Skip to main content
Springer logoLink to Springer
. 2021 Nov 2;388(3):1603–1635. doi: 10.1007/s00220-021-04248-y

Differences Between Robin and Neumann Eigenvalues

Zeév Rudnick 1, Igor Wigman 2,, Nadav Yesha 3
PMCID: PMC8599415  PMID: 34840338

Abstract

Let ΩR2 be a bounded planar domain, with piecewise smooth boundary Ω. For σ>0, we consider the Robin boundary value problem

-Δf=λf,fn+σf=0onΩ

where fn is the derivative in the direction of the outward pointing normal to Ω. Let 0<λ0σλ1σ be the corresponding eigenvalues. The purpose of this paper is to study the Robin–Neumann gaps

dn(σ):=λnσ-λn0.

For a wide class of planar domains we show that there is a limiting mean value, equal to 2length(Ω)/area(Ω)·σ and in the smooth case, give an upper bound of dn(σ)C(Ω)n1/3σ and a uniform lower bound. For ergodic billiards we show that along a density-one subsequence, the gaps converge to the mean value. We obtain further properties for rectangles, where we have a uniform upper bound, and for disks, where we improve the general upper bound.

Statement of Results

Let ΩR2 be a bounded planar domain, with piecewise smooth boundary Ω. For σ0, we consider the Robin boundary value problem

-Δf=λfonΩ,fn+σf=0onΩ

where fn is the derivative in the direction of the outward pointing normal to Ω. The case σ=0 is the Neumann boundary condition, and we use σ= as a shorthand for the Dirichlet boundary condition f|Ω=0.

Robin boundary conditions are used in heat conductance theory to interpolate between a perfectly insulating boundary, described by Neumann boundary conditions σ=0, and a temperature fixing boundary, described by Dirichlet boundary conditions corresponding to σ=+. To date, most studies concentrated on the first few Robin eigenvalues, with applications in shape optimization and related isoperimetric inequalities and asymptotics of the first eigenvalues (see [5]). Our goal is very different, aiming to study the difference between high-lying Robin and Neumann eigenvalues. There are very few studies addressing this in the literature, except for [4, 32] which aim at different goals.

We will take the Robin condition for a fixed and positive σ>0, when all eigenvalues are positive, one excuse being that a negative Robin parameter gives non-physical boundary conditions for the heat equation, with heat flowing from cold to hot; see however [17] for a model where negative σ is of interest, in particular σ- [10, 15, 21, 22]. Let 0<λ0σλ1σ be the corresponding eigenvalues. The Robin spectrum always lies between the Neumann and Dirichlet spectra (Dirichlet–Neumann bracketing) [5] :

λn0<λnσ<λn. 1.1

We define the Robin–Neumann difference (RN gaps) as

dn(σ):=λnσ-λn0

and study several of their properties. See Sect. 2 for some numerical experiments. This seems to be a novel subject, and the only related study that we are aware of is the very recent work of Rivière and Royer [28], which addresses the RN gaps for quantum star graphs.

The mean value

The first result concerns the mean value of the gaps:

Theorem 1.1

Let ΩR2 be a bounded, piecewise smooth domain. Then the mean value of the RN gaps exists, and equals

limN1Nn=1Ndn(σ)=2length(Ω)area(Ω)·σ.

Since the differences dn(σ)>0 are positive, we deduce by Chebyshev’s inequality:

Corollary 1.2

Let Ω be a bounded, piecewise smooth domain. Fix σ>0. Let Φ(n) be a function tending to infinity (arbitrarily slowly). Then for almost all n’s, dn(σ)Φ(n) in the sense that

#{nN:dn(σ)>Φ(n)}NΦ(N).

A lower bound

Recall that a domain Ω is “star-shaped with respect to a point xΩ" if the segment between x and every other point of Ω lies inside the domain; so convex means star-shaped with respect to any point; “star-shaped" just means that there is some x so that it is star-shaped with respect to x.

Theorem 1.3

Let ΩR2 be a bounded star-shaped planar domain with smooth boundary. Then the Robin–Neumann differences are uniformly bounded below: For all σ>0, C=C(Ω,σ)>0 so that

dn(σ)C.

Note that for quantum star graphs, this lower bound fails [28].

A general upper bound

We give a quantitative upper bound:

Theorem 1.4

Assume that Ω has a smooth boundary. Then C=C(Ω)>0 so that for all σ>0,

dn(σ)C(λn)1/3σ.

While quite poor, it is the best individual bound that we have in general. Below, we will indicate how to improve it in special cases.

Question 1.5

Are there planar domains where the differences dn(σ) are unbounded?

We believe that this happens in several cases, e.g. the disk, but at present can only show this for the hemisphere [30], which is not a planar domain.

Ergodic billiards

To a piecewise smooth planar domain one associates a billiard dynamics. When this dynamics is ergodic, as for the stadium billiard (see Fig. 2), we can improve on Corollary 1.2:

Fig. 2.

Fig. 2

The first 200 RN gaps for the ergodic quarter-stadium billiard (A), a quarter of the shape formed by gluing two half-disks to a square of sidelength 2, and for the uniformly hyperbolic billiard consisting of a quarter of the shape formed by the intersection of the exteriors of four disks (B)

Theorem 1.6

Let ΩR2 be a bounded, piecewise smooth domain. Assume that the billiard dynamics associated to Ω is ergodic. Then for every σ>0, there is a sub-sequence N=NσN of density one so that along that subsequence,

limnnNdn(σ)=2length(Ω)area(Ω)·σ.

If the billiard dynamics is uniformly hyperbolic, we expect that more is true, that all the gaps converge to the mean.

A key ingredient in the proofs of the above results is that they can be connected to L2 restriction estimates for eigenfunctions on the boundary via a variational formula for the gaps (Lemma 3.1)

dn(σ)=0σΩ|un,τ|2dsdτ

where un,τ is any L2(Ω)-normalized eigenfunction associated with λnτ.

Generalizations

Most of the above results easily extend to higher dimensions: The upper bound (Theorem 1.4), the mean value result (Theorem 1.1), and the almost sure convergence for ergodic billiards (Theorem 1.6). At this stage our proof of the lower bound (Theorem 1.3) is restricted to dimension 2.

In Sect. 7 we discuss extensions of the above results to the case of variable boundary conditions σ:ΩR.

Rectangles

For the special case of rectangles, we show that the RN gaps are bounded:

Theorem 1.7

Let Ω be a rectangle. Then for every σ>0 there is some CΩ(σ)>0 so that for all n,

dn(σ)CΩ(σ).

We use Theorem 1.7 to draw a consequence for the level spacing distribution of the Robin eigenvalues on a rectangle: Let x0x1x2 be a sequence of levels, and δn=xn+1-xn be the nearest neighbour gaps. We assume that xN=N+o(N) so that the average gap is unity. The level spacing distribution P(s) of the sequence is then defined as

0yP(s)ds=limN1N#{nN:δny}

(assuming that the limit exists).

It is well known that the level spacing distribution for the Neumann (or Dirichlet) eigenvalues on the square is a delta-function at the origin, due to large arithmetic multiplicities in the spectrum. Once we put a Robin boundary condition, we can show [31] that the multiplicities disappear for σ>0 sufficiently small, except for systematic doubling due to symmetry. Nonetheless, even after desymmetrizing (removing the systematic multiplicities) we show that the level spacing does not change:

Theorem 1.8

The level spacing distribution for the desymmetrized Robin spectrum on the square is a delta-function at the origin.

The disk

As we will explain, upper bounds for the gaps dn can be obtained from upper bounds for the remainder term in Weyl’s law for the Robin/Neumann problem. While this method will usually fall short of Theorem 1.4, for the disk it gives a better bound. In that case, Kuznetsov and Fedosov [16] (see also Colin de Verdiére [8]) gave an improved remainder term in Weyl’s law for Dirichlet boundary conditions, by relating the problem to counting (shifted) lattice points in a certain cusped domain. With some work, the argument can also be adapted to the Robin case (see Sect. 10.2 and “Appendix A”), which recovers Theorem 1.4 in this special case. The remainder term for the lattice count was improved by Guo, Wang and Wang [13], from which we obtain:

Theorem 1.9

For the unit disk, for any fixed σ>0, we have

dn(σ)=O(n1/3-δ),δ=1/990.

