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

One dimensional 𝖱𝖢𝖣\mathsf{RCD} spaces always satisfy the regular Weyl’s law

Akemi Iwahashi, Yu Kitabeppu, Akari Yonekura Kumamoto University 211d8001@st.kumamoto-u.ac.jp ybeppu@kumamoto-u.ac.jp 216d8018@st.kumamoto-u.ac.jp
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 𝖱𝖢𝖣\mathsf{RCD} 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 nn at least 2, there exists a compact 𝖱𝖢𝖣\mathsf{RCD} space of nn dimension fails to satisfy the regular Weyl’s law. In this short article we prove that one dimensional 𝖱𝖢𝖣\mathsf{RCD} 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 Ωn\Omega\subset\mathbb{R}^{n}([W]); precisely, let {λi}i\{\lambda_{i}\}_{i} be a sequence of all eigenvalues of the (minus) Dirichlet Laplacian Δ-\Delta, and let N(λ)N(\lambda) be the counting function of the eigenvalues, namely N(λ):=#{i;λiλ}N(\lambda):=\#\{i\;;\;\lambda_{i}\leq\lambda\}. It is known that λi\lambda_{i} are positive numbers and λi\lambda_{i}\rightarrow\infty. Then it holds that

limλN(λ)λn/2=ωn(2π)nn(Ω),\displaystyle\lim_{\lambda\rightarrow\infty}\frac{N(\lambda)}{\lambda^{n/2}}=\frac{\omega_{n}}{(2\pi)^{n}}\mathcal{H}^{n}(\Omega),

where ωn\omega_{n} is the volume of the unit ball in n\mathbb{R}^{n} and n\mathcal{H}^{n} is the nn-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 𝖱𝖢𝖣\mathsf{RCD} space (X,d,𝔪)(X,d,\mathfrak{m}) of the essential dimension nn(see Definition 2.13 for the definition), the Weyl’s law

(1.1) limλN(λ)λn/2=ωn(2π)nn(X)\displaystyle\lim_{\lambda\rightarrow\infty}\frac{N(\lambda)}{\lambda^{n/2}}=\frac{\omega_{n}}{(2\pi)^{n}}\mathcal{H}^{n}(X)

holds if and only if

(1.2) limr+0Xrn𝔪(Br(x))𝔪(dx)=Xlimr+0rn𝔪(Br(x))𝔪(dx).\displaystyle\lim_{r\rightarrow+0}\int_{X}\frac{r^{n}}{\mathfrak{m}(B_{r}(x))}\,\mathfrak{m}(dx)=\int_{X}\lim_{r\rightarrow+0}\frac{r^{n}}{\mathfrak{m}(B_{r}(x))}\,\mathfrak{m}(dx).

We give some remarks. A finite dimensional 𝖱𝖢𝖣\mathsf{RCD} space (X,d,𝔪)(X,d,\mathfrak{m}) 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 𝔪\mathfrak{m} is not n\mathcal{H}^{n} in general, the nn-dimensional Hausdorff measure appears in the right-hand side in (1.1).

Since it is known that limr+0rn/𝔪(Br(x))(0,)\lim_{r\rightarrow+0}r^{n}/\mathfrak{m}(B_{r}(x))\in(0,\infty) 𝔪\mathfrak{m}-a.e.([AHTweyl]), (1.2) seems to be held for all 𝖱𝖢𝖣\mathsf{RCD} spaces. However, Dai, Honda, Pan, and Wei prove the following.

Theorem 1.2 ([DHPW]).

For any β(2,)\beta\in(2,\infty), there exists a compact 𝖱𝖢𝖣(1,Nβ)\mathsf{RCD}(-1,N_{\beta}) space (X,d,𝔪)(X,d,\mathfrak{m}) of the essential dimension 2 such that

limλN(λ)λβ/2=cΓ(β+1)β(𝒮)(0,),\displaystyle\lim_{\lambda\rightarrow\infty}\frac{N(\lambda)}{\lambda^{\beta/2}}=\frac{c}{\Gamma(\beta+1)}\mathcal{H}^{\beta}(\mathcal{S})\in(0,\infty),

where 𝒮X\mathcal{S}\subset X is the singular set and cc is a canonical constant.

See Theorem 2.12 for the definition of the singular set and [DHPW] the explicit definition of the canonical constant cc. 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 𝖱𝖢𝖣(1,10)\mathsf{RCD}(-1,10) space such that

limλN(λ)λlogλ=14π.\displaystyle\lim_{\lambda\rightarrow\infty}\frac{N(\lambda)}{\lambda\log\lambda}=\frac{1}{4\pi}.

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:

Q¯\underline{Q}: Does the regular Weyl’s law hold for one-dimensional 𝖱𝖢𝖣\mathsf{RCD} spaces?

The following is the main theorem.

Theorem 1.4.

Let (X,d,𝔪)(X,d,\mathfrak{m}) be a non-trivial compact 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space for KK\in\mathbb{R}, N(1,)N\in(1,\infty). Then the followings are all equivalent;

  1. (1)

    𝖽𝗂𝗆𝖾𝗌𝗌(X,d,𝔪)=1\mathsf{dim}_{\mathsf{ess}}(X,d,\mathfrak{m})=1,

  2. (2)

    X=[0,]X=[0,\ell] or S1(r)S^{1}(r) for ,r>0\ell,r>0,

  3. (3)

    N(λ)λ1/2N(\lambda)\sim\lambda^{1/2},

  4. (4)

    limλ0N(λ)λ1/2=ω12π1(X)=1π1(X)\displaystyle{\lim_{\lambda\downarrow 0}\frac{N(\lambda)}{\lambda^{1/2}}=\frac{\omega_{1}}{2\pi}\mathcal{H}^{1}(X)=\frac{1}{\pi}\mathcal{H}^{1}(X)},

  5. (5)

    limλN(λ)λ(1+α)/2=0\displaystyle{\lim_{\lambda\rightarrow\infty}\frac{N(\lambda)}{\lambda^{(1+\alpha)/2}}}=0 for 0<α10<\alpha\leq 1.

