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

Benjamin-Feir instability of
Stokes waves in finite depth

Massimiliano Berti, Alberto Maspero, Paolo Ventura111 International School for Advanced Studies (SISSA), Via Bonomea 265, 34136, Trieste, Italy. Emails: berti@sissa.it, alberto.maspero@sissa.it, paolo.ventura@sissa.it
Abstract

Whitham and Benjamin predicted in 1967 that small-amplitude periodic traveling Stokes waves of the 2d-gravity water waves equations are linearly unstable with respect to long-wave perturbations, if the depth 𝚑{\mathtt{h}} is larger than a critical threshold 𝚑WB1.363\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}\approx 1.363. In this paper we completely describe, for any value of 𝚑>0\mathtt{h}>0, the four eigenvalues close to zero of the linearized equations at the Stokes wave, as the Floquet exponent μ\mu is turned on. We prove in particular the existence of a unique depth 𝚑WB\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, which coincides with the one predicted by Whitham and Benjamin, such that, for any 0<𝚑<𝚑WB0<\mathtt{h}<\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, the eigenvalues close to zero remain purely imaginary and, for any 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, a pair of non-purely imaginary eigenvalues depicts a closed figure “8”, parameterized by the Floquet exponent. As 𝚑𝚑WB+{\mathtt{h}}\to\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}^{\,+} this figure “8” collapses to the origin of the complex plane. The proof combines a symplectic version of Kato’s perturbative theory to compute the eigenvalues of a 4×44\times 4 Hamiltonian and reversible matrix, and KAM inspired transformations to block-diagonalize it. The four eigenvalues have all the same size 𝒪(μ)\mathcal{O}(\mu) –unlike the infinitely deep water case in [6]– and the correct Benjamin-Feir phenomenon appears only after one non-perturbative block-diagonalization step. In addition one has to track, along the whole proof, the explicit dependence of the entries of the 4×44\times 4 reduced matrix with respect to the depth 𝚑\mathtt{h}.

1 Introduction to main results

A classical problem in fluid dynamics, pioneered by the famous work of Stokes [38] in 1847, concerns the spectral stability/instability of periodic traveling waves –called Stokes waves– of the gravity water waves equations in any depth.

Benjamin and Feir [3], Lighthill [32] and Zakharov [43, 45] discovered in the sixties, through experiments and formal arguments, that Stokes waves in deep water are unstable, proposing an heuristic mechanism which leads to the disintegration of wave trains. More precisely, these works predicted unstable eigenvalues of the linearized equations at the Stokes wave, near the origin of the complex plane, corresponding to small Floquet exponents μ\mu or, equivalently, to long-wave perturbations. The same phenomenon was later predicted by Whitham [41] and Benjamin [2] for Stokes waves of wavelength 2πκ2\pi\kappa, in finite depth 𝚑\mathtt{h}, provided that κ𝚑>1.363\kappa\mathtt{h}>1.363 approximately. This phenomenon is nowadays called “Benjamin-Feir” –or modulational– instability, and it is supported by an enormous amount of physical observations and numerical simulations, see e.g. [15, 33]. We refer to [46] for an historical survey.

A serious difficulty for a rigorous mathematical proof of the Benjamin-Feir instability is that the perturbed eigenvalues bifurcate from the eigenvalue zero, which is defective, with multiplicity four. The first rigorous proof of a local branch of unstable eigenvalues close to zero for κ𝚑\kappa\mathtt{h} larger than the Whitham-Benjamin threshold 1.3631.363\ldots was obtained by Bridges-Mielke [9] in finite depth (see also the preprint by Hur-Yang [23]). Their method, based on a spatial dynamics and a center manifold reduction, breaks down in deep water. For dealing with this case Nguyen-Strauss [35] have recently developed a new approach, based on a Lyapunov-Schmidt decomposition. The novel spectral approach developed in Berti-Maspero-Ventura [6] allowed to fully describe, in deep water, the splitting of the four eigenvalues close to zero, as the Floquet exponent is turned on, proving in particular the conjecture that a pair of non-purely imaginary eigenvalues depicts a closed figure “8”, parameterized by the Floquet exponent.

The goal of this paper is to describe the full Benjamin-Feir instability phenomenon at any finite value of the depth 𝚑>0\mathtt{h}>0. This analysis has fundamental physical importance since real-life experiments are performed in water tanks (for example the original Benjamin and Feir experiments, in Feltham’s National Physical Laboratory, had Stokes waves of wavelength 2.2 m and bottom’s depth of 7.62 m, see [2]). We also remark that the Benjamin-Feir instability mechanism is a possible responsible of the emergence of rogue waves in the ocean, we refer to [28, 29] and references therein for a vast physical literature. A first mathematically rigorous treatment of large waves is given in [18], via a probabilistic analysis, in the case of NLS.

Along this paper, with no loss of generality, we consider 2π2\pi-periodic Stokes waves, i.e. with wave number κ=1\kappa=1. In Theorems 2.5 and 1.1 we prove the existence of a unique depth 𝚑WB\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, in perfect agreement with the Benjamin-Feir critical value 1.363…, such that:

  • Shallow water case: for any 0<𝚑<𝚑WB0<\mathtt{h}<\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}} the eigenvalues close to zero remain purely imaginary for Stokes waves of sufficiently small amplitude, see Figure 2(a)-left;

  • Sufficiently deep water case: for any 𝚑WB<𝚑<\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}<\mathtt{h}<\infty, there exists a pair of non-purely imaginary eigenvalues which traces a complete closed figure “8” (as shown in Figure 2(a)-right) parameterized by the Floquet exponent μ\mu. By further increasing μ\mu, the eigenvalues recollide far from the origin on the imaginary axis where then they keep moving. As 𝚑𝚑WB+{\mathtt{h}}\to\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}^{\,+} the set of unstable Floquet exponents shrinks to zero and the Benjamin-Feir unstable eigenvalues collapse to the origin, see Figure 3. This figure ‘8” was first numerically discovered by Deconink-Oliveras in [15].

We remark that our approach fully describes all the eigenvalues close to 0, providing a necessary and sufficient condition for the existence of unstable eigenvalues, i.e. the positivity of the Benjamin-Feir discriminant function ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon) defined in (1.6).

The results of Theorems 2.5 and 1.1 are complementary to those of [6]. In following the natural spectral approach developed in [6], we encounter a major difference in the proof, that we now anticipate. In the infinitely deep water ideal case it turns out that the “reduced” 4×44\times 4 matrix obtained by the Kato reduction procedure is a small perturbation of a block-diagonal matrix which possesses yet the correct Benjamin-Feir unstable eigenvalues. This is not the case in finite depth. The correct eigenvalues of the “reduced” 4×44\times 4 matrix emerge only after one non-perturbative step of block diagonalization. We shall explain in detail this point after the statement of Theorem 2.5. This is related with the fact that, in infinite deep water, among the four eigenvalues close to zero of the linearized operator at the Stokes wave, two are 𝒪(μ)\mathcal{O}(\mu), whereas the other two have much larger size 𝒪(μ)\mathcal{O}(\sqrt{\mu}), whereas in finite depth all four eigenvalues have size 𝒪(μ)\mathcal{O}(\mu). In addition, along the whole proof, one needs to carefully track the explicit dependence with respect to 𝚑\mathtt{h} of the entries of the reduced 4×44\times 4 matrix.

Let us now present rigorously our results.

Benjamin-Feir instability in finite depth

We consider the pure gravity water waves equations for a bidimensional fluid occupying a region with finite depth 𝚑\mathtt{h}. With no loss of generality we set the gravity g=1g=1, see Remark 2.4. We consider a 2π2\pi-periodic Stokes wave with amplitude 0<ϵ10<\epsilon\ll 1 and speed

cϵ=𝚌𝚑+𝒪(ϵ2),𝚌𝚑:=tanh(𝚑).c_{\epsilon}={\mathtt{c}}_{\mathtt{h}}+\mathcal{O}(\epsilon^{2})\,,\quad{\mathtt{c}}_{\mathtt{h}}:=\sqrt{\tanh(\mathtt{h})}\,.

The linearized water waves equations at the Stokes wave are, in the inertial reference frame moving with speed cϵc_{\epsilon}, a linear time independent system of the form ht=ϵhh_{t}=\mathcal{L}_{\epsilon}h where ϵ:=ϵ(𝚑)\mathcal{L}_{\epsilon}:=\mathcal{L}_{\epsilon}({\mathtt{h}}) is a linear operator with 2π2\pi-periodic coefficients, see (2.17) (the operator ϵ\mathcal{L}_{\epsilon} in (2.17) is actually obtained conjugating the linearized water waves equations in the Zakharov formulation at the Stokes wave via the “good unknown of Alinhac” (2.11) and the Levi-Civita (2.16) invertible transformations). The operator ϵ\mathcal{L}_{\epsilon} possesses the eigenvalue 0, which is defective, with multiplicity four, due to symmetries of the water waves equations. The problem is to prove that the linear system ht=ϵhh_{t}=\mathcal{L}_{\epsilon}h has solutions of the form h(t,x)=Re(eλteiμxv(x))h(t,x)=\text{Re}\left(e^{\lambda t}e^{\mathrm{i}\,\mu x}v(x)\right) where v(x)v(x) is a 2π2\pi-periodic function, μ\mu in \mathbb{R} is the Floquet exponent and λ\lambda has positive real part, thus h(t,x)h(t,x) grows exponentially in time. By Bloch-Floquet theory, such λ\lambda is an eigenvalue of the operator μ,ϵ:=eiμxϵeiμx\mathcal{L}_{\mu,\epsilon}:=e^{-\mathrm{i}\,\mu x}\,\mathcal{L}_{\epsilon}\,e^{\mathrm{i}\,\mu x} acting on 2π2\pi-periodic functions.

The main result of this paper proves, for any finite value of the depth 𝚑\mathtt{h}, the full splitting of the four eigenvalues close to zero of the operator μ,ϵ:=μ,ϵ(𝚑)\mathcal{L}_{\mu,\epsilon}:=\mathcal{L}_{\mu,\epsilon}(\mathtt{h}) when ϵ\epsilon and μ\mu are small enough, see Theorem 2.5. We first present Theorem 1.1 which focuses on the figure ``8"``8" formed by the Benjamin-Feir unstable eigenvalues.

We first need to introduce the “Whitham-Benjamin” function

𝚎WB:=𝚎WB(𝚑):=1𝚌𝚑[9𝚌𝚑810𝚌𝚑4+98𝚌𝚑61𝚑14𝚎122(1+1𝚌𝚑42+34(1𝚌𝚑4)2𝚌𝚑2𝚑)],\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}:=\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h}):=\frac{1}{{\mathtt{c}}_{\mathtt{h}}}\Big{[}\frac{9{\mathtt{c}}_{\mathtt{h}}^{8}-10{\mathtt{c}}_{\mathtt{h}}^{4}+9}{8{\mathtt{c}}_{\mathtt{h}}^{6}}-\frac{1}{\mathtt{h}-\frac{1}{4}\mathtt{e}_{12}^{2}}\Big{(}1+\frac{1-{\mathtt{c}}_{\mathtt{h}}^{4}}{2}+\frac{3}{4}\frac{(1-{\mathtt{c}}_{\mathtt{h}}^{4})^{2}}{{\mathtt{c}}_{\mathtt{h}}^{2}}\mathtt{h}\Big{)}\Big{]}\,, (1.1)

where 𝚌𝚑=tanh(𝚑){\mathtt{c}}_{\mathtt{h}}=\sqrt{\tanh(\mathtt{h})} is the speed of the linear Stokes wave, and

𝚎12:=𝚎12(𝚑):=𝚌𝚑+𝚌𝚑1(1𝚌𝚑4)𝚑>0,𝚑>0.\mathtt{e}_{12}:=\mathtt{e}_{12}(\mathtt{h}):={\mathtt{c}}_{\mathtt{h}}+{\mathtt{c}}_{\mathtt{h}}^{-1}(1-{\mathtt{c}}_{\mathtt{h}}^{4})\mathtt{h}>0\,,\quad\forall\mathtt{h}>0\,. (1.2)

The function 𝚎WB(𝚑)\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h}) is well defined for any 𝚑>0\mathtt{h}>0 because the denominator 𝚑14𝚎122>0\mathtt{h}-\tfrac{1}{4}\mathtt{e}_{12}^{2}>0 in (1.1) is positive for any 𝚑>0\mathtt{h}>0, see Lemma 5.7. The function (1.1) coincides, up to a non zero factor, with the celebrated function obtained by Whitham [41], Benjamin [2] and Bridges-Mielke [9] which determines the “shallow/sufficiently deep” threshold regime. In particular the Whitham-Benjamin function 𝚎WB(𝚑)\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h}) vanishes at 𝚑WB=1.363\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}=1.363..., it is negative for 0<𝚑<𝚑WB0<\mathtt{h}<\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, positive for 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}} and tends to 11 as 𝚑+\mathtt{h}\to+\infty, see Figure 1. We also introduce the positive coefficient

𝚎22:=𝚎22(𝚑):=(1𝚌𝚑4)(1+3𝚌𝚑4)𝚑2+2𝚌𝚑2(𝚌𝚑41)𝚑+𝚌𝚑4𝚌𝚑3>0,𝚑>0.\mathtt{e}_{22}:=\mathtt{e}_{22}(\mathtt{h}):=\dfrac{(1-{\mathtt{c}}_{\mathtt{h}}^{4})(1+3{\mathtt{c}}_{\mathtt{h}}^{4})\mathtt{h}^{2}+2{\mathtt{c}}_{\mathtt{h}}^{2}({\mathtt{c}}_{\mathtt{h}}^{4}-1)\mathtt{h}+{\mathtt{c}}_{\mathtt{h}}^{4}}{{\mathtt{c}}_{\mathtt{h}}^{3}}>0\,,\quad\forall\mathtt{h}>0\,. (1.3)

We remark that the functions 𝚎12(𝚑)>𝚌𝚑\mathtt{e}_{12}(\mathtt{h})>{\mathtt{c}}_{\mathtt{h}} and 𝚎22(𝚑)>0\mathtt{e}_{22}(\mathtt{h})>0 are positive for any 𝚑>0\mathtt{h}>0, tend to 0 as 𝚑0+\mathtt{h}\to 0^{+} and to 11 as 𝚑+\mathtt{h}\to+\infty, see Lemma 4.8.

Refer to caption
Figure 1: Plot of the Whitham-Benjamin function 𝚎WB(𝚑)\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h}). The red dot shows its unique root 𝚑WB=1.363\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}=1.363\dots. which is the celebrated “shallow/sufficiently deep” water threshold predicted independently by Whitham (cfr.[41] p.49) and Benjamin (cfr.[2] p.68), and recovered in the rigorous proof of Bridges-Mielke [9, p. 183].

Along the paper we denote by r(ϵm1μn1,,ϵmpμnp)r(\epsilon^{m_{1}}\mu^{n_{1}},\ldots,\epsilon^{m_{p}}\mu^{n_{p}}) a real analytic function fulfilling for some C>0C>0 and ϵ,μ\epsilon,\mu sufficiently small, the estimate |r(ϵm1μn1,,ϵmpμnp)|Cj=1p|ϵ|mj|μ|nj|r(\epsilon^{m_{1}}\mu^{n_{1}},\ldots,\epsilon^{m_{p}}\mu^{n_{p}})|\leq C\sum_{j=1}^{p}|\epsilon|^{m_{j}}|\mu|^{n_{j}}, where the constant C:=C(𝚑)C:=C(\mathtt{h}) is uniform for 𝚑\mathtt{h} in any compact set of (0,+)(0,+\infty).

Theorem 1.1.

(Benjamin-Feir unstable eigenvalues) For any 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, there exist ϵ1,μ0>0\epsilon_{1},\mu_{0}>0 and an analytic function μ¯:[0,ϵ1)[0,μ0)\underline{\mu}:[0,\epsilon_{1})\to[0,\mu_{0}), of the form

μ¯(ϵ)=𝚎𝚑ϵ(1+r(ϵ)),𝚎𝚑:=8𝚎WB(𝚑)𝚎22(𝚑),\underline{\mu}(\epsilon)=\mathtt{e}_{\mathtt{h}}\epsilon(1+r(\epsilon))\,,\quad\mathtt{e}_{\mathtt{h}}:=\sqrt{\frac{8\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h})}{\mathtt{e}_{22}(\mathtt{h})}}\,, (1.4)

such that, for any ϵ[0,ϵ1)\epsilon\in[0,\epsilon_{1}), the operator μ,ϵ\mathcal{L}_{\mu,\epsilon} has two eigenvalues λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) of the form

{i12𝚌˘𝚑μ+ir2(μϵ2,μ2ϵ,μ3)±18μ𝚎22(𝚑)(1+r(ϵ,μ))ΔBF(𝚑;μ,ϵ),μ[0,μ¯(ϵ))i12𝚌˘𝚑μ¯(ϵ)+ir(ϵ3),μ=μ¯(ϵ)i12𝚌˘𝚑μ+ir2(μϵ2,μ2ϵ,μ3)±i18μ𝚎22(𝚑)(1+r(ϵ,μ))|ΔBF(𝚑;μ,ϵ)|,μ(μ¯(ϵ),μ0)\begin{cases}\mathrm{i}\,\frac{1}{2}\breve{\mathtt{c}}_{\mathtt{h}}\mu+\mathrm{i}\,r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\pm\tfrac{1}{8}\mu\sqrt{\mathtt{e}_{22}(\mathtt{h})}(1+r(\epsilon,\mu))\sqrt{\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon)},&\forall\mu\in[0,\underline{\mu}(\epsilon))\!\!\!\\[4.2679pt] \mathrm{i}\,\frac{1}{2}\breve{\mathtt{c}}_{\mathtt{h}}\underline{\mu}(\epsilon)+\mathrm{i}\,r(\epsilon^{3}),&\mu=\underline{\mu}(\epsilon)\!\!\!\\[4.2679pt] \mathrm{i}\,\frac{1}{2}\breve{\mathtt{c}}_{\mathtt{h}}\mu+\mathrm{i}\,r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\pm\mathrm{i}\,\tfrac{1}{8}\mu\sqrt{\mathtt{e}_{22}(\mathtt{h})}(1+r(\epsilon,\mu))\sqrt{|\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon)|},&\forall\mu\in(\underline{\mu}(\epsilon),\mu_{0})\!\!\!\end{cases}\!\!\! (1.5)

where 𝚌˘𝚑:=2𝚌𝚑𝚎12(𝚑)>0\breve{\mathtt{c}}_{\mathtt{h}}:=2{\mathtt{c}}_{\mathtt{h}}-\mathtt{e}_{12}(\mathtt{h})>0 and ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon) is the “Benjamin-Feir discriminant” function

ΔBF(𝚑;μ,ϵ):=8𝚎WB(𝚑)ϵ2+r1(ϵ3,μϵ2)𝚎22(𝚑)μ2(1+r1′′(ϵ,μ)).\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon):=8\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h})\epsilon^{2}+r_{1}(\epsilon^{3},\mu\epsilon^{2})-\mathtt{e}_{22}(\mathtt{h})\mu^{2}\big{(}1+r_{1}^{\prime\prime}(\epsilon,\mu)\big{)}\,. (1.6)

Note that, for any 0<ϵ<ϵ10<\epsilon<\epsilon_{1} (depending on 𝚑\mathtt{h}) the function ΔBF(𝚑;μ,ϵ)>0\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon)>0 is positive, respectively <0<0, provided 0<μ<μ¯(ϵ)0<\mu<\underline{\mu}(\epsilon), respectively μ>μ¯(ϵ)\mu>\underline{\mu}(\epsilon).

Let us make some comments.
1. Benjamin-Feir unstable eigenvalues. For 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, according to (1.5), for values of the Floquet parameter 0<μ<μ¯(ϵ)0<\mu<\underline{\mu}(\epsilon), the eigenvalues λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) have opposite non-zero real part. As μ\mu tends to μ¯(ϵ)\underline{\mu}(\epsilon), the two eigenvalues λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) collide on the imaginary axis far from 0 (in the upper semiplane Im(λ)>0\text{Im}(\lambda)>0), along which they keep moving for μ>μ¯(ϵ)\mu>\underline{\mu}(\epsilon), see Figure 2(a). For μ<0\mu<0 the operator μ,ϵ{\mathcal{L}}_{\mu,\epsilon} possesses the symmetric eigenvalues λ1±(μ,ϵ)¯\overline{\lambda_{1}^{\pm}(-\mu,\epsilon)} in the semiplane Im(λ)<0\text{Im}(\lambda)<0. For μ[0,μ¯(ϵ)]\mu\in[0,\underline{\mu}(\epsilon)] we obtain the upper part of the figure “8”, which is well approximated by the curves

μ(±μ8𝚎228𝚎WBϵ2𝚎22μ2,12𝚌˘𝚑μ),\mu\mapsto\Big{(}\pm\frac{\mu}{8}\sqrt{\mathtt{e}_{22}}\sqrt{8\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}\epsilon^{2}-\mathtt{e}_{22}\mu^{2}},\ \tfrac{1}{2}\breve{\mathtt{c}}_{\mathtt{h}}\mu\Big{)}\,, (1.7)

in accordance with the numerical simulations by Deconinck-Oliveras [15]. Note that for μ>0\mu>0 the imaginary part in (1.7) is positive because 𝚌˘𝚑=𝚌𝚑1(tanh(𝚑)(1tanh2(𝚑))𝚑)>0\breve{\mathtt{c}}_{\mathtt{h}}={\mathtt{c}}_{\mathtt{h}}^{-1}(\tanh(\mathtt{h})-(1-\tanh^{2}(\mathtt{h}))\mathtt{h})>0 for any 𝚑>0\mathtt{h}>0. The higher order “side-band” corrections of the eigenvalues λ1±(μ,ϵ)\lambda_{1}^{\pm}(\mu,\epsilon) in (1.5), provided by the analytic functions r,r1,r1′′,r2r,r_{1},r_{1}^{\prime\prime},r_{2}, are explicitly computable. We finally remark that the eigenvalues (1.5) are not analytic in (μ,ϵ)(\mu,\epsilon) close to the value (μ¯(ϵ),ϵ)(\underline{\mu}(\epsilon),\epsilon) where λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) collide at the top of the figure ``8"``8" far from 0 (clearly they are continuous).

(a) *
(b) *
Figure 2: The picture on the left shows, in the “shallow” water regime 𝚑<𝚑WB\mathtt{h}<\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, the eigenvalues λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) and λ0±(μ,ϵ)\lambda^{\pm}_{0}(\mu,\epsilon) which are purely imaginary. The picture on the right shows, in the “sufficiently deep” water regime 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, the eigenvalues λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) in the complex λ\lambda-plane at fixed |ϵ|1|\epsilon|\ll 1 as μ\mu varies. This figure “8 ” depends on 𝚑\mathtt{h} and shrinks to 0 as 𝚑𝚑WB+\mathtt{h}\to\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}^{+}, see Figure 3. As 𝚑+\mathtt{h}\to+\infty the spectrum resembles the one in deep water found in [6].

[.4]Refer to caption          [.4] Refer to caption

2. Behaviour near the Whitham-Benjamin depth 𝚑WB\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}. As 𝚑𝚑WB+{\mathtt{h}}\to\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}^{+} the constant ϵ1:=ϵ1(𝚑)>0\epsilon_{1}:=\epsilon_{1}(\mathtt{h})>0 in Theorem 1.1 tends to zero, the set of unstable Floquet exponents (0,μ¯(ϵ))(0,\underline{\mu}(\epsilon)) with μ¯(ϵ)=𝚎𝚑ϵ(1+r(ϵ))\underline{\mu}(\epsilon)=\mathtt{e}_{\mathtt{h}}\epsilon(1+r(\epsilon)) given in (1.4) shrinks to zero and the figure “8” of Benjamin-Feir unstable eigenvalues collapse to zero, see Figure 3. In particular

maxμ[0,μ¯(ϵ)]Reλ1+(μ,ϵ)=Reλ1+(μmax,ϵ)=12𝚎WB(𝚑)ϵ2+r(ϵ3) and \max_{\mu\in[0,\underline{\mu}(\epsilon)]}\text{Re}\,\lambda_{1}^{+}(\mu,\epsilon)=\text{Re}\,\lambda_{1}^{+}(\mu_{\max},\epsilon)=\frac{1}{2}{\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}}(\mathtt{h})\epsilon^{2}+r(\epsilon^{3})\ \text{ and } (1.8)

tends to zero as 𝚑𝚑WB+\mathtt{h}\to\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}^{+}, since 0<ϵ<ϵ1(𝚑)0<\epsilon<\epsilon_{1}(\mathtt{h}) and ϵ1(𝚑)0+\epsilon_{1}(\mathtt{h})\to 0^{+}.

Refer to caption
Figure 3: The Benjamin-Feir eigenvalue λ1+(μmax,ϵ)\lambda^{+}_{1}(\mu_{\max},\epsilon) in (1.8) with maximal real part, as well as the whole figure “88” shrinks to zero as 𝚑𝚑WB+\mathtt{h}\to\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}^{+}.

3. Relation with Bridges-Mielke [9]. Bridges and Mielke describe the unstable eigenvalues very close to the origin, namely the cross amid the ‘8”. In order to make a precise comparison with our result let us spell out the relation of the functions 𝚎WB\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}, 𝚎12\mathtt{e}_{12} and 𝚎22\mathtt{e}_{22} with the coefficients obtained in [9]. The Whitham-Benjamin function 𝚎WB\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}} in (4.13) is 𝚎WB=(𝚌𝚑𝚑)1ν(F)\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}=({\mathtt{c}}_{\mathtt{h}}\mathtt{h})^{-1}\nu(F), where ν(F)\nu(F) is defined in [9, formula (6.17)] and F=𝚌𝚑𝚑12F={\mathtt{c}}_{\mathtt{h}}\mathtt{h}^{-\frac{1}{2}} is the Froude number, cfr. [9, formula (3.4)]. Moreover the term 𝚎12\mathtt{e}_{12} in (1.2) is 𝚎12=2cg\mathtt{e}_{12}=2c_{g}, where cg=12𝚌𝚑(1+F2sech2(𝚑))c_{g}=\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}\big{(}1+F^{-2}\text{sech}^{2}(\mathtt{h})\big{)} is the group velocity defined in Bridges-Mielke [9, formula (3.8)]. Finally 𝚎22(𝚑)c˙g\mathtt{e}_{22}(\mathtt{h})\propto\dot{c}_{g} where c˙g\dot{c}_{g} is the derivative of the group velocity defined in [9, formula (6.15)], which for gravity waves is negative in any depth.
4. Complete spectrum near 0. In Theorem 1.1 we have described just the two unstable eigenvalues of μ,ϵ\mathcal{L}_{\mu,\epsilon} close to zero for 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}. There are also two larger purely imaginary eigenvalues of order 𝒪(μ)\mathcal{O}(\mu), see Theorem 2.5. We remark that our approach describes all the eigenvalues of μ,ϵ{\mathcal{L}}_{\mu,\epsilon} close to 0 (which are 44).
5. Shallow water regime. In the shallow water regime 0<𝚑<𝚑WB0<\mathtt{h}<\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, we prove in Theorem 2.5 that all the four eigenvalues of μ,ϵ{\mathcal{L}}_{\mu,\epsilon} close to zero remain purely imaginary for ϵ\epsilon sufficiently small. The eigenvalue expansions of Theorem 2.5 become singular as 𝚑0+\mathtt{h}\to 0^{+}.
6. Behavior at the Whitham-Benjamin threshold 𝚑WB\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}. The analysis of Theorem 1.1 is not conclusive at the critical depth 𝚑=𝚑WB\mathtt{h}=\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}. The reason is that 𝚎WB(𝚑WB)=0\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}})=0 and the Benjamin-Feir discriminant function (1.6) reduces to

ΔBF(𝚑WB;μ,ϵ)=r(ϵ3)+r(μϵ2)𝚎22(𝚑WB)μ2(1+r1′′(ϵ,μ)).\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}};\mu,\epsilon)=r(\epsilon^{3})+r(\mu\epsilon^{2})-\mathtt{e}_{22}(\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}})\mu^{2}(1+r_{1}^{\prime\prime}(\epsilon,\mu))\,. (1.9)

Thus its quadratic expansion is not sufficient anymore to determine the sign of ΔBF(𝚑WB;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}};\mu,\epsilon). Note that (1.9) could be positive due to the cubic term r(ϵ3)=αϵ3+r(\epsilon^{3})=\alpha\epsilon^{3}+\dots for ϵ\epsilon and μ\mu small enough. The coefficient α\alpha could be explicitly computed taking into account the third order expansion of the Stokes waves.
7. Unstable Floquet exponents and amplitudes (μ,ϵ)(\mu,\epsilon). In Theorem 2.5 we actually prove that the expansion (1.5) of the eigenvalues of μ,ϵ\mathcal{L}_{\mu,\epsilon} holds for any value of (μ,ϵ)(\mu,\epsilon) in a larger rectangle [0,μ0)×[0,ϵ0)[0,\mu_{0})\times[0,\epsilon_{0}), and there exist Benjamin-Feir unstable eigenvalues if and only if the analytic function ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon) in (1.6) is positive. The zero set of ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon) is an analytic variety which, for 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, is, restricted to the rectangle [0,μ0)×[0,ϵ1)[0,\mu_{0})\times[0,\epsilon_{1}), the graph of the analytic function μ¯(ϵ)=𝚎𝚑ϵ(1+r(ϵ))\underline{\mu}(\epsilon)=\mathtt{e}_{\mathtt{h}}\epsilon(1+r(\epsilon)) in (1.4). This function is tangent at ϵ=0\epsilon=0 to the straight line μ=𝚎𝚑ϵ\mu=\mathtt{e}_{\mathtt{h}}\epsilon, and divides [0,μ0)×[0,ϵ1)[0,\mu_{0})\times[0,\epsilon_{1}) in the region where ΔBF(𝚑;μ,ϵ)>0\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon)>0 –and thus the eigenvalues of μ,ϵ{\mathcal{L}}_{\mu,\epsilon} have non-trivial real part–, from the “stable” one where all the eigenvalues of μ,ϵ{\mathcal{L}}_{\mu,\epsilon} are purely imaginary, see Figure 4. In the region [0,μ0)×[ϵ1,ϵ0)[0,\mu_{0})\times[\epsilon_{1},\epsilon_{0}) the higher order polynomial approximations of ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon) (which are computable) will determine the sign of ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon).

Refer to caption
Figure 4: The solid curve portrays the graph of the real analytic function μ¯(ϵ)\underline{\mu}(\epsilon) in (1.4) as 𝚑>𝚑WB\mathtt{h}>\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}. For values of μ\mu below this curve, the two eigenvalues λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) have non zero real part. For μ\mu above the curve, λ1±(μ,ϵ)\lambda^{\pm}_{1}(\mu,\epsilon) are purely imaginary. In the region [ϵ1,ϵ0)×[0,μ0)[\epsilon_{1},\epsilon_{0})\times[0,\mu_{0}) the eigenvalues are real/purely imaginary depending on the higher order corrections given by Theorem 2.5, which determine the sign of ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon).

8. Deep water limit. Theorems 1.1 and 2.5 do not pass to the limit as 𝚑+\mathtt{h}\to+\infty since the remainders in the expansions of the eigenvalues are uniform only on any compact set of 𝚑(0,+)\mathtt{h}\in(0,+\infty). From a mathematical point of view, the difference is evident in the asymptotic behavior of tanh(𝚑μ)\tanh(\mathtt{h}\mu) (and similar quantities) which, in the idealized deep water case 𝚑=+\mathtt{h}=+\infty, is identically equal to 11 for any arbitrarily small Floquet exponent μ\mu, whereas tanh(𝚑μ)=O(μ𝚑)\tanh(\mathtt{h}\mu)=O(\mu\mathtt{h}) for any 𝚑\mathtt{h} finite. However additional intermediate scaling regimes 𝚑μ1\mathtt{h}\mu\sim 1, 𝚑μ1\mathtt{h}\mu\ll 1, 𝚑μ1\mathtt{h}\mu\gg 1 are possible. It is well-known (e.g. see [14]) that intermediate long-wave regimes of the water-waves equations formally lead to different physically-relevant limit equations as Boussinesq, KdV, NLS, Benjamin-Ono, etc…

We shall describe in detail the ideas of proof and the differences with the deep water case below the statement of Theorem 2.5.
Further literature. Modulational instability has been studied also for a variety of approximate water waves models, such as KdV, gKdV, NLS and the Whitham equation by, for instance, Whitham [42], Segur, Henderson, Carter and Hammack [37], Gallay and Haragus [17], Haragus and Kapitula [19], Bronski and Johnson [11], Johnson [25], Hur and Johnson [21], Bronski, Hur and Johnson [10], Hur and Pandey [22], Leisman, Bronski, Johnson and Marangell [30]. Also for these approximate models, numerical simulations predict a figure “8” similar to that in Figure 2(a) for the bifurcation of the unstable eigenvalues close to zero. We expect the present approach can be adapted to describe the full bifurcation of the eigenvalues also for these models.

Finally we mention the nonlinear modulational instability result of Jin, Liao, and Lin [24] for several fluid model equations and the preprint by Chen-Su [12] for Stokes waves in deep water. Nonlinear transversal instability results of traveling solitary water waves in finite depth decaying at infinity on \mathbb{R} have been proved in [36] (in deep water no solitary wave exists [20, 27]).
Acknowledgments. Research supported by PRIN 2020 (2020XB3EFL001) “Hamiltonian and dispersive PDEs”.

2 The complete Benjamin-Feir spectrum in finite depth

In this section we present in detail the complete spectral Theorem 2.5. We first introduce the pure gravity water waves equations and the Stokes waves solutions.
The water waves equations. We consider the Euler equations for a 2-dimensional incompressible, irrotational fluid under the action of gravity. The fluid fills the region

𝒟η:={(x,y)𝕋×:𝚑y<η(t,x)},𝕋:=/2π,{\mathcal{D}}_{\eta}:=\left\{(x,y)\in\mathbb{T}\times\mathbb{R}\;:\;-\mathtt{h}\leq y<\eta(t,x)\right\}\,,\quad\mathbb{T}:=\mathbb{R}/2\pi\mathbb{Z}\,,

with finite depth and space periodic boundary conditions. The irrotational velocity field is the gradient of a harmonic scalar potential Φ=Φ(t,x,y)\Phi=\Phi(t,x,y) determined by its trace ψ(t,x)=Φ(t,x,η(t,x))\psi(t,x)=\Phi(t,x,\eta(t,x)) at the free surface y=η(t,x)y=\eta(t,x). Actually Φ\Phi is the unique solution of the elliptic equation ΔΦ=0\Delta\Phi=0 in 𝒟η{\mathcal{D}}_{\eta} with Dirichlet datum Φ(t,x,η(t,x))=ψ(t,x)\Phi(t,x,\eta(t,x))=\psi(t,x) and Φy(t,x,y)=0\Phi_{y}(t,x,y)=0 at y=𝚑y=-\mathtt{h}.

The time evolution of the fluid is determined by two boundary conditions at the free surface. The first is that the fluid particles remain, along the evolution, on the free surface (kinematic boundary condition), and the second one is that the pressure of the fluid is equal, at the free surface, to the constant atmospheric pressure (dynamic boundary condition). Then, as shown by Zakharov [44] and Craig-Sulem [13], the time evolution of the fluid is determined by the following equations for the unknowns (η(t,x),ψ(t,x))(\eta(t,x),\psi(t,x)),

ηt=G(η)ψ,ψt=gηψx22+12(1+ηx2)(G(η)ψ+ηxψx)2,\eta_{t}=G(\eta)\psi\,,\quad\psi_{t}=-g\eta-\dfrac{\psi_{x}^{2}}{2}+\dfrac{1}{2(1+\eta_{x}^{2})}\big{(}G(\eta)\psi+\eta_{x}\psi_{x}\big{)}^{2}\,, (2.1)

where g>0g>0 is the gravity constant and G(η):=G(η,𝚑)G(\eta):=G(\eta,\mathtt{h}) denotes the Dirichlet-Neumann operator [G(η)ψ](x):=Φy(x,η(x))Φx(x,η(x))ηx(x)[G(\eta)\psi](x):=\Phi_{y}(x,\eta(x))-\Phi_{x}(x,\eta(x))\eta_{x}(x). In the sequel, with no loss of generality, we set the gravity constant g=1g=1, see Remark 2.4.

The equations (2.1) are the Hamiltonian system

t[ηψ]=𝒥[ηψ],𝒥:=[0IdId0],\partial_{t}\begin{bmatrix}\eta\\ \psi\end{bmatrix}=\mathcal{J}\begin{bmatrix}\nabla_{\eta}\mathcal{H}\\ \nabla_{\psi}\mathcal{H}\end{bmatrix},\quad\quad\mathcal{J}:=\begin{bmatrix}0&\mathrm{Id}\\ -\mathrm{Id}&0\end{bmatrix}, (2.2)

where \nabla denote the L2L^{2}-gradient, and the Hamiltonian (η,ψ):=12𝕋(ψG(η)ψ+η2)dx\mathcal{H}(\eta,\psi):=\frac{1}{2}\int_{\mathbb{T}}\left(\psi\,G(\eta)\psi+\eta^{2}\right)\mathrm{d}x is the sum of the kinetic and potential energy of the fluid. In addition of being Hamiltonian, the water waves system (2.1) possesses other important symmetries. First of all it is time reversible with respect to the involution

ρ[η(x)ψ(x)]:=[η(x)ψ(x)],i.e. ρ=.\rho\begin{bmatrix}\eta(x)\\ \psi(x)\end{bmatrix}:=\begin{bmatrix}\eta(-x)\\ -\psi(-x)\end{bmatrix},\quad\text{i.e. }\mathcal{H}\circ\rho=\mathcal{H}\,. (2.3)

Moreover, the equation (2.1) is space invariant, since, being the bottom flat,

τθG(η)ψ=G(τθη)[τθψ],θ,whereτθu(x):=u(x+θ).\tau_{\theta}G(\eta)\psi=G(\tau_{\theta}\eta)[\tau_{\theta}\psi]\,,\quad\forall\theta\in\mathbb{R}\,,\quad\text{where}\quad\tau_{\theta}u(x):=u(x+\theta)\,.

In addition, the Dirichlet-Neumann operator satisfies G(η+m,𝚑)=G(η,𝚑+m)G(\eta+m,\mathtt{h})=G(\eta,\mathtt{h}+m), for any mm\in\mathbb{R}.
Stokes waves. The Stokes waves are traveling solutions of (2.1) of the form η(t,x)=η˘(xct)\eta(t,x)=\breve{\eta}(x-ct) and ψ(t,x)=ψ˘(xct)\psi(t,x)=\breve{\psi}(x-ct) for some real cc and 2π2\pi-periodic functions (η˘(x),ψ˘(x))(\breve{\eta}(x),\breve{\psi}(x)). In a reference frame in translational motion with constant speed cc, the water waves equations (2.1) become

ηt=cηx+G(η)ψ,ψt=cψxηψx22+12(1+ηx2)(G(η)ψ+ηxψx)2\eta_{t}=c\eta_{x}+G(\eta)\psi\,,\quad\psi_{t}=c\psi_{x}-\eta-\dfrac{\psi_{x}^{2}}{2}+\dfrac{1}{2(1+\eta_{x}^{2})}\big{(}G(\eta)\psi+\eta_{x}\psi_{x}\big{)}^{2} (2.4)

and the Stokes waves (η˘,ψ˘)(\breve{\eta},\breve{\psi}) are equilibrium steady solutions of (2.4).

The bifurcation result of small amplitude of Stokes waves is due to Struik [39] in finite depth, and Levi-Civita [31], and Nekrasov [34] in infinite depth. We denote by B(r):={x:|x|<r}B(r):=\{x\in\mathbb{R}\colon\ |x|<r\} the real ball with center 0 and radius rr.

Theorem 2.1.

(Stokes waves) For any 𝚑>0\mathtt{h}>0 there exist ϵ:=ϵ(𝚑)>0\epsilon_{*}:=\epsilon_{*}(\mathtt{h})>0 and a unique family of real analytic solutions (ηϵ(x),ψϵ(x),cϵ)(\eta_{\epsilon}(x),\psi_{\epsilon}(x),c_{\epsilon}), parameterized by the amplitude |ϵ|ϵ|\epsilon|\leq\epsilon_{*}, of

cηx+G(η)ψ=0,cψxηψx22+12(1+ηx2)(G(η)ψ+ηxψx)2=0,c\,\eta_{x}+G(\eta)\psi=0\,,\quad c\,\psi_{x}-\eta-\dfrac{\psi_{x}^{2}}{2}+\dfrac{1}{2(1+\eta_{x}^{2})}\big{(}G(\eta)\psi+\eta_{x}\psi_{x}\big{)}^{2}=0\,, (2.5)

such that ηϵ(x),ψϵ(x)\eta_{\epsilon}(x),\psi_{\epsilon}(x) are 2π2\pi-periodic; ηϵ(x)\eta_{\epsilon}(x) is even and ψϵ(x)\psi_{\epsilon}(x) is odd, of the form

ηϵ(x)=ϵcos(x)+ϵ2(η2[0]+η2[2]cos(2x))+𝒪(ϵ3),\displaystyle\eta_{\epsilon}(x)=\epsilon\cos(x)+\epsilon^{2}(\eta_{2}^{[0]}+\eta_{2}^{[2]}\cos(2x))+\mathcal{O}(\epsilon^{3}), (2.6)
ψϵ(x)=ϵ𝚌𝚑1sin(x)+ϵ2ψ2[2]sin(2x)+𝒪(ϵ3),\displaystyle\psi_{\epsilon}(x)=\epsilon{\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x)+\epsilon^{2}\psi_{2}^{[2]}\sin(2x)+\mathcal{O}(\epsilon^{3})\,,
cϵ=𝚌𝚑+ϵ2c2+𝒪(ϵ3)where𝚌𝚑=tanh(𝚑),\displaystyle c_{\epsilon}={\mathtt{c}}_{\mathtt{h}}+\epsilon^{2}c_{2}+\mathcal{O}(\epsilon^{3})\quad\text{where}\quad{\mathtt{c}}_{\mathtt{h}}=\sqrt{\tanh(\mathtt{h})}\,,

and

η2[0]:=𝚌𝚑414𝚌𝚑2,η2[2]:=3𝚌𝚑44𝚌𝚑6,ψ2[2]:=3+𝚌𝚑88𝚌𝚑7,\displaystyle\eta_{2}^{[0]}:=\frac{{\mathtt{c}}_{\mathtt{h}}^{4}-1}{4{\mathtt{c}}_{\mathtt{h}}^{2}}\,,\qquad\eta_{2}^{[2]}:=\frac{3-{\mathtt{c}}_{\mathtt{h}}^{4}}{4{\mathtt{c}}_{\mathtt{h}}^{6}}\,,\qquad\psi_{2}^{[2]}:=\frac{3+{\mathtt{c}}_{\mathtt{h}}^{8}}{8{\mathtt{c}}_{\mathtt{h}}^{7}}\,,\qquad (2.7)
c2:=910𝚌𝚑4+9𝚌𝚑816𝚌𝚑7+(1𝚌𝚑4)2𝚌𝚑η2[0]=2𝚌𝚑12+13𝚌𝚑812𝚌𝚑4+916𝚌𝚑7.\displaystyle c_{2}:=\frac{9-10{\mathtt{c}}_{\mathtt{h}}^{4}+9{\mathtt{c}}_{\mathtt{h}}^{8}}{16{\mathtt{c}}_{\mathtt{h}}^{7}}+{\frac{(1-{\mathtt{c}}_{\mathtt{h}}^{4})}{2{\mathtt{c}}_{\mathtt{h}}}}\eta_{2}^{[0]}=\frac{-2{\mathtt{c}}_{\mathtt{h}}^{12}+13{\mathtt{c}}_{\mathtt{h}}^{8}-12{\mathtt{c}}_{\mathtt{h}}^{4}+9}{16{\mathtt{c}}_{\mathtt{h}}^{7}}\,. (2.8)

More precisely for any σ0\sigma\geq 0 and s>52s>\frac{5}{2}, there exists ϵ>0\epsilon_{*}>0 such that the map ϵ(ηϵ,ψϵ,cϵ)\epsilon\mapsto(\eta_{\epsilon},\psi_{\epsilon},c_{\epsilon}) is analytic from B(ϵ)H𝚎𝚟σ,s(𝕋)×H𝚘𝚍𝚍σ,s(𝕋)×B(\epsilon_{*})\to H^{\sigma,s}_{\mathtt{ev}}(\mathbb{T})\times H^{\sigma,s}_{\mathtt{odd}}(\mathbb{T})\times\mathbb{R}, where H𝚎𝚟σ,s(𝕋)H^{\sigma,s}_{\mathtt{ev}}(\mathbb{T}), respectively H𝚘𝚍𝚍σ,s(𝕋)H^{\sigma,s}_{\mathtt{odd}}(\mathbb{T}), denote the space of even, respectively odd, real valued 2π2\pi-periodic analytic functions u(x)=kukeikxu(x)=\sum_{k\in\mathbb{Z}}u_{k}e^{\mathrm{i}\,kx} such that uσ,s2:=k|uk|2k2se2σ|k|<+\|u\|_{\sigma,s}^{2}:=\sum_{k\in\mathbb{Z}}|u_{k}|^{2}\langle k\rangle^{2s}e^{2\sigma|k|}<+\infty.

