{\tilde{y}\partial_{c}\tilde{y}}{U-c}\big{|}_{(\rho,k)=(0,k_{0})} in x2x_{2} for any α[0,1)\alpha\in[0,1). Since (3.23) implies x20δ+h>δx_{20}-\delta+h>\delta, with the above inequality, (3.22), and (3.33), we are ready to obtain the leading order term of I2=I3I_{2}=I_{3}

(3.36) |I2(y~cy~)(0,k0,U(h),x20)U(x20)U(h)+x200(x2x20)(y~cy~)(0,k0,U(h),x2)U′′(U(x2)U(h))2𝑑x2|(x20δx20+δ1δγ′′(x2x20δ)|y~cy~Uc(y~cy~)(x20)U(x20)c|dx2+C~δδh0|cy~|y~(Uc)2dx2)|(ρ,k,c)=(0,k0,U(h))C~x20δx20+δδα1γ′′(x2x20δ)𝑑x2+C~δδh0(1+|log(x2+h)|)𝑑x2C~δα.\begin{split}&\Big{|}I_{2}-\frac{(\tilde{y}\partial_{c}\tilde{y})(0,k_{0},U(-h),x_{20})}{U(x_{20})-U(-h)}+\int_{x_{20}}^{0}\frac{(x_{2}-x_{20})(\tilde{y}\partial_{c}\tilde{y})(0,k_{0},U(-h),x_{2})U^{\prime\prime}}{(U(x_{2})-U(-h))^{2}}dx_{2}\Big{|}\\ \leq&\Big{(}\int_{x_{20}-\delta}^{x_{20}+\delta}\frac{1}{\delta}\gamma^{\prime\prime}(\frac{x_{2}-x_{20}}{\delta})\Big{|}\frac{\tilde{y}\partial_{c}\tilde{y}}{U-c}-\frac{(\tilde{y}\partial_{c}\tilde{y})(x_{20})}{U(x_{20})-c}\Big{|}dx_{2}\\ &+\tilde{C}\delta\int_{\delta-h}^{0}\frac{|\partial_{c}\tilde{y}|\tilde{y}}{(U-c)^{2}}dx_{2}\Big{)}\Big{|}_{(\rho,k,c)=(0,k_{0},U(-h))}\\ \leq&\tilde{C}\int_{x_{20}-\delta}^{x_{20}+\delta}\delta^{\alpha-1}\gamma^{\prime\prime}(\frac{x_{2}-x_{20}}{\delta})dx_{2}+\tilde{C}\delta\int_{\delta-h}^{0}\big{(}1+|\log(x_{2}+h)|\big{)}dx_{2}\leq\tilde{C}\delta^{\alpha}.\end{split}

The last term I1I_{1} is handled similarly starting with

cρ(𝕌′′𝕌c)(x2)|ρ=0=1δγ′′(x2x20δ)(U(x2)c)22δγ(x2x20δ)U′′(x2)(U(x2)c)3.\partial_{c\rho}\Big{(}\frac{\mathbb{U}^{\prime\prime}}{\mathbb{U}-c}\Big{)}(x_{2})\Big{|}_{\rho=0}=\frac{\frac{1}{\delta}\gamma^{\prime\prime}(\frac{x_{2}-x_{20}}{\delta})}{(U(x_{2})-c)^{2}}-\frac{2\delta\gamma(\frac{x_{2}-x_{20}}{\delta})U^{\prime\prime}(x_{2})}{(U(x_{2})-c)^{3}}.

Moreover we can estimate using Lemma 3.8

|(y~02(UU(h))2)|=\displaystyle\Big{|}\Big{(}\frac{\tilde{y}_{0}^{2}}{(U-U(-h))^{2}}\Big{)}^{\prime}\Big{|}= 2y~0(UU(h))2|y~0Uy~0UU(h)|C~|x2+h|.\displaystyle\frac{2\tilde{y}_{0}}{(U-U(-h))^{2}}\Big{|}\tilde{y}_{0}^{\prime}-\frac{U^{\prime}\tilde{y}_{0}}{U-U(-h)}\Big{|}\leq\tilde{C}|x_{2}+h|.

Again, using (3.22) we obtain

|I1y~0(x20)2(U(x20)U(h))2+x2002(x2x20)U′′(x2)y~0(x2)2(U(x2)U(h))3𝑑x2|\displaystyle\Big{|}I_{1}-\frac{\tilde{y}_{0}(x_{20})^{2}}{(U(x_{20})-U(-h))^{2}}+\int_{x_{20}}^{0}\frac{2(x_{2}-x_{20})U^{\prime\prime}(x_{2})\tilde{y}_{0}(x_{2})^{2}}{(U(x_{2})-U(-h))^{3}}dx_{2}\Big{|}
\displaystyle\leq x20δx20+δ1δγ′′(x2x20δ)|y~02(UU(h))2y~0(x20)2(U(x20)U(h))2|𝑑x2+δh0C~δy~02(UU(h))3𝑑x2\displaystyle\int_{x_{20}-\delta}^{x_{20}+\delta}\frac{1}{\delta}\gamma^{\prime\prime}(\frac{x_{2}-x_{20}}{\delta})\Big{|}\frac{\tilde{y}_{0}^{2}}{(U-U(-h))^{2}}-\frac{\tilde{y}_{0}(x_{20})^{2}}{(U(x_{20})-U(-h))^{2}}\Big{|}dx_{2}+\int_{\delta-h}^{0}\frac{\tilde{C}\delta\tilde{y}_{0}^{2}}{(U-U(-h))^{3}}dx_{2}
\displaystyle\leq C~δ+C~δh0δx2+h𝑑x2C~δα,α[0,1).\displaystyle\tilde{C}\delta+\tilde{C}\int_{\delta-h}^{0}\frac{\delta}{x_{2}+h}dx_{2}\leq\tilde{C}\delta^{\alpha},\quad\;\forall\alpha\in[0,1).

Therefore (3.35) implies

(3.37) |cρ𝕐(0,k0,U(h))y~0(x20)2(U(x20)U(h))22y~0(x20)y~1(x20)U(x20)U(h)+x2002(x2x20)U′′(x2)(U(x2)U(h))2(y~0(x2)y~1(x2)+y~0(x2)2U(x2)U(h))dx2|C~δ12,\begin{split}\Big{|}\partial_{c\rho}\mathbb{Y}(0,&k_{0},U(-h))-\frac{\tilde{y}_{0}(x_{20})^{2}}{(U(x_{20})-U(-h))^{2}}-\frac{2\tilde{y}_{0}(x_{20})\tilde{y}_{1}(x_{20})}{U(x_{20})-U(-h)}\\ &+\int_{x_{20}}^{0}\frac{2(x_{2}-x_{20})U^{\prime\prime}(x_{2})}{(U(x_{2})-U(-h))^{2}}\Big{(}\tilde{y}_{0}(x_{2})\tilde{y}_{1}(x_{2})+\frac{\tilde{y}_{0}(x_{2})^{2}}{U(x_{2})-U(-h)}\Big{)}dx_{2}\Big{|}\leq\tilde{C}\delta^{\frac{1}{2}},\end{split}

where

y~1(x2)=cy~(0,k0,U(h),x2).\tilde{y}_{1}(x_{2})=\partial_{c}\tilde{y}(0,k_{0},U(-h),x_{2}).

From (3.31), (3.34), and (3.37) we obtain

(3.38) |ρ(c𝔽(ρ,k(ρ),U(h)))|ρ=0I(x20)|C~δ12|\partial_{\rho}\big{(}\partial_{c}\mathbb{F}(\rho,k_{*}(\rho),U(-h))\big{)}\big{|}_{\rho=0}-I(x_{20})|\leq\tilde{C}\delta^{\frac{1}{2}}

where

(3.39) I(x20)=A0A1x20+A2I4(x20)+(U(0)U(h))2I5(x20),I4(x20)=y~0(x20)2U(x20)U(h)x200(x2x20)U′′(x2)y~0(x2)2(U(x2)U(h))2𝑑x2I5(x20)=y~0(x20)2(U(x20)U(h))2+2y~0(x20)y~1(x20)U(x20)U(h),x2002(x2x20)U′′(x2)(U(x2)U(h))2(y~0(x2)y~1(x2)+y~0(x2)2U(x2)U(h))𝑑x2,\begin{split}I(x_{20})=&A_{0}-A_{1}x_{20}+A_{2}I_{4}(x_{20})+(U(0)-U(-h))^{2}I_{5}(x_{20}),\\ I_{4}(x_{20})=&\frac{\tilde{y}_{0}(x_{20})^{2}}{U(x_{20})-U(-h)}-\int_{x_{20}}^{0}\frac{(x_{2}-x_{20})U^{\prime\prime}(x_{2})\tilde{y}_{0}(x_{2})^{2}}{(U(x_{2})-U(-h))^{2}}dx_{2}\\ I_{5}(x_{20})=&\frac{\tilde{y}_{0}(x_{20})^{2}}{(U(x_{20})-U(-h))^{2}}+\frac{2\tilde{y}_{0}(x_{20})\tilde{y}_{1}(x_{20})}{U(x_{20})-U(-h)},\\ &-\int_{x_{20}}^{0}\frac{2(x_{2}-x_{20})U^{\prime\prime}(x_{2})}{(U(x_{2})-U(-h))^{2}}\Big{(}\tilde{y}_{0}(x_{2})\tilde{y}_{1}(x_{2})+\frac{\tilde{y}_{0}(x_{2})^{2}}{U(x_{2})-U(-h)}\Big{)}dx_{2},\\ \end{split}

and A0,A1A_{0},A_{1}, and A2A_{2} are defined right bellow (3.31). Clearly for (3.26) to hold for some x20(h,0)x_{20}\in(-h,0) and 0<δ10<\delta\ll 1, it suffices to show I(x20)I(x_{20}), which is independent of δ\delta, is not identically zero. This will be achieved by computing I′′(x20)I^{\prime\prime}(x_{20}).

Direct calculations and using the Rayleigh equation (3.13) yield

I4′′=\displaystyle I_{4}^{\prime\prime}= 2(y~0)2+2y~0y~0′′UU(h)4Uy~0y~0+U′′y~02(UU(h))2+2(U)2y~02(UU(h))3U′′y~02(UU(h))2\displaystyle\frac{2(\tilde{y}_{0}^{\prime})^{2}+2\tilde{y}_{0}\tilde{y}_{0}^{\prime\prime}}{U-U(-h)}-\frac{4U^{\prime}\tilde{y}_{0}\tilde{y}_{0}^{\prime}+U^{\prime\prime}\tilde{y}_{0}^{2}}{(U-U(-h))^{2}}+\frac{2(U^{\prime})^{2}\tilde{y}_{0}^{2}}{(U-U(-h))^{3}}-\frac{U^{\prime\prime}\tilde{y}_{0}^{2}}{(U-U(-h))^{2}}
=\displaystyle= 2(y~0)2+2k02y~02UU(h)4Uy~0y~0(UU(h))2+2(U)2y~02(UU(h))3\displaystyle\frac{2(\tilde{y}_{0}^{\prime})^{2}+2k_{0}^{2}\tilde{y}_{0}^{2}}{U-U(-h)}-\frac{4U^{\prime}\tilde{y}_{0}\tilde{y}_{0}^{\prime}}{(U-U(-h))^{2}}+\frac{2(U^{\prime})^{2}\tilde{y}_{0}^{2}}{(U-U(-h))^{3}}
=\displaystyle= 2k02y~02UU(h)+2UU(h)(y~0Uy~0UU(h))2.\displaystyle\frac{2k_{0}^{2}\tilde{y}_{0}^{2}}{U-U(-h)}+\frac{2}{U-U(-h)}\Big{(}\tilde{y}_{0}^{\prime}-\frac{U^{\prime}\tilde{y}_{0}}{U-U(-h)}\Big{)}^{2}.

Similar direct calculations lead to

I5′′=\displaystyle I_{5}^{\prime\prime}= 2(y~0)2+2y~0y~0′′(UU(h))28Uy~0y~0+2U′′y~02(UU(h))3+6(U)2y~02(UU(h))4\displaystyle\frac{2(\tilde{y}_{0}^{\prime})^{2}+2\tilde{y}_{0}\tilde{y}_{0}^{\prime\prime}}{(U-U(-h))^{2}}-\frac{8U^{\prime}\tilde{y}_{0}\tilde{y}_{0}^{\prime}+2U^{\prime\prime}\tilde{y}_{0}^{2}}{(U-U(-h))^{3}}+\frac{6(U^{\prime})^{2}\tilde{y}_{0}^{2}}{(U-U(-h))^{4}}
+4y~0y~1+2y~0′′y~1+2y~0y~1′′UU(h)4U(y~0y~1+y~0y~1)+2U′′y~0y~1(UU(h))2+4(U)2y~0y~1(UU(h))3\displaystyle+\frac{4\tilde{y}_{0}^{\prime}\tilde{y}_{1}^{\prime}+2\tilde{y}_{0}^{\prime\prime}\tilde{y}_{1}+2\tilde{y}_{0}\tilde{y}_{1}^{\prime\prime}}{U-U(-h)}-\frac{4U^{\prime}(\tilde{y}_{0}^{\prime}\tilde{y}_{1}+\tilde{y}_{0}\tilde{y}_{1}^{\prime})+2U^{\prime\prime}\tilde{y}_{0}\tilde{y}_{1}}{(U-U(-h))^{2}}+\frac{4(U^{\prime})^{2}\tilde{y}_{0}\tilde{y}_{1}}{(U-U(-h))^{3}}
2U′′y~0y~1(UU(h))22U′′y~02(UU(h))3.\displaystyle-\frac{2U^{\prime\prime}\tilde{y}_{0}\tilde{y}_{1}}{(U-U(-h))^{2}}-\frac{2U^{\prime\prime}\tilde{y}_{0}^{2}}{(U-U(-h))^{3}}.

Using the Rayleigh equations (3.13) and (3.14) it follows

I5′′=\displaystyle I_{5}^{\prime\prime}= 2(y~0)2+2k02y~02(UU(h))28Uy~0y~0(UU(h))3+6(U)2y~02(UU(h))4\displaystyle\frac{2(\tilde{y}_{0}^{\prime})^{2}+2k_{0}^{2}\tilde{y}_{0}^{2}}{(U-U(-h))^{2}}-\frac{8U^{\prime}\tilde{y}_{0}\tilde{y}_{0}^{\prime}}{(U-U(-h))^{3}}+\frac{6(U^{\prime})^{2}\tilde{y}_{0}^{2}}{(U-U(-h))^{4}}
+4y~0y~1+4k02y~0y~1UU(h)4U(y~0y~1+y~0y~1)(UU(h))2+4(U)2y~0y~1(UU(h))3\displaystyle+\frac{4\tilde{y}_{0}^{\prime}\tilde{y}_{1}^{\prime}+4k_{0}^{2}\tilde{y}_{0}\tilde{y}_{1}}{U-U(-h)}-\frac{4U^{\prime}(\tilde{y}_{0}^{\prime}\tilde{y}_{1}+\tilde{y}_{0}\tilde{y}_{1}^{\prime})}{(U-U(-h))^{2}}+\frac{4(U^{\prime})^{2}\tilde{y}_{0}\tilde{y}_{1}}{(U-U(-h))^{3}}

Reorganizing the terms we obtain

I5′′=\displaystyle I_{5}^{\prime\prime}= 2k02y~02(UU(h))2+2(UU(h))2(y~0Uy~0UU(h))(y~03Uy~0UU(h))\displaystyle\frac{2k_{0}^{2}\tilde{y}_{0}^{2}}{(U-U(-h))^{2}}+\frac{2}{(U-U(-h))^{2}}\Big{(}\tilde{y}_{0}^{\prime}-\frac{U^{\prime}\tilde{y}_{0}}{U-U(-h)}\Big{)}\Big{(}\tilde{y}_{0}^{\prime}-\frac{3U^{\prime}\tilde{y}_{0}}{U-U(-h)}\Big{)}
+4k02y~0y~1UU(h)+4UU(h)(y~0Uy~0UU(h))(y~1Uy~1UU(h)).\displaystyle+\frac{4k_{0}^{2}\tilde{y}_{0}\tilde{y}_{1}}{U-U(-h)}+\frac{4}{U-U(-h)}\Big{(}\tilde{y}_{0}^{\prime}-\frac{U^{\prime}\tilde{y}_{0}}{U-U(-h)}\Big{)}\Big{(}\tilde{y}_{1}^{\prime}-\frac{U^{\prime}\tilde{y}_{1}}{U-U(-h)}\Big{)}.

As A0,A1,A2A_{0},A_{1},A_{2} are independent of x20x_{20}, Lemma 3.8 and the above computations imply

limx20(h)+I′′(x20)=2k02(U(0)U(h))2y~0(h)2/(3U(h)2)>0,\lim_{x_{20}\to(-h)+}I^{\prime\prime}(x_{20})=2k_{0}^{2}(U(0)-U(-h))^{2}\tilde{y}_{0}^{\prime}(-h)^{2}/(3U^{\prime}(-h)^{2})>0,

and thus I(x20)I(x_{20}) is not a constant of x20x_{20}.

Therefore, according to (3.38), (3.26) holds if x20x_{20} is close to h-h and 0<δ10<\delta\ll 1. This contradicts (3.24) and thus (3.25) can not occur. This prove the lemma under the assumption kσ(k0,U(h))0\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))\neq 0.