This theorem is a complete answer to the question. And on the contrary for 𝖱𝖢𝖣\mathsf{RCD} spaces of dimension at least 2, the regular Weyl’s law always holds for ALL one-dimensional 𝖱𝖢𝖣\mathsf{RCD} spaces.

2. Preliminaries

Let (Y,dY)(Y,d_{Y}) be a metric space. We denote the set of all continuous functions on YY with bounded support by Cbs(Y)C_{bs}(Y). We call a function g:Yg:Y\rightarrow\mathbb{R} an LL-Lipschitz function for L>0L>0 if |g(y0)g(y1)|LdY(y0,y1)\left|g(y_{0})-g(y_{1})\right|\leq Ld_{Y}(y_{0},y_{1}) holds for any y0,y1Yy_{0},y_{1}\in Y. We denote the set of all Lipschitz functions in YY by LIP(Y){\bf\mathrm{LIP}}(Y). For a generic function f:Yf:Y\rightarrow\mathbb{R}, its local Lipschitz constant at yYy\in Y is defined by

lipf(y):={lim supzy|f(z)f(y)|dY(z,y)if y is not isolated,0otherwise.\displaystyle\mathrm{lip}f(y):=\begin{cases}\limsup_{z\rightarrow y}\frac{\left|f(z)-f(y)\right|}{d_{Y}(z,y)}&\text{if $y$ is not isolated},\\ 0&\text{otherwise}.\end{cases}

We denote the set of all Borel probability measures on YY by 𝒫(Y)\mathcal{P}(Y). We define

𝒫2(Y):={μ𝒫(Y);YdY2(o,y)μ(dy)<}.\displaystyle\mathcal{P}_{2}(Y):=\left\{\mu\in\mathcal{P}(Y)\;;\;\int_{Y}d_{Y}^{2}(~{}^{\exists}o,y)\,\mu(dy)<\infty\right\}.

For any two Borel probability measures μ,ν𝒫(Y)\mu,\nu\in\mathcal{P}(Y), the coupling ξ𝒫(Y×Y)\xi\in\mathcal{P}(Y\times Y) between them is defined as

{ξ(A×Y)=μ(A)ξ(Y×A)=ν(A)for any Borel subset AY.\displaystyle\begin{cases}\xi(A\times Y)=\mu(A)\\ \xi(Y\times A)=\nu(A)&\text{for any Borel subset }A\subset Y.\end{cases}

The set of all couplings between μ\mu and ν\nu is denoted by 𝖢𝗉𝗅(μ,ν)\mathsf{Cpl}(\mu,\nu), which is not empty since μν𝖢𝗉𝗅(μ,ν)\mu\otimes\nu\in\mathsf{Cpl}(\mu,\nu).

Definition 2.1 (L2L^{2}-Wasserstein space).

For given μ,ν𝒫2(Y)\mu,\nu\in\mathcal{P}_{2}(Y), the L2L^{2}-Wasserstein distance between them, W2(μ,ν)W_{2}(\mu,\nu), is defined by

W2(μ,ν):=inf{dL2(ξ);ξ𝖢𝗉𝗅(μ,ν)}.\displaystyle W_{2}(\mu,\nu):=\inf\left\{\left\|d\right\|_{L^{2}(\xi)}\;;\;\xi\in\mathsf{Cpl}(\mu,\nu)\right\}.

It is known that W2W_{2} is a metric on 𝒫2(Y)\mathcal{P}_{2}(Y). The metric space (𝒫2(Y),W2)(\mathcal{P}_{2}(Y),W_{2}) is called the L2L^{2}-Wasserstein space.

It is known that (𝒫2(Y),W2)(\mathcal{P}_{2}(Y),W_{2}) is complete separable if and only if so is (Y,dY)(Y,d_{Y}).

2.1. Convex functions on geodesic spaces

Let (Y,dY)(Y,d_{Y}) be a metric space. We call (Y,dY)(Y,d_{Y}) a geodesic space if for any two points y0,y1Yy_{0},y_{1}\in Y, there exists a continuous curve γ:[0,1]Y\gamma:[0,1]\rightarrow Y connecting them such that

dY(γs,γt)=|st|dY(y0,y1)\displaystyle d_{Y}(\gamma_{s},\gamma_{t})=\left|s-t\right|d_{Y}(y_{0},y_{1})

holds for any s,t[0,1]s,t\in[0,1]. For given KK\in\mathbb{R}, N(1,)N\in(1,\infty), we define the distortion coefficients σK,N(t)\sigma^{(t)}_{K,N} for t[0,1]t\in[0,1] by

σK,N(t)(θ):={if Kθ2Nπ2,sin(tθK/N)sin(θ)K/Nif 0<Kθ2<Nπ2,tif Kθ2=0,sinh(tθK/N)sinh(θK/N)if Kθ2<0.\displaystyle\sigma^{(t)}_{K,N}(\theta):=\begin{cases}\infty&\text{if }K\theta^{2}\geq N\pi^{2},\\ \frac{\sin(t\theta\sqrt{K/N})}{\sin(\theta)\sqrt{K/N}}&\text{if }0<K\theta^{2}<N\pi^{2},\\ t&\text{if }K\theta^{2}=0,\\ \frac{\sinh(t\theta\sqrt{-K/N})}{\sinh(\theta\sqrt{-K/N})}&\text{if }K\theta^{2}<0.\end{cases}

By using the distortion coefficients, we define the convexity of functions. For a given function g:Y{}g:Y\rightarrow\mathbb{R}\cup\{\infty\}, 𝒟(g):={yY;g(y)<}\mathcal{D}(g):=\{y\in Y\;;\;g(y)<\infty\}.

Definition 2.2 ((K,N)(K,N)-convex).

We say a function f:Y{}f:Y\rightarrow\mathbb{R}\cup\{\infty\}, (K,N)(K,N)-convex if for any y0,y1Yy_{0},y_{1}\in Y, there exists a geodesic γ:[0,1]Y\gamma:[0,1]\rightarrow Y connecting them such that

exp(1Nf(γt))\displaystyle\exp\left(-\frac{1}{N}f(\gamma_{t})\right)
σK,N(1t)(dY(y0,y1))exp(1Nf(y0))+σK,N(t)(dY(y0,y1))exp(1Nf(y1))\displaystyle\geq\sigma^{(1-t)}_{K,N}(d_{Y}(y_{0},y_{1}))\exp\left(-\frac{1}{N}f(y_{0})\right)+\sigma^{(t)}_{K,N}(d_{Y}(y_{0},y_{1}))\exp\left(-\frac{1}{N}f(y_{1})\right)

for any t[0,1]t\in[0,1].

