Aidan Young1University of North Carolina at Chapel Hill
1[email protected]
Abstract.
We extend the theory of ergodic optimization and maximizing measures to the non-commutative field of C*-dynamical systems. We then provide a result linking the ergodic optimizations of elements of a C*-dynamical system to the convergence of certain ergodic averages in a suitable seminorm. We also provide alternate proofs of several results in this article using the tools of nonstandard analysis.
One of the guiding questions of the field of ergodic optimization is the following: Given a topological dynamical system , and a real-valued continuous function , what values can take when is an invariant Borel probability measure on , and in particular, what are the extreme values it can take? In a joint work with I. Assani [Assani-Young, Section 3], and later in [PointwiseReductionHeuristic], we noticed that the field of ergodic optimization was relevant to the study of certain temporo-spatial differentiation problems. Hoping to extend these tools to the study of temporo-spatial differentiation problems in the setting of operator-algebraic dynamical systems, this article develops an operator-algebraic formalization of this question of ergodic optimization, re-interpreting it as a question about the values of invariant states on a C*-dynamical system.
Section 1 develops the theory of ergodic optimization in the context of C*-dynamical systems, where the role of “maximizing measures" is instead played by invariant states on a C*-algebra. The framework we adopt is in fact somewhat more general than the classical framework of maximizing measures, since we consider ergodic optimizations relative to a restricted class of invariant states, which we call relative ergodic optimizations. We also demonstrate that some of the basic results of that classical theory of ergodic optimization extend to the C*-dynamical setting.
In Section 2, we define a value called the gauge of a singly generated C*-dynamical system, a non-commutative generalization of the functional of the same name defined in [Assani-Young], and describe its connections to questions of ergodic optimization, as well as the ways in which it can be used to “detect" the unique ergodicity of C*-dynamical systems under certain Choquet-theoretic assumptions.
In Section 3, we extend the results of the previous section to the case where the phase group is a countable discrete amenable group. We also provide a characterization of uniquely ergodic C*-dynamical systems of countable discrete amenable groups in terms of various notions of convergence of ergodic averages.
In Section 4, we extend some fundamental identities of ergodic optimization to the noncommutative and relative setting. We also relate the convergence properties of certain ergodic averages to relative ergodic optimizations.
Finally, in Section 5, we provide alternate proofs of several results from this article using the toolbox of nonstandard analysis.
1. Ergodic Optimization in C*-Dynamical Systems
Given a unital C*-algebra , let denote the family of all *-automorphisms of . We endow with the point-norm topology, i.e. the topology induced by the pseudometrics
This topology makes a topological group [Blackadar, II.5.5.4].
We define a C*-dynamical system to be a triple consisting of a unital C*-algebra , a topological group (called the phase group), and a point-continuous left group action .
Notation 1.1.
Let be a C*-dynamical system, and let be a nonempty finite subset. We define by
Denote by the family of all states on endowed with the weak*-topology, and by the subfamily of all tracial states on . A state on is called -invariant (or simply invariant if the action is understood in context) if for all . Denote by the family of all -invariant states on , and by the family of all -invariant tracial states on . The set (resp. ) is weak*-compact in (resp. in ). Unless otherwise stated, whenever we deal with subspaces of , we consider these subspaces equipped with the weak*-topology.
We will assume for the remainder of this section that is a C*-dynamical system such that is separable, and also that . This framework will include every system of the form , where is a compact metrizable topological space, the group is countable, discrete, and amenable, and is of the form for all , where is a right action of on by homeomorphisms. Because of the correspondence between topological dynamical systems as we’ve defined them previously in Section LABEL:Topological_stuff and C*-dynamical systems over commutative C*-algebras, it is customary to call a C*-dynamical system a “non-commutative topological dynamical systems."
Before proceeding, we prove the following Krylov–Bogolyubov-type result, which will be useful to establish the -invariance of certain states later.
Lemma 1.2.
Let be a C*-dynamical system, and let be an amenable group. If is a sequence in , and is a right Følner sequence for , then any weak*-limit point of the sequence is -invariant. In particular, if is a nonempty, -invariant, weak*-compact, convex subset of , then .
Proof.
Let be a sequence of states, and fix . Then
Therefore, if is such that exists, then
Finally, let be a nonempty, -invariant, weak*-compact, convex subset of . Let be any state in , and consider the sequence . By the convexity and -invariance of , every term of this sequence is an element of , and since is compact, there exists a subsequence of this sequence which converges in . As has already been shown, that limit must be an element of .
∎
Remark 1.3.
Lemma 5.1 can be seen as a nonstandard-analytic analogue to Lemma 1.2.
Although our manner of proof of Lemma 1.2 is scarcely novel, the result as we have stated it here can be used to ensure the existence of invariant states with specific properties that might interest us, as seen for example in Corollary 1.4 and Proposition 1.14. Our standing hypothesis that be separable is not necessary for this proof of Lemma 1.2.
We denote by the real Banach space of all self-adjoint elements of , and denote by the space of all real self-adjoint bounded linear functionals on .
Definition 1.6.
Let be a locally convex topological real vector space, and let be a compact subset of which is contained in a hyperplane that does not contain the origin. We call a simplex if the positive cone defines a lattice ordering on with respect to the partial order .
Remark 1.7.
In Definition 1.6, the assumption that lives in a hyperplane that does not contain the origin is technically superfluous, but simplifies the theory somewhat (see [Phelps, Section 10]), and is satisfied by all the simplices that interest us here. Specifically, we know that (and by extension ) lives in the real hyperplane defined by the evaluation at .
We begin with the following lemma.
Lemma 1.8.
(i)
The spaces are compact and metrizable.
(ii)
If , then the space is a simplex.
Before proving this lemma, we need to introduce some terminology. Let be two positive linear functionals on a unital C*-algebra . We say that the two positive functionals are orthogonal, notated , if they satisfy either of the following two equivalent conditions:
(a)
.
(b)
For every exists positive of norm such that .
It is well-know that these conditions are equivalent [Pedersen, Lemma 3.2.3]. For every , there exist unique positive linear functionals such that , and , called the Jordan decomposition of [Blackadar, II.6.3.4].
Before proving Lemma 1.8, we demonstrate the following property of the Jordan decomposition of a tracial functional.
Lemma 1.9.
Let be a unital C*-algebra, and . Suppose that for all . Then for all .
Proof.
Let denote the group of unitary elements in . For a unitary element , let denote the inner automorphism
Let . We claim that is tracial if and only if for all unitaries .
Let be unitary, and an arbitrary element. Then
So if and only if .
In one direction, suppose that for all Fix . Then we can write for some and unitaries unitary. Then
Thus is tracial.
In the other direction, suppose there exists such that . Let such that , and let . Then
Therefore is not tracial.
Now, if is tracial, then for all . Then . But , so it follows that . Therefore is an orthogonal decomposition of , and so it is the Jordan decomposition. This means that . Since this is true for all , it follows that are tracial.
∎
This all follows because is a weak*-closed real subspace of the unit ball in the continuous dual of the separable Banach space , and the spaces are all closed subspaces of .
(ii)
It is a standard fact that if , then is a simplex [Blackadar, II.6.8.11]. Let
be the positive cone of , and let denote the (real) space of all bounded self-adjoint tracial linear functionals on . Let denote the (real) space of all bounded self-adjoint -invariant linear functionals on . We already know that lives in a hyperplane of defined by the evaluation functional . It will therefore suffice to show that , and that is a sub-lattice of .
Let be positive functionals on such that is tracial, and . By Lemma 1.9, we know that are tracial. We claim that if , then . To prove this, let , and consider that are both positive linear functionals such that .
We claim that . Fix . We know that there exists such that , and such that . Then is a positive element of norm such that
Therefore is a Jordan decomposition of , and since the Jordan decomposition is unique, it follows that , i.e. that . This means that .
We now want to show that is a sublattice of , i.e. that it is closed under the lattice operations. Let . For this calculation, we draw on the identities listed in [PositiveOperators, Theorem 1.3]. Then
Therefore, if is a real linear space and is closed under the operations , then it is also closed under the lattice operations. Thus is a sublattice of .
Hence, the subset is a compact metrizable simplex.
∎
In order to keep our treatment relatively self-contained, we define here several elementary concepts from Choquet theory that will be relevant in this section.
Definition 1.10.
Let be convex spaces. We call a map an affine map if for every , we have
In the case where , we call an affine functional.
Definition 1.11.
Let be a convex subset of a locally convex real topological vector space .
(a)
A point is called an extreme point of if for every pair of points and parameter such that , either or . In other words, we call extreme if there is no nontrivial way of expressing as a convex combination of elements of .
(b)
The set of all extreme points of is denoted .
(c)
A subset of is called a face if for every pair such that , we have that .
(d)
A face of is called an exposed face of if there exists a continuous affine functional such that for all , and for all .
(e)
A point is called an exposed point of if is an exposed face of .
(f)
Given a subset of , the closed convex hull of is written as .
We now introduce the basic concepts in our treatment of ergodic optimization.
Definition 1.12.
Let be a self-adjoint element, and let be a compact convex subset of . Define a value by
We say a state is -maximizing if . Let denote the set of all -maximizing states. A state is called uniquely -maximizing if .
Remark 1.13.
We note here a trivial inequality: If are compact convex subsets of , then , and in particular, we will always have .
We will single out one type of compact convex subset of which will prove important later. Given a subset , set
When is a -invariant closed ideal of , we have a bijective correspondence between the states in and the states on invariant under the action induced by . We will be referring to this set again in Sections 2 and 3, when values of the form come up in reference to certain ergodic averages. We observe that , and that . There is also no a priori guarantee that , since for example . However, Proposition 1.14 gives sufficient conditions for to be nonempty.
Proposition 1.14.
Let be such that for all . Suppose there exists a state on which vanishes on . Then . In particular, if is a proper closed two-sided ideal of for which for all , then .
Proof.
Let denote the family of all (not necessarily invariant) states on which vanish on . Then if and , then , so vanishes on . Therefore for all . It follows from Lemma 1.2 that .
Suppose is a proper closed two-sided ideal of for which for all , and let be the canonical quotient map. Let be the induced action of on by . Let be a -invariant state on . Then is a -invariant state on which vanishes on , i.e. .
∎
Proposition 1.15.
Let be a nonempty compact convex subset of , and let . Then is a nonempty, compact, exposed face of .
Proof.
To see that is nonempty, for each , let such that . Then since is compact, the sequence has a convergent subsequence. Let be the limit of a convergent subsequence of . Then is -maximizing.
To see that is compact, consider that
which is a closed subset of . As for being an exposed face, consider the continuous affine functional given by
Then the functional exposes , since it is nonpositive on all of and vanishes exactly on .
∎
The following result describes the ways in which some ergodic optimizations interact with equivariant *-homomorphisms of C*-dynamical systems.
Theorem 1.16.
Let be two C*-dynamical systems, and let be a surjective *-homomorphism such that
Let denote the space of -invariant states on . Then .
Proof.
Let denote the space of -invariant states on . We claim that there is a natural bijective correspondence between and . If is a -invariant state on , then we can pull it back to a -invariant state on by
This obviously vanishes on , and is -invariant by virtue of the equivariance property of . Conversely, if we start with a -invariant state on that vanishes on , then we can push it to a -invariant state on by
We claim now that
Let be a -maximizing state on . Then , so
On the other hand, if is -maximizing, then let be such that . Then , so
∎
The assumption in Theorem 1.16 that is surjective is actually superfluous, as shown in Corollary 3.8. We will later provide a proof of this stronger claim that uses the gauge functional, introduced in the context of actions of in Section 2 and in the context of actions of amenable groups in Section 3.
Moreover, the proof of Theorem 1.16 can be extended to establish a correspondence between ergodic optimization over certain compact convex subsets of and certain compact convex subsets of . For example under the same hypotheses, if , then the proof could be modified in a simple manner to establish that , where denotes the -invariant tracial states on . In lieu of stating Theorem 1.16 in greater generality, we content ourselves to state this special case (which we will use in future sections) and remark that the argument can be generalized further.
The following characterization of exposed faces in compact metrizable simplices will prove useful.
Lemma 1.17.
Let be a compact metrizable simplex. Then every closed face of is exposed.
Proof.
See [Davies, Theorem 7.4].
∎
The theorem we are building to in this section is as follows.
Theorem 1.18.
Let be a compact simplex. Then the closed faces of are exactly the sets of the form for some .
Before we can prove our main theorem of this section, we will need to prove the following result, which gives us a means by which to build an important linear functional.
Theorem 1.19.
Let be a compact simplex, and let be a continuous affine functional. Then there exists a continuous linear functional such that .
To prove this theorem, we break it up into several parts, attaining the extension as the final step of a few subsequent extensions of .
Lemma 1.20.
Let be a compact metrizable simplex, and let be a continuous affine functional. Let . Then there exists a continuous functional satisfying the following conditions for all :
(a)
,
(b)
,
(c)
.
Proof.
Note that every nonzero element of can be expressed uniquely as for some . As such we define
It is immediately clear that this satisfies conditions (a) and (c), leaving only (b) to check.
Now, suppose that for some . Consider first the case where at least one of are nonzero.
Then
[because is affine]
In the event that , then the additivity property attains trivially.
It remains now to show that is continuous. We will check continuity at nonzero points in , and then at . First, consider the case where , and . Suppose that is a sequence in converging in the weak*-topology to . We claim that in , and in the weak*-topology.
We first observe that , so converges in to , meaning in particular that for sufficiently large , we have that . Now, if is a norm-continuous linear functional, then
Therefore . Thus we can compute
where must be finite because is weak*-compact, and because is weak*-continuous.
Now, suppose that converges to . Then again we have that by the same argument used above (i.e. ). Therefore
We can thus conclude that is weak*-continuous.
∎
Lemma 1.21.
Let be as in Lemma 1.20, and let . Then there exists a continuous linear functional such that .
Proof.
Define by
where are meant in the sense of the lattice structure possesses by virtue of being a simplex.
Our first claim is that if such that , then . To see this, we observe that . Therefore
This makes linearity fairly straightforward to check. First, to confirm additivity, let . Then , where . Thus
To check homogeneity, let . If , then , and ; on the other hand, if , then , and . In both cases, homogeneity is straightforward to show. This proves that is linear.
It is also quick to show that , since if , then , so .
It remains now to show that is continuous. By [RudinFunctional, Theorem 1.18], it will suffice to show that is weak*-closed. To prove the kernel is closed, let be a sequence in converging in the weak*-topology to . By the Uniform Boundedness Principle, it follows that . By rescaling, we can assume without loss of generality that for all , and since the unit ball is weak*-closed by Banach-Alaoglu, we can infer that .
Since the unit ball is weak*-compact, it follows that the sequences have convergent subsequences. Let be a subsequence along which . Then if , then
Therefore , so
Therefore, we can conclude that is weak*-continuous.
∎
Let be a closed face of . By Lemma 1.17, the face is exposed, so let be a weak*-continuous affine functional such that
Set
and let be a continuous linear extension of to whose existence is promised by Theorem 1.19. We can then extend to a weak*-continuous linear functional [PositiveOperators, Theorem 3.6]. There thus exists some such that for all [Baggett, Theorem 5.2]. In particular, we have for all . Therefore .
In particular, we can recover the following corollary.
Corollary 1.22.
If , then there exists such that is uniquely -maximizing, i.e. such that .
Proof.
The singleton is a closed face, and by Lemma 1.17 is therefore an exposed face. Apply Theorem 1.18.
∎
We have developed the language of ergodic optimization here in a somewhat atypical way, where we speak not of -maximizing states simpliciter, but of a state that is maximizing relative to a compact convex subset of , especially a compact simplex . This notion of relative ergodic optimization has precedent in [ObservableMeasures]. For our purposes, this relative ergodic optimization means we can consider ergodic optimization problems over different types of states. In Section 4, we will broaden our scope somewhat to consider ergodic optimization in the noncommutative setting relative to a set of states that aren’t necessarily -invariant.
Since Theorem 1.18 applies in cases where is a simplex, we will conclude this section by describing some situations where is a compact metrizable simplex.
For each , let be the GNS representation corresponding to . Define a unitary representation of by
extending this from to . Set
Let be the orthogonal projection (in the functional-analytic sense) of onto . We call the C*-dynamical system a -abelian system if for every , the family of operators is mutually commutative.
We record here a handful of germane facts about -abelian systems.
Proposition 1.23.
If is -abelian, then is a simplex.
Proof.
See [Sakai, Theorem 3.1.14].
∎
Definition 1.24.
We call a system asymptotically abelian if there exists a sequence in such that
for all , where is the Lie bracket on .
Proposition 1.25.
If is asymptotically abelian, then it is also -abelian.
Proof.
See [Sakai, Proposition 3.1.16].
∎
2. Unique ergodicity and gauges: the singly generated setting
So far we have spoken about C*-dynamical systems, a noncommutative analog of a topological dynamical systems. But just as classical ergodic theory is often interested in the interplay between topological dynamical systems and the measure-theoretic dynamical systems they can be realized in, we are interested in questions about the interplay between C*-dynamical systems and the non-commutative measure-theoretic dynamical systems they can be realized in. To make this more precise, we introduce the notion of a W*-dynamical system.
A W*-probability space is a pair consisting of a von Neumann algebra and a faithful tracial normal state on . An automorphism of a W*-probability space is a *-automorphism such that , i.e. an automorphism of which preserves . A W*-dynamical system is a quadruple , where is a W*-probability space, and is a left action of a discrete topological group (called the phase group) on by -preserving automorphisms of , i.e. such that for all . Importantly, if is a W*-dynamical system, then is automatically a W*-dynamical system.
Remark 2.1.
In the literature, the term “W*-dynamical system" is sometimes used to refer to a more general construction, where the group is assumed to satisfy some topological conditions, and the action is assumed to be continuous in the strong operator topology, e.g. [NoncommutativeJoinings]. Other authors use a yet more general definition, e.g. [Blackadar, III.3.2]. Since we are only interested in actions of discrete groups, we adopt a narrower definition.
Definition 2.2.
Given a W*-probability space, we define to be the Hilbert space defined by completing with respect to the inner product , i.e. the Hilbert space associated with the faithful GNS representation of induced by .
Finally, we introduce the notion of a C*-model, intending to generalize the notion of a topological model from classical ergodic theory to this noncommutative setting.
Definition 2.3.
Let be a W*-dynamical system. A C*-model of is a quadruple consisting of a C*-dynamical system and a *-homomorphism such that
(a)
is dense in the weak operator topology of ,
(b)
for all , and
(c)
for all .
We call the C*-model faithful if is also injective.
We remark that we can turn any C*-model into a faithful C*-model through a quotienting process. If was not injective, then we could instead consider . In the case where is commutative, this quotienting process corresponds (via the Gelfand-Naimark Theorem) to taking a measure-theoretic dynamical system and restricting to the support of the resident probability measure. To see this, let , where is a compact metrizable topological space, and let for some Borel probability measure . Let be the (not necessarily injective) map that maps a continuous function on to its equivalence class in . It can be seen that if and only if the open set is of measure , or equivalently if , and in particular that is injective if and only if is strictly positive (i.e. assigns positive measure to all nonempty open sets). As such, we can identify with . Let denote the support of on , and let be the quotient map (which corresponds to a restriction from to , i.e. ). Then algebraically, we have a commutative diagram
So in the commutative case, we can make injective by looking at , i.e. by using the support to model .
Importantly, so long as is separable, any W*-dynamical system will admit a faithful separable C*-model. To construct such a C*-model, it suffices to take some separable C*-subalgebra which is dense in with respect to the weak operator topology, then let be the norm-closure of the span of . We then define and let be the inclusion map.
One last important concept in this section and the next will be unique ergodicity. A C*-dynamical system is called uniquely ergodic if is a singleton. As in the commutative setting, unique ergodicity can be equivalently characterized in terms of convergence properties of ergodic averages. To our knowledge, the strongest such characterization of unique ergodicity for singly generated C*-dynamical systems can be found in [AbadieDykema, Theorem 3.2], which describes unique ergodicity relative to the fixed point subalgebra. This characterization was then generalized to characterize unique ergodicity relative to the fixed point subalgebra for C*-dynamical systems over amenable phase groups in [DuvenhageStroeh, Theorem 5.2]; however, in Corollary 3.6, we provide a characterization of uniquely ergodic C*-dynamical systems in terms of ergodic averages that is not encompassed by [DuvenhageStroeh, Theorem 5.2].
Given a C*-dynamical system , let be a positive element. We define the gauge of to be
To prove this limit exists, it suffices to observe that the sequence is subadditive, since
Therefore, by the Subadditivity Lemma, the sequence converges, and we have the equality
We have the following characterization of in the language of ergodic optimization.
Theorem 2.4.
Let be a C*-dynamical system. Then if is a positive element, then .
Proof.
For each , choose a state on such that
Let , so
Let be a weak*-limit point of , and let be a subsequence such that in the weak*-topology. By Lemma 1.2, we know that is -invariant. Therefore , and is a -invariant state on , so
Now, we prove the opposite inequality. Let . Then
Therefore
This establishes the identity.
∎
Corollary 2.5.
Let be a W*-dynamical system, and let be a C*-model of . If is a positive element, then
Proof.
Write , and let be the action obtained by restricting to . Write for the space of -invariant states on .
We can write . By Theorem 2.4, we know that , and by Theorem 1.16, we know that .
∎
Remark 2.6.
Corollary 2.5 can be regarded as an operator-algebraic extension of Lemma 2.3 from [Assani-Young]. The assumption that is faithful can be understood as analogous to the assumption of strict positivity in that paper.
This value provides an alternative characterization of unique ergodicity, at least under some additional Choquet-theoretic hypotheses.
Theorem 2.7.
Let be a W*-dynamical system, and let be a faithful C*-model of . Then the following conditions are related by the implications (i)(ii)(iii).
(i)
The C*-dynamical system is uniquely ergodic.
(ii)
The C*-dynamical system is strictly ergodic.
(iii)
for all positive .
Further, if is a simplex, then (iii)(i).
Proof.
(i)(ii) Suppose that is uniquely ergodic. Then is an invariant state on , so it follows that is the unique invariant state on . But is also a faithful state on , so it follows that is strictly ergodic.
(ii)(i) Trivial.
(i)(iii) Suppose that is uniquely ergodic, and let be positive. Let be a -maximizing state for . Then , since both and are invariant states on , and is uniquely ergodic. Thus , so .
(iii)(i) Suppose that is a simplex, but that is not uniquely ergodic. By the Krein-Milman Theorem, there exists two distinct extreme points of , and in particular there exists an extreme point of distinct from . Then by Corollary 1.22, there exists self-adjoint such that . We can assume that is positive, since otherwise we could replace with for a sufficiently large positive real number , and . Then . But by the assumption that is uniquely -maximizing, it follows that . Therefore , meaning that (iii) does not attain. Thus (i)(iii).
∎
3. Unique ergodicity and gauge: the amenable setting
For the duration of this section, we assume that is a W*-dynamical system with separable. Assume further that is a C*-dynamical system such that is separable, and that is amenable. It follows from Corollary 1.4 that .
In this section, we expand upon some of the ideas presented in Section 2, generalizing from the case of actions of to actions of a countable discrete amenable group . We separate these two sections because our treatment of the more general amenable setting has some additional nuances to it.
Our first result of this section is a generalization of a classical result from ergodic theory regarding unique ergodicity, which is that a (singly generated) topological dynamical system is uniquely ergodic if and only if the averages of the continuous functions converge to a constant. This classical result is well-known, and can be found in many standard texts on ergodic theory, e.g. [DajaniDirksin, Thm 6.2.1], [EisnerOperators, Thm 10.6], [Walters, Thm 5.17], but the earliest example of a result like this that we could find was [OxtobyErgodic, 5.3]. Theorem 3.1 generalizes this classical result not only to the noncommutative setting, but to the setting where the phase group is amenable.
We define the weak topology on a C*-algebra to be the topology generated by the states on , i.e.
In other words, the weak topology is the topology in which a net converges to if and only if converges to for every state on . We say the net converges weakly to if it converges in the weak topology.
Theorem 3.1.
Let be a C*-dynamical system. Then the following conditions are equivalent.
(i)
is uniquely ergodic.
(ii)
There exists a right Følner sequences for and a linear functional such that for all , the sequence converges in norm to .
(iii)
There exists a left Følner sequences for and a linear functional such that for all , the sequence converges weakly to .
(iv)
There exists a state on such that for every right Følner sequence for , the sequence converges in norm to .
(v)
There exists a state on such that for every left Følner sequence for , the sequence converges weakly to .
Proof.
Assume throughout that any is nonzero.
(ii)(iii) Follows from the existence of two-sided Følner sequence.
(iv)(v) Follows from the existence of two-sided Følner sequence.
(iv)(ii) Trivial.
(v)(iii) Trivial.
(iii)(i) Suppose that weakly for all . We claim that is the unique invariant state of . First, we demonstrate that is -invariant. Fix , and fix . Choose such that
The exist because we know that in the weak topology, the functionals are both continuous, and exists by the amenability of . Let . Then if , then
Therefore is -invariant. To see that it is positive, it suffices to observe that , meaning that . To see that , we just observe that for all .
Now we show that is the unique -invariant state. Let be any invariant state. Then
Therefore , and so is uniquely ergodic.
(i)(iv) Fix a right Følner sequence , and assume for contradiction that is uniquely ergodic with -invariant state , but that there exists such that does not converge in norm to a scalar, and in particular does not converge in norm to . Since we can decompose into its real and imaginary parts, we can assume that . Fix for which there exists an infinite sequence such that . Then for each exists a state on such that .
Set
so . Then has a subsequence, call it which converges in the weak*-topology to some . This is also a state on , and by Lemma 1.2, we know is -invariant. But , since
This contradicts being uniquely ergodic.
∎
Remark 3.2.
Although [DuvenhageStroeh, Theorem 5.2] describes conditions under which unique ergodicity of an action of an amenable group on a C*-algebra can be related to the convergence of ergodic averages, that result is not a direct generalization of our Theorem 3.1.
In order to develop the gauge machinery from the previous section in the context of actions of amenable groups, we will need to use slightly different techniques, since we do not have access to the Subadditivity Lemma. The main results of the remainder of this section can be summarized as follows.
Main results 3.3.
Let be a right Følner sequence.
(a)
Let be a C*-dynamical system, and let be a right Følner sequence for . Then if is a positive element, then the sequence
converges to .
(b)
Let be a faithful C*-model of . Then the following conditions are related by the implications (i)(ii)(iii).
(i)
The C*-dynamical system is uniquely ergodic.
(ii)
The C*-dynamical system is strictly ergodic.
(iii)
for all positive .
Further, if is a simplex, then (iii)(i).
Theorem 3.4.
Let be a C*-dynamical system, and let be a right Følner sequence for . Then if is a positive element, then the sequence
converges to .
Proof.
For each , choose a state on such that
Let , so
This means that in order to show that converges to , it suffices to show that . So for the remainder of this proof, we are going to be looking instead at the sequence .
Let be some sequence such that converges in the weak*-topology to some . It follows from Lemma 1.2 that is -invariant. To see that converges to , it will suffice to show that every limit point of satisfies
This follows because if there existed a subsequence of such that , then by compactness, that subsequence would have some subsequence converging to some for which , meaning in particular that .
So let be some sequence such that converges in the weak*-topology to some . As has already been remarked, we have that , so . We prove the opposite inequality. Let . Then
Therefore . This establishes the desired identity.
∎
Remark 3.5.
An alternate proof of Theorem 3.4 using nonstandard analysis is presented in Section 5.
Corollary 3.6.
Let be a C*-dynamical system, and let be a right Følner sequence for . Let . Then is uniquely ergodic if and only if for every positive element , the sequence
converges to .
Proof.
() Suppose is uniquely ergodic. Then for all positive , so by Theorem 3.4 .
() We’ll prove the contrapositive. Suppose is not uniquely ergodic. Then there exists an extreme point of different from . By Corollary 1.22, there exists self-adjoint such that . We can assume that is positive, replacing by for a sufficiently large positive real number otherwise. Thus .
∎
Definition 3.7.
Given a C*-dynamical system , a positive element , and a right Følner sequence for , we define the gauge of to be the limit
Theorem 3.4 shows that the gauge exists, but Theorem 3.9 demonstrates the way that the gauge interacts with a W*-dynamical system and a C*-model. Moreover, the gauge is dependent only on , and independent of the right Følner sequence . As such, even though the gauge as we have described it is computed using a right Følner sequence , we do not need to include in our notation for .
Corollary 3.8.
Let be two C*-dynamical systems, and let be a *-homomorphism (not necessarily surjective) such that
Let denote the space of -invariant states on . Then .
Proof.
Let , and let be the action . Let denote the space of all -invariant states on . Then
∎
Corollary 3.9.
Let be a W*-dynamical system, and let be a C*-model of . Then if is a positive element, then
Proof.
Write , and let be the action obtained by restricting to . Write for the space of -invariant states on .
We know . By Theorem 3.4, we know that , and by Theorem 1.16, we know that
∎
This brings us to our characterization of unique ergodicity with respect to the gauge for C*-models.
Theorem 3.10.
Let be a W*-dynamical system, and let be a faithful C*-model of . Then the following conditions are related by the implications (i)(ii)(iii).
(i)
The C*-dynamical system is uniquely ergodic.
(ii)
The C*-dynamical system is strictly ergodic.
(iii)
for all positive .
Further, if is a simplex, then (iii)(i).
Proof.
(i)(ii) Suppose that is uniquely ergodic. Then is an invariant state on , so it follows that is the unique invariant state. But is also faithful, so it follows that is strictly ergodic.
(ii)(i) Trivial.
(i)(iii) Suppose that is uniquely ergodic, and let be positive. Let be an -maximizing state on . Then , since both are invariant states and is uniquely ergodic. Then , so .
(iii)(i) Suppose that is a simplex, but that is not uniquely ergodic. Let be an extreme point of different from . Then by Corollary 1.22, there exists self-adjoint such that . We can assume that is positive, since otherwise we could replace with for a sufficiently large positive real number , and . Then . But by the assumption that is uniquely -maximizing, it follows that . Therefore , meaning that (iii) does not attain. Thus (i)(iii).
∎
4. A noncommutative Herman ergodic theorem
For the duration of this section, we assume that is a C*-dynamical system such that is separable, and that is amenable.
Let be a right Følner sequence for . Write to denote the set of all limit points of sequences of the form , where for all . Because is right Følner, we know from Lemma 1.2 that if is nonempty, then will be a nonempty compact subset of . In particular, if , then for any choice of . Moreover, if is convex and -invariant, then .
Question 4.1.
Is dependent on in general?
We now define two quantities.
Notation 4.2.
Let be a right Følner sequence for , and a nonempty subset of . Let . Define
The values can be compared to the and quantities presented in Section 2 of [Jenkinson], respectively. Ergodic optimization is concerned with finding the extrema of sequences of ergodic averages of real-valued functions, but there are several ways we might attempt to formalize what an “extremum" of a sequence of ergodic averages would be. In [Jenkinson], O. Jenkinson proposes several different ways we might formalize this notion, then demonstrates that they are equivalent under reasonable conditions [Jenkinson, Proposition 2.1]. Our Proposition 4.3 is an attempt to extend some part of this result to the noncommutative and relative setting.
Proposition 4.3.
The quantities are well-defined when is compact, convex, and -invariant. Moreover, they satisfy
Proof.
We’ll prove that , as the proof that is very similar. We know a priori that .
Let be a sequence in such that for each , we have
We know that any limit point of is in . Let be chosen such that . We can assume that is weak*-convergent to a state , passing to a subsequence if necessary. Then
Assume for contradiction that . Let be such that . Then
Let such that converges to . Then
a contradiction. Therefore we conclude that . Thus
Thus we can conclude that is well-defined and equal to .
∎
Remark 4.4.
An alternate proof of Proposition 4.3 using nonstandard analysis is presented in Section 5.
To our knowledge, the first result like Theorem 4.5 is [Herman, Lemme on pg. 487]. Herman’s result can be understood as an extension of the classical result that a topological dynamical system is uniquely ergodic if and only if the ergodic averages of all continuous functions converge uniformly to a constant. To our knowledge, the first record of this classical result is [OxtobyErgodic, (5.3)]. If Oxtoby’s result can be understood as relating the uniform convergence properties of ergodic averages of all continuous functions to the ergodic optimization of all continuous functions, then Herman’s result relates the uniform convergence properties of ergodic averages of a single continuous function to its ergodic optimization. Our result extends Herman’s in a few directions. First, it extends Herman’s result to the setting of actions of amenable groups other than . Moreover, it extends the result to C*-dynamical systems. Finally, it allows us to relate convergence in certain seminorms to relative ergodic optimizations.
Let be a C*-dynamical system, where is an amenable group. Given a nonempty subset of , define the seminorm on by
Theorem 4.5.
Let be a right Følner sequence for , and . Let , and . Then the following are equivalent.
(i)
.
(ii)
.
Proof.
(i)(ii): We prove the contrapositive. Suppose there exists and such that
For each , choose such that . Then in particular we know that
By the weak*-compactness of , there must exist a weak*-convergent subsequence of . Assume without loss of generality that converges in the weak* topology, and write . Then
Therefore , meaning that .
(ii)(i): Suppose that . Let be a sequence in , and let be a weak*-convergent subsequence of with limit . Then
Therefore .
∎
Remark 4.6.
An alternate proof of Theorem 4.5 using nonstandard analysis is presented in Section 5.
Corollary 4.7.
Let be a right Følner sequence for . Let , and . Then the following are equivalent.
(i)
.
(ii)
.
Proof.
Apply Theorem 4.5 in the case where , implying that and .
∎
Corollary 4.7 strengthens the noncommutative analogue of Oxtoby’s characterization of unique ergodicity, as we see below.
Corollary 4.8(A noncommutative extension of Oxtoby’s characterization of unique ergodicity).