Finally, we shall complete the proof of the lemma in the case of kσ(k0,U(h))=0\partial_{k}\mathcal{F}_{\sigma}(k_{0},U(-h))=0. From Lemma 2.6, this can happen only if σ>0\sigma>0, namely, in the case of the linearized capillary gravity waves. Our strategy is to consider the problem with modified parameters

g~(ϵ)=g+ϵk02,σ~(ϵ)=σϵ, where  0<ϵ1.\tilde{g}(\epsilon)=g+\epsilon k_{0}^{2},\quad\tilde{\sigma}(\epsilon)=\sigma-\epsilon,\ \text{ where }\ 0<\epsilon\ll 1.

The corresponding function ~(ϵ,k,c)\tilde{\mathcal{F}}(\epsilon,k,c) associated to the eigenvalue problem becomes

~(ϵ,k,c)=(U(0)c)2Y(k,c)U(0)(U(0)c)g~σ~k2=σ(k,c)+ϵ(k2k02),\tilde{\mathcal{F}}(\epsilon,k,c)=(U(0)-c)^{2}Y(k,c)-U^{\prime}(0)(U(0)-c)-\tilde{g}-\tilde{\sigma}k^{2}=\mathcal{F}_{\sigma}(k,c)+\epsilon(k^{2}-k_{0}^{2}),

which satisfies

~(ϵ,k0,c)=σ(k0,c),k~(ϵ,k0,U(h))=2ϵk0.\tilde{\mathcal{F}}(\epsilon,k_{0},c)=\mathcal{F}_{\sigma}(k_{0},c),\quad\partial_{k}\tilde{\mathcal{F}}(\epsilon,k_{0},U(-h))=2\epsilon k_{0}.

Therefore, for any ϵσ\epsilon\leq\sigma, ~(ϵ,,)\tilde{\mathcal{F}}(\epsilon,\cdot,\cdot) satisfies the assumption on σ\mathcal{F}_{\sigma} along with k~(ϵ,k0,U(h))0\partial_{k}\tilde{\mathcal{F}}(\epsilon,k_{0},U(-h))\neq 0. Hence the above proof implies cσ(k0,U(h))=c~(ϵ,k0,U(h))0\partial_{c}\mathcal{F}_{\sigma}(k_{0},U(-h))=\partial_{c}\tilde{\mathcal{F}}(\epsilon,k_{0},U(-h))\neq 0. It completes the proof of the lemma. ∎

3.5. Bifurcation analysis

With the technical preparations of the previous subsections, we shall consider the bifurcation of the unstable eigenvalues from limiting neutral modes. In the following lemma we incorporate the bifurcation analysis of cc near U(h)U(-h) and inflection values of UU, which often leads to linear instability. The lemma is stated for σ(k,c)\mathcal{F}_{\sigma}(k,c) defined in (3.20) with σ0\sigma\geq 0 so that it also applies to linearized capillary gravity waves.

Lemma 3.11.

Suppose UCl0U\in C^{l_{0}} and (k0,x20)(×[h,0))(k_{0},x_{20})\in\big{(}\mathbb{R}\times[-h,0)\big{)} satisfy

σ(k0,c0)=0, where c0=U(x20), and cRσ(k0,c0)=cRF(k0,c0)0,\mathcal{F}_{\sigma}(k_{0},c_{0})=0,\;\text{ where }\;c_{0}=U(x_{20}),\;\;\text{ and }\;\;\partial_{c_{R}}\mathcal{F}_{\sigma}(k_{0},c_{0})=\partial_{c_{R}}F(k_{0},c_{0})\neq 0,

then there exist ϵ>0\epsilon>0, 0<ρU(0)c00<\rho\in U(0)-c_{0}, and 𝒞C1,α([k0ϵ,k0+ϵ],)\mathcal{C}\in C^{1,\alpha}\big{(}[k_{0}-\epsilon,k_{0}+\epsilon],\mathbb{C}\big{)} for any α[0,1)\alpha\in[0,1) if x20=hx_{20}=-h and l05l_{0}\geq 5, or 𝒞Cl02\mathcal{C}\in C^{l_{0}-2} if x20(h,0)x_{20}\in(-h,0) and l04l_{0}\geq 4, such that 𝒞(k0)=c0\mathcal{C}(k_{0})=c_{0},

𝒞(k0)=kσ(k0,c0)/cRF(k0,c0)𝒞I(k0)=kσcRFI/|cRF|2|(k,c)=(k0,c0),\mathcal{C}^{\prime}(k_{0})=-\partial_{k}\mathcal{F}_{\sigma}(k_{0},c_{0})/\partial_{c_{R}}F(k_{0},c_{0})\implies\mathcal{C}_{I}^{\prime}(k_{0})=\partial_{k}\mathcal{F}_{\sigma}\partial_{c_{R}}F_{I}/|\partial_{c_{R}}F|^{2}\big{|}_{(k,c)=(k_{0},c_{0})},

and

(3.40) σ(k,c)=0,|kk0|ϵ,|cRc0|,cI[0,ρ], iff c=𝒞(k),𝒞I(k)0.\mathcal{F}_{\sigma}(k,c)=0,\;\;|k-k_{0}|\leq\epsilon,\;|c_{R}-c_{0}|,c_{I}\in[0,\rho],\;\text{ iff }\;c=\mathcal{C}(k),\;\mathcal{C}_{I}(k)\geq 0.

Moreover, the following properties hold for k[k0ϵ,k0+ϵ]k\in[k_{0}-\epsilon,k_{0}+\epsilon] and some C~>0\tilde{C}>0 determined by α\alpha, k0k_{0} and UU.

  1. (1)

    Suppose cRReσ(k0,c0)=cRFR(k0,c0)0\partial_{c_{R}}\text{Re}\,\mathcal{F}_{\sigma}(k_{0},c_{0})=\partial_{c_{R}}F_{R}(k_{0},c_{0})\neq 0, then

    |𝒞I(k)+(U(0)c0)2YI(k,𝒞R(k))/cRFR(k0,c0)|C~|kk0|α|YI(k,𝒞R(k))|.\big{|}\mathcal{C}_{I}(k)+\big{(}U(0)-c_{0}\big{)}^{2}Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}/\partial_{c_{R}}F_{R}(k_{0},c_{0})\big{|}\leq\tilde{C}|k-k_{0}|^{\alpha}\big{|}Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}\big{|}.
  2. (2)

    Suppose kσ(k0,c0)=0\partial_{k}\mathcal{F}_{\sigma}(k_{0},c_{0})=0, then either i.) σ>0\sigma>0 or ii.) σ=0\sigma=0, k0=0k_{0}=0, and x20(h,0)x_{20}\in(-h,0). Moreover the following hold.

    1. (a)

      If k00k_{0}\neq 0, then

      |𝒞(k)+kσ(k,c0)|cRF(k0,c0)|2cRF(k0,c0)¯|C~|kk0|α(1+α).\big{|}\mathcal{C}^{\prime}(k)+\partial_{k}\mathcal{F}_{\sigma}(k,c_{0})|\partial_{c_{R}}F(k_{0},c_{0})|^{-2}\overline{\partial_{c_{R}}F(k_{0},c_{0})}\big{|}\leq\tilde{C}|k-k_{0}|^{\alpha(1+\alpha)}.
    2. (b)

      If k0=0k_{0}=0, then

      |𝒞(k)/k+2Kσ(k,c0)|cRF(0,c0)|2cRF(0,c0)¯|C~|K|α(1+β),\big{|}\mathcal{C}^{\prime}(k)/k+2\partial_{K}\mathcal{F}_{\sigma}(k,c_{0})|\partial_{c_{R}}F(0,c_{0})|^{-2}\overline{\partial_{c_{R}}F(0,c_{0})}\big{|}\leq\tilde{C}|K|^{\alpha(1+\beta)},

      where K=k2K=k^{2}, β=0\beta=0 if Kσ(0,c0)0\partial_{K}\mathcal{F}_{\sigma}(0,c_{0})\neq 0, and any β(0,1)\beta\in(0,1) if Kσ(0,c0)=0\partial_{K}\mathcal{F}_{\sigma}(0,c_{0})=0.

According to (3.40), clearly 𝒞(k)\mathcal{C}(k) is relevant if and only if 𝒞I(k)0\mathcal{C}_{I}(k)\geq 0. The formula for 𝒞(k0)\mathcal{C}^{\prime}(k_{0}) indicates the behavior of 𝒞I(k)\mathcal{C}_{I}(k) for |kk0|1|k-k_{0}|\ll 1 if kσ(k0,c0)cRFI(k0,c0)0\partial_{k}\mathcal{F}_{\sigma}(k_{0},c_{0})\partial_{c_{R}}F_{I}(k_{0},c_{0})\neq 0. The above statements (1) and (2) in combination are useful to provide such information in the degenerate cases including when c0=U(h)c_{0}=U(-h).

Remark 3.3.

a.) Due to Lemma 2.6(1), F(0,U(h))=F(k,U(0))=gF(0,U(-h))=F(k,U(0))=-g and thus x20=0x_{20}=0 and (k0,x20)=(0,h)(k_{0},x_{20})=(0,-h) are actually excluded.
b.) Assume cRFR(k0,c0)0\partial_{c_{R}}F_{R}(k_{0},c_{0})\neq 0, which in particular holds if x20=hx_{20}=-h due to Lemma 3.10. Statement (1) yields that 𝒞I(k)\mathcal{C}_{I}(k) has the opposite sign as cRFR(k0,c0)YI(k,𝒞R(k))\partial_{c_{R}}F_{R}(k_{0},c_{0})Y_{I}(k,\mathcal{C}_{R}(k)). Therefore, Lemma 2.4(5) implies that 𝒞I(k)=0\mathcal{C}_{I}(k)=0 if 𝒞R(k)U((h,0))\mathcal{C}_{R}(k)\notin U\big{(}(-h,0)\big{)} or U′′(U1(𝒞R(k)))=0U^{\prime\prime}\big{(}U^{-1}(\mathcal{C}_{R}(k))\big{)}=0. More importantly, (3.40) implies that whether σ(k,)\mathcal{F}_{\sigma}(k,\cdot) has zero points near c0c_{0} for 0<|kk0|10<|k-k_{0}|\ll 1 is determined by the sign of cRFR(k0,c0)YI(k,𝒞R(k))\partial_{c_{R}}F_{R}(k_{0},c_{0})Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}. A sufficient test for the latter is obviously the signs of U′′(x20)U^{\prime\prime}(x_{20}) and U′′′(x20)U^{\prime\prime\prime}(x_{20}).
c.) From Lemma 2.5(4), it holds that either x20=hx_{20}=-h, where cRImσ(k0,c0)=cRFI(k0,c0)=0\partial_{c_{R}}\text{Im}\,\mathcal{F}_{\sigma}(k_{0},c_{0})=\partial_{c_{R}}F_{I}(k_{0},c_{0})=0, or x20(h,0)x_{20}\in(-h,0) with U′′(x20)=0U^{\prime\prime}(x_{20})=0. In the latter case, U′′′(x20)0U^{\prime\prime\prime}(x_{20})\neq 0 is equivalent to cRFI(k0,c0)0\partial_{c_{R}}F_{I}(k_{0},c_{0})\neq 0 which is sufficient for cRσ(k0,c0)0\partial_{c_{R}}\mathcal{F}_{\sigma}(k_{0},c_{0})\neq 0.
d.) According to Lemmas 2.6(2) and 3.6 and (3.10), if a.) x20=hx_{20}=-h, or b.) x20(h,0)x_{20}\in(-h,0) and k0>0k_{0}>0 is the greatest solution to σ(k0,c0)=0\mathcal{F}_{\sigma}(k_{0},c_{0})=0, then K2F(k0,c0)<0\partial_{K}^{2}F(k_{0},c_{0})<0. Hence, when kσ(k0,c0)=0\partial_{k}\mathcal{F}_{\sigma}(k_{0},c_{0})=0, we have k2σ(k0,c0)<0\partial_{k}^{2}\mathcal{F}_{\sigma}(k_{0},c_{0})<0 and

(k0k)kσ(k,c0)C~1|kk0|2.(k_{0}-k)\partial_{k}\mathcal{F}_{\sigma}(k,c_{0})\geq\tilde{C}^{-1}|k-k_{0}|^{2}.

Hence in the above statement (2a), kσ(k,c0)/cRF(k0,c0)-\partial_{k}\mathcal{F}_{\sigma}(k,c_{0})/\partial_{c_{R}}F(k_{0},c_{0}) gives the leading order term of 𝒞I(k)\mathcal{C}_{I}(k) for kk near k0k_{0}.

Proof.

Since x200x_{20}\neq 0, when restricted to the upper half plane cI0c_{I}\geq 0, according to Lemma 2.5(1)(2) and Corollary 3.1.1, σ\mathcal{F}_{\sigma} is C1,αC^{1,\alpha} in kk and cc near (k0,c0)(k_{0},c_{0}) (actually Cl02C^{l_{0}-2} if x20(h,0)x_{20}\in(-h,0) due to Lemma 2.5(1)). Since FI=ImσF_{I}=\text{Im}\,\mathcal{F}_{\sigma} is not continuous at cU((h,0))c\in U\big{(}(-h,0)\big{)}\subset\mathbb{C} in general, let F~(k,c)=F~R+iF~I\tilde{F}(k,c)=\tilde{F}_{R}+i\tilde{F}_{I}\in\mathbb{C} be a C1,αC^{1,\alpha} extension of σ\mathcal{F}_{\sigma} into a neighborhood of (k0,c0)×(k_{0},c_{0})\in\mathbb{R}\times\mathbb{C} which coincides with σ\mathcal{F}_{\sigma} for cI0c_{I}\geq 0 (or Cl02C^{l_{0}-2} extension if x20(h,0)x_{20}\in(-h,0)). The 2×22\times 2 real Jacobian matrix of DcF~D_{c}\tilde{F} satisfies

DcF~(k0,c0)=(cRF~RcIF~RcRF~IcIF~I)|(k0,c0)=(cRReσcRImσcRImσcRReσ)|(k0,c0),D_{c}\tilde{F}(k_{0},c_{0})=\begin{pmatrix}\partial_{c_{R}}\tilde{F}_{R}&\partial_{c_{I}}\tilde{F}_{R}\\ \partial_{c_{R}}\tilde{F}_{I}&\partial_{c_{I}}\tilde{F}_{I}\end{pmatrix}\Big{|}_{(k_{0},c_{0})}=\begin{pmatrix}\partial_{c_{R}}\text{Re}\,\mathcal{F}_{\sigma}&-\partial_{c_{R}}\text{Im}\,\mathcal{F}_{\sigma}\\ \partial_{c_{R}}\text{Im}\,\mathcal{F}_{\sigma}&\partial_{c_{R}}\text{Re}\,\mathcal{F}_{\sigma}\end{pmatrix}\Big{|}_{(k_{0},c_{0})},

where we also used the Cauchy-Riemann equation satisfied by σ\mathcal{F}_{\sigma} when restricted to cI0c_{I}\geq 0. From

detDcF~(k0,c0)=|cσ(k0,c0)|20,\det D_{c}\tilde{F}(k_{0},c_{0})=|\partial_{c}\mathcal{F}_{\sigma}(k_{0},c_{0})|^{2}\neq 0,

the Implicit Function Theorem implies that all roots of F~(k,c)\tilde{F}(k,c) near (k0,c0)(k_{0},c_{0}) form the graph of a C1,αC^{1,\alpha} complex-valued function 𝒞(k)\mathcal{C}(k) which contains (k0,c0)(k_{0},c_{0}) (or 𝒞(k)Cl02\mathcal{C}(k)\in C^{l_{0}-2} if x20(h,0)x_{20}\in(-h,0)). This and (2.25) prove the existence and the basic properties of 𝒞(k)\mathcal{C}(k) and (3.40). In the rest of the proof, we study the properties of 𝒞(k)\mathcal{C}(k).

Suppose cRFR(k0,c0)0\partial_{c_{R}}F_{R}(k_{0},c_{0})\neq 0. For any k[k0ϵ,k0+ϵ]k\in[k_{0}-\epsilon,k_{0}+\epsilon], from the Mean Value Theorem, there exists τ\tau between 0 and 𝒞I(k)\mathcal{C}_{I}(k) such that

0=F~I(k,𝒞(k))=FI(k,𝒞R(k))+𝒞I(k)cIF~I(k,𝒞R(k)+iτ).0=\tilde{F}_{I}\big{(}k,\mathcal{C}(k)\big{)}=F_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}+\mathcal{C}_{I}(k)\partial_{c_{I}}\tilde{F}_{I}\big{(}k,\mathcal{C}_{R}(k)+i\tau\big{)}.