One can prove that each (K,N)(K,N)-convex function is locally Lipschitz in the interior of the geodesic. Hence (K,N)(K,N)-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 I:=(a,b)I:=(a,b) be an interval for a<ba<b and ff a (K,N1)(K,N-1)-convex function for K<0K<0 on II. Then any x0<x1Ix_{0}<x_{1}\in I,

(2.1) (sinh((bx1)K/N1)sinh((bx0)K/N1))N1ef(x1)ef(x0)(sinh((x1a)K/N1)sinh((x0a)K/N1))N1.\displaystyle\left(\frac{\sinh((b-x_{1})\sqrt{-K/N-1})}{\sinh((b-x_{0})\sqrt{-K/N-1})}\right)^{N-1}\leq\frac{e^{-f(x_{1})}}{e^{-f(x_{0})}}\leq\left(\frac{\sinh((x_{1}-a)\sqrt{-K/N-1})}{\sinh((x_{0}-a)\sqrt{-K/N-1})}\right)^{N-1}.

2.2. 𝖱𝖢𝖣\mathsf{RCD} spaces

We call a triplet (X,d,𝔪)(X,d,\mathfrak{m}) metric measure space if (X,d)(X,d) is a complete separable metric space and 𝔪\mathfrak{m} is a locally finite Borel measure on XX. Two metric measure spaces (X,dX,𝔪X)(X,d_{X},\mathfrak{m}_{X}) and (Y,dY,𝔪Y)(Y,d_{Y},\mathfrak{m}_{Y}) are isomorphic to each other if there exists an isometry f:𝗌𝗎𝗉𝗉𝔪X𝗌𝗎𝗉𝗉𝔪Yf:\mathsf{supp}\,\mathfrak{m}_{X}\rightarrow\mathsf{supp}\,\mathfrak{m}_{Y} such that f𝔪X=𝔪Yf_{*}\mathfrak{m}_{X}=\mathfrak{m}_{Y}. In this case we just denote (X,dX,𝔪X)=(Y,dY,𝔪Y)(X,d_{X},\mathfrak{m}_{X})=(Y,d_{Y},\mathfrak{m}_{Y}). For a metric measure space (X,d,𝔪)(X,d,\mathfrak{m}), (X,𝔪)(X,\mathfrak{m}) is a σ\sigma-finite measure space. Hence the Radon-Nikodym theorem holds, that is, if μ𝒫(X)\mu\in\mathcal{P}(X) is absolutely continuous with respect to 𝔪\mathfrak{m}, denoted by μ𝔪\mu\ll\mathfrak{m}, there exists an L1(𝔪)L^{1}(\mathfrak{m}) function ρ\rho such that μ=ρ𝔪\mu=\rho\mathfrak{m}. From now on, we always assume that there exists a constant C>0C>0 such that

(2.2) XeCd2(x0,x)𝔪(dx)<\displaystyle\int_{X}e^{-Cd^{2}(x_{0},x)}\,\mathfrak{m}(dx)<\infty

for a point x0Xx_{0}\in X. The relative entropy functional 𝖤𝗇𝗍𝔪:𝒫(X){±}\mathsf{Ent}_{\mathfrak{m}}:\mathcal{P}(X)\rightarrow\mathbb{R}\cup\{\pm\infty\} is defined by

𝖤𝗇𝗍𝔪(μ):={{ρ>0}ρlogρd𝔪if μ=ρ𝔪𝔪otherwise.\displaystyle\mathsf{Ent}_{\mathfrak{m}}(\mu):=\begin{cases}\int_{\{\rho>0\}}\rho\log\rho\,d\mathfrak{m}&\text{if }\mu=\rho\mathfrak{m}\ll\mathfrak{m}\\ \infty&\text{otherwise}.\end{cases}

By (2.2), 𝖤𝗇𝗍𝔪(μ)>\mathsf{Ent}_{\mathfrak{m}}(\mu)>-\infty for any μ𝒫2(X)\mu\in\mathcal{P}_{2}(X).

Definition 2.4 (𝖢𝖣e(K,N)\mathsf{CD}_{e}(K,N) space,[EKS]).

Let KK\in\mathbb{R}, N(1,)N\in(1,\infty). A metric measure space (X,d,𝔪)(X,d,\mathfrak{m}) with (2.2) is called a 𝖢𝖣e(K,N)\mathsf{CD}_{e}(K,N) space if 𝖤𝗇𝗍𝔪\mathsf{Ent}_{\mathfrak{m}} is (K,N)(K,N)-convex.

For fL2(𝔪)f\in L^{2}(\mathfrak{m}), we define the Cheeger energy of ff, 𝖢𝗁(f)\mathsf{Ch}(f), by

𝖢𝗁(f):=12inf{lim infnX(lipfn)2d𝔪;fnL2(𝔪)f,fnLIP(X)}.\displaystyle\mathsf{Ch}(f):=\frac{1}{2}\inf\left\{\liminf_{n\rightarrow\infty}\int_{X}(\mathrm{lip}f_{n})^{2}\,d\mathfrak{m}\;;\;f_{n}\xrightarrow{L^{2}(\mathfrak{m})}f,\,f_{n}\in{\bf\mathrm{LIP}}(X)\right\}.

It is known that for any f𝒟(𝖢𝗁)f\in\mathcal{D}(\mathsf{Ch}), there exists an L2(𝔪)L^{2}(\mathfrak{m})-function |df|\left|df\right| such that 2𝖢𝗁(f)=X|df|2𝑑𝔪2\mathsf{Ch}(f)=\int_{X}\left|df\right|^{2}\,d\mathfrak{m}. We define a Banach space W1,2(X,d,𝔪):=L2(𝔪)𝒟(𝖢𝗁)W^{1,2}(X,d,\mathfrak{m}):=L^{2}(\mathfrak{m})\cap\mathcal{D}(\mathsf{Ch}) equipped with the norm

f1,22:=fL2(𝔪)2+|df|L2(𝔪)2,\displaystyle\left\|f\right\|^{2}_{1,2}:=\left\|f\right\|^{2}_{L^{2}(\mathfrak{m})}+\left\|\left|df\right|\right\|^{2}_{L^{2}(\mathfrak{m})},

called the Sobolev space over (X,d,𝔪)(X,d,\mathfrak{m}).