Numerics

We present some numerical experiments on the fluctuation of the RN gaps. In all cases, we took the Robin constant to be σ=1. Displayed are the run sequence plots of the RN gaps. The solid (green) curve is the cumulative mean. The solid (red) horizontal line is the limiting mean value 2length(Ω)/area(Ω) obtained in Theorem 1.1.

In Fig. 1 we present numerics for two domains where the Neumann and Dirichlet problems are solvable, by means of separation of variables, the square and the disk. These were generated using Mathematica [35]. For the square, we are reduced to finding Robin eigenvalues on an interval as (numerical) solutions to a secular equation, see Sect. 8, and have used Mathematica to find these.

Fig. 1.

Fig. 1

The first 2000 RN gaps for the unit square (A) and for the unit disk (B)

The disk admits separation of variables, and as is well known the Dirichlet eigenvalues on the unit disk are the squares of the positive zeros of the Bessel functions Jn(x). The positive Neumann eigenvalues are squares of the positive zeros of the derivatives Jn(x), and the Robin eigenvalues are the squares of the positive zeros of xJn(x)+σJn(x). We generated these using Mathematica, see Fig. 1B.

For the remaining cases we used the finite elements package FreeFem [11, 20]. In Fig. 2 we display two ergodic examples, the quarter-stadium billiard and a uniformly hyperbolic, Sinai-type dispersing billiard which was investigated numerically by Barnett [2].

It is also of interest to understand rational polygons, that is simple plane polygons all of whose vertex angles are rational multiples of π (Fig. 3), when we expect an analogue of Theorem 1.6 to hold, compare [23].

Fig. 3.

Fig. 3

The first 200 RN gaps for two examples of rational polygons: An L-shaped billiard (A) made of 4 squares of sidelength 1/2, and a right triangle with an angle π/5 and a long side of length unity (B)

The case of dynamics with a mixed phase space, such as the mushroom billiard investigated by Bunimovich [6] (see also the survey [26]) also deserves study, see Fig. 4.

Fig. 4.

Fig. 4

The first 200 RN gaps for the mushroom billiard, with a half-disk of diameter 3 on top of a unit square, which has mixed (chaotic and regular) billiard dynamics

Generalities About the RN Gaps

Robin–Neumann bracketing and positivity of the RN gaps

We recall the min-max characterization of the Robin eigenvalues

λnσ=infMH1(Ω)dimM=nsup0uMΩ|u|2dx+Ωσu2dsΩu2dx

where H1(Ω) is the Sobolev space. This shows that λnσλn0 if σ>0. Likewise, there is a min-max characterization of the Dirichlet eigenvalues with H1(Ω) replaced by the subspace H01(Ω), the closure of functions vanishing near the boundary:

λn=infMH01(Ω)dimM=nsup0uMΩ|u|2dxΩu2dx.

This shows that λnσλn.

In fact, we have strict inequality,

λn0<λnσ<λn.

This is proved (in greater generality) in [29] using a unique continuation principle.

A variational formula for the gaps

Lemma 3.1

Let ΩRd be a bounded Lipschitz domain. Then

dn(σ):=λnσ-λn0=0σΩ|un,τ|2dsdτ 3.1

where un,τ is any L2(Ω)-normalized eigenfunction associated with λnτ.

Proof

According to [1, Lemma 2.11] (who attribute it as folklore), for any bounded Lipschitz domain ΩRd, and n1, the function σλnσ is strictly increasing for σ[0,), is differentiable almost everywhere in (0,), is piecewise analytic, and the non-smooth points are locally finite (i.e. finite in each bounded interval). It is absolutely continuous, and in particular its derivative dλnσ/dσ (which exists almost everywhere) is locally integrable, and for any 0α<β,

λnβ-λnα=αβdλnσdσdσ.

Moreover, there is a variational formula valid at any point where the derivative exists:

dλnσdσ=Ω|un,σ|2ds 3.2

where un,σ is any normalized eigenfunction associated with λnσ. Therefore

dn(σ)=λnσ-λn0=0σdλnτdτdτ=0σΩ|un,τ|2dsdτ.

We can ignore the finitely many points τ where (3.2) fails, as the derivative is integrable.

A general upper bound: Proof of Theorem 1.4

As a corollary, we can show that for the case of smooth boundary, we have an upper bound1

dn(σ)Ω,σ(λn)1/3.

Indeed, for the case of smooth boundary, [3, Proposition 2.4]2 give an upper bound on the boundary integrals of eigenfunctions

Ωun,σ2dsΩ(λnσ)1/3(λn)1/3,

uniformly in σ0.

As a consequence of the variational formula (3.1), we deduce

dn(σ)Ω(λn)1/3·σ

and in particular for planar domains, using Weyl’s law, we obtain for n1

dn(σ)Ωn1/3·σ.

The Mean Value

In this section we give a proof of Theorem 1.1, that

limN1NnNdn(σ)=2length(Ω)area(Ω)σ.

Denote

WN(σ):=1NnNΩun,σ2ds.

Using Lemma 3.1 gives

1Nn=1Ndn(σ)=0σ1NnNΩun,τ2dsdτ=0σWN(τ)dτ.

The local Weyl law [14] (valid for any piecewise smooth Ω) shows that for any fixed σ,

limNWN(σ)=2length(Ω)area(Ω)

so if we know that WN(τ)C is uniformly bounded for all τσ, then by the Dominated Convergence Theorem we deduce that

limN1Nn=1Ndn(σ)=0σlimNWN(τ)dτ=0σ2length(Ω)area(Ω)dτ=2length(Ω)area(Ω)·σ

as claimed.

It remains to prove a uniform upper bound for WN(σ).

Lemma 4.1

There is a constant C=C(Ω) so that for all σ>0 and all N1,

1NnNΩun,σ2dsC.

Proof

What we use is an upper bound on the heat kernel on the boundary. Let Kσ(x,y;t) be the heat kernel for the Robin problem. Then [14, Lemma 12.1],

Kσ(x,y;t)Ct-dimΩ/2exp(-δ|x-y|2/t) 4.1

where C,δ>0 depend only on the domain Ω. Moreover, on the regular part of the boundary,

Kσ(x,y;t)=n0e-tλnσun,σ(x)un,σ(y).

We have for nN that λnσλNσλN, so for Λ=λN,

nNΩun,σ2dsenNe-λnσ/ΛΩun,σ2dseΩKσx,x;1Λds.

By (4.1),

ΩKσx,x;1ΛdsΩΛdimΩ/2.

Thus we find a uniform upper bound

λnσΛΩun,σ2dsΩΛdimΩ/2N

on using Weyl’s law, that is for all σ>0

1NnNΩun,σ2dsC(Ω).

We note that the mean value result is valid in any dimension d2 for piecewise smooth domains ΩRd as in [14], in the form

limN1NnNdn(σ)=2vold-1(Ω)vold(Ω)σ.

Indeed [14] prove the local Weyl law in that context, and Lemma 4.1 is also valid in any dimension.

A Uniform Lower Bound for the Gaps

To obtain the lower bound of Theorem 1.3 for the gaps, we use the variational formula (3.1) to relate the derivative dλnσ/dσ to the boundary integrals Ωun,σ2ds, where un,σ is any eigenfunction with eigenvalue λnσ, and for that will require a lower bound on these boundary integrals.

A lower bound for the boundary integral

The goal here is to prove a uniform lower bound for the boundary data of Robin eigenfunctions on a star-shaped, smooth planar domain Ω.

