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

The Aubry set for the XY model and typicality of periodic optimization for 22-locally constant potentials

Yuika Kajihara Department of Mathematics, Kyoto University, Kitashirakawa Oiwake-cho, Sakyo-ku, Kyoto, 606-8502, Japan kajihara.yuika.6f@kyoto-u.ac.jp Shoya Motonaga Department of Mathematical Sciences, Ritsumeikan University, 1-1-1 Nojihigashi, Kusatsu, Shiga 525-8577, Japan motonaga@fc.ritsumei.ac.jp  and  Mao Shinoda Department of Mathematics, Ochanomizu University, 2-1-1 Otsuka, Bunkyo-ku, Tokyo, 112-8610, Japan shinoda.mao@ocha.ac.jp
Abstract.

We consider the Aubry set for the XY model, symbolic dynamics ([0,1]0,σ)([0,1]^{\mathbb{N}_{0}},\sigma) with the uncountable symbol [0,1][0,1], and study its action-optimizing properties. Moreover, for a potential function that depends on the first two coordinates we obtain an explicit expression of the set of optimal periodic measures and a detailed description of the Aubry set. We also show the typicality of periodic optimization for 2-locally constant potentials with the twist condition. Our approach combines the weak KAM method for symbolic dynamics and variational techniques for twist maps.

2020 Mathematics Subject Classification:
Primary 37B10, 37E40, 37A99, 37J51

1. Introduction

This paper serves as a bridge between ergodic optimization for symbolic dynamics and variational problems for twist maps. More precisely, we consider symbolic dynamics with the uncountable symbols [0,1], the so-called XY model, and investigate the associated action-optimizing sets for Lipschitz continuous potentials, in particular 2-locally constant functions, from the view points of the weak KAM method and variational approaches. It is well known that, in Aubry-Mather theory for Euler-Lagrange flows [Mather91, Mane1997], optimizing invariant probability measures are closely related with optimizing curves through the principle of least action. Although our system has no variational structure, we will see the advantages of the concept of “optimizing orbits” as in [BLL13] and of variational techniques based on [Ban88] in ergodic optimization. Before presenting our main results, we briefly give the background of our study and some key notions.

For a continuous map TT on a compact metric space 𝒳\mathcal{X} and a continuous function (called as a potential) φ:𝒳\varphi:\mathcal{X}\rightarrow\mathbb{R}, ergodic optimization investigates the optimal (minimum) ergodic average

αφ=infμT(𝒳)φ𝑑μ\displaystyle\alpha_{\varphi}=\inf_{\mu\in\mathcal{M}_{T}(\mathcal{X})}\int\varphi d\mu

where T(𝒳)\mathcal{M}_{T}(\mathcal{X}) is the set of TT-invariant Borel probability measures on 𝒳\mathcal{X} endowed with the weak*-topology. An invariant measure which attains the minimum is called an optimizing (minimizing) measure for φ\varphi and denote by min(φ)\mathcal{M}_{{\rm min}}(\varphi) the set of optimizing measures for φ\varphi. Since T(𝒳)\mathcal{M}_{T}(\mathcal{X}) is compact and φd():T(𝒳)\int\varphi d(\cdot):\mathcal{M}_{T}(\mathcal{X})\to\mathbb{R} is continuous, min(φ)\mathcal{M}_{{\rm min}}(\varphi)\neq\emptyset. Note that we consider the minimum ergodic average instead of the maximum one in order to describe a more natural connection with the minimizing method in variational problems (see Section 4). We also remark that Jenkinson’s formula [Jenkinson19] for the optimal ergodic average:

(1) αφ=infx𝒳lim infn1nSnφ(x)=lim infninfx𝒳1nSnφ(x)\displaystyle\alpha_{\varphi}=\inf_{x\in\mathcal{X}}\liminf_{n\to\infty}\frac{1}{n}S_{n}\varphi(x)=\liminf_{n\to\infty}\inf_{x\in\mathcal{X}}\frac{1}{n}S_{n}\varphi(x)

where

(2) Snφ=i=0n1φTi.\displaystyle S_{n}\varphi=\sum_{i=0}^{n-1}\varphi\circ T^{i}.

The uniqueness of optimizing measures and the “shape (complexity)” of their support are fundamental questions of ergodic optimization. This leads to the definition of the Mather set

φ=μmin(φ)supp(μ),\displaystyle\mathscr{M}_{\varphi}=\bigcup_{\mu\in\mathcal{M}_{{\rm min}}(\varphi)}{\rm supp}(\mu),

where supp(μ){\rm supp}(\mu) is the intersection of all compact sets with full measure with respect to μ\mu. There are several results on the uniqueness of the optimizing measures and on the low complexity of the Mather set for “typical” functions (see [Jenkinson19] and the references therein for more details). However, it is worth pointing out that there are few specific examples whose optimizing measures are well understood.

One of the fundamental approaches to extract the detailed description of the Mather set is to consider (calibrated) subactions for potentials. We do not touch the details of this notion here (see Section  for the precise definition and properties). We only remark that a certain value of the level set of an associated function contains the Mather set, and thus the existence and uniqueness of (calibrated) subactions are also interesting problems.

Another (but closely related) approach to obtain more precise information about the Mather set is to investigate “optimizing orbits” for potentials. Inspired by the weak KAM theory [Fathi14] due to Fathi, [BLL13] introduced several important notions related to “optimizing orbits” for symbolic dynamics with a finite set of symbols. Borrowing their ideas presented in [BLL13], we first study an “action-optimizing set” of Lipschitz continuous potentials for symbolic dynamics whose symbol is the interval [0,1][0,1]. Let 0\mathbb{N}_{0} be the set of non-negative integers and let X=[0,1]0X=[0,1]^{\mathbb{N}_{0}}. Set the metric on XX by

d(x¯,y¯)=i=0|xiyi|2i\displaystyle d(\underline{x},\underline{y})=\sum_{i=0}^{\infty}\frac{|x_{i}-y_{i}|}{2^{i}}

for x¯,y¯X\underline{x},\underline{y}\in X where |||\cdot| is the Euclidean distance in the interval [0,1][0,1]. Let σ:XX\sigma:X\rightarrow X be the left shift, i.e., (σ(x¯))i=xi+1(\sigma(\underline{x}))_{i}=x_{i+1} for all i0i\in\mathbb{N}_{0}, and we call the topological dynamical system (X,σ)(X,\sigma) the XY model. See [CR, LM14] for its details. From now on, we consider ergodic optimization for the case 𝒳=X\mathcal{X}=X and T=σT=\sigma. For a Lipschitz continuous potential φ:X\varphi:X\to\mathbb{R}, we define two functions the Mañé potential Sφ(,)S_{\varphi}(\cdot,\cdot) and Peierl’s barrier Hφ(,)H_{\varphi}(\cdot,\cdot) on X×XX\times X as

Sφ(x¯,y¯)\displaystyle S_{\varphi}(\underline{x},\underline{y}) =limε0inf{Sn(φαφ)(z¯)n,z¯B(x¯,y¯,n;ε)},\displaystyle=\lim_{\varepsilon\to 0}\inf\{S_{n}(\varphi-\alpha_{\varphi})(\underline{z})\mid n\in\mathbb{N},\underline{z}\in B(\underline{x},\underline{y},n;\varepsilon)\},
Hφ(x¯,y¯;ε)\displaystyle H_{\varphi}(\underline{x},\underline{y};\varepsilon) =limε0lim infn{Sn(φαφ)(z¯)z¯B(x¯,y¯,n;ε)},\displaystyle=\lim_{\varepsilon\to 0}\liminf_{n\to\infty}\{S_{n}(\varphi-\alpha_{\varphi})(\underline{z})\mid\underline{z}\in B(\underline{x},\underline{y},n;\varepsilon)\},

where

B(x¯,y¯,n;ε)\displaystyle B(\underline{x},\underline{y},n;\varepsilon) ={z¯Xd(x¯,z¯)<ε,d(σn(z¯),y¯)<ε}.\displaystyle=\{\underline{z}\in X\mid d(\underline{x},\underline{z})<\varepsilon,d(\sigma^{n}(\underline{z}),\underline{y})<\varepsilon\}.

See Section 2 for the details of SφS_{\varphi} and HφH_{\varphi}. Then we define the Aubry set Ωφ\Omega_{\varphi} as the zero-level set of S~φ\tilde{S}_{\varphi}, where S~φ(x¯)=Sφ(x¯,x¯)\tilde{S}_{\varphi}(\underline{x})=S_{\varphi}(\underline{x},\underline{x}). We remark that Ωφ\Omega_{\varphi} includes the Mather set φ\mathscr{M}_{\varphi} of φ\varphi.

Now we give our first main results concerning Peierl’s barrier. A positive-semi orbit {σn(x)}n0\{\sigma^{n}(x)\}_{n\in\mathbb{N}_{0}} is said to be φ\varphi-static if for any non-negative integers i<ji<j it holds that

n=ij1(φσn(x¯)αφ)=Sφ(σj(x¯),σi(x¯)).\sum_{n=i}^{j-1}\Big{(}\varphi\circ\sigma^{n}(\underline{x})-\alpha_{\varphi}\Big{)}=-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x})).

Define

Aφ\displaystyle A_{\varphi} :={σk(x¯)X{σn(x¯)}n0isφ-static,k0}.\displaystyle:=\{\sigma^{k}(\underline{x})\in X\mid\{\sigma^{n}(\underline{x})\}_{n\in\mathbb{N}_{0}}\ \text{is}\ \varphi\text{-static},k\in\mathbb{N}_{0}\}.

Note that this definition is motivated by Mañé’s work [Mane1997] for Lagrangian systems. See Section 3 for their details. We then obtain the following characterizations of the Aubry set and Peierl’s barrier as in [BLL13, Mane1997].

Main Theorem 1.

Let φ:X\varphi:X\rightarrow\mathbb{R} be a Lipschitz function and define H~φ(x¯)=Hφ(x¯,x¯)\tilde{H}_{\varphi}(\underline{x})={H}_{\varphi}(\underline{x},\underline{x}). Then, we have

Ωφ=H~φ1({0})=Aφ.\Omega_{\varphi}=\tilde{H}_{\varphi}^{-1}(\{0\})=A_{\varphi}.
Main Theorem 2.

For any x¯Ωφ\underline{x}\in\Omega_{\varphi}, Hφ(x¯,):X{H}_{\varphi}(\underline{x},\cdot)\colon X\to\mathbb{R} is a Lipschitz calibrated subaction. Moreover, the relation x¯y¯\underline{x}\sim\underline{y} given by

Hφ(x¯,y¯)+Hφ(y¯,x¯)=0H_{\varphi}(\underline{x},\underline{y})+H_{\varphi}(\underline{y},\underline{x})=0

is an equivalence relation if both x¯\underline{x} and y¯\underline{y} belong to Ωφ\Omega_{\varphi}.

Next, we restrict our attention to 2-locally constant potentials and make use of variational techniques. Although “action-optimizing sets” play important roles in the study of Lagrangian systems as well as symbolic dynamics with finite symbols, it is difficult to obtain explicit information about these sets in most cases. On the other hand, for the case of area-preserving twist maps on an annulus, Aubry and Mather originally developed their theory and it provides much detailed descriptions of “minimal” orbits (originally appeared in [Morse24] under the name “Class A”), i.e., minimizers under arbitrary two-point boundary conditions. There are several papers by them related to twist maps; for example, see [Aubry83, AD83, Mather82, Mather89, Mather91]. The best general reference here is Bangert’s survey article [Ban88]. Following [Ban88], we assume the twist condition for our 2-locally constant potentials and give the following result.

Main Theorem 3.

Suppose that φ(x¯)=h(x0,x1)\varphi(\underline{x})=h(x_{0},x_{1}) with D2D1h<0D_{2}D_{1}h<0, where h:[0,1]2h:[0,1]^{2}\to\mathbb{R} is a C2C^{2}-function on [0,1]2[0,1]^{2} and DiD_{i} means derivative for the ii-th component for i=1,2i=1,2. Let h=minx[0,1]h(x,x)h^{\ast}=\min_{x\in[0,1]}h(x,x) and m={a[0,1]h(a,a)=h}\mathrm{m}=\{a\in[0,1]\mid h(a,a)=h^{\ast}\}. Then we have (1)-(3):

  • (1)

    αφ=h\alpha_{\varphi}=h^{\ast}.

  • (2)

    min(φ)p={δaam}\mathcal{M}_{{\rm min}}(\varphi)\cap\mathcal{M}^{\mathrm{p}}=\{\delta_{a^{\infty}}\mid a\in\mathrm{m}\}, where δx¯\delta_{\underline{x}} is the Dirac measure supported at x¯\underline{x} and p\mathcal{M}^{\mathrm{p}} stands for the set of invariant probability measures supported on a single periodic orbit.

  • (3)

    Ωφm0\Omega_{\varphi}\subset\mathrm{m}^{\mathbb{N}_{0}}, i.e., for any x¯={xi}i0Ωφ\underline{x}=\{x_{i}\}_{i\in\mathbb{N}_{0}}\in\Omega_{\varphi} we have

    h(xi,xi)=hfor alli0.h(x_{i},x_{i})=h^{\ast}\ \text{for all}\ i\in\mathbb{N}_{0}.

If, in addition, h(x,x)h(x,x) has a unique minimum point aa_{\ast} in [0,1][0,1], then the Mather set of φ\varphi coincides with the Aubry set of φ\varphi and it consists of the single fixed point aa_{\ast}^{\infty}.

We call the assumption D2D1h<0D_{2}D_{1}h<0 the twist condition. Note that Main theorem 3 holds under weaker assumptions for Lipschitz continuous hh on [0,1]2[0,1]^{2} (see Section 4 for the details). We emphasize that, for 2-locally constant potentials with the twist condition, Main theorem 3 provides the explicit formulas of the optimal ergodic average and of the set of optimizing periodic measures, and in some cases it also completely determines the Mather set and the Aubry set.

Finally we turn to the typically periodic optimization (TPO) problem in the class of 22-locally constant potentials for the XY model. In the field of ergodic optimization, for “chaotic” systems, it is conjectured that the optimizing measure for a “typical” potential with a suitable regularity is a periodic measure, i.e., supported on a single periodic orbit. Many authors investigate the “typicality” of the periodic minimizing measures in many contexts (see [Jenkinson19] for more details). Recently Gao et.al established a TPO property of real analytic expanding circle maps for smooth potentials [GSZ]. Our result is also formulated for smooth potentials as described below. Consider the set of CrC^{r}-functions (r2r\geq 2) with the twist condition,

r={hCr([0,1]2;)D2D1h<0},\mathscr{H}^{r}=\{h\in C^{r}([0,1]^{2};\mathbb{R})\mid D_{2}D_{1}h<0\},

equipped with the CrC^{r}-norm. Using Theorem 3, we obtain the following TPO property.

Main Theorem 4 (TPO property for the XY model in the class of 2-locally constant functions).

Let r2r\geq 2 be an integer. For the XY model, we have the followings:

  • (i)

    There is a CrC^{r} open dense subset 𝒪\mathscr{O} in r\mathscr{H}^{r} such that for each h𝒪h\in\mathscr{O} the Mather set and the Aubry set of hh consist of a single fixed point.

  • (ii)

    For arbitrary hrh\in\mathscr{H}^{r} there is a CrC^{r} open dense subset 𝒱h\mathscr{V}_{h} in Cr([0,1];)C^{r}([0,1];\mathbb{R}) such that for each V𝒱hV\in\mathscr{V}_{h} the Mather set and the Aubry set of h+Vh+V consist of a single fixed point.

The structure of this paper is as follows. In Section 2, we extend the definitions of the Mañé potential and Peierl’s barrier for symbolic dynamics with finite alphabets to our setting. We also confirm that these functions satisfy similar properties presented in [BLL13]. In Section 3, we derives a characterization of the Aubry set in terms of Mañé’s action-optimizing approach [Mane1997]. In Section 4, we consider 2-locally constant potentials under several assumptions and determine the set of optimizing periodic measures as well as the elements in the Aubry set. Finally, we verify that the TPO property holds within our framework in Section 5.

2. Mañé potential and Peierl’s barrier

In this section, following [BLL13], we consider the Mañé potential, Peierl’s barrier, and the Aubry set. We begin with the definition of the Mañé potential.

Definition 2.1 (Mañé potential).

For φ:X\varphi:X\rightarrow\mathbb{R} and ε>0\varepsilon>0 define Sφ:X×X{}S_{\varphi}:X\times X\rightarrow\mathbb{R}\cup\{\infty\} by

Sφ(x¯,y¯;ε)=inf{Sn(φαφ)(z¯)n,z¯B(x¯,y¯,n;ε)}\displaystyle S_{\varphi}(\underline{x},\underline{y};\varepsilon)=\inf\{S_{n}(\varphi-\alpha_{\varphi})(\underline{z})\mid n\in\mathbb{N},\underline{z}\in B(\underline{x},\underline{y},n;\varepsilon)\}

where

B(x¯,y¯,n;ε)={z¯Xd(x¯,z¯)<ε,d(σn(z¯),y¯)<ε}.\displaystyle B(\underline{x},\underline{y},n;\varepsilon)=\{\underline{z}\in X\mid d(\underline{x},\underline{z})<\varepsilon,d(\sigma^{n}(\underline{z}),\underline{y})<\varepsilon\}.

