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

Internal waves in 2D domains with ergodic classical dynamics

Yves Colin de Verdière and Zhenhao Li zhenhao@mit.edu Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139 (U.S.)
Abstract.

We study a model of internal waves in an effectively 2D aquarium under periodic forcing. In the case when the underlying classical dynamics has sufficiently irrational rotation number, we prove that the energy of the internal waves remains bounded. This involves studying the spectrum of a related 0-th order pseudodifferential operator at spectral parameters corresponding to such dynamics. For the special cases of rectangular and elliptic domains, we give an explicit spectral description of that operator.

1. Introduction

Below the surface layers of the ocean, the density field can be be approximated by a stable-stratified field. This means that the density depends only on depth and increases slowly with it. A standard model for internal waves is given by considering linear perturbations of such stable-stratified fluids, and is a central topic in oceanography. These perturbations occur naturally and can arise mechanically or thermodynamically. For a more complete introduction to the physics behind internal waves, see Mass [Maa05] and Sibgatullin–Ermanyuk [SE19].

In this paper, we consider an open, bounded, and simply-connected domain Ωx1,x22\Omega\subset\mathbb{R}^{2}_{x_{1},x_{2}} with a smooth boundary. Internal waves are modeled by the equation

(t2Δ+x22)u=f(x)cos(λt),u|t=0=tu|t=0=0,u|Ω=0(\partial_{t}^{2}\Delta+\partial_{x_{2}}^{2})u=f(x)\cos(\lambda t),\quad u|_{t=0}=\partial_{t}u|_{t=0}=0,\quad u|_{\partial\Omega}=0 (1.1)

with λ(0,1)\lambda\in(0,1) and fC(Ω¯;)f\in C^{\infty}(\overline{\Omega};\mathbb{R}). This is the Poincaré equation [Poi85], also called the Sobolev equation [Sob54]. This problem comes from the study of internal waves in a 2D aquarium with a constant Brunt-Vaïsälä frequency which we take equal to 11. The solution uu represents the stream function of the fluid velocity, meaning the velocity field is given by

𝐯=(x2u,x1u).\mathbf{v}=(\partial_{x_{2}}u,-\partial_{x_{1}}u). (1.2)

Then (1.1) can be interpreted as the evolution of the stream function under periodic forcing in the interior of the domain with forcing profile ff. The Dirichlet boundary condition is simply saying that velocity of the fluid near the boundary must be tangent to the boundary, i.e. no forcing from the boundary. Let gC(Ω;)g\in C^{\infty}(\partial\Omega;\mathbb{R}), and let g~|Ω=g\tilde{g}|_{\partial\Omega}=g and Δg~=0\Delta\tilde{g}=0. Then the boundary forced equation

(t2Δ+x22)u=0,u|t=0=g~,tu|t=0=0,u|Ω=g(x)cos(λt)(\partial_{t}^{2}\Delta+\partial_{x_{2}}^{2})u=0,\quad u|_{t=0}=\tilde{g},\quad\partial_{t}u|_{t=0}=0,\quad u|_{\partial\Omega}=g(x)\cos(\lambda t)

can be easily reduced to (1.1) by replacing u(x,t)u(x,t) with u(x,t)g~(x)cos(λt)u(x,t)-\tilde{g}(x)\cos(\lambda t).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1. Plots of u=Δω1wu=\Delta^{-1}_{\omega}w (ΔΩ\Delta_{\Omega} is the Dirichlet Laplacian) where ww is an eigenfunction of the Poincaré operator P=x22ΔΩ1P=\partial_{x_{2}}^{2}\Delta_{\Omega}^{-1}, see §4.2. We also plot the velocity field of u2,5u_{2,5} given by (1.2). It is clear that the integral curves are simply the level sets of uk,Nu_{k,N}, the zero set (excluding the boundary) is highlighted in red above.

Rewriting (1.2) so that (1.1) reads like an evolution problem, we define

P:=x22ΔΩ1:H1(Ω)H1(Ω),u,wH1(Ω):=ΔΩ1u,ΔΩ1wL2(Ω).P:=\partial_{x_{2}}^{2}\Delta_{\Omega}^{-1}:H^{-1}(\Omega)\to H^{-1}(\Omega),\quad\langle u,w\rangle_{H^{-1}(\Omega)}:=\langle\nabla\Delta_{\Omega}^{-1}u,\nabla\Delta_{\Omega}^{-1}w\rangle_{L^{2}(\Omega)}. (1.3)

where ΔΩ\Delta_{\Omega} denotes the Dirichlet Laplacian. Then w:=Δuw:=\Delta u satisfies the equation

(t2+P)w=fcosλt,w|t=0=tw|t=0=0,fCc(Ω;),u=ΔΩ1w.(\partial_{t}^{2}+P)w=f\cos\lambda t,\quad w|_{t=0}=\partial_{t}w|_{t=0}=0,\quad f\in C^{\infty}_{\mathrm{c}}(\Omega;\mathbb{R}),\quad u=\Delta_{\Omega}^{-1}w. (1.4)

The operator PP is bounded and self-adjoint, see [Ral73]. The solution w(t)w(t) can then be written using the functional calculus:

w(t)=cos(tP)cos(tλ)λ2Pf.\begin{gathered}w(t)=\frac{\cos(t\sqrt{P})-\cos(t\lambda)}{\lambda^{2}-P}f.\end{gathered} (1.5)

We are interested in its long-time behavior which equivalent to the long time behavior of the solution uu to (1.1). As tt\to\infty the functional solution (1.5) becomes singular when the spectral parameter is equal to λ2\lambda^{2}. Therefore studying the long-time behavior is closely related top the spectral properties of PP at λ2\lambda^{2}.

Our main result concerns the behavior of uu when the underlying classical dynamics is ergodic. The relevant dynamics at forcing frequency λ\lambda is given by the chess billiard map, which is a λ\lambda-dependent family of circle diffeomorphism b(,λ):ΩΩb(\bullet,\lambda):\partial\Omega\to\partial\Omega (see (2.4) for details and §2.1 for motivation). To every b(,λ)b(\bullet,\lambda), we may assign a rotation number 𝐫(λ)[0,1]\mathbf{r}(\lambda)\in[0,1], which measures the average rotation of b(,λ)b(\bullet,\lambda), see (2.5) for the precise definition. When the rotation number 𝐫(λ)\mathbf{r}(\lambda) is irrational, the map b(,λ)b(\bullet,\lambda) is ergodic, which is the setting we will focus on.

Finally, we need a natural geometric assumption on the domain Ω\Omega, called λ\lambda-simplicity [DWZ21], see Definition 2.1. For now, we emphasize that all strictly convex domains satisfy this assumption for all λ(0,1)\lambda\in(0,1).

Theorem 1.

Let Ω2\Omega\subset\mathbb{R}^{2} be open, bounded, and simply connected. Assume that λ(0,1)\lambda\in(0,1) is such that Ω\Omega is λ\lambda-simple.

  1. (a)

    If the rotation number 𝐫(λ)\mathbf{r}(\lambda) is irrational, then λ2\lambda^{2} is not in the pure point spectrum of PP, i.e. (Pλ2):H1(Ω)H1(Ω)(P-\lambda^{2}):H^{-1}(\Omega)\to H^{-1}(\Omega) is injective.

  2. (b)

    If the rotation number 𝐫(λ)\mathbf{r}(\lambda) is Diophantine (see Definition 2.3), the solution u(t)u(t) to (1.1) remains bounded in energy space for all times, i.e. there exists a constant C>0C>0 such that

    u(t)H01(Ω)C\|u(t)\|_{H^{1}_{0}(\Omega)}\leq C (1.6)

    for all tt\in\mathbb{R}.

