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

Damping for fractional wave equations and applications to water waves

Thomas Alazard thomas.alazard@ens-paris-saclay.fr Université Paris-Saclay, ENS Paris-Saclay, CNRS, Centre Borelli UMR9010, avenue des Sciences, F-91190 Gif-sur-Yvette France Jeremy L. Marzuola marzuola@math.unc.edu Department of Mathematics, University of North Carolina, Chapel Hill, NC 27514  and  Jian Wang wangjian@email.unc.edu Department of Mathematics, University of North Carolina, Chapel Hill, NC 27514
Abstract.

Motivated by numerically modeling surface waves for inviscid Euler equations, we analyze linear models for damped water waves and establish decay properties for the energy for sufficiently regular initial configurations. Our findings give the explicit decay rates for the energy, but do not address reflection/transmission of waves at the interface of the damping. Still for a subset of the models considered, this represents the first result proving the decay of the energy of the surface wave models.

1. Introduction

Motivated by highly successful numerical methods for damping the surface water wave equations proposed in the work [CFGK05], we wish to establish a theory of absorbing boundary conditions/perfectly matched layers as an approach to a damped linear water wave models. Such methods are essential to ensure that one can numerically simulate long-time behaviors of wave-trains without boundary interference. While nonlinear damping mechanisms have been proposed using nonlinear properties of the water wave models in the work [Ala17, Ala18, ABHK18], implementation of such methods can be numerically very stiff since the nonlinear damping mechanisms involve many spatial derivatives of the underlying models. However, the proposed methods in the work [CFGK05] are extremely non-stiff, which we argue is strongly related to them arising due to linear mechanisms for damping. Such connections should be explored further to fully understand the efficacy of these existing methods. The connection between damped waves and absorbing boundary conditions has long been understood for models with local differential operators, see for instance [Joh21, Nat13] and references therein, so we endeavor here to extend our understanding of damping effects the inherently non-local models that arise the water wave problem.

Let us very briefly recall a Hamiltonian formulation of the evolution of a fluid interface in the gravity-capillary water wave system subject to an external pressure. This equations can be written in terms of the surface height, denoted η(x,t)\eta(x,t), and the velocity potential of the fluid restricted to the surface, denoted ϕ(x,t)\phi(x,t). In [ACM+22], the authors derive a robust method for numerically solving the Euler equations in very general geometric setting using the coordinate equations

ηt=G(η)ϕ,ϕt=(ϕα/s0,α)V+(ϕ/n)U12|ϕ|2gη0(α,t)+τθαs0,α+Pext.\displaystyle\eta_{t}=G(\eta)\phi,\ \ \phi_{t}=(\phi_{\alpha}/s_{0,\alpha})V+(\partial\phi/\partial n)U-\frac{1}{2}|\nabla\phi|^{2}-g\,\eta_{0}(\alpha,t)+\tau\frac{\theta_{\alpha}}{s_{0,\alpha}}+P_{ext}.

Here G(η)G(\eta) is the un-normalized Dirichlet-to-Neumann map, θ\theta is the tangent angle of the surface, sαs_{\alpha} is an arc-length parameter, U,VU,V are the normal and tangential derivatives at the surface and PextP_{ext} is an external pressure term in which we can introduce damping or forcing on the equations. We will consider especially a form of damping introduced by Clamond-Fructus-Frue-Kristiansen in [CFGK05].

In the case of non-zero surface tension (τ>0\tau>0) one can use a prescribed PextP_{ext} to stabilize small waves similar to the work of Alazard et al, see [ABHK18]. When τ=0\tau=0, we can think of these conditions as numerical boundary conditions that absorb energy and allow for as little reflection as possible. Generically, one numerically solves the water waves problem on a periodic domain of length 2π2\pi and take ω[0,2π)\omega\subset[0,2\pi) a connected interval on which we will damp the fluid with corresponding indicator function χω\chi_{\omega}. We will consider here the damping properties of the numerically effective damping term,

Pext=x1(χωϕx),P_{ext}=\partial_{x}^{-1}(\chi_{\omega}\phi_{x}),

as proposed in [CFGK05], which we will denote as Linear H1/2H^{1/2} Damping of the water wave problem related to the order of regularity required to establish the model equation (1.1) from a full paradifferential diagonalization of the water wave equations, see [ABZ11].

Using the paradifferential formulation of the water waves developed in for instance [ABZ11], one observes the following leading order linear model for damped gravity water waves

t2u+|D|u+χtu=0.\partial_{t}^{2}u+|D|u+\chi\partial_{t}u=0. (1.1)

This model can be studied from the classical point of view of scattering theory and perfectly matched layers, though the non-local nature of the operator |D||D| means that many known techniques fail and more refined tools are required. To that end, we study (1.1) here using propagation estimates in the study of semiclassical scattering operators, which have been developed quite thoroughly in the recent book [DZ19] for operators of the form Δ+V-\Delta+V. However, the non-locality of operators of the form |D|+V|D|+V results in some important modifications that we illuminate here. Much of our analysis should be extendable to other non-local wave equation models with appropriate modifications. The well-posedness of a nonlinear model related to (1.1) in the setting of the water waves with surface-tension was established in the recent work of [Moo22], but the strength and speed of damping that arises from such a method is not clear. Here, we are able to prove the polynomial decay of the energy for the linear model.

As discussed in [Ala17], there is a long-standing connection between damped wave equations, absorbing boundary conditions in numerical analysis and the notions of so-called control and observability estimates for a given equation on the support of the damping function.111See for instance [Ala17] or [BZ19] for careful definitions of control/observability estimates if the reader is unfamiliar. While our approach here does not use such an estimates directly, some important surveys and results in this direction for a variety of models that have similar proof strategies include [ABBG+12, BLR92, BZ19, BZ04, Mac21, Phu07, RT75, Zua05, Zua07]. We also highlight a 1d1d specific version of absorbing boundary condition was introduced in [JKR14], though we point out that the model we consider here can be easily generalized to higher dimensional water wave models.

A related damping model is of the form

t2u+|D|u+|D|12(χ|D|12tu)=0.\partial_{t}^{2}u+|D|u+|D|^{\frac{1}{2}}(\chi|D|^{\frac{1}{2}}\partial_{t}u)=0. (1.2)

This results from a similar paradifferential diagonalization of a damping that is guaranteed to lead to nonlinear damping by consideration of the Hamiltonian energy for the water wave equations, see [Ala18]. The techniques we apply here can likely be applied to study damping of this form with appropriate modifications, in particular with respect to the required regularity of the initial data. However, for the sake of smoothness of exposition, we focus only on equation (1.1) in our analysis below.

1.1. Main results

Here, we study a linear model for the damped water wave equation explicitly framed on a periodic domain, where we are able to give quantitative estimates on the damping rates of (1.1). To be precise, let 𝕋:=/2π\mathbb{T}:={\mathbb{R}}/2\pi\mathbb{Z} be the circle and χL(𝕋)\chi\in L^{\infty}(\mathbb{T}) satisfy χ0\chi\geq 0. For s>0s>0, we define the fractional Laplacian operator as follows

|D|s:Hs(𝕋)L2(𝕋),|D|su(x):=n|n|su^(n)einx, for uHs(𝕋),u^(n):=12π02πeinxu(x)𝑑x.\begin{gathered}|D|^{s}\colon H^{s}(\mathbb{T})\to L^{2}(\mathbb{T}),\ |D|^{s}u(x):=\sum_{n\in\mathbb{Z}}|n|^{s}\widehat{u}(n)e^{inx},\\ \text{ for }u\in H^{s}(\mathbb{T}),\ \widehat{u}(n):=\frac{1}{2\pi}\int_{0}^{2\pi}e^{-inx}u(x)dx.\end{gathered} (1.3)

For (u0,u1)H12(𝕋)×L2(𝕋)(u_{0},u_{1})\in H^{\frac{1}{2}}(\mathbb{T})\times L^{2}(\mathbb{T}), we consider the damped fractional wave equation

(t2+χt+|D|)u(t,x)=0,u(0,x)=u0(x),tu(0,x)=u1(x).(\partial_{t}^{2}+\chi\partial_{t}+|D|)u(t,x)=0,\ u(0,x)=u_{0}(x),\ \partial_{t}u(0,x)=u_{1}(x). (1.4)

The energy of the solution to (1.4) is defined by

E(u,t):=𝕋(||D|12u(t,x)|2+|tu(t,x)|2)𝑑x.E(u,t):=\int_{\mathbb{T}}\left(||D|^{\frac{1}{2}}u(t,x)|^{2}+|\partial_{t}u(t,x)|^{2}\right)dx. (1.5)

For a localized damping function χ\chi (meaning that suppχ𝕋\operatorname{supp}\chi\neq\mathbb{T}), the geometric control condition fails. In this case, we show that for any (u0,u1)H1(𝕋)×H12(𝕋)(u_{0},u_{1})\in H^{1}(\mathbb{T})\times H^{\frac{1}{2}}(\mathbb{T}), the energy of the solution decays as 1/t21/t^{2}. By constructing quasi-modes, we show that this polynomial rate is sharp for localized damping.

In order to explore the possibility of larger decay rates, we now consider damping with finitely many zeros. In this case, the energy decay rates depend on the “switching on” behavior of the damping function χ\chi near zeros of χ\chi. To quantify this connection, we introduce the following definition:

Definition 1.1.

We say χC(𝕋)\chi\in C^{\infty}(\mathbb{T}), χ0\chi\geq 0 has finite degeneracy, if χ\chi has finitely many zeros xkx_{k}, 1kn1\leq k\leq n, and for each xkx_{k}, there exists Nk>0N_{k}>0, such that

χ()(xk)=0, 02Nk1,χ(2Nk)(xk)>0.\chi^{(\ell)}(x_{k})=0,\ 0\leq\ell\leq 2N_{k}-1,\ \chi^{(2N_{k})}(x_{k})>0. (1.6)

Given β(0,1]\beta\in(0,1], we denote by C0,β(𝕋)C^{0,\beta}(\mathbb{T}) the Hölder space of those continuous periodic functions u:𝕋u\colon\mathbb{T}\to\mathbb{R} such that

uC0,β=supx𝕋|u(x)|+supxy|u(x)u(y)||xy|<+.\left\lVert u\right\rVert_{C^{0,\beta}}=\sup_{x\in\mathbb{T}}\left\lvert u(x)\right\rvert+\sup_{x\neq y}\frac{\left\lvert u(x)-u(y)\right\rvert}{\left\lvert x-y\right\rvert}<+\infty.

We are now ready to state the main results on the energy decay rates.

Theorem 1.

Suppose χC0,β(𝕋)\chi\in C^{0,\beta}(\mathbb{T}) with β>12\beta>\frac{1}{2}, χ0\chi\geq 0, χ0\chi\neq 0. Then there exists C>0C>0 such that for any (u0,u1)H1(𝕋)×H12(𝕋)(u_{0},u_{1})\in H^{1}(\mathbb{T})\times H^{\frac{1}{2}}(\mathbb{T}), if uu solves (1.4), then

E(u,t)Ct2(u0H12+u1H122),t>0.E(u,t)\leq\frac{C}{t^{2}}\left(\|u_{0}\|_{H^{1}}^{2}+\|u_{1}\|_{H^{\frac{1}{2}}}^{2}\right),\ t>0. (1.7)

Moreover, if χC(𝕋)\chi\in C^{\infty}(\mathbb{T}) has finite degeneracy as in Definition 1.1 and let NN be the maximal NkN_{k} there, then for any γ>0\gamma>0, there exists C>0C>0 such that for (u0,u1)H1(𝕋)×H12(𝕋)(u_{0},u_{1})\in H^{1}(\mathbb{T})\times H^{\frac{1}{2}}(\mathbb{T}), we have

E(u,t)Ct2+1Nγ(u0H12+u1H122),t>0.E(u,t)\leq\frac{C}{t^{2+\frac{1}{N}-\gamma}}\left(\|u_{0}\|_{H^{1}}^{2}+\|u_{1}\|_{H^{\frac{1}{2}}}^{2}\right),\ t>0. (1.8)

Remarks. 1. The 1/t21/t^{2} energy decay rate for localized damping (see (1.7)) is sharp. This follows from the sharpness of the resolvent bound (1.12) (see Remark after Theorem 2) and the equivalence between resolvent bounds and polynomial decay rates proved by Anantharaman–Léautaud (see [AL14, Proposition 2.4], which is also stated in §5 in the current paper).

2. Suppose χCk2(𝕋)\chi\in C^{\frac{k}{2}}(\mathbb{T}) for k2k\geq 2. If (u0,u1)Hk+12×Hk2(u_{0},u_{1})\in H^{\frac{k+1}{2}}\times H^{\frac{k}{2}}, then the polynomial rates in (1.7), (1.8) can be improved to Ckt2k\frac{C_{k}}{t^{2k}} and Ckt(2+1N)kγ\frac{C_{k}}{t^{(2+\frac{1}{N})k-\gamma}}, respectively. See Remark in §5.

3. The first result in Theorem 1 can be generalized to tori 𝕋n\mathbb{T}^{n} of higher dimensions n2n\geq 2. Indeed, suppose χC(𝕋n)\chi\in C^{\infty}(\mathbb{T}^{n}), χ0\chi\geq 0, χ0\chi\neq 0, and

 for any (x,ξ)S𝕋n, there exists T>0, such that x+Tξ{χ>0},\text{ for any }(x,\xi)\in S^{*}\mathbb{T}^{n},\text{ there exists }T>0,\text{ such that }x+T\xi\in\{\chi>0\}, (1.9)

where S𝕋n𝕋n×𝕊n1S^{*}\mathbb{T}^{n}\simeq\mathbb{T}^{n}\times\mathbb{S}^{n-1} is the cosphere bundle of 𝕋n\mathbb{T}^{n}. Then there exists C>0C>0 such that for any (u0,u1)H1(𝕋n)×H12(𝕋n)(u_{0},u_{1})\in H^{1}(\mathbb{T}^{n})\times H^{\frac{1}{2}}(\mathbb{T}^{n}), we have

E(u,t)Ct2(u0H12+u1H122).E(u,t)\leq\frac{C}{t^{2}}\left(\|u_{0}\|_{H^{1}}^{2}+\|u_{1}\|_{H^{\frac{1}{2}}}^{2}\right).

The proof is the same as in the 1D case presented in this paper.

4. For comparison, we recall the usual damped wave equation on 𝕋n\mathbb{T}^{n}

(t2+χtΔ)u=0.(\partial_{t}^{2}+\chi\partial_{t}-\Delta)u=0. (1.10)

It is known that under the same dynamical condition (1.9), the energy of the solution to (1.10) decays exponentially (see for instance [Zwo12, Theorem 5.10]). A result [Phu07] for a damped wave equation on a bounded domain results in polynomial rates for damped wave equations without geometric control conditions provided an observability estimate holds and with very minimal regularity requirements on the damping function.

The energy decay rates for the damped equation (1.4) are closely related to the resolvent estimates (via semi-group theory or Fourier transform, see for instance [AL14, Zwo12]) of the stationary operator

P(τ):=|D|iτχτ2,τ.P(\tau):=|D|-i\tau\chi-\tau^{2},\quad\ \tau\in{\mathbb{C}}. (1.11)

Theorem 1 follows from the following resolvent bounds.

Theorem 2.

Suppose χC0,β(𝕋)\chi\in C^{0,\beta}(\mathbb{T}) with β>12\beta>\frac{1}{2}, χ0\chi\geq 0, χ0\chi\neq 0. Let P(τ)P(\tau) be as in (1.11). Then there exists C>0C>0, such that τ\tau\in{\mathbb{R}}, |τ|>C|\tau|>C implies

P(τ)1L2L2C.\|P(\tau)^{-1}\|_{L^{2}\to L^{2}}\leq C. (1.12)

Moreover, if χC(𝕋)\chi\in C^{\infty}(\mathbb{T}) has finite degeneracy as in Definition 1.1 and NN denotes the maximum of NkN_{k} there, then for any γ>0\gamma^{\prime}>0, there exists C>0C>0 such that τ\tau\in{\mathbb{R}}, |τ|>C|\tau|>C implies

P(τ)1L2L2C|τ|12N+1+γ.\|P(\tau)^{-1}\|_{L^{2}\to L^{2}}\leq C|\tau|^{-\frac{1}{2N+1}+\gamma^{\prime}}. (1.13)

Remark. Let us prove that (1.12) is optimal for cut-off function χ\chi which does not have full support on 𝕋\mathbb{T} that is, χ\chi is a localized damping function. To see this, let aC(𝕋)a\in C^{\infty}(\mathbb{T}) be such that a0a\neq 0 and suppasuppχ=\operatorname{supp}a\cap\operatorname{supp}\chi=\emptyset. For all integers k>0k>0, define uk(x):=a(x)eikxC(𝕋)u_{k}(x):=a(x)e^{ikx}\in C^{\infty}(\mathbb{T}). Since suppasuppχ=\operatorname{supp}a\cap\operatorname{supp}\chi=\emptyset, we have

P(k)uk(x)=[|D|,a]eikx.P(\sqrt{k})u_{k}(x)=[|D|,a]e^{ikx}.

We will establish in §2.1 that the commutator [|D|,a]Ψ0(𝕋)[|D|,a]\in\Psi^{0}(\mathbb{T}). This means that [|D|,a]:L2(𝕋)L2(𝕋)[|D|,a]:L^{2}(\mathbb{T})\to L^{2}(\mathbb{T}) is a bounded operator and hence there is C>0C>0 such that

P(k)ukL2CeikL2=(C/aL2)ukL2.\|P(\sqrt{k})u_{k}\|_{L^{2}}\leq C\|e^{ik\bullet}\|_{L^{2}}=(C/\|a\|_{L^{2}})\|u_{k}\|_{L^{2}}.

Therefore, there is no such δ>0\delta>0 and C(δ)>0C(\delta)>0, such that τ\tau\in{\mathbb{R}}, |τ|>C(δ)|\tau|>C(\delta) implies

P(τ)1L2L2C(δ)|τ|δ.\|P(\tau)^{-1}\|_{L^{2}\to L^{2}}\leq C(\delta)|\tau|^{-\delta}.

