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

\nouppercaseheads\externaldocument

[book_new-]book_new[http://www.math.toronto.edu/ivrii/futurebook.pdf]

\addtopsmarks

headings\createmarkchapterrightshownumber.

Spectral asymptotics for the semiclassical Dirichlet to Neumann operatorthanks: 2010 Mathematics Subject Classification: 35P20, 58J50.thanks: Key words and phrases: Dirichlet-to-Neumann operator, semiclassical Dirichlet-to-Neumann operator, spectral asymptotics.

Andrew Hassell This research was supported in part by Australian Research Council Discovery Grant DP120102019    Victor Ivrii This research was supported in part by National Science and Engineering Research Council (Canada) Discovery Grant RGPIN 13827
Abstract

Let MM be a compact Riemannian manifold with smooth boundary, and let R(λ)R(\lambda) be the Dirichlet-to-Neumann operator at frequency λ\lambda. The semiclassical Dirichlet-to-Neumann operator R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda) is defined to be λ1R(λ)\lambda^{-1}R(\lambda). We obtain a leading asymptotic for the spectral counting function for R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda) in an interval [a1,a2)[a_{1},a_{2}) as λ\lambda\to\infty, under the assumption that the measure of periodic billiards on TMT^{*}M is zero. The asymptotic takes the form

𝖭(λ;a1,a2)=(κ(a2)κ(a1))vol(M)λd1+o(λd1),{\mathsf{N}}(\lambda;a_{1},a_{2})=\bigl{(}\kappa(a_{2})-\kappa(a_{1})\bigr{)}\operatorname{vol}^{\prime}(\partial M)\lambda^{d-1}+o(\lambda^{d-1}),

where κ(a)\kappa(a) is given explicitly by

κ(a)=ωd1(2π)d1(12π11(1η2)(d1)/2aa2+η2dη14+H(a)(1+a2)(d1)/2).\kappa(a)=\frac{\omega_{d-1}}{(2\pi)^{d-1}}\bigg{(}-\frac{1}{2\pi}\int_{-1}^{1}(1-\eta^{2})^{(d-1)/2}\frac{a}{a^{2}+\eta^{2}}\,d\eta\\ -\frac{1}{4}+H(a)(1+a^{2})^{(d-1)/2}\bigg{)}.

Chapter 1 Introduction

Let MM be a Riemannian manifold with boundary. The Dirichlet-to-Neumann operator is a family of operators defined on L2(M)L^{2}({\partial M}) depending on the parameter λ0\lambda\geq 0. It is defined as follows: given fL2(M)f\in L^{2}({\partial M}), we solve the equation (if possible)

(1.1) (Δλ2)u=0on M,u|M=f.(\Delta-\lambda^{2})u=0\quad\text{on\ \ }M,\qquad u|_{{\partial M}}=f.

Then the Dirichlet-to-Neumann operator at frequency λ\lambda is the map

(1.2) R(λ):fνu|M.R(\lambda):f\mapsto-\partial_{\nu}u|_{{\partial M}}.

Here ν\partial_{\nu} is the interior unit normal derivative, and Δ\Delta is the positive Laplacian on MM.

It is well known that R(λ)R(\lambda) is a self-adjoint, semi-bounded from below pseudodifferential operator of order 11 on L2(M)L^{2}({\partial M}), with domain H1(M)H^{1}(\partial M). It therefore has discrete spectrum accumulating only at ++\infty. The Dirichlet-to-Neumann operator and closely related operators are important in a number of areas of mathematical analysis including inverse problems (such as Calderón’s problem [3]), domain decomposition problems (such as the determinant gluing formula of Burghelea-Friedlander-Kappeler [2]), and spectral asymptotics (see e.g. [7]).

In this paper, we are interested in the spectral asymptotics of R(λ)R(\lambda) in the high-frequency limit, λ\lambda\to\infty. Let us recall standard spectral asymptotics for elliptic differential operators, for simplicity in the simplest case of a positive self-adjoint second order scalar operator. Suppose that QQ is such an operator on a manifold MM of dimension dd, with principal symbol qq. Then in the case that MM is closed, we have an asymptotic for the number 𝖭(λ){\mathsf{N}}(\lambda) of eigenvalues of QQ (counted with multiplicity) less than λ2\lambda^{2}:

(1.3) 𝖭(λ)=(2π)dvol{(x,ξ)TMq(x,ξ)λ2}+O(λd1)=(λ2π)dvol{(x,ξ)TMq(x,ξ)1}+O(λd1),{\mathsf{N}}(\lambda)=(2\pi)^{-d}\operatorname{vol}\{(x,\xi)\in T^{*}M\mid q(x,\xi)\leq\lambda^{2}\}+O(\lambda^{d-1})\\ =\big{(}\frac{\lambda}{2\pi}\big{)}^{d}\operatorname{vol}\{(x,\xi)\in T^{*}M\mid q(x,\xi)\leq 1\}+O(\lambda^{d-1}),

where the equality of the two expressions on the RHS is a simple consequence of the homogeneity of qq. Moreover, if the set of periodic geodesics has measure zero, then there is a two-term expansion of the form

(λ2π)dvol{(x,ξ)TMq(x,ξ)1}+λd1(2π)d{q=1}sub(Q)+o(λd1)\bigl{(}\frac{\lambda}{2\pi}\bigr{)}^{d}\operatorname{vol}\{(x,\xi)\in T^{*}M\mid q(x,\xi)\leq 1\}+\frac{\lambda^{d-1}}{(2\pi)^{d}}\int_{\{q=1\}}\operatorname{sub}(Q)+o(\lambda^{d-1})

where sub(Q)\operatorname{sub}(Q) is the subprincipal symbol of QQ [5]. This was generalised to the case of manifolds with boundary by the second author [9]. For simplicity we state the result in the case that Q=ΔQ=\Delta is the (positive) metric Laplacian, which satisfies sub(Δ)=0\operatorname{sub}(\Delta)=0. Then Δ\Delta is a self-adjoint operator under either Dirichlet ()(-) or Neumann (+)(+) boundary conditions, and if the set of periodic generalised bicharacteristics has measure zero, we get a two-term expansion for 𝖭Δ(λ){\mathsf{N}}_{\Delta}(\lambda) of the form

(1.4) (λ2π)dvolBM±14(λ2π)d1volBM+o(λd1).\bigl{(}\frac{\lambda}{2\pi}\bigr{)}^{d}\operatorname{vol}B^{*}M\pm\frac{1}{4}\big{(}\frac{\lambda}{2\pi}\big{)}^{d-1}\operatorname{vol}B^{*}\partial M+o(\lambda^{d-1}).

These statements can be generalized to the semiclassical setting. Consider a classical Schrödinger operator on MM, P=h2Δ+V(x)1P=h^{2}\Delta+V(x)-1, where h>0h>0 is a small parameter (“Planck’s constant”) and VV is a smooth real-valued function. We consider the asymptotic behaviour 𝖭h(P){\mathsf{N}}^{-}_{h}(P) of the number of negative eigenvalues of PP as h0h\to 0. This is equivalent to the problem above if h=λ1h=\lambda^{-1} and VV is identically zero. Define p(x,ξ)p(x,\xi) to be the semiclassical symbol of PP, i.e. p=|ξ|g(x)2+V(x)1p=|\xi|^{2}_{g(x)}+V(x)-1. Then, if MM is closed, under the assumption that the measure of periodic bicharacteristics of PP is zero in TMT^{*}M, and that 0 is a regular value for pp, we have

(1.5) 𝖭h(P)=(2πh)dvol{(x,ξ)TMp(x,ξ)0}+O(h1d).{\mathsf{N}}^{-}_{h}(P)=(2\pi h)^{-d}\operatorname{vol}\{(x,\xi)\in T^{*}M\mid p(x,\xi)\leq 0\}+O(h^{1-d}).

Moreover, for manifolds with boundary, we have an analogue of (1.4): under either Dirichlet ()(-) or Neumann (+)(+) boundary conditions, if the set of periodic generalised bicharacteristics has measure zero, we get a two-term expansion for 𝖭h(P){\mathsf{N}}^{-}_{h}(P) (where here we understand the self-adjoint realization of PP with either Dirichlet or Neumann boundary condition) of the form

(1.6) (2πh)dvol{(x,ξ)TMp(x,ξ)0}±14(2πh)1dvol+o(h1d),(2\pi h)^{-d}\operatorname{vol}\{(x,\xi)\in T^{*}M\mid p(x,\xi)\leq 0\}\pm\frac{1}{4}(2\pi h)^{1-d}\operatorname{vol}\mathcal{H}+o(h^{1-d}),

where T(M)\mathcal{H}\subset T^{*}(\partial M) is the hyperbolic region in the boundary, that is, the projection of the set {(x,ξ)p(x,ξ)0}TMM\{(x,\xi)\mid p(x,\xi)\leq 0\}\cap T^{*}_{\partial M}M to TMT^{*}\partial M.

