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

Constrained ergodic optimization for generic continuous functions

Shoya Motonaga Research Organization of Science and Technology, Ritsumeikan University, 1-1-1 Noji-higashi, Kusatsu, Shiga 525-8577, Japan [email protected]  and  Mao Shinoda Department of Mathematics, Ochanomizu University, 2-1-1 Otsuka, Bunkyo-ku, Tokyo, 112-8610, Japan [email protected]
Abstract.

One of the fundamental results of ergodic optimization asserts that for any dynamical system on a compact metric space with the specification property and for a generic continuous function ff every invariant probability measure that maximizes the space average of ff must have zero entropy. We establish the analogical result in the context of constrained ergodic optimization, which is introduced by Garibaldi and Lopes (2007).

Key words and phrases:
2010 Mathematics Subject Classification:
Primary 37E45, 37B10, 37A99

1. Introduction

Let T:XXT:X\rightarrow X be a continuous map on a compact metric space and T(X)\mathcal{M}_{T}(X) be the space of Borel probability measures endowed with the weak*-topology. Let C(X,)C(X,\mathbb{R}) be the space of real-valued continuous functions on XX with the supremum norm \|\cdot\|_{\infty}. For each fC(X,)f\in C(X,\mathbb{R}) we consider the maximum ergodic average

(1) β(f)=supμT(X)f𝑑μ\displaystyle\beta(f)=\sup_{\mu\in\mathcal{M}_{T}(X)}\int f\ d\mu

and the set of all maximizing measures of ff

(2) max(f)={μT(X):f𝑑μ=β(f)}.\displaystyle\mathcal{M}_{{\rm max}}(f)=\left\{\mu\in\mathcal{M}_{T}(X):\int f\ d\mu=\beta(f)\right\}.

The functional β\beta and the set max(f)\mathcal{M}_{{\rm max}}(f) are main objects in ergodic optimization, which has been actively studied for a decade (see for more details [Jen06survey, Jen2017]).

Constrained ergodic optimization, which is introduced by Garibaldi and Lopes in [GarLop07], investigates the analogical objects as (1) and (2) under some constraint. Introducing a continuous d\mathbb{R}^{d}-valued function φC(X,d)\varphi\in C(X,\mathbb{R}^{d}) playing the role of a constraint, we define the rotation set of φ=(φ1,,φd)\varphi=(\varphi_{1},\ldots,\varphi_{d}) by

Rot(φ)={rvφ(μ)d:μT(X)},\displaystyle{\rm Rot}(\varphi)=\left\{{\rm rv}_{\varphi}(\mu)\in\mathbb{R}^{d}:\mu\in\mathcal{M}_{T}(X)\right\},

where rvφ(μ)=(φ1𝑑μ,,φd𝑑μ){\rm rv}_{\varphi}(\mu)=(\int\varphi_{1}d\mu,\ldots,\int\varphi_{d}d\mu). For hRot(φ)h\in{\rm Rot}(\varphi), the fiber rvφ1(h){\rm rv}_{\varphi}^{-1}(h) is called the rotation class of hh. A point hRot(φ)h\in{\rm Rot}(\varphi) is called a rotation vector. The terminology comes from Poincaré’s rotation numbers for circle homeomorphisms. Many authors study the properties of rotation sets and characterize several dynamics in terms of rotation vectors [GM99, Jen01rotation, K92, K95, KW14, KW16, KW19].

We define the maximum ergodic average for a continuous function ff with constraint hRot(φ)h\in{\rm Rot}(\varphi) by

(3) βhφ(f)=supμrvφ1(h)f𝑑μ\displaystyle\beta^{\varphi}_{h}(f)=\sup_{\mu\in{\rm rv}_{\varphi}^{-1}(h)}\int f\ d\mu

and the set of all relative maximizing measures of ff with a rotation vector hRot(φ)h\in{\rm Rot}(\varphi) by

hφ(f)={μT(X):βhφ(f)=f𝑑μ,rvφ(μ)=h}.\displaystyle\mathcal{M}^{\varphi}_{h}(f)=\left\{\mu\in\mathcal{M}_{T}(X):\beta^{\varphi}_{h}(f)=\int f\ d\mu,\ {\rm rv}_{\varphi}(\mu)=h\right\}.

Note that this formulation is a generalization of (unconstrained) ergodic optimization, since for a constant constraint φh\varphi\equiv h and its unique rotation vector hh the rotation set of hh is T(X)\mathcal{M}_{T}(X).

We have much interest in the above constrained optimization because the constraints provide information about the asymptotic behavior of orbits. Such problem is originally studied by Mather [Mather91] and Mañé [Mane96] for Euler-Lagrange flows, where the asymptotic homological position of the trajectory in the configuration space is given as a constraint. Recently, Bochi and Rams [BochiRams16] proved that the Lyapunov optimizing measures for one-step cocycles of 2×22\times 2 matrices have zero entropy on the Mather sets under some conditions, which implies that a low complexity phenomena occurs in noncommutative setting such as a Lyapunov-optimization problem for one-step cocycles. A relative Lyapunov-optimization is mentioned for their future research but to the authors’ knowledge the classical commutative counterpart, typicality of zero entropy of relative maximizing measure, has not been established yet. We remark that Zhao [Zhao19] studies constrained ergodic optimization for asymptotically additive potentials for the application in the study of multifractal analysis.

It is natural to extend the fundamental results of unconstrained ergodic optimization to constrained one. In [GarLop07], several general results on constrained ergodic optimization are provided, especially, uniqueness of maximizing measures with any constraint for generic continuous functions is asserted. Moreover, prevalent uniqueness of maximizing measures for continuous functions in the constrained settings easily follows from [Mor21] (see Appendix B). However, in these studies, the differences between constrained ergodic optimization and unconstrained one are not mentioned explicitly. In some cases, constraints prevent existence of relatively maximizing periodic measures (see Remark 1.3), which should give rise to the problem of existence of periodic measures in a given rotation class. Moreover, in contrast to Morris’ theorem [Mor2010] which asserts that for any dynamical systems with the specification property every maximizing measure for a generic continuous function has zero entropy, we can easily verify that there exists a constraint such that for a generic potential its unique relatively maximizing measure has positive entropy (see Proposition A.1). Thus it is important to investigate the condition that the statement of Morris’s theorem holds for constrained ergodic optimization.

In this paper, we study the structure of rotation classes and the generic property in constrained ergodic optimization for symbolic dynamics. Our first main result is the density of periodic measures with some rational constraints. This is an analogical result to Sigmund’s work for a dynamical system with the specification [Sigmund]. The difficulty in the constrained case comes from the existence of a measure which is a convex combination of ergodic measures with different rotation vectors. This prevents us to use the ergodic decomposition in a given rotation class. To overcome this difficulty, our approach requires a certain finiteness for both of a subshift and a constraint function. Moreover, a detailed analysis is needed to construct a periodic measure approximating a given invariant one with the same rotation vector. Let (Ω,σ)(\Omega,\sigma) be an irreducible subshift of finite type with finite alphabets (See §2.2 below). A function φC(Ω,d)\varphi\in C(\Omega,\mathbb{R}^{d}) is said to be locally constant if for each i=1,,di=1,\ldots,d there exists ki0k_{i}\geq 0 such that

φi(x)=φi(y)ifxj=yjforj=0,,ki.\displaystyle\varphi_{i}(x)=\varphi_{i}(y)\quad\mbox{if}\quad x_{j}=y_{j}\ \mbox{for}\ j=0,\ldots,k_{i}.

Denote by σp(Ω)\mathcal{M}_{\sigma}^{p}(\Omega) the set of invariant measures supported on a single periodic orbit.

Theorem 1.

Let (Ω,σ)(\Omega,\sigma) be an irreducible subshift of finite type. Let φ=(φ1,,φd)C(Ω,d)\varphi=(\varphi_{1},\ldots,\varphi_{d})\in C(\Omega,\mathbb{Q}^{d}) be a locally constant function and hint(Rot(φ))dh\in{\rm int}({\rm Rot}(\varphi))\cap\mathbb{Q}^{d}. Then the set rvφ1(h)σp(Ω){\rm rv}_{\varphi}^{-1}(h)\cap\mathcal{M}_{\sigma}^{p}(\Omega) is dense in rvφ1(h){\rm rv}_{\varphi}^{-1}(h).

Remark 1.1.