  3. (c)

    If the rotation number 𝐫(λ)\mathbf{r}(\lambda) is Diophantine, the spectral measure μf\mu_{f} of ff satisfies, for all NN\in\mathbb{N},

    μf([λ2ε,λ2+ε])=O(εN)\mu_{f}\left([\lambda^{2}-\varepsilon,\lambda^{2}+\varepsilon]\right)=O\left(\varepsilon^{N}\right)

Parts (b) and (c) of Theorem 1 were first proved by the second named author [Li23] (Theorem 2 and Theorem 1 respectively). This was done by studying a boundary reduced 0-th order pseudo-differential operator on the circle, found in [DWZ21]. The assumptions in (a) are weaker than the corresponding assumption of [Li23, Lemma 6.1], which assumes the eigenfunction to be smooth. It was already proven in [Arn61, Theorem 10], see also [Joh41].

In this paper, we give a simpler proof by directly constructing an inverse to a related eigenvalue problem (see (2.1)). We also describe the spectrum in two simple cases: rectangular and elliptic domains; the latter is related to the recent work of the first named author and Vidal [CdVV23] which studied the case rotating fluids in ellipsoids.

The spectral measure of PP can be numerically approximated using the methods developed in [Col21, CHT21, Col22]. Theorem 1 still holds for the square even though the boundary is not smooth. In that case, the rotation number can be computed as an explicit smooth function of λ\lambda, and so the spectral bound can be seen numerically in Figure 2.

Two examples where we can explicitly compute the spectrum of PP are given by Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1] and Ω=𝔻\Omega=\mathbb{D}, the unit disk. For the square, the eigenfunctions are simply the Fourier modes. For the circle, we will give an explicit complete basis of H01H^{1}_{0} consisting of solutions to the eigenvalue problem (2.1), see (4.11) and Figure 1. Then by changing the coordinates, we have spectral information about elliptic and rectangular domains.

Refer to caption
Figure 2. Numerical evidence for the relevance of the Diophantine assumption in part (c) of Theorem 1. Here Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1]. The spectral measure of a 2ε2\varepsilon interval centered at λ2\lambda^{2} with 𝐫(λ)\mathbf{r}(\lambda) Diophantine decays much faster than if 𝐫(λ)\mathbf{r}(\lambda) is irrational but not Diophantine. It is clear that n=12n!\sum_{n=1}^{\infty}2^{-n!}, which is a base 2 version of Liouville’s constant, cannot satisfy Definition 2.3 for any choice of constants cc or β\beta. The data presented is courtesy of Matthew Colbrook.
Theorem 2.

If Ω\Omega is an ellipse or if Ω\Omega is a rectangle with sides parallel to the coordinate axis, the spectrum of PP is pure point dense in [0,1][0,1] and the eigenvalues are exactly the values of λ\lambda for which the rotation number 𝐫(λ)\mathbf{r}(\lambda) is rational.

For the theorem above, we emphasize that the ellipses we consider can be any nondegenerate linear transformation of a circle. This includes tilted ellipses with major and minor axis not necessarily parallel to the coordinate axis. On the other hand, the theorem fails for tilted rectangle; it is shown in [Ral73] that the spectrum contains an absolutely continuous part.

1.1. Related works

Analysis of (1.1) goes back to [Sob54]. The spectral properties of the operator PP defined in (1.3) were studied in [Ale60] and [Ral73]. The study of internal waves has motivated the mathematical analysis of 0-th order self-adjoint pseudodifferential operators. Such operators on closed surfaces in the presence of attractors were studied in [CdVSR20, CdV20] and then in [DZ19]. The viscosity limits of these 0-th order pseudodifferential operators were recently studied by Galkowski–Zworski [GZ22] and Wang [Wan22]. Spectral properties of 0-th order pseudodifferential operators on the circle were studied by Zhongkai Tao [Tao19] who produced examples of embedded eigenvalues.

Then for the case of 2D planar domains, [DWZ21] showed that when the underlying dynamics has hyperbolic attractors, there can be high concentration of the fluid velocity near these attractors. This phenomenon was predicted in the physics literature by Maas–Lam [ML95] in 1995, and has since been experimentally observed by Maas et al. [MBSL97], Hazewinkel et al. [HTMD10], and Brouzet [Bro16]. We also consider 2D planar domains in our work, but the underlying dynamical assumption is different, thus leading to different conclusions.

2. Preliminaries

In this section, we establish the necessary geometric assumptions so that the underlying classical dynamics governing the system can be reduced to the the boundary of Ω\Omega. We then provide an outline of the necessary results from one-dimensional dynamics required to analyze our internal waves model.

2.1. Geometric assumptions

From (1.5), it is clear that we need to understand the spectral properties of PP at λ2\lambda^{2}. We consider the eigenvalue problem

P(λ)u=0whereP(λ):=(Pλ2)Δ=(1λ2)x22λ2x12.P(\lambda)u=0\quad\text{where}\quad P(\lambda):=(P-\lambda^{2})\Delta=(1-\lambda^{2})\partial_{x_{2}}^{2}-\lambda^{2}\partial_{x_{1}}^{2}. (2.1)

Clearly, P(λ):H01(Ω)H1(Ω)P(\lambda):H^{1}_{0}(\Omega)\to H^{-1}(\Omega) is invertible if and only if λ2\lambda^{2} is not in the spectrum of the operator PP defined in (1.4). The problem has no nontrivial solutions if λ>1\lambda>1, because then the operator P(λ)P(\lambda) is elliptic and the result follows from the maximum principle. There is also no solution for λ=0\lambda=0 or λ=1\lambda=1 by direct inspection. It is also known that the spectrum of PP is precisely [0,1][0,1], see [Ral73].

The advantage of working with P(λ)P(\lambda) is that it is simply a (1+1)(1+1)-dimensional wave operator, and the symbol is given by the quadratic form (1λ2)ξ22+λ2ξ12-(1-\lambda^{2})\xi_{2}^{2}+\lambda^{2}\xi_{1}^{2}. The relevant classical dynamics here is given by the Hamiltonian flow of the symbol. The dual of this quadratic form can be factorized as

x12λ2+x221λ2=+(x,λ)(x,λ),±(x,λ):=±x1λ+x21λ2.-\frac{x_{1}^{2}}{\lambda^{2}}+\frac{x_{2}^{2}}{1-\lambda^{2}}=\ell^{+}(x,\lambda)\ell^{-}(x,\lambda),\quad\ell^{\pm}(x,\lambda):=\pm\frac{x_{1}}{\lambda}+\frac{x_{2}}{\sqrt{1-\lambda^{2}}}. (2.2)

In particular, the integral curves of the Hamiltonian flow of the symbol projected onto Ω\Omega are precisely the level sets of ±(x,λ)\ell^{\pm}(x,\lambda).

Definition 2.1.

Let λ(0,1)\lambda\in(0,1). Then Ω\Omega is called λ\lambda-simple if each ±(,λ):Ω\ell^{\pm}(\bullet,\lambda):\partial\Omega\to\mathbb{R} has exactly two distinct critical points which are both non-degenerate. See Figure 3.

Refer to caption
Figure 3. Definition of the involutions γ±\gamma^{\pm} as well as the chess billiard map b(x)b(x). Also illustrates the definition of λ\lambda-simplicity, since ±\ell^{\pm} is nondegenerate at the critical points xmin±x_{\mathrm{min}}^{\pm} and xmax±x_{\mathrm{max}}^{\pm}. The diagram is from [DWZ21], which considered the same dynamical system.

Under the λ\lambda-simple condition, the dynamics can be reduced to the boundary. In particular, there exist unique smooth orientation-reversing involutions γ±(,λ):ΩΩ\gamma^{\pm}(\bullet,\lambda):\partial\Omega\to\partial\Omega that satisfy

±(x)=±(γ±(x)).\ell^{\pm}(x)=\ell^{\pm}(\gamma^{\pm}(x)). (2.3)

Composing γ+\gamma^{+} and γ\gamma^{-}, we obtain an orientation preserving diffeomorphism of the boundary Ω\partial\Omega. This is known as the chess billiard map b(,λ):ΩΩb(\bullet,\lambda):\partial\Omega\to\partial\Omega, defined by

b:=γ+γ.b:=\gamma^{+}\circ\gamma^{-}. (2.4)

See Figure 3. We will often suppress the dependence on λ\lambda in the notation when there is no ambiguity.

