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

Balian-Bloch wave invariants for nearly degenerate orbits

Vadim Kaloshin 1 Illya Koval 1  and  Amir Vig 2 1 Institute of Science and Technology Austria, Klosterneuburg, Lower Austria 2 Department of Mathematics, University of Michigan, Ann Arbor, MI 48109, USA Vadim.Kaloshin@gmail.com Illya.Koval@ist.ac.at Vig@umich.edu
Abstract.

This paper is part I of a series in which we aim to show that the singular support of the wave trace and the length spectrum of a smooth, strictly convex, and bounded planar billiard table are generally distinct objects. We derive an asymptotic trace formula for the regularized resolvent which is dual to the wave trace and contains the same information. To do this, we consider a class of periodic orbits which have nearly degenerate Poincaré maps and study their leading order behavior as the deformation parameter goes to zero, generating large coefficients in the wave trace. We also keep careful track of the Maslov indices, which will allow us to match contributions of opposite signs in our subsequent paper. Each cancellation of coefficients in the resolvent trace corresponds to making the wave trace one degree smoother. The resolvent based approach is due to Balian and Bloch and was significantly expanded upon by Zelditch in a foundational series of papers [Zel09], [Zel04a], [Zel04c] and [Zel00].

1. Introduction

In this paper, we consider the Laplacian on a smooth, strictly convex and bounded planar domain Ω\Omega:

(1) {Δu=λ2uxΩBu=0,xΩ,\displaystyle\begin{cases}-\Delta u=\lambda^{2}u&x\in\Omega\\ Bu=0,&x\in\partial\Omega,\end{cases}

BB is a boundary operator corresponding to Dirichlet, Neumann or Robin boundary condtions. The even wave trace w(t)w(t) is defined by the real part of the distribution

(2) 12π+TreitΔφ^(t)𝑑t=Trφ(Δ),φ𝒮()\displaystyle\frac{1}{2\pi}\int_{-\infty}^{+\infty}\text{Tr}e^{it\sqrt{-\Delta}}\hat{\varphi}(t)dt=\text{Tr}\varphi(\sqrt{-\Delta}),\qquad\varphi\in\mathcal{S}(\mathbb{R})

and we write w(t)=Trcos(tΔ)w(t)=\text{Tr}\cos({t\sqrt{-\Delta}}). Denote by LSP(Ω)\text{LSP}(\Omega) the (unmarked) length spectrum, consisting of lengths of periodic billiard trajectories in Ω\Omega. The Poisson relation tells us that the singular support of w(t)w(t) is contained in the closure of the length spectrum together with 0 and its reflection about the origin. Assume LL is isolated in LSP(Ω)\text{LSP}(\Omega) with finite multiplicity and all corresponding periodic orbits are nondegenerate. Choose ρ^Cc\widehat{\rho}\in C_{c}^{\infty} to be identically equal to 11 in a neighborhood of LL and satisfy supp(ρ^)LSP(Ω)=L\text{supp}(\widehat{\rho})\cap\text{LSP}(\Omega)=L. We then have an asymptotic expansion of the regularized resolvent trace which has form

(3) 0ρ^(t)w(t)eikt𝑑tγ:length(γ)=L𝒟γj=0Bγ,jkj,\displaystyle\int_{0}^{\infty}\widehat{\rho}(t)w(t)e^{ikt}dt\sim\sum_{\gamma:\text{length}(\gamma)=L}\mathcal{D}_{\gamma}\sum_{j=0}^{\infty}B_{\gamma,j}k^{-j},

where the righthand side is a sum over all periodic orbits γ\gamma having length LL. We call 𝒟γ\mathcal{D}_{\gamma} the symplectic prefactor and Bγ,jB_{\gamma,j} the Balian-Bloch wave invariants.

Theorem 1.1.

Let Ωε\Omega_{\varepsilon}, ε[0,ε0]\varepsilon\in[0,\varepsilon_{0}] be a smooth one parameter family of domains which fixes the reflection points and angles of a period q3q\geq 3 billiard orbit γ\gamma. Assume further that 2\partial^{2}\mathcal{L} has rank q1q-1 in Ω0\Omega_{0}, where 2\partial^{2}\mathcal{L} is the Hessian of the length functional in arclength coordinates evaluated at γ\gamma, and that the perturbation makes γ\gamma nondegenerate in Ωε\Omega_{\varepsilon} when ε>0\varepsilon>0, with |det2|cγε|\det\partial^{2}\mathcal{L}|\sim c_{\gamma}\varepsilon for some cγ>0c_{\gamma}>0. Then, the jjth Balian-Bloch wave invariant Bj,γB_{j,\gamma} associated to γ\gamma in the regularized resolvent trace 3 is given by

eiπq42qLγ(i1,i2,i3=1qhi1hi2hi3i1,i2,i33)2j(cγε)3j=1qcosϑ|x(s)x(s+1)|(±i)jw(j)+j,\displaystyle\frac{e^{-\frac{i\pi q}{4}}}{2q}L_{\gamma}\left(\sum_{i_{1},i_{2},i_{3}=1}^{q}h_{i_{1}}h_{i_{2}}h_{i_{3}}\partial_{i_{1},i_{2},i_{3}}^{3}\mathcal{L}\right)^{2j}(c_{\gamma}\varepsilon)^{-3j}\prod_{\ell=1}^{q}\frac{\cos\vartheta_{\ell}}{|x(s_{\ell})-x(s_{\ell+1})|}\frac{(\pm i)^{j}}{w(j)}+\mathcal{R}_{j},

where

  • |x(s)x(s+1)||x(s_{\ell})-x(s_{\ell+1})| is the length of the \ellth link in the billiard trajectory.

  • ϑ\vartheta_{\ell} are the angles of reflection measured from the interior normal at \ellth impact point on the boundary.

  • w(j)=𝒢1w(𝒢)>0w(j)=\sum_{\mathcal{G}}\frac{1}{w(\mathcal{G})}>0 is a sum is over all 33-regular graphs on 2j2j vertices, with w(𝒢)w(\mathcal{G}) being the order of the automorphism group of a graph 𝒢\mathcal{G}. In particular, it is independent of both the domain and the orbit.

  • \mathcal{L} is the length functional in arclength coordinates (see Definition 3.1), LγL_{\gamma} is the length of γ\gamma and the functions hiC(Ω0)h_{i}\in C^{\infty}(\partial\Omega_{0}) are such that hi1hi2h_{i_{1}}h_{i_{2}} is, up to a sign, the (i1,i2)(i_{1},i_{2}) entry of the adjugate matrix of 2\partial^{2}\mathcal{L} at γ\gamma in Ω0\Omega_{0}, which is independent of the deformation. The factor

    i1,i2,i3hi1hi2hi3i1,i2,i33\sum_{i_{1},i_{2},i_{3}}h_{i_{1}}h_{i_{2}}h_{i_{3}}\partial_{i_{1},i_{2},i_{3}}^{3}\mathcal{L}

    is, modulo an O(ε)O(\varepsilon) error, the third directional derivative of \mathcal{L} along the null space of 2\partial^{2}\mathcal{L} in Ω0\Omega_{0}.

  • The signs ±\pm are given by ++ if the closest eigenvalue to 0 of 2\partial^{2}\mathcal{L} in Ωε\Omega_{\varepsilon} is positive and - otherwise.

  • j=(γ,ε,j,κ,κ(1),,κ(2j))\mathcal{R}_{j}=\mathcal{R}(\gamma,\varepsilon,j,\kappa,\kappa^{(1)},\cdots,\kappa^{(2j)}), with κ(i)\kappa^{(i)} being the iith derivative of the boundary curvature of Ωε\Omega_{\varepsilon}, is a remainder satisfying:

    • j=O(ε3j+1)\mathcal{R}_{j}=O(\varepsilon^{-3j+1}).

    • ε3j1j\varepsilon^{3j-1}\mathcal{R}_{j} is smooth in all parameters down to ε=0\varepsilon=0.

    • j\mathcal{R}_{j} depends on at most 2j2j derivatives of the boundary curvature in a small neighborhood of the reflection points of γ\gamma.

  • When j=0j=0, w(0)=1w(0)=1 and 0=0\mathcal{R}_{0}=0.

The symplectic prefactor associated to γ\gamma is

𝒟γ(k)=eikLγeiπsgn 2(Sγ)4|det2(Sγ)|.\displaystyle\mathcal{D}_{\gamma}(k)=\frac{e^{ikL_{\gamma}}e^{\frac{i\pi\text{sgn \,}\partial^{2}\mathcal{L}(S_{\gamma})}{4}}}{\sqrt{|\det\partial^{2}\mathcal{L}(S_{\gamma})|}}.

Our paper is part of a series in which we aim to show that the singular support of the wave trace and the length spectrum are in general distinct objects which encode different information. In part II, we will create cancellations in the wave trace by perturbing an initial domain having the properties in Theorem 1.1 and then using the asymptotic formulas above together with a careful matching of Maslov indices, which arise as complex phases. The possibility of a smooth wave trace was first posed in [DG75] and it was later remarked by Colin de Verdiere (see [Zel04b]) that apriori, there is no immediate reason why w(t)w(t) cannot be CC^{\infty} on \{0}\mathbb{R}\backslash\{0\}; it is known to be singular at t=0t=0. In a work in progress, Hezari and Zelditch construct domains for which pairs of bouncing ball orbits give infinite order cancellations in the wave trace, going further in this direction. Our result applies to a large class of convex domains, having a rather general class of degenerate orbits; see Section 5.

2. Background

Historically, the wave trace invariants have been exploited in the context of the inverse spectral problem, first posed by Mark Kac in [Kac66]: “Can one hear the shape of a drum?” Mathematically, this asks if one can recover the geometry of a domain from knowledge of its Laplace spectrum. In general the answer is no, as was shown in [Mil64], [GWW92], [Vig80], and [SM85] amongst others. However, there are many cases in which one expects the answer to be yes, perhaps the most tantalizing of which is the class of strictly convex, smooth planar domains, which we consider in the present article. There have been numerous works in this direction, perhaps the most significant of which are [Zel09] and [HZ22]. In the first article, Zelditch showed that generic families of analytic 2\mathbb{Z}_{2} symmetric planar domains can be recovered from their spectrum by a careful analysis of the wave invariants associated to iterates of a bouncing ball orbit. We follow parts of this approach closely in our paper. In the second, Hezari and Zelditch prove that ellipses of small eccentricity are spectrally determined amongst all smooth, strictly convex bounded domains by their spectrum. This is the first and only known example of noncircular domains which are spectrally determined. CC^{\infty} compactness of Laplace isospectral sets was shown in [OPS88c], [OPS88b] and [OPS88a]. Other recent progress can be found in [Zel00], [Zel98], [Zel04c], [HZ10], [HZ12], [Vig21], and [Vig23b].

The dynamical analogue of the inverse spectral problem is to determine the shape of a manifold with boundary from knowledge of the length spectrum, consisting of lengths of closed geodesics (billiard trajectories) which make elastic reflections at the boundary. For planar domains, recent progress in this direction was acheived by the first author and Sorrentino in [KS18], together with the works [ADSK16], [Kov21] and [dSKW17]. In the first three, it was shown that ellipses are isolated within the class of integrable domains. The latter two concern spectral rigidity of ellipses and nearly circular domains, respectively. CC^{\infty} compactness of marked length isospectral sets was demonstrated by the third author in [Vig23a]. Other significant results, including cases without boundary, are contained in [HKS18], [PT12], [Pop94], [PT11], [GK80], [CdV84] and [GL18].

3. Billiards

Before obtaining a singularity expansion for the wave trace, we first review the relevant background on billiards. This will be useful in our construction of the Balian-Bloch resolvent parametrix in Section 4.2. We denote by Ω\Omega a bounded and strictly convex region in 2\mathbb{R}^{2} with smooth boundary. This means that the curvature of Ω\partial\Omega is a strictly positive function. We can identify BΩB^{*}\partial\Omega with /×(0,π){\mathbb{R}}/{\ell\mathbb{Z}}\times(0,\pi), where =|Ω|\ell=|\partial\Omega| is the length of the boundary. Let x:[0,]2x:[0,\ell]\to\mathbb{R}^{2} be a unit speed parametrization of Ω\partial\Omega and denote by dsds the arclength measure on Ω\partial\Omega. The billiard map is defined on the coball bundle of the boundary, BΩ={s,σ)TΩ:|ζ|<1}B^{*}\partial\Omega=\{s,\sigma)\in T^{*}\partial\Omega:|\zeta|<1\}, which can be identified in two ways with the cosphere bundle SΩ2S_{\partial\Omega}^{*}\mathbb{R}{{}^{2}} over the boundary via the two orthogonal projection maps from inward (+)(+)/outward ()(-) pointing covectors. For σBΩ\sigma\in B^{*}\partial\Omega, define φ±\varphi_{\pm} to be the inward (+)(+)/outward ()(-) facing covectors in SΩ2S_{\partial\Omega}^{*}\mathbb{R}^{2} which project to σ\sigma.

Define the maps

β±(s,σ)=(s,σ),\displaystyle\beta_{\pm}(s,\sigma)=(s^{\prime},\sigma^{\prime}),

where ss^{\prime} is the arclength coordinate of the subsequent intersection of the line through x(s)x(s) with direction φ±\varphi_{\pm} and σ\sigma is the projection onto BΩB^{*}\partial\Omega of the parallel transport of φ±\varphi_{\pm} to x(s)x(s^{\prime}). We call β=β+\beta=\beta_{+} the billiard map. It is well known that β\beta preserves the canonical one form σds\sigma ds induced on B(Ω)B^{*}(\partial\Omega) and is differentiable there, extending continuously up to the boundary, with a square root type singularity at 0 and π\pi corresponding to a Whitney fold in its graph. As a consequence, β\beta also preserves the area form dσdsd\sigma\wedge ds. The maps β±n\beta_{\pm}^{n} are defined via iteration and it is clear that β±n=βn\beta_{\pm}^{-n}=\beta_{\mp}^{n} for nn\in\mathbb{Z}. It is also apparent that βϑ>0\frac{\partial\beta}{\partial\vartheta}>0, which means that vertical fibers in BΩB^{*}\partial\Omega are twisted upon iterating β\beta. Together, these properties make β\beta into an exact symplectic twist map. It has a particularly simple generating function

h(s,s)=|x(s)x(s)|,\displaystyle h(s,s^{\prime})=-|x(s)-x(s^{\prime})|,

by which mean that β(x,hs)=(s,hs)\beta\left(x,-\frac{\partial h}{\partial s}\right)=\left(s,\frac{\partial h}{\partial s^{\prime}}\right), which implies that ss and ss^{\prime} give local coordinates on the graph of β\beta. Geometrically, a billiard orbit corresponds to a union of line segments which are called links. A smooth closed curve 𝒞\mathcal{C} lying in Ω\Omega is called a caustic if any link drawn tangent to 𝒞\mathcal{C} remains tangent to 𝒞\mathcal{C} after an elastic reflection at the boundary of Ω\Omega. By elastic reflection, we mean that the angle of incidence equals the angle of reflection at an impact point on the boundary. We can map 𝒞\mathcal{C} onto phase space BΩB^{*}\partial\Omega to obtain a smooth closed curve which is invariant under β±\beta_{\pm}. If the dynamics are integrable, these invariant curves are precisely the Lagrangian tori which folliate phase space. A point PP in BΩB^{*}\partial\Omega is called qq-periodic (q2q\geq 2) if βq(P)=P\beta^{q}(P)=P. We define the rotation number of a qq-periodic orbit γ\gamma emanating from PB(Ω)¯P\in\overline{B^{*}(\partial\Omega)} by ω(γ)=ω(P)=pq\omega(\gamma)=\omega(P)=\frac{p}{q}, where pp is the winding number of γ\gamma, which we now define. There exists a unique lift β^\widehat{\beta} of the billiard map β\beta to the closure of the universal cover ×[0,π]\mathbb{R}\times[0,\pi] which is continuous, \ell periodic in first variable and satisfies β^(s~,0)=(s~,0)\widehat{\beta}(\widetilde{s},0)=(\widetilde{s},0). Given this normalization, for any point (s~,σ~)/×[0,π](\widetilde{s},\widetilde{\sigma})\in\mathbb{R}/\ell\mathbb{Z}\times[0,\pi] in the lift of a qq periodic orbit of β\beta, we see that β^q(s~,σ~)=(s~+p,σ~)\widehat{\beta}^{q}(\widetilde{s},\widetilde{\sigma})=(\widetilde{s}+p\ell,\widetilde{\sigma}) for some pp\in\mathbb{Z}. We define this pp to be the winding number of the orbit generated by π(s~,σ~)=(s,σ)BΩ¯\pi(\widetilde{s},\widetilde{\sigma})=(s,\sigma)\in\overline{B^{*}\partial\Omega}. Even if a point (s,σ)(s,\sigma) generates an orbit which is not periodic in the full phase space but is such that π1(βq(s,σ))=s\pi_{1}(\beta^{q}(s,\sigma))=s for some qq\in\mathbb{Z}, we can still define a winding number in this case. Such orbits are called (p,q)(p,q) geodesic loops, or loops for short. For a given periodic orbit, the winding number is independent of which point in the orbit is chosen, so we sometimes write ω(γ)=ω(P)\omega(\gamma)=\omega(P^{\prime}) for any P{P,β(P),,βq1(P)}P^{\prime}\in\{P,\beta(P),\cdots,\beta^{q-1}(P)\}. For deeper results and a more thorough introduction to the theory of dynamical billiards, we refer the reader to [Tab95], [Sib04], [Kat05], [Pop94] and [PT11].