Theorem 1 is motivated by Theorem 10 in [GarLop07]: It was shown that for a Walters potential AA on a subshift of finite type Ω\Omega, there exists a periodic measure μ\mu whose action A𝑑μ\int A\ d\mu approximates A𝑑ν\int A\ d\nu for νrvφ1(h)\nu\in{\rm rv}_{\varphi}^{-1}(h) with rvφ(μ)=h{\rm rv}_{\varphi}(\mu)=h if ν\nu is ergodic and the locally constant constraint φC(Ω,d)\varphi\in C(\Omega,\mathbb{Q}^{d}) is joint recurrent in relation to ν\nu (see [GarLop07] for the definition of the joint recurrence and their precise statement). Note that Theorem 10 in [GarLop07] was suggested by the fact that a circle homeomorphism with rational rotation number has a periodic point and a result of this kind for symbolic dynamics was studied by Ziemian (see Theorem 4.2 of [Zie95]).

Remark 1.2.

Although the properties of rotation sets are well-studied in [Zie95], to the authors’ knowledge, that of rotation classes have attracted little attention. We emphasize that Theorem 1 provides a more detailed description of Theorem 4.2 in [Zie95] under weaker assumptions.

Remark 1.3.

If a constraint φC(Ω,d)\varphi\in C(\Omega,\mathbb{Q}^{d}) is locally constant, it is easy to see that the rotation vector of a periodic measure should be rational, i.e.,

φ(σp(Ω))Rot(φ)d.\displaystyle\varphi(\mathcal{M}_{\sigma}^{p}(\Omega))\subset{\rm Rot}(\varphi)\cap\mathbb{Q}^{d}.

Hence in Theorem 1 we need to choose a rotation vector hh in Rot(φ)d{\rm Rot}(\varphi)\cap\mathbb{Q}^{d}.

Applying Theorem 1, we next investigate the property of relative maximizing measure for symbolic dynamics. Regarding (3) as a functional on C(Ω,)C(\Omega,\mathbb{R}), we can characterize a relative maximizing measure as a “tangent” measure. This characterization allows us to adapt argument in [Mor2010] for our constrained case (See §3 for more details) and we obtain the following.

Theorem 2.

Let (Ω,σ)(\Omega,\sigma) be an irreducible subshift of finite type. Let φ=(φ1,,φd)C(Ω,d)\varphi=(\varphi_{1},\ldots,\varphi_{d})\in C(\Omega,\mathbb{Q}^{d}) be a locally constant function and hint(Rot(φ))dh\in{\rm int}({\rm Rot}(\varphi))\cap\mathbb{Q}^{d}. Then for generic fC(Ω,)f\in C(\Omega,\mathbb{R}) every relative maximizing measure of ff with constraint hint(Rot(φ))dh\in{\rm int}({\rm Rot}(\varphi))\cap\mathbb{Q}^{d} has zero entropy. In particular, setting K(f):=μhφ(f)suppμK(f):=\bigcup_{\mu\in\mathcal{M}_{h}^{\varphi}(f)}{\rm supp}\mu, we have htop(K(f))=0h_{{\rm top}}(K(f))=0 for generic fC(Ω,)f\in C(\Omega,\mathbb{R}).

The remainder of this paper is organized as follows. In §2, we study the structure of rotation classes. In particular, we clarify our definitions and notations for symbolic dynamics and prove Theorem 1 in §2.2. In §3 we will illustrate that a relative maximizing measure is regarded as a tangent measure of (3) and prove Theorem 2.

2. Structure of rotation classes

2.1. Density of convex combinations of periodic measures

We first investigate the structure of rotation classes for a continuous map T:XXT:X\rightarrow X on a compact metric space XX such that Tp(X)\mathcal{M}_{T}^{p}(X) is dense in T(X)\mathcal{M}_{T}(X). As mentioned in Remark 1.3, a rotation class does not contain a periodic measure in some cases. Nevertheless, we have the density of convex combinations of periodic measures in a rotation class for every continuous constraint function and for every rotation vector in the interior of the rotation set.

For νT(X)\nu\in\mathcal{M}_{T}(X), a finite set FC(X,)F\subset C(X,\mathbb{R}) and ε>0\varepsilon>0 set

U(F,ε)(ν):={μT(X):|f𝑑μf𝑑ν|<ε,fF}.\displaystyle U_{(F,\varepsilon)}(\nu):=\left\{\mu\in\mathcal{M}_{T}(X):\left|\int fd\mu-\int fd\nu\right|<\varepsilon,f\in F\right\}.

Let

Δd\displaystyle\Delta^{d} ={(λ1,,λd)(0,1)d:i=1dλi=1},\displaystyle=\left\{(\lambda_{1},\ldots,\lambda_{d})\in(0,1)^{d}:\sum_{i=1}^{d}\lambda_{i}=1\right\},
𝒩T\displaystyle\mathcal{N}_{T} ={i=1d+1λiμi:μiTp(X),(λ1,,λd+1)Δd+1,\displaystyle=\Bigg{\{}\sum_{i=1}^{d+1}\lambda_{i}\mu_{i}:\ \mu_{i}\in\mathcal{M}_{T}^{p}(X),\quad(\lambda_{1},\ldots,\lambda_{d+1})\in\Delta^{d+1},
dimspan{rvφ(μ1)rvφ(μd+1),,rvφ(μd)rvφ(μd+1)}=d},\displaystyle\qquad\quad\dim{\rm span}\{{\rm rv}_{\varphi}(\mu_{1})-{\rm rv}_{\varphi}(\mu_{d+1}),\ldots,{\rm rv}_{\varphi}(\mu_{d})-{\rm rv}_{\varphi}(\mu_{d+1})\}=d\Bigg{\}},

where span{h1,,hd}{\rm span}\{h_{1},\ldots,h_{d}\} is the vector space spanned by {h1,,hd}\{h_{1},\ldots,h_{d}\}. We begin with the following proposition.

Proposition 2.1.

Let T:XXT:X\rightarrow X be a continuous map on a compact metric space XX such that Tp(X)\mathcal{M}_{T}^{p}(X) is dense in T(X)\mathcal{M}_{T}(X). Let φ=(φ1,,φd)C(X,d)\varphi=(\varphi_{1},\ldots,\varphi_{d})\in C(X,\mathbb{R}^{d}) and hint(Rot(φ))h\in{\rm int}({\rm Rot}(\varphi)). Then the set rvφ1(h)𝒩T{\rm rv}_{\varphi}^{-1}(h)\cap\mathcal{N}_{T} is dense in rvφ1(h){\rm rv}_{\varphi}^{-1}(h).

Proof.

Take νrvφ1(h)\nu\in{\rm rv}_{\varphi}^{-1}(h) and a open neighborhood UνU_{\nu} of ν\nu. Then there exists a finite set FC(X,)F\subset C(X,\mathbb{R}) and ε>0\varepsilon>0 such that U(F,ε)(ν)UνU_{(F,\varepsilon)}(\nu)\subset U_{\nu}.

Since hint(Rot(φ))h\in{\rm int}({\rm Rot}(\varphi)), there exists δ>0\delta>0 such that

Bdδ(h)Rot(φ).\displaystyle B_{\sqrt{d}\delta}(h)\subset{\rm Rot}(\varphi).

where Bdδ(h)B_{\sqrt{d}\delta}(h) is the open ball of radius dδ\sqrt{d}\delta centered at hh. Let hi=h+δeih^{\prime}_{i}=h+\delta e_{i} for i=1,,d+1i=1,\ldots,d+1 where {ei}i=1d\{e_{i}\}_{i=1}^{d} is the standard basis of d\mathbb{R}^{d} and ed+1=i=1deie_{d+1}=-\sum_{i=1}^{d}e_{i}. Then {hi}i=1d+1\{h^{\prime}_{i}\}_{i=1}^{d+1} forms a simplex with the barycenter hh. Let ξirvφ1(hi)\xi_{i}\in{\rm rv}_{\varphi}^{-1}(h^{\prime}_{i}).

Fix

0<t<ε4maxfFf\displaystyle 0<t<\frac{\varepsilon}{4\max_{f\in F}\|f\|_{\infty}}

and define νi:=(1t)ν+tξi\nu_{i}:=(1-t)\nu+t\xi_{i}. Then for fFf\in F we have

|f𝑑νif𝑑ν|\displaystyle\left|\int fd\nu_{i}-\int fd\nu\right| t|f𝑑ξif𝑑ν|2tf<ε2\displaystyle\leq t\left|\int fd\xi_{i}-\int fd\nu\right|\leq 2t\|f\|_{\infty}<\frac{\varepsilon}{2}

and νiU(F,ε/2)(ν)\nu_{i}\in U_{(F,\varepsilon/2)}(\nu). By the definition of νi\nu_{i} we have

rvφ(νi)=(1t)rvφ(ν)+trvφ(ξi)=(1t)h+thi=h+tδei.\displaystyle{\rm rv}_{\varphi}(\nu_{i})=(1-t){\rm rv}_{\varphi}(\nu)+t{\rm rv}_{\varphi}(\xi_{i})=(1-t)h+th^{\prime}_{i}=h+t\delta e_{i}.