2.2. One-dimensional dynamics

Let 𝐱:𝕊1=/Ω\mathbf{x}:\mathbb{S}^{1}=\mathbb{R}/\mathbb{Z}\to\partial\Omega be a positively oriented parametrization on Ω\partial\Omega. This gives rise to a covering map π~:Ω\tilde{\pi}:\mathbb{R}\to\partial\Omega given by π~(x)=𝐱(xmod1)\tilde{\pi}(x)=\mathbf{x}(x\bmod 1). Let 𝐛(,λ):\mathbf{b}(\bullet,\lambda):\mathbb{R}\to\mathbb{R} be a lift of b(,λ)b(\bullet,\lambda), i.e. 𝐛(,λ)\mathbf{b}(\bullet,\lambda) satisfies

π(𝐛(x,λ))=b(π(x),λ)\pi(\mathbf{b}(x,\lambda))=b(\pi(x),\lambda)

for all xx\in\mathbb{R}. Fix an initial point x0x_{0}\in\mathbb{R}. The rotation number of b(,λ)b(\bullet,\lambda) is then defined as

𝐫(λ):=limk𝐛k(x0,λ)x0k/.\mathbf{r}(\lambda):=\lim_{k\to\infty}\frac{\mathbf{b}^{k}(x_{0},\lambda)-x_{0}}{k}\in\mathbb{R}/\mathbb{Z}. (2.5)

The rotation number is simply the averaged rotation along an orbit, and it is a well-defined quantity:

Lemma 2.2.

The rotation number 𝐫(λ)\mathbf{r}(\lambda) as defined by the limit (2.5) exists. Furthermore, it is independent of the choice of initial point x0x_{0}, the parametrization 𝐱\mathbf{x}, and the lift 𝐛\mathbf{b}.

See e.g. [Wal99] for a proof of Lemma 2.2. We refer the reader to [dMvS93, Chapter 1] for a thorough discussion of circle homeomorphisms.

We wish to understand to what extent a circle diffeomorphism bb can be conjugated to a rotation. This is clearly impossible for rational rotations, and it was proved by Denjoy in [Den32] that when the rotation number is irrational, the circle diffeomorphism is topologically conjugate to a rotation. We will need more regularity than just topologically conjugate, which is known when the rotation number satisfies the following Diophantine condition:

Definition 2.3.

A number α\alpha\in\mathbb{R} is Diophantine if there exists constants c,β>0c,\,\beta>0 so that

|αpq|cq2+β\left|\alpha-\frac{p}{q}\right|\geq\frac{c}{q^{2+\beta}}

for all pp\in\mathbb{Z} and qq\in\mathbb{N}.

Roughly speaking, this is a “sufficiently irrational” condition; α\alpha is Diophantine if it is sufficiently far away from all rational numbers. Diophantine numbers exist and form a full measure set. Indeed, take some β>0\beta>0 and a small c>0c>0, and consider the set of numbers in (0,1) that are Diophantine with respect to β\beta and cc:

E={α(0,1):|αpq|cq2+β for all p and q}.E=\left\{\alpha\in(0,1):\left|\alpha-\frac{p}{q}\right|\geq\frac{c}{q^{2+\beta}}\,\text{ for all $p\in\mathbb{Z}$ and $q\in\mathbb{N}$}\right\}.

We can think of EE as the set remaining after stripping away small neighborhoods of rational numbers, so the measure of EE is at least

|E|1q=12qcq2+β.|E|\geq 1-\sum_{q=1}^{\infty}2q\cdot\frac{c}{q^{2+\beta}}.

The right hand side becomes arbitrarily close to 11 as c0c\to 0.

If the rotation number is irrational, then the topological conjugation can be upgraded to smooth conjugation to a rotation. This upgrade is due to Herman [Her79] and Yoccoz [Yoc84]. We collect the conjugation results from [Den32, Her79, Yoc84] in the following proposition.

Proposition 2.4.

Let b:𝕊1𝕊1b:\mathbb{S}^{1}\to\mathbb{S}^{1} be a smooth circle diffeomorphism with rotation number α\alpha.

  1. (a)

    if α\alpha is irrational, then there exists a homeomorphism ψ:𝕊1𝕊1\psi:\mathbb{S}^{1}\to\mathbb{S}^{1} such that (ψbψ1)(θ)=θ+α(\psi\circ b\circ\psi^{-1})(\theta)=\theta+\alpha.

  2. (b)

    if α\alpha is Diophantine, then there exists a smooth diffeomorphism ψ:𝕊1𝕊1\psi:\mathbb{S}^{1}\to\mathbb{S}^{1} such that (ψbψ1)(θ)=θ+α(\psi\circ b\circ\psi^{-1})(\theta)=\theta+\alpha.

A standard consequence of the topological conjugacy result is the unique ergodicity of circle diffeomorphisms with irrational rotation number.

Corollary 2.5.

Let b:𝕊1𝕊1b:\mathbb{S}^{1}\to\mathbb{S}^{1} be a smooth circle diffeomorphism with irrational rotation number. Then for every fC0(𝕊1)f\in C^{0}(\mathbb{S}^{1}), there exists a constant function c(f)c(f) such that

1Nn=0N1(b)nfc(f)in C0(𝕊1) as N.\frac{1}{N}\sum_{n=0}^{N-1}(b^{*})^{n}f\to c(f)\quad\text{in $C^{0}(\mathbb{S}^{1})$ as $N\to\infty$}.

Consequently, if bf=fb^{*}f=f, then ff must be constant.

Proof.

By Proposition 2.4, it suffices to prove the theorem in the case that b(θ)=θ+αb(\theta)=\theta+\alpha for some irrational α(0,1)\alpha\in(0,1). Furthermore, it suffices to prove the statement for a dense subset of C0(𝕊1)C^{0}(\mathbb{S}^{1}), so we simply check the statement for trigonometric polynomials. Indeed, it is trivial for constant functions, and for k{0}k\in\mathbb{Z}\setminus\{0\},

1Nn=0N1e2πik(θ+nα)=1Ne2πikθ1e2πikαN1e2πikα0\frac{1}{N}\sum_{n=0}^{N-1}e^{2\pi ik(\theta+n\alpha)}=\frac{1}{N}e^{2\pi ik\theta}\frac{1-e^{2\pi ik\alpha N}}{1-e^{2\pi ik\alpha}}\to 0

uniformly in θ\theta since 1e2πikα01-e^{2\pi ik\alpha}\neq 0 by the irrationality of α\alpha. This proves the statement for all trigonometric polynomials as desired. ∎

Eventually, we will need to solve cohomological equations of the form

v(θ)v(θ+α)=g(θ)v(\theta)-v(\theta+\alpha)=g(\theta) (2.6)

given gCc(𝕊1)g\in C^{\infty}_{\mathrm{c}}(\mathbb{S}^{1}) with vanishing integral. If α\alpha satisfies the Diophantine condition, then we can have explicit high frequency control of vv. In particular, if gg is smooth, then we have smooth solutions.

Lemma 2.6.

Let gC(𝕊1)g\in C^{\infty}(\mathbb{S}^{1}) with 𝕊1g=0\int_{\mathbb{S}^{1}}g=0, and assume that α\alpha is Diophantine. Then there exists vC(𝕊1)v\in C^{\infty}(\mathbb{S}^{1}) that solves (2.6) uniquely modulo constant functions.

Proof.