3.1. Parametrization and Notation

Following the notation in [MM82], we may translate and rotate Ω\Omega by a rigid motion so that there exists a point pΩp\in\partial\Omega at the point (0,0)(0,0) with unit tangent vector (1,0)(1,0). Let ss denote the arclength coordinate along Ω\partial\Omega, measured counterclockwise from pp. We then define the arclength parametrization

(4) x(s)=(x1(s),x2(s))=0s(cosφ(t),sinφ(t))𝑑t,\displaystyle x(s)=(x_{1}(s),x_{2}(s))=\int_{0}^{s}(\cos\varphi(t),\sin\varphi(t))dt,

where φ\varphi is the angle made by the tangent to a base point, dtdt is arclength on Ω\partial\Omega and φ(t)=κ(t)\varphi^{\prime}(t)=\kappa(t) is the curvature of Ω\partial\Omega. When dealing with successive points of reflection in the length functional, we use the capital letters S=(s1,,sq)S=(s_{1},\cdots,s_{q}) to denote the arclength coordinates of the reflection points and x(S)x(S) to denote the corresponding boundary points (x(s1),,x(sq))(x(s_{1}),\cdots,x(s_{q})) in 2\mathbb{R}^{2}, with xi(S)x_{i}(S) being the iith coordinate. Similarly, when dealing with a periodic orbit γ\gamma, we denote by SγS_{\gamma} the arclength coordinates of the reflection points and by γ\partial\gamma the reflection points in 2\mathbb{R}^{2}.

3.2. Lengths of Orbits

There are two natural functions for studying the length spectrum. If γ\gamma is a qq-periodic orbit of the billiard map, the length of γ\gamma is then defined to be

(5) Lγ=i=1q|xi+1xi|,\displaystyle L_{\gamma}=\sum_{i=1}^{q}|x_{i+1}-x_{i}|,

where xix_{i} are the reflection points of γ\gamma on the boundary, xq+1=x1x_{q+1}=x_{1} and |||\cdot| is the Euclidean distance function.

Definition 3.1.

For an ordered collection of points S=(s1,,sq)[0,)qS=(s_{1},\cdots,s_{q})\in[0,\ell)^{q}, we define the length functional (S)\mathcal{L}(S) to be i=1q|xi+1(S)xi(S)|\sum_{i=1}^{q}|x_{i+1}(S)-x_{i}(S)|, with xq+1=x1x_{q+1}=x_{1}. For any p,q>0p,q\in\mathbb{Z}_{>0} relatively prime, we define the (p,q)(p,q) loop function p,q(s)\ell_{p,q}(s) to be the length of the locally unique broken geodesic of rotation number p/qp/q based at x(s)x(s), assuming a well defined branch of it exists (for example, when p=1p=1 and qq is sufficiently large; see Theorem 3.1 in [Vig23b]).

In the same way that h(s,s)=|x(s)x(s)|h(s,s^{\prime})=-|x(s)-x(s^{\prime})| is a generating function for β\beta, we have that (s1,,sq)\mathcal{L}(s_{1},\cdots,s_{q}) is a generating function for βq\beta^{q}. The same holds p,q\ell_{p,q}, at least locally in phase space.

Definition 3.2.

The length spectrum of Ω\Omega, denoted by LSP(Ω)\text{LSP}(\Omega), is the union of all lengths of periodic orbits.

We will also need the notion of degeneracy for periodic orbits in the Section 6, where we apply the method of stationary phase with the length functional as the phase function of an oscillatory integral.

Definition 3.3.

We say that a periodic orbit γ\gamma is nondegenerate if det(1Pγ)0\det(1-P_{\gamma})\neq 0, or equivalently if 2(Sγ)0\partial^{2}\mathcal{L}(S_{\gamma})\neq 0 (see 6 and 7 below).

3.3. Hessian Invariants

Both generating functions \mathcal{L} and p,q\ell_{p,q} have second order invariants associated to them at a critical point corresponding to a periodic orbit γ\gamma. The first is the Maslov index, which is a universal affine function of sgn2mod8\text{sgn}\partial^{2}\mathcal{L}\mod 8 coming from the complex exponential of the signature the Hessian upon performing a stationary phase expansion in 8. From a symplectic geometry perspective, they are intersection numbers of vertical fibers upon iteration by β\beta, corresponding to the twist property mentioned above. It is related to the Morse index for the length variational problem with periodic boundary conditions. References can be found in [Bia93] and [WYZ23]. From a microlocal point of view, these indices arise as unitary phases when expanding via stationary phase a Lagrangian distribution associated to the quantization of the billiard map. They come from the integer valued exponents in the unitary transition functions for the Arnold-Keller-Maslov line bundle, sections of which, when tensored with a half density, invariantly define the notion of a principal symbol for Fourier integral operators. Their relavence in the trace formula comes from the fact that the wave propagator is a Fourier integral operator microlocally near nonglancing, transversally reflected rays. The nonpositivity of these complex phases will be important for cancelling contributions to the wave trace of different orbits in our subsequent paper.

The other second order invariant is the determinant of 2\partial^{2}\mathcal{L}. It is shown in [KT91] (Theorem 3, page 67) that

(6) det2=(1)q+1|det(IdPγ)|1qbi,\displaystyle\det\partial^{2}\mathcal{L}=(-1)^{q+1}|\det(\text{Id}-P_{\gamma})|\prod_{1}^{q}b_{i},

where PγP_{\gamma} is the linearized Poincaré map of β\beta and

(7) bi=2|xi(S)xi+1(S)|sisi+1=cosϑicosϑi+1|xi(S)xi+1(S)|.\displaystyle b_{i}=\frac{\partial^{2}|x_{i}(S)-x_{i+1}(S)|}{\partial s_{i}\partial s_{i+1}}=\frac{\cos\vartheta_{i}\cos\vartheta_{i+1}}{|x_{i}(S)-x_{i+1}(S)|}.

The bib_{i} above consist of cosines of angles of reflection for the billiard orbit γ\gamma at which \mathcal{L} is evaluated and |xi(S)xi+1(S)||x_{i}(S)-x_{i+1}(S)| is the length of a link connecting the iith and i+1i+1st reflection points in γ\gamma. The angles are measured from the inward pointing normal vector at the boundary. The Poincaré map will appear naturally in the symplectic prefactor for our stationary phase expansion of the regularized resolvent trace and the product 6 enters via the multiple reflection expansion with the Hankel functions in Proposition 4.5 below.

Remark 3.4.

The use of :Ωq\mathcal{L}:\partial\Omega^{q}\to\mathbb{R} as a phase function comes naturally from the layer potential formulas in Section 4. In [MM82], [Vig20] and [HZ22], the (p,q)(p,q) loop function was used instead. Both are generating functions for the qq-th iterate of the billiard map and each presents its own difficulties. One disadvantage of \mathcal{L} is that it has singularities on the diagonal which require regularization. It also has a large number of varaiables to keep track of together with Maslov indices. The loop function is simpler in that it depends only on the base point of an orbit, but is in general multivalued and only local branches exist. Here, we primarily use \mathcal{L} as it is globally well defined and works with potentially smaller qq.

4. Balian-Bloch Invariants

Let Ω\Omega be a compact, strictly convex domain having smooth boundary and a nondegenerate periodic orbit γ\gamma of period qq and length LγL_{\gamma}, which we assume to be isolated in the length spectrum. The connection between the length and Laplace spectrums come from the Poisson relation:

SingSuppTrcostΔ{0}±LSP(Ω)¯.\displaystyle\text{SingSupp}\text{Tr}\,\cos t\sqrt{-\Delta}\subset\{0\}\cup\pm\overline{\text{LSP}(\Omega)}.

The lefthand side is the singular support of the distribution 2, which is a purely spectral quantity, while the righthand side is geometric. In the case of bounded planar domains, this result is due to Anderson and Melrose [AM77]. If ρ𝒮()\rho\in\mathcal{S}(\mathbb{R}) is such that ρ^=1\hat{\rho}=1 in a neighborhood of LγL_{\gamma} and suppρ^±LSP(Ω)¯={Lγ}\text{supp}\hat{\rho}\cap\pm\overline{\text{LSP}(\Omega)}=\{L_{\gamma}\}, then the regularized resolvent at frequency kk is given by

(8) 0eiktρ^(t)w(t)𝑑t=1iTrρkGΩ,D(k,x,y),\displaystyle\int_{0}^{\infty}e^{ikt}\widehat{\rho}(t)w(t)dt=\frac{1}{i}\text{Tr}\rho\ast kG_{\Omega,D}(k,x,y),

where GΩ,DG_{\Omega,D} is the Green’s kernel, i.e. the Schwartz Kernel of the resolvent RΩ,D(k)=(ΔΩ,Dk2)1R_{\Omega,D}(k)=(-\Delta_{\Omega,D}-k^{2})^{-1} on Ω\Omega with Dirichlet boundary conditions. This identity comes from the formula

RΩ,D(k)=ik0eiktcostΔΩ,Ddt.\displaystyle R_{\Omega,D}(k)=\frac{i}{k}\int_{0}^{\infty}e^{ikt}\cos t\sqrt{-\Delta_{\Omega,D}}dt.

The Dirichlet resolvent RΩ,DR_{\Omega,D} on Ω\Omega extends meromorphically to \mathbb{C}, with poles at the spectrum of the Laplacian and corresponding residues equal to the orthogonal projectors onto finite dimensional eigenspaces. In view of the Paley-Weiner theorem, the regularized resolvent 8 is in fact an entire function of kk. Replacing kk by k+iτk+i\tau or k+iτlogkk+i\tau\log k, the maps τeτtρ^(t)w(t)\tau\mapsto e^{-\tau t}\widehat{\rho}(t)w(t) and τkτρ^(t)w(t)\tau\mapsto k^{-\tau}\widehat{\rho}(t)w(t) give rise to continuous families of compactly supported distributions. Consequently, the regularized resolvent trace has a well defined and smooth limit as τ0+\tau\to 0^{+}. When τ=0\tau=0, we have the smoothed density of states formula

(9) 0eiktρ^(t)w(t)𝑑tπ±n=1ρ(k±λn),\displaystyle\int_{0}^{\infty}e^{ikt}\widehat{\rho}(t)w(t)dt\sim\pi\sum_{\pm}\sum_{n=1}^{\infty}\rho(k\pm\lambda_{n}),

where λn2\lambda_{n}^{2} are the eigenvalues of Δ-\Delta, which provides a connection between the high frequency behavior of Δ-\Delta and the wave trace. From positivity of the spectrum, it is clear that the term in the sum with ρ(k+λn)=O(k)\rho(k+\lambda_{n})=O(k^{-\infty}) is asymptotically negligible. The order of a singularity in the wave trace at a particular length is related to the decay properties of the correspondingly regularized resolvent trace. It follows that the expression 8 has an asymptotic expansion in kk of the form

(10) Fγ(k)j=0B~γ,jkj,\displaystyle F_{\gamma}(k)\sum_{j=0}^{\infty}\widetilde{B}_{\gamma,j}k^{-j},

where

(11) Fγ(k)=(1)εB(γ)c0Lγ#eikLγeiπmγ/4|det(IPγ)|\displaystyle F_{\gamma}(k)=(-1)^{\varepsilon_{B}(\gamma)}\frac{c_{0}L_{\gamma}^{\#}e^{ikL_{\gamma}}e^{i\pi m_{\gamma}/4}}{\sqrt{|\det(I-P_{\gamma})}|}

is called the principal symplectic prefactor. Here, c0c_{0} is a universal constant, Lγ=LL_{\gamma}=L the length of an isolated nondegenerate periodic orbit γ\gamma, Lγ#L_{\gamma}^{\#} the primitive period of γ\gamma, mγm_{\gamma} the Maslov index of γ\gamma, PγP_{\gamma} the linearized Poincaré map and εB(γ)\varepsilon_{B}(\gamma) depends on boundary conditions (see [Zel09]). The coefficients B~γ,j\widetilde{B}_{\gamma,j} are called the principal Balian-Bloch invariants associated to γ\gamma, sometimes also called wave trace invariants. The resolvent trace expansion was originally investigated by physicists Balian and Bloch in [BB74], [BB71], [BB72]. The invariants B~j\widetilde{B}_{j} are clearly equivalent to usual wave trace invariants aγ,ka_{\gamma,k} associated to w(t)w(t) near the length spectrum in Theorem 4.2 below. They are independent of the choice of ρ\rho subject to the constraints above, since only the singularity of w(t)w(t) at LγL_{\gamma} affects asymptotics of 10 upon expanding via stationary phase expansion the integral 8.

Our end goal is to show that one can find domains Ω\Omega and lengths of high multiplicity, such that several families of corresponding orbits have Maslov factors of opposite signs. We can then perturb Ω\partial\Omega in such a way that the symplectic prefactors together with B~γ,j\widetilde{B}_{\gamma,j}’s cancel for arbitrarily large jj. If the asymptotic expansion 10 is O(k)O(k^{-\infty}), then w(t)w(t) is CC^{\infty} near LL. If the decay is only O(km)O(k^{-m}) for some m2m\geq 2, then wCm2w\in C^{m-2} locally near LγL_{\gamma}. One difficulty arises from the nonreality of the coefficients BjB_{j}, whose complex phases may also depend on the corresponding orbit and could potentially negate whatever complex phase appears from the Maslov index. This is known not to be the case for the leading order invariant B0B_{0}, but not necessarily for BjB_{j} with j1j\geq 1.

We introduce the following notation for simplicity:

Definition 4.1.

Define 𝒟B,γ\mathcal{D}_{B,\gamma} and Bγ,jB_{\gamma,j} by the formula

Fγ(k)j=0B~γ,jkj=𝒟γ(k)j=0Bγ,jkj,\displaystyle F_{\gamma}(k)\sum_{j=0}^{\infty}\widetilde{B}_{\gamma,j}k^{-j}=\mathcal{D}_{\gamma}(k)\sum_{j=0}^{\infty}{B}_{\gamma,j}k^{-j},

where

𝒟γ(k)=c0eikLγeiπsgn2/4|det2|\displaystyle\mathcal{D}_{\gamma}(k)=\frac{c_{0}e^{ikL_{\gamma}}e^{i\pi\text{sgn}\partial^{2}\mathcal{L}/4}}{\sqrt{|\det\partial^{2}\mathcal{L}|}}

and the Bγ,jB_{\gamma,j} are related to the B~γ,j\widetilde{B}_{\gamma,j} as explained above.

The Bγ,jB_{\gamma,j} are called modified Balian Bloch invariants and we will often refer to them simply as the Balian-Bloch invariants when there is no risk of confusion. The reason for introducing 𝒟B,γ\mathcal{D}_{B,\gamma} and Bγ,jB_{\gamma,j} is that they appear more naturally when expanding via stationary phase a microlocal parametrix for the resolvent in Section 6. The corresponding oscillatory integral for the regularized resolvent trace produces terms canonically equivalent to |det(IdPγ)||\det(\text{Id}-P_{\gamma})| and mγm_{\gamma} in 11, but they appear more naturally in terms of the coordinates x(S)x(S) introduced in Section 3.