The expansions (2.6)-(2.8) are derived in the Appendix B for completeness, although present in the literature (they coincide with [42, section 13, chapter 13] and [2, section 2]). Note that in the shallow water regime 𝚑0+\mathtt{h}\to 0^{+} the expansions (2.6)-(2.8) become singular. For the analiticity properties of the maps stated in Theorem 2.1 we refer to [8].

We also mention that more general time quasi-periodic traveling Stokes waves – which are nonlinear superpositions of multiple Stokes waves traveling with rationally independent speeds – have been recently proved for (2.1) in [5] in finite depth, in [16] in infinite depth, and in [4] for capillary-gravity water waves in any depth.
Linearization at the Stokes waves. In order to determine the stability/instability of the Stokes waves given by Theorem 2.1, we linearize the water waves equations (2.4) with c=cϵc=c_{\epsilon} at (ηϵ(x),ψϵ(x))(\eta_{\epsilon}(x),\psi_{\epsilon}(x)). In the sequel we closely follow [6] pointing out the differences of the finite depth case.

By using the shape derivative formula for the differential dηG(η)[η^]\mathrm{d}_{\eta}G(\eta)[\hat{\eta}] of the Dirichlet-Neumann operator one obtains the autonomous real linear system

[η^tψ^t]=[G(ηϵ)Bx(Vcϵ)G(ηϵ)1+B(Vcϵ)xBx(Vcϵ)BG(ηϵ)B(Vcϵ)x+BG(ηϵ)][η^ψ^]\begin{bmatrix}\hat{\eta}_{t}\\ \hat{\psi}_{t}\end{bmatrix}=\begin{bmatrix}-G(\eta_{\epsilon})B-\partial_{x}\circ(V-c_{\epsilon})&G(\eta_{\epsilon})\\ -1+B(V-c_{\epsilon})\partial_{x}-B\partial_{x}\circ(V-c_{\epsilon})-BG(\eta_{\epsilon})\circ B&-(V-c_{\epsilon})\partial_{x}+BG(\eta_{\epsilon})\end{bmatrix}\begin{bmatrix}\hat{\eta}\\ \hat{\psi}\end{bmatrix} (2.9)

where

V:=V(x):=B(ηϵ)x+(ψϵ)x,B:=B(x):=G(ηϵ)ψϵ+(ψϵ)x(ηϵ)x1+(ηϵ)x2=(ψϵ)xcϵ1+(ηϵ)x2(ηϵ)x.V:=V(x):=-B(\eta_{\epsilon})_{x}+(\psi_{\epsilon})_{x}\,,\ \ B:=B(x):=\frac{G(\eta_{\epsilon})\psi_{\epsilon}+(\psi_{\epsilon})_{x}(\eta_{\epsilon})_{x}}{1+(\eta_{\epsilon})_{x}^{2}}=\frac{(\psi_{\epsilon})_{x}-c_{\epsilon}}{1+(\eta_{\epsilon})_{x}^{2}}(\eta_{\epsilon})_{x}\,. (2.10)

The functions (V,B)(V,B) are the horizontal and vertical components of the velocity field (Φx,Φy)(\Phi_{x},\Phi_{y}) at the free surface. Moreover ϵ(V,B)\epsilon\mapsto(V,B) is analytic as a map B(ϵ0)Hσ,s1(𝕋)×Hσ,s1(𝕋)B(\epsilon_{0})\to H^{\sigma,s-1}(\mathbb{T})\times H^{\sigma,s-1}(\mathbb{T}).

The real system (2.9) is Hamiltonian, i.e. of the form 𝒥𝒜\mathcal{J}\mathcal{A} for a symmetric operator 𝒜=𝒜\mathcal{A}=\mathcal{A}^{\top}, where 𝒜\mathcal{A}^{\top} is the transposed operator with respect the standard real scalar product of L2(𝕋,)×L2(𝕋,)L^{2}(\mathbb{T},\mathbb{R})\times L^{2}(\mathbb{T},\mathbb{R}).

Moreover, since ηϵ\eta_{\epsilon} is even in xx and ψϵ\psi_{\epsilon} is odd in xx, then the functions (V,B)(V,B) are respectively even and odd in xx, and the linear operator in (2.9) is reversible, i.e. it anti-commutes with the involution ρ\rho in (2.3).

Under the time-independent “good unknown of Alinhac” linear transformation

[η^ψ^]:=Z[uv],Z=[10B1],Z1=[10B1],\begin{bmatrix}\hat{\eta}\\ \hat{\psi}\end{bmatrix}:=Z\begin{bmatrix}u\\ v\end{bmatrix}\,,\qquad Z=\begin{bmatrix}1&0\\ B&1\end{bmatrix},\quad Z^{-1}=\begin{bmatrix}1&0\\ -B&1\end{bmatrix}, (2.11)

the system (2.9) assumes the simpler form

[utvt]=~ϵ[uv],~ϵ:=[x(Vcϵ)G(ηϵ)1(Vcϵ)Bx(Vcϵ)x].\begin{bmatrix}u_{t}\\ v_{t}\end{bmatrix}=\widetilde{\mathcal{L}}_{\epsilon}\begin{bmatrix}u\\ v\end{bmatrix},\qquad\widetilde{\mathcal{L}}_{\epsilon}:=\begin{bmatrix}-\partial_{x}\circ(V-c_{\epsilon})&G(\eta_{\epsilon})\\ -1-(V-c_{\epsilon})B_{x}&-(V-c_{\epsilon})\partial_{x}\end{bmatrix}\,. (2.12)

Note that, since the transformation ZZ is symplectic, i.e. Z𝒥Z=𝒥Z^{\top}\mathcal{J}Z=\mathcal{J}, and reversibility preserving, i.e. Zρ=ρZZ\circ\rho=\rho\circ Z, the linear system (2.12) is Hamiltonian and reversible as (2.9).

Next we perform a conformal change of variables to flatten the water surface. Here the finite depth case induces a modification with respect to the deep water case. By [1, Appendix A], there exists a diffeomorphism of 𝕋\mathbb{T}, xx+𝔭(x)x\mapsto x+\mathfrak{p}(x), with a small 2π2\pi-periodic function 𝔭(x)\mathfrak{p}(x), and a small constant 𝚏ϵ\mathtt{f}_{\epsilon}, such that, by defining the associated composition operator (𝔓u)(x):=u(x+𝔭(x))(\mathfrak{P}u)(x):=u(x+\mathfrak{p}(x)), the Dirichlet-Neumann operator writes as [1, Lemma A.5]

G(ηϵ)=x𝔓1tanh((𝚑+𝚏ϵ)|D|)𝔓,G(\eta_{\epsilon})=\partial_{x}\circ\mathfrak{P}^{-1}\circ{\mathcal{H}}\circ\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D|\big{)}\circ\mathfrak{P}\,, (2.13)

where {\mathcal{H}} is the Hilbert transform, i.e. the Fourier multiplier operator

(eijx):=isign(j)eijx,j{0},(1):=0.\mathcal{H}(e^{\mathrm{i}\,jx}):=-\mathrm{i}\,\textup{sign}(j)e^{\mathrm{i}\,jx}\,,\quad\forall j\in\mathbb{Z}\setminus\{0\}\,,\quad\mathcal{H}(1):=0\,.

The function 𝔭(x)\mathfrak{p}(x) and the constant 𝚏ϵ\mathtt{f}_{\epsilon} are determined as a fixed point of (see [1, formula (A.15)])

𝔭=tanh((𝚑+𝚏ϵ)|D|)[ηϵ(x+𝔭(x))],𝚏ϵ:=12π𝕋ηϵ(x+𝔭(x))dx.\mathfrak{p}=\frac{\mathcal{H}}{\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D|\big{)}}[\eta_{\epsilon}(x+\mathfrak{p}(x))]\,,\qquad\mathtt{f}_{\epsilon}:=\frac{1}{2\pi}\int_{\mathbb{T}}\eta_{\epsilon}(x+\mathfrak{p}(x))\mathrm{d}x\,. (2.14)

By the analyticity of the map ϵηϵHσ,s\epsilon\to\eta_{\epsilon}\in H^{\sigma,s}, σ>0\sigma>0, s>1/2s>1/2, the analytic implicit function theorem implies the existence of a solution ϵ𝔭(x):=𝔭ϵ(x)\epsilon\mapsto\mathfrak{p}(x):=\mathfrak{p}_{\epsilon}(x), ϵ𝚏ϵ\epsilon\mapsto\mathtt{f}_{\epsilon}, analytic as a map B(ϵ0)Hs(𝕋)×B(\epsilon_{0})\to H^{s}(\mathbb{T})\times\mathbb{R}. Moreover, since ηϵ\eta_{\epsilon} is even, the function 𝔭(x)\mathfrak{p}(x) is odd. In Appendix B we prove the expansion

𝔭(x)=ϵ𝚌𝚑2sin(x)+ϵ2(1+𝚌𝚑4)(3+𝚌𝚑4)8𝚌𝚑8sin(2x)+𝒪(ϵ3),𝚏ϵ=ϵ2𝚌𝚑434𝚌𝚑2+𝒪(ϵ3).\mathfrak{p}(x)=\epsilon{\mathtt{c}}_{\mathtt{h}}^{-2}\sin(x)+\epsilon^{2}\frac{(1+{\mathtt{c}}_{\mathtt{h}}^{4})(3+{\mathtt{c}}_{\mathtt{h}}^{4})}{8{\mathtt{c}}_{\mathtt{h}}^{8}}\sin(2x)+\mathcal{O}(\epsilon^{3})\,,\quad\mathtt{f}_{\epsilon}=\epsilon^{2}\frac{{\mathtt{c}}_{\mathtt{h}}^{4}-3}{4{\mathtt{c}}_{\mathtt{h}}^{2}}+\mathcal{O}(\epsilon^{3})\,. (2.15)

Under the symplectic and reversibility-preserving map

𝒫:=[(1+𝔭x)𝔓00𝔓],\mathcal{P}:=\begin{bmatrix}(1+\mathfrak{p}_{x})\mathfrak{P}&0\\ 0&\mathfrak{P}\end{bmatrix}\,, (2.16)

the system (2.12) transforms, by (2.13), into the linear system ht=ϵhh_{t}=\mathcal{L}_{\epsilon}h where ϵ\mathcal{L}_{\epsilon} is the Hamiltonian and reversible real operator

ϵ:=𝒫~ϵ𝒫1\displaystyle\mathcal{L}_{\epsilon}:=\mathcal{P}\,\widetilde{\mathcal{L}}_{\epsilon}\,\mathcal{P}^{-1} =[x(𝚌𝚑+pϵ(x))|D|tanh((𝚑+𝚏ϵ)|D|)(1+aϵ(x))(𝚌𝚑+pϵ(x))x]\displaystyle=\begin{bmatrix}\partial_{x}\circ({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))&|D|\tanh((\mathtt{h}+\mathtt{f}_{\epsilon})|D|)\\ -(1+a_{\epsilon}(x))&({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))\partial_{x}\end{bmatrix} (2.17)
=𝒥[1+aϵ(x)(𝚌𝚑+pϵ(x))xx(𝚌𝚑+pϵ(x))|D|tanh((𝚑+𝚏ϵ)|D|)]\displaystyle=\mathcal{J}\begin{bmatrix}1+a_{\epsilon}(x)&-({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))\partial_{x}\\ \partial_{x}\circ({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))&|D|\tanh((\mathtt{h}+\mathtt{f}_{\epsilon})|D|)\end{bmatrix}

where

𝚌𝚑+pϵ(x):=cϵV(x+𝔭(x))1+𝔭x(x),1+aϵ(x):=1+(V(x+𝔭(x))cϵ)Bx(x+𝔭(x))1+𝔭x(x).{\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x):=\displaystyle{\frac{c_{\epsilon}-V(x+\mathfrak{p}(x))}{1+\mathfrak{p}_{x}(x)}}\,,\quad 1+a_{\epsilon}(x):=\displaystyle{\frac{1+(V(x+\mathfrak{p}(x))-c_{\epsilon})B_{x}(x+\mathfrak{p}(x))}{1+\mathfrak{p}_{x}(x)}}\,. (2.18)

By the analiticity results of the functions V,B,𝔭(x)V,B,\mathfrak{p}(x) given above, the functions pϵp_{\epsilon} and aϵa_{\epsilon} are analytic in ϵ\epsilon as maps B(ϵ0)Hs(𝕋)B(\epsilon_{0})\to H^{s}(\mathbb{T}). In the Appendix B we prove the following expansions.

Lemma 2.2.

The analytic functions pϵ(x)p_{\epsilon}(x) and aϵ(x)a_{\epsilon}(x) in (2.18) are even in xx, and

pϵ(x)=ϵp1(x)+ϵ2p2(x)+𝒪(ϵ3),aϵ(x)=ϵa1(x)+ϵ2a2(x)+𝒪(ϵ3),p_{\epsilon}(x)=\epsilon p_{1}(x)+\epsilon^{2}p_{2}(x)+\mathcal{O}(\epsilon^{3})\,,\qquad a_{\epsilon}(x)=\epsilon a_{1}(x)+\epsilon^{2}a_{2}(x)+\mathcal{O}(\epsilon^{3})\,, (2.19)

where

p1(x)\displaystyle p_{1}(x) =p1[1]cos(x),p1[1]:=2𝚌𝚑1,\displaystyle=p_{1}^{[1]}\cos(x)\,,\qquad\quad\quad p_{1}^{[1]}:=-2{\mathtt{c}}_{\mathtt{h}}^{-1}\,, (2.20)
p2(x)\displaystyle p_{2}(x) =p2[0]+p2[2]cos(2x),p2[0]:=9+12𝚌𝚑4+5𝚌𝚑82𝚌𝚑1216𝚌𝚑7,p2[2]:=3+𝚌𝚑42𝚌𝚑7,\displaystyle=p_{2}^{[0]}+p_{2}^{[2]}\cos(2x)\,,\quad p_{2}^{[0]}:=\frac{9+12{\mathtt{c}}_{\mathtt{h}}^{4}+5{\mathtt{c}}_{\mathtt{h}}^{8}-2{\mathtt{c}}_{\mathtt{h}}^{12}}{16{\mathtt{c}}_{\mathtt{h}}^{7}}\,,\quad p_{2}^{[2]}:=-\frac{3+{\mathtt{c}}_{\mathtt{h}}^{4}}{2{\mathtt{c}}_{\mathtt{h}}^{7}}\,, (2.21)

and

a1(x)\displaystyle a_{1}(x) =a1[1]cos(x),a1[1]:=(𝚌𝚑2+𝚌𝚑2),\displaystyle=a_{1}^{[1]}\cos(x)\,,\qquad\qquad a_{1}^{[1]}:=-({\mathtt{c}}_{\mathtt{h}}^{2}+{\mathtt{c}}_{\mathtt{h}}^{-2})\,, (2.22)
a2(x)\displaystyle a_{2}(x) =a2[0]+a2[2]cos(2x),a2[0]:=32+12𝚌𝚑4,a2[2]:=14𝚌𝚑4+9𝚌𝚑834𝚌𝚑8.\displaystyle=a_{2}^{[0]}+a_{2}^{[2]}\cos(2x)\,,\quad a_{2}^{[0]}:=\frac{3}{2}+\frac{1}{2{\mathtt{c}}_{\mathtt{h}}^{4}}\,,\quad a_{2}^{[2]}:=\frac{-14{\mathtt{c}}_{\mathtt{h}}^{4}+9{\mathtt{c}}_{\mathtt{h}}^{8}-3}{4{\mathtt{c}}_{\mathtt{h}}^{8}}\,. (2.23)

Bloch-Floquet expansion. Since the operator ϵ\mathcal{L}_{\epsilon} in (2.17) has 2π2\pi-periodic coefficients, Bloch-Floquet theory guarantees that

σL2()(ϵ)=μ[12,12)σL2(𝕋)(μ,ϵ)whereμ,ϵ:=eiμxϵeiμx.\sigma_{L^{2}(\mathbb{R})}(\mathcal{L}_{\epsilon})=\bigcup_{\mu\in[-\frac{1}{2},\frac{1}{2})}\sigma_{L^{2}(\mathbb{T})}(\mathcal{L}_{\mu,\epsilon})\qquad\text{where}\quad\qquad\mathcal{L}_{\mu,\epsilon}:=e^{-\mathrm{i}\,\mu x}\,\mathcal{L}_{\epsilon}\,e^{\mathrm{i}\,\mu x}\,.

The domain [12,12)[-\frac{1}{2},\frac{1}{2}) is called, in solid state physics, the “first zone of Brillouin”. In particular, if λ\lambda is an eigenvalue of μ,ϵ\mathcal{L}_{\mu,\epsilon} on L2(𝕋,2)L^{2}(\mathbb{T},\mathbb{C}^{2}) with eigenvector v(x)v(x), then h(t,x)=eλteiμxv(x)h(t,x)=e^{\lambda t}e^{\mathrm{i}\,\mu x}v(x) solves ht=ϵhh_{t}=\mathcal{L}_{\epsilon}h. We remark that:
1. If A=Op(a)A=\mathrm{Op}(a) is a pseudo-differential operator with symbol a(x,ξ)a(x,\xi), which is 2π2\pi periodic in the xx-variable, then Aμ:=eiμxAeiμx=Op(a(x,ξ+μ))A_{\mu}:=e^{-\mathrm{i}\,\mu x}Ae^{\mathrm{i}\,\mu x}=\mathrm{Op}(a(x,\xi+\mu)).
2. If AA is a real operator then Aμ¯=Aμ\overline{A_{\mu}}=A_{-\mu}. As a consequence the spectrum σ(Aμ)=σ(Aμ)¯\sigma(A_{-\mu})=\overline{\sigma(A_{\mu})} and we can study σ(Aμ)\sigma(A_{\mu}) just for μ>0\mu>0. Furthermore σ(Aμ)\sigma(A_{\mu}) is a 1-periodic set with respect to μ\mu, so one can restrict to μ[0,12)\mu\in[0,\frac{1}{2}).

By the previous remarks the Floquet operator associated with the real operator ϵ\mathcal{L}_{\epsilon} in (2.17) is the complex Hamiltonian and reversible operator

μ,ϵ:\displaystyle\mathcal{L}_{\mu,\epsilon}: =[(x+iμ)(𝚌𝚑+pϵ(x))|D+μ|tanh((𝚑+𝚏ϵ)|D+μ|)(1+aϵ(x))(𝚌𝚑+pϵ(x))(x+iμ)]\displaystyle=\begin{bmatrix}(\partial_{x}+\mathrm{i}\,\mu)\circ({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))&|D+\mu|\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D+\mu|\big{)}\\ -(1+a_{\epsilon}(x))&({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))(\partial_{x}+\mathrm{i}\,\mu)\end{bmatrix} (2.24)
=[0IdId0]=𝒥[1+aϵ(x)(𝚌𝚑+pϵ(x))(x+iμ)(x+iμ)(𝚌𝚑+pϵ(x))|D+μ|tanh((𝚑+𝚏ϵ)|D+μ|)]=:μ,ϵ.\displaystyle=\underbrace{\begin{bmatrix}0&\mathrm{Id}\\ -\mathrm{Id}&0\end{bmatrix}}_{\displaystyle{=\mathcal{J}}}\underbrace{\begin{bmatrix}1+a_{\epsilon}(x)&-({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))(\partial_{x}+\mathrm{i}\,\mu)\\ (\partial_{x}+\mathrm{i}\,\mu)\circ({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))&|D+\mu|\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D+\mu|\big{)}\end{bmatrix}}_{\displaystyle{=:\mathcal{B}_{\mu,\epsilon}}}\,.

We regard μ,ϵ\mathcal{L}_{\mu,\epsilon} as an operator with domain H1(𝕋):=H1(𝕋,2)H^{1}(\mathbb{T}):=H^{1}(\mathbb{T},\mathbb{C}^{2}) and range L2(𝕋):=L2(𝕋,2)L^{2}(\mathbb{T}):=L^{2}(\mathbb{T},\mathbb{C}^{2}), equipped with the complex scalar product

(f,g):=12π02π(f1g1¯+f2g2¯)dx,f=[f1f2],g=[g1g2]L2(𝕋,2).(f,g):=\frac{1}{2\pi}\int_{0}^{2\pi}\left(f_{1}\overline{g_{1}}+f_{2}\overline{g_{2}}\right)\,\text{d}x\,,\quad\forall f=\begin{bmatrix}f_{1}\\ f_{2}\end{bmatrix},\ \ g=\begin{bmatrix}g_{1}\\ g_{2}\end{bmatrix}\in L^{2}(\mathbb{T},\mathbb{C}^{2})\,. (2.25)

We also denote f2=(f,f)\|f\|^{2}=(f,f).

The complex operator μ,ϵ\mathcal{L}_{\mu,\epsilon} in (2.24) is Hamiltonian and Reversible, according to the following definition.

Definition 2.3.

(Complex Hamiltonian/Reversible operator) A complex operator :H1(𝕋,2)L2(𝕋,2)\mathcal{L}:H^{1}(\mathbb{T},\mathbb{C}^{2})\to L^{2}(\mathbb{T},\mathbb{C}^{2}) is
(ii) Hamiltonian, if =𝒥\mathcal{L}=\mathcal{J}\mathcal{B} where \mathcal{B} is a self-adjoint operator, namely =\mathcal{B}=\mathcal{B}^{*}, where \mathcal{B}^{*} (with domain H1(𝕋)H^{1}(\mathbb{T})) is the adjoint with respect to the complex scalar product (2.25) of L2(𝕋)L^{2}(\mathbb{T}).
(iiii) Reversible, if

ρ¯=ρ¯,\mathcal{L}\circ\overline{\rho}=-\overline{\rho}\circ\mathcal{L}\,, (2.26)

where ρ¯\overline{\rho} is the complex involution (cfr. (2.3))

ρ¯[η(x)ψ(x)]:=[η¯(x)ψ¯(x)].\overline{\rho}\begin{bmatrix}\eta(x)\\ \psi(x)\end{bmatrix}:=\begin{bmatrix}\overline{\eta}(-x)\\ -\overline{\psi}(-x)\end{bmatrix}\,. (2.27)

The property (2.26) for μ,ϵ\mathcal{L}_{\mu,\epsilon} follows because ϵ\mathcal{L}_{\epsilon} is a real operator which is reversible with respect to the involution ρ\rho in (2.3). Equivalently, since 𝒥ρ¯=ρ¯𝒥\mathcal{J}\circ\overline{\rho}=-\overline{\rho}\circ\mathcal{J}, the self-adjoint operator μ,ϵ\mathcal{B}_{\mu,\epsilon} is reversibility-preserving, i.e.

μ,ϵρ¯=ρ¯μ,ϵ.\mathcal{B}_{\mu,\epsilon}\circ\overline{\rho}=\overline{\rho}\circ\mathcal{B}_{\mu,\epsilon}\,. (2.28)

In addition (μ,ϵ)μ,ϵ(H1(𝕋),L2(𝕋))(\mu,\epsilon)\to\mathcal{L}_{\mu,\epsilon}\in\mathcal{L}(H^{1}(\mathbb{T}),L^{2}(\mathbb{T})) is analytic, since the functions ϵaϵ\epsilon\mapsto a_{\epsilon}, pϵp_{\epsilon} defined in (2.19) are analytic as maps B(ϵ0)H1(𝕋)B(\epsilon_{0})\to H^{1}(\mathbb{T}) and μ,ϵ{\mathcal{L}}_{\mu,\epsilon} is analytic with respect to μ\mu, since, for any μ[12,12)\mu\in[-\frac{1}{2},\frac{1}{2}),

|D+μ|tanh((𝚑+𝚏ϵ)|D+μ|)=(D+μ)tanh((𝚑+𝚏ϵ)(D+μ)).|D+\mu|\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D+\mu|\big{)}=(D+\mu)\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})(D+\mu)\big{)}\,. (2.29)

We also note that (see [35, Section 5.1])

|D+μ|=|D|+μ(sgn(D)+Π0),μ>0,|D+\mu|=|D|+\mu(\operatorname*{sgn}(D)+\Pi_{0})\,,\quad\forall\mu>0\,, (2.30)

where sgn(D)\operatorname*{sgn}(D) is the Fourier multiplier operator, acting on 2π2\pi-periodic functions, with symbol

sgn(k):=1k>0,sgn(0):=0,sgn(k):=1k<0,\operatorname*{sgn}(k):=1\ \forall k>0\,,\quad\operatorname*{sgn}(0):=0\,,\quad\operatorname*{sgn}(k):=-1\ \forall k<0\,, (2.31)

and Π0\Pi_{0} is the projector operator on the zero mode, Π0f(x):=12π𝕋f(x)dx.\Pi_{0}f(x):=\frac{1}{2\pi}\int_{\mathbb{T}}f(x)\mathrm{d}x.

Remark 2.4.

If (η(x),ψ(x),c)(\eta(x),\psi(x),c) solve the traveling wave equations (2.5) then the rescaled functions (η~(x),ψ~(x),c~):=(η(x),gψ(x),gc)(\widetilde{\eta}(x),\widetilde{\psi}(x),\widetilde{c}):=(\eta(x),\sqrt{g}\psi(x),\sqrt{g}c) solve the same equations with gravity constant gg instead of 11. The eigenvalues of the corresponding linearized operators (2.9) and (2.24) for a general gravity gg are those of the g=1g=1 case multiplied by g\sqrt{g}.

Our aim is to prove the existence of eigenvalues of μ,ϵ\mathcal{L}_{\mu,\epsilon} in (2.24) with non zero real part. We remark that the Hamiltonian structure of μ,ϵ\mathcal{L}_{\mu,\epsilon} implies that eigenvalues with non zero real part may arise only from multiple eigenvalues of μ,0\mathcal{L}_{\mu,0} (“Krein criterion”), because if λ\lambda is an eigenvalue of μ,ϵ\mathcal{L}_{\mu,\epsilon} then also λ¯-\overline{\lambda} is, and the total algebraic multiplicity of the eigenvalues is conserved under small perturbation. We now describe the spectrum of μ,0\mathcal{L}_{\mu,0}.
The spectrum of μ,0\mathcal{L}_{\mu,0}. The spectrum of the Fourier multiplier matrix operator

μ,0=[𝚌𝚑(x+iμ)|D+μ|tanh(𝚑|D+μ|)1𝚌𝚑(x+iμ)]\mathcal{L}_{\mu,0}=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}(\partial_{x}+\mathrm{i}\,\mu)&|D+\mu|\,\tanh\big{(}\mathtt{h}|D+\mu|\big{)}\\ -1&{\mathtt{c}}_{\mathtt{h}}(\partial_{x}+\mathrm{i}\,\mu)\end{bmatrix} (2.32)

consists of the purely imaginary eigenvalues {λk±(μ),k}\{\lambda_{k}^{\pm}(\mu)\;,\;k\in\mathbb{Z}\}, where

λk±(μ):=i(𝚌𝚑(±k+μ)|k±μ|tanh(𝚑|k±μ|)).\lambda_{k}^{\pm}(\mu):=\mathrm{i}\,\big{(}{\mathtt{c}}_{\mathtt{h}}(\pm k+\mu)\mp\sqrt{|k\pm\mu|\tanh(\mathtt{h}|k\pm\mu|)}\big{)}\,. (2.33)

For μ=0\mu=0 the real operator 0,0\mathcal{L}_{0,0} possesses the eigenvalue 0 with algebraic multiplicity 44,

λ0+(0)=λ0(0)=λ1+(0)=λ1(0)=0,\lambda_{0}^{+}(0)=\lambda_{0}^{-}(0)=\lambda_{1}^{+}(0)=\lambda_{1}^{-}(0)=0\,,

and geometric multiplicity 33. A real basis of the Kernel of 0,0\mathcal{L}_{0,0} is

f1+:=[𝚌𝚑1/2cos(x)𝚌𝚑1/2sin(x)],f1:=[𝚌𝚑1/2sin(x)𝚌𝚑1/2cos(x)],f0:=[01],\displaystyle f_{1}^{+}:=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{1/2}\cos(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-1/2}\sin(x)\end{bmatrix},\quad f_{1}^{-}:=\begin{bmatrix}-{\mathtt{c}}_{\mathtt{h}}^{1/2}\sin(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-1/2}\cos(x)\end{bmatrix},\qquad f_{0}^{-}:=\begin{bmatrix}0\\ 1\end{bmatrix}\,, (2.34)

together with the generalized eigenvector

f0+:=[10],0,0f0+=f0.\displaystyle f_{0}^{+}:=\begin{bmatrix}1\\ 0\end{bmatrix},\qquad\mathcal{L}_{0,0}f_{0}^{+}=-f_{0}^{-}\,. (2.35)

Furthermore 0 is an isolated eigenvalue for 0,0\mathcal{L}_{0,0}, namely the spectrum σ(0,0)\sigma\left(\mathcal{L}_{0,0}\right) decomposes in two separated parts

σ(0,0)=σ(0,0)σ′′(0,0)whereσ(0,0):={0}\sigma\left(\mathcal{L}_{0,0}\right)=\sigma^{\prime}\left(\mathcal{L}_{0,0}\right)\cup\sigma^{\prime\prime}\left(\mathcal{L}_{0,0}\right)\quad\text{where}\quad\sigma^{\prime}(\mathcal{L}_{0,0}):=\{0\} (2.36)

and

σ′′(0,0):={λkσ(0),k=0,1,σ=±}.\sigma^{\prime\prime}(\mathcal{L}_{0,0}):=\big{\{}\lambda_{k}^{\sigma}(0),\ k=0,1\,,\sigma=\pm\big{\}}\,.

We shall also use that, as proved in Theorem 4.1 in [35], the operator 0,ϵ{\mathcal{L}}_{0,\epsilon} possesses, for any sufficiently small ϵ0\epsilon\neq 0, the eigenvalue 0 with a four dimensional generalized Kernel, spanned by ϵ\epsilon-dependent vectors U1,U~2,U3,U4U_{1},\tilde{U}_{2},U_{3},U_{4} satisfying, for some real constant αϵ,βϵ\alpha_{\epsilon},\beta_{\epsilon},

0,ϵU1=0,0,ϵU~2=0,0,ϵU3=αϵU~2,0,ϵU4=U1βϵU~2,U1:=[01].{\mathcal{L}}_{0,\epsilon}U_{1}=0\,,\ \ {\mathcal{L}}_{0,\epsilon}\tilde{U}_{2}=0\,,\ \ {\mathcal{L}}_{0,\epsilon}U_{3}=\alpha_{\epsilon}\,\tilde{U}_{2}\,,\ \ {\mathcal{L}}_{0,\epsilon}U_{4}=-U_{1}-\beta_{\epsilon}\tilde{U}_{2}\,,\quad U_{1}:=\begin{bmatrix}0\\ 1\end{bmatrix}\,. (2.37)

By Kato’s perturbation theory (see Lemma 3.1 below) for any μ,ϵ0\mu,\epsilon\neq 0 sufficiently small, the perturbed spectrum σ(μ,ϵ)\sigma\left(\mathcal{L}_{\mu,\epsilon}\right) admits a disjoint decomposition as

σ(μ,ϵ)=σ(μ,ϵ)σ′′(μ,ϵ),\sigma\left(\mathcal{L}_{\mu,\epsilon}\right)=\sigma^{\prime}\left(\mathcal{L}_{\mu,\epsilon}\right)\cup\sigma^{\prime\prime}\left(\mathcal{L}_{\mu,\epsilon}\right)\,, (2.38)

where σ(μ,ϵ)\sigma^{\prime}\left(\mathcal{L}_{\mu,\epsilon}\right) consists of 4 eigenvalues close to 0. We denote by 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon} the spectral subspace associated with σ(μ,ϵ)\sigma^{\prime}\left(\mathcal{L}_{\mu,\epsilon}\right), which has dimension 4 and it is invariant by μ,ϵ\mathcal{L}_{\mu,\epsilon}. Our goal is to prove that, for ϵ\epsilon small, for values of the Floquet exponent μ\mu in an interval of order ϵ\epsilon, the 4×44\times 4 matrix which represents the operator μ,ϵ:𝒱μ,ϵ𝒱μ,ϵ\mathcal{L}_{\mu,\epsilon}:\mathcal{V}_{\mu,\epsilon}\to\mathcal{V}_{\mu,\epsilon} possesses a pair of eigenvalues close to zero with opposite non zero real parts.

Before stating our main result, let us introduce a notation we shall use through all the paper:

  • \bullet

    Notation: we denote by 𝒪(μm1ϵn1,,μmpϵnp)\mathcal{O}(\mu^{m_{1}}\epsilon^{n_{1}},\dots,\mu^{m_{p}}\epsilon^{n_{p}}), mj,njm_{j},n_{j}\in\mathbb{N} (for us :={1,2,}\mathbb{N}:=\{1,2,\dots\}), analytic functions of (μ,ϵ)(\mu,\epsilon) with values in a Banach space XX which satisfy, for some C>0C>0 uniform for 𝚑\mathtt{h} in any compact set of (0,+)(0,+\infty), the bound 𝒪(μmjϵnj)XCj=1p|μ|mj|ϵ|nj\|\mathcal{O}(\mu^{m_{j}}\epsilon^{n_{j}})\|_{X}\leq C\sum_{j=1}^{p}|\mu|^{m_{j}}|\epsilon|^{n_{j}} for small values of (μ,ϵ)(\mu,\epsilon). Similarly we denote rk(μm1ϵn1,,μmpϵnp)r_{k}(\mu^{m_{1}}\epsilon^{n_{1}},\dots,\mu^{m_{p}}\epsilon^{n_{p}}) scalar functions 𝒪(μm1ϵn1,,μmpϵnp)\mathcal{O}(\mu^{m_{1}}\epsilon^{n_{1}},\dots,\mu^{m_{p}}\epsilon^{n_{p}}) which are also real analytic.

Our complete spectral result is the following:

Theorem 2.5.

(Complete Benjamin-Feir spectrum) There exist ϵ0,μ0>0\epsilon_{0},\mu_{0}>0, uniformly for the depth 𝚑\mathtt{h} in any compact set of (0,+)(0,+\infty), such that, for any 0<μ<μ00\,<\,\mu<\mu_{0} and 0ϵ<ϵ00\leq\epsilon<\epsilon_{0}, the operator μ,ϵ:𝒱μ,ϵ𝒱μ,ϵ\mathcal{L}_{\mu,\epsilon}:\mathcal{V}_{\mu,\epsilon}\to\mathcal{V}_{\mu,\epsilon} can be represented by a 4×44\times 4 matrix of the form

(𝚄00𝚂),\begin{pmatrix}\mathtt{U}&\vline&0\\ \hline\cr 0&\vline&\mathtt{S}\end{pmatrix}, (2.39)

where 𝚄\mathtt{U} and 𝚂\mathtt{S} are 2×22\times 2 matrices, with identical diagonal entries each, of the form

𝚄=(i((𝚌𝚑12𝚎12)μ+r2(μϵ2,μ2ϵ,μ3))𝚎22μ8(1+r5(ϵ,μ))μϵ2𝚎WB+r1(μϵ3,μ2ϵ2)+𝚎22μ38(1+r1′′(ϵ,μ))i((𝚌𝚑12𝚎12)μ+r2(μϵ2,μ2ϵ,μ3))),\displaystyle\mathtt{U}={\begin{pmatrix}\mathrm{i}\,\big{(}({\mathtt{c}}_{\mathtt{h}}-\tfrac{1}{2}\mathtt{e}_{12})\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}&-\mathtt{e}_{22}\frac{\mu}{8}(1+r_{5}(\epsilon,\mu))\\ -\mu\epsilon^{2}\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}+r_{1}^{\prime}(\mu\epsilon^{3},\mu^{2}\epsilon^{2})+\mathtt{e}_{22}\frac{\mu^{3}}{8}(1+r_{1}^{\prime\prime}(\epsilon,\mu))&\mathrm{i}\,\big{(}({\mathtt{c}}_{\mathtt{h}}-\tfrac{1}{2}\mathtt{e}_{12})\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}\end{pmatrix}}\,,
𝚂=(i𝚌𝚑μ+ir9(μϵ2,μ2ϵ)tanh(𝚑μ)+r10(μϵ)μ+r8(μϵ2,μ3ϵ)i𝚌𝚑μ+ir9(μϵ2,μ2ϵ)),\displaystyle\mathtt{S}=\begin{pmatrix}\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu+\mathrm{i}\,{r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)}&\tanh(\mathtt{h}\mu)+{r_{10}(\mu\epsilon)}\\ -\mu+{r_{8}(\mu\epsilon^{2},\mu^{3}\epsilon)}&\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu+\mathrm{i}\,{r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)}\end{pmatrix}\,, (2.40)

where 𝚎WB\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}, 𝚎12,𝚎22\mathtt{e}_{12},\mathtt{e}_{22} are defined in (1.1), (1.2), (1.3). The eigenvalues of 𝚄\mathtt{U} have the form

λ1±(μ,ϵ)\displaystyle\lambda_{1}^{\pm}(\mu,\epsilon) =i12𝚌˘𝚑μ+ir2(μϵ2,μ2ϵ,μ3)±18μ𝚎22(𝚑)(1+r5(ϵ,μ))ΔBF(𝚑;μ,ϵ),\displaystyle=\mathrm{i}\,\frac{1}{2}\breve{\mathtt{c}}_{\mathtt{h}}\mu+\mathrm{i}\,r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\pm\tfrac{1}{8}\mu\sqrt{\mathtt{e}_{22}(\mathtt{h})(1+r_{5}(\epsilon,\mu))}\sqrt{\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon)}\,, (2.41)

where 𝚌˘𝚑:=2𝚌𝚑𝚎12(𝚑)\breve{\mathtt{c}}_{\mathtt{h}}:=2{\mathtt{c}}_{\mathtt{h}}-\mathtt{e}_{12}(\mathtt{h}) and ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon) is the Benjamin-Feir discriminant function (1.6) (with r1(ϵ3,μϵ2):=8r1(ϵ3,μϵ2)r_{1}(\epsilon^{3},\mu\epsilon^{2}):=-8r_{1}^{\prime}(\epsilon^{3},\mu\epsilon^{2})). As 𝚎22(𝚑)>0\mathtt{e}_{22}(\mathtt{h})>0, they have non-zero real part if and only if ΔBF(𝚑;μ,ϵ)>0\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon)>0.

The eigenvalues of the matrix 𝚂\mathtt{S} are a pair of purely imaginary eigenvalues of the form

λ0±(μ,ϵ)=i𝚌𝚑μ(1+r9(ϵ2,μϵ))iμtanh(𝚑μ)(1+r(ϵ)).\lambda_{0}^{\pm}(\mu,\epsilon)=\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu\big{(}1+{r_{9}(\epsilon^{2},\mu\epsilon)\big{)}}\mp\mathrm{i}\,\sqrt{\mu\tanh(\mathtt{h}\mu)}\big{(}1+{r(\epsilon)}\big{)}\,. (2.42)

For ϵ=0\epsilon=0 the eigenvalues λ1±(μ,0),λ0±(μ,0)\lambda_{1}^{\pm}(\mu,0),\lambda_{0}^{\pm}(\mu,0) coincide with those in (2.33).

Remark 2.6.

At ϵ=0\epsilon=0, the eigenvalues in (2.41) have the Taylor expansion

λ1±(μ,0)=i(𝚌𝚑12𝚎12(𝚑))μ±i𝚎22(𝚑)8μ2+𝒪(μ3),\lambda^{\pm}_{1}(\mu,0)=\mathrm{i}\,({\mathtt{c}}_{\mathtt{h}}-\frac{1}{2}\mathtt{e}_{12}(\mathtt{h}))\mu\pm\mathrm{i}\,\frac{\mathtt{e}_{22}(\mathtt{h})}{8}\mu^{2}+\mathcal{O}(\mu^{3})\,,

which coincides with the one of λ1±(μ)\lambda^{\pm}_{1}(\mu) in (2.33), in view of the coefficients 𝚎12(𝚑)\mathtt{e}_{12}(\mathtt{h}) and 𝚎22(𝚑)\mathtt{e}_{22}(\mathtt{h}) defined in (1.2), (1.3).