We define the Mañé potential SφS_{\varphi} by

Sφ(x¯,y¯)=limε0Sφ(x¯,y¯;ε),\displaystyle S_{\varphi}(\underline{x},\underline{y})=\lim_{\varepsilon\to 0}S_{\varphi}(\underline{x},\underline{y};\varepsilon),
Remark 2.2.

The Mañé potential originates from the context of the Aubry-Mather theory for Euler-Lagrange flows. Mañé [Mane1997] considered a function ϕ:M×M\phi\colon M\times M\to\mathbb{R} defined by

(3) ϕ(x,y)=infT>0infγC(x,y;T)0T(L(γ,γ˙)α~φ)𝑑t,\displaystyle\phi(x,y)=\inf_{T>0}\inf_{\gamma\in C(x,y;T)}\int_{0}^{T}(L(\gamma,\dot{\gamma})-\tilde{\alpha}_{\varphi})dt,

where MM is a closed Riemannian manifold, L:TML:TM\to\mathbb{R} is a Tonelli Lagrangian (see [Mane1997] for the precise definition), C(x,y;T)C(x,y;T) is the set of absolutely continuous curves γ:M\gamma\colon\mathbb{R}\to M with γ(0)=x,γ(T)=y\gamma(0)=x,\gamma(T)=y, and α~φ\tilde{\alpha}_{\varphi} is given by

α~φ=infμΦt(L)L(γ,γ˙)𝑑μ,\tilde{\alpha}_{\varphi}=\inf_{\mu\in\mathcal{M}_{\Phi^{t}(L)}}\int L(\gamma,\dot{\gamma})d\mu,

with the Euler-Lagrange flow Φt(L)\Phi^{t}(L). The right-side of (3)\eqref{eq:mane-el} corresponds to a minimizing method that provides trajectories with the energy α~φ\tilde{\alpha}_{\varphi} in the two-point boundary value problem. [BLL13] has rewritten this concept in the context of ergodic optimization for symbolic dynamics with finite symbols, and Definition 2.1 is an analogy of their definition.

Definition 2.3 (Aubry set).

The set Ωφ={x¯XSφ(x¯,x¯)=0}\Omega_{\varphi}=\{\underline{x}\in X\mid S_{\varphi}(\underline{x},\underline{x})=0\} is called the Aubry set of φ\varphi.

Note that we will see that SφS_{\varphi} does not take -\infty in Lemma 2.6 if φ\varphi is Lipschitz. The following notion, called (calibrated) subation, is important as a technical tool for ergodic optimization.

Definition 2.4 (Subaction, calibrated subaction).

A continuous function u:Xu:X\rightarrow\mathbb{R} is called a subaction of φ\varphi if

u(x¯)+φ(x¯)u(σx¯)+αφ\displaystyle u(\underline{x})+\varphi(\underline{x})\geq u(\sigma\underline{x})+\alpha_{\varphi}

for every x¯X\underline{x}\in X. Moreover, a subaction uu is called calibrated if

minσ(y¯)=x¯(φ(y¯)+u(y¯))=u(x¯)+αφ\displaystyle\min_{\sigma(\underline{y})=\underline{x}}(\varphi(\underline{y})+u(\underline{y}))=u(\underline{x})+\alpha_{\varphi}

for every x¯X\underline{x}\in X.

Remark 2.5.

There exists a calibrated subaction for a Walters function φ\varphi on a weakly-expanding topological dynamical system by [Bou01]. Here, ‘Walters function’ is a broader class of functions that includes Holder continuous functions. Moreover, we can get Lipschits calibrated subaction for Lipschitz φ\varphi. See [BCLMS11] for the details.

The next lemma gives the lower bound of SφS_{\varphi} using subactions.

Lemma 2.6.

Assume φ:X\varphi:X\rightarrow\mathbb{R} be Lipschitz. For a Lipschitz subaction uu of φ\varphi we have

(4) Sφ(x¯,y¯)u(y¯)u(x¯)\displaystyle S_{\varphi}(\underline{x},\underline{y})\geq u(\underline{y})-u(\underline{x})

for every x¯,y¯X\underline{x},\underline{y}\in X.

Proof.

Since φ\varphi is Lipschitz, its calibrated subaction uu is also Lipschitz. Then we have

φ(x¯)αφuσ(x¯)u(x¯)\varphi(\underline{x})-\alpha_{\varphi}\geq u\circ\sigma(\underline{x})-u(\underline{x})

for all x¯X\underline{x}\in X. Fix ε>0\varepsilon>0. Take n1n\geq 1 and z¯B(x¯,y¯,n;ε)\underline{z}\in B(\underline{x},\underline{y},n;\varepsilon). Then we have

Sn(φαφ)(z¯)\displaystyle S_{n}(\varphi-\alpha_{\varphi})(\underline{z}) Sn(uσu)(z¯)\displaystyle\geq S_{n}(u\circ\sigma-u)(\underline{z})
=u(σn(z¯))u(z¯)\displaystyle=u(\sigma^{n}(\underline{z}))-u(\underline{z})
u(y¯)u(x¯)Luε\displaystyle\geq u(\underline{y})-u(\underline{x})-L_{u}\varepsilon

and

Sφ(x¯,y¯;ε)u(y¯)u(x¯)Luε,S_{\varphi}(\underline{x},\underline{y};\varepsilon)\geq u(\underline{y})-u(\underline{x})-L_{u}\varepsilon,

where LuL_{u} is the Lipschitz constant of uu. Letting ε0\varepsilon\to 0, we have

Sφ(x¯,y¯)u(y¯)u(x¯).S_{\varphi}(\underline{x},\underline{y})\geq u(\underline{y})-u(\underline{x}).

Proposition 2.7.

Sφ(x¯,y¯)S_{\varphi}(\underline{x},\underline{y}) is lower semicontinuous on X×XX\times X.

Proof.

Fix y¯,z¯,w¯X\underline{y},\underline{z},\underline{w}\in X and ε>0\varepsilon>0. Since

d(σn(w¯),y¯)d(σn(w¯),z¯)+d(y¯,z¯)ε+d(y¯,z¯),d(\sigma^{n}(\underline{w}),\underline{y})\leq d(\sigma^{n}(\underline{w}),\underline{z})+d(\underline{y},\underline{z})\leq\varepsilon+d(\underline{y},\underline{z}),

it holds that B(x¯,y¯,n;ε+d(y¯,z¯))B(x¯,z¯,n;ε)B(\underline{x},\underline{y},n;\varepsilon+d(\underline{y},\underline{z}))\supset B(\underline{x},\underline{z},n;\varepsilon) and thus we obtain

Sφ(x¯,y¯;ε+d(y¯,z¯))Sφ(x¯,z¯;ε).S_{\varphi}(\underline{x},\underline{y};\varepsilon+d(\underline{y},\underline{z}))\leq S_{\varphi}(\underline{x},\underline{z};\varepsilon).

Moreover, we have

lim infz¯y¯(limε0Sφ(x¯,z¯;ε))=lim infz¯y¯Sφ(x¯,z¯)\liminf_{\underline{z}\to\underline{y}}\left(\lim_{\varepsilon\to 0}S_{\varphi}(\underline{x},\underline{z};\varepsilon)\right)=\liminf_{\underline{z}\to\underline{y}}S_{\varphi}(\underline{x},\underline{z})

and

lim infz¯y¯(limε0Sφ(x¯,y¯;ε+d(y¯,z¯)))=limε0Sφ(x¯,y¯;ε))=Sφ(x¯,y¯).\liminf_{\underline{z}\to\underline{y}}\left(\lim_{\varepsilon\to 0}S_{\varphi}(\underline{x},\underline{y};\varepsilon+d(\underline{y},\underline{z}))\right)=\lim_{\varepsilon^{\prime}\to 0}S_{\varphi}(\underline{x},\underline{y};\varepsilon^{\prime}))=S_{\varphi}(\underline{x},\underline{y}).

Thus it holds that

Sφ(x¯,y¯)lim infz¯y¯Sφ(x¯,z¯).S_{\varphi}(\underline{x},\underline{y})\leq\liminf_{\underline{z}\to\underline{y}}S_{\varphi}(\underline{x},\underline{z}).

Therefore, the map y¯Sφ(x¯,y¯)\underline{y}\mapsto S_{\varphi}(\underline{x},\underline{y}) is lower semicontinuous for each x¯X\underline{x}\in X. Similarly, we can verify that the map y¯Sφ(y¯,x¯)\underline{y}\mapsto S_{\varphi}(\underline{y},\underline{x}) is also lower semicontinuous for each x¯X\underline{x}\in X. ∎

The following proposition asserts that the Mather set is included in the Aubry set.

Proposition 2.8.

Let φ:X\varphi:X\to\mathbb{R} be a Lipschitz function. Then the Mather set φ\mathscr{M}_{\varphi} is a subset of the Aubry set Ωφ\Omega_{\varphi}.

Proof.

By the ergodic decomposition, it is sufficient to show that supp(μ)Ωφ\mathrm{supp}(\mu)\subset\Omega_{\varphi} for each ergodic μmin(φ)\mu\in\mathcal{M}_{{\rm min}}(\varphi). Take arbitrary ergodic μmin(φ)\mu\in\mathcal{M}_{{\rm min}}(\varphi) and a subaction uu for φ\varphi. Note that φ𝑑μ=αφ\int\varphi d\mu=\alpha_{\varphi}. Letting φu=φαφ+uuσ\varphi^{u}=\varphi-\alpha_{\varphi}+u-u\circ\sigma, we obtain

φu𝑑μ=(φαφ)𝑑μ=0\displaystyle\int\varphi^{u}d\mu=\int(\varphi-\alpha_{\varphi})d\mu=0

since uσ𝑑μ=u𝑑μ\int u\circ\sigma\ d\mu=\int u\ d\mu by the σ\sigma-invariance of μ\mu. From the definition of uu, it holds that φu0\varphi^{u}\geq 0, which implies that φu(x¯)=0\varphi^{u}(\underline{x})=0 for μ\mu-a.e. x¯X\underline{x}\in X. Since μ\mu is ergodic and φu\varphi^{u} is continuous, we obtain φu(x¯)=0\varphi^{u}(\underline{x})=0 on supp(μ)\mathrm{supp}(\mu). Hence, we see that

Sn(φαφ)(x¯)=uσn(x¯)u(x¯)S_{n}(\varphi-\alpha_{\varphi})(\underline{x})=u\circ\sigma^{n}(\underline{x})-u(\underline{x})

for each x¯supp(μ)\underline{x}\in\mathrm{supp}(\mu) and nn\in\mathbb{N}. Note that σ\sigma-invariance of supp(μ)\mathrm{supp}(\mu) implies σn(x¯)supp(μ)\sigma^{n}(\underline{x})\in\mathrm{supp}(\mu) for nn\in\mathbb{N} if x¯supp(μ)\underline{x}\in\mathrm{supp}(\mu). By Poincaré’s recurrence theorem, for μ\mu-a.e. x¯\underline{x}, there exists a monotone increasing sequence {nk}k\{n_{k}\}_{k\in\mathbb{N}} with nk+n_{k}\to+\infty as k+k\to+\infty such that d(x¯,σnk(x¯))<εd(\underline{x},\sigma^{n_{k}}(\underline{x}))<\varepsilon. This implies that, for μ\mu-a.e. x¯\underline{x}, we have

Sφ(x¯,x¯;ε)Snk(φαφ)(x¯)=uσnk(x¯)u(x¯)Lφd(x¯,σnkx¯)<Lφε,S_{\varphi}(\underline{x},\underline{x};\varepsilon)\leq S_{n_{k}}(\varphi-\alpha_{\varphi})(\underline{x})=u\circ\sigma^{n_{k}}(\underline{x})-u(\underline{x})\leq L_{\varphi}d(\underline{x},\sigma^{n_{k}}\underline{x})<L_{\varphi}\varepsilon,

i.e., Sφ(x¯,x¯)0S_{\varphi}(\underline{x},\underline{x})\leq 0. Therefore, by the lower semicontinuity of SφS_{\varphi} (Proposition 2.7) and the density of μ\mu-a.e. points in supp(μ)\mathrm{supp}(\mu), it holds that Sφ(x¯,x¯)0S_{\varphi}(\underline{x},\underline{x})\leq 0 on supp(μ)\mathrm{supp}(\mu). Using Lemma 2.6, we have Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0 on supp(μ)\mathrm{supp}(\mu), which completes the proof. ∎

Next, let us define Peierl’s barrier.

Definition 2.9 (Peierl’s barrier).

For a Lipschitz function φ:X\varphi:X\rightarrow\mathbb{R} and ε>0\varepsilon>0, define Hφ:X×X{}H_{\varphi}:X\times X\rightarrow\mathbb{R}\cup\{\infty\} by

Hφ(x¯,y¯;ε)=lim infn{Sn(φαφ)(z¯)z¯B(x¯,y¯,n;ε)}\displaystyle H_{\varphi}(\underline{x},\underline{y};\varepsilon)=\liminf_{n\to\infty}\{S_{n}(\varphi-\alpha_{\varphi})(\underline{z})\mid\underline{z}\in B(\underline{x},\underline{y},n;\varepsilon)\}

and

Hφ(x¯,y¯)=limε0Hφ(x¯,y¯;ε).\displaystyle H_{\varphi}(\underline{x},\underline{y})=\lim_{\varepsilon\to 0}H_{\varphi}(\underline{x},\underline{y};\varepsilon).
Remark 2.10.

Note that the Peierl’s barrier defined as above may take \infty. Indeed, for φ(x¯)=x0\varphi(\underline{x})=x_{0}, we see that Hφ(1,1)=H_{\varphi}(1^{\infty},1^{\infty})=\infty in the following way. Fix k1k\geq 1. Take nk+1n\geq k+1 and z¯X\underline{z}\in X such that d(1,z¯)<2(k+1)d(1^{\infty},\underline{z})<2^{-(k+1)} and d(σn(z¯),1)<2(k+1)d(\sigma^{n}(\underline{z}),1^{\infty})<2^{-(k+1)}. Since αφ=0\alpha_{\varphi}=0, we have

Snφ(z¯)nαφ\displaystyle S_{n}\varphi(\underline{z})-n\alpha_{\varphi} =i=0k1(zi1)+i=0k1(1αφ)+i=kn1(ziαφ)\displaystyle=\sum_{i=0}^{k-1}(z_{i}-1)+\sum_{i=0}^{k-1}(1-\alpha_{\varphi})+\sum_{i=k}^{n-1}(z_{i}-\alpha_{\varphi})
1+k\displaystyle\geq-1+k

and Hφ(1,1;2k)1+kH_{\varphi}(1^{\infty},1^{\infty};2^{-k})\geq-1+k for all k1k\geq 1. Letting kk\to\infty, we have Hφ(1,1)=H_{\varphi}(1^{\infty},1^{\infty})=\infty. Similarly, we can show that any point (x¯,y¯)X×X(\underline{x},\underline{y})\in X\times X of the form (x¯,y¯)=(a0al1,b0bm1)(\underline{x},\underline{y})=(a_{0}\ldots a_{l}1^{\infty},b_{0}\ldots b_{m}1^{\infty}) provides Hφ(x¯,y¯)=H_{\varphi}(\underline{x},\underline{y})=\infty. Note that any cylinder set

[a0,,al]×[b0,,bm]={(x¯,y¯)xi=ai(i=0,,l),yj=bj(j=0,,m)}[a_{0},\ldots,a_{l}]\times[b_{0},\ldots,b_{m}]=\{(\underline{x},\underline{y})\mid x_{i}=a_{i}\ (i=0,\ldots,l),y_{j}=b_{j}\ (j=0,\ldots,m)\}

contains {(a1al1,b1bm1)}\{(a_{1}\ldots a_{l}1^{\infty},b_{1}\ldots b_{m}1^{\infty})\} and this implies that the function

(x¯,y¯)X×XHφ(x¯,y¯){+}(\underline{x},\underline{y})\in X\times X\mapsto H_{\varphi}(\underline{x},\underline{y})\in\mathbb{R}\cup\{+\infty\}

takes ++\infty on a dense set in X×XX\times X for the case φ(x¯)=x0\varphi(\underline{x})=x_{0}. We remark that a similar phenomenon occurs even for a subshift with a finite alphabet, a point which seems to have been not explicitly discussed in the literature, e.g., [BLL13]. This omission, however, does not affect other arguments in [BLL13]. Below, we provide a sufficient condition for the finiteness of the Peierls barrier.

Now let us prove a part of Main Theorem 1. Note that the first equality of Main Theorem 1 says that the Aubry set defined in Definition 2.3 coincides with the zero-level set derived from Peierl’s barrier.

Theorem 2.11 (cf. Main Theorem 1).

Let φ:X\varphi:X\rightarrow\mathbb{R} be a Lipschitz function. For x¯Ωφ\underline{x}\in\Omega_{\varphi} and y¯X\underline{y}\in X we have Hφ(x¯,y¯)Lφd(x¯,y¯)H_{\varphi}(\underline{x},\underline{y})\leq L_{\varphi}d(\underline{x},\underline{y}). In particular, x¯Ωφ\underline{x}\in\Omega_{\varphi} holds if and only if Hφ(x¯,x¯)=0H_{\varphi}(\underline{x},\underline{x})=0.