4.1. Relation to wave invariants

We say that the length LL\in\mathbb{R} of a periodic orbit γ\gamma is simple if up to time reversal (ttt\mapsto-t), γ\gamma is the unique periodic orbit of length LL. Without length spectral simplicity, there is no way to deduce Laplace spectral information from the length spectrum alone. It is shown in [PS92] that generically, smooth convex domains have simple length spectrum and only nondegenerate periodic orbits. In that case, the following theorem holds:

Theorem 4.2 ([GM79], [PS17]).

Assume γ\gamma is a nondegenerate periodic billiard orbit in a bounded, strictly convex domain with smooth boundary and γ\gamma has length LL which is simple. Then near LL, the even wave trace has an asymptotic expansion of the form

TrcostΔRe{aγ,0(tL+i0)1+k=0aγ,k(tL+i0)klog(tL+i0)},\displaystyle\text{Tr}\cos t\sqrt{-\Delta}\sim\operatorname{Re}\left\{a_{\gamma,0}(t-L+i0)^{-1}+\sum_{k=0}^{\infty}a_{\gamma,k}(t-L+i0)^{k}\log(t-L+i0)\right\},

where the coefficients aγka_{\gamma k} are wave invariants associated to γ\gamma. The leading order term is given by

aγ,0=eiπmγ/4Lγ#|det(IPγ)|1/2.\displaystyle a_{\gamma,0}=\frac{e^{i\pi m_{\gamma}/4}L_{\gamma}^{\#}}{|\det(I-P_{\gamma})|^{1/2}}.

Putting the formula from 4.2 into 8, we recover the asymptotics in the Balian-Bloch expansion. An algebraic formula for one set of invariants in terms of the other can be found using basic identities in Fourier analysis.

4.2. Layer Potentials

Here we fix a spectral parameter λ=k+iτ+\lambda=k+i\tau\in\mathbb{C}_{+} with τ>0\tau>0 and follow [Tay13], [Zel09] in describing potential theory for the resolvent. As before, we let dsds denote arclength along the boundary Ω\partial\Omega. From potential theory in the plane (see Proposition 4.7), we have the following formula for the interior Dirichlet resolvent:

(12) RΩ,D(λ)=𝟙Ω(R0(λ)2𝒟(λ)(Id+𝒩(λ))1rΩ𝒮(λ))𝟙Ω,\displaystyle R_{\Omega,D}(\lambda)=\mathds{1}_{\Omega}\left(R_{0}(\lambda)-2\mathcal{D}\ell(\lambda)(\text{Id}+\mathcal{N}(\lambda))^{-1}r_{\partial\Omega}\mathcal{S}\ell^{\dagger}(\lambda)\right)\mathds{1}_{\Omega},

where

𝒮(λ)f(x)=ΩG0(λ,x,y)f(y)𝑑s(y),x2\Ω\displaystyle\mathcal{S}\ell(\lambda)f(x)=\int_{\partial\Omega}G_{0}(\lambda,x,y)f(y)ds(y),\qquad x\in\mathbb{R}^{2}\backslash\partial\Omega
𝒟(λ)f(x)=ΩνsG0(λ,x,y)f(y)ds(y),x2\Ω\displaystyle\mathcal{D}\ell(\lambda)f(x)=\int_{\partial\Omega}\partial_{\nu_{s}}G_{0}(\lambda,x,y)f(y)ds(y),\qquad x\in\mathbb{R}^{2}\backslash\partial\Omega

are the single and double layer potentials at frequency λ\lambda and the boundary operator 𝒩(λ)\mathcal{N}(\lambda) is given by

(13) 𝒩(λ)f(x)=2ΩνsG0(λ,x,y)f(y)ds(y).xΩ\displaystyle\mathcal{N}(\lambda)f(x)=2\int_{\partial\Omega}\partial_{\nu_{s^{\prime}}}G_{0}(\lambda,x,y)f(y)ds(y).\qquad x\in\partial\Omega

Here, G0(λ,x,y)G_{0}(\lambda,x,y) is the free Green’s function, i.e. the Schwartz kernel of the free outgoing resolvent on 2\mathbb{R}^{2} at frequency λ\lambda, evaluated at x2x\in\mathbb{R}^{2} and y=y(s)Ωy=y(s)\in\partial\Omega. The ±\pm notation indicates limits taken from within Ω\Omega (resp. Ωc\Omega^{c}). The following jump formula holds across Ω\partial\Omega:

𝒟(λ)f±=12(±Id+𝒩(λ))f.\displaystyle\mathcal{D}\ell(\lambda)f_{\pm}=\frac{1}{2}(\pm\text{Id}+\mathcal{N}(\lambda))f.
Remark 4.3.

Note that the signs of the double and boundary layer potentials depend on the choice of unit normal to the boundary. We will use the outward pointing unit normal moving forward, for both the interior and exterior domains.

The free Green’s function on 2\mathbb{R}^{2} is given by

(14) G0(λ,x,y)=i4H0(1)(λ|xy|),\displaystyle G_{0}(\lambda,x,y)=\frac{i}{4}H_{0}^{(1)}(\lambda|x-y|),

where Hν(1)H_{\nu}^{(1)} is a Hankel function of the first kind (of order ν\nu) given by Jν(1)+iYν(1)J_{\nu}^{(1)}+iY_{\nu}^{(1)}; Jν(1),Yν(1)J_{\nu}^{(1)},Y_{\nu}^{(1)} are Bessel functions of the first and second kind respectively. The free Greens function solves the problem

(Δ2λ2)G0(λ,x,y)=δ0(xy).\displaystyle(-\Delta_{\mathbb{R}^{2}}-\lambda^{2})G_{0}(\lambda,x,y)=\delta_{0}(x-y).
Remark 4.4.

In even dimensions, the free resolvent R0(λ)R_{0}(\lambda) is only holomorphic on the Riemann surface of the logarithm rather than all of \mathbb{C}, so taking λ=k+iτ\lambda=k+i\tau\in\mathbb{C} with Imλ>0\operatorname{Im}\lambda>0 allows us to consider only the the outgoing resolvent corresponding to H0(λ|xy|)H_{0}(\lambda|x-y|) as opposed to H0(λ|xy|)H_{0}(-\lambda|x-y|), since it is bounded on L2(2)L^{2}(\mathbb{R}^{2}). After using layer potentials to obtain an explicit parametrix for the interior Dirichlet resolvent on Ω\Omega, we can let τ0\tau\to 0 using meromorphy of RΩ,DR_{\Omega,D} and holomorphy of the regularized resolvent trace.

Proposition 4.5.

For Ω2\Omega\subset\mathbb{R}^{2} with the Euclidean Laplacian, we have

(15) 𝒩(λ,x(s),x(s))λ1/2eiλ|x(s)x(s)|+3πi/4|x(s)x(s)|1/2cosϑa1(λ|x(s)x(s)|),\displaystyle\begin{split}\mathcal{N}(\lambda,x(s),x(s^{\prime}))\sim\lambda^{1/2}e^{i\lambda|x(s)-x(s^{\prime})|+3\pi i/4}|x(s)-x(s^{\prime})|^{-1/2}\cos\vartheta a_{1}(\lambda|x(s)-x(s^{\prime})|),\end{split}

where ϑ\vartheta is the angle made by the link with the interior unit normal at x(s)x(s^{\prime}) and a1a_{1} is a semiclassical symbol having the asymptotic expansion

a1(z)\displaystyle a_{1}(z) m=0imcmzm,c0=12π,\displaystyle\sim\sum_{m=0}^{\infty}i^{m}c_{m}z^{-m},\qquad c_{0}=\frac{1}{\sqrt{2\pi}},
cm\displaystyle c_{m} =(412)(432)(4(2m1)2)m!8m2π.\displaystyle=\frac{(4-1^{2})(4-3^{2})\cdots(4-(2m-1)^{2})}{m!8^{m}\sqrt{2\pi}}.

The symbol a1(λ|x(s)x(s)|)a_{1}(\lambda|x(s)-x(s^{\prime})|) is holomorphic in λ\lambda on \(,0]\mathbb{C}\backslash(-\infty,0]. Tor any fixed ϱ(0,1/2)\varrho\in(0,1/2) and χCc()\chi\in C_{c}^{\infty}(\mathbb{R}) supported on [1,1][-1,1], (1χ(kϱz))a1((k+iτ)z)(1-\chi(k^{\varrho}z))a_{1}((k+i\tau)z) belongs to the symbol class Sϱ0()S_{\varrho}^{0}(\mathbb{R}):

|zαkβ((1χ(kϱz))a0((k+iτ)z))|Cα,β,ϱ,τ,Kkϱ|α||β|\displaystyle\left|\partial_{z}^{\alpha}\partial_{k}^{\beta}\left((1-\chi(k^{\varrho}z))a_{0}((k+i\tau)z)\right)\right|\leq C_{\alpha,\beta,\varrho,\tau,K}\langle k\rangle^{\varrho|\alpha|-|\beta|}

for all KK\subset\subset\mathbb{R} compact and α,β,τ\alpha,\beta,\tau.

Proof.

Using the well known formula

ddzHν(1)(z)=νHν(1)(z)zHν+1(1)(z),\displaystyle\frac{d}{dz}H_{\nu}^{(1)}(z)=\frac{\nu H_{\nu}^{(1)}(z)}{z}-H_{\nu+1}^{(1)}(z),

and differentiating G0G_{0} to obtain the boundary layer operator, we obtain

𝒩(λ,x(s),x(s))=iλ2cosϑH1(1)(λ|x(s)x(s)|).\displaystyle\mathcal{N}(\lambda,x(s),x(s^{\prime}))=-\frac{i\lambda}{2}\cos\vartheta H_{1}^{(1)}(\lambda|x(s)-x(s^{\prime})|).

The cosϑ\cos\vartheta term comes from the calculation

y|xy|=yx|xy|,\displaystyle\nabla_{y}|x-y|=\frac{y-x}{|x-y|},

which is the unit vector in the direction of the link xy¯\overline{xy}. The asymptotics of H1(1)(λ|x(s)x(s)|)H_{1}^{(1)}(\lambda|x(s)-x(s^{\prime})|), including the coefficients cmc_{m} and the phase 3π/4-3\pi/4, are given in NIST:

12H1(1)(z)eiz3πi/4zmimcmzm.\displaystyle\frac{1}{2}H_{1}^{(1)}(z)\sim\frac{e^{iz-3\pi i/4}}{\sqrt{z}}\sum_{m}i^{m}c_{m}z^{-m}.

Combining with i-i gives a total phase of +3πi/4+3\pi i/4. That a1S00a_{1}\in S_{0}^{0} follows immediately from the asymptotic expansion in the statement of the proposition. Differentiating term by term gives

|zαkβa0(kz)|m=max{α,β}cm(m!)2((mα)!(mβ)!)kmβzmαkβzα,\displaystyle\left|\partial_{z}^{\alpha}\partial_{k}^{\beta}a_{0}(kz)\right|\leq\sum_{m=\max{\{\alpha,\beta\}}}^{\infty}c_{m}\frac{(m!)^{2}}{((m-\alpha)!(m-\beta)!)}k^{-m-\beta}z^{-m-\alpha}\lesssim\langle k\rangle^{-\beta}z^{-\alpha},

as kk\to\infty, zz away from 0. Here, k=(1+k2)1/2\langle k\rangle=(1+k^{2})^{1/2} is the Japanese bracket. Differentiating (1χ(kϱz))(1-\chi(k^{\varrho}z)) clearly shows that it is in Sϱ0S_{\varrho}^{0}, so (1χ(kϱz))a1(kz)Sϱ0(1-\chi(k^{\varrho}z))a_{1}(kz)\in S_{\varrho}^{0}. ∎

Remark 4.6.

The cutoff factor was introduced to localize asymptotics away from zero, where the asymptotic expansion isn’t useful. The introduction of the parameter ϱ\varrho is important for the decomposition of 𝒩\mathcal{N} in Proposition 4.10 below, in which the resolvent is decomposed into a sum of microlocal (homogeneous) pseudodifferential operators and semiclassical Fourier integral operators at scale =1/k\hbar=1/k. This decomposition is only valid when working with symbol classes which have ϱ>0\varrho>0.

In calculations below, we will use the function a1a_{1} in formula 4.5 to avoid extra combinatorial constants while keeping track of complex phases.

4.3. Reduction to the boundary

We now derive the way in which layer potentials translate the interior problem to one defined only on the boundary, following the presentation in [Zel04a]. To do so, we adapt the Poisson integral solution of the boundary value problem considered in [Tay13] to the Dirichlet/Neumann resolvents:

Proposition 4.7.

The Dirichlet resolvent on Ω\Omega and the Neumann resolvent on the exterior domain Ωc\Omega^{c} are given by

RΩ,D(λ)=\displaystyle R_{\Omega,D}(\lambda)= 𝟙Ω(R0(λ)+2𝒟(λ)(Id+𝒩Ω(λ))1rΩR0(λ))𝟙Ω,\displaystyle\mathds{1}_{\Omega}\left(R_{0}(\lambda)+2\mathcal{D}\ell(\lambda)(\text{Id}+\mathcal{N}_{\Omega}(\lambda))^{-1}r_{\partial\Omega}R_{0}(\lambda)\right)\mathds{1}_{\Omega},
RΩc,N(λ)=\displaystyle R_{\Omega^{c},N}(\lambda)= 𝟙Ωc(R0(λ)2𝒮(λ)(Id𝒩Ωc#(λ)))1rΩcνΩcR0(λ))𝟙Ωc,\displaystyle\mathds{1}_{\Omega^{c}}\left(R_{0}(\lambda)-2\mathcal{S}\ell(\lambda)(\text{Id}-\mathcal{N}_{\Omega^{c}}^{\#}(\lambda)))^{-1}r_{\partial\Omega^{c}}\partial_{\nu_{\Omega^{c}}}R_{0}(\lambda)\right)\mathds{1}_{\Omega^{c}},

where

𝒩Ω#g(x)=2Ωνx,ΩG0(λ,x,y)g(y)ds(y),xΩ\displaystyle\mathcal{N}_{\Omega}^{\#}g(x)=2\int_{\partial\Omega}\partial_{\nu_{x,\Omega}}G_{0}(\lambda,x,y)g(y)ds(y),\qquad x\in\partial\Omega
𝒩Ωc#g(x)=2Ωcνx,ΩcG0(λ,x,y)g(y)ds(y).xΩc\displaystyle\mathcal{N}_{\Omega^{c}}^{\#}g(x)=2\int_{\partial\Omega^{c}}\partial_{\nu_{x,\Omega^{c}}}G_{0}(\lambda,x,y)g(y)ds(y).\qquad x\in\partial\Omega^{c}
Proof.

