Constrained ergodic optimization for generic continuous functions
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 every invariant probability measure that maximizes the space average of 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, 37A991. Introduction
Let be a continuous map on a compact metric space and be the space of Borel probability measures endowed with the weak*-topology. Let be the space of real-valued continuous functions on with the supremum norm . For each we consider the maximum ergodic average
(1) |
and the set of all maximizing measures of
(2) |
The functional and the set 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 -valued function playing the role of a constraint, we define the rotation set of by
where . For , the fiber is called the rotation class of . A point 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 with constraint by
(3) |
and the set of all relative maximizing measures of with a rotation vector by
Note that this formulation is a generalization of (unconstrained) ergodic optimization, since for a constant constraint and its unique rotation vector the rotation set of is .
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 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 be an irreducible subshift of finite type with finite alphabets (See §2.2 below). A function is said to be locally constant if for each there exists such that
Denote by the set of invariant measures supported on a single periodic orbit.
Theorem 1.
Let be an irreducible subshift of finite type. Let be a locally constant function and . Then the set is dense in .
Remark 1.1.
Theorem 1 is motivated by Theorem 10 in [GarLop07]: It was shown that for a Walters potential on a subshift of finite type , there exists a periodic measure whose action approximates for with if is ergodic and the locally constant constraint is joint recurrent in relation to (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 is locally constant, it is easy to see that the rotation vector of a periodic measure should be rational, i.e.,
Hence in Theorem 1 we need to choose a rotation vector in .
Applying Theorem 1, we next investigate the property of relative maximizing measure for symbolic dynamics. Regarding (3) as a functional on , 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 be an irreducible subshift of finite type. Let be a locally constant function and . Then for generic every relative maximizing measure of with constraint has zero entropy. In particular, setting , we have for generic .
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 on a compact metric space such that is dense in . 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 , a finite set and set
Let
where is the vector space spanned by . We begin with the following proposition.
Proposition 2.1.
Let be a continuous map on a compact metric space such that is dense in . Let and . Then the set is dense in .
Proof.
Take and a open neighborhood of . Then there exists a finite set and such that .
Since , there exists such that
where is the open ball of radius centered at . Let for where is the standard basis of and . Then forms a simplex with the barycenter . Let .
Fix
and define . Then for we have
and . By the definition of we have
Thus holds.
Since is dense in , for , there exists
Since are linearly independent, by the open property of linearly independence and the continuity of the map , we see that are also linearly independent for all sufficiently small . Moreover, we obtain
where is the standard norm of . Hence is contained in the open ball of radius centered at . Taking sufficiently small, we deduce that is an interior point of the simplex whose vertices are , which implies that there exists such that .
Let . Trivially, holds. Then for every we have
i.e., , and this is precisely the assertion of the proposition. ∎
Remark 2.2.
Corollary 2.3.
Assume the hypotheses of Proposition 2.1. Let be the entropy map of . Then the set
is a residual subset of .
Proof.
The argument is similar to [DGS][Proposition 22.16, p.223]. By Proposition 2.1, is dense in and thus is also. Moreover, upper semi-continuity of the entropy map implies for every
is nonempty, open and dense in . Hence is a residual set in . ∎
Lemma 2.4.
Assume the hypotheses of Proposition 2.1. Let such that and
(4) |
Let and . Then is rational for each if belongs to .
Proof.
Since and , we have
(5) |
Let be a -matrix given by
(6) |
and be a column vector given by for . It follows from (4) that the matrix is invertible. Therefore, by (5), we obtain
(7) |
Since and are in , each component of and that of its inverse are rational. Therefore, by (7), we deduce that is rational for each . ∎
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 the set of all non-negative integers. For a finite set we consider the one-sided infinite product equipped with the product topology of the discrete one.
Let be the shift map on (i.e., for each and ). When a subset of is -invariant and closed, we call it a subshift. Slightly abusing the notation we denote by the shift map restricted on .
For a subshift , let for each , and set . We also denote for , where denotes the length of , i.e., if . A word appears in if there exists such that . For , we use the juxtaposition to denote the word obtained by the concatenation and means a one-sided infinite sequence . We say that is irreducible if for any , we can find such that holds. A subshift is a subshift of finite type (SFT) if there exists a finite set such that no word from appears in any . The set is called a forbidden set of . 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 be a subshift of finite type with a forbidden set . Let . Let and such that . Then for every and sequences , we have
Proof.
Let , and . By the definition of we should only check the words in with length less than . However such a word is a subword of and . Since , there is no forbidden word in and , which complete the proof. ∎
In addition, an irreducible sofic shift satisfies the specification property and thus is dense in (see for example [Weiss73] and [Sigmund]). Hence we have the following by Proposition 2.1.
Corollary 2.6.
Let be an irreducible subshift of finite type with a forbidden set . Let be a locally constant function and . Then the set is dense in .