This shows that (1.12) is optimal for localized damping functions.

As we show in Lemma 5.2,

P(τ)1:L2(𝕋)H1(𝕋),τP(\tau)^{-1}:L^{2}(\mathbb{T})\to H^{1}(\mathbb{T}),\ \tau\in{\mathbb{C}}

is a meromorphic family of operators with finite rank poles. The poles of P(τ)1P(\tau)^{-1} are called resonances for P(τ)P(\tau). We denote the set of resonances by \mathscr{R}. Using Theorem 2 and Grushin problems, we give the following description of the distribution of resonances:

Theorem 3.

Suppose χC0,β(𝕋)\chi\in C^{0,\beta}(\mathbb{T}) with β>12\beta>\frac{1}{2}, χ0\chi\geq 0, χ0\chi\neq 0. Then

  1. 1.

    There exists C>0C>0, such that

    {|Rez|>C}{τ|Imz>C1|Reτ|1}.\mathscr{R}\cap\{|\operatorname{Re}z|>C\}\subset\left\{\tau\in{\mathbb{C}}\ \left|\ \operatorname{Im}z>-C^{-1}|\operatorname{Re}\tau|^{-1}\right.\right\}. (1.14)

    Moreover, if χC(𝕋)\chi\in C^{\infty}(\mathbb{T}) has finite degeneracy as in Definition 1.1, then for any γ′′>0\gamma^{\prime\prime}>0, there exists C=C(γ′′)>0C=C(\gamma^{\prime\prime})>0, such that

    {|Rez|>C}{τ|Imz>C1|Reτ|2N2N+1γ′′}.\mathscr{R}\cap\{|\operatorname{Re}z|>C\}\subset\{\tau\in{\mathbb{C}}\ |\ \operatorname{Im}z>-C^{-1}|\operatorname{Re}\tau|^{-\frac{2N}{2N+1}-\gamma^{\prime\prime}}\}. (1.15)
  2. 2.

    For ν>0\nu>0, we denote P(ν,τ):=|D|iντχτ2P(\nu,\tau):=|D|-i\nu\tau\chi-\tau^{2} and (ν)\mathscr{R}(\nu) the set of resonances of P(ν,τ)1P(\nu,\tau)^{-1}. Then for each kk\in\mathbb{Z}, k>0k>0, there exists νk>0\nu_{k}>0, an open neighborhood UkU_{k} of k\sqrt{k}\in{\mathbb{C}}, and τk,±C((0,νk);Uk)\tau_{k,\pm}\in C^{\infty}((0,\nu_{k});U_{k}), such that

    (ν)Uk={τk,+(ν),τk,(ν)},τk,±(ν)=kχ^(0)±|χ^(2k)|2iν+o(ν).\mathscr{R}(\nu)\cap U_{k}=\{\tau_{k,+}(\nu),\ \tau_{k,-}(\nu)\},\ \tau_{k,\pm}(\nu)=\sqrt{k}-\frac{\widehat{\chi}(0)\pm|\widehat{\chi}(2k)|}{2}i\nu+o(\nu). (1.16)

Remarks. 1. Despite the asymptotic expansion of the resonances in (1.16), Theorem 3 does not imply the existence of a resonance-free strip with constant width for P(τ)1P(\tau)^{-1}, because the expansions are not uniform in kk.

2. For the damped wave equation framed on a compact manifold,

t2u+χtuΔgu=0,\partial_{t}^{2}u+\chi\partial_{t}u-\Delta_{g}u=0,

where Δg\Delta_{g} is the usual Laplace-Beltrami operator, the distribution of resonances and corresponding energy decay rates have been studied in the works [MM82], [Sjö00] and [Ana10]. In [MM82], Markus–Matsaev established a Weyl law for the resonances in terms of counting how many resonances can exist at a given energy. In [Sjö00], Sjöstrand further proved that the imaginary parts of the resonances “concentrate” (in a suitable sense) on the half average (with respect to the Liouville measure on the cosphere bundle of the manifold) of the damping function. In [Ana10], Anantharaman proved a fractal Weyl law for the resonances and studied several inverse problems. It is an important topic of future work to see if analogous results hold for damped fractional wave equations of the form studied here.

1.2. Outline of Paper

In §2, we recall some properties of the semiclassical calculus for operators on the torus that we will require for our analysis. The propagation estimates required to prove the resolvent estimate and the resulting resolvent mapping properties are proven in §3. We prove a stronger resolvent estimate for damping functions that vanish to a given order at a finite number of points on 𝕋\mathbb{T} in §4. In §5, we give an overview of the proof from [AL14] (simplified in our particular setting) that the resolvent bounds proved are equivalent to energy decay bounds for the damped fractional wave equation. To give insight into the properties of the resolvent, in §6 we prove that the low energy resonances can be approximated well by a finite approximation that can be constructed explicitly using a Grushin problem. Finally, in §7 we provide some numerical simulations demonstrating various aspects of our theorems in practice, both for the exact linear fractional wave model, as well as for water wave models with Clamond Damping. This includes a means of approximating the low-energy resonances and comparing to the asymptotics in the previous section.

Acknowledgements. We would like to thank Jared Wunsch for many helpful discussions, and Ruoyu P. T. Wang for showing us useful references. J.L.M was supported in part by NSF Applied Math Grant DMS-1909035 and NSF Applied Math Grant DMS-2307384.

2. Semiclassical analysis on the circle

2.1. Semiclassical pseudodifferential operators

We consider the following symbol class Sk(T𝕋)S^{k}(T^{*}\mathbb{T})

Sk(T𝕋):={aC(T𝕋)||xαξβa(x,ξ)|Cαβξk|β|,Cαβ>0,α,β},S^{k}(T^{*}\mathbb{T}):=\left\{a\in C^{\infty}(T^{*}\mathbb{T})\ |\ |\partial_{x}^{\alpha}\partial_{\xi}^{\beta}a(x,\xi)|\leq C_{\alpha\beta}\langle\xi\rangle^{k-|\beta|},\ C_{\alpha\beta}>0,\ \forall\alpha,\beta\right\},

here T𝕋T^{*}\mathbb{T} is the cotangent bundle of 𝕋\mathbb{T}, T𝕋𝕋x×ξT^{*}\mathbb{T}\simeq\mathbb{T}_{x}\times{\mathbb{R}}_{\xi}, and ξ=1+ξ2\langle\xi\rangle=\sqrt{1+\xi^{2}}. With the best constants CαβC_{\alpha\beta} as semi-norms, the class SkS^{k} is a Fréchet space. For aSk(T𝕋)a\in S^{k}(T^{*}\mathbb{T}), which could depend on hh with semi-norms uniform in hh, we define its semiclassical quantization by

Oph(a)u(x):=eih(xy)ξa(h,x,ξ)u(y)𝑑y𝑑ξ.\operatorname{Op}_{h}(a)u(x):=\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}e^{\frac{i}{h}(x-y)\xi}a(h,x,\xi)u(y)dyd\xi. (2.1)

Let a^(k,ξ)\widehat{a}(k,\xi), u^(k)\widehat{u}(k) be the Fourier coefficients of a(,ξ)a(\bullet,\xi), uu as in (1.3). Then we have

Oph(a)u(x)=na^(h,n,h)u^()einx.\operatorname{Op}_{h}(a)u(x)=\sum_{n\in\mathbb{Z}}\sum_{\ell\in\mathbb{Z}}\widehat{a}(h,n-\ell,h\ell)\widehat{u}(\ell)e^{inx}. (2.2)

Similarly, we define its microlocal quantization by

Op(a)u(x):=ei(xy)ξa(h,x,ξ)u(y)𝑑y𝑑ξ=nZa^(h,n,)u^()einx.\begin{split}\operatorname{Op}(a)u(x):=\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}e^{i(x-y)\xi}a(h,x,\xi)u(y)dyd\xi=\sum_{n\in\mathbb{Z}}\sum_{\ell\in Z}\widehat{a}(h,n-\ell,\ell)\widehat{u}(\ell)e^{inx}.\end{split}

We also use the notations a(x,D)a(x,D), a(x,hD)a(x,hD) for the microlocal or semiclassical quantization of aa. Let Ψhk(𝕋)\Psi_{h}^{k}(\mathbb{T}) be the set of pseudodifferential operators that consists of semiclassical quantizations of all symbols in Sk(T𝕋)S^{k}(T^{*}\mathbb{T}). We define the semiclassical symbol map

σh:Ψhk(𝕋)Sk(T𝕋)/Sk1(T𝕋),A[a].\sigma_{h}:\Psi_{h}^{k}(\mathbb{T})\to S^{k}(T^{*}\mathbb{T})/S^{k-1}(T^{*}\mathbb{T}),\ A\mapsto[a].

Later we will identify [a][a] with aa if no ambiguity.

Remark. |D||D| is a microlocal operator with symbol a(ξ)S1(T𝕋)a(\xi)\in S^{1}(T^{*}\mathbb{T}) such that a(k)=|k|a(k)=|k|, kk\in\mathbb{Z}. However, h|D|h|D| is not a semiclassical operator, as its “semiclassical symbol” |ξ||\xi| is not a smooth function.

Consider a family of symbols a={a(h,x,ξ)| 0<h<h0}a=\{a(h,x,\xi)\,|\,0<h<h_{0}\} which is bounded in SkS^{k} and is CC^{\infty} in hh. Let AA be the semiclassical quantization of aa. The semiclassical wavefront set WFh(A)\operatorname{WF}_{h}(A) of AA is defined to be the essential support of aa, that is, (x0,ξ0)WFh(A)(x_{0},\xi_{0})\notin\operatorname{WF}_{h}(A) if and only if there exists a neighborhood UU of (x0,ξ0)(x_{0},\xi_{0}) in TT^{*}{\mathbb{R}}, such that for any α,β\alpha,\beta\in{\mathbb{N}},

xαξβa(h,x,ξ)=O(hNξN), for all (x,ξ)U,N.\partial_{x}^{\alpha}\partial_{\xi}^{\beta}a(h,x,\xi)=O(h^{N}\langle\xi\rangle^{-N}),\text{ for all }(x,\xi)\in U,\ N\in{\mathbb{N}}.

We also define the semiclassical elliptic set of AA by

ellh(A):={(x,ξ)|ξkσh(A)(x,ξ)0}.\operatorname{ell}_{h}(A):=\{(x,\xi)\ |\ \langle\xi\rangle^{-k}\sigma_{h}(A)(x,\xi)\neq 0\}.

We record the following formula for the symbol calculus of operator compositions: if aSk(T𝕋)a\in S^{k}(T^{*}\mathbb{T}), bS(T𝕋)b\in S^{\ell}(T^{*}\mathbb{T}), then

Oph(a)Oph(b)=Oph(a#b),a#bSk+(T𝕋),a#b=k=0N(ih)jj!ξja(x,ξ)xjb(x,ξ)+OSk+j(T𝕋)(hN+1),for allN.\begin{gathered}\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)=\operatorname{Op}_{h}(a\#b),\ a\#b\in S^{k+\ell}(T^{*}\mathbb{T}),\\ a\#b=\sum_{k=0}^{N}\frac{(-ih)^{j}}{j!}\partial_{\xi}^{j}a(x,\xi)\partial_{x}^{j}b(x,\xi)+O_{S^{k+\ell-j}(T^{*}\mathbb{T})}(h^{N+1}),\ \text{for all}\ N\in{\mathbb{N}}.\end{gathered}

In particular, we can compute the commutator of two pseudodifferential operators

[Oph(a),Oph(b)]hΨhk+1(𝕋),σh(h1[Oph(a),Oph(b)])=i{a,b}[\operatorname{Op}_{h}(a),\operatorname{Op}_{h}(b)]\in h\Psi^{k+\ell-1}_{h}(\mathbb{T}),\ \sigma_{h}(h^{-1}[\operatorname{Op}_{h}(a),\operatorname{Op}_{h}(b)])=-i\{a,b\} (2.3)

where {a,b}:=ξaxbxaξb\{a,b\}:=\partial_{\xi}a\partial_{x}b-\partial_{x}a\partial_{\xi}b.

2.2. Semiclassical Fourier multipliers

When the symbol aa in (2.2) does not depend on xx, we say Oph(a)\operatorname{Op}_{h}(a) is a semiclassical Fourier multiplier. In this section, we generalize the definition of semiclassical Fourier multipliers to bounded symbols and prove a commutator estimate for semiclassical Fourier multipliers and functions with Hölder continuity.

Given a bounded function a:a\colon\mathbb{R}\to\mathbb{R}, we define the semiclassical Fourier multiplier a(hD)a(hD) by

a(hD)u^(k)=a(hk)u^(k)\widehat{a(hD)u}(k)=a(hk)\widehat{u}(k)

for uL2(𝕋)u\in L^{2}(\mathbb{T}). We call aa the symbol of the semi-classical Fourier multiplier a(hD)a(hD).

Proposition 2.1.

Let aCb()a\in C^{\infty}_{b}(\mathbb{R}) be a smooth function, bounded together with all its derivatives, which in addition vanishes on a neighborhood of the origin. Consider two real numbers 0<α<β10<\alpha<\beta\leq 1. There exists a constant C>0C>0 such that, for all fC0,β(𝕋)f\in C^{0,\beta}(\mathbb{T}) and for all uL2(𝕋)u\in L^{2}(\mathbb{T}),

fa(hD)ua(hD)(fu)L2ChαfC0,βuL2.\left\lVert fa(hD)u-a(hD)(fu)\right\rVert_{L^{2}}\leq Ch^{\alpha}\left\lVert f\right\rVert_{C^{0,\beta}}\left\lVert u\right\rVert_{L^{2}}. (2.4)
Proof.

We introduce the symbols aha_{h} defined by ah(ξ)=a(hξ)a_{h}(\xi)=a(h\xi). The key point is to observe that {hαah0<h1}\left\{\,h^{-\alpha}a_{h}\mid 0<h\leq 1\,\right\} is a bounded family in SαS^{\alpha}. The wanted estimate (2.4) will then be a direct consequence of the following lemma.

Lemma 2.2.

Consider two real numbers mm and α\alpha such that 0<α<β10<\alpha<\beta\leq 1. For any bounded subset \mathcal{B} of SαS^{\alpha}, there exists a constant KK such that for all symbol qq\in\mathcal{B}, all fC0,β(𝕋)f\in C^{0,\beta}(\mathbb{T}), and all uL2(𝕋)u\in L^{2}(\mathbb{T}),

𝒬(fu)f𝒬uL2KfC0,β(𝕋)uL2,\left\lVert\mathcal{Q}(fu)-f\mathcal{Q}u\right\rVert_{L^{2}}\leq K\left\lVert f\right\rVert_{C^{0,\beta}(\mathbb{T})}\left\lVert u\right\rVert_{L^{2}}, (2.5)

where 𝒬\mathcal{Q} is the Fourier multiplier with symbol qq.

Proof.

To prove this result, it is convenient to use the paradifferential calculus of Bony [Bon81] and the Littlewood-Paley decomposition. We start by introducing some notations. Fix a function ΦC0()\Phi\in C^{\infty}_{0}(\mathbb{R}) with support in the interval [1,1][-1,1] and equal to 11 when |ξ|1/2|\xi|\leq 1/2. Then set ϕ(ξ)=Φ(ξ/2)Φ(ξ)\phi(\xi)=\Phi(\xi/2)-\Phi(\xi) which is supported in {ξ| 1/2|ξ|2}\{\xi\in\mathbb{R}\ |\ 1/2\leq|\xi|\leq 2\}. Then, for all ξ\xi\in\mathbb{R}, one has Φ(ξ)+jϕ(2jξ)=1\Phi(\xi)+\sum_{j\in\mathbb{N}}\phi(2^{-j}\xi)=1, which one can use to decompose tempered distribution (this setting includes in particular periodic function uL2(𝕋)u\in L^{2}(\mathbb{T})). For u𝒮()u\in\mathcal{S}^{\prime}(\mathbb{R}), we set

Δ1u=1(Φ(ξ)u^),Δju=1(ϕ(2jξ)u^)forj.\Delta_{-1}u=\mathcal{F}^{-1}(\Phi(\xi)\widehat{u}),\qquad\Delta_{j}u=\mathcal{F}^{-1}(\phi(2^{-j}\xi)\widehat{u})\quad\text{for}\quad j\in\mathbb{N}.

We also use the notation Sju=1pj1ΔpuS_{j}u=\sum_{-1\leq p\leq j-1}\Delta_{p}u for j0j\geq 0 (so that S0u=Δ1u=Φ(D)uS_{0}u=\Delta_{-1}u=\Phi(D)u).

Given a function ff, denote by ff^{\flat} denotes the multiplication operator ufuu\mapsto fu and denote by TfT_{f} the operator of paramultiplication by ff, defined by

Tfu=j1Sj1(f)Δju.T_{f}u=\sum_{j\geq 1}S_{j-1}(f)\Delta_{j}u.

Now rewrite the commutator [𝒬,f]\left[\mathcal{Q},f^{\flat}\right] as

[𝒬,Tf]+𝒬(fTf)(fTf)𝒬.\left[\mathcal{Q},T_{f}\right]+\mathcal{Q}(f^{\flat}-T_{f})-(f^{\flat}-T_{f})\mathcal{Q}.