We begin with solutions for the boundary value problems given in [Tay13], keeping in mind our convention on exterior normals. For brevity, put P(λ)=Δλ2P(\lambda)=-\Delta-\lambda^{2}:

{P(λ)u(x)=0,rΩu=f,u=2𝒟(λ)(Id+𝒩(λ))1f,(interior Dirichlet)\displaystyle\begin{cases}P(\lambda)u(x)=0,\\ r_{\partial\Omega}u=f,\end{cases}\implies u=-2\mathcal{D}\ell(\lambda)(\text{Id}+\mathcal{N}(\lambda))^{-1}f,\qquad(\text{interior Dirichlet})
{P(λ)v(x)=0,rΩcνv=φ,v=2𝒮(λ)(Id𝒩Ωc#(λ))1φ.(exterior Neumann)\displaystyle\begin{cases}P(\lambda)v(x)=0,\\ r_{\partial\Omega^{c}}\partial_{\nu}v=\varphi,\end{cases}\implies v=2\mathcal{S}\ell(\lambda)(\text{Id}-\mathcal{N}_{\Omega^{c}}^{\#}(\lambda))^{-1}\varphi.\qquad(\text{exterior Neumann})

Notice that 𝒩Ω#\mathcal{N}_{\Omega}^{\#} is defined with the outward pointing normal for Ωc\Omega^{c}, and hence the inward pointing normal for Ω\Omega. To obtain the resolvents, set G~=RΩ,D(λ)R0(λ)\widetilde{G}=R_{\Omega,D}(\lambda)-R_{0}(\lambda) and G~~(λ)=RΩc,N(λ)R0(λ)\widetilde{\widetilde{G}}(\lambda)=R_{\Omega^{c},N}(\lambda)-R_{0}(\lambda). These solve the interior Dirichlet and exterior Neumann problems with f=rΩR0(λ)f=-r_{\partial\Omega}R_{0}(\lambda) and φ=rΩνR0(λ)\varphi=-r_{\partial\Omega}\partial_{\nu}R_{0}(\lambda). Applying the layer potentials to each and adding back R0R_{0} completes the proof. ∎

To combine the interior and exterior resolvents into a similar form, note that for the exterior problem, the outer normal points in the opposite direction and 𝒩Ω(λ)=𝒩Ωc(λ)\mathcal{N}_{\Omega}(\lambda)=-\mathcal{N}_{\Omega^{c}}(\lambda). The symmetry G0(λ,x,y)=G0(λ,y,x)G_{0}(\lambda,x,y)=G_{0}(\lambda,y,x) coming from the explicit form of the resolvent in 14 implies that 𝒩Ω#=𝒩Ω\mathcal{N}_{\Omega}^{\#}=\mathcal{N}_{\Omega}^{\dagger} and 𝒩Ωc#=𝒩Ωc\mathcal{N}_{\Omega^{c}}^{\#}=\mathcal{N}_{\Omega^{c}}^{\dagger}. We also have that (rΩcνR0(λ))=𝒟(λ)(r_{\partial\Omega^{c}}\partial_{\nu}R_{0}(\lambda))^{\dagger}=\mathcal{D}\ell(\lambda) and rΩR0(λ)=𝒮(λ)r_{\partial\Omega}R_{0}(\lambda)=\mathcal{S}\ell^{\dagger}(\lambda). The free resolvent R0(λ)R_{0}(\lambda) is formally self adjoint, so taking the transpose of RΩc,N(λ)R_{\Omega^{c},N}(\lambda) and putting in cutoffs to the interior/exterior, we obtain

(16) RΩ,D(λ)=𝟙Ω(R0(λ)+2𝒟(λ)(Id+𝒩(λ))1rΩR0(λ))𝟙Ω,RΩc,N(λ)=𝟙Ωc(R0(λ)+2𝒟(λ)(Id+𝒩(λ))1rΩR0(λ))𝟙Ωc,\displaystyle\begin{split}R_{\Omega,D}(\lambda)=&\mathds{1}_{\Omega}\left(R_{0}(\lambda)+2\mathcal{D}\ell(\lambda)(\text{Id}+\mathcal{N}(\lambda))^{-1}r_{\partial\Omega}R_{0}(\lambda)\right)\mathds{1}_{\Omega},\\ R_{\Omega^{c},N}^{\dagger}(\lambda)=&\mathds{1}_{\Omega^{c}}\left(R_{0}(\lambda)+2\mathcal{D}\ell(\lambda)(\text{Id}+\mathcal{N}(\lambda))^{-1}r_{\partial\Omega}R_{0}(\lambda)\right)\mathds{1}_{\Omega^{c}},\end{split}

where rΩr_{\partial\Omega} is the operator restricting to the boundary and now all operators are defined using the outward pointing normal from Ω\Omega. Adding the two and subtracting the free resolvent yields

(17) Tr(RΩ,D(λ)+RΩc,N(λ)R0(λ))=2Tr(rΩR0(λ)(𝟙Ω2+𝟙Ωc2)𝒟(λ)(Id+𝒩(λ))1),\displaystyle\text{Tr}\,\left(R_{\Omega,D}(\lambda)+R_{\Omega^{c},N}^{\dagger}(\lambda)-R_{0}(\lambda)\right)=2\text{Tr}\,\left(r_{\partial\Omega}R_{0}(\lambda)(\mathds{1}_{\Omega}^{2}+\mathds{1}_{\Omega^{c}}^{2})\mathcal{D}\ell(\lambda)(\text{Id}+\mathcal{N}(\lambda))^{-1}\right),

where we commuted rΩR0(λ)r_{\partial\Omega}R_{0}(\lambda) through the trace to obtain a kernel on Ω×Ω\partial\Omega\times\partial\Omega. Note that the off diagonal terms are trace free. Observe also that the leftmost factor in the righthand side of 17 has the property that

2rΩR0(λ)𝒟(λ)\displaystyle 2r_{\partial\Omega}R_{0}(\lambda)\mathcal{D}\ell(\lambda) =2rΩR0(λ)ν2R0(λ)rΩ\displaystyle=2r_{\partial\Omega}R_{0}(\lambda)\circ\partial_{\nu_{2}}R_{0}(\lambda)r_{\partial\Omega}
=2rΩ12λddλν2R0(λ)rΩ=212λddλ(12𝒩(λ)),\displaystyle=2r_{\partial\Omega}\frac{1}{2\lambda}\frac{d}{d\lambda}\partial_{\nu_{2}}R_{0}(\lambda)r_{\partial\Omega}=2\frac{1}{2\lambda}\frac{d}{d\lambda}\left(\frac{1}{2}\mathcal{N}(\lambda)\right),

where we used the identity

ddλR0(λ)=2λR02(λ),\displaystyle\frac{d}{d\lambda}R_{0}(\lambda)=2\lambda R_{0}^{2}(\lambda),

together with the fact that ν2\partial_{\nu_{2}} is differentiating in only one of the variables. Hence, the righthand side of 17 adds up to

(18) 12λTrddλlog(Id+𝒩(λ))=12λddλlogdet(Id+𝒩(λ)).\displaystyle\frac{1}{2\lambda}\text{Tr}\,\frac{d}{d\lambda}\log(\text{Id}+\mathcal{N}(\lambda))=\frac{1}{2\lambda}\frac{d}{d\lambda}\log\det(\text{Id}+\mathcal{N}(\lambda)).

It is shown in [Zwo98], [Sjö97] and [Chr17] that the regularized exterior resolvent trace Tr(RΩc,N(λ)R0(λ))\text{Tr}(R_{\Omega^{c},N}(\lambda)-R_{0}(\lambda)) admits an asymptotic expansion in negative powers of λ\lambda for each periodic orbit in Ωc\Omega^{c}. By convexity of Ω\Omega, there are only gliding orbits along Ω\partial\Omega and no transversally reflected orbits in the exterior domain. In particular, for ρ^\widehat{\rho} supported away from |Ω|\mathbb{Z}|\partial\Omega|, we have

0eitkρ^(t)costΔΩc,N=O(k).\displaystyle\int_{0}^{\infty}e^{itk}\widehat{\rho}(t)\cos t\sqrt{-\Delta_{\Omega^{c},N}}=O(k^{-\infty}).

Convolving 18 with the test function iλρ(λ)-i\lambda\rho(\lambda) gives

(19) γ𝒟γ(k)j=0Bγ,jkj12iρ(kμ)μlogdet(Id+𝒩(μ+iτ))1dμ.\displaystyle\sum_{\gamma}\mathcal{D}_{\gamma}(k)\sum_{j=0}^{\infty}{B}_{\gamma,j}k^{-j}\sim\frac{1}{2i}\int_{-\infty}^{\infty}\rho(k-\mu)\frac{\partial}{\partial\mu}\log\det(\text{Id}+\mathcal{N}(\mu+i\tau))^{-1}d\mu.

Verification that the relevant functional analytic properties are satisfied to make the above formal computations legitimate (e.g. trace class, Fredholm, etc.) are contained in [Zel04a] and [Zel04c].

4.4. Regularization

As in the introduction, let ρ^Cc()\widehat{\rho}\in C_{c}^{\infty}(\mathbb{R}) be identically equal to 11 in a neighborhood of an isolated length LLSP(Ω)L\in\text{LSP}(\Omega), and zero outside of a small interval (Lε,L+ε)(L-\varepsilon,L+\varepsilon) such that suppρ^LSP¯={L}\text{supp}\widehat{\rho}\cap\overline{\text{LSP}}=\{L\}. Convolving the regularized trace above with ρ\rho and expanding 𝒩(λ)\mathcal{N}(\lambda) in a finite Neumann series yields the following formula connecting the resolvent trace asymptotics to billiard dynamics:

Proposition 4.8 ([Zel09], Proposition 3.6).

Let χγ(x,Re(λ)1Dx)\chi_{\partial\gamma}(x,\operatorname{Re}(\lambda)^{-1}D_{x}) be a semiclassical pseudodifferential operator microlocalizing in phase space near the projection of an orbit γ\gamma having length LL to BΩB^{*}\partial\Omega. Then, for each J+J\in\mathbb{N}^{+}, there exists M0(J)M_{0}(J) such that

(20) Trρ(kζ)NM0+1(ζ+iτ)(Id+𝒩(ζ+iτ))1𝒩(ζ+iτ)χγ(ζ)𝑑ζ=O(k(J+1)).\displaystyle\begin{split}\text{Tr}\int_{\mathbb{R}}\rho(k-\zeta)N^{M_{0}+1}(\zeta+i\tau)(\text{Id}+\mathcal{N}(\zeta+i\tau))^{-1}\mathcal{N}^{\prime}(\zeta+i\tau)\chi_{\partial\gamma}(\zeta)d\zeta=O(k^{-(J+1)}).\end{split}

In particular,

(21) 12iM=0M0(1)MTrρ(kζ)𝒩M(ζ+iτ)𝒩(ζ+iτ)χγ(ζ)𝑑ζ=𝒟B,γ(k+iτ)j=0JBγ,jkj+O(k(J+1)).\displaystyle\begin{split}\frac{1}{2i}\sum_{M=0}^{M_{0}}(-1)^{M}\text{Tr}&\int_{\mathbb{R}}\rho(k-\zeta)\mathcal{N}^{M}(\zeta+i\tau)\mathcal{N}^{\prime}(\zeta+i\tau)\chi_{\partial\gamma}(\zeta)d\zeta\\ &=\mathcal{D}_{B,\gamma}(k+i\tau)\sum_{j=0}^{J}B_{\gamma,j}k^{-j}+O(k^{-(J+1)}).\end{split}

This reduces to the computation of Bγ,jB_{\gamma,j} to a calculation of boundary integrals corresponding to powers of 𝒩\mathcal{N}. The boundary operator 𝒩\mathcal{N} does not have small operator norm and hence the infinite series isn’t actually convergent. However, the remainder can still be made arbitrarily small for sufficiently large MM, which is shown in [Zel04c]. This explains that while not an asymptotic expansion in the usual case, the Neuman series still provides an effective algorithm for computing the coefficients BjB_{j}. In [BB74], [BB71], [BB72], the physicists Balian and Bloch refer to the sum above as the multiple reflection expansion. The kernel 𝒩M(λ,s,s)\mathcal{N}^{M}(\lambda,s,s^{\prime}) is of the form

𝒩M(λ,x(s),x(s))\displaystyle\mathcal{N}^{M}(\lambda,x(s),x(s^{\prime})) =ΩM1𝒩(λ,s,s1)𝒩(λ,x(s1),x(s2))𝒩(λ,x(sM1),x(s))𝑑s1𝑑sM1\displaystyle=\int_{\partial\Omega^{M-1}}\mathcal{N}(\lambda,s,s_{1})\mathcal{N}(\lambda,x(s_{1}),x(s_{2}))\cdots\mathcal{N}(\lambda,x(s_{M-1}),x(s^{\prime}))ds_{1}\cdots ds_{M-1}
=ΩM1eik(x(s),x(s1),,x(sM1),s)a(λ,s,s1,,sM1,s)𝑑s1𝑑sM1,\displaystyle=\int_{\partial\Omega^{M-1}}e^{ik\mathcal{L}(x(s),x(s_{1}),\cdots,x(s_{M-1}),s^{\prime})}a(\lambda,s,s_{1},\cdots,s_{M-1},s^{\prime})ds_{1}\cdots ds_{M-1},

for the semiclassical amplitude

(22) a(λ,s,s1,,sM1,s)=λM/2e3πiM/4i=1Mcosϑi|x(si1)x(si)|12a1(λ|x(si1)x(si)|).\displaystyle a(\lambda,s,s_{1},\cdots,s_{M-1},s^{\prime})=\lambda^{M/2}e^{3\pi iM/4}\prod_{i=1}^{M}\frac{\cos\vartheta_{i}}{|x(s_{i-1})-x(s_{i})|^{\frac{1}{2}}}a_{1}(\lambda|x(s_{i-1})-x(s_{i})|).

The term multiple reflection refers to the phase function, which is a sum of consecutive generating functions for β\beta and hence generates the MM-fold iterate of the billiard map. Here, a1a_{1} is the semiclassical symbol in Proposition 4.5 and we use the convention that s=s0,s=sMs=s_{0},s^{\prime}=s_{M}. Integrating by parts the formula in Proposition 4.8 and reindexing, one obtains

12iM=1M0+1(1)MMTrρ(kζ)𝒩M(ζ+iτ)𝑑ζ=𝒟B,γ(k+iτ)j=0JBγ,jkj+O(k(J+1)).\displaystyle\begin{split}\frac{1}{2i}\sum_{M=1}^{M_{0}+1}\frac{(-1)^{M}}{M}\text{Tr}&\int_{\mathbb{R}}\rho^{\prime}(k-\zeta)\mathcal{N}^{M}(\zeta+i\tau)d\zeta\\ &=\mathcal{D}_{B,\gamma}(k+i\tau)\sum_{j=0}^{J}B_{\gamma,j}k^{-j}+O(k^{-(J+1)}).\end{split}

With the understanding that the following is not a true asymptotic expansion, but rather an algorithm for computing the wave invariants Bγ,jB_{\gamma,j} as above, we write

(23) 𝒟B,γ(k+iτ)j=0Bγ,jkj12iM=1(1)MMTrρ(kζ)𝒩M(ζ+iτ)𝑑ζ.\displaystyle\begin{split}\mathcal{D}_{B,\gamma}(k+i\tau)\sum_{j=0}^{\infty}B_{\gamma,j}k^{-j}\sim^{*}\frac{1}{2i}\sum_{M=1}^{\infty}\frac{(-1)^{M}}{M}\text{Tr}\int_{\mathbb{R}}\rho^{\prime}(k-\zeta)\mathcal{N}^{M}(\zeta+i\tau)d\zeta.\end{split}

The notation \sim^{*} is meant to emphasize the fact that the expansion is in the sense of Proposition 4.8. The operator 𝒩\mathcal{N} is both a microlocal pseudodifferential operater and a semiclassical Fourier integral operator. It has microlocally homogeneous singularities near the diagonal and away from the diagonal, it is a semiclassical Fourier integral operator quantizing the billiard map with semiclassical parameter =1/k\hbar=1/k. Hence, some care must be taken to apply the Ψ\PsiDO and FIO calculi. In [HZ04], Hassell and Zelditch introduced a microlocal decomposition 𝒩=𝒩0+𝒩1\mathcal{N}=\mathcal{N}_{0}+\mathcal{N}_{1},

𝒩0(x(s),x(s))\displaystyle\mathcal{N}_{0}(x(s),x(s^{\prime})) =χ(kρ|x(s)x(s))𝒩(λ,x(s),x(s)),\displaystyle=\chi(k^{\rho}|x(s)-x(s^{\prime}))\mathcal{N}(\lambda,x(s),x(s^{\prime})),
𝒩1(x(s),x(s))\displaystyle\mathcal{N}_{1}(x(s),x(s^{\prime})) =(1χ(kρ|x(s)x(s)|))𝒩(λ,x(s),x(s)),\displaystyle=(1-\chi(k^{\rho}|x(s)-x(s^{\prime})|))\mathcal{N}(\lambda,x(s),x(s^{\prime})),

where 𝒩0Ψ1(Ω)\mathcal{N}_{0}\in\Psi^{-1}(\partial\Omega) and 𝒩1I1/k0(Ω×Ω;Λβ)\mathcal{N}_{1}\in I_{1/k}^{0}(\partial\Omega\times\partial\Omega;\Lambda_{\beta}) is a Lagrangian distribution of order zero which is the Schwartz kernel of a semiclassical Fourier integral operator quantizing the billiard map. Here, Λβ\Lambda_{\beta} is the twisted graph of the billiard map:

Λβ={(s,s,σ,σ):(s,σ),(s,σ)B(Ω),β(s,σ)=(s,σ)},\displaystyle\Lambda_{\beta}=\{(s,s^{\prime},\sigma,-\sigma^{\prime}):(s,\sigma),(s^{\prime},\sigma^{\prime})\in B^{*}(\partial\Omega),\beta(s,\sigma)=(s^{\prime},\sigma^{\prime})\},

thought of as a Lagrangian submanifold of the product cotangent bundle. In this case, one has

𝒩M=(𝒩0+𝒩1)M=σ:M{0,1}𝒩σ(0)𝒩σ(1)𝒩σ(M).\displaystyle\mathcal{N}^{M}=(\mathcal{N}_{0}+\mathcal{N}_{1})^{M}=\sum_{\sigma:\mathbb{Z}_{M}\to\{0,1\}}\mathcal{N}_{\sigma(0)}\mathcal{N}_{\sigma{(1)}}\cdots\mathcal{N}_{\sigma(M)}.