Proof.

The second claim immediately follows from the first claim and

Hφ(x¯,x¯)Sφ(x¯,x¯)0H_{\varphi}(\underline{x}^{\prime},\underline{x}^{\prime})\geq S_{\varphi}(\underline{x}^{\prime},\underline{x}^{\prime})\geq 0

for all x¯X\underline{x}^{\prime}\in X by Lemma 2.6. Now we prove the first claim. Since x¯Ωφ\underline{x}\in\Omega_{\varphi}, we have Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0. It suffices to consider the following two cases:

  1. (1)

    For any θ>0\theta>0, there exists ε(0,θ)\varepsilon\in(0,\theta) such that

    Hφ(x¯,x¯;ε)=Sφ(x¯,x¯;ε)\displaystyle H_{\varphi}(\underline{x},\underline{x};\varepsilon)=S_{\varphi}(\underline{x},\underline{x};\varepsilon)
  2. (2)

    There exists θ>0\theta>0 such that for any ε(0,θ)\varepsilon\in(0,\theta), there exists a finite number N=N(ε)N=N(\varepsilon) such that

    Sφ(x¯,x¯;ε)=inf{SN(φαφ)(z¯):z¯B(x¯,x¯,N;ε)}.\displaystyle S_{\varphi}(\underline{x},\underline{x};\varepsilon)=\inf\{S_{N}(\varphi-\alpha_{\varphi})(\underline{z}):\underline{z}\in B(\underline{x},\underline{x},N;\varepsilon)\}.

The proof of Case (1)(1): Fix θ>0\theta>0. By (1) and Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0, we may assume there exists ε(0,θ/2)\varepsilon\in(0,\theta/2) such that Hφ(x¯,x¯;ε)=Sφ(x¯,x¯;ε)<θ/2H_{\varphi}(\underline{x},\underline{x};\varepsilon)=S_{\varphi}(\underline{x},\underline{x};\varepsilon)<\theta/2. Hence there exist an increasing sequence {ni}\{n_{i}\} with 2n1<ε2^{-n_{1}}<\varepsilon and z¯(i)B(x¯,x¯,ni;ε)\underline{z}^{(i)}\in B(\underline{x},\underline{x},n_{i};\varepsilon) such that Sni(φαφ)(z¯(i))<θS_{n_{i}}(\varphi-\alpha_{\varphi})(\underline{z}^{(i)})<\theta. Define w¯(ni)X\underline{w}^{(n_{i})}\in X by

w¯(ni)=z0(ni)z1(ni)zni1(ni)y¯.\displaystyle\underline{w}^{(n_{i})}=z_{0}^{(n_{i})}z_{1}^{(n_{i})}\cdots z_{n_{i}-1}^{(n_{i})}\underline{y}.

Then for each i1i\geq 1 we have

Sni(φαφ)(w¯(ni))\displaystyle S_{n_{i}}(\varphi-\alpha_{\varphi})(\underline{w}^{(n_{i})}) =Sni(φαφ)(z¯)+i=0ni1(φσi(w¯(ni))φσi(z¯))\displaystyle=S_{n_{i}}(\varphi-\alpha_{\varphi})(\underline{z})+\sum_{i=0}^{n_{i}-1}(\varphi\circ\sigma^{i}(\underline{w}^{(n_{i})})-\varphi\circ\sigma^{i}(\underline{z}))
θ+Lφi=0ni1d(σi(w¯(ni)),σi(z¯))\displaystyle\leq\theta+L_{\varphi}\sum_{i=0}^{n_{i}-1}d(\sigma^{i}(\underline{w}^{(n_{i})}),\sigma^{i}(\underline{z}))
θ+Lφi=1nid(y¯,σniz¯)2i\displaystyle\leq\theta+L_{\varphi}\sum_{i=1}^{n_{i}}\frac{d(\underline{y},\sigma^{n_{i}}\underline{z})}{2^{i}}
θ+Lφ(d(x¯,y¯)+ε)\displaystyle\leq\theta+L_{\varphi}(d(\underline{x},\underline{y})+\varepsilon)
<(1+Lφ/2)θ+Lφd(x¯,y¯).\displaystyle<(1+L_{\varphi}/2)\theta+L_{\varphi}d(\underline{x},\underline{y}).

It is easy to see d(x¯,w¯(ni))2εd(\underline{x},\underline{w}^{(n_{i})})\leq 2\varepsilon, d(σniw¯(ni),y¯)=0d(\sigma^{n_{i}}\underline{w}^{(n_{i})},\underline{y})=0 and w¯(ni)B(x¯,y¯,ni;2ε)B(x¯,y¯,ni;θ)\underline{w}^{(n_{i})}\in B(\underline{x},\underline{y},n_{i};2\varepsilon)\subset B(\underline{x},\underline{y},n_{i};\theta). Hence we have

Hφ(x¯,y¯;θ)(1+Lφ/2)θ+Lφd(x¯,y¯).\displaystyle H_{\varphi}(\underline{x},\underline{y};\theta)\leq(1+L_{\varphi}/2)\theta+L_{\varphi}d(\underline{x},\underline{y}).

Then letting θ0\theta\to 0 we have

Hφ(x¯,y¯)Lφd(x¯,y¯).\displaystyle H_{\varphi}(\underline{x},\underline{y})\leq L_{\varphi}d(\underline{x},\underline{y}).

The proof of Case (2)(2): Fix ε(0,θ)\varepsilon\in(0,\theta). Set a monotone decreasing positive sequence {εi}\{\varepsilon_{i}\} satisfying

iεi<ε.\sum_{i\in\mathbb{N}}\varepsilon_{i}<\varepsilon.

For {εi}\{\varepsilon_{i}\}, we define a positive integer sequence {Ni}\{N_{i}\} by Ni=N(εi)N_{i}=N(\varepsilon_{i}).

If {Ni}\{N_{i}\} is bounded, then there exist an integer MM and an infinite subsequence {Nij}\{N_{i_{j}}\} such that M=NijM=N_{i_{j}} for any jj\in\mathbb{N}. Then we can take a sequence {z¯(j)}\{\underline{z}^{(j)}\} such that z¯(j)B(x¯,x¯,M;εij)\underline{z}^{(j)}\in B(\underline{x},\underline{x},M;\varepsilon_{i_{j}}) and

limjSM(φαφ)(z¯(j))=0\lim_{j\to\infty}S_{M}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})=0

since x¯Ωφ\underline{x}\in\Omega_{\varphi} and εij0\varepsilon_{i_{j}}\to 0 as jj\to\infty. This yields σM(x¯)=x¯\sigma^{M}(\underline{x})=\underline{x} because if d(σM(x¯),x¯)>δd(\sigma^{M}(\underline{x}),\underline{x})>\delta for some δ>0\delta>0, then for εij<δ2(M+2)\varepsilon_{i_{j}}<\delta 2^{-(M+2)}, we obtain

δ<d(σM(x¯),σMz¯(j))+d(σMz¯(j),x¯)<2M+1d(σMz¯(j),x¯)<δ/2,\delta<d(\sigma^{M}(\underline{x}),\sigma^{M}\underline{z}^{(j)})+d(\sigma^{M}\underline{z}^{(j)},\underline{x})<2^{M+1}d(\sigma^{M}\underline{z}^{(j)},\underline{x})<\delta/2,

which is contradiction. Then we get

|SM(φαφ)(x¯)SM(φαφ)(z¯(j))|\displaystyle|S_{M}(\varphi-\alpha_{\varphi})(\underline{x})-S_{M}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})| Lφk=0M1d(σk(x¯),σk(z¯(j)))\displaystyle\leq L_{\varphi}\sum_{k=0}^{M-1}d(\sigma^{k}(\underline{x}),\sigma^{k}(\underline{z}^{(j)}))
Lφk=0M12kd(x¯,z¯(j))<Lφ2Mεij.\displaystyle\leq L_{\varphi}\sum_{k=0}^{M-1}2^{k}d(\underline{x},\underline{z}^{(j)})<L_{\varphi}2^{M}\varepsilon_{i_{j}}.

Combining this inequality and εij0\varepsilon_{i_{j}}\to 0 as jj\to\infty, we have

SM(φαφ)(x¯)=0.S_{M}(\varphi-\alpha_{\varphi})(\underline{x})=0.

For a MM-periodic sequence x¯\underline{x}, we define w¯(k)\underline{w}^{(k)} by

w(k)=(x0xM1)ky¯.w^{(k)}=(x_{0}\cdots x_{M-1})^{k}\underline{y}.

Notice that for nk=kMn_{k}=kM,

Snk(φαφ)(x¯)=0S_{n_{k}}(\varphi-\alpha_{\varphi})(\underline{x})=0

and we get

Snk(φαφ)(w¯(k))\displaystyle S_{n_{k}}(\varphi-\alpha_{\varphi})(\underline{w}^{(k)}) =Snk(φαφ)(w¯(k))Snk(φαφ)(x¯)\displaystyle=S_{n_{k}}(\varphi-\alpha_{\varphi})(\underline{w}^{(k)})-S_{n_{k}}(\varphi-\alpha_{\varphi})(\underline{x})
Lφj=0nk1d(σj(w¯(k)),σj(x¯))\displaystyle\leq L_{\varphi}\sum_{j=0}^{n_{k}-1}d(\sigma^{j}(\underline{w}^{(k)}),\sigma^{j}(\underline{x}))
Lφd(y¯,x¯).\displaystyle\leq L_{\varphi}d(\underline{y},\underline{x}).

Since w¯(k)B(x¯,y¯,nk;2nk)\underline{w}^{(k)}\in B(\underline{x},\underline{y},n_{k};2^{-n_{k}}) for every k1k\geq 1, we have Hφ(x¯,y¯)Lφd(x¯,y¯)H_{\varphi}(\underline{x},\underline{y})\leq L_{\varphi}d(\underline{x},\underline{y}).

Next, we assume that {Ni}\{N_{i}\} is not bounded. Then we can take an increasing subsequence {Nij}\{N_{i_{j}}\} with max{2Nij,εij}<εj\max\{2^{-N_{i_{j}}},\varepsilon_{i_{j}}\}<\varepsilon_{j} and a sequence z¯(j)\underline{z}^{(j)} satisfying

(5) SNij(φαφ)(z¯(j))<ε2j\displaystyle S_{N_{i_{j}}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})<\frac{\varepsilon}{2^{j}}

and z¯(j)B(x¯,x¯,Nij;εij)\underline{z}^{(j)}\in B(\underline{x},\underline{x},N_{i_{j}};\varepsilon_{i_{j}}) for any jj\in\mathbb{N}. Set {Mj}j0\{M_{j}\}_{j\geq 0} by M0=0M_{0}=0 and Mj=NijM_{j}=N_{i_{j}}. Set w¯(n)\underline{w}^{(n)} by

w¯(n)=(z0(1)zM11(1))(z0(2)zM21(2))(z0(n)zMn1(n))y¯.\underline{w}^{(n)}=(z_{0}^{(1)}\cdots z^{(1)}_{M_{1}-1})(z_{0}^{(2)}\cdots z^{(2)}_{M_{2}-1})\cdots(z_{0}^{(n)}\cdots z^{(n)}_{M_{n}-1})\underline{y}.

Notice that for jnj\leq n

d(σMj1++M0(w¯(n)),z¯(j))2Nij<εjd(\sigma^{M_{j-1}+\cdots+M_{0}}(\underline{w}^{(n)}),\underline{z}^{(j)})\leq 2^{-N_{i_{j}}}<\varepsilon_{j}

since the first Mj(=Nij)M_{j}(=N_{i_{j}}) coordinates of σMj1+M0w¯(n)\sigma^{M_{j-1}+\cdots M_{0}}\underline{w}^{(n)} coincide with that of z¯(j)\underline{z}^{(j)}. Let mk=M1++Mkm_{k}=M_{1}+\cdots+M_{k}. For any kk,

Smk(φαφ)(w¯(k))=j=1k1SMj(φαφ)(σMj1++M0(w¯(k)))\displaystyle S_{m_{k}}(\varphi-\alpha_{\varphi})(\underline{w}^{(k)})=\sum_{j=1}^{k-1}S_{M_{j}}(\varphi-\alpha_{\varphi})(\sigma^{{M_{j-1}+\cdots+M_{0}}}(\underline{w}^{(k)}))
=j=1kSMj(φαφ)(z¯(j))\displaystyle=\sum_{j=1}^{k}S_{M_{j}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})
+j=1kSMj(φαφ)(σMj1++M0(w¯(k)))SMj(φαφ)(z¯(j)).\displaystyle\quad+\sum_{j=1}^{k}S_{M_{j}}(\varphi-\alpha_{\varphi})(\sigma^{M_{j-1}+\cdots+M_{0}}(\underline{w}^{(k)}))-S_{M_{j}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)}).

Here by (5)

j=1kSMj(φαφ)(z¯(j))j=1kε2i<ε\sum_{j=1}^{k}S_{M_{j}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})\leq\sum_{j=1}^{k}\frac{\varepsilon}{2^{i}}<\varepsilon

for any kk. If j<kj<k,

SMj(φαφ)(σMj1++M0(w¯(k)))SMj(φαφ)(z¯(j))\displaystyle S_{M_{j}}(\varphi-\alpha_{\varphi})(\sigma^{{M_{j-1}+\cdots+M_{0}}}(\underline{w}^{(k)}))-S_{M_{j}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})
=i=0Mj(φ(σi+Mj1++M0(w¯(k)))φ(σi(z¯(j))))\displaystyle=\sum_{i=0}^{M_{j}}\left(\varphi(\sigma^{i+{M_{j-1}+\cdots+M_{0}}}(\underline{w}^{(k)}))-\varphi(\sigma^{i}(\underline{z}^{(j)}))\right)
Lφi=0Mjd(σi+Mj1++M0(w¯(k)),σi(z¯(j)))\displaystyle\leq L_{\varphi}\sum_{i=0}^{M_{j}}d(\sigma^{i+M_{j-1}+\cdots+M_{0}}(\underline{w}^{(k)}),\sigma^{i}(\underline{z}^{(j)}))
Lφi=1Mj12id(σMj+Mj1++M0(w¯(k)),σMj(z¯(j)))\displaystyle\leq L_{\varphi}\sum_{i=1}^{M_{j}}\frac{1}{2^{i}}d(\sigma^{M_{j}+M_{j-1}+\cdots+M_{0}}(\underline{w}^{(k)}),\sigma^{M_{j}}(\underline{z}^{(j)}))
Lφd(σMj+Mj1++M0(w¯(k)),σMj(z¯(j))\displaystyle\leq L_{\varphi}d(\sigma^{M_{j}+M_{j-1}+\cdots+M_{0}}(\underline{w}^{(k)}),\sigma^{M_{j}}(\underline{z}^{(j)})
Lφ(d(σMj+Mj1++M0(w¯(k)),z¯(j+1))+d(z¯(j+1),x¯)+d(x¯,σMj(z¯(j))))\displaystyle\leq L_{\varphi}(d(\sigma^{M_{j}+M_{j-1}+\cdots+M_{0}}(\underline{w}^{(k)}),\underline{z}^{(j+1)})+d(\underline{z}^{(j+1)},\underline{x})+d(\underline{x},\sigma^{M_{j}}(\underline{z}^{(j)})))
<Lφ(εj+1+εij+1+εij)<2Lφε.\displaystyle<L_{\varphi}(\varepsilon_{j+1}+\varepsilon_{i_{j+1}}+\varepsilon_{i_{j}})<2L_{\varphi}\varepsilon.

If j=kj=k, Thus we get:

SMk(φαφ)(σMk1++M0(w¯(k)))SMk(φαφ)(z¯(k))\displaystyle S_{M_{k}}(\varphi-\alpha_{\varphi})(\sigma^{{M_{k-1}+\cdots+M_{0}}}(\underline{w}^{(k)}))-S_{M_{k}}(\varphi-\alpha_{\varphi})(\underline{z}^{(k)})
Lφi=1n12id(y¯,σMk(z¯(k)))\displaystyle\leq L_{\varphi}\sum_{i=1}^{n}\frac{1}{2^{i}}d(\underline{y},\sigma^{M_{k}}(\underline{z}^{(k)}))
Lφ(d(x¯,y¯)+d(x¯,σMk(z¯(k))))\displaystyle\leq L_{\varphi}(d(\underline{x},\underline{y})+d(\underline{x},\sigma^{M_{k}}(\underline{z}^{(k)})))
Lφd(x¯,y¯)+Lφεik\displaystyle\leq L_{\varphi}d(\underline{x},\underline{y})+L_{\varphi}\varepsilon_{i_{k}}
<Lφd(x¯,y¯)+Lφε.\displaystyle<L_{\varphi}d(\underline{x},\underline{y})+L_{\varphi}\varepsilon.

Hence we get

Smk(φαφ)(w¯(k))\displaystyle S_{m_{k}}(\varphi-\alpha_{\varphi})(\underline{w}^{(k)}) <ε+3Lφε+Lφd(x¯,y¯).\displaystyle<\varepsilon+3L_{\varphi}\varepsilon+L_{\varphi}d(\underline{x},\underline{y}).