Let c,β>0c,\beta>0 be the Diophantine constants associated with α\alpha in Definition 2.3. Taking the Fourier series of both sides of (2.6),

g^(k)=(1e2πikα)v^(k).\hat{g}(k)=(1-e^{2\pi ik\alpha})\hat{v}(k).

For all k0k\neq 0,

|11e2πikα|c1|k|1+β.\left|\frac{1}{1-e^{2\pi ik\alpha}}\right|\leq c^{-1}|k|^{1+\beta}.

Since g^(0)=0\hat{g}(0)=0, vv can be solved uniquely up to choice of v^(0)\hat{v}(0), and we have the estimate

vHsCgHs+β+1\|v\|_{H^{s}}\leq C\|g\|_{H^{s+\beta+1}}

for every s>0s>0. Therefore vv is smooth and unique modulo constant functions. ∎

3. Internal waves in an ergodic setting

In this section we prove Theorem 1. The strategy is to construct a right inverse to P(λ)P(\lambda) on smooth functions for λ\lambda such that 𝐫(λ)\mathbf{r}(\lambda) is Diophantine (Definition 2.3). Define coordinates

y±=12±(x,λ)y_{\pm}=\frac{1}{2}\ell^{\pm}(x,\lambda) (3.1)

where ±\ell^{\pm} are defined in (2.2). In these coordinates, we have

P(λ)=2y+y.P(\lambda)=\frac{\partial^{2}}{\partial y_{+}\partial y_{-}}.

In this form, it is clear that solutions on the interior of Ω\Omega to the eigenvalue problem (2.1) can be found by integration up to boundary conditions, which effectively reduces the problem to the boundary.

3.1. Injectivity of the eigenvalue problem

We first show that solutions to the eigenvalue problem take a very convenient form.

Lemma 3.1.

Let Ω\Omega be λ\lambda-simple and suppose uH01(Ω)u\in H^{1}_{0}(\Omega) is a solution to the eigenvalue problem (2.1). Then using the coordinates (3.1), uu can be decomposed as

u(y+,y)=u+(y+)+u(y)u(y_{+},y_{-})=u_{+}(y_{+})+u_{-}(y_{-})

where u±H1(Ω)u_{\pm}\in H^{1}(\Omega) are functions on Ω\Omega that depend only on y±y_{\pm} respectively. In fact, u±C0(Ω¯)u_{\pm}\in C^{0}(\overline{\Omega}).

Proof.

Let u+:=y+uu^{\prime}_{+}:=\partial_{y_{+}}u. Observe that yu+=0\partial_{y_{-}}u^{\prime}_{+}=0, so u+u^{\prime}_{+} is a function of y+y_{+} only. The eigenvalue problem (2.1) is invariant under shifting of the domain, so we may assume without the loss of generality that the two nondegenerate critical points of +\ell^{+} are y+=0y_{+}=0 and y+=ay_{+}=a. Note that u+L2(Ω)u^{\prime}_{+}\in L^{2}(\Omega), so the nondegeneracy of the critical points imply that

01y+(ay+)|u+(y+)|2𝑑y+<.\int_{0}^{1}\sqrt{y_{+}(a-y_{+})}|u^{\prime}_{+}(y_{+})|^{2}\,dy_{+}<\infty.

By Cauchy-Schwarz, we then have

u+L1([0,a])(y+(1y+))1/4u+L2([0,a])(y+(1y+))1/4L2([0,a])<\|u^{\prime}_{+}\|_{L^{1}([0,a])}\leq\|(y_{+}(1-y_{+}))^{1/4}u^{\prime}_{+}\|_{L^{2}([0,a])}\|(y_{+}(1-y_{+}))^{-1/4}\|_{L^{2}([0,a])}<\infty

Put

u~+(y+):=0y+u+(s)𝑑s\tilde{u}_{+}(y_{+}):=\int_{0}^{y_{+}}u^{\prime}_{+}(s)\,ds

By a slight abuse of notation, we can view u~+\tilde{u}_{+} as an element of C0(Ω)C^{0}(\Omega) that depends only on y+y_{+}. Then we see that uu~+H1u-\tilde{u}_{+}\in H^{1} is a function that depends only on yy_{-}.

Running the same argument with yy_{-} instead of y+y_{+}, we also have u~C0(Ω)\tilde{u}_{-}\in C^{0}(\Omega) depending only on yy_{-} such that uu~u-\tilde{u}_{-} depends only on y+y_{+}. Modifying u~±\tilde{u}_{\pm} by constants, there must exist u±C0(Ω)u_{\pm}\in C^{0}(\Omega) that depends only on y±y_{\pm} such that u++u=0u_{+}+u_{-}=0. ∎

Proof of Theorem 1(a).

We prove that the operator P(λ)P(\lambda) has trivial kernel. Suppose uu is a solution the eigenvalue problem (2.1). Then we have the decomposition u(y+,y)=u+(y+)+u(y)u(y_{+},y_{-})=u_{+}(y_{+})+u_{-}(y_{-}) from Lemma 3.1. Define the boundary traces

U±:=u±|ΩC0(Ω).U_{\pm}:=u_{\pm}|_{\partial\Omega}\in C^{0}(\Omega). (3.2)

Observe that U++U=0U_{+}+U_{-}=0 and U±γ±=U±U_{\pm}\circ\gamma_{\pm}=U_{\pm}, so it follows that U±U_{\pm} are invariant under pullback by the chess-billiard map:

U±=bU±.U_{\pm}=b^{*}U_{\pm}. (3.3)

If the rotation number ρ(λ)\rho(\lambda) of b(,λ)b(\bullet,\lambda) is irrational, it follows from (3.3) and Corollary 2.5 that the functions U±U_{\pm} must be constant. The solution uu can then be recovered from the boundary data U±U_{\pm}, and we clearly have u=0u=0. ∎

3.2. Energy boundedness

We first construct the right inverse on smooth functions to P(λ)P(\lambda) when the rotation number defined in (2.5) is Diophantine.

Proposition 3.2.

Let Ω\Omega be λ\lambda-simple and assume that the rotation number 𝐫(λ)\mathbf{r}(\lambda) is Diophantine. Then there exists R(λ):C(Ω¯)C(Ω¯)H01(Ω)R(\lambda):C^{\infty}(\overline{\Omega})\to C^{\infty}(\overline{\Omega})\cap H^{1}_{0}(\Omega) such that P(λ)R(λ)=IdP(\lambda)R(\lambda)=\operatorname{Id}.

Proof.

1. Using the coordinates y±y_{\pm} defined in (3.1), we wish to solve

2uy+y=f,fC(Ω¯)\frac{\partial^{2}u}{\partial y_{+}\partial y_{-}}=f,\quad f\in C^{\infty}(\overline{\Omega}) (3.4)

for uH01(Ω)u\in H_{0}^{1}(\Omega). Let f~Cc(2)\tilde{f}\in C^{\infty}_{\mathrm{c}}(\mathbb{R}^{2}) be an extension of ff, i.e. f~|Ω=f\tilde{f}|_{\Omega}=f. Define

u0(y+,y):=y+yf~(η+,η)𝑑η𝑑η+,andU0:=u0|ΩC(Ω).u_{0}(y_{+},y_{-}):=\int_{-\infty}^{y_{+}}\int_{-\infty}^{y_{-}}\tilde{f}(\eta_{+},\eta_{-})\,d\eta_{-}d\eta_{+},\quad\text{and}\quad U_{0}:=u_{0}|_{\partial\Omega}\in C^{\infty}(\partial\Omega).

Therefore, it suffices to solve

2vy+y=0,v|Ω=U0\frac{\partial^{2}v}{\partial y_{+}\partial y_{-}}=0,\quad v|_{\partial\Omega}=U_{0} (3.5)

for vC(Ω)v\in C^{\infty}(\Omega) since u=u0vu=u_{0}-v would then be the solution to (3.4).