Theorem 5.1

Let ΩR2 be a star-shaped bounded planar domain with smooth boundary. Let f be an L2(Ω) normalized Robin eigenfunction associated with the n-th eigenvalue λnσ. Then there are constants C>0, A,B0 depending on Ω so that for all n1,

Ωf2ds1Aσ2+Bσ+C>0. 5.1

For σ=0 (Neumann problem), this is related to the L2 restriction bound of Barnett–Hassell–Tacy [3, Proposition 6.1].

The Neumann case σ=0

We first show the corresponding statement for Neumann eigenfunctions (which are Robin case with σ=0), which is much simpler. Let f be a Neumann eigenfunction, that is (Δ+λ)f=0 in Ω, fn=0 in Ω. We may assume that λ>0, the result being obvious for λ=0 when f is a constant function. After translation, we may assume that the domain is star-shaped with respect to the origin.

We start with a Rellich identity ([27, Eq 2]): Assume that ΩRd is a Lipschitz domain. Let L=Δ+λ, and A=j=1dxjxj. For every function f on Ω

Ω(Lf)(Af)dx=ΩfnAf-12Ω||f||2j=1dxjxjn+λ2Ωf2j=1dxjxjn-d2λΩf2dx+d2-1Ω||f||2dx. 5.2

Using (5.2) in dimension d=2 for a normalized eigenfunction, so that Lf=0 and Ωf2=1, and recalling that for Neumann eigenfunctions fn=0 on Ω, gives

0=-12Ω||f||2xxn+yyn+λ2Ωf2xxn+yyn-λ

or

Ωf2-1λ||f||2xxn+yynds=2.

The term xxn+yyn is the inner product n(x)·x between the outward unit normal n(x)=(xn,yn) at the point xΩ and the radius vector x=(x,y) joining x and the origin. Since the domain is star-shaped w.r.t. the origin, we have on the boundary Ω

xxn+yyn=n(x)·x0

so that we can drop3 the term with ||f||2 and get an inequality

Ω(n(x)·x)f2ds2.

Replacing (n(x)·x)2CΩ on Ω gives Theorem 5.1 for σ=0:

Ωf21CΩ.

The Robin case

Using the Rellich identity (5.2) in dimension d=2 for a normalized eigenfunction, so that Lf=0 and Ωf2=1, gives

0=ΩfnAf-12Ω||f||2(n(x)·x)+λ2Ωf2(n(x)·x)-λ.

Now n(x)·x0 on the boundary Ω since Ω is star-shaped with respect to the origin, and λ>0, so we may drop the term with ||f||2 and get an inequality

Ωf2(x)(n(x)·x)ds+2λΩfnAf2.

Due to the boundary condition, we may replace the normal derivative fn by -σf, and obtain, after using 0n(x)·x2C=2CΩ (we may take 2C to be the diameter of Ω), that

CΩf2-σλΩf(Af)1. 5.3

To proceed further, we need:

Lemma 5.2

Assume that Ω is smooth. There are numbers P,Q0, not both zero, depending only on Ω, so that for any normalized σ-Robin eigenfunction f,

Ωf(Af)ds(P+σQ)Ωf2ds. 5.4

Proof

Decompose the vector field A=xx+yy into its normal and tangential components along the boundary:

A=pn+qτ

where pq are functions on the boundary Ω. For example, for the circle x2+y2=ρ2, we have A=ρn and the normal derivative is just the radial derivative n=r, so that pρ, and q0.

Then using the Robin condition fn=-σf on Ω gives

Ωf(Af)ds=Ωfpfn+qfτds=-σΩpf2ds+Ωqffτds.

Setting P:=maxΩ|p|, we have

-σΩpf2dsσPΩf2ds

so it remains to bound Ωqffτds.

Let γ:[0,L]Ω be an arclength parameterization with γ(0)=γ(L). Then note that the tangential derivative of f at x0=γ(s0) is

fτ(x0)=ddsf(γ(s))|s=s0

and hence

ffτ=12(f2)τ=12ddsf(γ(s))2

so that abbreviating q(s)=q(γ(s)) and integrating by parts

Ωqffτds=120Lq(s)ddsf(γ(s))2ds=12q(s)f(γ(s))2|0L-120Lq(s)f(γ(s))2ds.

Because the curve is closed: γ(L)=γ(0), the boundary terms cancel out:

q(s)f(γ(s))2|0L=q(γ(L))f(γ(L))2-q(γ(0))f(γ(0))2=0

and so

Ωqffτds=120Lq(s)f(γ(s))2dsQΩf2ds

where Q=maxΩ|dqdτ|. Altogether we found that

Ωf(Af)ds(σP+Q)Ωf2ds.

We may now conclude the proof of Theorem 5.1 for σ>0: Take f=un,σ the n-th eigenfunction, with n1. Inserting (5.4) into (5.3) we find

1CΩf2-σλΩf(Af)C+σ(P+Qσ)λnσΩf2ds.

Hence we find, on replacing λnσλ1σλ10>0, that

Ωf2ds1C+σ(P+Qσ)/λ10>0

which is of the desired form.

Proof of Theorem 1.3

We use the variational formula (3.1) for n1 with the lower bound (5.1) of Theorem 5.1

dn(σ)=0σΩun,τ2dsdτ0σdτAτ2+Bτ+C=:c1(Ω,σ)>0.

For n=0, we just use positivity of the RN gap d0(σ)>0, and finally deduce that for all n0, and σ>0,

dn(σ)c(Ω,σ):=minc1(Ω,σ),d0(σ)>0.

Ergodic Billiards

In this section we give a proof of Theorem 1.6. By Chebyshev’s inequality, it suffices to show:

Proposition 6.1

Let ΩR2 be a bounded, piecewise smooth domain. Assume that the billiard map for Ω is ergodic. Then for every σ>0,

limN1NnNdn(σ)-2length(Ω)area(Ω)·σ=0. 6.1

Proof

We again use the variational formula (3.1)

dn(σ)=0σΩun,τ2dsdτ.

We have

dn(σ)-2length(Ω)area(Ω)σ=0σ(Ωun,τ2ds)dτ-2length(Ω)area(Ω)σ=0σ(Ωun,τ2ds-2length(Ω)area(Ω))dτ0σΩun,τ2ds-2length(Ω)area(Ω)dτ.

Therefore

1NnNdn(σ)-2length(Ω)area(Ω)σ0σ1NnNΩun,τ2ds-2length(Ω)area(Ω)dτ=:0σSN(τ)dτ

where

SN(τ):=1NnNΩun,τ2ds-2length(Ω)area(Ω).

Hassell and Zelditch [14, eq 7.1] (see also Burq [7]) show that if the billiard map is ergodic then for each σ0,

limN1NnNΩun,σ2ds-2length(Ω)area(Ω)2=0. 6.2

Therefore, by Cauchy–Schwarz, SN(τ) tends to zero for all τ0, by (6.2); by Lemma 4.1 we know that SN(τ)C is uniformly bounded for all τσ, so that by the Dominated Convergence Theorem we deduce that the limit of the integrals tends to zero, hence that

limN1NnNdn(σ)-2length(Ω)area(Ω)σ=0.

We note that Theorem 1.6 is valid in any dimension d2 for piecewise smooth domains ΩRd with ergodic billiard map as in [14], with the mean value interpreted as 2vold-1(Ω)vold(Ω)σ.

Variable Robin Function

In this section, we indicate extensions of our general results to the case of variable boundary conditions.

Variable boundary conditions

The general Robin boundary condition is obtained by taking a function on the boundary σ:ΩR which we assume is always non-negative: σ(x)0 for all xΩ. Thus we look for solutions of

Δu+λu=0onΩ,un(x)+σ(x)u(x)=0,xΩ

which is interpreted in weak form as saying that

Ωun·v+Ωσunv=λnΩunv

for all vH1(Ω). We will assume that σ is continuous. Then we obtain positive Robin eigenvalues

0<λ0σλ1σ

except that in the Neumann case σ0 we also have zero as an eigenvalue.