We conclude this section describing in detail our approach.
Ideas and scheme of proof. The proof follows the general ideas of the infinitely deep water case [6], although important differences arise in finite depth and require a different approach. The first step is to exploit Kato’s theory to prolong the unperturbed symplectic basis {f1±,f0±}\{f_{1}^{\pm},f_{0}^{\pm}\} of 𝒱0,0\mathcal{V}_{0,0} in (2.34)-(2.35) into a symplectic basis of the spectral subspace 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon} associated with σ(μ,ϵ)\sigma^{\prime}\left(\mathcal{L}_{\mu,\epsilon}\right) in (2.38), depending analytically on μ,ϵ\mu,\epsilon. The transformation operator Uμ,ϵU_{\mu,\epsilon} in Lemma 3.1 is symplectic, analytic in μ,ϵ\mu,\epsilon, and maps isomorphically 𝒱0,0\mathcal{V}_{0,0} into 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon}. The vectors fkσ(μ,ϵ):=Uμ,ϵfkσf^{\sigma}_{k}(\mu,\epsilon):=U_{\mu,\epsilon}f_{k}^{\sigma}, k=0,1k=0,1, σ=±\sigma=\pm, are the required symplectic basis of the symplectic subspace 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon}. Its expansion in μ,ϵ\mu,\epsilon is provided in Lemma 4.2. This procedure reduces our spectral problem to determine the eigenvalues of the 4×44\times 4 Hamiltonian and reversible matrix 𝙻μ,ϵ\mathtt{L}_{\mu,\epsilon} (cfr. Lemma 3.4), representing the action of the operator μ,ϵi𝚌𝚑μ\mathcal{L}_{\mu,\epsilon}-\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu on the basis {fkσ(μ,ϵ)}\{f_{k}^{\sigma}(\mu,\epsilon)\}. In Proposition 4.3 we prove that

𝙻μ,ϵ=𝙹4(EFFG)=(𝙹2E𝙹2F𝙹2F𝙹2G)where𝙹4=(𝙹200𝙹2),𝙹2=(0110),\mathtt{L}_{\mu,\epsilon}=\mathtt{J}_{4}\begin{pmatrix}E&F\\ F^{*}&G\end{pmatrix}=\begin{pmatrix}\mathtt{J}_{2}E&\mathtt{J}_{2}F\\ \mathtt{J}_{2}F^{*}&\mathtt{J}_{2}G\end{pmatrix}\qquad\text{where}\qquad\mathtt{J}_{4}=\begin{pmatrix}\mathtt{J}_{2}&0\\ 0&\mathtt{J}_{2}\end{pmatrix}\,,\ \ \mathtt{J}_{2}=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\,, (2.43)

and the 2×22\times 2 matrices E,G,FE,G,F have the expansions (4.10)-(4.12). In finite depth this computation is much more involved than in deep water, as we need to track the exact dependence of the matrix entries with respect to 𝚑\mathtt{h}. In particular the matrix EE is

E=(𝚎11ϵ2(1+r1(ϵ,μϵ))𝚎22μ28(1+r1′′(ϵ,μ))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))𝚎22μ28(1+r5(ϵ,μ)))E=\begin{pmatrix}\mathtt{e}_{11}\epsilon^{2}(1+r_{1}^{\prime}(\epsilon,\mu\epsilon))-\mathtt{e}_{22}\frac{\mu^{2}}{8}(1+r_{1}^{\prime\prime}(\epsilon,\mu))&\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}\\ -\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}&-\mathtt{e}_{22}\frac{\mu^{2}}{8}(1+r_{5}(\epsilon,\mu))\end{pmatrix} (2.44)

where the coefficients 𝚎11\mathtt{e}_{11} and 𝚎22\mathtt{e}_{22}, defined in (4.13) and (1.3), are strictly positive for any value of 𝚑>0\mathtt{h}>0. Thus the submatrix 𝙹2E\mathtt{J}_{2}E has a pair of eigenvalues with nonzero real part, for any value of 𝚑>0\mathtt{h}>0, provided 0<μ<μ¯(ϵ)ϵ0<\mu<\overline{\mu}(\epsilon)\sim\epsilon. On the other hand, it has to come out that the complete 4×44\times 4 matrix 𝙻μ,ϵ\mathtt{L}_{\mu,\epsilon} possesses unstable eigenvalues if and only if the depth exceeds the celebrated Whitham-Benjamin threshold 𝚑WB1.363\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}\sim 1.363\ldots. Indeed the correct eigenvalues of 𝙻μ,ϵ\mathtt{L}_{\mu,\epsilon} are not a small perturbation of those of (𝙹2E00𝙹2G)\footnotesize{\begin{pmatrix}\mathtt{J}_{2}E&0\\ 0&\mathtt{J}_{2}G\end{pmatrix}} and will emerge only after one non-perturbative step of block diagonalization. This was not the case in the infinitely deep water case [6], where at this stage the corresponding submatrix 𝙹2E\mathtt{J}_{2}E had already the correct Benjamin-Feir eigenvalues, and we only had to check their stability under perturbation.

Remark 2.7.

We also stress that (2.44) is not a simple Taylor expansion in μ,ϵ\mu,\epsilon: note that the (2,2)(2,2)-entry in (2.44) does not have any term 𝒪(ϵm)\mathcal{O}(\epsilon^{m}) nor 𝒪(μϵm)\mathcal{O}(\mu\epsilon^{m}) for any mm\in\mathbb{N}. These terms would be dangerous because they could change the sign of the entry (2,2)(2,2) which instead, in (2.44), is always negative (recall that 𝚎22(𝚑)>0\mathtt{e}_{22}(\mathtt{h})>0). We prove the absence of terms ϵm\epsilon^{m}, mm\in\mathbb{N}, fully exploiting (as in [6]) the structural information (2.37) concerning the four dimensional generalized Kernel of the operator 0,ϵ\mathcal{L}_{0,\epsilon} for any ϵ>0\epsilon>0, see Lemma 4.4. Moreover, in finite depth it turns out that there are no terms of order μϵm\mu\epsilon^{m}, mm\in\mathbb{N}, which instead are present in deep water, and were eliminated in [6] via a further change of basis. We also note that the 2×22\times 2 matrices 𝙹2E\mathtt{J}_{2}E and 𝙹2G\mathtt{J}_{2}G in (2.43) have both eigenvalues of size 𝒪(μ)\mathcal{O}(\mu). As already mentioned in the introduction, this is a crucial difference with the deep water case, where the eigenvalues of 𝙹2G\mathtt{J}_{2}G have the much larger size 𝒪(μ)\mathcal{O}(\sqrt{\mu}).

In order to determine the correct spectrum of the matrix 𝙻μ,ϵ\mathtt{L}_{\mu,\epsilon} in (2.43), we perform a block diagonalization of 𝙻μ,ϵ\mathtt{L}_{\mu,\epsilon} to eliminate the coupling term 𝙹2F\mathtt{J}_{2}F (which has size ϵ\epsilon, see (4.12)). We proceed, in Section 5, in three steps:
1. Symplectic rescaling. We first perform a symplectic rescaling which is singular at μ=0\mu=0, see Lemma 5.1, obtaining the matrix 𝙻μ,ϵ(1)\mathtt{L}_{\mu,\epsilon}^{(1)}. The effects are twofold: (i) the diagonal elements of

E(1)=(𝚎11μϵ2(1+r1(ϵ,μϵ))𝚎22μ38(1+r1′′(ϵ,μ))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))𝚎22μ8(1+r5(ϵ,μ)))E^{(1)}=\begin{pmatrix}\mathtt{e}_{11}\mu\epsilon^{2}(1+r_{1}^{\prime}(\epsilon,\mu\epsilon))-\mathtt{e}_{22}\frac{\mu^{3}}{8}(1+r_{1}^{\prime\prime}(\epsilon,\mu))&\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}\\ -\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}&-\mathtt{e}_{22}\frac{\mu}{8}(1+r_{5}(\epsilon,\mu))\end{pmatrix} (2.45)

have size 𝒪(μ)\mathcal{O}(\mu), as well as those of G(1)G^{(1)}, and (ii) the matrix F(1)F^{(1)} has the smaller size 𝒪(μϵ){\mathcal{O}(\mu\epsilon)}.
2. Non-perturbative step of block-diagonalization (Section 5.1). Inspired by KAM theory, we perform one step of block decoupling to decrease further the size of the off-diagonal blocks. This step modifies the matrix 𝙹2E(1)\mathtt{J}_{2}E^{(1)} in a substantial way, by a term 𝒪(μϵ2)\mathcal{O}(\mu\epsilon^{2}). Let us explain better this step. In order to reduce the size of 𝙹2F(1)\mathtt{J}_{2}F^{(1)}, we conjugate 𝙻μ,ϵ(1)\mathtt{L}_{\mu,\epsilon}^{(1)} by the symplectic matrix exp(S(1))\exp(S^{(1)}), where S(1)S^{(1)} is a Hamiltonian matrix with the same form of 𝙹2F(1)\mathtt{J}_{2}F^{(1)}, see (5.9). The transformed matrix 𝙻μ,ϵ(2)=exp(S(1))𝙻μ,ϵ(1)exp(S(1))\mathtt{L}_{\mu,\epsilon}^{(2)}=\exp(S^{(1)})\mathtt{L}_{\mu,\epsilon}^{(1)}\exp(-S^{(1)}) has the Lie expansion222recall that exp(S)Lexp(S)=n01n!adSn(L)\exp(S)L\exp(-S)=\sum_{n\geq 0}\frac{1}{n!}\textup{ad}_{S}^{n}(L), where adS0(L):=L\textup{ad}_{S}^{0}(L):=L, adSn(L)=[S,adSn1(L)]\textup{ad}_{S}^{n}(L)=[S,\textup{ad}_{S}^{n-1}(L)] for n1n\geq 1.

𝙻μ,ϵ(2)\displaystyle\mathtt{L}_{\mu,\epsilon}^{(2)} =(𝙹2E(1)00𝙹2G(1))\displaystyle=\begin{pmatrix}\mathtt{J}_{2}E^{(1)}&0\\ 0&\mathtt{J}_{2}G^{(1)}\end{pmatrix} (2.46)
+(0𝙹2F(1)𝙹2[F(1)]0)+[S(1),(𝙹2E(1)00𝙹2G(1))]\displaystyle\quad+\begin{pmatrix}0&\mathtt{J}_{2}F^{(1)}\\ \mathtt{J}_{2}[F^{(1)}]^{*}&0\end{pmatrix}+\left[S^{(1)}\,,\,\begin{pmatrix}\mathtt{J}_{2}E^{(1)}&0\\ 0&\mathtt{J}_{2}G^{(1)}\end{pmatrix}\right]
+12[S(1),[S(1),(𝙹2E(1)00𝙹2G(1))]]+[S(1),(0𝙹2F(1)𝙹2[F(1)]0)]+h.o.t.\displaystyle\quad+\frac{1}{2}\Big{[}S^{(1)},\Big{[}S^{(1)},\begin{pmatrix}\mathtt{J}_{2}E^{(1)}&0\\ 0&\mathtt{J}_{2}G^{(1)}\end{pmatrix}\Big{]}\Big{]}+\Big{[}S^{(1)},\begin{pmatrix}0&\mathtt{J}_{2}F^{(1)}\\ \mathtt{J}_{2}[F^{(1)}]^{*}&0\end{pmatrix}\Big{]}+\mbox{h.o.t.}

The first line in the right hand side of (2.46) is the previous block-diagonal matrix, the second line of (2.46) is a purely off-diagonal matrix and the third line is the sum of two block-diagonal matrices and “h.o.t.” collects terms of much smaller size. S(1)S^{(1)} is determined in such a way that the second line of (2.46) vanishes, and therefore the remaining off-diagonal matrices (appearing in the h.o.t. remainder) are smaller in size. Unlike the infinitely deep water case [6], the block-diagonal corrections in the third line of (2.46) are not perturbative and they modify substantially the block-diagonal part. More precisely we obtain that 𝙻μ,ϵ(2)\mathtt{L}_{\mu,\epsilon}^{(2)} has the form (5.10) with

E(2):=(μϵ2𝚎WB+r1(μϵ3,μ2ϵ2)𝚎22μ38(1+r1′′(ϵ,μ))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))𝚎22μ8(1+r5(ϵ,μ))).E^{(2)}:={\begin{pmatrix}\mu\epsilon^{2}\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}+r_{1}^{\prime}(\mu\epsilon^{3},\mu^{2}\epsilon^{2})-\mathtt{e}_{22}\frac{\mu^{3}}{8}(1+r_{1}^{\prime\prime}(\epsilon,\mu))&\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}\\ -\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}&-\mathtt{e}_{22}\frac{\mu}{8}(1+r_{5}(\epsilon,\mu))\end{pmatrix}}\,.

Note the appearance of the Whitham-Benjamin function 𝚎WB(𝚑)\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h}) in the (1,1)-entry of E(2)E^{(2)}, which changes sign at the critical depth 𝚑WB\mathtt{h}_{\scriptscriptstyle{\textsc{WB}}}, see Figure 1, unlike the coefficient 𝚎11(𝚑)>0\mathtt{e}_{11}(\mathtt{h})>0 in (2.45). If 𝚎WB(𝚑)>0\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h})>0 and ϵ\epsilon and μ\mu are sufficiently small, the matrix 𝙹2E(2)\mathtt{J}_{2}E^{(2)} has eigenvalues with non-zero real part (recall that 𝚎22(𝚑)>0\mathtt{e}_{22}(\mathtt{h})>0 for any 𝚑\mathtt{h}). On the contrary, if 𝚎WB(𝚑)<0\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}(\mathtt{h})<0, then the eigenvalues of 𝙹2E(2)\mathtt{J}_{2}E^{(2)} lay on the imaginary axis.
3. Complete block-diagonalization (Section 5.2). In Lemma 5.9 we completely block-diagonalize 𝙻μ,ϵ(2)\mathtt{L}^{(2)}_{\mu,\epsilon} by means of a standard implicit function theorem. By this procedure the original matrix 𝙻μ,ϵ\mathtt{L}_{\mu,\epsilon} is conjugated into the Hamiltonian and reversible matrix (2.39), proving Theorem 2.5.

3 Perturbative approach to the separated eigenvalues

We apply Kato’s similarity transformation theory [26, I-§4-6, II-§4] to study the splitting of the eigenvalues of μ,ϵ\mathcal{L}_{\mu,\epsilon} close to 0 for small values of μ\mu and ϵ\epsilon, following [6]. First of all, it is convenient to decompose the operator μ,ϵ\mathcal{L}_{\mu,\epsilon} in (2.24) as

μ,ϵ=i𝚌𝚑μ+μ,ϵ,μ>0,\mathcal{L}_{\mu,\epsilon}=\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu+\mathscr{L}_{\mu,\epsilon}\,,\qquad\mu>0\,, (3.1)

where, using also (2.30),

μ,ϵ:=[x(𝚌𝚑+pϵ(x))+iμpϵ(x)|D+μ|tanh((𝚑+𝚏ϵ)|D+μ|)(1+aϵ(x))(𝚌𝚑+pϵ(x))x+iμpϵ(x)].\mathscr{L}_{\mu,\epsilon}:=\begin{bmatrix}\partial_{x}\circ({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))+\mathrm{i}\,\mu\,p_{\epsilon}(x)&|D+\mu|\,\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D+\mu|\big{)}\\ -(1+a_{\epsilon}(x))&({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))\partial_{x}+\mathrm{i}\,\mu\,p_{\epsilon}(x)\end{bmatrix}\,. (3.2)

The operator μ,ϵ\mathscr{L}_{\mu,\epsilon} is still Hamiltonian, having the form

μ,ϵ=𝒥μ,ϵ,μ,ϵ:=[1+aϵ(x)(𝚌𝚑+pϵ(x))xiμpϵ(x)x(𝚌𝚑+pϵ(x))+iμpϵ(x)|D+μ|tanh((𝚑+𝚏ϵ)|D+μ|)]\mathscr{L}_{\mu,\epsilon}=\mathcal{J}\,{\mathcal{B}}_{\mu,\epsilon}\,,\quad{\mathcal{B}}_{\mu,\epsilon}:=\begin{bmatrix}1+a_{\epsilon}(x)&-({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))\partial_{x}-\mathrm{i}\,\mu\,p_{\epsilon}(x)\\ \partial_{x}\circ({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))+\mathrm{i}\,\mu\,p_{\epsilon}(x)&|D+\mu|\,\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D+\mu|\big{)}\end{bmatrix} (3.3)

with μ,ϵ{\mathcal{B}}_{\mu,\epsilon} selfadjoint, and it is also reversible, namely it satisfies, by (2.26),

μ,ϵρ¯=ρ¯μ,ϵ,ρ¯ defined in (2.27),\mathscr{L}_{\mu,\epsilon}\circ\overline{\rho}=-\overline{\rho}\circ\mathscr{L}_{\mu,\epsilon}\,,\qquad\overline{\rho}\mbox{ defined in }\eqref{reversibilityappears}\,, (3.4)

whereas μ,ϵ{\mathcal{B}}_{\mu,\epsilon} is reversibility-preserving, i.e. fulfills (2.28). Note also that 0,ϵ{\mathcal{B}}_{0,\epsilon} is a real operator.

The scalar operator i𝚌𝚑μi𝚌𝚑μId\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu\equiv\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu\,\text{Id} just translates the spectrum of μ,ϵ\mathscr{L}_{\mu,\epsilon} along the imaginary axis of the quantity i𝚌𝚑μ\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu, that is, in view of (3.1), σ(μ,ϵ)=i𝚌𝚑μ+σ(μ,ϵ).\sigma({\mathcal{L}}_{\mu,\epsilon})=\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu+\sigma(\mathscr{L}_{\mu,\epsilon})\,. Thus in the sequel we focus on studying the spectrum of μ,ϵ\mathscr{L}_{\mu,\epsilon}.

Note also that 0,ϵ=0,ϵ\mathscr{L}_{0,\epsilon}=\mathcal{L}_{0,\epsilon} for any ϵ0\epsilon\geq 0. In particular 0,0\mathscr{L}_{0,0} has zero as isolated eigenvalue with algebraic multiplicity 4, geometric multiplicity 3 and generalized kernel spanned by the vectors {f1+,f1,f0+,f0}\{f^{+}_{1},f^{-}_{1},f^{+}_{0},f^{-}_{0}\} in (2.34), (2.35). Furthermore its spectrum is separated as in (2.36). For any ϵ0\epsilon\neq 0 small, 0,ϵ\mathscr{L}_{0,\epsilon} has zero as isolated eigenvalue with geometric multiplicity 22, and two generalized eigenvectors satisfying (2.37).

We remark that, in view of (2.30), the operator μ,ϵ\mathscr{L}_{\mu,\epsilon} is analytic with respect to μ\mu. The operator μ,ϵ:YXX\mathscr{L}_{\mu,\epsilon}:Y\subset X\to X has domain Y:=H1(𝕋):=H1(𝕋,2)Y:=H^{1}(\mathbb{T}):=H^{1}(\mathbb{T},\mathbb{C}^{2}) and range X:=L2(𝕋):=L2(𝕋,2)X:=L^{2}(\mathbb{T}):=L^{2}(\mathbb{T},\mathbb{C}^{2}).

Lemma 3.1.

(Kato theory for separated eigenvalues) Let Γ\Gamma be a closed, counterclockwise-oriented curve around 0 in the complex plane separating σ(0,0)={0}\sigma^{\prime}\left(\mathscr{L}_{0,0}\right)=\{0\} and the other part of the spectrum σ′′(0,0)\sigma^{\prime\prime}\left(\mathscr{L}_{0,0}\right) in (2.36). There exist ϵ0,μ0>0\epsilon_{0},\mu_{0}>0 such that for any (μ,ϵ)B(μ0)×B(ϵ0)(\mu,\epsilon)\in B(\mu_{0})\times B(\epsilon_{0}) the following statements hold:
1. The curve Γ\Gamma belongs to the resolvent set of the operator μ,ϵ:YXX\mathscr{L}_{\mu,\epsilon}:Y\subset X\to X defined in (3.2).
2. The operators

Pμ,ϵ:=12πiΓ(μ,ϵλ)1dλ:XYP_{\mu,\epsilon}:=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{\mu,\epsilon}-\lambda)^{-1}\mathrm{d}\lambda:X\to Y (3.5)

are well defined projectors commuting with μ,ϵ\mathscr{L}_{\mu,\epsilon}, i.e. Pμ,ϵ2=Pμ,ϵP_{\mu,\epsilon}^{2}=P_{\mu,\epsilon} and Pμ,ϵμ,ϵ=μ,ϵPμ,ϵP_{\mu,\epsilon}\mathscr{L}_{\mu,\epsilon}=\mathscr{L}_{\mu,\epsilon}P_{\mu,\epsilon}. The map (μ,ϵ)Pμ,ϵ(\mu,\epsilon)\mapsto P_{\mu,\epsilon} is analytic from B(μ0)×B(ϵ0)B({\mu_{0}})\times B({\epsilon_{0}}) to (X,Y)\mathcal{L}(X,Y).
3. The domain YY of the operator μ,ϵ\mathscr{L}_{\mu,\epsilon} decomposes as the direct sum

Y=𝒱μ,ϵKer(Pμ,ϵ),𝒱μ,ϵ:=Rg(Pμ,ϵ)=Ker(IdPμ,ϵ),Y=\mathcal{V}_{\mu,\epsilon}\oplus\text{Ker}(P_{\mu,\epsilon})\,,\quad\mathcal{V}_{\mu,\epsilon}:=\text{Rg}(P_{\mu,\epsilon})=\text{Ker}(\mathrm{Id}-P_{\mu,\epsilon})\,,

of closed invariant subspaces, namely μ,ϵ:𝒱μ,ϵ𝒱μ,ϵ\mathscr{L}_{\mu,\epsilon}:\mathcal{V}_{\mu,\epsilon}\to\mathcal{V}_{\mu,\epsilon}, μ,ϵ:Ker(Pμ,ϵ)Ker(Pμ,ϵ)\mathscr{L}_{\mu,\epsilon}:\text{Ker}(P_{\mu,\epsilon})\to\text{Ker}(P_{\mu,\epsilon}). Moreover

σ(μ,ϵ){z inside Γ}=σ(μ,ϵ|𝒱μ,ϵ)=σ(μ,ϵ),\displaystyle\sigma(\mathscr{L}_{\mu,\epsilon})\cap\{z\in\mathbb{C}\mbox{ inside }\Gamma\}=\sigma(\mathscr{L}_{\mu,\epsilon}|_{{\mathcal{V}}_{\mu,\epsilon}})=\sigma^{\prime}(\mathscr{L}_{\mu,\epsilon}),
σ(μ,ϵ){z outside Γ}=σ(μ,ϵ|Ker(Pμ,ϵ))=σ′′(μ,ϵ),\displaystyle\sigma(\mathscr{L}_{\mu,\epsilon})\cap\{z\in\mathbb{C}\mbox{ outside }\Gamma\}=\sigma(\mathscr{L}_{\mu,\epsilon}|_{Ker(P_{\mu,\epsilon})})=\sigma^{\prime\prime}(\mathscr{L}_{\mu,\epsilon})\ ,

proving the “semicontinuity property” (2.38) of separated parts of the spectrum.
4. The projectors Pμ,ϵP_{\mu,\epsilon} are similar one to each other: the transformation operators333 The operator (IdR)12(\mathrm{Id}-R)^{-\frac{1}{2}} is defined, for any operator RR satisfying R(Y)<1\|R\|_{{\mathcal{L}}(Y)}<1, by the power series (IdR)12:=k=0(1/2k)(R)k=Id+12R+38R2+𝒪(R3).\displaystyle(\mathrm{Id}-R)^{-\frac{1}{2}}:=\sum_{k=0}^{\infty}{-1/2\choose k}(-R)^{k}=\mathrm{Id}+\frac{1}{2}R+\frac{3}{8}R^{2}+\mathcal{O}(R^{3})\,. (3.6)

Uμ,ϵ:=(Id(Pμ,ϵP0,0)2)1/2[Pμ,ϵP0,0+(IdPμ,ϵ)(IdP0,0)]U_{\mu,\epsilon}:=\big{(}\mathrm{Id}-(P_{\mu,\epsilon}-P_{0,0})^{2}\big{)}^{-1/2}\big{[}P_{\mu,\epsilon}P_{0,0}+(\mathrm{Id}-P_{\mu,\epsilon})(\mathrm{Id}-P_{0,0})\big{]} (3.7)

are bounded and invertible in YY and in XX, with inverse

Uμ,ϵ1=[P0,0Pμ,ϵ+(IdP0,0)(IdPμ,ϵ)](Id(Pμ,ϵP0,0)2)1/2,U_{\mu,\epsilon}^{-1}=\big{[}P_{0,0}P_{\mu,\epsilon}+(\mathrm{Id}-P_{0,0})(\mathrm{Id}-P_{\mu,\epsilon})\big{]}\big{(}\mathrm{Id}-(P_{\mu,\epsilon}-P_{0,0})^{2}\big{)}^{-1/2}\,,

and Uμ,ϵP0,0Uμ,ϵ1=Pμ,ϵU_{\mu,\epsilon}P_{0,0}U_{\mu,\epsilon}^{-1}=P_{\mu,\epsilon} as well as Uμ,ϵ1Pμ,ϵUμ,ϵ=P0,0U_{\mu,\epsilon}^{-1}P_{\mu,\epsilon}U_{\mu,\epsilon}=P_{0,0}.

The map (μ,ϵ)Uμ,ϵ(\mu,\epsilon)\mapsto U_{\mu,\epsilon} is analytic from B(μ0)×B(ϵ0)B(\mu_{0})\times B(\epsilon_{0}) to (Y)\mathcal{L}(Y).
5. The subspaces 𝒱μ,ϵ=Rg(Pμ,ϵ)\mathcal{V}_{\mu,\epsilon}=\text{Rg}(P_{\mu,\epsilon}) are isomorphic one to each other: 𝒱μ,ϵ=Uμ,ϵ𝒱0,0.\mathcal{V}_{\mu,\epsilon}=U_{\mu,\epsilon}\mathcal{V}_{0,0}. In particular dim𝒱μ,ϵ=dim𝒱0,0=4\dim\mathcal{V}_{\mu,\epsilon}=\dim\mathcal{V}_{0,0}=4, for any (μ,ϵ)B(μ0)×B(ϵ0)(\mu,\epsilon)\in B(\mu_{0})\times B(\epsilon_{0}).

Proof.

For any λ\lambda\in\mathbb{C} we decompose μ,ϵλ=0,0λ+μ,ϵ\mathscr{L}_{\mu,\epsilon}-\lambda=\mathscr{L}_{0,0}-\lambda+{\mathcal{R}}_{\mu,\epsilon} where 0,0=[𝚌𝚑x|D|tanh(𝚑|D|)1𝚌𝚑x]\footnotesize\mathscr{L}_{0,0}=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}\partial_{x}&|D|\tanh(\mathtt{h}|D|)\\ -1&{\mathtt{c}}_{\mathtt{h}}\partial_{x}\end{bmatrix} and

μ,ϵ:=μ,ϵ0,0=[(x+iμ)pϵ(x)fμ,ϵ(D)aϵ(x)pϵ(x)(x+iμ)]:YX,{\mathcal{R}}_{\mu,\epsilon}:=\mathscr{L}_{\mu,\epsilon}-\mathscr{L}_{0,0}=\begin{bmatrix}(\partial_{x}+\mathrm{i}\,\mu)p_{\epsilon}(x)&f_{\mu,\epsilon}(D)\\ -a_{\epsilon}(x)&p_{\epsilon}(x)(\partial_{x}+\mathrm{i}\,\mu)\end{bmatrix}:Y\to X\,,

having used also (2.30) and setting

fμ,ϵ(D):=|D+μ|tanh((𝚑+𝚏ϵ)|D+μ|)|D|tanh(𝚑|D|)(Y),fμ,ϵ(D)(Y)=𝒪(μ,ϵ).f_{\mu,\epsilon}(D):=|D+\mu|\,\tanh\big{(}(\mathtt{h}+\mathtt{f}_{\epsilon})|D+\mu|\big{)}-|D|\tanh(\mathtt{h}|D|)\in\mathcal{L}(Y)\,,\ \ \ {\|f_{\mu,\epsilon}(D)\|}_{\mathcal{L}(Y)}=\mathcal{O}(\mu,\epsilon)\,.

For any λΓ\lambda\in\Gamma, the operator 0,0λ\mathscr{L}_{0,0}-\lambda is invertible and its inverse is the Fourier multiplier matrix operator

(0,0λ)1=Op(1(i𝚌𝚑kλ)2+|k|tanh(𝚑|k|)[i𝚌𝚑kλ|k|tanh(𝚑|k|)1i𝚌𝚑kλ]):XY.(\mathscr{L}_{0,0}-\lambda)^{-1}=\text{Op}\left(\frac{1}{(\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}k-\lambda)^{2}+|k|\tanh(\mathtt{h}|k|)}\begin{bmatrix}\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}k-\lambda&-|k|\tanh(\mathtt{h}|k|)\\ 1&\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}k-\lambda\end{bmatrix}\right):X\to Y\,.

Hence, for |ϵ|<ϵ0|\epsilon|<\epsilon_{0} and |μ|<μ0|\mu|<\mu_{0} small enough, uniformly on the compact set Γ\Gamma, the operator (0,0λ)1μ,ϵ:YY(\mathscr{L}_{0,0}-\lambda)^{-1}{\mathcal{R}}_{\mu,\epsilon}:Y\to Y is bounded, with small operatorial norm. Then μ,ϵλ\mathscr{L}_{\mu,\epsilon}-\lambda is invertible by Neumann series and Γ\Gamma belongs to the resolvent set of μ,ϵ\mathscr{L}_{\mu,\epsilon}. The remaining part of the proof follows exactly as in Lemma 3.1 in [6]. ∎

The Hamiltonian and reversible nature of the operator μ,ϵ\mathscr{L}_{\mu,\epsilon}, see (3.3) and (3.4), imply additional algebraic properties for spectral projectors Pμ,ϵP_{\mu,\epsilon} and the transformation operators Uμ,ϵU_{\mu,\epsilon}. By Lemma 3.2 in [6] we have that:

Lemma 3.2.

For any (μ,ϵ)B(μ0)×B(ϵ0)(\mu,\epsilon)\in B(\mu_{0})\times B(\epsilon_{0}), the following holds true:
(ii) The projectors Pμ,ϵP_{\mu,\epsilon} defined in (3.5) are skew-Hamiltonian, namely 𝒥Pμ,ϵ=Pμ,ϵ𝒥\mathcal{J}P_{\mu,\epsilon}=P_{\mu,\epsilon}^{*}\mathcal{J}, and reversibility preserving, i.e. ρ¯Pμ,ϵ=Pμ,ϵρ¯\overline{\rho}P_{\mu,\epsilon}=P_{\mu,\epsilon}\overline{\rho}.
(ii) The transformation operators Uμ,ϵU_{\mu,\epsilon} in (3.7) are symplectic, namely Uμ,ϵ𝒥Uμ,ϵ=𝒥U_{\mu,\epsilon}^{*}\mathcal{J}U_{\mu,\epsilon}=\mathcal{J}, and reversibility preserving.
(iii) P0,ϵP_{0,\epsilon} and U0,ϵU_{0,\epsilon} are real operators, i.e. P0,ϵ¯=P0,ϵ\overline{P_{0,\epsilon}}=P_{0,\epsilon} and U0,ϵ¯=U0,ϵ\overline{U_{0,\epsilon}}=U_{0,\epsilon}.

By the previous lemma, the linear involution ρ¯\overline{\rho} commutes with the spectral projectors Pμ,ϵP_{\mu,\epsilon} and then ρ¯\overline{\rho} leaves invariant the subspace 𝒱μ,ϵ=Rg(Pμ,ϵ)\mathcal{V}_{\mu,\epsilon}=\text{Rg}(P_{\mu,\epsilon}).
Symplectic and reversible basis of 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon}. It is convenient to represent the Hamiltonian and reversible operator μ,ϵ:𝒱μ,ϵ𝒱μ,ϵ\mathscr{L}_{\mu,\epsilon}:\mathcal{V}_{\mu,\epsilon}\to\mathcal{V}_{\mu,\epsilon} in a basis which is symplectic and reversible, according to the following definition.

Definition 3.3.

(Symplectic and reversible basis) A basis 𝙵:={𝚏1+,𝚏1,𝚏0+,𝚏0}\mathtt{F}:=\{\mathtt{f}^{+}_{1},\,\mathtt{f}^{-}_{1},\,\mathtt{f}^{+}_{0},\,\mathtt{f}^{-}_{0}\} of 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon} is

  • symplectic if, for any k,k=0,1k,k^{\prime}=0,1,

    (𝒥𝚏k,𝚏k+)=1,(𝒥𝚏kσ,𝚏kσ)=0,σ=±;ifkkthen(𝒥𝚏kσ,𝚏kσ)=0,σ,σ=±.\left(\mathcal{J}\mathtt{f}_{k}^{-}\,,\,\mathtt{f}_{k}^{+}\right)=1\,,\ \ \big{(}\mathcal{J}\mathtt{f}_{k}^{\sigma},\mathtt{f}_{k}^{\sigma}\big{)}=0\,,\ \forall\sigma=\pm\,;\ \ \text{if}\ k\neq k^{\prime}\ \text{then}\ \big{(}\mathcal{J}\mathtt{f}_{k}^{\sigma},\mathtt{f}_{k^{\prime}}^{\sigma^{\prime}}\big{)}=0\,,\ \forall\sigma,\sigma^{\prime}=\pm\,. (3.8)
  • reversible if

    ρ¯𝚏1+=𝚏1+,ρ¯𝚏1=𝚏1,ρ¯𝚏0+=𝚏0+,ρ¯𝚏0=𝚏0,i.e. ρ¯𝚏kσ=σ𝚏kσ,σ=±,k=0,1.\overline{\rho}\mathtt{f}^{+}_{1}=\mathtt{f}^{+}_{1},\quad\overline{\rho}\mathtt{f}^{-}_{1}=-\mathtt{f}^{-}_{1},\quad\overline{\rho}\mathtt{f}^{+}_{0}=\mathtt{f}^{+}_{0},\quad\overline{\rho}\mathtt{f}^{-}_{0}=-\mathtt{f}^{-}_{0},\quad\text{i.e. }\overline{\rho}\mathtt{f}_{k}^{\sigma}=\sigma\mathtt{f}_{k}^{\sigma}\,,\ \forall\sigma=\pm,k=0,1\,. (3.9)

We use the following notation along the paper: we denote by even(x)even(x) a real 2π2\pi-periodic function which is even in xx, and by odd(x)odd(x) a real 2π2\pi-periodic function which is odd in xx.

By the definition of the involution ρ¯\overline{\rho} in (2.27), the real and imaginary parts of a reversible basis 𝙵={𝚏k±}\mathtt{F}=\{\mathtt{f}^{\pm}_{k}\}, k=0,1k=0,1, enjoy the following parity properties (cfr. Lemma 3.4 in [6])

𝚏k+(x)=[even(x)+iodd(x)odd(x)+ieven(x)],𝚏k(x)=[odd(x)+ieven(x)even(x)+iodd(x)].\mathtt{f}_{k}^{+}(x)=\begin{bmatrix}even(x)+\mathrm{i}\,odd(x)\\ odd(x)+\mathrm{i}\,even(x)\end{bmatrix},\quad\mathtt{f}_{k}^{-}(x)=\begin{bmatrix}odd(x)+\mathrm{i}\,even(x)\\ even(x)+\mathrm{i}\,odd(x)\end{bmatrix}. (3.10)

By Lemmata 3.5 and 3.6 in [6] we have the following result.

Lemma 3.4.

The 4×44\times 4 matrix that represents the Hamiltonian and reversible operator μ,ϵ=𝒥μ,ϵ:𝒱μ,ϵ𝒱μ,ϵ\mathscr{L}_{\mu,\epsilon}=\mathcal{J}{\mathcal{B}}_{\mu,\epsilon}:\mathcal{V}_{\mu,\epsilon}\to\mathcal{V}_{\mu,\epsilon} with respect to a symplectic and reversible basis 𝙵={𝚏1+,𝚏1,𝚏0+,𝚏0}\mathtt{F}=\{\mathtt{f}_{1}^{+},\mathtt{f}_{1}^{-},\mathtt{f}_{0}^{+},\mathtt{f}_{0}^{-}\} of 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon} is

𝙹4𝙱μ,ϵ,𝙹4:=(𝙹200𝙹2),𝙹2:=(0110),where 𝙱μ,ϵ=𝙱μ,ϵ\displaystyle\mathtt{J}_{4}\mathtt{B}_{\mu,\epsilon}\,,\quad\mathtt{J}_{4}:=\begin{pmatrix}\mathtt{J}_{2}&\vline&0\\ \hline\cr 0&\vline&\mathtt{J}_{2}\end{pmatrix},\quad{\small\mathtt{J}_{2}:=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}},\quad\text{where }\quad\mathtt{B}_{\mu,\epsilon}=\mathtt{B}_{\mu,\epsilon}^{*} (3.11)

is the self-adjoint matrix

𝙱μ,ϵ=((μ,ϵ𝚏1+,𝚏1+)(μ,ϵ𝚏1,𝚏1+)(μ,ϵ𝚏0+,𝚏1+)(μ,ϵ𝚏0,𝚏1+)(μ,ϵ𝚏1+,𝚏1)(μ,ϵ𝚏1,𝚏1)(μ,ϵ𝚏0+,𝚏1)(μ,ϵ𝚏0,𝚏1)(μ,ϵ𝚏1+,𝚏0+)(μ,ϵ𝚏1,𝚏0+)(μ,ϵ𝚏0+,𝚏0+)(μ,ϵ𝚏0,𝚏0+)(μ,ϵ𝚏1+,𝚏0)(μ,ϵ𝚏1,𝚏0)(μ,ϵ𝚏0+,𝚏0)(μ,ϵ𝚏0,𝚏0)).\mathtt{B}_{\mu,\epsilon}=\begin{pmatrix}\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{1}}\ ,\mathtt{f}^{+}_{1}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{1}}\ ,\mathtt{f}^{+}_{1}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{0}}\ ,\mathtt{f}^{+}_{1}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{0}}\ ,\mathtt{f}^{+}_{1}\right)\\ \left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{1}}\ ,\mathtt{f}^{-}_{1}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{1}}\ ,\mathtt{f}^{-}_{1}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{0}}\ ,\mathtt{f}^{-}_{1}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{0}}\ ,\mathtt{f}^{-}_{1}\right)\\ \left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{1}}\ ,\mathtt{f}^{+}_{0}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{1}}\ ,\mathtt{f}^{+}_{0}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{0}}\ ,\mathtt{f}^{+}_{0}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{0}}\ ,\mathtt{f}^{+}_{0}\right)\\ \left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{1}}\ ,\mathtt{f}^{-}_{0}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{1}}\ ,\mathtt{f}^{-}_{0}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{+}_{{0}}\ ,\mathtt{f}^{-}_{0}\right)&\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{-}_{{0}}\ ,\mathtt{f}^{-}_{0}\right)\\ \end{pmatrix}. (3.12)

The entries of the matrix 𝙱μ,ϵ\mathtt{B}_{\mu,\epsilon} are alternatively real or purely imaginary: for any σ=±\sigma=\pm, k=0,1k=0,1,

(μ,ϵ𝚏kσ,𝚏kσ) is real,(μ,ϵ𝚏kσ,𝚏kσ) is purely imaginary.\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{\sigma}_{k}\,,\,\mathtt{f}^{\sigma}_{k^{\prime}}\right)\text{ is real},\qquad\left({\mathcal{B}}_{\mu,\epsilon}\,\mathtt{f}^{\sigma}_{k}\,,\,\mathtt{f}^{-\sigma}_{k^{\prime}}\right)\text{ is purely imaginary}\,. (3.13)

It is convenient to give a name to the matrices of the form obtained in Lemma 3.4.

Definition 3.5.

A 2n×2n2n\times 2n, n=1,2,n=1,2, matrix of the form 𝙻=𝙹2n𝙱\mathtt{L}=\mathtt{J}_{2n}\mathtt{B} is
1. Hamiltonian if 𝙱\mathtt{B} is a self-adjoint matrix, i.e. 𝙱=𝙱\mathtt{B}=\mathtt{B}^{*};
2. Reversible if 𝙱\mathtt{B} is reversibility-preserving, i.e. ρ2n𝙱=𝙱ρ2n\rho_{2n}\circ\mathtt{B}=\mathtt{B}\circ\rho_{2n}, where

ρ4:=(ρ200ρ2),ρ2:=(𝔠00𝔠),\rho_{4}:=\begin{pmatrix}\rho_{2}&0\\ 0&\rho_{2}\end{pmatrix},\qquad\rho_{2}:=\begin{pmatrix}\mathfrak{c}&0\\ 0&-\mathfrak{c}\end{pmatrix}, (3.14)

and 𝔠:zz¯\mathfrak{c}:z\mapsto\overline{z} is the conjugation of the complex plane. Equivalently, ρ2n𝙻=𝙻ρ2n\rho_{2n}\circ\mathtt{L}=-\mathtt{L}\circ\rho_{2n}.

In the sequel we shall mainly deal with 4×44\times 4 Hamiltonian and reversible matrices. The transformations preserving the Hamiltonian structure are called symplectic, and satisfy

Y𝙹4Y=𝙹4.\displaystyle Y^{*}\mathtt{J}_{4}Y=\mathtt{J}_{4}\,. (3.15)

If YY is symplectic then YY^{*} and Y1Y^{-1} are symplectic as well. A Hamiltonian matrix 𝙻=𝙹4𝙱\mathtt{L}=\mathtt{J}_{4}\mathtt{B}, with 𝙱=𝙱\mathtt{B}=\mathtt{B}^{*}, is conjugated through YY in the new Hamiltonian matrix

𝙻1=Y1𝙻Y=Y1𝙹4YY𝙱Y=𝙹4𝙱1where 𝙱1:=Y𝙱Y=𝙱1.\mathtt{L}_{1}=Y^{-1}\mathtt{L}Y=Y^{-1}\mathtt{J}_{4}Y^{-*}Y^{*}\mathtt{B}Y=\mathtt{J}_{4}\mathtt{B}_{1}\quad\text{where }\quad\mathtt{B}_{1}:=Y^{*}\mathtt{B}Y=\mathtt{B}_{1}^{*}\,. (3.16)

Note that the matrix ρ4\rho_{4} in (3.14) represents the action of the involution ρ¯:𝒱μ,ϵ𝒱μ,ϵ\overline{\rho}:{\mathcal{V}}_{\mu,\epsilon}\to{\mathcal{V}}_{\mu,\epsilon} defined in (2.27) in a reversible basis (cfr. (3.9)). A 4×44\times 4 matrix 𝙱=(𝙱ij)i,j=1,,4\mathtt{B}=(\mathtt{B}_{ij})_{i,j=1,\dots,4} is reversibility-preserving if and only if its entries are alternatively real and purely imaginary, namely 𝙱ij\mathtt{B}_{ij} is real when i+ji+j is even and purely imaginary otherwise, as in (3.13). A 4×44\times 4 complex matrix 𝙻=(𝙻ij)i,j=1,,4\mathtt{L}=(\mathtt{L}_{ij})_{i,j=1,\ldots,4} is reversible if and only if 𝙻ij\mathtt{L}_{ij} is purely imaginary when i+ji+j is even and real otherwise.

