One dimensional spaces always satisfy the regular Weyl’s law
Abstract.
Ambrosio, Honda, and Tewodrose proved that the regular Weyl’s law is equivalent to a mild condition related to the infinitesimal behavior of the measure of balls in compact finite dimensional spaces. Though that condition is seemed to always hold for any such spaces, however, Dai, Honda, Pan, and Wei recently show that for any integer at least 2, there exists a compact space of dimension fails to satisfy the regular Weyl’s law. In this short article we prove that one dimensional spaces always satisfy the regular Weyl’s law.
1. Introduction
In 1911, H. Weyl proved an asymptotic formula of the distribution of the eigenvalues of the Dirichlet Laplacian in the bounded domain ([W]); precisely, let be a sequence of all eigenvalues of the (minus) Dirichlet Laplacian , and let be the counting function of the eigenvalues, namely . It is known that are positive numbers and . Then it holds that
where is the volume of the unit ball in and is the -dimensional Hausdorff measure. This result has been generalized to many other situations(see for instance [I100] for a brief history and the generalization of that theorems). Here we focus on non-smooth setting. In [AHTweyl], Ambrosio, Honda, and Tewodrose proved the following theorem.
Theorem 1.1.
For a compact finite dimensional space of the essential dimension (see Definition 2.13 for the definition), the Weyl’s law
(1.1) |
holds if and only if
(1.2) |
We give some remarks. A finite dimensional space is a metric measure space with Ricci curvature bounded from below and dimension bounded from above in the synthetic sense(see Definition 2.6 for precise definition). Under (1.2), though the reference measure is not in general, the -dimensional Hausdorff measure appears in the right-hand side in (1.1).
Since it is known that -a.e.([AHTweyl]), (1.2) seems to be held for all spaces. However, Dai, Honda, Pan, and Wei prove the following.
Theorem 1.2 ([DHPW]).
For any , there exists a compact space of the essential dimension 2 such that
where is the singular set and is a canonical constant.
See Theorem 2.12 for the definition of the singular set and [DHPW] the explicit definition of the canonical constant . One notably fact is that the limit of the counting function is different from the ordinary ones. Moreover, they show a more pathological example.
Theorem 1.3 ([DHPW]).
There exists a compact space such that
In this example, the asymptotic behavior of the counting function is not even polynomial growth.
Motivated by these theorem, (1.1) is called the regular Weyl’s law. So a natural question is raised:
: Does the regular Weyl’s law hold for one-dimensional spaces?
The following is the main theorem.
Theorem 1.4.
Let be a non-trivial compact space for , . Then the followings are all equivalent;
-
(1)
,
-
(2)
or for ,
-
(3)
,
-
(4)
,
-
(5)
for .
This theorem is a complete answer to the question. And on the contrary for spaces of dimension at least 2, the regular Weyl’s law always holds for ALL one-dimensional spaces.
2. Preliminaries
Let be a metric space. We denote the set of all continuous functions on with bounded support by . We call a function an -Lipschitz function for if holds for any . We denote the set of all Lipschitz functions in by . For a generic function , its local Lipschitz constant at is defined by
We denote the set of all Borel probability measures on by . We define
For any two Borel probability measures , the coupling between them is defined as
The set of all couplings between and is denoted by , which is not empty since .
Definition 2.1 (-Wasserstein space).
For given , the -Wasserstein distance between them, , is defined by
It is known that is a metric on . The metric space is called the -Wasserstein space.
It is known that is complete separable if and only if so is .
2.1. Convex functions on geodesic spaces
Let be a metric space. We call a geodesic space if for any two points , there exists a continuous curve connecting them such that
holds for any . For given , , we define the distortion coefficients for by
By using the distortion coefficients, we define the convexity of functions. For a given function , .
Definition 2.2 (-convex).
We say a function , -convex if for any , there exists a geodesic connecting them such that
for any .
One can prove that each -convex function is locally Lipschitz in the interior of the geodesic. Hence -convex functions on the intervals are differentiable almost everywhere by Rademacher’s theorem. The following results, that is a key proposition in this short article, are found in [CMnew] (cf. [CMonge]).
Proposition 2.3 ([CMnew]).