We denote these compositions by 𝒩σ\mathcal{N}_{\sigma} and write |σ|=#{i:σ(i)=0}|\sigma|=\#\{i:\sigma(i)=0\}. The composition rules for Fourier integral operators tell us that the product 𝒩0𝒩1\mathcal{N}_{0}\mathcal{N}_{1} has canonical relation contained in ΔΩΛβ=Λβ\Delta_{\partial\Omega}\circ\Lambda_{\beta}=\Lambda_{\beta}, and the fact that 𝒩0Ψ1\mathcal{N}_{0}\in\Psi^{-1} reduces the order of the symbol by 11. Hence, 𝒩σ\mathcal{N}_{\sigma} has order |σ|-|\sigma| and quantizes the M|σ|thM-|\sigma|^{th} iterate of the billiard map. As we have microlocalized near an orbit γ\gamma with primitive period qq, only the terms with M|σ|=rqM-|\sigma|=rq for rr\in\mathbb{Z} potentially contribute to the leading order asymptotics near the lengths rLrL of a p/qp/q periodic orbit iterated rr times. This is shown rigorously in [Zel09], where it is also demonstrated that for any specified order RR and sufficiently large corresponding MM, the terms 𝒩0M\mathcal{N}_{0}^{M} do not contribute to the regularized resolvent trace asymptotics modulo O(kR)O(k^{-R}). In addition, it is shown there that if L=LγL=L_{\gamma} is simple, then the trace is unchanged modulo O(k)O(k^{-\infty}) when the semiclassical cutoff in 4.8 is removed. The presence of a microlocal cutoff is what is meant in Theorem 1.1 regarding the contributions to the resolvent trace of a specific orbit γ\gamma. The reason for a small remainder in Proposition 4.8 is that after microlocalizing near the projection of γ\gamma to B(Ω)B^{*}(\partial\Omega) and convolving with ρ\rho, ρ^\widehat{\rho} being supported near LγL_{\gamma}, 𝒩σ\mathcal{N}_{\sigma} can have at most qq factors of 𝒩1\mathcal{N}_{1} without making the phase nonstationary. For large enough MM, at least MqRM-q\gtrsim R factors of 𝒩0\mathcal{N}_{0} must enter the composition, each of which reduces the order by 11.

The only possible terms contributing to Bγ,jB_{\gamma,j} have qMq+jq\leq M\leq q+j; if MM were larger, the terms would be at most of order kj1k^{-j-1} and if MM were smaller than qq, the phase would be nonstationary on the support of ρ^\widehat{\rho}, in which case the asymptotics would be O(k)O(k^{-\infty}). Moreover, we will see later that only the compositions with |σ|=0|\sigma|=0 generate useful terms containing

  1. (1)

    highest powers of a yet to be specified deformation parameter ε1\varepsilon^{-1} and

  2. (2)

    highest order derivatives of the boundary curvature upon performing a stationary phase expansion.

4.5. Principal symbol calculation

To compute 23, we use the Fourier inversion formula to see that

(24) Trρ(kζ)𝒩σ(ζ+iτ)𝑑ζ=12πΩSMei(kζ)titρ^(t)×i=1Mei(ζ+iτ)|x(si)x(si+1)|+3πi/4χi(ζϱ|x(si)x(si+1)|)×(ζ+iτ)|x(si)x(si+1)|a1((ζ+iτ)|x(si)x(si+1)|)cosϑidtdζdS,\displaystyle\begin{split}&\text{Tr}\,\int\rho^{\prime}(k-\zeta)\mathcal{N}_{\sigma}(\zeta+i\tau)d\zeta=\frac{1}{2\pi}\int_{{\partial\Omega}_{S}^{M}}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}e^{i(k-\zeta)t}it\hat{\rho}(t)\\ &\times\prod_{i=1}^{M}e^{i(\zeta+i\tau)|x(s_{i})-x(s_{i+1})|+3\pi i/4}\chi_{i}(\zeta^{\varrho}|x(s_{i})-x(s_{i+1})|)\\ &\times\sqrt{\frac{(\zeta+i\tau)}{|x(s_{i})-x(s_{i+1})|}}a_{1}((\zeta+i\tau)|x(s_{i})-x(s_{i+1})|)\cos\vartheta_{i}dtd\zeta dS,\end{split}

where S=(s1,,sM)S=(s_{1},\cdots,s_{M}), sM+1=s1s_{M+1}=s_{1} and

χi(z)={χ(z)σ(i)=01χ(z)σ(i)=1,\displaystyle\chi_{i}(z)=\begin{cases}\chi(z)&\sigma(i)=0\\ 1-\chi(z)&\sigma(i)=1,\end{cases}

for a cutoff χ\chi equal to 11 on [1/2,+1/2][-1/2,+1/2] and zero outside of [1,+1][-1,+1] (cf. Proposition 4.5). Changing variables ζζ/k\zeta\mapsto\zeta/k, formula 24 can be written as the oscillatory integral

(25) kΩSMeikΦ(ζ+iτ,S,t)Bσ(kζ+iτ,t,S)𝑑ζ𝑑t𝑑S,\displaystyle\begin{split}k\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}\int_{{\partial\Omega}_{S}^{M}}e^{ik\Phi({\zeta}+i{\tau},S,{t})}B_{\sigma}(k{\zeta}+i{\tau},t,S)d{\zeta}d{t}dS,\end{split}

where the phase is given by

(26) Φ(ζ+iτ,S,t)=(1ζ)t+ζ(S)\displaystyle\Phi({\zeta}+i{\tau},S,{t})=(1-\zeta)t+{\zeta}\mathcal{L}(S)

and the amplitude by

(27) Bσ(λ,S)=e3Mπi/42πitρ^(t)λM2eτ(S)i=1Mχ~i(λ,S)×cosϑi|x(si)x(si+1)|1/2a1(λ|x(si)x(si+1)|),\displaystyle\begin{split}B_{\sigma}(\lambda,S)=&\frac{e^{3M\pi i/4}}{2\pi}it\hat{\rho}({t})\lambda^{\frac{M}{2}}e^{-\tau\mathcal{L}(S)}\prod_{i=1}^{M}\widetilde{\chi}_{i}(\lambda,S)\\ &\times\frac{\cos\vartheta_{i}}{|x(s_{i})-x(s_{i+1})|^{1/2}}a_{1}(\lambda|x(s_{i})-x(s_{i+1})|),\end{split}

where we use the notation

χ~i(λ,S)={χ(Re(λ)ϱ|x(si)x(si+1)|)σ(i)=01χ(Re(λ)ϱ|x(si)x(si+1)|)σ(i)=1.\displaystyle\widetilde{\chi}_{i}(\lambda,S)=\begin{cases}\chi(\operatorname{Re}(\lambda)^{\varrho}|x(s_{i})-x(s_{i+1})|)&\sigma(i)=0\\ 1-\chi(\operatorname{Re}(\lambda)^{\varrho}|x(s_{i})-x(s_{i+1})|)&\sigma(i)=1.\end{cases}

Recall that ϑi\vartheta_{i} is the angle between the interior normal to the boundary and the link (x(si+1)x(si))(x(s_{i+1})-x(s_{i})). When we later set |σ|=0|\sigma|=0 and M=qM=q to obtain leading order asymptotics, all of the χi\chi_{i} will be one.

Remark 4.9.

Note as above that a priori, BσSϱM2(ΩM)B_{\sigma}\in S_{\varrho}^{\frac{M}{2}}(\partial\Omega^{M}). The condition ϱ(0,12)\varrho\in(0,\frac{1}{2}) guarantees that one can integrate out the diagonal singularities corresponding to terms with σ(i)=0\sigma(i)=0 to obtain a Fourier integral kernel of lower order quantizing βM|σ|\beta^{M-|\sigma|}, as was done in [Zel09]. We will make use of this in Section 6.

An application of stationary phase in the variables ζ,t{\zeta},{t} of 25 gives the following:

Proposition 4.10 ([Zel09]).

For each σ:M{0,1}\sigma:\mathbb{Z}_{M}\to\{0,1\}, Trρ𝒩σ(k+iτ)\text{Tr}\rho^{\prime}\ast\mathcal{N}_{\sigma}(k+i\tau) is given modulo kk^{-\infty} by

(28) e3Mπi/4iΩSMei(k+iτ)(S)a0σ(k+iτ,S)𝑑S\displaystyle e^{3M\pi i/4}i\int_{{\partial\Omega}_{S}^{M}}e^{i(k+i\tau)\mathcal{L}(S)}a_{0}^{\sigma}(k+i\tau,S)dS

where

a0σ(k+iτ,S)=ρ^((S))((S)Aσ(k+iτ,S)1ikAσ(k+iτ,S))SϱM2(ΩM)\displaystyle a_{0}^{\sigma}(k+i\tau,S)=\hat{\rho}(\mathcal{L}(S))\left(\mathcal{L}(S)A_{\sigma}(k+i\tau,S)-\frac{1}{i}\frac{\partial}{\partial k}A_{\sigma}(k+i\tau,S)\right)\in S_{\varrho}^{\frac{M}{2}}(\partial\Omega^{M})

and Aσ(k+iτ,S)A_{\sigma}(k+i\tau,S) is

(k+iτ)M/2i=1Mχi(k+iτ,S)cosϑi|x(si)x(si+1)|1/2a1((k+iτ)|x(si)x(si+1)|).\displaystyle(k+i\tau)^{M/2}\prod_{i=1}^{M}\chi_{i}(k+i\tau,S)\frac{\cos\vartheta_{i}}{|x(s_{i})-x(s_{i+1})|^{1/2}}a_{1}((k+i\tau)|x(s_{i})-x(s_{i+1})|).

The complex phase is given by (3M+2)π/4(3M+2)\pi/4 and a1a_{1} is the Hankel amplitude from Proposition 4.5.

Proof.

See Equation 32 in Section 6 below for the stationary phase formula. In our case, the stationary points occur at t=,ζ=1t=\mathcal{L},{\zeta}=1. The Hessian is given by

ζ,t2Φ=(0110),\displaystyle\partial_{{\zeta},{t}}^{2}\Phi=\begin{pmatrix}0&-1\\ -1&0\end{pmatrix},

from which we obtain ζ,t2Φ1ζ,t,ζ,t=2ζt\langle\partial_{{\zeta},{t}}^{2}\Phi^{-1}\partial_{{\zeta},{t}},\partial_{{\zeta},{t}}\rangle=-2\partial_{{\zeta}}\partial_{{t}}. The signature of the Hessian is 0 and the modulus of its determinant is one. The total prefactor is then 2πkeik\frac{2\pi}{k}e^{ik\mathcal{L}}, the coefficient of which cancels with that from Fourier inversion and the change of variables above. Since the phase is quadratic, only the amplitude is differentiated (g=0g=0, μ=0\mu=0 in the notation of formula 32 below). Since the amplitude factors into a locally linear function of tt and one of ζ\zeta, there are only two nonzero terms. Near the critical point corresponding to the length LγL_{\gamma} of a periodic orbit, ρ^\hat{\rho} is constant and hence t\partial_{t} only lands on tt. The kk coming from ζ\partial_{{\zeta}} is canceled by the k1k^{-1} in stationary phase as is the factor of 2π2\pi. The factor eτ(S)e^{-\tau\mathcal{L}(S)} is unchanged by the Hessian operator and can be reinserted into the phase function. ∎

In the next section, we will prove that the |σ|1|\sigma|\geq 1 terms do not contribute to the leading order deformation asymptotics or highest order derivatives in the curvature jet of the boundary. Hence, we can set χi=1\chi_{i}=1 in the expression for AσA_{\sigma} in 4.10. As 𝒩0Ψ1(Ω)\mathcal{N}_{0}\in\Psi^{-1}(\partial\Omega), the symbol a0σa_{0}^{\sigma} in Proposition 4.10 can be integrated to yield a symbol in SϱM3|σ|2(ΩM|σ|)S_{\varrho}^{\frac{M-3|\sigma|}{2}}(\partial\Omega^{M-|\sigma|}); see Theorem 3.8 in [Zel09]. Moreover, since a1SϱM2(ΩM)a_{1}\in S_{\varrho}^{\frac{M}{2}}(\partial\Omega^{M}), AσSϱM21A_{\sigma}^{\prime}\in S_{\varrho}^{\frac{M}{2}-1}. Hence, for |σ|=0|\sigma|=0, M=qM=q and SS near SγS_{\gamma}, we have

(29) a00(λ,S)=(λ2π)q/2(S)i=1qcosϑi|x(si)x(si+1)|1/2+OSϱ0(Re(λ)q21),\displaystyle a_{0}^{0}(\lambda,S)=\left(\frac{\lambda}{2\pi}\right)^{q/2}\mathcal{L}(S)\prod_{i=1}^{q}\frac{\cos\vartheta_{i}}{|x(s_{i})-x(s_{i+1})|^{1/2}}+O_{S_{\varrho}^{0}}\left(\operatorname{Re}(\lambda)^{\frac{q}{2}-1}\right),

where we used the leading order asymptotics for a1a_{1} in formula 4.5.

Remark 4.11.

Recall that by formula 6, the product in the leading order expansion of a0a_{0} is precisely

i=1qcosϑi|x(si)x(si+1)|1/2=det2|det(1Pγ)|,\displaystyle\prod_{i=1}^{q}\frac{\cos\vartheta_{i}}{|x(s_{i})-x(s_{i+1})|^{1/2}}=\sqrt{\frac{\det\partial^{2}\mathcal{L}}{|\det(1-P_{\gamma})|}},

with PγP_{\gamma} being the Poincaré map associated to γ\gamma.

Combining Proposition 4.10 and 23, we obtain the following:

Corollary 4.12.

The coefficients in the regularized resolvent trace expansion can be determined from the oscillatory integrals in Proposition 4.10:

𝒟γ(λ)j=0Bγ,jRe(λ)j12M=1σ:M{0,1}eMπi/4MΩSMeiλ(S)a0σ(λ,S)dS.\displaystyle\mathcal{D}_{\gamma}(\lambda)\sum_{j=0}^{\infty}B_{\gamma,j}\operatorname{Re}(\lambda)^{-j}\sim^{*}\frac{1}{2}\sum_{M=1}^{\infty}\sum_{\sigma:\mathbb{Z}_{M}\to\{0,1\}}\frac{e^{-M\pi i/4}}{M}\int_{{\partial\Omega}_{S}^{M}}e^{i\lambda\mathcal{L}(S)}a_{0}^{\sigma}(\lambda,S)dS.

5. Perturbation Thoery

Starting with a domain Ω0\Omega_{0} and a degenerate periodic orbit γ\gamma, we introduce a one parameter family of domains Ωε\Omega_{\varepsilon} specified by a smooth function μ\mu:

Ωε={x+μ(x,ε)νx:xΩ0},\displaystyle\Omega_{\varepsilon}=\{x+\mu(x,\varepsilon)\nu_{x}:x\in\partial\Omega_{0}\},

where νx\nu_{x} is the outward pointing normal to Ω0\Omega_{0} at xx. We impose the following constraints on Ωε\Omega_{\varepsilon}:

  1. (1)

    Ω0=Ω\Omega_{0}=\Omega.

  2. (2)

    Ωε\Omega_{\varepsilon} and Ω0\Omega_{0} make first order contact at the reflection points xiγx_{i}\in\partial\gamma.

  3. (3)

    Near each xix_{i}, Ωε\Omega_{\varepsilon} is given by x+μ(x,ε)ν(x)x+\mu(x,\varepsilon)\nu(x), where xΩx\in\partial\Omega and ν(x)\nu(x) is the unit normal at xx.

  4. (4)

    μ\mu is CC^{\infty} in all parameters.