Definition 2.5 (Infinitesimal Hilbertianity, [Gdiffstr]).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be a metric measure space. We say that (X,d,𝔪)(X,d,\mathfrak{m}) is infinitesimally Hilbertian if W1,2(X,d,𝔪)W^{1,2}(X,d,\mathfrak{m}) is a Hilbert space.

By the infinitesimal Hilbertianity, we define the inner product of the differentials for f,g𝒟(𝖢𝗁)f,g\in\mathcal{D}(\mathsf{Ch}) by

Xdf,dg𝑑𝔪:=12(𝖢𝗁(f+g)𝖢𝗁(fg)).\displaystyle\int_{X}\left\langle df,dg\right\rangle\,d\mathfrak{m}:=\frac{1}{2}\left(\mathsf{Ch}(f+g)-\mathsf{Ch}(f-g)\right).
Definition 2.6 (𝖱𝖢𝖣\mathsf{RCD} space).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be a metric measure space with the volume growth condition (2.2), and KK\in\mathbb{R}, N(1,)N\in(1,\infty). We call (X,d,𝔪)(X,d,\mathfrak{m}) an 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space if the following two conditions are hold;

  1. (1)

    (X,d,𝔪)(X,d,\mathfrak{m}) is a 𝖢𝖣e(K,N)\mathsf{CD}_{e}(K,N) space,

  2. (2)

    (X,d,𝔪)(X,d,\mathfrak{m}) is infinitesimally Hilbertian.

The nn-dimensional Riemannian manifolds with RicK\mathrm{Ric}\geq K are 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) spaces if nNn\leq N, and Ricci limit spaces, measured Gromov-Hausdorff limit of manifolds with bounded dimension from above and bounded Ricci curvature from below, are also 𝖱𝖢𝖣\mathsf{RCD} 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 𝖢𝖣(K,)\mathsf{CD}(K,\infty) and 𝖢𝖣(0,N)\mathsf{CD}(0,N) for finite NN, [Stmms1] for 𝖢𝖣(K,)\mathsf{CD}(K,\infty), [Stmms2] for 𝖢𝖣(K,N)\mathsf{CD}(K,N) for finite NN). In order to get the tensorial and localization property, Bacher and Sturm introduced the reduced curvature-dimension condition 𝖢𝖣(K,N)\mathsf{CD}^{*}(K,N) [BS]. Non-Riemannian Finsler manifolds can be 𝖢𝖣\mathsf{CD} 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 𝖱𝖢𝖣(K,)\mathsf{RCD}(K,\infty) for compact metric measure spaces [AGSRiem], afterwards, Ambrosio, Gigli, Mondino, Rajala defined the same notion for σ\sigma-finite cases [AGMR]. For finite NN, Erbar, Kuwada, Sturm [EKS] and Ambrosio, Mondino, Savaré [AMSnon] defined 𝖱𝖢𝖣(K,N)\mathsf{RCD}^{*}(K,N) space independently. Under essentially non-branching assumption(see [RaS] for the definition), 𝖢𝖣\mathsf{CD} condition and 𝖢𝖣\mathsf{CD}^{*} condition are equivalent to each other[CMil]. The metric measure spaces that satisfy the conditions in Definition 2.6 was called 𝖱𝖢𝖣(K,N)\mathsf{RCD}^{*}(K,N) spaces. However, by [CMil], we call these spaces 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) spaces now.

In order to emphasize NN being finite, we say that 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space is finite dimensional.

Remark 2.8.

The 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) condition is the synthetic notion of lower bound of Ricci curvature(K\geq K), and upper bound of dimension(N\leq N). Actually, 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}) also satisfies 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K^{\prime},N^{\prime}) condition for KKK^{\prime}\leq K and NNN^{\prime}\geq N.

For 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}), (X,ad,b𝔪)(X,ad,b\mathfrak{m}) is a 𝖱𝖢𝖣(a2K,N)\mathsf{RCD}(a^{-2}K,N) space for a,b>0a,b>0.

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space.

Definition 2.9 (Laplacian).

A function f𝒟(𝖢𝗁)f\in\mathcal{D}(\mathsf{Ch}) belongs to 𝒟(Δ)\mathcal{D}(\Delta) if there exists an L2(𝔪)L^{2}(\mathfrak{m})-function hh such that

Xdf,dg𝑑𝔪=Xhg𝑑𝔪\displaystyle\int_{X}\left\langle df,dg\right\rangle\,d\mathfrak{m}=-\int_{X}hg\,d\mathfrak{m}

holds for any g𝒟(𝖢𝗁)g\in\mathcal{D}(\mathsf{Ch}). In this case, the Laplacian of ff is denoted by Δf:=h\Delta f:=h.

The Laplacian is a densely defined nonpositively definite self-adjoint operator in L2(𝔪)L^{2}(\mathfrak{m}). When (X,d,𝔪)(X,d,\mathfrak{m}) is a compact finite dimensional 𝖱𝖢𝖣\mathsf{RCD} space, all the spectrum of Δ-\Delta are eigenvalues, and the multiplicity of zero spectrum is 1 since the resolvent operators are compact. We denote the non-zero eigenvalues by 0<λ1λ20<\lambda_{1}\leq\lambda_{2}\leq\cdots\rightarrow\infty with multiplicity.