Thus 1d+1i=1d+1rvφ(νi)=h\frac{1}{d+1}\sum_{i=1}^{d+1}{\rm rv}_{\varphi}(\nu_{i})=h holds.

Since Tp(X)\mathcal{M}_{T}^{p}(X) is dense in T(X)\mathcal{M}_{T}(X), for r>0r>0, there exists

μiU(F,ε/2)(νi)U({φ},r/d)(νi)Tp(X).\displaystyle\mu_{i}\in U_{(F,\varepsilon/2)}(\nu_{i})\cap U_{(\{\varphi\},r/\sqrt{d})}(\nu_{i})\cap\mathcal{M}_{T}^{p}(X).

Since rvφ(ν1)rvφ(νd+1),rvφ(ν2)rvφ(νd+1),,rvφ(νd)rvφ(νd+1){\rm rv}_{\varphi}(\nu_{1})-{\rm rv}_{\varphi}(\nu_{d+1}),{\rm rv}_{\varphi}(\nu_{2})-{\rm rv}_{\varphi}(\nu_{d+1}),\ldots,{\rm rv}_{\varphi}(\nu_{d})-{\rm rv}_{\varphi}(\nu_{d+1}) are linearly independent, by the open property of linearly independence and the continuity of the map rvφ{\rm rv}_{\varphi}, we see that rvφ(μ1)rvφ(μd+1),rvφ(μ2)rvφ(μd+1),,rvφ(μd)rvφ(μd+1){\rm rv}_{\varphi}(\mu_{1})-{\rm rv}_{\varphi}(\mu_{d+1}),{\rm rv}_{\varphi}(\mu_{2})-{\rm rv}_{\varphi}(\mu_{d+1}),\ldots,{\rm rv}_{\varphi}(\mu_{d})-{\rm rv}_{\varphi}(\mu_{d+1}) are also linearly independent for all sufficiently small r>0r>0. Moreover, we obtain

1d+1i=1d+1rvφ(μi)hd=1d+1i=1d+1rvφ(μi)1d+1i=1d+1rvφ(νi)d\displaystyle\left|\left|\frac{1}{d+1}\sum_{i=1}^{d+1}{\rm rv}_{\varphi}(\mu_{i})-h\right|\right|_{\mathbb{R}^{d}}=\left|\left|\frac{1}{d+1}\sum_{i=1}^{d+1}{\rm rv}_{\varphi}(\mu_{i})-\frac{1}{d+1}\sum_{i=1}^{d+1}{\rm rv}_{\varphi}(\nu_{i})\right|\right|_{\mathbb{R}^{d}}
1d+1i=1d+1rvφ(μi)rvφ(νi)d<d+1d+1r=r,\displaystyle\leq\frac{1}{d+1}\sum_{i=1}^{d+1}\left|\left|{\rm rv}_{\varphi}(\mu_{i})-{\rm rv}_{\varphi}(\nu_{i})\right|\right|_{\mathbb{R}^{d}}<\frac{d+1}{d+1}r=r,

where d\|\cdot\|_{\mathbb{R}^{d}} is the standard norm of d\mathbb{R}^{d}. Hence hh is contained in the open ball of radius rr centered at 1d+1i=1d+1rvφ(μi)\frac{1}{d+1}\sum_{i=1}^{d+1}{\rm rv}_{\varphi}(\mu_{i}). Taking r>0r>0 sufficiently small, we deduce that hh is an interior point of the simplex whose vertices are rvφ(μ1),,rvφ(μd+1){\rm rv}_{\varphi}(\mu_{1}),\ldots,{\rm rv}_{\varphi}(\mu_{d+1}), which implies that there exists (λ1,,λd+1)Δd+1(\lambda_{1},\ldots,\lambda_{d+1})\in\Delta^{d+1} such that h=i=1d+1λirvφ(μi)h=\sum_{i=1}^{d+1}\lambda_{i}{\rm rv}_{\varphi}(\mu_{i}).

Let ν~=i=1d+1λiμi\tilde{\nu}=\sum_{i=1}^{d+1}\lambda_{i}\mu_{i}. Trivially, rvφ(ν~)=h{\rm rv}_{\varphi}(\tilde{\nu})=h holds. Then for every fFf\in F we have

|f𝑑ν~f𝑑ν|\displaystyle\left|\int fd\tilde{\nu}-\int fd\nu\right| =|i=1d+1λif𝑑μii=1d+1λif𝑑ν|\displaystyle=\left|\sum_{i=1}^{d+1}\lambda_{i}\int fd\mu_{i}-\sum_{i=1}^{d+1}\lambda_{i}\int fd\nu\right|
i=1d+1λi|f𝑑μif𝑑ν|\displaystyle\leq\sum_{i=1}^{d+1}\lambda_{i}\left|\int fd\mu_{i}-\int fd\nu\right|
i=1d+1λi(|f𝑑μif𝑑νi|+|f𝑑νif𝑑ν|)\displaystyle\leq\sum_{i=1}^{d+1}\lambda_{i}\left(\left|\int fd\mu_{i}-\int fd\nu_{i}\right|+\left|\int fd\nu_{i}-\int fd\nu\right|\right)
<i=1d+1λi(ε2+ε2)=ε,\displaystyle<\sum_{i=1}^{d+1}\lambda_{i}\left(\frac{\varepsilon}{2}+\frac{\varepsilon}{2}\right)=\varepsilon,

i.e., ν~=i=1d+1λiμiU(F,ε)(ν)\tilde{\nu}=\sum_{i=1}^{d+1}\lambda_{i}\mu_{i}\in U_{(F,\varepsilon)}(\nu), and this is precisely the assertion of the proposition. ∎

Remark 2.2.

In the proof of Proposition 2.1, we have μiU(F,ε)(ν)\mu_{i}\in U_{(F,\varepsilon)}(\nu) for every i{1,,d+1}i\in\{1,\ldots,d+1\} since we have νiU(F,ε/2)(ν)\nu_{i}\in U_{(F,\varepsilon/2)}(\nu) and μiU(F,ε/2)(νi)\mu_{i}\in U_{(F,\varepsilon/2)}(\nu_{i}). We will use this fact in the proof of Theorem 1.

Corollary 2.3.

Assume the hypotheses of Proposition 2.1. Let HμH_{\mu} be the entropy map of TT. Then the set

𝒵={μrvφ1(h):Hμ=0}\displaystyle\mathcal{Z}=\{\mu\in{\rm rv}_{\varphi}^{-1}(h):H_{\mu}=0\}

is a residual subset of rvφ1(h){\rm rv}_{\varphi}^{-1}(h).

Proof.

The argument is similar to [DGS][Proposition 22.16, p.223]. By Proposition 2.1, 𝒩Trvφ1(h)\mathcal{N}_{T}\cap{\rm rv}_{\varphi}^{-1}(h) is dense in rvφ1(h){\rm rv}_{\varphi}^{-1}(h) and thus 𝒵\mathcal{Z} is also. Moreover, upper semi-continuity of the entropy map μHμ\mu\mapsto H_{\mu} implies for every n1n\geq 1

𝒵n:={μrvφ1(h):0Hμ<1n}𝒩Trvφ1(h)\displaystyle\mathcal{Z}_{n}:=\left\{\mu\in{\rm rv}_{\varphi}^{-1}(h):0\leq H_{\mu}<\frac{1}{n}\right\}\supset\mathcal{N}_{T}\cap{\rm rv}_{\varphi}^{-1}(h)

is nonempty, open and dense in rvφ1(h){\rm rv}_{\varphi}^{-1}(h). Hence 𝒵=n1𝒵n\mathcal{Z}=\bigcap_{n\geq 1}\mathcal{Z}_{n} is a residual set in rvφ1(h){\rm rv}_{\varphi}^{-1}(h). ∎

Lemma 2.4.

Assume the hypotheses of Proposition 2.1. Let μ1,,μd+1Tp(X)\mu_{1},\ldots,\mu_{d+1}\in\mathcal{M}_{T}^{p}(X) such that rvφ(μ1),,rvφ(μd+1)d{\rm rv}_{\varphi}(\mu_{1}),\ldots,{\rm rv}_{\varphi}(\mu_{d+1})\in\mathbb{Q}^{d} and

(4) dimspan{rvφ(μ1)rvφ(μd+1),,rvφ(μd)rvφ(μd+1)}=d.\displaystyle\dim{\rm span}\{{\rm rv}_{\varphi}(\mu_{1})-{\rm rv}_{\varphi}(\mu_{d+1}),\ldots,{\rm rv}_{\varphi}(\mu_{d})-{\rm rv}_{\varphi}(\mu_{d+1})\}=d.