The claim then follows from the bounds

σ,[𝒬,Tf]HσHσα+βc1(q,σ)fC0,β,\displaystyle\forall\sigma\in\mathbb{R},\quad\quad\quad\left\lVert\left[\mathcal{Q},T_{f}\right]\right\rVert_{H^{\sigma}\rightarrow H^{\sigma-\alpha+\beta}}\leq c_{1}(q,\sigma)\left\lVert f\right\rVert_{C^{0,\beta}}, (2.6)
σ[α,0],fTfHσHσ+αc2(σ)fC0,β,\displaystyle\forall\sigma\in[-\alpha,0],\quad\big{\lVert}f^{\flat}-T_{f}\big{\rVert}_{H^{\sigma}\rightarrow H^{\sigma+\alpha}}\leq c_{2}(\sigma)\left\lVert f\right\rVert_{C^{0,\beta}}, (2.7)
σ,𝒬HσHσαsupξ|ξαq(ξ)|,\displaystyle\forall\sigma\in\mathbb{R},\quad\quad\quad\left\lVert\mathcal{Q}\right\rVert_{H^{\sigma}\rightarrow H^{\sigma-\alpha}}\lesssim\sup\nolimits_{\xi}\left\lvert\langle\xi\rangle^{-\alpha}q(\xi)\right\rvert, (2.8)

where c1(,σ):Sα+c_{1}(\cdot,\sigma)\colon S^{\alpha}\rightarrow\mathbb{R}_{+} maps bounded sets to bounded sets. The estimate (2.8) is a direct result of Plancherel Theorem. We refer the reader to [Hör97, Proposition 10.2.2] and [Hör97, Theorem 9.6.4] for the proof of (2.6). Let us now prove (2.7). To do so, observe that

fgTfg=j,p1(Δjf)(Δpg)1jp2(Δjf)(Δpg)=j1(Sj+2g)Δjf.fg-T_{f}g=\sum_{j,p\geq-1}(\Delta_{j}f)(\Delta_{p}g)-\sum_{-1\leq j\leq p-2}(\Delta_{j}f)(\Delta_{p}g)=\sum_{j\geq-1}(S_{j+2}g)\Delta_{j}f.

We then use the Bernstein’s inequality and the characterization of Sobolev and Hölder spaces in terms of Littlewood-Paley decomposition, to write

fgTfgHσ+α\displaystyle\left\lVert fg-T_{f}g\right\rVert_{H^{\sigma+\alpha}} (Sj+2g)(Δjf)Hσ+α\displaystyle\leq\sum\left\lVert(S_{j+2}g)(\Delta_{j}f)\right\rVert_{H^{\sigma+\alpha}}
2j(σ+α)(Sj+2g)(Δjf)L2\displaystyle\lesssim\sum 2^{j(\sigma+\alpha)}\left\lVert(S_{j+2}g)(\Delta_{j}f)\right\rVert_{L^{2}}\qquad (since σ+α0)\displaystyle(\text{since }\sigma+\alpha\geq 0)
2j(σ+α)Sj+2gL2ΔjfL\displaystyle\leq\sum 2^{j(\sigma+\alpha)}\left\lVert S_{j+2}g\right\rVert_{L^{2}}\left\lVert\Delta_{j}f\right\rVert_{L^{\infty}}
2jαgHσ2jαfC0,β\displaystyle\lesssim\sum 2^{j\alpha}\left\lVert g\right\rVert_{H^{\sigma}}2^{-j\alpha}\left\lVert f\right\rVert_{C^{0,\beta}} (since σ0)\displaystyle(\text{since }\sigma\leq 0)
gHσfC0,β,\displaystyle\lesssim\left\lVert g\right\rVert_{H^{\sigma}}\left\lVert f\right\rVert_{C^{0,\beta}},

where we used the assumption α<β\alpha<\beta to insure that the series 2j(αβ)\sum 2^{j(\alpha-\beta)} converges. ∎

This concludes the proof of Proposition 2.1. ∎

3. Resolvent estimates for localized damping

This section is devoted to proving the resolvent bound for localized damping, that is, the first part of Theorem 2. This resolvent bound gives 1/t21/t^{2} energy decay for solutions to (1.4) when (u0,u1)H1×H12(u_{0},u_{1})\in H^{1}\times H^{\frac{1}{2}} using [AL14], and the proof of the energy decay is streamlined in §5.

To take advantage of semiclassical analysis, we introduce the semiclassical rescaling

τ=zh,h>0,z,\tau=\frac{z}{\sqrt{h}},\ h>0,\ z\in{\mathbb{C}}, (3.1)

and define

𝒫(h,z):=h|D|ihzχz2.\mathcal{P}(h,z):=h|D|-i\sqrt{h}z\chi-z^{2}. (3.2)

We notice that

P(τ)=h1𝒫(h,z).P(\tau)=h^{-1}\mathcal{P}(h,z).

We start by stating an equivalent version of the first part of Theorem 2 in the semiclassical scale.

Proposition 3.1.

Suppose χC0,β(𝕋)\chi\in C^{0,\beta}(\mathbb{T}) with β>12\beta>\frac{1}{2}. Let 𝒫(h,z){\mathcal{P}}(h,z) be as in (3.2). Then there exist h0>0h_{0}>0, δ>0\delta>0 and C>0C>0 such that, for all 0<h<h00<h<h_{0}, all complex number zSδ={x+iy|x(1δ,1+δ),y(δh,δh)}z\in S_{\delta}=\{x+iy\ |\ x\in(1-\delta,1+\delta),y\in(-\delta h,\delta h)\} and for all uC(𝕋)u\in C^{\infty}(\mathbb{T}), there holds

uL2Ch1𝒫(h,z)uL2.\left\lVert u\right\rVert_{L^{2}}\leq Ch^{-1}\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}. (3.3)
Proof of Proposition 3.1.

We denote by CC various constants independent of hh and whose value may change from line to line. We write ABA\lesssim B to say that ACBA\leq CB for such a constant CC.

We first assume that zz is a real number with z(1δ,1+δ)z\in(1-\delta,1+\delta) for some δ>0\delta>0 sufficiently small.

1. Estimate of χu\chi u.

We claim that

χuL22h1𝒫(h,z)uL2uL2.\left\lVert\chi u\right\rVert_{L^{2}}^{2}\lesssim h^{-1}\|{\mathcal{P}}(h,z)u\|_{L^{2}}\left\lVert u\right\rVert_{L^{2}}. (3.4)

To see this, observe that, by definition of 𝒫(h,z)\mathcal{P}(h,z) we have

χu=iz1h12(h|D|u𝒫(h,z)uz2u).\chi u=-iz^{-1}h^{-\frac{1}{2}}\left(h\left\lvert D\right\rvert u-\mathcal{P}(h,z)u-z^{2}u\right).

Consequently, by taking the L2(𝕋)L^{2}(\mathbb{T})-scalar product with uu, we get

χu,u=Reχu,u=z1h12Im𝒫(h,z)u,u.\langle\chi u,u\rangle=\operatorname{Re}\langle\chi u,u\rangle=-z^{-1}h^{-\frac{1}{2}}\operatorname{Im}\langle{\mathcal{P}}(h,z)u,u\rangle.

Therefore, it follows from the Cauchy-Schwarz inequality that

0χu,u11δh12𝒫(h,z)uL2uL2.0\leq\langle\chi u,u\rangle\leq\frac{1}{1-\delta}h^{-\frac{1}{2}}\left\lVert{\mathcal{P}}(h,z)u\right\rVert_{L^{2}}\left\lVert u\right\rVert_{L^{2}}.

Since χuL22χLχu,u\left\lVert\chi u\right\rVert_{L^{2}}^{2}\leq\left\lVert\chi\right\rVert_{L^{\infty}}\langle\chi u,u\rangle, this immediately implies the wanted estimate (3.4).

2. Propagation estimates.

The estimate of the remaining component (1χ)u(1-\chi)u is divided into two steps. We begin with the most delicate part, which consists in estimate the microlocal component of (1χ)u(1-\chi)u where the operator 𝒫(h,z)\mathcal{P}(h,z) is not elliptic. The analysis will therefore rely on a propagation argument.

More precisely, consider a cut-off function gC(;[0,1])g\in C^{\infty}(\mathbb{R};[0,1]) satisfying

g(ξ)={0for |ξ|15 or |ξ|5,1for 14|ξ|4.g(\xi)=\left\{\begin{aligned} &0\quad&&\text{for }\ \ |\xi|\leq\tfrac{1}{5}\text{ or }|\xi|\geq 5,\\ &1\quad&&\text{for }\tfrac{1}{4}\leq|\xi|\leq 4.\end{aligned}\right.

We want to estimate the L2L^{2}-norm of g(hD)((1χ)u)g(hD)((1-\chi)u). We claim that, for some exponent ν>0\nu>0,

g(hD)((1χ)u)L22χuL22+h1𝒫(h,z)uL2uL2+hνuL22.\left\lVert g(hD)\big{(}(1-\chi)u\big{)}\right\rVert_{L^{2}}^{2}\lesssim\left\lVert\chi u\right\rVert_{L^{2}}^{2}+h^{-1}\|{\mathcal{P}}(h,z)u\|_{L^{2}}\left\lVert u\right\rVert_{L^{2}}+h^{\nu}\left\lVert u\right\rVert_{L^{2}}^{2}. (3.5)

To prove this claim we use two different kinds of localization.

Localization in frequency. We further decompose the problem into waves traveling to the left and waves traveling to the right. To do so, consider two cut-off functions g±Cc(;[0,1])g_{\pm}\in C_{c}^{\infty}(\mathbb{R};[0,1]) with g(ξ)=g+(ξ)g_{-}(\xi)=g_{+}(-\xi) and such that

g+(ξ)={0for ξ(,15)(5,),1for ξ[14,4].g_{+}(\xi)=\left\{\begin{aligned} &0\quad&&\text{for }\ \xi\in(-\infty,\tfrac{1}{5})\cup(5,\infty),\\ &1\quad&&\text{for }\xi\in[\tfrac{1}{4},4].\end{aligned}\right.

To obtain the wanted estimate (3.5), it is sufficient to prove that g+(hD)((1χ)u)L2\left\lVert g_{+}(hD)\big{(}(1-\chi)u\big{)}\right\rVert_{L^{2}} and g(hD)((1χ)u)L2\left\lVert g_{-}(hD)\big{(}(1-\chi)u\big{)}\right\rVert_{L^{2}} are bounded by the right-hand side of (3.5).

These two terms will be treated similarly (see the explanations at the end of this step) and for notational simplicity we focus on the estimate of g+(hD)((1χ)u)g_{+}(hD)((1-\chi)u). Our aim is thus to prove that, for some exponent ν>0\nu>0, we have

g+(hD)((1χ)u)L22χuL22+h1𝒫(h,z)uL2uL2+hνuL22.\left\lVert g_{+}(hD)\big{(}(1-\chi)u\big{)}\right\rVert_{L^{2}}^{2}\lesssim\left\lVert\chi u\right\rVert_{L^{2}}^{2}+h^{-1}\left\lVert{\mathcal{P}}(h,z)u\right\rVert_{L^{2}}\left\lVert u\right\rVert_{L^{2}}+h^{\nu}\left\lVert u\right\rVert_{L^{2}}^{2}. (3.6)

Localization in space. In addition to the previous localization in frequency, we see that to prove (3.6), by using a suitable partition of unity, it is sufficient to prove that, for any point x0supp(1χ)x_{0}\in\operatorname{supp}(1-\chi), there exist ν>0\nu>0 and a function bC(𝕋;[0,1])b\in C^{\infty}(\mathbb{T};[0,1]) with b(x0)>0b(x_{0})>0 such that

g+(hD)(bu)L2χuL22+h1𝒫(h,z)uL2uL2+hνuL22.\left\lVert g_{+}(hD)(bu)\right\rVert_{L^{2}}\lesssim\left\lVert\chi u\right\rVert_{L^{2}}^{2}+h^{-1}\left\lVert{\mathcal{P}}(h,z)u\right\rVert_{L^{2}}\left\lVert u\right\rVert_{L^{2}}+h^{\nu}\left\lVert u\right\rVert_{L^{2}}^{2}. (3.7)

We will use suitable cut-off functions bb, as given by the following

Lemma 3.2.

Assume that χ(x0)=0\chi(x_{0})=0. Then there exist two functions φ,bC(𝕋)\varphi,b\in C^{\infty}(\mathbb{T}) such that

1φ3, 0b1,b(x0)>0,1\leq\varphi\leq 3,\ 0\leq b\leq 1,\ b(x_{0})>0,

and moreover φ\varphi is such that its derivative satisfies φ=ψ+ψ\varphi^{\prime}=\psi_{+}-\psi_{-} for some functions ψ±C(𝕋)\psi_{\pm}\in C^{\infty}(\mathbb{T}) satisfying ψ±0\psi_{\pm}\geq 0 and

suppψ+suppψ=,χ|suppψ+>0,ψ|suppb>0.\operatorname{supp}\psi_{+}\cap\operatorname{supp}\psi_{-}=\emptyset,\ \ \chi|_{\operatorname{supp}\psi_{+}}>0,\ \psi_{-}|_{\operatorname{supp}b}>0.
Refer to caption
Figure 1. Auxiliary functions φ\varphi, bb constructed in Lemma 3.2 with frequencies localized near ξ=1\xi=1. The arrow indicates the direction of the Hamiltonian flow for p=|ξ|p=|\xi| near ξ=1\xi=1.
Proof of Lemma 3.2.

Introduce a 2π2\pi-periodic function κC(𝕋;[0,1])\kappa\in C^{\infty}(\mathbb{T};[0,1]) which is even and such that

κ(x)={1for |x|14,0for 12|x|π.\kappa(x)=\left\{\begin{aligned} &1\quad&&\text{for }\ \left\lvert x\right\rvert\leq\tfrac{1}{4},\\ &0\quad&&\text{for }\tfrac{1}{2}\leq\left\lvert x\right\rvert\leq\pi.\end{aligned}\right.

Given three parameters α,α\alpha,\alpha^{\prime} and rr to be determined, define

ψ(x)=ακ(xx02r),b(x)=ακ(xx0r).\psi_{-}(x)=\alpha\kappa\left(\frac{x-x_{0}}{2r}\right),\quad b(x)=\alpha^{\prime}\kappa\left(\frac{x-x_{0}}{r}\right).

Now pick x1x_{1} such that χ(x)>0\chi(x)>0 for all x[x12r,x1+2r]x\in[x_{1}-2r,x_{1}+2r] and set

ψ+(x)=ακ(xx12r).\psi_{+}(x)=\alpha\kappa\left(\frac{x-x_{1}}{2r}\right).

Now let θ\theta be the unique function θC(𝕋)\theta\in C^{\infty}(\mathbb{T}) with mean value 0 and such that θ=ψ+ψ\theta^{\prime}=\psi_{+}-\psi_{-}. We then set φ(x)=2+θ(x)\varphi(x)=2+\theta(x) and choose α,α\alpha,\alpha^{\prime} and rr small enough. ∎

Commutator argument. Given the functions φ=φ(x)\varphi=\varphi(x) and g+=g+(ξ)g_{+}=g_{+}(\xi) as introduced above, we consider the operator G+G_{+} defined by

G+u=φg+(hD)u.G_{+}u=\varphi g_{+}(hD)u. (3.8)

The idea is to exploit the fact that G+G+G_{+}^{*}G_{+} is self-adjoint to write Im𝒫(h,z)u,G+G+u\operatorname{Im}\langle{\mathcal{P}}(h,z)u,G_{+}^{*}G_{+}u\rangle under the form of a commutator:

Im𝒫(h,z)u,G+G+u\displaystyle\operatorname{Im}\langle{\mathcal{P}}(h,z)u,G_{+}^{*}G_{+}u\rangle =12i(𝒫(h,z)u,G+G+uG+G+u,𝒫(h,z)u)\displaystyle=\frac{1}{2i}\Big{(}\langle{\mathcal{P}}(h,z)u,G_{+}^{*}G_{+}u\rangle-\langle{G_{+}^{*}G_{+}u,\mathcal{P}}(h,z)u\rangle\Big{)}
=12i(G+G+𝒫(h,z)u,u𝒫(h,z)G+G+u,u)\displaystyle=\frac{1}{2i}\Big{(}\langle G_{+}^{*}G_{+}\mathcal{P}(h,z)u,u\rangle-\langle{\mathcal{P}}(h,z)^{*}G_{+}^{*}G_{+}u,u\rangle\Big{)}
=12i[h|D|,G+G+]u,uzhReχu,G+G+u.\displaystyle=\big{\langle}-\tfrac{1}{2i}[h|D|,G_{+}^{*}G_{+}]u,u\big{\rangle}-z\sqrt{h}\operatorname{Re}\langle\chi u,G_{+}^{*}G_{+}u\rangle. (3.9)

We then notice that by the assumption on the support of g+g_{+}, we have |hD|g+(hD)=hDg+(hD)\left\lvert hD\right\rvert g_{+}(hD)=hDg_{+}(hD) which in turn implies that

|D|G+=DG+andG+|D|=G+Dwith D=1ix.\left\lvert D\right\rvert G_{+}^{*}=DG_{+}^{*}\quad\text{and}\quad G_{+}\left\lvert D\right\rvert=G_{+}D\quad\text{with }D=\frac{1}{i}\partial_{x}.

We thus end up with

12i[h|D|,G+G+]=12i[hD,G+G+].-\frac{1}{2i}[h|D|,G_{+}^{*}G_{+}]=-\frac{1}{2i}[hD,G_{+}^{*}G_{+}].

Once this formula is established, one can compute this commutator using only the Leibniz formula. Indeed, directly from the definitions of D=x/iD=\partial_{x}/i and G+G_{+} (see (3.8)), we have (noticing that g+(hD)=g+(hD)g_{+}(hD)^{*}=g_{+}(hD))

12i[hD,G+G+]u\displaystyle-\frac{1}{2i}[hD,G_{+}^{*}G_{+}]u =h2i1i[x,g+(hD)(φ2g+(hD))]u\displaystyle=-\frac{h}{2i}\frac{1}{i}\big{[}\partial_{x},g_{+}(hD)\big{(}\varphi^{2}g_{+}(hD)\cdot\big{)}\big{]}u
=hg+(hD)(φφg+(hD)u).\displaystyle=hg_{+}(hD)\big{(}\varphi\varphi^{\prime}g_{+}(hD)u\big{)}.