From the semiclassical point of view, since R(λ)R(\lambda) is a first order operator, it makes sense to consider R𝗌𝖼𝗅(λ):=λ1R(λ)R_{\mathsf{scl}}(\lambda):=\lambda^{-1}R(\lambda) (for λ>0\lambda>0), which we call the semiclassical Dirichlet-to-Neumann operator. Like R(λ)R(\lambda), it is a self-adjoint, semi-bounded from below operator on L2(M)L^{2}({\partial M}), with discrete spectrum accumulating only at ++\infty. The goal of this paper is to investigate the spectral asymptotics of R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda), that is, the asymptotics of

(1.7) 𝖭(λ;a1,a2):=#{β:β is an eigenvalue of R𝗌𝖼𝗅(λ),a1β<a2},{\mathsf{N}}(\lambda;a_{1},a_{2})\mathrel{\mathop{:}}=\#\{\beta:\,\beta\text{\ is an eigenvalue of \ }R_{\mathsf{scl}}(\lambda),\ a_{1}\leq\beta<a_{2}\},

the number of eigenvalues of R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda) in the interval [a1,a2)[a_{1},a_{2}), as λ\lambda\to\infty.

Both R(λ)R(\lambda) and R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda) have the disadvantage that they are undefined whenever λ2\lambda^{2} is a Dirichlet eigenvalue, since then (1.1) is not solvable for arbitrary fH1(M)f\in H^{1}(M). Indeed, when λ2\lambda^{2} is a Dirichlet eigenvalue, a necessary condition for solvability of (1.1) is that ff is orthogonal to the normal derivatives of Dirichlet eigenfunctions at frequency λ\lambda. To overcome this issue, we introduce the Cayley transform of R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda): we define

(1.8) C(λ)=(R𝗌𝖼𝗅(λ)i)(R𝗌𝖼𝗅(λ)+i)1.C(\lambda)=(R_{\mathsf{scl}}(\lambda)-i)(R_{\mathsf{scl}}(\lambda)+i)^{-1}.

This family of operators is related to impedance boundary conditions: we have C(λ)f=gC(\lambda)f=g if and only if there is a function uu on MM satisfying

(1.9) (Δλ2)u=0\displaystyle(\Delta-\lambda^{2})u=0
and
(1)1,2\textup{(\ref*{eq-1-10})}_{1,2} 12(λ1νuiu)=f,12(λ1νu+iu)=g.\displaystyle\frac{1}{2}(\lambda^{-1}\partial_{\nu}u-iu)=f,\qquad\frac{1}{2}(\lambda^{-1}\partial_{\nu}u+iu)=g.

As observed in [1], C(λ)C(\lambda) is a well-defined analytic family of operators for λ\lambda in a neighbourhood of the positive real axis, which is unitary on the real axis. In particular, it is well-defined even when λ2\lambda^{2} is a Dirichlet eigenvalue of the Laplacian on MM. As a unitary operator, C(λ)C(\lambda), λ>0\lambda>0, has its spectrum on the unit circle, and as R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda) has discrete spectrum accumulating only at \infty, it follows that the spectrum of C(λ)C(\lambda) is discrete on the unit circle except at the point 11. Our question can be formulated in terms of C(λ)C(\lambda): given two angles θ1,θ2\theta_{1},\theta_{2} satisfying 0<θ1<θ2<2π0<\theta_{1}<\theta_{2}<2\pi, what is the leading asymptotic for

(1.11) 𝖭~(λ;θ1,θ2):=#{eiθ:eiθ is an eigenvalue of C(λ),θ1θ<θ2}\tilde{\mathsf{N}}(\lambda;\theta_{1},\theta_{2})\mathrel{\mathop{:}}=\#\{e^{i\theta}:\,e^{i\theta}\text{\ is an eigenvalue of \ }C(\lambda),\ \theta_{1}\leq\theta<\theta_{2}\}

the number of eigenvalues of C(λ)C(\lambda) in the interval {eiθ:θ[θ1,θ2)}\{e^{i\theta}:\,\theta\in[\theta_{1},\theta_{2})\} of the unit circle, as λ\lambda\to\infty. Clearly, we have

(1.12) 𝖭~(λ;θ1,θ2)=𝖭(λ;a1,a2),where eiθj=ajiaj+i, i.e. aj=cot(θj2).\tilde{{\mathsf{N}}}(\lambda;\theta_{1},\theta_{2})={\mathsf{N}}(\lambda;a_{1},a_{2}),\ \text{where }\ e^{i\theta_{j}}=\frac{a_{j}-i}{a_{j}+i},\text{ i.e. }a_{j}=-\cot\bigl{(}\frac{\theta_{j}}{2}\bigr{)}.

To answer this question we relate it to a standard semiclassical eigenvalue counting problem on MM. To state the next result, we first define the self-adjoint operator Pa,hP_{a,h} on L2(M)L^{2}(M) by

(1.13) 𝔇(Pa,h)={uH2(M):(hν+a)u=0 at M},\displaystyle\mathfrak{D}(P_{a,h})=\{u\in H^{2}(M):\,(h\partial_{\nu}+a)u=0\text{\ \ at\ \ }{\partial M}\},
(1.14) Pa,h(u)=(h2Δ1)u,u𝔇(Pa,h).\displaystyle P_{a,h}(u)=(h^{2}\Delta-1)u,\quad u\in\mathfrak{D}(P_{a,h}).

It is the self-adjoint operator associated to the semi-bounded quadratic form

(1.15) h2uM2uM2hauM2.h^{2}\|\nabla u\|_{M}^{2}-\|u\|_{M}^{2}-ha\|u\|^{2}_{\partial M}.

The operator Pa,hP_{a,h} is linked with the semiclassical Dirichlet-to-Neumann operator as follows: if ff is an eigenfunction of R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda) with eigenvalue aa, then the corresponding Helmholtz function uu defined by (1.1) is in the domain (1.13) of Pa,hP_{a,h}, and Pa,hu=0P_{a,h}u=0 (where h=λ1h=\lambda^{-1}).

Then we have the following result, proved in Section 2.

Proposition 1.1.

Let h=λ1h=\lambda^{-1}. Assume 0<θ1<θ2<2π0<\theta_{1}<\theta_{2}<2\pi. Then the number of eigenvalues of C(λ)C(\lambda) in the interval Jθ1,θ2:={eiθ:θ[θ1,θ2)}J_{\theta_{1},\theta_{2}}:=\{e^{i\theta}:\,\theta\in[\theta_{1},\theta_{2})\} is equal to

(1.16) 𝖭~(λ;θ1,θ2)=𝖭(λ;a1,a2)=𝖭h(a2)𝖭h(a1),\displaystyle\tilde{{\mathsf{N}}}(\lambda;\theta_{1},\theta_{2})={\mathsf{N}}(\lambda;a_{1},a_{2})={\mathsf{N}}_{h}^{-}(a_{2})-{\mathsf{N}}^{-}_{h}(a_{1}),
where aj=cot(θj/2)a_{j}=-\cot(\theta_{j}/2) and
(1.17) 𝖭h(a):=#{μ:μ is an eigenvalue of Pa,h,μ<0}.\displaystyle{\mathsf{N}}_{h}^{-}(a)\mathrel{\mathop{:}}=\#\{\mu:\,\mu\text{\ is an eigenvalue of\ \ }P_{a,h},\ \mu<0\}.

Having thus reduced the problem to a standard question about semiclassical spectral asymptotics, we obtain (after some calculations in Section 3) our main result.