Robin to Neumann bracketing is still valid here, in the following form: if σ1,σ2C(Ω) are two continuous functions with 0σ1σ2 and such that there is some point x0Ω such that there is strict inequality σ1(x0)<σ2(x0) (by continuity this therefore holds on a neighborhood of x0), then we have a strict inequality [29]

λnσ1<λnσ2,n1. 7.1

Fix such a Robin function σC(Ω), which is positive: σ(x)>0 for all xΩ. We are interested in the Robin–Neumann gaps

dn(σ):=λnσ-λn0

which are positive by (7.1).

Extension of general results

The lower and upper bounds of Theorems 1.3 and 1.4 remain valid for variable σ by an easy reduction to the constant case: Let

σmin=minxΩσ(x),σmax=maxxΩσ(x)

so that 0<σminσmax (with equality only if σ is constant). Using (7.1) gives

λn0<λnσminλnσλnσmax

so that

dn(σmin)dn(σ)dn(σmax).

For instance, the universal lower bound for star-shaped domains (Theorem 1.3) follows because dn(σ)dn(σmin)C(σmin)>0, etcetera.

The existence of mean values (Theorem 1.1) and the almost sure convergence of the gaps to the mean value in the ergodic case (Theorem 1.6) require an adjustment of the variational formula (Lemma 3.1) which is provided in Sect. 7.3. Once that is in place, the result is

limN1Nn=1Ndn(σ)=2Ωσ(s)dsarea(Ω). 7.2

Given the mean value formula (7.2), Theorem 1.6 (almost sure convergence of the RN gaps to the mean in the ergodic case) also follows.

A variational formula

Let ΩRd be a bounded Lipschitz domain. Fix a continuous, positive Robin function σ:ΩR>0. We consider a one-parameter deformation of the boundary value problem Δu+λu=0,

un(x)+ασ(x)u(x)=0,xΩ 7.3

with a real parameter α0. Denote the corresponding eigenvalues by

λ1(α)λ2(α)λn(α)

By Robin–Neumann bracketing, if 0α1<α2 then

λn(α1)<λn(α2),n1.

The previous RN gaps dn(σ) are precisely λn(1)-λn(0). The variational formula for the RN gaps is:

Lemma 7.1

Let ΩRd be a bounded Lipschitz domain. Then

dn(σ)=01Ω|un,α|2dsdα

where un,α is any L2(Ω)-normalized eigenfunction associated with λn(α).

The proof is identical to that of Lemma 3.1, except that we need a reformulation of [1, Lemma 2.11] to this context4:

Lemma 7.2

Let ΩRd be a bounded Lipschitz domain and σ a continuous function on the boundary Ω which is positive: σ(x)>0 for all xΩ. For α0, let λn(α) be the eigenvalues of the Robin eigenvalue problem (7.3)

Then for n1, λn(α) is an absolutely continuous and strictly increasing function of α[0,), which is differentiable almost everywhere in (0,). Where it exists, its derivative is given by

ddαλn(α)=Ωσun,α2Ωun,α2 7.4

where un,αH1(Ω) is any eigenfunction associated with λn(α).

Proof

The proof is verbatim that of [1, Lemma 2.11] where σ1. As is explained there, each eigenvalue depends locally analytically on α, with at most a locally finite set of splitting points. We just repeat the computation of the derivative at any α which is not a splitting point for λn(α): We use the weak formulation of the boundary condition, as saying that for all vH1(Ω),

Ωun,α·v+Ωασ(s)un,α(s)v(s)ds=λn(α)Ωun,αv. 7.5

In particular, applying (7.5) with v=un,β gives

Ωun,α·un,β+Ωασun,αun,βds=λn(α)Ωun,αun,β. 7.6

Changing the roles of α and β gives

Ωun,α·un,β+Ωβσun,αun,βds=λn(β)Ωun,αun,β. 7.7

Subtracting (7.6) from (7.7) gives

λn(β)-λn(α)β-α=Ωσ(s)un,α(s)un,β(s)dsΩun,αun,β.

Taking the limit βα and assuming that un,βun,α in H1(Ω) as βα, as verified in [1, Lemma 2.11] so that in particular the denominator is eventually nonzero, gives (7.4).

Boundedness of RN Gaps for Rectangles

We consider the rectangle QL=[0,1]×[0,L], with L(0,1] the aspect ratio. We denote by λ0σλ1σ the ordered Robin eigenvalues. We will prove Theorem 1.7, that

0<λnσ-λn0<CL(σ).

The one-dimensional case

Let σ>0 be the Robin constant. The Robin problem on the unit interval is -un=kn2un, with the one-dimensional Robin boundary conditions

-u(0)+σu(0)=0,u(1)+σu(1)=0.

The eigenvalues of the Laplacian on the unit interval are the numbers -kn2 where the frequencies kn=kn(σ) are the solutions of the secular equation (k2-σ2)sink=2kσcosk, or

tan(k)=2σkk2-σ2 8.1

(see Fig. 5) and the corresponding eigenfunctions are

un(x)=kncos(knx)+σsin(knx).

Fig. 5.

Fig. 5

The secular equation (8.1) for σ=4. Displayed are plots of tank versus 2σkk2-σ2

As a special case5 of Dirichlet–Neumann bracketing (1.1), we know that given σ>0, for each n0 there is a unique solution kn=kn(σ) of the secular equation (8.1) with

kn(nπ,(n+1)π),n0.

Note that kn(0)=nπ.

From (8.1), we have as n,

kn(σ)=nπ+arctan2σkn(σ)11-σ2kn(σ)2=nπ+2σkn(σ)+Okn(σ)-3

so that

kn(σ)2-kn(0)24σ,n. 8.2

We can interpret, for Ω being the unit interval, 4=2#Ω/lengthΩ so that we find convergence of the RN gaps to their mean value in this case.

From (8.2) we deduce:

Lemma 8.1

For every σ>0, there is some C(σ)>0 so that

kn(σ)2-kn(0)2C(σ),n0. 8.3

Proof of Theorem 1.7

The frequencies for the interval [0, L] are 1L·km(σ·L). Hence the Robin energy levels of QL are the numbers

Λn,m(σ)=kn(σ)2+1L2·km(σ·L)2,n,m0. 8.4

We have

0Λn,m(σ)-Λn,m(0)=(kn(σ)2-kn(0)2)+1L2·km(σ·L)2-km(0)2.

From the one-dimensional result (8.3), we deduce that

Λn,m(σ)-Λn,m(0)C(σ)+1L2C(Lσ)=CL(σ).

We now pass from the Λm,n(σ) to the ordered eigenvalues {λkσ:k=0,1,}. We know that λkσλk0, and want to show that λkσλk0+CL(σ). For this it suffices to show that the interval Ik:=[0,λk0+CL(σ)] contains at least k+1 Robin eigenvalues, since then it will contain λ0σ,,λkσ and hence we will find λkσλk0+CL(σ).

The interval Ik contains the interval [0,λk0] and so certainly contains the first k+1 Neumann eigenvalues λ00,,λk0, which are of the form Λm,n(0) with (mn) lying in a set Sk. Since Λm,n(σ)Λm,n(0)+CL(σ), the interval Ik must contain the k+1 eigenvalues {Λm,n(σ):(m,n)Sk}, and we are done.

Application of Boundedness of the RN Gaps to Level Spacings

In this section, we show that the level spacing distribution of the Robin eigenvalues for the desymmetrized square is a delta function at the origin, as is the case with Neumann or Dirichlet boundary conditions.

Recall the definition of the level spacing distribution: We are given a sequence of levels x0x1x2. We assume that xN=cN+o(N), as is the case of the eigenvalues of a planar domain. Let δn=(xn+1-xn)/c be the normalized nearest neighbour gaps. so that the average gap is unity. The level spacing distribution P(s) of the sequence is then defined as

0yP(s)ds=limN1N#{nN:δny}

(assuming that the limit exists).