Let (λ1,,λd+1)Δd+1(\lambda_{1},\ldots,\lambda_{d+1})\in\Delta^{d+1} and ν~=i=1d+1λiμiT(X)\tilde{\nu}=\sum_{i=1}^{d+1}\lambda_{i}\mu_{i}\in\mathcal{M}_{T}(X). Then λi\lambda_{i} is rational for each i=1,,d+1i=1,\ldots,d+1 if rvφ(ν~){\rm rv}_{\varphi}(\tilde{\nu}) belongs to d\mathbb{Q}^{d}.

Proof.

Since ν~=i=1d+1λiμi\tilde{\nu}=\sum_{i=1}^{d+1}\lambda_{i}\mu_{i} and i=1d+1λi=1\sum_{i=1}^{d+1}\lambda_{i}=1, we have

(5) rvφ(ν~)=rvφ(μd+1)+i=1dλi(rvφ(μi)rvφ(μd+1)).\displaystyle{\rm rv}_{\varphi}(\tilde{\nu})={\rm rv}_{\varphi}(\mu_{d+1})+\sum_{i=1}^{d}\lambda_{i}\left({\rm rv}_{\varphi}(\mu_{i})-{\rm rv}_{\varphi}(\mu_{d+1})\right).

Let VV be a d×dd\times d-matrix given by

(6) V=(rvφ(μd+1)rvφ(μ1),,rvφ(μd+1)rvφ(μd))\displaystyle V=\Big{(}{\rm rv}_{\varphi}(\mu_{d+1})-{\rm rv}_{\varphi}(\mu_{1}),\ldots,{\rm rv}_{\varphi}(\mu_{d+1})-{\rm rv}_{\varphi}(\mu_{d})\Big{)}

and λ\lambda be a column vector given by (λ)i=λi(\lambda)_{i}=\lambda_{i} for i=1,,di=1,\ldots,d. It follows from (4) that the matrix VV is invertible. Therefore, by (5), we obtain

(7) λ=V1(rvφ(μd+1)rvφ(ν~)).\displaystyle\lambda=V^{-1}\left({\rm rv}_{\varphi}(\mu_{d+1})-{\rm rv}_{\varphi}(\tilde{\nu})\right).

Since rvφ(μ1),,rvφ(μd+1){\rm rv}_{\varphi}(\mu_{1}),\ldots,{\rm rv}_{\varphi}(\mu_{d+1}) and rvφ(ν~){\rm rv}_{\varphi}(\tilde{\nu}) are in d\mathbb{Q}^{d}, each component of VV and that of its inverse V1V^{-1} are rational. Therefore, by (7), we deduce that λi\lambda_{i} is rational for each i=1,,d+1i=1,\ldots,d+1. ∎

2.2. Symbolic dynamics and density of periodic measures

We next consider symbolic dynamics. In this particular case, under some assumptions, we can prove the density of periodic measures in a given rotation class. Denote by 0\mathbb{N}_{0} the set of all non-negative integers. For a finite set 𝒜\mathcal{A} we consider the one-sided infinite product 𝒜0\mathcal{A}^{\mathbb{N}_{0}} equipped with the product topology of the discrete one.

Let σ\sigma be the shift map on 𝒜0\mathcal{A}^{\mathbb{N}_{0}} (i.e., (σ(x¯))i=xi+1(\sigma(\underline{x}))_{i}=x_{i+1} for each i0i\in{\mathbb{N}_{0}} and x¯=(xi)i0𝒜0\underline{x}=(x_{i})_{i\in{\mathbb{N}_{0}}}\in\mathcal{A}^{\mathbb{N}_{0}}). When a subset Ω\Omega of 𝒜0\mathcal{A}^{\mathbb{N}_{0}} is σ\sigma-invariant and closed, we call it a subshift. Slightly abusing the notation we denote by σ\sigma the shift map restricted on Ω\Omega.

For a subshift Ω\Omega, let [u]={x¯Ω:u=x0xn1}[u]=\left\{\underline{x}\in\Omega:u=x_{0}\cdots x_{n-1}\right\} for each u𝒜nu\in\mathcal{A}^{n}, n1n\geq 1 and set (Ω)={un1𝒜n:[u]}\mathcal{L}(\Omega)=\left\{u\in\bigcup_{n\geq 1}\mathcal{A}^{n}:[u]\neq\emptyset\right\}. We also denote n(Ω):={u(Ω):|u|=n}\mathcal{L}_{n}(\Omega):=\{u\in\mathcal{L}(\Omega):|u|=n\} for n1n\geq 1, where |u||u| denotes the length of uu, i.e., |u|=n|u|=n if u=u0un1𝒜nu=u_{0}\cdots u_{n-1}\in\mathcal{A}^{n}. A word u𝒜nu\in\mathcal{A}^{n} appears in x¯𝒜0\underline{x}\in\mathcal{A}^{\mathbb{N}_{0}} if there exists k0k\geq 0 such that xkxk+n1=u0un1x_{k}\cdots x_{k+n-1}=u_{0}\cdots u_{n-1}. For u,v(Ω)u,v\in\mathcal{L}(\Omega), we use the juxtaposition uvuv to denote the word obtained by the concatenation and uu^{\infty} means a one-sided infinite sequence uuu𝒜0uuu\cdots\in\mathcal{A}^{\mathbb{N}_{0}}. We say that Ω\Omega is irreducible if for any i,jAi,j\in A, we can find u(Ω)u\in\mathcal{L}(\Omega) such that iuj(Ω)iuj\in\mathcal{L}(\Omega) holds. A subshift Ω\Omega is a subshift of finite type (SFT) if there exists a finite set n1𝒜n\mathcal{F}\subset\bigcup_{n\geq 1}\mathcal{A}^{n} such that no word from \mathcal{F} appears in any x¯Ω\underline{x}\in\Omega. The set \mathcal{F} is called a forbidden set of Ω\Omega. Note that different forbidden sets may define the same subshift of finite type.

In order to prove Theorem 1 we shall show that on a subsfhit of finite type periodic orbits which share the same word can be concatenated without extra gap words.

Lemma 2.5.

Let (Ω,σ)(\Omega,\sigma) be a subshift of finite type with a forbidden set \mathcal{F}. Let κ=max{|u|:u}\kappa=\max\{|u|:u\in\mathcal{F}\}. Let uκ(Ω)u\in\mathcal{L}_{\kappa}(\Omega) and v,w(Ω)v,w\in\mathcal{L}(\Omega) such that (vu),(wu)Ω(vu)^{\infty},(wu)^{\infty}\in\Omega. Then for every k1k\geq 1 and sequences {mi}i=1k,{ni}i=1k\{m_{i}\}_{i=1}^{k},\{n_{i}\}_{i=1}^{k}\subset\mathbb{N}, we have

((vu)m1(wu)n1(vu)m2(vu)mk(wu)nk)Ω.\displaystyle((vu)^{m_{1}}(wu)^{n_{1}}(vu)^{m_{2}}\cdots(vu)^{m_{k}}(wu)^{n_{k}})^{\infty}\in\Omega.
Proof.

Let k1k\geq 1, {mi}i=1k\{m_{i}\}_{i=1}^{k} and {ni}i=1k\{n_{i}\}_{i=1}^{k}\subset\mathbb{N}. By the definition of κ\kappa we should only check the words in ((vu)m1(wu)n1(vu)m2(vu)mk(wu)nk)((vu)^{m_{1}}(wu)^{n_{1}}(vu)^{m_{2}}\cdots(vu)^{m_{k}}(wu)^{n_{k}})^{\infty} with length less than κ\kappa. However such a word is a subword of vu,uv,wuvu,uv,wu and uwuw. Since (vu),(wu)Ω(vu)^{\infty},(wu)^{\infty}\in\Omega, there is no forbidden word in vu,uv,wuvu,uv,wu and uwuw, which complete the proof. ∎

In addition, an irreducible sofic shift satisfies the specification property and thus σp(Ω)\mathcal{M}_{\sigma}^{p}(\Omega) is dense in σ(Ω)\mathcal{M}_{\sigma}(\Omega) (see for example [Weiss73] and [Sigmund]). Hence we have the following by Proposition 2.1.

Corollary 2.6.