The C1,αC^{1,\alpha} regularity of F~\tilde{F} and 𝒞(k)\mathcal{C}(k) and the Cauchy-Riemann equation imply

𝒞I(k)=\displaystyle\mathcal{C}_{I}(k)= FI(k,𝒞R(k))cIF~I(k,𝒞R(k)+iτ)=(U(0)𝒞R(k))2YI(k,𝒞R(k))cIF~I(k,𝒞R(k))+O(|𝒞I(k)|α)\displaystyle-\frac{F_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}}{\partial_{c_{I}}\tilde{F}_{I}\big{(}k,\mathcal{C}_{R}(k)+i\tau\big{)}}=-\frac{\big{(}U(0)-\mathcal{C}_{R}(k)\big{)}^{2}Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}}{\partial_{c_{I}}\tilde{F}_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}+O\big{(}|\mathcal{C}_{I}(k)|^{\alpha}\big{)}}
=\displaystyle= (U(0)𝒞R(k))2YI(k,𝒞R(k))cIFI(k,𝒞R(k))+O(|𝒞I(k)|α)=(U(0)c0+O(|kk0|))2YI(k,𝒞R(k))cRFR(k0,c0)+O(|kk0|α).\displaystyle-\frac{\big{(}U(0)-\mathcal{C}_{R}(k)\big{)}^{2}Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}}{\partial_{c_{I}}F_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}+O\big{(}|\mathcal{C}_{I}(k)|^{\alpha}\big{)}}=-\frac{\big{(}U(0)-c_{0}+O(|k-k_{0}|)\big{)}^{2}Y_{I}\big{(}k,\mathcal{C}_{R}(k)\big{)}}{\partial_{c_{R}}F_{R}(k_{0},c_{0})+O\big{(}|k-k_{0}|^{\alpha}\big{)}}.

The desired estimate on 𝒞I(k)\mathcal{C}_{I}(k) in statement (1) follows immediately.

In the rest of the proof, we consider the case of kσ(k0,c0)=0\partial_{k}\mathcal{F}_{\sigma}(k_{0},c_{0})=0. In this situation, Lemma 2.6(1) and 3.4 (and (3.9) as well) implies either i.) σ>0\sigma>0 or ii.) σ=0\sigma=0, k0=0k_{0}=0, and x20(h,0)x_{20}\in(-h,0), where we used k=2kK\partial_{k}=2k\partial_{K}. The analysis relies on the following equality obtained from the Implicit Function Theorem

𝒞(k)=DcF~(k,𝒞(k))1kF~(k,𝒞(k)).\mathcal{C}^{\prime}(k)=-D_{c}\tilde{F}(k,\mathcal{C}(k))^{-1}\partial_{k}\tilde{F}(k,\mathcal{C}(k)).

Let us first consider the case of k00k_{0}\neq 0. Without loss of generality, we may assume k0>0k_{0}>0. The C1,αC^{1,\alpha} regularity of F~\tilde{F} and 𝒞\mathcal{C} yields

|kF~(k,𝒞(k))|,|kF~(k,c0)|,|𝒞(k)|C~|kk0|α,|𝒞(k)c0|C~|kk0|1+α,|\partial_{k}\tilde{F}(k,\mathcal{C}(k))|,|\partial_{k}\tilde{F}(k,c_{0})|,|\mathcal{C}^{\prime}(k)|\leq\tilde{C}|k-k_{0}|^{\alpha},\quad|\mathcal{C}(k)-c_{0}|\leq\tilde{C}|k-k_{0}|^{1+\alpha},

and thus

|𝒞(k)+kσ(k,c0)|cRF(k0,c0)|2cRF(k0,c0)¯|=\displaystyle\big{|}\mathcal{C}^{\prime}(k)+\partial_{k}\mathcal{F}_{\sigma}(k,c_{0})|\partial_{c_{R}}F(k_{0},c_{0})|^{-2}\overline{\partial_{c_{R}}F(k_{0},c_{0})}\big{|}= |𝒞(k)+kF~(k,c0)/cRF~(k0,c0)|\displaystyle\big{|}\mathcal{C}^{\prime}(k)+\partial_{k}\tilde{F}(k,c_{0})/\partial_{c_{R}}\tilde{F}(k_{0},c_{0})\big{|}
\displaystyle\leq C~|kk0|α(1+α).\displaystyle\tilde{C}|k-k_{0}|^{\alpha(1+\alpha)}.

In the case of k0=0k_{0}=0, where x20(h,0)x_{20}\in(-h,0) must hold. Since the dependence of YY and FF on kk is actually through K=k2K=k^{2}, the same conclusions in Lemmas 2.3(8) and 2.5(1) still hold that FF is C3C^{3} in both KK and cc near (0,c0)(0,c_{0}) when restricted to cI0c_{I}\geq 0. Therefore 𝒞\mathcal{C} can also be viewed as a function of KK and we have

K𝒞=DcF~(k,𝒞)1KF~(k,𝒞).\partial_{K}\mathcal{C}=-D_{c}\tilde{F}(k,\mathcal{C})^{-1}\partial_{K}\tilde{F}(k,\mathcal{C}).

Much as in the above, we first obtain

|KF~(k,𝒞(k))|,|KF~(k,c0)|,|K𝒞|C~|K|β,|𝒞(k)c0|C~|K|1+β,|\partial_{K}\tilde{F}(k,\mathcal{C}(k))|,|\partial_{K}\tilde{F}(k,c_{0})|,|\partial_{K}\mathcal{C}|\leq\tilde{C}|K|^{\beta},\quad|\mathcal{C}(k)-c_{0}|\leq\tilde{C}|K|^{1+\beta},

where β=0\beta=0 if Kσ(0,c0)0\partial_{K}\mathcal{F}_{\sigma}(0,c_{0})\neq 0 and for any β(0,1)\beta\in(0,1) if Kσ(0,c0)=0\partial_{K}\mathcal{F}_{\sigma}(0,c_{0})=0. Along with the evenness of F~\tilde{F} and 𝒞\mathcal{C} in kk, it implies, for |k|(0,ϵ]|k|\in(0,\epsilon],

|𝒞(k)/k+2Kσ(k,c0)|cRF(0,c0)|2cRF(0,c0)¯|=\displaystyle\big{|}\mathcal{C}^{\prime}(k)/k+2\partial_{K}\mathcal{F}_{\sigma}(k,c_{0})|\partial_{c_{R}}F(0,c_{0})|^{-2}\overline{\partial_{c_{R}}F(0,c_{0})}\big{|}= 2|K𝒞+KF~(k,c0)/cRF~(0,c0)|\displaystyle 2\big{|}\partial_{K}\mathcal{C}+\partial_{K}\tilde{F}(k,c_{0})/\partial_{c_{R}}\tilde{F}(0,c_{0})\big{|}
\displaystyle\leq C~|K|α(1+β).\displaystyle\tilde{C}|K|^{\alpha(1+\beta)}.

It completes the proof of the lemma. ∎

In the following lemma, we consider a special case of bifurcation of unstable modes from an interior non-degenerate inflection value.

Lemma 3.12.

Suppose UC4U\in C^{4} and c0=U(x20)U((h,0))c_{0}=U(x_{20})\in U((-h,0)) satisfy

U′′(x20)=0>U′′′(x20),g0(U(0)c0)2Y(0,c0)U(0)(U(0)c0)>0,U^{\prime\prime}(x_{20})=0>U^{\prime\prime\prime}(x_{20}),\quad g_{0}\triangleq(U(0)-c_{0})^{2}Y(0,c_{0})-U^{\prime}(0)(U(0)-c_{0})>0,

then there exists ϵ0,δ>0\epsilon_{0},\delta>0 determined by UU and c0c_{0} such that for any g=g0+ϵ(g0,g0+ϵ0)g=g_{0}+\epsilon\in(g_{0},g_{0}+\epsilon_{0}), there exist C>0C>0 and k(ϵ)>0k(\epsilon)>0 depending on UU and c0c_{0} only and 𝒞(ϵ,k)=𝒞R(ϵ,k)+𝒞I(ϵ,k)\mathcal{C}(\epsilon,k)=\mathcal{C}_{R}(\epsilon,k)+\mathcal{C}_{I}(\epsilon,k), |k|k(ϵ)|k|\leq k(\epsilon), C1C^{1} in ϵ>0\epsilon>0 and kk and even in kk, such that

𝒞(ϵ,k(ϵ))=c0,k𝒞I(ϵ,k)/(2k)+(KFcRFI)|(0,c0)0 as ϵ0+ uniformly,\mathcal{C}(\epsilon,k(\epsilon))=c_{0},\quad\partial_{k}\mathcal{C}_{I}(\epsilon,k)/(2k)+\big{(}\partial_{K}F\partial_{c_{R}}F_{I}\big{)}\big{|}_{(0,c_{0})}\to 0\;\text{ as }\;\epsilon\to 0+\text{ uniformly},

where K=k2K=k^{2}, hence 𝒞I(ϵ,k)>0\mathcal{C}_{I}(\epsilon,k)>0 for |k|<k(ϵ)|k|<k(\epsilon), and moreover

F(ϵ,k,c)=0,|k|k(ϵ),|cc0|δ,F(\epsilon,k,c)=0,\quad|k|\leq k(\epsilon),\;|c-c_{0}|\leq\delta,

iff c=𝒞(ϵ,k)c=\mathcal{C}(\epsilon,k) or c=𝒞(ϵ,k)¯c=\overline{\mathcal{C}(\epsilon,k)}. Here FF is defined as in (2.23) in terms of UU and g=g0+ϵg=g_{0}+\epsilon.

According to Lemma 3.7(2), the assumption g0>0g_{0}>0 holds only if there exists kC>0k_{C}>0 such that y(kC,c0,0)=0y_{-}(k_{C},c_{0},0)=0. This lemma shows that the branch of the unstable eigenvalues bifurcating from the interior inflection value c0c_{0}, starting at the smaller wave number k(ϵ)k(\epsilon) which makes (k(ϵ),c0)(k(\epsilon),c_{0}) a singular neutral mode, connect back to (k(ϵ),c0)(-k(\epsilon),c_{0}).

Proof.

Apparently y(k,c0,x2)y_{-}(k,c_{0},x_{2}) and y(k,c0,x2)y_{-}^{\prime}(k,c_{0},x_{2}), and thus Y(k,c0)Y(k,c_{0}) and F(k,c0)F(k,c_{0}) as well if y(k,c0,0)0y_{-}(k,c_{0},0)\neq 0, are smooth in K=k2K=k^{2}. From KF(k,c0)>0\partial_{K}F(k,c_{0})>0 given by (3.9), there exists k(ϵ)>0k(\epsilon)>0 such that k(0)=0k(0)=0, F(ϵ,k(ϵ),c0)=0F(\epsilon,k(\epsilon),c_{0})=0, and k(ϵ)2k(\epsilon)^{2} is smooth in 0ϵ10\leq\epsilon\ll 1. Hence there exists C>0C>0 depending only on UU and c0c_{0} such that C1ϵk(ϵ)CϵC^{-1}\sqrt{\epsilon}\leq k(\epsilon)\leq C\sqrt{\epsilon}. Lemma 3.11 and the assumption U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0 imply that, for any ϵ1\epsilon\ll 1, F(ϵ,k,c)=0F(\epsilon,k,c)=0 has roots near (k(ϵ),c0)(k(\epsilon),c_{0}) with 0<k(ϵ)|k|10<k(\epsilon)-|k|\ll 1 and Imc>0\text{Im}\,c>0. In order to obtain a global branch for each small ϵ>0\epsilon>0, let

F~(ϵ,τ,c)=F(ϵ,k(ϵ)τ,c),|τ|1,|cc0|1, 0ϵ1.\tilde{F}(\epsilon,\tau,c)=F(\epsilon,k(\epsilon)\tau,c),\quad|\tau|\leq 1,\;|c-c_{0}|\ll 1,\;0\leq\epsilon\ll 1.

According to Lemma 2.4, FF is C2C^{2} in cc (when restricted to cI0c_{I}\geq 0 near c0c_{0}) and K=k2K=k^{2}, while k(ϵ)2k(\epsilon)^{2} is smooth in ϵ\epsilon, F~\tilde{F} is also C2C^{2} in such cc, ϵ\epsilon, and (also even in) τ\tau. We also extend FF and F~\tilde{F} as C2C^{2} functions defined on a whole neighborhood of c0c_{0}. Since F~(0,τ,c0)=0\tilde{F}(0,\tau,c_{0})=0 and the Jacobian satisfies

detDcF~(0,0,c0)=detDcF(0,0,c0)=|cF(0,0,c0)|2|cRFI(0,0,c0)|2>0,\det D_{c}\tilde{F}(0,0,c_{0})=\det D_{c}F(0,0,c_{0})=|\partial_{c}F(0,0,c_{0})|^{2}\geq|\partial_{c_{R}}F_{I}(0,0,c_{0})|^{2}>0,

the Implicit Function Theorem yields the (even in τ\tau) roots 𝒞~(ϵ,τ)\tilde{\mathcal{C}}(\epsilon,\tau) of F~(ϵ,τ,c)\tilde{F}(\epsilon,\tau,c). Clearly 𝒞~(ϵ,±1)=c0\tilde{\mathcal{C}}(\epsilon,\pm 1)=c_{0} due to the definition of k(ϵ)k(\epsilon). One may compute, for τ(0,1)\tau\in(0,1) and 0<ϵ10<\epsilon\ll 1,

τ𝒞~I\displaystyle\partial_{\tau}\tilde{\mathcal{C}}_{I} =Im((DcF~)1τF~)=2k(ϵ)2τIm((DcF~)1KF)\displaystyle=-\text{Im}\big{(}(D_{c}\tilde{F})^{-1}\partial_{\tau}\tilde{F}\big{)}=-2k(\epsilon)^{2}\tau\text{Im}\big{(}(D_{c}\tilde{F})^{-1}\partial_{K}F\big{)}
k(ϵ)2τ(KFcRFI/|cF|2)|(0,c0)<0,\displaystyle\leq k(\epsilon)^{2}\tau\big{(}\partial_{K}F\partial_{c_{R}}F_{I}/|\partial_{c}F|^{2}\big{)}\big{|}_{(0,c_{0})}<0,

where U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0 is used. Hence 𝒞~I>0\tilde{\mathcal{C}}_{I}>0 for |τ|<1|\tau|<1. Obviously 𝒞(ϵ,k)=𝒞~(ϵ,k/k(ϵ))\mathcal{C}(\epsilon,k)=\tilde{\mathcal{C}}(\epsilon,k/k(\epsilon)) satisfies the desired properties. ∎

3.6. Eigenvalues for |k|1|k|\gg 1

A major difference between the eigenvalue distributions of the linearized gravity waves and that of the capillary gravity waves is when |k|1|k|\gg 1. We shall work on 𝐅\mathbf{F} and FF of the linearized gravity waves. The analysis is divided into several steps starting with some rough bounds of the zeros of 𝐅(k,c)\mathbf{F}(k,c).

Lemma 3.13.

Assume l03l_{0}\geq 3, here exists k0>0k_{0}>0 depending only on UU, such that, for any |k|k0|k|\geq k_{0}, solutions to (2.22) (which is 𝐅(k,c)=0\mathbf{F}(k,c)=0) belong to SkLSkRS_{k}^{L}\cup S_{k}^{R} where

SkL,R={c=cR+cI±(|k|/g)12(U(0)cR)[1/2,2],|cI|k78}.S_{k}^{L,R}=\{c=c_{R}+c_{I}\in\mathbb{C}\mid\pm(|k|/g)^{\frac{1}{2}}(U(0)-c_{R})\in[1/2,2],\ |c_{I}|\leq k^{-\frac{7}{8}}\}.
Proof.

From Lemma 2.2,

k0>0, s. t. |y(k,c,x2)|(μ/2)sinhμ1(x2+h)>0,|k|k0,c,x2(h,0],\exists k_{0}>0,\text{ s. t. }|y_{-}(k,c,x_{2})|\geq(\mu/2)\sinh\mu^{-1}(x_{2}+h)>0,\quad\forall|k|\geq k_{0},\,c\in\mathbb{C},\,x_{2}\in(-h,0],

where we recall μ=(1+k2)12\mu=(1+k^{2})^{-\frac{1}{2}}. Hence we are able to work on (2.23) with F(k,c)F(k,c) and Y(k,c)Y(k,c). Due to the evenness of the problem in kk, we only consider k>0k>0.

From Lemma 2.4(3), for any α(0,12)\alpha\in(0,\frac{1}{2}), there exist k0,C>0k_{0},C>0 depending only on α\alpha and UU, such that for any k>k0k>k_{0},

|F(k,c)(U(0)c)2k+U(0)(U(0)c)+g|Ck1α(|U(0)c|74+|U(0)c|2).\big{|}F(k,c)-(U(0)-c)^{2}k+U^{\prime}(0)(U(0)-c)+g\big{|}\leq Ck^{1-\alpha}\big{(}|U(0)-c|^{\frac{7}{4}}+|U(0)-c|^{2}\big{)}.

On the one hand, if |U(0)c|C1k12|U(0)-c|\leq C_{1}k^{-\frac{1}{2}} where C1(0,1)C_{1}\in(0,1), then

|F(k,c)|gC12CC1k18α.|F(k,c)|\geq g-C_{1}^{2}-CC_{1}k^{\frac{1}{8}-\alpha}.

Taking α=13\alpha=\frac{1}{3} and sufficiently large k0k_{0} and C11C_{1}^{-1} determined by UU, we obtain F(k,c)0F(k,c)\neq 0 if kk0k\geq k_{0} and |U(0)c|C1k12|U(0)-c|\leq C_{1}k^{-\frac{1}{2}}.