Let be an interval for and a -convex function for on . Then any ,
(2.1) |
2.2. spaces
We call a triplet metric measure space if is a complete separable metric space and is a locally finite Borel measure on . Two metric measure spaces and are isomorphic to each other if there exists an isometry such that . In this case we just denote . For a metric measure space , is a -finite measure space. Hence the Radon-Nikodym theorem holds, that is, if is absolutely continuous with respect to , denoted by , there exists an function such that . From now on, we always assume that there exists a constant such that
(2.2) |
for a point . The relative entropy functional is defined by
By (2.2), for any .
Definition 2.4 ( space,[EKS]).
Let , . A metric measure space with (2.2) is called a space if is -convex.
For , we define the Cheeger energy of , , by
It is known that for any , there exists an -function such that . We define a Banach space equipped with the norm
called the Sobolev space over .
Definition 2.5 (Infinitesimal Hilbertianity, [Gdiffstr]).
Let be a metric measure space. We say that is infinitesimally Hilbertian if is a Hilbert space.
By the infinitesimal Hilbertianity, we define the inner product of the differentials for by
Definition 2.6 ( space).
Let be a metric measure space with the volume growth condition (2.2), and , . We call an space if the following two conditions are hold;
-
(1)
is a space,
-
(2)
is infinitesimally Hilbertian.
The -dimensional Riemannian manifolds with are spaces if , and Ricci limit spaces, measured Gromov-Hausdorff limit of manifolds with bounded dimension from above and bounded Ricci curvature from below, are also spaces. See Theorem 2.10.
Remark 2.7 (Historical remarks).
The synthetic notion of Ricci curvature bound for metric measure spaces is first introduced by Sturm [Stmms1, Stmms2] and Lott, Villani [LV] independently([LV] for and for finite , [Stmms1] for , [Stmms2] for for finite ). In order to get the tensorial and localization property, Bacher and Sturm introduced the reduced curvature-dimension condition [BS]. Non-Riemannian Finsler manifolds can be spaces. To get rid of such a class of spaces, Gigli defined the infinitesimal Hilbertianity [Gdiffstr] and Ambrosio, Gigli, Savaré defined the Riemannian curvature-dimension condition for compact metric measure spaces [AGSRiem], afterwards, Ambrosio, Gigli, Mondino, Rajala defined the same notion for -finite cases [AGMR]. For finite , Erbar, Kuwada, Sturm [EKS] and Ambrosio, Mondino, Savaré [AMSnon] defined space independently. Under essentially non-branching assumption(see [RaS] for the definition), condition and condition are equivalent to each other[CMil]. The metric measure spaces that satisfy the conditions in Definition 2.6 was called spaces. However, by [CMil], we call these spaces spaces now.
In order to emphasize being finite, we say that space is finite dimensional.
Remark 2.8.
The condition is the synthetic notion of lower bound of Ricci curvature(), and upper bound of dimension(). Actually, space also satisfies condition for and .
For space , is a space for .
Let be an space.
Definition 2.9 (Laplacian).
A function belongs to if there exists an -function such that
holds for any . In this case, the Laplacian of is denoted by .
The Laplacian is a densely defined nonpositively definite self-adjoint operator in . When is a compact finite dimensional space, all the spectrum of are eigenvalues, and the multiplicity of zero spectrum is 1 since the resolvent operators are compact. We denote the non-zero eigenvalues by with multiplicity.
2.3. Infinitesimal structure on spaces
Let be a family of pointed metric measure spaces. We say that converges to in the measured Gromov-Hausdorff sense (mGH for short) if the following conditions are satisfied; there exist sequences of positive numbers , , and of Borel maps
-
(1)
for any , any . And ,
-
(2)
,
-
(3)
in duality with .
In this case we denote by . condition is stable under mGH convergence and a set of pointed metric measure spaces with the same condition is sequentially precompact with respect to mGH convergence(see [GMS]).
Theorem 2.10 ([GMS, EKS]).
Let be a sequence of spaces and . Assume and assume . Then is an space. Moreover, any sequence of spaces has a convergent subsequence in mGH sense.