Recall that the Robin spectrum has systematic double multiplicities Λm,n(σ)=Λn,m(σ) (see (8.4) with L=1), which forces half the gaps to vanish for a trivial reason. To avoid this issue, one takes only the levels Λm,n(σ) with mn, which we call the desymmetrized Robin spectrum.

Theorem 9.1

For every σ0, the level spacing distribution for the desymmetrized Robin spectrum on the square is a delta-function at the origin.

In other words, if we denote by λ0σλ1σ the ordered (desymmetrized) Robin eigenvalues, then the cumulant of the level spacing distribution satisfies: For all y>0,

0yP(s)ds=limN1N#nN:12area(Ω)4π(λn+1σ-λnσ)y=1.

Proof

The Neumann spectrum for the square consists of the numbers m2+n2 (up to a multiple), with m,n0. There is a systematic double multiplicity, manifested by the symmetry (m,n)(n,m). We remove it by requiring mn. Denote the integers which are sums of two squares by

s1=0<s2=1<s3=2<s4=4<s5=5<<s14=25<

We define index clusters Ni as the set of all indices of desymmetrized Neumann eigenvalues which coincide with si:

Ni={n:λn0=si}

For instance, s0=0=02+02 has multiplicity one, and gives the index set N1={1}; s1=1=02+12 has multiplicity 1 (after desymmetrization) and gives N2={2}; s3=2=12+12 giving N3={3}, s14=25=02+52=32+42, N14={14,15}, etcetera. Then these are sets of consecutive integers which form a partition of the natural numbers {1,2,3,}, and if i<j then the largest integer in Ni is smaller than the smallest integer in Nj.

Denote by λnσ the ordered desymmetrized Robin eigenvalues: λ0σλ1σ, so for σ=0 these are just the integers si repeated with multiplicity #Ni. For each σ0, we define clusters Ci(σ) as the set of all desymmetrized Robin eigenvalues λnσ with nNi:

Ci(σ)={λnσ:nNi}.

Now use the boundedness of the RN gaps (Theorem 1.7): 0λnσ-λn0C(σ), to deduce that the clusters have bounded diameter:

diamCi(σ)C(σ).

If #Ni=1 then diamCi(σ)=0, so we may assume that #Ni2 and write

Ni={n-,n-+1,,n+},n+=maxNi,n-=minNi.

Then

diamCi(σ)=λn+σ-λn-σ=(λn+σ-si)+(si-λn-σ)=(λn+σ-λn+0)-(λn-σ-λn-0)C(σ)-0=C(σ).

For the first N eigenvalues, the number I of clusters containing them is the number of the si involved, which is at most the number of siλNσN. A classical result of Landau [18] states that the number of integers N which are sums of two squares is about N/logN, in particular6 is o(N). Hence

I#{i:siN}=o(N).

We count the number of nearest neighbour7 gaps δnσ=λn+1σ-λnσ of size bigger than y. Of these, there are at most I such that λn+1σ and λnσ belong to different clusters, and since I=o(N) their contribution is negligible. For the remaining ones, we group them by cluster to which they belong:

#{nN:δnσ>y}=i=1I#n:λn+1σ,λnσCi(σ)&δnσ>y+o(N).

We have

#n:λn+1σ,λnσCi(σ)&δnσ>y=#nNi,n<maxNi,δnσ>y=nNin<maxNiδnσ>yyy<nNin<maxNiδnσ>yδnσy1ynNin<maxNiδnσ.

The sum of nearest neighbour gaps in each cluster is

nNin<maxNiδnσ=nNin<maxNi(λn+1σ-λnσ)=λmaxNiσ-λminNiσ=diamCi(σ)C(σ).

Thus we find

#{n:λn+1σ,λnσCi(σ)&δnσ>y}C(σ)y

so that

#{nN:δnσ>y}i=1IC(σ)y+o(N)=C(σ)yI+o(N).

Since I=o(N), and C, y are fixed, we conclude that

1N#{nN:δnσ>y}=o(1).

Thus the cumulant of the level spacing distribution satisfies: For all y>0,

0yP(s)ds=limN1N#{nN:δnσy}=1

so that P(s) is a delta function at the origin.

Note that the claim is not that all gaps λn+1σ-λnσ tend to zero. On the contrary, it is possible to produce thin sequences {n} so that λn+1σ-λnσ tend to infinity. Looking at the proof of Theorem 9.1, these correspond to the rare cases when λnσ and λn+1σ belong to neighboring “clusters” which are far apart from each other.

The Unit Disk

Upper bounds for dn via Weyl’s law

In this section we prove Theorem 1.9. We first show how to obtain upper bounds for the gaps dn from upper bounds in Weyl’s law for the Robin/Neumann problem. The result is that

Lemma 10.1

Let Ω be a bounded planar domain. Assume that there is some θ(0,1/2) so that

Nσ(x):=#{λnσx}=area(Ω)4πx+length(Ω)4πx+Oσ(xθ). 10.1

and the same result holds for σ=0. Then we have

dn(σ)σnθ.

Proof

We first note that (10.1) gives

N0(λn0)=n+O(nθ), 10.2

and likewise for the Robin counting function, as will be explained below. Now compare the counting functions Nσ(λnσ) and N0(λn0) for the Robin and Neumann spectrum using (10.1) and (10.2):

n+O(nθ)=Nσ(λnσ)=area(Ω)4πλnσ+length(Ω)4πλnσ+Oσ(nθ)

and

n+O(nθ)=N0(λn0)=area(Ω)4πλn0+length(Ω)4πλn0+O(nθ).

Subtracting the two gives

(λnσ-λn0)·area(Ω)+length(Ω)λnσ+λn0=Oσ(nθ),

and therefore

dn(σ)=λnσ-λn0=Oσ(nθ).

To show (10.2), denote λ=λn0, and pick ε(0,1) sufficiently small so that in the interval [λ-ε2,λ+ε2] there are no eigenvalues other than λ, which is repeated with multiplicity K1. Then

Nλ+ε2-Nλ-ε2=K.

On the other hand, by Weyl’s law (with A=area(Ω)/4π, B=length(Ω)/4π)

K=Nλ+ε2-Nλ-ε2=Aλ+ε2+Bλ+ε2+Oλ+ε2θ-Aλ-ε2+Bλ-ε2+Oλ-ε2θ=Aε+Oελ+O(λθ).

Now use |N(λn0)-n|Kλθnθ which gives (10.2).

Below we implement this strategy for the disk to obtain Theorem 1.9.

Relating Weyl’s law and a lattice point count

The eigenvalues of the Laplacian on the disk are squares of zeros of Bessel functions and understanding Weyl’s law leads to requiring knowledge of the semiclassical asymptotics of these Bessel zeros; the nature of these asymptotics leads to an exotic lattice point problem, as was exploited by Kuznetsov and Fedosov [16] and Colin De Verdière [8].

Define the domain

D=x,y:x-1,1,max0,-xygx

where

gx=1π1-x2-xarccosx. 10.3

Fig. 6.

Fig. 6

The domain D

Let

NDμ:=#n,k:n,k+max(0,-n)-34μD

and

Ndisk,σx:=#λnσx.

Proposition 10.2

Fix σ0. Then

NDμ-Cμ3/7-Cμ4/7Ndisk,σμ2NDμ+Cμ3/7+Cμ4/7.

The argument extends [8, Theorem 3.1], [12] (who fix a flaw in the argument of [8]) to Robin boundary conditions.

We can now prove Theorem 1.9. We use the result of [13]8

NDμ=area(D)μ2+μ2+O(μ2(1/3-δ))

where δ=1/990. Noting that

area(D)=area(Ω)4π=14,length(Ω)4π=12

we obtain from Proposition 10.2 that

Ndisk,σx=area(Ω)4πx+length(Ω)4πx+O(x1/3-δ).

Applying Lemma 10.1 gives

dn(σ)=O(n1/3-δ)

which proves Theorem 1.9.