On the other hand, if |U(0)c|2g/k|U(0)-c|\geq\sqrt{2g/k}, then

|F(k,c)|\displaystyle|F(k,c)|\geq |U(0)c|2k|U(0)(U(0)c)+g|Ck1α(|U(0)c|74+|U(0)c|2)\displaystyle|U(0)-c|^{2}k-|U^{\prime}(0)(U(0)-c)+g|-Ck^{1-\alpha}\big{(}|U(0)-c|^{\frac{7}{4}}+|U(0)-c|^{2}\big{)}
\displaystyle\geq |U(0)c|2(kg|U(0)c|2C(k1α(1+|U(0)c|14)+|U(0)c|1))\displaystyle|U(0)-c|^{2}\big{(}k-g|U(0)-c|^{-2}-C\big{(}k^{1-\alpha}(1+|U(0)-c|^{-\frac{1}{4}})+|U(0)-c|^{-1}\big{)}\big{)}
\displaystyle\geq |U(0)c|2(k/2C(k98α+k12)).\displaystyle|U(0)-c|^{2}\big{(}k/2-C(k^{\frac{9}{8}-\alpha}+k^{\frac{1}{2}})\big{)}.

Again, taking α=13\alpha=\frac{1}{3} and k0k_{0} sufficiently large determined by UU, we obtain F(k,c)0F(k,c)\neq 0, either.

The above analysis implies

(3.41) F(k,c)=0,kk0k12|cU(0)|[C1,2g],F(k,c)=0,\;k\geq k_{0}\implies k^{\frac{1}{2}}|c-U(0)|\in[C_{1},\sqrt{2g}],

which also yields the desired upper bound of |U(0)cR||U(0)-c_{R}|. Let us take the real part of F(k,c)=0F(k,c)=0, namely,

((U(0)cR)2cI2)YR=g+U(0)(U(0)cR)2(U(0)cR)cIYI.\big{(}(U(0)-c_{R})^{2}-c_{I}^{2})Y_{R}=g+U^{\prime}(0)(U(0)-c_{R})-2(U(0)-c_{R})c_{I}Y_{I}.

For kk0k\geq k_{0} and cc satisfying (3.41), a simple bound of YY can be derived from Lemma 2.4(3),

|Y(k,c)k|Ck1α,α(0,1/2)k((U(0)cR)2cI2)g/2,\big{|}Y(k,c)-k\big{|}\leq Ck^{1-\alpha},\;\alpha\in(0,1/2)\implies k\big{(}(U(0)-c_{R})^{2}-c_{I}^{2})\geq g/2,

which implies the desired lower bound of |U(0)cR||U(0)-c_{R}|. Taking the imaginary part of F(k,c)=0F(k,c)=0, we obtain

(3.42) ((U(0)cR)2cI2)YI2(U(0)cR)cIYR+U(0)cI=0cI=(U(0)cR)2YI(k,c)2(U(0)cR)YR+cIYIU(0),\begin{split}&\big{(}(U(0)-c_{R})^{2}-c_{I}^{2})Y_{I}-2(U(0)-c_{R})c_{I}Y_{R}+U^{\prime}(0)c_{I}=0\\ \implies&c_{I}=\frac{(U(0)-c_{R})^{2}Y_{I}(k,c)}{2(U(0)-c_{R})Y_{R}+c_{I}Y_{I}-U^{\prime}(0)},\end{split}

and thus the estimate on cIc_{I} follows from (3.41), the above bounds on |Yk||Y-k| and |U(0)cR||U(0)-c_{R}| and letting α=2/5\alpha=2/5 and k01k_{0}\gg 1. ∎

Due to the evenness and conjugacy property (2.25), we shall only need to consider

Sk+L,R={c=cR+cISkL,RcI0}.S_{k+}^{L,R}=\{c=c_{R}+c_{I}\in S_{k}^{L,R}\mid c_{I}\geq 0\}.

Before we proceed to obtain the roots of F(k,)F(k,\cdot), we first establish some estimates on YY and cY\partial_{c}Y for cSk+L,Rc\in S_{k+}^{L,R}.

Lemma 3.14.

Assume l05l_{0}\geq 5. There exist k0,C>0k_{0},C>0 depending only on UU, such that, for any |k|k0|k|\geq k_{0}, cSk+c\in S_{k+}^{\dagger} where =L,R\dagger=L,R, and 0jl050\leq j\leq l_{0}-5, it holds

|cj(Y(k,c)|k|)|C|k|j12.|\partial_{c}^{j}(Y(k,c)-|k|)|\leq C|k|^{\frac{j-1}{2}}.
Proof.

Due to the evenness we shall focus on k>0k>0. We first consider cSk+Lc\in S_{k+}^{L}.

One may compute, for cU([h,0])c\in U([-h,0]),

cRYI(k,c)=\displaystyle\partial_{c_{R}}Y_{I}(k,c)= c(πU′′(x2c)U(x2c))y(k,c,x2c)2|y(k,c,0)|2+2πU′′(x2c)y(k,c,x2c)c(y(k,c,x2c))U(x2c)|y(k,c,0)|2\displaystyle\partial_{c}\left(\frac{\pi U^{\prime\prime}(x_{2}^{c})}{U^{\prime}(x_{2}^{c})}\right)\frac{y_{-}(k,c,x_{2}^{c})^{2}}{|y_{-}(k,c,0)|^{2}}+\frac{2\pi U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})\partial_{c}\big{(}y_{-}(k,c,x_{2}^{c})\big{)}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{2}}
2πU′′(x2c)y(k,c,x2c)2(y(k,c,0)cy(k,c,0))U(x2c)|y(k,c,0)|4,\displaystyle-\frac{2\pi U^{\prime\prime}(x_{2}^{c})y_{-}(k,c,x_{2}^{c})^{2}\big{(}y_{-}(k,c,0)\cdot\partial_{c}y_{-}(k,c,0)\big{)}}{U^{\prime}(x_{2}^{c})|y_{-}(k,c,0)|^{4}},

and cRjYI(c)\partial_{c_{R}}^{j}Y_{I}(c) can be computed in a similar fashion, where we recall x2c=U1(c)x_{2}^{c}=U^{-1}(c). Using Lemmas 2.2 and 2.3(8b)(9), we obtain that there exist C,k0>0C,k_{0}>0 depending only on UU such that, for kk0k\geq k_{0} and 0jl040\leq j\leq l_{0}-4,

|cjYI(k,c)|Ckje2kx2c,c[U(h)+2(ρ0k)1,U(0)2(ρ0k)1].|\partial_{c}^{j}Y_{I}(k,c)|\leq Ck^{j}e^{2kx_{2}^{c}},\quad\forall c\in[U(-h)+2(\rho_{0}k)^{-1},U(0)-2(\rho_{0}k)^{-1}].

Here we recall that ρ0\rho_{0} was given in (2.7). Due to the singularity in the integral representation of Y(k,c)Y(k,c) given in Lemma 2.4(7), we use a cut-off function γC(,[0,1])\gamma\in C^{\infty}(\mathbb{R},[0,1]) satisfying

γ|[14,3]=1,γ|[18,4]c=0.\gamma|_{[\frac{1}{4},3]}=1,\quad\gamma|_{[\frac{1}{8},4]^{c}}=0.

For cU([h,0])c\in U([-h,0]), let

f1(c)=YI(k,c)γ((k/g)12(U(0)c)),f2(c)=YI(k,c)f1(c).f_{1}(c)=Y_{I}(k,c)\gamma\big{(}(k/g)^{\frac{1}{2}}(U(0)-c)\big{)},\quad f_{2}(c)=Y_{I}(k,c)-f_{1}(c).

We rewrite Lemma 2.4(7) for cSk+Lc\in S_{k+}^{L} and kk0k\geq k_{0},

Y(k,c)kcothkh=Y1(k,c)+Y2(k,c),Y(k,c)-k\coth kh=Y^{1}(k,c)+Y^{2}(k,c),

where

Y1(k,c)=1πf1(c)cc𝑑c,Y2(k,c)=1πU(h)U(0)f2(c)cc𝑑c,Y^{1}(k,c)=\frac{1}{\pi}\int_{\mathbb{R}}\frac{f_{1}(c^{\prime})}{c^{\prime}-c}dc^{\prime},\quad Y^{2}(k,c)=\frac{1}{\pi}\int_{U(-h)}^{U(0)}\frac{f_{2}(c^{\prime})}{c^{\prime}-c}dc^{\prime},

and, for cU([h,0])Sk+Lc\in U([-h,0])\cap S_{k+}^{L}, Y1(k,c)=limcI0+Y1(k,c+icI)Y^{1}(k,c)=\lim_{c_{I}\to 0+}Y^{1}(k,c+ic_{I}) is understood. Using the above estimates on YIY_{I} and the definition of f1f_{1}, it is straight forward to see, for jl03j\leq l_{0}-3,

|f1(j)(c)|Ckje2kx2c, if (k/g)12(U(0)c)[1/8,4], and f1(c)=0, otherwise.|f_{1}^{(j)}(c)|\leq Ck^{j}e^{2kx_{2}^{c}},\;\text{ if }\;(k/g)^{\frac{1}{2}}(U(0)-c)\in[1/8,4],\;\text{ and }\;f_{1}(c)=0,\;\text{ otherwise}.

It implies, for all cc\in\mathbb{C} with cI0c_{I}\geq 0,

|cjY1(k,)|L2()2C|f1(j)|L2()2Ck2jU(0)4g/kU(0)18g/ke4kU1(c)𝑑cCk2j1ekC,|\partial_{c}^{j}Y^{1}(k,\cdot)|_{L^{2}(\mathbb{R})}^{2}\leq C|f_{1}^{(j)}|_{L^{2}(\mathbb{R})}^{2}\leq Ck^{2j}\int_{U(0)-4\sqrt{g/k}}^{U(0)-\frac{1}{8}\sqrt{g/k}}e^{4kU^{-1}(c^{\prime})}dc^{\prime}\leq Ck^{2j-1}e^{-\frac{\sqrt{k}}{C}},

where the substitution c=U(x2)c^{\prime}=U(x_{2}) was also used in the last step. From interpolation, we obtain, for jl05j\leq l_{0}-5,

|cjY1(k,)|L()CkjekC,kk0.|\partial_{c}^{j}Y^{1}(k,\cdot)|_{L^{\infty}(\mathbb{R})}\leq Ck^{j}e^{-\frac{\sqrt{k}}{C}},\quad\forall k\geq k_{0}.

Concerning Y2(k,c)Y^{2}(k,c), we have, for k0k\geq 0, cSk+Lc\in S_{k+}^{L}, and j0j\geq 0,

|cjY2(k,c)|\displaystyle|\partial_{c}^{j}Y^{2}(k,c)|\leq 1π(U(h)U(0)3g/k+U(0)g/16kU(0))|YI(k,c)||cc|j+1dc\displaystyle\frac{1}{\pi}\Big{(}\int_{U(-h)}^{U(0)-3\sqrt{g/k}}+\int_{U(0)-\sqrt{g/16k}}^{U(0)}\Big{)}\frac{|Y_{I}(k,c^{\prime})|}{|c^{\prime}-c|^{j+1}}dc^{\prime}
\displaystyle\leq Ckj+12U(h)U(0)e2kU1(c)𝑑cCkj12,\displaystyle Ck^{\frac{j+1}{2}}\int_{U(-h)}^{U(0)}e^{2kU^{-1}(c^{\prime})}dc^{\prime}\leq Ck^{\frac{j-1}{2}},

where the substitution c=U(x2)c^{\prime}=U(x_{2}) was used as well. Therefore we obtain the desired estimate on YY in Sk+LS_{k+}^{L}.

The estimate of Y(k,c)Y(k,c) for cSk+Rc\in S_{k+}^{R} is similar, but simpler. Actually in this case, the norm of denominator ccc^{\prime}-c in the integrand of the integral representation in Lemma 2.4(7) of YY is bounded below by O(|k|)O(\sqrt{|k|}). Hence YY can be estimated much as the above Y2(k,c)Y^{2}(k,c) and we omit the details. ∎

In the following we analyze possible zeros of 𝐅(k,c)=0\mathbf{F}(k,c)=0 for k1k\gg 1 in the rectangles Sk+L,RS_{k+}^{L,R}.

Lemma 3.15.

Assume l06l_{0}\geq 6. There exist k0,C>0k_{0},C>0 depending only on UU, such that there exist an analytic function c+(k)(U(0),+)c^{+}(k)\in(U(0),+\infty) and a Cl02C^{l_{0}-2} function c(k)=cR(k)+icI(k)c^{-}(k)=c_{R}^{-}(k)+ic_{I}^{-}(k) both defined and even for |k|k0|k|\geq k_{0} and the following hold.

  1. (1)

    For any |k|k0|k|\geq k_{0},

    𝐅(k,c)=0,|k|k0,cI0 iff c{c+(k),c(k)},Imc0.\mathbf{F}(k,c)=0,\;\;|k|\geq k_{0},\;\;c_{I}\geq 0\text{ iff }\;c\in\{c^{+}(k),\,c^{-}(k)\},\;\text{Im}\,c\geq 0.

    Namely, for |k|k0|k|\geq k_{0}, the roots of 𝐅(k,)\mathbf{F}(k,\cdot) are either only c+(k)c^{+}(k) if cI(k)<0c_{I}^{-}(k)<0, or {c+(k),c(k),c(k)¯}\{c^{+}(k),\,c^{-}(k),\,\overline{c^{-}(k)}\} if cI(k)0c_{I}^{-}(k)\geq 0.

  2. (2)

    c±(k)c^{\pm}(k) satisfy the estimates

    |c±(k)U(0)+U(0)2|k|g|k|(1+U(0)24g|k|)|Ck2,\Big{|}c^{\pm}(k)-U(0)+\frac{U^{\prime}(0)}{2|k|}\mp\sqrt{\frac{g}{|k|}\Big{(}1+\frac{U^{\prime}(0)^{2}}{4g|k|}}\Big{)}\Big{|}\leq Ck^{-2},
    |cI(k)g/(4|k|3)YI(k,cR(k))|Ck2|YI(k,cR(k))|,\big{|}c_{I}^{-}(k)-\sqrt{g/(4|k|^{3})}Y_{I}(k,c_{R}^{-}(k))\big{|}\leq Ck^{-2}|Y_{I}(k,c_{R}^{-}(k))|,
    C1|U′′(U1(cR(k))||k|32e2g|k|U(0)|cI(k)|C|U′′(U1(cR(k))|.C^{-1}\big{|}U^{\prime\prime}(U^{-1}(c_{R}^{-}(k)\big{)}\big{|}\leq|k|^{\frac{3}{2}}e^{\frac{2\sqrt{g|k|}}{U^{\prime}(0)}}|c_{I}^{-}(k)|\leq C\big{|}U^{\prime\prime}(U^{-1}(c_{R}^{-}(k)\big{)}\big{|}.

    Moreover, the following inequality holds for c(k)c^{-}(k) if cI(k)0c_{I}^{-}(k)\geq 0 and c+(k)c^{+}(k),

    |DcF(k,c±(k))2g|k||C|k|12.|D_{c}F(k,c^{\pm}(k))\mp 2\sqrt{g|k|}|\leq C|k|^{-\frac{1}{2}}.
Proof.

Again we only need to consider F(k,c)=0F(k,c)=0 for k1k\gg 1. As in the bifurcation analysis in the proof of Lemma 3.11, we consider a Cl02C^{l_{0}-2} extension Y~(k,c)\tilde{Y}(k,c) of Y(k,c)Y(k,c) from the domain cSk+L,Rc\in S_{k+}^{L,R} to SkL,RS_{k}^{L,R}, which is different from the original Y(k,c)Y(k,c) defined on SkL,RSk+L,RS_{k}^{L,R}\setminus S_{k+}^{L,R}. The extension Y~(k,c)\tilde{Y}(k,c) can be defined as, e.g.,

Y~(k,c)=Y~R(k,c)+iY~I(k,c)=l=1l01alY(k,cRilcI), for c¯Sk+L,R,\tilde{Y}(k,c)=\tilde{Y}_{R}(k,c)+i\tilde{Y}_{I}(k,c)=\sum_{l=1}^{l_{0}-1}a_{l}Y(k,c_{R}-ilc_{I}),\;\text{ for }\;\bar{c}\in S_{k+}^{L,R},

where a1,,al01a_{1},\ldots,a_{l_{0}-1} can be chosen so that all derivatives up to the order of l02l_{0}-2 are matched at cI=0c_{I}=0. Extend F(k,c)F(k,c) to F~(k,c)\tilde{F}(k,c) accordingly and consider F~(k,c)=0\tilde{F}(k,c)=0, namely,

(3.43) F~(k,c)=(U(0)c)2Y~(k,c)U(0)(U(0)c)g=0,cSkL,R.\tilde{F}(k,c)=(U(0)-c)^{2}\tilde{Y}(k,c)-U^{\prime}(0)(U(0)-c)-g=0,\quad c\in S_{k}^{L,R}.

Treating it as a quadratic equation of U(0)cU(0)-c, its two branches of roots must satisfy

c=f±(k,c)U(0)U(0)2Y~(k,c)±U(0)24Y~(k,c)2+gY~(k,c).c=f_{\pm}(k,c)\triangleq U(0)-\frac{U^{\prime}(0)}{2\tilde{Y}(k,c)}\pm\sqrt{\frac{U^{\prime}(0)^{2}}{4\tilde{Y}(k,c)^{2}}+\frac{g}{\tilde{Y}(k,c)}}.