Let be an space. Given , and for is defined by
Since is an space, combining Remark 2.8 with Theorem 2.10, we are able to find a convergent subsequence . Therefore we reach the following definition.
Definition 2.11 (Tangent cone).
For a given point , we define the tangent cone at by
We often denote it by for short.
The -dimensional regular set is defined by
where is the standard Euclidean distance and is the normalized Lebesgue measure on , this means,
It is known that if for spaces.
Brue and Semola proved the following result.
Theorem 2.12 ([BSconst]).
Let be an space. Then there exists an integer such that .
Definition 2.13 (Essentially dimension).
We call the integer in Theorem 2.12 the essential dimension, and denote it by and the singular set in .
Remark 2.14 (Hausdorff dimension and Essential dimension).
In [Stmms2], Sturm proved that the Hausdorff dimension is at most for spaces. By the behavior of measure on the regular set (see [AHTweyl]), it is clear that . The coincident of these two notion of dimension was open. However, recently Pan and Wei proved that there exists an space whose Hausdorff dimension is strictly larger than essential one ([PW]).
Theorem 2.12 does not guarantee the non-existence of points belonging to another dimensional regular set(Non-existence of higher dimensional regular point is proven in [K]). One dimensional case is much simpler than the situation for other dimension.
Theorem 2.15 ([KL, CMisop, CMnew]).
Let be an space. Then the following are all equivalent to each other;
-
(1)
,
-
(2)
,
-
(3)
is isometric to either , , , or for , .
Moreover the reference measure is equivalent to (denote it by ), which is of the form , and its density function is -convex. Hence satisfies (2.1).
Recall we say that two measures and are equivalent to each other if both and hold.
Remark 2.16.
-convexity of the density function is proven in [KL]. The improvement version of the convexity is proven by [CMonge, CMil]. The density function is continuous.
2.4. Weyl’s law on finite dimensional compact spaces
In this subsection, we always assume the metric measure space is a compact space. As aforementioned before, all the spectrum of are eigenvalues, and holds. We define the counting function by
Ambrosio, Honda, and Tewodrose [AHTweyl] proves the following result.
Theorem 2.17 ([AHTweyl] cf. [ZZweyl]).
Let be a compact space with . Then
if and only if
Remark 2.18.
Independently [ZZweyl] also proves a similar result.
For , define the subset by
It is known that (see [AHTweyl]). In order to prove the main result, we need the following results.
Theorem 2.19 ([AHTweyl]).
Let be a compact space and . Then we have
where is the -dimensional Hausdorff measure on .
The so-called Abelian theorem is also important for our main result.
Theorem 2.20 (Abelian theorem cf.[AHTweyl]).
Let be a nonnegative and -finite Borel measure on . Assume that there exist and such that
Then
In the next section, we use the above theorem for . Note that
in this case.
3. Proof of the main theorem
Now let be a compact space with . Without loss of generality, we may assume and for simplicity.
The equivalence between (1) and (2) is proven in [KL]. The implication (4) to (3), (3) to (5) are trivial.
(2) (4): When , then the density function for is continuous(see Remark 2.16). Since and the continuity of , never vanish. Put and . Since
holds, we have
(3.1) |
by the dominated convergence theorem.
Let us consider the case for . Without loss of generality, we may assume . When the density function for has a positive minimum, then a similar argument as implies the consequence. So, we assume . Take a small positive number and fix it. For , we have
Thus we get
Therefore
We have
By using the Taylor expansion for , we have and for sufficiently small , further, applying to on , we obtain
Also we get
Finally we have
On the other hand, for , we have
Applying (2.1) to on , we get the estimate
Then
The upper bound depends on neither nor . We apply the same argument for near . Then by the dominated convergence theorem,
(3.2) |
Both cases, , , we have (4) by combining (3.1) and (3.2) with Theorem 4.3 in [AHTweyl].
(5) (1): By combining the assumption and Abelian theorem, we have
Let . Then by Theorem 2.19, we obtain
This implies . Since and is an integer, has to be 1.
Acknowledgement
The authors would like to thank Professor Shouhei Honda for telling us the problem and his helpful comments and fruitful discussion. Y.K. is partly supported by JSPS KAKENHI Grant Numbers JP18K13412 and JP22K03291.