Theorem 1.2.
  1. (i)

    The following estimate for the quantity (1.11) holds:

    (1.18) 𝖭(λ;a1,a2)=O(λd1);{\mathsf{N}}(\lambda;a_{1},a_{2})=O(\lambda^{d-1});
  2. (ii)

    Further, if the set of periodic billiards on MM has measure 0 then the following asymptotic holds as λ+\lambda\to+\infty:

    (1.19) 𝖭(λ;a1,a2)=(κ(a2)κ(a1))vol(M)λd1+o(λd1),{\mathsf{N}}(\lambda;a_{1},a_{2})=\bigl{(}\kappa(a_{2})-\kappa(a_{1})\bigr{)}\operatorname{vol}^{\prime}(\partial M)\lambda^{d-1}+o(\lambda^{d-1}),

    where κ(a)\kappa(a) is given explicitly by

    (1.20) κ(a)=ωd1(2π)d1(12π11(1η2)(d1)/2aa2+η2dη14+H(a)(1+a2)(d1)/2).\kappa(a)=\frac{\omega_{d-1}}{(2\pi)^{d-1}}\bigg{(}-\frac{1}{2\pi}\int_{-1}^{1}(1-\eta^{2})^{(d-1)/2}\frac{a}{a^{2}+\eta^{2}}\,d\eta\\ -\frac{1}{4}+H(a)(1+a^{2})^{(d-1)/2}\bigg{)}.

    Here H()H(\cdot) is the Heaviside function, ωd\omega_{d} is the volume of the unit ball in d\mathbb{R}^{d}, and vol(M)\operatorname{vol}(M) and vol(M)\operatorname{vol}^{\prime}(\partial M) are dd-dimensional volume of MM and (d1)(d-1)-dimensional volume of M\partial M respectively.

  3. (iii)

    In the case d=3d=3, we can evaluate this integral exactly and we find that

    (1.21) κ(a)=14π(141πarccot(a)(1+a2)+(1+a2)+1πa)\kappa(a)=\frac{1}{4\pi}\Bigl{(}-\frac{1}{4}-\frac{1}{\pi}\operatorname{arccot}(a)(1+a^{2})+(1+a^{2})+\frac{1}{\pi}a\Bigr{)}

    where arccot\operatorname{arccot} has range (0,π)(0,\pi). This is simpler expressed in terms of θ\theta. Defining κ~(θ)=κ(a)\tilde{\kappa}(\theta)=\kappa(a) where a=cot(θ/2)=cot(πθ/2)a=-\cot(\theta/2)=\cot(\pi-\theta/2), we have (still under the zero-measure assumption on periodic billiards)

    (1.22) 𝖭~(λ;θ1,θ2)=(κ~(θ2)κ~(θ1))vol(M)λ2+o(λ2),\displaystyle\tilde{{\mathsf{N}}}(\lambda;\theta_{1},\theta_{2})=\bigl{(}\tilde{\kappa}(\theta_{2})-\tilde{\kappa}(\theta_{1})\bigr{)}\operatorname{vol}^{\prime}(\partial M)\lambda^{2}+o(\lambda^{2}),
    (1.23) κ~(θ)=14π(14+12π(θsinθsin2(θ/2))).\displaystyle\tilde{\kappa}(\theta)=\frac{1}{4\pi}\Big{(}-\frac{1}{4}+\frac{1}{2\pi}\big{(}\frac{\theta-\sin\theta}{\sin^{2}(\theta/2)}\big{)}\Big{)}.
Remark 1.3.
  1. (i)

    It looks disheartening that the remainder is just “oo” in comparison with the principal part and only under geometric condition of the global nature but it is the nature of the beast. Consider MM a hemisphere; then for λn2=n(n+d1)\lambda_{n}^{2}=n(n+d-1) with n+n\in\mathbb{Z}^{+} the operator R(λ)R(\lambda) has eigenvalue 0 of multiplicity nd1\asymp n^{d-1} and therefore we do not have even an asymptotic but just an estimate.

  2. (ii)

    We can explain this by pointing out that the problem we consider is truly dd-dimensional, in contrast, for example, to the problem of distribution of negative eigenvalues of H=ΔH=\Delta with

    𝔇(H)={uH2(X):νu|M+aΔb1/2u|M=0}\mathfrak{D}(H)=\{u\in H^{2}(X):\,\partial_{\nu}u|_{\partial M}+a\Delta_{b}^{1/2}u|_{\partial M}=0\}

    with Δb\Delta_{b} positive Laplacian on M\partial M and a>1a>1 which is in fact (d1)(d-1)-dimensional.

  3. (iii)

    On the other hand, calculations of Section 4 imply that for a0a\neq 0 quantum periodicity will be broken (see [11], Section LABEL:book_new-sect-8-3) and one can get remainder estimate o(λd1)o(\lambda^{d-1}) albeit with the oscillating principal part.

  4. (iv)

    Under certain assumptions about geodesic billiard flow remainder (see [11], Section LABEL:book_new-sect-7-4) estimate (3.2) could be improved to O(h1d/|logh|)O(h^{1-d}/|\log h|) or even O(h1d+δ)O(h^{1-d+\delta}) with a small exponent δ>0\delta>0 and therefore remainder estimate (1.19) could be improved to O(λd1/logλ)O(\lambda^{d-1}/\log\lambda) or even O(λd1δ)O(\lambda^{d-1-\delta}).

  5. (v)

    Despite different expressions for a<0a<0 and a>0a>0 one can observe that κ(a)\kappa(a) is monotone increasing and continuous.

  6. (vi)

    In particular, κ(0)=14(2π)dωd1\kappa(0)=\frac{1}{4}(2\pi)^{-d}\omega_{d-1} and κ()=14(2π)dωd1\kappa(-\infty)=-\frac{1}{4}(2\pi)^{-d}\omega_{d-1}. Since a=0a=0 corresponds to the Neumann boundary condition and a=a=-\infty to Dirichlet (see Proposition 4.1), this agrees with (1.4).

  7. (vii)

    One can consider eigenvalues of operator ρλ1R(λ)\rho\lambda^{-1}R(\lambda) with ρ>0\rho>0 smooth on M\partial M; then estimates (3.1), (1.18) and asymptotics (3.2), (1.19) hold in the frameworks of Statement (i) and (ii) of Theorem 1.2 respectively albeit with κ(a)vol(M)\kappa(a)\operatorname{vol}^{\prime}(\partial M) replaced by

    Mκ(ρ(x)a)𝑑σ\int_{\partial M}\kappa(\rho(x^{\prime})a)\,d\sigma

    where dσd\sigma is a natural measure on M\partial M; however without this condition ρ>0\rho>0 problem may be much more challenging; even self-adjointness is by no means guaranteed.

  8. (viii)

    Operators of the form Pa,hP_{a,h} were considered by Frank and Geisinger [6]. They showed that the trace of the negative part of Pa,hP_{a,h} has a two-term expansion as h0h\to 0 regardless of dynamical assumptions111The fact that Frank and Geisinger obtain a second term regardless of dynamical assumptions is simply due to the fact that they study Trf(Pa,h)\operatorname{Tr}f(P_{a,h}) with f(λ)=λH(λ)f(\lambda)=-\lambda H(-\lambda) (HH is the Heaviside function), which is one order smoother than f(λ)=H(λ)f(\lambda)=H(-\lambda)., and the second term in their expansion (the Ld(2)L_{d}^{(2)} term of [6, Theorem 1.1]) is closely related to κ(a)\kappa(a) — see Remark 3.4.

Remark 1.4.

We can rephrase Theorem 1.2 in terms of a limiting measure on the unit circle. For each λ>0\lambda>0, let μ(h)\mu(h), h=λ1h=\lambda^{-1}, denote the atomic measure determined by the spectrum of C(λ)C(\lambda):

(1.24) μ(h)=(2πh)d1eiθjspecC(h1)δ(θθj),\mu(h)=(2\pi h)^{d-1}\sum_{e^{i\theta_{j}}\in\mathrm{spec}C(h^{-1})}\delta(\theta-\theta_{j}),

where we include each eigenvalue according to its multiplicity as usual. Then Theorem 1.2 can be expressed in the following way: the measures μ(h)\mu(h) converge in the weak-* topology as h0h\to 0 to the measure

(1.25) ωd1vol(M)ddθκ~(θ)dθon (0,2π), that is on S1{1}.\omega_{d-1}\operatorname{vol}^{\prime}(\partial M)\frac{d}{d\theta}\tilde{\kappa}(\theta)d\theta\quad\text{on }(0,2\pi),\text{ that is on }S^{1}\setminus\{1\}.