Since w¯(k)B(x¯,y¯,mk;εi1)B(x¯,y¯,mk;ε)\underline{w}^{(k)}\in B(\underline{x},\underline{y},m_{k};\varepsilon_{i_{1}})\subset B(\underline{x},\underline{y},m_{k};\varepsilon) for every kk, we have

Hφ(x¯,y¯;ε)(1+3Lφ)ε+Lφd(x¯,y¯).\displaystyle H_{\varphi}(\underline{x},\underline{y};\varepsilon)\leq(1+3L_{\varphi})\varepsilon+L_{\varphi}d(\underline{x},\underline{y}).

Letting ε0\varepsilon\to 0, we have

Hφ(x¯,y¯)Lφd(x¯,y¯).H_{\varphi}(\underline{x},\underline{y})\leq L_{\varphi}d(\underline{x},\underline{y}).

Next, we prove the first half of Main Theorem 2.

Theorem 2.12 (Analogy of Theorem 4.1 in [BLL13], cf. Main Theorem 2).

For any x¯Ωφ\underline{x}\in\Omega_{\varphi}, the map Xy¯Hφ(x¯,y¯)X\ni\underline{y}\mapsto H_{\varphi}(\underline{x},\underline{y}) is a Lipschitz calibrated subaction.

Proof.

Note that Hφ(x¯,):X{+}H_{\varphi}(\underline{x},\cdot):X\to\mathbb{R}\cup\{+\infty\} does not take ++\infty by Theorem 2.11 and x¯Ωφ\underline{x}\in\Omega_{\varphi}. We first check the Lipschitz property. Fix ε>0\varepsilon>0. Take sequences {w¯(n)},{w¯(n)}X\{\underline{w}^{(n)}\},\{\underline{w}^{\prime(n)}\}\subset X and a monotone increasing sequence {Nn}\{N_{n}\} satisfying

limnSNn(φαφ)(w¯(n))\displaystyle\lim_{n\to\infty}S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{w}^{(n)}) =Hφ(x¯,y¯;ε),\displaystyle=H_{\varphi}(\underline{x},\underline{y};\varepsilon),
w¯(n)\displaystyle\underline{w}^{(n)} B(x¯,y¯,Nn;ε),\displaystyle\in B(\underline{x},\underline{y},N_{n};\varepsilon),

and

limnSNn(φαφ)(w¯(n))\displaystyle\lim_{n\to\infty}S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{w}^{\prime(n)}) =Hφ(x¯,y¯;2ε),\displaystyle=H_{\varphi}(\underline{x},\underline{y}^{\prime};2\varepsilon),
w¯(n)\displaystyle\underline{w}^{\prime(n)} B(x¯,y¯,Nn;2ε)\displaystyle\in B(\underline{x},\underline{y}^{\prime},N_{n};2\varepsilon)

respectively. Set

z¯(n)=w0(n)w1(n)wn1(n)y¯.\underline{z}^{(n)}=w_{0}^{(n)}w_{1}^{(n)}\cdots w_{n-1}^{(n)}\underline{y}^{\prime}.

Since z¯(n)B(x¯,y¯,Nn;2ε)\underline{z}^{(n)}\in B(\underline{x},\underline{y}^{\prime},N_{n};2\varepsilon), (by replacing a subsequence if necessary) we take NN such that if nNn\geq N,

SNn(φαφ)(z¯(n))SNn(φαφ)(w¯(n))ε.S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{z}^{({n})})\geq S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{w}^{\prime({n})})-\varepsilon.

Then we obtain

SNn(φαφ)(w¯(n))\displaystyle S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{w}^{(n)}) =SNn(φαφ)(z¯(n))+i=0Nn1(φ(σi(w¯(n)))φ(σi(z¯(n))))\displaystyle=S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{z}^{(n)})+\sum_{i=0}^{N_{n}-1}(\varphi(\sigma^{i}(\underline{w}^{(n)}))-\varphi(\sigma^{i}(\underline{z}^{(n)})))
SNn(φαφ)(z¯(n))Lφd(σNn(w¯(n)),y)\displaystyle\geq S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{z}^{(n)})-L_{\varphi}d(\sigma^{N_{n}}(\underline{w}^{(n)}),y^{\prime})
SNn(φαφ)(w¯(n))εLφ(d(y,y)+d(σNn(w¯(n)),y))\displaystyle\geq S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{w}^{\prime(n)})-\varepsilon-L_{\varphi}(d(y,y^{\prime})+d(\sigma^{N_{n}}(\underline{w}^{(n)}),y))
SNn(φαφ)(w¯(n))Lφd(y,y)(Lφ+1)ε.\displaystyle\geq S_{N_{n}}(\varphi-\alpha_{\varphi})(\underline{w}^{\prime(n)})-L_{\varphi}d(y,y^{\prime})-(L_{\varphi}+1)\varepsilon.

Letting nn\to\infty and ε0\varepsilon\to 0, we have

Hφ(x¯,y¯)Hφ(x¯,y¯)Lφ(y¯,y¯).\displaystyle H_{\varphi}(\underline{x},\underline{y^{\prime}})-H_{\varphi}(\underline{x},\underline{y})\leq L_{\varphi}(\underline{y},\underline{y^{\prime}}).

Since the opposite inequality can be obtained by swapping the roles of {w¯(n)}\{\underline{w}^{(n)}\} and {w¯(n)}\{\underline{w^{\prime}}^{(n)}\}, the map Xy¯Hφ(x¯,y¯)X\ni\underline{y}\mapsto H_{\varphi}(\underline{x},\underline{y}) is Lipschitz for any x¯Ωφ\underline{x}\in\Omega_{\varphi}.

Next, we check the property of calibrated subaction. Fix ε>0\varepsilon>0 and we take a sequence w¯(k)\underline{w}^{(k)} and a monotone increasing sequence {nk}\{n_{k}\} such that

w¯(k)B(x¯,y¯,nk;ε)\underline{w}^{(k)}\in B(\underline{x},\underline{y},n_{k};\varepsilon)

satisfying

Hφ(x¯,y¯,ε)=limnkSnk(φαφ)(w¯(k)).H_{\varphi}(\underline{x},\underline{y},\varepsilon)=\lim_{n_{k}\to\infty}S_{n_{k}}(\varphi-\alpha_{\varphi})(\underline{w}^{(k)}).

Here, we can assume that Hφ(x¯,y¯,ε)H_{\varphi}(\underline{x},\underline{y},\varepsilon) is finite since x¯Ωφ\underline{x}\in\Omega_{\varphi}. Notice that

w¯(k)B(x¯,σy¯,nk+1;2ε).\underline{w}^{(k)}\in B(\underline{x},\sigma\underline{y},n_{k}+1;2\varepsilon).

Taking sufficiently large NN, for any kNk\geq N,

Hφ(x¯,y¯,ε)+ε\displaystyle H_{\varphi}(\underline{x},\underline{y},\varepsilon)+\varepsilon
>Snk+1(φαφ)(w¯(k))(φ(σnkw¯(k))αφ)\displaystyle>S_{n_{k}+1}(\varphi-\alpha_{\varphi})(\underline{w}^{(k)})-(\varphi(\sigma^{n_{k}}\underline{w}^{(k)})-\alpha_{\varphi})
>Snk+1(φαφ)(w¯(k))(φ(y¯)αφ)Lφε\displaystyle>S_{n_{k}+1}(\varphi-\alpha_{\varphi})(\underline{w}^{(k)})-(\varphi(\underline{y})-\alpha_{\varphi})-L_{\varphi}\varepsilon
>Hφ(x¯,σy¯,2ε)ε(φ(y¯)αφ)Lφε\displaystyle>H_{\varphi}(\underline{x},\sigma\underline{y},2\varepsilon)-\varepsilon-(\varphi(\underline{y})-\alpha_{\varphi})-L_{\varphi}\varepsilon

Hence we get

Hφ(x¯,σ(y¯))Hφ(x¯,y¯)+φ(y¯)αφ.H_{\varphi}(\underline{x},\sigma(\underline{y}))\leq H_{\varphi}(\underline{x},\underline{y})+\varphi(\underline{y})-\alpha_{\varphi}.

for any y¯X\underline{y}\in X, and it implies

Hφ(x¯,y¯)miny¯^σ1{y¯}{Hφ(x¯,y¯^)+φ(y¯^)αφ}.\displaystyle H_{\varphi}(\underline{x},{\underline{y}})\leq\min_{\hat{\underline{y}}\in\sigma^{-1}\{{\underline{y}}\}}\{H_{\varphi}(\underline{x},\hat{\underline{y}})+\varphi(\hat{\underline{y}})-\alpha_{\varphi}\}.

To show the opposite inequality, let z[0,1]z\in[0,1] be a limit of a convergent subsequence {wkj1(kj)}\{w^{(k_{j})}_{k_{j}-1}\} of {wk1(k)}[0,1]\{w^{(k)}_{k-1}\}\subset[0,1]. Letting y¯=zy¯\underline{y^{\prime}}=z\underline{y}, for sufficiently large jj, we have

w¯(kj)B(x¯,y¯,kj1;2ε)\underline{w}^{(k_{j})}\in B(\underline{x},\underline{y}^{\prime},k_{j}-1;2\varepsilon)

and

Hφ(x¯,y¯;2ε)Skj1(φαφ)(w¯(kj))+ε.H_{\varphi}(\underline{x},\underline{y^{\prime}};2\varepsilon)\leq S_{k_{j}-1}(\varphi-\alpha_{\varphi})(\underline{w}^{(k_{j})})+\varepsilon.

Moreover, for large jj, we have

Hφ(x¯,y¯,ε)+ε\displaystyle H_{\varphi}(\underline{x},\underline{y},\varepsilon)+\varepsilon >Skj1(φαφ)(w¯(kj))+φ(y¯)αφLφε\displaystyle>S_{k_{j}-1}(\varphi-\alpha_{\varphi})(\underline{w}^{(k_{j})})+\varphi(\underline{y}^{\prime})-\alpha_{\varphi}-L_{\varphi}\varepsilon
>Hφ(x¯,y¯,2ε)ε+φ(y¯)αφLφε.\displaystyle>H_{\varphi}(\underline{x},\underline{y}^{\prime},2\varepsilon)-\varepsilon+\varphi(\underline{y}^{\prime})-\alpha_{\varphi}-L_{\varphi}\varepsilon.

Letting ε0\varepsilon\to 0 we have

Hφ(x¯,y¯)\displaystyle H_{\varphi}(\underline{x},\underline{y}) Hφ(x¯,y¯)+φ(y¯)αφ\displaystyle\geq H_{\varphi}(\underline{x},\underline{y^{\prime}})+\varphi(\underline{y^{\prime}})-\alpha_{\varphi} miny¯^σ1{y¯}{Hφ(x¯,y¯^)+φ(y¯^)αφ},\displaystyle\geq\min_{\hat{\underline{y}}\in\sigma^{-1}\{{\underline{y}}\}}\{H_{\varphi}(\underline{x},\hat{\underline{y}})+\varphi(\hat{\underline{y}})-\alpha_{\varphi}\},

which completes the proof. ∎

The next theorem is the second half of Main Theorem 2.

Theorem 2.13 (cf. Main Theorem 2).

For x¯\underline{x} and y¯Ωφ\underline{y}\in\Omega_{\varphi} define x¯y¯\underline{x}\sim\underline{y} by

Hφ(x¯,y¯)+Hφ(y¯,x¯)=0.\displaystyle H_{\varphi}(\underline{x},\underline{y})+H_{\varphi}(\underline{y},\underline{x})=0.

Then this is an euqivalence relation on Ω\Omega.

Before the proof of Theorem 2.13, we show the following lemma.

Lemma 2.14 (Analogy of Lemma 4.2 in [BLL13]).

For any x¯,y¯,z¯X\underline{x},\underline{y},\underline{z}\in X

(6) Hφ(x¯,y¯)Hφ(x¯,z¯)+Hφ(z¯,y¯).\displaystyle H_{\varphi}(\underline{x},\underline{y})\leq H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{y}).
Proof.

We remark that both sides of (6) may become ++\infty.

Fix θ>0\theta>0. Let ε>0\varepsilon>0 and N1N\geq 1 s.t. 2Lφεθ2L_{\varphi}\varepsilon\leq\theta,

Hφ(x¯,y¯)\displaystyle H_{\varphi}(\underline{x},\underline{y}) infnN{Sn(φαφ)(w¯):w¯B(x¯,y¯,n;2ε)+θ,\displaystyle\leq\inf_{n\geq N}\{S_{n}(\varphi-\alpha_{\varphi})(\underline{w}):\underline{w}\in B(\underline{x},\underline{y},n;2\varepsilon)+\theta,
Hφ(x¯,z¯)\displaystyle H_{\varphi}(\underline{x},\underline{z}) infnN{Sn(φαφ)(w¯):w¯B(x¯,z¯,n;ε)}θ\displaystyle\geq\inf_{n\geq N}\{S_{n}(\varphi-\alpha_{\varphi})(\underline{w}):\underline{w}\in B(\underline{x},\underline{z},n;\varepsilon)\}-\theta

and

Hφ(z¯,y¯)infnN{Sn(φαφ)(w¯):w¯B(z¯,y¯,n;ε)}θ.\displaystyle H_{\varphi}(\underline{z},\underline{y})\geq\inf_{n\geq N}\{S_{n}(\varphi-\alpha_{\varphi})(\underline{w}):\underline{w}\in B(\underline{z},\underline{y},n;\varepsilon)\}-\theta.

Then there exist n1Nn_{1}\geq N and w¯(1)B(x¯,z¯,n1;ε)\underline{w}^{(1)}\in B(\underline{x},\underline{z},n_{1};\varepsilon) s.t.

Hφ(x¯,z¯)Sn1(φαφ)(w¯(1))2θ\displaystyle H_{\varphi}(\underline{x},\underline{z})\geq S_{n_{1}}(\varphi-\alpha_{\varphi})(\underline{w}^{(1)})-2\theta

and there exist n2Nn_{2}\geq N and w¯(2)B(z¯,y¯,n2;ε)\underline{w}^{(2)}\in B(\underline{z},\underline{y},n_{2};\varepsilon) s.t.

Hφ(z¯,y¯)Sn2(φαφ)(w¯(2))2θ.\displaystyle H_{\varphi}(\underline{z},\underline{y})\geq S_{n_{2}}(\varphi-\alpha_{\varphi})(\underline{w}^{(2)})-2\theta.

Let

w¯=w0(1)wn11(1)w¯(2),i.e.w¯[w0(1)wn11(1)]σn1{w¯(2)}.\displaystyle\underline{w}=w^{(1)}_{0}\cdots w^{(1)}_{n_{1}-1}\underline{w}^{(2)},\quad\mbox{i.e.}\quad\underline{w}\in[w^{(1)}_{0}\cdots w^{(1)}_{n_{1}-1}]\cap\sigma^{-n_{1}}\{\underline{w}^{(2)}\}.

Then we have

Sn1+n2(φαφ)(w¯)\displaystyle S_{n_{1}+n_{2}}(\varphi-\alpha_{\varphi})(\underline{w})
Sn1(φαφ)(w¯(1))+Sn2(φαφ)(w¯(2))+Lφi=0n11d(σi(w¯),σiw¯(1))\displaystyle\leq S_{n_{1}}(\varphi-\alpha_{\varphi})(\underline{w}^{(1)})+S_{n_{2}}(\varphi-\alpha_{\varphi})(\underline{w}^{(2)})+L_{\varphi}\sum_{i=0}^{n_{1}-1}d(\sigma^{i}(\underline{w}),\sigma^{i}\underline{w}^{(1)})
Hφ(x¯,z¯)+Hφ(z¯,y¯)+Lφi=1n12id(σn1w¯,σn1w¯(1))+4θ\displaystyle\leq H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{y})+L_{\varphi}\sum_{i=1}^{n_{1}}2^{-i}d(\sigma^{n_{1}}\underline{w},\sigma^{n_{1}}\underline{w}^{(1)})+4\theta
Hφ(x¯,z¯)+Hφ(z¯,y¯)+Lφ(d(w¯(2),z¯)+d(z¯,σn1w¯(1)))+4θ\displaystyle\leq H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{y})+L_{\varphi}\left(d(\underline{w}^{(2)},\underline{z})+d(\underline{z},\sigma^{n_{1}}\underline{w}^{(1)})\right)+4\theta
Hφ(x¯,z¯)+Hφ(z¯,y¯)+2Lφε+4θ\displaystyle\leq H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{y})+2L_{\varphi}\varepsilon+4\theta
Hφ(x¯,z¯)+Hφ(z¯,y¯)+5θ.\displaystyle\leq H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{y})+5\theta.