Let (Ω,σ)(\Omega,\sigma) be an irreducible subshift of finite type with a forbidden set \mathcal{F}. Let φ=(φ1,,φd)C(Ω,d)\varphi=(\varphi_{1},\ldots,\varphi_{d})\in C(\Omega,\mathbb{R}^{d}) be a locally constant function and hint(Rot(φ))h\in{\rm int}({\rm Rot}(\varphi)). Then the set rvφ1(h)𝒩σ{\rm rv}_{\varphi}^{-1}(h)\cap\mathcal{N}_{\sigma} is dense in rvφ1(h){\rm rv}_{\varphi}^{-1}(h).

Using Lemma 2.5 and Corollary 2.6, we can prove Theorem 1.

Proof of Theorem 1.

Take νrvφ1(h)\nu\in{\rm rv}_{\varphi}^{-1}(h) and a open neighborhood UνU_{\nu} of ν\nu. There exists a finite set FC(Ω,)F\subset C(\Omega,\mathbb{R}) and ε>0\varepsilon>0 such that U(F,ε)(ν)UνU_{(F,\varepsilon)}(\nu)\subset U_{\nu}. Without loss of generality we may assume every fFf\in F is locally constant. Let κ=max{|u|:u}\kappa=\max\{|u|:u\in\mathcal{F}\} and u=u0uκ1κ(Ω)u=u_{0}\cdots u_{\kappa-1}\in\mathcal{L}_{\kappa}(\Omega) such that ν([u])>0\nu([u])>0. Replace FF and ε\varepsilon with F{χ[u]}F\cup\{\chi_{[u]}\} and min{ε,ν([u])/2}\min\{\varepsilon,\nu([u])/2\} respectively, where χ[u]\chi_{[u]} is the characteristic function of [u][u].

By Corollary 2.6, there is ν~U(F,ε/2)(ν)rvφ1(h)\tilde{\nu}\in U_{(F,\varepsilon/2)}(\nu)\cap{\rm rv}_{\varphi}^{-1}(h) of the form ν~=i=1d+1λiμi\tilde{\nu}=\sum_{i=1}^{d+1}\lambda_{i}\mu_{i} where μiσp(Ω)\mu_{i}\in\mathcal{M}_{\sigma}^{p}(\Omega) and (λ1,,λd+1)Δd+1(\lambda_{1},\ldots,\lambda_{d+1})\in\Delta^{d+1} with (4). Note that rvφ(μi)d(i=1,,d+1){\rm rv}_{\varphi}(\mu_{i})\in\mathbb{Q}^{d}\ (i=1,\ldots,d+1) as stated in Remark 1.3. For each i=1,,d+1i=1,\ldots,d+1, we will denote by aia_{i}^{\infty} the corresponding periodic orbits to μiσp(Ω)\mu_{i}\in\mathcal{M}_{\sigma}^{p}(\Omega) where aia_{i} is the word with lengths of periods. Moreover, by Lemma 2.4, each λi\lambda_{i} can be written as λi=qi/Q\lambda_{i}=q_{i}/Q where q1,,qd+1,Qq_{1},\ldots,q_{d+1},Q\in\mathbb{N} with q1++qd+1=Qq_{1}+\ldots+q_{d+1}=Q since (λ1,,λd+1)Δd+1d+1(\lambda_{1},\ldots,\lambda_{d+1})\in\Delta^{d+1}\cap\mathbb{Q}^{d+1}.

Let \ell be the maximum length of the words on which elements of F{φ1,,φd}F\cup\{\varphi_{1},\ldots,\varphi_{d}\} depend. We can assume

<min{|a1|,,|ad+1|}\displaystyle\ell<\min\{|a_{1}|,\ldots,|a_{d+1}|\}

by replacing ai(i=1,,d+1)a_{i}\ (i=1,\ldots,d+1) with concatenations aika_{i}^{k} for some kk\in\mathbb{N} if necessary. As stated in Remark 2.2, we have μiU(F,ε)(ν)\mu_{i}\in U_{(F,\varepsilon)}(\nu) for every i=1,,d+1i=1,\ldots,d+1, which implies μi([u])>ν([u])εν([u])/2>0\mu_{i}([u])>\nu([u])-\varepsilon\geq\nu([u])/2>0. Hence uu is a subword of each aia_{i} and without loss of generality we may assume ai,0ai,|u|1=ua_{i,0}\cdots a_{i,|u|-1}=u. For gF{φ}g\in F\cup\{\varphi\} define δg,k(k=1,,d+1)\delta_{g,k}\ (k=1,\ldots,d+1) by

δg,k=i=01{\displaystyle\delta_{g,k}=\sum_{i=0}^{\ell-1}\Big{\{} g(σ|ak|+i(akak+1))g(σ|ak|+i(akak))}\displaystyle g(\sigma^{|a_{k}|-\ell+i}(a_{k}a_{k+1}))-g(\sigma^{|a_{k}|-\ell+i}(a_{k}a_{k}))\Big{\}}

where ad+2=a1a_{d+2}=a_{1}. Set δg=δg,1++δg,d+1\delta_{g}=\delta_{g,1}+\ldots+\delta_{g,d+1}.

Let VV be a d×dd\times d-matrix given by (6). As stated in the proof of Lemma 2.4, the matrix VV is invertible by (4). Since each component of VV and δφ\delta_{\varphi} is rational, we denote V1δφ=v/RV^{-1}\delta_{\varphi}=v/R where vdv\in\mathbb{Z}^{d} and RR\in\mathbb{N}. Let vi=(v)i(i=1,,d)v_{i}=(v)_{i}\ (i=1,\ldots,d).

Now we construct a periodic measure near ν\nu with the rotation vector hh. Since a1,,ad+1a_{1},\ldots,a_{d+1} share the same word uu, by Lemma 2.5, we can concatenate them without extra gap words. Hence let

A\displaystyle A =|a1||ad+1|,C=maxfF,i=1,,d+1|f𝑑μi|,δ=maxfF|δf|,\displaystyle=|a_{1}|\ldots|a_{d+1}|,\quad C=\max_{f\in F,\ i=1,\ldots,d+1}\left|\int f\ d\mu_{i}\right|,\quad\delta^{*}=\max_{f\in F}|\delta_{f}|,
mi\displaystyle m_{i}^{\prime} =Avi/|ai|(i=1,,d),md+1=A(i=1dvi)/|ad+1|,\displaystyle=Av_{i}/|a_{i}|\ (i=1,\ldots,d),\quad m_{d+1}^{\prime}=-A(\sum_{i=1}^{d}v_{i})/|a_{d+1}|,
Mj\displaystyle M_{j}^{\prime} =Aqj/|aj|,Mj=tARMj+mj(j=1,,d+1),\displaystyle=Aq_{j}/|a_{j}|,\quad M_{j}=tARM_{j}^{\prime}+m_{j}^{\prime}\ (j=1,\ldots,d+1),
y\displaystyle y =a1tM1+m1a2tM2+m2ad+1tMd+1+md+1,\displaystyle=a_{1}^{tM_{1}^{\prime}+m_{1}^{\prime}}a_{2}^{tM_{2}^{\prime}+m_{2}^{\prime}}\cdots a_{d+1}^{tM_{d+1}^{\prime}+m_{d+1}^{\prime}},
z\displaystyle z =a1tM1a2tM2ad+1tMd+1,x=yz(AR1),\displaystyle=a_{1}^{tM_{1}^{\prime}}a_{2}^{tM_{2}^{\prime}}\cdots a_{d+1}^{tM_{d+1}^{\prime}},\quad x=yz^{(AR-1)},

where tt\in\mathbb{N} is large enough to satisfy

(8) AR|x|δ+Ci=1d+1|Mi|ai||x|λi|<ε2\displaystyle\frac{AR}{|x|}\delta^{*}+C\sum_{i=1}^{d+1}\left|\frac{M_{i}|a_{i}|}{|x|}-\lambda_{i}\right|<\frac{\varepsilon}{2}

and tMd+1A(i=1dvi)/|ad+1|>0tM_{d+1}^{\prime}-A(\sum_{i=1}^{d}v_{i})/|a_{d+1}|>0. Note that such tt\in\mathbb{N} exists since

|x|\displaystyle|x| =i=1d+1Mi|ai|=tARi=1d+1Mi|ai|,\displaystyle=\sum_{i=1}^{d+1}M_{i}|a_{i}|=tAR\sum_{i=1}^{d+1}M_{i}^{\prime}|a_{i}|,
λi\displaystyle\lambda_{i} =qij=1d+1qj=Mi|ai|j=1d+1Mj|aj|\displaystyle=\frac{q_{i}}{\sum_{j=1}^{d+1}q_{j}}=\frac{M_{i}^{\prime}|a_{i}|}{\sum_{j=1}^{d+1}M_{j}^{\prime}|a_{j}|}

hold and the left hand side of (8) tends to 0 as t+t\to+\infty. Let μ\mu be the periodic measure supported on xx^{\infty}.