By combining the previous identities and using again the fact that g+(hD)=g+(hD)g_{+}(hD)^{*}=g_{+}(hD), we conclude that

12i[h|D|,G+G+]u,u=hφφg+(hD)u,g+(hD)u.\big{\langle}-\tfrac{1}{2i}[h|D|,G_{+}^{*}G_{+}]u,u\big{\rangle}=h\big{\langle}\varphi\varphi^{\prime}g_{+}(hD)u,g_{+}(hD)u\big{\rangle}.

Now we use the special form of the function φ\varphi, that is the fact that φ=ψ+ψ\varphi^{\prime}=\psi_{+}-\psi_{-}. Introduce the semiclassical operators E+E_{+} and F+F_{+} defined by

E+u:=φψ+g+(hD)u,F+u:=φψg+(hD)u.E_{+}u:=\varphi\psi_{+}g_{+}(hD)u\quad,\quad F_{+}u:=\varphi\psi_{-}g_{+}(hD)u.

Then, we have

12i[h|D|,G+G+]u,u=hE+u,g+(hD)uhF+u,g+(hD)u.\langle-\tfrac{1}{2i}[h|D|,G_{+}^{*}G_{+}]u,u\rangle=h\langle E_{+}u,g_{+}(hD)u\rangle-h\langle F_{+}u,g_{+}(hD)u\rangle. (3.10)

Having analyzed the first term in the right-side of (3.9), we now estimate the second one. To do so, we begin by writing the latter under the form

Reχu,G+G+u=χG+u,G+u+Re(G+[G+,χ])u,u.\operatorname{Re}\langle\chi u,G_{+}^{*}G_{+}u\rangle=\langle\chi G_{+}u,G_{+}u\rangle+\langle\operatorname{Re}(G_{+}^{*}[G_{+},\chi])u,u\rangle. (3.11)

Since χ0\chi\geq 0, one has the obvious inequality

χG+u,G+u0.\langle\chi G_{+}u,G_{+}u\rangle\geq 0.

To estimate the commutator [G+,χ]u[G_{+},\chi]u we shall make use of the assumption that χ\chi belongs to the Hölder space C0,β(𝕋)C^{0,\beta}(\mathbb{T}) for some exponent β(1/2,1]\beta\in(1/2,1]. Let 0<ε<β1/20<\varepsilon<\beta-1/2. It follows from Proposition 2.1 applied with α=1/2+ε\alpha=1/2+\varepsilon that

[G+,χ]uL2Ch12+εuL2.\left\lVert[G_{+},\chi]u\right\rVert_{L^{2}}\leq Ch^{\frac{1}{2}+\varepsilon}\|u\|_{L^{2}}. (3.12)

Since G+L2L2\left\lVert G_{+}\right\rVert_{L^{2}\to L^{2}} is bounded uniformly in hh, it follows that

zhReχu,G+G+uCh1+εuL22.z\sqrt{h}\operatorname{Re}\langle\chi u,G_{+}^{*}G_{+}u\rangle\geq-Ch^{1+\varepsilon}\left\lVert u\right\rVert_{L^{2}}^{2}. (3.13)

Now, by combining (3.9) together with (3.10) and (3.13), we find

hF+u,g+(hD)uhE+u,g+(hD)u+|𝒫(h,z)u,G+G+u|+O(h1+ε)uL22.h\langle F_{+}u,g_{+}(hD)u\rangle\leq h\langle E_{+}u,g_{+}(hD)u\rangle+|\langle{\mathcal{P}}(h,z)u,G_{+}^{*}G_{+}u\rangle|+O(h^{1+\varepsilon})\left\lVert u\right\rVert^{2}_{L^{2}}.

By using again the fact that G+L2L2\left\lVert G_{+}\right\rVert_{L^{2}\to L^{2}} is bounded uniformly in hh, by dividing each side of the previous inequality by hh and using the Cauchy-Schwarz inequality, this yields

F+u,g+(hD)uE+u,g+(hD)u+h1𝒫(h,z)uL2uL2+O(hε)uL22.\langle F_{+}u,g_{+}(hD)u\rangle\leq\langle E_{+}u,g_{+}(hD)u\rangle+h^{-1}\left\lVert{\mathcal{P}}(h,z)u\right\rVert_{L^{2}}\left\lVert u\right\rVert_{L^{2}}+O(h^{\varepsilon})\left\lVert u\right\rVert^{2}_{L^{2}}.

It remains to bound F+u,u\langle F_{+}u,u\rangle (resp. E+u,u\langle E_{+}u,u\rangle) from below (resp. above). To do so, set f=ψf=\sqrt{\psi_{-}}. Since ff belongs to C0,1/2(𝕋)C^{0,1/2}(\mathbb{T}), it follows from Proposition 2.1 applied with β=1/2\beta=1/2 and α\alpha any arbitrary real number in (0,β)(0,\beta), that

F+u,g+(hD)u=φg+(hD)(fu),g+(hD)(fu)+O(hα)uL22.\langle F_{+}u,g_{+}(hD)u\rangle=\langle\varphi g_{+}(hD)(fu),g_{+}(hD)(fu)\rangle+O(h^{\alpha})\left\lVert u\right\rVert_{L^{2}}^{2}.

Since φ1\varphi\geq 1, we deduce that

F+u,g+(hD)ug+(hD)(ψu)L22+O(hα)uL22.\langle F_{+}u,g_{+}(hD)u\rangle\geq\left\lVert g_{+}(hD)(\sqrt{\psi_{-}}u)\right\rVert_{L^{2}}^{2}+O(h^{\alpha})\left\lVert u\right\rVert_{L^{2}}^{2}.

Now consider the function bb as given by Lemma 3.2. Since bb can be written under the form

b=fψwithf=bψC(𝕋),b=f\sqrt{\psi_{-}}\quad\text{with}\quad f=\frac{b}{\sqrt{\psi_{-}}}\in C^{\infty}(\mathbb{T}),

we have

g+(hD)(bu)=fg+(hD)(ψu)+[g+(hD),f]u.g_{+}(hD)(bu)=fg_{+}(hD)(\sqrt{\psi_{-}}u)+\big{[}g_{+}(hD),f\big{]}u.

Using again Proposition 2.1 to estimate the commutator, we get that

F+u,g+(hD)ug+(hD)(bu)L22+O(hα)uL22.\langle F_{+}u,g_{+}(hD)u\rangle\geq\left\lVert g_{+}(hD)(bu)\right\rVert_{L^{2}}^{2}+O(h^{\alpha})\left\lVert u\right\rVert_{L^{2}}^{2}.

On the other hand, by applying Proposition 2.1 with f=φψ+C0,1/2(𝕋)f=\sqrt{\varphi\psi_{+}}\in C^{0,1/2}(\mathbb{T}), we deduce that

E+u,g+(hD)u=g+(hD)(φψ+u)L22+O(hα)uL22.\langle E_{+}u,g_{+}(hD)u\rangle=\left\lVert g_{+}(hD)(\sqrt{\varphi\psi_{+}}u)\right\rVert_{L^{2}}^{2}+O(h^{\alpha})\left\lVert u\right\rVert_{L^{2}}^{2}.

This immediately implies that

E+u,g+(hD)uφψ+uL22+O(hα)uL22CχuL22+ChαuL22.\langle E_{+}u,g_{+}(hD)u\rangle\leq\left\lVert\sqrt{\varphi\psi_{+}}u\right\rVert_{L^{2}}^{2}+O(h^{\alpha})\left\lVert u\right\rVert_{L^{2}}^{2}\leq C\left\lVert\chi u\right\rVert_{L^{2}}^{2}+Ch^{\alpha}\left\lVert u\right\rVert_{L^{2}}^{2}.

Consequently, we end up with

g+(hD)(bu)L2χuL22+h1𝒫(h,z)uL2uL2+(hα+hε)uL22.\left\lVert g_{+}(hD)(bu)\right\rVert_{L^{2}}\lesssim\left\lVert\chi u\right\rVert_{L^{2}}^{2}+h^{-1}\|{\mathcal{P}}(h,z)u\|_{L^{2}}\left\lVert u\right\rVert_{L^{2}}+(h^{\alpha}+h^{\varepsilon})\left\lVert u\right\rVert_{L^{2}}^{2}. (3.14)

This completes the proof of (3.7) which in turn completes the proof of (3.16).

Lastly, to prove similar estimates for g(hD)((1χ)u)L2\|g_{-}(hD)((1-\chi)u)\|_{L^{2}}, it suffices to construct auxiliary functions φ\varphi, bb as in Lemma 3.2 with mere changes (see Figure 2)

φ=ψψ+,χ|suppψ>0,ψ+|suppb>0.\varphi^{\prime}=\psi_{-}-\psi_{+},\ \chi|_{\operatorname{supp}\psi_{-}}>0,\ \psi_{+}|_{\operatorname{supp}b}>0.
Refer to caption
Figure 2. Auxiliary functions φ\varphi, bb used in the proof of the propagation estimates with frequencies localized near ξ=1\xi=-1. The arrow indicates the direction of the Hamiltonian flow for p=|ξ|p=|\xi| near ξ=1\xi=-1.

3. Elliptic estimates.

Having estimated the main contribution of the frequencies of size near 11 in the semiclassical scale, we now turn to the estimation of the low and high frequency components. More precisely, we want to estimate the L2L^{2}-norms of G0uG_{0}u and GuG_{\infty}u where G0=g0(hD)G_{0}=g_{0}(hD) and G=g(hD)G_{\infty}=g_{\infty}(hD) are semiclassical Fourier multipliers with symbols g0=g0(ξ)g_{0}=g_{0}(\xi) and g=g(ξ)g_{\infty}=g_{\infty}(\xi) in C(;[0,1])C^{\infty}(\mathbb{R};[0,1]), such that

g0(ξ)=1for|ξ|1/5,g0(ξ)=0for|ξ|1/4;g(ξ)=0for|ξ|4,g(ξ)=1for|ξ|5.\begin{gathered}g_{0}(\xi)=1\ \text{for}\ |\xi|\leq 1/5,\ g_{0}(\xi)=0\ \text{for}\ |\xi|\geq 1/4;\\ g_{\infty}(\xi)=0\ \text{for}\ |\xi|\leq 4,\ g_{\infty}(\xi)=1\ \text{for}\ |\xi|\geq 5.\end{gathered} (3.15)

High frequency estimate. We begin by estimating GuG_{\infty}u. To do so, we write that, by the definition of 𝒫(h,z)\mathcal{P}(h,z),

G|hD|u=G𝒫(h,z)u+z2Gu+izGhχu.G_{\infty}\left\lvert hD\right\rvert u=G_{\infty}\mathcal{P}(h,z)u+z^{2}G_{\infty}u+izG_{\infty}\sqrt{h}\chi u.

Since GL2L21\left\lVert G_{\infty}\right\rVert_{L^{2}\to L^{2}}\leq 1, we have

G|hD|uL2𝒫(h,z)uL2+(1+δ)2GuL2+(1+δ)hχLuL2.\left\lVert G_{\infty}\left\lvert hD\right\rvert u\right\rVert_{L^{2}}\leq\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}+(1+\delta)^{2}\left\lVert G_{\infty}u\right\rVert_{L^{2}}+(1+\delta)\sqrt{h}\left\lVert\chi\right\rVert_{L^{\infty}}\left\lVert u\right\rVert_{L^{2}}.

On the other hand, since |ξ|4\left\lvert\xi\right\rvert\geq 4 on the support of g(ξ)g_{\infty}(\xi), directly from the Plancherel theorem, we have

4GuL2G|hD|uL2.4\left\lVert G_{\infty}u\right\rVert_{L^{2}}\leq\left\lVert G_{\infty}\left\lvert hD\right\rvert u\right\rVert_{L^{2}}.

Therefore, for hh and δ\delta small enough, we get

GuL22𝒫(h,z)uL22+13uL22.\left\lVert G_{\infty}u\right\rVert_{L^{2}}^{2}\leq\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}^{2}+\tfrac{1}{3}\left\lVert u\right\rVert_{L^{2}}^{2}.

Low frequency estimates. The low frequency component G0uG_{0}u is estimated in a similar way. Namely, we write

z2G0u=G0|hD|uG0𝒫(h,z)uizG0hχu,z^{2}G_{0}u=G_{0}\left\lvert hD\right\rvert u-G_{0}\mathcal{P}(h,z)u-izG_{0}\sqrt{h}\chi u,

and then observe that G0L2L21\left\lVert G_{0}\right\rVert_{L^{2}\to L^{2}}\leq 1, to obtain

(1δ)2G0uL2G0|hD|uL2+𝒫(h,z)uL2+(1+δ)hχLuL2.(1-\delta)^{2}\left\lVert G_{0}u\right\rVert_{L^{2}}\leq\left\lVert G_{0}\left\lvert hD\right\rvert u\right\rVert_{L^{2}}+\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}+(1+\delta)\sqrt{h}\left\lVert\chi\right\rVert_{L^{\infty}}\left\lVert u\right\rVert_{L^{2}}.

Since |ξ|1/4\left\lvert\xi\right\rvert\leq 1/4 on the support of g0(ξ)g_{0}(\xi), directly from the Plancherel theorem, we have

G0|hD|uL214G0uL2.\left\lVert G_{0}\left\lvert hD\right\rvert u\right\rVert_{L^{2}}\leq\tfrac{1}{4}\left\lVert G_{0}u\right\rVert_{L^{2}}.

Therefore, for hh, δ\delta small enough, we get

G0uL22𝒫(h,z)uL22+13uL22.\left\lVert G_{0}u\right\rVert_{L^{2}}^{2}\leq\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}^{2}+\tfrac{1}{3}\left\lVert u\right\rVert_{L^{2}}^{2}.

4. End of the proof when zz is a real number.

The Plancherel theorem implies that

uL22g(hD)uL22+G0uL22+GuL22.\left\lVert u\right\rVert_{L^{2}}^{2}\leq\left\lVert g(hD)u\right\rVert_{L^{2}}^{2}+\left\lVert G_{0}u\right\rVert_{L^{2}}^{2}+\left\lVert G_{\infty}u\right\rVert_{L^{2}}^{2}.

Therefore, it follows from the previous L2L^{2}-estimates for G0uG_{0}u and GuG_{\infty}u that

uL22g(hD)uL22+2𝒫(h,z)uL22+23uL22,\left\lVert u\right\rVert_{L^{2}}^{2}\leq\left\lVert g(hD)u\right\rVert_{L^{2}}^{2}+2\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}^{2}+\tfrac{2}{3}\left\lVert u\right\rVert_{L^{2}}^{2},

and hence we have

uL223g(hD)uL22+6𝒫(h,z)uL22.\left\lVert u\right\rVert_{L^{2}}^{2}\leq 3\left\lVert g(hD)u\right\rVert_{L^{2}}^{2}+6\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}^{2}.

Now, it follows from (3.4) and (3.5) that

g(hD)uL22h1𝒫(h,z)uL2uL2+hνuL22.\left\lVert g(hD)u\right\rVert_{L^{2}}^{2}\lesssim h^{-1}\|{\mathcal{P}}(h,z)u\|_{L^{2}}\left\lVert u\right\rVert_{L^{2}}+h^{\nu}\left\lVert u\right\rVert_{L^{2}}^{2}. (3.16)

We conclude that

uL22h1𝒫(h,z)uL2uL2+𝒫(h,z)uL22+hνuL22.\left\lVert u\right\rVert_{L^{2}}^{2}\lesssim h^{-1}\|{\mathcal{P}}(h,z)u\|_{L^{2}}\left\lVert u\right\rVert_{L^{2}}+\left\lVert\mathcal{P}(h,z)u\right\rVert_{L^{2}}^{2}+h^{\nu}\left\lVert u\right\rVert_{L^{2}}^{2}.

Consequently, taking hh small enough, we obtain

uL2h1𝒫(h,z)uL2.\left\lVert u\right\rVert_{L^{2}}\lesssim h^{-1}\|{\mathcal{P}}(h,z)u\|_{L^{2}}.

This concludes the proof of the desired result (3.3) when z(1δ,1+δ)z\in(1-\delta,1+\delta).

5. End of the proof when zz is a complex number. Lastly, if zz is a complex number, we notice that

𝒫(h,z)=𝒫(h,Rez)+(Imz)(hχ+Imz2iRez).\mathcal{P}(h,z)=\mathcal{P}(h,\operatorname{Re}z)+(\operatorname{Im}z)\left(\sqrt{h}\chi+\operatorname{Im}z-2i\operatorname{Re}z\right).

Thus for any complex number zz such that |Rez1|δ|\operatorname{Re}z-1|\leq\delta and |Imz|δh|\operatorname{Im}z|\leq\delta h, we have

𝒫(h,z)uL2𝒫(h,Rez)uL2δh(hχL+δh+2(1+δ))uL2(C1Cδ)huL2.\begin{split}\|\mathcal{P}(h,z)u\|_{L^{2}}\geq&\|\mathcal{P}(h,\operatorname{Re}z)u\|_{L^{2}}-\delta h(\sqrt{h}\|\chi\|_{L^{\infty}}+\delta h+2(1+\delta))\|u\|_{L^{2}}\\ \geq&(C^{-1}-C\delta)h\|u\|_{L^{2}}.\end{split}

Consequently, for all δ\delta and hh small enough, the estimate (3.3) still holds. ∎

4. resolvent bounds for finitely degenerate damping

Let 𝒫(h,z)\mathcal{P}(h,z) be as in (3.2). We also introduce

P±(τ):=±Diτχτ2,τ,P_{\pm}(\tau):=\pm D-i\tau\chi-\tau^{2},\ \tau\in{\mathbb{C}},

and the corresponding rescaled operators

𝒫±(h,z):=±hDihzχz2.\mathcal{P}_{\pm}(h,z):=\pm hD-i\sqrt{h}z\chi-z^{2}. (4.1)

The goal of this section is to prove the second part of Theorem 2. We start by rewriting the resolvent bounds in the semiclassical scale.

Proposition 4.1.