Moreover d(σn1+n2w¯,y¯)=d(σn2w¯(2),y¯)<εd(\sigma^{n_{1}+n_{2}}\underline{w},\underline{y})=d(\sigma^{n_{2}}\underline{w}^{(2)},\underline{y})<\varepsilon and

d(w¯,x¯)\displaystyle d(\underline{w},\underline{x}) d(w¯,w¯(1))+d(w¯(1),x¯)\displaystyle\leq d(\underline{w},\underline{w}^{(1)})+d(\underline{w}^{(1)},\underline{x})
2n1d(σn1w¯,σn1w¯(1))+ε\displaystyle\leq 2^{-n_{1}}d(\sigma^{n_{1}}\underline{w},\sigma^{n_{1}}\underline{w}^{(1)})+\varepsilon
2n1(d(w¯(2),z¯)+d(z¯,σn1w¯(1)))+ε\displaystyle\leq 2^{-n_{1}}\left(d(\underline{w}^{(2)},\underline{z})+d(\underline{z},\sigma^{n_{1}}\underline{w}^{(1)})\right)+\varepsilon
(2n1+1+1)ε2ε,\displaystyle\leq\left(2^{-n_{1}+1}+1\right)\varepsilon\leq 2\varepsilon,

which implies

Hφ(x¯,y¯)\displaystyle H_{\varphi}(\underline{x},\underline{y}) Sn1+n2(φαφ)(w¯)+θ\displaystyle\leq S_{n_{1}+n_{2}}(\varphi-\alpha_{\varphi})(\underline{w})+\theta
Hφ(x¯,z¯)+Hφ(z¯,y¯)+6θ.\displaystyle\leq H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{y})+6\theta.

Since θ>0\theta>0 is arbitrary, we complete the proof. ∎

Proof of Theorem 2.13.

It suffices to show the transitive relation. Take x¯,y¯\underline{x},\underline{y} and z¯Ω\underline{z}\in\Omega with x¯y¯\underline{x}\sim\underline{y} and y¯z¯\underline{y}\sim\underline{z}. By (6) we have

Hφ(x¯,z¯)+Hφ(z¯,x¯)\displaystyle H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{x}) Hφ(x¯,y¯)+Hφ(y¯,z¯)+Hφ(z¯,y¯)+Hφ(y¯,x¯)=0.\displaystyle\leq H_{\varphi}(\underline{x},\underline{y})+H_{\varphi}(\underline{y},\underline{z})+H_{\varphi}(\underline{z},\underline{y})+H_{\varphi}(\underline{y},\underline{x})=0.

It also yields

Hφ(x¯,z¯)+Hφ(z¯,x¯)Hφ(x¯,x¯)=0.\displaystyle H_{\varphi}(\underline{x},\underline{z})+H_{\varphi}(\underline{z},\underline{x})\geq H_{\varphi}(\underline{x},\underline{x})=0.

Proof of Main Theorem 2.

It follows immediately from Theorem 2.12 and 2.13. ∎

At the end of this section, we present the invariance of the above equivalent classes.

Proposition 2.15.

For any x¯Ωφ\underline{x}\in\Omega_{\varphi} we have

Hφ(x¯,σ(x¯))+Hφ(σ(x¯),x¯)=0.\displaystyle H_{\varphi}(\underline{x},\sigma(\underline{x}))+H_{\varphi}(\sigma(\underline{x}),\underline{x})=0.

In particular a equivalence class [x¯][\underline{x}] of the relation satisfies σ[x¯][x¯]\sigma[\underline{x}]\subset[\underline{x}].

Proof.

By Theorem 2.11 and Lemma 2.14 we have

Hφ(x¯,σ(x¯))+Hφ(σ(x¯),x¯)Hφ(x¯,x¯)=0.\displaystyle H_{\varphi}(\underline{x},\sigma(\underline{x}))+H_{\varphi}(\sigma(\underline{x}),\underline{x})\geq H_{\varphi}(\underline{x},\underline{x})=0.

Thus we will show that it is non-negative. Since Hφ(x¯,)H_{\varphi}(\underline{x},\cdot) is a subaction, we have

Hφ(x¯,σ(x¯))\displaystyle H_{\varphi}(\underline{x},\sigma(\underline{x})) =minσ(y¯)=σ(x¯){Hφ(x¯,y¯)+φ(y¯)αφ}\displaystyle=\min_{\sigma(\underline{y})=\sigma(\underline{x})}\{H_{\varphi}(\underline{x},\underline{y})+\varphi(\underline{y})-\alpha_{\varphi}\}
(7) Hφ(x¯,x¯)+φ(x¯)αφ=φ(x¯)αφ.\displaystyle\leq H_{\varphi}(\underline{x},\underline{x})+\varphi(\underline{x})-\alpha_{\varphi}=\varphi(\underline{x})-\alpha_{\varphi}.

Let θ>0\theta>0. Take 0<ε<θ/Lφ\displaystyle 0<\varepsilon<\theta/L_{\varphi} and N1N\geq 1 s.t.

Hφ(σ(x¯),x¯)infnN{Sn(φαφ)(z¯):z¯B(σ(x¯),x¯,n;2ε)}+θ\displaystyle H_{\varphi}(\sigma(\underline{x}),\underline{x})\leq\inf_{n\geq N}\{S_{n}(\varphi-\alpha_{\varphi})(\underline{z}):\underline{z}\in B(\sigma(\underline{x}),\underline{x},n;2\varepsilon)\}+\theta

and

Hφ(x¯,x¯)infnN+1{Sn(φαφ)(z¯):z¯B(x¯,x¯,n;ε)}θ.\displaystyle H_{\varphi}(\underline{x},\underline{x})\geq\inf_{n\geq N+1}\{S_{n}(\varphi-\alpha_{\varphi})(\underline{z}):\underline{z}\in B(\underline{x},\underline{x},n;\varepsilon)\}-\theta.

Then there exist nN+1n\geq N+1 and z¯B(x¯,x¯,n;ε)\underline{z}\in B(\underline{x},\underline{x},n;\varepsilon) s.t.

Hφ(x¯,x¯)Sn(φαφ)(z¯)2θ.\displaystyle H_{\varphi}(\underline{x},\underline{x})\geq S_{n}(\varphi-\alpha_{\varphi})(\underline{z})-2\theta.

Then we have d(σ(z¯),σ(x¯))=2(d(z¯,x¯)|z0x0|2)2ε\displaystyle d(\sigma(\underline{z}),\sigma(\underline{x}))=2\left(d(\underline{z},\underline{x})-\frac{|z_{0}-x_{0}|}{2}\right)\leq 2\varepsilon and

Sn1(φαφ)(σ(z¯))\displaystyle S_{n-1}(\varphi-\alpha_{\varphi})(\sigma(\underline{z})) =Sn(φαφ)(z¯)φ(z¯)+αφ\displaystyle=S_{n}(\varphi-\alpha_{\varphi})(\underline{z})-\varphi(\underline{z})+\alpha_{\varphi}
Hφ(x¯,x¯)φ(z¯)+αφ+2θ\displaystyle\leq H_{\varphi}(\underline{x},\underline{x})-\varphi(\underline{z})+\alpha_{\varphi}+2\theta
=φ(x¯)+αφφ(z¯)+φ(x¯)+2θ\displaystyle=-\varphi(\underline{x})+\alpha_{\varphi}-\varphi(\underline{z})+\varphi(\underline{x})+2\theta
φ(x¯)+αφ+Lφd(x¯,z¯)+2θ\displaystyle\leq-\varphi(\underline{x})+\alpha_{\varphi}+L_{\varphi}d(\underline{x},\underline{z})+2\theta
φ(x¯)+αφ+Lφε+2θ\displaystyle\leq-\varphi(\underline{x})+\alpha_{\varphi}+L_{\varphi}\varepsilon+2\theta
φ(x¯)+αφ+3θ.\displaystyle\leq-\varphi(\underline{x})+\alpha_{\varphi}+3\theta.

Combining (7), we have

Hφ(σ(x¯),x¯)+Hφ(x¯,σ(x¯))+4θ,\displaystyle H_{\varphi}(\sigma(\underline{x}),\underline{x})+H_{\varphi}(\underline{x},\sigma(\underline{x}))\leq+4\theta,

which complete the proof. ∎

3. Another characterization of the Aubry set

In this section, we describe another characterization of the Aubry set. We assume that the potential φ:X\varphi:X\to\mathbb{R} is Lipschitz. A positive-semi orbit {σn(x)}n0\{\sigma^{n}(x)\}_{n\in\mathbb{N}_{0}} is said to be

  • φ\varphi-semi-static: if for any non-negative integers i<ji<j

    n=ij1(φσn(x¯)αφ)=Sφ(σi(x¯),σj(x¯)).\sum_{n=i}^{j-1}\Big{(}\varphi\circ\sigma^{n}(\underline{x})-\alpha_{\varphi}\Big{)}=S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x})).
  • φ\varphi-static: if for any non-negative integers i<ji<j

    n=ij1(φσn(x¯)αφ)=Sφ(σj(x¯),σi(x¯)).\sum_{n=i}^{j-1}\Big{(}\varphi\circ\sigma^{n}(\underline{x})-\alpha_{\varphi}\Big{)}=-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x})).

We call the sets

Nφ\displaystyle N_{\varphi} :={σk(x¯)X{σn(x¯)}n0isφ-semi-static,k0},\displaystyle:=\{\sigma^{k}(\underline{x})\in X\mid\{\sigma^{n}(\underline{x})\}_{n\in\mathbb{N}_{0}}\ \text{is}\ \varphi\text{-semi-static},k\in\mathbb{N}_{0}\},
Aφ\displaystyle A_{\varphi} :={σk(x¯)X{σn(x¯)}n0isφ-static,k0}\displaystyle:=\{\sigma^{k}(\underline{x})\in X\mid\{\sigma^{n}(\underline{x})\}_{n\in\mathbb{N}_{0}}\ \text{is}\ \varphi\text{-static},k\in\mathbb{N}_{0}\}

as the φ\varphi-semi-static set and the φ\varphi-static set respectively. It is trivial that NφN_{\varphi} and AφA_{\varphi} are σ\sigma-invariant since they are sets of positive semi-orbits. We will see that AφNφA_{\varphi}\subset N_{\varphi} (Proposition 3.2), and that AφA_{\varphi} (and hence NφN_{\varphi}) is not empty (Theorem 3.5). Moreover, by Proposition 2.7, we deduce that the φ\varphi-static set AφA_{\varphi} is closed and hence compact. We begin with the following.

Lemma 3.1.

For any x¯,y¯,z¯X\underline{x},\underline{y},\underline{z}\in X, we have

Sφ(x¯,y¯)Sφ(x¯,z¯)+Sφ(z¯,y¯).\displaystyle S_{\varphi}(\underline{x},\underline{y})\leq S_{\varphi}(\underline{x},\underline{z})+S_{\varphi}(\underline{z},\underline{y}).

In particular,

Sφ(x¯,y¯)+Sφ(y¯,x¯)0\displaystyle S_{\varphi}(\underline{x},\underline{y})+S_{\varphi}(\underline{y},\underline{x})\geq 0

holds for all x¯,y¯X\underline{x},\underline{y}\in X.

Proof.

We can prove the claim as in the proof of Lemma 2.14. Note that we do not assume that NN is sufficiently large in the discussion. ∎

Proposition 3.2.

For each φC(X)\varphi\in C(X), it holds that AφNφA_{\varphi}\subset N_{\varphi}.

Proof.

Fix x¯X\underline{x}\in X. Since σi(x¯)B(σi(x¯),σj(x¯),ji;ε)\sigma^{i}(\underline{x})\in B(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x}),j-i;\varepsilon) holds for any ε>0\varepsilon>0, it is trivial that

n=ij1(φσn(x¯)αφ)Sφ(σi(x¯),σj(x¯)).\sum_{n=i}^{j-1}\Big{(}\varphi\circ\sigma^{n}(\underline{x})-\alpha_{\varphi}\Big{)}\geq S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x})).

By Lemma 3.1, we have Sφ(σi(x¯),σj(x¯))Sφ(σj(x¯),σi(x¯))S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x}))\geq-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x})). Hence, for any x¯X\underline{x}\in X, we have

n=ij1(φσn(x¯)αφ)Sφ(σi(x¯),σj(x¯))Sφ(σj(x¯),σi(x¯)).\sum_{n=i}^{j-1}\Big{(}\varphi\circ\sigma^{n}(\underline{x})-\alpha_{\varphi}\Big{)}\geq S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x}))\geq-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x})).

Therefore x¯Aφ\underline{x}\in A_{\varphi} implies x¯Nφ\underline{x}\in N_{\varphi}. ∎

Remark 3.3.

In the Aubry-Mather theory for Euler-Lagrange flows, the corresponding object of AφA_{\varphi} (resp. NφN_{\varphi}) is called as the Aubry set (resp. the Mañé set), which looks different from our terminology “Aubry set” (Definition 2.3). In Theorem 3.5 below, we see that these notions are equivalent in our setting.

Before we state our main result in this section, we give the following lemma.

Lemma 3.4.

If x¯X\underline{x}\in X satisfies Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0, then

Sφ(σ(x¯),x¯)φ(x¯)+αφ.\displaystyle S_{\varphi}(\sigma(\underline{x}),\underline{x})\leq-\varphi(\underline{x})+\alpha_{\varphi}.
Proof.

When x¯\underline{x} is a fixed point of σ\sigma, the Dirac measure at x¯\underline{x} must be the optimal measure for φ\varphi and thus we have φ(x¯)=αφ\varphi(\underline{x})=\alpha_{\varphi}, which implies

Sφ(σ(x¯),x¯)=Sφ(x¯,x¯)=0=φ(x¯)+αφ.S_{\varphi}(\sigma(\underline{x}),\underline{x})=S_{\varphi}(\underline{x},\underline{x})=0=-\varphi(\underline{x})+\alpha_{\varphi}.

Now we consider the case that x¯\underline{x} satisfies σ(x¯)x¯\sigma(\underline{x})\neq\underline{x}. Fix ε>0\varepsilon>0. Take n(j)n^{(j)}\in\mathbb{N} and z¯(j)B(x¯,x¯,n(j);ε)\underline{z}^{(j)}\in B(\underline{x},\underline{x},n^{(j)};\varepsilon) s.t.

limj+Sn(j)(φαφ)(z¯(j))=Sφ(x¯,x¯;ε)Sφ(x¯,x¯)=0.\lim_{j\to+\infty}S_{n^{(j)}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})=S_{\varphi}(\underline{x},\underline{x};\varepsilon)\leq S_{\varphi}(\underline{x},\underline{x})=0.

Since z¯(j)B(x¯,x¯,n(j);ε)\underline{z}^{(j)}\in B(\underline{x},\underline{x},n^{(j)};\varepsilon), we have

d(x¯,z¯(j))<ε,d(σnj(z¯(j)),x)<ε.d(\underline{x},\underline{z}^{(j)})<\varepsilon,\quad d(\sigma^{n_{j}}(\underline{z}^{(j)}),x)<\varepsilon.

By the property of the shift map, the inequality

d(σ(x¯),σ(z¯(j)))<2εd(\sigma(\underline{x}),\sigma(\underline{z}^{(j)}))<2\varepsilon

holds and thus we obtain

σ(z¯(j))B(σ(x¯),x¯,n(j)1;2ε).\sigma(\underline{z}^{(j)})\in B(\sigma(\underline{x}),\underline{x},n^{(j)}-1;2\varepsilon).

Note that n(j)n^{(j)} is greater than 1 since σ(x¯)x¯\sigma(\underline{x})\neq\underline{x}. Moreover, the Lipschitz continuity of φ\varphi implies that

|φ(x¯)φ(z¯(j))|Lφd(x¯,z¯(j)).|\varphi(\underline{x})-\varphi(\underline{z}^{(j)})|\leq L_{\varphi}d(\underline{x},\underline{z}^{(j)}).

We compute

Sφ(σ(x¯),x¯;2ε)\displaystyle S_{\varphi}(\sigma(\underline{x}),\underline{x};2\varepsilon) limj+Sn(j)1(φαφ)(σ(z¯(j)))\displaystyle\leq\lim_{j\to+\infty}S_{n^{(j)}-1}(\varphi-\alpha_{\varphi})(\sigma(\underline{z}^{(j)}))
=limj+(Sn(j)(φαφ)(z¯(j))φ(z¯(j))+αφ)\displaystyle=\lim_{j\to+\infty}\Big{(}S_{n^{(j)}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})-\varphi(\underline{z}^{(j)})+\alpha_{\varphi}\Big{)}
limj+(Sn(j)(φαφ)(z¯(j))φ(x)+Lφd(x¯,z¯(j))+αφ)\displaystyle\leq\lim_{j\to+\infty}\Big{(}S_{n^{(j)}}(\varphi-\alpha_{\varphi})(\underline{z}^{(j)})-\varphi(x)+L_{\varphi}d(\underline{x},\underline{z}^{(j)})+\alpha_{\varphi}\Big{)}
φ(x¯)+Lφε+αφ,\displaystyle\leq-\varphi(\underline{x})+L_{\varphi}\varepsilon+\alpha_{\varphi},

which yields

Sφ(σ(x¯),x¯)φ(x¯)+αφ.S_{\varphi}(\sigma(\underline{x}),\underline{x})\leq-\varphi(\underline{x})+\alpha_{\varphi}.

Theorem 3.5 (cf. Main Theorem 1).

For each Lipschitz function φ:X\varphi:X\to\mathbb{R}, we have

Aφ=Ωφ.\displaystyle A_{\varphi}=\Omega_{\varphi}.
Proof.