Proof of Proposition 10.2

Fix a Robin parameter σ0. Separating variables in polar coordinates (r,θ) and inserting the boundary conditions, we find a basis of eigenfunctions of the form

fn,k(r,θ)=Jn(κn,kr)einθ,nZ,k=1,2,

with eigenvalues κn,k2, where κn,k is the k-th positive zero of xJnx+σJnx. In particular, for the Neumann case (σ=0), we get zeros of the derivative Jn(x), denoted by jn,k; since zero is a Neumann eigenvalue we use the standard convention that x=0 is counted as the first zero of J0(x).

Let

S=x,y:ymax0,-x,

and let F:SR be the degree 1 homogeneous function satisfying F1 on the graph of g. Obviously,

Fn,k+max0,-n-34μn,k+max0,-n-34μD;

on the other hand, as will be shown in Lemma 10.3 below, the numbers κn,k are well approximated by Fn,k+max0,-n-34. This will give the desired connection between Weyl’s law on the disk and the lattice count problem in dilations of D.

Lemma 10.3

Fix σ0, and let c>0 be a constant.

1. As n, uniformly for kn/c, we have

κn,k=F(n,k-34)+Oc,σn1/3k4/3. 10.4

2. As k, uniformly for nc·k, we have

κn,k=Fn,k+max0,-n-34+Oc,σ1k. 10.5

The proof of Lemma 10.3 will be given in “Appendix A”.

It will be handy to derive an explicit formula for the function F, which we will now do. Let ζ=ζz be the solution to the differential equation

dζdz2=1-z2ζz2 10.6

which for z1 is given by

23-ζ3/2=z2-1-arccos1z 10.7

(see [24, Eq. 10.20.3]). The interval z1 is bijectively mapped to the interval ζ0; denote by z=zζ the inverse function.

Lemma 10.4

For x>0, we have

F(x,y)=xz(-x-2/33π2y2/3). 10.8

Additionally, for y0 we have F0,y=πy, and for -x,yS we have

F-x,y=F(x,y-x). 10.9

Proof

Let x>0, and denote t=Fx,yx. Then F1t,ytx=1 so that the point 1t,ytx lies on the graph of g,  and therefore

yx=1πt2-1-arccos1t=1π23-ζt3/2

so that

t=z(-x-2/33π2y2/3).

The other claims are also straightforward from the definitions.

We proceed towards the proof of Proposition 10.2 by following the ideas of [8, Sec. 3]. Let

ND1μ=#n,k:n,k+max0,-n-34μD,n<c·k,ND2μ=#n,k:n,k-34μD,nc·k,

and

Ndisk,σ1μ2=#n,k:κn,kμ,n<c·k,Ndisk,σ2μ2=#n,k:κn,kμ,nc·k,

so that

NDμ=ND1μ+2ND2μ

and

Ndisk,σμ2=Ndisk,σ1μ2+2Ndisk,σ2μ2,

where we used (10.9) and the relation κ-n,k=κn,k. We first compare ND1μ and Ndisk,σ1μ2:

Lemma 10.5

There exists a constant C=Cc,σ>0 such that

ND1μ-CμNdisk,σ1μ2ND1μ+Cμ.

Proof

Assume that |n|<c·k. By (10.9) and the homogeneity of F we have

Fn,k+max(0,-n)-34=F|n|,k-34=kF|n|k,1-34k,

and since 1F|n|k,1-34kc1, we conclude that

kFn,k+max(0,-n)-34ck.

Hence, if Fn,k+max(0,-n)-34μ, then kcμ. Combining this with Lemma 10.3, we see that

Ndisk,σ1μ2#n,k:F(n,k+max(0,-n)-34)μ+Ck,n<c·k=#n,k:F(n,k+max0,-n-34)μ,n<c·k+#n,k:μ<F(n,k+max0,-n-34)μ+Ck,n<c·k#n,k:F(n,k+max0,-n-34)μ+Cμ,n<c·k=ND1μ+Cμ.

The proof of the other inequality is similar.

We will now compare between ND2μ and Ndisk,σ2μ2. To this end, for fixed k1, we denote

Nkμ=#n:n,k-34μD,nc·kNkμ=#n:κn,kμ,nc·k.

Lemma 10.6

Given a sufficiently large c>0, there exists a constant C=Cc,σ>0 such that

Nkμ-Cμ1/3k4/3-1NkμNkμ+Cμ1/3k4/3+1.

Proof

Let

Akμ:=#n:μ<Fn,k-34μ+Cμ1/3k4/3,nc·k,

and recall the inequality (see (A.1)) njn,kκn,k, so in particular if κn,kμ, then nμ. Thus, Lemma 10.3 gives

Nkμ#n:Fn,k-34μ+Cn1/3k4/3,μnc·k#n:Fn,k-34μ,nc·k+Akμ=Nkμ+Akμ.

When xc·k, we have Fx,k-34=xF1,k-3/4x, and therefore (note that F(1,y)1 for all y0)

Fxx,k-34=F1,k-3/4x-k-3/4xFy1,k-3/4x1

when c is taken sufficiently large. In particular, F~(x):=F(x,k-34) is strictly increasing for xc·k, and so Ak(μ) is bounded above by the number of integer points in the interval

I:=F~-1max(μ,F~(c·k)),F~-1μ+Cμ1/3k4/3,

which in turn is bounded above by length(I)+1; by the mean value theorem, keeping in mind that (F~-1)x=F~x-1, we conclude that

length(I)Cμ1/3k4/3·maxxI1F~x(x)Cμ1/3k4/3.

The proof of the other inequality is similar.

Remark 10.7

The +1 factor was missing in [8].

For large values of k we will use the following estimate:

Lemma 10.8

There exists a constant C=Cc,σ>0 such that for k>μ4/7, we have

Nkμ-Cμ3/7NkμNkμ+Cμ3/7.

Proof

By Lemma 10.3,

Nkμ#n:Fn,k-34μ+Cμ3/7,nc·k=Nkμ+Cμ3/7

and likewise

NkμNkμ-Cμ3/7.

Proof of Proposition 10.2

By Lemma 10.6 (applied for kμ4/7) and Lemma 10.8 (applied for k>μ4/7), we get that

Ndisk,σ2μ2=k1Nkμk1Nkμ+Cμ3/7+Cμ4/7=ND2μ+Cμ3/7+Cμ4/7

and likewise

Ndisk,σ2μ2ND2μ-Cμ3/7-Cμ4/7.

This, together with Lemma 10.5 gives the claim.

Acknowledgments

Open Access

This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Appendix A. Proof of Lemma 10.3

The goal of this appendix is to prove the asymptotic formulas (10.4) and (10.5) for the zeros κn,k of xJnx+σJnx where σ0. More generally, we will work with Bessel functions Jνx of real order ν. Many properties of the zeros κν,k are well-known, e.g. for all σ>0, ν0 and k1 we have (see e.g. [33, Eq. (III.6)])

νjν,k<κν,k<jν,k A.1

where jν,k (resp. jν,k) is the k-th positive zero of Jνx (resp. Jνx, with the convention that x=0 is counted as the first zero of J0(x)); for σ0 and fixed ν0 we have the asymptotic formula [33, Eq. (IV.9)]

κν,k=jν,k+σjν,k+-13σ3-12σ2+ν2σjν,k3+Oσjν,k-5k.

Recall the function ζz defined above by (10.6) which satisfies (10.7) for z1, with an inverse zζ. Denote hζ=4ζ1-z21/4. We have the following asymptotic expansion for Jννz as ν [24, Eq. 10.20.4]

JννzhζAiν2/3ζν1/3j=0Ajζν2j+Aiν2/3ζν5/3j=0Bjζν2j A.2

which holds uniformly for z>0, where Aiz is the Airy function, and the coefficients Ajζ and Bjζ are given by [24, Eq. 10.2.10, 10.20.11] and the remark following these equations. Likewise, we have the asymptotic expansion [24, Eq. 10.20.7]

Jννz-2zhζAiν2/3ζν4/3j=0Cjζν2j+Aiν2/3ζν2/3j=0Djζν2j A.3

uniformly for z>0, where the coefficients Cjζ and Djζ are given by [24, Eq. 10.2.12, 10.20.13] and the remark which follows them. Each of the coefficients Ajζ,Bjζ,Cjζ,Djζ, j=0,1,2, is bounded near ζ=0; we have A0ζ=D0ζ=1.