Proof of Theorem 1.
Take and a open neighborhood of . There exists a finite set and such that . Without loss of generality we may assume every is locally constant. Let and such that . Replace and with and respectively, where is the characteristic function of .
By Corollary 2.6, there is of the form where and with (4). Note that as stated in Remark 1.3. For each , we will denote by the corresponding periodic orbits to where is the word with lengths of periods. Moreover, by Lemma 2.4, each can be written as where with since .
Let be the maximum length of the words on which elements of depend. We can assume
by replacing with concatenations for some if necessary. As stated in Remark 2.2, we have for every , which implies . Hence is a subword of each and without loss of generality we may assume . For define by
where . Set .
Let be a -matrix given by (6). As stated in the proof of Lemma 2.4, the matrix is invertible by (4). Since each component of and is rational, we denote where and . Let .
Now we construct a periodic measure near with the rotation vector . Since share the same word , by Lemma 2.5, we can concatenate them without extra gap words. Hence let
where is large enough to satisfy
(8) |
and . Note that such exists since
hold and the left hand side of (8) tends to as . Let be the periodic measure supported on .
First we check . Since is locally constant,
Next we check . Let . We compute
which completes the proof. ∎
Remark 2.7.
In the proof of Theorem 1, we think the word 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 does not depend on in our case. For a subshift with the specification condition, we can concatenate the words with some gap words but the error term in such case depends on , which implies we cannot choose a suitable corrective word for the error term 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 be a Markov shift with an adjacency matrix
(i.e., ) and define by
Then its rotation set is the polyhedron whose extremal points are and , where is the standard basis of . Take a rotation vector from the open side whose vertices are and , i.e., . If there exists a periodic measure , the corresponding periodic orbit should contain both of and . Since there is no sequence including and in , the word must contain the symbol . Hence we have and , 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 be a continuous map on a compact metric space. Denote by the set of all ergodic measures on . By the Riesz representation theorem a Borel probability measure on can be regarded as a bounded linear functional on . Hence we use the operator norm for an invariant measure .
First we characterize a relative maximizing measures by tangency to (3). Let and . Note that here we do not need to assume that .
Lemma 3.1.
iff is tangent to at .
Proof.
Let . Then for every we have
Let be tangent to at . For every we have
Then we can show that is an invariant probability measure in the same way as unconstrained case (See for example [Shi2018, Bre08]). We now see that takes the rotation vector . For we have
which yields .
Finally we check . Indeed,
Multiplying , we have
∎
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 and . If such that , then there exists such that and .
Proof.
Applying Bishop-Phelps’s theorem [Israel, Theorem V.1.1.] to , and we have , such that is tangent to at , and
By Lemma 3.1 we have . If , the proof is complete. Let . We can conclude and are not mutually singular by in the same way as in [Mor2010][Lemma 2.2]. Hence by the Lebesgue decomposition theorem, there exist and such that
where and . By a standard argument using the Radon-Nikodym theorem, it is easy to see (See for example [Walters]). Then for each
Hence we have
which yields . Since , we have
They should be equality since . Therefore, we obtain . ∎
Set . In the reminder of this subsection we assume
(9) |
Since the constraint requires no further discussion in the next two lemmas, we omit the proofs.
Lemma 3.3.
Let be an open set. Then
is open in .
Lemma 3.4.
Let be an open set. Then
is open in .
At the end of this subsection, we give the following two lemmas.
Lemma 3.5.
Let be a dense subset of . Then
is dense in .
Proof.
Let be an open set. Set
By Lemma 3.4, is open in . Since is dense in , . Take and let such that . Note that . Since , by Jenkinson’s theorem [Jen06][Theorem 1], there exists such that . This implies since . Hence for we have and . For sufficiently small we have , which complete the proof. ∎
Lemma 3.6.
Let be a dense subset. Then
is dense in .
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
is nonempty, open and dense subset in . Moreover, by Theorem 1, we have
which implies (9).
Fix . Since (9) holds, we can apply Lemmas 3.3 and 3.5 to our symbolic case. Therefore, we see that
is open in and
is dense in . Hence is an open dense subset of .
Then the set
is a residual subset of , 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 be a continuous map on a compact metric space with an ergodic invariant probability measure having positive entropy. Then there exist and such that for generic every relative maximizing measure of with the constraint has positive entropy.
Proof.
By Jenkinson’s theorem [Jen06][Theorem 1], there exists such that . Let . Then for all . Since is not empty, we have , which implies the unique relative maximizing measure of with constraint 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 of is prevalent if there exists a compactly supported Borel probability measure on such that the set has full -measure for every .
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 be a continuous map on a compact metric space . For a continuous constraint and a rotation vector , the set
is prevalent.
Proof.
Since is a nonempty compact set in , 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.