We first prove that x¯Aφ\underline{x}\in A_{\varphi} implies Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0. For each x¯Aφ\underline{x}\in A_{\varphi}, we have

Sn(φαφ)(x¯)+Sφ(σn(x¯),x¯)=0\displaystyle S_{n}(\varphi-\alpha_{\varphi})(\underline{x})+S_{\varphi}(\sigma^{n}(\underline{x}),\underline{x})=0

for nn\in\mathbb{N} by the definition of φ\varphi-static set. Moreover, by the definition of the Mañé potential, it holds that

Sn(φαφ)(x¯)Sφ(x¯,σn(x¯)).S_{n}(\varphi-\alpha_{\varphi})(\underline{x})\geq S_{\varphi}(\underline{x},\sigma^{n}(\underline{x})).

Therefore, using the triangle inequality for the Mañé potential (Lemma 3.1), we obtain

0\displaystyle 0 =Sn(φαφ)(x¯)+Sφ(σn(x¯),x¯)\displaystyle=S_{n}(\varphi-\alpha_{\varphi})(\underline{x})+S_{\varphi}(\sigma^{n}(\underline{x}),\underline{x})
Sφ(x¯,σn(x¯))+Sφ(σn(x¯),x¯)\displaystyle\geq S_{\varphi}(\underline{x},\sigma^{n}(\underline{x}))+S_{\varphi}(\sigma^{n}(\underline{x}),\underline{x})
Sφ(x¯,x¯)0.\displaystyle\geq S_{\varphi}(\underline{x},\underline{x})\geq 0.

Note that in the last inequality we use the fact that Sφ(y¯,y¯)0S_{\varphi}(\underline{y},\underline{y})\geq 0 for any y¯X\underline{y}\in X. Thus we have Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0.

Next we see that Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0 implies xAφx\in A_{\varphi}. Assume that Sφ(x¯,x¯)=0S_{\varphi}(\underline{x},\underline{x})=0. Since each xΩφx\in\Omega_{\varphi} satisfies

Hφ(x¯,σ(x¯))+Hφ(σ(x¯),x¯)=0,H_{\varphi}(\underline{x},\sigma(\underline{x}))+H_{\varphi}(\sigma(\underline{x}),\underline{x})=0,

we obtain

0Hφ(σ(x¯),σ(x¯))Hφ(x¯,σ(x¯))+Hφ(σ(x¯),x¯)=0\displaystyle 0\leq H_{\varphi}(\sigma(\underline{x}),\sigma(\underline{x}))\leq H_{\varphi}(\underline{x},\sigma(\underline{x}))+H_{\varphi}(\sigma(\underline{x}),\underline{x})=0

By the equivalence between Hφ(σ(x¯),σ(x¯))=0H_{\varphi}(\sigma(\underline{x}),\sigma(\underline{x}))=0 and Sφ(σ(x¯),σ(x¯))=0S_{\varphi}(\sigma(\underline{x}),\sigma(\underline{x}))=0, we conclude that Sφ(σ(x¯),σ(x¯))=0S_{\varphi}(\sigma(\underline{x}),\sigma(\underline{x}))=0. Repeating the same discussion, we obtain Sφ(σk(x¯),σk(x¯))=0S_{\varphi}(\sigma^{k}(\underline{x}),\sigma^{k}(\underline{x}))=0 for k0k\in\mathbb{N}_{0}.

By Lemma 3.4, we have

Sφ(σ(x¯),x¯))φ(x¯)+αφ.S_{\varphi}(\sigma(\underline{x}),\underline{x}))\leq-\varphi(\underline{x})+\alpha_{\varphi}.

Trivially it holds that

Sφ(x¯,σ(x¯))φ(x¯)αφ,S_{\varphi}(\underline{x},\sigma(\underline{x}))\leq\varphi(\underline{x})-\alpha_{\varphi},

since x¯B(x¯,σ(x¯),n=1;ε)\underline{x}\in B(\underline{x},\sigma(\underline{x}),n=1;\varepsilon) holds for any ε>0\varepsilon>0 and thus we obtain

Sφ(x¯,σ(x¯))+Sφ(σ(x¯),x¯)0.S_{\varphi}(\underline{x},\sigma(\underline{x}))+S_{\varphi}(\sigma(\underline{x}),\underline{x})\leq 0.

Combining Lemma 3.1, we see that

Sφ(x¯,σ(x¯))+Sφ(σ(x¯),x¯)=0S_{\varphi}(\underline{x},\sigma(\underline{x}))+S_{\varphi}(\sigma(\underline{x}),\underline{x})=0

and deduce that

Sφ(x¯,σ(x¯))=φ(x¯)αφ,Sφ(σ(x¯),x¯)=φ(x¯)+αφ.S_{\varphi}(\underline{x},\sigma(\underline{x}))=\varphi(\underline{x})-\alpha_{\varphi},\quad S_{\varphi}(\sigma(\underline{x}),\underline{x})=-\varphi(\underline{x})+\alpha_{\varphi}.

Similarly, from the identities Sφ(σk(x¯),σk(x¯))=0S_{\varphi}(\sigma^{k}(\underline{x}),\sigma^{k}(\underline{x}))=0 for k0k\in\mathbb{N}_{0}, we have

Sφ(σk(x¯),σk+1(x¯))\displaystyle S_{\varphi}(\sigma^{k}(\underline{x}),\sigma^{k+1}(\underline{x})) =φ(σk(x¯))αφ,\displaystyle=\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi},
Sφ(σk+1(x¯),σk(x¯))\displaystyle S_{\varphi}(\sigma^{k+1}(\underline{x}),\sigma^{k}(\underline{x})) =φ(σk(x¯))+αφ\displaystyle=-\varphi(\sigma^{k}(\underline{x}))+\alpha_{\varphi}

for k0k\in\mathbb{N}_{0}. Take arbitrary non-negative integers i,ji,j with i<ji<j. From the triangle inequality for the Mañé potential (Lemma 3.1),

Sφ(σj(x¯),σi(x¯))k=ij1Sφ(σk+1(x¯),σk(x¯))=k=ij1(φ(σk(x¯))αφ)S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x}))\leq\sum_{k=i}^{j-1}S_{\varphi}(\sigma^{k+1}(\underline{x}),\sigma^{k}(\underline{x}))=-\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}

holds and thus we obtain

k=ij1(φ(σk(x¯))αφ)Sφ(σj(x¯),σi(x¯))Sφ(σi(x¯),σj(x¯)).\displaystyle\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}\leq-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x}))\leq S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x})).

Since

Sφ(σi(x¯),σj(x¯))k=ij1(φ(σk(x¯))αφ)S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x}))\leq\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}

trivially holds, we deduce that

k=ij1(φ(σk(x¯))αφ)=Sφ(σj(x¯),σi(x¯))=Sφ(σi(x¯),σj(x¯)),\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}=-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x}))=S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x})),

which yields that x¯Aφ\underline{x}\in A_{\varphi}. ∎

Proof of Main Theorem 1.

It follows immediately from Theorem 2.11 and 3.5. ∎

Lastly, we show the relationship between Lipschitz subactions and AφA_{\varphi}.

Proposition 3.6.

Let uu be a Lipschitz subaction of a Lipschitz function φ:X\varphi:X\to\mathbb{R}. If x¯Aφ\underline{x}\in A_{\varphi}, then

(8) u(σk+1(x¯))u(σk(x¯))=φ(σk(x¯))αφ\displaystyle u(\sigma^{k+1}(\underline{x}))-u(\sigma^{k}(\underline{x}))=\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}

for all k0k\in\mathbb{N}_{0}. Conversely, if (8) holds for each k0k\in\mathbb{N}_{0}, then x¯Nφ\underline{x}\in N_{\varphi}.

Proof.

Assume that x¯Aφ\underline{x}\in A_{\varphi}. Since x¯AφNφ\underline{x}\in A_{\varphi}\subset N_{\varphi}, we have

k=ij1(φ(σk(x¯))αφ)=Sφ(σj(x¯),σi(x¯))=Sφ(σi(x¯),σj(x¯))\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}=-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x}))=S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x}))

for all non-negative integers i<ji<j. By Lemma 2.6, we obtain

u(σj(x¯))u(σi(x¯))Sφ(σi(x¯),σj(x¯))u(\sigma^{j}(\underline{x}))-u(\sigma^{i}(\underline{x}))\leq S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x}))

and

Sφ(σj(x¯),σi(x¯))u(σj(x¯))u(σi(x¯)).-S_{\varphi}(\sigma^{j}(\underline{x}),\sigma^{i}(\underline{x}))\leq u(\sigma^{j}(\underline{x}))-u(\sigma^{i}(\underline{x})).

Therefore, it holds that

k=ij1(φ(σk(x¯))αφ)=u(σj(x¯))u(σi(x¯)).\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}=u(\sigma^{j}(\underline{x}))-u(\sigma^{i}(\underline{x})).

We rewrite

u(σj(x¯))u(σi(x¯))=k=ij1(u(σk+1(x¯))u(σk(x¯)))u(\sigma^{j}(\underline{x}))-u(\sigma^{i}(\underline{x}))=\sum_{k=i}^{j-1}\big{(}u(\sigma^{k+1}(\underline{x}))-u(\sigma^{k}(\underline{x}))\big{)}

and compute

(9) k=ij1{(φ(σk(x¯))αφ)(u(σk+1(x¯))u(σk(x¯)))}=0.\displaystyle\sum_{k=i}^{j-1}\Big{\{}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}-\big{(}u(\sigma^{k+1}(\underline{x}))-u(\sigma^{k}(\underline{x}))\big{)}\Big{\}}=0.

Since uu is a calibrated subaction of φ\varphi, we have

(φ(σk(x¯))αφ)(u(σk+1(x¯))u(σk(x¯)))0,k0\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}-\big{(}u(\sigma^{k+1}(\underline{x}))-u(\sigma^{k}(\underline{x}))\big{)}\geq 0,\quad k\in\mathbb{N}_{0}

and thus we deduce that

(φ(σk(x¯))αφ)(u(σk+1(x¯))u(σk(x¯)))=0,k0\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}-\big{(}u(\sigma^{k+1}(\underline{x}))-u(\sigma^{k}(\underline{x}))\big{)}=0,\quad k\in\mathbb{N}_{0}

from (9).

Conversely, assume that (8) holds for each k0k\in\mathbb{N}_{0}. Then we have

k=ij1(φ(σk(x¯))αφ)\displaystyle\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)} =k=ij1(u(σk+1(x¯))u(σk(x¯)))\displaystyle=\sum_{k=i}^{j-1}\big{(}u(\sigma^{k+1}(\underline{x}))-u(\sigma^{k}(\underline{x}))\big{)}
=u(σj(x¯))u(σi(x¯))Sφ(σi(x¯),σj(x¯))\displaystyle=u(\sigma^{j}(\underline{x}))-u(\sigma^{i}(\underline{x}))\leq S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x}))

by Lemma 2.6. Since k=ij1(φ(σk(x¯))αφ)Sφ(σi(x¯),σj(x¯))\sum_{k=i}^{j-1}\big{(}\varphi(\sigma^{k}(\underline{x}))-\alpha_{\varphi}\big{)}\geq S_{\varphi}(\sigma^{i}(\underline{x}),\sigma^{j}(\underline{x})) is trivial, we see that x¯Nφ\underline{x}\in N_{\varphi}. ∎

4. Variational method applied to the Aubry set

In this section, we consider the case that the potential function φ:X\varphi:X\to\mathbb{R} is 2-locally constant, i.e.,

φ(x¯)=h(x0,x1),x¯=x0x1x2,\varphi(\underline{x})=h(x_{0},x_{1}),\quad\underline{x}=x_{0}x_{1}x_{2}\ldots,

for some Lipschitz continuous function h:[0,1]2h:[0,1]^{2}\to\mathbb{R} under suitable assumptions (see below for the precise assumptions for hh). In this case, we can obtain much more explicit information about the elements in the Mather set and the Aubry set for φ\varphi by using variational techniques developed in [Ban88, Yu22].

Firstly, the following proposition is easily shown:

Proposition 4.1.

Let φ:X\varphi:X\to\mathbb{R} be a function depending on only the first two coordinates, i.e., for all x¯=x0x1x2\underline{x}=x_{0}x_{1}x_{2}\ldots, φ(x¯)=φ(x0,x1)\varphi(\underline{x})=\varphi(x_{0},x_{1}). Then φ\varphi is Lipschitz continuous as a function on XX with respect to the metric dd on XX if and only if φ\varphi is Lipschitz continuous as a function on [0,1]2[0,1]^{2} with respect to the Euclidian metric d2d_{\mathbb{R}^{2}} on [0,1]2[0,1]^{2}.

Proof.

Suppose that there exists Lφ>0L_{\varphi}>0 such that |φ(x¯)φ(y¯)|Lφd(x¯,y¯)|\varphi(\underline{x})-\varphi(\underline{y})|\leq L_{\varphi}d(\underline{x},\underline{y}) for all x¯,y¯X\underline{x},\underline{y}\in X. Taking arbitrary x0,x1,y0,y1[0,1]x_{0},x_{1},y_{0},y_{1}\in[0,1], we have

|φ(x0,x1)φ(y0,y1)|\displaystyle|\varphi(x_{0},x_{1})-\varphi(y_{0},y_{1})| =|φ(x0x10)φ(y0y10)|\displaystyle=|\varphi(x_{0}x_{1}0^{\infty})-\varphi(y_{0}y_{1}0^{\infty})|
Lφ(|x0y0|+|x1y1|/2)\displaystyle\leq L_{\varphi}(|x_{0}-y_{0}|+|x_{1}-y_{1}|/2)
Lφ2d2((x0,x1),(y0,y1)).\displaystyle L_{\varphi}\sqrt{2}d_{\mathbb{R}^{2}}((x_{0},x_{1}),(y_{0},y_{1})).

Conversely, suppose that there exists Lφ>0L_{\varphi}>0 such that |φ(x0,x1)φ(y0,y1)|Lφd2((x0,x1),(y0,y1))|\varphi(x_{0},x_{1})-\varphi(y_{0},y_{1})|\leq L_{\varphi}d_{\mathbb{R}^{2}}((x_{0},x_{1}),(y_{0},y_{1})) for all (x0,x1),(y0,y1))[0,1]2(x_{0},x_{1}),(y_{0},y_{1}))\in[0,1]^{2}. Then

|φ(x¯)φ(y¯)|\displaystyle|\varphi(\underline{x})-\varphi(\underline{y})| =|φ(x0,x1)φ(y0,y1)|\displaystyle=|\varphi(x_{0},x_{1})-\varphi(y_{0},y_{1})|
Lφd2((x0,x1),(y0,y1))\displaystyle\leq L_{\varphi}d_{\mathbb{R}^{2}}((x_{0},x_{1}),(y_{0},y_{1}))
2Lφ(|x0y0|+|x1y1|/2)2Lφd(x¯,y¯),\displaystyle\leq 2L_{\varphi}(|x_{0}-y_{0}|+|x_{1}-y_{1}|/2)\leq 2L_{\varphi}d(\underline{x},\underline{y}),

which complete the proof. ∎

Remark 4.2.

For a subshift of finite type with a finite set of symbols, locally constant functions are always Lipschitz continuous with respect to a natural metric on its symbolic space. However, for the symbolic dynamics with uncountable symbols [0,1][0,1], locally constant functions are not always Lipschitz continuous with respect to the metric dd on XX.

Hereafter, we always assume (H3)(H_{3}) and (H4)(H_{4}) for the Lipschitz continuous function φ(x¯)=h(x0,x1)\varphi(\underline{x})=h(x_{0},x_{1}), where the assumptions (H3)(H_{3}) and (H4)(H_{4}) are defined by:

  • (H3)(H_{3})

    If ξ1<ξ2\xi_{1}<\xi_{2} and η1<η2\eta_{1}<\eta_{2}, then

    h(ξ1,η1)+h(ξ2,η2)<h(ξ1,η2)+h(ξ2,η1).h(\xi_{1},\eta_{1})+h(\xi_{2},\eta_{2})<h(\xi_{1},\eta_{2})+h(\xi_{2},\eta_{1}).
  • (H4)(H_{4})

    If both (x1,x0,x1)(x_{-1},x_{0},x_{1}) and (x1,x0,x1)(x^{{}^{\prime}}_{-1},x_{0},x^{{}^{\prime}}_{1}) with (x1,x0,x1)(x1,x0,x1)(x_{-1},x_{0},x_{1})\neq(x^{{}^{\prime}}_{-1},x_{0},x^{{}^{\prime}}_{1}) are minimal, then

    (x1x1)(x1x1)<0.(x_{-1}-x^{{}^{\prime}}_{-1})(x_{1}-x^{{}^{\prime}}_{1})<0.

Here, we give the definition of the word minimal :

Definition 4.3 (Minimal).

Fix k,l0k,l\in\mathbb{N}_{0} with k<lk<l arbitrarily. A finite word {xi}i=kl\{x_{i}\}_{i=k}^{l} is said to be minimal if, for any {yi}i=kl\{y_{i}\}_{i=k}^{l} with yk=xky_{k}=x_{k} and yl=xly_{l}=x_{l}, we have:

i=kl1h(xi,xi+1)i=kl1h(yi,yi+1).\sum_{i=k}^{l-1}h(x_{i},x_{i+1})\leq\sum_{i=k}^{l-1}h(y_{i},y_{i+1}).