For the Robin parameter σ0, if we denote B-1ζ=0 and let

αjσζ:=Cjζ-σAjζh2ζ2βjσζ:=Djζ-σBj-1ζh2ζ2,

then (A.2) and (A.3) give

ϕννz:=Jννz+σνzJννz-2zhζAiν2/3ζν2/3j=0βjσzν2j+Aiν2/3ζν4/3j=0αjσzν2j

uniformly for z>0. Note that α0σζ=C0ζ-σh2ζ2. Using the derivation of [25, p. 345] with αjσ, βjσ instead of Cjζ, Djζ (the latter were used to establish the asymptotic expansion of the zeros of Jνz corresponding to σ=0), we get the following uniform asymptotic formula for κν,k as ν :

Lemma A.1

Fix σ0, let ak be the k-th zero of Aiz (all of these zeros are real and negative), and let ζ=ζν,k=ν-2/3ak. Then in the above notation, uniformly for k1, we have

κν,k=νzζ-zζC0ζ-σh2ζ2ζν+Oσ1ν. A.4

In particular, for σ=0 we reconstructed the formula [24, Eq. 10.21.43]

jν,k=νzζ-zζC0ζζν+O1ν

uniformly for k1 (note the identity zζ=-zζh2ζ2). We remark that the secondary term in (A.4) is necessary because of the ζ factor in the denominator which may be as small as ν-2/3 when k is small. This phenomenon does not occur for the zeros of Jνz (see [25, Sec. 7]), which satisfy the more compact uniform expansion

jν,k=νzζ+O1ν

where ζ=ν-2/3ak and ak is the k-th zero of Aiz [24, Eq. 10.21.41].

Recall that the zeros ak of Aiz satisfy the asymptotic formula [24, Eq. 9.9.8]

ak=-3π2k-342/3+Ok-4/3. A.5

Proof of Lemma 10.3, first part

Assume that kν/c, where ν. The functions zζ and hζ are bounded near ζ=0, and therefore inserting (A.5) into (A.4) gives

νzζ=νz-ν-2/33π2k-342/3+Oc,σν1/3k4/3

and

zζC0ζ-σh2ζ2ζνc,σ1ν1/3k2/3cν1/3k4/3.

Also note that 1νcν1/3k4/3 when kν/c. By (10.8) we have

z-ν-2/33π2k-342/3=Fν,k-34,

and therefore the above estimates yield

κν,k=Fν,k-34+Oc,σν1/3k4/3 A.6

which gives (10.4).

In order to prove the second part of Lemma 10.3, we require the following lemma. Recall the function gx defined in (10.3).

Lemma A.2

Fix σ0, and let C>0 be a constant. Let ϕνx:=Jνx+σxJνx. As x, uniformly for 0νx/1+C, we have

ϕνx=-2π1/2x2-ν21/4x-1sinπxgν/x-π4+OC,σx-1. A.7

Proof

We use the standard integral representation [24, Eq. 10.9.6]

Jνx=1π0πcosxsint-νtdt-sinνππ0e-xsinht-νtdt A.8

(for integer ν the second integral in (A.8) vanishes). Assume that x1+Cν0, and denote r=ν/x11+C<1. The first integral in (A.8) is equal to the real part of

Ir1x:=1π0πeixsint-rtdt.

By the method of stationary phase [24, Eq. 2.3.23], we have the asymptotics:

Ir1x=eπixgr-142π1-r2x1/2+e-irπxiπ1+rx+iπ1-rx+OCx-3/2.

Hence

ReIr1x=cosπxgr-π42π1-r2x1/2+sinπrxπ1+rx+OCx-3/2.

The second integral in (A.8) is equal to

Ir2x:=sinπrxπ0e-xsinht+rtdt

and can be evaluated by the Laplace method [24, Eq. 2.3.15]:

Ir2x=sinπrxπ1+rx+OCx-2.

We obtain

Jνx=2π1/2x2-ν2-1/4cosπxgν/x-π4+OCx-1; A.9

a similar procedure gives

Jνx=-2π1/2x2-ν21/4x-1sinπxgν/x-π4+OCx-1. A.10

The formula (A.7) now follows upon combining (A.9) and (A.10).

Proof of Lemma 10.3, second part

Let 0νc·k, where k. Clearly, the condition 0νc·k implies that κν,k1+Cν0 for some constant C=Cc>0 and that κν,kck (e.g. by the analogous well-known inequalities for the Bessel zeros jν,k, see (5.3) in [9], together with (A.1)). By Lemma A.2, we have

sinπκν,kgν/κν,k-π4=Oc,σk-1

so there exists an integer m such that

κν,kgν/κν,k=m-34+Oc,σk-1. A.11

This in particular gives mc,σκν,k. We have 1 Fyc,σ1 when yc,σx, as can be seen by differentiating (10.8) and combining with (10.6). This, together with the equality κν,k=F(ν,κν,kg(ν/κν,k)) and (A.11) gives

κν,k=Fν,m-34+Oc,σk-1=Fν,m-34+Oc,σk-1.

We will now show that m=k: indeed, fix k,ν and the corresponding m, and assume that ν is close to ν, so there exists an integer m such that

κν,k=Fν,m-34+Oc,σk-1.

By the mean value theorem

|m-m|Fν,m-34-Fν,m-34Fν,m-34-Fν,m-34+|κν,k-κν,k|+Oc,σk-1. A.12

Note that κν,k is a continuous function of ν0 (in fact, it is differentiable in ν in this regime, see e.g. [19]), and so is Fν,m-34 as a function of ν. Therefore the right-hand-side of (A.12) can be made arbitrarily small when k is sufficiently large and ν is sufficiently close to ν, and hence m=m. We see that map νκν,km is well-defined and it is locally constant and hence constant for 0νc·k. But we know by (A.6) that for νk we have m=k; hence m=k for any 0νc·k. This gives (10.5) when n0; for n<0, (10.4) follows from the relations κ-n,k=κn,k and (10.9).

Footnotes

1

Here and in the sequel we use AB as an alternative to A=O(B).

2

Their Proposition 2.4 is stated only for the Neumann case, but as is pointed out in Remark 2.7, the proof applies to Robin case as well, uniformly in σ0; and they attribute it to Tataru [34, Theorem 3]. Note that their convention for the normal derivative is different than ours.

3

If we also allow negative Robin constant σ<0, we may have a finite number of negative eigenvalues and this part of the argument would not work for these.

4

[1, Lemma 2.11] allows ΩRN to be any bounded Lipschitz domain and takes σ1.

5

Of course, in this case it directly follows from the secular equation (8.1).

6

This is much easier to show using a sieve.

7

For simplicity we replace 12area(Ω)4π by 1, that is we don’t bother normalizing so as to have mean gap unity; the result is independent of this normalization.

8

They treat the shifted lattice Z2-(0,14) but as they say [13, Remark 6.5], the arguments also work for the shift by (0,34). See [12] for an a further improvement in the Dirichlet case to 131/416=1/3-23/1248=0.314904.

We thank Michael Levitin and Iosif Polterovich for the their comments. This research was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant Agreement No. 786758) and by the ISRAEL SCIENCE FOUNDATION (Grant No. 1881/20).

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Zeév Rudnick, Email: rudnick@tauex.tau.ac.il.

Igor Wigman, Email: igor.wigman@kcl.ac.uk.

Nadav Yesha, Email: nyesha@univ.haifa.ac.il.