  5. (5)

    det2cγε\det\partial^{2}\mathcal{L}\sim c_{\gamma}\varepsilon for cγ0c_{\gamma}\neq 0.

  6. (6)

    2\partial^{2}\mathcal{L} has rank q1q-1 on Ω0\Omega_{0}.

  7. (7)

    The third and higher order derivatives of \mathcal{L} are O(1)O(1) near γ\partial\gamma.

In particular, the perturbation preserves periodicity of γ\gamma and removes its degeneracy. It will be important for us that the Hessian of \mathcal{L} on Ω0\partial\Omega_{0} be minimally degenerate in the sense that it has rank q1q-1. Theorem 1.1 applies to any family of domains which satisfy the conditions above. To demonstrate the robustness of our main theorem, we now give a large class of examples to which it applies.

Definition 5.1.

Given a domain Ω\Omega and a (p,q)(p,q)-periodic orbit γ\gamma, we say that γ\gamma satisfies the injectivity condition if its relfection points are all distinct.

Proposition 5.2.

Let 𝒞1\mathcal{C}_{1} be the class of smooth, bounded strictly convex planar domains Ω\Omega which have a degenerate periodic orbit γ\gamma satisfying the injectivity condition. Within 𝒞1\mathcal{C}_{1}, there is an open and dense set in the CC^{\infty} topology on which the Hessian of \mathcal{L} has rank q1q-1.

Proposition 5.3.

Given a domain Ω0\Omega_{0} and a degenerate orbit γ\gamma having rank(2(Sγ))=q1\text{rank}(\partial^{2}\mathcal{L}(S_{\gamma}))=q-1, consider the class 𝒞2\mathcal{C}_{2} of deformations μ:Ω×[0,ε0)\mu:\partial\Omega\times[0,\varepsilon_{0})\to\mathbb{R} which preserve γ\gamma and the angles of reflection. Within 𝒞2\mathcal{C}_{2}, those which make 2\partial^{2}\mathcal{L} nondegenerate at SγS_{\gamma} in arclength coordinates form an open set. If in addition, the orbit satisfies the injectivity condition, then they are also dense in 𝒞2\mathcal{C}_{2}.

The preceding two propositions follow from the multijet transversality theorem, which is a generalization due to Mather of Thom’s transversality theorem. It is easy to check that for ellipses, the Hessian has rank q1q-1. However, degenerate homoclinic orbits such as those constructed in [Cal22] could have lower rank Hessians. In [KZ18], it is shown that domains with a rational caustic of rotation number 1/q1/q are quantitatively dense in the space of all convex domains. For analytic domains, the density is exponential in qq, whereas for domains of finite smoothness, the density is polynomial. For such domains, it is easy to see that the length functional evaluated at reflection points of an orbit tangent to a corresponding rational caustic has Hessian of rank q1q-1. Together with the cancellations in our forthcoming paper, we see that the any finite order singular support of the wave trace is generically distinct from the length spectrum in the sense above.

5.1. Nonsingular perturbations

We first show that the jet of the length functional is algebraically equivalent to that of the boundary curvature at reflection points and in particular, demonstrate that arbitrarily small perturbations of curvature can result in nondegeneracy. In what follows, write detε1=Mε\det\mathcal{L}_{\varepsilon}^{-1}=M_{\varepsilon}.

Proposition 5.4.

If the rank of 2\partial^{2}\mathcal{L} is q1q-1, then the condition det2ε0\det\partial^{2}\mathcal{L}_{\varepsilon}\neq 0 imposes a codimension 11 constraint on the qq-fold 0-jet of the boundary curvature at reflection points. For each point xiγx_{i}\in\partial\gamma and each 1iq1\leq i\leq q, the data ϑi\vartheta_{i}, |xixi+1||x_{i}-x_{i+1}| and i1,,imm\partial_{i_{1},\cdots,i_{m}}^{m}\mathcal{L}, with |ijik|1|i_{j}-i_{k}|\leq 1, uniquely determine the qq-fold m2m-2-jet of the curvature at xiγx_{i}\in\partial\gamma.

Proof.

It is shown in [KT91] that

(30) δi:=22si=2cosϑi(cosϑiiκi),αi:=2sisi+1=cosϑicosϑi+1i,\displaystyle\begin{split}\delta_{i}&:=\frac{\partial^{2}\mathcal{L}}{\partial^{2}s_{i}}=2\cos\vartheta_{i}\left(\frac{\cos\vartheta_{i}}{\ell_{i}}-\kappa_{i}\right),\\ \alpha_{i}&:=\frac{\partial^{2}\mathcal{L}}{\partial s_{i}\partial s_{i+1}}=\frac{\cos\vartheta_{i}\cos\vartheta_{i+1}}{\ell_{i}},\end{split}

where i=|xi(S)xi+1(S)|\ell_{i}=|x_{i}(S)-x_{i+1}(S)|, ϑi\vartheta_{i} is the angle of incidence with the inward pointing normal at xi(S)x_{i}(S), and κi\kappa_{i} is the curvature of Ω\partial\Omega at xi(S)x_{i}(S). It follows that 2\partial^{2}\mathcal{L} is tridiagonal and cyclic with δi\delta_{i}s on the diagonal and αi\alpha_{i}s on the off diagonals. The Hessian matrix is of the form

(31) 2=(δ1α100αqα1δ2α2000α2δ3α3000α3δ40αq000αq1δq)\displaystyle\partial^{2}\mathcal{L}=\begin{pmatrix}\delta_{1}&\alpha_{1}&0&0&\cdots&\alpha_{q}\\ \alpha_{1}&\delta_{2}&\alpha_{2}&0&\cdots&0\\ 0&\alpha_{2}&\delta_{3}&\alpha_{3}&\cdots&0\\ 0&0&\alpha_{3}&\delta_{4}&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots&\vdots\\ \alpha_{q}&0&0&0&\alpha_{q-1}&\delta_{q}\end{pmatrix}

and the equivalence of curvature and length functional jets follows immediately. Note that the angles of incidence and lengths of links are preserved under deformations which make first order contact with Ω0\partial\Omega_{0} at γ\partial\gamma. However, the curvature κi\kappa_{i} in the formula for δi\delta_{i} can change. If the perturbed curvature has an asymptotic expansion of the form

κi(ε)j=0μjiεj\displaystyle\kappa_{i}(\varepsilon)\sim\sum_{j=0}^{\infty}\mu_{j}^{i}\varepsilon^{j}

for some smooth functions μji(x)\mu_{j}^{i}(x), then we have

det2ε\displaystyle\det\partial^{2}\mathcal{L}_{\varepsilon} =0+εTr(adj(20)ε2ε|ε=0)+O(ε2)\displaystyle=0+\varepsilon\text{Tr}\left(\text{adj}\left(\partial^{2}\mathcal{L}_{0}\right)\partial_{\varepsilon}\partial^{2}\mathcal{L}_{\varepsilon}\big{|}_{\varepsilon=0}\right)+O(\varepsilon^{2})
=Tr(Mμ)ε+O(ε2),\displaystyle=-\text{Tr}(M\mu)\varepsilon+O(\varepsilon^{2}),

where μ\mu is the diagonal matrix given by μ1iδij\mu_{1}^{i}\delta_{ij} and MM is the adjugate of 20\partial^{2}\mathcal{L}_{0} consisting of principal minors. For any matrix, its rank is given by the maximal size for which there is a nonzero minor. If 2\partial^{2}\mathcal{L} has rank q1q-1, then the rank of MM is 11. Hence, the condition TrMμ0\text{Tr}M\mu\neq 0 imposes a codimension one, linear constraint on the perturbation matrix μ\mu or equivalently, on the curvature at reflection points. ∎

From the formulas 30 and 31 above, one sees that the rank of 2\partial^{2}\mathcal{L} is always at least q2q-2 due to the off diagonal terms being nonzero for nonglancing orbits. If it is in fact equal to q2q-2, then the linearized Poincaré map is equal to the identity. If the rank of 2\partial^{2}\mathcal{L} is q1q-1, the linearized Poincaré map has 11 as an eigenvalue but differs from the identity and if the rank is qq, then the orbit is nondegenerate.

5.2. Analysis of the Hessian

Proposition 5.5.

If 2\partial^{2}\mathcal{L} has rank q1q-1 in arclength coordinates at a degenerate periodic trajectory γ\gamma in Ω0\Omega_{0}, then its adjugate matrix has rank 11 and its elements are of the form hi1,i2(S,0)=±hi1(S,0)hi2(S,0)h_{i_{1},i_{2}}(S,0)=\pm h_{i_{1}}(S,0)h_{i_{2}}(S,0) for some smooth functions hi(S,0)h_{i}(S,0). If a deformation Ωε\Omega_{\varepsilon} which preserves the points and angles of reflection makes |det2|cγε|\det\partial^{2}\mathcal{L}|\sim c_{\gamma}\varepsilon, cγ>0c_{\gamma}>0, then the (i1,i2)(i_{1},i_{2}) entry of 21\partial^{2}\mathcal{L}^{-1} is h~i1i2=σh~i1h~i2+O(1)\widetilde{h}_{i_{1}i_{2}}=\sigma\cdot\widetilde{h}_{i_{1}}\widetilde{h}_{i_{2}}+O(1) with each h~i\widetilde{h}_{i} having the form (cγε)1/2hi(c_{\gamma}\varepsilon)^{-1/2}h_{i} for some smooth functions hiC(/×[0,ε0))h_{i}\in C^{\infty}(\mathbb{R}/\ell\mathbb{Z}\times[0,\varepsilon_{0})) which are uniformly bounded in ε\varepsilon. Here, σ\sigma is +1+1 if closest eigenvalue to zero of 2\partial^{2}\mathcal{L} is positive and 1-1 otherwise. If Ω0\Omega_{0} is a circle, then hi1=hi2h_{i_{1}}=h_{i_{2}} for all 1i1,i2q1\leq i_{1},i_{2}\leq q.

Proof.

From Proposition 5.4 above, we can choose a perturbation μ\mu such that det2cγε\det\partial^{2}\mathcal{L}\sim c_{\gamma}\varepsilon, cγ0c_{\gamma}\neq 0. Denote by H=Hε=2H=H_{\varepsilon}=\partial^{2}\mathcal{L} the Hessian of \mathcal{L} on Ωε\partial\Omega_{\varepsilon}. The adjugate matrix MεM_{\varepsilon} of 2\partial^{2}\mathcal{L} in Ωε\Omega_{\varepsilon} consists of complimentary minors depending on κC0(Ωε)\|\kappa\|_{C^{0}(\partial\Omega_{\varepsilon})}. If the perturbation μ(ε,)\mu(\varepsilon,\cdot) makes detεcγε0\det\mathcal{L}_{\varepsilon}\sim c_{\gamma}\varepsilon\neq 0, then the inverse Hessians satisfy

211cγεMε,\displaystyle\partial^{2}\mathcal{L}^{-1}\sim\frac{1}{c_{\gamma}\varepsilon}M_{\varepsilon},

where the cofactor matrices are C(q2×[0,ε0])C^{\infty}(\mathbb{R}^{q^{2}}\times[0,\varepsilon_{0}]) and uniformly bounded in ε\varepsilon all the way down to ε=0\varepsilon=0. In particular, 21(cγε)1M0+O(1)\partial^{2}\mathcal{L}^{-1}\sim(c_{\gamma}\varepsilon)^{-1}M_{0}+O(1). Under the rank q1q-1 assumption, M0M_{0} is rank 11, which together with symmetry about the diagonal implies that 21σ(cγε)1hi1hi2\partial^{2}\mathcal{L}^{-1}\sim\sigma\cdot(c_{\gamma}\varepsilon)^{-1}{h}_{i_{1}}{h}_{i_{2}} as in the statement of the theorem.

If Ω0\Omega_{0} is a circle, then we have action angle variables in which the angles are just the usual arclength coordinates on Ω0\partial\Omega_{0} and the actions are the radii of concentric circles to which corresponding orbits are tangent. It is easy to see in this case that 20\partial^{2}\mathcal{L}_{0} is negative semi-definite with rank q1q-1 and has kernel v\mathbb{R}v, for v=(1,,1)v=(1,\cdots,1) corresponding to rotation along a concentric circle. We claim that each row of M0M_{0} is a multiple of vv. Let MiM^{i} denote the iith row of M0M_{0} and choose any ww orthogonal to vv. Then, then using the definition of MM in terms of complimentary minors, we have

M1,w=det(w1w2wqH021H022H02qH0q1H0q2H0qq),\displaystyle\langle M^{1},w\rangle=\det\begin{pmatrix}w_{1}&w_{2}&\cdots&w_{q}\\ H_{0}^{21}&H_{0}^{22}&\cdots&H_{0}^{2q}\\ \vdots&\ldots&\ddots&\vdots\\ H_{0}^{q1}&H_{0}^{q2}&\ldots&H_{0}^{qq}\end{pmatrix},

where H0ijH_{0}^{ij} are the entries of 20\partial^{2}\mathcal{L}_{0}. Since wkerH0=kerH0w\perp\ker H_{0}=\ker H_{0}^{*}, wSpan{H01,H02,,H0q}=Span{H02,,H0q}w\in\text{Span}\{H_{0}^{1},H_{0}^{2},\cdots,H_{0}^{q}\}=\text{Span}\{H_{0}^{2},\cdots,H_{0}^{q}\}, given that the rows sum to zero and are hence linearly dependent. Therefore, M1,w=0\langle M^{1},w\rangle=0. We have shown that Span(v)Span(M1)\text{Span}(v)^{\perp}\subset\text{Span}(M^{1})^{\perp} which implies Span(M1)=Span(v)\text{Span}(M^{1})=\text{Span}(v) since they have equal dimensions; the same holds if one replaces 11 by any 1iq1\leq i\leq q. Each row of MM is a multiple of vv and by symmetry about the diagonal, the multiples must be the same. In particular,

21σcγεh(1,,1)T(1,,1)+O(1),\displaystyle\partial^{2}\mathcal{L}^{-1}\sim\frac{\sigma}{c_{\gamma}\varepsilon}h(1,\cdots,1)^{T}(1,\cdots,1)+O(1),

for a single smooth function hh on Ω0\Omega_{0}. Since 2\partial^{2}\mathcal{L} has q1q-1 negative eigenvalues on the disk, σ\sigma only depends on ellipticity or hyperbolicity of the corresponding perturbed orbit. ∎

6. Stationary Phase and Feynman Diagrams

We now apply the method of stationary phase to the integral appearing in Theorem 4.10, but on the perturbed domain Ωε\Omega_{\varepsilon} from Section 5, so as to make the phase nondegenerate. In general, if Φ\Phi has a unique stationary point at x0x_{0} with nondegenerate Hessian and uu is a compactly supported smooth function, the stationary phase formula in [Hör03] reads:

(32) nu(x)eikΦ(x)𝑑x(2π/k)n/2eikΦ(x0)eiπsgn2Φ(x0)/4|det2Φ(x0)|1/2j=0kjLju(x0),\displaystyle\int_{\mathbb{R}^{n}}u(x)e^{ik\Phi(x)}dx\sim(2\pi/k)^{n/2}\frac{e^{ik\Phi(x_{0})}e^{i\pi\text{sgn}\partial^{2}\Phi(x_{0})/4}}{|\det\partial^{2}\Phi(x_{0})|^{1/2}}\sum_{j=0}^{\infty}k^{-j}L_{j}u(x_{0}),

where the LjL_{j} are differential operators of order 2j2j having the form

(33) Lju(x0)=νμ=j2ν3μij2ν2Φ(x0)1,ν(gμu(x0))/μ!ν!.\displaystyle L_{j}u(x_{0})=\sum_{\nu-\mu=j}\sum_{2\nu\geq 3\mu}i^{-j}2^{-\nu}\langle\partial^{2}\Phi(x_{0})^{-1}\partial,\partial\rangle^{\nu}(g^{\mu}u(x_{0}))/\mu!\nu!.

Here, g=Φ(x)Φ(x0)Φ(x0)(xx0)Φ′′(x0)(xx0)2g=\Phi(x)-\Phi(x_{0})-\Phi^{\prime}(x_{0})(x-x_{0})-\Phi^{\prime\prime}(x_{0})(x-x_{0})^{2} is the higher order part of the Taylor expansion of Φ\Phi, having polynomial order 3\geq 3. We in fact need a version of stationary phase which allows for dependence on auxilary parameters. If the phase Φ\Phi and amplitude uu depend smoothly on a parameter τT\tau\in T with uu being supported in a compact set which is independent of τ\tau, then expansion 32 is uniform in τ\tau (see [Dui96]). There are several asymptotics to keep track of simultaneously: powers of kk and ε\varepsilon together with the number of boundary curvature derivatives.