Suppose χC(𝕋)\chi\in C^{\infty}(\mathbb{T}) has finite degeneracy as in Definition 1.1. Then for any 0<ϵ<14N+20<\epsilon<\frac{1}{4N+2}, there exists h0>0h_{0}>0, δ>0\delta>0, C>0C>0, such that for 0<h<h00<h<h_{0}, |Rez1|δ|\operatorname{Re}z-1|\leq\delta, |Imz|δh4N+14N+2+ϵ|\operatorname{Im}z|\leq\delta h^{\frac{4N+1}{4N+2}+\epsilon}, such that for any uC(𝕋)u\in C^{\infty}(\mathbb{T}), we have

uL2Ch4N+14N+2ϵ𝒫(h,z)L2.\|u\|_{L^{2}}\leq Ch^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}(h,z)\|_{L^{2}}.

We reduce the proof of the resolvent bound for 𝒫\mathcal{P} to the resolvent bounds for 𝒫±\mathcal{P}_{\pm}. The basic idea is that near the characteristic set of 𝒫\mathcal{P}, the frequency is comparable to ±1\pm 1, hence we can “replace” 𝒫\mathcal{P} by 𝒫±\mathcal{P}_{\pm}. Away from the characteristic set of 𝒫\mathcal{P}, we have the ellipticity of 𝒫\mathcal{P}. 𝒫±\mathcal{P}_{\pm} is easier to analyze as it is an ordinary differential operator and we can use integrating factors to simplify 𝒫±\mathcal{P}_{\pm}.

Refer to caption
Figure 3. Resonances of 𝒫±(h,z)\mathcal{P}_{\pm}(h,z), and zz-domains of estimations in Proposition 4.1 (red) and Proposition 4.2 (green).
Proposition 4.2.

Suppose χ\chi has finite degeneracy as in Definition 1.1 and 𝒫±\mathcal{P}_{\pm} are as in (4.1). Then

  1. 1.

    For h>0h>0, all resonances of 𝒫±(h,z)\mathcal{P}_{\pm}(h,z) lie on the lines

    Rez=0 or Imz=χ^(0)2hi.\operatorname{Re}z=0\text{ or }\operatorname{Im}z=-\frac{\widehat{\chi}(0)}{2}\sqrt{h}i.
  2. 2.

    For any 0<γ<χ^(0)20<\gamma<\frac{\widehat{\chi}(0)}{2}, there exists C>0C>0 such that for any |Rez1|δ|\operatorname{Re}z-1|\leq\delta, Imzγh\operatorname{Im}z\geq-\gamma\sqrt{h}, and any uC(𝕋)u\in C^{\infty}(\mathbb{T}), we have

    uL2CeCh𝒫±(h,z)uL2.\|u\|_{L^{2}}\leq Ce^{\frac{C}{\sqrt{h}}}\|\mathcal{P}_{\pm}(h,z)u\|_{L^{2}}.
  3. 3.

    For any 0<ϵ<14N+20<\epsilon<\frac{1}{4N+2}, there exists h0>0h_{0}>0, δ>0\delta>0, such that for 0<h<h00<h<h_{0}, |Rez1|δ|\operatorname{Re}z-1|\leq\delta, |Imz|δh4N+14N+2+ϵ|\operatorname{Im}z|\leq\delta h^{\frac{4N+1}{4N+2}+\epsilon}, there exists C>0C>0 such that for any uC(𝕋)u\in C^{\infty}(\mathbb{T}), we have

    uL2Ch4N+14N+2ϵ𝒫±(h,z)L2.\|u\|_{L^{2}}\leq Ch^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}_{\pm}(h,z)\|_{L^{2}}.
Proof of Proposition 4.1 using Proposition 4.2.

Let G0G_{0}, GG_{\infty}, G±G_{\pm} be as in the proof of Proposition 3.1. Then for z(1δ,1+δ)z\in(1-\delta,1+\delta), uC(𝕋)u\in C^{\infty}(\mathbb{T}), we have

G0(h|D|z2)u=G0𝒫(h,z)uizhG0u.G_{0}(h|D|-z^{2})u=G_{0}\mathcal{P}(h,z)u-iz\sqrt{h}G_{0}u.

Thus for δ>0\delta>0 sufficiently small, using Plancherel theorem, we have

G0uL2𝒫(h,z)uL2+huL2.\|G_{0}u\|_{L^{2}}\lesssim\|\mathcal{P}(h,z)u\|_{L^{2}}+\sqrt{h}\|u\|_{L^{2}}. (4.2)

A similar argument shows that

GuL2𝒫(h,z)uL2+huL2.\|G_{\infty}u\|_{L^{2}}\lesssim\|\mathcal{P}(h,z)u\|_{L^{2}}+\sqrt{h}\|u\|_{L^{2}}. (4.3)

Notice that we have the identity

𝒫±(h,z)G±=G±𝒫(h,z)izh[χ,G±].\mathcal{P}_{\pm}(h,z)G_{\pm}=G_{\pm}\mathcal{P}(h,z)-iz\sqrt{h}[\chi,G_{\pm}].

Use Proposition 4.2 and we have

G±uL2h4N+14N+2ϵ𝒫±(h,z)G±uL2h4N+14N+2ϵG±𝒫(h,z)L2+hN+12N+1ϵh1[χ,G±]uL2.\begin{split}\|G_{\pm}u\|_{L^{2}}\lesssim&h^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}_{\pm}(h,z)G_{\pm}u\|_{L^{2}}\\ \lesssim&h^{-\frac{4N+1}{4N+2}-\epsilon}\|G_{\pm}\mathcal{P}(h,z)\|_{L^{2}}+h^{\frac{N+1}{2N+1}-\epsilon}\|h^{-1}[\chi,G_{\pm}]u\|_{L^{2}}.\end{split}

Since h1[χ,G±]Ψ1(𝕋)h^{-1}[\chi,G_{\pm}]\in\Psi^{-1}(\mathbb{T}), we know h1[χ,G±]L2L2C\|h^{-1}[\chi,G_{\pm}]\|_{L^{2}\to L^{2}}\leq C. Therefore

G±uL2h4N+14N+2ϵ𝒫(h,z)uL2+hN+12N+1ϵuL2.\|G_{\pm}u\|_{L^{2}}\lesssim h^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}(h,z)u\|_{L^{2}}+h^{\frac{N+1}{2N+1}-\epsilon}\|u\|_{L^{2}}. (4.4)

Gathering estimates (4.2), (4.3), (4.4), we find that

uL2h4N+14N+2ϵ𝒫(h,z)uL2+huL2.\|u\|_{L^{2}}\lesssim h^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}(h,z)u\|_{L^{2}}+\sqrt{h}\|u\|_{L^{2}}.

Thus when hh is sufficiently small, we have

uL2h4N+14N+2ϵ𝒫(h,z)uL2.\|u\|_{L^{2}}\lesssim h^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}(h,z)u\|_{L^{2}}.

This proves Proposition 4.1 when z(1δ,1+δ)z\in(1-\delta,1+\delta). The proof of Proposition 4.1 is completed by applying the same triangle inequality argument as in the last step of the proof of Proposition 3.1 when zz is complex. ∎

The rest of this section is devoted to proving Proposition 4.2, that is, the spectral gap and the resolvent bound for 𝒫±\mathcal{P}_{\pm}. The main idea is to consider a second microlocalization near zeros of χ\chi: away from the zeros of χ\chi, χ\chi has lower bounds; near zeros of χ\chi, we use the smallness of the second microlocalization and the explicit Green’s formula for 𝒫±\mathcal{P}_{\pm}.

We start by recording a property of χ\chi when it has finite degeneracy.

Lemma 4.3.

Let χ\chi be as in Definition 1.1, then there exists 0<C1<C20<C_{1}<C_{2}, δ>0\delta>0, such that

  1. 1.

    For any |xxk|<2δ|x-x_{k}|<2\delta, 1kn1\leq k\leq n, we have

    C1χ(x)(xxk)2NkC2,C1χ(x)(xxk)2Nk1C2.C_{1}\leq\frac{\chi(x)}{(x-x_{k})^{2N_{k}}}\leq C_{2},\ C_{1}\leq\frac{\chi^{\prime}(x)}{(x-x_{k})^{2N_{k}-1}}\leq C_{2}.
  2. 2.

    If y𝕋y\in\mathbb{T} is a local minimum of χ\chi such that χ(y)0\chi(y)\neq 0, then we have

    χ(y)>10maxxk[xk2δ,xk+2δ]χ(x).\chi(y)>10\max_{x\in\cup_{k}[x_{k}-2\delta,x_{k}+2\delta]}\chi(x).
Proof.

The first conclusion follows from Taylor expansion

χ(x)=1(2Nk)!χ(2Nk)(xk)(xxk)2Nk+O(|xxk|2Nk+1),|xxk|0\chi(x)=\frac{1}{(2N_{k})!}\chi^{(2N_{k})}(x_{k})(x-x_{k})^{2N_{k}}+O(|x-x_{k}|^{2N_{k}+1}),\ |x-x_{k}|\to 0

and a similar expansion for χ\chi^{\prime}.

For the second claim, we notice that χ\chi is a non-vanishing continuous function on J:=𝕋k(xk2δ,xk+2δ)J:=\mathbb{T}\setminus\cup_{k}(x_{k}-2\delta,x_{k}+2\delta). Since JJ is compact, there exists c>0c>0 such that χ(x)>c\chi(x)>c for xJx\in J. Thus the second claim can be achieved by shrinking the value of δ\delta. ∎

Proof of Proposition 4.2.

We only prove for 𝒫+\mathcal{P}_{+}, the proof for 𝒫\mathcal{P}_{-} is similar.

1. Resonances of 𝒫±\mathcal{P}_{\pm}.

To study the resonances of P+P_{+}, we introduce the integrating factor

ezhρ(x)C(𝕋),ρ(x):=0x(χ(y)χ^(0))𝑑yC(𝕋).e^{\frac{z}{\sqrt{h}}\rho(x)}\in C^{\infty}(\mathbb{T}),\ \rho(x):=\int_{0}^{x}(\chi(y)-\widehat{\chi}(0))dy\in C^{\infty}(\mathbb{T}).

One can check that

𝒫+(h,z)=ezhρ(hDizhχ^(0)z2)ezhρ.\mathcal{P}_{+}(h,z)=e^{-\frac{z}{\sqrt{h}}\rho}\left(hD-iz\sqrt{h}\widehat{\chi}(0)-z^{2}\right)e^{\frac{z}{\sqrt{h}}\rho}.

Notice that hDhD has eigenvalues hkhk\in\mathbb{Z}, and izhχ^(0)z2-iz\sqrt{h}\widehat{\chi}(0)-z^{2} is a constant function. Therefore, for h>0h>0, the resonances zz for 𝒫+(h,z)\mathcal{P}_{+}(h,z) must satisfy

z2+izhχ^(0)hk=0,k.z^{2}+iz\sqrt{h}\widehat{\chi}(0)-hk=0,\ k\in\mathbb{Z}.

We can solve

z=(χ^(0)2i±kχ^(0)24)h,kχ^(0)24;z=(χ^(0)2±χ^(0)24k)hi,k<χ^(0)24.\begin{gathered}z=\left(-\frac{\widehat{\chi}(0)}{2}i\pm\sqrt{k-\frac{\widehat{\chi}(0)^{2}}{4}}\right)\sqrt{h},\ k\geq\frac{\widehat{\chi}(0)^{2}}{4};\\ z=\left(-\frac{\widehat{\chi}(0)}{2}\pm\sqrt{\frac{\widehat{\chi}(0)^{2}}{4}-k}\right)\sqrt{h}i,\ k<\frac{\widehat{\chi}(0)^{2}}{4}.\end{gathered}

Thus resonances of 𝒫\mathcal{P} lie in {Rez=0}{Imz=χ^(0)2h}\{\operatorname{Re}z=0\}\cup\{\operatorname{Im}z=-\frac{\widehat{\chi}(0)}{2}\sqrt{h}\}.

2. Estimates for 𝒫±\mathcal{P}_{\pm} away from zeros of χ\chi.

It suffices to prove the resolvent bound for z(1δ,1+δ)z\in(1-\delta,1+\delta). When zz is complex we apply the same triangle inequality argument as in the last step of the proof of Proposition 3.1.

Let xkx_{k}, 1kN1\leq k\leq N, be the zeros of χ\chi. For r>0r>0, we denote

Ik(r):=[xkr,xk+r],I(r):=kIk(r).I_{k}(r):=[x_{k}-r,x_{k}+r],\ I(r):=\bigcup_{k}I_{k}(r).

Let φC(𝕋;[0,1])\varphi\in C^{\infty}(\mathbb{T};[0,1]) such that

suppφ[2,2],φ|[1,1]=1.\operatorname{supp}\varphi\subset[-2,2],\ \varphi|_{[-1,1]}=1.

For β>0\beta>0, we introduce cut-off functions ψk,φ0,φ1C(𝕋)\psi_{k},\varphi_{0},\varphi_{1}\in C^{\infty}(\mathbb{T})

ψk(x):=φ(xxkhβ),φ0(x):=kψk(x),φ1(x):=1kφ(2(xxk)hβ).\begin{gathered}\psi_{k}(x):=\varphi\left(\frac{x-x_{k}}{h^{\beta}}\right),\ \varphi_{0}(x):=\sum_{k}\psi_{k}(x),\ \varphi_{1}(x):=1-\sum_{k}\varphi\left(\frac{2(x-x_{k})}{h^{\beta}}\right).\end{gathered}

Functions φ0\varphi_{0}, φ1\varphi_{1} satisfies the following conditions

suppφ0I(2hβ),φ0|I(hβ)=1;suppφ1𝕋I(hβ/2),φ1|𝕋I(hβ)=1.\operatorname{supp}\varphi_{0}\subset I(2h^{\beta}),\ \varphi_{0}|_{I(h^{\beta})}=1;\ \operatorname{supp}\varphi_{1}\subset\mathbb{T}\setminus I(h^{\beta}/2),\ \varphi_{1}|_{\mathbb{T}\setminus I(h^{\beta})}=1.
Refer to caption
Figure 4. Second microlocalization near zeros of χ\chi.

We first consider the estimates for φ1u\varphi_{1}u. Notice that

Im𝒫+(h,z)φ1u,φ1u=zh𝕋χ|φ1u|2𝑑xCh12+2Nβφ1uL22.\operatorname{Im}\langle\mathcal{P}_{+}(h,z)\varphi_{1}u,\varphi_{1}u\rangle=-z\sqrt{h}\int_{\mathbb{T}}\chi|\varphi_{1}u|^{2}dx\leq-Ch^{\frac{1}{2}+2N\beta}\|\varphi_{1}u\|_{L^{2}}^{2}.

Here we used the fact that

χ|𝕋I(hβ/2)ch2Nβ,c>0, 0<h1,\chi|_{\mathbb{T}\setminus I(h^{\beta}/2)}\geq ch^{2N\beta},\ c>0,\ 0<h\ll 1,

which follows from Lemma 4.3. Now we find that

φ1uL22Ch122Nβ|𝒫+(h,z)φ1u,φ1u|Ch122Nβ𝒫+(h,z)φ1uL2φ1uL2.\begin{split}\|\varphi_{1}u\|_{L^{2}}^{2}\leq&Ch^{-\frac{1}{2}-2N\beta}|\langle\mathcal{P}_{+}(h,z)\varphi_{1}u,\varphi_{1}u\rangle|\\ \leq&Ch^{-\frac{1}{2}-2N\beta}\|\mathcal{P}_{+}(h,z)\varphi_{1}u\|_{L^{2}}\|\varphi_{1}u\|_{L^{2}}.\end{split}

Use Cauchy’s inequality and we find

φ1uL2Ch122Nβ𝒫+(h,z)φ1uL2.\|\varphi_{1}u\|_{L^{2}}\leq Ch^{-\frac{1}{2}-2N\beta}\|\mathcal{P}_{+}(h,z)\varphi_{1}u\|_{L^{2}}.

Now notice that

𝒫+(h,z)φ1u=φ1𝒫+(h,z)uihφ1u,|φ1|Chβ,\mathcal{P}_{+}(h,z)\varphi_{1}u=\varphi_{1}\mathcal{P}_{+}(h,z)u-ih\varphi_{1}^{\prime}u,\ |\varphi_{1}^{\prime}|\leq Ch^{-\beta},

and we conclude

φ1uL2Ch122Nβ𝒫+(h,z)uL2+Ch12(2N+1)βuL2.\|\varphi_{1}u\|_{L^{2}}\leq Ch^{-\frac{1}{2}-2N\beta}\|\mathcal{P}_{+}(h,z)u\|_{L^{2}}+Ch^{\frac{1}{2}-(2N+1)\beta}\|u\|_{L^{2}}. (4.5)

3. Estimates for 𝒫±\mathcal{P}_{\pm} near zeros of χ\chi.

We now turn to estimate φ0u\varphi_{0}u. We first solve

ψk(x)u(x)=ih1xk2hβxK(x,y)𝒫+(h,z)ψku(y)𝑑y,K(x,y)=ezhyxχ(θ)𝑑θ+iz2h(xy).\psi_{k}(x)u(x)=ih^{-1}\int_{x_{k}-2h^{\beta}}^{x}K(x,y)\mathcal{P}_{+}(h,z)\psi_{k}u(y)dy,\ K(x,y)=e^{-\frac{z}{\sqrt{h}}\int_{y}^{x}\chi(\theta)d\theta+\frac{iz^{2}}{h}(x-y)}.

Since χ0\chi\geq 0, we know |K(x,y)|1|K(x,y)|\leq 1, when x,yIk(2hβ)x,y\in I_{k}(2h^{\beta}), y<xy<x. Hence

|ψku|h1Ik(2hβ)|𝒫+(h,z)ψku(y)|𝑑yCh1+β2𝒫+(h,z)ψkuL2.|\psi_{k}u|\leq h^{-1}\int_{I_{k}(2h^{\beta})}|\mathcal{P}_{+}(h,z)\psi_{k}u(y)|dy\leq Ch^{-1+\frac{\beta}{2}}\|\mathcal{P}_{+}(h,z)\psi_{k}u\|_{L^{2}}.