In particular, this measure is absolutely continuous, and finite away from eiθ=1e^{i\theta}=1 with an infinite accumulation of mass as θ2π\theta\uparrow 2\pi. In this form, we can compare our result with results on the semiclassical spectral asymptotics of scattering matrices. In [4] and [8], the scattering matrix Sh(E)S_{h}(E) at energy EE for the Schrödinger operator h2Δ+V(x)h^{2}\Delta+V(x) on d\mathbb{R}^{d} was studied in the semiclassical limit h0h\to 0. Assuming that VV is smooth and compactly supported, that EE is a nontrapping energy level, and that the set of periodic trajectories of the classical scattering transformation on TSd1T^{*}S^{d-1} has measure zero, it was shown that the measure μ(h)\mu(h) defined by (1.24) converged weak-* to a uniform measure on S1{1}S^{1}\setminus\{1\}, with an atom of infinite mass at the point 11. On the other hand, for polynomially decaying potentials, it was shown by Sobolev and Yafaev [12] in the case of central potentials and by Gell-Redman and the first author more generally (work in progress) that there is a limiting measure which is nonuniform, and is qualitatively similar to the measure for C(h1)C(h^{-1}) above in that it is finite away from 11, with an infinite accumulation of mass at 11 from one side.

Chapter 2 Reduction to semiclassical spectral asymptotics

In this section we prove Proposition 1.1. This result actually follows directly from the Birman-Schwinger principle. As some readers may not be familiar with this, we give the details.

Proof of Proposition 1.1.

We begin by recalling that the operator Pa,hP_{a,h} is the self-adjoint operator associated to the quadratic form (1.15), that is,

Qa,h(u):=h2uM2uM2hauM2.Q_{a,h}(u)\mathrel{\mathop{:}}=h^{2}\|\nabla u\|_{M}^{2}-\|u\|_{M}^{2}-ha\|u\|^{2}_{\partial M}.

We recall the min-max characterization of eigenvalues: the nnth eigenvalue μn(a,h)\mu_{n}(a,h) of Pa,hP_{a,h} is equal to the infimum of

supvV,v=1Qa,h(v)\sup_{v\in V,\|v\|=1}Q_{a,h}(v)

over all subspaces VH1(M)V\in H^{1}(M) of dimension nn. The monotonicity of Qa,hQ_{a,h} in aa, for fixed hh, shows that the eigenvalues are monotone nonincreasing with aa. In fact, they are strictly decreasing, which follows from the fact that eigenfunctions of Pa,hP_{a,h} cannot vanish at the boundary. Indeed, the eigenfunctions satisfy the boundary condition hνu=auh\partial_{\nu}u=-au, which shows that if uu vanishes at the boundary, so does νu\partial_{\nu}u, which is impossible.

The eigenvalues μn(a,h)\mu_{n}(a,h) are thus continuous, strictly decreasing functions of aa. Let a1<a2a_{1}<a_{2} be real numbers. The Birman-Schwinger principle [10, Prop. 9.2.7] says that the number of negative eigenvalues of Pa2,hP_{a_{2},h} is equal to the number of negative eigenvalues of Pa1,hP_{a_{1},h}, plus the number of eigenvalues μn(a,h)\mu_{n}(a,h) of Pa,hP_{a,h} that change from nonnegative to negative as aa varies from a1a_{1} to a2a_{2}. A diagram makes this clear: see Figure 1.

aaμ\mua=a1a=a_{1}a=a2a=a_{2}
Figure 1: Diagram showing the variation of eigenvalues μ(a,h)\mu(a,h) of Pa,hP_{a,h} as a function of aa for fixed hh. The eigenvalues are strictly decreasing in aa. Consequently, the number of negative eigenvalues of Pa2,hP_{a_{2},h} is equal to the number of negative eigenvalues of Pa1,hP_{a_{1},h} together with the number that cross the aa-axis between a=a1a=a_{1} and a=a2a=a_{2}.

The strict monotonicity of μ(a,h)\mu(a,h) in aa shows that the number of eigenvalues μn(a,h)\mu_{n}(a,h) of Pa,hP_{a,h} that change from nonnegative to negative as aa varies from a1a_{1} to a2a_{2} is the same as the number of μ(a,h)\mu(a,h) (counted with multiplicity) equal to zero, for a[a1,a2)a\in[a_{1},a_{2}). Next, we observe that the space of eigenfunctions un(a,h)u_{n}(a,h) of Pa,hP_{a,h} with zero eigenvalue, i.e. μn(a,h)=0\mu_{n}(a,h)=0 is in one-to-one correspondence with the space of eigenfunctions of C(λ)C(\lambda), λ=h1\lambda=h^{-1}, with eigenvalue (ai)(a+i)1(a-i)(a+i)^{-1}, or equivalently eiθe^{i\theta} where a=cot(θ/2)a=-\cot(\theta/2). Indeed, whenever unu_{n} is such an eigenfunction of Pa,hP_{a,h}, then

(2.1) f:=12(hνuiu)|Mf\mathrel{\mathop{:}}=\frac{1}{2}(h\partial_{\nu}u-iu)\big{|}_{\partial M}

is an eigenfunction of C(λ)C(\lambda), with eigenvalue (ai)(a+i)1(a-i)(a+i)^{-1}. Conversely, if ff is an eigenfunction of C(λ)C(\lambda) with eigenvalue (ai)(a+i)1(a-i)(a+i)^{-1}, then by definition there exists a Helmholtz function uu such that uu is related to ff according to (2.1), and we have (hν+a)u=0(h\partial_{\nu}+a)u=0 at M\partial M. This completes the proof. ∎

Remark 2.1.

We can apply similar arguments for λδρ1R(λ)\lambda^{-\delta}\rho^{-1}R(\lambda) as ρ>0\rho>0 is a smooth function on M\partial M and then plug corresponding parameters in the boundary conditions coming again to equality (1.16).

We next digress to prove that the eigenvalues of C(λ)C(\lambda) are monotonic (that is, they move monotonically around the unit circle) in λ\lambda. This plays no role in the remainder of our proof, but is (in the authors’ opinion) of independent interest.

Proposition 2.2.

The eigenvalues of C(λ)C(\lambda) rotate clockwise around the unit circle as λ\lambda increases.

Remark 2.3.

This implies that the eigenvalues of R𝗌𝖼𝗅(λ)R_{\mathsf{scl}}(\lambda) are monotone decreasing in λ\lambda.

Proof.

As discussed in the previous proof, C(λ)C(\lambda) has eigenvalue eiθe^{i\theta} if and only if Pa,hP_{a,h} has a zero eigenvalue, where a=a(θ)=cot(θ/2)a=a(\theta)=-\cot(\theta/2). Thus, as a function of h=λ1h=\lambda^{-1}, θ(h)\theta(h) is defined implicitly by the condition

μ(a(θ),h)=0.\mu(a(\theta),h)=0.

Since aa is a strictly increasing function of θ\theta, and we have just seen that μ\mu is a strictly decreasing function of aa, it suffices to show that when μ=0\mu=0, μ\mu is a strictly increasing function of hh, hence a strictly decreasing function of λ\lambda.

We now compute the derivative of μ\mu with respect to hh, at a value of aa and hh where μ(a,h)=0\mu(a,h)=0. We have

ddhμ(a,h)=ddh((h2Δ1)u(h),u(h))M=2h(Δu,u)M+((h2Δ1)u(h),u(h))M+((h2Δ1)u(h),u(h))M=2h(Δu,u)M+((h2Δ1)u,u)M(u,(h2Δ1)u)M.\frac{d}{dh}\mu(a,h)=\frac{d}{dh}((h^{2}\Delta-1)u(h),u(h))_{M}\\[3.0pt] =2h(\Delta u,u)_{M}+((h^{2}\Delta-1)u^{\prime}(h),u(h))_{M}+((h^{2}\Delta-1)u(h),u^{\prime}(h))_{M}\\[3.0pt] =2h(\Delta u,u)_{M}+((h^{2}\Delta-1)u^{\prime},u)_{M}-(u^{\prime},(h^{2}\Delta-1)u)_{M}.