We use Feynman diagrams to organize the terms hierarchically, inspired by but using a different structure than those in [Zel09].

6.1. Feynman Calculus

We review the procedure first used by Feynman to evaluate asymptotic integrals. We will follow the conventions on notation as in [Axe97].

Definition 6.1.

A graph 𝒢\mathcal{G} with VV vertices and II edges is called a Feynman diagram of order jj if

  1. (1)

    IV=jI-V=j,

  2. (2)

    there are VV closed (indistinguishable) vertices, each of which has valency at least 33,

  3. (3)

    there is one open, marked vertex with any possible valency, including 0 (corresponding to an isolated vertex), and

  4. (4)

    each edge is labeled by a function \ell, which assigns two indices (i1,i2)(i_{1},i_{2}) corresponding two its two endpoints.

Remark 6.2.

The quantity IV=jI-V=j is closely related to the Euler characteristic χ(𝒢)\chi(\mathcal{G}), but does not include faces; not all diagrams are assumed to be planar graphs. We will sometimes denote 𝒢\mathcal{G} by 𝒢\mathcal{G}_{\ell} to emphasize \ell, the edge labeling.

Definition 6.3.

Let 𝒢\mathcal{G} be a Feynman diagram of order jj with labeling \ell.

  • For the open vertex, we associate the quantity

    vuxi1xiv(x0),\frac{\partial^{v}u}{\partial x_{i_{1}}\cdots\partial x_{i_{v}}}(x_{0}),

    where i1,,ivi_{1},\cdots,i_{v} are the indices labeling the incident edges (vv is the valency of the vertex).

  • To each closed vertex corresponds a factor of

    ikvΦxi1xiv(x0),ik\frac{\partial^{v}\Phi}{\partial x_{i_{1}}\cdots\partial x_{i_{v}}}(x_{0}),

    where again i1,,ivi_{1},\cdots,i_{v} are the indices of the incident edges.

  • For each edge with indices i1,i2i_{1},i_{2}, we assign the inverse Hessian element

    ik(2Φ)i1i21:=ikh~i1i2.\frac{i}{k}(\partial^{2}\Phi)_{i_{1}i_{2}}^{-1}:=\frac{i}{k}\widetilde{h}_{i_{1}i_{2}}.

The Feynman amplitude F𝒢,F_{\mathcal{G},\ell} corresponding to 𝒢\mathcal{G} and label \ell is given by the product of each factor above.

Proposition 6.4 ([Axe97]).

Let ZkZ_{k} denote the integral 32. As kk\to\infty,

(34) Zk(2πk)n/2eikΦ(x0)eiπsgn2Φ|det2Φ|1/2I=0V=0𝒢,F𝒢w(𝒢),\displaystyle Z_{k}\sim\left(\frac{2\pi}{k}\right)^{n/2}\frac{e^{ik\Phi(x_{0})}e^{i\pi\text{sgn}\partial^{2}\Phi}}{|\det\partial^{2}\Phi|^{1/2}}\sum_{I=0}^{\infty}\sum_{V=0}^{\infty}\sum_{\mathcal{G},\ell}\frac{F_{\mathcal{G}}}{w(\mathcal{G})},

where the inner sum is over all labeled labeled Feynman diagrams with VV internal vertices, 11 external vertex, and II edges. ω(𝒢)\omega(\mathcal{G}) is the order of the automorphism group of 𝒢\mathcal{G}, F𝒢,F_{\mathcal{G},\ell} is the Feynman amplitude and \ell assigns to each endpoint of each edge an index in {1,,n}\{1,\cdots,n\}.

We refer to [Axe97] for an elementary proof of the equivalence of 32 and Proposition 6.4. It remains evaluate each of these Feynman amplitudes in coordinates to determine their contributions to Bγ,jB_{\gamma,j}.

6.2. Contributing Graphs

Here, we examine the Feynman diagrams which contribute maximal inverse powers of ε\varepsilon to the trace expansion in Proposition 4.10. It is clear that the amplitudes a0σ(k+iτ,S;ε)a_{0}^{\sigma}(k+i\tau,S;\varepsilon) are uniformly bounded in ε\varepsilon. The main contributions come from the closed vertices (derivatives of the phase) and their incident edges (inverse Hessians). To keep track of leading order terms in ε\varepsilon, we must isolate those Feynman diagrams which produce the most inverse Hessians, each of which produces a power of ε1\varepsilon^{-1}.

Lemma 6.5.

For each jj, the Feynman amplitudes having maximal products of inverse Hessian elements correspond to the class of 33-regular graphs on 2j2j closed vertices with one additional open vertex having valency zero.

Proof.

In the usual non-diagrammatic stationary phase notation, maximal inverse Hessians arise from the terms where νμ=j\nu-\mu=j and 2ν=3μ2\nu=3\mu, which yields ν=2j\nu=2j and μ=3j\mu=3j. In Proposition 6.4, VI=jV-I=j and 𝒢\mathcal{G} has I=3jI=3j edges. For a general graph,

I=u02+12i=1Vvi,I=\frac{u_{0}}{2}+\frac{1}{2}\sum_{i=1}^{V}v_{i},

where u0u_{0} is the valency of the open vertex and viv_{i} are the valencies of unmarked vertices. As each vi3v_{i}\geq 3 by assumption, we have

I32V\displaystyle I-\frac{3}{2}V\geq 0,\displaystyle 0,
32I+32V=\displaystyle-\frac{3}{2}I+\frac{3}{2}V= 32j,\displaystyle-\frac{3}{2}j,

from which it follows that I3jI\leq 3j, which is the maximal number of inverse Hessians. This is acheived when V=2jV=2j and I=3jI=3j, corresponding to each vi=3v_{i}=3. If the open vertex were not isolated, consider the graph 𝒢̊=𝒢\{open vertex and all its incident edges}\mathring{\mathcal{G}}=\mathcal{G}\backslash\{\text{open vertex and all its incident edges}\}. The formula above would then force there to be vertices with valency strictly less than 33. ∎

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1. Some admissible Feynman diagrams given by 33 regular graphs on 2j2j vertices.
Corollary 6.6.

Modulo an error bounded uniformly in ε\varepsilon, the Feynman amplitudes for contributing graphs with maximal Hessians consist of a0(Sγ)(±i)jkja_{0}(S_{\gamma})(\pm i)^{j}k^{-j} and products of terms of the form

(cγε)32hi1hi2hi3i1,i2,i3i1i2i33,\displaystyle(c_{\gamma}\varepsilon)^{-\frac{3}{2}}h_{i_{1}}h_{i_{2}}h_{i_{3}}\sum_{i_{1},i_{2},i_{3}}\partial_{i_{1}i_{2}i_{3}}^{3}\mathcal{L},

for some indices 1i1,i2,i3q1\leq i_{1},i_{2},i_{3}\leq q and a0a_{0} the amplitude defined in Proposition 4.10.

6.3. Critical points on the diagonal

The integral in Theorem 4.10 has two types of critical points. There are those on the large diagonals xi=xi+1x_{i}=x_{i+1} and regular critical points which are off diagonal.

Lemma 6.7.

[[Zel09]] The amplitude AσA_{\sigma} of 𝒩σ\mathcal{N}_{\sigma} is a semiclassical amplitude of order |σ|-|\sigma|.

Hence, the critical points on the large diagonal have lower order and we can focus on the portion of NN which quantizes the MM-fold billiard map on S+ΩS_{+}^{*}\partial\Omega.

Corollary 6.8.

The coefficient of kjk^{j} in the asymptotic expansion of the integral coming from 𝒩σ\mathcal{N}_{\sigma} in Theorem 4.10 is O(ε3j+|σ|)O(\varepsilon^{-3j+|\sigma|}).

Proof.

By Lemma 6.7, Trρ𝒩σ(k)\text{Tr}\rho^{\prime}\ast\mathcal{N}_{\sigma}(k) has an asymptotic expansion in negative powers of kk begining with k|σ|k^{-|\sigma|}. Hence, the coefficient of kjk^{-j} corresponds to the j|σ|j-|\sigma| term in the usual stationary phase expansion 32 and the maximal contribution of inverse Hessians to such a term is 3(j|σ|)3j3(j-|\sigma|)\leq 3j. ∎

When |σ|>0|\sigma|>0, these terms are lower order in 1/ε1/\varepsilon and can be included in the remainder so it suffices to only consider the contributions of regular critical points in the limit ε0\varepsilon\to 0. Notice also that the AσA_{\sigma}^{\prime} term in the amplitude a0a_{0} appearing in Theorem 4.10 is a semiclassical symbol of order 1|σ|-1-|\sigma| and can be discarded from the leading order asymptotics in kk by the same reasoning as in Corollary 6.8.

6.4. Proof of Theorem 1.1

We can now complete the derivation of the regularized resolvent trace formula appearing in Theorem 1.1.

Proof.

We will analyze the expression

𝒟γ(λ)j=0Bγ,jRe(λ)j12M=1σ:M{0,1}eMπi/4MΩSMeiλ(S)a0σ(λ,S)dS\displaystyle\mathcal{D}_{\gamma}(\lambda)\sum_{j=0}^{\infty}B_{\gamma,j}\operatorname{Re}(\lambda)^{-j}\sim^{*}\frac{1}{2}\sum_{M=1}^{\infty}\sum_{\sigma:\mathbb{Z}_{M}\to\{0,1\}}\frac{e^{-M\pi i/4}}{M}\int_{{\partial\Omega}_{S}^{M}}e^{i\lambda\mathcal{L}(S)}a_{0}^{\sigma}(\lambda,S)dS

in Proposition 4.10 by first expanding the amplitude a0σa_{0}^{\sigma} in inverse powers of kk from the representation 4.5 and then replacing each term by its stationary phase expansion. Recall that

a0σ(k+iτ,S)=ρ^((S))((S)Aσ(k+iτ,S)1ikAσ(k+iτ,S)),\displaystyle a_{0}^{\sigma}(k+i\tau,S)=\hat{\rho}(\mathcal{L}(S))\left(\mathcal{L}(S)A_{\sigma}(k+i\tau,S)-\frac{1}{i}\frac{\partial}{\partial k}A_{\sigma}(k+i\tau,S)\right),

with Aσ(k+iτ,S)A_{\sigma}(k+i\tau,S) being

(k+iτ)M/2i=1Mχiσ(k+iτ,S)|x(si)x(si+1)|1/2a1((k+iτ)|x(si)x(si+1)|)cosϑi,\displaystyle(k+i\tau)^{M/2}\prod_{i=1}^{M}\frac{\chi_{i}^{\sigma}(k+i\tau,S)}{|x(s_{i})-x(s_{i+1})|^{1/2}}a_{1}((k+i\tau)|x(s_{i})-x(s_{i+1})|)\cos\vartheta_{i},

and a1a_{1} the Hankel amplitude which has the asymptotic expansion 4.5. It is shown in [Zel09] (Corollary 3.9 and the proof of Proposition 3.10, together with Lemma 5.8) that

ΩMei(k+iτ)(S)a0σ((k+iτ),S)𝑑S=ΩM|σ|ei(k+iτ)(S)b0σ((k+iτ),S)𝑑S,\displaystyle\int_{\partial\Omega^{M}}e^{i(k+i\tau)\mathcal{L}(S)}a_{0}^{\sigma}((k+i\tau),S)dS=\int_{\partial\Omega^{M-|\sigma|}}e^{i(k+i\tau)\mathcal{L}(S)}b_{0}^{\sigma}((k+i\tau),S)dS,

for a symbol b0σSϱ|σ|(ΩM|σ|)b_{0}^{\sigma}\in S_{\varrho}^{-|\sigma|}(\partial\Omega^{M-|\sigma|}) having an asymptotic expansion of the form

b0σ(S,k+iτ)pb0,pσ(S)kσp,\displaystyle b_{0}^{\sigma}(S,k+i\tau)\sim\sum_{p}b_{0,p}^{\sigma}(S)k^{-\sigma-p},

with each b0,pσb_{0,p}^{\sigma} depending on at most pp derivatives of the boundary curvature. From this, an application of the stationary phase lemma gives

ΩMeik(S)a0σ((k+iτ),S)𝑑S(2πk)M|σ|2eik(Sγ)eiπsgn 2(Sγ)/4|det2(Sγ)|jcj,σkj|σ|\displaystyle\int_{\partial\Omega^{M}}e^{ik\mathcal{L}(S)}a_{0}^{\sigma}((k+i\tau),S)dS\sim\left(\frac{2\pi}{k}\right)^{\frac{M-|\sigma|}{2}}\frac{e^{ik\mathcal{L}(S_{\gamma})}e^{i\pi\text{sgn \,}\partial^{2}\mathcal{L}(S_{\gamma})/4}}{\sqrt{|\det\partial^{2}\mathcal{L}(S_{\gamma})|}}\sum_{j}c_{j,\sigma}k^{-j-|\sigma|}

where cj,σc_{j,\sigma} has 3j3j inverse Hessians. Having localized ρ^\hat{\rho} near LγL_{\gamma}, the integral is O(k)O(k^{-\infty}) unless M|σ|=qM-|\sigma|=q, in which case the coefficient of any kj0k^{-j_{0}} in the expansion has 3(j0|σ|)3(j_{0}-|\sigma|) inverse Hessians. After separating out the prefactor and using the fact that the derivatives of angles at reflection points depend only on the first order curvature jet, we see that the nonzero σ\sigma terms contribute 3j3|σ|3j-3|\sigma| inverse Hessians multiplied by a coefficient which depends smoothly on parameters and a priori on the first 6(j|σ|)6(j-|\sigma|) derivatives of the boundary curvature. However, it is shown in [Zel09] that this can be reduced to 2j2j derivatives of boundary curvature, essentially because in 32 and 33, powers of the higher order phase expansion gμg^{\mu} vanish to order 3μ3\mu. It takes 3μ3\mu derivatives to remove this zero, which leaves at most 2jμ2j-\mu derivatives left to act on the phase or amplitude. The amplitude only depends on the 0-jet of κ\kappa and gg depends on the 11-jet, but only appears when μ0\mu\neq 0. In either case, at most 2j2j derivatives of κ\kappa emerge.

We now restrict our attention to the case |σ|=0|\sigma|=0. Since |x(si)x(si+1)|diam(Ω)|Ω||x(s_{i})-x(s_{i+1})|\leq\text{diam}(\Omega)\leq|\partial\Omega|, we have that for SS near SγS_{\gamma},

a0(k+iτ,S)=b00(S,k+iτ)=(k+iτ2π)q2(S)i=1qcosϑi|x(si)x(si+1)|+a0,0(k+iτ,S),\displaystyle a_{0}(k+i\tau,S)=b_{0}^{0}(S,k+i\tau)=\left(\frac{k+i\tau}{2\pi}\right)^{\frac{q}{2}}\mathcal{L}(S)\prod_{i=1}^{q}\frac{\cos\vartheta_{i}}{|x(s_{i})-x(s_{i+1})|}+\mathcal{R}_{a_{0},0}(k+i\tau,S),

where the remainder satisfies

|zαkβa0,0(k+iτ,S)|Cα,β,ϱ,|Ω|,qkϱ|α||β|q21\displaystyle|\partial_{z}^{\alpha}\partial_{k}^{\beta}\mathcal{R}_{a_{0},0}(k+i\tau,S)|\leq C_{\alpha,\beta,\varrho,|\partial\Omega|,q}\langle k\rangle^{\varrho|\alpha|-|\beta|-\frac{q}{2}-1}

uniformly and with smooth dependence on parameters, including ε\varepsilon and the full jet of κ\kappa. Here, we have chosen only leading order terms in the asymptotic expansion of a1a_{1}. The contributions of a0,0\mathcal{R}_{a_{0},0} and kAσSϱ1\frac{\partial}{\partial k}A_{\sigma}\in S_{\varrho}^{-1} to the kjk^{-j} term in the stationary phase expansion have submaximal inverse Hessians for essentially the same reason as the |σ|1|\sigma|\geq 1 terms: an asymptotic expansion of either which starts with kj~k^{-\widetilde{j}} (j~1\widetilde{j}\geq 1) contributes to the kjk^{-j} term in stationary phase expansion through the operator Ljj~L_{j-\widetilde{j}}, which has at most 3(jj~)3(j-\widetilde{j}) inverse Hessians.