Integrate over 𝕋\mathbb{T} and notice that ψk\psi_{k} is supported in Ik(2hβ)I_{k}(2h^{\beta}), and we find

ψkuL2Ch1+β𝒫+(h,z)ψkuL2.\|\psi_{k}u\|_{L^{2}}\leq Ch^{-1+\beta}\|\mathcal{P}_{+}(h,z)\psi_{k}u\|_{L^{2}}.

Again notice

𝒫+(h,z)ψku=ψk𝒫+(h,z)uihψku,|ψk|Chβ,suppψk{φ1=1},\mathcal{P}_{+}(h,z)\psi_{k}u=\psi_{k}\mathcal{P}_{+}(h,z)u-ih\psi_{k}^{\prime}u,\ |\psi^{\prime}_{k}|\leq Ch^{-\beta},\ \operatorname{supp}\psi_{k}^{\prime}\subset\{\varphi_{1}=1\},

and we conclude that

ψkuL2Ch1+β𝒫+(h,z)uL2+Cφ1uL2.\|\psi_{k}u\|_{L^{2}}\leq Ch^{-1+\beta}\|\mathcal{P}_{+}(h,z)u\|_{L^{2}}+C\|\varphi_{1}u\|_{L^{2}}.

It follows by the definition of φ0\varphi_{0} now that

φ0uL2Ch1+β𝒫+(h,z)uL2+Cφ1uL2.\|\varphi_{0}u\|_{L^{2}}\leq Ch^{-1+\beta}\|\mathcal{P}_{+}(h,z)u\|_{L^{2}}+C\|\varphi_{1}u\|_{L^{2}}. (4.6)

4. End of the proof.

It remains to glue estimates (4.5) and (4.6) to obtain

uL2Cmax(h1+β,h122Nβ)𝒫+(h,z)uL2+Ch12(2N+1)βuL2.\|u\|_{L^{2}}\leq C\max(h^{-1+\beta},h^{-\frac{1}{2}-2N\beta})\|\mathcal{P}_{+}(h,z)u\|_{L^{2}}+Ch^{\frac{1}{2}-(2N+1)\beta}\|u\|_{L^{2}}.

For 0<ϵ<14N+20<\epsilon<\frac{1}{4N+2}, we take β=14N+2ϵ>0\beta=\frac{1}{4N+2}-\epsilon>0, then direct computations show that

1+β=4N+14N+2ϵ<4N+14N+2,122Nβ=4N+14N+2+2Nϵ>4N+14N+2,12(2N+1)β=(2N+1)ϵ>0.\begin{gathered}-1+\beta=-\tfrac{4N+1}{4N+2}-\epsilon<-\tfrac{4N+1}{4N+2},\ -\tfrac{1}{2}-2N\beta=-\tfrac{4N+1}{4N+2}+2N\epsilon>-\tfrac{4N+1}{4N+2},\\ \tfrac{1}{2}-(2N+1)\beta=(2N+1)\epsilon>0.\end{gathered}

Therefore we have

uL2Ch4N+14N+2ϵ𝒫+(h,z)uL2+Ch(2N+1)ϵuL2.\|u\|_{L^{2}}\leq Ch^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}_{+}(h,z)u\|_{L^{2}}+Ch^{(2N+1)\epsilon}\|u\|_{L^{2}}.

For a fixed ϵ>0\epsilon>0, there exists h0=h0(ϵ)h_{0}=h_{0}(\epsilon), such that Ch(2N+1)ϵ<12Ch^{(2N+1)\epsilon}<\frac{1}{2} when 0<h<h00<h<h_{0}, and in this situation we obtain

uL2Ch4N+14N+2ϵ𝒫+(h,z)uL2.\|u\|_{L^{2}}\leq Ch^{-\frac{4N+1}{4N+2}-\epsilon}\|\mathcal{P}_{+}(h,z)u\|_{L^{2}}.

This concludes the proof of Proposition 4.2. ∎

5. Polynomial energy decay

Here, for the benefit of the reader, we give a sketch proof of the equivalence between the resolvent estimates proved above and the corresponding energy and semigroup bounds for the damped wave equation. This gives an overview of the works of [AL14, BD08] taking A=|D|A=|D| and B=χB=\sqrt{\chi}, which builds on ideas developed in [Leb94, LR97]. Since we are working in a very explicit setting with relatively simple operators, connecting the resolvent estimate to the damped wave decay can be done in a fairly explicit manner.

To start with, we rewrite the damped fractional wave equation as

tU=𝐀U,U(0)=U0,𝐀:=(0I|D|χ),U=(utu),\partial_{t}U=\mathbf{A}U,\ U(0)=U_{0},\ \mathbf{A}:=\begin{pmatrix}0&I\\ -|D|&-\chi\end{pmatrix},\ U=\begin{pmatrix}u\\ \partial_{t}u\end{pmatrix}, (5.1)

where U0=(u0u1)H12×L2U_{0}=\begin{pmatrix}u_{0}\\ u_{1}\end{pmatrix}\in H^{\frac{1}{2}}\times L^{2}, 𝐀\mathbf{A} is an operator on :=H12×L2\mathscr{H}:=H^{\frac{1}{2}}\times L^{2} with domain 𝒟(𝐀)=H1×H12\mathcal{D}(\mathbf{A})=H^{1}\times H^{\frac{1}{2}}. The space \mathscr{H} is equipped with a seminorm |(w1w2)|:=|D|12w1L2+w2L2\left|\begin{pmatrix}w_{1}\\ w_{2}\end{pmatrix}\right|_{\mathscr{H}}:=\||D|^{\frac{1}{2}}w_{1}\|_{L^{2}}+\|w_{2}\|_{L^{2}}.

The following proposition established by Anantharaman–Léautaud [AL14] connects the resolvent estimates and energy decay rates of the damped fractional wave equation.

Proposition 5.1 ([AL14, Proposition 2.4]).

Suppose χL(𝕋)\chi\in L^{\infty}(\mathbb{T}). Let P(τ)P(\tau) be as in (1.11). Then for α>0\alpha>0, the following statements are equivalent

  1. 1.

    There exists C>0C>0 such that for any (u0,u1)H1(𝕋)×H12(𝕋)(u_{0},u_{1})\in H^{1}(\mathbb{T})\times H^{\frac{1}{2}}(\mathbb{T}), there holds

    E(u,t)Ct2α|𝐀(u0u1)|2.E(u,t)\leq\frac{C}{t^{2\alpha}}\left|\mathbf{A}\begin{pmatrix}u_{0}\\ u_{1}\end{pmatrix}\right|^{2}_{\mathscr{H}}.

    Notice that |𝐀(u0u1)|2C(|D|u0L22+u1H122)C(u0H12+u1H122)\left|\mathbf{A}\begin{pmatrix}u_{0}\\ u_{1}\end{pmatrix}\right|^{2}_{\mathscr{H}}\leq C\left(\||D|u_{0}\|_{L^{2}}^{2}+\|u_{1}\|_{H^{\frac{1}{2}}}^{2}\right)\leq C\left(\|u_{0}\|_{H^{1}}^{2}+\|u_{1}\|_{H^{\frac{1}{2}}}^{2}\right).

  2. 2.

    There exists C>0C>0, such that for any τ\tau\in{\mathbb{R}}, |τ|>C|\tau|>C, there holds

    P(τ)1L2L2C|τ|1α1.\|P(\tau)^{-1}\|_{L^{2}\to L^{2}}\leq C|\tau|^{\frac{1}{\alpha}-1}.

One would like to use the semi-group of 𝐀\mathbf{A} to solve the damped wave system (5.1), hence it is important to understand the spectrum of 𝐀\mathbf{A}. As we see below, the set of resonances of 𝐀\mathbf{A} is the same set of resonances of P(τ)P(\tau). Hence it suffices to study the resonances of P(τ)P(\tau).

Lemma 5.2.

Suppose χL(𝕋)\chi\in L^{\infty}(\mathbb{T}) and P(τ)P(\tau) is as in (1.11). Then the resolvent

P(τ)1:L2(𝕋)H1(𝕋),τP(\tau)^{-1}:L^{2}(\mathbb{T})\to H^{1}(\mathbb{T}),\ \tau\in{\mathbb{C}}

is a meromorphic family of operators with finite rank poles. Moreover, P(τ)1P(\tau)^{-1} is holomorphic in the region 𝒪{\mathbb{C}}\setminus\mathcal{O}, where

𝒪:=([χL,0]i)({0}+[12χL,0]i).\mathcal{O}:=\left(\left[-\|\chi\|_{L^{\infty}},0\right]i\right)\cup\left({\mathbb{R}}\setminus\{0\}+\left[-\tfrac{1}{2}\|\chi\|_{L^{\infty}},0\right]i\right).
Proof.

Take τ0\tau_{0}\in{\mathbb{R}}, τ00\tau_{0}\neq 0. We have the following resolvent identity

P(τ)=(I(iτχ+τ2+τ02)(|D|+τ02)1)(|D|+τ02).P(\tau)=(I-(i\tau\chi+\tau^{2}+\tau_{0}^{2})(|D|+\tau_{0}^{2})^{-1})(|D|+\tau_{0}^{2}). (5.2)

Notice that |D|+τ02:H1(𝕋)L2(𝕋)|D|+\tau_{0}^{2}:H^{1}(\mathbb{T})\to L^{2}(\mathbb{T}) is invertible. Let

Q(τ):=(iτχ+τ2+τ02)(|D|+τ02)1:L2(𝕋)L2(𝕋),τ.Q(\tau):=(i\tau\chi+\tau^{2}+\tau_{0}^{2})(|D|+\tau_{0}^{2})^{-1}:L^{2}(\mathbb{T})\to L^{2}(\mathbb{T}),\tau\in{\mathbb{C}}.

Then Q(τ)Q(\tau) is a family of compact operators, hence I+Q(τ)I+Q(\tau) is a family of Fredholm operators. For τ01\tau_{0}\gg 1, we have

Q(iτ0)L2(𝕋)L2(𝕋)χL(𝕋)/|τ0|<1.\|Q(i\tau_{0})\|_{L^{2}(\mathbb{T})\to L^{2}(\mathbb{T})}\leq\|\chi\|_{L^{\infty}(\mathbb{T})}/|\tau_{0}|<1.

Then for Imτ1\operatorname{Im}\tau\gg 1, I+Q(τ)I+Q(\tau) is a family of invertible Fredholm operators. The analytic Fredholm theory (see for instance [DZ19, Theorem C.8]) implies that

(I+Q(τ))1:L2(𝕋)L2(𝕋),τ(I+Q(\tau))^{-1}:L^{2}(\mathbb{T})\to L^{2}(\mathbb{T}),\ \tau\in{\mathbb{C}}

is a meromorphic family of operators with finite rank poles. This and (5.2) show that P(τ)1:L2(𝕋)H1(𝕋)P(\tau)^{-1}:L^{2}(\mathbb{T})\to H^{1}(\mathbb{T}) is a meromorphic family of operators with finite rank poles.

To see that P(τ)1P(\tau)^{-1} is holomorphic in 𝒪{\mathbb{C}}\setminus\mathcal{O}, we write τ=λ+iγ\tau=\lambda+i\gamma, λ,γ\lambda,\gamma\in{\mathbb{R}}, then

P(λ+iγ)=|D|λ2+γ(χ+γ)iλ(χ+2γ).P(\lambda+i\gamma)=|D|-\lambda^{2}+\gamma(\chi+\gamma)-i\lambda(\chi+2\gamma).

If λ=0\lambda=0, γ>0\gamma>0 or γ<χL\gamma<-\|\chi\|_{L^{\infty}}, then there exists c=c(γ)>0c=c(\gamma)>0 such that for uH1u\in H^{1}, we have

P(iγ)u,u|D|12uL22+cuL22.\langle P(i\gamma)u,u\rangle\geq\||D|^{\frac{1}{2}}u\|^{2}_{L^{2}}+c\|u\|_{L^{2}}^{2}.

Therefore P(iγ)u=0P(i\gamma)u=0 implies u=0u=0. Hence τ=λ+iγ\tau=\lambda+i\gamma can not be a resonance in this case.

If λ0\lambda\neq 0, γ>0\gamma>0 or γ<12χL\gamma<-\frac{1}{2}\|\chi\|_{L^{\infty}}, then for uH1u\in H^{1}, we compute

ImP(λ+iγ)u,u=λ(χ+2γ)u,u\operatorname{Im}\langle P(\lambda+i\gamma)u,u\rangle=\lambda\langle(\chi+2\gamma)u,u\rangle

Since either χ+2γ>0\chi+2\gamma>0 or χ+2γ<0\chi+2\gamma<0 for all x𝕋x\in\mathbb{T}, we again see that P(λ+iγ)u=0P(\lambda+i\gamma)u=0 implies that u=0u=0. ∎

The same result holds for 𝐀\mathbf{A}. In fact,

Lemma 5.3.

Let χL(𝕋)\chi\in L^{\infty}(\mathbb{T}), 𝐀\mathbf{A} be as in (5.1). Then the resolvent

(iτI𝐀)1:𝒟(𝐀),τ(-i\tau I-\mathbf{A})^{-1}:\mathscr{H}\to\mathcal{D}(\mathbf{A}),\ \tau\in{\mathbb{C}}

is a meromorphic family of operators with finite rank poles and the poles (resonances) are exactly the resonances of P(τ)P(\tau). In particular, 0 is a resonance of 𝐀\mathbf{A} and Ker(𝐀)={(a0)|a}\mathrm{Ker}_{\mathscr{H}}(\mathbf{A})=\left\{\left.\begin{pmatrix}a\\ 0\end{pmatrix}\ \right|\ a\in{\mathbb{C}}\right\}.

Proof.

It suffices to recall the resolvent identity

(iτI𝐀)1=(P(τ)1(χiτI)P(τ)1P(τ)1(iτχτ2I)iτP(τ)1).(-i\tau I-\mathbf{A})^{-1}=\begin{pmatrix}P(\tau)^{-1}(\chi-i\tau I)&P(\tau)^{-1}\\ P(\tau)^{-1}(-i\tau\chi-\tau^{2}I)&-i\tau P(\tau)^{-1}\end{pmatrix}. (5.3)

from [Leb94]. ∎

We now sketch the proof of Proposition 5.1 below. For further details of the proof, we refer to [AL14, §4].

Sketch proof of Proposition 5.1.

Notice that 0 is the only resonance of 𝐀\mathbf{A} with a nonnegative imaginary part – which correspond to the nondecaying part in the solution. Hence it is natural to split the eigenspace of τ=0\tau=0. For that, we let Π0:\Pi_{0}:\mathscr{H}\to\mathscr{H} be the orthogonal projection onto Ker(𝐀)\mathrm{Ker}_{\mathscr{H}}(\mathbf{A}). We define

̊:=(IΠ0),Ů:=|U|,𝐀̊:=𝐀|̊.\mathring{\mathscr{H}}:=(I-\Pi_{0})\mathscr{H},\ \|U\|_{\mathring{\mathscr{H}}}:=|U|_{\mathscr{H}},\ \mathring{\mathbf{A}}:=\mathbf{A}|_{\mathring{\mathscr{H}}}.

Then the solution to the damped wave system (5.1) can be expressed in terms of the semi-group of 𝐀̊\mathring{\mathbf{A}}:

U(t)=et𝐀̊(IΠ0)U0+Π0U0.U(t)=e^{t\mathring{\mathbf{A}}}(I-\Pi_{0})U_{0}+\Pi_{0}U_{0}. (5.4)

Notice now that

E(u,t)=|U(t)|2=et𝐀̊(IΠ0)U0̊2,E(u,t)=|U(t)|^{2}_{\mathscr{H}}=\|e^{t\mathring{\mathbf{A}}}(I-\Pi_{0})U_{0}\|^{2}_{\mathring{\mathscr{H}}},

Thus the first statement in Proposition 5.1 is equivalent to the following semi-group bound: there exists C>0C>0 such that

et𝐀̊𝐀̊1̊̊Ctα,t>C.\|e^{t\mathring{\mathbf{A}}}\mathring{\mathbf{A}}^{-1}\|_{\mathring{\mathscr{H}}\to\mathring{\mathscr{H}}}\leq\frac{C}{t^{\alpha}},\ t>C. (5.5)

We now recall the result [BT10, Theorem 2.4] by Borichev–Tomilov and realize that (5.5) is equivalent to the following resolvent estimate for 𝐀̊\mathring{\mathbf{A}}: there exists C>0C>0 such that

(iτI𝐀̊)1̊̊C|τ|1α,|τ|>C.\|(-i\tau I-\mathring{\mathbf{A}})^{-1}\|_{\mathring{\mathscr{H}}\to\mathring{\mathscr{H}}}\leq C|\tau|^{\frac{1}{\alpha}},\ |\tau|>C. (5.6)

Use the following identity on \mathscr{H} for τ0\tau\neq 0

(iτI𝐀̊)1(IΠ0)=(iτI𝐀)1(IΠ0)=(iτI𝐀)1+Π0iτ(-i\tau I-\mathring{\mathbf{A}})^{-1}(I-\Pi_{0})=(-i\tau I-\mathbf{A})^{-1}(I-\Pi_{0})=(-i\tau I-\mathbf{A})^{-1}+\frac{\Pi_{0}}{i\tau}

and we find that

|(iτI𝐀̊)1̊̊(iτI𝐀)1|C|τ|,|τ|>1.\left|\|(-i\tau I-\mathring{\mathbf{A}})^{-1}\|_{\mathring{\mathscr{H}}\to\mathring{\mathscr{H}}}-\|(-i\tau I-\mathbf{A})^{-1}\|_{\mathscr{H}\to\mathscr{H}}\right|\leq\frac{C}{|\tau|},\ |\tau|>1.