2.3. Infinitesimal structure on 𝖱𝖢𝖣\mathsf{RCD} spaces

Let {(Xi,di,𝔪i,xi0)}i{}\{(X_{i},d_{i},\mathfrak{m}_{i},x_{i}^{0})\}_{i\in\mathbb{N}\cup\{\infty\}} be a family of pointed metric measure spaces. We say that (Xi,di,𝔪i,xi0)(X_{i},d_{i},\mathfrak{m}_{i},x^{0}_{i}) converges to (X,d,𝔪,x0)(X_{\infty},d_{\infty},\mathfrak{m}_{\infty},x^{0}_{\infty}) in the measured Gromov-Hausdorff sense (mGH for short) if the following conditions are satisfied; there exist sequences of positive numbers εi0\varepsilon_{i}\downarrow 0, RiR_{i}\uparrow\infty, and of Borel maps φi:BRi(xi0)X\varphi_{i}:B_{R_{i}}(x_{i}^{0})\rightarrow X_{\infty}

  1. (1)

    |di(x,y)d(φi(x),φi(y))|<εi\left|d_{i}(x,y)-d_{\infty}(\varphi_{i}(x),\varphi_{i}(y))\right|<\varepsilon_{i} for any ii\in\mathbb{N}, any x,yBRi(xi0)x,y\in B_{R_{i}}(x_{i}^{0}). And BRiεi(φi(xi))Bεi(φi(BRi(xi0)))B_{R_{i}-\varepsilon_{i}}(\varphi_{i}(x_{i}))\subset B_{\varepsilon_{i}}(\varphi_{i}(B_{R_{i}}(x_{i}^{0}))),

  2. (2)

    φi(xi0)x0\varphi_{i}(x_{i}^{0})\rightarrow x^{0}_{\infty},

  3. (3)

    (φi)𝔪i𝔪(\varphi_{i})_{*}\mathfrak{m}_{i}\rightharpoonup\mathfrak{m}_{\infty} in duality with Cbs(X)C_{bs}(X_{\infty}).

In this case we denote by (Xi,di,𝔪i,xi0)mGH(X,d,𝔪,x0)(X_{i},d_{i},\mathfrak{m}_{i},x_{i}^{0})\xrightarrow{mGH}(X_{\infty},d_{\infty},\mathfrak{m}_{\infty},x_{\infty}^{0}). 𝖱𝖢𝖣\mathsf{RCD} condition is stable under mGH convergence and a set of pointed metric measure spaces with the same 𝖱𝖢𝖣\mathsf{RCD} condition is sequentially precompact with respect to mGH convergence(see [GMS]).

Theorem 2.10 ([GMS, EKS]).

Let {(Xi,di,𝔪i)}i\{(X_{i},d_{i},\mathfrak{m}_{i})\}_{i\in\mathbb{N}} be a sequence of 𝖱𝖢𝖣(Ki,N)\mathsf{RCD}(K_{i},N) spaces and xi0𝗌𝗎𝗉𝗉𝔪ix_{i}^{0}\in\mathsf{supp}\,\mathfrak{m}_{i}. Assume (Xi,di,𝔪i,xi0)mGH(X,d,𝔪,x0)(X_{i},d_{i},\mathfrak{m}_{i},x_{i}^{0})\xrightarrow{mGH}(X_{\infty},d_{\infty},\mathfrak{m}_{\infty},x^{0}_{\infty}) and assume KiKK_{i}\rightarrow K. Then (X,d,𝔪)(X_{\infty},d_{\infty},\mathfrak{m}_{\infty}) is an 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space. Moreover, any sequence of 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) spaces has a convergent subsequence in mGH sense.

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space. Given r>0r>0, dr:=r1dd_{r}:=r^{-1}d and 𝔪rx\mathfrak{m}_{r}^{x} for x𝗌𝗎𝗉𝗉𝔪x\in\mathsf{supp}\,\mathfrak{m} is defined by

𝔪rx:=(Br(x)1d(x,)rd𝔪)1𝔪.\displaystyle\mathfrak{m}_{r}^{x}:=\left(\int_{B_{r}(x)}1-\frac{d(x,\cdot)}{r}\,d\mathfrak{m}\right)^{-1}\mathfrak{m}.

Since (X,dr,𝔪rx)(X,d_{r},\mathfrak{m}_{r}^{x}) is an 𝖱𝖢𝖣(r2K,N)\mathsf{RCD}(r^{2}K,N) space, combining Remark 2.8 with Theorem 2.10, we are able to find a convergent subsequence {(X,dri,𝔪rix,x)}i\{(X,d_{r_{i}},\mathfrak{m}^{x}_{r_{i}},x)\}_{i}. Therefore we reach the following definition.

Definition 2.11 (Tangent cone).

For a given point x𝗌𝗎𝗉𝗉𝔪x\in\mathsf{supp}\,\mathfrak{m}, we define the tangent cone at xx by

𝖳𝖺𝗇(X,d,𝔪,x)\displaystyle\mathsf{Tan}(X,d,\mathfrak{m},x)
:={(Y,dY,𝔪Y,y);(X,dri,𝔪rix,x)mGH(Y,dY,𝔪Y,y) for a sequence ri0}.\displaystyle:=\left\{(Y,d_{Y},\mathfrak{m}_{Y},y)\;;\;(X,d_{r_{i}},\mathfrak{m}^{x}_{r_{i}},x)\xrightarrow{mGH}(Y,d_{Y},\mathfrak{m}_{Y},y)\text{ for a sequence }r_{i}\downarrow 0\right\}.

We often denote it by 𝖳𝖺𝗇(X,x)\mathsf{Tan}(X,x) for short.

The \ell-dimensional regular set \mathcal{R}_{\ell} is defined by

:={xX;𝖳𝖺𝗇(X,x)={(,dE,¯,0)}},\displaystyle\mathcal{R}_{\ell}:=\left\{x\in X\;;\;\mathsf{Tan}(X,x)=\{(\mathbb{R}^{\ell},d_{E},\underline{\mathcal{L}^{\ell}},0)\}\right\},

where dEd_{E} is the standard Euclidean distance and ¯\underline{\mathcal{L}^{\ell}} is the normalized Lebesgue measure on \mathbb{R}^{\ell}, this means,

B1(0)1|x|d¯(dx)=1.\displaystyle\int_{B_{1}(0)}1-\left|x\right|\,d\underline{\mathcal{L}^{\ell}}(dx)=1.

It is known that =\mathcal{R}_{\ell}=\emptyset if >[N]\ell>[N] for 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) spaces.