Thus, for the purposes of calculating wave invariants modulo lower order inverse Hessians, we may replace the integral in Proposition 4.10 by

(35) (k+iτ2π)q2Ωqei(k+iτ)(S)(S)i=1qcosϑi|x(si)x(si+1)|dS,\displaystyle\left(\frac{k+i\tau}{2\pi}\right)^{\frac{q}{2}}\int_{\partial\Omega^{q}}e^{i(k+i\tau)\mathcal{L}(S)}\mathcal{L}(S)\prod_{i=1}^{q}\frac{\cos\vartheta_{i}}{|x(s_{i})-x(s_{i+1})|}dS,

to which we apply the stationary phase lemma with parameter τ[0,1]\tau\in[0,1]. As noted in Section 4, holomorphy of the regularized resolvent trace allows us to take τ0+\tau\to 0^{+}. Denote by a~(λ,S)\widetilde{a}(\lambda,S) the integrand of 35, corresponding to the amplitude with |σ|=0|\sigma|=0 and highest powers of ε1\varepsilon^{-1}. To find explicitly the stationary phase coefficients, we apply the Feynman rules. In Corollary 6.6, we saw the terms which appear in each Feynman amplitide. Most labeled diagrams contribute a Feynman amplitude of zero, since the derivatives of \mathcal{L} are only nonzero if max{|i1i2|,|i1i3|,|i2i3|}1\max\{|i_{1}-i_{2}|,|i_{1}-i_{3}|,|i_{2}-i_{3}|\}\leq 1. Hence, the combinatorial constant depends only on the order of the automorphism group of 33-regular graphs on 2j2j vertices. Plugging in a~\widetilde{a} and \mathcal{L} to the formula in Proposition 6.4, and considering a 33-regular graph 𝒢\mathcal{G}, we sum over all labelings to obtain the Feynman amplitude

F𝒢,=\displaystyle\sum_{\ell}F_{\mathcal{G},\ell}= a~(Sγ)edgesvertices(iki1(),i2(),i3()3)(ikh~j1()j2())\displaystyle\sum_{\ell}\widetilde{a}(S_{\gamma})\prod_{\text{edges}}\prod_{\text{vertices}}\left(ik\partial_{i_{1}(\ell),i_{2}(\ell),i_{3}(\ell)}^{3}\mathcal{L}\right)\left(\frac{i}{k}\widetilde{h}_{j_{1}(\ell)j_{2}(\ell)}\right)
=\displaystyle= (±i)jkj(cγε)3ja~(Sγ)(i1,i2,i3hi1hi2hi3i1,i2,i33(Sγ))2j.\displaystyle(\pm i)^{j}k^{-j}\left(c_{\gamma}\varepsilon\right)^{-3j}\widetilde{a}(S_{\gamma})\left(\sum_{i_{1},i_{2},i_{3}}h_{i_{1}}h_{i_{2}}h_{i_{3}}\partial_{i_{1},i_{2},i_{3}}^{3}\mathcal{L}(S_{\gamma})\right)^{2j}.

Putting M=qM=q, |σ|=0|\sigma|=0, dividing by the order of the automorphism groups, summing over all Feynman diagrams and collecting remainder terms finishes the proof. ∎

7. Future directions

As mentioned in the introduction, this paper is part of a series in which we aim to show that the singular support of the wave trace and the length spectrum are inherently different objects. Our subsequent paper will produce cancellations to arbitrarily high orders in the wave trace. It would also be interesting to both derive degenerate asymptotics and create cancellations in:

  • the wave trace for Riemannian manifolds without boundary,

  • the semiclassical trace formula for the density of states of a Schrödinger operator on a Riemannian manifold with or without boundary, and

  • and the semiclassical trace formula for eigenvalues of large graphs.

In the setting of convex planar domains considered in the present paper, the billiard ball map forms a sort of global Poincaré expression for the billiard flow, which consists of broken bicharacteristics. Formally, our result can then be interpreted as an asymptotic expansion of the trace of a certain boundary operator which quantizes the Poincaré map, analogous to the approach in [SZ02] and [ISZ02]. The authors there use an abstract reduction to a Grushin problem and in future work, it would be interesting to compare their method with the layer potential approach we employ here.

8. Acknowledgements

The first and second authors acknowledge the partial support of the ERC Grant #885707. The third author is grateful to IST Austria and the Schrödinger Institute in Vienna for hosting him during much of the writing of this paper. The authors would also like to thank Hamid Hezari for suggesting the Balian-Bloch approach and the late Steven Morris Zelditch for many helpful conversations in the begining of this project. We are grateful to have known Zelditch for many years and are deeply saddened by his passing. His contributions have made a profound impact on many fields in mathematics. His energy, passion and enthusiasm uplifted all those around him. He was a great inspiration to many mathematicians, including ourselves.

References

  • [ADSK16] Artur Avila, Jacopo De Simoi, and Vadim Kaloshin. An integrable deformation of an ellipse of small eccentricity is an ellipse. Ann. of Math. (2), 184(2):527–558, 2016.
  • [AM77] K. G. Andersson and R. B. Melrose. The propagation of singularities along gliding rays. Invent. Math., 41(3):197–232, 1977.
  • [Axe97] Scott Axelrod. Overview and warmup example for perturbation theory with instantons. In Geometry and physics (Aarhus, 1995), volume 184 of Lecture Notes in Pure and Appl. Math., pages 321–338. Dekker, New York, 1997.
  • [BB71] R. Balian and C. Bloch. Distribution of eigenfrequencies for the wave equation in a finite domain. II. Electromagnetic field. Riemannian spaces. Ann. Physics, 64:271–307, 1971.
  • [BB72] R. Balian and C. Bloch. Distribution of eigenfrequencies for the wave equation in a finite domain. III. Eigenfrequency density oscillations. Ann. Physics, 69:76–160, 1972.
  • [BB74] R. Balian and C. Bloch. Errata: “Distribution of eigenfrequencies for the wave equation in a finite domain. I. Three-dimensional problem with smooth boundary surface” (Ann. Physics 60 (1970), 401–447). Ann. Physics, 84:559, 1974.
  • [Bia93] Misha Bialy. Convex billiards and a theorem by E. Hopf. Math. Z., 214(1):147–154, 1993.
  • [Cal22] Keagan G. Callis. Absolutely Periodic Billiard Orbits of Arbitrarily High Order. arXiv e-prints, page arXiv:2209.11721, September 2022.
  • [CdV84] Y. Colin de Verdière. Sur les longueurs des trajectoires périodiques d’un billard. In South Rhone seminar on geometry, III (Lyon, 1983), Travaux en Cours, pages 122–139. Hermann, Paris, 1984.
  • [Chr17] T. J. Christiansen. A sharp lower bound for a resonance-counting function in even dimensions. Ann. Inst. Fourier (Grenoble), 67(2):579–604, 2017.
  • [DG75] Johannes J. Duistermaat and Victor W. Guillemin. The spectrum of positive elliptic operators and periodic bicharacteristics. Invent. Math., 29(1):39–79, 1975.
  • [dSKW17] Jacopo de Simoi, Vadim Kaloshin, and Qiaoling Wei. Dynamical spectral rigidity among 2\mathbb{Z}_{2}-symmetric strictly convex domains close to a circle (appendix b coauthored with h. hezari). Ann. of Math. (2), 186(1):277–314, 2017. Appendix B coauthored with H. Hezari.
  • [Dui96] Johannes J. Duistermaat. Fourier integral operators, volume 130 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1996.
  • [GK80] V. Guillemin and D. Kazhdan. Some inverse spectral results for negatively curved 22-manifolds. Topology, 19(3):301–312, 1980.
  • [GL18] Colin Guillarmou and Thibault Lefeuvre. The marked length spectrum of Anosov manifolds. ArXiv e-prints, June 2018.
  • [GM79] Victor Guillemin and Richard Melrose. The Poisson summation formula for manifolds with boundary. Adv. in Math., 32(3):204–232, 1979.
  • [GWW92] Carolyn Gordon, David L. Webb, and Scott Wolpert. One cannot hear the shape of a drum. Bull. Amer. Math. Soc. (N.S.), 27(1):134–138, 1992.
  • [HKS18] Guan Huang, Vadim Kaloshin, and Alfonso Sorrentino. On the marked length spectrum of generic strictly convex billiard tables. Duke Math. J., 167(1):175–209, 01 2018.
  • [Hör03] Lars Hörmander. The analysis of linear partial differential operators. I. Classics in Mathematics. Springer-Verlag, Berlin, 2003. Distribution theory and Fourier analysis, Reprint of the second (1990) edition [Springer, Berlin; MR1065993 (91m:35001a)].
  • [HZ04] Andrew Hassell and Steve Zelditch. Quantum ergodicity of boundary values of eigenfunctions. Comm. Math. Phys., 248(1):119–168, 2004.
  • [HZ10] Hamid Hezari and Steve Zelditch. Inverse spectral problem for analytic (/2)n(\mathbb{Z}/2\mathbb{Z})^{n}-symmetric domains in n\mathbb{R}^{n}. Geom. Funct. Anal., 20(1):160–191, 2010.
  • [HZ12] Hamid Hezari and Steve Zelditch. CC^{\infty} spectral rigidity of the ellipse. Anal. PDE, 5(5):1105–1132, 2012.
  • [HZ22] Hamid Hezari and Steve Zelditch. One can hear the shape of ellipses of small eccentricity. Ann. of Math. (2), 196(3):1083–1134, 2022.
  • [ISZ02] Alexei Iantchenko, Johannes Sjöstrand, and Maciej Zworski. Birkhoff normal forms in semi-classical inverse problems. arXiv preprint math/0201190, 2002.
  • [Kac66] Mark Kac. Can one hear the shape of a drum? Amer. Math. Monthly, 73(4, part II):1–23, 1966.
  • [Kat05] Anatole B. Katok. Billiard table as a playground for a mathematician. In Surveys in modern mathematics, volume 321 of London Math. Soc. Lecture Note Ser., pages 216–242. Cambridge Univ. Press, Cambridge, 2005.
  • [Kov21] Illya Koval. Local strong Birkhoff conjecture and local spectral rigidity of almost every ellipse. arXiv e-prints, page arXiv:2111.12171, November 2021.
  • [KS18] Vadim Kaloshin and Alfonso Sorrentino. On the local Birkhoff conjecture for convex billiards. Ann. of Math. (2), 188(1):315–380, 2018.
  • [KT91] Valeriĭ V. Kozlov and Dmitriĭ V. Treshchëv. Billiards, volume 89 of Translations of Mathematical Monographs. American Mathematical Society, Providence, RI, 1991. A genetic introduction to the dynamics of systems with impacts, Translated from the Russian by J. R. Schulenberger.
  • [KZ18] Vadim Kaloshin and Ke Zhang. Density of convex billiards with rational caustics. Nonlinearity, 31(11):5214–5234, 2018.
  • [Mil64] J. Milnor. Eigenvalues of the laplace operator on certain manifolds. Proceedings of the National Academy of Sciences, 51(4):542–542, 1964.
  • [MM82] Shahla Marvizi and Richard Melrose. Spectral invariants of convex planar regions. J. Differential Geom., 17(3):475–502, 1982.
  • [OPS88a] Brad Osgood, Ralph Phillips, and Peter Sarnak. Compact isospectral sets of plane domains. Proc. Nat. Acad. Sci. U.S.A., 85(15):5359–5361, 1988.
  • [OPS88b] Brad Osgood, Ralph Phillips, and Peter Sarnak. Compact isospectral sets of surfaces. J. Funct. Anal., 80(1):212–234, 1988.
  • [OPS88c] Brad Osgood, Ralph Phillips, and Peter Sarnak. Extremals of determinants of Laplacians. J. Funct. Anal., 80(1):148–211, 1988.
  • [Pop94] Georgi Popov. Invariants of the length spectrum and spectral invariants of planar convex domains. Comm. Math. Phys., 161(2):335–364, 1994.
  • [PS92] Vesselin M. Petkov and Luchezar N. Stoyanov. Geometry of reflecting rays and inverse spectral problems. Pure and Applied Mathematics (New York). John Wiley & Sons, Ltd., Chichester, 1992.
  • [PS17] Vesselin M. Petkov and Luchezar N. Stoyanov. Geometry of the generalized geodesic flow and inverse spectral problems. John Wiley & Sons, Ltd., Chichester, second edition, 2017.
  • [PT11] Georgi Popov and Petar Topalov. On the integral geometry of Liouville billiard tables. Comm. Math. Phys., 303(3):721–759, 2011.
  • [PT12] Georgi Popov and Peter Topalov. Invariants of isospectral deformations and spectral rigidity. Comm. Partial Differential Equations, 37(3):369–446, 2012.
  • [Sib04] Karl Friedrich Siburg. The principle of least action in geometry and dynamics. Number 1844. Springer Science & Business Media, 2004.
  • [Sjö97] J. Sjöstrand. A trace formula and review of some estimates for resonances. In Microlocal analysis and spectral theory (Lucca, 1996), volume 490 of NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci., pages 377–437. Kluwer Acad. Publ., Dordrecht, 1997.
  • [SM85] Toshikazu Sunada and Yozō Matsushima. Riemannian coverings and isospectral manifolds. Annals of Mathematics, 121:169, 1985.
  • [SZ02] Johannes Sjöstrand and Maciej Zworski. Quantum monodromy and semi-classical trace formulae. Journal de mathématiques pures et appliquées, 81(1):1–33, 2002.
  • [Tab95] Serge Tabachnikov. Billiards. (No Title), 1995.
  • [Tay13] Michael Taylor. Partial differential equations II: Qualitative studies of linear equations, volume 116. Springer Science & Business Media, 2013.
  • [Vig80] Marie-France Vignéras. Variétés riemanniennes isospectrales et non isométriques. Ann. of Math. (2), 112(1):21–32, 1980.
  • [Vig20] Amir B. Vig. The Inverse Spectral Problem for Convex Planar Domains. ProQuest LLC, Ann Arbor, MI, 2020. Thesis (Ph.D.)–University of California, Irvine.
  • [Vig21] Amir Vig. Robin spectral rigidity of the ellipse. J. Geom. Anal., 31(3):2238–2295, 2021.
  • [Vig23a] Amir Vig. Compactness of marked length isospectral sets of birkhoff billiard tables. arXiv preprint arXiv:2310.05426, 2023.
  • [Vig23b] Amir Vig. The wave trace and birkhoff billiards. Journal of Spectral Theory, 12(3):877–938, 2023.
  • [WYZ23] Jared Wunsch, Mengxuan Yang, and Yuzhou Zou. The Morse index theorem for mechanical systems with reflections. arXiv e-prints, page arXiv:2308.16162, August 2023.
  • [Zel98] Steve Zelditch. The inverse spectral problem for surfaces of revolution. J. Differential Geom., 49(2):207–264, 1998.
  • [Zel00] S. Zelditch. Spectral determination of analytic bi-axisymmetric plane domains. Geom. Funct. Anal., 10(3):628–677, 2000.
  • [Zel04a] Steve Zelditch. Inverse resonance problem for 2\mathbb{Z}_{2}-symmetric analytic obstacles in the plane. In Christopher B. Croke, Michael S. Vogelius, Gunther Uhlmann, and Irena Lasiecka, editors, Geometric Methods in Inverse Problems and PDE Control, pages 289–321, New York, NY, 2004. Springer New York.
  • [Zel04b] Steve Zelditch. The inverse spectral problem. In Surveys in differential geometry. Vol. IX, volume 9 of Surv. Differ. Geom., pages 401–467. Int. Press, Somerville, MA, 2004. With an appendix by Johannes Sjöstrand and Maciej Zworski.
  • [Zel04c] Steve Zelditch. Inverse spectral problem for analytic domains. I. Balian-Bloch trace formula. Comm. Math. Phys., 248(2):357–407, 2004.
  • [Zel09] Steve Zelditch. Inverse spectral problem for analytic domains. II. 2\mathbb{Z}_{2}-symmetric domains. Ann. of Math. (2), 170(1):205–269, 2009.
  • [Zwo98] Maciej Zworski. Poisson formula for resonances in even dimensions. Asian J. Math., 2(3):609–617, 1998.