First we check rvφ(μ)=h(=rvφ(ν~)){\rm rv}_{\varphi}(\mu)=h(={\rm rv}_{\varphi}(\tilde{\nu})). Since φ\varphi is locally constant,

rvφ(μ)\displaystyle{\rm rv}_{\varphi}(\mu) =1|x|S|x|φ(x)\displaystyle=\frac{1}{|x|}S_{|x|}\varphi(x)
=1|x|(ARδφ+i=1d+1Mi|ai|φ𝑑μi)\displaystyle=\frac{1}{|x|}\left(AR\delta_{\varphi}+\sum_{i=1}^{d+1}M_{i}|a_{i}|\int\varphi d\mu_{i}\right)
=1|x|ARδφ+1|x|i=1d+1mi|ai|rvφ(μi)+i=1d+1λirvφ(μi)\displaystyle=\frac{1}{|x|}AR\delta_{\varphi}+\frac{1}{|x|}\sum_{i=1}^{d+1}m_{i}^{\prime}|a_{i}|{\rm rv}_{\varphi}(\mu_{i})+\sum_{i=1}^{d+1}\lambda_{i}{\rm rv}_{\varphi}(\mu_{i})
=h+A|x|Vv+1|x|i=1d+1mi|ai|rvφ(μi)\displaystyle=h+\frac{A}{|x|}Vv+\frac{1}{|x|}\sum_{i=1}^{d+1}m_{i}^{\prime}|a_{i}|{\rm rv}_{\varphi}(\mu_{i})
=h+A|x|{Vv+i=1dvirvφ(μi)(j=1dvj)rvφ(μd+1)}=h.\displaystyle=h+\frac{A}{|x|}\left\{Vv+\sum_{i=1}^{d}v_{i}{\rm rv}_{\varphi}(\mu_{i})-(\sum_{j=1}^{d}v_{j}){\rm rv}_{\varphi}(\mu_{d+1})\right\}=h.

Next we check μU(F,ε)(ν)\mu\in U_{(F,\varepsilon)}(\nu). Let fFf\in F. We compute

|1|x|S|x|ff𝑑ν|\displaystyle\left|\frac{1}{|x|}S_{|x|}f-\int f\ d\nu\right|
|1|x|S|x|ff𝑑ν~|+|f𝑑ν~f𝑑ν|\displaystyle\leq\left|\frac{1}{|x|}S_{|x|}f-\int f\ d\tilde{\nu}\right|+\left|\int f\ d\tilde{\nu}-\int f\ d\nu\right|
=|1|x|(ARδf+i=1d+1Mi|ai|f𝑑μi)j=1d+1λif𝑑μi|+ε2\displaystyle=\left|\frac{1}{|x|}\Big{(}AR\delta_{f}+\sum_{i=1}^{d+1}M_{i}|a_{i}|\int fd\mu_{i}\Big{)}-\sum_{j=1}^{d+1}\lambda_{i}\int fd\mu_{i}\right|+\frac{\varepsilon}{2}
i=1d+1|(Mi|ai||x|λi)f𝑑μi|+AR|x||δf|+ε2\displaystyle\leq\sum_{i=1}^{d+1}\left|\left(\frac{M_{i}|a_{i}|}{|x|}-\lambda_{i}\right)\int fd\mu_{i}\right|+\frac{AR}{|x|}|\delta_{f}|+\frac{\varepsilon}{2}
<ε2+ε2=ε,\displaystyle<\frac{\varepsilon}{2}+\frac{\varepsilon}{2}=\varepsilon,

which completes the proof. ∎

Remark 2.7.

In the proof of Theorem 1, we think the word yy as a corrective one to attain the desired rotation vector. A similar approach is used in Theorem 5 of [Jen01rotation] in a different setting but our construction is more explicit than it.

Remark 2.8.

Note that the error term δφ\delta_{\varphi} does not depend on mi(i=1,,d+1)m^{\prime}_{i}\ (i=1,\ldots,d+1) in our case. For a subshift with the specification condition, we can concatenate the words a1m1,,ad+1md+1a_{1}^{m^{\prime}_{1}},\ldots,a_{d+1}^{m^{\prime}_{d+1}} with some gap words but the error term in such case depends on mi(i=1,,d+1)m^{\prime}_{i}\ (i=1,\ldots,d+1), which implies we cannot choose a suitable corrective word yy for the error term δφ\delta_{\varphi} in such case. So we have to overcome this difficulty to extend Theorem 1 to the case of a subshift with the specification condition.

Remark 2.9.

For a rotation vector in the boundary of a rotation set, there may exist no periodic measure in the rotation class. Let Ω{1,2,3}0\Omega\subset\{1,2,3\}^{\mathbb{N}_{0}} be a Markov shift with an adjacency matrix

A=(110111011)\displaystyle A=\begin{pmatrix}1&1&0\\ 1&1&1\\ 0&1&1\end{pmatrix}

(i.e., Ω={x¯{1,2,3}0:Axixi+1=1for alli0}\Omega=\{\underline{x}\in\{1,2,3\}^{\mathbb{N}_{0}}:A_{x_{i}x_{i+1}}=1\ \mbox{for all}\ i\in\mathbb{N}_{0}\}) and define φ=(φ1,φ2,φ3):Ω3\varphi=(\varphi_{1},\varphi_{2},\varphi_{3}):\Omega\rightarrow\mathbb{Q}^{3} by

φi(x¯)={1x0=i0else.\displaystyle\varphi_{i}(\underline{x})=\left\{\begin{array}[]{cc}1&x_{0}=i\\ 0&\mbox{else.}\end{array}\right.

Then its rotation set Rot(φ){\rm Rot}(\varphi) is the polyhedron whose extremal points are e1,e2e_{1},e_{2} and e3e_{3} , where {e1,e2,e3}\{e_{1},e_{2},e_{3}\} is the standard basis of 3\mathbb{R}^{3}. Take a rotation vector hh from the open side whose vertices are e1e_{1} and e3e_{3}, i.e., h{te1+(1t)e3:t(0,1)}h\in\{te_{1}+(1-t)e_{3}:t\in(0,1)\}. If there exists a periodic measure μrvφ1(h)\mu\in{\rm rv}_{\varphi}^{-1}(h), the corresponding periodic orbit uu^{\infty} should contain both of 11 and 33. Since there is no sequence including 1313 and 3131 in Ω\Omega, the word uu must contain the symbol 22. Hence we have rvφ2(μ)>0{\rm rv}_{\varphi_{2}}(\mu)>0 and h=rvφ(μ){te1+(1t)e3:t(0,1)}h={\rm rv}_{\varphi}(\mu)\notin\{te_{1}+(1-t)e_{3}:t\in(0,1)\}, which is a contradiction.

3. Generic property for constraint ergodic optimization

In this section, we prove Theorem 2. Our proof is based on the approach presented by Morris [Mor2010] for the unconstrained case but we need to pay careful attention to the constraint. Moreover, density of periodic measures in the rotation class (i.e., Theorem 1) plays an important role to obtain the argument.

3.1. Characterization by tangency

We turn to a general dynamical system in this subsection. Let T:XXT:X\rightarrow X be a continuous map on a compact metric space. Denote by Te(X)\mathcal{M}_{T}^{e}(X) the set of all ergodic measures on XX. By the Riesz representation theorem a Borel probability measure on XX can be regarded as a bounded linear functional on C(X,)C(X,\mathbb{R}). Hence we use the operator norm μ=sup{|μ(f)|:fC(X,)withf=1}\|\mu\|=\sup\{|\mu(f)|:f\in C(X,\mathbb{R})\ \mbox{with}\ \|f\|_{\infty}=1\} for an invariant measure μT(X)\mu\in\mathcal{M}_{T}(X).

First we characterize a relative maximizing measures by tangency to (3). Let φC(X,d)\varphi\in C(X,\mathbb{R}^{d}) and hRot(φ)h\in{\rm Rot}(\varphi). Note that here we do not need to assume that hint(Rot(φ))dh\in{\rm int}({\rm Rot}(\varphi))\cap\mathbb{Q}^{d}.

Lemma 3.1.

μhφ(f)\mu\in\mathcal{M}^{\varphi}_{h}(f) iff μ\mu is tangent to βhφ\beta^{\varphi}_{h} at ff.