In the third line, we used the fact that (h2Δ1)u=0(h^{2}\Delta-1)u=0 when μ(h)=0\mu(h)=0. Note the second term is not zero, as uu^{\prime} is not in the domain of the operator due to the changing boundary condition, so we cannot move the operator to the right hand side of the inner product without incurring boundary terms. We use the Gauss-Green formula to express the last two terms as a boundary integral:

μ(h)=2h(Δu,u)M+h(hνu,u)Mh(u,hνu)M\displaystyle\mu^{\prime}(h)=2h(\Delta u,u)_{M}+h(h\partial_{\nu}u^{\prime},u)_{\partial M}-h(u^{\prime},h\partial_{\nu}u)_{\partial M}
=2h(Δu,u)M+h(hνu,u)M+ha(u,u)M.\displaystyle\phantom{\mu^{\prime}(h}=2h(\Delta u,u)_{M}+h(h\partial_{\nu}u^{\prime},u)_{\partial M}+ha(u^{\prime},u)_{\partial M}.
Differentiating the boundary condition we find that
(hν+a)u=hνu at M.\displaystyle(h\partial_{\nu}+a)u^{\prime}=-h\partial_{\nu}u\text{ at }\partial M.
Substituting that in we get
μ(h)=2h(Δu,u)Mh(u,νu)M.\displaystyle\mu^{\prime}(h)=2h(\Delta u,u)_{M}-h(u,\partial_{\nu}u)_{\partial M}.
Applying Gauss-Green again, we get
μ(h)=h(Δu,u)M+huM2\displaystyle\mu^{\prime}(h)=h(\Delta u,u)_{M}+h\|\nabla u\|_{M}^{2}
=h1(uL2(M)2+huL2(M)2)>0.\displaystyle\phantom{\mu^{\prime}(h)+1+11}=h^{-1}\Big{(}\|u\|_{L^{2}(M)}^{2}+\|h\nabla u\|_{L^{2}(M)}^{2}\Big{)}>0.

Chapter 3 Semiclassical spectral asymptotics

In this section, we prove Theorem 1.2. Essentially, we have arrived at a rather standard semiclassical spectral asymptotics problem and results are due to [10], Chapter 5 or [11], Chapter LABEL:book_new-sect-7.

Proposition 3.1.
  1. (i)

    Let 𝖭h(a){\mathsf{N}}^{-}_{h}(a) be as in (1.17). The following asymptotic holds as h+0h\to+0:

    (3.1) 𝖭h(a)=(2πh)dωdvol(M)+O(h1d){\mathsf{N}}^{-}_{h}(a)=(2\pi h)^{-d}\omega_{d}\operatorname{vol}(M)+O(h^{1-d})
  2. (ii)

    Further, if the set of periodic billiards on MM has measure 0 then as h+0h\to+0:

    (3.2) 𝖭h(a)=(2πh)dωdvol(M)+h1dκ(a)vol(M)+o(h1d)\displaystyle{\mathsf{N}}^{-}_{h}(a)=(2\pi h)^{-d}\omega_{d}\operatorname{vol}(M)+h^{1-d}\kappa(a)\operatorname{vol}^{\prime}(\partial M)+o(h^{1-d})

    with κ(a)\kappa(a) given by (1.20).

Proof.

One can check easily that the operator Pa,hP_{a,h} is microhyperbolic at energy level 0 at each point (x,ξ)TM(x,\xi)\in T^{*}M in the direction ξ\xi; further, the boundary value problem is microhyperbolic at each point (x;ξ)TM(x^{\prime};\xi^{\prime})\in T^{*}\partial M at energy level 0 in the multidirection (ξ,ξ1,ξ1+)(\xi^{\prime},\xi_{1}^{-},\xi_{1}^{+}) with ξ1=ξ1±\xi_{1}=\xi_{1}^{\pm} roots of gjkξjξk=0\sum g^{jk}\xi_{j}\xi_{k}=0; finally, the boundary value problem is elliptic at each point of the elliptic zone (TM\subset T^{*}\partial M) if a0a\leq 0, and either elliptic or microhyperbolic in the direction ξ\xi^{\prime} at each point of the elliptic zone (TM\subset T^{*}\partial M) if a>0a>0 — see definitions in Chapters LABEL:book_new-sect-2, LABEL:book_new-sect-3 of [11]. Then statements (1.18), (1.19) follow from Theorems LABEL:book_new-thm-7-3-11 and LABEL:book_new-thm-7-4-1 of [11].

We now assume that the set of periodic billiards on MM has measure zero, and compute the second term in the spectral asymptotic explicitly. Similar calculations appear in [6].

To do this, one can use method of freezing coefficients (see f.e. [11], LABEL:book_new-sect-7-2) which results in

(3.3) h1dκ(a)=+(e(0,x1;0,x1;1)(2πh)dωd)𝑑x1h^{1-d}\kappa(a)=\int_{\mathbb{R}^{+}}\bigl{(}e(0,x_{1};0,x_{1};1)-(2\pi h)^{-d}\omega_{d}\bigr{)}\,dx_{1}

where e(x,x1;y,y1;τ)e(x^{\prime},x_{1};y^{\prime},y_{1};\tau) is the Schwartz kernel of the spectral projector E(τ)E(\tau) of the operator Ha=h2ΔH_{a}=h^{2}\Delta in half-space d1×+(x,x1)\mathbb{R}^{d-1}\times\mathbb{R}^{+}\ni(x^{\prime},x_{1}) with domain 𝔇(Ha)={uH2:(hx1+a)u|x1=0=0}\mathfrak{D}(H_{a})=\{u\in H^{2}:\ (h\partial_{x_{1}}+a)u|_{x_{1}=0}=0\}.

We obtain this spectral projector by integrating the spectral measure. This in turn is obtained via Stone’s formula

(3.4) dEL(σ)=12πi((L(σ+i0))1(L(σi0)1)dσ.dE_{L}(\sigma)=\frac{1}{2\pi i}\Big{(}(L-(\sigma+i0))^{-1}-(L-(\sigma-i0)^{-1}\Big{)}\,d\sigma.

Consider the resolvent for HaH_{a}, (Haσ)1(H_{a}-\sigma)^{-1}, for σ\sigma\in\mathbb{C}\setminus\mathbb{R}. Using the Fourier transform in the xx^{\prime} variables, we can write the Schwartz kernel of this resolvent in the form

(3.5) (2πh)1dei(xy)ξ(Ta+|ξ|2σ)1(x1,y1)𝑑ξ.(2\pi h)^{1-d}\int e^{i(x^{\prime}-y^{\prime})\cdot\xi^{\prime}}(T_{a}+|\xi^{\prime}|^{2}-\sigma)^{-1}(x_{1},y_{1})\,d\xi^{\prime}.

Here TaT_{a} is the one-dimensional operator Ta=h22+|ξ|2T_{a}=-h^{2}\partial^{2}+|\xi^{\prime}|^{2} on L2(+)L^{2}(\mathbb{R}_{+}) under the boundary condition (h+a)u|x1=0=0(h\partial+a)u|_{x_{1}=0}=0. The spectral projector EHa(1)E_{H_{a}}(1) is therefore given by

(3.6) (2πh)1d1ei(xy)ξ𝑑ETa(σ|ξ|2)(x1,y1)𝑑ξ𝑑σ.(2\pi h)^{1-d}\int_{-\infty}^{1}\int e^{i(x^{\prime}-y^{\prime})\cdot\xi^{\prime}}dE_{T_{a}}(\sigma-|\xi^{\prime}|^{2})(x_{1},y_{1})\,d\xi^{\prime}\,d\sigma.

Thus, we need to find the spectral measure for TaT_{a}. Write σ|ξ|2=η2\sigma-|\xi^{\prime}|^{2}=\eta^{2}, where we take η\eta to be in the first quadrant of \mathbb{C} for Imσ>0\operatorname{Im}\sigma>0, and in the fourth quadrant for Imσ<0\operatorname{Im}\sigma<0.

Lemma 3.2.

Suppose that Imη>0\operatorname{Im}\eta>0 and Reη0\operatorname{Re}\eta\geq 0. Then the resolvent kernel (Taη2)1(T_{a}-\eta^{2})^{-1} takes the form

(3.7) (Taη2)(x,y)={i2hη(eiη(xy)/h+iηaiη+aeiη(x+y)/h),x>yi2hη(eiη(yx)/h+iηaiη+aeiη(x+y)/h),x<y.(T_{a}-\eta^{2})(x,y)=\begin{cases}\frac{i}{2h\eta}\Big{(}e^{i\eta(x-y)/h}+\frac{i\eta-a}{i\eta+a}e^{i\eta(x+y)/h}\Big{)},\quad x>y\\[3.0pt] \frac{i}{2h\eta}\Big{(}e^{i\eta(y-x)/h}+\frac{i\eta-a}{i\eta+a}e^{i\eta(x+y)/h}\Big{)},\quad x<y.\end{cases}

If Imη<0\operatorname{Im}\eta<0 and Reη0\operatorname{Re}\eta\geq 0, then the resolvent kernel (Taη2)1(T_{a}-\eta^{2})^{-1} takes the form

(3.8) (Taη2)(x,y)={i2hη(eiη(yx)/h+iη+aiηaeiη(x+y)/h),x>yi2hη(eiη(xy)/h+iη+aiηaeiη(x+y)/h),x<y.(T_{a}-\eta^{2})(x,y)=\begin{cases}-\frac{i}{2h\eta}\Big{(}e^{i\eta(y-x)/h}+\frac{i\eta+a}{i\eta-a}e^{-i\eta(x+y)/h}\Big{)},\quad x>y\\[3.0pt] -\frac{i}{2h\eta}\Big{(}e^{i\eta(x-y)/h}+\frac{i\eta+a}{i\eta-a}e^{-i\eta(x+y)/h}\Big{)},\quad x<y.\end{cases}
Proof.