Moreover, an infinite word {xi}i0\{x_{i}\}_{i\in\mathbb{N}_{0}} is said to be minimal if {xi}i=kl\{x_{i}\}_{i=k}^{l} is minimal for any k,lk,l with k<lk<l.

Remark 4.4.

We state some remarks about the settings mentioned above.

  1. (1)

    In (H3)(H_{3}), the equality h(ξ1,η1)+h(ξ2,η2)=h(ξ1,η2)+h(ξ2,η1)h(\xi_{1},\eta_{1})+h(\xi_{2},\eta_{2})=h(\xi_{1},\eta_{2})+h(\xi_{2},\eta_{1}) holds if ξ1=ξ2\xi_{1}=\xi_{2} or η1=η2\eta_{1}=\eta_{2}.

  2. (2)

    The assumptions (H3)(H_{3}) and (H4)(H_{4}) hold if hh satisfies the twist condition D1D2h<0D_{1}D_{2}h<0. In fact, for ξ1<ξ2\xi_{1}<\xi_{2} and η1<η2\eta_{1}<\eta_{2}, we have:

    0\displaystyle 0 >η1η2ξ1ξ2D1D2H(x,y)𝑑x𝑑y\displaystyle>\int_{\eta_{1}}^{\eta_{2}}\int_{\xi_{1}}^{\xi_{2}}D_{1}D_{2}H(x,y)dxdy
    =η1η2D2H(ξ2,y)D2H(ξ1,y)dy\displaystyle=\int_{\eta_{1}}^{\eta_{2}}D_{2}H(\xi_{2},y)-D_{2}H(\xi_{1},y)dy
    =H(ξ1,η1)+H(ξ2.η2)H(ξ1,η2)H(ξ2,η1).\displaystyle=H(\xi_{1},\eta_{1})+H(\xi_{2}.\eta_{2})-H(\xi_{1},\eta_{2})-H(\xi_{2},\eta_{1}).

    This inequality implies (H3)(H_{3}). Next, we show (H4)(H_{4}). Suppose that both (x1,x0,x1)(x_{-1},x_{0},x_{1}) and (x1,x0,x1)(x^{\ast}_{-1},x_{0},x^{\ast}_{1}) are minimal and

    (x1,x0,x1)(x1,x0,x1).(x_{-1},x_{0},x_{1})\neq(x^{\ast}_{-1},x_{0},x^{\ast}_{1}).

    Clearly,

    D1H(x0,x1)+D2H(x1,x0)\displaystyle D_{1}H(x_{0},x_{1})+D_{2}H(x_{-1},x_{0}) =0,and\displaystyle=0,and
    D1H(x0,x1)+D2H(x1,x0)\displaystyle D_{1}H(x_{0},x_{1}^{\ast})+D_{2}H(x^{\ast}_{-1},x_{0}) =0.\displaystyle=0.

    Since D1D2H<0D_{1}D_{2}H<0, D1H(x,y)D_{1}H(x,y) is monotonically decreasing with respect to yy, and D2H(x,y)D_{2}H(x,y) is monotonically decreasing with respect to xx. Hence, if two minimal segment satisfies x1x1<0x_{-1}-x_{-1}^{\ast}<0 and x1x1<0x_{1}-x_{1}^{\ast}<0, then

    D1H(x0,x1)+D2H(x1,x0)>D1H(x0,x1)+D2H(x1,x0),\displaystyle D_{1}H(x_{0},x_{1})+D_{2}H(x_{-1},x_{0})>D_{1}H(x_{0},x_{1}^{\ast})+D_{2}H(x^{\ast}_{-1},x_{0}),

    which is contradiction.

  3. (3)

    The above notations and labels are derived from [Ban88]. In his paper, hh is considered as a fuction on 2\mathbb{R}^{2} satisfying (H1)(H4)(H_{1})-(H_{4}), where (H1)(H_{1}) and (H2)(H_{2}) are given by:

    • (H1)(H_{1})

      h(ξ,η)=h(ξ+1,η+1)h(\xi,\eta)=h(\xi+1,\eta+1) for all (ξ,η)2(\xi,\eta)\in\mathbb{R}^{2}, and

    • (H2)(H_{2})

      lim|η|h(ξ,ξ+η)=\displaystyle{\lim_{|\eta|\to\infty}h(\xi,\xi+\eta)=\infty}  uniformly in  ξ\xi.

    Note that the condition (H2)(H_{2}) holds if hh satisfies (H1)(H_{1}) and the twist condition. For the proof, integrate D2D1hD_{2}D_{1}h over the triangular region bounded by three points (ξ,ξ)(\xi,\xi), (ξ,ξ+η)(\xi,\xi+\eta), and (ξ+η,ξ+η)(\xi+\eta,\xi+\eta). In addition, the fact that the differentiability of hh is no longer needed is useful when considering geodesics on 𝕋2\mathbb{T}^{2}, for example (see Section 66 in [Ban88] for the detail).

Let h=minx[0,1]h(x,x)h^{\ast}=\min_{x\in[0,1]}h(x,x). Set X(n)(X=[0,1]0)X(n){(\subset X=[0,1]^{\mathbb{N}_{0}})} and m([0,1])\mathrm{m}\ {(\subset[0,1])} by

X(n)={x¯Xxi=xn+ifor alli0},X(n)=\{\underline{x}\in X\mid x_{i}=x_{n+i}\ \text{for all}\ i\in\mathbb{N}_{0}\},

and

m={a[0,1]h(a,a)=h}.\mathrm{m}=\{a\in[0,1]\mid h(a,a)=h^{\ast}\}.

Since hh is non-constant by (H3)(H_{3}) and hh is continuous, the set m\mathrm{m} is nonempty, compact and m[0,1]\mathrm{m}\subsetneq[0,1]. Firstly, we introduce the result of [Ban88] and [Yu22].

Lemma 4.5.

Fix nn\in\mathbb{N}. Then i=0n1(h(xi,xi+1)h)0\displaystyle{\sum_{i=0}^{n-1}(h(x_{i},x_{i+1})-h^{\ast})\geq 0} for any x¯X(n)\underline{x}\in X(n). Moreover, the equality is true if and only if there exists ama\in\mathrm{m} satisfying x¯i=a\underline{x}_{i}=a for i=0,,n1i=0,\cdots,n-1.

This claim is essentially the same as Lemma 2.5 of [Yu22]. The proof is almost the same as the proof of the case p=0p=0 in Theorem 3.3 of [Ban88].

Proof of Lemma 4.5.

Fix nn\in\mathbb{N} arbitrarily. It is easily seen that if x¯\underline{x}^{\ast} is a minimizer in X(n)X(n), i.e., satisfies:

Snφ(x¯)=minx¯X(n)Snφ(x¯),S_{n}\varphi(\underline{x}^{\ast})=\min_{\underline{x}\in X(n)}S_{n}\varphi(\underline{x}),

so is σk(x¯)\sigma^{k}(\underline{x}^{\ast}) for any kk\in\mathbb{N}. Set mm and MM with 0m,Mn10\leq m,M\leq n-1 for each xX(n)x\in X(n) by:

(10) xm=min0in1xi,xM=max0in1xi\displaystyle x_{m}=\min_{0\leq i\leq n-1}x_{i},\ x_{M}=\max_{0\leq i\leq n-1}x_{i}

For the proof of the claim, it suffices to show the following claim:

Claim 4.6.

xm=xMx_{m}=x_{M} if x¯\underline{x} is a minimizer in X(n)X(n).

Suppose that the claim is failed, i.e., x¯\underline{x} is a minimizer in X(n)X(n) and xm<xMx_{m}<x_{M}. We consider only the case m=0m=0. The other cases can be shown in a similar way.

Firstly, we introduce the definition of cross. We say the segments {ξi}0n1\{\xi_{i}\}_{0}^{n-1} and {ηi}0n1\{\eta_{i}\}_{0}^{n-1} cross if (ξlηl)(ξl+1ηl+1)0(\xi_{l}-\eta_{l})(\xi_{l+1}-\eta_{l+1})\leq 0 for some l{0,,n1}l\in\{0,\ldots,n-1\}. Let y¯=σM(x¯)\underline{y}=\sigma^{M}(\underline{x}), i.e., yi=xi+My_{i}=x_{i+M} for i0i\in\mathbb{N}_{0}. Note that {xi}0n1\{x_{i}\}_{0}^{n-1} and {yi}0n1\{y_{i}\}_{0}^{n-1} cross at least once because if not we have xiyix_{i}\leq y_{i} for all i{1,,n1}i\in\{1,\ldots,n-1\} and thus the inequality x0<y0x_{0}<y_{0} yields j=0n1xj<j=0n1yj\sum_{j=0}^{n-1}x_{j}<\sum_{j=0}^{n-1}y_{j}, but the periodicity of x¯\underline{x} and y¯\underline{y} yields j=0n1yi=j=0n1xi+M=j=0n1xi\sum_{j=0}^{n-1}y_{i}=\sum_{j=0}^{n-1}x_{i+M}=\sum_{j=0}^{n-1}x_{i}.

Suppose that {xi}0n1\{x_{i}\}_{0}^{n-1} and {yi}0n1\{y_{i}\}_{0}^{n-1} cross only once. The other cases can be shown by repeating the following discussion. Let ll^{*} be the largest number among {0,,n1}\{0,\ldots,n-1\} such that xkykx_{k}\leq y_{k} for all k{1,,l}k\in\{1,\ldots,l^{*}\}. Then we have xlylx_{l^{*}}\leq y_{l^{*}} and xl+1>yl+1x_{l^{*}+1}>y_{l^{*}+1}. Define w¯,z¯X(n)\underline{w},\underline{z}\in X(n) by wi=min{xi,yi},zi=max{xi,yi}w_{i}=\min\{x_{i},y_{i}\},z_{i}=\max\{x_{i},y_{i}\} for all i0i\in\mathbb{N}_{0}. When xlylx_{l^{*}}\neq y_{l^{*}}, by (H3)(H_{3}), xl<ylx_{l^{*}}<y_{l^{*}} and yl+1<xl+1y_{l^{*}+1}<x_{l^{*}+1} give

h(wl,wl+1)+h(zl,zl+1)\displaystyle h(w_{l^{*}},w_{l^{*}+1})+h(z_{l^{*}},z_{l^{*}+1}) =h(xl,yl+1)+h(yl,xl+1)\displaystyle=h(x_{l^{*}},y_{l^{*}+1})+h(y_{l^{*}},x_{l^{*}+1})
<h(xl,xl+1)+h(yl,yl+1).\displaystyle<h(x_{l^{*}},x_{l^{*}+1})+h(y_{l^{*}},y_{l^{*}+1}).

As a result, it holds that:

Snφ(w¯)+Snφ(z¯)<Snφ(x¯)+Snφ(σM(x¯))=2minx¯X(n)Snφ(x¯).S_{n}\varphi(\underline{w})+S_{n}\varphi(\underline{z})<S_{n}\varphi(\underline{x})+S_{n}\varphi(\sigma^{M}(\underline{x}))=2\min_{\underline{x}^{\prime}\in X(n)}S_{n}\varphi(\underline{x}^{\prime}).

Therefore, at least one of the following inequalities holds:

Snφ(w¯)<minx¯X(n)Snφ(x¯)orSnφ(z¯)<minx¯X(n)Snφ(x¯),S_{n}\varphi(\underline{w})<\min_{\underline{x}^{\prime}\in X(n)}S_{n}\varphi(\underline{x}^{\prime})\quad\text{or}\quad S_{n}\varphi(\underline{z})<\min_{\underline{x}^{\prime}\in X(n)}S_{n}\varphi(\underline{x}^{\prime}),

which is a contradiction (note that w¯,z¯X(n)\underline{w},\underline{z}\in X(n)). When xl=ylx_{l^{*}}=y_{l^{*}}, we have (xl1yl1)(xl+1yl+1)0(x_{l^{*}-1}-y_{l^{*}-1})(x_{l^{*}+1}-y_{l^{*}+1})\leq 0, but the equality cannot occur by (H4)(H_{4}) and the minimality of x¯\underline{x} and y¯\underline{y}. Thus xl1<yl1x_{l^{*}-1}<y_{l^{*}-1} and xl+1>yl+1x_{l^{*}+1}>y_{l^{*}+1} hold and we have:

h(wl1,wl)+h(zl1,zl)\displaystyle h(w_{l^{*}-1},w_{l^{*}})+h(z_{l^{*}-1},z_{l^{*}}) =h(xl1,xl)+h(yl,yl),and\displaystyle=h(x_{l^{*}-1},x_{l^{*}})+h(y_{l^{*}},y_{l^{*}}),\ \text{and}
h(wl,wl+1)+h(zl,zl+1)\displaystyle h(w_{l^{*}},w_{l^{*}+1})+h(z_{l^{*}},z_{l^{*}+1}) =h(yl,yl+1)+h(xl,xl+1),\displaystyle=h(y_{l^{*}},y_{l^{*}+1})+h(x_{l^{*}},x_{l^{*}+1}),

as seen in Remark 4.4 (1). Therefore, we obtain

Snφ(w¯)+Snφ(z¯)=Snφ(x¯)+Snφ(σM(x¯))=2minx¯X(n)Snφ(x¯).S_{n}\varphi(\underline{w})+S_{n}\varphi(\underline{z})=S_{n}\varphi(\underline{x})+S_{n}\varphi(\sigma^{M}(\underline{x}))=2\min_{\underline{x}^{\prime}\in X(n)}S_{n}\varphi(\underline{x}^{\prime}).

This implies that both w¯\underline{w} and z¯\underline{z} are also minimal, which is a contradiction for (H4)(H_{4}). ∎

Using Lemma 4.5, we can easily show a part of our main theorem.

Lemma 4.7 (cf. Main Theorem 3. (1)).

αφ=h\alpha_{\varphi}=h^{\ast}

Proof.

The continuity of hh and Lemma 4.5 imply that for any x¯X\underline{x}\in X,

1nSnφ(x¯)\displaystyle\frac{1}{n}S_{n}\varphi(\underline{x}) =1n(h(x0,x1)++h(xn1,x0)+h(xn1,xn)h(xn1,x0))\displaystyle=\frac{1}{n}(h(x_{0},x_{1})+\cdots+h(x_{n-1},x_{0})+h(x_{n-1},x_{n})-h(x_{n-1},x_{0}))
1n(h(x0,x1)++h(xn1,x0))1n(hmaxhmin)\displaystyle\geq\frac{1}{n}(h(x_{0},x_{1})+\cdots+h(x_{n-1},x_{0}))-\frac{1}{n}(h_{\max}-h_{\min})
h1n(hmaxhmin).\displaystyle\geq h^{\ast}-\frac{1}{n}(h_{\max}-h_{\min}).\

By (1)\eqref{eq:Jenkinson} in Section 1, we have

αφ=infx¯Xlim infn1nSnφ(x¯)h.\alpha_{\varphi}=\inf_{\underline{x}\in X}\liminf_{n\to\infty}\frac{1}{n}S_{n}\varphi(\underline{x})\geq h^{\ast}.

Moreover, taking x¯=a\underline{x}^{\ast}=a^{\infty} for some ama\in\mathrm{m}, we obtain

1nSnφ(x¯)=h,\displaystyle\frac{1}{n}S_{n}\varphi(\underline{x}^{\ast})=h^{\ast},

which implies that αφ=h.\alpha_{\varphi}=h^{\ast}.

Applying Lemma 4.5 and 4.7, we immediately obtain an explicit formula for optimizing periodic measures.

Theorem 4.8 (cf. Main Theorem 3. (2)).

Let φ:X\varphi:X\to\mathbb{R} be a 2-locally constant function φ(x¯)=h(x0,x1)\varphi(\underline{x})=h(x_{0},x_{1}) where h:[0,1]2h:[0,1]^{2}\to\mathbb{R} is a Lipschitz continuous function on [0,1]2[0,1]^{2} with (H3)(H_{3}) and (H4)(H_{4}). Then

min(φ)p={δaam},\mathcal{M}_{{\rm min}}(\varphi)\cap\mathcal{M}^{\mathrm{p}}=\{\delta_{a^{\infty}}\mid a\in\mathrm{m}\},

where p\mathcal{M}^{\mathrm{p}} stands for the set of invariant probability measures supported on a single periodic orbit.

Next, we give an explicit characterization for elements in the Aubry set of φ\varphi. Consider the distance between a point x[0,1]x\in[0,1] to the closed set m\mathrm{m}:

d(x,m)=infam|xa|.d_{\mathbb{R}}(x,\mathrm{m})=\inf_{a\in\mathrm{m}}|x-a|.

The following lemma plays a key role in our statement.

Lemma 4.9 (A slightly extended version of Lemma 2.7 of [Yu22]).

Set

ϕ(δ)=infnϕ(δ;n)\phi(\delta)=\inf_{n\in\mathbb{N}}\phi(\delta;n)

where

ϕ(δ;n)=inf{Sn(φαφ)(x¯)x¯X(n),max0in1d(xi,m)δ)}.\phi(\delta;n)=\inf\{S_{n}(\varphi-\alpha_{\varphi})(\underline{x})\mid\underline{x}\in X(n),\ \max_{0\leq i\leq n-1}d_{\mathbb{R}}({x}_{i},\mathrm{m})\geq\delta)\}.