We finally mention that the flow of a Hamiltonian reversibility-preserving matrix is symplectic and reversibility-preserving (see Lemma 3.8 in [6]).

4 Matrix representation of μ,ϵ\mathscr{L}_{\mu,\epsilon} on 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon}

Using the transformation operators Uμ,ϵU_{\mu,\epsilon} in (3.7), we construct the basis of 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon}

:={f1+(μ,ϵ),f1(μ,ϵ),f0+(μ,ϵ),f0(μ,ϵ)},fkσ(μ,ϵ):=Uμ,ϵfkσ,σ=±,k=0,1,\mathcal{F}:=\big{\{}f_{1}^{+}(\mu,\epsilon),\ f_{1}^{-}(\mu,\epsilon),\ f_{0}^{+}(\mu,\epsilon),\ f_{0}^{-}(\mu,\epsilon)\big{\}}\,,\quad f_{k}^{\sigma}(\mu,\epsilon):=U_{\mu,\epsilon}f_{k}^{\sigma}\,,\ \sigma=\pm\,,\,k=0,1\,, (4.1)

where

f1+=[𝚌𝚑1/2cos(x)𝚌𝚑1/2sin(x)],f1=[𝚌𝚑1/2sin(x)𝚌𝚑1/2cos(x)],f0+=[10],f0=[01],f_{1}^{+}=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{1/2}\cos(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-1/2}\sin(x)\end{bmatrix},\quad f_{1}^{-}=\begin{bmatrix}-{\mathtt{c}}_{\mathtt{h}}^{1/2}\sin(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-1/2}\cos(x)\end{bmatrix},\quad f_{0}^{+}=\begin{bmatrix}1\\ 0\end{bmatrix},\quad f_{0}^{-}=\begin{bmatrix}0\\ 1\end{bmatrix}\,, (4.2)

form a basis of 𝒱0,0=Rg(P0,0)\mathcal{V}_{0,0}=\mathrm{Rg}(P_{0,0}), cfr. (2.34)-(2.35). Note that the real valued vectors {f1±,f0±}\{f_{1}^{\pm},f_{0}^{\pm}\} form a symplectic and reversible basis for 𝒱0,0\mathcal{V}_{0,0}, according to Definition 3.3. Then, by Lemma 3.2 and 3.1 we deduce that (cfr. Lemma 4.1 in [6]):

Lemma 4.1.

The basis \mathcal{F} of 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon} defined in (4.1), is symplectic and reversible, i.e. satisfies (3.8) and (3.9). Each map (μ,ϵ)fkσ(μ,ϵ)(\mu,\epsilon)\mapsto f^{\sigma}_{k}(\mu,\epsilon) is analytic as a map B(μ0)×B(ϵ0)H1(𝕋)B(\mu_{0})\times B(\epsilon_{0})\to H^{1}(\mathbb{T}).

In the next lemma we expand the vectors fkσ(μ,ϵ)f_{k}^{\sigma}(\mu,\epsilon) in (μ,ϵ)(\mu,\epsilon). We denote by even0(x)even_{0}(x) a real, even, 2π2\pi-periodic function with zero space average. In the sequel 𝒪(μmϵn)[even(x)odd(x)]\mathcal{O}(\mu^{m}\epsilon^{n})\footnotesize\begin{bmatrix}even(x)\\ odd(x)\end{bmatrix} denotes an analytic map in (μ,ϵ)(\mu,\epsilon) with values in H1(𝕋,2)H^{1}(\mathbb{T},\mathbb{C}^{2}), whose first component is even(x)even(x) and the second one odd(x)odd(x); similar meaning for 𝒪(μmϵn)[odd(x)even(x)]\mathcal{O}(\mu^{m}\epsilon^{n})\footnotesize\begin{bmatrix}odd(x)\\ even(x)\end{bmatrix}, etc…

Lemma 4.2.

(Expansion of the basis \mathcal{F}) For small values of (μ,ϵ)(\mu,\epsilon) the basis \mathcal{F} in (4.1) has the expansion

f1+(μ,ϵ)\displaystyle f^{+}_{1}(\mu,\epsilon) =[𝚌𝚑12cos(x)𝚌𝚑12sin(x)]+iμ4γ𝚑[𝚌𝚑12sin(x)𝚌𝚑12cos(x)]+ϵ[α𝚑cos(2x)β𝚑sin(2x)]\displaystyle=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\cos(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\sin(x)\end{bmatrix}+\mathrm{i}\,\frac{\mu}{4}\gamma_{\mathtt{h}}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\sin(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\cos(x)\end{bmatrix}+\epsilon\begin{bmatrix}\alpha_{\mathtt{h}}\cos(2x)\\ \beta_{\mathtt{h}}\sin(2x)\end{bmatrix} (4.3)
+𝒪(μ2)[even0(x)+iodd(x)odd(x)+ieven0(x)]+𝒪(ϵ2)[even0(x)odd(x)]+iμϵ[odd(x)even(x)]+𝒪(μ2ϵ,μϵ2),\displaystyle+\mathcal{O}(\mu^{2})\begin{bmatrix}even_{0}(x)+\mathrm{i}\,odd(x)\\ odd(x)+\mathrm{i}\,even_{0}(x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix}+\mathrm{i}\,\mu\epsilon\begin{bmatrix}odd(x)\\ even(x)\end{bmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu\epsilon^{2})\,,
f1(μ,ϵ)\displaystyle f^{-}_{1}(\mu,\epsilon) =[𝚌𝚑12sin(x)𝚌𝚑12cos(x)]+iμ4γ𝚑[𝚌𝚑12cos(x)𝚌𝚑12sin(x)]+ϵ[α𝚑sin(2x)β𝚑cos(2x)]\displaystyle=\begin{bmatrix}-{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\sin(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\cos(x)\end{bmatrix}+\mathrm{i}\,\frac{\mu}{4}\gamma_{\mathtt{h}}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\cos(x)\\ -{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\sin(x)\end{bmatrix}+\epsilon\begin{bmatrix}-\alpha_{\mathtt{h}}\sin(2x)\\ \beta_{\mathtt{h}}\cos(2x)\end{bmatrix} (4.4)
+𝒪(μ2)[odd(x)+ieven0(x)even0(x)+iodd(x)]+𝒪(ϵ2)[odd(x)even(x)]+iμϵ[even(x)odd(x)]+𝒪(μ2ϵ,μϵ2),\displaystyle+\mathcal{O}(\mu^{2})\begin{bmatrix}odd(x)+\mathrm{i}\,even_{0}(x)\\ even_{0}(x)+\mathrm{i}\,odd(x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\begin{bmatrix}odd(x)\\ even(x)\end{bmatrix}+\mathrm{i}\,\mu\epsilon\begin{bmatrix}even(x)\\ odd(x)\end{bmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu\epsilon^{2})\,,
f0+(μ,ϵ)\displaystyle f^{+}_{0}(\mu,\epsilon) =[10]+ϵδ𝚑[𝚌𝚑12cos(x)𝚌𝚑12sin(x)]+𝒪(ϵ2)[even0(x)odd(x)]+iμϵ[odd(x)even0(x)]+𝒪(μ2ϵ,μϵ2),\displaystyle=\begin{bmatrix}1\\ 0\end{bmatrix}+\epsilon\delta_{\mathtt{h}}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\cos(x)\\ -{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\sin(x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix}+\mathrm{i}\,\mu\epsilon\begin{bmatrix}odd(x)\\ even_{0}(x)\end{bmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu\epsilon^{2})\,, (4.5)
f0(μ,ϵ)\displaystyle f^{-}_{0}(\mu,\epsilon) =[01]+iμϵ[even0(x)odd(x)]+𝒪(μ2ϵ,μϵ2),\displaystyle=\begin{bmatrix}0\\ 1\end{bmatrix}+\mathrm{i}\,\mu\epsilon\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu\epsilon^{2})\,, (4.6)

where the remainders 𝒪()\mathcal{O}() are vectors in H1(𝕋)H^{1}(\mathbb{T}) and

α𝚑:=12𝚌𝚑112(3+𝚌𝚑4),β𝚑:=14𝚌𝚑132(1+𝚌𝚑4)(3𝚌𝚑4),γ𝚑:=1+𝚑(1𝚌𝚑4)𝚌𝚑2,δ𝚑:=3+𝚌𝚑44𝚌𝚑52.\alpha_{\mathtt{h}}:=\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{11}{2}}(3+{\mathtt{c}}_{\mathtt{h}}^{4})\,,\quad\beta_{\mathtt{h}}:=\frac{1}{4}{\mathtt{c}}_{\mathtt{h}}^{-\frac{13}{2}}(1+{\mathtt{c}}_{\mathtt{h}}^{4})(3-{\mathtt{c}}_{\mathtt{h}}^{4})\,,\quad\gamma_{\mathtt{h}}:=1+\frac{\mathtt{h}(1-{\mathtt{c}}_{\mathtt{h}}^{4})}{{\mathtt{c}}_{\mathtt{h}}^{2}}\,,\quad\delta_{\mathtt{h}}:=\frac{3+{\mathtt{c}}_{\mathtt{h}}^{4}}{4{\mathtt{c}}_{\mathtt{h}}^{\frac{5}{2}}}\,. (4.7)

For μ=0\mu=0 the basis {fk±(0,ϵ),k=0,1}\{f_{k}^{\pm}(0,\epsilon),k=0,1\} is real and

f1+(0,ϵ)=[even0(x)odd(x)],f1(0,ϵ)=[odd(x)even(x)],f0+(0,ϵ)=[10]+[even0(x)odd(x)],f0(0,ϵ)=[01].f^{+}_{1}(0,\epsilon)=\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix},\ f^{-}_{1}(0,\epsilon)=\begin{bmatrix}odd(x)\\ even(x)\end{bmatrix},\ f^{+}_{0}(0,\epsilon)=\begin{bmatrix}1\\ 0\end{bmatrix}+\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix}\,,\ f^{-}_{0}(0,\epsilon)=\begin{bmatrix}0\\ 1\end{bmatrix}\,. (4.8)
Proof.

The long calculations are given in Appendix A. ∎

We now state the main result of this section.

Proposition 4.3.

The matrix that represents the Hamiltonian and reversible operator μ,ϵ:𝒱μ,ϵ𝒱μ,ϵ\mathscr{L}_{\mu,\epsilon}:\mathcal{V}_{\mu,\epsilon}\to\mathcal{V}_{\mu,\epsilon} in the symplectic and reversible basis \mathcal{F} of 𝒱μ,ϵ\mathcal{V}_{\mu,\epsilon} defined in (4.1), is a Hamiltonian matrix 𝙻μ,ϵ=𝙹4𝙱μ,ϵ\mathtt{L}_{\mu,\epsilon}=\mathtt{J}_{4}\mathtt{B}_{\mu,\epsilon}, where 𝙱μ,ϵ\mathtt{B}_{\mu,\epsilon} is a self-adjoint and reversibility preserving (i.e. satisfying (3.13)) 4×44\times 4 matrix of the form

𝙱μ,ϵ=(EFFG),E=E,G=G,\mathtt{B}_{\mu,\epsilon}=\begin{pmatrix}E&F\\ F^{*}&G\end{pmatrix},\qquad E=E^{*}\,,\ \ G=G^{*}\,, (4.9)

where E,F,GE,F,G are the 2×22\times 2 matrices

E:=(𝚎11ϵ2(1+r1(ϵ,μϵ))𝚎22μ28(1+r1′′(ϵ,μ))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))𝚎22μ28(1+r5(ϵ,μ)))\displaystyle E:=\begin{pmatrix}\mathtt{e}_{11}\epsilon^{2}(1+r_{1}^{\prime}(\epsilon,\mu\epsilon))-\mathtt{e}_{22}\frac{\mu^{2}}{8}(1+r_{1}^{\prime\prime}(\epsilon,\mu))&\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}\\ -\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}&-\mathtt{e}_{22}\frac{\mu^{2}}{8}(1+r_{5}(\epsilon,\mu))\end{pmatrix} (4.10)
G:=(1+r8(ϵ2,μ2ϵ)ir9(μϵ2,μ2ϵ)ir9(μϵ2,μ2ϵ)μtanh(𝚑μ)+r10(μ2ϵ))\displaystyle G:=\begin{pmatrix}1+r_{8}(\epsilon^{2},\mu^{2}\epsilon)&-\mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)\\ \mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)&\mu\tanh(\mathtt{h}\mu)+r_{10}(\mu^{2}\epsilon)\end{pmatrix} (4.11)
F:=(𝚏11ϵ+r3(ϵ3,μϵ2,μ2ϵ)iμϵ𝚌𝚑12+ir4(μϵ2,μ2ϵ)ir6(μϵ)r7(μ2ϵ)),\displaystyle F:=\begin{pmatrix}\mathtt{f}_{11}\epsilon+r_{3}(\epsilon^{3},\mu\epsilon^{2},\mu^{2}\epsilon)&\mathrm{i}\,\mu\epsilon{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}+\mathrm{i}\,r_{4}({\mu\epsilon^{2}},\mu^{2}\epsilon)\\ \mathrm{i}\,r_{6}(\mu\epsilon)&r_{7}(\mu^{2}\epsilon)\end{pmatrix}\,, (4.12)

with 𝚎12\mathtt{e}_{12} and 𝚎22\mathtt{e}_{22} given in (1.2) and (1.3) respectively, and

𝚎11\displaystyle\mathtt{e}_{11} :=9𝚌𝚑810𝚌𝚑4+98𝚌𝚑7=9(1𝚌𝚑4)2+8𝚌𝚑48𝚌𝚑7>0,𝚏11:=12𝚌𝚑32(1𝚌𝚑4).\displaystyle:=\dfrac{9{\mathtt{c}}_{\mathtt{h}}^{8}-10{\mathtt{c}}_{\mathtt{h}}^{4}+9}{8{\mathtt{c}}_{\mathtt{h}}^{7}}=\dfrac{9(1-{\mathtt{c}}_{\mathtt{h}}^{4})^{2}+8{\mathtt{c}}_{\mathtt{h}}^{4}}{8{\mathtt{c}}_{\mathtt{h}}^{7}}>0\,,\qquad\mathtt{f}_{11}:=\tfrac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{3}{2}}(1-{\mathtt{c}}_{\mathtt{h}}^{4})\,. (4.13)

The rest of this section is devoted to the proof of Proposition 4.3.

We decompose μ,ϵ{\mathcal{B}}_{\mu,\epsilon} in (3.3) as

μ,ϵ=ϵ++,{\mathcal{B}}_{\mu,\epsilon}={\mathcal{B}}_{\epsilon}+{\mathcal{B}}^{\flat}+{\mathcal{B}}^{\sharp}\,,

where ϵ{\mathcal{B}}_{\epsilon}, {\mathcal{B}}^{\flat}, {\mathcal{B}}^{\sharp} are the self-adjoint and reversibility preserving operators

ϵ:=0,ϵ:=[1+aϵ(x)(𝚌𝚑+pϵ(x))xx(𝚌𝚑+pϵ(x))|D|tanh((𝚑+𝚏ϵ)|D|)],\displaystyle{\mathcal{B}}_{\epsilon}:={\mathcal{B}}_{0,\epsilon}:=\left[\begin{array}[]{cc}1+a_{\epsilon}(x)&-({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))\partial_{x}\\ \partial_{x}\circ({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))&|D|\tanh((\mathtt{h}+\mathtt{f}_{\epsilon})|D|)\end{array}\right], (4.16)
:=[000|D+μ|tanh((𝚑+𝚏ϵ)|D+μ|)|D|tanh((𝚑+𝚏ϵ)|D|)],\displaystyle{\mathcal{B}}^{\flat}:=\begin{bmatrix}0&0\\ 0&|D+\mu|\tanh((\mathtt{h}+\mathtt{f}_{\epsilon})|D+\mu|)-|D|\tanh((\mathtt{h}+\mathtt{f}_{\epsilon})|D|)\end{bmatrix},\, (4.17)
:=μ[0ipϵipϵ0].\displaystyle{\mathcal{B}}^{\sharp}:=\mu\begin{bmatrix}0&-\mathrm{i}\,p_{\epsilon}\\ \mathrm{i}\,p_{\epsilon}&0\end{bmatrix}\,. (4.18)

In view of (2.29) the operator {\mathcal{B}}^{\flat} is analytic in μ\mu.

Lemma 4.4.

(Expansion of 𝙱ϵ\mathtt{B}_{\epsilon}) The self-adjoint and reversibility preserving matrix 𝙱ϵ:=𝙱ϵ(μ)\mathtt{B}_{\epsilon}:=\mathtt{B}_{\epsilon}(\mu) associated, as in (3.12), with the self-adjoint and reversibility preserving operator ϵ{\mathcal{B}}_{\epsilon} defined in (4.16), with respect to the basis \mathcal{F} of 𝒱μ,ϵ{\mathcal{V}}_{\mu,\epsilon} in (4.1), expands as

𝙱ϵ=(𝚎11ϵ2+ζ𝚑μ2+r1(ϵ3,μϵ3)ir2(μϵ2)𝚏11ϵ+r3(ϵ3,μϵ2)ir4(μϵ3)ir2(μϵ2)ζ𝚑μ2ir6(μϵ)0𝚏11ϵ+r3(ϵ3,μϵ2)ir6(μϵ)1+r8(ϵ2,μϵ2)ir9(μϵ2)ir4(μϵ3)0ir9(μϵ2)0)+𝒪(μ2ϵ,μ3),\displaystyle\mathtt{B}_{\epsilon}=\begin{pmatrix}\mathtt{e}_{11}\epsilon^{2}+\zeta_{\mathtt{h}}\mu^{2}+r_{1}(\epsilon^{3},\mu\epsilon^{3})&\mathrm{i}\,r_{2}(\mu\epsilon^{2})&\vline&\mathtt{f}_{11}\epsilon+r_{3}(\epsilon^{3},\mu\epsilon^{2})&\mathrm{i}\,r_{4}(\mu\epsilon^{3})\\ -\mathrm{i}\,r_{2}(\mu\epsilon^{2})&\zeta_{\mathtt{h}}\mu^{2}&\vline&\mathrm{i}\,r_{6}(\mu\epsilon)&0\\ \hline\cr\mathtt{f}_{11}\epsilon+r_{3}(\epsilon^{3},\mu\epsilon^{2})&-\mathrm{i}\,r_{6}(\mu\epsilon)&\vline&1+r_{8}(\epsilon^{2},\mu\epsilon^{2})&\mathrm{i}\,r_{9}(\mu\epsilon^{2})\\ -\mathrm{i}\,r_{4}(\mu\epsilon^{3})&0&\vline&-\mathrm{i}\,r_{9}(\mu\epsilon^{2})&0\\ \end{pmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3})\,, (4.19)

where 𝚎11\mathtt{e}_{11}, 𝚏11\mathtt{f}_{11} are defined respectively in (4.13), and

ζ𝚑:=18𝚌𝚑γ𝚑2.\zeta_{\mathtt{h}}:=\tfrac{1}{8}{\mathtt{c}}_{\mathtt{h}}\gamma_{\mathtt{h}}^{2}\,. (4.20)
Proof.

We expand the matrix 𝙱ϵ(μ)\mathtt{B}_{\epsilon}(\mu) as

𝙱ϵ(μ)=𝙱ϵ(0)+μ(μ𝙱ϵ)(0)+μ22(μ2𝙱0)(0)+𝒪(μ2ϵ,μ3).\mathtt{B}_{\epsilon}(\mu)=\mathtt{B}_{\epsilon}(0)+\mu(\partial_{\mu}\mathtt{B}_{\epsilon})(0)+\frac{\mu^{2}}{2}(\partial_{\mu}^{2}\mathtt{B}_{0})(0)+\mathcal{O}(\mu^{2}\epsilon,\mu^{3})\,. (4.21)

The matrix 𝙱ϵ(0)\mathtt{B}_{\epsilon}(0). The main result of this long paragraph is to prove that the matrix 𝙱ϵ(0)\mathtt{B}_{\epsilon}(0) has the expansion (4.25). The matrix 𝙱ϵ(0)\mathtt{B}_{\epsilon}(0) is real, because the operator ϵ{\mathcal{B}}_{\epsilon} is real and the basis {fk±(0,ϵ)}k=0,1\{f_{k}^{\pm}(0,\epsilon)\}_{k=0,1} is real. Consequently, by (3.13), its matrix elements (𝙱ϵ(0))i,j(\mathtt{B}_{\epsilon}(0))_{i,j} are real whenever i+ji+j is even and vanish for i+ji+j odd. In addition f0(0,ϵ)=[01]f^{-}_{0}(0,\epsilon)=\footnotesize\begin{bmatrix}0\\ 1\end{bmatrix} by (4.8), and, by (4.16), we get ϵf0(0,ϵ)=0{\mathcal{B}}_{\epsilon}f^{-}_{0}(0,\epsilon)=0, for any ϵ\epsilon. We deduce that the self-adjoint matrix 𝙱ϵ(0)\mathtt{B}_{\epsilon}(0) has the form

𝙱ϵ(0)=(ϵfkσ(0,ϵ),fkσ(0,ϵ))k,k=0,1,σ,σ=±=(E11(0,ϵ)0F11(0,ϵ)00E22(0,ϵ)00F11(0,ϵ)0G11(0,ϵ)00000),\mathtt{B}_{\epsilon}(0)=\left({\mathcal{B}}_{\epsilon}\,f^{\sigma}_{k}(0,\epsilon),\,f^{\sigma^{\prime}}_{k^{\prime}}(0,\epsilon)\right)_{k,k^{\prime}=0,1,\sigma,\sigma^{\prime}=\pm}=\begin{pmatrix}E_{11}(0,\epsilon)&0&\vline&F_{11}(0,\epsilon)&0\\ 0&E_{22}(0,\epsilon)&\vline&0&0\\ \hline\cr F_{11}(0,\epsilon)&0&\vline&G_{11}(0,\epsilon)&0\\ 0&0&\vline&0&0\end{pmatrix}\,, (4.22)

where E11(0,ϵ)E_{11}(0,\epsilon), E22(0,ϵ)E_{22}(0,\epsilon), G11(0,ϵ)G_{11}(0,\epsilon), F11(0,ϵ)F_{11}(0,\epsilon) are real. We claim that E22(0,ϵ)=0E_{22}(0,\epsilon)=0 for any ϵ\epsilon. As a first step, following [6], we prove that

 either E22(0,ϵ)0, or E11(0,ϵ)0F11(0,ϵ).\text{ either }\ E_{22}(0,\epsilon)\equiv 0\,,\qquad\text{ or }\ E_{11}(0,\epsilon)\equiv 0\equiv F_{11}(0,\epsilon)\,. (4.23)

Indeed, by (2.37), the operator 0,ϵ0,ϵ\mathscr{L}_{0,\epsilon}\equiv{\mathcal{L}}_{0,\epsilon} possesses, for any sufficiently small ϵ0\epsilon\neq 0, the eigenvalue 0 with a four dimensional generalized Kernel 𝒲ϵ:=span{U1,U~2,U3,U4}\mathcal{W}_{\epsilon}:=\text{span}\{U_{1},\tilde{U}_{2},U_{3},U_{4}\}, spanned by ϵ\epsilon-dependent vectors U1,U~2,U3,U4U_{1},\tilde{U}_{2},U_{3},U_{4}. By Lemma 3.1 it results that 𝒲ϵ=𝒱0,ϵ=Rg(P0,ϵ)\mathcal{W}_{\epsilon}={\mathcal{V}}_{0,\epsilon}=\text{Rg}(P_{0,\epsilon}) and by (2.37) we have 0,ϵ2=0\mathscr{L}_{0,\epsilon}^{2}=0 on 𝒱0,ϵ\mathcal{V}_{0,\epsilon}. Thus the matrix

𝙻ϵ(0):=𝙹4𝙱ϵ(0)=(0E22(0,ϵ)00E11(0,ϵ)0F11(0,ϵ)00000F11(0,ϵ)0G11(0,ϵ)0),\mathtt{L}_{\epsilon}(0):=\mathtt{J}_{4}\mathtt{B}_{\epsilon}(0)=\begin{pmatrix}0&E_{22}(0,\epsilon)&\vline&0&0\\ -E_{11}(0,\epsilon)&0&\vline&-F_{11}(0,\epsilon)&0\\ \hline\cr 0&0&\vline&0&0\\ -F_{11}(0,\epsilon)&0&\vline&-G_{11}(0,\epsilon)&0\end{pmatrix}\,, (4.24)

which represents 0,ϵ:𝒱0,ϵ𝒱0,ϵ\mathscr{L}_{0,\epsilon}:\mathcal{V}_{0,\epsilon}\to\mathcal{V}_{0,\epsilon}, satisfies 𝙻ϵ2(0)=0\mathtt{L}^{2}_{\epsilon}(0)=0, namely

𝙻ϵ2(0)=((E11E22)(0,ϵ)0(F11E22)(0,ϵ)00(E11E22)(0,ϵ)0000000(F11E22)(0,ϵ)00)=0\mathtt{L}^{2}_{\epsilon}(0)=-\begin{pmatrix}(E_{11}E_{22})(0,\epsilon)&0&\vline&(F_{11}E_{22})(0,\epsilon)&0\\ 0&(E_{11}E_{22})(0,\epsilon)&\vline&0&0\\ \hline\cr 0&0&\vline&0&0\\ 0&(F_{11}E_{22})(0,\epsilon)&\vline&0&0\end{pmatrix}=0

which implies (4.23). We now prove that the matrix 𝙱ϵ(0)\mathtt{B}_{\epsilon}(0) defined in (4.22) expands as

𝙱ϵ(0)=(𝚎11ϵ2+r(ϵ3)0𝚏11ϵ+r(ϵ3)00000𝚏11ϵ+r(ϵ3)01+r(ϵ2)00000)\mathtt{B}_{\epsilon}(0)=\begin{pmatrix}\mathtt{e}_{11}\epsilon^{2}+{r(\epsilon^{3})}&0&\vline&\mathtt{f}_{11}\epsilon+r(\epsilon^{3})&0\\ 0&0&\vline&0&0\\ \hline\cr\mathtt{f}_{11}\epsilon+r(\epsilon^{3})&0&\vline&1+r(\epsilon^{2})&0\\ 0&0&\vline&0&0\end{pmatrix} (4.25)

where 𝚎11\mathtt{e}_{11} and 𝚏11\mathtt{f}_{11} are in (4.31) and (4.34). We expand the operator ϵ{\mathcal{B}}_{\epsilon} in (4.16) as

ϵ=0+ϵ1+ϵ22+𝒪(ϵ3),0:=[1𝚌𝚑x𝚌𝚑x|D|tanh(𝚑|D|)],\displaystyle{\mathcal{B}}_{\epsilon}={\mathcal{B}}_{0}+\epsilon{\mathcal{B}}_{1}+\epsilon^{2}{\mathcal{B}}_{2}+\mathcal{O}(\epsilon^{3}),\quad{\mathcal{B}}_{0}:=\begin{bmatrix}1&-{\mathtt{c}}_{\mathtt{h}}\partial_{x}\\ {\mathtt{c}}_{\mathtt{h}}\partial_{x}&|D|\tanh(\mathtt{h}|D|)\end{bmatrix}\,, (4.26)
1:=[a1(x)p1(x)xxp1(x)0],2:=[a2(x)p2(x)xxp2(x)𝚏2x2(1tanh2(𝚑|D|))],\displaystyle{\mathcal{B}}_{1}:=\begin{bmatrix}a_{1}(x)&-p_{1}(x)\partial_{x}\\ \partial_{x}\circ p_{1}(x)&0\end{bmatrix}\,,\ \;{\mathcal{B}}_{2}:=\begin{bmatrix}a_{2}(x)&-p_{2}(x)\partial_{x}\\ \partial_{x}\circ p_{2}(x)&-\mathtt{f}_{2}\partial_{x}^{2}\big{(}1-\tanh^{2}(\mathtt{h}|D|)\big{)}\end{bmatrix}\,,

where the remainder term 𝒪(ϵ3)(Y,X)\mathcal{O}(\epsilon^{3})\in\mathcal{L}(Y,X), the functions a1a_{1}, p1p_{1}, a2a_{2}, p2p_{2} are given in (2.20)-(2.23) and, in view of (2.15), 𝚏2:=14𝚌𝚑2(𝚌𝚑43)\mathtt{f}_{2}:=\tfrac{1}{4}{\mathtt{c}}_{\mathtt{h}}^{-2}({\mathtt{c}}_{\mathtt{h}}^{4}-3).

\bullet Expansion of E11(0,ϵ)=𝚎11ϵ2+r(ϵ3)E_{11}(0,\epsilon)=\mathtt{e}_{11}\epsilon^{2}+r(\epsilon^{3}). By (4.3) we split the real function f1+(0,ϵ)f_{1}^{+}(0,\epsilon) as

f1+(0,ϵ)=f1++ϵf11++ϵ2f12++𝒪(ϵ3),\displaystyle\qquad\qquad f_{1}^{+}(0,\epsilon)=f_{1}^{+}+\epsilon f_{1_{1}}^{+}+\epsilon^{2}f_{1_{2}}^{+}+\mathcal{O}(\epsilon^{3})\,, (4.27)
f1+=[𝚌𝚑12cos(x)𝚌𝚑12sin(x)],f11+:=[α𝚑cos(2x)β𝚑sin(2x)],f12+:=[even0(x)odd(x)],\displaystyle f_{1}^{+}=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\cos(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\sin(x)\end{bmatrix},\ f_{1_{1}}^{+}:=\begin{bmatrix}\alpha_{\mathtt{h}}\cos(2x)\\ \beta_{\mathtt{h}}\sin(2x)\end{bmatrix}\,,\ f_{1_{2}}^{+}:=\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix},

where both f12+f_{1_{2}}^{+} and 𝒪(ϵ3)\mathcal{O}(\epsilon^{3}) are vectors in H1(𝕋)H^{1}(\mathbb{T}). Since 0f1+=𝒥10,0f1+=0{\mathcal{B}}_{0}f_{1}^{+}=\mathcal{J}^{-1}\mathscr{L}_{0,0}f_{1}^{+}=0, and both 0{\mathcal{B}}_{0}, 1{\mathcal{B}}_{1} are self-adjoint real operators, it results

E11(0,ϵ)\displaystyle E_{11}(0,\epsilon) =(ϵf1+(0,ϵ),f1+(0,ϵ))\displaystyle=\left({\mathcal{B}}_{\epsilon}f^{+}_{1}(0,\epsilon)\,,\,f^{+}_{1}(0,\epsilon)\right)
=ϵ(1f1+,f1+)+ϵ2[(2f1+,f1+)+2(1f1+,f11+)+(0f11+,f11+)]+𝒪(ϵ3).\displaystyle=\epsilon\left({\mathcal{B}}_{1}f_{1}^{+}\,,\,f_{1}^{+}\right)+\epsilon^{2}\left[\left({\mathcal{B}}_{2}f_{1}^{+}\,,\,f_{1}^{+}\right)+2\left({\mathcal{B}}_{1}f_{1}^{+}\,,\,f_{1_{1}}^{+}\right)+\left({\mathcal{B}}_{0}f_{1_{1}}^{+}\,,\,f_{1_{1}}^{+}\right)\right]+\mathcal{O}(\epsilon^{3})\,. (4.28)

By (4.26) one has

1f1+=[A1(1+cos(2x))B1sin(2x)],2f1+=[A2cos(x)+A3cos(3x)B2sin(x)+B3sin(3x)],0f11+=[A4cos(2x)B4sin(2x)],\displaystyle{\mathcal{B}}_{1}f_{1}^{+}=\begin{bmatrix}A_{1}(1+\cos(2x))\\ B_{1}\sin(2x)\end{bmatrix},\ \ {\mathcal{B}}_{2}f_{1}^{+}=\begin{bmatrix}A_{2}\cos(x)+A_{3}\cos(3x)\\ B_{2}\sin(x)+B_{3}\sin(3x)\end{bmatrix},\ \ {\mathcal{B}}_{0}f_{1_{1}}^{+}=\begin{bmatrix}A_{4}\cos(2x)\\ B_{4}\sin(2x)\end{bmatrix}\,, (4.29)

with

A1:=12(a1[1]𝚌𝚑12p1[1]𝚌𝚑12),B1:=p1[1]𝚌𝚑12,\displaystyle A_{1}:=\tfrac{1}{2}(a_{1}^{[1]}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}-p_{1}^{[1]}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}),\qquad B_{1}:=-p_{1}^{[1]}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\,, (4.30)
A2:=𝚌𝚑12a2[0]𝚌𝚑12p2[0]+12𝚌𝚑12a2[2]12𝚌𝚑12p2[2],A4:=α𝚑2β𝚑𝚌𝚑,\displaystyle A_{2}:={\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}a_{2}^{[0]}-{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}p_{2}^{[0]}+\tfrac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}a_{2}^{[2]}-\tfrac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}p_{2}^{[2]}\,,\qquad A_{4}:=\alpha_{\mathtt{h}}-2\beta_{\mathtt{h}}{\mathtt{c}}_{\mathtt{h}}\,,
B2:=𝚌𝚑12p2[0]12𝚌𝚑12p2[2]+𝚌𝚑12𝚏2(1𝚌𝚑4),B4:=2α𝚑𝚌𝚑+4𝚌𝚑21+𝚌𝚑4β𝚑.\displaystyle B_{2}:=-{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}p_{2}^{[0]}-\tfrac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}p_{2}^{[2]}+{{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}}\mathtt{f}_{2}(1-{\mathtt{c}}_{\mathtt{h}}^{4})\,,\qquad B_{4}:=-2\alpha_{\mathtt{h}}{\mathtt{c}}_{\mathtt{h}}+{\frac{4{\mathtt{c}}_{\mathtt{h}}^{2}}{1+{\mathtt{c}}_{\mathtt{h}}^{4}}}{\beta_{\mathtt{h}}}\,.

By (4.29) and (4.27), we deduce

E11(0,ϵ)=𝚎11ϵ2+r(ϵ3),𝚎11:=12(A2𝚌𝚑12+B2𝚌𝚑12+2α𝚑A1+2B1β𝚑+α𝚑A4+β𝚑B4).E_{11}(0,\epsilon)=\mathtt{e}_{11}\epsilon^{2}+r(\epsilon^{3})\,,\quad\mathtt{e}_{11}:=\frac{1}{2}\big{(}A_{2}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}+B_{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}+2\alpha_{\mathtt{h}}A_{1}+2B_{1}\beta_{\mathtt{h}}+\alpha_{\mathtt{h}}A_{4}+\beta_{\mathtt{h}}B_{4}\big{)}\,. (4.31)

By (4.31), (4.30), (4.7), (2.20)-(2.23) we obtain (4.13). Since 𝚎11>0\mathtt{e}_{11}>0 the second alternative in (4.23) is ruled out, implying E22(0,ϵ)0E_{22}(0,\epsilon)\equiv 0.
\bullet Expansion of G11(0,ϵ)=1+r(ϵ2)G_{11}(0,\epsilon)=1+r(\epsilon^{2}). By (4.5) we split the real-valued function f0+(0,ϵ)f_{0}^{+}(0,\epsilon) as

f0+(0,ϵ)=f0++ϵf01++ϵ2f02++𝒪(ϵ3),f0+=[10],f01+:=δ𝚑[𝚌𝚑12cos(x)𝚌𝚑12sin(x)],f02+:=[even0(x)odd(x)].f_{0}^{+}(0,\epsilon)=f_{0}^{+}+\epsilon f_{0_{1}}^{+}+\epsilon^{2}f_{0_{2}}^{+}+\mathcal{O}(\epsilon^{3})\,,\ \ f_{0}^{+}=\begin{bmatrix}1\\ 0\end{bmatrix}\,,\ f_{0_{1}}^{+}:=\delta_{\mathtt{h}}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\cos(x)\\ -{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\sin(x)\end{bmatrix}\,,\ f_{0_{2}}^{+}:=\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix}\,. (4.32)

Since, by (2.34) and (4.26), 0f0+=f0+{\mathcal{B}}_{0}f_{0}^{+}=f_{0}^{+}, using that 0{\mathcal{B}}_{0}, 1{\mathcal{B}}_{1} are self-adjoint real operators, and f0+=1\|f_{0}^{+}\|=1, (f0+,f01+)(f_{0}^{+},f_{0_{1}}^{+}), we have G11(0,ϵ)=(ϵf0+(0,ϵ),f0+(0,ϵ))=1+ϵ(1f0+,f0+)+r(ϵ2)G_{11}(0,\epsilon)=\left({\mathcal{B}}_{\epsilon}f^{+}_{0}(0,\epsilon)\,,\,f^{+}_{0}(0,\epsilon)\right)=1+\epsilon\left({\mathcal{B}}_{1}f_{0}^{+}\,,\,f_{0}^{+}\right)+r(\epsilon^{2}). By (4.26) and (2.20)-(2.23) one has

1f0+=[a1[1]cos(x)p1[1]sin(x)]{\mathcal{B}}_{1}f_{0}^{+}=\begin{bmatrix}a_{1}^{[1]}\cos(x)\\ -p_{1}^{[1]}\sin(x)\end{bmatrix} (4.33)

and, by (4.32), we deduce G11(0,ϵ)=1+r(ϵ2)G_{11}(0,\epsilon)=1+r(\epsilon^{2}).
\bullet Expansion of F11(0,ϵ)=𝚏11ϵ+r(ϵ3)F_{11}(0,\epsilon)=\mathtt{f}_{11}\epsilon+r(\epsilon^{3}). By (4.26), (4.27), (4.32), using that 0,1{\mathcal{B}}_{0},{\mathcal{B}}_{1} are self-adjoint and real, and 0f1+=0{\mathcal{B}}_{0}f_{1}^{+}=0, 0f0+=f0+{\mathcal{B}}_{0}f_{0}^{+}=f_{0}^{+}, we obtain

F11(0,ϵ)\displaystyle F_{11}(0,\epsilon) =ϵ[(1f1+,f0+)+(f11+,f0+)]\displaystyle=\epsilon\left[\left({\mathcal{B}}_{1}f_{1}^{+}\,,\,f_{0}^{+}\right)+\left(f_{1_{1}}^{+}\,,\,f_{0}^{+}\right)\right]
+ϵ2[(2f1+,f0+)+(1f1+,f01+)+(1f0+,f11+)+(f12+,f0+)+(0f11+,f01+)]+r(ϵ3).\displaystyle\quad+\epsilon^{2}\big{[}\left({\mathcal{B}}_{2}f_{1}^{+}\,,\,f_{0}^{+}\right)+\left({\mathcal{B}}_{1}f_{1}^{+}\,,\,f_{0_{1}}^{+}\right)+\left({\mathcal{B}}_{1}f_{0}^{+}\,,\,f_{1_{1}}^{+}\right)+\left(f_{1_{2}}^{+}\,,\,f_{0}^{+}\right)+\left({\mathcal{B}}_{0}f_{1_{1}}^{+}\,,\,f_{0_{1}}^{+}\right)\big{]}+r(\epsilon^{3})\,.

By (4.27), (4.29), (4.30), (4.32), (4.33), all these scalar products vanish but the first one, and then

F11(0,ϵ)=𝚏11ϵ+r(ϵ3),𝚏11:=A1=12(a1[1]𝚌𝚑12p1[1]𝚌𝚑12),F_{11}(0,\epsilon)=\mathtt{f}_{11}\epsilon+r(\epsilon^{3})\,,\quad\mathtt{f}_{11}:=A_{1}=\tfrac{1}{2}(a_{1}^{[1]}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}-p_{1}^{[1]}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}})\,, (4.34)

which, by substituting the expressions of a1[1]a_{1}^{[1]}, p1[1]p_{1}^{[1]} in Lemma 2.2, gives the expression in (4.13).

The expansion (4.25) in proved.
Linear terms in μ\mu. We now compute the terms of 𝙱ϵ(μ)\mathtt{B}_{\epsilon}(\mu) that are linear in μ\mu. It results

μ𝙱ϵ(0)=X+XwhereX:=(ϵfkσ(0,ϵ),(μfkσ)(0,ϵ))k,k=0,1,σ,σ=±.\partial_{\mu}\mathtt{B}_{\epsilon}(0)=X+X^{*}\qquad\text{where}\qquad X:=\big{(}{\mathcal{B}}_{\epsilon}f_{k}^{\sigma}(0,\epsilon),(\partial_{\mu}f^{\sigma^{\prime}}_{k^{\prime}})(0,\epsilon)\big{)}_{k,k^{\prime}=0,1,\sigma,\sigma^{\prime}=\pm}\,. (4.35)

We now prove that

X=(𝒪(ϵ3)0𝒪(ϵ2)0𝒪(ϵ2)0𝒪(ϵ)0𝒪(ϵ3)0𝒪(ϵ2)0𝒪(ϵ3)0𝒪(ϵ2)0).X=\begin{pmatrix}\mathcal{O}(\epsilon^{3})&0&\vline&\mathcal{O}(\epsilon^{2})&0\\ \mathcal{O}(\epsilon^{2})&0&\vline&\mathcal{O}(\epsilon)&0\\ \hline\cr\mathcal{O}(\epsilon^{3})&0&\vline&\mathcal{O}(\epsilon^{2})&0\\ \mathcal{O}(\epsilon^{3})&0&\vline&\mathcal{O}(\epsilon^{2})&0\end{pmatrix}. (4.36)

The matrix 𝙻ϵ(0)\mathtt{L}_{\epsilon}(0) in (4.24) where E22(0,ϵ)=0E_{22}(0,\epsilon)=0, represents the action of the operator 0,ϵ:𝒱0,ϵ𝒱0,ϵ\mathscr{L}_{0,\epsilon}:\mathcal{V}_{0,\epsilon}\to\mathcal{V}_{0,\epsilon} in the basis {fkσ(0,ϵ)}\{f^{\sigma}_{k}(0,\epsilon)\} and then we deduce that 0,ϵf1(0,ϵ)=0\mathscr{L}_{0,\epsilon}f_{1}^{-}(0,\epsilon)=0, 0,ϵf0(0,ϵ)=0\mathscr{L}_{0,\epsilon}f_{0}^{-}(0,\epsilon)=0. Thus also ϵf1(0,ϵ)=0{\mathcal{B}}_{\epsilon}f_{1}^{-}(0,\epsilon)=0, ϵf0(0,ϵ)=0{\mathcal{B}}_{\epsilon}f_{0}^{-}(0,\epsilon)=0, and the second and the fourth column of the matrix XX in (4.36) are zero. To compute the other two columns we use the expansion of the derivatives. In view of (4.3)-(4.6) and by denoting with a dot the derivative w.r.t. μ\mu, one has

f˙1+(0,ϵ)=i4γ𝚑[𝚌𝚑12sin(x)𝚌𝚑12cos(x)]+iϵ[odd(x)even(x)]+𝒪(ϵ2),f˙0+(0,ϵ)=iϵ[odd(x)even0(x)]+𝒪(ϵ2),\displaystyle\dot{f}^{+}_{1}(0,\epsilon)=\frac{\mathrm{i}\,}{4}\gamma_{\mathtt{h}}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\sin(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\cos(x)\end{bmatrix}+\mathrm{i}\,\epsilon\begin{bmatrix}odd(x)\\ even(x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\,,\quad\dot{f}^{+}_{0}(0,\epsilon)=\mathrm{i}\,\epsilon\begin{bmatrix}odd(x)\\ even_{0}(x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\,, (4.37)
f˙1(0,ϵ)=i4γ𝚑[𝚌𝚑12cos(x)𝚌𝚑12sin(x)]+iϵ[even(x)odd(x)]+𝒪(ϵ2),f˙0(0,ϵ)=iϵ[even0(x)odd(x)]+𝒪(ϵ2).\displaystyle\dot{f}^{-}_{1}(0,\epsilon)=\frac{\mathrm{i}\,}{4}\gamma_{\mathtt{h}}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\cos(x)\\ -{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\sin(x)\end{bmatrix}+\mathrm{i}\,\epsilon\begin{bmatrix}even(x)\\ odd(x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\,,\ \ \dot{f}^{-}_{0}(0,\epsilon)=\mathrm{i}\,\epsilon\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\,.