In the regions x<yx<y and x>yx>y, the resolvent kernel must be a linear combination of eiηx/he^{i\eta x/h} and eiηx/he^{-i\eta x/h}. Moreover, for Imη>0\operatorname{Im}\eta>0, we can only have the e+iηx/he^{+i\eta x/h} term, as xx\to\infty, as the other would be exponentially increasing. So we can write the kernel in the form

(3.9) {c1e+iηx/h,x>yc2e+iηx/h+c3eiηx/h,x<y.\left\{\begin{aligned} &c_{1}e^{+i\eta x/h},&&x>y\\ &c_{2}e^{+i\eta x/h}+c_{3}e^{-i\eta x/h},&&x<y.\end{aligned}\right.

We apply the boundary condition, and the two connection conditions at x=yx=y, namely continuity, and a jump in the derivative of 1/h-1/h, in order to obtain the delta function δ(xy)\delta(x-y) after applying TaT_{a}. These three conditions determine the cic_{i} uniquely, and we find that, in the case Imη>0\operatorname{Im}\eta>0,

(3.9)1(\ref*{eq-4-10})_{1} c1=i2hη(eiηy/h+iηaiη+ae+iηy/h),\displaystyle c_{1}=\frac{i}{2h\eta}\big{(}e^{-i\eta y/h}+\frac{i\eta-a}{i\eta+a}e^{+i\eta y/h}\big{)},
(3.9)2,3(\ref*{eq-4-10})_{2,3} c2=i2hηiηaiη+ae+iηy/h,c3=i2hηe+iηy/h,\displaystyle c_{2}=\frac{i}{2h\eta}\frac{i\eta-a}{i\eta+a}e^{+i\eta y/h},\qquad c_{3}=\frac{i}{2h\eta}e^{+i\eta y/h},

yielding (3.7). A similar calculation yields (3.8). ∎

We now apply (3.4) to find the Schwartz kernel of the spectral measure for TaT_{a}.

Lemma 3.3.

The spectral measure dETa(τ)dE_{T_{a}}(\tau) is given by the following.

  1. (i)

    For τ0\tau\geq 0, τ=η2\tau=\eta^{2}

    (3.11) dETa(τ)=14πhη(eiη(xy)/h+eiη(yx)/h+iηaiη+aeiη(x+y)/h+iη+aiηaeiη(x+y)/h) 2ηdη.dE_{T_{a}}(\tau)\\ =\frac{1}{4\pi h\eta}\Big{(}e^{i\eta(x-y)/h}+e^{i\eta(y-x)/h}+\frac{i\eta-a}{i\eta+a}e^{i\eta(x+y)/h}+\frac{i\eta+a}{i\eta-a}e^{-i\eta(x+y)/h}\Big{)}\,2\eta d\eta.
  2. (ii)

    For τ<0\tau<0, the spectral measure dE(τ)dE(\tau) vanishes for a0a\leq 0, while for a>0a>0

    (3.12) dETa(τ)=2aheax/heay/hδ(τ+a2)dτ.dE_{T_{a}}(\tau)=\frac{2a}{h}e^{-ax/h}e^{-ay/h}\delta(\tau+a^{2})d\tau.
Proof.

This follows directly from Lemma 3.2 and Stone’s formula, (3.4). The extra term for a>0a>0 arises from the pole in the denominator, iη+ai\eta+a for Imη>0\operatorname{Im}\eta>0 and iηai\eta-a for Imη<0\operatorname{Im}\eta<0 in the expressions (3.7), (3.8), which only occurs for a>0a>0. For τ\tau negative, we need to set η=iτ+0\eta=i\sqrt{-\tau}+0 in (3.7) and η=iτ+0\eta=-i\sqrt{-\tau}+0 in (3.8), and subtract. Then everything cancels except at the pole, where we obtain a delta function 2πiδ(τa)-2\pi i\delta(\sqrt{-\tau}-a), which arises from (τ+i0+a)1(τi0+a)1(\sqrt{-\tau}+i0+a)^{-1}-(\sqrt{-\tau}-i0+a)^{-1}. This term arises from the negative eigenvalue a2-a^{2} which occurs for a>0a>0, corresponding to the eigenfunction 2a/heax/h\sqrt{2a/h}\,e^{-ax/h}. ∎

Plugging this into (3.6), and making use of the fact that dσdξ=2ηdηdξd\sigma d\xi^{\prime}=2\eta d\eta d\xi^{\prime}, we find that the Schwartz kernel of EHa(1)E_{H_{a}}(1) is given by

(3.13) (2πh)d01H(1|ξ|2η2)ei(xy)ξ×(eiη(x1y1)/h+eiη(y1x1)/h+iηaiη+aeiη(x1+y1)/h+iη+aiηaeiη(x1+y1)/h)dξdη(2\pi h)^{-d}\int_{0}^{1}\int H(1-|\xi^{\prime}|^{2}-\eta^{2})e^{i(x^{\prime}-y^{\prime})\cdot\xi^{\prime}}\\ \times\Big{(}e^{i\eta(x_{1}-y_{1})/h}+e^{i\eta(y_{1}-x_{1})/h}+\frac{i\eta-a}{i\eta+a}e^{i\eta(x_{1}+y_{1})/h}+\frac{i\eta+a}{i\eta-a}e^{-i\eta(x_{1}+y_{1})/h}\Big{)}\,d\xi^{\prime}\,d\eta

for a0a\leq 0 while for a>0a>0, it is given by the sum of (3.13) and

(3.14) (2πh)1d2aheax1/heay1/h1ei(xy)ξδ(σ|ξ|2+a2)𝑑ξ𝑑σ.(2\pi h)^{1-d}\frac{2a}{h}e^{-ax_{1}/h}e^{-ay_{1}/h}\int_{-\infty}^{1}\int e^{i(x^{\prime}-y^{\prime})\cdot\xi^{\prime}}\delta(\sigma-|\xi^{\prime}|^{2}+a^{2})\,d\xi^{\prime}\,d\sigma.

We are actually interested in the value on the diagonal. Setting x=yx=y, and performing the trivial ξ\xi^{\prime} integral, we find that the Schwartz kernel of the spectral projector EHa(1)(x,x)E_{H_{a}}(1)(x,x) on the diagonal is given by

(3.15) ωd1(2πh)d01(1η2)(d1)/2(2+iηaiη+ae2iηx1/h+iη+aiηae2iηx1/h)𝑑η+H(a)(d1)ωd1(2πh)d1ahe2ax1/ha21(σ+a2)(d3)/2𝑑σ.\frac{\omega_{d-1}}{(2\pi h)^{d}}\int_{0}^{1}(1-\eta^{2})^{(d-1)/2}\Bigg{(}2+\frac{i\eta-a}{i\eta+a}e^{2i\eta x_{1}/h}+\frac{i\eta+a}{i\eta-a}e^{-2i\eta x_{1}/h}\Bigg{)}\,d\eta\\ +H(a)\frac{(d-1)\omega_{d-1}}{(2\pi h)^{d-1}}\frac{a}{h}e^{-2ax_{1}/h}\int_{-a^{2}}^{1}(\sigma+a^{2})^{(d-3)/2}\,d\sigma.