Proof.

Let μhφ(f)\mu\in\mathcal{M}^{\varphi}_{h}(f). Then for every gC(X,)g\in C(X,\mathbb{R}) we have

βhφ(f+g)βhφ(f)f+gdμf𝑑μg𝑑μ.\beta^{\varphi}_{h}(f+g)-\beta^{\varphi}_{h}(f)\geq\int f+g\ d\mu-\int f\ d\mu\geq\int g\ d\mu.

Let μ\mu be tangent to βhφ\beta^{\varphi}_{h} at ff. For every gC(X,)g\in C(X,\mathbb{R}) we have

g𝑑μ\displaystyle\int g\ d\mu βhφ(f+g)βhφ(f)\displaystyle\leq\beta^{\varphi}_{h}(f+g)-\beta^{\varphi}_{h}(f)
=βhφ(g)+βhφ(f+g)βhφ(f)βhφ(g)\displaystyle=\beta^{\varphi}_{h}(g)+\beta^{\varphi}_{h}(f+g)-\beta^{\varphi}_{h}(f)-\beta^{\varphi}_{h}(g)
βhφ(g).\displaystyle\leq\beta^{\varphi}_{h}(g).

Then we can show that μ\mu is an invariant probability measure in the same way as unconstrained case (See for example [Shi2018, Bre08]). We now see that μ\mu takes the rotation vector hh. For i=1,,di=1,\ldots,d we have

φi𝑑μβhφ(φi)=hi,φi𝑑μβhφ(φi)=hi,\displaystyle\int\varphi_{i}\ d\mu\leq\beta^{\varphi}_{h}(\varphi_{i})=h_{i},\quad-\int\varphi_{i}\ d\mu\leq\beta^{\varphi}_{h}(-\varphi_{i})=-h_{i},

which yields rvφ(μ)=h{\rm rv}_{\varphi}(\mu)=h.

Finally we check f𝑑μ=βhφ(f)\int f\ d\mu=\beta^{\varphi}_{h}(f). Indeed,

fdμβhφ(ff)βhφ(f)=βhφ(f).\displaystyle\int-f\ d\mu\leq\beta^{\varphi}_{h}(f-f)-\beta^{\varphi}_{h}(f)=-\beta_{h}^{\varphi}(f).

Multiplying 1-1, we have

f𝑑μβhφ(f).\displaystyle\int fd\mu\geq\beta^{\varphi}_{h}(f).

Now Lemma 3.1 allows us to adopt the techniques for unconstrained ergodic optimization. While the rest of this subsection is similar to [Mor2010], attention should be paid to the constraint and thus we give proofs of Lemmas 3.2, 3.5, 3.6 below.

Lemma 3.2.

Let fC(X,)f\in C(X,\mathbb{R}) and ε>0\varepsilon>0. If νrvφ1(h)Te(X)\nu\in{\rm rv}_{\varphi}^{-1}(h)\cap\mathcal{M}_{T}^{e}(X) such that βhφ(f)f𝑑ν<ε\beta^{\varphi}_{h}(f)-\int f\ d\nu<\varepsilon, then there exists gC(X,)g\in C(X,\mathbb{R}) such that fg<ε\|f-g\|_{\infty}<\varepsilon and νhφ(g)\nu\in\mathcal{M}^{\varphi}_{h}(g).

Proof.

Applying Bishop-Phelps’s theorem [Israel, Theorem V.1.1.] to ff, ν\nu and ε=1\varepsilon^{\prime}=1 we have gC(X,)g\in C(X,\mathbb{R}), ηrvφ1(h)\eta\in{\rm rv}_{\varphi}^{-1}(h) such that η\eta is tangent to βhφ\beta^{\varphi}_{h} at gg, νη<ε=1\|\nu-\eta\|<\varepsilon^{\prime}=1 and

fg<1ε(βhφ(f)f𝑑ν)<ε.\|f-g\|_{\infty}<\frac{1}{\varepsilon^{\prime}}\left(\beta^{\varphi}_{h}(f)-\int fd\nu\right)<\varepsilon.

By Lemma 3.1 we have ηhφ(g)\eta\in\mathcal{M}^{\varphi}_{h}(g). If η=ν\eta=\nu, the proof is complete. Let ην\eta\neq\nu. We can conclude ν\nu and η\eta are not mutually singular by νη<1\|\nu-\eta\|<1 in the same way as in [Mor2010][Lemma 2.2]. Hence by the Lebesgue decomposition theorem, there exist ν^,η^\hat{\nu},\hat{\eta}\in\mathcal{M} and λ(0,1)\lambda\in(0,1) such that

η=(1λ)η^+λν^\eta=(1-\lambda)\hat{\eta}+\lambda\hat{\nu}

where η^ν\hat{\eta}\perp\nu and ν^ν\hat{\nu}\ll\nu. By a standard argument using the Radon-Nikodym theorem, it is easy to see ν^=ν\hat{\nu}=\nu (See for example [Walters]). Then for each i=1,,di=1,\ldots,d

hi=φi𝑑η\displaystyle h_{i}=\int\varphi_{i}\ d\eta =(1λ)φi𝑑η^+λφi𝑑ν\displaystyle=(1-\lambda)\int\varphi_{i}\ d\hat{\eta}+\lambda\int\varphi_{i}\ d\nu
=(1λ)φi𝑑η^+λhi.\displaystyle=(1-\lambda)\int\varphi_{i}\ d\hat{\eta}+\lambda h_{i}.

Hence we have

(1λ)(φi𝑑η^hi)=0(i=1,,m),\displaystyle(1-\lambda)\left(\int\varphi_{i}\ d\hat{\eta}-h_{i}\right)=0\quad(i=1,\ldots,m),

which yields η^rvφ1(h)\hat{\eta}\in{\rm rv}_{\varphi}^{-1}(h). Since η^,νrvφ1(h)\hat{\eta},\nu\in{\rm rv}_{\varphi}^{-1}(h), we have

g𝑑η^βhφ(g)andg𝑑νβhφ(g).\int g\ d\hat{\eta}\leq\beta^{\varphi}_{h}(g)\quad\mbox{and}\quad\int g\ d\nu\leq\beta^{\varphi}_{h}(g).

They should be equality since g𝑑η=βhφ(g)\int g\ d\eta=\beta^{\varphi}_{h}(g). Therefore, we obtain νhφ(g)\nu\in\mathcal{M}^{\varphi}_{h}(g). ∎

Set =rvφ1(h)Te(X)\mathcal{E}={\rm rv}_{\varphi}^{-1}(h)\cap\mathcal{M}^{e}_{T}(X). In the reminder of this subsection we assume

(9) ¯=rvφ1(h)\displaystyle\overline{\mathcal{E}}={\rm rv}_{\varphi}^{-1}(h)

Since the constraint requires no further discussion in the next two lemmas, we omit the proofs.

Lemma 3.3.

Let 𝒰rvφ1(h)\mathcal{U}\subset{\rm rv}_{\varphi}^{-1}(h) be an open set. Then

U:={fC(X,):hφ(f)𝒰}U:=\{f\in C(X,\mathbb{R}):\mathcal{M}^{\varphi}_{h}(f)\subset\mathcal{U}\}

is open in C(X,)C(X,\mathbb{R}).

Lemma 3.4.

Let UC(X,)U\subset C(X,\mathbb{R}) be an open set. Then

𝒰:=fUhφ(f)\mathcal{U}:=\mathcal{E}\cap\bigcup_{f\in U}\mathcal{M}^{\varphi}_{h}(f)

is open in \mathcal{E}.

At the end of this subsection, we give the following two lemmas.

Lemma 3.5.

Let 𝒰\mathcal{U} be a dense subset of \mathcal{E}. Then

U:={fC(X,):hφ(f)={μ}for someμ𝒰}U:=\{f\in C(X,\mathbb{R}):\mathcal{M}^{\varphi}_{h}(f)=\{\mu\}\ \mbox{for some}\ \mu\in\mathcal{U}\}

is dense in C(X,)C(X,\mathbb{R}).

Proof.

Let VC(X,)V\subset C(X,\mathbb{R}) be an open set. Set

𝒱=fVhφ(f).\mathcal{V}=\mathcal{E}\cap\bigcup_{f\in V}\mathcal{M}^{\varphi}_{h}(f).