References

  • 1.Antunes PRS, Freitas P, Kennedy JB. Asymptotic behaviour and numerical approximation of optimal eigenvalues of the Robin Laplacian. ESAIM Control Optim. Calculus Var. 2013;19(2):438–459. doi: 10.1051/cocv/2012016. [DOI] [Google Scholar]
  • 2.Barnett AH. Asymptotic rate of quantum ergodicity in chaotic Euclidean billiards. Commun. Pure Appl. Math. 2006;59(10):1457–1488. doi: 10.1002/cpa.20150. [DOI] [Google Scholar]
  • 3.Barnett AH, Hassell A, Tacy M. Comparable upper and lower bounds for boundary values of Neumann eigenfunctions and tight inclusion of eigenvalues. Duke Math. J. 2018;167(16):3059–3114. doi: 10.1215/00127094-2018-0031. [DOI] [Google Scholar]
  • 4.Berry MV, Dennis MR. Boundary-condition-varying circle billiards and gratings: the Dirichlet singularity. J. Phys. A. 2008;41(13):5203. doi: 10.1088/1751-8113/41/13/135203. [DOI] [Google Scholar]
  • 5.Bucur, D., Freitas, P., Kennedy, J.: The Robin problem. In: Shape Optimization and Spectral Theory, pp. 78–119. De Gruyter Open, Warsaw (2017)
  • 6.Bunimovich LA. Mushrooms and other billiards with divided phase space. Chaos. 2001;11:802–808. doi: 10.1063/1.1418763. [DOI] [PubMed] [Google Scholar]
  • 7.Burq N. Quantum ergodicity of boundary values of eigenfunctions: a control theory approach. Can. Math. Bull. 2005;48(1):3–15. doi: 10.4153/CMB-2005-001-3. [DOI] [Google Scholar]
  • 8.Colin De Verdière Y. On the remainder in the Weyl formula for the Euclidean disk. Sémin. Théo. Spectr. Géom. 2010;29:1–13. [Google Scholar]
  • 9.Colin De Verdière, Y., Guillemin, V., Jerison, D.: Singularities of the wave trace near cluster points of the length spectrum. arXiv:1101.0099 [math.AP]
  • 10.Daners D, Kennedy J. On the asymptotic behaviour of the eigenvalues of a Robin problem. Differ. Integral Equ. 2010;23(7–8):659–669. [Google Scholar]
  • 11.Hecht F. New development in FreeFem++ J. Numer. Math. 2012;20(3–4):251–266. [Google Scholar]
  • 12.Guo, J., Müller, W., Wang, W., Wang, Z.: The Weyl formula for planar annuli. arXiv:1907.03669 [math.SP]
  • 13.Guo J, Wang W, Wang Z. An improved remainder estimate in the Weyl formula for the planar disk. J. Fourier Anal. Appl. 2019;25(4):1553–1579. doi: 10.1007/s00041-018-9637-z. [DOI] [Google Scholar]
  • 14.Hassell A, Zelditch S. Quantum ergodicity of boundary values of eigenfunctions. Commun. Math. Phys. 2004;248(1):119–168. doi: 10.1007/s00220-004-1070-2. [DOI] [Google Scholar]
  • 15.Khalile M. Spectral asymptotics for Robin Laplacians on polygonal domains. J. Math. Anal. Appl. 2018;461(2):1498–1543. doi: 10.1016/j.jmaa.2018.01.062. [DOI] [Google Scholar]
  • 16.Kuznetsov NV, Fedosov BV. An asymptotic formula for eigenvalues of a circular membrane. Differ. Uravn. 1965;1(12):1682–1685. [Google Scholar]
  • 17.Lacey AA, Ockendon JR, Sabina J. Multidimensional reaction diffusion equations with nonlinear boundary conditions. SIAM J. Appl. Math. 1998;58(5):1622–1647. doi: 10.1137/S0036139996308121. [DOI] [Google Scholar]
  • 18.Landau E. Über die Einteilung der positiven ganzen Zahlen in vier Klassen nach der Mindeszahl der zu ihrer additiven Zusammensetzung erforderlichen Quadrate. Arch. Math. Phys. 1908;13:305–312. [Google Scholar]
  • 19.Landau LJ. Ratios of Bessel functions and roots of αJν(x)+xJν(x)=0. J. Math. Anal. Appl. 1999;240(1):174–204. doi: 10.1006/jmaa.1999.6608. [DOI] [Google Scholar]
  • 20.Levitin, M.: A very brief introduction to eigenvalue computations with FreeFem. http://michaellevitin.net/Papers/freefem++.pdf (2019)
  • 21.Levitin M, Parnovski L. On the principal eigenvalue of a Robin problem with a large parameter. Math. Nachr. 2008;281(2):272–281. doi: 10.1002/mana.200510600. [DOI] [Google Scholar]
  • 22.Lou Y, Zhu M. A singularly perturbed linear eigenvalue problem in C1 domains. Pac. J. Math. 2004;214:323–334. doi: 10.2140/pjm.2004.214.323. [DOI] [Google Scholar]
  • 23.Marklof J, Rudnick Z. Almost all eigenfunctions of a rational polygon are uniformly distributed. J. Spectr. Theor. 2012;2:107–113. doi: 10.4171/JST/23. [DOI] [Google Scholar]
  • 24.Olver, F.W.J., Olde Daalhuis, A.B., Lozier, D.W., Schneider, B.I., Boisvert, R.F., Clark, C.W., Miller, B.R., Saunders, B.V., Cohl, H.S., McClain, M.A. (eds.) NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/. Release 1.0.27 of 2020-06-15
  • 25.Olver FWJ. The asymptotic expansion of Bessel functions of large order. Philos. Trans. R. Soc. Lond. Ser. A. 1954;247:328–368. doi: 10.1098/rsta.1954.0021. [DOI] [Google Scholar]
  • 26.Porter MA, Lansel S. Mushroom billiards. Not. Am. Math. Soc. 2006;53(3):334–337. [Google Scholar]
  • 27.Rellich F. Darstellung der Eigenwerte von Δu+λu=0 durch ein Randintegral. Math. Z. 1940;46:635–636. doi: 10.1007/BF01181459. [DOI] [Google Scholar]
  • 28.Rivière G, Royer J. Spectrum of a non-selfadjoint quantum star graph. J. Phys. A. 2020;53(49):495202. doi: 10.1088/1751-8121/abbfbe. [DOI] [Google Scholar]
  • 29.Rohleder J. Strict inequality of Robin eigenvalues for elliptic differential operators on Lipschitz domains. J. Math. Anal. Appl. 2014;418(2):978–984. doi: 10.1016/j.jmaa.2014.04.036. [DOI] [Google Scholar]
  • 30.Rudnick Z, Wigman I. On the Robin spectrum for the hemisphere (special issue in honor of A. Shnirelman) Ann. Math. Qué. 2021 doi: 10.1007/s40316-021-00155-9. [DOI] [Google Scholar]
  • 31.Rudnick, Z., Wigman, I.: The Robin problem on rectangles, accepted for publication in J. Math. Phys, special collection of papers honoring Freeman Dyson, available online arXiv:2103.15129
  • 32.Sieber M, Primack H, Smilansky U, Ussishkin I, Schanz H. Semiclassical quantization of billiards with mixed boundary conditions. J. Phys. A. 1995;28(17):5041–5078. doi: 10.1088/0305-4470/28/17/032. [DOI] [Google Scholar]
  • 33.Spigler, R.: Sulle radici dell’equazione: AJν(x)+BxJν(x)=0. Atti Sem. Mat. Fis. Univ. Modena 24(2), 399–419 (1975, 1976)
  • 34.Tataru D. On the regularity of boundary traces for the wave equation. Ann. Sc. Norm. Super. Pisa Cl. Sci. (4) 1998;26:185–206. [Google Scholar]
  • 35.Wolfram Research, Inc. Mathematica, Version 12.1, Champaign (2020)

Articles from Communications in Mathematical Physics are provided here courtesy of Springer

RESOURCES