Brue and Semola proved the following result.

Theorem 2.12 ([BSconst]).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space. Then there exists an integer nn such that 𝔪(Xn)=0\mathfrak{m}(X\setminus\mathcal{R}_{n})=0.

Definition 2.13 (Essentially dimension).

We call the integer nn in Theorem 2.12 the essential dimension, and denote it by 𝖽𝗂𝗆𝖾𝗌𝗌(X,d,𝔪)\mathsf{dim}_{\mathsf{ess}}(X,d,\mathfrak{m}) and 𝒮:=Xn\mathcal{S}:=X\setminus\mathcal{R}_{n} the singular set in XX.

Remark 2.14 (Hausdorff dimension and Essential dimension).

In [Stmms2], Sturm proved that the Hausdorff dimension is at most NN for 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) spaces. By the behavior of measure on the regular set (see [AHTweyl]), it is clear that 𝖽𝗂𝗆𝖾𝗌𝗌𝖽𝗂𝗆H\mathsf{dim}_{\mathsf{ess}}\leq\mathsf{dim}_{H}. The coincident of these two notion of dimension was open. However, recently Pan and Wei proved that there exists an 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) 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 (X,d,𝔪)(X,d,\mathfrak{m}) be an 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space. Then the following are all equivalent to each other;

  1. (1)

    1\mathcal{R}_{1}\neq\emptyset,

  2. (2)

    𝖽𝗂𝗆𝖾𝗌𝗌(X,d,𝔪)=1\mathsf{dim}_{\mathsf{ess}}(X,d,\mathfrak{m})=1,

  3. (3)

    (X,d)(X,d) is isometric to either \mathbb{R}, 0\mathbb{R}_{\geq 0}, [0,][0,\ell], or S1(r)S^{1}(r) for >0\ell>0, r>0r>0.

Moreover the reference measure 𝔪\mathfrak{m} is equivalent to 1\mathcal{H}^{1}(denote it by 𝔪1\mathfrak{m}\sim\mathcal{H}^{1}), which is of the form 𝔪=ef1\mathfrak{m}=e^{-f}\mathcal{H}^{1}, and its density function ff is (K,N1)(K,N-1)-convex. Hence h:=efh:=e^{-f} satisfies (2.1).

Recall we say that two measures σ\sigma and τ\tau are equivalent to each other if both στ\sigma\ll\tau and τσ\tau\ll\sigma hold.

Remark 2.16.

(K,N)(K,N)-convexity of the density function is proven in [KL]. The improvement version of the convexity is proven by [CMonge, CMil]. The density function h:=efh:=e^{-f} is continuous.

2.4. Weyl’s law on finite dimensional compact 𝖱𝖢𝖣\mathsf{RCD} spaces

In this subsection, we always assume the metric measure space (X,d,𝔪)(X,d,\mathfrak{m}) is a compact 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space. As aforementioned before, all the spectrum of Δ-\Delta are eigenvalues, and 0=λ0<λ1λ20=\lambda_{0}<\lambda_{1}\leq\lambda_{2}\leq\cdots\rightarrow\infty holds. We define the counting function N(λ)N(\lambda) by

N(λ):=#{i;λiλ}.\displaystyle N(\lambda):=\#\left\{i\;;\;\lambda_{i}\leq\lambda\right\}.

Ambrosio, Honda, and Tewodrose [AHTweyl] proves the following result.

Theorem 2.17 ([AHTweyl] cf. [ZZweyl]).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be a compact 𝖱𝖢𝖣\mathsf{RCD} space with 𝖽𝗂𝗆𝖾𝗌𝗌(X,d,𝔪)=n\mathsf{dim}_{\mathsf{ess}}(X,d,\mathfrak{m})=n. Then

Xlimr+0r𝔪(Br(x))𝔪(dx)=limr+0Xr𝔪(Br(x))𝔪(dx)<\displaystyle\int_{X}\lim_{r\rightarrow+0}\frac{r}{\mathfrak{m}(B_{r}(x))}\,\mathfrak{m}(dx)=\lim_{r\rightarrow+0}\int_{X}\frac{r}{\mathfrak{m}(B_{r}(x))}\mathfrak{m}(dx)<\infty

if and only if

limλN(λ)λn/2=ωn(2π)nn(n).\displaystyle\lim_{\lambda\rightarrow\infty}\frac{N(\lambda)}{\lambda^{n/2}}=\frac{\omega_{n}}{(2\pi)^{n}}\mathcal{H}^{n}(\mathcal{R}_{n}).
Remark 2.18.

Independently [ZZweyl] also proves a similar result.

For kk\in\mathbb{N}, define the subset k\mathcal{R}_{k}^{*} by

k:={xk;limr0𝔪(Br(x))ωkrk(0,)}.\displaystyle\mathcal{R}_{k}^{*}:=\left\{x\in\mathcal{R}_{k}\;;\;~{}^{\exists}\lim_{r\downarrow 0}\frac{\mathfrak{m}(B_{r}(x))}{\omega_{k}r^{k}}\in(0,\infty)\right\}.

It is known that 𝔪(kk)=0\mathfrak{m}(\mathcal{R}_{k}\setminus\mathcal{R}_{k}^{*})=0 (see [AHTweyl]). In order to prove the main result, we need the following results.

Theorem 2.19 ([AHTweyl]).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be a compact 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space and k=𝖽𝗂𝗆𝖾𝗌𝗌(X,d,𝔪)k=\mathsf{dim}_{\mathsf{ess}}(X,d,\mathfrak{m}). Then we have

lim inft+0(tk/2ieλit)1(4π)k/2k(k)>0,\displaystyle\liminf_{t\rightarrow+0}\left(t^{k/2}\sum_{i}e^{-\lambda_{i}t}\right)\geq\frac{1}{(4\pi)^{k/2}}\mathcal{H}^{k}(\mathcal{R}_{k}^{*})>0,

where k\mathcal{H}^{k} is the kk-dimensional Hausdorff measure on (X,d)(X,d).

The so-called Abelian theorem is also important for our main result.

Theorem 2.20 (Abelian theorem cf.[AHTweyl]).