By Lemma 3.4, 𝒱\mathcal{V} is open in \mathcal{E}. Since 𝒰\mathcal{U} is dense in \mathcal{E}, 𝒰𝒱\mathcal{U}\cap\mathcal{V}\neq\emptyset. Take μ𝒰𝒱\mu\in\mathcal{U}\cap\mathcal{V} and let fVf\in V such that μhφ(f)\mu\in\mathcal{M}^{\varphi}_{h}(f). Note that μrvφ1(h)\mu\in{\rm rv}_{\varphi}^{-1}(h). Since μTe(X)\mu\in\mathcal{E}\subset\mathcal{M}_{T}^{e}(X), by Jenkinson’s theorem [Jen06][Theorem 1], there exists gC(X,)g\in C(X,\mathbb{R}) such that max(g)={μ}\mathcal{M}_{\rm max}(g)=\{\mu\}. This implies hφ(g)={μ}\mathcal{M}^{\varphi}_{h}(g)=\{\mu\} since μrvφ1(h)\mu\in{\rm rv}_{\varphi}^{-1}(h). Hence for δ>0\delta>0 we have hφ(f+δg)={μ}\mathcal{M}^{\varphi}_{h}(f+\delta g)=\{\mu\} and f+δgUf+\delta g\in U. For sufficiently small δ\delta we have f+δgVf+\delta g\in V, which complete the proof. ∎

Lemma 3.6.

Let UC(X,)U\subset C(X,\mathbb{R}) be a dense subset. Then

𝒰:=fUhφ(f)\mathcal{U}:=\mathcal{E}\cap\bigcup_{f\in U}\mathcal{M}^{\varphi}_{h}(f)

is dense in \mathcal{E}.

Proof.

Take a nonempty open subset 𝒱¯\mathcal{V}\subset\overline{\mathcal{E}}. We show 𝒱𝒰\mathcal{V}\cap\mathcal{U}\neq\emptyset.

By Lemma 3.3,

V:={fC(X,):hφ(f)𝒱}V:=\{f\in C(X,\mathbb{R}):\mathcal{M}^{\varphi}_{h}(f)\subset\mathcal{V}\}

is open in C(X,)C(X,\mathbb{R}). Since 𝒱\mathcal{V} is nonempty, there exists μ𝒱rvφ1(h)Te(X)\mu\in\mathcal{V}\cap\mathcal{E}\subset{\rm rv}_{\varphi}^{-1}(h)\cap\mathcal{M}_{T}^{e}(X). Since μ\mu is ergodic, as stated in Lemma 3.5, there exists gC(X,)g\in C(X,\mathbb{R}) such that max(g)={μ}\mathcal{M}_{\rm max}(g)=\{\mu\} and we have hφ(g)={μ}𝒱\mathcal{M}^{\varphi}_{h}(g)=\{\mu\}\subset\mathcal{V} by Jenkinson’s Theorem [Jen06][Theorem 1]. Hence VV is nonempty. Since UU is dense in C(X,)C(X,\mathbb{R}), UVU\cap V\neq\emptyset and 𝒰𝒱\mathcal{U}\cap\mathcal{V}\neq\emptyset. ∎

3.2. Generic zero entropy for a symbolic dynamics

In this subsection, we apply the lemmas in the previous subsection to the symbolic case and give the proof of Theorem 2.

Proof of Theorem 2.

Suppose the hypotheses of Theorem 2 hold. As in Corollary 2.3, for each n1n\geq 1

𝒵n:={μrvφ1(h):0Hμ<1n}𝒩σrvφ1(h)\displaystyle\mathcal{Z}_{n}:=\left\{\mu\in{\rm rv}_{\varphi}^{-1}(h):0\leq H_{\mu}<\frac{1}{n}\right\}\supset\mathcal{N}_{\sigma}\cap{\rm rv}_{\varphi}^{-1}(h)

is nonempty, open and dense subset in rvφ1(h){\rm rv}_{\varphi}^{-1}(h). Moreover, by Theorem 1, we have

rvφ1(h)σp(Ω)¯=rvφ1(h),\displaystyle\overline{{\rm rv}_{\varphi}^{-1}(h)\cap\mathcal{M}_{\sigma}^{p}(\Omega)}={\rm rv}_{\varphi}^{-1}(h),

which implies (9).

Fix n1n\geq 1. Since (9) holds, we can apply Lemmas 3.3 and 3.5 to our symbolic case. Therefore, we see that

Un:={fC(Ω,):hφ(f)𝒵n}\displaystyle U_{n}:=\{f\in C(\Omega,\mathbb{R}):\mathcal{M}^{\varphi}_{h}(f)\subset\mathcal{Z}_{n}\}

is open in C(Ω,)C(\Omega,\mathbb{R}) and

U^n:={fC(Ω,):hφ(f)={μ}for someμ𝒵n}Un\displaystyle\widehat{U}_{n}:=\{f\in C(\Omega,\mathbb{R}):\mathcal{M}_{h}^{\varphi}(f)=\{\mu\}\ \mbox{for some}\ \mu\in\mathcal{Z}_{n}\}\subset U_{n}

is dense in C(Ω,)C(\Omega,\mathbb{R}). Hence UnU_{n} is an open dense subset of C(Ω,)C(\Omega,\mathbb{R}).

Then the set

R:=n1Un={fC(Ω,):hφ(f)n1𝒵n}\displaystyle R:=\bigcap_{n\geq 1}U_{n}=\left\{f\in C(\Omega,\mathbb{R}):\mathcal{M}_{h}^{\varphi}(f)\subset\bigcap_{n\geq 1}\mathcal{Z}_{n}\right\}

is a residual subset of C(Ω,)C(\Omega,\mathbb{R}), and the proof is complete. ∎

Appendix A On positive entropy

In this appendix, we see that for generic continuous function every relative maximizing measure has positive entropy under some constraints, which is a trivial consequence of ergodic optimization.

Proposition A.1.

Let T:XXT:X\rightarrow X be a continuous map on a compact metric space XX with an ergodic invariant probability measure μ\mu having positive entropy. Then there exist φC(X,)\varphi\in C(X,\mathbb{R}) and hRot(φ)h\in{\rm Rot}(\varphi) such that for generic fC(X,)f\in C(X,\mathbb{R}) every relative maximizing measure of ff with the constraint hRot(φ)h\in{\rm Rot}(\varphi) has positive entropy.

Proof.

By Jenkinson’s theorem [Jen06][Theorem 1], there exists φC(X,)\varphi\in C(X,\mathbb{R}) such that max(φ)={μ}\mathcal{M}_{\rm max}(\varphi)=\{\mu\}. Let h:=rvφ(μ)h:={\rm rv}_{\varphi}(\mu). Then hφ(f)rvφ1(h)={μ}\mathcal{M}^{\varphi}_{h}(f)\subset{\rm rv}_{\varphi}^{-1}(h)=\{\mu\} for all fC(X,)f\in C(X,\mathbb{R}). Since hφ(f)\mathcal{M}^{\varphi}_{h}(f) is not empty, we have hφ(f)={μ}\mathcal{M}^{\varphi}_{h}(f)=\{\mu\}, which implies the unique relative maximizing measure of ff with constraint hRot(φ)h\in{\rm Rot}(\varphi) has positive entropy. ∎

Appendix B Prevalent uniqueness

In this paper we focus on generic property of continuous functions in constrained setting. On the other hand, measure-theoretic “typicality” which is known as prevalence is also important. A subset 𝒫\mathcal{P} of C(X,)C(X,\mathbb{R}) is prevalent if there exists a compactly supported Borel probability measure mm on C(X,)C(X,\mathbb{R}) such that the set f+𝒫f+\mathcal{P} has full mm-measure for every fC(X,)f\in C(X,\mathbb{R}).

Prevalent uniqueness of maximizing measures for continuous functions in the literature on (unconstrained) ergodic optimization is recently established by Morris [Mor21, Theorem 1]. Moreover the result is generalized to more abstract statement, which yields prevalent uniqueness in constrained setting.

Corollary B.1.

Let T:XXT:X\rightarrow X be a continuous map on a compact metric space XX. For a continuous constraint φ:Xd\varphi:X\rightarrow\mathbb{R}^{d} and a rotation vector hRot(φ)h\in{\rm Rot}(\varphi), the set

{fC(X,):hφ(f)is a singleton}\displaystyle\{f\in C(X,\mathbb{R}):\mathcal{M}^{\varphi}_{h}(f)\ \mbox{is a singleton}\}

is prevalent.

Proof.

Since rvφ1(h)T(X){\rm rv}_{\varphi}^{-1}(h)\subset\mathcal{M}_{T}(X) is a nonempty compact set in C(X,)C(X,\mathbb{R})^{*}, the statement follows from Theorem 2 in [Mor21]. ∎

Acknowledgement.  The first author was partially supported by JSPS KAKENHI Grant Number 22H01138 and the second 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