In view of (2.2), (4.3)-(4.6), (4.24), (4.8), (4.31),(4.34), and since ϵfkσ(0,ϵ)=𝒥ϵfkσ(0,ϵ){\mathcal{B}}_{\epsilon}f_{k}^{\sigma}(0,\epsilon)=-\mathcal{J}\mathscr{L}_{\epsilon}f_{k}^{\sigma}(0,\epsilon), we have

ϵf1+(0,ϵ)\displaystyle{\mathcal{B}}_{\epsilon}f_{1}^{+}(0,\epsilon) =E11(0,ϵ)𝒥f1(0,ϵ)+F11(0,ϵ)𝒥f0=ϵ[𝚏110]+ϵ2𝚎11[𝚌𝚑12cos(x)𝚌𝚑12sin(x)]+𝒪(ϵ3),\displaystyle=E_{11}(0,\epsilon)\,\mathcal{J}f_{1}^{-}(0,\epsilon)+F_{11}(0,\epsilon)\,\mathcal{J}f_{0}^{-}=\epsilon\begin{bmatrix}\mathtt{f}_{11}\\ 0\end{bmatrix}+\epsilon^{2}\mathtt{e}_{11}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\cos(x)\\ {\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\sin(x)\end{bmatrix}+\mathcal{O}(\epsilon^{3})\,, (4.38)
ϵf0+(0,ϵ)\displaystyle{\mathcal{B}}_{\epsilon}f_{0}^{+}(0,\epsilon) =F11(0,ϵ)𝒥f1(0,ϵ)+G11(0,ϵ)𝒥f0=[10]+ϵ𝚏11[𝚌𝚑12cos(2x)𝚌𝚑12sin(2x)]+𝒪(ϵ2).\displaystyle=F_{11}(0,\epsilon)\,\mathcal{J}f_{1}^{-}(0,\epsilon)+G_{11}(0,\epsilon)\,\mathcal{J}f_{0}^{-}=\begin{bmatrix}1\\ 0\end{bmatrix}+\epsilon\mathtt{f}_{11}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\cos(2x)\\ {\mathtt{c}}_{\mathtt{h}}^{\frac{1}{2}}\sin(2x)\end{bmatrix}+\mathcal{O}(\epsilon^{2})\,.

We deduce (4.36) by (4.37) and (4.38).
Quadratic terms in μ\mu. By denoting with a double dot the double derivative w.r.t. μ\mu, we have

μ2𝙱0(0)=(0fkσ,f¨kσ(0,0))+(f¨kσ(0,0),0fkσ)+2(0f˙kσ(0,0),f˙kσ(0,0))=:Y+Y+2Z.\partial_{\mu}^{2}\mathtt{B}_{0}(0)=\left({\mathcal{B}}_{0}f_{k}^{\sigma}\,,\,\ddot{f}_{k^{\prime}}^{\sigma^{\prime}}(0,0)\right)+\left(\ddot{f}_{k}^{\sigma}(0,0)\,,\,{\mathcal{B}}_{0}f_{k}^{\sigma^{\prime}}\right)+2\left({\mathcal{B}}_{0}\dot{f}_{k}^{\sigma}(0,0)\,,\,\dot{f}_{k^{\prime}}^{\sigma^{\prime}}(0,0)\right)=:Y+Y^{*}+2Z\,. (4.39)

We claim that Y=0Y=0. Indeed, its first, second and fourth column are zero, since 0fkσ=0{\mathcal{B}}_{0}f_{k}^{\sigma}=0 for fkσ{f1+,f1,f0}f_{k}^{\sigma}\in\{f_{1}^{+},f_{1}^{-},f_{0}^{-}\}. The third column is also zero by noting that 0f0+=f0+{\mathcal{B}}_{0}f_{0}^{+}=f_{0}^{+} and

f¨1+(0,0)=[even0(x)+iodd(x)odd(x)+ieven0(x)],f¨1(0,0)=[odd(x)+ieven0(x)even0(x)+iodd(x)],f¨0+(0,0)=f¨0(0,0)=0.\ddot{f}_{1}^{+}(0,0)=\begin{bmatrix}even_{0}(x)+\mathrm{i}\,odd(x)\\ odd(x)+\mathrm{i}\,even_{0}(x)\end{bmatrix},\ \ \ddot{f}_{1}^{-}(0,0)=\begin{bmatrix}odd(x)+\mathrm{i}\,even_{0}(x)\\ even_{0}(x)+\mathrm{i}\,odd(x)\end{bmatrix},\ \ \ddot{f}_{0}^{+}(0,0)=\ddot{f}_{0}^{-}(0,0)=0\,.

We claim that

Z=(0f˙kσ(0,0),f˙kσ(0,0))k,k=0,1,σ,σ=±=(ζ𝚑0000ζ𝚑0000000000),\displaystyle Z=\left({\mathcal{B}}_{0}\dot{f}_{k}^{\sigma}(0,0)\,,\,\dot{f}_{k^{\prime}}^{\sigma^{\prime}}(0,0)\right)_{\begin{subarray}{c}k,k^{\prime}=0,1,\\ \sigma,\sigma^{\prime}=\pm\end{subarray}}=\begin{pmatrix}\zeta_{\mathtt{h}}&0&\vline&0&0\\ 0&\zeta_{\mathtt{h}}&\vline&0&0\\ \hline\cr 0&0&\vline&0&0\\ 0&0&\vline&0&0\\ \end{pmatrix}\,, (4.40)

with ζ𝚑\zeta_{\mathtt{h}} as in (4.20). Indeed, by (4.37), we have f˙0+(0,0)=f˙0(0,0)=0\dot{f}^{+}_{0}(0,0)=\dot{f}^{-}_{0}(0,0)=0. Therefore the last two columns of ZZ, and by self-adjointness the last two rows, are zero. By (4.26), (4.37) we obtain the matrix (4.40) with

ζ𝚑:=(0f˙1+(0,0),f˙1+(0,0))=(0f˙1(0,0),f˙1(0,0))=18𝚌𝚑γ𝚑2.\zeta_{\mathtt{h}}:=\left({\mathcal{B}}_{0}\dot{f}^{+}_{1}(0,0)\,,\,\dot{f}^{+}_{1}(0,0)\right)=\left({\mathcal{B}}_{0}\dot{f}^{-}_{1}(0,0)\,,\,\dot{f}^{-}_{1}(0,0)\right)=\tfrac{1}{8}{\mathtt{c}}_{\mathtt{h}}\gamma_{\mathtt{h}}^{2}\,.

In conclusion (4.21), (4.35), (4.36), (4.39), the fact that Y=0Y=0 and (4.40) imply (4.19), using also the selfadjointness of 𝙱ϵ\mathtt{B}_{\epsilon} and (3.13). ∎

We now consider {\mathcal{B}}^{\flat}.

Lemma 4.5.

(Expansion of 𝙱\mathtt{B}^{\flat}) The self-adjoint and reversibility-preserving matrix 𝙱\mathtt{B}^{\flat} associated, as in (3.12), to the self-adjoint and reversibility-preserving operator {\mathcal{B}}^{\flat}, defined in (4.17), with respect to the basis \mathcal{F} of 𝒱μ,ϵ{\mathcal{V}}_{\mu,\epsilon} in (4.1), admits the expansion

𝙱=(μ24b𝚑i(μ2𝚎12+r2(μϵ2))00i(μ2𝚎12+r2(μϵ2))μ24b𝚑ir6(μϵ)00ir6(μϵ)00000μtanh(𝚑μ))+𝒪(μ2ϵ,μ3)\mathtt{B}^{\flat}=\begin{pmatrix}-\frac{\mu^{2}}{4}\text{{b}}_{\mathtt{h}}&\mathrm{i}\,(\frac{\mu}{2}\mathtt{e}_{12}+r_{2}(\mu\epsilon^{2}))&\vline&0&0\\ -\mathrm{i}\,(\frac{\mu}{2}\mathtt{e}_{12}+r_{2}(\mu\epsilon^{2}))&-\frac{\mu^{2}}{4}\text{{b}}_{\mathtt{h}}&\vline&\mathrm{i}\,r_{6}(\mu\epsilon)&0\\ \hline\cr 0&-\mathrm{i}\,r_{6}(\mu\epsilon)&\vline&0&0\\ 0&0&\vline&0&\mu\tanh(\mathtt{h}\mu)\end{pmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3}) (4.41)

where 𝚎12\mathtt{e}_{12} is defined in (1.2)  and

b𝚑:=γ𝚑𝚌𝚑+𝚌𝚑1𝚑(1𝚌𝚑4)(γ𝚑2(1𝚌𝚑2𝚑)).\text{{b}}_{\mathtt{h}}:=\gamma_{\mathtt{h}}{\mathtt{c}}_{\mathtt{h}}+{\mathtt{c}}_{\mathtt{h}}^{-1}\mathtt{h}(1-{\mathtt{c}}_{\mathtt{h}}^{4})(\gamma_{\mathtt{h}}-2(1-{\mathtt{c}}_{\mathtt{h}}^{2}\mathtt{h}))\,\,. (4.42)
Proof.

We have to compute the expansion of the matrix entries (fkσ(μ,ϵ),fkσ(μ,ϵ))({\mathcal{B}}^{\flat}f^{\sigma}_{k}(\mu,\epsilon),f^{\sigma^{\prime}}_{k^{\prime}}(\mu,\epsilon)). First, by (4.6), (4.17) and since 𝚏ϵ=O(ϵ2)\mathtt{f}_{\epsilon}=O(\epsilon^{2}) (cfr. (2.15)) we have

f0(μ,ϵ)=[0μtanh(𝚑μ)]+[0𝒪(μ2ϵ)].\displaystyle{\mathcal{B}}^{\flat}f^{-}_{0}(\mu,\epsilon)=\begin{bmatrix}0\\ \mu\tanh\big{(}\mathtt{h}\mu\big{)}\end{bmatrix}+\begin{bmatrix}0\\ \mathcal{O}(\mu^{2}\epsilon)\end{bmatrix}\,.

Hence, by (4.3)-(4.6), the entries of the last column (and row) of 𝙱\mathtt{B}^{\flat} are

(f0(μ,ϵ),f1+(μ,ϵ))=𝒪(μ2ϵ),(f0(μ,ϵ),f1(μ,ϵ))=μtanh(𝚑μ)𝒪(ϵ2)+𝒪(μ2ϵ2)=𝒪(μ2ϵ2)\displaystyle\big{(}{\mathcal{B}}^{\flat}f^{-}_{0}(\mu,\epsilon),f^{+}_{1}(\mu,\epsilon)\big{)}=\mathcal{O}(\mu^{2}\epsilon)\ ,\quad\big{(}{\mathcal{B}}^{\flat}f^{-}_{0}(\mu,\epsilon),f^{-}_{1}(\mu,\epsilon)\big{)}=\mu\tanh(\mathtt{h}\mu)\mathcal{O}(\epsilon^{2})+\mathcal{O}(\mu^{2}\epsilon^{2})=\mathcal{O}(\mu^{2}\epsilon^{2})
(f0(μ,ϵ),f0+(μ,ϵ))=𝒪(μ2ϵ,μ3),(f0(μ,ϵ),f0(μ,ϵ))=μtanh(𝚑μ)+𝒪(μ2ϵ),\displaystyle\big{(}{\mathcal{B}}^{\flat}f^{-}_{0}(\mu,\epsilon),f^{+}_{0}(\mu,\epsilon)\big{)}=\mathcal{O}(\mu^{2}\epsilon,\mu^{3})\ ,\quad\big{(}{\mathcal{B}}^{\flat}f^{-}_{0}(\mu,\epsilon),f^{-}_{0}(\mu,\epsilon)\big{)}=\mu\tanh(\mathtt{h}\mu)+\mathcal{O}(\mu^{2}\epsilon)\,,

in agreement with (4.41).

In order to compute the other matrix entries we expand {\mathcal{B}}^{\flat} in (4.17) at μ=0\mu=0, obtaining

=μ1(0)+μ(ϵ)+μ22+𝒪(μ2ϵ,μ3),where\displaystyle{\mathcal{B}}^{\flat}=\mu{\mathcal{B}}^{\flat}_{1}{(0)}+\mu{\mathcal{R}}^{\flat}(\epsilon)+\mu^{2}{\mathcal{B}}^{\flat}_{2}+{\mathcal{O}(\mu^{2}\epsilon,\mu^{3})\,,\quad\text{where}} (4.43)
1(0):=[𝚑D(1tanh2(𝚑|D|))+sgn(D)tanh(𝚑|D|)]ΠII,ΠII:=[000Id],\displaystyle{\mathcal{B}}^{\flat}_{1}(0):=\Big{[}\mathtt{h}D\big{(}1-\tanh^{2}(\mathtt{h}|D|)\big{)}+\operatorname*{sgn}(D)\tanh(\mathtt{h}|D|)\Big{]}\Pi_{{\mathrm{I\!I}}}\,,\quad\Pi_{{\mathrm{I\!I}}}:=\begin{bmatrix}0&0\\ 0&\mathrm{Id}\end{bmatrix}\,,
(ϵ):=𝒪(ϵ2)ΠII,2:=[𝚑(1tanh2(𝚑|D|))(1𝚑tanh(𝚑|D|)|D|)]ΠII.\displaystyle{\mathcal{R}}^{\flat}(\epsilon):=\mathcal{O}(\epsilon^{2})\Pi_{{\mathrm{I\!I}}}\,,\qquad{\mathcal{B}}^{\flat}_{2}:={\Big{[}\mathtt{h}\big{(}1-\tanh^{2}(\mathtt{h}|D|)\big{)}\big{(}1-\mathtt{h}\tanh(\mathtt{h}|D|)|D|\big{)}\Big{]}\Pi_{{\mathrm{I\!I}}}\,.}

We note that

μ((ϵ)fkσ(μ,ϵ),fkσ(μ,ϵ))=μ(fkσ(0,ϵ),fkσ(0,ϵ))+𝒪(μ2ϵ2)={𝒪(μ2ϵ2)if σ=σ,𝒪(μϵ2)if σσ.\mu\big{(}{\mathcal{R}}^{\flat}(\epsilon)f^{\sigma}_{k}(\mu,\epsilon),f^{\sigma^{\prime}}_{k^{\prime}}(\mu,\epsilon)\big{)}=\mu\big{(}{\mathcal{R}}^{\flat}f^{\sigma}_{k}(0,\epsilon),f^{\sigma^{\prime}}_{k^{\prime}}(0,\epsilon)\big{)}+\mathcal{O}(\mu^{2}\epsilon^{2})=\begin{cases}\mathcal{O}(\mu^{2}\epsilon^{2})&\mbox{if }\sigma=\sigma^{\prime}\,,\\ \mathcal{O}(\mu\epsilon^{2})&\mbox{if }\sigma\neq\sigma^{\prime}\,.\end{cases} (4.44)

Indeed, if σ=σ\sigma=\sigma^{\prime}, (fkσ(0,ϵ),fkσ(0,ϵ))\big{(}{\mathcal{R}}^{\flat}f^{\sigma}_{k}(0,\epsilon),f^{\sigma^{\prime}}_{k^{\prime}}(0,\epsilon)\big{)} is real by (3.13), but purely imaginary444 An operator 𝒜\mathcal{A} is purely imaginary if 𝒜¯=𝒜\overline{\mathcal{A}}=-\mathcal{A}. A purely imaginary operator sends real functions into purely imaginary ones. too, since the operator {\mathcal{R}}^{\flat} is purely imaginary (as {\mathcal{B}}^{\flat} is) and the basis {fk±(0,ϵ)}k=0,1\{f_{k}^{\pm}(0,\epsilon)\}_{k=0,1} is real. The terms (4.44) contribute to r2(μϵ2)r_{2}(\mu\epsilon^{2}) and r6(ϵμ)r_{6}(\epsilon\mu) in (4.41).

Next we compute the other scalar products. By (4.3), (4.43), and the identities sgn(D)sin(kx)=icos(kx)\operatorname*{sgn}(D)\sin(kx)=-\mathrm{i}\,\cos(kx) and sgn(D)cos(kx)=isin(kx)\operatorname*{sgn}(D)\cos(kx)=\mathrm{i}\,\sin(kx) for any kk\in\mathbb{N}, we have

μ1(0)f1+(μ,ϵ)=iμ1[0cos(x)]μ24γ𝚑1[0sin(x)]iμϵ2[0cos(2x)]+i𝒪(μϵ2)[0even0(x)]+𝒪(μ2ϵ,μ3)\mu{\mathcal{B}}^{\flat}_{1}(0)f^{+}_{1}(\mu,\epsilon)=\footnotesize-\mathrm{i}\,\mu\text{\large\bf$\flat$}_{1}\begin{bmatrix}0\\ \cos(x)\end{bmatrix}-\frac{\mu^{2}}{4}\gamma_{\mathtt{h}}\text{\large\bf$\flat$}_{1}\begin{bmatrix}0\\ \sin(x)\end{bmatrix}-\mathrm{i}\,\mu\epsilon\text{\large\bf$\flat$}_{2}\begin{bmatrix}0\\ \cos(2x)\end{bmatrix}+\mathrm{i}\,\mathcal{O}(\mu\epsilon^{2})\begin{bmatrix}0\\ even_{0}(x)\end{bmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3})

where

1:=𝚌𝚑12(𝚌𝚑2+(1𝚌𝚑4)𝚑)\displaystyle\text{\large\bf$\flat$}_{1}:={\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}({\mathtt{c}}_{\mathtt{h}}^{2}+(1-{\mathtt{c}}_{\mathtt{h}}^{4})\mathtt{h}) (4.45)
2:=β𝚑(tanh(2𝚑)+2𝚑(1tanh2(2𝚑)))=β𝚑(2𝚌𝚑21+𝚌𝚑4+2𝚑(14𝚌𝚑4(1+𝚌𝚑4)2)).\displaystyle\text{\large\bf$\flat$}_{2}:=\beta_{\mathtt{h}}\Big{(}\tanh(2\mathtt{h})+2\mathtt{h}(1-\tanh^{2}(2\mathtt{h}))\Big{)}=\beta_{\mathtt{h}}\Big{(}\frac{2{\mathtt{c}}_{\mathtt{h}}^{2}}{1+{\mathtt{c}}_{\mathtt{h}}^{4}}+2\mathtt{h}\big{(}1-\frac{4{\mathtt{c}}_{\mathtt{h}}^{4}}{(1+{\mathtt{c}}_{\mathtt{h}}^{4})^{2}}\big{)}\Big{)}\,.

Similarly μ22f1+(μ,ϵ)=μ23[0sin(x)]+𝒪(μ2ϵ,μ3)\mu^{2}{\mathcal{B}}^{\flat}_{2}f^{+}_{1}(\mu,\epsilon)=\mu^{2}\text{\large\bf$\flat$}_{3}\footnotesize{\begin{bmatrix}0\\ \sin(x)\end{bmatrix}}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3}), where

3:=𝚑(1tanh2(𝚑))(1tanh(𝚑)𝚑)𝚌𝚑12=𝚑(1𝚌𝚑4)(1𝚌𝚑2𝚑)𝚌𝚑12.\text{\large\bf$\flat$}_{3}:=\mathtt{h}\big{(}1-\tanh^{2}(\mathtt{h})\big{)}\big{(}1-\tanh(\mathtt{h})\mathtt{h}\big{)}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}=\mathtt{h}(1-{\mathtt{c}}_{\mathtt{h}}^{4})(1-{\mathtt{c}}_{\mathtt{h}}^{2}\mathtt{h}){\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\,. (4.46)

Analogously, using (4.4),

μ1(0)f1(μ,ϵ)=iμ1[0sin(x)]μ24γ𝚑1[0cos(x)]+iμϵ3[0sin(2x)]+i𝒪(μϵ2)[0odd(x)]+𝒪(μ2ϵ,μ3),\footnotesize\mu{\mathcal{B}}^{\flat}_{1}(0)f^{-}_{1}(\mu,\epsilon)=\mathrm{i}\,\mu\text{\large\bf$\flat$}_{1}\begin{bmatrix}0\\ \sin(x)\end{bmatrix}-\frac{\mu^{2}}{4}\gamma_{\mathtt{h}}\text{\large\bf$\flat$}_{1}\begin{bmatrix}0\\ \cos(x)\end{bmatrix}+\mathrm{i}\,\mu\epsilon\text{\large\bf$\flat$}_{3}\begin{bmatrix}0\\ \sin(2x)\end{bmatrix}+\mathrm{i}\,\mathcal{O}(\mu\epsilon^{2})\begin{bmatrix}0\\ odd(x)\end{bmatrix}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3})\,,

and μ22f1(μ,ϵ)=μ23[0cos(x)]+𝒪(μ2ϵ,μ3)\mu^{2}{\mathcal{B}}^{\flat}_{2}f^{-}_{1}(\mu,\epsilon)=\mu^{2}\text{\large\bf$\flat$}_{3}\footnotesize{\begin{bmatrix}0\\ \cos(x)\end{bmatrix}}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3}), with j\text{\large\bf$\flat$}_{j}, j=1,2,3j=1,2,3, defined in (4.45) and (4.46). In addition, by (4.5)-(4.6), we get that

μ1(0)f0+(μ,ϵ)=iμϵδ𝚑1[0cos(x)]+i𝒪(μϵ2)[0even0(x)]+𝒪(μ2ϵ),μ22f0+(μ,ϵ)=[0𝒪(μ2ϵ)]\mu{\mathcal{B}}^{\flat}_{1}(0)f^{+}_{0}(\mu,\epsilon)=\mathrm{i}\,\mu\epsilon\delta_{\mathtt{h}}\text{\large\bf$\flat$}_{1}\begin{bmatrix}0\\ \cos(x)\end{bmatrix}+\mathrm{i}\,\mathcal{O}(\mu\epsilon^{2})\begin{bmatrix}0\\ even_{0}(x)\end{bmatrix}+\mathcal{O}(\mu^{2}\epsilon)\,,\ \ \mu^{2}{\mathcal{B}}^{\flat}_{2}f^{+}_{0}(\mu,\epsilon)=\begin{bmatrix}0\\ \mathcal{O}(\mu^{2}\epsilon)\end{bmatrix}\,

with 1\text{\large\bf$\flat$}_{1} in (4.45). By taking the scalar products of the above expansions of fkσ(μ,ϵ){\mathcal{B}}^{\flat}f^{\sigma}_{k}(\mu,\epsilon) with the functions fkσ(μ,ϵ)f^{\sigma^{\prime}}_{k^{\prime}}(\mu,\epsilon) expanded as in (4.3)-(4.6) we obtain that (recall that the scalar product is conjugate-linear in the second component)

(μ1(0)f1+(μ,ϵ),f1+(μ,ϵ)),(μ1(0)f1(μ,ϵ),f1(μ,ϵ))=μ24γ𝚑1𝚌𝚑12+𝒪(μ2ϵ,μ3)\displaystyle\big{(}\mu{\mathcal{B}}^{\flat}_{1}(0)f^{+}_{1}(\mu,\epsilon),f^{+}_{1}(\mu,\epsilon)\big{)}\,,\ \big{(}\mu{\mathcal{B}}^{\flat}_{1}(0)f^{-}_{1}(\mu,\epsilon),f^{-}_{1}(\mu,\epsilon)\big{)}={-\frac{\mu^{2}}{4}\gamma_{\mathtt{h}}\text{\large\bf$\flat$}_{1}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3})
(μ22f1+(μ,ϵ),f1+(μ,ϵ)),(μ22f1(μ,ϵ),f1(μ,ϵ))=μ223𝚌𝚑12+𝒪(μ2ϵ,μ3)\displaystyle\big{(}\mu^{2}{\mathcal{B}}^{\flat}_{2}f^{+}_{1}(\mu,\epsilon),f^{+}_{1}(\mu,\epsilon)\big{)}\,,\ \big{(}\mu^{2}{\mathcal{B}}^{\flat}_{2}f^{-}_{1}(\mu,\epsilon),f^{-}_{1}(\mu,\epsilon)\big{)}={\frac{\mu^{2}}{2}\text{\large\bf$\flat$}_{3}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}}+\mathcal{O}(\mu^{2}\epsilon,\mu^{3})

and, recalling (4.43), (4.45), (4.46), we deduce the expansion of the entries (1,1)(1,1) and (2,2)(2,2) of the matrix 𝙱\mathtt{B}^{\flat} in (4.41) with b𝚑=𝚌𝚑12(γ𝚑123)\text{{b}}_{\mathtt{h}}={\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}(\gamma_{\mathtt{h}}\text{\large\bf$\flat$}_{1}-2\text{\large\bf$\flat$}_{3}) in (4.42). Moreover

(μ1(0)f1(μ,ϵ),f1+(μ,ϵ))=iμ2𝚎12+𝒪(μϵ2,μ2ϵ,μ3),(μ22f1(μ,ϵ),f1+(μ,ϵ))=𝒪(μ3,μ2ϵ),\big{(}\mu{\mathcal{B}}^{\flat}_{1}(0)f^{-}_{1}(\mu,\epsilon),f^{+}_{1}(\mu,\epsilon)\big{)}={\mathrm{i}\,\frac{\mu}{2}\mathtt{e}_{12}}+\mathcal{O}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\,,\ \ \big{(}\mu^{2}{\mathcal{B}}^{\flat}_{2}f^{-}_{1}(\mu,\epsilon),f^{+}_{1}(\mu,\epsilon)\big{)}=\mathcal{O}(\mu^{3},\mu^{2}\epsilon)\,,

where 𝚎12:=1𝚌𝚑12\mathtt{e}_{12}:=\text{\large\bf$\flat$}_{1}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}} is equal to (1.2). Finally we obtain

(μ(1(0)+μ2)f1(μ,ϵ),f0+(μ,ϵ))=𝒪(μϵ,μ3)\displaystyle\big{(}\mu({\mathcal{B}}^{\flat}_{1}(0)+\mu{\mathcal{B}}^{\flat}_{2})f^{-}_{1}(\mu,\epsilon),f^{+}_{0}(\mu,\epsilon)\big{)}=\mathcal{O}(\mu\epsilon,\mu^{3})
(μ(1(0)+μ2)f1+(μ,ϵ),f0+(μ,ϵ))=𝒪(μ3,μ2ϵ),\displaystyle(\mu({\mathcal{B}}^{\flat}_{1}(0)+\mu{\mathcal{B}}^{\flat}_{2})f^{+}_{1}(\mu,\epsilon),f^{+}_{0}(\mu,\epsilon))=\mathcal{O}(\mu^{3},\mu^{2}\epsilon)\,,
(μ(1(0)+μ2)f0+(μ,ϵ),f0+(μ,ϵ))=𝒪(μ2ϵ2).\displaystyle\big{(}\mu({\mathcal{B}}^{\flat}_{1}(0)+\mu{\mathcal{B}}^{\flat}_{2})f^{+}_{0}(\mu,\epsilon),f^{+}_{0}(\mu,\epsilon)\big{)}=\mathcal{O}(\mu^{2}\epsilon^{2})\,.

The expansion (4.41) is proved. ∎

Finally we consider {\mathcal{B}}^{\sharp}.

Lemma 4.6.

(Expansion of 𝙱\mathtt{B}^{\sharp}) The self-adjoint and reversibility-preserving matrix 𝙱\mathtt{B}^{\sharp} associated, as in (3.12), to the self-adjoint and reversibility-preserving operators {\mathcal{B}}^{\sharp}, defined in (4.18), with respect to the basis \mathcal{F} of 𝒱μ,ϵ{\mathcal{V}}_{\mu,\epsilon} in (4.1), admits the expansion

𝙱=(0ir2(μϵ2)0iμϵ𝚌𝚑12+ir4(μϵ2)ir2(μϵ2)0ir6(μϵ)00ir6(μϵ)0ir9(μϵ2)iμϵ𝚌𝚑12ir4(μϵ2)0ir9(μϵ2)0)+𝒪(μ2ϵ).\mathtt{B}^{\sharp}=\begin{pmatrix}0&\mathrm{i}\,r_{2}(\mu\epsilon^{2})&\vline&0&\mathrm{i}\,\mu\epsilon{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}+\mathrm{i}\,r_{4}(\mu\epsilon^{2})\\ -\mathrm{i}\,r_{2}(\mu\epsilon^{2})&0&\vline&-\mathrm{i}\,r_{6}(\mu\epsilon)&0\\ \hline\cr 0&\mathrm{i}\,r_{6}(\mu\epsilon)&\vline&0&-\mathrm{i}\,r_{9}(\mu\epsilon^{2})\\ -\mathrm{i}\,\mu\epsilon{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}-\mathrm{i}\,r_{4}(\mu\epsilon^{2})&0&\vline&\mathrm{i}\,r_{9}(\mu\epsilon^{2})&0\end{pmatrix}+\mathcal{O}(\mu^{2}\epsilon)\,. (4.47)
Proof.

Since =iμpϵ𝒥{\mathcal{B}}^{\sharp}=-\mathrm{i}\,\mu p_{\epsilon}\mathcal{J} and pϵ=𝒪(ϵ)p_{\epsilon}=\mathcal{O}(\epsilon) by (2.19), we have the expansion

(fkσ(μ,ϵ),fkσ(μ,ϵ))=(fkσ(0,ϵ),fkσ(0,ϵ))+𝒪(μ2ϵ).\big{(}{\mathcal{B}}^{\sharp}f_{k}^{\sigma}(\mu,\epsilon),f_{k^{\prime}}^{\sigma^{\prime}}(\mu,\epsilon)\big{)}=\big{(}{\mathcal{B}}^{\sharp}f_{k}^{\sigma}(0,\epsilon),f_{k^{\prime}}^{\sigma^{\prime}}(0,\epsilon)\big{)}+\mathcal{O}(\mu^{2}\epsilon)\,. (4.48)

The matrix entries (fkσ(0,ϵ),fkσ(0,ϵ))({\mathcal{B}}^{\sharp}f^{\sigma}_{k}(0,\epsilon),f^{\sigma}_{k^{\prime}}(0,\epsilon)), k,k=0,1k,k^{\prime}=0,1, σ={±}\sigma=\{\pm\} are zero, because they are simultaneously real by (3.13), and purely imaginary, being the operator {\mathcal{B}}^{\sharp} purely imaginary and the basis {fk±(0,ϵ)}k=0,1\{f_{k}^{\pm}(0,\epsilon)\}_{k=0,1} real. Hence 𝙱\mathtt{B}^{\sharp} has the form

𝙱=(0iβ0iδiβ0iγ00iγ0iηiδ0iη0)+𝒪(μ2ϵ)where{(f1(0,ϵ),f1+(0,ϵ))=:iβ,(f1(0,ϵ),f0+(0,ϵ))=:iγ,(f0(0,ϵ),f1+(0,ϵ))=:iδ,(f0(0,ϵ),f0+(0,ϵ))=:iη,\mathtt{B}^{\sharp}=\begin{pmatrix}0&\mathrm{i}\,\beta&\vline&0&\mathrm{i}\,\delta\\ -\mathrm{i}\,\beta&0&\vline&-\mathrm{i}\,\gamma&0\\ \hline\cr 0&\mathrm{i}\,\gamma&\vline&0&\mathrm{i}\,\eta\\ -\mathrm{i}\,\delta&0&\vline&-\mathrm{i}\,\eta&0\end{pmatrix}+\mathcal{O}(\mu^{2}\epsilon)\quad\text{where}\quad\left\{\begin{matrix}\left({\mathcal{B}}^{\sharp}f_{1}^{-}(0,\epsilon)\,,\,f_{1}^{+}(0,\epsilon)\right)=:\mathrm{i}\,\beta\,,\\ \left({\mathcal{B}}^{\sharp}f_{1}^{-}(0,\epsilon)\,,\,f_{0}^{+}(0,\epsilon)\right)=:\mathrm{i}\,\gamma\,,\\ \left({\mathcal{B}}^{\sharp}f_{0}^{-}(0,\epsilon)\,,\,f_{1}^{+}(0,\epsilon)\right)=:\mathrm{i}\,\delta\,,\\ \left({\mathcal{B}}^{\sharp}f_{0}^{-}(0,\epsilon)\,,\,f_{0}^{+}(0,\epsilon)\right)=:\mathrm{i}\,\eta\,,\end{matrix}\right. (4.49)

and α\alpha, β\beta, γ\gamma, δ\delta are real numbers. As =𝒪(μϵ){\mathcal{B}}^{\sharp}=\mathcal{O}(\mu\epsilon) in (Y)\mathcal{L}(Y), we deduce that γ=r(μϵ)\gamma=r(\mu\epsilon). Let us compute the expansion of β\beta, δ\delta and η\eta. By (2.20) and (2.2) we write the operator {\mathcal{B}}^{\sharp} in (4.18) as

=iμϵ1+𝒪(μϵ2),1:=2𝚌𝚑1cos(x)[0IdId0],{\mathcal{B}}^{\sharp}=\mathrm{i}\,\mu\epsilon{\mathcal{B}}_{1}^{\sharp}+\mathcal{O}(\mu\epsilon^{2})\,,\quad{\mathcal{B}}_{1}^{\sharp}:=2{\mathtt{c}}_{\mathtt{h}}^{-1}\cos(x)\begin{bmatrix}0&\mathrm{Id}\\ -\mathrm{Id}&0\end{bmatrix}\,, (4.50)

with 𝒪(μϵ2)(Y)\mathcal{O}(\mu\epsilon^{2})\in\mathcal{L}(Y). In view of (4.3)-(4.6), f1±(0,ϵ)=f1±+𝒪(ϵ)f_{1}^{\pm}(0,\epsilon)=f_{1}^{\pm}+\mathcal{O}(\epsilon), f0+(0,ϵ)=f0++𝒪(ϵ)f_{0}^{+}(0,\epsilon)=f_{0}^{+}+\mathcal{O}(\epsilon), f0(0,ϵ)=[01]f_{0}^{-}(0,\epsilon)=\footnotesize\begin{bmatrix}0\\ 1\end{bmatrix}, where fkσf_{k}^{\sigma} are in (4.2). By (4.50) we have 1f1=[𝚌𝚑32(1+cos(2x))𝚌𝚑12sin(2x)]\footnotesize{\mathcal{B}}_{1}^{\sharp}f_{1}^{-}=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{-\frac{3}{2}}(1+\cos(2x))\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\sin(2x)\end{bmatrix}, 1f0=[2𝚌𝚑1cos(x)0]\footnotesize{\mathcal{B}}_{1}^{\sharp}f_{0}^{-}=\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1}\cos(x)\\ 0\end{bmatrix} and then

β=μϵ(1f1,f1+)+r(μϵ2)=r(μϵ2),\displaystyle\beta=\mu\epsilon\left({\mathcal{B}}_{1}^{\sharp}f_{1}^{-}\,,\,f_{1}^{+}\right)+r(\mu\epsilon^{2})=r(\mu\epsilon^{2})\,,
δ=μϵ(1f0,f1+)+r(μϵ2)=μϵ𝚌𝚑12+r(μϵ2),\displaystyle\delta=\mu\epsilon\left({\mathcal{B}}_{1}^{\sharp}f_{0}^{-}\,,\,f_{1}^{+}\right)+r(\mu\epsilon^{2})=\mu\epsilon{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}+r(\mu\epsilon^{2})\,,
η=μϵ(1f0,f0+)+r(μϵ2)=r(μϵ2).\displaystyle\eta=\mu\epsilon\left({\mathcal{B}}_{1}^{\sharp}f_{0}^{-}\,,\,f_{0}^{+}\right)+r(\mu\epsilon^{2})=r(\mu\epsilon^{2})\,.

This proves (4.47). ∎

Lemmata 4.4, 4.5, 4.6 imply (4.9) where the matrix EE has the form (4.10) and

𝚎22:=2(b𝚑4ζ𝚑)=2γ𝚑𝚌𝚑+2𝚌𝚑1𝚑(1𝚌𝚑4)(γ𝚑2(1𝚌𝚑2𝚑))𝚌𝚑γ𝚑2,\mathtt{e}_{22}:=2(\textbf{b}_{\mathtt{h}}-4\zeta_{\mathtt{h}})=2\gamma_{\mathtt{h}}{\mathtt{c}}_{\mathtt{h}}+2{\mathtt{c}}_{\mathtt{h}}^{-1}\mathtt{h}(1-{\mathtt{c}}_{\mathtt{h}}^{4})(\gamma_{\mathtt{h}}-2(1-{\mathtt{c}}_{\mathtt{h}}^{2}\mathtt{h}))-{\mathtt{c}}_{\mathtt{h}}\gamma_{\mathtt{h}}^{2}\,,

with b𝚑\textbf{b}_{\mathtt{h}} in (4.42) and ζ𝚑\zeta_{\mathtt{h}} in (4.20). The term 𝚎22\mathtt{e}_{22} has the expansion in (1.3). Moreover

G:=G(μ,ϵ)=(1+r8(ϵ2,μ2ϵ,μ3)ir9(μϵ2,μ2ϵ,μ3)ir9(μϵ2,μ2ϵ,μ3)μtanh(𝚑μ)+r10(μ2ϵ,μ3))\displaystyle G:=G(\mu,\epsilon)=\begin{pmatrix}1+r_{8}(\epsilon^{2},\mu^{2}\epsilon,\mu^{3})&-\mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\\ \mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})&\mu\tanh(\mathtt{h}\mu)+r_{10}(\mu^{2}\epsilon,\mu^{3})\end{pmatrix} (4.51)
F:=F(μ,ϵ)=(𝚏11ϵ+r3(ϵ3,μϵ2,μ2ϵ,μ3)iμϵ𝚌𝚑12+ir4(μϵ2,μ2ϵ,μ3)ir6(μϵ,μ3)r7(μ2ϵ,μ3)).\displaystyle F:=F(\mu,\epsilon)=\begin{pmatrix}\mathtt{f}_{11}\epsilon+r_{3}(\epsilon^{3},\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})&\mathrm{i}\,\mu\epsilon{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}+\mathrm{i}\,r_{4}({\mu\epsilon^{2}},\mu^{2}\epsilon,\mu^{3})\\ \mathrm{i}\,r_{6}(\mu\epsilon,\mu^{3})&r_{7}(\mu^{2}\epsilon,\mu^{3})\end{pmatrix}\,. (4.52)

In order to deduce the expansion (4.11)-(4.12) of the matrices F,GF,G we exploit further information for

μ,0:=𝒥μ,0,μ,0:=[1𝚌𝚑x𝚌𝚑x|D+μ|tanh(𝚑|D+μ|)].\mathscr{L}_{\mu,0}:=\mathcal{J}{\mathcal{B}}_{\mu,0}\,,\quad{\mathcal{B}}_{\mu,0}:=\begin{bmatrix}1&-{\mathtt{c}}_{\mathtt{h}}\partial_{x}\\ {\mathtt{c}}_{\mathtt{h}}\partial_{x}&|D+\mu|\,\tanh\big{(}\mathtt{h}|D+\mu|\big{)}\end{bmatrix}\,. (4.53)

We have

Lemma 4.7.

At ϵ=0\epsilon=0 the matrices are F(μ,0)=0F(\mu,0)=0 and G(μ,0)=(100μtanh(𝚑μ))G(\mu,0)=\begin{pmatrix}1&0\\ 0&\mu\tanh(\mathtt{h}\mu)\end{pmatrix}.

Proof.

By Lemma A.5 and (4.53) we have μ,0f0+(μ,0)=f0+{\mathcal{B}}_{\mu,0}f_{0}^{+}(\mu,0)=f_{0}^{+} and μ,0f0(μ,0)=μtanh(𝚑μ)f0{\mathcal{B}}_{\mu,0}f_{0}^{-}(\mu,0)=\mu\tanh(\mathtt{h}\mu)f_{0}^{-}, for any μ\mu. Then the lemma follows recalling (3.12) and the fact that f1+(μ,0)f_{1}^{+}(\mu,0) and f1(μ,0)f_{1}^{-}(\mu,0) have zero space average by Lemma A.5. ∎

In view of Lemma 4.7 we deduce that the matrices (4.51) and (4.52) have the form (4.11) and (4.12). This completes the proof of Proposition 4.3.

We now show that the constant 𝚎22\mathtt{e}_{22} in (1.3) is positive for any depth 𝚑>0\mathtt{h}>0.

Lemma 4.8.

For any 𝚑>0\mathtt{h}>0 the term 𝚎22\mathtt{e}_{22} in (1.3) is positive, 𝚎220\mathtt{e}_{22}\to 0 as 𝚑0+\mathtt{h}\to 0^{+} and 𝚎221\mathtt{e}_{22}\to 1 as 𝚑+\mathtt{h}\to+\infty. As a consequence for any 𝚑0>0\mathtt{h}_{0}>0 the term 𝚎22\mathtt{e}_{22} is bounded from below uniformly in 𝚑>𝚑0\mathtt{h}>\mathtt{h}_{0}.

Proof.