For kk0k\geq k_{0}, the argument of the last square root is close to g/|k|g/|k| and the ±\pm signs correspond to the square roots close to ±g/|k|\pm\sqrt{g/|k|}. From Lemma 3.14 and the definition of the extension Y~\tilde{Y}, there exist some k0,C>0k_{0},C>0 determined by UU such that, for k>k0k>k_{0} and cSkL,Rc\in S_{k}^{L,R}, it holds

|(U(0)24Y~(k,c)2+gY~(k,c))gk(1+U(0)24gk)|Ck52.\Big{|}\Big{(}\frac{U^{\prime}(0)^{2}}{4\tilde{Y}(k,c)^{2}}+\frac{g}{\tilde{Y}(k,c)}\Big{)}-\frac{g}{k}\Big{(}1+\frac{U^{\prime}(0)^{2}}{4gk}\Big{)}\Big{|}\leq Ck^{-\frac{5}{2}}.

The other term U(0)/(2Y~(k,c))U^{\prime}(0)/(2\tilde{Y}(k,c)) can be handled similarly and thus for such (k,c)(k,c) we obtain,

|f±(k,c)U(0)+U(0)2kgk(1+U(0)24gk)|Ck2f±(k,c)SkR,L.\Big{|}f_{\pm}(k,c)-U(0)+\frac{U^{\prime}(0)}{2k}\mp\sqrt{\frac{g}{k}\Big{(}1+\frac{U^{\prime}(0)^{2}}{4gk}}\Big{)}\Big{|}\leq Ck^{-2}\implies f_{\pm}(k,c)\in S_{k}^{R,L}.

Moreover, from Lemma 3.14 we also have the derivative estimates of f±f_{\pm}

|Dcf±(k,c)|=\displaystyle|D_{c}f_{\pm}(k,c)|= |DcY~(k,c)||U(0)2Y~(k,c)2(U(0)24Y~(k,c)2+gY~(k,c))12(U(0)24Y~(k,c)3+g2Y~(k,c)2)|\displaystyle|D_{c}\tilde{Y}(k,c)|\Big{|}\frac{U^{\prime}(0)}{2\tilde{Y}(k,c)^{2}}\mp\Big{(}\frac{U^{\prime}(0)^{2}}{4\tilde{Y}(k,c)^{2}}+\frac{g}{\tilde{Y}(k,c)}\Big{)}^{-\frac{1}{2}}\Big{(}\frac{U^{\prime}(0)^{2}}{4\tilde{Y}(k,c)^{3}}+\frac{g}{2\tilde{Y}(k,c)^{2}}\Big{)}\Big{|}
\displaystyle\leq Ck32.\displaystyle Ck^{-\frac{3}{2}}.

Therefore, for kk01k\geq k_{0}\gg 1, f(k,)f_{-}(k,\cdot) is a Cl02C^{l_{0}-2} contraction on SkLS_{k}^{L} and f+(k,)f_{+}(k,\cdot) an analytic contraction on SkRS_{k}^{R}. Let c±(k)SkR,Lc^{\pm}(k)\in S_{k}^{R,L} denote their unique fixed points, which satisfy the same above leading order asymptotics as f±f_{\pm} and are Cl02C^{l_{0}-2} in kk.

It is straight forward to compute

DcF~(k,c)=2(cU(0))Y~(k,c)+(U(0)c)2DcY~(k,c)+U(0).D_{c}\tilde{F}(k,c)=2(c-U(0))\tilde{Y}(k,c)+(U(0)-c)^{2}D_{c}\tilde{Y}(k,c)+U^{\prime}(0).

From Lemma 3.14 and the leading order estimates of c±(k)c^{\pm}(k) which yields

|c±(k)U(0)g/k+U(0)/(2k)|Ck32,|c^{\pm}(k)-U(0)\mp\sqrt{g/k}+U^{\prime}(0)/(2k)|\leq Ck^{-\frac{3}{2}},

we obtain the estimate on DcF~(k,c)D_{c}\tilde{F}(k,c)

|DcF~(k,c±(k))2gk|Ck12.|D_{c}\tilde{F}(k,c^{\pm}(k))\mp 2\sqrt{gk}|\leq Ck^{-\frac{1}{2}}.

To complete the proof, we consider the imaginary parts of c±(k)c^{\pm}(k). Using (3.43) we have

((U(0)cR)2cI2)Y~I(k,c)2(U(0)cR)cIY~R(k,c)+U(0)cI=0, at c=c±(k).\big{(}(U(0)-c_{R})^{2}-c_{I}^{2})\tilde{Y}_{I}(k,c)-2(U(0)-c_{R})c_{I}\tilde{Y}_{R}(k,c)+U^{\prime}(0)c_{I}=0,\;\text{ at }\;c=c^{\pm}(k).

From the Mean Value Theorem, there exists τ±\tau^{\pm} between cR±(k)c_{R}^{\pm}(k) and c±(k)c^{\pm}(k) such that

Y~I(k,c±(k))YI(k,cR±(k))=cI±(k)cIY~I(k,τ±),\tilde{Y}_{I}(k,c^{\pm}(k))-Y_{I}(k,c_{R}^{\pm}(k))=c_{I}^{\pm}(k)\partial_{c_{I}}\tilde{Y}_{I}(k,\tau^{\pm}),

which, along with Lemma 3.14, yields, at c=c±(k)c=c^{\pm}(k),

(U(0)cR)2(YI(k,cR)+cIcIY~I(k,τ±))cI2Y~I(k,c)2(U(0)cR)cIY~R(k,c)+U(0)cI=0.(U(0)-c_{R})^{2}\big{(}Y_{I}(k,c_{R})+c_{I}\partial_{c_{I}}\tilde{Y}_{I}(k,\tau^{\pm})\big{)}-c_{I}^{2}\tilde{Y}_{I}(k,c)-2(U(0)-c_{R})c_{I}\tilde{Y}_{R}(k,c)+U^{\prime}(0)c_{I}=0.

Therefore we obtain

cI±(k)=(U(0)cR)2YI(k,cR)(U(0)cR)(2Y~R(k,c)(U(0)cR)cIY~I(k,τ±))U(0)+cIY~I(k,c)|c=c±(k),c_{I}^{\pm}(k)=\frac{(U(0)-c_{R})^{2}Y_{I}(k,c_{R})}{(U(0)-c_{R})\big{(}2\tilde{Y}_{R}(k,c)-(U(0)-c_{R})\partial_{c_{I}}\tilde{Y}_{I}(k,\tau^{\pm})\big{)}-U^{\prime}(0)+c_{I}\tilde{Y}_{I}(k,c)}\Big{|}_{c=c^{\pm}(k)},

which, along with Lemma 3.14 and the asymptotics of c±(k)c^{\pm}(k), implies the desired estimates on cI±(k)c_{I}^{\pm}(k) in term of YI(k,cR±(k))Y_{I}(k,c_{R}^{\pm}(k)). In particular, since cR+(k)>U(0)c_{R}^{+}(k)>U(0), we have cI+(k)=0c_{I}^{+}(k)=0 and c+(k)>U(0)c^{+}(k)>U(0). As c+(k)c^{+}(k) belongs to the domain of analyticity of F(k,c)F(k,c), it is also an analytic function of kk. Finally, the upper and lower bounds of cI(k)c_{I}^{-}(k) follow from Lemmas 2.4(5) and 2.2 and U1(cR(k))=g/(U(0)k)+O(k1)U^{-1}(c_{R}^{-}(k))=-\sqrt{g}/(U^{\prime}(0)\sqrt{k})+O(k^{-1}). ∎

3.7. Eigenvalue distribution

With the above preparations, we are ready to prove the main theorems on the eigenvalue distribution of the linearized gravity water waves.

\bullet Proof of Theorem 1.1. Throughout the proof, we recall that 𝐅(k,c)\mathbf{F}(k,c) and F(k,c)F(k,c) have the same zero sets (Lemma 2.5(2)) and we often mix them in the arguments.

Most of statement (1) has been proved in Lemma 3.15 except for the global extension etc. of c+(k)c^{+}(k) in (1c) which will be proved here.

Since cF(0,c)>0\partial_{c}F(0,c)>0 for c[U(0),+)c\in[U(0),+\infty) which can be verified directly using its integral formula given in Lemma 2.6(1), there exists a unique c1>U(0)c_{1}>U(0) such that

(3.44) F(0,c1)=0,cF(0,c1)>0, and F(0,c)(cc1)>0,c[U(0),+){c1}.F(0,c_{1})=0,\;\partial_{c}F(0,c_{1})>0,\;\text{ and }\;F(0,c)(c-c_{1})>0,\;c\in[U(0),+\infty)\setminus\{c_{1}\}.

Let

Ω1={ccR(U(0),c1+1),cI(1,1)}.\Omega_{1}=\{c\in\mathbb{C}\mid c_{R}\in(U(0),c_{1}+1),\,c_{I}\in(-1,1)\}.

Because F(k,c)F(k,c) is also strictly increasing in k0k\geq 0 for any cU(0)c\geq U(0) (Lemma 2.6(2)), we have F(k,c)>0F(k,c)>0 for all c>c1c>c_{1} and kk\in\mathbb{R}. Therefore, according to Lemma 2.6(1) and the semicircle theorem [28], 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for any kk\in\mathbb{R} and cΩ1c\in\partial\Omega_{1} and c1c_{1} is the only root of F(0,)F(0,\cdot) in Ω1\Omega_{1}, which is also simple. It implies Ind(𝐅(k,),Ω1)=1\big{(}\mathbf{F}(k,\cdot),\Omega_{1}\big{)}=1 for all kk\in\mathbb{R}. On the one hand, from Lemma 2.7 and Remark 2.3, the unique root c(k)c(k) of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω1\Omega_{1} is simple and depends on kk\in\mathbb{R} analytically and evenly. On the other hand, Lemma 3.15 implies that the root c+(k)Ω1c^{+}(k)\in\Omega_{1} for |k|k0|k|\geq k_{0}. Therefore c+(k)c^{+}(k) coincides with c(k)c(k) for |k|k0|k|\geq k_{0} and thus c(k)c(k) serves as its extension for all kk\in\mathbb{R} as the only root of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω1\Omega_{1}, which is simple. It is also the only root of 𝐅(k,)\mathbf{F}(k,\cdot) in (U(0),+)(U(0),+\infty) since F(k,c)>0F(k,c)>0 for all c>c1c>c_{1} and kk\in\mathbb{R} obtained in the above. Finally, since cF(k,c1)>0\partial_{c}F(k,c_{1})>0, cF(k,c(k))\partial_{c}F(k,c(k)) does not change sign as c(k)c(k) is a simple root for all kk, and KF(k,c+(k))>0\partial_{K}F(k,c^{+}(k))>0 where K=k2K=k^{2} (Lemma 2.6(2)), we obtain c(k)<0c^{\prime}(k)<0 for k>0k>0. This completes the proof of statement (1c).

In statement (2), the existence and uniqueness of kk_{-} has been obtained in Lemma 3.3. From Lemma 2.6, there exist c2<U(h)c_{2}<U(-h) unique in (,U(h)](-\infty,U(-h)], such that

(3.45) F(0,c2)=0=F(k,U(h)),cF(0,c2)<0,F(0,c)(cc2)>0,c(,U(h)]{c2}.\begin{split}&F(0,c_{2})=0=F(k_{-},U(-h)),\;\partial_{c}F(0,c_{2})<0,\\ &F(0,c)(c-c_{2})>0,\ \forall c\in(-\infty,U(-h)]\setminus\{c_{2}\}.\end{split}

Due to the semi-circle theorem, F(0,)F(0,\cdot) has exactly two roots c2c_{2} and c1=c+(0)c_{1}=c^{+}(0) in Ω2\Omega_{2}, both of which are simple, and c2c_{2} the only one in Ω3\Omega_{3}, where

Ω2={c|c(U(h)+U(0))/2|>|U(0)U(h)|/2},\Omega_{2}=\{c\in\mathbb{C}\mid|c-(U(-h)+U(0))/2|>|U(0)-U(-h)|/2\},
Ω3={ccR(c21,U(h)),cI(1,1)}.\Omega_{3}=\{c\in\mathbb{C}\mid c_{R}\in(c_{2}-1,U(-h)),\,c_{I}\in(-1,1)\}.

The defintions of c2c_{2} as well as that of c1c_{1}, the monotonicity of F(k,c)F(k,c) in k>0k>0 for cU((h,0])c\in\mathbb{R}\setminus U((-h,0]), the local monotonicity of F(0,c)F(0,c) in cU((h,0])c\in\mathbb{R}\setminus U((-h,0]), and the semi-circle theorem also imply that a.) for any kk\in\mathbb{R}, the roots of 𝐅(k,c)\mathbf{F}(k,c) in Ω2\Omega_{2} have to belong to Ω1Ω3\Omega_{1}\cup\Omega_{3} and b.) the only root of 𝐅(k,c)\mathbf{F}(k,c) for cΩ1Ω3c\in\partial\Omega_{1}\cup\partial\Omega_{3} is (k,U(h))(k_{-},U(-h)). Consequently, the total number of roots (counting the multiplicity) of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω2\Omega_{2} can change only at k=kk=k_{-}. From the same argument as in the above proof of statement (1c), we obtain that, for |k|<k|k|<k_{-}, the only root of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω3\Omega_{3}, which is also the only one in (,U(h))(-\infty,U(-h)), is given by some c(k)[c2,U(h))c_{-}(k)\in[c_{2},U(-h)). It is analytic and even in kk and satisfies

c(0)=c2,cF(k,c(k))<0,c(k)>0,F(k,c)(c(k)c)>0,|k|<k,cU(h).c_{-}(0)=c_{2},\;\partial_{c}F(k,c_{-}(k))<0,\;c_{-}^{\prime}(k)>0,\;F(k,c)(c_{-}(k)-c)>0,\;\forall|k|<k_{-},\,c\leq U(-h).

Here the last property is due to the monotonicity of F(k,c)F(k,c) in k>0k>0 for cU((h,0])c\in\mathbb{R}\setminus U((-h,0]) (Lemma 2.6(2)) and that of c(k)c_{-}(k). Hence c+(k)c^{+}(k) and c(k)c_{-}(k) are the only roots of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω2\Omega_{2} for all |k|<k|k|<k_{-}. Moreover, for |k|>k0|k|>k_{0}, the only root of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω2\Omega_{2} (and Ω1Ω3\Omega_{1}\cup\Omega_{3} as well) is c+(k)c^{+}(k) implies that this also holds for all |k|>k|k|>k_{-}. The limit c((k))c_{-}((k_{-})-) exists due to its boundedness and monotonicity. It has to be equal to U(h)U(-h), otherwise, if c((k))<U(h)c_{-}((k_{-})-)<U(-h), by Lemma 2.7 c(k)c_{-}(k) could be continued beyond into k>kk>k_{-} and then 𝐅(k,)\mathbf{F}(k,\cdot) would have at least two roots c+(k)c^{+}(k) and c(k)c_{-}(k) in Ω2\Omega_{2} for 1kk>01\gg k-k_{-}>0, which contradicts the above analysis. Finally, statement (2c) follows from Lemma 3.11(1) and the signs kF>0\partial_{k}F>0 and cF<0\partial_{c}F<0 at (k,U(h))(k_{-},U(-h)) (Lemma 3.10).

Statement (3) regarding cc near the inflection values of UU has been proved in Lemmas 3.7 and 3.11, see also Remark 3.3.

Statement (4) on the linear instability is a direct corollary of statements (1)–(3). \square

We proceed to prove Theorem 1.2 assuming either U′′>0U^{\prime\prime}>0 or U′′0U^{\prime\prime}\leq 0 on (h,0)(-h,0).