2. By Proposition 2.4, we may choose smooth coordinates θ\theta on the boundary Ω\partial\Omega so that b(θ)=θ+𝐫(λ)b(\theta)=\theta+\mathbf{r}(\lambda). We claim that

ΩU0(γ±(θ))U0(θ)dθ=0.\int_{\partial\Omega}U_{0}(\gamma^{\pm}(\theta))-U_{0}(\theta)\,d\theta=0. (3.6)

Recall by (2.4), we have bγ+=γ+b1b\circ\gamma^{+}=\gamma^{+}\circ b^{-1}. In the coordinate θ\theta this gives γ+(θ)+𝐫(λ)=γ+(θ𝐫(λ))\gamma^{+}(\theta)+\mathbf{r}(\lambda)=\gamma^{+}(\theta-\mathbf{r}(\lambda)). Differentiating in θ\theta, we get

θγ+(θ)=θγ+(θ𝐫(λ)).\partial_{\theta}\gamma^{+}(\theta)=\partial_{\theta}\gamma^{+}(\theta-\mathbf{r}(\lambda)).

Since 𝐫(λ)\mathbf{r}(\lambda) is irrational and γ+\gamma^{+} is an orientation reversing involution, we see that θγ+(θ)=1\partial_{\theta}\gamma^{+}(\theta)=-1. Similarly, θγ(θ)=1\partial_{\theta}\gamma^{-}(\theta)=-1, hence (3.6) follows.

3. Therefore, it follows from Lemma 2.6 that there exist unique V±C(Ω)V_{\pm}\in C^{\infty}(\partial\Omega) such that

(b±1)V±V±=(γ)U0U0andΩV±(θ)𝑑θ=12ΩU0(θ)𝑑θ.(b^{\pm 1})^{*}V_{\pm}-V_{\pm}=(\gamma^{\mp})^{*}U_{0}-U_{0}\quad\text{and}\quad\int_{\partial\Omega}V_{\pm}(\theta)\,d\theta=\frac{1}{2}\int_{\partial\Omega}U_{0}(\theta)\,d\theta. (3.7)

Observe that

(V++V)(b1)(V++V)\displaystyle(V_{+}+V_{-})-(b^{-1})^{*}(V_{+}+V_{-}) =(b1)((γ)U0U0)((γ+)U0U0)\displaystyle=(b^{-1})^{*}((\gamma^{-})^{*}U_{0}-U_{0})-((\gamma^{+})^{*}U_{0}-U_{0})
=U0(b1)U0.\displaystyle=U_{0}-(b^{-1})^{*}U_{0}.

Solutions to the cohomological equation are unique up to constants, and it follows from (3.7) that V++V=U0\int V_{+}+V_{-}=\int U_{0}, so we must have

V++V=U0.V_{+}+V_{-}=U_{0}. (3.8)

Furthermore, applying γ\gamma^{\mp} to both sides of (3.7), we see that

((γ±)V±)(b±1)((γ±)V±)=U0(γ)U0.((\gamma^{\pm})^{*}V_{\pm})-(b^{\pm 1})^{*}((\gamma^{\pm})^{*}V_{\pm})=U_{0}-(\gamma^{\mp})^{*}U_{0}.

Adding this to (3.6),

[(γ±)V±V±](b±1)[(γ±)V±V±]=0.[(\gamma^{\pm})^{*}V_{\pm}-V_{\pm}]-(b^{\pm 1})^{*}[(\gamma^{\pm})^{*}V_{\pm}-V_{\pm}]=0.

Again using Lemma 2.6, we find that

(γ±)V±=V±.(\gamma^{\pm})^{*}V_{\pm}=V_{\pm}. (3.9)

By the λ\lambda-simplicity assumption, the coordinate functions restricted to boundary, y±|Ωy_{\pm}|_{\partial\Omega}, have nondegenerate critical points. In particular, up to a smooth change of coordinates, we may assume that y+=0y_{+}=0 is a critical point of y+|Ωy_{+}|_{\partial\Omega}, near which the boundary can be parameterized by yy_{-} and is given by {y+=y2}\{y_{+}=y_{-}^{2}\}. We may further assume that the γ+\gamma^{+}-invariance of V+V_{+} from (3.9) in these coordinates reads V+(y)=V+(y)V_{+}(y_{-})=V_{+}(-y_{-}) near the critical point. Therefore V+V_{+} is a smooth function of y+=y2y_{+}=y_{-}^{2} near the critical point. Similar analysis holds for all other critical points, so, there exists v±C(Ω¯)v_{\pm}\in C^{\infty}(\overline{\Omega}) such that v±v_{\pm} depends only on y±y_{\pm} and v±|Ω=V±v_{\pm}|_{\partial\Omega}=V_{\pm}. Then v=v++vv=v_{+}+v_{-} solves (3.5) and the boundary conditions are satisfied due to (3.8). ∎

The obstruction to energy boundedness in the functional calculus solution (1.5) is the singularity at z=λ2z=\lambda^{2} that appears as tt\to\infty. This can be cancelled out using the right inverse to P(λ)P(\lambda) constructed in the previous proposition.

Proof of Theorem 1(b).

Let fC(Ω¯;)f\in C^{\infty}(\overline{\Omega};\mathbb{R}) and let R(λ)R(\lambda) be as in Proposition 3.2. Then g:=ΔΩR(λ)fg:=\Delta_{\Omega}R(\lambda)f lies in C(Ω¯)H1(Ω)C^{\infty}(\overline{\Omega})\subset H^{-1}(\Omega). Then the evolution problem (1.4) can be rewritten as

(t2+P)w=(Pλ2)gcosλt,w|t=0=tw|t=0=0,u=ΔΩ1w.(\partial_{t}^{2}+P)w=(P-\lambda^{2})g\cos\lambda t,\quad w|_{t=0}=\partial_{t}w|_{t=0}=0,\quad u=\Delta_{\Omega}^{-1}w.

Using the functional calculus solution formula for w(t)w(t) given in (1.5), we have

w(t)=(cos(tP)cos(tλ))g.w(t)=(\cos(t\sqrt{P})-\cos(t\lambda))g. (3.10)

Since |cos(tz)cos(tλ)|2\left|\cos(t\sqrt{z})-\cos(t\lambda)\right|\leq 2 for all z[0,1]z\in[0,1] and tt\in\mathbb{R}, it follows from the spectral theorem that w(t)w(t) given by (3.10) is uniformly bounded in H1(Ω)H^{-1}(\Omega) for all tt. Therefore the solution to the internal waves equation (1.1) given by u(t)=ΔΩ1w(t)u(t)=\Delta_{\Omega}^{-1}w(t) is uniformly bounded in H01(Ω)H^{1}_{0}(\Omega) for all tt. ∎

Remark.  In fact, it is easy to see that smoothness of ff is not required. It suffices to have fHN(Ω)f\in H^{N}(\Omega) for sufficiently large N>0N>0 depending on the constants in Definition 2.3 of Diophantine numbers. The proof of Lemma 2.6 gives explicit estimates for the regularity of the solution to the cohomological equation. This means that the boundary traces V±V_{\pm} from Proposition 3.2 lies in HNk(Ω)H^{N-k}(\partial\Omega) for some kk depending on the Diophantine constants. Since the critical points of the coordinate functions y±y_{\pm} restricted to the boundary are nondegenerate and V±V_{\pm} are invariant under pullback by γ±\gamma^{\pm} respectively, it follows that v±HN/2k(Ω)v_{\pm}\in H^{N/2-k}(\Omega) for a possibly different kk depending only on the Diophantine constants.

3.3. Spectral estimate near λ2\lambda^{2}

We now use the right inverse R(λ)R(\lambda) in Proposition 3.2 to obtain bounds on the spectral measure near λ\lambda.

Proof of Theorem 1(c).