The quantity z:=𝚌𝚑2=tanh(𝚑)z:={\mathtt{c}}_{\mathtt{h}}^{2}=\tanh(\mathtt{h}) is in (0,1)(0,1) for any 𝚑>0\mathtt{h}>0. Then the quadratic polynomial (0,+)𝚑(1z2)(1+3z2)𝚑2+2z(z21)𝚑+z2(0,+\infty)\ni\mathtt{h}\mapsto(1-z^{2})(1+3z^{2})\mathtt{h}^{2}+2z(z^{2}-1)\mathtt{h}+z^{2} is positive because its discriminant 4z4(1z2)-4z^{4}(1-z^{2}) is negative as 0<z2<10<z^{2}<1. The limits for 𝚑0+\mathtt{h}\to 0^{+} and 𝚑+\mathtt{h}\to+\infty follow by inspection. ∎

5 Block-decoupling and emergence of the Whitham-Benjamin function

In this section we block-decouple the 4×44\times 4 Hamiltonian matrix 𝙻μ,ϵ=𝙹4𝙱μ,ϵ\mathtt{L}_{\mu,\epsilon}=\mathtt{J}_{4}\mathtt{B}_{\mu,\epsilon} obtained in Proposition 4.3.

We first perform a singular symplectic and reversibility-preserving change of coordinates.

Lemma 5.1.

(Singular symplectic rescaling) The conjugation of the Hamiltonian and reversible matrix 𝙻μ,ϵ=𝙹4𝙱μ,ϵ\mathtt{L}_{\mu,\epsilon}=\mathtt{J}_{4}\mathtt{B}_{\mu,\epsilon} obtained in Proposition 4.3 through the symplectic and reversibility-preserving 4×44\times 4-matrix

Y:=(Q00Q)withQ:=(μ1200μ12),μ>0,Y:=\begin{pmatrix}Q&0\\ 0&Q\end{pmatrix}\quad\text{with}\quad Q:=\begin{pmatrix}\mu^{\frac{1}{2}}&0\\ 0&\mu^{-\frac{1}{2}}\end{pmatrix}\,,\ \ \mu>0\,, (5.1)

yields the Hamiltonian and reversible matrix

𝙻μ,ϵ(1):=Y1𝙻μ,ϵY=𝙹4𝙱μ,ϵ(1)=(𝙹2E(1)𝙹2F(1)𝙹2[F(1)]𝙹2G(1))\displaystyle\mathtt{L}_{\mu,\epsilon}^{(1)}:=Y^{-1}\mathtt{L}_{\mu,\epsilon}Y=\mathtt{J}_{4}\mathtt{B}^{(1)}_{\mu,\epsilon}=\begin{pmatrix}\mathtt{J}_{2}E^{(1)}&\mathtt{J}_{2}F^{(1)}\\ \mathtt{J}_{2}[F^{(1)}]^{*}&\mathtt{J}_{2}G^{(1)}\end{pmatrix} (5.2)

where 𝙱μ,ϵ(1)\mathtt{B}_{\mu,\epsilon}^{(1)} is a self-adjoint and reversibility-preserving 4×44\times 4 matrix

𝙱μ,ϵ(1)=(E(1)F(1)[F(1)]G(1)),E(1)=[E(1)],G(1)=[G(1)],\mathtt{B}_{\mu,\epsilon}^{(1)}=\begin{pmatrix}E^{(1)}&F^{(1)}\\ [F^{(1)}]^{*}&G^{(1)}\end{pmatrix},\quad E^{(1)}=[E^{(1)}]^{*}\,,\ G^{(1)}=[G^{(1)}]^{*}\,, (5.3)

where the 2×22\times 2 reversibility-preserving matrices E(1)E^{(1)}, G(1)G^{(1)} and F(1)F^{(1)} extend analytically at μ=0\mu=0 with the following expansion

E(1)=(𝚎11μϵ2(1+r1(ϵ,μϵ))𝚎22μ38(1+r1′′(ϵ,μ))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))𝚎22μ8(1+r5(ϵ,μ))),\displaystyle E^{(1)}=\begin{pmatrix}\mathtt{e}_{11}\mu\epsilon^{2}(1+r_{1}^{\prime}(\epsilon,\mu\epsilon))-\mathtt{e}_{22}\frac{\mu^{3}}{8}(1+r_{1}^{\prime\prime}(\epsilon,\mu))&\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}\\ -\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}&-\mathtt{e}_{22}\frac{\mu}{8}(1+r_{5}(\epsilon,\mu))\end{pmatrix}\,, (5.4)
G(1)=(μ+r8(μϵ2,μ3ϵ)ir9(μϵ2,μ2ϵ)ir9(μϵ2,μ2ϵ)tanh(𝚑μ)+r10(μϵ)),\displaystyle G^{(1)}=\begin{pmatrix}\mu+r_{8}(\mu\epsilon^{2},\mu^{3}\epsilon)&-\mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)\\ \mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)&\tanh(\mathtt{h}\mu)+r_{10}(\mu\epsilon)\end{pmatrix}\,, (5.5)
F(1)=(𝚏11μϵ+r3(μϵ3,μ2ϵ2,μ3ϵ)iμϵ𝚌𝚑12+ir4(μϵ2,μ2ϵ)ir6(μϵ)r7(μϵ))\displaystyle F^{(1)}=\begin{pmatrix}\mathtt{f}_{11}\mu\epsilon+r_{3}(\mu\epsilon^{3},\mu^{2}\epsilon^{2},\mu^{3}\epsilon)&\mathrm{i}\,\mu\epsilon{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}+\mathrm{i}\,r_{4}(\mu\epsilon^{2},\mu^{2}\epsilon)\\ \mathrm{i}\,r_{6}(\mu\epsilon)&r_{7}(\mu\epsilon)\end{pmatrix} (5.6)

where 𝚎11,𝚎12,𝚎22,𝚏11\mathtt{e}_{11},\mathtt{e}_{12},\mathtt{e}_{22},\mathtt{f}_{11} are defined in (4.13), (1.2), (1.3).

Remark 5.2.

The matrix 𝙻μ,ϵ(1)\mathtt{L}_{\mu,\epsilon}^{(1)}, a priori defined only for μ0\mu\neq 0, extends analytically to the zero matrix at μ=0\mu=0. For μ0\mu\neq 0 the spectrum of 𝙻μ,ϵ(1)\mathtt{L}_{\mu,\epsilon}^{(1)} coincides with the spectrum of 𝙻μ,ϵ\mathtt{L}_{\mu,\epsilon}.

Proof.

The matrix YY is symplectic, i.e. (3.15) holds, and since μ\mu is real, it is reversibility preserving, i.e. satisfies (3.13). By (3.16),

𝙱μ,ϵ(1)=Y𝙱μ,ϵY=(E(1)F(1)[F(1)]G(1)),\mathtt{B}_{\mu,\epsilon}^{(1)}=Y^{*}\mathtt{B}_{\mu,\epsilon}Y=\begin{pmatrix}E^{(1)}&F^{(1)}\\ [F^{(1)}]^{*}&G^{(1)}\end{pmatrix},

with, QQ being self-adjoint, E(1)=QEQ=[E(1)]E^{(1)}=QEQ=[E^{(1)}]^{*}, G(1)=QGQ=[G(1)]G^{(1)}=QGQ=[G^{(1)}]^{*} and F(1)=QFQF^{(1)}=QFQ. In view of (4.10)-(4.12), we obtain (5.4)-(5.6). ∎

5.1 Non-perturbative step of block-decoupling

We first verify that the quantity D𝚑:=𝚑14𝚎122D_{\mathtt{h}}:=\mathtt{h}-\tfrac{1}{4}\mathtt{e}_{12}^{2} is nonzero for any 𝚑>0\mathtt{h}>0. In view of the comment 3 after Theorem 1.1, we have that D𝚑=𝚑cg2D_{\mathtt{h}}=\mathtt{h}-c_{g}^{2}. The non-degeneracy property D𝚑0D_{\mathtt{h}}\neq 0 corresponds to that in Bridges-Mielke [9, p.183] and [41, p.409].

Lemma 5.3.

For any 𝚑>0\mathtt{h}>0 it results

𝙳𝚑:=𝚑14𝚎122>0,andlim𝚑0+𝙳𝚑=0.\mathtt{D}_{\mathtt{h}}:=\mathtt{h}-\tfrac{1}{4}\mathtt{e}_{12}^{2}>0\,,\quad\text{and}\quad\lim_{\mathtt{h}\to 0^{+}}\mathtt{D}_{\mathtt{h}}=0\,. (5.7)
Proof.

We write 𝙳𝚑=(𝚑+12𝚎12)(𝚑12𝚎12)\mathtt{D}_{\mathtt{h}}=(\sqrt{\mathtt{h}}+\frac{1}{2}\mathtt{e}_{12})(\sqrt{\mathtt{h}}-\frac{1}{2}\mathtt{e}_{12}) whose first factor is positive for 𝚑>0\mathtt{h}>0. We claim that also the second factor is positive. In view of (1.2) it is equal to 12𝚌𝚑1f(𝚑)\tfrac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-1}f(\mathtt{h}) with

f(𝚑):=(𝚑tanh(𝚑)𝚑+tanh(𝚑))(𝚑tanh(𝚑)+𝚑tanh(𝚑))=:q(𝚑)p(𝚑).\displaystyle f(\mathtt{h}):=\big{(}\sqrt{\mathtt{h}}\tanh(\mathtt{h})-\sqrt{\mathtt{h}}+\sqrt{\tanh(\mathtt{h})}\big{)}\big{(}\sqrt{\mathtt{h}}\tanh(\mathtt{h})+\sqrt{\mathtt{h}}-\sqrt{\tanh(\mathtt{h})}\big{)}=:q(\mathtt{h})p(\mathtt{h})\,.

The function p(𝚑)p(\mathtt{h}) is positive since 𝚑>tanh(𝚑)\mathtt{h}>\tanh(\mathtt{h}) for any 𝚑>0\mathtt{h}>0. We claim that also the function q(𝚑)q(\mathtt{h}) is positive. Indeed its derivative

q(𝚑)=1tanh(𝚑)2𝚑tanh(𝚑)(tanh(𝚑)+𝚑+𝚑tanh(𝚑))+𝚑(1tanh2(𝚑))>0q^{\prime}(\mathtt{h})=\frac{1-\tanh(\mathtt{h})}{2\sqrt{\mathtt{h}}\sqrt{\tanh(\mathtt{h})}}\Big{(}-\sqrt{\tanh(\mathtt{h})}+\sqrt{\mathtt{h}}+\sqrt{\mathtt{h}}\,{\tanh(\mathtt{h})}\Big{)}+\sqrt{\mathtt{h}}\big{(}1-\tanh^{2}(\mathtt{h})\big{)}>0

for any 𝚑>0\mathtt{h}>0. Since q(0)=0q(0)=0 we deduce that q(𝚑)>0q(\mathtt{h})>0 for any 𝚑>0\mathtt{h}>0. This proves the lemma. ∎

We now state the main result of this section.

Lemma 5.4.

(Step of block-decoupling) There exists a 2×22\times 2 reversibility-preserving matrix XX, analytic in (μ,ϵ)(\mu,\epsilon), of the form

X\displaystyle X :=(x11ix12ix21x22)withxij,i,j=1,2,\displaystyle:=\begin{pmatrix}x_{11}&\mathrm{i}\,x_{12}\\ \mathrm{i}\,x_{21}&x_{22}\end{pmatrix}\qquad\qquad\qquad\qquad\text{with}\quad x_{ij}\in\mathbb{R}\,,\ i,j=1,2\,, (5.8)
=(r11(ϵ)ir12(ϵ)i12𝙳𝚑1(𝚎12𝚏11+2𝚌𝚑12)ϵ+ir21(ϵ2,μϵ)12𝙳𝚑1(𝚌𝚑12𝚎12+2𝚑𝚏11)ϵ+r22(ϵ2,μϵ)),\displaystyle=\begin{pmatrix}r_{11}(\epsilon)&\mathrm{i}\,\,r_{12}(\epsilon)\\ -\mathrm{i}\,\frac{1}{2}\mathtt{D}_{\mathtt{h}}^{-1}(\mathtt{e}_{12}\mathtt{f}_{11}+2{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}})\epsilon+\mathrm{i}\,r_{21}(\epsilon^{2},\mu\epsilon)&\frac{1}{2}\mathtt{D}_{\mathtt{h}}^{-1}({\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\mathtt{e}_{12}+2\mathtt{h}\mathtt{f}_{11})\epsilon+r_{22}(\epsilon^{2},\mu\epsilon)\end{pmatrix}\,,

where 𝚎12\mathtt{e}_{12}, 𝚏11\mathtt{f}_{11} are defined in (1.2), (4.13) and 𝙳𝚑\mathtt{D}_{\mathtt{h}} is the positive constant in (5.7), such that the following holds true. By conjugating the Hamiltonian and reversible matrix 𝙻μ,ϵ(1)\mathtt{L}_{\mu,\epsilon}^{(1)}, defined in (5.2), with the symplectic and reversibility-preserving 4×44\times 4 matrix

exp(S(1)), where S(1):=𝙹4(0ΣΣ0),Σ:=𝙹2X,\exp\left(S^{(1)}\right)\,,\quad\text{ where }\qquad S^{(1)}:=\mathtt{J}_{4}\begin{pmatrix}0&\Sigma\\ \Sigma^{*}&0\end{pmatrix}\,,\qquad\Sigma:=\mathtt{J}_{2}X\,, (5.9)

we get the Hamiltonian and reversible matrix

𝙻μ,ϵ(2):=exp(S(1))𝙻μ,ϵ(1)exp(S(1))=𝙹4𝙱μ,ϵ(2)=(𝙹2E(2)𝙹2F(2)𝙹2[F(2)]𝙹2G(2)),\mathtt{L}_{\mu,\epsilon}^{(2)}:=\exp\left(S^{(1)}\right)\mathtt{L}_{\mu,\epsilon}^{(1)}\exp\left(-S^{(1)}\right)=\mathtt{J}_{4}\mathtt{B}_{\mu,\epsilon}^{(2)}=\begin{pmatrix}\mathtt{J}_{2}E^{(2)}&\mathtt{J}_{2}F^{(2)}\\ \mathtt{J}_{2}[F^{(2)}]^{*}&\mathtt{J}_{2}G^{(2)}\end{pmatrix}\,, (5.10)

where the reversibility-preserving 2×22\times 2 self-adjoint matrix [E(2)]=E(2)[E^{(2)}]^{*}=E^{(2)} has the form

E(2)=(μϵ2𝚎WB+r1(μϵ3,μ2ϵ2)𝚎22μ38(1+r1′′(ϵ,μ))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))i(12𝚎12μ+r2(μϵ2,μ2ϵ,μ3))𝚎22μ8(1+r5(ϵ,μ))),\displaystyle E^{(2)}=\begin{pmatrix}\mu\epsilon^{2}\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}+r_{1}^{\prime}(\mu\epsilon^{3},\mu^{2}\epsilon^{2})-\mathtt{e}_{22}\frac{\mu^{3}}{8}(1+r_{1}^{\prime\prime}(\epsilon,\mu))&\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}\\ -\mathrm{i}\,\big{(}\frac{1}{2}\mathtt{e}_{12}\mu+r_{2}(\mu\epsilon^{2},\mu^{2}\epsilon,\mu^{3})\big{)}&-\mathtt{e}_{22}\frac{\mu}{8}(1+r_{5}(\epsilon,\mu))\end{pmatrix}\,, (5.11)

where

𝚎WB=𝚎11𝙳𝚑1(𝚌𝚑1+𝚑𝚏112+𝚎12𝚏11𝚌𝚑12)\displaystyle\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}=\mathtt{e}_{11}-\mathtt{D}_{\mathtt{h}}^{-1}\big{(}{\mathtt{c}}_{\mathtt{h}}^{-1}+\mathtt{h}\mathtt{f}_{11}^{2}+\mathtt{e}_{12}\mathtt{f}_{11}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\big{)} (5.12)

(with constants 𝚎11\mathtt{e}_{11}, 𝙳𝚑\mathtt{D}_{\mathtt{h}}, 𝚏11\mathtt{f}_{11}, 𝚎12\mathtt{e}_{12}, defined in (4.13), (5.7), (1.2)), is the Whitham-Benjamin function defined in (1.1), the reversibility-preserving 2×22\times 2 self-adjoint matrix [G(2)]=G(2)[G^{(2)}]^{*}=G^{(2)} has the form

G(2)=(μ+r8(μϵ2,μ3ϵ)ir9(μϵ2,μ2ϵ)ir9(μϵ2,μ2ϵ)tanh(𝚑μ)+r10(μϵ)),G^{(2)}=\begin{pmatrix}\mu+r_{8}(\mu\epsilon^{2},\mu^{3}\epsilon)&-\mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)\\ \mathrm{i}\,r_{9}(\mu\epsilon^{2},\mu^{2}\epsilon)&\tanh(\mathtt{h}\mu)+r_{10}(\mu\epsilon)\end{pmatrix}\,, (5.13)

and

F(2)=(r3(μϵ3)ir4(μϵ3)ir6(μϵ3)r7(μϵ3)).F^{(2)}=\begin{pmatrix}r_{3}(\mu\epsilon^{3})&\mathrm{i}\,r_{4}(\mu\epsilon^{3})\\ \mathrm{i}\,r_{6}(\mu\epsilon^{3})&r_{7}(\mu\epsilon^{3})\end{pmatrix}\,. (5.14)

The rest of the section is devoted to the proof of Lemma 5.4. For simplicity let S=S(1)S=S^{(1)}.

The matrix exp(S)\text{exp}(S) is symplectic and reversibility-preserving because the matrix SS in (5.9) is Hamiltonian and reversibility-preserving, cfr. Lemma 3.8 in [6]. Note that SS is reversibility preserving since XX has the form (5.8).

We now expand in Lie series the Hamiltonian and reversible matrix 𝙻μ,ϵ(2)=exp(S)𝙻μ,ϵ(1)exp(S)\mathtt{L}_{\mu,\epsilon}^{(2)}=\exp(S)\mathtt{L}_{\mu,\epsilon}^{(1)}\exp(-S).

We split 𝙻μ,ϵ(1)\mathtt{L}_{\mu,\epsilon}^{(1)} into its 2×22\times 2-diagonal and off-diagonal Hamiltonian and reversible matrices

𝙻μ,ϵ(1)=D(1)+R(1),\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\mathtt{L}_{\mu,\epsilon}^{(1)}=D^{(1)}+R^{(1)}\,, (5.15)
D(1):=(D100D0):=(𝙹2E(1)00𝙹2G(1)),R(1):=(0𝙹2F(1)𝙹2[F(1)]0),\displaystyle D^{(1)}:=\begin{pmatrix}D_{1}&0\\ 0&D_{0}\end{pmatrix}:=\begin{pmatrix}\mathtt{J}_{2}E^{(1)}&0\\ 0&\mathtt{J}_{2}G^{(1)}\end{pmatrix},\quad R^{(1)}:=\begin{pmatrix}0&\mathtt{J}_{2}F^{(1)}\\ \mathtt{J}_{2}[F^{(1)}]^{*}&0\end{pmatrix},

and we perform the Lie expansion

𝙻μ,ϵ(2)=exp(S)𝙻μ,ϵ(1)exp(S)=D(1)+[S,D(1)]+12[S,[S,D(1)]]+R(1)+[S,R(1)]\displaystyle\mathtt{L}_{\mu,\epsilon}^{(2)}=\exp(S)\mathtt{L}_{\mu,\epsilon}^{(1)}\exp(-S)=D^{(1)}+\left[S\,,\,D^{(1)}\right]+\frac{1}{2}[S,[S,D^{(1)}]]+R^{(1)}+[S,R^{(1)}] (5.16)
+1201(1τ)2exp(τS)adS3(D(1))exp(τS)dτ+01(1τ)exp(τS)adS2(R(1))exp(τS)dτ\displaystyle+\frac{1}{2}\int_{0}^{1}(1-\tau)^{2}\exp(\tau S)\text{ad}_{S}^{3}(D^{(1)})\exp(-\tau S)\,\mathrm{d}\tau+\int_{0}^{1}(1-\tau)\,\exp(\tau S)\,\text{ad}_{S}^{2}(R^{(1)})\,\exp(-\tau S)\,\mathrm{d}\tau

where adA(B):=[A,B]:=ABBA\text{ad}_{A}(B):=[A,B]:=AB-BA denotes the commutator between the linear operators A,BA,B.

We look for a 4×44\times 4 matrix SS as in (5.9) that solves the homological equation R(1)+[S,D(1)]=0R^{(1)}+\left[S\,,\,D^{(1)}\right]=0, which, recalling (5.15), reads

(0𝙹2F(1)+𝙹2ΣD0D1𝙹2Σ𝙹2[F(1)]+𝙹2ΣD1D0𝙹2Σ0)=0.\begin{pmatrix}0&\mathtt{J}_{2}F^{(1)}+\mathtt{J}_{2}\Sigma D_{0}-D_{1}\mathtt{J}_{2}\Sigma\\ \mathtt{J}_{2}{[F^{(1)}]}^{*}+\mathtt{J}_{2}\Sigma^{*}D_{1}-D_{0}\mathtt{J}_{2}\Sigma^{*}&0\end{pmatrix}=0\,. (5.17)

Note that the equation 𝙹2F(1)+𝙹2ΣD0D1𝙹2Σ=0\mathtt{J}_{2}F^{(1)}+\mathtt{J}_{2}\Sigma D_{0}-D_{1}\mathtt{J}_{2}\Sigma=0 implies also 𝙹2[F(1)]+𝙹2ΣD1D0𝙹2Σ=0\mathtt{J}_{2}{[F^{(1)}]}^{*}+\mathtt{J}_{2}\Sigma^{*}D_{1}-D_{0}\mathtt{J}_{2}\Sigma^{*}=0 and viceversa. Thus, writing Σ=𝙹2X\Sigma=\mathtt{J}_{2}X, namely X=𝙹2ΣX=-\mathtt{J}_{2}\Sigma, the equation (5.17) amounts to solve the “Sylvester” equation

D1XXD0=𝙹2F(1).D_{1}X-XD_{0}=-\mathtt{J}_{2}F^{(1)}\,. (5.18)

We write the matrices E(1),F(1),G(1)E^{(1)},F^{(1)},G^{(1)} in (5.2) as

E(1)=(E11(1)iE12(1)iE12(1)E22(1)),F(1)=(F11(1)iF12(1)iF21(1)F22(1)),G(1)=(G11(1)iG12(1)iG12(1)G22(1))E^{(1)}=\begin{pmatrix}E_{11}^{(1)}&\mathrm{i}\,E_{12}^{(1)}\\ -\mathrm{i}\,E_{12}^{(1)}&E_{22}^{(1)}\end{pmatrix}\,,\quad F^{(1)}=\begin{pmatrix}F_{11}^{(1)}&\mathrm{i}\,F_{12}^{(1)}\\ \mathrm{i}\,F_{21}^{(1)}&F_{22}^{(1)}\end{pmatrix}\,,\quad G^{(1)}=\begin{pmatrix}G_{11}^{(1)}&\mathrm{i}\,G_{12}^{(1)}\\ -\mathrm{i}\,G_{12}^{(1)}&G_{22}^{(1)}\end{pmatrix} (5.19)

where the real numbers Eij(1),Fij(1),Gij(1)E_{ij}^{(1)},F_{ij}^{(1)},G_{ij}^{(1)}, i,j=1,2i,j=1,2, have the expansion in (5.4), (5.5), (5.6). Thus, by (5.15), (5.8) and (5.19), the equation (5.18) amounts to solve the 4×44\times 4 real linear system

(G12(1)E12(1)G11(1)E22(1)0G22(1)G12(1)E12(1)0E22(1)E11(1)0G12(1)E12(1)G11(1)0E11(1)G22(1)G12(1)E12(1))=:𝒜(x11x12x21x22)=:x=(F21(1)F22(1)F11(1)F12(1))=:f.\displaystyle\underbrace{\begin{pmatrix}G_{12}^{(1)}-E_{12}^{(1)}&G_{11}^{(1)}&E_{22}^{(1)}&0\\ G_{22}^{(1)}&G_{12}^{(1)}-E_{12}^{(1)}&0&-E_{22}^{(1)}\\ E_{11}^{(1)}&0&G_{12}^{(1)}-E_{12}^{(1)}&-G_{11}^{(1)}\\ 0&-E_{11}^{(1)}&-G_{22}^{(1)}&G_{12}^{(1)}-E_{12}^{(1)}\end{pmatrix}}_{=:{\mathcal{A}}}\underbrace{\begin{pmatrix}x_{11}\\ x_{12}\\ x_{21}\\ x_{22}\end{pmatrix}}_{=:\vec{x}}=\underbrace{\begin{pmatrix}-F_{21}^{(1)}\\ F_{22}^{(1)}\\ -F_{11}^{(1)}\\ F_{12}^{(1)}\end{pmatrix}}_{=:\vec{f}}. (5.20)

We solve this system using the following result, verified by a direct calculus.

Lemma 5.5.

The determinant of the matrix

A:=(abc0da0ce0ab0eda)A:=\begin{pmatrix}a&b&c&0\\ d&a&0&-c\\ e&0&a&-b\\ 0&-e&-d&a\end{pmatrix} (5.21)

where a,b,c,d,ea,b,c,d,e are real numbers, is

detA=a42a2(bd+ce)+(bdce)2=(bda2)22ce(a2+bd12ce).\det A=a^{4}-2a^{2}(bd+ce)+(bd-ce)^{2}=(bd-a^{2})^{2}-2ce\big{(}a^{2}+bd-\frac{1}{2}ce\big{)}\,. (5.22)

If detA0\det A\neq 0 then AA is invertible and

A1=1detA(a(a2bdce)b(a2+bdce)c(a2+bdce)2abcd(a2+bdce)a(a2bdce)2acdc(a2bd+ce)e(a2+bdce)2abea(a2bdce)b(a2bd+ce)2adee(a2bd+ce)d(a2bd+ce)a(a2bdce)).\displaystyle A^{-1}=\footnotesize{\frac{1}{\det A}\left(\begin{array}[]{cccc}\!a\left(a^{2}-bd-ce\right)&\!b\left(-a^{2}+bd-ce\right)&-c\left(a^{2}+bd-ce\right)&\!-2abc\\ \!d\left(-a^{2}+bd-ce\right)&\!a\left(a^{2}-bd-ce\right)&2acd&\!-c\left(-a^{2}-bd+ce\right)\\ \!-e\left(a^{2}+bd-ce\right)&\!2abe&a\left(a^{2}-bd-ce\right)&\!b\left(a^{2}-bd+ce\right)\\ \!-2ade&\!-e\left(-a^{2}-bd+ce\right)&d\left(a^{2}-bd+ce\right)&\!a\left(a^{2}-bd-ce\right)\end{array}\right)}\,. (5.27)

The Sylvester matrix 𝒜\mathcal{A} in (5.20) has the form (5.21) where, by (5.4)-(5.6) and since tanh(𝚑μ)=𝚑μ+r(μ3)\tanh(\mathtt{h}\mu)=\mathtt{h}\mu+r(\mu^{3}),

a=G12(1)E12(1)=𝚎12μ2(1+r(ϵ2,μϵ,μ2)),b=G11(1)=μ+r8(μϵ2,μ3ϵ),\displaystyle a=G_{12}^{(1)}-E_{12}^{(1)}=-\mathtt{e}_{12}\frac{\mu}{2}\big{(}1+r(\epsilon^{2},\mu\epsilon,\mu^{2})\big{)}\,,\ b=G_{11}^{(1)}=\mu+r_{8}(\mu\epsilon^{2},\mu^{3}\epsilon)\,, (5.28)
c=E22(1)=𝚎22μ8(1+r5(ϵ,μ)),d=G22(1)=μ𝚑+r(μϵ,μ3),e=E11(1)=r(μϵ2,μ3),\displaystyle c=E_{22}^{(1)}=-\mathtt{e}_{22}\frac{\mu}{8}(1+r_{5}(\epsilon,\mu))\,,\ d=G_{22}^{(1)}=\mu\mathtt{h}+r(\mu\epsilon,\mu^{3})\,,\ e=E_{11}^{(1)}=r(\mu\epsilon^{2},\mu^{3})\,,

where 𝚎12\mathtt{e}_{12} and 𝚎22\mathtt{e}_{22}, defined respectively in (1.2), (1.3), are positive for any 𝚑>0\mathtt{h}>0.

By (5.22), the determinant of the matrix 𝒜{\mathcal{A}} is

det𝒜=(bda2)2+r(μ4ϵ2,μ6)=μ4𝙳𝚑2(1+r(ϵ,μ2))\displaystyle\det{\mathcal{A}}=(bd-a^{2})^{2}+r(\mu^{4}\epsilon^{2},\mu^{6})=\mu^{4}\mathtt{D}_{\mathtt{h}}^{2}(1+r(\epsilon,\mu^{2}))\, (5.29)

where 𝙳𝚑\mathtt{D}_{\mathtt{h}} is defined in (5.7). By (5.27), (5.28), (5.29) and, since 𝙳𝚑=𝚑14𝚎122\mathtt{D}_{\mathtt{h}}=\mathtt{h}-\frac{1}{4}\mathtt{e}_{12}^{2}, we obtain

𝒜1=(1+r(ϵ,μ))1μ𝙳𝚑2(12𝚎12𝙳𝚑𝙳𝚑132𝚎22(𝚎122+4𝚑)18𝚎12𝚎22𝚑𝙳𝚑12𝚎12𝙳𝚑18𝚎12𝚎22𝚑132𝚎22(𝚎122+4𝚑)r(ϵ2,μ2)r(ϵ2,μ2)12𝚎12𝙳𝚑𝙳𝚑r(ϵ2,μ2)r(ϵ2,μ2)𝚑𝙳𝚑12𝚎12𝙳𝚑).\displaystyle{\mathcal{A}}^{-1}=(1+r(\epsilon,\mu))\displaystyle{\frac{1}{\mu\mathtt{D}^{2}_{\mathtt{h}}}}\,\begin{pmatrix}\frac{1}{2}{\mathtt{e}_{12}}\mathtt{D}_{\mathtt{h}}&\mathtt{D}_{\mathtt{h}}&\frac{1}{32}\mathtt{e}_{22}(\mathtt{e}_{12}^{2}+4\mathtt{h})&-\frac{1}{8}{\mathtt{e}_{12}}\,\mathtt{e}_{22}\\ \mathtt{h}\mathtt{D}_{\mathtt{h}}&\frac{1}{2}{\mathtt{e}_{12}}\mathtt{D}_{\mathtt{h}}&\frac{1}{8}\mathtt{e}_{12}\mathtt{e}_{22}\mathtt{h}&-\frac{1}{32}\mathtt{e}_{22}\,(\mathtt{e}_{12}^{2}+4\mathtt{h})\\ r(\epsilon^{2},\mu^{2})&r(\epsilon^{2},\mu^{2})&\frac{1}{2}{\mathtt{e}_{12}}\mathtt{D}_{\mathtt{h}}&-{\mathtt{D}_{\mathtt{h}}}\\ r(\epsilon^{2},\mu^{2})&r(\epsilon^{2},\mu^{2})&-\mathtt{h}\mathtt{D}_{\mathtt{h}}&\frac{1}{2}{\mathtt{e}_{12}}\mathtt{D}_{\mathtt{h}}\end{pmatrix}\,. (5.30)

Therefore, for any μ0\mu\neq 0, there exists a unique solution x=𝒜1f\vec{x}={\mathcal{A}}^{-1}\vec{f} of the linear system (5.20), namely a unique matrix XX which solves the Sylvester equation (5.18).

Lemma 5.6.

The matrix solution XX of the Sylvester equation (5.18) is analytic in (μ,ϵ)(\mu,\epsilon), and admits an expansion as in (5.8).

Proof.

By (5.20), (5.30), (5.19), (5.6) we obtain, for any μ0\mu\neq 0

(x11x12x21x22)=1𝙳𝚑2(12𝚎12𝙳𝚑𝙳𝚑132𝚎22(𝚎122+4𝚑)18𝚎12𝚎22𝚑𝙳𝚑12𝚎12𝙳𝚑18𝚎12𝚎22𝚑132𝚎22(𝚎122+4𝚑)r(ϵ2,μ2)r(ϵ2,μ2)12𝚎12𝙳𝚑𝙳𝚑r(ϵ2,μ2)r(ϵ2,μ2)𝚑𝙳𝚑12𝚎12𝙳𝚑)(r(ϵ)r(ϵ)𝚏11ϵ+r(ϵ3,μϵ2,μ2ϵ)𝚌𝚑12ϵ+r(ϵ2,μϵ))(1+r(ϵ,μ)),\displaystyle\footnotesize\begin{pmatrix}x_{11}\\ x_{12}\\ x_{21}\\ x_{22}\end{pmatrix}\footnotesize=\frac{1}{\mathtt{D}^{2}_{\mathtt{h}}}\begin{pmatrix}\frac{1}{2}{\mathtt{e}_{12}}\mathtt{D}_{\mathtt{h}}&\mathtt{D}_{\mathtt{h}}&\frac{1}{32}\mathtt{e}_{22}(\mathtt{e}_{12}^{2}+4\mathtt{h})&-\frac{1}{8}{\mathtt{e}_{12}}\,\mathtt{e}_{22}\\ \mathtt{h}\mathtt{D}_{\mathtt{h}}&\frac{1}{2}{\mathtt{e}_{12}}\mathtt{D}_{\mathtt{h}}&\frac{1}{8}\mathtt{e}_{12}\mathtt{e}_{22}\mathtt{h}&-\frac{1}{32}\mathtt{e}_{22}\,(\mathtt{e}_{12}^{2}+4\mathtt{h})\\ r(\epsilon^{2},\mu^{2})\quad&r(\epsilon^{2},\mu^{2})&\frac{1}{2}{\mathtt{e}_{12}}\mathtt{D}_{\mathtt{h}}&-{\mathtt{D}_{\mathtt{h}}}\\ r(\epsilon^{2},\mu^{2})\quad&r(\epsilon^{2},\mu^{2})&-\mathtt{h}\mathtt{D}_{\mathtt{h}}&\frac{1}{2}\mathtt{e}_{12}\mathtt{D}_{\mathtt{h}}\end{pmatrix}\begin{pmatrix}r(\epsilon)\\ r(\epsilon)\\ -\mathtt{f}_{11}\epsilon+r(\epsilon^{3},\mu\epsilon^{2},\mu^{2}\epsilon)\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\epsilon+r(\epsilon^{2},\mu\epsilon)\end{pmatrix}(1+r(\epsilon,\mu))\,,

which proves (5.8). In particular each xijx_{ij} admits an analytic extension at μ=0\mu=0. Note that, for μ=0\mu=0, one has E(2)=G(2)=F(2)=0E^{(2)}=G^{(2)}=F^{(2)}=0 and the Sylvester equation reduces to tautology. ∎

Since the matrix SS solves the homological equation [S,D(1)]+R(1)=0\left[S\,,\,D^{(1)}\right]+R^{(1)}=0, identity (5.16) simplifies to

𝙻μ,ϵ(2)=D(1)+12[S,R(1)]+1201(1τ2)exp(τS)adS2(R(1))exp(τS)dτ.\mathtt{L}_{\mu,\epsilon}^{(2)}=D^{(1)}+\frac{1}{2}\left[S\,,\,R^{(1)}\right]+\frac{1}{2}\int_{0}^{1}(1-\tau^{2})\,\exp(\tau S)\,\text{ad}_{S}^{2}(R^{(1)})\,\exp(-\tau S)\mathrm{d}\tau\,. (5.31)

The matrix 12[S,R(1)]\frac{1}{2}\left[S\,,\,R^{(1)}\right] is, by (5.9), (5.15), the block-diagonal Hamiltonian and reversible matrix

12[S,R(1)]\displaystyle\frac{1}{2}\left[S\,,\,R^{(1)}\right] (5.32)
=(12𝙹2(Σ𝙹2[F(1)]F(1)𝙹2Σ)0012𝙹2(Σ𝙹2F(1)[F(1)]𝙹2Σ))=(𝙹2E~00𝙹2G~),\displaystyle=\begin{pmatrix}\frac{1}{2}\mathtt{J}_{2}(\Sigma\mathtt{J}_{2}[F^{(1)}]^{*}-F^{(1)}\mathtt{J}_{2}\Sigma^{*})&0\\ 0&\frac{1}{2}\mathtt{J}_{2}(\Sigma^{*}\mathtt{J}_{2}F^{(1)}-[F^{(1)}]^{*}\mathtt{J}_{2}\Sigma)\end{pmatrix}=\begin{pmatrix}\mathtt{J}_{2}\tilde{E}&0\\ 0&\mathtt{J}_{2}\tilde{G}\end{pmatrix},

where, since Σ=𝙹2X\Sigma=\mathtt{J}_{2}X,

E~:=Sym(𝙹2X𝙹2[F(1)]),G~:=Sym(XF(1)),\tilde{E}:=\text{{Sym}}\big{(}\mathtt{J}_{2}X\mathtt{J}_{2}[F^{(1)}]^{*}\big{)}\,,\qquad\tilde{G}:=\text{{Sym}}\big{(}X^{*}F^{(1)}\big{)}\,, (5.33)

denoting Sym(A):=12(A+A)\text{{Sym}}(A):=\frac{1}{2}(A+A^{*}).

Lemma 5.7.

The self-adjoint and reversibility-preserving matrices E~,G~\tilde{E},\ \tilde{G} in (5.33) have the form

E~=(𝚎~11μϵ2+r~1(μϵ3,μ2ϵ2)ir~2(μϵ2)ir~2(μϵ2)r~5(μϵ2)),G~=(r~8(μϵ2)ir~9(μϵ2)ir~9(μϵ2)r~10(μϵ2)),\displaystyle\tilde{E}=\begin{pmatrix}\tilde{\mathtt{e}}_{11}\mu\epsilon^{2}+\tilde{r}_{1}(\mu\epsilon^{3},\mu^{2}\epsilon^{2})&\mathrm{i}\,\tilde{r}_{2}(\mu\epsilon^{2})\\ -\mathrm{i}\,\tilde{r}_{2}(\mu\epsilon^{2})&\tilde{r}_{5}(\mu\epsilon^{2})\end{pmatrix}\,,\quad\tilde{G}=\begin{pmatrix}\tilde{r}_{8}(\mu\epsilon^{2})&\mathrm{i}\,\tilde{r}_{9}(\mu\epsilon^{2})\\ -\mathrm{i}\,\tilde{r}_{9}(\mu\epsilon^{2})&\tilde{r}_{10}(\mu\epsilon^{2})\end{pmatrix}\,, (5.34)
𝚎~11:=𝙳𝚑1(𝚌𝚑1+𝚑𝚏112+𝚎12𝚏11𝚌𝚑12).\displaystyle\tilde{\mathtt{e}}_{11}:=-\mathtt{D}_{\mathtt{h}}^{-1}\big{(}{\mathtt{c}}_{\mathtt{h}}^{-1}+\mathtt{h}\mathtt{f}_{11}^{2}+\mathtt{e}_{12}\mathtt{f}_{11}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}\big{)}\,.
Proof.

For simplicity we set F=F(1)F=F^{(1)}. By (5.8), (5.6), one has

𝙹2X𝙹2F\displaystyle\mathtt{J}_{2}X\mathtt{J}_{2}F^{*} =(x21F12x22F11i(x21F22+x22F21)i(x11F12+x12F11)x11F22+x12F21)=(𝚎~11μϵ2+r(μϵ3,μ2ϵ2)ir(μϵ2)ir(μϵ2)r(μϵ2))\displaystyle=\begin{pmatrix}x_{21}F_{12}-x_{22}F_{11}&\mathrm{i}\,(x_{21}F_{22}+x_{22}F_{21})\\ \mathrm{i}\,(x_{11}F_{12}+x_{12}F_{11})&-x_{11}F_{22}+x_{12}F_{21}\end{pmatrix}=\begin{pmatrix}\tilde{\mathtt{e}}_{11}\mu\epsilon^{2}+r(\mu\epsilon^{3},\mu^{2}\epsilon^{2})&\mathrm{i}\,r(\mu\epsilon^{2})\\ \mathrm{i}\,r(\mu\epsilon^{2})&r(\mu\epsilon^{2})\end{pmatrix}

with 𝚎~11\tilde{\mathtt{e}}_{11} defined in (5.34). The expansion of E~\tilde{E} in (5.34) follows in view of (5.33). Since X=𝒪(ϵ)X=\mathcal{O}(\epsilon) by (5.8) and F=O(μϵ)F=O(\mu\epsilon) by (5.6) we deduce that XF=𝒪(μϵ2)X^{*}F=\mathcal{O}(\mu\epsilon^{2}) and the expansion of G~\tilde{G} in (5.34) follows. ∎

Note that the term 𝚎~11μϵ2\tilde{\mathtt{e}}_{11}\mu\epsilon^{2} in the matrix E~\tilde{E} in (5.33)-(5.34), has the same order of the (1,1)(1,1)-entry of E(1)E^{(1)} in (5.4), thus will contribute to the Whitham-Benjamin function 𝚎WB\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}} in the (1,1)(1,1)-entry of E(2)E^{(2)} in (5.11). Finally we show that the last term in (5.31) is small.

Lemma 5.8.

The 4×44\times 4 Hamiltonian and reversibility matrix

1201(1τ2)exp(τS)adS2(R(1))exp(τS)dτ=(𝙹2E^𝙹2F(2)𝙹2[F(2)]𝙹2G^)\frac{1}{2}\int_{0}^{1}(1-\tau^{2})\,\exp(\tau S)\,\textup{ad}_{S}^{2}(R^{(1)})\,\exp(-\tau S)\,\mathrm{d}\tau=\begin{pmatrix}\mathtt{J}_{2}\widehat{E}&\mathtt{J}_{2}F^{(2)}\\ \mathtt{J}_{2}[F^{(2)}]^{*}&\mathtt{J}_{2}\widehat{G}\end{pmatrix} (5.35)

where the 2×22\times 2 self-adjoint and reversible matrices E^\widehat{E}, G^\widehat{G} have entries

E^ij,G^ij=r(μϵ3),i,j=1,2,\widehat{E}_{ij}\ ,\widehat{G}_{ij}=r(\mu\epsilon^{3})\,,\quad i,j=1,2\,, (5.36)

and the 2×22\times 2 reversible matrix F(2)F^{(2)} admits an expansion as in (5.14).

Proof.