\bullet Proof of Theorem 1.2. Suppose U′′>0U^{\prime\prime}>0 on (h,0)(-h,0), then Theorem 1.1(2) and Lemmas 2.3(4) and 2.5(4) imply that F(k,c)=0F(k,c)=0 with cU([h,0])c\in U([-h,0]) iff (k,c)=(k,U(h))(k,c)=(k_{-},U(-h)). From Lemma 3.15 (or Theorem 1.1(1)), there exists k0>kk_{0}>k_{-} such that, for any |k|>k0|k|>k_{0}, there are exactly three roots of F(k,)F(k,\cdot), which are simple and given by c+(k)(U(0),c+(0)]c^{+}(k)\in(U(0),c^{+}(0)] (already extended for all kk\in\mathbb{R} in Theorem 1.1(1)), c(k)=cR(k)+icI(k)c^{-}(k)=c_{R}^{-}(k)+ic_{I}^{-}(k), and c(k)¯\overline{c^{-}(k)}, where cI(k)>0c_{I}^{-}(k)>0 due to U′′>0U^{\prime\prime}>0. For any k1(k,k0]k_{1}\in(k_{-},k_{0}], 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all k[k1,k0]k\in[k_{1},k_{0}] and cU([h,0])c\in U([-h,0]). The conjugacy property (2.24) and the continuity of 𝐅\mathbf{F} (restricted to cI0c_{I}\geq 0) imply that there exists δ0>0\delta_{0}>0 such that 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 if k[k1,k0]k\in[k_{1},k_{0}] and dist(c,U([h,0]))δ0dist\big{(}c,U([-h,0])\big{)}\leq\delta_{0}. For any δ(0,δ0)\delta\in(0,\delta_{0}), let

Ω4={c|c12(U(h)+U(0)))|<δ+12(U(0)U(h)),dist(c,U([h,0]))>δ},\Omega_{4}=\big{\{}c\in\mathbb{C}\mid\big{|}c-\tfrac{1}{2}(U(-h)+U(0))\big{)}\big{|}<\delta+\tfrac{1}{2}(U(0)-U(-h)),\ dist\big{(}c,U([-h,0])\big{)}>\delta\big{\}},

and Ω4±\Omega_{4}^{\pm} be its (disjoint) upper and lower connected components with ±cI>0\pm c_{I}>0. Clearly c(k0)Ω4+c^{-}(k_{0})\in\Omega_{4}^{+}, c(k0)¯Ω4\overline{c^{-}(k_{0})}\in\Omega_{4}^{-}, c+(k)Ω4c^{+}(k)\notin\Omega_{4} and 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all k[k1,k0]k\in[k_{1},k_{0}] and cΩ4=Ω4+Ω4c\in\partial\Omega_{4}=\partial\Omega_{4}^{+}\cup\partial\Omega_{4}^{-} due to the semi-circle theorem and the choice of δ\delta. Therefore

Ind(𝐅(k,),Ω4±)=Ind(𝐅(k0,),Ω4±)=1,k[k1,k0].\text{Ind}(\mathbf{F}(k,\cdot),\Omega_{4}^{\pm})=\text{Ind}(\mathbf{F}(k_{0},\cdot),\Omega_{4}^{\pm})=1,\quad\forall k\in[k_{1},k_{0}].

From Lemma 2.7 and Remark 2.3, the simple root c(k0)c^{-}(k_{0}) of 𝐅(k0,)\mathbf{F}(k_{0},\cdot) can be extended as a simple root c(k)Ω4+c^{-}(k)\in\Omega_{4}^{+} of 𝐅(k,)\mathbf{F}(k,\cdot) for all k[k1,k0]k\in[k_{1},k_{0}], which is the only root of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω4+\Omega_{4}^{+}. Hence, by taking all k1(k,k0]k_{1}\in(k_{-},k_{0}], c(k)c^{-}(k) can be extended for all k(k,+)k\in(k_{-},+\infty) as the only root (counting the multiplicity) of 𝐅(k,)\mathbf{F}(k,\cdot) in Ω2c{cI>0}\Omega_{2}^{c}\cap\{c_{I}>0\}. From the semi-circle theorem and Theorem 1.1(1b), c±(k)c^{\pm}(k) and c(k)¯\overline{c^{-}(k)} are the only roots of 𝐅(k,)\mathbf{F}(k,\cdot) for |k|>k|k|>k_{-}, which are also simple.

Recall the branch c(k)c_{-}(k) of the root of 𝐅(k,)\mathbf{F}(k,\cdot) for k[kϵ,k+ϵ]k\in[-k_{-}-\epsilon,k_{-}+\epsilon] obtained in Theorem 1.1(2). The assumption U′′>0U^{\prime\prime}>0 yields Imc(k+ϵ)>0\text{Im}\,c_{-}(k_{-}+\epsilon)>0. Choose δ>0\delta>0 sufficiently small so that 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all k[k+ϵ,k0]k\in[k_{-}+\epsilon,k_{0}] and dist(c,U([h,0]))δdist\big{(}c,U([-h,0])\big{)}\leq\delta. Hence the semi-circle theorem implies c(k+ϵ)Ω4+c_{-}(k_{-}+\epsilon)\in\Omega_{4}^{+} where Ω4±\Omega_{4}^{\pm} is defined in the same form as in the above. While Ind(𝐅(k+ϵ,),Ω4+)=1(\mathbf{F}(k_{-}+\epsilon,\cdot),\Omega_{4}^{+})=1, both c(k+ϵ)c^{-}(k_{-}+\epsilon) and c(k+ϵ)c_{-}(k_{-}+\epsilon) are roots of 𝐅(k+ϵ,)\mathbf{F}(k_{-}+\epsilon,\cdot) in Ω4+\Omega_{4}^{+}. Therefore we obtain c(k)=c(k)c^{-}(k)=c_{-}(k) for k>kk>k_{-}.

A similar argument based on the index, continuation, and the Semi-circle Theorem starting at k=k+ϵk=k_{-}+\epsilon also yields that the roots of 𝐅(k,)\mathbf{F}(k,\cdot) for |kk|1|k-k_{-}|\ll 1 are either c+(k)c^{+}(k) or close to U(h)U(-h). Hence Theorem 1.1(2c) implies that c+(k),c(k)c^{+}(k),c_{-}(k)\in\mathbb{R} are all the roots for k[kϵ,k]k\in[k_{-}-\epsilon,k_{-}]. There exists β0>0\beta_{0}>0 such that 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all |k|kϵ|k|\leq k_{-}-\epsilon and dist(c,U([h,0]))β0dist\big{(}c,U([-h,0])\big{)}\leq\beta_{0}. Hence 𝐅(k,c)0\mathbf{F}(k,c)\neq 0 for all β(0,β0)\beta\in(0,\beta_{0}), |k|kϵ|k|\leq k_{-}-\epsilon, and cΩ5c\in\partial\Omega_{5} where

Ω5={ccR(c21,c1+1),|cI|U(0)U(h),dist(c,U([h,0]))>β},\Omega_{5}=\{c\in\mathbb{C}\mid c_{R}\in(c_{2}-1,c_{1}+1),\,|c_{I}|\leq U(0)-U(-h),\,dist\big{(}c,U([-h,0])\big{)}>\beta\},

and c1c_{1} and c2c_{2} have been given in (3.44) and (3.45), respectively. Clearly, for all |k|kϵ|k|\leq k_{-}-\epsilon, Ind(𝐅(k,),Ω5)=\big{(}\mathbf{F}(k,\cdot),\Omega_{5})=Ind(𝐅(kϵ,),Ω5)=2\big{(}\mathbf{F}(k_{-}-\epsilon,\cdot),\Omega_{5})=2 and the two roots are c+(k)c^{+}(k) and c(k)c_{-}(k). For |k|kϵ|k|\leq k_{-}-\epsilon, 𝐅(k,)\mathbf{F}(k,\cdot) has no other roots outside Ω5\Omega_{5} due to the semi-circle theorem and the choice of β\beta and c1,2c_{1,2}. Therefore c±(k)c^{\pm}(k) are also the only roots of 𝐅(k,)\mathbf{F}(k,\cdot) for all |k|k|k|\leq k_{-}. It completes the proof of the case of U′′>0U^{\prime\prime}>0.

Assume U′′0U^{\prime\prime}\leq 0 on (h,0)(-h,0). Firstly, we show the spectral stability following the most standard technique, which had actually been obtained in [28] (see also [10]). Suppose (k,c)(k,c) is an unstable mode (i.e. cI>0c_{I}>0) with the eigenfunction y(x2)y_{-}(x_{2}). Multiplying the Rayleigh equation by y(x2)¯\overline{y_{-}(x_{2})}, integrating by parts over (h,0)(-h,0), taking the imaginary part, and using the boundary condition of (1.5c), we have

cIh0U′′|y|2|Uc|2𝑑x2=\displaystyle c_{I}\int_{-h}^{0}\frac{U^{\prime\prime}|y_{-}|^{2}}{|U-c|^{2}}dx_{2}= Im(y(0)y(0)¯)=Im(U(0)(U(0)c)+g(U(0)c)2|y(0)|2)\displaystyle\text{Im}\big{(}y_{-}^{\prime}(0)\overline{y_{-}(0)}\big{)}=\text{Im}\Big{(}\frac{U^{\prime}(0)(U(0)-c)+g}{(U(0)-c)^{2}}|y_{-}(0)|^{2}\Big{)}
=\displaystyle= cI(U(0)|U(0)c|2+2g(U(0)cR)|U(0)c|4)|y(0)|2.\displaystyle c_{I}\Big{(}\frac{U^{\prime}(0)}{|U(0)-c|^{2}}+\frac{2g(U(0)-c_{R})}{|U(0)-c|^{4}}\Big{)}|y_{-}(0)|^{2}.

The semi-circle theorem implies cRU([h,0])c_{R}\in U([-h,0]) and thus Lemma 2.5(2) yields cI=0c_{I}=0 which is a contradiction. Hence the spectral stability follows. Due to Lemmas 2.5(2)(4) and 2.6(1), the only singular or non-singular modes inside the circle (1.8) have to be at inflection values of UU. The remaining statements in this case has been proved in Theorem 1.1. \square

Remark 3.4.

Suppose, instead of U>0U^{\prime}>0 on [h,0][-h,0], only U(0)>0U^{\prime}(0)>0 is assumed along with U′′0U^{\prime\prime}\leq 0 on [h,0][-h,0]. It is easy to see that Lemma 2.5(2) still holds for any cU([h,0])c\in\mathbb{C}\setminus U([-h,0]). The same above argument still applies to imply the spectral stability as already proved in [28].

Finally we prove Theorem 1.3 assuming UU has exactly one non-degenerate inflection point x20x_{20} on (h,0)(-h,0).

\bullet Proof of Theorem 1.3. Suppose (1.6) holds with U′′(x20)=0U^{\prime\prime}(x_{20})=0 and c0=U(x20)c_{0}=U(x_{20}). The same type of argument as above based on the index and the continuation will be employed extensively below on the semi-disk domain

Ω6={c(cR12(U(0)+U(h)))2+cI2<14(U(0)U(h))2,cI>0}.\Omega_{6}=\big{\{}c\in\mathbb{C}\mid\big{(}c_{R}-\tfrac{1}{2}(U(0)+U(-h))\big{)}^{2}+c_{I}^{2}<\tfrac{1}{4}(U(0)-U(-h))^{2},\ c_{I}>0\}.

Let us start with the case of U′′′(x20)>0U^{\prime\prime\prime}(x_{20})>0 which implies U′′Uc0>0\frac{U^{\prime\prime}}{U-c_{0}}>0 on [h,0][-h,0]. Lemma 2.3 yields y(k,c0,x2)y_{-}(k,c_{0},x_{2})\in\mathbb{R} over [h,0][-h,0] for any kk\in\mathbb{R}. Multiplying the Rayleigh equation by yy_{-}, integrating by parts over (h,0)(-h,0), and using U′′Uc0>0\frac{U^{\prime\prime}}{U-c_{0}}>0, we obtain y(k,c0,0)>0y_{-}(k,c_{0},0)>0 for all kk\in\mathbb{R}. According to Lemma 3.7, there exists k0>0k_{0}>0, unique among k0k\geq 0, such that 𝐅(k0,c0)=0\mathbf{F}(k_{0},c_{0})=0. Along with the Semi-circle Theorem111From the standard proof of the Semi-circle Theorem, one could see that unstable modes can not occur on the boundary semi-circle of (1.8) when U>0U^{\prime}>0 on [h,0][-h,0]., Lemmas 2.5(4) and 3.3, it implies that (±k,U(h))(\pm k_{-},U(-h)) and (±k0,c0)(\pm k_{0},c_{0}) are the only roots of 𝐅\mathbf{F} with cΩ6c\in\partial\Omega_{6}. On the one hand, since U′′<0U^{\prime\prime}<0 near x2=hx_{2}=-h, Lemma 3.11 implies that there are never singular or non-singular modes with cΩ6c\in\Omega_{6} near U(h)U(-h) for any kk\in\mathbb{R}. Hence the index Ind(𝐅(k,),Ω6)(\mathbf{F}(k,\cdot),\Omega_{6}) remains a constant for k>k0k>k_{0}. On the other hand, since U′′>0U^{\prime\prime}>0 near x2=0x_{2}=0, Lemma 3.15 implies that (k,c(k))(k,c^{-}(k)) with cI(k)>0c_{I}^{-}(k)>0 is the only unstable mode for k1k\gg 1, which implies Ind(𝐅(k,),Ω6)=1(\mathbf{F}(k,\cdot),\Omega_{6})=1 for all k>k0k>k_{0}. Again from Lemma 3.11, there exists a branch of unstable modes (k,𝒞(k))(k,\mathcal{C}(k)) with 𝒞(k0)=c0\mathcal{C}(k_{0})=c_{0} and 𝒞I(k)>0\mathcal{C}_{I}(k)>0 for 1kk0>01\gg k-k_{0}>0, which are the only singular or non-singular modes near (k0,c0)(k_{0},c_{0}) with cΩ6c\in\Omega_{6}. From Lemma 2.7, this branch (k,𝒞(k))(k,\mathcal{C}(k)) can be continued for all kk0k\geq k_{0}. Since 𝐅\mathbf{F} has exactly one root with cΩ6c\in\Omega_{6} for k>k0k>k_{0}, we obtain that the continuations of 𝒞(k)\mathcal{C}(k) and c(k)c^{-}(k) must coincide and form the only singular or non-singular modes with cΩ6¯c\in\overline{\Omega_{6}}. The local bifurcation given by Lemma 3.11 implies that Ind(𝐅(k,),Ω6)=0(\mathbf{F}(k,\cdot),\Omega_{6})=0 for |k|<k0|k|<k_{0} and thus there are no unstable modes if |k|<k0|k|<k_{0}. Along with Theorem 1.1(2b), it completes the proof of this case.

In the rest of the proof we assume U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0. Lemma 3.6 yields k0>kk_{0}>k_{-}. Much as the above, the index Ind(𝐅(k,),Ω6)(\mathbf{F}(k,\cdot),\Omega_{6}) remains a constant except for k{±k,±k0,±k1}k\in\{\pm k_{-},\pm k_{0},\pm k_{1}\} (no k1k_{1} if it does not exist). Note (1.6) and U′′′(x20)<0U^{\prime\prime\prime}(x_{20})<0 implies U′′<0U^{\prime\prime}<0 on (x20,0)(x_{20},0) and U′′>0U^{\prime\prime}>0 on (h,0)(-h,0). From Lemmas 3.15

(3.46) Ind(𝐅(k,),Ω6)=0, if |k|>k0.\text{Ind}(\mathbf{F}(k,\cdot),\Omega_{6})=0,\quad\text{ if }\;|k|>k_{0}.

By a similar argument to that in the proof of Theorem 1.2, Lemma 3.11, which gives all the unstable modes near the singular neutral modes at c0c_{0} and U(h)U(-h), implies that the index increases by 1 as kk increases through kk_{-} and decreases by 1 as kk increases through k0k_{0}, as well as through k1k_{1} if k1k_{1} exists. Therefore we have

(3.47) Ind(𝐅(k,),Ω6)=1, if |k|(max{k,k1},k0),\text{Ind}(\mathbf{F}(k,\cdot),\Omega_{6})=1,\quad\text{ if }\;|k|\in(\max\{k_{-},k_{1}\},k_{0}),

where max{k,k1}=k\max\{k_{-},k_{1}\}=k_{-} is understood if k1k_{1} does not exist. In the case k1k_{1} does not exist, we also have

(3.48) Ind(𝐅(k,),Ω6)=0,|k|<k,\text{Ind}(\mathbf{F}(k,\cdot),\Omega_{6})=0,\qquad|k|<k_{-},

and otherwise

(3.49) Ind(𝐅(k,),Ω6)={2,|k|(k,k1),0,|k|(k1,k),1,|k|<min{k,k1}.\text{Ind}(\mathbf{F}(k,\cdot),\Omega_{6})=\begin{cases}2,\quad&|k|\in(k_{-},k_{1}),\\ 0,\quad&|k|\in(k_{1},k_{-}),\\ 1,&|k|<\min\{k_{-},k_{1}\}.\end{cases}

The linearized system (1.3) is unstable when the above index is greater than zero. By a similar argument to that in the proof of Theorem 1.2 bases on Lemma 2.7 and the index, the branches of unstable modes bifurcating from the singular neutral modes can be continued in kk unless kk approaches ±k\pm k_{-}, ±k0\pm k_{0}, and ±k1\pm k_{1} or it collides with another branch, the latter of which happens only possibly for |k|(k1,k)|k|\in(k_{1},k_{-}). Let us consider the different cases (2)–(4) in the theorem separately.

Suppose k1k_{1} does not exist and thus ±k0\pm k_{0} are the only wave numbers which make c0c_{0} a singular neutral mode. From Lemma 2.7 and (3.47), the branch of unstable modes c(k)c_{-}(k) obtained in Theorem 1.1(2) (or Lemma 3.11 more directly bifurcating from (±k,U(h))(\pm k_{-},U(-h))) can be continued for all |k|<k0|k|<k_{0}. Due to (3.47), this branch has to coincide with the branch 𝒞(k)\mathcal{C}(k) bifurcating from (±k0,c0)(\pm k_{0},c_{0}) (again by Lemma 3.11) for |k|k0|k|\leq k_{0} and with c(±k0)=c0c_{-}(\pm k_{0})=c_{0}. From the Semi-circle Theorem and (3.48), (k,c(k))(k,c_{-}(k)) is the only singular or non-singular modes with cΩ6¯c\in\overline{\Omega_{6}} and statement (2) is proved.

Suppose k1k_{1} exists and k1kk_{1}\leq k_{-}. By continuity, the branch c(k)c_{-}(k) satisfies c(k)c0c_{-}(k)\neq c_{0} for 1|k|k>01\gg|k|-k_{-}>0. The same argument as above also yields that the branch c(k)c_{-}(k) connects to c0=c(±k0)c_{0}=c_{-}(\pm k_{0}) and is the only singular or non-singular modes with cΩ6¯c\in\overline{\Omega_{6}} and |k|k|k|\geq k_{-}. Due to (3.49), there are no singular or non-singular modes with cΩ6¯c\in\overline{\Omega_{6}} and |k|(k1,k)|k|\in(k_{1},k_{-}). From Lemmas 3.11 and 2.7, both branches of unstable modes bifurcating from (±k1,c0)(\pm k_{1},c_{0}) for |k|<k1|k|<k_{1} can be continued for all |k|<k1|k|<k_{1} and have to coincide due to (3.49). This branch is apparently even in kk and the only singular or non-singular modes with cΩ6¯c\in\overline{\Omega_{6}} and |k|<k1|k|<k_{1}. It completes the proof of statement (3).

To prove the last case, suppose k1k_{1} exists and k1>kk_{1}>k_{-}. From the Semi-circle Theorem and the assumptions, it is clear

𝒮={(k,c)cΩ6¯,|k|k0,𝐅(k,c)=0},𝒮¯=𝒮.\mathcal{S}=\{(k,c)\mid c\in\overline{\Omega_{6}},\ |k|\leq k_{0},\ \mathbf{F}(k,c)=0\},\quad\overline{\mathcal{S}}=\mathcal{S}.

\bullet Claim. For any connected component 𝒮~\tilde{\mathcal{S}} of 𝒮\mathcal{S}, it holds

k=sup{kcΩ6¯,(k,c)𝒮~}{k,k1,k0}&c{c0,U(h)},(k,c)𝒮~,k^{*}=\sup\{k\mid\exists c\in\overline{\Omega_{6}},\ (k,c)\in\tilde{\mathcal{S}}\}\in\{-k_{-},k_{1},k_{0}\}\;\;\&\;\;\exists c^{*}\in\{c_{0},U(-h)\},\ (k^{*},c^{*})\in\tilde{\mathcal{S}},
k=inf{kcΩ6¯,(k,c)𝒮~}{k,k1,k0}&c{c0,U(h)},(k,c)𝒮~.k_{*}=\inf\{k\mid\exists c\in\overline{\Omega_{6}},\ (k,c)\in\tilde{\mathcal{S}}\}\in\{k_{-},-k_{1},-k_{0}\}\;\;\&\;\;\exists c_{*}\in\{c_{0},U(-h)\},\ (k_{*},c_{*})\in\tilde{\mathcal{S}}.

We shall prove the infimum case and the supremum case is similar. In fact, since the compactness of 𝒮\mathcal{S} yields that of 𝒮~\tilde{\mathcal{S}}, there exists cΩ6¯c_{*}\in\overline{\Omega_{6}} such that (k,c)𝒮~(k_{*},c_{*})\in\tilde{\mathcal{S}}. If cΩ6c_{*}\in\Omega_{6} where 𝐅\mathbf{F} is analytic, then cc_{*} is an isolated zero of 𝐅(k,)\mathbf{F}(k_{*},\cdot). Since, as kk varies locally, the set of zero points of analytic functions can always be extended continuously beyond kk_{*} (even though splitting may occur if cc_{*} is not simple), it contradicts with the definition of kk_{*}. Therefore cΩ6c_{*}\in\partial\Omega_{6} and thus k{±k,±k1,±k0}k_{*}\in\{\pm k_{-},\pm k_{1},\pm k_{0}\}. Lemma 3.11 implies that (k0,c0)(k_{0},c_{0}) is the end point of a branch of unstable modes with k<k0k<k_{0}, therefore k0k_{0} can not be the infimum. Another two possibilities (k,c)=(k1,c0)(k_{*},c_{*})=(k_{1},c_{0}) or (k,c)=(k,U(h))(k_{*},c_{*})=(-k_{-},U(-h)) can be excluded in the same way as (k0,c0)(k_{0},c_{0}). The claim is proved.

Let 𝒮1\mathcal{S}_{1} be the connected component of 𝒮\mathcal{S} containing (k,U(h))(k_{-},U(-h)) and 𝒮1={(k,c)(k,c)𝒮1}\mathcal{S}_{-1}=\{(-k,c)\mid(k,c)\in\mathcal{S}_{1}\}. From (3.49),

(3.50) |c0Ω6 s. t. (0,c0)𝒮.\exists|c^{0}\in\Omega_{6}\;\text{ s. t. }\;(0,c^{0})\in\mathcal{S}.

Applying the above claim to the supremum to 𝒮1\mathcal{S}_{1}, there are two possibilities.

#\# Case A. k1=sup(k,c)𝒮1kk_{1}=\sup_{(k,c)\in\mathcal{S}_{1}}k and thus (k1,c0)𝒮1(k_{1},c_{0})\in\mathcal{S}_{1}. In this case, let 𝒮0\mathcal{S}_{0} be the connected component of 𝒮\mathcal{S} containing (k0,c0)(k_{0},c_{0}). As k1k_{1} is the supremum of 𝒮1\mathcal{S}_{1} and k1<k0k_{1}<k_{0}, we have 𝒮0𝒮1=\mathcal{S}_{0}\cap\mathcal{S}_{1}=\emptyset. This implies that inf(k,c)𝒮0kk\inf_{(k,c)\in\mathcal{S}_{0}}k\neq k_{-}, otherwise (k,U(h))𝒮0(k_{-},U(-h))\in\mathcal{S}_{0} which would yield 𝒮0=𝒮1\mathcal{S}_{0}=\mathcal{S}_{1}. Again the above claim implies either (k1,c0)𝒮0(-k_{1},c_{0})\in\mathcal{S}_{0} or (k0,c0)𝒮0(-k_{0},c_{0})\in\mathcal{S}_{0}. From (3.50) and the connectedness of 𝒮0\mathcal{S}_{0}, it holds (0,c0)𝒮0(0,c^{0})\in\mathcal{S}_{0}. By the even symmetry of 𝐅\mathbf{F} in kk, immediately we derive the same even symmetry of 𝒮0\mathcal{S}_{0} about kk and thus (k0,c0)𝒮0(-k_{0},c_{0})\in\mathcal{S}_{0}. Since all points (k,c)𝒮(k,c)\in\mathcal{S} with c{c0,U(h)}c\in\{c_{0},U(-h)\} belong to 𝒮1𝒮0𝒮1\mathcal{S}_{-1}\cup\mathcal{S}_{0}\cup\mathcal{S}_{1}, it follows from the above claim that these three sets are all the connected component of 𝒮\mathcal{S} and thus 𝒮=𝒮1𝒮0𝒮1\mathcal{S}=\mathcal{S}_{-1}\cup\mathcal{S}_{0}\cup\mathcal{S}_{1}. The desired statement (4) in case (4b) is obtained.

#\# Case B. k0=sup(k,c)𝒮1kk_{0}=\sup_{(k,c)\in\mathcal{S}_{1}}k and thus (k0,c0)𝒮1(k_{0},c_{0})\in\mathcal{S}_{1}. Accordingly we let 𝒮0\mathcal{S}_{0} be the connected component of 𝒮\mathcal{S} containing (k1,c0)(k_{1},c_{0}). Again we consider the two possibilities separately.

* Case B.1. (k1,c0)𝒮1(k_{1},c_{0})\in\mathcal{S}_{1}, too. In this case, we have 𝒮0=𝒮1\mathcal{S}_{0}=\mathcal{S}_{1}. Again since all points (k,c)𝒮(k,c)\in\mathcal{S} with c{c0,U(h)}c\in\{c_{0},U(-h)\} belong to 𝒮1𝒮1\mathcal{S}_{-1}\cup\mathcal{S}_{1}, it follows from the above claim (0,c0)𝒮=𝒮1𝒮1(0,c^{0})\in\mathcal{S}=\mathcal{S}_{-1}\cup\mathcal{S}_{1}. Due to the symmetry, we obtain (0,c0)𝒮1𝒮1(0,c^{0})\in\mathcal{S}_{1}\cap\mathcal{S}_{-1} and thus 𝒮=𝒮1=𝒮0=𝒮1\mathcal{S}=\mathcal{S}_{-1}=\mathcal{S}_{0}=\mathcal{S}_{1}. Hence the desired statement (4) in case (4b) holds.

* Case B.2. (k1,c0)𝒮1(k_{1},c_{0})\notin\mathcal{S}_{1}, which yields 𝒮0𝒮1=\mathcal{S}_{0}\cap\mathcal{S}_{1}=\emptyset. The above claim implies inf(k,c)𝒮0k<k\inf_{(k,c)\in\mathcal{S}_{0}}k<k_{-} and thus (k1,c0)𝒮0(-k_{1},c_{0})\in\mathcal{S}_{0} or (k0,c0)𝒮0(-k_{0},c_{0})\in\mathcal{S}_{0}. The desired statement (4) in case (4a) follows from the same argument as in the above Case A. The proof of the theorem is complete. \square

According to Lemma 3.7, if kC>0k_{C}>0 exists, i.e. UU is unstable for the channel flow, then k1>0k_{1}>0 exists if g>F0(c0)g>F^{0}(c_{0}) defined in (3.12). From Lemma 3.12 which is under a weaker condition, the closed branch 𝒞(k)\mathcal{C}(k) of unstable modes in case (3) bifurcates from c0c_{0}. As this branch grows, we can not exclude the possibility that it intersects with the branch c(k)c_{-}(k) connecting (k,U(h))(k_{-},U(-h)) and (k0,c0)(k_{0},c_{0}) obtained in case (2).

3.8. The linearization of the capillary gravity water waves at monotonic shear flow revisited

When the surface tension is also considered, the boundary condition (1.1d) is replaced by

(3.51) p(t,x)=σκ(t,x),xSt,p(t,x)=\sigma\kappa(t,x),\qquad x\in S_{t},

where σ>0\sigma>0, κ(t,x)=ηx1x1(1+ηx12)32\kappa(t,x)=-\frac{\eta_{x_{1}x_{1}}}{(1+\eta_{x_{1}}^{2})^{\frac{3}{2}}} is the mean curvature of StS_{t} at xx. Shear flow v(x)=(U(x2),0)Tv(x)=(U(x_{2}),0)^{T} under the flat surface St={x2=0}S_{t}=\{x_{2}=0\} is also a stationary solution. The linearization of the capillary gravity water waves at such a shear flow is given by a system similar to (1.3) only with the linearized boundary condition (1.3d) replaced accordingly by

(3.52) p(t,x)=gησx22η, at x2=0.p(t,x)=g\eta-\sigma\partial_{x_{2}}^{2}\eta,\qquad\;\text{ at }x_{2}=0.

Following the same procedure, we obtain that (k,c)(k,c) corresponds to an eigenvalue ikc-ikc in the kk-th Fourier modes of x1x_{1} if and only if σ(k,c)=0\mathcal{F}_{\sigma}(k,c)=0, or equivalently, 𝐅σ(k,c)=0\mathbf{F}_{\sigma}(k,c)=0, where σ\mathcal{F}_{\sigma} was defined (3.20) and 𝐅σ(k,c)=y(k,c,0)σ(k,c)\mathbf{F}_{\sigma}(k,c)=y_{-}(k,c,0)\mathcal{F}_{\sigma}(k,c).

In [17], we obtained results on the eigenvalue distribution and the inviscid damping of the linearized capillary gravity waves. In particular, the semi-circle theorem still holds for unstable modes, namely, all non-singular modes cc outside the circle (1.8) belong to U((h,0))\mathbb{R}\setminus U((-h,0)). Moreover, there exists k0>0k_{0}>0 such that 𝐅σ(k,)\mathbf{F}_{\sigma}(k,\cdot) has exactly two roots c±(k)c^{\pm}(k) for |k|k0|k|\geq k_{0}, which are simple roots, even and analytic in kk, and satisfy

c+(k)>U(0),c(k)<U(h),lim|k|c±(k)/σ|k|=±1.c^{+}(k)>U(0),\quad c^{-}(k)<U(-h),\quad\lim_{|k|\to\infty}c^{\pm}(k)/\sqrt{\sigma|k|}=\pm 1.

The branch c+(k)c^{+}(k) can be continued as simple roots for each kk\in\mathbb{R} with c+(k)>U(0)c^{+}(k)>U(0) and even and analytic in kk. The continuation of the other branch c(k)c^{-}(k) may or may not reach U(h)U(-h) depending on whether

g<g#(σ)\displaystyle g<g_{\#}(\sigma)\triangleq max{Y(k,U(h))(U(0)U(h))2U(0)(U(0)U(h))σk2k}\displaystyle\max\big{\{}Y\big{(}k,U(-h)\big{)}\big{(}U(0)-U(-h)\big{)}^{2}-U^{\prime}(0)\big{(}U(0)-U(-h)\big{)}-\sigma k^{2}\mid k\in\mathbb{R}\big{\}}
=\displaystyle= max{σ(k,U(h))+gk}σ(0,U(h))+g=0.\displaystyle\max\big{\{}\mathcal{F}_{\sigma}\big{(}k,U(-h)\big{)}+g\mid k\in\mathbb{R}\big{\}}\geq\mathcal{F}_{\sigma}(0,U(-h))+g=0.

If it does, the bifurcation at c=U(h)c=U(-h) was analyzed under the assumption U′′0U^{\prime\prime}\neq 0 on [h,0][-h,0], which was only used to ensure a.) the C1,αC^{1,\alpha} regularity of σ\mathcal{F}_{\sigma} near c=U(h)c=U(-h) and b.) the sign cσ(k0,U(h))<0\partial_{c}\mathcal{F}_{\sigma}(k_{0},U(-h))<0 if σ(k0,U(h))=0\mathcal{F}_{\sigma}(k_{0},U(-h))=0. More details can be found in Theorem 1.1 in [17]. Thanks to Corollary 3.1.1 and Lemma 3.10, these two key properties of σ\mathcal{F}_{\sigma} also hold without assuming U′′0U^{\prime\prime}\neq 0. Therefore the following proposition on the eigenvalue distribution of the linearized capillary gravity waves holds with weakened assumptions.

Proposition 3.16.

Assume UC6U\in C^{6} and U>0U^{\prime}>0 on [h,0][-h,0], then the following hold.

  1. (1)

    If g=g#(σ)g=g_{\#}(\sigma), then there exist k#>0k_{\#}>0, unique in [0,+)[0,+\infty), and an even C1,αC^{1,\alpha} function (for any α[0,1)\alpha\in[0,1)) c(k)c^{-}(k) defined for all kk\in\mathbb{R} such that

    1. (a)

      σ(k,U(h))=0\mathcal{F}_{\sigma}(k,U(-h))=0 iff k=±k#k=\pm k_{\#};

    2. (b)

      c(k)<U(h)c^{-}(k)<U(-h) is analytic in k±k#k\neq\pm k_{\#}, and c(±k#)=U(h)c^{-}(\pm k_{\#})=U(-h);

    3. (c)

      c+(k)c^{+}(k) and c(k)c^{-}(k) are simple roots of σ(k,)\mathcal{F}_{\sigma}(k,\cdot) which include all roots of σ(k,)\mathcal{F}_{\sigma}(k,\cdot) outside the disk (1.8); and

    4. (d)

      there exists ϵ,ρ>0\epsilon,\rho>0 that σ(k,c)=0\mathcal{F}_{\sigma}(k,c)=0 with |k±k#|ϵ|k\pm k_{\#}|\leq\epsilon and |cU(h)|ρ|c-U(-h)|\leq\rho iff c=c(k)c=c^{-}(k).

  2. (2)

    If g<g#(σ)g<g_{\#}(\sigma), then there exist k#+>k#>0k_{\#}^{+}>k_{\#}^{-}>0, ϵ,ρ>0\epsilon,\rho>0 and even C1,αC^{1,\alpha} functions (for any α[0,1)\alpha\in[0,1)) c(k)c^{-}(k) defined for |k|k#+ϵ|k|\geq k_{\#}^{+}-\epsilon and c(k)c_{-}(k) for |k|k#+ϵ|k|\leq k_{\#}^{-}+\epsilon satisfying the following.

    1. (a)

      σ(k,U(h))=0\mathcal{F}_{\sigma}(k,U(-h))=0 iff k=±k#±k=\pm k_{\#}^{\pm}.

    2. (b)

      c(k)<U(h)c^{-}(k)<U(-h) is analytic in kk if |k|>k#+|k|>k_{\#}^{+} and c(k)<U(h)c_{-}(k)<U(-h) is analytic in kk if |k|<k#|k|<k_{\#}^{-}. Moreover, c+(k)c^{+}(k), c(k)c^{-}(k) with |k|>k#+|k|>k_{\#}^{+}, and c(k)c_{-}(k) with |k|<k#|k|<k_{\#}^{-} are the only roots of σ\mathcal{F}_{\sigma} outside the disk (1.8), which are all simple.

    3. (c)

      c(±k#+)=U(h)=c(±k#)c^{-}(\pm k_{\#}^{+})=U(-h)=c_{-}(\pm k_{\#}^{-}), (cR)(±k#+)>0\mp(c_{R}^{-})^{\prime}(\pm k_{\#}^{+})>0, ±Rec(±k#)>0\pm\text{Re}\,c_{-}^{\prime}(\pm k_{\#}^{-})>0.