Let fC(Ω¯)H1(Ω)f\in C^{\infty}(\overline{\Omega})\subset H^{-1}(\Omega). The spectral measure dμfd\mu_{f} satisfies

φ𝑑μf=φ(P)f,f.\int\varphi\,d\mu_{f}=\langle\varphi(P)f,f\rangle.

Put fk=(ΔΩR(λ))kff_{k}=(\Delta_{\Omega}R(\lambda))^{k}f. Note that fkC(Ω¯)f_{k}\in C^{\infty}(\overline{\Omega}) and (Pλ2)kfk=f(P-\lambda^{2})^{k}f_{k}=f. Therefore,

|λ2ελ2+ε𝑑μf|=|λ2ελ2+ε(xλ2)2k𝑑μfk(x)|Ckε2k\left|\int_{\lambda^{2}-\varepsilon}^{\lambda^{2}+\varepsilon}\,d\mu_{f}\right|=\left|\int_{\lambda^{2}-\varepsilon}^{\lambda^{2}+\varepsilon}(x-\lambda^{2})^{2k}\,d\mu_{f_{k}}(x)\right|\leq C_{k}\varepsilon^{2k}

as desired. ∎

4. Examples

In this section we give the explicit computations related to the spectrum of the square and the disk, from which Theorem 2 follows immediately. Note that the square does not satisfy the hypothesis of Theorem 1 since the boundary is not smooth. Nevertheless, the conclusions of Theorem 1 holds, and we verify this directly.

4.1. The square

We consider the square domain Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1]. Clearly the Fourier modes provide a basis of eigenfunctions and it is easy to see that the eigenvalues are dense in [0,1][0,1], so Theorem 2 for the square is clear.

The square thus has the advantage that (1.1) can be solved directly in Fourier series, so we can verify the contents of Theorem 1 for the square domain, despite it having corners.

The chess billiard flow on the square is the same as the standard billiard flow, so the rotation number function 𝐫(λ)\mathbf{r}(\lambda) defined in (2.5) is smooth and can be written down explicitly. It is given by

𝐫(λ)=λ1λ2+λ.\mathbf{r}(\lambda)=\frac{\lambda}{\sqrt{1-\lambda^{2}}+\lambda}.

See [Zhu22] for the full derivation. We can formally write

u(t,x)=𝐤2u^(t,𝐤)sin(πk1x1)sin(πk2x2)u(t,x)=\sum_{\mathbf{k}\in\mathbb{N}^{2}}\hat{u}(t,\mathbf{k})\sin(\pi k_{1}x_{1})\sin(\pi k_{2}x_{2})

where 𝐤=(k1,k2)2\mathbf{k}=(k_{1},k_{2})\in\mathbb{N}^{2}. Only in this subsection, we use the hat to denote Fourier transform with respect to the Dirichlet sine basis. If u(t,x)u(t,x) is a solution to (1.1), the coefficients must satisfy the periodically driven harmonic oscillator equation

π2(k12+k22)t2u^(t,𝐤)+π2k22u^(t,𝐤)=f^(𝐤)cosλt,wheref^(𝐤)=[0,1]2f(x)sin(πk1x1)sin(πk2x2)𝑑x1𝑑x2\begin{gathered}-\pi^{2}(k_{1}^{2}+k_{2}^{2})\partial_{t}^{2}\hat{u}(t,\mathbf{k})+\pi^{2}k_{2}^{2}\hat{u}(t,\mathbf{k})=\hat{f}(\mathbf{k})\cos\lambda t,\\ \text{where}\quad\hat{f}(\mathbf{k})=\int_{[0,1]^{2}}f(x)\sin(\pi k_{1}x_{1})\sin(\pi k_{2}x_{2})\,dx_{1}dx_{2}\end{gathered} (4.1)

This has solution

u^(t,𝐤)=f^(𝐤)λ2k12(1λ2)k22[cos(λt)cos(k2|𝐤|t)]\hat{u}(t,\mathbf{k})=\frac{\hat{f}(\mathbf{k})}{\lambda^{2}k_{1}^{2}-(1-\lambda^{2})k_{2}^{2}}\left[\cos(\lambda t)-\cos\left(\frac{k_{2}}{|\mathbf{k}|}t\right)\right] (4.2)

If 𝐫(λ)\mathbf{r}(\lambda) is Diophantine, then the result of Theorem 1 holds. In fact, in this case, we get that u(t)C(Ω)u(t)\in C^{\infty}(\Omega) uniformly in all the seminorms for all tt\in\mathbb{R}. Indeed, if 𝐫(λ)\mathbf{r}(\lambda) is Diophantine, then there exist constants c,β>0c,\,\beta>0 such that

|q𝐫(λ)p|cq1+β|q\cdot\mathbf{r}(\lambda)-p|\geq\frac{c}{q^{1+\beta}}

for any pp\in\mathbb{Z} and qq\in\mathbb{N}. Rewriting this condition in terms of λ\lambda, we find that

|(qp)λp1λ2|cq1+β|(q-p)\lambda-p\sqrt{1-\lambda^{2}}|\geq\frac{c}{q^{1+\beta}}

where cc is a possibly different constant. Put q=k1+k2q=k_{1}+k_{2} and p=k2p=k_{2} to find that

|k1λk21λ2|cq1+β1|𝐤|1+β.|k_{1}\lambda-k_{2}\sqrt{1-\lambda^{2}}|\geq\frac{c}{q^{1+\beta}}\gtrsim\frac{1}{|\mathbf{k}|^{1+\beta}}. (4.3)

where the hidden constant is independent of 𝐤\mathbf{k}. We clearly have

|k1λ+k21λ2|1,|k_{1}\lambda+k_{2}\sqrt{1-\lambda^{2}}|\geq 1, (4.4)

so combining (4.3) and (4.4) yields

|λ2k12(1λ2)k22|c|𝐤|1+β.|\lambda^{2}k_{1}^{2}-(1-\lambda^{2})k_{2}^{2}|\geq\frac{c}{|\mathbf{k}|^{1+\beta}}. (4.5)

for a possibly different constant c>0c>0. Since fCc(Ω;)f\in C^{\infty}_{\mathrm{c}}(\Omega;\mathbb{R}), f^(𝐤)\hat{f}(\mathbf{k}) is rapidly decreasing in 𝐤\mathbf{k}, which tempers the denominator in (4.2) to give uniform smoothness of u(t)u(t) in time.

The above analysis also exhibits the spectral result part (c) of Theorem 1. Note that κ(𝐤)sin(πk1x1)sin(πk2x2)\kappa(\mathbf{k})\sin(\pi k_{1}x_{1})\sin(\pi k_{2}x_{2}), 𝐤2\mathbf{k}\in\mathbb{N}^{2} form a complete orthonormal basis for H1(Ω)H^{-1}(\Omega), where κ(𝐤)=𝒪(|k|)\kappa(\mathbf{k})=\mathcal{O}(|k|) are simply normalizing constants. This basis consists of eigenfunctions with eigenvalues k1k1+k2\frac{k_{1}}{k_{1}+k_{2}} for the operator PP defined in (1.4). In particular, sin(πk1x1)sin(πk2x2)\sin(\pi k_{1}x_{1})\sin(\pi k_{2}x_{2}) has eigenvalue k22k12+k22\frac{k_{2}^{2}}{k_{1}^{2}+k_{2}^{2}}. If 𝐫(λ)\mathbf{r}(\lambda) is Diophantine, then (4.5) gives a characterization of the eigenvalues near λ2\lambda^{2}:

|k22k12+k22λ2|ε1|𝐤|3+βε1\left|\frac{k_{2}^{2}}{k_{1}^{2}+k_{2}^{2}}-\lambda^{2}\right|\leq\varepsilon\implies\frac{1}{|\mathbf{k}|^{3+\beta}}\gtrsim\varepsilon^{-1}

Therefore the spectral measure μf,f\mu_{f,f} satisfies part (c) of Theorem 1 near λ2\lambda^{2}. Indeed, using the fact that the coefficients f^(k)\hat{f}(k) of fCc(Ω)f\in C^{\infty}_{\mathrm{c}}(\Omega) defined in (4.1) are rapidly decreasing, we have

μf,f((λ2ε,λ2+ε))𝐤:|k2k12+k22λ2|ε1f^(𝐤)2κ(𝐤)2<Cdεd\mu_{f,f}((\lambda^{2}-\varepsilon,\lambda^{2}+\varepsilon))\lesssim\sum_{\mathbf{k}:\big{|}\frac{k_{2}}{k_{1}^{2}+k_{2}^{2}}-\lambda^{2}\big{|}\leq\varepsilon^{-1}}\hat{f}(\mathbf{k})^{2}\kappa(\mathbf{k})^{-2}<C_{d}\varepsilon^{d}

for any dd\in\mathbb{N}. See Figure 2.