Thus (5.6) is equivalent to the bound for the resolvent of 𝐀\mathbf{A}: there exists C>0C>0 such that

(iτI𝐀)1C|τ|1α,|τ|>C.\|(-i\tau I-\mathbf{A})^{-1}\|_{\mathscr{H}\to\mathscr{H}}\leq C|\tau|^{\frac{1}{\alpha}},\ |\tau|>C. (5.7)

It remains to use (5.3) again to build the equivalence between (5.7) and the second statement of Proposition 5.1. ∎

Remark. Iterating (5.5) (see also [BD08]), we have for k1k\geq 1,

et𝐀̊𝐀̊k̊̊=(etk𝐀̊𝐀̊1)k̊̊etk𝐀̊𝐀̊1̊̊kCkkkαtkα.\|e^{t\mathring{\mathbf{A}}}\mathring{\mathbf{A}}^{-k}\|_{\mathring{\mathscr{H}}\to\mathring{\mathscr{H}}}=\left\|\left(e^{\frac{t}{k}\mathring{\mathbf{A}}}\mathring{\mathbf{A}}^{-1}\right)^{k}\right\|_{\mathring{\mathscr{H}}\to\mathring{\mathscr{H}}}\leq\|e^{\frac{t}{k}\mathring{\mathbf{A}}}\mathring{\mathbf{A}}^{-1}\|^{k}_{\mathring{\mathscr{H}}\to\mathring{\mathscr{H}}}\leq\frac{C^{k}k^{k\alpha}}{t^{k\alpha}}.

Hence if we further assume χCk2(𝕋)\chi\in C^{\frac{k}{2}}(\mathbb{T}) and (u0,u1)Hk+12(𝕋)×Hk2(𝕋)(u_{0},u_{1})\in H^{\frac{k+1}{2}}(\mathbb{T})\times H^{\frac{k}{2}}(\mathbb{T}), then (5.4) implies that

E(u,t)Ckt2kα𝐀̊k(u0u1)̊Ckt2kα(u0Hk+122+u1Hk22).E(u,t)\leq\frac{C_{k}}{t^{2k\alpha}}\left\|\mathring{\mathbf{A}}^{k}\begin{pmatrix}u_{0}\\ u_{1}\end{pmatrix}\right\|_{\mathring{\mathscr{H}}}\leq\frac{C_{k}}{t^{2k\alpha}}\left(\|u_{0}\|_{H^{\frac{k+1}{2}}}^{2}+\|u_{1}\|^{2}_{H^{\frac{k}{2}}}\right).

This means that the more regular the initial data is, the faster the energy decays.

6. asymptotics of the resonances for small damping

This section is devoted to proving Theorem 3. Since the first statement is a direct result of the resolvent bounds in Theorem 2, we focus on the proof of the second statement. The main tools we use are Grushin problems – we refer to [SZ07] for an introduction and applications of Grushin problems.

Recall the notation

P(ν,τ)=|D|iντχτ2:L2(𝕋)L2(𝕋),ν>0.P(\nu,\tau)=|D|-i\nu\tau\chi-\tau^{2}:L^{2}(\mathbb{T})\to L^{2}(\mathbb{T}),\ \nu>0.

The operator P(0,τ)P(0,\tau) has resonances ±k\pm\sqrt{k}, kk\in\mathbb{Z}, k0k\geq 0. Notice that if τ\tau is a resonance of P(ν,)P(\nu,\bullet), then τ¯-\overline{\tau} is also a resonance of P(ν,)P(\nu,\bullet). Thus we only need to consider the resonances τ\tau with Reτ0\operatorname{Re}\tau\geq 0.

Proof of the second statement of Theorem 3.

For kk\in\mathbb{Z}, k>0k>0, we propose the following Grushin problem for (ν,τ)(\nu,\tau) in a neighborhood of (0,k)2(0,\sqrt{k})\in{\mathbb{C}}^{2}:

𝒫(ν,τ):=(P(ν,τ)RR+0):H1(𝕋)×2L2(𝕋)×2,\mathscr{P}(\nu,\tau):=\begin{pmatrix}P(\nu,\tau)&R_{-}\\ R_{+}&0\end{pmatrix}:H^{1}(\mathbb{T})\times{\mathbb{C}}^{2}\to L^{2}(\mathbb{T})\times{\mathbb{C}}^{2},

where R±R_{\pm} are given by

R:2L2(𝕋),R(a1,a1):=a1eikx+a2eikx,R+:H1(𝕋)2,R+(v):=(v,eikx,v,eikx).\begin{gathered}R_{-}:{\mathbb{C}}^{2}\to L^{2}(\mathbb{T}),\ R_{-}(a_{1},a_{1}):=a_{1}e^{ikx}+a_{2}e^{-ikx},\\ R_{+}:H^{1}(\mathbb{T})\to{\mathbb{C}}^{2},\ R_{+}(v):=(\langle v,e^{ikx}\rangle,\langle v,e^{-ikx}\rangle).\end{gathered}

Here the inner product on L2(𝕋)L^{2}(\mathbb{T}) is defined by

u,v:=12π𝕋u(x)v(x)¯𝑑x.\langle u,v\rangle:=\frac{1}{2\pi}\int_{\mathbb{T}}u(x)\overline{v(x)}dx.

For k>0k>0, let Mk(τ)M_{k}(\tau) be the operator defined by

Mk(τ):=ΠP(0,τ)1Π:L2(𝕋)L2(𝕋),τ{±|,0}.M_{k}(\tau):=\Pi_{\perp}P(0,\tau)^{-1}\Pi_{\perp}:L^{2}(\mathbb{T})\to L^{2}(\mathbb{T}),\ \tau\in{\mathbb{C}}\setminus\{\pm\sqrt{\ell}\ |\ \ell\in\mathbb{Z},\ \ell\geq 0\}.

where Π:L2(𝕋)L2(𝕋)\Pi_{\perp}:L^{2}(\mathbb{T})\to L^{2}(\mathbb{T}) is the orthogonal projection onto the orthogonal complement of KerL2(P(0,k))=span(eikx,eikx)\mathrm{Ker}_{L^{2}}(P(0,\sqrt{k}))=\mathrm{span}(e^{ikx},e^{-ikx}). We have the following explicit formula for Mk(τ)M_{k}(\tau):

Mk(τ)u(x)=n±kan|n|τ2einx, when u(x)=naneinx.M_{k}(\tau)u(x)=\sum_{n\neq\pm k}\tfrac{a_{n}}{|n|-\tau^{2}}e^{inx},\ \text{ when }u(x)=\sum_{n\in\mathbb{Z}}a_{n}e^{inx}. (6.1)

From (6.1) we know Mk(τ)M_{k}(\tau) is in fact defined and holomorphic in a neighborhood of τ=k\tau=\sqrt{k}. Moreover, for τ\tau near k\sqrt{k}, Mk(τ):L2(𝕋)H1(𝕋)M_{k}(\tau):L^{2}(\mathbb{T})\to H^{1}(\mathbb{T}) is a bounded operator. Let Πk\Pi_{k} be the orthogonal projection onto KerL2(P(0,k))\mathrm{Ker}_{L^{2}}(P(0,\sqrt{k})), then we have

P(0,τ)1=Mk(τ)+Πkkτ2.P(0,\tau)^{-1}=M_{k}(\tau)+\frac{\Pi_{k}}{k-\tau^{2}}.

A direct computation then shows that 𝒫(0,τ)\mathscr{P}(0,\tau) has an inverse

(τ):=(Mk(τ)E+EE+(τ)):L2(𝕋)×2H1(𝕋)×2,\mathscr{E}(\tau):=\begin{pmatrix}M_{k}(\tau)&E_{+}\\ E_{-}&E_{-+}(\tau)\end{pmatrix}:L^{2}(\mathbb{T})\times{\mathbb{C}}^{2}\to H^{1}(\mathbb{T})\times{\mathbb{C}}^{2},

where E+(τ)=(kτ2)I2E_{-+}(\tau)=(k-\tau^{2})I_{2} and E±E_{\pm} are given by

E+:2H1(𝕋),E+(a1,a2):=a1eikx+a2eikx,E:L2(𝕋)2,E(v):=(v,eikx,v,eikx).\begin{gathered}E_{+}:{\mathbb{C}}^{2}\to H^{1}(\mathbb{T}),\ E_{+}(a_{1},a_{2}):=a_{1}e^{ikx}+a_{2}e^{-ikx},\\ E_{-}:L^{2}(\mathbb{T})\to{\mathbb{C}}^{2},\ E_{-}(v):=(\langle v,e^{ikx}\rangle,\langle v,e^{-ikx}\rangle).\end{gathered}

Notice that

P(ν,τ)P(0,k)=iντχ+kτ2.P(\nu,\tau)-P(0,\sqrt{k})=-i\nu\tau\chi+k-\tau^{2}.

Hence for (ν,τ)(\nu,\tau) in an open neighborhood of (0,k)2(0,\sqrt{k})\in{\mathbb{C}}^{2}, there holds

maxν,τ{Mk(τ)(P(ν,τ)P(0,k))H1H1(P(ν,τ)P(0,k)Mk(τ)L2L2}<12.\max_{\nu,\tau}\begin{Bmatrix}\|M_{k}(\tau)(P(\nu,\tau)-P(0,\sqrt{k}))\|_{H^{1}\to H^{1}}\\ \|(P(\nu,\tau)-P(0,\sqrt{k})M_{k}(\tau)\|_{L^{2}\to L^{2}}\end{Bmatrix}<\frac{1}{2}.

Now by [DZ19, Lemma C.3], we know 𝒫(ν,τ)\mathscr{P}(\nu,\tau) is invertible for (ν,τ)(\nu,\tau) in a neighborhood of (0,k)2(0,\sqrt{k})\in{\mathbb{C}}^{2}. We denote the inverse of 𝒫(ν,τ)\mathscr{P}(\nu,\tau) by

(ν,τ):=(E(ν,τ)E+(ν,τ)E(ν,τ)E+(ν,τ)):L2(𝕋)×2H1(𝕋)×2.\mathscr{E}(\nu,\tau):=\begin{pmatrix}E(\nu,\tau)&E_{+}(\nu,\tau)\\ E_{-}(\nu,\tau)&E_{-+}(\nu,\tau)\end{pmatrix}:L^{2}(\mathbb{T})\times{\mathbb{C}}^{2}\to H^{1}(\mathbb{T})\times{\mathbb{C}}^{2}.

Then (ν,τ)\mathscr{E}(\nu,\tau) is holomorphic near (0,k)(0,\sqrt{k}) and (0,τ)=(τ)\mathscr{E}(0,\tau)=\mathscr{E}(\tau). Recall that the invertibility of P(ν,τ):H1(𝕋)L2(𝕋)P(\nu,\tau):H^{1}(\mathbb{T})\to L^{2}(\mathbb{T}) is equivalent to the invertibility of E+(ν,τ)E_{-+}(\nu,\tau). In fact, there holds the following Schur complement formula

P1=EE+E+1E,E+1=R+P1R.P^{-1}=E-E_{+}E_{-+}^{-1}E_{-},\ E_{-+}^{-1}=-R_{+}P^{-1}R_{-}.

Since E+(ν,τ)E_{-+}(\nu,\tau) is a 2×22\times 2 matrix, its invertibility is much easier to characterize. Let (ν,τ):=detE+(ν,τ)\mathcal{L}(\nu,\tau):=\det E_{-+}(\nu,\tau), then we know \mathcal{L} is holomorphic near (0,k)(0,\sqrt{k}) and τ(ν)\tau\in\mathscr{R}(\nu) if and only if (ν,τ)=0\mathcal{L}(\nu,\tau)=0. Since E+(0,τ)=(kτ2)I1E_{-+}(0,\tau)=(k-\tau^{2})I_{1}, we know (0,τ)=(kτ2)2\mathcal{L}(0,\tau)=(k-\tau^{2})^{2}. By the Weierstrass Preparation Theorem, we know that for (ν,τ)(\nu,\tau) in an open neighborhood UkU_{k} of (0,k)2(0,\sqrt{k})\in{\mathbb{C}}^{2}, there exist holomorphic functions g0(ν)g_{0}(\nu), g1(ν)g_{1}(\nu) and 𝒩(ν,τ)\mathcal{N}(\nu,\tau) such that

(ν,τ)=𝒩(ν,τ)((τk)2+g0(ν)(τk)+g1(ν)),g0(0)=g1(0)=0,𝒩(ν,τ)0,(ν,τ)Uk.\begin{gathered}\mathcal{L}(\nu,\tau)=\mathcal{N}(\nu,\tau)((\tau-\sqrt{k})^{2}+g_{0}(\nu)(\tau-\sqrt{k})+g_{1}(\nu)),\\ g_{0}(0)=g_{1}(0)=0,\ \mathcal{N}(\nu,\tau)\neq 0,\ (\nu,\tau)\in U_{k}.\end{gathered}

Thus in UkU_{k}, the zeros (ν,τ(ν))(\nu,\tau(\nu)) of (ν,τ)\mathcal{L}(\nu,\tau) are

τk,±(ν)=kg0(ν)±g0(ν)24g1(ν)2.\tau_{k,\pm}(\nu)=\sqrt{k}-\frac{g_{0}(\nu)\pm\sqrt{g_{0}(\nu)^{2}-4g_{1}(\nu)}}{2}.

Since g0g_{0}, g1g_{1} are holomorphic functions, we have the Taylor expansion

g0(ν)24g1(ν)=νr(a0+a1ν+),r,r>0,a00.g_{0}(\nu)^{2}-4g_{1}(\nu)=\nu^{r}(a_{0}+a_{1}\nu+\cdots),\ r\in\mathbb{Z},\ r>0,\ a_{0}\neq 0. (6.2)

Thus one of the following statement is true:

  1. 1.

    Either both τk,±\tau_{k,\pm} are holomorphic functions of ν\nu for ν\nu near 00\in{\mathbb{C}}, that is,

    τk,±(ν)=k+=1a,±ν,a,±;\tau_{k,\pm}(\nu)=\sqrt{k}+\sum_{\ell=1}^{\infty}a_{\ell,\pm}\nu^{\ell},\ a_{\ell,\pm}\in{\mathbb{C}}; (6.3)
  2. 2.

    Or both τk,±\tau_{k,\pm} have power series expansion in terms of ν\sqrt{\nu} when ν\nu is near 0 and ν0\nu\neq 0:

    τk,±(ν)=k+=1b(±ν),b.\tau_{k,\pm}(\nu)=\sqrt{k}+\sum_{\ell=1}^{\infty}b_{\ell}(\pm\sqrt{\nu})^{\ell},\ b_{\ell}\in{\mathbb{C}}. (6.4)

On the other hand, we have the following expansion of E+E_{-+} (see for instance [DZ19, Lemma C.3]):

E+(ν,τ)=(kτ2)I2=1(iντ)T(τ),T(τ):=(χ(Mk(τ)χ)1eikx,eikxχ(Mk(τ)χ)1eikx,eikxχ(Mk(τ)χ)1eikx,eikxχ(Mk(τ)χ)1eikx,eikx).\begin{gathered}E_{-+}(\nu,\tau)=(k-\tau^{2})I_{2}-\sum_{\ell=1}^{\infty}(i\nu\tau)^{\ell}T_{\ell}(\tau),\\ T_{\ell}(\tau):=\begin{pmatrix}\langle\chi(M_{k}(\tau)\chi)^{\ell-1}e^{ikx},e^{ikx}\rangle&\langle\chi(M_{k}(\tau)\chi)^{\ell-1}e^{ikx},e^{-ikx}\rangle\\ \langle\chi(M_{k}(\tau)\chi)^{\ell-1}e^{-ikx},e^{ikx}\rangle&\langle\chi(M_{k}(\tau)\chi)^{\ell-1}e^{-ikx},e^{-ikx}\rangle\end{pmatrix}.\end{gathered} (6.5)

In particular, we know ν1E+(ν,k)\nu^{-1}E_{-+}(\nu,\sqrt{k}) is holomorphic near ν=0\nu=0. This indicates that ν2(ν,k)=ν2𝒩(ν,k)g1(ν)\nu^{-2}\mathcal{L}(\nu,\sqrt{k})=\nu^{-2}\mathcal{N}(\nu,\sqrt{k})g_{1}(\nu) is holomorphic. Therefore ν2g1(ν)\nu^{-2}g_{1}(\nu) is holomorphic near ν=0\nu=0 and in the expansion (6.2), we must have r2r\geq 2. As a result, in either (6.3) or (6.4), we must have

τk,±(ν)=k+c±ν+o(ν).\tau_{k,\pm}(\nu)=\sqrt{k}+{c_{\pm}}\nu+o(\nu). (6.6)

Inserting (6.6) in (6.5) gives

E+(ν,τk,±)=(2kc±ν+o(ν))I2(ikν+o(ν))(χ^(0)χ^(2k)¯χ^(2k)χ^(0))+S(ν),E_{-+}(\nu,\tau_{k,\pm})=(-2\sqrt{k}c_{\pm}\nu+o(\nu))I_{2}-(i\sqrt{k}\nu+o(\nu))\begin{pmatrix}\widehat{\chi}(0)&\overline{\widehat{\chi}(2k)}\\ \widehat{\chi}(2k)&\widehat{\chi}(0)\end{pmatrix}+S(\nu),

where S(ν)S(\nu) is a 2×22\times 2 matrix and each entry of S(ν)S(\nu) is of o(ν)o(\nu). Use the fact that (ν,τk,±(ν))=0\mathcal{L}(\nu,\tau_{k,\pm}(\nu))=0 and we find

det(2c±I2+i(χ^(0)χ^(2k)¯χ^(2k)χ^(0)))=0,\det\left(2c_{\pm}I_{2}+i\begin{pmatrix}\widehat{\chi}(0)&\overline{\widehat{\chi}(2k)}\\ \widehat{\chi}(2k)&\widehat{\chi}(0)\end{pmatrix}\right)=0,

from which we solve

c±=χ^(0)±|χ^(2k)|2i.c_{\pm}=-\frac{\widehat{\chi}(0)\pm|\widehat{\chi}(2k)|}{2}i.

This completes the proof. ∎

Remarks. 1. The expansion of τk,±\tau_{k,\pm} as in (6.3), (6.4) are special cases of Puiseux series expansion. See [Wan22, §2.4] and the references there for a brief introduction to Puiseux series and their applications to spectral problems.