Then ϕ(δ)>0\phi(\delta)>0 if δ>0\delta>0.

Remark 4.10.

Let

X(n;δ)={x¯X(n)|max0in1d(xi,m)δ}.X(n;\delta)=\{\underline{x}\in X(n)|\max_{0\leq i\leq n-1}d_{\mathbb{R}}({x}_{i},\mathrm{m})\geq\delta\}.

Then ϕ(δ)\phi(\delta) should be defined as ++\infty if X(n;δ)=X(n;\delta)=\emptyset. By the definition, if δ<δ\delta^{\prime}<\delta then X(n;δ)X(n;δ)X(n;\delta^{\prime})\supset X(n;\delta) and ϕ(δ)<ϕ(δ)\phi(\delta^{\prime})<\phi(\delta).

Proof of Lemma 4.9.

Lemma 2.7 in [Yu22] corresponds to the case of #m=2\#\mathrm{m}=2, i.e., m={u0,u1}\mathrm{m}=\{u_{0},u_{1}\} for some u0,u1[0,1]u_{0},u_{1}\in[0,1] and

d(xi,m)=minj=0,1|xiuj|,d_{\mathbb{R}}({x}_{i},\mathrm{m})=\min_{j=0,1}|x_{i}-u_{j}|,

but our proof is almost the same as his proof. It is sufficient to prove that for δ>0\delta>0,

  1. (1)

    ϕ(δ;1)>0\phi(\delta;1)>0, and

  2. (2)

    ϕ(δ;n)ϕ(δ;1)\phi(\delta;n)\geq\phi(\delta;1) for all nn\in\mathbb{N}.

The first claim is clear since x¯\underline{x} is given by x0x_{0}^{\infty} for some x0mx_{0}\not\in\mathrm{m}. We show the second using induction. The case of n=1n=1 is trivial. Assume that it holds if nm1n\leq m-1. Take any x¯X(m;δ)\underline{x}\in X(m;\delta). When xjx0x_{j}\neq x_{0} for any j{1,,m1}j\in\{1,\ldots,m-1\}, there exists an integer k{1,,m1}k\in\{1,\ldots,m-1\} such that

(xkxk1)(xkxk+1)0(x_{k}-x_{k-1})(x_{k}-x_{k+1})\geq 0

since x0=xmx_{0}=x_{m}. From this inequality, (H3)(H_{3}), and Remark 4.4 (1), we have

h(xk,xk)+h(xk1,xk+1)h(xk1,xk)+h(xk,xk+1).h(x_{k},x_{k})+h(x_{k-1},x_{k+1})\leq h(x_{k-1},x_{k})+h(x_{k},x_{k+1}).

Set

u¯=xk,v¯=(x0xk1xk+1xm1).\underline{u}=x_{k}^{\infty},\qquad\underline{v}=(x_{0}\cdots x_{k-1}x_{k+1}\cdots x_{m-1})^{\infty}.

Then

Sm(φαφ)(x¯)S1(φαφ)(u¯)+Sm1(φαφ)(v¯).S_{m}(\varphi-\alpha_{\varphi})(\underline{x})\geq S_{1}(\varphi-\alpha_{\varphi})(\underline{u})+S_{m-1}(\varphi-\alpha_{\varphi})(\underline{v}).

Clearly, u¯X(1)\underline{u}\in X(1) and v¯X(m1)\underline{v}\in X(m-1) hold and thus we obtain the following two inequalities by Lemma 4.5:

S1(φαφ)(u¯)0,Sm1(φαφ)(v¯)0.S_{1}(\varphi-\alpha_{\varphi})(\underline{u})\geq 0,\qquad S_{m-1}(\varphi-\alpha_{\varphi})(\underline{v})\geq 0.

Moreover, at least one of

d(xk,m)δd_{\mathbb{R}}({x}_{k},\mathrm{m})\geq\delta

and

maxi[0,m]\{k}d(xi,m)δ\max_{i\in[0,m]\backslash\{k\}}d_{\mathbb{R}}({x}_{i},\mathrm{m})\geq\delta

is true, which yields at least one of the following inequalities:

S1(φαφ)(u¯)ϕ(δ;1),Sm1(φαφ)(v¯)ϕ(δ;m1).S_{1}(\varphi-\alpha_{\varphi})(\underline{u})\geq\phi(\delta;1),\qquad S_{m-1}(\varphi-\alpha_{\varphi})(\underline{v})\geq\phi(\delta;m-1).

By the assumption of induction ϕ(δ;n)ϕ(δ;1)>0\phi(\delta;n)\geq\phi(\delta;1)>0 for n{1,,m1}n\in\{1,\ldots,m-1\}, we have

Sm(φαφ)(x¯)\displaystyle S_{m}(\varphi-\alpha_{\varphi})(\underline{x}) S1(φαφ)(u¯)+Sm1(φαφ)(v¯)ϕ(δ;1)>0.\displaystyle\geq S_{1}(\varphi-\alpha_{\varphi})(\underline{u})+S_{m-1}(\varphi-\alpha_{\varphi})(\underline{v})\geq\phi(\delta;1)>0.

When there exists jj such that xj=x0x_{j}=x_{0}, then we get

Sm(φαφ)(x¯)ϕ(δ;j)+ϕ(δ;mj)2ϕ(δ;1)>0.\displaystyle S_{m}(\varphi-\alpha_{\varphi})(\underline{x})\geq\phi(\delta;j)+\phi(\delta;m-j)\geq 2\phi(\delta;1)>0.

The proof is complete. ∎

Using these lemmata, we can show:

Theorem 4.11 (cf. Main Theorem 3. (3)).

Ωφm0\Omega_{\varphi}\subset\mathrm{m}^{\mathbb{N}_{0}}.

Proof.

It suffices to show that x¯Ωφ\underline{x}\not\in\Omega_{\varphi} if x¯X\m0\underline{x}\in X\backslash\mathrm{m}^{\mathbb{N}_{0}}. There exists an integer NN such that

max0iNd(xi,m)δ.\max_{0\leq i\leq N}d_{\mathbb{R}}({x}_{i},\mathrm{m})\geq{\delta}.

since x¯X\m0\underline{x}\in X\backslash\mathrm{m}^{\mathbb{N}_{0}}. It follows from the compactness of XX and continuity of hh that φ\varphi is uniformly continuous. Thus we can take ε0>0\varepsilon_{0}>0 satisfying if |η1η2|<ε0|\eta_{1}-\eta_{2}|<\varepsilon_{0}, then:

|h(ξ,η1)h(ξ,η2)|<12ϕ(δ/2),|h(\xi,\eta_{1})-h(\xi,\eta_{2})|<\frac{1}{2}\phi(\delta/2),

where the function ϕ\phi is given in Lemma 4.9. Fix ε(0,min{2N,2(N+2)δ,ε0/2})\varepsilon\in(0,\min\{{2^{-N}},{2^{-(N+2)}\delta},{\varepsilon_{0}}/{2}\}). Let kk be a large number satisfying 2k<ε2^{-k}<\varepsilon and B(x¯,x¯,k;ε)B(\underline{x},\underline{x},k;\varepsilon)\neq\emptyset. For z¯B(x¯,x¯,k;ε)\underline{z}\in B(\underline{x},\underline{x},k;\varepsilon), set w¯=w¯(k)\underline{w}=\underline{w}(k) by

w¯(k)=(z0zk1).\underline{w}(k)=(z_{0}\cdots z_{k-1})^{\infty}.

Immediately,

|z0zk|d(z¯,σk(z¯))d(z¯,x¯)+d(x¯,σk(z¯))<2ϵ<ϵ0\displaystyle|z_{0}-z_{k}|\leq d(\underline{z},\sigma^{k}(\underline{z}))\leq d(\underline{z},\underline{x})+d(\underline{x},\sigma^{k}(\underline{z}))<2\epsilon<\epsilon_{0}

and

|Sk(φαφ)(z¯)Sk(φαφ)(w¯)|\displaystyle|S_{k}(\varphi-\alpha_{\varphi})(\underline{z})-S_{k}(\varphi-\alpha_{\varphi})(\underline{w})| =|h(zk1,zk)h(zk1,z0)|<12ϕ(δ/2).\displaystyle=|h(z_{k-1},z_{k})-h(z_{k-1},z_{0})|<\frac{1}{2}\phi(\delta/2).

Thus we get

(11) Sk(φαφ)(z¯)>Sk(φαφ)(w¯(k))12ϕ(δ/2).\displaystyle S_{k}(\varphi-\alpha_{\varphi})(\underline{z})>S_{k}(\varphi-\alpha_{\varphi})(\underline{w}(k))-\frac{1}{2}\phi(\delta/2).

It is seen that w¯B(x¯,x¯,k;2ε)X(k)\underline{w}\in B(\underline{x},\underline{x},k;2\varepsilon)\cap X(k) since kk satisfies 12k<ε\frac{1}{2^{k}}<\varepsilon. Moreover, by ε<min{12N,δ2N+2}\varepsilon<\min\{\frac{1}{2^{N}},\frac{\delta}{2^{N+2}}\} and kk with 2k<ε2^{-k}<\varepsilon (so kNk\geq N), we see that

max0iNd(wi,m)δ2.\max_{0\leq i\leq N}d_{\mathbb{R}}({w}_{i},\mathrm{m})\geq\frac{\delta}{2}.

Hence the estimate (11) shows:

Sk(φαφ)(z¯)>12ϕ(δ/2)>0S_{k}(\varphi-\alpha_{\varphi})(\underline{z})>\frac{1}{2}\phi(\delta/2)>0

and we obtain

H(x¯,x¯)>0.H(\underline{x},\underline{x})>0.

This is the desired inequality. ∎

Proof of Main Theorem 3.

We already have (1)-(3) in Main Theorem 3. We now prove the rest part. Suppose that h(x,x)h(x,x) has a unique minimum point aa_{*} in [0,1][0,1]. Then m={a}\mathrm{m}=\{a_{*}\}, and thus Ωφ={a}\Omega_{\varphi}=\{a_{*}^{\infty}\}. Since the Mather set φ\mathscr{M}_{\varphi} is not empty and included in Ωφ\Omega_{\varphi}, we obtain the desired result. ∎

5. TPO property

In this section, we discuss typical properties of optimal measures. We still consider the case that the potential function is 2-locally constant.

As stated in the last part of Section 4, by Theorem 4.11, we see that if x¯Ωφ\underline{x}\in\Omega_{\varphi} then x¯m0\underline{x}\in\mathrm{m}^{\mathbb{N}_{0}}. Therefore, if m={a[0,1]h(a,a)=minx[0,1]h(x,x)}\mathrm{m}=\{a\in[0,1]\mid h(a,a)=\min_{x\in[0,1]}h(x,x)\} for φ=h(x,y)\varphi=h(x,y) is a singleton {a}\{a\}, we have Ωh={a}\Omega_{h}=\{a^{\infty}\} and it must coincide with the Mather set of hh. As described below, In this section, we will see that such hh is typical.

Let r2r\geq 2 be a positive integer. Consider the set of CrC^{r}-variational structure with the twist condition,

r={hCr([0,1]2;)D2D1h<0},\mathscr{H}^{r}=\{h\in C^{r}([0,1]^{2};\mathbb{R})\mid D_{2}D_{1}h<0\},

equipped with the CrC^{r}-norm, i.e.,

hCr=|β|rsup(x,y)[0,1]2|βh(x,y)|.||h||_{C^{r}}=\sum_{|\beta|\leq r}\sup_{(x,y)\in[0,1]^{2}}|\partial^{\beta}h(x,y)|.

For hrh\in\mathscr{H}^{r}, let h~(x)=h(x,x)\tilde{h}(x)=h(x,x) for x[0,1]x\in[0,1].

Proposition 5.1.

The subset 𝒪\mathscr{O} in r\mathscr{H}^{r} given by

𝒪={hr|h~has a unique minimum pointx(h)s.t.h~′′(x(h))>0}\mathscr{O}=\{h\in\mathscr{H}^{r}|\ \tilde{h}\ \text{has a unique minimum point}\ x^{*}(h)\ s.t.\ \tilde{h}^{\prime\prime}(x^{*}(h))>0\}

is CrC^{r} open and dense in r\mathscr{H}^{r}.

Proof.

(Dense): Take hrh\in\mathscr{H}^{r}. Let a[0,1]a\in[0,1] be a minimum point of h~\tilde{h}. Take arbitrary ε>0\varepsilon>0. Set

hε(x,y):=h(x,y)+εV(x)h_{\varepsilon}(x,y):={h}(x,y)+\varepsilon V(x)

where VCr([0,1];)V\in C^{r}([0,1];\mathbb{R}) with a unique minimum point at x=ax=a s.t. V(a)0,V′′(a)>0V(a)\geq 0,V^{\prime\prime}(a)>0 (e.g., V(x)=cos(2π(xa)+π)V(x)=\cos(2\pi(x-a)+\pi) or (xa)2(x-a)^{2}). Then hεhC2=εVC2||{h}_{\varepsilon}-{h}||_{C^{2}}=\varepsilon||V||_{C^{2}} and D2D1hε=D2D1h<0D_{2}D_{1}h_{\varepsilon}=D_{2}D_{1}h<0. Moreover, for xax\neq a

hε(x,x)=h(x,x)+εV(x)h(a,a)+εV(a)>h(a,a)=hε(a,a).{h}_{\varepsilon}(x,x)=h(x,x)+\varepsilon V(x)\geq h(a,a)+\varepsilon V(a)>h(a,a)=h_{\varepsilon}(a,a).

Therefore, hε𝒪h_{\varepsilon}\in\mathscr{O} holds.

(Open): Let h𝒪h\in\mathscr{O}. Denote the corresponding minimum point xhx_{h} of h~\tilde{h}. Take ε(0,h~′′(xh)/2)\varepsilon\in(0,\tilde{h}^{\prime\prime}(x_{h})/2) Let hεh_{\varepsilon}\in\mathscr{H} s.t. hεhCr<ε||h_{\varepsilon}-h||_{C^{r}}<\varepsilon. Then we have

|(h~εh~)′′(xh)|\displaystyle|(\tilde{h}_{\varepsilon}-\tilde{h})^{\prime\prime}(x_{h})| =|D1D1(hεh)(xh,xh)+D1D2(hεh)(xh,xh)\displaystyle=|D_{1}D_{1}({h}_{\varepsilon}-{h})(x_{h},x_{h})+D_{1}D_{2}({h}_{\varepsilon}-{h})(x_{h},x_{h})
+D2D1(hεh)(xh,xh)+D2D2(hεh)(xh,xh)|\displaystyle\qquad+D_{2}D_{1}({h}_{\varepsilon}-{h})(x_{h},x_{h})+D_{2}D_{2}({h}_{\varepsilon}-{h})(x_{h},x_{h})|
hεhCr<ε,\displaystyle\leq||h_{\varepsilon}-h||_{C^{r}}<\varepsilon,

i.e.,

h~ε′′(xh)h~′′(xh)ε>h~′′(xh)/2>0.\tilde{h}_{\varepsilon}^{\prime\prime}(x_{h})\geq{\tilde{h}^{\prime\prime}(x_{h})}-\varepsilon>\tilde{h}^{\prime\prime}(x_{h})/2>0.

This also implies that hεh_{\varepsilon} has a unique minimum point at x=xhx=x_{h}, which yields hε𝒪h_{\varepsilon}\in\mathscr{O}. ∎

Similarly we easily obtain the following perturbation result in the sense of Mañé. Note that for any hrh\in\mathscr{H}^{r} and fCr([0,1];)f\in C^{r}([0,1];\mathbb{R}) the function h(x,y)+f(x)h(x,y)+f(x) satisfies the twist condition, i.e., D2D1(h+f)=D2D1h<0D_{2}D_{1}(h+f)=D_{2}D_{1}h<0.

Proposition 5.2.

For arbitrary hrh\in\mathscr{H}^{r}, the set

𝒱h\displaystyle\mathscr{V}_{h} ={VCr([0,1];)|h~+Vhas a unique minimum point\displaystyle=\{V\in C^{r}([0,1];\mathbb{R})|\ \widetilde{h}+V\ \text{has a unique minimum point}
x(h~+V)s.t.(h~+V)′′(x(h~+V))>0}\displaystyle\qquad\qquad\qquad\qquad\qquad x^{*}(\tilde{h}+V)\ s.t.\ (\widetilde{h}+V)^{\prime\prime}(x^{*}(\tilde{h}+V))>0\}

is CrC^{r} open and dense in 𝒱h\mathscr{V}_{h} in Cr([0,1];)C^{r}([0,1];\mathbb{R}).

Proof of Main Theorem 4.

Combining Propositions 5.1,5.2 and Main Theorem 3, we obtain Main Theorem 4. ∎


Acknowledgement.  The first author was partially supported by JSPS KAKENHI Grant Number 23H01081 and 23K19009. The third author was partially supported by JSPS KAKENHI Grant Number 21K13816.


Data Availability.  Data sharing not applicable to this article as no datasets were generated or analyzed during the current study.

References

  • \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry