The Aubry set for the XY model and typicality of periodic optimization for -locally constant potentials
Abstract.
We consider the Aubry set for the XY model, symbolic dynamics with the uncountable symbol , 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, 37J511. 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 on a compact metric space and a continuous function (called as a potential) , ergodic optimization investigates the optimal (minimum) ergodic average
where is the set of -invariant Borel probability measures on endowed with the weak*-topology. An invariant measure which attains the minimum is called an optimizing (minimizing) measure for and denote by the set of optimizing measures for . Since is compact and is continuous, . 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) |
where
(2) |
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
where is the intersection of all compact sets with full measure with respect to . 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 . Let be the set of non-negative integers and let . Set the metric on by
for where is the Euclidean distance in the interval . Let be the left shift, i.e., for all , and we call the topological dynamical system the XY model. See [CR, LM14] for its details. From now on, we consider ergodic optimization for the case and . For a Lipschitz continuous potential , we define two functions the Mañé potential and Peierl’s barrier on as
where
See Section 2 for the details of and . Then we define the Aubry set as the zero-level set of , where . We remark that includes the Mather set of .
Now we give our first main results concerning Peierl’s barrier. A positive-semi orbit is said to be -static if for any non-negative integers it holds that
Define
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 be a Lipschitz function and define . Then, we have
Main Theorem 2.
For any , is a Lipschitz calibrated subaction. Moreover, the relation given by
is an equivalence relation if both and belong to .
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 with , where is a -function on and means derivative for the -th component for . Let and . Then we have (1)-(3):
-
(1)
.
-
(2)
, where is the Dirac measure supported at and stands for the set of invariant probability measures supported on a single periodic orbit.
-
(3)
, i.e., for any we have
If, in addition, has a unique minimum point in , then the Mather set of coincides with the Aubry set of and it consists of the single fixed point .
We call the assumption the twist condition. Note that Main theorem 3 holds under weaker assumptions for Lipschitz continuous on (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 -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 -functions () with the twist condition,
equipped with the -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 be an integer. For the XY model, we have the followings:
-
(i)
There is a open dense subset in such that for each the Mather set and the Aubry set of consist of a single fixed point.
-
(ii)
For arbitrary there is a open dense subset in such that for each the Mather set and the Aubry set of 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 and define by
where
We define the Mañé potential by
Remark 2.2.
The Mañé potential originates from the context of the Aubry-Mather theory for Euler-Lagrange flows. Mañé [Mane1997] considered a function defined by
(3) |
where is a closed Riemannian manifold, is a Tonelli Lagrangian (see [Mane1997] for the precise definition), is the set of absolutely continuous curves with , and is given by
with the Euler-Lagrange flow . The right-side of corresponds to a minimizing method that provides trajectories with the energy 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 is called the Aubry set of .
Note that we will see that does not take in Lemma 2.6 if 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 is called a subaction of if
for every . Moreover, a subaction is called calibrated if
for every .
Remark 2.5.
There exists a calibrated subaction for a Walters function 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 . See [BCLMS11] for the details.
The next lemma gives the lower bound of using subactions.
Lemma 2.6.
Assume be Lipschitz. For a Lipschitz subaction of we have
(4) |
for every .
Proof.
Since is Lipschitz, its calibrated subaction is also Lipschitz. Then we have
for all . Fix . Take and . Then we have
and
where is the Lipschitz constant of . Letting , we have
∎
Proposition 2.7.
is lower semicontinuous on .
Proof.
Fix and . Since
it holds that and thus we obtain
Moreover, we have
and
Thus it holds that
Therefore, the map is lower semicontinuous for each . Similarly, we can verify that the map is also lower semicontinuous for each . ∎
The following proposition asserts that the Mather set is included in the Aubry set.
Proposition 2.8.
Let be a Lipschitz function. Then the Mather set is a subset of the Aubry set .
Proof.
By the ergodic decomposition, it is sufficient to show that for each ergodic . Take arbitrary ergodic and a subaction for . Note that . Letting , we obtain
since by the -invariance of . From the definition of , it holds that , which implies that for -a.e. . Since is ergodic and is continuous, we obtain on . Hence, we see that
for each and . Note that -invariance of implies for if . By Poincaré’s recurrence theorem, for -a.e. , there exists a monotone increasing sequence with as such that . This implies that, for -a.e. , we have
i.e., . Therefore, by the lower semicontinuity of (Proposition 2.7) and the density of -a.e. points in , it holds that on . Using Lemma 2.6, we have on , which completes the proof. ∎
Next, let us define Peierl’s barrier.
Definition 2.9 (Peierl’s barrier).
For a Lipschitz function and , define by
and
Remark 2.10.
Note that the Peierl’s barrier defined as above may take . Indeed, for , we see that in the following way. Fix . Take and such that and . Since , we have
and for all . Letting , we have . Similarly, we can show that any point of the form provides . Note that any cylinder set
contains and this implies that the function
takes on a dense set in for the case . 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 be a Lipschitz function. For and we have . In particular, holds if and only if .
Proof.
The second claim immediately follows from the first claim and
for all by Lemma 2.6. Now we prove the first claim. Since , we have . It suffices to consider the following two cases:
-
(1)
For any , there exists such that
-
(2)
There exists such that for any , there exists a finite number such that
The proof of Case : Fix . By (1) and , we may assume there exists such that . Hence there exist an increasing sequence with and such that . Define by
Then for each we have
It is easy to see , and . Hence we have
Then letting we have
The proof of Case : Fix . Set a monotone decreasing positive sequence satisfying
For , we define a positive integer sequence by .
If is bounded, then there exist an integer and an infinite subsequence such that for any . Then we can take a sequence such that and
since and as . This yields because if for some , then for , we obtain
which is contradiction. Then we get
Combining this inequality and as , we have
For a -periodic sequence , we define by
Notice that for ,
and we get
Since for every , we have .
Next, we assume that is not bounded. Then we can take an increasing subsequence with and a sequence satisfying
(5) |
and for any . Set by and . Set by
Notice that for
since the first coordinates of coincide with that of . Let . For any ,
Here by (5)
for any . If ,
If , Thus we get:
Hence we get
Since for every , we have
Letting , we have
∎
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 , the map is a Lipschitz calibrated subaction.
Proof.
Note that does not take by Theorem 2.11 and . We first check the Lipschitz property. Fix . Take sequences and a monotone increasing sequence satisfying
and
respectively. Set
Since , (by replacing a subsequence if necessary) we take such that if ,
Then we obtain
Letting and , we have
Since the opposite inequality can be obtained by swapping the roles of and , the map is Lipschitz for any .
Next, we check the property of calibrated subaction. Fix and we take a sequence and a monotone increasing sequence such that
satisfying
Here, we can assume that is finite since . Notice that
Taking sufficiently large , for any ,
Hence we get
for any , and it implies
To show the opposite inequality, let be a limit of a convergent subsequence of . Letting , for sufficiently large , we have
and
Moreover, for large , we have
Letting we have
which completes the proof. ∎
The next theorem is the second half of Main Theorem 2.
Theorem 2.13 (cf. Main Theorem 2).
For and define by
Then this is an euqivalence relation on .
Before the proof of Theorem 2.13, we show the following lemma.
Lemma 2.14 (Analogy of Lemma 4.2 in [BLL13]).
For any
(6) |
Proof.
We remark that both sides of (6) may become .
Fix . Let and s.t. ,
and
Then there exist and s.t.
and there exist and s.t.
Let
Then we have
Moreover and
which implies
Since is arbitrary, we complete the proof. ∎
Proof of Theorem 2.13.
At the end of this section, we present the invariance of the above equivalent classes.
Proposition 2.15.
For any we have
In particular a equivalence class of the relation satisfies .
3. Another characterization of the Aubry set
In this section, we describe another characterization of the Aubry set. We assume that the potential is Lipschitz. A positive-semi orbit is said to be
-
•
-semi-static: if for any non-negative integers
-
•
-static: if for any non-negative integers
We call the sets
as the -semi-static set and the -static set respectively. It is trivial that and are -invariant since they are sets of positive semi-orbits. We will see that (Proposition 3.2), and that (and hence ) is not empty (Theorem 3.5). Moreover, by Proposition 2.7, we deduce that the -static set is closed and hence compact. We begin with the following.
Lemma 3.1.
For any , we have
In particular,
holds for all .
Proof.
We can prove the claim as in the proof of Lemma 2.14. Note that we do not assume that is sufficiently large in the discussion. ∎
Proposition 3.2.
For each , it holds that .
Proof.
Fix . Since holds for any , it is trivial that
By Lemma 3.1, we have . Hence, for any , we have
Therefore implies . ∎
Remark 3.3.
Before we state our main result in this section, we give the following lemma.
Lemma 3.4.
If satisfies , then
Proof.
When is a fixed point of , the Dirac measure at must be the optimal measure for and thus we have , which implies
Now we consider the case that satisfies . Fix . Take and s.t.
Since , we have
By the property of the shift map, the inequality
holds and thus we obtain
Note that is greater than 1 since . Moreover, the Lipschitz continuity of implies that
We compute
which yields
∎
Theorem 3.5 (cf. Main Theorem 1).
For each Lipschitz function , we have
Proof.
We first prove that implies . For each , we have
for by the definition of -static set. Moreover, by the definition of the Mañé potential, it holds that
Therefore, using the triangle inequality for the Mañé potential (Lemma 3.1), we obtain
Note that in the last inequality we use the fact that for any . Thus we have .
Next we see that implies . Assume that . Since each satisfies
we obtain
By the equivalence between and , we conclude that . Repeating the same discussion, we obtain for .
By Lemma 3.4, we have
Trivially it holds that
since holds for any and thus we obtain
Combining Lemma 3.1, we see that
and deduce that
Similarly, from the identities for , we have
for . Take arbitrary non-negative integers with . From the triangle inequality for the Mañé potential (Lemma 3.1),
holds and thus we obtain
Since
trivially holds, we deduce that
which yields that . ∎
Lastly, we show the relationship between Lipschitz subactions and .
Proposition 3.6.
Let be a Lipschitz subaction of a Lipschitz function . If , then
(8) |
for all . Conversely, if (8) holds for each , then .
4. Variational method applied to the Aubry set
In this section, we consider the case that the potential function is 2-locally constant, i.e.,
for some Lipschitz continuous function under suitable assumptions (see below for the precise assumptions for ). In this case, we can obtain much more explicit information about the elements in the Mather set and the Aubry set for by using variational techniques developed in [Ban88, Yu22].
Firstly, the following proposition is easily shown:
Proposition 4.1.
Let be a function depending on only the first two coordinates, i.e., for all , . Then is Lipschitz continuous as a function on with respect to the metric on if and only if is Lipschitz continuous as a function on with respect to the Euclidian metric on .
Proof.
Suppose that there exists such that for all . Taking arbitrary , we have
Conversely, suppose that there exists such that for all . Then
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 , locally constant functions are not always Lipschitz continuous with respect to the metric on .
Hereafter, we always assume and for the Lipschitz continuous function , where the assumptions and are defined by:
-
If and , then
-
If both and with are minimal, then
Here, we give the definition of the word minimal :
Definition 4.3 (Minimal).
Fix with arbitrarily. A finite word is said to be minimal if, for any with and , we have:
Moreover, an infinite word is said to be minimal if is minimal for any with .
Remark 4.4.
We state some remarks about the settings mentioned above.
-
(1)
In , the equality holds if or .
-
(2)
The assumptions and hold if satisfies the twist condition . In fact, for and , we have:
This inequality implies . Next, we show . Suppose that both and are minimal and
Clearly,
Since , is monotonically decreasing with respect to , and is monotonically decreasing with respect to . Hence, if two minimal segment satisfies and , then
which is contradiction.
-
(3)
The above notations and labels are derived from [Ban88]. In his paper, is considered as a fuction on satisfying , where and are given by:
-
for all , and
-
uniformly in .
Note that the condition holds if satisfies and the twist condition. For the proof, integrate over the triangular region bounded by three points , , and . In addition, the fact that the differentiability of is no longer needed is useful when considering geodesics on , for example (see Section in [Ban88] for the detail).
-
Let . Set and by
and
Since is non-constant by and is continuous, the set is nonempty, compact and . Firstly, we introduce the result of [Ban88] and [Yu22].
Lemma 4.5.
Fix . Then for any . Moreover, the equality is true if and only if there exists satisfying for .
This claim is essentially the same as Lemma 2.5 of [Yu22]. The proof is almost the same as the proof of the case in Theorem 3.3 of [Ban88].
Proof of Lemma 4.5.
Fix arbitrarily. It is easily seen that if is a minimizer in , i.e., satisfies:
so is for any . Set and with for each by:
(10) |
For the proof of the claim, it suffices to show the following claim:
Claim 4.6.
if is a minimizer in .
Suppose that the claim is failed, i.e., is a minimizer in and . We consider only the case . The other cases can be shown in a similar way.
Firstly, we introduce the definition of cross. We say the segments and cross if for some . Let , i.e., for . Note that and cross at least once because if not we have for all and thus the inequality yields , but the periodicity of and yields .
Suppose that and cross only once. The other cases can be shown by repeating the following discussion. Let be the largest number among such that for all . Then we have and . Define by for all . When , by , and give
As a result, it holds that:
Therefore, at least one of the following inequalities holds:
which is a contradiction (note that ). When , we have , but the equality cannot occur by and the minimality of and . Thus and hold and we have:
as seen in Remark 4.4 (1). Therefore, we obtain
This implies that both and are also minimal, which is a contradiction for . ∎
Using Lemma 4.5, we can easily show a part of our main theorem.
Lemma 4.7 (cf. Main Theorem 3. (1)).
Proof.
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 be a 2-locally constant function where is a Lipschitz continuous function on with and . Then
where 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 . Consider the distance between a point to the closed set :
The following lemma plays a key role in our statement.
Lemma 4.9 (A slightly extended version of Lemma 2.7 of [Yu22]).
Set
where
Then if .
Remark 4.10.
Let
Then should be defined as if . By the definition, if then and .
Proof of Lemma 4.9.
Lemma 2.7 in [Yu22] corresponds to the case of , i.e., for some and
but our proof is almost the same as his proof. It is sufficient to prove that for ,
-
(1)
, and
-
(2)
for all .
The first claim is clear since is given by for some . We show the second using induction. The case of is trivial. Assume that it holds if . Take any . When for any , there exists an integer such that
since . From this inequality, , and Remark 4.4 (1), we have
Set
Then
Clearly, and hold and thus we obtain the following two inequalities by Lemma 4.5:
Moreover, at least one of
and
is true, which yields at least one of the following inequalities:
By the assumption of induction for , we have
When there exists such that , then we get
The proof is complete. ∎
Using these lemmata, we can show:
Theorem 4.11 (cf. Main Theorem 3. (3)).
.
Proof.
It suffices to show that if . There exists an integer such that
since . It follows from the compactness of and continuity of that is uniformly continuous. Thus we can take satisfying if , then:
where the function is given in Lemma 4.9. Fix . Let be a large number satisfying and . For , set by
Immediately,
and
Thus we get
(11) |
It is seen that since satisfies . Moreover, by and with (so ), we see that
Hence the estimate (11) shows:
and we obtain
This is the desired inequality. ∎
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 then . Therefore, if for is a singleton , we have and it must coincide with the Mather set of . As described below, In this section, we will see that such is typical.
Let be a positive integer. Consider the set of -variational structure with the twist condition,
equipped with the -norm, i.e.,
For , let for .
Proposition 5.1.
The subset in given by
is open and dense in .
Proof.
(Dense): Take . Let be a minimum point of . Take arbitrary . Set
where with a unique minimum point at s.t. (e.g., or ). Then and . Moreover, for
Therefore, holds.
(Open): Let . Denote the corresponding minimum point of . Take Let s.t. . Then we have
i.e.,
This also implies that has a unique minimum point at , which yields . ∎
Similarly we easily obtain the following perturbation result in the sense of Mañé. Note that for any and the function satisfies the twist condition, i.e., .
Proposition 5.2.
For arbitrary , the set
is open and dense in in .
Proof of 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