Since

ωd1012(1η2)(d1)/2𝑑η=ωd,\omega_{d-1}\int_{0}^{1}2(1-\eta^{2})^{(d-1)/2}\,d\eta=\omega_{d},

we see by comparing with (3.3) that this term disappears in the expression for κ(a)\kappa(a) and we have, after performing the x1x_{1} integral as in (3.3)

(3.16) h1dκ(a)=ωd1(2πh)d01(1η2)(d1)/2×(iηaiη+a(ih2(η+i0)1)iη+aiηa(ih2(ηi0)1))dη+H(a)(d1)ωd12(2πh)d1a21(σ+a2)(d3)/2𝑑σ.h^{1-d}\kappa(a)=\frac{\omega_{d-1}}{(2\pi h)^{d}}\int_{0}^{1}(1-\eta^{2})^{(d-1)/2}\\ \times\Bigg{(}\frac{i\eta-a}{i\eta+a}\Big{(}\frac{ih}{2}(\eta+i0)^{-1}\Big{)}\ -\frac{i\eta+a}{i\eta-a}\Big{(}\frac{ih}{2}(\eta-i0)^{-1}\Big{)}\Bigg{)}\,d\eta\\ +H(a)\frac{(d-1)\omega_{d-1}}{2(2\pi h)^{d-1}}\int_{-a^{2}}^{1}(\sigma+a^{2})^{(d-3)/2}\,d\sigma.

Simplifying a bit, and performing the σ\sigma integral, we have

(3.17) κ(a)=iωd12(2π)d11(1η2)(d1)/2(iηa)2a2+η2(η+i0)1𝑑η+H(a)ωd1(2π)d1(1+a2)(d1)/2.\kappa(a)=-\frac{i\omega_{d-1}}{2(2\pi)^{d}}\int_{-1}^{1}(1-\eta^{2})^{(d-1)/2}\frac{(i\eta-a)^{2}}{a^{2}+\eta^{2}}(\eta+i0)^{-1}\,d\eta\\ +H(a)\frac{\omega_{d-1}}{(2\pi)^{d-1}}(1+a^{2})^{(d-1)/2}.

We further simplify this expression by expanding (iηa)2=a22iaηη2(i\eta-a)^{2}=a^{2}-2ia\eta-\eta^{2}, and noting that the contribution of the η2-\eta^{2} term is zero, as this gives an odd integrand in the η\eta integral. A similar statement can be made for the a2a^{2} term, except that there is a contribution from the pole in this case. This leads to the expression

(3.18) κ(a)=ωd1(2π)d1(12π11(1η2)(d1)/2aa2+η2dη14+H(a)(1+a2)(d1)/2).\kappa(a)=\frac{\omega_{d-1}}{(2\pi)^{d-1}}\bigg{(}-\frac{1}{2\pi}\int_{-1}^{1}(1-\eta^{2})^{(d-1)/2}\frac{a}{a^{2}+\eta^{2}}\,d\eta\\ -\frac{1}{4}+H(a)(1+a^{2})^{(d-1)/2}\bigg{)}.

Although not immediately apparent, this formula is continuous at a=0a=0. In fact, the function a(a2+η2)1a(a^{2}+\eta^{2})^{-1} has a distributional limit (sgna)πδ(η)(\operatorname{sgn}a)\pi\delta(\eta) as aa tends to zero from above or below. The change of sign as aa crosses 0 means that the integral in (3.18) has a jump of 1-1 as aa crosses zero from negative to positive. That exactly compensates the jump in the final term.

In odd dimensions, we can compute this integral exactly. In particular, in dimension d=3d=3, we find that

(3.19) κ(a)=ω2(2π)2(14+aπ+(1+a2)(1arccotaπ)).\kappa(a)=\frac{\omega_{2}}{(2\pi)^{2}}\Big{(}-\frac{1}{4}+\frac{a}{\pi}+(1+a^{2})\big{(}1-\frac{\operatorname{arccot}a}{\pi}\big{)}\Big{)}.

Proof of Theorem 1.2.

This follows immediately from Proposition 3.1 and Proposition 1.1. ∎

Remark 3.4.

The second term of the expansion in [6, Theorem 1.1] is obtained by computing

(3.20) (2πh)1d1(1σ)ei(xy)ξ𝑑ETa(σ|ξ|2)(x1,y1)𝑑ξ𝑑σ(2\pi h)^{1-d}\int_{-\infty}^{1}(1-\sigma)\int e^{i(x^{\prime}-y^{\prime})\cdot\xi^{\prime}}dE_{T_{a}}(\sigma-|\xi^{\prime}|^{2})(x_{1},y_{1})\,d\xi^{\prime}\,d\sigma

instead of (3.6).

Chapter 4 Relation to Dirichlet boundary condition

In this section we observe that the limit aa\to-\infty corresponds to the Dirichlet boundary condition. More precisely, we have

Proposition 4.1.

Let 𝖭h(){\mathsf{N}}_{h}^{-}(-\infty) denote the limit

𝖭h():=lima𝖭h(a),{\mathsf{N}}_{h}^{-}(-\infty)\mathrel{\mathop{:}}=\lim_{a\to-\infty}{\mathsf{N}}_{h}^{-}(a),

where 𝖭h(a){\mathsf{N}}_{h}^{-}(a) is given by (1.17). Then we have

(4.1) 𝖭h()=#{λjh1λj2 is a Dirichlet eigenvalue of Δ}.{\mathsf{N}}_{h}^{-}(-\infty)=\#\{\lambda_{j}\leq h^{-1}\mid\lambda^{2}_{j}\text{ is a Dirichlet eigenvalue of }\Delta\}.
Remark 4.2.

Because the quadratic form (1.15) is monotone in aa, the counting function 𝖭h(a){\mathsf{N}}_{h}^{-}(a) is monotone in aa. Hence the limit above exists.

Proof.

We use the min-max characterisation of eigenvalues. Let 𝖭~D(λ)\tilde{{\mathsf{N}}}_{D}(\lambda) denote the number of Dirichlet eigenvalues (counted with multiplicity) less than or equal to λ=h1\lambda=h^{-1}. This is equal to the maximal dimension of a subspace of H01(M)H^{1}_{0}(M) on which the quadratic form QDQ_{D}, given by

(4.2) QD(u,u)=h2u22u22Q_{D}(u,u)=h^{2}\|\nabla u\|_{2}^{2}-\|u\|_{2}^{2}

is negative semidefinite. On the other hand, 𝖭~h(a)\tilde{{\mathsf{N}}}_{h}^{-}(a) is equal to the maximal dimension of a subspace of H1(M)H^{1}(M) on which the quadratic form QaQ_{a} given by (1.15) is (strictly) negative definite.

We first show that 𝖭~D(h1)𝖭~h()\tilde{{\mathsf{N}}}_{D}(h^{-1})\leq\tilde{{\mathsf{N}}}_{h}^{-}(-\infty). Let VV be the vector space spanned by Dirichlet eigenfunctions with eigenvalue λ2\leq\lambda^{2}. Clearly, the quadratic form QaQ_{a} is negative semidefinite on VV, and if λ2\lambda^{2} is not a Dirichlet eigenvalue, then it is negative definite, proving the assertion. In the case that λ2\lambda^{2} is a Dirichlet eigenvalue, we perturb VV to VϵV_{\epsilon}, a vector space of H1(M)H^{1}(M) of the same dimension as VV, so that, for for ϵ\epsilon sufficiently small depending on aa, QaQ_{a} is negative definite on VϵV_{\epsilon}. For simplicity we only do this in the case that the λ2\lambda^{2}-eigenspace is one dimensional, leaving the general case to the reader. To do this, we choose an orthonormal basis of VV (with respect to the L2L^{2} inner product) of Dirichlet eigenfunctions v1,,vkv_{1},\dots,v_{k} with eigenvalues λ12λk2\lambda_{1}^{2}\dots\lambda_{k}^{2}, where λk=λ\lambda_{k}=\lambda. Then we perturb only vkv_{k}, leaving the others fixed. We choose sH01(M)s\in H^{1}_{0}(M)^{\perp}, the orthogonal complement of H01(M)H^{1}_{0}(M) in H1(M)H^{1}(M) (with respect to the inner product in H1(M)H^{1}(M)), so that

(4.3) Qa(vi,s)=0,i<k and Qa(vk,s)>0.Q_{a}(v_{i},s)=0,\quad i<k\text{ and }Q_{a}(v_{k},s)>0.

We check that this is possible. Notice that sH01(M)s\in H^{1}_{0}(M)^{\perp} implies that (Δ+1)s=0(\Delta+1)s=0 in MM. Then as viv_{i} has zero boundary data, we have

(4.4) (λi2+1)(vi,s)M=(Δvi,s)M(vi,Δs)M=νvi,sM.(\lambda_{i}^{2}+1)(v_{i},s)_{M}=(\Delta v_{i},s)_{M}-(v_{i},\Delta s)_{M}\\ =\langle\partial_{\nu}v_{i},s\rangle_{\partial M}.