Finally, we mention that by tilting the square by η\eta, the set of λ\lambda for which λ\mathbf{\lambda} is Diophantine is no longer a full measure set. More specifically, we consider the square domain specified by the vertices

{(0,0),(cosη,sinη),(sinη,cosη),2(cos(η+π4),sin(η+π4))}.\big{\{}(0,0),\,(\cos\eta,\sin\eta),\,(-\sin\eta,\cos\eta),\,\sqrt{2}(\cos(\eta+\tfrac{\pi}{4}),\sin(\eta+\tfrac{\pi}{4}))\big{\}}.

Then graph of 𝐫(λ)\mathbf{r}(\lambda) is constant near values of λ\lambda for which 𝐫(λ)\mathbf{r}(\lambda) is rational. See Figure 4 for an illustration and [DWZ21, §2.5] for details.

Refer to caption
Refer to caption
Figure 4. Graphs of the rotation number function 𝐫(λ)\mathbf{r}(\lambda) (defined in (2.5)) for various domains. Left: The blue curve is for the untilted square [0,1]2[0,1]^{2}. The orange curve is the square tilted by η=π/20\eta=\pi/20. Note that it has rational plateaus. Right: The domain is a disk, so the boundary is a round circle.

4.2. The disk

We finally consider the case when Ω=𝔻\Omega=\mathbb{D} is the unit disk. This has previously been studied in [Ale60, Section 9]. The 3-dimensional case of a triaxial ellipsoid has been studied in [CdVV23], following the physics work of [IJW15, BR17].

We may assume 𝔻\mathbb{D} is centered at origin. Parameterize the boundary 𝔻\partial\mathbb{D} counterclockwise by arclength with ϕ=0\phi=0 being the point (1,0)(1,0). The boundary is then identified with the circle /2π\mathbb{R}/2\pi\mathbb{Z}. This is a different convention from the previous sections, and we switch conventions to avoid carrying factors of 2π2\pi. The mod 1 rotation number as defined in 2.5 is given by

𝐫(λ)=12πarccos(λ),\mathbf{r}(\lambda)=1-\frac{2}{\pi}\arccos(\lambda), (4.6)

see [Li23, §2.2.2] for details and Figure 4. It will be convenient to rescale the rotation number and define

α(λ)=π2𝐫(λ).\alpha(\lambda)=\frac{\pi}{2}\mathbf{r}(\lambda). (4.7)

We henceforth drop λ\lambda from the notation when there is no ambiguity. The terms of the factorization (2.2) and the chess billiard map take the explicit forms

±(x1,x2)=x1cosα±x2sinα,b(ϕ)=ϕ+4α.\ell^{\pm}(x_{1},x_{2})=x_{1}\cos\alpha\pm x_{2}\sin\alpha,\qquad b(\phi)=\phi+4\alpha. (4.8)

We will provide an explicit complete basis of eigenfunctions. Recall from (2.2) that +(x,λ)(x,λ)\ell^{+}(x,\lambda)\ell^{-}(x,\lambda) is dual to the symbol of P(λ)P(\lambda), so uH01(Ω)u\in H^{1}_{0}(\Omega) is a solution to the eigenvalue problem if and only if

u(x)=u+(+(x))+u((x))andU++U=0whereU+:=(u++)|Ω,U:=(u)|Ω.\begin{gathered}u(x)=u_{+}(\ell^{+}(x))+u_{-}(\ell^{-}(x))\quad\text{and}\\ U_{+}+U_{-}=0\quad\text{where}\quad U_{+}:=(u_{+}\circ\ell^{+})|_{\partial\Omega},\quad U_{-}:=(u_{-}\circ\ell^{-})|_{\partial\Omega}.\end{gathered} (4.9)

for some u±C0u_{\pm}\in C^{0} and u±±H1(Ω)u_{\pm}\circ\ell^{\pm}\in H^{1}(\Omega). See Lemma 3.1 for discussion of the regularity of u±u_{\pm}. Using the explicit formulas (4.8), we can compute

U±(ϕ)=u±(cosϕcosα±sinϕsinα)=u±(cos(ϕα)),U±(ϕ)=U±(ϕ+4α)\begin{gathered}U_{\pm}(\phi)=u_{\pm}(\cos\phi\cos\alpha\pm\sin\phi\sin\alpha)=u_{\pm}(\cos(\phi\mp\alpha)),\\ U_{\pm}(\phi)=U_{\pm}(\phi+4\alpha)\end{gathered} (4.10)

Let NN\in\mathbb{N} and k{1,,N1}k\in\{1,\dots,N-1\}. Then define

αk,N:=α(λk,N)=π2kNwhereλk,N(0,1)uniquely satisfies𝐫(λk,N)=kN.\alpha_{k,N}:=\alpha(\lambda_{k,N})=\frac{\pi}{2}\frac{k}{N}\quad\text{where}\quad\lambda_{k,N}\in(0,1)\quad\text{uniquely satisfies}\quad\mathbf{r}(\lambda_{k,N})=\frac{k}{N}.

We wish to construct a solution to every eigenvalue problem P(λk,N)u=0P(\lambda_{k,N})u=0. Collecting the symmetries in (4.9) and (4.10), we must have

U+(ϕ+2αk,N)=U+(ϕ)=U+(ϕ+2αk,N),U=U+.U_{+}(-\phi+2\alpha_{k,N})=U_{+}(\phi)=U_{+}(-\phi+2\alpha_{k,N}),\quad U_{-}=-U_{+}.

U+U_{+} is thus invariant under the action of a dihedral group, so we consider the Fourier modes on an interval of the boundary which is a fundamental domain, and we are led to functions of the form

U±(ϕ)=(±1)k+1cos(N(ϕαk,N)),U_{\pm}(\phi)=(\pm 1)^{k+1}\cos(N(\phi\mp\alpha_{k,N})),

which precisely satisfies the conditions given in (4.10). Let TNT_{N} be the Chebyshev polynomials, which are defined by TN(cosϕ)=cos(Nϕ)T_{N}(\cos\phi)=\cos(N\phi). Therefore, we have solutions to the eigenvalue problem (2.1) at λ=λk,N\lambda=\lambda_{k,N} given by

uk,N(x1,x2)=TN(x1cosαk,N+x2sinαk,N)(1)kTN(x1cosαk,Nx2sinαk,N)for everyN,k{1,,N1}.\begin{gathered}u_{k,N}(x_{1},x_{2})=T_{N}(x_{1}\cos\alpha_{k,N}+x_{2}\sin\alpha_{k,N})-(-1)^{k}T_{N}(x_{1}\cos\alpha_{k,N}-x_{2}\sin\alpha_{k,N})\\ \text{for every}\quad N\in\mathbb{N},\quad k\in\{1,\dots,N-1\}.\end{gathered} (4.11)

The above forms a complete basis of H01(𝔻)H_{0}^{1}(\mathbb{D}) since for every NN, the set {uk,N}k=1N1\{u_{k,N}\}_{k=1}^{N-1} consists of N1N-1 linearly independent degree NN polynomials vanishing at the boundary. Linear independence follows immediately from the fact that Δuk,N\Delta u_{k,N} is an eigenfunction of PP with eigenvalue λk,N\lambda_{k,N}. See Figure 1. Furthermore, we see that there are infinitely many solutions to the eigenvalue problem (2.1) for every λ\lambda such that 𝐫(λ)\mathbf{r}(\lambda) is rational, since there are infinitely many ways to represent a rational number as k/Nk/N.