Let ν\nu be a nonnegative and σ\sigma-finite Borel measure on [0,)[0,\infty). Assume that there exist γ[0,)\gamma\in[0,\infty) and C[0,)C\in[0,\infty) such that

limaν([0,a])aγ=C.\displaystyle\lim_{a\rightarrow\infty}\frac{\nu([0,a])}{a^{\gamma}}=C.

Then

limt+0tγ[0,)etx𝑑ν(x)=CΓ(γ+1).\displaystyle\lim_{t\rightarrow+0}t^{\gamma}\int_{[0,\infty)}e^{-tx}\,d\nu(x)=C\Gamma(\gamma+1).

In the next section, we use the above theorem for ν=iδλi\nu=\sum_{i}\delta_{\lambda_{i}}. Note that

[0,)etx𝑑ν(x)=ieλit\displaystyle\int_{[0,\infty)}e^{-tx}\,d\nu(x)=\sum_{i}e^{-\lambda_{i}t}

in this case.

3. Proof of the main theorem

Now let (X,d,𝔪)(X,d,\mathfrak{m}) be a compact 𝖱𝖢𝖣(K,N)\mathsf{RCD}(K,N) space with 𝔪(X)=1\mathfrak{m}(X)=1. Without loss of generality, we may assume K<0K<0 and K=N1-K=N-1 for simplicity.

The equivalence between (1) and (2) is proven in [KL]. The implication (4) to (3), (3) to (5) are trivial.

(2) \Rightarrow (4): When X=S1(r)X=S^{1}(r), then the density function hh for 𝔪=h1\mathfrak{m}=h\mathcal{H}^{1} is continuous(see Remark 2.16). Since 𝔪1\mathfrak{m}\sim\mathcal{H}^{1} and the continuity of hh, hh never vanish. Put c:=infh>0c:=\inf h>0 and C:=suph<C:=\sup h<\infty. Since

0<r𝔪(Br(x))h(x)=rh(x)Br(x)h(t)𝑑trh(x)2rcC2c<\displaystyle 0<\frac{r}{\mathfrak{m}(B_{r}(x))}h(x)=\frac{rh(x)}{\int_{B_{r}(x)}h(t)\,dt}\leq\frac{rh(x)}{2rc}\leq\frac{C}{2c}<\infty

holds, we have

(3.1) limr+0Xr𝔪(Br(x))𝔪(dx)=Xlimr+0r𝔪(Br(x))𝔪(dx)<\displaystyle\lim_{r\rightarrow+0}\int_{X}\frac{r}{\mathfrak{m}(B_{r}(x))}\,\mathfrak{m}(dx)=\int_{X}\lim_{r\rightarrow+0}\frac{r}{\mathfrak{m}(B_{r}(x))}\,\mathfrak{m}(dx)<\infty

by the dominated convergence theorem.

Let us consider the case for X=[0,]X=[0,\ell]. Without loss of generality, we may assume =π\ell=\pi. When the density function hh for 𝔪=h1\mathfrak{m}=h\mathcal{H}^{1} has a positive minimum, then a similar argument as X=S1X=S^{1} implies the consequence. So, we assume h(0)=h(π)=0h(0)=h(\pi)=0. Take a small positive number r>0r>0 and fix it. For x[0,r)x\in[0,r), we have

h(t(x+r))1/N1\displaystyle h(t(x+r))^{1/N-1} σK,N1(1t)(x+r)h(0)1/N1+σK,N1(t)(x+r)h(x+r)1/N1\displaystyle\geq\sigma^{(1-t)}_{K,N-1}(x+r)h(0)^{1/N-1}+\sigma^{(t)}_{K,N-1}(x+r)h(x+r)^{1/N-1}
=sinh(t(x+r))sinh(x+r)h(x+r)1/N1.\displaystyle=\frac{\sinh(t(x+r))}{\sinh(x+r)}h(x+r)^{1/N-1}.

Thus we get

𝔪(Br(x))=0x+rh(y)𝑑y=(x+r)01h(t(x+r))𝑑t\displaystyle\mathfrak{m}(B_{r}(x))=\int_{0}^{x+r}h(y)\,dy=(x+r)\int_{0}^{1}h(t(x+r))\,dt
(x+r)01sinhN1(t(x+r))sinhN1(x+r)h(x+r)𝑑t.\displaystyle\geq(x+r)\int_{0}^{1}\frac{\sinh^{N-1}(t(x+r))}{\sinh^{N-1}(x+r)}h(x+r)\,dt.

Therefore

0<r𝔪(Br(x))rsinhN1(x+r)(x+r)h(x+r)(01sinhN1(s(x+r))𝑑s)1.\displaystyle 0<\frac{r}{\mathfrak{m}(B_{r}(x))}\leq\frac{r\sinh^{N-1}(x+r)}{(x+r)h(x+r)}\left(\int_{0}^{1}\sinh^{N-1}(s(x+r))\,ds\right)^{-1}.

We have

rh(x)𝔪(Br(x))\displaystyle\frac{rh(x)}{\mathfrak{m}(B_{r}(x))}
rh(x)sinhN1(x+r)(x+r)h(x+r)(01sinhN1(s(x+r))𝑑s)1\displaystyle\leq\frac{rh(x)\sinh^{N-1}(x+r)}{(x+r)h(x+r)}\left(\int_{0}^{1}\sinh^{N-1}(s(x+r))\,ds\right)^{-1}
=rx+rh(x)h(x+r)sinhN1(x+r)01sinhN1(s(x+r))𝑑s.\displaystyle=\frac{r}{x+r}\frac{h(x)}{h(x+r)}\frac{\sinh^{N-1}(x+r)}{\int_{0}^{1}\sinh^{N-1}(s(x+r))\,ds}.