We choose ss so that νvi,sM\langle\partial_{\nu}v_{i},s\rangle_{\partial M} vanishes for i<ki<k and is positive for i=ki=k. This is possible: in fact, due to the unique solvability of the boundary value problem

(4.5) (Δ+1)s=0,s|M=fH1/2(M),(\Delta+1)s=0,\quad s|_{\partial M}=f\in H^{1/2}(M),

for sH1(M)s\in H^{1}(M), we see that ss can have any boundary value in H1/2(M)H^{1/2}(\partial M) which is dense in L2(M)L^{2}(\partial M). Then using (4.4) we see that νvk,sM>0\langle\partial_{\nu}v_{k},s\rangle_{\partial M}>0 implies that (vk,s)M>0(v_{k},s)_{M}>0.

We now define VϵV_{\epsilon} to be the span of v1,,vk1v_{1},\dots,v_{k-1} and vk+ϵsv_{k}+\epsilon s. Then we have

(4.6) Qa(vi,vk+ϵs)=0,i<kQ_{a}(v_{i},v_{k}+\epsilon s)=0,\ i<k

and

(4.7) Qa(vk+ϵs,vk+ϵs)\displaystyle Q_{a}(v_{k}+\epsilon s,v_{k}+\epsilon s) =Qa(vk,vk)+2ϵQa(vk,s)+ϵ2Qa(s,s)\displaystyle=Q_{a}(v_{k},v_{k})+2\epsilon Q_{a}(v_{k},s)+\epsilon^{2}Q_{a}(s,s)
=2ϵQa(vi,si)+ϵ2Qa(si,si)\displaystyle=2\epsilon Q_{a}(v_{i},s_{i})+\epsilon^{2}Q_{a}(s_{i},s_{i})
=2ϵ(h2+1)(vi,si)M+O(ϵ2a2)\displaystyle=-2\epsilon(h^{2}+1)(v_{i},s_{i})_{M}+O(\epsilon^{2}a^{2})

which is strictly negative for ϵa2\epsilon a^{2} small enough. It follows that QaQ_{a} is negative definite on VkV_{k} when ϵa2\epsilon a^{2} is small enough. A similar construction can be made when λ2\lambda^{2} has multiplicity greater than 11.

We next show that 𝖭~D(h1)𝖭h()\tilde{{\mathsf{N}}}_{D}(h^{-1})\geq{\mathsf{N}}_{h}^{-}(-\infty). We argue by contradiction: if not, then for any aa, there is a vector space WW of dimension k+1\geq k+1 on which QaQ_{a} is negative definite. Then there is a nonzero vector wWw\in W orthogonal (in the H1(M)H^{1}(M) inner product) to VV. We can write w=w+sw=w^{\prime}+s where wH01(M)w^{\prime}\in H^{1}_{0}(M) and sH01(M)s\in H^{1}_{0}(M)^{\perp}. Then ww^{\prime} is a linear combination of Dirichlet eigenfunctions with eigenvalue λ>λ\geq\lambda^{\prime}>\lambda, where λ\lambda^{\prime} is the smallest eigenvalue larger than λ\lambda. We then have

(4.8) 0>Qa(w+s,w+s)=Qa(w,w)+2Qa(w,s)+Qa(s,s)(λλ)w222(h2+1)w2sL2(M)hasL2(M)2.0>Q_{a}(w^{\prime}+s,w^{\prime}+s)=Q_{a}(w^{\prime},w^{\prime})+2Q_{a}(w^{\prime},s)+Q_{a}(s,s)\\ \geq(\lambda^{\prime}-\lambda)\|w^{\prime}\|_{2}^{2}-2(h^{2}+1)\|w^{\prime}\|_{2}\|s\|_{L^{2}(M)}-ha\|s\|_{L^{2}(\partial M)}^{2}.

However, some standard potential theory shows that sL2(M)\|s\|_{L^{2}(M)} is bounded by a constant times sL2(M)\|s\|_{L^{2}(\partial M)}. To see this, extend MM to a larger manifold M~\tilde{M} of the same dimension, and let G(x,y)G(x,y) be the Schwartz kernel of the inverse of (ΔM~+1)1(\Delta_{\tilde{M}}+1)^{-1} on L2(M~)L^{2}(\tilde{M}), with Dirichlet boundary conditions at M~\partial\tilde{M}. We can write ss as MdνyG(x,y)h(y)𝑑y\int_{\partial M}d_{\nu_{y}}G(x,y)h(y)\,dy where (1/2+D)h=s|M(1/2+D)h=s|_{\partial M} and DD is the double layer operator on M\partial M determined by GG. Standard arguments show that (1/2+D)(1/2+D) has a bounded inverse on L2(M)L^{2}(\partial M) and dνyG(x,y)d_{\nu_{y}}G(x,y) is a bounded integral operator from L2(M)L^{2}(\partial M) to L2(M)L^{2}(M). So we can write, for a<0a<0,

0>Qa(w+s,w+s)(λλ)w222C(h2+1)w2sL2(M)+h|a|sL2(M)220>Q_{a}(w^{\prime}+s,w^{\prime}+s)\\ \geq(\lambda^{\prime}-\lambda)\|w^{\prime}\|_{2}^{2}-2C(h^{2}+1)\|w^{\prime}\|_{2}\|s\|_{L^{2}(\partial M)}+h|a|\|s\|_{L^{2}(\partial M)^{2}}^{2}

and the RHS is clearly positive for |a||a| large enough, giving us the desired contradiction.

References

  • [1] A. Barnett and A. Hassell, Boundary quasi-orthogonality and sharp inclusion bounds for large Dirichlet eigenvalues. SIAM J. Numer. Anal. 49 (2011), no. 3, 1046–1063.
  • [2] D. Burghelea, L. Friedlander, T. Kappeler, Meyer-Vietoris type formula for determinants of elliptic differential operators. J. Funct. Anal. 107 (1992), no. 1, 34–65.
  • [3] A. P. Calderón, On an inverse boundary value problem. Seminar on Numerical Analysis and its Applications to Continuum Physics (Rio de Janeiro, 1980), pp. 65–73, Soc. Brasil. Mat., Rio de Janeiro, 1980.
  • [4] K. Datchev, J. Gell-Redman, A. Hassell, P. Humphries,, Approximation and equidistribution of phase shifts: spherical symmetry. Comm. Math. Phys. 326 (2014), no. 1, 209–236.
  • [5] J. J. Duistermaat, V. W. Guillemin, The spectrum of positive elliptic operators and periodic bicharacteristics. Invent. Math. 29 (1975), no. 1, 39–79.
  • [6] R. L. Frank, and L. Geisinger, Semi-classical analysis of the Laplace operator with Robin boundary conditions. Bull. Math. Sci. 2 (2012), no. 2, 281–319.
  • [7] L. Friedlander, Some inequalities between Dirichlet and Neumann eigenvalues. Arch. Rational Mech. Anal. 116 (1991), no. 2, 153–160.
  • [8] J. Gell-Redman, A. Hassell, S. Zelditch, Equidistribution of phase shifts in semiclassical potential scattering. J. Lond. Math. Soc. (2) 91 (2015), no. 1, 159–179.
  • Ivr [80] V. Ivrii, The second term of the spectral asymptotics for a Laplace-Beltrami operator on manifolds with boundary. (Russian) Funktsional. Anal. i Prilozhen. 14 (1980), no. 2, 25–34. English translation in Functional Analysis and Its Applications, 14:2 (1980), 98–106.
  • Ivr [1] V. Ivrii. Microlocal Analysis and Precise Spectral Asymptotics, Springer-Verlag, SMM, 1998, xv+731.
  • Ivr [2] V. Ivrii. Microlocal Analysis and Sharp Spectral Asymptotics, in progress: available online at
    http://www.math.toronto.edu/ivrii/futurebook.pdf
  • [12] A. V. Sobolev and D. R. Yafaev, Phase analysis in the problem of scattering by a radial potential. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 147:155–178, 206, 1985.
Andrew Hassell, Victor Ivrii
Mathematical Sciences Institute, Department of Mathematics
Australian National University, University of Toronto,
40, St.George Str.,
Canberra, ACT 0200 Toronto, Ontario M5S 2E4
Australia Canada
Andrew.Hassell@anu.edu.au ivrii@math.toronto.edu