Since SS and R(1)R^{(1)} are Hamiltonian and reversibility-preserving then adSR(1)=[S,R(1)]\textup{ad}_{S}R^{(1)}=[S,R^{(1)}] is Hamiltonian and reversibility-preserving as well. Thus each exp(τS)adS2(R(1))exp(τS)\exp(\tau S)\,\textup{ad}_{S}^{2}(R^{(1)})\,\exp(-\tau S) is Hamiltonian and reversibility-preserving, and formula (5.35) holds. In order to estimate its entries we first compute adS2(R(1))\textup{ad}_{S}^{2}(R^{(1)}). Using the form of SS in (5.9) and [S,R(1)][S,R^{(1)}] in (5.32) one gets

adS2(R(1))=(0𝙹2F~𝙹2F~0)whereF~:=2(Σ𝙹2G~E~𝙹2Σ)\textup{ad}_{S}^{2}(R^{(1)})=\begin{pmatrix}0&\mathtt{J}_{2}\tilde{F}\\ \mathtt{J}_{2}\tilde{F}^{*}&0\end{pmatrix}\qquad\text{where}\qquad\tilde{F}:=2\left(\Sigma\mathtt{J}_{2}\tilde{G}-\tilde{E}\mathtt{J}_{2}\Sigma\right) (5.37)

and E~\tilde{E}, G~\tilde{G} are defined in (5.33). Since E~,G~=𝒪(μϵ2)\tilde{E},\tilde{G}=\mathcal{O}(\mu\epsilon^{2}) by (5.34), and Σ=𝙹2X=𝒪(ϵ)\Sigma=\mathtt{J}_{2}X=\mathcal{O}(\epsilon) by (5.8), we deduce that F~=𝒪(μϵ3)\tilde{F}=\mathcal{O}(\mu\epsilon^{3}). Then, for any τ[0,1]\tau\in[0,1], the matrix exp(τS)adS2(R(1))exp(τS)=adS2(R(1))(1+𝒪(μ,ϵ))\exp(\tau S)\,\textup{ad}_{S}^{2}(R^{(1)})\,\exp(-\tau S)=\textup{ad}_{S}^{2}(R^{(1)})(1+\mathcal{O}(\mu,\epsilon)). In particular the matrix F(2)F^{(2)} in (5.35) has the same expansion of F~\tilde{F}, namely F(2)=𝒪(μϵ3)F^{(2)}=\mathcal{O}(\mu\epsilon^{3}), and the matrices E^\widehat{E}, G^\widehat{G} have entries as in (5.36). ∎

Proof of Lemma 5.4..

It follows by (5.31)-(5.32), (5.15) and Lemmata 5.7 and 5.8. The matrix E(2):=E(1)+E~+E^E^{(2)}:=E^{(1)}+\tilde{E}+\widehat{E} has the expansion in (5.11), with 𝚎WB=𝚎11+𝚎~11\mathtt{e}_{\scriptscriptstyle{\textsc{WB}}}=\mathtt{e}_{11}+\tilde{\mathtt{e}}_{11} as in (5.12). Similarly G(2):=G(1)+G~+G^G^{(2)}:=G^{(1)}+\tilde{G}+\widehat{G} has the expansion in (5.13). ∎

5.2 Complete block-decoupling and proof of the main results

We now block-diagonalize the 4×44\times 4 Hamiltonian and reversible matrix 𝙻μ,ϵ(2)\mathtt{L}_{\mu,\epsilon}^{(2)} in (5.10). First we split it into its 2×22\times 2-diagonal and off-diagonal Hamiltonian and reversible matrices

𝙻μ,ϵ(2)=D(2)+R(2),\displaystyle\qquad\qquad\qquad\qquad\quad\mathtt{L}_{\mu,\epsilon}^{(2)}=D^{(2)}+R^{(2)}\,,
D(2):=(𝙹2E(2)00𝙹2G(2)),R(2):=(0𝙹2F(2)𝙹2[F(2)]0).\displaystyle D^{(2)}:=\begin{pmatrix}\mathtt{J}_{2}E^{(2)}&0\\ 0&\mathtt{J}_{2}G^{(2)}\end{pmatrix},\quad R^{(2)}:=\begin{pmatrix}0&\mathtt{J}_{2}F^{(2)}\\ \mathtt{J}_{2}[F^{(2)}]^{*}&0\end{pmatrix}. (5.38)
Lemma 5.9.

There exist a 4×44\times 4 reversibility-preserving Hamiltonian matrix S(2):=S(2)(μ,ϵ)S^{(2)}:=S^{(2)}(\mu,\epsilon) of the form (5.9), analytic in (μ,ϵ)(\mu,\epsilon), of size 𝒪(ϵ3)\mathcal{O}(\epsilon^{3}), and a 4×44\times 4 block-diagonal reversible Hamiltonian matrix P:=P(μ,ϵ)P:=P(\mu,\epsilon), analytic in (μ,ϵ)(\mu,\epsilon), of size 𝒪(μϵ6){\mathcal{O}(\mu\epsilon^{6})} such that

exp(S(2))(D(2)+R(2))exp(S(2))=D(2)+P.\exp(S^{(2)})(D^{(2)}+R^{(2)})\exp(-S^{(2)})=D^{(2)}+P\,. (5.39)
Proof.

We set for brevity S=S(2)S=S^{(2)}. The equation (5.39) is equivalent to the system

{ΠD(eS(D(2)+R(2))eS)D(2)=PΠ(eS(D(2)+R(2))eS)=0,\begin{cases}\Pi_{D}\big{(}e^{S}\big{(}D^{(2)}+R^{(2)}\big{)}e^{-S}\big{)}-D^{(2)}=P\\ \Pi_{\varnothing}\big{(}e^{S}\big{(}D^{(2)}+R^{(2)}\big{)}e^{-S}\big{)}=0\,,\end{cases} (5.40)

where ΠD\Pi_{D} is the projector onto the block-diagonal matrices and Π\Pi_{\varnothing} onto the block-off-diagonal ones. The second equation in (5.40) is equivalent, by a Lie expansion, and since [S,R(2)][S,R^{(2)}] is block-diagonal, to

R(2)+[S,D(2)]+Π01(1τ)eτSadS2(D(2)+R(2))eτSdτ=:(S)=0.R^{(2)}+\left[S\,,\,D^{(2)}\right]+\underbrace{\Pi_{\varnothing}\int_{0}^{1}(1-\tau)e^{\tau S}\text{ad}_{S}^{2}\big{(}D^{(2)}+R^{(2)}\big{)}e^{-\tau S}\mathrm{d}\tau}_{=:\mathcal{R}(S)}=0\,. (5.41)

The “nonlinear homological equation” (5.41),

[S,D(2)]=R(2)(S),[S,D^{(2)}]=-R^{(2)}-\mathcal{R}(S)\,, (5.42)

is equivalent to solve the 4×44\times 4 real linear system

𝒜x=f(μ,ϵ,x),f(μ,ϵ,x)=μv(μ,ϵ)+μg(μ,ϵ,x){\mathcal{A}}\vec{x}=\vec{f}(\mu,\epsilon,\vec{x})\,,\quad\vec{f}(\mu,\epsilon,\vec{x})=\mu\vec{v}(\mu,\epsilon)+\mu\vec{g}(\mu,\epsilon,\vec{x}) (5.43)

associated, as in (5.20), to (5.42). The vector μv(μ,ϵ)\mu\vec{v}(\mu,\epsilon) is associated with R(2)-R^{(2)} where R(2)R^{(2)} is in (5.38). The vector μg(μ,ϵ,x)\mu\vec{g}(\mu,\epsilon,\vec{x}) is associated with the matrix (S)-\mathcal{R}(S), which is a Hamiltonian and reversible block-off-diagonal matrix (i.e of the form (5.15)). The factor μ\mu is present in D(2)D^{(2)} and R(2)R^{(2)}, see (5.11), (5.13), (5.14) and the analytic function g(μ,ϵ,x)\vec{g}(\mu,\epsilon,\vec{x}) is quadratic in x\vec{x} (for the presence of adS2\text{ad}_{S}^{2} in (S)\mathcal{R}(S)). In view of (5.14) one has

μv(μ,ϵ):=(F21(2),F22(2),F11(2),F12(2)),Fij(2)=r(μϵ3).\mu\vec{v}(\mu,\epsilon):=(-F^{(2)}_{21},F^{(2)}_{22},-F^{(2)}_{11},F^{(2)}_{12})^{\top},\quad F^{(2)}_{ij}=\,{r(\mu\epsilon^{3})}\,. (5.44)

System (5.43) is equivalent to x=𝒜1f(μ,ϵ,x)\vec{x}={\mathcal{A}}^{-1}\vec{f}(\mu,\epsilon,\vec{x}) and, writing 𝒜1=1μ(μ,ϵ){\mathcal{A}}^{-1}=\frac{1}{\mu}{\mathcal{B}}(\mu,\epsilon) (cfr. (5.30)), to

x=(μ,ϵ)v(μ,ϵ)+(μ,ϵ)g(μ,ϵ,x).\vec{x}={\mathcal{B}}(\mu,\epsilon)\vec{v}(\mu,\epsilon)+{\mathcal{B}}(\mu,\epsilon)\vec{g}(\mu,\epsilon,\vec{x})\,.

By the implicit function theorem this equation admits a unique small solution x=x(μ,ϵ)\vec{x}=\vec{x}(\mu,\epsilon), analytic in (μ,ϵ)(\mu,\epsilon), with size 𝒪(ϵ3){\mathcal{O}(\epsilon^{3})} as v\vec{v} in (5.44). Then the first equation of (5.40) gives P=[S,R(2)]+ΠD01(1τ)eτSadS2(D(2)+R(2))eτSdτP=[S,R^{(2)}]+\Pi_{D}\int_{0}^{1}(1-\tau)e^{\tau S}\text{ad}_{S}^{2}\big{(}D^{(2)}+R^{(2)}\big{)}e^{-\tau S}\mathrm{d}\tau, and its estimate follows from those of SS and R(2)R^{(2)} (see (5.14)). ∎

Proof of Theorems 2.5 and 1.1. By Lemma 5.9 and recalling (3.1) the operator μ,ϵ:𝒱μ,ϵ𝒱μ,ϵ\mathcal{L}_{\mu,\epsilon}:\mathcal{V}_{\mu,\epsilon}\to\mathcal{V}_{\mu,\epsilon} is represented by the 4×44\times 4 Hamiltonian and reversible matrix

i𝚌𝚑μ+exp(S(2))𝙻μ,ϵ(2)exp(S(2))=i𝚌𝚑μ+(𝙹2E(3)00𝙹2G(3))=:(𝚄00𝚂),\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu+\exp(S^{(2)})\mathtt{L}_{\mu,\epsilon}^{(2)}\exp(-S^{(2)})=\mathrm{i}\,{\mathtt{c}}_{\mathtt{h}}\mu+\begin{pmatrix}\mathtt{J}_{2}E^{(3)}&0\\ 0&\mathtt{J}_{2}G^{(3)}\end{pmatrix}=:\begin{pmatrix}\mathtt{U}&0\\ 0&\mathtt{S}\end{pmatrix}\,,

where the matrices E(3)E^{(3)} and G(3)G^{(3)} expand as in (5.11), (5.13). Consequently the matrices 𝚄\mathtt{U} and 𝚂\mathtt{S} expand as in (2.40). Theorem 2.5 is proved. Theorem 1.1 is a straight-forward corollary. The function μ¯(ϵ)\underline{\mu}(\epsilon) in (1.4) is defined as the implicit solution of the function ΔBF(𝚑;μ,ϵ)\Delta_{\scriptscriptstyle{\textsc{BF}}}(\mathtt{h};\mu,\epsilon) in (1.6) for ϵ\epsilon small enough, depending on 𝚑\mathtt{h}.

Appendix A Expansion of the Kato basis

In this appendix we prove Lemma 4.2. We provide the expansion of the basis fk±(μ,ϵ)=Uμ,ϵfk±f_{k}^{\pm}(\mu,\epsilon)=U_{\mu,\epsilon}f_{k}^{\pm}, k=0,1k=0,1, in (4.1), where fk±f_{k}^{\pm} defined in (4.2) belong to the subspace 𝒱0,0:=Rg(P0,0)\mathcal{V}_{0,0}:=\text{Rg}(P_{0,0}). We first Taylor-expand the transformation operators Uμ,ϵU_{\mu,\epsilon} defined in (3.7). We denote ϵ\partial_{\epsilon} with an apex and μ\partial_{\mu} with a dot.

Lemma A.1.

The first jets of Uμ,ϵP0,0U_{\mu,\epsilon}P_{0,0} are

U0,0P0,0\displaystyle U_{0,0}P_{0,0} =P0,0,U0,0P0,0=P0,0P0,0,U˙0,0P0,0=P˙0,0P0,0,\displaystyle=P_{0,0}\,,\quad U_{0,0}^{\prime}P_{0,0}=P_{0,0}^{\prime}P_{0,0}\,,\quad\dot{U}_{0,0}P_{0,0}=\dot{P}_{0,0}P_{0,0}\,, (A.1)
U˙0,0P0,0\displaystyle\dot{U}_{0,0}^{\prime}P_{0,0} =(P˙0,012P0,0P˙0,0)P0,0,\displaystyle=\big{(}\dot{P}_{0,0}^{\prime}-\tfrac{1}{2}P_{0,0}\dot{P}_{0,0}^{\prime}\big{)}P_{0,0}\,, (A.2)

where

P0,0\displaystyle P_{0,0}^{\prime} =12πiΓ(0,0λ)10,0(0,0λ)1dλ,\displaystyle=\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathscr{L}_{0,0}^{\prime}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathrm{d}\lambda\,, (A.3)
P˙0,0\displaystyle\dot{P}_{0,0} =12πiΓ(0,0λ)1˙0,0(0,0λ)1dλ,\displaystyle=\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\dot{\mathscr{L}}_{0,0}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathrm{d}\lambda\,, (A.4)

and

P˙0,0\displaystyle\dot{P}_{0,0}^{\prime} =12πiΓ(0,0λ)1˙0,0(0,0λ)10,0(0,0λ)1dλ\displaystyle=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\dot{\mathscr{L}}_{0,0}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathscr{L}_{0,0}^{\prime}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathrm{d}\lambda (A.5a)
12πiΓ(0,0λ)10,0(0,0λ)1˙0,0(0,0λ)1dλ\displaystyle\qquad-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathscr{L}_{0,0}^{\prime}(\mathscr{L}_{0,0}-\lambda)^{-1}\dot{\mathscr{L}}_{0,0}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathrm{d}\lambda (A.5b)
+12πiΓ(0,0λ)1˙0,0(0,0λ)1dλ.\displaystyle\qquad+\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\dot{\mathscr{L}}_{0,0}^{\prime}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathrm{d}\lambda\,. (A.5c)

The operators 0,0\mathscr{L}_{0,0}^{\prime} and ˙0,0\dot{\mathscr{L}}_{0,0} are

0,0=[xp1(x)0a1(x)p1(x)x],˙0,0=[0sgn(D)m(D)00],\mathscr{L}_{0,0}^{\prime}=\begin{bmatrix}\partial_{x}\circ p_{1}(x)&0\\ -a_{1}(x)&p_{1}(x)\circ\partial_{x}\end{bmatrix},\qquad\dot{\mathscr{L}}_{0,0}=\begin{bmatrix}0&\textup{sgn}(D)m(D)\\ 0&0\end{bmatrix}, (A.6)

where sgn(D)\operatorname*{sgn}(D) is defined in (2.31) and m(D)m(D) is the real, even operator

m(D):=tanh(𝚑|D|)+𝚑|D|(1tanh2(𝚑|D|))m(D):=\tanh(\mathtt{h}|D|)+\mathtt{h}|D|(1-\tanh^{2}(\mathtt{h}|D|)) (A.7)

and a1(x)a_{1}(x) and p1(x)p_{1}(x) are given in Lemma 2.2.

The operator ˙0,0\dot{\mathscr{L}}_{0,0}^{\prime} is

˙0,0=[ip1(x)00ip1(x)].\dot{\mathscr{L}}_{0,0}^{\prime}=\begin{bmatrix}\mathrm{i}\,p_{1}(x)&0\\ 0&\mathrm{i}\,p_{1}(x)\end{bmatrix}\,. (A.8)
Proof.

By (3.7) and (3.6) one has the Taylor expansion in (Y)\mathcal{L}(Y)

Uμ,ϵP0,0=Pμ,ϵP0,0+12(Pμ,ϵP0,0)2Pμ,ϵP0,0+𝒪(Pμ,ϵP0,0)4,U_{\mu,\epsilon}P_{0,0}=P_{\mu,\epsilon}P_{0,0}+\frac{1}{2}(P_{\mu,\epsilon}-P_{0,0})^{2}P_{\mu,\epsilon}P_{0,0}+\mathcal{O}(P_{\mu,\epsilon}-P_{0,0})^{4}\,,

where 𝒪(Pμ,ϵP0,0)4=𝒪(ϵ4,ϵ3μ,ϵ2μ2,ϵμ3,μ4)(Y)\mathcal{O}(P_{\mu,\epsilon}-P_{0,0})^{4}=\mathcal{O}(\epsilon^{4},\epsilon^{3}\mu,\epsilon^{2}\mu^{2},\epsilon\mu^{3},\mu^{4})\in\mathcal{L}(Y). Consequently one derives (A.1), (A.2), using also the identity P˙0,0P0,0P0,0+P0,0P˙0,0P0,0=P0,0P˙0,0P0,0\dot{P}_{0,0}P_{0,0}^{\prime}P_{0,0}+P_{0,0}^{\prime}\dot{P}_{0,0}P_{0,0}=-P_{0,0}\dot{P}_{0,0}^{\prime}P_{0,0}, which follows differentiating Pμ,ϵ2=Pμ,ϵP_{\mu,\epsilon}^{2}=P_{\mu,\epsilon}. Differentiating (3.5) one gets (A.3)-(A.5c). Formulas (A.6)-(A.8) follow by (3.2) using also that the Fourier multiplier Π0(tanh(𝚑|D|)+𝚑|D|(1tanh2(𝚑|D|)))=0\Pi_{0}\big{(}\tanh(\mathtt{h}|D|)+\mathtt{h}|D|\big{(}1-\tanh^{2}(\mathtt{h}|D|)\big{)}\big{)}=0. ∎

By the previous lemma we have the Taylor expansion

fkσ(μ,ϵ)=fkσ+ϵP0,0fkσ+μP˙0,0fkσ+μϵ(P˙0,012P0,0P˙0,0)fkσ+𝒪(μ2,ϵ2).f_{k}^{\sigma}(\mu,\epsilon)=f_{k}^{\sigma}+\epsilon P_{0,0}^{\prime}f_{k}^{\sigma}+\mu\dot{P}_{0,0}f_{k}^{\sigma}+\mu\epsilon\big{(}\dot{P}_{0,0}^{\prime}-\frac{1}{2}P_{0,0}\dot{P}_{0,0}^{\prime}\big{)}f_{k}^{\sigma}+\mathcal{O}(\mu^{2},\epsilon^{2})\,. (A.9)

In order to compute the vectors P0,0fkσP_{0,0}^{\prime}f_{k}^{\sigma} and P˙0,0fkσ\dot{P}_{0,0}f_{k}^{\sigma} using (A.3) and (A.4), it is useful to know the action of (0,0λ)1(\mathscr{L}_{0,0}-\lambda)^{-1} on the vectors

fk+:=[𝚌𝚑1/2cos(kx)𝚌𝚑1/2sin(kx)],fk:=[𝚌𝚑1/2sin(kx)𝚌𝚑1/2cos(kx)],\displaystyle f_{k}^{+}:=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{1/2}\cos(kx)\\ {\mathtt{c}}_{\mathtt{h}}^{-1/2}\sin(kx)\end{bmatrix}\,,\quad f_{k}^{-}:=\begin{bmatrix}-{\mathtt{c}}_{\mathtt{h}}^{1/2}\sin(kx)\\ {\mathtt{c}}_{\mathtt{h}}^{-1/2}\cos(kx)\end{bmatrix}\,, (A.10)
fk+:=[𝚌𝚑1/2cos(kx)𝚌𝚑1/2sin(kx)],fk:=[𝚌𝚑1/2sin(kx)𝚌𝚑1/2cos(kx)],k.\displaystyle f_{-k}^{+}:=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{1/2}\cos(kx)\\ -{\mathtt{c}}_{\mathtt{h}}^{-1/2}\sin(kx)\end{bmatrix}\,,\quad f_{-k}^{-}:=\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}^{1/2}\sin(kx)\\ {\mathtt{c}}_{\mathtt{h}}^{-1/2}\cos(kx)\end{bmatrix}\,,\quad k\in\mathbb{N}\,.
Lemma A.2.

The space H1(𝕋)H^{1}(\mathbb{T}) decomposes as H1(𝕋)=𝒱0,0𝒰𝒲H1H^{1}(\mathbb{T})=\mathcal{V}_{0,0}\oplus\mathcal{U}\oplus{\mathcal{W}_{H^{1}}}, with 𝒲H1=k=2𝒲k¯H1\mathcal{W}_{H^{1}}=\overline{\bigoplus\limits_{k=2}^{\infty}\mathcal{W}_{k}}^{H^{1}} where the subspaces 𝒱0,0,𝒰\mathcal{V}_{0,0},\mathcal{U} and 𝒲k\mathcal{W}_{k}, defined below, are invariant under 0,0\mathscr{L}_{0,0} and the following properties hold:

  • (i)

    𝒱0,0=span{f1+,f1,f0+,f0}\mathcal{V}_{0,0}=\text{span}\{f^{+}_{1},f^{-}_{1},f^{+}_{0},f^{-}_{0}\} is the generalized kernel of 0,0\mathscr{L}_{0,0}. For any λ0\lambda\neq 0 the operator 0,0λ:𝒱0,0𝒱0,0\mathscr{L}_{0,0}-\lambda:\mathcal{V}_{0,0}\to\mathcal{V}_{0,0} is invertible and

    (0,0λ)1f1+=1λf1+,(0,0λ)1f1=1λf1,(0,0λ)1f0=1λf0,\displaystyle(\mathscr{L}_{0,0}-\lambda)^{-1}f_{1}^{+}=-\frac{1}{\lambda}f_{1}^{+}\,,\quad(\mathscr{L}_{0,0}-\lambda)^{-1}f_{1}^{-}=-\frac{1}{\lambda}f_{1}^{-},\quad(\mathscr{L}_{0,0}-\lambda)^{-1}f_{0}^{-}=-\frac{1}{\lambda}f_{0}^{-}\,, (A.11)
    (0,0λ)1f0+=1λf0++1λ2f0.\displaystyle(\mathscr{L}_{0,0}-\lambda)^{-1}f_{0}^{+}=-\frac{1}{\lambda}f_{0}^{+}+\frac{1}{\lambda^{2}}f_{0}^{-}\,. (A.12)
  • (ii)

    𝒰:=span{f1+,f1}\mathcal{U}:=\text{span}\left\{f_{-1}^{+},f_{-1}^{-}\right\}. For any λ±2i\lambda\neq\pm 2\mathrm{i}\, the operator 0,0λ:𝒰𝒰\mathscr{L}_{0,0}-\lambda:\mathcal{U}\to\mathcal{U} is invertible and

    (0,0λ)1f1+=1λ2+4𝚌𝚑2(λf1++2𝚌𝚑f1),\displaystyle(\mathscr{L}_{0,0}-\lambda)^{-1}f_{-1}^{+}=\frac{1}{\lambda^{2}+4{\mathtt{c}}_{\mathtt{h}}^{2}}\left(-\lambda f_{-1}^{+}+2{\mathtt{c}}_{\mathtt{h}}f_{-1}^{-}\right), (A.13)
    (0,0λ)1f1=1λ2+4𝚌𝚑2(2𝚌𝚑f1+λf1).\displaystyle(\mathscr{L}_{0,0}-\lambda)^{-1}f_{-1}^{-}=\frac{1}{\lambda^{2}+4{\mathtt{c}}_{\mathtt{h}}^{2}}\left(-2{\mathtt{c}}_{\mathtt{h}}f_{-1}^{+}-\lambda f_{-1}^{-}\right)\,.
  • (iii)

    Each subspace 𝒲k:=span{fk+,fk,fk+,fk}\mathcal{W}_{k}:=\text{span}\left\{f_{k}^{+},\ f_{k}^{-},f_{-k}^{+},\ f_{-k}^{-}\right\} is invariant under 0,0\mathscr{L}_{0,0}. Let 𝒲L2=k=2𝒲k¯L2\mathcal{W}_{L^{2}}=\overline{\bigoplus\limits_{k=2}^{\infty}\mathcal{W}_{k}}^{L^{2}}. For any |λ|<δ(𝚑)|\lambda|<\delta(\mathtt{h}) small enough, the operator 0,0λ:𝒲H1𝒲L2{\mathscr{L}_{0,0}-\lambda:\mathcal{W}_{H^{1}}\to\mathcal{W}_{L^{2}}} is invertible and for any f𝒲L2f\in{\mathcal{W}_{L^{2}}}

    (0,0λ)1f=(𝚌𝚑2x2+|D|tanh(𝚑|D|))1[𝚌𝚑x|D|tanh(𝚑|D|)1𝚌𝚑x]f+λφf(λ,x),(\mathscr{L}_{0,0}-\lambda)^{-1}f=\big{(}{\mathtt{c}}_{\mathtt{h}}^{2}\partial_{x}^{2}+|D|\tanh(\mathtt{h}|D|)\big{)}^{-1}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}\partial_{x}&-|D|\tanh(\mathtt{h}|D|)\\ 1&{\mathtt{c}}_{\mathtt{h}}\partial_{x}\end{bmatrix}f+\lambda\varphi_{f}(\lambda,x)\,, (A.14)

    for some analytic function λφf(λ,)H1(𝕋,2)\lambda\mapsto\varphi_{f}(\lambda,\cdot)\in H^{1}(\mathbb{T},\mathbb{C}^{2}).

Proof.

By inspection the spaces 𝒱0,0\mathcal{V}_{0,0}, 𝒰\mathcal{U} and 𝒲k{\mathcal{W}_{k}} are invariant under 0,0\mathscr{L}_{0,0} and, by Fourier series, they decompose H1(𝕋,2)H^{1}(\mathbb{T},\mathbb{C}^{2}). Formulas (A.11)-(A.12) follow using that f1+,f1,f0f_{1}^{+},f_{1}^{-},f_{0}^{-} are in the kernel of 0,0\mathscr{L}_{0,0}, and 0,0f0+=f0\mathscr{L}_{0,0}f_{0}^{+}=-f_{0}^{-}. Formula (A.13) follows using that 0,0f1+=2𝚌𝚑f1\mathscr{L}_{0,0}f^{+}_{-1}=-2{\mathtt{c}}_{\mathtt{h}}f^{-}_{-1} and 0,0f1=2𝚌𝚑f1+\mathscr{L}_{0,0}f^{-}_{-1}=2{\mathtt{c}}_{\mathtt{h}}f^{+}_{-1}. Let us prove item (iii)(iii). Let 𝒲:=𝒲H1\mathcal{W}:=\mathcal{W}_{H^{1}}. The operator (0,0λId)|𝒲{{\left.\kern-1.2pt(\mathscr{L}_{0,0}-\lambda\mathrm{Id})\vphantom{\big{|}}\right|_{\mathcal{W}}}} is invertible for any λ{±i|k|tanh(𝚑|k|)±ik𝚌𝚑,k2,k}\lambda\notin\{\pm\mathrm{i}\,\sqrt{|k|\tanh{(\mathtt{h}|k|)}}\pm\mathrm{i}\,k{\mathtt{c}}_{\mathtt{h}},k\geq 2,k\in{\mathbb{N}}\} and

(0,0|𝒲)1=(𝚌𝚑2x2+|D|tanh(𝚑|D|))1[𝚌𝚑x|D|tanh(𝚑|D|))1𝚌𝚑x]|𝒲.\footnotesize({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}=\left({\mathtt{c}}_{\mathtt{h}}^{2}\partial_{x}^{2}+|D|\tanh(\mathtt{h}|D|)\right)^{-1}\begin{bmatrix}{\mathtt{c}}_{\mathtt{h}}\partial_{x}&-|D|\tanh(\mathtt{h}|D|))\\ 1&{\mathtt{c}}_{\mathtt{h}}\partial_{x}\end{bmatrix}_{|\mathcal{W}}\,.

By Neumann series, for any λ\lambda such that |λ|(0,0|𝒲)1(𝒲,H1(𝕋))<1|\lambda|\|({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}\|_{{\mathcal{L}(\mathcal{W},H^{1}(\mathbb{T}))}}<1 we have

(0,0|𝒲λ)1=(0,0|𝒲)1(Idλ(0,0|𝒲)1)1=(0,0|𝒲)1k0((0,0|𝒲)1λ)k.({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}}-\lambda)^{-1}=({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}\big{(}\mathrm{Id}-\lambda({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}\big{)}^{-1}=({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}\sum_{k\geq 0}(({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}\lambda)^{k}\,.

Formula (A.14) follows with φf(λ,x):=(0,0|𝒲)1k1λk1[(0,0|𝒲)1]kf\varphi_{f}(\lambda,x):=({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}\sum_{k\geq 1}\lambda^{k-1}[({\left.\kern-1.2pt\mathscr{L}_{0,0}\vphantom{\big{|}}\right|_{\mathcal{W}}})^{-1}]^{k}f. ∎

We shall also use the following formulas obtained by (A.6), (A.7) and (4.2):

0,0f1+=[2𝚌𝚑1/2sin(2x)12𝚌𝚑5/2(1𝚌𝚑4)(1+cos(2x))],0,0f1=[2𝚌𝚑1/2cos(2x)12𝚌𝚑5/2(1𝚌𝚑4)sin(2x)],\displaystyle\mathscr{L}_{0,0}^{\prime}f_{1}^{+}=\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1/2}\,\sin(2x)\\ \frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{5/2}(1-{\mathtt{c}}_{\mathtt{h}}^{-4})(1+\cos(2x))\end{bmatrix}\,,\qquad\mathscr{L}_{0,0}^{\prime}f_{1}^{-}=\begin{bmatrix}2\,{\mathtt{c}}_{\mathtt{h}}^{-1/2}\,\cos(2x)\\ -\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{5/2}(1-{\mathtt{c}}_{\mathtt{h}}^{-4})\sin(2x)\end{bmatrix}\,, (A.15)
0,0f0+=[2𝚌𝚑1sin(x)(𝚌𝚑2+𝚌𝚑2)cos(x)],0,0f0=0,\displaystyle\mathscr{L}_{0,0}^{\prime}f_{0}^{+}=\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x)\\ \left({\mathtt{c}}_{\mathtt{h}}^{2}+{\mathtt{c}}_{\mathtt{h}}^{-2}\right)\cos(x)\end{bmatrix}\,,\qquad\mathscr{L}_{0,0}^{\prime}f_{0}^{-}=0\,,
˙0,0f1+=ib(𝚑)[cos(x)0],˙0,0f1=ib(𝚑)[sin(x)0],b(𝚑):=𝚌𝚑1/2(𝚌𝚑2+𝚑(1𝚌𝚑4)),\displaystyle\dot{\mathscr{L}}_{0,0}f_{1}^{+}=-\mathrm{i}\,b(\mathtt{h})\begin{bmatrix}\cos(x)\\ 0\end{bmatrix}\,,\qquad\dot{\mathscr{L}}_{0,0}f_{1}^{-}=\mathrm{i}\,b(\mathtt{h})\begin{bmatrix}\sin(x)\\ 0\end{bmatrix}\,,\quad b(\mathtt{h}):={\mathtt{c}}_{\mathtt{h}}^{-1/2}\big{(}{\mathtt{c}}_{\mathtt{h}}^{2}+\mathtt{h}(1-{\mathtt{c}}_{\mathtt{h}}^{4})\big{)}\,,
˙0,0f0+=0,˙0,0f0=0.\displaystyle\dot{\mathscr{L}}_{0,0}f_{0}^{+}=0\,,\qquad\dot{\mathscr{L}}_{0,0}f_{0}^{-}=0\,.
Remark.

In deep water we have ˙0,0f0=f0+\dot{\mathscr{L}}_{0,0}f_{0}^{-}=f_{0}^{+} (cfr. formula (A.14) in [6]). In finite depth instead ˙0,0f0=0\dot{\mathscr{L}}_{0,0}f_{0}^{-}=0 because the Fourier multiplier sgn(D)m(D)\operatorname*{sgn}(D)m(D) in (A.7) vanishes on the constants.

We finally compute P0,0fkσP_{0,0}^{\prime}f_{k}^{\sigma} and P˙0,0fkσ\dot{P}_{0,0}f_{k}^{\sigma}.

Lemma A.3.

One has

P0,0f1+=[12𝚌𝚑112(3+𝚌𝚑4)cos(2x)14𝚌𝚑132(1+𝚌𝚑4)(3𝚌𝚑4)sin(2x)],P0,0f1=[12𝚌𝚑112(3+𝚌𝚑4)sin(2x)14𝚌𝚑132(1+𝚌𝚑4)(3𝚌𝚑4)cos(2x)],\displaystyle P_{0,0}^{\prime}f^{+}_{1}={\begin{bmatrix}\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{11}{2}}(3+{\mathtt{c}}_{\mathtt{h}}^{4})\,\cos(2x)\\ \frac{1}{4}{\mathtt{c}}_{\mathtt{h}}^{-\frac{13}{2}}(1+{\mathtt{c}}_{\mathtt{h}}^{4})(3-{\mathtt{c}}_{\mathtt{h}}^{4})\sin(2x)\end{bmatrix}}\,,\quad P_{0,0}^{\prime}f^{-}_{1}={\begin{bmatrix}-\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{11}{2}}(3+{\mathtt{c}}_{\mathtt{h}}^{4})\,\sin(2x)\\ \frac{1}{4}{\mathtt{c}}_{\mathtt{h}}^{-\frac{13}{2}}(1+{\mathtt{c}}_{\mathtt{h}}^{4})(3-{\mathtt{c}}_{\mathtt{h}}^{4})\cos(2x)\end{bmatrix}}\,, (A.16)
P0,0f0+=14𝚌𝚑52(3+𝚌𝚑4)f1+,P0,0f0=0,P˙0,0f0+=0,P˙0,0f0=0,\displaystyle P_{0,0}^{\prime}f^{+}_{0}=\tfrac{1}{4}{\mathtt{c}}_{\mathtt{h}}^{-\frac{5}{2}}(3+{\mathtt{c}}_{\mathtt{h}}^{4})f^{+}_{-1}\,,\quad P_{0,0}^{\prime}f^{-}_{0}=0\,,\quad\dot{P}_{0,0}f_{0}^{+}=0\,,\quad\dot{P}_{0,0}f_{0}^{-}=0\,,
P˙0,0f1+=i4(1+𝚌𝚑2𝚑(1𝚌𝚑4))f1,P˙0,0f1=i4(1+𝚌𝚑2𝚑(1𝚌𝚑4))f1+.\displaystyle\dot{P}_{0,0}f_{1}^{+}=\frac{\mathrm{i}\,}{4}\big{(}1+{\mathtt{c}}_{\mathtt{h}}^{-2}\mathtt{h}(1-{\mathtt{c}}_{\mathtt{h}}^{4})\big{)}f^{-}_{-1}\,,\quad\dot{P}_{0,0}f_{1}^{-}=\frac{\mathrm{i}\,}{4}\big{(}1+{\mathtt{c}}_{\mathtt{h}}^{-2}\mathtt{h}(1-{\mathtt{c}}_{\mathtt{h}}^{4})\big{)}f^{+}_{-1}\,.
Proof.

We first compute P0,0f1+P_{0,0}^{\prime}f_{1}^{+}. By (A.3), (A.11) and (A.15) we deduce

P0,0f1+=12πiΓ1λ(0,0λ)1[2𝚌𝚑1/2sin(2x)12𝚌𝚑5/2(1𝚌𝚑4)(1+cos(2x))]dλ.P_{0,0}^{\prime}f_{1}^{+}=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}\frac{1}{\lambda}(\mathscr{L}_{0,0}-\lambda)^{-1}\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1/2}\,\sin(2x)\\ \frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{5/2}(1-{\mathtt{c}}_{\mathtt{h}}^{-4})(1+\cos(2x))\end{bmatrix}\mathrm{d}\lambda\,.

We note that [2𝚌𝚑1/2sin(2x)12𝚌𝚑5/2(1𝚌𝚑4)(1+cos(2x))]=12𝚌𝚑5/2(1𝚌𝚑4)f0+𝒲\footnotesize\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1/2}\,\sin(2x)\\ \frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{5/2}(1-{\mathtt{c}}_{\mathtt{h}}^{-4})(1+\cos(2x))\end{bmatrix}=\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{5/2}(1-{\mathtt{c}}_{\mathtt{h}}^{-4})f_{0}^{-}+\mathcal{W}. Therefore by (A.11) and (A.14) there is an analytic function λφ(λ,)H1(𝕋,2)\lambda\mapsto\varphi(\lambda,\cdot)\in H^{1}(\mathbb{T},\mathbb{C}^{2}) so that

P0,0f1+=12πiΓ1λ(𝚌𝚑5/2(1𝚌𝚑4)2λf01+𝚌𝚑44𝚌𝚑6[2𝚌𝚑𝚌𝚑12(3+𝚌𝚑4)1+𝚌𝚑4cos(2x)𝚌𝚑12(3𝚌𝚑4)sin(2x)]+λφ(λ))dλ,P_{0,0}^{\prime}f_{1}^{+}=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}\frac{1}{\lambda}\Big{(}-\dfrac{{\mathtt{c}}_{\mathtt{h}}^{5/2}(1-{\mathtt{c}}_{\mathtt{h}}^{-4})}{2\lambda}f_{0}^{-}{-\frac{1+{\mathtt{c}}_{\mathtt{h}}^{4}}{4{\mathtt{c}}_{\mathtt{h}}^{6}}\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}\frac{{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}(3+{\mathtt{c}}_{\mathtt{h}}^{4})}{1+{\mathtt{c}}_{\mathtt{h}}^{4}}\cos(2x)\\ {\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}(3-{\mathtt{c}}_{\mathtt{h}}^{4})\sin(2x)\end{bmatrix}}+\lambda\varphi(\lambda)\Big{)}\,\mathrm{d}\lambda\,,

where we exploited the identity tanh(2𝚑)=2𝚌𝚑21+𝚌𝚑4\tanh(2\mathtt{h})=\frac{2{\mathtt{c}}_{\mathtt{h}}^{2}}{1+{\mathtt{c}}_{\mathtt{h}}^{4}} in applying (A.14). Thus, by means of residue Theorem we obtain the first identity in (A.16). Similarly one computes P0,0f1P_{0,0}^{\prime}f_{1}^{-}. By (A.3), (A.11) and (A.15), one has P0,0f0=0P_{0,0}^{\prime}f_{0}^{-}=0. Next we compute P0,0f0+P_{0,0}^{\prime}f_{0}^{+}. By (A.3), (A.11), (A.12) and (A.15) we get

P0,0f0+=12πiΓ1λ(0,0λ)1[2𝚌𝚑1sin(x)(𝚌𝚑2+𝚌𝚑2)cos(x)]dλ.P_{0,0}^{\prime}f_{0}^{+}=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}\frac{1}{\lambda}(\mathscr{L}_{0,0}-\lambda)^{-1}\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x)\\ ({\mathtt{c}}_{\mathtt{h}}^{2}+{\mathtt{c}}_{\mathtt{h}}^{-2})\cos(x)\end{bmatrix}\mathrm{d}\lambda\,.

Next we decompose [2𝚌𝚑1sin(x)(𝚌𝚑2+𝚌𝚑2)cos(x)]=12𝚌𝚑32(𝚌𝚑4+3)=:αf1+12𝚌𝚑32(𝚌𝚑41)=:βf1\footnotesize\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x)\\ ({\mathtt{c}}_{\mathtt{h}}^{2}+{\mathtt{c}}_{\mathtt{h}}^{-2})\cos(x)\end{bmatrix}=\underbrace{{\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{3}{2}}({\mathtt{c}}_{\mathtt{h}}^{4}+3)}}_{=:\alpha}f^{-}_{-1}+\underbrace{{\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{3}{2}}({\mathtt{c}}_{\mathtt{h}}^{4}-1)}}_{=:\beta}f^{-}_{1}. By (A.15) and (A.13) we get

P0,0f0+=12πiΓ(2α𝚌𝚑λ(λ2+4𝚌𝚑2)f1+αλ2+4𝚌𝚑2f1+βλ2f1)dλ=α2𝚌𝚑f1+,P_{0,0}^{\prime}f_{0}^{+}=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}\Big{(}-\frac{2\alpha{\mathtt{c}}_{\mathtt{h}}}{\lambda(\lambda^{2}+4{\mathtt{c}}_{\mathtt{h}}^{2})}f_{-1}^{+}-\frac{\alpha}{\lambda^{2}+4{\mathtt{c}}_{\mathtt{h}}^{2}}f_{-1}^{-}+\frac{\beta}{\lambda^{2}}f^{-}_{1}\Big{)}\mathrm{d}\lambda=\frac{\alpha}{2{\mathtt{c}}_{\mathtt{h}}}f^{+}_{-1}\,,

where in the last step we used the residue theorem. We compute now P˙0,0f1+\dot{P}_{0,0}f^{+}_{1}. First we have P˙0,0f1+=i2πib(𝚑)Γ1λ(0,0λ)1[cos(x)0]dλ\dot{P}_{0,0}f_{1}^{+}=\ \frac{\mathrm{i}\,}{2\pi\mathrm{i}\,}b({\mathtt{h}})\oint_{\Gamma}\frac{1}{\lambda}(\mathscr{L}_{0,0}-\lambda)^{-1}\footnotesize\begin{bmatrix}\cos(x)\\ 0\end{bmatrix}\mathrm{d}\lambda, where b(𝚑)b(\mathtt{h}) is in (A.15), and then, writing [cos(x)0]=12𝚌𝚑12(f1++f1+)\footnotesize\begin{bmatrix}\cos(x)\\ 0\end{bmatrix}=\frac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-\frac{1}{2}}(f_{1}^{+}+f_{-1}^{+}) and using (A.13), we conclude using again the residue theorem P˙0,0f1+=i4(1+𝚑(1𝚌𝚑4)𝚌𝚑2)f1\dot{P}_{0,0}f_{1}^{+}=\frac{\mathrm{i}\,}{4}\big{(}1+{\mathtt{h}}(1-{\mathtt{c}}_{\mathtt{h}}^{4}){\mathtt{c}}_{\mathtt{h}}^{-2}\big{)}f^{-}_{-1}. The computation of P˙0,0f1\dot{P}_{0,0}f^{-}_{1} is analogous. Finally, in view of (A.15), we have

P˙0,0f0+=12πiΓ(0,0λ)1˙0,0(1λ2f01λf0+)dλ=0,\displaystyle\dot{P}_{0,0}f^{+}_{0}=\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathcal{L}_{0,0}-\lambda)^{-1}\dot{\mathcal{L}}_{0,0}\big{(}\frac{1}{\lambda^{2}}f_{0}^{-}-\frac{1}{\lambda}f_{0}^{+}\big{)}\mathrm{d}\lambda=0\,,
P˙0,0f0=12πiΓ1λ(0,0λ)1˙0,0f0dλ=0.\displaystyle\dot{P}_{0,0}f^{-}_{0}=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}\frac{1}{\lambda}(\mathcal{L}_{0,0}-\lambda)^{-1}\dot{\mathcal{L}}_{0,0}f_{0}^{-}\mathrm{d}\lambda=0\,.