2. By inserting (6.6) in (6.5), one can actually obtain successively the coefficient for any order terms in the expansions (6.3) or (6.4).

7. Numerical Simulations

We run our numerical approximations on three models of damping to highlight the theoretical results above. The three damping functions we consider are:

χ1(x)\displaystyle\chi_{1}(x) =14cos2(x),χ2(x)=e2(xπ)2,\displaystyle=\frac{1}{4}\cos^{2}(x),\quad\ \chi_{2}(x)=e^{-2(x-\pi)^{2}},
χ3(x)\displaystyle\chi_{3}(x) =18[tanh(20(xπ+14))tanh(20(xπ14))],\displaystyle=\frac{1}{8}\left[\tanh\left(20\left(x-\pi+\frac{1}{4}\right)\right)-\tanh\left(20\left(x-\pi-\frac{1}{4}\right)\right)\right],

where the function χ1\chi_{1} is a low-frequency damping, χ2\chi_{2} is a Gaussian damping with exponential decay at high frequency, and χ3\chi_{3} is a compactly supported damping function with slowly decaying Fourier modes.

The equations to (1.1) are rather straightforward to discretize as ODE systems in Fourier space, all the time dependent solvers are run by using Fourier pseudospectral methods in space with 292^{9} spatial grid points and integrating in time using the ode45 package in MATLAB with relative and absolute tolerance levels set at 10910^{-9} to ensure high accuracy of the solutions.

7.1. Demonstration of Polynomial Decay rates

We can observe that decay rates for the damped fractional wave dynamics appear to be numerically very close to the polynomial rates predicted by our estimates for well-chosen initial data. Indeed, in Figure 5, we observe relative decay properties comparing to the polynomial rate t2t^{-2} by plotting the evolution of t2E(t)t^{2}E(t) for (1.1) using initial conditions that are high frequency and localized far from the damping. In particular, we take

U(0,x)=χ3(x+π)cos(10x)U(0,x)=\chi_{3}(x+\pi)\cos(10x)

and observe that the decay rates are quite remarkably close to our predicted polynomial rates.

To highlight this polynomial behavior, in Figure 6 we plot logE/logt\log E/\log t for both χ1\chi_{1} damping and χ3\chi_{3} damping with the same initial condition as above. For the low frequency damping, χ1\chi_{1}, we observe similar decay rates to those predicted in Theorem 1, namely close to t3t^{-3}.

Refer to caption
Refer to caption
Figure 5. (Left) A plot of t2E(t)t^{2}E(t) in the setting of χ3\chi_{3} damping and (Right) a zoom in near the large tt asymptotics for U(0,x)=χ3(x+π)cos(10x)U(0,x)=\chi_{3}(x+\pi)\cos(10x).
Refer to caption
Refer to caption
Figure 6. (Left) The numerically observed values of log(E(t))/log(t)\log(E(t))/\log(t) demonstrating that the rate of decay is very close to t2t^{-2} for χ3\chi_{3} damping. (Right) A plot of log(E(t))/log(t)\log(E(t))/\log(t) in the setting of χ1\chi_{1} damping for demonstrating that the rate of decay is approaching t3t^{-3}. In both, we have taken U(0,x)=χ3(x+π)cos(10x)U(0,x)=\chi_{3}(x+\pi)\cos(10x).

7.2. Approximation of the resonances at low frequency

To begin, by implementing a low frequency approximation scheme, we can analyze the behavior of the resonances at low energy. For that we define

LN:=span{eikx||k|N},πN:L2(𝕋)LN is the orthogonal projection. L_{N}:=\mathrm{span}_{{\mathbb{C}}}\{e^{ikx}\ |\ |k|\leq N\},\ \pi_{N}:L^{2}(\mathbb{T})\to L_{N}\text{ is the orthogonal projection. }

We introduce the discretization PN(τ)P_{N}(\tau) of P(τ)P(\tau) using Fourier modes:

PN(τ):=|D|iτπNχτ2:LNLN.P_{N}(\tau):=|D|-i\tau\pi_{N}\chi-\tau^{2}:L_{N}\to L_{N}.

Since LNL_{N} is a linear space of dimension 2N+12N+1, we know for each τ\tau\in{\mathbb{C}}, PN(τ)P_{N}(\tau) is a (2N+1)×(2N+1)(2N+1)\times(2N+1) matrix. Indeed, if we expand χ\chi using Fourier series

χ(x)=nχ^(n)einx,\chi(x)=\sum_{n\in\mathbb{Z}}\widehat{\chi}(n)e^{inx},

then using the basis {eikx}|k|N\{e^{ikx}\}_{|k|\leq N} of LNL_{N}, PN(τ)P_{N}(\tau) becomes a matrix

PN(τ)=diag(|n|)|n|N(τ2+iχ^(0)τ)I2N+1iτn=12N(χ^(n)J2N+1n+χ^(n)¯(J2N+1t)n),P_{N}(\tau)=\mathrm{diag}(|n|)_{|n|\leq N}-(\tau^{2}+i\widehat{\chi}(0)\tau)I_{2N+1}-i\tau\sum_{n=1}^{2N}\left(\widehat{\chi}(n)J_{2N+1}^{n}+\overline{\widehat{\chi}(n)}(J_{2N+1}^{t})^{n}\right),

where J2N+1J_{2N+1} is the (2N+1)×(2N+1)(2N+1)\times(2N+1) upper-triangular Jordan block.

Because P(τ)P(\tau) is a quadratic polynomial of τ\tau with matrix coefficients, we know det(PN(τ))\det(P_{N}(\tau)) is polynomial of τ\tau of order 2(2N+1)2(2N+1). Therefore det(PN(τ))1\det(P_{N}(\tau))^{-1} is meromorphic function with 2(2N+1)2(2N+1) poles (counting multiplicities). As a result, the inverse matrix

PN(τ)1:LNLN,τP_{N}(\tau)^{-1}:L_{N}\to L_{N},\ \tau\in{\mathbb{C}}

is a meromorphic family of (2N+1)×(2N+1)(2N+1)\times(2N+1) matrices with 2(2N+1)2(2N+1) poles (counting multiplicities). We denote the set of poles of PN(τ)P_{N}(\tau), i.e., zeros of det(PN(τ))\det(P_{N}(\tau)), by N\mathscr{R}_{N}. We regard N\mathscr{R}_{N} as a low frequency approximation to \mathscr{R}.

For a given damping function χ\chi, we can thus numerically construct function detPN(τ){\rm det}P_{N}(\tau) whose form we can compute symbolically in MATLAB. We can then find the roots of this polynomial equation in τ\tau.

We uniformly take N=12N=12 as an approximation, and observe the following each of our three experiments. Depending upon the potential, we have a number of resonances concentrating around Im(z)=χ^j(0)\mathrm{Im}(z)=\widehat{\chi}_{j}(0) for j=1,2,3j=1,2,3 as predicted by the asymptotics in §6. The observed resonances are computed using N=12N=12. The resonance with smallest imaginary part is Im(z)=.0312,.0378,4.48×104\mathrm{Im}(z)=-.0312,-.0378,-4.48\times 10^{-4} for χj\chi_{j} with j=1,2,3j=1,2,3 respectively. We observed that these resonances were stable under refinement of the approximation parameter NN. Of course, there is 0 resonance for every example corresponding to the constant solution.

We can compare our asymptotics from §6 to the computed approximate resonances with N=12N=12 for a variety of damping functions and parameters ν\nu. We consider again three experiments again motivated by low frequency damping (χ1\chi_{1}), Gaussian damping (χ2\chi_{2}) and compactly supported damping (χ3\chi_{3}), but with varying amplitude depending upon the parameter ν\nu:

χ1,ν(x)\displaystyle\chi_{1,\nu}(x) =νcos2(x),χ2,ν(x)=νe2(xπ)2,\displaystyle=\nu\cos^{2}(x),\ \ \chi_{2,\nu}(x)=\nu e^{-2(x-\pi)^{2}},
χ3,ν(x)\displaystyle\chi_{3,\nu}(x) =ν2[tanh(20(xπ+14))tanh(20(xπ14))].\displaystyle=\frac{\nu}{2}\left[\tanh\left(20\left(x-\pi+\frac{1}{4}\right)\right)-\tanh\left(20\left(x-\pi-\frac{1}{4}\right)\right)\right].

In particular, our experiments above corresponded to ν=.25,1,.25\nu=.25,1,.25 for j=1,2,3j=1,2,3 respectively. Figures 7-9 demonstrate quite well that our asymptotics remain quite robust for each of these potentials for small ν\nu and still give a fair amount of insight especially into the real part of the resonances even for large ν\nu. These resonances give insight into how states that are low-frequency and overlapping with the damping function decay in a significant fashion under the evolution of (1.1).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7. Clockwise from top left, a plot of the low-frequency approximation of the damping function compared to the original followed by the computed approximate resonances compared to prediction generated by χ1,ν\chi_{1,\nu} for ν=.125,.25,.5,1\nu=.125,.25,.5,1 respectively.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8. Clockwise from top left, a plot of the low-frequency approximation of the damping function compared to the original followed by the computed approximate resonances compared to prediction generated by χ2,ν\chi_{2,\nu} for ν=.125,.25,.5,1\nu=.125,.25,.5,1 respectively.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9. Clockwise from top left, a plot of the low-frequency approximation of the damping function compared to the original followed by the computed approximate resonances compared to prediction generated by χ3,ν\chi_{3,\nu} for ν=.125,.25,.5,1\nu=.125,.25,.5,1 respectively.

7.3. Improved decay for higher regularity initial data

Given our approximate resonance values, if we had a resonance free strip otherwise, we should have that E(t)12eIm(z)tE(t)^{\frac{1}{2}}\sim e^{\mathrm{Im}(z)t} where EE is defined as in (1.5). We have seen that this is not expected given our observed polynomial decay rates above. However, given highly regular data, we can prove that our decay rate will be better than any polynomial as in the Remark at the end of §5. However, we cannot provide such resonance free strips with our existing propagation estimates and based upon our estimates believe that no such strip exists at high energies. Even so, we can test the evolution of (1.1) for χ1,2,3\chi_{1,2,3} and compare the observed rate of convergence for E12E^{\frac{1}{2}} to those suggested by our computed resonances with very regular initial data. These findings are displayed in Figure 10, where from left to right we plot the numerically observed decay rates of E(t)12E(t)^{\frac{1}{2}} for each χj\chi_{j}, j=1,2,3j=1,2,3. Specifically, we plot log(E(t))/2t-\log(E(t))/2t vs. tt and compare to the exponential decay rate we would expect from the low-energy resonances. To ensure we observe slow decay rates in particular, in each of these simulations, the initial data is taken to be

U(0,x)=sin(x),U(0,x)=\sin(x),

which by our approximation theory is the function that dominates the low energy resonances we can compute that are closest to the real axis. As suggested by our results, the decay rate does not fit an explicit asymptotic profile related to a specific resonance, though it does decay strongly and appears to decay at close to the expected exponential rate corresponding to the computed resonances at low energy.

Refer to caption
Refer to caption
Refer to caption
Figure 10. The numerically observed value of log(E(t))/t-\log(E(t))/t for the solution to (1.1) with U(0,x)=sin(x)U(0,x)=\sin(x) (red) compared to the resonance prediction (blue) for (Left) χ1\chi_{1} damping. (Middle) χ2\chi_{2} damping. (Right) χ3\chi_{3} damping.

7.4. The full water wave problem

To demonstrate how effective the damping we present here can be in the full model however, we also have included a model wave train from a forced-damped water wave model solved with very high precision using the techniques of [ACM+22]. We illustrate this in Figure 11. As can be seen clearly, the damping appears to have a very strong local effect and allows for nonlinear wave trains to exist stably far from the damping.

Refer to caption
Figure 11. A time slice of a forced-damped water wave train.

References

  • [ABBG+12] Fatiha Alabau-Boussouira, Roger Brockett, Olivier Glass, Jérôme Le Rousseau, Enrique Zuazua, Sylvain Ervedoza, and Enrique Zuazua. The wave equation: Control and numerics. Control of Partial Differential Equations: Cetraro, Italy 2010, Editors: Piermarco Cannarsa, Jean-Michel Coron, pages 245–339, 2012.
  • [ABHK18] Thomas Alazard, Pietro Baldi, and Daniel Han-Kwan. Control of water waves. Journal of the European Mathematical Society, 20(3):657–745, 2018.
  • [ABZ11] Thomas Alazard, Nicolas Burq, and Claude Zuily. On the water-wave equations with surface tension. Duke Mathematical Journal, 158(3):413–499, 2011.
  • [ACM+22] David M Ambrose, Roberto Camassa, Jeremy L Marzuola, Richard M McLaughlin, Quentin Robinson, and Jon Wilkening. Numerical algorithms for water waves with background flow over obstacles and topography. Advances in Computational Mathematics, 48(4):1–62, 2022.
  • [AL14] Nalini Anantharaman and Matthieu Léautaud. Sharp polynomial decay rates for the damped wave equation on the torus. Analysis & PDE, 7(1):159–214, 2014.
  • [Ala17] Thomas Alazard. Stabilization of the water-wave equations with surface tension. Annals of PDE, 3:1–41, 2017.
  • [Ala18] Thomas Alazard. Stabilization of gravity water waves. Journal de Mathématiques Pures et Appliquées, 114:51–84, 2018.
  • [Ana10] Nalini Anantharaman. Spectral deviations for the damped wave equation. Geom. Funct. Anal., 20:593–626, 2010.
  • [BD08] Charles J. K. Batty and Thomas Duyckaerts. Non-uniform stability for bounded semi-groups on Banach spaces. J. Evol. Equ., 8(4):765–780, 2008.
  • [BLR92] Claude Bardos, Gilles Lebeau, and Jeffrey Rauch. Sharp sufficient conditions for the observation, control, and stabilization of waves from the boundary. SIAM journal on control and optimization, 30(5):1024–1065, 1992.
  • [Bon81] Jean-Michel Bony. Calcul symbolique et propagation des singularités pour les équations aux dérivées partielles non linéaires. Annales scientifiques de l’École normale supérieure, 14(2):209–246, 1981.
  • [BT10] Alexander Borichev and Yuri Tomilov. Optimal polynomial decay of functions and operator semigroups. Mathematische Annalen, 347:455–478, 2010.
  • [BZ04] Nicolas Burq and Maciej Zworski. Geometric control in the presence of a black box. Journal of the American Mathematical Society, 17(2):443–471, 2004.
  • [BZ19] Nicolas Burq and Maciej Zworski. Rough controls for schrödinger operators on 2-tori. Annales Henri Lebesgue, 2:331–347, 2019.
  • [CFGK05] Didier Clamond, Dorian Fructus, John Grue, and Øyvind Kristiansen. An efficient model for three-dimensional surface wave simulations. part ii: Generation and absorption. Journal of computational physics, 205(2):686–705, 2005.
  • [DZ19] Semyon Dyatlov and Maciej Zworski. Mathematical theory of scattering resonances, volume 200. American Mathematical Soc., 2019.
  • [Hör97] Lars Hörmander. Lectures on nonlinear hyperbolic differential equations, volume 26. Springer Science & Business Media, 1997.
  • [JKR14] Geri I Jennings, Smadar Karni, and JB Rauch. Water wave propagation in unbounded domains. part i: nonreflecting boundaries. Journal of Computational Physics, 276:729–739, 2014.
  • [Joh21] Steven G Johnson. Notes on perfectly matched layers (pmls). arXiv preprint arXiv:2108.05348, 2021.
  • [Leb94] G. Lebeau. Équations des ondes amorties. In Séminaire sur les Équations aux Dérivées Partielles, 1993–1994, pages Exp. No. XV, 16. École Polytech., Palaiseau, 1994.
  • [LR97] Gilles Lebeau and Luc Robbiano. Stabilisation de l’équation des ondes par le bord. Duke Math. J., 86(3):465–491, 1997.
  • [Mac21] Fabricio Macià. Observability results related to fractional schrödinger operators. Vietnam Journal of Mathematics, 49(3):919–936, 2021.
  • [MM82] A. S. Markus and V. I. Matsaev. Comparison theorems for spectra of linear operators and spectral asymptotics. Trudy Moskov. Mat. Obshch., 45:133–181, 1982.
  • [Moo22] Gary Moon. A toy model for damped water waves. arXiv preprint arXiv:2203.16645, 2022.
  • [Nat13] Frédéric Nataf. Absorbing boundary conditions and perfectly matched layers in wave propagation problems. Direct and inverse problems in wave propagation and applications, 14:219–231, 2013.
  • [Phu07] Kim Dang Phung. Polynomial decay rate for the dissipative wave equation. Journal of Differential Equations, 240(1):92–124, 2007.
  • [RT75] Jeffrey Rauch and Michael Taylor. Decay of solutions to nondissipative hyperbolic systems on compact manifolds. Comm. Pure Appl. Math., 28(4):501–523, 1975.
  • [Sjö00] Johannes Sjöstrand. Asymptotic distribution of eigenfrequencies for damped wave equations. Publ. Res. Inst. Math. Sci., 30(5):573–611, 2000.
  • [SZ07] Johannes Sjöstrand and Maciej Zworski. Elementary linear algebra for advanced spectral problems. Ann. I. Fourier, 57(7):2095–2141, 2007.
  • [Wan22] Jian Wang. Dynamics of resonances for 0th order pseudodifferential operators. Commun. Math. Phys., 391:643–668, 2022.
  • [Zua05] Enrique Zuazua. Propagation, observation, and control of waves approximated by finite difference methods. SIAM review, 47(2):197–243, 2005.
  • [Zua07] Enrique Zuazua. Controllability and observability of partial differential equations: some results and open problems. In Handbook of differential equations: evolutionary equations, volume 3, pages 527–621. Elsevier, 2007.
  • [Zwo12] Maciej Zworski. Semiclassical analysis, volume 138. American Mathematical Society, 2012.