The explicit formulas (4.11) for the solutions of the eigenvalue problem directly completes the proof of Theorem 2 for the unit disk. The proof of Theorem 2 for ellipses and rectangles is now a purely geometric problem.

Proof of Theorem 2.

Let Ω\Omega be an ellipse. Then there exists a symmetric nondegenerate 2×22\times 2 matrix AA and v2v\in\mathbb{R}^{2} such that Ω=A𝔻+v\Omega=A\mathbb{D}+v where 𝔻\mathbb{D} is the unit disk. Then observe that there exists a family of rotation matrices R(λ)R(\lambda) that depends smoothly on λ\lambda such that under the coordinate change

y=R(λ)A1(xv),y=R(\lambda)A^{-1}(x-v),

P(λ)P(\lambda) on Ω\Omega becomes c(λ)P(σ(λ))c(\lambda)P(\sigma(\lambda)) on 𝔻\mathbb{D}, for some c(λ)>0c(\lambda)>0 and σ:[0,1][0,1]\sigma:[0,1]\to[0,1] smooth and monotonically increasing with σ(0)=0\sigma(0)=0 and σ(1)=1\sigma(1)=1. The result then follows from the explicit basis of eigenfunctions corresponding to eigenvalues dense in (0,1)(0,1) given in (4.11) for the circle.

The rectangle case follows from the square case simply by a linear change of coordinates by a diagonal matrix independent of λ\lambda. ∎

Acknowledgements. The authors would like to thank Semyon Dyatlov, Maciej Zworski, and Leo Maas for insightful discussions. They would also like to thank Matthew Colbrook for providing the numerical data used in Figure 2. The second author is partially supported by Semyon Dyatlov’s NSF CAREER grant DMS-1749858.

References

  • [Ale60] R. A. Aleksandryan. Spectral properties of operators arising from systems of differential equations of Sobolev type. Tr. Mosk. Mat. Obs., 9:455–505, 1960.
  • [Arn61] V. I. Arnol’d. Small denominators I. mapping the circle onto itself. Izv. Akad. Nauk SSSR Ser. Mat., 25:21–86, 1961.
  • [BR17] G. Backus and M. Rieutord. Completeness of inertial modes of an incompressible inviscid fluid in a corotating ellipsoid. Phys. Rev. E, 95:053116, May 2017.
  • [Bro16] C. Brouzet. Internal wave attractors : from geometrical focusing to non-linear energy cascade and mixing. Theses, Université de Lyon, July 2016.
  • [CdV20] Y. Colin de Verdière. Spectral theory of pseudodifferential operators of degree 0 and an application to forced linear waves. Analysis & PDE, 13(5):1521–1537, jul 2020.
  • [CdVSR20] Y. Colin de Verdière and L. Saint-Raymond. Attractors for two-dimensional waves with homogeneous hamiltonians of degree 0. Communications on Pure and Applied Mathematics, 73(2):421–462, 2020.
  • [CdVV23] Y. Colin de Verdière and J. Vidal. The spectrum of the poincaré operator in an ellipsoid, 2023. Preprint; arXiv:2305.01369.
  • [CHT21] M. Colbrook, A. Horning, and A. Townsend. Computing spectral measures of self-adjoint operators. SIAM Review, 63(3):489–524, 2021.
  • [Col21] M. Colbrook. Computing spectral measures and spectral types. Commun. Math. Phys., 384:433–501, 2021.
  • [Col22] M. Colbrook. Specsolve: Spectral methods for spectral measures, 2022. Preprint; arXiv:2201.01314.
  • [Den32] A. Denjoy. Sur les courbes definies par les équations différentielles à la surface du tore. Journal de Mathématiques Pures et Appliqués, 11:333–376, 1932.
  • [dMvS93] W. de Melo and S. van Strien. One-Dimensional Dynamics. Springer Berlin, Heidelberg, 1993.
  • [DWZ21] S. Dyatlov, J. Wang, and M. Zworski. Mathematics of internal waves in a 2d aquarium, 2021. Preprint; arXiv:2112.10191.
  • [DZ19] S. Dyatlov and M. Zworski. Microlocal analysis of forced waves. Pure and Applied Analysis, 1(3):359–384, jul 2019.
  • [GZ22] J. Galkowski and M. Zworski. Viscosity limits for zeroth-order pseudodifferential operators. Communications on Pure and Applied Mathematics, 75(8):1798–1869, 2022.
  • [Her79] M. R. Herman. Sur la conjugaison differentiable des difféomorphismes du cercle a des rotations. Publ. Math. I.H.E.S., 49:5–234, 1979.
  • [HTMD10] J. Hazewinkel, C. Tsimitri, L. R. M. Maas, and S. B. Dalziel. Observations on the robustness of internal wave attractors to perturbations. Physics of Fluids, 22(10):107102, 2010.
  • [IJW15] D. J. Ivers, A. Jackson, and D. Winch. Enumeration, orthogonality and completeness of the incompressible coriolis modes in a sphere. Journal of Fluid Mechanics, 766:468–498, 2015.
  • [Joh41] F. John. The dirichlet problem for a hyperbolic equation. American Journal of Mathematics, 63(1):141–154, 1941.
  • [Li23] Z. Li. 2d internal waves in an ergodic setting, 2023. Preprint; arXiv:2301.12365.
  • [Maa05] L. R. M. Maas. Wave attractors: Linear yet nonlinear. International Journal of Bifurcation and Chaos, 15(09):2757–2782, 2005.
  • [MBSL97] L. R. M. Maas, D. Benielli, J. Sommeria, and F.-P. A. Lam. Observation of an internal wave attractor in a confined, stably stratified fluid. Nature, 388(6642):557–561, 1997.
  • [ML95] L. R. M. Maas and F.-P. A. Lam. Geometric focusing of internal waves. Journal of Fluid Mechanics, 300:1–41, 1995.
  • [Poi85] Henri Poincaré. Sur l’équilibre d’une masse fluide animée d’un mouvement de rotation. Acta Math., 7:259–380, 1885.
  • [Ral73] J. V. Ralston. On stationary modes in inviscid rotating fluids. Journal of Mathematical Analysis and Applications, 44(2):366–383, 1973.
  • [SE19] I. N. Sibgatullin and E. V. Ermanyuk. Internal and inertial wave attractors: A review. J Appl Mech Tech Phy, 60(2):284–302, 2019.
  • [Sob54] S. L. Sobolev. On a new problem of mathematical physics. Izv. Akad. Nauk SSSR Ser. Mat., 18(1):3–50, 1954.
  • [Tao19] Z. Tao. 0-th order pseudo-differential operator on the circle, 2019. to appear in Proc. Amer. Math. Soc.; arXiv:1909.06316.
  • [Wal99] J. A. Walsh. The dynamics of circle homeomorphisms: a hands-on introduction. Math. Mag., 72(1):3–13, 1999.
  • [Wan22] J. Wang. Dynamics of resonances for 0th order pseudodifferential operators. Commun. Math. Phys., 391:643–668, 2022.
  • [Yoc84] J.-C. Yoccoz. Conjugaison différentiable des difféomorphismes du cercle dont le nombre de rotation vérifie une condition diophantienne. Annales scientifiques de l’École Normale Supérieure, 17(3):333–359, 1984.
  • [Zhu22] S. Zhu. On the smoothness and regularity of the chess billiard flow and the poincaré problem, 2022. Preprint; arXiv:2210.13384.