    4. (d)

      cI(k)c_{I}^{-}(k) with 0k#+|k|ϵ0\leq k_{\#}^{+}-|k|\leq\epsilon and Imc(k)\text{Im}\,c_{-}(k) with 0|k|k#ϵ0\leq|k|-k_{\#}^{-}\leq\epsilon have the same sign (+,+,-, or 0) as U′′(U1(cR(k)))U^{\prime\prime}\big{(}U^{-1}(c_{R}^{-}(k))\big{)} and U′′(U1(Rec(k)))U^{\prime\prime}\big{(}U^{-1}(\text{Re}\,c_{-}(k))\big{)}, respectively, and for such kk,

      C1|U′′(U1(cR(k)))|(|k|k#+)2|cI(k)|C|U′′(U1(cR(k)))|(|k|k#+)2,C^{-1}\big{|}U^{\prime\prime}(U^{-1}(c_{R}^{-}(k)))\big{|}(|k|-k_{\#}^{+})^{2}\leq|c_{I}^{-}(k)|\leq C\big{|}U^{\prime\prime}(U^{-1}(c_{R}^{-}(k)))\big{|}(|k|-k_{\#}^{+})^{2},
      C1|U′′(U1(Rec(k)))|(|k|k#)2|Imc(k)|C|U′′(U1(Rec(k)))|(|k|k#)2.C^{-1}\big{|}U^{\prime\prime}(U^{-1}(\text{Re}\,c_{-}(k)))\big{|}(|k|-k_{\#}^{-})^{2}\leq|\text{Im}\,c_{-}(k)|\leq C\big{|}U^{\prime\prime}(U^{-1}(\text{Re}\,c_{-}(k)))\big{|}(|k|-k_{\#}^{-})^{2}.
    5. (e)

      σ(k,c=cR+icI)=0\mathcal{F}_{\sigma}(k,c=c_{R}+ic_{I})=0 with ||k|k#±|ϵ\big{|}|k|-k_{\#}^{\pm}\big{|}\leq\epsilon, |cRU(h)|ρ|c_{R}-U(-h)|\leq\rho, cI0c_{I}\geq 0 iff c=c(k)c=c^{-}(k) or c=c(k)c=c_{-}(k).

Remark 3.5.

a.) The bifurcation near inflection values of UU can be found in Lemma 3.11.
b.) If g>g#(σ)g>g_{\#}(\sigma), it had been proved in [17] that c(k)c^{-}(k) can be extended for all kk\in\mathbb{R} as a simple root of σ(k,c)˙\mathcal{F}_{\sigma}(k,c\dot{)} which is the only root other than c+(k)c^{+}(k) outside the disk (1.8)). It also satisfies c(k)<U(h)c^{-}(k)<U(-h). An explicit necessary and sufficient condition for g#(σ)=0g_{\#}(\sigma)=0 was also given there. Under the assumption U′′0U^{\prime\prime}\neq 0 a complete picture of the eigenvalue distribution in the fashion of Theorem 1.1(5) was also obtained in [17], where the continuations of c(k)c^{-}(k) and c(k)c_{-}(k) coincide.
c.) Following from the same proof as in statement (6) of Theorem 1.1, one can prove that the spectral stability of the linearized capillary gravity waves.

Proof.

Due to the concavity of σ\mathcal{F}_{\sigma} in K=k2K=k^{2} (Lemma 2.6(2)), for gg#(σ)g\leq g_{\#}(\sigma), the existence of k#>0k_{\#}>0 or k#±>0k_{\#}^{\pm}>0 as the only zero points of σ(,U(h))\mathcal{F}_{\sigma}(\cdot,U(-h)) follows from the definition of g#(σ)g_{\#}(\sigma). Much in the proof of Theorem 1.1(2), the branch c(k)c_{-}(k) starts with the simple root c(0)=c2<U(h)c_{-}(0)=c_{2}<U(-h) of σ(0,)\mathcal{F}_{\sigma}(0,\cdot). For |k|1|k|\gg 1, another branch c(k)<U(h)c^{-}(k)<U(-h) had been constructed in Theorem 1.1 in [17]. Applying Lemma 2.7 and the semi-circle theorem, c(k)c_{-}(k) and c(k)c^{-}(k) can be continued in (,U(h))(-\infty,U(-h)) as the only roots of σ(k,)\mathcal{F}_{\sigma}(k,\cdot) outside the disk (1.8) for all |k|<k#|k|<k_{\#}^{-} and |k|>k#+|k|>k_{\#}^{+}, respectively (k#±=k#k_{\#}^{\pm}=k_{\#} is understood if g=g#(σ)g=g_{\#}(\sigma)), where σ(k,U(h))=0\mathcal{F}_{\sigma}(k,U(-h))=0 for the first kk in the continuation. It suffices to prove c(k#)=c(k#+)=U(h)c_{-}(k_{\#}^{-})=c^{-}(k_{\#}^{+})=U(-h) and the rest of the proposition would follow from the same proof as of Theorem 1.1.

Our strategy is to vary the parameter σ0\sigma\geq 0 starting from 0. From the definitions of σ\mathcal{F}_{\sigma} and g#()g_{\#}(\cdot), if g#(σ)>0g_{\#}(\sigma^{\prime})>0, then it is strictly decreasing in σ\sigma^{\prime} and there exists σ0σ\sigma_{0}\geq\sigma such that g#(σ0)=gg_{\#}(\sigma_{0})=g. We use the notations k#±(σ)k_{\#}^{\pm}(\sigma^{\prime}) for σ(0,σ0)\sigma^{\prime}\in(0,\sigma_{0}) and k#k_{\#} for g=g#(σ)g=g_{\#}(\sigma^{\prime}) which occurs only when σ=σ0\sigma^{\prime}=\sigma_{0}. From the concavity of σ(,U(h))\mathcal{F}_{\sigma^{\prime}}(\cdot,U(-h)) in K=k2K=k^{2}, we have

σk#±(σ)=±(σσ/kσ)(k#±(σ),U(h))>0,σ(0,σ0).\mp\partial_{\sigma}k_{\#}^{\pm}(\sigma^{\prime})=\pm(\partial_{\sigma}\mathcal{F}_{\sigma^{\prime}}/\partial_{k}\mathcal{F}_{\sigma^{\prime}})(k_{\#}^{\pm}(\sigma^{\prime}),U(-h))>0,\quad\sigma^{\prime}\in(0,\sigma_{0}).

They also clearly satisfy

k#+(σ)>k#>k#(σ)>0,limσσ0k#±(σ)=k#(σ0),limσ0+k#(σ)=k,limσ0+k#+(σ)=+,k_{\#}^{+}(\sigma^{\prime})>k_{\#}>k_{\#}^{-}(\sigma^{\prime})>0,\;\;\lim_{\sigma^{\prime}\to\sigma_{0}-}k_{\#}^{\pm}(\sigma^{\prime})=k_{\#}(\sigma_{0}),\;\;\lim_{\sigma^{\prime}\to 0+}k_{\#}^{-}(\sigma^{\prime})=k_{-},\;\;\lim_{\sigma^{\prime}\to 0+}k_{\#}^{+}(\sigma^{\prime})=+\infty,

where k>0k_{-}>0 is given in Theorem 1.1(2) satisfying F(k,U(h))=0F(k_{-},U(-h))=0.

Let us use c(σ,k)c^{-}(\sigma^{\prime},k) and c(σ,k)c_{-}(\sigma^{\prime},k) to denote the branches of the roots in (,U(h))(-\infty,U(-h)) of σ(k,)\mathcal{F}_{\sigma^{\prime}}(k,\cdot) for σ(0,σ0]\sigma^{\prime}\in(0,\sigma_{0}], while we skip σ\sigma^{\prime} if there is no confusion. According to Theorem 1.1(2), c(0,0)<U(h)c_{-}(0,0)<U(-h) can be extended as the only root c(0,k)c_{-}(0,k) in (,U(h)](-\infty,U(-h)] of 0(k,)=F(k,)\mathcal{F}_{0}(k,\cdot)=F(k,\cdot) for all |k|<k|k|<k_{-} and limk(k)c(0,k)=U(h)\lim_{k\to(k_{-})-}c_{-}(0,k)=U(-h). The continuation in the direction of σ\sigma^{\prime} through the Implicit Function Theorem applied in a neighborhood of U(h)U(-h) implies that, for each σ[0,σ0)\sigma^{\prime}\in[0,\sigma_{0}), c(σ,k)c_{-}(\sigma^{\prime},k) can be extended until it reaches U(h)U(-h) which has to occur at k=k#(σ)k=k_{\#}^{-}(\sigma^{\prime}).

Let

Ω={ccR(,U(h)),cI(1,1)}.\Omega=\{c\in\mathbb{C}\mid c_{R}\in(-\infty,U(-h)),\,c_{I}\in(-1,1)\}.

Lemma 3.2(2) implies that σ(k,)\mathcal{F}_{\sigma^{\prime}}(k,\cdot) has no roots in Ω\Omega for cR1-c_{R}\gg 1. For any σ(0,σ0)\sigma^{\prime}\in(0,\sigma_{0}), the only roots of σ\mathcal{F}_{\sigma^{\prime}} on ×Ω\mathbb{R}\times\partial\Omega are (k#±(σ),U(h))(k_{\#}^{\pm}(\sigma),U(-h)) due to the semi-circle theorem. Hence from an index argument applied to sufficiently large compact subsets of Ω\Omega, the total number of roots (counting the multiplicity) of σ(k,)\mathcal{F}_{\sigma^{\prime}}(k,\cdot) in Ω\Omega are constants for kk in [0,k#(σ))[0,k_{\#}^{-}(\sigma^{\prime})), (k#(σ),k#+(σ))(k_{\#}^{-}(\sigma^{\prime}),k_{\#}^{+}(\sigma^{\prime})), and (k#+(σ),+)(k_{\#}^{+}(\sigma^{\prime}),+\infty), respectively. Since c(0,0)c_{-}(0,0) is the only root in Ω\Omega for (σ,k)=(0,0)(\sigma^{\prime},k)=(0,0) and c(σ,k)c^{-}(\sigma^{\prime},k) the only one for σ(0,σ0)\sigma^{\prime}\in(0,\sigma_{0}) and k1k\gg 1, the total numbers of roots of σ(k,)\mathcal{F}_{\sigma^{\prime}}(k,\cdot) in Ω\Omega is 11 for kk in both [0,k#(σ))[0,k_{\#}^{-}(\sigma^{\prime})) and (k#+(σ),+)(k_{\#}^{+}(\sigma^{\prime}),+\infty). As c(σ,k#(σ))=U(h)c_{-}(\sigma^{\prime},k_{\#}^{-}(\sigma^{\prime}))=U(-h), by the bifurcation analysis based on Lemma 3.11 and the signs cσ(k#(σ),U(h))<0\partial_{c}\mathcal{F}_{\sigma^{\prime}}(k_{\#}^{-}(\sigma^{\prime}),U(-h))<0 (Lemma 3.10) and kσ(k#(σ),U(h))>0\partial_{k}\mathcal{F}_{\sigma^{\prime}}(k_{\#}^{-}(\sigma^{\prime}),U(-h))>0 (from the concavity of σ\mathcal{F}_{\sigma^{\prime}} in K=k2K=k^{2}), σ(k,)\mathcal{F}_{\sigma^{\prime}}(k,\cdot) has no roots in Ω\Omega for 1kk#(σ)>01\gg k-k_{\#}^{-}(\sigma^{\prime})>0. Hence σ(k,)\mathcal{F}_{\sigma^{\prime}}(k,\cdot) has no roots in Ω\Omega for k(k#(σ),k#+(σ))k\in(k_{\#}^{-}(\sigma^{\prime}),k_{\#}^{+}(\sigma^{\prime})). For the total number of roots in Ω\Omega of σ(k,)\mathcal{F}_{\sigma^{\prime}}(k,\cdot) to change from 0 to 11 as kk increases through k#+(σ)k_{\#}^{+}(\sigma^{\prime}), it must hold c(σ,k#+(σ))=U(h)c^{-}(\sigma^{\prime},k_{\#}^{+}(\sigma^{\prime}))=U(-h), otherwise c(σ,k#+(σ))c^{-}(\sigma^{\prime},k_{\#}^{+}(\sigma^{\prime})) would be continued to a root in Ω\Omega for k<k#+(σ)k<k_{\#}^{+}(\sigma^{\prime}) which is a contradiction.

Finally, c(σ0,k#)=c(σ0,k#)=U(h)c^{-}(\sigma_{0},k_{\#})=c_{-}(\sigma_{0},k_{\#})=U(-h) follows from taking the limit as σσ0\sigma^{\prime}\to\sigma_{0}-. ∎

A direct corollary of the proposition is a sufficient condition for the linear instability.

Corollary 3.16.1.

If g<g#g<g_{\#} and there exists a sequence x2,n(h,0)x_{2,n}\in(-h,0) converging to h-h as n+n\to+\infty such that U′′(x2,n)>0U^{\prime\prime}(x_{2,n})>0 for all nn, then the linearized capillary gravity wave system is linearly unstable.

Remark 3.6.

For the linearized capillary gravity wave problem, a non-degenerate inflection value c0c_{0} of UU does not necessarily leads to instability as a strong surface tension may prevent c0c_{0} from becoming a neutral mode at all.

References

  • [1] T. B. Benjamin. The solitary wave on a stream with an arbitrary distribution of vorticity. J. Fluid Mech., 12:97–116, 1962.
  • [2] D. Bresch and M. Renardy. Kelvin-Helmholtz instability with a free surface. Z. Angew. Math. Phys., 64(4):905–915, 2013.
  • [3] O. Buhler, J. Shatah, S. Walsh, and C. Zeng. On the wind generation of water waves. Arch. Ration. Mech. Anal., 222:827 – 878, 2016.
  • [4] J. C. Burns. Long waves in running water. Proc. Cambridge Philos. Soc., 49:695–706, 1953. With an appendix by M. J. Lighthill.
  • [5] P. Drazin and W. Reid. Hydrodynamic stability. Cambridge: Cambridge University Press., 2004.
  • [6] 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.
  • [7] S. Friedlander and L. Howard. Instability in parallel flows revisited. Stud. Appl. Math., 101(1):1–21, 1998.
  • [8] L. Howard. Note on a paper of john w. miles. J. Fluid. Mech., 10:509–512, 1961.
  • [9] J. N. Hunt. Gravity waves in flowing water. Proc. Roy. Soc. London Ser. A, 231:496–504, 1955.
  • [10] V. Hur and Z. Lin. Unstable surface waves in running water. Comm. Math Phys., 282:733 – 796, 2008.
  • [11] V. M. Hur and Z. Lin. Erratum to: Unstable surface waves in running water [mr2426143]. Comm. Math. Phys., 318(3):857–861, 2013.
  • [12] P. Janssen. The interaction of ocean waves and wind. Cambridge University Press, 2004.
  • [13] A. Kaffel and M. Renardy. Surface modes in inviscid free surface shear flows. ZAMM Z. Angew. Math. Mech., 91(8):649–652, 2011.
  • [14] C. C. Lin. The theory of hydrodynamic stability. Cambridge, at the University Press, 1955.
  • [15] Z. Lin. Instability of some ideal plane flows. SIAM J. Math. Anal., 35(2):318–356, 2003.
  • [16] X. Liu. Instability and spectrum of the linearized two-phase fluids interface problem at shear flows. arXiv:2208.11159, 2022.
  • [17] X. Liu and C. Zeng. Capillary gravity water waves linearized at monotone shear flows: eigenvalues and inviscid damping. arXiv:2110.12604, 2021.
  • [18] M. S. Longuet-Higgins. Instabilities of a horizontal shear flow with a free surface. J. Fluid Mech., 364:147–162, 1998.
  • [19] C. Marchioro and M. Pulvirenti. Mathematical theory of incompressible nonviscous fluids, volume 96 of Applied Mathematical Sciences. Springer-Verlag, New York, 1994.
  • [20] J. Miles. On the generation of surface waves by shear flows. J. Fluid Mech., 3:185–204, 1957.
  • [21] J. W. Miles. On the generation of surface waves by shear flows. II. J. Fluid Mech., 6:568–582, 1959.
  • [22] J. W. Miles. On the generation of surface waves by shear flows. III. Kelvin-Helmholtz instability. J. Fluid Mech., 6:583–598. (1 plate), 1959.
  • [23] L. Rayleigh. On the stability or instability of certain fluid motions. Pro. London Math. Soc., 9:57–70, 1880.
  • [24] M. Renardy and Y. Renardy. On the stability of inviscid parallel shear flows with a free surface. Math. Fluid Mech., 15:129–137, 2013.
  • [25] V. I. Shrira. Surface waves on shear currents: solution of the boundary value problem. J. Fluid Mech., 252:565–584, 1993.
  • [26] W. L. K. Thomson. Hydrokinetic solutions and observations. Philosophical Magazine Series 4, (42):362–377, 1871.
  • [27] W. Tollmien. Ein allgemeines kriterium der instabititat laminarer geschwindigkeitsverteilungen. Nachr. Ges. Wiss. Gottingen Math. Phys., 50:79–114, 1935.
  • [28] C.-S. Yih. Surface waves in flowing water. J. Fluid. Mech., 51:209–220, 1972.