Let be a right Følner sequence for . A C*-dynamical system is uniquely ergodic if and only if converges in norm to an element of for all .
Proof.
: By taking real and imaginary parts, we can reduce to the case where is self-adjoint. If is uniquely ergodic, then is singleton, so by Corollary 4.7 the averages will converge to a scalar.
: Conversely, if is not uniquely ergodic, then there exist two states for which there exists such that , implying that is not singleton. Corollary 4.7 then tells us that doesn’t converge in norm.
∎
5. Applications of nonstandard analysis to noncommutative ergodic optimization
The tools of nonstandard analysis can be used to provide alternate proofs of some results in this article. In this section, we assume that the reader is familiar with the basic tools and vocabulary of nonstandard analysis. See [Goldblatt] for references. Since some of the terminology of the field is not entirely universal, we define some of the less universal terms here.
We will assume throughout this section that is a C*-dynamical system, and that is a universe that contains . Assume that is a countably saturated universe embedding. We say that is unlimited if for all , and limited otherwise. Let denote the external ring of limited elements of . For , we write if for all . This is an equivalence relation on . We define the shadow to be the -linear functional mapping to the unique (standard) complex number for which . The shadow is also order-preserving on . Let denote the unlimited hypernaturals.
We have the following nonstandard analogue of Lemma 1.2.
Lemma 5.1.
Let be a C*-dynamical system, and let be an amenable group. Consider a sequence in in , and a right Følner sequence for . Let be an unlimited hypernatural, and define a state by
Then is a well-defined -invariant state, and is a limit point of the sequence .
Proof.
First, we take up the well-definedness of . If , then
and so by the Transfer Principle
In particular, it follows that , meaning that . Thus is well-defined. We can similarly prove that is positive and unital by applying the Transfer Principle to the sentences
To prove the -invariance of , we recall from a familiar argument (see proof of Lemma 1.2) that if , then
It follows from a classical result of nonstandard analysis [Goldblatt, Theorem 6.1.1] that , meaning that .
Finally, we argue that is a limit point of . For , consider the sentence given by
Then is true for all , witnessed by . Therefore, it follows from the Transfer Principle that is true for all . We know that
is a neighborhood basis for in the weak* topology. Thus we have shown that is a limit point of the sequence .
∎
We might ask whether Lemma 5.1 is strictly weaker than Lemma 1.2, since Lemma 5.1 also asserts that the state it describes is a limit point of the sequence that generates it. In fact, the two lemmas are equivalent in the sense that for a sequence in , every limit point of the sequence can be written as for some . To see this, choose such that exists. Let be a countable neighborhood basis for in the weak*-topology, and for each , let be the set
Then has the finite intersection property, and so by the countable saturation of our universe embedding, it follows that there exists such that
which is necessarily unlimited. Then for any , we have that for all , so
This correspondence can be generalized in the following result.
Proposition 5.2.
Let be a compact Hausdorff topological space, and let be a countably saturated extension of a universe containing and . Let be the equivalence relation on defined by
Define a map that sends to the unique such that , and let be a sequence in . Then the map is well-defined.
Further, set
Then
In addition, if is -saturated for some uncountable cardinal , where is some topological basis of , then
Proof.
The fact that in a compact topological space, for every exists some such that can be found in [AppliedNSA, Theorem 1.6 of Chapter 3]. As for the uniqueness, assume for contradiction that there existed such that . Using the Hausdorff property, choose open neighborhoods such that . Then , so by the Transfer Principle . Thus , because . Therefore by the Transfer Principle, a contradiction. Thus is well-defined.
Let , and consider . Let , where is a topological basis for , and consider for the sentence defined by
Then is true for all , since and for all , so it follows that is true for all . Since forms a neighborhood basis for , it follows that .
Now suppose that is -saturated for some uncountable cardinal , and let . Let . For , consider the set
Then has the finite intersection property, and thus there exists . Thus for all , and . Thus .
∎
Remark 5.3.
Our definitions of and in the statement of Proposition 5.2 is consistent with our definition of on in the following sense. We can write . If , then there exists such that . Then . The set is compact, and the definition of on that compact space in the sense of Proposition 5.2 will agree with our definition of on from the start of this section.
In light of Theorem 5.2, several compactness arguments in this article can be proven alternatively in the language of nonstandard analysis. Here we provide a few examples.
(i)(ii): Suppose . For each , choose such that . Fix , and let be the state
Lemma 5.1 tells us that . Thus . Therefore for all , meaning a classical result of nonstandard analysis [Goldblatt, Theorem 6.1.1] tells us that . But because for all , we can conclude that .
(ii)(i): Suppose that . Let be a sequence in , and let be the state
This paper is written as part of the author’s graduate studies. He is grateful to his beneficent advisor, professor Idris Assani, for no shortage of helpful guidance.
An earlier version of this paper referred to “tempero-spatial differentiations." Professor Mark Williams pointed out that the more correct portmanteau would be “temporo-spatial." We thank Professor Williams for this observation.