By using the Taylor expansion for sinh\sinh, we have sinhN1(s(x+r))sN1(x+r)N1\sinh^{N-1}(s(x+r))\geq s^{N-1}(x+r)^{N-1} and sinhN1z2N1zN1\sinh^{N-1}z\leq 2^{N-1}z^{N-1} for sufficiently small z>0z>0, further, applying (2.1)(\ref{eq:hratiioest}) to hh on (0,x+2r)(0,x+2r), we obtain

h(x)h(x+r)sinhN1((x+2r)x)sinhN1((x+2r)(x+r))=sinhN1(2r)sinhN1(r)(4r)N1rN1=4N1.\displaystyle\frac{h(x)}{h(x+r)}\leq\frac{\sinh^{N-1}((x+2r)-x)}{\sinh^{N-1}((x+2r)-(x+r))}=\frac{\sinh^{N-1}(2r)}{\sinh^{N-1}(r)}\leq\frac{(4r)^{N-1}}{r^{N-1}}=4^{N-1}.

Also we get

sinhN1(x+r)(01sinhN1(s(x+r))𝑑s)1\displaystyle\sinh^{N-1}(x+r)\left(\int_{0}^{1}\sinh^{N-1}(s(x+r))\,ds\right)^{-1}
2N1(x+r)N1((x+r)N1N)1\displaystyle\leq 2^{N-1}(x+r)^{N-1}\left(\frac{(x+r)^{N-1}}{N}\right)^{-1}
N2N1.\displaystyle\leq N2^{N-1}.

Finally we have

rh(x)𝔪(Br(x))N8N1.\displaystyle\frac{rh(x)}{\mathfrak{m}(B_{r}(x))}\leq N8^{N-1}.

On the other hand, for x(r,π/2)x\in(r,\pi/2), we have

𝔪(Br(x))=xrx+rh(y)𝑑y=01h((1t)(xr)+t(x+r))2r𝑑t\displaystyle\mathfrak{m}(B_{r}(x))=\int_{x-r}^{x+r}h(y)\,dy=\int_{0}^{1}h\left((1-t)(x-r)+t(x+r)\right)\cdot 2r\,dt
2r01(σK,N1(1t)(2r)h(xr)1/N1+σK,N1(t)(2r)h(x+r)1/N1)N1𝑑t\displaystyle\geq 2r\int_{0}^{1}\left(\sigma^{(1-t)}_{K,N-1}(2r)h(x-r)^{1/N-1}+\sigma^{(t)}_{K,N-1}(2r)h(x+r)^{1/N-1}\right)^{N-1}\,dt
2r01σK,N1(t)(2r)N1h(x+r)𝑑t\displaystyle\geq 2r\int_{0}^{1}\sigma^{(t)}_{K,N-1}(2r)^{N-1}h(x+r)\,dt
=2rh(x+r)sinhN1(2r)01sinhN1(2tr)𝑑t\displaystyle=\frac{2rh(x+r)}{\sinh^{N-1}\left(2r\right)}\int_{0}^{1}\sinh^{N-1}(2tr)\,dt
2rh(x+r)(4r)N101(2tr)N1𝑑t\displaystyle\geq\frac{2rh(x+r)}{(4r)^{N-1}}\int_{0}^{1}(2tr)^{N-1}\,dt
2rh(x+r)N2N1.\displaystyle\geq\frac{2rh(x+r)}{N2^{N-1}}.

Applying (2.1) to hh on (xr,x+2r)(x-r,x+2r), we get the estimate

h(x)h(x+r)(sinh((x+2r)x)sinh((x+2r)(x+r)))N1=sinhN1(2r)sinhN1(r)(4r)N1rN1=4N1.\displaystyle\frac{h(x)}{h(x+r)}\leq\left(\frac{\sinh((x+2r)-x)}{\sinh((x+2r)-(x+r))}\right)^{N-1}=\frac{\sinh^{N-1}(2r)}{\sinh^{N-1}(r)}\leq\frac{(4r)^{N-1}}{r^{N-1}}=4^{N-1}.

Then

rh(x)𝔪(Br(x))\displaystyle\frac{rh(x)}{\mathfrak{m}(B_{r}(x))} rh(x)N2N12rh(x+r)=N2N12h(x)h(xr)\displaystyle\leq\frac{rh(x)N2^{N-1}}{2rh(x+r)}=\frac{N2^{N-1}}{2}\cdot\frac{h(x)}{h(x-r)}
N2N124N1N8N1.\displaystyle\leq\frac{N2^{N-1}}{2}\cdot 4^{N-1}\leq N8^{N-1}.

The upper bound N8N1N8^{N-1} depends on neither x(0,π/2)x\in(0,\pi/2) nor r>0r>0. We apply the same argument for near x=πx=\pi. Then by the dominated convergence theorem,

(3.2) limr0Xr𝔪(Br(x))𝔪(dx)=Xlimr0r𝔪(Br(x))𝔪(dx)<.\displaystyle\lim_{r\downarrow 0}\int_{X}\frac{r}{\mathfrak{m}(B_{r}(x))}\,\mathfrak{m}(dx)=\int_{X}\lim_{r\downarrow 0}\frac{r}{\mathfrak{m}(B_{r}(x))}\,\mathfrak{m}(dx)<\infty.

Both cases, X=S1(r)X=S^{1}(r), [0,][0,\ell], we have (4) by combining (3.1) and (3.2) with Theorem 4.3 in [AHTweyl].

(5) \Rightarrow (1): By combining the assumption and Abelian theorem, we have

limt+0t1+α2ieλit=0.\displaystyle\lim_{t\rightarrow+0}t^{\frac{1+\alpha}{2}}\sum_{i}e^{-\lambda_{i}t}=0.

Let k=𝖽𝗂𝗆𝖾𝗌𝗌(X,d,𝔪)k=\mathsf{dim}_{\mathsf{ess}}(X,d,\mathfrak{m}). Then by Theorem 2.19, we obtain

limt+0t1+α2ieλit\displaystyle\lim_{t\rightarrow+0}t^{\frac{1+\alpha}{2}}\sum_{i}e^{-\lambda_{i}t} =0<1(4π)k/2k(k)\displaystyle=0<\frac{1}{(4\pi)^{k/2}}\mathcal{H}^{k}(\mathcal{R}_{k}^{*})
lim inft+0tk2ieλit.\displaystyle\leq\liminf_{t\rightarrow+0}t^{\frac{k}{2}}\sum_{i}e^{-\lambda_{i}t}.

This implies 1+α>k1+\alpha>k. Since α1\alpha\leq 1 and kk is an integer, kk 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.

References