In conclusion all the formulas in (A.16) are proved. ∎

So far we have obtained the linear terms of the expansions (4.3), (4.4), (4.5), (4.6). We now provide further information about the expansion of the basis at μ=0\mu=0. The proof of the next lemma follows as that of Lemma A.4 in [6].

Lemma A.4.

The basis {fkσ(0,ϵ),k=0,1,σ=±}\{f_{k}^{\sigma}(0,\epsilon),\,k=0,1,\sigma=\pm\} is real. For any ϵ\epsilon it results f0(0,ϵ)f0f_{0}^{-}(0,\epsilon)\equiv f_{0}^{-}. The property (4.8) holds.

We now provide further information about the expansion of the basis at ϵ=0\epsilon=0. The following lemma follows as Lemma A.5 in [6]. The key observation is that the operator μ,0|𝒵{\left.\kern-1.2pt\mathscr{L}_{\mu,0}\vphantom{\big{|}}\right|_{\mathcal{Z}}}, where 𝒵\mathcal{Z} is the invariant subspace 𝒵:=span{f0+,f0}\mathcal{Z}:=\text{span}\{f_{0}^{+},\,f_{0}^{-}\}, has the two eigenvalues ±iμtanh(𝚑μ)\pm\mathrm{i}\,\sqrt{\mu\tanh(\mathtt{h}\mu)}, which, for small μ\mu, lie inside the loop Γ\Gamma around 0 in (3.5).

Lemma A.5.

For any small μ\mu, we have f0+(μ,0)f0+f_{0}^{+}(\mu,0)\equiv f_{0}^{+} and f0(μ,0)f0f_{0}^{-}(\mu,0)\equiv f_{0}^{-}. Moreover the vectors f1+(μ,0)f_{1}^{+}(\mu,0) and f1(μ,0)f_{1}^{-}(\mu,0) have both components with zero space average.

We finally consider the μϵ\mu\epsilon term in the expansion (A.9).

Lemma A.6.

The derivatives (μϵfkσ)(0,0)=(P˙0,012P0,0P˙0,0)fkσ(\partial_{\mu}\partial_{\epsilon}f_{k}^{\sigma})(0,0)=\left(\dot{P}_{0,0}^{\prime}-\frac{1}{2}P_{0,0}\dot{P}_{0,0}^{\prime}\right)f_{k}^{\sigma} satisfy

(μϵf1+)(0,0)=i[odd(x)even(x)],(μϵf1)(0,0)=i[even(x)odd(x)],\displaystyle(\partial_{\mu}\partial_{\epsilon}f_{1}^{+})(0,0)=\mathrm{i}\,\begin{bmatrix}odd(x)\\ even(x)\end{bmatrix},\qquad(\partial_{\mu}\partial_{\epsilon}f_{1}^{-})(0,0)-=\mathrm{i}\,\begin{bmatrix}even(x)\\ odd(x)\end{bmatrix}, (A.17)
(μϵf0+)(0,0)=i[odd(x)even0(x)],(μϵf0)(0,0)=i[even0(x)odd(x)].\displaystyle(\partial_{\mu}\partial_{\epsilon}f_{0}^{+})(0,0)=\mathrm{i}\,\begin{bmatrix}odd(x)\\ even_{0}(x)\end{bmatrix},\qquad(\partial_{\mu}\partial_{\epsilon}f_{0}^{-})(0,0)=\mathrm{i}\,\begin{bmatrix}even_{0}(x)\\ odd(x)\end{bmatrix}\,.
Proof.

We prove that P˙0,0=(A.5a)+(A.5b)+(A.5c)\dot{P}^{\prime}_{0,0}=\eqref{Pmisto1}+\eqref{Pmisto2}+\eqref{Pmisto3} is purely imaginary, see footnote 4. This follows since the operators in (A.5a)\eqref{Pmisto1}, (A.5b)\eqref{Pmisto2} and (A.5c)\eqref{Pmisto3} are purely imaginary because ˙0,0\dot{\mathscr{L}}_{0,0} is purely imaginary, 0,0\mathscr{L}_{0,0}^{\prime} in (A.6) is real and ˙0,0\dot{\mathscr{L}}_{0,0}^{\prime} in (A.8) is purely imaginary (argue as in Lemma 3.2-(iii)(iii) of [6]). Then, applied to the real vectors fkσf^{\sigma}_{k}, k=0,1k=0,1, σ=±\sigma=\pm, give purely imaginary vectors.

The property (3.10) implies that (μϵfkσ)(0,0)(\partial_{\mu}\partial_{\epsilon}f_{k}^{\sigma})(0,0) have the claimed parity structure in (A.17). We shall now prove that (μϵf0±)(0,0)(\partial_{\mu}\partial_{\epsilon}f_{0}^{\pm})(0,0) have zero average. We have, by (A.12) and (A.15)

(A.5a)f0+:=12πiΓ(0,0λ)1˙0,0(0,0λ)11λ[2𝚌𝚑1sin(x)(𝚌𝚑2+𝚌𝚑2)cos(x)]dλ\displaystyle\eqref{Pmisto1}f_{0}^{+}:=\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\dot{\mathscr{L}}_{0,0}(\mathscr{L}_{0,0}-\lambda)^{-1}\frac{1}{\lambda}\begin{bmatrix}2{\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x)\\ \left({\mathtt{c}}_{\mathtt{h}}^{2}+{\mathtt{c}}_{\mathtt{h}}^{-2}\right)\cos(x)\end{bmatrix}\,\mathrm{d}\lambda

and since the operators (0,0λ)1(\mathscr{L}_{0,0}-\lambda)^{-1} and ˙0,0\dot{\mathscr{L}}_{0,0} are both Fourier multipliers, hence they preserve the absence of average of the vectors, then (A.5a)f0+\eqref{Pmisto1}f_{0}^{+} has zero average. Next (A.5b)f0+=0\eqref{Pmisto2}f_{0}^{+}=0 since ˙0,0f0±=0\dot{\mathscr{L}}_{0,0}f_{0}^{\pm}=0, cfr. (2.31). Finally, by (A.12) and (A.8) where p1(x)=p1[1]cos(x)p_{1}(x)=p_{1}^{[1]}\cos(x),

(A.5c)f0+=ip1[1]2πiΓ(0,0λ)1(1λ[cos(x)0]+1λ2[0cos(x)])dλ\eqref{Pmisto3}f_{0}^{+}=\frac{\mathrm{i}\,p_{1}^{[1]}}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\Big{(}-\frac{1}{\lambda}\begin{bmatrix}\cos(x)\\ 0\end{bmatrix}+\frac{1}{\lambda^{2}}\begin{bmatrix}0\\ \cos(x)\end{bmatrix}\Big{)}\,\mathrm{d}\lambda

is a vector with zero average. We conclude that P˙0,0f0+\dot{P}_{0,0}^{\prime}f_{0}^{+} is an imaginary vector with zero average, as well as (μϵf0+)(0,0)(\partial_{\mu}\partial_{\epsilon}f_{0}^{+})(0,0) since P0,0P_{0,0} sends zero average functions in zero average functions. Finally, by (3.10), (μϵf0+)(0,0)(\partial_{\mu}\partial_{\epsilon}f_{0}^{+})(0,0) has the claimed structure in (A.17).

We finally consider (μϵf0)(0,0)(\partial_{\mu}\partial_{\epsilon}f_{0}^{-})(0,0). By (A.11) and 0,0f0=0\mathscr{L}_{0,0}^{\prime}f_{0}^{-}=0 (cfr. (A.15)), it results

(A.5a)f0=12πiΓ(0,0λ)1λ˙0,0(0,0λ)10,0f0dλ=0.\eqref{Pmisto1}f_{0}^{-}=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}\frac{(\mathscr{L}_{0,0}-\lambda)^{-1}}{\lambda}\dot{\mathscr{L}}_{0,0}(\mathscr{L}_{0,0}-\lambda)^{-1}\mathscr{L}_{0,0}^{\prime}f_{0}^{-}\mathrm{d}\lambda=0\,.

Next by (A.11) and ˙0,0f0=0\dot{\mathscr{L}}_{0,0}f_{0}^{-}=0 we get (A.5b)f0=0\eqref{Pmisto2}f_{0}^{-}=0. Finally by (A.11) and (A.8)

(A.5c)f0=12πiΓ(0,0λ)11λ[0ip1[1]cos(x)]dλ\displaystyle\eqref{Pmisto3}f_{0}^{-}=-\frac{1}{2\pi\mathrm{i}\,}\oint_{\Gamma}(\mathscr{L}_{0,0}-\lambda)^{-1}\frac{1}{\lambda}\begin{bmatrix}0\\ \mathrm{i}\,p_{1}^{[1]}\cos(x)\end{bmatrix}\mathrm{d}\lambda

has zero average since (0,0λ)1(\mathscr{L}_{0,0}-\lambda)^{-1} is a Fourier multiplier (and thus preserves average absence). ∎

This completes the proof of Lemma 4.2.

Appendix B Expansion of the Stokes waves in finite depth

In this Appendix we provide the expansions (2.6)-(2.7), (2.15), (2.20)-(2.23).
Proof of (2.6)-(2.7). Writing

ηϵ(x)=ϵη1(x)+ϵ2η2(x)+𝒪(ϵ3),ψϵ(x)=ϵψ1(x)+ϵ2ψ2(x)+𝒪(ϵ3),cϵ=𝚌𝚑+ϵc1+ϵ2c2+𝒪(ϵ3),\begin{aligned} &\eta_{\epsilon}(x)=\epsilon\eta_{1}(x)+\epsilon^{2}\eta_{2}(x)+\mathcal{O}(\epsilon^{3})\,,\\ &\psi_{\epsilon}(x)=\epsilon\psi_{1}(x)+\epsilon^{2}\psi_{2}(x)+\mathcal{O}(\epsilon^{3})\,,\end{aligned}\qquad\quad c_{\epsilon}={\mathtt{c}}_{\mathtt{h}}+\epsilon c_{1}+\epsilon^{2}c_{2}+\mathcal{O}(\epsilon^{3})\,, (B.1)

where ηi\eta_{i} is even(x)even(x) and ψi\psi_{i} is odd(x)odd(x) for i=1,2i=1,2, we solve order by order in ϵ\epsilon the equations (2.5), that we rewrite as

{cψx+η+ψx22ηx22(1+ηx2)(cψx)2=0cηx+G(η)ψ=0,\begin{cases}-c\,\psi_{x}+\eta+\dfrac{\psi_{x}^{2}}{2}-\dfrac{\eta_{x}^{2}}{2(1+\eta_{x}^{2})}(c-\psi_{x})^{2}=0\\ c\,\eta_{x}+G(\eta)\psi=0\,,\end{cases} (B.2)

having substituted G(η)ψG(\eta)\psi with cηx-c\,\eta_{x} in the first equation. We expand the Dirichlet-Neumann operator G(η)=G0+G1(η)+G2(η)+𝒪(η3)G(\eta)=G_{0}+G_{1}(\eta)+G_{2}(\eta)+\mathcal{O}(\eta^{3}) where, according to [13][formula (2.14)],

G0\displaystyle G_{0} :=Dtanh(𝚑D)=|D|tanh(𝚑|D|),\displaystyle:=D\tanh(\mathtt{h}D)=|D|\tanh(\mathtt{h}|D|)\,, (B.3)
G1(η)\displaystyle G_{1}(\eta) :=D(ηtanh(𝚑D)ηtanh(𝚑D))D=xηx|D|tanh(𝚑|D|)η|D|tanh(𝚑|D|),\displaystyle:=D\big{(}\eta-\tanh(\mathtt{h}D)\eta\tanh(\mathtt{h}D)\big{)}D=-\partial_{x}\eta\partial_{x}-|D|\tanh(\mathtt{h}|D|)\eta|D|\tanh(\mathtt{h}|D|),
G2(η)\displaystyle G_{2}(\eta) :=12D(Dη2tanh(𝚑D)+tanh(𝚑D)η2D2tanh(𝚑D)ηDtanh(𝚑D)ηtanh(𝚑D))D.\displaystyle:=-\frac{1}{2}D\Big{(}D{\eta}^{2}\tanh(\mathtt{h}D)+\tanh(\mathtt{h}D){\eta}^{2}D-2\tanh(\mathtt{h}D)\eta D\tanh(\mathtt{h}D)\eta\tanh(\mathtt{h}D)\Big{)}D\,.

First order in ϵ\epsilon. Substituting in (B.2) the expansions in (B.1), we get the linear system

{𝚌𝚑(ψ1)x+η1=0𝚌𝚑(η1)x+G0ψ1=0,i.e.[η1ψ1]Ker 0 with 0:=[1𝚌𝚑x𝚌𝚑xG0],\left\{\begin{matrix}-{\mathtt{c}}_{\mathtt{h}}(\psi_{1})_{x}+\eta_{1}=0\\ {\mathtt{c}}_{\mathtt{h}}(\eta_{1})_{x}+G_{0}\psi_{1}=0\,,\end{matrix}\right.\quad\text{i.e.}\,\begin{bmatrix}\eta_{1}\\ \psi_{1}\end{bmatrix}\in\text{Ker }\mathcal{B}_{0}\text{ with }\mathcal{B}_{0}:=\begin{bmatrix}1&-{\mathtt{c}}_{\mathtt{h}}\partial_{x}\\ {\mathtt{c}}_{\mathtt{h}}\partial_{x}&G_{0}\end{bmatrix}, (B.4)

where η1\eta_{1} is even(x)even(x) and ψ1\psi_{1} is odd(x)odd(x).

Lemma B.1.

The kernel of the linear operator 0\mathcal{B}_{0} in (B.4) is

Ker 0=span{[cos(x)𝚌𝚑1sin(x)]}.\text{Ker }\mathcal{B}_{0}=\text{span}\,\Big{\{}\begin{bmatrix}\cos(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x)\end{bmatrix}\Big{\}}. (B.5)
Proof.

The action of 0\mathcal{B}_{0} on each subspace span{[cos(kx)0],[0sin(kx)]}\footnotesize{\,\Big{\{}\begin{bmatrix}\cos(kx)\\ 0\end{bmatrix},\begin{bmatrix}0\\ \sin(kx)\end{bmatrix}\Big{\}}}, kk\in\mathbb{N}, is represented by the 2×22\times 2 matrix [1𝚌𝚑k𝚌𝚑kktanh(𝚑k)]\footnotesize{\begin{bmatrix}1&-{\mathtt{c}}_{\mathtt{h}}k\\ -{\mathtt{c}}_{\mathtt{h}}k&k\tanh(\mathtt{h}k)\end{bmatrix}}. Its determinant ktanh(𝚑k)𝚌𝚑2k2=k2(tanh(𝚑k)ktanh(𝚑))k\tanh(\mathtt{h}k)-{\mathtt{c}}_{\mathtt{h}}^{2}k^{2}=k^{2}\Big{(}\frac{\tanh(\mathtt{h}k)}{k}-\tanh(\mathtt{h})\Big{)} vanishes if and only if k=1k=1. Indeed the function xtanh(𝚑x)xx\mapsto\frac{\tanh(\mathtt{h}x)}{x} is monotonically decreasing for x>0x>0, since its derivative 2x𝚑sinh(2𝚑x)2cosh2(𝚑x)x2\frac{2x\mathtt{h}-\sinh(2\mathtt{h}x)}{2\cosh^{2}(\mathtt{h}x)x^{2}} is negative for x>0x>0. For k=1k=1 we obtain the kernel of 0\mathcal{B}_{0} given in (B.5). For k=0k=0 it has no kernel since ψ1(x)\psi_{1}(x) is odd. ∎

We set η1(x):=cos(x)\eta_{1}(x):=\cos(x), ψ1(x):=𝚌𝚑1sin(x)\psi_{1}(x):={\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x) in agreement with (2.6).
Second order in ϵ\epsilon. By (B.2), and since 𝚌𝚑2(η1)x2=(G0ψ1)2{\mathtt{c}}_{\mathtt{h}}^{2}(\eta_{1})_{x}^{2}=(G_{0}\psi_{1})^{2}, we get the linear system

0[η2ψ2]=[c1(ψ1)x12(ψ1)x2+12(G0ψ1)2c1(η1)xG1(η1)ψ1],\mathcal{B}_{0}\begin{bmatrix}\eta_{2}\\ \psi_{2}\end{bmatrix}=\begin{bmatrix}c_{1}(\psi_{1})_{x}-\frac{1}{2}(\psi_{1})_{x}^{2}+\frac{1}{2}(G_{0}\psi_{1})^{2}\\ -c_{1}(\eta_{1})_{x}-G_{1}(\eta_{1})\psi_{1}\end{bmatrix}\,, (B.6)

where 0\mathcal{B}_{0} is the self-adjoint operator in (B.4). System (B.6) admits a solution if and only if its right-hand term is orthogonal to the Kernel of 0\mathcal{B}_{0} in (B.5), namely

([c1(ψ1)x12(ψ1)x2+12(G0ψ1)2c1(η1)xG1(η1)ψ1],[cos(x)𝚌𝚑1sin(x)])=0.\Big{(}\begin{bmatrix}c_{1}(\psi_{1})_{x}-\frac{1}{2}(\psi_{1})_{x}^{2}+\frac{1}{2}(G_{0}\psi_{1})^{2}\\ -c_{1}(\eta_{1})_{x}-G_{1}(\eta_{1})\psi_{1}\end{bmatrix}\;,\;\begin{bmatrix}\cos(x)\\ {\mathtt{c}}_{\mathtt{h}}^{-1}\sin(x)\end{bmatrix}\Big{)}=0\,. (B.7)

In view of the first order expansion (2.6), (B.3) and the identity tanh(2𝚑)=2𝚌𝚑21+𝚌𝚑4\tanh(2\mathtt{h})=\displaystyle{\frac{2{\mathtt{c}}_{\mathtt{h}}^{2}}{1+{\mathtt{c}}_{\mathtt{h}}^{4}}}, it results [G0ψ1](x)=𝚌𝚑sin(x)[G_{0}\psi_{1}](x)={\mathtt{c}}_{\mathtt{h}}\sin(x), [G1(η1)ψ1](x)=1𝚌𝚑4𝚌𝚑(1+𝚌𝚑4)sin(2x)\big{[}G_{1}(\eta_{1})\psi_{1}\big{]}(x)=\frac{1-{\mathtt{c}}_{\mathtt{h}}^{4}}{{\mathtt{c}}_{\mathtt{h}}(1+{\mathtt{c}}_{\mathtt{h}}^{4})}\sin(2x) so that (B.7) implies c1=0c_{1}=0, in agrement with (2.6). Equation (B.6) reduces to

[1𝚌𝚑x𝚌𝚑xG0][η2ψ2]=[14(𝚌𝚑2𝚌𝚑2)14(𝚌𝚑2+𝚌𝚑2)cos(2x)1𝚌𝚑4𝚌𝚑(1+𝚌𝚑4)sin(2x)].\displaystyle\begin{bmatrix}1&-{\mathtt{c}}_{\mathtt{h}}\partial_{x}\\ {\mathtt{c}}_{\mathtt{h}}\partial_{x}&G_{0}\end{bmatrix}\begin{bmatrix}\eta_{2}\\ \psi_{2}\end{bmatrix}=\begin{bmatrix}-\frac{1}{4}({\mathtt{c}}_{\mathtt{h}}^{-2}-{\mathtt{c}}_{\mathtt{h}}^{2})-\frac{1}{4}({\mathtt{c}}_{\mathtt{h}}^{-2}+{\mathtt{c}}_{\mathtt{h}}^{2})\cos(2x)\\ -\frac{1-{\mathtt{c}}_{\mathtt{h}}^{4}}{{\mathtt{c}}_{\mathtt{h}}(1+{\mathtt{c}}_{\mathtt{h}}^{4})}\sin(2x)\end{bmatrix}. (B.8)

Setting η2=η2[0]+η2[2]cos(2x)\eta_{2}=\eta_{2}^{[0]}+\eta_{2}^{[2]}\cos(2x) and ψ2=ψ2[2]sin(2x)\psi_{2}=\psi_{2}^{[2]}\sin(2x), system (B.8) amounts to

{η2[0]+(η2[2]2𝚌𝚑ψ2[2])cos(2x)=14(𝚌𝚑2𝚌𝚑2)14(𝚌𝚑2+𝚌𝚑2)cos(2x)(2𝚌𝚑η2[2]+2ψ2[2]tanh(2𝚑))sin(2x)=1𝚌𝚑4𝚌𝚑(1+𝚌𝚑4)sin(2x),\displaystyle\left\{\begin{matrix}\eta_{2}^{[0]}+\big{(}\eta_{2}^{[2]}-2{\mathtt{c}}_{\mathtt{h}}\psi_{2}^{[2]}\big{)}\cos(2x)=-\frac{1}{4}\left({\mathtt{c}}_{\mathtt{h}}^{-2}-{\mathtt{c}}_{\mathtt{h}}^{2}\right)-\frac{1}{4}\left({\mathtt{c}}_{\mathtt{h}}^{-2}+{\mathtt{c}}_{\mathtt{h}}^{2}\right)\cos(2x)\\ (-2{\mathtt{c}}_{\mathtt{h}}\eta_{2}^{[2]}+2\psi_{2}^{[2]}\tanh(2\mathtt{h}))\sin(2x)=-\frac{1-{\mathtt{c}}_{\mathtt{h}}^{4}}{{\mathtt{c}}_{\mathtt{h}}(1+{\mathtt{c}}_{\mathtt{h}}^{4})}\sin(2x)\,,\end{matrix}\right.

which leads to the expansions of η2[0]\eta_{2}^{[0]}, η2[2]\eta_{2}^{[2]}, ψ2[2]\psi_{2}^{[2]} given in (2.6)-(2.7).

Third order in ϵ\epsilon. It remains to determine c2c_{2} in (2.8). We get the linear system

0[η3ψ3]=[c2(ψ1)x(ψ1)x(ψ2)x(η1)x2(ψ1)x𝚌𝚑+(η1)x(η2)x𝚌𝚑2c2(η1)xG1(η1)ψ2G1(η2)ψ1G2(η1)ψ1].\mathcal{B}_{0}\begin{bmatrix}\eta_{3}\\ \psi_{3}\end{bmatrix}=\begin{bmatrix}c_{2}(\psi_{1})_{x}-(\psi_{1})_{x}(\psi_{2})_{x}-(\eta_{1})_{x}^{2}(\psi_{1})_{x}{\mathtt{c}}_{\mathtt{h}}+(\eta_{1})_{x}(\eta_{2})_{x}{\mathtt{c}}_{\mathtt{h}}^{2}\\ -c_{2}(\eta_{1})_{x}-G_{1}(\eta_{1})\psi_{2}-G_{1}(\eta_{2})\psi_{1}-G_{2}(\eta_{1})\psi_{1}\end{bmatrix}\,. (B.9)

System (B.9) has a solution if and only if the right hand side is orthogonal to the Kernel of 0\mathcal{B}_{0} given in (B.5). This condition determines uniquely c2c_{2}. Denoting Π1\Pi_{1} the L2L^{2}-orthogonal projector on span{cos(x),sin(x)}\,\{\cos(x),\sin(x)\}, it results

c2(ψ1)x=c2𝚌𝚑1cos(x),c2(η1)x=c2sin(x),Π1[(ψ1)x(ψ2)x]=ψ2[2]𝚌𝚑1cos(x),\displaystyle c_{2}(\psi_{1})_{x}=c_{2}{\mathtt{c}}_{\mathtt{h}}^{-1}\cos(x)\,,\quad c_{2}(\eta_{1})_{x}=-c_{2}\sin(x)\,,\quad\Pi_{1}[(\psi_{1})_{x}(\psi_{2})_{x}]=\psi_{2}^{[2]}{\mathtt{c}}_{\mathtt{h}}^{-1}\cos(x)\,,
Π1[𝚌𝚑(η1)x2(ψ1)x]=14cos(x),Π1[𝚌𝚑2(η1)x(η2)x]=η2[2]𝚌𝚑2cos(x),\displaystyle\Pi_{1}[{\mathtt{c}}_{\mathtt{h}}(\eta_{1})_{x}^{2}(\psi_{1})_{x}]=\tfrac{1}{4}\cos(x)\,,\quad\Pi_{1}[{\mathtt{c}}_{\mathtt{h}}^{2}(\eta_{1})_{x}(\eta_{2})_{x}]=\eta_{2}^{[2]}{\mathtt{c}}_{\mathtt{h}}^{2}\cos(x)\,,

and, in view of (B.3), and (2.6), (2.7),

Π1[G1(η1)ψ2]\displaystyle\Pi_{1}[G_{1}(\eta_{1})\psi_{2}] =ψ2[2]1𝚌𝚑41+𝚌𝚑4sin(x),Π1[G2(η1)ψ1]=𝚌𝚑3𝚌𝚑414(1+𝚌𝚑4)sin(x),\displaystyle=\psi_{2}^{[2]}\frac{1-{\mathtt{c}}_{\mathtt{h}}^{4}}{1+{\mathtt{c}}_{\mathtt{h}}^{4}}\sin(x)\,,\quad\Pi_{1}[G_{2}(\eta_{1})\psi_{1}]={\mathtt{c}}_{\mathtt{h}}\frac{3{\mathtt{c}}_{\mathtt{h}}^{4}-1}{4(1+{\mathtt{c}}_{\mathtt{h}}^{4})}\sin(x)\,,
Π1[G1(η2)ψ1]\displaystyle\Pi_{1}[G_{1}(\eta_{2})\psi_{1}] =𝚌𝚑1(η2[0](1𝚌𝚑4)+12η2[2](1+𝚌𝚑4))sin(x).\displaystyle={\mathtt{c}}_{\mathtt{h}}^{-1}\Big{(}\eta_{2}^{[0]}(1-{\mathtt{c}}_{\mathtt{h}}^{4})+\tfrac{1}{2}\eta_{2}^{[2]}(1+{\mathtt{c}}_{\mathtt{h}}^{4})\Big{)}\sin(x)\,.

Therefore the orthogonality condition proves (2.8).

Proof of (2.15). We expand the function 𝔭(x)=ϵ𝔭1(x)+ϵ2𝔭2(x)+𝒪(ϵ3)\mathfrak{p}(x)=\epsilon\mathfrak{p}_{1}(x)+\epsilon^{2}\mathfrak{p}_{2}(x)+\mathcal{O}(\epsilon^{3}) defined by the fixed point equation (2.14). We first note that the constant 𝚏ϵ=𝒪(ϵ2)\mathtt{f}_{\epsilon}=\mathcal{O}(\epsilon^{2}) because η1(x)=cos(x)\eta_{1}(x)=\cos(x) has zero average. Then 𝔭(x)=tanh(𝚑|D|)[ϵη1+ϵ2(η2+(η1)x𝔭1)+𝒪(ϵ3)]\mathfrak{p}(x)=\frac{\mathcal{H}}{\tanh(\mathtt{h}|D|)}\big{[}\epsilon\eta_{1}+\epsilon^{2}\big{(}\eta_{2}+(\eta_{1})_{x}\mathfrak{p}_{1}\big{)}+\mathcal{O}(\epsilon^{3})\big{]}, and, using that cos(kx)=sin(kx)\mathcal{H}\cos(kx)=\sin(kx), for any kk\in\mathbb{N}, we get

𝔭1(x)\displaystyle\mathfrak{p}_{1}(x) =tanh(𝚑|D|)cos(x)=𝚌𝚑2sin(x),\displaystyle=\frac{\mathcal{H}}{\tanh(\mathtt{h}|D|)}\cos(x)={\mathtt{c}}_{\mathtt{h}}^{-2}\sin(x)\,, (B.10)
𝔭2(x)\displaystyle\mathfrak{p}_{2}(x) =tanh(𝚑|D|)((η1)x𝔭1+η2)=(1+𝚌𝚑4)(𝚌𝚑4+3)8𝚌𝚑8sin(2x).\displaystyle=\frac{\mathcal{H}}{\tanh(\mathtt{h}|D|)}((\eta_{1})_{x}\mathfrak{p}_{1}+\eta_{2})=\frac{(1+{\mathtt{c}}_{\mathtt{h}}^{4})({\mathtt{c}}_{\mathtt{h}}^{4}+3)}{8{\mathtt{c}}_{\mathtt{h}}^{8}}\sin(2x)\,. (B.11)

Finally

𝚏ϵ\displaystyle\mathtt{f}_{\epsilon} =ϵ22π𝕋(η2+(η1)x𝔭1)dx+𝒪(ϵ3)=ϵ2(η2[0]12𝚌𝚑2)+𝒪(ϵ3)=(2.7)ϵ2𝚌𝚑434𝚌𝚑2+𝒪(ϵ3).\displaystyle=\frac{\epsilon^{2}}{2\pi}\int_{\mathbb{T}}\big{(}\eta_{2}+(\eta_{1})_{x}\mathfrak{p}_{1}\big{)}\mathrm{d}x+\mathcal{O}(\epsilon^{3})=\epsilon^{2}\big{(}\eta_{2}^{[0]}-\tfrac{1}{2}{\mathtt{c}}_{\mathtt{h}}^{-2}\big{)}+\mathcal{O}(\epsilon^{3})\stackrel{{\scriptstyle\eqref{expcoef}}}{{=}}\epsilon^{2}\frac{{\mathtt{c}}_{\mathtt{h}}^{4}-3}{4{\mathtt{c}}_{\mathtt{h}}^{2}}+\mathcal{O}(\epsilon^{3})\,.

The expansion (2.15) is proved.
Proof of Lemma 2.2. In view of (2.6)-(2.7), the expansions of the functions BB, VV in (2.10) are

B=:ϵB1(x)+ϵ2B2(x)+𝒪(ϵ3)=ϵ𝚌𝚑sin(x)+ϵ232𝚌𝚑42𝚌𝚑5sin(2x)+𝒪(ϵ3)\displaystyle B=:\epsilon B_{1}(x)+\epsilon^{2}B_{2}(x)+\mathcal{O}(\epsilon^{3})=\epsilon{\mathtt{c}}_{\mathtt{h}}\sin(x)+\epsilon^{2}\frac{3-2{\mathtt{c}}_{\mathtt{h}}^{4}}{2{\mathtt{c}}_{\mathtt{h}}^{5}}\sin(2x)+\mathcal{O}(\epsilon^{3}) (B.12)

and

V=:ϵV1(x)+ϵ2V2(x)+𝒪(ϵ3)=ϵ𝚌𝚑1cos(x)+ϵ2[𝚌𝚑2+3𝚌𝚑84𝚌𝚑7cos(2x)]+𝒪(ϵ3).\displaystyle V=:\epsilon V_{1}(x)+\epsilon^{2}V_{2}(x)+\mathcal{O}(\epsilon^{3})=\epsilon{\mathtt{c}}_{\mathtt{h}}^{-1}\cos(x)+\epsilon^{2}\Big{[}\frac{{\mathtt{c}}_{\mathtt{h}}}{2}+\frac{3-{\mathtt{c}}_{\mathtt{h}}^{8}}{4{\mathtt{c}}_{\mathtt{h}}^{7}}\cos(2x)\Big{]}+\mathcal{O}(\epsilon^{3})\,. (B.13)

In view of (2.18), denoting derivatives w.r.t xx with an apex and suppressing dependence on xx when trivial, we have

𝚌𝚑+pϵ(x)\displaystyle{\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x) =(𝚌𝚑+ϵ2c2V(x)V(x)𝔭(x)+𝒪(ϵ3))(1𝔭(x)+(𝔭(x))2+𝒪(ϵ3))\displaystyle=({\mathtt{c}}_{\mathtt{h}}+\epsilon^{2}c_{2}-V(x)-V^{\prime}(x)\mathfrak{p}(x)+\mathcal{O}(\epsilon^{3}))(1-\mathfrak{p}^{\prime}(x)+(\mathfrak{p}^{\prime}(x))^{2}+\mathcal{O}(\epsilon^{3}))
=𝚌𝚑+ϵ(V1𝚌𝚑𝔭1)=:p1+ϵ2(c2+V1𝔭1V2V1𝔭1𝚌𝚑𝔭2+𝚌𝚑(𝔭1)2)=:p2+𝒪(ϵ3).\displaystyle={\mathtt{c}}_{\mathtt{h}}+\epsilon\underbrace{(-V_{1}-{\mathtt{c}}_{\mathtt{h}}\mathfrak{p}_{1}^{\prime})}_{=:p_{1}}+\epsilon^{2}\underbrace{\big{(}c_{2}+V_{1}\mathfrak{p}_{1}^{\prime}-V_{2}-V_{1}^{\prime}\mathfrak{p}_{1}-{\mathtt{c}}_{\mathtt{h}}\mathfrak{p}_{2}^{\prime}+{\mathtt{c}}_{\mathtt{h}}(\mathfrak{p}_{1}^{\prime})^{2}\big{)}}_{=:p_{2}}+\,\mathcal{O}(\epsilon^{3})\,. (B.14)

Similarly by (2.18)

1+aϵ(\displaystyle 1+a_{\epsilon}( x):=11+𝔭x(x)(𝚌𝚑+pϵ(x))Bx(x+𝔭(x))\displaystyle x):=\frac{1}{1+\mathfrak{p}_{x}(x)}-({\mathtt{c}}_{\mathtt{h}}+p_{\epsilon}(x))B_{x}(x+\mathfrak{p}(x))
=\displaystyle= 1+ϵ(𝔭1𝚌𝚑B1)=:a1+ϵ2((𝔭1)2𝔭2𝚌𝚑B2𝚌𝚑B1′′𝔭1(x)+B1V1+𝚌𝚑B1𝔭1)=:a2+𝒪(ϵ3).\displaystyle 1+\epsilon\underbrace{\big{(}-\mathfrak{p}_{1}^{\prime}-{\mathtt{c}}_{\mathtt{h}}B_{1}^{\prime}\big{)}}_{=:a_{1}}+\epsilon^{2}\underbrace{\big{(}(\mathfrak{p}_{1}^{\prime})^{2}-\mathfrak{p}_{2}^{\prime}-{\mathtt{c}}_{\mathtt{h}}B_{2}^{\prime}-{\mathtt{c}}_{\mathtt{h}}B_{1}^{\prime\prime}\mathfrak{p}_{1}(x)+B_{1}^{\prime}V_{1}+{\mathtt{c}}_{\mathtt{h}}B_{1}^{\prime}\mathfrak{p}_{1}^{\prime}\big{)}}_{=:a_{2}}+\mathcal{O}(\epsilon^{3})\,. (B.15)

By (B.13), (B.10), (2.6), (B.11), (B.12) we deduce that the functions p1p_{1}, p2p_{2}, a1a_{1}, a2a_{2} in (B.14) and (B.15) have an expansion as in (2.20)-(2.23).∎

References

  • [1] P. Baldi, M. Berti, E. Haus and R. Montalto, Time quasi-periodic gravity water waves in finite depth. Inventiones Math. 214 (2): 739–911, 2018.
  • [2] T. Benjamin Instability of periodic wavetrains in nonlinear dispersive systems, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. Volume 299 Issue 1456, 1967.
  • [3] T. Benjamin and J. Feir. The disintegration of wave trains on deep water, Part 1. Theory. J. Fluid Mech. 27(3): 417-430, 1967.
  • [4] M. Berti, L. Franzoi and A. Maspero. Traveling quasi-periodic water waves with constant vorticity, Archive for Rational Mechanics, 240: 99–202, 2021.
  • [5] M. Berti, L. Franzoi and A. Maspero. Pure gravity traveling quasi-periodic water waves with constant vorticity, arXiv:2101.12006, 2021, to appear in Comm. Pure Appl. Math.
  • [6] M. Berti, A. Maspero and P. Ventura. Full description of Benjamin-Feir instability of Stokes waves in deep water, arXiv:2109.11852, 2021, to appear in Inventiones Math.
  • [7] M. Berti, A. Maspero and P. Ventura. Benjamin-Feir instability of Stokes waves, to appear on Rend. Lincei Mat. Appl., 2022.
  • [8] M. Berti, A. Maspero and P. Ventura. On the analyticity of the Dirichlet-Neumann operator and Stokes waves, arXiv:2201.04675, to appear on Rend. Lincei Mat. Appl., 2022.
  • [9] T. Bridges and A. Mielke. A proof of the Benjamin-Feir instability. Arch. Rational Mech. Anal. 133(2): 145–198, 1995.
  • [10] J. Bronski, V. Hur and M. Johnson. Modulational Instability in Equations of KdV Type. In: Tobisch E. (eds) New Approaches to Nonlinear Waves. Lecture Notes in Physics, vol 908. Springer, 2016.
  • [11] J. Bronski and M. Johnson. The modulational instability for a generalized Korteweg-de Vries equation. Arch. Ration. Mech. Anal. 197(2): 357–400, 2010.
  • [12] G. Chen and Q. Su. Nonlinear modulational instabililty of the Stokes waves in 2d full water waves. arXiv:2012.15071.
  • [13] W. Craig and C. Sulem. Numerical simulation of gravity waves. J. Comput. Phys., 108(1): 73–83, 1993.
  • [14] W. Craig, P. Guyenne and H. Kalisch. Hamiltonian long-wave expansions for free surfaces and interfaces. Comm. Pure Appl. Math. 58, no. 12, 1587–1641, 2005.
  • [15] B. Deconinck and K. Oliveras. The instability of periodic surface gravity waves. J. Fluid Mech., 675: 141–167, 2011.
  • [16] R. Feola and F. Giuliani. Quasi-periodic traveling waves on an infinitely deep fluid under gravity. arXiv:2005.08280, to appear on Memoires American Mathematical Society.
  • [17] T. Gallay and M. Haragus. Stability of small periodic waves for the nonlinear Schrödinger equation. J. Differential Equations, 234: 544–581, 2007.
  • [18] M.A. Garrido, R. K. Grande, M. Kurianski and G. Staffilani, Large deviations principle for the cubic NLS equation, https://arxiv.org/abs/2110.15748, to appear in Comm. Pure Appl. Math.
  • [19] M. Haragus and T. Kapitula. On the spectra of periodic waves for infinite-dimensional Hamiltonian systems. Phys. D, 237: 2649–2671, 2008.
  • [20] V. Hur. No solitary waves exist on 2D deep water. Nonlinearity 25, no. 12, 3301–3312, 2012.
  • [21] V. Hur and M. Johnson. Modulational instability in the Whitham equation for water waves. Stud. Appl. Math. 134(1): 120–143, 2015.
  • [22] V. Hur and A. Pandey. Modulational instability in nonlinear nonlocal equations of regularized long wave type. Phys. D, 325: 98–112, 2016.
  • [23] V. Hur and Z. Yang. Unstable Stokes waves. arXiv:2010.10766.
  • [24] J. Jin, S. Liao and Z. Lin. Nonlinear modulational instability of dispersive PDE models. Arch. Ration. Mech. Anal. 231(3): 1487-–1530, 2019.
  • [25] M. Johnson. Stability of small periodic waves in fractional KdV type equations. SIAM J. Math. Anal. 45: 2529–3228, 2013.
  • [26] T. Kato. Perturbation theory for linear operators. Die Grundlehren der mathematischen Wissenschaften, Band 132 Springer-Verlag New York, Inc., New York, 1966.
  • [27] M. Ifrim and D. Tataru. No solitary waves in 2D gravity and capillary waves in deep water. Nonlinearity 33 5457, 2020.
  • [28] P. Janssen and M. Onorato, The Intermediate Water Depth Limit of the Zakharov Equation and Consequences for Wave Prediction, Journal of Physical Oceanography, Vol. 37, 2389-2400, 2007.
  • [29] M. Onorato and P. Suret. Twenty years of progresses in oceanic rogue waves: the role played by weakly nonlinear models. Nat Hazards 84, 541-548, 2016.
  • [30] K. Leisman, J. Bronski, M. Johnson, and R. Marangell. Stability of Traveling Wave Solutions of Nonlinear Dispersive Equations of NLS Type. Arch. Rational Mech. Anal., 240: 927-–969, 2021.
  • [31] T. Levi-Civita. Détermination rigoureuse des ondes permanentes d’ ampleur finie, Math. Ann. 93, 264-314, 1925.
  • [32] M. J. Lighthill, Contribution to the theory of waves in nonlinear dispersive systems, IMA Journal of Applied Mathematics, 1, 3, 269-306, 1965.
  • [33] A. O. Korotkevich, A. I. Dyachenko and V. E. Zakharov, Numerical simulation of surface waves instability on a homogeneous grid, Physica D: Nonlinear Phenomena, Volumes 321-322, 51-66, 2016.
  • [34] A. Nekrasov. On steady waves. Izv. Ivanovo-Voznesenk. Politekhn. 3, 1921.
  • [35] H. Nguyen and W. Strauss. Proof of modulational instability of Stokes waves in deep water. To appear in Comm. Pure Appl. Math., 2020.
  • [36] F. Rousset and N. Tzvetkov. Transverse instability of the line solitary water-waves. Inventiones Math. 184: 257-388, 2011.
  • [37] H. Segur, D. Henderson, J. Carter and J. Hammack. Stabilizing the Benjamin-Feir instability. J. Fluid Mech. 539: 229–271, 2005.
  • [38] G. Stokes. On the theory of oscillatory waves. Trans. Cambridge Phil. Soc. 8: 441–455, 1847.
  • [39] D. Struik. Détermination rigoureuse des ondes irrotationelles périodiques dans un canal á profondeur finie. Math. Ann. 95: 595–634, 1926.
  • [40] G.B. Whitham. A general approach to linear and nonlinear dispersive waves using a Lagrangian. J. Fluid Mech. Volume 22 pp. 273–283, 1965.
  • [41] G.B. Whitham. Non-linear dispersion of water waves. J. Fluid Mech, volume 26 part 2 pp. 399-412, 1967.
  • [42] G.B. Whitham. Linear and Nonlinear Waves. J. Wiley-Sons, New York, 1974.
  • [43] V. Zakharov. The instability of waves in nonlinear dispersive media. J. Exp.Teor.Phys. 24 (4), 740-744, 1967.
  • [44] V. Zakharov. Stability of periodic waves of finite amplitude on the surface of a deep fluid. Zhurnal Prikladnoi Mekhaniki i Teckhnicheskoi Fiziki 9(2): 86–94, 1969.
  • [45] V. Zakharov and V. Kharitonov. Instability of monochromatic waves on the surface of a liquid of arbitrary depth. J Appl Mech Tech Phys 11, 747-751, 1970.
  • [46] V. Zakharov and L. Ostrovsky. Modulation instability: the beginning. Phys. D, 238(5): 540–548, 2009.